Skip to main content

Full text of "The Wilson bulletin"

See other formats


HARVARD  UNIVERSITY 


Ernst  Mayr  Library 
of  the  Museum  of 
Comparative  Zoolop'y 


-e 


Tlie  WIson  Bulletin 

PUBLISHED  BY  THE  WILSON  ORNITHOLOGICAL  SOCIETY 
VOL.  Ill,  NO.  1 MARCH  1999  PAGES  1-156 

(ISSN  (KU3-5643) 


THE  WILSON  ORNITHOLOGICAL  SOCIETY 
FOUNDED  DECEMBER  3,  1888 

Named  after  ALEXANDER  WILSON,  the  first  American  Ornithologist. 

President— Edward  H.  Burtt,  Jr.,  Department  of  Biology,  Ohio  Wesleyan  University,  Delaware,  Ohio 
43015;  E-mail:  EHBurtt@cc.owu.edu. 

First  Vice-President— John  C.  Kricher,  Biology  Department,  Wheaton  College,  Norton,  Massachusetts 
02766;  E-mail:  JKricher@wheatonma.edu. 

Second  Vice-President— William  E.  Davis,  Jr.,  College  of  General  Studies,  871  Commonwealth  Ave., 
Boston  University,  Boston,  Massachusetts  02215;  E-mail:  WEDavis@bu.edu. 

Editor— Robert  C.  Beason,  Department  of  Biology,  State  University  of  New  York,  1 College  Circle, 
Geneseo,  New  York  14454;  E-mail;  WilsonBull@geneseo.edu. 

Secretary— John  A.  Smallwood,  Department  of  Biology,  Montclair  State  University,  Upper  Montclair, 
New  Jersey  07043;  E-mail:  Smallwood@satum.montclair.edu. 

Treasurer— Doris  J.  Watt,  Department  of  Biology,  Saint  Mary’s  College,  Notre  Dame,  Indiana  46556; 
E-mail:  DWatt@saintmarys.edu. 

Elected  Council  Members — Peter  C.  Frederick  and  Danny  J.  Ingold  (terms  expire  1999),  Charles  R.  Blem 
and  Sara  R.  Morris  (terms  expire  2000),  Jonathan  L.  Atwood  and  James  L.  Ingold  (terms  expire 

2001). 

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  Bulletin  is  sent  to  all  members  not  in  arrears  for  dues. 

THE  JOSSELYN  VAN  TYNE  MEMORIAL  LIBRARY 
The  Josselyn  Van  Tyne  Memorial  Library  of  the  Wilson  Ornithological  Society,  housed  m the  University 
of  Michigan  Museum  of  Zoology,  was  established  in  concurrence  with  the  University  of  Michigan  m 
1930.  Until  1947  the  Library  was  maintained  entirely  by  gifts  and  bequests  of  books,  reprints,  and  orni- 
thological 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  administered  by  the  Library  Committee.  William  A.  Lunk,  University  of 
Michigan  is  Chairman  of  the  Committee.  The  Library  currently  receives  over  200  periodicals  as  gifts  and 
in  exchange  for  The  Wilson  Bulletin.  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  University 
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 
University  of  Michigan,  1 109  Geddes  Ave.,  Ann  Arbor,  MI  48109-1079  USA.  Contributions  to  the  New 
Book  Fund  should  be  sent  to  the  Treasurer. 


THE  WILSON  BULLETIN 
(ISSN  0043-5643) 

THE  WILSON  BULLETIN  (ISSN  004.^-5643)  i.s  published  quarterly  in  March.  June,  September,  and  December  by 
the  Wilson  Ornithological  Society.  810  East  10th  Street.  Lawrence,  KS  66044-8897.  The  .subscription  price,  both  in  the 
United  States  and  elsewhere,  is  S40.00  per  year.  Periodicals  postage  paid  at  Lawrence.  KS.  POSTMASTER:  Send  address 
changes  to  THE  WILSON  BULLETIN.  P.O.  Box  1897,  Lawrence,  KS  66044-8897. 

Back  issues  or  single  copies  are  available  for  $12.00  each.  Most  back  issues  of  the  Bulletin  are  available  and  may  be 
ordered  from  the  Treasurer.  Special  prices  will  be  quoted  for  quantity  orders. 

All  articles  and  communications  for  publications,  books  and  publications  for  reviews  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  addre.ss  and  claims  for  undelivered  copies  should  be  .sent  to  the  OSNA,  P.O.  Box  1897. 
Lawrence.  KS  66044-8897.  Phone:  (785)  843-1221;  FAX;  (785)  843-1274. 

© Copyright  1999  by  the  Wilson  Ornithological  Society 

Printed  by  Allen  Press,  Inc.,  Lawrence,  Kansas  66044,  U.S.A. 


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


THE  WILSON  BULLETIN 

A QUARTERLY  JOURNAL  OF  ORNITHOLOGY 
Published  by  the  Wilson  Ornithological  Society 


VOL.  Ill,  NO.  1 MARCH  1999  PAGES  1-156 


Wilson  Bull.,  111(1),  1999,  pp.  1—6 


ANNUAL  SURVIVAL  RATES  OF  FEMALE  HOODED 
MERGANSERS  AND  WOOD  DUCKS  IN 
SOUTHEASTERN  MISSOURI 


KATIE  M.  DUGGER,'  23  BRUCE  D.  DUGGER,'  ^ AND  LEIGH  H.  FREDRICKSON' 


ABSTRACT. — Successful  conservation  and  management,  particularly  of  harvested  species,  relies  on  accurate 
estimates  of  population  demographics.  In  addition,  estimates  of  survival  and  longevity  allow  more  accurate 
modeling  of  evolutionary  life-history  trade-offs  within  and  between  species.  We  estimated  survival  rates  for  box 
nesting  female  Hooded  Mergansers  {Lophodytes  cucullatus)  and  Wood  Ducks  (Ai.x  sponsa)  in  southeastern 
Missouri  during  1987-1997  and  1987-1993,  respectively.  Hooded  Merganser  survival  rates  varied  annually  and 
ranged  from  0.42-1.0  (jc  = 0.66  ± 0.04).  Wood  Duck  survival  did  not  vary  significantly  over  time  and  averaged 
0.63  (±  0.02).  Mean  annual  survival  rates  and  capture  probabilities  were  similar  for  the  two  species  (x^  = 0.49, 
df  = \,  P > 0.05;  = 0.02,  df  = 1,  P > 0.05).  Annual  variation  in  Hooded  Merganser  survival  rates  was  an 

important  component  of  this  species’  population  ecology,  but  was  not  related  to  winter  weather  conditions, 
harvest  rates,  breeding  season  rainfall,  or  nesting  parameters.  Our  female  Wood  Duck  survival  rates  were  higher 
than  survival  estimates  for  other  adult  females  in  the  north-central  subpopulation,  but  were  comparable  to  some 
estimates  for  adult  females  that  breed  in  southern  and  mid- Atlantic  states.  Received  12  May  1998,  accepted  5 


Sept.  1998. 

Estimates  of  annual  survival  are  important 
for  comparative  studies  of  life-history  strate- 
gies between  species  (Krementz  et  al.  1989) 
and  for  modeling  population  demographics  for 
conservation  and  management.  Annual  sur- 
vival rate  estimates  are  available  for  certain 
geese  (tribe  Anserini)  and  dabbling  ducks 
(tribe  Anatini)  traditionally  important  to  hunt- 
ers (Johnson  et  al.  1992).  Except  for  the  Com- 
mon Eider  {Somateria  molUssima-,  Krementz 
et  al.  1996),  survival  estimates  based  on  mod- 
em survival  estimation  procedures  are  com- 
pletely lacking  for  most  seaducks  (tribe  Mer- 
gini),  despite  increased  harvest  pressure  in  re- 


‘ Gaylord  Memorial  Laboratory,  School  of  Natural 
Resources,  Univ.  of  Missouri,  Puxico,  MO  63690. 

^ Present  address:  Cooperative  Wildlife  Research 
Laboratory,  Southern  Illinois  Univ.,  Carbondale,  IL 
62901. 

^ Corresponding  author;  E-mail: 
tinamou  @ midwest.net 


cent  years  (U.S.  Fish  and  Wild.  Serv.,  unpubl. 
data). 

Hooded  Mergansers  {Lophodytes  cuculla- 
tus) are  among  the  least  studied  of  all  Mergini 
(Dugger  et  al.  1994).  Attempts  to  estimate  sur- 
vival rates  are  difficult  because  of  their  low 
harvest  rate,  secretive  behavior,  year-round 
occurrence  in  low  densities,  and  preference 
for  forested  wetlands.  However,  Hooded  Mer- 
gansers nest  in  man-made  boxes,  and  capture 
of  these  females  can  provide  mark-recapture 
samples  large  enough  to  estimate  survival 
rates  (Dugger  et  al.  1994). 

Over  much  of  their  range  Hooded  Mergan- 
sers occur  sympatrically  with  Wood  Ducks 
{Aix  sponsa,  tribe  Anatini;  Livezey  1986)  pro- 
viding an  opportunity  for  comparison  of  sur- 
vival estimates.  The  two  species  are  similar  in 
body  size  (HM,  540-725  g;  WD,  530—680  g; 
Bellrose  and  Holm  1994),  both  nest  in  tree 
cavities,  and  both  rely  on  forested  wetlands 


1 


2 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


during  the  breeding  season.  However,  these 
species  differ  considerably  in  many  aspects  of 
their  biology  and  might  be  expected  to  exhibit 
differences  in  annual  survival  rates.  Hooded 
Mergansers  are  carnivorous,  forage  by  diving, 
and  exhibit  delayed  maturation  (first  breed  at 
^2  years  old;  Dugger  et  al.  1994).  In  contrast. 
Wood  Ducks  are  omnivorous,  forage  near  the 
water’s  surface,  and  most  breed  as  yearlings 
(Bellrose  and  Holm  1994).  Based  on  the  dif- 
ference in  age  at  first  breeding  and  phylogeny 
(Krementz  et  al.  1997),  we  predict  that  Hood- 
ed Mergansers  experience  higher  annual  sur- 
vival rates  than  Wood  Ducks  (Ricklefs  1973, 
Wittenberger  1979).  In  this  paper  we  estimate 
annual  survival  rates  of  box  nesting  female 
Hooded  Mergansers  and  Wood  Ducks  in 
southeastern  Missouri  and  compare  mean  an- 
nual survival  rates  of  these  two  species. 

STUDY  AREA  AND  METHODS 

The  study  was  conducted  on  the  Duck  Creek  Wild- 
life Conservation  Area  (WCA)  and  Mingo  National 
Wildlife  Refuge  (NWR)  in  southeastern  Missouri. 
These  adjoining  areas  comprise  Mingo  Swamp,  the 
largest  contiguous  block  of  bottomland  hardwood  for- 
est in  Missouri  (1 1,174  ha).  For  a detailed  description 
of  the  habitat  types  available  in  Mingo  Swamp  see 
Heitmeyer  and  coworkers  (1989).  Approximately  85 
boxes  were  available  to  nesting  females  in  all  years  of 
the  study  on  Mingo  NWR,  and  100-120  nest  boxes 
were  available  for  nesting  birds  on  Duck  Creek  WCA 
during  1987-1993,  and  most  nest  boxes  were  equipped 
with  predator  guards  to  reduce  predation  by  raccoons 
(Procyort  lotor).  After  the  nesting  season  in  1993,  ap- 
proximately 50  boxes  were  removed  on  Duck  Creek 
WCA  and  the  Wood  Duck  nesting  study  was  termi- 
nated. Capture  of  Hooded  Mergansers  continued  on 
both  Mingo  NWR  and  Duck  Creek  WCA  through 
1997. 

Nest  boxes  were  cleaned,  repaired  when  necessary, 
and  filled  with  wood  chips  before  each  nesting  season. 
Hooded  Mergansers  and  Wood  Ducks  were  captured 
in  nest  boxes  between  1 February  and  15  August  each 
year  during  1987-1997  and  1987-93,  respectively.  We 
checked  boxes  at  2-4  week  intervals  and  captured 
nesting  females  of  both  species  during  the  third  week 
of  incubation.  Unmarked  females  were  banded  with 
U.S.  Fish  and  Wildlife  Service  leg  bands,  and  band 
numbers  of  previously  marked  birds  were  recorded. 
All  breeding  females  of  both  species  captured  in  nest 
boxes  were  u.sed  in  our  analysis,  including  Wood 
Ducks  that  bred  as  yearlings.  Although  both  Hooded 
Mergan.ser  and  Wood  Duck  ducklings  were  web- 
tagged  in  boxes  at  hatching,  .sample  sizes  of  known- 
age  birds  were  too  small  to  analyze  by  age  class. 

We  constructed  mark-recapture  matrices  for  Hooded 
Mcrgan.sers  ( I 1 years)  and  Wood  Ducks  (7  years)  and 


TABLE  1.  The  number  of  Hooded  Merganser 
nesting  attempts  and  nest  success  in  southeastern  Mis- 
souri 1987-1997. 


Year 

Nest  attempts* 

Nest  success 
rate*’  (%) 

1987 

17 

88.2 

1988 

32 

84.4 

1989 

37 

62.2 

1990 

35 

77.1 

1991 

37 

73.0 

1992 

59 

35.6 

1993 

48 

39.6 

1994 

64 

51.5 

1995 

38 

73.7 

1996 

43 

58.1 

1997 

50 

56.0 

“ Total  number  of  nests  initiated  by  Hooded  Mergansers  in  southeast  Mis- 
souri. 

^ Hooded  Merganser  nest  success  (number  of  successful  nests/total  num- 
ber of  nest  attempts). 


used  Jolly-Seber  mark-recapture  models  for  open  pop- 
ulations to  estimate  survival  for  both  species  (Pollock 
et  al.  1990).  Program  JOLLY  computes  point  esti- 
mates, their  associated  variances,  goodness-of-fit  tests, 
and  likelihood  ratio  tests  for  five  open  population  mod- 
els (Pollock  et  al.  1990).  We  used  model  goodness-of- 
fit  tests  and  likelihood  ratio  tests  between  models  to 
select  the  model  that  provided  the  best  fit  for  each  data 
set  (Pollock  et  al.  1990).  We  then  compared  Wood 
Duck  and  Hooded  Merganser  mean  annual  survival 
rates  and  capture  probabilities  using  the  program  Con- 
trast (Hines  and  Sauer  1989)  and  the  methods  de- 
scribed by  Sauer  and  Williams  (1989).  All  analyses 
were  performed  on  an  IBM  computer  under  DOS. 

Our  analysis  suggested  time  dependent  variation 
was  an  important  component  of  Hooded  Merganser 
survival,  so  we  attempted  to  identify  factors  that  might 
be  correlated  with  merganser  annual  survival  rates.  We 
correlated  Hooded  Merganser  survival  estimates  for 
each  year  with  annual  harvest  (USFWS,  unpubl.  data), 
winter  weather  conditions  (rainfall,  temperature)  in  the 
Mississippi  Alluvial  Valley,  and  rainfall  in  Mingo 
Swamp  from  March  through  September.  We  also  cor- 
related Hooded  Merganser  survival  with  the  number 
of  nesting  attempts  and  nest  success  on  Mingo  NWR 
and  Duck  Creek  WCA  (Table  1 ).  These  reproductive 
variables  might  be  expected  to  index  local  Hooded 
Merganser  nesting  density,  and  therefore  represent  the 
potential  for  permanent  emigration  from  our  study 
sites. 

Although  specific  wintering  areas  for  birds  in  our 
population  arc  largely  unknown,  the  Mississippi  Al- 
luvial Valley  is  the  closest  region  with  suitable  win- 
tering habitat,  and  females  nesting  in  Mingo  Swamp 
have  been  recovered  from  this  region  (Dugger  et  al. 
1994).  We  used  mean  daily  temperature  and  monthly 
rainfall  totals  recorded  at  two  sites  in  each  of  three 
states  (Arkansas,  Mississippi,  and  Louisiana)  October 


Du}i}>er  et  al.  • MERGANSER  AND  WOOD  DUCK  SURVIVAL 


3 


TABLE  2.  Annual  survival  estimates  and  associated  capture  probabilities  for  female  Hooded  Mergansers 
and  Wood  Ducks  nesting  in  southeastern  Missouri. 


Survival  probability  (SE) 

Capture  probability  (SE) 

Year 

Hooded  Merganser 

Wood  Duck 

Hooded  Merganser 

Wood  Duck 

1987 

0.58  (0.17) 

a 

1988 

0.86  (0.16) 

a 

0.78  (0.19) 

0.71  (0.07) 

1989 

1.00  (0.25) 

a 

0.43  (0.13) 

0.77  (0.05) 

1990 

0.42  (0. 1 1 ) 

a 

0.37  (0.12) 

0.55  (0.05) 

1991 

0.97  (0.15) 

a 

0.53  (0.12) 

0.48  (0.05) 

1992 

0.54  (0.12) 

a 

0.69  (0.13) 

0.51  (0.06) 

1993 

0.55  (0.12) 

a 

0.59  (0.13) 

0.66  (0.08) 

1994 

0.52  (0.13) 

0.71  (0.14) 

1995 

0.45  (0.12) 

0.65  (0.15) 

1996 

0.66  (0.16) 

Mean 

0.66  (0.04) 

0.63  (0.02) 

0.60  (0.05) 

0.61  (0.04) 

^ Recapture  data  from  these  years  were  available  to  estimate  constant  survival  rale  for  Wood  Ducks. 

through  January  for  each  year  (National  Climatic  Data 
Center)  to  index  winter  habitat  conditions.  We  used 
total  rainfall  during  March  through  September  collect- 
ed at  Advance,  Missouri  (National  Climatic  Data  Cen- 
ter) to  index  local  habitat  conditions  during  the  breed- 
ing and  post-breeding  season  when  Hooded  Mergan- 
sers are  present  in  Mingo  Swamp. 

RESULTS 

We  used  individual  capture  histories  of  151 
Hooded  Merganser  and  512  Wood  Duck  fe- 
males to  estimate  annual  survival.  Model  A, 
from  Program  JOLLY,  with  time-dependent 
capture  probabilities  and  survival  rates  pro- 
vided the  best  fit  for  the  Hooded  Merganser 
data  (x^  = 19.61,  df  = 12,  P > 0.05).  Model 
A also  fit  the  Wood  Duck  data  (x^  = 17.52, 
df  = 12,  P > 0.05)  as  did  Model  B,  a reduced 
parameter  model  with  constant  survival  rates 
and  time-dependant  capture  probability  (x^  = 
18.41,  df  = 16,  P > 0.05).  The  likelihood  ra- 
tio test  between  Models  A and  B (x^  = 0.90, 
df  = 4,  P > 0.05)  suggested  Model  B provid- 
ed the  most  parsimonious  fit  for  the  Wood 
Duck  data.  Jolly-Seber  models  estimate  sur- 
vival through  sample  k-2,  and  capture  proba- 
bility for  samples  2 through  k-l  (model  A)  or 
k (model  B).  Thus,  we  had  9 estimates  of  an- 
nual survival  and  capture  probability  for 
Hooded  Mergansers  (1987-1995)  and  6 esti- 
mates of  capture  probability  (1988-1992)  for 
Wood  Ducks  with  a single  estimate  of  con- 
stant survival  (Table  2).  Hooded  Merganser 
survival  rates  ranged  from  a low  of  0.42  in 
1990  to  a high  of  1.0  in  1989  with  a mean  of 
0.66  (±  0.04  SE)  with  95%  confidence  limits 


of  0.59-0.73  (Table  2).  Wood  Duck  annual 
survival  was  0.63  (2;  0.02)  with  95%  confi- 
dence limits  of  0.59—0.68  (Table  2).  The  mean 
annual  survival  of  Hooded  Mergansers  was 
not  significantly  different  than  Wood  Duck  an- 
nual survival  (x^  = 0.49,  df  = 1,  P > 0.05). 

Capture  probability  for  Hooded  Mergansers 
varied  annually  from  a low  of  0.37  in  1990  to 
a high  of  0.78  in  1988  (Table  2).  Wood  Duck 
capture  rates  also  exhibited  annual  variation, 
ranging  from  0.51  in  1992  to  0.77  in  1989 
(Table  2).  Mean  capture  probabilities  did  not 
differ  between  species  ~ 0.02,  df  = 1,  P 
> 0.05).  In  addition,  we  observed  no  signifi- 
cant correlation  between  Hooded  Merganser 
survival  rates  and  annual  harvest,  winter 
weather  conditions  (temperature  and  rainfall), 
breeding  season  rainfall,  or  Hooded  Mergan- 
ser nesting  parameters  (all  P > 0.05). 

DISCUSSION 

Hooded  Merganser  and  Wood  Duck  annual 
survival  estimates  from  our  study  were  gen- 
erally higher  than  those  reported  for  other 
duck  species  (Johnson  et  al.  1992).  However, 
because  our  estimates  were  for  birds  using 
nest  boxes  with  some  protection  from  preda- 
tory raccoons,  comparisons  with  other  species 
must  be  made  with  caution.  Nevertheless, 
Hooded  Mergansers  in  our  study  had  substan- 
tially lower  mean  annual  survival  rates  than 
the  Common  Eider  (x  = 0.87;  Krementz  et  al. 
1996),  the  only  other  member  of  Mergini  for 
which  estimates  are  available.  Yearly  survival 
rates  for  Hooded  Mergansers  during  1988, 


4 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  1,  March  1999 


1989,  and  1991  were  comparable  with  those 
for  Common  Eiders  (Krementz  et  al.  1996); 
rates  for  other  years  were  substantially  lower. 
Wood  Duck  survival  in  our  study  was  higher 
than  or  comparable  to  other  estimates  avail- 
able for  female  Wood  Ducks  (Johnson  et  al. 
1986,  Nichols  and  Johnson  1990,  Kelley 
1997),  including  estimates  from  a South  Car- 
olina box-nesting  population  (T  = 0.55,  Hepp 
et  al.  1987).  Wood  Duck  females  are  strongly 
philopatric  to  nest  sites  (nearly  100%),  so 
mark-recapture  survival  estimates  that  include 
a measure  of  capture  site  fidelity  (nest  boxes 
in  this  case)  can  be  comparable  to  band  re- 
covery estimates  (Hepp  et  al.  1987).  Whether 
our  high  survival  rates  reflect  general  regional 
differences  in  survival  of  eastern  Wood  Ducks 
(Nichols  and  Johnson  1990)  or  a survival  ben- 
efit associated  with  box-nesting  remains  un- 
clear. Wood  Ducks  breeding  in  Missouri  are 
included  in  the  “north-central”  sub-popula- 
tion of  Bowers  and  Martin  (1975)  and  Kelley 
(1997),  but  exhibit  survival  rates  much  higher 
than  band  recovery  estimates  for  adult  females 
in  this  region  (Kelley  1997).  Our  Wood  Duck 
females  exhibited  survival  rates  most  similar 
to  adult  females  in  the  southern  population  of 
the  Mississippi  Fly  way  {x  = 0.61;  Kelley 
1997)  and  the  mid- Atlantic  population  of  the 
Atlantic  Flyway  (T  = 0.63;  Kelley  1997). 

Mean  survival  did  not  differ  between 
Hooded  Mergansers  and  Wood  Ducks,  al- 
though in  three  of  nine  years  Hooded  Mer- 
ganser survival  was  higher  than  the  constant 
rate  estimated  for  Wood  Ducks  (Table  2).  This 
is  inconsistent  with  life-history  theory  which 
predicts  that  birds  with  delayed  maturation 
should  experience  higher  annual  survival 
(Wittenberger  1979),  but  consistent  with  anal- 
yses showing  survival  rates  are  correlated 
with  body  size  and  breeding  latitude  in  wa- 
terfowl (Arnold  1988).  Maybe  more  important 
than  the  comparison  of  mean  survival  rates 
was  our  observation  that  annual  variation  was 
an  important  component  of  Hooded  Mergan- 
ser, but  not  Wood  Duck,  survival.  Differences 
in  diet,  foraging  method,  and  habitat  require- 
ments may  make  Hooded  Mergansers  more 
.sensitive  to  local  fluctuations  in  food  resourc- 
es or  water  conditions  during  reproduction  or 
winter,  with  increa.sed  mortality  or  emigration 
during  years  when  habitat  conditions  are  poor. 
None  of  the  harvest,  breeding  sea.son  rainfall. 


density-dependent  factors,  or  winter  weather 
variables  we  investigated  were  significantly 
related  to  variation  in  Hooded  Merganser  sur- 
vival. However,  we  did  not  quantify  food 
availability  on  the  breeding  grounds  directly 
and  because  information  on  migration  patterns 
and  winter  site  fidelity  for  Hooded  Mergan- 
sers is  lacking  we  may  not  have  compared  our 
Hooded  Merganser  survival  rates  with  the 
most  appropriate  winter  or  breeding  season 
weather  conditions.  Very  little  information  ex- 
ists concerning  Hooded  Merganser  foraging 
ecology  or  foraging  habitat  characteristics 
(Dugger  et  al.  1994),  consequently,  determin- 
ing the  climatic  factors  that  index  habitat  con- 
ditions throughout  the  Hooded  Merganser  an- 
nual cycle  will  be  difficult. 

We  believe  our  survival  estimates  for  both 
species  are  unbiased,  but  some  behavioral  re- 
sponses such  as  permanent  emigration  from 
the  study  area  cannot  be  distinguished  from 
“deaths”  by  Jolly-Seber  models  and  can  re- 
sult in  negatively  biased  survival  estimates 
(Pollock  et  al.  1990).  Wood  Ducks  are  strong- 
ly philopatric  (Hepp  et  al.  1987),  but  data  are 
lacking  to  estimate  philopatry  for  Hooded 
Mergansers  or  to  make  direct  quantitative 
comparisons  with  Wood  Ducks.  We  believe 
that  Hooded  Mergansers  are  strongly  philo- 
patric to  general  nesting  areas  (e.g.,  Zicus 
1990),  but  not  as  philopatric  as  Wood  Ducks 
to  specific  nesting  boxes.  Factors  that  might 
have  caused  Hooded  Merganser  females  to 
have  left  the  study  site  or  chosen  not  to  nest 
in  boxes  in  subsequent  years  (low  nest  success 
or  high  breeding  density)  were  not  correlated 
with  annual  survival  as  we  might  expect  if 
permanent  emigration  were  common.  Further- 
more, in  Minnesota,  distances  moved  by 
Hooded  Mergansers  between  nesting  sites 
each  year  were  not  related  to  nest  success  (Zi- 
cus 1990),  suggesting  that  variation  in  nest 
success  does  not  affect  philopatry.  Habitat 
conditions  in  Mingo  Swamp  could  have  af- 
fected Hooded  Merganser  use  of  boxes,  but  it 
is  unlikely  that  these  effects  would  be  per- 
manent. Finally,  we  do  not  believe  that  nest 
boxes  were  limiting  for  Hooded  Mergansers 
or  that  competition  for  nest  sites  led  to  higher 
permanent  emigration  by  Hooded  Mergansers 
from  Mingo  Swamp.  Hooded  Merganser  pop- 
ulations were  substantially  lower  than  Wood 
Duck  populations  each  year  (Dugger  1991) 


Dui>}>er  el  al.  • MERGANSER  AND  WOOD  DUCK  SURVIVAL 


5 


and  the  annual  number  of  Hooded  Merganser 
nest  attempts  remained  low  in  relation  to  box 
availability.  In  addition.  Hooded  Mergansers 
initiate  nests  on  average  3-4  weeks  earlier 
than  Wood  Ducks,  thereby  reducing  the  po- 
tential for  nest  site  competition  between  the 
two  species  (Dugger  et  al.  1994). 

Overall  mean  Hooded  Merganser  survival 
in  this  study  was  lower  than  a previous  esti- 
mate (Dugger  et  al.  1994),  and  this  earlier 
analysis  did  not  detect  any  significant  annual 
variation  in  survival  rates.  As  more  data  were 
collected  and  added  to  the  analysis,  estimate 
precision  increased,  mean  annual  survival  de- 
creased, and  annual  variation  became  an  im- 
portant component  of  Hooded  Merganser  sur- 
vival. The  increased  variation  in  annual  sur- 
vival estimates  and  changes  in  the  long-term 
mean  associated  with  additional  years  of  study 
on  Hooded  Mergansers  illustrate  the  impor- 
tance of  long-term  data  sets.  Continued  band- 
ing effort  and  more  information  concerning 
Hooded  Merganser  habitat  use,  foraging  ecol- 
ogy, and  age-specific  survival  rates  are  needed 
to  understand  the  factors  affecting  annual  var- 
iation in  survival  of  this  species. 

ACKNOWLEDGMENTS 

We  thank  L.  Bollman,  R Blum,  J.  H.  Gammonley, 
J.  R.  Kelley,  M.  Shannon,  J.  Ware,  and  J.  S.  Wortham 
for  assistance  with  data  collection  and  the  staffs  at 
Duck  Creek  WCA  and  Mingo  NWR  for  access  to  their 
nest  boxes.  Comments  provided  by  P.  Blums,  J.  Gam- 
monley, R.  Kennamer  and  an  anonymous  reviewer 
greatly  improved  earlier  drafts  of  this  manuscript.  This 
research  was  funded  by  Gaylord  Memorial  Laboratory, 
Patuxent  Wildlife  Research  Center,  and  the  Missouri 
Cooperative  Wildlife  Research  Unit  at  the  University 
of  Missouri,  Columbia.  This  is  a contribution  from  the 
Missouri  Agricultural  Experiment  Station,  Journal  Se- 
ries Number  12,  771. 

LITERATURE  CITED 

Arnold,  T A.  1988.  Life  histories  of  North  American 
game  birds;  a reanalysis.  Can.  J.  Zool.  66:1906- 
1912. 

Bellrose,  F.  C.  and  D.  j.  Holm.  1994.  Ecology  and 
management  of  the  Wood  Duck.  Stackpole  Books, 
Harrisburg,  Pennsylvania. 

Bowers,  E.  E and  E W.  Martin.  1975.  Managing 
Wood  Ducks  by  population  units.  Trans.  N.  Am. 
Wildl.  Nat.  Resour.  Conf.  40:300-324. 

Dugger.  B.  D.,  K.  M.  Dugger,  and  L.  H.  Fredrick- 
son. 1994.  Hooded  Merganser  (Lophodytes  cu- 
cullatus).  In  The  birds  of  North  America,  no.  98 
(A.  Poole  and  F Gill,  Eds.).  The  Academy  of  Nat- 


ural Sciences,  Philadelphia,  Pennsylvania;  The 
American  Ornithologists’  Union,  Washington, 
DC. 

Dugger,  K.  M.  1991.  Nesting  parameters  and  popu- 
lation dynamics  of  box-nesting  female  wood 
ducks  in  Mingo  Swamp.  M.S.  thesis.  Univ.  of 
Missouri,  Columbia. 

Heitmeyer,  M.  E.,  L.  H.  Fredrickson,  and  G.  F. 
Krause.  1989.  Water  and  habitat  dynamics  of  the 
Mingo  Swamp  in  southeastern  Missouri.  Fish 
Wildl.  Res.  6:1-26. 

Hepp,  G.  R.,  R.  T.  Hoppe,  and  R.  A.  Kennamer.  1987. 
Population  parameters  and  philopatry  of  breeding 
female  Wood  Ducks.  J.  Wildl.  Manage.  51:401  — 
404. 

Hines,  J.  E.  and  J.  R.  Sauer.  1989.  Program  CON- 
TRAST— a general  program  for  the  analysis  of 
several  survival  or  recovery  rate  estimates.  Fish 
Wildl.  Tech.  Report  24:1-7. 

Johnson,  D.  H.,  J.  D.  Nichols,  and  M.  D.  Schwartz. 
1992.  Population  dynamics  of  breeding  waterfowl. 
Pp.  446-485  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.  Kadlec,  and  G.  L.  Krapu,  Eds.).  Univ.  of 
Minnesota  Press,  Minneapolis. 

Johnson,  F.  A.,  J.  E.  Hines,  F.  Montalbano,  III,  and 
J.  D.  Nichols.  1986.  Effects  of  liberalized  harvest 
regulations  on  Wood  Ducks  in  the  Atlantic  Fly- 
way. Wildl.  Soc.  Bull.  14:383-388. 

Kelley,  J.  R.,  Jr.  1997.  Wood  duck  population  mon- 
itoring initiative;  final  report.  Atlantic  Flyway 
Council,  Mississippi  Flyway  Council,  and  U.S. 
Fish  and  Wildlife  Service,  Laurel,  Maryland. 

Krementz,  D.  G.,  R.  j.  Barker,  and  J.  D.  Nichols. 
1997.  Sources  of  variation  in  waterfowl  survival 
rates.  Auk  114:93-102. 

Krementz,  D.  G.,  J.  E.  Hines,  and  D.  F Caithamer. 
1996.  Survival  and  recovery  rates  of  American 
Eiders  in  eastern  North  America.  J.  Wildl.  Man- 
age. 60:855-862. 

Krementz,  D.  G.,  J.  R.  Sauer,  and  J.  D.  Nichols. 
1989.  Model-based  estimates  of  annual  survival 
rate  are  preferable  to  observed  maximum  lifespan 
statistics  for  use  in  comparative  life-history  stud- 
ies. Oikos  56:203—208. 

Livezey,  B.  C.  1986.  A phylogenetic  analysis  of  recent 
anseriform  genera  using  morphological  characters. 
Auk  103:737-754. 

Nichols,  J.  D.  and  F A.  Johnson.  1990.  Wood  Duck 
population  dynamics:  a review.  Pp.  83-105  in 
Proc.  1988  North  Am.  Wood  Duck  Symp.  (L.  H. 
Fredrickson,  G.  V.  Burger,  S.  P.  Havera,  D.  A. 
Grabcr,  R.  E.  Kirby,  and  T.  S.  Taylor.  Eds.).  St. 
Louis,  Missouri. 

Pollock,  K.  H.,  J.  D.  Nichols,  C.  Brownie,  and  J.  E. 
Hines.  1990.  Statistical  inference  of  capture-re- 
capture  experiments.  Wildl.  Monogr.  107;  1—97. 

Ricklefs,  R.  E.  1973.  Fecundity,  mortality  and  avian 


6 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


demography.  Pp.  366—435  in  Breeding  biology  of 
birds  (D.  S.  Earner,  Ed.).  National  Academy  of 
Sciences,  Washington,  D.C. 

Sauer,  J.  R.  and  B.  K.  Williams.  1989.  Generalized 
procedures  for  testing  hypotheses  about  survival 
or  recovery  rates.  J.  Wildl.  Manage.  53:137-142. 


WiTTENBERGER,  J.  E 1979.  A model  for  delayed  repro- 
duction in  iteroparous  animals.  Am.  Nat.  114: 
439-446. 

Zicus,  M.  C.  1990.  Nesting  biology  of  Hooded  Mer- 
gansers using  nest  boxes.  J.  Wildl.  Manage.  54: 
637-643. 


Wilson  Bull:  111(1),  1999,  pp.  7-14 


COMPARATIVE  NEST  SITE  HABITATS  IN  SHARP-SHINNED  AND 

COOPER’S  HAWKS  IN  WISCONSIN 

DALE  R.  TREXEL,'  ROBERT  N.  ROSENFIELD,^  ^ JOHN  BIELEFELDT\  AND 

EUGENE  A.  JACOBS^ 


ABSTRACT. — From  an  analysis  of  nest  site  habitat  data  at  24  Sharp-shinned  Hawk  (Accipiter  striatus)  and 
52  Cooper’s  Hawk  (A.  cooperii)  nests  in  Wisconsin,  we  conclude  that  Cooper’s  Hawks  tend  to  nest  in  stands 
with  lower  densities  of  taller  and  larger  trees  than  do  Sharp-shinned  Hawks,  and  that  Cooper  s Hawks  also  tend 
to  nest  in  sites  with  a greater  proportion  of  hardwood  cover  than  Sharp-shinned  Hawks.  Significant  interspecific 
differences  were  found  in  combined  habitat  types  (hardwoods,  mixed  conifer-hardwoods,  and  conifer  plantations) 
for  nest  tree  height  and  nest  tree  DBH  (diameter  at  breast  height);  nest  height;  nest  height  relative  to  tree  height; 
canopy  height;  canopy  cover;  tall  shrub  density;  tree  density;  and  mean  DBH.  Nest  sites  of  the  two  species  were 
similar  in  terms  of  understory  canopy  cover,  ground  cover,  low  shrub  index,  understory  tree  density,  basal  area, 
distance  to  nearest  forest  opening,  and  distance  to  water.  We  detected  few  significant  intraspecific  differences  in 
nest  site  habitat,  and  these  only  in  the  Cooper’s  Hawk.  Received  23  Oct.  1997,  accepted  4 Nov.  1998. 


Although  the  Sharp-shinned  Hawk  {Accip- 
iter striatus)  and  the  Cooper’s  Hawk  (A.  coop- 
erii) breed  sympatrically  in  many  parts  of  the 
United  States  and  southern  Canada,  their  nest 
site  habitats  have  been  compared  in  only  four 
published  quantitative  studies.  These  conge- 
ners are  sometimes  assumed  to  partition  nest- 
ing habitat  by  way  of  interspecific  competition 
and/or  predation  (Siders  and  Kennedy  1996). 
With  one  exception  in  Missouri  (Wiggers  and 
Kritz  1991),  these  studies  were  conducted  in 
the  western  United  States  (Oregon:  Reynolds 
et  al.  1982,  Moore  and  Henny  1983;  New 
Mexico:  Siders  and  Kennedy  1996). 

Such  geographically  restricted  results  may 
be  difficult  to  extrapolate  to  other  areas  of 
sympatry  because  of  regional  differences  in 
vegetational  composition  and  structure.  Each 
of  the  previously  published  comparisons  of 
these  hawks’  nest  site  habitats  was  derived 
from  upland  forests  with  relatively  homoge- 
neous vegetation,  principally  montane  conifer 
forests  in  New  Mexico  and  Oregon,  and  co- 
nifer plantations  or  oak-hickory  forests  in 
Missouri.  Our  study  area  (Fig.  1)  was  the  state 


‘ Dept,  of  Ecology,  Evolution,  and  Behavior,  100 
Ecology  Building,  1987  Upper  Buford  Circle,  Univ.  of 
Minnesota,  St.  Paul,  MN  55108. 

^ Dept,  of  Biology,  Univ.  of  Wisconsin,  Stevens 
Point,  WI  5448 1 . 

^ Park  Planning,  Racine  County  Public  Works  Di- 
vision, Sturtevant,  WI  53177. 

■*  Linwood  Springs  Research  Station,  1601  Brown 
Deer  Lane,  Stevens  Point,  WI  54481. 

’Corresponding  author;  E-mail:  rrosenfi@uwsp.edu 


of  Wisconsin  (145,000  km^).  The  ecologically 
diverse  set  of  available  woodland  nesting  hab- 
itats on  this  statewide  scale  includes  boreal 
conifer  forests  (plus  conifer  swamps  of  boreal 
affinity  over  much  of  the  state),  conifer  plan- 
tations, mixed  conifer-hardwood  forests,  pure- 
ly deciduous  woodlands  on  upland  and  low- 
land sites,  and  highly  fragmented  or  urban 
woodlands  (Rosenfield  et  al.  1996)  as  well  as 
extensive  forests.  For  further  details  on  Wis- 
consin forests  see  Curtis  (1959). 

Potentially  conflicting  results  among  past 
studies  may  also  limit  their  utility  in  unstudied 
areas  of  sympatry.  In  New  Mexico,  for  ex- 
ample, Siders  and  Kennedy  (1996)  found  sig- 
nificant differences  between  Sharp-shinned 
Hawks  and  Cooper’s  Hawks  in  the  majority  of 
nest  site  variables  tested,  while  in  Oregon, 
both  Reynolds  and  coworkers  (1982)  and 
Moore  and  Henny  (1983)  found  few  discern- 
ible differences  in  nest  site  characterisitics  be- 
tween these  accipiters.  Furthermore,  Siders 
and  Kennedy  (1996)  have  suggested  that  in- 
terpretations of  previous  results  may  be  ham- 
pered by  small  sample  sizes,  especially  for 
Sharp-shinned  Hawks  (n  < 18  nests  in  prior 
studies),  and  by  possible  biases  in  nest  search 
methods  or  methods  of  selecting  search  areas. 

We  compare  habitat  at  24  Sharp-shinned 
Hawk  nests  and  52  Cooper’s  Hawk  nests  in 
Wisconsin,  1980-1994,  all  discovered  by  un- 
biased means.  Previous  comparative  work  on 
nest  site  habitats  of  these  two  hawks  has  em- 
phasized interspecific  differences  within  rela- 
tively uniform  habitat  types.  We  expand  this 


7 


8 


THE  WILSON  BULLETIN  • VoL  111,  No.  I.  March  1999 


FIG.  1.  Distribution  by  county  of  nest  sites  sampled  for  Accipiter  cooperii  (circled)  and  A.  striatus  (not 
circled)  in  Wisconsin. 


emphasis  to  include  intraspecific  similarities 
as  well  as  interspecific  differences  across  hab- 
itat types  (i.e.,  combined  habitats)  at  a land- 
scape scale.  Intraspecific  nest  site  features 
held  in  common  across  habitat  types  may  aid 
land  management  agencies  in  assessing  and 
conserving  a range  of  usable  breeding  habitats 
for  Sharp-shinned  and  Cooper’s  hawks.  Our 
results  seem  timely  and  pertinent  to  the  recent 
Birds  in  Forested  Landscape  project  for  North 
America  (Cornell  Lab  of  Ornithology),  which 
focuses  in  part  on  the  nesting  habitats  of  these 
two  hawks,  and  is  designed  to  develop  man- 
agement and  conservation  strategies  on  their 
behalf  (Anonymous  1997). 

METHODS 

Ne.st  locations. — Nest  site  locations  were  considered 
unbia.sed  if  they  were  discovered  by  one  of  two  meth- 
ods: ( 1 ) incidental  or  random  locations  obtained  by 
cooperators  during  any  activity  other  than  .searching 
for  accipiter  nests,  and  (2)  locations  resulting  from 
Cooper's  Hawk  density  studies  in  which  objectively 


drawn  study  areas  were  completely  searched  regardless 
of  their  perceived  suitability  for  nesting  and  without 
foreknowledge  of  current  or  historical  nest  sites  on 
these  areas.  By  these  methods,  we  located  Cooper’s 
Hawk  nests  in  Wisconsin  (see  Fig.  1)  on  52  widely 
separated,  independent  nesting  areas,  as  defined  in  Ro- 
senfield  and  Bielefeldt  (1992,  1996).  All  24  Sharp- 
shinned  Hawk  nests  occurred  in  independent  nesting 
areas;  therefore  each  was  included  in  our  analyses. 

Data  collection  and  analyses. — Habitat  measure- 
ments (Table  1 ) were  made  postfiedging  at  each  nest 
site  within  a 0.04  ha  circular  plot  centered  on  the  nest 
tree  following  the  technique  of  James  and  Shugart 
(1970)  as  modified  by  Titus  and  Mosher  (1981). 

All  variables  were  tested  for  normality  with  Lillie- 
fors  test,  further  statistical  analyses  were  performed  on 
SYSTAT  (Wilkinson  1992).  /-tests  were  used  exclu- 
sively to  examine  interspecific  differences  among  sev- 
en habitat  varables  that  exhibited  normal  distributions 
in  combined  habitats  (i.e.,  tree  height,  tree  DBH,  nest 
height,  nest  percent,  canopy  height,  total  canopy,  and 
mean  DBH;  Table  2).  We  used  the  Mann-Whitney  La- 
test for  all  other  inferential  comparisons  because  all 
other  variables  were  not  normally  distributed. 

To  examine  inter-  and  intraspecific  differences  and 


Trexel  et  al.  • ACCIPITER  NEST  SITE  HABITATS 


9 


TABLE  1.  Habitat  variables  and  measurement  tecliniques  at  Accipiter  strialus  and  A.  coopcrii  nest  sites  in 
Wisconsin. 


Variable 


Description 


Tree  height 
Tree  DBH 
Nest  height 
Nest  percent 
Canopy  height 
Total  canopy 
Deciduous  can. 
Coniferous  can. 
Understory  can. 
Ground  cover 
Shrub  density 
Shrub  index 
Tree  density 
Under,  dens. 
Basal  area 
Mean  DBH 
Dist.  to  water 
Dist.  to  open. 


Height  (m)  of  nest  tree  (Haga  altimeter) 

Diameter  (cm)  at  breast  height  of  nest  tree 
Height  (m)  of  nest  (meter  tape  or  Haga  altimeter) 

(Nest  height/Tree  height)  X 100 

Mean  height  (m)  of  five  canopy  trees  in  study  plot  (Haga  altimeter) 

Percent  of  area  over  study  plot  occluded  by  overstory  foliage'’ 

Percent  of  area  over  plot  (not  of  total  canopy)  occluded  by  deciduous  overstory  foliage'' 
Percent  of  area  over  plot  occluded  by  evergreen  overstory  foliage" 

Percent  of  area  over  plot  occluded  by  understory  foliage" 

Percent  of  ground  in  plot  covered  by  ground-layer  foliage" 

Index  of  tall  shrubs  < 3 cm  DBH  and  ^ shoulder  height* ** 

Index  of  low  shrubs  < 3 cm  DBH  between  knee  and  shoulder  height** 

Number  of  canopy  trees  ^ 9 cm  DBH  per  hectare 
Number  of  understory  trees  ^ 9 cm  DBH  per  hectare 
mV  ha  of  canopy  trees 

Mean  DBH  (cm)  of  canopy  trees  in  study  plot 

Distance  (m)  to  nearest  permanent  water  source  (pacing  or  USGS  7.5  min.  quadrangles) 
Distance  (m)  to  nearest  forest  opening  > 5 ha  (pacing  or  USGS  7.5  min.  quadrangles) 


“ 40  ocular  tube  readings. 

**  Sum  of  four  plot  radii. 


similarities  among  habitat  types,  we  separated  our  nest 
site  samples  into  three  categories  based  on  trees  pre- 
sent within  the  0.04  ha  plot.  We  first  divided  nest  site 
samples  between  those  occurring  within  conifer  plan- 
tations and  those  not  in  plantation  habitats.  (While  de- 
ciduous trees  occurred  in  some  conifer  plantations,  no 
nest  sites  occurred  in  hardwood  plantations.)  We  then 
divided  non-plantation  nest  sites  into  those  situated  in 
pure  hardwood  stands  (where  no  trees  within  the  study 
plot  were  conifers)  and  those  in  mixed  conifer-hard- 
wood stands.  In  keeping  with  our  statewide  sample, 
these  three  habitat  categories  should  be  construed  as 
physiognomic  types  that  do  not  necessarily  exhibit  oth- 
er internal  similarities  in  vegetational  attributes.  In 
hardwood  stands,  for  example,  dominant  or  prominent 
tree  species  might  include  oaks  (Quercus  spp.),  maples 
(Acer  spp.),  aspen  (Populus  spp.),  and  other  species  of 
varied  ages,  management  histories,  and  moisture  re- 
gimes. Mixed  woodlands  might  include  lowland  co- 
nifers such  as  tamarack  (Lxirix  larcina)  and  black 
spruce  (Picea  rnariana)  or  upland  conifers  such  as 
pines  (Pinus  spp.)  as  well  as  deciduous  species.  For 
compositional  variety  among  nest  tree  species  (and  sci- 
entific names)  see  Table  3. 

Because  there  was  only  one  pure  hardwood  site  used 
by  Sharp-shinned  Hawks,  we  examined  interspecific 
differences  only  within  conifer  plantations  and  mixed 
conifer-hardwood  habitats.  Likewise,  we  could  only 
test  for  intraspecific  differences  between  mixed  and 
plantation  habitats  among  Sharp-shinned  Hawk  nest 
sites.  We  used  the  nonparametric  Kniskall-Wallis  test 
to  examine  interspecific  differences  among  the  three 
habitat  types  used  by  Cooper’s  Hawks.  Because  of  the 
number  of  multiple  univariate  comparisons  (Table  2), 


we  calculated  that  an  alpha  of  0.001  was  the  appro- 
priate level  of  significance  for  both  inter-  and  intraspe- 
cific inferences  (Sokal  and  Rohlf  1981). 

RESULTS 

The  majority  of  the  18  nest  site  variables 
compared  for  Sharp-shinned  and  Cooper’s 
hawks  in  combined  habitats  showed  statisti- 
cally significant  interspecific  differences  (Ta- 
ble 2).  Nest  tree  DBH  in  conifer  plantations, 
and  nest  height  and  canopy  height  in  mixed 
conifer-hardwood  stands  were  significantly 
different  between  species  across  uncombined 
habitat  types  (Table  2). 

Of  the  18  variables  examined  only  four  ex- 
hibited significant  intraspecific  differences 
across  habitat  types,  and  only  in  Cooper’s 
Hawk  (Table  2).  Although  intraspecific  nest 
site  selection  itself  might  vary  among  habitats, 
we  speculate  that  these  statistical  differences 
instead  are  attributable  to  inherent  vegetation- 
al contrasts  among  habitat  types  as  circum- 
scribed here.  In  the  most  transparent  example, 
cross-habitat  intraspecific  differences  in  per- 
centages of  coniferous  and  deciduous  canopy 
covers  at  Cooper’s  Hawk  nest  sites  in  hard- 
wood stands  versus  pine  plantations  (Table  2) 
are  a predictable  outcome  of  our  habitat  cat- 
egories. The  more  interesting  result  of  intra- 


10 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I.  March  1999 


TABLE  2.  Mean  values 
Wisconsin. 

of  habitat  variables  measured 

at  Accipiter  striatus  and  A. 

cooperii  nest  sites  in 

Variable 

A.  striatus  (mean  ± SE) 

Combined 

Mixed 

Plantation 

(n  = 24)“ 

(n  = 11) 

(»  = 12) 

Tree  height  (m) 

15.1  ± 0.6 

14.8  ± 0.6 

15.8  ± 1.0 

Tree  DBH  (cm) 

23.8  ± 1.4 

26.2  ± 2.3 

21.4  ± 1.6 

Nest  height  (m) 

9.1  ± 0.6‘> 

7.9  ± 0.7” 

10.4  ± 0.8 

Nest  percent  (%) 

59.7  ± 2.7'’ 

54.1  ± 4.3” 

65.5  ± 2.7 

Canopy  height  (m) 

15.2  ± 0.6 

15.0  ± 0.6 

15.8  ± 0.9 

Total  canopy  (%) 

76.5  ± 2.3 

74.3  ± 2.5 

80.4  ± 3.3 

Deciduous  can.  (%) 

15.2  ± 4.4 

24.3  ± 7.7 

3.8  ± 1.7 

Coniferous  can.  (%) 

61.3  ± 5.6 

50.0  ± 8.1 

76.7  ± 3.6 

Understory  can.  (%) 

32.7  ± 5.8 

42.7  ± 8.7 

19.8  ± 6.2 

Ground  cover  (%) 

39.0  ± 6.2 

53.4  ± 8.0 

22.9  ± 7.8 

Shrub  density 

61.6  ± 9.3 

73.4  ± 17.5 

50.3  ± 9.2 

Shrub  index 

90.3  ± 13.7 

79.0  ± 18.4 

98.1  ± 21.9 

Tree  density  (trees/ha) 

1071  ± 95'= 

914  ± 96 

1037  ± 13U 

Under,  dens,  (trees/ha) 

334  ± 71” 

375  ± 120” 

231  ± 57 

Basal  area  (m’/ha) 

28.9  ± 3.0 

22.7  ± 2.7 

37.5  ± 3.9^' 

Mean  DBH  (cm) 

17.6  ± 0.9^' 

16.9  ± 1.2 

19.1  ± 1.3'= 

Dist.  to  water  (m) 

260  ± 71 

88  ± 28 

440  ± 120 

Dist.  to  open,  (m) 

58.9  ± 16.2 

72.7  ± 26.2 

50.8  ± 21.8 

* P s 0.()0I. 

**  P s 0.00()5. 

“ Combined  data  for  A.  striaius  includes  one  hardwood  nest  site  in  addition  to  mixed  and  plantation  nest  sites. 
Missing  data  at  one  nest  site  (n  = 23  combined,  n = 10  mixed). 

Missing  data  at  one  nest  site  (n  = 23  combined,  ;i  = 1 1 plantation). 


specific  analyses  may  lie  in  the  variables  that 
did  not  differ  significantly  across  habitats, 
such  as  nest  tree  height,  nest  height,  canopy 
height,  and  mean  tree  DBH — each  of  which 
differed  between  species  (see  Discussion). 

For  combined  habitats.  Cooper’s  Hawks 
nested  in  a wider  array  of  tree  species  than 
Sharp-shinned  Hawks  (Table  3).  This  varia- 
tion however,  occurred  mostly  within  hard- 
wood sites;  within  mixed  stands  and  conifer 
plantations  Sharp-shinned  Hawks  used  a 
greater  variety  of  tree  species.  Of  the  conif- 
erous nest  trees  used  by  Cooper’s  Hawks  {n 
= 29),  only  Pinus  was  represented  in  this 
sample,  while  Sharp-shinnned  Hawks  {n  = 
23)  used  five  genera.  With  only  one  exception 
[a  Cooper’s  Hawk  nest  in  a white  ash  (Frax- 
inus  americana)],  both  species  consistently 
used  conifers  for  nesting  in  mixed  sites  where 
both  hardwoods  and  conifers  were  present  in 
the  canopy.  For  both  species,  nest  trees  in  co- 
nifer plantations  were  all  conifers,  despite  the 
presence  of  canopy-level  hardwoods  in  60% 
of  Cooper’s  Hawk  and  42%  of  Sharp-shinned 
Hawk  plantation  sites. 


DISCUSSION 

Our  comparative  analyses  of  nest  site  hab- 
itat at  52  Cooper’s  Hawk  and  24  Sharp- 
shinned  Hawk  nests  in  Wisconsin  did  not  pro- 
vide data  on  nest  site  use  relative  to  avail- 
ability, and  we  cannot  contend  that  numbers 
of  nests  in  our  three  habitat  categories  are  nec- 
essarily proportional  to  use  of  these  habitat 
types.  Nevertheless,  our  sample  involves  in- 
dependent nests  discovered  by  unbiased 
means  on  a statewide  scale  in  compositionally 
diverse  woodland  habitats:  upland  and  low- 
land sites;  coniferous,  hardwood,  and  mixed 
forests;  urban  and  rural  woodlands  of  varied 
sizes;  and  both  managed  and  unmanaged  for- 
ests including  conifer  plantations.  Thus  we 
suggest  that  our  data  set  provides  a reasonably 
thorough  and  representative  sample  of  the 
range  of  nest  site  habitats  used  by  these  hawks 
in  Wisconsin. 

If  interspecific  differences  in  nest  site  char- 
acteristics of  these  congeners  occur  on  a finer 
within-habitat  scale,  as  some  prior  work  has 
indicated  (Siders  and  Kennedy  1996),  then 


I'rexel  el  al.  • ACCIPITER  NEST  SITE  HABITATS 


11 


TABLE  2.  Extended. 

A.  cooperii  (mean  ± SE) 

Interspecific  differences 

Inira.specific 

differences 

Combined 

Hardwood 

Mi.xed 

Plantation 

Combined  Mixed  Plantation 

A.  siriatus  A.  cooperii 

II 

CA 

to 

(/7  = 22) 

(n  = 10) 

in  = 20) 

19.1  ± 0.6 

20.5  ± 1.0 

17.5  ± 0.8 

18.1  ± 0.8 

** 

32.6  ± 1.2 

36.2  ± 1.6 

29.7  ± 3.1 

30.0  ± 1.6 

**  * 

13.1  ± 0.4 

13.3  ± 0.7 

13.2  ± 0.7 

12.9  ± 0.5 

**  * 

69.8  ± 1.4 

66.1  ± 2.9 

75.0  ± 1.9 

72.3  ± 2.1 

* 

19.5  ± 0.5 

20.9  ± 0.9 

18.6  ± 0.6 

18.3  ± 0.7 

**  * 

84.9  ± 1.3 

86.3  ± 2.1 

79.3  ± 3.4 

86.3  ± 1.8 

* 

54.9  ± 4.8 

86.1  ± 2.1 

46.3  ± 7.3 

24.8  ± 5.9 

** 

** 

30.0  ± 4.5 

0.1  ±0.1 

33.0  ± 6.2 

61.5  ± 5.5 

** 

** 

37.8  ± 3.6 

48.8  ± 5.1 

30.3  ± 6.0 

29.5  ± 6.1 

47.8  ± 3.0 

53.6  ± 4.2 

54.5  ± 5.7 

38.1  ± 5.2 

30.0  ± 4.4 

26.2  ± 5.4 

38.6  ± 15.8 

29.9  ± 6.4 

* 

71.5  ± 8.4 

65.5  ± 10.7 

60.0  ± 15.2 

84.0  ± 16.8 

623  ± 48 

438  ± 38 

623  ± 103 

826  ± 87 

** 

* 

307  ± 28 

340  ± 40 

383  ± 68 

233  ± 44 

31.6  ± 2.8 

27.4  ± 3.4 

24.0  ± 2.0 

39.9  ± 5.8 

* 

25.6  ± 0.9 

27.4  ± 1.5 

22.4  ± 1.7 

25.2  ± 1.4 

** 

320  ± 56 

412  ± 87 

277  ± 111 

468  ± 115 

56.7  ± 8.6 

86.8  ±17.4 

33.9  ± 9.6 

39.1  ± 7.3 

differences  might  also  exist  on  a coarser  scale 
among  more  broadly  defined  and  heteroge- 
neous habitat  types.  Such  differences  might 
furthermore  emerge  on  a landscape  scale 
among  woodland  habitats  in  general. 

The  variables  we  measured  are  not  inde- 
pendent indicators  of  interspecific  differences 
in  nest  site  habitat;  many  of  them  seem  to  be 
related  to  stand  age  or  successional  stage.  Tree 
age  was  not  measured  in  this  study,  but  it  ap- 
pears that  Cooper’s  Hawks  tended  to  use  older 
stands  with  a lower  density  of  taller  and  larger 
trees.  Sharp-shinned  Hawks,  on  the  other 
hand,  tended  to  use  younger  stands  with  a 
higher  density  of  smaller,  shorter  trees.  Reyn- 
olds and  coworkers  (1982)  and  Moore  and 
Henny  (1983)  also  have  suggested  that  differ- 
ences in  accipiter  nest  site  habitat  are  corre- 
lated with  stand  age  or  successional  stage, 
with  Cooper’s  Hawks  using  older  stands  than 
Sharp-shinned  Hawks. 

Interspecific  differences  in  combined  habi- 
tats seldom  seem  the  result  of  contrasting  pro- 
portions of  habitats  used  on  the  intraspecific 
level.  The  lower  percent  coniferous  canopy  in 
combined  Cooper’s  Hawk  habitats  versus 
Sharp-shinned  Hawk  habitats  (30%  vs  61%,  F 
< 0.0005;  Table  2)  appears  to  be  the  result  of 
the  disproportionate  number  of  Cooper’s 


Hawk  sites  in  hardwoods  (42%  of  52  nests) 
compared  to  the  one  Sharp-shinned  Hawk  at 
a hardwood  site  (4%  of  24  nests).  The  differ- 
ence in  deciduous  canopy  cover  between  Coop- 
er’s Hawks  versus  Sharp-shinned  Hawks  in 
combined  habitats  (55%  vs  15%,  P < 0.0005; 
Table  2)  also  seems  to  be  a result  of  contrast- 
ing proportions  of  habitat  use. 

In  addition  to  having  proportionally  more 
nests  in  hardwood  stands.  Cooper’s  Hawks 
nested  in  conifer  plantations  that  had  substan- 
tially greater  deciduous  canopies  than  those 
used  by  Sharp-shinned  Hawks  (25%  vs  4%; 
Table  2).  In  mixed  conifer-hardwood  habitats 
the  deciduous  canopy  cover  percentage  again 
was  greater  for  Cooper’s  Hawks  than  for 
Sharp-shinned  Hawks  (46%  vs  24%;  Table  2). 
Although  neither  of  these  within-habitat  dif- 
ferences was  statistically  significant,  they  are 
clearly  consistent  with  a significant  difference 
in  deciduous  canopy  in  combined  habitats.  On 
a landscape-scale  continuum  from  Wiscon- 
sin’s northern  coniferous  forests  (plus  conifer 
plantations)  to  mixed  and  southern  deciduous 
woodlands  (see  Curtis  1959),  nest  habitat  thus 
appears  to  be  comprised  more  of  deciduous 
sites  or  elements  for  the  Cooper’s  Hawk  and 
coniferous  elements  for  the  Sharp-shinned 
Hawk,  albeit  with  considerable  overlap  in 


TABLE  3.  Nest  tree  species  used  by  Accipiter  striatus  and  A.  cooperii  in  Wisconsin. 


12 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


^ o 


(N'^'^CN(N(N(NCN'^CO’^(N 


(N 


lO  — <N  CNl  -- 


CN 

fN 


fNj  — 


C> 


(N 


SO  fO 


IT) 


(N  (N 


:§ 

s 

A 


g 

-2 

la 

-Q 

"5 

la 

c 

Ot 

<3 

llg 

2 

g 

M 

a 

"a 

- 

o 

Ol  ^ 

cS 

yi 

■Si 

c3 

o 

"a 

cd 

o 

5 

u 

s 

■a 

V 

o 

33 

iS 

TJ 

(U 

a; 

S 

0^ 

a 2 

i| 

a s 

a 

ii  K 

"n 

S 8 

b ^ 

W)  V5 

3 ca 


S 

Ci. 

■S 

2 

oa 

JZ 

u 

Ui 

i5 


3 

2 

5 

Oc 


<U  dJ 

^ p 

cd  5 

Cu  < 


00— '-H-rt— 100— -00  00 


(N  'Tf  ro 


■b"  (N  ro  (N  (N 


■b"  — 


(N  — 


m (N  (N 


!< 

s 

a 


2 

.2 

io 

c 

a 

-S5 


s: 

<u 

c 


.5^ 

sa 


^ «3 


s; 

K 

<u 

c 


4= 

m ^ ^ 


<D 

CJ 

5 

a 
• S oo 
Q. 

x: 

2 ^ 
o o 
o 5 
on  Z 


Q 
<0 

2 

ns 

-Cl 
' '>) 

Q -ii 
^ -Cl 

8 ^ 

2 •- 

au. 

6 

4^  cd 
O (A 

-2  -3 

m oQ 


G 

o 

o 

<3 

'S' 

S 

!a 

T3 

lU 

U 

o 

ic 


■S 

.c 

■2 


W) 

cd 

c 

c/5 


1(4%)  11(46%)  12(50%) 24  22  (42%)  10  (19%)  20  (38%) 


Trexel  et  al.  • ACCIPITER  NEST  SITE  HABITATS 


13 


mixed  forests.  This  divergence  seems  unap- 
parent  in  western  montane  environments 
(Reynolds  1983,  Fischer  1986). 

Siders  and  Kennedy  (1996)  also  found  that 
Cooper’s  Hawks  used  significantly  taller  nest 
trees  with  greater  diameters  and  nest  sites  with 
lower  tree  densities  than  did  Sharp-shinned 
Hawks.  However,  they  reported  that  Sharp- 
shinned  Hawk  nest  sites  had  significantly 
higher  basal  areas  and  canopy  closures  than 
did  those  of  Cooper’s  Hawks.  Reynolds  and 
coworkers  (1982)  found,  as  we  did  that  Coop- 
er’s Hawks  had  greater  nest  heights  in  eastern 
Oregon,  and  used  habitats  at  lower  tree  den- 
sities in  northwestern  Oregon  than  did  Sharp- 
shinned  Hawks.  As  did  Siders  and  Kennedy 
(1996),  they  found  that  Sharp-shinned  Hawk 
nest  sites  had  greater  canopy  closure  than 
those  of  Cooper’s  Hawks.  Moore  and  Henny 
(1983)  noted  that  Cooper’s  Hawk  nests  were 
significantly  higher  than  those  of  Sharp- 
shinned  Hawks,  but  again  in  contrast  to  our 
results  they  found  that  Sharp-shinned  Hawk 
nest  sites  had  significantly  higher  canopy  clo- 
sure than  sites  used  by  Cooper’s  Hawks. 

It  seems  that  significantly  higher  tree  den- 
sities at  Sharp-shinned  Hawk  nest  sites  would 
usually  lead  to  significantly  greater  canopy 
closure,  as  reported  for  mostly  coniferous  hab- 
itats in  New  Mexico  and  Oregon,  but  this  was 
not  the  case  for  combined  habitats  in  Wiscon- 
sin. Our  results  show  lesser  canopy  closure  in 
association  with  greater  tree  densities  at  nest 
sites  of  Sharp-shinned  versus  Cooper’s  hawks 
(Table  2).  This  seeming  contradiction  is  prob- 
ably the  result  of  disproportional  use  of  broad- 
leaved hardwood  forests  (vs  needle-leaved  co- 
nifer forests)  by  Cooper’s  Hawks  and  conse- 
quent effects  of  leaf  surface  on  measures  of 
canopy  closure.  Overstory  canopy  measures 
might  also  be  influenced  by  lower  foliage  den- 
sities (e.g.,  tamarack)  or  strongly  conical 
growth  forms  (e.g.,  black  spruce)  in  some  nest 
tree  species  used  by  Sharp-shinned  Hawks 
(Table  3). 

In  Missouri,  Wiggers  and  Kritz  (1991)  used 
the  most  similar  set  of  habitat  measures  and 
techniques  for  analyzing  those  measures,  yet 
they  reported  no  significant  differences  in  nest 
site  characteristics  for  these  two  accipiters. 
However,  they  divided  their  nest  sites  into 
habitat  types  differently  than  we  did  and  were 
able  to  make  interspecific  comparisons  only 


for  pine  dominated  habitat  (“>  50%  of  over- 
story trees  were  pines”).  Still,  with  small  data 
sets  (Table  2)  and  the  same  alpha  level  (0.001) 
we  detected  significant  differences  in  nest 
height  and  average  canopy  height  for  nest 
sites  in  mixed  conifer-hardwoods,  and  in  nest 
tree  DBH  in  conifer  plantations.  Wiggers  and 
Kritz  (1991)  reported  significant  intraspecific 
differences  between  pine  habitat  and  hard- 
wood habitat  for  Cooper’s  Hawks;  had  they 
combined  these  habitats  they  might  have 
found  overall  interspecific  differences  as  we 
did.  Their  ability  to  detect  significant  differ- 
ences may  also  have  been  hampered  by  the 
fact  that  87%  of  their  nests  were  located  by 
searching  habitat  (especially  coniferous  habi- 
tat) that  was  assumed  a priori  to  be  suitable 
for  one  or  both  species  (Siders  and  Kennedy 
1996).  Consequently,  92%  of  nests  in  conifers 
(n  = 50)  and  77%  of  all  nests  (n  = 60)  were 
situated  in  pine  plantations  of  similar  age  and 
vegetational  structure. 

In  Wisconsin  there  appear  to  be  numerous 
interspecific  differences  in  nest  site  habitats  of 
Cooper’s  and  Sharp-shinned  hawks.  Such  in- 
terspecific differences,  within  and  across  di- 
vergent habitat  types,  may  provide  guidance 
in  identifying  and  managing  the  respective 
nesting  habitats  of  these  birds,  one  or  both  of 
which  have  been  listed  as  species  of  conser- 
vation concern  in  several  midwestern  states 
(Rosenfield  et  al.  1991,  Rosenfield  and  Bie- 
lefeld! 1993).  Many  of  the  nest  tree  and  nest 
site  variables  differing  significantly  between 
species  (tree  heights,  densities,  diameters,  and 
coniferous  components)  are  routinely  and  eas- 
ily estimated  measures  of  woodland  habitats 
among  resource  managers. 

Intraspecific  analyses  of  nest  site  variables 
across  habitat  types  may  also  be  useful  to 
management  and  conservation.  Significant  in- 
traspecific differences  among  habitats  in  the 
Cooper’s  Hawk  would  seem  to  portray  the 
breadth  of  acceptable  nesting  habitat(s).  Var- 
iables that  do  not  differ  intraspecifically 
across  habitat  types  (e.g.,  nest  tree  height  or 
mean  DBH  of  nest  site  trees)  may  serve  as 
focal  points  for  managers  in  identifying  po- 
tentially usable  nesting  habitats,  whether  or 
not  these  features  actually  provide  proximate 
cues  to  nest  site  use  for  the  birds  themselves. 

We  examined  habitat  characteristics  only  at 
the  nest  tree  level  and  in  a small  area  (0.04 


14 


THE  WILSON  BULLETIN 


Vol.  Ill,  No.  1,  March  1999 


ha)  immediately  surrounding  the  nest.  We  did 
not  deal  with  other  habitats  used  by  these  ac- 
cipiters  such  as  hunting  areas  or  non-breeding 
habitats.  Recent  studies  of  nest  site  habitat  in 
the  Cooper’s  Hawk  in  North  Dakota  (M.  Nen- 
neman,  pers.  comm.)  suggest  that  existing  an- 
alyses of  breeding  habitats  from  disparate  ar- 
eas and  woodland  types  may  not  be  general- 
izable  to  other  regions.  Management  impli- 
cations drawn  from  our  Wisconsin  data  should 
therefore  be  cautious. 

ACKNOWLEDGMENTS 

We  gratefully  acknowledge  the  field  assistance  of  T. 
Doolittle,  M.  Fuller,  J.  Partelow,  and  M.  Thwaits,  as 
well  as  the  computer  and  plant  ecology  knowledge  of 
D.  Hillier.  Advice  on  statistical  analyses  and  interpre- 
tation was  provided  by  M.  Bozek,  W.  Gould,  and  B. 
Rogers.  K.  Beal,  P.  Kennedy  and  especially  K.  Kritz 
provided  helpful  reviews  of  this  paper.  Primary  fund- 
ing was  provided  by  the  Wisconsin  Department  of  Nat- 
ural Resources,  the  University  of  Wisconsin-Stevens 
Point,  and  the  Society  for  Tympanuchus  Cupido  Pin- 
natus.  Ltd.  The  Personnel  Development  Committee  at 
the  University  of  Wisconsin-Stevens  Point  provided 
support  for  publication  and  sabbatical  leave  to  RNR. 

LITERATURE  CITED 

Anonymous.  1997.  Birds  in  forested  landscapes.  Wild- 
lifer  281:21-44. 

Curtis,  J.  T.  1959.  The  vegetation  of  Wisconsin:  an 
ordination  of  plant  communities.  Univ.  of  Wis- 
consin Press,  Madison. 

Fischer,  D.  L.  1986.  Daily  activity  patterns  and  habitat 
use  of  coexisting  Accipiter  hawks  in  Utah.  Ph.D. 
diss.,  Brigham  Young  Univ.,  Provo,  Utah. 

James,  F.  C.  and  H.  H.  Shugart,  Jr.  1970.  A quanti- 
tative method  of  habitat  description.  Audubon 
Field  Notes  24:727-736. 

Moore,  K.  R.  and  C.  J.  Henny.  1983.  Nest  site  char- 


acteristics of  three  coexisting  accipiter  hawks  in 
northeastern  Oregon.  Raptor  Res.  17:65-76. 

Reynolds,  R.  T.  1983.  Management  of  western  conif- 
erous forest  habitat  for  nesting  accipiter  hawks. 
USDA  For.  Serv.  Gen.  Tech.  Rep.  RM-102:1— 7. 

Reynolds,  R.  T,  E.  C.  Meslow,  and  H.  M.  Wight. 
1982.  Nesting  habitat  of  coexisting  Accipiter  in 
Oregon.  J.  Wildl.  Manage.  46:124—138. 

Rosenfield,  R.  N.  and  J.  Bielefeldt.  1992.  Natal  dis- 
persal and  inbreeding  in  the  Cooper’s  Hawk.  Wil- 
son Bull.  104:182-184. 

Rosenfield,  R.  N.  and  J.  Bielefeldt.  1993.  Cooper’s 
Hawk  {Accipiter  cooperii).  In  The  birds  of  North 
America,  no.  75  (A.  Poole  and  E Gill,  Eds.).  The 
Academy  of  Natural  Sciences,  Philadelphia,  Penn- 
sylvania; The  American  Ornithologists’  Union, 
Washington,  D.C. 

Rosenfield,  R.  N.  and  J.  Bielefeldt.  1996.  Lifetime 
nesting  area  fidelity  in  male  Cooper’s  Hawks  in 
Wisconsin.  Condor  98:165-167. 

Rosenfield,  R.  N.,  J.  Bielefeldt,  R.  K.  Anderson, 
AND  J.  M.  Papp.  1991.  Status  reports:  accipiters. 
Natl.  Wildl.  Fed.  Sci.  Tech.  Serv.  15:42—49. 

Rosenfield,  R.  N.,  J.  Bielefeldt,  J.  L.  Affeldt,  and 
D.  J.  Beckmann.  1996.  Urban  nesting  biology  of 
Cooper’s  Hawks  in  Wisconsin.  Pp.  41-44  in  Rap- 
tors in  human  landscapes  (D.  M.  Bird,  D.  E.  Var- 
land,  and  J.  J.  Negro,  Eds.).  Academic  Press,  Lon- 
don, U.K. 

Siders,  M.  S.  and  P.  L.  Kennedy.  1996.  Forest  struc- 
tural characteristics  of  accipiter  nesting  habitat:  is 
there  an  allometric  relationship?  Condor  98:123- 
132. 

SoKAL,  R.  R.  AND  F.  J.  Rohlf.  1981.  Biometry,  second 
ed.  W.  H.  Freeman  and  Co.,  San  Francisco,  Cali- 
fornia. 

Titus,  K.  and  J.  A.  Mosher.  1981.  Nest-site  habitat 
selected  by  woodland  hawks  in  the  central  Ap- 
palachians. Auk  98:270-281. 

Wiggers,  E.  P.  and  K.  j.  Kritz.  1991.  Comparison  of 
nesting  habitat  of  coexisting  Sharp-shinned  and 
Cooper’s  hawks  in  Missouri.  Wilson  Bull.  103: 
568-577. 

Wilkinson,  L.  1992.  SYSTAT:  the  system  for  statis- 
tics. SYSTAT  Inc.,  Evanston,  Illinois. 


Wilson  Bull.,  111(1),  1999,  pp.  15-21 


MADAGASCAR  FISH-EAGLE  PREY  PREFERENCE  AND 

FORAGING  SUCCESS 

JAMES  BERKELMAN,"^  JAMES  D.  FRASER,'  AND  RICHARD  T.  WATSON^ 


ABSTRACT. — We  investigated  Madagascar  Fish-Eagle  (Haliaeenis  vociferoides)  foraging  ecology  to  deter- 
mine prey  preference  and  the  effect  of  fish  abundance  on  fish-eagle  foraging  rates  and  foraging  success.  We 
observed  fish-eagle  foraging  behavior  at  nine  lakes  in  western  Madagascar  from  May  to  August  1996.  We 
sampled  the  fish  population  at  each  lake  using  gill  nets  and  recorded  fish  weights  and  species.  Introduced  tilapia, 
Oreochromis  spp.  and  Tilapia  spp.,  made  up  the  majority  of  both  the  gill  net  (66.3%)  and  fish-eagle  catch 
(64.7%)  in  simiku"  proportion,  suggesting  that  the  fish-eagle  is  an  opportunistic  predator.  Consequently,  replace- 
ment of  native  fish  species  by  exotics  probably  has  not  been  detrimental  to  the  island’s  fish-eagle  population. 
Male  fish-eagle  foraging  success  was  positively  correlated  (P  < 0.001)  with  number  of  fish  species,  suggesting 
that  fish  species  diversity  may  affect  fish-eagle  foraging  effectiveness.  Received  24  July  1997,  accepted  2 Nov. 
1998. 


Prey  availability  influences  breeding  area 
selection  (Swenson  et  al.  1986),  breeding  den- 
sity (Dzus  and  Gerrard  1989),  reproductive 
success  (Grubb  1995),  and  date  of  breeding 
(Hansen  1987)  in  Bald  Eagles  {Haliaeetus 
leucocephalus)  and  productivity  in  White- 
tailed Eagles  {Haliaeetus  albicilla\  Helander 
1985).  It  also  affects  distribution  and  density 
of  Bald  Eagles  at  wintering  sites  (Griffin  and 
Baskett  1985,  Sabine  and  Klimstra  1985, 
Keister  et  al.  1987,  Hunt  et  al.  1992b)  and 
migratory  stopovers  (Fraser  et  al.  1985,  Ben- 
netts and  McClelland  1991). 

Although  prey  availability  is  clearly  impor- 
tant to  Haliaeetus  eagles,  there  has  been  little 
research  aimed  at  quantitatively  determining 
prey  abundance  and  its  effects  on  prey  selec- 
tion, foraging  rates,  and  foraging  success  in 
the  genus.  Steenhof  (1976),  Mersmann 
(1989),  and  Hunt  and  coworkers  (1992a)  used 
gill  nets  to  inventory  relative  fish  abundance 
and  determined  that  the  most  frequently  netted 
fish  species  made  up  the  greatest  proportion 
of  the  Bald  Eagle’s  diet.  Wintering  Bald  Ea- 
gles in  New  Mexico  fed  most  frequently  on 
big  game  carrion  when  it  was  the  most  abun- 


‘ Dept,  of  Fisheries  and  Wildlife  Sciences,  Virginia 
Polytechnic  Inst,  and  State  Univ.,  Blacksburg,  VA 
24061-0321. 

^ The  Peregrine  Fund,  566  West  Flying  Hawk  Lane, 
Boise,  ID  83709. 

^ Present  address:  Dept,  of  Wildlife  Ecology,  Univ. 
of  Wisconsin,  226  Russell  Labs,  1630  Linden  Dr., 
Madison,  WI  53706-1598. 

■*  Corresponding  author;  Email: 
jberkelman@facstaff.wisc.edu 


dant  prey  source  (Grubb  1984).  There  is  a 
positive  relationship  between  prey  abundance 
and  foraging  success  of  wintering  Bald  Eagles 
both  between  locations  (Stalmaster  and  Plett- 
ner  1992)  and  between  years  (Brown  1993). 
Knight  and  Knight  (1983)  found  a negative 
correlation  between  search  time  and  relative 
prey  abundance  of  Bald  Eagles  wintering  in 
Washington,  but  Mersmann  (1989)  did  not 
find  a correlation  between  Bald  Eagle  forag- 
ing rates  and  gill  net  catch  rates  on  the  north- 
ern Chesapeake  Bay. 

Langrand  and  Meyburg  (1989)  and  Raza- 
findramanana  (1995)  have  documented  fish 
species  eaten  by  Madagascar  Fish-Eagles 
{Haliaeetus  vociferoides),  but  there  has  been 
no  previous  attempt  to  quantitatively  assess 
the  eagle’s  diet.  The  objectives  of  this  study 
were  (1)  to  describe  the  diet  and  foraging  be- 
havior of  the  Madagascar  Fish-Eagle  at  lakes 
in  western  Madagascar,  (2)  to  determine  fish- 
eagle  prey  preference,  and  (3)  to  determine 
whether  fish-eagle  foraging  rates  and  foraging 
success  are  dependent  on  prey  abundance. 

STUDY  AREA  AND  METHODS 

We  observed  Madagascar  Fish-Eagle  foraging  ecol- 
ogy from  22  May  to  4 August  1996  at  nine  lakes  in 
the  Tsiribihina,  Manambolo,  and  Beboka  river  drain- 
ages between  the  Bongolava  escarpment  and  the  Mo- 
zambique Channel  in  western  Madagascar  (Table  1 ). 
We  selected  lakes  that  we  felt  would  offer  the  best 
conditions  for  viewing  eagles  throughout  the  day  from 
among  32  lakes  with  resident  Madagascar  Fish-Eagle 
pairs  that  we  studied  in  1995  (Berkelman  1997). 

We  observed  fish-eagle  foraging  behavior  through- 
out daylight  hours  from  06:00  to  18:00  (GMT  + 3 h) 
for  six  or  seven  days  at  each  lake.  We  recorded  both 


15 


16 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


TABLE  1.  Lakes  included  in  Madagascar  fish-eagle  foraging  ecology  study  in  western  Madagascar,  May- 
August,  1996. 

Lake 

Coordinate.s 

Dates  observed 

Ambereny 

18° 

55'  S, 

44° 

23'  E 

22-28  May 

Bejijo 

19° 

13'  S, 

44° 

32'  E 

30  May-5  June 

Ankazomena 

19° 

42'  S, 

45° 

23'  E 

8-15  June 

Asonjo 

19° 

50'  S, 

45° 

26'  E 

16—23  June 

Ampamandrika 

19° 

46'  S, 

44° 

34'  E 

27  June— 3 July 

Befotaka 

19° 

1'  S,  44°  24'  E 

7-12,  20  July 

Masama 

18° 

50'  S, 

44° 

27'  E 

13-19  July 

Bevoay 

19° 

9'  S,  44°  24'  E 

22-28  July 

Tsiandrora 

18° 

58'  S, 

44° 

38'  E 

30  July— 4 August 

observer-time,  the  number  of  hours  spent  observing 
eagles,  and  eagle-time,  the  product  of  observer-time 
and  the  number  of  eagles  observed  during  each  hour, 
tor  each  lake.  At  the  three  lakes  that  had  more  than 
one  resident  fish-eagle  pair,  we  randomly  selected  one 
of  the  pairs  for  observation  during  the  period.  We 
watched  from  an  inflatable  kayak  or  from  the  shore 
using  10  X 50  binoculars  and  a 15-45  X spotting 
.scope  on  a rifle  mount.  We  concentrated  on  the  resi- 
dent eagle  pair  while  also  noting  the  behavior  of  any 
other  eagles  that  we  could  see,  including  immatures 
and  other  adults  that  were  associated  with  the  resident 
pair.  We  distinguished  adult  fish-eagle  sexes  by  the 
smaller  size  and  higher  pitched  vocalizations  of  the 
males.  We  distinguished  adults  from  immatures  by 
their  vocalizations  and  by  the  completely  white  plum- 
age on  the  tails  and  the  faces  of  the  adults  (Langrand 
and  Meyburg  1989). 

We  recorded  prey  searches  when  eagles  flew  low 
over  the  water  looking  down  (Stalmaster  and  Plettner 
1992)  and  kills  when  they  picked  up  a fish  from  the 
water.  We  al.so  noted  instances  of  fish-eagles  scaveng- 
ing dead  fish  from  the  shoreline  or  pirating  fish  from 
Black  Kites  (Milvu.s  migrans).  We  identified  fish  to 
species  whenever  possible.  If  we  could  not  identify  the 
fish  while  the  eagle  was  in  flight,  we  looked  for  prey 
remains  on  the  ground  beneath  the  eagle’s  feeding 
perch. 

We  .set  two  monofilament  gill  nets  for  3 h at  each 
lake  starting  at  06:00-06:15.  The  gill  nets  had  a foam 
core  float  rope  and  a lead  core  bottom  rope,  were  0.91 
m deep  by  45.7  m long,  and  were  divided  into  three 
15.2  m panels  of  2.5,  3.8,  and  5.1  cm  mesh  size.  We 
attached  floats  to  the  first  net  and  .set  it  parallel  to  the 
shore  in  water  about  0.9  m deep.  We  set  the  second 
net  on  the  bottom,  parallel  to  the  shore,  in  water  about 
1 .8  m deep.  Thus,  we  sampled  fish  from  among  the 
first  and  second  0.9  m of  the  water  column.  If  the  lake 
was  less  than  1.8  m deep,  we  .set  the  .second  net  in  the 
deepest  water  within  200  m of  where  we  had  set  the 
first  net.  We  placed  nets  adjacent  to  the  nest  or,  where 
we  did  not  find  a nest  (/;  = I ),  adjacent  to  a frequently 
used  perch.  We  believed  that  the.se  sites  were  repre- 
sentative of  fish-eagle  foraging  areas  becau.se  we  ob- 


served the  eagles  forage  most  frequently  within  300  m 
of  the  nest  at  eight  of  the  nine  lakes. 

We  identified  each  fish  caught  in  the  gill  nets  using 
keys  (Arnoult  1959,  Kiener  1963,  Glaw  and  Vences 
1994)  and  weighed  it  to  the  nearest  g and  measured 
total  fish  length  to  the  nearest  cm.  We  combined  the 
data  for  the  three  days  that  we  sampled  each  lake  (nine 
hours  total)  and  calculated  total  number  of  fish  caught, 
total  weight  (kg)  of  fish  catch,  average  fish  weight  (g), 
and  number  of  species.  We  did  not  include  fish  that 
weighed  over  1.5  kg  in  these  calculations  because  we 
did  not  see  fish-eagles  capture  larger  fish. 

We  conducted  the  test  of  equal  proportions  to 
determine  it  fish-eagle  use  of  fish  species  was  different 
from  expected  use  based  on  gill  net  samples  using  SAS 
on  an  IBM  compatible  computer  (PROC  EREQ,  SAS 
Institute  Inc.  1990).  We  excluded  unidentified  prey 
from  this  analysis.  After  finding  a significant  {P  < 
0.05)  overall  difference,  we  tested  the  hypothesis  of  no 
difference  between  use  and  availability  of  each  fish 
species,  following  Marcum’s  and  Loftsgaarden’s 
(1980)  technique.  We  calculated  Spearman  correlation 
coefficients  between  fish-eagle  foraging  variables  and 
fish  variables  (PROC  CORR,  SAS  Institute  Inc.  1990). 
For  all  analyses,  we  used  an  overall  confidence  level 
of  a — 0.05  and  a confidence  level  of  a.lk,  where  k 
was  the  number  of  significance  values  calculated,  fol- 
lowing the  Bonferroni  approach  (Miller  1966). 

RESULTS 

Foraging  behavior. — There  were  extra 
adult  Madagascar  Fish-Eagles  associated  with 
three  pairs  and  immatures  with  another  three 
of  the  nine  resident  pairs  that  we  studied.  Al- 
together, we  observed  1 1 adult  males,  10  adult 
females,  and  3 immatures. 

Hunting  methods  were  similar  to  those  used 
by  other  .sea  eagles  (Brown  1980,  Love  1983, 
Stalmaster  1987).  The  fish-eagles  we  observed 
hunted  from  perches  and  either  stooped  di- 
rectly from  a perch  or  searched  low  over  the 
water,  generally  returning  to  perch  within  5 


Herkelimin  et  al.  • MADAGASCAR  FISH-EAGLE  FORAGING 


17 


TABLE  2.  Male  Madagascar  Fish-Eagle  foraging 
at  nine  lakes  in  western  Madagascar,  May-Aiigust, 
1996. 


Variable 

.X  (n  = 11) 

SE 

Searches 

36.1 

8.0 

Kills 

5.9 

1.7 

Searches/hour/eagle 

0.68 

0.15 

Kills/hour/eagle 

0.10 

0.03 

Kills/search 

0.16 

0.04 

TABLE  3.  Number  of  Hsh,  total  fish  weight,  av- 
erage fish  weight,  and  number  of  hsh  species  caught 
in  gill  nets  at  nine  lakes  occupied  by  Madagascar  Fish- 
Eagles  in  western  Madagascar,  May-August,  1996. 


Variable 

X 

SE 

Range 

Number  of  fish 

30.1 

7.3 

4-66 

Total  weight,  kg 

4.6 

1.6 

0.2-15.9 

Average  weight,  g 

139.0 

23.7 

55.3-269.3 

Number  of  species 

3.9 

0.4 

2-6 

min  of  leaving.  When  striking,  the  eagles  en- 
tered the  water  feet  first  at  a low  angle  and 
only  took  fish  that  were  at  or  just  below  the 
surface. 

We  watched  eagles  for  669.5  h observer- 
time and  1030.98  h eagle-time,  including 
490.25  h (47.6%)  male  eagle-time,  526.0  h 
(51.0%)  female  eagle-time,  and  14.73  h 
(1.4%)  immature  eagle-time.  We  recorded  67 
occurrences  of  eagles  obtaining  fish,  including 
60  (89.6%)  occasions  when  they  captured  fish 
in  open  water,  3 (4.4%)  when  they  scavenged 
dead  fish  from  the  shoreline,  and  4 (6.0%) 
when  they  stole  fish  from  Black  Kites.  We 
also  recorded  32  occurrences  of  eagles  eating 
fish  or  delivering  fish  to  their  mates  when  we 
did  not  see  an  eagle  obtain  the  fish.  On  one 
occasion  we  observed  an  eagle  eating  a do- 
mestic duckling  (Anas  sp.).  We  did  not  see  the 
eagle  capture  the  duckling,  but  the  local  peo- 
ple claimed  that  the  same  eagle  pair  had  killed 
domestic  ducklings  and  turkey  (Meleagris  sp.) 
poults  at  the  same  lake  on  several  occasions 
in  1996. 

Of  the  67  occasions  when  we  saw  eagles 
obtain  fish,  the  eagles  were  adult  males  on  53 
(79.1%)  occasions,  adult  females  on  13 
(19.4%),  and  an  immature  on  1 (1.5%)  occa- 
sion. Nine  (69.2%)  of  13  adult  females  that 
we  saw  capture  fish  were  not  nesting  at  the 
time.  The  other  four  (31.8%)  were  incubating 
eggs.  All  32  occasions  on  which  we  saw  ea- 
gles eating  or  delivering  fish  but  did  not  see 
them  catch  the  fish  involved  adult  male  ea- 
gles. All  four  instances  of  piracy  from  kites 
occurred  at  the  same  lake  and  involved  two 
cooperating  adult  male  eagles  associated  with 
the  same  territory.  In  each  case,  the  eagles  ha- 
rassed a kite  until  it  dropped  its  fish,  which 
one  of  the  eagles  then  retrieved. 

Foraging  rates  and  fish  abundance. — Male 


fish-eagle  kills/search  was  positively  correlat- 
ed with  number  of  fish  species  caught  in  gill 
nets  (p  = 0.909,  P < 0.001).  There  were  no 
other  significant  correlations  between  fish-ea- 
gle foraging  rates  (Table  2)  and  fish  variables 
(Table  3).  We  only  analyzed  male  foraging  be- 
cause we  rarely  saw  females  forage. 

The  Madagascar  Fish-Eagle  search  rate 
peaked  in  the  early  morning  and  again,  at  a 
higher  level,  in  the  early  afternoon  (Fig.  lA). 
Foraging  success,  expressed  as  the  proportion 
of  prey  searches  that  resulted  in  kills,  was 
highest  before  10:00  and  after  16:00  but  lower 
between  these  times  (Fig.  IB). 

Dietary  preference. — Of  the  99  observed 
occurrences  of  fish-eagles  capturing,  carrying, 
or  eating  fish,  we  were  able  to  identify  68 
(68.7%)  either  to  species  or  to  a closely  re- 
lated group  of  species  (Table  4).  We  were  un- 
able to  identify  eagle-caught  tilapia  to  species 
or  to  distinguish  between  the  closely  related 
Oreochromis  and  Tilapia  genera.  In  our  gill 
net  samples,  we  caught  271  fish  of  12  species, 
including  four  species  of  tilapia  and  eight  oth- 
er species.  The  total  weight  of  the  catch  at  all 
nine  lakes  was  41.1  kg. 

We  combined  all  native  fish  species  into  a 
single  group  because  our  catches  of  each  spe- 
cies were  too  small  to  analyze  separately  (Ta- 
ble 4).  The  proportions  of  fish  species  differed 
significantly  between  the  fish-eagle  catch  and 
the  gill  net  catch  (x^  = 41.97,  df  = 4,  P = 
0.001).  The  95%  confidence  limits  for  the  dif- 
ference between  the  proportion  used  and  the 
proportion  available  suggested  that  fish-eagles 
catch  Ophicephalus  striatus  in  greater  propor- 
tion, Cyprinus  carpio  in  lesser  proportion,  and 
tilapia,  Heterotis  niloticus,  and  native  species 
in  equal  proportion  to  their  relative  abun- 
dance. 


18 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  1.  March  1999 


A. 


B. 


£ 

u 

re 

(1) 

w 


jfl 

5 


0.3  . 


Hour  of  day 


FIG.  1.  Madagascar  Fish-Eagle  prey 
by  time  of  day  at  nine  lakes  in  western 


searches  per  hour  per  eagle  (A)  and  prey  searches 
Madagascar,  May-August,  1996. 


resulting  in  kills  (B) 


DISCUSSION 

Foraging  behavior. — Our  observation  that 
4.4%  of  fish  taken  by  Madagascar  Fish-Eagles 
were  scavenged  was  lower  than  scavenging 
rates  that  Mersmann  (1989)  and  Brown  (1993) 
reported  for  Bald  Eagles  (25%  and  7.7%,  re- 
spectively) but  comparable  to  the  4%  reported 
by  Stalmaster  and  Plettner  (1992).  We  did  not 
.see  fish-eagles  take  floating  dead  fish  from  the 
surface  in  open  water,  but  it  is  possible  that 
some  of  the  fish  that  we  observed  eagles  catch 
from  a distance  were  dead  fish  floating  below 
the  water  surface. 

We  are  unaware  of  previous  reports  of  pi- 


racy or  capture  of  avian  prey  by  the  Mada- 
gascar Fish-Eagle.  It  is  unclear  why  we  ob- 
served four  instances  of  piracy  at  one  of  the 
lakes  and  none  at  the  other  eight  lakes.  We 
saw  numerous  other  fish-eating  birds  at  all  the 
lakes,  including  Black  Kites,  herons,  storks, 
anhingas,  and  cormorants. 

Although  we  observed  adult  male  eagles  in- 
cubating eggs  and  tending  nestlings,  it  appears 
that  the  male  does  most  of  the  foraging  for  the 
pair,  at  least  during  early  nesting.  The  four 
instances  in  which  we  observed  incubating  fe- 
males catch  fish  occurred  near  the  nest  when 
males  were  not  present.  The  most  advanced 


Berkelmcm  et  al.  • MADAGASCAR  FISH-EAGLE  FORAGING 


19 


TABLE  4.  Fish  caught  (number  and  % of  total)  and  number  of  lakes  where  fish  were  caught  (out  of  nine) 
by  Madagascar  Fish-Eagles  and  in  gill  nets  in  western  Madagascar,  May-August,  1996. 

Fish  species“,  family 

Fish-eagle  catch 

Gill  nei  catch 

No.  of 
Hsh 

% of 
fish 

No.  of 
lakes 

No,  of 

Hsh 

% of 
fish 

No.  of 
lakes*’ 

Exotic 

Tilapia'-',  Cichlidae 

44 

64.7 

9 

183 

67.5 

9 

Heteroti.'!  niloticu.s,  Osteoglossidae 

7 

10.3 

3 

17 

6.3 

3 

Cyprinus  carpio,  Cyprinidae 

2 

2.9 

2 

33 

12.2 

1 (2) 

Ophicephalus  striatus,  Channidae 

10 

14.7 

6 

1 

0.4 

1 (6) 

Native 

Megalops  cyprinoicles,  Megalopidae 

2 

2.9 

2 

18 

6.6 

6 (1) 

Arius  madagascariensis,  Ariidae 

3 

4.4 

3 

14 

5.2 

5 

Glossogobius  giuris,  Gobiidae 

0 

0 

0 

2 

0.7 

1 

Ambassis  gymnocephalus,  Ambassidae 

0 

0 

0 

2 

0.7 

1 

Scatophagus  tetracanthus,  Scatophagidae 

0 

0 

0 

1 

0.4 

1 

TOTAL 

68 

100 

271 

100 

“ Unidentified  fish  that  fish-eagles  caught  (n  = 31)  were  excluded. 

^Number  in  parentheses  represents  additional  lakes  where  each  fish  species  was  known  to  be  present  either  from  1995  gill  net  sampling  or  from  fish 
catches  of  local  fishermen. 

Tilapia  species  included  in  order  of  decreasing  gill  net  catch:  Oreocliromis  macrochir,  Tilapia  zilii.  O.  mossamhicus,  and  O.  niloticus. 


nesting  attempt  we  observed  had  a 2-3  week 
old  downy  chick,  so  we  were  unable  to  doc- 
ument whether  female  fish-eagle  foraging 
rates  change  as  nesting  progresses. 

Foraging  rates  and  fish  abundance. — The 
strong  positive  correlation  between  fish-eagle 
kills/search  and  number  of  fish  species  may 
indicate  that  the  eagles  forage  most  effectively 
at  lakes  that  have  the  highest  fish  species  di- 
versity. In  a previous  study,  Berkelman  (1997) 
found  that  fish  species  diversity,  along  with 
shoreline  perch  density,  was  one  of  the  best 
predictors  of  fish-eagle  lake  use,  lending  fur- 
ther support  to  the  importance  of  fish  species 
diversity.  However,  the  strength  of  the  corre- 
lation between  foraging  success  and  fish  spe- 
cies diversity  may  be  related  to  the  low  range 
of  variability  in  number  of  fish  species  caught 
(2  to  6)  at  lakes  in  this  study. 

The  early  morning  peak  in  search  rate  that 
we  observed  also  was  reported  for  Madagas- 
car Fish-Eagles  by  Razafindramanana  (1995) 
and  for  Bald  Eagles  (Steenhof  et  al.  1980, 
Mersmann  1989)  and  Ospreys  (Pandion  hali- 
aetus;  Flemming  and  Smith  1990).  This  peak 
may  result  from  hunger  after  fasting  overnight 
or  from  eagles  taking  advantage  of  greater  fish 
availability  and  calmer  weather  during  the 
early  morning  hours.  The  early  afternoon 
search  rate  peak  may  reflect  eagles  that  have 
digested  the  morning  food  and  are  hungry 
again.  Whitfield  and  Blaber  (1978)  observed 


a midday  foraging  peak  in  African  Fish-Ea- 
gles (Haliaeetus  vocifer)  and  suggested  that 
the  eagles  were  taking  advantage  of  thermals 
at  this  time,  but  the  Madagascar  Fish-Eagles 
that  we  observed  foraged  mostly  low  over  the 
water  from  a perch.  Daily  weather  patterns 
varied  little  during  the  study,  so  differences 
among  eagle  pairs  in  foraging  rates  and  suc- 
cess probably  were  not  related  to  weather. 

Dietary  preference. — The  results  suggest 
that  Madagascar  Fish-Eagles  prefer  Ophice- 
phalus  striatus  to  other  fish  and  avoid  Cypri- 
nus  carpio.  Ophicephalus  striatus  was  the 
largest  fish  species  that  we  saw  fish-eagles 
capture.  We  estimated  the  largest  ones  caught 
by  fish-eagles  to  be  between  1 and  1.5  kg. 
This  species  is  a predatory  fish  that  was  intro- 
duced to  Madagascar  in  1978  (Reinthal  and 
Stiassny  1991).  In  field  experiments.  Bald  Ea- 
gles selected  large  fish  more  often  than  small- 
er fish  during  the  breeding  season,  but  not  dur- 
ing the  non-breeding  season  (Jenkins  and 
Jackman  1995);  Madagascar  Fish-Eagle  pref- 
erence for  O.  striatus  may  reflect  the  eagle’s 
greater  energy  requirements  during  the  breed- 
ing season.  Fish-eagles  showed  no  preference 
for  Heterotis  niloticus,  another  large  intro- 
duced fish  species,  but  it  was  present  in  only 
three  (33.3%)  of  the  nine  lakes.  Cyprinus  car- 
pio may  not  be  used  because  this  species  feeds 
on  the  bottom  of  lakes  (Scott  and  Crossman 
1973)  where  it  is  difficult  for  eagles  to  catch. 


20 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  1,  March  1999 


The  results  also  may  indicate  biases  in  the 
fish  abundance  data.  Ophicephalus  striatus,  a 
visually  orienting  predator,  may  be  better  at 
avoiding  entanglement  in  gill  nets  than  the 
other  fish  species.  All  of  the  C.  carpio  abun- 
dance data  were  from  a single  lake  where  we 
caught  33  individuals.  This  lake  was  so  shal- 
low (0.9  m)  that  both  gill  nets  extended  to  the 
bottom  and  consequently  were  more  likely  to 
catch  bottom-dwelling  fish  such  as  C.  carpio. 

The  use  and  relative  abundance  data  for  ti- 
lapia,  the  most  abundant  fish  in  all  of  the 
lakes,  were  similar  (64.7%  of  identifiable  fish- 
eagle  catch  and  67.5%  of  gill  net  catch).  Ti- 
lapia  were  introduced  to  Madagascar  for  aqua- 
culture in  the  1950s  (Kiener  1963)  and  have 
since  spread  to  most  bodies  of  freshwater 
throughout  the  island.  The  predominance  of 
tilapia  in  the  fish-eagles’  diet  in  this  study 
suggests  that  the  Madagascar  Fish-Eagle  is  an 
opportunistic  predator  that  catches  whatever 
prey  species  are  most  abundant.  Thus,  the 
marked  change  in  species  composition  of 
Madagascar’s  freshwater  fish  fauna  resulting 
from  exotic  species  introductions  (Loiselle 
1993,  Reinthal  et  al.  1995)  probably  has  not 
been  detrimental  to  the  island’s  fish-eagle  pop- 
ulation. The  positive  relationship  between 
fish-eagle  foraging  success  and  number  of  fish 
species  suggests  that  the  fish-eagle  population 
may  be  sensitive  to  declines  in  fish  species 
diversity. 

ACKNOWLEDGMENTS 

We  thank  C.  Razafimahatratra,  G.  Tohaky,  E Paul, 
and  M.  Philbert  for  technical  a.ssistance.  Thanks  to  Y. 
Rakotonirina  for  his  driving  and  mechanical  skills. 
Thanks  to  J.  Rajesy,  R.  Rabarisoa,  and  M.  Razafin- 
drakoto  for  administrative  and  logistical  support.  We 
thank  C.  A.  Haas,  J.  J.  Ney,  R.  G.  Oderwald,  D.  E 
Stauffer,  and  R.  Thorstrom  for  comments  on  the  man- 
uscript. We  received  funding  from  the  National  Geo- 
graphic Society,  the  Raptor  Research  Foundation,  The 
International  Osprey  Foundation,  the  World  Nature  As- 
•sociation,  the  American  Museum  of  Natural  History, 
and  the  Cooper  Ornithological  Society.  We  conducted 
this  study  under  the  auspices  of  The  Peregrine  Fund’s 
Madaga.scar  Fish-Eagle  and  Wetland  Conservation 
Project  in  Madaga.scar. 

LITERATURE  CITED 

Arnoult,  j.  1959.  Pois.son  des  eaux  donees.  Faune  de 
Madaga.scar  10.  ORSTOM  and  CNRS,  Paris, 
France. 

Bknnktt.s,  R.  E.  and  B.  R.  McCi.klland.  1991.  Dif- 


ferences in  the  distribution  of  adult  and  immature 
Bald  Eagles  at  an  autumn  concentration  in  Mon- 
tana. Northwest  Sci.  65:223-230. 

Berkelman,  j.  1997.  Habitat  requirements  and  forag- 
ing ecology  of  the  Madagascar  Fish-Eagle.  Ph.D. 
diss.,  Virginia  Polytechnic  Inst,  and  State  Univ., 
Blacksburg. 

Brown,  B.  T.  1993.  Winter  foraging  ecology  of  Bald 
Eagles  in  Arizona.  Condor  95:132—138. 

Brown,  L.  1980.  The  African  Fish  Eagle.  Bailey  Bros, 
and  Swinfen  Ltd,  Folkstone,  U.K. 

Dzus,  E.  H.  AND  J.  M.  Gerrard.  1989.  Interlake  var- 
iations of  Bald  Eagle,  Haliaeetiis  leucocephalus, 
populations  in  north-central  Saskatchewan.  Can. 
Field-Nat.  103:29-33. 

Flemming,  S.  P.  and  P.  C.  Smith.  1990.  Environmental 
influences  on  Osprey  foraging  in  northeastern 
Nova  Scotia.  J.  Raptor  Res.  24:64-67. 

Fraser,  J.  D.,  L.  D.  Frenzel,  J.  E.  Mathisen,  and  M. 
E.  Shough.  1985.  Seasonal  distribution  of  sub- 
adult Bald  Eagles  in  three  Minnesota  habitats. 
Wilson  Bull.  97:365-366. 

Glaw,  E and  M.  Vences.  1994.  A fieldguide  to  the 
amphibians  and  reptiles  of  Madagascar  including 
mammals  and  freshwater  fish,  second  ed.  Moos 
Dmck,  Leverkusen  and  Farbo,  Cologne,  Germany. 

Griffin,  C.  R.  and  T.  S.  Baskett.  1985.  Food  avail- 
ability and  winter  range  sizes  of  immature  and 
adult  Bald  Eagles.  J.  Wildl.  Manage.  49:592-594. 

Grubb,  T.  G.  1984.  Winter  activity  of  Bald  Eagles  Hali- 
aeetus  leucocephalus  at  Navajo  Lake,  New  Mex- 
ico. Southwest.  Nat.  29:335-342. 

Grubb,  T.  G.  1995.  Food  habits  of  Bald  Eagles  breed- 
ing in  the  Arizona  desert.  Wilson  Bull.  107:258- 
274. 

Hansen,  A.  J.  1987.  Regulation  of  Bald  Eagle  repro- 
ductive rates  in  southeast  Alaska.  Ecology  68- 
1387-1392. 

Helander,  B.  1985.  Reproduction  of  the  White-tailed 
Sea  Eagle  Haliaeetiis  alhicilla  in  Sweden.  Ho- 
larct.  Ecol.  8:21  1-227. 

Hunt,  W.  G.,  J.  M.  Jenkins,  R.  E.  Jackman,  C.  G. 
Thelander,  and  a.  T.  Gerstell.  1992a.  Foraging 
ecology  of  Bald  Eagles  on  a regulated  river.  J. 
Raptor  Res.  26:243-256. 

Hunt,  W.  G.,  B.  S.  Johnson,  and  R.  E.  Jackman. 
1992b.  Carrying  capacity  for  Bald  Eagles  winter- 
ing along  a northwestern  river.  J.  Raptor  Res  '’b' 
49-60. 

Jenkins,  J.  M.  and  R.  E.  Jackman.  1995.  Field  exper- 
iments m prey  selection  by  resident  Bald  Eagles 
m the  breeding  and  non-breeding  season.  J.  Field 
Ornithol.  65:441-446. 

Kelster,  G.  R,  Jr.,  r.  g.  Anthony,  and  E.  J.  O'Neill. 
1987.  Use  of  communal  roosts  and  foraging  areas 
by  Bald  Eagles  wintering  in  the  Klamath  Basin. 

J.  Wildl.  Manage.  51:415-420. 

Kiener,  A.  1963.  Poissons,  peche  et  pisciculture  a 
Madaga.scar.  Centre  Technique  Forestier  Tropical 
24,  Nogent-sur-Marne,  France. 

Knight,  S.  K.  and  R.  L.  Knight.  1983.  Aspects  of 


Berkeltmm  el  cil.  • MADAGASCAR  FISH-EAGLE  FORAGING 


21 


food  finding  by  wintering  Bald  Eagles.  Auk  100: 
477-484. 

Langrand,  O.  and  B.-U.  Meyburg.  1989.  Range,  sta- 
tus, and  biology  of  the  Madagascar  Sea  Eagle 
Haliaeetiis  vociferoides.  Pp.  269-277  in  Raptors 
in  the  modern  world  (B.-U.  Meyburg  and  R.  D. 
Chancellor,  Eds.).  World  Working  Group  for  Birds 
of  Prey,  Berlin,  Germany. 

Loiselle,  P.  V.  1993.  Madagascar  update.  Aquatic  Sur- 
vival 2:1-3. 

Love,  J.  A.  1983.  The  return  of  the  Sea  Eagle.  Cam- 
bridge Univ.  Press,  Cambridge,  U.K. 

Marcum,  C.  L.  and  D.  O.  Loftsgaarden.  1980.  A 
nonmapping  technique  for  studying  habitat  pref- 
erences. J.  Wildl.  Manage.  44:963-968. 

Mersmann,  T.  j.  1989.  Eoraging  ecology  of  Bald  Ea- 
gles on  the  northern  Chesapeake  Bay  with  an  ex- 
amination of  techniques  used  in  the  study  of  Bald 
Eagle  food  habits.  M.S.  thesis,  Virginia  Polytech- 
nic Inst,  and  State  Univ.,  Blacksburg. 

Miller,  R.  1966.  Simultaneous  statistical  inference. 
McGraw-Hill  Book  Company,  New  York. 

Razafindramanana,  S.  1995.  Contribution  a I’etude 
de  la  biologie  de  Haliaeetus  vociferoides,  Desmur 
1845  (Pygargue  de  Madagascar):  reproduction  et 
domaine  vital.  Memoire  de  Diplome  d’Etudes  Ap- 
profondies,  Univ.  of  Antananarivo,  Madagascar. 

Reinthal,  P.  N.  and  M.  L.  J.  Stiassny.  1991.  The 
freshwater  fishes  of  Madagascar:  a study  of  an 
endangered  fauna  with  recommendations  for  a 
conservation  strategy.  Conserv.  Biol.  5:231-243. 

Reinthal,  P.  N.,  J.  Sparks,  and  K.  Riseng.  1995.  Ecol- 


ogy, vulnerability,  and  temporal  changes  in  fresh- 
water fish  communities.  P.  20  in  Environmental 
change  in  Madagascar  (B.  D.  Patterson,  S.  M. 
Goodman,  and  J.  L.  Sedlock,  Eds.).  Eield  Museum 
of  Natural  History,  Chicago,  Illinois. 

Sabine,  N.  and  W.  D.  Klim.stra.  1985.  Ecology  of 
Bald  Eagles  wintering  in  southern  Illinois.  Trans. 
111.  State  Acad.  Sci.  78:13-24. 

SAS  Institute  Inc.  1990.  SAS  user’s  guide:  statistics, 
version  6,  fourth  ed.  SAS  Institute  Inc.,  Cary, 
North  Carolina. 

Scott,  W.  B.  and  E.  J.  Crossman.  1973.  Ereshwater 
fishes  of  Canada.  Eish.  Res.  Board  Can.  Bull.  184: 
1-966. 

Stalmaster,  M.  V.  1987.  The  Bald  Eagle.  Universe 
Books,  New  York. 

Stalmaster,  M.  V.  and  R.  G.  Plettner.  1992.  Diets 
and  foraging  effectiveness  of  Bald  Eagles  during 
extreme  winter  weather  in  Nebraska.  J.  Wildl. 
Manage.  56:355-367. 

Steenhof,  K.  1976.  The  ecology  of  wintering  Bald 
Eagles  in  southeastern  South  Dakota.  M.Sc.  the- 
sis, Univ.  of  Missouri,  Columbia. 

Steenhof,  K.,  S.  S.  Berlinger,  and  L.  H.  Eredrick- 
SON.  1980.  Habitat  use  by  wintering  Bald  Eagles 
in  South  Dakota.  J.  Wildl.  Manage.  44:798-805. 

Swenson,  J.  E.,  K.  L.  Alt,  and  R.  L.  Eng.  1986.  Ecol- 
ogy of  Bald  Eagles  in  the  greater  Yellowstone 
ecosystem.  Wildl.  Monogr.  95:1-46. 

Whitfield,  A.  K.  and  S.  J.  M.  Blaber.  1978.  Eeeding 
ecology  of  piscivorous  birds  at  Lake  St.  Lucia: 
part  I,  diving  birds.  Ostrich  49:185-198. 


Wilson  Bull,  111(1),  1999,  pp.  22-29 


THE  RELATIONSHIP  BETWEEN  SPOTTED  OWL  DIET  AND 
REPRODUCTIVE  SUCCESS  IN  THE  SAN  BERNARDINO 
MOUNTAINS,  CALIFORNIA 

RICHARD  B.  SMITH,'  - M.  ZACHARIAH  PEERY,'  ^ ^ r.  j.  GUTIERREZ,'  AND 

WILLIAM  S.  L AH  AYE'-* 


ABSTRACT. — We  analyzed  the  breeding  season  diets  of  California  Spotted  Owls  {Strix  occidentalis  occi- 
dentalis)  in  the  San  Bernardino  Mountains  from  1987  through  1991  to  estimate  the  relative  importance  of 
individual  prey  species  to  owl  reproduction.  We  identified  a total  of  8441  prey  remains  from  109  unique  terri- 
tories, which  represents  the  largest  collection  of  prey  remains  from  a single  Spotted  Owl  population.  Dusky- 
footed  woodrats  (Neotoma  fuscipes)  and  Jerusalem  crickets  {Stenopelmatus  fuscus)  were  the  most  frequently 
consumed  taxa  (42.2%  and  20.7%,  respectively),  but  dusky-footed  woodrats  dominated  Spotted  Owl  diets  by 
biomass  (74.0%).  Spotted  owls  consumed  primarily  mammals  by  both  frequency  (66.4%)  and  biomass  (95.3%). 
After  excluding  territories  with  less  than  20  prey  remains,  we  compared  the  diets  of  24  nonnesting,  24  unsuc- 
cessfully nesting,  and  58  successfully  nesting  pairs  of  Spotted  Owls  from  56  unique  territories;  estimated  diet 
along  a large  elevational  gradient;  and  controlled  for  interterritorial  and  annual  variation  in  diet.  A significant 
relationship  existed  between  reproductive  status  and  the  percent  biomass  of  woodrats  in  Spotted  Owl  diets  where 
successful  nesters  consumed  a greater  percent  biomass  of  woodrats  (x  = 81.8)  than  nonnesters  (.v  = 74.1)  but 
not  unsuccessful  nesters  (x  = 75.5).  Unsuccessful  nesters  and  nonnesters  did  not  consume  a significantly  different 
percent  biomass  of  woodrats.  The  percentage  of  woodrat  biomass  in  Spotted  Owl  diets  increased  with  elevation 
but  did  not  differ  among  territories  or  years.  We  hypothesized  that  breeding  Spotted  Owls  were  able  to  meet 
the  increased  energetic  demands  associated  with  producing  young  by  consuming  primarily  large,  energetically 
profitable  prey  such  as  woodrats.  Received  6 May  1998,  accepted  21  Oct.  1998. 


The  Spotted  Owl  (Strix  occidentalis)  preys 
on  a wide  range  of  vertebrate  and  invertebrate 
taxa,  but  primarily  on  a few  species  of  small 
mammals  (Gutierrez  et  al.  1995).  The  distri- 
bution of  these  small  mammals  has  an  impor- 
tant influence  on  the  owl’s  home  range  size 
(Carey  et  al.  1992,  Zabel  et  al.  1995),  habitat 
use  patterns  (Carey  et  al.  1992,  Carey  and 
Peeler  1995,  Zabel  et  al.  1995,  Ward  et  al. 
1998),  and  demographic  rates  (Franklin  1997, 
Ward  et  al.  1998).  In  particular,  prey  abun- 
dance positively  influences  Spotted  Owl  re- 
production. Ward  and  coworkers  (1998)  found 
that  dusky-footed  woodrats  (Neotoma  fusci- 
pes) were  more  abundant  in  the  territories  of 
breeding  Northern  Spotted  Owls  (S.  o.  cauri- 
na)  than  in  the  territories  of  nonbreeding  owls. 
Although  this  difference  was  not  statistically 


' Dept,  of  Wildlife,  Humboldt  State  Univ.,  Areata, 
CA  95521. 

2 Present  Address;  P.O.  Box  266,  Mattituck,  NY 
I 1952. 

’ Present  Address:  927  Lincoln  Way,  San  Francisco, 
CA  94122. 

^ Present  Address;  PO.  Box  523,  Big  Bear  City,  CA 
92314. 

'Corresponding  author;  E-mail: 
mz.pl  @axc. humboldt.edu 


significant,  the  authors  suggested  that  high 
variation  in  woodrat  abundances  resulted  in 
low  statistical  power  for  rejecting  the  hypoth- 
esis of  no  difference  between  breeders  and 
nonbreeders.  Franklin  (1997)  showed  that  the 
distribution  of  woodrat  habitat  explained  a 
large  amount  of  interterritorial  variation  in 
Northern  Spotted  Owl  reproductive  success. 
In  addition,  studies  of  other  strigids  demon- 
strate convincingly  that  reproduction  for  most 
owl  species  is  limited,  at  least  in  part,  by  prey 
availability  (see  Verner  et  al.  1992  for  a re- 
view). 

Assuming  that  Spotted  Owl  reproductive 
success  is  determined  in  part  by  food  avail- 
ability, food  habit  studies  based  on  prey  re- 
mains from  egested  pellets  can  be  used  to  ex- 
amine the  relative  importance  of  individual 
prey  species  for  reproduction.  Such  studies 
have  shown  that  successful  breeders  consume 
a greater  proportion  of  large  prey  than  unsuc- 
cessful breeders  (Barrows  1985,  1987; 
Thradkill  and  Bias  1989;  White  1996),  al- 
though Ward  and  coworkers  (1998)  found  no 
difference.  These  studies  have  been  based 
upon  relatively  few  prey  items  or  owl  terri- 
tories which  has  resulted  in  (1)  data  being 
pooled  among  territories  or  (2)  data  from  a 


22 


Smith  et  ill.  • SPOTTED  OWL  DIET  AND  REPRODUCTION 


23 


particular  territory  being  considered  indepen- 
dent from  year  to  year.  The  first  approach 
weights  all  territories  equally  regardless  of  the 
number  of  prey  items  collected.  For  example, 
a breeding  territory  from  which  one  prey  item 
was  collected  is  given  the  same  weight  as  a 
breeding  territory  with  hundreds  of  prey 
items.  The  second  approach  results  in  pseu- 
doreplication and  can  dramatically  inflate  the 
probability  of  rejecting  the  null  hypothesis 
(Hurlbert  1984)  that  breeding  and  nonbreed- 
ing Spotted  Owls  have  the  same  diets.  This  is 
especially  likely  because  Spotted  Owl  diets 
exhibit  interterritorial  variation  (Laymon 
1988). 

Because  previous  diet  studies  were  based 
on  relatively  few  territories  sampled  from 
large,  open  populations,  inferences  about  the 
relationship  between  diet  patterns  and  the 
breeding  ecology  of  the  Spotted  Owl  are  lim- 
ited. We  have  been  studying  the  entire  popu- 
lation of  California  Spotted  Owls  (S.  o.  occi- 
dentalis)  in  the  San  Bernardino  Mountains  of 
southern  California  since  1987.  It  is  the  larg- 
est subpopulation  within  the  southern  Califor- 
nia owl  metapopulation  (Noon  and  McKelvey 
1992,  LaHaye  et  al.  1994)  and  occupies  a di- 
verse array  of  habitats  along  a large  elevation 
gradient.  This  allowed  us  to  estimate  diet  over 
a range  of  ecological  conditions  and  reliably 
evaluate  its  relationship  with  reproduction.  In 
this  paper  we  compare  the  diets  of  nonnesting, 
unsuccessfully  nesting,  and  successfully  nest- 
ing Spotted  Owls  throughout  the  breeding  sea- 
son (March  through  October),  enumerate  the 
food  habits  of  individuals  of  this  population, 
and  compare  our  results  to  other  food  habit 
studies  of  the  California  Spotted  Owl. 

STUDY  AREA  AND  METHODS 

The  San  Bernardino  Mountains  Study  Area  was  lo- 
cated approximately  140  km  east  of  Los  Angeles,  Cal- 
ifornia and  encompassed  1890  km^  with  elevations 
ranging  from  800-3500  m.  Mean  annual  precipitation 
ranged  from  less  than  20  cm  to  more  than  100  cm  and 
was  strongly  influenced  by  elevation  and  topography 
(Minnich  1988).  The  vegetation  was  diverse,  ranging 
from  Mojave  Desert  scrub  (Vasek  and  Barbour  1988) 
at  lower  elevations  to  alpine  (Major  and  Taylor  1988) 
on  San  Gorgonio  Mountain.  Most  Spotted  Owls  oc- 
cupied conifer  dominated  forest  (Sawyer  et  al.  1988, 
Thome  1988)  between  1000  m and  2500  m elevation. 

Owl  survey  methods. — We  surveyed  the  study  area 
for  Spotted  Owls  following  methods  described  by 
Franklin  and  coworkers  (1996)  during  the  breeding 


season  (March-October),  1987-1991.  We  conducted 
1113  nocturnal  surveys  during  which  we  spent  15  min 
at  an  individual  point  or  walked  along  designated 
routes,  using  vocal  imitations  of  Spotted  Owl  calls  to 
survey  the  forested  habitat  within  the  study  area.  Sur- 
vey points  and  routes  were  placed  so  that  all  of  this 
potentially  suitable  habitat  was  surveyed  each  year.  We 
conducted  1659  diurnal  surveys  at  territories  that  were 
known  to  be  occupied  in  order  to  locate  nests,  locate 
roosts,  collect  regurgitated  pellets,  and  assess  repro- 
ductive status.  We  conducted  initial  diurnal  surveys  for 
each  territory  in  March  or  April  and  conducted  follow- 
up diurnal  surveys  every  3 to  5 weeks.  We  collected 
pellets  by  thoroughly  searching  areas  underneath  Spot- 
ted Owl  nests  and  roosts  during  most  diurnal  surveys. 
We  assessed  nesting  status  by  feeding  owls  live  mice 
{Mus  musculus)  during  diurnal  surveys  (Franklin  et  al. 
1996).  We  considered  an  owl  pair  not  to  have  nested 
if  one  member  of  the  pair  ate  or  cached  four  consec- 
utive mice  during  a single  diurnal  survey  prior  to  31 
May.  Although  only  one  formal  survey  was  conducted 
to  assess  nesting  status,  multiple  surveys  were  con- 
ducted at  each  territory  to  band  and  resight  owls.  If  a 
nest  was  located  during  one  of  these  surveys,  the  owls 
at  that  particular  territory  were  of  course  considered  to 
be  nesting.  When  a nest  was  located,  we  used  the  same 
method  to  locate  fledglings  or  to  determine  nest  fail- 
ure. Owls  that  did  not  take  sufficient  mice  were  not 
included  in  the  study. 

Quantifying  Spotted  Owl  diets. — We  dissected  all 
collected  pellets  and  isolated  all  identifiable  prey  re- 
mains. Identifiable  remains  included  skulls  (birds  and 
mammals),  mandibles  (mammals,  reptiles,  and  inver- 
tebrates), legs  (birds  and  invertebrates),  claws  (inver- 
tebrates), and  bills  (birds).  Remains  collected  during 
each  diurnal  survey  were  enumerated  to  the  lowest 
possible  taxonomic  level,  and  the  highest  count  was 
taken  as  the  total  number  of  prey  items  for  that  survey. 
If,  for  example,  pellets  collected  during  a single  survey 
contained  one  woodrat  skull,  two  left  mandibles,  and 
three  right  mandibles,  three  woodrats  were  considered 
to  be  present.  The  percent  frequency  of  each  taxa 
(Marti  1987)  was  then  calculated  for  each  territory 
pooled  among  years.  The  percent  biomass  (Marti 
1987)  was  calculated  for  each  territory  using  the  mean 
prey  weights  in  Table  1 . The  mean  percent  frequency 
and  biomass  among  territories  were  used  as  estimates 
of  the  overall  diet  composition.  This  approach  weights 
all  territories  equally  so  that  overall  percentages  were 
not  biased  towards  territories  with  many  prey  remains 
(Swanson  et  al.  1974). 

Statistical  analyses. — Statistical  analyses  were  per- 
formed only  on  the  biomass  data  1988-1991  because 
of  small  sample  sizes  in  1987  and  because  we  felt 
biomass  more  accurately  represented  the  energetic  im- 
portance of  each  prey  taxa. 

We  used  a mixed-model  ANOVA  approach  (Proc 
MIXED  of  pc  SAS,  version  6.12;  Littel  et  al.  1996)  to 
estimate  variation  in  Spotted  Owl  diet,  where  each  diet 
variable  was  considered  as  the  dependent  variable  in  a 
separate  model.  We  treated  reproductive  status  (non- 


24 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


TABLE  1. 

Diet  of  Spotted  Owls 

in  the  San  Bernardino  Mountains,  California. 

Mean  prey 
weight  (g) 

Source^ 

% of  prey 
by  frequency 
(n  = 8,169) 

% of  prey 
by  biomass 
(707,193  g) 

Mammals 

Neotonia  fu.scipe.s 

173.7 

1 

42.2 

74.0 

Thomoniy.'i  hottae 

124.4 

1 

7.4 

10.4 

GUiiicomy.s  sahrinus 

121.5 

2 

2.1 

3.0 

Peromy.'icu.s  spp. 

22.1 

1 

11.3 

4.0 

Microtus  spp. 

60.2 

1,2 

1.8 

1.3 

Sylvilagus  spp. 

538.3 

2 

0.2 

1.0 

Sciurus  griseii.s 

785.0 

2 

0,1 

1.0 

D ipodomy.s  me  ni  a ini 

42.1 

2 

0.1 

<0.1 

Scapaiui.s  latimanus 

55.2 

2 

0.4 

0.3 

Spennophilu.s  lateralis 

153.4 

1 

0.1 

0.2 

Tamius  merriami 

67.6 

1,2 

0.1 

0.1 

Antrozous  pallidiis 

21.5 

2 

<0.1 

<0.1 

Eptesiciis  fuscus 

14.0 

2 

<0.1 

<0.1 

Lasiiirius  cinereus 

25.5 

2 

0.1 

<0.1 

My  Otis  spp. 

4.9 

2 

0.1 

<0.1 

Perognathus  spp. 

14.4 

1 

0.1 

<0.1 

Sore.x  ornatus 

5.1 

2 

0.5 

<0.1 

Subtotal 

66.4 

95.3 

Birds 

64.4 

3 

4.3 

3.5 

Invertebrates 

Stenopelnuit us  fuscus 

2.0 

4 

20.7 

0.9 

Prionus  californicus 

2.0 

3 

3.6 

0.1 

Scorpionida 

2.0 

4 

2.7 

0.1 

Unidentified 

2.0 

4 

2.1 

0.1 

Subtotal 

29.1 

1.2 

Reptiles 

Sceloporus  spp. 

10.0 

3 

0.1 

<0.1 

■'  Sources  were  ( I ) specimens  from  the  San  Bernardino  Mountains  in  the  Museum  of  Vertebrate  Zoology  at  University  of  California,  Berkeley  collections 
( specimens  from  nearest  available  location  to  the  San  Bernardino  Mountains  in  the  Museum  of  Vertebrate  Zoology  at  University  of  California.  Berkeley 
collections.  (3)  Foreman  et  al.  ( 1984;  biomass  of  birds  divided  by  the  number  of  birds  in  tables  12-14),  (4)  estimated  based  on  prey  size. 


nesters,  unsuccessful  nesters,  and  successful  nesters) 
and  year  as  fixed  effects,  and  elevation  (at  the  center 
of  activity  for  each  territory)  as  a covariate.  We  u.sed 
territory  as  a random  blocking  factor  to  estimate  var- 
iation in  diet  among  territories  and  control  for  non- 
independence of  samples  within  territories.  Two-way 
interactions  between  reproductive  status  and  the  other 
main  effects  were  included  in  the  model.  Main  effect 
by  territory  interactions  were  assumed  to  be  nonexis- 
tent which  resulted  in  all  effects  being  tested  over  the 
residual  mean  square  (Newman  et  al.  1997).  Signifi- 
cant tixed  effects  were  further  tested  with  /-tests  on 
least  squares  means  using  sequential  Bonferonni  ad- 
justments on  critical  values  (Rice  1989).  This  proce- 
dure involves  testing  each  comparison,  starting  with 
the  most  significant  and  stopping  at  the  first  nonsig- 
nificant comparison,  using  a/{  \ + k - i)  as  the  critical 
value  where  k is  the  number  of  post-hoc  compari.sons 
and  /'  is  the  number  of  the  comparison. 

Proper  use  of  an  ANOVA  model  requires  a normally 
distributed  dependent  variable  and  equality  of  varianc- 
es among  treatment  levels  (Zar  1984).  Because  the  per- 
cent biomass  of  woodrats  was  a proportional  variable. 


and  hence  formed  a binomial  distribution,  we  used  the 
square  root-arcsine  transformation  to  normalize  the 
data  (Zar  1984).  Normality  was  then  tested  with  the 
Shapiro- Wilk  statistic  (Zar  1984).  Equality  of  variances 
among  reproductive  groups  was  tested  with  an  F-test 
(Zar  1984). 

By  using  diet  as  the  dependent  variable  we  do  not 
imply  that  diet  is  a function  of  reproductive  status;  the 
mixed  modeling  approach  simply  provides  a conve- 
nient way  of  examining  variation  in  diet  in  relation  to 
other  variables  (including  reproduction).  We  believe 
that  this  approach  is  justified  because  ( 1 ) the  goal  of 
the  study  was  to  determine  if  diet  differed  among  non- 
nesting, unsuccessfully  ne.sting,  and  successfully  nest- 
ing Spotted  Owls,  and  (2)  it  is  appropriate  to  analyze 
a correlative  relationship  with  a linear  model  because 
the  resultant  F-statistic  provides  a test  of  the  null  hy- 
pothesis that  the  multiple  correlation  coefficient  R is 
zero  (Zar  1984). 

RESULTS 

Spotted  Owl  diets.~We  iijentified  a total  of 
8441  prey  items  from  109  unique  Spotted  Owl 


Smilh  el  al.  • SPOTTED  OWL  DIET  AND  REPRODUCTION 


25 


TABLE  2.  The  effect  of 

reproductive  status,  elevation,  and  year  on  the  percent  biomass 

of  woodrats  in 

Spotted  Owl  diets  in  the  San 

Bernardino 

Mountains. 

Results  are  from 

a mixed-model  ANOVA  (Littel  ct  al. 

1996)  where  reproductive  status  and  year 

were  fixed 

effects,  elevation 

was  a covariatc,  and 

territory^'  was  a 

random  blocking  factor  effect. 

Effecl 

ncir 

(Jdf 

p 

Reproductive  status 

2 

36 

3.65 

0.04 

Elevation 

1 

36 

34.93 

<0.01 

Year 

3 

36 

1.87 

0.15 

Reproductive  status*year 

3 

36 

1.17 

0.35 

Reproductive  status*elevation 

1 

36 

3.16 

0.06 

“Territory  was  not  significant  (Z  = 1.43,  P = 0.15). 


territories.  This  represents,  to  the  best  of  our 
knowledge,  the  largest  collection  of  prey 
items  recorded  from  a single  Spotted  Owl 
population.  When  estimating  the  overall  diet 
composition,  we  excluded  territories  from 
which  fewer  than  20  prey  remains  were  col- 
lected during  the  entire  study  period  (this  re- 
sulted in  a subsample  of  8,169  prey  remains 
from  71  territories).  Dusky-footed  woodrats 
and  Jerusalem  crickets  (Stenopelmatus  fuscus) 
were  the  most  common  taxa  by  frequency 
(42.2%  and  20.7%,  respectively).  White-foot- 
ed mice  (Peromyscus  spp.)  and  northern  pock- 
et gophers  (Thomomys  bottae)  were  less  com- 
mon by  frequency  (11.3%  and  7.4%,  respec- 
tively). No  other  taxa  contributed  more  than 
4.3%  to  the  total  number  of  prey  items  (Table 
1). 

Dusky-footed  woodrats  dominated  Spotted 
Owl  diet  by  biomass  (74.0%).  followed  by 
pocket  gophers  (10.4%).  Dusky-footed  wood- 
rats were  the  largest  (173.7  g)  of  the  common 
prey  items.  Western  grey  squirrels  (Sciurus 
griseus)  and  cottontails  (Sylvilagus  spp.)  were 
larger,  but  represented  only  2.0%  of  the  total 
biomass.  No  other  taxa  contributed  more  than 
4.0%  to  the  total  biomass  consumed  (Table  1). 
Mammals  contributed  66.4%  and  95.3%  to  the 
total  number  of  prey  items  and  the  total  bio- 
mass, respectively.  Invertebrates  contributed 
29.1%  to  the  total  number  of  prey  items,  but 
only  1.2%  to  the  total  biomass. 

Variation  in  Spotted  Owl  diets. — For  statis- 
tical analyses,  we  excluded  territories  repre- 
sented by  fewer  than  20  prey  items  in  any  one 
year.  In  doing  so,  we  retained  a large  sample 
size  (106  samples  from  56  unique  Spotted 
Owl  territories;  24  nonnesters,  24  unsuccess- 
ful nesters,  and  58  successful  nesters)  and 
were  able  to  estimate  the  percent  biomass  of 


woodrats  with  reasonable  precision  on  a ter- 
ritory by  territory  basis  (mean  CV  = 0.08, 
maximum  CV  = 0.27).  The  percent  biomass 
of  pocket  gophers  was  estimated  with  consid- 
erably less  precision  (mean  CV  = 0.50,  max- 
imum CV  = 1.00).  In  addition,  the  percent 
biomass  of  pocket  gophers  was  highly  and 
negatively  correlated  with  the  percent  biomass 
of  woodrats  (r  = —0.67,  P < 0.01,  n = 106). 
For  these  reasons,  we  did  not  model  the  per- 
cent biomass  of  pocket  gophers  statistically. 

Although  the  percent  biomass  of  woodrats 
was  not  distributed  normally  (W  = 0.93,  P < 
0.01)  the  variance  in  the  percent  biomass  of 
woodrats  did  not  differ  between  the  least  var- 
iable (successful  nesters)  and  the  most  vari- 
able (unsuccessful  nesters)  reproductive 
groups  (7^23  57  = 1.46,  P > 0.05).  Because  AN- 
OVA  is  sensitive  to  heterogeneity  of  variances 
among  treatment  levels  but  robust  to  depar- 
tures from  normality  (Hicks  1993),  we  as- 
sumed the  data  met  the  assumptions  of  the 
model. 

A significant  relationship  existed  between 
reproductive  status  and  the  percent  biomass  of 
woodrats  in  Spotted  Owl  diets  (Table  2).  Suc- 
cessful nesters  consumed  a greater  percent 
biomass  of  woodrats  (x  = 81.8  ± 1.5  SE)  than 
nonnesters  (x  = 74.1  ± 2.4;  ^0017.36  ^ 2.49,  P 
= 0.017)  but  not  unsuccessful  nesters  (.v  = 
75.5  ± 2.4;  02.3,36  = 2.08,  P - 0.044),  al- 

though the  difference  between  successful  and 
unsuccessful  nesters  was  nearly  significant. 
Unsuccessful  nesters  and  nonnesters  did  not 
consume  a different  percent  biomass  of  wood- 
rats (t()  050.36  “ 0.04,  P > 0.05).  The  percentage 
of  woodrat  biomass  in  Spotted  Owl  diets  did 
not  differ  among  territories  but  increased  with 
elevation  (Table  2;  Fig.  1 ).  The  interaction  be- 
tween elevation  and  reproductive  status  was 


20  -I ^ 1— 

1000  1200  1400 


1600  1800  2000 
ELEVATION  (m) 


2200  2400  2600 


FIG.  1 . The  relationship  between  elevation  and  the  percent  biomass  of  woodrats  in  the  diets  of  Spotted  Owls 
in  the  San  Bernardino  Mountains,  California,  1988-1991  (h  = 0.000085,  R~  = 0.26,  df  = 104,  F < 0.01). 


not  significant  (Table  2),  indicating  that  the 
percent  biomass  of  woodrats  increased  with 
elevation  at  a constant  rate  for  nonnesters,  un- 
successful nesters,  and  successful  nesters.  The 
percentage  of  woodrat  biomass  in  the  owl’s 
diet  did  not  differ  among  years  and  no  inter- 
action existed  between  year  and  reproductive 
status  (Table  2). 

DISCUSSION 

For  an  owl  pair  to  produce  young,  females 
must  acquire  sufficient  fat  reserves  prior  to 
nesting,  and  males  need  to  provide  sufficient 
amounts  of  food  to  the  female  and  nestlings 
during  the  nesting  period  (Hirons  1985). 
Dusky-footed  woodrats  are  relatively  large 
and  may  have  provided  an  energetically  prof- 
itable food  source  that  enabled  Spotted  Owls 
to  reproduce  successfully.  This  idea  is  sup- 
ported by  Ward  and  coworkers  (1998)  who 
showed  that  selection  of  dusky-footed  wood- 
rats provided  an  indirect  benefit  to  Spotted 
Owl  fitness  by  reducing  the  amount  of  habitat 
needed  to  reproduce  successfully.  Further,  Or- 
ians  and  Pearson  (1979)  predict  that  central 


place  foragers  such  as  Spotted  Owls  (Carey 
and  Peeler  1995)  can  increase  the  rate  of  en- 
ergy return  to  the  central  place  (nest),  and 
hence  fitness,  by  consuming  large  prey.  By 
consuming  large  prey,  the  male  can  minimize 
the  number  of  flights  from  the  point  of  capture 
to  the  nest,  allowing  more  time  for  hunting. 

Since  the  diet  of  unsuccessful  nesters  was 
more  similar  to  nonnesters  than  successful 
nesters,  consuming  woodrats  may  be  more  im- 
portant for  incubation  and  brooding  than  for 
nest  initiation.  This  is  supported  by  the  fact 
that  the  energetic  cost  of  egg  production  is 
relatively  small  compared  to  the  amount  of 
energy  needed  to  provision  the  female  and 
young  during  the  nesting  period  (Ward  et  al 
1998). 

Our  results  support  Barrows  (1985,  1987), 
Thrailkill  and  Bias  (1989),  and  White  (1996) 
who  found  that  Spotted  Owls  that  fledged 
young  consumed  a higher  proportion  of  large 
prey  than  Spotted  Owls  that  did  not  fledge 
young.  Despite  the  sample  size  limitations  of 
these  previous  studies,  it  appears  that  a true 
difference  in  diet,  particularly  in  terms  of  prey 


Smith  et  al.  • SPOTTED  OWL  DIET  AND  REPRODUCTION 


27 


size,  exists  between  breeding  and  nonbreeding 
Spotted  Owls.  The  importance  of  prey  size 
has  also  been  observed  for  Black-shouldered 
Kite  (Elanus  caeruleus;  Slotow  and  Perrin 
1992)  and  Common  Kestrel  {Falco  tinnuncu- 
lus;  Korpimaki  1986)  reproduction,  suggest- 
ing that  this  may  be  a common  pattern  for 
raptorial  species. 

It  is  possible  that  the  difference  between 
successful  nesters  and  nonnesters  was  statis- 
tically significant,  but  not  biologically  signif- 
icant. An  8%  difference  in  the  percent  bio- 
mass of  woodrats  is  relatively  small  when 
compared  to  the  additional  amount  of  food  a 
male  must  procure  to  provision  a female  and 
even  just  one  nestling.  Based  upon  Ward’s  and 
coworkers’  (1998)  energetic  calculations,  a 
male  must  increase  the  amount  of  food  he  pro- 
cures by  276%  in  order  to  provision  the  fe- 
male and  a Juvenile  from  egglaying  to  fledg- 
ing. However,  the  dietary  difference  observed 
between  breeders  and  nonbreeders  may  be 
only  one  of  the  factors  that  allows  an  owl  pair 
to  produce  young.  For  example,  the  variation 
in  diet  caused  by  breeding  status  may  simply 
reflect  higher  prey  availabilities  in  the  terri- 
tories of  breeding  and  nonbreeding  owls  (see 
below). 

The  percent  biomasses  of  the  prey  species 
presented  here  are  not  necessarily  estimates  of 
the  total  amount  of  prey  taken  because  it  is 
unlikely  that  we  collected  all  regurgitated  pel- 
lets. Hence  we  cannot  evaluate  the  effect  of 
food  supply  on  reproduction.  It  is  possible  that 
even  though  successfully  nesting  owls  con- 
sumed a greater  percent  biomass  of  woodrats, 
they  obtained  less  total  biomass  from  wood- 
rats than  nonnesters  or  unsuccessful  nesters. 
However,  this  seems  unlikely  because  breed- 
ing Spotted  Owls  need  to  take  considerably 
more  total  prey  to  feed  their  juveniles. 

Without  measuring  prey  abundances,  we 
cannot  be  certain  if  the  relationship  between 
Spotted  Owl  diet  and  reproduction  was  the  re- 
sult of  differences  in  prey  availability  or  prey 
selection  among  territories.  Spotted  Owls  se- 
lect dusky-footed  woodrats  more  than  would 
be  expected  based  on  availability  in  both 
northern  and  southern  California  (Hedlund 
1996,  Ward  et  al.  1998).  Further,  Barrows 
(1987)  found  that  Spotted  Owl  pairs  shift  from 
large  to  small  prey  after  nest  failure  and  sug- 
gested that  this  change  in  prey  selection  was 


in  response  to  reduced  energy  requirements.  If 
Barrow’s  hypothesis  was  true  for  our  popula- 
tion, unsuccessful  nesters  should  have  had  an 
intermediate  percentage  of  woodrat  biomass 
in  their  diets.  Since  this  did  not  occur,  we  con- 
sider it  more  likely  that  the  difference  in  the 
percent  biomass  of  woodrats  between  owls 
that  fledged  young  and  owls  that  did  not 
fledge  young  was  the  result  of  differences  in 
prey  availability  among  territories.  Optimal 
foraging  theory  (see  Pyke  et  al.  1977  for  a 
review)  predicts,  and  raptor  field  studies  show 
(Korpimaki  1986,  Steenhof  and  Kochert 
1988),  that  when  the  density  of  primary  prey 
is  high  within  the  landscape,  the  percentage  of 
that  prey  in  the  predator’s  diet  is  also  high 
(i.e.,  a functional  response).  Hence,  if  wood- 
rats occurred  at  higher  densities  in  the  terri- 
tories of  successfully  reproducing  owls 
(Franklin  1997,  Ward  et  al.  1998)  one  would 
expect  these  owls  to  consume  a greater  per- 
centage of  woodrats. 

Bull  and  coworkers  (1989)  found  that  male 
Great  Gray  Owls  (S.  nebulosa)  preferentially 
consumed  small  prey  items  at  the  point  of 
capture  and  brought  large  prey  items  back  to 
the  nest.  Although  this  presents  a potential 
source  of  bias,  we  do  not  believe  that  it  affects 
the  conclusions  of  this  study.  For  nesting 
owls,  our  pellets  were  probably  biased  to- 
wards males  because  we  (1)  often  located 
male  roosts  and  (2)  females  frequently  flew 
away  from  the  nest  to  egest  pellets  (personal 
observation).  If  Bull’s  and  coworkers’  (1989) 
findings  are  true  for  Spotted  Owls,  we  may 
have  underestimated  the  percent  biomass  of 
woodrats  in  the  diets  of  nesting  owls.  This 
bias,  however,  would  have  decreased  the  dif- 
ference between  owls  that  fledged  young  and 
owls  that  did  not  fledge  young. 

Another  potential  bias  is  that  pellets  may 
not  represent  a random  sample  of  the  owl’s 
diet.  In  particular,  single  prey-item  Bam  Owl 
{Tyto  alba)  pellets  are  more  likely  to  contain 
large  prey  than  expected  by  chance  (Yom-Tov 
and  Wool  1997).  If  this  was  tme  for  our  Spot- 
ted Owl  population,  the  percent  frequency  and 
biomass  of  relatively  large  prey  such  as  wood- 
rats, pocket  gophers,  and  northern  flying 
squirrels  {Glaucomys  sabrinus)  would  be  pos- 
itively biased,  while  the  percent  frequency  and 
biomass  of  relatively  small  prey  such  as 
white-footed  mice  would  be  negatively  bi- 


28 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


ased.  However,  these  biases  should  be  consis- 
tent for  nonnesting,  unsuccessfully  nesting, 
and  successfully  nesting  owls  and  should  not 
affect  the  relationship  between  reproductive 
status  and  diet  that  we  observed. 

Although  some  similarities  existed  between 
Spotted  Owl  diets  in  the  San  Bernardino 
Mountains  and  other  localities  in  southern 
California,  some  marked  differences  existed 
as  well.  For  example,  while  the  percent  fre- 
quencies of  woodrats  and  pocket  gophers  in 
Spotted  Owl  diets  reported  in  this  study  were 
similar  to  those  reported  by  Barrows  (1980, 
1987),  white-footed  mice  comprised  a consid- 
erably smaller  percentage  (11.3%  versus 
40.0%)  and  invertebrates  a greater  percentage 
(29.1%  versus  18.0%)  of  the  total  prey  items 
in  this  study.  The  percent  biomasses  of  wood- 
rats,  pocket  gophers,  and  most  other  prey  taxa 
were  remarkably  similar  to  those  reported  by 
Barrows  (1980,  1987). 

In  the  Sierra  Nevada,  woodrats  are  the 
Spotted  Owl’s  primary  prey  at  low  elevations 
but  are  almost  completely  replaced  by  flying 
squirrels  at  high  elevations  (Verner  et  al. 
1992).  In  the  San  Bernardino  Mountains, 
woodrats  increased  in  Spotted  Owl  diets  with 
elevation  and  flying  squirrels  were  uncommon 
(2.1%).  Flying  squirrels  are  probably  more 
abundant  in  the  upper  mixed-conifer  and  red 
fir  {Abies  magnified)  zones  of  the  Sierra  Ne- 
vada (see  Waters  and  Zabel  1995)  than  they 
are  in  the  San  Bernardino  Mountains  where 
they  reach  the  southern  edge  of  their  range 
(Hall  and  Kelson  1959).  The  San  Bernardino 
Mountains  apparently  lack  a relatively  large 
alternative  prey  species  at  higher  elevations 
and  it  is  possible  that  most  of  the  other  prey 
species  decrease  in  elevation  as  well.  Wood- 
rats are  an  important  part  of  the  Spotted  Owl’s 
diet  in  southern  California,  both  in  terms  of 
biomass  and  reproduction,  and  we  recommend 
that  future  management  of  forested  habitat  in 
this  region  promote  high  woodrat  densities. 

ACKNOWLEDGMENTS 

Wc  thank  M.  Seamans,  P.  Carlson.  A.  Franklin,  P. 
Ward,  E.  Forsman,  and  an  anonymous  reviewer  for 
many  u.seful  comments  on  earlier  drafts.  We  thank  the 
many  dedicated  people  who  provided  field  assistance 
including  A.  Kirn,  D.  Call.  D.  Cousins,  D.  Roberts,  E. 
Pausch.  G.  Caulkins.  J.  Gronski,  J.  Stephenson,  J. 
Schmid.  K.  Norgaard.  M.  Hollister,  R.  Long,  R.  Tan- 
ner. S.  Bailey,  S.  Stroich.  and  K.  Dimcheff.  We  thank 


B.  Stein  of  the  Museum  of  Vertebrate  Zoology  at  the 
University  of  California,  Berkeley  for  access  to  mam- 
mal collections.  Funding  was  provided  by  the  San  Ber- 
nardino National  Forest,  Region  5 USDA-Forest  Ser- 
vice, Southern  California  Edison,  California  Depart- 
ment of  Fish  and  Game,  Snow  Summit  Ski  Coopera- 
tion, San  Bernardino  County  Audubon  Society,  the 
California  State  Resource  Agency,  Pacific  Southwest 
Forest  and  Range  Experiment  Station,  Bear  Mountain, 
Ltd.,  the  Army  Corps  of  Engineers,  and  Salad  King 
Inc. 

LITERATURE  CITED 

Barrows,  C.  W.  1980.  Feeding  ecology  of  the  Spotted 
Owl  in  California.  Raptor  Res.  14:73-78. 
Barrows,  C.  W.  1985.  Breeding  success  relative  to 
fluctuations  in  diet  for  Spotted  Owls  in  California. 
U.S.D.A.  Pacific  Northwest  Forest  and  Range  Ex- 
periment Station.  Gen.  Tech.  Rept.  PNW-GTR- 
185:50-54 

Barrows,  C.  W.  1987.  Diet  shifts  in  breeding  and  non- 
breeding Spotted  Owls.  Raptor  Res.  21:95-97. 
Bull,  E.  L.,  M.  G.  Henjum,  and  R.  S.  Rohweder. 
1989.  Diet  and  optimal  foraging  of  Great  Grey 
Owls.  J.  Wildl.  Manage.  53:47—50. 

Carey,  A.  B.,  S.  P.  Horton,  and  B.  L.  Biswell.  1992. 
Northern  Spotted  Owls:  influence  of  prey  base  and 
landscape  character.  Ecol.  Monogr.  62:223-250. 
Carey,  A.  B and  K.  C.  Peeler.  1995.  Spotted  Owls: 
resource  and  space  use  in  mosaic  landscapes.  J. 
Raptor  Res.  29:223-239. 

Forsman,  E.  D.,  E.  C.  Meslow,  and  H.  M.  Wight. 
1984.  Distribution  and  biology  of  the  Spotted  Owl 
in  Oregon.  Wildl.  Monogr.  87:1—64. 

Franklin,  A.  B.  1997.  Factors  affecting  temporal  and 
spatial  variation  in  Northern  Spotted  Owl  popu- 
lations in  northwest  California.  Ph.D.  diss.,  Col- 
orado State  Univ.,  Fort  Collins. 

Franklin,  A.  B.,  D.  R.  Anderson,  E.  D.  Forsman,  K. 

P Burnham,  and  F F Wagner.  1996.  Methods 
for  collecting  and  analyzing  demographic  data  on 
the  Northern  Spotted  Owl.  Stud.  Avian  Biol  17- 
12-26. 

Gutierrez,  R.J.,  A.  B.  Franklin,  and  W.  S.  LaHaye. 
1995.  Spotted  Owl  {Strix  occidenlalis).  In  The 
birds  of  North  America,  no.  179  (A.  Poole  and  F 
Gill,  Eds.).  The  Academy  of  Natural  Sciences, 
Philadelphia,  Pennsylvania;  The  American  Orni- 
thologists’ Union,  Washington,  D.C. 

Hall,  R.  and  K.  R.  Kelson.  1959.  The  mammals  of 
North  America.  Ronald  Press  Co.,  New  York 
Hedlund,  C.  D.  1996.  Relative  proportion  of  prey  spe- 
cies m the  diet  of  California  Spotted  Owls  Stri.x 
occicIenlali.K  occiclentalis  in  the  San  Gabriel 
Mountains  of  Southern  California.  M.Sc.  thesis, 
California  State  Polytechnic  Univ.,  Pomona. 

Hicks,  C.  R.  1993.  Fundamental  concepts  in  the  design 
ot  experiments.  Saunders  College  Publishing 
New  York. 

H.RONs,  G.  j.  M.  1985.  The  importance  of  body  re- 


Smith  et  ill.  • SPOTTED  OWL  DIET  AND  REPRODUCTION 


29 


serves  for  successful  reproduction  in  the  Tawny 
owl  (Strix  aluco).  J.  Zool.  (Lond.)  1:1-20. 

Hurlbert,  S.  H.  1984.  Pseudoreplication  and  the  de- 
sign of  ecological  field  experiments.  Ecol.  Mon- 
ogr.  54: 1 87—2 1 I . 

Korpimaki,  E.  1986.  Diet  variation,  hunting  habitat, 
and  reproductive  output  in  the  Kestrel  Falco  lin- 
niinculus  in  the  light  of  optimum  diet  theory.  Or- 
nis  Fennica  63:84-90. 

LaHaye,  W.  S.,  R.  J.  Gutierrez,  and  H.  R.  Akqa- 
KAYA.  1994.  Spotted  Owl  metapopulation  dynam- 
ics in  southern  California.  J.  Anim.  Ecol.  63:775- 
785. 

Laymon,  S.  D.  1988.  The  ecology  of  the  Spotted  Owl 
in  the  central  Sierra  Nevada,  California.  Ph.D. 
diss.,  Univ.  of  California,  Berkeley. 

Littel,  R.  C.,  G.  a.  Milliken,  W.  W.  Stroup,  and  R. 
D.  WoLFiNGER.  1996.  SAS  system  for  mixed  mod- 
els. SAS  Institute,  Cary,  North  Carolina. 

Major,  J.  and  D.  W.  Taylor.  1988.  Alpine.  Pp.  601- 
678  in  Terrestrial  vegetation  of  California  (M.  G. 
Barbour  and  J.  Major,  Eds.).  California  Native 
Plant  Society,  Sacramento,  California. 

Marti,  C.  D.  1987.  Raptor  food  habits  studies.  Pp.  67— 
81  in  Raptor  management  techniques  manual  (B. 
A.  G.  Pendleton,  B.  A.  Millsap,  K.  W.  Cline,  and 
D.  M.  Bird,  Eds.).  National  Wildlife  Federation, 
Washington,  D.C. 

Minnich,  R.  a.  1988.  The  biogeography  of  fire  in  the 
San  Bernardino  Mountains  of  California.  Univ.  of 
California  Press,  Los  Angeles. 

Newman,  J.  A.,  J.  Bergelson,  and  A.  Grafen.  1997. 
Blocking  factors  and  hypothesis  tests  in  ecology: 
is  your  statistics  text  wrong?  Ecology  78:1312- 
1320. 

Noon,  B.  R.  and  K.  S.  McKelvey.  1992.  Stability 
properties  of  the  Spotted  Owl  metapopulation  in 
Southern  California.  U.S.D.A.  Pacific  Southwest 
Research  Station.  Gen.  Tech.  Rep.  PSW-GTR- 
133:187-206. 

Orians,  G.  H.  and  N.  E.  Pearson.  1979.  On  the  theory 
of  central  place  foraging.  Pp.  155-177  in  Analysis 
of  ecological  systems  (D.  J.  Horn,  R.  D.  Mitchell, 
and  G.  R.  Stairs,  Eds.).  Ohio  Univ.  Press,  Colum- 
bus. 

Pyke,  G.  H.,  R.  H.  Pulliam,  and  E.  L.  Charnov.  1977. 
Optimal  foraging:  a selective  review  of  theory  and 
tests.  Quart.  Rev.  Biol.  52:137-154. 

Rice,  W.  R.  1989.  Analyzing  tables  of  statistical  tests. 
Evolution  43:223-225. 

Sawyer,  T.  O.,  D.  A.  Thornburgh,  and  T.  R.  Griffin. 
1988.  Mixed  evergreen  forest.  Pp.  359-382  in 


Terrestrial  vegetation  of  California  (M.  G.  Bar- 
bour and  J.  Major,  Eds.).  California  Native  Plant 
Society,  Sacramento,  California. 

Slotow,  R.  and  M.  R.  Perrin.  1992.  The  importance 
of  large  prey  for  Blackshouldered  Kite  reproduc- 
tion. Ostrich  63:180-182. 

Steenhof,  K.  and  M.  N.  Kochert.  1988.  Dietary  re- 
sponses of  three  raptor  species  to  changing  prey 
densities  in  a natural  environment.  J.  Anim.  Ecol. 
57:37-48. 

Swanson,  G.  A.,  G.  L.  Krapu,  J.  C.  Bartonek,  J.  R. 
Serie,  and  D.  H.  Johnson.  1974.  Advantages  in 
mathematically  weighting  waterfowl  food  habits 
data.  J.  Wildl.  Manage.  38:302-307. 

Thorne,  R.  E 1988.  Montane  and  subalpine  forests  of 
the  Transverse  and  Peninsular  Ranges.  Pp.  537- 
558  in  Terrestrial  vegetation  of  California  (M.  G. 
Barbour  and  J.  Major,  Eds.).  California  Native 
Plant  Society,  Sacramento,  California. 

Thrailkill,  j.  and  M.  A.  Bias.  1989.  Diets  of  breeding 
and  nonbreeding  California  Spotted  Owls.  J.  Rap- 
tor Res.  23:39-41 . 

Vasek,  F.  C.  and  M.  G.  Barbour.  1988.  Mojave  De- 
sert scrub  vegetation.  Pp.  835-867  in  Terrestrial 
vegetation  of  California  (M.  G.  Barbour  and  J. 
Major,  Eds.).  California  Native  Plant  Society,  Sac- 
ramento, California. 

Verner,  j.,  R.  j.  Gutierrez,  and  G.  I.  Gould,  Jr. 
1992.  The  California  Spotted  Owl:  general  biol- 
ogy and  ecological  relations.  U.S.D.A.  Pacific 
Southwest  Research  Station.  Gen.  Tech.  Rep. 
PSW-GTR- 133:55-77. 

Ward,  J.  R,  Jr.,  R.  J.  Gutierrez,  and  B.  R.  Noon. 
1998.  Habitat  selection  by  Northern  Spotted  Owls: 
the  consequences  of  prey  selection  and  distribu- 
tion. Condor  100:79-92. 

Waters,  J.  R.  and  C.  J.  Zabel.  1995.  Northern  flying 
squirrel  densities  in  fir  forests  of  northeastern  Cal- 
ifornia. J.  Wildl.  Manag.  59:858—866. 

White,  K.  1996.  Comparison  of  fledging  success  and 
sizes  of  prey  consumed  by  Spotted  Owl  in  north- 
western California.  J.  Raptor.  Res.  30:234-236. 

Yom-Tov,  Y.  and  D.  Wool.  1997.  Do  the  contents  of 
Barn  Owl  pellets  accurately  represent  the  propor- 
tion of  prey  species  in  the  field?  Condor  99:972— 
976. 

Zabel,  C.  J.,  K.  McKelvey,  and  J.  P.  Ward,  Jr.  1995. 
Influence  of  primary  prey  on  home-range  size  and 
habitat-use  patterns  of  Northern  Spotted  Owls 
(Strix  occidentaUs  caurina).  Can.  J.  Zool.  73:433- 
439. 

Zar,  j.  H.  1984.  Biostatistical  analysis.  Second  cd. 
Prentice-Hall,  Inc.,  Englewood  Cliffs,  New  Jersey. 


Wilson  Bull.,  111(1),  1999,  pp.  30-37 


FOOD,  FORAGING,  AND  TIMING  OF  BREEDING  OF  THE  BLACK 

SWIFT  IN  CALIFORNIA 

MANUEL  MARIN*  2 


ABSTRACT. — The  nestling  diet  and  breeding  seasonality  of  the  Black  Swift  {Cypseloides  niger)  were  studied 
in  southern  California  1990-1992.  The  peak  (40%)  of  egg  laying  was  in  mid-June  and  the  peak  of  fledging 
(60%)  was  mid-  to  late  August  {n  = 87  nests).  Winged  ants  comprised  91%  (n  = 1179  prey  items,  10  boluses) 
of  nestling  diet.  Three  main  prey  size  classes  were  found;  6,  8,  and  13  mm.  Food  bolus  mass  increased  and 
number  of  trips  per  day  to  feed  the  nestlings  decreased  with  nestling  age.  The  parents  made  short  and  long 
foraging  trips  during  early  morning  hours  and  long  trips  from  early  to  late  afternoon.  Short  trips  were  observed 
only  during  the  first  half  of  the  nestling  period.  During  the  last  half  of  the  nestling  period,  parent  swifts  made 
a single  foraging  trip  per  day  that  lasted  about  12  hrs.  Perhaps  the  short  foraging  bouts  are  for  feeding  the 
young,  whereas  the  long  foraging  bouts  are  not  only  for  feeding  the  young  but  also  for  parental  energy  storage. 
The  single  foraging  bout,  during  the  mid-  and  late  nestling  period,  might  also  serve  to  store  fat  for  migration 
by  the  adults.  Received  13  Feb.  I998\  accepted  24  Oct.  1998. 


The  Black  Swift  {Cypseloides  niger)  is  a 
member  of  the  subfamily  Cypseloidinae 
which  consists  of  12-13  species,  most  of 
which  are  tropical  or  subtropical  in  their 
breeding  distribution.  The  Black  Swift  is 
found  locally  in  the  West  Indies,  Middle 
America,  and  north  through  much  of  western 
North  America  to  southeastern  Alaska  (Bent 
1940;  AOU  1957,  1983).  For  a species  with 
such  a wide  latitudinal  distribution,  quantita- 
tive data  on  diet  and  timing  of  breeding  are 
rare.  Most  of  what  is  known  about  the  Black 
Swift  is  limited  to  breeding  and  distributional 
records  (e.g.,  Vrooman  1901,  1905;  Michael 
1927;  Dixon  1935;  Knorr  1961;  Foerster 
1987;  Foerster  and  Collins  1990;  Stiles  and 
Negret  1994).  This  reflects  the  difficulty  of 
studying  this  species  because  of  its  aerial  life 
style  and  its  usually  inaccessible  nest  sites. 
Here  1 present  new  information  on  diet  and 
timing  of  breeding  of  this  species. 

The  Black  Swift  is  a summer  breeding  vis- 
itor to  western  North  America,  and  like  many 
migratory  species  there,  has  a restricted  breed- 
ing season.  Furthermore,  it  has  a proportion- 
ately large  egg,  a single-egg  clutch,  is  single 
brooded,  and  has  a long  incubation  and  nest- 
ling period  (Marin  1997).  These  factors 


' Dept,  of  Biological  Science.s  and  Museum  of  Nat- 
ural .Science,  I 19  Foster  Hall,  Louisiana  State  Univ., 
Baton  Rouge,  LA  70803;  E-mail;  Zomari@LSU.edu 
^ Present  address;  Western  Foundation  of  Vertebrate 
Zoology.  439  Callc  San  Pablo,  Carmarillo,  CA  93012; 
E-mail;  MMA@wfv/..org 

30 


should  constrain  variation  in  the  timing  of 
breeding  of  the  Black  Swift. 

Swifts  catch  airborne  insects  and  ballooning 
spiders  (Lack  and  Owen  1955,  Whitacre 
1991).  During  the  reproductive  season,  breed- 
ing cypseloidine  swifts  accumulate  insects  and 
arachnids  in  the  back  of  the  throat  continuing 
into  the  esophagus  and  bind  them  with  saliva 
to  produce  a sticky  assortment  of  insects.  This 
insect  conglomerate  or  food  bolus  is  produced 
exclusively  to  feed  the  nestling(s)  and  have 
never  been  reported  outside  the  breeding  sea- 
son. Alive  or  dead  these  insects  are  complete, 
making  them  ideal  for  identifying  and  quan- 
tifying diets.  Hespenheide  (1975)  pointed  out 
that  one  difficulty  in  analyzing  bird  diets  is 
knowing  which  is  more  important:  the  number 
or  the  size  of  food  items.  Some  diet  studies 
of  New  World  swifts  have  addressed  both  pa- 
rameters [e.g.,  Whitacre  (1991)];  however, 
most  have  emphasized  only  one  or  they  had 
small  sample  sizes  (e.g.,  Collins  and  Landy 
1968,  Hespenheide  1975,  Foerster  1987,  Bull 
and  Beckwith  1993,  Marm  and  Stiles  1993). 
Other  authors  have  considered  only  the  num- 
ber of  prey  items  (e.g.,  Beebe  1949;  Rathbun 
1925;  Rowley  and  Orr,  1962,  1965;  Marin  and 
Stiles  1992).  One  problem  in  quantifying  swift 
diets  is  the  source  of  prey  samples,  either 
stomach  contents  or  food  boluses.  Stomach 
contents  of  adults  are  prey  items  that  the  adult 
bird  has  fed  upon,  whereas  food  boluses  are 
prey  fed  to  nestlings.  Thus,  examination  of 
stomach  contents  versus  food  boluses  might 
produce  different  results.  Because  my  focus 


Mcuin  • DIET  AND  BREEDING  OE  BLACK  SWIFT  IN  SOUTHERN  CALIFORNIA 


31 


was  on  the  diet  of  Black  Swift  nestlings,  1 
analyzed  only  food  boluses. 

METHODS 

Most  data  were  gathered  during  a study  of  the 
breeding  biology  of  the  Black  Swift  in  the  San  Jacinto 
Mountains,  Riverside  Co.,  California.  The  study  site  is 
at  1500  m elevation,  and  the  surrounding  area  is  mon- 
tane forest.  The  study  site  was  visited  40  times  from 
1990  to  1992,  between  the  months  of  May  and  Sep- 
tember. Observations  were  made  in  blocks  of  4 to  7 
hours,  from  05:00  to  12:00,  12:00  to  19:00,  or  19:00 
to  23:00  PST,  throughout  the  breeding  season.  For 
more  detailed  information  on  the  study  site  and  distri- 
bution of  visits  see  Marin  (1997)  and  references  there- 
in. 

Data  on  nestling  diet  were  obtained  from  regurgi- 
tated food  boluses  from  adult  swifts  captured  upon  ar- 
rival at  the  nest.  The  boluses  were  weighed  immedi- 
ately (to  nearest  0.1  g;  Pesola  scale)  and  placed  in  a 
vial  containing  alcohol.  Insects  in  the  boluses  were 
counted  and  measured  to  the  nearest  0.1  millimeter 
with  a micrometer  under  a microscope  in  the  lab. 

Ten  boluses  (n  = 1179  prey  items)  were  collected 
from  different  adults  on  different  dates  and  years 
throughout  the  study  to  minimize  nest  disturbance. 
Seven  boluses  were  collected  in  1991  and  three  during 
1992.  Bolus  mass  and  time  of  collection  were  recorded 
for  nine  of  them.  I measured  the  length  of  15  randomly 
selected  individual  prey  items  per  prey  species  per  bo- 
lus, and  I used  the  average  length  as  the  mean  of  that 
species  in  the  specific  food  bolus.  Insects  were  sorted 
to  morphospecies  and  identified  to  families  using  Bor- 
ror  and  Delong  (1970),  Borror  and  White  (1970),  and 
Powell  and  Hogue  (1979). 

To  assess  timing  of  breeding,  I combined  field  nest 
data  (n  = 20)  with  archived  nest  and  egg  data  cards 
(n  = 67;  see  Acknowledgments  for  list  of  sources), 
along  with  museum  study  skins  and  the  literature.  To 
determine  length  of  the  breeding  season,  I used  egg 
laying,  hatching,  and  fledgling  periods.  I restricted  anal- 
yses of  museum  egg  data  cards  to  two  areas  in  Cali- 
fornia: mountains  (San  Jacinto  area)  and  coast  (Santa 
Cruz  Co.).  I compared  the  estimated  hatching  date 
from  the  museum  egg  data  cards  to  my  own  field  data 
on  hatching  dates  gathered  at  San  Jacinto  to  look  for 
potential  date  discrepancies  between  actual  and  esti- 
mated data.  The  incubation  stage  given  in  the  egg  data 
cards  (e.g.,  fresh,  commenced,  advanced,  etc.)  was  ex- 
trapolated using  the  known  days  of  the  incubation 
stage  from  San  Jacinto  (24  days;  Marin  1997).  I can- 
died  16  eggs  at  San  Jacinto  and  determined  that  no 
egg  of  this  species  could  be  blown  without  signifi- 
cantly damaging  or  destroying  the  shell  by  day  16-18. 
Thus,  any  museum  egg  specimen  of  this  species  was 
unlikely  to  have  been  collected  beyond  18  days  of  in- 
cubation. I estimated  the  duration  of  each  stage  visible 
through  candling  as:  “fresh”  (0-4,  5 days)  “veins  and 
small  embryo”  (5-10  days),  “embryo”  (11-14  days), 
and  “large  embryo”  (>15  days).  These  data  were  used 


TABLE  1. 

Contents  of  food  boluses  (n  = 10)  of 

Cyp.seloides  niger  from  San  Jacinto,  California. 

Order 

Family 

Number 

Isoptera 

Hodotermitidae 

1 

1 

Hemiptera 

27 

Pentatomidae 

3 

Coreidae 

10 

Miridae 

2 

Reduviidae 

11 

Nabidae 

1 

Homoptera 

54 

Cicadidae 

17 

Cicadellidae 

36 

Aphidae 

1 

Neuroptera 

8 

Hemerobiidae 

1 

Myrmeleonitidae 

3 

Corydalidae 

1 

Family? 

3 

Coleoptera 

1 

Buprestidae 

1 

Lepidoptera 

3 

Pyralidae 

1 

Family? 

2 

Diptera 

Family? 

11 

Hymenoptera 

1074 

Formicidae 

1074 

Arachnidea 

1 

TOTAL 

1179 

to  infer  incubation  stage  (from  the  data  cards)  and  to 
estimate  egg  laying  dates.  Using  the  known  incubation 
and  fledgling  period  (24  and  48  days,  respectively; 
Marin  1997),  the  estimated  laying,  hatching,  and  fledg- 
ing dates  were  compared  and  then  combined  with  field 
data.  The  dates  were  separated  by  month,  and  each 
month  was  subdivided  into  early,  middle,  and  late. 

Rainfall  data  were  gathered  for  each  month  for  the 
years  of  study  (1990-1992)  and  also  1963-1973  from 
the  Idyllwild  Fire  Department  weather  station  about 
1.5  km  from  study  area  (National  Oceanic  and  At- 
mospheric Administration  1991,  1992). 

I was  able  to  identify  adults  individually  because 
they  were  already  banded  by  C.  Collins  and  K.  Foers- 
ter  because  this  was  the  main  study  site  that  Foerster 
{ 1987)  used  for  his  MS  thesis  work. 

RESULTS 

Nestling  diet  and  foraging. — All  boluses 
were  composed  of  one  predominant  prey  spe- 
cies, suggesting  that  the  birds  had  fed  on 
swarming  species  or  highly  localized  prey. 
Winged  ants  were  the  majority  of  prey  items 
(91%;  Table  1).  In  10  boluses  the  average  pro- 
portion of  female  winged  ants  was  79.5  % 


32 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


1 I 

0-10  ■ ■ 

qI  Imbj-I 

1 2 3 4 5 6 7 8 9 10  11  12  13  14 

LENGTH  (mm) 

FIG.  1.  Frequency  distribution  of  prey  sizes  taken  by  the  Black  Swift  {Cypseloides  niger)  in  southern 
Calitornia.  Data  are  from  10  food  boluses  (/;  = 1154  prey  items);  prey  size  categories  are;  0.5— 1.5  mm  = I; 
1.6-2. 5 mm  = 2;  2. 6-3. 5 mm  = 3;  etc. 


(range  20-100%;  n = 1179  prey  items).  Sex- 
ual dimorphism  in  ants  accounted  for  prey 
size  differences  among  boluses.  Male  ants 
were  smaller  and  ranged  2-6  mm,  whereas  fe- 
males ranged  6-14  mm.  Sometimes  a bolus 
contained  a small  number  of  female  ants  but 
those  ants  were  the  largest  prey.  Thus,  female 
winged  ants  had  the  highest  volume/prey  in 
all  boluses.  The  average  measurable  prey  size 
was  7.4  mm  (range  1.8-14.5  mm;  n = 1 154). 
Three  main  size  classes  were  found  in  the  bo- 
luses: 6 mm  (33%),  8 mm  (39%),  and  13  mm 


(12.2%;  Fig.  1).  Two  length  categories  (6  and 
8 mm)  made  up  the  bulk  of  the  samples.  Be- 
cause only  a few  prey  taxa  were  represented 
in  the  diet,  size  frequencies  for  all  prey  spe- 
cies follow  a similar  pattern. 

Nestling  age  and  bolus  mass  were  positive- 
ly correlated  (r"  = 0.93,  P < 0.001,  /?  = 10; 
Fig.  2).  No  correlation  was  found  between 
mean  prey  size  per  bolus  and  chick  age  (r-  = 
0.08,  P > 0.05)  or  mass  per  bolus  (r^  = 0.04, 
P > 0.05,  ?i  = 10).  No  individual  prey  item 
was  weighed,  but  female  winged  ants  were 


CO 

CO 

< 


(O 

z> 

_l 

o 

m 


0 


••  • • 


50 


0 10  20  30  40 

AGE  OF  YOUNG  (days) 

FIG.  2.  Relation.ship  between  bolu.s  mass  and  nestling  age  (F  = 0.93;  P < 0 001  n 


= 10). 


Mcuin  • DIET  AND  BREEDING  OE  BLACK  SWIFT  IN  SOUTHERN  CALIFORNIA 


33 


FIG.  3.  Time  of  day  young  were  fed  versus  their  age.  None  of  the  adult  birds  were  observed  feeding  young 
between  the  two  feeding  clusters  or  during  early  morning  after  30  days. 


undoubtedly  the  heaviest  prey  items  because 
they  were  the  largest.  After  day  30  I never  saw 
any  adults  feed  young  in  the  morning;  they 
were  fed  mainly  late  in  the  evening  (Fig.  3). 
Other  nestlings  not  included  in  the  analysis 
because  they  were  inaccessible  and  of  un- 
known age,  but  at  least  30  days  old,  were  also 
observed  being  fed  between  18:30—20:00.  The 
overall  pattern  seems  to  be  that  as  age  in- 
creased, feeding  rate  decreased,  but  bolus 
mass  increased  (Figs.  2,  3). 

Data  gathered  from  adults  feeding  young  at 
the  nest  showed  two  clusters  of  feeding  times: 
between  8:30  and  12:30  and  after  18:30  (Fig. 
3).  The  birds  usually  left  the  cave  at  about  05: 
30.  In  two  instances,  however,  some  departed 
earlier  unnoticed  because  of  the  darkness. 
This  implies  that  they  spent  3-7  hours  search- 
ing for  food  for  the  first  feeding  bout.  For  the 
second  bout,  the  birds  were  away  from  the 
nest  longer:  6-8  hrs  (Fig.  3).  I never  observed 
nestlings  being  fed  between  12:30-18:30.  I 
did  not  gather  data  late  in  the  evening  or  at 
night  during  the  early  nestling  stage  so  late 
arrivals  and  feeding  at  that  stage  are  possible. 
If  the  intervals  between  feeding  bouts  were 
consistent  through  the  season,  then  older  nest- 
lings often  waited  more  than  12  hrs  between 
meals  when  they  were  well  grown  or  more 
than  30  days  of  age. 

Breeding  season. — Hatching  dates  were  the 


main  variable  I used  to  compare  the  timing  of 
breeding  between  a mountain  site  (San  Jacinto 
area,  San  Bernardino  Co.,  California)  and  a 
coastal  site  (Santa  Cruz  area,  Santa  Cruz  Co., 
California).  The  San  Jacinto  data  were  pri- 
marily from  my  observations,  whereas  the 
Santa  Cruz  data  were  taken  from  museum  nest 
and  egg  data  cards.  I found  no  significant  dif- 
ference between  the  coastal  and  mountain 
sites  in  timing  of  hatching  (Fisher’s  exact  Test 
(2-tail):  P > 0.05).  Therefore,  I concluded  that 
it  was  safe  to  pool  both  field  and  museum  data 
for  coastal  and  interior  southern  California. 
Most  eggs  were  laid  during  mid-June  (40%), 
with  30%  during  late  June  (Fig.  4).  The  ear- 
liest laying  date  was  estimated  to  be  18  May, 
from  an  egg  set  collected  in  1960  near  Santa 
Cruz,  California.  The  latest  date  for  egg  laying 
on  the  data  cards  was  estimated  to  be  12  July 
1921,  from  the  same  site  as  the  earliest  date. 
Some  of  the  observed  variation  might  be  due 
to  inter-year  differences,  which  are  difficult  to 
evaluate  with  the  present  data.  Nevertheless, 
most  eggs  (81%)  had  an  estimated  laying  date 
in  June.  The  earliest  estimated  date  for  hatch- 
ing was  about  1 1 June  (same  nest  as  above) 
and  the  latest  date  was  about  5 August  from 
the  same  site.  In  total,  89%  of  the  hatching 
dates  were  in  July;  24%  of  the  estimated 
hatching  dates  were  during  the  first  10  days 
of  July  and  53%  during  the  middle  third  of 


34 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  1,  March  1999 


FIG.  4.  Distribution  of  estimated  dates  of  egg  laying  (open  bars),  hatching  (black  bais),  and  fledging  (stipled 
on  bars)  of  Black  Swifts  in  southern  California. 


July.  Sixty  percent  of  the  young  were  esti- 
mated to  fledge  during  mid-  to  late  August 
and  29%  during  the  first  10  days  of  Septem- 
ber. The  highest  proportion  of  fledging  (78%) 
was  estimated  to  occur  between  late  August 
and  early  September  (Fig.  4). 

DISCUSSION 

Contrary  to  most,  if  not  all  tropical  cypse- 
loidines,  the  southern  California  populations 
of  the  Black  Swift  breeds  during  the  dry  sea- 
son. For  the  small,  tropical  cypseloidine 
swifts,  rainfall  itself  can  be  as  important  a 
stimulus  as  food  in  initiating  breeding.  Mois- 
ture is  needed  to  keep  the  appropriate  condi- 
tions for  nest  “growth”  and  maintenance 
(Marm  and  Stiles  1992).  Many  Black  Swifts 
did  not  build  a nest  at  all,  instead  layed  eggs 
directly  on  ledges,  especially  in  the  coastal 
sites  (Man'll  1997).  This  might  be  related  to 
the  lack  of  the  proper  nesting  materials  (moss- 
es and  liverwoths). 

The  breeding  season  of  the  Black  Swift  in 
southern  California  is  spread  over  4.5-5 
months  (Foerster  1987,  Marin  1997).  Lack 
(1954,  1968)  observed  that  breeding  in  most 
species  of  birds  is  timed  to  occur  when  food 
is  most  abundant,  especially  in  temperate  re- 
gions. 

In  the  western  United  States,  Chapman 
( 1 954)  noted  that  ants  swarmed  from  May 


through  September  and  that  the  peak  of  ant 
swarming  was  July.  The  observed  peak  of  egg 
hatching  in  Black  Swifts  was  also  July  coin- 
ciding with  the  peak  of  ant  swarming  (Fig.  5). 
These  data  support  Holroyd’s  and  Jalkotzy’s 
(in  Campbell  et  al.  1990)  suggestion  that  the 
breeding  of  the  Black  Swift  in  southwestern 
Canada  was  timed  to  the  swarming  of  flying 
ants  (Hymenoptera).  In  the  western  U.S.  ants 
swarm  in  large  numbers  on  mountain  and 
ridge  tops  for  several  days  (Chapman  1954). 
The  peak  time  of  ant  swarming  observed  by 
Chapman  (1954)  was  from  07:00  to  14:00  and 
coincides  with  the  first  period  of  shorter  feed- 
ing bouts  in  the  Black  Swifts  (Fig.  3). 

Foerster  (1987)  reported  average  prey  sizes 
from  two  boluses  (n  = 289  prey  items)  as  9.9 
and  10.2  mm,  slightly  larger  than  my  aver- 
ages. He  did  not  report  sizes  smaller  than  7 
mm  or  larger  than  13  mm.  Foster  (1987)  spec- 
ulated on  possible  size  selection  by  the  swifts; 
however,  I observed  46.5%  percent  of  prey 
items  below  and  above  those  categories  (n  = 
1179  prey  items,  10  boluses).  The  . data  from 
this  and  other  studies  (Collins  and  Landy 
1968,  Foerster  1987)  suggest  that  prey  items 
given  to  the  nestlings  are  selected  not  by  size 
but  by  insect  taxon.  This  is  probably  a con- 
sequence of  feeding  on  insect  swarms. 

The  mam  diet  of  Black  Swift  nestlings  at 
San  Jacinto  was  winged  ants,  which  have  a 


Marin  • DIET  AND  BREEDING  OE  BLACK  SWIFT  IN  SOUTHERN  CALIFORNIA 


35 


> 

O 

z 

UJ 

3 

a 

LU 

IT 


> 

GC 

LU 

C/3 

CQ 

O 


100 


80 


60 


40 


20 


□ ANT  SWARMING 
■ SWIFT  HATCHING 


JUN 


JUL  AUG 

MONTH 


SEP 


FIG.  5.  Frequency  of  Black  Swift  hatching  and  ant  swarming  in  southern  California.  Data  for  ant  swarming 
are  from  Chapman  (1954). 


high  fat  content.  The  large  preponderance  of 
winged  ants  in  the  nestling  diet  is  similar  that 
of  other  cypseloidine  swifts  (Whitacre  1991). 
The  percent  fat  per  dry  weight  in  alate  ants 
ranges  from  23.8  to  59.5%  in  females  and 
from  3.3  to  9.6%  in  males  (Taylor  1975,  Red- 
ford  and  Dorea  1984).  A nestling  of  any  bird 
species  fed  a diet  rich  in  energy  could  accu- 
mulate large  amounts  of  subcutaneous  fat.  Be- 
fore fledging  the  young  Black  Swift  accumu- 
lates much  visible  subcutaneous  fat  and  at- 
tains up  to  148%  of  adult  body  mass;  it  reach- 
es adult  mass  at  day  15-16  of  the  nestling 
period  (Marin  1997).  The  limited  inter-year 
sampling  by  Foerster  (1987)  and  myself  sug- 
gests that  the  swifts  at  San  Jacinto,  during  the 
breeding  season,  may  specialize  in  exploiting 
local  concentrations  of  2-3  ant  species  (Cam- 
ponotus  spp.).  Winged  ants  are  a temporarily 
superabundant,  patchy,  and  ephemeral,  but 
lipid-rich  food  source.  Other  important  prey 
items  included  Hemiptera  and  Homoptera  (Ta- 
ble 1). 

From  scattered  observations,  (e.g.,  Michael 
1927,  Smith  1928,  Bent  1940,  Collins  1998, 
Collins  and  Peterson  1998)  there  is  a general 
agreement  that  Black  Swift  nestlings  are  fed 
at  long  intervals,  primarily  early  in  the  morn- 
ing and  late  in  the  afternoon  or  at  night.  My 
data  corroborate  those  conclusions  (Fig.  3). 
The  alternation  of  long  and  short  foraging 


trips  resembles  the  strategy  of  energy  expen- 
diture described  for  foraging  and  food  deliv- 
ery in  pelagic  seabirds.  Charurand  and  Wei- 
merskirch  (1994)  and  Weimerskirch  and  co- 
workers (1994)  showed  that  long  trips  were 
primarily  for  parental  food  storage  as  well  as 
nestling  food  gathering,  whereas  short  trips 
were  used  to  deliver  food  to  the  nestlings.  Al- 
though the  duration  of  seabirds’  trips  is  days, 
instead  of  hours  as  in  swifts,  they  might  well 
serve  analogous  purposes. 

Like  seabirds,  the  Black  Swift  might  gain 
weight  on  the  long  trips  and  lose  it  overnight. 
Black  Swifts  have  a high  metabolic  rate  and 
lose  on  average  7.9  % of  body  mass  overnight 
(Marin,  unpubl.  data).  Thus,  the  need  for  the 
long  foraging  bout  is  in  accordance  with  the 
energy  storage  hypothesis  (Chaurand  and 
Weimerskirch  1994,  Weimerskirch  et  al. 
1994).  Black  Swift  migration  occurs  imme- 
diately after  the  nestlings  fledge;  other  species 
of  swifts  (e.g.,  Chaetura  spp.)  stay  a few 
months  after  breeding,  probably  to  store  some 
energy  for  migration  (Marm  1997).  Accord- 
ingly, this  long  single  foraging  bout  might 
also  serve  to  store  energy  for  migration,  par- 
ticularly during  the  later  part  of  the  breeding 
season. 

As  the  Black  Swift  nestlings  increase  in 
age,  it  seems  that  the  adults  feed  them  only 
late  at  night.  Quantitative  data  on  feeding 


36 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  1,  March  1999 


rates  or  number  of  trips  per  day  with  respect 
to  nestling  age  are  scarce;  however,  declines 
in  the  number  of  trips  per  day  with  nestling 
age  have  been  reported  in  other  swift  species 
(Malacame  et  al.  1992,  Oniki  et  al.  1992). 

Lack  (1954,  1968)  suggested  that  seabirds 
with  long  nestling  periods  and  single  egg 
clutches  were  energy  limited.  This  energy  lim- 
itation was  in  food  finding,  food  delivery,  or 
both.  Some  swifts,  particularly  cypseloidines, 
have  life  history  parameters  similar  to  procel- 
lariiform  seabirds  (Lack  and  Lack  1951;  Lack 
1956,  1968;  Marin  and  Stiles  1992;  Marin 
1993).  For  example,  the  Black  Swift  rears  a 
single,  slowly  growing  nestling,  which  sug- 
gests a constraint  in  either  finding  or  deliver- 
ing food.  Because  Black  Swift  nestlings  re- 
quire a highly  specialized  diet  in  order  to  have 
an  initial  fast  growth  and  acquire  a size  larger 
than  the  adult  quickly.  1 predict  that  the  major 
constraint  is  food  finding. 

ACKNOWLEDGMENTS 

I thank  R.  Corado,  L.  Lyon.  D.  MacLean,  J.  Schmitt. 
C.  Sumida.  and  W.  Wehtje  for  their  companionship  and 
field  help.  This  paper  benefited  from  comments  by  C. 
T.  Collins.  K,  Naoki.  A.  G.  Navarro  S..  J.  V.  Remsen. 
A.  Styring.  D.  Whitacre.  and  an  anonymous  reviewer. 

I thank  the  following  museums  for  access  to  specimens 
and  egg  card  data:  R Sweet.  M.  LeCroy,  and  E Vuil- 
leumier.  American  Museum  of  Natural  History,  New 
York;  N.  Johnson  and  C.  Cicero,  Museum  of  Verte- 
brate Zoology  University  of  California,  Berkeley;  R. 
Prys-Jones  and  M.  Walters,  Natural  History  Museum. 
Tring  (ex-British  Museum  of  Natural  History);  R Col- 
lins, Santa  Barbara  Museum  of  Natural  History,  Santa 
Barbara;  L.  Kiff,  C.  Sumida,  R.  Corado,  and  W.  We- 
htje, Western  Foundation  of  Vertebrate  Zoology,  Cam- 
arillo, The  Western  Foundation  of  Vertebrate  Zoology 
helped  with  funds  for  the  field  work  and  with  a small 
grant  to  cover  some  preparation  costs,  and  AVINET 
provided  some  field  equipment.  Financial  assistance 
was  also  provided  by  a Collection  Study  Grant  of  the 
American  Museum  of  Natural  History.  I am  grateful 
to  all  those  egg  and  bird  collectors  who  seldom  receive 
recognition  for  what  they  did. 

LITERATURE  CITED 

American  Ornithologi.sts'  Union.  1957.  Check-list 
of  North  American  birds,  fifth  ed.  American  Or- 
nithologists' Union.  Port  City  Press,  Baltimore, 
Maryland. 

American  Ornithologi.sts'  Union.  1983.  Check-list 
of  North  American  birds,  sixth  ed.  Allen  Press, 
Lawrence,  Kan.sas. 

Beebe.  W,  1949,  The  swifts  of  Rancho  Grande,  north 


central  Venezuela,  with  special  reference  to  mi- 
gration. Zoologica  34;53-62. 

Bent,  A.  C.  1940.  Life  histories  of  North  American 
cuckoos,  goatsuckers,  hummingbirds,  and  their  al- 
lies. U.S.  Nat.  Mus.  Bull.  176:1-506. 

Borror,  D.  L.  and  D.  M.  DeLong.  1970.  An  intro- 
duction to  the  study  of  insects.  Holt,  Rinehart  and 
Winston,  New  York. 

Borror,  D.  L.  and  R.  E.  White.  1970.  A field  guide 
to  the  insects.  Houghton  Mifflin  Co.,  Boston, 
Massachusetts. 

Bull,  E.  L.  and  R.  C.  Beckwith.  1993.  Diet  and  for- 
aging behavior  of  Vaux's  Swifts  in  northeastern 
Oregon.  Condor  95:1016-1023. 

Campbell,  R.  W,  N.  K.  Dawe,  I.  McTaggart-Cowan, 
J.  M.  Cooper,  G.  W.  Kaiser,  and  M.  C.  E. 
McNall.  1990.  The  birds  of  British  Columbia. 
Vol.  11,  nonpasserines.  Royal  British  Columbia 
Museum,  Victoria. 

Chapman,  J.  A.  1954.  Swarming  of  ants  on  western 
United  States  mountain  summits.  Pan-Pacific  En- 
tomol.  30:93-102. 

Chaurand,  T.  and  H.  Weimerskirch.  1994.  The  reg- 
ular alternation  of  short  and  long  foraging  trips  in 
the  Blue  Petrel  Halohaeno  caerulea'.  a previously 
undescribed  strategy  of  food  provisioning  in  a pe- 
lagic seabird.  J.  Anim.  Ecol.  63:275-282. 

Collins,  C.  T.  1998.  Food  delivery  and  chick  provi- 
sioning in  cypseloidine  swifts.  Bull.  Brit.  Orni- 
thol.  Club  1 18:108-1 12. 

Collins,  C.  T.  and  M.  J.  Landy.  1968.  Breeding  of 
the  Black  Swift  in  Veracruz,  Mexico.  Bull.  South. 
Calif.  Acad.  Sci.  67:266-268. 

Collins,  C.  T.  and  B.  M.  Peterson.  1998.  Nocturnal 
chick  provisioning  by  Black  Swifts.  West  Birds 
29:227-228. 

Dixon,  J.  S.  1935.  Nesting  of  the  Black  Swift  in  Se- 
quoia National  Park.  Condor  37:265-267. 

Foerster,  K.  S.  1987.  The  distribution  and  breeding 
biology  of  the  Black  Swift  {Cypseloides  niger)  in 
southern  California.  M.Sc.  thesis.  California  State 
Univ.,  Long  Beach. 

Foerster.  K.  S.  and  C.  T.  Collins.  1990.  Breeding 
distribution  of  the  Black  Swift  in  southern  Cali- 
fornia. West.  Birds  21:1-9. 

Hespenheide,  H.  a.  1975.  Prey  characteristics  and 
predator  niche  width.  Pp.  158-180  in  Ecology  and 
evolution  of  communities  (M.  L.  Cody  and  J.  M. 
Diamond,  Eds.).  Belknap  Press,  Cambridge,  Mas- 
sachusetts. 

Knorr,  O.  a.  1961.  The  geographical  and  ecological 
di,stribution  of  the  Black  Swift  in  Colorado.  Wil- 
son Bull.  73:155-170. 

Lack.  D.  1954.  The  natural  regulation  of  animal  num- 
bers. Clarendon  Press,  Oxford,  U.K. 

Lack,  D.  1956.  Swifts  in  a tower.  Methuen  and  Co 
London.  U.K. 

Lack,  D.  1968.  Ecological  adaptations  for  breeding  in 
birds.  Methuen  and  Co.,  London,  U.K 

Lack  D^ AND  E.  Lack.  1951.  The  breeding  biology  of 
the  Swift  Apu.s'  apus.  Ibis  93:501-546. 


Mann  • DIET  AND  BREEDING  OE  BLACK  SWIFT  IN  SOUTHERN  CALIFORNIA 


37 


Lack,  D.  and  D.  E Owen.  1955.  The  food  of  the  swift. 
J.  Anim.  Ecology.  24:120-136. 

Malacarne,  G.,  M.  Cucco,  and  G.  Orecchia.  1992. 
Nest  attendance,  parental  roles  and  breeding  suc- 
cess in  the  Pallid  Swift  (Apus  pallidus).  Vogel- 
warte  36:203-210. 

Marin,  M.  1993.  Patterns  of  distribution  of  swifts  in 
the  Andes  of  Ecuador.  Avocetta  17:117-123. 

Marin,  M.  1997.  Some  aspects  of  the  breeding  biology 
of  the  Black  Swift.  Wilson  Bull.  109:290-306. 

Marin,  M.  and  E G.  Stiles.  1992.  On  the  biology  of 
five  species  of  swifts  (Apodidae,  Cypseloidinae) 
in  Costa  Rica.  Proc.  West.  Found.  Vertebr.  Zool. 
4:287-351. 

Marin,  M.  and  E G.  Stiles.  1993.  Notes  on  the  bi- 
ology of  the  Spot-fronted  Swift.  Condor  95:479- 
483. 

Michael,  C.  W.  1927.  Black  Swifts  nesting  in  Yosem- 
ite  National  Park.  Condor  24:89-97. 

National  Oceanic  and  Atmospheric  Administra- 
tion. 1991.  Climatological  data  California.  95(6- 
9):  1-28. 

National  Oceanic  and  Atmospheric  Administra- 
tion. 1992.  Climatological  data  California.  96(6— 
9):  1-28. 

Oniki,  Y,  E.  O.  Willis,  and  M.  M.  Willis.  1992. 
Chaetura  andrei  ( Apodiformes,  Apodidae):  as- 
pects of  nesting.  Ornitol.  Neotrop.  3:65-68. 

Powell,  J.  E.  and  C.  L.  Hogue.  1979.  California  in- 
sects. Univ.  of  California  Press,  Berkeley. 

Rathbun,  S.  F.  1925.  The  Black  Swift  and  its  habits. 
Auk  42:497-516. 

Redford,  K.  H.  and  j.  G.  Dorea.  1984.  The  nutri- 


tional value  of  invertebrates  with  emphasis  on  ants 
and  termites  as  food  for  mammals.  J.  Zool. 
(Lond.)  203:385-395. 

Rowley,  J.  S.  and  R.  T.  Orr.  1962.  Nesting  and  feed- 
ing habits  of  the  White-naped  Swift.  Condor  64: 
361-367. 

Rowley,  J.  S.  and  R.  T.  Orr.  1965.  Nesting  and  feed- 
ing habits  of  the  White-collared  Swift.  Condor  67: 
449-456. 

Smith,  E.  1928.  Black  Swifts  nesting  behind  a water- 
fall. Condor  30:136-138. 

Stiles,  F.  G.  and  A.  J.  Negret.  1994.  The  nonbreeding 
distribution  of  the  Black  Swift:  a clue  from  Co- 
lombia and  unsolved  problems.  Condor  96:1091  — 
1094. 

Taylor,  R.  L.  1975.  Butterflies  in  my  stomach.  Or: 
insects  in  human  nutrition.  Woodbridge  Press, 
Santa  Barbara,  California. 

Vrooman,  a.  G.  1901.  Discovery  of  the  egg  of  the 
Black  Swift  Cypseloides  niger  borealis.  Auk  18: 
394-395. 

Vrooman,  A.  G.  1905.  Discovery  of  a second  egg  of 
the  Black  Swift.  Condor  7:176—177. 

Weimerskirch,  H.,  O.  Chastel,  L.  Ackermann,  T. 
Chaurand,  F.  Cuenot-Chaillet,  X.  Hindermey- 
ER,  AND  J.  Judas.  1994.  Alternate  long  and  short 
foraging  trips  in  pelagic  seabird  parent.  Anim.  Be- 
hav.  47:472-476. 

Whit  ACRE,  D.  F.  1991.  Studies  of  the  ecology  of  the 
White-collared  Swift  (Streptoprocne  zonaris)  and 
White-naped  Swift  (S.  semicollaris),  and  of  pat- 
terns of  adaptation  among  the  swifts  (Aves:  Apod- 
idae). Ph.D.  diss.,  Univ.  of  California,  Davis. 


Wilson  Bull.,  111(1),  1999,  pp.  38-45 


FACTORS  THAT  INFLUENCE  TRANSLOCATION  SUCCESS  IN  THE 

RED-COCKADED  WOODPECKER 

KATHLEEN  E.  FRANZREB' 


ABSTRACT— To  restore  a population  that  had  declined  to  4 individuals  by  late  1985,  54  Red-cockaded 
Woodpeckers  (Picoides  borealis)  were  translocated  at  the  Savannah  River  Site  in  South  Carolina  between  1986 
and  1995.  Translocation  success  was  evaluated  by  sex,  age,  and  distance  between  the  capture  and  release  site. 
For  moves  involving  females,  the  presence  of  a resident  male  and  the  status  of  the  male  (breeder,  inexperienced, 
or  helper)  also  was  assessed.  Of  the  factors  I evaluated,  only  the  distance  of  the  move  was  statistically  significant 
with  increasing  success  associated  with  increasing  distance.  The  presence  of  a resident  male  at  the  female’s 
release  site  led  to  no  more  success  than  releasing  the  female  concunently  with  a male;  nor  did  the  male’s  status 
appear  to  play  a significant  role  in  female  translocation  success.  Overall,  31  of  49  (excluding  nestlings)  trans- 
located birds  remained  at  or  near  the  release  site  for  at  least  30  days,  resulting  in  a success  rate  of  63.2%.  Of 
the  birds  that  were  successfully  translocated,  51.0%  had  reproduced  by  July  1996.  Received  2 March  1998 
accepted  15  Oct.  1998. 


Endemic  to  the  open  pine  woodlands  of  the 
South,  Red-cockaded  Woodpeckers  {Picoides 
borealis)  are  cooperative  breeders  whose 
groups  usually  consist  of  a breeding  pair  and 
often  one  or  more  helpers,  usually  male  off- 
spring (U.S.  Fish  and  Wildlife  Service  1985). 
A series  of  cavity  trees  occupied  by  such  a 
group  is  referred  to  as  a cluster.  These  cavities 
are  used  year  round  for  night  roosting  and  as 
nest  sites  during  the  breeding  season  (Steirly 
1957).  Since  1970  the  species  has  been  con- 
sidered Federally  endangered  primarily  be- 
cause of  widespread  habitat  loss,  which  has 
fragmented  the  original  population  into  many 
subunits,  some  quite  small  and/or  isolated 
(U.S.  Fish  and  Wildlife  Service  1985).  One 
such  small  population  occupies  the  Savannah 
River  Site  in  South  Carolina. 

By  late  1985  the  number  of  Red-cockaded 
Woodpeckers  had  dwindled  to  one  breeding 
pair  and  two  single  males  (DeFazio  et  al. 
1987),  and  the  Forest  Service  began  intensive 
management  to  prevent  extirpation  on  the  site 
(Gaines  et  al.  1995).  With  the  nearest  known 
Red-cockaded  Woodpecker  population  32  km 
away,  natural  recruitment  of  and  colonization 
by  new  individuals  was  considered  unlikely. 
Because  Red-cockaded  Woodpeckers  prefer 
older,  live  pine  trees  for  constructing  their 
cavities  (Steirly  1957,  Jackson  et  al.  1979, 
Conner  and  O’Halloran  1987,  Rudolph  and 
Conner  1991)  and  few  trees  of  sufficient  age 

' Southern  Re.scarch  Station-USDA  Forest  Service, 
Dept,  of  Forest  Re.sourees,  Clemson  Univ.,  Clemson, 
.SC  29634—1003;  E-mail:  KFRANZR@clcmson.edu 


and  diameter  were  available.  Forest  Service 
personnel  installed  305  artificial  cavities  (see 
Allen  1991  for  details  on  artificial  cavity  con- 
struction and  installation).  Other  management 
activities  have  included:  (1)  restricting  cavity 
access  by  other  larger  woodpecker  species 
with  metal  “restrictor”  plates  (Carter  et  al. 
1989),  (2)  removing  southern  flying  squirrels 
{Glaucomys  volans)  encountered  while  moni- 
toring cavities  and  squirrel  nest  boxes,  and  (3) 
improving  habitat  quality  by  controlling  the 
hardwood  midstory  vegetation  that  causes 
woodpeckers  to  abandon  their  cavities  (Con- 
ner and  Rudolph  1989,  Costa  and  Escano 
1989,  Hooper  et  al.  1991,  Loeb  et  al.  1992). 

In  an  effort  to  stabilize  and  eventually  in- 
crease the  population  at  the  site,  the  Forest 
Service  began  a program  of  translocating 
woodpeckers  from  populations  outside  of  and 
within  the  site.  The  objectives  were  to  in- 
crease the  number  of  breeding  pairs,  bolster 
the  overall  population  size,  and  minimize  po- 
tential adverse  genetic  consequences  arising 
from  small  population  size  (Allen  et  al.  1993, 
Gaines  et  al.  1995).  Here  I assess  the  results 
of  10  years  of  Red-cockaded  Woodpecker 
translocations  at  the  Savannah  River  Site  to 
determine  the  variables  most  likely  to  contrib- 
ute to  successful  translocations,  an  important 
strategy  in  the  recovery  of  small,  isolated  pop- 
ulations. 

STUDY  AREA  AND  METHODS 

Study  «m^_The  Savannah  River  Site  lies  within 
the  Upper  Coastal  Plain  physiographic  region  in  Ai- 
ken, Allendale,  and  Barnwell  counties  in  South  Caro- 


38 


Fnmzreh  • RED-COCKADED  WOODPECKER  TRANSLOCATIONS 


39 


lina.  By  the  early  195()s,  most  of  the  site  was  in  ag- 
ricultural use  or  had  been  harvested  for  timber.  In 
1951,  the  Department  of  Energy  (DOE)  acquired 
80,269  ha  of  contiguous  land  to  develop  the  area  as  a 
nuclear  production  facility.  Under  an  interagency 
agreement,  the  Savannah  River  Natural  Resource  Man- 
agement and  Research  Institute  (U.S.  Department  of 
Agriculture,  Forest  Service)  has  managed  the  natural 
resources  on  the  site  for  DOE  since  1952.  Today  ap- 
proximately 69.000  ha  on  the  site  are  in  pine  stands 
(Workman  and  McLeod  1990),  most  of  which  are  less 
than  50  years  old  although  there  are  some  residual  old- 
er pine  trees.  The  area  managed  for  the  woodpecker 
contains  3 1,970  ha  of  pine  forest  consisting  of  longleaf 
(Pimis  paliisiris;  Yl .1%  of  the  pine  acreage),  loblolly 
(P.  taedo:  45.4%),  slash  {P.  elliotti;  13.4%),  and  other 
(0.2%)  pines  in  addition  to  pine-hardwoods  (3.3%;  G. 
Gaines,  unpubl.  data). 

Field  methods. — Beginning  in  1980,  Red-cockaded 
Woodpeckers  at  the  Savannah  River  Site  were  banded 
with  a U.S.  Fish  and  Wildlife  Service  aluminum  leg 
band  and  with  a unique  color  plastic  leg  band  combi- 
nation for  field  identification.  Birds  were  banded  as 
nestlings  on  the  site,  when  first  captured  as  adults,  or 
just  prior  to  release  if  they  were  from  an  offsite  pop- 
ulation. 

Blood  samples  were  taken  from  adults  at  the  Savan- 
nah River  Site  to  determine  the  relatedness  of  individ- 
uals of  unknown  heritage  and  level  of  genetic  hetero- 
zygosity (Stangel  et  al.  1992,  Haig  et  al.  1993).  The 
results  helped  to  provide  the  genealogical  pedigree  and 
verification  of  parentage  needed  to  determine  which 
individuals  should  be  matched  for  mating. 

Individual  translocations  either  provided  a mate  for 
an  established  breeding  bird  (e.g.,  to  replace  a lost 
mate)  or  established  a new  pair  in  unoccupied  territory. 
If  a translocated  bird  remained  in  the  vicinity  of  the 
release  site  for  at  least  30  days,  I regarded  the  release 
as  successful.  Preference  was  given  to  abandoned  clus- 
ters that  were  more  than  1 km  from  other  active  clus- 
ters to  minimize  interference  by  other  Red-cockaded 
Woodpeckers. 

Translocated  individuals  and  pairs  were  introduced 
into  groups  with  an  unpaired,  resident  bird  or  into 
abandoned  clusters.  Trapping,  transportation,  and  re- 
lease followed  the  methods  described  by  DeFazio  and 
coworkers  (1987)  and  Allen  and  coworkers  (1993). 
The  age,  group,  sex,  status  (such  as  helper  or  breeding 
female),  previous  breeding  experience,  distance  from 
the  capture  to  release  site,  and  location  of  capture  site 
were  recorded  for  each  bird  translocated.  Transloca- 
tions involved  moving  an  independent  subadult  (one 
year  or  less  in  age)  or  adult  female  to  a bachelor  male, 
moving  an  unpaired  female  and/or  male  to  unoccupied 
habitat,  moving  a family  unit  (mated  pair  and  nest- 
lings) to  unoccupied  habitat,  and  cross-fostering  nest- 
lings. 

Observations  of  translocated  birds  lasted  approxi- 
mately 8-30  hrs  per  bird  the  week  immediately  fol- 
lowing release  and  were  repeated  at  least  once  per 
week  during  the  breeding  season  and  once  per  month 


during  the  non-breeding  season  to  monitor  each  bird’s 
status.  If  a translocated  bird  could  not  be  relocated,  a 
thorough  search  was  made  in  clusters  within  approxi- 
mately 0.8  km.  For  birds  captured  on  the  site,  the 
search  included  previous  roost  trees  even  if  they  were 
beyond  0.8  km  of  the  release  site. 

Analytical  methods. — How  a bird  responded  to 
translocation  (e.g.,  stayed  at  release  site,  returned 
home,  disappeared),  whether  or  not  it  eventually  re- 
produced in  the  vicinity  of  the  release  site,  and  the 
number  of  fledglings  produced  was  recorded  for  each 
bird.  Because  the  distance  between  the  capture  and  re- 
lease site  was  found  to  influence  the  results,  the  data 
were  examined  separately  for  moves  of  various  dis- 
tances 1 km,  19-23  km,  182-483  km). 

To  evaluate  if  sex  of  the  translocated  bird  affected 
the  outcome  of  a move,  translocation  success  was 
compared  for  all  males  to  all  females,  adult  males  to 
adult  females,  and  subadult  males  to  subadult  females. 
To  determine  if  there  was  a period  of  time  shortly  after 
fledging  when  younger  subadult  females  were  more 
likely  to  remain  at  the  release  site,  the  translocation 
success  of  subadult  females  5—7  months  of  age  was 
compared  to  those  7—12  months  old. 

Distance  between  the  capture  and  release  site  was 
evaluated  by  examining  translocation  success  for  short 
1 km),  moderate  (19—23  km),  and  long  (182-483 
km)  distance  moves.  There  are  two  subpopulations  of 
Red-cockaded  Woodpeckers  at  the  Savannah  River 
Site.  Moves  within  either  subpopulation  were  no  more 
than  7 km.  The  two  subpopulations  are  separated  by 
about  19  km.  Hence,  translocations  onsite  between  the 
two  subpopulations  involved  distances  of  19-23  km. 
All  offsite  populations  were  at  least  182  km  from  the 
Savannah  River  Site.  Moves  from  offsite  were  done 
for  7 of  10  years  between  1986  and  1995.  Capture  sites 
on  the  Savannah  River  Site  were  monitored  to  deter- 
mine if  released  birds  returned  home.  Similar  moni- 
toring was  not  undertaken  at  offsite  populations  be- 
cause they  were  too  far  from  the  Savannah  River  Site 
to  check  routinely. 

To  determine  if  the  presence  of  a resident  male  af- 
fected the  translocation  success  of  a female,  I com- 
pared responses  of  females  who  were  moved  to  clus- 
ters with  a resident  male  (regardless  of  his  reproduc- 
tive experience)  versus  a “co-move"  in  which  a male 
(captured  in  a separate  cluster)  was  translocated  si- 
multaneously with  a female  to  a new  site.  To  evaluate 
the  possible  influence  of  distance  from  the  capture  to 
release  site,  the  translocation  success  for  females 
moved  to  resident  males  and  those  moved  with  a male 
were  compared  with  respect  to  distance. 

The  possible  effect  of  male  status  (breeder,  helper, 
or  inexperienced)  on  female  relocation  success  was  ex- 
amined for  females:  ( I ) moved  to  a resident  male.  (2) 
moved  simultaneously  with  a male,  and  (3)  for  both 
situations  combined.  An  “inexperienced"  male  had  no 
known  experience  as  a breeder  or  helper.  Female  trans- 
location succe.ss  with  respect  to  male  status  was  seg- 
regated further  by  distance  moved. 

All  statistical  comparisons  were  made  using  Fi.sher's 


40 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


Exact  Test  (Sokal  and  Rohlf  1995)  with  the  level  of 
significance  defined  as  P < 0.05  and  executed  with 
SAS  (version  6.12,  Windows  95,  IBM-compatible; 
SAS  Institute  1990).  Fisher’s  Exact  Test  for  contingen- 
cy tables  was  used  because  in  most  cases  cell  frequen- 
cies were  too  small  to  support  standard  y-  tests.  Unless 
otherwise  noted,  the  results  for  the  five  translocated 
nestlings  are  excluded  from  these  comparisons. 

RESULTS 

From  1986  to  1995,  54  Red-cockaded 
Woodpeckers  were  translocated,  at  first  from 
populations  off  the  site,  but  later  from  onsite 
as  their  numbers  increased.  Beginning  in 
1986,  21  birds  were  taken  from  offsite  popu- 
lations. 7 females,  1 male,  and  5 nestlings 
from  the  Francis  Marion  National  Forest  in 
South  Carolina  (about  192  km  away);  5 fe- 
males from  the  Apalachicola  National  Forest 
in  Florida  (483  km  away);  1 female  from  Fort 
Bragg  in  North  Carolina  (266  km  away);  and 
2 females  from  the  Carolina  Sandhills  Nation- 
al Wildlife  Refuge  in  South  Carolina  (182  km 
away).  Offsite  locations  were  selected  because 
they  contained  relatively  large  numbers  of 
Red-cockaded  Woodpeckers.  Thirty-three 
birds  were  translocated  within  the  site  begin- 
ning in  1987. 

Of  the  24  subadult  females,  6 remained  at 
the  release  site  and  bred,  1 died  after  remain- 
ing  more  than  30  days,  6 moved  to  clusters 
near  the  release  site  and  bred  with  nearby 
males,  1 returned  to  the  capture  site,  and  10 
disappeared.  Of  the  10,  4 remained  at  the  re- 
lease site  for  more  than  30  days,  I was  chased 
away  by  other  Red-cockaded  Woodpeckers, 
and  another  reappeared  five  months  later  ap- 
proximately 20  km  away  and  became  the 
breeding  female  in  that  cluster.  Of  9 adult  fe- 
males, 2 remained  at  the  release  site  and  bred, 

5 moved  to  nearby  clusters  and  bred,  1 re- 
turned to  her  original  cluster,  and  1 remained 
at  the  release  site  but  did  not  breed. 

Of  10  subadult  males,  2 stayed  at  the  re- 
lease site  and  bred,  2 moved  to  a nearby  clus- 
ter and  bred,  1 remained  at  the  release  site  for 
four  months  then  disappeared,  4 disappeared 
soon  after  release,  and  1 returned  home  after 
30  days.  Two  adult  males  remained  where  re- 
leased or  close  by  and  bred,  1 disappeared  in 
less  than  two  days,  and  3 returned  home  im- 
mediately. 

Five  nestlings  were  relocated.  The  first 
three  nestlings  were  moved  with  their  parents 


to  the  Savannah  River  Site  from  the  Francis 
Marion  National  Forest  in  1988  and  later  died 
from  parental  neglect  (Allen  et  al.  1993).  The 
other  two  nestlings  were  fostered  in  1987; 
both  successfully  fledged  after  being  placed  in 
a Red-cockaded  Woodpecker  nest.  The  female 
disappeared  after  five  months,  and  the  male 
became  a breeder  in  a nearby  cluster  and 
eventually  produced  two  fledglings. 

Ten  of  the  49  birds  (excluding  the  5 nest- 
lings) that  were  moved  consisted  of  pairs  of 
subadult  females  and  subadult  males  moved 
concurrently.  Overall,  31  of  49  (63.2%)  adults 
and  subadults  remained  at  or  near  the  release 
site  for  at  least  30  days  after  release  and  25 
(51.0%)  eventually  reproduced  (Table  1).  The 
number  of  birds  represented  in  the  various 
combinations  of  moves  segregated  by  age  and 
cluster  status  are  shown  in  Table  1. 

Translocation  was  successful  for  61.8%  of 
subadults  (21  of  34)  and  66.7%  of  adults  (10 
of  15;  Table  2).  There  were  no  significant  dif- 
ferences in  success  measured  either  by  the 
number  that  stayed  or  by  the  number  that  re- 
produced for  adult  males  compared  to  adult 
females  for  short,  moderate,  or  long  distance 
moves  (Table  2;  Fisher’s  Exact  Tests:  all  P > 
0.05).  Nor  was  there  a difference  in  success 
of  subadult  males  compared  to  subadult  fe- 
males for  any  of  the  distance  classes  (Fisher’s 
Exact  Tests:  all  P > 0.05). 

Of  189  fledglings  produced  from  1986- 
1996,  104  (55.0%)  had  at  least  one  parent  that 
had  been  translocated.  The  number  of  fledg- 
lings excludes  the  young  produced  by  birds 
that  were  translocated  but  did  not  remain  in 
the  vicinity  of  the  release  site  to  breed  (Table 
2). 

Translocation  success  of  younger  subadult 
females  (5-7  months  of  age)  did  not  differ 
significantly  from  those  that  were  older  (7-12 
months  of  age;  Fisher’s  Exact  Test:  P > 0.05 
for  all  comparisons).  There  were  no  short  dis- 
tance moves  involving  younger  subadult  fe- 
males. 

Because  sex  and  age  did  not  appear  to  in- 
fluence success  (Table  2),  I pooled  the  data 
and  tested  for  a distance  effect.  Translocated 
birds  were  more  likely  to  stay  with  increasing 
distance  from  their  capture  .site:  25.0%  suc- 
cess for  translocations  less  than  7 km,  71.4% 
for  19-23  km,  and  81.3%  for  182-483  km 
moves.  The  distance  a bird  was  moved  had  a 


h'nmzreh  • RED-COCKADED  WOODPECKER  TRANSL(3CATIONS 


41 


TABLE  1.  Succes.s  of  translocated  Red-cockaded  Woodpeckers  by  age  and  type  of  move 
River  Site  (1986-1995). 

at  the  Savannah 

Translocated 

Number  of  birds 

Stayed  > 30  days 

Reproduced 

Translocated  to  resident  male: 

Adult  female 

5 

5 

5 

Subadult  female 

17 

13 

9 

Translocated  to  unoccupied  cluster: 

Adult  breeding  male 

1 

0 

0 

Subadult  male 

5 

3 

2 

Adult  female 

1 

1 

1 

Adult  male/adult  female-' 

4 

2 

2 

Adult  niale/sLibadult  female*’ 

4 

2 

2 

Subadult  male/subadult  female'- 

10 

4 

4 

Adult  male/adult  female/nestings'*  (family  unit) 

5 

1 

0 

Fostered  nestlings 

2 

2 

1 

Total  adults  and  subadults 

49 

31  (63.2%) 

25  (51.0%) 

Total  including  nestlings 

54 

33  (61.1%) 

26  (48.1%) 

^ In  one  move,  the  male  remained;  in  the  other,  only  the  female  remained. 

In  one  move,  neither  bird  remained;  in  the  other,  both  remained. 

In  three  cases,  neither  the  male  nor  female  remained;  in  two  ca.ses,  only  the  male  remained;  in  three  ca.ses,  only  the  female  remained;  in  two  cases, 
both  birds  remained. 

Only  the  female  remained. 


highly  significant  effect  on  whether  the  bird 
remained  more  than  30  days  (Fisher’s  Exact 
Test:  P = 0.01),  but  was  not  significant  for 
birds  that  eventually  reproduced  (Fisher’s  Ex- 
act Test:  P = 0.12;  Table  3).  Birds  moved  a 
short  distance  were  more  likely  to  return  home 
[41.7%  (n  = 12)  for  short  versus  4.8%  (n  = 
21)  for  moderate  distance  moves;  Fisher’s  Ex- 
act Test:  P = 0.02J.  There  was  no  significant 
difference  in  the  success  rate  of  a bird  moved 
a moderate  versus  a long  distance  (Fisher’s 
Exact  Test:  P > 0.05  stay,  P > 0.05  repro- 
duce, n — 21  and  16,  respectively).  Nor  was 
there  a significant  difference  in  rate  of  return 
for  males  versus  females  moved  a short  (Fish- 
er’s Exact  Test:  P > 0.05)  or  moderate  dis- 
tance (Fisher’s  Exact  Test:  P > 0.05). 

Eighteen  of  22  females  (81.8%)  that  were 
moved  to  resident  males  were  successful 
(stayed),  whereas  5 of  10  females  (50.0%) 
succeeded  that  were  moved  concurrently  with 
a male.  Of  the  10  co-moves,  3 females  re- 
mained after  the  male  left  and  2 males  stayed 
even  though  the  female  departed.  In  two  cas- 
es, both  male  and  female  remained.  In  one  of 
the  three  instances  when  both  members  of  the 
co-move  left,  the  female  left  first  and  in  the 
other  two  cases  it  could  not  be  determined 
which  of  the  birds  was  the  first  to  leave.  Of 
the  six  cases  in  which  the  male  left,  the  female 


remained  behind  in  three  of  them.  Although 
moving  a female  to  a site  where  a male  al- 
ready was  established  was  thought  to  be  ad- 
vantageous, the  success  rate  was  not  signifi- 
cantly different  from  situations  in  which  the 
female  was  moved  simultaneously  with  a 
male  for  either  moderate  (Fisher’s  Exact  Test: 
P > 0.05  for  stay,  P > 0.05  for  reproduce,  n 
= 14)  or  long  distance  moves  (Eisher’s  Exact 
Test:  P > 0.05  for  stay,  P > 0.05  for  repro- 
duce, n = 14).  Nor  was  there  a significant 
difference  in  success  of  females  moved  either 
to  a resident  male  or  with  a male  when  trans- 
locations of  moderate  and  long  distances  were 
combined  (Fisher’s  Exact  Test:  P > 0.05  stay, 
P > 0.05  reproduce,  n = 28).  No  short  dis- 
tance moves  of  a female  to  a resident  male 
were  undertaken. 

Because  female  success  was  not  influenced 
by  whether  the  male  already  was  on  the  re- 
lease site  or  whether  he  was  moved  simulta- 
neously with  her,  these  data  were  pooled.  Fe- 
males had  a success  rate  of  87.5%  (seven  suc- 
cesses in  eight  cases)  if  the  male  involved  was 
an  experienced  breeding  male,  40.0%  (/;  = 5) 
if  he  was  a helper,  and  73.7%  (/?  = 19)  if  the 
male  was  inexperienced.  Because  there  were 
only  four  short  distance  moves  and  none  of 
these  involved  a breeder  male,  the  effect  of 
male  status  could  not  be  assessed  for  females 


42 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


TABLE  2.  Translocation  success  by  sex  and  age  of  Red-cockaded  Woodpeckers  with  respect  to  distance 
moved  at  the  Savannah  River  Site  (1986-1995).^ 


Di.stance  moved 

No.  fledglings 

Sex/age  class 

< 7 km 

19-23  km 

182-483  km 

Total 

produced 

(No./sex-age 

class) 

Adult  females: 

29  (3.2) 

Number  translocated 

1 

3 

5 

9 

Number  stayed  (%) 

0 (0.0%) 

3 (100.0%) 

5 (100.0%) 

8 (88.9%) 

Number  reproduced  (%) 
Adult  males: 

0 (0.0%) 

3 (100.0%) 

4 (80.0%) 

7 (77.8%) 

4 (0.7) 

Number  translocated 

4 

1 

1 

6 

Number  stayed  (%) 

2 (50.0%) 

0 (0.0%) 

0 (0.0%) 

2 (33.3%) 

Number  reproduced  (%) 
Subadult  females: 

2 (50.0%) 

0 (0.0%) 

0 (0.0%) 

2 (33.3%) 

58  (2.4) 

Number  translocated 

3 

1 1 

10 

24 

Number  stayed  (%) 

1 (33.3%) 

7 (63.6%) 

8 (80.0%) 

16  (66.7%) 

Number  reproduced  (%) 
Subadult  males: 

1 (33.3%) 

5 (45.4%) 

6 (60.0%) 

12  (50.0%) 

20  (2.0) 

Number  moved 

4 

6 

0 

10 

Number  stayed  (%) 

0 (0.0%) 

5 (83.3%) 

b 

5 (50.0%) 

Number  reproduced  (%) 

All  females: 

0 (0.0%) 

4 (66.7%) 

b 

4 (40.0%) 

87  (2.6) 

Number  translocated 

4 

14 

15 

33 

Number  stayed  {%) 

1 (25.0%) 

10  (71.4%) 

13  (86.7%) 

24  (72.7%) 

Number  reproduced  (%) 

All  males: 

1 (25.0%) 

8 (57.1%) 

10  (66.7%) 

19  (57.6%) 

24  (1.5) 

Number  translocated 

8 

7 

1 

16 

Number  .stayed  (%) 

2 (25.0%) 

5 (71.4%) 

0 (0.0%) 

7 (43.8%) 

Number  reproduced  (%) 
Nestlings: 

2 (25.0%) 

4 (57.1%) 

0 (0.0%) 

6 (37.5%) 

2 (0.4) 

Number  moved 

b 

b 

5 

5 

Number  stayed  (%) 

b 

b 

2 (40.0%) 

2 (40.0%) 

Number  reproduced  (%) 
Total  includes  nestlings 

Total  excludes  nestlings 

b 

b 

1 (20.0%) 

1 (20.0%) 
54 

49 

104'^ 

102“^ 

“ None  of  the  results  from  Fisher’s  Exact  Tests  was  significant  at  A*  < 0.05. 

h — = not  applicable;  no  tests  of  this  type  were  made. 

fledJtgTprSucedl"  translocated;  total  figure  includes 


moved  a short  distance.  For  all  moves,  there 
was  no  significant  difference  in  female  suc- 
cess when  comparing  breeder,  helper,  or  in- 
experienced males  (Fisher’s  Exact  Test:  all  P 
> 0.05). 

DISCUSSION 

The  first  reported  Red-cockaded  Wood- 
pecker translocations  involved  a 1981  relo- 
cation of  12  birds  from  5 groups  at  the  Fort 
Stewart  Army  Base  to  St.  Catherines  Island, 
both  in  Georgia  (Odum  et  al.  1982).  Five  of 
these  birds  survived  at  least  eight  months  and 
two  produced  one  fledgling  in  1981.  In  1984 
and  1986,  two  pairs  and  one  single  male  were 
moved  from  private  land  to  the  St.  Marks  Na- 


tional Wildlife  Refuge  and  adjacent  Ochlock- 
onee  River  State  Park  in  Florida  in  an  attempt 
to  enhance  the  three  active  groups  at  the  re- 
lease site  (Reinman  1995).  One  female  re- 
mained and  nested  successfully  for  four  con- 
secutive years,  one  male  returned  to  the  cap- 
ture site,  one  male  died,  and  the  fate  of  the 
other  birds  is  unknown.  Other  translocations 
have  been  conducted  to  establish  a group  at  a 
site  occupied  by  a single  bird  (Allen  et  al. 
1993)  and  to  establish  new  groups  (Rudolph 
et  al.  1992,  Allen  et  al.  1993). 

Working  with  data  collated  from  143  Red- 
cockaded  Woodpecker  translocations  under  a 
wide  range  of  circumstances,  Costa  and  Ken- 
nedy (1994)  found  various  definitions  of 


Fnmzreh  • RED-COCKADED  WOODPECKER  TRANSLOCATIONS 


43 


TABLE  3.  Effect  of  distance  between 

capture  and  release 

site  on  number  of  successful 

translocations  of 

Red-cockaded  Woodpeckers  at  the  Savannah  River  Site  (1986- 

1995). 

Distance  moved 

Number  of  birds  < 7 km 

19-23  km 

182-483  km 

Total 

Moved  1 2 

21 

16 

49“ 

Stayed  (%)  3 (25.0%) 

15  (71.4%) 

13  (81.3%) 

31  (63.3%) 

Reproduced  (%)  3 (25.0%) 

12  (57.1%) 

10  (62.5%) 

25  (51.0%) 

Returned  home  (%)  5 (41.7%) 

1 (4.8%) 

b 

6 (7.5%) 

Fisher's  exact  test:  P value 

Distance  moved 

Stayed 

Reproduce 

Returned  home 

All  distances 

0.01 

0.12 

Short  vs  moderate  distance 

0.01 

0.15 

Moderate  vs  long  distance 

Short  vs  moderate  distance 

0.70 

1.00 

0.02 

“ Excludes  nestlings. 
^ — = not  available. 


translocation  success  ranging  from  “interacted 
well”  to  “fledged  young.”  They  noted  suc- 
cessful moves  66%  of  the  time  for  subadult 
females  {n  = 44)  and  58%  of  the  time  for 
adult  females  (n  = 33).  My  study  showed  an 
overall  female  success  rate  of  67%  for  sub- 
adults and  89%  for  adults.  However,  Costa’s 
and  Kennedy’s  results  are  difficult  to  compare 
to  mine  because  they  contain  a variety  of  cri- 
teria for  translocation  success.  Moreover,  in 
my  study  there  was  no  significant  difference 
in  the  success  rate  based  on  age  (subadult  ver- 
sus adult)  for  either  females  or  males  when 
considering  the  distance  of  the  move.  Some  of 
my  comparisons  involve  small  sample  sizes 
and  it  is  possible  that  a larger  data  set  may 
have  revealed  some  significant  differences. 
Additional  work  is  needed  to  explore  more 
fully  any  possible  differences  in  success  rate 
based  on  age  of  the  bird. 

My  study  showed  a greater  tendency  for 
birds  being  moved  a moderate  (19-23  km)  or 
long  (182-483  km)  distance  to  remain  at  the 
release  site  and  reproduce  than  birds  that  were 
moved  short  distances  (<  7 km).  Because 
there  were  no  moves  between  7-19  km,  it  is 
not  known  at  what  distance  the  success  would 
equal  that  of  moves  more  than  19  km.  There- 
fore, at  the  present  time,  it  is  recommended 
that  translocations  involve  distances  of  at  least 
7 km  (preferably  more)  between  the  capture 
and  release  sites  to  discourage  homing  by  the 
birds. 

DeFazio  and  coworkers  (1987),  Hess  and 
Costa  (1995),  and  Reinman  (1995)  suggest 


that  the  most  successful  translocations  of  fe- 
males are  those  in  which  the  release  site  con- 
tains established  single  males — a finding  sup- 
ported by  earlier  translocations  of  16  females 
at  the  Savannah  River  Site  (Allen  et  al.  1993). 
The  success  rate  for  translocations  to  areas 
that  contained  single  established  males  was 
63.2%  for  Costa  and  Kennedy  (1994)  and 
81.0%  for  my  study.  However,  I found  that 
when  the  release  site  contained  a resident 
male,  female  success  was  no  greater  than 
when  a female  was  moved  concurrently  with 
a male  for  moderate  and  long  distances. 

Costa  and  Kennedy  (1994)  recommend  us- 
ing a two  level  standardized  definition  of  suc- 
cess. One  level  reflects  primary  evidence  of 
breeding  (e.g.,  copulation,  etc.)  and  the  other 
that  the  bird  has  become  attached  to  the  site 
(e.g.,  roosting  in  a cluster,  etc.).  For  any  trans- 
location effort  to  succeed,  the  first  major  hur- 
dle is  for  the  bird  to  remain  at  the  release  site. 
In  my  study,  the  presence  of  a translocated 
bird  at  the  release  site  after  30  days  was  con- 
sidered evidence  that  the  bird  had  accepted  the 
site  and  was  likely  to  breed  once  a suitable 
mate  became  available.  Because  disease  and 
predation  may  prevent  some  of  these  birds 
from  surviving  long  enough  to  reproduce,  the 
use  of  breeding  as  the  criterion  of  transloca- 
tion success  may  be  overly  conservative.  If 
producing  at  least  one  fledgling  is  used  to 
measure  translocation  success,  then  51.0%  of 
the  translocated  birds  in  this  study  were  suc- 
cessful. The  success  rate  was  63.2%  if  defined 


44 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  1,  March  1999 


as  the  number  of  birds  remaining  near  the  re- 
lease site  for  at  least  30  days. 

Translocations  at  the  Savannah  River  Site 
have  played  an  instrumental  role  in  restoring 
the  Red-cockaded  Woodpecker  population 
from  4 individuals  in  1985  to  99  individuals 
(56  adults  and  43  young-of-the-year)  and  19 
breeding  pairs  in  1996  (Franzreb  1997).  Clear- 
ly, the  use  of  translocations  as  a management 
tool  has  been  an  integral  part  in  the  recovery 
of  this  nearly  extirpated  population. 

ACKNOWLEDGMENTS 

Thi.s  research  was  funded  by  the  Department  of  En- 
ergy (DOE),  Savannah  River  Site,  and  its  cooperation 
is  gratefully  acknowledged.  G.  Gaines,  J.  Blake,  and 
R.  Hooper  provided  important  discussion  and  com- 
ments. I thank  R.  Conner,  J.  Reinman,  T.  Engstrom, 
M.  Reed,  R Doerr,  B.  Wigley  and  R.  Hooper  for  their 
insightful  reviews.  In  addition,  P.  Jackson  (DOE)  and 
Savannah  River  Natural  Resource  Management  and 
Research  Institute  staff  (especially  J.  Irwin,  J.  Blake, 
E.  LeMaster,  and  W.  Jarvis)  provided  support  through- 
out the  course  of  this  work.  M.  Lennartz  deserves  spe- 
cial credit  for  involvement  through  1990.  I am  grateful 
to  the  numerous  hard-working  research  field  support 
staff,  C.  Dachelet,  D.  Allen,  K.  Laves,  J.  Edwards,  P. 
Johnston,  D.  Ussery,  and  K.  Shinn,  as  well  as  wildlife 
biologists,  foresters,  and  technicians  at  the  donor  pop- 
ulation forests  for  their  outstanding  efforts  on  behalf 
of  this  project.  And  I thank  statistician  W.  Pepper  who 
was  instrumental  in  providing  statistical  data  analysis. 

LITERATURE  CITED 

Allen,  D.  H.  1991.  An  insert  technique  for  construct- 
ing artificial  Red-cockaded  Woodpecker  cavities. 
U.S.D.A.  Southeast.  For.  Res.  Sta.  Gen.  Tech. 
Rep.  SE-73;1-19. 

Allen,  D.  H.,  K.  E.  Franzreb,  and  R.  E.  E Escano. 
1993.  Efficacy  of  translocation  strategies  for  Red- 
cockaded  Woodpeckers.  Wildl.  Soc.  Bull.  21:155- 
159. 

Carter,  J.  H.,  Ill,  J.  R.  Walters,  S.  H.  Everhart, 
AND  P.  D.  Doerr.  1989.  Restrictors  for  Red-cock- 
aded Woodpecker  cavities.  Wildl.  Soc.  Bull.  17- 
68-72. 

Conner,  R.  N.  and  K.  A.  O’Halloran.  1987.  Cavity 
tree  .selection  by  Red-cockaded  Woodpeckers  as 
related  to  growth  dynamics  of  southern  pines. 
Wilson  Bull.  99:398-412. 

Conner.  R.  N.  and  D.  C.  Rudolph.  1989.  Red-cock- 
aded Woodpecker  colony  status  and  trends  on  the 
Angelina,  Davy  Crockett,  and  Sabine  National 
Forests.  U.S.D.A.  .South.  Res.  Sta.  Res.  Pap.  SO- 
250:1-15. 

Co.STA,  R.  AND  R.  E.  E E.SCANO.  1989.  Red-cockaded 
Woodpecker:  status  and  management  in  the  south- 


ern region  in  1986.  U.S.D.A.  For.  Serv.  Tech.  Pub. 
R8-TP  12:1-71. 

Costa,  R.  and  E.  Kennedy.  1994.  Red-cockaded 
Woodpecker  translocations  1989-1994:  state-of- 
our-knowledge.  Pp.  74-81  in  Proc.  Annu.  Conf. 
Am.  Zoo  Aquarium  Assoc.,  Atlanta,  Georgia. 

DeFazio,  j.  T,  Jr.,  M.  A.  Hunnicutt,  M.  R.  Lennartz, 
G.  L.  Chapman,  and  J.  A.  Jackson.  1987.  Red- 
cockaded  Woodpecker  translocation  experiments 
in  South  Carolina.  Proc.  Annu.  Conf.  Southeast. 
Assoc.  Fish  Wildl.  Agen.  41:311-317. 

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.  Pp. 
81—88  in  Red-cockaded  Woodpecker:  recovery, 
ecology  and  management  (D.  L.  Kulhavy,  R.  G. 
Hooper,  and  R.  Costa,  Eds.).  Center  for  Applied 
Studies,  College  of  Forestry,  Stephen  E Austin 
State  Univ.,  Nacogdoches,  Texas. 

Franzreb,  K.  E.  1997.  Success  of  intensive  manage- 
ment of  a critically  imperiled  population  of  Red- 
cockaded  Woodpeckers  in  South  Carolina.  J.  Field 
Omithol.  68:458-470. 

Haig,  S.  M.,  J.  R.  Belthoff,  and  D.  H.  Allen.  1993. 
Examination  of  population  structure  in  Red-cock- 
aded Woodpeckers  using  DNA  profiles.  Evolution 
47:185-194. 

Hess,  C.  A.  and  R.  Costa.  1995.  Augmentation  from 
the  Apalachicola  National  Forest:  the  develop- 
ment of  a new  management  technique.  Pp.  385- 
388  in  Red-cockaded  Woodpecker:  recovery, 
ecology  and  management  (D.  L.  Kulhavy,  R.  G. 
Hooper,  and  R.  Costa,  Eds.).  Center  for  Applied 
Studies,  College  of  Forestry,  Stephen  F.  Austin 
State  Univ.,  Nacogdoches,  Texas. 

Hooper,  R.  G.,  D.  L.  Krusac,  and  D.  L.  Carlson. 
1991.  An  increase  in  a population  of  Red-cock- 
aded Woodpeckers.  Wildl.  Soc.  Bull.  19:277-286. 

Jackson,  J.  A.,  M.  R.  Lennartz,  and  R.  G.  Hooper. 
1979.  Tree  age  and  cavity  initiation  by  Red-cock- 
aded Woodpeckers.  J.  For.  77:102-103. 

Loeb,  S.  C.,  W.  D.  Pepper,  and  A.  T.  Doyle.  1992. 
Habitat  characteristics  of  active  and  abandoned 
Red-cockaded  Woodpecker  colonies.  South.  J. 
Appl.  For.  16:120-125. 

Odum,  R.  R.,  J,  Rappole,  J.  Evans,  D.  Charbonneau, 
AND  D.  Palmer.  1982.  Red-cockaded  Woodpecker 
relocation  experiment  in  coastal  Georgia.  Wildl. 
Soc.  Bull.  10:197-203. 

Reinman,  J.  P 1995.  Status  and  management  of  Red- 
cockaded  Woodpeckers  on  St.  Marks  National 
Wildlife  Refuge  1980-1992.  Pp.  106-111  in  Red- 
cockaded  Woodpecker:  recovery,  ecology  and 
management  (D.  L.  Kulhavy,  R.  G.  Hooper,  and 
R.  Co.sta,  Eds.).  Center  for  Applied  Studies,  Col- 
lege of  Forestry,  Stephen  F.  Austin  State  Univ., 
Nacogdoches,  Texas. 

Rudolph,  D.  C.  and  R.  N.  Conner.  1991.  Cavity  tree 
■selection  by  Red-cockaded  Woodpeckers  in  rela- 
tion to  tree  age.  Wilson  Bull.  103:458-467. 


Franzreh  • RED-COCKADED  WOODPECKER  TRANSLOCATIONS 


45 


Rudolph,  D.  C.,  R.  N.  Conner,  D.  K.  Carrie,  and  R. 
R.  Schaefer.  1992.  Experimental  reintroduction 
of  Red-cockaded  Woodpeckers.  Auk  109:914- 
916. 

SAS  Institute  Inc.  1990.  SAS  procedures  guide,  ver- 
sion 6.  Third  ed.  SAS  Institute,  Cary,  North  Car- 
olina. 

SoKAL,  R.  R.  AND  E J.  Rohlf.  1995.  Biometry.  Third  ed., 
W.  H.  Freeman  and  Co.,  San  Francisco,  California. 
Stangel,  P.  W.,  M.  R.  Lennartz,  and  M.  H.  Smith. 
1992.  Genetic  variation  and  population  structure 


of  Red-cockaded  Woodpeckers.  Conserv.  Biol.  6: 
283-292. 

Steirly,  C.  C.  1957.  Nesting  ecology  of  the  Red-cock- 
aded Woodpecker  in  Virginia.  Atl.  Nat.  12:280-292. 

U.S.  Fish  and  Wildlife  Service.  1985.  Recovery  plan 
for  the  Red-cockaded  Woodpecker.  Region  4,  At- 
lanta, Georgia. 

Workman,  S.  W.  and  K.  W.  McLeod.  1990.  Vegeta- 
tion of  the  Savannah  River  Site:  major  community 
types.  U.S.  Dept,  of  Energy,  Savannah  River  Site, 
SRO-NERP- 19: 1-137. 


Wilson  Bull.,  1 1 1(1).  1999,  pp.  46-55 


BANDING  RETURNS,  ARRIVAL  PATTERN,  AND  SITE-FIDELITY 

OF  WHITE-EYED  VIREOS 

S.  L.  HOPP,'^  A.  KIRBY, 2 and  C.  A.  BOONE^ 


ABSTRACT. — We  present  nine  years  of  return  data  for  individually  color-banded  White-eyed  Vireos  and 
describe  patterns  of  arrival  and  territory  use.  Of  all  opportunities  for  annual  return.  48.3%  of  males  and  50%  of 
females  were  resighted.  Most  males  arrived  between  17  and  30  April,  with  a median  arrival  date  of  24  April, 
while  most  females  arrived  between  21  April  and  1 May  with  a median  arrival  date  of  26  April;  males  arrived 
significantly  earlier  than  females.  Older  males  arrived  significantly  earlier  than  younger,  as  has  been  reported 
for  several  other  species.  The  arrival  dates  for  individual  males  were  consistent  across  years;  an  individual's 
arrival  date  in  one  year  reliably  predicted  its  arrival  date  in  the  next  year.  Thus,  the  timing  of  arrival  co-varied 
with  three  factors:  sex,  age,  and  individual.  Nearly  all  males  remained  faithful  to  previous  territories,  although 
some  shifted  so  that  the  new  territory  overlapped  the  old.  Aspects  of  our  data  and  those  of  others  suggest  our 
return  rates  are  likely  a low  estimate  of  survivorship  for  the  species;  the  actual  survival  rate  is  probably  higher. 
Received  3 May  1998,  accepted  23  Sept.  1998. 


Recent  studies  have  strongly  indicated  that 
populations  of  many  species  of  migratory 
birds  in  North  America  are  declining.  This 
recognition  has  invigorated  research  efforts 
aimed  at  documenting  various  aspects  of  pop- 
ulation dynamics  among  different  species  and 
determining  the  factors  underlying  these  de- 
clines (Terborgh  1980,  Lovejoy  and  Oren 
1981,  Hagan  and  Johnston  1992,  DeGraaf  and 
Rappole  1995,  Rappole  1995,  Sauer  et  al. 
1996).  These  studies  are  useful  in  establishing 
current  population  status  for  various  species, 
for  monitoring  relationships  between  popula- 
tion status  and  ecological  and  demographic 
factors,  and  for  identifying  future  research  and 
conservation  goals.  Several  approaches  have 
been  employed,  including  point  counts  of 
breeding  populations,  mist-netting  of  birds  at 
migration  stopovers,  breeding  bird  surveys 
(BBS),  and  studies  of  banded  populations  of 
birds  on  both  wintering  and  breeding  grounds 
(Askins  et  al.  1990,  Payne  and  Payne  1990, 
Bibby  et  al.  1992). 

Among  the  groups  of  birds  in  decline  are 
migratory  species  in  the  family  Vireonidae 
(Robbins  et  al.  1989).  Two  species  of  vireo 
are  federally  endangered  in  the  United  States: 
the  Black-capped  Vireo  (Vireo  atricapillm) 
found  in  Texas  and  Oklahoma  (Grzybowski  et 


' Dept,  of  Ecology  and  Evolutionary  Biology,  Univ. 
of  Arizona.  Tuc.son.  AZ  85721. 

^ P.O.  Box  193,  Independence,  OR  97351. 

' Mu.seum  of  the  Middle  Appalachians,  123  Palmer 
Avc.,  .Saltville,  VA,  24370. 

^ Correspt)nding  author;  E-mtiil:  SHOPP@u.tirizona.cdu 


al.  1986,  USFWS  1991),  and  a race  of  Bell’s 
Vireo  in  California,  the  Least  Bell’s  Vireo  (V. 
bellii  pusillus;  USFWS  1986,  Franzreb  1989). 
Several  other  species  of  vireos  have  declined 
in  numbers,  as  determined  by  the  USGS 
Breeding  Bird  Survey  (BBS;  Robbins  et  al. 
1989,  Sauer  and  Droege  1992).  A number  of 
studies  have  addressed  different  aspects  of 
population  dynamics  in  various  vireo  species 
including  migration  patterns  (Remsen  et  al. 
1996,  Woodrey  and  Chandler  1997,  Woodrey 
and  Moore  1997),  population  structure  (Grzy- 
bowski 1991),  aspects  of  breeding  (Graber  et 
al.  1985,  Grzybowski  et  al.  1994,  Marvil  and 
Cruz  1989,  Barber  and  Martin  1997),  and  win- 
tering ecology  (Greenberg  1992;  Greenberg  et 
al.  1993,  1995).  Studies  of  these  types  are  im- 
portant for  gauging  long-term  changes  in  mi- 
gratory vireo  populations  and  the  factors  that 
affect  them  (Holmes  et  al.  1989;  Lynch  1989 
1992). 

One  of  these  Nearctic  migrant  species  is  the 
White-eyed  Vireo  (Vireo  griseus),  a small 
passerine  that  occupies  secondary  deciduous 
habitat,  thickets,  and  forest-edge  in  the  eastern 
United  States.  Its  winter  range  extends  across 
the  southern  US  from  Texas  to  South  Caroli- 
na, south  through  the  West  Indies,  and  along 
the  eastern  coast  of  Mexico  (Barlow  1980, 
Hopp  et  al.  1995).  The  northern  subspecies, 
V.  g.  noveboracemis,  is  fully  migratory.  The 
southeastern  subspecies,  V.  g.  griseus,  has 
been  reported  to  migrate  (Barlow  1980).  How- 
ever, Bradley  (1981)  reported  that  a popula- 
tion near  Gainesville,  Florida  was  sedentary, 


46 


Hopp  el  ill.  • VIREO  RETURNS 


47 


with  individuals  remaining  through  the  winter. 
The  remaining  subspecies,  V.  g.  micrus,  V.  g. 
bermudianus,  V.  g.  maynardi,  and  V.  g.  per- 
quisitor,  are  apparently  nonmigratory  (Barlow 
1980,  Hopp  et  al.  1995).  Analyses  of  the  data 
from  the  BBS  suggest  that  White-eyed  Vireos 
have  been  declining  in  parts  of  their  breeding 
range  (Robbins  et  al.  1989,  Sauer  and  Droege 
1992).  Efforts  to  document  these  declines  and 
determine  factors  that  influence  population 
structure  and  territory  use  are  important  for 
monitoring  long-term  changes  in  vireo  popu- 
lations. 

In  the  present  study  we  report  on  returns  of 
banded  White-eyed  Vireos  for  nine  years  in 
southwestern  Virginia.  Banded  White-eyed 
Vireos  have  been  reported  to  show  site  fidelity 
in  their  breeding  range  (Hopp  et  al.  1995)  and 
on  winter  territories  (Rappole  and  Warner 
1980).  We  used  this  fidelity  to  territories  to 
directly  measure  return  rates  of  banded  indi- 
viduals. In  addition  we  report  on  the  pattern 
of  arrival  in  the  spring  and  outline  the  use  of 
territories  by  individuals  during  the  breeding 
season. 

METHODS 

Most  birds  were  banded  on  two  study  sites  in  Wash- 
ington and  Smyth  counties  in  southwestern  Virginia. 
The  two  sites,  each  about  35  ha  and  located  600-800 
m above  sea  level  comprised  tracts  of  secondary  de- 
ciduous growth,  typically  favored  by  this  species  (Con- 
ner et  al.  1983,  Graber  et  al.  1985,  Hopp  et  al.  1995). 
Both  areas  contained  12-16  contiguous  White-eyed 
Vireo  territories.  Other  birds  were  banded  in  the  same 
counties  at  smaller  tracts  of  3—12  ha,  each  with  1-6 
territories;  2 of  these  smaller  areas  were  within  5 km 
of  the  main  study  area  in  Washington  County.  For  all 
of  these  study  areas  the  habitat  type  was  mixed,  with 
approximately  60-80%  of  the  areas  constituting  suit- 
able habitat  for  the  vireos.  We  attempted  to  locate  and 
band  birds  at  all  smaller  appropriate  tracts  within  about 
15  km  of  the  site  in  Washington  County.  The  principal 
habitat  types  in  this  region  are  deciduous  forest  or 
open  pastureland;  the  presence  of  habitat  appropriate 
for  White-eyed  Vireos  is  limited  and  typically  exists 
in  small  areas  supporting  only  a few  individuals.  All 
but  one  male  in  our  study  shared  at  least  one  territorial 
boundary  with  another  male;  most  birds  shared  bound- 
aries with  several  other  males. 

We  captured  most  birds  soon  after  their  arrival  in 
late  April  or  early  May.  Males  were  easily  taken  in 
mist  nets  as  they  approached  tape  playbacks  of  con- 
specific  song.  Our  attempts  to  capture  females  by  in- 
tercepting approaches  to  nests  clearly  disrupted  nesting 
activities,  so  we  discontinued  those  attempts.  On  a few 
occasions  females  were  captured  with  males,  if  they 


followed  them  into  nets  during  playbacks.  All  birds 
were  banded  under  permit  with  a unique  configuration 
of  USFWS  aluminum  and  plastic  color  bands. 

We  began  checking  For  arrivals  in  known  territories 
and  neighboring  suitable  habitat  in  the  first  week  ol' 
April.  Monitoring  consisted  of  listening  for  singing 
males  in  known  territories  for  at  least  15  minutes;  often 
we  used  recorded  song  to  attract  males.  When  detect- 
ed, males  were  followed  to  determine  whether  they 
were  banded,  and  to  identify  the  color(s)  and  config- 
uration for  banded  individuals.  Study  areas  were 
checked  daily  until  mid-May;  unoccupied  territories 
were  then  checked  at  least  twice  weekly  until  mid- 
June.  We  also  checked  several  other  known  popula- 
tions of  birds  within  15  km  of  the  study  areas  to  po- 
tentially detect  dispersed  birds.  We  were  able  to  as.sess 
returns  without  recapture  by  identifying  the  color-band 
configurations.  Because  the  males  of  this  species  are 
vocally  prolific,  it  was  easy  to  locate  newly  arrived 
males.  The  arrival  dates  used  in  this  analysis  are  re- 
stricted to  subjects  for  which  we  spent  at  least  1 5 min- 
utes in  the  area  on  the  day  prior  to  their  first  detection, 
i.e.,  we  were  confident  they  were  not  present  on  the 
previous  day.  Determining  arrival  dates  for  females 
was  more  difficult  because  they  are  behaviorally  cryp- 
tic, and  because  they  sometimes  changed  locations  af- 
ter a day  or  two.  Most  reliably,  finding  males  allowed 
us  to  locate  females.  Whenever  a male  was  located, 
we  observed  him  long  enough  to  determine  whether  a 
female  was  also  present.  In  most  cases,  changes  in  the 
males’  behavior  was  indicative  of  pairing  status;  un- 
paired males  usually  sang  at  high  rates  from  high,  ex- 
posed perches,  while  paired  males  typically  spent  more 
time  in  lower  areas  and  sang  at  lower  rates  (Hopp  et 
al.  1995).  Because  newly  paired  males  stay  close  to 
females  it  was  relatively  easy  to  assess  pairing  status. 
Because  females  were  mostly  unbanded,  determining 
whether  a particular  female  was  a new  arrival  or  had 
moved  from  a neighboring  male’s  territory  was  not 
possible.  Arrival  dates  for  females  could  be  unambig- 
uously determined  by  assessing  the  total  number  of 
females  on  the  entire  study  site,  with  changes  between 
successive  days  indicating  new  arrivals.  Arrival  dates 
for  females  used  in  this  analysis  are  only  those  we 
could  definitively  determine  to  be  new  arrivals. 

We  considered  an  individual’s  tenitory  to  be  the  to- 
tal area  the  bird  was  observed  to  occupy  throughout 
the  course  of  a season  without  outside  influence  either 
by  other  birds  or  by  the  researchers.  Site  fidelity  was 
then  defined  as  use  of  an  area  that  overlapped  a terri- 
tory from  the  previous  year  by  at  least  50%.  Approx- 
imately two-thirds  of  the  males  also  served  as  focal 
subjects  for  studies  of  vocal  communication,  and  hence 
provided  more  complete  data  on  territory  use. 

We  also  examined  the  encounter  records  for  White- 
eyed Vireos  from  the  Bird  Banding  Laboratory  (BBL) 
through  1996.  These  records  include  encounters  within 
the  same  10'  block  (prior  to  1958)  and  document  in- 
dividuals encountered  outside  the  original  10'  block 
where  they  were  banded.  These  records  potentially 
provide  information  about  the  dispersal  of  both  hatch- 


48 


THE  WILSON  BULLETIN 


Vol.  Ill,  No.  I.  March  1999 


TABLE  1.  Returns  of  male  White-eyed  Vireos 
known  to  be  at  least  one  year  old  when  banded. 

Banding 

year 

Number 

banded 

Returns  i 

in  years  following  banding 

One 

Two 

Three 

Four 

Five 

Six 

1985 

8 

3 

2 

2 

1 

0 



1986 

4 

3 

2 

1 

1 

1 

0 

1987 

1 1 

7 

4 

1 

1 

1 

0 

1988 

1 I 

5 

3 

0 

— 

— 

— 

1989 

14 

10 

1 

1 

0 

— 

— 

1990 

13 

8 

6 

0 

— 

— 

— 

1991 

13 

4 

1 

0 

— 

— 

— 

Total 

74 

40 

19 

5 

3 

2 

0 

Percent 

54 

26 

7 

4 

3 

0 

^ Calculations  based  on  percent  returns  relative  to  banding  year. 


ing  year  (HY)  and  adult  (after  hatching  year;  AHY) 
birds.  To  assure  that  we  were  assessing  dispersal  rather 
than  migratory  movements,  we  considered  only  birds 
with  both  banding  and  encounter  dates  between  16 
May  and  16  September  inclusive.  These  dates  are  well 
within  the  average  spring  arrival  and  fall  departure 
dates  for  this  species  throughout  most  of  its  range  (see 
appendix  1 in  Hopp  et  al.  1995). 

RESULTS 

Returns  of  banded  birds. — During  the 
breeding  seasons  of  1985-1991,  we  banded 
74  adult  male  birds  used  in  this  analysis.  We 
also  banded  5 females  and  42  nestlings/fledg- 
lings, most  of  which  were  known  to  have  sur- 
vived until  they  were  capable  of  sustained 
flight.  Returns  of  males  for  each  year  from 
1986—1994  individually  and  combined  are 
given  in  Table  1.  The  percentage  returns  are 
based  on  combined  data  from  nine  years, 
computing  the  number  that  return  relative  to 
the  banding  year.  Because  these  subjects  were 
all  banded  as  adults  (AHY),  these  percentages 
give  an  indication  of  survivorship  in  the  suc- 
cessive years  following  banding.  The  two  in- 
dividuals surviving  five  seasons  after  banding 
were  thus  at  least  six  years  old.  Our  oldest 
bird  (not  included  in  this  analysis)  arrived  as 
an  AHY  bird  in  1990  and  was  still  alive  in 
the  summer  of  1998,  making  him  at  least  9 
years  old. 

Calculating  the  probability  of  return  given 
that  a bird  was  alive  in  the  previous  year,  40 
of  74  (54.1%)  returned  the  year  following 
banding,  19  of  40  (47.5%)  returned  the  second 
year,  5 of  19  (26.3%)  returned  the  third  year, 
3 of  5 (60.0%)  returned  the  fourth,  2 of  3 
(66.7%)  returned  the  fifth  year,  and  none  of  2 


10  20  30  10  20  30  9 

April  May  June 

Date 


FIG.  1.  Box  plot  of  arrival  dates  for  all  females 
and  males.  Males  are  further  divided  into  various  age 
clas.ses  (see  text  for  explanations).  For  each  plot  the 
total  box  encompasses  the  second  and  third  quartiles 
for  each  distribution,  with  the  central  vertical  line 
showing  the  median  arrival  date.  Horizontal  lines  show 
the  1 0—90%  range  of  dates,  and  triangles  indicate  in- 
dividual arrival  dates  in  the  outlying  10%  ranges. 

returned  the  sixth  year  following  banding. 
Over  the  nine  year  period,  there  were  143  op- 
portunities for  return  and  69  documented  re- 
turns. Thus  the  percentage  of  male  White- 
eyed Vireos  returning  in  any  year,  given  the 
bird  was  known  to  be  alive  the  previous  year, 
was  48.3%.  Of  the  five  females  banded,  three 
returned  in  the  year  following  banding  and 
one  individual  returned  in  the  second  and  third 
years  following  banding.  Of  the  ten  opportu- 
nities for  female  returns,  five  (50.0%)  were 
resighted.  We  did  not  recover  any  of  the  birds 
banded  as  nestlings  or  fledglings. 

Male  arrival  pattern. — We  obtained  90  ar- 
rival dates  from  63  individuals.  The  combined 
dates  ranged  from  12  April  to  11  June  with  a 
median  arrival  date  of  23  April  (Fig.  1).  Most 
birds  arrived  during  the  two  weeks  between 
16  April  and  30  April  (75  of  90:  83.3  %). 
There  were  six  arrival  dates  prior  to  1 6 April, 
with  three  of  these  from  the  same  banded  in- 
dividual in  successive  years.  Eight  arrival 
dates  were  after  30  April,  five  ol  these  in  early 
May  (2  May,  and  two  each  on  5 May  and  6 
May)  and  three  were  well  outside  the  distri- 


Hopp  et  al.  • VIREO  RETURNS 


49 


bution  of  other  arrivals — 28,  29  May  and  1 1 
June.  These  three  extremely  late  unhanded  in- 
dividuals likely  represent  relocations  rather 
than  arrival  dates  (see  below). 

Female  arrival  pattern. — We  obtained  37 
arrival  dates  for  females.  Of  these  only  4 were 
from  returning,  banded  individuals,  the  re- 
mainder from  unbanded  individuals.  The  com- 
bined dates  ranged  from  17  April  to  9 May, 
with  a median  arrival  date  of  26  April  (Fig. 
1).  Most  of  the  arrival  dates  fell  in  the  ten  day 
period  between  21  April  and  1 May  (28  of  37: 
75.7%).  Two  of  the  four  earliest  arrival  dates 
(17  and  20  April)  were  from  returning  banded 
individuals.  Because  females  are  cryptic  in 
their  plumage  and  behavior  and  because  we 
assessed  their  arrival  primarily  by  pairing  with 
males,  our  sampling  was  likely  biased  toward 
the  earlier  arrival  dates.  Despite  this,  the  me- 
dian arrival  date  for  females  was  later  than  all 
of  the  individual  male  arrival  date  categories 
(Mann- Whitney  Test:  U = 3.43,  P < 0.001; 
see  Fig.  1).  We  typically  observed  several 
males  on  territory  before  any  females  arrived. 
In  several  years,  most  male  territories  were 
occupied  before  any  females  were  seen. 

Age-related  male  returns. — We  recorded  49 
arrival  dates  from  returning  banded  males  and 
41  from  unbanded  males  that  were  later  band- 
ed. Of  the  unbanded  birds,  21  were  of  un- 
known age.  The  remaining  20  arrival  dates 
were  from  unbanded  birds  that  occupied  ter- 
ritories previously  occupied  by  non-returning 
banded  individuals.  Given  the  high  site  fidel- 
ity for  returning  individuals,  this  second  cat- 
egory of  unbanded  replacement  birds  possibly 
were  second-year  adults  (SY;  first  breeding 
season)  which  were  treated  separately  in  our 
analysis  of  age-related  returns. 

Figure  1 shows  the  distributions  of  arrival 
dates  for  various  age-related  classes  of  males. 
Because  the  sample  sizes  in  the  three  oldest 
age  categories  were  too  small  to  permit  statis- 
tical comparison  across  all  age  categories,  we 
combined  all  dates  of  birds  returning  after 
three  years  into  one  category  (four  years  or 
older,  n = 8).  To  avoid  pseudoreplication,  in 
analysis  we  averaged  dates  for  individuals  that 
contributed  more  than  one  arrival  date  to  this 
combined  category,  yielding  an  effective  sam- 
ple size  of  5.  A comparison  of  the  four  male 
groups,  previously  unbanded  birds,  first  re- 
turns, second  returns,  and  third  + returns,  was 


significant  (Kruskal-Wallis  Test:  H = 1 1.27,  P 
< 0.05).  In  post  hoc  (Bonferroni)  pairwise 
comparisons,  the  only  significant  difference 
was  between  the  new  (unbanded)  arrivals  and 
the  first  return  year  (Z  = 2.77,  P < 0.05), 
indicating  that  the  gain  in  arrival  is  between 
the  first  (banding)  year  and  the  first  return 
year.  In  a more  conservative  version  of  this 
comparison,  we  compared  all  returning,  band- 
ed birds  to  all  unbanded  birds.  This  compar- 
ison was  also  significant,  showing  arrival 
dates  of  all  banded  birds  to  be  earlier  than 
arrival  dates  of  all  unbanded  birds  (Mann- 
Whitney  Test:  U = 3.66,  P < 0.001).  In  a 
more  direct  test,  we  compared  successive 
dates  from  individuals  for  whom  we  obtained 
arrival  dates  in  two  consecutive  years  (n  = 32 
pairs).  This  comparison  was  only  marginally 
significant  when  using  all  consecutive  arrival 
pairs  (Wilcoxon  Test:  Z = 1.465,  P = 0.074). 
When  this  analysis  was  restricted  to  consec- 
utive-year pairs  starting  with  the  first  (un- 
banded) year,  the  comparison  was  significant 
(Wilcoxon  Test:  Z = 1.80,  P < 0.05,  n — 14), 
showing  that  birds  arrived  earlier  in  the  sec- 
ond of  these  two  consecutive  years,  and  re- 
inforcing the  finding  that  the  gain  in  arrival 
dates  is  apparent  only  between  the  banding 
and  subsequent  year.  For  the  32  pairs  of  con- 
secutive year  arrival  dates,  we  also  found  that 
the  arrival  date  of  individuals  was  a significant 
predictor  of  its  arrival  date  in  the  subsequent 
year  (r  = +0.616,  P < 0.001).  Thus,  while  a 
portion  of  the  variability  seen  in  arrival  dates 
can  be  attributed  to  age,  a substantial  portion 
can  be  attributed  to  individual-specific  differ- 
ences in  arrival,  with  early  and  late  arrivers 
remaining  early  and  late  arrivers  respectively 
across  seasons  (Fig.  2). 

Site  fidelity. — Of  the  returning  banded 
males,  67  of  69,  or  97%  of  returns  were  to 
their  previous  territory.  The  two  individuals 
observed  to  move  to  non-overlapping  territo- 
ries between  years  both  remained  within  800 
m of  their  original  territory.  Several  birds  en- 
larged their  territories  in  subsequent  years,  oc- 
cupying areas  that  included  their  previous  ter- 
ritory. Seven  individuals  disappeared  during 
the  course  of  a season;  three  returning  birds 
and  four  birds  in  the  year  they  were  banded. 
None  of  these  seven  birds  was  seen  in  sub- 
sequent years.  One  of  the  three  returning  birds 
lost  his  territory  to  human  habitat  clearing 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  1,  March  1999 


50 


Individual  Return  Dates  in  Successive  Years 


FIG.  2.  Scatter  plot  showing  return  dates  obtained 
from  individual  males  in  successive  years.  Line  depicts 
best-fitting  regression  line;  squares  denote  2 overlap- 
ping data  points. 


during  the  winter.  He  was  seen  in  adjacent  and 
nearby  territories  for  only  five  days  following 
his  arrival.  Two  of  the  four  birds  that  disap- 
peared during  the  year  they  were  banded  re- 
appeared later  in  the  season.  One  of  these  ar- 
rived on  22  April  in  a territory  previously  oc- 
cupied by  a non-retuming  banded  individual. 
He  was  banded  on  24  April,  disappeared  on  2 
May,  then  reappeared  on  the  same  territory  7 
June  where  he  remained  for  the  season. 

For  females,  all  returns  were  to  the  same 
study  area  where  they  were  first  banded,  but 
none  to  the  same  territory;  none  paired  with 
the  same  male  in  more  than  one  year.  The 
nearest  movement  was  to  an  adjacent  territory. 
The  farthest  moved  three  territories  away,  i.e., 
two  intervening  territories  from  the  previous 
year,  a distance  of  about  750  meters. 

Territory  use. — Territory  use  in  this  species 
is  a typical  “type  A”  territory  (Nice  1941), 
with  males  aggressively  defending  territories 
against  other  males,  and  with  the  area  used 
for  mating,  nesting,  and  feeding  for  both 
adults  and  young  during  the  breeding  season. 
The  defense  of  territory  is  usually  by  a series 
of  behavioral  displays,  including  singing,  and 
less  frequently  by  direct  contact  (see  descrip- 
tion by  Bradley  1980).  Most  territorial  en- 


counters occurred  early  in  the  season,  prior  to 
the  arrival  and  pairing  with  females;  following 
the  onset  of  nesting  activities  these  territorial 
encounters  were  rare. 

Individual  males  could  easily  be  tracked 
during  the  season,  and  adults  remained  on  ter- 
ritory into  September;  the  latest  date  we  re- 
corded a bird  on  territory  in  autumn  was  6 
October.  During  the  last  few  weeks  in  the  fall 
the  adults  increased  their  singing  rate  and 
again  became  quite  responsive  to  tape-record- 
ed song. 

Female  use  of  territory  was  less  easy  to  ob- 
serve. On  several  occasions  we  observed  two 
females  within  the  same  territory,  always  ear- 
ly in  a season.  In  no  case  did  we  observe  in- 
teractions between  females  indicating  territory 
defense.  On  one  occasion  we  observed  two 
females  within  a few  centimeters  of  each  other 
with  no  overt  behavioral  response. 

As  with  females,  HY  birds  were  difficult  to 
track  carefully  during  a season.  Typically, 
banded  young  could  be  found  in  their  natal 
territories  for  4-5  weeks;  often  birds  from  the 
same  brood  were  found  together.  After  this 
time,  banded  HY  birds  began  to  disperse  from 
their  natal  territories  mid-August  through  ear- 
ly September,  often  being  found  in  other  ter- 
ritories within  the  study  site.  A few  HY  birds 
would  defend  small  territories,  responding  ag- 
gressively to  tape-recordings  with  approach 
and/or  singing.  Most  HY  birds  left  the  region 
on  migration  prior  to  the  onset  of  banded 
AHY  bird  departures. 

USFWS  encounter  data. — There  were  81 
total  White-eyed  Vireo  encounter  records 
from  the  Bird  Banding  Laboratory.  Of  these, 
22  were  both  banded  and  encountered  within 
the  16  May  to  16  September  period,  all  of 
these  encountered  during  the  breeding  season 
in  the  first  year  after  banding.  Of  these  22,  18 
were  AHY  birds,  three  were  HY,  and  one  was 
local  status  (incapable  of  sustained  flight;  pre- 
sumably nestling).  Of  the  18  adults,  14  were 
encountered  in  the  same  10'  block  in  which 
they  were  banded.  Of  the  four  adults  encoun- 
tered outside  their  initial  banding  10'  block, 
three  were  encountered  in  the  next  10'  block 
of  latitude  (approximately  19  km  on  average) 
and  one  was  encountered  in  the  adjacent  10' 
block  of  longitude  (approximately  15  km).  Of 
the  three  HY  birds,  one  was  encountered  in 
the  original  block,  and  the  other  two  were  en- 


Hopp  et  ill.  • VIREO  RETURNS 


51 


countered  one  10'  block  of  latitude  and  one 
10'  block  of  longitude  respectively  from 
where  originally  banded.  The  local-status  bird 
was  encountered  in  the  same  10'  block  where 
it  was  initially  banded.  Two  other  records  de- 
serve mention.  One  HY  bird  was  banded  on 
17  August  and  encountered  approximately 
155  km  away  in  the  following  year  on  14 
May.  The  other,  a local  bird,  was  banded  on 
12  June  and  encountered  the  following  13 
May,  approximately  190  km  from  initial  band- 
ing location.  As  both  of  these  birds  were 
banded  and  encountered  in  states  near  the 
northern  edge  of  the  species’  range  (Massa- 
chusetts and  Maryland  respectively),  these  en- 
counters in  the  middle  part  of  May  likely  rep- 
resent an  encounter  on  their  breeding  ground 
rather  than  a point  in  their  migration  paths. 

DISCUSSION 

Returns  of  banded  birds. — Our  overall  per- 
centage returns  of  48.3%  for  males  and  50% 
for  females  gives  an  indication  of  the  baseline 
returns  expected  for  this  species.  These  can  be 
compared  to  return  rates  reported  for  Black- 
capped  and  Bell’s  vireos.  In  a long-term  study 
of  Black-capped  Vireos,  Grzybowski  (1991) 
reported  male  returns  of  65%  for  a study  pop- 
ulation of  approximately  250  birds.  Perhaps 
more  directly  comparable  were  his  return  rates 
for  approximately  130  males  from  scattered 
study  locations.  For  these  he  reported  a return 
rate  of  58%,  slightly  higher  than  in  our  study. 
For  females,  he  reported  52%  and  41%  en- 
countered for  larger  and  more  scattered  study 
groups,  respectively.  Similarly,  for  Bell’s  Vir- 
eos, Greaves  and  Labinger  (1998)  reported 
male  returns  of  62.5%  and  58.6%  for  two  sep- 
arate study  areas,  and  57.1%  and  58.8%  for 
females  from  the  same  two  study  groups.  Two 
notable  differences  between  the  studies  of 
both  these  species  and  our  study  are  the  sam- 
ple sizes  and  the  study  areas.  Grzybowski 
(1991)  reported  his  percentage  returns  on 
samples  of  approximately  250  and  130  males 
in  the  main  and  scattered  study  areas.  The 
study  by  Greaves  and  Labinger  (1998)  re- 
ported on  returns  from  a smaller  number  ot 
males:  approximately  40  and  30.  The  study 
areas,  however,  were  considerably  larger  than 
in  our  study.  Many  researchers  have  noted  th.it 
the  percentages  of  birds  re-encountered  in 
banding  studies  is  directly  related  to  the  size 


of  the  study  population,  with  larger  popula- 
tions affording  a higher  re-encounter  percent- 
age, and  that  site  fidelity  is  greater  in  larger 
study  groups  (e.g..  Temple  and  Cary  1988, 
Payne  and  Payne  1990,  Grzybowski  1995). 
Applied  to  this  study,  then,  our  reported  return 
rates  likely  represent  a low  return  estimate  for 
the  species.  The  data  presented  here  provide 
a baseline  rate  of  returns  for  adult  White-eyed 
Vireos.  Because  the  population  densities  in 
this  region  are  relatively  low  (Price  et  al. 
1995),  comparative  studies  in  other  areas  of 
this  species’  range  are  needed  to  determine 
which  aspects  of  the  data  observed  here  are 
shared  in  other  regions,  and  what  factors 
might  affect  return  rates  and  survival  of 
White-eyed  Vireos. 

Arrival  patterns. — Our  finding  that  older 
birds  arrive  sooner  than  younger  birds  in 
spring  is  consistent  with  similar  reports  in  oth- 
er species  (e.g.,  Nolan  1978,  Bedard  and 
LaPointe  1984,  Hill  1989,  Morton  and  Der- 
rickson  1990).  An  earlier  arrival  could  provide 
either  a longer  potential  breeding  season  or 
better  chances  of  obtaining  a mate  (see  Mpller 
1990).  However,  the  variation  we  saw  was 
modest;  the  only  significant  difference  was 
approximately  2 days  gained  between  the  first 
(banding)  year  and  the  first  return  year,  a pat- 
tern seen  with  both  the  between-subject  and 
within-subject  comparisons.  It’s  difficult  to  ar- 
gue strongly  for  a significant  pairing  advan- 
tage, particularly  because  most  males  arrive 
before  most  females.  We  have  found  that  the 
arrival  date  itself  is  not  a significant  predictor 
of  either  pairing  date  or  probability  of  obtain- 
ing a mate.  Rather,  the  age  of  the  individual 
is  more  likely  a factor,  i.e.,  females  are  more 
likely  to  pair  with  older  (returning)  males,  re- 
gardless of  aiTival  date  (SLH,  unpubl.  data). 
However,  the  relation  between  arrival  date  and 
reproductive  success  in  White-eyed  Vireos  is 
unknown. 

Several  studies  have  identified  reasons  for 
earlier  arrival  dates  by  older  birds,  and  several 
of  these  might  pertain  to  White-eyed  Vireos 
(see  Ketterson  and  Nolan  1983,  Woodrey  and 
Chandler  1997  for  summaries).  First,  older 
birds  might  winter  further  north  than  younger 
birds.  However,  the  site  fidelity  of  White-eyed 
Vireos  to  winter  territories  (Rappole  and 
Warner  1980)  argues  against  this.  Second,  old- 
er birds  might  leave  earlier  for  northern  mi- 


52  THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


gration  in  the  spring.  Third,  older  birds  might 
be  better  at  finding  their  territory  once  they 
have  arrived  on  the  breeding  grounds.  Finally, 
it  is  possible  that  older  birds  travel  faster  dur- 
ing spring  migration,  either  through  superior 
navigational  skills  or  more  efficient  foraging 
while  enroute. 

Our  finding  that  the  arrival  date  for  an  in- 
dividual in  a given  year  predicted  its  arrival 
in  the  subsequent  year  is  a pattern  that  to  our 
knowledge  has  not  been  previously  reported 
for  any  species.  The  variation  seen  as  a result 
of  individual  differences  accounts  for  more  of 
the  arrival  date  variability  than  that  seen  for 
age  related  return  dates.  The  factors  outlined 
above  for  age  related  returns  may  also  be  used 
to  explain  this  result,  but  with  individual  rath- 
er than  age  related  differences  in  these  abili- 
ties being  applied  to  individuals  initiating 
spring  migration  from  a common  wintering 
location.  Alternatively,  this  pattern  could  re- 
sult from  individuals  commencing  from  a 
wide  range  of  geographic  origins,  with  more 
northerly-wintering  individuals  arriving  earli- 
er. The  relationship  between  the  winter  and 
breeding  locations  of  individuals  is  unknown. 

Philopatry  and  territory  use. — Male  White- 
eyed Vireos  show  a high  degree  of  territory 
fidelity.  Some  researchers  have  distinguished 
between  site  fidelity,  where  birds  return  to  the 
study  area  but  to  a different  territory,  and  ter- 
ritory fidelity,  where  birds  return  to  their  pre- 
vious territory  (Greenwood  1980,  Greenwood 
and  Harvey  1982,  Holmes  and  Sherry  1992). 
The  extent  to  which  birds  move  between  ter- 
ritories in  consecutive  years  has  been  called 
breeding  dispersal  by  Greenwood  (1980).  In 
our  study,  the  attachment  to  particular  terri- 
tories is  remarkably  high  and  the  two  tenden- 
cies, site  and  territory  fidelity,  appear  essen- 
tially the  same.  While  perhaps  uncommon, 
three  lines  of  evidence  may  indicate  that 
breeding  dispersal  movements  occur  in  White- 
eyed Vireos.  First,  on  two  occasions  we  ob- 
■served  individuals  disappear  only  to  reappear 
in  the  same  territory  later  within  the  same  sea- 
son. While  we  did  not  locate  these  birds  on 
other  study  areas,  they  obviously  relocated  for 
at  least  a portion  of  that  season.  Grzybowski 
(1995)  reported  that  male  Black-capped  Vir- 
eos  sometimes  sequentially  occupied  two  non- 
contiguous territories,  and  this  might  also  oc- 
cur in  White-eyed  Vireos.  Second,  records 


from  the  BBL  show  four  relocations  of  adult 
White-eyed  Vireos  between  breeding  seasons, 
with  distances  of  between  15  and  20  km  for 
each.  While  these  records  are  limited,  they 
nevertheless  show  that  adult  White-eyed  Vir- 
eos are  known  to  relocate  between  breeding 
seasons.  Third,  we  observed  three  arrival 
dates  of  individual  males  in  late  May  and  ear- 
ly June,  well  outside  of  the  migration  dates 
known  for  this  species  (Hopp  et  al.  1995, 
Remsen  et  al.  1996).  While  circumstantial, 
these  late  dates  likely  reflect  instances  of  re- 
locating individuals  rather  than  first-time  ar- 
rivals. Taken  together,  these  lines  of  evidence 
suggest  that  at  least  some  portion  of  adult 
birds  tend  to  disperse  from  one  breeding  ter- 
ritory to  another,  both  within  and  between  sea- 
sons. This  also  suggests  that  our  return  per- 
centages represent  a low  estimate  of  survival 
for  the  species. 

We  observed  that  HY  birds  depart  from  the 
breeding  grounds  before  AHY  birds.  This  ear- 
lier departure  of  young  may  be  related  to  an 
unusual  pattern  of  molt  described  for  White- 
eyed Vireos,  where  juveniles  exhibit  a partial 
replacement  of  primaries  prior  to  fall  migra- 
tion, perhaps  facilitating  migration  in  HY 
birds  (Lloyd-Evans  1983).  This  pattern  of 
molt,  however,  may  vary  geographically 
(George  1973).  Whether  departures  of  HY 
birds  precede  adults  in  other  parts  of  their 
range  is  unknown.  The  timing  of  fall  migra- 
tion between  HY  and  AHY  White-eyed  Vir- 
eos is  not  significantly  different  in  the  south- 
ern United  States.  Woodrey  and  Moore  (1997) 
found  that  AHY  and  HY  White-eyed  Vireos 
did  not  differ  in  their  distribution  of  arrivals. 
This  timing  pattern  is  in  contrast  to  Red-eyed 
Vireos  (V.  olivaceus)  whose  adults  depart  sig- 
nificantly earlier  than  the  young,  and  the  two 
age-classes  apparently  migrate  at  different 
rates  (Woodrey  and  Chandler  1997).  Perhaps 
while  young  White-eyed  Vireos  depart  earlier, 
the  adults  minimize  the  timing  differences  by 
the  time  the  two  groups  reach  the  southern 
United  States.  Alternatively,  our  observed  ear- 
lier departure  of  HY  birds  may  signal  the  on- 
set of  a pre-migration  dispersal  by  yoiing 
birds,  followed  by  their  actual  southerly  mi- 
gration some  time  later. 

Our  failure  to  recover  any  SY  birds  banded 
as  young  may  stem  from  several  factors.  First, 
it  is  possible  that  young  birds  tend  not  to  re- 


Hopp  et  cil.  • VIREO  RETURNS 


53 


turn  to  the  natal  area.  For  many  species,  a dis- 
persal of  young  provides  an  effective  mecha- 
nism for  avoiding  inbreeding  (Greenwood  and 
Harvey  1982,  Davis  and  Howe  1992),  al- 
though some  young  of  other  vireo  species  re- 
turn to  their  natal  area  (see  below).  Alterna- 
tively, since  White-eyed  Vireos  inhabit  a suc- 
cessional  window  of  habitat  it  may  be  in  the 
best  interest  of  young  birds  to  disperse  to  oth- 
er areas,  because  the  optimum  successional 
window  will  eventually  close  in  their  natal 
area.  Young  birds  may  disperse  from  their  na- 
tal areas  to  locate  suitable  breeding  habitat  for 
the  following  spring  (e.g..  Brewer  and  Harri- 
son 1975).  Second,  it  is  possible  that  the  com- 
bination of  a modest  number  of  birds  banded 
and  relatively  small  study  areas  simply  pre- 
cluded our  re-encounter  with  young.  In  the 
BBL  encounter  records  there  were  six  report- 
ed recoveries  of  banded  HY  birds  in  later 
years;  two  were  encountered  in  the  original 
10'  block,  two  were  encountered  15—20  km  of 
initial  banding  location,  and  two  were  en- 
countered 150-200  km  away  from  the  original 
banding  block.  These  records  are  equivocal  in 
suggesting  a modal  pattern  of  natal  dispersal 
for  the  species.  Instead  the  pattern  appears 
complex,  with  distance  depending  perhaps  on 
other  factors,  such  as  habitat  size  and  com- 
position, bird  densities,  or  hatching  time  of 
year.  In  Black-capped  Vireos,  young  birds 
show  a gradient  of  natal  dispersal  to  at  least 
21  km  (Grzybowski  1991).  For  this  species, 
the  return  percentages  were  directly  related  to 
size  of  the  study  area,  with  fewer  returns  in 
smaller  areas  (Grzybowski  1995).  In  Bell’s 
Vireos,  natal  dispersal  is  complex,  with  some 
young  returning  to  their  natal  area,  and  others 
dispersing  up  to  300  km  (Kus  1995,  Greaves 
and  Labinger  1998).  Greaves  and  Labinger 
(1998)  found  that  early  and  late  (before  and 
after  15  June,  respectively)  cohorts  of  HY 
Bell’s  Vireos  exhibit  significantly  different  de- 
grees of  natal  philopatry,  with  earlier  young 
returning  at  higher  rates.  For  both  Black- 
capped  and  Bell’s  vireos,  reported  return  per- 
centages of  young  from  various  studies  were 
15-27%  (Greaves  1987;  Grzybowski  1991, 
1995;  Greaves  and  Labinger  1998).  Our  lack 
of  returns  provides  us  with  little  information 
on  natal  dispersal  and  survival,  and  instead 
raises  questions  about  dispersal  dynamics  of 
this  species. 


ACKNOWLEDGMENTS 

We  thank  J.  C.  Barlow,  R.  Greenberg,  J.  A.  Grzy- 
bowski, T.  Keller,  B.  Kingsolver,  K.  Kliinkiewicz,  B. 
McClellan,  T.  Melcer,  C.  Qualls,  and  two  anonymous 
reviewers  for  comments  on  earlier  versions  of  this 
manuscript.  We  thank  the  Bird  Banding  Laboratory  for 
providing  encounter  records  for  White-eyed  Vireos. 
We  thank  T.  Lloyd-Evans  and  K.  Klimkiewicz  for  per- 
mission to  use  individual  encounter  records.  We  grate- 
fully acknowledge  Mr.  and  Mrs.  L.  R Collins,  III  for 
their  kind  permission  to  conduct  our  field  work  on 
their  land.  We  thank  G.  McLaren,  D.  Kegley  and  C. 
Carlin  for  field  assistance.  The  work  was  partially  sup- 
ported by  a Mednick  Research  fellowship  and  two 
Mellon  Faculty  Development  grants  to  SLH,  and  a J.  J. 
Munay  research  award  to  CAB.  This  research  was 
conducted  while  SLH  was  on  the  faculty  of  Emory  and 
Henry  College  in  Emory,  Virginia. 

LITERATURE  CITED 

Askins,  R.  a.,  j.  F.  Lynch,  and  R.  Greenberg.  1990. 
Population  declines  in  migratory  birds  in  eastern 
North  America.  Curr.  Ornithol.  7:1-57. 

Barber  D.  R.  and  T.  E.  Martin.  1997.  Influence  of 
alternate  host  densities  on  Brown-headed  Cowbird 
parasitism  rates  in  Black-capped  Vireos.  Condor 
99:595-604. 

Barlow,  J.  C.  1980.  Patterns  of  ecological  interactions 
among  migrant  and  resident  vireos  on  the  winter- 
ing grounds.  Pp.  79-107  in  Migrant  birds  in  the 
Neotropics:  ecology,  behavior,  distribution  and 
conservation  (A.  Keast  and  E.  S.  Morton,  Eds.). 
Smith.sonian  Institution  Press,  Washington,  D.C. 
Bedard,  J.  and  G.  LaPointe.  1984.  Banding  returns, 
arrival  times,  and  site  fidelity  in  the  Savannah 
Sparrow.  Wilson  Bull.  96:196—205. 

Bibby,  C.  j.,  N.  D.  Burgess,  and  D.  A.  Hill.  1992. 
Bird  census  techniques.  Academic  Press,  London, 
U.K. 

Bradley,  R.  A.  1980.  Vocal  and  territorial  behavior  in 
the  white-eyed  vireo.  Wilson  Bull.  92:302-31 1. 
Bradley,  R.  A.  1981.  Song  variation  within  a popu- 
lation of  White-eyed  Vireos  (Vireo  gri.sens).  Auk 
98:80-87. 

Brewer,  R.  and  K.  G.  Harrison.  1975.  The  time  of 
habitat  selection  by  birds.  Ibis  117:521-522. 
Conner,  R.  N.,  J.  G.  Dickson,  B.  A.  Locke,  and  C. 
A.  Segelquist.  1983.  Vegetation  characteristics 
important  to  common  .songbirds  in  East  Texas. 
Wilson  Bull.  95:349-361. 

Davis.  G.  J.  and  R.  W.  Howe.  1992.  Juvenile  dispersal, 
limited  breeding  sites,  and  the  dynamics  of  meta- 
populations. Thcor.  Pop.  Biol.  41:184-207. 
DeGraaf,  R.  and  j.  H.  Rappole.  1995.  Neotropical 
migratory  birds.  Cornell  Univ.  Press,  Ithaca,  New 
York. 

Franzreb,  K.  E.  1989.  Endangered  status  and  strate- 
gies for  protection  of  the  Least  Bell's  Vireo  (Vireo 
helm  pusiUus)  in  California.  West.  Birds  1 8:43-49. 


54 


THE  WILSON  BULLETIN 


Vol.  Ill,  No.  I,  March  1999 


George.  W.  G.  1973.  Molt  of  juvenile  White-eyed  Vir- 
eos.  Wihson  Bull.  85:327-330. 

Graber,  J.  W.,  R.  R.  Graber,  and  E.  L.  Kirk.  1985. 
Illinois  birds:  vireos.  111.  Nat.  Hist.  Surv.  Biol. 
Notes  124:1-38. 

Greaves,  J.  M.  1987.  Nest-site  tenacity  of  Least  Bell’s 
Vireos.  West.  Birds  18:50-54. 

Greaves,  J.  M.  and  Z.  Labinger.  1998.  Site  tenacity 
and  dispersal  of  Least  Bell’s  Vireos.  Trans.  West. 
Sect.  Wildl.  Soc.  33:18-23. 

Greenberg,  R.  1992.  Forest  migrants  in  non-forest 
habitats  on  the  Yucatan  Peninsula.  Pp.  273-286  in 
Ecology  and  conservation  of  neotropical  migrant 
landbirds  (J.  M.  Hagan,  III  and  D.  W.  Johnston, 
Eds.).  Smithsonian  Institution  Press,  Washington, 
DC. 

Greenberg,  R.,  D.  K.  Niven,  S.  L.  Hopp,  and  C. 
Boone.  1993.  Fmgivory  and  coexistence  in  a res- 
ident and  a migratory  vireo  on  the  Yucatan  pen- 
insula. Condor  95:990-999. 

Greenberg,  R.,  M.  S.  Foster,  and  L.  Marquez-Val- 
DELAMAR.  1995.  The  role  of  the  White-eyed  Vireo 
in  the  dispersal  of  Biirsera  fruit  on  the  Yucatan 
peninsula.  J.  Trop.  Ecol.  11:619-639. 

Greenwood,  P.  J.  1980.  Mating  systems,  philopatry, 
and  dispersal  in  birds  and  mammals.  Anim.  Be- 
hav.  28:1 140-1 162. 

Greenwood,  P J.  and  P.  H.  Harvey.  1982.  The  natal 
and  breeding  dispersal  of  birds.  Annu.  Rev.  Ecol. 
Syst.  13:1-21. 

Grzybowski,  j.  a.  1991.  Survivorship,  dispersal  and 
population  structure  of  Black-capped  Vireos  at  the 
Kerr  Wildlife  Management  Area,  Texas.  Resour. 
Protection  Div.,  Texas  Parks  Wildlife  Dept.,  Aus- 
tin. 

Grzybowski,  J.  A.  1995.  The  Black-capped  Vireo 
(Vireo  alricapillus).  In  The  birds  of  North  Amer- 
ica, no.  181  (A.  Poole  and  F.  Gill,  Eds.).  The 
Academy  of  Natural  Sciences,  Philadelphia;  The 
American  Ornithologists'  Union,  Washington, 
D.C. 

Grzybowski,  J.  A.,  R.  B.  Clapp,  and  J.  T.  Marshall. 
1986.  Hi.story  and  current  status  of  the  Black- 
capped  Vireo  in  Oklahoma.  Am.  Birds  34:742- 
745. 

Grzybowski,  J.  A.,  D.  J.  Tazik,  and  G.  D.  Schnell. 
1994.  Regional  analysis  of  Black-capped  Vireo 
breeding  habitats.  Condor  96:512-544. 

Hagan,  J.  M.  and  D.  W.  Johnston.  1992.  Ecology  and 
con.servation  of  Neotropical  migrant  landbirds. 
Smith.sonian  Institution  Press,  Washington,  D.C. 

Hill,  G.  E.  1989.  Late  spring  arrival  and  dull  nuptial 
plumage:  aggression  avoidance  by  yearling 
males?  Anim.  Behav.  37:665-673. 

Holmes,  R.  T,  T.  W.  Sherry,  and  L.  Reitsma.  1989. 
Population  structure,  territoriality  and  overwinter 
survival  of  two  migrant  warbler  species  in  Jamai- 
ca. Condor  91:545-561. 

Holmes,  R.  T.  and  T.  W.  Sherry.  1992.  Site  fidelity 
of  migratory  warblers  in  temperate  breeding  and 
Neotropical  wintering  areas:  implications  for  pop- 


ulation dynamics,  habitat  selection,  and  conser- 
vation. Pp.  563-575  in  Ecology  and  conservation 
of  neotropical  migrant  landbirds.  (J.  M.  Hagan,  III 
and  D.  W.  Johnston,  Eds.).  Smithsonian  Institution 
Press,  Washington,  D.C. 

Hopp,  S.  L.,  A.  Kirby,  and  C.  A.  Boone.  1995.  White- 
eyed Vireo  (Vireo  griseiis).  In  The  birds  of  North 
America,  no.  168  (A.  Poole  and  F.  Gill,  Eds.).  The 
Academy  of  Natural  Sciences,  Philadelphia;  The 
American  Ornithologists’  Union;  Washington, 
DC. 

Ketterson,  E.  D.  and  V.  Nolan,  Jr.  1983.  The  evo- 
lution of  differential  bird  migration.  Curr.  Orni- 
thol.  1:357-402. 

Kus,  B.  E.  1995.  Status  of  the  Least  Bell’s  Vireo  and 
southwestern  Willow  Flycatcher  at  Marine  Corps 
Base  Camp  Pendleton,  California,  in  1995.  Envi- 
ronmental and  Natural  Resources  Office,  Marine 
Corps  Base  Camp  Pendleton,  California. 

Lloyd-Evans,  T.  L.  1983.  Incomplete  molt  of  juvenile 
White-eyed  Vireos  in  Massachusetts.  J.  Field  Or- 
nithol.  54:50-57. 

Lovehoy,  T.  E.,  Ill  AND  D.  C.  Oren.  1981.  The  mini- 
mum critical  size  of  ecosystems.  Pp.  7-12  in  For- 
est island  dynamics  in  man-dominated  landscapes 
(R.  L.  Burgess  and  D.  M.  Sharpe,  Eds.).  Springer- 
Verlag,  New  York. 

Lynch,  J.  F.  1989.  Distribution  of  overwintering  Ne- 
arctic  migrants  in  the  Yucatan  peninsula,  I:  gen- 
eral patterns  of  occurrence.  Condor  91:515-544. 

Lynch,  J.  F.  1992.  Distribution  of  overwintering  Ne- 
arctic  migrants  in  the  Yucatan  peninsula,  II:  use 
of  native  and  human-modified  vegetation.  Pp. 
178-196  in  Ecology  and  conservation  of  neotrop- 
ical migrant  landbirds  (J.  M.  Hagan,  III  and  D.  W. 
Johnston.  Eds.).  Smithsonian  Institution  Press, 
Washington,  D.C. 

Marvil,  R.  E.  and  a.  Cruz.  1989.  Impact  of  Brown- 
headed Cowbird  parasitism  on  reproductive  suc- 
cess of  the  Solitary  Vireo.  Auk  106:476-480. 

M0LLER,  A.  P.  1990.  Male  tail  length  and  female  mate 
choice  in  the  monogamous  swallow  Hirundo  rus- 
tica.  Anim.  Behav.  39:458-465. 

Morton,  E.  S.  and  K.  C.  Derrickson.  1990.  The  bi- 
ological significance  of  age-specific  return  sched- 
ules in  breeding  Purple  Martins.  Condor  92: 1040- 
1050. 

Nice,  M.  M.  1941.  The  role  of  territory  in  bird  life. 
Am.  Midi.  Nat.  26:441-487. 

Nolan,  V.,  Jr.  1978.  The  ecology  and  behavior  of  the 
Prairie  Warbler  (Dendroica  discolor).  Ornithol. 
Monogr.  26:1-595. 

Payne,  R.  B.  and  L.  L.  Payne.  1990.  Survival  esti- 
mates ot  Indigo  Buntings:  comparisons  of  banding 
recoveries  and  local  observations.  Condor  92' 
938-946. 

Price,  J.,  S.  Droege,  and  A.  Price.  1995.  The  summer 
atlas  of  North  American  birds.  Academic  Press, 
London,  U.K. 

Rappole,  j.  H.  1995.  The  ecology  of  migrant  birds:  a 


Hopp  el  al.  • VIREO  RETURNS 


55 


Neotropical  perspective.  Smithsonian  Institution 
Press,  Washington,  D.C. 

Rappole,  J.  H.  and  D.  W.  Warner.  1980.  Ecological 
aspects  of  migrant  bird  behavior  in  Veracruz, 
Mexico.  Pp.  353-393  in  Migrant  birds  in  the  Neo- 
tropics: ecology,  behavior,  distribution  and  con- 
servation (A.  Keast  and  E.  S.  Morton,  Eds.). 
Smithsonian  Institution  Press,  Washington,  D.C. 

Remsen,  J.  V,  Jr.,  S.  W.  Cardiff,  and  D.  L.  Ditt- 
MANN.  1996.  Timing  of  migration  and  status  of 
vireos  (Vireonidcie)  in  Louisiana.  J.  Eield  Orni- 
thol.  67:119-140. 

Robbins,  C.  S.,  J.  S.  Sauer,  R.  S.  Greenberg,  and  S. 
Droege.  1989.  Population  declines  in  North 
American  birds  that  migrate  to  the  Neotropics. 
Proc.  Natl.  Acad.  Sci.  USA  86:7658-7662. 

Sauer,  J.  R.  and  S.  Droege.  1992.  Geographic  pat- 
terns in  population  trends  of  Neotropical  migrants 
in  North  America.  Pp.  26-42  in  Ecology  and  con- 
servation of  Neotropical  migrant  landbirds  (J.  M. 
Hagan,  III  and  D.  W.  Johnston,  Eds.).  Smithsonian 
Institution  Press,  Washington,  D.C. 

Sauer,  J.  R.,  G.  W.  Pendleton,  and  B.  G.  Peterjohn. 
1996.  Evaluating  causes  of  population  change  in 


North  American  insectivorous  songbirds.  Con- 
serv.  Biol.  10:465-478. 

Temple,  S.  A.  and  J.  R.  Cary.  1988.  Modeling  dy- 
namics of  habitat-interior  bird  populations  in  frag- 
mented landscapes.  Conserv.  Biol.  2:340-347. 

Terborgh,  j.  W.  1980.  The  conservation  status  of  Neo- 
tropical migrants:  present  and  future.  Pp.  21-30  in 
Migrant  birds  in  the  Neotropics:  ecology,  behav- 
ior, distribution  and  conservation  (A.  Keast  and  E. 
S.  Morton,  Eds.).  Smithsonian  Institution  Press, 
Washington,  D.C. 

USFWS  (United  States  Fish  and  Wildlife  Service). 
1986.  Endangered  and  threatened  wildlife  and 
plants;  determination  of  endangered  status  for  the 
Least  Bell’s  Vireo.  Federal  Register  51:16474— 
16483. 

USFWS  (United  States  Fish  and  Wildlife  Service). 
1991.  Black-capped  Vireo  {Vireo  atricapillus)  re- 
covery plan.  Austin,  Texas. 

WOODREY,  M.  S.  AND  C.  R.  CHANDLER.  1997.  Age- 
related  timing  of  migration:  geographic  and  inter- 
specific patterns.  Wilson  Bull.  109:52—67. 

WooDREY,  M.  S.  AND  E R.  MooRE.  1997.  Age-related 
differences  in  the  stopover  of  fall  landbird  mi- 
grants on  the  coast  of  Alabama.  Auk  114:695-707. 


Wilson  Bull.,  111(1),  1999,  pp.  56-60 


RESPONSE  OF  BROWN-HEADED  NUTHATCHES  TO  THINNING 

OF  PINE  PLANTATIONS 

MICHAEL  D.  WILSON^  - AND  BRYAN  D.  WATTS' 


ABSTRACT. — Brown-headed  Nuthatches  (Sitta  pusilla)  reached  their  highest  abundance  within  loblolly  pine 
(Finns  taecki)  plantations  in  the  hrst  year  after  thinning  and  declined  in  subsequent  years.  Commercial  thinning 
of  plantations  resulted  in  a reduction  of  canopy  cover,  hardwood  basal  area,  and  understory  density.  Overall, 
the  detection  rates  of  nuthatches  were  low  (19%  of  points  surveyed)  and  no  nuthatches  were  detected  in  stands 
before  thinning.  Nuthatches  were  more  than  three  times  as  likely  to  be  detected  within  survey  points  containing 
snags  compared  to  those  that  did  not.  However,  snag  density  did  not  vary  signihcantly  between  stand  ages. 
These  patterns  suggest  that  nuthatch  distribution  within  stands  may  be  influenced  by  snag  distribution  but  that 
distribution  among  stands  may  be  determined  by  the  density  and  height  of  understory  vegetation.  Received  13 
March  1998,  accepted  15  Sept.  1998. 


Prior  to  European  settlement  of  North 
America,  the  Southeastern  Coastal  Plain  was 
characterize<J  by  old-growth  pine  forests  that 
covered  more  than  24  million  ha  (Croker 
1979).  This  ecosystem  was  maintained  by 
low-intensity  ground  fires  caused  by  lightning 
strikes  (Komarek  1964,  1974)  and  indigenous 
people  (Bartram  1791,  Ware  et  al.  1993).  Fires 
occurred  over  vast  areas  at  approximately  3- 
5 year  intervals  (Chapman  1932,  Krusac  et  al. 
1995)  and  maintained  forests  with  an  open 
midstory  and  dense  ground  cover  of  forbs  and 
grasses  (Platt  et  al.  1991). 

Land  clearing  for  agriculture,  harvesting  of 
longleaf  pine  (Pinus  palustris)  for  the  naval 
stores  industry,  and  the  suppression  of  wild- 
fires severly  reduced  the  extent  of  the  south- 
eastern pine  ecosystem  by  the  early  1800s 
(Ashe  1894,  1915;  Pinchot  and  Ashe  1897). 
Currently,  natural  stands  of  longleaf  pine  are 
restricted  to  only  about  1%  of  their  former 
range  (Ware  et  al.  1993). 

Brown-headed  Nuthatches  {Sitta  pusilla) 
are  among  a small  group  of  species  including 
the  Red-cockaded  Woodpecker  (Picoides  ho- 
realis)  and  the  Bachman’s  Sparrow  (Aimophi- 
la  aestivalis)  that  are  endemic  to  the  south- 
eastern pine  ecosystem  (Jackson  1988).  The 
Red-cockaded  Woodpecker  and  the  Bach- 
man’s Sparrow  have  experienced  significant 
population  declines  within  the  southeast  re- 
gion (Lennartz  and  Henry  1985,  Dunning 
1993);  however,  both  have  benefited  from 


' Center  F*or  Con.servation  Biology,  College  of  Wil- 
liam and  Mary.  Williamsburg,  VA  23187-8795. 
^Corresponding  author;  E-mail:  mdwils@mail.wm.edu 


management  practices  that  produce  a habitat 
structure  similar  to  the  historic  southeastern 
pine  ecosystem  (Gobris  1992,  Plentovich  et  al. 
1998). 

The  Brown-headed  Nuthatch  has  also  ex- 
perienced a contraction  of  its  former  range 
(Jackson  1988),  and  according  to  data  from 
the  U.S.  Fish  and  Wildlife  Service’s  Breeding 
Bird  Survey  has  been  declining  at  a rate  of 
more  than  1.5%  per  year  throughout  much  of 
the  Southeast  (Sauer  et  al.  1997).  Very  little 
is  known  about  the  ecology  and  habitat  re- 
quirements of  the  Brown-headed  Nuthatch 
and  even  less  is  known  about  how  current  for- 
est management  practices  may  affect  its  dis- 
tribution. The  purpose  of  this  paper  is  to  pre- 
sent some  information  on  the  use  of  pine  plan- 
tations by  Brown-headed  Nuthatches  relative 
to  stand  age  and  commercial  thinning. 

METHODS 

This  study  was  conducted  in  managed  loblolly  pine 
(Finns  taeda)  plantations  in  eastern  North  Carolina 
(approximately  35°  50'  N,  77°  00'  W).  These  planta- 
tions are  managed  for  pulpwood  and  sawtimber  pro- 
duction on  a 30—35  year  rotation.  After  canopy  clo- 
sure, the  plantations  are  thinned  twice  before  final  har- 
vest. Thinnings  reduce  the  number  of  trees,  open  the 
forest  canopy,  and  allow  for  growth  of  understory  veg- 
etation. 

We  selected  stands  that  repre.sented  seven  different 
ages  and  relation  to  thinning:  (1)9-11  year  old  stands 
with  closed  canopies,  (2)  13—16  year  old  stands  within 
one  year  after  the  first  commercial  thinning,  (3)  l6-f8 
year  old  stands  that  were  three  years  after  first  thin- 
ning, (4)  19-21  year  old  stands  that  were  5 years  after 
first  thinning,  (5)  22-26  year  old  stands  that  were  with- 
in I year  after  .second  (hinning,  (6)  28-29  year  old 
stands  that  were  3 years  after  second  thinning,  and  (7) 
30—35  year  old  stands  that  were  5 years  after  second 


56 


VV/7.sv;/(  and  Walls  • BROWN-HEADHD  NUTHATCH  HABn'AT  USE 


57 


thinning.  Six  replicate  stands  (each  > 24  ha)  were  se- 
lected for  each  stand  type.  Within  each  stand  age, 
stands  were  chosen  to  minimize  variation  in  planted 
stocking  level  and  basal  area  of  pine.  Stands  within 
each  type  were  separated  by  at  least  500  m. 

Seven  minute,  rtxed-radiiis  (50  m)  point  counts  were 
used  to  measure  the  density  and  frequency  of  occur- 
rence of  Brown-headed  Nuthatches  within  study  plan- 
tations. Four  point  counts  were  established  within  each 
stand  and  distributed  evenly  between  edge  and  interior 
locations.  Edge  points  were  positioned  50  m from  the 
stand  edge  such  that  the  plot  perimeter  was  tangential 
to  the  stand  edge.  For  all  stands,  edge  points  were 
positioned  on  stand  edges  that  were  adjacent  to  log- 
ging roads.  Interior  points  were  positioned  150  m from 
the  stand  edge.  Stands  were  surveyed  three  times  be- 
tween 1 June  and  4 July  1997.  Surveys  were  initiated 
0.5  hr  after  sunrise  and  concluded  within  four  hours. 

The  vegetation  was  sampled  within  all  point  count 
plots  to  determine  ( 1 ) vegetation  changes  across  the 
growing  period,  (2)  vegetation  responses  to  thinning, 
and  (3)  relationships  between  nuthatch  distribution  and 
vegetation.  Linear  transects  were  used  for  vegetation 
sampling  parallel  to  the  long,  regularly  distributed  can- 
opy openings  created  by  row  thinning.  The  length  of 
vegetation  transects  was  standardized  to  25  m and  the 
width  varied  between  4 and  7 m to  accommodate  var- 
iation in  thinned  and  non-thinned  longitudinal  rows 
within  stands.  Four  vegetation  transects  were  estab- 
lished within  each  point  count  and  equally  distributed 
between  thinned  and  non-thinned  rows. 

Habitat  data  were  collected  at  two  levels  within  tran- 
sects. Counts  of  all  large  woody  plants  (>  8 cm  dbh) 
and  dead  standing  stems  (snags)  by  type  (hardwood  vs 
pine)  and  stem  diameter  class  (8-23,  24—38,  > 38  cm 
dbh)  were  made  over  the  entire  25  m transect.  Pine 
and  hardwood  basal  areas  were  estimated  using  the 
midpoint  dbh  for  the  two  smaller  diameter  classes  and 
38  cm  for  the  larger  class  (few  trees  were  larger  than 
38  cm  dbh).  Additional  information  was  collected 
within  2 X 2 m quadrats  established  at  opposite  ends 
of  each  transect.  Information  collected  included  can- 
opy cover  (measured  in  four  cardinal  directions  of  a 
compass  by  convex  densiometer)  and  canopy  height 
(measured  using  a clinometer),  groundcover  height  and 
counts  of  all  stems,  shrubs,  and  saplings  (>  0.5  m in 
height  and  < 8 cm  dbh).  Counts  were  summed  to  rep- 
resent total  groundcover  density  (stem.s/m^). 

A Kruskal-Wallis  test  was  used  to  test  for  the  influ- 
ence of  stand  age  on  all  habitat  variables  except  for 
counts  of  snags.  Because  of  the  many  zero  values  for 
counts  of  both  Brown-headed  Nuthatches  and  snags 
(i.e.,  data  were  distributed  as  a negative  binomial),  fre- 
quency of  occurrence  values  were  used  to  assess  pat- 
terns among  stand  types.  The  relationship  between  nut- 
hatches and  habitat  variables  was  assessed  at  the  level 
of  the  point  count  using  Kendall’s  rank  correlation. 
Nuthatches  were  not  detected  in  9-10  year  old  stands, 
so  this  stand  age  was  eliminated  from  all  analy.ses  and 
used  only  for  descriptive  purposes. 


RESULTS 

Stand  age  had  a significant  influence  (Krus- 
kal-Wallis test:  df  = 5,  P 0.01)  on  all  hab- 
itat variables  measured  except  the  density  of 
snags  (Table  1).  Canopy  height  and  ground 
cover  height  were  positively  related  to  stand 
age  whereas  pine  density  was  negatively  re- 
lated to  stand  age.  All  other  significant  vari- 
ables increased  with  stand  age  but  were  also 
influenced  by  commercial  thinning. 

Detection  rates  for  Brown-headed  Nut- 
hatches within  pine  plantations  were  relatively 
low.  Nuthatches  were  detected  in  15  of  42 
(35.7%)  pine  stands  included  in  the  study  and 
32  of  168  (19%)  individual  point  counts  sur- 
veyed. Stand  age  had  a significant  influence 
on  the  detection  of  Brown-headed  Nuthatches 

(XNates  Correction  = 12-3,  df  = 5,  P < 0.05;  Fig. 
1).  No  nuthatches  were  detected  in  forest 
patches  prior  to  first  thinning.  The  number  of 
points  where  nuthatches  were  detected  was 
greatest  in  the  year  immediately  following 
thinning  and  declined  with  time  after  thinning. 
Using  survey  points  as  statistical  units,  nut- 
hatches were  significantly  associated  with 
habitat  variables  that  were  directly  influenced 
by  thinning  events.  For  example,  nuthatch 
abundance  was  negatively  correlated  with 
canopy  cover  (Kendall  t = —0.12,  77  = 144, 
P < 0.03),  hardwood  density  (t  = —0.14,  n 
= 144,  P < 0.02),  and  basal  area  of  hard- 
woods (t  = —0.13,  n = 144,  P < 0.02).  In 
addition,  nuthatch  abundance  was  positively 
correlated  with  groundcover  density  (t  = 
0.19,  n = 144,  P < 0.001).  Nuthatch  density 
was  not  significantly  correlated  with  canopy 
height  (t  = —0.04,  n = 144,  P > 0.05),  pine 
density  (t  = 0.04,  n — 144,  P > 0.05)  or  pine 
basal  area  (t  = —0.02,  n = 144,  P > 0.05). 

Although  stand  type  did  not  have  a signif- 
icant influence  on  the  number  of  survey  plots 
containing  standing  snags,  and  snags  did  not 
appear  to  result  from  thinning,  nuthatches 
were  positively  correlated  with  standing  snags 
(T  = 0.15,  77  = 144,  P < 0.009).  In  fact,  nut- 
hatches were  over  three  times  more  likely  to 
be  detected  within  survey  plots  containing 
standing  snags  (12  of  32  plots,  37.5%)  com- 
pared to  plots  that  did  not  (13  of  122  plots, 
11.6%;  = 7.35,  df  = 1,  P < 0.007). 

DISCUSSION 

It  is  generally  thought  that  partially  rotted 
wood  is  a prerequisite  for  cavity  excavation 


TABLE  1.  Habitat  characteristics  of  7 stand  types  selected  for  study  («  — 6 each).  Median  values  are  presented  for  all  habitat  variables  except  for  snags  where 
frequency  of  occurrence  is  presented. 


58 


THE  WILSON  BULLETIN  • Vol.  111.  No.  1.  March  1999 


— — — oo  — lO  — — 
cqqqooooo 
dcddddddd 

VVVVVVAVV 


I DJ 
O'  >s 


— VO  r-  rj  (N 


^ O O O O On  (N 
(N  On  r-  (N  ID  — 


m O' 


O On  O On  O'  cD 
P O O »D  NO  CN 


ID 

(N  ON  On  fN  CN 


Dj  O (N  O NO 
CD  (N  O <N  O 
(N  00  so  — 


00 


^ o 
d (N 


ON  Tf  O 00  o 

— ‘ no  d — d 

n 00  — — ON 

(N 


rD  ID 
nO  CD 


^ ON  o p o o 

On  d — - d 

— 00  CD  O 

fD  — 


q 

00  00 
— 00 


ID  p O CD 
NO  d ^ 
— On  CD 
O 


^ o 
d 


— ID 
nO  i 


Tf  «D 

r-  00  — ^ 


p O 00  O 
d d »D  ' 


•o 

B 

o 

o 


eg  eg 

'S  =5 

C/)  (N 

-■  r* 

c 


eg  o 
6 


flj  rM 

■n  E 
.S  E 

o B 

D.  c/) 
u-  ^ • 
O ^ 

=tt  • 


JZ 

Of) 

‘S 

-C 

>N 

a 

o 

c 


Vi  eg 

w (U 
u 

>>  e^ 


o c/; 

>1  S 

CL  T3 
C3  ^ 


« rt 
C 

<U  C3 
■V 

■o  -o 
O O 
O O 

■a  •2 

ca 


U U cu  ix  j:  H 


c .2? 

g'  T3  J= 
C u u 
(U  o OJ 
3 > > 

cr  O o 

<U  O CJ 
^"0-0 
^ c c 

00  O 3 

eg  o o 
c ^ ^ 

c/5  o a 


by  Brown-headed  Nuthatches,  and  the  major- 
ity of  cavities  reported  have  been  located  in 
snags  (McNair  1984).  In  general  the  popula- 
tion density  of  cavity-nesting  birds  is  posi- 
tively related  to  snag  density  (Cunningham  et 
al.  1980,  O’Meara  1984,  Raphael  and  White 
1984).  In  Florida,  a large  percentage  of  the 
variation  in  the  density  of  cavity-nesting  birds 
(including  Brown-headed  Nuthatches)  was  ex- 
plained by  snag  density  and  dispersion  (Land 
et  al.  1989).  Snag  density  has  been  shown  to 
be  lower  in  pine  plantations  than  in  natural 
stands  (McComb  et  al.  1986),  and  was  low  in 
the  plantations  we  surveyed.  Brown-headed 
Nuthatches  were  significantly  more  likely  to 
be  detected  within  survey  points  that  con- 
tained snags.  The  possibility  that  snag  density 
may  serve  to  limit  overall  nuthatch  density 
within  loblolly  pine  plantations  requires  fur- 
ther investigation.  Because  nuthatch  density 
was  influenced  by  thinning  and  snag  density, 
snag  density  alone  does  not  explain  nuthatch 
distribution  among  pine  stands. 

Brown-headed  Nuthatches  exhibited  a rapid 
response  to  thinning.  Nuthatches  were  not  de- 
tected within  pine  plantations  prior  to  the  first 
thinning  but  reached  their  highest  densities 
within  the  first  year  after  thinning.  This  re- 
sponse suggests  that  thinning  activities  may, 
in  some  way,  enhance  habitat  structure  for 
nuthatches.  Thinning  activities  were  shown  to 
reduce  canopy  cover,  reduce  the  density  and 
basal  area  of  hardwoods,  and  increase  ground- 
cover  density. 

Although  the  importance  of  canopy  cover 
to  the  use  of  pinelands  by  Brown-headed  Nut- 
hatches has  not  been  explored,  Engstrom  and 
coworkers  (1984)  reported  that  nuthatch  abun- 
dance declined  as  the  density  of  midstory 
hardwoods  increased.  This  result  is  consistent 
with  our  observations  that  nuthatches  were 
less  common  in  years  after  thinning,  as  the 
density  and  basal  area  of  hardwoods  in- 
creased. 

The  effects  of  burning  hardwoods  on  stand 
use  by  Brown-headed  Nuthatches  is  similar  to 
that  of  thinning.  Nuthatches  used  (45  nut- 
hatches/km^)  mature  longleaf  pine  stands  that 
were  regularly  burned  in  Florida  (Repenning 
and  Labisky  1985),  but  not  stands  with  well 
developed  understories  (Hirth  et  al.  1991). 
Nuthatch  density  decreased  with  time  follow- 
ing burning  (Engstrom  et  al.  1984,  Wilson  et 


Wilson  and  Watts  • BROWN-HEADED  NUTHATCH  HABITAT  USE 


59 


FIG.  1.  Detection  frequency  (#  of  point  counts)  of  Brown-headed  Nuthatches  in  commercially  thinned  pine 
plantations  in  eastern  North  Carolina  (n  = 24  point  counts  for  each  stand  type).  Frequency  distribution  between 
stand  ages  was  significantly  different  from  an  even  distribution  (x^vaies correction  = 12.3,  df  = 5,  P < 0.05). 


al.  1995),  sirnilar  to  the  decline  in  nuthatches 
we  observed  following  thinning. 

One  possible  explanation  for  the  inverse  re- 
lationship between  the  density  of  understory 
vegetation  and  numbers  of  Brown-headed 
Nuthatches  is  that  vegetation  may  obscure  po- 
tential cavity  locations.  Brown-headed  Nut- 
hatch cavities  are  frequently  excavated  in  rel- 
atively low  positions;  usually  below  3.66  m 
(n  = 309;  McNair  1984).  Most  (68%)  cavities 
were  located  in  tree  stumps  (McNair  1984) 
suggesting  that  the  potential  for  increasing 
cavity  height  may  be  limited.  Regeneration  of 
understory  and  groundcover  vegetation  was 
rapid  in  the  current  study  such  that  low  cavity 
positions  could  be  obscured  quickly,  but  the 
impact  on  patch  use  by  Brown-headed  Nut- 
hatches is  unknown. 

ACKNOWLEDGMENTS 

We  thank  T.  Melchiors  for  administrative  support 
and  assistance  through  all  aspects  of  the  field  work  and 
for  providing  editorial  comments  on  the  manuscript. 
We  also  thank  B.  Barber  and  J.  Hughes  for  educating 
us  about  forest  tracts  and  silvicultural  techniques  and 
for  assistance  in  selecting  study  sites.  D.  Bradshaw,  G. 
Levandoski,  and  B.  Piccolo  provided  valuable  field  as- 
sistance, and  T.  Rafiq  helped  with  data  management. 
This  study  was  funded  through  a cooperative  agree- 
ment between  the  Weyerhaeuser  Company  and  the 
College  of  William  and  Mary. 


LITERATURE  CITED 

Ashe,  W.  W.  1894.  The  forests,  forest  lands,  and  forest 
products  of  eastern  North  Carolina.  N.C.  Geol. 
Surv.  5:1-128. 

Ashe,  W.  W.  1915.  Loblolly  or  North  Carolina  pine. 

N.C.  Geol.  Surv.  and  Econ.  Surv.  24:1—176. 
Bartram,  W.  1791.  Travels  through  North  and  South 
Carolina,  Georgia,  east  and  west  Florida.  Dover 
Publications,  New  York. 

Chapman,  H.  H.  1932.  Is  the  longleaf  type  a climax? 
Ecology  13:328-334. 

Conner,  R.  N.  and  D.  C.  Rudolph.  1991.  Effects  of 
midstory  reduction  and  thinning  in  Red-cockaded 
Woodpecker  cavity  clusters.  Wildl.  Soc.  Bull.  19: 
63-66. 

Croker,  T.  C.,  Jr.  1979.  Longleaf  pine.  J.  For.  Hist. 
23:32-43. 

Cunningham,  J.  B.,  R.  P.  Balda,  and  W.  S.  Gaud. 
1980.  Selection  and  use  of  snags  by  secondary 
cavity-nesting  birds  of  the  ponderosa  pine  forest. 
U.S.  For.  Serv.  Res.  Pap.  RM-222:1-15. 

Dunning,  J.  G.  Jr.  1993.  Bachman's  Sparrow  in  The 
birds  of  North  America,  no.  38  (A.  Poole.  P.  Set- 
tenheim,  and  F.  Gill,  Eds.).  The  Academy  of  Nat- 
ural Sciences,  Philadelphia,  Pennsylvania;  The 
American  Ornithologists'  Union,  Washington, 
D.C. 

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  Bull.  96:437-450. 
Gobris,  N.  M.  1992.  Habitat  occupancy  during  the 
breeding  season  by  Bachman's  Sparrow  at  Pied- 


60 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  1,  March  1999 


mont  National  Wildlife  Refuge  in  central  Georgia. 

M. Sc.  Thesis,  Univ.  of  Georgia,  Athens. 

Hirth,  D.  H.,  L.  D.  Harris,  and  R.  E Noss.  1991. 

Avian  community  dynamics  in  a peninsular  Flor- 
ida longleaf  pine  forest.  Fla.  Field.  Nat.  19:33-48. 
Jackson,  J.  A.  1988.  The  southeastern  pine  forest  eco- 
system and  its  birds:  past,  present,  and  future.  Pp. 
119-159  in  Bird  conservation  3 (J.  A.  Jackson, 
Ed.).  Univ.  of  Wisconsin  Press,  Madison. 
Komarek,  E.  V.,  Sr.  1964.  The  natural  history  of  light- 
ning. Proc.  Tall  Timbers  Fire  Ecol.  Conf.  3:139- 
183. 

Komarek,  E.  V.,  Sr.  1974.  Effects  of  fire  on  temperate 
forests  and  related  ecosystems:  southeastern  Unit- 
ed States.  Pp.  251-278  in  Fire  and  ecosystems  (T. 
T.  Kowalski  and  C.  E.  Ahlgren,  Eds.).  Academic 
Press,  New  York. 

Krusac,  D.  L.,  j.  M.  Dabney,  and  J.  J.  Petrick.  1995. 
An  ecological  approach  to  recovering  the  Red- 
cockaded  Woodpecker  on  southern  National  For- 
ests. Pp.  61-66  in  Red-cockaded  Woodpecker:  re- 
cover, ecology  and  management  (D.  L.  Kulhavy, 
R.  G.  Hooper,  and  R.  Costa,  Eds).  Center  for  Ap- 
plied Studies  in  Forestry,  College  of  Forestry,  Ste- 
phen E Austin  State  Univ.,  Nacogdoches,  Texas. 
Land,  D.,  W.  R.  Marion,  and  T.  E.  O’Meara.  1989. 
Snag  availability  and  cavity  nesting  birds  in  slash 
pine  plantations.  J.  Wildl.  Manage.  53: 1 165-1171. 
Lennartz,  M.  R.  and  V.  G.  Henry.  1985.  Recovery 
plan  for  the  Red-cockaded  Woodpecker.  U.S.  Fish 
and  Wildlife  Service,  Atlanta,  Georgia. 

McComb.  W.  C.,  S.  a.  Bonney,  R.  M.  Sheffield,  and 

N.  D.  Cost.  1986.  Snag  re.sources  in  Florida — are 
they  sufficient  for  average  populations  of  primary 
cavity  nesters?  Wildl.  Soc.  Bull.  14:40-48. 

McNair,  D.  B.  1984.  Clutch-size  and  nest  placement 
in  the  Brown-headed  Nuthatch.  Wilson  Bull.  96: 
296-301. 


O’Meara,  T.  E.  1984.  Habitat-island  effects  on  the  avi- 
an community  in  cypress  ponds.  Proc.  Annu. 
Conf.  Southeast.  Assoc.  Fish  W'ildl.  Agencies  38: 
97-110. 

Platt,  W.  J.,  J.  S.  Glitzenstein,  and  D.  R.  Streng. 
1991.  Evaluating  pyrogenicity  and  its  effects  on 
vegetation  in  longleaf  pine  savannahs.  Proc.  Tall 
Timbers  Fire  Ecol.  Conf.  17:143-161. 

PiNCHOT,  G.  AND  W.  W.  AsHE.  1897.  Timber  trees  and 
forests  of  North  Carolina.  N.C.  Geol.  Surv.  6:1- 
227. 

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. J.  Wild.  Manage.  62:347-354. 

Raphael,  M.  G.  and  M.  White.  1984.  Use  of  snags 
by  cavity-nesting  birds  in  the  Sierra  Nevada. 
Wildl.  Monogr.  86:1-66. 

Repenning,  R.  W.  and  R.  F.  Labisky.  1985.  Effects  of 
even-age  timber  management  on  bird  communi- 
ties of  the  longleaf  pine  forest  in  northern  Florida. 
J.  Wildl.  Manage.  49:1088-1098. 

Sauer,  J.  R.,  J.  E.  Hines,  G.  Gough,  I.  Thomas,  and 
B.  G.  Peterjohn.  1997.  The  North  American 
breeding  bird  survey  results  and  analysis,  version 
96.4.  Patuxent  Wildlife  Research  Center,  Laurel, 
Maryland.  URL  = http://www.mbr.nbs.gov/bbs/ 
bbs.html 

Ware,  S.,  C.  Frost,  and  P D.  Doerr.  1993.  Southern 
mixed  hardwood  forest:  the  former  longleaf  pine 
forest.  Pp.  447-494  in  Biodiversity  of  the  south- 
eastern United  States  (W.  H.  Martin,  S.  G.  Boyce, 
and  A.  C.  Echternacht,  Eds.).  John  Wiley  and 
Sons,  Inc.,  New  York. 

Wilson,  C.  W.,  R.  E.  Masters,  and  G.  A.  Bukenho- 
FER.  1995.  Breeding  bird  response  to  pine-grass- 
land community  restoration  for  Red-cockaded 
Woodpeckers.  J.  Wildl.  Manage.  59:56-67. 


Wilson  Bull.,  111(1),  1999,  pp.  61-69 


DIFFERENCES  IN  MIGRATORY  TIMING  AND  ENERGETIC 
CONDITION  AMONG  SEX/AGE  CLASSES  IN  MIGRANT 
RUBY-CROWNED  KINGLETS 

DAVID  L,  SWANSON,'^  ERIC  T.  LIKNES,'  AND  KURTIS  L.  DEAN' 


ABSTRACT. — Ruby-crowned  Kinglets,  Kegulus  calendula,  are  small  temperate  zone  passerine  migrants  that 
breed  in  conifer  forests  of  Canada  and  the  western  United  States  and  winter  along  the  west  coast  and  southern 
United  States  into  Mexico.  Previous  studies  have  revealed  that  male  kinglets  precede  females  during  spring 
migration  and  that  females  precede  males  during  fall  migration  at  various  sites  in  eastern  North  America.  We 
used  mist  net  capture  data  to  document  sex/age  structure,  fat  loads,  and  morphometries  of  kinglets  passing 
through  southeastern  South  Dakota  during  spring  and  fall  migrations  from  1992-1995.  Males  migrated  signifi- 
cantly earlier  than  females  in  spring;  median  passage  dates  differed  by  at  least  8 days.  These  data  are  consistent 
with  U.S.  Geological  Survey  Bird  Banding  Laboratory  records  for  central  and  eastern  North  America,  which 
indicate  that  passage  of  males  before  females  during  spring  migration  is  a widespread  phenomenon.  Bird  Banding 
Laboratory  data  also  indicate  that  males  winter  significantly  farther  north  than  females.  We  tested  whether  the 
differential  winter  distribution  of  the  sexes  could  account  for  the  differential  pattern  of  spring  migration  and 
found  that  differential  winter  distributions  do  not  fully  account  for  the  differential  timing  of  spring  migration 
between  the  sexes.  Sex-specific  migration  rates  do  not  differ  over  the  entire  spring  migration  route,  so  differences 
in  the  onset  of  migration  apparently  contribute  to  differential  spring  migration.  Males  migrated  significantly  later 
in  the  fall  than  females,  with  median  dates  in  southeastern  South  Dakota  differing  by  7 days.  We  also  examined 
whether  differences  in  energetic  condition  were  associated  with  differences  in  migratory  timing.  Visible  fat 
scores  did  not  differ  between  males  and  females  in  spring,  although  males  had  a significantly  higher  mean 
condition  index  (mass/wing  chord).  Sexes  within  the  same  age  class  did  not  differ  in  visible  fat  scores  in  fall, 
but  adults  tended  to  be  fatter  than  juveniles,  significantly  so  in  some  cases.  No  significant  differences  in  condition 
index  were  apparent  among  sex/age  classes  in  fall,  although  the  adult-juvenile  difference  approached  significance 
{P  < 0.09),  with  adults  having  higher  values.  This  tendency  toward  elevated  fat  and  condition  index  in  adults 
relative  to  juveniles  in  fall  is  consistent  with  more  efficient  foraging  by  adults  than  by  inexperienced  juveniles 
at  migratory  stopover  sites,  although  differences  in  predation  pressure  or  migratory  routes  and  destinations  might 
also  be  involved.  Received  12  May  1998,  accepted  24  Oct.  1998. 


Differential  timing  of  migration  of  sexes 
an(j/or  age  classes  has  been  reported  for  nu- 
merous birds  (see  Gauthreaux  1982,  Ramos 
1988  for  reviews).  For  passerine  birds,  males 
generally  precede  females  in  spring,  but  the 
timing  of  migration  in  fall  is  more  complex 
and  varied  between  sexes  or  age  classes  (see 
Hall  1981,  Gauthreaux  1982,  Francis  and 
Cooke  1986,  Ramos  1988,  Nolan  and  Ketter- 
son  1990,  Hussell  1991,  Winker  and  Rappole 
1992).  Earlier  passage  of  males  than  females 
in  spring  presumably  relates  to  advantages  in 
territory  acquisition  that  accrue  to  those  in- 
dividuals arriving  early  on  the  breeding 
grounds,  regardless  of  whether  this  early  ar- 
rival results  from  sexual  differences  in  win- 
tering areas,  departure  dates,  or  rates  of  mi- 
gration (Myers  1981,  Chandler  and  Mulvihill 


' Dept,  of  Biology,  Univ.  of  South  Dakota,  Vermil- 
lion, SD  57069-2390. 

^ Corresponding  author;  E-mail: 
dlswanso(g)  sunflowr.usd.edu 


1990).  Differential  timing  of  fall  migration 
may  result  from  differences  in  departure  dates 
or  rates  of  migration  as  a result  of  ecological 
or  physiological  factors  (e.g.,  juvenile  devel- 
opment, parental  responsibilities,  molt,  fat  de- 
position), and/or  from  differential  distribution 
of  age/sex  classes  throughout  the  wintering 
range  (Gauthreaux  1982,  Ketterson  and  Nolan 
1976,  Morton  1984,  Prescott  and  Middleton 
1990,  Woodrey  and  Chandler  1997). 

Ruby-crowned  Kinglets,  Regulus  calendu- 
la, are  small  passerines  breeding  in  coniferous 
or  mixed  coniferous-deciduous  forests  over 
most  of  northern  and  western  North  America 
and  wintering  throughout  the  western  and 
southern  United  States,  Mexico,  and  south  to 
Guatemala  (American  Ornithologists’  Union 
1983).  The  winter  range  is  apparently  restrict- 
ed to  regions  with  a relatively  mild  winter  cli- 
mate (Lepthein  and  Bock  1976,  Root  1988). 
Differential  timing  of  migration  in  males  and 
females  has  been  reported  from  Iowa  (Crim 
1976)  and  Ontario  (Fairfield  and  Shirokoff 


61 


62 


THE  WILSON  BULLETIN 


Vol.  Ill,  No.  I,  March  1999 


1978),  but  migration  timing  of  sexes  in  the 
Great  Plains  has  not  been  previously  reported. 
In  addition,  male  kinglets  apparently  winter 
farther  north,  on  average,  than  females  (Fair- 
field  and  Shirokoff  1978),  but  the  extent  to 
which  differences  in  winter  distribution  affect 
differential  migration  is  unknown.  No  previ- 
ous study  has  been  attempted  to  define  the  rel- 
ative roles  of  geographic  origin,  departure 
dates,  and  rates  of  migration  in  shaping  dif- 
ferential migration  patterns  in  kinglets. 

We  used  capture  data  from  spring  and  fall 
migration  periods  from  1992-1995  in  south- 
eastern South  Dakota  to  document  patterns  of 
migratory  movement  in  Ruby-crowned  King- 
lets and  to  determine  whether  timing  of  mi- 
gration and  energetic  condition  differed  be- 
tween sexes  in  either  season.  We  compared 
these  results  with  banding  data  from  the  U.S. 
Geological  Survey  Bird  Banding  Laboratory 
(BBL)  for  central  and  eastern  North  America 
(east  of  107°  W longitude)  to  determine  if  pat- 
terns of  migration  between  sexes  in  south- 
eastern South  Dakota  are  consistent  with  mi- 
gratory patterns  in  Ruby-crowned  Kinglets  in 
eastern  North  America.  We  also  tested  the  hy- 
pothesis that  differential  winter  distributions 
could  fully  account  for  differential  migratory 
timing. 

METHODS 

Collection  of  capture  data. — Banding  was  conduct- 
ed during  spring  and  fall  migration  from  1992-1995 
at  four  deciduous  woodland  study  sites  in  southeastern 
South  Dakota.  Two  sites  were  located  in  riparian  hab- 
itats along  the  Missouri  River  (42°  45'  N,  97°  00'  W; 
and  42°  46'  N,  97°  07'  W),  one  study  site  was  in  a 
riparian  woodland  along  the  Big  Sioux  River  (42°  45' 
N,  96°  37'  W),  and  one  study  site  included  both  ripar- 
ian and  upland  woodlands  along  Brule  Creek  (42°  55' 
N.  96°  46'  W),  a tributary  of  the  Big  Sioux  River.  Ri- 
parian habitats  along  the  Missouri  River  consisted  of 
deciduous  forest  dominated  by  cottonwood  (Populu.t 
deltoide.s).  boxelder  (.Acer  nef>undo),  American  elm 
(Ulmu.s-  americana),  mulberry  (Moru.s  alba),  and  dog- 
wood (Cannes  spp.).  The  Big  Sioux  River  site  was 
dominated  by  boxelder,  silver  maple  (Acer  .sacchari- 
nunt),  and  cottonwood.  The  riparian  forest  at  the  Brule 
Creek  site  consisted  mainly  of  boxelder  and  American 
elm,  while  the  upland  forest  was  dominated  by  bur  oak 
(Quercu.s  macrocarpa)  with  American  elm  and  hack- 
berry  (Celti.s  occidentali.s)  also  present.  The  Missouri 
River  study  sites  have  a generally  west-east  orienta- 
tion, while  the  Big  Sioux  River  and  Brule  Creek  sites 
are  oriented  north-south.  Extensive  deciduous  wood- 
lands in  this  area  of  .South  Dakota,  and  indeed  over 


most  of  the  northern  Great  Plains,  are  mainly  restricted 
to  river  courses  (South  Dakota  Ornithologists’  Union 
1991). 

Kinglets  were  captured  using  mist  nets  during  spring 
(15  April-26  May)  and  fall  (5  September-2  Novem- 
ber) migration  periods.  Capture  dates  were  chosen  to 
coincide  with  the  major  migratory  movements  for 
Neotropical  migrants,  not  kinglets  specifically,  so  the 
distribution  of  capture  effort  did  not  cover  the  kinglet 
migration  evenly.  However,  the  bulk  of  the  kinglet  mi- 
gration through  South  Dakota  occurs  during  our  cap- 
ture periods  (South  Dakota  Ornithologists’  Union 
1991).  From  15  April  through  26  May  in  spring  and 
from  5 September  to  5 October  in  fall,  nets  were 
opened  daily  and  cycled  on  a 4 day  rotation  among 
the  four  study  sites  so  that  each  study  site  was  sampled 
every  4 days.  Later  in  the  fall,  nets  were  not  opened 
every  day  and  most  capture  effort  was  concentrated  at 
the  Brule  Creek  study  site.  We  erected  2-7  mist  nets 
(10  m X 2.6  m,  30  mm  mesh)  each  day,  depending  on 
weather  and  available  personnel.  Net  placement  at  the 
various  study  sites  was  consistent  among  seasons  and 
years.  We  opened  the  nets  at  sunrise  and  nets  remained 
open  until  approximately  1 1 :00  CST  during  both 
spring  and  fall  migration.  Upon  capture  we  measured 
the  bird’s  mass  to  the  nearest  0.1  g with  an  Ohaus 
Model  LS200  balance,  unflattened  wing  chord  to  the 
nearest  0.1  mm,  and  visible  fat  on  a scale  of  0-5 
(Helms  and  Drury  1960).  Fat  scoring  was  performed 
by  all  three  authors,  but  we  regularly  checked  each 
other  on  individual  birds  to  ensure  that  we  were  scor- 
ing fat  similarly.  During  fall  migration  birds  were  aged 
by  skull  ossification  as  “after  hatching-year’’  (AHY, 
adults)  or  “hatching-year’’  (HY,  juveniles),  although 
this  method  may  be  inaccurate  after  1 October  because 
some  hatching-year  birds  may  complete  ossification  af- 
ter this  date  (Leberman  1970,  Pyle  et  al.  1987).  Retrix 
shape  was  not  an  effective  measure  of  age  in  our  pop- 
ulation because  most  birds  showed  an  intermediate 
condition  (Pyle  et  al.  1987).  Sex  was  determined  by 
plumage  differences  in  both  seasons.  Recaptures  were 
very  rare  and  were  not  included  in  our  analyses. 

We  used  banding  data  from  the  BBL  from  1986- 
1995  to  determine  winter  distributions  and  timing  of 
spring  migration  in  males  and  females.  Data  were 
grouped  into  5°  latitude  zones  for  comparisons.  Only 
banding  records  east  of  107°W  longitude  were  used 
in  these  comparisons  to  eliminate  populations  in  west- 
ern North  America  that  breed  at  lower  latitudes  and 
winter  at  higher  latitudes  than  populations  in  central 
and  eastern  North  America.  We  considered  records 
from  1 December-29  February  as  winter  records,  and 
from  1 April-31  May  as  spring  records.  Banding  data 
were  provided  as  number  of  captures  for  consecutive 
5 day  periods. 

To  determine  whether  differences  in  winter  distri- 
bution of  male  and  female  Ruby-crowned  Kinglets  are 
sufficient  to  account  for  spring  migration  patterns,  we 
u.sed  a method  developed  by  Chandler  and  Mulvihill 
(1990)  to  calculate  a predicted  pattern  of  differential 
migration  based  on  winter  distributions  of  the  sexes. 


Swanson  et  al.  • DIFFERENTIAL  MIGRATION  IN  KINGLETS 


63 


TABLE  1.  Winter  (December-Eebruary)  abundance  of  male  and  female  Ruby-crowned  Kinglets  in  eastern 
and  central  North  America  from  BBL  records  and  results  of  Goodness  of  Fit  tests  for  equal  abundance  at 
each  latitude. 


L.ititude“  (°N) 

Males 

Females 

% Males 

Xc^ 

Significance 

< 25 

14 

14 

50.0 

0.04 

P > 0.05 

25-29°  59' 

30 

28 

51.7 

0.02 

P > 0.05 

30-34°  59' 

169 

133 

56.0 

4.06 

P < 0.05 

35-39°  59' 

54 

17 

76.1 

18.25 

P < 0.001 

“ Latitudes  north  of  40°  N had  too  few  records  (n  = 5)  to  adequately  assess  distribution  of  sexes. 


Chandler  and  Mulvihill  (1990)  assumed  a linear 
change  in  the  percentage  of  males  in  the  population 
from  the  northern  to  the  southern  extent  of  the  win- 
tering range  for  their  winter  distribution;  they  predict- 
ed a linear  change  in  the  percentage  of  males  in  the 
population  throughout  migration  from  the  first  appear- 
ance to  the  last  appearance  of  migrating  males.  They 
then  used  linear  regression  to  calculate  the  predicted 
percentage  of  males  throughout  the  migratory  period 
and  compared  this  regression  against  a linear  regres- 
sion of  the  observed  percentage  of  males  over  the  mi- 
gratory period. 

We  used  BBL  data  for  wintering  Ruby-crowned 
Kinglets  to  calculate  a predicted  differential  pattern  of 
spring  migration  between  the  sexes.  The  predicted  per- 
centage of  males  in  the  migratory  population  was  cal- 
culated using  the  percentage  of  males  in  wintering 
populations  from  BBL  data  for  5°  latitude  zones  over 
the  entire  kinglet  wintering  range  in  eastern  and  central 
North  America  (Table  1).  We  then  divided  the  spring 
migratory  period  for  males  into  an  equal  number  of 
intervals  (4  intervals  at  5 days  each)  and  assumed  that 
the  percent  males  during  each  successive  5-day  inter- 
val (from  first  to  last  appearance)  should  equal  the  per- 
cent males  in  consecutive  5°  latitude  zones  from  north 
to  south. 

Statistics. — Median  passage  dates  for  sexes  were 
compared  by  median  tests  (Zar  1996).  Because  capture 
effort  was  not  evenly  distributed  over  the  entire  kinglet 
migration  period,  median  tests  may  be  misleading.  To 
control  for  uneven  capture  effort,  we  compared  pas- 
sage dates  for  sexes  by  Kolmogorov-Smirnov  test  for 
frequency  distributions  (Sokal  and  Rohlf  1995)  as  a 
function  of  cumulative  net  hours  over  all  years  pooled. 
Morphometries,  condition  index  [mass/wing  chord 
(Winker  1995)],  and  visible  fat  scores  for  males  and 
females  in  spring  were  compared  by  Student’s  r-test  or, 
if  sample  variances  were  unequal  (determined  by  F- 
test),  Mann-Whitney  test.  In  fall,  morphometries,  con- 
dition index,  and  visible  fat  scores  of  sex/age  classes 
were  compared  by  one-way  ANOVA,  or  by  Kruskal- 
Wallis  test  if  sample  variances  differed  (F-test).  Be- 
cause skull  ossification  provides  unreliable  age  dis- 
tinction after  1 October  in  kinglets  (Leberman  1970), 
fall  comparisons  of  sex/age  classes  were  conducted 
twice,  excluding  data  after  30  September  for  birds  with 
completely  ossified  skulls  (aged  AHY)  and  excluding 
all  data  (HY  and  AFIY)  after  30  September.  If  signif- 


icant differences  among  sex/age  classes  were  detected 
by  ANOVA  or  Kruskal-Wallis  tests,  Student’s  r-tests 
or  Mann-Whitney  tests  were  used  to  identify  which 
means  differed  with  a sequential  Bonferroni  procedure 
employed  for  a-level  adjustment  to  protect  against  in- 
flated type  1 error  rates  in  multiple  comparisons  (Rice 
1989).  This  procedure  involved  six  comparisons  (k  = 
6)  among  the  four  sex/age  classes  for  each  variable 
measured  (mass,  wing  chord,  fat  score,  and  condition 
index).  For  a-level  adjustment,  P values  from  individ- 
ual r-tests  for  each  comparison  were  ranked  from 
smallest  (P,)  to  largest  (Pg).  The  smallest  P value  was 
considered  significant  only  if  it  was  less  than  a/k  (0.05/ 
6 in  this  case).  If  Pj  was  significant,  then  Pj  was  con- 
sidered significant  only  if  less  than  a/k  — 1,  P3  was 
considered  significant  only  if  less  than  a/k  — 2,  and 
so  on,  until  the  equality  Pj  ^ a/(l-t-/:  — i)  was  not  met 
(Rice  1989).  Once  a P value  was  found  to  be  not  sig- 
nificant, all  larger  P values  for  that  comparison  were 
also  considered  nonsignificant. 

For  winter  BBL  data,  distributions  of  males  and  fe- 
males in  5°  latitude  zones  were  compared  by  2-tailed 

Goodness  of  Fit  test  with  a null  hypothesis  of  equal 
distribution  (1:1)  in  each  latitude  zone.  Bird  Banding 
Laboratory  data  for  timing  of  spring  migration  in 
males  and  females  were  compared  by  Kolmogorov- 
Smirnov  test  for  frequency  distribution  as  a function 
of  cumulative  captures  during  the  migratory  season 
pooled  over  all  years  (1986-1995).  To  determine 
whether  winter  distributions  of  sexes  could  fully  ac- 
count for  differential  migration  in  spring,  predicted  and 
observed  percentages  of  males  in  the  migratory  pop- 
ulation were  compared  by  ANCOVA.  Because  neither 
the  change  in  the  percentage  of  males  in  the  wintering 
population  as  a function  of  latitude  nor  the  percentage 
of  males  in  the  migratory  population  through  the 
spring  migratory  period  varied  in  a linear  fashion  (Fig. 
1),  data  for  percent  males  were  arcsine  transformed 
prior  to  ANCOVA.  In  addition,  we  conducted  Fisher’s 
exact  tests  on  observed  versus  predicted  percent  males 
for  each  5 day  interval  over  the  spring  migration  of 
male  kinglets.  Statistical  comparisons  of  mass,  wing 
chord,  condition  index,  and  visible  fat  scores  were  con- 
ducted with  Number  Cruncher  Statistical  System  (Ver- 
sion 4.1,  Kaysville,  Utah).  Other  statistical  tests  were 
performed  with  SAS  (PC  Version  6.03,  SAS  Institute 
1988). 


64 


THE  WILSON  BULLETIN 


Vo/.  Ill,  No.  J,  March  1999 


FIG.  1 . Observed  and  predicted  percentages  of 
male  Ruby-crowned  Kinglets  in  the  population  during 
successive  5-day  intervals  over  the  spring  migratory 
period  as  measured  by  capture  data  (first  captures 
only).  Predicted  values  were  calculated  based  upon 
U.S.  Geological  Survey  Bird  Banding  Laboratory  data 
from  1985-1995  for  wintering  distributions  of  male 
and  females  kinglets  east  of  107°  W longitude. 


RESULTS 

Southeastern  South  Dakota  is  not  a part  of 
the  breeding  range  for  Ruby-crowned  Kinglets 
(South  Dakota  Ornithologists’  Union  1991), 
so  all  birds  in  our  samples  were  transients. 
Median  passage  dates  for  all  years  pooled  for 
male  and  female  kinglets  in  spring  were  23 
April  and  1 May,  respectively  (Fig.  2).  For  fall 
migration,  median  passage  dates  for  all  years 
pooled  were  29  September  for  females  and  6 
October  for  males  (Fig.  2).  Median  tests  in- 
dicated that  males  migrated  significantly  ear- 
lier (x-  = 45.74,  df  = \,  P < O.OOl)  than 
females  in  spring  and  significantly  later  (y^  = 
5.30,  df  = \,  P < 0.01)  than  females  in  fall. 

Total  net  hours  ( 1 net  hour  = 1 net  open 
for  I h)  were  2,081  in  spring  and  1,369  in  fall. 
Our  capture  effort  increased  gradually  over 
the  spring  migratory  period  (as  days  became 
longer)  and  similarly  decreased  gradually  over 
the  early  part  of  fall  migration  (5  September- 
5 October).  Capture  effort  was  reduced  after 
5 October  and  only  12%  of  all  fall  net  hours 
occurred  after  this  date.  This  reduced  capture 
effort  late  in  fall  migration  may  have  reduced 
the  relative  number  of  captures  of  late  mi- 
grating kinglets,  mostly  males,  so  the  differ- 


Females  - 

(44) 


Males  - 

(60) 


SPRING 


10  Apr  20  Apr  30  Apr  10  May  20  May  30  May 


Date 

FIG.  2.  Distribution  of  capture  dates  (first  captures 
only)  for  male  and  females  Ruby-crowned  Kinglets 
during  spring  and  fall  migration.  Box  plots  display  the 
median  (vertical  line  within  box),  the  interquartile 
range  (extent  of  box),  the  10th-90th  percentile  range 
(extent  of  horizontal  lines),  and  the  data  points  falling 
below  the  10th  or  above  the  90th  percentiles  (dots). 
For  spring  females,  the  median  passage  date  and  the 
lower  interquartile  boundary  were  coincident  on  1 
May.  The  numbers  included  in  parenthesis  for  males 
and  females  represent  sample  sizes  for  each  season. 


ence  in  median  passage  dates  between  sexes 
in  fall  that  we  report  is  probably  conservative. 
Because  capture  effort  was  not  even  through- 
out migration  periods,  we  tested  whether  fre- 
quency distributions  of  sexes  differed  as  a 
function  of  cumulative  capture  effort.  Again, 
these  tests  indicated  that  males  migrated  sig- 
nificantly earlier  in  spring  (Kolmogorov-Smir- 
nov  test:  D = 0.673,  n = 104,  P < 0.001)  and 
significantly  later  in  fall  (D  = 0.425,  n = 132, 
P < 0.001)  than  females. 

Data  from  the  BBL  indicated  that  at  all  lat- 
itudes males  migrated  significantly  earlier  in 
spring  than  females  (Table  2).  Median  passage 
dates  differed  between  the  sexes  by  10-15 
days  in  the  different  5°  latitude  zones.  In  win- 
ter, BBL  data  indicated  that  males  had  a sig- 
nificantly higher  relative  abundance  than  fe- 
males at  30-35°  N (y^  = 4.06,  df  = 1,  P < 


Swanson  et  al.  • DIFFEREN  I’lAL  MIGRATION  IN  KINGLETS 


65 


TABLE  2.  Median  migration  dates  (5-day  intervals)  for  male  and  female  Ruby-crowned  Kinglets  in  eastern 
and  eentral  North  America  from  BBL  records  for  April  and  May  1985-1996.  Males  migrated  signiheantly  earlier 
(Kolmogorov-Smirnov  test:  F < O.OOl)  than  females  at  all  latitudes.  Sample  sizes  are  given  in  parentheses. 


Latitude-'  (°N)  Males  Females  I)  F 


30-34°  59' 

6-10  April  (69) 

16-20  April  (115) 

0.333 

< 0.001 

35-39°  59' 

16-20  April  (275) 

1-5  May  (319) 

0.566 

< 0.001 

40-44°  59' 

21-25  April  (1647) 

6-10  May  (1691) 

0.502 

< 0.001 

45-50° 

26-30  April  (230) 

6-10  May  (297) 

().45() 

< O.OOl 

® 5°  latitude  zones  to  the  south  and  north  of  those  above  had  loo  few  total  records  27)  to  adequately  quantity  medians  tor  each  sex. 


0.05)  and  35-40°  N (x“  - 18.25,  df  = 1,  P < 
0.001),  but  sexes  were  evenly  distributed 
south  of  30°  N latitude  (Table  1).  Thus,  male 
Ruby-crowned  Kinglets  winter  farther  north 
than  females. 

ANCOVA  revealed  that  slopes  for  equa- 
tions describing  observed  and  predicted  per- 
cent males  in  the  spring  migrant  population 
differed  significantly  (F,  4 = 9.46,  P = 0.037). 
Least-squares  regression  yielded  the  following 
equations; 

arcsin  Predicted  % Males  = 65.76  — 
5.22(Time  Interval)  (n  = 4,  P-  = 0.99,  P < 
0.001) 

arcsin  Observed  % Males  = 111.05  — 
19.54(Time  Interval)  (n  = 4,  P-  = 0.89,  P = 

0.05) 

where  Time  Interval  refers  to  successive  5- 
day  intervals  (1-4)  from  first  (15  April)  to  last 
(2  May)  capture  of  migratory  males.  For  20- 
24  April  a higher  percentage  of  males  was 
captured  than  predicted  (Fisher’s  exact  test,  P 
— 0.039)  and  for  30  April-4  May  a lower 
percentage  of  males  was  captured  than  pre- 
dicted (Fisher’s  exact  test,  P = 0.04);  signifi- 
cant differences  were  not  detected  for  other 
intervals.  These  data  indicate  that  migratory 
passage  of  the  sexes  during  the  early  season 
is  more  biased  toward  males  than  expected  on 
the  basis  of  their  wintering  distribution. 

Rates  of  migration  can  also  be  estimated  for 
males  and  females  by  dividing  the  distance 
between  two  successive  5°  latitude  zones  (as- 
suming 1°  equals  1 1 1 km)  by  the  number  of 
days  difference  between  median  passage  dates 
for  each  zone  as  determined  from  BBL  data 
(Ellegren  1990,  Woodrey  and  Chandler  1997). 
Estimating  migration  rates  in  this  manner 
shows  that  both  males  and  females  increase 
migration  speed  as  they  near  their  breeding 


grounds.  For  latitudes  south  of  our  study  sites 
migration  rates  for  males  and  females  were  74 
and  56  km  day',  respectively,  but  for  all  lat- 
itudes overall  migration  speed  was  83  km 
day'  for  both  sexes. 

Male  kinglets  were  significantly  heavier 
than  females  (t,02  = 5.76,  P < 0.001)  in 
spring.  Visible  fat  scores  did  not  differ  signif- 
icantly between  males  and  females  in  spring, 
but  spring  males  had  significantly  longer 
wings  (^102  = 9.49,  P < 0.001)  and  signifi- 
cantly higher  condition  index  (Mann- Whitney 
test:  Z102  = 3.20,  P = 0.001)  than  spring  fe- 
males (Table  3).  When  age  classes  were 
pooled  by  sex  in  fall,  males  were  significantly 
heavier  than  females  ((,30  = 4.80,  P < 0.001) 
and  had  significantly  longer  wings  (r,3o 
= 11.79,  P < 0.001).  Neither  visible  fat  nor 
condition  index  differed  significantly  between 
males  and  females  in  fall  when  age  classes 
were  pooled. 

In  fall,  when  AHY  data  after  30  September 
were  excluded  from  comparisons,  AHY  males 
(^38  = 3.39,  P = 0.002)  and  HY  males  (t,,  = 
3.93,  P < 0.001)  were  both  significantly 
heavier  than  HY  females.  Similarly,  when  all 
data  after  30  September  were  excluded,  AHY 
males  were  significantly  heavier  than  HY  fe- 
males (^24  = 3.035,  P = 0.006).  Mass  did  not 
vary  significantly  among  other  sex/age  classes 
in  fall  (Table  3).  Visible  furcular  fat  scores 
were  significantly  greater  in  AHY  females 
than  in  juveniles  in  fall  when  AHY  data  after 
30  September  were  excluded  from  compari- 
sons (Mann-Whitney  test;  Z41,  = 3.14,  P = 
0.002;  Z35  = 2.99,  P = 0.003  for  HY  females 
and  HY  males,  respectively).  Likewise,  when 
all  data  after  30  September  were  excluded 
from  analyses,  AHY  females  carried  signifi- 
cantly more  furcular  fat  than  HY  females 
(Mann-Whitney  test:  Z34  = 2.840,  P = 0.005). 


66 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  I,  March  1999 


TABLE  3.  Means  (±  SD)  of  mass,  wing  chord,  fat  class,  and  condition  index  (mass/wing  chord)  in  migrant 
Ruby-crowned  Kinglets.  Age  classes  (AHY  = adults,  HY  = hatching-year)  refer  to  fall  migrants. 


Sex/age  class  n Mass  (g)  Wing  chord  (nini)  Furcular  fat  Abdominal  fat  Condition  index 


Spring  males 

60 

6.6 

-h 

0.6 

57.7 

-F 

1.3 

2.2 

-F 

1.1 

2.1 

-F 

1.4 

1.15 

-F 

0.10 

Spring  females 

44 

6.0 

-h 

0.4 

55.0 

-F 

1.6 

2.1 

-F 

0.9 

2.1 

-F 

1.0 

1.09 

0.07 

Eall  males 

AHY 

5 

6.5 

-F 

0.6 

58.5 

-F 

1.5 

3.2 

-F 

0.8 

3.2 

-F 

0.8 

1.1 1 

-F 

0.10 

HY  (before  1 Oct.)"' 

5 

6.3 

-F 

0.4 

57.6 

-F 

1.4 

1.6 

-F 

1.1 

1.6 

-F 

1.1 

1.08 

-F 

0.06 

HY  (all  data)*’ 

22 

6.3 

-F 

0.4 

57.6 

-F 

1.2 

1.9 

-F 

1.0 

1.8 

-F 

1.0 

1.09 

-F 

0.07 

All  males'’ 

66 

6.4 

-F 

0.4 

57.7 

-F 

1.4 

2.0 

-F 

1.1 

2.1 

-F 

1.0 

1.10 

-F 

0.07 

Eall  females 

AHY 

15 

6.1 

-F 

0.4 

55.1 

-h 

1.4 

2.9 

-F 

0.7 

2.8 

-F 

0.7 

1.12 

-F 

0.07 

HY  (before  1 Oct.)" 

21 

5.9 

-F 

0.4 

54.8 

-F 

1.2 

2.0 

-F 

0.8 

2.0 

-F 

0.7 

1.08 

-F 

0.06 

HY  (all  data)*’ 

36 

5.9 

-F 

0.3 

54.7 

-F 

1.4 

2.0 

-F 

0.8 

2.1 

-F 

0.8 

1.08 

-F 

0.06 

All  females' 

66 

6.0 

-F 

0.3 

54.9 

-F 

1.3 

2.3 

-F 

0.8 

2.3 

-F 

0.8 

1.10 

-F 

0.06 

“ Data  for  HY  birds  prior  to  1 October  only. 

^ Data  for  HY  birds  over  the  entire  fall  season. 

^ Includes  birds  with  completely  ossified  skulls  after  30  September  that  were  of  indeterminate  age. 


Visible  abdominal  fat  scores  were  significant- 
ly greater  in  AHY  females  than  in  HY  birds 
when  AHY  data  after  30  September  were  ex- 
cluded from  comparisons  (Z35  = 3.11,  P = 
0.002;  Z49  = 2.84,  P = 0.005  for  HY  males 
and  HY  females,  respectively).  When  all  data 
after  30  September  were  excluded  from  com- 
parisons, AHY  females  had  significantly  more 
abdominal  fat  than  HY  females  (Z34  = 2.695, 
P = 0.007).  The  same  trend  occurred  for  AHY 
males,  although  differences  were  nonsignifi- 
cant (Table  3).  Neither  furcular  nor  abdominal 
fat  varied  significantly  between  males  and  fe- 
males within  the  same  age  class.  No  signifi- 
cant differences  among  age/sex  classes  were 
detected  for  condition  index  in  fall,  although 
when  male  and  female  data  for  each  age  class 
were  pooled  the  adult-juvenile  difference  ap- 
proached significance  = 1.74,  P > 0.05 
when  AHY  data  after  30  September  were  ex- 
cluded; ^43  = 1.85,  P > 0.05  when  all  data 
after  30  September  were  excluded). 

Wing  chord  did  not  differ  significantly 
among  age  classes  in  fall,  but  males  of  both 
age  classes  had  significantly  longer  wings 
than  females  when  AHY  data  after  30  Sep- 
tember were  excluded  (P  < 0.001)  and  when 
all  data  after  30  September  were  excluded  (P 
^ 0.005,  Table  3).  Test  statistics  for  between 
sex  wing  chord  comparisons  were;  AHY 
males  vs.  AHY  females  = 4.35),  AHY 
males  vs.  HY  females  = 5.59),  HY  males 
vs.  AHY  females  (/„  = 5.59),  and  HY  males 
vs.  HY  lemales  = 7.94)  for  comparisons 


with  AHY  data  after  30  September  excluded. 
For  comparisons  with  all  data  after  30  Sep- 
tember excluded,  between  sex  wing  chord 
comparison  test  statistics  were:  AHY  males 
vs.  AHY  females  (r.g  = 4.35),  AHY  males  vs. 
HY  females  (r24  = 5.99),  HY  males  vs.  AHY 
females  (6g  = 3.23),  and  HY  males  vs.  HY 
females  (U4  = 4.59). 

DISCUSSION 

Male  Ruby-crowned  Kinglets  migrated  ear- 
lier in  spring  than  females  in  both  southeast- 
ern South  Dakota  and  across  eastern  North 
America  (Crim  1976,  Fairfield  and  Shirokoff 
1978,  BBL  data).  Bird  Banding  Laboratory 
data  indicate  that  median  passage  dates  for  the 
latitudes  of  our  study  sites  (approximately 
43°  N)  in  eastern  North  America  are  21-25 
April  for  males  and  6—10  May  for  females. 
These  dates  agree  very  closely  with  median 
passage  dates  for  male  and  female  kinglets 
from  Iowa,  which  occurred  from  20-25  April 
for  males  and  from  5—10  May  for  females 
(Crim  1976).  The  median  date  for  males  in 
southeastern  South  Dakota  was  similar  to  me- 
dian dates  in  Iowa  and  to  median  dates  de- 
rived from  BBL  data  for  40-45°  N latitude  in 
central  and  eastern  North  America  (Table  2). 
The  median  date  for  females  in  our  study  is 
several  days  earlier  than  the  median  date  from 
Iowa  and  from  the  median  date  derived  from 
BBL  data  for  the  latitudes  of  our  study  sites. 
However,  capture  effort  was  not  quantified  in 
Crim  (1976)  and  was  not  available  for  BBL 


Swanson  et  al.  • DIFFERENTIAL  MIGRATION  IN  KINGLETS 


67 


data,  so  median  passage  dates  may  not  be  di- 
rectly comparable. 

Earlier  spring  passage  of  males  than  fe- 
males is  a common  pattern  among  passerines, 
presumably  because  of  the  advantages  that 
early  arrival  provides  to  the  sex  establishing 
territory,  which  among  passerines  is  usually 
the  male  (Gauthreaux  1982,  Francis  and 
Cooke  1986).  Postponing  arrival  on  the  breed- 
ing grounds  until  after  males  have  established 
territories  might  be  beneficial  to  females  if 
they  compete  for  mates  (Francis  and  Cooke 
1986).  Moreover,  because  temperatures  and 
food  availability  increase  throughout  spring  in 
northern  latitudes,  late  arrival  on  the  breeding 
grounds  might  also  benefit  females  by  provid- 
ing more  favorable  conditions  for  breeding. 
Consistent  with  this  latter  argument,  male 
kinglets  were  larger  (i.e.,  heavier  and  longer 
wings),  had  higher  condition  index,  and  were 
more  cold  tolerant  (Swanson  and  Dean  1999) 
than  females  in  spring. 

Because  juvenile  Ruby-crowned  Kinglets 
may  show  complete  skull  ossification  as  early 
as  1 October  (Leberman  1970),  documenta- 
tion of  differential  migration  patterns  of  age 
classes  in  fall  is  problematic.  In  addition,  we 
found  retrix  shape  (Pyle  et  al.  1987)  to  be  un- 
reliable for  aging  kinglets  in  our  population. 
Thus,  we  were  unable  to  compare  passage 
dates  for  age  classes  during  fall  migration. 
However,  males  migrated  significantly  later 
than  females  during  fall  migration  in  south- 
eastern South  Dakota.  This  pattern  of  males 
migrating  later  in  the  fall  than  females  is  con- 
sistent with  the  observations  of  Fairfield  and 
Shirokoff  (1978)  for  Ruby-crowned  Kinglets 
from  Ontario. 

Prescott  (1980)  found  that  adult  female 
kinglets  were  fatter  than  adult  males  in  fall  in 
New  Jersey,  but  sexes  did  not  differ  in  fat 
scores  during  fall  migration  in  our  study. 
Adult  kinglets  in  New  Jersey  (Prescott  1980) 
had  a higher  percentage  of  individuals  with 
either  no  fat  or  heavy  fat  than  juveniles,  while 
juveniles  had  higher  percentages  of  individu- 
als with  intermediate  fat  loads.  In  our  study 
males  were  heavier  than  juvenile  females  in 
fall,  but  not  adult  females.  Moreover,  adults 
in  our  study  carried  more  fat  than  juveniles  in 
fall,  significantly  so  for  females.  Woodrey  and 
Moore  (1997)  reviewed  several  possible  ex- 
planations for  elevated  fat  levels  in  adults 


compared  to  juveniles  during  migratory  stop- 
over. These  include:  ( 1 ) less  efficient  foraging 
by  juveniles  at  stopover  sites  because  of  in- 
experience or  lower  social  status,  (2)  juveniles 
carry  less  fat  to  increase  mobility  for  escape 
from  predators  because  they  may  be  more  vul- 
nerable to  predation,  and  (3)  possible  differ- 
ences in  migratory  routes  or  destinations  af- 
fecting fattening. 

Proximate  factors  regulating  differential 
timing  of  migration  among  sex/age  classes  in- 
clude differences  in  geographic  origin,  timing 
of  the  onset  of  migration,  and  rates  of  migra- 
tion (Chandler  and  Mulvihill  1990).  Fairfield 
and  Shirokoff  (1978)  analyzed  North  Ameri- 
can banding  data  from  1972-1975  and  found 
that  male  kinglets  winter  farther  north  than  fe- 
males on  average.  Bird  Banding  Laboratory 
data  from  1986-1995  also  indicate  that  male 
kinglets  winter  farther  north  than  females  (Ta- 
ble 1)  so  different  geographic  origins  un- 
doubtedly contribute  to  differential  migration 
of  the  sexes  in  the  spring.  However,  passage 
of  sexes  during  the  early  portion  of  spring  mi- 
gration is  more  biased  toward  males  than  ex- 
pected on  the  basis  of  the  differential  winter- 
ing distributions.  Sample  sizes  from  BBL  data 
for  1986-1995  for  wintering  populations  of 
kinglets  in  eastern  and  central  North  America 
were  relatively  small,  particularly  south  of 
30°  N latitude,  even  though  kinglets  regularly 
winter  south  of  this  latitude  (AOU  1983,  In- 
gold and  Wallace  1994).  Bird  Banding  Labo- 
ratory data  indicate  approximately  equal  num- 
bers of  males  and  females  at  southern  latitudes 
in  the  wintering  range,  whereas  females 
should  outnumber  males  in  the  southern  part 
of  the  wintering  range,  assuming  equal  sex  ra- 
tios on  breeding  grounds.  This  could  influence 
calculation  of  predicted  percent  males  during 
the  latter  part  of  the  spring  migration  of  males, 
but  should  not  affect  comparisons  during  the 
early  portion  of  the  migration.  Because  a 
higher  percentage  of  males  was  observed  than 
predicted  based  on  wintering  distribution  dur- 
ing the  early  part  of  spring  migration,  the 
small  sample  size  from  southern  portions  of 
the  wintering  range  should  not  influence  our 
conclusions  that  wintering  distributions  do  not 
fully  account  for  differential  migration  of  the 
sexes  in  spring. 

This  suggests  that  males  either  migrate  at  a 
faster  rate  or  initiate  migration  earlier  than  fe- 


68 


THE  WILSON  BULLETIN 


Vol.  Ill,  No.  1,  March  1999 


males.  Calculation  of  spring  migratory  rates 
from  BBL  data  revealed  that  male  kinglets  mi- 
grate at  a faster  pace  than  females  south  of 
45°  N latitude,  but  that  rates  over  the  entire 
migratory  range  do  not  differ  between  sexes. 
Thus,  elevated  rates  of  migration  in  males  rel- 
ative to  females  may  contribute  to  their  early 
arrival  at  our  study  sites,  but  not  at  higher 
latitudes.  This  suggests  that  the  onset  of 
spring  migration  is  earlier  in  male  Ruby- 
crowned  Kinglets  than  in  females  and  that  this 
difference  contributes  substantially  to  differ- 
ential spring  migration  of  the  sexes  in  this 
species. 

For  fall  migration,  where  all  sex/age  classes 
presumably  initiate  migration  from  breeding 
grounds,  differential  timing  of  migration 
should  reflect  differences  in  onset  or  rate  of 
migration,  assuming  that  sex/age  classes  ex- 
hibit similar  migration  routes  and  that  little 
postbreeding  dispersal  away  from  breeding 
sites  occurs  prior  to  southward  migration. 
Fairfield  and  Shirokoff  (1978)  suggested  that 
differential  migration  of  the  sexes  in  fall  re- 
sults from  earlier  departure  of  females  from 
breeding  grounds  for  two  possible  reasons. 
First,  females  are  smaller  and  presumably  less 
hardy  than  males,  so  they  might  depart  before 
food  availability  and  temperatures  decline 
markedly  in  the  fall.  However,  cold  tolerance 
of  male  and  female  kinglets  during  fall  mi- 
gration through  South  Dakota  did  not  differ 
significantly  (Swanson  and  Dean  1999).  Sec- 
ond, females  might  depart  earlier  than  males 
because  they  winter  farther  south  and,  there- 
fore, must  travel  farther.  If  we  assume  that  all 
sex/age  classes  initiate  migration  from  breed- 
ing grounds,  then  differential  migration  of 
sex/age  classes  might  result  not  only  from  dif- 
ferences in  departure,  but  also  from  differences 
in  rates  of  migration  (Chandler  and  Mulvihill 
1990).  Because  Fairfield  and  Shirokoff  (1978) 
did  not  directly  test  whether  early  departure 
or  rates  of  migration  differed  between  the  sex- 
es in  fall,  their  suggestion  must  remain  ten- 
tative. In  our  study  males  migrated  later  than 
females  in  fall  but  comparisons  among  age 
classes  were  confounded  because  juvenile 
kinglets  could  not  safely  be  distinguished 
from  adults  after  1 October  (Leberman  1970). 

ACKNOWLEDGMENTS 

This  work  was  supported,  in  part,  by  a grant  from 
the  U..S.  Fish  and  Wildlife  Service  Neotropical  Migra- 


tory Bird  Program  to  DLS.  We  thank  the  many  vol- 
unteers who  helped  us  with  mist-netting,  prominent 
among  them  were  M.  Dutenhoffer,  J.  Martin,  and  D. 
Weinacht.  We  also  thank  K.  Olmstead  for  help  with 
statistics,  and  J.  Ingold,  M.  Woodrey,  and  S.  Morris 
for  helpful  comments  on  an  earlier  version  of  this  man- 
uscript. J.  Ingold  also  provided  access  to  hard-to-find 
literature.  Finally,  we  thank  K.  Klimkiewicz  of  the 
BBL  for  providing  access  to  banding  records  for  king- 
lets. 

LITERATURE  CITED 

American  Ornithologists’  Union.  1983.  Checklist  of 
North  American  birds,  sixth  ed.  Allen  Press, 
Lawrence,  Kansas. 

Chandler,  C.  R.  and  R.  S.  Mulvihill.  1990.  Inter- 
preting differential  timing  of  capture  of  sex  clas- 
ses during  spring  migration.  J.  Field  Ornithol.  61: 
85-89. 

Crim,  G.  B.  1976.  Kinglets  in  Iowa.  Iowa  Bird  Life 
76:45-47. 

Ellegren,  H.  1990.  Autumn  migration  speed  in  Scan- 
dinavian Bluethroats  Luscinia  .s.  svecica.  Ringing 
& Migr.  11:121-131. 

Fairfield,  D.  M.  and  P.  A.  Shirokoff.  1978.  Migra- 
tory patterns  and  winter  distribution  in  the  Ruby- 
crowned  Kinglet  (Regiilus  calendula).  Blue  Bill 
(suppl.)  25:22-25. 

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  Biol.  6: 
93-168. 

Hall,  G.  A.  1981.  Fall  migration  patterns  of  wood 
warblers  in  the  southern  Appalachians.  J.  Field 
Ornithol.  52:43-49. 

Helms,  C.  W.  and  W.  H.  Drury,  Jr.  1960.  Winter  and 
migratory  weight  and  fat  field  studies  on  some 
North  American  buntings.  Bird-Banding  3 1 : 1-40. 
Hussell,  D.  j.  T.  1991.  Fall  migrations  of  Alder  and 
Willow  flycatchers  in  southern  Ontario.  J.  Field 
Ornithol.  62:260-270. 

Ingold,  J.  L.  and  G.  E.  Wallace.  1994.  Ruby- 
crowned  Kinglet  (Regains  calendula),  no.  1 19.  In 
The  birds  of  North  America  (A.  Poole  and  F.  Gill, 
Eds.).  The  Academy  of  Natural  Sciences,  Phila- 
delphia, Pennsylvania;  The  American  Ornitholo- 
gists’ Union,  Washington,  D.C. 

Ketter.son,  E.  D.  and  V.  Nolan,  Jr.  1976.  Geographic 
variation  and  its  climatic  correlates  in  the  sex  ratio 
of  eastern-wintering  Dark-eyed  Juncos  (.lunco 
hyemalis  hyemalis).  Ecology  57:679-693. 
Leberman,  R.  C.  1970.  Pattern  and  timing  of  skull 
pneumatization  in  the  Ruby-crowned  Kinglet-. 
Bird-Banding  41: 121-124. 

Lepthein,  L.  W.  and  C.  E.  Bock.  1976.  Winter  abun- 
dance patterns  of  North  American  kinglets.  Wil- 
•son  Bull.  88:483-485. 

Morton,  M.  L.  1984.  Sex  and  age  ratios  in  wintering 
White-crowned  Sparrows.  Condor  86:85-87. 


Swanson  el  al.  • DIFFERENTIAL  MIGRATION  IN  KINGLETS 


69 


Myers,  J.  P.  1981.  A test  of  three  hypotheses  for  lat- 
itudinal segregation  of  the  sexes  in  wintering 
birds.  Can.  J.  Zool.  59:1527-1534. 

Nolan,  V.,  Jr.  and  E.  D.  Ketterson.  1990.  Timing  of 
autumn  migration  and  its  relation  to  winter  distri- 
bution in  Dark-eyed  Juncos.  Ecology  71:1267- 
1278. 

Prescott,  D.  R.  C.  and  A.  L.  A.  Middleton.  1990. 
Age  and  sex  differences  in  winter  distribution  ot 
American  Goldfinches  in  eastern  North  America. 
Omis  Scand.  21:99-104. 

Prescott,  K.  W.  1980.  Weight,  fat  class,  and  wing 
measurements  of  Ruby-crowned  Kinglets  during 
migration.  Inland  Bird  Banding  52:1—7. 

Pyle.  P,  S.  N.  G.  Howell,  R.  P.  Yunick,  and  D.  F. 
DeSante.  1987.  Identification  guide  to  North 
American  passerines.  Slate  Creek  Press,  Bolinas, 
California. 

Ramos,  M.  A.  1988.  Eco-evolutionary  aspects  of  bird 
movements  in  the  northern  Neotropical  region. 
Proc.  Int.  Omithol.  Congr.  19:251-293. 

Rice,  W.  R.  1989.  Analyzing  tables  of  statistical  tests. 
Evolution  43:223-225. 

Root,  T.  1988.  Atlas  of  wintering  North  American 
birds.  Univ.  of  Chicago  Press,  Chicago,  Illinois. 


SAS  Institute.  1988.  SAS  user’s  guide.  SAS  Institute, 
Inc.,  Cary,  North  Carolina. 

SOKAL,  R.  R.  AND  F.  J.  Rohle.  1995.  Biometry,  third 
ed.  Freeman,  New  York. 

South  Dakota  Ornithologists’  Union.  1991.  The 
birds  of  South  Dakota,  second  ed.  Northern  State 
Univ.  Press,  Aberdeen,  South  Dakota. 

Swanson,  D.  L.  and  K.  L.  Dean.  1999.  Migration- 
induced  variation  in  thermogenic  capacity  in  mi- 
gratory passerines.  J.  Avian  Biol.  30:In  press. 

Winker,  K.  1995.  Autumn  stopover  on  the  Isthmus  ol 
Tehuantepec  by  woodland  Nearctic-Neotropic  mi- 
grants. Auk  112:690—700. 

Winker,  K.  and  J.  H.  Rappole.  1992.  The  autumn  pas- 
sage of  Yellow-bellied  Flycatchers  in  South  Texas. 
Condor  94:526-529. 

WOODREY,  M.  S.  AND  C.  R.  CHANDLER.  1997.  Age- 
related  timing  of  migration:  geographic  and  inter- 
specific patterns.  Wilson  Bull.  109:52-67. 

WooDREY,  M.  S.  AND  F.  R.  MooRE.  1997.  Age-related 
differences  in  the  stopover  of  fall  landbird  mi- 
grants on  the  coast  of  Alabama.  Auk  1 14:695— 
707. 

Zar,  j.  H.  1996.  Biostatistical  analysis,  third  ed.  Pren- 
tice Hall,  Upper  Saddle  River,  New  Jersey. 


Wilson  Bull.,  111(1),  1999,  pp.  70-75 


SCALE-DEPENDENT  HABITAT  SELECTION  BY  AMERICAN 
REDSTARTS  IN  ASPEN-DOMINATED  FOREST  FRAGMENTS 

NAVJOT  S.  SODHI,'  23  CYNTHIA  A.  PASZKOWSKI,'  AND  SHANNON  KEEHN' 


ABSTRACT. — We  examined  scale-dependent  site  occupancy  of  American  Redstarts  {Setophaga  ruticilla)  in 
forest  fragments  (2-140  ha  in  area)  dominated  by  mature  trembling  aspen  (Populus  tremuloides)  in  central 
Alberta,  Canada.  Vegetation  within  territories,  both  adjacent  to  and  away  from  nests,  differed  from  vegetation 
in  unoccupied  fragments  and  within  occupied  fragments  outside  of  territories.  Territories  contained  higher  den- 
sities of  willow  (Salix  sp.),  a taller  shrub-layer,  and  lower  densities  of  trembling  aspen  than  other  sites.  Willow 
was  the  most  frequently  used  plant  species  for  nesting  and  foraging.  Our  results  indicate  that  even  within  mature 
forest  patches,  American  Redstarts  select  disturbed  areas  dominated  by  early  successional  plant  species.  Received 
6 March  1998,  accepted  30  Aug.  1998. 


Numerous  researchers  have  investigated 
habitat  selection  by  American  Redstarts  {Se- 
tophaga  ruticilla),  but  debate  continues  re- 
garding the  preferred  age  of  forest  stands  used 
for  breeding.  Although  most  studies  reported 
that  American  Redstarts  prefer  early  succes- 
sional forests  (Martin  1960,  Collins  et  al. 
1982,  DeGraaf  1991,  Westworth  and  Telfer 
1993,  Huffman  1997),  they  have  also  been 
commonly  found  in  mid-  to  late  successional 
stands  (Bond  1957,  Ficken  and  Ficken  1967, 
Crawford  et  al.  1981,  Morgan  and  Freedman 
1986,  Thompson  and  Capen  1988).  Scale-de- 
pendent  habitat  requirements  have  been  poor- 
ly documented  for  the  American  Redstart.  Be- 
cause abiotic  and  biotic  factors  may  affect  a 
species  differently  at  various  spatial  scales 
(Bock  1987,  Wiens  et  al.  1987),  some  docu- 
mented differences  in  habitat  selection  may  be 
due  to  scale  effects.  Thus,  a holistic  under- 
standing of  a species’  habitat  use  patterns  may 
require  that  data  be  collected  and  analyzed  at 
several  spatial  scales  (e.g.,  patch,  territory, 
and  nest  site;  Bergin  1992,  Steele  1992,  Kelly 
1993).  Considering  that  many  forest  patches 
are  heterogeneous  in  age  and  structure  (e.g., 
woody  vegetation  may  increase  in  age  and 
height  from  edge  to  interior),  the  investigation 
of  scale-dependent  habitat  use  patterns  of  the 
American  Redstart  could  clarify  the  species’ 
requirements  and  reconcile  some  of  the  incon- 
sistencies in  reported  habitat  preferences. 


' Dept,  of  Biological  Sciences,  Univ.  of  Alberta,  Ed- 
monton, AB.  Canada  T6G  2E9. 

^ Prc.sent  address:  Dept,  of  Biological  Sciences,  Na- 
tional Univ.  of  Singapore,  Kent  Ridge,  Singapore 
1 19260:  Email:  dbsns(®mis.edu.sg 
' Corresponding  author. 


We  studied  scale-dependent  habitat  selec- 
tion of  the  American  Redstart  in  forest  frag- 
ments dominated  by  mature  trembling  aspen 
{Populus  tremuloides)  at  the  northwestern 
edge  of  the  species’  range  in  central  Alberta, 
Canada.  We  asked:  (1)  do  American  Redstarts 
occupy  available  forest  fragments  based  on 
particular  habitat  characteristics  (vegetation 
structure  and  composition),  and  (2)  within  an 
occupied  fragment,  do  American  Redstarts  se- 
lect territories  and  nest  sites  based  on  habitat 
characteristics? 

STUDY  AREA  AND  METHODS 

The  study  was  conducted  around  the  Meanook  Bi- 
ological Research  Station  (54°  37'  N,  1 13°  20'  W)  near 
Athabasca,  Alberta,  between  9 May  and  6 July  1994. 
Upland  forests  in  this  region  are  primarily  a mosaic  of 
trembling  aspen  and  white  spruce  (Picea  glauca) 
stands  which  was  historically  created  and  maintained 
by  fire  (Rowe  1972).  We  examined  eight  forest  frag- 
ments dominated  by  mature  (>80  yr  old)  trembling 
aspen  interspersed  with  patches  of  willow  {Salix  sp.) 
and  alder  (Alnus  spp.).  A fragment  was  defined  as  a 
wooded  area  separated  from  other  wooded  areas  by 
more  than  30  m of  cropland  or  pasture  (see  Villard  et 
al.  1995).  Eorest  fragments  varied  from  2 to  140  ha 
(areas  were  calculated  from  1:30,000  aerial  photo- 
graphs using  a Placon®  digital  planimeter).  All  frag- 
ments were  located  within  a 9-km^  area  and  repre- 
sented the  range  of  fragment  sizes  available  in  the 
landscape  (for  details  see  Sodhi  and  Paszkowski  1997). 
The  selection  of  fragments  was  constrained  by  various 
factors  including  stand  age,  access,  presence  of  cattle, 
and  landowner  cooperation. 

Each  fragment  was  flagged  into  100  X 100  m grids 
and  was  surveyed  three  times  to  locate  redstart  terri- 
tories. During  surveys,  we  walked  along  the  flagged 
gridlines  and  visually  located  all  redstarts  (primarily 
by  following  singing  males).  Locations  of  redstarts, 
along  with  characterization  of  individual  plumage  pat- 
terns (see  below)  were  recorded  on  a gridded  map. 


70 


Sodhi  et  al.  • REDSTART  HABITAT  SELECTION 


71 


Unoccupied  fragmenls  were  surveyed  again  in  late 
June  by  playing  American  Redstart  territorial  songs  I'or 
2 min  at  50-m  intervals  and  waiting  5 min  tor  a re- 
sponse. No  new  individuals  were  detected  during  these 
surveys.  Individual  males  were  identified  based  on  lo- 
cation of  ten  itory  and  on  patterns  of  plumage  colora- 
tion on  the  breast,  abdomen,  neck,  and  head  (Sherry 
and  Holmes  1989,  Lemon  et  al.  1992).  Because  second 
year  male  redstarts  can  differ  from  older  males  in  hab- 
itat selection  (e.g..  Sherry  and  Holmes  1989,  Hunt 
1996),  we  collected  habitat  data  only  for  males  in  full 
adult  plumage. 

The  composition  of  vegetation  was  measured  within 
territories,  outside  of  territories,  adjacent  to  the  nest 
tree,  and  in  unoccupied  habitat  fragments.  We  sampled 
in  early  to  mid-July  after  the  breeding  season.  Vege- 
tation plots  were  located  at  random  with  the  following 
exceptions:  plots  a,ssociated  with  nest  trees  were  dic- 
tated by  nest  location  and  plots  associated  with  teiri- 
tories  were  always  placed  at  least  10  m inside  a bound- 
ary. A total  of  88  vegetation  plots  were  sampled:  33 
within  13  territories,  17  outside  of  temtories  but  within 
4 occupied  fragments,  17  immediately  around  17  nest 
trees,  and  21  plots  in  4 unoccupied  fragments. 

We  sampled  vegetation  using  a modification  of  the 
circular  sample-plot  method  (James  and  Shugart 
1970).  For  each  of  the  four  site  types,  we  recorded 
four  sets  of  information:  (1)  tree  (woody  plants  more 
than  1.75  m tall)  species,  number,  height  (using  a cli- 
nometer), and  diameter  at  breast  height  (using  a dbh 
measuring  tape)  within  a 22.4  m diameter  circular  plot, 
(2)  shrub  (woody  plant  less  than  1.75  m tall)  species, 
number,  and  maximum  height  within  a 10  m diameter 
circular  plot.  (3)  presence/absence  of  canopy  (above  5 
m)  at  20  random  points  within  the  22.4  m diameter 
plot  using  a 4 cm  diameter  ocular  tube,  and  (4)  pres- 
ence/absence of  ground  cover  at  the  same  20  random 
points  using  the  ocular  tube.  The  last  two  sets  of  in- 
formation gave  a measure  of  relative  canopy  and 
ground  cover,  respectively. 

We  used  Principal  Components  Analysis  (PCA)  to 
compare  vegetation  composition  within  and  among  the 
four  site-types  (CANOCO;  ter  Braak  1991)  tor  all  88 
plots.  Principal  Components  Analysis  is  recommended 
and  frequently  used  for  analysis  of  habitat  data  (Don- 
caster et  al.  1996,  Hunt  1996).  The  27  vegetation  var- 
iables recorded  for  each  plot  (Table  1)  were  square- 
root  transformed  to  approximate  a normal  distribution. 
A univariate  analysis,  Kruskal-Wallis  test,  was  used  to 
determine  if  PCA  scores  for  the  first  two  ordination 
axes  differed  among  site  types.  If  the  Kruskal-Wallis 
test  was  significant  (P  < 0.05),  Multiple  Comparisons 
tests  were  performed  to  determine  which  characteris- 
tics differed  (Siegel  and  Castellan  1988).  All  statistical 
tests  were  done  using  STATVIEW  version  4.1  on  a 
Macintosh-compatible  computer. 

RESULTS 

Redstarts  occurred  in  four  forest  fragments 
(2,  50,  107  and  140  ha  in  area)  but  were  ab- 


sent from  four  others  (4,  6,  9 and  32  ha).  Prin- 
cipal Components  Analysis  based  on  vegeta- 
tion characteristics  clearly  grouped  and  sepa- 
rated unoccupied  plots  from  plots  within  ter- 
ritories (Fig.  1).  Loadings  of  vegetation 
variables  on  PCA  Axis  1 indicated  that  unoc- 
cupied fragments  and  areas  outside  of  terri- 
tories within  occupied  fragments  contained 
larger  trees  and  higher  densities  of  trembling 
aspen,  accompanied  by  prickly  wild  rose 
(Rosa  acicularis),  red-osier  dogwood  {Cornus 
stolonifera),  and  honeysuckle  (Lonicera  spp.) 
(Table  1).  Areas  within  territories  were  char- 
acterized by  high  densities  of  willow  (both 
trees  and  saplings),  cherry  (Prunus  sp.),  sas- 
katoon berry  (Amelanchier  alnifolia)  and  bal- 
sam poplar  {Populus  balsamifera)  saplings,  as 
well  a tall  shrub-layer  (loadings  on  Axis  1, 
Table  1).  Scores  on  Axis  I differed  signifi- 
cantly among  site  types  {KW  = 42.51,  df  = 
3,  P < 0.001).  There  was  no  significant  dif- 
ference between  nest  sites  and  plots  within 
territories  (Multiple  Comparisons  test:  P > 
0.05),  but  both  differed  significantly  from 
plots  in  unoccupied  fragments  and  plots  out- 
side of  territories  within  occupied  fragments 
(Multiple  Comparisons  tests:  P < 0.05).  Plots 
in  unoccupied  fragments  and  those  outside  of 
territories  in  occupied  fragments  had  similar 
PCA  I scores  (Multiple  Comparisons  test:  P 
> 0.05). 

Nest  sites  were  associated  with  high  den- 
sities of  white  birch  {Betula  papyrifera,  both 
trees  and  saplings),  willow,  cherry,  and  bal- 
sam poplar  saplings,  as  well  as  Viburnum 
spp.,  which  contributed  to  a tall  shrub  layer 
(high  scores  and  loadings  on  Axis  II).  Nest 
sites  were  also  clustered  on  PCA  Axis  II  {KW 
= 17.79,  df  = 3,  P < 0.001).  On  this  axis, 
only  scores  for  nest  sites  differed  significantly 
from  scores  for  plots  from  unoccupied  frag- 
ments (Multiple  Comparisons  tests,  P < 0.05). 
However,  there  was  similarity  in  vegetation 
characteristics  in  some  plots  within  territories 
and  those  at  nest  sites  (Fig.  1). 

DISCUSSION 

Principal  Components  Analysis  and  subse- 
quent univariate  analyses  indicate  that  red- 
starts did  not  establish  territories  randomly 
within  or  among  forest  patches,  but  based 
their  occupancy  on  vegetation  characteristics. 
A high  abundance  of  willow  was  a particular- 


72 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  I,  March  1999 


TABLE  1.  Loadings  of  vegetation  variables  on 
PC  II). 

the  first  and  second  Principal  Components  axes  (PC  I and 

Variables 

PC  1 

PC  11 

Eigenvalues 

0.42 

0.12 

Variation  explained  (%) 

42.4 

11.1 

No.  trees  (woody  plants  > 1.75  ni  tall) 

Willow’  (Sali.x  sp.) 

-121 

-289 

Trembling  aspen  (Populit.s  tremuloides) 

338 

-654 

Balsam  poplar  {P.  halsamifera) 

38 

19 

White  birch  (Betula  papyrifera) 

-101 

556 

Alder  (Alnu.s  spp.) 

29 

195 

White  spruce  {Piceti  glauca) 

28 

133 

Tree  sizes 

Mean  tree  height  (m) 

-95 

-68 

Mean  dbh  (mm) 

203 

318 

No.  shrubs  (woody  plants  0.5-1.75  m tall) 

Red-osier  dogwood  (Cornu.'i  .stolonifera) 

224 

106 

Wild  gooseberry  (Rihe.s  o.xyacanthoide.'i) 

-55 

103 

Wild  red  raspberry  (Ruhii.s  idaeu.s) 

75 

-229 

Honeysuckle  (Lonicera  spp.) 

183 

-418 

Prickly  wild  rose  {Rosa  acicidaris) 

996 

81 

Snowberry  (Symphoricarpos  spp.) 

203 

88 

Caragana  sp. 

-74 

-121 

Viburnum  spp. 

91 

522 

Primus  spp. 

-223 

448 

Saskatoon  berry  (Amelanchier  alnifolia) 

-139 

-5 

Maple  sapling  {Acer  sp.) 

-11 

0 

Balsam  poplar  sapling 

-122 

266 

Willow  .sapling 

-158 

380 

Trembling  a.spen  sapling 

39 

196 

White  birch  sapling 

-119 

769 

Alder  sapling 

142 

172 

Shrub  height 

Mean  maximum  shrub  height  (mm) 

-206 

563 

Cover' 

Canopy 

-97 

177 

Ground 

-37 

5 

^ See  Methods  for  estimation  of  these  variables. 


ly  distinctive  feature  of  areas  within  territories 
and  around  nest  sites.  Willow  was  the  most 
frequently  used  foraging  site  (mean  ± SE  = 
65.8  ± 8.9%  of  observation  time,  n ~ 13 
males  for  which  we  had  at  least  10  min  of 
observations)  and  nesting  site  (13  of  20  nests; 
Sodhi  and  Paszkowski,  unpubl.  data).  Areas 
within  territories  and  nest  sites  were  charac- 
terized not  only  by  willows,  but  also  by  a tall 
shrub  layer  in  general.  American  Redstarts 
have  been  found  to  be  positively  associated 
with  shrub  cover  in  Minnesota  (Huffman 
1997).  The  extra  cover  and  structural  com- 
plexity provided  by  taller  shrubs  may  have 
protected  nests  from  predation  or  parasitism 
and  thus  enhanced  reproductive  success  (Mor- 


ris and  Lemon  1987).  Nest  sites  also  support- 
ed higher  densities  of  white  birches,  which 
have  wider  leaves  than  willows  and  may  im- 
prove nest  cover.  In  our  study  area,  only  18% 
{n  = 22)  of  the  nests  were  parasitized  by 
Brown-headed  Cowbirds  (Molothrus  ater)  and 
only  18%  {n  = 22)  were  depredated  (Sodhi 
and  Paszkowski,  unpubl.  data).  Both  nest  par- 
asitism and  depredation  rates  at  our  sites  were 
relatively  low  compared  to  other  studies  (e.g., 
Freedman  1929,  Sherry  and  Holmes  1992)! 
The  structural  complexity  of  understory  veg- 
etation in  earlier  successional  stages  could  of- 
fer appropriate  nest  protection  for  redstarts. 

Certain  plant  species,  most  notably  trem- 
bling aspen,  were  underrepresented  on  sites 


Sodhi  et  al.  • REDSTART  HABITAT  SELECTION 


73 


400 


300 


200 


0) 

o 

o 

(/) 


100  -- 


o 

0. 


0 


-100 


-200 


-300  -I— 
-250 


o 


ONest  Site  (n  = 17) 

□ Within  Territories  (n  = 33) 
^Outside  Territories  (n  = 17) 

• Unoccupied  Fragments  (n  = 21) 


A 


-200  -150  -100  -50  0 50  100  150  200  250 

PC  I Score 


FIG.  1 . Comparison  of  sites  occupied  and  unoccupied  by  American  Redstarts  based  on  the  first  two  Principal 
Component  scores  of  plot  vegetation.  The  Principal  Components  are  based  on  27  vegetation  variables  collected 
from  88  plots.  Note:  one  plot  near  a nest  was  the  only  plot  dominated  by  conifers  and  it  contained  little 
undergrowth,  therefore  it  is  distinct  from  all  other  plots. 


occupied  by  redstarts.  The  lack  of  use  of  hab- 
itat dominated  by  common,  and  even  domi- 
nant, tree  and  shrub  species  at  both  the  scale 
of  the  forest  patch  and  the  territory  may  be 
related  to  constraints  imposed  on  redstart  for- 
aging behavior  by  the  leaf  and  branch  mor- 
phology of  these  plant  species  (Holmes  et  al. 
1978,  Sedgewick  and  Knopf  1992).  The  ex- 
clusion and  inclusion  of  certain  shrub  species 
on  redstart  territories  could  also  be  an  artifact 
of  their  co-occurrence  with  overstory  domi- 
nants, especially  trembling  aspen.  The  canopy 
forming  woody  species  associated  with  habitat 
used  by  redstarts,  i.e.,  willow  and  white  birch, 
are  all  species  typically  found  on  wetter  sites 
in  central  Alberta  (Rowe  1972).  Thus,  it  is 
also  possible  that  redstarts  were  choosing  lo- 
cations in  this  relatively  dry  landscape  that 
had  higher  soil  moisture  levels  and  were 
therefore  more  productive  in  terms  of  plant 
and  invertebrate  biomass  (Adams  and  Morri- 
son 1993).  Our  surveys  showed  that  even 
within  stands  dominated  by  mature  trembling 


aspen,  redstarts  prefered  areas  that  supported 
more  willows,  which  are  typically  early  suc- 
cessional  species.  The  presence  of  thickets  of 
willow  and  other  shrubs  in  these  stands  might 
be  maintained  by  periodic  localized  flooding 
resulting  from  vernal  soil  saturation  following 
heavy  snow  cover  or  from  beaver  activity  (So- 
dhi and  Paszkowski,  pers.  obs.). 

American  Redstarts  are  abundant  in  har- 
vested aspen  forests  in  Minnesota  (Hoffman 
1997).  Other  studies  also  show  that  redstarts 
attain  maximum  densities  in  stands  in  early 
successional  stages  of  different  forest  types 
(DeGraaf  1991,  Westworth  and  Telfer  1993, 
Hunt  1996).  It  might  be  argued,  based  on  pat- 
terns reported  for  some  other  passerines 
(Vickery  et  al.  1992),  that  although  redstarts 
occur  at  high  densities  in  early  successional 
forests,  they  reproduce  poorly  here  compared 
to  mid-  to  late  successional  forests.  We  found 
that  90.6%  (n  = 32)  of  older  adult  redstart 
males  were  paired  in  our  study  area  in  1994 
(Sodhi,  unpubl.  data);  this  figure  is  similar  to 


74 


THE  WILSON  BULLETIN 


Vol.  ///,  No.  /,  March  1999 


values  for  older  males  in  various  successional 
forests  and  in  a continuous  forest  tract  in  New 
Hampshire  (Sherry  and  Holmes  1989,  Hunt 
1996).  Assuming  that  male  pairing  success  is 
correlated  with  reproductive  success,  pockets 
of  disturbed,  productive  early  successional 
vegetation  nested  within  mature  forest  stands 
may  offer  good  breeding  conditions  for  red- 
starts. 

Every  avian  habitat  can  be  represented  as  a 
spatially  based  hierarchy  that  ranges  from  the 
level  of  landscape  to  nest  site.  The  behavioral 
decision-making  processes  behind  habitat  use 
can  be  related  to  these  hierarchical  habitat 
units  (Kolasa  1989,  Kotliar  and  Wiens  1990). 
Studies  of  scale-dependent  habitat  selection 
have  revealed  that  the  decision-making  pro- 
cess varies  among  species  and  can  sometimes 
operate  at  multiple  scales  within  a single  spe- 
cies (Bergin  1992).  For  example.  Black- 
throated  Blue  Warblers  (Dendroica  ccientles- 
cens)  were  more  selective  at  the  habitat  patch 
(or  stand)  level  than  at  the  territory  level 
(Steele  1992).  In  contrast,  habitat  selection  in 
Dusky  Flycathers  {Empidoncix  oberholseri) 
was  primarily  based  on  nest-site  characteris- 
tics rather  than  on  territorial  features  (Kelly 

1993) .  In  our  study,  habitat  selection  by 
American  Redstarts  operated  most  clearly  at 
the  territory  level.  Males  defended  areas  of 
forest  with  a distinct  plant  composition  and 
structure.  Habitat  selection  at  the  territorial 
level  was  manifested,  in  turn,  at  the  next  high- 
er spatial  scale,  as  males  did  not  occupy  frag- 
ments that  contained  insufficient  appropriate 
vegetation.  At  the  finest  spatial  scale,  evi- 
dence for  nest  site  selection  was  present  but 
relatively  weak.  Nest  sites  were  not  signifi- 
cantly different  from  other  locations  within 
territories,  possibly  because  the  willows  and 
other  tall  shrubs  used  as  foraging  sites  also 
offered  good  nest  protection. 

Some  que.stion  remains  as  to  whether  birds 
actually  differentiate  between  potential  terri- 
tories ba.sed  on  the  environmental  parameters 
that  researchers  perceive  to  be  important 
(Morse  1989).  Most  authors  agree  however, 
that,  at  some  level,  territory  establishment  is 
a behavioral  response  to  certain  vegetation 
characteristics  (Maurer  and  Whitmore  1981, 
Smith  and  Shugart  1987,  Parrish  and  Sherry 

1994) .  We  concur  that  American  Redstarts  in 
central  Alberta  establish  territories  within  a 


definable  type  of  vegetation  that  appears  to 
offer  appropriate  conditions  for  successful  for- 
aging and  nesting.  Based  on  our  research,  as 
well  as  previous  studies,  redstarts  appear  to 
prefer  early  successional  forest  stands  or  dis- 
turbed sites  embedded  within  older  stands. 
However,  with  correlative  data,  it  remains 
possible  that  redstarts  choose  sites  based  on 
soil  moisture  or  food  availability,  which  are 
in  turn  associated  with  particular  vegetation 
characteristics.  In  light  of  reported  population 
declines  for  the  species  in  parts  of  its  range 
(Sauer  et  al.  1996),  the  value  of  such  habitats 
should  be  assessed  on  a regional  basis  and 
their  use  by  redstarts  integrated  into  conser- 
vation strategies. 

ACKNOWLEDGMENTS 

We  thank  S.  Jamieson  for  field  assistance.  This 
study  was  supported  by  the  Alberta  Sports,  Recreation, 
Parks  and  Wildlife  Foundation,  Canadian  Circumpolar 
Institute,  Natural  Sciences  and  Engineering  Research 
Council  of  Canada  (grant  to  CAP),  and  James  Ander- 
son McAfee  Postdoctoral  Fellowship  (to  NSS).  We 
thank  two  anonymous  reviewers  for  making  construc- 
tive comments  on  an  earlier  draft. 

LITERATURE  CITED 

Adams.  E.  M.  and  M.  L.  Morrison.  1993.  Effects  of 
forest  stand  structure  and  composition  on  Red- 
breasted Nuthatches  and  Brown  Creepers.  J. 
Wildl.  Manage.  57:616-629. 

Bergin,  T.  1992.  Habitat  selection  by  the  Western 
Kingbird  in  western  Nebraska:  a hierarchical  anal- 
ysis. Condor  94:903-91  1. 

Bock,  C.  1987.  Distribution-abundance  relationships 
of  some  Arizona  land  birds:  a matter  of  scale? 
Ecology  68:124-129. 

Bond,  R.  R.  1957.  Ecological  distribution  of  breeding 
birds  in  upland  forests  of  southern  Wisconsin. 
Ecol.  Monogr.  27:351-384. 

Collins,  S.  L.,  E C.  James,  and  P.  G.  Ris.ser.  1982. 
Habitat  relationships  of  wood  warblers  (Parulidae) 
in  northern  central  Minnesota.  Oikos  39:50-58. 
Crawford,  H.  S.,  R.  G.  Hooper,  and  R.  W.  Titter- 
INGTON.  1981.  Songbird  population  response  to 
silvicultural  practices  in  central  Appalachian  hard- 
woods. J.  Wildl.  Manage.  45:680-692. 

DeGraaf,  R.  M.  1991.  Breeding  bird  assemblages  in 
managed  northern  hardwood  forests  in-  New  Eng- 
land. Pp.  153-171  in  Wildlife  and  habitats  in  man- 
aged landscapes  (J.  E.  Rodiek  and  E.  G.  Bolen, 
Eds.).  Island  Press,  Covelo,  California. 

Donca.ster,  C.  R,  T.  Micol,  and  S.  P.  Jensen.  1996. 
Determining  minimum  habitat  requirements  in 
theory  and  practice.  Oikos  75:335-339. 

Ficken,  M.  S.  and  R.  W.  Ficken.  1967.  Age-specific 
differences  in  the  br'eeding  behavior  and  ecology 


Sodhi  et  al.  • REDSTART  HABITAT  SELECTION 


75 


of  the  American  Redstart.  Wilson  Bull.  79:188- 
199. 

Freedman,  H.  1929.  The  cowbird:  a study  of  social 
parasitism.  Charles  C.  Thomas,  Springfield,  Illi- 
nois. 

Holmes,  R.  T,  T.  W.  Sherry,  and  S.  E.  Bennett. 
1978.  Diurnal  and  individual  variability  in  the  for- 
aging behavior  of  American  Redstarts.  Oecologia 
36:141-149. 

Huffman,  R.  D.  1997.  Effects  of  residual  overstory  on 
bird  use  and  aspen  regeneration  in  aspen  harvest 
sites  in  Tamarac  National  Wildlife  Refuge,  Min- 
nesota. M.Sc.  thesis.  West  Virginia  Univ.,  Mor- 
gantown. 

Hunt,  R D.  1996.  Habitat  selection  by  American  Red- 
starts along  a successional  gradient  in  northern 
hardwoods  forest:  evaluation  of  habitat  quality. 
Auk  113:875-888. 

James,  F.  C.  and  H.  H.  Shugart.  1970.  A quantitative 
method  of  habitat  description.  Audubon  Field 
Notes  24:727-736. 

Kelly,  J.  P.  1993.  The  effect  of  nest  predation  on  hab- 
itat selection  by  Dusky  Flycatchers  in  limber  pine- 
juniper  woodland.  Condor  95:83—93. 

Kolasa,  J.  1989.  Ecological  systems  in  hierarchical 
perspective:  breaks  in  community  structure  and 
other  consequences.  Ecology  70:36—47. 

Kotliar,  N.  B.  and  j.  a.  Wiens.  1990.  Multiple  scales 
of  patchiness  and  patch  structure:  a hierarchical 
framework  for  the  study  of  heterogeneity.  Oikos 
59:253-260. 

Lemon,  R.  E.,  D.  M.  Weary,  and  K.  J.  Norris.  1992. 
Male  morphology  and  behavior  correlates  with  re- 
productive success  in  the  American  Redstart  (Se- 
tophaga  ruticilla).  Behav.  Ecol.  Sociobiol.  29: 
399-403. 

Martin,  N.  D.  1960.  An  analysis  of  bird  populations 
in  relation  to  forest  succession  in  Algonquin  Pro- 
vincial Park,  Ontario.  Ecology  41:126-140. 

Maurer,  B.  A.  and  R.  C.  Whitmore.  1981.  Foraging 
of  five  bird  species  in  two  forests  with  different 
vegetation  structure.  Wilson  Bull.  93:478—490. 

Morgan,  K.  and  B.  Freedman.  1986.  Breeding  bird 
communities  in  a hardwood  forest  succession  in 
Nova  Scotia.  Can.  Field-Nat.  100:506—519. 

Morris,  M.  M.  J.  and  R.  E.  Lemon.  1987.  American 
Redstart  nest  placement  in  southwestern  New 
Brunswick.  Can.  J.  Zool.  66:212—216. 

Morse,  D.  H.  1989.  American  warblers.  Harvard  Univ. 
Press,  Cambridge,  Massachusetts. 

Parrish,  J.  D.  and  T.  W.  Sherry.  1994.  Sexual  seg- 
regation by  American  Redstarts  wintering  in  Ja- 
maica: importance  of  resource  seasonality.  Auk 
111:38-49. 

Rowe,  J.  F.  1972.  Forest  regions  of  Canada.  Depart- 


ment of  Environment.  Can.  For.  Serv.  Publ.  1300: 
1-171. 

Sauer,  J.  R.,  S.  Schwartz,  B.  G.  Peterjohn,  J.  E. 
Hines,  and  S.  Droege.  1996.  The  North  American 
breeding  bird  survey.  Version  95.1.  Patuxent 
Wildlife  Research  Center,  Laurel,  Maryland.  URL 
= http://www.nbs. go  v./bbs/bbs.html. 

Sedgewick,  j.  a.  and  F L.  Knopf.  1992.  Describing 
Willow  Flycatcher  habitats:  scale  perspectives  and 
gender  differences.  Condor  94:720-733. 

Siegel,  S.  and  N.  J.  Castellan,  Jr.  1988.  Nonpara- 
metric  statistics  for  the  behavioral  sciences.  Mc- 
Graw-Hill Book  Company,  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.  Behav.  Ecol.  Sociobiol.  25: 
327-333. 

Sherry,  T.  W.  and  R.  T.  Holmes.  1992.  Population 
fluctuation  in  a long-distance  neotropical  migrant: 
demographic  evidence  for  the  importance  of 
breeding  season  events  in  the  American  Redstart. 
Pp.  431-442  in  Ecology  and  conservation  of  neo- 
tropical migrant  landbirds  (J.  M.  Hagan  and  D.  W. 
Johnston,  Eds.).  Smithsonian  Institution  Press, 
Washington,  D.C. 

Smith,  T.  M.  and  H.  H.  Shugart.  1987.  Territory  size 
variation  in  the  Ovenbird:  the  role  of  habitat  struc- 
ture. Ecology  68:695-704. 

Sodhi,  N.  S.  and  C.  A.  Paszkowski.  1997.  The  pairing 
success  of  male  Black-and-white  Warblers,  Mni- 
otilta  varia,  in  forest  fragments  and  a continuous 
forest.  Can.  Field-Nat.  111:457-458. 

Steele,  B.  B.  1992.  Habitat  selection  by  breeding 
Black-throated  Blue  Warblers  at  two  spatial 
scales.  Ornis  Scand.  23:33-42. 

TER  Braak,  C.  j.  E 1991.  Program  CANOCO,  version 
3.2.  Agriculture  Mathematics  Group,  DLO,  AC 
Wageningen,  The  Netherlands. 

Thompson,  F.  R.,  Ill  and  D.  E.  Capen.  1988.  Avian 
assemblages  in  serai  stages  of  a Vermont  forest. 
J.  Wildl.  Manage.  52:771-777. 

Vickery,  P.  D.,  M.  L.  Hunter.  Jr.,  and  J.  W.  Wells. 
1992.  Is  density  an  indicator  of  breeding  success? 
Auk  109:706-710. 

ViLLARD,  M-A.,  G.  Merriam.  and  B.  a.  Maurer. 
1995.  Dynamics  in  subdivided  populations  of  neo- 
tropical migratory  birds  in  fragmented  temperate 
forest.  Ecology  76:27—40. 

Wiens,  J.  A..  J.  T.  Rotenberry,  and  B.  Van  Horne. 
1987.  Habitat  occupancy  patterns  of  North  Amer- 
ican shnibstcppe  birds:  the  effect  of  spatial  scale. 
Oikos  48:132-147. 

Westworth,  D.  a.  and  E.  S.  Telfer.  1993.  Summer 
and  winter  bird  populations  associated  with  five 
age-classes  of  aspen  forest  in  Alberta.  Can.  J.  For- 
est. Res.  23:1830-1836. 


Wilson  Bull.,  1 1 1(1),  1999,  pp.  76-83 


FEMALE  MATE  CHOICE  IN  NORTHERN  CARDINALS:  IS  THERE 
A PREFERENCE  FOR  REDDER  MALES? 

L.  LAREESA  WOLEENBARGER*  ^ 


ABSTRACT. — 1 tested  whether  female  Northern  Cardinals  iCardinalis  cardinalis)  associated  with  redder 
males  in  two  laboratory  experiments,  one  using  males  with  unaltered  plumage  and  the  other  using  males  with 
plumage  altered  by  a lightened  or  reddened  treatment.  Females  exhibited  no  preference  for  redder  males.  Given 
the  long  duration  over  which  pair  formation  can  occur  in  natural  populations  and  the  importance  of  territory 
quality  to  reproductive  success,  a female  may  choose  a mate  based  on  other  morphological  characteristics  or 
aspects  of  his  territory  rather  than  on  only  red  coloration.  Received  27  May  1998,  accepted  22  Oct.  1998. 


Ornate  plumage  characteristics  in  birds  are 
generally  assumed  to  have  arisen  through  sex- 
ual selection  and,  specifically,  through  female 
mating  preferences  or  competitive  interactions 
between  males  (Darwin  1871,  Andersson 
1994).  A growing  number  of  studies  has  dem- 
onstrated that  females  prefer  males  with 
brighter  coloration  (Hill  1990,  Sastre  et  al. 
1994,  Sundberg  1995)  as  well  as  with  other 
exaggerated  plumage  characteristics,  such  as 
longer  crests  (Jones  and  Hunter  1993)  and 
longer  tails  (Andersson  1982,  1992;  Mpller 
1988,  1992;  Evans  and  Hatch  well  1992). 

Northern  Cardinals  {Cardinalis  cardinalis) 
are  highly  dichromatic.  Males  vary  from  or- 
ange to  scarlet  red.  Male  coloration  varies 
both  in  overall  hue  and  in  evenness  of  breast 
coloration.  In  contrast,  females  are  primarily 
a light  brown  but  have  small  and  variable 
amounts  of  red  in  the  crest  and  breast;  all  fe- 
males also  have  red  coloration  on  the  wings 
and  tail. 

In  addition  to  being  highly  dichromatic, 
cardinals  are  territorial  and  socially  monoga- 
mous. Males  feed  females  substantially  during 
nest  construction,  a time  coinciding  with  egg 
production  (Kinser  1973).  Males  also  feed 
nestlings  at  higher  rates  than  do  females  (Fil- 
liater  and  Breitwi.sch  1997,  but  see  Linville  et 
al.  1998).  Because  coloration  derived  from  ca- 
rotenoids is  dependent  on  diet  in  birds  (Good- 
win 1950,  Hill  1992),  red  coloration  in  car- 
dinals could  signal  information  to  the  female 


' Section  of  Neurobiology  and  Behavior,  Cornell 
Univ.,  Ithaca,  NY  14853-2702. 

^ Present  address:  Dept,  of  Biology,  Univ.  of  Mary- 
land. College  Park.  MD  20742; 

E-mail:  LW 1 37@umail.umd.edu 


regarding  a male’s  foraging  abilities,  age,  or 
overall  phenotypic  quality. 

In  this  study,  I address  whether  female  pref- 
erences may  account  for  the  maintenance  of 
red  coloration  in  male  Northern  Cardinals.  I 
used  two  laboratory  experiments  to  test 
whether  female  cardinals  preferred  to  associ- 
ate with  redder  males  during  the  breeding  sea- 
son. 

METHODS 

The  experiments  were  conducted  on  the  Cornell 
University  campus,  Ithaca,  New  York,  from  mid  Feb- 
ruary through  April  1995.  Males  and  females  used  in 
the  experiments  were  captured  between  14  January 
and  16  April  1995  using  baited  traps  and  mist  nets  at 
6 sites  within  Tompkins  County,  New  York  (42°  N, 
76°  W).  Individuals  were  marked  uniquely  with  one 
color  band.  Because  previous  studies  have  demonstrat- 
ed that  red  color  bands  can  influence  behavior  of  in- 
dividuals in  other  species  (Burley  et  al.  1982,  Hagan 
and  Reed  1988,  Metz  and  Weatherhead  1991),  I did 
not  use  red  or  orange  color  bands. 

Prior  to  trials  males  were  housed  in  two  indoor  flight 
aviaries  (3.6  X 3.6  X 3.0  m)  with  skylights  and  incan- 
descent lighting  synchronized  to  dawn  and  dusk.  Fe- 
males were  housed  in  a similar  third  aviary  (5.2  X 3.6 
X 3.0  m).  Aviary  rooms  were  both  visually  and  acous- 
tically isolated  from  each  other.  The  temperature  of 
each  aviary  ranged  between  10  and  12°  C.  Food  and 
water  were  provided  ad  libitum. 

Measuring  coloration. — To  measure  male  colora- 
tion, I used  methods  described  elsewhere  (Wolfenbar- 
ger  in  press).  Briefly,  I used  the  color  chip  series  of 
the  Methuen  Handbook  of  Color  (Kornerup  1967)  to 
quantify  breast  coloration  of  males.  The  color  chip  se- 
ries provides  a measure  of  three  components  of  color: 
hue,  tone,  and  intensity.  The  hue  component  ranked 
color  on  a scale  from  yellow  (5)  to  intense  red  (11). 
The  tone  component  quantified  the  amount  of  black 
present  1 1 (all  black)  to  6 (none)],  the  intensity  com- 
ponent indicated  the  degree  of  saturation  of  pigment 
[from  little  ( 1 ) to  complete  (8)|.  I used  a grid  that 
divided  the  breast  into  eight  1 X 4 em  rectangles  and 


76 


Wolfenhcirfier  • RED  COLORATION  AND  FEMALE  PREFERENCE 


77 


TABLE  1.  Male  color  and  morphology  differences  in 

male  cardinals  used 

in  natural  trials. 

Relative  color 

score 

Lower  .f  ± SD  (n) 

Higher  .f  ± .SD  (h) 

Z (Cf 

Bright  breast  hue 

73.9  ± 7.0  (8) 

81.7  ± 4.3  (8) 

2.81  (0.005) 

Dull  breast  hue 

65.9  ± 6.4  (8) 

71.9  ± 4.0  (8) 

2.81  (0.005) 

Bright  breast  intensity 

62.9  ± 0.9  (3) 

63.2  ± 0.6  (3) 

b 

Dull  breast  intensity 

61.1  ± 3.1  (7) 

62.7  ± 0.9  (7) 

2.64  (0.008) 

Morphological  traits  (in  mm): 

Tarsus  length 

24.3  ± 0.7  (10) 

24.1  ± 1.0  (10) 

0.82  (>  0.05) 

Crest  length 

36.8  ± 3.5  (10) 

35.8  ± 1.9  (10) 

0.46  (>  0.05) 

Tail  length 

99.9  ± 5.0  (10) 

98.4  ± 3.7  (10) 

0.46  (>  0.05) 

Black  bib  length 

19.8  ± 4.6  (10) 

20.2  ± 4.3  (10) 

0.65  (>  0.05) 

Black  bib  width 

21.8  ± 1.1  (10) 

22.0  ±1.1  (10) 

0.46  (>  0.05) 

“Z-value  is  from  Wilcoxon  matched  pairs  signed  rank  test. 

^ In  7 trials,  male  scores  for  bright  breast  intensity  were  identical. 


placed  it  immediately  posterior  to  the  black  bib  and 
measured  coloration  in  each  of  these  regions.  Because 
male  cardinals  often  had  clumps  of  orange,  yellow,  or 
even  tan  feathers  interspersed  within  a background  of 
red  feathers,  I recorded  the  highest  (“Bright”  breast 
score)  and  lowest  color  score  (“Dull”  breast  score)  for 
hue,  tone  and  intensity  in  each  of  the  8 regions.  For 
all  of  these  components  a higher  score  indicates  a red- 
der or  brighter  color.  Among  the  birds  I used,  there 
was  greater  variation  among  males  in  the  color  of  dull 
regions  relative  to  bright  regions  (Wolfenbarger,  in 
press).  I tested  the  repeatability  of  the  color  scoring 
method  by  using  specimens  in  the  Cornell  Vertebrate 
Collections,  and  found  the  method  to  be  highly  re- 
peatable (Wolfenbarger  1996). 

For  analyses  of  the  association  between  measures  of 
female  association  and  male  coloration,  I summed  each 
component  of  color  (i.e.,  hue,  tone,  intensity)  for  the 
eight  regions  and  used  separate  scores  for  the  bright 
and  dull  color  measurements.  In  this  experiment,  all 
males  except  two  exhibited  the  maximum  tone  scores 
possible  for  both  bright  and  dull  breast  measurements; 
therefore,  the  relationship  between  color  tone  and  fe- 
male association  was  not  tested.  Similarly,  there  was 
very  little  variation  in  bright  breast  intensity.  There- 
fore, I used  three  variables;  bright  breast  hue,  dull 
breast  hue,  and  dull  breast  intensity  to  characterize 
male  coloration. 

Although  other  studies  have  created  a composite 
score  from  color  chips  (Linville  et  al.  1998),  1 used 
these  components  separately  for  two  reasons.  Hue  and 
intensity  were  not  consistently  correlated  in  this  pop- 
ulation of  cardinals  (Wolfenbarger  1996),  and  combin- 
ing these  scores  would  result  in  identical  scores  for 
males  that  actually  had  different  hue  and  intensity 
scores.  Second,  the  relative  importance  of  hue,  inten- 
sity and  tone  for  color  perception  of  cardinals  is  not 
known.  Any  weighting  of  these  components  becomes 
problematic  for  the  interpretation  of  negative  results 
since  one  obvious  alternative  would  be  that  the  weight- 
ing factor  might  be  incorrect. 

Morphological  measurements. — 1 measured  the  fol- 


lowing on  males:  tarsus  length,  tail  length,  crest  length, 
maximum  length  of  black  bib  (after  Mpller  1987),  and 
width  of  black  bib.  All  were  measured  to  the  nearest 
0.1  mm  except  tail  length  which  was  measured  to  the 
nearest  0.5  mm. 

Experiment  I:  natural  plumage  trials. — Ten  trials 
were  conducted  to  test  whether  females  spent  more 
time  with  males  having  higher  natural  plumage  scores. 
Males  and  females  in  each  trial  were  captured  at  least 
4 km  apart  to  reduce  the  possibility  that  males  and 
females  had  interacted  prior  to  the  experiment.  Males 
and  females  were  tested  in  the  order  that  they  were 
captured  from  the  field  so  that  males  spent  similar 
lengths  of  time  in  captivity  (approximately  7 days  for 
natural  trials  and  18  days  for  manipulated  trials),  but 
within  a trial  males  and  females  had  spent  similar 
amounts  of  time  in  captivity.  No  apparent  change  in 
behavior  was  associated  with  when  males  and  females 
were  captured  or  how  long  they  spent  in  captivity. 

In  each  trial  males  differed  primarily  in  hue  scores. 
Within  a trial  males  differed  significantly  in  bright  and 
dull  breast  hue  (Wilcoxon  matched-pairs  test;  Bright 
hue;  Z = 2.81,  P < 0.05;  Dull  hue:  Z = 2.81.  P < 
0.05;  Table  1).  There  were  also  significant  but  small 
differences  in  dull  breast  intensity  between  males  in  a 
trial  (Z  = 2.64,  P < 0.05;  Table  1).  The  lack  of  dif- 
ferences in  bright  breast  intensity  (Table  1 ) reflects  the 
limited  natural  variation  of  breast  intensity  in  male  car- 
dinals. 1 analyzed  bright  and  dull  breast  measurements 
separately  because  male  rankings  based  on  bright  and 
dull  measurements  were  not  identical. 

Trials  were  conducted  in  a rectangular  experimental 
aviary  (5.0  X 1.4  X 1.75  m)  divided  by  netting  into 
three  main  compartments.  Males  were  placed  at  op- 
posite ends  of  the  aviary  and  a female  was  placed  in 
the  center  compartment  (2.5  X 1.4  X 1.75  m).  The 
female  compartment  was  divided  into  three  equal  areas 
so  that  females  could  associate  with  either  male  or 
spend  time  in  a middle  area  where  she  could  not  in- 
teract visually  with  either  male.  Opaque  baniers  within 
the  female  area  were  used  to  divide  the  compartment. 
These  prevented  a male  from  observing  a female  in- 


78 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


TABLE  2.  Male  color  and  morphology  differences  in 

male  cardinals  used  in 

manipulated  trials. 

Lightened  male 

Reddened  male 

X ± SD  (n) 

^±50  (n) 

Z (P)“ 

Color  score  before  manipulation: 

Bright  breast  hue 

75.7  ± 7.4  (10) 

80.5  ± 6.5  (10) 

1.89  (0.059) 

Dull  breast  hue 

68.9  ± 4.9  (10) 

70.9  ± 3.5  (10) 

1.99  (0.046) 

Bright  breast  intensity 

63.0  ± 0.7  (10) 

63.1  ± 0.9  (10) 

0.30  (>  0.05) 

Dull  breast  intensity 

61.5  ± 3.1  (10) 

62.3  ± 0.9  (10) 

0.34  (>  0.05) 

Color  score  after  manipulation^: 

Bright  breast  hue 

72.9  ± 5.4  (9) 

86.1  ± 4.3  (9) 

2.69  (0.0072) 

Dull  breast  hue 

66.6  ± 6.4  (8) 

77.3  ± 4.8  (8) 

2.52  (0.011) 

Bright  breast  intensity 

62.8  ± 0.9  (5) 

63.8  ± 0.7  (5) 

0.68  (>  0.05) 

Dull  breast  intensity 

61.2  ± 0.9  (9) 

63.0  ± 0.8  (9) 

2.69  (0.0072) 

Morphological  traits  (in  mm): 

Tarsus  length 

24.0  ± 0.9  (10) 

24.4  ± 0.8  (10) 

0.83  (>  0.05) 

Crest  length 

35.4  ± 2.5  (10) 

37.1  ± 2.9  (10) 

1.17  (>  0.05) 

Tail  length 

99.7  ± 4.3  (10) 

98.6  ± 4.5  (10) 

0.35  (>  0.05) 

Black  bib  length 

20.4  ± 4.1  (10) 

19.6  ± 4.8  (10) 

0.56  (>  0.05) 

Black  bib  width 

21.7  ± 0.9  (10) 

22.1  ± 1.3  (10) 

0.82  (>  0.05) 

“ Z-value  is  from  Wilcoxon  matched  pairs  signed  rank  test. 

Males  that  have  identical  scores  are  not  included  in  means  of  color  scores  after  manipulation. 


teracting  with  the  male  on  the  opposite  side,  but  did 
not  restrict  female  movement.  I considered  a female  as 
interacting  with  a male  when  she  spent  time  in  the  area 
adjacent  to  the  male  compartment  as  opposed  to  the 
center  area.  To  minimize  the  possibility  of  side  pref- 
erences, all  aspects  of  the  male  compartments  as  well 
as  the  female  compartment  were  symmetrical.  A single 
perch  was  provided  in  each  male  compartment  and 
three  perches  in  the  female  compartment,  one  in  each 
area.  Food  and  water  were  provided  ad  libitum  in  each 
male  compartment  and  in  the  middle  area  of  the  female 
compartment. 

Each  trial  lasted  29  hours.  For  each  trial  a female 
was  placed  in  the  experimental  aviary  for  a one  hour 
acclimation  period  without  the  presence  of  males.  A 
male  was  introduced  to  each  side  of  the  aviary  and  the 
experiment  proceeded  for  26  hours.  At  the  end  of  26 
hours,  the  positions  ot  the  two  males  were  switched 
and  the  female  interacted  with  males  for  an  additional 
two  hours. 

Females  were  videotaped  during  four  different  ob- 
servation periods:  ( 1 ) for  one  hour  with  no  males  pres- 
ent (Pretrial  Period),  (2)  for  the  first  two  hours  (0  Hour 
Period)  after  introduction  of  the  males,  (3)  from  24- 
26  hours  after  introduction  of  the  males  (24  Hour  Pe- 
riod), and  (4)  from  26-28  hours  of  the  trial  (Reversed 
Period).  In  sum,  each  female  was  taped  for  6 hours 
with  the  two  males  present  and  for  one  hour  prior  to 
the  introduction  of  males.  After  completion  of  a trial, 
males  and  females  were  returned  to  their  respective 
aviaries. 

E.xperiment  2:  manipulated  plumaf’e  trials. — The  30 
individuals  used  in  the  natural  plumage  experiment 
were  also  used  in  the  manipulated  plumage  experi- 
ment. but  within  each  experiment  individuals  were 
used  only  once.  Males  and  females  for  each  trial  were 
unfamiliar  with  each  other  because  they  were  housed 


in  different  aviaries  prior  to  the  trial  and  were  intro- 
duced in  novel  combinations  for  the  two  experiments. 
For  each  pair  of  males  in  a manipulated  trial,  one  was 
randomly  assigned  to  a reddened  plumage  treatment 
and  the  other  to  a lightened  plumage  treatment.  Trials 
proceeded  as  in  experiment  one.  At  the  conclusion  of 
the  manipulated  trials,  all  individuals  were  released  at 
the  original  site  of  capture. 

Plumage  manipulation  methods. — Prior  to  the  ma- 
nipulation, males  were  anesthetized  with  1.5  mg/kg  of 
Midazolam  injected  into  the  pectoralis  muscle.  This 
dosage  induced  a state  of  light  anesthesia:  males  closed 
their  eyes  but  opened  them  in  response  to  external 
stimuli  such  as  having  a wing  extended.  No  mortality 
was  associated  with  using  the  anesthesia. 

For  the  “reddened”  treatment,  a mixture  of  1 part 
Divina  20  Volume  Creme  Developer  and  3 parts  Clai- 
rol Professional  Hi  Power  Tint  670®  was  placed  on  the 
feathers  for  25  minutes.  A “lightened”  treatment  con- 
sisted of  applying  Clairol  Professional  7th  Stage 
Creme  Hair  Lightener  mixed  with  Divina  20  Volume 
Creme  Developer  in  a 1:1  mixture  to  the  feathers  for 
25  minutes.  In  both  treatments  males  were  rinsed  and 
were  dried  with  a hair  dryer.  The  males  were  alert 
within  an  hour  of  the  injection  and  were  released  back 
into  an  aviary  within  2 hours.  Trials  were  conducted 
at  least  four  days  after  the  males  were  manipulated. 

Manipulated  plumage  coloration  scores. — Although 
treatments  were  assigned  randomly,  plumage  hue 
scores  before  the  manipulation  were  significantly  high- 
er for  males  in  the  reddened  treatment  group  (Wilcox- 
on matched-pairs  test:  Natural  bright  hue:  Z = 1 .89,  P 
= 0.06;  Natural  dull  hue:  Z = 1.99,  P < 0.05;  Table 
2).  There  was  no  significant  difference  in  the  intensity 
scores  prior  to  the  manipulation  (Natural  bright  inten- 
sity. Z = 0.30,  P > 0.05,  Natural  dull  intensity:  Z = 
0.034,  P > 0.05;  Table  2).  After  plumage  manipula- 


\V<>lfenhari>er  • RED  COLORATION  AND  FEMALE  PREFERENCE 


79 


tions,  differences  between  line  seores  for  reddened  and 
lightened  treatment  groups  were  significant  (Manipu- 
lated bright  hue:  Z = 2.69.  P < {).05,  Manipulated  dull 
hue:  Z = 2.52,  P < ().()5;  Table  2).  Males  in  the  red- 
dened treatment  had  signihcantly  higher  intensity  for 
the  dull  breast  measurement  but  not  for  the  bright  mea- 
surement (Manipulated  bright  intensity:  Z = 0.68,  P > 
0.05;  Manipulated  dull  intensity:  Z = 2.69,  P < 0.05; 
Table  2).  Little  variation  in  manipulated  tone  seores 
e.xisted:  all  except  two  indi\'iduals  had  the  maximum 
possible  tone  scores.  Reflectance  spectra  (from  280- 
750  nm)  of  manipulated  plumage  were  within  the 
range  of  natural  variation  in  plumage  reflectance  that 
occurs  at  these  wavelengths  (Wolfenbarger,  unpubl. 
data).  As  in  experiment  one.  dull  and  bright  breast 
measurements  were  analyzed  separately. 

Analyses  of  experiments. — For  the  periods  video- 
taped, the  time  a female  spent  in  each  area  of  the  ex- 
perimental aviary  was  measured  either  during  the  trial 
via  a monitor  connected  to  one  camera  or  after  the  trial 
from  the  videotapes.  The  monitor  was  located  in  a 
room  visually  and  acoustically  isolated  from  the  ex- 
perimental aviary.  For  each  sampling  period.  1 deter- 
mined the  number  of  trials  in  which  the  female  spent 
more  time  with  the  male  having  the  higher  color  score 
for  each  period  videotaped  (0  Flour.  24  Hour,  Re- 
versed). I used  the  0 Hour  and  24  Hour  samplings  to 
assess  female  preferences  and  the  pretrial  and  reversed 
sampling  periods  to  assess  whether  females  had  site 
preferences  in  the  aviary. 

For  the  0 Hour  and  24  Hour  periods,  I used  a one- 
tailed  binomial  test  to  determine  whether  the  number 
of  trials  in  which  a female  spent  more  time  with  the 
redder/brighter  male  was  significantly  greater  than  ex- 
pected by  chance  (50%;  Conover  1980).  The  test  sta- 
tistic T refers  to  the  number  of  trials  in  which  the 
female  spent  more  time  with  the  male  with  the  higher 
color  score  (Conover  1980). 

Because  males  were  randomly  assigned  to  trials, 
there  were  trials  in  which  hue  or  intensity  scores  were 
identical.  I eliminated  these  from  analyses  because  nei- 
ther male  was  redder  or  brighter  using  my  measure- 
ments. I also  eliminated  .sampling  periods  in  which  the 
difference  in  amount  of  time  spent  with  males  was  less 
than  2 minutes  (i.e..  the  female  showed  no  preference 
for  a particular  male). 

Among  trials,  differences  in  male  coloration  scores 
varied  widely  (range  in  differences;  bright  breast  = 0- 
21.  dull  breast  hue  = 0-24,  dull  breast  intensity  = 0- 
8).  Females  may  exhibit  strong  preferences  (as  mea- 
sured by  time  spent  with  male)  when  differences  be- 
tween male  coloration  are  large;  whereas  females  may 
spend  equal  amounts  of  time  with  males  whose  color 
scores  are  similar.  1 used  regression  analysis  to  test 
whether  the  magnitude  t)f  the  difference  in  color  was 
related  to  the  difference  in  the  time  a female  spent  with 
a particular  male.  In  particular,  I tested  whether  there 
was  a positive  relationship  between  the  difference  in 
color  scores  between  the  two  males  and  the  difference 
in  the  amount  of  time  spent  between  the  two  males  in 
the  0 Hour  and  24  Hour  .sampling  periods  (Wilkinson 


(n  = 8)  (n  = 8)  (n  = 7) 

FIG.  1.  Number  of  natural  trials  in  which  female 
spent  more  time  with  male  having  lower  or  higher  col- 
or score  during  the  0 Hour  Sampling  period  (Binomial 
test:  all  P > 0.05). 


et  al.  1992).  Because  the  predicted  difference  in  the 
amount  of  time  spent  with  males  of  the  same  color  is 
zero,  the  regression  line  was  forced  through  the  origin. 

I tested  whether  females  exhibited  two  types  of  po- 
tential site  preferences;  ( 1 ) a general  preference  among 
females  for  either  the  east  or  west  side  of  the  experi- 
mental aviary  or  (2)  the  likelihood  that  an  individual 
female  stayed  on  the  same  side  of  the  aviary  between 
consecutive  sampling  periods.  I used  the  pretrial  and 
reversed  periods  to  assess  the  consistency  of  females’ 
preferences  for  a particular  side  or  male.  Two-tailed 
binomial  tests  were  used  to  determine  whether  females 
remained  on  the  same  side  of  the  aviary  between  sub- 
sequent sampling  periods  in  more  than  half  of  the  trials 
(Conover  1980).  I used  a sequential  Bonferroni  ad- 
justment for  multiple  comparisons  (Rice  1989)  because 
I used  both  dull  and  bright  breast  measures  in  analyses. 
A Wilcoxon  matched  pairs  signed  rank  test  was  used 
to  compare  moiphological  variables  between  paired 
males  (Wilkinson  et  al.  1992). 

RESULTS 

Natural  and  manipulated  plumage  experi- 
ment.— Of  the  59  observation  periods  during 
the  two  experiments,  females  spent  an  equal 
amount  of  time  (±2  min)  with  both  males  in 
only  7 periods.  During  the  remaining  52  pe- 
riods, females  spent  an  average  of  45.8  (SE  = 
4.0)  minutes  more  with  one  male  than  the  oth- 
er (range  = 12-110  min,  total  possible  = 120 
min). 

In  the  0 Hour  and  24  Hour  sampling  peri- 
ods, females  were  as  likely  to  associate  with 
the  relatively  dull  males  as  with  brighter,  red- 
der males  (Fig.  1,  binomial  test:  0 Hour:  T 
Bright  hue  = 3,  P > 0.05,  n—  8;  T Dull  hue 
= 3,  P > 0.05,  n = 8;  T Dull  Intensity  = 4, 
P > 0.05,  n = 7;  24  Hour:  T Bright  hue  = 5, 
P > 0.05,  n=  9;  T Dull  hue  = 3,  P > 0.05, 


80 


THE  WILSON  BULLETIN 


Vol.  Ill,  No.  1,  March  1999 


FIG.  2.  Number  of  manipulated  trials  in  which  fe- 
male spent  more  time  with  males  having  lower  or 
higher  color  .scores  during  the  0 Hour  Sampling  period 
(Binomial  test;  all  P > 0.05). 

/2  = 9;  T Dull  Intensity  — 6,  P > 0.05,  n = 
7).  Similarly,  in  the  manipulated  plumage  tri- 
als where  color  differences  between  males 
were  greater,  females  did  not  consistently  as- 
sociate with  males  having  higher  coloration 
scores  (Fig.  2;  binomial  test:  0 Hour:  T Bright 
hue  = 3,  P > 0.05,  n = 9;  T Dull  hue  = 4, 
P > 0.05,  n = 8;  T Dull  intensity  = 2,  P > 
0.05,  n = 9;  24  Hour:  T Bright  hue  = 3,  P > 
0.05,  n = 9;  T Dull  hue  = 4,  P > 0.05,  n = 
8;  T Dull  intensity  = 2,  P > 0.05,  n = 9). 

The  power  of  binomial  tests  at  the  critical 
value  is  95%  at  sample  sizes  of  8 and  9,  and 
70%  with  a sample  size  of  7.  However,  the 
power  of  all  of  these  tests  at  the  P-values  of 
the  results  is  less  (25—36%)  because  of  the 
small  effect  observed  in  the  experiments.  Re- 
sults from  at  least  150  trials  would  be  needed 
to  hnd  significant  differences  with  such  a mi- 
nor effect  (Conover  1980). 

Magnitude  of  color  differences  and  female 
association. — Within  trials  of  both  experi- 
ments, males  varied  widely  in  color  differenc- 
es, and  it  is  possible  that  females  only  asso- 
ciated with  redder  or  brighter  males  when  col- 
or differences  were  large.  If  so,  then  the  dif- 
ference between  coloration  of  males  should  be 
positively  related  to  the  difference  in  time  a 
female  spent  with  males  within  a trial.  In  con- 
trast to  this  prediction,  there  were  no  positive 
or  significant  relationships  between  color 
score  differences  and  differences  in  time  spent 
with  males  in  a trial  for  the  natural  or  manip- 
ulated experiment  during  the  0 Hour  and  24 
Hour  periods.  In  fact,  all  of  the  slopes  were 
near  zero  or  negative  (range  of  h - -0.36  to 
0.08,  n = 10,  all  P > 0.05).  Therefore  as  color 


differences  between  males  increased,  females 
did  not  spend  a greater  amount  of  time  with 
redder  or  brighter  males. 

Female  behavior  in  the  experimental  avi- 
ary.— Males  and  females  rapidly  adjusted  to 
the  experimental  aviary  and  females  spent  the 
majority  of  time  during  trials  on  the  three 
perches  rather  than  on  the  floor  or  netting. 
Typically  within  the  first  15  minutes  after 
males  were  introduced,  females  had  visited 
each  of  the  side  compartments.  Singing  oc- 
curred in  7 of  the  10  natural  trials  and  8 of 
the  10  manipulated  trials,  but  I was  unable  to 
determine  from  the  videotapes  which  male  (or 
female)  sang.  In  any  one  trial  singing  occurred 
for  less  than  5 minutes.  When  adjacent  to  a 
male  compartment,  females  sat  on  perches  for 
relatively  long  times  but  regularly  interrupted 
these  periods  by  flights  back  and  forth  be- 
tween and  within  the  two  side  male  compart- 
ments. Such  flights  also  occurred  in  the  pre- 
trial periods. 

Site  preferences  by  females. — Females  did 
not  spend  more  time  consistently  on  the  east 
or  west  side  of  the  experimental  aviary  (Wil- 
coxon  test,  two-tailed:  natural  trials:  Pretrial  Z 
= 1.13,  P > 0.05,  n — 10,  0 Hours  Z = 0.06, 
P > 0.05;  24  Hours  Z = 0.05,  P > 0.05;  Re- 
versed Z = 1.07,  P > 0.05;  manipulated  trials: 
Pretrial  Z = 0.41,  P > 0.05;  0 Hours  Z = 1.38, 
P > 0.05;  24  Hour  Z = 1.07,  P > 0.05;  Re- 
versed Z = 0.97,  P > 0.05).  Between  the  pre- 
trial and  0 H periods,  females  did  not  exhibit 
a tendency  to  remain  on  the  same  side  of  the 
experimental  aviary  (binomial  test,  natural  tri- 
als: T=  3,  P > 0.05,  n = 8,  manipulated  trials: 
T = 5,  n = 10,  P > 0.05).  In  the  0 H and  24 
H periods,  most  females  remained  on  the 
same  side  (binomial  test,  natural  trials:  T = 5, 
n = 10,  P > 0.05;  manipulated  trials:  T = 8, 
n = 10,  P = 0.01)  indicating  that  females 
were  likely  to  associate  with  the  same  male  in 
the  0 H and  24  H sampling  periods.  However, 
in  the  24  H and  the  reversed  sampling  periods, 
females  again  spent  the  majority  of  time  on 
the  same  side  (binomial  test,  natural  trials:  T 
= 7,  /?  = 9,  P = 0.02;  manipulated  trials:  T 
= 6,  n = 8,  P = 0.035),  indicating  that  fe- 
males did  not  consistently  associate  with  the 
same  male  once  the  males’  positions  were  re- 
versed. 

Morphological  variables. — In  the  trials  us- 
ing natural  plumage  coloration,  males  with  the 


Wol/enharffer  • RED  COLORATION  AND  FEMALE  PREFERENCE 


81 


higher  color  score  did  not  have  a longer  tarsus 
length,  crest  length,  tail  length,  or  black  bib 
size  (Table  1;  Wilcoxon  rank  tests:  all  P > 
0.05).  Likewise,  males  in  the  reddened  and 
lightened  groups  did  not  differ  significantly  in 
morphological  measurements  (Table  2,  Wil- 
coxon rank  tests,  all  P > 0.05). 

DISCUSSION 

During  the  majority  of  observation  periods, 
females  spent  significantly  more  time  with 
one  male.  However,  the  two  experiments  pro- 
vided no  evidence  to  support  the  hypothesis 
that  female  Northern  Cardinals  prefer  males 
with  redder  or  brighter  coloration.  In  trials  us- 
ing natural  plumage  coloration,  females  were 
as  likely  to  spend  more  time  with  males  hav- 
ing low  color  scores  as  with  those  having  high 
color  scores  (Fig.  1).  Similarly  in  the  manip- 
ulated trials,  where  average  plumage  differ- 
ences between  males  were  greater  than  in  the 
natural  trials,  females  still  did  not  spend  more 
time  with  redder  or  brighter  males  (Fig.  2). 

These  results  differ  from  other  studies  of 
female  preference  in  passerines.  Under  aviary 
conditions  analogous  to  the  conditions  I used, 
females  of  many  species  showed  a preference 
for  brightly  colored  males  (Johnson  1988,  Hill 
1990,  Enstrom  1993,  Johnson  et  al.  1993, 
Saetre  et  al.  1994,  Sundberg  1995;  but  see 
Rohwer  and  Rpskaft  1989,  Alatalo  et  al.  1990, 
Butcher  1991).  Although  bright  plumage  col- 
ors do  not  seem  important  in  gallinaceous  spe- 
cies, females  show  preferences  for  other  male 
ornaments  (Zuk  et  al.  1992,  Buchholz  1995, 
Ligon  and  Zwartjes  1995,  Mateos  and  Carran- 
za 1995). 

The  individuals  adapted  well  to  captivity 
and  appeared  to  be  in  breeding  condition.  The 
experiments  were  conducted  at  a seasonally 
appropriate  time  when  pair  formation  occurs 
in  central  New  York  (pers.  obs.).  Behaviors 
associated  with  pair  formation  such  as  singing 
and  slow  flight  displays  occurred  in  the  hous- 
ing aviaries  and  during  experimental  trials, 
suggesting  that  males  were  responding  to  fe- 
male presence. 

Assigning  female  preference  based  on  the 
total  time  spent  with  males  has  become  a stan- 
dard method  in  avian  studies  of  female  choice 
(Burley  et  al.  1982,  Burley  1986,  Hill  1990, 
Enstrom  1993,  Johnson  et  al.  1993,  Sundberg 
1995).  Observations  of  pair  formation  in  the 


field  indicate  that  male  and  female  cardinals 
interact  extensively,  consequently  time  spent 
with  males  should  be  a reasonable  indication 
of  mate  preference  (Kinser  1973). 

While  males  in  this  study  appeared  to  ex- 
hibit some  behaviors  associated  with  pair  for- 
mation, female  behavior  was  more  ambigu- 
ous. Females  flew  toward  the  male  area  re- 
gardless of  whether  the  male  was  present  or 
not.  The  side  of  the  aviary  on  which  females 
spent  more  time  changed  from  the  pretrial  to 
0 hour  sampling  period,  indicating  that  fe- 
males did  not  immediately  establish  a prefer- 
ence for  one  side  of  the  experimental  aviary 
and  may  have  responded  to  the  addition  of  the 
males.  However,  females  preferred  the  same 
side  of  the  aviary  during  the  24  hour  and  re- 
versed sampling  periods,  even  though  the 
males  had  switched  sides.  This  suggests  that 
after  24  hours  individual  males  did  not  strong- 
ly influence  where  females  spent  more  time. 

With  the  sample  sizes  used,  the  power  of 
the  binomial  test  is  high  at  the  critical  value. 
It  is  noteworthy  that  other  choice  experiments 
of  this  design  have  found  significant  female 
preferences  with  similar  samples  sizes  (rang- 
ing from  7 to  21  females;  Hill  1990,  Enstrom 
1993,  Saetre  et  al.  1994,  Sundberg  1995).  This 
suggests  that  if  a preference  for  red  color  ex- 
ists, it  is  weak  in  comparison  to  color  pref- 
erences found  in  other  passerines. 

The  dietary  basis  of  red  coloration  in  male 
cardinals  suggests  that  females  could  acquire 
information  about  a male’s  foraging  abilities 
or  ability  to  defend  resources  using  coloration. 
Male  color  and  his  absolute  effort  in  feeding 
nestlings  are  not  related  in  cardinals  (Linville 
et  al.  1998),  emphasizing  the  limitation  of  us- 
ing red  coloration  alone  to  assess  mates.  How- 
ever, females  mated  to  brighter  males  fed  nest- 
lings less  (Linville  et  al.  1998),  suggesting 
that  there  may  be  advantages  to  pairing  with 
brighter  males. 

My  results  provided  no  evidence  for  female 
preference  for  brighter  or  redder  male  plum- 
age coloration  as  a single  criterion  for  mate 
choice;  however,  the  possibility  remains  that 
female  cardinals  assess  mates  using  a combi- 
nation of  factors,  including  coloration,  as  has 
been  found  in  other  species  (Zuk  et  al.  1990; 
Omland  1996a,  b;  Scheffer  et  al.  1996;  Mpller 
et  al.  1998).  The  prolonged  opportunities  for 
direct  interactions  between  males  and  females 


82 


THE  WILSON  BULLETIN 


Vol.  Ill,  No.  1,  March  1999 


prior  to  pairing  (Kinser  1973)  suggests  that 
females  may  be  able  to  use  other  male  char- 
acteristics in  mate  choice,  such  as  song  or 
courtship  behaviors.  If  females  assess  male 
condition,  multiple  ornaments  or  traits  may 
provide  more  accurate  information  either 
through  redundancy  or  because  some  orna- 
ments are  unreliable  indicators  of  condition 
(Mpller  and  Pomiankowski  1993).  It  also  re- 
mains possible  that  females  assess  male  col- 
oration in  some  circumstances  but  not  other. 
For  example,  females  may  use  different  cri- 
teria when  choosing  social  mates  and  when 
choosing  extra-pair  mates. 

Lastly,  territorial  resources  may  influence 
mate  choices,  but  coloration  may  be  relatively 
more  important  in  mediating  competition  be- 
tween males.  Redder  male  cardinals  acquire 
territories  with  denser  vegetation  density,  and 
pairs  on  these  territories  produce  more  off- 
spring, most  likely  because  of  reduced  nest 
predation  (Wolfenbarger  in  press).  Given  the 
importance  of  territory  quality,  females  may 
directly  assess  a male’s  territory  during  pair 
formation  rather  than  relying  on  plumage  col- 
oration alone  for  mate  choices. 

ACKNOWLEDGMENTS 

For  assistance  in  designing  and  conducting  these  ex- 
periments, I thank  S.  Barker,  C.  Brown,  L.  Carbone, 

R.  Cocroft,  S.  Emlen,  M.  Martino,  J.  McCarty,  K. 
Reeve,  J.  Roth,  P.  Sherman,  S.  Smedley,  and  D.  Wink- 
ler. The  manuscript  was  improved  by  comments  from 

S.  Emlen,  J.  McCarty,  P.  Sherman,  D.  Winkler,  and  two 
anonymous  reviewers.  The  work  was  conducted  in  ac- 
cordance with  federal  (PRT-792153),  NY  state 
(LCP94-681),  and  Cornell  University's  Institutional 
Animal  Care  and  Use  Committee  (94-129)  regulations. 
This  work  was  supported  by  a Dissertational  Improve- 
ment Grant  from  the  NSF  (IBN-9321801 ). 

LITERATURE  CITED 

Alatalo,  R.  V.,  A.  Lundberg,  and  J.  Sundberg.  1990. 
Can  female  preference  explain  sexual  dichroma- 
tism in  the  Pied  Flycatcher,  Ficedula  hypoleuca? 
Anim.  Behav.  39:244-252. 

Andersson,  M.  1982.  Female  choice  selects  for  ex- 
treme tail  length  in  a widowbird.  Nature  (Lond.) 
299:818-820. 

Andersson,  M.  1994.  Sexual  selection.  Princeton 
Univ.  Press,  Princeton,  New  Jersey. 

Andersson,  S.  1992.  Female  preferences  for  long  tails 
in  lekking  Jackson's  Widowbirds:  experimental 
evidence.  Anim.  Behav.  43:379-388. 

Buchholz,  R.  1995.  Female  choice,  parasite  load  and 


male  ornamentation  in  Wild  Turkeys.  Anim.  Be- 
hav. 50:929-943. 

Burley,  N.  1986.  Comparison  of  the  band-colour  pref- 
erences of  two  species  of  estrildid  finches.  Anim. 
Behav.  34:1732-1741. 

Burley,  N.,  G.  Krantzberg,  and  P.  Radman.  1982. 
Influence  of  colour-banding  on  the  conspecific 
preferences  of  Zebra  Finches.  Anim.  Behav.  30: 
444-455. 

Butcher,  G.  S.  1991.  Mate  choice  in  female  Northern 
Orioles  with  a consideration  of  the  role  of  the 
black  male  coloration  in  female  choice.  Condor 
93:82-88. 

Conover,  W.  J.  1980.  Practical  non-parametric  statis- 
tics, second  edition.  John  Wiley  and  Sons,  Inc., 
New  York. 

Darwin,  C.  1871.  The  descent  of  man,  and  selection 
in  relation  to  sex.  Princeton  Univ.  Press,  Prince- 
ton, New  Jersey. 

Enstrom,  D.  a.  1993.  Female  choice  for  age-specific 
plumage  in  the  Orchard  Oriole:  implications  for 
delayed  plumage  maturation.  Anim.  Behav.  45: 
435-442. 

Evans,  M.  R.  and  B.  J.  Hatchwell.  1992.  An  exper- 
imental study  of  male  adornment  in  the  Scarlet- 
tufted  Malachite  Sunbird:  II.  The  role  of  the  elon- 
gated tail  in  mate  choice  and  experimental  evi- 
dence for  a handicap.  Behav.  Ecol.  Sociobiol.  29: 
421-427. 

Filliater,  T.  S.  and  R.  Breitwisch.  1997.  Nestling 
provisioning  by  the  extremely  dichromatic  North- 
ern Cardinal.  Wilson  Bull.  109:145-153. 

Goodwin,  T.  W.  1950.  Carotenoids  and  reproduction. 
Biol.  Rev.  25:391-413. 

Hagan,  J.  M.  and  J.  M.  Reed.  1988.  Red  color  bands 
reduce  fledging  success  in  Red-cockaded  Wood- 
peckers. Auk  105:498-503. 

Hill,  G.  E.  1990.  Female  House  Finches  prefer  col- 
ourful males:  sexual  selection  for  a condition-de- 
pendent trait.  Anim.  Behav.  40:563-572. 

Hill,  G.  E.  1992.  Proximate  basis  of  variation  in  ca- 
rotenoid pigmentation  in  male  House  Finches. 
Auk  109:1-12. 

Johnson,  K.  1988.  Sexual  selection  in  Pinyon  Jays  I: 
female  choice  and  male-male  competition.  Anim. 
Behav.  36:1038-1047. 

Johnson,  K.,  R.  Dalton,  and  N.  Burley.  1993.  Pref- 
erences of  female  American  Goldfinches  {Car- 
cluelis  iri.stis)  for  natural  and  artificial  male  traits. 
Behav.  Ecol.  4:138-143. 

Jones,  I.  L.  and  F.  M.  Hunter.  1993.  Mutual  .sexual 
selection  in  a monogamous  seabird.  Nature 
(Lond.)  362:238-239. 

Kinser,  G.  W.  1973.  Ecology  and  behavior  of  the  car- 
dinal Richmomlemi  cardinaUs  (L).  in  southern  In- 
diana. Ph.D.  diss.,  Indiana  Univ.,  Bloomington. 

Kornerup,  a.  1967.  Methuen  handbook  of  color.  Thiru 
ed.  Methuen,  London,  U.K. 

Ligon,  j.  D.  and  P.  W.  Zwartjes.  1995.  Ornate  phm 
age  of  male  Red  Jungle  Fowl  does  not  influeiu 


Wolfenbarger  • RED  COLORATION  AND  FEMALE  PREFERENCE 


83 


mate  choice  by  females.  Anim.  Behav.  49:117- 
125. 

Linville,  S.U.,  R.  Breitwisch,  and  A.  J.  Schilling. 
1998.  Plumage  brightness  as  an  indicator  of  pa- 
rental care  in  Northern  Cardinals.  Anim.  Behav. 
55:  119-127. 

Mateos,  C.  and  J.  Carranza.  1995.  Female  choice 
for  morphological  features  of  male  Ring-necked 
Pheasants.  Anim.  Behav.  49:737-748. 

Metz,  K.  J.  and  P.  J.  Weatherhead.  1991.  Color 
bands  function  as  secondary  sexual  traits  in  male 
Red-winged  Blackbirds.  Behav.  Ecol.  Sociobiol. 
28:23-27. 

Moller,  a.  P.  1987.  Variation  in  badge  size  in  male 
House  Sparrows  Passer  domesticus:  evidence  for 
status  signalling.  Anim.  Behav.  35:1637—1644. 

Moller,  a.  P.  1988.  Female  choice  selects  for  male 
sexual  tail  ornaments  in  the  monogamous  swal- 
low. Nature  (Lond.)  332:640—642. 

Moller,  A.  P.  1992.  Female  swallow  preference  for 
symmetrical  male  sexual  ornaments.  Nature 
(Lond.)  357:238-240. 

Moller,  A.  P.  and  A.  Pomiankowski.  1993.  Why  have 
birds  got  multiple  sexual  ornaments?  Behav.  Ecol. 
Sociobiol.  32:167-176. 

Moller,  A.  P,  N.  Saino,  G.  Taramino,  P.  Galeotti, 
AND  S.  Ferrario.  1998.  Paternity  and  multiple  sig- 
naling: effects  of  a secondary  sexual  character  and 
song  on  paternity  in  the  Bam  Swallow.  Am.  Nat. 
151:236-242. 

Omland,  K.  E.  1996a.  Female  Mallard  mating  pref- 
erences for  multiple  male  ornaments.  1.  Natural 
variation.  Behav.  Ecol.  Sociobiol.  39:353-360. 

Omland,  K.  E.  1996b.  Female  Mallard  mating  prefer- 
ences for  multiple  male  ornaments.  II.  Experimen- 
tal variation.  Behav.  Ecol.  Sociobiol.  39:361—366. 


Rice,  W.  R.  1989.  Analyzing  tables  of  statistical  tests. 
Evolution  43:223-225. 

Rohwer,  S.  and  E.  Roskalt.  1989.  Results  of  dyeing 
male  Yellow-headed  Blackbirds  solid  black:  im- 
plications for  the  arbitrary  identity  badge  hypoth- 
esis. Behav.  Ecol.  Sociobiol.  25:39-49. 

S/ETRE,  G.-P,  S.  Dale,  and  T.  Slagsvold.  1994.  Fe- 
male Pied  Flycatchers  prefer  brightly  coloured 
males.  Anim.  Behav.  48:1407-1416. 

Scheffer,  S.  J.,  G.  W.  Uetz,  and  G.  E.  Stratton. 
1996.  Sexual  selection,  male  morphology,  and  the 
efficacy  of  courtship  signalling  in  two  wolf  spi- 
ders (Araneae:  Lycosidae).  Behav.  Ecol.  Socio- 
biol. 38:17-23. 

SuNDBERG,  J.  1995.  Female  Yellowhammers  {Emheriza 
citrinella)  prefer  yellower  males:  a laboratory  ex- 
periment. Behav.  Ecol.  Sociobiol.  37:275-282. 

Wilkinson,  L.,  M.  Hill,  S.  Miceli,  P.  Howe,  and  E. 
Vang.  1992.  SYSTAT  version  5.2.  SYSTAT,  Inc., 
Evanston,  Illinois. 

WoLFENBARGER,  L.  L.  1996.  Fitness  effects  associated 
with  red  coloration  of  male  Northern  Cardinals 
(Cardinalis  cardinalis).  Ph.D.  diss.,  Cornell  Univ., 
Ithaca,  New  York. 

WOLFENBARGER,  L.  L.  In  press.  Red  coloration  of  male 
northern  cardinals  correlates  with  mate  quality 
and  territory  quality.  Behav.  Ecol. 

ZuK,  M.,  J.  D.  Ligon,  and  R.  Thornhill.  1992.  Effects 
of  experimental  manipulation  of  male  secondary 
sex  characters  on  female  mate  preference  in  Red 
Jungle  Fowl.  Anim.  Behav.  44:999-1006. 

ZuK,  M.,  R.  Thornhill,  J.  D.  Ligon,  K.  Johnson,  S. 
Austad,  S.  Ligon,  N.  Thornhill,  and  C.  Costin. 
1990.  The  role  of  male  ornaments  and  courtship 
behavior  in  female  choice  of  Red  Jungle  Fowl. 
Am.  Nat.  136:459-473. 


Wilson  Bull.,  111(1),  1999,  pp.  84-88 


FRUIT  SUGAR  PREFERENCES  OF  HOUSE  FINCHES 

MICHAEL  L.  AVERY,'  3 CARRIE  L.  SCHREIBER,' - AND  DAVID  G.  DECKER' 


ABSTRACT. — In  a series  of  choice  tests,  we  determined  the  relative  preferences  of  House  Finches  (Carpo- 
dacus  mexicamis)  for  equicaloric  aqueous  solutions  of  hexoses  (1:1  mixture  of  fructose  and  glucose)  and  sucrose. 
At  2%  (m/v),  birds  consumed  each  sugar  solution  equally  and  in  amounts  similar  to  plain  water.  Consumption 
of  hexose  but  not  sucrose  increased  at  4%  sugar  concentration.  At  6%  and  10%,  finches  displayed  consistent, 
strong  preferences  for  the  hexoses  over  sucrose.  In  other  passerine  species,  strong  hexose  preference  has  been 
linked  to  the  absence  of  sucrase,  the  enzyme  needed  for  digestion  of  sucrose.  Fecal  sugar  readings  from  the 
House  Finches,  however,  indicated  approximately  equal  assimilation  of  hexose  and  sucrose,  so  the  hexose  pref- 
erence apparently  is  not  due  to  sucrase  deficiency.  Rather,  energetics  may  determine  the  finches’  sugar  prefer- 
ences: hexoses  are  rapidly  processed  because  the  6-carbon  sugars  are  readily  assimilable  whereas  sucrose  must 
first  be  hydrolyzed.  Received  22  Jem.  1998,  accepted  30  Aug.  1998. 


Physiology  imposes  major  constraints  on 
the  digestion  of  sugars  by  some  fruit-eating 
birds.  These  constraints  in  turn  affect  species’ 
food  selection  behavior.  Species  of  Stumidae 
(e.g.,  European  Starling,  Sturnus  vulgaris)  and 
Turdidae  (e.g.,  American  Robin,  Turdus  mig- 
ratorius)  are  unable  to  digest  sucrose  because 
they  lack  the  enzyme  sucrase  needed  to  hy- 
drolyze sucrose  into  6-carbon  sugars,  glucose 
and  fructose  that  can  be  assimilated  (Martinez 
del  Rio  and  Stevens  1989,  Karasov  and  Levey 
1990).  Ingestion  of  high  concentrations  of  su- 
crose by  these  species  produces  osmotic  di- 
arrhea and,  in  extreme  cases,  death  (Martinez 
del  Rio  et  al.  1988,  Brugger  and  Nelms  1991). 
Consequently,  in  feeding  and  drinking  trials 
starlings  and  robins  learn  to  avoid  sucrose 
(Schuler  1983,  Martinez  del  Rio  et  al.  1988, 
Brugger  1992). 

Although  Cedar  Waxwings  (Bombycilla 
cedrorum)  can  digest  sucrose,  in  choice  tests 
they  also  prefer  hexoses  to  sucrose  (Martinez 
del  Rio  et  al.  1989,  Avery  et  al.  1995).  Wax- 
wings  exhibit  very  rapid  gut  passage  rates 
(Levey  and  Grajal  1991).  As  a result,  sucrose 
is  not  in  the  gut  long  enough  to  be  completely 
hydrolyzed  and  is  therefore  inefficiently  as- 
similated relative  to  hexose  sugars  (Martinez 
del  Rio  et  al.  1989). 

In  the  Icteridae  and  Emberizidae,  two  fru- 


' U.vS.  Department  of  Agriculture,  Animal  and  Plant 
Health  In.spcction  .Service.  National  Wildlife  Research 
Center.  Florida  Field  Station,  2820  East  University  Av- 
enue, Gainesville,  FL  .12641. 

’ Present  addre.ss:  716  NE  4th  Avenue,  Gainesville, 
FL  .12601. 

'Corresponding  author;  E-mail:  dwrc-ffs@afn.org 


givorous  species,  the  Yellow-winged  Cacique 
{Cacicus  melanicterus)  and  the  Yellow-breast- 
ed Chat  (Icteria  virens)  preferred  15%  (by 
mass)  hexose  solution  over  sucrose  solution 
and  displayed  relatively  inefficient  sucrose  di- 
gestion (Martinez  del  Rio  and  Restrepo  1993). 
Conversely,  Red-winged  Blackbirds  (Agelaius 
phoeniceus)  and  Common  Crackles  {Quisca- 
lus  quiscula),  granivorous  icterids,  preferred 
sucrose  solutions  to  water  but  did  not  distin- 
guish between  0.175M  and  0.35M  hexose  so- 
lutions and  water  (Martinez  del  Rio  et  al. 
1988). 

The  House  Finch  (Carpodacus  mexicanus) 
is  primarily  granivorous  (Martin  et  al.  1951) 
but  feeds  opportunistically  on  cultivated  fruit 
(Tobin  and  DeHaven  1984,  Avery  et  al.  1992). 
To  our  knowledge  the  sugar  preferences  of 
House  Finches  and  other  Fringillidae  have  not 
been  evaluated.  Responses  of  House  Finches 
to  fruit  sugars  are  pertinent  to  the  develop- 
ment of  high-sucrose  fruit  cultivars  for  poten- 
tially reducing  bird  damage  to  fruit  crops 
(Brugger  et  al.  1993,  Darnell  et  al.  1994). 
Thus,  our  objectives  were  (1)  to  document 
House  Finch  consumption  of  sucrose  and  hex- 
ose in  equicaloric  aqueous  solutions  across  a 
range  of  sugar  concentrations  typically  found 
in  cultivated  fruit  and  (2)  to  measure  fecal 
sugar  to  determine  relative  digestion  of  su- 
crose and  hexoses. 

METHODS 

House  Finches  were  from  a captive  population 
maintained  at  the  Florida  Field  Station  of  the  U.S. 
Dept,  of  Agriculture’s  National  Wildlife  Research  Cen- 
ter m Gainesville,  Florida.  We  maintained  birds  on  a 
mixed  .seed  diet  supplemented  three  days/week  with 


84 


Aven’  et  cil.  • HOUSE  FINCH  SUGAR  PREFERENCES 


85 


apples  and  lettuce.  Testing  occurred  during  October— 
November  1995.  After  testing,  birds  were  returned  to 
their  home  cages. 

We  removed  birds  from  communal  enclosures  (2  X 
1.5  X 2.2  m)  and  placed  them  into  individual,  visually 
isolated  test  cages  (45  cm  on  a side)  in  a roofed  out- 
door aviary.  To  acclimate  the  birds,  we  offered  plain 
water  tinted  with  red  food  coloring  in  clear  glass  tubes 
(8  mm  diameter)  4-5  days  before  testing.  We  fixed  two 
tubes,  5 cm  apart,  to  the  front  of  each  cage.  During 
acclimation,  we  measured  water  consumption  after  6 
h and  24  h daily  to  determine  baseline  fluid  intake  and 
to  accustom  the  birds  to  disturbances. 

We  prepared  test  solutions  by  dissolving  20,  40,  60, 
or  100  g of  sucrose  or  hexose  sugars  (Sigma  Chemical 
Company,  St.  Louis,  Missouri)  in  1 L of  distilled  wa- 
ter. The  hexose  solution  contained  equal  amounts  of 
fructose  and  glucose.  We  then  conducted  separate  tests 
at  each  of  4 sugar  concentrations  (m/v):  2%,  4%,  6%, 
and  10%.  Tests  lasted  4 days  and  there  were  6 birds/ 
group.  One  hexose  tube  and  one  sucrose  tube,  5 cm 
apart,  were  available  during  each  test.  For  each  cage, 
we  first  randomly  determined  the  position  of  the  su- 
crose tube  and  then  alternated  sucrose  and  hexose  po- 
sitions daily.  We  removed  maintenance  food  and  water 
at  08:00  and  presented  the  tubes  with  sugar  solutions 
from  09:00  until  15:00.  Maintenance  food  and  water 
were  then  returned  to  the  cages. 

We  measured  the  amount  of  solution  missing  from 
each  tube  to  the  nearest  mm,  and  then  converted  to 
amount  of  sugar  (g)  ingested  for  analyses.  We  assessed 
sugar  consumption  in  a 3-way  analysis  of  variance, 
with  sugar  concentration  as  the  independent  factor,  and 
repeated  measures  over  sugar  type  and  days.  We  used 
Tukey’s  HSD  test  (Steel  and  Torrie  1980)  to  isolate 
differences  {P  < 0.05)  among  means. 

To  determine  relative  digestion  by  finches  of  sucrose 
and  hexoses,  we  analysed  fecal  sugar  with  a hand-held 
refractometer  (Hainesworth  1974,  Brugger  et  al.  1993). 
We  offered  six  birds  a 10%  (m/v)  agar-sucrose  mixture 
(Avery  et  al.  1995)  for  6 hours  and  offered  similar 
food  made  with  hexose  (equal  amounts  of  glucose  and 
fructose)  for  6 hours  the  next  day.  We  measured  three 
fresh  defecations  from  each  bird  with  each  sugar  treat- 
ment, and  compared  mean  values  in  a paired  r-test 
against  a null  hypothesis  of  no  difference  between  sug- 
ars. Refractometer  readings  are  expressed  as  degrees 
Brix  which  corresponds  to  the  percentage  of  sugar  pre- 
sent in  the  sample  on  a mass : mass  basis  (Bolten  et  al. 
1979). 

RESULTS 

Total  sugar  consumption  varied  (F,  20  = 
22.77,  P < 0.001)  with  concentration.  Sugar 
ingestion  at  6%  (mean  ± SE,  x = 1.07  ±0.12 
g/bird)  and  10%  (jc  = 1.66  ±0.11  g/bird)  ex- 
ceeded that  at  2%  (x  = 0.03  ± 0.01  g/bird) 
and  4%  (T  = 0.29  ± 0.08  g/bird).  Overall, 
hexose  consumption  (0.67  ± 0.07  g/bird)  ex- 
ceeded (F,2o  = 93.55,  P < 0.001)  sucrose 


FIG.  1 . Mean  consumption  of  hexoses  and  sucrose 
by  House  Finch  groups  (6  birds/group)  exposed  to  two 
tubes  of  aqueous  sugar  solutions  for  4 days,  6 hours 
per  day.  Vertical  bars  denote  1 SE.  Note  that  the  y axis 
is  logarithmic. 


consumption  (0.09  ± 0.02  g/bird).  Finches 
consumed  less  sugar  (F^^eo  = 2.83,  P = 0.046) 
on  day  1 (0.61  ± 0.17  g/bird)  than  on  days 
2-4  (mean  consumption  0.79-0.83  g/bird). 

Across  the  range  of  test  concentrations, 
finches  responded  differently  (F3  20  = 24.78,  P 
< 0.001)  to  the  two  types  of  sugars  (Fig.  1). 
Sucrose  consumption  was  consistently  low 
(mean  consumption  0.02-0.22  g/bird)  and  did 
not  differ  from  hexose  consumption  at  2% 
(0.02  ± 0.01  g/bird)  and  4%  (0.22  ± 0.07  g/ 
bird).  Hexose  consumption  increased  (P  < 
0.05)  substantially,  however,  at  6%  (1.01  ± 
0.11  g/bird)  and  at  10%  (1.43  ± 0.10  g/bird). 

The  interaction  between  type  of  sugar  and 
test  day  affected  consumption  (F^^o  ~ 16.59, 
P < 0.001).  Sucrose  consumption  did  not  dif- 
fer across  the  4 test  days,  and  on  day  1,  mean 
sucrose  consumption  (0.22  ± 0.10  g/bird) 
equalled  hexose  consumption  (0.39  ± 0.10  g/ 
bird).  Hexose  consumption  increased  thereaf- 
ter and  averaged  0.73  to  0.79  g/bird  on  days 
2-4. 

The  3-way  interaction  (Fg^,,,  = 5.31,  P < 
0.001)  reflected  differing  daily  consumption 
patterns  of  the  two  sugar  types  as  sugar  con- 
centration varied  (Fig.  2).  At  2%,  consump- 
tion of  both  types  of  sugar  remained  low 
throughout  the  test.  At  4%,  mean  hexose  con- 
sumption increased  each  day  but  not  suffi- 
ciently to  achieve  statistical  significance  (F  > 
0.05).  At  6%,  mean  hexose  consumption  in- 
creased (F  < 0.05)  from  day  1 (0.66  ± 0.21 


86 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


Day  1 Day  2 Day  3 Day  4 

FIG.  2.  Mean  daily  consumption  of  hexoses  (open  bars)  and  sucrose  (solid  bars)  by  House  Finches  (6  birds/ 
trial)  exposed  to  two  tubes  of  aqueous  sugar  solutions  for  6 hours  each  day.  Vertical  bars  denote  1 SE. 


g/bird)  to  day  4 (1.25  ± 0.22  g/bird).  On  day 
1,  finches  consumed  equal  amounts  of  hexose 
and  sucrose  at  the  10%  level,  but  consumption 
diverged  {P  < 0.05)  on  day  2 and  remained 
so  through  day  4. 

During  pretest  days,  hourly  consumption  of 
water  averaged  0.22  ± 0.10  and  0.33  ± 0.07 
ml/bird  for  the  2%  and  4%  groups,  respec- 
tively, similar  to  their  total  consumption  of 
2%  (0.25  ± 0.03  ml/bird)  and  4%  (1.19  ± 
0.30  ml/bird)  sugar  solutions.  Fecal  sugar 
analysis  from  six  birds  revealed  no  difference 
(/  = 1.63,  P > 0.05)  between  sugars.  Hexose 
readings  averaged  4.2  ± 0.7°  Brix  compared 
to  an  average  of  2.8  ± 0.3°  Brix  for  sucrose. 

DISCUSSION 

In  the  range  of  concentrations  we  tested, 
preference  for  hexose  over  sucrose  has  not 
previously  been  demonstrated,  even  in  species 
lacking  sucrase.  Rejection  of  sucrose  by  Eu- 
ropean Starlings  and  American  Robins  oc- 
curred at  concentrations  in  excess  of  10% 
(Schuler  1983,  Martmez  del  Rio  et  al.  1988, 
Brugger  1992).  Other  species  are  either  indif- 
ferent (domestic  hen,  Kare  and  Medway  1959; 


Rock  Dove,  Columba  livia,  Duncan  1960; 
Common  Raven,  Corpus  corax,  Harriman  and 
Fry  1990)  or  prefer  sucrose  (Common  Crack- 
le, Red-winged  Blackbird;  Martmez  del  Rio  et 
al.  1988).  In  choice  tests,  hummingbirds  pre- 
fer sucrose  and  reject  fructose  (Stiles  1976), 
but  when  fructose  is  offered  alone,  humming- 
birds consume  it  at  a rate  no  different  from 
sucrose.  Other  nectarivorous  species  also  se- 
lect sucrose  preferentially  over  equimolar 
fructose  and  glucose  solutions  (Downs  and 
Perrin  1996). 

The  sugar  solutions  we  offered  appeared 
alike  to  us  and  their  relative  positions  were 
switched  daily.  At  2%,  it  appeared  that  finches 
did  not  distinguish  dilute  sugar  solutions  from 
plain  water;  consumption  was  low  and  re- 
mained so  throughout  the  trial.  Finches  re- 
sponded to  sugar  at  the  4%  level,  and  mean 
consumption  of  hexose  increased  steadily 
across  the  4-day  trial  while  sucrose  consump- 
tion remained  low.  At  6%,  hexose  consump- 
tion increased  markedly  over  that  at  2%  and 
4%,  while  sucrose  consumption  did  not  differ 
from  that  at  lower  concentrations.  Discrimi- 
nation between  sugars  was  more  rapid  at  10%, 


Avcrv  et  al.  • HOUSE  FINCH  SUGAR  PREFERENCES 


87 


as  finches  decisively  selected  hexose  over  su- 
crose after  one  trial.  The  birds  apparently 
tracked  the  position  of  the  hexose  tube 
through  a nonvisual  cue.  The  mechanism  by 
which  they  discriminated  hexose  from  sucrose 
is  unclear,  but  the  rapidity  of  the  discrimina- 
tion increased  with  sugar  concentration. 

We  hypothesize  that  finches  chose  hexoses 
in  response  to  an  increased  rate  of  energy  gain 
relative  to  sucrose  solutions  during  the  6-h 
drinking  trials.  Birds  are  sensitive  to  differ- 
ences in  rates  of  energy  assimilation  (Witmer 
1994),  and  the  extra  step,  hydrolysis  of  the 
sucrose  molecule  required  for  sucrose  diges- 
tion imposes  a constraint  on  the  potential  rate 
of  energy  assimilation.  In  our  choice  tests, 
finches  responded  facultatively  and  selected 
the  more  energetically  efficient  food  source. 

Martinez  del  Rio  and  coworkers  (1988)  pre- 
dicted that  granivores  should  have  high  su- 
crase  activity  and  prefer,  or  at  least  tolerate, 
sucrose.  This  follows  from  the  facts  that  malt- 
ose is  the  major  constituent  of  complex  car- 
bohydrates found  in  seeds,  granivorous  spe- 
cies show  high  intestinal  maltase  activity,  and 
the  activity  of  sucrase  seems  to  vary  with  that 
of  maltase  and  isomaltase  (Martinez  del  Rio 
1990,  Martinez  del  Rio  et  al.  1995).  Although 
House  Finches  are  basically  granivorous,  they 
strongly  favored  moderate  hexose  sugar  so- 
lutions over  sucrose  (Figs.  1,  2).  We  did  not 
determine  intestinal  enzyme  activity  directly, 
but  fecal  sugar  analyses  indicated  that  the 
preference  for  hexoses  was  not  because  of  ab- 
sence of  sucrase.  House  Finches  prefer  hexose 
sugars  but  are  “sucrose  tolerant”  granivores, 
consistent  with  the  hypothesis  of  Martinez  del 
Rio  and  coworkers  (1988).  Comparative  stud- 
ies of  House  Finches  and  other  granivores  will 
help  to  define  more  clearly  the  physiological 
basis  underlying  their  food  selection  behavior. 

Development  of  high-sucrose  fruit  cultivars 
could  represent  one  nonlethal  component  of 
an  integrated  plan  to  manage  bird  damage  to 
berry  crops  (Brugger  et  al.  1993,  Darnell  et 
al.  1994).  Such  an  approach  will  most  likely 
be  effective  against  species  such  as  the  Eu- 
ropean Starling  and  American  Robin  that  lack 
sucrase  and  are  thus  unable  to  digest  sucrose. 
For  sucrose  tolerant  species  such  as  the  House 
Finch,  elevated  sucrose  concentrations  in  fruit 
will  probably  not  reduce  crop  damage  unless 
alternative  food  sources  are  readily  available. 


Rather,  because  of  inefficient  energy  assimi- 
lation from  sucrose  ingestion,  sucrose  tolerant 
species  might  compensate  by  increasing  fruit 
consumption,  thereby  causing  greater  damage 
(Avery  et  al.  1995). 

ACKNOWLEDGMENTS 

We  thank  K.  L.  Roca  and  C.  C.  McClester  for  caring 
for  the  House  Finches.  Our  paper  was  improved  by  the 
constructive  comments  of  J.  R.  Belthoff,  K.  E.  Brug- 
ger, D.  J.  Levey,  C.  Martinez  del  Rio,  and  two  anon- 
ymous reviewers. 

LITERATURE  CITED 

Avery,  M.  L.,  J.  W.  Nelson,  and  M.  A.  Cone.  1992. 
Survey  of  bird  damage  to  blueberries  in  North 
America.  Proc.  East.  Wildl.  Damage  Control 
Conf.  5:105-1 10. 

Avery,  M.  L.,  D.  G.  Decker,  J.  S.  Humphrey,  A.  A. 
Hayes,  and  C.  C.  Laukert.  1995.  Color,  size,  and 
location  of  artificial  fruits  affect  sucrose  avoidance 
by  Cedar  Waxwings  and  European  Starlings.  Auk 
112:436-444. 

Bolten,  a.  B.,  P.  Feinsinger,  H.  G.  Baker,  and  1. 
Baker.  1979.  On  the  calculation  of  sugar  concen- 
tration in  flower  nectar.  Oecologia  41:301-304. 
Brugger,  K.  E.  1992.  Repellency  of  sucrose  to  captive 
American  Robins  (Turdus  migratorius).  J.  Wildl. 
Manage.  56:794-799. 

Brugger,  K.  E.  and  C.  O.  Nelms.  1991.  Sucrose 
avoidance  by  American  Robins  {Turdus  migrato- 
rius): implications  for  control  of  bird  damage  in 
fruit  crops.  Crop  Prot.  10:455-460. 

Brugger,  K.  E.,  P.  Nol,  and  C.  I.  Phillips.  1993.  Su- 
crose repellency  to  European  Starlings:  will  high- 
sucrose  cultivars  deter  bird  damage  to  fruit?  Ecol. 
Applic.  3:256—261. 

Darnell,  R.  L.,  R.  Cano-Medrano,  K.  E.  Koch,  and 
M.  L.  Avery.  1994.  Differences  in  sucrose  me- 
tabolism relative  to  accumulation  of  bird-deterrent 
sucrose  levels  in  faiits  of  wild  and  domestic  Vac- 
cinium  species.  Physiol.  Plant.  92:336—342. 
Downs,  C.  T.  and  M.  R.  Perrin.  1996.  Sugar  prefer- 
ences of  some  southern  African  nectarivorous 
birds.  Ibis  138:455-459. 

Duncan,  C.  J.  1960.  The  sense  of  taste  in  birds.  Ann. 
Appl.  Biol.  48:409-414. 

Hainesworth,  E R.  1974.  Food  quality  and  foraging 
efficiency:  the  efficiency  of  sugar  assimilation  by 
hummingbirds.  J.  Comp.  Physiol.  88:425-431. 
Harriman,  a.  E.  and  E.  G.  Fry.  1990.  Solution  ac- 
ceptance by  Common  Ravens  (Con-us  cora.x)  giv- 
en two-bottle  preference  tests.  Psychol.  Rep.  67: 
19-26. 

Karasov,  W.  H.  and  D.  j.  Levey.  1990.  Digestive  sys- 
tem trade-offs  and  adaptations  of  frugivorous  pas- 
serine birds.  Physiol.  Zool.  63:1248—1270. 

Kare,  M.  R.  and  W.  Medway.  1959.  Discrimination 
between  carbohydrates  by  the  fowl.  Poultry  Sci. 
38:1119-1  127. 


88 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  I,  March  1999 


Levey,  D.  J.  and  A.  Grajal.  1991.  Evolutionary  im- 
plications of  fruit-processing  limitations  in  Cedar 
Waxwings.  Am.  Nat.  138:171-189. 

Martin,  A.  C.,  H.  S.  Zim,  and  A.  L.  Nelson.  1951. 
American  wildlife  and  plants.  McGraw-Hill  Book 
Co.,  Inc.,  New  York. 

Martinez  del  Rio,  C.  1990.  Dietary  and  phylogenetic 
correlates  of  intestinal  sucrase  and  maltase  activity 
in  birds.  Physiol.  Zool.  63:987-1011. 

Martinez  del  Rio,  C.,  K.  E.  Brugger,  J.  L.  Rios,  M. 
E.  Vergara,  and  M.  Witmer.  1995.  An  experi- 
mental and  comparative  study  of  dietary  modu- 
lation of  intestinal  enzymes  in  European  Starlings 
(Sturnu.'i  vulgari.s).  Physiol.  Zool.  68:490-511. 

Martinez  del  Rio,  C.,  W.  H.  Karasov,  and  D.  J.  Lev- 
ey. 1989.  Physiological  basis  and  ecological  con- 
sequences of  sugar  preferences  in  Cedar  Wax- 
wings.  Auk  106:64-71. 

Martinez  del  Rio,  C.  and  C.  Restrepo.  1993.  Eco- 
logical and  behavioral  consequences  of  digestion 
in  frugivorous  animals.  Vegetatio  107/108:205- 
216. 


Martinez  del  Rio,  C.  and  B.  R.  Stevens.  1989.  Phys- 
iological constraint  on  feeding  behavior:  intestinal 
membrane  disaccharidases  of  the  starling.  Science 
243:794-796. 

Martinez  del  Rio,  C.,  B.  R.  Stevens,  D.  E.  Daneke, 
AND  P.  T.  Andreadis.  1988.  Physiological  corre- 
lates of  preference  and  aversion  for  sugars  in  three 
species  of  birds.  Physiol.  Zool.  61:222—229. 

Schuler,  W.  1983.  Responses  to  sugars  and  their  be- 
havioural mechanisms  in  the  starling  {Sturnus  vul- 
garis L.).  Behav.  Ecol.  Sociobiol.  13:243-251. 

Steel,  R.  G.  D.  and  J.  H.  Torrie.  1980.  Principles  and 
procedures  of  statistics.  Second  ed.,  McGraw-Hill 
Book  Co.,  New  York. 

Stiles,  E G.  1976.  Taste  preferences,  color  preferences 
and  flower  choice  in  hummingbirds.  Condor  78: 
10-26. 

Tobin,  M.  E.  and  R.  W.  DeHaven.  1984.  Repellency 
of  methiocarb-treated  grapes  to  three  species  of 
birds.  Agric.  Ecosyst.  Environ.  11:291-297. 

Witmer,  M.  C.  1994.  Contrasting  digestive  strategies 
of  frugivorous  birds.  Ph.D.  diss.,  Cornell  Univ., 
Ithaca,  New  York. 


Wilson  Bull.,  111(1),  1999,  pp.  89-99 


HIERARCHICAL  COMPARISONS  OF  BREEDING  BIRDS  IN  OLD- 
GROWTH  CONIFER-HARDWOOD  FOREST  ON  THE 
APPALACHIAN  PLATEAU 

J.  CHRISTOPHER  HANEY' 


ABSTRACT. — I compared  relative  abundances  of  breeding  birds  in  old-growth  forest  (>300  years  old)  to 
surrounding  landscapes  using  data  from  the  Breeding  Bird  Census  (BBC)  and  Breeding  Bird  Atlas  (BBA). 
Eleven  study  plots  (148  ha  total)  were  established  in  relict,  presettlement  hemlock-white  pine-northern  hardwood 
(Tsiiga  canodensis-Pinus  strobus)  forest  on  the  northern  Appalachian  Plateau,  Pennsylvania.  Of  56  breeding 
species  recorded  in  old-growth  forest,  34%  were  either  uncommon  (:S25%  of  BBA  blocks)  or  rare  (^10%  of 
BBA  blocks)  in  adjacent  landscape  units.  A species  accumulation  curve  indicated  that  about  40  species  recurred 
in  old-growth  habitat.  This  avian  community  included  species  less  likely  to  occur  in  oldgrowth,  forest  interior 
species  showing  a statistically  neutral  relationship  to  oldgrowth,  and  habitat  specialists  more  likely  to  reside  in 
oldgrowth  than  in  the  landscape  at  large.  The  last  group  included  several  taxa  linked  to  structural  features  of 
oldgrowth  elsewhere  in  North  America:  Hairy  Woodpecker  (Picoides  villosus).  Red-breasted  Nuthatch  (Sitta 
canadensis).  Brown  Creeper  (Certhia  americana).  Winter  Wren  (Troglodytes  troglodytes).  Golden-crowned 
Kinglet  (Regidus  satrapa),  Empidona.x  flycatchers,  and  several  species  of  arboreal  Dendroica  warblers.  Received 
14  July  1998,  accepted  4 Nov.  1998. 


Old-growth  forests  possess  unique  ecolog- 
ical characteristics  that  can  exert  profound  in- 
fluence on  some  bird  populations  and  com- 
munities (Hunter  et  al.  1995,  Dellasala  et  al. 
1996).  Ecological  importance  of  oldgrowth  to 
birds  is  poorly  known  in  much  of  North 
America,  largely  because  late  successional 
forest  outside  the  Pacific  Northwest  now  oc- 
curs only  in  relict  patches  (Davis  1996).  Ide- 
ally, the  role  of  old-growth  forest  in  facilitat- 
ing avian  diversity  could  be  best  evaluated  by 
comparing  species  occurrences  within  entire 
landscapes  made  up  of  many  different  habi- 
tats. 

What  is  the  best  way  to  evaluate  bird  dis- 
tributions and  abundances  over  multiple  spa- 
tial scales?  Results  of  studies  on  species  oc- 
currences at  any  one  scale  may  conflict  with 
results  at  alternative  scales  (Conroy  and  Noon 
1996).  In  avian  ecology,  this  concern  may  find 
expression  as  a tradeoff  among  within-habitat 
(a),  between-habitat  ((3),  and  landscape  (y)  di- 
versity (Whittaker  1977,  Wiens  1989).  For 
avian  conservation,  management  actions  at  lo- 
cal scales  must  be  weighed  against  their  con- 
sequences at  broader  scales  in  order  to  opti- 
mize benefits  of  land  use  (Flather  1996).  A 
hierarchical  framework  is  the  method  usually 


‘ Ecology  and  Economics  Research  Dept.,  The  Wil- 
derness Society,  900  17th  Street,  NW,  Washington, 
D.C.  20006;  E-mail:  Jchris_haney(3)tws.org 


recommended  to  address  such  scale  depen- 
dency (Kotliar  and  Wiens  1990). 

I used  a landscape  hierarchy  to  evaluate 
bird  distribution  and  abundance  in  old-growth 
forest,  once  a widespread  vegetation  type  in 
eastern  North  America.  Local  species’  occur- 
rences in  old-growth  conifer-hardwood  forest 
were  compared  to  occurrences  in  the  sur- 
rounding landscape  using  data  from  the 
Breeding  Bird  Census  and  Breeding  Bird  At- 
las. Three  questions  were  posed:  (1)  can  a 
metric  be  devised  to  compare  bird  species  oc- 
currences across  different  spatial  scales,  (2) 
does  old-growth  forest  harbor  birds  deter- 
mined independently  to  be  uncommon  or  rare 
in  larger  landscape  units,  and  (3)  which  indi- 
vidual species  are  more  likely  to  occur  in  old- 
growth than  in  the  landscape  as  a whole? 

METHODS 

Study  area. — Bird  communities  in  oldgrowth  were 
studied  within  three  permanent  forest  reserves  in  Penn- 
sylvania: Cook  Forest  State  Park.  Heart's  Content,  and 
Tionesta  Scenic  and  Research  natural  areas  in  the  Al- 
legheny National  Forest  (41°  2()'-41°  42'  N,  78°  56'- 
79°  15'  W).  Tract  sizes  of  old-growth  habitat  varied 
from  1000  ha  at  Tionesta  to  60  ha  at  Heart's  Content; 
each  of  these  reserves  is  embedded  within  much  larger 
contiguous  tracts  of  younger  managed  forest.  Land- 
scape fragmentation  is  greater  in  Cook  Forest  where 
developed  and  agricultural  lands  virtually  surround  this 
3000  ha  reserve  (which  includes  some  200  ha  of  old- 
growth in  three  sites).  All  reserves  are  located  on  the 
northern  Appalachian  Plateau  (212Ga:  Allegheny  High 


89 


90 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  1,  March  1999 


Plateau  Subsection,  Northern  Unglaciated  Allegheny 
Plateau  Section;  Keys  et  al.  1995),  a region  character- 
ized by  broad,  flattened  ridges  (500-700  m)  and  dis- 
sected by  deep,  V-shaped  valleys.  Higher  precipitation 
and  greater  cloud  cover  create  a cooler,  more  humid 
climate  compared  to  adjacent  regions  (Whitney  1990). 

Each  reserve  possesses  relict  stands  of  true  old- 
growth,  stands  with  extreme  ecological  maturity  (Lev- 
erett  1996).  The  forest  consists  of  hemlock- white  pine- 
northem  hardwoods  (Nichols  1935),  a cover  type  that 
most  closely  resembles  USDA  Forest  Service  CISC 
(continuous  inventory  of  stand  condition)  types  4 and 
8,  Society  of  American  Foresters  forest  type  code  22 
(Eyre  1980),  and  International  Classification  of  Eco- 
logical Communities  I.B.8.N.b.  150  (USDA  1997). 
Canopy  dominants  include  eastern  hemlock  (Tsuga 
canadensis),  eastern  white  pine  (Pinas  strohu.'i),  yellow 
birch  (Betula  alleghaniensis),  black  birch  (Betula  len- 
ta),  sugar  maple  (Acer  saccharum),  red  maple  (Acer 
ruhrum),  and  American  beech  (Fagus  grandifolia\ 
Whitney  1990). 

Understories  are  generally  sparse  with  little  herba- 
ceous ground  cover.  Canopy  trees  at  each  site  are  300- 
530  years  old  (Hough  and  Forbes  1943;  Abrams  and 
Orwig  1996;  Stable  1996;  C.  Nowack,  pers.  comm.), 
and  so  are  beyond  the  threshold  (275  years)  at  which 
unique  structure  begins  to  develop  in  this  community 
type  (Tyrrell  and  Crow  1994).  Typical  of  forest  in  pre- 
settlement condition,  most  stands  have  ecological  at- 
tributes that  are  rare  or  absent  in  younger,  managed 
forests  (e.g.,  57  metric  tons  of  coarse  woody  debris 
ha  ';  >Vs  of  stand  basal  area  in  trees  >70  cm  diameter 
at  breast  height;  Haney,  unpubl.  data). 

Since  the  late  1800s  conifer-hardwood  forest  in  the 
eastern  U.S.  has  been  fragmented  into  isolated  blocks, 
markedly  reduced  in  area,  and  converted  into  cover 
types  dominated  by  younger,  shade  intolerant  hard- 
woods. On  the  northern  Appalachian  Plateau,  the  old- 
growth  hemlock-hardwood  forest  once  covered  2.4 
million  ha  (Bjorkblom  and  Larson  1977),  but  today  it 
is  reduced  to  no  more  than  20%  of  its  presettlement 
extent  (Whitney  1990,  Abrams  and  Ruffner  1995). 

Data  collection. — I used  the  Breeding  Bird  Census 
(BBC;  Hall  1964)  to  evaluate  bird  species  occurrences 
within  old-growth  habitat.  Breeding  Bird  Census  meth- 
odology is  u.sed  primarily  to  assess  local  population 
density  by  counting  the  number  of  breeding  territories 
on  a few  ha.  Five  BBC  plots  (15-18  ha  each)  were 
located  in  Cook  Forest,  two  ( 10-1 2 ha)  in  Heart’s  Con- 
tent, and  four  (12  ha)  in  Tionesta.  Individual  plots 
within  sites  at  each  reserve  ranged  from  200  m to  2500 
m apart.  Breeding  birds  were  counted  during  eight  or 
more  visits  to  each  plot  during  May  and  June  1994 
using  .standard  protocols  (Hall  1964,  Lowe  1993). 
Each  plot  was  visited  on  a different  day,  usually  within 
a few  minutes  of  sunrise.  Two  visits  were  made  at 
dusk.  Each  visit  la.sted  about  two  hours,  which  resulted 
in  a census  speed  of  about  9 min  ha  ' and  is  compa- 
rable to  speeds  deemed  appropriate  for  relatively  open 
forests  (Engstrom  and  James  1984). 

On  each  visit,  an  observer  walked  slowly  along  a 


flagged  census  line  through  the  plot,  delineating  all 
bird  territories  on  grid  maps.  Birds  were  detected  both 
visually  and  acoustically,  but  most  detections  were 
acoustic.  The  census  line  was  configured  to  place  the 
observer  no  more  than  50  m from  any  part  of  the  plot 
so  as  to  reduce  detection  bias  from  acoustic  attenuation 
(Schieck  1997).  Numbers  of  territories  were  then  cal- 
culated from  grid  maps  using  standard  spot-mapping 
procedures  (Hall  1964). 

The  Pennsylvania  Breeding  Bird  Atlas  project 
(1983-1989)  was  a grid-based  survey  using  techniques 
developed  originally  in  Britain  and  Ireland,  with  stan- 
dards modified  for  the  northeastern  U.S.  (Laughlin 
1982).  Atlas  projects  are  used  primarily  for  broad  map- 
ping of  avian  distributions  and  rely  upon  a network  of 
volunteer  field  ornithologists  to  document  breeding  ev- 
idence at  three  levels  of  certainty  (“possible,”  “prob- 
able,” and  “confirmed”).  In  Pennsylvania,  the  basic 
sampling  units  consisted  of  7.5'  U.S.  Geological  Sur- 
vey topographic  maps  divided  into  six  equal-size 
blocks  formed  longitudinally  by  3.75'  intervals  and 
latitudinally  by  2.5'  intervals  (Brauning  1992).  Atlas 
efforts  were  undertaken  in  both  summer  and  winter 
within  known  “safe”  dates  for  nesting  activity  of  all 
species.  Based  on  previous  theoretical  and  empirical 
work,  blocks  were  considered  adequately  covered  if 
75-80%  of  the  expected  species  were  found,  10-20 
hours  of  survey  effort  were  expended,  or  70  or  more 
species  were  recorded. 

Community  level  analyses. — To  test  whether  sam- 
pling effort  was  adequate  for  characterizing  the  total 
species  complement  (5„,3J  of  the  old-growth  bird  com- 
munity, I conducted  two  analyses  on  the  area  curve  of 
cumulative  species  richness  (5).  The  shape  of  species 
accumulation  curves  depends  on  the  order  in  which 
samples  are  added,  a feature  not  modeled  well  with 
parametric  methods  (Bunge  and  Fitzpatrick  1993).  I 
used  non-parametric  routines  to  randomize  sample  or- 
der (PISCES  1.2  software,  Windows  95  version;  Hen- 
derson and  Seahy  1997).  For  greater  resolution  in  con- 
structing the  species  accumulation  curve,  I first  sub- 
sampled the  BBC  data  at  a scale  of  3 ha.  From  each 
and  all  of  the  1 1 original  study  plots,  I randomly  se- 
lected and  ordered  3 ha  subplots  and  scored  bird  spe- 
cies occurrences  and  territorial  densities  using  methods 
identical  to  those  used  in  the  original  large  plots. 

Accurate  estimation  of  is  possible  only  if  the 
species  accumulation  curve  is  derived  from  a homoge- 
nous community  (Henderson  and  Seahy  1997).  I first 
compared  the  mean  randomized  curve  (1000  itera- 
tions) with  a curve  expected  if  all  individual  birds  re- 
corded over  all  the  samples  were  assigned  randomly 
to  individual  samples  (Colwell  and  Coddington  1994). 
If  the  expected  curve  (Coleman  et  al.  1982)  rises  more 
sharply  from  its  origin,  then  heterogeneity  is  greater 
than  can  be  explained  by  chance.  Such  a result  could 
indicate  that  the  samples  were  a combination  of  dis- 
tinct bird  communities  or  derived  from  different  hab- 
itats (Flather  1996). 

Asymptotic  models  of  species  accumulation  curves 
are  usually  appropriate  for  homogenous  communities 


Haney  • OLD-GROWTH  BIRD  COMMUNITY 


91 


(Henderson  and  Seahy  1997).  I calculated  using  a 
non-parametric  maximum  likelihood  estimater  (Raa- 
ijmakers  1987)  in  which  sampling  is  assumed  to  be 
complete  when  the  asymptotic  estimate  is  equal  to  or 
less  than  the  observed.  This  procedure  was  applied  in- 
crementally to  larger  combinations  of  randomly  shuf- 
fled 3 ha  subplots  (1000  iterations  each)  until  the 
“stopping  rule”  indicated  that  sampling  of  the  old- 
growth  bird  community  was  sufficient. 

Species  level  analyses. — I used  incidence  (frequency 
in  a set  of  samples;  Wright  1991)  as  the  metric  to 
compare  individual  species’  occurrences  in  oldgrowth 
to  their  occurrence  in  landscape  units.  Incidence  in 
old-growth  samples  was  calculated  by  dividing  the 
number  of  plots  containing  each  species  by  11.  For 
Pennsylvania  and  the  northern  Appalachian  Plateau,  I 
used  the  proportion  of  BBA  blocks  recording  that  spe- 
cies for  each  of  the  two  landscape  divisions.  The  atlas 
program  covered  a total  of  4928  and  2027  BBA  blocks 
state  and  province  wide,  respectively  (Brauning  1992). 
Species  recorded  in  less  than  25%  and  10%  of  BBA 
blocks  in  either  landscape  division  were  considered 
uncommon  and  rare,  respectively. 

I compared  incidence  in  oldgrowth  (Iqg)  to  inci- 
dence statewide  (1st)  and  province-wide  (I^p)  with  the 
normal  deviate,  Z,  where: 

z = (I,  - y/(i  X (1  - I)(1/N,  + i/N2))°'5, 

and  I and  1 — I are  the  joint  probabilities  of  the  com- 
bined incidences  in  the  two  sample  proportions  of  find- 
ing and  not  finding  that  species,  respectively  (Snedecor 
and  Cochran  1980).  I used  Pearson’s  product  moment 
correlation  to  test  whether  incidence  was  related  to  the 
natural  log  of  population  density  (number  of  breeding 
territories).  Log  transformations  on  population  density 
were  used  to  smooth  variances  in  data  composed  of 
whole  integers  (Snedecor  and  Cochran  1980).  Values 
of  test  statistics  were  considered  significant  at  P < 
0.05  unless  otherwise  indicated. 

RESULTS 

Community  composition. — Fifty-six  species 
were  recorded  in  148  ha  of  old-growth  forest 
across  the  11  study  plots  (Table  1).  Thirteen 
species  were  found  in  only  one  plot  (incidence 
value  = 0.091).  Another  species,  Downy 
Woodpecker  (Picoides  pubescens),  was  re- 
corded in  two  plots,  but  less  than  one  full  ter- 
ritory was  recorded  in  each  plot.  Without  con- 
sidering these  14  species,  a recurring  comple- 
ment of  42  species  was  identified  in  which  full 
breeding  territories  were  established  in  two  or 
more  of  the  1 1 study  plots  (Table  2). 

The  observed  species  accumulation  curve 
(Fig.  1)  did  not  differ  from  the  curve  expected 
in  a homogenous  community  (x^  = 0.25,  P > 
0.05,  df  = 27).  Thus,  this  analysis  gave  no 
indication  that  more  than  one  bird  community 


was  being  sampled.  The  estimated  asymptotic 
value  (38.6)  for  species  richness  fell  below  the 
observed  value  (39)  after  1000  randomiza- 
tions of  28  3-ha  subplots.  This  level  of  effort 
corresponded  to  84  ha  (57%)  of  the  total  area 
actually  sampled  in  this  study. 

Species  groups. — About  one-third  of  all 
species  recorded  as  breeders  in  old-growth  co- 
nifer-hardwood were  either  uncommon  or  rare 
over  broad  spatial  scales  (Table  1).  Nineteen 
species  (34%)  were  more  likely  to  occur  in 
oldgrowth  than  in  the  landscape  unit  consist- 
ing of  the  entire  state.  Sixteen  species  (29%) 
were  more  likely  to  occur  in  oldgrowth  than 
in  the  landscape  unit  of  the  northern  Appala- 
chian Plateau.  Fifteen  individual  species  were 
more  likely  to  occur  in  oldgrowth  than  in  the 
landscape  at  both  state  and  province  levels 
(Table  2).  Red-shouldered  Hawk  (Buteo  linea- 
tus)  and  Barred  Owl  (Strix  varia)  were  more 
likely  to  occur  in  oldgrowth  than  in  the  land- 
scape unit  consisting  of  the  entire  state  but  not 
the  northern  Appalachian  Plateau. 

Seventeen  species  were  less  likely  to  occur 
in  old-growth  forest  than  in  the  landscape  at 
large  (Table  2).  This  group  included  perma- 
nent resident,  habitat  generalists  [e.g.,  Amer- 
ican Crow  (Corvus  brachyrhynchos)]  as  well 
as  some  Neotropical  migrants  with  more  spe- 
cific habitat  preferences  [e.g.,  cavity-nesting 
Great  Crested  Flycatcher  {Myiarchus  crini- 
tus)\. 

No  species  showing  negative  association 
with  oldgrowth  (Table  2)  was  rare  at  the  state 
level,  and  none  of  the  species  in  this  group 
was  either  rare  or  uncommon  at  the  level  of 
the  physiographic  province  (Table  1).  Only 
one  species.  Black-throated  Blue  Warbler 
(Dendroica  caerulescens),  was  uncommon  at 
the  state  level  (15%  of  BBA  blocks).  Most 
species  negatively  associated  with  oldgrowth 
were  very  widespread  within  broad  landscape 
units,  occurring  in  50-90%  of  the  BBA 
blocks. 

Based  on  statistical  criteria,  10  species  were 
neither  more  nor  less  likely  to  occur  in  old- 
growth than  in  at  least  one  of  the  larger  land- 
scape units  (Tables  1 and  2).  All  species  in 
this  group  rely  upon  forest  interior  habitat,  in- 
cluding the  raptors  Red-shouldered  Hawk  and 
Barred  Owl,  and  Neotropical  migrant  song- 
birds such  as  Red-eyed  Vireo  (Vireo  oliva- 


TABLE  1.  Relative  occurrence  of  breeding  birds  in  old-growth  hemlock-white  pine-hardwood  forest  on  the  northern  Appalachian  Plateau,  Pennsylvania.  Incidence 
(I)  of  each  species  in  oldgrowth  (proportion  of  study  plots,  n = 11)  is  compared  to  incidence  across  the  entire  state  and  within  the  physiographic  province  [proportion 
of  Breeding  Birds  Atlas  (BBA)  plots  recording  the  species;  n = 4928  and  2027  blocks,  respectively]. 


92 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


N 


0.2 
a.  r3 

<5: 


^ H 

V) 


O O 

*0 


in  lo 

o o o 


o o 

A A 


in 

o 

d 

A 


in 

o 

d 

A 


o 

o 

d 

V 


tn 

ro 

o 


o 

o 

d 

V 


VO  o 
in  ov 

d d 
I 


ON  ro 

' rn 


CN  ro  VO  Tf  vD 

— ' m (N  00  vO 

' d ■ ■ 


0 

1 


I 


o 

o 

d 

V 


Ov 

I 


o o 


VO 

o 

o 

d 


o 

o 


o 

o 

d 

V 


o 

o 

d 

V 


o 

o 

d 

V 


o 

o 

d 

V 


-vt  (N  in  ro 
in  in  ^ _ 


— ' 00 

00  in  in 


cn 


<n  m 
I I 


-It 


in  vD 
vq  C' 
d 00  ri 
I 


■'tr^mvoinin-rj- 
O 00  cn  ^ (N 
r-'-^ddr^vDodm 

- I r I 


VD 

in  in  (N 
‘ o o 
d d 
A 


o 

d 

A 


in 

o 

d 

A 


o 

o 

d 

V 


in 

o 

d 

A 


o 

o 

d 

V 


o 

o 

d 

V 


vO 

o 

o 


o 

o 

d 

V 


o 

o 

d 

V 


VO 

O 

o 


o 

o 

d 

V 


ov  in  00  o ro  in  in 

OV  O'  VO  ov  (N  — ^ 


0 — 

1 


cn 

I 


o (N  in 

O'  00  O 

— ' O'  d 00  in  rn 


— VO  00  O'  O' 
_ _ _ (N|  00  O' 


O' 

o in 


I 


I 


I 


d •sf 


oommmO'cnt^O't^r' 

^0(Nin'^m''^^ino 

cNvodiNdirioddinin 

I rsi  I (N  I 


o m in  C' 

O (N  m 
n)  m 00  (N 

d d d d 


•'ifO'inmmvooovo 
O— 'in^t^vovoo 
t^rooovoooinooo 
dddddddd 


'It  VO  O 
m in  It 


O'  It 
m VO  O' 


O 
-It 


cNvqvqovO'(Nvoo' 

dddddddd 


00  tn 

d d 


O'  CN  (N  O' 

O'  O'  00  —I 

O 00  O'  O' 


o o o o 


Or)— ;o^(Nt''.— -oovDO'inooo<nrnr':0^0;-Hooooooor'iooo(NO'0'^cN 

oooooooodddddddddddddddddddddddd 


— '■^minrrmno' 
O'O'vot^ini^t^ooo 
oorrn^nr)— 'O' 
ddddddddd 


in  — -H00(Nrivonr)m 
'fO'O'  — oooomoooot^ 
inoooq  — — VO  — — r) 

dddddddddd 


O'  <N  VO 

o 00  m 


O' 

d 


VO 

d 


(N  vt  n 

00  VO  00 

— m — 

d d d 


=o 

s: 

£0 


if 

2 '3 


2 


C) 


A 5 


a 

X 

“O 

u 

c 

c 


a 

X 

T3 

O 

u« 

1) 

2 

3 

o 

-C 

c/5 


a d 

? '3 

A > 

N 

^ s 


d 

’■2  u, 
OX)  B 

c ^ 

'c 

C 3 
3 1/3 

3 rv 
3 ct 

X ^ 
s .JJ 
o 


c 


5u  s; 

cx  -i: 


9^ 

s 


a *0 
x: 

cn  oC 


<u 

X) 

V ^ 


1-1  W 

3 <U 

V.  ^ 

I ° 

>>  ^ 
C >, 
^ .b 

O 3 

Q X 


I 

§ 

1-1 

O 

u 


- 2 
E 


§■  E 
o 2 
C‘  5 
Q U 

U (U 
<D  (D 

O ^ 

a ^ 
V -6 


.>1 

j-i 

I 


ii 

? t 

3 


In 

'S. 


o . 


c 

B £ 
S Si 

D c/5 
rz  CT3 
CL  W 


O u 

>>  (U 

E o 

4-* 

•O  3 

D O 

"B  tL 
X3 


_ g 

, ^ S 

S-  3 ^ 
S:  o 25 

^ ^ -S 

^ E 2 
u o 
-o  s: 

JC  1)  Q 

o ^ 

c/5 

C3  (U  U 

pi'  ^ >v 
pL,  t-  Cd 

(U  •-5 

^ ti 
c/5  C3 

(U  3 

<U  t 
J O CQ 


s: 

■2 

2 

;n 

E 

O 

o 


o 

C) 

■S  d 

^ r 
-3  ^ 

"5  ^ r- 

^11 


o 

Q 

ij 


tn 


cn 


^ =3 


(u  n 

C4l  U 
3 


'g  -S 
S "S 
Jr* 

t;  ^ 


2 a 

c/5  (U 
C3  (U 
D »-) 

i I 

E p 


c e -p 


oo  a ^ 
r ^ ^ 

b rSi  2 

*-  Q;  ^ 

^ W <3 

^ -a 

^ (U  ■*** 

^ OOCJ 
,2  3 0 

2?  2 
p ^ t? 

d T3  3 

fc  ^ I 

S 

^ V S 


s ^ 

2 ^ 

1 i 
2l 

a -2 
■s;  Cl 

2 -2 


5 

2 

2 

b! 

C 

c 


S 2 


o 

In 

G O 
d 2 

■CS  -i; 

S 


o 1) 
3 2 

o 


s: 

2 ^ 
jz  2 

4-1  C 

*S 

^ O 

0)  P 
X ^ 


.2  J= 

^ 3 

z 

T3 

6 

Q.  1/3 

V 2 

2;  E 
- o 5 

C4-(  cy  T3 

H03a:^cQ^Ocna:^<um 


•S  > 

X 

$ 


TABLE  1.  CONTINUED 


Haney  • OLD-GROWTH  BIRD  COMMUNITY 


93 


a." 

O.  <T3 

<a: 


4) 


lo 

O 


1 

rp 

r<P 

1 

1 

1 

, 

1 

rp 

OP 

OP 

o 

o 

op 

OP 

O 

o 

o 

O 

o 

o 

OP 

o 

op 

op 

o 

o 

op 

o 

o 

O 

o 

O 

O 

o 

o 

O 

O 

O 

o 

1 ° 

O 

o 

o 

O 

o 

O 

O 

o 

o 

O 

o 

o 

O 

o 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

A 

A 

V 

V 

A 

A 

V 

V 

V 

V 

A 

V 

A 

A 

V 

V 

A 

V 

V 

o\ 

CTv 

so 

so 

in 

00 

o 

(N 

ro 

so 

VO 

OS 

r- 

in 

so 

<n 

VO 

O 

(N 

00 

m 

cn 

1 ^ 

r-; 

o 

q 

(N 

rp 

q 

so 

— 

00 

d 

d 

d 

ri 

1 

CN 

un 

d 

in 

d 

r-J 

d 

cn 

ri 

d 

I I 


I I I 


OP 

OP 

o 

(N 

O 

OP 

o 

o 

o 

CN 

fN 

O 

(N 

O 

OP 

o 

OP 

OP 

OS 

(N 

o 

OP 

o 

o 

o 

o 

o 

o 

o 

o 

o 

o 

1 

o 

o 

o 

o 

o 

o 

o 

O 

o 

o 

o 

o 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

A 

A 

V 

A 

V 

V 

V 

A 

V 

A 

A 

V 

A 

V 

V 

o 

<n 

(N 

(N 

so 

so 

o 

SO 

00 

SO 

Os 

in 

Os 

O 

cn 

CN 

C^l 

so 

o 

Ov 

00 

CC 

ro 

q 

Ov 

1 ^ 

(N 

00 

O) 

SO 

00 

m 

CN 

O) 

m 

q 

"=t 

d 

d 

<N 

ri 

d 

vd 

1 

(N 

(N 

CN 

d 

CN 

d 

c^i 

in 

mi 

CN 

I I 


I I I 


I r I 


Os 

Ov 

vO 

m 

CN 

VO 

o 

r' 

so 

r- 

m 

m 

rp 

m 

VO 

00 

m 

»n 

Ov 

00 

r- 

(N 

00 

op 

1 

op 

o 

Os 

CN 

O 

m 

cn 

C^l 

cn 

00 

CN 

Os 

O 

VO 

cn 

CN 

in 

rp 

1 

o 

in 

r- 

00 

o 

m 

Ov 

00 

Os 

Ov 

m 

00 

OS 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

o 

so 

VO 

o 

00 

r- 

so 

00 

o 

00 

00 

Os 

so 

cn 

o 

o 

m 

m 

VO 

so 

00 

1 

VO 

CN 

o 

so 

00 

VO 

CO 

r- 

o^ 

os 

O 

m 

1 

o 

so 

r- 

o 

Os 

CN 

00 

so 

Os 

ov 

CN 

00 

CN 

Os 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

o 

o 

m 

o 

o 

3" 

CN 

'T 

CN 

o 

Ov 

o 

'T 

Os 

o 

Ov 

ov 

o 

in 

Ov 

o 

o 

OS 

VO 

Os 

00 

SO 

o 

00 

so 

o 

so 

Os 

o 

c 

SO 

o 

Os 

q 

o 

o 

q 

o 

q 

q 

O 

rp 

o 

cn 

o 

m 

q 

CO 

o 

Ov 

q 

cn 

Os 

o 

d 

d 

d 

d 

"“V 

d 

d 

d 

d 

d 

d 

d 

d 

-V 

d 

d 

d 

d 

d 

d 

“2 

C 

y-S  ^ 

2 d 

>} 

^3 

§2 

"S 

Sj 

£ 

C JV 

'3 

Sj 

So 

13 

C 

2 § 

5 

5 Q 

£ 

13 

C3 

Q 

c to 
"S  u 

SJ 

^3 

5 

C 

Q 

U 

Q -B 

s 

2 I 
02  >1 

2 


a 

c 

a 

> 


<3  -ii 


-s:  a 

-S-f 

^ .=0 


•t  c 
> Q, 

0 

1)  "5 

0- 
u £ 

(U 

V j= 

•a  t 

(U  O 

01  Z 


“ m 


OJ  ^ 

S Q 

U ■>-' 

rt  . 

•o  ■£  -a 
<L>  d (U 

:h  ^ s 

c/5  O 

' d 5- 

3 - ■£ 

C O I 

c ^ 

S 2P 

— d 


X5 

d 


c 

•D 

<j  (L) 

"S 

a 

n 


"2 

S!  ^ 


^ (U 

■e  3 

b Ui 


PQ 


Q, 

2 « 

s 3 

^ !3 

13  ^ 

s ^ 

Cl  B 

1-  ^ 
(U  ? 

^ s 

^ d 


5^  ^ 

i'l 

^ 2 o 


CO  2 
^ ^3 

t:  ^ 

d 5 

-o  -2 

V ^ 

OeC  52 

d u 

cj  .- 


?:  O 

ScB 

o ^ 

^ t; 


u. 

u t; 


I 

0{) 

c 


w --  qj  c-# 

iH  « c « 

m CQ  ix  pq 


o 

=:  xi 
c ^ 

O XJ 
(U 
ID 
O 
O 


5 


^ c: 

CJ  o 
'D  5 

§ 

?3  ^ 
00  ^ 
c*  d 

2 ^ 

^ £ 
o 

'U 
00 
C3 


.c 
'Z'  ^ 

c 

^ S 


C 

i2 


S 
E 

U X ^ 1^ 


~ lU 
C3  « 


c •!:/ 

tb  ^ 

^ s 

a 

Q.  ^ 
^ o 
00  t 

C C3 

■a  a 

C C/P 

m ^ 

o '5. 

.°‘^  a 

V 3 
3 U 


5U  O 

■^'5 

^ TO 

13 

o 

o u 

o 

c -o 
3 U 
■—I  "3 
■n  ^ 
3 


^ 'S 

1^ 

S.  a 

s ;3 
■2  ^ 
^ <— 
^ O 

2 c 

"O 


V c 

d u 

Q CO 


-9  ^ 


^ Test  statistic  (normal  deviate)  on  the  difference  between  proportions:  positive  values  of  Z indicate  a greater  incidence  of  that  species  in  oldgrowth,  negative  values  indicate  lower  incidence  in  oldgrowth. 

^ Unrecorded  during  the  state  s 5 year  BBA  project.  From  1993-1995,  an  irruption  of  the  elm  spanworm  Ennomus  subsignarius  affected  much  of  Pennsylvania’s  forests.  Several  birds  influenced  by  geometrid  outbreaks  were 
recorded  during  these  breeding  seasons  for  the  first  time:  Bay-breasted  Warbler,  Blackpoll  Warbler  (Dendroica  striata)  and  Evening  Grosbeck  {Coccothraustes  vespertinus). 


94 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  1,  March  1999 


TABLE  2.  Number  of  territories  and  population  densities  (number  territories/ 1 0 ha)  of  the  core  community 
of  breeding  birds  in  all  plots  {n  = 1 1 ) of  old-growth  hemlock-white  pine-hardwood  forest  on  the  northern 
Appalachian  Plateau,  Pennsylvania. 

Territorial  density 

Association:  Species“  Total  territories  Mean  SE 


Positive  old-growth  affinity: 


Hairy  Woodpecker  (Picoides  villosus) 

12.5 

0.50 

0.42 

Acadian  Llycatcher  (Empidonax  virescens) 

24.5 

1.27 

1.57 

Red-breasted  Nuthatch  (Sitta  canadensis) 

10 

0.40 

0.36 

Brown  Creeper  (Certhia  americana) 

31 

1.58 

0.71 

Winter  Wren  {Troglodytes  troglodytes) 

30.5 

1.73 

0.98 

Golden-crowned  Kinglet  (Regulus  satrapa) 

4 

0.10 

0.23 

Swainson’s  Thrush  (Catharus  ustulatus) 

35 

2.08 

2.42 

Hermit  Thrush  {Catharus  guttatus) 

36.5 

1.90 

1.49 

Blue-headed  Vireo  {Vireo  solitarius) 

88 

4.57 

2.26 

Magnolia  Warbler  {Dendroica  magnolia) 

185 

11.10 

2.70 

Black-throated  Green  Warbler  {Dendroica  virens) 

176 

9.41 

5.36 

Blackburnian  Warbler  (Dendroica  fusca) 

410 

23.94 

8.25 

Pine  Warbler  {Dendroica  pinu.s) 

5 

0.17 

0.25 

Dark-eyed  Junco  {Junco  hyemalis) 

83 

4.09 

0.92 

Purple  Pinch  {Carpodacus  purpureas) 

17 

0.72 

0.52 

Neutral  oldgrowth  affinity: 

Red-shouldered  Hawk  {Buteo  lineatusf 

2.5 

0.09 

0.13 

Barred  Owl  (Strix  varia)'° 

4.5 

0.18 

0.24 

Yellow-bellied  Sapsucker  {Sphyrapicus  varius) 

4 

0.14 

0.26 

Pileated  Woodpecker  {Dryocopus  pileatus) 

5.5 

0.25 

0.31 

Common  Raven  {Corvus  corax) 

1.5 

0.07 

0.19 

Black-capped  Chickadee  {Poecile  atricapillus) 

24 

1.14 

0.62 

Red-eyed  Vireo  {Vireo  olivaceus) 

110 

5.22 

3.06 

Hooded  Warbler  {Wilsonia  citrina) 

15 

0.87 

1.40 

Scarlet  Tanager  {Piranga  olivacea) 

42.5 

2.23 

1.08 

Chipping  Sparrow  {Spizella  passerina) 

20 

1.00 

0.63 

Negative  old-growth  affinity: 

Mourning  Dove  {Zenaida  macroura) 

3 

0.09 

0.21 

Ruby-throated  Hummingbird  {Archilochus  coluhris) 

2 

0.06 

0.11 

Least  Llycatcher  {Empidonax  minima.^) 

3.5 

0.18 

0.53 

Great  Crested  Llycatcher  {Myiarchus  crinitus) 

1.5 

0.08 

0.18 

Blue  Jay  {Cyanocitta  cristata) 

10.5 

0.45 

0.35 

American  Crow  {Corx’us  brachyrhynchos) 

1.5 

0.06 

0.14 

Tufted  Titmouse  {Baeolophus  hicolor) 

2 

0.11 

0.19 

White-breasted  Nuthatch  {Sitta  carolinensis) 

7.5 

0.27 

0.31 

Wood  Thrush  {Hylocichla  mustelina) 

1.5 

0.09 

0.25 

American  Robin  (Tiirdus  migratorius) 

4.5 

0.20 

0.40 

Cedar  Waxwing  (Bombycilla  cedrorum) 

2 

0.10 

0.23 

Black-throated  Blue  Warbler  {Dendroica  caerulescens) 

18 

1.08 

1.46 

American  Redstart  {Setophaga  ruticilla) 

7.5 

0.43 

1.28 

Ovenbird  (Seiurus  aurocapillus) 

8 

0.46 

0.95 

Common  Yellowthroat  {Geothlypis  tricha.'i) 

2 

0.06 

0.14 

Rose-breasted  Grosbeak  (Pheucticus  ludovicianus) 

5 

0.24 

0.44 

Brown-headed  Cowbird  {Molothrus  ater) 

9 

0.32 

0.58 

Does  not  include  species  found  only  in  one  plot  or  for  which  less  than  one  full  territory  was  recorded  (Table  1). 
^ Positively  associated  with  oldgrowth  at  landscape  level  of  entire  state  but  not  at  level  of  physiographic  province. 


Hane\  • OLD-GROWTH  BIRD  COMMUNITY 


95 


FIG.  1.  Mean  randomized  accumulation  curve 
(1000  iterations)  of  species  richness  in  the  bird  com- 
munity of  old-growth  conifer-hardwood  forest  on  the 
northern  Appalachian  Plateau,  Pennsylvania.  Horizon- 
tal line  indicates  the  putative  asymptote  of  species 
richness  as  a function  of  area  sampled.  The  asymptote 
was  identified  with  a maximum  likelihood  estimator. 


cells).  Hooded  Warbler  (Wilsonia  citrina),  and 
Scarlet  Tanager  (Piranga  olivacea). 

Incidence  versus  population  density.- — Bird 
species  that  were  rare  or  uncommon  at  land- 
scape levels  typically  had  low  population  den- 
sities locally  as  well.  Incidence  explained 
slightly  more  than  80%  of  the  variation  in  the 
natural  log  of  population  size  as  assessed  by 
territorial  density  (Fig.  2). 

DISCUSSION 

Old-growth  affinities. — At  both  province 
and  state  levels,  more  than  one-third  of  bird 
species  were  more  likely  to  occur  in  old- 
growth  conifer-hardwood  forest  than  in  the 
broader  landscape.  These  species  included 
some  of  Pennsylvania’s  rarest  breeding  birds. 
Yellow-bellied  Flycatcher  (Empidonax  flavi- 
ventris)  and  Swainson’s  Thrush  (Catharus  us- 
tulatus)  are  listed  in  the  state  as  threatened 
and  candidate-rare,  respectively  (D.  A.  Gross, 
pers.  comm.).  Pending  investigation  of  specif- 
ic habitat  preferences  for  individual  species, 
the  15  birds  in  this  group  (Table  2)  are  best 
regarded  as  old-growth  associates  rather  than 
old-growth  obligates.  Nevertheless,  it  is  no- 
table that  these  species  have  diverse  habitat 
affinities,  including  conifer  [e.g..  Red-breasted 
Nuthatch  {Sitta  canadensis)],  hardwood  [e.g.. 
Blue-headed  Vireo  (Vireo  solitarius)],  and 
mixed  forest  cover  types  [e.g..  Hairy  Wood- 
pecker (Picoides  villosus)]. 

Several  taxa  identified  as  old-growth  asso- 
ciates in  this  study  have  been  linked  repeat- 
edly to  late  successional  forest  elsewhere 
throughout  North  America.  Hairy  Woodpeck- 


4 

>- 
in  -t-> 

.2?  m 2 
o c 
0)  0) 

Q.  -O 

c 0 

l4_  O 

o ::: 
ro 

3 -2 

c a. 

— o 
a. 

-4 

0 0.2  0.4  0.6  0.8  1.0 

Species  landscape  incidence 

FIG.  2.  Relationship  between  natural  log  of  pop- 
ulation density  (number  of  breeding  territories)  and 
landscape  incidence  of  all  bird  species  (n  = 56)  found 
in  old-growth  conifer-hardwood  forest  of  the  northern 
Appalachian  Plateau,  Pennsylvania.  Landscape  inci- 
dence is  based  on  the  proportion  of  old-growth  study 
plots  (n  = 11;  Table  1 ) in  which  the  species  was  re- 
corded. Log  of  population  density  (y)  is  related  to 
landscape  incidence  (x)  by:  y = 4.74 lx  — 3.513,  (r^ 
= 0.805,  P < 0.001). 

ers  rely  on  mature  forests  with  large  snags  and 
logs  (Anthony  et  al.  1996,  Shackelford  and 
Conner  1998).  Across  broad  geographic  do- 
mains, Red-breasted  Nuthatch,  Brown  Creep- 
er (Certhia  americana).  Winter  Wren  {Trog- 
lodytes troglodytes),  and  Golden-crowned 
Kinglet  (Regulus  satrapa)  form  a recurring 
group  that  exploits  coniferous  cover  and  com- 
plex structure  typical  of  older  natural  forests 
(DeGraaf  et  al.  1992,  Hansen  et  al.  1995, 
Schieck  et  al.  1995,  Anthony  et  al.  1996,  Del- 
lasala  et  al.  1996). 

In  other  cases,  regional  counterparts  of  gen- 
era exhibited  a common  tendency  to  prefer 
mature  forest.  Acadian  Flycatcher  {Empidon- 
ax virescens)  showed  a greater  likelihood  of 
occurring  in  mesic  old-growth  forest  relative 
to  adjacent  landscapes  (Table  1),  similar  to 
Hammond’s  {E.  hammondii)  and  Pacific-slope 
{E.  difficilis)  flycatchers  in  Oregon  and  British 
Columbia  (Hansen  et  al.  1995,  Schieck  et  al. 
1995).  Like  their  western  congeners  Hermit 
{Dendroica  occidentalis)  and  Townsend’s 
warblers  {D.  townsendi;  Hansen  et  al.  1995, 
Schieck  et  al.  1995),  several  species  of  arbo- 
real Dendroica  warblers  were  far  more  likely 
to  occur  in  old-growth  hemlock-white  pine- 
northem  hardwood  forest  than  in  adjacent 


96 


THE  WILSON  BULLETIN  • VoL  III,  No.  I,  March  1999 


landscape  units  (Tables  1 and  2).  Populations 
of  Blackburnian  Warblers  (D.  fitsca)  achieve 
particularly  high  densities  in  the  oldest  coni- 
fer-hardwood forests  of  this  region  (Haney 
and  Schaadt  1996:  fig.  6.1). 

Two  species,  Red-shouldered  Hawk  and 
Barred  Owl,  were  more  likely  to  occur  in  old- 
growth  than  across  the  state  as  a whole  (Table 
1).  Red-shouldered  Hawks  depend  on  mature 
forests  with  large  trees  for  nest  sites  (Titus  and 
Mosher  1981,  Moorman  and  Chapman  1996). 
Barred  Owls  exhibit  greater  territorial  occu- 
pancy and  breeding  propensity  in  this  region’s 
old-growth  forest  (Haney  1997). 

Despite  directional  biases  in  comparisons  of 
incidence  across  spatial  scales  (see  Sampling 
adequacy),  negative  associations  of  bird  spe- 
cies with  oldgrowth  may  have  had  biological 
causes.  Mourning  Dove  {Zenaida  macroura). 
Blue  Jay  (Cyanocitta  cristata),  American 
Crow,  and  Brown-headed  Cowbird  (Moloth- 
rus  ater.  Table  2)  all  typically  exploit  land- 
scapes with  extensive  anthropogenic  distur- 
bance (Martin  1988,  Hoover  and  Brittingham 
1993,  Seitz  and  Zegers  1993,  Rodenhouse  et 
al.  1995).  Consequently,  they  would  be  less 
expected  to  occur  in  mature  tracts  of  reserved 
forest.  Least  Flycatcher  {Empidonax  mini- 
mus), Black-throated  Blue  Warbler,  American 
Redstart  {Setophaga  ruticilla),  and  Rose- 
breasted Grosbeak  (Pheucticus  ludovicianus) 
more  commonly  exploit  the  deciduous  habi- 
tats (Sherry  and  Holmes  1988,  Steele  1993, 
Yahner  1993)  typical  of  younger,  regenerating 
forest  now  prevalent  in  this  region  (Alerich 
1993).  Ovenbirds  (Seiurus  aurocapillus)  were 
probably  scarce  because  of  their  preference 
for  heavy  ground  cover  (Burke  and  Nol  1998), 
a microhabitat  virtually  absent  in  the  old- 
growth  forest  studied  here. 

Sampling  adequacy. — Deletion  of  periph- 
eral species  and  analysis  of  the  asymptote  on 
the  species  accumulation  curve  gave  similar 
values  for  total  species:  42  and  39  species, 
respectively.  I conclude  that  census  effort  was 
adequate  for  characterizing  the  avian  com- 
munity in  old-growth  hemlock-white  pine- 
northern  hardwood  forest  on  Pennsylvania’s 
northern  Appalachian  Plateau.  Numerical  an- 
alyses indicated  that  the  community  sampled 
was  in  fact  homogenous  and  an  asymptotic 
limit  to  species  richness  (5’, „;,;<)  was  achieved 
with  little  more  than  half  the  sampling  effort 


actually  undertaken  (Fig.  1).  Support  for  sam- 
pling adequacy  is  reassuring  because  limited 
amounts  and  local  distributions  of  eastern  old- 
growth  forest  often  preclude  obtaining  larger 
sample  sizes  and  greater  sample  dispersion  in 
this  scarce  habitat  type. 

Use  of  bird  species  richness  {S  or  S^^)  to 
evaluate  avian  habitat  can  be  problematic  un- 
less studies  account  for:  (1)  “core”  members 
of  the  avifauna,  (2)  quantity  and  quality  of 
sampling  effort,  (3)  number  of  habitat  types 
within  areas,  and  (4)  proximity  of  other  hab- 
itats (Remsen  1994,  Elphick  1997).  The  BBC 
method  itself  purposefully  discounts  non-ter- 
ritory holders,  thereby  eliminating  nonbreed- 
ing species.  I also  established  study  plots 
within  interiors  of  old-growth  forest  so  as  to 
avoid  inflating  or  confounding  species  rich- 
ness caused  by  proximity  of  different  habitat 
patch  types  (Flather  1996). 

Although  BBC  and  BBA  methods  have  dis- 
tinct purposes  and  can  have  different  quanti- 
ties and  qualities  of  observer  effort,  several 
factors  facilitated  comparisons  of  data  from 
the  two  techniques  in  this  study.  First,  species 
occurrence  data  from  both  methods  were  com- 
parable by  developing  a common  incidence 
metric.  Second,  both  methods  rely  to  some  ex- 
tent upon  a measure  of  saturation  in  the  cu- 
mulative number  of  species  recorded  in  order 
to  guage  whether  sampling  is  adequate.  Third, 
both  methods  had  similar  levels  of  observer 
effort  as  measured  by  survey  duration.  All 
BBCs  took  16.7—20.1  h to  complete  versus  an 
average  of  17  field-h  per  atlas  block  (Brauning 
1992). 

The  BBC  method’s  reliance  on  three  or 
more  records  to  score  territorial  occupancy, 
however,  is  more  restrictive  in  tallying  species 
occurrences  than  the  BBA  method.  The  latter 
includes  “possible,”  “probable,”  and  “con- 
firmed” categories  of  breeders,  and  is  there- 
fore likely  to  include  more  species  per  unit 
effort.  Greater  numbers  of  species  may  also 
be  detected  with  the  BBA  method  because  of 
the  substantially  larger  areas  covered  (poten- 
tially hundreds  or  thousands  of  ha  per  block 
versus  the  tens  of  ha  in  most  BBCs). 

As  a consequence  of  differences  in  the 
scope  of  effort  between  BBC  and  BBA  meth- 
ods, comparisons  of  incidence  values  (I)  for 
individual  species  (Table  1)  may  be  biased 
against  detecting  greater  occurrence  (and  to- 


Haiie\  • OLD-GROWTH  BIRD  COMMUNITY 


97 


wards  detecting  lower  occurrence)  in  old- 
growth  habitat  than  in  the  landscape  at  large. 
Findings  of  positive  old-growth  association  by 
individual  species  (Table  2)  are  more  robust 
as  a result.  Negative  and  neutral  associations 
with  oldgrowth  should  be  interpreted  cau- 
tiously because  more  liberal  listing  of  species 
under  BBA  methodology  could  elevate  rela- 
tive incidence  values  at  state  and  province  lev- 
els, thereby  leading  to  false  conclusions  that 
no  differences  in  species  occurrences  existed 
across  spatial  scales  (Type  II  error). 

Hierarchical  comparisons. — Although 
scale  is  viewed  as  essential  for  interpreting 
distributional  data  in  birds  (Lacy  and  Bock 
1986),  logistical  constraints  and  methodolog- 
ical inconsistencies  often  prevent  hierarchical 
or  multi-tiered  approaches.  Comparing  local 
density  of  bird  populations  to  density  in  a re- 
gion as  large  as  an  entire  state  is  impossible 
because  the  BBC  method  requires  large  in- 
vestments in  time  for  limited  spatial  coverage. 
Proportions  are  easy  to  derive  from  virtually 
any  kind  of  sample,  however,  and  a metric 
based  on  incidence  enabled  direct  comparison 
of  species  occurrences  in  BBC  plots  and  atlas 
blocks  (Table  1). 

Comprehensive  coverage  in  Pennsylvania’s 
atlas  program  also  enabled  more  reliable  com- 
parisons of  birds  in  oldgrowth  to  the  wider 
landscape:  all  blocks,  including  those  on  the 
state’s  borders,  were  censused  (Brauning 
1992).  Synoptic  coverage  allows  evaluation  of 
the  likely  impacts  of  potential  actions  on 
groups  of  bird  species  within  a wider  context. 
It  would  be  easy  to  scale  down  from  the  eco- 
physiographic  province  or  state  levels  used  in 
this  study  to  some  smaller  region  of  interest 
(e.g.,  county,  national  forest,  watershed).  Al- 
ternatively, BBA  data  from  adjacent  states 
could  be  aggregated  to  examine  individual 
species  occurrences  across  even  larger  land- 
scape units.  This  spatial  flexibility  should  en- 
able better  evaluation  of  potential  consequenc- 
es of  local  management  prescriptions  on  the 
regional  distributions  of  birds. 

I used  an  incidence  metric  as  a reasonable 
proxy  for  population  size  (Fig.  2).  Several  re- 
searchers have  documented  a general  relation- 
ship between  abundance  and  range  size  in 
birds  (Bock  and  Ricklefs  1983,  Lacy  and 
Bock  1986,  Mauer  and  Hey  wood  1993).  This 
relationship  may  not  indicate  the  existence  of 


a particular  ecological  hypothesis  (Wright 
1991).  Nevertheless,  the  generality  that  spe- 
cies with  sparse  distributions  also  have  low 
population  densities  was  confirmed  in  this 
study  by  documenting  regional  scarcity  in 
several  bird  species  that  use  a rare  and  very 
local  habitat  type. 

ACKNOWLEDGMENTS 

I thank  B.  Allison,  J.  Cheek,  L.  Hepfner,  R.  Kauf- 
mann,  J.  Lydic,  C.  Schaadt,  J.  Seachrist,  J.  Smreker,  S. 
Weilgosz,  S.  Wetzel,  and  R.  Williams  for  their  help  in 
conducting  the  Breeding  Bird  Censuses  and  vegetation 
surveys;  D.  DeCalesta,  C.  Nowack,  J.  Palmer,  S.  Stout, 
C.  Schlentner,  L.  Lentz,  J.  Sowl,  and  D.  Wright  for 
facilitating  logistic  anangements;  and  Cook  Forest 
State  Park  and  the  USD  A Forest  Service  for  research 
access.  Financial  support  was  provided  by  the  Wilder- 
ness Society,  Dodge  Foundation,  Johnson  and  John- 
son, Sweet  Water  Trust,  William  P.  Wharton  Trust, 
Pennsylvania  Wild  Resource  Conservation  Fund, 
Pennsylvania  Game  Commission,  Center  for  Rural 
Pennsylvania,  DuBois  Educational  Foundation  Fund 
for  Academic  Excellence,  and  Pennsylvania  State  Uni- 
versity Research  and  Development  funds. 

LITERATURE  CITED 

Abrams,  M.  D.  and  D.  A.  Orwig.  1996.  A 300-year 
history  of  disturbance  and  canopy  recruitment  for 
co-occurring  white  pine  and  hemlock  on  the  Al- 
legheny Plateau,  USA.  J.  Ecol.  84:353—363. 
Abrams,  M.  D.  and  C.  M.  Ruffner.  1995.  Physio- 
graphic analysis  of  witness-tree  distribution 
(1765-1798)  and  present  forest  cover  through 
north  central  Pennsylvania.  Can.  J.  For.  Res.  25: 
659-668. 

Alerich,  C.  L.  1993.  Forest  statistics  for  Pennsylva- 
nia-1978 and  1989.  U.S.  Dept.  Agric.  For.  Serv. 
Res.  Bull.  NE- 126: 1-244. 

Anthony,  R.  G.,  G.  A.  Green,  E.  D.  Forsman,  and 
S.  K.  Nelson.  1996.  Avian  abundance  in  riparian 
zones  of  three  forest  types  in  the  Cascade  Moun- 
tains, Oregon.  Wilson  Bull.  108:280—291. 
Bjorkblom,  j.  C.  and  R.  G.  Larson.  1977.  The  Tio- 
nesta  scenic  and  research  natural  areas.  USDA 
For.  Serv.  Gen.  Tech.  Rept.  NE-31:1— 24. 

Bock,  C.  E.  and  R.  E.  Ricklefs.  1983.  Range  size  and 
local  abundance  of  some  North  American  song- 
birds: a positive  coirelation.  Am.  Nat.  122:295- 
299. 

Brauning,  D.  W,  (Ed.).  1992.  Atlas  of  breeding  birds 
in  Pennsylvania.  Univ.  of  Pittsburgh  Press.  Pitts- 
burgh. Pennsylvania. 

Bunge,  J.  and  M.  Fitzpatrick.  1993.  Estimating  the 
number  of  species:  a review.  J.  Am.  Stat.  Assoc. 
88:364-373. 

Burke,  D.  M.  and  E.  Nol.  1998.  Influence  of  food 
abundance,  nest-site  habitat,  and  forest  fragmen- 
tation on  breeding  Ovenbirds.  Auk  115:96-104. 


98 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  I.  March  1999 


Coleman,  M.  D.,  M.  D.  Mares,  M.  R.  Willig,  and 
Y.-H.  Hsieh.  1982.  Randomness,  area,  and  species 
richness.  Ecology  63:1121-1133. 

Colwell.  R.  K.  and  J.  A.  Coddington.  1994.  Esti- 
mating terrestrial  biodiversity  through  extrapola- 
tion. Phil.  Trans.  Royal  Soc.  (B)  345:101-118. 

Conroy,  M.  J.  and  B.  R.  Noon.  1996.  Mapping  of 
species  richness  for  conservation  of  biological  di- 
versity: conceptual  and  methodological  issues. 
Ecol.  Appl.  6:763-773. 

Davis,  M.  B.  1996.  Extent  and  location.  Pp.  18-32  in 
Eastern  old-growth  forests:  prospects  for  rediscov- 
ery and  recovery  (M.  B.  Davis,  Ed.).  Island  Press, 
Washington,  D.C. 

DeGraaf,  R.  M.,  M.  Yamasaki.  W.  B.  Leak,  and  J. 
W.  L.anier.  1992.  New  England  wildlife:  manage- 
ment of  forested  habitats.  US  DA  Eor.  Serv., 
Northeast  For.  Exp.  Station,  Gen.  Tech.  Rept.  NE- 
144:92. 

Dellasala,  D.  a.,  j.  C.  H.agar,  K.  A.  Engel,  W.  C. 
McComb.  R.  L.  Fairbanks,  and  E.  G.  Campbell. 
1996.  Effects  of  silvicultural  modifications  of  tem- 
perate rainforest  on  breeding  and  wintering  bird 
communities.  Prince  of  Wales  Island,  southeast 
Alaska.  Condor  98:706-721. 

Elphick,  C.  S.  1997.  Correcting  avian  richness  esti- 
mates for  unequal  sample  effort  in  atlas  studies. 
Ibis  139:189-190. 

Engstrom,  R.  T.  and  E C.  James.  1984.  An  evaluation 
of  methods  used  in  the  Breeding  Bird  Census. 
Am.  Birds  38:19-23. 

Eyre,  E H.  (Ed.).  1980.  Forest  cover  types  of  the  Unit- 
ed States  and  Canada.  Society  of  American  For- 
esters, Washington,  D.C. 

Feather,  C.  H.  1996.  Fitting  species-accumulation 
functions  and  asse.ssing  regional  land  use  impacts 
on  avian  diversity.  J.  Biogeogr.  23:155-168. 

Hall,  G.  A.  1964.  Breeding  bird  censuse.s — why  and 
how.  Audubon  Field-Notes  18:413-416. 

Haney,  J.  C.  1997.  Spatial  incidence  of  Baired  Owl 
iStrix  varia)  reproduction  in  old-growth  forest  of 
the  Appalachian  Plateau.  J.  Raptor  Res.  31:241- 
252. 

Haney,  J.  C and  C.  P.  Schaadt.  1996.  Functional  roles 
of  eastern  old  growth  in  promoting  forest  bird  di- 
versity. Pp.  76-88  in  Eastern  old-growth  forests: 
prospects  for  rediscovery  and  recovery  (M.  B.  Da- 
vis, Ed.).  Island  Press,  Washington,  D.C. 

Hansen,  A.  J..  W.  C.  McComb,  R.  Vega,  M.  G.  Ra- 
phael, AND  M.  Hunter.  1995.  Bird  habitat  rela- 
tionships in  natural  and  managed  forests  in  the 
west  Cascades  of  Oregon.  Ecol.  Appl.  5:555-569. 

Hender.son,  P.  a.  and  R.  M.  H.  Seahy.  1997.  Species 
diversity  and  richness,  version  1.2.  PISCES  Con- 
servation Ltd.,  Lymington,  U.K. 

Hoover,  J.  P.  and  M.  C.  Brittingham.  1993.  Regional 
variation  in  cowbird  parasitism  of  Wood  Thrush- 
es. Wifson  Bull.  105:228-238. 

Hough,  A.  E and  R.  D.  Forbes.  1943.  The  ecology 
and  silvics  of  Pennsylvania  high-plateau  forests. 
Ecol.  Monogr.  13:299-320. 


Hunter,  J.  E..  R.  J.  Gutierrez,  and  A.  B.  Franklin. 
1995.  Habitat  configuration  around  Spotted  Owl 
sites  in  northwestern  California.  Condor  97:684- 
693. 

Keys,  J.  E.,  Jr.,  W.  H.  McNab,  and  C.  A.  Carpenter 
(Eds.).  1995.  Ecological  units  of  the  eastern  Unit- 
ed States — first  approximation.  USDA,  Forest 
Service.  Atlanta,  Georgia. 

Kotliar,  N.  B.  and  j.  a.  Wiens.  1990.  Multiple  scales 
of  patchiness  and  patch  structure:  a hierarchical 
framework  for  the  study  of  heterogeneity.  Oikos 
59:253-260. 

Lac^  . R.  C.  AND  C.  E.  Bock.  1986.  The  correlation 
between  range  size  and  local  abundance  of  some 
North  American  birds.  Ecology  67:258-260. 

Lalighlin,  S.  B.  (Ed.).  1982.  Proceedings  of  the  north- 
eastern breeding  bird  atlas  conference.  Vermont 
Institute  of  Natural  Sciences,  Wood.stock,  Ver- 
mont. 

Leverett,  R.  1996.  Definitions  and  history.  Pp.  3—17 
in  Eastern  old-growth  forests:  prospects  for  redis- 
covery and  recovery  (M.  B.  Davis,  Ed.).  Island 
Press,  Washington.  D.C. 

Lowe,  J.  D.  1993.  Resident  bird  counts.  1992.  J.  Field 
Ornithol.  64( Supplement ):3-5. 

Martin,  T.  E.  1988.  Habitat  and  area  effects  on  forest 
bird  assemblages:  is  nest  predation  an  influence? 
Ecology  69:74-84. 

Maurer,  B.  A.  and  S.  G.  Heywood.  1993.  Geographic 
range  fragmentation  and  abundance  in  Neotropical 
migratory  birds.  Conserv.  Biol.  7:501-509. 

Moorman,  C.  E.  and  B.  R.  Chapman.  1996.  Nest-site 
selection  of  Red-shouldered  Hawks  in  managed 
forest.  Wilson  Bull.  108:357-368. 

Nichols.  G.  E.  1935.  The  hemlock-white  pine-north- 
ern hardwood  region  of  eastern  North  America. 
Ecology  16:403-422. 

Raaijmakers.  j.  G.  W.  1987.  Statistical  analysis  of  the 
Michaelis-Menten  equation.  Biometrics  40:119- 
129. 

Remsen,  j.  V.,  Jr.  1994.  Use  and  misuse  of  bird  lists 
in  community  ecology  and  conservation.  Auk 
1 I 1 :225-227. 

Rodenhouse,  N.  L.,  L.  B.  Best,  R.  J.  O'Connor,  and 
E.  K.  Bollinger.  1995.  Effects  of  agricultural 
practices  and  farmland  structures.  Pp.  269—293  in 
Ecology  and  management  of  Neotropical  migra- 
tory birds  (T.  E.  Martin  and  D.  M.  Finch.  Eds.). 
Oxford  Univ.  Press,  Oxford.  U.K. 

Schieck,  j.  1997.  Biased  detection  of  bird  vocaliza- 
tions affects  comparisons  of  bird  abundance 
among  forested  habitats.  Condor  99:179-190. 

.Schieck.  J.,  K.  Lertzman,  B.  Nyberg,  a'nd  R.  Page. 
1995.  Effects  of  patch  size  on  birds  in  old-growth 
montane  forests.  Conserv.  Biol.  9:1072-1084. 

Seitz,  L.  C.  and  D.  A.  Zegers.  1993.  An  experimenlal 
study  of  nest  predation  in  adjacent  deciduous,  co- 
niferous and  successional  habitats.  Condor  95: 
297-304. 

Shackelford.  C.  E.  and  R.  N.  Conner.  1998.  Wood- 


Hanex  • OLD-GROWTH  BIRD  COMMUNITY 


99 


pecker  abundance  and  habitat  use  in  three  forest 
types  in  eastern  Texas.  Wilson  Bull.  109:614-629. 

Sherry,  T.  W.  and  R.  T.  Holmes.  1988.  Habitat  selec- 
tion by  breeding  American  Redstarts  in  response 
to  a dominant  competitor,  the  Least  Flycatcher. 
Auk  105:350-364. 

Snedecor,  G.  W.  and  W.  G.  Cochran.  1980.  Statistical 
methods,  seventh  ed.  Iowa  State  Univ.  Press, 
Ames. 

St.ahle,  D.  W.  1996.  Tree  rings  and  ancient  forest  his- 
tory. Pp.  321-343  in  Eastern  old-growth  forests: 
prospects  for  rediscovery  and  recovery  (M.  B.  Da- 
vis, Ed.).  Island  Press,  Washington,  D.C. 

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- 
579. 

Titus,  K.  and  J.  A.  Mosher.  1981.  Nest-site  habitat 
selected  by  woodland  hawks  in  the  Central  Ap- 
palachians. Auk  98:270-281. 

T'i  rrell,  L.  E.  and  T.  R.  Crow.  1994.  Structural  char- 


acteristics of  old-growth  hemlock-hardwood  for- 
ests in  relation  to  stand  age.  Ecology  75:370-386. 

USDA.  1997.  Guidance  for  conserving  and  re.storing 
old-growth  forest  communities  on  national  forests 
in  the  Southern  Region.  U.S.  Department  of  Ag- 
riculture, Forest  Service,  Southern  Region,  Atlan- 
ta, GA.  Forestry  Rept.  R8-FR  62:35-38. 

Whitney,  G.  G.  1990.  The  history  and  status  of  hem- 
lock-hardwood forests  of  the  Allegheny  Plateau. 
J.  Ecol.  78:443-458. 

Whittaker,  R.  H.  1977.  Evolution  of  species  diversity 
in  land  communities.  Evol.  Biol.  10:1-67. 

Wiens,  J.  A.  1989.  The  ecology  of  bird  communities: 
foundations  and  patterns.  Cambridge  Univ.  Press, 
Cambridge,  U.K. 

Wright,  D.  H.  1991.  Correlations  between  incidence 
and  abundance  are  expected  by  chance.  J.  Bio- 
geogr.  18:463-466. 

Yahner,  R.  H.  1993.  Effects  of  long-term  forest  clear- 
cutting  on  wintering  and  breeding  birds.  Wilson 
Bull.  105:239-255. 


Wilson  Bull..  111(1),  1999,  pp.  100-104 


EFFECTS  OF  WIND  TURBINES  ON  UPLAND  NESTING  BIRDS  IN 
CONSERVATION  RESERVE  PROGRAM  GRASSLANDS 

KRECIA  L.  LEDDY,'  3 KENNETH  E HIGGINS,^^  and  DAVID  E.  NAUGLE'^ 


ABSTRACT. — Grassland  passerines  were  surveyed  during  summer  1995  on  the  Buffalo  Ridge  Wind  Resource 
Area  in  southwestern  Minnesota  to  determine  the  relative  influence  of  wind  turbines  on  overall  densities  of 
upland  nesting  birds  in  Conservation  Reserve  Program  (CRP)  grasslands.  Birds  were  surveyed  along  40  m fixed 
width  transects  that  were  placed  along  wind  turbine  strings  within  three  CRP  fields  and  in  three  CRP  fields 
without  turbines.  Conservation  Reserve  Program  grasslands  without  turbines  and  areas  located  180  m from 
turbines  supported  higher  densities  (261.0—312.5  males/ 100  ha)  of  grassland  birds  than  areas  within  80  m of 
turbines  (58.2-128.0  males/100  ha).  Human  disturbance,  turbine  noise,  and  physical  movements  of  turbines 
during  operation  may  have  distrurbed  nesting  birds.  We  recommend  that  wind  turbines  be  placed  within  cropland 
habitats  that  support  lower  densities  of  grassland  passerines  than  those  found  in  CRP  grasslands.  Received  9 
Sept.  1997,  accepted  5 Oct.  1998. 


Technological  advances  that  have  reduced 
the  cost  of  electricity  generated  from  wind- 
plants  have  enabled  the  wind-power  industry 
to  expand  from  California  into  the  eastern 
United  States  and  Canada  (Nelson  and  Curry 
1995).  Wind  power  has  received  strong  public 
support  as  an  alternative  energy  source  despite 
the  potential  threats  that  the  presence  of  wind 
turbines  may  pose  to  avian  species.  Recent  re- 
search has  indicated  that  raptor  mortality  from 
collisions  with  wind  turbines  varies  greatly 
from  no  mortality  (Higgins  et  al.  1996;  Us- 
gaard  et  al.,  in  press)  to  substantial  mortality 
(Orloff  and  Flannery  1992).  In  addition  to  di- 
rect mortality  from  collisions,  research  also 
has  indicated  that  waterfowl,  wading  bird,  and 
raptor  densities  near  turbines  were  lower  com- 
pared to  densities  in  similar  habitats  away 
from  turbines  (Winkelman  1990;  Pedersen 
and  Poulsen  1991;  Usgaard  et  al.,  in  press). 
The  influence  of  wind  turbines  on  grassland 
nesting  passerine  species  has  not  been  previ- 
ously measured. 

Recent  construction  of  the  first  windplant 
facility  in  the  midwestern  United  States  pro- 


' Dept,  of  Wildlife  and  Fisheries  Sciences,  Box 
2I40B,  South  Dakota  State  Univ.,  Brookings,  SD 
57007. 

^ South  Dakota  Cooperative  Fish  and  Wildlife  Re- 
.search  Unit,  USGS-BRD,  South  Dakota  State  Univ., 
Box  2140B,  Brookings,  SD  57007. 

’ Pre.sent  address;  Natural  Re.sources  Conservation 
Service,  RR  1 Box  740,  Webster,  SD  57274. 

Present  address:  College  of  Natural  Resources, 
Univ.  of  Wisconsin-Stevens  Point,  Stevens  Point,  WI 
5448 1 . 

' Corresponding  author. 


vided  a unique  opportunity  to  study  the  effects 
of  wind  turbines  on  grassland  nesting  passer- 
ines. Several  midwestern  grassland  passerine 
species  have  declined  in  abundance  (Johnson 
and  Schwartz  1993)  in  response  to  agricultural 
tillage,  grazing,  and  invasive  woody  species 
that  have  destroyed  or  degraded  most  of  the 
remaining  grasslands  (Kantrud  1981,  Castrale 
1985).  Although  Conservation  Reserve  Pro- 
gram (CRP;  Young  and  Osborn  1990)  grass- 
lands provide  habitat  for  grassland  nesting 
birds  (Johnson  and  Schwartz  1993,  Igl  and 
Johnson  1995,  Johnson  and  Igl  1995,  King 
and  Savidge  1995,  Millenbah  et  al.  1996),  the 
potential  impact  of  wind  turbines  in  CRP 
fields  could  negate  those  benefits.  The  objec- 
tive of  this  study  was  to  determine  whether 
density  of  upland  nesting  passerines  in  CRP 
grasslands  was  influenced  by  the  presence  of 
wind  turbines.  We  hypothesized  that  bird  den- 
sity in  CRP  grasslands  would  not  differ  in  re- 
lation to  distance  from  wind  turbines. 

STUDY  AREA  AND  METHODS 

Study  area. — The  Buffalo  Ridge  Wind  Resource 
Area  (WRA)  in  southwestern  Minnesota  is  located 
along  a 100  km  segment  of  the  Bemis  Moraine  near 
Lake  Benton,  Minnesota.  Elevation  is  546-610  m. 
Wind  turbines  cover  32  km^  of  the  293  km^  Buffalo 
Ridge  WRA.  Additional  lands  within,  the  Buffalo 
Ridge  WRA  have  been  leased  as  future  wind-turbine 
development  sites.  The  windplant  contains  73  opera- 
tional wind  turbines  that  are  arranged  in  10  turbine 
strings,  with  3-20  turbine.s/string.  Turbines  are  91-183 
m apart  within  strings.  Turbines  (model  KVS-33;  KE- 
NETECH  Windpower,  Inc.),  which  operate  at  wind 
speeds  of  14-104  km/h,  consist  of  a 33  m diameter 
rotor  mounted  on  a 37  m tubular  tower. 


100 


Lcdd\  et  cd  • EFFECTS  OF  WIND  TURBINES  ON  BIRDS 


101 


Upland  grassland  bird  nesting  habitat  within  the 
Buffalo  Ridge  WRA  consisted  primarily  of  CRP  grass- 
lands, mostly  planted  with  a mixture  of  smooth  brome 
(Bronius  inermis)  and  alfalfa  (Medica^’o  saliva)  or 
switchgrass  (Panicmn  vir^atum).  Habitats  surrounding 
CRP  grasslands  were  agricultural  lands  dominated  by 
corn  {Zea  mays)  and  soybeans  {Glycine  max)  with 
smaller  areas  of  haylands,  pasturelands,  and  scattered 
woodlands  near  farmsteads  and  in  ravines. 

Methods. — Bird  survey  transects  were  placed  along 
wind-turbine  strings  within  three  CRP  fields  and  in 
three  CRP  fields  without  turbines  (i.e.,  control;  Feddy 
1996).  We  selected  CRP  fields  that  were  7-8  years  of 
age  to  minimize  effects  of  field  age  on  diversity  and 
density  of  avain  species  (Millenbah  et  al.  1996).  Visual 
obstruction  readings  (Robel  et  al.  1970,  Higgins  and 
Barker  1982)  did  not  differ  between  CRP  grasslands 
with  and  without  turbines,  indicating  that  vegetation 
structure  in  experimental  and  control  fields  was  similar 
(Feddy  1996).  Transects  were  surveyed  weekly  in  ran- 
dom order  from  15  May  to  1 July  1995.  Multiple  sur- 
veys of  a single  transect  were  averaged  into  one  bird 
density  to  avoid  pseudoreplication  (Hurlbert  1984).  Six 
40-m  fixed  width  transects  (Wakeley  1987)  paralleling 
each  turbine  string  were  used  per  field.  One  transect 
ran  directly  underneath  turbine  strings.  Two  additional 
transects  on  each  side  of  the  turbine  string  paralleled 
the  string  at  distances  of  40  and  80  m;  the  sixth  tran- 
sect was  placed  180  m from  the  turbine  string.  Tran- 
sects varied  in  length  according  to  field  size  and  were 
placed  at  least  30  m from  field  borders  and  wetlands 
to  minimize  bias  associated  with  edges  (Arnold  and 
Higgins  1986.  Reese  and  Ratti  1988).  One  transect  was 
established  at  a random  location  in  each  of  the  three 
control  CRP  fields  without  turbines. 

Inconsistencies  among  surveys  attributable  to  peri- 
odic bird  inactivity  (Skirvin  1981,  Verner  and  Ritter 
1986)  were  minimized  by  conducting  surveys  from 
sunrise  to  10:00  CST  We  recorded  all  birds  seen  or 
heard  while  walking  transects  at  1.0- 1.5  km/h  (Mikol 
1980,  Wakeley  1987);  only  perched  and/or  singing 
males  were  used  in  statistical  analyses.  Flushed  birds 
seen  leaving  transects  were  counted  (Burnham  et  al. 
1980),  whereas  birds  seen  entering  transects  or  flying 
overhead  were  not  counted.  Surveys  were  not  con- 
ducted during  heavy  rain  or  high  winds  (^20  km/h; 
Ralph  et  al.  1993).  Birds  were  surveyed  in  CRP  fields 
with  turbines  when  turbines  were  operational  and  non- 
operational  because  turbines  began  operating  during 
surveys  when  wind  speeds  reached  14—20  km/h.  We 
compared  surveys  that  were  conducted  during  opera- 
tional and  non-operational  periods  to  determine  wheth- 
er noise  produced  during  turbine  operation  biased  sur- 
veys. 

An  index  of  total  breeding  bird  density  was  calcu- 
lated by  dividing  the  number  of  perched  and/or  singing 
males  by  transect  area.  Percent  species  composition 
was  calculated  by  dividing  the  number  of  perched  and/ 
or  singing  males  of  a particular  species  by  the  total 
number  of  males.  Species  richness  was  defined  as  the 
number  of  species  (Koford  et  al.  1994). 


Analysis  of  covariance  (SAS  1989)  was  used  to  de- 
termine whether  bird  density  across  transects  was  re- 
lated to  noise  produced  during  wind  turbine  operation. 
We  used  turbine  operational  status  (i.e.,  running  versus 
idle)  to  determine  whether  the  slope  of  bird  densities 
differed.  Analysis  of  variance  (ANOVA;  SAS  1989) 
was  used  to  determine  whether  bird  density  in  CRP 
grasslands  without  turbines  differed  from  that  in  CRP 
grasslands  containing  turbines.  An  ANOVA  also  was 
used  to  determine  whether  bird  density  was  related  to 
distance  from  wind  turbines.  A Least  Significant  Dif- 
ference Multiple  Comparisons  test  was  used  to  deter- 
mine where  differences  in  bird  density  occurred  among 
transects. 

RESULTS 

Ten  upland  grassland  bird  species  occurred 
in  CRP  grasslands  with  and  without  turbines 
(Table  1).  Bobolinks  (Dolichonyx  oryzivorus). 
Red-winged  Blackbirds  (Agelaius  phoeni- 
ceus),  and  Savannah  Sparrows  (Passerculus 
sandwichensis)  comprised  74.5%  of  the  birds 
in  CRP  grasslands  with  turbines  (Table  1). 
Bobolinks,  Sedge  Wrens  (Cistothorus  platen- 
sis),  and  Savannah  Sparrows  comprised 
80.0%  of  the  individuals  in  CRP  fields  with- 
out turbines  (Table  1). 

Mean  bird  densities  from  surveys  conduct- 
ed while  wind  turbines  were  operational  {x  = 
4.7  ± 0.88  SE)  and  non-operational  {x  = 5.4 
± 0.94  SE)  were  pooled  because  slopes  of 
bird  densities  among  transects  did  not  differ 
(F  = 0.39,  1,30  df,  P > 0.05).  Total  bird  den- 
sity was  lower  in  CRP  grasslands  containing 
turbines  than  in  CRP  grasslands  without  tur- 
bines (F  = 17.36,  6,14  df,  P = 0.001;  Table 
2).  Bird  density  was  lower  (F  = 12.37,  1,10 
df,  P = 0.006)  in  the  0 and  40  m transects 
compared  to  density  in  transects  80  m or  more 
from  turbines  (Table  2).  Bird  density  also  was 
lower  (F=  13.10,  l,10df,  F = 0.001)  in  tran- 
sects within  80  m of  the  turbines  compared  to 
180  m from  turbines  (Table  2).  Bird  density 
180  m from  turbines  did  not  differ  (F  = 0.10. 
1,10  df,  P > 0.05)  from  that  in  CRP  grass- 
lands without  turbines  (Table  2).  A linear  re- 
lationship existed  (r-  = 0.746,  n = 18,  F < 
0.001)  between  bird  density  and  transect  dis- 
tance from  turbines  (Fig.  1). 

DISCUSSION 

Conservation  Reserve  Program  grasslands 
without  turbines  and  areas  located  180  m from 
turbines  supported  mean  densities  of  grassland 
birds  that  were  four  times  higher  than  those 


102 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  1,  March  1999 


TABLE  1.  Number  (/?)  and  percent  (%)  composition  of  breeding 

grassland  birds 

in  Conservation  Reserve 

Program  grasslands  with  and  without  turbines  at  the  Buffalo  Ridge  Wind  Resource  Area,  Minnesota,  May— July 

1995. 

Turbines 

No  turbines 

Species 

n 

% 

n 

% 

Bobolink  {Dolichonyx  oryzivorus) 

139 

36.6 

48 

32.0 

Red-winged  Blackbird  (Agelaius  phoeniceus) 

85 

22.4 

7 

4.7 

Savannah  Sparrow  (Passerculus  sandwichensis) 

59 

15.5 

33 

22.0 

Common  Yellowthroat  (Geothlypis  trichas) 

36 

9.5 

7 

4.7 

Dickcissel  (Spiza  americana) 

22 

5.8 

2 

1.3 

Le  Conte’s  Sparrow  (Ammodramus  leconteii) 

10 

2.6 

Brown-headed  Cowbird  (Molothriis  ater) 

9 

2.4 

6 

4.0 

Western  Meadowlark  (Strunella  neglecta) 

5 

1.3 

1 

0.7 

Grasshopper  Sparrow  (Ammodramus  savannarum) 

4 

1.1 

4 

2.7 

Sedge  Wren  (Cistothorus  platensis) 

1 

0.3 

39 

26.0 

Clay-colored  Sparrow  (Spizella  pallida) 

1 

0.7 

Unknown 

9 

2.3 

2 

1.3 

Total  species 

10 

10 

in  grasslands  nearer  to  turbines.  Three  of  four 
species  that  composed  at  least  74.5%  of  the 
bird  community  composition  (Bobolink,  Sa- 
vannah Sparrow,  Sedge  Wrens)  in  CRP  fields 
with  and  without  turbines  are  area-sensitive 
species  (Herkert  1994a,  b;  Swanson  1996)  that 
require  large  tracts  of  tall,  dense  vegetation  for 
nesting  (Wiens  1969,  Herkert  1994a).  Minor 
differences  in  overall  bird  species  richness 
and  composition  were  likely  related  to  subtle 
structural  differences  in  grassland  stand  types. 
Leddy  and  coworkers  (in  press)  found  that 
Clay-colored  Sparrows  (Spizella  pallida)  and 
Sedge  Wrens  using  CRP  grasslands  on  the 
Buffalo  Ridge  WRA  preferred  dense  stands  of 
switchgrass  while  Dickcissels  (Spiza  ameri- 
cana)  and  Bobolinks  usually  used  stands  of 
smooth  brome  and  alfalfa. 


TABLE  2.  Species  richness  and  mean  density  of 
upland  grassland  birds/100  ha  at  varying  distances 
from  wind  turbines  in  Conservation  Reserve  Program 
grasslands  at  the  Buffalo  Ridge  Wind  Resource  Area, 
Minnesota,  May-July  1995. 


Breeding  males 


Transect 

n 

Species 

richness 

Mean 

density^ 

SK 

0 m 

3 

6 

58.2  A 

26.3 

40  m 

6 

8 

66.0  A 

17.1 

80  m 

6 

7 

128.0  B 

19.6 

180  m 

3 

9 

261.0  C 

12.0 

CRP  Control 

3 

10 

312.5  C 

15.7 

Little  evidence  has  been  found  linking  avi- 
an mortality  to  collisions  with  wind  turbines 
on  the  Buffalo  Ridge  WRA  (Higgins  et  al. 
1996).  Although  wind  turbines  may  not  di- 
rectly cause  mortality,  the  presence  of  wind 
turbines  may  indirectly  affect  local  grassland 
bird  populations  by  decreasing  the  area  of 
grassland  habitat  available  to  breeding  birds. 
Comparison  of  bird  density  and  species  rich- 
ness among  transects  indicated  that  bird  use 
of  grasslands  1 80  m from  turbines  was  similar 
to  that  in  CRP  fields  without  turbines  (Table 
2).  Although  research  in  the  Netherlands  also 


FIG.  1.  Linear  relationship  (Density  = 32.30  + 
1.22  X Distance;  H = 0.746)  between  breeding  bird 
density  (males/100  ha)  and  distance  (0-180  m)  from 
wind  turbines  in  Conservation  Reserve  Program  grass- 
lands at  the  Buffalo  Ridge  Wind  Resource  Area  in 
southwestern  Minnesota,  May-July  1995. 


•'  Means  deni)ted  by  the  same  letter  do  not  differ  (F  £ O.O.S). 


Leckh  et  al  • EFFECTS  OF  WIND  TURBINES  ON  BIRDS 


103 


has  indicated  that  the  presence  of  turbines  has 
prevented  waterfowl  and  wading  bird  species 
from  using  otherwise  suitable  habitat  (Win- 
kelman  1990,  Pedersen  and  Poulsen  1991), 
mechanisms  inhibiting  birds  from  exploiting 
grasslands  near  turbines  have  not  yet  been 
identified.  In  addition  to  human  disturbance 
and  noise,  the  physical  movements  of  the  tur- 
bines when  they  are  operating  may  have  dis- 
turbed nesting  birds.  Maintenance  trails  be- 
tween turbines  that  are  driven  daily  may  have 
further  decreased  the  availiability  of  grassland 
habitat  adjacent  to  turbines. 

Construction  of  windplants  within  mid- 
western  grassland  habitats  may  soon  become 
an  additional  source  of  habitat  degradation  as 
demands  for  wind  generated  power  increase. 
Current  grazing  and  tillage  practices  on  many 
privately  owned  lands  that  are  less  conducive 
to  grassland  bird  production  increase  the  im- 
portance of  remaining  grasslands  to  prairie 
nesting  birds  (Johnson  and  Schwartz  1993; 
Johnson  and  Igl  1995;  Leddy  et  al.,  in  press). 
Until  additional  research  is  conducted,  we  rec- 
ommend that  wind  turbines  be  placed  within 
cropland  habitats  that  support  lower  densities 
of  grassland  passerines  than  those  found  in 
CRP  grasslands  (Leddy  et  al.,  in  press).  We 
also  recommend  that  additional  research  be 
conducted  in  other  geographic  regions  where 
wind  generated  power  is  currently  used  to  fur- 
ther assess  possible  effects  of  wind  turbines 
on  avian  habitats. 

ACKNOWLEDGMENTS 

We  thank  L.  D.  Flake,  D.  H.  Johnson,  and  anony- 
mous referees  for  reviews  of  our  manuscript.  We  also 
thank  R D.  Evenson  for  assisting  with  statistical  anal- 
yses and  V.  J.  Swier  and  R.  G.  Osborn  for  conducting 
field  work.  Project  funding  was  provided  by  Kenetech 
Windpower,  Inc.,  and  the  South  Dakota  Cooperative 
Fish  and  Wildlife  Research  Unit,  in  cooperation  with 
South  Dakota  Department  of  Game,  Fish  and  Parks, 
Natural  Resources  Conservation  Service,  Wildlife 
Management  Institute  and  South  Dakota  State  Univer- 
sity. 

Mention  of  trade  names  does  not  constitute  any  en- 
dorsement, guarantee,  or  warranty  of  any  trademark 
proprietary  product  by  the  authors.  South  Dakota  State 
University,  the  Department  of  Wildlife  and  Fisheries 
Sciences,  or  the  South  Dakota  Cooperative  Fish  and 
Wildlife  Research  Unit  (United  States  Geological  Sur- 
vey, Biological  Resources  Division). 


LITERATURE  CITED 

Arnold,  T.  W.  and  K.  F.  Higgins.  1986.  Effects  of 
shrub  coverages  on  birds  of  North  Dakota  mixed- 
grass  prairie.  Can.  Field-Nat.  100:10-14. 

Burnham,  K.  R,  D.  R.  Anderson,  and  J.  L.  Laake. 
1980.  Estimation  of  density  from  line  transect 
sampling  of  biological  populations.  Wildl.  Mon- 
ogr.  72:1-202. 

Castrale,  j.  S.  1985.  Responses  of  wildlife  to  various 
tillage  conditions.  Trans.  N.  Am.  Wildl.  Nat.  Re- 
sour. Conf.  50:142-156. 

Herkert,  j.  R.  1994a.  Breeding  bird  communities  of 
midwestern  prairie  fragments:  the  effects  of  pre- 
scribed burning  and  habitat-area.  Nat.  Areas  J.  14: 
128-135. 

Herkert,  J.  R.  1994b.  The  effects  of  habitat  fragmen- 
tation on  midwestern  grassland  bird  communities. 
Ecol.  Appl.  4:461-471. 

Higgins,  K.  F and  W.  T.  Barker.  1982.  Changes  in 
vegetation  structure  in  seeded  nesting  cover  in  the 
Prairie  Pothole  Region.  U.S.  Fish  Wildl.  Serv. 
Spec.  Sci.  Rep.  242:1-26. 

Higgins,  K.  F,  R.  E.  Usgaard,  and  C.  D.  Dieter. 
1996.  Monitoring  seasonal  bird  activity  and  mor- 
tality at  the  Buffalo  Ridge  Windplant,  MN.  KE- 
NETECH Windpower,  Inc.,  Cooperative  Fish  and 
Wildlife  Research  Unit,  South  Dakota  State  Univ., 
Brookings,  South  Dakota. 

Hurlbert,  S.  H.  1984.  Pseudoreplication  and  the  de- 
sign of  ecological  field  experiments.  Ecol.  Mon- 
ogr.  54:187-211. 

Igl,  L.  D.  and  D.  H.  Johnson.  1995.  Dramatic  increase 
of  Le  Conte’s  Sparrow  in  Conservation  Reserve 
Program  fields  in  the  northern  Great  Plains.  Prairie 
Nat.  27:89-94. 

Johnson,  D.  H.  and  L.  D.  Igl.  1995.  Contributions  of 
the  Conservation  Reserve  Program  to  populations 
of  breeding  birds  in  North  Dakota.  Wilson  Bull. 
107:709-718. 

Johnson,  D.  H.  and  M.  D.  Schwartz.  1993.  The  Con- 
servation Reserve  Program:  habitat  for  grassland 
birds.  Great  Plains  Res.  3:273-295. 

Kantrud,  H.  a.  1981.  Grazing  intensity  effects  on  the 
breeding  avifauna  of  North  Dakota  native  grass- 
lands. Can.  Field-Nat.  95:404-417. 

King,  J.  W.  and  J.  A.  Savidge.  1995.  Effects  of  the 
Conservation  Reserve  Program  on  wildlife  in 
southeast  Nebraska.  Wildl.  Soc.  Bull.  23:377-385. 

Koford,  R.  R.,  j.  B.  Dunning,  Jr.,  C.  A.  Ribic,  and 
D.  M.  Finch.  1994.  A glossary  for  avian  conser- 
vation biology.  Wilson  Bull.  106:121  — 137. 

Leddy,  K.  L.  1996.  Effects  of  wind  turbines  on  non- 
game birds  in  Conservation  Reserve  Program 
grasslands  in  southwestern  Minnesota.  M.S.  the- 
sis, South  Dakota  State  Univ..  Brookings. 

Leddy,  K.  L.,  K.  F.  Higgins,  and  D.  E.  Naugle.  In 
press.  The  importance  of  Conservation  Reserve 
Program  fields  to  breeding  grassland  birds  at  Buf- 
falo Ridge,  Minnesota.  S.D.  Acad.  Sci. 

Mikol,  S.  a.  1980.  Field  guidelines  for  using  transects 


104 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  /,  March  1999 


to  sample  nongame  bird  populations.  U.S.  Fish 
and  Wildl.  Serv.  Bio.  Serv.  Prog.  OBS-80— 58:1  — 
26. 

Millenbah,  K.  E,  S.  R.  WiNTERSTEiN,  H.  Campa,  III, 
L.  T.  Furrow,  and  R.  B.  Minnis.  1996.  Effects  of 
Conservation  Reserve  Program  field  age  on  avian 
relative  abundance,  diversity,  and  productivity. 
Wilson  Bull.  108:760-770. 

Nelson,  H.  K.  and  R.  C.  Curry.  1995.  Assessing  avi- 
an interactions  with  windplant  development  and 
operation.  Trans.  N.  Am.  Wildl.  Nat.  Resour. 
Conf.  60:266-277. 

Orloff,  S.  G.  and  a.  W.  Flannery.  1992.  Wind  tur- 
bine effects  on  avian  activity,  habitat  use,  and 
mortality  in  (the)  Altamont  Pass  and  Solano 
County  Wind  Resource  Areas,  1989-1991.  Ala- 
meda, Contra  Costa  and  Solano  Counties,  Cali- 
fornia Energy  Commission,  Biosystems  Analysis, 
Inc.,  Tiburon,  California. 

Pedersen,  M.  B.  and  E.  Poulsen.  1991.  Avian  re- 
sponses to  the  implementation  of  the  Tjaereborg 
Wind  Turbine  at  the  Danish  Wadden  Sea.  Dan. 
Wildtundersogelser  47:1-44. 

Ralph,  C.  J.,  G.  R.  Geupel,  P.  Pyle,  T.  E.  Martin, 
AND  D.  E DeSante.  1993.  Handbook  of  field 
methods  for  monitoring  landbirds.  Gen.  Tech. 
Rep.  PSW-GTR- 144: 1-41. 

Ree.se,  K.  P.  and  J.  T.  Ratti.  1988.  Edge  effect:  a con- 
cept under  scrutiny.  Trans.  N.  Am.  Wildl.  Nat.  Re- 
sour. Conf.  53:127-136. 

Robel,  R.  j.,  j.  N.  Briggs,  A.  D.  Dayton,  and  L.  C. 
Hulbert.  1970.  Relationships  between  visual  ob- 


struction measurements  and  weight  of  grassland 
vegetation.  J.  Range  Manage.  23:295-297. 

SAS  Institute,  Inc.  1989.  SAS/STAT  user’s  guide, 
version  6,  fourth  ed.  SAS  Institute,  Inc.  Cary, 
North  Carolina. 

Skirvin,  a.  a.  1981.  Effect  of  time  of  day  and  time 
of  season  on  number  of  observations  and  density 
estimates  of  breeding  birds.  Stud.  Avian  Biol.  6: 
271-274. 

Swanson,  D.  A.  1996.  Nesting  ecology  and  nesting 
habitat  requirements  of  Ohio’s  grassland-nesting 
birds:  a literature  review.  Ohio  Fish  Wildl.  Rep. 
13:1-60. 

UsGAARD,  R.  E.,  D.  E.  Naugle,  K.  E Higgins,  and  R. 
G.  Osborn.  In  press.  Effects  of  wind  turbines  on 
nesting  raptors  at  Buffalo  Ridge  in  southwestern 
Minnesota.  S.D.  Acad.  Sci. 

Verner,  j.  and  L.  V.  Ritter.  1986.  Hourly  variation 
in  morning  point  counts  of  birds.  Auk  103:117— 
124. 

Wakeley,  j.  S.  1987.  Avian  line-transect  methods. 
Section  6.3.2,  U.S.  Army  Corps  Eng.  Wildl.  Re- 
sour. Manage.  Manual.  Tech.  Rep.  EL-87-5: 1—21. 

Wiens,  J.  A.  1969.  An  approach  to  the  study  of  eco- 
logical relationships  among  grassland  birds.  Or- 
nithol.  Monogr.  8:1-93. 

Winkelman,  E.  1990.  Impact  of  the  wind  park  near 
Urk,  Netherlands,  on  birds:  bird  collision  victims 
and  disturbance  of  wintering  fowl.  Int.  Ornithol. 
Cong.  20:402-403. 

Young,  E.  C.  and  C.  T.  Osborn.  1990.  Costs  and  ben- 
efits of  the  Conservation  Reserve  Program.  J.  Soil 
Water  Conserv.  45:370—373. 


Wilson  Bull.,  111(1),  1999,  pp.  105-1  14 


AVIAN  USE  OF  PURPLE  LOOSESTRIFE  DOMINATED  HABITAT 
RELATIVE  TO  OTHER  VEGETATION  TYPES  IN  A LAKE  HURON 

WETLAND  COMPLEX 

MICHAEL  B.  WHITT,'  HAROLD  H.  PRINCE,'  AND  ROBERT  R.  COX,  JR.- 

ABSTRACT. — Purple  loosestrife  (Lythriini  salicaria),  native  to  Eurasia,  is  an  introduced  perennial  plant  in 
North  American  wetlands  that  displaces  other  wetland  plants.  Although  not  well  studied,  purple  loosestrife  is 
widely  believed  to  have  little  value  as  habitat  for  birds.  To  examine  the  value  of  purple  loosestrife  as  avian 
breeding  habitat,  we  conducted  early,  mid-,  and  late  season  bird  surveys  during  two  years  (1994  and  1995)  at 
258  18-m  (0. 1 ha)  fixed-radius  plots  in  coastal  wetlands  of  Saginaw  Bay,  Lake  Huron.  We  found  that  loosestrife- 
dominated  habitats  had  higher  avian  densities,  but  lower  avian  diversities  than  other  vegetation  types.  The  six 
most  commonly  observed  bird  species  in  all  habitats  combined  were  Sedge  Wren  (Cislolhorus  plaiensis).  Marsh 
Wren  (C.  palustris).  Yellow  Warbler  (Deiulroica  petechia).  Common  Yellowthroat  (Geothylpis  trichas).  Swamp 
Sparrow  (Melospiza  georf>iana),  and  Red-winged  Blackbird  (Agelaius  phoeniceus).  Swamp  Spanow  densities 
were  highest  and  Marsh  Wren  densities  were  lowest  in  loosestrife  dominated  habitats.  We  observed  ten  breeding 
species  in  loosestrife  dominated  habitats.  We  conclude  that  avian  use  of  loo.sestrife  warrants  further  quantitative 
investigation  because  avian  use  may  be  higher  than  is  commonly  believed.  Received  27  May  1998,  accepted  26 
Aug.  1998. 


Purple  loosestrife  (Lythrum  salicaria)  is  an 
exotic,  broad-leaved,  herbaceous  perennial 
that  is  common  in  North  American  freshwater 
wetland  habitats  north  of  35°  N latitude 
(Thompson  1989).  Loosestrife  is  native  to 
Eurasia  where  it  occurs  in  freshwater  marshes, 
open  stream  margins,  and  alluvial  floodplains; 
it  invades  similar  habitats  in  North  America 
(Thompson  1989).  Common  plant  associates 
of  loosestrife  in  North  American  wetland  hab- 
itats such  as  cattails  (Typha  spp.),  reed  canary 
grass  (Phalaris  anmdinacea),  sedges  (Carex 
spp.),  and  rushes  (Jimcus  spp.)  closely  resem- 
ble its  associates  in  Eurasian  wetlands 
(Thompson  et  al.  1987).  Loosestrife  out  com- 
petes and  partially  or  completely  replaces  na- 
tive emergent  vegetation  (Thompson  1989). 
Loosestrife  often  pioneers  in  disturbed  areas 
such  as  drainage  ditches  (Wilcox  1995)  and 
displaces  moist-soil  species  such  as  smart- 
weeds  {Polygonum  spp.)  and  millets  (Panicum 
spp.)  on  mudflats  (Thompson  et  al.  1987). 
Species  of  wetland  plants  become  distributed 
along  a wetland  gradient  and  are  good  indi- 


' Michigan  State  Univ.,  Dept,  of  Fisherie.s  and  Wild- 
life, East  Lansing,  Ml  48824. 

-Northern  Prairie  Wildlife  Research  Center.  8711 
37th  Street  SE,  Jamestown,  ND  58401. 

^ Present  address:  Svoboda  Ecological  Resources, 
2477  Shadywood  Road.  Excelsior,  MN  55331; 

E-mail:  michaeLwhitt@hotmail.com 
^ CoiTesponding  author. 


cators  of  long-term  hydrology  and  other  abi- 
otic factors  (Keddy  and  Reznicek  1985).  Wet- 
land vegetation  types  generally  grade  from 
forested  wetland  to  shrub-scrub,  to  wet  mead- 
ow, to  strand  (or  mudflat),  to  emergent  marsh, 
and  finally,  to  open  water  (Cowardin  et  al. 
1979,  Keddy  and  Reznicek  1985).  Loosestrife 
occupies  zones  near  the  strand  including 
emergent  and  wet  meadow  zones. 

Avian  use  of  loosestrife  is  not  well  studied 
(Thompson  et  al.  1987).  Prince  and  Flegel 
(1995)  found  no  records  in  the  literature  of 
loosestrife  as  avian  food  or  nesting  habitat  in 
Lake  Huron  wetlands.  In  New  York  wetlands. 
Rawinski  and  Malecki  (1984)  observed  that 
Marsh  Wrens  (Cistothorus  palustris)  preferred 
cattail  for  nesting,  whereas  Red-winged 
Blackbirds  (Agelaius  phoeniceus)  preferred 
loosestrife  for  nesting.  Rawinski  and  Malecki 
(1984)  also  noted  that  Black-crowned  Night- 
herons  (Nycticorax  nycticorax)  roosted  in 
loosestrife,  and  Pied-billed  Grebes  {Podilyin- 
hus  podiceps)  nested  in  one-  and  two-year-old 
emergent  loosestrife  stands.  Kiviat  (1996) 
found  15  American  Goldfinch  (Carduelis  tris- 
tis)  nests  in  loosestrife  during  a 23-year  study 
of  birds  in  the  Hudson  Valley.  Swift  and  co- 
workers (1988)  observed  Least  Bitterns  {l.xo- 
hrychus  exilis)  and  other  birds  in  Hudson  Riv- 
er wetlands  that  consisted  of  cattail,  river  bul- 
rush (Scirpus  fluviatilis),  loosestrife,  and  com- 
mon reed  (Phragmites  australis). 


105 


106 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


PaJustrine®P 


buttonbush 

{Cephalanthus  ocadentalis) 
silky  dogwood 
{Cortms  amonum) 
sedges 
(Carer  spp.) 


sedges  blue-joint  reedgrass 

{Carex  spp.)  (Ca/amagrostis  canadensis) 
blue-joint  reedgrass  red-osier  dogwood 

(Calamagroslis  canadensis)  (Cormis  slolonifera) 
red-osier  dogwood  purple  loosestrife 

{Cormis  stolonifera)  {Lyihrum  salicaria) 


blue-joint  reedgrass  blue-joint  reedgrass 

{Calamagroslis  canadensis)  {Calamagroslis  canadensis) 
sedges  purple  loosestrife 

(Carer  spp.)  {Lyihrum  salicaria) 


cattail 

{Typhaspp.) 


bulrush 

{Scirpus  amencana) 


Seasonal « C 


Saturated  =>  B 


Saturated  = B 


Saturated  = B 


Saturated  = B 


Semi- 
permanent = F 


Inicrmittenly 
exposed  « G 


Intermittenly 
exposed  = G 


PSSIC 

PEMl/SSlB 

PEMI/SSIB 

PEMIB 

Senib-shrub 

Wet  meadow/ 

Wet  meadow/ 

Wet 

scrub-shrub 

scrub-shrub/ 

meadow 

loosestrife 

PEMIB 

PEMIF 

PEMIG 

PEMIG 

Wet 

Inland 

Coastal 

Coastal 

Meadow/ 

cattail 

cattail 

bulrush 

loosestrife 

FIG.  1.  Characteristics  of  surveyed  vegetation  types  in  Saginaw  Bay  wetlands,  1994-1995,  based  on  Na- 
tional Wetlands  Inventory  (Cowardin  et  al.  1979)  classification  system.  PSSIC  = Palustrine,  broad-leafed  de- 
ciduous scrub-shrub,  and  seasonally  flooded;  PEMI/SSIB  = Palustrine,  persistent  emergent/broad-leafed  decid- 
uous scrub-shrub,  and  saturated;  PEMIB  = Palustrine,  persistent  emergent,  and  saturated;  PEMIF  = Palustrine, 
persistent  emergent,  and  semi-permanently  flooded;  PEMIG  = Palustrine,  persistent  emergent,  and  intermittently 
exposed  (Cowardin  et  al.  1979). 


Minnesota  established  the  hrst  statewide 
loosestrife  control  program  in  1987  with  the 
goal  of  broadening  public  awareness,  con- 
ducting inventories,  developing  control  meth- 
ods, and  initiating  control  work  (Skinner  et  al. 
1994).  Minnesota  has  spent  $US  1.75  million 
since  the  beginning  of  the  program  (Skinner, 
pers.  comm.).  Other  state  and  federal  agencies 
also  have  spent  considerable  money  and  effort 
to  control  loosestrife,  in  part,  because  wildlife 
values  of  this  plant  are  widely  regarded  to  be 
limited.  Methods  of  control  have  included  use 
of  chemicals,  water  manipulation,  mowing, 
tillage,  planting  robust  mudflat  species  such  as 
Japanese  millet  (Thompson  1989),  and,  most 
recently,  biological  control  using  insects  (Ma- 
lecki  et  al.  1993). 

Our  objective  was  to  compare  avian  use  of 
vegetation  zones  dominated  by  loosestrife 
with  other  wetland  zones  where  loosestrife 
was  absent  or  not  dominant.  Comparison  of 
avian  breeding  species  richness,  density,  and 
diversity  is  a necessary  hrst  step  to  assess  the 
value  of  loosestrife-dominated  habitats  to 
birds,  and  ultimately  to  evaluate  costs  and 
benehts  of  loosestrife  control. 


METHODS 

We  conducted  field  work  during  1994  and  1995  in 
Bay,  Tuscola,  and  Huron  counties  adjacent  to  Saginaw 
Bay,  Lake  Huron,  Michigan.  Saginaw  Bay  comprises 
the  majority  of  remaining  wetland  habitat  on  Lake  Hu- 
ron because  unsuitable  shore  morphology  (e.g.,  cliffs) 
prohibited  wetland  formation,  and  development  pres- 
sures (mostly  agricultural)  eliminated  presettlement 
wetland  habitats  (Prince  and  Flegel  1995).  Although 
this  area  has  experienced  a 50%  overall  wetland  loss 
(Dahl  1990),  70%  of  inland  wetlands  and  99%  of  lake- 
plain  prairies  have  been  drained  and  converted  to  other 
uses  (Comer  1996).  Most  existing  Saginaw  Bay  wet- 
lands are  disturbed  by  adjacent  urban  and  agricultural 
development,  diking,  and  exotic  flora  and  fauna. 

We  surveyed  birds  on  18-m  fixed-radius  plots  in 
eight  vegetation  types  based  on  hydrology  and  plant 
form  and  structure:  scruh-shrub,  wet  meadow/scrub- 
shrub,  wet  rneadow/scrub-shrub/loosestrife,  wet  mead- 
ow, wet  meadow/loo.sestrife,  inland  cattail,  coastal  cat- 
tail, and  coastal  bulrush  (Sc/r/JZ/.v  spp.).  Our  habitat 
classifications  were  based  on  Cowardin  and  coworkers 
(1979);  dominant  plants  had  greater  than  30%  cover 
(Fig.  1).  We  used  a split  class  (e.g.,  broad-leafed  de- 
ciduous scrub-shrub/persistent  emergent;  National 
Wetlands  Inventory)  to  classify  two  vegetation  types 
because  scattered  shrubs  of  at  least  30%  cover  were 
present.  We  separated  cattail  sites  into  coastal  and  in- 
land because  hydrologies  differed;  coastal  sites  were 


Whitt  et  id.  • AVIAN  USE  OF  PURPLE  LOOSESTRIFE 


107 


TABLE  1.  Mean  cover  (Robel)  height  (cm)  ± 
type  in  Saginaw  Bay  wetlands,  1994-1995. 

SE  and 

mean  water  depth  (cm) 

± SE  by  period  and  vegetation 

Period  1 

Period 

Period  3 

Year  and  site^ 

Cover  height 

Water  depth 

Cover  height 

Water  depth 

Cover  height 

Water  depth 

1994  SS 

27.8  ± 

4.3 

27.5  ± 0.9 

55.0 

•+- 

5.9 

12.1  ± 3.2 

75.7 

-h 

8.1 

25.4  ± 2.9 

1995  SS 

35.8  ± 

3.4 

24.4  ± 0.8 

68.8 

-h 

4.3 

14.4  ± 0.7 

74.6 

-h 

5.5 

1.9  ± 0.6 

1994  WM/SS 

b 

— 

62.8 

-h 

4.2 

saturated 

56.7 

-h 

3.5 

saturated 

1995  WM/SS 

24.4  ± 

4.1 

0.2  ± 0.2 

67.8 

7.8 

saturated 

100.8 

-H 

8.7 

saturated 

1994  WM/SS/LS 

— 

— 

107.5 

1 1.0 

saturated 

101.1 

-h 

3.9 

saturated 

1995  WM/SS/LS 

30.0  ± 

8.5 

saturated 

48.9 

4.2 

saturated 

87.8 

H- 

9.5 

0.5  ± 1.7 

1994  WM/LS 

— 

84.4 

-+- 

8.8 

saturated 

106.7 

-h 

8.2 

2.2  ± 0.4 

1995  WM/LS 

38.9  ± 

4.8 

saturated 

63.3 

3.3 

saturated 

87.8 

-h 

6.4 

saturated 

1994  WM 

— 

— 

81.7 

5.1 

0.3  ± 0.2 

83.9 

-h 

2.6 

1.8  ± 0.8 

1995  WM 

30.6  ± 

2.6 

saturated 

61.1 

-E 

4.3 

saturated 

68.3 

-t- 

6.7 

saturated 

1994  IC 

80.0  ± 

3.2 

29.9  ± 2.0 

70.7 

-h 

4.1 

21.0  ± 1.7 

1 10.6 

-h 

7.2 

26.4  ± 1.4 

1995  IC 

41.4  ± 

6.6 

17.1  ± 3.4 

52.4 

9.0 

13.3  ± 2.5 

94.8 

H- 

6.4 

7.7  ± 1.9 

1994  CC 

— 

— 

71.2 

-H 

7.5 

22.0  ± 2.6 

144.3 

-1- 

8.2 

31.6  ± 2.4 

1995  CC 

60.3  ± 

4.7 

9.1  ± 1.4 

82.2 

-h 

6.3 

12.5  ± 1.5 

127.9 

8.3 

22.2  ± 1.7 

1994  CB 

— 

— 

11.3 

-h 

3.4 

33.5  ± 1.4 

34.7 

5.5 

36.8  ± 1.2 

1995  CB 

— 

— 

8.9 

-+■ 

2.7 

23.7  ± 0.8 

33.2 

3.6 

30.1  ± 1.1 

“ Vegetation  types:  SS  = scrub-shrub,  WM/SS  = wet  meadow/scrub-shrub.  WM/SS/LS  = wet  meadow/scrub-shrub/loosestrife.  WM/LS  = wet  meadow/ 
loosestrife.  WM  = wet  meadow.  IC  = inland  cattail.  CC  = coastal  cattail.  CB  = coastal  bulrush. 

Dashes  ( — ) indicate  insufficient  or  lack  of  data. 


intermittently  exposed,  whereas  inland  sites  were  semi- 
permanently flooded  by  groundwater  and  precipitation. 

Sampling  periods  were  divided  into  an  early  season 
during  the  second  and  third  weeks  of  May,  a mid- 
season during  the  first  and  second  weeks  of  June,  and 
a late  season  during  the  last  week  of  June  and  first 
week  of  July.  We  conducted  surveys  between  sunrise 
and  10:00  EST.  Surveys  were  not  conducted  if  sus- 
tained winds  exceeded  24  km/h  or  during  heavy  rain. 

We  selected  plots  using  the  following  protocol:  first, 
an  azimuth  was  determined  that  traversed  the  habitat. 
The  center  of  the  first  plot  was  placed  at  least  18  m 
from  the  outer  boundary  of  the  vegetation  on  that  az- 
imuth. The  center  of  the  next  plot  was  70  m from  the 
first  plot  on  the  same  azimuth.  This  procedure  was 
continued  until  observers  surveyed  three  or  more  plots 
or  reached  a different  vegetation  type.  If  fewer  than  3 
plots  were  established  on  the  first  azimuth,  we  estab- 
lished a second  azimuth,  approximately  perpendicular 
to  the  first  azimuth,  that  traversed  the  vegetation  type 
and  permitted  plot  placement  at  least  70  m from  other 
plots.  Plots  were  set  on  this  azimuth  in  the  same  man- 
ner as  on  the  first  azimuth.  Plots  were  placed  in  dif- 
ferent locations  at  the  same  site  among  time  periods 
to  avoid  resampling  the  same  plots  and  recounting  the 
same  nests.  Coastal  bulrush  plots  were  not  surveyed 
during  the  first  periods  of  each  year  because  they 
lacked  structure;  new  vegetative  growth  was  not  yet 
established  and  the  previous  year's  growth  was  elimi- 
nated by  ice  action.  Neither  did  we  survey  three  veg- 
etation types  (wet  meadow/scrub-shrub/loosestrife,  wet 
meadow,  wet  meadow/loosestrife)  during  the  first  pe- 
riod of  1994.  We  surveyed  258  plots  in  8 wetland  hab- 
itats. 

Observers  waited  5 min  for  normal  bird  activity  to 


resume  after  arriving  at  a survey  plot.  We  recorded  all 
birds  seen  or  heard  on  plots  during  a 7-min  observa- 
tion period.  We  recorded  flying  birds  if  their  flight 
originated  or  terminated  within  the  plot  and  we  tallied 
individual  birds  only  once.  We  played  tape-recorded 
calls  (Peterson  1990)  of  five  secretive  species  [Amer- 
ican Bittern  (Botauru.s  lentiginosus).  Least  Bittern, 
King  Rail  (Rcdlus  elegcm.s)^  Virginia  Rail  {R.  limicola), 
and  Sora  (Porzana  Carolina)]  during  the  last  3 min 
using  portable  cassette  recorders  (Johnson  et  al.  1981, 
Marion  et  al.  1981,  Johnson  and  Dinsmore  1986).  We 
played  calls  for  25-30  sec  followed  by  10  sec  of  si- 
lence. We  measured  water  depth  and  vertical  cover  4 
m from  the  plot  center  at  0°.  120°,  and  240°  (Table  1). 
Observers  measured  vertical  cover  to  the  nearest  10 
cm  using  a 2-m  Robel  pole  placed  at  plot  center  and 
viewed  while  maintaining  eye  level  1 m above  the  wa- 
ter surface  or  ground  level  and  looking  back  toward 
plot  center  (Higgins  et  al.  1994).  Workers  returned  to 
plots  later  that  day  and  searched  the  innermost  13-m 
radius  (0.05  ha)  portion  for  nests.  A bird  species  was 
designated  as  breeding  when  nests  or  flightless  young 
were  observed  in  one  or  more  periods  or  when  adults 
were  observed  in  two  of  three  periods  (Brown  and 
Dinsmore  1986).  A nest  verified  breeding  status  when 
eggs,  young,  or  strong  evidence  of  use  such  as  egg 
shell  fragments,  down,  or  fecal  sacs  were  pre.sent.  We 
considered  predated  nests  as  breeding  evidence  when 
prey  species  could  be  determined.  We  also  tallied  spe- 
cies as  breeding  if  they  were  ob.served  within  the  sam- 
pled vegetation  type  but  outside  of  plot  boundaries  on 
two  of  three  visits. 

We  tallied  breeding  species  richness  (i.e.,  number  of 
breeding  species)  for  each  vegetation  type.  We  calcu- 
lated avian  diversities  for  each  plot  using  the  Shannon- 


108 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  1,  March  1999 


TABLE  2.  Distribution  of  breedin 

ig  birds  by  vegetation  type  in  Saginaw  Bay  wetlands. 

1994- 

1995. 

Breeding 

status  based  on  observation  of  adults 

on  at  least  two  of  three  visits,  or  a nest  or  fli 

ghtless 

young 

on  at  least  one 

\isit. 

Species 

SS“ 

WM/SS“ 

WM/SS/LS“  WM/LS“ 

WM"* 

tea 

CC“ 

CB“ 

Pied-billed  Grebe^ 

American  Bittern 

X 

X 

X 

X 

Least  Bittern 

X 

X 

Canada  Goose 

Wood  Duck*’ 

X 

X 

X 

American  Black  Duck 

X 

Mallard 

X 

X 

X X 

X 

X 

Blue-winged  Teal 

RedheaeP 

Northern  Harrier'’ 

X 

X 

X 

X 

Ring-necked  Pheasant 

Virginia  Rail 

X 

X 

X X 

X 

X 

X 

Sora 

X 

X 

X 

Common  Moorhen/American  CooL 

X 

X 

X 

Forster's  Tern'’ 

X 

Black  Tern'’ 

Northern  Flicker 

X 

X 

X 

Eastern  Wood-pewee 

Willow  Flycatcher 

X 

X 

Great  Crested  Flycatcher 

X 

Eastern  Kingbird 

X 

X 

Tree  Swallow 

X 

X 

Sedge  Wren 

Marsh  Wren 

X 

X 

X 

X 

X 

Gray  Catbird 

X 

X 

Yellow  Warbler 

X 

X 

X 

Common  Yellowthroat 

X 

X 

X 

X 

X 

X 

Rose-breasted  Grosbeak 

Savannah  Sparrow 

X 

X 

X 

Song  Sparrow 

X 

X 

Swamp  Sparrow 

X 

X 

X X 

X 

X 

X 

Bobolink 

Red-winged  Blackbird 

X 

X 

X X 

X 

X 

X 

Yellow-headed  Blackbird 

Brewer's  Blackbird 

X 

X 

Brown-headed  Cowbird 

X 

X 

Baltimore  Oriole 

American  Goldfinch 

X 

X 

X 

Total 

20 

15 

9 5 

9 

13 

13 

2 

^ Vegetation  types:  SS  = scmb-shnib.  WM/SS  = 

wet  meadow/.scrub-.shrub,  WM/SS/LS  = wet  meadow/scrub-shrub/loose.strife,  WM/LS  = 

wet  meadow/ 

loosestrife.  WM  = wet  meadow.  1C  = coastal  cattail.  CB  = coa.stal  bulrush. 

^.Species  observed  within  the  sampled  vegetation  type  but  not  on  plots. 

American  Coot  and  Common  Moorhen  were  grouped  together  because  these  species  were  most  often  observed  by  call  only  and  their  calls  are  difficult 
to  distinguish. 


Weiner  diversity  index.  Density  was  the  number  of 
birds  (both  sexes)  observed  on  a plot  multiplied  by  10 
to  obtain  density  per  hectare. 

We  used  ANOVA  (PROC  GLM;  SAS  1990;  SAS 
6. 1 2 for  Windows)  to  as.sess  fixed  effects  of  vegetation, 
period,  year,  tmd  their  interactions  on  avian  density  and 
diversity.  Residuals  were  normally  distributed,  but  var- 
iiinces  were  not  homogeneous  because  we  never  ob- 
served some  species  in  one  or  more  habitats  (resulting 


in  means  and  variances  of  zero).  However,  the  overall 
F-statistic  from  ANOVA  is  robust  to  violations  in  as- 
sumptions of  homogeneous  variances  (Sokal  and  Rohlf 
1981 ).  Early-period  observations  were  eliminated  from 
all  analyses  because  of  missing  data.  We  considered 
plots  as  the  experimental  units  becau.se  we  decided  a 
priori  to  restrict  our  inference  to  Saginaw  Bay  wet- 
lands. We  used  a = 0.05  for  all  statistical  comparisons. 
We  initially  analyzed  fully  specified  models  (all  main 


Whitt  et  cil.  • AVIAN  USE  OF  PURPLE  LOOSESTRIFE 


109 


etfects  and  interactions  included).  We  fitted  eacli  mod- 
el using  a backward,  stepwise  procedure  by  eliminat- 
ing non-significant  {P  > 0.05)  effects,  beginning  with 
highest-order  interactions.  Thus,  our  final  models  in- 
cluded only  significant  effects  or  interactions,  and 
main  effects  or  interactions  contained  in  significant 
higher-order  interactions.  We  used  Fisher's  protected 
least  significant  difference  test  to  isolate  differences 
antong  least-square  means  (LSMEANS,  SAS  1990)  for 
significant  effects  in  the  ANOVA  (Milliken  and  John- 
son 1984).  We  compared  density  and  diversity  of  birds 
in  loosestrife-dominated  vegetation  types  (wet  mead- 
ow/scrub-shrub/loosestrife and  wet  meadow/loose- 
strite)  to  those  in  other  vegetation  types  using  orthog- 
onal contrasts  (PROC  GLM;  SAS  1990),  and  estimated 
least-square  means  using  estimate  statements  (PROC 
GLM;  SAS  1990).  We  developed  similar  models  for 
abundance  of  the  six  most  commonly  observed  bird 
species:  Sedge  Wren  (Cistothoriis  platensis).  Marsh 
Wren.  Yellow  Warbler  (Dendroica  petechia).  Common 
Yellowthroat  (Geothlypis  trichas).  Swamp  Sparrow 
(Meiospiza  georgiana),  and  Red-winged  Blackbird. 

Standard  errors  reported  are  for  least-square  means 
(SAS  1990).  Because  multiple  comparison  of  means 
with  heterogenous  variances  may  be  misleading  (Sokal 
and  Rohlf  1981),  we  further  examined  comparisons  of 
non-zero  means  to  means  of  zero  using  confidence  in- 
tervals. For  each  mean  of  zero,  we  constructed  a 90% 
upper  confidence  limit  after  assigning  the  highest  stan- 
dard deviation  associated  with  any  mean  in  the  model. 
We  then  compared  90%  lower  confidence  intervals  for 
nonzero  means  to  90%  upper  confidence  intervals  for 
zero  means;  we  considered  failure  of  these  intervals  to 
overlap  as  statistically  significant.  Resulting  confi- 
dence intervals  for  zero  means  are  likely  overestimat- 
ed, yielding  a con.servative  comparison.  We  note  in 
tables  instances  where  confidence  interval  compari- 
.sons  did  not  corroborate  multiple  comparisons  using 
Fisher's  least  significant  difference  test. 

RESULTS 

As  the  season  progressed  water  depths  at 
coastal  sites  (coastal  cattail  and  coastal  bul- 
rush) increased  and  those  at  inland  sites  de- 
creased while  vertical  cover  generally  in- 
creased at  all  sites  (Table  1).  We  surveyed  258 
plots  and  observed  39  breeding  bird  species 
in  Saginaw  Bay  wetland  habitats  (Table  2). 
Six  breeding  species  were  observed  in  the 
sampled  vegetation  type,  but  not  on  survey 
plots:  Pied-billed  Grebe,  Wood  Duck  (Aix 
sponsa).  Redhead  (Aythya  americana).  North- 
ern Harrier  {Circus  cyaneus),  Forster’s  Tern 
(Sterna  forsteri),  and  Black  Tern  (Chlidonias 
niger).  We  also  observed  10  species  breeding 
in  loosestrife  dominated  habitats  (Table  2). 

Marsh  Wren  (n  = 20),  Swamp  Sparrow  (/; 
= 16),  and  Red-winged  Blackbird  {u  = 21) 


were  the  most  commonly  observed  nests  on 
all  plots  (Table  3).  We  observed  Mallard, 
Blue-winged  Teal  (Anas  discors),  Virginia 
Rail,  and  Red-winged  Blackbird  nests  while 
traversing  between  plots  in  loosestrife-domi- 
nated vegetation  zones,  but  not  on  the  plots. 

Avian  density  and  diversity. — Our  final 
model  indicated  that  avian  density  differed 
only  in  relation  to  vegetation  (ANOVA:  F = 
14.45,  df  = 7,  181,  P < 0.001;  Table  4).  Avi- 
an density  was  higher  (orthogonal  contrast:  F 
= 8.87,  df  = 1,  181,  P = 0.003)  in  loosestrife- 
dominated  vegetation  types  [46.9  ± 3.8  (SE) 
birds/haj  than  in  other  vegetation  types  (34.7 
± 1.6).  Avian  diversity  also  differed  only  in 
relation  to  vegetation  (ANOVA:  F = 12.76, 
df  = 7,  181,  P < 0.001;  Table  4).  Avian  di- 
versity was  lower  (orthogonal  contrast:  F = 
4.74,  df  = 1,  181,  P = 0.03)  in  loosestrife- 
dominated  vegetation  types  (0.42  ± 0.08)  than 
in  other  vegetation  types  (0.60  ± 0.03).  Ef- 
fects of  year,  period,  and  all  interactions  were 
not  significant  (P  > 0.05  for  all  tests)  for  both 
avian  density  and  diversity.  Scrub-shrub  con- 
tained the  highest  bird  species  diversity  and 
wet  meadow/loosestrife  and  coastal  bulrush 
the  lowest  (Table  4). 

Species  abundance. — The  vegetation  X pe- 
riod X year  interaction  was  significant  (AN- 
OVA: F = 2.34,  df  = 7,  157,  P = 0.03)  in 
our  initial  Sedge  Wren  model.  Thus,  vegeta- 
tion related  differences  in  Sedge  Wren  abun- 
dance were  not  consistent  among  periods  and 
years  (Table  5).  Within  periods  and  years, 
Sedge  Wren  abundance  did  not  differ  (orthog- 
onal contrasts:  P > 0.05  for  all  tests)  between 
loosestrife  dominated  vegetation  types  and 
other  vegetation  types. 

Marsh  Wren  abundance  differed  among 
vegetation  types  (ANOVA:  F = 30.72,  df  = 
7,  181,  P < 0.001;  Table  5).  Marsh  Wren 
abundance  was  lower  (orthogonal  contrast:  F 
— 10.73,  df  = 1,  181,  P = 0.001)  in  loose- 
strife-dominated  vegetation  types  (0  ± 1.8) 
than  in  other  vegetation  types  (6.2  ± 0.7). 
Yellow  Warbler  abundance  differed  among 
vegetation  types,  but  differences  were  not 
consistent  between  mid-  and  late  periods  (AN- 
OVA: vegetation  X period  interaction,  F = 
2.08,  df  = 7,  173,  P = 0.048;  Table  5).  The 
interaction  was  due  to  significantly  higher  (P 
< 0.001)  numbers  of  Yellow  Warblers  ob- 
served in  late  period  scrub-shrub  compared 


110 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


TABLE  3.  Number  of  nests  and  percent 
commonly  observed  bird  species  were  found 

of  plots  within  vegetation  types  where  nests 
in  Saginaw  Bay  wetlands,  1994—1995. 

of  the  three  most 

Vegetation  type“ 

Marsh  Wren 

Swamp  Sparrow 

Red-winged  Blackbird 

Total  plots*’ 

SS 

0 

4 (9%) 

13  (30%) 

43 

WM/SS 

0 

2 (6%) 

0 

31 

WM/SS/LS 

0 

2 (12%) 

0 

16 

WM/LS 

0 

4 (27%) 

0 

15 

WM 

0 

1 (4%) 

3 (11%) 

27 

IC 

8 (23%) 

3 (8%) 

5 (14%) 

35 

CC 

12  (17%) 

0 

0 

71 

CB 

0 

0 

0 

20 

Total  nests 

20 

16 

21 

— 

“ Vegetation  types:  SS  = scrub-shrub.  WM/SS  = wet  meadow/scrub-shrub,  WM/SS/LS  = wet  meadow/scrub-shrub/loosestrife,  WM/LS  = wet  meadow/ 
loosestrife.  WM  = wet  meadow.  1C  = inland  cattail.  CC  = coastal  cattail.  CB  = coastal  bulrush. 

Includes  early,  mid-,  and  late  season  surveys. 


with  mid-period  scrub-shrub  (Table  5).  Yellow 
Warbler  abundance  did  not  differ  (orthogonal 
contrast:  P > 0.05  for  both  tests)  between 
loosestrife-dominated  and  other  vegetation 
types  in  either  period.  Common  Yellowthroat 
abundance  differed  among  vegetation  types 
(ANOVA:  F = 6.04,  df  = 7,  181,  P < 0.001; 
Table  5).  Common  Yellowthroat  abundance 
did  not  differ  (orthogonal  contrast:  F = 1.20, 
df  = 1,  181,  P > 0.05)  between  loosestrife- 
dominated  and  other  vegetation  types. 

Swamp  Sparrow  abundance  differed  among 
vegetation  types  (ANOVA:  F = 39.03,  df  = 
7,  180,  P < 0.0001;  Table  5)  and  between 
periods  (ANOVA:  F = 6.88,  df  = 1,  180,  P 
= 0.009).  Swamp  Sparrow  abundance  was 
higher  during  the  late  period  (19.1  ± 1.1 
birds/ha)  compared  with  the  mid-period  (15.2 
± 1.1  birds/ha).  Swamp  Sparrow  abundance 
was  higher  (orthogonal  contrast:  F = 133.06, 
df  = 1,  180,  P < 0.001)  in  loosestrife-domi- 
nated vegetation  types  (36.0  ± 2.0)  than  in 
other  vegetation  types  (10.8  ± 0.8).  Swamp 
Sparrows  accounted  for  95%  and  65%  of  the 
overall  avian  density  at  wet  meadow/loose- 
strife and  wet  meadow/scrub-shrub/loosestrife 
plots,  respectively.  Abundance  of  Red-winged 
Blackbird  differed  among  vegetation  types, 
but  differences  were  not  consistent  between 
mid-  and  late  periods  (ANOVA:  vegetation  X 
period  interaction,  F = 2.14,  df  = 7,  173,  P 
= 0.04;  Table  5).  The  interaction  was  due  to 
significantly  higher  (P  < 0.001)  numbers  of 
Red-winged  Blackbirds  observed  in  mid-pe- 
riod scrub-shrub  compared  with  late  period 
scrub-shrub  (Table  5).  Red-winged  Blackbird 
abundance  did  not  differ  (orthogonal  con- 


trasts: P > 0.05  for  both  tests)  between  loose- 
strife dominated  and  other  vegetation  types  in 
either  period. 

DISCUSSION 

Weller  and  Spatcher  (1965),  Kantrud  and 
Stewart  (1984),  and  Burger  (1985)  concluded 
that  plant  form  and  structure,  rather  than  tax- 
onomic composition,  play  key  roles  in  habitat 
selection  by  marsh-nesting  birds.  The  struc- 
ture of  loosestrife  consists  of  stout,  wood-like 
persistent  growth  and  herbaceous  new  growth, 
similar  to  shrubs.  Overall,  species  richness  in 
loosestrife  was  slightly  lower  than  that  in  oth- 
er vegetation  types  except  coastal  bulrush  (Ta- 
ble 2).  Scrub-shrub  habitat  contained  the  high- 
est breeding  species  richness  and  diversity, 
but  these  values  may  be  explained  in  part  by 
the  location  of  scrub-shrub  as  an  ecotone  be- 
tween forest  and  emergent  wetland.  Several 
scrub-shrub  breeding  birds  were  not  wetland- 
dependent  species  but  instead  birds  of  forest 
edge  and  gaps  such  as  Northern  Flicker  (Co- 
laptes  aiiratus',  Moore  1995),  Eastern  Wood- 
pewee  (Contopus  virens;  McCarty  1996), 
Great  Crested  Flycatcher  (Myiarchus  crinitus; 
Lanyon  1997),  and  Brown-headed  Cowbird 
(Molothrus  ater\  Lowther  1993). 

Swamp  Sparrow  nests  were  most  abundant 
in  vegetation  types  where  loosestrife  was 
dominant  (Table  3).  Reinert  and  Golet  (1986) 
determined  that  breeding  Swamp  Sparrows 
principally  required  shallow  standing  water, 
low  (<1.5  m)  dense  cover,  and  elevated  song- 
posts,  similar  to  our  loosestrife-dominated 
sites.  Swamp  Sparrows  constructed  nests  us- 
ing fine-stemmed  sedges  and  grasses  anchored 


Whitt  et  al.  • AVIAN  USE  OF  PURPLE  LOOSESTRIFE 


TABLE  4.  Mean  avian  density  (no. /ha)  ± SE.  avian  diversity  (Shannon-Weaver)  ± SE,  and  number  of 
seeond  and  third  period  plots  by  {n)  vegetation  type  in  Saginaw  Bay  wetlands,  1994-1995. 


Vegetation  type*' 

Density'-' 

Diversity'-' 

SS 

30 

51.33  ± 3.4  A 

1.05  ± 0.08  A 

WM/SS 

19 

38.95  ± 4.3  BC 

0.63  ± 0.10  B 

WM/SS/LS 

13 

44.62  ± 5.2  ABC 

0.59  ± 0.12  B 

WM/LS 

12 

49.17  ± 5.4  AB 

0.22  ± 0.12  C 

WM 

21 

39.52  ±4.1  BC 

0.62  ± 0.09  B 

IC 

23 

41.74  ± 3.9  ABC 

0.74  ± 0.09  B 

CC 

51 

36.27  ± 2.6  C 

0.56  ± 0.06  B 

CB 

20 

0.5  ± 4.2  D 

0 C 

“ Vegetation  types:  SS  = .scrub-shrub,  WM/SS  = wet  meadow/scrub-shrub,  WM/SS/LS  = wet  meadow/,scrub-shrub/loo.sestrife.  WM/LS  = wet  meadow/ 
loosestrife.  WM  = wet  meadow,  1C  = inland  cattail,  CC  = coastal  cattail.  CB  = coastal  bulrush. 

I’  Excludes  early  period  surveys  becau.se  of  missing  data. 

Means  within  columns  followed  by  the  same  letter  do  not  differ  (F  > 0.05)  as  determined  by  ANOVA  and  Fisher's  lea.st  significant  difference. 


in  persistent  loosestrife  stalks.  We  also  ob- 
served Mallard,  Blue-winged  Teal,  Virginia 
Rail,  and  Red-winged  Blackbird  nests  at  our 
loosestrife-dominated  sites,  and  found  Amer- 
ican Bittern,  Sedge  Wren,  Yellow  Warbler, 
Common  Yellowthroat,  and  American  Gold- 
finch breeding  based  on  our  criteria.  Pied- 
billed Grebe  (Rawinski  and  Malecki  1984), 
Least  Bittern  (Swift  et  al.  1988),  Red-winged 
Blackbird  (Rawinski  and  Malecki  1984),  and 
American  Goldfinch  (Kiviat  1996)  were  ob- 
served nesting  in  loosestrife  habitats  previous 
to  this  study. 

Rawinski  and  Malecki  (1984)  observed  that 
Marsh  Wrens  preferred  cattail  habitats,  but 
Red-winged  Blackbirds  preferred  loosestrife 
habitats.  We  also  found  that  nesting  Marsh 
Wrens  used  cattail  habitats,  but  we  observed 
Red-winged  Blackbird  nests  most  frequently 
in  scrub-shrub  zones  (Table  5).  Inconsisten- 
cies in  vegetation  type,  period,  and  year  ef- 
fects (i.e.,  significant  three-way  interaction)  on 
Sedge  Wren  abundance  may  reflect  this  spe- 
cies’ variable  breeding  site  selection  (Table  5). 
Bums  (1982)  observed  that  Sedge  Wrens 
show  little  site  fidelity;  this  characteristic  may 
be  due  to  the  ephemeral  nature  of  wet  mead- 
ow habitats  (Kroodsma  and  Verner  1978).  We 
believe  that  Sedge  Wren  abundance  may  de- 
cline as  loosestrife  increases  in  wet  meadow 
canopies.  We  observed  greater  areal  cover  of 
loosestrife  at  the  wet  meadow/loosestrife  site 
compared  with  the  wet  meadow/scrub-shrub/ 
loosestrife  site  and  Sedge  Wren  abundance 
was  significantly  higher  in  two  of  four  sam- 
pling periods  at  the  site  with  less  loosestrife 
(Table  5). 


The  avian  diversity  in  loosestrife  dominated 
habitats  was  lower  on  average  than  that  of 
other  wetland  habitats  that  we  surveyed,  in- 
dicating uneven  distributions  of  fewer  species. 
We  found  higher  avian  densities  in  loosestrife- 
dominated  habitats  compared  to  other  vege- 
tation types,  although  Swamp  Sparrows  com- 
prised the  majority  of  overall  density  in  loose- 
strife habitats.  Swamp  Sparrows  accounted  for 
59%  of  the  overall  wet  meadow  density. 
Swamp  Sparrow  densities  reported  in  other 
studies  ranged  up  to  8.78  individuals/ha 
(Mowbray  1997)  and  are  considerably  lower 
than  our  densities  in  several  vegetation  types. 
We  observed  a significant  increase  in  Swamp 
Sparrow  density  between  mid-  and  late  peri- 
ods, which  may  be  explained,  in  part,  by  the 
addition  of  juveniles  from  early  nests  (Peck 
and  James  1987,  Beaver  1991,  Mowbray 
1997).  Swamp  Sparrows  prefer  open  wetlands 
of  sedges,  grasses  (i.e.,  wet  meadow),  and  cat- 
tail during  the  breeding  season  (Beaver  1991, 
Mowbray  1997).  Principally,  loosestrife  oc- 
curs in  the  wet  meadow,  strand,  and  emergent 
portions  of  a typical  wetland  profile,  which 
are  the  areas  where  Swamp  Sparrows  reach 
their  highest  abundance  (Beaver  1991,  Mow- 
bray 1997). 

Nesting  female  and  young  Swamp  Spar- 
rows satisfy  their  high  protein  requirements 
by  consuming  invertebrates.  Wetherbee 
(1968)  determined  that  88%  of  Swamp  Spar- 
row diets  during  spring  and  early  summer 
consisted  of  insects.  Arroll  (1995)  found  that 
aquatic  invertebrate  abundance  in  loosestrife 
in  central  Washington  was  similar  to  that  in 
cattail  and  bulrush.  Arroll  (1995)  found  only 


112 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  1,  March  1999 


TABLE  5 
in  Saginaw 

Mean  density  (no./ha)  ± 
Bay  wetlands,  1994-1995. 

SE  of  the  six  most  commonly  observed  bird  species  by  vegetation  type 

Vegetation  type^ 

Sedge  Wren 

Marsh  Wren*’ 

1994 

1995 

Period  2’’ 

Period  S*" 

Period  2*’ 

Period  3^ 

SS 

0 c 

0 c 

0 B 

0 B 

0 c 

WM/SS 

6.7  ± 1.6  B 

3.0  ± 1.6  BC 

0 B 

10.0  ± 1.9  A 

0 c 

WM/SS/LS 

10.0  ± 1.9  AB 

0 C 

6.7  ± 2.2  A^ 

3.3  ± 2.2  B 

0 c 

WM/LS 

0 C 

6.7  ± 2.2  B-^ 

0 B 

0 B 

0 c 

WM 

13.3  ± 2.2  A 

13.3  ± 1.6  A 

3.3  ± 1.6  AB 

3.3  ± 1.6  B 

0.5  ± 1.9  C 

IC 

0 C 

0 C 

0 B 

0 B 

15  .6  ± 1.8  B 

CC 

0 C 

0 C 

0 B 

0 B 

20.6  ± 1.2  A 

CB 

0 C 

0 C 

0 B 

0 B 

0.5  ± 2.0  C 

“ Vegetation  types:  SS  = scrub-shrub.  WM/SS  = wet  meadow/scrub-shrub.  WM/SS/LS  = wet  meadow/scrub-shrub/loosestrife,  WM/LS  = wet  meadow/ 
loosestrife.  WM  = wet  meadow.  1C  = inland  cattail.  CC  = coastal  cattail,  CB  = coa.stal  bulrush. 

^ Means  within  columns  followed  by  the  same  letter  do  not  differ  (P  > 0.05)  as  determined  by  ANOVA  and  Fisher's  least  significant  difference. 

Fisher's  least  significant  difference  multiple  comparisons  were  not  corroborated  by  90%  confidence  interval  comparisons  with  WM/SS/LS  and  IC  (see 
METHODS). 

Fisher's  least  significant  difference  multiple  compari.son  were  not  corroborated  by  90%  confidence  interval  comparisons  with  WM/SS  and  WM/LS  (see 
METHODS). 

' Fisher's  lea.st  significant  difference  multiple  comparison  were  not  corroborated  by  90%  confidence  interval  comparisons  with  WM/LS  (see  METHODS). 

Fisher's  lea.st  significant  difference  multiple  comparisons  were  not  corroborated  by  90%  confidence  interval  compari.son  with  WM/SS.  WM/LS,  and 
CB  (.see  METHODS). 

8 Fisher's  lea.st  significant  difference  multiple  comparisons  were  not  corroborated  by  90%  confidence  interval  compari.son  with  WM/SS.  WM/LS.  and 
CB  (see  METHODS). 


nine  statistically  significant  results  in  1 1 1 in- 
dividual comparisons  of  aquatic  invertebrates 
associated  with  macrophyte  stems  (using  stem 
vacuum),  sediment  (using  sediment  core),  and 
the  water  column  (using  activity  traps).  Of  the 
four  statistically  different  comparisons  involv- 
ing loosestrife,  two  showed  higher  Diptera 
and  Ostracoda  abundance  in  cattail  compared 
with  loosestrife,  and  two  showed  higher  Co- 
pepod  abundance  in  loosestrife  compared  with 
cattail  (Arroll  1995).  Thus,  invertebrate  food 
items  during  the  breeding  season  do  not  ap- 
pear limiting  in  loosestrife  habitat,  although 
quantitative  data  from  the  Northeast  are  need- 
ed. 

Loosestrife  is  an  anathema  to  wetland  man- 
agers because  it  often  replaces  seed-producing 
mudflat  species  managed  to  attract  waterfowl. 
Water  level  manipulations  such  as  early  sea- 
son drawdowns  encourage  loosestrife  estab- 
lishment (Thompson  1989).  Loosestrife  forms 
dense  stands  that  are  difficult  for  some  bird 
species  to  negotiate  and  this  may  be  especially 
true  for  larger  birds  such  as  waterfowl  or  spe- 
cies that  walk  on  the  ground  such  as  bitterns 
and  rails.  Our  study  demonstrates  that  loose- 
strife may  provide  suitable  habitat  for  some 
passerines. 

Many  researchers  have  observed  that  habi- 
tat diversity  leads  to  faunal  diversity  in  wet- 


lands (Weller  and  Spatcher  1965,  Weller  and 
Fredrickson  1974,  Weller  1978,  Kantrud  and 
Stewart  1984,  Burger  1985).  The  highest  avi- 
an density,  diversity,  and  productivity  in 
marshes  occurs  where  emergent  vegetation  is 
interspersed  1:1  with  open  water  (Weller  and 
Spatcher  1965,  Weller  and  Fredrickson  1974, 
Fredrickson  and  Reid  1988).  Wetland  man- 
agers manipulate  vegetative  interspersion  in 
marshes  using  artificial  drawdowns,  muskrat 
management,  and  other  means  (Fredrickson 
and  Reid  1988).  Kaminski  and  Prince  (1981) 
observed  increased  waterfowl  density  and  di- 
versity coincident  with  increased  abundance, 
biomass,  and  diversity  of  macroinvertebrates 
in  manipulated  emergent  wetland  habitat.  Our 
loosestrife  sites  contained  few  openings.  We 
suspect  that  manipulated  loosestrife  habitat  (to 
create  interspersion)  could  result  in  higher 
bird  diversity. 

Loosestrife  was  widespread  in  Saginaw  Bay 
coastal  wetlands  and  dominated  canopies  at 
several  sites.  Although  diversity  was  low, 
loosestrife  provided  nesting  and  brood  rearing 
habitat  to  birds  in  Saginaw  Bay  wetlands 
where  alternative  habitat  choices  were  avail- 
able. Some  species,  such  as  Marsh  Wren,  may 
be  disadvantaged  as  loosestrife  displaces  other 
plant  forms  (e.g.,  cattail  and  bulrush).  Swamp 
Sparrows  may  prefer  loosestrife  habitat  where 


Whin  et  III.  • AVIAN  USE  OF  PURPLE  LOOSESTRIFE 


113 


TABLE  5.  Extended. 

Yellow  Warbler 

Common 

Yellowihroat*’ 

Swamp 

Sparrow^’ 

Red-winged  Blackbird 

Period  2*' 

Period  .2*’ 

Period  2*’ 

Period  3*’ 

9.3 

± 1.0  A 

14.7  ± 1.0  A 

2.7  ± 0.6  B' 

1 1.3  ± 1.8  C 

17.3  ± 1.8  A 

6.0 

± 1.8  A? 

3.3 

± 1.3  B 

2.0  ± 1.2  B 

5.3  ± 0.8  A 

21.5  ± 2.3  B 

0 C 

0 B 

2.8 

± 1.4  B 

1.7  ± 1.6  B 

1.5  ± 1.0  BC 

27.8  ± 2.8  B 

7.1  ± 2.7  Bf 

1.7 

± 2.9  AB 

0 B 

0 B 

0 C 

44.2  ± 2.9  A 

0 C 

0 B 

0 B 

0 B 

1.4  ± 0.8  BC 

21.2  ± 2.2  B 

6.7  ± 2.4  B' 

4.2 

± 2.0  AB 

0.7 

± 1.0  B 

0 B 

0 C 

9.6  ± 2.1  C 

5.7  ± 1.9  BC 

4.4 

± 2.4  AB 

0 B 

0 B 

0.2  ± 0.5  C 

1.5  ± 1.4  D 

1.8  ± 1.4  BC 

1.2 

± 1.4  B 

0 B 

0 B 

0 C 

0 D 

0 C 

0 B 

nest-building  materials  (fine-stemmed  grasses 
and  sedges)  are  available.  We  conclude  that 
avian  use  of  loosestrife  warrants  further  quan- 
titative investigation  because  avian  use  may 
be  higher  than  is  commonly  believed. 

ACKNOWLEDGMENTS 

The  Michigan  Agricultural  Experiment  Station  and 
the  Michigan  Department  of  Natural  Resources  pro- 
vided funding.  T.  M.  Burton  and  J.  Burley,  both  of 
Michigan  State  University,  provided  advice  and  guid- 
ance. L.  D.  Igl  and  D.  H.  Johnson,  both  of  Northern 
Prairie  Wildlife  Research  Center  (NPWRC)  in  James- 
town, North  Dakota,  W.  Scharf,  E.  Kiviat,  and  two 
anonymous  reviewers  provided  comments  on  drafts.  D, 
H.  Johnson,  W.  E.  Newton,  and  G.  A.  Sargeant  (all  of 
NPWRC)  provided  statistical  advice.  L.  A.  Jagger  and 
D.  Ford  assisted  in  field  observations. 

LITERATURE  CITED 

Arroll,  S.  G.  1995.  Effects  of  purple  loosestrife 
(Lythnim  salicaria)  and  its  control  on  wetlands  in 
central  Washington  state.  Ph.D.  diss.,  Univ.  of 
Washington,  Seattle. 

Beaver,  D.  L.  1991.  Swamp  Sparrow.  Pp.  486—487  in 
The  atlas  of  breeding  birds  of  Michigan  (R.  Brew- 
er, G.  A.  McPeek,  and  R.  L.  Adams,  Jr.,  Eds.). 
Michigan  State  Univ.  Press,  East  Lansing. 

Brown,  M.  and  J.  J.  Dinsmore.  1986.  Implications  of 
marsh  size  and  isolation  for  marsh  bird  manage- 
ment. J.  Wildl.  Manage.  50:392-397. 

Burger,  J.  1985.  Habitat  selection  in  temperate  marsh- 
nesting birds.  Pp.  253-281  in  Habitat  selection  in 
birds  (M.  L.  Cody,  Ed.).  Academic  Press,  Orlan- 
do, Florida. 

Burns,  J.  T.  1982.  Nests,  territories,  and  reproduction 
of  Sedge  Wrens  {Cistothonis  platen.'ii.s).  Wilson 
Bull.  94:338-349. 

Comer,  P.  J.  1996.  Wetland  trends  in  Michigan  since 
1800:  a preliminary  assessment.  U.S.  Environ- 
mental Protection  Agency,  Water  Division  and 
Wildlife  Division,  Michigan  Department  of  Nat- 


ural Resources.  Michigan  Natural  Features  Inven- 
tory, Lansing. 

CowARDiN,  L.  M.,  V.  Carter,  F C.  Golet,  and  E.  T. 
LaRoe.  1979.  Classification  of  wetlands  and  deep- 
water habitats  of  the  United  States.  U.S.  Dept.  In- 
terior, FWS/OBS-79/3 1:1-103. 

Dahl,  T.  E.  1990.  Wetland  losses  in  the  United  States: 
1780s  to  1980s.  U.S.  Dept.  Interior,  U.S.  Fish 
Wildl.  Serv.,  Washington,  D.C. 

Fredrickson,  L.  H.  and  F.  A.  Reid.  1988.  Preliminary 
considerations  for  manipulating  vegetation.  U.S. 
Fish  Wildl.  Serv.,  leaflet  13.4.9. 

Higgins,  K.  F,  J.  L.  Oldemeyer,  K.  J.  Jenkins,  G.  K. 
Clambey,  and  R.  F.  Harlow.  1994.  Vegetation 
sampling  and  methods.  Pp.  567-591  in  Research 
and  management  techniques  for  wildlife  and  hab- 
itats, fifth  ed.  The  Wildlife  Society,  Bethesda,  Mary- 
land. 

Johnson,  R.  R.,  B.  T.  Brown,  L.  T.  Haight,  and  J.  M. 
Simpson.  1981.  Playback  recordings  as  a special 
avian  censusing  technique.  Stud.  Avian  Biol.  6: 
68-75. 

Johnson.  R.  R.  and  J.  J.  Dinsmore.  1986.  The  use  of 
tape-recorded  calls  to  count  Virginia  Rails  and  So- 
ras.  Wilson  Bull.  98:303-306. 

Kaminski,  R.  M.  and  H.  H.  Prince.  1981.  Dabbling 
duck  and  aquatic  macroinvertebrate  responses  to 
manipulated  wetland  habitat.  J.  Wildl.  Manage. 
45:1-15. 

Kantrud,  H.  a.  and  R.  E.  Stewart.  1984.  Ecological 
distribution  and  crude  density  of  breeding  birds 
on  prairie  wetlands.  J.  Wildl.  Manage.  48:426- 
437. 

Keddy,  P.  a.  and  a.  a.  Reznicek.  1985.  Vegetation 
dynamics,  buried  .seeds,  and  water  level  fluctua- 
tions on  the  shorelines  of  the  Great  Lakes.  Pp.  33- 
58  in  Coastal  wetlands  (H.  H.  Prince  and  F M. 
DTtri,  Eds.).  Lewis  Publishers,  Chelsea.  Michi- 
gan. 

Kiviat,  E.  1996.  American  Goldfinch  nests  in  purple 
loosestrife.  Wilson  Bull.  108:182-186. 

Kroodsma,  D.  and  j.  Verner.  1978.  Complex  singing 
behavior  among  Cistothoni.s  wrens.  Auk  95:703- 
716. 


114 


THE  WILSON  BULLETIN  • Vol.  Ul,  No.  I,  March  1999 


Lanyon,  W.  E.  1997.  Great  Crested  Flycatcher  (Myiar- 
chus  crinitiis).  In  The  birds  of  North  America,  no. 
300  (A.  Poole  and  E Gill,  Eds.).  The  Academy  of 
Natural  Sciences,  Philadelphia,  Pennsylvania;  The 
American  Ornithologists’  Union,  Washington, 
D.C. 

Lowther,  P.  E.  1993.  Brown-headed  Cowbird  (Mol- 
othriis  ater).  In  The  birds  of  North  America,  no. 
47  (A.  Poole  and  F.  Gill,  Eds.).  The  Academy  of 
Natural  Sciences,  Philadelphia,  Pennsylvania;  The 
American  Ornithologists’  Union,  Washington, 
DC. 

Malecki,  R.  a.,  B.  Blossey,  S.  D.  Hight,  D.  Schroe- 
DER,  L.  T.  Kok,  and  J.  R.  Coulson.  1993.  Bio- 
logical control  of  purple  loosestrife.  BioScience 
43:680-686. 

Marion,  W.  R.,  T.  E.  O’Meara,  and  D.  S.  Maehr. 
1981.  Use  of  playback  recordings  in  sampling  elu- 
sive or  secretive  birds.  Stud.  Avian  Biol.  6:81-85. 

McCarty,  J.  P.  1996.  Eastern  Wood-pewee  {Contopu.s 
virens).  In  The  birds  of  North  America,  no.  245 
(A.  Poole  and  F.  Gill,  Eds.).  The  Academy  of  Nat- 
ural Sciences,  Philadelphia,  Pennsylvania;  The 
American  Ornithologists’  Union,  Washington, 
DC. 

Milliken,  G.  a.  and  D.  E.  Johnson.  1984.  Analysis 
of  messy  data,  vol.  1:  designed  experiments.  Van 
Nostrand  Reinhold  Co.,  New  York. 

Moore,  W.  S.  1995.  Northern  Flicker  (Colapte.s'  aii- 
ratiis).  In  The  birds  of  North  America,  no.  166 
(A.  Poole  and  E Gill,  Eds.).  The  Academy  of  Nat- 
ural Sciences,  Philadelphia,  Pennsylvania;  The 
American  Ornithologists’  Union,  Washington, 
DC. 

Mowbray,  T.  B.  1997.  Swamp  Sparrow  (Melo.spiza 
georgiana).  In  The  birds  of  North  America,  no. 
279  (A.  Poole  and  E Gill,  Eds.).  The  Academy  of 
Natural  Sciences,  Philadelphia,  Pennsylvania;  The 
American  Ornithologists’  Union,  Washington, 
D.C. 

Peck,  G.  and  R.  James.  1987.  Breeding  birds  of  On- 
tario: nidiology  and  distribution.  Vol.  2.  Passer- 
ines. Royal  Ontario  Museum,  Toronto. 

PETER.SON,  R.  T.  1990.  Peterson's  held  guides,  eastern/ 
central  bird  songs.  Houghton  Mifflin  Co.,  Boston, 
Massachusetts. 

Prince,  H.  H.  and  C.  S.  Flegel.  1995.  Breeding  avi- 


fauna of  Lake  Huron.  Pp.  247—272  in  The  Lake 
Huron  ecosystem:  ecology,  hsheries  and  manage- 
ment (M.  Munwar,  T.  Edsall,  and  J.  Leach,  Eds.). 

5. P.B.  Academic  Publishing,  Amsterdam,  The 
Netherlands. 

Rawinski,  T.  j.  and  R.  A.  Malecki.  1984.  Ecological 
relationships  among  puiple  loosestrife,  cattail,  and 
wildlife  at  the  Montezuma  National  Wildlife  Ref- 
uge. N.Y.  Fish  Game  J.  31:81-87. 

Reinert,  S.  E.  and  E C.  Golet.  1986.  Breeding  ecol- 
ogy of  the  Swamp  Sparrow  in  a southern  Rhode 
Island  peatland.  Northeast  Sect.  Wildl.  Soc.  1986: 
1-13. 

SAS  Institute.  1990.  SAS/STAT  user’s  guide.  Version 

6,  fourth  ed.  Volume  2.  SAS  Institute,  Cary,  North 
Carolina. 

Skinner,  L.  C.,  W.  J.  Rerdall,  and  E.  L.  Fuge.  1994. 
Minnesota’s  purple  loosestrife  program:  history, 
hndings,  and  management  recommendations. 
Minnesota  Dept.  Nat.  Resour.  Spec.  Publ.  145:1  — 
36. 

SOKAL,  R.  R.  AND  E J.  Rohlf.  1981.  Biometry,  second 
ed.  W.  H.  Freeman  and  Company,  New  York. 
Swift,  B.  L.,  S.  R.  Orman,  and  J.  W.  Ozard.  1988. 
Response  of  Least  Bitterns  to  tape-recorded  calls. 
Wilson  Bull.  100:496-499. 

Thompson,  D.  Q.  1989.  Control  of  purple  loosestrife. 

U.S.  Fish  Wildl.  Serv.,  leaflet  13.4.11. 

Thompson,  D.  Q.,  R.  L.  Stuckey,  and  E.  B.  Thomp- 
son. 1987.  Spread,  impact,  and  control  of  puiple 
loosestrife.  U.S.  Fish  Wildl.  Serv.,  Fish  Wildl. 
Res.  2:1-55. 

Wetherbee,  D.  K.  1968.  Southern  Swamp  Sparrow. 

U.S.  Natl.  Mus.  Bull.  237:1475-1490. 

Weller,  M.  W.  1978.  Management  of  freshwater  wet- 
lands for  wildlife.  Pp.  267-284  in  Freshwater  wet- 
lands (R.  E.  Good,  D.  E Whigham,  and  R.  L. 
Simpson,  Eds.).  Academic  Press,  New  York. 
Weller,  M.  W.  and  C.  E.  Spatcher.  1965.  Role  of 
habitat  in  the  distribution  and  abundance  of  marsh 
birds.  Iowa  State  Univ.  Agr.  Home  Econ.  Exp. 
Sta.,  Spec.  Rept.  43:1-31. 

Weller,  M.  W.  and  L.  H.  Fredrickson.  1974.  Avian 
ecology  of  a managed  glacial  marsh.  Living  Bird 
12:269-291. 

Wilcox,  D.  A.  1995.  Wetland  and  aquatic  macrophytes 
as  indicators  of  anthropogenic  disturbance.  Nat. 
Areas  J.  15:240-248. 


Short  Communications 


Wilson  Bull.,  111(1),  1999,  pp.  115-116 


Bald  Eagle  Predation  on  Common  Loon  Chick 


James  D.  Paruk,'-^  Dean  Seanfieldr  and  Tara  Mack^ 


ABSTRACT. — We  report  predation  of  a Common 
Loon  (Gavia  immer)  chick  by  an  adult  Bald  Eagle 
(Haliaeetiis  leucocephaliis)  in  northern  Wisconsin.  Re- 
ceived 27  Feb.  1998,  accepted  26  Sept.  1998. 


Common  Loon  (Gavia  immer)  chicks  may 
be  vulnerable  to  aerial  and  underwater  pred- 
ators, including  Bald  Eagle  (Haliaeetus  leu- 
cocephalus).  Common  Raven  (Corvus  corax). 
Herring  Gull  (Larus  argentatus),  snapping 
turtle  (Chelydra  serpentina),  northern  pike 
(Esox  niger),  and  muskellunge  (Esox  masqui- 
nongy,  Yonge  1981,  McIntyre  1988,  J.  Wilson 
and  M.  Meyer,  pers.  comm.).  However,  few 
observers  have  actually  observed  such  pre- 
dation (McIntyre  1988).  Here  we  describe  the 
first  documented  observation  of  a Bald  Eagle 
killing  a loon  chick. 

On  26  July  1996,  while  collecting  data  on 
parental  effort  on  a pair  of  color-marked  adult 
Common  Loons  at  the  Turtle  Flambeau  Flow- 
age  in  north  central  Wisconsin  (46°  0'  N, 
90°  10'  W),  we  observed  an  adult  Bald  Eagle 
capture  a 15  day-old  loon  chick.  Loon  chicks 
are  most  prone  to  predation  during  the  first 
two  weeks  after  hatching,  but  continue  to  be 
vulnerable  up  to  4-5  weeks  of  age  (Yonge 
1981,  McIntyre  1983,  pers.  obs.).  At  10:00 
CST  an  adult  eagle  circled  the  territory  sev- 
eral times  before  perching  near  the  top  a lone, 
8 m tall  tamarack  (Larix  laricina)  on  a small 
island  (0.1  ha).  An  adult  eagle(s)  had  been 
observed  within  the  loons’  territory  on  2 pre- 
vious occasions  (19,  22  July).  When  the  rest- 
ing adult  loons  saw  the  eagle,  they  wailed  and 


' Dept,  of  Biological  Sciences,  Idaho  State  Univ., 
Pocatello,  ID  83209. 

^ Earthwatch,  680  Mt.  Auburn  St.,  Watertown,  MA 
02272. 

^ Sigurd  Olson  Env.  Institute,  Northland  College, 
Ashland,  WI  54806. 

Corresponding  author;  E-mail;  parujame@isu.edu 


tremoloed  several  times.  The  chick,  initially 
with  the  adults,  disappeared  from  our  sight, 
apparently  beneath  some  overhanging  alder 
(Alnus  sp.)  along  the  island’s  edge.  By  10:20 
the  adult  loons  stopped  vocalizing;  the  male 
resumed  foraging  dives,  and  the  female  re- 
mained alert  on  the  surface.  However,  at 
10:30,  two  non-resident  loons  intruded  on  the 
established  pair’s  territory,  triggering  aggres- 
sive interactions  for  10-12  min  until  the  in- 
truding pair  left  the  area.  The  territorial  adults 
remained  alert  on  the  surface  with  the  chick 
still  out  of  sight. 

At  10:58  the  eagle  swooped  down  and 
grabbed  something  offshore  near  the  island 
(because  of  the  density  of  the  alder  we  could 
not  positively  identify  it).  The  adult  loons  im- 
mediately started  giving  3-note  wails  and  3- 
note  tremolos,  and  the  male  let  out  several 
yodels.  At  1 1 :00  the  eagle  attempted  to  fly 
with  the  loon  chick  (now  visible)  in  its  talons, 
but  had  trouble  becoming  airborne  in  the  alder 
thicket.  Thirty  seconds  later,  the  eagle  attained 
flight  without  the  chick  in  its  talons  and  re- 
turned to  its  previous  perch  in  the  tamarack. 
It  remained  there  for  10  min  before  it  flew 
from  the  area.  The  adult  loons  continued  to 
wail  and  tremolo  for  20  min  while  swimming 
around  the  island.  They  remained  in  the  area 
until  12:25,  then  swam  about  300  m to  the 
other  side  of  their  territory.  We  searched  for 
the  chick  3 hours  later  and  found  it  dead,  with 
puncture  wounds  to  the  head  and  a crushed 
skull.  We  skinned  and  mounted  the  chick  as  a 
study  specimen,  but  did  not  record  its  body 
mass. 

ACKNOWLEDGMENTS 

We  thank  Earthwatch  and  Biodiversity  Research  In- 
stitute for  financial  support,  and  both  Earthwatch  and 
Loon  Watch  for  providing  volunteers.  Jeff  Wilson  pro- 
vided logistical  and  moral  support  throughout  the  field 
sea,son.  Additionally,  we  thank  D.  Gorde,  T.  Gerstell 


115 


116 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  1,  March  1999 


and  two  anonymous  reviewers  for  improving  the  man- 
uscript. 

LITERATURE  CITED 

McIntyre,  J.  W.  1983.  Nurseries:  a consideration  of 
habitat  requirements  during  the  early  chick-rear- 
ing period  in  Common  Loons.  J.  Field  Omithol. 
54:247-253. 


McIntyre,  J.  W.  1988.  The  Common  Loon:  spirit  of 
northern  lakes.  Univ.  of  Minnesota  Press,  Min- 
neapolis. 

Yonge,  K.  S.  1981.  The  breeding  cycle  and  annual 
production  of  the  Common  Loon  (Govia  immer) 
in  the  boreal  forest  region.  M.S.  thesis,  Univ.  of 
Manitoba,  Winnipeg. 


Wilson  Bull.,  111(1),  1999,  pp.  116—117 


Territorial  takeover  in  Common  Loons  {Gavia  immer) 

James  D.  Paruk' 


ABSTRACT. — Breeding  Common  Loons  (Gavia 
immer)  are  well  known  for  vigorously  defending  their 
territory  from  conspecihcs.  Territory  holders  are  not 
previously  known  to  be  supplanted  by  loons  during  the 
breeding  season.  I observed  a pair  of  adult  Common 
Loons  displace  a resident  pair  from  their  territory 
shortly  after  a territorial  conflict;  the  takeover  coincid- 
ed with  the  death  of  the  resident  loons’  chick  caused 
by  an  adult  Bald  Eagle  (Haliaeetus  leucocephalu.s). 
Received  27  Feb.  1998,  accepted  7 Sept.  1998. 


Common  Loons  (Gavia  immer)  are  philo- 
patric,  but  territory  switching  occurs  infre- 
quently (Piper  et  al.  1997;  Evers,  Reaman, 
Kaplan,  and  Paruk,  unpubl.  data).  Our  under- 
standing of  territory  switching  in  Common 
Loons  remains  largely  unknown  (Piper  et  al. 
1997).  In  1995—1997,  while  coordinating 
studies  of  parental  effort  and  social  flocking 
in  Common  Loons  at  the  Turtle  Flambeau 
Flowage  in  northern  Wisconsin  (46°  00'  N, 
90°  10'  W),  I observed  a territory  takeover. 

The  Turtle  Flambeau  Flowage  is  a large  im- 
poundment (5798  ha)  that  contains  24-26 
loon  territories.  Territories  are  generally  well 
delineated  by  coves  or  islands,  but  there  are 
several  places  where  the  presence  of  small  is- 
lands makes  it  difficult  to  distinguish  individ- 
ual territorial  boundaries. 

On  26  July  1996,  at  10:30  CST  a pair  of 
unbanded  loons  entered  the  territory  (Long  Is- 


'Dcpt.  of  Biological  Sciences,  Idaho  State  Univ.,  Po- 
catello, ID  83209;  E-mail:  parujame@isu.edu. 


land,  LI)  of  an  established  color-banded  pair. 
All  four  birds  moved  behind  an  island  and 
were  out  of  sight  for  2 min.  A territorial  bird 
(sex  could  not  be  determined  because  of  an 
obstruction)  “surface  rushed”  one  of  the  in- 
truders driving  it  onto  an  island  with  repeated 
bill  thrusts.  At  10:42,  the  territorial  pair  was 
reunited  and  the  intruders  were  no  longer  in 
sight. 

At  10:58,  an  adult  Bald  Eagle  (Haliaeetus 
leucocephalus)  killed  the  resident  pairs’  15 
day-old  chick  (Paruk  et  al.  1999).  At  this  age, 
loon  chicks  are  still  dependent  upon  adults  for 
food  and  protection  from  predators  (Dulin 
1988,  McIntyre  1988).  The  LI  pair  remained 
near  the  location  where  the  eagle  killed  the 
chick  until  12:25,  at  which  time  they  swam  to 
the  other  side  of  their  territory.  I searched  the 
territory  for  the  banded  loons  from  15:25- 
16:00,  but  did  not  observe  them.  Instead,  I 
observed  two  unbanded  loons  in  the  territory. 

On  27  July,  I observed  two  unbanded, 
paired  adults  in  the  LI  territory.  The  original 
pair  was  not  observed  on  27  July,  but  on  28 
July  an  assistant  spotted  them  several  hundred 
meters  south  of  their  former  territory. 
Throughout  the  rest  of  the  summer,  until  18 
August,  several  observations  (n  = 9)  con- 
firmed that  the  original  territorial  pair  had 
been  supplanted  by  two  unbanded  individuals. 
The  unbanded  birds  were  observed  foraging, 
resting  and  preening  in  the  LI  territory  until 
the  end  of  the  observation  period  on  18  Au- 
gust. 

Prior  to  the  takeover,  a pair  of  unbanded 


SHORT  COMMUNICATIONS 


loons  engaged  the  territorial  pair  on  four  con- 
secutive days  (22-25  July),  with  several  ago- 
nistic encounters.  On  both  24  and  25  July  the 
LI  pair  left  their  chick  and  engaged  in  ritual- 
ized behavior  (jerk  diving,  facing  away)  with 
conspecifics  for  20  and  26  min  respectively. 
Whether  these  were  the  same  individuals  that 
took  over  the  territory  is  unknown.  It  seems 
likely  that  there  was  a territorial  dispute  prior 
to  the  death  of  the  LI  pair’s  chick  and  the 
subsequent  takeover. 

Conmion  Loons  do  not  typically  abandon 
their  territory  after  the  loss  of  a chick,  al- 
though they  will  often  show  less  aggression 
towards  conspecifics  and  may  wander  more 
frequently  than  loon  pairs  with  chicks  (Evers, 
pers.  comm.;  pers.  obs.).  Thus,  it  is  unlikely 
that  the  resident  LI  pair  simply  abandoned 
their  territory  after  the  loss  of  their  chick. 

In  1997,  the  former  LI  pair  remained  to- 
gether and  occupied  a new  territory  400  m 
south  of  their  original  territory.  Two  unbanded 
loons  nested  in  the  former  LI  territory.  Zack 
and  Stutchbury  (1992)  proposed  that  non- 
breeders are  likely  to  acquire  territories  they 
visit  frequently  and  Piper  and  coworkers 
(1997)  proposed  that  mid-  to  late  seasonal 
movements  observed  in  loons  may  be  partly 
explained  by  their  searching  for  new  or  un- 
occupied territories  (reconnaissance  hypothe- 
sis). The  lack  of  distinct  physical  barriers  sep- 
arating loon  territories  and  the  high  number  of 
nonbreeders  present  on  the  Turtle  Flambeau 
Flowage  (Belant  1989,  pers.  obs.)  may  result 
in  higher  intrusion  rates  and  more  interterri- 
torial interaction,  lowering  territorial  stability 
(Strong  and  Bissonette  1988;  Belant  1991; 
Piper  et  al.  1997;  Evers  et  al.,  unpubl.  data). 
The  timing  of  the  observed  supplanting/take- 
over supports  the  reconnaissance  hypothesis 
for  loon  movements  during  mid-  to  late  sum- 


1 17 

mer,  and  suggests  that  Common  Loons  may 
actively  engage  in  territory  acquisition  during 
the  breeding  season.  To  what  extent  this  take- 
over was  precipitated  by  the  death  of  the  res- 
ident pair’s  chick  remains  unknown. 

ACKNOWLEDGMENTS 

The  behavioral  project,  of  which  this  observation 
was  a part,  was  funded  by  Earthwatch  and  Biodiversity 
Research  Institute.  I am  particularly  indebted  to  ail  the 
volunteers  for  their  assistance  in  gathering  the  data,  but 
e.specially  so  to  R Hart,  M.  Lockman,  T.  Mack,  D. 
Seefeldt,  A.  Turpen  and  M.  Wiranowski.  J.  Wilson  pro- 
vided logistical  and  moral  support  throughout  the  in- 
vestigation. I also  thank  T.  Ford,  T.  Gerstell,  J.  Mc- 
Intyre and  an  anonymous  reviewer  for  their  comments 
and  suggestions  on  this  manuscript. 

LITERATURE  CITED 

Belant,  J.  L.  1989.  Common  Loon  productivity  and 
brood  habitat  use  in  northern  Wisconsin.  M.S.  the- 
sis, Univ.  of  Wisconsin,  Stevens  Point. 

Belant,  J.  L.  1991.  Territorial  activities  of  Common 
Loons  on  single-pair  lakes.  Passenger  Pigeon.  54; 
115-118. 

Dulin,  G.  S.  1988.  Pre-fledging  feeding  behavior  and 
sibling  rivalry  in  Common  Loons.  M.S.  thesis. 
Central  Michigan  Univ.,  Mt.  Pleasant. 

McIntyre,  J.  W.  1988.  The  Common  Loon:  spirit  of 
northern  lakes.  Univ.  of  Minnesota  Press,  Min- 
neapolis. 

Paruk,  j.  D.,  D.  Seanfield,  and  T.  Mack.  1999.  Bald 
Eagle  predation  on  Common  Loon  chick.  Wilson 
Bull.  111:116-117. 

Piper,  W.  H.,  J.  D.  Paruk,  D.  C.  Evers,  M.  W.  Meyer, 
K.  B.  Tischler,  M.  Klich,  and  J.  J.  Hartigan. 
1997.  Local  movements  of  color-marked  Common 
Loons.  J.  Wildl.  Manage.  61:1253-1261. 

Strong,  P.  and  J.  Bissonette.  1988.  Territorial  activ- 
ities of  Common  Loons  on  multiple-pair  lakes.  Pp. 
19-24  in  Papers  from  1987  conference  on  loon 
research  and  management  (P.  Strong,  Ed.).  North 
American  Loon  Fund,  Meredith,  New  Hampshire. 
Zack,  S.  and  B.  J.  Stutchbury.  1992.  Delayed  breeding 
in  avian  social  systems:  tlie  role  of  territory  quality 
and  “floater”  tactics.  Behaviour  123:195-219. 


118 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


Wilson  Bull.,  1 1 1(1),  1999,  pp.  118-119 


Courtship  Behavior  of  the  Buff-necked  Ibis  (Theristicus  caudatus) 

Nathan  H.  Rice 


ABSTRACT — Buff-necked  Ibis  (Theristicus  cau- 
datus) courtship  displays  include  “Bill  Popping”  and 
grasping  twigs,  behaviors  reported  for  other  species  of 
ibis.  Received  4 Dec.  1997,  accepted  20  Aug.  1998. 


The  courtship  displays  of  the  ibises  (Thres- 
kiornithidae)  are  poorly  known  and  little  doc- 
umented, including  those  of  the  South  Amer- 
ican Buff-necked  Ibis  (Theristicus  caudatus-, 
del  Hoyo  et  al.  1992,  Hancock  et  al.  1992). 
Here,  1 present  field  observations  of  apparent 
courtship  displays  by  this  species  made  at 
Parque  Nacional  San  Luis  (22°  40'  S,  57°  21' 
W),  depto.  Concepcion,  on  25  October  1996, 
in  the  semihumid  forest/savanna  region  of 
northern  Paraguay. 

At  approximately  09:00  I heard  and  ob- 
served two  Buff-necked  Ibises  at  the  top  of  a 
dead  tree,  approximately  30  m above  the 
ground  making  snapping  noises  with  their 
bills.  The  presumed  male  and  female  would, 
in  turn,  grasp  dead  twigs  (2-3  cm  diameter) 
with  their  bills  and  release  them,  never  break- 
ing the  twigs  from  the  branch.  Occasionally, 
the  birds  would  make  low  grunts.  This  con- 
tinued for  about  3 min  until  the  birds  abruptly 
stopped,  faced  each  other,  pointed  their  bills 
vertically  (similar  to  photo  in  del  Hoyo  et  al. 
1992:480)  and  gave  a loud  squawking  call. 
Then  they  slapped  their  bills  together  horizon- 
tally, making  a sound  similar  to  two  hollow 
pieces  of  bamboo  hitting  each  other,  and  flew 
away  together  to  the  northwest. 

About  30  min  later,  approximately  1.5  km 
northwest  of  the  first  site,  I heard  the  same 
snapping  sound  and  quickly  located  a pair  of 
ibises,  perhaps  the  same  birds,  displaying  in  a 
similar  manner.  Again,  the  behavior  lasted 
about  3 min  and  ended  with  the  birds  slapping 
their  heads  together  and  flying  away.  On  30 


Natural  History  Museum.  Univ.  of  Kansas, 
Lawrence,  K.S  6604.5; 

E-mail:  nricc@falcon.cc.ukans.edu 


October  1997,  a male  (testes  18  X 11  mm, 
KU  #88342)  and  female  (ovary  20  X 14  mm, 
largest  ovum  3X3  mm,  oviduct  convoluted 
3 mm,  specimen  deposited  in  Museo  Nacional 
Historia  Natural  del  Paraguay)  were  collected 
that  may  be  the  pair  I observed  earlier.  Based 
on  the  specimen  gonad  sizes,  the  birds  were 
in  breeding  condition. 

The  behaviors  I observed  resemble  court- 
ship and  breeding  displays  of  other  ibises  (del 
Hoyo  et  al.  1992,  Hancock  et  al.  1992).  Un- 
mated male  ibises  use  Bill  Popping  (Hancock 
et  al.  1992)  when  soliciting  females.  This  be- 
havior involves  the  bird  snapping  its  gaped 
bill  shut,  occasionally  making  a popping 
sound.  Some  ibis  grab  a twig  and  shake  it  dur- 
ing this  behavior.  The  initial  “snapping”  noise 
that  I heard  may  have  been  a product  of  Bill 
Popping  by  the  male.  Although  the  pair  I ob- 
served never  removed  twigs  from  the  tree, 
they  did  grab  sticks  during  the  encounter — a 
further  indication  that  this  was  the  Bill  Pop- 
ping behavior  described  in  Hancock  and  co- 
workers (1992). 

Males  will  also  respond  to  females  entering 
their  territories  with  a ritualized  form  of  Spar- 
ring display  (Hancock  et  al.  1992).  Sparring 
behavior  consists  of  one  bird,  in  this  case  the 
male,  lunging  at  the  other.  The  female  will  flee 
and  not  fight  back.  Both  of  the  behavioral  se- 
ries I observed  ended  with  the  birds  confront- 
ing one  another  (i.e.,  slapping  bills  together) 
and  then  departing.  Perhaps  this  was  a modi- 
fication of  the  Sparring  behavior  described  in 
Hancock  et  al.  (1992). 

ACKNOWLEDGMENTS 

Director  O.  Romero  of  Departamento  del  Inventario 
Biologico  Nacional  and  Museo  Nacional  Historia  Nat- 
ural del  Paraguay  helped  in  innumerable  ways.  C.  Eox, 
director  of  Dirreccion  de  Barques  Nacionales  y Vida 
Silvestre  graciously  granted  permits  for  the  work  at 
San  Luis.  A.  L.  Aquino,  Director  of  CITES,  kindly 
provided  logistical  help  in  getting  to  San  Luis.  I thank 
San  Luis  park  guard,  A.  Acosta,  for  accommodating 


SHORT  COMMUNICATIONS 


1 19 


us.  Special  thanks  to  R.  Faucett,  M.  Robbins  and  J. 
Simmons  for  field  assistance.  This  manuscript  was  im- 
proved by  comments  from  A.  T.  Peterson,  M.  Robbins 
and  two  anonymous  reviewers.  This  work  was  partially 
funded  by  a Panorama  Society  grant  from  the  Univer- 
sity of  Kansas  Natural  History  Museum,  and  by  the 
generous  support  of  the  Ornithology  Interest  Group  at 
the  University  of  Kansas. 


LITERATURE  CITED 

DEL  Hoyo,  J.,  a.  Elliot,  and  J.  Sargatal.  (Eds.y 
1992.  Handbook  of  the  birds  of  the  world,  vol.  1. 
Lynx  Edicions,  Barcelona,  Spain. 

Hancock,  J.  A.,  J.  A.  Kushlan,  and  M.  P.  Kahl.  1992. 
Storks,  ibises,  and  spoonbills  of  the  world.  Aca- 
demic Press,  San  Diego,  California. 


Wilson  Bull,  111(1),  1999,  pp.  119-121 

Habitat  Use  by  Masked  Ducks  Along  the  Gulf  Coast  of  Texas 


James  T.  Anderson ‘ and  Thomas  C.  Tacha'-^ 


ABSTRACT. — We  counted  47  Masked  Ducks  {No- 
monyx  dominiciis)  in  seven  flocks  during  the  fall  and 
winter  of  1992-1993  on  1009  64.75-ha  plots  in  the 
Coastal  Plains  of  Texas.  Among  the  three  wetland  sub- 
classes used  by  Masked  Ducks,  bird  densities  were 
higher  on  lacustrine  littoral  aquatic-bed  rooted  vascular 
and  lacustrine  littoral  aquatic-bed  floating  vascular 
than  palustrine  scrub-shrub  broad-leaved  deciduous 
wetlands.  These  wetlands  provide  important  habitat 
even  though  they  are  not  the  most  abundant  wetlands 
in  the  region.  Received  23  June  1998,  accepted  25 
Aug.  1998. 


Masked  Ducks  (Nomonyx  dominicus)  are 
small,  scarce,  and  reclusive  inhabitants  of 
wetlands  throughout  eastern  South  America 
and  north  into  Texas  and  Florida  (Johnsgard 
and  Carbonell  1996,  Lockwood  1997,  Todd 
1997).  Little  ecological  data  exist  for  this  spe- 
cies anywhere,  but  particularly  at  the  northern 
extent  of  its  range.  Appropriate  habitat  has 
been  subjectively  defined  as  overgrown 
swamps  and  marshes,  where  aquatic  plants 
like  water  hyacinth  (Eichornia  crcissipes)  and 
water  lilies  (Nymphciceae  spp.)  occur  (Johns- 
gard and  Carbonell  1996,  Todd  1997).  Our  ob- 


‘ Caesar  Kleberg  Wildlife  Research  Institute,  Cam- 
pus Box  218,  Texas  A&M  Univ.— Kingsville,  Kings- 
ville, TX  78363. 

^ Present  Address:  Wildlife  and  Fisheries  Program, 
Division  of  Forestry,  West  Virginia  Univ.,  P.O.  Box 
6125,  Morgantown,  WV  26505-6125. 

^ Deceased. 

■*  Corresponding  author. 


Jective  was  to  quantify  habitat  use  by  Masked 
Ducks  in  the  Coastal  Plains  of  Texas. 

The  study  area  covered  5.5  million  ha  from 
Galveston  Bay,  Texas  south  to  the  Rio  Grande 
River  (Anderson  et  al.  1996,  1998).  The  re- 
gion is  dominated  by  coastal  prairie  and  sandy 
plains  in  the  southeast,  and  rice  fields  and 
coastal  marsh  in  the  northeast  (Anderson  et  al. 
1996).  Palustrine  and  estuarine  wetlands  (Co- 
wardin  et  al.  1979)  are  the  most  abundant  of 
the  wetland  systems  (Muehl  et  al.  1994). 

We  conducted  ground  based  surveys  of  all 
wetlands  located  on  512  quarter-sections 
(64.75-ha  plots)  in  1991-1992  and  1009  in 
1992-1993  (Anderson  et  al.  1996,  1998).  Sur- 
veys for  Masked  Ducks  on  wetlands  were 
conducted  during  September,  November,  Jan- 
uary, and  March.  Wetlands  were  classified  ac- 
cording to  Cowardin  and  coworkers  (1979). 
Surveys  were  part  of  a larger  project  address- 
ing waterbird  habitat  use  (Anderson  1 994,  An- 
derson et  al.  1996),  waterbird  abundance  (An- 
derson et  al.  1998),  and  wetland  abundance 
(Muehl  et  al.  1994). 

We  compared  densities  (no./ha)  of  Masked 
Ducks  among  wetland  types  on  which  they 
occurred  using  ANOVA  and  Scheffe's  proce- 
dure as  the  mean  separation  technique  with  a 
= 0.05  (SAS  Institute  Inc.  1988).  We  included 
in  the  analysis  all  wetlands  of  a type  on  which 
Masked  Ducks  were  observed  (Anderson  et 
al.  1996).  We  compared  microsite  habitat  use 
in  wetlands  with  two-way  contingency  tables 
and  a G-test  (Sokal  and  Rohlf  1995).  Count 


120 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  1.  March  1999 


periods  were  considered  independent  because 
counts  were  at  least  two  months  apart,  wet- 
lands were  dynamic  (Muehl  et  al.  1994),  and 
the  number  of  birds  varied  among  count  pe- 
riods (Anderson  et  al.  1996,  1998).  All 
Masked  Duck  density  data  were  rank  trans- 
formed (Conover  and  Iman  1981,  Potvin  and 
Roff  1993)  because  of  the  large  number  of 
wetlands  that  had  no  Masked  Ducks.  Data 
were  back  transformed  for  presentation. 

We  did  not  observe  any  Masked  Ducks  dur- 
ing 1991-1992.  During  1992-1993,  we  count- 
ed 47  Masked  Ducks  (September  6;  Novem- 
ber 4;  January  34;  March  3)  in  7 flocks  in  4 
separate  basins.  Masked  Ducks  occupied  0.3% 
of  quarter-sections  surveyed  during  1992- 
1993.  All  observations  were  made  in  the 
coastal  and  other  crop  strata  of  the  area  re- 
ferred to  as  the  Texas  Mid-coast  (Anderson  et 
al.  1996,  1998).  Masked  Duck  flocks  averaged 
6.7  birds  (SE  = 3.16;  range  1-25).  Sixty-four 
percent  (n  = 22)  of  undisturbed  Masked 
Ducks  were  observed  feeding. 

Masked  Duck  densities  (no./ha)  on  lacus- 
trine littoral  aquatic-bed  rooted  vascular  (x  = 
0.93;  SE  — 0.52)  and  lacustrine  littoral  aquat- 
ic-bed floating  vascular  (x  = 0.40;  SE  = 0.40) 
wetlands  were  not  different,  but  densities  on 
both  were  greater  than  densities  on  palustrine 
scrub-shrub  broad-leaved  deciduous  (x  = 
0.16;  SE  = 0.15)  wetlands  (ANOVA:  F = 
10.23;  df  = 2,  199;  P < 0.001).  Masked 
Ducks  did  not  occur  on  the  other  79  wetland 
subclasses  that  were  surveyed.  Masked  Ducks 
were  equally  likely  to  occur  in  open  water 
(43%)  and  in  emergent  vegetation  microsites 
within  these  three  wetland  types  (57%;  G-test: 
G = 0.2;  P > 0.05). 

Masked  Ducks  occupied  wetlands  that  av- 
eraged 8.25  ha  (SE  = 1.94)  in  area.  All  wet- 
lands were  seasonally  or  semipermanently 
flooded  with  fresh  water  and  had  emergent 
vegetation  interspersed  with  open  water  [i.e., 
cover  type  two  (Stewart  and  Kantrud  1971)J. 
Rooted  vascular  vegetation  on  occupied  wet- 
lands was  primarily  yellow  lotus  {Nelumbo  lu- 
tea),  but  yellow  waterlily  (Nuphar  mexicana) 
was  also  present.  Floating  vascular  wetlands 
were  dominated  by  water  hyacinth.  Scrub- 
shrub  vegetation  was  primarily  huisache  (Aca- 
cia .smallii)  and  sesbania  (Sesbania  drummon- 
dii). 

Although  few  Masked  Ducks  were  ob- 


served, it  was  apparent  that  they  prefered  wet- 
lands with  abundant  vegetation,  particularly 
aquatic-bed  and  scrub-shrub  wetlands.  No 
Masked  Ducks  were  observed  in  emergent 
wetlands,  as  has  often  been  stated  (Johnsgard 
and  Carbonell  1996,  Lockwood  1997,  Todd 
1997).  Masked  Ducks  also  are  known  to  occur 
in  flooded  rice  fields  in  Venezuela  (Gomez- 
Dallmeier  and  Cringan  1990),  but  none  were 
observed  in  the  Texas  rice  fields  we  surveyed. 

It  is  interesting  to  note  that  Masked  Ducks 
were  not  found  on  smaller  (palustrine)  aquat- 
ic-bed wetlands,  which  are  more  common 
than  lacustrine  littoral  aquatic-bed  wetlands  in 
the  area  (Muehl  et  al.  1994).  Their  absence 
from  these  wetlands  may  be  related  to  pref- 
erence for  larger  (lacustrine)  wetlands,  which 
provide  greater  habitat  diversity,  increased 
protection  from  predators,  and  more  food  re- 
sources (Anderson  et  al.  1996).  However, 
Weller  (1968)  and  Todd  (1997)  indicated  that 
Masked  Ducks  can  use  smaller  wetlands  than 
other  stiff-tailed  ducks  because  they  can  take- 
off vertically  like  dabbling  ducks. 

Previously,  no  specific  information  existed 
on  Masked  Duck  densities  (Johnsgard  and 
Carbonell  1996).  Anderson  and  coworkers 
(1998)  estimated  3817  Masked  Ducks  oc- 
curred in  coastal  Texas  during  January  1993, 
but  only  354  during  March  1993.  Masked 
Ducks  are  not  as  abundant  or  wide-spread  as 
other  waterfowl  species  in  the  study  area  and 
are  rare  throughout  their  range  (Johnsgard  and 
Carbonell  1996).  The  presence  of  Masked 
Ducks  in  Texas  may  be  a temporary  phenom- 
enon (Johnsgard  and  Hagemeyer  1969,  Blan- 
kenship and  Anderson  1993)  or  they  may  al- 
ways be  present,  but  seldom  seen,  as  a result 
of  their  rarity,  secretive  nature,  and  the  pre- 
ponderance of  private  property  in  Texas. 

ACKNOWLEDGMENTS 

We  appreciate  the  landowners  for  allowing  us  access 
to  their  properties.  We  thank  G.  T.  Muehl  for  assisting 
with  data  collection.  S.  L.  Beasom,  E S.  Guthery,  R. 
Kennamer,  and  an  anonymous  referee  reviewed  the 
manuscript.  Funding  was  provided  by  the  Caesar  Kle- 
berg Foundation  for  Wildlife  Conservation  and'  the 
Texas  Prairie  Wetlands  Project. 

LITERATURE  CITED 

Anderson,  J.  T.  1994.  Wetland  use  and  selection  by 
waterfowl  wintering  in  coastal  Texas.  M.S.  thesis, 
Texas  A&M  Univ.— Kingsville,  Kingsville. 


SHORT  COMMUNICATIONS 


121 


Anderson,  J.  T,  T C.  Tacha,  G.  T.  Muehl,  and  D. 
Lobpries.  1996.  Wetland  use  by  waterbirds  that 
winter  in  coastal  Texas.  Natl.  Biol.  Serv.  Inf.  Tech. 
Rep.  8: 1-40. 

Anderson,  J.  T,  G.  T.  Muehl,  and  T.  C.  Tacha.  1998. 
Distribution  and  abundance  of  waterbirds  in  coast- 
al Texas.  Bird  Pop.  4:1-15. 

Blankenship,  T.  L.  and  J.  T.  Anderson.  1993.  A large 
concentration  of  Masked  Ducks  (Oxyura  domini- 
ca)  on  the  Welder  Wildlife  Refuge,  San  Patricio 
County,  Texas.  Bull.  Tex.  Ornithol.  Soc.  26:19- 
21. 

Conover,  W.  J.  and  R.  L.  Iman.  1981.  Rank  transfor- 
mations as  a bridge  between  parametric  and  non- 
parametric  statistics.  Am.  Stat.  35:124-129. 

CowARDiN,  L.  M.,  V.  Carter,  F.  C.  Golet,  and  E.  T. 
Laroe.  1979.  Classification  of  wetlands  and  deep- 
water habitats  of  the  United  States.  U.S.  Fish 
Wildl.  Serv.  FWS/OBS-79/3 1 : 1-103. 

Gomez-Dallmeier,  F.  and  a.  Cringan.  1990.  Biology, 
conservation  and  management  of  waterfowl  in 
Venezuela.  Editorial  Ex  Libris,  Caracas,  Venezue- 
la. 

Johnsgard,  P.  a.  and  M.  Carbonell.  1996.  Ruddy 
Ducks  and  other  stifftails:  their  behavior  and  bi- 
ology. Univ.  of  Oklahoma  Press,  Norman. 


JOHN.SGARD,  P.  A.  AND  D.  Hacjemeyer.  1969.  The 
Masked  Duck  in  the  United  States.  Auk  86:691- 
695. 

Lockwood,  M.  W.  1997.  A closer  look:  Masked  Duck. 
Birding  29:386-390. 

Muehl,  G.  T,  T.  C.  Tacha,  and  J.  T.  Anderson.  1994. 
Distribution  and  abundance  of  wetlands  in  coastal 
Texas.  Texas  J.  Ag.  Nat.  Resour.  7:85-106. 

PoTviN,  C.  AND  D.  A.  Roff.  1993.  Distribution-free 
and  robust  statistical  methods:  viable  alternatives 
to  parametric  statistics?  Ecology  74:1617-1628. 

SAS  Institute  Inc.  1988.  SAS/STAT  user’s  guide,  re- 
lease 6.03  edition.  SAS  Institute  Inc.,  Cary,  North 
Carolina. 

SoKAL,  R.  R.  and  F.  j.  Rohlf.  1995.  Biometry,  third 
edition.  W.  H.  Freeman  and  Company,  New  York. 

Stewart,  R.  E.  and  H.  A.  Kantrud.  1971.  Classifi- 
cation of  natural  ponds  and  lakes  in  the  glaciated 
prairie  region.  U.S.  Fish  Wildl.  Serv.,  Resour. 
Publ.  92:1-23. 

Todd,  F.  S.  1997.  Natural  history  of  the  waterfowl.  Ibis 
Publishing,  Vista,  California. 

Weller,  M.  W.  1968.  Notes  on  some  Argentine  ana- 
tids.  Wilson  Bull.  80:189-212. 


Wilson  Bull,  111(1),  1999,  pp.  121-123 


Gizzarii  Contents  of  Piping  Plover  Chicks  in  Northern  Michigan 

Francesca  J.  Cuthbert,*-^  Brian  Scholtens,^  Lauren  C.  Wemmer,’  and  Robyn  McLain^-"* 


ABSTRACT. — The  diet  of  Piping  Plovers  (Char- 
adrius  melodiis)  is  not  well  known  and  information  on 
diet  requirements  will  enhance  food  resource  assess- 
ment and  identification  of  suitable  habitat  for  this  rare 
species.  Discovery  of  four  dead  Piping  Plover  chicks 
at  Grand  Marais,  Michigan,  allowed  us  to  examine 
their  digestive  tracts  for  identifiable  prey.  Gizzard  con- 
tents represented  16  families  in  6 orders  of  freshwater 
and  terrestrially  occurring  insects  confirming  behav- 
ioral observations  that  plover  chicks  opportunistically 
capture  insects  in  shallow  water  and  along  shorelines. 
The  most  commonly  taken  orders  were  Hymenoptera, 


' Dept,  of  Fisheries  and  Wildlife,  Univ.  of  Minne- 
sota, St.  Paul,  MN  55108. 

^ Dept,  of  Biology,  College  of  Charleston,  Charles- 
ton, SC  29424. 

^ Univ.  of  Michigan  Biological  Station,  Pellston,  Ml 
49769. 

^ Present  address:  5005  Elderben  y Dr.,  Reading,  PA 
19606. 

^ Corresponding  author; 

E-mail:  cuthbOOl  (3)maroon. tc.umn.edu 


Coleoptera,  and  Diptera.  Received  6 May  1998,  ac- 
cepted 30  Aug.  1998. 


Little  is  known  about  the  diet  or  foraging 
behavior  of  the  Piping  Plover  (Charadrius 
melodus)  during  any  part  of  its  annual  cycle. 
Federal  threatened  and  endangered  status 
(U.S.  Fish  and  Wildlife  Service  1985)  and 
sensitivity  to  human  disturbance  preclude  col- 
lection of  birds  for  stomach  content  analysis 
and  require  use  of  nondisruptive  techniques  to 
sample  food  while  plovers  are  present.  Be- 
cause food  availability  is  critical  to  shorebird 
reproductive  success,  migration,  and  over- 
winter survival  (Howe  1983,  Helmers  1992), 
assessment  of  food  resources  is  an  important 
component  of  conservation  efforts  for  this 
species.  Direct  observations  of  food  prefer- 
ence and  foraging  ecology  are  needed  to  im- 


122 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


prove  assessment  of  food  resources  and  allow 
identification  of  areas  appropriate  for  critical 
habitat  designation  and  for  reintroduction  ef- 
forts (U.S.  Fish  and  Wildlife  Service  1988). 

The  Piping  Plover  is  a visual  rather  than 
tactile  feeder,  capturing  invertebrates  moving 
on  the  beach  surface.  Information  on  diet  has 
been  derived  from  gizzard  contents  (Bent 
1929),  direct  observation  of  feeding  (Cairns 
1977),  sampling  of  organisms  present  in  the 
habitat  (Whyte  1985,  Nordstrom  1990,  Loe- 
gering  1992,  Nordstrom  and  Ryan  1996)  and 
fecal  analysis  (Nicholls  1989,  Shaffer  and  La- 
porte  1994);  however,  studies  of  prey  actually 
consumed  by  Piping  Plovers  have  been  re- 
ported only  for  marine  environments.  General 
diet  for  the  species  is  described  as  freshwater 
and  marine  invertebrates  washed  up  on  the 
shore  and  terrestrial  invertebrates  (Haig  1992). 
Insects  appear  to  be  a major  dietary  compo- 
nent in  most  or  all  habitats  occupied  by  Piping 
Plovers  throughout  the  year. 

While  monitoring  plover  nests  in  Michigan 
during  1996  and  1997,  we  salvaged  carcasses 
of  four  chicks  and  examined  gizzard  contents. 
We  believe  this  is  the  first  direct  information 
on  diet  reported  for  Piping  Plovers  from  the 
Great  Lakes  population. 

STUDY  AREA  AND  METHODS 

In  1996  three  pairs  of  Piping  Plovers  nested  near  the 
mouth  of  the  Sucker  River  east  of  Grand  Marais, 
Michigan  (46°  40'  N,  85°  56'  W)  on  the  shore  of  Lake 
Superior.  Two  chicks  from  one  of  these  pairs  disap- 
peared at  six  days  of  age.  The  carcass  of  one  was  dis- 
covered approximately  1.5  weeks  later  and  the  other 
2.5  weeks  later.  The  fresh  carcass  of  the  third  chick 
from  this  brood  was  discovered  a few  hours  after  its 
disappearance  at  19  days  of  age.  In  1997,  a fourth 
Piping  Plover  chick  carcass  was  found  in  the  same 
general  area.  This  chick  disappeared  from  its  brood 
when  it  was  one  week  old  and  was  found  three  days 
later.  The  digestive  tract  from  each  chick  was  removed, 
cut  open  and  Hushed  with  70%  ethyl  alcohol.  Only 
gizzards  yielded  identifiable  samples.  The  contents  of 
the  four  gizzards  were  examined  with  a dissecting  mi- 
croscope and  identified  to  family  by  an  entomologist 
(Scholtens)  familiar  with  in.sects  of  the  region.  Num- 
bers of  individual  prey  were  estimated. 

RESULTS  AND  DISCUSSION 

Parts  of  adult  and  larval  insects  were  the 
only  prey  identified  in  the  gizzards.  Prey  rep- 
resented 16  families  in  6 orders;  Hymenoptera 
(32%),  Coleoptera  (29%),  Diptera  (28%),  He- 


miptera  and  Homoptera  (10%),  and  Ephem- 
eroptera  (1%).  Based  on  the  natural  history  of 
these  families  in  northern  Michigan,  they  can 
be  characterized  as  inhabiting  shoreline/wet 
sand  (Dolichopodidae,  Ephydridae),  shallow 
water/wet  sand  (Corixidae,  Dytiscidae,  Hali- 
plidae),  beach  vegetation/sand  surface  (Aphi- 
didae,  Braconidae,  Carabidae,  Cicadellidae, 
Curculionidae,  Ichneumonidae)  and  general 
shoreline  habitat  (Superfamily  Chalcidoidea, 
Chironomidae,  Formicidae,  Muscidae,  un- 
identified Ephemeropteran  family).  The  only 
previous  information  related  to  the  prey  of  the 
Great  Lakes  Piping  Plover  was  found  in  a 
study  of  invertebrates  present  within  National 
Park  lands  being  assessed  as  potential  reintro- 
duction sites  (Nordstrom  1990,  Nordstrom 
and  Ryan  1996).  Nordstrom  (1990)  found  48 
families  of  9 orders  of  insects  and  1 family  of 
arachnid  on  the  shore  of  Lake  Superior  in  Pic- 
tured Rocks  National  Lakeshore  (approxi- 
mately 20  km  west). 

In  the  Great  Lakes  region.  Piping  Plovers 
nest  on  wide  sandy  beaches  and  forage  along 
the  water  line  of  Lake  Michigan  and  Lake  Su- 
perior. Birds  occasionally  glean  insects  from 
beach  vegetation  and  at  some  sites  forage 
along  the  edges  of  creeks  and  shallow  beach 
ponds.  At  the  Grand  Marais  site,  adults  and 
chicks  were  observed  foraging  primarily 
along  the  edge  of  the  Sucker  River  and  in 
shallow  pools  of  water  and  wet  depressions  in 
the  sand  along  the  river.  Aquatic  insects  in  the 
chick  gizzards  are  consistent  with  this  obser- 
vation, and  the  presence  of  terrestrial,  phy- 
tophagous insects  indicates  that  chicks  also 
gleaned  insects  from  beach  vegetation.  Pres- 
ence of  aquatic  algae-eating  beetles  (Halipli- 
dae)  in  the  gizzards  suggests  that  chicks  pick- 
ed insects  from  algae  on  the  river  edges. 
While  Nordstrom  and  Ryan  (1996)  reported  a 
predominance  of  Dipterans  in  the  Lake  Su- 
perior habitat  they  sampled,  we  found  pre- 
dominantly Hymenopterans  and  Coleopterans 
in  the  chick  gizzards.  We  identified  two  fam- 
ilies (Corixidae  and  Dytiscidae)  in  the  giz- 
zards that  were  not  reported  by  Nordstrom 
(1990).  Members  of  both  families  inhabit 
streams,  ponds,  and  stagnant  pools  associated 
with  beaches. 

Given  the  constraints  on  disturbance  and 
collecting,  opportunistic  discovery  of  dead 
plovers  and  subsequent  study  of  their  diges- 


SHORT  COMMUNICATIONS 


123 


live  tracts  may  contribute  information  vital  to 
understanding  the  diet  of  this  endangered  spe- 
cies. For  example,  insects  gleaned  from  veg- 
etation, algal  mats,  and  the  water  surface 
probably  would  not  be  revealed  by  traditional 
methods  (e.g.,  sticky  traps)  used  to  sample  in- 
vertebrates in  the  habitat.  The  information  on 
foraging  behavior  and  prey  selection  that  both 
gizzard  and  fecal  analyses  provide  is  needed 
to  increase  accuracy  of  methods  used  to  sam- 
ple invertebrates  from  the  habitat.  It  is  impor- 
tant to  note  that  all  three  methods  of  quanti- 
fying plover  food  resources  (gizzard  analysis, 
fecal  analysis,  and  sampling  from  habitat) 
may  be  greatly  affected  by  the  time  samples 
are  obtained  because  of  temporal  variation  in 
insect  abundance.  Fecal  analysis  offers  the  ad- 
vantage that  numerous  samples  can  be  col- 
lected from  the  same  individuals  to  reveal 
temporal  patterns  of  prey  selection;  however, 
this  method  underestimates  soft-bodied  inver- 
tebrates (Shaffer  and  Laporte  1994).  Because 
gizzard  contents  have  undergone  less  diges- 
tion, they  are  presumably  less  biased  in  this 
regard,  but  this  has  not  been  confirmed. 

ACKNOWLEDGMENTS 

Funds  and  permission  to  monitor  and  salvage  plo- 
vers were  provided  by  the  Endangered  Species  Pro- 
gram, Michigan  Department  of  Natural  Resources  and 
the  Michigan  Field  Office  and  Region  3 Endangered 
Species  Office,  U.S.  Fish  and  Wildlife  Service.  Addi- 
tional support  was  received  from  the  University  of 
Michigan  Biological  Station  and  the  University  of 
Michigan  Research  Experiences  for  Undergraduates 
Program  (National  Science  Foundation  Grant  #94- 
0512).  We  thank  employees  of  Whitefish  Point  Bird 
Observatory  and  Grand  Marais  landowners  for  their 
cooperation  with  Piping  Plover  monitoring  efforts. 

LITERATURE  CITED 

Bent,  A.  C.  1929.  Charadrius  melodus.  U.S.  Natl. 

Mus.  Bull.  146:236-246. 


Cairns,  W.  E.  1977.  Breeding  biology  and  behavior  of 
the  Piping  Plover  {Cluiradriiis  melodus)  in  south- 
ern Nova  Scotia.  M.S.  thesis,  Dalhousie  Univ., 
Halifax,  Nova  Scotia. 

Haig,  S.  M.  1992.  Piping  Plover  {Charadrius  melo- 
dus). In  The  birds  of  North  America,  no.  2 (A. 
Poole,  P.  Stettenheim,  and  E Gill,  Eds.).  The 
Academy  of  Natural  Sciences,  Philadelphia,  Penn- 
sylvania; The  American  Ornithologists’  Union, 
Washington,  D.C. 

Helmers,  D.  L.  1992.  Shorebird  management  manual. 
Western  Hemisphere  Shorebird  Reserve  Network. 
Manomet.  Massachusetts. 

Howe,  M.  1983.  Breeding  ecology  of  North  American 
shorebirds:  patterns  and  constraints.  Proc.  West. 
Hemisphere  Waterfowl  Waterbirds  Symp.  1 :95- 
100. 

Loegering,  J.  P.  1992.  Piping  Plover  breeding  biology, 
foraging  ecology  and  behavior  on  Assateague  Is- 
land National  Seashore,  Maryland.  M.S.  thesis, 
Virginia  Polytechnic  Inst,  and  State  Univ.,  Blacks- 
burg. 

Nicholes,  J.  L.  1989.  Distribution  and  other  ecological 
aspects  of  Piping  Plovers  (Charadrius  melodus) 
wintering  along  the  Atlantic  and  Gulf  Coasts. 
M.S.  thesis.  Auburn  Univ.,  Auburn,  Alabama. 

Nordstrom,  L.  H.  1990.  Assessment  of  habitat  suit- 
ability for  reestablishment  of  Piping  Plovers  in  the 
Great  Lakes  National  Lakeshores.  M.S.  thesis, 
Univ.  of  Missouri,  Columbia. 

Nordstrom,  L.  H.  and  M.  R.  Ryan.  1996.  Invertebrate 
abundance  at  occupied  and  potential  Piping  Plover 
nesting  beaches:  Great  Plains  alkali  wetlands  vs. 
the  Great  Lakes.  Wetlands  16:429—435. 

Shaffer,  F.  and  P.  Laporte.  1994.  Diet  of  Piping  Plo- 
vers on  the  Magdalen  Islands,  Quebec.  Wilson 
Bull.  106:531-536. 

U.S.  Fish  and  Wildlife  Service.  1985.  Determination 
of  endangered  and  threatened  status  for  the  Piping 
Plover.  Federal  Register  50:50726-50734. 

U.S.  Fish  and  Wildlife  Service.  1988.  Recovery  plan 
for  Piping  Plovers  breeding  on  the  Great  Lakes 
and  Northern  Great  Plains.  U.S.  Fish  and  Wildlife 
Service.  Twin  Cities,  Minnesota. 

Whyte,  A.  J.  1985.  Breeding  ecology  of  the  Piping 
Plover  (Charadrius  melodus)  in  central  Saskatch- 
ewan. M.S.  thesis,  Saskatchewan  Univ..  Saska- 
toon. 


124 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  I,  March  1999 


Wilson  Bull.,  II  Id),  1999,  pp.  I24-I28 


Nesting  of  Four  Poorly-Known  Bird  Species  on  the  Caribbean  Slope  of 

Costa  Rica 

Bruce  E.  Young'-^'^  and  James  R.  Zook- 


ABSTRACT — We  describe  the  nests  of  four  species 
of  birds  from  the  Caribbean  slope  of  Costa  Rica.  A 
Great  Potoo  (Nyctibius  grand  is,  nest  previously  un- 
known from  Mesoamerica)  nest  was  nothing  more  than 
a crevice  in  a high  branch  of  a large  tree,  similar  to 
those  reported  in  South  America.  A nest  of  the  Torrent 
Tyrannulet  [Serpophaga  cinerea)  was  found  along  a 
river  at  35  m elevation,  much  lower  than  previous 
breeding  reports  for  this  normally  montane  species. 
Also,  we  confirm  systematists’  predictions  that  the 
Tawny-chested  Llycatcher  (Aplianotriccus  capilalis),  a 
species  of  near-threatened  conservation  status,  is  a sec- 
ondary cavity  nester.  Finally,  we  report  on  the  second 
known  nest  of  the  Sooty-faced  Finch  (Lysurus  crassi- 
rostris)  from  montane  forest.  Received  20  May  199H, 
accepted  7 Oct.  1998. 

RESUMEN. — Describimos  los  nidos  de  cuatro  es- 
pecies  de  aves  de  la  vertiente  del  caribe  en  Costa  Rica. 
Un  nido  de  Nyctibius  grandis  (nido  era  anteriormente 
desconocido  en  Mesoamerica)  era  nada  mas  que  una 
grieta  en  una  rama  alta  de  un  arbol  grande,  muy  pa- 
recido  a los  nidos  de  la  misma  especie  encontrados  en 
America  del  Sur.  Describimos  un  nido  de  Serpophaga 
cinerea,  el  cual  fue  encontrado  en  un  rio  a los  35 
msnm,  mucho  mas  bajo  que  los  otros  registros  de  esta 
especie  del  bosque  montano.  Describimos  unas  obser- 
vaciones  que  indican  que  el  Aphanotriccus  capitalis, 
una  especie  ligeramente  amenazada  desde  la  punta  de 
vista  conservacionista,  anida  secondariamente  en  los 
huecos  dentro  de  los  arboles.  Finalmente,  presentamos 
la  segunda  descripcion  del  nido  de  Lysurus  crassiros- 
tris,  el  cual  fue  encontrado  en  un  bosque  montano. 


Despite  decades  of  intensive  ornithological 
study,  the  nests  and  eggs  of  a number  of  Cen- 
tral American  bird  species  are  poorly  known 
(Skutch  1954,  1960;  Stiles  and  Skutch  1989). 
Especially  enigmatic  are  species  inhabiting 


' Organizacion  para  Estudios  Tropicales,  Apartado 
676-2050.  .San  Pedro  de  Montes  de  Oca,  Costa  Rica. 

^Apartado  182-4200.  Naranjo  de  Alajuela,  Costa 
Rica. 

' Current  Address:  Latin  America  and  Ctiribbean  Di- 
vision. The  Nature  Conservancy,  4245  North  Fairfax  Dr.. 
Arlington.  Virginia  22203;  E-mail:  byoung@tnc.org 
■*  Corresponding  author. 


the  relatively  inaccessible  habitats  of  the  hu- 
mid Caribbean  slope.  The  lack  of  reproductive 
information  about  these  species  hinders  phy- 
logenetic studies  of  the  relationships  among 
avian  lineages,  studies  of  intraspecific  varia- 
tion, general  analyses  of  reproductive  behav- 
ior, and  the  development  of  management  prac- 
tices for  conserving  avian  biodiversity.  Here 
we  describe  the  first  Mesoamerican  nest  of  the 
Great  Potoo  (Nyctibius  grandis),  the  first  low- 
land nest  of  the  Torrent  Tyrannulet  {Serpo- 
phaga cinerea),  the  first  nest  of  the  Tawny- 
chested  Flycatcher  (Aphanotriccus  capitalis), 
and  the  second  nest  of  the  Sooty-faced  Finch 
(Lysurus  crassirostris). 

STUDY  AREA 

Our  observations  were  made  in  the  45,000  ha  La 
Selva-Braulio  Canillo  National  Park  reserve  complex 
in  Heredia  Province,  northeastern  Costa  Rica.  The  re- 
serve complex,  the  largest  protected  elevational  tran- 
sect in  Central  America,  extends  from  montane  rain 
forest  surrounding  the  Barva  Volcano  at  2,900  m down 
to  lowland  wet  forest  at  the  La  Selva  Biological  Sta- 
tion at  35  m elevation  on  the  Caribbean  slope  (Timm 
et  al.  1989). 

Observations  were  made  in  1997  during  routine  bird 
monitoring  activities  at  La  Selva  (10°  26'  N,  83°  59' 
W)  and  at  a remote  campsite  at  1070  m elevation  in 
Braulio  Carrillo  National  Park  (10°  16'  N,  84°  5'  W). 
Annual  rainfall  at  La  Selva  averages  3962  mm,  with  a 
relative  dry  period  between  January  and  March  during 
most  years  (Sanford  et  al.  1994).  Although  precipita- 
tion data  are  .scarce  for  higher  elevations,  annual  rain- 
fall may  average  over  5000  mm  at  the  1070-m  site, 
where  clouds  frequently  bathe  the  premontane  rain  for- 
est in  mist  (Hartshorn  and  Peralta  1988).  Average  can- 
opy height  varies  from  28-38  m at  La  Selva  to  22-36 
m at  1070  m (Lieberman  et  al.  1996). 

NEST  DESCRIPTIONS  AND  DISCUSSION 

Nyctibius  grandis. — The  Great  Potoo  is  one 
of  three  Central  American  species  of  the  ge- 
nus Nyctibius,  the  only  genus  in  the  exclu- 
sively Neotropical  family  Nyctibiidae.  Al- 
though difficult  to  observe  in  daylight  because 
it  roosts  motionless  on  canopy  branches,  the 


SHORT  COMMUNICATIONS 


125 


FIG.  1.  Adult  and  nestling  Great  Potoo  (Aycr/Wi/A- 
grandis)  on  the  branch  of  a large  Hernatulia  didyman- 
tha  tree.  La  Selva  Biological  Station,  Costa  Rica,  April 

1997. 

Great  Potoo  is  readily  detectable  at  night  by 
its  characteristic  calls  (Perry  1979,  Slud 
1979).  It  ranges  from  southern  Mexico  to 
southeastern  Brazil  and  central  Bolivia  (AOU 

1998,  Howell  and  Webb  1995). 

On  5 February,  a visitor  to  the  station  (K. 
McGowan)  found  a single  Great  Potoo 
perched  34  m high  on  the  branch  of  a 44  m 
tall  Hernandia  didyrnantha  (Hemandiaceae) 
tree.  The  tree  was  located  500  m inside  of  old 
growth  forest,  at  the  edge  of  a large  treefall 
gap  on  a steep  hill.  The  bird  was  perched  on 
an  upward  sloping  section  of  an  S-shaped 
branch  approximately  20  cm  in  diameter. 
Thereafter  on  daily  visits,  we  observed  a bird 
in  exactly  the  same  position. 

On  4 April,  we  observed  for  the  first  time 
a fully  feathered  chick  on  the  branch  in  front 
of  the  adult.  The  chick  was  paler  in  coloration 
than  the  adult  and  about  one  quarter  the  size 
(Fig.  1).  On  subsequent  days,  the  chick  moved 
between  a hidden  position  under  the  adult’s 
breast  feathers  and  the  branch  immediately  in 
front  of  the  adult.  The  adult  remained  almost 
motionless  and  never  moved  along  the  branch. 
We  last  saw  the  birds  on  20  April.  Despite 
extensive  searches  on  all  neighboring  branch- 
es in  the  following  days  and  months,  neither 
adult  nor  young  was  seen  again. 

The  motionless  adult  we  saw  was  probably 


incubating  an  egg  initially  and  then  brooding 
a chick,  although  we  could  not  determine  the 
hatching  date.  Although  collections  of  adults 
on  eggs  in  Brazil  and  observations  of  N.  gri- 
seus  in  Costa  Rica  indicate  that  males  incu- 
bate during  the  day  (Skutch  1970,  Sick  1993), 
we  could  not  identify  the  sex  of  the  individual 
(or  individuals)  we  saw  because  sexes  are 
similar  in  outward  appearance  in  Great  Potoos 
(Land  and  Schultz  1963,  Wetmore  1968).  No 
nest  was  visible  and,  from  our  vantagepoint 
30  m from  the  tree  and  level  with  the  nest, 
only  a slight  crevice  in  the  branch  was  visible. 
Although  the  location  appeared  precarious,  re- 
ports from  Brazil  suggest  that  a notch  in  the 
nest  branch  can  securely  hold  an  egg  (Sick 

1993) . 

At  a nest  in  Venezuela,  a chick  remained 
with  its  parent  for  a month,  and  then  alone  for 
almost  another  month  before  growing  to  a size 
greater  than  two-thirds  that  of  the  adult  and 
dispersing  (Vanderwerf  1988).  This  observa- 
tion suggests  that  the  chick  we  observed, 
which  never  attained  half  the  size  of  the  adult, 
did  not  survive.  In  addition,  the  wings  did  not 
appear  sufficiently  developed  for  sustained 
flight.  Despite  an  extensive  ground  search,  we 
found  no  evidence  of  its  having  fallen  from 
the  nest.  The  chick  may  have  been  taken  by 
an  arboreal  predator  such  as  a monkey  {Cebus 
capucinus,  Ateles  geojfroyi,  or  Alouatta  pal- 
liata),  tayra  (Eira  barbara),  or  Collared  For- 
est-Falcon (Micrastur  sernitorquatus),  all 
common  in  the  area.  Even  though  the  attempt 
was  probably  unsuccessful,  its  daily  survivor- 
ship rate  in  the  (presumed)  egg  and  chick 
stage  of  98.7%  is  substantially  higher  than  the 
93%  rate  measured  for  understory  cup-nesting 
birds  the  same  year  in  the  same  area  (B. 
Young,  unpubl.  data). 

Serpophaga  cinerea. — The  Torrent  Tyran- 
nulet  is  a conspicuous  resident  of  highland 
rivers  from  Costa  Rica  to  Venezuela  and  Bo- 
livia (AOU  1998).  Their  cup-shaped  nest  at- 
tached to  vegetation  above  rivers  is  well 
known  (Skutch  1960).  The  elevational  distri- 
bution of  the  species  is  variously  described  as 
250-2500  m in  different  parts  of  its  range 
(Meyer  de  Schauensee  and  Phelps  1978,  Hilty 
and  Brown  1986,  Stiles  and  Skutch  1989,  Rid- 
gely  and  Gwynne  1993,  Ridgely  and  Tudor 

1994) . 

We  first  detected  Torrent  Tyrannulets  on  the 


126 


THE  WILSON  BULLETIN  • Vol.  HI.  No.  I,  March  1999 


Sarapiqui  River  in  1994  in  Chilamate,  5 km 
west  of  La  Selva  at  an  elevation  of  40  m.  In 
March  1997,  we  began  seeing  individuals 
along  the  same  river  where  it  passes  through 
La  Selva  at  35  m elevation.  These  were  the 
first  observations  of  the  species  in  La  Selva 
in  four  decades  of  ornithological  investigation 
(Levey  and  Stiles  1994).  During  the  first  week 
of  April,  a pair  of  tyrannulets  began  construc- 
tion of  a nest  0.5  m above  the  water  level  in 
shrub  vegetation  growing  in  the  middle  of  a 
small  island  in  the  river,  which  is  approxi- 
mately 80  m wide  and  1 m deep  at  this  point. 
The  nest  was  typical  for  the  species  in  being 
cup  shaped,  supported  by  vertical  branches, 
with  feathers  and  moss  woven  into  the  struc- 
ture (Skutch  1960).  We  monitored  the  nest  ev- 
ery 2-5  days  until  22  April  when  two  eggs 
were  found  in  the  nest.  During  each  visit,  two 
birds  were  active  near  the  nest,  adding  mate- 
rial and  adjusting  its  structure.  On  our  next 
visit,  on  30  April,  the  nest  had  disappeared 
and  the  birds  were  not  present. 

Despite  its  conspicuousness  on  rivers,  pre- 
vious reports  of  the  species  occurring  in  the 
lowlands  are  of  scattered  observations  of  in- 
dividuals as  low  as  100  m elevation  (Hilty  and 
Brown  1986,  Ridgely  and  Gwynne  1993,  Rid- 
gely  and  Tudor  1994).  The  Torrent  Tyrannu- 
lets we  observed  may  have  strayed  from  their 
higher  elevation  habitat  because  of  a hydro- 
electric project  in  progress  higher  up  the  Sar- 
apiqui River. 

Aphanotriccus  capitalis. — The  Tawny- 
chested  Flycatcher  occurs  in  second  growth 
and  disturbed  forest  in  Nicaragua  and  Costa 
Rica  (AOU  1998).  Nests  of  both  this  species 
and  its  congener  are  apparently  undescribed 
(Lanyon  and  Lanyon  1986),  and  both  have 
near  threatened  conservation  status  as  a result 
of  their  small,  fragmented  ranges  (Collar  et  al. 
1994).  We  provide  two  observations  to  sug- 
gest the  species  nests  in  either  cavities  or 
crevices  in  trees. 

On  23  April,  a group  of  birdwatchers  spot- 
ted a pair  of  Tawny-chested  Flycatchers  build- 
ing a nest  in  the  hollow  of  a dead  branch  stub 
in  an  otherwise  live  Alchornea  costaricense 
(Euphorbiaceae)  tree.  On  30  April,  we  again 
observed  two  birds  carrying  fine  nesting  ma- 
terial, including  moss,  to  the  hollow  for  about 
30  min.  The  birds  apparently  abandoned  the 
attempt,  as  we  never  saw  them  there  again. 


The  rectangular  hollow  was  oriented  slightly 
upward,  8 cm  deep  and  25  X 8 cm  wide,  1.5 
m up  in  the  36  cm  dbh  tree.  The  tree  was  in 
a small,  shady  clearing  10  m away  from  a 
small  (0.25  ha)  patch  of  second  growth  forest. 

A local  naturalist  guide,  E.  Castro,  reported 
finding  an  active  nest  of  this  species  in  a hol- 
low section  of  a 30  m diameter  clump  of 
Asian  bamboo  {Guadua  sp.)  in  a patch  of  sec- 
ond growth  forest  2 km  north  of  La  Selva. 
Castro  reported  seeing  adults  carrying  food  to 
the  nest  and  later  feeding  a fledgling  in  the 
vicinity.  The  nest  was  5.9  m above  the 
ground,  13  cm  in  diameter  (the  diameter  of 
the  bamboo),  and  entered  through  a 5 cm  high 
by  2 cm  wide  teardrop-shaped  opening.  The 
bottom  of  the  opening  was  3 cm  above  a node, 
leaving  a shallow  area  for  the  nest.  This  find- 
ing confirms  predictions  that  Aphanotriccus 
builds  nests  in  crevices  based  on  its  phylo- 
genetic closeness  to  LMthrotriccus  and  Cne- 
motriccus,  two  genera  known  to  build  nests  in 
crevices  (Lanyon  1986,  Lanyon  and  Lanyon 
1986).  This  crevice  nesting  habit  may  aid  in 
the  conservation  of  the  species;  Guadua  bam- 
boo is  widely  introduced  in  the  region  for  use 
in  supporting  banana  trees  on  plantations. 

Lysurus  crassirostris. — The  Sooty-faced 
Finch  occurs  in  dense  vegetation  in  wet,  mid- 
elevation forest  in  Costa  Rica  and  Panama 
(AOU  1998).  Its  one  congener,  the  Olive 
Finch  (L.  castaneiceps),  occurs  in  humid  mon- 
tane forests  along  the  coastal  Andes  moun- 
tains from  Colombia  to  Peru  (Ridgely  and  Tu- 
dor 1989). 

On  7 May  we  discovered  two  adult  Sooty- 
faced  Finches  entering  and  exiting  a nest  2.1 
m high  attached  to  the  side  of  a large  Sorn- 
mera  sp.  (Rubiaceae)  tree  (dbh  = 46  cm)  next 
to  our  camp  at  1070  m,  several  kilometers  in- 
side undisturbed  forest.  The  nest  tree  was  lo- 
cated at  the  side  of  a little  used  trail  at  the 
edge  of  a stream  crossing.  The  nest  itself  was 
woven  into  a thick  epiphyte  mat  on  a section 
of  the  trunk  directly  above  the  3 m wide 
stream.  The  nest  was  a bulky,  covered  dome 
with  a side  entrance.  The  nest  was  almost  en- 
tirely constructed  of  fresh  moss  with  a lining 
of  thin,  black,  stringy  fungal  rhizomorphs  and 
strips  of  dried  bamboo  (Chusquea  sp.)  leaves. 
The  trunk  of  the  nest  tree  was  covered  with 
the  same  moss  as  was  used  to  construct  the 


SHORT  COMMUNICATIONS 


127 


nest,  causing  the  nest  to  be  fairly  inconspic- 
uous. 

On  1 1 May,  after  observing  the  adults 
spending  long  periods  of  time  in  the  nest,  we 
exainined  the  eggs.  The  two  eggs  were  whit- 
ish and  speckled  with  lavender.  The  speckles 
were  densest  around  the  thick  ends  of  the 
eggs.  Assuming  the  adults  were  incubating 
and  that  the  female  had  finished  laying,  the 
clutch  size  for  this  nest  was  two,  typical  for 
birds  of  humid  tropical  forests  (Skutch  1985). 

This  nest  was  very  similar  to  the  one  other 
nest  described  for  the  species,  although  the 
latter  was  built  into  the  side  of  a fern  stem 
and  apparently  was  not  associated  with  water 
(Barrantes  1994).  These  two  nests  were  sim- 
ilar to  the  single  nest  of  the  congeneric  Olive 
Finch  described  from  Ecuador  (Schulenberg 
and  Gill  1987).  All  nests  of  the  two  species 
were  bulky  and  dome-shaped,  constructed  pri- 
marily of  mosses  on  the  outside,  cryptically 
situated  in  moss-dominated  vegetation,  and,  in 
two  cases,  located  over  moving  water.  The 
nests  of  the  two  species  differ  in  that  the  Olive 
Finch  nest  was  built  on  the  side  of  a rock  in- 
stead of  a tree  and  had  a lining  of  dry  leaves 
instead  of  rhizomorphs.  The  eggs  differ  sub- 
stantially in  that  the  eggs  of  the  Olive  Finch 
were  immaculate  white  (Schulenberg  and  Gill 
1987)  compared  with  the  spotted  eggs  found 
in  both  Sooty-faced  Finch  nests  (Barrantes 
1994).  Interestingly,  eggs  in  two  collections 
attributed  to  the  Olive  Finch  are  spotted 
(Schulenberg  and  Gill  1987),  suggesting  that 
the  Ecuadorian  discovery  of  all  white  eggs 
may  have  been  atypical  for  the  group. 

ACKNOWLEDGMENTS 

We  thank  R.  Tenorio  of  the  Area  de  Conservacion 
Cordillera  Volcanica  Central  and  the  Organization  for 
Tropical  Studies  for  facilitating  our  studies.  We  are 
grateful  for  the  field  observations  of  R.  Alvarado,  A. 
Downs,  Y.  Hernandez,  and  A.  Hies,  and  to  K.  Mc- 
Gowan for  finding  the  potoo.  R.  Thiele  photographed 
the  potoo.  J.  Blake  and  two  anonymous  reviewers  pro- 
vided many  useful  comments  to  improve  the  manu- 
script. Financial  support  for  our  studies  was  generously 
provided  by  H.  M.  S.  Ambassador  Michael  Jackson 
and  the  British  Embassy  in  Costa  Rica  (for  work  in 
Braulio  Carrillo)  and  by  the  Chiquita  Brands  Company 
(for  work  in  La  Selva). 

LITERATURE  CITED 

American  Ornithologists’  Union.  1998.  Check-li.st 
of  North  American  birds,  seventh  ed.  American 
Ornithologists’  Union,  Washington,  D.C. 


Barrantes,  G.  1994.  First  description  of  the  nest  and 
eggs  of  the  Sooty-faced  Finch.  Wilson  Bull.  106: 
574. 

Collar,  N.  J.,  M.  J.  Crosby,  and  A.  J.  Stattersfield. 
1994.  Birds  to  watch  2:  the  world  list  of  threat- 
ened birds.  Smithsonian  Press,  Washington,  D.C. 

Hartshorn,  G.  S.  and  R.  Peralta.  1988.  Preliminary 
description  of  primary  forests  along  the  La  Selva- 
Volcan  Barva  altitudinal  transect.  Pp.  281-296  in 
Tropical  rainforests:  diversity  and  conservation  (F. 
Alameda  and  C.  M.  Pringle,  Eds.).  California 
Academy  of  Sciences,  San  Francisco. 

Hilty,  S.  L.  and  W.  L.  Brown.  1986.  A guide  to  the 
birds  of  Colombia.  Princeton  Univ.  Press,  Prince- 
ton, New  Jersey. 

Howell,  S.  N.  G.  and  S.  Webb.  1995.  The  birds  of 
Mexico  and  northern  Central  America.  Oxford 
Univ.  Press,  New  York. 

Land,  H.  C.  and  W.  L.  Schultz.  1963.  A proposed 
subspecies  of  the  Great  Potoo,  Nyctibius  grandis 
(Gmelin).  Auk  80:195-196. 

Lanyon,  W.  E.  1986.  A phylogeny  of  the  thirty-three 
genera  in  the  Empidonax  assemblage  of  tyrant  fly- 
catchers. Am.  Mus.  Novit.  2846:1-64. 

Lanyon,  W.  E.  and  S.  M.  Lanyon.  1986.  Generic  sta- 
tus of  Euler’s  Flycatcher:  a morphological  and 
biochemical  study.  Auk  103:341—350. 

Levey,  D.  J.  and  F.  G.  Stiles.  1994.  Birds:  ecology, 
behavior,  and  taxonomic  affinities.  Pp.  217-228  in 
La  Selva:  ecology  and  natural  history  of  a Neo- 
tropical rain  forest  (L.  A.  McDade,  K.  S.  Bawa, 
H.  A.  Hespenheide,  and  G.  S.  Hartshorn,  Eds.). 
Univ.  of  Chicago  Press,  Chicago,  Illinois. 

Lieberman,  D.,  M.  Lieberman,  R.  Peralta,  and  G.  S. 
Hartshorn.  1996.  Tropical  forest  structure  and 
composition  on  a large-scale  altitudinal  gradient 
in  Costa  Rica.  J.  Trop.  Ecol.  84:137—152. 

Meyer  de  Schauensee,  R.  and  W.  H.  Phelps.  Jr. 
1978.  A guide  to  the  birds  of  Venezuela.  Princeton 
Univ.  Press,  Princeton,  New  Jersey. 

Perry,  D.  R.  1979.  The  Great  Potoo  in  Costa  Rica. 
Condor  81:320-321. 

Ridgely,  R.  S.  and  j.  a.  Gwynne,  Jr.  1993.  Guia  de 
las  aves  de  Panama.  ANCON,  Panama  City,  Pan- 
ama. 

Ridgely,  R.  S.  and  G.  Tudor.  1989.  The  birds  of 
South  America,  vol.  1 . Univ.  of  Texas  Press,  Aus- 
tin. 

Ridgely,  R.  S.  and  G.  Tudor.  1994.  The  birds  of 
South  America,  vol.  2.  Univ.  of  Texas  Press,  Aus- 
tin. 

Sanford,  R.  L.,  Jr.,  P.  Paaby.  J.  C.  Luvall,  and  E. 
Phillips.  1994.  Climate,  geomorphology,  and 
aquatic  systems.  Pp.  19-33  in  La  Selva:  ecology 
and  natural  history  of  a Neotropical  rain  forest  (L. 
A.  McDade,  K.  S.  Bawa,  H.  A.  Hespenheide,  and 
G.  S.  Hartshorn,  Eds.).  Univ.  of  Chicago  Press, 
Chicago,  Illinois. 

Schulenberg,  T.  S.  and  F.  B.  Gill.  1987.  First  de- 
scription of  the  nest  of  the  Olive  Finch,  Lysurus 
castaneiceps.  Condor  89:673-674. 


128 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


Sick,  H.  1993.  Birds  in  Brazil.  Princeton  Univ.  Press, 
Princeton,  New  Jersey. 

Skutch,  a.  E 1954.  Life  histories  of  Central  American 
birds.  Pac.  Coast  Avifauna  31:1-448. 

Skutch,  A.  E 1960.  Life  histories  of  Central  American 
birds  11.  Pac.  Coast  Avifauna  34:1—593. 

Skutch,  A.  E 1970.  Life  history  of  the  Common  Po- 
too.  Living  Bird  9:265-280. 

Skutch,  A.  E 1985.  Clutch  size,  nesting  success,  and 
predation  on  nests  of  Neotropical  birds  reviewed. 
Ornithol.  Monogr.  36:575-594. 

Slud,  P.  1979.  Calls  of  the  Great  Potoo.  Condor  81: 
322. 


Stiles,  E G.  and  A.  E Skutch.  1989.  A guide  to  the 
birds  of  Costa  Rica.  Cornell  Univ.  Press,  Ithaca, 
New  York. 

Timm,  R.  M.,  D.  E.  Wilson,  B.  L.  Clauson,  R.  K. 
LaVal,  and  C.  S.  Vaughan.  1989.  Mammals  of 
the  La  Selva-Braulio  Carrillo  Complex,  Costa 
Rica.  N.  Am.  Fauna  75:1-162. 

Vanderwerf,  E.  a.  1988.  Observations  on  the  nesting 
of  the  Great  Potoo  {Nyctibiu.s  grandis)  in  central 
Venezuela.  Condor  90:948—950. 

Wetmore,  a.  1968.  The  birds  of  the  Republic  of  Pan- 
ama. Smithson.  Misc.  Collect.  150(2):  1—605. 


Wilson  Bull.  111(1),  1999,  pp.  128-130 


Sexual  Dimoq^hism  in  the  Song  of  Sumichrast’s  Wren 

Monica  Perez-Villafana,'  Hector  Gomez  de  Silva  and 
Atahualpa  DeSucre-Medrano^ 


ABSTRACT. — We  report  on  a song-like  vocaliza- 
tion of  female  Sumichrast’s  Wren  (Hylorchilus  sumi- 
chrasti).  The  female  song  is  a series  of  similar  sylla- 
bles, all  at  the  same  low  pitch,  that  varies  in  length. 
Thus,  it  differs  strongly  from  the  rich  and  complex 
songs  of  male  Sumichrast’s  Wrens  and  of  most  other 
wrens.  Received  27  Feb.  1998,  accepted  25  Aug.  1998. 


There  is  increasing  evidence  that  female 
song  is  not  as  rare  in  birds  as  previously 
thought  (Langmore  1998).  Among  most 
wrens,  females  have  songs  that  are  similar  to 
and  sometimes  combine  with  those  of  their 
mates  in  antiphonal  duets  (Skutch  1940, 
1960).  Until  recently,  the  only  reported  case 
of  strong  sexual  difference  in  wren  songs  was 
in  the  southern  House  Wren  {Troglodytes  ae- 
don)  among  which  the  females  give  a simple 
twittering  and/or  a short  trill,  at  least  in  Costa 
Rica  and  Panama,  generally  countersinging 


' Calle  1537-3,  Col.  San  Juan  de  Aragon,  Seccion 
6,  C.P.  07918,  Mexico,  D.E,  Mexico. 

^ Instituto  de  Ecologia,  UNAM,  Apartado  Po.stal  70- 
275,  Ciudad  Univ.,  UNAM,  C.P.  04510,  Mexico,  D.E, 
Mexico. 

’ UNAM  Campus  Iztacala,  Laboratorio  de  Zoologfa, 
Apartado  Postal  314,  Tlalnepantla,  Estado  de  Mexico, 
Ctkligo  Postal  54500,  Mexico. 

* Corresponding  author; 

E-mail:  hgomez@nosferatu. ecologia. imam. mx 


with  the  males  (Chapman  1929;  Skutch  1940, 
1953).  Distinct  female  songs  have  more  re- 
cently been  recorded  in  other  Troglodytes 
wrens — one  population  of  northern  House 
Wren  (Johnson  and  Kermott  1990)  and  So- 
corro Wren  {T.  sissonii;  Howell  and  Webb 
1995).  During  fieldwork  on  the  life-history  of 
Sumichrast's  Wren  {Hylorchilus  sumichrasti) 
in  Cerro  de  Oro,  Oaxaca  (18°  02'  N,  96°  15' 
W;  Perez-Villafana  1997),  we  recorded  the 
previously  unknown  song  of  a female  H.  sum- 
ichrasti. 

Sumichrast’s  Wren  is  sexually  monomor- 
phic  in  plumage.  The  birds  we  observed  were 
not  color-banded;  however,  during  direct  ob- 
servation of  the  members  of  a single  pair  from 
March  to  July  1994,  we  realized  that  the  pre- 
viously unrecorded  song  was  always  made  by 
the  bird  that  did  not  emit  the  more  complex 
song  described  by  Howell  and  Webb  (1995) 
and  Gomez  de  Silva  (1997).  By  analogy  with 
other  wrens  that  have  strong  sexual  differenc- 
es in  song  (in  which  the  female’s  song  is  the 
simpler  one),  and  from  the  birds’  behaviors, 
we  concluded  that  this  previously  unrecorded 
song  was  the  song  of  the  female.  We  subse- 
quently have  heard  this  “female  song’’  at  dif- 
ferent points  along  a 738  m transect  at  Cerro 
de  Oro,  and  throughout  the  range  of  Sumi- 
chrast’s Wren:  2 km  south  of  Amatlan  (18°  50' 


SHORT  COMMUNICATIONS 


129 


kHz 


B 





1 Seconds 

FIG.  1.  Songs  of  female  (A)  and  male  (B)  Sumichrast’s  Wren  (Hylorchilus  sumichrasti).  Recorded  by  S.  N. 
G.  Howell  2 km  south  of  Amatlan,  Veracruz. 


N,  96°  55'  W),  Agua  Escondida  (18°  32'  N, 
96°  47'  W),  OaxacaA^eracruz  border  on  the 
road  to  San  Juan  del  Rio  (17°  32'  N,  95°  44' 
W),  and  2 km  south  of  Bethania  (17°  56'  N, 
96°  01'  W). 

The  distinction  between  songs  and  calls  is 
sometimes  unclear.  In  general,  calls  comprise 
one  or  two  syllables  whereas  songs  are  longer 
vocalizations  comprising  multiple  syllables 
(Langmore  1998).  The  “female  song”  of 
Sumichrast’s  Wren  is  a simple  phrase  consist- 
ing of  a single  repeated  syllable  (fundamental 
< 2 kHz).  In  this  respect,  it  resembles  the 
main  song  of  the  Cactus  Wren  (Campy lorhyn- 
chus  brunneicapillus)  rather  than  the  rich  and 
complex  songs  of  male  Sumichrast’s  and  of 
most  other  wrens.  The  female  song  had  4-22 
such  syllables  per  song.  The  pause  between 
the  first  and  second  syllables,  and  to  a lesser 
extent  the  pause  before  the  last  syllable,  are 
the  longest.  This  simple  song  of  uniform  fre- 
quency contrasts  with  the  males’  complex 


songs  which  spans  a range  of  frequencies  and 
contains  syllables  of  variable  form  (Fig.  1,  see 
also  other  sonograms  of  male  songs  in  Gomez 
de  Silva  1997  and  Atkinson  et  al.  1993). 

Females  sing  less  frequently  than  males. 
Along  the  transect  at  Cerro  de  Oro,  censused 
twice  a month  between  April  1994  and  March 
1995,  only  28.5-75%  as  many  females  were 
recorded  singing  per  morning  as  males.  Fe- 
males sometimes  countersang  or  sang  at  the 
same  time  as  males.  In  the  one  focal  pair,  the 
female  countersang  with  the  male  46.2%  of 
the  time.  Females  usually  sang  with  their  bod- 
ies held  upright  and  their  tails  pointing  down- 
ward, the  same  position  as  singing  males  in 
Cerro  de  Oro. 

Nonantiphonal  female  songs  may  be  more 
widespread  in  wrens  than  previously  thought. 
The  few  records  may  be  due  to  a scarcity  of 
detailed  observation.  Carmona  (1989)  ob- 
served that  female  Canyon  Wrens  (Catherpes 
rnexicanus)  produce  a vocalization  that  is  dif- 


130 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  1,  March  1999 


ferent  from  the  male’s  but  gave  no  details 
about  the  vocalization.  A systematic  survey  of 
female  songs  (or  their  general  absence)  in 
wrens  appears  to  be  an  interesting  possibility 
for  research. 

ACKNOWLEDGMENTS 

We  are  grateful  to  T.  Altamirano  of  the  Museo  de 
las  Ciencias  Biologicas  Iztacala,  for  partial  support  of 
M.  P.  V.’s  fieldwork  in  Cerro  de  Oro.  S.  N.  G.  Howell 
kindly  provided  valuable  comments  and  information 
as  well  as  recorded  the  songs  and  made  the  sonagrams 
shown  here.  L.  F.  Baptista  kindly  made  available  his 
sound  analysis  laboratory  at  the  California  Academy 
of  Sciences  and  offered  helpful  comments.  The  sug- 
gestions of  three  anonymous  reviewers  helped  improve 
the  manuscript. 

LITERATURE  CITED 

Atkinson,  P.  W.,  M.  Whittingham,  H.  Gomez  de  Sil- 
va Garza,  A.  M.  Kent  and  R.  T.  Maier.  1993. 
Notes  on  the  conservation,  ecology  and  taxonomic 
status  of  Hylorchilus  wrens.  Bird  Conserv.  Int.  3: 
75-85. 

Carmona,  R.  1989.  Contribucion  al  conocimiento  de 
la  historia  natural  de  Catherpes  mexicanus  (Trog- 
lodytidae:  Aves)  en  la  Reserva  Ecologica  del  Ped- 
regal  de  San  Angel,  Mexico,  D.F.  B.Sc.  thesis, 
UNAM — Campus  Iztacala,  Mexico,  D.F. 


Chapman,  F.  M.  1929.  My  tropical  air  castle.  Appleton, 
New  York. 

Gomez  de  Silva  G.,  H.  1997.  Comparative  analysis  of 
the  vocalizations  of  Hylorchilus  wrens  (Troglo- 
dytidae).  Condor  99:  981—984. 

Hardy,  J.  W.  and  D.  J.  Delaney.  1987.  The  vocali- 
zations of  the  Slender-billed  Wren  {Hylorchilus 
sumichrasti):  who  are  its  close  relatives?  Auk  104: 
528-530. 

Howell,  S.  N.  G.  and  S.  Webb.  1995.  A guide  to  the 
birds  of  Mexico  and  northern  Central  America. 
Oxford  Univ.  Press,  Oxford,  U.K. 

Johnson,  L.  S.  and  L.  H.  Kermott.  1990.  Structure 
and  context  of  female  song  in  a north-temperate 
population  of  House  Wrens.  J.  Field  Ornithol.  61: 
273-284. 

Langmore,  N.  E.  1998.  Functions  of  duet  and  solo 
songs  of  female  birds.  Trends  Ecol.  Evol.  13:136— 
140. 

Perez-Villafana,  M.  1997.  Contribucion  al  conoci- 
miento de  la  historia  de  vida  de  Hylorchilus  sum- 
ichrasti (Aves:  Troglodytidae)  en  el  norte  del  Es- 
tado  de  Oaxaca.  B.Sc.  thesis,  UNAM — Campus 
Iztacala,  Mexico,  D.F. 

Skutch,  a.  F.  1940.  Social  and  sleeping  habits  of  Cen- 
tral American  wrens.  Auk  57:293—312. 

Skutch,  A.  F.  1953.  Life  history  of  the  Southern  House 
Wren.  Condor  55:121—149. 

Skutch,  A.  F.  1960.  Life  histories  of  Central  American 
birds  II.  Cooper  Ornithological  Society,  Berkeley, 
California. 


Wilson  Bull,  111(1),  1999,  pp.  130-132 

An  Incident  of  Female-Female  Aggression  in  the  House  Wren 

Tom  Alworth'  and  Isabella  B.  R.  ScheibeF 


ABSTRACT. — In  this  paper  we  describe  one  ex- 
ample of  female-female  aggression  in  the  House  Wren 
{Troglodytes  aedon).  An  intruding  female  usurped  the 
resident  female  and  paired  with  the  resident  male. 
House  Wrens  are  known  for  committing  infanticide  as 
well  as  puncturing  and  removing  eggs  of  conspecifics 
and  other  species.  These  behaviors  have  been  mainly 
attributed  to  resident  and  floating  males,  but  we  sug- 
gest that  females  may  also  be  responsible.  Received  22 
.Inly  I99H.  accepted  3 Nov.  1998. 


' The  E.  N.  Huyck  Pre.serve  and  Biological  Research 
Station,  Rensselaerville,  NY  12147. 

^ Dept,  of  Biological  Sciences,  Univ.  at  Albany,  Al- 
bany, NY  12222. 

’ Corresponding  author; 

E-mail:  is5()4l  @cnsunix. albany.edu 


In  many  passerine  bird  species,  males  es- 
tablish breeding  territories  in  the  spring, 
which  they  defend  against  intruders.  This 
form  of  sexual  competition  among  males  has 
been  recognized  as  one  of  the  driving  forces 
behind  mating  patterns  and  parental  care  (Da- 
vies 1991,  Andersson  1994).  Aggression 
among  females  has  received  much  less  atten- 
tion, although  it  has  recently  been  shown  to 
be  more  common  among  birds  than  initially 
assumed  (Lenington  1980,  Leffelaar  and  Rob- 
ertson 1985,  Searcy  1986,  Martin  et  al.  1990, 
Slagsvold  1993,  Hansson  et  al.  1997,  Liker 


SHORT  COMMUNICATIONS 


131 


and  Szekely  1997).  Female-female  aggression 
may  affect  several  aspects  of  mating  systems 
and  parental  care,  for  example,  maintaining 
monogamy  (Slagsvold  1993)  or  reducing  har- 
em size  (Hurly  and  Robertson  1985).  Female- 
female  aggression  in  Red-winged  Blackbirds 
{Agelaius  phoeniceus’,  Beletsky  1996)  as  well 
as  Lapwings  (Vanellus  vonellus;  Liker  and 
Szekely  1997)  is  strongest  early  in  the  breed- 
ing season  when  females  first  settle  on  the  ter- 
ritories. One  evolutionary  force  behind  fe- 
male-female aggression  in  polygynous  mating 
systems  is  the  conflict  between  females  for  the 
male’s  parental  investment  (Slagsvold  and  Li- 
fjeld  1994).  However,  female-female  aggres- 
sion is  not  limited  to  polygynous  mating  sys- 
tems (Slagsvold  1993).  Here  we  report  an  in- 
cident of  female-female  aggression  in  the 
House  Wren  {Troglodytes  aedon). 

We  have  been  studying  a population  of 
House  Wrens  on  the  E.  N.  Huyck  Preserve 
and  Biological  Research  Station  in  Rensse- 
laer ville,  New  York  since  1992  and  have  color 
banded  all  individuals  since  1995.  On  17  May 
1997,  07:15  EST  we  observed  a fight  between 
two  females  that  lasted  for  30  minutes.  Fe- 
male A,  who  had  been  paired  to  the  resident 
male  of  the  territory  since  12  May  was  chased 
in  circles  both  in  the  air  and  on  the  ground  by 
female  B.  Female  B had  been  the  resident  fe- 
male of  the  same  territory  in  1996  but  was 
paired  with  a different  male.  The  1996  male 
was  not  seen  in  1997.  The  fight  included  chas- 
es by  female  B with  occasional  aggressive  in- 
teractions that  included  bodily  contact  and 
pecking.  Neither  female  vocalized  during  the 
encounter.  The  resident  male  was  perched  and 
visible  during  the  whole  fight;  he  sang  but  did 
not  participate  in  the  fight.  Female  B eventu- 
ally usurped  the  territory  from  female  A, 
paired  with  the  resident  male,  and  took  over 
the  nest  that  was  close  to  completion.  Female 
A was  not  seen  again  during  the  1997  breed- 
ing season.  We  suggest  that  female  B was 
probably  fighting  for  the  territory  rather  than 
for  the  resident  male.  One  of  the  most  suc- 
cessful males  in  our  study  population,  who 
was  polygynous  in  1996  and  1997,  occupied 
the  adjacent  territory  and  was  at  that  time  un- 
paired. Female  B did  not  pair  with  this  un- 
mated male  but  returned  to  the  territory  with 
which  she  was  familiar. 

Many  researchers  strongly  suspect  that  res- 


ident birds  and  probably  non-resident  floaters 
as  well  (Johnson  and  Kermott  1993)  routinely 
enter  territories  not  their  own  and  kill  and/or 
remove  eggs  or  young  from  nests  (Belles-lsles 
and  Pieman  1986,  1987;  Quinn  and  Holroyd 
1989;  Kermott  et  al.  1991).  Until  recently 
these  birds  have  been  assumed  to  be  male 
(Quinn  and  Holroyd  1989,  Kermott  et  al. 
1991),  but  it  now  appears  that  residents  need 
to  be  concerned  about  intruding  females  as 
well.  This  and  other  observations  of  female- 
female  aggression  (Freed  1986,  Johnson  and 
Searcy  1996)  demonstrate  that  female  House 
Wrens  may  play  an  equally  important  role  as 
the  males  in  the  selection  of  nest  sites,  terri- 
torial defense,  and  intraspecific  aggression. 

ACKNOWLEDGMENTS 

We  thank  K.  R Able  for  his  continuing  support,  and 
all  the  landowners  including  the  E.  N.  Huyck  Preserve 
for  allowing  us  to  put  up  our  nest  boxes.  We  are  grate- 
ful to  our  field  assistants.  This  paper  significantly  ben- 
efitted  from  the  review  and  comments  of  L.  S.  John- 
son. This  study  has  been  funded  by  the  E.  N.  Huyck 
Preserve  and  Biological  Research  Station  (to  T A.  and 
1.  S.)  and  the  New  York  State  Museum  Biological  Sur- 
vey (to  I.  S.) 

LITERATURE  CITED 

Andersson,  M.  1994.  Sexual  selection.  Princeton 
Univ.  Press,  Princeton,  New  Jersey. 

Belles-Isles,  J.  C.  and  J.  Picman.  1986.  House  Wren 
nest  destroying  behavior.  Condor  88:190-193. 
Belles-Isles,  J.  C.  and  J.  Picman.  1987.  Suspected 
intraspecific  killing  by  House  Wrens.  Wilson  Bull. 
99:497-498. 

Beletsky,  L.  1996.  The  Red-winged  Blackbird.  Aca- 
demic Press,  Ltd.,  London,  U.K. 

Davies,  N.  B.  1991.  Mating  systems.  Pp.  263-293  in 
Behavioral  ecology:  an  evolutionary  approach  (J. 
R.  Krebs  and  N.  B.  Davies,  Eds.).  Blackball  Sci- 
entific Publications,  London,  U.K. 

Freed,  L.  A.  1986.  Territory  takeover  and  sexually  se- 
lected infanticide  in  tropical  House  Wrens.  Behav. 
Ecol.  Sociobiol.  19:197-206. 

Hansson,  B.,  S.  Bensch,  and  D.  Hasselquist.  1997. 
Infanticide  in  Great  Reed  Warblers:  secondary  fe- 
males destroy  eggs  of  primary  females.  Anim.  Be- 
hav. 54:297-304. 

Hurly,  T.  A.  and  R.  J.  Robertson.  1985.  Do  female 
Red-winged  Blackbirds  limit  harem  size?  A re- 
moval experiment.  Auk  102:205-209. 

Johnson,  L.  S.  and  L.  H.  Kermott.  1993.  Why  is 
reduced  male  parental  assistance  detrimental  to 
the  reproductive  success  of  secondary  female 
House  Wrens?  Anim.  Behav.  46:1 1 1 1-1 120. 
Johnson,  L.  S.  and  W.  A.  Searcy  1996.  Female  at- 


132 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I.  March  1999 


traction  to  male  song  in  House  Wrens  (Trof^lo- 
dytes  aedon).  Behaviour  133:357-366. 

Kaufman,  J.  H.  1983.  On  the  dehnitions  and  functions 
of  dominance  and  territoriality.  Biol.  Rev.  58:1- 
20. 

Kermott,  L.  H.,  L.  S.  Johnson,  and  M.  S.  Merkle. 
1991.  Experimental  evidence  for  the  function  of 
mate  replacement  and  infanticide  by  males  in  a 
north-temperate  population  of  House  Wrens.  Con- 
dor 93:630-636. 

Leffelaar,  D.  and  R.  J.  Robertson.  1985.  Nest  usur- 
pation and  female  competition  for  breeding  op- 
portunities by  Tree  Swallows.  Wilson  Bull.  97: 
221-224. 

Lenington,  S.  1980.  Female  choice  and  polygyny  in 
Red-winged  Blackbirds.  Anim.  Behav.  28:347— 
361. 


Liker,  a.  and  T.  Szekely.  1997.  Aggression  among 
female  Lapwings,  Vanellu.s  vanellu.s.  Anim.  Be- 
hav. 54:797-802. 

Martin,  K.,  S.  J.  Hannon,  and  S.  Lord.  1990.  Fe- 
male-female aggression  in  White-tailed  Ptarmigan 
and  Willow  Ptarmigan  during  pre-incubation  pe- 
riod. Wilson  Bull.  102:532—536. 

Quinn,  M.  S.  and  G.  L.  Holroyd.  1989.  Nestling  and 
egg  destruction  by  House  Wrens.  Condor  91:206- 
207. 

Searcy,  W.  A.  1986.  Are  female  Red-winged  Black- 
birds tenitorial?  Anim.  Behav.  34:1381-1391. 

Slagsvold.  T.  1993.  Female-female  aggression  and 
monogamy  in  Great  Tits  Panes  major.  Ornis. 
Scand.  24:155-158. 

Slagsvold,  T.  and  T.  J.  Lifjeld.  1994.  Polygyny  in 
birds:  the  role  of  competition  between  females  for 
male  parental  eare.  Am.  Nat.  143:59-94. 


Wil.son  Bull.,  111(1),  1999,  pp.  132-133 


Nest  Reuse  by  Wood  Thrushes  and  Rose-breasted  Grosbeaks 

Lyle  E.  Friesen,'  - Valerie  E.  Wyatt,'  and  Michael  D.  Cadman' 


ABSTRACT. — We  report  on  two  instances  of  nest 
reuse  by  Wood  Thrushes  (Hylocichia  mu.stelina)  within 
the  same  breeding  season,  and  three  cases  of  nest  reuse 
in  successive  years,  two  by  Wood  Thrushes  and  one 
by  Rose-breasted  Grosbeaks  (Pheucticu.s  liidovici- 
anit.s).  In  each  of  the  five  cases  of  nest  reuse,  host 
young  were  successfully  fledged  in  the  original  nesting 
episode  and  in  the  .second  nesting  episode.  Although 
occasional  nest  reuse  within  a single  breeding  season 
has  been  reported  before,  our  study  is  the  first  to  doc- 
ument reuse  of  the  same  nest  in  successive  years  by 
Wood  Thrushes  and  Rose-breasted  Grosbeaks.  Re- 
ceived 29  April  1998,  accepted  4 Oct.  1998. 


Open  nesting  passerines,  with  the  possible 
exception  of  tyrannid  flycatchers  (Curson  et 
al.  1996),  seldom  reuse  nests  within  and  be- 
tween breeding  seasons  (Briskie  and  Sealy 
1988).  Earlier  authorities  (Weaver  1949, 
Brackbill  1958)  were  unaware  of  nest  reuse 
by  Wood  Thrushes  (Hylocichia  mu.stelina). 
Roth  and  coworkers  (1996)  documented  oc- 


' Canadian  Wildlife  Service.  75  Farquhar  Street, 
Ciuelph,  ON,  Canada  N I H 3N4. 

^ Corresponding  author: 

E-mail:  lyle.frie.sen@sympatico.ca 


casional  nest  reuse  by  Wood  Thrushes  during 
the  same  breeding  season  (three  cases  out  of 
389  first  nests).  However,  they  did  not  report 
any  nest  reuse  between  years,  describing  such 
an  event  as  unlikely  because  nests  usually  dis- 
integrate after  the  nesting  season.  We  report 
on  two  cases  of  nest  reuse  by  Wood  Thrushes 
within  a breeding  season  and  three  cases  of 
nest  reuse  in  successive  years,  two  by  Wood 
Thrushes  and  one  by  Rose-breasted  Grosbeaks 
(Pheucticus  lucloviciaitus). 

Data  presented  in  this  study  were  gathered 
in  the  course  of  a larger  study  of  the  nesting 
success  of  Wood  Thrushes  and  Rose-breasted 
Grosbeaks  conducted  in  1996  and  1997  in 
Waterloo  Region,  a fragmented  agricultural 
landscape  located  in  southwestern  Ontario 
(see  Friesen  et  al.,  in  press  for  a description 
of  the  landscape).  In  these  two  years,  154 
Wood  Thrush  nests  and  63  Rose-breasted 
Grosbeak  nests  were  found  and  regularly 
monitored  to  determine  their  outcome. 

Two  Wood  Thrush  nests  (one  each  year) 
were  reused  during  the  same  breeding  season, 
with  young  successfully  fledging  in  all  four 
nesting  attempts  (see  Wyatt  1997  for  a de- 


SHORT  COMMUNICA  TIONS 


133 


tailed  account  of  one  of  the  renests).  Neither 
of  these  nesting  attempts  were  parasitized  by 
Brown-headed  Cowbirds  (Molothriis  citer)  al- 
though 47%  of  Wood  Thrush  nests  on  our 
study  sites  contained  cowbird  eggs  or  young 
(Friesen  et  al.,  in  press).  In  1996,  20  days 
elapsed  between  the  fledging  of  the  first  brood 
and  the  initiation  of  the  second  clutch  (June 
20  to  July  10);  in  1997,  this  interval  was  13 
days  (June  25  to  July  8).  Neither  of  the  orig- 
inal nests  appeared  to  have  been  relined  or 
refurbished  prior  to  its  second  use.  It  is  likely 
that  the  same  pairs  reused  each  of  the  nests 
but  this  could  not  be  confirmed  because  the 
birds  were  not  color-banded.  Studies  of  band- 
ed birds  in  Waterloo  Region  in  1998  showed 
that  at  least  half  of  the  pairs  attempted  two 
broods  in  a nesting  season  (Friesen,  unpubl. 
data). 

We  mapped  and  marked  all  of  the  Wood 
Thrush  nests  found  in  1996  (?i  = 61)  and  ob- 
served, through  visits  to  the  sites  the  follow- 
ing spring  prior  to  the  breeding  season,  that 
five  (8%)  of  them  survived  the  winter  seem- 
ingly intact.  Two  of  these  nests  were  subse- 
quently reused  in  1997:  one  nest  which 
fledged  three  Wood  Thrushes  and  one  cowbird 
in  1996,  fledged  four  thrushes  and  two  cow- 
birds  the  following  year;  the  other  nest  fledged 
two  thrushes  and  one  cowbird  in  each  of  the 
years.  Neither  nest  appeared  to  have  been  sig- 
nificantly renovated  in  the  second  year,  al- 
though both  were  in  poor  repair  by  the  time 
the  young  fledged  in  1997. 

Rose-breasted  Grosbeaks  are  typically  sin- 
gle brooded  in  southern  Ontario  (Friesen  et 
al.,  in  press)  and  we  found  no  evidence  of  nest 
reuse  within  the  same  breeding  season.  Three 
(12%)  of  the  24  grosbeak  nests  we  found  in 
1996  survived  the  winter.  One  of  these,  in 
which  three  young  were  fledged  in  1996,  was 
reused  in  1997  and  again  fledged  three  young. 

Our  results  suggest  that  nest  reuse  is  a con- 


sistent, albeit  infrequent,  breeding  strategy. 
The  reuse  of  old  nests  may  have  resulted  from 
a shortage  of  suitable  nesting  sites  although  it 
seemed  to  us  that  apparently  suitable  alterna- 
tive sites  were  present  nearby.  It  may  also  be 
that  the  birds  reusing  nests  recognized  the  lat- 
ter as  being  of  high  quality,  borne  out  by  the 
fact  that  all  five  nest  reuses  resulted  in  fledged 
host  young.  Our  study  is  a reminder  to  re- 
searchers of  the  importance  of  monitoring  the 
status  of  used  nests  both  within  and  between 
breeding  seasons. 

ACKNOWLEDGMENTS 

We  thank  our  tielci  assi.stants  D.  Dieboldt.  E.  Mc- 
Leish,  M.  Nighswander.  R.  Nonas,  B.  Pollock.  A. 
Spender,  E.  Stephens,  N.  Wessely,  and  S.  Zaheer.  Parts 
of  this  project  were  supported  by  the  Ontario  Region 
of  Environment  Canada's  Canadian  Wildlife  Service. 
Human  Resources  Development  Canada,  Environmen- 
tal Youth  Corps-Ontario,  Long  Point  Bird  Observato- 
ry, and  the  Regional  Municipality  of  Waterloo. 

LITERATURE  CITED 

Brackbill,  H,  1958.  Nesting  behavior  of  the  Wood 
Thrush.  Wilson  Bull.  70:70-89. 

Briskie,  J.  V.  AND  S.  G.  Sealy.  1988.  Nest  reuse  and 
egg  burial  in  the  Least  Elycatcher  Empidonax  inin- 
iimis.  Can.  Eield-Nat.  102:729-730. 

CuRSON,  D.  R..  C.  B.  Goguen,  and  N.  E.  Mathews. 
1996.  Nest-site  reu.se  in  the  Western  Wood-Pewee. 
Wilson  Bull.  108:378-380. 

Friesen,  L.  E.,  M.  D.  Cadman,  and  R.  J.  MacKav.  In 
press.  Ne.sting  success  of  Neotropical  migrant 
songbirds  in  a highly  fragmented  landscape.  Con- 
serv.  Biol. 

Roth,  R.  R.,  M.  S.  John.son,  and  T.  J.  Underwood. 
1996.  Wood  Thrush  (Hylocichla  miistelina).  In 
The  birds  of  North  America,  no.  246  (A.  Poole 
and  E Gill,  Eds.).  The  Academy  of  Natural  Sci- 
ences, Philadelphia,  Pennsylvania;  The  American 
Ornithologists’  Union.  Washington.  D.C. 

Weaver,  F.  G.  1949.  Wood  Thrush.  U.S.  Natl.  Mus. 
Bull.  196:101-123. 

Wyatt,  V.  E.  1997.  Nest  reused  by  Wood  Thrush.  Ont. 
Birds  15(l):36-37. 


134 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


Wilson  Bull.,  111(1),  1999,  pp.  134-137 


Singing  in  a Mated  Female  Wilson’s  Warbler 

William  M.  Gilbert'-^  and  Adele  F.  Carroll- 


ABSTRACT. — A female  Wilson’s  Warbler,  Wilson- 
ia  pusilla,  was  heard  singing  regularly  on  the  territory 
of  a male  in  middle  inner-coastal  California  during  ear- 
ly April,  1996,  and  occasionally  after  mid-April.  Based 
on  their  behavior,  the  resident  male  and  the  singing 
female  were  paired.  The  female  averaged  about  four 
songs/min  during  singing  bouts,  and  was  heard  about 
30%  of  the  time  during  early  April.  The  female’s  song 
was  high  pitched,  and  did  not  resemble  typical  male 
“chatter”  song.  In  contrast  to  the  functions  of  female 
song  in  many  tropical  and  some  temperate  parulids, 
this  song  seemed  to  serve  as  a simple  contact  vocali- 
zation between  mates,  as  call  notes  might.  A single 
female  song  heard  in  a newly  formed  pair  in  1997 
raises  the  possibility  that  such  songs  might  function  in 
pair  formation.  Received  10  April  1998,  accepted  25 
Aii^.  1998. 


Singing  by  female  birds,  while  common 
and  perhaps  characteristic  among  tropical  spe- 
cies, occurs  much  less  commonly  among  tem- 
perate species  (Morton  1996).  Tropical  and 
temperate  wood  warblers  follow  this  pattern. 
Songs  of  tropical  parulid  females  commonly 
are  used  in  “duets”  with  mates  and  may  func- 
tion in  pair  formation,  communication  with 
mate,  and  territorial  defense  (Spector  1992). 
To  serve  such  functions,  consistency  in  oc- 
currence and  stereotypy  in  form  of  song  seem 
to  be  required.  Song  in  temperate  parulid  fe- 
males, where  reported,  has  typically  occurred 
in  few  females  within  a population,  and  in 
some  cases  song  patterns  have  varied  among 
females  (e.g.,  Nolan  1978,  Hobson  and  Sealy 
1990).  This  suggests  that  song  in  temperate 
parulid  females  is  idiosyncratic,  and/or  serves 
very  limited  and  infrequent  functions.  To  our 
knowledge,  female  song  has  been  reported  in 
ten  temperate  parulid  species  from  six  genera: 
Vermivora,  Panda,  Dendroica,  Setophaga, 
Seiurus,  and  Geothlypis  (Spector  1992,  Mol- 
denhauer  and  Regelski  1996).  Here  we  report 


' 4630  Driftwood  Ct.,  El  Sobrante,  CA  94803; 
E-mail:  wmglbrt@aol.com 

2 I 147  Fre.sno  Ave.,  Berkeley,  CA  94707. 

' Corresponding  author. 


singing  in  a mated  female  Wilson’s  Warbler, 
Wdsonia  pusdla. 

We  first  heard  the  female  song  on  3 April, 
1996,  from  the  territory  (ca  0.2  ha)  of  a color- 
banded  resident  male  (the  “male”)  Wilson’s 
Warbler  in  the  Nature  Study  Area  of  Tilden 
Regional  Park,  Contra  Costa  Co.,  California. 
We  subsequently  heard  this  unusual  (com- 
pared with  typical  male  “chatter”  song;  Fig. 
IB)  and  distinctive  song  at  various  times  be- 
tween 07:15  and  11:40  PST  through  18  April. 
The  high  frequency  song  sounded  “sharp” 
and  “squeaky”  (Fig.  lA,  C).  It  was  delivered 
at  a rate  of  4.0  ± 0.6  songs/min  (range  = 2- 
6,  n = 7)  within  singing  bouts  (we  considered 
a singing  bout  to  be  continuous  singing  with 
no  pause  greater  than  one  minute  between 
songs).  On  four  separate  occasions,  we  ob- 
served the  beak  of  a Wilson’s  Warbler  to  open 
and  move  as  the  female  song  was  heard,  con- 
firming that  the  song  came  from  the  species. 
Behavioral  observations  (see  below)  indicate 
that  the  singer  was  the  resident  female  (the 
“female”)  in  the  territory  she  occupied.  This 
female  also  frequently  chipped  within  this  ter- 
ritory (Fig.  1C). 

From  3 through  12  April  we  located  the  fe- 
male during  30%  of  our  observation  time  (490 
min)  based  on  hearing  her  song  and/or  sight- 
ing the  singing  bird.  We  confirmed  that  the 
male  was  in  her  proximity  20%  of  our  obser- 
vation time  (the  male  usually  did  not  sing  and 
often  was  more  difficult  to  locate).  Bouts  of 
female  song  that  we  monitored  lasted  11.5  ± 

3.0  min  (range  = 1.5-27,  n = 11),  and  male- 
female  separation  distance  was  6.8  ± 1.2  m 
(range  = 2-20,  n = 22).  From  13  through  30 
April,  we  located  the  female  just  5%  of  our 
observations  time  (825  min),  and  located  the 
male  in  her  proximity  only  2%  of  that  time. 
Bouts  of  female  song  that  we  monitored  lasted 

3.1  ± 1.3  min  (range  = 1. 5-7.0,  n = 4),  and 
mean  male-female  separation  distance  was  4.4 
± 1.2  m (range  = 0-12,  n = 11).  Three  con- 
secutive singing  bouts  monitored  on  18  April 


SHORT  COMMUNICATIONS 


135 


FIG.  1.  Wilson’s  Warblers  vocalizations  recorded  in  April,  1996,  within  a breeding  territory  at  a study  site 
in  Contra  Costa  Co.,  California.  A.  Eight-note  song,  sung  by  resident  female,  with  resident  male  in  close 
proximity.  B.  A typical  “chatter  song”  of  resident  male.  C.  Three  call  notes,  followed  by  a 6-note  song,  by 
resident  female.  Recordings  made  by  W.  M.  Gilbert  using  a SONY  TCD-D8  recorder  and  a Sennheiser  K3U 
microphone.  Spectrograms  made  with  Canary  1.2.1  software  (Cornell  Laboratory  of  Ornithology)  using  a Mac- 
intosh 7.5  system  computer. 


had  separation  intervals  of  30  and  44  min.  At 
no  time  during  our  observations  did  we  notice 
unusual  resident  male  behavior  (compared 
with  other  mated  males)  that  might  have  elic- 
ited singing  in  the  female. 

We  relied  on  several  contextual  clues  to  de- 
termine the  sex  of  the  bird  singing  the  unusual 
song  (the  sexes  of  west  coast  Wilson’s  War- 


blers often  are  indistinguishable  in  the  field), 
and  that  bird’s  relationship  to  the  resident 
male  on  whose  territory  it  sang:  (1)  the  male 
was  mated  (although  we  found  no  direct  evi- 
dence of  breeding,  unpaired  males  tend  to  sing 
persistently,  often  from  exposed  perches, 
while  this  male  sang  sporadically,  often  from 
undergrowth);  (2)  we  consistently  heard  the 


136 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


unusual  song  from  within  the  male’s  territory, 
but  not  from  beyond  its  borders;  (3)  about 
two-thirds  of  the  time  that  we  heard  the  un- 
usual song,  the  male  was  sighted  within  25  m 
of  the  singer,  and  the  male  and  the  singer  often 
foraged  in  the  same  tree,  in  adjacent  trees,  or 
in  the  same  restricted  area  of  undergrowth;  (4) 
even  though  the  two  birds  frequently  were 
close,  they  never  were  in  conflict;  (5)  we  nev- 
er simultaneously  sighted  nor  heard  the  two 
birds  at  opposite  ends  of  the  territory;  and  (6) 
we  never  consistently  sighted  a third  Wilson’s 
Warbler  within  the  male’s  territory,  as  we 
would  have  had  the  male  been  mated  to  a bird 
other  than  the  one  singing  the  unusual  song. 
Based  on  this  evidence,  we  concluded  that  the 
bird  singing  the  unusual  song  was  a female 
mated  with  the  resident  male  on  the  territory 
where  she  sang. 

On  27  March,  1997,  WMG  observed  a 
newly  formed  (<3  h)  pair  of  Wilson’s  War- 
blers on  a territory  adjacent  to  that  in  which 
we  heard  the  female  song  in  1996.  The  new 
pair  was  in  view  for  41  min  and  the  color- 
banded  male  followed  the  unbanded  female 
through  vegetation  ranging  from  undergrowth 
to  tree  canopy.  Amid  continuous  chipping 
from  the  pair,  WMG  heard  one  song  indistin- 
guishable by  ear  from  the  female  songs  heard 
in  1996.  It  could  not  be  determined  which  bird 
delivered  the  song,  nor  if  the  female  of  the 
pair  was  the  same  bird  that  sang  the  female 
songs  in  1996. 

Information  on  female  song  in  temperate 
parulid  species  is  limited.  In  female  Prairie 
Warblers  (Dendroica  discolor),  a limited  num- 
ber of  females’  songs  heard  were  all  simple 
and  perhaps  “primitive,”  highly  variable 
among  individuals,  and  unlike  normal  male 
songs  (although  some  were  identifiable  to  spe- 
cies). These  songs  were  delivered  early  in  the 
.season,  were  heard  sporadically,  and  may 
have  been  delivered  by  older  females  display- 
ing more  male-like  behavior  (Nolan  1978).  In 
Yellow  Warblers  (Dendroica  petechia),  Hob- 
son and  Sealy  (1990)  suggest  that  female  song 
can  function  in  intrasexual  conflicts  within 
very  dense  breeding  populations. 

The  singing  we  heard  from  a female  Wil- 
son’s Warbler(s)  occurred  early  in  the  season 
and  possibly  came  from  an  older  bird(s),  as 
would  be  consistent  with  some  findings  of  No- 
lan (1978)  for  the  Prairie  Warbler  and  with 


some  correlates  of  female  song  in  other  spe- 
cies where  usually  only  the  male  sings  (Nice 
1943).  Other  aspects  of  the  singing  we  heard 
appear  to  differ  from  what  occurs  in  females 
of  some  other  parulids,  however.  The  songs 
we  heard  were  not  used  in  duets  with  mates, 
as  in  many  tropical  species  (Spector  1992). 
There  was  no  evidence  that  they  functioned  in 
intrasexual  conflicts  within  dense  populations 
(Hobson  and  Sealy  1990).  Finally,  we  heard 
the  female  consistently  for  more  than  a week 
in  early  April,  1996,  as  opposed  to  isolated 
bouts  of  singing  heard  on  single  days  (Nolan 
1978). 

The  female  songs  we  heard  in  1996  (as  well 
as  the  single  song  heard  in  1997)  were  inter- 
spersed between  chip  notes  and  usually  deliv- 
ered with  the  resident  male  close  to  the  fe- 
male. This  suggests  the  songs  may  have  func- 
tioned to  communicate  with  a mate.  The  sin- 
gle song  heard  in  1997  (if  delivered  by  a 
female  and  one  different  from  the  singer  of 
1996)  introduces  the  additional  possibility  that 
the  song  could  serve  a special  communicatory 
function  during  early  Wilson’s  Warbler  pair- 
ing. If  so,  then  the  persistent  female  singing 
heard  in  1996  would  have  been  an  abnormal 
carry-over  of  that  behavior  into  the  nesting 
period.  Singing  in  that  female  may  have  re- 
flected an  abnormal  hormonal  balance,  similar 
to  effects  of  testosterone  injection  in  stimu- 
lating song  in  female  birds  that  normally  don’t 
sing  (e.g.,  Baptista  and  Morton  1988). 

ACKNOWLEDGMENTS 

We  thank  the  East  Bay  Regional  Park  District  for 
authorization  to  conduct  research  in  the  Tilden  Nature 
Area  of  Tilden  Regional  Park.  We  thank  D.  Kroodsma 
and  an  anonymous  reviewer  for  their  comments  and 
critical  review  of  this  paper. 

LITERATURE  CITED 

Baptista,  L.  E and  M.  L.  Morton.  1988.  Song  learn- 
ing in  montane  White-crowned  SpaiTOWs;  from 
whom  and  when.  Anim.  Behav.  36:1753—1764. 
Hobson,  K.  A.  and  S.  G.  Sealy.  1990.  Female  song 
in  the  Yellow  Warbler.  Condor  92:259—261. 
Moldenhauer,  R.  R.  and  D.  J.  Regelski.  1996.  North- 
ern Parula  (Parula  americami).  In  The  birds  of 
North  America,  no.  215  (A.  Poole  and  E Gill, 
Eds.).  The  Academy  of  Natural  Sciences,  Phila- 
delphia, Pennsylvania;  The  American  Ornitholo- 
gists’ Union,  Washington,  D.C. 

Morton,  E.  S.  1996.  A comparison  of  vocal  behavior 
among  tropical  and  temperate  passerine  birds.  Pp. 


SHORT  COMMUNICATIONS 


137 


258-28 1 /■/(  Ecology  and  evolution  of  acoustic  coni- 
iminication  in  birds  (D.  Kroodsma  and  E.  Miller, 
Eds.).  Cornell  LIniv.  Press,  Ithaca,  New  York. 

Nice,  M.  M.  1943.  Studies  in  the  life  history  of  the 
Song  SpaiTow,  part  2.  Trans.  Linnaean  Soc.  N.Y. 
6:1-328. 


Nolan,  V.,  Jk.  1978.  fhe  ecology  and  behavior  ol  the 
Prairie  Warbler  Deiulroiia  discolor.  Ornithol. 
Monogr.  26:1-595. 

Spector,  D.  a.  1992.  Wt)od-warbler  song  systems:  a 
review  of  paruline  singing  behaviors.  Curr.  Orni- 
thol. 9:199-243. 


Wilson  Bull..  111(1),  1999,  pp.  137-139 


Laying  Time  of  the  Bronzed  Cowbird 

Brian  D.  Peer'  --^  and  Spencer  G.  Sealy' 


ABSTRACT. — We  report  the  first  observations  of 
egg  laying  by  the  parasitic  Bronzed  Cowbird  (Molotli- 
rus  aeneus).  Three  direct  observations  and  two  esti- 
mates of  laying  times  were  made  at  two  Northern  Car- 
dinal (Cardincdis  cardinalis)  nests.  Bronzed  Cowbirds 
laid  at  18.2  min  ± 1.7  (SE)  before  sunrise  (range  14— 
24  min).  Laying  lasted  5-10  seconds.  Although  the 
parasitic  Brown-headed  Cowbird  (M.  ater)  and  some- 
times Shiny  Cowbirds  (M.  bonariensis)  also  lay  before 
sunrise,  direct  observations  of  laying  by  other  cow- 
birds are  required  before  it  can  be  concluded  that  pre- 
sunrise laying  is  an  adaptation  for  brood  parasitism. 
Received  9 June  1998,  accepted  5 Sept.  1998. 


Avian  brood  parasites  that  are  surreptitious 
when  parasitizing  nests  may  avoid  detection 
by  their  hosts.  Indeed,  they  often  lay  their 
eggs  in  a matter  of  seconds;  Sealy  and  co- 
workers (1995)  found  this  behavior  to  be 
unique  to  the  diverse  groups  of  brood  para- 
sites. The  parasitic  Brown-headed  Cowbird 
(Molothrus  ater)  generally  lays  in  the  minutes 
prior  to  sunrise  and  it  has  been  suggested  that 
laying  at  this  time,  presumably  when  hosts  are 
less  likely  to  be  at  their  nests,  is  an  adaptation 
for  brood  parasitism  (Chance  and  Hann  1942). 
Scott  (1991)  found  that  female  Brown-headed 
Cowbirds  lay  their  eggs  an  average  of  9 min 
before  sunrise,  whereas  seven  potential  host 
species  all  lay  their  eggs  after  sunrise.  Shiny 
Cowbirds  (M.  bonariensis)  and,  possibly,  the 
nonparasitic  Bay-winged  Cowbird  (M.  bacbu.s) 


' Dept,  of  Zoology,  Univ.  of  Manitoba,  Winnipeg, 
MB.  R3T  2N2,  Canada. 

^Present  address:  3163  5th  St.,  East  Moline,  IL 
61244;  E-mail:  bdpcowbird(5> aol.com 
^ Corresponding  author. 


also  sometimes  lay  before  sunrise  (see  Scott 
1991),  but  the  data  available  to  Scott  (1991) 
were  insufficient  to  conclude  that  sunrise  lay- 
ing is  an  adaptation  for  brood  parasitism. 
There  were  no  direct  observations  of  laying 
for  the  Bronzed  Cowbird  (M.  aeneus).  Carter 
(1986)  stated  only  that  this  brood  parasite  lays 
“during  dawn  hours”.  Here  we  report,  to  our 
knowledge,  the  first  recorded  observations  of 
laying  times  for  the  Bronzed  Cowbird. 

METHODS 

Our  observations  were  made  at  the  Welder  Wildlife 
Refuge  in  San  Patricio  County,  Texas  (28°  0'  N,  97°  5' 
W)  in  1994.  Both  Bronzed  and  Brown-headed  cow- 
birds were  present  during  the  breeding  season.  After 
locating  a nest  at  which  a host  apparently  had  not  com- 
pleted laying,  we  watched  it  the  following  morning 
beginning  approximately  30  min  before  sunrise.  We 
hid  far  enough  away  so  that  hosts  or  visiting  cowbirds 
were  not  disturbed.  The  nests  were  observed  with  bin- 
oculars when  necessary.  Sunrise  (SR)  times  were  ob- 
tained from  the  website  of  the  United  States  Naval 
Observatory  Astronomical  Applications  Department 
(http://aa.usno.navy.mil/AA/).  All  times  are  Central 
Standard  Time. 

RESULTS 

Three  Bronzed  Cowbird  laying  events  were 
observed  directly,  all  at  Northern  Cardinal 
{Cardinalis  cardinalis)  nests.  On  30  May 
1994  we  located  a cardinal  nest  (94-16)  con- 
taining one  cardinal  egg.  The  following  morn- 
ing, BDP  arrived  at  this  nest  at  05:14  (SR  — 
20  min)  and  found  a Bronzed  Cowbird  egg 
that  was  slimy,  suggesting  it  had  been  laid  re- 
cently, plus  one  cracked  cardinal  egg.  Later 
the  same  day  the  nest  contained  two  cardinal 
eggs  plus  the  cowbird  egg.  At  05:06  (SR  - 


138 


THE  WILSON  BULLETIN  • Vol.  111.  No.  I.  March  1999 


28  min)  on  1 June  the  nest  contents  were  the 
same,  but  by  05:14  a second  Bronzed  Cow- 
bird  egg  had  been  laid.  Later  that  day,  the 
damaged  cardinal  egg  was  gone  and  a third 
cardinal  egg  had  been  laid.  On  the  morning  of 
2 June  the  female  cardinal  was  accidentally 
flushed  from  the  nest,  and  the  nest  contents 
were  the  same  as  the  day  before.  At  05:14  (SR 
— 20  min)  BDP  watched  a female  Bronzed 
Cowbird  fly  directly  to  the  nest,  lay  an  egg, 
and  fly  away  in  5-10  s.  BDP  left  momentarily 
at  05:15,  but  upon  returning  at  05:19,  found 
four  Bronzed  Cowbird  eggs  in  the  nest.  No 
bird  species  is  known  to  lay  more  than  one 
egg  per  day  (Sturkie  1976),  thus  we  assumed 
a second  female  had  laid  an  egg  in  this  nest. 

Nest  94-22  was  found  with  one  cardinal  egg 
on  14  June  1994.  On  15  June  a female 
Bronzed  Cowbird  looked  into  the  nest  at  5:19 
(SR  — 14  min),  but  she  did  not  lay.  On  16 
June  K.  Stewart  observed  two  Bronzed  Cow- 
birds  parasitize  this  nest,  one  at  05:17  (SR  — 
16  min)  and  a second  at  05:19  (SR  — 14  min). 
Neither  cardinal  was  present  during  the  laying 
events  and  both  laying  bouts  lasted  5-10  s.  In 
addition  to  the  three  laying  events  observed, 
we  estimated  the  two  other  laying  times  by 
taking  the  midpoints  of  repeated  visits  to  the 
nests  (Scott  1991)  and  found  that  Bronzed 
Cowbirds  laid  their  eggs  18.2  min  ± 1.7  SE 
before  sunrise  (range,  SR  — 14  to  24  min). 

DISCUSSION 

Like  the  Brown-headed  Cowbird,  and 
sometimes  the  Shiny  Cowbird  (Scott  1991), 
Bronzed  Cowbirds  lay  prior  to  sunrise  and, 
similar  to  other  brood  parasites,  they  lay  rap- 
idly (Sealy  et  al.  1995).  It  is  undoubtedly  ad- 
vantageous for  brood  parasites  to  lay  their 
eggs  when  hosts  are  absent  and  to  lay  as  rap- 
idly as  possible  (reviewed  in  Sealy  et  al. 
1995).  Indeed,  Neudorf  and  Sealy  (1994) 
found  that  hosts  of  the  Brown-headed  Cow- 
bird at  Delta  Marsh,  Manitoba,  that  did  not 
roost  on  their  nests  overnight  typically  arrived 
at  the  nests  in  the  morning  after  cowbird  par- 
asitism would  have  occurred.  Female  Bronzed 
Cowbirds  have  been  observed  entering  host 
nests  at  various  times  of  the  day,  but  it  is  un- 
known whether  eggs  were  laid  (Thurber  and 
Villeda  1980;  T.  Brush,  pers.  comm.).  While 
it  is  possible  that  eggs  were  laid  during  these 
visits,  these  females  may  have  been  inspecting 


nests  (see  below),  or  they  may  have  punctured 
host  eggs  (Carter  1986,  Peer  1998). 

The  female  Bronzed  Cowbird  observed  vis- 
iting a nest  prior  to  sunrise  without  laying 
may  have  been  inspecting  this  nest  to  deter- 
mine whether  it  was  active  and  ready  to  be 
parasitized  (see  also  Mayfield  1961,  Nolan 
1978).  This  nest  was  parasitized  by  two 
Bronzed  Cowbirds  the  following  morning. 
The  cowbirds  were  clearly  aware  of  the  nest 
beforehand  because  they  flew  directly  to  it. 
Similar  behavior  has  been  reported  for 
Brown-headed  Cowbirds  (Hann  1941,  Neu- 
dorf and  Sealy  1994). 

The  three  cowbird  species  mentioned  above 
are  the  only  icterids  known  to  lay  before  sun- 
rise. The  nonparasitic  Bay-winged  Cowbird 
may  also  lay  prior  to  sunrise  (see  Scott  1991). 
Direct  observations  of  laying  by  the  Bay- 
winged, Giant  {Scaphidura  oryzivora),  and 
Screaming  (M.  rufoaxillaris)  cowbirds  are  re- 
quired before  it  can  be  concluded  that  pre- 
sunrise laying  in  cowbirds  is  an  adaptation  for 
parasitism. 

ACKNOWLEDGMENTS 

The  Welder  Wildlife  Refuge  provided  accommoda- 
tion and  logistical  assistance  along  with  M.  L.  Peer. 
We  are  grateful  to  K.  Stewart  who  observed  two 
Bronzed  Cowbirds  lay  in  the  same  nest  in  one  morn- 
ing. Constructive  comments  by  T.  Brush,  D.  Burhans 
and  two  anonymous  reviewers  improved  the  manu- 
script. This  research  was  supported  by  a research  grant 
from  the  Natural  Sciences  and  Engineering  Research 
Council  of  Canada  to  S.G.S.  and  a G.  A.  Lubinsky 
Memorial  Scholarship  from  the  Department  of  Zool- 
ogy, University  of  Manitoba  to  B.D.P. 

LITERATURE  CITED 

Chance,  E.  P.  and  H.  W.  Hann.  1942.  The  European 
Cuckoo  and  the  cowbird.  Bird-Banding  13:99- 
103. 

Carter,  M.  D.  1986.  The  parasitic  behavior  of  the 
Bronzed  Cowbird  in  south  Texas.  Condor  88:1  1- 
25. 

Hann,  H.  W.  1941.  The  cowbird  at  the  nest.  Wilson 
Bull.  53:21  1-221. 

Mayfield,  H.  E 1961.  Vestiges  of  a proprietary  inter- 
est in  nests  by  the  Brown-headed  Cowbird  para- 
sitizing the  Kirtland's  Warbler.  Auk  78:162-166. 
Neudorf,  D.  L.  and  S.  G.  Sealy.  1994.  Sunrise  nest 
attentiveness  in  cowbird  hosts.  Condor  96:162— 
169. 

Nolan,  V.,  Jr.  1978.  The  ecology  and  behavior  of  the 
Prairie  Warhler  Deiulroica  di.scolor.  Ornithol. 
Monogr.  26:1-595. 


SHORT  COMMUNICATIONS 


139 


Peer,  B.  D.  1998.  An  experimental  investigation  of 
egg  rejection  behavior  in  the  grackles  (Quiscatus). 
Ph.D.  diss.,  Univ.  of  Manitoba,  Winnipeg. 

Scott,  D.  M.  1991.  The  time  of  day  of  egg  laying  by 
the  Brown-headed  Cowbird  and  other  icterines. 
Can.  J.  Zool.  69:2093-2099. 

Sealy,  S.  G.,  D.  L.  Neudorf,  and  D.  P.  Hill.  1995. 


Rapid  laying  by  Brown-headed  Cowbirds  Moloih- 
riis  ater  and  other  parasitic  birds.  Ibis  137:76-84. 

Sturkie,  P.  D.  1976.  Avian  physiology,  third  ed. 
Springer- Verlag,  New  York. 

Thurber,  W.  a.  and  a.  Villeda.  1980.  Notes  on  par- 
asitism by  Bronzed  Cowbirds  in  El  Salvador.  Wil- 
son Bull.  92:1 12-1  13. 


Wilson  Bull.,  111(1),  1999,  pp.  139-143 


Temporal  Differences  in  Point  Counts  of  Bottomland  Forest  Landbirds 

Winston  Paul  Smith and  Daniel  J.  Twedt- 


ABSTRACT. — We  compared  number  of  avian  spe- 
cies and  individuals  in  morning  and  evening  point 
counts  during  the  breeding  season  and  during  winter  in 
a bottomland  hardwood  forest  in  west-central  Mississip- 
pi, USA.  In  both  seasons,  more  species  and  individuals 
were  recorded  during  morning  counts  than  during  even- 
ing counts.  We  also  compared  morning  and  evening  de- 
tections for  18  species  during  the  breeding  season  and 
9 species  during  winter.  Blue  Jay  {Cyanocitta  cristata). 
Mourning  Dove  (Zenciida  macroura),  and  Red-bellied 
Woodpecker  (Melanerpes  caroliniLs)  were  detected  sig- 
nificantly more  often  in  morning  counts  than  in  evening 
counts  during  the  breeding  season.  Tufted  Titmouse 
IBaeoloplms  bicolor)  was  recorded  more  often  in  morn- 
ing counts  than  evening  counts  during  the  breeding  sea- 
son and  during  winter.  No  species  was  detected  more 
often  in  evening  counts.  Thus,  evening  point  counts  of 
birds  during  either  the  breeding  season  or  winter  will 
likely  underestimate  species  richness,  overall  avian 
abundance,  and  the  abundance  of  some  individual  spe- 
cies in  bottomland  hardwood  forests.  Received  15  Nov. 
1997.  accepted  20  Aug.  1998. 


Improvement  and  standardization  of  assess- 
ment techniques  for  monitoring  bird  popula- 
tions has  received  considerable  attention  (e.g., 
Ralph  et  al.,  1993,  1995a,  b;  Hamel  et  al. 
1996).  Although  most  studies  of  avian  popu- 
lation assessment  techniques  have  focused  on 
breeding  birds,  some  have  evaluated  winter 


' United  States  Department  of  Agriculture,  Forest 
Service,  Pacific  Northwest  Re.search  Station,  Forestry 
Sciences  Laboratory,  2770  Sherwood  Lane — Suite  2A, 
Juneau,  AK  99801-8545;  E-mail:  wpaulsmith@aol.com 
^ USGS  Patuxent  Wildlife  Research  Center.  2524 
South  Frontage  Road,  Vicksburg,  MS  39180. 

^ Corresponding  author. 


bird  populations  (Rollfinke  and  Yahner  1990; 
Gutzwiller  1991,  1993a,  b).  Detecting  statis- 
tically significant  changes  in  avian  popula- 
tions may  require  an  extensive  monitoring 
network  (Smith  et  al.  1993,  Hamel  et  al. 
1996).  To  achieve  monitoring  objectives  using 
limited  resources,  protocols  that  reduce  costs 
and  maximize  efficiency  are  required  (Smith 
et  al.  1993).  Unfortunately,  many  factors  that 
influence  survey  efficiency  are  beyond  the 
control  of  investigators.  For  example,  detec- 
tion varies  among  species,  among  census  tech- 
niques (e.g.,  Grue  et  al.  1981,  Rollfinke  and 
Yahner  1990),  and  may  be  influenced  by  phys- 
ical or  biological  factors  (Gutzwiller  1993a, 
b). 

If  detection  probabilities  were  constant  over 
time,  the  efficiency  of  avian  surveys  could  be 
increased  by  providing  a greater  window  of 
opportunity  during  which  surveys  could  be 
conducted.  However,  most  species  exhibit  diel 
and  seasonal  variation  in  detectability.  Thus, 
to  optimize  sampling  effort  and  reduce  sam- 
pling variances,  monitoring  should  be  focused 
on  periods  when  species  are  most  frequently 
detected  (Gutzwiller  1993a). 

To  assess  optimal  periods  of  detection,  in- 
vestigators have  compared  point  counts  from 
different  times  of  the  morning  during  the 
breeding  season  (Shields  1977,  Grue  et  al. 
1981,  Robbins  1981,  Skirvin  1981)  or  winter 
(Gutzwiller  1993a).  Only  Rollfinke  and  Yah- 
ner (1990),  using  tran,sect  counts,  compared 
morning  counts  to  evening  counts  during  win- 
ter. Although  birds  are  generally  assumed  to 


140 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  /,  March  1999 


be  more  detectable  during  morning  than  dur- 
ing evening,  we  suspected  that  some  species 
were  equally  detectable  during  both  periods. 
If  true,  monitoring  efforts  that  focused  only 
on  these  species  could  greatly  expand  the  time 
during  which  surveys  could  be  conducted.  To 
evaluate  the  differences  in  detection  of  avian 
species  between  morning  and  evening,  we 
conducted  paired  morning  and  evening  point 
counts  during  the  breeding  season  and  during 
winter.  In  this  paper,  we  report  the  resulting 
estimates  of  avian  species  richness,  overall 
abundance,  and  abundance  of  selected  species. 
We  also  assess  the  relationship  between  de- 
tection of  individual  species  and  the  observed 
variation  between  morning  and  evening 
counts. 

STUDY  SITE  AND  METHODS 

Point  counts  were  conducted  on  the  1050  ha  Delta 
Experimental  Forest,  near  Stoneville,  Mississippi 
(33°  29'  N,  90°  55'  W).  Surrounded  largely  by  agri- 
culture, this  bottomland  hardwood  forest  was  heavily 
logged  from  1910  to  1920  with  additional  research  and 
commercial  harvests  continuing  though  the  1960s. 
There  was  no  timber  harvest  on  Delta  Experimental 
Forests  between  the  early  1960s  and  the  time  of  this 
study. 

We  conducted  morning  and  evening  point  counts 
during  the  breeding  season  (30  May— 12  June  1991;  8— 
21  May  1992)  and  winter  (4-14  February  1991;  9-29 
January  1992)  at  25  stations  in  each  of  4 forest  stands. 
Forest  stands  were  similar  in  habitat  but  were  subjeet- 
ed  to  different  silvicultural  management.  We  generally 
followed  standardized  protocols  for  conducting  point 
counts  (Ralph  et  al.  1993,  Hamel  et  al.  1996)  but  used 
4-min  sampling  periods  instead  of  5-min  and  20-m 
hxed  radius  circular  plots  instead  of  50-m.  We  reduced 
the  sampling  period  based  on  species  detection  curves 
from  preliminary  survey  data  and  we  restricted  the  plot 
radius  to  2()-m  because  of  a concurrent  effort  to  model 
habitat  using  these  same  data.  Points  were  visited  so 
that  each  visit  occurred  at  a different  time  during  the 
3-h  periods  following  sunri.se  (morning)  and  preceding 
sun.set  (evening).  Over  the  two  years  of  this  study,  we 
made  a total  of  10  morning  and  10  evening  visits  to 
each  of  the  4 stands  during  the  breeding  .season.  Within 
the  same  time  interval,  we  made  a total  of  6 morning 
and  6 evening  visits  to  each  .stand  during  winter.  De- 
tection probability  (Gutzwiller  1993a)  was  e.stimated 
for  each  species  as  the  proportion  of  total  point  counts 
during  which  the  species  was  detected.  During  the 
breeding  season  and  during  winter  we  eompared  the 
number  of  species  and  individuals  detected  during 
morning  and  evening  visits  using  a split  plot,  repeated 
measures  analysis  of  variance;  each  stand  (an  experi- 
mental design  block)  was  split  into  morning  and  even- 
ing treatment  periods  with  visits  (dates)  constituting 


the  repeated  measure.  All  statistical  analysis  were  per- 
formed using  the  SAS  System  for  Windows  (Release 
6.11,  SAS  Institute,  Inc.,  Cary,  NC,  USA).  We  sub- 
sequently compared  the  abundance  of  selected  individ- 
ual species  between  morning  and  evening  counts  using 
the  same  experimental  design.  However,  individual 
species  abundances  were  compared  only  if  the  overall 
variability  of  the  species  allowed  detection  of  at  least 
0.25  individuals  when  the  power  of  the  test  (1  — (3) 
was  at  least  0.80  with  a = 0.10  (Hamel  et  al.  1996). 
Furthermore,  because  we  conducted  multiple  tests 
when  comparing  individual  species,  we  used  Bonfer- 
roni’s  correction  which  reduced  the  probability  re- 
quired for  signihcance  of  these  tests  to  a < 0.006. 

RESULTS 

We  recorded  57  forest  landbird  species  dur- 
ing the  breeding  season  and  36  species  during 
the  winter.  More  species  (F,  3 = 383.35,  P < 
0.01)  and  total  individuals  (F,  3 = 597.38,  P 

< 0.01)  were  detected  in  morning  counts  (x 
± SE;  10.05  ± 0.06  species,  11.66  ± 0.08 
individuals)  than  in  evening  counts  (.v  ± SE; 
7.77  ± 0.07  species,  8.46  ± 0.09  individuals) 
during  the  breeding  season.  During  winter,  we 
again  detected  more  species  (F,  3 = 82.38,  P 

< 0.01)  and  total  individuals  (F,  3 = 26.59,  P 
= 0.01)  in  morning  counts  (T  ± SE;  6.12  ± 
0.07  species;  9.36  ± 0.14  individuals)  than  in 
evening  counts  (.v  ± SE;  4.44  ± 0.08  species, 
6.45  ±0.13  individuals). 

During  the  breeding  season,  16  of  57  spe- 
cies met  our  criteria  for  comparing  morning 
and  evening  counts  of  individual  species  (Ta- 
ble 1).  We  detected  significantly  (F,  3 > 46.83, 
P < 0.006)  more  individuals  during  morning 
counts  than  during  evening  counts  for  four 
species:  Blue  Jay  {Cyanocitta  cristata).  Tufted 
Titmouse  (Baeolophus  bicolor).  Mourning 
Dove  (Zenaida  macroiira),  and  Red-bellied 
Woodpecker  (Melcinerpes  carolinus).  No  sig- 
nificant differences  (P  > 0.01)  were  detected 
between  morning  and  evening  counts  for  the 
other  12  species  (Table  1).  Of  the  11  species 
eligible  for  comparison  during  winter,  only 
Tufted  Titmouse  was  detected  significantly 
(F,.3  = 50.6,  P < 0.006)  more  during  morning 
than  during  evening.  As  with  the  breeding 
season,  the  detection  of  the  remaining  species 
did  not  differ  significantly  (P  > 0.01)  between 
morning  and  evening  counts  (Table  1). 

Detection  probability  of  individual  species 
ranged  from  less  than  0.01  to  0.70  during  win- 
ter and  from  less  than  0.01  to  0.82  during  the 
breeding  sea.son.  There  was  a significant  cor- 


TABLE  1.  Abundance  {.x  ± SE)  and  detection  probability  of  individual  species  from  morning  and  evening  point  counts  conducted  during  the  breeding 
and  during  winter  of  1991  and  1992  on  Delta  Experimental  Forest,  Stoneville,  Mississippi,  USA. 


SHORT  COMMUNICATIONS 


c 

o 

<U 

(41 


2,g 


=»  ^1 

QO 

C 

c 

O 


c >> 
.2  ~ 
u ^ 
2 2 
w o 


<u 

o 

00 

00 

lO 

ON 

cn 

X 

o^ 

X 

ON 

(N 

o 

o 

X 

o 

c 

•-H 

(N 

(N 

(N 

(N 

CN 

(N 

CN 

CN 

a 

•o 

o 

o 

O 

o 

o 

O 

O 

O 

O 

o 

O 

o 

o 

O 

O 

O 

c 

3 

X) 

3 

± SE 

d 

+1 

d 

+ 1 1 

d 

+1 

d 

X 

+1 

d 

+1 

d 

+1 

d 

+1 

d 

+1 

d 

+1 

d 

X 

+1 

d 
! +1 

o o 

+ 1 O +1 

d 

+1 

d 

+1 

d 

+1  o 

'c 

(N 

O 

»r) 

(N 

00 

(N 

r- 

(N 

o 

C4 

On 

'c 

1— H 

m 

X 

(N 

On 

0-) 

r- 

(N 

00 

•n 

CO 

ON 

On 

(U 

> 

00 

00 

00 

CO 

o 

UJ 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

C 'H 


O 

d o 


o 

O 

VO 

00 

Tj- 

CO 

vO 

u 

CO 

X 

o 

n 

ON 

o 

Ov 

00 

C^l 

CO 

On 

in 

ri 

X 

m 

CO 

lO 

m 

lo 

’t 

o 

o 

d 

d 

d 

d 

o 

d 

d 

d 

d 

o 

d 

o 

o 

d 

o 

o 

d 

o 

d 

o 

o 

CO 

VO 

CN 

m 

CO 

o 

in 

X 

CO 

CO 

(N 

OJ 

m 

CN 

n) 

lo 

m 

o 

o 

O 

o 

o 

O 

o 

o 

o 

o 

o 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

+1 

+1 

+1 

+1 

o 

+ 1 

+1 

+1 

+1 

o 

+1 

o 

o 

+ 1 

o 

o 

X 

o 

+1 

(N 

o 

o 

o 

o 

o 

r^ 

(N 

CO 

o 

CO 

in 

CN 

o 

o 

00 

(N 

CO 

m 

ON 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

o 

(^1 

CN 

ov 

00 

00 

00 

Ov 

CO 

(N 

O') 

CO 

m 

CO 

(N 

00 

ni 

o 

o 

O 

o 

o 

o 

o 

O 

o 

o 

o 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

+1 

+1 

+1 

+1 

o 

+1 

+1 

+1 

+1 

o 

+1 

o 

o 

+ 1 

o 

o 

X 

o 

+1 

ON 

00 

r- 

(N 

CO 

00 

X 

in 

X 

CN 

o 

o 

ON 

■3- 

NO 

in 

r- 

a^ 

n) 

CO 

(N 

VO 

00 

q 

00 

VO 

(N 

On 

d 

d 

d 

d 

-i 

d 

d 

d 

d 

•n  ON  00  CN  ^ 

r-  — ' r-  NO  o CN 

o r-  o 

d d d d d d 


OONOOOvO'^CNmOO'O  »riOONOO 
OinCNsOOfNONOON 

dddddddddoddddo 


« 

o 

CO 

00 

in 

« 

CN 

ON 

ra 

00 

a\ 

in 

NO 

X 

in 

CN 

CN 

CN 

1—1 

1-H 

CN 

CN 

(N 

(N 

,—1 

CN 

CN 

O 

O 

o 

o 

o 

O 

O 

O 

o 

o 

o 

o 

o 

o 

o 

o 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

+1 

+1  ^1 

+1 

+1  ■“! 

+1 

+1 

+1 

+1 

+1 

+1  ■°| 

+1 

+ ! 

o +1 

+1 

+1 

+1  o 

00 

in 

ON 

00 

r- 

in 

r- 

CO 

o 

m 

CN 

X 

f-H 

in 

NO 

00 

C^l 

r- 

X 

"=t 

r- 

o 

o 

00 

r-* 

CO 

CO 

X 

in 

CN 

CN 

00 

CN 

CN 

m 

CN 

CN 

CN 

CO 

CO 

in 

d 

d 

— * 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

^1? 


5 S 

'>5  ^ 


5 

5 

>3 

c 

t3 

>, 

5 

§ 


<v 


3 o 
S ° 

3 

^ u 
Q 1 


5 

o 


in 

M 

o rj  “ 
O D u 
Oh  Q,  <U 

■o  -a 
° o ^ 

III 

_ o 
T3  -a  o 

^ ^ ^ 
C3  ^ 

(U  (U 

-C  X) 


II 

p (U 

-S  X 
U a 

w o 
u,  ^ 

^ E 

“O 
r-  (U 


^ -S 
-S  Q 
o 2 .S 

■SI  ^ 


c 

TJ  "6  ^ 
dj  <U  O 

a:  ai  Q 


o ^ 

z o 


c:  1) 

II 

U -c 
- u 

>N 

« c 


§■  -2 

^I'-S 
u b :c’ 

c/5  ^ w 

2^5 

H g fS 


(U  o 
o S 

— 4 Cd  P 

CQ  U H U > 


o S3  ^ 


k. 

2 ^ ^ 

Q 5 S 
^ 

Oc  ^ ^ 
K § 

5 o "5 
"3^0 
3 S S! 

- o S 

.3  Ih  O 
XI  ^ JJ 

o ^ .3 
(i  X i> 

o V ^ 

'C  (U  (U 

E e o 
a! 


3 ^ 

S ^ 

S 5i 

C -3 
o o 


u w 
? ^ 
I ■£ 

■u  ^ 
w 

a.  >, 

I « 

S o 
? o 

O 4= 


5 2 


a s; 
ac  -5 

3 S 

u 

(D  C 


3 

ID  -j- 

c h 

'■'  J 

a N 


' -o 
u. 
c3 

u 


Q 3 

Q,  a 
oo 

00  “O 

c S 


c/^  Z 


r-  '•'* 

E O 

^ u 
QQ  X 

O 

00  ^ 
••g  X 
E ^ 


V 

o. 

nO 


o 

e 


a ^ 


141 


Morning  and  evening  counts  not  compared. 


142 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


relation  between  detection  probability  and 
variation  in  abundance  during  the  breeding 
season  (r^  = 0.74,  P < 0.01)  but  not  during 
winter  (r^  = 0.32,  P > 0.05).  Six  of  1 1 species 
found  at  our  study  site  throughout  the  year. 
Red-bellied  Woodpecker,  Downy  Woodpecker 
{Picoides  pubescens).  Blue  Jay,  Tufted  Tit- 
mouse, Carolina  Wren  {Thryothorus  ludovi- 
cicmus),  and  Northern  Cardinal  {Cardinalis 
cardinalis),  had  greater  detection  probabilities 
iP  < 0.05)  during  summer  (Table  1).  Con- 
versely, Red-headed  Woodpecker  {Melaner- 
pes  erythrocephalus).  Northern  Flicker  (Co- 
laptes  auratiis)  and  American  Robin  (Turdus 
migratorius)  had  greater  detection  probabili- 
ties during  winter. 

DISCUSSION 

During  the  breeding  season,  morning  point 
counts  yielded  more  species  and  more  indi- 
viduals than  did  evening  counts.  Furthermore, 
when  significant  differences  existed  for  indi- 
vidual species,  morning  counts  were  consis- 
tently higher  than  evening  counts.  Forest  land- 
birds  have  long  been  presumed  to  be  more 
detectable  during  early  morning  than  at  other 
times  of  the  day  and  many  observers  restrict 
breeding  bird  censuses  to  morning  hours 
(Skirvin  1981).  In  studies  of  diel  variation 
(Shields  1977,  Skirvin  1981),  more  species 
and  individuals  were  detected  during  the  ini- 
tial 2 h after  sunrise  than  at  other  times.  In 
floodplain  forests,  Robbins  (1981)  found  that, 
although  the  total  number  of  birds  recorded 
diminished  beyond  2 h after  sunrise,  the  num- 
ber of  species  detected  remained  nearly  uni- 
form for  up  to  5 h after  sunrise.  Although 
there  have  been  few  comparisons  of  early 
morning  and  late  evening  censuses,  Grue  and 
coworkers  (1981),  working  in  desert  habitats 
during  the  breeding  season,  found  more  spe- 
cies and  individuals  during  morning  point 
counts  than  during  evening  counts.  Our  data 
provide  further  empirical  evidence  to  support 
presumed  temporal  differences  in  avian  detec- 
tions between  morning  and  evening  counts 
during  the  breeding  season. 

During  winter,  we  also  detected  more  spe- 
cies and  individuals  on  morning  point  counts 
than  on  evening  counts.  Rollfinke  and  Yahner 
(1990)  also  reported  more  species  and  more 
individuals  on  early  morning  transects  than  on 
evening  tran.sects  during  winter.  Although  we 


detected  only  Tufted  Titmouse  significantly 
more  on  morning  counts  than  on  evening 
counts  during  winter,  Gutzwiller  (1993a) 
found  that  five  species  had  higher  detection 
probabilities  on  point  counts  between  07:00 
and  13:45  than  at  other  times  of  the  day.  Over- 
all, our  data  suggest  that  evening  point  counts 
during  either  the  breeding  season  or  winter 
will  probably  underestimate  species  richness, 
overall  avian  abundance,  and  the  abundance 
of  at  least  some  species  in  bottomland  hard- 
wood forests. 

ACKNOWLEDGMENTS 

These  data  were  collected  while  W.  P.  Smith  was  a 
biologist  with  the  Southern  Hardwoods  Laboratory, 
USDA  Forest  Service,  Southern  Research  Station, 
Stoneville,  Mississippi.  J.  H.  McGuiness  and  T D. 
McCarthey  provided  valuable  assistance  with  point 
counts  on  Delta  Experimental  Forest. 

LITERATURE  CITED 

Grue,  C.  E.,  R.  P.  Balda,  and  C.  D.  Johnson.  1981. 
Diurnal  activity  patterns  and  population  estimates 
of  breeding  birds  within  a disturbed  and  undis- 
turbed desert  scrub  community.  Stud.  Avian  Biol. 
6:287-291. 

Gutzwiller,  K.  J.  1991.  Estimating  winter  species 
richness  with  unlimited-distance  point  counts. 
Auk  108:853-862. 

Gutzwiller,  K.  J.  1993a.  Refining  the  use  of  point 
counts  for  winter  studies  of  individual  species. 
Wilson  Bull.  105:612—627. 

Gutzwiller,  K.  J.  1993b.  Avian  responses  to  observer 
clothing:  caveats  from  winter  point  counts.  Wilson 
Bull.  105:628-636. 

Hamel,  P.  B.,  W.  P.  Smith,  D.  J.  Twedt,  J.  R.  Woehr, 
E.  Morris,  R.  H.  Hamilton,  and  R.  J.  Cooper. 
1996.  A land  manager’s  guide  to  point  counts  of 
birds  in  the  Southeast.  USDA  For.  Ser.  Gen.  Tech. 
Rep.  SO- 120: 1-39. 

Ralph,  C.  J.,  G.  R.  Guepel,  P.  Pyle,  T.  E.  Martin, 
AND  D.  E Desante.  1993.  Handbook  of  field 
methods  for  monitoring  landbirds.  USDA  For.  Ser. 
Gen.  Tech.  Rep.  PSW-GTR-144,  Pacific  South- 
west Res.  Stn.,  Albany,  California. 

Ralph,  C.  J.,  J.  R.  Sauer,  and  S.  Droege.  (Eds.). 
1995a.  Monitoring  bird  populations  by  point 
counts.  USDA  For.  Ser.  Gen.  Tech.  Rep.  PSW- 
GTR-149,  Pacific  Southwest  Res.  .Stn.,  Albany, 
California. 

Ralph,  C.  J.,  S.  Droege,  and  J.  R.  Sauer.  1995b. 
Managing  and  monitoring  birds  using  point 
counts:  .standards  and  applications.  Pp.  161-168 
in  Monitoring  bird  populations  by  point  counts  (C. 
J.  Ralph,  J.  R.  Sauer,  and  S.  Droege,  Ed.).  USDA 
For.  Ser.  Gen.  Tech.  Rep.  PSW-GTR-149,  Pacific 
Southwest  Res.  Stn.,  Albany,  California. 


SHORT  COMMUNICATIONS 


143 


Robbins,  C.  S.  1981.  EITcct  of  nine  of  day  on  bird 
activity.  Stud.  Avian  Biol.  6:273-286. 

Rollfinke,  F.  B.  and  R.  H.  Yahner.  1990.  Effects  of 
time  ot  day  and  season  on  winter  bird  counts. 
Condor  92:213-219. 

Shields,  W.  M.  1977.  The  effect  of  time  of  day  on 
avian  census  results.  Auk  94:380-383. 

Skirvin,  a.  a.  1981.  Effect  of  time  of  day  and  time 
of  season  on  the  number  of  observations  and  den- 


sity estimates  of  breeding  birds.  Stud.  Avian  Biol. 
6:271-274. 

Smith,  W.  R,  D.  J.  Twedi,  D.  A.  Wiedenfeld,  P.  B. 
Hamel,  R.  P.  Ford,  and  R.  J.  Cooper.  1993.  Point 
counts  ot  birds  in  bottomland  hardwood  forests  of 
the  Mississippi  Alluvial  Valley:  duration,  mini- 
mum sample  size,  and  points  versus  visits.  USDA 
For.  Sen  Res.  Paper  SO-274,  Southern  For.  Exp. 
Stn.,  New  Orleans,  Louisiana. 


Wilson  Bull.,  111(1),  1999,  pp.  144-156 


Ornithological  Literature 

Edited  by  William  E.  Davis,  Jr. 


THE  BIRD  COLLECTORS.  By  Barbara 
and  Richard  Meams.  Academic  Press,  San  Di- 
ego and  London.  1998:  xviii  + 472  pages, 
many  unnumbered  figures,  4 maps  of  expe- 
dition itineraries.  $49.95  (cloth). — This  is  the 
third  book  by  the  Mearnses  in  10  years.  The 
first  two  consisted  of  biographies  of  persons 
after  whom  birds  have  been  named;  first,  for 
the  Western  Palearctic  (“Biographies  for  bird 
watchers”,  1988,  reviewed  in  Wilson  Bulletin 
101:658-659),  and  then  for  North  America 
(“Audubon  to  Xantus”,  1992,  reviewed  in 
Wilson  Bulletin  105:701-702).  Richard  is  a 
countryside  ranger  in  Scotland;  Barbara  is  a 
professional  occupational  therapist,  but  in  re- 
cent years  has  spent  more  and  more  time  on 
her  avocation,  biohistorical  research. 

Their  new  book,  although  biographical  in 
large  part  (and  thus  overlapping  slightly  with 
the  first  two  books),  has  quite  a different  bal- 
ance. The  Mearnses  are  well  aware  that  the 
collecting  of  bird  specimens  in  the  late  20th 
Century  is  subject  to  much  debate,  often  ac- 
rimonious. The  arguments  and  the  social  and 
political  pressures  of  the  current  anti-collect- 
ing camp  are  phenomena  seldom  if  ever  faced 
by  the  historical  collectors  whose  exploits 
form  most  of  the  book’s  subject  matter.  The 
Mearnses  make  their  position  plain  in  a brief 
preface;  neither  of  them  has  ever  deliberately 
killed  a bird  (they  salvage  accidentally  killed 
birds  for  the  Royal  Museum  of  Scotland). 
They  point  out,  however,  that  anyone  who,  in 
their  words,  turns  “pale  at  the  mere  thought 
of  killing  birds”  must  realize  that  “anyone 
who  drives  a car,  uses  products  of  the  petro- 
chemical industries,  owns  a cat,  has  glass  in 
the  windows  of  their  home,  buys  paper,  or 
consumes  electricity  will  be  responsible  for 
killing  birds.”  This  subject  is  dealt  with  at 
greater  length  in  their  Chapter  17,  “The  im- 
portance of  old  and  new  bird  collections,” 
which  relies  heavily  on  the  important  paper 
by  Rem.sen  (1995). 

The  first  three  chapters  bounce  around  a 
good  bit,  dealing  with  reasons  for  killing  birds 
other  than  for  museum  collections,  bird  books 


and  journals,  human  casualties,  labeling  and 
note-taking,  problems  in  the  field,  in  shipment 
of  specimens,  and  in  the  museum,  etc.  By  the 
fourth  chapter,  the  emphasis  becomes  primar- 
ily historical,  reviewing  collecting  and  collec- 
tors chronologically.  But  the  arrangement  is 
not  strictly  chronological  as  a whole;  chapters 
on  specific  kinds  of  collectors  are  internally 
chronological.  Examples  include  “Bird  artists 
as  collectors,”  “Government-sponsored  col- 
lecting,” “Army  officers,”  “The  medical  pro- 
fession,” and  “Clergymen  and  missionaries.” 
A chapter  entitled  “The  great  accumulators” 
treats  the  owners  of  large  private  collections, 
beginning  with  the  13th  Earl  of  Derby  (1775— 
1851)  and  ending  with  the  notorious  Colonel 
Richard  Meinertzhagen  (see  Knox  1993).  A 
chapter  on  “The  professional  field  collectors” 
is  a hodgepodge,  including  scientists  such  as 
Alfred  Russell  Wallace,  whose  collections 
were  indeed  sold,  and  contract  collectors 
working  for  museums,  such  as  Rollo  Beck. 
This  chapter  is  much  too  short,  as  major  por- 
tions of  the  holdings  of  several  of  the  large 
museums  were  made  by  collectors  under  con- 
tract; in  South  America,  for  example,  some  of 
the  most  prolific  were  Samuel  Klages,  the 
Steinbachs  (father  and  son),  M.  A.  Carriker, 
Jr.,  and  more  recently  William  H.  Partridge. 
Of  these,  only  Carriker  is  mentioned,  thrice; 
half  of  a sentence  in  the  “Professional  field 
collectors”  chapter  and  half  of  a paragraph  on 
his  two  wives  in  a chapter  on  “Women  in  the 
field”! 

The  Mearnses  admit  in  their  Preface  that 
their  “approach  has  been  rather  Anglo-cen- 
tric,” after  which  they  list  by  name  17  “great 
collectors  [who]  have  been  omitted  or  not 
mentioned  in  detail” — the  third  reference  to 
Carriker  is  the  presence  of  his  name  in  this 
list.  The  relative  neglect  of  some  parts  of  the 
world  is  obvious  all  through  the  book,  and  I’m 
not  sure  this  can  be  wholly  excused  by  their 
admission  quoted  above.  They  appear  to  have 
been  obsessed  by  the  history  of  collecting  in 
central  Asia,  as  this  is  the  subject  of  a 27-page 
chapter  called  “Terra  Incognita”,  in  which  IV2 


144 


ORNITHOLOGICAL  LITERATURE 


145 


pages,  a map  and  three  portraits  are  devoted 
to  the  exploits  of  General  Nicholas  M.  Prjev- 
alsky  (of  Prjevalsky’s  Horse),  and  slightly 
more  text  plus  a map  and  a portrait  to  those 
of  Armand  David  (of  Pere  David’s  Deer). 
Surely  most  of  the  South  American  continent 
was  at  least  as  “incognita”. 

In  the  “Professional  field  collectors”  chap- 
ter, Lord  Rothschild  is  quoted  as  having  stated 
that  William  Doherty  (1857-1901)  was  “un- 
questionably the  best  collector  for  the  last  fifty 
years.”  To  anybody  who  has  seen  bird  skins 
from  Doherty's  expeditions,  this  statement  is 
incomprehensible;  he  never  learned  to  skin 
birds  for  himself,  turning  them  over  to  Indian 
servants,  according  to  the  Mearnses.  Carnegie 
Museum  of  Natural  History  has  a collection 
Doherty  made  in  Kenya  six  months  before  his 
death  in  1901  (Holland  1905).  The  skins  are 
mediocre  and  the  original  labels  bear  nothing 
but  a pencilled  sex  mark;  the  data  (“10  miles 
W of  Mombasa,  September-October  1900”) 
were  apparently  supplied  by  Doherty  to  Dr. 
Holland,  Director  of  the  museum,  who  had 
purchased  the  collection.  All  of  this  suggests 
that  the  Mearnses  were  probably  correct  in 
suggesting  that  Rothschild’s  high  praise  of 
Doherty  may  have  been  based  on  his  “servic- 
es to  entomology.” 

A chapter  on  “Women  in  the  field”  con- 
tains 9 short  biographies.  Of  the  women  thus 
honored,  only  3 [Emilie  Snethlage,  Elizabeth 
Kozlova,  and  Beryl  P.  (Pat)  Hall]  contributed 
to  ornithology  primarily  through  their  collect- 
ing activities,  at  a level  comparable  to  that  of 
most  of  the  males  featured  in  the  rest  of  the 
book.  Here  is  where  we  find  the  Mearnses’ 
most  appalling  omission;  the  late  Maria  Koep- 
cke  (1924-1971),  whose  name  appears  as  the 
only  woman  in  the  Preface  list  of  collectors 
who  “deserve  more  space  than  we  could  give 
them”.  Maria  Koepcke  may  justifiably  be  said 
to  be  one  of  the  true  pioneers  in  the  20th  Cen- 
tury study  of  the  ornithology  of  Peru,  a small 
country  whose  avifauna  numbers  about  IVS 
times  that  of  the  entire  Palearctic.  She  and  her 
husband  Hans-Wilhelm,  trained  as  a hydrobi- 
ologist, founded  Casa  Humboldt  (^^Humboldt 
House)  in  Lima  in  1957;  this  became  the  con- 
venient base  for  many  expeditions  throughout 
Peru  involving  scientists  from  several  nations. 
Maria  conducted  avifaunal  surveys  in  areas 
ranging  from  the  desert  coast  to  the  Amazo- 


nian rain  forest.  Her  bibliography  lists  29  ti- 
tles, plus  12  co-authored  with  her  husband. 
She  described  3 new  species  and  1 3 new  sub- 
species. 

An  Appendix  lists,  in  sequence  of  size,  the 
world’s  69  largest  collections  of  bird  speci- 
mens, together  with  the  most  significant  com- 
ponents included  therein.  Unfortunately  there 
are  many  errors  and  omissions  in  this  list, 
partly  because  for  some  museums  the  authors’ 
information  was  a quarter-century  out  of  date, 
based  on  Banks  and  coworkers  (1973).  It  is  a 
pity  that  they  felt  no  need  to  enter  into  cor- 
respondence to  get  more  recent  figures.  Attri- 
bution of  components  is  irregular;  parts  of  the 
dismantled  collections  of  the  Cleveland  Mu- 
seum of  Natural  History,  for  example,  are  list- 
ed for  the  Field  Museum;  the  Museum  of  Zo- 
ology, University  of  Michigan;  and  the  Pea- 
body Museum,  Yale  University,  but  not  the 
Carnegie  Museum  of  Natural  History,  which 
may  have  the  largest  number  of  former  Cleve- 
land birds.  No  components  are  listed  for  the 
Delaware  Museum  of  Natural  History,  which 
holds  one  of  the  most  important  U.S.  collec- 
tions of  Philippine  birds,  and  much  of  the  for- 
mer collections  of  John  E.  duPont,  George  M. 
Sutton,  and  Allan  R.  Phillips.  Numerous  other 
omissions  could  be  mentioned,  such  that  this 
Appendix  is  not  as  valuable  as  it  could  have 
been. 

Any  heterogeneous  work  of  this  sort  is 
bound  to  induce  comments  and  corrections.  A 
few  are  listed  below: 

p.  13.  The  authors  mention  the  extinction  of 
the  endemic  flowerpecker  {Dicaeum  quadri- 
color)  of  Cebu  owing  to  the  deforestation  of 
that  Philippine  island;  in  fact,  no  fewer  than 
9 endemic  forms  of  Cebu  were  thought  to 
have  been  extirpated  (Rabor  1959),  although 
small  remnant  populations  of  a few  of  these 
have  subsequently  been  found, 
p.  111.  The  authors  refer  to  “remote,  little- 
known  islands  such  as  Whitsunday  Atoll, 
Clarion  Island  and  the  Revillagigedo  group.” 
Clarion  is,  in  fact,  one  of  the  islands  in  the 
Revillagigedo  group. 

p.  149.  For  a more  complete  account  of  the 
fate  of  the  Gould  Australian  collection,  see 
Meyer  de  Schauensee  1957. 
p.  182.  The  Ruwenzori  mountains  are  on  the 
western,  not  the  eastern  border  of  Uganda, 
p.  244.  British  writers  should  know  that  the 


146 


THE  WILSON  BULLETIN  • Vnl.  III.  No.  I.  March  1999 


authorship  of  the  classical  Handbook  of  Brit- 
ish Birds  was  in  the  sequence  Witherby,  Ti- 
cehurst,  Jourdain  and  Tucker;  the  Mearnses 
list  Jourdain  last. 

p.  270.  A comment  would  have  been  appro- 
priate after  the  statement  about  Prjevalsky’s 
first  expedition;  “there  would  have  been  more 
[bird  specimens]  but  most  were  moulting  so 
nine-tenths  of  all  of  the  birds  shot  were  dis- 
carded.” This  atrocious  practice  has,  of  course, 
proved  to  be  the  bane  of  students  of  molts  and 
plumages;  I encountered  the  same  kind  of 
statement  in  connection  with  some  19th  Cen- 
tury British  collectors  of  Philippine  birds, 
p.  366.  Brina  Kessel  and  1 have  been  friends 
and  colleagues  for  fifty  years,  but  I can’t  un- 
derstand the  rationale  for  including  her  in  a 
book  about  collectors.  The  Mearnses  mention 
that  “her  first  specimen-based  research  was 
for  her  Ph.D.  dissertation  on  European  Star- 
lings and  involved  the  preparation  of  over  500 
skins.”  True  enough,  except  that  the  majority 
of  the  starlings  were  collected  and  prepared 
by  Robert  W.  Dickerman. 

1 need  hardly  say  that  any  reader  of  this 
book  will  find,  as  1 did,  that  the  descriptions 
of  the  exploits  of  many  collectors  previously 
known  to  us  as  little  more  than  names  at- 
tached to  bird  species  have  been  brought  to 
vivid  life  by  the  Mearnses.  Their  book  is  high- 
ly readable,  and  their  attitude  toward  collect- 
ing as  fair-minded  as  one  might  ask  of  writers 
not  directly  involved  themselves  in  collecting. 

I am  indebted  to  Manuel  Plenge  for  sending 
me  biographical  materials  on  Maria  Koep- 
cke.— KENNETH  C.  PARKES. 

LITERATURE  CITED 

Banks,  R.  C.,  M.  H.  Clench,  and  J.  C.  Barlow.  1973. 
Bird  collections  in  the  United  States  and  Canada. 
Auk  90:136-170. 

Holland,  W.  J.  1905.  A list  of  the  birds  collected  near 
Mombasa.  East  Africa,  by  William  Doherty.  Ann. 
Carnegie  Mus.  3:453—463. 

Kno.x,  a.  G.  1993.  Richard  Mcinert/.hagen — a ca.se  of 
fraud  examined.  Ibis  135:320-325. 

Mi:yer  de  Schauensee,  R.  1957.  On  some  avian  types, 
principally  Gould's,  in  the  collection  of  the  Acade- 
my. Proc.  Acad.  Nat.  Sci.  Philadelphia  109:123-246. 
Rabor,  D.  S.  1959.  The  impact  of  deforestation  on 
birds  of  Cebu.  Philippines,  with  new  records  for 
that  island.  Auk  76:37-43. 

Rem.sen  .1.  V.,  Jr.  1995.  The  importance  of  continued 
collecting  of  bird  specimens  to  ornithology  and 
bird  conservation.  Bird  Conserv.  Int.  5:145-180. 


LIFE  OF  THE  FLYCATCHER.  By  Alex- 
ander E Skutch.  Illustrated  by  Dana  Gardner. 
University  of  Oklahoma  Press,  Norman, 
Oklahoma.  1997:  xiii  + 162  pp.,  16  color 
plates,  32  black-and-white  drawings,  4 tables, 
bibliography,  index.  $40.00.  ISBN  0-8061- 
2919-0. — Alexander  Skutch  brings  forth  a 
popular  review  of  the  Tyrannidae  in  Life  of 
the  Flycatcher.  This  volume  continues 
Skutch’s  Life  of.  . . series,  each  reviewing  var- 
iation in  life  history  within  a bird  family  (e.g., 
Tanager,  Pigeon,  Woodpecker,  Hummingbird). 
The  Tyrannidae,  subject  of  this  current  vol- 
ume, is  a large  family  (380  species)  of  subos- 
cines  limited  to  the  New  World.  The  text  is 
successful  in  showing  the  great  degree  of  di- 
versity displayed  by  this  family.  If  your 
knowledge  of  New  World  flycatchers  is  lim- 
ited to  North  American  species,  the  diversity 
described  for  the  family  in  Life  of  the  Fly- 
catcher will  amaze  you.  The  text  may  offer 
inspiration  for  evolutionary  inquires  and  di- 
rections for  further  study. 

The  text  begins  with  an  overview  of  the 
family,  and  continues  with  chapters  on  food, 
daily  life,  song,  courtship,  nest,  eggs,  young 
and  breeding  success.  Skutch’s  personal  ex- 
perience with  this  group  during  his  long  career 
in  the  American  tropics  is  evident.  The  book 
ends  with  138  references  organized  by  chap- 
ters and  an  index.  The  bibliography  provides 
some  direction  to  seeking  further  information. 

Skutch  takes  some  exception  to  the  illogic 
often  apparent  in  birds’  names  and  uses  “fly- 
catcher” in  common  names  rather  than  vari- 
ous forms  of  “tyrant”  (as  in  “marsh-tyrant” 
or  “tyrannulet”);  this  practice  may  delay 
tracking  down  specific  species  in  other  texts. 
The  only  error  of  some  note  I found  was  in 
the  index:  Alder  Flycatcher  was  identified  as 
""  Empidonax  minimum." — PETER  E. 
LOWTHER. 


A NEOTROPICAL  COMPANION:  AN 
INTRODUCTION  TO  THE  ANIMALS, 
PLANTS,  AND  ECOSYSTEMS  OF  THE 
NEW  WORLD  TROPICS.  Second  edition,  re- 
vised and  expanded.  By  John  Kricher.  Prince- 
ton University  Press,  Princeton,  New  Jersey. 
1997:  451  pp.,  177  color  photographs,  86  line 
drawings.  $29.95  (cloth). — This  sturdy  vol- 


ORNITHOLOGICAL  LITERATURE 


147 


ume  accompanied  me  for  10,000  miles  this 
spring,  enlightening  me  during  airport  vigils, 
on  long  flights,  and  during  hot  nights  in  Gua- 
temala and  elsewhere.  Except  for  a warped 
cover  it  is  still  in  excellent  condition.  Despite 
many  prior  trips  to  the  American  tropics,  I 
found  this  book  fascinating  reading  as  well  as 
a valuable  reference  to  other  literature  sourc- 
es. 

This  is  a series  of  authoritative  essays  on 
tropical  ecology,  organized  in  14  chapters  and 
followed  by  an  Appendix,  a list  of  Acronyms, 
a Glossary,  35  pages  of  References,  and  an 
Index  that  includes  every  species,  genus,  and 
topic  mentioned.  The  Appendix,  “And,  Hey, 
Let’s  be  Careful  Out  There,”  should  catch  the 
reader’s  eye  before  (s)he  leaves  home.  If  you 
are  a tropical  bookworm  you  will  appreciate 
the  miniabstracts  of  his  40  book-length  ref- 
erences. 

Chapters  are  arranged  in  a logical  sequence, 
beginning  with  textbook  descriptions  of  cli- 
mates and  ecosystems,  and  rainforest  struc- 
ture, diversity,  and  function.  Chapters  on  evo- 
lutionary patterns  and  coevolution/ecology  of 
fruit  track  relationships  among  animals  and 
plants  from  the  research  of  Darwin  to  inves- 
tigators of  the  mid  1990s.  After  brief  chapters 
on  the  Neotropics  as  a pharmacy  and  on  living 
off  the  land,  are  chapters  on  ecosystems  (riv- 
ers, mountains,  savannas,  dry  forests,  man- 
groves, coral  reefs)  and  on  Neotropical  birds 
and  mammals,  culminating  with  a lively  dis- 
cussion on  deforestation  and  biodiversity.  A 
strong  conservation  theme  teases  the  reader  to 
take  appropriate  action. 

Professor  Kricher  writes  in  an  easy  conver- 
sational style,  tempting  the  student,  the  re- 
searcher, or  the  vacationer  to  read  on  and  on. 
His  wide  field  experience  in  Central  and  South 
America,  his  extensive  knowledge  of  the  trop- 
ical literature,  his  long  academic  career,  and 
his  gift  for  writing  combine  to  make  this  pub- 
lication a gem  for  the  tropical  explorer.  State- 
ments in  the  text  are  supported  by  close  to  a 
thousand  references  to  the  scientific  literature. 
Terms  defined  in  the  glossary  are  italicized  in 
the  text.  Dozens  of  delightful  line  drawings  of 
birds,  mammals,  and  lesser  life  forms  from  the 
pen  of  Ted  Davis  grace  the  pages.  A new  fea- 
ture in  this  edition  is  a collection  of  177  color 
photos,  all  cross-referenced  from  the  text. 

I was  shocked  to  read  in  the  conversion  ta- 


ble (p.  xviii)  that  1 square  mile  = 2,590 
square  km  and  (p.  35)  that  1 cm  = 2.5  in. 
And  I regret  to  report  9 scientific  names  of 
birds  were  misspelled  one  or  more  times,  in- 
cluding “Beautiogallus”  (3  times).  These  pre- 
ventable accidents  aside,  the  book  is  a gold- 
mine of  information  that  will  greatly  enrich 
one’s  tropical  experience. 

The  Neotropical  Companion  should  be  in 
every  high  school  and  college  library,  in  the 
travel  section  of  your  public  library,  and  in 
the  carryon  luggage  of  all  students  and  birders 
bound  for  the  American  tropics. — CHAND- 
LER S.  ROBBINS. 


THE  BIRDS  OF  ST.  LUCIA,  WEST  IN- 
DIES. By  Allan  R.  Keith.  British  Ornitholo- 
gists’ Union,  c/o  The  Natural  History  Muse- 
um, Tring,  Herts  HP23  6AP,  United  Kingdom. 
1997;  176  pp.,  40  color  plates  with  captions, 
7 text  figures,  3 tables,  7 appendices,  14  £ 
(cloth). — In  recent  years,  the  British  Ornithol- 
ogists’ Union  has  done  the  ornithological 
community  a tremendous  service  by  sponsor- 
ing the  production  of  a variety  of  check-list 
style  books  specializing  on  unusual  or  exotic 
regions  of  the  world.  Compared  to  Gambia, 
Nigeria,  Cyprus,  The  Philippines,  or  even  the 
Southern  Bahamas,  St.  Lucia  is  a tiny  area 
that  supports  a fairly  small  avifauna.  Never- 
theless, this  is  a splendid  book  that  provides 
both  the  detailed  observational  information 
one  expects  from  a check-list  plus  a vast 
amount  of  other  information  that  shows  us 
why  St.  Lucia  is  an  important  place  in  the 
ornithological  world. 

The  nitty-gritty  of  any  check-list  lies  in  its 
species  accounts,  which  in  this  case  cover  162 
species  that  currently  occur  on  St.  Lucia  or 
have  reliable  records  from  the  past.  The  author 
does  these  accounts  in  a spartan  60  pages.  To 
this  can  be  added  short  appendices  covering 
species  of  uncertain  occurrence,  where  mu- 
seum specimens  from  St.  Lucia  reside,  recov- 
eries of  banded  birds  in  St.  Lucia,  a gazetteer, 
and  the  origins  of  resident  breeding  birds.  A 
major  appendix  ( 1 8 pages)  provides  more  in- 
formation on  the  3 endemic  species  and  3 en- 
demic sub-species  found  on  the  island.  The 
author  seems  to  have  left  no  leaf,  library,  or 


148 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


museum  unturned  in  gathering  all  the  details 
available  about  St.  Lucian  birds. 

Were  the  above  material  all  of  the  book,  I 
would  readily  recommend  it  for  anyone  plan- 
ning a visit  to  the  island,  but  1 would  question 
its  value  to  the  general  reader.  Allan  Keith’s 
introductory  material  (9  chapters  adding  up  to 
60  pages)  does  an  excellent  job  of  introducing 
the  island  on  which  the  bird  records  were 
made.  Chapters  include  political  history,  ge- 
ology and  geography,  climate,  vegetation,  his- 
tory of  ornithological  exploration,  migration, 
breeding,  zoogeography  and  conservation.  All 
of  this  information  makes  the  avifauna  of  St. 
Lucia  come  much  more  alive  as  part  of  one 
island  within  the  West  Indian  region.  Keith 
also  does  an  outstanding  job  of  pointing  out 
the  unusual  role  this  island  has  played  in  West 
Indian  conservation,  as  a result  of  the  com- 
bined efforts  of  a local  biologist  (the  late  Ga- 
briel Charles)  and  an  English  vagrant  (Paul 
Butler)  in  developing  a conservation  plan  that 
was  very  successful  on  St.  Lucia  and  has  been 
exported  elsewhere. 

This  combination  of  detailed  observational 
work  in  the  species  accounts  with  the  broad 
biogeographic  and  ecological  perspective 
found  in  the  introduction  allows  the  reader  to 
not  only  find  out  what  species  occur  on  St. 
Lucia,  but  to  get  a feeling  about  why  it  is  this 
number  and  how  the  island  compares  to  its 
neighbors.  The  author’s  synthesis  of  current 
scientific  thought  is  exemplary  (although  a 
few  sections  may  need  to  be  revised  when  the 
current  work  being  done  on  molecular  genet- 
ics of  West  Indian  birds  is  finally  released). 
This  is  a great  way  to  both  find  out  what  is 
on  St.  Lu^ia  and  to  get  a bit  of  a feeling  about 
why  this  is  so. — JOHN  FAABORG. 


MUNIAS  AND  MANNIKINS.  By  Robin 
Restall,  illus.  by  the  author.  Yale  University 
Press,  New  Haven,  Connecticut.  1997:  264 
pp.,  16  colored  plates  with  captions,  64  color 
plates  with  measured  drawings,  innumerable 
text  figures.  $60.00 — Robin  Restall  is  a life- 
long enthusiast  of  caged  finches,  whose  em- 
ployment with  an  international  advertising 
agency  has  taken  him  to  most  parts  of  the 
world  where  he  could  learn  about  many  of  the 
species  of  finches  in  their  native  contexts.  His 


particular  interest  is  with  the  genus  Lonchura, 
which  now  includes  all  of  the  birds  tradition- 
ally referred  to  as  munias,  mannikins,  and  the 
Java  and  Timor  Sparrows  which  were  former- 
ly considered  in  the  genus  Padda. 

This  is  a masterful  collection  of  virtually 
every  thing  known  about  this  assemblage  of 
estrildine  finches.  The  book  starts  off  with  a 
review  of  the  taxonomy  and  relationships  of 
the  genus  Lonchura,  continues  with  an  over- 
view of  their  natural  history,  and  concludes 
with  a detailed  accounting  of  all  pertinent  in- 
formation about  each  individual  species  (in- 
cluding one  apparently  new,  undescribed  spe- 
cies). 

Each  species  account  includes  headings  for 
field  characters,  status,  habitat,  habits  and  be- 
haviors, food  and  feeding,  movements,  call, 
song,  courtship  and  display,  breeding,  distri- 
bution, description  (including  all  known  sub- 
specific variants),  hybrids,  conservation,  and 
a list  of  relevant  references.  There  are  color 
plates  for  each  species,  showing  all  of  the 
most  obvious  sex,  age,  and  geographic  vari- 
ants and  an  even  more  impressive  collection 
of  color  plates  of  measured  drawings  that 
show  the  dorsal  and  ventral  views  with  one 
wing  extended. 

1 found  this  to  be  an  excellent  summary  of 
information  on  a group  of  birds  about  which 
I knew  relatively  little.  Yale  University  Press 
has  done  a fine  job  of  editing  (I  found  very 
few  typographic  errors),  and  the  color  repro- 
duction of  the  plates  appears  to  be  first-rate.  I 
recommend  the  volume  highly. — HERBERT 
T.  HENDRICKSON. 


FOREST  PATCHES  IN  TROPICAL 
LANDSCAPES.  By  John  Schelhas  and  Rus- 
sell Greenberg,  Eds.  Island  Press,  Washington, 
D.C.  1996:  426  pp.,  maps,  tables,  black-and- 
white  figures.  $30.00  (paper). — It  is  obvious 
to  anyone  who  studies  any  aspect  of  global 
tropical  ecology  that  forest  fragmentation  is 
increasing  annually.  The  complex  ecological 
effects  of  fragmentation  are  only  now  becom- 
ing known,  an  increasing  database  generated 
from  the  efforts  of  numerous  researchers  in 
tropical  regions  around  the  world.  Most  trop- 
ical ecologists  are  familiar  with  the  ongoing 
study  known  now  as  the  Biological  Dynamics 


ORNITHOLOGICAL  LITERATURE 


149 


of  Forest  Fragments  Project,  located  north  of 
Manaus,  Brazil,  but  this  multi-authored  vol- 
ume includes  data  from  numerous  studies  in 
other  regions  (as  welt  as  one  chapter  dealing 
with  the  BDFFP). 

The  book  is  divided  into  four  parts:  chang- 
ing forests,  regional  landscapes,  human  di- 
mensions, and  management.  The  introduction, 
authored  by  the  editors,  provides  a concise  but 
thorough  overview  of  the  issues  covered  in 
depth  throughout  the  volume.  Birds  are  the 
focus  of  but  one  of  the  19  chapters,  but  are 
discussed  to  varying  degrees  in  numerous  oth- 
er chapters.  Most  of  the  chapters  deal  with 
South  and  Central  America  though  there  is 
one  chapter  on  Africa,  one  largely  on  Indo- 
nesia, and  one  on  India.  Each  chapter  is  ref- 
erenced from  the  primary  literature. 

This  volume,  broad  in  scope,  and  excel- 
lently edited,  is  an  important  resource  for 
tropical  ecologists,  particularly  those  whose 
research  is  focused  on  biodiversity  preserva- 
tion and  ecologically  sound  management  pol- 
icies.—JOHN  C.  KRICHER. 


WHERE  TO  WATCH  BIRDS  IN  ASIA.  By 
Nigel  Wheatley.  Princeton  Univ.  Press, 
Princeton,  New  Jersey,  1996:  463  pp.,  51  line 
drawings,  8 figs.,  105  maps.  $35.00. — This  re- 
markable book  compresses  an  impressive 
amount  of  birding  data  from  our  largest  con- 
tinent into  a relatively  compact  volume.  Asia 
(including  the  island  nations  of  Indonesia  and 
the  Philippines)  harbors  just  under  2,700  bird 
species.  Considering  that  various  parts  of  the 
continent  have  been  off  limits — or  inaccessi- 
ble— to  outsiders  for  much  of  our  lifetimes,  I 
was  delighted  to  see  how  many  places  have 
now  been  surveyed  by  keen  birders.  The  au- 
thor is  to  be  congratulated  on  gathering  these 
scattered  data  into  one  volume.  The  book  is 
not  meant  as  an  encyclopedic  reference  to  all 
sites  or  to  all  species  in  a given  country  and 
the  author  rightly  suggests  that  the  informa- 
tion may  be  best  used  as  a first  “guiding 
light”,  a starting  point  in  travel  planning.  The 
book  fills  this  function  admirably  and  by 
studying  material  presented  here  you  can 
quickly  focus  on  desirable  Asian  locations  and 
the  species  found  there.  The  format  is  ap- 
pealing and  easy  to  follow,  while  the  text. 


composed  in  a clear  9 point  Cheltenham 
Light,  is  remarkably  free  of  typographical  er- 
rors. 

The  book  is  divided  into  three  parts:  a gen- 
eral introduction,  the  main  text  (organized  by 
countries),  and  additional  suggestions  and  in- 
dexes. The  introduction  is  necessarily  brief 
but  explains  the  book’s  layout  and  then  pro- 
vides much  useful  information  relating  to 
birds  and  birding  in  Asia  and  includes  notes 
on  habitat  diversity,  bird  diversity  (at  the  fam- 
ily and  species  levels),  and  how  Asia  com- 
pares with  other  continents  (400  species  more 
than  Africa  and  400  less  than  South  America). 
Conservation  is  given  a separate  six  paragraph 
section  where  Wheatley  is  brave  enough  to 
identify  human  population  growth  as  a major 
conservation  problem  and  to  call  for  growth 
stabilization,  or  better  yet,  “we  should  aim  to 
reduce”  population  growth.  Under  General 
Tips  we  learn  of  familiar  techniques  for  good 
birding  (some  Asian  birds  are  very  shy)  and 
of  various  dangers  to  travelers.  Altitude  sick- 
ness (in  the  higher  Himalayas  and  Tibet)  is 
mentioned  but  I find  the  subject  treated  too 
lightly.  The  recommended  one  night  at  3000 
m.  (9,843')  before  ascending  to  4500  m. 
(14,764')  the  next  night  is  far  too  fast  for 
many  people  hiking  in  the  Himalayas.  Simi- 
larly, it  is  suggested  that  if  one  “turns  blue 
and  coughs  up  pink  mucus”  one  should  im- 
mediately descend  to  below  3000  m.  (9,843'). 
Actually  one  should  have  descended  well  be- 
fore this  grim  stage  is  reached.  Keep  in  mind, 
that  “descending  immediately”  from  some  lo- 
cations in  Tibet  is  not  possible. 

In  the  main  text,  forty  countries  are  sum- 
marized in  varying  detail  (one  paragraph  for 
the  Maldives  to  57  pages  for  Indonesia;  North 
Asia  is  not  covered).  English  names  follow 
those  used  by  James  Clements  Birds  of  the 
World:  A Check  List  (Fourth  edition,  1991  and 
Supplements)  and  is  a most  useful  correlation. 
Yet  it  is  a pity  that,  except  for  189  species 
given  in  an  appendix,  scientific  names  are  not 
recorded  somewhere  in  the  book. 

Each  country  report  follows  a standard  for- 
mat that  starts  with  a general  map,  followed 
by  a summary  and  then  short  notes  on  size, 
transport  within  the  country,  accommodations, 
health  and  safety,  climate  and  timing,  habitats, 
conservation,  special  birds,  a note  on  endem- 
ics, and  how  many  species  one  might  expect 


150 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  I,  March  1999 


to  see  in  a stated  time.  These  sections  are 
helpful  planning  tools.  Accommodations  and 
travel  within  the  country  are  only  lightly 
touched  upon  as  vast  descriptions  are  avail- 
able in  various  general  guides:  accommoda- 
tions mentioned  are  usually  aimed  at  the  in- 
dividual traveling  on  a limited  budget. 

Following  this  general  introduction,  impor- 
tant birding  sites  are  covered  in  a clear  format. 
This  material  starts  with  an  overall  note  and 
sometimes  a map  (of  some  250  sites,  184  are 
not  mapped)  and  then  moves  to  a list  of  en- 
demics seen  at  that  particular  site,  followed  by 
specialties,  others,  and  finally  other  wildlife. 
Maps  are  so  helpful  that  I would  like  to  have 
seen  more.  The  51  line  drawings  (mostly  of 
excellent  quality)  sprinkled  throughout  add  vi- 
sual appeal  but  do  not  enhance  the  usefulness 
of  the  text.  Dropping  line  drawings  to  add 
maps  would  make  the  book  more  utilitarian 
albeit  less  attractive.  Each  country  section 
ends  with  Additional  Information  that  in- 
cludes addresses  of  local  bird  clubs  and  nature 
societies,  suggested  readings,  a complete  list 
of  all  the  country’s  endemics,  and  finally  a 
paragraph  giving  near-endemics. 

The  book  nears  the  end  with  yet  additional 
addresses  that  include  various  general  socie- 
ties and  clubs  dealing  with  Asia,  where  one 
may  obtain  trip  reports,  and  of  16  companies 
that  do  birding  tours.  After  a selection  of  gen- 
eral book  titles  there  are  three  pages  of  three- 
column  fine  print  that  give  scientific  names 
that  correspond  with  Clements  English  names, 
and  then  with  other  English  names  used  in 
Asian  bird  books  (where  these  differ  from  the 
names  used  by  Clements).  The  book  con- 
cludes with  two  indexes. 

With  any  volume  of  this  magnitude  readers 
will  have  varying  opinions  on  the  coverage 
and  the  presentation  of  the  material.  Perhaps 
the  most  serious  omission  of  the  book,  to  my 
mind,  is  that  there  is  no  clue  as  to  the  abun- 
dance of  the  species  listed.  Thus  there  is  no 
distinction  between  a bird  that  has  been  seen 
once  at  that  site  or  another  that  is  recorded  in 
numbers  every  day.  Even  a two-tier  indication 
giving  an  “r”  for  very  rare  species  and  an 
“a”  for  an  abundant  species  would  be  helpful. 
Similarly,  some  birds  move  seasonally  and  at 
times  the  text  gives  no  idea  as  to  when  the 
bird  might  be  at  the  site  described.  Tickell’s 
Leaf  Warbler  {Phylloscopus  affinis),  for  ex- 


ample, is  listed  for  Corbett  National  Park  (In- 
dia), but  for  much  of  the  year  this  migratory 
species  is  not  in  the  park. 

A curious  paragraph  called  “near-endem- 
ics” appears  at  the  end  of  each  country  sum- 
mary. To  save  space,  this  section  could  easily 
be  dropped  or  at  least  the  definition  of  “near- 
endemic” tightened.  It  is  hard  to  see  how  the 
Nepal  Fulvetta  {Alcippe  nipalensis)  is  a “near- 
endemic” when  it  is  listed  for  Bangladesh, 
Bhutan,  Burma,  Nepal,  and  northeast  India. 

As  common  to  many  first  editions  dealing 
with  this  much  detail,  there  are  a number  of 
minor  factual  errors.  For  example,  the  Purple- 
rumped  Sunbird  (India)  is  listed  under  “more 
or  less  throughout”  while  the  bird  does  not 
occur  in  the  north.  The  plural  of  genus,  p.25, 
is  genera  (not  genuses).  Similarly,  Padang  en 
route  to  Kirinci-Seblat  National  Park  (Indo- 
nesia) is  twenty-four  hours  (not  six)  by  bus 
from  Berestagi  (Brastagi) — twenty-four,  that 
is,  if  one  is  lucky.  The  Khunjerab  Pass  (Pak- 
istan) is  close  to  16,000  feet  (not  5575  m., 
18,290').  On  page  38  we  learn  that  “most  of 
the  pristine  forest  which  is  left  in  the  eastern 
Himalayas  is  in  . . . Bhutan.”  In  reality,  there 
is  far  more  eastern  Himalayan  forest  in  Arun- 
achal  Pradesh  than  in  Bhutan  but  much  of 
Arunachal  is  still  off  limits  to  outsiders.  Nam- 
dapha  National  Park  (India)  is  accorded  “the 
greatest  altitudinal  range  of  any  park  in  the 
world”  but  the  Sagarmatha  National  Park  and 
the  Makalu-Barun  National  Park  and  Conser- 
vation Area  in  Nepal  and  the  Quomolungma 
Nature  Reserve  in  Tibet  cover  substantially 
more  altitude  than  does  Namdapha. 

I fear  this  book  will  not  sell  well  in  the 
Maldives  for  on  page  124  the  Maldives  dis- 
appears as  a country  only  to  resurface  as  a 
single  paragraph  under  India.  This  treatment 
would  surprise  the  citizens  of  this  Indian 
Ocean  nation.  Agreed,  the  Maldives  may  not 
be  particularly  good  for  birding,  but  it  has 
world  class  coral  reefs  and  scuba  diving  and 
should  be  accorded  full  country  status  in  the 
next  edition. 

These  errors,  however,  are  of  a minor  nature 
and  do  not  detract  from  the  importance  of  this 
effort  and  1 strongly  recommend  the  book  to 
anyone  who  is  even  remotely  thinking  about 
those  nearly  2700  Asian  species  that  are  just 
waiting  to  be  seen.-ROBERT  L.  FLEMING, 
JR. 


ORNITHOLOGICAL  LITERATURE 


151 


THE  BIRDS  OF  SULAWESI.  By  Derek 
Holmes  and  Karen  Phillipps.  Oxford  Univer- 
sity Press,  Oxford,  U.K.  1997:  86  pp.,  20  col- 
or plates  illustrating  142  species,  22  black- 
and-white  illustrations,  one  table.  $24.95 
(cloth). — The  island  of  Sulawesi,  part  of  the 
vast  country  of  Indonesia,  has  only  relatively 
recently  been  frequented  by  ecotourists  and 
birders.  Sulawesi  is  perhaps  best  known  for 
Torajaland,  where  the  local  people  practice 
elaborate  funeral  rites  and  the  dead  are  placed 
in  cliffside  alcoves,  commemorated  with 
unique  statues.  Sulawesi  was  of  great  interest 
to  Alfred  Russel  Wallace,  as  it  sits  almost 
astride  “Wallace’s  Line,”  separating  two  bio- 
geographic realms.  The  authors  of  this  guide 
provide  a brief  table  that  compares  the  avifau- 
na of  Kalimantan,  on  one  side  of  Wallace’s 
Line,  with  Sulawesi,  on  the  other  side.  They 
note  that  of  the  380  species  found  on  Sulawesi 
and  its  near  neighbor  islands  (what  the  authors 
call  the  “Sulawesi  region”),  96  are  endemic 
to  the  region,  and  115  are  endemic  to  Indo- 
nesia. 

While  this  guide  is  admittedly  not  compre- 
hensive, it  will  prove  very  useful.  The  species 
treated  are  those  most  commonly  seen,  the  il- 
lustrations are  of  good  quality,  and  the  text 
descriptions  are  adequate  for  identification.  It 
would  be  helpful  if  some  maps  were  included 
but  none  are.  The  authors  state  that  “For  the 
bird-watcher,  Sulawesi  is  unequalled.”  That  is 
certainly  not  true.  Much  of  the  island  is  de- 
voted to  rice  farming  and  birds  other  than  mu- 
nias  and  some  herons,  are  sparse.  To  get  any 
real  sense  of  the  endemic  avifauna  one  must 
visit  one  or  more  of  the  nature  reserves  and 
national  parks,  where  some  natural  forest  sur- 
vives. Like  much  of  the  rest  of  Indonesia, 
roadside  birding  is  disappointing,  with  a sur- 
prising paucity  of  birdlife. 

The  book  is  indexed  and  provides  a check- 
list of  resident  land  birds  in  the  Sulawesi  Fau- 
nal Region.— JOHN  C.  KRICHER. 


THE  WHOOPING  CRANE:  NORTH 
AMERICA’S  SYMBOL  OF  CONSERVA- 
TION. By  Jerome  J.  Pratt.  Castle  Rock  Pub- 
lishing, Prescott,  Arizona.  1996:  171  pp.,  46 
photographs.  $12.95  (paper). — The  Whooping 
Crane  (Grus  americcma)  is  a species  that  has 


been  close  to  extinction  throughout  my  life- 
time and  has  rightly  come  to  symbolize  the 
conservation  movement  in  North  America. 
This  book  is  an  effort  to  relate  the  history  of 
the  species  and  the  efforts  made  to  prevent  it 
from  disappearing  completely.  The  author,  Je- 
rome Pratt,  was  one  of  the  charter  members 
of  the  Whooping  Crane  advisory  group  estab- 
lished in  1956,  and  thus  provides  an  “insid- 
er’s” perspective  on  how  the  recovery  project 
has  unfolded.  This  may  be  the  book’s  greatest 
short-coming.  Saving  endangered  species  ap- 
pears to  have  some  strong  similarities  to  the 
making  of  sausage;  you  don’t  necessarily  want 
to  know  everything  that  goes  into  it. 

Pratt  clearly  has  some  strong  opinions  on 
how  the  recovery  program  should  have  pro- 
ceeded, which  differ  from  the  way  it  actually 
unfolded.  However,  I found  it  difficult  in 
reading  the  text  to  determine  precisely  what 
the  points  of  difference  were.  Perhaps  some- 
one with  less  personal  involvement  would 
have  been  able  to  present  a more  clear  ex- 
planation of  exactly  how  the  differing  phi- 
losophies interacted  in  the  varying  political 
environments  to  determine  the  decisions  that 
were  made. 

I found  it  troublesome  that  nowhere  in  the 
text  is  there  a simple  graph  describing  the 
numbers  of  Whooping  Cranes  extant  in  the 
wild  and/or  in  captivity.  Much  of  the  data  is 
mentioned  in  the  text,  but  it  does  not  appear 
to  be  complete  and  it  is  extremely  difficult  to 
locate.  It  is  clear  that  there  are  presently  many 
more  Whooping  Cranes  than  there  were  in, 
say  1954,  but  exactly  how  the  various  man- 
agement techniques  that  have  been  applied  re- 
late to  the  number  of  birds  is  not. 

There  appears  to  have  been  only  minimal 
editing  done  on  the  original  manuscript  and 
I found  a large  number  of  spelling  errors  and 
other  typographic  mistakes  all  over  the  book. 
Most  of  the  time  these  were  merely  annoy- 
ing, but  when  numbers  were  transposed  in 
years,  these  mistakes  were  greatly  mislead- 
ing. 

The  strongest  feature  of  this  book  is  its  28- 
page  bibliography.  This  contains  essentially 
all  of  the  relevant  literature  on  Whooping 
Cranes  and  is  a must  for  anyone  who  wishes 
to  try  writing  the  definitive  history  of  Whoop- 
ing Crane  conservation  in  North  America. — 
HERBERT  T.  HENDRICKSON. 


152 


THE  WILSON  BULLETIN  • Vol.  11],  No.  1,  March  1999 


A PASSION  FOR  BIRDS.  AMERICAN 
ORNITHOLOGY  AFTER  AUDUBON.  By 
Mark  V.  Barrow,  Jr.  Princeton  University 
Press,  Princeton,  New  Jersey.  1998:  326  pp., 
33  unnumbered  text  figures.  ISBN  0-691- 
04402-3.  $39.50  (cloth). — Histories  of  orni- 
thology, and  especially  of  American  ornithol- 
ogy are  rare  and  those  by  professional  histo- 
rians of  science  are  rarer  still  as  shown  by  a 
perusal  of  the  bibliography  of  this  excellent 
new  book  by  Mark  Barrow.  Barrow’s  A Pas- 
sion for  Birds  does  not  provide  such  a com- 
plete history  of  North  American  ornithology 
as  this  was  not  his  purpose  in  writing  this  par- 
ticular analysis;  ornithologists  who  are  look- 
ing for  a full  history  of  their  field  may  be 
disappointed,  but  should  not  be.  Rather  the 
goal  of  Barrow’s  book  is  to  examine  the  rise 
of  professionalization  in  ornithology  from  the 
death  of  Audubon  in  1851  to  1940.  It  is  a most 
excellent  treatment  of  this  important  aspect  of 
ornithology,  one  which  is  well  worth  the  close 
attention  of  everyone  interested  in  this  biolog- 
ical discipline.  A Passion  for  Birds  is  a su- 
perbly excellent  history  of  science  with  full 
documentation  and  an  exhaustive  bibliogra- 
phy. Barrow  chose  to  begin  his  analysis  at 
1850  because  this  coincides  with  Audubon’s 
death  in  1851  and  continues  Farber’s  analysis 
which  ended  in  1850.  Barrow  undertook  a 
most  intensive  study  of  archives  and  the  lit- 
erature, as  demonstrated  by  his  citations,  and 
presents  a tremendous  amount  of  information 
on  the  history  of  North  American  ornithology 
which  is  of  interest  to  all  ornithologists,  pro- 
fessional and  amateur  alike.  One  does  not 
have  to  be  a historian  of  science  to  enjoy  read- 
ing A Passion  for  Birds  and  to  learn  much 
from  it. 

Barrow  stres.ses  three  topics  in  his  analysis 
of  professionalization  of  North  American  or- 
nithology; these  are:  (a)  collecting  and  sys- 
tematics;  (b)  the  American  Ornithologists’ 
Union  founded  in  New  York  City  in  Septem- 
ber, 1883;  and  (c)  bird  conservation.  Quite 
clearly,  the  pathway  for  most  early  North 
Americans  into  ornithology,  whether  they  re- 
mained amateurs  or  became  professionals,  was 
via  the  accumulation  of  a collection  of  bird 
skins  or  eggs. 

Ornithology  during  the  19th  century  was 
characterized  by  inten.se  activity  amassing  col- 
lections and  describing  the  diversity  of  North 


American  birds,  first  species  and  then  subspe- 
cies, followed  by  those  in  the  rest  of  the  New 
World  and  finally  the  Old  World.  This  descrip- 
tive work  led  to  two  different  check-lists  of 
North  American  birds,  namely  by  Elliott 
Coues  and  by  Robert  Ridgway.  It  was  largely 
the  differences  between  these  check-lists 
which  led  to  the  founding  of  the  American 
Ornithologists’  Union  in  1883. 

One  of  the  important  aspects  in  the  profes- 
sionalization of  a science,  as  emphasized  by 
Barrow,  is  the  founding  of  a national  society 
and  the  publication  of  a scholarly  journal  by 
that  society.  This  was  certainly  true  in  the 
course  of  professionalization  of  ornithology  in 
North  America  even  if  the  initial  goals  of  the 
American  Ornithologists’  Union  were  not 
quite  so  noble.  Invitations  were  sent  to  a small 
group  of  ornithologists  with  the  clear  purpose 
of  establishing  a society  with  the  primary  goal 
of  solving  the  check-list  problem  and  with  a 
structure  which  kept  control  of  the  society  in 
the  hands  of  a small  group  of  leading  orni- 
thologists. The  new  Union  solved  splendidly 
the  classification  and  nomenclature  problems 
in  short  order,  publishing  the  first  edition  of 
its  check-list  and  its  code  of  nomenclature  in 
1886.  Professionalization  was  further  achieved 
with  the  decision  to  accept  the  offer  to  take 
over  the  Bulletin  of  the  Nuttall  Ornithological 
Club  as  The  Auk. 

However,  the  AOU  was  less  successful  in 
dealing  with  other  ornithological  matters.  The 
formal  hierarchical  structure  of  membership 
classes,  designed  to  keep  the  large  masses  of 
amateur  and  other  ornithologists  out  of  the 
running  of  the  Union,  caused  problems  from 
the  beginning  which  became  more  serious  as 
the  decades  passed;  this  archaic  system  is  still 
in  place  as  the  20th  century  draws  to  a close. 
Moreover,  the  Union  did  not  deal  readily  with 
other  ornithological  questions  such  as  migra- 
tion and  bird  protection,  largely  because  of  the 
preoccupation  of  the  leading  members  with 
matters  of  collecting  and  systematics. 

Barrow  discusses  in  detail  the  development 
of  conservation  and  bird  protection,  which  be- 
came a central  issue  for  the  Union  from  its 
origin.  He  showed  that  although  several  prom- 
inent members  of  the  AOU  had  central  roles 
in  the  development  of  bird  protection  in  North 
America,  the  Union  did  little  in  this  area.  A 
large  part  of  the  problem  stemmed  from  con- 


ORNITHOLOGICAL  LITERATURE 


153 


flicts  between  member  in  favor  of  collecting 
birds  for  scientific  (as  opposed  to  commercial) 
purposes  and  those  in  favor  of  bird  protection. 
Barrow  does  an  excellent  job  in  summiirizing 
these  arguments  and  showing  their  conse- 
quences for  the  Union  and  for  conservation 
groups;  but  he  fails  to  analyze  these  argu- 
ments and  therefore  does  not  show  that  these 
two  groups  were  largely  arguing  past  one  an- 
other. By  1900,  it  was  clear  that  collecting 
birds  had  nothing  to  do  with  the  decline  and 
extinction  of  avian  species.  Rather  it  was  clear 
that  the  primary  factors  were  market  hunting 
and  habitat  destruction. 

The  second  major  aspect,  and  perhaps  the 
more  difficult  one  to  analyze,  in  the  profes- 
sionalization of  a science  is  the  balance  be- 
tween amateurs  and  professionals.  Barrow 
states  (p.  5)  that  ornithology  is  “a  classic  ex- 
ample of  an  inclusive  scientific  field”  but  does 
not  clarify  this  concept.  He  pays  most  careful 
attention  to  the  professional-amateur  distinc- 
tion in  his  introduction  providing  numerous 
citations  to  the  literature  in  this  field,  but  does 
not  come  to  a clear  resolution  because  of  the 
quandaries  in  defining  an  amateur  (including 
whether  a single  definition  would  serve  equal- 
ly well  in  1880,  1930,  and  1980,  and  in  char- 
acterizing particular  persons  as  an  amateur  or 
as  a professional.  Was  Sir  William  Herschel 
(1738-1822)  an  amateur  astronomer  when  he 
discovered  the  planet  Uranus  in  1781,  a time 
when  he  was  still  earning  his  keep  as  a mu- 
sician? Were  William  Brewster,  Elliott  Coues, 
Margaret  M.  Nice,  or  Lord  Walter  Rothschild 
amateurs?  Possibly  this  quandary  will  never 
be  settled.  Major  questions  still  exist,  such  as: 
How  much  interest  in  ornithology  does  a per- 
son have  to  have  to  be  identified  as  an  amateur 
ornithologist?  Is  every  person  who  maintains 
a bird  feeding  station  and  possesses  a pair  of 
binoculars  and  a bird  identification  guide,  an 
amateur  ornithologist?  Perhaps  a distinction 
should  be  made  between  serious  amateur  or- 
nithologists and  all  others  (the  “hobby-orni- 
thologists” as  expressed  in  German).  What 
were  the  roles  of  amateurs  to  ornithology 
since  1850?  Since  1900?  Since  1940,  the  end 
of  Barrow’s  analysis?  And  is  ornithology  the 
science  in  which  professionals  are  so  outnum- 
bered by  amateurs  (see  statement  by  Frank 


Chapman,  p.  5)7  1 think  not.  1 suspect  that 
ornithologists  have  over-emphasized  the  num- 
ber and  importance  of  amateurs  in  their  sci- 
ence. Comparative  studies  are  needed.  But  1 
suspect  that  there  have  been  and  still  are  more 
serious  amateur  astronomers  than  serious  am- 
ateur ornithologists,  and  that  these  amateur  as- 
tronomers have  and  continue  to  contribute 
more  to  astronomy  than  amateur  ornitholo- 
gists do  to  ornithology. 

Nevertheless,  Barrow  was  able  to  demon- 
strate that  by  1940,  the  termination  of  his 
analysis.  North  American  ornithology  had  be- 
come fully  professionalized  from  its  largely 
non-professional  status  in  1850.  By  1940,  a 
good  majority  of  active  ornithologists  were 
professionals  based  on  publishing  scholarly 
works  in  avian  biology  and  earning  their  liv- 
ing in  a position  requiring  some  involvement 
in  the  study  and/or  conservation  of  birds.  Fifty 
years  later,  serious  amateur  ornithologists 
have  all  but  disappeared.  Clearly  profession- 
alization in  ornithology  and  other  sciences  is 
dependent  on  the  rise  of  paid  positions  in  the 
field,  but  this  aspect  has  not  been  examined 
for  ornithology  in  any  detail  by  Barrow  or 
specifically  mentioned  in  the  end  notes,  pos- 
sibly because  it  is  too  obvious.  Yet,  it  would 
be  most  interesting  to  have  a detailed  analysis 
of  the  employment  opportunities  for  North 
Americans  in  all  aspects  of  avian  biology 
from  teaching  and  research  to  conservation 
and  protection  as  well  as  the  diversity  of  pos- 
sibilities in  government,  publishing,  and  in- 
dustry, as  well  as  the  changes  in  these  em- 
ployment opportunities  over  the  decades. 

1 would  like  to  congratulate  Mark  Barrow 
on  his  excellent  analysis  of  the  professionali- 
zation of  North  American  ornithology  and  for 
presenting  it  in  the  clear  and  lively  style  used 
in  A Passion  for  Birds.  1 thoroughly  enjoyed 
reading  this  book  and  learned  a great  deal 
about  the  history  of  North  American  orni- 
thology and  the  course  of  professionalization 
of  a scientific  field.  Without  any  hesitation,  1 
can  recommend  most  firmly  A Passion  for 
Birds  to  everyone  with  any  interest  in  orni- 
thology. 1 would  like  to  thank  Keir  B.  Sterling 
for  answering  urgent  questions  on  the  fine 
points  of  American  ornithology  and  for  read- 
ing the  manuscript. — WALTER  J.  BOCK. 


154 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


THE  EBCC  ATLAS  OF  EUROPEAN 
BREEDING  BIRDS:  THEIR  DISTRIBU- 
TION AND  ABUNDANCE.  W.  J.  M.  Hage- 
meijer  and  M.  J.  Blair  (editors).  T.  and  A.  D. 
Poyser,  London.  1997:  cxli  + 903  pp. 
$88.00. — Twenty-six  years  in  the  making,  this 
book  is  a monumental  work  and  one  of  the 
three  most  significant  bird  books  to  emerge 
from  Europe  in  recent  years.  Weighing  in  at 
over  six  pounds  (if  my  bathroom  scales  are 
correct),  its  900-plus  large  format  (81/2"  X 
12W')  pages  report  on  the  distribution  and  rel- 
ative abundance  of  birds  across  the  European 
continent,  ranging  from  Gibralter  north  to  Ost- 
end,  and  east  to  Svalbard  and  Novaya  Zemlya 
and  Franz  Josef  Land  and  south  again  to  the 
eastern  Mediterranean.  Significant  parts  of 
Russia  and  the  Ukraine  are  covered,  as  is  Ma- 
deira and  the  Azores  but  the  Mediterranean 
islands  administered  from  North  Africa,  as 
well  as  Cyprus  and  the  Canary  Islands  are 
omitted.  In  addition,  an  important  part  of  the 
Western  Palearctic,  the  North  African  coast,  is 
also  omitted.  Within  this  area  some  10,000 
plus  ornithologists  from  every  European 
country  obtained  presence/absence  data  on 
some  495  species  in  over  4400  50  X 50  km^, 
often  with  estimates  of  the  order  of  magnitude 
of  each  population.  Coverage  is  nevertheless 
regionally  uneven,  with  large  areas  of  former 
USSR  countries  lacking  any  data  at  all  and 
with  Norway,  Poland,  and  parts  of  four  west- 
ern countries  lacking  the  quantitative  popula- 
tion estimates  available  for  most  countries. 
The  area  to  be  covered  was  only  about  half 
again  as  large  as  the  conterminous  United 
States  and  the  number  of  countries  (ca  40,  de- 
pending on  which  year  one  considers)  was 
about  the  same  as  the  number  of  states  in  the 
U.S.  but  was  complicated  by  the  presence  of 
about  40  different  languages  and  the  occa- 
sional armed  conflict  between  countries! 
These  political,  language,  and  cultural  differ- 
ences meant  that  a large  part  of  the  success  of 
the  project  depended  on  the  use  of  relatively 
simple  methods  that  could  be  adopted  and  im- 
plemented relatively  rigorously  by  observers 
of  diverse  background. 

The  book  is  divided  into  two  major  sec- 
tions, one  of  141  (i-cxli)  introductory  pages, 
followed  by  903  substantive  pages.  A Fore- 
word by  the  eminent  biogeographer  K.  H. 
Voous,  a 3-page  Preface,  and  an  8-page  En- 


glish language  Introduction  describe  the  pur- 
pose and  structure  of  the  Atlas.  Figure  1 is 
particularly  important  in  that  it  documents  the 
completeness  of  coverage.  I could  find  no 
quantitative  figures  as  to  the  coverage 
achieved,  but  it  looks  as  if  about  two-thirds  of 
the  squares  received  data  for  at  least  75%  of 
the  breeding  species  expected  to  be  found 
there.  However,  a large  chunk  of  the  former 
USSR  received  no  data  at  all  and  there  are 
also  gaps  in  Albania  and  some  of  the  outer 
island  groups.  The  remaining  squares,  partic- 
ularly concentrated  in  eastern  Europe,  re- 
ceived data  for  fewer  than  75%  of  the  breed- 
ing species  anticipated.  What  could  have  been 
usefully  included  here  is  a political  map  of  the 
region:  country  boundaries  are  shown  but  not 
identified,  leaving  readers  unfamiliar  with  the 
political  geography  of  Europe  to  guess  which 
country  is  which. 

The  Atlas  is  mapped  on  a Universal  Trans- 
verse Mercator  (UTM)  projection,  chosen  be- 
cause it  covered  the  intended  area  of  the  Atlas, 
was  familiar  in  most  European  countries,  and 
was  compatible  with  national  map  projections 
for  each  country.  Each  UTM  100  km  X 100 
km  square  was  subdivided  into  four  smaller 
50  X 50  km  grid  squares  to  parallel  an  earlier 
botanical  Atlas.  Because  lines  of  longitude 
converge  towards  the  poles,  gradual  reduction 
of  the  number  of  50  X 50  km  squares  on 
northern  lines  of  latitude  was  necessary.  De- 
spite this  the  visual  effect  on  the  final  Atlas  is 
very  acceptable,  with  one  having  to  look 
closely  to  find  areas  where  the  local  density 
of  dots  was  not  regularly  spaced. 

Originally  intended  for  1985-1988,  Atlas 
field  work  actually  took  far  longer.  Data  for 
Spain,  for  example,  spanned  from  1970  to 
1992,  those  from  Finland  and  from  Moldova 
spanned  1986—90,  those  from  Georgia  were 
for  1992,  and  the  few  species  that  were  cov- 
ered in  Azerbaijan  came  from  1994.  For  some 
countries  visitors’  records  contributed  signifi- 
cantly to  the  information  available.  A standard 
form  allowed  recording  the  breeding  status  of 
about  440  species  recorded  for  each  square. 
Seven  classes  of  information  (e.g.,  distraction 
displaying,  egg  shells  found,  fledged  young, 
nests  seen)  confirmed  breeding,  another  of 
seven  classes  (e.g.,  pairs  observed  in  suitable 
nesting  habitat  in  the  breeding  season,  court- 
ship display  observed,  nest  being  seen,  etc.) 


ORNITHOLOGICAL  LITERATURE 


155 


indicates  probable  breeding,  and  two  catego- 
ries (species  observed  in  possible  nesting  hab- 
itat in  the  breeding  season,  and  seeing  males 
present  in  breeding  season)  indicated  possible 
breeding.  The  validity  of  the  data  was  re- 
viewed through  a hierarchy  of  subsequent 
checking,  ranging  from  EBCC  national  and 
regional  coordinators  through  species  experts 
to  the  authors  of  the  species  texts.  In  addition 
to  the  presence/absence  data,  the  Atlas  sought 
to  include  logarithmic  population  size  esti- 
mates as  semi-quantitative  information  of 
population  levels.  However,  organizers  in 
some  countries  (Norway  and  Poland)  refused 
entirely  to  provide  such  estimates  and  small 
or  large  parts  of  several  other  countries  (Ice- 
land, France,  Italy,  Spain)  likewise  lack  such 
estimates.  Not  surprisingly  with  armed  con- 
flict there  “the  project  lost  contact  with  Bos- 
nia and  Serbia,  ...”  and  had  to  make  use  of 
earlier  presence/absence  data  from  that  region. 
Estimates  from  border  squares  between  coun- 
tries were  merged  to  the  higher  value. 

The  main  body  of  the  Atlas  consists  of  spe- 
cies accounts.  Part  1 includes  496  species  for 
which  the  mapping  data  were  satisfactory  in 
quality;  a second  group  of  17  poorly  covered 
species  is  covered  in  a series  of  briefer  ac- 
counts. Each  species  account  covers  as  far  as 
possible  a list  of  standard  topics,  including 
world  distribution,  breeding  habitat,  distribu- 
tion and  abundance  in  Europe,  recent  changes 
in  status,  and  migration  patterns.  The  accounts 
do  not  cover  breeding  biology  and  conserva- 
tion status  since  these  two  topics  are  respec- 
tively covered  in  detail  in  the  standard  work 
on  the  region  (Cramp  et  al."s  1977-94  nine- 
volume  Birds  of  the  Western  Palearctic)  and 
in  Tucker  and  Heath’s  (1994)  Birds  in  Eu- 
rope— their  conservation  status.  The  status  of 
subspecies  is  referenced  briefly  in  the  species 
section.  For  each  species  a population  size  es- 
timate is  provided  as  the  geometric  mean  of 
the  population  size  in  pairs,  together  with  es- 
timates of  the  minimum  and  maximum  pop- 
ulation, the  year  of  the  estimate,  and  the  pop- 
ulation trend  in  numbers  and  range  size.  Pop- 
ulation change  estimates  cover  five  categories 
of  decreases  or  increases  of  more  than  50%, 
of  20-50%,  and  stable  (±  20%). 

The  English  introduction  is  followed  by  a 
translation  of  the  introduction  into  thirteen 
languages,  namely  Czech,  German,  Castilian, 


French,  Finnish,  Greek,  Hungarian,  Italian, 
Dutch,  Portuguese,  Polish,  Russian,  and 
Swedish.  Then  follow  two  figures  showing  the 
maximum  data  quality  and  the  minimum  data 
quality  for  each  square,  for  the  best  coverage 
and  worst  coverage  species  respectively. 
These  are,  in  my  view,  rather  uninformative 
figures,  given  the  extreme  nature  of  the  out- 
liers. A copy  of  the  recording  form  then  pre- 
cedes a four  page  account  of  the  project’s  or- 
ganization and  background.  Reading  between 
the  lines,  it  is  evident  that  two  countries,  Brit- 
ain and  the  Netherlands,  were  the  major  driv- 
ers of  the  project,  providing  both  initial  fund- 
ing and  much  in-kind  support.  Intriguingly,  a 
significant  amount  of  support  came  from  a 
“sponsor  a species”  campaign  in  which  in- 
dividuals, organizations,  and  commercial 
firms  undertook  to  sponsor  the  cost  of  analysis 
and  writing-up  species  accounts.  The  512  spe- 
cies required  authors  from  37  countries,  typi- 
cally involving  people  from  two  countries  at 
opposite  ends  of  the  species  range,  thus  pro- 
moting national  collaboration  and  a broad  per- 
spective on  the  treatment  of  the  species. 

The  remainder  of  the  introductory  material 
includes  an  8-page  account  of  the  evolution 
and  history  of  the  European  bird  fauna  by 
Jacques  Blondel,  an  acknowledgments  sec- 
tion, a bibliography  of  national  and  major  re- 
gional bird  books  for  Europe,  and  a 2-page 
introduction  to  the  individual  species  ac- 
counts. 

The  Atlas  maps  and  accompanying  species 
accounts  form  the  core  of  the  volume.  Each 
Atlas  map  contains  insets  for  six  island  groups 
(Svalbard,  Franz  Josef  Land,  Novaya  Zemlya, 
the  Azores,  the  Madeiran  archipelago,  and  the 
Selvagens).  However,  one  is  expected  to 
memorize  the  identity  of  each  of  these  insets, 
the  only  key  being  on  page  cxl.  Within  each 
map  different  colored  dots  distinguish  unsur- 
veyed squares  from  absence  and  mere  pres- 
ence from  estimated  abundance,  and  show 
whether  breeding  was  confirmed  or  probable 
or  merely  possible.  Dot  size  characterizes  log- 
arithmic abundance  (1-9  breeding  pairs,  10- 
99  pairs,  etc.,  through  “more  than  100,000 
breeding  pairs”).  For  parts  of  Russia  colored 
shading  indicates  extrapolation  of  presence 
from  earlier  literature  was  necessary.  Most 
species  accounts  in  Part  1 also  include  a graph 
and  pie  chart  providing  information  on  pop- 


156 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  I,  March  1999 


ulation  size  and  trends  in  each  of  the  major 
countries,  a tally  of  the  total  number  of  Eu- 
ropean countries  in  which  the  species  breeds, 
and  an  estimate  of  its  European  population 
(excluding  Russia  and  Turkey).  Most  species 
accounts  are  illustrated  by  a black  and  white 
drawing  of  the  species  by  a variety  of  Euro- 
pean artists  and  each  has  a list  of  the  species 
name  in  14  European  languages.  The  accounts 
for  the  17  irregular  or  rare  breeding  species  in 
Part  2 typically  include  an  illustration  and  a 
short  text  commenting  on  its  status. 

The  species  accounts  are  followed  by  a 
short  (5-page)  account  of  the  conservation  sta- 
tus of  European  birds,  adapted  by  Melanie 
Heath  and  Graham  Tucker  from  their  1994 
book  Birds  in  Europe — their  conservation  sta- 
tus. A short  summary  provides  an  overview 
for  the  European  avifauna  as  a whole.  The  re- 
maining pages  are  essentially  technical  mis- 
cellania,  with  a 65-page  bibliography  and  in- 
dices of  scientific  names  and  indices  in  the 
various  languages  of  the  book  concluding  the 
work. 

There  is  an  old  joke  in  computer  science  to 
the  effect  that  it  pays  to  be  first  or  third:  if 
you  are  first,  you  get  the  credit;  if  you  are 
third,  you  get  something  that  actually  works. 
This  volume  is  undoubtedly  in  the  first  cate- 
gory. It  provides  an  excellent  approximation 
to  a continent-wide  atlas,  and  one  that  is  in- 
finitely superior  to  the  range  maps  drawn  by 


guesstimate  in  earlier  regional  or  European 
avifauna  and  field  identification  guides.  How- 
ever, the  differing  effort  afforded  by  different 
countries,  the  variation  in  timing  of  field 
work,  the  differing  national  perspectives  as  to 
the  inclusion  of  population  estimates,  and  the 
major  uncertainties  about  the  status  of  species 
within  the  former  USSR,  limit  the  scientific 
quality  of  this  work.  Nevertheless,  the  cover- 
age of  some  4400  50  X 50  km  squares  by  a 
network  of  volunteer  observers  organized  on 
essentially  a shoestring  budget  is  a stunning 
achievement.  For  a North  American  audience 
this  work  takes  away  any  excuse  for  not  hav- 
ing a Canada/United  States/Mexico  atlas  of 
bird  distributions,  the  major  gap  in  our  current 
spectrum  of  ornithological  resources.  For  Eu- 
ropean biogeographers  it  creates  a tremendous 
new  database  that  will  undoubtedly  fuel  in- 
novative analyses  of  large-scale  bird  distri- 
butions on  a scale  previously  impossible.  And 
for  a European  Union  that  is  increasingly 
moving  to  becoming  a “United  States  of  Eu- 
rope” this  work  provides  a remarkable  con- 
servation resource  that  should  guide  pan-Eu- 
ropean  conservation  strategies,  at  least  for 
birds,  over  the  next  few  decades.  For  orni- 
thologists of  all  nationalities  the  book  is  a 
beautifully  produced  volume  that  will  invite 
browsing,  stimulate  comparisons,  and  provoke 
thought  for  years  to  come — RAYMOND  J. 
O’CONNOR. 


'Ihis  issue  of  rhe  Wilson  linllelin  was  published  on  I March  1999. 


THE  WILSON  BULLETIN 


Editor  ROBERT  C.  REASON 


Editorial  Board  KATHY  G.  BEAL 


Department  of  Biology 

State  University  of  New  York 

1 College  Circle 

Geneseo,  NY  14454 

E-mail:  WilsonBull@geneseo.edu 


CLAIT  E.  BRAUN 
RICHARD  N.  CONNER 


Review  Editor  WILLIAM  E.  DAVIS,  JR. 


127  East  Street 

Foxboro,  Massachusetts  02035 


Editorial  Assistants  TARA  BAIDEME 


Index  Editor  KATHY  G.  BEAL 


JOHN  LAMAR 
DANTE  THOMAS 
DORIS  WATT 


616  Xenia  Avenue 
Yellow  Springs,  Ohio  45387 


SUGGESTIONS  TO  AUTHORS 

See  Wilson  Bulletin,  110:152-154,  1998  for  more  detailed  “Instructions  to  Authors.” 
http://www.ummz.lsa.umich.edu/birds/wilsonbull.html 
Submit  four  copies  of  manuscripts  intended  for  publication  in  The  Wilson  Bulletin,  neatly  typewritten, 
double-spaced,  with  at  least  3 cm  margins,  and  on  one  side  only  of  good  quality  white  paper.  Do  not 
submit  xerographic  copies  that  are  made  on  slick,  heavy  paper.  Tables  should  be  typed  on  separate  sheets, 
and  should  be  narrow  and  deep  rather  than  wide  and  shallow.  Follow  the  AOU  Check-list  (Seventh  Edition, 
1998)  insofar  as  scientific  names  of  U.S.,  Canadian,  Mexican,  Central  American,  and  West  Indian  birds 
are  concerned.  Abstracts  should  be  brief  but  quotable.  Where  fewer  than  5 papers  are  cited,  the  citations 
may  be  included  in  the  text.  Follow  carefully  the  style  used  in  this  issue  in  listing  the  literature  cited; 
otherwise,  follow  the  “CBE  Scientific  Style  and  Format  Manual”  (AIBS  1994).  Photographs  for  illustra- 
tions should  have  good  contrast  and  be  on  glossy  paper.  Submit  prints  unmounted  and  provide  a brief  but 
adequate  legend  for  each  figure  with  all  captions  on  a single  page.  Do  not  write  heavily  on  the  backs  of 
photographs.  Diagrams  and  line  drawings  should  be  in  black  ink  and  their  lettering  large  enough  to  permit 
reduction.  Original  figures  or  photographs  submitted  must  be  smaller  than  22  X 28  cm.  Alterations  in 
copy  after  the  type  has  been  set  must  be  charged  to  the  author. 


NOTICE  OF  CHANGE  OF  ADDRESS 


If  your  address  changes,  notify  the  Society  immediately.  Send  your  complete  new  address  to  Ornitho- 
logical Societies  of  North  America,  P.O.  Box  1897,  Lawrence,  KS  66044-8897. 

The  permanent  mailing  address  of  the  Wilson  Ornithological  Society  is:  c/o  The  Museum  of  Zoology, 
The  University  of  Michigan,  Ann  Arbor,  Michigan  48109.  Persons  having  business  with  any  of  the  officers 
may  address  them  at  their  various  addresses  given  on  the  back  of  the  front  cover,  and  all  matters  pertaining 
to  the  Bulletin  should  be  sent  directly  to  the  Editor. 


MEMBERSHIP  INQUIRIES 


Membership  inquiries  should  be  sent  to  Laurie  J.  Goodrich,  Route  2 Box  301  A,  New  Ringgold,  PA 
17960-9445;  E-mail:  goodrich@haukmountain.org. 


CONTENTS 


MAJOR  PAPERS 

ANNUAL  SURVIVAL  RATES  OF  FEMALE  HOODED  MERGANSERS  AND  WOOD  DUCKS  IN 

SOUTHEASTERN  MISSOURI  — Katie  M.  Dugger,  Bruce  D.  Dugger,  and  Leigh  H.  Fredrickson  1 
COMPARATIVE  NEST  SITE  HABITATS  IN  SHARP-SHINNED  AND  COOPER’S  HAWKS  IN  WIS- 
CONSIN   Dale  R.  Tre.xel,  Robert  N.  Rosenfield,  John  Bielefeldt,  and  Eugene  A.  Jacobs  1 

MADAGASCAR  FISH-EAGLE  PREY  PREFERENCE  AND  FORAGING  SUCCESS  

James  Berkehnan,  James  D.  Fraser,  and  Richard  T.  Watson  1 5 

THE  RELATIONSHIP  BETWEEN  SPOTTED  OWL  DIET  AND  REPRODUCTIVE  SUCCESS  IN  THE 

SAN  BERNARDINO  MOUNTAINS,  CALIFORNIA  

Richard  B.  Smith,  M.  Zachariah  Peery,  R.  J.  Gutierrez,  and  William  S.  Lcdiaye  22 

FOOD,  FORAGING,  AND  TIMING  OF  BREEDING  OF  THE  BLACK  SWIFT  IN  CALIFORNIA 

— Manuel  Marin  30 

FACTORS  THAT  INFLUENCE  TRANSLOCATION  SUCCESS  IN  THE  RED-COCKADED  WOOD- 
PECKER   Kathleen  E.  Eranzreb  38 

BANDING  RETURNS,  ARRIVAL  PATTERN,  AND  SITE-FIDELITY  OF  WHITE-EYED  VIREOS  

- S.  L.  Hopp,  A.  Kirby,  and  C.  A.  Boone  46 

RESPONSE  OF  BROWN-HEADED  NUTHATCHES  TO  THINNING  OF  PINE  PLANTATIONS  

Michael  D.  Wilson  and  Bryan  D.  Watts  56 

DIFFERENCES  IN  MIGRATORY  TIMING  AND  ENERGETIC  CONDITION  AMONG  SEX/AGE 

CLASSES  IN  MIGRANT  RUBY-CROWNED  KINGLETS  , 

- — David  L.  Swanson,  Eric  T.  Liknes,  and  Kurtis  L.  Dean  61 

SCALE-DEPENDENT  HABITAT  SELECTION  BY  AMERICAN  REDSTARTS  IN  ASPEN-DOMINAT- 
ED FOREST  FRAGMENTS  Navjot  S.  Sodhi,  Cynthia  A.  Paszkowski,  and  Shannon  Keehn  70 

FEMALE  MATE  CHOICE  IN  NORTHERN  CARDINALS:  IS  THERE  A PREFERENCE  FOR  REDDER 

MALES?  L.  Lareesa  Wolfenbarger  lb 

FRUIT  SUGAR  PREFERENCES  OF  HOUSE  FINCHES  '..... 

Michael  L.  Avery,  Carrie  L.  Schreiber,  and  David  G.  Decker  84 

HIERARCHICAL  COMPARISONS  OF  BREEDING  BIRDS  IN  OLD-GROWTH  CONIFER-HARD- 

WOOD  FOREST  ON  THE  APPALACHIAN  PLATEAU  J.  Christopher  Haney  89 

EFFECTS  OF  WIND  TURBINES  ON  UPLAND  NESTING  BIRDS  IN  CONSERVATION  RESERVE 

PROGRAM  GRASSLANDS  Krecia  L.  Leddy,  Kenneth  E.  Higgins,  and  David  E.  Naugle  100 

AVIAN  USE  OF  PURPLE  LOOSESTRIFE  DOMINATED  HABITAT  RELATIVE  TO  OTHER  VEGE- 
TATION TYPES  IN  A LAKE  HURON  WETLAND  COMPLEX  

— — Michael  B.  Whitt.  Harold  H.  Prince,  and  Robert  R.  Cox,  Jr.  105 

SHORT  COMMUNICATIONS 

BALD  EAGLE  PREDATION  ON  COMMON  LOON  CHICK  

James  D.  Paruk,  Dean  Seanfield,  and  Tara  Mack  1 1 5 

TERRITORIAL  TAKEOVER  IN  COMMON  LOONS  (GAV/A  IMMER) James  D.  Paruk  116 

COURTSHIP  BEHAVIOR  OF  THE  BUFF-NECKED  IBIS  (THERISTICUS  CAUDATUS) 

Nathan  H.  Rice  1 1 8 

HABITAT  USE  BY  MASKED  DUCKS  ALONG  THE  GULF  COAST  OF  TEXAS  

- James  T.  Anderson  and  Thomas  C.  Tacha  1 19 

GIZZARD  CONTENTS  OF  PIPING  PLOVER  CHICKS  IN  NORTHERN  MICHIGAN 

Erance.sca  ./.  Cuthhert.  Brian  Scholtens.  Uiuren  C.  Wemnier.  and  Robxn  McLain  121 
NESTING  OF  FOUR  POORLY-KNOWN  BIRD  SPECIES  ON  THE  CARIBBEAN  SLOPE  OF 

COSTA  RICA  Bruce  E.  Young  and  James  R.  Zook  124 

SEXUAL  DIMORPHISM  IN  THE  SONG  OF  SUMICHRAST'S  WREN 

— Monica  Perez-Villafaha,  Hector  Gomez  de  Silva  G..  and  Atahualpa  DeSucre-Medrano  128 

AN  INCIDENT  OF  FEMALE-FEMALE  AGGRESSION  IN  THE  HOUSE  WREN  .. 

- lorn  Ahvorth  and  Isabella  B.R.  Scheiber  130 

NEST  REUSE  BY  WOOD  THRUSHES  AND  ROSE-BREASTED  GROSBEAKS  

- - Lyle  E.  Eriesen.  Valerie  E.  Wyatt,  and  Michael  D.  Cadman  132 

SINGING  IN  A MATED  FEMALE  WILSON’S  WARBLER  

William  M.  Gilbert  and  Adele  F.  Carroll  134 

LAYING  TIME  OF  THE  BRONZED  COWBIRD  Brian  D.  Peer  and  Spencer  G.  Sealy  137 

TEMPORAL  DIFFERENCES  IN  POINT  COUNTS  OF  BOTTOMLAND  FOREST  LANDBIRDS 

Winston  Paul  .Smith  and  Daniel  J.  Twedt  1 39 

144 


ORNITHOLOGICAL  LITERATURE 


JhcWilsonBulktin 

PUBLISHED  BY  THE  WILSON  ORNITHOLOGICAL  SOCIETY 


VOL.  Ill,  NO.  2 JUNE  1999  PAGES  157-302 

(ISSN  (XM3-5643) 

r 


c r. 


■ * ^ 


i 0 


THE  WILSON  ORNITHOLOGICAL  SOCIETY 
FOUNDED  DECEMBER  3,  1888 

Named  after  ALEXANDER  WILSON,  the  first  American  Ornithologist. 

President  Edward  H.  Burtt,  Jr.,  Department  of  Biology,  Ohio  Wesleyan  University,  Delaware,  Ohio 
43015;  E-mail:  EHBurtt@cc.owu.edu. 

First  Vice-President — John  C.  Kricher,  Biology  Department,  Wheaton  College,  Norton,  Massachusetts 
02766;  E-mail:  JKricher@wheatonma.edu. 

Second  Vice-President — William  E.  Davis,  Jr.,  College  of  General  Studies,  871  Commonwealth  Ave., 
Boston  University,  Boston,  Massachusetts  02215;  E-mail:  WEDavis@bu.edu. 

Editor  Robert  C.  Beason,  Department  of  Biology,  State  University  of  New  York,  1 College  Circle, 
Geneseo,  New  York  14454;  E-mail:  WilsonBull@geneseo.edu. 

Secretary — John  A.  Smallwood,  Department  of  Biology,  Montclair  State  University,  Upper  Montclair, 
New  Jersey  07043;  E-mail:  Smallwood@saturn.montclair.edu. 

Treasurer — Doris  J.  Watt,  Department  of  Biology,  Saint  Mary’s  College,  Notre  Dame,  Indiana  46556; 
E-mail:  DWatt@saintmarys.edu. 

Elected  Council  Members — Peter  C.  Frederick  and  Danny  J.  Ingold  (terms  expire  1999),  Charles  R.  Blem 
and  Sara  R.  Morris  (terms  expire  2000),  Jonathan  L.  Atwood  and  James  L.  Ingold  (terms  expire 
2001). 

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  Bulletin  is  sent  to  all  members  not  in  airears  for  dues. 

THE  JOSSELYN  VAN  TYNE  MEMORIAL  LIBRARY 
The  Josselyn  Van  Tyne  Memorial  Library  of  the  Wilson  Ornithological  Society,  housed  in  the  University 
of  Michigan  Museum  of  Zoology,  was  established  in  concurrence  with  the  University  of  Michigan  in 
1930.  Until  1947  the  Library  was  maintained  entirely  by  gifts  and  bequests  of  books,  reprints,  and  orni- 
thological 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  administered  by  the  Library  Committee.  William  A.  Lunk,  University  of 
Michigan  is  Chairman  of  the  Committee.  The  Library  currently  receives  over  200  periodicals  as  gifts  and 
in  exchange  for  The  Wilson  Bulletin.  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  University 
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 
University  of  Michigan,  1 109  Geddes  Ave.,  Ann  Arbor,  MI  48109-1079  USA.  Contributions  to  the  New 
Book  Fund  should  be  sent  to  the  Treasurer. 


THE  WILSON  BULLETIN 
(ISSN  0043-5643) 

THE  WILSON  BULLETIN  (ISSN  0043-5643)  is  published  quarterly  in  March.  June,  September,  and  December  by 
the  Wilson  Ornithological  Society,  810  East  lOth  Street,  Lawrence.  KS  66044-8897.  The  subscription  price,  both  in  the 
United  States  and  elsewhere,  is  .S40.00  per  year.  Periodicals  postage  paid  at  Lawrence.  KS.  POSTM.^STER:  Send  address 
changes  to  THE  WILSON  BULLETIN,  PO.  Box  1897,  Lawrence.  KS  66044-8897. 

Back  issues  or  single  copies  are  available  for  $12.00  each.  Most  back  issues  of  the  Bulletin  are  available  and  may  be 
ordered  from  the  Treasurer.  Special  prices  will  be  quoted  for  quantity  orders. 

All  articles  and  communications  for  publications,  books  and  publications  for  reviews  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  the  OSNA.  P.O.  Box  1897, 
Lawrence,  KS  66044-8897.  Phone:  (785)  843-1221;  FAX:  (785)  843-1274. 

© Copyright  1999  by  the  Wilson  Ornithological  Society 
Printed  by  Allen  Press,  Inc.,  Lawrence.  Kansas  66044,  U.S.A. 


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


1. 


FRONTISPIECE.  Left  from  top  to  bottom;  Myrmollienila  luiemutoiioiu  haematonota  male,  Mvrmoilienila 
liaeimitoiiota  pyrrhonoUi  male,  Mynuothenila  haematonota  pyrrhonota  Female,  Myrmothenila  spodionota  spo- 
dionota  male,  Myrmothenila  spodionota  spodionota  female.  Right  from  top  to  bottom:  Mvrmotherula  fjeldsaai 
male  type,  Myrmothenila  fjeldsaai  female,  Myrmothenila  lencophthalma  lencophthalnia  male,  Mvnnotherida 
leiieophthalma  lencophthalma  female.  Water  color  painting  by  J.  Fjeldsa. 


THE  WILSON  BULLETIN 

A QUARTERLY  JOURNAL  OF  ORNITHOLOGY 
Published  by  the  Wilson  Ornithological  Society 

VOL.  Ill,  NO.  2 JUNE  1999  PAGES  157-302 


Wilson  Bull.,  111(2),  1999,  pp.  157-165 


A NEW  SPECIES  IN  THE  MYRMOTHERULA  HAEMATONOTA 
SUPERSPECIES  (AVES;  THAMNOPHILIDAE)  FROM  THE 
WESTERN  AMAZONIAN  LOWLANDS  OF  ECUADOR  AND  PERU 

NIELS  KRABBE,'  MORTON  L.  ISLER,^  PHYLLIS  R.  ISLER,^ 

BRET  M.  WHITNEY, 3 JOSE  ALVAREZ  A.,^  AND  PAUL  J.  GREENFIELD^ 


ABSTRACT. — A new  species  of  antwren  (Myrmotherula  fjeldsaai)  closely  related  to  Myrmotherula  haema- 
tonota  is  described  from  the  lower  tropical  zone  of  eastern  Ecuador  and  immediately  adjacent  Peru.  It  primarily 
differs  from  M.  h.  haematonota  by  its  brown  instead  of  red  back  in  both  sexes.  New  distributional  data  for 
nominate  M.  h.  haematonota  shows  that  it  meets  the  new  species  north  of  the  Ri'o  Marafion,  between  the  Rios 
Napo  and  Pastaza,  with  no  apparently  significant  physical  barrier  between  them.  Received  6 March  1998,  ac- 
cepted 30  Dec.  1998. 


Two  species  of  “stipple-throated”  ant- 
wrens,  Ornate  Antwren  {Myrmotherula  orna- 
ta)  and  White-eyed  Antwren  {Myrmotherula 
leucophthalma)  show  distinct  geographical 
variation  in  back  color.  In  some  populations 
of  each  species  the  back  is  rufous;  in  others  it 
is  gray  or  olive-brown.  Separated  by  large  riv- 
ers, these  populations  are  not  in  physical  con- 
tact and  have  traditionally  been  ranked  as  sub- 
species. Here  we  describe  similar  variation  in 
Myrmotherula  haematonota,  but,  because  the 
two  forms  are  known  to  be  in  contact  and  to 


' Zoological  Museum,  Univ.  of  Copenhagen,  Uni- 
versitetsparken  15,  DK-2100,  Denmark. 

- Division  of  Birds,  National  Museum  of  Natural 
History,  Smithsonian  Institution,  Washington,  D.C. 
20560. 

^ Museum  of  Natural  Science,  1 19  Foster  Hall,  Lou- 
isiana State  Univ.,  Baton  Rouge,  Louisiana  70803. 

^ Instituto  de  Investigaciones  de  la  Amazom'a  Peru- 
ana-IIAP,  Avenida  A.  Quinones  km  2.5,  Apartado  784, 
Iquitos,  Peru. 

161-162  Calle  6,  El  Bosque,  Quito,  Ecuador. 

® Present  address:  Cas.  17-21-791,  Quito,  Ecuador; 
E-mail;  NKrabbe@pi.pro.ec 
’ Corresponding  author. 


retain  their  integrity,  we  propose  to  rank  them 
as  species. 

Zimmer  (1932)  defined  Myntiotherula  hae- 
matonota to  encompass  both  the  rufous 
backed  forms  that  occupy  Amazonian  low- 
lands and  the  gray  backed  forms  of  Andean 
foothills.  He  believed  two  specimens  from  the 
lowlands  of  Loreto,  Peru,  to  be  intermediate 
between  these  forms.  More  recently,  Hilty  and 
Brown  (1986)  and  Parker  and  Remsen  (1987) 
considered  the  foothill  forms  as  a distinct  spe- 
cies, Myrmotherula  spodionota  (including  so- 
roria),  but  did  not  address  the  issue  of  the 
apparently  intermediate  specimens. 

METHODS 

In  1992  PJG  noticed  a male  specimen  with  a brown 
back  in  the  collection  of  the  Museo  Ecuatoriano  de 
Ciencias  Naturales  (MECN),  labeled  as  M.  h.  haema- 
tonota, taken  at  Rfo  Bufeo  in  the  lowlands  of  Pastaza 
by  R.  Olalla  on  3 February  1963.  This  specimen  was 
referred  to  Myrmotherula  leucophthalma  by  Ortiz- 
Crespo  and  coworkers  ( 1990)  and  by  Ridgely  and  Tu- 
dor (1994). 

In  1994  NK  tape  recorded  and  collected  a male  and 
a female  “stipple-throated”  antwren  with  brown  backs 
near  Pompeya,  Napo,  Ecuador.  These  specimens  and 
the  Rfo  Bufeo  male  were  compared  directly  with  the 


157 


158 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


male  from  the  “Mouth  of  Rfo  Curaray”  at  the  Amer- 
ican Museum  of  Natural  History  (AMNH).  The  male 
is  one  of  two  specimens  Zimmer  (1932)  identified  as 
intermediate  between  the  lowland  and  foothill  forms. 
Six  additional  specimens  of  the  brown  backed  form 
were  later  collected  or  located  in  the  Western  Foun- 
dation of  Vertebrate  Zoology  (WFVZ),  Louisiana  State 
University,  Museum  of  Zoology  (LSUMZ),  Museum 
of  Natural  History,  University  of  Kansas  (MNHUK), 
and  Museo  de  Historia  Natural  de  San  Marcos 
(MUSM).  Four  records  were  available  only  as  photo- 
graphs. Thus,  12  specimens  (4  of  them  photographic) 
were  compared:  7 males  and  5 females.  All  males  were 
compared  directly  with  the  Rfo  Curaray  specimen. 
Two  females  were  compared  with  each  other  by  NK, 
three  by  MLI.  Both  red  backed  and  brown  backed 
forms  were  collected  along  Rfo  Tigre,  Loreto,  Peru  in 
1995.  An  apparently  intermediate  specimen  (not  lo- 
cated) from  an  unspecified  locality  along  Rfo  Tigre 
was  described  as  having  the  back  “dark  olive-brown, 
some  of  the  feathers  of  the  middle  of  the  upper  back 
with  rufous  brown  edges”  (Hellmayr  1910),  suggest- 
ing some  gene  flow  between  the  two  forms.  This  was 
the  second  specimen  that  Zimmer  (1932),  without  ex- 
amining it,  had  considered  intermediate  between  M.  h. 
haematonotci  and  M.  h.  spodionota. 

RESULTS 

Specimens  of  each  sex  of  the  brown  backed 
birds  were  found  to  be  essentially  identical, 
suggesting  a homogenous  population  rather 
than  birds  variously  intermediate  between  M. 
haematonota  and  M.  spodionota.  The  grayish 
edges  of  the  inner  webs  of  the  remiges  in  the 
males  (reddish  in  M.  leucophthalma)  and  the 
coloration  of  the  underparts  of  the  female  in- 
dicate that  the  brown  backed  specimens  are 
more  closely  related  to  M.  haematonota  and 
M.  spodionota  than  to  M.  leucophthalma.  An 
analysis  of  vocalizations  (unpubl.  data)  con- 
firms this  relationship.  Nominate  M.  h.  hae- 
matonota was  found  to  be  parapatric  with 
brown  backed  birds  in  seemingly  uniform 
habitat  (see  Fig.  1).  Brown  backed  birds  are 
uniform  in  plumage  over  a large  area  and  thus 
clearly  represent  a valid  taxon.  They  are  re- 
placed sharply  by  red-backed  birds  in  similar 
habitat,  suggesting  species  rank  of  the  new 
taxon,  which  we  propose  to  name: 

Brown-backed  Antwren 
Myrmotherula  fjeldsaai,  new  species 

Holotype. — MECN  6924,  adult  male  ob- 
tained by  N.  Krabbe  16  July  1994  near  Rio 
Tiputini,  37  road  km  south-southwest  of  Pom- 
peya,  Provincia  de  Napo,  Ecuador;  0°  38'  S 


76°  26'  W,  altitude  275  m.  Blood  sample 
(NK  14- 16.7.94)  deposited  at  Zoological  Mu- 
seum, University  of  Copenhagen.  Vocalization 
recordings  (LNS  65998)  archived  at  the  Li- 
brary of  Natural  Sounds,  Cornell  Laboratory 
of  Ornithology. 

Diagnosis. — Capitalized  names  and  num- 
bers of  colors  follow  Munsell  Soil  Color  Chart 
(Kollmorgen  Instruments  Corp.,  1994  edi- 
tion). Size,  shape,  plumage  pattern,  and  col- 
oration similar  to  those  of  M.  haematonota 
haematonota,  except  that  the  mantle  and  back 
are  between  Dark  Yellowish  Brown  (10YR3/ 
4)  and  Dark  Brown  (10YR3/3).  In  M.  h.  hae- 
matonota the  back  is  Dark  Red  (varying  be- 
tween 2.5YR4/8  and  2.5YR3/6).  The  male  of 
M.  fjeldsaai  differs  from  M.  h.  pyrrhonota  by 
having  brown  instead  of  red  back,  paler  flanks 
and  tail,  and  larger  and  pale  buff  instead  of 
pure  white  spots  on  tips  of  median  and  some 
lesser  coverts.  Female  differs  from  M.  h.  pyr- 
rhonota by  having  a red  back,  pale  (mostly 
white)  throat  streaked  with  black,  and  buffy 
brown  breast  and  belly;  in  M.  h.  pyrrhonota 
the  throat  is  yellow  ochre  and  usually  un- 
streaked, and  the  breast  is  reddish  brown  (Ta- 
ble 1).  Male  differs  from  M.  spodionota  by 
having  a brown  instead  of  pure  gray  back; 
larger,  more  buffy,  and  rounded  wing  covert 
spots,  distinctly  lighter  gray  underparts  with 
darker  olive  brown  sides  and  flanks,  and  by 
not  showing  the  tendency  in  many  individuals 
of  M.  spodionota  to  have  the  white  streaks  of 
the  throat  that  continue  onto  the  breast  and 
sometimes  even  the  belly.  Female  differs  from 
M.  spodionota  by  having  a red  back  and  being 
considerably  paler  throughout  with  whitish, 
black-streaked  throat  and  virtually  uniform 
buffy  brown  breast  and  belly.  Myrmotherula 
spodionota  has  a somewhat  flammulated  yel- 
low ochre  throat,  breast  and  belly  (with  throat 
lightly  marked;  Table  1).  Females  differ  from 
the  two  known  females  of  M.  h.  haematonota 
from  north  of  the  Rfo  Maranon  by  throat  color 
(Table  1),  but  this  difference  falls  within  the 
variation  seen  in  large  samples  of  M.  h.  hae- 
matonota and  M.  h.  amazonica.  Myrmotherula 
fjeldsaai  differs  from  brown  backed  forms  of 
M.  leucophthalma  by  having  darker  general 
coloration  and  smaller,  paler,  and  decidedly 
more  rounded  wing  spots;  male  has  grayish  as 
opposed  to  buffy  inner  webs  of  remiges,  and 
usually  has  smaller  throat  spots;  female  has 


Krcibhe  el  cil.  • NEW  ANTWREN  FROM  WESTERN  AMAZONIA 


159 


FIG.  1 . Distribution  of  taxa  of  the  Myrmotherula  [haematonota]  superspecies  in  the  eastern  Ecuador-northern 
Peru  region.  Heavy  lines  = coast  and  national  boundaries.  Dotted  line  = continental  divide.  Black  stars  = 
Myrmotherula  fjeldsaai.  Open  circles  = M.  spodionota  (including  sororia).  Open  squares  = M.  h.  haematonota. 
Open  triangles  = M.  h.  pyrrhonota.  Circle  surrounding  a “U”  = species  unknown.  Identification  of  locations 
discussed  in  text  and  type  localities  (identification  of  other  locations  on  map  available  from  MLI):  1.  Rio  Tiputini 
(0°  38'  S,  76°  26'  W),  Napo;  type  locality  of  M.  fjeldsaai.  2.  “Sunka  1”  (0°  42'  S,  75°  51'  W),  Napo.  3.  Tzapino 
(ca  01°  1 r S,  77°  44'  W),  Pastaza.  4.  Ri'o  Bufeo  (ca  02°  12'  S,  76°  48'  W),  Pastaza.  5.  Teniente  Lopez  (ca  02°  32' 
S,  76°  14'  W),  Loreto.  6.  San  Jacinto  (ca  02°  21'  S,  75°  43'  W),  Loreto.  7.  Mouth  of  Rio  Curaray  (ca  02°  24'  S, 
74°  04'  W,  exact  location  of  collecting  station  uncertain),  Loreto.  8.  Cocha  Hildalgo,  left  bank  Rio  Tigre  above 
mouth  of  Rio  Pacacuro,  Loreto.  9.  Vicinity  of  Intuto,  right  bank  Rio  Tigre  (03°  16'  S,  75°  04'  W),  Loreto.  10. 
Rio  (Quebrada)  Pavayacu,  Loreto.  11.  Andoas,  Loreto.  12.  Ri'o  Pacacuro,  left  bank  near  mouth,  Loreto.  13. 
Santa  Andrea,  left  bank  Ri'o  Tigre  approximately  half  way  between  mouth  of  Ri'o  Pacacuro  and  Intuto.  14.  Nuevo 
Manchuria  (03°  50'  S,  74°  19'  W)  and  Nuevo  Tarma  (03°  48'  S,  74°  21'  W),  left  bank  Ri'o  Tigre,  Loreto.  15. 
Puerto  Indiana,  Loreto.  16.  Libertad,  Loreto.  17.  Chamicuros,  Loreto;  type  locality  of  M.  h.  haematonota.  18. 
Huampami  (200  m),  Amazonas.  19.  Ri'o  Kagka  (800  m),  Amazonas.  20.  Sarayacu,  Pastaza;  type  locality  of  M. 
spodionota.  21.  Tigre  Playa,  Sucumbi'os  (MECN  6750  and  6751). 


mottled  cheeks,  whitish-streaked  throat  and 
huffy  brown  breast,  as  opposed  to  uniform 
bright  buffy  yellow  cheeks,  throat  and  breast 
in  M.  leucophthalma. 

Description  of  holotype.— Above,  including 
most  of  crown,  between  Dark  Yellowish 
Brown  (10YR3/4)  and  Dark  Brown  (10YR3/ 
3),  edge  of  tail  and  1-2  mm  wide  tips  of  feath- 
ers of  back  a more  reddish  Dark  Brown 
(7.5YR3/3).  Wing  coverts  brownish  black 
with  pale  tips,  forming  three  distinct  rows  of 


pale  Reddish  Yellow  (7.5YR8/6)  spots.  Inner 
webs  and  basal  half  of  outer  web  of  inner 
greater  secondary  coverts  Dark  Brown 
(10YR3/3)  to  Dark  Yellowish  Brown  (10YR3/ 
4),  inner  webs  washed  with  Dark  Brown 
(7.5YR3/3)  on  their  tips.  Primary  coverts 
blackish  brown  with  minute,  barely  discern- 
ible, Reddish  Yellow  (7.5YR6/8)  tips.  Pale 
tips  of  the  two  alula  feathers  as  large  as  on 
secondary  coverts,  but  lighter,  whitish  on  out- 
er web.  Median  coverts  and  the  largest  of  the 


160 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


TABLE  1.  Diagnosis  of  plumage  differences  of  females  of  Myrmotherula  fjeldsaai 
lations  in  the  M.  [haematonota]  superspecies. 

and  neighboring  popu- 

M.  spodionota 
« = >10 

M.  fjeldsaai 
ii  = 5 

M.  h.  haemainnola  from 
north  of  Rio  Maranon; 
n = 2 

M.  h.  pyrrhonota  from 
north  bank  of  Rio  Napo; 

« = 2 

Chin  & throat 

Buffy  Yellow  to 
Yellow  Ochre 

10  YR8/6  to 
10YR6/8 

Mostly  white; 
sides  faintly 
tinged  yellow 
brown 

Pale  Buff 

10YR8/2 

Buffy  Yellow 
10YR8/6  to  Yel- 
low Ochre  be- 
tween 10YR7/6 
and  7.5YR7/6 

Streaks  on  throat 

Streaks  are  mini- 
mal, but  black- 
ish feather  bases 
show  through 

Moderate  to  heavy 
blackish  streaks 
over  blackish 
feather  bases 

Moderate  to  heavy 
blackish  streaks 
over  blackish 
feather  bases 

Usually  no  streaks, 
sometimes  a 
few  fine  streaks, 
but  blackish 
feather  bases 
not  apparent 

Center  of  breast  to 

Yellow  Ochre 

Buffy  Brown  to 

Buffy  Brown  to 

Reddish-yellow 

upper  center  of 

10YR7/6  to 

Light  Yellowish 

Light  Yellowish 

7.5YR6/8,  suf- 

belly 

10YR6/8  spots; 
lateral  feather 
edges  like  sides 
giving  a spotty 
appearance 

Brown  10YR7/4 
to  10YR6/4  suf- 
fused with  Light 
Olive  Brown 
2.5Y5/3 

Brown  10YR7/4 
to  10YR6/4  suf- 
fused with  Light 
Olive  Brown 
2.5Y5/3 

fused  with  Light 
Olive  Brown 
2.5Y5/3 

Crown  & anterior 

Very  Dark  Grayish 

Dark  Yellowish 

Olive  Brown 

Olive  Brown 

mantle 

Brown  2.5Y3/2 

Brown  between 
10YR4/4  & 
10YR4/6 

2.5Y4/3 

2.5Y4/3 

Posterior  mantle  to 
tail  coverts 

Same  as  crown 

Same  as  crown 

Reddish  Brown 
2.5YR3/6  to  4/6 
to  Dark  Reddish 
Brown 

2.5YR4/8 

Reddish  Brown 
2.5YR3/6  to  4/6 
to  Dark  Red- 
dish Brown 
2.5YR4/8 

lesser  coverts  blackish  brown,  spots  palest  on 
the  distal  coverts.  The  smallest  of  the  lesser 
wing  coverts  mainly  Bluish  Gray  (5PB5/1), 
each  with  a small,  black  bordered  whitish  dot 
at  the  tip.  Forehead,  cheeks,  breast,  and  belly 
Bluish  Gray  (5PB5/1,  but  slightly  lighter), 
flanks  and  vent  between  Olive  Brown  (2.5Y4/ 
4)  and  Light  Olive  Brown  (2.5Y5/4).  Throat 
black,  feathers  with  pale  shafts  and  white  tips, 
forming  white,  triangular  spots  pointing  an- 
teriorly. Inner  webs  of  remiges  gray.  Twelve 
rectrices,  tail  strongly  graduated,  with  tips  of 
the  three  outer  rectrices  16,  8 and  4.5  mm 
from  the  tips  of  the  central  pair  of  rectrices. 
Body  mass  9.8  g.  hides  grayish  brown;  bill 
blackish  with  thin  gray  blue  line  along  cutting 
edge;  feet  gray  blue.  Skull  100%  ossified.  No 
Bursa  Fabricii  found.  Largest  testis  3 X 1 mm. 
No  fat.  Stomach  contents:  small  arthropods 
(saved).  Netted  in  tongue  of  flooded  forest  in 
hill  country. 

Variation  in  males. — Specimens  from  the 


Rio  Bufeo,  the  mouth  of  Rio  Curaray,  and  one 
hand  held  bird  photographed  near  the  type  lo- 
cality in  1994  (C.  Canaday  photo)  were  ex- 
amined by  NK  and  found  to  be  similar  to  the 
type.  The  two  old  specimens  were  slightly 
lighter  gray,  equivalent  to  the  difference  seen 
between  old  and  fresh  material  of  M.  spo- 
dionota.  MLI  made  the  same  conclusions 
when  he  compared  the  Rio  Curaray  specimen 
to  the  Teniente  Lopez  specimen  and  to  the  de- 
scription of  the  Rio  Tigre  specimen  collected 
by  JA.  The  backs  of  the  males  were  slightly 
brighter  (Dark  Yellowish  Brown,  10YR3/4) 
than  the  description  of  the  type  specimen; 
backs  of  two  specimens  were  more  Olive 
Brown  (between  10YR3/4  and  2.5Y4/4);  and 
all  lacked  the  reddish  brown  feather  tips  de- 
scribed for  the  type  specimen. 

Additional  male  specimens  examined. — 
Myrmotherula  fjeldsaai:  ECUADOR:  Napo, 
Rio  Bufeo  (MECN  2181).  PERU:  Loreto, 
mouth  of  Rio  Curaray  (AMNH  255780);  Ten- 


Krahhe  el  al.  • NEW  ANTWREN  EROM  WESTERN  AMAZONIA 


161 


iente  Lopez,  Ri'o  Conientes  (MNHUK  uncata- 
logued, skull  incompletely  ossified,  collected 
July  1993  by  Aucca);  above  Intuto,  Rio  Tigie 
(collected  Januaiy  1995  by  JA,  to  be  depos- 
ited in  MUSM).  Mynnothenda  h.  haematon- 
ota  and  M.  h.  amazonica  (Museum  abbrevia- 
tions aie  followed  by  number  of  specimens): 
PERU:  Loreto,  Lores,  Rio  Tigre  (collected 
January  1995  by  JA,  to  be  deposited  in 
MUSM,  2);  Libeitad,  S bank  Napo  (LSUMZ, 
2);  Puerto  Indiana,  N bank  Amazon  (AMNH, 
1);  Quebrada  Vainilla,  S bank  Amazon 
(LSUMZ,  5);  Orosa  (AMNH,  3);  Sai’ayacu 
(AMNH,  2);  15  km  E Puerto  Maldonado 
(MNHUK,  1).  BOLIVIA:  (LSUMZ,  5).  BRA- 
ZIL: (AMNH,  10).  Mynnothenda  h.  pyrrhon- 
ota:  ECUADOR:  Sucumbios,  about  14  km  N 
Tigre  Playa  (MECN,  1).  PERU:  Loreto,  N of 
Rio  Napo,  157  km  by  river  NNE  of  Iquitos 
(LSUMZ,  2);  Rio  Yanayacu  (LSUMZ,  2); 
Quebrada  Oran  (LSUMZ,  2).  COLOMBIA: 
(AMNH,  1).  BRAZIL:  (AMNH,  8;  USNM, 
10).  VENEZUELA:  (AMNH,  26;  USNM,  6). 
Mynnothenda  spodionota  including  soroha: 
ECUADOR:  Napo,  San  Jose  Abajo  (AMNH, 
5;  USNM,  1);  above  San  Jose  (ANSP,  1);  Avi- 
la (ANSP,  1);  above  Avila  (AMNH,  4);  Chon- 
ta  Urcu  (ANSP,  1);  Morona-Santiago,  Cutucii 
(AMNH,  1).  PERU:  (AMNH,  8;  LSUMZ,  23). 

Female. — A topotypical  young  female 
(MECN  6925)  with  2%  ossified  skull  is  sim- 
ilar to  the  male  above,  but  slightly  lighter  and 
yellower  (between  10YR4/4  and  10YR4/6), 
with  tail  feathers  having  Strong  Brown 
(7.5YR4/6)  lateral  edges.  Sides  of  head  like 
crown,  but  with  ill-defined  buff  mottling.  Be- 
low Light  Yellowish  Brown  (between  2.5Y6/ 
4 and  10YR6/4),  somewhat  browner  on  sides 
and  flanks,  breast  with  faint  pale  flammula- 
tions.  Throat  white  tinged  Pale  Yellow 
(2.5Y8/3),  irregularly  but  conspicuously 
streaked  by  black  edges  (but  not  tips)  on  some 
feathers.  Irides  gray  brown,  bill  like  the  type, 
feet  slaty;  body  mass  9.0  g. 

Another  female  netted  and  released  at  the 
type  locality  in  1994  by  NK  was  similar,  but 
with  whitish  irides  and  with  nairower  black 
edges  on  the  feathers  of  the  whitish  throat; 
body  mass  10.7  g.  A third  individual  netted 
and  photographed  near  the  type  locality  in 
1994  (C.  Canaday  photo)  was  found  by  NK 
to  be  similar.  It  had  a whitish  throat  with  nar- 
row dark  feather  edges,  appeai'ed  to  have  a 


fully  ossified  skull,  brown  irides,  and  blue 
gray  feet;  body  mass  9.5  g.  The  foregoing  de- 
scription of  MECN  6925  was  found  by  MLI 
to  match  females  taken  at  “Sunka  1”,  Tza- 
pino,  and  San  Jacinto,  as  well  as  the  descrip- 
tion and  photographs  of  the  specimen  collect- 
ed upstream  from  Intuto  on  the  Rio  Tigre  by 
JA,  except  that  central  feathers  of  throats  were 
white  (untinged)  and  lateral  edges  of  tail 
feathers  were  the  same  as  the  color  of  the  back 
(rather  than  reddish  brown). 

Additional  female  specimens  examined. — 
Myrmothenda  fjeldsaai:  ECUADOR:  Napo, 
“Sunka  1”,  40  km  S Coca  (WFVZ  45662, 
collected  November  1988).  Pastaza,  Tzapino 
(LSUMZ  83109,  collected  in  May  1976  by 
Tallman).  PERU:  Loreto,  San  Jacinto,  upper 
Rio  Tigre  (MNHUK  uncatalogued;  collected 
in  July  1993  by  Aucca);  above  Intuto,  Rio  Ti- 
gre (taken  by  JA  in  January  1995  and  to  be 
deposited  at  MHNJP).  Mynnothenda  h.  hae- 
matonota  and  M.  h.  amazonica'.  PERU:  Lor- 
eto, Libertad,  S bank  Napo  (LSUMZ,  2); 
Puerto  Indiana,  N bank  Amazon  (AMNH,  1); 
the  vicinity  of  Huampami,  Amazonas,  Peru, 
about  200  m elevation  (LSUMZ,  1);  Quebrada 
Vainilla,  S bank  Amazon  (LSUMZ,  6);  Orosa 
(AMNH,  1);  Sarayacu  (AMNH,  1);  “mouth  of 
Rio  Umbamba”  (AMNH,  1).  Madre  del  Dios, 
15  km  E Puerto  Maldonado  (MNHUK,  1). 
BOLIVIA:  (LSUMZ,  5).  BRAZIL:  (AMNH, 
7).  Myrmotherula  h.  pyrrhonota:  ECUADOR: 
Sucumbios,  about  14  km  N Tigre  Playa 
(MECN,  1).  PERU:  Loreto,  Quebrada  Oran 
(LSUMZ,  1).  COLOMBIA:  (AMNH,  1). 
BRAZIL:  (AMNH,  16;  USNM,  7).  VENE- 
ZUELA: (AMNH,  26;  USNM,  5).  Mynnothe- 
nda spodionota  including  soioria:  ECUA- 
DOR: Napo,  San  Jose  Abajo  (AMNH,  2; 
USNM,  1);  Morona-Santiago:  Chiguaza.  Cu- 
tucu  (ANSP,  1);  Zamora-Chinchipe:  Zamora 
(AMNH,  1).  PERU:  (AMNH,  4;  LSUMZ,  20). 

Mensural  variation. — Measurements  are 
given  in  Table  2.  As  a whole,  measurements 
of  populations  in  M.  haematonota  and  M.  spo- 
dionota do  not  differ  significantly  (Table  2). 
The  only  measurement  whose  ranges  do  not 
overlap  are  the  wing  measurements  for  the 
few  known  specimens  of  M.  h.  haematonota 
from  north  of  the  Amazon  and  M.  spodionota, 
although  there  is  only  slight  overlap  of  some 
wing  and  tail  measurements  of  M.  spodionota 
compared  to  other  populations.  Using  a rule 


162 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


TABLE  2.  Measurements  of  six  populations  in  the  Myrmotherula  [haematonota]  superspecies.  Sexes  are 
combined.  Bill  is  measured  from  the  anterior  nostril;  wing  is  unflattened  wing  chord. 

Measurement 

M.  fjeldsaai 

M.  h.  haematonota 
north  of  Amazon 

M.  h.  haematonota 
south  of  Amazon 

M.  h.  pyrrhonota 
Venezuela  & 
northern  Brazil 

M.  spodionota 

Bill  Width 

n = 7 

n = 6 

n = S 

n = 10 

n = \0 

Range 

3.6-4.0 

3. 2-4.0 

3.5-4. 1 

3. 7-4. 2 

3.6-3.9 

Mean  ± SD 

3.79  ± 0.14 

3.58  ± 0.34 

3.86  ± 0.18 

4.05  ±0.18 

3.75  ± 0.13 

Bill  Depth 

n — 1 

n = 6 

n = 8 

n = 10 

n = 10 

Range 

3. 6-3. 9 

3.6-4.2 

3. 8-4.4 

3. 6-4.4 

3.8-4.3 

Mean  ± SD 

3.72  ±0.15 

3.83  ± 0.22 

4.04  ± 0.21 

4.03  ± 0.23 

4.09  ±0.19 

Bill  Length 

n = 10 

n = 1 

n = 10 

n = 10 

n = 10 

Range 

8.4-9. 5 

8. 3-8.9 

8.4-9.6 

8. 8-9. 7 

8. 2-9.4 

Mean  ± SD 

8.97  ± 0.32 

8.57  ± 0.30 

9.07  ± 0.36 

9.19  ± 0.31 

8.81  ± 0.43 

Tarsus  Length 

n = 8 

= 6 

n = 8 

n = 10 

n = 10 

Range 

15-16 

16-17 

15-17 

16-17 

16-18 

Mean  ± SD 

15.7  ± 0.5 

16.2  ± 0.4 

15.9  ± 0.6 

16.1  ± 0.3 

16.8  ± 0.6 

Tail  Length 

n = 10 

n = 1 

n = 10 

n = 10 

/!  = 10 

Range 

33-39 

34-37 

34-39 

34-38 

37-40 

Mean  ± SD 

35.8  ± 1.5 

35.3  ± 1.1 

35.1  ± 1.8 

35.9  ± 1.3 

38.4  ± 1.2 

Wing  Chord 

n = 10 

n = 1 

n = 10 

/?  = 10 

n = 10 

Range 

47-54 

47-50 

48-53 

47-51 

51-55 

Mean  ± SD 

49.8  ± 1.9 

48.1  ± 1.1 

50.3  ± 1.6 

49.2  ± 1.4 

52.6  ± 1.4 

similar  to  one  proposed  by  Amadon  (1949), 
comparing  means  of  measurements  plus  or 
minus  three  standard  deviations,  differences 
between  wing  and  tail  measurements  of  M. 
spodionota  compared  with  M.  h.  haematonota 
north  of  the  Amazon  again  stand  out.  Another, 
larger  set  of  measurements  (mean  followed  by 
range)  confirms  this  tendency:  the  wing  (53.8 
mm;  51-56)  and  tail  (38.9  mm;  33-43)  mea- 
surements of  M.  spodionota  (n  = 29)  com- 
pares to  wing  (51.0  mm;  46-54)  and  tail 
(36.4;  33-41)  for  all  populations  of  M.  h.  pyr- 
rhonota  (n  = 86)  and  wing  (52.4  mm;  50-56) 
and  tail  (38.9  mm;  33-43)  for  all  populations 
of  M.  h.  haematonota  (n  = 40).  Differences 
in  bill  among  these  larger  samples  were  found 
to  be  insignificant. 

Distribution. — South  and  west  of  the  Rfo 
Napo  in  the  lowlands  of  eastern  Ecuador  and 
extreme  northern  Peru.  Geographic  range  ex- 
tends from  the  type  locality  and  from  nearby 
“Sunka  1 ” in  Napo,  Ecuador,  south  to  the 
central  portion  of  the  Rio  Tigre,  a short  dis- 
tance upriver  from  Intuto,  Loreto,  Peru,  ap- 
proximately 360  km  south-south-east  of  the 
type  locality  (Fig.  1).  A sight  record  (Willis 
1988)  from  Andoas,  Loreto,  may  also  be  of 
this  species. 

Habitat. — At  the  type  locality  NK  observed 
M.  fjeld.saai  in  tangled,  but  fairly  open  under- 


story to  lower  canopy  in  humid  primary  terra 
firma  forest  and  into  adjacent  tongues  of  var- 
zea  forest.  Neai'  Sachacocha  in  the  middle  Rio 
Tigre,  Loreto,  Peru,  BMW  and  JA  found  M. 
fjeldsaai  in  tall,  closed  canopy  (except  for 
scattered  treefalls),  teira  firme  forest  with  an 
abundance  of  palms  in  the  understory.  Known 
from  about  150  to  300  m elevation. 

Vocalizations. — The  loudsongs  (Isler  et  al. 
1998)  of  nominate  M.  haematonota,  M.  jjeld- 
saai,  and  M.  spodionota  (Fig.  2)  are  similar, 
but  easily  distinguishable  from  those  of  M. 
leucophthalma  (Isler,  Isler,  and  Whitney,  un- 
publ.  data).  We  do  not  provide  a detailed  anal- 
ysis of  differences  among  the  loudsongs  illus- 
trated in  Fig.  2 because  the  sample  size  for  M. 
fjeldsaai  is  inadequate.  However,  pace  (notes 
per  sec),  an  element  that  appears  to  vary 
among  the  loudsongs  of  M.  haematonota 
complex,  has  been  shown  to  be  one  of  the 
most  important  characteristics  distinguishing 
loudsongs  of  closely  related  and  syntopic 
pairs  of  Thamnophilidae  (Isler  et  al.  1998). 

Behavior. — Foraging  behavior  (NK,  BMW, 
JA,  C.  Canaday,  G.  Rivadaneira,  and  R.  S. 
Ridgely,  unpubl.  data),  of  M.  fjeldsaai  appears 
to  resemble  the  other  dead  leaf  specialists  in 
the  group  as  described  by  Gradwohl  and 
Greenberg  (1984),  Rosenberg  (1990,  1993, 
1997),  and  Whitney  (1994). 


Krahhe  et  ul.  • NEW  ANTWREN  FROM  WESTERN  AMAZONIA 


163 


A 

1 0 - 

B 

-<  -N  ^ A -X  ^ 



' ' ' ' \ \ 

1 1 1 

2 “ 

0 - 

1 1 1 

0 0.5  1.0  1.5  2.0  0 0.5  1.0  1.5  2.0 


N 

X 

U 

c 

3 

O' 

a 

u 

b 


C 

1 0 - 

8 - 

6 - 

4 - 

D 

iiiiiirTd:].  iti  ^ 

1 \ 1 

2 “ 

0 - 

1 1 1 

0 0.5  1.0  1.5  2.0  0 0.5  1.0  1.5  2.0 


FIG.  2.  Vocalizations  of  Myrmothenila  fjeldsoai  and  loudsongs  representative  of  neighboring  populations  of 
other  taxa  in  the  M.  [haematonota]  superspecies.  A.  Myrmotherula  fjeldsaai  loudsong  from  the  type  locality 
(recorded  by  N.  Krabbe;  LNS  65998).  B.  Myrmotherula  fjeldsaai  loudsong  from  Sachacocha,  Loreto,  Peru  (B. 
M.  Whitney;  to  be  archived  at  LNS).  C.  Myrmotherula  h.  pyrrhonota  loudsong  from  Quebrada  Sucusari,  Loreto, 
Peru  (T.  A.  Parker,  III;  LNS  33798).  D.  Myrmotherula  h.  haematonota  loudsong  from  Lores,  Loreto,  Peru  (B. 
M.  Whitney;  to  be  archived  at  LNS).  E.  Myrmotherula  s.  spodionota  loudsong  from  30  km  west  of  Loreto, 
Napo,  Ecuador  (B.  M.  Whitney;  to  be  archived  at  LNS).  E,  G.,  and  H.  Myrmotherula  fjeldsaai  calls  and  rattle 
from  the  type  locality  (N.  Krabbe;  LNS  65998) — compare  with  the  rattle  of  M.  spodionota  presented  by  Whitney 
(1994,  fig.  1). 


164 


THE  WILSON  BULLETIN  • VoL  HI.  No.  2,  June  1999 


Taxonomic  rank. — The  possibility  that  M. 
fjeldsaai  and  M.  haematonota  occasionally 
hybridize  is  raised  by  the  Rio  Tigre  specimen 
described  by  Hellmayr  (1910).  However,  the 
fact  that  the  two  maintain  their  integrity,  de- 
spite being  in  contact  in  an  area  with  exten- 
sive floodplain  dynamics,  makes  it  most  prob- 
able that  they  are  correctly  ranked  as  full  spe- 
cies. 

Speciation. — In  addition  to  being  parapatric 
with  M.  h.  haematonota,  M.  fjeldsaai  might 
meet  with  M.  h.  pyrrhonota  and  M.  spodiono- 
ta,  but  there  is  no  evidence  of  intergradation. 
To  the  southeast  (Fig.  1)  the  ranges  of  M. 
fjeldsaai  and  M.  h.  pyrrhonota  appear  to  be 
separated  by  the  middle  Rio  Napo,  but  both 
may  occur  farther  upstream  where  the  rivers 
Napo  and  Aguarico  are  narrow  enough  for 
them  to  cross.  To  the  west,  the  ranges  of  M. 
fjeldsaai  and  M.  spodionota  appear  to  be  sep- 
arated altitudinally,  with  a narrow  elevational 
band  that  neither  species  occupies.  Neither 
was  among  the  505  species  recorded  at  Jatun- 
sacha,  at  450  m during  26  months  of  field 
work  (B.  Bocham,  unpubl.  data).  Nor  has 
BMW  found  either  between  400-600  m dur- 
ing extensive  field  work  around  the  village  of 
Loreto,  Napo,  Ecuador.  Additionally,  neither 
was  among  the  six  species  of  Myrmotherula 
recorded  by  NK  during  1 2 days  of  field  work 
at  Canelos,  Pastaza,  500-700  m.  Because  the 
forest  is  continuous  between  their  elevational 
limits,  further  field  work  might  show  the  two 
species’  ranges  to  be  in  contact. 

Conservation. — The  known  range  of  Myr- 
motheriila  fjeldsaai  encompasses  the  Yasuni 
National  Park  (7281  km-)  in  Ecuador  (whence 
comes  the  type).  Given  that  this  huge  area  re- 
mains effectively  protected,  M.  fjeldsaai  ap- 
pears not  to  be  at  risk  at  present. 

Etymology. — We  take  the  pleasure  of  nam- 
ing this  species  in  honor  of  Prof.  Jon  Fjeldsa 
of  the  Zoological  Museum,  University  of  Co- 
penhagen. Through  his  countless  publications, 
most  based  on  results  obtained  during  field 
trips  to  the  most  hostile  of  environments,  he 
has  inspired  a large  number  of  biologists  to 
leave  their  desks  and  get  into  the  field.  Among 
his  achievements  should  also  be  mentioned 
his  most  recent  work  (with  C.  Rahbek)  in  de- 
limiting areas  of  top  priority  for  conservation 
in  South  America  and  Africa,  work  that  could 
eventually  save  a number  of  species  from  ex- 


tinction. Apart  from  his  impressive  profes- 
sional knowledge  and  self  discipline,  his  many 
achievements  were  possible  only  through  his 
great  understanding  of  human  nature,  his  gen- 
erosity and  helpfulness,  and  his  unfailing  habit 
of  treating  everybody  as  an  equal,  legendary 
from  the  jungles  of  South  America  to  Africa. 
The  English  name  refers  to  the  distinctive  col- 
or of  the  back. 

ACKNOWLEDGMENTS 

We  are  indebted  to  K.  V.  Rosenberg,  J.  M.  Bates, 
and  D.  Stotz  for  fruitful  discussions;  to  E Vuilleumier, 
AMNH,  for  helping  NK  with  fund  raising;  to  M.  Mo- 
reno, MECN,  M.  LeCroy,  AMNH,  and  D.  Agro,  ANSP 
for  help  in  the  use  of  collections  in  their  care;  to  G. 
Erisk,  Naturhistoriske  Riksmuseum,  Stockholm,  J.  V. 
Remsen,  LSUMZ,  R Colston,  BMNH,  M.  B.  Robbins 
and  R.  O.  Prum,  UKMNS,  and  C.  Samida,  WEVZ,  for 
the  loan  of  specimens  under  their  care;  to  G.  Budney, 
LNS,  for  assistance  and  use  of  recordings;  to  the  late 
T.  A.  Parker,  III,  who  recorded  much  of  the  material 
used  to  compare  vocalizations  of  M.  leucophthalma 
and  M.  haematonota-,  and  to  P.  Coopmans  and  J.  Row- 
lett who  supplied  additional  recordings  of  M.  spodion- 
ota. Krabbe’s  field  work  and  studies  in  the  USA  were 
generously  funded  by  ECUAAMBIENTE,  Quito;  Zoo- 
logical Museum,  University  of  Copenhagen;  and  a 
Chapman  grant  from  the  AMNH.  M.  Islet's  work  at 
AMNH  was  supported  by  a Chapman  collection  study 
grant.  J.  M.  Bates,  J.  EJeldsa,  G.  R.  Graves,  D.  Stotz, 
J.  P.  O'Neill,  and  an  anonymous  referee  kindly  com- 
mented on  early  drafts  of  the  manuscript. 

LITERATURE  CITED 

Amadon,  D.  1949.  The  seventy-five  per  cent  rule  for 
subspecies.  Condor  51:250-258. 

Cory,  C.  B.  and  C.  E.  Hellmayr.  1924.  Catalogue  of 
birds  of  the  Americas.  Part  III.  Eield  Mus.  Nat. 
Hist.  Zool.  Sen  13  (Publ.  223):  1-369. 

Gradwohl,  j.  and  R.  Greenberg.  1984.  Search  be- 
havior of  the  Checker-throated  Antwren  foraging 
in  aerial  leaf-litter.  Behav.  Ecol.  Sociobiol.  15: 
281-285. 

Hellmayr,  C.  E.  1910.  The  birds  of  the  Rio  Madeira. 
Novit.  Zool.  17:257-428. 

Hilty,  S.  L.  and  W.  L.  Brown.  1986.  A guide  to  the 
birds  of  Colombia.  Princeton  Univ.  Press,  Prince- 
ton, New  Jersey. 

ISLER.  M.  L.,  P.  R.  ISLER,  AND  B.  M.  Whitney.  1998. 
Use  of  vocalizations  to  establish  species  limits  in 
antbirds  (Passeriformes:  Thamnophilidae).  Auk 
I 15:577-590. 

Ortiz-Crespo,  E I.,  P.  J.  Greenfield,  and  J.  C.  Ma- 
THEUS.  1990.  Aves  del  Ecuador  Continente  y Ar- 
chipielago  de  Galdpagos.  FEPROTUR  and  CE- 
CIA,  Quito,  Ecuador. 

Parker,  T.  A.,  Ill  and  J.  V.  Remsen,  Jr.  1987.  Fifty- 


Krahhe  et  al.  • NEW  ANTWREN  FROM  WESTERN  AMAZONIA 


165 


two  Amazonian  bird  species  new  to  Bolivia.  Bull. 
Brit.  Orn.  Club  107:94-107. 

Ridgely,  R.  S.  and  G.  Tudor.  1994.  The  birds  of 
South  America.  Vol.  2.  The  suboscine  passerines. 
Univ.  of  Texas  Press,  Austin,  Texas. 

Rosenberg,  K.  V.  1990.  Dead-leaf  foraging  speciali- 
zation in  tropical  forest  birds.  Ph.D.  diss.,  Loui- 
siana State  Univ.,  Baton  Rouge. 

Rosenberg,  K.  V.  1993.  Diet  selection  in  Amazonian 
antwrens.  Auk  110:361-375. 

Rosenberg,  K.  V.  1997.  Ecology  of  dead-leaf  foraging 
specialists  and  their  contribution  to  Amazonian 
bird  diversity.  Ornithol.  Monogr.  48:673—700. 


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

Whitney,  B.  M.  1994.  Behavior,  vocalizations,  and 
possible  relationships  of  four  Myrmolherula  ant- 
wrens (Formicariidae)  from  eastern  Ecuador.  Auk 
1 1 1:469-475. 

Willis,  E.  O.  1988.  Behavioral  notes,  breeding  re- 
cords, and  range  extensions  for  Colombian  birds. 
Rev.  Acad.  Colombiana  Cienc.  Exactas  Fisicas 
Nat.  16:137-150. 

Zimmer,  J.  T.  1932.  Studies  of  Peruvian  birds.  III.  The 
genus  Myrmolherula  in  Peru,  with  notes  on  extra- 
limital  forms.  Part  1.  Am.  Mus.  Novit.  523:1-19. 


Wilson  Bull.,  111(2),  1999,  pp.  166-180 


COMPARATIVE  SPRING  HABITAT  AND  FOOD  USE  BY  TWO 

ARCTIC  NESTING  GEESE 

SUZANNE  CARRIERE,'  2 5 ROBERT  G.  BROMLEY,'^  AND  GILLES  GAUTHIER' 

ABSTRACT. — The  timing  of  egg  laying  is  generally  constrained  by  female  condition,  which  is  partly  deter- 
mined by  the  food  available  to  her  before  laying.  Although  it  was  generally  believed  that  geese  rely  exclusively 
on  internal  nutrient  reserves  for  egg  production,  spring  feeding  is  intensive  in  many  populations  of  geese, 
significantly  adding  nutrients  necessary  for  egg  production  and  incubation.  We  compared  the  spring  feeding 
ecology  of  Greater  White-fronted  Geese  {Anser  albifrons  frontalis)  and  Canada  Geese  {Brantci  canadensis  hutch- 
insii)  on  a shared  nesting  ground  on  the  Kent  Peninsula,  NWT  (68°  N,  108°  W),  where  pairs  feed  intensively 
from  arrival  until  incubation.  Live  plant  biomass  did  not  significantly  increase  within  specific  habitats  during 
preincubation,  but  the  total  available  biomass  was  greater  after  snow  melt  because  habitats  with  higher  biomass 
became  available.  Live  plant  biomass  available  in  pond  margins  (30—60  g/m-)  was  4-15  times  higher  than  in 
habitats  that  were  available  earlier,  i.e.,  mud-flats  and  hummocks  (4-8  g/m-).  Before  snow  melt,  both  species 
shared  the  1-20%  of  the  study  area  that  was  snow  free  (max.  density  600  pairs/km-),  opportunistically  used  the 
only  two  available  habitats,  mud-flats  and  hummocks,  and  primarily  ate  (50-70%)  tillers  of  Puccinellia  spp. 
During  snow  melt,  pairs  dispersed,  pair  density  decreased  (max.  of  40  pairs/km-),  and  interspecific  differences 
in  habitat  and  food  use  appeared.  White-fronted  Geese  used  pond  margins  and  ponds  more  often  than  Canada 
Geese.  After  snow  melt.  White-fronted  Geese  predominantly  fed  in  ponds  on  Carex  spp.  and  Dupontia  fisheri 
rhizomes  and  basal  stems;  Canada  Geese  continued  feeding  opportunistically,  pecking  leaves  in  all  habitats  and 
grubbing  rhizomes  in  pond  margins  and  ponds.  White-fronted  Geese  used  the  grubbing  technique  more  often 
than  Canada  Geese  in  all  habitats  and  periods.  Received  13  Feb.  1998.  accepted  17  Nov.  1998. 


Energy  investment  by  females  in  reproduc- 
tion is  highest  during  the  period  of  egg  for- 
mation in  birds  with  precocial  young,  such  as 
geese  (King  1973).  Clutch  size  and  timing  of 
laying  are  potentially  constrained  by  female 
condition,  which  is  partly  determined  by  the 
amount  of  energy  and  nutrients  available  to 
her  before  egg  formation  (Drent  and  Daan 
1980,  Winkler  and  Walters  1983). 

Early  nesting  is  critical  for  Arctic  nesting 
geese  because  of  the  short  summer  and  the 
rapid  seasonal  decline  in  components  of  re- 
productive success  such  as  gosling  growth  and 
probability  of  producing  recruits  (Barry  1962, 
Cooke  et  al.  1984,  Cooch  et  al.  1991,  Sedinger 
and  Flint  1991,  Lindholm  et  al.  1994).  One 


' Dept,  de  Biologic  et  Centre  d’Etudes  Nordiques, 
Univ.  Laval,  Quebec,  QC,  GIK  7P4,  Canada. 

’ Present  address:  Wildlife  and  Fisheries  Division, 
Dept,  of  Resources,  Wildlife,  and  Economic  Devel- 
opment, Government  of  the  Northwest  Territories,  600 
5102-50  Avenue,  Yellowknife,  NT  XI A 3,S8,  Canada. 

’ Wildlife  and  Fisheries  Division,  Dept,  of  Resourc- 
es, Wildlife,  and  Economic  Development,  Government 
of  the  Northwest  Territories,  600  5102-50  Avenue, 
Yellowknife,  NT  XI A 3S8,  Canada. 

Present  address:  Box  1 177,  Yellowknife,  NT  XI A 
2N8,  Canada. 

^ Corresponding  author; 

E-mail:  suzanne_carriere@gov.nt.ca 


Strategy  to  facilitate  early  nesting  is  to  carry 
nutrient  reserves  accumulated  during  migra- 
tion to  the  breeding  grounds  (Ankney  and 
Macinnes  1978,  Wypkema  and  Ankney  1979, 
Ankney  1984,  Budeau  et  al.  1991,  Bromley 
and  Jarvis  1993,  Choiniere  and  Gauthier 
1995).  There  is,  however,  an  upper  limit  to  the 
amount  of  reserves  that  can  be  economically 
carried  during  migration  (Lindstrom  and  Al- 
erstam  1992). 

In  migratory  birds,  the  timing  of  rapid  fol- 
licular development  (RED)  initiation  with  re- 
spect to  spring  migration  may  affect  the  rel- 
ative contribution  to  egg  production  of  nutri- 
ents acquired  en  route  versus  those  acquired 
on  nesting  grounds.  Timing  of  RED  initiation 
directly  determines  laying  date  of  the  first  egg 
(reviewed  by  Rohwer  1992).  Because  some 
female  geese  in  some  locations  typically  nest 
soon  (3-6  days)  after  their  arrival  on  the  nest- 
ing ground  {Anser  rossii:  Ryder  1970;  Anser 
caerulescens  caerulescens:  Ankney  1977,  An- 
kney and  Macinnes  1978;  Anser  canagica: 
Thompson  and  Raveling  1987;  Branta  berni- 
cla  bernicla:  Spaans  et  al.  1993),  RED  is  ini- 
tiated before  arrival,  hence  the  date  of  nest 
initiation  is  independent  of  food  availability 
on  the  nesting  ground  (Raveling  1978,  Ank- 
ney 1984,  Newton  1977;  but  see  Prop  and  de 
Vries  1993). 


166 


Carriere  el  al.  • SPRING  FEEDING  BY  ARCTIC  GEESE 


167 


In  other  populations  most  female  geese  lay 
at  least  12  days  after  an'ival  on  the  nesting 
grounds,  long  enough  for  the  completion  of 
RFD  of  the  first  egg  [Branta  bernicla  nigri- 
cans: Raveling  1978;  B.  canadensis:  Macln- 
nes  et  al.  1974,  Bromley  1984;  A.  albifrons: 
Fox  and  Madsen  1981,  Budeau  et  al.  1991;  A. 
caenilescens  caerulescens  (at  La  Perouse 
Bay):  Findlay  and  Cooke  1982;  A.  c.  atlantica: 
Gauthier  and  Tardif  1991].  In  this  case,  food 
availability  on  the  nesting  ground  can  affect 
the  date  of  RFD  initiation,  the  date  of  nest 
initiation,  and  potentially  clutch  size  and  con- 
stancy of  incubation.  Most  of  the  energy  nec- 
essary for  egg  formation  and  laying  was  met 
by  food  on  the  nesting  grounds  for  A.  albi- 
frons frontalis  (Budeau  et  al.  1991),  B.  can- 
adensis occidentalis  (Bromley  and  Jarvis 
1993),  and  A.  caerulescens  atlantica  (Choin- 
iere  and  Gauthier  1995).  Intermediate  cases 
may  occur  where  the  time  available  for  feed- 
ing between  arrival  on  the  nesting  ground  and 
laying  varies  greatly  among  females  and/or 
among  years  within  populations  (Raveling 
1978). 

Food  availability  in  the  Arctic  in  spring  was 
traditionally  thought  to  be  so  low  that,  al- 
though female  geese  could  feed,  they  could 
not  meet  their  energy  requirements  for  daily 
maintenance  or  egg  production  (Barry  1962, 
Ryder  1970;  reviewed  by  Rohwer  1992).  For 
some  species,  body  mass  of  females  generally 
increases  before  or  during  egg  production 
(Wypkema  and  Ankney  1979,  Budeau  et  al. 
1991,  Bromley  and  Jarvis  1993,  Choiniere 
and  Gauthier  1995)  indicating  that  energy  in- 
take during  these  periods  could  at  least  meet 
requirements  for  daily  maintenance  (Ganter 
and  Cooke  1996,  Carriere  1996).  Further,  var- 
iation in  timing  of  nesting  can  be  related  to 
variation  in  food  availability  prior  to  egg  for- 
mation (Prop  and  de  Vries  1993).  Thus,  food 
availability  and  use  in  the  Arctic  in  spring 
clearly  is  significant. 

In  the  Arctic,  food  availability  during  pre- 
incubation is  highly  variable.  This  variation 
hinders  interspecific  comparisons  of  feeding 
ecology  during  preincubation  because  we  can- 
not differentiate  between  factors  that  are  site 
specific  (e.g.,  weather,  snow  melt  patterns, 
plant  phenology)  and  species  specific  (e.g., 
body  size,  bill  moiphology;  Prevett  et  al. 
1985,  Fox  et  al.  1992). 


We  compared  the  feeding  ecology  of  Great- 
er White-fronted  Geese  (Anser  albifrons  fron- 
talis) and  Canada  Geese  {Branta  canadensis 
hutchinsii)  during  preincubation.  These  spe- 
cies have  similai'  body  mass  (White-fronted 
Goose  8%  > Canada  Goose;  R.G.B.,  unpubl. 
data),  reproductive  chronology,  and  share  the 
same  Arctic  spring  feeding  and  nesting 
grounds  on  the  Kent  Peninsula,  NWT,  Canada. 

Our  objectives  were  to  determine  (1)  how 
snow  melt  affected  the  availability  of  feeding 
habitats  and  plant  biomass,  (2)  whether 
White-fronted  and  Canada  geese  differed  in 
their  use  of  habitat  and  food  during  preincu- 
bation on  a shared  nesting  ground,  and  (3) 
how  changes  in  habitat  and  food  availability 
resulting  from  snow  melt  affected  dispersal, 
habitat  and  food  use,  and  the  timing  of  nesting 
in  these  two  species  of  geese. 

STUDY  AREA  AND  METHODS 

The  study  was  conducted  on  the  Walker  Bay  Study 
Area  (68°  22'  N,  108°  04'  W),  southwest  Kent  Penin- 
sula, Northwest  Territories  (Eig.  1),  as  part  of  longterm 
studies  on  the  breeding  ecology  of  White-fronted  and 
Canada  geese  (see  Bromley  et  al.  1995).  The  shallow 
plain  of  the  river  valley  is  a high  density  nesting  area 
for  both  species  of  geese  (Bromley  et  al.  1995). 

Studies  of  the  feeding  ecology  during  preincubation 
were  made  on  an  intensive  study  area  positioned  to 
permit  semicontinuous  observation  of  geese  in  spring 
while  minimizing  disturbance  caused  by  human  move- 
ments (Eig.  1 ).  The  area  was  further  divided  into  two 
sites:  site  A (0.72  km-)  was  representative  of  the  hab- 
itats found  all  along  the  river  and  typical  of  areas  in 
which  arriving  geese  concentrated  until  snow  melt;  site 
B (4.33  km-)  was  representative  of  the  rest  of  the  gen- 
eral study  area  where  nesting  typically  occurred.  Site 
A,  a raised  levee  paralleling  the  river,  was  drier  than 
site  B.  Habitats  available  to  geese  were  classified  from 
dry  to  wet:  hummock,  mud-fiat,  pond  margin,  and 
pond  (see  plant  list  per  habitat  in  Carriere  1996).  Mud- 
fiat  habitats  were  sparsely  vegetated  flats  of  exposed 
glacial  marine  sediments  (ca  3000  yrs  ago;  Dyke  and 
Dredge  1989),  largely  saline  clays  and  silts.  Hummock 
habitats  were  formed  by  frost  heaves  (5-30  cm  high) 
and  covered  by  thin  soil  where  Salix  spp.  dominated. 
Pond  margins  were  the  edges  of  depressions  that  were 
wet  from  snow  melt  to  early  June,  had  low  salinity, 
and  formed  meadows  dominated  by  graminoids  and 
forbs.  Pond  habitats  were  depressions  0.10-1.00  m 
deep,  inundated  at  least  until  early  July,  and  dominated 
by  hydrophilic  forbs.  All  habitats  formed  a fine  grained 
mosaic  with  patches  (i.e.,  continuous  areas  of  same 
habitat)  of  various  shapes.  Most  patches  covered  50- 
900  m-,  with  some  mud-flats  outside  sites  A and  B 
extending  more  than  1 km-. 

General  phenology. — We  recorded  the  phenology  of 


168 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


Nests 

White-fronted  geese 

1993 

1994 

Canada  geese 

1993  * 

1994  ■ 

Drumlin  — — 
Lake/Pond  ^ 

T ransect  

Tower  ■ 


FIG.  1.  Walker  Bay  Study  Area  is  situated  in  a valley  near  Walker  Bay  on  the  Kent  Peninsula,  NWT, 
Canada.  A section  selected  for  detailed  study  was  divided  into  2 sites:  site  A was  snow-free  first  in  spring  and 
was  typical  of  habitats  available  along  the  river  banks.  Snow  melt  phenology  in  site  B was  typical  of  the  rest 
of  the  study  area. 


snow  melt,  goose  arrival,  and  goose  dispersal  along 
four  1 km  transects  (Fig.  1).  At  2-3  day  intervals  we 
visually  estimated  snow  cover  (±5%)  in  200  X 200  m 
areas,  situated  in  pairs,  one  on  each  side  of  the  transect 
at  each  of  five  stations  200  m apart.  Thus,  we  covered 


11%  of  site  A (2  stations)  and  7%  of  site  B (18  sta- 
tions). At  each  station,  we  used  binoculars  to  count 
and  locate  all  geese  that  could  be  identified  to  species. 

Preincubation  was  divided  into  three  periods  defined 
by  snow  melt  phenology  on  site  A:  (1)  before  snow 


Carhere  el  al.  • SPRING  FEEDING  BY  ARCTIC  GEESE 


169 


melt  (snow  cover  >80%).  (2)  during  snow  melt  (snow 
cover  = 20-80%)- — a period  of  rapid  change,  and  (3) 
after  snow  melt  (snow  cover  <20%). 

Nesting  phenology. — Nests  were  located  by  observ- 
ing nest  building  behavior  from  towers.  Date  of  the 
first  egg  laid  (i.e.,  date  of  nest  initiation)  was  estimated 
only  for  nests  found  during  laying  by  subtracting  1 .3 
d per  egg  (R.G.B.,  unpubl.  data)  already  laid.  Nests 
outside  site  A and  B were  located  during  ground  sur- 
veys of  randomly  selected  1 km-  plots  (Bromley  et  al. 
1995).  We  compared  median  laying  dates  for  combi- 
nations of  years  and  species  using  Kruskal-Wallis  AN- 
OVA  on  Ranks  tests,  followed  by  non-parametric  mul- 
tiple comparison  tests  using  Excel  97  following  Daniel 
(1978). 

Habitat  availability. — We  measured  habitat  avail- 
ability in  the  areas  at  two  towers  (see  below)  by  map- 
ping habitat  patches  with  a 20  X 20  m grid  in  the  field 
and  using  aerial  photographs.  Availability  of  a habitat 
was  defined  by  the  cumulative  area  of  all  patches  of 
that  habitat  as  a proportion  of  the  total  snow-free  area. 
Only  snow-free  areas  were  considered  accessible  to 
feeding  geese  (Hall  et  al.  1997). 

Use  and  selection  of  habitats. — We  observed  goose 
pairs  throughout  preincubation  from  two  towers  in 
1993  and  one  tower  in  1994.  Daily  observations  were 
conducted  from  one  tower  at  a time  depending  on  the 
distribution  of  geese  in  the  study  area.  We  used  the 
scan  sampling  method  (Altman  1974),  with  scans  con- 
ducted every  2-6  hr  during  the  24  hr  cycle.  During  a 
scan,  we  observed  all  pairs  present  within  400  m of 
the  observation  tower.  Sections  of  sites  A and  B were 
visible  from  both  towers  (Eig.  1).  Eor  each  pair,  we 
noted  sex,  behavior,  and  habitat  used  by  each  pair 
member.  Sex  was  determined  using  relative  neck  size 
(larger  in  males),  abdominal  profile  (larger  in  females 
with  developing  follicles;  Owen  1981,  Eox  et  al. 
1995),  and  alert  position  (male  usually  standing  high- 
er). Pairs  were  recorded  as  feeding  when  either  the 
female  or  both  members  were  grubbing  (on  below- 
ground plant  parts),  pecking  (on  above-ground  plant 
parts),  searching  for  food  (moving  with  head  down), 
or  drinking. 

We  calculated  habitat  use  for  each  scan  as  the  pro- 
portion of  observed  feeding  pairs  using  each  available 
habitat.  Pond  and  pond  margin  use  data  were  pooled 
as  wet  habitats  to  obtain  sufficiently  large  samples  for 
a C-test  (Sokal  and  Rohlf  1981).  For  analysis,  we 
pooled  habitat  use  data  within  snow  melt  period  and 
tower.  Habitat  use  data  were  assumed  to  be  indepen- 
dent within  and  between  scans.  This  seemed  reason- 
able for  two  reasons.  First,  preliminary  observations 
on  focal  individuals  indicated  that  pairs  showed  little 
synchronization  in  movements  between  habitat  patches 
(i.e.,  no  group  behavior).  Second,  because  of  the  ex- 
treme patchiness  of  habitats,  pairs  could  use  all  avail- 
able habitats  within  2 hr  (minimum  time  interval  be- 
tween scans);  58%  of  Canada  Geese  (n  = 96)  and  43% 
of  White-fronted  Geese  (/;  = 69)  individuals  observed 
in  focals  used  more  than  I habitat  patch  within  10  min 
(S.C.,  unpubl.  data).  The  constant  movement  of  pairs 


between  the  observed  and  unobserved  areas  during 
preincubation  minimized  pseudoreplication.  For  ex- 
ample, no  individual  with  a coded  neck  collar  (from 
1987-1994  banding  operations;  Bromley  et  al.  1995) 
was  observed  in  more  than  six  different  scans  (only 
30%  of  collared  individuals  were  observed  more  than 
twice). 

For  each  year  and  period,  interspecific  differences 
in  habitat  use  were  analyzed  without  reference  to  hab- 
itat availability  using  G-tests  of  independence,  fol- 
lowed, when  significant  differences  were  detected,  by 
pairwise  unplanned  comparison  tests  (Sokal  and  Rohlf 
1981 ; using  Excel  97). 

Habitat  use  was  compared  to  availability  before  and 
after  snow  melt,  but  not  during  melting,  because  avail- 
ability changed  too  rapidly,  i.e.,  from  20%  to  80% 
snow-free  area  in  <7  d.  We  used  X"  goodness-of-fit 
tests,  followed  by  Bonferroni  simultaneous  confidence 
interval  tests  (Sokal  and  Rohlf  1981,  Neu  et  al.  1974; 
using  Excel  97)  to  detect  sources  of  significant  differ- 
ences (design  1 in  Thomas  and  Taylor  1990). 

Plant  availability  within  habitats. — We  estimated 
food  plant  availability  in  snow-free  habitats  by  sam- 
pling randomly  20  X 20  cm  quadrats  within  habitat 
patches  in  site  A and  B,  4 times  between  17  May-14 
June  throughout  preincubation  (1993:  n = 120;  1994: 
n = 218).  We  collected  all  above  and  below  ground 
vegetation  0. 1-3  cm  deep  then  froze  each  sample  for 
transportation.  Ground  vegetation  below  3 cm  was 
never  thawed  and  was  considered  unavailable  to  geese. 
In  the  laboratory,  live  above  and  below  ground  vege- 
tation was  sorted  by  species,  dried  at  45°  C to  constant 
mass,  and  weighed.  We  pooled  plant  species  according 
to  food  plant  categories  used  in  the  analysis  of  feces 
(see  below).  Availability  of  each  food  plant  category 
in  each  habitat  was  defined  as  the  average  proportion 
of  total  dry  biomass  represented  by  that  category  in 
each  habitat  during  each  snow  melt  period  (Eig.  2). 

We  analyzed  total  (pooled  species)  above  and  below 
ground  biomass  separately.  We  first  analyzed  differ- 
ences in  total  biomass  among  habitats  and  sampling 
dates  for  each  year  separately  using  Kruskal-Wallis 
ANOVA  on  Ranks  tests,  followed  by  multiple  com- 
parison tests  to  detect  pairwise  differences  (Daniel 
1978).  We  then  pooled  sampling  dates  and  analyzed 
differences  in  biomass  among  habitats  and  years  with 
Two  Way  ANOVA  using  square-root  transformed  data 
(Sokal  and  Rohlf  1981). 

Diet. — We  determined  diet  using  two  complemen- 
tary methods:  (I)  microhistological  analyses  of  feces 
and  (2)  detailed  observations  of  feeding  techniques. 
Additional  data  were  available  from  the  esophageal 
content  of  16  female  Canada  Geese  collected  in  1994. 

Throughout  preincubation,  we  collected  all  feces 
that  we  could  assign  to  an  individual  bird  (i.e.,  a bird 
that  was  observed  defecating).  Feces  were  individually 
frozen  for  transportation,  dried  in  the  laboratory,  and 
analyzed  using  microhistological  techniques  (see  John- 
son 1982).  We  sampled  four  slides  per  feces,  with  20 
observation  fields  per  slide.  We  identified  most  plant 
fragments  (80-100%  per  feces)  to  genus.  We  grouped 


170 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


(/) 

CO 
CT3 

E 

g 

1q 

40 

O 20 
^ 0 
^ 20 
^ 40 

60 


40 

20 

0 

20 

40 

60 


A 


Pond 


I After  (20) 


/ 


Mud-flat 


’ ^ 

Q' 

# i 

S 

' -3 

d 

e 

1 1 

^ i 

■'s/ 

Si 

0) 

<tr 

J 1 

i 

s- 

(3 

§ 

U7  S' 

_l  L 


Hummock 


Margin 


Pond 


FIG.  2.  Relative  availability  of  plants  in  each  snow-free  habitat  before,  during,  and  after  snow  melt,  measured 
as  the  average  percent  food  in  dry  biomass  on  the  Walker  Bay  Study  Area,  NWT,  May-June  1993-1994.  Sample 
sizes  are  given  in  parentheses,  years  pooled.  Availability  of  aboveground  plant  parts  is  given  above  the  zero 
line  and  of  below  ground  plant  parts  below  the  zero  line. 


some  rare  food  plants  of  the  same  genus  or  family  that 
were  growing  in  the  same  habitat  (Table  1,  Fig.  2). 
The  validity  of  the  identification  of  plant  fragments  in 
feces  was  assessed  by  analyzing  plant  mixtures  of 
known  composition  (Holechek  and  Gross  1982)  and 
by  comparing  fecal  results  to  esophageal  contents  of 
female  Canada  Geese  (Carriere  1996).  We  could  not 
easily  differentiate  Sali.x  arctica  from  Potentilla  nivea, 
nor  discriminate  Dupontia  fisheri  and  Eriophorum  spp. 
from  some  fragments  of  other  Gramineae  and  Cyper- 
aceae,  respectively.  Potentilla  spp.  and  Eriophorum 
spp.,  however,  were  relatively  rare  on  the  study  area 
(Fig.  2,  see  other  Dicots  and  Eriophorum  spp.).  Some 
below  ground  plant  parts  could  not  be  identified  reli- 
ably (Carriere  1996).  Feces  analyses  consequently  pro- 
vided only  an  estimate  of  use  of  common  above 
ground  food  plants  on  our  study  area.  Coefficients  of 
variation  of  the  proportion  of  food  plants  among  slides 


within  feces  ranged  from  less  than  5%  for  graminoids 
to  10—20%  for  dicots.  We  defined  diet  as  the  average 
percent  of  each  food  plant  (%  of  fragments)  present  in 
the  feces  (frequency  of  occurrence  method;  Johnson 
1982).  Diet  was  determined  for  each  goose  species  and 
snow  melt  period  by  pooling  years,  with  individual 
feces  representing  sampling  units. 

We  compared  use  of  each  food  plant  among  all  pos- 
sible combinations  of  goose  species  and  snow  melt 
periods  using  Kruskall-Wallis  ANOVA  on  Ranks  tests 
followed,  if  significant,  by  multiple  comparison  tests 
(Daniel  1978)  using  Excel  97. 

Feeding  techniques. — Use  of  below  ground  plant 
parts  was  indexed  by  the  proportion  of  feeding  pairs 
grubbing  in  each  habitat  during  scans.  Conversely, 
pairs  were  assumed  to  eat  above  ground  plant  parts 
while  pecking. 

We  used  a Paired  ?-test  to  determine  if  the  propor- 


TABLE  1 . Diet  of  White-fronted  and  Canada  geese  before,  during,  and  after  snow-melt,  from  arrival  on  the  nesting  grounds  to  incubation.  Walker  Bay  Study 
Area,  Kent  Peninsula,  NWT,  1993—94.  “Arriving”  females  were  collected  before  snow-melt,  “pre-laying”  females  during  and  after  snow-melt,  and  the  incubating 
female  after  snow-melt. 


Carriere  et  ui  • SPRING  FEEDING  BY  ARCTIC  GEESE 


o 


o 


'sO 


GN 


00  ^00 

— sO  ^ 


\n  ^ — o 

so 


SO  so 
Ov 


_ 

(N 

_ 

OI 

n 

o 

m 

O 

O 

o 

o 

o 

o 

o 

o 

o 

O 

o 

o 

o 

d 

d 

d 

d 

d 

d 

d 

d 

V 

A 

V 

V 

V 

ro 

o 

o^ 

ov 

vO 

VO 

VO 

ot 

n 

n 

CQ 

U 

02 

< 

U 

02 

< U 

02 

(N 

lo 

in 

in 

•rt  n 

Ov 

Urn 

ov 

m 

1^ 

n n 

(N 

m 

U 

m 

< 

02 

u < 

00 

ro 

— 

(N 

— 

— 

n 

u IT) 


■rt 

(N 

o 

u 

O 

r^i 

(N 

02 

U 

< 

< 

in 

m 

— 

— 

< 

< 

in 

m 

u 

m 

in 

(■■I 

u 

SO 

r- 

U 

CQ 

U 

CQ 

u 

< 

< 

02 

< 

02 

PP 

r't 

in 

— 

in 

— 

VO 

in 

— 

in 

u 

cs 

o 

in 

in 

(N 

(^1 

pa 

PP 

(J 

< 

U 

< 

< 

02 

OP 

so 

— 

< 

r' 

rn 

00 

o 

On 

ri 

u 

tn 

ro 

" 

n 

U 

U 

PP 

u 

— 

fn 

o 

ri 

A 

A 

< 

r- 

ri 

so 

n 

os 

X- 

ri 

in 

ri 

c/) 

<U 

CJ 

OX) 


O ic: 


<u 

P 

5 

N 


1) 

C3 

G 

C 

E 

03 

u 

o 


o 

cj 


•t:  i:  t:  'ij 


(u  C 
= 5; 
■“  Si 
d.^ 
a 


V 2 

^ S- 


UD  "O 

■o  o 

^ S 

« - 

OJ 

"O 


a,k)O'^O^Q03:GD 


Q. 

£ 


"O 
3 ■ 
42 


§• 

O 


) ^ 

5 ' 


■ • : 

; ^ Cl  , 


-S 


^ O. 

i=  a. 

5 t/5 


S =>, 


“■S3 

I Q I 
■2 

^✓5  V)  . . 


^ -S 
S -SJ- 
s£ 


: <n  i 
“ o S ^ 

; s cL  S ^ 

: O S' 


5 

X. 


O O 

o T3  y 


171 


f Hippiiris  sp.:  W.  vulfturis  or  //.  lelniphylla.  Senecio  sp.:  5.  conges/us,  includes  rhizomes,  young  shoots,  and  leaves. 

Unidentified  includes  fragments  of  dicots,  of  Eriophorum  spp..  and  all  rhizomes  except  those  of  Etymus  sp..  and  Hippiiris  spp. 
' tr  = trace. 


172 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


Date 


10 


8 

6 

4 

2 

0 


■O 

.2 

c 

<0 

to 


a> 

c 

o 

% 


E 

D 


z 


10 


6 

4 

2 

0 


0) 

E 

3 


Z 


LIG.  3.  Number  of  individual  White-fronted  Geese  (WL)  and  Canada  Geese  (CG)  counted  daily  at  each 
station  on  sites  A (/;  — 2)  and  B (//  = 18;  see  Lig.  1),  from  arrival  to  early  incubation,  and  number  of  nests 
initiated  daily  by  each  species  on  the  Walker  Bay  Study  Area,  NWT,  1993-1994.  Arrows  indicate  duration  of 
rapid  follicular  development  (RED)  for  a first  egg  layed  at  the  median  laying  date  each  year.  Shaded  areas 
indicate  periods  of  snow  melt,  defined  as  80-20%  snow  cover.  Snow  melt  pattern  on  site  A was  used  to  define 
periods  for  all  analyses. 


tion  of  time  spent  grubbing  differed  between  goose 
species,  pairing  observations  within  habitat-period- 
year  combinations.  Lor  each  species,  we  used  linear 
regre.ssion  to  analyze  how  the  proportion  (arcsine 
square-root  transformed  in  degrees;  Sokal  and  Rohlf 
1981)  of  time  spent  grubbing  changed  with  the  pro- 
portion of  below  ground  biomass  available  in  different 
habitats,  snow  melt  periods,  and  years.  The  short,  rapid 
melting  period  in  1994  was  combined  with  the  after- 
melt  period. 

Unless  otherwise  stated,  statistical  analyses  were 
performed  using  SigmaStat®  (version  1.0;  Jandel  Sci- 
entific Software  1993)  in  the  PC  DOS  and  Microsoft® 
Windows®  operating  system.  Statistical  significance 
was  established  at  P < 0.05. 


RESULTS 

Annual  phenology. — Upon  our  airival  on 
20  May  1993  and  17  May  1994,  sites  A and 
B were  95%  and  99%  snow  covered,  respec- 
tively. Hummock  and  mud-flat  habitats  locat- 
ed along  the  river  were  the  only  habitats  avail- 
able. Relative  to  1987-1996  (R.G.B.,  unpubl. 
data),  snow  melt  was  early  in  1994  and  av- 
erage in  1993  (Fig.  3).  In  general,  snow  melt- 
ed first  on  dry  habitats  (mud-flat,  hummock), 
then  on  wet  ones  (pond  margin,  pond),  and 
earlier  on  site  A than  on  site  B (Fig.  3).  The 
periods  before  snow  melt  were  10  days  (1993) 


Carriere  el  al.  • SPRING  FEEDING  BY  ARCTIC  GEESE 


173 


and  7 days  (1994);  melting  periods  were  7 
days  (1993)  and  3 days  (1994),  and  periods 
after  snow  melt  [end  of  melting  to  onset  of 
incubation  (median  laying  date  + 4 days)] 
were  9 days  (1993)  and  13  days  (1994).  The 
total  length  of  the  preincubation  periods  on 
the  nesting  ground  were  thus  26  d in  1993  and 
23  d in  1994. 

Geese  were  present  along  the  river  (on  site 
A or  outside  the  intensive  study  area)  upon 
our  aiTival  in  1993  {n  = 50-80  pairs)  and 
1994  {n  = 140-150  pairs),  but  major  anivals 
occuned  aiound  27  May  1 993  and  on  or  prior 
to  20  May  1994  (see  transect  surveys:  Fig.  3, 
site  A).  During  snow  melt,  goose  numbers  de- 
creased on  site  A (and  from  other  areas  along 
the  river)  and  increased  slightly  on  site  B as 
geese  dispersed  from  areas  along  the  river  to 
the  rest  of  the  study  area  (Figs.  1,  3). 

We  are  confident  that  the  pairs  observed 
during  this  study  were  part  of  the  locally  nest- 
ing population  because  about  Vi  of  the  collared 
individuals  we  found  nesting  in  the  study  area 
{n  > 1 1 of  each  species  per  year)  were  ob- 
served on  site  A and  other  areas  along  the 
river.  Individuals  collared  in  other  study  areas 
were  never  observed  on  Walker  Bay  Study 
Area  (R.G.B.,  pers.  obs.). 

The  first  nests  were  found  on  5 June  1993 
and  on  28  May  1994.  Median  laying  dates 
were  11  June  1993  (White-fronted  Goose:  n 
= 38,  Canada  Goose:  /r  = 35)  and  2 June 
1994  (White-fronted  Goose:  n = 26,  Canada 
Goose:  n = 21)  for  both  species  (Kruskal- 
Wallis  ANOVA  on  Ranks:  H = 77.4,  P < 
0.001;  medians  were  different  between  years 
only;  Fig.  3).  The  minimum  intervals  between 
goose  arrival  and  median  laying  date  were  14 
days  in  1993  and  13  days  in  1994. 

Use  and  selection  of  habitats. — White- 
fronted  and  Canada  geese  did  not  differ  in 
their  use  of  habitats  before  snow  melt  in  either 
year  (G-tests  of  independence:  1993:  G.,jj  = 
0.083,  1 df,  F > 0.05;  1994:  G,„  = 0.031,  1 
df,  P > 0.05;  Fig.  4A,  B).  Both  species  also 
used  habitats  according  to  their  availability 
during  that  period  (y^  Goodness-of-fit:  all 
< 3.08,  P > 0.05). 

Habitat  use  differed  between  species  during 
snow  melt  in  both  years  (G-tests  of  indepen- 
dence: 1993:  G,jj  = 18.14,  P < 0.001;  1994: 
Gajj  = 39.26,  P < 0.001).  Both  species  fed  in 
newly  available  wet  habitats  (pond  margin 


and  pond)  during  snow-melt,  but  White- 
fronted  Geese  used  them  more  often  than  did 
Canada  Geese. 

Pairwise  compaiison  tests  showed  that  in 
both  years  Canada  Geese  used  hummock  and 
mud-flat  habitats  significantly  more  often  than 
White-fronted  Geese  and  that  White-fronted 
Geese  used  wet  habitats  significantly  more  of- 
ten than  Canada  Geese  (Fig.  4C,  D). 

Similar  interspecific  differences  in  habitat 
use  were  observed  after  snow  melt  in  both 
years  (G-test  of  independence:  1993:  G^^j  = 
23.03,  P < 0.001;  1994  sites  pooled:  G = 
69.44,  P < 0.001).  Pairwise  comparison  tests 
showed  that  White-fronted  Geese  used  ponds 
(1993  and  1994)  and  pond  margins  (1994) 
significantly  more  often  than  Canada  Geese 
(Fig.  4E,  F,  G). 

In  1993,  habitat  selection  (i.e.,  use  vs  avail- 
ability) by  both  species  differed  after  snow 
melt  (x"  Goodness-of-fit:  Canada  Goose:  x^  = 

94.8,  P < 0.001;  White-fronted  Goose:  yf  = 
117.9,  P < 0.001).  Canada  Geese  preferred 
(i.e.,  use  > available)  mud-flat,  pond  margin, 
and  pond,  and  avoided  hummock  habitats 
(Fig.  4E).  White-fronted  Geese  preferred  pond 
and  avoided  hummock  habitat  (Eig.  4E).  In 
1994,  most  geese  fed  neai'  tower  1 after  snow 
melt  (Fig.  1),  where  the  relative  availability  of 
habitats  differed  greatly  between  site  A and  B 
(Fig.  4).  Consequently,  we  analyzed  habitat 
use  in  each  site  separately.  Data  in  dry  (mud- 
flat/hummock)  and  wet  (pond  maigin/pond) 
habitats  were  pooled  at  site  B to  obtain  suf- 
ficient sample  sizes  for  x^  tests.  Neai'  the  river 
(site  A),  Canada  Geese  used  habitats  accord- 
ing to  their  availability  (x'  Goodness-of-fit:  x" 
= 0.92,  P > 0.05;  Fig.  4F),  while  White-front- 
ed Geese  prefeired  pond  maigin  and  pond 
habitats  and  avoided  mud-flat  habitats  (x“  = 

49.8,  P < 0.001).  On  site  B,  Canada  Geese 
again  used  habitats  according  to  their  avail- 
ability (x^  = 0.22,  P > 0.05;  Fig.  4G)  and 
White-fronted  Geese  still  prefeired  wet  habi- 
tats (pond  margin/pond)  and  avoided  dry  ones 
(mud-flat/hummock;  x‘  = 23.4,  P < 0.001). 

Diet. — Nine  food  types  were  recognized  in 
feces  (Table  1 ).  The  number  of  different  food 
types  detected  increased  as  snow  melted,  as 
expected  because  of  the  increase  in  available 
habitats.  A major  food  before  snow  melt  was 
Puccinellia  spp.  tillers,  which  accounted  for 
52%  and  73%  of  White-fronted  and  Canada 


174 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  2,  June  1999 


>s 

B 

ro 

CD 

> 

CD 

TD 

C 

(D 

0) 

(0 

D 

ro 

B 

CD 

X 


1993  1994 


Mud  Hummock  Mud  Hummock 


100 


Mud  Hummock  Margin  Pond  Mud  Hummock  Margin  Pond 


I I Canada  Geese 
El  White-fronted  Geese 
H Availability  (expected  use) 


Mud/Hummock  Margin/Pond 


FIG.  4.  Use  (%  of  feeding  pairs)  of  mud-flat,  hummock,  pond  margin,  and  pond  habitats  by  White-fronted 
(WE)  and  Canada  geese  (CG),  and  availability  of  these  habitats  during  preincubation  on  the  Walker  Bay  Study 
Area,  NWT,  1993-1994.  (a-b)  Before  snow  melt  1993  (CG:  n = 332;  WE:  n = 228)  and  1994  (CG:  /;  = 1265; 
WE;  n = 383).  (c-d)  During  snow  melt  1993  (CG:  n = 532;  WF:  n = 137)  and  1994  (CG:  n = 637;  WE:  n = 
168);  habitat  availability  could  not  be  estimated,  (e)  After  snow  melt  at  site  A in  1993  (CG:  /;  = 120;  WF;  n 
= 87).  (f-g)  After  snow  melt  in  1994,  site  A (CG:  n = 50;  WF;  n = 33)  and  site  B (CG;  n = 22;  WF:  n = 
29).  A -I-  indicates  use  > availability,  indicates  use  < availability  and  no  symbol  indicates  use  = 

availability  (P  > 0.05). 


goose  diet  respectively.  After  snow  melt,  more 
than  50%  of  the  diets  of  White-fronted  and 
Canada  geese  were  composed  of  Carex  spp. 
and  Diipontia  fisheri.  For  both  species,  most 
(85%)  feces  contained  food  plants  represen- 
tative of  at  least  2 habitats.  Consequently, 
changes  in  diet  reflected  changes  in  both  hab- 
itat use  and  food  use  within  habitats. 

The  greater  diversity  of  the  diet  found  in 
the  feces  of  Canada  Geese  relative  to  esophagi 
(Table  1 ) is  probably  caused  by  the  accumu- 
lation of  food  fragments  in  the  digestive  sys- 
tem. Nevertheless,  analyses  of  esophageal 
content  indicated  that  some  Gramineae  (frag- 
ments not  identified  at  the  genus  level)  could 
be  overestimated  in  fecal  analyses,  whereas 


Stellaria  humifusa,  Diipontia  fisheri,  and  Carex 
seeds  could  be  underestimated  (see  methods, 
Cairiere  1996).  Young  leaves  and  open  buds 
of  Salix  spp.  were  found  only  in  female  Can- 
ada Geese  collected  during  incubation. 

Available  biomass. — Plant  biomass  varied 
among  combinations  of  habitats  and  sampling 
dates  for  both  above  ground  (Kruskal- Wallis 
ANOVA  on  ranks:  1993;  H = 46.0,  P < 
0.001;  1994:  H = 152.1,  P < 0.001)  and  be- 
low ground  plant  parts  (1993:  H = 25.7,  P = 
0.012;  1994:  H = 58.7,  P < 0.001;  Fig.  5). 
There  was  little  seasonal  increase  in  total  dry 
biomass  within  each  habitat,  except  after  the 
median  laying  date  lor  above  ground  biomass 
in  pond  margins  (1994  only)  and  below 


Carriere  et  al.  • SPRING  FEEDING  BY  ARCTIC  GEESE 


175 


21  28  'June  6 14 

MUD-FLAT  HUMMOCK  MARGIN  POND 


FIG.  5.  Live  above  and  below  ground  dry  biomass 
(mean  ± SE)  available  to  geese  in  4 habitats  during 
preincubation  on  the  Walker  Bay  Study  Area,  NWT, 
1993-1994.  Values  with  same  letter  within  graphs  do 
not  significantly  differ  (multiple  comparison  tests  after 
a Kruskall-Wallis  ANOVA  on  Ranks  test  for  each  year, 
above  and  below  ground  biomass  tested  separately,  all 
tests  P < 0.02).  Grey  areas  indicate  periods  of  snow 
melt  (80-20%  snow  cover  on  site  A).  Number  of  sam- 
pled 20  X 20  cm  quadrats  per  sampling  date  were: 
(1993)  mud-flat  = 3,  5,  17,  14;  hummock  = 6,  9,  11, 
1 1;  pond  margin  = NA,  3,  6,  6;  pond  = NA,  NA,  3, 
26  and  (1994)  mud-flat  = 12,  16,  24,  23;  hummock  = 
10,  16,  16,  16;  pond  margin  = NA,  17,  16,  16;  pond 
= NA,  NA,  16,  20.  NA  = datum  was  not  available, 
under  snow. 


ground  biomass  in  pond  margins  (1993)  and 
pond  (1994).  Seasonal  changes  in  below 
ground  biomass  were  mostly  the  result  of  a 
gradual  deepening  of  the  active  layer.  We 
pooled  sampling  dates  and  found  differences 
in  above  ground  biomass  among  habitats 
(Two  Way  ANOVA:  F = 1 12.1,  P < 0.001) 
and  years  (F  = 4.25,  P = 0.04;  habitat-year 
interaction,  F = 7.25,  P < 0.001).  Above 
ground  biomass  in  pond  margins  was  higher 
than  in  any  other  habitat,  particularly  in  1994. 
We  found  a tendency  for  below  ground  bio- 
mass to  be  lower  in  dry  (0-1.08  g/m^)  than  in 
wet  habitats  (1.33-24.4  g/m-;  Two  Way  AN- 
OVA: F = 2.61,  P = 0.05),  but  no  difference 


EIG.  6.  Relationship  between  below  ground  bio- 
mass available  and  proportion  of  feeding  time  spent 
grubbing  by  female  White-fronted  (WE)  and  Canada 
geese  (CG)  during  preincubation  on  the  Walker  Bay 
Study  Area,  NWT,  1993-1994.  Each  point  is  the  pro- 
portion of  scanned  pairs  observed  grubbing  (arcsine 
square-root  transformed)  in  a habitat  during  a specific 
period  and  year.  Closed  symbols  are  Canada  Geese, 
open  symbols  are  White-fronted  Geese.  Habitats  are 
mud-flat  (circles),  hummock  (squares),  pond  margin 
(triangles),  and  pond  (diamonds).  Linear  regressions: 
for  WE:  y = 0.78  -F  0.91x  [0.50-1.32;  6(95%C1)],  F 
= 18.0,  P = 0.001;  for  CG:  y = 0.28  + 0.72x  [0.27- 
1.17;  B(95%CI)],  F = 9.9,  P = 0.009. 

between  years  {F  — 1.53,  P > 0.05;  habitat- 
year  interaction,  F = 4.09,  P = 0.007). 

Use  of  below  ground  food  plants. — Canada 
Goose  pairs  grubbed  less  often  than  White- 
fronted  Geese  (Canada  Goose:  12.4%,  White- 
fronted  Goose:  56.9%;  Paired  f-test:  t — 
— 14.2,  P < 0.001;  Fig.  6),  which  grubbed  ex- 
tensively even  in  habitats  where  below-ground 
biomass  was  very  low  (i.e.,  in  mud-flat/hum- 
mock),  before  widespread  snow  melt  and 
thawing  of  the  ground  surface.  Spatial  distri- 
bution of  below  ground  vegetation  (mostly 
Elymus  sp.  rhizomes)  was  clumped  in  these 
habitats  and  available  below  ground  biomass 
may  have  effectively  been  much  higher  in 
some  patches  used  by  foraging  individuals.  In 
all  habitats,  before  and  during  snow  melt, 
most  of  the  below  ground  biomass  was  frozen 
under  the  active  layer  and  hence  was  not 
available  to  geese. 

There  was  a significant  positive  relationship 
between  the  proportion  of  below  ground  bio- 
mass in  different  habitat-year-period  combi- 
nations and  the  proportion  of  feeding  time 


176 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


geese  spent  grubbing  in  these  habitats  (linear 
regression:  White-fronted  Geese:  F = 18.0,  P 
= 0.001;  Canada  Geese:  F = 9.9,  P = 0.009; 
Fig.  6).  This  indicates  that  both  White-fronted 
and  Canada  geese  modified  their  feeding  be- 
havior with  changes  in  below  ground  food 
availability. 

DISCUSSION 

Arrival,  phenology,  and  timing  of  laying. — 
The  minimum  interval  between  peak  arrival 
and  laying  was  1-4  day  longer  than  necessary 
for  rapid  follicular  development  (Alisauskas 
and  Ankney  1992a),  which  is  typical  in  most 
goose  populations  during  eaily  and  average 
years  (Raveling  1978,  Fox  and  Madsen  1981, 
Budeau  et  al.  1991,  Gauthier  and  Tardif  1991, 
Bromley  and  Jarvis  1993).  Anival  and  laying 
initation  were  both  earlier  in  1994  than  in 
1993  and  peak  initiation  of  laying  occurred 
about  15  days  after  peak  arrival  in  1993,  but 
apparently  only  12-13  days  after  airival  in 
1994.  The  longer  interval  in  1993  likely  was 
due  to  later  snow  melt  that  yeai"  nest  initiation 
closely  followed  snow  melt  in  site  B.  In  con- 
trast, during  the  early  snow  melt  of  1994, 
nesting  sites  became  available  (i.e.,  when 
snow  cover  <50%  in  site  B)  5-8  days  before 
peak  nest  initiation,  indicating  that  completion 
of  RFD,  rather  than  snow  melt,  constrained 
nesting  that  year.  These  relationships  were  re- 
markably similar  in  White-fronted  and  Canada 
geese. 

The  close  proximity  of  prenesting  feeding 
sites  to  nesting  sites  conveys  benefits  to  geese 
arriving  on  the  Arctic  nesting  grounds  before 
RFD.  During  both  years,  pairs  of  both  species 
dispersed  from  early  exposed  locations  near 
the  river  to  additional  areas  exposed  late  dur- 
ing snow  melt.  This  resulted  in  their  using 
feeding  areas  that  were  increasingly  distant 
from  the  river.  Such  short  distance  movement 
by  breeding  pairs  from  feeding  areas  to  nearby 
nesting  areas  has  been  observed  in  other 
goose  populations  (Gauthier  1993,  Prop  and 
de  Vries  1993).  This  local  dispersal  is  similar 
to  the  final  migratory  flight  from  staging  areas 
to  the  breeding  grounds  in  populations  where 
most  females  nest  soon  after  arrival  in  the 
Arctic  (Ankney  and  Machines  1978,  Wypke- 
ma  and  Ankney  1979).  For  the  latter  case, 
however,  the  flight  occurs  late  in  RFD  (Wyp- 
kema  and  Ankney  1979).  Unlike  geese  nesting 


adjacent  to  feeding  sites,  this  flight  necessarily 
draws  largely  upon  body  reserves  during  RFD 
because  long  migratory  flights  would  use 
more  energy  than  local  dispersal.  Further- 
more, earlier  anival  on  the  breeding  ground 
enables  pairs  to  directly  assess  nesting  con- 
ditions (Wypkema  and  Ankney  1979,  Peterson 

1992,  Rohwer  1992,  Ganter  and  Cooke  1996). 

Although  little  plant  growth  occurred,  food 

availability  rapidly  and  greatly  increased  dur- 
ing preincubation.  The  change  in  total  avail- 
able biomass  was  due  to  the  rapid  exposure 
of  habitats  with  large  plant  biomasses  during 
snow  melt,  a common  phenomenon  in  tundra 
ecosystems  (Wielgolaski  et  al.  1981).  This 
may  play  an  important  role  in  determining  in- 
dividual reproductive  decisions  in  many  Arc- 
tic nesting  geese  (Gauthier  1993,  Prop  and  de 
Vries  1993)  because  increasing  food  avail- 
ability before  and  during  egg-production 
could  improve  individual  condition  (as  in- 
dexed by  body  mass;  Bromley  and  Jaivis 

1993,  Choiniere  and  Gauthier  1995,  Ganter 
and  Cooke  1996).  Which  components  of  re- 
productive success  (laying  date,  clutch  size, 
nest  attentiveness,  condition  at  hatching,  or  a 
combination  of  these)  may  be  affected  by  a 
female’s  improved  body  condition  depends  on 
the  timing  of  changes  in  food  availability  with 
respect  to  RFD  initiation  in  individual  females 
(e.g.,  Bolton  et  al.  1993,  Dalhaug  et  al.  1996). 

Interspecific  differences  in  feeding  ecolo- 
gy.— Snow  melt,  and  the  concomitant  changes 
in  habitat  and  food  availability,  enabled  us  to 
detect  similarities  and  differences  in  resource 
use  and  selection  between  White-fronted  and 
Canada  geese  during  preincubation.  In  both 
years,  both  species  shared  feeding  sites  and 
had  the  same  average  date  of  nest  initiation 
(i.e.,  RFD  initiation),  hence  they  could  expe- 
rience the  same  changes  in  the  availability  of 
resources  at  similar  times  relative  to  their  re- 
productive process. 

Both  species  used  habitats  opportunistically 
before  snow  melt  when  pair  density  was  high- 
est (up  to  600  pairs/km^  in  snow-free  areas) 
and  habitat  availability  lowest  (1—20%  of 
study  area  was  snow  free).  With  onset  of  snow 
melt,  both  species  dispersed  and  overall  pair 
density  decreased  (<40  pairs/km'). 

Alter  snow  melt,  Canada  Geese  generally 
used  habitats  more  opportunistically  than 
White-fronted  Geese,  which  consistently  se- 


Caniere  et  al.  • SPRING  FEEDING  BY  ARCTIC  GEESE 


177 


lected  wet  habitats  with  high  plant  biomass. 
Below  ground  food  plants  were  used  more  ex- 
tensively by  White-fronted  Geese  than  by 
Canada  Geese  in  all  habitats.  These  differenc- 
es are  similar  to  those  reported  between  Bran- 
ta  canadensis  interior  and  Anser  caeridescens 
caeridescens  on  a common  staging  area  on 
James  Bay  (Prevett  et  al.  1985),  but  unlike  our 
study  those  differences  could  be  explained  in 
part  by  segregation  of  feeding  areas. 

Although  we  could  not  test  for  density  de- 
pendent habitat  selection  because  of  synchro- 
nous changes  in  habitat  availability  and  in  pair 
density  with  snow  melt,  the  evidence  indicates 
that  interspecific  differences  in  the  relative 
suitability  of  habitats  for  feeding  exists.  Con- 
sistent with  a release  from  population  density 
effects  in  a constant  environment  (Rosen- 
zweig  1985,  Morse  1990),  White-fronted 
Geese  changed  from  opportunistic  to  selective 
use  of  habitats  with  the  highest  plant  biomass. 
In  contrast,  Canada  Geese  exhibited  a rela- 
tively weak  shift. 

We  suggest  three  possible  non-exclusive 
explanations  for  the  interspecific  differences 
in  habitat  use  and  diet  we  observed. 

1.  Pond  margin  and  pond  habitats  may  be 
more  profitable  to  White-fronted  Geese  than 
to  Canada  Geese  because  of  the  longer  and 
apparently  more  robust  bill  morphology  of  the 
former  (WF  culmen  is  35.3%  and  skull  is 
15.4%  longer  than  CG’s;  Bolen  and  Rylander 
1978,  Gawlik  and  Slack  1996,  R.G.B.,  unpubl. 
data).  A longer  bill  may  enable  White-fronted 
Geese  to  be  more  efficient  at  grubbing  in  the 
ground,  presumably  giving  them  easier  access 
to  resources  buried  in  a frozen  and  dry  sub- 
strate. There  is  a spectrum  of  feeding  tech- 
niques used  by  geese  to  obtain  food  (Bolen 
and  Rylander  1978,  Bellrose  1980,  Prevett  et 
al.  1985,  Ganter  and  Cooke  1996)  but  we  still 
know  little  of  how  morphology  affects  the  rel- 
ative efficiency  of  these  techniques  among 
species. 

2.  White-fronted  and  Canada  geese  may 
differ  in  their  food  use  because  their  nutrient 
requirements  for  reproduction  differ.  Protein, 
fat,  and  calcium  are  the  most  important  nutri- 
ents required  for  egg  formation  (Robbins 
1993,  Alisauskas  and  Ankney  1992a).  Geese 
generally  switch  from  a simple,  carbohydrate- 
rich  diet  before  spring  migration  to  a protein- 
rich  one  during  spring  migration  and  egg  pro- 


duction (Mainguy  and  Thomas  1985,  Prevett 
et  al.  1985,  Budeau  et  al.  1991,  Alisauskas 
and  Ankney  1992b,  Bromley  and  Jarvis  1993, 
Gauthier  1993),  suggesting  that  protein  is  a 
limiting  nutrient  to  egg  formation  in  geese 
(Krapu  and  Reinecke  1992).  Incubating  fe- 
males mainly  require  large  fat  reserves  for 
maintenance  (Raveling  1979,  Le  Maho  et  al. 
1981,  Boismenu  et  al.  1992).  Nutrient  require- 
ments for  egg  formation  and  incubation  there- 
fore differ,  and  whether  a female  should  select 
for  protein  or  energy  rich  food  during  egg  pro- 
duction will  depend  on  her  initial  nutrient  re- 
serves, and  on  how  requirements  are  met 
through  food  intake  and  reserve  reallocation. 

How  females  allot  nutrients  ingested  during 
preincubation  to  short  (egg  production)  and 
long  term  (incubation)  requirements  may  be 
reflected  in  their  incubation  behavior  (Brom- 
ley 1984,  Thompson  and  Raveling  1987). 
White-fronted  Geese  have  cryptic  nesting  be- 
havior, and  like  Emperor  and  Giant  Canada 
{B.  c.  maxima)  geese,  rely  mostly  on  stored 
nutrients  for  incubation  (mean  feeding  time 
per  day  = 1-8  min;  Thompson  and  Raveling 
1987;  R.G.B.,  unpubl.  data).  In  the  central  Ca- 
nadian Arctic,  Canada  Geese  nest  openly,  de- 
fend their  nests,  and  take  frequent  recesses  to 
feed  (mean:  40  min  per  day;  Jarvis  and  Brom- 
ley, in  press;  R.G.B.,  unpubl.  data).  Little 
feeding  during  incubation  by  White-fronted 
Geese  suggests  a greater  requirement  for  en- 
ergy rich  food  prior  to  incubation.  Longer  in- 
cubation recesses  by  Canada  Geese  imply  a 
greater  reliance  on  foraging  to  meet  nutrient 
requirements  during  incubation,  allowing  for 
more  nutrients  (e.g.,  protein)  ingested  prior  to 
incubation  for  egg  production  (see  Thompson 
and  Raveling  1987).  Thompson  and  Raveling 
(1987)  suggested  that  greater  incubation  atten- 
tiveness may  be  related  to  larger  body  size  in 
geese,  because  they  are  vulnerable  to  different 
types  of  predators  and  have  a greater  fasting 
endurance  (Calder  1974,  Boismenu  et  al. 
1992). 

3.  For  Canada  Geese,  feeding  pair  density 
may  not  entirely  reflect  the  suitability  of  the 
habitat  for  feeding  during  laying  (Van  Home 
1983).  Canada  Geese  typically  nest  on  pond 
margins  (R.G.B.,  unpubl.  data),  and  as  laying 
time  approaches  pairs  may  show  increasing 
tenitorial  behavior  in  some  patches  of  this 
habitat.  Intraspecific  aggressive  behavior  was 


178 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


observed,  and  could  exclude  conspecifics 
from  some  patches  of  pond  margins.  In  con- 
trast, White-fronted  Geese  prefer  hummock 
habitat  to  nest  in,  and  their  territorial  behavior 
appears  much  weaker  (R.G.B.  and  S.C.,  pers. 
obs.). 

Potential  for  competition. — Interspecific 
competition  for  habitat  and  food  in  spring  is 
believed  to  be  minimal  in  geese  because  spe- 
cies usually  segregate  either  spatially  (differ- 
ent ranges:  Bellrose  1980,  Ebbinge  et  al. 
1982,  Owen  and  Black  1990;  different  feeding 
habitats:  Fox  et  al.  1992)  or  temporally  (dif- 
ferent timing  of  breeding:  Fox  et  al.  1992).  We 
documented  differences  in  the  feeding  ecolo- 
gy of  two  sympatric  goose  species  that  nest  in 
synchrony  and  differ  mostly  in  bill  moiphol- 
ogy,  slightly  in  body  mass,  and  in  incubation 
behavior.  We  did  not  examine  niche  shifts 
with  changes  in  the  relative  densities  of  each 
species  during  spring  (Madsen  and  Mortensen 
1987)  because  both  species  experienced  sim- 
ilar decreases  in  pair  densities  with  snow  melt. 
Nonetheless,  we  showed  that  both  species 
overlapped  in  habitat  use  but  each  differed  in 
its  preferred  feeding  technique  and  hence  in 
its  selection  of  plant  parts  (“niche  comple- 
mentarity”; see  Nudds  1992).  On  our  study 
area,  potential  for  both  intra-  and  interspecific 
competition  for  food  is  highest  before  snow 
melt,  when  food  availability  and  diversity  are 
low,  and  pair  density  is  high. 

Interactions  between  nutrient  requirements, 
foraging  ability  and  food  profitabilities  (hence 
habitat  selection:  Stephens  and  Krebs  1986)  in 
goose  species  with  similar  body  sizes  but  with 
different  bill  morphologies  wanant  further 
study. 

ACKNOWLEDGMENTS 

This  study  was  funded  by  the  Government  of  the 
Northwest  Territories,  Department  of  Resources,  Wild- 
life and  Economic  Development;  by  a grant  from  the 
National  Science  and  Engineering  Research  Council  of 
Canada  to  G.G.;  and  by  grants  to  S.C.  from  the  Fonds 
pour  la  Formation  des  Chercheurs  et  I’Aide  a la  Re- 
cherche, Ministere  de  I'Education  du  Quebec;  the  Ca- 
nadian Wildlife  Foundation;  the  Canadian  Wildlife 
Service,  Government  of  Canada;  Indian  and  Northern 
Affairs,  Canada;  Ducks  Unlimited  Canada,  Yellow- 
knife, NT;  and  the  Centre  d’Etudes  Nordiques,  Univer- 
sitc  Laval.  Quebec.  We  thank  Polar  Continental  Shelf 
Project  for  transportation  and  logistical  support,  and 
the  Umingmaktok  and  Ekaluktutiak  Hunters’  and  Trap- 
pers’ Associations  for  general  support.  Laboratory  sup- 


port was  provided  by  the  Departement  de  Biologic, 
Universite  Laval.  We  are  grateful  to  C.  Brodie,  B. 
Croft,  R.  DeKoster,  M.-J.  Hotte,  Y.  Richard,  and  D. 
Wilson  for  their  help  in  the  field  and  to  E.  Thibeault, 
for  his  help  in  the  laboratory.  We  thank  Drs.  J.  H. 
Bedard,  A.  D.  Fox,  J.  Huot,  E Cooke,  E.  G.  Cooch 
and  two  anonymous  reviewers  for  their  comments  on 
the  manuscript. 

LITERATURE  CITED 

Alisauskas,  R.  T.  and  C.  D.  Ankney.  1992a.  The  cost 
of  egg  laying  and  its  relationship  to  nutrient  re- 
serves in  waterfowl.  Pp.  31-61  in  Ecology  and 
management  of  breeding  waterfowl  (B.  D.  J.  Batt, 
A.  Afton,  M.  G.  Anderson,  C.  D.  Ankney,  D.  H. 
Johnson,  J.  A.  Kaldec,  and  G.  L.  Krapu,  Eds.). 
Univ.  of  Minnesota  Press,  Minneapolis. 
Alisauskas,  R.  T.  and  C.  D.  Ankney.  1992b.  Spring 
habitat  use  and  diet  of  mid-continent  adult  Lesser 
Snow  Geese.  J.  Wildl.  Manage.  56:43-54. 
Altmann,  j.  1974.  Observational  study  of  behavior: 

sampling  methods.  Behaviour  49:227—267. 
Ankney,  C.  D.  1977.  The  use  of  nutrient  reserves  by 
breeding  male  Lesser  Snow  Geese.  Can.  J.  Zool. 
55:1984-1987. 

Ankney,  C.  D.  1984.  Nutrient  reserve  dynamics  of 
breeding  and  molting  Brant.  Auk  101:361-370. 
Ankney,  C.  D.  and  C.  D.  MacInnes.  1978.  Nutrient 
reserves  and  reproductive  performance  of  female 
Lesser  Snow  Geese.  Auk  95:459-471. 

Barry,  T.  W.  1962.  Effects  of  late  seasons  on  Atlantic 
Brant  reproduction.  J.  Wildl.  Manage.  26:19-26. 
Bellrose,  F.  C.  1980.  Ducks,  geese,  and  swans  of 
North  America.  Stackpole  Books,  Harrisburg, 
Pennsylvania. 

Boismenu,  C.,  G.  Gauthier,  and  J.  Larochelle.  1992. 
Physiology  of  prolonged  fasting  in  Greater  Snow 
Geese.  Auk  109:51 1-521. 

Bolen,  E.  G.  and  M.  K.  Rylander.  1978.  Feeding 
adaptations  in  the  Lesser  Snow  Goose  (Anser  ca- 
enilescens).  Southwest.  Nat.  23:158-161. 

Bolton,  M,  P.  Monaghan,  and  D.  C.  Houston.  1993. 
Proximate  determination  of  clutch  size  in  Lesser 
Black-backed  Gulls:  the  roles  of  food  supply  and 
body  condition.  Can.  J.  Zool.  71:273-279. 
Bromley,  R.  G.  1984.  The  energetics  of  migration  and 
reproduction  of  Dusky  Canada  Geese  (Branta 
canadensis  occidentalis).  Ph.D.  diss.,  Oregon 
State  Univ.,  Corvallis,  Oregon. 

Bromley,  R.  G.,  D.  C.  Heard,  and  B.  Croft.  1995. 
Visibility  bias  in  aerial  surveys  relating  to  nest 
success  of  arctic  geese.  J.  Wildl.  Manage.  59:364- 
371. 

Bromley,  R.  G.  and  R.  L.  Jarvis.  1993.  The  energet- 
ics ol  migration  and  reproduction  of  Dusky  Can- 
ada Geese.  Condor  95:193-210. 

Budeau,  D.  a.,  j.  T.  Ratti,  and  C.  R.  Ely.  1991. 
Energy  dynamics,  foraging  ecology,  and  behavior 
of  prenesting  Greater  White-fronted  Geese.  J. 
Wildl.  Manage.  55:556-563. 

Calder,  W.  a.  1974.  Consequences  of  body  size  for 


Caniere  el  al.  • SPRING  FEEDING  BY  ARCTIC  GEESE 


179 


avian  energetics.  Publ.  Nullall  Ornilhol.  Club  15: 
86-144. 

Carriere,  S.  1996.  Ecologie  de  I’alimentation  printan- 
iere  chez  des  Anserini  {Auser  albifrons  et  Brantci 
canadensis)  nichant  en  sympatrie  dans  I’Arctique. 
Ph.D.  diss.,  Univ.  Laval,  Quebec. 

Choiniere,  L.  and  G.  Gauthier.  1995.  Energetics  of 
reproduction  in  female  and  male  Greater  Snow 
Geese.  Oecologia  103:379-389. 

CoocH,  E.  G.,  D.  B.  Lank,  R.  E Rockwell,  and  E 
Cooke.  1991.  Long-term  decline  in  body  size  in 
a Snow  Goose  population:  evidence  of  environ- 
mental degradation?  J.  Anim.  Ecol.  60:483-496. 

Cooke,  E,  C.  S.  Eindlay,  and  R.  F.  Rockwell.  1984. 
Recruitment  and  the  timing  of  reproduction  in 
Lesser  Snow  Geese  {Chen  caeritlescens  caerules- 
cens).  Auk  101:451-458. 

Dalhaug,  L.,  I.  M.  Tombre,  and  K.  E.  Erikstad. 
1996.  Seasonal  decline  in  clutch  size  of  the  Bar- 
nacle Goose  in  Svalbard.  Condor  98:42-47. 

Daniel,  W.  W.  1978.  Applied  nonparametric  statistics. 
Houghton  Mifflin  Company,  Boston,  Massachu- 
setts. 

Drent,  H.  R.  and  S.  Daan.  1980.  The  prudent  parent: 
energetic  adjustment  in  avian  breeding.  Ardea  68: 
225-252. 

Dyke,  A.  S.  and  L.  A.  Dredge.  1989.  Quaternary  ge- 
ology of  the  northwestern  Canadian  shield.  Pp. 
188-214  in  Quaternary  geology  of  Canada  and 
Greenland  (R.  J.  Fulton,  Ed.).  Geological  Survey 
of  Canada,  Ottawa,  Ontario. 

Ebbinge,  B.  S.,  a.  M.  St.  Joseph,  P.  Prokosch,  and 
B.  Spaans.  1982.  The  importance  of  spring  stag- 
ing areas  for  Arctic-breeding  geese,  wintering  in 
western  Europe.  Aquila  89:249-258. 

Findlay,  C.  S.  and  F.  Cooke.  1982.  Breeding  syn- 
chrony in  the  Lesser  Snow  Goose  (Anser  caeru- 
lescens  caendescens):  I.  Genetic  and  environmen- 
tal component  of  hatch  date  variability  and  their 
effects  on  hatch  synchrony.  Evolution  36:342— 
351. 

Eox,  A.  D.,  H.  Boyd,  and  R.  G.  Bromley.  1995.  Mu- 
tual benefits  of  associations  between  breeding  and 
non-breeding  White-fronted  Geese  Anser  albi- 
frons. Ibis  137:151-156. 

Fox,  A.  D.,  H.  Boyd,  and  S.  M.  Warren.  1992.  Spa- 
tial and  temporal  feeding  segregation  of  two  Ice- 
landic goose  species  during  the  spring  pre-nesting 
period.  Ecography  15:289-295. 

Fox,  A.  D.  AND  J.  Madsen.  1981.  The  pre-breeding 
behaviour  of  the  Greenland  White-fronted  Goose. 
Wildfowl  32:48-54. 

Ganter,  B.  and  F.  Cooke.  1996.  Pre-incubation  feed- 
ing activities  and  energy  budgets  of  Snow  Geese: 
can  food  on  the  breeding  ground  influence  fecun- 
dity? Oecologia  106:153-165. 

Gauthier,  G.  1993.  Feeding  ecology  of  nesting  Great- 
er Snow  Geese.  J.  Wildl.  Manage.  57:216—223. 

Gauthier,  G.  and  J.  Tardif.  1991.  Female  feeding  and 
male  vigilance  during  nesting  in  Greater  Snow 
Geese.  Condor  93:701—71 1. 


Gawlik,  D.  E.  and  R.  D.  Slack.  1996.  Comparative 
foraging  behavior  of  sympatrie  Snow  Geese, 
Greater  White-fonted  Geese,  and  Canada  Geese 
during  the  non-breeding  season.  Wilson  Bull.  108: 
154-159. 

Hall,  L.  S.,  P.  R.  Krausman,  and  M.  L.  Morrison. 
1997.  The  habitat  concept  and  a plea  for  standard 
terminology.  Wildl.  Soc.  Bull.  25:173—182. 

Holechek,  j.  L.  and  B.  Gross.  1982.  Training  needed 
for  quantifying  simulated  diets  from  fragmented 
range  plants.  J.  Range  Manage.  35:644-647. 

Jandel  Scientific  Software.  1993.  SigmaStat®*  sta- 
tistical software — user’s  manual.  Jandel  Scientific 
GmbH,  Erkrath,  Germany. 

Jarvis,  R.  L.  and  R.  G.  Bromley.  In  press.  Incubation 
behaviour  of  Richardson’s  Canada  Geese  (Branta 
canadensis  butchinsii)  on  Victoria  Island,  North- 
west Territories,  Canada.  Can.  Wildl.  Serv.  Occas. 
Pap. 

Johnson,  M.  K.  1982.  Frequency  sampling  for  micro- 
scope analysis  of  botanical  compositions.  J.  Range 
Manage.  35:541-542. 

King,  J.  R.  1973.  Energetics  of  reproduction  in  birds. 
Pp.  78-107  in  Breeding  biology  of  birds  (D.  S. 
Farner,  Ed.).  National  Academy  of  Science,  Wash- 
ington, D.C. 

Krapu,  G.  L.  and  K.  j.  Reinecke.  1992.  Eoraging  ecol- 
ogy and  nutrition.  Pp.  1-29  in  Ecology  and  man- 
agement of  breeding  waterfowl  (B.  D.  J.  Batt,  A. 
Afton,  M.  G.  Anderson,  C.  D.  Ankney,  D.  H. 
Johnson,  J.  A.  Kaldec,  and  G.  L.  Krapu,  Eds.). 
Univ.  of  Minnesota  Press,  Minneapolis. 

LeMaho,  Y.,  H.  Vu  Van  Kha,  H.  Koubi,  G.  Dewas- 
MES,  J.  Girard,  P.  Eerre,  and  M.  Gagnard.  1981. 
Body  composition,  energy  expenditure,  and  plas- 
ma metabolites  in  long-temi  fasting  geese.  Am.  J. 
Physiol.  241:E342-354. 

Lindholm,  a,  G.  Gauthier,  and  A.  Desrochers. 
1994.  Effects  of  hatch  date  and  food  supply  on 
gosling  growth  in  Arctic-nesting  Greater  Snow 
Geese.  Condor  96:898-908. 

Lindstrom,  a.  and  T.  Alerstam.  1992.  Optimal  fat 
loads  in  migrating  birds:  a test  of  the  time-mini- 
mization hypothesis.  Am.  Nat.  140:477-491. 

MacInnes,  C.  D.,  R.  a.  Davis,  R.  N.  Jones,  B.  C. 
Lieff,  and  a.  j.  Pakulak.  1974.  Reproductive  ef- 
ficiency of  McConnell  River  small  Canada  Geese. 
J.  Wildl.  Manage.  38:686-707. 

Madsen,  J.  and  C.  E.  Mortensen.  1987.  Habitat  ex- 
ploitation and  interspecific  competition  of  moult- 
ing geese  in  east  Greenland.  Ibis  129:25-44. 

Mainguy,  S.  K.  and  V.  G.  Thomas.  1985.  Compari- 
sons of  body  reserve  buildup  and  use  in  several 
groups  of  Canada  Geese.  Can.  J.  Zool.  63:1765- 
1772. 

Morse,  D.  H.  1990.  Eood  exploitation  by  birds:  some 
current  problems  and  future  goals.  Stud.  Avian 
Biol.  13:134-143. 

Neu,  C.  W.,  C.  R.  Byers,  and  J.  M.  Peek.  1974.  A 
technique  for  analysis  of  utilization-availability 
data.  J.  Wildl.  Manage.  38:541-545. 


180 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


Newton,  I.  1977.  Timing  and  success  of  breeding  in 
tundra-nesting  geese.  Pp.  3—26  in  Evolutionary 
ecology  (B.  Stonehouse  and  C.  Perrins,  Eds.). 
Univ.  Park  Press,  Baltimore,  Maryland. 

Nudds,  T.  D.  1992.  Patterns  in  breeding  waterfowl 
communities.  Pp.  540—567  in  Ecology  and  man- 
agement of  breeding  waterfowl  (B.  D.  J.  Batt,  A. 
Afton,  M.  G.  Anderson,  C.  D.  Ankney,  D.  H. 
Johnson,  J.  A.  Kaldec,  and  G.  L.  Krapu,  Eds.). 
Univ.  of  Minnesota  Press,  Minneapolis. 

Owen,  M.  1981.  Abdominal  profile — a condition  index 
for  wild  geese  in  the  field.  J.  Wildl.  Manage.  45: 
227-230. 

Owen,  M.  and  J.  M.  Black.  1990.  Waterfowl  ecology. 

Chapman  and  Hall,  New  York. 

Petersen,  M.  R.  1992.  Reproductive  ecology  of  Em- 
peror Geese:  annual  and  individual  variation  in 
nesting.  Condor  94:383-397. 

Prevett,  j.  R,  I.  E Marshall,  and  V.  G.  Thomas. 
1985.  Spring  foods  of  Snow  and  Canada  geese  at 
James  Bay.  J.  Wildl.  Manage.  49:558-563. 

Prop,  J.  and  J.  deVries.  1993.  Impact  of  snow  and 
food  conditions  on  the  reproductive  performance 
of  Barnacle  Geese  Branta  leucopsis.  Ornis  Scand. 
24:1 10-121. 

Raveling,  D.  G.  1978.  The  timing  of  egg  laying  by 
northern  geese.  Auk  95:294-303. 

R.aveling,  D.  G.  1979.  The  annual  cycle  of  body  com- 
position of  Canada  Geese  with  special  reference 
to  control  of  reproduction.  Auk  96:234-252. 
Robbins,  C.  T.  1993.  Wildlife  feeding  and  nutrition. 

Academic  Press,  San  Diego,  California. 

Rohwer,  E C.  1992.  The  evolution  of  reproductive 
patterns  in  waterfowl.  Pp.  486-539  in  Ecology 
and  management  of  breeding  waterfowl  (B.  D.  J. 
Batt,  A.  Afton,  M.  G.  Anderson,  C.  D.  Ankney, 
D.  H.  Johnson,  J.  A.  Kaldec,  and  G.  L.  Krapu, 
Eds.).  Univ.  of  Minnesota  Press,  Minneapolis, 
Minnesota. 


Rosenzweig,  M.  L.  1985.  Some  theoretical  aspects  of 
habitat  selection.  Pp.  517-539  in  Habitat  selection 
in  birds  (M.  L.  Cody,  Ed.).  Academic  Press,  Lon- 
don, U.K. 

Ryder,  J.  P.  1970.  A possible  factor  in  the  evolution 
of  clutch  size  in  Ross’s  Geese.  Wilson  Bull.  82: 
5-13. 

Sedinger,  j.  S.  and  P.  L.  Flint.  1991.  Growth  rate  is 
negatively  correlated  with  hatch  date  in  Black 
Brant.  Ecology  72:496-502. 

SoKAL,  R.  R.  AND  E J.  Rohlf.  1981.  Biometry.  W.  H. 
Freeman  and  Company,  New  York. 

Spaans,  B.,  M.  Stock,  A.  St  Joseph,  H.-H.  Bergmann, 
AND  B.  S.  Ebbinge.  1993.  Breeding  biology  of 
dark-bellied  Brant  Geese  Branla  h.  bernicla.  Polar 
Res.  12:117-130. 

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

Thomas,  D.  L.  and  E.  J.  Taylor.  1990.  Study  designs 
and  tests  for  comparing  resource  use  and  avail- 
ability. J.  Wildl.  Manage.  54:322-330. 

Thompson,  S.  C.  and  D.  G.  Raveling.  1987.  Incuba- 
tion behavior  of  Emperor  Geese  compared  with 
other  geese:  interactions  of  predation,  body  size, 
and  energetics.  Auk  104:707-716. 

Van  Horne.  B.  1983.  Density  as  a misleading  indi- 
cator of  habitat  quality.  J.  Wildl.  Manage.  47:893- 
901. 

WlELGOLASKl,  E E.,  L.  C.  Bliss,  j.  Svoboda,  and  G. 
Doyle.  1981.  Primary  production  of  tundra.  Pp. 
187-226  in  Tundra  ecosystems:  a comparative 
analysis  (L.  C.  Bliss,  O.  W.  Heal,  and  J.  J.  Moore, 
Eds.).  Cambridge  Univ.  Press,  Cambridge.  U.K. 

Winkler,  D.  W.  and  J.  R.  Walters.  1983.  The  deter- 
mination of  clutch  size  in  precocial  birds.  Curr. 
Ornithol.  1:329-356. 

Wypkkma,  C.  P.  and  C.  D.  Ankney.  1979.  Nutrient 
reserve  dynamics  of  Lesser  Snow  Geese  staging 
at  James  Bay,  Ontario.  Can.  J.  Zool.  57:213-219. 


Wilson  Bull..  I I 1(2),  1999,  pp.  181-187 


A TEST  OF  THE  CONDITION-BIAS  HYPOTHESIS  YIELDS 
DIFFERENT  RESULTS  FOR  TWO  SPECIES  OF 
SPARROWHAWKS  {ACCIPITER) 

EDNA  GORNEY,'  WILLIAM  S.  CLARK,^  AND  YORAM  YOM-TOV'-^ 


ABSTRACT. — The  determination  of  body  condition  of  birds  is  important  for  many  field  studies.  However, 
when  using  trapping  methods  based  on  food  as  a lure,  the  sample  of  trapped  birds  could  be  biased  toward 
individuals  in  poor  physical  condition.  We  provide  information  on  body  mass,  body  condition,  and  sex  and  age 
ratio  of  Levant  Sparrowhawks  (Accipiter  hrevipes)  and  Eurasian  Sparrowhawks  (Accipiter  nisns)  caught  in  Elat, 
southern  Israel,  during  spring  migration.  We  compared  physical  condition  of  birds  trapped  in  baited  traps  to 
physical  condition  of  birds  trapped  in  mist  nets  (no  bait).  The  body  mass  and  index  of  physical  condition  of 
migrating  Levant  Sparrowhawks  trapped  in  baited  traps  was  lower  than  birds  trapped  in  mist  nets.  By  comparison 
no  differences  were  detected  in  body  mass  and  condition  index  of  migrating  Eurasian  Sparrowhawks  caught  by 
the  different  trapping  methods.  The  differences  found  in  condition  of  Levant  Sparrowhawks  trapped  with  and 
without  food  support  the  predictions  of  the  condition-bias  hypothesis;  however,  data  from  the  Eurasian  Sparrow- 
hawk  do  not.  The  extent  to  which  biases  occur  may  be  different  even  for  closely  related  species.  Received  1 
Oct.  1998.  accepted  7 Jan.  1999. 


Using  food  to  trap  animals,  especially  pred- 
ators, is  probably  the  most  widespread  capture 
method  used  by  biologists.  The  capture  of 
birds  of  prey  involves  many  methods,  most  of 
which  use  small  birds  and  rodents  as  lures 
(Clark  1981,  Bloom  1987).  These  trapping 
methods  are  vulnerable  to  sampling  bias  be- 
cause hungry  birds  in  poor  condition  are  more 
likely  to  overcome  their  fear  of  entering  traps 
compared  to  birds  in  good  condition  (Weath- 
erhead  and  Greenwood  1981).  Considerable 
support  for  this  condition-bias  hypothesis 
comes  from  studies  of  songbirds  (Weather- 
head  and  Greenwood  1981,  Dufour  and 
Weatherhead  1991)  and  ducks  (Greenwood  et 
al.  1986,  Hepp  et  al.  1986,  Reinecke  and 
Shaiffer  1988,  Conroy  et  al.  1989).  We  know 
of  no  such  studies  conducted  on  birds  of  prey. 

Two  species  of  spaiTOwhawks  occur  regu- 
larly during  spring  migration  at  Elat,  Israel. 
The  Levant  Spanowhawk  {Accipiter  brevipes) 
breeds  in  Russia,  eastern  Europe,  and  the  Bal- 
kans. It  migrates  in  large  flocks  to  Africa  in 
September,  returning  in  May  (Cramp  and  Sim- 
mons 1980).  In  autumn  spanowhawks  pass 
over  northern  Israel  between  mid-September 
and  the  beginning  of  October;  up  to  41,700 
birds  were  counted  during  one  autumn  (Lesh- 
em  and  Yom-Tov  1996).  During  spring  they 


' Dept,  of  Zoology,  Tel  Aviv  Univ.,  Tel  Aviv  69978, 
Israel. 

- 7800  Dassett  Cl.,  Annandale.  VA.,  22003. 

' Corresponding  author;  E-mail;  yomlov@post.tau.ac.il 


return  through  southern  Israel  between  20 
April  and  the  beginning  of  May,  with  up  to 
49,800  birds  counted  during  one  season  (Clark 
et  al.  1986,  Shirihai  1996).  The  Eurasian  Spar- 
rowhawk  {Accipiter  nisus)  breeds  across  the 
Palearctic  region  (Newton  1986).  In  contrast 
to  the  Levant  Spanowhawk  only  a few  hun- 
dred Eurasian  Sparrowhawks  pass  through  Is- 
rael during  autumn  (Dovrat  1986,  Leshem  and 
Yom-Tov  1996)  and  spring  (Clark  et  al.  1986, 
Shirihai  1996). 

Physical  condition  during  spring  migration 
is  significant,  not  only  for  the  successful 
completion  of  the  migratory  journey,  but  also 
for  the  timing  of  amval  at  the  breeding 
grounds  and  the  condition  of  nutrient  reserves 
required  for  successful  breeding.  The  breeding 
success  of  birds  in  the  temperate  zone  typi- 
cally declines  as  the  breeding  season  progress- 
es. Body  condition  and  amount  of  nutrient  re- 
serves have  been  shown  to  affect  breeding 
success  in  a variety  of  species  (Korschgen 
1977,  Moss  and  Watson  1984,  Newton  1986). 
McLandress  and  Raveling  (1981)  demonstrat- 
ed the  importance  of  spring  nutrient  reserves 
for  migration  as  well  as  for  reproduction  and 
their  effect  on  clutch  size  in  Canada  Geese 
{Branta  canadensis).  Price  and  coworkers 
(1988)  suggested  that  the  ability  to  accumu- 
late sufficient  reserves  required  for  breeding 
may  prevent  the  evolution  of  progressively 
earlier  breeding  dates.  Spring  migrants  anive 
in  Elat  after  crossing  desert  areas  where  feed- 


181 


182 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2.  June  1999 


ing  opportunities  ai'e  probably  limited.  Thus, 
it  is  important  to  investigate  the  occunence  of 
condition  biases  which  may  go  undetected  and 
lead  to  misinteipretation  of  data  collected 
from  trapped  birds  (Weatherhead  and  Green- 
wood 1981,  Gomey  and  Yom-Tov  1994). 

In  conjunction  with  a spring  migration  rap- 
tor banding  project,  we  captured  Eurasian  and 
Levant  sparrowhawks  using  a variety  of  trap- 
ping methods.  Our  objective  was  to  test  the 
condition-bias  hypothesis  in  birds  of  prey  and 
to  compare  the  physical  condition  of  spring 
migrants  to  condition  of  pre-breeding  spai- 
rowhawks. 

METHODS 

Migrating  sparrowhawks  were  captured  as  part  of  a 
raptor  banding  project  during  five  consecutive  spring 
migration  seasons  between  March  and  May  in  1984- 
1988,  in  the  agricultural  areas  just  north  of  Elat 
(29°  33'  N 34°  55'  E),  Israel  (Clark  et  al.  1986).  Birds 
were  caught  in  mist  nets,  bal-chatri  traps,  and  bow  nets 
(Bloom  1987)  commonly  used  by  hawk  researchers  in 
North  America  and  conforming  to  regulations  applied 
there.  Only  bal-chatri  and  bow-nets  were  placed  on  the 
ground  using  food  as  a lure.  Three-meter  tall,  4 shelf 
mist  nets  were  placed  in  palm  groves  along  rows  of 
trees  used  for  roosting  by  sparrowhawks.  The  distance 
between  the  lured  traps  and  mist  nets  was  generally  at 
least  500  m.  Occasionally  small  birds  were  caught  in 
mist  nets  and  they  may  have  lured  some  sparrowhawks 
into  the  nets.  Such  cases  were  excluded  from  the  data 
analysis  when  detected. 

All  birds  were  measured,  weighed  to  the  nearest  I 
g,  and  banded  before  release.  Wing  chord  (straight  line 
from  wrist  to  wing  tip,  without  stretching  or  flattening 
the  wing)  was  measured  with  a ruler  to  the  nearest  I 
mm.  Culmen  length  was  measured  to  0.1  mm  using 
Manostat  calipers.  Age  and  sex  were  determined  by 
plumage,  eye  color,  and  size  (Cramp  and  Simmons 
1980).  Age  was  determined  as  either  adult  (two  years 
old  or  more)  or  yearling  (approximately  one  year  old, 
i.e.,  birds  in  their  second  calendar  year). 

We  compared  the  body  mass  and  body  condition 
index  (body  mass/wing  chord  X culmen  length)  of  Le- 
vant and  Eurasian  sparrowhawks  that  were  caught  in 
mist  nets  to  those  that  were  caught  in  baited  traps  (bal- 
chatri  and  bow-nets).  Comparisons  were  conducted  on 
four  groups;  adult  and  yearling  females,  and  adult  and 
yearling  males.  Sample  sizes  vary  slightly  because  we 
lack  measurements  for  several  individuals.  This  partic- 
ular physical  condition  index  was  used  because  of  its 
high  correlation  with  fat  reserves  in  migrating  Steppe 
Buzzards  (fiuteo  huteo  vulpinii.'i;  Corney  and  Yom-Tov 
1994).  We  found  similar  results  however,  using  the 
more  commonly  employed  index  of  body  mass  divided 
by  wing  length. 

To  test  whether  birds  in  poorer  body  condition  arc 
more  likely  to  be  attracted  to  baited  traps  than  birds  in 


better  condition,  we  compared  for  each  age  and  sex 
category  the  body  mass  and  condition  index  of  birds 
trapped  with  baited  traps  and  birds  trapped  in  mist  nets 
(no  bait).  We  also  compared  the  number  of  individuals 
from  each  age  and  sex  class  that  were  trapped  in  baited 
and  non-baited  traps. 

We  used  Abstat  and  Minitab  6.1.1  programs  for 
IBM  compatable  computer  for  statistical  analysis.  All 
variables  were  tested  for  normality.  When  assumptions 
were  met  we  used  separate  SD  /-tests.  If  normality 
assumptions  were  not  met  we  used  Mann-Whitney 
tests.  All  probabilities  are  two-tailed. 

RESULTS 

Levant  Sparrowhawk. — Although  Levant 
Sparrowhawks  pass  through  Elat  during  thi'ee 
weeks,  we  captured  90%  of  all  age  and  sex 
categories  within  a mean  of  9.4  (yearling 
males)  and  10.5  (adult  females)  days  (means 
calculated  for  5 years  of  the  study).  Although 
the  passage  for  this  species  is  concentrated 
and  rapid,  adults  were  trapped  earlier  than 
yearlings  with  median  capture  dates  of  24 
April  for  adult  males,  25  April  for  adult  fe- 
males and  26  April  for  yearling  males  and  fe- 
males. We  caught  425  Levant  SpaiTowhawks, 
404  of  which  we  sexed  and  aged.  Slightly 
more  yearlings  were  trapped  (52%)  than 
adults  (48%)  at  Elat.  Sex  ratio  deviated  from 
the  expected  1 : 1 ratio  in  favor  of  males  in 
both  adults  (61.1%  males,  38.9%  females;  n 
= 193;  X'  = 9.58,  P = 0.002),  and  yearlings 
(56.9%  males,  43.1%  females;  n = 211;  x'  = 
3.98,  P = 0.046). 

Body  masses  of  adult  males  and  females 
were  significantly  larger  than  those  of  yeai- 
lings  (Table  1 ).  We  found  no  significant  dif- 
ferences between  condition  indices  of  adult 
and  yearling  birds  (Table  1).  Thus,  much  of 
the  difference  in  mass  appeared  to  be  due  to 
the  smaller  size  of  yearling  birds  (Table  1). 
We  also  found  no  significant  associations 
within  age  and  sex  groups  between  date  of 
migration  and  physical  condition,  nor  did  the 
physical  condition  of  trapped  birds  vary  with 
time  of  day. 

Mean  condition  index  of  adult  females  was 
significantly  higher  than  condition  index  of 
adult  males  (Mann-Whitney  U = 1862,  P = 
0.001,  two-tailed  test).  Similarly  yearling  fe- 
males were  in  better  condition  than  yearling 
males  (Mann-Whitney  U = 1740,  P = 0.001, 
two-tailed  test). 

We  found  no  significant  differences  be- 


Gornex  el  at.  • BODY  CONDITION  OF  SPARROWHAWKS 


183 


c 

crj 

> 

IT) 

IT) 

(U 

o 

o 

0. 

d 

d 

V 

V 

c 

in 

-o 

— 

d 

X 

(U 

to 

• S 

E 

c 

00 

in 

0 

ON 

00 

IT) 

WJ 

•3 

c 

c 

o 

~ 

ca 

a; 

U 

r-- 

r-- 

rv 

o 

o 

o 

o 

rri 

o 

o 

o 

o 

•a 

c 

d 

d 

d 

d 

03 

w 

D 

(N 

in 

■o 

§ 

O 

vO 

VO 

o 

o 

o 

O 

O 

c 

d 

d 

d 

W) 

C 

o 

o 

a. 

o 

o 

c 

d 

d 

£ 

V 

V 

d 

X 

On 

it 

•D 

u 

B 

d 

d 

0 

E 

— 

— 

-C 

o 

•o 

Cifj 

£ 

c 

'■j 

00 

o 

o 

o 

01} 

s: 

— 

rj 

ON 

c 

— 

— 

C/D 

C/D 

C3 

P 

Q 

ITj 

lO 

c/5 

■o 

o 

£ 

C 

r-* 

V 

■o 

o 

— 

o 

(N 

s 

n 

ri 

Ol 

(N 

c 

c 

o 

— 

• "" 

O 

O 

"O 

c 

o. 

o 

o 

o 

d 

d 

o 

V 

V 

■o 

q 

-t 

Tt- 

£ 

Oii 

■o 

u 

00 

00 

— 

00 

o 

c^ 

00 

o 

s: 

•o 

s; 

— 

ri 

00 

o 

Os 

G 

£ 

OJ) 

1 

1 

-i- 

00 

c^ 

Q 

o 

C>0 

— 

— 

ri 

n 

cz 

C/5 

UJ 

C/5 

03 

c 

f3 

o^ 

tT) 

ON 

c 

U 

a 

— 

ON 

C 

— 

— 

ri 

— 

>> 

*n 

OX) 

o 

3 

CQ 

03 

U 

R 

c/5 

o 

03 

f— 

— 

c. 

03 

C 

03 

jr 

3 

03 

R 

UJ 

O 

03 

C 

— 

£ 

nJ 

o 

t 

O 

OX) 

OXJ 

CQ 

< 

c 

£ 

9— 

c 

in 

CZ 

Q_ 

3 

■a 

C3 

(U 

3 

■o 

03 

(U 

C/0 

< 

>“  < >■ 

tween  males  and  females  (x"  — 0.313,  P > 
0.05,  df  = 1 ) nor  between  adults  and  yearlings 
(X“  = 0.024,  P > 0.05,  df  = 1 ) in  proportion 
of  birds  trapped  in  baited  and  non-baited 
traps.  The  mean  body  mass  and  mean  condi- 
tion index  of  adult  males  and  yearling  females 
trapped  in  mist  nets  was  higher  than  the  mean 
body  mass  and  condition  index  of  birds 
trapped  in  baited  traps  (P  < 0.05;  Table  2). 
Adult  females  and  yearling  males  demonstrat- 
ed no  difference  in  mean  condition  index  be- 
tween both  kinds  of  traps  (Table  2). 

Eurasian  Sparrowhawks. — Most  Eurasian 
Spanowhawks  (90%)  pass  through  Elat  in  five 
weeks.  As  with  Levant  Sparrowhawks,  adults 
were  trapped  before  yearlings.  Median  capture 
date  was  14  April  for  adult  males  and  females, 
20  April  for  yearling  females  and  22  April  for 
yearling  males.  We  trapped  72  Eurasian  Spar- 
rowhawks, with  slightly  more  yearlings  (53%) 
than  adults  (47%).  Sex  ratio  among  adults  (n 
= 35;  43%  males,  75%  females),  did  not  de- 
viate from  the  expected  1:1  ratio  (x'  = 1.48, 
P > 0.05),  but  did  deviate  among  young  birds 
in  favor  of  females  (n  = 37;  27%  males,  73% 
females;  X“  “ 7.81,  P = 0.005). 

Mean  body  mass  of  adult  females  was  sig- 
nificantly larger  than  that  of  yearling  females 
(P  < 0.05),  and  their  wing  chord  was  signif- 
icantly longer  (P  < 0.01;  Table  3).  However, 
the  large  body  mass  of  adult  females  was  not 
due  to  size  difference  alone  since  their  phys- 
ical condition  index  was  significantly  higher 
than  that  of  yearling  females  (P  < 0.05;  Table 
3).  We  also  found  no  significant  difference  in 
mean  body  mass  or  physical  condition  index 
between  adult  and  yearling  males  (Table  3). 
However,  the  wing  chord  of  adult  males  was 
significantly  longer  than  that  of  yearling  males 
(P  < 0.01;  Table  3).  Females  in  both  age 
groups  had  significantly  higher  condition  in- 
dices than  males  (Mann-Whitney  U = 13,  P 
< 0.001;  U = 65,  P ==  0.04,  for  adults  and 
yearlings,  respectively).  No  association  was 
found  between  index  of  physical  condition 
and  between  time  of  day  and  date  of  capture. 

We  found  no  significant  differences  in  body 
mass  and  condition  indices  between  Eurasian 
Spanowhawks  trapped  in  lured  traps  and  in 
mist  nets.  Significantly  more  females  (33  of 
48)  than  males  (6  of  18)  were  trapped  in  bait- 
ed traps  (x'  = 12.3,  P < 0.001,  df  = 1).  A 
similar  proportion  of  adults  (16  of  35)  and 


184 


THE  WILSON  BULLETIN  • VoL  III,  No.  2,  June  1999 


TABLE  2.  Comparison  of  body  mass  and  condition  index  of  Levant  Sparrowhawks  caught  in  mist  nets  (no 
lures)  and  in  lured  traps,  Elat  1984-1988. 


Mislnets 

Lured  traps 

Comparison 

n 

Mean 

SD 

n 

Mean 

SD 

f 

P 

Body  mass 

Adult  females 

19 

213 

27 

39 

219 

23 

0.74 

>0.05 

Adult  males 

35 

175 

18 

61 

167 

17 

2.06 

0.043 

Yearling  females 

22 

208 

16 

49 

197 

21 

2.42 

0.019 

Yearling  males 

29 

162 

19 

55 

157 

14 

1.14 

<0.05 

Mann-Whitney 


VT  P 


Condition  index 


Adult  females 

19 

0.065 

0.007 

39 

0.065 

0.007 

1 142 

>0.05 

Adult  males 

34 

0.064 

0.008 

60 

0.061 

0.006 

2568 

0.027 

Yearling  females 

21 

0.068 

0.005 

44 

0.063 

0.006 

1261 

0.008 

Yearling  males 

29 

0.062 

0.007 

54 

0.060 

0.006 

2124 

>0.05 

yearlings  (23  of  37)  were  trapped  in  baited 
traps  (x^  = 1.96;  P > 0.05,  df  = 1). 

DISCUSSION 

Physical  condition  of  migrants  trapped  with 
food  as  a lure  may  be  subject  to  biases 
( Weatherhead  and  Greenwood  1981).  One  of 
the  prominent  differences  we  found  between 
the  two  Accipiter  species  is  that  the  condition 
index  of  Levant  Sparrowhawks  captured  with 
and  without  food  as  a lure  supports  the  con- 
dition-bias hypothesis,  but  we  found  no  evi- 
dence for  a condition-bias  for  Eurasian  Spar- 
rowhawks. Thus,  the  occurrence  and  extent  of 
a condition-bias  may  be  different  even  for 
closely  related  species  of  approximately  the 
same  body  size  trapped  during  the  same  study 
using  the  same  traps.  Although  few  data  are 
available  to  determine  the  cause  of  this  dif- 
ference, several  aspects  of  their  migration 
strategies  may  be  pertinent.  Physical  condition 
of  birds  on  the  wintering  grounds,  and  food 
availability  could  differ  for  the  two  species.  In 
addition,  they  may  differ  in  their  tendencies 
to  hunt  during  migration.  Levant  SpaiTow- 
hawks  migrate  mainly  in  large  flocks  whereas 
Eurasian  Sparrowhawks  migrate  singly 
(Cramp  and  Simmons  1980,  Shirihai  1996). 
Levant  Sparrowhawks  might  not  regularly 
hunt  in  areas  like  Elat  where  many  birds 
would  be  unlikely  to  And  food.  Therefore  only 
weak  individuals  would  hunt  and  the  species 
would  demonstrate  a condition-bias.  On  the 
other  hand,  all  Eurasian  Sparrowhawks  may 


hunt  regularly  as  has  been  described  from  oth- 
er migrating  populations  and  species  (Rude- 
beck  1950,  Cochran  1972,  Newton  1986)  and 
thus  show  no  evidence  for  a condition-bias. 

When  baited  traps  are  used,  age  bias  can 
occur  with  more  younger  birds  trapped  than 
older  individuals  (Nass  1964).  Indeed,  more 
yearling  than  adult  Steppe  Buzzards  were 
trapped  in  Elat  during  the  same  years  (Gomey 
and  Yom-Tov  1994).  However,  we  detected  no 
age  bias  in  the  capturing  of  the  two  Accipiter 
species. 

Migrating  Eurasian  SpaiTowhawks  do  not 
appear  to  store  extra  fat  for  migration.  Moritz 
and  Vauk  ( 1976)  found  that  birds  trapped  dur- 
ing autumn  and  spring  migration  at  Helgoland 
weighed  the  same  as  non-migrating  individu- 
als caught  at  the  same  time.  Extra  fattening  is 
probably  unnecessary  because  spaiTowhawks 
in  some  regions  migrate  with  their  prey  spe- 
cies, and  thus  can  obtain  food  along  the  way 
(Rudebeck  1950,  Cochran  1972,  Newton 
1986).  The  spring  migrants  in  our  study 
weighed  considerably  less  than  birds  at  other 
times  of  year.  The  mean  body  masses  of  adult 
Eurasian  Sparrowhawks  in  our  study  were  33 
g (females)  and  14  g (males),  lower  than  the 
mean  for  European  birds  (Cramp  and  Sim- 
mons 1980)  and  20  g lower  for  both  males 
and  lemales  than  birds  from  the  former  Soviet 
Union  (Dementiev  1966).  The  mass  diffei  ence 
between  migrants  and  non-migrants  is  not  the 
result  of  size  differences  between  populations. 


Gorney  et  al.  • BODY  CONDITION  OF  SPARROWHAWKS 


185 


c 

t/5 

eg 

tn 

ro 

U 

3 

a. 

o 

o 

UJ 

d 

d 

A 

— 

C3 

C 

r- 

r- 

•o 

in 

c 

eg 

X 

u 

<u 

-o 

c 

eg 

f— 

c 

o 

O 

E 

o 

— 

(N 

(N 

COj 

c 

'■o 

c 

o 

u 

u 

eg 

NO 

NO 

r- 

NO 

0) 

Q 

o 

o 

o 

c/2 

o 

o 

o 

o 

T3 

d 

d 

d 

d 

C 

eg 

3 

uo 

CN 

o 

in 

IT) 

NO 

r- 

NO 

eg 

S 

o 

o 

o 

o 

O 

d 

d 

d 

d 

sz 

0^) 

c 

On 

o 

a. 

o 

o 

c 

OJ 

p 

d 

d 

u> 

"E 

CJ 

ON 

(N 

X 

1— 

«— 

n 

NO 

“O 

£ 

u 

B 

0 

sz 

-o 

o 

o 

o 

m 

OD 

(j 

— 

(N 

n 

c 

Oi) 

c 

5 

t/5 

c/2 

eg 

E 

SD 

NO 

NO 

NO 

■o 

o 

Si 

c 

w 

NO 

r- 

(N 

1) 

o 

On 

m 

X 

V 

•o 

O') 

n 

c 

o 

On 

a. 

O 

O 

TD 

C 

O 

d 

V 

d 

o 

"O 

c 

eg 

in 

c 

r- 

On 

00 

C 

in 

P 

00 

3 

V5 

■o 

u 

f3 

£ 

O 

00 

O 

o 

r- 

s 

00 

•o 

n 

n 

o 

c^ 

o 

CD 

oi:) 

1 

c 

00 

ON 

Q 

o 

n 

o 

— 

c/2 

— 

s 

E 

c/2 

PJ 

C/2 

eg 

c 

c 

o 

On 

p 

• 

i; 

ro 

cn 

w 

S 

ri 

sz 

■o 

Oij 

o 

3 

PQ 

eg 

u 

c/2 

zz. 

eg 

c 

rs 

PJ 

eg 

-C 

CL 

3 

2 

— 

eg 

eg 

£ 

E 

C 

u 

CQ 

< 

o 

fc 

o 

£ 

OD 

C 

OD 

C 

"u 

eg 

3 

eg 

3 

eg 

D- 

“O 

<u 

“O 

<u 

00 

< 

Mean  wing  length  of  adult  Eurasian  Sparrow- 
hawks  in  our  study  was  the  same  as  reported 
by  Cramp  and  Simmons  ( 1 980)  and  Demen- 
tiev (1966).  The  lower  body  mass  of  birds 
during  migration  likely  reflects  reserves  used 
for  migration  over  deserts  where  chances  for 
feeding  are  probably  minimal.  A similar  com- 
parison of  body  mass  of  Levant  Sparrow- 
hawks  during  migration  and  at  other  times  of 
year  was  not  possible  because  body  mass  data 
are  not  available  for  non-migration  periods. 
Levant  SpaiTowhawk’s  wing  chord  data  from 
our  study  are  similar  to  data  of  Cramp  and 
Simmons  (1980)  and  Dementiev  (1966),  also 
to  data  from  a 1996  study  conducted  in  Elat 
(Clark  and  Yosef  1997). 

Our  finding  that  Eurasian  Sparrowhawks 
migrating  during  spring  weigh  less  than  birds 
at  other  times  of  year  is  not  unique  to  this 
species.  The  mean  mass  of  spring  migrating 
Steppe  Buzzards  in  Elat  was  significantly  low- 
er than  for  the  same  species  on  their  wintering 
grounds  in  southern  Africa  (Gomey  and  Yom- 
Tov  1994).  Adult  Eleonora’s  Falcons  {Falco 
elenorae)  during  migration  weighed  100  g 
less  than  adults  on  their  breeding  grounds 
(Cramp  and  Simmons  1980).  Although  data 
were  not  available  for  Levant  Spairowhawks, 
it  is  likely  that  they  also  lose  weight  during 
migration.  In  general  birds  are  more  likely  to 
demonstrate  considerable  mass  declines  dur- 
ing spring  migration  in  Elat  because  they  must 
cover  large  stretches  of  desert  on  their  way 
from  Africa.  Upon  entering  the  Meditenanean 
area  north  of  Elat  these  species  probably  begin 
replenishing  their  reserves  (Yom-Tov  and 
Ben-Shahar  1995).  Indeed,  Levant  SpaiTow- 
hawks  in  autumn  were  not  attracted  to  the 
same  food  in  bal-chatri  traps  that  attracted 
them  during  spring  migration  (E.G.,  pers. 
obs.). 

Adults  migrate  in  spring  before  young  birds 
in  most  raptor  species  probably  because  of 
benefits  they  receive  from  early  airival  to  the 
breeding  tenitories  (Newton  1979,  Kerlinger 
1989).  Their  earlier  passage  is  undoubtedly 
aided  by  better  physical  condition.  In  support 
of  this,  several  researchers  found  larger  fat  re- 
serves in  adults  than  in  immature  birds  (Dunn 
et  al.  1988,  Serie  and  Shai*p  1989,  Alerstam 
and  Lindstrom  1990,  Morton  et  al.  1990,  Gor- 
ney and  Yom-Tov  1994;  but  see  Alerstam  and 
Lindstrom  1990).  Our  data  from  this  study  of 


186 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


two  Accipiter  species  also  demonstrate  earlier 
trapping  dates  of  adult  than  yearling  birds. 
However,  we  found  no  difference  in  physical 
condition  of  Levant  Sparrowhawks  between 
adults  and  yearlings.  In  addition,  Eurasian 
Sparrowhawk  adult  females  were  in  signifi- 
cantly better  condition  than  yearling  females; 
however,  we  found  no  such  trend  among 
males. 

ACKNOWLEDGMENTS 

We  thank  K.  Titus,  B.  Millsap  and  one  anonymous 
reviewer  for  their  comments  on  a previous  draft  of  this 
paper.  We  thank  Y.  Leshem  for  his  help  and  encour- 
agement, B.  Millsap  for  useful  comments  on  the  man- 
uscript. and  N.  Paz  for  editorial  comments.  We  are  also 
indebted  to  many  tireless  field  workers  of  whom  we 
can  mention  but  a few:  M.  Britain,  K.  Duffy,  O.  Hat- 
zofe,  Z.  Labinger,  J.  Mason,  M.  McGrady,  C.  Mc- 
Intyre, R.  Parslow,  T.  Sbocbat,  and  C.  Scbultz.  Thanks 
to  Kibbutz  Elot  for  their  generous  hospitality.  This 
study  was  funded  by  the  Israel  Raptor  Information 
Center  of  the  Society  for  the  Protection  of  Nature  in 
Israel,  the  Inter-University  Ecological  Fund  of  the  Jew- 
ish National  Fund,  and  the  Elat  Ornithological  Center. 

LITERATURE  CITED 

Alkrstam,  T.  and  a.  Lindstrom.  1990.  Optimal  bird 
migration:  the  relative  importance  of  time,  energy, 
and  safety.  Pp.  331-351  in  Bird  migration  (E. 
Gwinner,  Ed.).  Springer- Verlag,  Berlin,  Germany. 
Bloom,  P.  H.  1987.  Capturing  and  handling  raptors. 
Pp.  99-124  in  Raptor  management  techniques 
manual  (B.  A.  Giron  Pendelton,  B.  A.  Millsap,  K. 
W.  Cline  and  D.  M.  Bird,  Eds.).  National  Wildlife 
Federation,  Washington,  D.C. 

Clark,  W.  S.  1981.  A modified  dho-gaza  for  use  at  a 
raptor  banding  station.  J.  Wildl.  Manage.  45: 
1043-1044. 

Clark,  W.  S.,  K.  Duhfy,  E.  Gorney,  M.  McGrady, 
AND  C.  SCHLfLTZ.  1986.  Raptor  ringing  at  Elat,  Is- 
rael. Sandgrouse  7:21-28. 

Clark,  W.  S.  and  R.  Yosef.  1997.  Migrant  Levant 
Sparrowhawks  (Accipiter  hrevipes)  at  Elat,  Israel: 
measurements  and  timing.  J.  Raptor  Res.  31:317— 
320. 

Cochran,  W.  W.  1972.  A few  days  of  the  fall  migra- 
tion of  a Sharp-shinned  Hawk.  Haw'k  Chalk  I I: 
39-44. 

Conroy,  M.  J..  G.  R.  Co.stanzo,  and  D.  B.  Stotts. 
1989.  Winter  survival  of  female  American  Black 
Ducks  on  the  Atlantic  coast.  J.  Wildl.  Manage.  53: 
99-109. 

Cramp.  S.  and  K.  E.  L.  Simmons.  1980.  The  birds  of 
the  western  Palearctic  Vol.  2.  Oxford  Univ.  Press, 
Oxford,  U.K. 

Dementiev,  G.  P.  1966.  Birds  of  the  Soviet  Union  Vol. 
I.  Israel  Program  for  Scientific  Translations.  Je- 
rusalem. Israel. 


Dovrat,  E.  1986.  Kafr-Qasim  cross  Samaria  raptor 
migration  survey,  autumn.  1985.  Torgos  5:28-60. 

Dufour,  K.  W.  and  P.  j.  Weatherhead.  1991.  A test 
of  the  condition-bias  hypothesis  using  Brown- 
headed Cowbirds  trapped  during  the  breeding  sea- 
son. Can.  J.  Zool.  69:2686—2692. 

Dunn,  P.  O.,  T.  A.  May,  M.  A.  McCollough,  and  M. 
A.  Howe.  1988.  Length  of  stay  and  fat  content  of 
migrant  Semipalmated  Sandpipers  in  eastern 
Maine.  Condor  90:824-835. 

Ginn,  H.  B.  and  D.  S.  Melville.  1983.  Moult  in  birds. 
British  Trust  for  Ornithology,  Hertfordshire,  U.K. 

Gorney,  E.  and  Y.  Yom-Tov.  1994.  Fat,  hydration, 
and  moult  of  Steppe  Buzzards  Buteo  buteo  vul- 
pinus  on  spring  migration.  Ibis  136:185-192. 

Greenwood,  H.,  R.  G.  Clark,  and  P.  J.  Weather- 
head.  1986.  Condition  bias  of  bunter-shot  Mal- 
lards. Can.  J.  Zool.  64:599-601. 

Hepp,  G.  R.,  R.  j.  Blohm,  R.  E.  Reynolds,  J.  E.  Hines, 
AND  J.  D.  Nichols.  1986.  Physiological  condition 
of  autumn  banded  Mallards  and  its  relationship  to 
hunting  vulnerability.  J.  Wildl.  Manage.  50:177- 
183. 

Kerlinger,  P.  1989.  Flight  strategies  of  migrating 
hawks.  Univ.  of  Chicago  Press,  Chicago,  Illinois. 

Korschgen,  C.  E.  1977.  Breeding  stress  of  female  ei- 
ders in  Maine.  J.  Wildl.  Manage.  41:360-373. 

Leshem,  Y.  and  Y.  Yom-Tov.  1996.  The  magnitude 
and  timing  of  migration  by  soaring  raptors,  peli- 
cans and  storks  over  Israel.  Ibis  138:188-203. 

McLandress  and  D.  G.  Raveling.  1981.  Changes  in 
diet  and  body  composition  of  Canada  Geese  be- 
fore spring  migration.  Auk  98:65-79. 

Moritz,  D.  and  G.  Vauk.  1976.  Der  Zug  des  Sperbers 
(Accipiter  nisiis)  auf  Helgoland.  J.  Ornithol.  117: 
317-328. 

Morton,  J.  M.,  R.  L.  Kirkpatrick,  and  M.  R. 
Vaughan.  1990.  Changes  in  body  composition  of 
American  Black  Ducks  wintering  at  Chinco- 
teague,  Virginia.  Condor  92:598-605. 

Moss,  R.  and  a.  Watson.  1984.  Maternal  nutrition, 
egg  quality  and  breeding  success  of  Scottish  Ptar- 
migan Lagopus  niutus.  Ibis  126:212-220. 

Nass,  R.  D.  1964.  Sex  and  age  ratio  bias  of  cannon- 
netted  geese.  J.  Wildl.  Manage.  28:522-527. 

Newton,  I.  1979.  Population  ecology  of  raptors.  T & 
A D Poy.ser  Ltd.,  Berkhamstcd,  U.K. 

Newton,  I.  1986.  The  Sparrowhawk.  T & A D Poyser 
Ltd.,  Waterhouses,  U.K. 

Price,  T,  M.  Kirkpatrick,  and  S.  J.  Arnold.  1988. 
Directional  selection  and  the  evolution  of  breed- 
ing date  in  birds.  Science  240:798-799. 

Reinecke,  K.  j.  and  C.  W.  Shaiffer.  1988.  A field  test 
lor  dillerences  in  condition  among  trapped  and 
shot  Mallards.  J.  Wildl.  Manage.  52:227-232. 

Rudebeck,  G.  1950.  The  choice  of  prey  and  modes  of 
hunting  ol  predatory  birds  with  special  reference 
to  their  selective  effect.  Oikos  2:65-88. 


Gonie\  et  al.  • BODY  CONDITION  OF  SPARROWHAWKS 


187 


Serie,  J.  R.  and  D.  E.  Sharp.  1989.  Body  weight  and 
composition  dynamics  of  fall  migrating  Canvas- 
backs.  J.  Wildl.  Manage.  53:431-441. 

Shirihai,  H.  1996.  The  birds  of  Israel.  Academic  Press 
Ltd.,  London,  U.K. 

Weatherhead,  P.  j.  and  H.  Greenwood.  1981.  Age 


and  condition  bias  of  decoy-trapped  birds.  J.  Field 
Ornithol.  52: 10-15. 

Yom-Tov,  Y.  and  R.  Ben-Shahar.  1995.  Seasonal 
body  mass  and  habitat  selection  of  some  migra- 
tory passerines  occurring  in  Israel.  Israel  J.  Zool. 
41:443-454. 


Wilson  Bull.,  111(2),  1999,  pp.  188-194 


THE  DEVELOPMENT  OF  A VOCAL  THERMOREGULATORY 
RESPONSE  TO  TEMPERATURE  IN  EMBRYOS  OF  THE 

DOMESTIC  CHICKEN 

SHAWN  C.  BUGDEN'  2 AND  ROGER  M.  EVANS' 


ABSTRACT. — We  examined  the  vocal  responsiveness  of  chicken  (Callus  gallu.s)  embryos  at  the  pipped  egg 
stage  to  determine  if  they  were  able  to  regulate  their  thermal  environment  by  soliciting  heat  from  a surrogate 
parent.  There  was  no  overall  effect  on  vocalizations  of  exposure  to  20°  C or  45°  C relative  to  the  normal 
incubation  temperature  of  37.8°  C.  There  was,  however,  a general  trend  towards  increased  calling  as  the  time  of 
hatching  approached.  There  was  also  some  indication  that  embryos  tested  in  the  late  stages  of  hatching  (ringing) 
vocalized  more  in  the  cold,  then  became  relatively  silent  when  rewarmed.  When  cold-challenged  embryos  were 
given  2 mm  of  rewamriing  (surrogate  brooding)  in  response  to  their  calls  body  temperature  was  slightly  but 
significantly  elevated  above  cold  only  exposed  controls.  Unlike  previously  reported  anecdotal  evidence  sug- 
gesting a strong  vocal  response  to  cold,  our  results  suggest  chicken  embryos  show  only  weak  incipient  vocal 
response  to  temperature  that  begins  to  increase  late  in  incubation  and  becomes  fully  functional  only  after  hatch- 
ing. Unlike  other  species  tested  to  date,  the  developmental  progression  of  behavioral  and  metabolic  thermoreg- 
ulation appear  to  be  tightly  linked  in  this  species.  Received  27  Feb.  1998,  accepted  5 Oct.  1998. 


The  embryos  of  many  avian  species  are  ca- 
pable of  vocalizations  prior  to  hatch  (Freeman 
and  Vince  1974;  reviewed  in  Evans  1988a). 
Embryonic  vocalizations  may  facilitate  the 
transition  from  incubating  eggs  to  brooding 
and  feeding  young  (Impekoven  1973;  Temple- 
ton 1983;  Evans  1988a,  b),  may  function  as 
antecedents  to  post-hatching  social  behavior 
(Tuculescu  and  Griswold  1983),  or  may  func- 
tion in  soliciting  care  from  parents.  The  em- 
bryos of  several  species  respond  to  cold  by 
increasing  vocalizations  (Evans  1990a;  Bug- 
den  and  Evans  1991;  Evans  et  al.  1994,  1995; 
Brua  et  al.  1996).  These  vocalizations  affect 
parental  behavior  and  may  elicit  more  atten- 
tive incubation  (Evans  1989,  1990b,  1992; 
Brua  1996). 

When  cold-induced  vocalizations  trigger  re- 
warming by  a surrogate  parent  in  altricial  pel- 
icans and  semiprecocial  gulls,  cold-challenged 
embryos  can  vocally  regulate  their  body  tem- 
peratures at  relatively  safe  levels  (Evans 
1990a;  Evans  et  al.  1994,  1995).  In  these  spe- 
cies such  vocal  behavioral  thermoregulation 
evidently  precedes  their  ability  to  thermoreg- 
ulate  endothermically;  altricial  and  semipre- 
cocial species  show  no  apparent  metabolic  re- 
sponse to  cooling  before  hatching  (Matsunaga 


' Dept,  of  Zoology,  Univ.  of  Manitoba,  Winnipeg, 
MB  R3T  2N2,  Canada. 

^ Current  address:  Mowbray  Research  Station, 
McElroy  Mouse,  645  Thornhill  .St.,  Morden,  MB  R6M 
IU4,  Canada. 


et  al.  1989,  Kuroda  et  al.  1990).  However,  in 
precocial  species  there  is  evidence  that  grad- 
ual cooling  of  late  stage  embryos  brings  about 
a small  but  measurable  incipient  endothermic 
response  (Ereeman  1964,  Tazawa  et  al.  1988, 
Kuroda  et  al.  1990).  The  relationship  between 
the  emergence  of  endothermy  and  the  timing 
of  a vocal  response  to  cold  in  precocial  spe- 
cies is  unknown.  We  examined  this  issue  in 
the  precocial  domestic  chicken  (Gallus  gal- 
lus). 

Chicken  embryos  are  capable  of  vocaliza- 
tions 2-3  days  prior  to  hatch,  but  these  vo- 
calizations become  more  frequent  during  the 
final  24  hours  before  hatching  (Gottlieb  and 
Vandenbergh  1968,  Dawes  1981,  Tuculescu 
and  Griswold  1983).  In  response  to  the  loud 
calls  (“distress  calls”,  Collias  1987)  of  em- 
bryos,  hens  will  vocalize  or  move  around  on 
the  nest  (Tuculescu  and  Griswold  1983).  Fol- 
lowing the  maternal  response  the  embryos  be- 
come silent  or  emit  soft  trill  calls  (“pleasure 
calls”,  Collias  1987). 

Study  of  the  vocal  response  of  chicken  em- 
bryos to  cold  has  produced  variable  results. 
Early  investigation  suggested  that  a pipped 
egg  that  was  alternately  cooled  and  warmed 
would  give  “di.stress”  calls  or  “pleasure” 
calls  in  close  conespondence  to  the  tempera- 
ture changes  (Collias  1952).  Subsequent  stud- 
ies have  shown  decreased  vocal  activity  (Op- 
penheim  and  Levin  1974),  increased  vocal  ac- 
tivity (Evans  1988a),  and  inconsistent  vocal 


188 


Buiiden  cind  Evans  • VOCAL  RESPONSE  OF  CHICKEN  EMBRYOS 


189 


responses  (Dawes  1981).  Some  evidence  also 
suggests  that  chicken  embryos  increase  their 
rate  of  vocalization  during  exposure  to  high 
temperatures  (Oppenheim  and  Levin  1974). 

This  study  was  designed  to  systematically 
reassess  the  vocal  response  of  late  stage  chick- 
en embryos  to  temperature.  We  examined  the 
effects  of  temperatures  both  above  and  below 
the  normal  incubation  temperature,  and  the  ef- 
fectiveness of  call-induced  rewarming  bouts 
in  the  regulation  of  temperature  during  cold 
challenge. 

METHODS 

White  Leghorn  Chicken  eggs  were  incubated  in  a 
forced  air  commercial  poultry  incubator  (Perersime 
Model  no.  1 ) that  maintained  conditions  within  a suit- 
able range  (37.8  ± 0.5°  C and  65  ± 5%  relative  hu- 
midity). Only  externally  pipped  eggs,  which  are  known 
to  be  capable  of  vocalizations  (Tuculescu  and  Gris- 
wold 1983),  were  selected  for  study. 

Effects  of  continuous  chilling  and  heating. — To  ex- 
amine the  vocal  response  of  embryos  to  low  or  high 
temperature,  pipped  eggs  were  placed  singly,  pip  hole 
up,  within  an  environmental  chamber  that  consisted  of 
a coil  of  copper  tubing  surrounded  by  insulating  Styro- 
foam. The  temperature  in  the  chamber  was  controlled 
by  pumping  water  from  controlled  water  baths 
(±0.5°  C)  through  the  coil  surrounding  the  egg.  Testing 
began  with  a 10  min  pre-test  at  the  control  temperature 
(37.8°  C).  The  coil  temperature  was  then  changed  to 
the  experimental  (20°  C or  45°  C)  or  left  at  the  control 
temperature  (37.8°  C)  for  30  minutes.  This  was  fol- 
lowed by  a 10  min  post-test  period  at  37.8°  C. 

The  body  temperature  of  the  embryo  was  measured 
with  a thermocouple  placed  approximately  I cm  di- 
rectly into  the  pip  hole.  The  thermocouple  was  sur- 
rounded by  deep  lying  portions  of  the  embryo’s  body, 
away  from  the  outer  shell.  It  was  held  in  place  by 
porous  adhesive  tape  (“Micropore”)  applied  to  the  ex- 
terior of  the  shell.  The  body  temperature  of  the  embryo 
and  the  coil  temperature  were  recorded  to  the  nearest 
0.1°  C every  30  s by  a data  logger  [Grant  Instrument 
(Cambridge)  Model  1203).  Calls  with  a minimum  in- 
tensity of  78  dB  (2.5  cm  from  the  pip  hole,  B-fast 
scale)  were  recorded  by  a microphone  set  in  the  plex- 
iglass lid  of  the  chamber,  connected  to  a sound  oper- 
ated relay  and  an  Esterline  Angus  event  recorder. 

Vocal  regulation  of  temperature  during  cold  chal- 
lenge.— The  apparatus  was  similar  to  that  described 
above  except  that  calling  of  the  embryo  triggered  a 
period  of  rewarming  (illustrated  in  Evans  1990a).  The 
embryo  faced  a continuous  cold  challenge  at  20°  C un- 
til 5 calls  were  given.  The  fifth  call  then  triggered  a 2 
min  period  of  rewarming  with  water  at  37.8°  C being 
pumped  through  the  coil  surrounding  the  egg.  This  pe- 
riod of  rewarming  was  followed  by  a return  to  default 
chilling  at  20°  C until  another  bout  of  calling  was  ini- 
tiated. If  an  embryo  called  in  response  to  each  succes- 


sive period  of  cold  challenge  it  would  in  elTect  be  ca- 
pable of  regulating  ambient,  and  hence  body  temper- 
ature (Evans  1990a,  Bugden  and  Evans  1997).  Control 
embryos  were  placed  in  the  same  apparatus  and  held 
at  a constant  37.8°  C throughout.  Calls  in  the  control 
situation  triggered  a mock  warming  bout  where  the 
same  timer  and  pumps  were  activated  as  in  the  cold 
challenge  situation  but  the  water  circulating  through 
the  coil  remained  at  37.8°  C.  All  temperatures  were 
recorded  as  in  the  first  experiment.  An  Esterline  event 
recorder  recorded  both  individual  calls  and  the  warm- 
ing and  mock  warming  bouts.  Control  and  cold  chal- 
lenge tests  lasted  for  1 hour  on  separate  samples  of 
eggs. 

To  determine  when  during  the  pip-to-hatch  interval 
the  voeal  response  to  cold  might  develop,  the  timing 
of  pipping,  hatching,  and  testing  were  reeorded  at  4 h 
intervals.  Short  term  exposure  of  pipped  eggs  to  mod- 
erate cold  can  delay  hatching  in  domestic  chickens 
(Evans  1990c).  Testing  of  the  eggs  at  20°  C in  this 
experiment  thus  could  potentially  affect  the  timing  of 
their  hatching  and  so  distort  the  interpretation  of  the 
developmental  onset  of  the  vocal  response  to  cold.  To 
control  for  this  possibility,  the  pip-to-hatch  intervals  of 
cold  challenged  and  control  embryos  were  compared 
with  a separate  sample  of  embryos  (untested  control 
embyros)  that  were  not  tested  and  left  to  hatch  nor- 
mally in  the  incubator.  Statistical  tests  were  done  with 
STATISTIX  (version  4.1,  Analytical  Software,  IBM 
platform). 

RESULTS 

Effects  of  continuous  chilling  and  heat- 
ing.— Seven  chicken  embryos  exposed  to  a 30 
minute  period  of  chilling  at  20°  C experienced 
a fall  in  body  temperature  of  8.0  ± 0.3°  C 
(mean  ± SE)  from  36.9  ± 0.2°  C at  the  start 
to  28.8  ± 0.3°  C at  the  end  of  the  exposure 
period.  These  embryos  had  a mean  calling  rate 
of  10.2  ± 6.8  calls  per  minute.  This  result  was 
skewed  by  two  highly  vocal  embryos  that 
were  nearly  hatched  (ringing  stage,  Ereeman 
and  Vince  1973)  by  the  end  of  the  test.  The 
remaining  five  embryos  were  completely  si- 
lent during  their  exposure  to  cold,  resulting  in 
a median  call  rate  of  0.0  calls  per  minute.  The 
body  temperature  of  embryos  held  at  the  con- 
trol temperature  of  37.8°  C shifted  by  0.5  ± 
0.1°  C,  from  37.0  ± 0.2°  C to  37.5  ± 0.2°  C. 
None  of  the  7 control  embryos  reached  the 
ringing  stage  and  all  were  relatively  quiet  dur- 
ing the  test  period  with  a mean  calling  rate  of 
0.1  ± 0.04  (median  of  0.1)  calls  per  minute. 

Body  temperature  of  seven  embryos  ex- 
posed to  45°  C rose  by  an  average  of  4. 1 ± 
0.3°  C,  from  36.3  ± 0.2°  C to  40.4  ± 0.4°  C. 
Call  rate  averaged  1.2  ± 0.7  calls  per  minute 


190 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


with  a median  of  0.5  calls  per  minute.  While 
2 of  these  embryos  were  in  the  ringed  stage, 
their  call  rates  (1.7  and  0.5  calls/min)  were 
similar  to  the  overall  average  call  rate  for  this 
group.  A Kruskal-Wallis  one  way  nonpara- 
metric  ANOVA  showed  no  overall  differences  ^ 
in  calling  rates  of  chicken  embryos  at  the  3 
three  temperatures  (H  = 4.14,  P > 0.05,  df  = m 
2).  b 

Vocal  regulation  of  temperature  during  2 
cold  challenge. — Thirty-three  embryos  were  < 
tested  under  cold  challenge  experimental  con-  ^ 
ditions  (20°  C)  and  34  were  tested  as  warm-  ° 
only  (37.8°  C)  controls.  There  was  no  signif-  ^ 
leant  difference  in  the  time  from  pipping  to  § 
the  time  of  hatching  in  cold  challenged,  con-  ^ 
trol,  and  12  untested  control  embryos  (One  < 
way  ANOVA:  7^2,76  = 1-34,  P > 0.05).  For  ^ 
additional  analyses  embryos  were  grouped  ac- 
cording to  time  between  testing  and  hatching. 
There  were  8,  13,  and  12  embryos  tested  un- 
der cold  challenge  and  13,  12,  and  9 embryos 
tested  as  warm  only  controls  for  three  devel- 
opmental categories  0-2.5,  2. 5-7. 5 and  >7.5 
h before  hatching.  The  vocal  response  of  cold 
challenged  embryos  was  greater  than  that  of 
the  warm  only  controls  in  all  developmental 
categories  (Fig.  1).  However,  a priori  two- 
sample  t-tests  between  cold  experimental  and 
warm  controls  showed  no  significant  differ- 
ence in  any  of  the  developmental  categories 
(0-2.5:  t = 0.93,  P > 0.05;  2.5-1. 5\  t = 0.44, 

P > 0.05;  >7.5:  t = 1.63,  P > 0.05). 

Within  2.5  h of  hatching,  the  mean  number 
of  warming  or  mock  warming  bouts  increased 
in  both  the  cold  challenged  embryos  and  in 
the  warm  only  controls  (Fig.  1 ).  Since  cold 
exposure  produced  no  significant  difference  in 
vocal  response,  all  data  were  combined  for 
further  comparison  of  vocal  response  with  re- 
spect to  time  before  hatch.  Embryos  vocalized 
significantly  more  frequently  as  the  time  of 
hatching  approached  (Kj'uskal-Wallis  ANO- 
VA: H = 1 1.10,  P < 0.01,  df  = 2). 

The  pattern  of  calling  during  the  2 min  re- 
warming and  mock  rewarming  bouts  provided 
an  additional  measurement  of  the  response  to 
temperature  (Table  1 ).  Because  of  a 6 s delay 
of  the  chamber  to  temperature  changes,  cold 
challenged  embryos  were  still  experiencing 
temperatures  well  below  the  incubation/con- 
trol temperature  of  37.8°  C at  the  start  of  each 
rewarming  period.  Cold  challenged  embryos 


LIG.  1.  Mean  (±SE)  number  of  vocally  generated 
warming  (cold  challenged)  and  mock-warming  (con- 
trol) bouts  at  three  pipped-egg  developmental  stages 
(cold  challenge  n equals  8,  13,  and  12;  warm-only  con- 
trol n equals  13,  12  and  9 for  0-2.5,  2. 5-7. 5 and  >7.5 
h before  hatching  respectively). 


in  the  two  groups  vocalized  more  often  than 
controls  at  this  time,  significantly  so  in  the 
2. 5-7. 5 h age  group  (Table  1).  By  the  final 
minute  of  rewarming,  the  cold  challenged  em- 
bryos were  almost  silent,  but  this  was  also 
true  for  most  of  the  control  embryos  whose 
temperatures  had  not  changed.  Control  em- 
bryos that  were  less  than  2.5  hours  from  hatch 
were  an  exception.  These  embryos  vocalized 
at  a relatively  high  rate  during  the  final  minute 
of  the  rewarming  bout  and  maintained  a call 
rate  significantly  greater  than  the  experimental 
embryos. 

Body  temperatures  maintained  by  experi- 
mental embryos  during  vocal  regulation  tests 
increased  by  about  1°C  as  hatching  time  ap- 
proached, but  this  increase  was  not  statistical- 
ly significant  (Fig.  1,  Table  2).  Body  temper- 
atures of  experimentals  were  significantly 
lower  than  those  of  control  embryos  held  at 
37.8°  C (Table  2),  reflecting  the  general  low 
level  of  call-induced  rewarming  periods  in  the 
experimentals.  Body  temperature  of  the  ex- 
perimental embryos  by  the  end  of  30  min  of 
testing  was  significantly  higher  (29.9  ± 


Bit}>den  and  Evans  • VOCAL  RESPONSE  OF  CHICKEN  EMBRYOS 


191 


TABLE  1.  Median  number  of  calls  per  minute  given  by  chicken  embryos  in  pipped  eggs  at  the  start  and 
end  of  vocally-generated  rewarming  (experimental)  and  mock  rewarming  (control)  bouts.  Listed  are  medians  of 
calls  given  per  embryo  per  bout  (1st  and  3rd  quartiles  in  parentheses). 


Stage 
(h  before 
hatch) 

Eirst  20  s 

Final  minute 

Experimental 

Control 

Experimental 

Control 

<2.5 

9.6  (6.1-1  1.8) 

6.6  (2.0-17.4) 

0.2  (0.0-2.2) 

3.1  ( 1.0-8. 7)“ 

/i  = 8 

/;  = 12 

/;  = 8 

/;  = 12 

2.5-7.5 

9.0  (4.7-13.1) 

3.0  (0.0-6.9)^ 

0.0  (0.0-L4) 

0.5  (0.0-2. 3) 

n = 12 

/;  = 1 1 

n = 12 

/7  = 1 1 

>7.5 

3.0  (1.7-5. 3) 

3.0  (0.0-6.0) 

0.0  (0.0-0.5) 

0.0  (0.0-0.5) 

;t  = 8 

n = 1 

/t  = 8 

n = 1 

7.03 

3.61 

1.13 

10.67 

df 

2 

2 

2 

2 

P 

<0.05 

>0.05 

>0.05 

<0.01 

^ Experimenlals  and  controls  differ  significantly  (P  < 0.05:  Mann-Whitney  U test). 
H.  Kruskal-Wallis  ANOVA  statistic,  distributed  as  x". 


0.4°  C,  all  stages  combined;  t = 2.28,  P < 
0.05,  df  = 23)  than  embryos  that  were  ex- 
posed to  constant  chilling  for  30  min  in  ex- 
periment 1 (28.8  ± 0.3°  C).  This  suggests  a 
slight  warming  effect  of  vocalizations  in  ex- 
periment 2. 

DISCUSSION 

Exposure  of  chicken  embryos  to  continuous 
cold  (20°  C)  and  continuous  hot  (45°  C)  en- 
vironments did  not  significantly  increase  their 
rates  of  vocalization.  The  embryos  remained 
relatively  silent  in  spite  of  exposure  to  envi- 
ronmental temperatures  that  altered  body  tem- 
perature to  a level  which,  if  continued,  would 
be  expected  to  result  in  death  of  the  embryo 
(Webb  1987).  While  the  embryos  were  clearly 
capable  of  vocalizing,  the  close  correspon- 


dence of  calling  and  temperature  suggested  in 
the  literature  (e.g.,  Collias  1952)  was  not  ev- 
ident. Only  two  cold  challenged  embryos,  in 
the  process  of  ringing  prior  to  hatching, 
showed  a strong  vocal  response.  The  vocal 
regulation  experiments  also  showed  an  in- 
crease in  vocal  response  in  chilled  embryos 
that  were  near  to  hatching  (Fig.  1)  but  this 
trend  was  also  seen  in  control  embryos  that 
were  not  exposed  to  cold.  Tuculescu  and  Gris- 
wold (1983)  have  also  noted  a general  in- 
crease in  the  rate  of  vocalization  in  the  few 
hours  just  prior  to  hatching. 

Despite  increased  vocalization  rates  as 
hatching  approached,  cold  challenged  chicken 
embryos  in  the  vocal  regulation  appaiatus 
were  not  able  to  elevate  their  body  tempera- 
tures to  safe,  near  normal  incubation  temper- 


TABLE  2.  Median  body  temperature  of  chicken  embryos  during  experimental  and  control  vocal  regulation 
test.  Listed  are  medians  of  the  average  body  temperature  maintained  during  the  one-hour  test  period  (1st  and 
3rd  quartiles  in  parentheses). 


Body  temperature  of  embryos  (°C) 

(h  before  hatch) 

Experimental 

Control 

ph 

<2.5 

31.5  (30.0-35.5) 

37.8  (37.2-38.0) 

<0.002 

n = 1-' 

n = 7“ 

2.5-1.5 

30.5  (30.0-31.0) 

38.1  (37.8-38.3) 

<0.001 

n = 13 

/)  = 12 

>7.5 

30.4  (29.4-30.1) 

37.6  (36.9-38.1) 

<0.001 

n = 12 

n = 9 

df 

2.79 

4.61 

P 

2 

2 

>0.05 

>0.05 

“ E.xcludes  I experimenlal  and  6 control  embryos  that  displaced  the  thermocouple  during  ringing. 
Comparison  of  experimentals  and  controls  at  each  stage — Mann-Whitney  U test. 

H,  Kruskal-Wallis  ANOVA  statistic,  distributed  as  x~- 


192 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


ature,  at  least  under  the  conditions  employed 
here  (Table  2).  This  contrasts  strikingly  with 
a significant  increase  in  vocal  regulatory  ca- 
pability of  hatched  domestic  chicks  given  2 
min  periods  of  rewarming  in  response  to  cold- 
induced  calling  on  the  day  of  hatching  (Bug- 
den  and  Evans  1997).  The  finding  that  body 
temperature  after  30  min  of  testing  of  vocal 
regulation  in  20°  C cold  challenged  embryos 
was  significantly  different  than  cold  only  ex- 
posed embryos  in  experiment  1 is  consistent 
with  the  presence  of  some  incipient  effects  of 
vocally  elicited  rewarming  bouts.  Our  results 
thus  raise  the  possibility  that  the  marginally 
higher  rate  of  calling  in  cold  challenged  em- 
bryos, especially  as  they  neared  hatching  (Fig. 
1 ),  may  represent  the  beginnings  of  a prehatch 
vocal  thermoregulatory  response  to  cold. 

Examination  of  the  patterns  of  calling  (Ta- 
ble 1 ) suggests  that  immediately  before  hatch 
(<2.5  h),  rewarming  was  associated  with  de- 
creased calling  in  experimental  (cold  chal- 
lenged) embryos,  while  the  calling  of  control 
embryos  during  the  final  minute  of  the  2 min 
mock  rewarming  period  continued  at  a signif- 
icantly higher  level.  This  difference  is  also 
suggestive  of  an  incipient  vocal  response  to 
temperature  by  chick  embryos  during  the  final 
hours  before  hatching.  An  increase  in  vocal 
response  to  temperature  as  the  time  of  hatch- 
ing approaches  has  also  been  noted  during  vo- 
cal regulation  studies  in  semiprecocial  Ring- 
billed Gulls  (iMrus  delawarensis',  Evans  et  al. 
1994)  and  Herring  Gulls  {Lams  argentatus\ 
Evans  et  al.  1995). 

The  subtle  beginnings  of  the  vocal  response 
to  temperature  found  here  parallels  the  incip- 
ient development  of  endothermy  in  late  stage 
chicken  embryos.  The  prehatching  endother- 
mic response  is  not  robust.  It  has  been  noted 
only  when  exposure  to  cold  was  limited  to 
gradual  cooling  of  late  stage  embryos  (Tazawa 
et  al.  1988).  Incipient  endothermy  is  thought 
to  be  limited  initially  by  conductance  of  O, 
through  the  eggshell  (Tazawa  et  al.  1989)  and 
then  by  the  embryo’s  limited  endothermic 
power  (Tazawa  et  al.  1988,  Whittow  and  Ta- 
zawa 1991).  While  these  physiological  results 
suggest  that  incipient  endothermy  occurs  dur- 
ing the  latter  stages  of  embryonic  develop- 
ment, the  dramatic  increase  in  oxygen  con- 
sumption at  the  time  of  hatching  (Kuroda  et 
al.  1990,  Whittow  and  Tazawa  1991)  indicates 


that  endothermy  becomes  functional  at  that 
time  (Freeman  1971,  1983).  Our  present  re- 
sults suggest  that  there  is  a similar  incipient 
vocal  thermoregulatory  response  of  late  stage 
embryos  to  cold  that  becomes  fully  functional 
at  or  soon  after  hatching  (Tuculescu  and  Gris- 
wold 1983,  Espira  and  Evans  1996,  Bugden 
and  Evans  1997).  The  developmental  onset  of 
vocal  and  metabolic  thermoregulation  thus  ap- 
pear to  be  closely  linked  in  precocial  domestic 
chicks. 

In  contrast  to  domestic  chicks,  the  vocal 
thermoregulatory  system  of  altricial  pelicans 
and  semiprecocial  gulls  is  well  developed  pri- 
or to  hatch  (Evans  1988a,  1990a;  Evans  et  al. 
1994,  1995).  Altricial  and  semiprecocial  spe- 
cies evidently  do  not  show  any  endothermic 
response  before  hatch  (Matsunaga  et  al.  1989, 
Kuroda  et  al.  1990,  Whittow  and  Tazawa 
1991)  suggesting  that  behavioral  and  meta- 
bolic thermoregulation  are  developmentally 
uncoupled  in  these  species,  unlike  the  appar- 
ent linkage  in  the  domestic  chicken. 

Pelicans  and  gulls  both  exhibit  asynchro- 
nous hatching,  and  later  hatching  eggs  poten- 
tially experience  significant  levels  of  incuba- 
tion neglect  as  the  parents  attend  to  the  chicks 
that  have  already  hatched  (Evans  1990d,  Lee 
et  al.  1993,  Evans  et  al.  1995).  Although 
chickens  display  some  level  of  hatching  asyn- 
chrony (mean  of  15  h in  Burmese  Junglefowl; 
Meijer  and  Siemers  1994),  the  chicken  is  not 
known  to  neglect  its  eggs  and  will  normally 
remain  on  the  nest  for  the  first  12-24  h after 
hatching  (McBride  et  al.  1969,  Miller  1978, 
Meijer  and  Siemers  1994).  A vocal  response 
to  cold  thus  may  not  be  a functionally  useful 
behavioral  response  for  chicken  embryos.  Af- 
ter hatching,  the  situation  changes  dramatical- 
ly when  mobile  chicks  are  potentially  exposed 
to  colder  ambient  temperatures,  especially 
during  foraging  bouts  (McBride  et  al.  1969, 
Sherry  1981).  At  that  time  calling  to  solicit 
brooding  warmth  becomes  an  important  part 
of  their  behavioral  response  to  cold  (Kaufman 
and  Hinde  1961,  McBride  et  al.  1969,  Sheny 
1981).  Taken  together,  results  to  date  suggest 
that  in  precocial  chickens,  vocal  and  endo- 
thermic thermoregulation  both  show  incipient, 
but  largely  nonfunctional,  development  prior 
to  hatching  and  are  both  turned  on  rapidly  as 
the  chicks  hatch  and  thermoregulation  be- 
comes a highly  adaptive  capability. 


Bugden  and  Evans  • VOCAL  RESPONSE  OF  CHICKEN  EMBRYOS 


193 


ACKNOWLEDGMENTS 

This  study  was  supported  financially  by  an  operating 
grant  to  R.M.E.  from  the  Natural  Sciences  and  Engi- 
neering Research  Council,  Ottawa,  Canada.  R.  Wilson 
and  R.  A.  McArthur  provided  helpful  comments  on  the 
manuscript. 

LITERATURE  CITED 

Brua,  R.  B.  1996.  Impact  of  embryonic  vocalizations 
on  the  incubation  behaviour  of  Eared  Grebes.  Be- 
haviour 133:145-160. 

Brua,  R.  B.,  G.  L.  Nuechtkrlein,  and  D.  Buitron. 
1996.  Vocal  response  of  Eared  Grebe  embryos  to 
egg  cooling  and  egg  turning.  Auk  1 13:525-533. 
Bugden,  S.  C.  and  R.  M.  Evans.  1991.  Vocal  respon- 
siveness to  chilling  in  embryonic  and  neonatal 
American  Coots.  Wilson  Bull.  103:712-717. 
Bugden,  S.  C.  and  R.  M.  Evans.  1997.  Vocal  solici- 
tation of  heat  as  an  integral  component  of  the  de- 
veloping thermoregulatory  system  in  young  do- 
mestic chickens.  Can.  J.  Zool.  75:1949-1954. 
COLLiAS,  N.  E.  1952.  The  development  of  social  be- 
havior in  birds.  Auk  69:127—159. 

COLLiAS,  N.  E.  1987.  The  vocal  repertoire  of  the  Red 
Junglefowl:  a spectrographic  classification  and  the 
code  of  communication.  Condor  89:510-524. 
Dawes,  C.  M.  1981.  The  effects  of  cooling  the  egg  on 
the  respiratory  movements  of  the  hatching  fowl. 
Callus  g.  domesticus,  with  a note  on  vocalization. 
Comp.  Biochem.  Physiol.  68A:399— 404. 

Espira,  a.  and  R.  M.  Evans.  1996.  Energy  savings 
from  vocal  regulation  of  ambient  temperature  by 
3 day-old  domestic  chicks.  Can.  J.  Zool.  74:599- 
605. 

Evans,  R.  M.  1988a.  Embryonic  vocalizations  as  care- 
soliciting  signals,  with  particular  reference  to  the 
American  White  Pelican.  Proc.  Int.  Ornith.  Congr. 
19:1467-1475. 

Evans,  R.  M.  1988b.  Embryonic  vocalizations  and  the 
removal  of  foot  webs  from  pipped  eggs  in  the 
American  White  Pelican.  Condor  90:721—723. 
Evans,  R.  M.  1989.  Egg  temperatures  and  parental  be- 
havior during  the  transition  from  incubation  to 
brooding  in  the  American  White  Pelican.  Auk 
106:26-33. 

Evans,  R.  M.  1990a.  Vocal  regulation  of  temperature 
by  avian  embryos:  a laboratory  study  with  pipped 
eggs  of  the  American  White  Pelican.  Anim.  Be- 
hav.  40:969-979. 

Evans,  R.  M.  1990b.  Embryonic  fine  tuning  of  pipped 
egg  temperature  in  the  American  White  Pelican. 
Anim.  Behav.  40:963-968. 

Evans,  R.  M.  1990c.  Effects  of  low  incubation  tem- 
peratures during  the  pipped  egg  stage  on  hatch- 
ability  and  hatching  limes  in  domestic  chickens 
and  Ring-billed  Gulls.  Can.  J.  Zool.  68:836—840. 
Evans,  R.  M.  1990d.  Terminal  egg  neglect  in  the 
American  White  Pelican.  Wilson  Bull.  102:684— 
692. 

Evans,  R.  M.  1992.  Embryonic  and  neonatal  vocal 


elicitation  of  parental  brooding  and  feeding  re- 
sponses in  American  While  Pelicans.  Anim.  Be- 
hav. 44:667—675. 

Evans,  R.  M.,  A.  Whitaker,  and  M.  O.  Wiehe.  1994. 
Development  of  vocal  regulation  of  temperature 
by  embryos  in  pipped  eggs  of  Ring-billed  Gulls. 
Auk  1 1 1 :596-604. 

Evans,  R.  M.,  M.  O.  Wiebe,  S.  C.  Lee,  and  S.  C. 
Bugden.  1995.  Embryonic  and  parental  preferenc- 
es for  incubation  temperature  in  Herring  Gulls: 
implications  for  parent-offspring  conflict.  Behav. 
Ecol.  Sociobiol.  36:17—23. 

Freeman,  B.  M.  1964.  The  emergence  of  the  homeo- 
thermic  metabolic  response  in  the  fowl  {.Callus 
domesticus).  Comp.  Biochem.  Physiol.  13:413- 
422. 

Freeman,  B.  M.  1971.  Body  temperature  and  ther- 
moregulation. Pp.  1115-1151  in  Physiology  and 
biochemistry  of  the  domestic  fowl  (D.  M.  Bell  and 
B.  M.  Freeman,  Eds.).  Academic  Press,  London. 
U.K. 

Freeman,  B.  M.  1983.  Body  temperature  and  ther- 
moregulation. Pp.  365-377  in  Physiology  and  bio- 
chemistry of  the  domestic  fowl.  Vol.  4 (D.  M.  Bell 
and  B.  M.  Freeman,  Eds.).  Academic  Press,  Lon- 
don, U.K. 

Freeman,  B.  M.  and  M.  A.  Vince.  1974.  Development 
of  the  avian  embryo.  Chapman  and  Hall,  London, 
U.K. 

Gottlieb,  G.  and  J.  G.  Vandenbergh.  1968.  Ontog- 
eny of  vocalization  in  duck  and  chick  embryos.  J. 
Exp.  Zool.  168:307-326. 

Impekoven,  M.  1973.  The  response  of  incubating 
Laughing  Gulls  {Earns  alricilla  L.)  to  calls  of 
hatching  chicks.  Behaviour  46:94-1 13. 

Kaufman,  I.  C.  and  R.  A.  Hinde.  1961.  Factors  influ- 
encing distress  calling  in  chicks,  with  special  ref- 
erence to  temperature  changes  and  social  isola- 
tion. Anim.  Behav.  9:197-204. 

Kuroda,  O.,  C.  Matsunaga,  G.  C.  Whittow,  and  H. 
Tazawa.  1990.  Comparative  metabolic  responses 
to  prolonged  cooling  in  precocial  duck  {Anas  do- 
mesticiis)  and  altricial  pigeon  {Coluinba  doniesti- 
ca)  embryos.  Comp.  Biochem.  Physiol.  95A:407— 
410. 

Lee,  S.  C.,  R.  M.  Evans,  and  S.  C.  Bugden.  1993. 
Benign  neglect  of  terminal  eggs  in  Herring  Gulls. 
Condor  95:507-514. 

Matsunaga,  C.,  P.  M.  Matiiili,  G.  C.  Whittow,  and 
H.  Tazawa.  1989.  Oxygen  consumption  of  Brown 
Noddy  {Anous  stolidu.s)  embryos  in  a quasiequi- 
librium stale  at  lowered  ambient  temperatures. 
Comp.  Biochem.  Physiol.  93A:707— 710. 

McBride,  G.,  1.  P.  Parer,  and  F.  Foenander.  1969. 
The  social  organization  and  behaviour  of  the  feral 
domestic  fowl.  Anim.  Behav.  Monogr.  2:127-181. 

Meijer,  T.  and  I.  SiEMERS.  1994.  Incubation  develop- 
ment and  asynchronous  hatching  in  junglefowl. 
Behaviour  127:309-322. 

Miller,  D.  B.  1978.  Early  parent-young  interaction  in 


194 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2.  June  1999 


Red  Junglefowl:  earlobe  pecking.  Condor  80:503- 
504. 

Oppenhein,  R.  W.  and  H.  L.  Levin.  1975.  Short-term 
changes  in  incubation  temperature:  behavioral  and 
physiological  effects  in  the  chick  embryo  from  6 
to  20  days.  Dev.  Psychobiol.  8:103-115. 

Sherry,  D.  E 1981.  Parental  care  and  the  development 
of  thermoregulation  in  Red  Junglefowl.  Behaviour 
76:250-279. 

Tazawa,  H.,  H.  Wakayama,  J.  S.  Turner,  and  C.  V. 
Paganelli.  1988.  Metabolic  compensation  for 
gradual  cooling  in  developing  chick  embryos. 
Comp.  Biochem.  Physiol.  89A:  125-129. 


Tazawa,  H.,  G.  C.  Whittow,  J.  S.  Turner,  and  C.  V. 
Paganelli.  1989.  Metabolic  responses  to  gradual 
cooling  in  chicken  eggs  treated  with  thiourea  and 
oxygen.  Comp.  Biochem.  Physiol.  92A:619-622. 
Templeton,  R.  K.  1983.  Why  do  Herring  Gull  chicks 
vocalize  in  the  shell?  Bird  Study  30:73-74. 
Tuculescu,  R.  a.  and  j.  G.  Griswold.  1983.  Pre- 
hatching interactions  in  domestic  chickens.  Anim. 
Behav.  3 1 : 1-10. 

Webb,  D.  R.  1987.  Thermal  tolerance  of  avian  embryos: 

a review.  Condor  89:874-898. 

Whittow,  G.  C.  and  H.  Tazawa.  1991.  The  early  de- 
velopment of  themioregulation  in  birds.  Physiol. 
Zool.  64:1371-1390. 


Wilson  Bull..  111(2),  1999,  pp.  195-209 


BEHAVIOR  AND  VOCALIZATIONS  OF  THE  CAURA  AND  THE 

YAPACANA  ANTBIRDS 

KEVIN  J,  ZIMMER'  2 


ABSTRACT. — The  first  detailed  information  on  the  vocalizations  (including  the  first  sound  spectrograms)  and 
natural  history  of  the  Caura  Antbird  (Percnostola  caurensis)  and  the  Yapacana  Antbird  (Myrmeciza  clLsjuncta) 
are  presented.  The  Caura  Antbird  was  studied  in  the  Serrania  de  la  Cerbatana,  edo.  BoKvar,  Venezuela,  where 
it  inhabits  humid  foothill  forest  dominated  by  large  rocks.  Caura  Antbirds  specialized  in  foraging  on  or  beneath 
rocks,  a behavior  unusual  among  the  Thamnophilidae.  The  Yapacana  Antbird  was  studied  at  a site  along  the 
south  bank  of  the  Rio  Ventuari,  edo.  Amazonas,  Venezuela.  These  antbirds  were  locally  abundant  in  a specialized 
stunted  woodland  that  grows  on  white  sand  soils.  Based  on  newly  described  vocal  characters,  the  closest  relatives 
of  P.  caurensis  appear  to  be  P.  leucostigrna  and  P.  schistacea,  whereas  M.  disjuncta  has  no  apparent  close 
relatives  and  probably  merits  placement  in  a monotypic  genus.  Received  9 July  1998,  accepted  5 Jan.  1999. 


Among  the  least  known  members  of  the 
large  antbird  family  Thamnophilidae  are  the 
Caura  Antbird  {Percnostola  caurensis)  and 
the  Yapacana  Antbird  {Myrmeciza  disjuncta). 
Both  species  are  nearly  endemic  to  south- 
western Venezuela  and  have  remained  rela- 
tively unobserved  by  modem  field  ornitholo- 
gists. The  most  extensive  collections  of  both 
species  (36  specimens  of  P.  caurensis  and  5 
specimens  of  M.  disjuncta)  reside  in  the  Co- 
lecion  Omitologia  Phelps  (COP),  Caracas, 
Venezuela.  Single  specimens  of  each  species 
collected  near  Pico  Neblina  (edo.  Amazonas, 
Venezuela)  in  1984  by  Field  Museum  of  Nat- 
ural History  (FMNH)  personnel  represent  the 
only  specimens  of  P.  caurensis  and  M.  dis- 
juncta collected  anywhere  since  1972  and 
1981  respectively.  There  is  essentially  no  pub- 
lished information  on  habitat  or  behavior  of 
the  two  species,  and  nothing  is  known  of  their 
vocalizations  (Ridgely  and  Tudor  1994). 

In  February  1998  I observed  the  habitats 
and  behaviors  and  tape-recorded  the  vocali- 
zations of  Caura  Antbirds  in  the  Senania  de 
la  Cerbatana,  edo.  Bolivar,  Venezuela,  and  of 
Yapacana  Antbirds  in  Yapacana  National 
Park,  edo.  Amazonas,  Venezuela.  This  is  the 
first  detailed  information  on  the  natural  his- 
tory and  vocalizations  of  these  species  and  al- 
lows a more  informed  assessment  of  their  pos- 
sible generic  affinities. 


' Los  Angeles  County  Museum  of  Natural  History, 
900  Exposition  Blvd.,  Los  Angeles,  CA  90007. 

-Correspondence  address:  1665  Garcia  Rd.,  Atas- 
cadero, CA  93422;  E-mail:  kjzsrzC^tcsn.net 


STUDY  AREAS  AND  METHODS 

I observed  Caura  Antbirds  10-15  February,  1998  in 
the  Serrania  de  la  Cerbatana  near  Hato  Las  Nieves 
(6°  34'  N,  66°  12'  W),  edo.  Bolivar,  Venezuela  (Fig. 
1).  The  Serrania  de  la  Cerbatana  rings  a large  valley 
vegetated  mostly  by  a mixture  of  savanna  and  tropical 
dry  forest,  transected  by  narrow  bands  of  gallery  forest 
along  occasional  streams,  and  dotted  with  groves  of 
Mauritia  palms  in  poorly  drained  areas.  Three  main 
rivers,  Cano  Las  Nieves,  Rio  Agua  Fria,  and  Rio  Dan- 
ta,  drain  the  valley.  The  Serrania  mountain  range  rises 
dramatically  from  the  valley,  and  on  the  southern  and 
western  borders  is  covered  with  humid  forest.  The 
south  facing  slopes  of  the  mountains  to  the  north  of 
the  valley  are  noticeably  drier.  The  tallest  peak  in  the 
chain  is  Pico  Las  Nieves  at  2080  m.  I studied  Caura 
Antbirds  along  a 1.5  km  trail  that  began  in  humid  for- 
est at  280  m elevation  and  extended  up  the  side  of  a 
ridge  to  400  m. 

I studied  Yapacana  Antbirds  22-27  February  1998 
along  a trail  (hereafter  the  “Picua  Trail”)  near  the  set- 
tlement of  Picua  (4°  5'  N,  66°  45'  W)  in  Yapacana  Na- 
tional Park,  edo.  Amazonas,  Venezuela  (Fig.  1).  Picua 
is  a small  settlement  of  Piaroa  and  Mako  Indians,  on 
the  south  (left)  bank  of  the  Rfo  Ventuari  at  about  150 
m elevation.  The  surrounding  area  contains  a mosaic 
of  different  soil  types  that  support  a patchwork  of  dis- 
tinct vegetation  types.  The  Ventuari  and  its  many  small 
tributaries  are  flanked  by  bands  (of  varying  width)  of 
tall  forest  (15-25  m)  that  grow  on  yellow  clay  soils 
and  are  seasonally  flooded.  This  varzea  forest  is  char- 
acterized by  a closed  canopy  with  a fairly  open  un- 
derstory and  an  abundance  of  vines  and  lianas.  On 
higher  banks  above  the  river  the  varzea  grades  into  a 
taller  transitional  forest  with  a denser  understory. 
Much  of  the  area  farther  removed  from  the  river  is 
dominated  by  white  sand  soils  on  which  grow  lower 
stature  woodlands  and  grassy  savannas.  The  savannas 
range  in  size  from  less  than  1 ha  to  about  1 km-,  with 
scattered  shrubs  and  small  trees.  Larger  savannas  in 
the  area  contained  stands  of  Mauritia  palms.  Large, 
isolated  patches  of  red  clay  soils  support  tall  (>30  m). 


195 


196 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


tola  caurensis)  and  Yapacana  Antbird  (Myrmeciza  cHs- 
juncta):  open  circles  = confirmed  sites  for  P.  caurenxis 
(confirmation  based  on  specimen  records);  open 
squares  = confirmed  sites  for  Myrmeciz.a  disjuncta\ 
open  star  = site  near  the  base  of  Pico  de  la  Neblina 
where  both  species  have  been  collected  in  close  prox- 
imity. The  black  circle  and  #1  indicates  the  site  of  the 
present  study  in  the  Serrania  de  la  Cerbatana,  edo.  Bo- 
livar, Venezuela  where  P.  caurensis  was  studied.  The 
black  square  and  #2  indicates  the  site  of  the  present 
study  near  Picua,  in  Yapacana  National  Park,  edo. 
Amazonas,  Venezuela  where  M.  disjuncta  was  studied. 


lush  “islands”  of  humid  tropical  forest  that  are  not 
seasonally  flooded  (=  terra  firme  forest).  Quartzite 
dome-like  sandstone  hills  (cerros)  and  low  outcrop- 
pings are  scattered  throughout  the  region. 

Whenever  individuals  or  pairs  of  P.  caurensis  or  M. 
di.sjuncta  were  located,  1 followed  them  for  as  long  as 
possible,  tape-recording  as  many  vocalizations  as  1 
could  and  summarizing  foraging  and  other  behaviors 
on  cassette  tape.  Some  of  these  behaviors  were  also 
documented  on  videotape.  I used  tape-playback  of 
these  recordings  to  assess  presence  or  absence  of  ant- 
birds  in  places  where  no  spontaneously  vocalizing 
birds  were  heard  and  to  determine  the  limits  of  terri- 
torial boundaries.  All  measurements  included  in  such 
summaries  (height  above  ground,  territory  size,  dis- 
tances, times,  etc.)  are  estimates.  Terminology  for  for- 
aging behavior  follows  Remsen  and  Robinson  (1990). 

Tape  recordings  were  made  with  a Sony  TCM-50()0 
recorder  and  ,Sennheiser  MKH-70  microphone.  All  re- 
cordings have  been  or  will  be  archived  at  the  Library 
of  Natural  Sounds.  Cornell  Univ.,  Ithaca,  New  York. 
Spectrograms  were  made  by  Phyllis  Isler  on  a Power 
Macintosh  7500  computer  using  Canary  version  1.2.1 
(Bioacoustics  Research  Program,  Cornell  Laboratory 
of  Ornithology,  Ithaca.  New  York).  Morton  Isler  com- 
piled a comprehensive  list  of  distributional  records  of 
the  two  species  of  antbirds  as  documented  by  speci- 


mens, tape  recordings,  or  photographs  (Isler  1997;  Pig. 
1). 

PERCNOSTOLA  CAURENSIS 

Distribution  and  habitat. — The  Caura  Ant- 
bird is  known  only  from  the  western  portion 
of  the  “Pantepui”  region  (Mayr  and  Phelps 
1967)  south  of  the  Rfo  Orinoco  in  the  Vene- 
zuelan states  of  Bolivar  and  Amazonas  and  in 
extreme  northern  Brazil  (Fig.  1).  Percnostola 
caurensis  is  well  represented  in  museum  col- 
lections, with  36  specimens  (the  most  recent 
collected  in  1972)  in  the  Colecion  Omitologia 
Phelps  (Caracas,  Venezuela)  alone  (C.  Rodner, 
pers.  comm.).  The  most  recent  substantiated 
record  was  of  a male  collected  in  1984  at  1250 
m near  the  base  of  Pico  Neblina  (edo.  Ama- 
zonas, Venezuela;  Willard  et  al.  1991).  The 
occurrence  of  P.  caurensis  in  the  Serrania  de 
la  Cerbatana  represents  a slight  range  exten- 
sion to  the  northwest  (Fig.  1). 

The  forest  along  the  first  500  m of  the  trail 
was  tall  (ca  30  m),  with  an  open  understory 
dominated  by  slender  palms.  The  terrain  was 
flat  and  nearly  devoid  of  large  rocks.  I could 
locate  only  one  territory  of  P.  caurensis  along 
this  portion  of  the  trail,  and  it  abutted  the  bot- 
tom of  the  hill.  The  hillside  forest  beyond  500 
m was  also  fairly  open,  with  an  intermittent 
canopy  of  about  20  m.  Few  trees  were  larger 
than  30  cm  dbh,  and  woody  vines  were  abun- 
dant. Large  stands  of  a naiTow-leaved,  non- 
spiny  bamboo  (1-2  m in  height)  occupied 
most  light  gaps.  The  entire  slope  was  extreme- 
ly rocky,  with  numbers  of  boulders  up  to  8 m 
tall  and  15  m along  their  longest  axis.  These 
boulders  were  typically  moss  and  fern  cov- 
ered, with  terrestrial  bromeliads,  cacti,  and 
bamboo  growing  over  their  tops  and  in  the 
crevices  (Fig.  2).  Many  were  topped  with 
small  trees,  the  gnarled  roots  of  which  draped 
off  the  sides  of  the  rocks  like  tendrils,  trapping 
leaf  litter  and  organic  debris.  I located  6 pairs 
of  P.  caurensis  along  about  1 km  of  trail 
through  this  rocky,  hillside  forest. 

The  SeiTania  de  la  Cerbatana  was  extremely 
dry  in  February  1998.  Typical  dry  season  con- 
ditions appeared  exacerbated  by  ongoing  El 
Nino  related  events.  Leaf  litter  throughout  the 
forest  was  extremely  dry  and  many  trees  had 
shed  large  numbers  of  leaves.  This  was  par- 
ticularly evident  in  the  hillside  forest,  where 
large  patches  (0.5-1  ha)  of  deciduous  vege- 


Zimmer  • CAURA  AND  YAPACANA  ANTBIRDS 


197 


FIG.  2.  Rocky  hillside  forest  in  the  Serrania  de  la  Cerbatana,  edo.  Bolivar,  Venezuela.  (A)  Relatively  open 
forest,  with  a broken  canopy  of  about  20  m.  The  rocks  were  2—3  m tall  and  4 ni  in  diameter.  (B)  A rocky 
alluvial  fan  along  which  were  located  two  Caura  Antbird  territories.  Note  the  highly  deciduous  state  of  the 
vegetation  in  this  light  gap,  and  the  abundance  of  leaf  litter  trapped  in  the  roots  and  vines  overtopping  the  rocks. 


tation  were  conspicuously  scattered  across  the 
slope,  usually  coincident  with  the  rock  strewn 
alluvial  fans  at  the  bottom  of  ravines. 

Morphology. — Soft-part  colors  were  iden- 
tical for  both  sexes.  The  iris  was  reddish- 
brown,  the  legs  and  feet  were  slate  gray  (a 
shade  paler  than  the  bill),  and  the  bill  was 
blackish.  Plumage  was  as  described  by  Ridge- 
ly  and  Tudor  (1994). 

Vocalizations. — Caura  Antbirds  were  gen- 
erally quiet  during  my  fieldwork,  as  were 
most  other  species  of  insectivorous  birds. 
Dawn  choruses  were  both  unremarkable  and 
short,  suggesting  a low  level  of  breeding  ac- 
tivity for  most  species  during  the  height  of  the 
dry  season.  There  was  a sustained  rain  during 
the  early  morning  hours  of  13  February;  the 
two  following  mornings  I noted  increased 
spontaneous  song  from  Caura  Antbirds. 

I recorded  over  120  loudsongs  (as  defined 
by  Isler  et  al.  1997)  and  900  calls  from  12 


individual  antbirds.  The  loudsong  of  P.  cau- 
rensis  is  a far  carrying  series  of  7-15  modu- 
lated and  well  spaced  notes  (Fig.  3A).  The 
first  notes  are  widely  spaced,  and  the  terminal 
notes  are  closer  together  and  drop  in  pitch. 
Female  songs  (Fig.  3B)  were  similar  in  patterm 
to  male  songs,  but  differed  in  other  character- 
istics such  as  mean  number  of  notes,  mean 
frequency,  etc.  Females  sang  less  frequently 
than  males. 

In  response  to  tape  playback  and  during  ter- 
ritorial encounters,  both  sexes  gave  loud, 
buzzy  “zhew”  calls  at  varying  levels  of  fre- 
quency modulation  (Figs.  3C-E).  On  a few 
occasions,  birds  involved  in  tenitoiial  dis- 
putes uttered  sharp  “quip”  notes  (Fig.  3F) 
when  neighboring  birds  approached  closely. 
The  most  frequently  heard  vocalization,  and 
one  given  by  birds  startled  along  the  trail,  was 
an  abbreviated  loud  rattle  (Figs.  3G,  H),  sim- 
ilar to  the  alar'm  calls  of  P.  leucostigma  and 


198 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


G 

H 

H 1 

l| 

III 

^ 1 1 — 1 1 

1 

1 1 ^ 

1 1 1 1 

' I T I I 1 1 1— 

0 0.5  1.0  1.5  2.0  2.5  3.0  3.5  4,0  4 5 


Time  (seconds) 

PIG.  3.  Spectrograms  ot  Caura  Antbird  {Percnosloki  caurensi.s)  vocalizations:  (A)  male  loudson'g,  (B)  female 
loLidsong,  (C,  D,  E)  “zhew”  calls,  which  probably  function  as  aggression  calls,  at  various  levels  of  frequency 
modulation,  (E)  “quip”  call,  (G)  male  alarm  rattles,  and  (H)  female  alarm  rattles. 


several  species  of  Myrmeciza  anlbirds  (pers. 
obs.).  A loud  “chikit”  (not  tape  recorded)  was 
often  given  by  Caura  Antbirds  at  the  moment 
they  took  flight.  Birds  alarmed  by  my  pres- 


ence gave  this  call  immediately  before  flush- 
ing, as  did  birds  chasing  one  another  about  in 
an  aggressive  boundary  dispute. 

Behavior. — Caura  Antbirds  were  encoun- 


Zimmer  • CAURA  AND  YAPACANA  ANTBIRDS 


199 


tered  singly  or  in  pairs,  but  did  not  associate 
with  mixed-species  flocks.  Mates  foraging  to- 
gether were  typically  within  15  m of  one  an- 
other, and  alarm  calls  from  one  bird  elicited 
an  immediate  vocal  response  from  the  other. 
Foraging  birds  maintained  a neaily  horizontal 
posture,  with  the  head  held  higher  than  the 
axis  of  the  body.  All  individuals  continuously 
raised  their  tails  a few  degrees  above  the  plane 
of  the  body  and  then  wagged  them  slowly 
downward  in  an  arc  20-30°  below  the  plane 
of  the  body.  This  tail  movement  was  most  ex- 
aggerated when  the  antbirds  clung  laterally  to 
elevated  perches  or  responded  to  tape  play- 
back, but  was  also  used  during  tenestrial  for- 
aging. Mostly  the  tail  appeared  to  be  slightly 
fanned.  Tail  movements  were  occasionally  ac- 
companied by  a simultaneous  wing-flick. 

Caura  Antbirds  foraged  mostly  on  rocks  or 
the  ground.  Both  sexes  spent  long  times 
creeping  over  the  large  boulders,  often  cling- 
ing laterally  to  nearly  vertical  rock  faces  and 
going  in  and  out  of  the  numerous  crevices  (of- 
ten for  minutes  at  a time)  in  the  manner  of 
Slaty  Bristlefronts  (Meridaxis  ater,  M.  Isler 
and  P.  Isler,  pers.  comm.).  Rock  Wrens  {Sal- 
pinctes  obsoletus)  or  Canyon  Wrens  {Cather- 
pes  mexicamis).  While  creeping  about  the 
rocks,  the  antbirds  frequently  probed  in  the 
mosses  and  small  ferns  covering  the  surface, 
but  spent  most  of  their  time  inspecting  the  leaf 
litter  trapped  between  the  roots  and  vine  tan- 
gles of  trees  overtopping  the  boulders.  Arthro- 
pod prey  (primarily  orthopterans  and  hemip- 
terans,  as  well  as  many  arthropods  too  small 
to  be  identified)  were  gleaned  from  root  and 
vine  surfaces  with  quick  stabbing  motions. 
Curled  dead  leaves  were  carefully  probed  with 
the  bill.  Some  antbirds  inspected  dead  leaves 
without  tossing  them,  others  picked  up  leaves 
with  the  bill  before  tossing  them  aside.  The 
antbirds  routinely  squeezed  themselves  into 
small  spaces  between  the  rock  surface  and 
overlying  roots  and  vines,  remaining  in  these 
somewhat  “canopied”  niches  to  forage  for  up 
to  60  s.  When  foraging  on  rocks,  the  antbirds 
tended  to  spend  most  of  their  time  in  some- 
what protected  locations,  within  the  interior  of 
vine  tangles  and  root  masses  overtopping  the 
rocks,  within  crevices  in  the  rocks,  or  beneath 
rocky  ledges  and  overhangs.  The  antbirds 
moved  steadily  over  open  rock  faces,  pro- 
gressing by  short  hops  of  5—10  cm  and  oc- 


casional longer  wing-assisted  hops  or  short, 
abrupt  flights,  often  to  low,  overhanging 
branches  or  ledges.  They  frequently  hopped 
from  a rock  up  to  a low  branch  or  sapling  to 
scan  for  1-5  s before  dropping  back  to  the 
rock.  Birds  often  dropped  1-5  m from  the 
rocks  to  the  ground.  There,  they  hopped  be- 
neath the  overhangs,  probed  in  leaf  litter  and 
inspected  rock  surfaces.  Antbirds  spent  many 
minutes  inspecting  fissures,  crevices,  and  gaps 
within  and  between  rocks,  often  retracing  their 
routes. 

One  male  observed  foraging  for  more  than 
30  minutes  spent  the  bulk  of  this  time  forag- 
ing over,  under,  and  between  rocks,  probing  in 
moss  and  vine  tangles.  On  three  occasions  the 
bird  dropped  to  the  ground  at  the  base  of  large 
rocks  and  spent  1-5  minutes  vigorously  toss- 
ing dead  leaves  in  the  manner  of  a leaftosser 
{Sclerurus  spp.).  Large  leaves  (many  larger 
than  the  bird)  were  picked  up  with  the  bill  and 
tossed.  Smaller  leaves  were  frequently  flipped 
by  inserting  the  bill  beneath  the  leaf  and  then 
giving  a quick  upward  flaking  motion.  He  also 
made  an  upward  sally  of  20-25  cm  from  the 
ground  to  take  an  unidentified  arthropod  from 
the  underside  of  a green  leaf.  Other  individ- 
uals occasionally  made  similar  short  upward 
sallies  to  glean  prey  from  overhanging  rocks. 

One  female  antbird  spent  several  minutes 
hopping  ar'ound  the  periphery  of  a large  emer- 
gent swarm  of  small,  winged  ants.  Although 
she  picked  at  the  ground  several  times,  it  was 
not  clear  whether  she  was  feeding  on  the  ants. 
The  bird  flew  off  and  returned  to  the  ant- 
swarm  twice.  The  male  antbird  was  foraging 
nearby,  but  did  not  attend  the  swarm. 

Territories  appeared  to  be  about  150—200  m 
in  diameter.  I witnessed  only  one  territorial 
conflict.  This  was  a prolonged  encounter  be- 
tween pairs  whose  territories  bounded  a large 
rock-slide  along  the  center  of  a ravine.  The 
conflict  was  marked  by  several  advances  and 
retreats  by  both  pairs.  Both  members  of  each 
pair  countersang  at  length  while  gradually  ap- 
proaching their  counterparts.  Whenever  the 
pairs  approached  to  within  about  20  m of  one 
another,  they  tended  to  substitute  “zhew” 
calls  for  songs.  When  they  were  in  visual  con- 
tact or  close  auditory  contact,  the  pairs  ex- 
changed harsh  calls  for  a few  minutes  before 
one  individual  or  pair  retreated,  often  with  its 
rival  in  close  pursuit.  These  abrupt  retreats 


200 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


were  always  preceded  by  the  loud,  hard 
“chikit”  call.  The  dispute  lasted  for  more  than 
30  minutes.  Eventually  the  pairs  retreated  to 
opposite  ends  of  the  rockslide,  periodically 
singing. 

Most  species  of  passerines  besides  P.  cau- 
rensis  were  encountered  in  mixed-species  can- 
opy and  mid-level  flocks.  The  apparent  ab- 
sence of  terrestrial  or  semi-terrestrial  antbirds 
(other  than  P.  caurensis)  or  fumariids  is  note- 
worthy, although  my  failure  to  detect  such 
species  could  be  an  artifact  of  depressed  song 
activity  during  the  dry  season.  The  taller,  less- 
deciduous  forest  below  the  mountains  had  a 
more  diverse  avifauna. 

The  presence  of  large  rocks  appeared  to  be 
a critical  component  of  P.  caurensis  habitat. 
Of  the  7 territories,  6 were  in  areas  with  abun- 
dant large  rocks.  The  seventh  territory  was 
close  enough  to  the  bottom  of  the  hillside  to 
possibly  have  included  some  rocky  terrain  as 
well.  The  rocks  provided  microhabitats  in 
which  a diversity  of  plants  flourished,  and 
these  plants,  in  turn,  provided  a wealth  of  po- 
tential foraging  strata  for  a small  terrestrial  in- 
sectivore  such  as  P.  caurensis.  Caura  Antbirds 
spent  more  than  80%  of  their  foraging  time 
on  rocks  and  overtopping  vegetation,  or  on  the 
ground  directly  beneath  overhanging  rocks. 
The  importance  of  rocks  as  foraging  sites  for 
Caura  Antbirds  may  be  increased  during  the 
dry  season,  when  favorable  moisture-retaining 
microclimates  are  created  along  rock  edges  or 
crevices.  Furthermore,  the  tendency  for  over- 
topping root  masses  and  vine  tangles  to  trap 
leaf  litter  is  accentuated  during  the  dry  season, 
when  many  trees  drop  their  leaves.  Such  ac- 
cumulations of  organic  litter  may  provide  at- 
tractive sites  for  arthropods  when  the  forest  is 
water-stressed.  It  is  interesting  that  no  other 
bird  species  was  observed  to  exploit  the  extra 
resource  dimensions  created  by  the  large 
rocks.  The  presence  of  Guianan  Cock-of-Rock 
(Rupicola  rupicola)  in  the  area  is  almost  cer- 
tainly dependent  on  the  availability  of  large 
rocks  for  nest  sites  (Snow  1982),  but  only  the 
Caura  Antbirds  seemed  to  use  the  rocks  as 
foraging  substrates. 

MYRMECfZA  DISJUNCTA 

Distrihution  and  habitat. — The  Yapacana 
Antbird  was  described  in  1945  by  H.  Fried- 
mann from  two  specimens  collected  near  the 


base  of  Cerro  Yapacana  (edo.  Amazonas,  Ven- 
ezuela) in  April  1931.  Five  more  specimens 
were  collected  from  the  same  general  locality 
in  April-May  1947  (specimens  COP).  Cerro 
Yapacana  is  an  isolated  outlier  of  the  western 
Tepuis,  rising  steeply  above  the  Rio  Orinoco 
to  an  elevation  of  1 340  m.  Meyer  de  Schauen- 
see  and  Phelps  (1978:218),  perhaps  describing 
the  general  habitat  surrounding  Cerro  Yapa- 
cana, listed  the  habitat  of  M.  disjuncta  as 
“High  rain  forest  at  about  100  m in  under- 
growth and  low  bushes.”  Subsequent  to  its 
description,  M.  disjuncta  has  been  document- 
ed from  only  three  additional  sites  (Fig.  1).  In 
March  1981,  while  working  in  sandy-belt  for- 
est near  Puerto  Inirida,  depto.  Guainia,  Co- 
lombia, J.  Dunning  mist-netted  an  antbird  lat- 
er identified  from  photos  as  a female  Yapa- 
cana Antbird  (Hilty  and  Brown  1986;  ANJP 
specimen  175723,  R.  Ridgely,  pers.  comm.; 
photo  on  file  at  VIREO).  In  Februaiy  1984  a 
single  female  M.  disjuncta  was  collected  at 
140  m near  the  left  bank  of  the  Rio  Baria  on 
the  Venezuelan-Brazilian  border  (Willard  et 
al.  1991).  This  extended  the  known  range  of 
M.  disjuncta  about  350  km  south.  There  is  no 
published  description  of  the  habitat  in  which 
this  Yapacana  Antbird  was  collected,  but  the 
nearby  base  camp  was  in  “tall  seasonal  rain 
forest  drained  by  both  black-water  and  white- 
water  streams”  (Willard  et  al.  1991).  J.  Coons 
and  D.  Stejskal  (pers.  comm.)  were  the  first  to 
find  M.  disjuncta  near  Picua  in  January  1997. 
They  reported  seeing  or  hearing  several  indi- 
viduals in  savanna  woodland  on  white  sand 
soils  along  the  Picua  Trail. 

The  only  habitat  of  the  Ventuari  in  which  I 
found  M.  disjuncta  was  what  the  local  people 
refer  to  as  “monte  cenado.”  I found  this  hab- 
itat only  on  the  south  bank  of  the  Rio  Ventuari 
along  the  Picua  trail.  This  is  a stunted,  virtu- 
ally impenetrable  woodland  that  grows  on 
fine,  compacted  white  sand  soils  that  are  sea- 
sonally saturated  (Fig.  4).  It  is  similar  to  the 
“savanna  woodland”  described  from  Cam- 
pamento  Junglaven  located  farther  north  (up- 
stream) along  the  Ventuari  (Zimmer  and  Hilty 
1997)  but  has  a greater  density  of  vines,  along 
with  abundant  sawgrass  and  bamboo  scattered 
through  the  understory.  The  canopy  varies 
from  6-10  m and  is  of  generally  uniform 
height  with  only  occasional  emergent  trees  of 


Zimmer  • CAURA  AND  YAPACANA  ANTBIRDS 


201 


FIG.  4.  (A)  Monte  cerrado  woodland  along  the  Picua  Trail,  Yapacana  National  Park,  edo.  Amazonas,  Ven- 

ezuela. (B)  Interior  of  the  monte  cerrado,  depicting  a typical  Yapacana  Antbird  tenitory. 


10-15  m.  Few  trees  in  this  habitat  have  trunks 
thicker  than  10  cm  dbh. 

A different  type  of  “sandy  belt  forest”  or 
savanna  woodland  occurred  on  the  north  bank. 
This  woodland  was  partly  deciduous,  less  di- 
verse, and  even  more  stunted  than  the  monte 
cerrado.  The  understory  was  more  open  and 
lacked  both  bamboo  and  sawgrass.  This  forest 
grew  on  coarser,  well-drained  white  sand  soils 
atop  low  ridges  or  rocky  outcroppings.  Myr- 
meciza  disjuncta  and  many  other  species  typ- 
ical of  the  monte  cenado  were  absent  from 
this  scrub  woodland. 

Using  tape  playback  1 located  at  least  24 
pairs  of  M.  disjuncta  along  1 350  m of  the  Pi- 
cua Trail.  No  birds  were  detected  farther  than 
50  m from  the  trail.  Tenitories  were  evenly 
spaced  along  both  sides  of  the  trail  and  ap- 
peared to  be  no  more  than  50—75  m in  di- 
ameter. Near  the  savanna  edge  the  monte  cer- 
rado was  particularly  stunted,  with  a more 
open  canopy  and  more  sawgrass  in  the  under- 
story. I found  only  two  tenitories  of  Yapacana 


Antbirds  along  more  than  150  m of  trail  tran- 
secting this  more  grassy  woodland. 

Morphology. — There  has  been  some  con- 
fusion in  the  literature  regarding  plumage 
characters  of  M.  disjuncta.  Central  to  the  con- 
fusion is  the  type  specimen  of  M.  disjuncta, 
an  immature  male  molting  into  adult  plumage 
(Friedmann  1945,  1948),  which  displayed  a 
combination  of  adult  and  immature  plumage 
characters.  Thus,  Meyer  de  Schauensee  (1970: 
249)  wrote  that  the  male  has  “uppeipaits 
blackish-gray,  crown  and  nape  tinged  brown” 
and  “sides  of  head  gray,  chin  white;  rest  of 
undeiparts  white,  strongly  tinged  ochraceous 
buff;  center  of  abdomen  white.”  Similarly, 
Meyer  de  Schauensee  and  Phelps  (1978:218) 
described  the  male  plumage  as  having  the 
“throat  and  breast  white  suffused  with  ochra- 
ceous tawny,  strongly  so  on  breast  and  sides 
of  throat,  middle  of  abdomen  white,  undertail 
coverts  dark  gray.”  They  described  the  female 
as  differing  from  the  male  by  “daik  ochra- 
ceous buff  spots  on  wing  coverts  and  ochra- 


202 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


ceous  under  tail  coverts.”  This  description 
was  essentially  repeated  in  Hilty  and  Brown 
(1986).  Ridgely  and  Tudor  (1994:332)  were 
much  more  nearly  correct  in  both  their  illus- 
tration of  the  male  and  in  the  description  of 
the  plumage,  but  still  suggested  that  the  un- 
derparts of  the  male  are  “tinged  with  creamy 
buff.” 

None  of  the  20  or  more  male  Yapacana 
Antbirds  that  I saw  showed  any  hint  of  tawny, 
ochraceous  buff,  or  creamy  buff  color  on  the 
underparts,  nor  did  they  have  the  crown  or 
nape  tinged  brown.  Instead  they  were  uni- 
formly dark  gray  (with  an  almost  steely  blue 
cast)  above  (except  for  the  usually  concealed 
white  interscapular  patch)  and  on  the  sides  of 
the  face,  with  the  chin,  throat,  breast,  belly, 
and  undertail  coverts  white.  The  sides  and 
flanks  were  washed  with  gray,  but  these  areas 
were  often  concealed  by  the  wings.  The  wing 
coverts  were  blackish,  with  the  lesser  and  me- 
dian wing  coverts  fringed  white.  The  tail  was 
blackish  and  the  rectrices  were  either  narrowly 
tipped  or  fringed  white.  This  character  is  not 
mentioned  in  any  of  the  previous  descriptions, 
possibly  because  the  narrow  white  tips/fringes 
were  lost  to  wear  in  the  few  existing  speci- 
mens. The  legs  were  pinkish  gray,  the  iris 
blackish  brown,  the  maxilla  blackish,  and  the 
mandible  whitish.  Females  differed  in  having 
a slightly  brownish  cast  to  the  upperparts,  par- 
ticularly the  crown  and  nape,  which  contrasted 
with  the  gray  sides  of  the  face.  The  underparts 
were  a bright  ochraceous  buff  (almost  pale  or- 
ange) from  the  throat  to  the  undertail  coverts, 
and  were  only  slightly  paler  on  the  chin  and 
upper  throat.  I did  not  note  pale  tips  or  fringes 
on  the  rectrices  of  females,  nor  could  I con- 
firm the  presence  of  a contrastingly  colored 
interscapular  patch. 

All  birds  that  I saw  had  proportionately 
short,  broad  tails  that  appeared  graduated. 
This  could  have  been  influenced  by  molt,  but, 
if  so,  all  of  the  individuals  in  this  area  were 
highly  synchronous  in  the  regrowth  of  their 
outer  rectrices. 

Vocalizations. — J.  Coons  and  D.  Stejskal 
(pers.  comm.)  tape  recorded  some  single-note 
harsh  calls  of  M.  disjimcta  during  their  Janu- 
ary 1997  visit  to  Picua,  but  did  not  encounter 
singing  birds.  Yapacana  Antbirds  were  also 
generally  quiet  during  the  period  of  my  field 
work.  Even  at  dawn  there  was  little  sponta- 


neous singing,  suggesting  that  the  level  of 
breeding  activity  was  low.  However,  birds 
were  highly  responsive  to  tape  playback, 
which  often  elicited  singing  from  one  or  more 
neighboring  pairs  in  addition  to  the  resident 
pair.  Both  males  and  females  responded  vo- 
cally to  playbacks,  although  males  were  much 
more  aggressive  in  approaching  the  speaker. 
All  types  of  vocalizations  that  I recorded  were 
given  by  both  sexes  of  Yapacana  Antbirds.  I 
tape  recorded  over  350  loudsongs  and  230 
calls  from  34  individuals. 

The  typical  loudsong  of  M.  disjimcta  con- 
sisted of  two  prolonged,  harsh,  heavily  fre- 
quency modulated  notes  separated  by  a short 
pause,  into  which  were  inserted  one  or  two 
“pip”  notes  (Fig.  5A).  The  first  harsh  note 
was  the  longest,  and  increased  in  intensity 
while  rising  in  pitch.  The  second  harsh  note 
was  shorter  and  had  a more  uniform  ampli- 
tude. A typical  loudsong  could  be  transcribed 
as  “cchhhhhhh  pipizhhhh”.  On  many  occa- 
sions birds  sang  songs  with  no  discernible 
“pip”  notes  between  the  harsh  elements,  al- 
though spectrograms  of  such  songs  reveal  a 
distinct  spike  at  the  beginning  of  the  second 
harsh  note  (Fig.  5B).  Less  frequently  birds 
sang  songs  with  thi'ee  harsh  elements  instead 
of  two  (Figs.  5C,  D).  The  third  harsh  note  in 
such  series  was  usually  the  shortest.  Loud- 
songs varied  in  duration,  depending  largely  on 
the  number  of  harsh  elements  included.  I 
could  find  no  consistent  differences  between 
male  and  female  songs,  although  males  more 
frequently  inserted  two  “pip”  notes  between 
harsh  elements  and  females  more  frequently 
sang  songs  with  no  “pip”  notes. 

The  most  commonly  heard  calls  were  long, 
harsh,  single  notes,  at  a somewhat  higher  pitch 
than  the  harsh  elements  of  the  song,  and  with 
a peculiar,  slightly  nasal  quality  (Figs.  5E,  E). 
These  “harsh  calls”  were  given  by  both  sexes, 
although  those  of  females  were  higher 
pitched.  “Harsh  calls”  seemed  to  be  aggres- 
sion calls  given  in  response  to  tape  playback 
or  by  a bird  disturbed  by  my  presence.  Birds 
occasionally  gave  a soft  rattle  (Eig.  5G)  in  a 
similar  context.  When  Yapacana  Antbirds  of 
either  sex  were  suddenly  startled  or  strongly 
agitated,  they  gave  one  or  more  sharp  and 
somewhat  squeaky  “squip”  notes  (Fig.  6A). 
These  notes  were  similar  in  tonal  quality  to 
the  notes  inserted  between  the  primaiy  harsh 


Frequency  (kHz) 


Zimmer  • CAURA  AND  YAPACANA  ANTBIRDS 


203 


7 

6 

5 - 
4 - 
3 

2 - 
1 - 


‘HiV  Jk  r ' u 


0 0.5  1.0  1.5  2.0  2.5  3.0  3.5  4.0  4.5  5.0  5.5  6.0 


Time  (seconds) 

FIG.  5.  Spectrograms  of  Yapacana  Antbird  (Myrmecizo  disjuncta)  vocalizations:  (A)  male  loudsong  with 
two  “pip”  notes  inserted  between  the  primary  harsh  elements,  (B)  loudsong  variant  with  no  discernible  “pip" 
notes  between  harsh  elements  (sex  of  singing  bird  unknown),  (C)  loudsong  variant  with  three  harsh  elements 
and  single  “pip”  notes  (sex  of  singing  bird  unknown),  (D)  loudsong  variant  with  three  harsh  elements  and  no 
“pip”  notes  (sex  of  singing  bird  unknown),  (E)  male  harsh  call,  probably  an  aggression  call,  (F)  female  harsh 
call,  probably  an  aggression  call,  and  (G)  soft  rattle  call,  given  in  an  aggressive  context. 


204 


THE  WILSON  BULLETIN  • Vol.  J 11,  No.  2,  June  1999 


FIG.  6.  Spectrograms  of  Yapacana  Antbird  (Myrmeciza  disjuncta)  calls:  (A)  “squip”  notes  given  by  a 
startled  or  alarmed  bird,  (B)  rarely  heard  complex  call  (sex  of  calling  bird  unknown),  and  (C)  variant  complex 
call  (sex  of  calling  bird  unknown). 


elements  of  most  loudsongs.  Occasionally 
birds  of  either  sex  gave  a more  complex  call 
that  began  with  several  “squip”  notes  and 
ended  with  a descending  series  of  soft,  whis- 
tled “wheee”  or  “whew”  notes  (Figs.  6B,  C). 
These  complex  calls  had  a distinct  tailing  off 
quality,  as  in  “squip  squip  squip  wheee  wheee 
wheee  whew  whew.”  The  function  of  these 
calls  was  not  clear. 

Behavior  and  sociality. — Yapacana  Ant- 
birds  were  encountered  singly  or  in  pairs  and 
did  not  associate  with  mixed-species  flocks. 
Given  the  relatively  small  size  of  their  terri- 
tories, members  of  pairs  were  rarely  far  from 
their  mates.  An  alarm  call  or  song  from  one 
bird  almost  invariably  brought  an  immediate 
vocal  response  from  its  mate.  I did  not  witness 
any  confrontations  between  neighboring  pairs 
of  antbirds,  although  on  several  occasions  a 
singing  pair  of  birds  stimulated  an  adjacent 
pair  to  approach  the  apparent  boundary  and 
countersing  for  several  minutes. 

Yapacana  Antbirds  typically  maintained  a 
horizontal  posture,  with  the  head  held  higher 
than  the  axis  of  the  body.  Singing  birds  usu- 
ally maintained  a more  upright  posture.  The 
tail  was  held  within  a few  degrees  above  or 
below  horizontal,  and  often  was  kept  slightly 
fanned.  Foraging  birds  often  quickly  flicked 
the  tail  up  and  down  in  a shallow  arc  of  less 
than  10°,  but  Just  as  frequently  dipped  the  tail 
slowly  downward  at  about  a 30°  angle  before 
flicking  it  back  up  more  rapidly.  Singing  birds 
frequently  shivered  the  tail  up  and  down  more 
rapidly  throughout  a song.  Some  individuals 
wagged  their  tail  sideways  in  a slow,  some- 


what jerky  manner.  Such  motions  involved  the 
entire  tail  being  swung  a few  degrees  away 
from  the  axis  of  the  body,  held  briefly  in  that 
position,  then  swung  still  further  in  the  same 
direction  before  being  swung  back  into  align- 
ment. This  jerky  motion  is  similar  to  some  of 
the  tail  movements  employed  by  the  Silvered 
Antbird  {Sclateria  naevia;  pers.  obs.).  Wheth- 
er foraging  or  singing,  Yapacana  Antbirds 
flicked  their  wings  at  least  once  during  virtu- 
ally every  pause  between  hops.  Wing-flicks 
occurred  both  independent  of  and  in  synchro- 
ny with  tail  movements.  Singing  birds  rarely 
sang  consecutive  songs  from  the  same  perch. 
In  response  to  tape  playback  males  often  ex- 
posed a white  interscapular  patch. 

Yapacana  Antbirds  foraged  mostly  on  or 
neai"  the  ground,  always  lower  than  1.5  m. 
They  were  restless,  active  foragers,  moving  by 
short  hops  (often  wing  assisted)  and  seldom 
pausing  for  more  than  2 s in  one  spot.  When 
moving  above  the  ground  they  clung  laterally 
to  slender  vertical  saplings  or  perched  across 
horizontal  limbs  and  vines,  progressed  in  an 
often  enatic,  zigzag  course,  and  frequently 
moved  up  and  down.  They  were  adept  at 
clinging  to  the  thinnest  stems,  including  slen- 
der bamboo  stalks  and  vines.  On  a few  oc- 
casions I saw  birds  hop  headfirst  down  nesuiy 
vertical  stems  or  branches,  almost  in  the  man- 
ner of  a nuthatch  (Sitta  spp.).  Birds  frequently 
took  several  hops  on  the  ground  before  jump- 
ing up  to  a low  perch  and  then  back  down  to 
the  ground.  Small  arthropod  prey  were 
gleaned  from  stems  and  from  tops  and  bot- 
toms of  live  leaves  by  reaching  out,  up,  or 


Zimmer  • CAURA  AND  YAPACANA  ANTBIRDS 


205 


down  on  extended  legs  and  with  neck  craned. 
Prey  were  captured  with  a quick  stabbing  mo- 
tion of  the  bill  and  swallowed  entire.  Larger 
prey  items  were  bashed  against  the  perch  and 
mandibled  one  or  more  times  before  being 
swallowed.  Antbirds  typically  wiped  their  bill 
on  the  perch  after  swallowing  prey.  Most  prey 
items  that  could  be  identified  were  small  or- 
thopterans  (katydids  and  crickets),  hemipter- 
ans,  and  geometrid  larvae. 

I encountered  Yapacana  Antbirds  foraging 
in  the  open  along  the  main  trail  on  only  three 
occasions.  Two  of  these  encounters  involved 
pairs,  and  the  other  involved  a lone  female 
plumaged  bird.  In  each  case  the  birds  were 
working  the  edge  of  the  dense  monte  cenado 
vegetation,  as  well  as  shrubs  and  clumps  of 
grass  growing  in  parts  of  the  trail.  All  foraged 
mostly  on  the  ground,  progressing  by  a series 
of  short  hops  with  minimal  pauses  in  between, 
and  always  with  wings  (and  frequently  the 
tail)  flicking.  The  most  frequent  attack  ma- 
neuvers were  gleans  from  the  surface  of  the 
leaf  litter  or  brief  probes  with  the  bill  beneath 
the  leaf  litter.  The  next  most  frequent  tech- 
nique was  reaching  up  to  glean  from  the  un- 
dersides of  overhanging  green  leaves  and 
grass  blades.  On  several  occasions  birds 
jumped  6-15  cm  upward  to  glean  prey  from 
the  undersides  of  leaves.  The  two  pairs  of  ant- 
birds encountered  in  the  open  were  found  in 
the  early  morning  before  the  sun  had  illumi- 
nated the  trail.  The  lone  female  plumaged  bird 
(possibly  a subadult  male)  was  found  in  mid- 
moming,  when  the  entire  trail  was  sunlit  and 
temperatures  were  already  above  30°  C.  I fol- 
lowed this  bird  as  it  foraged  steadily  at  the 
edge  of  the  woodland  for  more  than  20  min- 
utes and  covered  more  than  50  m.  It  crossed 
at  least  one  known  territorial  boundary,  but 
remained  silent  and  did  not  attract  attention 
from  any  other  antbird. 

On  one  occasion  I found  a female  Yapacana 
Antbird  attending  a foraging  swarm  of  army 
ants  {Eciton  sp.)  within  the  monte  ceirado.  I 
observed  this  bird  over  20  minutes  during 
which  it  was  the  only  bird  attending  the 
swarm.  The  female  antbird  employed  two 
strategies  in  the  vicinity  of  the  ants.  Pait  of 
the  time  she  scanned  the  swarm  from  perches 
within  0.3  m of  the  ground,  dropped  to  the 
ground  to  seize  fleeing  arthropods  (orthopter- 
ans,  hemipterans,  and  spiders)  and  then  re- 


turned to  a low  perch  to  beat  the  prey  on  a 
branch  before  swallowing.  Slightly  more  time 
was  spent  hopping  on  the  ground  between  the 
columns  of  ants  and  tossing  dead  leaves  in  the 
manner  of  a leaftosser  (Sclerurus  spp.).  Most 
leaves  were  tossed  by  inserting  the  bill  be- 
neath the  leaf  and  lifting  it  with  a quick  flak- 
ing motion.  Occasionally  the  bird  picked  up  a 
leaf  in  its  bill  and  tossed  it  aside.  David  Wolf 
(pers.  comm.)  observed  another  female  ant- 
bird (away  from  ants)  that  remained  in  one 
spot  tossing  leaves  in  a similar  manner  for  1- 
2 minutes. 

I observed  no  other  species  of  terrestrial  or 
semi-terrestrial  antbirds  in  the  monte  cerrado. 
The  Black- throated  Antbird  (Myrmeciza  atro- 
thorax)  and  the  Black-chinned  Antbird  {Hy- 
pocnemoides  melanopogon),  both  of  which 
routinely  forage  below  1.5  m (Hilty  and 
Brown  1986;  pers.  obs.),  were  locally  com- 
mon in  nearby  forest  or  edge  habitats,  but 
were  not  found  in  the  monte  cerrado.  The  only 
other  passerine  (besides  M.  disjuncta)  in  this 
habitat  that  I found  foraging  below  1 .5  m was 
the  Buff-breasted  Wren  (Thryothorus  leuco- 
tis),  which  foraged  everywhere  from  the 
ground  to  the  canopy. 

DISCUSSION 

Habitat  and  conservation. — My  field  work 
indicates  that  both  the  Caura  Antbird  and  the 
Yapacana  Antbird  are  habitat  specialists,  oc- 
cuiTing  in  subtypes  of  more  widely  distributed 
macrohabitats.  Percnostola  caurensis  has 
been  recorded  over  a broad  elevational  range, 
100-1300  m (Meyer  de  Schauensee  and 
Phelps  1978).  In  the  Pantepui  region  this 
range  of  elevations  often  spans  the  distance 
between  tall,  seasonal  humid  forest  and  elfin 
cloud  forest.  An  antbird  that  occurs  across 
such  a spectrum  of  habitats  might  normally  be 
considered  an  ecological  generalist.  However, 
if  the  critical  ecological  factor  determining  its 
distribution  is  the  presence  of  lai'ge  rocks 
within  forest  regardless  of  elevation,  then  P. 
caurensis  is  very  much  a specialist.  This 
could,  in  part,  account  for  the  absence  of  the 
species  from  so  many  seemingly  suitable  low- 
land sites  in  Bolivai'  and  Amazonas  (Campa- 
mento  Junglaven:  Zimmer  and  Hilty  1997; 
Brazo  Casiquiare:  Paynter  1982). 

Percnostola  caurensis  may  have  evolved  as 
something  of  a rock-specialist  to  occupy  a 


206 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  2,  June  1999 


niche  that  is  locally  abundant  in  parts  of  the 
highly  eroded  Guianan  Shield.  Mayr  and 
Phelps  (1967:277)  described  the  tepuis  (table 
top  mountains)  of  this  region:  “many  are  ac- 
tually strongly  dissected  and  strewn  with  iso- 
lated blocks,  some  more  than  100  m high,  and 
with  a large  variety  of  other  rock  forms.” 
Most  Caura  Antbirds  have  been  collected 
from  the  slopes  of  tepuis  (Zimmer  and  Phelps 
1947,  Phelps  and  Phelps  1963,  Meyer  de 
Schauensee  and  Phelps  1978).  The  most  re- 
cently collected  specimen,  from  1250  m at  the 
base  of  Pico  Maguire  (edo.  Amazonas,  Ven- 
ezuela), was  from  a site  described  as  “A 
hanging  valley  in  dense  cloud  forest  with 
moderately  tall  trees,  and  rocky  forest  floor 
covered  with  thick  moss”  (Willard  et  al.  1991: 
6).  The  apparent  absence  of  P.  caurensis  from 
much  of  the  eastern  portion  of  the  tepui  region 
of  Bolivar  is  a mystery. 

Myrmeciza  disjimcta  appeal's  to  be  restrict- 
ed to  woodlands  growing  on  white  sand  soils. 
White  sand  habitats  are  widely  but  patchily 
distributed  throughout  Amazonia,  with  their 
center  of  distribution  in  the  upper  Rfo  Negro 
region  (Pires  1974,  Stotz  et  al.  1996).  White 
sand  soils  support  many  different  types  of 
vegetation,  from  scrub  to  tall  forest  (Anderson 
1981).  Several  distinctly  different  types  of 
vegetation  were  found  growing  on  white  sand 
soils  in  the  Picua  region  but  I found  M.  dis- 
jimcta only  in  the  monte  cerrado.  Similarly, 
surveys  of  two  nearby  sites  in  Amazonas 
(Campamento  Junglaven  and  Pto.  Ayacucho) 
have  failed  to  record  M.  disjimcta  in  spite  of 
the  prevalence  of  white  sand  woodlands  (Zim- 
mer and  Hilty  1997).  The  Yapacana  Antbird 
may  therefore  be  restricted  only  to  a particular 
type  of  white  sand  woodland,  the  monte  cer- 
rado. This  would  indicate  an  even  patchier 
distribution  than  previously  suspected  and 
would  help  explain  how  this  species  has  es- 
caped detection  for  so  long. 

Both  the  Caura  Antbird  and  the  Yapacana 
Antbird  appear  to  be  locally  common  within 
their  prefeired  habitats.  The  distributions  of 
both  species  are  centered  in  the  lowlands  or 
foothills  of  Amazonas  and  western  Bolivar, 
which  are  among  the  least  populated  regions 
in  Amazonia.  Thus,  neither  species  is  under 
immediate  threat  of  extinction.  However,  their 
patchy  distributions  and  apparent  restriction  to 
particular  microhabitats  make  them  more  vul- 


nerable than  most  other  birds  of  the  Guianan 
lowlands.  Stotz  and  co-workers  (1996:4)  have 
noted  that  “.  . . the  first  major  waves  of  ex- 
tinctions in  the  Neotropics  are  not  occurring 
in  centers  of  diversity  such  as  the  Amazon. 
Rather,  extinctions  are  occuiring  within  cen- 
ters of  local  endemism,  especially  among  spe- 
cies that  have  evolved  ecological  specializa- 
tions that  limit  their  ability  to  adapt  to  human 
modifications  of  their  habitats.” 

Before  the  conservation  threats  facing  P. 
caurensis  and  M.  disjimcta  can  be  adequately 
assessed,  we  must  first  confirm  their  depen- 
dence on  or  preference  for  the  microhabitats 
in  which  I found  them,  then  attempt  to  quan- 
tify just  how  much  appropriate  habitat  exists 
within  their  ranges.  More  comprehensive  sur- 
vey work  within  the  region  is  cleaidy  needed. 
Continued  protection  of  existing  parks  or  re- 
serves, such  as  Yapacana  National  Paik,  is  vi- 
tal, pai'ticularly  with  regaid  to  threats  posed 
by  illegal  gold-mining. 

Intrafamilial  relationships. — The  Caura 
Antbird  was  described  by  Hellmayr  (1906) 
and  placed  in  the  genus  Sclateria.  The  Caura 
Antbird  was  subsequently  transfened  to  Schis- 
tocichla  (Zimmer  and  Phelps  1947),  which 
was  later  subsumed  into  Percnostola  without 
elaboration  by  Peters  (1951).  Subsequent  au- 
thors (e.g.,  Meyer  de  Schauensee  1966,  1970; 
Sibley  and  Mom'oe  1990;  Monroe  and  Sibley 
1993)  have  continued  this  treatment,  recog- 
nizing five  species:  P.  rufifrons,  P.  schistacea, 
P.  leiicostigma,  P.  caurensis,  and  P.  lophotes. 
Ridgely  and  Tudor  (1994  unected  the  ge- 
nus Schistocichla  for  schista  a,  leiicostigma, 
and  caurensis  on  the  basis  of  their  rounder, 
uncrested  heads  (crested  in  P.  rufifrons  and  P. 
lophotes)  and  spotted  rather  than  fringed  wing 
coverts. 

On  purely  moiphological  grounds,  P.  cau- 
rensis, P.  leiicostigma,  and  P.  schistacea 
would  appear  to  comprise  a natural  grouping. 
Plumage  differences  between  P.  caurensis  and 
P.  leiicostigma  are  especially  subtle,  with  size 
and  soft  part  coloration  being  the  most  im- 
portant field  characters  for  visually  distin- 
guishing the  two  species  (pers.  obs.).  Vocal 
similarities  are  less  apparent,  in  part  because 
of  pronounced  geographic  variation  in  the  vo- 
calizations of  the  various  named  subspecies  of 
P.  leiicostigma  (pers.  obs.).  Indeed,  some  vo- 
cal differences  within  the  P.  leiicostigma  com- 


Zimmer  • CAURA  AND  YAPACANA  ANTBIRDS 


207 


plex  are  as  great  as  the  between  species  dif- 
ferences in  the  Schistocichla  group.  Given 
this,  I feel  that  resolution  of  the  intrageneric 
relationships  of  the  five  species  cunently  in- 
cluded in  Percnostola  should  await  molecular 
comparisons,  as  well  as  a closer  evaluation  of 
vocal  and  morphological  differences  as  they 
relate  to  the  P.  leucostigma  group. 

Almost  since  its  description,  there  has  been 
speculation  regarding  the  placement  of  the  Ya- 
pacana  Antbird  in  Myrrneciza.  Friedmann 
(1948:478)  offered  that  “the  species  is  not  too 
distantly  related  to  Myrrneciza  atrothorax  but 
is  clearly  specifically  distinct  from  that  form.” 
He  went  on  to  note  that  Zimmer  had  examined 
the  type  and  the  paratype  and  had  pointed  out 
that  “the  general  plumage  has  about  the  tex- 
ture of  Cercornacra  carbonaria”  (Friedmann 
1948:478).  Friedmann  (1948)  further  noted 
Zimmer’s  suggestion  that  a fully  adult  male 
M.  disjimcta  might  show  a closer  relationship 
to  Cercornacra  than  was  suggested  by  the  type 
specimen.  He  also  commented  that  “The  pat- 
tern of  the  markings  of  the  upperwing  coverts 
is  very  like  that  of  some  forms  of  Cercornacra 
(serva  for  example),  but  the  bill  is  that  of 
Myrrneciza”  (Friedmann  1948:478).  Peters 
(1951)  alluded  to  the  seemingly  polyphyletic 
nature  of  Myrrneciza  as  he  defined  it  and  made 
several  recommendations  for  the  placement  of 
various  species,  but  did  not  mention  M.  dis- 
jimcta. Ridgely  and  Tudor  (1994:333)  also 
noted  the  heterogeneous  nature  of  Myrrneciza 
and  suggested  specifically  that  M.  disjuncta 
may  not  belong  in  the  genus  and  “perhaps  is 
more  closely  allied  to  Sclateria” . They  also 
seemed  to  suggest  somewhat  indirectly,  that 
M.  disjuncta  was  or  should  be  included  in  the 
formerly  recognized  genus  Myrmoderus. 
However,  I can  find  no  evidence  that  M.  dis- 
juncta was  included  in  the  various  shifts  of 
species  between  Myrrneciza  and  Myrmoderus 
by  Hellmayr  (in  Cory  and  Hellmayr  1924), 
Todd  (1927),  or  Peters  (1951). 

Morphological,  vocal,  and  behavioral  char- 
acters offer  contradictory  clues  to  the  possible 
generic  affiliations  of  the  Yapacana  Antbird. 
Myrmoderus  is  not  currently  recognized,  but 
both  Todd  ( 1 927)  and  Peters  (1951)  advocated 
that  it  be  reserved  for  [Myrrneciza]  loricata 
and  squamosa.  The  latter  are  clearly  sibling 
species  that  share  several  distinctive  morpho- 
logical, vocal,  and  behavioral  characters,  and 


they  are  distant  in  all  respects  from  M.  dis- 
juncta (pers.  obs.).  In  some  moiphological  re- 
spects, M.  disjuncta  is  reminiscent  of  the 
monotypic  Sclateria,  as  suggested  by  Ridgely 
and  Tudor  (1994),  but  it  has  fringed  rather 
than  spotted  wing  coverts,  white  tail  tips,  an 
interscapular  patch,  and  differs  greatly  in  both 
vocal  and  behavioral  characters.  The  song  of 
M.  disjuncta  is  mildly  reminiscent  of  that  of 
various  members  of  the  Cercornacra  nigricans 
group  (as  defined  by  Fitzpatrick  and  Willard 
1990),  and  the  plumage  pattern  of  males, 
lai'gely  gray  with  white-fringed  wingbars  and 
white  tail-tips,  and  females  ochraceous  below, 
fits  several  members  of  the  C.  tyrannina 
group.  However,  Cercornacra  antbirds  tend  to 
be  slender  and  proportionately  long-tailed 
(Ridgely  and  Tudor  1994),  whereas  M.  dis- 
juncta is  relatively  compact  and  short-tailed. 
No  Cercornacra  approaches  the  white  under- 
parts of  male  M.  disjuncta,  the  genus  as  a 
whole  has  gray  or  blackish  underparts.  More 
importantly,  male-female  antiphonal  duets  are 
an  important  component  of  the  vocal  reper- 
toires of  virtually  all  species  of  Cercornacra 
(Zimmer  et  al.  1997),  but  are  not  found  in  M. 
disjuncta. 

In  size,  proportions,  and  some  aspects  of 
plumage  Myrrneciza  disjuncta  is  somewhat 
suggestive  of  Hypocnemoides.  However,  nei- 
ther species  of  Hypocnemoides  is  nearly  as 
sexually  dimoiphic  as  is  M.  disjuncta,  and 
they  lack  any  suggestion  of  the  ochraceous 
coloration  found  in  female  M.  disjuncta.  Myr- 
meciza  disjuncta  lacks  the  black  throat  and 
pale  eye  found  in  Hypocnemoides.  Vocal  dif- 
ferences between  M.  disjuncta  and  Hypocne- 
moides are  much  greater  than  the  moipholog- 
ical  differences  (pers.  obs.).  By  themselves, 
the  two  species  of  Hypocnemoides  form  a nat- 
ural grouping,  with  great  similarities  in  plum- 
age, voice,  foraging  behavior,  and  habitat  use. 
Almost  none  of  these  characters  are  shared 
with  M.  disjuncta. 

I feel  that  Myrrneciza  as  cunently  con- 
structed is  paraphyletic,  with  various  sub- 
groupings that  do  not  appear  to  be  closely  al- 
lied on  the  basis  of  morphological,  vocal,  or 
behavioral  characters  (e.g.,  Peters  1951, 
Ridgely  and  Tudor  1994).  None  of  these  sub- 
groups is  a good  fit  for  M.  disjuncta.  Myr- 
meciza  atrothorax  has  been  suggested  as  a 
close  relative  of  M.  disjuncta  by  Friedmann 


208 


THE  WILSON  BULLETIN 


VoL  111,  No.  2,  June  1999 


(1948),  but  the  males  of  the  two  species  differ 
dramatically  in  plumage,  and  the  two  species 
share  no  vocal  similarities  that  I can  detect. 
No  compelling  morphological  similarities 
clearly  ally  M.  disjuncta  with  any  other  Myr- 
meciza,  nor  with  any  other  antbird.  Similarly, 
I have  compared  songs  and  calls  of  M.  dis- 
juncta to  the  other  1 8 species  currently  placed 
in  Myrmeciza  (Isler  and  Whitney  1999;  Zim- 
mer, unpubl.  data)  and  can  find  nothing  to  sug- 
gest a close  relationship  between  M.  disjuncta 
and  any  of  the  other  species.  Relationships 
suggested  by  one  or  two  morphological  char- 
acters in  one  sex  are  contradicted  by  morpho- 
logical characters  in  the  other  sex,  vocal  char- 
acters, behavioral  characters,  or  by  some  com- 
bination of  the  three.  Although  past  descrip- 
tions of  genera  have  been  based  largely  on 
plumage  characters  which  may  or  may  not 
have  phylogenetic  relevance,  the  addition  of 
vocal  and  behavioral  data  adds  important  ev- 
idence in  redefining  these  relationships  (e.g, 
comments  in  Remsen  1997,  Remsen  and 
Schulenberg  1997).  In  the  absence  of  a mo- 
lecular based  phylogeny  the  most  conserva- 
tive approach  would  be  to  leave  M.  disjuncta 
where  it  is,  as  yet  another  poor  fit  in  a genus 
understood  to  be  heterogeneous.  However,  in 
my  opinion,  the  sum  of  morphological,  vocal, 
and  behavioral  evidence  would  suggest  that 
the  Yapacana  Antbird  is  monotypic,  deserving 
of  its  own  genus. 

ACKNOWLEDGMENTS 

Primary  thanks  go  to  M.  and  P.  Isler  for  generously 
lending  their  time  and  talents  to  the  many  figures  in 
this  paper.  Phyllis  produced  the  many  spectrograms  af- 
ter reviewing  several  hours  of  my  tape  recordings,  and 
Mort  compiled  the  list  of  distributional  records  and 
produced  the  map  used  in  Eigure  1.  The  Islers  made 
many  helpful  comments  on  an  early  draft  of  this  man- 
uscript and  have  been  an  invaluable  source  of  advice 
and  ideas  with  all  aspects  of  my  research  on  antbirds. 
M.  Lentino,  C.  Rodner,  and  R.  Restall  of  the  Colecion 
Ornitoldgica  Phelps  generously  supplied  important 
collection  information.  T.  Schulenberg  provided  valu- 
able advice  as  well  as  access  to  several  references.  J. 
V.  Remsen  reviewed  an  early  draft  of  the  manuscript 
and  made  numerous  helpful  comments  that  improved 
it.  1 am  most  grateful  to  J.  Coons  for  sharing  infor- 
mation on  his  experiences  in  the  Picua  region,  partic- 
ularly for  bringing  Alechiven  Lodge  to  my  attention. 
D.  Cooper  was  the  first  to  report  the  probable  presence 
of  Caura  Antbirds  in  the  Serrania  dc  la  Cerbetana  and 
his  advice  was  helpful  in  planning  my  trip  to  that  re- 
gion. Thanks  also  to  the  staffs  of  Hato  Las  Nieves  and 


Alechiven  Lodge,  my  two  bases  of  operation  for  this 
research.  L.  Oleaga  was  of  particular  help  in  cutting 
trails  and  facilitating  my  work  at  Hato  Las  Nieves.  C. 
Herrera,  M.  Ruiz  and  D.  Eorbes  of  Caracas  provided 
invaluable  assistance  with  trip  logistics. 

Special  thanks  must  go  to  D.  Wolf,  who  accompa- 
nied me  on  my  trip  to  Yapacana  National  Park.  Dave 
shared  in  the  excitement  of  finding  the  Yapacana  Ant- 
birds and  taping  their  vocalizations;  his  good  humor 
and  knowledgeable  insights  made  the  heat  and  the  bit- 
ing jejenes  more  tolerable.  Pinal  thanks  go  to  V.  Eman- 
uel and  Victor  Emanuel  Nature  Tours,  Inc.,  for  provid- 
ing me  with  travel  opportunities  that  made  much  of 
my  research  possible. 

LITERATURE  CITED 

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

Cory,  C.  B.  and  C.  E.  Hellmayr.  1924.  Catalogue  of 
birds  of  the  Americas.  Pteroptochidae,  Conopo- 
phagidae,  Pormicariidae.  Pield  Mus.  Nat.  Hist. 
(Zool.  Series)  13(3):  1-369. 

PiTZPATRiCK,  J.  W.  AND  D.  E.  WiLLARD.  1990.  Ccrcom- 
acra  manu,  a new  species  of  antbird  from  south- 
western Amazonia.  Auk  107:239-245. 
pRiEDMANN,  H.  1945.  A new  ant-thrush  from  Venezue- 
la. Proc.  Biol.  Soc.  Wash.  58:83-84. 
pRiEDMANN,  H.  1948.  Birds  collected  by  the  National 
Geographic  Society’s  expeditions  to  northern  Bra- 
zil and  southern  Venezuela.  Proc.  U.S.  Nat.  Mus. 
97:373-569. 

Hellmayr,  C.  E.  1906.  [Descriptions  of  new  South 
American  birds.]  Bull.  Brit.  Ornithol.  Club.  19: 
8-9. 

Hilty,  S.  L.  and  W.  L.  Brown.  1986.  A guide  to  the 
birds  of  Colombia.  Princeton  Univ.  Press,  Prince- 
ton, New  Jersey. 

Isler,  M.  L.  1997.  A sector-based  ornithological  geo- 
graphic information  system  for  the  Neotropics. 
Ornithol.  Monogr.  48:345-354. 

Isler,  M.  L.,  P.  R.  Isler,  and  B.  M.  Whitney.  1997. 
Biogeography  and  systematics  of  the  Thamnophi- 
lus  punctatiis  (Thamnophilidae)  complex.  Orni- 
thol. Monogr.  48:355—382. 

Isler,  P.  R.  and  B.  M.  Whitney.  1999.  Voices  of  the 
antbirds:  Thamnophilidae,  Pormicariidae,  and 
Conopophagidae.  Library  of  Natural  Sounds,  Cor- 
nell Laboratory  of  Ornithology,  Ithaca,  New  York. 
Mayr,  E.  and  W.  H.  Phelps,  Jr.  1967.  The  origin  of 
the  bird  fauna  of  the  south  Venezuelan  highlands. 
Bull.  Am.  Mus.  Nat.  Hist.  136:269-328. 

Meyer  de  Schauen.see,  R.  1966.  The  species  of  birds 
of  South  America  with  their  distribution.  Living- 
ston Press,  Narbcth,  Pennsylvania. 

Meyer  de  Schauensee,  R.  1970.  A guide  to  the  birds 
of  South  America.  Livingston  Press,  Narbeth, 
Pennsylvania. 

Meyer  de  Schauensee,  R.  and  W.  H.  Phelps,  Jr. 
1978.  A guide  to  the  birds  of  Venezuela.  Princeton 
Univ.  Press,  Princeton,  New  Jersey. 

Monroe,  B.  L.,  Jr.  and  C.  G.  Sibley.  1993.  A world 


Zimmer  • CAURA  AND  YAPACANA  ANTBIRDS 


209 


checklist  of  birds.  Yale  Univ.  Press,  New  Haven, 
Connecticut. 

Parker,  T.  A.,  Ill,  D.  F.  Stot/,  and  J.  W.  Fitzpatrick. 
1996.  Ecological  and  distributional  databases.  Pp. 
1 13-436  in  Neotropical  birds:  ecology  and  con- 
servation (D.  F Stotz,  J.  W.  Fitzpatrick,  T.  A. 
Parker,  III,  and  D.  K.  Moscovits,  Eds.).  Univ.  of 
Chicago  Press,  Chicago,  Illinois. 

Paynter,  R.  a.  1982.  Ornithological  gazetteer  of  Ven- 
ezuela. Museum  of  Comparative  Zoology,  Cam- 
bridge, Massachussetts. 

Peters,  J.  L.  1951.  Checklist  of  birds  of  the  world. 
Vol.  VII.  Harvard  Univ.  Press,  Cambridge,  Mas- 
sachussetts. 

Phelps,  W.  H.  and  W.  H.  Phelps,  Jr.  1963.  Lista  de 
las  aves  de  Venezuela  con  su  distribucion.  Parte 
II.  Passeriformes,  second  ed.  Bol.  Soc.  Venez. 
Cienc.  Nat.  24:1—479. 

PiRES,  J.  M.  1974.  Tipos  de  vegetacao  da  Amazonia. 
Publ.  Avul.  Mus.  Goeldi.  20:  179-202. 

Remsen,  j.  V.,  Jr.  1997.  A new  genus  for  the  Yellow- 
shouldered Grosbeak.  Ornithol.  Monogr.  48:89— 
90. 

Remsen,  J.  V.,  Jr.  and  S.  K.  Robinson.  1990.  A clas- 
sihcation  scheme  for  foraging  behavior  of  birds  in 
terrestrial  habitats.  Stud.  Avian  Biol.  13:144—160. 

Remsen,  J.  V.,  Jr.  and  T.  S.  Schulenberg.  1997.  The 
pervasive  influence  of  Ted  Parker  on  Neotropical 
field  ornithology.  Ornithol.  Monogr.  48:7-20. 


Ridgely,  R.  S.  and  G.  Tudor.  1994.  The  birds  of 
South  America.  Vol.  2.  Univ.  of  Texas  Press.  Aus- 
tin. 

Sibley,  C.  G.  and  B.  L.  Monroe.  1990.  Distribution 
and  taxonomy  of  birds  of  the  world.  Yale  Univ. 
Press,  New  Haven,  Connecticut. 

Snow,  D.  1982.  The  cotingas.  Cornell  Univ.  Press,  Ith- 
aca, New  York. 

Stotz,  D.  E,  J.  W.  Fitzpatrick,  T.  A.  Parker,  III,  and 
D.  K.  Moskovits.  1996.  Neotropical  birds:  ecol- 
ogy and  conservation.  Univ.  of  Chicago  Press, 
Chicago,  Illinois. 

Todd,  W.  E.  C.  1927.  New  gnateaters  and  antbirds 
from  tropical  America,  with  a revision  of  the  ge- 
nus Myrmeciza  and  its  allies.  Proc.  Biol.  Soc. 
Wash.  40:149-178. 

Willard,  D.  E.,  M.  S.  Foster,  G.  F.  Barrowcloligh, 
R.  W.  Dickerman,  P.  F.  Cannell,  S.  L.  Coats,  J. 
L.  Cracraft,  and  j.  P.  O’Neill.  1991.  The  birds 
of  Cerro  de  la  Neblina,  Territorio  Federal  Ama- 
zonas, Venezuela.  Fieldiana  Zool.  65:1-90. 

Zimmer,  J.  T.  and  W.  H.  Phelps.  1947.  Seven  new 
subspecies  of  birds  from  Venezuela  and  Brazil. 
Am.  Mus.  Novit.  1338:1-7. 

Zimmer,  K.  J.  and  S.  L.  Hilty.  1997.  Avifauna  of  a 
locality  in  the  upper  Orinoco  drainage  of  Ama- 
zonas, Venezuela.  Ornithol.  Monogr.  48:865-886. 

Zimmer,  K.  J.,  A.  Whittaker,  and  D.  F.  Stotz.  1997. 
Vocalizations,  behavior  and  distribution  of  the  Rio 
Branco  Antbird.  Wilson  Bull.  109:663—678. 


Wilson  Bull.,  111(2),  1999,  pp.  210-215 


HABITAT  PATCH  SIZE  AND  NESTING  SUCCESS  OF 
YELLOW-BREASTED  CHATS 

DIRK  E.  BURHANS'  - AND  FRANK  R,  THOMPSON  IIP 


ABSTRACT. — We  measured  vegetation  at  shrub  patches  used  for  nesting  by  Yellow-breasted  Chats  {Icteria 
virens)  to  evaluate  the  importance  of  nesting  habitat  patch  features  on  nest  predation,  cowbird  parasitism,  and 
nest  site  selection.  Logistic  regression  models  indicated  that  nests  in  small  patches  (average  diameter  <5.5  m) 
that  were  parasitized  by  Brown-headed  Cowbirds  (Molothrus  ater)  experienced  higher  predation  than  unpara- 
sitized nests  in  large  patches.  Nests  in  large  patches  were  more  likely  to  become  parasitized  by  cowbirds,  as 
were  nests  with  more  large  stems  (>10  cm  dbh)  nearby.  Patches  used  by  chats  for  nesting  had  larger  average 
diameters  than  unused  patches  and  tended  to  contain  more  small  stems.  Chats  appeared  to  prefer  large  patches 
and  experienced  lower  nest  predation  there.  Although  they  might  experience  higher  brood  parasitism  frequencies 
in  large  patches,  losses  to  parasitism  were  balanced  by  higher  nesting  success  because  the  mean  number  of  chat 
young  that  fledged  did  not  differ  between  nests  in  small  versus  large  patches.  Received  12  Jan.  1998,  accepted 
28  Dec.  1998. 


The  nest  “patch”  has  been  defined  as  the 
habitat  patch  immediately  sunounding  the 
nest  (Martin  and  Roper  1988).  Characteristics 
of  the  songbird  nesting  patch  may  differ  from 
the  habitat  available  (Martin  and  Roper  1988; 
Kelly  1993;  Steele  1993;  Kligo  et  al.  1996a, 
b)  and  there  may  be  differences  between  suc- 
cessful and  unsuccessful  nests  according  to 
nest  patch  characteristics  (Martin  and  Roper 
1988,  Kelly  1993,  Norment  1993,  Tarvin  and 
Smith  1995).  However,  there  is  no  consensus 
on  exactly  what  determines  a nest  patch.  Pe- 
tersen and  Best  (1985)  and  Martin  and  Roper 
( 1988)  defined  the  nest  patch  as  the  area  with- 
in 5 m of  the  nest,  a criterion  that  other  studies 
since  have  adopted  (Kligo  et  al.  1996a,  b;  Bar- 
ber and  Martin  1997).  Other  workers  have 
evaluated  nest  patches  based  upon  other  pre- 
determined sizes  (Conner  et  al.  1986,  Kelly 
1993,  Norment  1993,  Tarvin  and  Smith  1995), 
multiple  radius  patch  sizes  (Petit  et  al.  1988, 
Holway  1991,  With  1994),  or  stem  density 
(Hoi way  1991,  Knopf  and  Sedgewick  1992). 
Knopf  and  Sedgwick  ( 1992)  based  their  patch 
definition  upon  vegetation  height  and  radius 
descriptors  rather  than  upon  pre-determined 
size,  and  concluded  that  individual  plants 
probably  are  functionally  indistinguishable  to 
Yellow  Warblers  (Dendroicci  petechia),  which 


' North  Central  Research  .Station,  USDA  Forest  Ser- 
vice, 202  Natural  Resources  Building,  Univ.  of  Mis- 
souri, Columbia.  MO  6521  I. 

- Corresponding  author; 

E-mail:  dburhan,s/nc_co@fs.fcd.us 


select  nests  based  on  patch  characteristics 
rather  than  the  nest  plant. 

We  examined  the  relationship  between  nest 
patch  characteristics  and  nest  predation,  brood 
parasitism,  and  nest  site  selection  for  the  Yel- 
low-breasted Chat  {Icteria  virens).  Yellow- 
breasted Chats  are  a common  songbird  of 
shrub  habitats  (Nolan  1963,  Thompson  and 
Nolan  1973)  and  at  our  sites  often  nested  in 
conspicuous  dense  thickets  of  shrubs.  We 
combined  two  approaches  by  measuring  veg- 
etation structure  in  a fixed-radius  plot  centered 
on  the  nest  and  measuring  dimensions  of  the 
shrub  patch  in  which  the  nest  was  located.  Our 
principle  questions  were:  ( 1 ) ai'e  chat  nests  in 
large  thickets,  or  patches,  more  likely  to 
fledge  young  than  nests  in  small  patches  or 
single  shrubs  and  trees?  and  (2)  are  chat  nests 
that  are  placed  further  from  the  edge  of  the 
nesting  patch  more  likely  to  fledge  young?  We 
predicted  that  chats  nesting  in  larger  patches 
at  greater  distances  from  the  patch  edge  would 
be  more  likely  to  avoid  predation  because 
large  patches  may  impede  the  movements  of 
predators  (Bowman  and  Hanis  1980,  Holway 
1991).  Additionally,  we  predicted  that  nests 
near  greater  numbers  of  trees  would  experi- 
ence higher  frequencies  of  cowbird  paiasitism 
because  Brown-headed  Cowbirds  {Molothrus 
ater)  use  trees  to  aid  in  finding  nests  (Ander- 
son and  Storer  1976,  Romig  and  Crawford 
1995,  Clotfelter  1998).  We  also  predicted  that 
size  of  nest  patches  would  differ  from  the  size 
of  patches  selected  at  random.  We  tested  these 
predictions  by  monitoring  nest  success  and 


210 


Bmiums  and  Thompson  • YELLOW-BREASTED  CHAT  NESTING  I’ATCH 


21 1 


cowbird  parasitism  of  chats  and  by  measuring 
vegetation  at  nest  sites  and  unused  sites. 

METHODS 

We  found  Yellow-breasted  Chat  nests  at  Thomas 
Baskett  Wildlife  Research  and  Education  Center  near 
Ashland,  (Boone  County)  Missouri,  from  1992-1994 
as  part  of  a study  of  shrubland  birds.  Study  sites  were 
six  old  fields  ranging  from  2.4  to  16.3  ha  and  surround- 
ed by  oak-hickory  forest  (see  Burhans  1997  for  de- 
tailed site  description).  We  monitored  nests  every  3-4 
days  and  daily  toward  the  end  of  the  nestling  period. 
We  considered  nests  that  avoided  predation  and  suc- 
ceeded in  Hedging  either  chat  or  cowbird  young  as 
"fledged”.  In  most  cases  fledged  nests  were  identified 
by  observing  adults  carrying  food  or  scolding,  or  by 
observing  fledglings.  Nests  that  were  empty  on  the 
fledging  day  (day  8,  where  day  of  hatching  = day  0) 
were  classified  as  Hedged  if  they  were  active  the  day 
before.  We  classified  nests  that  were  empty  prior  to 
this  time  as  depredated  unless  there  were  signs  of  pre- 
mature Hedging,  such  as  nearby  fiedglings  or  adult 
feeding  activity.  Parasitism  status  was  determined  for 
all  nests  and  only  those  nests  that  were  initiated  during 
the  period  of  cowbird  parasitism  (before  the  second 
week  of  July)  were  considered  in  the  parasitism  anal- 
ysis. 

Vegetation  samples  were  taken  at  nest  sites  and  un- 
used sites  at  the  end  of  the  nesting  season.  We  mea- 
sured nest  height  to  the  bottom  of  the  nest  cup.  We 
also  measured  nest  “patch”,  which  was  defined  as  in- 
terlocking leafy  shrub  or  tree  vegetation  at  nest  height 
within  which  the  nest  plant  was  situated.  Nest  patches 
varied  in  size  from  the  single  nest  tree  or  shrub  to  an 
entire  fencerow.  We  measured  length  and  width  of 
patches  to  the  nearest  0. 1 m for  distances  within  3 m 
and  paced  (calibrated  at  1 m/pace)  to  the  nearest  m for 
greater  distances.  “Average  patch  diameter”  was  the 
sum  of  the  length  of  the  nest  vegetation  clump  plus 
the  width  of  the  clump  divided  by  two.  Nest  patch 
diameter  varied  greatly  among  patches  (median  = 5.5 
m,  range  0.3-65  m)  so  we  grouped  nests  into  “large” 
(S:5.5  m)  or  “small”  patches  for  analyses  (see  below). 
“Patch-edge  distance”  was  the  distance  (to  the  nearest 
0.1  m)  from  the  outside  rim  of  the  nest  cup  to  the 
nearest  leafy  edge  of  the  nest  patch.  In  order  to  further 
characterize  patches  and  evaluate  potential  cowbird 
perches,  we  counted  woody  stems  1 1—20,  21—50,  and 
greater  than  50  cm  dbh  (diameter  at  breast-height)  in 
an  11.3  m radius  circle  centered  on  each  nest.  We 
counted  shrub  and  sapling  stems  (^1  m high)  in  a 5 
m radius  circle  around  each  nest  in  categories  less  than 
2,  2-5,  and  at  least  5 cm  dbh.  Many  chat  nests  were 
placed  in  large  blackberry  (Rnbus  allegheniensis) 
patches  in  which  it  was  difficult  to  count  stems.  For 
large  blackberry  patches  (>10%  of  the  circle)  we  es- 
timated number  of  blackberry  stems  by  counting  the 
number  of  stems  in  a square  meter  and  extrapolating 
to  the  proportion  of  the  5 m circle  that  was  blackberry. 

Unused  sites  were  located  by  pacing  in  a randomly 


determined  compass  direction  to  the  first  plant  en- 
countered of  the  same  species  and  size  category  as  the 
nest  plant  (at  least  40  m from  the  nest).  As  with  nest 
sites,  we  took  patch  diameter  and  stem  count  mea- 
surements for  unused  patches.  We  did  not  sample  veg- 
etation for  10  nests  destroyed  by  Hooding  in  1993  and 
storms  in  1994  and  did  not  include  these  nests  in  the 
analysis.  We  also  omitted  2 nests  found  immediately 
before  Hedging  where  it  was  not  possible  to  inspect 
chicks  to  determine  parasitism  status  without  forcing 
Hedging. 

Data  analyses. — We  evaluated  Hedging  success  us- 
ing both  simple  nesting  success  (number  of  successful 
nests/total  nests)  and  the  Mayfield  method  (Mayfield 
1961,  1975).  For  the  Mayfield  method  half  the  number 
of  days  between  subsequent  visits  over  which  a nest 
was  empty  were  added  to  the  number  of  previous  days 
the  nest  survived  to  obtain  the  total  number  of  obser- 
vation days  for  a nest.  When  calculating  daily  survival 
probabilities  we  only  included  mortality  caused  by  nest 
predation.  We  calculated  survival  probabilities  and 
variances  with  standard  errors  according  to  Johnson 
(1979).  We  compared  survival  probabilities  using 
CONTRAST  (DOS;  Sauer  and  Williams  1989).  An- 
other species  that  nested  at  this  site  (Indigo  Bunting; 
Passerina  cyanea)  suffered  higher  predation  at  para- 
sitized nests  (Dearborn  in  press),  so  we  compared  daily 
survival  probabilities  between  parasitized  and  unpar- 
asitized nests.  Simple  nesting  success  was  used  for  lo- 
gistic regression  models  (below).  Nests  that  Hedged  at 
least  one  chick  (chat  or  cowbird)  were  considered 
“Hedged”.  When  calculating  mean  number  of  chat 
young  Hedged,  we  assumed  that  the  number  success- 
fully Hedged  was  equal  to  the  number  of  chicks  last 
counted  in  the  nest.  We  compared  mean  number  of 
chat  chicks  Hedged  from  nests  in  large  versus  small 
patches  with  an  independent  sample  two-tailed  r-test. 

We  analyzed  both  nest  predation  and  nest  parasitsm 
with  logistic  regression  models.  Nest  height,  patch  dis- 
tance. average  patch  diameter,  stems  defined  as  above, 
total  stems  10  cm  dbh  or  smaller  (“total  small  stems”), 
and  cowbird  parasitism  status  (parasitized  or  not)  were 
evaluated  in  the  nest  predation  model.  Frequency  of 
parasitism  has  been  related  to  nest  height  and  nest  veg- 
etation (Hahn  and  Hatfield  1995,  Britlingham  and 
Temple  1996),  so  we  similarly  used  logistic  regression 
to  analyze  parasitism  against  nest  height,  average  patch 
diameter,  patch  distance,  mean  stems  at  least  10-20, 
21-50,  greater  than  50  cm  dbh.  and  combined  stems 
greater  than  10  cm  dbh  (“total  large  stems").  Model 
building  for  both  nest  predation  and  parasitism  models 
followed  the  method  of  Hosmer  and  Lemeshow  ( 1989) 
and  consisted  of  running  univariate  logistic  regression 
models  and  retaining  variables  with  P-values  of  0.25 
or  less  in  a full  model.  The  final  reduced  models  in- 
cluded those  variables  with  P < 0.05.  Decisions  about 
which  variables  should  be  left  in  final  models  were 
based  on  probability  values  for  individual  variables 
from  a set  of  alternative  multivariate  models.  We  per- 
formed Hosmer  and  Lemeshow  ( 1989)  goodness-of-fit 
tests  on  the  final  models. 


212 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


TABLE  1.  Parameter  estimates,  Wald  X"  statistics  and  probability  levels  for  final  nest  predation  (top)  and 
nest  parasitism  (bottom)  logistic  regression  models. 

ModelA'ariable 

Parameter 

-) 

X- 

p 

Nest  predation 

Parasitism 

-2.23 

5.44 

0.02 

Average  patch  diameter 

-1.52 

4.24 

0.04 

Nest  parasitism 

Average  patch  diameter 

2.28 

5.50 

0.02 

Total  large  stems  (>10  cm  dbh) 

0.25 

5.36 

0.02 

Nest  sites  and  unused  sites  were  compared  with  lo- 
gistic regression  rather  than  discriminant  function  be- 
cause of  the  presence  of  binary  explanatory  variables 
and  non-normal  variance  of  other  variables  (Press  and 
Wilson  1978,  James  and  McCullogh  1990).  Nest  and 
unused  site  variables  were  screened  with  univariate  lo- 
gistic regression  models,  and  multivariate  models  were 
developed  similarly  to  the  predation  and  brood  para- 
sitism models  (above).  Results  for  statistical  tests  are 
reported  as  mean  ± SE. 

RESULTS 

Nesting  success  and  predation/parasitism 
models. — Daily  survival  estimates  of  Yellow- 
breasted Chat  nests  did  not  differ  among  years 
from  1992-1994  (1992:  0.96  ± 0.01,  1993: 
0.94  ± 0.02,  1994:  0.95  ± 0.02;  x'  = 0.7,  df 
= 2,  P > 0.05)  so  data  from  all  nests  were 
pooled  for  the  predation  analysis  (0.95  ± 

0. 01;  n = 48  nests).  Brood  parasitism  fre- 

quency was  33%  (n  = 15  nests),  36%  (n  = 
14  nests)  and  23%  (n  = 13  nests)  for  1992, 
1993,  and  1994  and  did  not  differ  between 
years  (Fisher  exact  test;  P > 0.05).  Cowbird 
parasitism  averaged  3 1 % over  all  years  during 
the  seasonal  period  of  parasitism  (n  = 42 
nests).  Parasitized  nests  did  not  have  signifi- 
cantly different  survival  rates  than  unparasit- 
ized nests  (parasitized  nests  0.94  ± 0.02;  un- 
parasitized nests  0.96  ± 0.01;  ~ 1-2,  df  = 

1,  P > 0.05). 

Nest  predation  was  best  explained  by  a final 
logistic  regression  model  including  parasitism 
status  and  average  patch  diameter  (Table  1; 
Log  likelihood  for  model  = 51.8,  x^  = 8.7,  df 
= 2,  P = 0.01).  Nests  that  were  parasitized 
and  in  small  patches  were  more  likely  to  suf- 
fer predation.  However,  the  nest  parasitism 
model  indicated  that  nests  in  large  patches 
were  more  likely  to  become  parasitized.  The 
nest  parasitism  model  included  the  variables 
average  patch  diameter  and  total  large  stems 
(Table  1;  Log  likelihood  for  model  = 40.9,  x^ 


= 1 1.1,  df  = 2,  P = 0.004).  Nests  with  more 
large  stems  were  more  likely  to  be  parasitized, 
but  large  patches  did  not  have  greater  mean 
values  for  total  large  stems  than  did  small 
patches  (large  patches  2.27  ± 0.67;  small 
patches  3.45  ± 1.01;  / = 0.99,  df  = 40,  P > 
0.05).  Distance  from  the  nest  to  the  edge  of 
the  patch  tended  to  be  greater  for  paiasitized 
nests  (Table  2),  but  was  eliminated  from  the 
parasitism  models  because  of  the  higher  prob- 
ability values  associated  with  average  patch 
diameter,  with  which  patch-edge  distance  was 
positively  conelated  prior  to  transformation  of 
the  former  variable  (r  = 0.39,  P = 0.009). 
Mean  number  of  chat  young  fledged  did  not 
vary  between  nests  in  small  versus  large 
patches  (small  patches:  1.04  ± 0.34  chat 
young  per  nest;  large  patches  1.43  ± 0.36  chat 
young  per  nest;  t = —0.79,  df  = 44,  P > 
0.05). 

Nest  sites  versus  unused  sites. — Univariate 
logistic  regression  models  indicated  that  nest 
sites  were  situated  in  larger  patches  than  un- 
used sites  (Table  3).  When  variables  were 
combined  in  the  multivariate  model  only  av- 
erage patch  diameter  was  significant  (Log 
likelihood  for  model  = 121.07,  X“  = 12.0,  P 
= 0.001). 

DISCUSSION 

As  predicted,  logistic  regression  models  in- 
dicated that  Yellow-breasted  Chats  experi- 
enced less  predation  in  larger  nest  patches.  As 
with  Indigo  Buntings  at  these  sites  (Dearborn, 
in  press),  predation  was  related  to  pai'asitism 
status  at  Yellow-breasted  Chat  nests;  nests  that 
were  parasitized  were  more  likely  to  experi- 
ence predation.  Chats  tended  to  place  nests  in 
larger  patches  with  more  small  stems  than 
those  in  unused  sites.  Nests  that  were  placed 
faither  from  the  patch  edge  were  more  sus- 


Means  (±  standard  error),  parameter  estimates,  Wald  statistics,  and  probability  levels  for  individual  variables  from  logistic  regressions  on  predation 


Burhciiis  and  Thompson  • YELLOW-BREASTED  CHAT  NESTING  PATCH 


213 


NO 

in 

fN 

X 

r-- 

CL. 

1 

O 

o 

(N 

o 

On 

o 

d 

d 

d 

d 

d 

d 

d 

ON 

r^ 

o 

(N 

(N 

O 

in 

<N 

X 

1 

in 

q 

O 

m 

on 

d 

rn 

d 

rA 

w 

m 

r-- 

r- 

00 

fO 

X 

£ 

1 

in 

a^ 

tn 

q 

1 

d 

d 

d 

d 

d 

d 

CQ 

Cl 

ON 

r- 

00 

On 

o 

r-- 

•o 

o 

o 

m 

(N 

o 

in 

- 

'(A 

CO 

1 

d 

+1 

d 

+1 

1 

1 

1 

1 

d 
+ 1 

d 

+1 

d 

+1 

d 

-t-l 

CL 

o 

r- 

00 

On 

o 

r- 

c 

r-' 

NO 

q 

X 

o 

q 

£ 

d 

d 

d 

d 

(N 

Vi 

e9 

o 

in 

00 

00 

Cl» 

m 

o 

r- 

r- 

q 

(n 

N 

'(A 

1 

d 

+1 

d 

+1 

1 

1 

1 

1 

d 

+1 

d 

+1 

-1-1 

-M 

CfJ 

o 

00 

CL 

(N 

r- 

fo 

q 

q 

q 

d 

(N 

(N 

'Ct 

r-- 

00 

On 

in 

NO 

(N 

o 

— 

in 

r- 

o 

NO 

1 

j 

j 

d 

d 

d 

d 

d 

d 

d 

d 

00 

On 

in 

Cb 

00 

On 

(N 

— 

in 

o 

1 

1 

1 

1 

fO 

d 

d 

d 

rn 

d 

d 

1) 

in 

m 

(N 

o 

in 

r- 

o 

V 

£ 

ca 

in 

00 

m 

in 

o 

o 

o 

1 

1 

1 

1 

— 

d 

d 

d 

d 

d 

d 

d 

1 

1 

1 

1 

?3 

1 

Q. 

in 

m 

00 

rj 

r- 

NO 

00 

o 

00 

(N 

in 

Vw^ 

00 

•a 

11% 

63% 

d 

d 

(N 

d 

u 

013 

-o 

V 

+1 

+1 

+1 

+ 1 

+1 

+1 

1 

1 

1 

1 

u. 

'i- 

00 

00 

1— • 

(N 

r- 

NO 

q 

— 

»— .* 

d 

NO 

X 

NO 

r- 

00 

c 

g 

(N 

T3 

u 

Cl 

r-- 

m 

NO 

•o 

in 

vO 

o 

ON 

in 

m 

cn 

d 

d 

d 

d 

-a 

D 

r- 

+ 1 

+1 

+ 1 

+ 1 

+ 1 

+1 

1 

1 

1 

1 

a 

1* 

rn 

00 

ON 

o 

ON 

m 

Q 

r- 

NO 

q 

q 

q 

q 

d 

d 

rn 

X 

X 

in 

m 

X 

X 

X 

X 

*o 

TD 

c/3 

c 

c 

<U 

s: 

LL 

u 

c 

a 

Variable 

c/3 

c/3 

V 

u 

V 

V 

B 

o 

S 

D, 

? 

o 

X 

o 

X 

X 

X 

X 

r* 

O 

A 

c 

Lln 

O 

03 

<L) 

00 

u 

03 

u 

c 

s 

-C 

X) 

*0 

X 

X 

*o 

X 

■o 

c 

c/3 

E 

T3 

E 

"O 

E 

X 

•o 

c/3 

B 

E 

C/3 

£ 

-C 

o 

C 

c/3 

'-B 

e 

E 

(j 

c 

u 

w 

C/3 

o 

o 

CJ 

O 

E 

o 

o 

c/3 

C/3 

p 

Q. 

c/3 

<u 

00 

OD 

o 

in 

1 

o 

1 

a 

in 

o 

in 

<U 

OX) 

03 

u 

c3 

D- 

c^ 

*C/3 

U 

CdD 

03 

L- 

C/3 

O 

c 

O 

-C 

'G 

s: 

V 

c/3 

f— 

c/3 

r“ 

inj 

c/3 

c 

E 

c/3 

c/3 

c 

(N 

c/3 

c 

A 

c/3 

C 

u 

— 

*0 

c 

03 

u 

03 

OJ 

> 

o 

w 

03 

to 

c 

o 

•w 

C 

C 

o 

o 

c 

c 

o 

c 

<D 

s 

03 

CL 

< 

0. 

Z 

00 

55 

H 

55 

55 

55 

H 

ceptible  to  parasitism  (Table  2);  however,  we 
were  unable  to  separate  the  importance  of 
patch-edge  distance  from  the  size  of  the  patch 
itself  (patch  diameter).  Although  nests  in  large 
patches  were  more  likely  to  become  parasit- 
ized, higher  nesting  success  in  large  patches 
compensated  for  decrements  in  fitness  caused 
cowbird  parasitism  because  the  number  of 
host  young  that  fledged  was  equal  between 
small  and  large  patches. 

Petersen  and  Best  (1985),  Knopf  and  Sedg- 
wick (1992),  and  Holway  (1991)  found  that 
birds  selected  large  shrubs  or  shrub  stands  for 
nest  placement.  Holway  (1991)  and  Knopf 
and  Sedgwick  (1992)  suggested  that  large 
patches  offer  improved  nest  concealment; 
Holway  (1991)  also  believed  that  large  patch- 
es could  impede  the  movements  of  mammals, 
and  could  contain  more  potential  nest  sites  for 
predators  to  search  (see  also  Martin  and  Roper 
1988). 

Several  researchers  have  found  that  birds 
place  nests  in  denser  cover  than  in  unused 
sites  (Knopf  and  Sedwick  1992,  Sedgwick  and 
Knopf  1992).  Holway  (1991)  and  Steele 
(1993)  found  higher  foliage  and  shrub  density 
at  nests  of  Black-throated  Blue  Warblers 
(Dendroica  caerulescens)  than  at  random 
points.  Wray  and  Whitmore  (1979)  and  Nor- 
ment  (1993)  found  that  successful  Vesper 
Sparrow  (Pooecetes  gramineus)  and  Harris 
Sparrow  (Zonotrichia  querula)  nests  tended  to 
be  placed  in  denser  cover  than  unsuccessful 
nests. 

Chat  nests  parasitized  by  Brown-headed 
Cowbirds  were  more  likely  to  become  dep- 
redated. Dearborn  (in  press)  found  higher  dai- 
ly mortality  at  parasitized  nests  of  Indigo  Bun- 
tings in  a five  year  study  from  the  same  sites 
we  used.  He  suggested  that  louder  vocaliza- 
tion by  cowbird  nestlings  was  partly  the 
cause,  although  daily  mortality  was  also  high- 
er at  parasitized  bunting  nests  during  the  in- 
cubation stage.  In  our  study,  the  sample  size 
of  parasitized  nests  with  cowbird  chicks  was 
too  small  (/?  = 4)  to  adequately  compare  daily 
mortality  between  nests  with  cowbird  chicks 
and  those  without  (Hensler  and  Nichols  1981). 

Parasitized  nests  had  more  potential  cow- 
bird perches  (laige  stems)  than  unparasitized 
nests.  Recent  studies  have  documented  the  im- 
portance of  perch  proximity  in  brood  parasit- 
ism in  both  cuckoos  (Cuculus  canorus\  Al- 


214 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


TABLE  3.  Means  (±  standard  error),  parameter  estimates,  Wald  X‘  statistics,  and  probability  levels  for 
individual  variables  from  univariate  logistic  regressions  comparing  nest  sites  and  unused  sites. 


Variable 

Nest  sites 

Unused  sites 

Parameter 

X- 

p 

Average  patch  diameter 

52% 

19% 

1.55 

10.91 

0.001 

(%  in  large  patches) 

Stems  <2  cm  dbh 

313.23 

± 108.47 

60.19  ± 

1 1.49 

0.00 

2.47 

0.12 

Stems  2-5  cm  dbh 

5.44 

± 1.14 

4.17  ± 

1.01 

0.02 

0.68 

0.41 

Stems  5-10  cm  dbh 

1.48 

± 0.29 

1.10  ± 

0.24 

0.1  1 

0.99 

0.32 

Stems  1 1-20  cm  dbh 

1.58 

± 0.31 

2.15  ± 

0.42 

-0.09 

1.12 

0.29 

Stems  21-50  cm  dbh 

1.06 

± 0.29 

1.21  ± 

0.32 

-0.03 

0.12 

0.73 

Stems  >50  cm  dbh 

0.04 

± 0.03 

0.02  ± 

0.02 

0.71 

0.33 

0.57 

Total  small  stems  (<10  cm  dbh) 

320.14 

± 108.19 

65.45  ± 

1 1.51 

0.00 

2.77 

0.10 

Total  large  stems  (>10  cm  dbh) 

2.69 

± 0.54 

3.38  ± 

0.65 

-0.04 

0.66 

0.42 

varez  1993,  0ien  et  al.  1996)  and  cowbirds 
(Romig  and  Crawford  1995,  Clotfelter  1998; 
see  also  Anderson  and  Storer  1979).  Previous 
studies  (Burhans  1997)  on  Field  Sparrows 
{Spizella  pusilla)  and  Indigo  Buntings  nesting 
at  Thomas  Baskett  Wildlife  Research  and  Ed- 
ucation Center  indicated  no  direct  relationship 
between  perches  near  the  nest  and  frequency 
of  parasitism.  However,  Yellow-breasted  Chat 
nest  sites  generally  are  situated  in  patches 
with  more  trees  and  shrubs  than  old  field  nests 
of  Indigo  Buntings  and  Field  Sparrows.  The 
higher  frequency  of  parasitism  in  large  patch- 
es was  not  an  artifact  of  patch  size,  because 
large  patches  did  not  necessarily  contain  more 
total  large  stems. 

Nest  site  selection,  nesting  success,  and  fre- 
quency of  cowbird  parasitism  at  Yellow- 
breasted Chat  nests  appear  to  be  influenced  by 
patch  size.  However,  relaxed  predation  in 
large  patches  did  not  improve  host  fledging 
success,  because  chats  were  more  likely  to  be- 
come parasitized  in  large  patches  and  fledge 
fewer  of  their  own  young.  Although  predation 
and  parasitism  appeared  to  differ  across  patch 
sizes,  the  effects  of  patch  size  on  host  fitness 
appear  to  cancel  each  other  out.  Future  studies 
should  look  further  at  interactions  between 
site  selection,  brood  parasitism,  and  predation, 
and  investigate  tradeoffs  in  reproductive  suc- 
cess associated  with  these  factors  according  to 
different  types  of  nest  sites. 

ACKNOWLEDGMENTS 

Sugge.stions  from  D.  Dearborn,  and  an  anonymous 
reviewer  greatly  improved  the  manuscript.  We  are 
grateful  to  C.  Lrceman,  D.  Dearborn,  D.  Martasian,  M. 
Alexander,  and  R.  Kun/.a  for  help  in  the  field  and  J. 


Demand  for  help  with  data  entry.  We  thank  J.  Laaborg 

for  his  assistance.  C.  Lreiling  kindly  allowed  us  to 

work  on  his  property.  This  study  was  funded  by  the 

USDA  Eorest  Service  North  Central  Research  Station. 

LITERATURE  CITED 

Alvarez,  F.  1993.  Proximity  of  trees  facilitates  para- 
sitism by  Cuckoos  Cuciilus  canoni.s  on  Rufous 
Warblers  Cercotrichas  galacfofes.  Ibis  135:331. 

Anderson,  W.  L.  and  R.  W.  Storer.  1976.  Factors 
influencing  Kirtland’s  Warbler  nesting  suceess. 
Jaek-Pine  Warbler  54:105-1 15. 

Barber,  D.  R.  and  T.  E.  Martin.  1997.  Influenee  of 
alternate  host  densities  on  Brown-headed  Cowbird 
parasitism  rates  in  Black-eapped  Vireos.  Condor 
99:595-604. 

Bowman,  G.  B.  and  L.  D.  Harris.  1980.  Effect  of 
spatial  heterogeneity  on  ground-nest  depredation. 
J.  Wildl.  Manage.  44:806-813. 

Brittingham,  M.  C.  and  S.  A.  Temple.  1996.  Vege- 
tation around  parasitized  and  non-parasitized  nests 
within  deeiduous  forests.  J.  Field  Ornith.  67:406- 
413. 

Burhans,  D.  E.  1997.  Habitat  and  microhabitat  fea- 
tures associated  with  cowbird  parasitism  in  two 
forest  edge  cowbird  hosts.  Condor  99:866-872. 

Clotfelter,  E.  D.  1998.  What  cues  do  Brown-headed 
Cowbirds  use  to  locate  Red-winged  Blackbird 
host  nests?  Anim.  Behav.  55:1181-1189. 

Conner,  R.  N.,  M.  E.  Anderson,  and  J.  G.  DtCKSON. 
1986.  Relationships  among  territory  size,  song, 
and  nesting  success  of  Northern  Cardinals.  Auk 
103:23-31. 

Dearborn,  D.  C.  In  Press.  Brown-headed  Cowbird 
nestling  vocalizations  and  the  risk  of  nest  preda- 
tion. Auk. 

Hahn,  D.  C.  and  J.  S.  Hatfield.  1995.  Parasitism  at 
the  landscape  scale:  cowbirds  prefer  forests.  Con- 
■serv.  Biol.  9:1415-1424. 

Hensler,  G.  L.  and  j.  D.  Nichols.  1981.  The  Mayfield 
method  of  estimating  nesting  success:  a model, 
estimators  and  simulation  results.  Wilson  Bull.  93: 
42-53. 


limiuins  anil  Thompson  • YELLOW-BREASTED  CHAT  NESTINCj  PATCH 


215 


Holway,  D.  a.  1991.  Nest-site  seleetion  and  the  im- 
portance of  nest  concealment  in  the  Black-throat- 
ed Blue  Warbler.  Condor  93:575-581. 

Hosmer,  D.  W.,  Jr.  and  S.  Lemeshow.  1989.  Applied 
logistic  regression.  John  Wiley  & Sons,  New 
York. 

J.AMES,  E C.  AND  C.  E.  McCulloch.  1990.  Multivariate 
analysis  in  ecology  and  systematics:  panacea  or 
pandora’s  box?  Annu.  Rev.  Ecol.  Syst.  21:129- 
166. 

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

Kelly,  J.  P.  1993.  The  effect  of  nest  predation  on  hab- 
itat selection  by  Dusky  Flycatchers  in  limber  pine- 
juniper  woodland.  Condor  95:83-93. 

Kligo,  J.  C.,  R.  a.  Sargent,  B.  R.  Chapman,  and  K. 
V.  Miller.  1996a.  Nest-site  selection  by  Hooded 
Warblers  in  bottomland  hardwoods  of  South  Car- 
olina. Wilson  Bull.  108:53-60. 

Kligo,  J.  C.,  R.  A.  Sargent,  K.  V.  Miller,  and  B.  R. 
Chapman.  1996b.  Nest  sites  of  Kentucky  Warblers 
in  bottomland  hardwoods  of  South  Carolina.  J. 
Field  Ornith.  67:300-306. 

Knopf,  E L.  and  J.  A.  Sedgwick.  1992.  An  experi- 
mental study  of  nest-site  selection  by  Yellow  War- 
blers. Condor  94:734-742. 

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. 

Mayfield,  H.  1961.  Nesting  success  calculated  from 
exposure.  Wilson  Bull.  73:255-261. 

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

Nolan,  V.,  Jr.  1963.  Reproductive  success  of  birds  in 
a deciduous  scrub  habitat.  Ecology  44:305-313. 

Norment,  C.  j.  1993.  Nest-site  characteristics  and  nest 
predation  in  Harris’  Sparrows  and  White-Crowned 
Sparrows  in  the  northwest  territories,  Canada.  Auk 
1 10:769-777. 


0IEN,  1.  J.,  M.  HoNZA,  a.  MoKSNES,  and  E.  R0.SKAET. 
1996.  The  risk  of  parasitism  in  relation  to  the  dis- 
tance from  Reed  Warbler  nests  to  cuckoo  perches. 
J.  Anim.  Ecol.  65:147-153. 

Petersen,  K.  L.  and  L.  B.  Best.  1985.  Nest-site  se- 
lection by  Sage  Sparrows.  Condor  87:217-221. 

Petit,  K.  E.,  D.  R.  Petit,  and  L.  J.  Petit.  1988.  On 
measuring  vegetation  characteristics  in  bird  terri- 
tories: nest  sites  vs.  perch  sites  and  the  effect  of 
plot  size.  Am.  Midland  Nat.  1 19:209-215. 

Press,  S.  J.  and  S.  Wilson.  1978.  Choosing  between 
logistic  regression  and  discriminant  analysis.  J. 
Am.  Stat.  Assoc.  73:699-705. 

Romig,  G.  P.  and  R.  D.  Crawford.  1995.  Clay-colored 
Sparrows  in  North  Dakota  parasitized  by  Brown- 
headed Cowbirds.  Prairie  Nat.  27:193-203. 

Sauer,  J.  R.  and  B.  K.  Williams.  1989.  Generalized 
procedures  for  testing  hypotheses  about  survival 
or  recovery  rates.  J.  Wildl.  Manage.  53:137-142. 

Sedgwick,  J.  A.  and  F.  L.  Knopf.  1992.  Describing 
Willow  Flycatcher  habitats:  scale  perspectives  and 
gender  differences.  Condor  94:720—733. 

Steele,  B.  B.  1993.  Selection  of  foraging  and  nesting 
sites  by  Black-thoated  Blue  Warblers:  their  rela- 
tive influence  on  habitat  choice.  Condor  95:568- 
579. 

Tarvin,  K.  a.  and  K.  G.  Smith.  1995.  Microhabitat 
factors  influencing  predation  and  success  of  sub- 
urban Blue  Jay  Cyanocitta  cristata  nests.  J.  Avian. 
Biol.  26:296-304. 

Thompson,  C.  F.  and  V.  Nolan,  Jr.  1973.  Population 
biology  of  the  Yellow-breasted  Chat  (Icteria  vi- 
rens  L.)  in  southern  Indiana.  Ecol.  Monog.  43: 
145-171. 

With,  K.  A.  1994.  The  hazards  of  nesting  near  shrubs 
for  a grassland  bird,  the  McCown’s  Longspun 
Condor  96:1009-1019. 

Wray,  T,  II  and  R.  C.  Whitmore.  1979.  Effects  of 
vegetation  on  nesting  success  of  Vesper  SpaiTows. 
Auk  96:802-805. 


Wilson  Bull.,  1 1 1(2),  1999,  pp.  216-228 


AVIFAUNA  OF  A PARAGUAYAN  CERRADO  LOCALITY:  PARQUE 
NACIONAL  SERRANIA  SAN  LUIS,  DEPTO.  CONCEPCION 

MARK  B.  ROBBINS,'  3 ROB.  C.  FAUCETT,^  AND  NATHAN  H.  RICE'^ 

ABSTRACT. — We  recorded  181  avian  species  at  the  Paraguayan  Cerrado  site,  Parque  Nacional  Serram'a  San 
Luis,  depto.  Concepcion,  including  the  first  record  of  Veery  (Catharus  fuscescens)  for  the  country.  We  obtained 
further  evidence  of  hybridization  between  White-bellied  (Basileiiterus  hypoleucus)  and  Golden-crowned  [B. 
ciilicivorus)  warblers.  Our  results  combined  with  those  of  earlier  workers  document  a total  of  219  species  for 
this  area.  Pronounced  differences  in  species  composition  exist  between  San  Luis  and  a nearby  Cerrado  locality. 
Three  threatened  and  four  near-threatened  species  were  recorded  at  San  Luis.  Since  our  inventory  the  area  that 
we  worked  was  traded  for  an  adjacent,  less  human  impacted  sector  of  similar  size.  The  new  park  boundary 
contains  relatively  pristine  campo  (grassland),  which  is  an  important  habitat  for  a number  of  resident  and  migrant 
species  suspected  to  be  suffering  serious  population  declines.  Received  29  Jan.  1998,  accepted  14  Nov.  1998. 

RESUMEN. — Se  registraron  181  especies  de  aves  en  el  Parque  Nacional  Serrania  San  Luis,  en  el  cerrado  del 
Paraguay,  depto.  Concepcion.  La  lista  incluye  el  primer  registro  de  Catharus  fuscescens  para  el  pafs,  y mas 
evidencia  de  hibridizacidn  entre  los  parulidos  Basileiiterus  hypoleucus  y B.  culiciviorus.  En  combinacion  con 
los  resultados  de  investigadores  anteriores,  han  sido  registrado  219  especies  de  aves  para  la  zona.  Existen 
diferencias  marcadas  en  la  composicion  avifaum'stica  entre  San  Luis  y otra  localidad  cercana  en  cerrado.  Se 
registraron  tres  especies  amenazadas  y cuatro  casi-amenazadas  en  la  zona.  Desde  que  se  hizo  el  presente  estudio, 
se  cambio.  el  area  de  estudio  por  otro  adyacente,  menos  impactado  y de  extenso  similar;  esta  zona  abarca  mas 
del  campo  (pastizal)  en  buen  estado  de  conservacion,  el  cual  es  un  refugio  para  varias  especies  residentes  y 
migratorias  que  se  cuentran  en  declives  poblacionales. 

The  Cerrado  is  the  second  largest  ecologi- 
cal region  in  South  America  (Ah’ Saber  1977), 
but  has  only  recently  attracted  the  attention  of 
avian  biogeographers  (Haffer  1985;  Silva 
1995a,  b,  c).  Silva  (1995a)  demonstrated  that 
about  70%  of  the  Cerrado  is  inadequately 
sampled  and  that  the  southern  component  es- 
pecially is  poorly  known  and  most  heavily  im- 
pacted by  human  activities.  The  Cerrado 
reaches  its  southern  terminus  in  northeastern 
Paraguay,  where  it  interdigitates  with  two  oth- 
er physiogeographic  and  biotic  regions:  the 
Chaco  and  the  Atlantic  Forest  (Hayes  1995a). 

The  lack  of  a quantitative  inventory  of  this 
region,  coupled  with  the  potential  for  exten- 
sive biotic  interchange  among  the  contiguous 
regions,  make  the  Paraguayan  Cerrado  impor- 
tant for  its  unique  contribution  to  the  biodi- 
versity of  the  Cerrado.  In  this  paper,  we  pre- 
sent a preliminary  avifaunal  inventory  for  the 


' Division  of  Ornithology,  Natural  History  Museum, 
Univ.  of  Kansas,  Lawrence,  KS  66045. 

- Museum  of  Natural  Science,  Foster  Hall,  Louisiana 
State  Univ.,  Baton  Rouge,  LA  70803. 

’ Corresponding  author; 

E-mail:  mrobbins@falcon.cc.ukans.edu 

••  Present  address:  Academy  of  Natural  Sciences, 
Dept,  of  Ornithology,  1900  Benjamin  Franklin  Park- 
way, Philadelphia,  PA  19103. 


10,273  ha  Parque  Nacional  Serrama  San  Luis, 
created  in  1991  to  preserve  a representative 
sample  of  the  Paraguayan  Cerrado  (Direccion 
de  Parques  Nacionales  y Vida  Silvestre  1993). 
We  have  incorporated  results  from  earlier 
work  in  the  region  to  provide  a more  compre- 
hensive list  for  this  area.  The  San  Luis  list  is 
compared  to  another  well-surveyed  Paiaguay- 
an  Cerrado  locality  to  examine  avian  species 
turnover.  Selected  species  accounts  are  pro- 
vided where  additional  comment  is  merited  to 
clarify  status,  distribution,  migration,  food 
habits,  hybridization,  and  plumage  characters. 

STUDY  AREA  AND  METHODS 

We  worked  at  the  southern  end  of  Parque  Nacional 
Serram'a  San  Luis  (22°  40'  S,  57°  21'  W;  taken  with 
Global  Positioning  System  at  park  headquarters;  Fig. 
1)  from  19-31  October  1996,  and  RCF  revisited  the 
site  from  29  November  to  7 December  1996.  The  park 
is  located  in  the  Serrania  San  Luis,  an  isolated,  low- 
lying  (max.  elevation  500  m)  set  of  limestone  hills 
mostly  covered  with  semihumid  forest.  Approximately 
5,300  ha  is  cultivated  grassland  with  about  70%  of  this 
habitat  now  occupied  by  the  aggressive  African  grass 
Hyparrhenia  rufa  (Poaceae).  The  dominant  savannah 
tree,  Tahehuia  aurea  (Bignoniaceae),  was  often  asso- 
ciated with  Astronium  urundeuva  (Anacardiaceae),  A/i- 
adenanthera  coluhrina  (Leguminosae),  and  Rhainni- 
dium  elasocarpum  (Rhamnaceae;  Consorcio  Parelc- 
Foragro-Porto  Real,  unpubl.  report).  Hillsides  were 


216 


Rohhins  el  al.  • PARAGUAYAN  CERRADO  AVIFAUNA 


217 


FIG.  1 . Locations  of  Parque  Nacional  Serrania  San 
Luis  and  Parque  Nacional  Cerro  Cora. 


covered  with  shallow,  rocky  soil,  resulting  in  shorter, 
xeric  forest;  the  more  level  areas  along  two  arroyos 
bordering  the  western  and  eastern  sides  of  the  park 
held  taller,  more  humid  forest.  Terrestrial  bromeliads 
and  cacti  were  common  in  the  dense,  thorn-covered 
understory  on  slopes  and  on  isolated  woodland  in  the 
campo  (open  grassland).  The  largest  area  of  campo, 
about  4X2  km,  had  a few  woodlots,  some  only  a few 
meters  square.  The  upland  forest  and  isolated  wood- 
lands had  uneven  canopies  and  were  relatively  short, 
with  few  trees  exceeding  20  m.  Prominent  trees  in- 
cluded; Amburana  cearensis  (Leguminosae),  Aspido- 
sperma  pyrifolium  (Apocynaceae),  Calycophyllum 
midtiflorum  (Rubiaceae),  and  Astroniiim  urundeuva 
(Anacardiaceae).  The  understory  was  dominated  by 
trees  of  tbe  genus  Trichilia  (Meliaceae). 

Forests  along  the  arroyos,  especially  east  of  Tagatlya 
Guazu  arroyo  at  the  headquarters,  were  taller  and  more 
humid,  with  some  trees  exceeding  40  m.  These  forests 
were  heavily  disturbed  by  selective  logging,  and  tree- 
fall  gaps  with  dense  understory  were  common.  Vine 
tangles  were  also  more  common  than  in  the  upland 
forest,  especially  about  2 km  east  of  the  park  head- 
quarters. Tall  arborescent  bamboo  (Gucidua  spp.)  was 
primarily  restricted  to  disturbed  areas  along  the  ar- 
royos. 

This  area  is  very  seasonal  with  a cool,  dry  climate 
from  April  through  most  of  September  when  temper- 
atures occasionally  drop  to  near  0°  C.  Temperature  and 
rainfall  increase  in  October  and  November  with  De- 
cember through  March  being  relatively  hot  and  humid. 
Average  annual  rainfall  is  1300  mm  with  considerable 
variation  (A.  Acosta,  pers.  comm.),  and  average  annual 
temperature  is  24°  C (Consorcio  Parelc-Foragro-Porto 
Real,  unpubl.  report).  During  our  stay,  weather  was 
highly  variable.  On  the  evening  of  19  October  a major 
storm  system  from  the  southeast  brought  heavy  rains 
(70  mm)  and  strong  winds.  Precipitation  also  occurred 
on  25  October,  12  mm;  26  October,  2 mm;  and  27 
October,  6 mm.  Apparently  it  did  not  rain  between  our 


October  and  December  surveys  (A.  Acosta,  pers. 
comm.);  however,  it  rained  five  of  nine  days  during 
the  latter  inventory  period.  Temperature  highs  ranged 
from  24-38°  C.  The  low  was  I8°C  at  dawn  on  23 
October.  We  did  not  record  the  temperature  during  No- 
vember and  December. 

Four  mist-nets  were  opened  on  the  afternoon  of  19 
October,  and  6 on  20  October;  15  nets  were  maintained 
from  21-30  October,  for  a total  of  about  1 1,400  mist- 
net-hr.  All  mist-nets  were  positioned  in  mesic  forest 
bordering  Tagatlya  Guazu  arroyo;  the  most  distant  nets 
were  about  1.5  km  east  of  the  headquarters.  Most  nets 
were  moved  every  three  to  four  days.  Nets  were  typ- 
ically opened  at  dawn  and  closed  about  noon,  rarely 
at  sunset.  Our  inventory  was  confined  to  about  4 km 
of  forest  trail  east  of  the  headquarters,  about  5 km  of 
dirt  road  to  the  north  of  the  headquarters,  and  about  2 
km  of  dirt  road  from  the  headquarters  to  the  southwest 
entrance.  On  most  mornings,  RCF  and  NHR  worked 
the  nets  and  adjacent  forest,  while  MBR  made  inde- 
pendent surveys  of  other  areas  from  predawn  until 
11:00  or  12:00.  Occasional  forays  were  made  in  the 
afternoon.  Nocturnal  fieldwork  was  limited  to  nights 
with  a full  moon  in  October. 

Specimens  were  deposited  at  the  University  of  Kan- 
sas Natural  History  Museum  (KU),  Lawrence,  Kansas 
and  Museo  Nacional  de  Historia  Natural  del  Paraguay 
(MNHNP),  Asuncion,  Paraguay.  Tissue  samples  were 
taken  from  every  nonfluid  preserved  specimen  and  de- 
posited at  KU.  Tape  recordings  will  be  deposited  at  the 
Library  of  Natural  Sounds,  Cornell  University,  Ithaca, 
New  York. 

To  provide  a more  complete  inventory  of  the  Ser- 
rania de  San  Luis  region,  we  have  integrated  our  spe- 
cies list  (Appendix)  with  results  from  earlier  work  in 
San  Luis  de  La  Sierra  that  was  conducted  in  Septem- 
ber-October  1931  by  Hans  Krieg  (referred  to  as  “Apa- 
Bergland”;  summarized  in  Laubmann  1939,  1940). 
Results  from  Kreig’s  expeditions  to  Zanja  Morotf, 
Centurion,  and  Estrella,  Concepci'on  were  not  included 
because  these  areas  are  farther  removed  from  San  Luis, 
and  the  Estrella  site  was  at  the  Ri'o  Apa.  We  presume 
that  the  riparian  habitat  and  associated  fauna  along  the 
Rio  Apa  is  somewhat  distinct  from  that  at  San  Luis. 
We  also  excluded  specimens  taken  by  Emil  Kaempfer 
at  La  Fonciere,  Concepcion  (deposited  at  American 
Museum  of  Natural  History;  Floyd  Hayes  and  Paul 
Sweet,  pers.  comm.).  Except  where  noted  otherwise, 
we  follow  Hayes  (1995a)  for  taxonomy  and  nomen- 
clature. 

RESULTS 

We  recorded  a total  of  181  species  (Appen- 
dix), of  which  three  are  designated  as  threat- 
ened and  four  as  near- threatened  (Collar  et  al. 
1992,  1994).  Three  of  the  29  species  identified 
by  Silva  (1995c)  as  endemic  to  the  Ceixado, 
Planalto  Foliage-gleaner  (Philydor  dimidia- 
tus).  Curl-crested  Jay  {Cyanocorax  cristatel- 
lus),  and  Black-throated  Saltator  {Saltator 


218 


THE  WILSON  BULLETIN  • Voi  111,  No.  2,  June  1999 


atricolUs),  have  been  recorded  at  San  Luis. 
We  recorded  four  species  that  Silva  (1995c) 
did  not  list  for  the  Cerrado:  South  American 
Painted-Snipe  (Rostratula  semicollaris), 
White-naped  Xenopsaris  {Xenopsaris  albimi- 
cha).  Red-crested  Cardinal  (Paroaria  coron- 
ata),  and  Golden-winged  Cacique  {Caciciis 
chrysopterus).  Hayes  (1995a)  did  not  list  R. 
semicollaris,  X.  albinucha,  Golden-green 
Woodpecker  {Piculiis  chrysochloros),  nor 
Dark-throated  Seedeater  (Sporophila  ruficol- 
lis)  as  occurring  in  his  “Campos  Cerrados” 
category. 

At  least  64  of  the  92  species  we  collected 
had  enlarged  gonads  indicating  reproduction, 
and  nests  or  recently  fledged  young  were  ob- 
served for  12  additional  species  that  were  not 
collected  (Appendix).  Based  on  vocal  activity 
and  behavior,  breeding  was  suspected  for  sev- 
eral species  not  collected,  such  as  Bare-faced 
Currasow  {Crax  fasciolata).  Red-legged  Ser- 
iema  (Cariama  cristata).  Short-tailed  Night- 
hawk  {Lurocalis  semitorquatus),  and  Crowned 
Slaty  Flycatcher  {Griseotyrannus  aurantioa- 
trocristatus).  Hence,  at  least  80  species  were 
actively  nesting  at  the  time  of  our  studies 
(mid-October-early  December).  One  Neaictic, 
one  intratropical,  and  at  least  32  Austral  mi- 
grants were  recorded  (Appendix;  migrant  ter- 
minology follows  Hayes  1995b). 

Nocturnal  avian  vocal  activity  was  most 
pronounced  shortly  after  dusk  and  in  the  hour 
before  dawn,  with  the  greatest  activity  on 
nights  with  the  brightest  moon  light.  The  Fer- 
ruginous Pymy-Owl  (Glaucidium  brasilian- 
um)  and  both  screech-owls  (Otus)  were  heard 
nightly,  as  were  Pauraque  (Nyctidromus  albi- 
collis).  Rufous  Nightjar  (Caprimidgus  rufus), 
and  Common  Potoo  (Nyctibius  griseus).  The 
Little  Nightjar  (C.  parvulus)  was  never  heard, 
even  though  individuals  and  a nest  with  an 
egg  and  a day-old  young,  were  collected  at 
the  edge  of  camp. 

Rodent  eating  hawks  and  owls  were  not 
well  represented  in  the  San  Luis  avifauna.  A 
possible  causal  explanation  is  that  rodent  di- 
versity and  density  were  low.  During  October, 
we  observed  no  evidence  of  rodents.  Squirrels 
are  unknown  from  this  part  of  Paraguay;  dur- 
ing seven  nights  in  late  November— early  De- 
cember Texas  Tech  University  mammalogists 
captured  few  mammals  in  a variety  of  habitats 
(M.  Gorresen,  pers.  comm.).  Raptor  species 


expected,  but  not  detected  include:  Great 
Black  Hawk  {Buteogallus  urubitinga),  Har- 
ris’s Hawk  {Parabuteo  unicinctus),  all  Buteo 
species.  Spectacled  Owl  (Pulsatrix  perspicil- 
lata).  Great  Homed  Owl  {Bubo  virginianus), 
and  Asio  spp.  Only  the  Savannah  Hawk  {Bu- 
teogallus meridionalis)  and  Ornate  Hawk-Ea- 
gle {Spizaetus  ornatus)  were  recorded  (both 
rarely)  among  the  species  of  raptor  whose  di- 
ets include  a relatively  high  percent  (>25%) 
of  rodents  (Brown  and  Amadon  1968,  del 
Hoyo  et  al.  1994). 

SPECIES  ACCOUNTS 

Snail  Kite  (Rostrhamus  sociabilis). — On  22 
and  23  October,  we  observed  two  flocks  of 
154  and  33  individuals  (mostly  adults),  re- 
spectively, pass  through  the  campo.  The  birds 
appeared  from  the  north  and  landed  on  the 
ground,  in  bushes,  and  in  isolated  palm  trees, 
flying  only  a few  meters  between  each  perch. 
When  the  groups  came  to  the  forested  Tagat- 
lya  Guazu  aiToyo  at  the  south  end  of  the  park, 
they  rose  as  a group  into  the  air  and  disap- 
peared to  the  south.  Snail  Kites  were  not  ob- 
served during  late  November-early  December. 
Our  observations  coincide  with  prior  obser- 
vations of  migratory  movements  of  this  spe- 
cies in  Pai'aguay  (Hayes  et  al.  1994). 

Crowned  Solitary  Eagle  { Harpy haliaetus 
coronatus). — An  adult  with  a tegu  lizard  (Tei- 
idae;  Tupinambis  sp.)  in  its  talons  was  flushed 
from  a large  tree  at  the  forest/campo  edge  on 
19  October.  It  landed  in  the  top  of  a neai’by 
tree  where  it  sat  for  about  3 min  before  it  dis- 
appeared with  the  lizard  in  its  talons.  This  spe- 
cies is  considered  vulnerable  (“taxa  believed 
likely  to  move  into  the  endangered  category 
in  the  near  future  if  the  causal  factors  continue 
operating”;  Collar  et  al.  1992)  and  has  been 
recorded  few  times  from  Paraguay  with  only 
one  prior  record  for  depto.  Concepcion  (Hayes 
1995a).  Very  little  is  known  about  this  eagle’s 
prey  (Collar  et  al.  1992),  and  our  observation 
is  apparently  the  first  of  it  taking  a. tegu.  It  is 
known  to  take  small  mammals  (Collar  et  al. 
1992,  del  Hoyo  et  al.  1994),  but  we  saw  ho 
signs  of  rodents.  Del  Hoyo  and  coworkers 
(1994)  noted  that  this  eagle  has  short  toes 
characteristic  of  snake  eaters. 

Bare-faced  Curassow  (Crax  fasciolata). — 
Given  that  cracids  are  usually  one  of  the  first 
avian  species  to  disappear  because  of  exces- 


Robbins  el  at.  • PARAGUAYAN  CERRADO  AVIFAUNA 


219 


sive  hunting,  we  were  encouraged  to  find  at 
least  seven  temtorial  males  calling  along 
about  4 km  of  trail  on  the  east  side  of  Tagatlya 
Guazu  aiToyo.  Males  called  daily  in  October 
from  predawn  until  about  07:00,  occasionally 
until  10:00.  We  did  not  hear’  males  in  the  drier 
forest;  however,  we  surveyed  only  a relatively 
small  area  of  this  forest  type  at  the  optimal 
time  of  day.  No  calling  was  heard  in  late  No- 
vember—early  December. 

Reddish-bellied  Parrot  (Pyrrhura  fronta- 
lis).— This  species  was  common  in  both  forest 
types.  Birds  collected  (KU  88346,  MNHNP) 
at  San  Luis  were  typical  of  the  race  P.  f chi- 
ripepe,  and  showed  no  sign  of  hybridization 
with  P.  devillei  as  has  been  reported  in  areas 
just  northwest  of  San  Luis  (Short  1975). 

Rufous  Nightjar  (Caprimulgus  rufus). — The 
single  male  (testes  11X6  mm;  KU  88350) 
obtained  was  not  assignable  to  either  the  nom- 
inate race  or  to  southern  C.  r.  rutilus.  Like 
birds  from  central  Brazil  (Goias,  Bahia;  Rob- 
bins and  Parker  1997),  the  San  Luis  speci- 
men’s wing  and  tail  measurements  (173.6  and 
119.7  mm,  respectively)  fall  within  the  range 
of  variation  of  the  nominate  race,  but  plumage 
color  and  pattern  are  closest  to  C.  r.  rutilus. 

Blue-crowned  Motmot  (Momotus  momo- 
ta). — Our  specimens  (KU  88580,  MNHNP) 
from  San  Luis  represent  only  the  fifth  locality 
in  Paraguay  (Hayes  1995a,  Ericson  and  Amar- 
illa  1997),  and  the  first  for  depto.  Concepcion. 
Both  specimens,  a probable  pair,  lack  the  blu- 
ish-green throat  of  eastern  Bolivian  and  Ar- 
gentinian M.  m.  pilcomajensis;  the  greenish 
underparts  were  washed  with  cinnamon,  es- 
pecially on  the  abdomen,  which  is  more  typ- 
ical of  southern  Brazilian  M.  m.  simplex. 
Chapman  (1923)  was  uncertain  in  assigning  a 
specimen  from  western  Sao  Paulo  to  either  of 
the  above  races  and  concluded  that  there  was 
considerable  variability  in  this  species  south 
of  the  Amazon.  Ericson  and  Amarilla  (1997) 
believed  that  specimens  they  obtained  from 
Parque  Nacional  Defensores  del  Chaco,  depto. 
Chaco  were  intergrades  between  the  above 
forms. 

Planalto  Foliage-gleaner  (Philydor  dimi- 
diatus). — Known  in  most  of  the  literature  as 
the  Russet-mantled  Foliage-gleaner,  we  follow 
Ridgely’s  and  Tudor’s  (1994)  appropriate  sug- 
gestion for  an  English  name.  Two  pairs  of  this 
poorly  known  foliage-gleaner  were  encoun- 


tered about  2 km  east  of  the  headquarters  (KU 
88362,  88363,  MNHNP).  Both  pairs  were  ob- 
served foraging  2—4  m above  the  ground  in 
relatively  dense  understory.  One  pair  was  as- 
sociated with  a understory  mixed-species 
flock.  Ridgely’s  and  Tudor’s  (1994)  descrip- 
tion of  the  song  and  call  accurately  describe 
what  we  recorded,  with  the  San  Luis  birds 
sounding  very  similar  to  birds  from  Patios  de 
Minas,  Minas  Gerais  (tape  recording  by  An- 
drew Whittaker).  Plumage  differentiation  and 
taxonomy  of  this  species  will  be  presented 
elsewhere. 

Bare-throated  Bellbird  (Procnias  nudicol- 
lis). — San  Luis  lies  at  the  western  limit  of  this 
species’  range  (Snow  1982,  Hayes  1995a).  In 
October,  two  birds  were  recorded.  An  adult 
male  (KU  88873)  sang  for  more  than  15  min 
during  mid-moming  on  24  October  from  a 
leafless,  uppermost  branch  of  one  of  the  tallest 
trees  (ca  40  m)  in  the  mesic  forest  east  of  the 
headquarters.  Only  one  or  two  other  individ- 
uals were  heard  in  October,  both  gave  the 
bock  call  a few  times,  on  two  separate  occa- 
sions. In  November  and  December,  RCF  heard 
at  least  12  displaying  males  daily  along  about 
4 km  of  trail  east  of  Tagatlya  Guazu  an'oyo. 
An  immature  male  (KU  88387:  testes  8X4 
mm)  was  one  year  old  based  on  plumage 
(Snow  1973)  and  bursa  size  (8X8  mm).  The 
plumage  of  this  specimen  resembles  closely 
that  of  an  adult  female  (Snow  1982,  Sick 
1993),  except  that  the  head  is  entirely  black 
and  the  abdomen  and  crissum  are  primarily 
white  instead  of  pale  yellow.  It  was  in  mod- 
erate to  heavy  body  molt,  with  #5  primai-y  in 
molt  (primaries  1-4  fresh).  Secondaiy  7 or  8 
(one  is  missing  on  each  side)  was  in  molt. 
Secondary  9 was  very  worn,  and  the  others 
appeared  lightly  worn.  The  central  pair  of  rec- 
trices  was  in  molt  with  all  other  rectrices 
heavily  worn. 

This  bellbird  has  been  documented  as  mi- 
gratory in  the  southeastern  part  of  its  Brazilian 
range  (Belton  1985,  Sick  1993);  in  Misiones, 
Argentina  it  is  apparently  only  transient  (M. 
Barnett  in  Lowen  et  al.  1996b).  In  Paraguay 
it  is  thought  to  occur  year  round  in  depto.  Ca- 
nendiyii  and  probably  depto.  Amambay  (Ma- 
drono and  Esquivel  1995).  Its  status  elsewhere 
in  Paraguay  is  uncleai'  (Lowen  et  al.  1996b). 
Our  limited  fieldwork  at  San  Luis  suggests  it 
may  occur  there  only  seasonally.  If  this  bell- 


220 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


bird  covers  great  distances  in  its  seasonal 
movements,  then  it  may  be  especially  vulner- 
able to  habitat  fragmentation.  Willis  (1979)  re- 
garded large  canopy  frugivores  to  be  vulner- 
able because  of  their  reliance  on  spatially  and 
temporally  patchy  resources.  This  bellbird  and 
the  Red-ruffed  Fruit  Crow  {Pyroderus  scuta- 
tus)  completely  disappeared  from  forests  that 
had  been  fragmented  (largest  fragment  198 
ha)  in  the  Lagoa  Santa  area  of  Brazil  (Chris- 
tiansen and  Fitter  1997). 

Veery  (Catharus  fuscescens). — We  netted 
an  immature  male  (KU  88484)  on  30  October. 
This  record  is  the  first  for  Paraguay,  and  rep- 
resents one  of  the  southernmost  records  for 
the  continent  (McFarlane  1974;  Remsen  and 
Traylor  1983;  Ridgely  and  Tudor  1989;  Willis 
and  Oniki  1993;  D.  Stotz,  pers.  comm.).  The 
specimen  is  referable  to  the  nominate  subspe- 
cies. All  three  Bolivian  specimens  originally 
identified  as  C.  f.  salicicola  by  Remsen  and 
Traylor  (1983)  are  now  thought  to  be  of  either 
the  nominate  race  or  C.  f.  fuliginosa  (V.  Rem- 
sen, D.  Stotz,  pers.  comm.).  No  subspecific 
determinations  have  been  made  for  southeast- 
ern Brazilian  specimens. 

White-bellied  Warbler  (Basileuterus  hypo- 
leucus). — The  presence  of  yellow  in  the  lower 
underparts  of  a few  specimens  from  several 
areas  and  observation  of  mixed-species  pairs 
document  the  occasional  hybridization  be- 
tween B.  hypoleucus  and  the  Golden-crowned 
Warbler  (B.  culicivorus;  Hellmayr  1935,  Wil- 
lis 1986,  Remsen  and  Traylor  1989,  Silva 
1991),  although  this  literature  has  been  over- 
looked in  recent  paruline  summations  (Ridge- 
ly and  Tudor  1989,  Curson  et  al.  1994).  All 
four  specimens  prepared  as  skins  from  San 
Luis  show  signs  of  hybridization  with  varying 
amounts  of  pale  yellow  on  the  center  of  the 
abdomen  and  lower  flanks  (compared  with 
color  plate  and  descriptions  in  Silva  1991). 
We  agree  with  prior  assessments  that  this  col- 
or is  indicative  of  gene  flow  between  B.  hy- 
poleucus and  B.  culicivorus,  and  not  with  B. 
flaveolus  as  implied  in  Hayes  (1995a)  for  the 
following  reasons.  Contrary  to  S.  LaBar  (in 
Hayes  1995a),  the  vocalizations  of  B.  hypo- 
leucus are  not  more  similar  to  those  of  B.  flav- 
eolus than  to  those  of  B.  culicivorus.  In  fact, 
the  vocalizations  of  B.  hypoleucus  and  B.  cul- 
icivorus are  so  similar  that  both  species  re- 
spond to  play-back  recordings  of  each  other’s 


song  (Silva  1991).  Neither  of  these  taxa’s  song 
resembles  those  of  B.  flaveolus,  whose  vocal- 
izations and  behavior  clearly  align  it  with  the 
Phaeothlypis  assemblage  (Ridgely  and  Tudor 
1989;  MBR,  pers.  obs.).  Furthermore,  the 
plumages  and  behaviors  of  B.  hypoleucus  and 
B.  culicivorus  are  very  similar  (Silva  1991; 
MBR,  pers.  obs.),  unlike  the  phaeothlypine- 
like  B.  flaveolus.  We  originally  interpreted 
Hayes’  (1995a)  statement  of  “a  mixed  family 
of  B.  hypoleucus  and  B.  flaveolus  (two  of 
each)  at  Estancia  Fonciere  [depto.  Concep- 
cion]” as  meaning  mixed-species  pairs;  his  in- 
tent, however,  was  to  report  “two  of  each  spe- 
cies intermingling  together”  (F.  E.  Hayes, 
pers.  comm.).  Willis  (1986)  observed  only  a 
single  mixed  pair  of  B.  hypoleucus  and  B.  cul- 
icivorus-, all  other  pairs  were  conspecific.  Nev- 
ertheless, Silva  (1991)  documented  more  ex- 
tensive hybridization  in  southeastern  Brazil 
and  treated  the  two  species  as  conspecific.  To 
our  knowledge,  all  apparent  hybrid  morpho- 
types  are  B.  hypoleucus-\\ke,,  i.e.,  ventrally 
yellow  B.  culicivorus-XWe  birds  with  some 
white  are  unknown.  Presumably  the  situation 
in  Paraguay  is  similar  to  that  in  Brazil  (Willis 
1986,  Silva  1991),  with  hybridization  appar- 
ently occumng  only  at  ecotones  where  the  dry 
forest  inhabiting  B.  hypoleucus  meets  the 
moist  forest  inhabiting  B.  culicivorus.  Typical 
B.  culicivorus  have  been  collected  at  Parque 
Nacional  Cerro  Cora,  depto.  Amambay,  only 
about  130  km  east  of  San  Luis  (Fig.  1),  where 
patches  of  moist  forest  are  present  (MNHNP 
specimens,  Hayes  and  Scharf  1995). 

Screaming  Cowbird  (Molothrus  rufoaxil- 
laris). — Throughout  most  of  its  range,  this 
brood  parasite  has  been  documented  to  spe- 
cialize on  Bay-winged  Cowbirds  (Molothrus 
badius',  Friedmann  1963,  Fraga  1986);  how- 
ever, Sick  (1993)  and  Fraga  (1996)  demon- 
strated that  M.  rufoaxillaris  pai'asitizes  the 
Chopi  Blackbird  {Gnorimopsar  chopi)  in  the 
absence  of  M.  badius.  Our  limited  observa- 
tions at  San  Luis  also  suggest  that.  M.  rufoax- 
illaris is  parasiting  a host  other  than  M.  bad- 
ius. Male  M.  rufoaxillaris  and  Shining  Cow- 
birds  {Molothrus  bonariensis)  were  observed 
displaying  within  15  m of  each  other  to  black- 
ish-appearing female  cowbirds  in  areas  where 
the  open  forest  and  campo  interdigitated.  Giv- 
en that  no  M.  badius  were  observed  at  San 
Luis,  we  presume  that  M.  rufoaxillaris  was 


Rohhins  et  al.  • PARAGUAYAN  CERRADO  AVIFAUNA 


221 


paiasitizing  Gnoriniopsar  chopi,  which  was 
fairly  common  in  the  area.  Unless  male  M. 
nifoaxillahs  and  M.  bonariensis  are  vocaliz- 
ing, they  ai'e  extremely  difficult  to  distinguish 
under  field  conditions  (Ridgely  and  Tudor 
1989).  However,  Ridgely  and  Tudor  (1989) 
stated  that  pronounced  plumage  differences  in 
females  (grayish-brown  in  M.  bonahensis\ 
blackish  in  M.  nifoaxillaris)  could  aid  in  iden- 
tification in  areas  of  sympatry.  Our  work  at 
San  Luis,  coupled  with  museum  and  literature 
reviews,  demonstrates  that  an  unknown  pro- 
portion of  female  M.  bonariensis  in  southern 
South  America  also  have  blackish  plumage. 
Two  adult  female  M.  bonariensis  collected  at 
San  Luis  (KU  88485,  MNHNP;  others  ob- 
served) were  glossy,  bluish-black  on  all  but 
the  lower  ventral  surface  and  tail.  Under  field 
conditions,  these  females  closely  resembled 
accompanying  males;  only  under  excellent 
light  conditions  were  plumage  differences  de- 
tected. This  male-like  plumage  in  female  M. 
bonariensis  is  not  limited  to  Paraguay,  as  blu- 
ish-black females  that  are  very  similar  to  the 
San  Luis  birds  have  also  been  collected  from 
southeastern  Brazil  and  extreme  northeastern 
Argentina  (Friedmann  1927,  Sick  1993;  KU 
65019).  Observers  should  exercise  caution  in 
identifying  these  cowbirds  in  areas  of  sym- 
patry based  on  female  plumage  patterns.  Data 
are  needed  on  what  proportion  of  nominate 
female  M.  bonariensis  have  male-like  plum- 
age. 

DISCUSSION 

Silva  (1995a)  used  two  criteria — a mini- 
mum of  100  species  recorded  and  at  least  80 
specimens  collected — for  defining  a Cerrado 
locality  as  minimally  sampled.  Based  on  his 
criteria,  our  results  (181  species,  220  speci- 
mens) would  be  classified  as  a relatively  thor- 
ough Cerrado  inventory;  however,  our  results 
are  clearly  preliminary.  The  inclusion  of 
Krieg’s  results  (Laubmann  1939,  1940)  puts 
the  Serrania  de  la  San  Luis  species  list  at  219. 
With  additional  work,  we  predict  that  the 
Parque  Nacional  Serrania  San  Luis  avifauna 
will  surpass  250  species.  Surveys  are  needed 
at  the  drier  northern  end  of  the  park,  where 
such  species  as  Rusty-backed  Antwren  {For- 
micivora  rufa)  and  Rufous-sided  Pygmy-Ty- 
rant {Euscarthmus  rufomarginatus)  may  oc- 
cur. From  our  limited  observations,  it  is  clear 


that  San  Luis  is  an  important  site  for  grassland 
and  forest  Austral  migrants,  and  we  predict 
that  the  majority  of  the  species  to  be  added  to 
the  San  Luis  list  will  be  from  this  component. 
Year  round  surveys  are  needed  to  clarify  the 
status  of  many  species. 

Although  San  Luis  is  one  of  the  Cerrado 
sites  closest  to  the  Chaco,  only  one  species 
primarily  restricted  to  the  Chaco  was  found 
there:  the  Great  Rufous  Woodcreeper  (Xiph- 
ocolaptes  major).  Considerable  plumage  var- 
iation in  the  two  specimens  collected  of  this 
species  obfuscates  subspecific  determination. 
For  the  Olivaceous  Woodcreeper  (Sittasomus 
griseicapillus),  the  nominate  form,  principally 
of  the  Chaco,  was  present;  for  the  Narrow- 
billed Woodcreeper  (Lepidocolaptes  angusti- 
rostris)  specimens  collected  were  assignable 
to  the  ventrally  heavily  streaked  nominate 
form  than  to  the  ventrally  unmarked  Cerrado 
form,  L.  a.  bivittatus. 

The  Atlantic  Forest  avifauna  is  also  poorly 
represented  at  San  Luis,  with  only  six  species 
recorded:  Reddish-bellied  Parakeet  (Pyrrhura 
frontalis).  Variable  Screech-Owl  (Otus  atri- 
capillus),  Rufous-capped  Motmot  (Bary- 
phthengus  ruficapillus).  Ochre-breasted  Fo- 
liage-gleaner (Philydor  lichtensteini).  Eared 
Pygmy-Tyrant  {Myiornis  auricularis),  and  the 
Bare-throated  Bellbird.  The  minimal  Atlantic 
Forest  avifaunal  contribution  to  San  Luis  con- 
trasts with  that  of  the  only  other  Paiaguyan 
Cenado  site  that  has  been  adequately  inven- 
toried. A total  of  201  species  have  been  re- 
corded at  nearby  Parque  Nacional  CeiTO  Cora 
(Hayes  and  Scharf  1995;  R.  Clay,  pers. 
comm.).  Twenty-five  Atlantic  Forest  species, 
including  all  six  known  from  San  Luis,  have 
been  reported  from  Cerro  Cora.  Eliminating 
water  inhabiting  species  (herons,  waterfowl, 
migratory  shorebirds,  kingfishers,  marsh 
dwelling  passerines),  61  Ceno  Cora  species 
have  not  been  recorded  at  San  Luis,  whereas 
62  species  have  been  recorded  at  San  Luis  and 
not  Cerro  Cora.  Additional  surveys  at  both 
sites  undoubtedly  will  reduce  the  uniqueness 
of  each  site;  nevertheless,  the  presence  of  hu- 
mid forest  patches  at  Ceno  Cora  and  campo 
at  San  Luis  explain  genuine  species  differenc- 
es between  these  sites.  The  humid  forest  ele- 
ment at  Cerro  Cora  is  composed  not  only  of 
species  primarily  restricted  to  Atlantic  Forest, 
but  also  of  species  more  widely  distributed  in 


222 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


humid  forest  across  much  of  South  America, 
such  as  Plain  Antvireo  {Dysithamnus  menta- 
lis)  and  Red-rumped  Cacique  {Cacicus  hae- 
morrhous).  Moreover,  the  difference  in  the 
number  of  hummingbird  species  between  the 
two  areas  is  striking,  with  ten  species  [Ru- 
fous-throated Sapphire  {Hylocharis  sapphiri- 
na)  being  added  since  Hayes  and  Scharf  1995; 
R.  Clay,  pers.  comm.]  known  from  Cerro 
Cora,  but  only  two  from  San  Luis.  The  pro- 
nounced avifaunal  differences  between  these 
nearby  localities  support  Silva’s  (1995c)  as- 
sertion that  additional  Cerrado  inventories  are 
needed. 

Only  3 of  the  29  Cenado  endemic  species 
(Silva  1995c)  have  been  found  at  San  Luis. 
We  predict  that  additional  work  in  the  area 
will  document  at  least  6 more  Ceirado  endem- 
ics: White-winged  Nightjar  {Caprimulgiis 
candicans),  White-lored  Spinetail  {Synallaxis 
albilora).  Chestnut-capped  Foliage-gleaner 
(Hylocryptus  rectirostris).  Collared  Cres- 
centchest  {Melanopareia  torquata).  Reiser’s 
Tyrannulet  (Phyllomyias  reiseri),  and  Helmet- 
ed  Manakin  (Antilophia  galeata),  which  have 
been  found  at  other  Paraguayan  Cerrado  sites 
(Laubmann  1939,  1940;  Hayes  1995a;  Lowen 
et  al.  1996a).  Hence,  maximum  Cerrado  en- 
demic species  diversity  at  San  Luis  is  expect- 
ed to  be  about  nine  species.  Many  of  the  re- 
maining Cerrado  endemic  species  have  re- 
stricted geographic  ranges;  six  have  extremely 
small  distributions  (Silva  1995c).  Indeed,  one 
generalization  that  can  be  made  of  the  Cerrado 
endemics  is  that  most  are  not  widespread  in 
the  region.  Even  in  the  Cerrado’s  geographic 
center  (eastern  Mato  Grosso/west-central 
Goias),  only  about  two-thirds  of  the  endemics 
have  been  found.  Several  factors  may  contrib- 
ute to  the  somewhat  reduced  species  diversity 
in  Paraguay,  (1)  less  than  an  estimated  5%  of 
the  total  Cerrado  habitat  is  found  in  Paraguay; 
(2)  northern  Paraguay  is  at  the  southern  limit 
of  this  habitat’s  cunent  distribution;  and  (3) 
Paraguay  has  not  had  the  geographic  isolating 
mechanisms  that  have  promoted  speciation  in 
other  parts  of  the  CeiTado.  However,  compa- 
rable numbers  of  endemics  to  Paraguay  are 
found  in  other  peripheral  Cenado  areas,  such 
as  western  Mato  Grosso  and  northern  Goias; 
hence,  the  Paraguayan  Cenado  is  not  espe- 
cially depauperate  when  compared  across  the 
range  of  the  habitat. 


The  importance  of  Parque  Nacional  San 
Luis  cannot  be  overly  stressed.  Since  our  in- 
ventory the  area  that  we  inventoried  was  trad- 
ed for  an  adjacent,  less  human  impacted  sec- 
tor, formerly  known  as  Estancia  Garay  Kue 
(R.  Clay  and  A.  Madrono,  pers.  comm.).  The 
new  park  boundary  apparently  encloses  about 
4500  ha  of  relatively  pristine  grassland.  This 
takes  on  special  conservation  significance  giv- 
en that  virtually  all  grassland  habitat  in  Pai'a- 
guay  has  been  moderately  impacted  by  hu- 
mans (Clay  et  al.  1998).  Indeed,  grasslands 
are  one  of  the  most  threatened  habitats  on  the 
continent  (Goriup  1988,  Bates  et  al.  1992, 
Stotz  et  al.  1996,  Silva  et  al.  1997).  The  grass- 
lands at  San  Luis  are  critical  to  several  avian 
species,  both  resident  and  migrant.  Three  of 
the  neai-threatened  avian  species  are  grass- 
land-inhabiting: Greater  Rhea  {Rhea  ameri- 
cana).  Cock-tailed  Tyrant  {Alectnirus  tricol- 
or), and  Dark-throated  Seedeater  {Sporophila 
ruficollis).  Moreover,  we  suspect  that  the 
poorly-known  and  critically  endangered 
White-winged  Nightjar  {Caprimulgiis  candi- 
cans) occurs  in  the  San  Luis  area.  This  spec- 
tacular nightjar  was  only  recently  documented 
in  Paraguay  (Lowen  et  al.  1996a),  and  quite 
suiprisingly,  in  eastern  Bolivia  (Davis  and 
Flores  1994).  Prior  to  these  recent  records  it 
was  known  only  from  a few  sites  in  central 
Brazil  (Collar  et  al.  1992).  Further  surveys  of 
this  pristine  grassland  may  yield  additional 
grassland  specialists,  e.g..  Bearded  Tachuri 
{Polystictus  pectoralis)  and  Sharp-tailed  Ty- 
rant {Cidicivora  caiidacuta),  that  are  declining 
(Collar  and  Wege  1995,  Stotz  et  al.  1996, 
Parker  and  Willis  1997). 

We  commend  the  Paraguayan  government 
for  having  the  foresight  to  establish  San  Luis 
and  Cerro  Cora  national  parks.  We  hope  that 
our  data  will  help  underscore  the  importance 
of  these  parks,  and  that  this  information  may 
be  used  in  setting  conservation  priorities  for 
the  rapidly  disappearing  Cenado. 

ACKNOWLEDGMENTS 

Director  O.  Romero  of  Departamento  del  Inventario 
Biologico  Nacional  and  MNHNP  was  indispensable  in 
too  many  ways  to  mention.  C.  Fox.  director  of  Direc- 
cion  de  Parques  Nacionales  y Vida  Silvestre  graciously 
granted  permits  for  our  work  al  San  Luis  National 
Park.  A.  L.  Aquino,  Director  of  CITES-Paraguay, 
kindly  provided  logistical  help  in  getting  to  San  Luis. 
We  thank  San  Luis  park  guard,  A.  Acosta,  for  accom- 


Rohhins  et  al.  • PARAGUAYAN  CERRADO  AVIFAUNA 


223 


modating  us.  KU  herpelologist  J.  Simmons  provided  a 
litany  of  stories,  recorded  weather  data,  and  kept  us 
from  harm’s  way  of  the  Neotropical  rattlesnake.  D. 
Stotz,  T.  Schulenberg,  M.  Adams,  R.  Clay,  and  P. 
Sweet  provided  information  from  specimen  material 
not  available  to  us.  A.  Whittaker  kindly  shared  his  re- 
cordings of  Philydor  dimidiatiis.  K.  McVay  translated 
German  literature  and  C.  Freeman  clarified  vegetation 
nomenclature.  R.  Clay,  M.  Foster,  F.  Hayes,  A.  Madro- 
no, J.  M.  C.  da  Silva,  and  T.  Peterson  provided  valu- 
able comments  on  the  manuscript.  We  thank  the  Field 
Museum  of  Natural  History  (FMNH;  J.  Bates,  D.  Wil- 
lard) and  the  Fos  Angeles  County  Museum  of  Natural 
History  (FACMNH;  K.  GaiTett)  for  loan  of  specimens. 
Funding  in  part  was  through  the  Kansas  University 
Ornithology  Interest  Group.  We  especially  want  to 
thank  the  Burroughs  Audubon  Society  for  their  gen- 
erous support. 

LITERATURE  CITED 

Ab’Saber,  a.  N.  1977.  Os  Dommios  morfoclimaticos 
da  America  do  Sul.  Primerira  Aproxima9ao.  Geo- 
morfologia  52:1-21. 

Bates,  J.  M.,  T.  A.  Parker,  III,  A.  P.  Capparella,  and 
T.  J.  Davis.  1992.  Observations  on  the  campo,  cer- 
rado,  and  forest  avifaunas  of  eastern  Dpto.  Santa 
Cruz,  Bolivia,  including  21  species  new  to  the 
country.  Bull.  Brit.  Ornithol.  Club  1 12:86-98. 
Belton,  W.  1985.  Birds  of  Rio  Grande  do  Sul,  Brazil. 
Part  2.  Formicariidae  through  Corvidae.  Bull.  Am. 
Mus.  Nat.  Hist.  180:1-242. 

Brown,  L.  H.  and  D.  Amadon.  1968.  Eagles,  hawks, 
and  falcons  of  the  world.  McGraw-Hill  Book  Co., 
New  York. 

Chapman,  F.  M.  1923.  The  distribution  of  the  motmots 
of  the  genus  Momotus.  Bull.  Am.  Mus.  Nat.  Hist. 
48:27-59. 

Chesser,  R.  T.  1997.  Patterns  of  seasonal  and  geo- 
graphical distribution  of  Austral  migrant  flycatch- 
ers (Tyrannidae)  in  Bolivia.  Ornithol.  Monogr.  48: 
171-204. 

Christiansen,  M.  B.  and  E.  Fitter.  1997.  Species  loss 
in  forest  bird  community  near  Lagoa  Santa  in 
southeastern  Brazil.  Biol.  Conserv.  80:23-32. 
Clay,  R.  R,  D.  R.  Capper,  J.  M.  Barnett,  I.  J.  Bur- 
field,  E.  Z.  Esquivel,  R.  Farina,  C.  P.  Kennedy, 
M.  Perrens,  and  R.  G.  Pople.  1998.  White- 
winged Nightjars  Caprimidgus  ccindiccins  and  cer- 
rado  conservation:  the  key  findings  of  Project 
Aguara  N.  1997.  Cotinga  9:52-56. 

Collar,  N.  J.,  L.  P.  Gonzaga,  N.  Krabbe,  A.  Mad- 
rono Nieto,  L.  G.  Naranjo,  T.  A.  Parker,  III, 
AND  D.  C.  Wege.  1992.  Threatened  birds  of  the 
Americas.  The  ICBP/IUCN  Red  Data  Book,  third 
ed.  Part  2.  Inter.  Council  for  Bird  Preservation, 
Cambridge,  U.K. 

Collar,  N.  J.  and  D.  C.  Wege.  1995.  The  distribution 
and  conservation  status  of  the  Bearded  Tachuri 
Polystictus  pectondis.  BirdLife  Int.  5:367-390. 
Collar,  N.  J.,  M.  J.  Crosby,  and  A.  J.  Stattersfield. 
1994.  Birds  to  watch  2:  the  world  list  of  threat- 


ened birds.  BirdLife  International,  Cambridge, 
U.K. 

CuRSON,  J.,  D.  Quinn,  and  D.  Beadle.  1994.  Warblers 
of  the  Americas.  Houghton  Mifflin  Co.,  Boston, 
Massachusetts. 

Davis,  S.  E.  and  E.  Flores.  1994.  Firsi  record  of 
White-winged  Nightjar  Caprimulgus  cundicans 
for  Bolivia.  Bull.  Brit.  Ornithol.  Club  114:127— 
128. 

Del  Hoyo,  J.,  A.  Elliot,  and  J.  Sargatal  (Eds.). 
1994.  Handbook  of  birds  of  the  world.  Vol.  2. 
New  World  vultures  to  guineafowl.  Lynx  Edi- 
cions,  Barcelona,  Spain. 

Direccion  de  Barques  Nacionales  y Vida  Silvestre. 
1993.  Plan  estrategico  del  Sistema  Nacional  de 
Areas  Silvestres  Protegidas  del  Paraguay  (SINA- 
SIP).  Ministerio  de  Agricultura  y Ganaden'a,  Sub- 
secretan'a  de  Estado  de  Recursos  Naturales  y Me- 
dio Ambiente,  Direccion  de  Barques  Nacionales  y 
Vida  Silvestre,  Asuncion,  Paraguay. 

Ericson,  P.  G.  P.  and  L.  a.  Amarilla.  1997.  First  ob- 
servations and  new  distributional  data  for  birds  in 
Paraguay.  Bull.  Brit.  Ornithol.  Club  1 17:60—67. 

Fraga,  R.  M.  1986.  The  Bay-winged  Cowbird  (Mol- 
othrus  badius)  and  its  brood  parasites:  interac- 
tions, coevolution  and  cooperative  efficiency. 
Ph.D.  diss.,  Univ.  of  California,  Santa  Barbara. 

Fraga,  R.  M.  1996.  Further  evidence  of  parasitism  of 
Chopi  Blackbirds  (Gnorimopsar  chopi)  by  the 
specialized  Screaming  Cowbird  (Molothrus  ru- 
foaxillaris).  Condor  98:866-867. 

Friedmann,  H.  1927.  A revision  of  the  classification 
of  the  cowbirds.  Auk  44:495-508. 

Friedmann,  H.  1963.  Host  relations  of  the  parasitic 
cowbirds.  U.S.  Natl.  Mus.  Bull.  223:1-276. 

Goriup,  P.  D.  (Ed.).  1988.  Ecology  and  conservation 
of  grassland  birds.  Inter.  Council  for  Bird  Preser- 
vation and  Smithsonian  Institution  Press,  Wash- 
ington, D.C. 

Haeter,  j.  1985.  Avian  zoogeography  of  the  Neotrop- 
ical lowlands.  Ornithol.  Monogr.  36:113—146. 

Hayes,  F.  E.  1995a.  Status,  distribution,  and  biogeog- 
raphy of  the  birds  of  Paraguay.  Monogr.  Field  Or- 
nithol. 1:1-230. 

Hayes,  F.  E.  1995b.  Definitions  for  migrant  birds:  what 
is  a Neotropical  migrant?  Auk  112:521—523. 

Hayes,  F.  E.  and  P.  A.  Scharf.  1995.  The  birds  of 
Parque  Nacional  Ceno  Cora,  Paraguay.  Cotinga 
4:20-24. 

Hayes,  F.  E.,  Scharf,  P.  A.,  and  R.  S.  Ridgely.  1994. 
Austral  bird  migrants  in  Paraguay.  Condor  96:83- 
97. 

Hellmayr,  C.  E.  1935.  Catalogue  of  birds  of  the 
Americas.  Field  Mus.  Nat.  Hist.,  Zool.  Ser.  13(8): 
1-541. 

Laubmann,  a.  1939.  Die  Vogel  von  Paraguay.  Vol.  1. 
Strecker  und  Schroder,  Stuttgart,  Germany. 

Laubmann,  A.  1940.  Die  Vogel  von  Paraguay.  Vol.  2. 
Strecker  und  Schroder,  Stuttgart.  Germany. 

Lowen,  j.  C.,  L.  Bartrina,  T.  M.  Brooks,  R.  P.  Clay, 
AND  J.  A.  Tobias.  1996a.  Project  YACUTINGA 


224 


THE  WILSON  BULLETIN  • Vol.  HI.  No.  2.  June  1999 


'95;  bird  surveys  and  conservation  priorities  in 
eastern  Paraguay.  Cotinga  5:14-19. 

Lowen,  J.  C.,  L.  Bartrina,  R.  P.  Clay,  and  J.  A.  To- 
bias. 1996b.  Biological  surveys  and  conservation 
priorities  in  eastern  Paraguay.  CSB  Conservation 
Publications,  Cambridge,  U.K. 

Madrono  Nieto,  A.  and  E.  Z.  Esquivel.  1995.  Re- 
serva  Natural  del  Bosque  Mbaracayu:  su  impor- 
tancia  en  la  conservacion  de  aves  amenazadas, 
cuasi-amenazadas  y endemicas  del  bosque  Atlan- 
tico.  Cotinga  4:52-57. 

Marin,  M.  1997.  Species  limits  and  distribution  of 
some  New  World  spine-tailed  swifts  (Chaetura 
spp.).  Ornithol.  Monogr.  48:431-444. 

McEarlane,  R.  W.  1974.  Unusual  avian  migrants  in 
Tarapaca.  Idesia  3:181-184. 

Parker,  T.  A.,  Ill  and  E.  O.  Willis.  1997.  Notes  on 
three  tiny  grassland  flycatchers,  with  comments  on 
the  disappearance  of  South  American  fire-diver- 
sified savannas.  Ornithol.  Monogr.  48:549—555. 

Remsen,  j.  V.,  Jr.  and  M.  A.  Traylor,  Jr.  1983.  Ad- 
ditions to  the  avifauna  of  Bolivia  Part  2.  Condor 
85:95-98. 

Remsen,  J.  V.,  Jr.  and  M.  A.  Traylor,  Jr.  1989.  An 
annotated  list  of  the  birds  of  Bolivia.  Buteo 
Books,  Vermillion,  South  Dakota. 

Ridgely,  R.  S.  and  G.  Tudor.  1989.  The  birds  of 
South  America.  Vol.  1.  The  oscine  passerines. 
Univ.  of  Texas  Press,  Austin. 

Ridgely,  R.  S.  and  G.  Tudor.  1994.  The  birds  of 
South  America.  Vol.  2.  The  suboscine  passerines. 
Univ.  of  Texas  Press,  Austin. 

Robbins,  M.  B.  and  T.  A.  Parker,  III.  1997.  Voice 
and  taxonomy  of  Caprimulgus  (rufus)  otiosus 
(Caprimulgidae),  with  a reevaluation  of  Capri- 
miilgii.s  rufus  subspecies.  Ornithol.  Monogr.  48: 
601-607. 

Short,  L.  L.  1975.  A zoogeographic  analysis  of  the 


South  American  Chaco  avifauna.  Bull.  Am.  Mus. 
Nat.  Hist.  154:167-352. 

Sick,  H.  1993.  Birds  in  Brazil.  Princeton  Univ.  Press, 
Princeton,  New  Jersey. 

Silva,  J.  M.  C.  da.  1995a.  Avian  inventory  of  the  Cer- 
rado  Region,  South  America:  implications  for  bi- 
ological conservation.  Bird  Conserv.  Int.  5:291  — 
304. 

Silva,  J.  M.  C.  da.  1995b.  Biogeographic  analysis  of 
the  South  American  cerrado  avifauna.  Steenstru- 
pia  21 :49-67. 

Silva,  J.  M.  C.  da.  1995c.  Birds  of  the  cerrado  region. 
South  America.  Steenstrupia  21:69—92. 

Silva,  J.  M.  C.  da,  D.  C.  Oren,  J.  C.  Roma,  and  L. 
M.  P.  Henriques.  1997.  Composition  and  distri- 
bution of  the  avifauna  of  an  Amazonian  upland 
savanna,  Amapa,  Brazil.  Ornithol.  Monogr.  48: 
743-762. 

Silva,  W.  R.  1991.  Padroes  ecologicos,  bioacusticos, 
biogeograficos  e filogeneticos  do  complex  Basi- 
leuterus  culicivorus  (Aves,  Parulidae)  e demais 
especies  brasileiras  do  genero.  Ph.D.  diss.,  Univ. 
Estadual  de  Campinas,  Brazil. 

Snow,  D.  W.  1973.  Distribution,  ecology,  and  evolu- 
tion of  the  bellbirds  (Procnias,  Cotingidae).  Bull. 
Brit.  Mus.  (Nat.  Hist.)  Zool.  25:369-391. 

Snow,  D.  1982.  The  cotingas.  Cornell  Univ.  Press,  Ith- 
aca, New  York. 

Stotz,  D.  E,  j.  W.  Eitzpatrick,  T.  A.  Parker,  III,  and 
D.  K.  Moskovits.  1996.  Neotropical  birds:  ecol- 
ogy and  conservation.  Univ.  of  Chicago  Press, 
Chicago,  Illinois. 

Willis,  E.  O.  1979.  The  composition  of  avian  com- 
munities in  remanescent  woodlots  in  southern 
Brazil.  Pap.  Avulsos  Zool.  (Sao  Paulo)  33:1-25. 

Willis,  E.  O.  1986.  Vireos,  wood  warblers  and  war- 
blers as  ant  followers.  Gerfaut  76:177—186. 

Willis,  E.  O.  and  Y.  Oniki.  1993.  New  and  confimied 
birds  from  the  state  of  Sao  Paulo,  Brazil,  with 
notes  on  disappearing  species.  Bull.  Brit.  Ornithol. 
Club  113:23-34. 


Rohhins  el  cil.  • PARAGUAYAN  CERRADO  AVIFAUNA 


225 


APPENDIX.  Avian  species 
cepcion,  Paraguay. 

and  their  relative  abundance  in 

Parque  Nacional  Serrania  San  Luis,  depto.  Con- 

Relative 

abundance^ 

Status^ 

Habitat‘s 

Documentation** 

Rhea  americana'^ 

u 

p 

I 

P 

Crypturellus  imdulatiis 

c 

p 

2 

V 

Cryptiirellus  parvirostris 

c 

p* 

I 

c,v 

Rhynchotus  rufescens 

u 

p 

I 

V 

Nothura  maculosa 

c 

p 

s 

Phalacrocorax  brasilianus 

R 

V 

4 

s 

Syrigma  sibilatrix 

R 

p* 

I 

c 

Ardea  alba 

L 

Bubulcus  ibis 

C 

V 

I 

s 

Butorides  striatus 

L 

Nycticorax  nycticorax 

L 

Phimosus  infuscatus 

C 

V 

I 

S 

Theristiciis  caiidatus 

u 

p* 

2 

C 

Mycteria  americana 

R 

V 

4 

s 

Corcigyps  atratus 

U 

p 

2 

s 

Cathartes  aura 

u 

p 

2 

s 

Sarcoramphus  papa 

u 

p 

3 

s 

Cairina  moschata 

R 

p 

I 

s 

Leptodon  cayanensis 

R 

p 

3 

c 

Elanus  leucurus 

R 

p 

I 

s 

Rostrhamus  sociabilis 

R 

AM 

I 

s 

Ictinia  plumbea 

R 

B 

2 

s 

Accipiter  bicolor 

U 

P 

2 

c 

Buteogallus  meridionalis 

X 

P 

1 

s 

Harpyhaliaetus  coronatus 

X 

P 

2 

s 

Buteo  magnirostris 

L 

Spizaetus  ornatus 

R 

P 

3 

s 

Caracara  plane us 

u 

P 

I 

s 

Milvago  chimachima 

R 

P 

I 

L 

Herpetotheres  cachinnans 

R 

P 

2 

L 

Micrastur  ruficollis 

L 

Falco  sparverius 

R 

P 

I 

L 

Crax  fasciolata 

U 

P 

3 

V 

Aramides  cajanea 

L 

Porzana  albicollis 

L 

Cariama  cristata 

u 

P 

I 

P 

Vanellus  chilensis 

u 

Pn 

I 

P 

Rostratula  semicollaris 

X 

V? 

I 

S 

Tringa  solitaria 

L 

Calidris  fuscicollis 

L 

Gallinago  paraguaiae 

L 

Columba  cayennensis 

c 

P 

3 

V 

Columbina  squammata 

u 

P 

1 

S 

Columbina  talpacoti 

c 

P 

1 

V 

Columbina  picui 

X 

p* 

1 

C 

Claravis  pretiosa 

c 

p* 

2 

c,v 

Leptotila  verreauxi 

c 

P 

1 

S 

Leptotila  rufaxilla 

u 

p* 

3 

c 

Anodorhynchus  hyacinthinus 

X 

P 

2 

s 

Ara  chloropterus 

L 

Ara  maracana 

L 

Aratinga  leucophthalmus 

c 

P* 

3 

c,v 

Aratinga  aurea 

L 

Pyrrhura  frontalis 

c 

P 

3 

c,v 

Brotogeris  chiriri 

L 

Pionus  maximiliani 

u 

P 

2 

S 

Amazona  aestiva 

u 

P 

3 

V 

226 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


APPENDIX.  Continued 

Relalive 

abundance^ 

Status*’ 

Habitat*-' 

Documentation** 

Coccyziis  melacof-yphus 

u 

B* 

2 

c 

Piayci  cayana 

u 

P 

2 

L 

Tupera  naevia 

u 

P 

2 

V 

Crotophaga  major 

R 

B 

2 

L 

Crotophaga  ani 

C 

P 

1 

V 

Guira  guira 

C 

P 

1 

V 

Otus  atricapilliis 

U 

p* 

3 

C 

Otus  choliba 

U 

P 

2 

s 

Piilsatri.x  per.spicillata 

Bubo  virginianu.s 

Glaiicidium  bra.'iilianiim 

U 

p* 

2 

L 

L 

c,v 

Speotylo  cimicularia 

u 

Py 

1 

P 

Lurocalis  seniitorquatu.s 

c 

P? 

3 

V 

Chordeile.s  minor 

Podager  nacunda 

Nyctidromii.s  albicolli.s 

u 

p* 

2 

L 

L 

c,v 

Caprimulgits  rufiis 

u 

P* 

3 

c,v 

Caprimulgiis  parx  iilas 

u 

P*ny 

1 

C 

Nyctibius  gri.seiis 

u 

p* 

2 

c,v 

Chaetura  meridionali.s^ 

u 

IM 

4 

c,v 

TIudurania  furcata 

R 

P 

3 

C 

Hylocluiris  duy.stira 

u 

P 

2 

c,v 

Trogon  ciirucui 

u 

P 

3 

V 

Momotus  momota 

R 

p* 

3 

c,v 

Baryphthengii.s  ruficapilliis 

U 

p* 

3 

C 

Chloroceryle  americana 

R 

P 

2 

s 

Chlorocerxle  inda 

R 

P 

2 

s 

Ny.stalus  chacuru 

R 

P 

2 

V 

Ny.slalus  macular u.s 

R 

P 

2 

c 

Pteroglossu.s  castanotis 

Ramphastos  toco 

R 

P 

2 

L 

V 

Picumnu.s  cirratus 

X 

P 

2 

S 

MeUmerpe.s  candid  u.s- 

U 

P 

2 

V 

Picoide.s  mi.xtus 

Veniliorni.s  pa.s.serinu.s 

R 

P 

2 

L 

c,v 

Picul  u.s  ch  ry.soch  loros 

R 

P 

3 

s 

Colapte.s  campestris 

R 

P 

1 

L 

Celeus  lugubris 

R 

p* 

3 

C 

Dryocopu.s  1 ineatu.s 

R 

P 

2 

L 

Campepliilus  melanoleucos 

R 

P 

3 

L 

Furnarius  rufu.s 

Sclioeniopbyla.x  pluyganophila 

U 

P 

2 

L 

c,v 

S VI  la  lla.xi.s  a ! be  seen  s 

X 

P 

1 

L 

Pbacellodomu.s  rufifrons 

X 

p* 

2 

C 

Philydor  dimidiatiis 

R 

p* 

3 

c,v 

Philvdor  Uchtensteini 

U 

p* 

3 

c,v 

Philydor  rufu.s 

U 

p* 

3 

c,v 

Sirtasomus  griseicapiUus 

C 

p* 

3 

c,v 

Xiphocolapte.s  major 

R 

p* 

2 

C,V 

Dendrocolaptes  platyro.stris 

U 

p* 

3 

c,v 

Lepidocolapte.s  an  gusli  rostris 

U 

p* 

2 

• c,v 

ThamnophUus  caerule.scen.s 

C 

p* 

3 

c,v 

HerpsHochmus  atricapilliis 

U 

P 

3 

C,V  ' 

Camptostoma  ob.soletum 

U 

P?y 

2 

c,v 

Pliaeomyia.s  murina 

u 

B 

2 

V 

Suiriri  suiriri 

Myiopagis  caniceps 

u 

P? 

3 

L 

S 

Mviopagis  viridicata 

u 

p9* 

3 

c,v 

Flaenia  jlavogaster 

u 

p 

2 

V 

Robbins  el  cil.  • PARAGUAYAN  CERRADO  AVIF'AUNA 


227 


APPENDIX.  Conlinued. 

Kelalive 

abundance*' 

•Slams*’ 

Hahital^ 

Documentation^* 

Elaenia  albiceps 

R 

AM 

2 

c 

Elaenia  pcin  irostris 

X 

P? 

2 

s 

Inezici  inonuita 

R 

P?* 

2 

c,v 

Leptopogon  amciiirocephaliis 

U 

p* 

3 

c,v 

Corythopis  lielalandi 

U 

p* 

3 

c,v 

My iorn is  atiriciilaris 

L 

Hemitriccus  margaritacei venter 

u 

P 

2 

c,v 

Tohnomvius  sulphiirescens 

u 

p* 

3 

c,v 

Platvrinchiis  mvstaceus 

c 

p* 

3 

c,v 

Myiophobus  fasciatiis 

X 

P? 

s 

Contopns  cinereus 

R 

Pn 

2 

L 

Lathrotriceus  euleri 

U 

P?* 

3 

c,v 

CnemotriccHS  fiiscatus 

U 

P?* 

3 

c,v 

Py  roceph  ulus  rub  inns 

p* 

L 

Xobnis  cinerea 

u 

1 

C 

Xolmis  velata 

u 

p* 

1 

C 

Hymenops  perspicillatus 

L 

Eluvicola  leucocephala 

L 

Alectrurus  tricolor 

X 

P?* 

1 

C 

Gubernetes  yetcipa 

u 

p* 

1 

c,v 

Machelornis  rixosus 

u 

p* 

1 

c.v 

Casiornis  riifa 

R 

P 

3 

c,v 

Sirvstes  sibilator 

u 

p* 

3 

c,v 

Myiarchus  swciinsoni 

u 

P?* 

3 

c,v 

Myiarchiis  tyrannulus 

c 

p* 

2 

c,v 

Pitangus  sulphiiratus 

u 

Pn 

2 

s 

Megarynchus  pitangua 

L 

Myiodynastes  macidatiis 

c 

B* 

2 

C.V 

Legatus  leucophaius 

u 

B 

2 

V 

Empidonornus  variiis 

u 

B 

2 

L 

Griseotyrcinniis  aiirantioatrocristatus 

u 

B 

2 

C.V 

Tyrannus  melancholicus 

c 

By 

1 

L 

Tyrannus  scivana 

c 

Bn 

1 

V 

Xenopsaris  albinucha 

X 

P? 

3 

S 

Pachyramphus  viridis 

u 

Pn 

2 

C.V 

Pachyramphiis  castaneiis 

R 

P 

2 

s 

Pachyramphus  polychopterus 

R 

Bn 

2 

V 

Pachyramphus  validus 

R 

P?n 

2 

V 

Tityra  cayana 

R 

P 

3 

s 

Tityra  inquisitor 

u 

p* 

3 

c.v 

Pvroderus  sculatus 

u 

P 

3 

c 

Procnias  nudicollis 

R,C 

P?* 

3 

c.v 

Pipra  fasciicauda 

R 

P 

3 

c 

Oxynmcus  cristatus 

X 

P 

3 

c 

Progne  chalybea 

u 

P? 

1 

L 

Phaeoprocne  tapera 

1. 

Tachycineta  leucorrhoa 

R 

B 

1 

s 

Stelgidopteryx  ruficollis 

R 

B 

1 

L 

Cyanocorax  cyanomelas 

U 

P 

3 

C.V 

Cyanocorax  cristatelhis 

L 

Cyanocorax  chrysops 

X 

P 

3 

C 

Troglodytes  aedon 

c 

P 

2 

V 

Polioptila  dumicola 

R 

P 

2 

C.V 

Turdus  riifiventris 

U 

2 

C.V 

Turdus  albicollis 

C 

p* 

3 

C.V 

Turdus  amaurochalinus 

u 

P 

2 

L 

Catharus  fiiscescens 

X 

NM 

3 

C 

Mimus  saturninus 

u 

P 

1 

L 

228 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


APPENDIX.  Continued. 

Relative 

abundance' 

Status'’ 

Habitat' 

Documentation‘* 

Mimus  triurus 

L 

Vireo  olivaceus  chivi 

c 

P?* 

3 

c,v 

Cyclarhis  gujanensis 

u 

P 

2 

V 

Parula  pitiayumi 

u 

P 

3 

V 

Geothlypis  aequinoctialis 

R 

P 

1 

V 

Basileuterus  ftaveolus 

u 

P* 

3 

c,v 

Basileuterus  hypoleucus 

c 

P*y 

3 

c,v 

Conirostrum  speciosum 

L 

Euphonia  chlorotica 

L 

Thraupis  sayaca 

c 

P 

2 

V 

Eucometis  penicillata 

R 

p* 

3 

c,v 

Tachyphonus  rufus 

X 

p* 

2 

c 

Piranga  flava 

u 

P 

2 

L 

Trichothraupis  melanops 

R 

p* 

3 

C 

Cypsnagra  hirundinacea 

L 

Nemosia  pileata 

U 

P 

3 

V 

Hemithraupis  guira 

u 

p* 

3 

c,v 

Saltator  similis 

u 

p* 

2 

c,v 

Saltator  atricollis 

u 

p* 

1 

c,v 

Paroaria  coronata 

X 

P 

2 

s 

Coryphospingus  cucullatus 

c 

p* 

2 

c,v 

Arremon  flavirostris 

R 

p* 

3 

c,v 

Volatinia  jacarina 

C 

P 

1 

s 

Sporophila  plumbea 

u 

P? 

1 

c,v 

Sporophila  collaris 

L 

Sporophila  caerulescens 

c 

P? 

1 

c 

Sporophila  bouvreuil 

R 

p* 

1 

c 

Sporophila  hypoxantha 

X 

P? 

1 

c 

Sporophila  ruficollis 

u 

P? 

1 

c 

Sicalis  flaveola 

X 

P 

1 

s 

Embernagra  platensis 

L 

Emberizoides  herbicola 

c 

p* 

1 

c,v 

Ammodramus  humeralis 

c 

p* 

1 

c,v 

Zonotrichia  capensis 

R 

P 

1 

o 

s 

Gnorimopsar  chopi 

C 

P 

1 

V 

Pseudoleistes  guirahuro 

L 

Leistes  superciliaris 

L 

Molothrus  bonarien.sis 

u 

p* 

1 

C 

Molothrus  rufoaxillaris 

u 

p* 

1 

c,v 

Scaphidura  oryzivora 

2 

L 

Icterus  cayanensis 

u 

P 

2 

c,v 

Cacicus  chrysopterus 

u 

P* 

c,v 

Psarocolius  decumanus 

u 

Pn 

2 

V 

Carduelis  magellanica 

u 

P 

1 

s 

Passer  domesticus 

R 

P 

1 

s 

^ Relative  abundance  criteria  based  on  our  work.  X = single  observation;  R = rare,  not  recorded  daily  when  in  appropriate  habitat,  and  only  in  small 
numbers  when  recorded,  <5  individuals/day;  U = uncommon,  recorded  in  small  numbers  daily  when  in  appropriate  habitat,  l-IO  individual.s/day;  C = 
common,  recorded  daily  in  large  numbers  when  in  appropriate  habitat,  >10  individuals/day. 

•’  Status:  P = permanent  resident;  a breeder  with  at  least  a few  individuals  present  throughout  the  year.  Most  species  given  this  designation  are  not 
known  to  make  seasonal  movements;  virtually  all  with  P?  are  presumed  to  have  Austral  migrant  populations  present  at  appropriate  seasons;  B = breeder 
only.  Austral  migrant;  * = at  least  one  individual/species  had  enlarged  gonads  indicative  of  breeding,  n = nest  found,  y = young  observed;  V = visitor; 
nonbreeder,  but  may  breed  in  adjacent  areas  to  the  park;  IM  = Intratropical  migrant;  AM  = Austral  migrant;  NM  = Nearctic  migrant.  Migrant  terminology 
follows  Hayes  (l99.Sb);  ? = status  uncertain.  Hayes  et  al.  (1994)  and  Chesser  (1997)  were  consulted  for  Austral  migrant  status. 

Habitat  based  on  our  work;  I = open  gra.ssland  (campo);  2 = isolated  woodlots  in  grassland,  forest  edge;  3 = forest;  4 = aerial. 

‘'Documentation:  C = collected;  V = voice  recorded;  P = photographed;  we  use  this  designation  only  when  a species  was  photographed  but  not  "C” 
or  “V".  .S  = sight  observation(s)  only.  L = based  on  specimens  by  the  Kreig  expedition  to  this  region;  referenced  in  Laubmann  (1939,  1940);  .see  Methods 
for  explanation. 

'.Species  in  boldface  are  threatened  or  near-threatened  (Collar  et  al.  1992,  1994). 

f Following  Marin  (1997). 


Wilson  Bull.,  1 1 1(2),  1999,  pp.  229-235 


NOTES  ON  THE  AVIFAUNA  OF  TABASCO 

KEVIN  WINKER,'  STEFAN  ARRIAGA  WEISS, ^ JUANA  LOURDES  TREJO  P.,^ 

AND  PATRICIA  ESCALANTE  P' 


ABSTRACT. — Tabasco,  a Mexican  state  nearly  half  the  size  of  the  country  of  Costa  Rica,  lies  juxtaposed 
between  the  Isthmus  of  Tehuantepec  and  the  Yucatan  Peninsula.  This  state  hosts  a diverse  Neotropical  resident 
avifauna,  is  a significant  wintering  area  for  Nearctic-Neotropic  migrants,  and  has  important  biogeographic  sig- 
nificance. Surprisingly  little  recent  ornithological  study  has  occurred  in  Tabasco;  the  last  major  publication  treats 
data  from  1939.  Field  work  in  March  1996  and  recent  specimens  add  nine  species  to  the  state  list:  Caprimulgus 
vociferus,  Chaetura  vaiixi.  Campy lopterus  excellens,  Enipidonax  albigularis,  Thryothorus  modestus,  Turdus  in- 
fuscatus.  Myadestes  unicolor,  Limnothlypis  swainsonii,  and  Vermivora  ruficapilla.  The  status  of  26  other  species 
is  discussed.  Further  evidence  of  lowland  forests  being  used  as  temporary  refugia  by  birds  from  higher  elevations 
is  also  considered.  Received  20  May  1998,  accepted  19  Nov.  1998. 

“Compared  to  many  of  the  Mexican  States,  Tabasco  has  been  slighted  ornithologically.  Anything  collectors 
could  obtain  in  Tabasco  could  also  be  found  in  southern  Veracruz  with  less  effort.  If  the  collector  were  to 
exert  the  effort  necessary  to  enter  Tabasco,  he  might  as  well  go  the  entire  distance  and  travel  on  to  the 
Yucatan  Peninsula,  an  area  with  many  unique  and  peculiar  forms.”  Berrett  (1962:4). 


Tabasco,  a state  of  approximately  24,600 
km^,  is  situated  on  the  northeastern  side  of  the 
Isthmus  of  Tehuantepec  on  the  Gulf  of  Mex- 
ico (Fig.  1).  The  ornithological  neglect  of  Ta- 
basco noted  by  Berrett  (1962)  has  generally 
continued  for  the  past  35  years.  Berrett  (1962) 
added  166  species  to  those  previously  known 
from  Tabasco  and  reported  a total  of  457  spe- 
cies from  the  state.  The  bibliographic  and  mu- 
seum database  research  of  Centeno  ( 1 994) 
added  73  species  to  the  total  reported  by  Ber- 
rett (1962),  but  the  knowledge  of  avian  dis- 
tribution within  Tabasco  and  specimen  repre- 
sentation of  the  birds  of  the  state  remain  rel- 
atively poor.  Avian  specimens  apparently  do 
not  even  exist  from  3 of  the  state’s  17  muni- 
cipios  (the  equivalent  of  counties  in  the 
U.S.A.;  Centeno  1994). 

Perhaps  because  the  two  major  works  on 
Tabasco  birds  in  the  last  55  years  are  not  read- 
ily available  (Berrett  1962,  Centeno  1994)  and 


' Conservation  & Research  Center,  NZP,  Smithson- 
ian Institution,  1500  Remount  Rd.,  Front  Royal,  VA 
22630. 

^ Division  Academica  de  Ciencias  Biologicas,  Univ. 
Juarez  Autdnoma  de  Tabasco,  Km  0.5  Car.  Villahcr- 
mosa-Cardenas,  Villahermosa,  Tabasco  86000,  Mexi- 
co. 

^ Instituto  de  Biologi'a,  Depto.  de  Zoologfa,  Univ. 
Nacional  Autonoma  de  Mexico,  Apartado  Postal  70- 
153,  Mexico,  DF  04510. 

■*  Present  address:  Univ.  of  Alaska  Museum,  907  Yu- 
kon Dr.,  Fairbanks,  AK  99775-6960. 

■’Corresponding  author;  E-mail;  ffksw@uaf.edu 


because  specimen  representation  is  sparse,  er- 
rors and  omissions  regarding  the  distribution 
and  occurrence  of  birds  in  the  state  have  been 
fairly  common  (e.g.,  Peterson  and  Chalif 
1973,  Howell  and  Webb  1995).  The  last  major 
publication  on  the  birds  of  Tabasco  is  the  out- 
dated report  by  Brodkorb  (1943),  which  was 
based  on  a collection  made  in  the  region  by 
the  botanist  Eizi  Matuda  and  his  assistants  in 
1939. 

In  this  paper  we  report  on  our  recent  efforts 
to  increase  existing  knowledge  of  the  birds  of 
Tabasco.  In  March  1996  we  held  a collabo- 
rative field  workshop  in  the  municipios  of 
Centla  and  Huimanguillo.  In  Centla  (the 
northeastern  most  municipio  in  the  state,  bor- 
dering Campeche),  our  field  studies  were  con- 
ducted from  4—13  March  on  the  Rio  Grijalva 
in  the  Reserva  de  la  Biosfera  Pantanos  de 
Centla  (Fig.  1).  Our  studies  were  focused 
mainly  on  two  sites  in  the  area  of  18°  29'  N, 
92°  38'  W.  The  first  site  was  a remnant  tract 
of  mangrove  {Rhizophora  mangle)  and  puktal 
(Bucida  buceras)  forest  (a  selva  perennifolia, 
or  evergreen  forest)  along  the  banks  of  An  oyo 
Polo,  a tributary  of  the  Rio  Grijalva  about  4 
km  south  of  Frontera.  The  other  area  was  a 
partially  flooded,  shrubby  pasture  on  the 
banks  of  the  Rio  Grijalva  and  Rio  San  Pedrito, 
about  14  km  south  of  Frontera.  Additional  ob- 
servations were  made  on  water  and  land  be- 
tween these  two  sites. 

In  Huimanguillo,  the  southwestemmost 


229 


230 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  2,  June  1999 


FIG.  1.  The  Mexican  state  of  Tabasco,  with  its  capital  city  of  Villaheimosa  and  the  location  of  our  study 
sites  at  the  Biosphere  Reserve  of  Pantanos  de  Centla  and  near  the  ejido  of  Malpasito. 


municipio  (bordering  Veracruz  and  Chiapas), 
our  field  studies  were  conducted  from  14-22 
March  in  the  southwesternmost  corner  of  the 
state,  below  Cerro  La  Pava,  west  of  Ejido 
Malpasito,  and  approximately  5 km  west  of 
the  state’s  border  with  Chiapas  (ca  17°  20'  N, 
93°  36'  W,  300-500  m elevation;  Fig.  1).  This 
site  consisted  of  open  fields,  pastures,  small 
fruit  and  coffee  plantations,  acahual  (second 
growth  forest),  and  remnant  rainforest  patches 
restricted  to  mountain  crevices  and  aiToyos. 

In  addition  to  specimens  and  observations 
obtained  during  our  field  work,  we  include 
some  additional  specimen-based  information 
for  the  birds  of  Tabasco  from  the  national  or- 
nithological collection  of  Mexico  (Coleccion 
Nacional  dc  las  Aves — CNAV),  which  is 
housed  at  the  Instituto  de  Biologfa,  Departa- 


mento  de  Zoologia,  Universidad  Nacional  Au- 
tonoma de  Mexico,  in  Mexico  City.  Our  spec- 
imens representing  new  records  for  the  state 
were  also  deposited  in  CNAV.  A number  of 
records  we  report  as  new  for  the  state  were 
predicted  by  Benett  (1962).  Several  of  our  re- 
cords expand  the  known  distribution  of  par- 
ticular species  within  the  state;  many  of  these 
were  also  foreseen  by  Beirett  (1962). 

SPECIES  ACCOUNTS 

Broad-winged  Hawk  (Biiteo  platyptenis). — 
Berrett  ( 1962)  recorded  only  two  observations 
of  this  species  in  Tabasco,  both  of  migratory 
flocks  (28  March  1960,  17  April  1961).  What 
appears  to  be  the  first  state  specimen,  a female 
in  migration,  was  taken  at  Balneario  Agua 


Winker  el  al.  • TABASCO  AVIFAUNA 


231 


Blanca,  Macuspana,  on  13  March  1989 
(CNAV  13,523). 

Ruddy  Crake  (Laterallus  ruber). — BeiTCtt 
(1962)  noted  that  the  species  had  been  re- 
ported from  only  two  areas  in  the  state  (near 
Tenosique  and  Teapa),  but  that  it  was  probably 
more  widespread.  We  found  the  species  to  be 
very  common  at  Pantanos  de  Centla,  hearing 
its  distinctive  vocalizations  daily  during  our 
stay  in  the  aiea. 

Common  Snipe  (Gallinago  gallinago). — 
Berrett  ( 1 962)  reported  few  Tabasco  records 
of  this  wintering  migrant,  including  only  three 
sightings  of  his  own.  All  but  one  of  the  re- 
cords (from  Sanchez  Magallanes)  were  from 
southern,  inland  areas  of  the  state.  We  ob- 
served at  least  one  individual  per  day  at  Pan- 
tanos de  Centla  on  8-10  March. 

Sandwich  Tern  (Sterna  sandvicensis). — 
Berrett  (1962)  reported  collecting  the  first 
specimens  for  the  state  at  Sanchez  Magallanes 
and  Miramar;  his  observations  of  the  species 
did  not  extend  beyond  these  areas.  We  found 
this  tern  to  be  common  on  the  Rio  Grijalva 
from  Frontera  to  an  area  about  5 km  to  the 
south.  Two  to  approximately  40  individuals 
were  seen  almost  daily  during  our  stay  in  this 
area. 

Least  Tern  (Sterna  antillarum). — Berrett 
(1962)  reported  only  a single  specimen  from 
the  state  taken  by  D.  M.  Lay  on  2 May  1960 
near  Jonuta  on  the  Rio  Usumacinta.  Apart 
from  the  many  individuals  Lay  observed  near 
Jonuta  on  6 April  and  2 May  1960,  only  two 
other  observations  (both  of  single  individuals) 
were  reported  by  Berrett:  one  from  Chable  on 
the  Rio  Usumacinta  and  the  other  from  near 
Chontalpa  (Huimanguillo)  on  the  R(o  Grijal- 
va. These  are  all  inland  records.  On  12  March 
at  Pantanos  de  Centla  we  had  the  opportunity 
to  closely  observe  approximately  40  individ- 
uals feeding  over  the  Rio  Grijalva  near  Fron- 
tera in  loose  aggregation  with  individuals  of 
the  preceding  species. 

Mangrove  Cuckoo  (Coccyzus  minor). — 
Berrett  ( 1 962)  noted  few  Tabasco  records  and 
felt  that  the  species  was  restricted  to  coastal 
areas.  We  note  a male  taken  from  a decidedly 
inland  locality  6 km  S of  Huimanguillo  on  6 
March  1984  by  H.  Munoz  (CNAV  2,904). 

Common  Parauque  (Nyctidromus  albicol- 
lis). — Berrett  (1962)  noted  that  his  specimens 
of  this  permanent  resident  taken  between  1 1 


March  and  7 May  were  in  breeding  condition. 
An  individual  found  during  the  day  in  river- 
bank  mangrove/selva  forest  on  5 and  6 March 
acted  as  though  it  was  nesting  when  we 
flushed  it.  We  did  not  find  a nest,  but  on  7 
March  found  a single  egg  laid  on  the  leaf  litter 
in  one  of  our  mist  net  lanes.  We  captured  and 
banded  the  bird  when  it  returned  to  this  nest. 
A second  egg  was  laid  by  12  March  (our  next 
visit  to  the  site). 

Whip-poor-will  (Caprimulgus  vociferus). — 
New  record.  Berrett  (1962)  noted  only  a sin- 
gle report  of  this  species  from  Tabasco,  that 
of  Rovirosa  (1887).  Rovirosa’s  observations 
are  notoriously  suspect,  however;  Brodkorb 
(1943:8)  stated  that  “ . . . scarcely  any  reli- 
ance can  be  placed  on  his  records  which  have 
not  been  confirmed  by  other  workers.”  In  fact, 
Rovirosa  (1887)  simply  pooled  all  of  his  ca- 
primulgid  observations  under  this  single  spe- 
cific epithet,  apparently  not  knowing  what 
species  he  had  observed  and  perhaps  liking 
this  particular  name.  Under  “Antrostomus  vo- 
ciferus"  he  stated  that  various  species  of  ca- 
primulgid  occurred  in  the  state  from  January 
to  May.  Thus,  based  on  present  evidence  (i.e., 
Berrett  1962),  it  seems  likely  that  Rovirosa 
was  discussing  one  or  more  of  the  four  other 
caprimulgids  that  have  since  been  found  to  oc- 
cur in  the  state.  Given  its  commonness  and 
vocal  habits,  his  observations  probably  con- 
sisted mostly  of  Nyctidromus  albicollis,  but 
this  is  speculative,  and  does  not  explain  his 
lack  of  records  from  the  second  half  of  the 
year. 

KW  distinctly  heard  a Whip-poor-will  sing- 
ing at  Pantanos  de  Centla  on  9 Maixh.  The 
song  was  of  the  “Mexican”  subspecific  group 
(C.  V.  arizonae/oaxacae/chiapensis),  which 
Howell  and  Webb  (1995)  suggested  might  oc- 
cur in  lowlands  during  the  nonbreeding  sea- 
son. Subsequently,  we  discovered  that  a spec- 
imen of  this  subspecies  group  had  been  taken 
on  16  March  1984,  10  km  S of  Chontalpa, 
Municipio  de  Huimanguillo,  by  F.  Ornelas 
(CNAV  2,917). 

Vaux’s  Swift  (Chaetura  vaiixi). — New  Re- 
cord; no  specimen.  A flock  of  approximately 
15  individuals  was  seen  and  heard  at  close 
range  above  Malpasito,  Huimanguillo  on  15 
March,  and  a flock  of  35-40  was  seen  each 
day  16-18  March.  The  first  migrant  Chaetura 
pelagica  of  the  spring  (a  species  that  might  be 


232 


THE  WILSON  BULLETIN 


Vol.  Ill,  No.  2,  June  1999 


confused  with  C.  vaiixi)  were  seen  and  heard 
on  21  March. 

Violet  Sabrewing  (Campylopterus  hemileu- 
curus). — The  four  specimens  recorded  by 
Berrett  (1962)  for  the  state  all  came  from  near 
Teapa.  The  single  additional  sight  record  not- 
ed was  from  near  Chontalpa.  We  collected  two 
females  on  17  and  19  March  at  Malpasito. 
Neither  had  yet  entered  reproductive  condi- 
tion. 

Long-tailed  Sabrewing  (Campylopterus  ex- 
cellens). — New  Record.  An  adult  female  of 
this  species  was  taken  at  Malpasito  in  remnant 
forest  at  the  edge  of  a small  arroyo.  This  is 
the  first  record  of  this  species  for  the  state, 
and  also  the  northeastemmost  occurrence  of 
the  species,  which  until  recently  has  been  con- 
sidered to  have  a very  restiicted  range  (see 
Winker  et  al.  1992a,  Howell  and  Webb  1995). 
Howell  and  Webb  (1995)  based  their  exten- 
sion of  the  species’  range  into  Chiapas  on  a 
previously  misidentified  specimen  in  the 
American  Museum  of  Natural  History.  Reex- 
amination of  this  specimen  (KW)  suggests 
that  Howell  and  Webb  (1995)  were  correct;  it 
is  a Campylopterus  excellens,  and  not  a C. 
curvipennis  as  originally  identified.  Together, 
these  two  specimens  suggest  an  approximate 
doubling  of  the  species’  total  geographic 
range. 

Pygmy  Kingfisher  (Chloroceryle  aenea). — 
Hitherto,  the  only  records  for  this  species  in 
the  state  were  from  southern,  inland  localities 
(Balancan  and  Ocuapan:  Benett  1962;  Chon- 
talpa: CNAV  2,933).  We  collected  a non- 
breeding female  at  Pantanos  de  Centla  on  a 
small,  well-forested  tributary  (Arroyo  Polo)  of 
the  Rio  Grijalva  on  6 March. 

Wedge-billed  Woodcreeper  (Glyphorhyn- 
chus  spirurus). — Recorded  by  Beirett  (1962) 
as  an  uncommon  resident  of  the  rainforest  belt 
of  the  state,  records  were  lacking  from  the 
westernmost  rainforest  in  the  state  (Municipio 
de  Huimanguillo).  We  collected  a nonbreeding 
female  in  a patch  of  acahual  (second  growth 
forest)  at  the  base  of  the  local  mountains  at 
Malpasito  on  17  March. 

Ochre-bellied  Flycatcher  (Mionectes  olea- 
gineus). — Although  Berrett  (1962)  recorded 
this  species  near  Chontalpa,  he  did  not  find  it 
to  be  particularly  common.  We  found  it  to  be 
rather  common  at  Malpasito,  collecting  seven 
individuals  19-21  March.  Their  abundance  at 


the  site  was  associated  with  the  weather;  they 
were  undetected  before  a relatively  cold  norte 
(cold,  wet  weather  system  from  the  north)  ar- 
rived on  19  March. 

Sepia-capped  Flycatcher  (Leptopogon 
amaurocephalus). — A male  taken  on  20 
March  at  Malpasito  (testes  moderately  en- 
larged) extends  the  range  of  this  species  with- 
in the  state  to  the  westernmost  occurrence  of 
the  rainforest  belt. 

Sulphur-rumped  Flycatcher  (Myiobius  sul- 
phureipygius). — As  with  the  previous  species, 
Berrett  (1962)  documented  this  species  as  a 
permanent  resident  of  the  rainforest  belt,  but 
lacked  records  for  the  westernmost  rainforest 
in  the  state.  We  collected  a male  with  unen- 
larged testes  in  acahual  at  Malpasito  on  20 
March. 

White-throated  Flycatcher  (Empidonax  al- 
bigularis). — New  Record.  We  collected  what 
appear  to  be  the  first  specimens  for  the  state 
at  Pantanos  de  Centla  on  10  March  in  a shrub- 
by pasture.  Both  birds  were  males  with  no  fat 
and  a heavy  molt  that  included  body,  wings, 
and  tail. 

Great  Crested  Flycatcher  (Myiarchus  crin- 
itus). — Berrett  (1962)  noted  a single  specimen 
and  a few  sightings  of  this  species  in  Tabasco, 
all  from  May  1961  and  neai'  Balancan.  We 
note  two  other  specimens,  both  males,  taken 
on  13  Mai'ch  1984  and  29  September  1965 
(indicating  a presence  in  fall  migration  as  well 
as  spring),  near  Comalco  and  on  the  Rio  San 
Pablo,  Municipio  de  San  Pablo  by  F.  Ornelas 
and  R.  W.  Dickerman,  respectively  (CNAV 
3,491  and  17,215). 

Gray-collared  Becard  (Pachyramphus  ma- 
jor).— An  adult  male,  appai'ently  the  second 
specimen  for  the  state  (Benett  1962),  was  tak- 
en in  8-10  m acahual  on  20  March  at  Mal- 
pasito. The  first  specimen,  also  a male,  was 
taken  in  the  eastern  part  of  the  state  near  Re- 
forma, Balancan,  on  28  May  1939  (Brodkorb 
1943). 

Mangrove  Vireo  (Vireo  pallens). — BeiTett 
(1962)  collected  the  only  previous  specimen 
for  the  state  (male,  13  April  1961)  near  Villa- 
hermosa  and  noted  three  additional  sightings: 
two  near  Villahermosa  and  another  on  the  Rio 
Usumacinta  at  Emiliano  Zapata  (all  in  April 
1961 ).  Peterson  and  Chalif  ( 1973)  and  Howell 
and  Webb  (1995)  overlooked  these  records. 
We  collected  two  more  individuals  in  low- 


Winker  et  al.  • TABASCO  AVIFAUNA 


233 


lying,  shrubby  pasture  on  the  Rio  Grijalva  at 
Pantanos  de  Centla  on  9 and  10  March.  Both 
were  females  with  no  fat  and  unenlarged  ova- 
ries. At  least  one  more  individual  was  later 
seen  in  this  same  area. 

Blue-headed  Vireo  (Vireo  solitarius). — 
Berrett  (1962)  reported  only  two  observations 
of  this  species  in  the  state  (March  1959, 
March  1960),  both  near  Teapa.  We  collected 
what  are  apparently  the  first  two  state  speci- 
mens at  Malpasito  on  16  and  17  March,  a fe- 
male with  no  fat  and  a male  with  little  fat. 

Green  Jay  (Cyanocorax  yncas). — Berrett 
(1962)  recorded  specimens  and  sightings  of 
this  species  from  only  four  localities  in  the 
municipios  of  Balancan,  Centla,  and  Teapa. 
Our  sightings  of  two  individuals  each  on  15 
and  21  March  at  Malpasito  indicate  that  the 
species  also  occurs  in  the  western  part  of  the 
state. 

Tree  Swallow  (Tachycineta  bicolor). — Ber- 
rett (1962)  recorded  only  one  specimen  for  the 
state  but  noted  a few  additional  sightings,  in- 
cluding several  large  flocks.  Although  it  re- 
mains unclear  whether  the  species  spends  the 
winter  or  is  only  a transient  in  migration,  we 
add  the  following  records  from  Pantanos  de 
Centla:  three  specimens,  all  females  with  little 
fat,  taken  on  8 March  1996,  and  sightings  of 
1-150  individuals  almost  daily  from  4—12 
March.  In  addition,  a female  was  taken  by  R. 
W.  Dickerman  53  km  W of  Villahermosa  on 
3 April  1973  (CNAV  17,723). 

Plain  Wren  (Thryothorus  modestus). — New 
record.  An  adult  female  was  taken  at  Bal- 
neario  Agua  Blanca,  Macuspana,  in  selva  me- 
diana,  by  E.  Diaz  I.  on  14  March  1989  (CNAV 
13,389). 

Slate-colored  Solitaire  (Myadestes  unicoT 
or). — New  record.  We  collected  a male  on  20 
March  at  Malpasito  during  the  same  norte  as- 
sociated with  the  occurrence  of  the  Black 
Robin  (below)  in  the  area.  Movement  of 
Slate-colored  Solitaires  to  lowland  forest  dur- 
ing nortes  is  a frequent  occurrence  in  the  Si- 
erra de  Los  Tuxtlas  in  southern  Veracruz  (Ra- 
mos 1983;  KW,  pers.  obs.). 

Black  Robin  (Turdus  infuscatus). — New  re- 
cord; no  specimen.  A female  or  immature 
male  of  this  highland  species  was  seen  feeding 
in  a fruiting  tree  on  20  March  at  Malpasito  on 
the  last  day  of  a relatively  cold  norte.  It  had 
probably  descended  to  this  lower  elevation  (ca 


450  m)  to  escape  less  suitable  conditions  in 
the  highlands,  a common  occurrence  in  some 
areas  of  southern  Mexico  (see  Winker  et  al. 
1992b). 

White-throated  Robin  (Turdus  assimilis). — 
Berrett  ( 1 962)  noted  only  three  records  of  this 
species  in  Tabasco,  one  specimen  and  two 
sightings  from  near  Teapa  and  near  Chontalpa, 
both  southern  localities.  A male  was  taken  at 
Comalcalco  (in  the  northern  part  of  the  state) 
on  13  March  1984  by  F.  Ornelas  (CNAV 
6,776). 

Nashville  Warbler  (Vermivora  ruficapil- 
la). — New  record.  Berrett  (1962)  reported  a 
brief  glimpse  of  a bird  that  was  probably  this 
species  at  Balancan.  We  collected  the  first 
state  specimen,  a female  with  no  fat,  at  Mal- 
pasito on  20  March. 

Yellow-rumped  Warbler  (Dendroica  coron- 
ata). — Berrett  (1962)  reported  only  a few  rec- 
ords for  this  species,  all  from  December.  We- 
ber (1945)  reported  it  at  La  Venta  in  spring 
1943.  We  took  three  females  with  moderate  to 
heavy  fat  in  shrubby  pasture  at  Pantanos  de 
Centla  on  9-11  March,  and  observed  up  to 
three  other  individuals  on  these  days. 

Swainson’s  Warbler  (Limnothlypis  swain- 
sonii). — New  record.  The  first  state  specimen, 
a male  with  no  fat  and  with  tail  characteristics 
(uneven  growth  bars)  of  an  after  second  year 
(ASY)  individual,  was  taken  in  a mixed  forest 
of  mangrove  and  selva  near  the  bank  of  Ar- 
royo Polo  (Rio  Grijalva)  near  Frontera  on  6 
March.  This  bird  was  not  in  migratory  con- 
dition and  the  habitat  it  occupied  was  emi- 
nently suitable  for  wintering  (KW,  pers.  obs.; 
see  Graves  1998). 

White-winged  Tanager  (Piranga  leucop- 
tera). — Berrett  (1962)  reported  a few  individ- 
uals of  this  species  in  early  and  mid-Novem- 
ber 1961  near  Tenosique  and  surmised  that 
they  may  have  been  wintering  birds  from 
higher  elevations.  We  saw  a single  female  at 
Malpasito  on  19  March,  during  the  same  norte 
associated  with  the  presence  of  the  Black 
Robin  and  Slate-colored  Solitaire  noted 
above. 

Orange-billed  Sparrow  (Arremon  aurantii- 
rostris). — Benett  (1962:375)  noted  six  speci- 
mens from  the  state  and  referred  to  the  species 
as  a “rarely  recorded  permanent  resident  of 
the  humid  rain  forest  undergrowth.”  Although 
we  worked  intensively  in  the  Malpasito  ai'ea 


234 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


for  four  days  before  encountering  any  individ- 
uals of  this  species,  during  the  relatively  se- 
vere norte  of  19  and  20  March  we  collected 
four  individuals  (three  females,  one  male)  in 
our  mist  nets  that  had  been  in  the  same  lo- 
cations the  entire  time  in  a small  patch  of  rem- 
nant rainforest  at  the  base  of  the  local  moun- 
tains. This  striking  movement  pattern  of  in- 
creased captures  during  nortes  is  typical  of 
highland  species  at  lowland  sites,  but  not  of 
lowland  species.  We  surmise  that  at  this  site 
these  typically  lower-elevation  birds  might  oc- 
cupy less  disturbed  forest  at  slightly  higher 
elevations  than  the  remnant  forest  available 
where  our  field  work  was  conducted. 

Olive  Sparrow  (Arremonops  riifivirga- 
tus). — Benett  (1962)  reported  only  a single 
specimen,  but  noted  that  the  species  was  a 
common  permanent  resident  north  of  Balan- 
can.  These  records  were  overlooked  by  How- 
ell and  Webb  (1995).  We  found  that  the  spe- 
cies also  occurs  in  the  westernmost  region  of 
the  state.  We  collected  a female  with  an  un- 
enlarged ovary  on  18  March  at  Malpasito  and, 
in  addition,  found  what  appeared  to  be  two 
pairs  (two  singing  males  with  associated  in- 
dividuals not  singing)  in  a brushy  area  of 
young  second  growth. 

Savannah  Sparrow  (Passerculus  sandwich- 
ensis). — Four  individuals  were  seen  on  14 
March  beside  scrubby  pasture  at  Pantanos  de 
Centla  on  the  Rio  Grijalva,  supplementing  the 
single  specimen  and  several  sightings  reported 
by  Berrett  (1962)  from  a southern  part  of  the 
state  (Huastecas,  Teapa,  ca  29  km  NE  of  Tea- 
pa). 

Lincoln’s  Sparrow  (Melospiza  lincolnii). — 
Berrett  (1962)  included  this  species  as  an  un- 
common wintering  bird  in  the  state,  reporting 
only  a single  specimen  and  scattered  sight  re- 
cords from  largely  southern,  inland  sites.  Our 
records  add  northern,  near-coastal  records  for 
the  state.  We  observed  a single  individual  on 
4 March  at  Pantanos  de  Centla  in  roadside 
scrub,  and  collected  a male  with  light  fat  on 
1 1 March  in  shrubby  pasture  on  the  Rfo  Gri- 
jalva. 

DISCUSSION 

Our  records  consist  of  10  migrant  and  25 
resident  species.  Of  our  nine  new  records  for 
the  state,  only  two  are  of  migrant  species;  sev- 
en are  resident.  This  is  unusual  for  an  area  in 


the  northern  Neotropics,  where  at  the  end  of 
the  twentieth  century  one  would  expect  our 
knowledge  of  the  avifauna  to  be  sufficiently 
strong  that  most  new  records  would  be  gen- 
erated by  individuals  of  migrant  species.  We 
consider  the  preponderance  of  resident  species 
in  this  report  to  be  a reflection  of  Tabasco’s 
status  as  a poorly  known  Neotropical  region. 
Further  ornithological  study  of  this  region  is 
fully  warranted.  Tabasco  is  nearly  half  the  size 
of  Costa  Rica.  It  occupies  a geographic  posi- 
tion making  it  important  as  a wintering  ground 
for  Nearctic-Neotropic  migrants  and  as  a 
breeding  and  nonbreeding  area  for  a diverse 
Neotropical  resident  avifauna.  Further,  it  has 
twofold  biogeographic  significance:  one  as  a 
transition  zone  between  the  Yucatan  Peninsula 
and  the  Isthmus  of  Tehuantepec  (including  a 
significant  portion  of  the  rainforest  belt  at  the 
peninsula’s  base)  and  another  in  its  position  at 
the  northern  limits  of  Neotropical  families 
such  as  Heliomithidae,  Eurypygidae,  Galbu- 
lidae,  Bucconidae,  and  Pipridae. 

Our  data  provide  important  new  distribu- 
tional information  for  both  resident  and  mi- 
grant species  in  this  poorly  known  region.  For 
example,  our  records  of  Swainson’s  Warbler 
and  the  Long-tailed  Sabrewing  constitute  sig- 
nificant wintering  and  probable  breeding 
range  extensions  (respectively)  for  rare  spe- 
cies of  conservation  concern  (see  Howell  and 
Webb  1995,  Winker  et  al.  1992a). 

In  addition,  our  data  show  inteimittent  use 
of  lowland  forest,  indicating  movements 
among  “resident”  birds.  The  relatively  low- 
land sites  we  studied  near  Malpasito  suddenly 
became  home  to  individuals  of  a number  of 
species  that  were  not  detected  during  the  four 
days  of  intensive  field  study  prior  to  the  arriv- 
al of  a norte  on  19  March.  At  least  five  species 
(including  two  new  to  the  state)  showed 
movement  patterns  that  were  strongly  tied  to 
the  inclement  weather:  Mionectes  oleagineus, 
Turdus  infuscatus,  Myadestes  unicolor,  Pir- 
anga  leucoptera,  and  Arrenion  aurantiirostris. 
All  of  these  species  probably  occupy  forest  at 
higher  elevations  in  this  area.  An  increase  in 
overall  capture  rates  and  census  detections 
(unpubl.  data)  suggested  increased  numbers  of 
individuals  of  many  species  already  present  at 
the  lowland  sites.  These  individuals  may  also 
have  come  from  the  local  highlands. 

Temporary  use  of  remnant  lowland  forest 


Winker  et  al.  • TABASCO  AVIFAUNA 


235 


during  nortes  is  a relatively  common  phenom- 
enon in  southern  Veracruz  (Ramos  1983, 
Winker  et  al.  1997)  and  is  an  issue  of  grave 
conservation  concern;  when  lowland  forests 
have  diminished  to  a degree  of  scarcity,  they 
are  no  longer  available  to  individuals  seeking 
temporary  refuge  from  inclement  conditions 
in  the  highlands. 

ACKNOWLEDGMENTS 

We  thank  the  U.S.  National  Science  Foundation 
(NSF  INT-9403053)  and  CONACyT  (El 20)  for  sup- 
porting our  field  studies,  and  SEMARNAP  for  issuing 
the  necessary  permits.  J.  M.  Arias  R.,  C.  A.  Cordero 
M.,  O.  E.  Escobar  R,  J.  L.  Gonzalez  A.,  A.  C.  IbaiTa 
M.,  G.  Lopez  S.,  M.  de  J.  Mendez  G.,  L.  Montanez 
G.,  M.  Ramirez  L.,  R.  E.  Sobrino  R,  and  M.  Suarez  I. 
proved  to  be  excellent  field  companions  and  contrib- 
uted to  a stimulating  workshop  and  field  experience. 
We  also  thank  J.  C.  Romero,  C.  A.  Jimenez  B.,  P.  del 
Valle,  R.  del  Valle  Reyna,  and  the  staffs  of  the  Pan- 
tanos  de  Centla  Field  Station  and  Aguaima  for  their 
help  and  hospitality  during  our  visits.  F.  Gonzalez  G. 
and  J.  H.  Rappole  provided  helpful  comments  on  an 
earlier  draft. 

LITERATURE  CITED 

Berrett,  D.  G.  1962.  The  birds  of  the  Mexican  state 
of  Tabasco.  Ph.D.  diss.,  Louisiana  State  Univ.,  Ba- 
ton Rouge. 

Brodkorb,  P.  1943.  Birds  from  the  lowlands  of  south- 
ern Mexico.  Misc.  Publ.  Mus.  Zool.  Univ.  Mich. 
55:1-88. 

Centeno  A.,  B.  E.  1994.  Estado  actual  del  conoci- 


miento  de  la  avifauna  de  Tabasco:  revision  biblio- 
grafica.  Tesis  Licenciado  en  Biologia,  Univ.  Jua- 
rez Autonoma  de  Tabasco,  Villahermosa. 

Graves,  G.  R.  1998.  Stereotyped  foraging  behavior  of 
the  Swainson’s  Warbler.  J.  Field  Ornithol.  69: 121  — 
127. 

Howell,  S.  N.  G.  and  S.  Webb.  1995.  A guide  to  the 
birds  of  Mexico  and  northern  Central  America. 
Oxford  Univ.  Press,  New  York. 

Peterson,  R.  T.  and  E.  L.  Chalie.  1973.  A field  guide 
to  Mexican  birds.  Houghton  Mifflin  Co.,  Boston, 
Massachusetts. 

Ramos,  M.  A.  1983.  Seasonal  movements  of  bird  pop- 
ulations at  a neotropical  study  site  in  southern  Ve- 
racruz, Mexico.  Ph.D.  diss.,  Univ.  of  Minnesota, 
Minneapolis. 

Rovirosa,  j.  N.  1887.  Apuntes  para  la  zoologia  de 
Tabasco:  vertebrados  observados  en  el  Territorio 
de  Macuspana.  Naturaleza  7:345-389. 

Weber,  W.  A.  1945.  Wildlife  of  Tabasco  and  Veracruz. 
Nat.  Geog.  87(2):  187-2 16. 

Winker,  K.,  M.  A.  Ramos,  J.  H.  Rappole,  and  D.  W. 
Warner.  1992a.  A note  on  Campylopterus  excel- 
lens  in  southern  Veracruz,  with  a guide  to  sexing 
captured  individuals.  J.  Field  Ornithol.  62:339- 
343. 

Winker,  K.,  R.  J.  Oehlenschlager,  M.  A.  Ramos,  R. 
M.  Zink,  J.  H.  Rappole,  and  D.  W.  Warner. 
1992b.  Bird  distribution  and  abundance  records 
for  the  Sierra  de  Los  Tuxtlas,  Veracruz,  Mexico. 
Wilson  Bull.  104:699-718. 

Winker,  K.,  P.  Escalante,  J.  H.  Rappole,  M.  A.  Ra- 
mos, R.  J.  Oehlenschlager,  and  D.  W.  Warner. 
1997.  The  evolution  and  conservation  of  Wet- 
more’s  Bush-Tanager:  periodic  migration  and  low- 
land forest  refugia  in  a “sedentary”  neotropical 
bird.  Conserv.  Biol.  11:692-697. 


Wilson  Bull.,  111(2),  1999,  pp.  236-242 


PREDATION  OF  SMALL  EGGS  IN  ARTIFICIAL  NESTS:  EFFECTS 
OF  NEST  POSITION,  EDGE,  AND  POTENTIAL  PREDATOR 
ABUNDANCE  IN  EXTENSIVE  FOREST 

RICHARD  M.  DEGRAAF,'^^  THOMAS  J.  MAIER,'  AND  TODD  K.  FULLERS 


ABSTRACT. — Alter  photographic  observations  in  the  field  and  laboratory  tests  indicated  that  small  rodents 
might  be  significant  predators  on  small  eggs,  we  conducted  a field  study  in  central  Massachusetts  to  compare 
predation  ot  House  Sparrow  (Passer  domesticus)  eggs  in  artificial  nests  near  to  (5-15  m)  and  far  from  (100- 
120  m)  forest  edges  and  between  ground  and  shrub  nests.  As  in  earlier  studies  in  managed  northeastern  forest 
landscapes  that  used  larger  quail  eggs,  predation  rates  on  small  eggs  in  nests  at  the  forest  edge  did  not  differ 
(P  > 0.05)  from  those  in  the  forest  interior  for  either  ground  nests  (edge  = 0.80  vs  interior  = 0.90)  or  shrub 
nests  (edge  = 0.38  vs  interior  = 0.28)  after  12  days  of  exposure.  However,  predation  rates  on  eggs  in  ground 
nests  were  significantly  higher  (P  < 0.001)  than  in  shrub  nests  at  both  the  edge  and  interior.  There  were  no 
significant  (P  > 0.05)  differences  in  the  frequency  of  capture  of  the  6 most  common  small  mammal  species 
between  forest  edge  and  interior.  Logistic  regression  analyses  indicated  a highly  significant  (P  < 0.001)  nest 
placement  effect  but  very  little  location  or  small  mammal  effect.  Predation  of  small  eggs  by  small-mouthed 
ground  predators  such  as  white-footed  mice  (Peromyscus  leiicopns)  has  not  been  documented  as  a major  factor 
in  egg  predation  studies,  but  use  of  appropriately-sized  eggs  and  quantification  of  predator  species  presence  and 
abundance  seems  essential  to  future  studies.  Received  31  March  1998,  accepted  5 Jan.  1999. 


Previously  published  evidence  for  elevated 
nest  predation  rates  at  forest  edges  in  the 
northeastern  U.S.  is  not  consistent.  For  ex- 
ample, in  Maine,  predation  rates  were  higher 
for  artificial  nests  placed  in  shrubs  at  edges 
than  in  forest  interiors,  but  the  distance  to 
edge  had  no  effect  on  predation  of  ground 
nests  (Rudnicky  and  Hunter  1993).  Also,  nei- 
ther the  edge: area  ratio  of  forest  patches  nor 
the  distance  from  edge  affected  artificial 
ground  nest  predation  rates  (Small  and  Hunter 
1988).  Predation  rates  of  artificial  nests  were 
higher  in  extensive  industrial  forests  than  in 
fragments,  but  within  fragments,  shrub  nests 
near  edges  were  depredated  at  a higher  rate 
than  those  farther  from  edges.  Furthermore, 
the  predation  rate  in  clearcuts  was  lower  than 
that  in  forest  fragments  or  plantations,  and 
within  plantations,  predation  rates  increased 
with  increasing  distance  from  the  edge  (Van- 
der  Haegen  and  DeGraaf  1996).  In  Pennsyl- 
vania, Yahner  and  Scott  ( 1988)  reported  a di- 
rect relationship  between  amount  of  forest 


' USDA  F'orest  Service,  Northeastern  Research  Sta- 
tion, University  of  Massachusetts,  Amherst,  MA 
01 003  USA. 

’ Department  of  Forestry  and  Wildlife  Management 
and  Graduate  Program  in  Organismic  and  Evolution- 
ary F3iology,  University  of  Massachusetts,  Amherst, 
MA  01003-4210. 

’ Corresponding  author;  E-mail: 
rdegraaf@forwild.umass.edu 


fragmentation  caused  by  clearcutting  and  pre- 
dation rates  on  artificial  nests,  yet  Yahner  and 
CO  workers  (1993)  did  not  find  greater  preda- 
tion rates  on  such  nests  despite  greater  frag- 
mentation resulting  from  additional  cleai'cut- 
ting  on  the  same  study  area.  In  sum,  the  re- 
sults of  previous  studies  in  the  northeastern 
U.S.  are  inconsistent,  perhaps  because  the 
large  quail  (Coturnix  sp.)  or  chicken  {Gallus 
sp.)  eggs  used  do  not  sample  the  entire  pred- 
ator community  (Haskell  1995).  Would  the 
use  of  eggs  of  approximately  the  same  size  as 
most  forest  passerines  shed  light  on  patterns 
of  predation  on  artificial  nests  in  relation  to 
forest  edge? 

White-footed  mice  {Peromyscus  leucopus) 
were  frequently  recorded  by  remotely-trig- 
gered cameras  at  ground  and  shrub  nests  con- 
taining eggs  of  Japanese  Quail  {Coturnix  ja- 
ponica)-,  many  of  these  same  nests  appeared 
to  be  undisturbed  at  the  end  of  the  exposure 
period  and  thus  were  not  classified  as  visited 
by  predators  (Danielson  et  al.  1997).  Similar- 
ly, Northern  Bobwhite  {Colinus  9irginianus) 
eggs  at  ai'tificial  nests  in  Minnesota  that  were 
visited  (as  determined  by  photographs)  by 
red-backed  voles  {Clethrionomys  gapperi)  and 
deer  mice  {Peromyscus  maniculatus)  were  not 
damaged,  although  those  nests  were  classified 
as  depredated  by  Fenske-Crawford  and  Niemi 
(1997).  Small  mammalian  predators  are  clear- 


236 


DeGniafet  at.  • PREDATION  ON  SMAl.L  EGGS  IN  ARTIFICIAE  NESTS 


237 


ly  able  to  locate  artificial  nests,  but  have  lim- 
ited ability  to  destroy  quail  eggs  in  these  nests. 
Quail  eggs  are  not  representative  of  the  sizes 
of  eggs  of  most  temperate  forest  passerines, 
especially  those  of  Neotropical  migratory  spe- 
cies (Haskell  1995,  DeGraaf  and  Maier  1996). 
If  appropriately-sized  eggs  were  not  available 
to  potentially  common  predators,  then  results 
of  previous  studies  to  estimate  nest  predation 
rates  for  forest  songbirds  may  have  been  bi- 
ased, contributing  to  the  inconsistency  of  re- 
sults in  extensive  northeastern  forests. 

Egg  size  is  potentially  important  in  nest 
predation  studies;  even  though  small  rodents 
such  as  mice  and  eastern  chipmunks  (Tamias 
striatus)  may  be  egg  predators  (e.g.,  Maxon 
and  Oring  1978,  Reitsma  et  al.  1990,  respec- 
tively), they  apparently  cannot  readily  open 
and  consume  the  larger  eggs  of  quails  and 
chickens  (Roper  1992,  Haskell  1995;  but  see 
Craig  1998).  Roper  (1992)  showed  that  pred- 
ators did  not  respond  to  quail  eggs  as  they  did 
to  native  birds’  eggs  in  Panama  because  most 
mammalian  nest  predators  were  too  small  to 
eat  quail  eggs.  Such  eggs,  however,  are  vir- 
tually the  only  ones  that  have  been  used  in 
artificial  nest  predation  studies  (Major  and 
Kendal  1996). 

These  facts  led  us  to  conduct  a laboratory 
experiment  of  mouse  predation  on  large  (C. 
japonica)  and  small  (Zebra  Finch,  Taeniopy- 
gia  guttata)  eggs  (DeGraaf  and  Maier  1996). 
Mouse  predation  on  small  eggs  was  immedi- 
ate but  did  not  occur  on  the  large  eggs.  Sim- 
ilar laboratory  trials  (Maier  and  DeGraaf,  un- 
publ.  data)  indicated  that  white-footed  mice, 
including  juveniles,  could  open  House  Spar- 
row (Passer  domesticus)  eggs;  we  conducted 
a field  study  to  evaluate  egg  predation  in  ar- 
tificial nests  containing  such  eggs. 

We  attempted  to  assess  the  effects  of  nest 
location  (edge  vs  interior),  placement  (shioib 
vs  ground),  and  the  relative  abundance  of 
small  mammals  on  the  predation  of  small 
eggs.  We  hypothesized  that  small  mammals 
were  equally  abundant  at  edges  and  in  forest 
interiors  (Heske  1995),  that  no  edge-related 
differences  in  nest  predation  would  be  found 
for  either  ground  or  shrub  nests  (Major  and 
Kendall  1996),  and  that  predation  would  be 
greater  on  ground  nests  than  on  shmb  nests 
because  small  mammals  such  as  mice  and 
chipmunks  spend  the  majority  of  their  forag- 


ing time  on  the  ground  (Madison  1977,  Elliot 
1978,  Graves  et  al.  1988). 

METHODS 

We  placed  artilicial  nests  near  (5-15  ni)  and  far 
( 100-120  m)  from  stand  edges  in  40  mature  stands  in 
an  extensive  managed  mixed-wood  forest  in  central 
Massachusetts  during  June  to  15  July  1997.  All  stands 
were  at  least  80  years  old  and  of  the  red  oak  ( Quercus 
n/^ra)-white  pine  (Finns  strobus)-re.d  maple  (Acer 
ritbrum)  forest-cover  type  (Eyre  1980);  edges  were 
formed  by  small  (2-4  ha)  clearcuts  1-6  years  old.  We 
placed  two  ground  nests  and  two  shrub  nests  in  each 
stand,  one  of  each  type  near  and  far  from  the  edge  and 
at  least  100  m from  each  other  (Fig.  1).  Nests  (160 
total)  were  wicker  baskets  10  cm  in  diameter  and  6 
cm  deep,  weathered  for  3 weeks  before  use,  and  con- 
tained one  fresh  House  Sparrow  egg.  To  minimize  hu- 
man scent  at  nests,  we  wore  rubber  boots  and  clean 
cotton  gloves  during  nest  placement  (Whelan  et  al. 
1994).  Ground  nests  were  set  into  the  surface  litter; 
shrub  nests  were  wired  1-1.5  m above  the  ground  in 
crotches  or  forks  of  branches  of  shrubs  or  small  sap- 
lings. All  nests  were  checked  after  12  days,  approxi- 
mately the  mean  incubation  time  for  small  forest  pas- 
serines. Eggs  found  out  of  the  nest,  destroyed  in  the 
nest,  or  missing  were  classified  as  predations. 

We  analyzed  the  nest  predation  data  as  paired-sam- 
ple nest  types  within  stands  (Zar  1996:163)  and  per- 
formed statistical  tests  using  SYSTAT  7.0  for  Win- 
dows. Sign  tests  were  used  to  detect  differences  in  the 
number  of  nest  predations  among  edge  and  interior 
nests  on  the  ground  and  in  shrubs  (Zar  1996:536). 

We  assessed  the  relative  abundance  of  small  mam- 
mal species  at  edge  and  interior  sites  using  3-day  re- 
moval trapping  (Miller  and  Getz  1977)  at  each  site 
immediately  after  the  nest  predation  experiments.  Two 
circular  trapping  arrays  (20  traps/20  m diameter  array) 
were  set  in  each  of  the  40  stands,  one  midway  between 
edge  nests  and  one  between  interior  nests  (Fig.  1 ).  Dif- 
ferent types  of  small  mammal  traps  are  more  efficient 
for  trapping  certain  species  under  varying  conditions, 
e.g.,  weather  (Williams  and  Braun  1983,  Bury  and 
Corn  1987,  Mengak  and  Guynn  1987);  we  used  four 
types  of  traps  in  an  attempt  to  more  completely  sample 
the  small  mammal  community  (Pelikan  et  al.  1977). 
Small  Victor  snap  traps  with  expanded  pedals.  Muse- 
um Special  snap  traps  with  expanded  pedals,  large 
(approx.  8 X 8 X 24  cm)  Sherman  traps,  and  modified 
large  Sherman  live  traps  with  circular  glass  windows 
(5.5  cm  diameter)  in  the  rear  door  were  used  at  each 
array  in  equal  numbers.  All  traps  were  baited  with  a 
mixture  of  peanut  butter,  oatmeal,  bacon,  and  black 
sunflower  seed  and  were  checked  daily.  All  small 
mammals  collected  were  deposited  in  the  Vertebrate 
Museum  of  the  University  of  Massachusetts,  Amherst, 
Massachusetts.  We  followed  the  guidelines  for  the  cap- 
ture and  handling  of  mammals  approved  by  the  Amer- 
ican Society  of  Mammalogists  (American  Society  of 
Mammalogists  1998). 


238 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2.  June  1999 


Clearcut  g 

(2  - 4 ha)  m 


Edge  ground  nest 


bd 


a c 


bd 


a c 


a c 
b d 


a c 


b d 
a c 
b d 


10  m 


Edge  shrub  nest 


Forest 


100-  120  m 
\ 


Interior  shrub  nest 


a c 

bd  bd 


- bd 


b d b d 

a c 


20  m dia 


Interior  ground  nest 


IV 


FIG.  1.  Placement  of  small  mammal  trapping  arrays  and  artificial  nests  in  central  Massachusetts,  June-August. 
1997.  Four  types  of  traps  include:  (a)  small  Victor  snap,  (b)  Museum  Special,  (c)  large  Sherman,  (d)  large 
Sherman  with  glass  window.  (Figure  not  to  scale.) 


Small  mammal  capture  counts  were  compared  by 
species  between  edge  and  interior  using  the  Wilcoxon 
paired-sample  test  for  species  with  sufficient  n (Zar 
1996:167).  Logistic  regression  (PROC  LOGISTIC, 
SAS  1989)  was  used  to  assess  the  relationship  between 
small  mammal  counts  and  nest  predation;  we  used  a 
model  with  nest  placement  (ground  or  shrub),  location 
(edge  or  interior),  and  small  mammal  abundance  ef- 
fects; the  first  two  were  treated  as  categorical  variables 
and  the  third  as  a continuous  variable.  The  tests  for 
whether  a coefficient  is  zero  were  carried  out  using  Z 
= (estimated  coefficient/standard  error)  with  the  P-val- 
ue  obtained  using  the  standard  normal  distribution 
(Hosmerand  Lemeshow  1989:17). 

RESULTS 

We  did  not  detect  any  significant  differenc- 
es between  the  number  of  nest  predations  at 
the  forest  edge  and  those  in  the  forest  interior 
for  either  ground  nests  (Sign  test:  ties  = 28, 
4 “-b”,  8 critical  value  = 2,  P > 0.05) 
or  shrub  nests  (Sign  test:  ties  = 24,  10  “+”, 
6 “ — critical  value  = 3,  P > 0.05).  How- 
ever, the  number  of  nest  predations  on  ground 
nests  at  both  the  edge  (Sign  test:  ties  = 15, 
21  + 4 critical  value  = 7,  P < 0.001) 

and  interior  (Sign  test:  ties  = 15,  25  “+”,  0 
critical  value  = 1,  P < 0.001)  were  sig- 
nificantly higher  than  those  on  shrub  nests. 


Twelve  species  of  small  mammals  were  de- 
tected; six  species  represented  99%  of  cap- 
tures at  both  forest  edge  and  interior.  The  dis- 
tributions of  the  6 most  commonly  detected 
small  mammal  species  did  not  differ  signifi- 
cantly (Wilcoxon  paired-sample  tests:  P > 
0.05)  between  stand  edges  and  interiors  (Table 
1).  White-footed  mice  were  detected  more 
than  all  other  species  combined  in  both  stand 
edges  and  inteiiors  and  were  the  only  small 
mammal  species  detected  in  all  40  stands.  Lo- 
gistic regression  analyses  confirmed  nest 
placement  (ground,  shrub)  effects  but  showed 
no  effect  of  small  mammal  abundance  or  lo- 
cation (edge,  interior)  on  nest  predation  rate 
(Table  2). 

DISCUSSION 

Because  this  is  the  first  study  that  we  know 
of  to  systematically  evaluate  artificial  nest 
predation  in  relation  to  forest  edge  using  small 
eggs,  comparison  with  other  studies  where 
larger  eggs  were  used  is  difficult.  In  a recent 
review  of  studies  in  both  agricultural  and  set- 
tled landscapes  in  North  America  and  Europe, 
Major  and  Kendal  (1996)  showed  that  egg 
predation  (on  large  eggs)  was  higher  near  the 


DeGrcuifei  al.  • PREDATION  ON  SMALL  EGGS  IN  ARTILICIAL  NESTS 


239 


TABLE  1.  Numbers  of  small  mammals  captured  near  edges  and  interiors  of  40  stands  in  extensive  forest 
in  central  Massachusetts,  July  and  August  1997.  Wilcoxon  paired-sample  results  for  most  commonly  detected 
species;  N = stands  species  detected  in,  n = differences  (N  minus  ties),  T = smallest  sum  of  ranks,  Tuo-iij,,,,  = 
critical  value  (Zar  1996:  table  B.12), 


Species^ 

Edge 

Inlerior 

Total 

N 

n 

T 

p 

White-footed  mouse  (PeroiHyscus  teiicopus) 

251 

235 

486 

40 

34 

262.5 

182 

>0.05 

Red-backed  vole  (Clethriononiys  gapperi) 

74 

63 

137 

25 

21 

90.5 

58 

>0.05 

Northern  short-tailed  shrew  (Blarina  brevicaiida) 

61 

65 

126 

35 

28 

202.0 

1 16 

>0.05 

Masked  shrew  (Sorex  cinereus) 

16 

22 

38 

17 

15 

44.0 

25 

>0.05 

Eastern  chipmunk  (Tamias  striatiis) 

14 

5 

19 

15 

15 

28.5 

25 

>0.05 

Smoky  shrew  (Sorex  fiimeus) 

7 

3 

10 

9 

9 

13.5 

5 

>0.05 

^ Species  delected  in  ^3  siands:  woodland  jumping  mouse  {Napcteozcipus  insifinis),  flying  squirrel  {Glaucontys  sp.).  long-tailed  weasel  {Musiela  frenata), 
red  squirrel  {Tamiasciurus  huJsonicus),  pine  vole  {Microtus  pineforum),  meadow  vole  {Microius  pennsylvcmicus). 


forest  edge  in  three  studies,  higher  away  from 
the  edge  in  one  study,  and  equal  in  seven  stud- 
ies, Predation  of  artificial  nests  containing 
Northern  Bobwhite  eggs  in  Wisconsin  pine 
barrens  savannah  patches  was  coirelated  with 
proximity  to  the  edge  (Niesmuth  and  Boyce 
1997),  Predation  on  artificial  nests  containing 
small  chicken,  Japanese  Quail,  and  plasticine 
eggs  in  Alberta  was  highest  in  larger  woodlots 
and  showed  no  edge  effect  (Hannon  and  Cot- 
terill  1998).  Two  additional  studies  (which 
used  Japanese  Quail  eggs)  in  the  northeastern 
U.S.  did  not  detect  any  difference  in  predation 
rates  between  edges  and  interiors  of  exten- 
sive-managed forests  (Table  3;  Vander  Haegan 
and  DeGraaf  1996)  or  suburban/agricultural/ 
forest  landscapes  (Danielson  et  al.  1997). 
Along  with  our  current  results,  these  vaiiable 
findings  suggest  either  that  the  “edge”  effect 
as  related  to  egg  predation  (Andren  and  An- 
gelstam  1988)  is  not  a widespread  phenome- 


TABLE  2.  Parameter  estimates  and  statistics  from 
logistic  regression  analysis  of  placement  (ground/ 
shrub),  location  (edge/interior),  and  small  mammal 
abundance  in  relation  to  predation  of  small  eggs  in 
artificial  nests  in  central  Massachusetts,  1997.  The  tests 
for  whether  a coefficient  is  zero  were  carried  out  using 
Z = (est.  coefficient/SE)  with  the  P-value  obtained  us- 
ing the  standard  normal  distribution  (equivalent  to  C 
= Zr  with  the  P-value  based  on  X'  distribution  with  I 
df). 


Variable 

Estimated 

coefficient 

Standard 

error 

p 

INTERCEPT 

-1.7320 

0.5080 

0.0007 

LOCATION 

0.0002 

0.3805 

0.9996 

W-F  MICE 

-0.0005 

0.0602 

0.9940 

PLACEMENT 

2.4655 

0.3937 

0.0001 

non  or  that  not  all  forest  edges  are  the  same; 
i.e.,  forest-clearcut,  forest-agriculture,  and  for- 
est-suburb edges  differ  in  the  predators  pre- 
sent (Danielson  et  al.  1997). 

Equally  variable  are  the  results  of  nest 
placement  studies  (i.e.,  ground  vs  shrub/ele- 
vated nests).  Major  and  Kendal  (1996)  re- 
ported higher  predation  at  elevated  nests  in  six 
studies,  higher  predation  at  ground  nests  in 
four  studies,  and  equal  predation  rates  in  three 
studies.  Ground  nests  containing  Japanese 
Quail  and  plasticine  eggs  had  increased  pre- 
dation along  farm  edge  and  interior  in  Sas- 
katchewan, but  there  were  no  detectable  dif- 
ferences in  predation  rate  between  ground  and 
shrub  nests  at  logged  edge,  logged  interior,  or 
contiguous  forest  (Bayne  and  Hobson  1997). 
Although  two  studies  in  the  noitheastem  U.S. 
did  not  detect  any  difference  in  predation  rates 
between  ground  and  shrub  nests  (Vander  Hae- 
gan and  DeGraaf  1996,  Danielson  et  al.  1997), 
we  found  a strong  placement  effect  (high  pre- 
dation on  ground  nests)  using  small  eggs. 

Where  edge  or  nest  placement  effects  oc- 
cuiTed,  generalist  predators  commonly  were 
presumed  to  depredate  specific  nest  types  dis- 
proportionately. The  variability  in  results 
among  studies  may  reflect  differences  in  nest 
predator  guilds  or  the  abundance  of  particular 
species  in  study  areas  (e.g..  Pieman  1988).  At- 
tempts to  identify  individual  egg  predators  in- 
clude characterizations  of  predation  remains 
of  real  eggs  (Gottfried  and  Thompson  1978, 
but  see  Marini  and  Melo  1998),  impressions 
in  plasticine  (Bayne  et  al.  1997),  and  clay 
eggs  (Donovan  et  al.  1997),  hair  catchers 
(Baker  1980),  and  remotely  triggered  cameras 
(DeGraaf  1995).  Nevertheless,  egg  predation 


240 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


TABLE  3.  Predation  rates  (%)  on  eggs  in  artificial  nests  exposed  for  12—14  days  at  forest  edge  and  interior 
(>50  m)  in  the  northeastern  U.S. 


Nest 

placement 

Egg 

size 

Edge 

Interior 

Study 

location 

Reference 

Rate 

n 

Rate 

n 

p 

Ground 

Sparrow 

0.80 

40 

0.90 

40 

>0.05 

Massachusetts 

This  study 

Quail 

0.45 

20 

0.41 

80 

>0.05 

Massachusetts 

DeGraaf,  unpubl.  data“ 

0.20 

50 

0.25 

48 

>0.05 

Maine 

Vander  Haegan  and  DeGraaf  1996*’ 

0.29 

42 

0.21 

42 

>0.05 

Maine 

Rudnicky  and  Hunter  1993 

Shrub 

Sparrow 

0.38 

40 

0.28 

40 

>0.05 

Massachusetts 

This  study 

Quail 

0.60 

20 

0.51 

80 

>0.05 

Massachusetts 

DeGraaf,  unpubl.  data“ 

0.16 

50 

0.14 

50 

>0.05 

Maine 

Vander  Haegan  and  DeGraaf  1996'’ 

0.55 

42 

0.29 

42 

0.015 

Maine 

Rudnicky  and  Hunter  1993 

^ Recalculated  from  data  used  by  Danielson  et  al.  1997. 

’’  From  Table  1 ; edge  = 5 m and  interior  = 200  m from  edge. 


studies  almost  never  include  surveys  of  the 
predator  community  in  the  study  area  or  index 
predator  abundance  (Yahner  1996).  We  in- 
dexed small  mammal  abundance;  the  overall 
abundance  and  distribution  of  small  mammals 
at  forest  edge  and  interior  were  similar,  as 
were  the  distributions  of  depredated  ground 
and  shrub  nests,  but  the  abundance  of  small 
mammals  was  not  related  to  nest  predation  for 
either  nest  type  or  location  in  one  season.  Ei- 
ther the  small  mammals  that  we  detected  were 
not  major  nest  predators,  or  they  did  not  vary 
sufficiently  in  abundance  in  a homogeneous 
landscape  in  one  season  to  show  a relationship 
with  nest  predation.  A relationship  between 
nest  predation  and  small  mammal  abundance 
may  be  detectable  only  over  time;  small  mam- 
mals vary  greatly  from  year  to  year  with  food 
abundance  (Elkinton  et  al.  1996).  Long-term 
studies  are  needed  to  determine  if  this  is  the 
case. 

Small-mouthed  nest  predators  such  as  Per- 
omyscus  were  abundant  in  our  study  area  ( 10— 
40/ha;  Elkinton  et  al.  1996)  compared  to  larg- 
er generalist  predators  such  as  fishers  (Martes 
pennanti;  21/100  km-  in  central  Massachu- 
setts; York  1996)  that  have  been  shown  to 
depredate  artificial  nests  in  northern  New 
England  (DeGraaf  1995).  If  small  eggs  that 
are  susceptible  to  depredation  by  all  potential 
predators  are  used  in  artificial  nests,  then 
ubiquitous,  abundant  predators  (e.g.,  small 
mammals)  may  swamp  the  effect  of  larger 
generalist  predators,  even  if  the  latter  are  more 
abundant  along  forest  edges  (apparently  not 
the  case  in  the  northeastern  U.S.).  Moreover, 
our  data  suggest  that  ground  nests  may  be  par- 


ticularly vulnerable  to  predators  such  as  mice 
and  chipmunks  (Haskell  1995,  Bayne  et  al. 
1997),  which  spend  more  time  foraging  on  the 
ground  than  in  shrubs  or  trees  (Madison  1977, 
Elliot  1978,  Graves  et  al.  1988).  Hence,  the 
hypothesis  that  egg  predation  rates  are  elevat- 
ed at  forest  edges  may,  in  large  part,  be  an 
artifact  of  egg  size.  Virtually  all  studies  to  date 
have  used  quail  eggs  (see  Paton  1994,  Major 
and  Kendal  1996;  but  see  George  1987)  which 
apparently  cannot  be  opened  by  the  most 
abundant  small-mouthed  predators  in  temper- 
ate forests. 

Do  natural  nests  containing  small  eggs 
show  edge  related  predation  in  the  extensively 
forested  northeastern  U.S.?  In  a 2-yeai'  study 
of  ground  nesting  Ovenbird  (Seiiirus  aurocap- 
illus)  reproductive  success  in  New  Hampshire 
(King  et  al.  1996),  nests,  tenitories,  and  ter- 
ritorial males  were  equally  distributed  in  edge 
(0-200  m)  and  forest  interior  (201-400  m); 
nest  survival  was  higher  in  the  forest  interior 
in  year  1 , but  not  in  year  2.  The  proportion  of 
pairs  fledging  at  least  1 young,  fledgling 
weight,  and  fledgling  wing  chord  did  not  dif- 
fer between  edge  and  interior  over  the  course 
of  the  study. 

In  extensive  mixed- wood  forests  in  New 
England,  edge  related  differences  in  artificial 
nest  predation  rates  have  not  been  consistently 
demonstrated.  In  our  study  predation  rates 
were  substantially  higher  on  artificial  ground 
nests  that  contained  small  eggs  than  those  in 
studies  that  used  quail  eggs  (Table  3).  All  po- 
tential predators  can  open  small  eggs,  and 
their  use  should  result  in  higher  predation 
rates  because  small-mouthed  predators  are 


DeGnulfci  al.  • PREDATION  ON  SMALL  EGGS  IN  ARTILICIAL  NESTS 


241 


more  abundant  than  large  nest  predators.  Pre- 
dation rates  of  artificial  nests  often  have  been 
assumed  to  track  those  of  natural  nests  (Major 
and  Kendal  1996),  but  they  may  not  unless 
egg  sizes  closely  approximate  those  of  the 
species  of  concern.  For  example,  nest  survival 
of  natural  nests  was  lower  than  that  of  exper- 
imental nests  containing  quail  eggs  in  Panama 
because  of  the  abundance  of  small-mouthed 
nest  predators  (Roper  1992).  Predation  rates 
in  quail  egg  experiments  (e.g.,  Loiselle  and 
Hoppes  1983;  Martin  1987,  1988)  may  be 
useful  to  compaie  local  habitats,  but  may  be 
inappropriate  for  estimating  natural  predation 
rates  or  for  comparing  ai’eas  inhabited  by  dif- 
ferent predators  (Roper  1992).  Nest  predation 
is  a dominant  factor  in  avian  reproductive  suc- 
cess (Ricklefs  1969,  Martin  1988);  results  of 
experiments  that  exclude  major  sources  of 
mortality  (i.e.,  small-mouthed  predators)  may 
not  be  representative  (Roper  1992). 

Only  if  appropriate  egg  sizes  are  used  can 
predation  rates  in  relation  to  habitat  edge  or 
placement  be  generalized  or  approximated  for 
natural  nests.  Even  then,  effects  such  as  nest 
defense  and  appearance  (Martin  1987)  are  dif- 
ficult to  address.  Our  data  suggest  that  egg 
predation  rates  may  be  strongly  related  to  egg 
size,  other  factors  being  equal,  because  the 
most  abundant  predators  can  only  open  small 
eggs. 

ACKNOWLEDGMENTS 

We  thank  M.  Stoddard  for  assistance  with  field 
work,  J.  Buonaccorsi  for  help  with  statistical  analyses, 
R.  Askins,  D.  Haskell,  D.  King,  R Sievert,  and  R.  Yah- 
ner  for  their  critical  reviews,  and  M.  A.  Sheremeta  for 
typing  the  manuscript. 

LITERATURE  CITED 

American  Society  of  Mammalogists.  1998.  Guide- 
lines for  the  capture,  handling,  and  care  of  mam- 
mals as  approved  by  the  American  Society  of 
Mammalogists.  J.  Mamm.  79:1404-1409. 
Andrew,  H.  and  P.  Angel.stam.  1988.  Elevated  pre- 
dation rates  as  an  edge  effect  in  habitat  islands: 
experimental  evidence.  Ecology  69:.544-547. 
Baker,  B.  W.  1980.  Hair-catchers  aid  in  identifying 
mammalian  predators  of  ground-nesting  birds. 
Wildl.  Soc.  Bull.  8:257-259. 

Bayne,  E.  M.  and  K.  A.  Hob.son.  1997.  Comparing 
the  effects  of  landscape  fragmentation  by  forestry 
and  agriculture  on  predation  of  artificial  nests. 
Conserv.  Biol.  11:1418—1429. 

Bayne,  E.  M.,  K.  A.  Hobson,  and  P.  Eargey.  1997. 


Predation  on  artificial  nests  in  relation  to  forest 
type:  contrasting  the  use  of  quail  and  plasticine 
eggs.  Ecography  20:233-239. 

Bury,  R.  B.  and  P.  S.  Corn.  1987.  Evaluation  of  pitfall 
trapping  in  northwestern  forests:  trap  arrays  with 
drift  fences.  J.  Wildl.  Manage.  51:112-119. 

Craig,  D.  P.  1998.  Chipmunks  use  leverage  to  eat 
oversized  eggs:  support  for  the  use  of  quail  eggs 
in  artificial  nest  studies.  Auk  115:486-489. 

Danielson,  W.,  R.  M.  DeGraae,  and  T.  K.  Fuller. 
1997.  Rural  and  suburban  forest  edges:  effect  on 
egg  predators  and  nest  predation  rates.  Landscape 
Urban  Plann.  38:25-36. 

DeGraae,  R.  M.  1995.  Nest  predation  rates  in  man- 
aged and  reserved  extensive  northern  hardwood 
forests.  For.  Ecol.  Manage.  79:227-234. 

DeGraae,  R.  M.  and  T.  J.  Maier.  1996.  Effect  of  egg 
size  on  predation  by  white-footed  mice.  Wilson 
Bull.  108:535-539. 

Donovan,  T.  M.,  P.  W.  Jones,  E.  M.  Annand,  and  F. 
R.  Thompson,  III.  1997.  Variation  in  local-scale 
edge  effects:  mechanisms  and  landscape  context. 
Ecology  78:2064-2075. 

Elkinton,  j.  S.,  W.  M.  Healy,  J.  P.  Buonaccorsi,  G. 
H.  Boettner,  a.  M.  Hazzard,  H.  R.  Smith,  and 
A.  M.  Liebhold.  1996.  Interactions  among  gypsy 
moths,  white-footed  mice,  and  acorns.  Ecology 
77:2332-2342. 

Elliot,  L.  1978.  Social  behavior  and  foraging  ecology 
of  the  eastern  chipmunk  (Tamis  striatus)  in  the 
Adirondack  Mountains.  Smithson.  Contrib.  Zool. 
265:1-107. 

Eyre,  F.  H.  1980.  Forest  cover  types  of  the  United 
States  and  Canada.  Society  of  American  Foresters, 
Washington,  D.C. 

Fenske-Crawford,  T.  J.  and  G.  J.  Niemi.  1997.  Pre- 
dation of  artificial  ground  nests  at  two  types  of 
edges  in  a forest-dominated  landscape.  Condor  99: 
14-24. 

George,  T.  L.  1987.  Greater  land  bird  densities  on  is- 
land vs  mainland:  relation  to  nest  predation  level. 
Ecology  68:1393-1400. 

Gon'ERiED,  B.  M.  AND  C.  F.  Thompson.  1978.  Exper- 
imental analysis  of  nest  predation  in  an  old-field 
habitat.  Auk  95:304-312. 

Graves,  S.,  J.  Maldonado,  and  J.  O.  Wolfe.  1988. 
Use  of  ground  and  arboreal  microhabitats  by  Per- 
otnvscus  leiicopiis  and  Peromyscits  inaniculaliis. 
Can.  J.  Zool.  66:277-288. 

Hannon,  S.  J.  and  S.  E.  Cotthrill.  1998.  Nest  pre- 
dation in  aspen  woodlots  in  an  agricultural  area  in 
Alberta:  the  enemy  from  within.  Auk  1 15:16-25. 

Haskei.l,  D.  G.  1995.  Forest  fragmentation  and  nest 
predation:  are  experiments  with  Japanese  Quail 
eggs  misleading?  Auk  1 12:767-770. 

Hhskk,  E.  j.  1995.  Mammals  in  abundances  on  forest- 
farm  edges  versus  forest  interiors  in  southern  Il- 
linois: is  there  an  edge  effect  ? J.  Mamm.  76:562- 
568. 

Hosmer,  D.  W.,  Jr.  and  S.  Lemeshow.  1989.  Applied 


242 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


logistic  regression.  John  Wiley  and  Sons,  New 
York. 

King,  D.  I.,  C.  R.  Griffin,  and  R.  M.  DeGraaf.  1996. 
Effects  of  clearcutting  on  habitat  use  and  repro- 
ductive success  of  the  Ovenbird  in  forested  land- 
scapes. Conserv.  Biol.  10:1380-1386. 

Loiselle,  B.  a.  and  W.  a.  Hoppes.  1983.  Nest  pre- 
dation in  insular  and  mainland  lowland  rainforest 
in  Panama.  Condor  85:93-95. 

Madison,  D.  M.  1977.  Movements  and  habitat  use 
among  interacting  Peromyscus  leiicopiis  as  re- 
vealed by  radio  telemetry.  Can.  Lield-Nat.  91: 
273-281. 

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. 

Marini,  M.  A.  and  C.  Melo.  1998.  Predators  of  quail 
eggs,  and  the  evidence  of  the  remains:  implica- 
tions for  nest  predation  studies.  Condor  100:395- 
399. 

Martin,  T.  E.  1987.  Artificial  nest  experiments:  effects 
of  nest  appearance  and  types  of  predators.  Condor 
89:925-928. 

Martin,  T.  E.  1988.  Habitat  and  area  effects  on  forest 
bird  assemblages:  is  nest  predation  an  influence? 
Ecology  69:74-84. 

Maxon,  S.  J.  and  L.  W.  Oring.  1978.  Mice  as  a source 
of  egg  loss  among  ground-nesting  birds.  Auk  95: 
582-584. 

Mengak,  M.  T.  and  D.  C.  Guynn,  Jr.  1987.  Pitfalls 
and  snap  traps  for  sampling  small  mammals  and 
herpetofauna.  Am.  Midi.  Nat.  118:284—288. 

Miller,  D.  H.  and  L.  L.  Getz.  1977.  Factors  influ- 
encing local  distribution  and  species  diversity  of 
forest  small  mammals  in  New  England.  Can.  J. 
Zool.  55:806-814. 

Niesmuth,  N.  D.  and  M.  S.  Boyce.  1997.  Edge-related 
nest  losses  in  Wisconsin  pine  barrens.  J.  Wildl. 
Manage.  61:1234-1239. 

Paton,  P.  W.  C.  1994.  The  effect  of  edge  on  avian 
nesting  success:  how  strong  is  the  evidence?  Con- 
serv. Biol.  8: 1 7-26. 

Pelikan,  j.,  j.  Zejda,  and  V.  Holisova.  1977.  Effi- 
ciency of  different  traps  in  catching  small  mam- 
mals. Folia  Zool.  26:1-13. 

PiCMAN,  J.  1988.  Experimental  study  of  predation  on 


eggs  of  ground-nesting  birds:  effects  of  habitat 
and  nest  distribution.  Condor  90:124—131. 

Reitsma,  L.  R.,  R.  T.  Holmes,  and  T.  W.  Sherry. 
1990.  Effects  of  removal  of  red  squirrels,  Tamias- 
ciurus  hudsonicus,  and  eastern  chipmunk,  Tamias 
striatiis,  on  nest  predation  in  a northern  hardwood 
forest:  an  artificial  nest  experiment.  Oikos  57: 
375-380. 

Ricklefs,  R.  E.  1969.  An  analysis  of  nesting  mortality 
in  birds.  Smithson.  Contrib.  Zool.  9:1-48. 

Roper,  J.  J.  1992.  Nest  predation  experiments  with 
quail  eggs:  too  much  to  swallow?  Oikos  65:528- 
530. 

Rudnicky,  T.  C.  and  M.  L.  Hunter,  Jr.  1993.  Avian 
nest  predation  in  clearcuts,  forests,  and  edges  in  a 
forest-dominated  landscape.  J.  Wildl.  Manage.  57: 
358-364. 

SAS  Institute  Inc.  1989.  SAS/STAT'  user’s  guide, 
version  6,  fourth  ed.  Cary,  North  Carolina. 

Small,  M.  E and  M.  L.  Hunter.  1988.  Forest  frag- 
mentation and  avian  nest  predation  in  forested 
landscapes.  Oecologia  76:62-64. 

Vander  Haegan,  W.  M.  and  R.  M.  DeGraaf.  1996. 
Predation  on  artificial  nests  in  forested  riparian 
buffer  strips.  J.  Wildl.  Manage.  60:542-550. 

Whelan,  C.  J.,  M.  L.  Dilger,  D.  Robson,  N.  Hallyn, 
AND  S.  Dilger.  1994.  Effects  of  olfactory  cues  on 
artificial-nest  experiments.  Auk  1 1 1 :945-952. 

Williams,  D.  E and  S.  E.  Braun.  1983.  Comparison 
of  pitfall  and  conventional  traps  for  sampling 
small  animal  populations.  J.  Wildl.  Manage.  47: 
841-845. 

Yahner,  R.  H.  1996.  Forest  fragmentation,  artificial 
nest  studies,  and  predator  abundance.  Conserv. 
Biol.  10:672-673. 

Yahner,  R.  H.,  C.  G.  Mahan,  and  C.  A.  DeLong. 
1993.  Dynamics  of  depredation  on  artificial 
ground  nests  in  habitat  managed  for  Ruffed 
Grouse.  Wilson  Bull.  105:172-179. 

Yahner,  R.  H.  and  D.  P.  Scott.  1988.  Effects  of  forest 
fragmentation  on  depredation  of  artificial  nests.  J. 
Wildl.  Manage.  52:158-161. 

York,  E.  C.  1996.  Fisher  population  dynamics  in 
north-central  Massachusetts.  M.Sc.  thesis,  Univ. 
of  Massachusetts,  Amherst. 

Zar,  j.  H.  1996.  Biostatistical  Analysis,  third  ed.  Pren- 
tice-Hall, Englewood  Cliffs,  New  Jersey. 


Wilson  Bull.,  111(2),  1999,  pp.  243-250 


BIRD  USE  OF  BURNED  AND  UNBURNED  CONIFEROUS  FORESTS 

DURING  WINTER 

KAREN  J.  KREISEL'  2^  AND  STEVEN  J.  STEIN' 


ABSTRACT. — Cavity-nesting  bird  species  have  been  shown  to  be  associated  with  early  post-fire  habitat  during 
the  breeding  season  but  little  study  has  been  done  of  birds  in  the  non-breeding  season.  We  compared  bird 
composition  and  foraging  behavior  during  the  winter  in  burned  and  unburned  forests.  We  conducted  point  counts 
during  four  consecutive  winters  immediately  following  a stand  replacement  fire.  Burned  and  unburned  forests 
had  similar  numbers  of  bird  species,  yet  species  composition  was  distinctly  different.  Trunk  and  branch  foraging 
species  were  2.5  times  more  abundant  in  burned  forest  than  in  unburned  forest.  Within  burned  forests,  trunk 
and  branch  foraging  species  significantly  decreased  from  the  first  winter  post-fire  to  the  fourth  winter  post-fire. 
We  conducted  foraging  observations  of  four  woodpecker  species  within  burned  forests  only.  Woodpeckers  used 
western  larch  (Lari.x  occidentalis),  ponderosa  pine  {Pinus  ponderosa)  and  Douglas-fir  (Pseudotsuga  menziesii) 
snags  that  were  greater  than  23  cm  in  diameter.  Stand  replacement  fires  may  play  an  important  role  in  maintaining 
populations  of  trunk  and  branch  foraging  species  in  mixed  coniferous  forests  in  northeastern  Washington.  Re- 
ceived 24  August  1998,  accepted  28  Dec.  1998. 


Wildfire  plays  a major  role  in  determining 
landscape  patterns  by  creating  large  mosaics 
of  burned  habitat  intermixed  with  unbumed 
habitat.  Prior  to  fire  suppression  policies  in  the 
early  1900s,  wildfires  were  more  frequent  and 
widespread  in  the  western  United  States 
(Agee  1994,  Hejl  1994).  Mixed-conifer  forests 
of  the  Pacific  Northwest  have  a history  of 
stand  replacement  fires,  in  which  most  trees 
are  killed,  occurring  every  140-340  years 
(Agee  1994).  Stand  replacement  fires  create  a 
unique  habitat  of  large  patches  of  standing 
dead  trees  that  host  great  numbers  of  bark  and 
wood-boring  beetles  (Fumiss  1965,  Amman 
and  Ryan  1991)  which  serve  as  food  for  birds 
inhabiting  recently  burned  forests  (Spring 
1965,  Wickman  1965).  This  food  resource  is 
thought  to  decrease  dramatically  2-5  years 
post-fire  (Koplin  1972). 

Cavity-nesting  bird  species  are  associated 
with  early  post-fire  forests  (1-9  years  post- 
fire, Hutto  1995).  Many  researchers  have 
compared  bird  abundance  in  burned  and  un- 
bumed forest  during  the  breeding  season  and 
found  increased  numbers  of  cavity-nesting 
bird  species  in  early  post-fire  forests  (Bock 
and  Lynch  1970,  Raphael  and  White  1984, 
Raphael  et  al.  1987,  Hutto  1995,  Caton  1996, 
Hitchcox  1996).  Some  species,  including  the 


' Dept,  of  Biology,  Eastern  Washington  Univ.,  Che- 
ney, WA  99004. 

- Present  address:  2802  W.  Depot  Springs  Rd.,  Span- 
gle, WA  99031;  E-mail:  KKREISEL@hotmail.com 
’ Corresponding  author. 


Black-backed  Woodpecker  {Picoides  arcti- 
cus).  Three-toed  Woodpecker  (Picoides  tri- 
dactylus),  and  Mountain  Bluebird  (Sialia  mex- 
icana)  have  been  shown  to  be  more  common 
in  the  first  few  years  after  a fire  than  later 
(Hutto  1995).  How  long  this  high  abundance 
of  cavity-nesting  bird  species  persists  in  post- 
fire habitat  is  unclear. 

There  are  few  data  on  bird  species  com- 
position of  burned  forests  during  the  non- 
breeding season.  In  Montana,  Blackford 
(1955)  anecdotally  noted  numerous  wood- 
pecker species  in  burned  forests  during  the 
winter.  Blake  (1982)  compared  stand  replace- 
ment burned  and  unbumed  ponderosa  pine 
(Pinus  ponderosa)  forests  during  the  non- 
breeding season  and  found  bark  insectivores 
(including  woodpeckers)  to  be  more  abundant 
in  burned  forests.  Non-breeding  season  habitat 
may  play  a significant  role  in  determining 
overall  survival  and  numbers  of  individuals 
that  breed  (Conner  1979,  Graber  and  Graber 
1983,  Klein  1988)  and  is  probably  as  impor- 
tant as  breeding  season  habitat  for  the  persis- 
tence of  avian  populations.  Many  resident  spe- 
cies present  in  burned  forest  during  the  non- 
breeding season  may  not  remain  in  burned 
forest  year-round  but  seasonally  migrate  short 
distances  to  different  habitats,  thereby  maxi- 
mizing their  use  of  available  resources. 

Several  studies  have  examined  the  chai'ac- 
teristics  of  trees  used  for  nesting  by  cavity- 
nesting birds  in  burned  forests.  In  general, 
cavity-nesters  seem  to  prefer  broken  top  snags 


243 


244 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


and  snags  greater  than  23  cm  in  diameter  for 
nesting  (Raphael  and  White  1984,  Hutto  1995, 
Caton  1996,  Hitchcox  1996).  Fewer  research- 
ers have  examined  the  snag  characteristics 
used  for  foraging  by  cavity-nesting  birds  in 
burned  forests  (Hutto  1995,  Caton  1996),  even 
though  food  availability  may  be  as  important 
to  excavating  cavity-nesting  species  as  the 
availability  of  nest  sites,  especially  in  burned 
forests  where  nest  snags  may  be  abundant 
(Caton  1996).  Because  snags  used  for  nesting 
and  foraging  by  cavity-nesting  birds  may  dif- 
fer, information  on  the  characteristics  of  snags 
used  for  nesting  would  be  of  particular  interest 
to  managers  to  maintain  fire-dependent  bird 
populations. 

To  investigate  the  use  of  burned  forests  dur- 
ing the  winter,  we  examined  bird  community 
composition  and  foraging  behaviors  in  burned 
and  unburned  forests  during  four  consecutive 
winters  in  northeastern  Washington. 

STUDY  AREA  AND  METHODS 

Study  urea. — The  study  sites  were  located  in  the 
Kettle  River  Range  on  the  Colville  National  Lorest  in 
the  northeastern  corner  of  Washington,  26  km  north  of 
Republic  (48°  65'  N,  118°  73'  W).  This  area  is  pre- 
dominantly a mixed-conifer  forest  consisting  of  sub- 
alpine  fir  (Abies  lasiocarpa),  Douglas-fir  (Pseudotsuga 
menziesii),  Engelmann  spruce  (Picea  engelmannii), 
ponderosa  pine,  western  larch  (Lari.x  occidentcdis),  and 
lodgepole  pine  (Pinus  contorta),  with  a minor  under- 
story component  of  ninebark  (Physocctrpus  nuilva- 
ceus).  Both  the  burned  and  unburned  study  sites  had 
similar  tree  species  composition  and  were  located  in 
unlogged  areas. 

In  August  1994  the  Copper  Butte  fire  was  ignited 
by  lightning  and  burned  4000  ha,  resulting  in  large 
mosaics  of  high  intensity  burned  areas  mixed  with  low 
intensity  and  unburned  areas  (U.S.  Dept,  of  Agric. 
1995).  We  chose  two  stand  replacement  burned  sites 
of  greater  than  80  ha  within  the  Copper  Butte  fire  area, 
with  elevations  ranging  from  1320-1650  m.  The  un- 
burned site  was  located  20  km  north  of  the  burned 
sites,  0.6  km  from  the  burn  boundary.  The  80  ha  un- 
burned site  had  similar  slope  and  aspect  to  the  burned 
sites  but  was  only  1030-1320  m in  elevation.  Average 
yearly  rainfall  for  this  area  is  384  cm  and  average  year- 
ly temperature  is  6.1°C.  During  November-Lebruary, 
when  all  of  the  data  were  collected,  the  average  month- 
ly precipitation  was  42  cm  (predominantly  snow)  and 
the  average  temperature  was  — 2.1°C. 

Avian  sampling. — We  used  the  point  count  method 
to  quantify  birds  in  burned  and  unburned  forests 
(Blondel  ct  al.  1981,  Hutto  et  al.  1986).  Point  ctxmt 
stations  were  systematically  laid  out  at  least  200  m 
apart,  100  m from  a change  in  habitat,  and  50  m from 


roads  and  creeks.  We  conducted  ten-minute  counts,  re- 
cording all  birds  seen  and  heard  within  100  m (Hutto 
1995).  Unfortunately,  we  could  not  always  distinguish 
between  Hairy  (P.  villosus).  Three-toed  (P.  tridacty- 
lus),  or  Black-backed  (P.  arcticus)  woodpeckers  so  we 
created  a category  of  “unknown  woodpecker”  to  ac- 
count for  these  detections.  Counts  were  conducted  on 
fair  weather  days  when  winds  were  less  than  25  kph, 
with  little  or  no  precipitation,  and  temperatures  greater 
than  -9°  C.  Winter  counts  were  conducted  between 
08:00  PST  (I  h after  sunrise)  and  16:00  from  8 No- 
vember to  4 Lebruary  1994-1997. 

Over  the  four  years  of  the  study,  the  number  of  point 
count  stations  surveyed  was  as  follows:  first  winter  to 
fourth  winter  post-fire,  burned  (8,  9,  13,  13)  and  un- 
burned (0,  9,  9,  9).  The  number  of  stations  on  the 
burned  site  varied  over  the  four  years  because  of  log- 
ging that  occurred  in  1996  at  four  of  the  stations. 
These  four  stations  were  no  longer  used,  and  eight  new 
stations  were  added  in  a nearby  similarly  burned  area. 
The  difference  in  the  number  of  count  stations  in 
burned  and  unburned  forest  affected  bird  diversity 
very  little.  The  four  point  count  stations  in  addition  to 
nine  on  the  burned  study  site  added  only  one  bird  spe- 
cies that  was  not  detected  in  the  first  nine  stations.  On 
both  study  sites  there  were  two  visits  to  each  point 
count  station  during  a winter  and  these  were  averaged. 
We  used  the  mean  number  of  birds  per  point  in  com- 
paring burned  and  unburned  forests  and  number  of 
points  as  sample  size. 

Bird  species  were  assigned  to  three  foraging  guilds 
based  on  Ehrlich  and  coworkers  (1988)  and  Hutto 
(1995):  (1)  trunk  and  branch  foragers  (timber  drillers 
and  timber  gleaners),  (2)  foliage  foragers  (including 
aerial  foragers),  and  (3)  ground  foragers. 

Foraging  sampling. — We  recorded  foraging  obser- 
vations of  four  woodpecker  species  in  the  burned  for- 
est only,  during  and  after  point  count  surveys.  A bird 
was  considered  foraging  when  it  appeared  to  be  ac- 
tively searching  for  and/or  obtaining  food  (i.e.,  digging 
or  pecking).  Only  the  first  foraging  observation  was 
recorded  for  each  individual  (Hejl  et  al.  1990).  The 
following  foraging  data  were  recorded:  bird  species, 
maneuver  (pecking  = tapping  on  the  surface,  flaking 
= removing  bark,  drilling  = excavating  into  wood), 
zone  of  foraging  (lower  trunk,  middle  trunk,  upper 
trunk,  branches),  tree  species,  tree  dbh  class  (8-22,  23- 
37,  38-53,  >53  cm),  burn  severity  of  trees  (alive  and 
green,  possibly  alive  with  some  green  present,  dead 
with  brown  needles  present,  dead  and  severely 
burned),  and  top  condition  (broken,  crooked,  double, 
intact).  We  combined  the  foraging  data  for  each  spe- 
cies over  four  winters  iti  order  to  increase  sample  size 
(Morrison  1984).  Consequently,  we  do  not  have  infor- 
mation on  between  year  differences  in  foraging. 

Tree  sampling. — Several  tree  characteristics  within 
the  burned  forest  were  recorded  at  count  stations  and 
between  stations  in  ().()4-ha  circular  samples  for  a total 
of  24  samples  (Martin  1994):  tree  species,  tree  dbh, 
burn  severity,  bark  cover,  branch  condition,  and  top 
condition. 


Kreise!  and  Stein  • BURNED  FOREST  BIRDS  IN  WINTER 


245 


TABLE  1.  Species  detecled  in  burned  and  imbiirned  ear 
ington  during  the  winters,  1994-1997. 

ly  post-lire  mixed-conifer 

forests 

in  northeast  Wash- 

Species 

Forage 

guilcU 

Nest 

guild^ 

Burned^ 

Unburncd^ 

/xl 

Year 

Mean/pl 

Year 

Mcan/pt 

Blue  Grouse,  Dendragapiis  ohscnnts 

G 

o 

2 

0.083 

— 

0 

>0.05 

Downy  Woodpecker,  Picoides  puhescens 

T 

c 

1,2,4 

0.175 

— 

0 

>0.05 

Hairy  Woodpecker,  Picoides  viUosns 

T 

c 

1, 2,3,4 

0.410 

2,4 

0.056 

0.035 

Three-toed  Woodpecker,  Picoides  tridactyins 

T 

c 

1,2 

0.042 

— 

0 

0.047 

Black-backed  Woodpecker,  Picoides  arcticiis 

T 

c 

1, 2,3,4 

0.406 

2 

0.019 

0.001 

Pileated  Woodpecker,  Drycocopiis  pi  lea  tits 

T 

c 

— 

0 

3 

0.019 

>0.05 

Unknown  Woodpecker 

T 

c 

1,2 

0.144 

2 

0.019 

0.05 

Gray  Jay,  Perisorius  canadensis 

F 

o 

1,2 

0.075 

2,3 

0.093 

>0.05 

Steller’s  Jay,  Cyanocitta  stelleri 

F 

o 

— 

0 

4 

0.019 

>0.05 

Clark’s  Nutcracker,  Nucifraga  coiumbiana 

F 

o 

1,2 

0.016 

— 

0 

>0.05 

Black-billed  Magpie,  Pica  pica 

G 

o 

2 

0.014 

— 

0 

>0.05 

Common  Raven,  Coreas  corax 

G 

o 

1 

0.016 

4 

0.037 

>0.05 

Black-capped  Chickadee,  Pants  atricapiUiis 

F 

c 

1,2 

0.031 

2 

0.074 

>0.05 

Mountain  Chickadee,  Pants  gambeli 

F 

c 

1, 2,3,4 

0.179 

2,3,4 

0.352 

>0.05 

White-breasted  Nuthatch,  Sitta  carolinensis 

T 

c 

4 

0.048 

— 

0 

>0.05 

Red-breasted  Nuthatch,  Sitta  canadensis 

T 

c 

— 

0 

2,3,4 

0.389 

0.001 

Brown  Creeper,  Certhis  aniericana 

T 

c 

3 

0.019 

— 

0 

>0.05 

Winter  Wren,  Troglodytes  troglodytes 

G 

o 

— 

0 

4 

0.019 

>0.05 

Golden-crowned  Kinglet,  Regiilits  satrapa 

F 

o 

— 

0 

2,3,4 

0.260 

0.001 

Varied  Thrush,  Ixoreus  naeviiis 

G 

o 

— 

0 

4 

0.056 

>0.05 

Red  Crossbill,  Loxia  curvirostra 

F 

o 

2,4 

0.067 

2 

0.148 

>0.05 

^Forage  guild:  T = trunk  and  branch.  F = foliage.  G = ground. 

^ Nest  guild:  O = open.  C = cavity. 

^ Burned  and  Unburned:  year:  year  post-fire  that  bird  species  were  present  during  the  winter,  mean/pl:  mean  number  of  birds  per  point  averaged  over  4 
years. 

^ /^-value:  2-way  ANOVA,  treatment  effect  (burned-unburned). 


Statistical  analyses. — The  mean  numbers  of  birds 
per  point  in  burned  and  unburned  forests  were  com- 
pared with  a 2-way  Analysis  of  Variance  (ANOVA). 
Bird  abundance  between  years  within  a treatment  was 
compared  using  a Kruskal-Wallis  nonparametric  1-way 
ANOVA  and  pairwise  comparisons  were  made  using 
Tukey’s  BSD  procedure.  Bird  abundances,  grouped  by 
foraging  guilds,  were  compared  between  burned  and 
unburned  forests  using  a Mann-Whitney  (/-test.  To  de- 
termine trends  in  abundance  over  four  years,  a Spear- 
man rank  order  correlation  was  performed  (Zar  1996). 
Comparisons  of  foraging  observations  and  available 
vegetation  were  made  using  the  Goodness  of  Fit  test 
for  categorical  data.  For  example,  if  the  distribution  of 
tree  species  used  by  Downy  Woodpeckers  was  signif- 
icantly different  than  the  proportion  of  available  tree 
species  the  Goodness  of  Fit  test  would  have  a.  P < 
0.05.  P values  of  less  than  0.05  were  considered  sig- 
nificant for  all  tests.  Statistical  analyses  were  per- 
formed using  SPSS  (SPSS  1993). 

RESULTS 

Winter  bird  assemblage  in  burned  and  im- 
biirned  forests. — During  the  winter,  20  bird 
species  were  detected  in  burned  and  unburned 
forests  combined  (Table  1).  Bird  species  com- 
position differed  between  the  two  forests.  In 


burned  forests  7 of  the  14  species  detected 
wei'e  restricted  to  burned  forests  and  in  un- 
bumed  forests,  6 of  the  14  species  detected 
were  restricted  to  unburned  forests.  Averaged 
over  four  years  tmnk  and  branch  foraging  spe- 
cies were  2.5  times  more  abundant  in  burned 
forest  {U  = 2.0,  df  = 1,  P > 0.05),  and  foliage 
foraging  species  were  3 times  more  abundant 
in  unburned  forest  ((/  = 2.0,  df  = 1,  P > 0.05; 
Fig.  1).  Woodpecker  species  combined  were 
10  times  more  abundant  in  burned  forests  than 
in  unburned  forest.  The  four  most  abundant 
species  detected  in  burned  forest  in  descend- 
ing order  were  the  Hairy  Woodpecker,  Black- 
backed  Woodpecker,  Mountain  Chickadee 
(Pants  gambeli),  and  Downy  Woodpecker 
(Picoides  piibescens),  and  in  unburned  forest 
the  Red-breasted  Nuthatch  (Sitta  canadensis). 
Mountain  Chickadee,  Golden-crowned  King- 
let (Regidus  satrapa),  and  Red  Crossbill  (Lox- 
ia  curvirostra).  Black-backed  Woodpeckers 
(2-way  ANOVA:  F = 11.26,  df  = 1,  P = 
0.001)  and  Hairy  Woodpeckers  (P  = 4.62,  df 
= 1,  P = 0.035)  were  significantly  more  abun- 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


246 
2.0  - 


K 

Z 

1.8  - 

2 

1.6  - 

CO 

1.4  - 

o 

K 

1.2 

m 

1.0  - 

u. 

O 

-I-- 

CO 

d 

0.6  ^ 

1 

0 

1 

lU 

S 

0.2  1 
0.0  f 

■ Burned 
H Unburned 


T T 


y/A/j 


TRUNK  & FOLIAGE  GROUND 

BRANCH 


LIG.  1.  Mean  (±SD)  number  of  birds  per  point 
over  four  years,  by  foraging  guild  in  burned  and  un- 
burned forests  during  winter  (Mann-Whitney  t/-test: 
for  each  guild  all  P > 0.05). 


LIG.  2.  Trunk  and  branch  foraging  species  abun- 
dance over  four  years,  in  burned  and  unburned  forest 
during  winter  (Spearman  rank:  Burned  s^  = -1.0,  df 
= 3,  P = 0.001;  Unburned  s,  = 0,  df  = 2,  P > 0.05). 


dant  in  burned  forest  regardless  of  year  (Table 
1).  The  Three-toed  Woodpecker  was  signifi- 
cantly more  abundant  (F  = 4. 12,  df  = I,  P = 
0.047)  in  burned  forest  only  during  the  second 
winter  post-fire  (Kruskal- Wallis:  = 11-9,  df 
= 3,  P = 0.008;  Tukey  P = 0.042).  The  Red- 
breasted Nuthatch  (F  = 13.10,  df  = 1,  F = 
0.001)  and  Golden-crowned  Kinglet  (F  = 
11.80,  df  = 1,  F = 0.001)  were  significantly 
more  abundant  in  unbumed  forest. 

Change  in  bird  abundance  over  four 
years. — Bird  species  composition  and  abun- 
dance in  burned  forests  during  the  winter 
changed  from  1994-1997.  Eleven  species 
were  present  the  first  or  second  winter  post- 
fire and  were  absent  by  the  third  winter  post- 
fire (Table  1).  The  Brown  Creeper  {Certhia 
americana)  was  not  present  until  the  third 
winter  post-fire  and  the  White-breasted  Nut- 
hatch (Sitta  canadensis)  was  not  present  until 
the  fourth  winter  post-fire.  The  abundance  of 
trunk  and  branch  foraging  species  decreased 
3.8  times  from  winter  1994—1997  and  showed 
a significant  negative  trend  (Spearman  rank:  r^ 
= — 1.0,  F = 0.001;  Fig.  2).  Although  all  four 
woodpecker  species  present  in  burned  forest 
during  the  winter  had  declined  by  the  fourth 
winter  post-fire,  none  of  the  changes  was  sig- 
nificant (all  F > 0.05).  Within  burned  forest 
there  was  no  significant  trend  in  the  abun- 
dance of  foliage  and  ground  foraging  species 
from  1994—1997.  Within  unburned  forest 
there  was  a significant  decreasing  trend  of  fo- 
liage foraging  species  (u  = —1.0,  F = 0.001) 
during  the  four  years. 

Foraging. — In  burned  forest  Downy,  Hairy, 
Three-toed,  and  Black-backed  woodpeckers 


foraged  upon  standing  dead  trees  99%  of  the 
time  and  1%  of  the  time  on  logs  (n  = 145). 
Woodpeckers  used  burned  trees  with  brown 
needles  51%  of  the  time,  significantly  differ- 
ent than  the  proportion  available  (20%;  Good- 
ness of  Fit:  x‘  = 14.98,  df  = 2,  F = 0.001). 
They  foraged  predominantly  on  Douglas-fir 
(61%),  western  larch  (38%),  and  ponderosa 
pine  (38%;  Fig.  3).  Western  larch  (13%  avail- 
able) and  ponderosa  pine  (2%  available)  were 
used  for  foraging  significantly  more  than  ex- 
pected (x^  = 60.58,  df  = 4,  F = 0.001).  Trees 
greater  than  23  cm  in  diameter  were  used 
(84%)  significantly  more  than  the  proportion 
available  (36%;  Fig.  4;  X"  = 85.86,  df  = 3,  F 
= 0.001).  Broken  top  snags  were  used  (14%) 
in  similar  proportions  to  their  availability 
(12%;  x^  = 0.28,  df  = 1,  F > 0.05).  Available 
snags  consisted  mainly  of  severely  burned 
(80%),  intact  top  (83%),  Douglas-fir  (78%), 
western  larch  (13%),  or  sub-alpine  fir  (6%) 
with  a mean  diameter  of  22  cm. 

The  four  woodpecker  species  foraged  dif- 
ferently. Downy  Woodpeckers  foraged  pre- 
dominantly by  pecking  while  Hairy,  Three- 
toed, and  Black-backed  woodpeckers  foraged 
predominantly  by  flaking  and  drilling.  Hairy 
and  Three-toed  woodpeckers  foraged  on  sim- 
ilar tree  species  (Fig.  3)  and  on  similai'  parts 
of  trees  (Fig.  5).  Black-backed  Woodpeckers 
foraged  on  western  larch  and  Douglas-fir  (Fig. 
3)  and  foraged  predominantly  on  the  middle 
and  lower  trunks,  of  trees  (Fig.  5).  Downy 
Woodpeckers  foraged  most  frequently  on 
branches  of  ponderosa  pine  (Figs.  3,  5). 


Kreise!  anil  Stein  • BURNED  FOREST  BIRDS  IN  WINTER 


247 


FIG.  3.  Proportion  of  tree  species  available  and  used  for  foraging  by  four  woodpecker  species  during  winter 
in  burned  forests  (Goodness  of  Fit,  the  tree  species  distribution  used  by  all  woodpeckers  combined  was  signif- 
icantly different  than  the  available  distribution,  x"  = 60.58,  df  = 4,  P = 0.001;  each  bird  species  used  a 
significantly  different  tree  species  distribution  than  what  was  available  (all  P < 0.001)  and  each  bird  species 
foraged  on  different  tree  species  than  each  other  (all  P < 0.003)  except  Hairy  and  Three-toed  woodpeckers 
which  foraged  similarly  (x‘  = 3.573,  df  = 4,  P > 0.05)). 


DISCUSSION 

Trunk  and  branch  foraging  species  were 
more  abundant  in  recently  burned  forests  than 
in  unbumed  forests.  Other  studies  of  burned 
forests  had  similar  results  (Raphael  and  White 
1984,  Hutto  1995).  This  may  be  due  to  chang- 
es in  forest  structure  and  related  food  resourc- 
es as  a result  of  stand  replacement  fire.  Stand 
replacement  fire  changes  the  structure  of  a for- 
est from  a dense  canopy  cover,  with  shrub  un- 
dergrowth and  few  standing  dead  trees  to  little 
canopy  cover,  few  shrubs,  and  numerous 
standing  dead  trees.  Food  in  unbumed  forests 
include  a variety  of  seeds  and  insects  on  fo- 
liage, bark,  and  shrubs.  During  the  winter, 
food  in  burned  forests  may  be  limited  to  seeds 


8-22  23-37 

TREE  DIAME 


FIG.  4.  Proportion  of  tree  diameters  available  (n  — 
in  burned  forests.  Significantly  different  tree  diameters 
= 85.86,  df  = 3,  P = 0.001). 


from  fire  opened  cones  and  bark  and  wood- 
boring beetle  larvae  in  fire-killed  trees  (Hutto 
1995).  Bird  species  that  forage  on  the  ground 
and  foliage  probably  have  more  food  available 
in  unbumed  forests,  and  species  that  forage  on 
tmnks  and  branches  of  trees  probably  have 
more  food  available  in  burned  forests.  Bark 
and  wood-boring  beetle  larvae  that  are  abun- 
dant in  fire-killed  trees  are  the  major  food 
source  of  woodpeckers,  especially  during  win- 
ter (Brawn  et  al.  1982).  Therefore,  fire-killed 
trees  may  be  cmcial  in  supplying  year  round 
food  for  tmnk  and  branch  foraging  species. 
Large  diameter,  thick  barked  snags  are  typical 
locations  for  bark  and  wood-boring  beetle  lar- 
vae (Otvos  1965).  Woodpecker  species  in  this 

■ Available 

H Foraged 


38-52  > 53 

R CATEGORIES  (CM) 

173)  and  used  for  foraging  (/?  = 145)  by  woodpeckers 
were  used  than  what  was  available  (Goodness  of  Fit,  x' 


248 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


□ Downy  □ Hairy  ■ Three-toed  0 Black-backed 


I I 


Lower  Trunk  Middle  Trunk  Upper  Trunk 


Branches 


FIG.  5.  Foraging  zones  used  by  four  woodpecker  species  during  winter  in  burned  forests  [Goodness  of  Fit: 
Downy  (n  = 26,  x’  = 20.89,  df  = 3,  P = 0.001)  and  Black-backed  (n  = 42,  x"  = 14.83,  df  = 3,  P = 0.002) 
woodpeckers  used  zones  significantly  different  from  an  even  distribution.  Each  woodpecker  species  foraged  in 
significantly  different  zones  from  each  other  (all  P < 0.012)  except  Hairy  (n  = 55)  and  Three-toed  (n  = 12) 
woodpeckers  which  foraged  in  similar  zones  (x“  = 6.01,  df  = 3,  F > 0.05)]. 


Study  used  large  snags  between  23  and  37  cm 
dbh  disproportionately  more  often  for  forag- 
ing in  burned  forest.  In  Montana,  the  average 
foraging  snag  diameter  was  10-30  cm  (Caton 
1996).  Snag  species  used  most  often  for  for- 
aging in  this  study  included  thick  barked  west- 
ern larch,  ponderosa  pine,  and  Douglas-fir. 
Hutto  (1995)  and  Caton  (1996)  also  found  that 
woodpeckers  foraged  on  these  same  snag  spe- 
cies in  burned  forests  in  Montana. 

Within  burned  forest  there  were  several 
changes  in  the  bird  community  from  the  first 
winter  post-fire  to  the  fourth.  Some  species 
were  present  only  the  first  or  second  winter 
and  were  absent  by  the  third  winter.  Wood- 
peckers, however,  decreased  in  abundance 
from  the  first  winter  to  levels  equal  to  that 
found  in  unburned  forests  by  the  fourth  winter. 
In  the  Sierra  Nevada  Mountains,  Bock  and 
Lynch  ( 1 970)  found  woodpeckers  to  be  in  low 
numbers  by  six  years  post-fire.  Hutto  (1995) 
found  Three-toed  and  Black-backed  wood- 
peckers to  be  most  common  1-9  years  post- 
fire. Post-fire  forests  appear  to  support  a suc- 
cession of  bird  species,  with  early  post-fire 
conditions  being  conducive  to  specific  wood- 
pecker species.  The  decrease  in  abundance  of 
the  Downy,  Hairy,  Three-toed,  and  Black- 
backed  woodpeckers  may  have  occurred  as  a 
result  of  a decrease  in  food  resources.  Bark 
and  wood-boring  beetle  larvae  dramatically 
decrease  2-5  years  post-fire  (Koplin  1972). 
Food  resources  may  be  as  important  to  the 
distribution  of  these  species  as  nest-site  avail- 


ability. Caton  (1996)  suggested  that  the  avail- 
ability of  suitable  foraging  snags  played  a 
large  role  in  the  distribution  of  woodpecker 
species  in  burned  forests  in  Montana.  Further 
studies  on  food  resource  availability  in  burned 
and  unbumed  forests  may  better  help  explain 
their  distribution. 

Non-breeding  season  habitat  quality  is 
probably  as  important  to  the  persistence  of 
bird  populations  as  breeding  season  habitat 
(Conner  1979)  because  most  bird  mortality 
occurs  during  the  non-breeding  season  (Gra- 
ber  and  Graber  1983).  Favorable  winter  hab- 
itat can  increase  winter  survival  and  result  in 
more  birds  breeding  the  following  season.  Lo- 
cal movements  to  burned  habitats  by  resident 
bird  species  may  increase  survival.  The  use  of 
burned  forests  during  the  winter  by  wood- 
peckers may  also  increase  their  overwinter 
survivorship  by  increasing  food  and  decreas- 
ing predation.  Different  foraging  techniques 
observed  by  woodpeckers  in  burned  forest 
may  allow  many  species  to  coexist.  All  of 
these  factors  may  contribute  to  the  increased 
capability  of  recently  burned  forests  to  sup- 
port bird  species,  including  woodpeckers,  dur- 
ing the  non-breeding  season. 

The  100  m radius  point  counts  seemed,  to 
be  an  adequate  way  to  compare  the  relative 
abundance  of  birds  in  these  two  forests.  Visual 
detections  of  birds  in  the  burned  and  unburned 
forests  were  probably  different,  and  audible 
detections,  which  most  of  ours  were,  were 
probably  similar  in  the  two  forests.  Differenc- 


Kreisel  and  Stein  • BURNED  FOREST  BIRDS  IN  WINTER 


249 


es  in  auditory  detections  were  most  likely 
minimal  during  the  winter  when  it  was  ex- 
tremely quiet.  Dellasala  and  coworkers  (1996) 
found  little  difference  in  detection  rates  of 
birds  in  100  m radius  point  counts  in  young 
(20  yr)  and  old  growth  forests. 

Fire  suppression  over  the  past  100  years 
(Agee  1994,  Hejl  1994)  has  probably  had  a 
major  effect  on  bird  communities  in  mixed- 
conifer  forests.  Stand  replacement  fires  may 
help  to  increase  populations  of  cavity-nesting 
and  trunk  and  branch  foraging  species.  Pop- 
ulations of  Hairy,  Three-toed,  and  Black- 
backed  woodpeckers  might  be  maintained  by 
periodic  occurrences  of  stand  replacement 
fires  throughout  the  landscape.  The  spatial  and 
temporal  pattern  of  stand  replacement  fires 
needed  to  maintain  bird  populations  needs  fur- 
ther investigation. 

Management  implications. — Forest  manag- 
ers can  increase  cavity-nesting  bird  popula- 
tions by  relaxing  fire  suppression  policies  and/ 
or  by  initiating  prescribed  burning  programs. 
To  manage  for  cavity-nesting  birds  in  burned 
forests  snags  of  western  larch,  ponderosa  pine 
and  Douglas-fir  larger  than  23  cm  in  diameter 
should  be  present.  If  managers  can  delay  log- 
ging of  burned  forests  three  to  four  years,  the 
habitat  will  be  less  suitable  for  trunk  and 
branch  foraging  species  and  logging  may  have 
less  impact  on  these  species.  Future  studies 
investigating  snag  densities  and  patch  sizes  of 
burned  forests  required  to  maintain  popula- 
tions of  cavity-nesting  species  during  summer 
and  winter  are  needed  to  help  managers  pro- 
vide for  these  species. 

ACKNOWLEDGMENTS 

We  thank  D.  J.  Kinateder  for  invaluable  field  work 
and  helpful  discussion.  We  give  special  thanks  to  S.  J. 
Hejl  and  M.  A.  O’Connell  for  suggestions  and  com- 
ments on  the  manuscript.  Funding  was  provided  by 
Kettle  Range  Conservation  Group,  Northwest  Fund  for 
the  Environment,  Kongsgaard-Goldman  Foundation, 
Eastern  Washington  University,  and  the  Barry  Me- 
morial Research  Grant  from  the  Turnbull  Laboratory 
for  Ecological  Studies. 

LITERATURE  CITED 

Agee,  J.  A.  1994.  Fire  ecology  of  Pacific  Northwest 
forests.  Island  Press,  Washington,  D.C. 

Amman,  G.  D.  and  K.  C.  Ryan.  1991.  Insect  infesta- 
tions of  fire-injured  trees  in  the  greater  Yellow- 
stone area.  Intermountain  Research  Station,  Re- 


search Note  lNT-398.  U.S.  Forest  Service,  Ogden, 
Utah. 

Bi.ackeord,  j.  L.  1955.  Woodpecker  concentrations  in 
burned  forests.  Condor  57:28-30. 

Blake,  J.  G.  1982.  Influence  of  fire  and  logging  on 
non-breeding  bird  communities  of  ponderosa  pine 
forests.  J.  Wildl.  Manage.  46:404-415. 

Blondel,  j.,  C.  Ferry,  and  B.  Frochot.  1981.  Point 
counts  with  unlimited  distance.  Stud.  Avian  Biol. 
6:414-420. 

Bock,  C.  E.  and  J.  F.  Lynch.  1970.  Breeding  bird  pop- 
ulations of  burned  and  unburned  conifer  forests  in 
the  Sierra  Nevada.  Condor  72:182-189. 

Brawn,  J.  D.,  W.  H.  Elder,  and  K.  E.  Evans.  1982. 
Winter  foraging  by  cavity  nesting  birds  in  an  oak- 
hickory  forest.  Wildl.  Soc.  Bull.  10:271-275. 

Caton,  E.  L.  1996.  Effects  of  fire  and  salvage-logging 
on  a cavity-nesting  bird  community  in  northwest- 
ern Montana.  Ph.D.  diss.,  Univ.  of  Montana,  Mis- 
soula. 

Conner,  R.  N.  1979.  Seasonal  changes  in  woodpecker 
foraging  methods:  strategies  for  winter  survival. 
Pp.  95-105  in  The  role  of  insectivorous  birds  in 
forest  ecosystems  (W.  Lafeber,  Ed.).  Academic 
Press,  Orlando,  Florida. 

Dellasala,  D.  A.,  J.  C.  Hagar,  K.  A.  Engel,  W.  C. 
McComb,  R.  L.  Fairbanks,  and  E.  G.  Campbell. 
1996.  Effects  of  silvicultural  modifications  of  tem- 
perate rainforest  on  breeding  and  wintering  bird 
communities.  Prince  of  Wales  Island,  southeast 
Alaska.  Condor  98:706-721. 

Ehrlich,  P.  R.,  D.  S.  Dobkin,  and  D.  Wheye.  1988. 
The  birder’s  handbook:  field  guide  to  the  natural 
history  of  North  American  birds.  Simon  and 
Schuster,  New  York. 

Furniss,  M.  M.  1965.  Susceptibility  of  fire-injured 
Douglas-fir  to  bark  beetle  attack  in  southern  Ida- 
ho. J.  For.  63:8-1 1. 

Graber,  j.  W.  and  R.  R.  Graber.  1983.  Expectable 
decline  of  forest  bird  populations  in  severe  and 
mild  winters.  Wilson  Bull.  95:682—690. 

Hejl,  S.  J.  1994.  Human-induced  changes  in  bird  pop- 
ulations in  coniferous  forests  in  western  North 
America  during  the  past  100  years.  Stud.  Avian 
Biol.  15:232-246. 

Hejl,  S.  J.,  J.  Verner,  and  G.  W.  Bell.  1990.  Se- 
quential versus  initial  observations  in  studies  of 
avian  foraging.  Stud.  Avian  Biol.  13:166-173. 

Hitchcox,  S.  M.  1996.  Abundance  and  nesting  success 
of  cavity-nesting  birds  in  unlogged  and  salvage- 
logged  burned  forest  in  northwestern  Montana. 
M.S.  thesis,  Univ.  of  Montana,  Missoula. 

Hutto,  R.  L.  1995.  Composition  of  bird  communities 
following  stand  replacement  fires  in  northern 
Rocky  Mountain  conifer  forests.  Conserv.  Biol.  9: 
1041-1058. 

Huito,  R.  L.,  S.  M.  Pletschht,  and  P.  Hendricks. 
1986.  A fixed-radius  point  count  method  for  non- 
breeding and  breeding  season  use.  Auk  103:593— 
602. 


250 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


Klein,  B.  C.  1988.  Weather-dependent  mixed-species 
flocking  during  the  winter.  Auk  105:583-584. 

Koplin,  J.  R.  1972.  Measuring  predator  impact  of 
woodpeckers  on  spruce  beetles.  J.  Wildl.  Manage. 
36:308-320. 

Martin,  T.  E.  1994.  BBIRD  field  protocol.  Montana 
Cooperative  Wildlife  Research  Unit,  Univ.  of 
Montana,  Missoula. 

Morrison,  M.  L.  1984.  Influence  of  sample  size  and 
sampling  design  on  analysis  of  avian  foraging  be- 
havior. Condor  86:146-150. 

Otvos,  I.  S.  1965.  Studies  on  avian  predators  of  Den- 
droctonus  breviconiis  LeConte  (Coleoptera:  Scol- 
ytidae)  with  special  reference  to  Picidae.  Can. 
Ent.  97:1 184-1 199. 

Raphael,  M.  G.  and  M.  White.  1984.  Use  of  snags 
by  cavity-nesting  birds  in  the  Sierra  Nevada. 
Wildl.  Monogr.  86:1-66. 


Raphael,  M.  G.,  M.  L.  Morrison,  and  M.  R Yoder- 
WiLLiAMS.  1987.  Breeding  bird  populations  during 
25  years  of  postfire  succession  in  the  Sierra  Ne- 
vada. Condor  89:614-626. 

Spring,  L.  W.  1965.  Climbing  and  pecking  adaptations 
in  some  North  American  woodpeckers.  Condor 
67:457-488. 

SPSS.  1993.  SPSS  for  Windows  6.0.  SPSS,  Inc.,  Chi- 
cago, Illinois. 

U.S.  Dept,  op  Agric.  1995.  Pinal  supplement  to  final 
environmental  impact  statement  for  east  Curlew 
Creek  area  timber  sales.  Copper  Butte  fire  salvage. 
Republic  Ranger  District,  Republic,  Washington. 

WiCKMAN,  B.  E.  1965.  Black-backed  Three-toed  pre- 
dation of  Monochamus  oregonensis  (Coleoptera: 
Cerambycidae).  Pan-Pacific  Ent.  41:162-164. 

Zar,  j.  H.  1996.  Biostatistical  analysis,  third  ed.  Simon 
and  Schuster,  Upper  Saddle  River,  New  Jersey. 


Wilson  Bull.,  I 1 1(2),  1999,  pp.  251-256 


NEST  PREDATORS  OF  OPEN  AND  CAVITY  NESTING  BIRDS  IN 

OAK  WOODLANDS 

KATHRYN  L.  PURCELL'  ^ AND  JARED  VERNER' 


ABSTRACT. — Camera  setups  revealed  at  least  three  species  of  rodents  and  seven  species  of  birds  as  potential 
predators  at  artificial  open  nests.  Surprisingly,  among  avian  predators  identified  at  open  nests,  one  third  were 
Bullock’s  Orioles  (Icterus  hullockii).  Two  rodent  species  and  three  bird  species  were  potential  predators  at 
artificial  cavity  nests.  This  high  predator  diversity  was  consistent  with  previous  studies,  although  the  number  of 
avian  predators  at  open  nests  was  higher  than  expected.  Received  31  March  1998,  accepted  22  Nov.  1998. 


As  the  primary  source  of  nest  failure  among 
birds  (Lack  1968,  Ricklefs  1969),  predation  is 
a likely  factor  affecting  species’  coexistence, 
habitat  selection,  and  conservation  (Zimmer- 
man 1984;  Martin  1988a,  b).  When  nest  pre- 
dation differs  among  species,  habitats,  and  lo- 
cations, it  can  influence  life  history  traits  such 
as  clutch  size,  nest  placement,  developmental 
period,  and  number  of  broods  (Ricklefs  1969; 
Martin  1988c,  1995).  Avian  ecologists  gen- 
erally agree  that  predation  rates  differ  among 
species  nesting  in  cavities  and  open  (cup) 
nests  (Lack  1954,  Nice  1957,  Ricklefs  1969). 
Predators  may  differ  as  well,  but  little  is 
known  about  predators  of  bird  nests  because 
predation  is  rarely  observed,  and  observations 
are  biased  toward  diurnal  predators.  Some  re- 
searchers have  made  assumptions  about  broad 
classes  of  predators  based  on  the  appearance 
of  the  depredated  nest,  but  few  data  exist  to 
support  those  assumptions,  and  authors  dis- 
agree on  evidence  used  to  assign  depredated 
nests  to  predator  groups  and  the  reliability  of 
the  evidence  (Best  1978,  Best  and  Stauffer 
1980,  Wray  et  al.  1982,  Boag  et  al.  1984,  Her- 
nandez et  al.  1998a,  Marini  and  Melo  1998). 
Here  we  report  results  of  a camera  study  at 
both  artificial  open  and  cavity  nests.  The  pri- 
mary objective  of  our  study  was  to  identify 
nest  predators  as  part  of  a larger  study  of  re- 
productive success  among  birds  in  oak-pine 
woodlands  in  the  west-central  foothills  of  the 
Sierra  Nevada  of  California. 

STUDY  AREA  AND  METHODS 

The  study  was  done  at  the  San  Joaquin  Experimen- 
tal Range,  approximately  40  km  north  of  Fresno,  Cal- 


' USDA  Forest  Service,  Pacific  Southwest  Research 
Station,  2081  E.  Sierra  Ave.,  Fresno,  CA  93710. 

^ Corresponding  author; 

E-mail:  kpurcell/psw_fresno@fs.fed.us 


ifornia.  The  San  Joaquin  Experimental  Range  covers 
about  1875  ha  and  ranges  in  elevation  from  215  to  520 
m.  Climate  is  Mediterranean,  with  cool,  wet  winters 
and  hot,  dry  summers.  A sparse  woodland  overstory 
of  blue  oak  (Querciis  doiiglasii),  interior  live  oak  (Q. 
wislizenii),  and  foothill  pine  (Pinus  sabiniana)  covers 
most  of  the  San  Joaquin  Experimental  Range.  A scat- 
tered understory  of  shrubs  includes  mainly  wedgeleaf 
ceanothus  (Ceanothus  cuneatus),  chaparral  whitethorn 
(C.  leucodennis),  redberry  (Rhamnus  crocea),  and 
mariposa  manzanita  (Arctostaphylos  viscida  maripo- 
sa).  The  San  Joaquin  Experimental  Range  has  been 
lightly  to  moderately  grazed  since  about  1900  and  is 
surrounded  on  all  sides  by  similar  habitat. 

Using  nests  of  California  Towhees  (Pipilo  crissalis) 
collected  at  the  end  of  the  previous  field  season,  we 
situated  artificial  open  nests  low  in  small  trees  or 
shrubs  in  positions  similar  to  those  known  to  be  used 
by  California  Towhees  (on  a forked  branch  or  sup- 
ported by  several  twigs).  At  cavity  setups,  eggs  were 
placed  with  a “pick-up”  tool,  using  cavities  known  to 
be  deep  enough  for  cavity-nesting  species  at  the  San 
Joaquin  Experimental  Range.  Most  cavities  were  ex- 
cavated by  primary  cavity  nesters,  but  some  natural 
cavities  previously  used  for  nesting  were  also  used.  A 
fiberscope  (Purcell  1997)  was  used  to  guide  the  place- 
ment of  eggs  in  cavities,  to  monitor  eggs  for  possible 
predation,  and  to  measure  cavity  depth.  To  avoid  leav- 
ing olfactory  cues  at  nests,  field  personnel  washed  their 
hands  before  going  into  the  field  with  a soap  developed 
to  remove  human  scent  and  sprayed  their  boots  with  a 
scent  masker.  We  avoided  dead-end  trails  and  did  not 
create  paths  that  might  lead  predators  to  nests. 

One  experimental  egg  was  placed  in  open  nests,  and 
one  or  two  eggs  were  placed  in  cavity  nests,  the  num- 
ber and  type  depending  on  availability.  Most  eggs  used 
in  open  nests  were  from  wild  House  Sparrows  (Passer 
domesticus)  or  captive  Ringed  Turtle-Doves  (Strepto- 
pelia  risoria)-,  most  eggs  in  cavities  were  from  captive 
Zebra  Finches  (Poephila  gidlata).  We  sometimes  used 
a Buttonquail  (Turni.x  sp.)  egg  as  the  second  egg  in  a 
cavity  nest.  House  Sparrow  eggs  were  slightly  smaller 
and  Ringed  Turtle-dove  eggs  slightly  larger  than  those 
of  California  Towhees  (see  Baicich  and  Harrison 
1997).  Although  similar  in  size  to  eggs  of  the  Plain 
Titmouse  (Baeolophus  inornatus),  the  Zebra  Finch 


251 


252 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


eggs  were  smaller  than  eggs  of  all  cavity-nesting  spe- 
cies in  our  study  area. 

Predation  at  open  nests  was  monitored  mechanical- 
ly, with  an  egg  encircled  by  a loop  of  wire  attached 
within  a nest  of  the  California  Towhee.  Removal  of 
the  egg  activated  an  electrical  signal  to  a solenoid,  trip- 
ping a camera  mounted  nearby.  We  used  inexpensive, 
autofocus,  autoflash.  Keystone  550D  or  590AP  cam- 
eras, allowing  identification  of  both  diurnal  and  noc- 
turnal predators.  (Trade  names  and  commercial  prod- 
ucts are  mentioned  for  information  only;  no  endorse- 
ment by  the  U.S.  Department  of  Agriculture  is  im- 
plied.) Details  on  the  mechanical  system  for  open  nests 
are  available  from  KLP.  At  cavity  nests,  we  used  Trail- 
master  Active  Infrared  trail  monitors  with  weather- 
proof, autoflash  35  mm  cameras  to  monitor  predation. 
One  box  that  transmitted  (12.1  cm  L X 8.3  cm  W X 
4.6  cm  D)  and  one  that  received  (19.1  cm  L X 8.9  cm 
W X 5.3  cm  D)  the  infrared  beam  were  placed  on  each 
side  of  the  cavity  so  that  an  animal  entering  it  would 
break  the  beam,  triggering  the  camera  positioned  on  a 
nearby  branch  with  a good  view  of  the  cavity.  Because 
Trailmaster  units  are  designed  to  be  set  up  horizontally 
across  trails,  we  modified  the  boxes  so  they  could  be 
attached  easily  to  the  tree  bole  or  limb  with  bungee 
cords.  Sensitivity  was  set  at  the  minimum  delay  of  0.5 
s (one  pulse)  before  an  event  was  recorded,  and  the 
camera  delay  between  photos  was  set  at  the  minimum 
of  6 s. 

Based  on  the  nesting  seasons  of  cavity-  and  open- 
nesting species,  cameras  were  set  up  from  March 
through  June  1995  (cavities)  and  April  through  June 
1995  (open).  We  used  10  open-nest  setups  to  monitor 
70  open  nests,  and  7 Trailmaster  monitors  at  61  cavity 
nests.  Some  data  on  cavity  nests  were  also  included 
from  the  1993  and  1994  field  seasons  (eight  each  year). 
All  setups  were  checked  about  every  4 days.  If  an  egg 
was  taken,  or  not  taken  after  14  days,  the  setup  was 
dismantled  and  moved  to  another  location  and  installed 
using  fresh  eggs. 

All  artificial  nests  of  the  same  nest  type  were  sep- 
arated by  at  least  200  m in  an  effort  to  reduce  the 
chance  of  visitation  by  the  same  animal  at  two  or  more 
setups.  This  distance  was  thought  to  be  enough  to  as- 
sure independent  samples  of  the  small  mammals  iden- 
tified as  predators  in  this  study.  Based  on  spot  mapping 
at  the  San  Joaquin  Experimental  Range  (unpublished 
data),  territories  of  the  Western  Scrub-Jay  (Apheloco- 
ma  californica),  a common  nest  predator,  were  ap- 
proximately 120-210  m in  diameter.  Mean  territory 
diameters  of  other  common  bird  species  ranged  from 
180  m (California  Towhee)  to  310  m (Western  King- 
bird, Tyranniis  verticali.s).  Some  cavity  setups  were 
closer  than  200  m to  open  set-ups,  but  cameras  and 
eggs  were  not  placed  concurrently  at  the  two  nest 
types. 

At  open  nests,  we  measured  nest  height  and  the 
height  and  diameter  of  the  shrub  or  small  tree  contain- 
ing the  nest.  Diameter  was  measured  as  the  mean  of 
the  maximum  crown  diameter  and  the  widest  diameter 
perpendicular  to  the  maximum  diameter.  At  cavity 


nests,  we  measured  nest  height,  cavity  depth,  and  hor- 
izontal and  vertical  entrance  diameters.  We  tested  dif- 
ferences in  these  attributes  between  predated  and  un- 
predated nests  using  two-tailed  /-tests  (SAS  version 
6.12  for  Windows,  SAS  Institute  1988),  a = 0.05,  and 
Bonferroni  adjustments  for  multiple  tests.  We  calcu- 
lated power  according  to  Abramowitz  and  Stegun 
(1964)  based  on  specified  effect  sizes,  an  a of  0.05, 
and  two-tailed  tests  using  an  inhouse  SAS  program. 

RESULTS 

Open  nests. — Eggs  were  removed  from  39 
of  the  70  open  nest  setups,  but  we  could  iden- 
tify the  animals  at  only  29  of  those.  Rodents 
were  identified  at  four  (14%);  deer  mouse 
{Peromyscus  spp.,  two  cases),  California 
ground  squirrel  (Spermophilus  beecheyi,  one 
case),  and  Merriam’s  chipmunk  {Eutamias 
merriami,  one  case).  At  least  five  bird  species 
were  photographed  at  the  remaining  25  nests 
(86%)  from  which  an  egg  was  taken:  Western 
Scrub-Jay  (12  cases),  Bullock’s  Oriole  {Icter- 
us bullockii,  1 cases).  Acorn  Woodpecker 
{Melanerpes  fonnicivorus,  1 case).  Western 
Kingbird  (1  case,  a pair),  and  California  To- 
whee (1  case).  We  could  not  identify  the  bird 
species  at  the  three  remaining  setups. 

In  three  additional  cases,  eggs  were  pecked, 
chewed,  or  otherwise  damaged  but  not  re- 
moved. A pair  of  Plain  Titmice  pecked  a large 
hole  in  the  egg  at  one  nest;  a female  Brown- 
headed Cowbird  {Molothrus  ater)  punctured 
the  egg  in  another  nest;  and  either  a dusky- 
footed  woodrat  (Neotoma  fuscipes)  or  a West- 
ern Scmb-Jay  chewed  or  pecked  another  egg 
(both  species  were  photographed). 

We  may  have  underestimated  nocturnal  pre- 
dation. Only  one  photo  of  a deer  mouse  was 
taken  at  night.  In  2 of  the  10  cases  with  no 
identifiable  predator,  photos  were  taken  at 
night  but  were  dark,  perhaps  because  the  cam- 
era’s flash  was  too  far  from  the  nest  or  failed 
to  operate  properly  (see  also  Hernandez  et  al. 
1998b). 

No  attribute  measured  at  open  nest  setups 
differed  significantly  between  predated  and 
nonpredated  nests  (Table  1;  P > 0.05  in  all 
cases,  P < 0.017  required  for  Bonfenoni  ad- 
justment for  multiple  tests). 

Cavity  nests. — Photos  were  taken  at  47  of 
69  cavity  nests  where  the  egg  was  removed 
or  pecked  open.  Interpretation  of  the  photos 
was  complicated,  however.  First,  the  mini- 
mum camera  delay  did  not  allow  a photo  each 


Purcell  and  Verner  • NEST  PREDATORS 


253 


TABLE  1.  Nest  site  variables  and  results  of  r-tests  for  nonpredated  (n  = 18)  and  predated  (n  — 48)  open 
nests  at  the  San  Joaquin  Experimental  Range. 


Nonpredated 
open  nests 

Mean  (SE) 

Predated 
open  nests 

Mean  (SE) 

p. 

Power 

Nest  height  (in) 

1.21  (0.11) 

1.11  (0.06) 

0.48 

0.97'’ 

Substrate  height  (m) 

4.61  (0.81) 

4.25  (0.41) 

0.67 

0.9L 

Substrate  diameter  (m) 

6.03  (1.34) 

4.90  (0.40) 

0.43 

0.82” 

“ A F-value  of  0.017  is  needed  for  significance  at  a = 0.05  after  Bonferroni  adjustment  for  multiple  comparisons. 
’’  Based  on  an  effect  size  of  0.5m. 

Based  on  an  effect  size  of  3 m. 


time  the  infrared  beam  was  broken  after  an 
initial  photo  was  taken.  Consequently,  photos 
rarely  showed  animals  leaving  a cavity  nest 
and  none  showed  one  “caught  in  the  act”  of 
leaving  with  an  egg.  We  had  to  assume  that 
an  animal  in  a photo  consumed  the  missing 
egg.  Second,  in  seven  cases  the  egg  was  gone 
and  no  animal  was  evident  in  the  photo  (see 
also  Brooks  1996).  The  departure  of  some 
predators  from  a cavity  may  have  been  too 
rapid  for  it  to  be  caught  in  the  photo,  or  photos 
with  no  animal  may  have  resulted  from  direct 
sunlight  entering  the  receiver  window  (Kucera 
and  Barrett  1993).  Third,  in  15  instances  more 
than  one  species  entered  the  cavity  and  tripped 
the  camera  before  we  found  that  the  egg  was 
gone.  Fourth,  in  16  cases  nest  material  was 
added  to  the  cavity  by  birds  or  mammals  so 
we  could  not  ascertain  whether  the  eggs  had 
been  eaten  or  simply  buried.  We  did  not  in- 
clude these  cases  in  our  comparisons  of  pre- 
dated and  nonpredated  nests. 

In  nine  cavity  setups  with  missing  eggs, 
only  one  species  appeared  in  the  photos.  The 
assumed  predators  were  European  Starling 
(Sturnus  vulgaris;  four  cases).  House  Wren 
{Troglodytes  aedon;  one  case).  Western  Blue- 


bird (Sialia  mexicana',  one  case),  deer  mouse 
(one  case),  and  unidentified  squirrels  (proba- 
bly California  ground  squirrels;  two  cases). 
All  photos  were  taken  during  daylight  hours 
except  that  of  the  deer  mouse. 

No  attribute  measured  at  cavity  nests  dif- 
fered significantly  between  predated  and  non- 
predated nests  (Table  2,  P > 0.05  in  all  cases, 
P < 0.01  required  after  adjusting  for  multiple 
tests). 

DISCUSSION 

Our  results  are  consistent  with  other  studies 
using  artificial  nests  in  finding  a high  diversity 
of  nest  predators,  ranging  from  six  to  nine 
species  (Henry  1969,  Wilcove  1985,  Reitsma 
et  al.  1990,  Leimgruber  et  al.  1994).  We  iden- 
tified eight  species  at  open  nests,  and  two 
(possibly  three)  other  species  pecked  or 
pierced  eggs.  Pieman  and  Schriml  (1994) 
found  only  one  or  two  major  predator  species 
in  each  of  four  vegetation  types,  although 
predator  diversity  ranged  from  four  (marsh)  to 
nine  species  (scrubland  and  forest).  Lack  of 
independence  of  the  setups  in  their  study  may 
have  overestimated  the  importance  of  some 
predator  species  (see  below).  Interestingly,  all 


TABLE  2.  Nest  site  variables  and  results  of  /-tests  for  predated  and  nonpredated  cavity  nests  at  the  San 
Joaquin  Experimental  Range. 


Nonpredated  cavities  Predated  cavities 


Mean  (SE) 

n 

Mean  (SE) 

n 

pa 

Power 

Nest  height  (m) 

3.82  (0.24) 

35 

3.78  (0.23) 

32 

0.90 

0.98'’ 

Depth  (cm) 

25.1  (1.94) 

35 

31.0  (2.42) 

31 

0.06 

0.95” 

Vertical  entrance  diameter  (cm) 

5.68  (0.67) 

34 

5.06  (0.16) 

30 

0.37 

1.00*' 

Horizontal  entrance  diameter  (cm) 

5.19  (0.19) 

34 

5.14  (0.13) 

30 

0.83 

0.98“ 

Minimum  entrance  diameter  (cm) 

4.80  (0.15) 

34 

4.94  (0.14) 

30 

0.51 

0.99“ 

“ A P- value  of  0.010  is  needed  for  significance  at  a = 0.05  after  Bonferroni  adjustment  for  multiple  comparisons. 
Based  on  an  effect  size  of  1 m. 

Based  on  an  effect  size  of  0.5  cm. 

Based  on  an  effect  size  of  1.0  cm. 


254 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


but  one  (Acom  Woodpecker)  of  the  avian 
predators  at  open  nests  in  our  study  were  open 
nesters  and  all  avian  predators  at  cavity  nests 
were  cavity  nesters. 

We  know  of  only  two  efforts  to  study  pre- 
dation at  experimental  cavity  nests  (Wilcove 
1985,  Sandstrdm  1991),  neither  of  which 
identified  predators.  Although  some  species 
photographed  at  our  cavity  setups  may  have 
been  only  reconnoitering  potential  nest  sites, 
they  may  also  opportunistically  eat  eggs  thus 
encountered  in  a cavity.  The  importance  of 
birds  as  predators  at  active  cavity  nests,  de- 
fended by  nesting  birds,  is  unknown  but  prob- 
ably significant  in  some  instances.  European 
Starlings  are  known  to  usurp  nest  sites  from 
other  bird  species  (Troetschler  1976,  Ingold 
1989)  and,  although  the  contents  of  the  cavi- 
ties were  unknown,  eggs  were  likely  present 
in  some  when  usurpation  occurred  late  enough 
in  the  nesting  cycle  for  egg-laying  or  incu- 
bation to  have  begun.  Our  results  suggest  that 
starlings  probably  consume  eggs  in  the  course 
of  usurping  nest  sites. 

Although  several  species  of  corvids  are  be- 
lieved to  be  important  predators  at  open  nests 
(Yahner  and  Wright  1985,  Yahner  and  Scott 
1988,  Andren  1992,  Pieman  and  Schriml 
1994,  Hannon  and  Cotterill  1998),  nest  pre- 
dation at  open  nests  by  noncorvids  has  not 
often  been  documented.  Gates  and  Gysel 
(1978)  reported  anecdotal  evidence  of  nest 
predation  by  an  Eastern  Screech-Owl  {Otus 
asio).  Pieman  (1987)  photographed  Marsh 
Wrens  (Cistothorus  palustnis).  House  Wrens, 
and  Gray  Catbirds  (Diimetella  carolinensis) 
predating  nests.  Pieman  and  Schriml  (1994) 
recorded  predation  events  by  Broad-winged 
Hawks  (Buteo  platypterus).  Eastern  Mead- 
owlarks {Sturnella  magna),  and  Red-winged 
Blackbirds  (Agelaius  phoeniceus).  The  Marsh 
Wren  was  the  only  predator  recorded  at  nests 
of  Yellow-headed  Blackbirds  (Xanthocephal- 
us  xanthocephalus'.  Pieman  and  Isabelle 
1995).  Predation  of  open  nests  by  woodpeck- 
ers has  been  documented  rarely  (Bent  1939; 
Watt  1980;  Hernandez  et  al.  1998a,  b;  Robert 
Cooper,  unpubl.  data),  and  never  by  the  Acorn 
Woodpecker.  Egg  removal  by  Bullock’s  Ori- 
oles was  surprisingly  common  in  this  study. 
Both  sexes  of  this  oriole  are  known  to  be  ejec- 
tors of  Brown-headed  Cowbird  eggs,  and  they 
sometimes  consume  the  eggs  before  removing 


the  shells  (Sealy  and  Neudorf  1995).  Al- 
though some  of  these  “predators”  may  de- 
stroy nests  or  eggs  with  no  nutritional  moti- 
vation, predation  of  open  nests  by  noncorvids 
may  be  more  frequent  than  previously 
thought. 

Leimgruber  and  coworkers  (1994)  and  Pie- 
man and  Schriml  (1994)  also  found  that  po- 
tential predators  visiting  nests  did  not  always 
eat  the  eggs.  Consistent  with  the  findings  of 
Pieman  and  Schriml,  mammals  in  our  study 
nearly  always  took  the  egg.  Pieman  and 
Schriml  (1994)  classified  as  “accidental  visi- 
tors” several  bird  species  that  visited  nests, 
including  Red-winged  Blackbirds,  that  appar- 
ently ate  eggs  at  6 of  29  nests  visited.  Re- 
gardless of  motivation,  the  effect  of  egg  re- 
moval on  the  nesting  birds  is  the  same. 

Artificial  nest  studies  are  just  that — artifi- 
cial; some  biases  are  certainly  involved  (Mar- 
tin 1987,  Reitsma  et  al.  1990,  Whelan  et  al. 
1994,  Marini  and  Melo  1998,  Wilson  et  al. 
1998).  As  visual  predators,  birds  may  key  in 
on  cameras  or  unrealistic  nest  placements,  or 
they  may  follow  field  workers  and  leai'n  to 
associate  conspicuous  mai'kers  at  nests  with 
food  (Picozzi  1975,  Gotmark  1992).  At  open 
nests  we  were  able  to  attain  a realistic  nest 
placement  or  a good  camera  view,  but  usually 
not  both.  At  cavity  nests,  the  transmitters  and 
receivers  may  have  inhibited  predators  from 
going  to  a cavity  or  may  have  attracted  curi- 
ous predators.  Predation  also  may  have  been 
more  or  less  likely  at  artificial  nests  than  real 
nests  defended  by  adult  birds,  but  even  real 
nests  are  left  unattended  at  regular  intervals. 
In  spite  of  these  potential  biases,  we  feel  that 
useful  data  can  be  obtained  from  studies  of 
artificial  nests  because  they  establish  baseline 
data  in  an  arena  where  so  little  is  known. 

We  believe  that  the  value  of  studies  using 
artificial  nests  can  be  substantially  increased 
if  studies  are  designed  more  cai'efully  to  re- 
duce potential  biases.  For  example,  most  of 
the  studies  we  reviewed  used  distances  be- 
tween setups  ranging  from  20  to  60  m.  Such 
short  intervals  risk  detection  of  the  same  in- 
dividual predator  at  two  or  more  setups,  vio- 
lating assumptions  of  independence.  Ideally, 
the  distance  between  artificial  nests  should  ex- 
ceed the  largest  home-range  diameter  of  the 
suite  of  likely  predators.  One  might  argue  that 
shorter  distances  are  appropriate  for  examin- 


Purcell  ami  Verner  • NEST  PREDATORS 


255 


ing  predation  rates,  although  rates  from  arti- 
ficial nest  studies  are  of  questionable  value  for 
extrapolation  to  natural  conditions  (Martin 
1987,  Whelan  et  al.  1994,  Wilson  et  al.  1998). 
Replication  over  large  areas  is  required  to 
characterize  the  suite  of  predators  for  a given 
vegetation  type,  since  predators  are  often  un- 
evenly distributed  in  space  and  time. 

Further  problems  of  independence  may 
have  occurred  in  studies  that  replaced  eggs  in 
nests  that  had  been  predated  previously.  Nour 
and  coworkers  (1993)  suggested  that  such  egg 
replacement  may  not  be  a problem  in  studies 
using  plasticine  eggs  or  eggs  made  from  mod- 
eling clay  because  the  eggs  are  not  eaten  and 
provide  the  predator  no  incentive  to  return, 
although  predators  could  avoid  nests  with  clay 
eggs  because  of  prior  negative  conditioning. 

ACKNOWLEDGMENTS 

We  thank  L.  and  S.  Garner  for  instructions  to  con- 
struct the  camera  apparatus  to  photograph  nest  preda- 
tors at  open  nests.  We  were  aided  in  the  field  by  K. 
Kalin,  R.  Miller,  and,  especially,  by  D.  Cubanski  and 
J.  Ohanesian,  who  improved  the  open-nest  apparatus. 
The  manuscript  benefitted  from  reviews  by  S.  Hejl,  C. 
Maguire,  C.  Meslow,  W.  Laudenslayer,  L.  Reitsma,  C. 
Whelan,  and  three  anonymous  reviewers. 

LITERATURE  CITED 

Abramowitz,  M.  and  I.  A.  Stegun.  1964.  Handbook 
of  mathematical  functions.  Dover  Publications, 
New  York. 

Andren,  H.  1992.  Corvid  density  and  nest  predation 
in  relation  to  forest  fragmentation:  a landscape 
perspective.  Ecology  73:794-804. 

Baicich,  P.  J.  and  C.  J.  O.  Harrison.  1997.  A guide 
to  the  nests,  eggs,  and  nestlings  of  North  Ameri- 
can birds,  second  ed.  Academic  Press,  San  Diego, 
California. 

Bent,  A.  C.  1939.  Life  histories  of  North  American 
woodpeckers.  U.S.  Nat.  Mus.  Bull.  174:1—334. 
Best,  L.  B.  1978.  Field  Sparrow  reproductive  success 
and  nesting  ecology.  Auk  95:9-22. 

Best,  L.  B.  and  D.  E Staubter.  1980.  Factors  affect- 
ing nesting  success  in  riparian  bird  communities. 
Condor  82:149-158. 

Boag,  D.  a.,  S.  G.  Reebs,  and  M.  A.  Schroeder. 
1984.  Egg  loss  among  Spruce  Grouse  inhabiting 
lodgepole  pine  forests.  Can.  J.  Zool.  62:1034- 
1037. 

Brooks,  R.  T.  1996.  Assessment  of  two  camera-based 
systems  for  monitoring  arboreal  wildlife.  Wildl. 
Soc.  Bull.  24:298-300. 

Gates,  J.  E.  and  L.  W.  Gysel.  1978.  Avian  nest  dis- 
persion and  fledging  success  in  field-forest  eco- 
tones.  Ecology  59:871-883. 


Gotmark,  F.  1992.  The  effects  of  investigator  distur- 
bance on  nesting  birds.  Curr.  Ornithol.  9:63-104. 

Hannon,  S.  J.  and  S.  E.  Cotterii.l.  1998.  Nest  pre- 
dation in  aspen  woodlols  in  an  agricultural  area  in 
Alberta:  the  enemy  from  within.  Auk  I 15:16-25. 

Henry,  V.  G.  1969.  Predation  on  dummy  nests  of 
ground-nesting  birds  in  the  southern  Appala- 
chians. J.  Wildl.  Manage.  33:169-172. 

Hernandez,  E,  D.  Roluns,  and  R.  Cantu.  1998a. 
Evaluating  evidence  to  identify  ground-nest  pred- 
ators in  west  Texas.  Wildl.  Soc.  Bull.  25:826-831. 

Hernandez,  E,  D.  Rollins,  and  R.  Cantu.  1998b.  An 
evaluation  of  Trailmaster  camera  systems  for  iden- 
tifying ground-nest  predators.  Wildl.  Soc.  Bull. 
25:848-853. 

Kucera,  T.  E.  and  R.  H.  Barrett.  1993.  The  Trail- 
master  camera  system  for  detecting  wildlife. 
Wildl.  Soc.  Bull.  21:505-508. 

Lack,  D.  1954.  The  natural  regulation  of  animal  num- 
bers. Oxford  Univ.  Press,  London,  U.K. 

Lack,  D.  1968.  Ecological  adaptations  for  breeding  in 
birds.  Methuen,  London,  U.K. 

Leimgruber,  R,  W.  j.  McShea,  and  J.  H.  Rappole. 
1994.  Predation  on  artificial  nests  in  large  forest 
blocks.  J.  Wildl.  Manage.  58:254-260. 

Marini,  M.  A.  and  C.  Melo.  1998.  Predators  of  quail 
eggs,  and  the  evidence  of  the  remains:  implica- 
tions for  nest  predation  studies.  Condor  100:395- 
399. 

Martin,  T.  E.  1987.  Artificial  nest  experiments:  effects 
of  nest  appearance  and  type  of  predator.  Condor 
89:925-928. 

Martin,  T.  E.  1988a.  Processes  organizing  open-nest- 
ing bird  assemblages:  competition  or  nest  preda- 
tion? Evol.  Ecol.  2:37-50. 

Martin,  T.  E.  1988b.  On  the  advantage  of  being  dif- 
ferent: nest  predation  and  the  coexistence  of  bird 
species.  Proc.  Nat.  Acad.  Sci.,  USA  85:2196- 
2199. 

Martin,  T.  E.  1988c.  Nest  placement:  implications  for 
selected  life-history  traits,  with  special  reference 
to  clutch  size.  Am.  Nat.  132:900-910. 

Martin,  T.  E.  1995.  Avian  life  history  evolution  in 
relation  to  nest  sites,  nest  predation,  and  food. 
Ecol.  Monogr.  65:101-127. 

Nice,  M.  M.  1957.  Nesting  success  in  altricial  birds. 
Auk  74:305-321. 

Nour,  N.,  E.  Matthysen,  and  A.  A.  Dhondt.  1993. 
Artificial  nest  predation  and  habitat  fragmentation: 
different  trends  in  bird  and  mammal  predators. 
Ecography  16:111-116. 

PiCMAN,  J.  1987.  An  inexpensive  camera  setup  for  the 
study  of  egg  predation  at  artificial  nests.  J.  Field 
Ornithol.  58:372-382. 

PiCMAN,  J.  AND  A.  Isabelle.  1995.  Sources  of  mortality 
and  correlates  of  nesting  success  in  Yellow-head- 
ed Blackbirds.  Auk  112:183-191. 

PiCMAN,  J.  AND  L.  M.  ScHRiML.  1994.  A Camera  study 
of  temporal  patterns  of  nest  predation  in  different 
habitats.  Wilson  Bull.  106:456-465. 


256 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  2,  June  1999 


Picozzi,  N.  1975.  Crow  predation  on  marked  nests.  J. 
Wildl.  Manage.  39:151-155. 

Purcell,  K.  L.  1997.  Use  of  a fiberscope  for  exam- 
ining cavity  nests.  J.  Lield  Ornithol.  68:283-286. 

Reitsma,  L.  R.,  R.  T.  Holmes,  and  T.  W.  Sherry. 
1990.  Effects  of  removal  of  red  squirrels,  Tamias- 
ciurus  hudsonicus,  and  eastern  chipmunks,  Tam- 
ias  striatus,  on  nest  predation  in  a northern  hard- 
wood forest:  an  artificial  nest  experiment.  Oikos 
57:375-380. 

Ricklefs,  R.  E.  1969.  An  analysis  of  nesting  mortality 
in  birds.  Smithson.  Contrib.  Zool.  9:1-48. 

Sandstrom,  U.  1991.  Enhanced  predation  rates  on 
cavity  bird  nests  at  deciduous  forest  edges — an 
experimental  study.  Ornis  Fenn.  68:93-98. 

SAS  Institute.  1988.  SAS/STAT  user’s  guide:  statis- 
tics. Release  6.03.  SAS  Institute,  Inc.,  Cary,  North 
Carolinia. 

Sealy,  S.  G.  and  D.  L.  Neudorf.  1995.  Male  Northern 
Orioles  eject  cowbird  eggs:  implications  for  the 
evolution  of  rejection  behavior.  Condor  97:369— 
375. 

Watt,  D.  J.  1980.  Red-bellied  Woodpecker  predation 


on  nestling  American  Redstarts.  Wilson  Bull.  92: 
249. 

Whelan,  C.  J.,  M.  L.  Dilger,  D.  Robson,  N.  Hallyn, 
AND  S.  Dilger.  1994.  Effects  of  olfactory  cues  on 
artificial-nest  experiments.  Auk  111:945-952. 

Welcove,  D.  S.  1985.  Nest  predation  in  forest  tracts 
and  the  decline  of  migratory  songbirds.  Ecology 
66:1211-1214. 

Wilson,  G.  R.,  M.  C.  Brittingham,  and  L.  J.  Good- 
rich. 1998.  How  well  do  artificial  nests  estimate 
success  of  real  nests?  Condor  100:357—364. 

Wray,  T.  II,  K.  A.  Strait,  and  R.  C.  Whitmore.  1982. 
Reproductive  success  of  grassland  sparrows  on  a 
reclaimed  surface  mine  in  West  Virginia.  Auk  99: 
157-164. 

Yahner,  R.  H.  and  D.  P.  Scott.  1988.  Effects  of  forest 
fragmentation  on  depredation  of  artificial  nests.  J. 
Wildl.  Manage.  52:158-161. 

Yahner,  R.  H.  and  A.  L.  Wright.  1985.  Depredation 
on  artificial  ground  nests:  effects  of  edge  and  plot 
age.  J.  Wildl.  Manage.  49:508-513. 

Zimmerman,  J.  L.  1984.  Nest  predation  and  its  rela- 
tionship to  habitat  and  nest  density  in  Dickcissels. 
Condor  86:68-72. 


Short  Communications 


Wilson  Bull..  111(2).  1999,  pp.  257-261 

Juvenile  Marbled  Murrelet  Nurseries  and  the  Productivity  Index 

Katherine  J.  Kuletz'-^  and  John  F.  Piatt- 


ABSTRACT. — Late  summer  counts  of  juveniles  at 
sea  are  used  as  an  index  of  Marbled  Murrelet  (Bra- 
chyramplius  mannoratu.s)  reproductive  success,  but  lit- 
tle is  known  about  juvenile  dispersal  or  habitat  use. 
Further,  it  is  not  known  whether  these  counts  accu- 
rately rellect  absolute  breeding  success.  To  address 
these  questions  we  conducted  five  boat  surveys  for 
Marbled  Murrelets  and  Pigeon  Guillemots  (Cepphus 
coliimba)  in  Kachemak  Bay,  Alaska  between  7-24  Au- 
gust 1996.  Juvenile  murrelet  distribution  in  the  bay 
was  patchy,  and  we  identified  a juvenile  Marbled  Mur- 
relet ‘nursery’  area  in  the  outer  bay.  Fifty-three  of  61 
juvenile  murrelets  were  in  this  area,  whereas  after- 
hatch-year (AHY)  murrelets  were  dispersed  throughout 
the  bay,  as  were  juvenile  and  AHY  Pigeon  Guillemots. 
The  murrelet  nursery  was  characterized  by  water  in- 
side of  or  at  the  edge  of  a 20  m deep  contour,  semi- 
protected  seas,  productive  waters,  and  a large  bed  of 
Nereocystis  kelp.  Juveniles  comprised  16.1%  of  all 
murrelets  and  24.8%  of  all  guillemots  observed  at  sea. 
These  data  suggest  a maximum  reproductive  success 
of  0.32  chicks/pair  if  all  AHY  murrelets  were  breeding 
and  0.46  chicks/pair  if  only  70%  of  AHY  murrelets 
were  breeding.  For  guillemots,  maximum  productivity 
estimated  from  at-sea  counts  was  0.50  chicks/pair  if 
all  AHY  were  breeding  and  0.71  chicks/pair  if  only 
70%  were  breeding.  The  guillemot  estimate  was  sim- 
ilar to  that  obtained  by  concurrent  studies  at  nine  guil- 
lemot colonies  in  the  bay  (0.56  chicks/pair).  These  re- 
sults suggest  that  at  sea  surveys  in  late  summer  provide 
a reasonable  index  of  local  productivity  for  nearshore 
alcids.  Further,  if  murrelet  nursery  areas  can  be  found, 
at  sea  counts  may  provide  a valid  measure  of  absolute 
productivity.  Received  Jl  June  1998,  accepted  7 Jan. 
1999. 


Nests  of  the  Marbled  Murrelet  (Brachyram- 
phus  marmoratus)  are  difficult  to  find  or 
study,  and  reproductive  success  is  known  only 
from  widely  scattered  nests  studied  over  many 
years.  Because  of  the  murrelet’s  threatened 


' U.S.  Fish  and  Wildlife  Service,  1011  E.  Tudor  Rd., 
Anchorage,  AK  99503. 

-Alaska  Biological  Sciences  Center,  U.S.G.S.,  101 1 
E.  Tudor  Rd.,  Anchorage,  AK  99503. 

^ Corresponding  author; 

E-mail;  kathy_kuletz@fws.gov 


status  from  British  Columbia  to  California 
(Ralph  et  al.  1995),  considerable  effort  has 
been  devoted  to  finding  alternate  means  of  es- 
timating murrelet  reproductive  success.  The 
most  practical  approach  is  to  use  a productiv- 
ity index  based  on  surveys  at  sea,  which  uses 
the  ratio  of  juveniles  to  adults  or  juvenile  den- 
sities during  the  fledging  period  as  indices  of 
production  (Ralph  and  Long  1995,  Strong 
1995,  Kuletz  and  Kendall  1998).  To  be  ac- 
curate, surveys  require  some  knowledge  of 
fledgling  dispersal  at  sea,  but  little  is  known 
about  juvenile  movements  or  habitat  use. 

Anecdotal  evidence  suggests  that  juvenile 
Marbled  Murrelets  sometimes  congregate  in 
“nursery  areas”,  often  near  shore  or  in  exten- 
sive kelp  beds  (Sealy  1975,  Beissinger  1995, 
Strachan  et  al.  1995,  Strong  et  al.  1995).  If 
juveniles  gather  in  specific  habitats  after 
fledging,  productivity  surveys  could  be  im- 
proved by  identifying  their  location  and  time 
of  use.  Here,  we  report  on  a juvenile  murrelet 
nursery  and  describe  associated  habitat  fea- 
tures. We  estimate  muixelet  productivity  from 
the  ratio  of  juveniles  to  adults  at  sea,  and  com- 
pare this  with  Pigeon  Guillemot  {Cepphus  col- 
umba)  productivity  estimates  obtained  by  both 
counts  at  sea  and  local  colony  studies. 

STUDY  AREA  AND  METHODS 

We  conducted  surveys  in  Kachemak  Bay.  southcen- 
tral Alaska  on  five  days  between  7-24  August  1996 
(Eig.  1).  We  surveyed  the  south  side  of  Kachemak  Bay 
because  Marbled  Munelet  densities  are  highest  on  the 
south  side,  which  has  deep  water,  many  side  bays,  and 
a predominantly  rocky,  convoluted  shoreline  (Agler  et 
al.  1998). 

From  a 10  m vessel  we  counted  all  Marbled  Mur- 
relets and  Pigeon  Guillemots  within  100  m either  side 
of  the  boat.  Two  observers  used  8 X 42  and  10  X 50 
binoculars  to  identify  species  and  plumages.  Juvenile 
murrelets,  which  resemble  adults  in  basic  plumage, 
were  identified  using  characteristics  described  in  Cart- 
er and  Stein  (1995)  and  Kuletz  and  Kendall  (1998).  A 
third  person  entered  observations  into  a laptop  com- 
puter using  DLOG  (Ecological  Consulting  Inc.,  Port- 


257 


258 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


FIG.  1.  Survey  routes  (a-j)  in  Kachemak  Bay,  Alaska,  surveyed  by  boat  on  five  days  on  7-24  August  1996. 


land,  Oregon).  The  DLOG  data  entry  program  was 
linked  with  a Global  Positioning  System  and  every 
observation  had  an  associated  latitude  and  longitude. 
Survey  routes  followed  a path  parallel  to  shore.  For 
most  of  the  survey  we  used  radar  to  maintain  a dis- 
tance of  100  m from  shore.  In  rocky  or  shallow  sec- 
tions we  surveyed  outside  the  20  m depth  contour. 
From  the  head  of  the  bay  to  Glacier  Spit,  and  from 
Kasitsna  Bay  to  Seldovia  Bay,  we  also  surveyed  0.5- 
1.0  km  offshore  (Fig.  1).  The  vessel  traveled  at  speeds 
of  about  7 km/hr,  but  because  this  was  a reconnais- 
sance survey,  we  temporarily  paused  or  left  our  path 
to  observe  potential  juvenile  murrelets  or  guillemots 
(birds  in  black  and  white  plumages). 

We  surveyed  a linear  distance  of  214  km  on  10  dif- 


ferent survey  routes  over  five  days  for  a total  area  sur- 
veyed of  36.6  km-  (Table  1).  We  refer  to  the  area  from 
the  head  of  Kachemak  Bay  to  China  Foot  Bay  as  the 
inner  bay  and  the  area  west  of  China  Foot  to  Seldovia 
Bay  as  the  outer  bay.  Our  main  objective  was  to  de- 
scribe the  spatial  distribution  of  murrelets  during  the 
fledging  period,  but  we  obtained  some  temporal  cov- 
erage. Portions  of  the  survey  routes  overlapped  on  dif- 
ferent days  and  all  regions  of  the  bay  were  surveyed 
both  early  and  late  in  the  fledging  period  (Table  1). 
Survey  dates  (7-24  August)  encompassed  the  main 
and  peak  fledging  period  for  mumelets,  based  on  five 
replicate  surveys  conducted  independently  between  7 
August  and  4 September  1996  near  Kasitsna  Bay  by 
KJK  and  J.  Figurski.  These  dates  correspond  to  the 


TABLE  1.  Numbers  of  adult  (after-hatch-year)  and  juvenile  Marbled  Murrelets  and  Pigeon  Guillemots  ob- 
served on  survey  routes  in  Kachemak  Bay,  Alaska,  in  August  1996.  Area  (km-)  was  calculated  from  the  survey 
route  length  X width. 

Bay  area 

Survey 

route 

Date 

Area 

(kni2) 

No.  Marbled  Murrelets 

No.  Pigeon  Guillemots 

Adults 

Juveniles 

Adults 

Juveniles 

Inner 

a 

8-13 

2.86 

72 

0 

8 

4 

Inner 

h 

8-24 

3.44 

83 

1 

5 

16 

Inner 

c 

8-13 

3.64 

19 

0 

25 

3 

Inner 

cl 

8-24 

1.52 

12 

1 

10 

1 

Total 

1 1.46 

186 

2 

48 

24 

Outer 

e 

8-13 

2.89 

37 

I 

21 

1 

Outer 

f 

8-07 

5.25 

43 

2 

12 

1 

Outer 

8-07 

1.26 

7 

3 

94 

3 

Outer 

h 

8-12 

1.54 

3 

0 

0 

1 

Outer 

i 

8-12 

4.66 

7 

23 

0 

0 

Outer 

j 

8-23 

9.57 

34 

30 

12 

32 

Total 

25.17 

131 

59 

139 

38 

SHORT  COMMUNICATIONS 


259 


FIG.  2.  Distribution  of  adult  (after-hatch-year)  and  juvenile  Marbled  Murrelets  on  surveys  conducted  in 
Kachemak  Bay,  Alaska,  on  7-24  August  1996. 


fledging  period  in  nearby  Prince  William  Sound,  Alas- 
ka (Kuletz  and  Kendall  1998).  Pigeon  Guillemot  fledg- 
ing dates  were  similar  and  were  verified  from  local 
colony  studies  (Piatt  et  al.  1997). 

Because  we  wanted  to  describe  the  general  distri- 
bution of  murrelets  and  our  survey  routes  varied,  we 
pooled  all  bird  counts  for  a single  tally.  For  both  Mar- 
bled Murrelets  and  Pigeon  Guillemots  we  determined 
the  ratio  at  sea  of  juveniles  to  adults  and  subadults 
(after-hatching-year  birds;  AHY).  We  also  calculated 
an  index  of  juveniles/pair  based  on  counts  of  juveniles 
and  half  the  number  of  adults  counted  on  the  same 
surveys.  Piatt  and  coworkers  (1997)  obtained  detailed 
observations  of  Pigeon  Guillemots  on  60  nests  in  9 
colonies  distributed  along  the  south  shore  of  Kache- 
mak Bay  from  Glacier  Spit  to  Seldovia  Bay. 

RESULTS 

Fifty-nine  of  61  juvenile  Marbled  Murrelets 
were  found  in  outer  Kachemak  Bay  and  two 
were  found  in  the  inner  bay  (Table  1).  Most 
of  the  juveniles  in  the  outer  bay  were  concen- 
trated 0. 5-1.0  km  offshore,  near  the  mouth  of 
Seldovia  Bay  (Fig.  2).  This  area  has  an  exten- 
sive kelp  bed  {Nereocystis  sp.)  and  covers  an 
underwater  shelf  less  than  20  m deep.  Adult 
Marbled  Murrelets  (n  = 317;  5 in  basic  plum- 
age) were  distributed  throughout  Kachemak 
Bay,  with  highest  densities  in  the  inner  bay 
between  Glacier  Spit  and  the  bay  head  (Fig. 
2).  We  found  Pigeon  Guillemots,  both  adults 
(n  = 249)  and  juveniles  (n  = 62),  disti'ibuted 
throughout  Kachemak  Bay  (Table  1 ). 


Juveniles  represented  16.1%  of  all  Marbled 
Murrelets  and  24.8%  of  all  Pigeon  Guillemots 
counted.  If  all  of  the  AHY  Marbled  Murrelets 
were  breeding,  our  counts  suggest  a maximum 
reproductive  success  of  0.32  chicks/pair.  A 
more  conservative  estimate  is  that  only  70% 
of  AHY  birds  were  breeding  (Piatt  and  Ford 
1993),  and  therefore  maximum  productivity  is 
calculated  as  0.46  chicks/pair.  For  Pigeon 
Guillemots,  the  maximum  productivity  would 
be  0.50  chicks/pair  if  all  AHY  birds  were 
breeding,  and  0.71  chick/pair  if  only  70% 
were  breeding  adults. 

DISCUSSION 

Juvenile  Marbled  Munelets  in  Kachemak 
Bay  showed  a clear  preference  for  the  kelp 
beds  approximately  4 km  on  either  side  of  the 
mouth  of  Seldovia  Bay.  This  distribution  con- 
trasts shaiply  with  the  distribution  of  adult 
mun-elets  found  throughout  the  bay.  Adult 
munelets  forage  on  Pacific  sand  lance  (Am- 
modytes  hexaptenis)  in  the  inner  bay  (J.  Piatt, 
unpubl.  data),  suggesting  that  the  distribution 
of  forage  fish  was  not  limiting  the  distribution 
of  juvenile  munelets. 

Why  were  juvenile  munelets  concentrated 
along  the  shore  in  outer  Kachemak  Bay  and 
in  extensive,  dense  beds  of  Nereocystis  kelp? 
Although  exposed  relative  to  the  inner  bay, 
the  orientation  of  the  shoreline  in  this  area 


260 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


provided  protection  from  prevailing  south- 
westerly winds.  The  southwest  portion  of  Ka- 
chemak  Bay  receives  up  welled  waters  from 
the  Alaska  Coastal  Current  entering  Cook  In- 
let from  the  southeast,  and  gyres  in  the  outer 
bay  retain  nutrients  and  promote  high  local 
productivity  (Trasky  et  al.  1977).  The  pres- 
ence of  Nereocystis,  which  attach  to  rocky 
substrate  and  grow  in  water  20-40  m deep 
where  fast  currents  or  upwelling  occurs,  is  of- 
ten associated  with  productive  waters  (Lalli 
and  Parsons  1993).  Thus,  shallow  water,  semi- 
protected  seas,  the  presence  of  kelp,  and  lo- 
cally productive  waters  appear  to  combine 
here  to  create  a favorable  nursery  area  for 
newly-fledged  murrelets.  In  addition,  the  kelp 
made  it  difficult  to  see  the  juveniles,  and  so 
may  provide  protection  from  avian  predators 
such  as  gulls  and  Bald  Eagles  {Haliaeetus  leii- 
cocephalus),  which  are  common  in  this  area. 
Large  Nereocystis  kelp  beds  are  not  common 
elsewhere  in  Kachemak  Bay  so  this  feature 
may  be  the  primary  defining  characteristic  of 
the  nursery. 

Juvenile  murrelets  may  use  the  inner  bay 
temporarily  after  fledging,  and  if  fledging 
peaked  early  in  August  1996,  it  is  possible 
that  we  missed  seeing  them  before  they  emi- 
grated to  the  outer  bay.  It  is  also  possible  that 
juveniles  were  absent  from  the  inner  bay  be- 
cause few  murrelets  may  breed  there  now  as 
the  result  of  extensive  damage  to  mature  for- 
ests from  spruce  beetle  (Dendroctonus  nifi- 
pennis).  However,  the  middle  portion  of  the 
bay  (China  Foot  to  Kasitsna  Bay)  still  has 
largely  intact  forests,  and  while  the  inner  bay 
is  clearly  an  important  foraging  area  for 
adults,  most  juveniles  were  found  in  the  outer 
bay.  The  use  of  kelp  beds  in  the  outer  bay  by 
juvenile  murrelets  appears  to  be  a recurring 
event;  we  have  observed  juvenile  murrelets  in 
this  area  in  previous  years.  Surveys  of  the  en- 
tire bay  throughout  the  fledging  period  would 
be  necessary  to  determine  whether,  and  if  so 
when,  juveniles  from  throughout  the  bay 
move  to  the  kelp  beds. 

Estimates  of  Pigeon  Guillemot  productivity 
obtained  from  juvenile  surveys  at  sea  com- 
pared well  to  the  productivity  of  guillemots 
measured  from  colony-based  reproductive 
studies.  Pigeon  Guillemots  at  nine  Kachemak 
Bay  colonies  in  1996  produced  0.56  chicks/ 
pair,  which  falls  within  the  range  we  estimated 


from  counts  at  sea.  The  estimate  of  production 
we  obtained  for  Marbled  Murrelets  also  ap- 
proximates that  found  for  Marbled  Murrelets 
throughout  their  range  (0.28  chicks/pair), 
based  on  32  nests  followed  to  completion 
(Nelson  and  Hamer  1995).  It  is  noteworthy 
that  our  estimate  of  murrelet  production  in 
Kachemak  Bay  is  much  higher  than  those  cal- 
culated from  surveys  at  sea  in  areas  south  of 
Alaska  (e.g.,  0.001-0.1 1 chicks/pair),  even  af- 
ter adjustments  (0.01-0.17  chicks/pair)  for  the 
timing  of  surveys  (Beissinger  1995).  This  is 
undoubtedly  because  we  located  the  nursery 
area  near  Seldovia  Bay,  which  accounted  for 
53  of  61  juveniles  we  observed  on  surveys. 

While  the  possibility  of  juvenile  murrelet 
nurseries  has  been  suggested  in  some  areas 
(Sealy  1975,  Strachan  et  al.  1995),  they  have 
never  been  documented,  and  munelet  distri- 
bution may  not  always  be  as  patchy  as  it  ap- 
pears to  be  in  Kachemak  Bay.  In  southeast 
Alaska,  VanVliet  (pers.  comm.)  observed  ju- 
venile murrelets  clustered  near  or  in  kelp  beds 
in  late  August  in  discrete  areas  of  Port  Al- 
thorp,  whereas  adults  were  distributed  from 
Inian  Pass  to  Icy  Strait.  In  Prince  William 
Sound,  Alaska,  however,  juvenile  murrelets 
were  evenly  dispersed  in  nearshore  waters 
(relative  to  local  murrelet  abundance),  with 
the  exception  of  highly  exposed  shoreline 
where  they  were  absent  (Kuletz  et  al.  1997). 
The  areas  surveyed  in  Prince  William  Sound 
did  not  have  large  kelp  beds  and  were  char- 
acterized by  convoluted,  rocky  shorelines  with 
numerous  protected  bays  and  coves.  In  addi- 
tion, the  juveniles  in  Prince  William  Sound 
may  not  travel  fai'  in  the  first  two  weeks  after 
fledging  (Kuletz  and  Marks  1997;  Kuletz,  un- 
publ.  data). 

These  results  confirm  that  surveys  at  sea 
provide  a reasonable  index  of  productivity  for 
nearshore  seabirds  such  as  Pigeon  Guillemots 
and  Marbled  Murrelets.  However,  it  is  impor- 
tant to  determine  the  post-fledging  movements 
of  adults  and  juveniles  for  any  given  area  of 
study  because  adultrjuvenile  ratios  are  sensi- 
tive to  late  summer  movements  of  adults  and 
subadults  (Kuletz  and  Kendall  1998).  Al- 
though our  temporal  data  were  limited  in  this 
study,  we  did  not  find  obvious  declines  in 
adult  numbers  in  August,  such  as  occurs  in 
Prince  William  Sound.  If  juvenile  munelet 
nurseries  can  be  located,  it  would  facilitate  the 


SHORT  COMMUNICATIONS 


261 


use  of  juvenile  densities  to  measure  produc- 
tivity and  thus  avoid  problems  associated  with 
using  adult:juvenile  ratios  (Kuletz  and  Ken- 
dall 1998).  If  adults  remain  in  an  area  during 
the  main  fledging  period  where  muirelet  nurs- 
eries exist,  surveys  at  sea  may  provide  a valid 
measure  of  absolute  productivity,  and  not  just 
an  index  of  production  (e.g.,  Beissinger  1995). 

ACKNOWLEDGMENTS 

This  work  was  conducted  with  funding  from  the 
Exxon  Valdez  Oil  Spill  Trustee  Council,  the  U.S.  Geo- 
logical Survey,  and  the  U.S.  Fish  and  Wildlife  Service. 
We  are  grateful  to  the  Alaska  Maritime  National  Wild- 
life Refuge  for  logistic  support,  to  B.  Keitt  and  J.  Fi- 
gurski  for  assistance  with  surveys,  and  to  M.  Litzow 
for  use  of  unpublished  data  on  Pigeon  Guillemot  pro- 
ductivity. 

LITERATURE  CITED 

Agler,  B.  a.,  S.  J.  Kendall,  and  D.  B.  Irons.  1998. 
Abundance  and  distribution  of  Marbled  and  Kitt- 
litz’s  murrelets  in  southcentral  and  southeast  Alas- 
ka. Condor  100:254-265. 

Beissinger,  S.  R.  1995.  Population  trends  of  the  Mar- 
bled Murrelet  projected  from  demographic  anal- 
yses. USDA  For.  Serv.  Gen.  Tech.  Rep.  PSW- 
GTR- 152:385-394. 

Carter,  H.  R.  and  J.  L.  Stein.  1995.  Molts  and  plum- 
ages in  the  annual  cycle  of  the  Marbled  Murrelet. 
USDA  For.  Serv.  Gen.  Tech.  Rep.  PSW-GTR-152: 
99-112. 

Kuletz,  K.  J.,  S.  Kendall,  and  D.  Nigro.  1997.  Rel- 
ative abundance  of  adult  and  juvenile  Marbled 
Murrelets  in  Prince  William  Sound,  Alaska:  de- 
veloping a productivity  index.  Exxon  Valdez  Oil 
Spill  Restoration  Proj.  Final  Rep.  95031.  U.S.  Fish 
and  Wildl.  Serv.,  Anchorage,  Alaska. 

Kuletz,  K.  J.  and  D.  K.  Marks.  1997.  Post-fledging 
behavior  of  a radio-tagged  juvenile  Marbled  Mur- 
relet. J.  Field  Ornithol.  68:421-425. 

Kuletz,  K.  J.  and  S.  J.  Kendall.  1998.  A productiv- 


ity index  for  Marbled  Murrelets  in  Alaska  based 
on  surveys  at  sea.  J.  Wildl.  Manage.  62:446-460. 
Lalli,  C.  M.  and  T.  R.  Parsons.  1993.  Biological 
oceanography:  an  introduction.  Pergamon  Press, 
New  York. 

Nelson,  S.  K.  and  T.  E.  Hamer.  1995.  Nest  success 
and  the  effects  of  predation  on  Marbled  Murrelets. 
USDA  For.  Serv.  Gen.  Tech.  Rep.  PSW-GTR-152: 
89-98. 

Piatt,  J.  F.  and  R.  G.  Ford.  1993.  Distribution  and 
abundance  of  Marbled  Murrelets  in  Alaska.  Con- 
dor 95:662—669. 

Piatt,  J.  F,  M.  Robards,  S.  Zador,  M.  Litzow,  and 
G.  Drew.  1997.  Cook  Inlet  seabirds  and  forage 
fish  studies.  Exxon  Valdez  Oil  Spill  Restoration 
Proj.  Ann.  Rep.  96163M,  Biol.  Resour.  Div.,  U.S. 
Geological  Survey,  Anchorage,  Alaska. 

Ralph,  C.  J.,  G.  L.  Hunt,  Jr.,  M.  G.  Raphael,  and  J. 
F.  Piatt.  1995.  Ecology  and  conservation  of  the 
Marbled  Murrelet.  U.S.  For.  Serv.  Gen.  Tech. 
Rep.,  PSW-GTR-152: 1-5. 

Ralph,  C.  J.  and  L.  L.  Long.  1995.  Productivity  of 
Marbled  Murrelets  in  California  from  observa- 
tions of  young  at  sea.  USDA  For.  Serv.  Gen.  Tech. 
Rep.  PSW-GTR-152:371-380. 

Sealy,  S.  G.  1975.  Aspects  of  the  breeding  biology 
of  the  Marbled  Murrelet  in  British  Columbia. 
Bird-Banding  46:141-154. 

Strachan,  G.,  M.  McAllister,  and  C.  J.  Ralph. 
1995.  Marbled  Murrelet  at-sea  and  foraging  be- 
havior. USDA  For.  Serv.  Gen.  Tech.  Rep.  PSW- 
GTR-152:247-253. 

Strong,  C.  S.  1995.  Distribution  of  Marbled  Murre- 
lets along  the  Oregon  coast  in  1992.  Northwest. 
Nat.  76:99-105. 

Strong,  C.  S.,  B.  S.  Keitt,  W.  R.  McIver,  C.  J.  Palm- 
er, AND  I.  Gaffney.  1995.  Distribution  and  pop- 
ulation estimates  of  Marbled  Murrelets  at  sea  in 
Oregon  during  the  summers  of  1992  and  1993. 
U.S.  For.  Serv.  Gen.  Tech.  Rep.,  PSW-GTR-152: 
339-352. 

Trasky,  L.  L.,  L.  B.  Flagg,  and  D.C.  Burbank.  1977. 
Environmental  studies  of  Kachemak  Bay  and 
Lower  Cook  Inlet.  Vol.  1.  Impact  of  oil  on  the 
Kachemak  Bay  environment.  Alaska  Dept,  of  Fish 
and  Game,  Marine  Coastal  Habitat  Management, 
Anchorage,  Alaska. 


262 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


Wilson  Bull.,  111(2),  1999,  pp.  262-265 


“Snorkeling”  by  the  Chicks  of  the  Wattled  Jacana 

Carlos  Bosque'-^  and  Emilio  A.  Herrera- 


ABSTRACT — The  chicks  of  the  Wattled  Jacana 
(Jacana  jacana)  exhibited  an  unusual  predator  escape 
behavior  in  the  floodplains  of  Venezuela.  When  ap- 
proached by  a human,  chicks  dove  and  remained  com- 
pletely immobile  while  entirely  submerged  with  just 
the  beak  protruding  vertically  above  the  water  line. 
Since  breathing  should  continue  while  hiding,  we  rea- 
soned that  it  would  be  advantageous  for  the  bird  to 
have  the  nostrils  placed  in  a forward  position  along 
the  bill  to  facilitate  breathing  while  submerged.  To  ex- 
amine this  expectation  we  compared  the  relative  po- 
sition of  the  nostrils  of  the  Wattled  Jacana  with  those 
of  species  belonging  to  phylogenetically  related  Scol- 
opacidae.  In  accordance  with  expectations.  Wattled  Ja- 
canas  have  nostrils  that  are  placed  significantly  more 
forward  along  the  bill  than  all  species  of  Scolopacidae 
measured.  Nostril  placement  in  species  belonging  to 
other  phylogenetically  related  families,  Thinocoridae, 
Pedionomidae,  and  Rostratulide  is  also  basal  as  in 
Scolopacidae.  Lorward  placement  of  nostrils  seems  to 
be  a derived  character  in  Jacanas.  The  “snorkeling” 
behavior  of  Wattled  Jacana  chicks  is  a behaviorally 
elaborate  predator  escape  mechanism,  seemingly  ac- 
companied by  anatomical  adaptations.  Received  16 
June  1998,  accepted  15  Dec.  1998. 


Predation  is  often  the  major  mortality  factor 
of  young  birds  and  is  therefore  an  important 
selective  force  shaping  the  behavior  of  young 
birds  and  their  parents  (Rickleffs  1969,  Martin 
1992).  Capabilities  of  precocial  chicks  to  es- 
cape predation  or  defend  themselves  from 
predators  are  limited.  Their  chances  of  pre- 
dation are  largely  dependent  on  actions  taken 
by  their  parents.  When  predators  approach 
their  chicks,  adults  frequently  perform  “dis- 
traction displays”  that  often  divert  the  atten- 
tion of  the  predator  away  from  their  offspring 
(Skutch  1976).  Safety  of  precocial  chicks  also 
depends  upon  their  small  size,  concealing  col- 
oration and  immobility  (Skutch  1976).  In  this 
note  we  describe  an  unusual  predator  escape 


' Dept.  Biologia  de  Organismos,  Univ.  Simon  Bo- 
h'var.  Aparlado  89. ()()(),  Caracas  1080,  Venezuela. 

- Dept.  Estudios  Ambientales,  Univ.  Simdn  Bohvar, 
Apartado  89.000,  Caracas  1080,  Venezuela. 

’ Corresponding  author:  E-mail;  caiiosb@usb.ve 


behavior  exhibited  by  Wattled  Jacana  {Jacana 
jacana)  chicks  and  extend  information  on  the 
hiding  behavior  of  Jacana  chicks. 

The  Wattled  Jacana  inhabits  freshwater 
wetlands  with  floating  and  emergent  vegeta- 
tion throughout  its  range  from  northern  South 
America  and  Panama  to  central  Argentina  (del 
Hoyo  et  al.  1996).  Like  several  other  Jacanas, 
Wattled  Jacanas  show  sex-role  reversal  and  a 
polyandrous  mating  system  (Osborne  1982). 
Although  few  details  have  been  reported  on 
Wattled  Jacanas  (see  Osborne  and  Bourne 
1977,  Osborne  1982),  it  is  generally  assumed 
to  have  similai'  behavioral  traits  to  the  North- 
ern Jacana  (7.  spinosa;  del  Hoyo  et  al.  1996). 
Chicks  are  highly  precocious,  leaving  the  nest 
soon  after  hatching,  but  ai'e  tended  by  the 
male  parent.  Parental  care  includes  brooding, 
attending  and  defending,  but  not  feeding  the 
chicks  (Jenni  and  Collier  1972,  Stephens 
1984a,  Betts  and  Jenni  1991).  Females  aid 
males  in  defending  offspring  from  potential 
predators  or  conspecifics  (Jenni  and  Collier 
1972,  Stephens  1984a).  Heavy  predation  pres- 
sure on  eggs  and  chicks  appears  to  be  impor- 
tant in  both  Wattled  and  Northern  Jacanas 
(Jenni  1974;  Osborne  and  Bourne  1977;  Ste- 
phens 1984a,  b). 

Our  observations  were  made  on  a savanna 
flooded  approximately  40  cm  deep,  covered 
with  floating  and  emergent  vegetation  (Eic- 
chornia  sp.,  Hymenachne  amplexicaiilis)  at 
Hato  El  Frio,  a cattle  ranch  and  biological  re- 
serve in  the  southern  Llanos  (floodplains)  of 
Venezuela  (7°  46'  N,  68°  57'  W).  As  E.H.  ap- 
proached a pair  of  Jacana  adults  with  three 
chicks,  one  of  the  adults,  presumably  the 
male,  performed  a typical  “broken  wing”  dis- 
play, Jumping  and  apparently  attempting  un- 
successfully to  fly.  As  the  observer  continued 
to  approach,  the  parents  flew  away.  When  we 
looked  among  the  vegetation  for  the  chicks 
they  were  nowhere  to  be  seen.  While  search- 
ing, we  found  that  what  seemed  to  be  an  odd 
looking  slender  yellowish  flower  was  in  fact 


SHORT  COMMUNICATIONS 


263 


the  bill  of  one  of  the  young  jacanas.  The  chick 
remained  completely  immobile  while  entirely 
submerged  with  just  the  beak  protruding  ver- 
tically above  the  water.  By  bringing  a hand 
from  below  we  were  able  to  pick  up  the  bird, 
which  made  no  attempt  to  escape  nor  showed 
any  defense  behavior.  The  chick  was  in  the 
downy  stage. 

This  hiding  behavior  should  reduce  the 
chances  of  detection  by  predators  or  aggres- 
sive conspecifics.  For  it  to  be  effective,  the 
chick  should  remain  motionless  underwater 
for  an  unpredictable  length  of  time.  In  order 
to  continue  breathing  while  submerged  it 
would  be  advantageous  to  have  the  nostrils 
placed  toward  the  tip  of  the  bill. 

To  examine  this  prediction,  we  compared 
the  relative  position  of  the  nostrils  of  the  Wat- 
tled Jacana  with  those  of  phylogenetically  re- 
lated species.  Of  the  four  other  families  in  the 
same  parvorder  (Sibley  and  Monroe  1990) 
only  Scolopacidae  occur  in  Venezuela  and 
specimens  were  available  in  bird  collections. 
We  selected  at  random  6 of  the  12  genera  of 
Scolopacidae  that  occur  in  the  country  (Meyer 
de  Schauensee  and  Phelps  1978)  and  one  spe- 
cies from  each  of  these,  except  for  speciose 
genus  Calidris  from  which  we  chose  two  spe- 
cies. In  10  individuals  of  each  species  we 
measured  the  length  of  the  exposed  culmen 
(EC)  and  the  distance  between  the  posterior 
margin  of  the  right  nostril  and  the  tip  of  the 
bill  (NT).  Specimens,  selected  at  random, 
were  measured  at  the  Phelps  Ornithological 
Museum  in  Caracas  with  calipers  to  0.1  mm. 
For  jacanas  we  measured  separately  adults,  ju- 
veniles, and  the  one  downy  chick  in  the  col- 
lection. The  exposed  culmen  was  measured 
from  the  base  of  the  frontal  shield  where  it 
rises  more  abruptly.  From  those  measurements 
we  calculated  an  index  (I)  to  describe  the  rel- 
ative position  of  the  nostrils  along  the  bill: 

I = (EC  - NT)/EC 

The  value  of  this  index  should  be  zero  in 
those  species  in  which  the  nostrils  are  placed 
at  the  base  of  the  bill  and  closer  to  one  the 
nearer  the  nostrils  are  to  the  bill  tip.  Statistical 
analyses  were  done  with  SYSTAT  7.0  for 
Windows  (Wilkinson  1997). 

Mean  relative  position  of  the  nostrils  dif- 
fered between  species  (Fig.  1;  Single  Factor 
ANOVA  on  the  arcsine-transformed  data:  P 


0.4 


0.3 


2 0.2 
Q 


■ f S ‘ 

0.0  — — 

CH  GG  CF  NP  At  CS  BL  JJa  JJj  JJc 
SPECIES 

FIG.  1.  A comparison  of  the  relative  position  of 
the  nostrils  along  the  bill  between  Jacano  jacana  and 
several  species  of  Scolopacidae.  Mean  Index  values  (± 
SD),  see  text  for  definition  of  the  Index.  Species  are: 
CH  = Calidris  himantopiis,  GG  = GaUinago  gallin- 
ago,  CF  = Calidris  fuscicollis,  NP  = Numenius  phaeo- 
pus,  AI  = Arenaria  interpres,  CS  = Catoptrophorus 
semipalmatus,  BL  = Bartramia  longicauda,  JJa  = Ja- 
cana jacana  (adults),  JJj  = J.  jacana  (juveniles),  JJc 
= J.  jacana  chick,  n = 10  in  all  cases,  except  for  the 
single  J.  jacana  chick.  Horizontal  lines  above  species 
labels  indicate  similarity  between  the  index  mean  of 
those  species  (from  Tukey  HSD  test).  Jacana  jacana 
juveniles  and  the  chick  were  not  included  in  the  test. 

< 0.001,  F = 143.632,  df  = 7);  the  single 
jacana  chick  and  juveniles  were  not  included 
in  this  comparison.  A Tukey  HSD  a posteriori 
test  revealed  a number  of  significant  differ- 
ences between  species  {P  < 0.05;  Fig.  1). 
Wattled  Jacana  nostrils  are  placed  significant- 
ly more  forward  along  the  bill  than  all  species 
of  Scolopacidae  measured  (Fig.  1 ).  Nostrils 
were  also  placed  forward  on  juveniles  and  one 
newly  hatched  downy  chick. 

Nostril  placement  in  species  belonging  to 
families  phylogenetically  related  to  the  Jacan- 
idae,  but  not  available  to  us,  can  be  seen  in 
photographs  in  del  Hoyo  and  coworkers 
(1996).  Both  species  of  Rostratulidae  (belong- 
ing in  the  same  superfamily  with  Jacanidae) 
have  nostrils  placed  basally  (del  Hoyo  et  al. 
1996:293-299).  In  the  Plains-wanderer  {Pe- 
dionomus  torquatus,  Pedionomidae),  the  pos- 
terior end  of  its  longish  naiina  is  clearly  set 


264 


THE  WILSON  BULLETIN 


Vol.  Ill,  No.  2,  June  1999 


back  (del  Hoyo  et  al.  1996:535-536).  The  cor- 
naceous  flaps  covering  the  nostrils  of  the 
Thinocopridae  {Thinocorus  spp.,  Attagis  gayi) 
appear  to  be  placed  at  the  base  of  the  bill  also 
(del  Hoyo  et  al.  1996:539-543). 

It  appears  that  basal  placement  of  the  nos- 
trils is  ancestral  in  the  group  and  that  forward 
displacement  in  jacanas  is  a derived  character. 
We  cannot  assert  that  breathing  while  sub- 
merged was  an  important  selective  force  in  the 
forward  displacement  of  jacana  nostrils,  but 
their  current  position  should  facilitate  it. 

Young  of  jacanas  are  known  to  avoid  pred- 
ators or  aggressive  conspecifics  by  taking  to 
water  and  snorkeling  behavior  is  shared  with 
other  species.  The  chicks  of  the  Northern  Ja- 
cana, which  is  considered  to  form  a superspe- 
cies with  the  Wattled  Jacana  (del  Hoyo  et  al. 
1996),  have  been  reported  to  swim  (Gilliard 
1967),  to  hide  in  the  water  (Stephens  1984a), 
and  to  submerge  with  only  their  bills  and  the 
tops  of  their  heads  showing  above  the  surface 
(Miller  1931).  The  hiding  behavior  of  young 
Pheasent-tailed  Jacanas  {Hydrophasianus  chi- 
rurgus)  from  India  and  Asia  seems  similar; 
chicks  “may  freeze  while  hiding  under  a leaf 
or  even  while  completely  submerged  except 
for  the  bill”  (Johnsgard  1981:40).  The  chicks, 
and  sometimes  even  the  adults,  of  the  Lesser 
Jacana  {Microparra  capensis)  of  Africa  are 
also  known  to  submerge  with  only  their  bills 
out  of  the  water  (Maclean  1972).  Since  snor- 
keling seems  to  be  accompanied  by  anatomi- 
cal adaptations,  it  is  likely  that  those  species 
have  forward  displaced  nostrils  also.  In  fact, 
this  seems  to  be  the  case  in  the  Northern  (del 
Hoyo  et  al.  1996:277)  and  Pheasent-tailed  Ja- 
cana (drawing  in  del  Hoyo  et  al.  1996:288). 
For  the  Lesser  Jacana  it  is  difficult  to  judge 
from  available  pictures.  Other  species  of  ja- 
cana for  which  we  do  not  have  information  on 
snorkeling  behavior  also  have  nostrils  placed 
in  forward  position  along  the  bill,  most  no- 
tably the  AWcan  Jacana  (Actophilornis  afri- 
canus-,  del  Hoyo  et  al.  1996:281).  A broader 
survey  would  be  necessary  to  establish  if 
snorkeling  and  forward  placement  of  the  nos- 
trils is  shared  by  all  species  of  the  family. 

In  similar  aquatic  habitats  in  the  central 
Llanos  of  Venezuela,  the  young  of  at  least 
three  other  species  swim  or  dive,  although 
adults  do  not:  the  altricial  nestlings  of  Hoa- 
tzins  (Opisthocomus  hoazin-,  Strahl  1987)  and 


Greater  Anis  {Crotophaga  major,  Lau  et  al., 
in  press)  jump  from  their  nest  and  swim  or 
dive  in  the  water  below;  the  precocial  chicks 
of  Purple  Gallinules  {Porphirula  martinica) 
dive  when  threatened  (Zaida  Tarano,  pers. 
comm.).  Interestingly,  the  latter  and  conge- 
neric P.  flavirostris  have  nostrils  displaced 
forward  along  the  bill  (pers.  obs.). 

ACKNOWLEDGMENTS 

We  gratefully  acknowledge  Dr.  I.  D.  Maldonado  for 
his  hospitality  and  permission  to  work  at  Hato  El  Erio. 
M.  Lentino,  curator  of  the  Phelps  Ornithological  Mu- 
seum, allowed  us  to  measure  specimens  and  made 
helpful  suggestions.  I.  Carreno  helped  measure  speci- 
mens. Reviewers  and  editorial  staff  of  The  Wilson  Bul- 
letin improved  the  manuscript. 

LITERATURE  CITED 

Betts,  B.  J.  and  D.  A.  Jenni.  1991.  Time  budgets  and 
the  adaptaviness  of  polyandry  in  Northern  Jaca- 
nas. Wilson  Bull.  103:578-597. 
del  Hoyo,  J.,  A.  Elliot,  and  J.  Sargatal  (Eds.). 
1996.  Handbook  of  the  birds  of  the  world.  Vol. 
3.  Lynx  Editions,  Barcelona,  Spain. 

Gilliard,  E.  T.  1967.  Living  birds  of  the  world.  Dou- 
bleday & Company,  Inc.,  Garden  City,  New  York. 
Jenni,  D.  A.  1974.  Evolution  of  polyandry  in  birds. 
Am.  Zool.  14:129-144. 

Jenni,  D.  A.  and  G.  Collier.  1972.  Polyandry  in  the 
American  Jacana  {Jacana  spinosa).  Auk  89:743- 
765. 

Johnsgard,  P.  A.  1981.  The  plovers,  sandpipers  and 
snipes  of  the  world.  Univ.  of  Nebraska,  Lincoln. 
Lau,  P.  a.,  C.  Bosque,  and  S.  Strahl.  In  press.  Nest 
predation  in  relation  to  nest  placement  in  the 
Greater  Ani  {Crotophaga  major).  Ornitol.  Neo- 
trop.  In  press. 

Maclean,  G.  1972.  Waders  of  waterside  vegetation. 

African  Wildl.  26:163-167. 

Martin,  T.  E.  1992.  Interaction  of  nest  predation  and 
food  limitation  in  reproductive  strategies.  Curr. 
Ornithol.  9:163—197. 

Meyer  De  Schauensee,  R.  and  W.  H.  Phelps,  Jr. 
1978.  A guide  to  the  birds  of  Venezuela.  Prince- 
ton Univ.  Press,  Princeton,  New  Jersey. 

Miller,  A.  H.  1931.  Observations  on  the  incubation 
and  the  care  of  the  young  in  the  jacana.  Condor 
33:32-33. 

Osborne,  D.  R.  1982.  Replacement  nesting  and  poly- 
andry in  the  Wattled  Jacana.  Wilson  Bull.  94:206— 
208. 

Osborne,  D.  R.  and  G.  R.  Bourne.  1977.  Breeding 
behavior  and  food  habits  of  the  Wattled  Jacana. 
Condor  79:98-105. 

Rickleffs,  R.  E.  1969.  An  analysis  of  nesting  mor- 
tality in  birds.  Smithson.  Contrib.  Zool.  9:1-48. 


SHORT  COMMUNICATIONS 


265 


Sibley,  C.  G.  and  B.  L.  Monroe,  Jr.  1990.  Distri- 
bution and  taxonomy  of  birds  of  the  world.  Yale 
Univ.  Press,  New  Haven,  Connecticut. 

Skutch,  a.  F.  1976.  Parent  birds  and  their  young. 
Univ.  of  Texas  Press,  Austin. 

Stephens,  M.  1984a.  Interspecific  aggresive  behavior 
of  the  polyandrous  Northern  Jacana  (Jacana  spi- 
nosa).  Auk  101:508-518. 


Stephens,  M.  1984b.  Intraspecific  distraction  displays 
of  the  polyandrous  Northern  Jacana,  Jacana  spi- 
nosa.  Ibis  126:70-72. 

Strahl,  S.  D.  1987.  The  social  organization  and  be- 
haviour of  the  Hoatzin  Opisihocomus  hoazin  in 
central  Venezuela.  Ibis  130:483-502. 

Wilkinson,  L.  1997.  SYSTAT  7.0:  the  system  for  sta- 
tistics. SYSTAT  Inc.,  Evanston,  Illinois. 


Wilson  Bull.,  1 1 1(2),  1999,  pp.  265-268 


Rapid  Long-distance  Colonization  of  Lake  Gatun,  Panama, 

by  Snail  Kites 

George  R.  Angehr' 


ABSTRACT. — The  distribution  of  the  Snail  Kite 
(Rostrhamus  sociabilis)  is  closely  tied  to  that  of  apple 
snails  (Pomacea  spp.),  its  nearly  exclusive  food.  Be- 
fore the  early  1990s,  the  species  occurred  in  Panama 
primarily  as  a vagrant.  Apple  snails  were  introduced 
to  Lake  Gatun  in  central  Panama  in  the  late  1980s,  and 
by  1994  Snail  Kites  had  colonized  the  lake  from  pop- 
ulation sources  at  least  350  km  away  and  initiated 
breeding.  Since  1994  the  population  has  increased  rap- 
idly and  the  species  can  now  be  found  throughout  the 
lake.  Received  7 Oct.  1998,  accepted  6 Jan.  1999. 


The  Snail  Kite  {Rostrhamus  sociabilis)  is  a 
highly  specialized  raptor  that  ranges  from 
southern  Florida  and  Mexico,  through  Central 
America,  to  Bolivia,  northern  Argentina  and 
Uruguay  (Beissinger  1988).  Its  distribution  is 
closely  tied  to  that  of  apple  snails  {Pomacea 
spp.),  which  form  its  diet  almost  exclusively 
(Beissinger  1988).  The  kite  uses  its  exception- 
ally thin  upper  mandible  to  extract  snails  from 
their  shells  (Snyder  and  Snyder  1969).  Other 
species  of  snails,  turtles,  crabs  and  other  items 
are  taken  on  occasion  (Beissinger  1990a,  Sny- 
der and  Kale  1983,  Sykes  and  Kale  1974). 
The  species  is  nomadic,  moving  in  response 
to  changes  in  the  availability  of  its  favored 
prey  because  of  fluctuating  water  levels 
(Sykes  1979,  1983;  Beissinger  and  Takekawa 
1983,  Takekawa  and  Beissinger  1989). 

Snail  Kites  are  rare  in  southern  Central 


' Smithsonian  Tropical  Research  Institute,  Unit 
0948,  APO  AA  34002-0948; 

E-mail:  angehrg@tivoli.si.edu 


America,  and  there  have  been  only  seven  pre- 
vious reports  from  Panama  (Ridgely  and 
Gwynne  1989).  The  sole  Panama  specimen, 
an  immature  female,  was  collected  near  the 
Colombian  border  at  Perme,  near  Puerto  Ob- 
aldia,  San  Bias  Province,  in  1929  (Wetmore 
1965).  There  were  four  reports  of  single  birds, 
all  either  females  or  immatures,  at  marshes 
near  Panama  City,  in  1971,  1973,  1977,  and 
1979.  There  were  two  reports  from  Chiriqui 
province  in  western  Panama;  an  adult  male 
and  a female  or  immature  near  Gualaca  in 
1965,  and  several  pairs  and  a nest  in  a marsh 
near  Remedies  in  1973,  the  latter  being  the 
only  previous  report  of  breeding  activity  in 
Panama  (Ridgely  and  Gwynne  1989).  The 
closest  significant  populations  of  Snail  Kites 
to  Panama  are  in  western  Colombia,  on  the 
west  side  of  the  Gulf  of  Uraba  (Hilty  and 
Brown  1986),  about  350  km  from  the  Panama 
Canal  area,  and  the  Tempisque  Basin  in  north- 
ern Costa  Rica  (Stiles  and  Skutch  1989),  ap- 
proximately 650  km  away. 

STUDY  AREA 

Lake  Gatun  (420  kmO  is  an  artificial  lake  created 
by  the  damming  of  the  Chagres  River  to  form  the  cen- 
tral part  of  the  Panama  Canal  in  1914.  The  lake  level 
is  controlled  by  the  Panama  Canal  Commission,  and 
may  vary  several  meters  between  wet  and  dry  seasons 
as  water  is  released  when  ships  pass  through  the  locks. 
The  Canal  area  experiences  a strong  four-month  long 
dry  season  from  mid-December  to  mid-April. 

The  introduced  aquatic  plant  Hydrilla  verticillata  is 
a major  problem  in  the  lake.  It  apparently  first  became 
established  in  the  late  1920s  or  early  1930s,  and  had 


266 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


become  a significant  problem  by  the  1960s  (Maturell 
and  Salazar  1994).  Rooting  in  shallow  water,  it  now 
occurs  in  huge  beds  in  many  parts  of  the  lake. 

The  apple  snail  Pomacea  lattrei  was  introduced  to 
Panama  from  Guatemala  in  the  early  1980s  by  Pana- 
ma’s National  Directorate  of  Aquaculture.  Snails  were 
established  at  a research  station  at  Divisa,  Herrera 
Province,  approximately  130  km  west  of  Lake  Gatun, 
for  use  in  rural  aquaculture  programs.  Between  1986 
and  1988,  the  owner  of  a sportfishing  business  in  the 
town  of  La  Arenosa  on  the  southwestern  arm  of  Lake 
Gatun  obtained  15  snails  from  the  Divisa  station  and 
introduced  them  to  a small  artificial  pond  next  to  Lake 
Gatun.  After  they  had  reproduced,  she  distributed 
some  to  neighbors  and  introduced  others  into  the  lake. 
A few  months  later  it  was  noticed  that  the  snails  were 
feeding  on  and  reducing  the  abundant  Hydrilla  around 
the  community’s  public  dock  and  in  surrounding  areas. 
Residents  of  other  lakeside  communities  began  delib- 
erately introducing  snails  to  their  areas  in  an  effort  to 
control  Hydrilla  (Maturell  and  Salazar  1994),  a major 
impediment  to  small  boat  travel.  The  snails  are  also 
used  for  food  by  some  local  people. 

Surveys  by  the  Panama  Canal  Commission  found 
that  the  snails  were  distributed  throughout  the  south- 
eastern arm  of  the  lake  by  1991,  were  in  the  north- 
western part  by  1992,  and  had  reached  the  town  of 
Gamboa  at  the  east  end  of  the  lake  by  1993.  In  some 
areas  Hydrilla  cover  was  reduced  by  as  much  as  94% 
in  three  years.  Deliberate  introduction  by  humans  ev- 
idently allowed  the  snail  to  disperse  quickly  around 
the  lake.  The  snails  also  spread  by  floating  in  currents 
and  by  egg  masses  fixed  to  floating  vegetation,  logs, 
and  boats  (Maturell  and  Salazar  1994). 

A native  species  of  apple  snail,  Pomacea  cumingi, 
occurs  in  the  Canal  area.  It  is  a bottom-dwelling  spe- 
cies that  typically  occurs  in  low  densities.  It  mostly 
inhabits  small  streams,  but  is  sometimes  found  in  larg- 
er rivers.  The  combination  of  habitat,  bottom  dwelling 
habit,  and  low  density  evidently  makes  this  species 
unsuitable  as  prey  for  Snail  Kites  (L  G.  Thompson, 
pers.  comm.). 

RESULTS 

M.  Santamaria  (pers.  comm.),  a game  war- 
den at  the  Barro  Colorado  Nature  Monument, 
a reserve  managed  by  the  Smithsonian  Trop- 
ical Research  Institute,  first  observed  unusual 
hawklike  birds  eating  snails  in  Gigante  Bay, 
south  of  Barro  Colorado  Island,  in  May  or 
June  1994.  On  3 February  1995  Santamaria 
and  I visited  an  area  where  he  had  recently 
seen  birds  building  nests  in  Guindilla  Cove,  a 
narrow  inlet  about  1500  m long  on  the  south 
shore  of  Gigante  Bay.  We  saw  at  least  14  Snail 
Kites  in  the  cove  at  that  time,  including  at 
least  two  adult  males  in  black  plumage.  The 
remaining  birds  were  in  brown  plumage.  Im- 


matures  and  adult  females  have  brown  plum- 
age and  cannot  reliably  be  distinguished  in  the 
field  (S.  Beissinger,  pers.  comm.). 

At  four  locations  we  observed  nests  con- 
sisting of  loose  platforms  of  twigs  at  various 
stages  of  construction.  One  small  vine-cov- 
ered tree  standing  in  water  (Site  1)  had  seven 
nests  while  the  other  sites  had  one  nest  each. 
No  eggs  or  chicks  were  evident  at  that  time, 
although  we  could  not  see  the  contents  of  high 
nests.  Santamaria  recently  had  seen  birds 
bringing  twigs  to  add  to  the  nests. 

The  area  was  visited  again  on  20  May  1994 
by  D.  and  L.  Engleman.  They  did  not  see  any 
active  nests  at  Site  1,  but  at  least  17  kites  were 
soaring  or  perched  in  the  area.  They  observed 
five  nests  at  another  site  (Site  2)  about  300  m 
north  of  Site  1 . These  were  located  on  a small 
island  composed  of  Annona  glabra  shrubs 
overgrown  with  vines.  Three  nests  were  com- 
plete, one  with  two  eggs  and  another  with  at 
least  one  egg  and  perhaps  more.  A third  nest 
appeared  to  have  eggs  but  they  could  not  be 
counted.  The  remaining  two  nests  were  under 
construction  and  were  visited  by  birds  in 
brown  plumage  caiTying  twigs.  At  least  20 
kites  were  present  at  this  site,  yielding  a total 
of  at  least  37  in  the  cove  (D.  and  L.  Engleman, 
pers.  comm.). 

I visited  Guindilla  Cove  again  on  15  July 
1994.  At  that  time  seven  nests  were  present  at 
Site  2.  Two  nests  had  three  well-feathered 
nestlings,  while  two  more  fledglings  were 
perched  together  on  a branch  near  a third  nest. 
Three  other  birds  which  took  flight  from  the 
island  had  very  short  tails  and  appeared  to 
have  recently  fledged.  On  27  August  1994  two 
nests  at  Site  1 each  had  a single  lai'ge  downy 
young.  No  active  nests  were  present  at  Site  2, 
but  I saw  two  birds  that  apparently  had  re- 
cently fledged.  On  14  October  no  activity  was 
seen  at  Site  1,  but  three  apparently  recently 
fledged  birds  were  present  at  Site  2.  On  this 
date  I also  surveyed  several  other  ai’eas  in  Gi- 
gante Bay  that  I had  not  previously  investi- 
gated. One  nest  with  two  small  nestlings  was 
found  near  an  island  at  the  mouth  of  Guindilla 
Cove,  and  a second  nest  with  two  fledglings 
on  the  verge  of  flying  was  found  on  a large 
Annona  glabra  island  about  3 km  west  of 
Guindilla.  Approximately  15  adults  and  three 
apparently  recently  fledged  young  were  also 
seen  at  this  site. 


SHORT  COMMUNICATIONS 


267 


DISCUSSION 

Apple  snails  probably  reached  the  area  of 
Gigante  Bay  in  1991  (Matured  and  Salazar 
1994).  I am  certain  no  kites  were  present  in 
the  area  before  1992,  because  I made  three 
surveys  of  the  shoreline  of  parts  of  the  bay  by 
canoe  in  1991  and  1992,  and  surveyed  Guin- 
dilla  Cove  itself  on  23  August  1992.  Guindilla 
Cove  probably  was  colonized  by  kites  in  ei- 
ther 1993  or  1994. 

In  1995,  in  the  central  part  of  the  lake  neai' 
the  Panama  Canal  channel.  Snail  Kites  were 
restricted  to  Gigante  Bay,  based  on  informa- 
tion from  Smithsonian  game  wardens  and  re- 
searchers who  worked  on  the  lake.  Kites  were 
rarely  if  ever  seen  at  the  Smithsonian  research 
station  on  the  north  side  of  Barro  Colorado 
Island  4 km  away. 

Since  1995,  kites  have  been  seen  more  of- 
ten in  other  parts  of  Lake  Gatun  and  the  Canal 
area  in  general,  ranging  from  the  northern  end 
of  the  lake  near  the  Gatun  Locks  to  Miraflores 
Locks  near  the  Pacific  entrance  to  the  Panama 
Canal.  Although  no  comprehensive  surveys  of 
kite  distribution  and  numbers  have  been  made. 
Snail  Kites  appear  to  have  spread  throughout 
the  lake  and  occur  wherever  suitable  habitat 
is  present. 

In  1995  nest  construction  evidently  began 
in  January,  during  the  early  dry  season.  A few 
small  nestlings  were  still  present  in  mid-Oc- 
tober, during  the  late  rainy  season.  These  latter 
birds  would  probably  have  fledged  in  Novem- 
ber, so  that  in  1995  the  breeding  season  was 
at  least  10  months  long.  In  Florida,  the  main 
nesting  season  is  January-August,  peaking  in 
February-June,  although  in  yeais  of  high  wa- 
ter breeding  may  begin  in  December  and  ex- 
tend to  September  (Beissinger  1986,  1988; 
Snyder  et  al.  1989).  Nesting  seasonality  is 
poorly  known  elsewhere  in  the  tropics,  but 
may  be  tied  to  the  rainy  season.  Nesting  takes 
place  during  the  wettest  period  in  Surinam, 
Argentina  and  Venezuela  (Beissinger  1988). 

At  Gigante,  clutch  or  brood  size  ranged 
from  one  to  three.  In  Florida,  clutch  size  pres- 
ently ranges  from  one  to  three  with  the  latter 
much  more  frequent.  However,  clutch  size  ap- 
pears to  have  declined  historically  in  Florida, 
with  four-egg  clutches  previously  having  been 
common  (Beissinger  1986).  The  decline  may 
be  attributable  to  habitat  deterioration.  Al- 


though there  are  few  data,  clutch  size  else- 
where in  the  kite’s  range  seems  typically  to  be 
two  to  three,  although  four-egg  clutches  often 
occur  in  Argentina  (Beissinger  1990b)  and 
clutch  size  may  range  up  to  six  (Beissinger 
1988).  Despite  a presumably  superabundant 
food  supply  at  Gigante  in  1995,  there  was  no 
evidence  that  kites  increased  their  clutch  size 
in  response. 

The  Snail  Kite  colony  in  Guindilla  cove  in- 
creased from  no  birds  in  1992  to  at  least  37 
in  1995,  an  extraordinarily  high  rate  of  in- 
crease if  only  a single  founding  pair  had  been 
involved.  Because  Kites  aie  highly  sociable, 
initial  colonization  could  have  been  by  a small 
flock  of  birds  rather  than  a single  pair.  It  is 
possible  that  Guindilla  Cove  was  not  the  ini- 
tial site  of  colonization  of  Lake  Gatun.  Snails 
were  present  in  the  southwestern  arm  of  the 
lake  by  the  late  1980s.  Kites  could  have  col- 
onized that  area  first,  and  then  spread  to  Guin- 
dilla. The  southwestern  arm  of  the  lake  is 
large  and  remote  from  population  centers,  so 
a colony  there  could  easily  have  gone  unde- 
tected. 

Nevertheless,  Snail  Kite  populations  in- 
crease rapidly  under  favorable  conditions 
(Snyder  et  al.  1989,  Beissinger  1995),  such  as 
the  essentially  unlimited  food  supply  Lake 
Gatun  would  have  offered  the  first  colonists. 
The  age  at  first  reproduction  in  Snail  Kites  is 
very  low,  with  some  females  nesting  at  10 
months.  In  favorable  years  in  Florida,  the 
breeding  season  may  last  up  to  10  months,  and 
some  kites  may  re-nest  and  raise  second 
broods  (Beissinger  1986,  Snyder  et  al.  1989). 
Parents  of  either  sex  may  desert  a nest  leaving 
the  other  member  of  the  pair  to  continue  to 
raise  the  young  alone  (Beissinger  and  Snyder 
1987).  The  remaining  parent  almost  always  is 
able  to  rear  the  young  to  independence  by  it- 
self, while  the  deserting  parent  has  the  oppor- 
tunity to  re-nest  with  another  paitner. 

The  rapid  colonization  by  Snail  Kites  of 
Lake  Gatun  and  subsequent  population  expan- 
sion provides  an  interesting  example  of  the 
dispersal  capabilities  of  this  highly  opportu- 
nistic species.  The  most  likely  source  of  col- 
onizing birds  would  have  been  the  Colombian 
population,  approximately  350  km  from  Lake 
Gatun.  This  population  is  much  closer  than 
the  one  in  Costa  Rica,  and  there  ai'e  several 
large  areas  of  freshwater  habitat,  including 


268 


THE  WILSON  BULLETIN 


Vol.  Ill,  No.  2,  June  1999 


Lake  Bayano  and  several  large  rivers,  in  the 
intervening  area.  Dispersers  discovered  the 
newly  suitable  habitat  of  Lake  Gatun  within 
only  a few  years  after  the  introduction  of  their 
preferred  food,  and  rapidly  spread  throughout 
the  available  area. 

ACKNOWLEDGMENTS 

I would  particularly  like  to  thank  M.  Santamaria, 
who  first  told  me  of  the  kites,  D.  and  L.  Engleman, 
who  provided  additional  information,  and  D.  George, 
who  organized  two  of  our  visits  to  the  colony.  K. 
Aparicio,  S.  Lollett,  K.  Kaufmann,  and  J.  Nason  also 
assisted  with  field  observations.  I thank  L G.  Thomp- 
son for  identifying  snail  specimens  and  providing  ad- 
ditional information  on  Pomacea  biology.  S.  R.  Beis- 
singer  and  R W.  Sykes  Jr.  provided  helpful  reviews  that 
improved  the  manuscript.  I would  also  like  to  thank 
ANAM,  Panama’s  environmental  authority,  for  grant- 
ing the  permits  under  which  this  research  was  con- 
ducted. 

LITERATURE  CITED 

Beissinger,  S.  R.  1986.  Demography,  environmental 
uncertainty,  and  the  evolution  of  mate  desertion 
in  the  Snail  Kite.  Ecology  67:1445-1459. 
Beissinger,  S.  R.  1988.  Snail  Kite.  Pp.  148-165  in 
Handbook  of  North  American  birds.  Vol.  4 (R.  S. 
Palmer,  Ed.).  Yale  Univ.  Press,  New  Haven,  Con- 
necticut. 

Beissinger,  S.  R.  1990a.  Alternative  foods  of  a diet 
specialist,  the  Snail  Kite.  Auk  107:327—333. 
Beissinger,  S.  R.  1990b.  Experimental  brood  manip- 
ulations and  the  monoparental  threshold  in  Snail 
Kites.  Am.  Nat.  136:20-38. 

Beissinger,  S.  R.  1995.  Modeling  extinction  in  peri- 
odic environments:  everglades  water  levels  and 
Snail  Kite  population  viability.  Ecol.  Applications 
5:618-631. 

Beissinger,  S.  R.  and  N.  E R.  Snyder.  1987.  Mate 
desertion  in  the  Snail  Kite.  Anim.  Behav.  35:477- 
487. 


Beissinger,  S.  R.  and  J.  E.  Takekawa.  1983.  Habitat 
use  and  dispersal  by  Snail  Kites  in  Florida  during 
drought  conditions.  Fla.  Field  Nat.  11:89-106. 

Haverschmidt,  F.  1970.  Notes  on  the  Snail  Kite  in 
Surinam.  Auk  87:580-584. 

Hilty,  S.  L.  and  W.  L.  Brown.  1986.  A guide  to  the 
birds  of  Colombia.  Princeton  Univ.  Press,  Prince- 
ton, New  Jersey. 

Maturell,  j.  C.  and  a.  J.  Salazar.  1994.  Aspectos 
de  la  introduccion  y diseminacion  del  caracol  gi- 
gante  Pomacea  sp.  en  el  Lago  Gatun  y sus  efectos 
sobre  la  abundancia  de  Hydrilla  verticillata.  Pan- 
ama Canal  Commission,  Bureau  of  Engineering 
and  Construction,  Balboa,  Panama. 

Ridgely,  R.  and  j.  Gwynne.  1989.  A guide  to  the 
birds  of  Panama.  Princeton  Univ.  Press,  Princeton, 
New  Jersey. 

Snyder,  N.  E and  H.  A.  Snyder.  1969.  A compara- 
tive study  of  mollusc  predation  by  Limpkins,  Ev- 
erglade Kites,  and  Boat-tailed  Grackles.  Living 
Bird  8:177-233. 

Snyder,  N.  F.  and  H.  W.  Kale,  II.  1983.  Mollusk 
predation  by  Snail  Kites  in  Colombia.  Auk  100: 
93-97. 

Snyder,  N.  E,  S.  R.  Beissinger,  and  R.  F.  Chandler. 
1989.  Reproduction  and  demography  of  the  Flor- 
ida Everglade  (Snail)  Kite.  Condor  91:300—316. 

Stiles,  F.  G.  and  A.  F.  Skutch.  1989.  A guide  to  the 
birds  of  Costa  Rica.  Cornell  Univ.  Press,  Ithaca, 
New  York. 

Sykes,  P.  W.,  Jr.  1979.  Status  of  the  Everglade  Kite 
in  Florida  1968-1978.  Wilson  Bull.  91:495-511. 

Sykes,  P.  W.,  Jr.  1983.  Recent  population  trends  of 
the  Everglade  Kite  in  Florida  and  its  relationship 
to  water  levels.  J.  Field  Ornithol.  54:237-246. 

Sykes,  P.  W.,  Jr.  and  H.  W.  Kale,  II.  1974.  Ever- 
glades Kites  feed  on  non-snail  prey.  Auk  91:818- 
820. 

Takekawa,  J.  E.  and  S.  R.  Beissinger.  1989.  Cyclic 
drought,  dispersal,  and  the  conservation  of  the 
Snail  Kite  in  Florida:  lessons  in  critical  habitat. 
Conserv.  Biol.  3:302—311. 

Wetmore,  a.  1965.  The  birds  of  the  Republic  of  Pan- 
ama. Vol.  I.  Tinamidae  to  Rynchopidae.  Smith- 
sonian Institution  Press,  Washington,  D.C. 


SHORT  COMMUNICATIONS 


269 


Wilson  Bull.,  1 1 1(2),  1999,  pp.  269-271 


The  “Significant  Others”  of  American  Kestrels: 
Cohabitation  with  Arthropods 

Jeffrey  P Neubig'  - and  John  A.  Smallwood'’^ 


ABSTRACT. — We  examined  the  arthropod  fauna 
that  coexists  in  nest  boxes  with  American  Kestrel 
chicks  (Folco  sparverius)  in  northwestern  New  Jersey. 
Of  the  seven  arthropod  species  present,  five  were  scav- 
enging beetles,  including  carrion  beetles  (Silpha  in- 
aeqiialis),  hister  beetles  {Atholus  arnericanus  and  Phel- 
isier  subrotundus),  dermestid  beetles  (Dermestes  ca- 
ninus),  and  skin  beetles  (Trox  foveicollis),  which  ap- 
parently were  attracted  to  prey  remains  that 
accumulated  in  the  nest  boxes.  Arthropod  density  and 
species  richness  were  significantly  greater  for  nest  box- 
es in  which  kestrels  bred  than  for  unoccupied  nest  box- 
es. Received  30  June  1998,  accepted  16  Nov.  1998. 


Studies  of  the  association  between  Ameri- 
can Kestrels  {Falco  sparverius)  and  arthro- 
pods have  focused  primarily  on  ( 1 ) the  occur- 
rence of  these  invertebrates  in  kestrel  diets 
and  the  predatory  behavior  kestrels  direct  to- 
ward them  and  (2)  parasitic  arthropods  re- 
ported to  infest  kestrels  (for  reviews  see  Sher- 
rod 1978  and  Smallwood  and  Bird  in  press, 
respectively).  However,  arthropods  interact 
with  kestrels  in  another  functional  role,  that  of 
symbiotic  scavengers.  Kestrels  are  a cavity 
nesting  species;  typical  brood  size  is  four  or 
five,  and  the  nestling  period  lasts  about  30 
days  (Johnsgard  1990).  During  this  time  prey 
remains,  regurgitated  pellets,  and  other  organ- 
ic material  accumulates  in  the  nest  cavity 
(Balgooyen  1976;  Smallwood,  pers.  obs.). 
Balgooyen  (1976)  observed  dermestid  beetles 
{Dermestes  spp.)  in  each  of  approximately  40 
kestrel  nests  in  northern  California  and  com- 
mented on  their  role  in  nest  sanitation.  The 
objectives  of  the  present  study  were  to  ex- 
amine the  arthropod  community  that  coexists 
with  kestrels  breeding  in  nest  boxes  in  New 


' Dept,  of  Biology,  Montclair  State  Univ.,  Upper 
Montclair,  NJ  07043. 

- Current  address:  Dept,  of  Zoology,  The  Ohio  State 
Univ.,  Columbus,  OH  43210. 

^ Corresponding  author; 

E-mail:  smallwood@saturn.montclair.edu 


Jersey  and  to  compare  them  to  the  invertebrate 
fauna  in  nest  boxes  not  occupied  by  kestrels. 

STUDY  AREA  AND  METHODS 

The  study  area  was  located  in  rural  northwestern 
New  Jersey,  bordered  to  the  north  and  west  by  the 
Kittatinny  Ridge  and  Delaware  River,  and  to  the  east 
and  south  by  residential  and  commercial  development. 
This  area  is  characterized  by  mixed  agriculture,  in- 
cluding corn,  hay,  and  cattle  production,  and  forestland 
in  the  ridge  and  valley  physiographic  region  (Sauer  et 
al.  1997).  Sixty  wooden  nest  boxes  (internal  dimen- 
sions: 20  X 23  cm  floor,  ca  34  cm  in  height)  were 
erected  in  open  habitats  in  Sussex  County  (centered 
approximately  41°  11'  N,  74°  38'  W)  between  1 April 
1995  and  25  April  1997,  and  69  nest  boxes  in  Warren 
County  (approximately  40°  47'  N,  75°  04'  W)  between 
5 August  1995  and  26  March  1997.  Because  kestrels 
do  not  bring  any  nesting  material  into  the  nest  cavity 
(Bird  and  Palmer  1988),  we  covered  the  floor  of  each 
nest  box  with  approximately  6 cm  of  wood  shavings 
to  provide  a cushion  and  insulation  for  the  eggs. 

Each  nest  box  was  monitored  at  4-week  intervals 
from  23  March  through  13  July  1997  to  determine  oc- 
cupancy status;  a nest  box  was  considered  occupied  by 
breeding  kestrels  if  at  least  one  egg  was  observed. 
Kestrels  require  7-9  days  to  produce  and  28-30  days 
to  incubate  a clutch  (Bird  and  Palmer  1988);  thus,  all 
kestrel  nesting  attempts  were  discovered  during  the 
laying  or  incubation  stage.  Nest  boxes  in  which  kes- 
trels bred  (herein  “kestrel  nest  boxes”)  were  visited 
3-7  times  throughout  the  nesting  attempt.  During  the 
final  visit  chicks  were  banded  and  the  bedding  (in- 
cluding cohabiting  arthropods,  prey  remains,  pellets, 
and  any  other  material)  was  collected  and  replaced 
with  fresh  wood  shavings.  Pinal  visits  occurred  be- 
tween 21  June  and  1 August  1997  when  nestlings  were 
16-23  days  old  (75%  were  20-22  days  old). 

In  addition  to  kestrels,  other  vertebrates  that  bred  in 
the  nest  boxes  included  five  avian  species  (Great  Crest- 
ed Flycatcher,  Myiorchus  crinitus;  Tree  Swallow, 
Tachycineta  hicolor.  Eastern  Bluebird,  Siolia  sialis\ 
European  Starling,  Sturnus  vulgaris;  and  House  Spar- 
row, Baser  domesticus)  and  four  species  of  mammals 
(eastern  gray  squirrel,  Sciurus  carolinensis;  red  squir- 
rel, Tamiasciurus  hudsonicus;  southern  flying  squirrel, 
Glaucomys  volans;  and  white-footed  mouse,  Peromys- 
cus  leucopu.'i).  Only  four  nest  boxes  remained  unoc- 
cupied by  any  vertebrate  species,  including  kestrels, 
during  the  breeding  season  (herein  “unoccupied  nest 


270 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


TABLE  1.  Taxonomy  and  abundance  of  arthropods 
in  northwestern  New  Jersey. 

co-inhabiting  nest  boxes  with  American  Kestrel  broods 

Nest  boxes 

Individuals  per  occupied^  nest  box 

Taxonomy 

occupied 

{%) 

Mean 

SD 

Range 

Arachnida 

Araneida 

Aglenidae 

Tegenaria  derhamii,  funnel  spider 

6.3 

1.0 

0.0 

1-1 

Insecta 

Coleoptera 

Silphidae 

Silpha  inaequalis,  carrion  beetle 

68.8 

41.0 

30.8 

3-114 

Histeridae 

Atholus  americanus,  hister  beetle 

31.3 

4.8 

3.4 

2-9 

Phelister  suhrotimdiis,  hister  beetle 

37.5 

8.8 

11.1 

1-24 

Dermestidae 

Denneste.s  caninus,  dermestid  beetle 

12.5 

5.0 

5.7 

1-9 

Trogidae 

Trox  foveicollis,  skin  beetle 

6.3 

3.0 

0.0 

3-3 

Hymenoptera 

Formicidae 

Camponotus  pennsylvanicus,  black  carpenter  ant 

6.3 

1.0 

0.0 

1-1 

“ Nest  boxes  in  which  a particular  arthropod  species  was  present. 


boxes”).  The  bedding  was  collected  from  unoccupied 
nest  boxes  on  11  and  13  July  1997. 

All  bedding  samples  were  stored  at  — 22°C  in  air- 
tight plastic  bags  to  preserve  any  arthropod  specimens 
present.  We  subsequently  extracted  all  arthropods  vis- 
ible at  1.75X  magnification,  preserved  them  in  70% 
ethanol,  and  identified  them  with  information  from 
Comstock  and  Gertsch  (1948),  Emerton  and  Lrost 
(1961),  Borrer  and  White  (1970),  Headstrom  (1977), 
Kaston  (1978),  Milne  and  Milne  (1980),  and  Arnett 
(1985). 

Data  on  arthropod  density  and  richness  were  tested 
for  normality.  Because  significant  deviations  were  de- 
tected, we  used  nonparametric  statistical  treatments  ex- 
clusively, including  the  Wilcoxon  rank  sum  test, 
Spearman’s  correlation  coefficient,  and  Eisher’s  exact 
test  (Snedacor  and  Cochran  1980).  Analyses  were  per- 
formed using  SAS  6.12  on  a Sun  Solaris  2.6  platform. 

RESULTS 

A total  of  567  individual  arthropods  were 
extracted  from  the  bedding  collected  from  16 
kestrel  (Table  1)  and  four  unoccupied  nest 
boxes.  Arthropod  density  ranged  from  0 (for 
two  kestrel  and  two  unoccupied  nest  boxes) 
to  115  individuals  in  one  kestrel  nest  box. 
Mean  arthropod  density  for  kestrel  nest  boxes 
(35.3  ± 31.0  SD)  was  significantly  greater 
than  that  for  unoccupied  nest  boxes  (0.50  ± 
0.58  SD;  two-tailed  Wilcoxon  rank  sums  test: 
Z = 2.421,  P = 0.016).  Maximum  arthropod 
density  in  unoccupied  nest  boxes  was  1:  a 


crab  spider  (Thomsidae,  Xysticus  triguttus)  in 
one  nest  box  and  a European  earwig  (Forfi- 
eulidae,  Forficula  aiiriciilaria)  in  one  other. 

Species  richness  per  nest  box  ranged  from 
0-3  and  mean  richness  for  kestrel  nest  boxes 
(1.69  ± 0.87  SD)  was  significantly  greater 
than  that  for  unoccupied  nest  boxes  (0.50  ± 
0.58  SD;  two-tailed  Wilcoxon  rank  sum  test: 
Z = 2.258,  P = 0.024).  No  significant  corre- 
lation was  detected  between  arthropod  species 
richness  and  kestrel  brood  size  (r^  = 0.17,  P 
> 0.05,  n = 16),  or  between  kestrel  brood  size 
and  arthropod  density  (r^  = -0.14,  P > 0.05, 
n = 16). 

DISCUSSION 

Of  the  seven  arthropod  species  observed  in 
kestrel  nest  boxes  in  New  Jersey,  five  (all  the 
beetles)  are  considered  scavengers  (Head- 
strom 1977),  and  the  caipenter  ant  is  known 
to  forage  primarily  on  insects  (Palmer  1975). 
The  dermestid  beetle  (D.  caninus),  found  in 
two  of  the  nest  boxes,  was  the  same  species 
commonly  used  in  museums  for  cleaning  flesh 
from  skeletal  specimens  (Headstrom  1977). 

The  mean  and  maximum  densities  of  ar- 
thropods in  the  kestrel  nest  boxes  in  New  Jer- 
sey were  considerably  less  than  those  ob- 
served in  northcentral  Florida  (Smallwood, 


SHORT  COMMUNICATIONS 


271 


pers.  obs.).  In  Florida,  dermestid  beetles  and 
other  species  typically  occuned  in  large  num- 
bers, such  that  by  the  end  of  the  nestling  pe- 
riod the  substrate  upon  which  the  chicks  stood 
visibly  pulsated  as  the  result  of  the  motion  of 
these  arthropods.  In  an  experiment  to  deter- 
mine if  high  arthropod  densities  in  nest  cavi- 
ties discourages  second  clutches,  Smallwood 
(unpubl.  data)  replaced  the  bedding  for  a ran- 
domly selected  group  of  nest  boxes  after  kes- 
trel chicks  had  fledged.  Kestrels  renested  in 
11.1%  of  the  cleaned  nest  boxes  (n  = 27)  and 
14.0%  of  the  control  nest  boxes  {n  = 50); 
these  percentages  were  not  significantly  dif- 
ferent (one-tailed  Fisher’s  exact  test,  P = 
0.76). 

We  saw  no  evidence  of  kestrel  chicks  bitten 
or  otherwise  harmed  by  the  arthropods  living 
in  the  New  Jersey  nest  boxes.  Rather,  the  re- 
moval of  much  of  the  decaying  organic  ma- 
terial (i.e.,  uneaten  scraps  of  insects,  small 
birds,  and  rodents)  may  benefit  kestrels  by  re- 
ducing the  risk  of  disease  or  infestation  with 
parasites.  The  significant  difference  in  the 
number  and  species  of  arthropods  present  be- 
tween kestrel  and  unoccupied  nest  boxes,  and 
the  fact  that  nearly  all  the  arthropods  were 
scavengers,  suggest  that  breeding  kestrels  are 
producing  the  conditions  that  attract  these  in- 
vertebrates. 

ACKNOWLEDGMENTS 

We  thank  GPU  Energy  and  Sussex  Rural  Electric 
Cooperative  for  allowing  us  to  erect  nest  boxes  on 
their  utility  poles;  A.  Bonner,  J.  Bullis,  D.  Darlington, 
C.  Goodwin,  S.  Kamphausen,  S.  Kindler,  A.  Kinney, 
J.  Mershon,  C.  Natale,  J.  Trinca,  P.  Wargo,  N.  War- 
hoftig,  R.  Wyman,  and  K.  Zientko  for  monitoring  nest 
boxes.  Release  time  for  JAS  was  provided  by  the  Fac- 
ulty Scholarship  Incentive  Program,  Montclair  State 
University. 


LITERATURE  CITED 

Arnett,  R.  H.,  Jr.  1985.  American  insects.  A hand- 
book of  the  insects  of  America  north  of  Mexico. 
Van  Nostrand  Reinhold  Co.,  New  York. 

Balgooyen,  T.  G.  1976.  Behavior  and  ecology  of  the 
American  Kestrel  {Falco  sparveriits  L.)  in  the  Si- 
erra Nevada  of  California.  Univ.  Calif.  Publ.  Zool. 
103:1-83. 

Bird,  D.  M.  and  R.  S.  Palmer.  1988.  American  Kes- 
trel. Pp.  253-290  in  Handbook  of  North  American 
birds.  Vol.  5.  Diurnal  raptors.  Part  2 (R.  S.  Palmer, 
Ed.).  Yale  Univ.  Press,  New  Haven,  Connecticut. 

Borrer,  D.  j.  and  R.  E.  White.  1970.  A field  guide 
to  the  insects  of  America  north  of  Mexico.  Hough- 
ton Mifflin  Co.,  Boston,  Massachusetts. 

Comstock,  J.  H.  and  W.  J.  Gert.sch.  1948.  The  spider 
book.  Cornell  Univ.  Press,  Ithaca,  New  York. 

Emerton,  j.  H.  and  S.  W.  Frost.  1961.  The  common 
spiders  of  the  United  States.  Dover  Publications, 
Inc.,  New  York. 

Headstrom,  R.  1977.  The  beetles  of  America.  A.  S. 
Barnes  and  Co.,  Cranbury,  New  Jersey. 

Johnsgard,  P.  A.  1990.  Hawks,  eagles,  and  falcons  of 
North  America:  biology  and  natural  history. 
Smithsonian  Institution  Press,  Washington,  D.C. 

Kaston,  B.  j.  1978.  How  to  know  the  spiders,  third 
ed.  W.  C.  Brown  and  Co.,  Dubuque,  Iowa. 

Milne,  L.  and  M.  Milne.  1980.  The  Audubon  Society 
field  guide  to  North  American  insects  and  spiders. 
Alfred  Knopf,  New  York. 

Palmer,  E.  L.  1975.  Fieldbook  of  natural  history,  sec- 
ond ed.  McGraw-Hill  Book  Co.,  New  York. 

Sauer,  J.  R.,  J.  E.  Hines,  G.  Gough,  I.  Thomas,  and 
B.  G.  Peterjohn.  1997.  The  breeding  bird  survey 
results  and  analysis,  version  96.3.  Patuxent  Wild- 
life Research  Center,  Laurel,  Maryland.  URL  = 
http://www.mbr.nbs.gov/bbs/bbs.html. 

Sherrod,  S.  K.  1978.  Diets  of  North  American  Fal- 
coniformes.  Raptor  Res.  12:49—121. 

Smallwood,  J.  A.  and  D.  M.  Bird.  In  press.  Ameri- 
can Kestrel  (Falco  sparverius).  In  The  birds  of 
North  America  (A.  Poole  and  E Gill,  Eds.).  Acad- 
emy of  Natural  Sciences,  Philadelphia,  Pennsyl- 
vania; American  Ornithologists'  Union,  Washing- 
ton, D.C. 

Snedacor,  G.  W.  and  W.  G.  Cochran.  1980.  Statis- 
tical methods.  Seventh  ed.  Iowa  State  Univ.  Press, 
Ames. 

Wilson,  E.  O.  and  W.  H.  Bossert.  1971.  A primer 
of  population  biology.  Sinauer  Associates.  Inc., 
Sunderland,  Massachusetts. 


272 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


Wilson  Bull.,  111(2),  1999,  pp.  272-273 


Barred  Owl  Nest  in  Attic  of  Shed 

C.  Stuart  Houston' 


ABSTRACT — Barred  Owls  (Strix  varia)  nested  in 
the  attic  of  a shed  during  12  out  of  13  years  at  Llotten 
Lake,  Saskatchewan.  Twelve  nestlings  were  banded  in 
five  seasons.  Although  two  species  of  Strix  are  known 
to  nest  in  buildings  in  Europe,  this  appears  to  be  the 
first  such  instance  in  North  America.  Received  7 Oct. 
1998,  accepted  7 Jan.  1999. 


For  twelve  out  of  thirteen  years  (1980— 
1992),  Barred  Owls  (Strix  varia)  nested  in  the 
same  attic  of  a shed  along  Flotten  Lake,  Sas- 
katchewan. Prior  to  this.  Barred  Owls  had 
never  been  reported  to  use  a building  for  nest- 
ing purposes.  Furthermore,  this  example  rep- 
resents the  longest  continuous  use  of  one  nest 
site  by  this  species,  reported  to  date  in  Sas- 
katchewan. 

Barred  Owls  apparently  moved  into  central 
Saskatchewan  from  adjacent  Manitoba  in  the 
1950s.  They  occupied  a relatively  narrow 
band  of  mixed  forest,  most  often  near  lakes 
and  rivers.  The  first  Saskatchewan  nest  was 
located  in  1961  (Houston  1959,  1961).  In  the 
only  Barred  Owl  nesting  study  conducted  in 
Saskatchewan,  Mazur  and  coworkers  (1997) 
found  nests  only  in  mature  forest:  6 in  broken 
tree  snags,  4 in  broken  limbs,  2 on  squirrel 
platforms,  2 on  stick  nests,  and  1 on  a witch’s 
broom  platform.  Of  these  nests,  10  were  in 
deciduous  trees  and  5 were  in  conifers. 

On  15  May  1988,  I first  visited  an  aban- 
doned shed  (Fig.  1),  built  in  1946  in  a 1.6-ha 
man-made  clearing  located  in  mixed  old- 
growth  forest  near  the  shores  of  Flotten  Lake, 
Saskatchewan  (54°  30'  N,  108°  30'  W).  Dur- 
ing the  visit  I banded  three  half-grown  Barred 
Owl  nestlings.  The  property  owners,  Mr.  and 
Mrs.  D.  Mazuren,  had  observed  Barred  Owls 
nesting  on  the  attic  floor  in  this  building  ever-y 
year  except  one  since  1980.  Consequently, 
1988  was  the  eighth  year  of  use  in  nine  years. 
During  other  limes  of  year,  the  owls  were  seen 


' 863  University  Drive,  Saskatoon,  SK  S7N  ()J8, 
Canada;  E-mail:  houstons@duke.u.sask.ca 


in  nearby  trees  but  never  perched  on  or  in  the 
buildings.  On  3 June  1989,  we  again  banded 
three  young  and  took  photographs.  On  1 1 
June  1990,  we  banded  the  single  nestling 
raised  that  year;  on  28  May  1991,  three  nest- 
lings; and  on  30  May  1992,  two  nestlings 
(only  2 eggs  had  been  present  on  19  April). 
The  shed  attic  has  not  been  used  since. 

The  12  years  of  use  is  similar  to  the  10 
consecutive  years  that  Baned  Owls  used  a 
deep  cavity  in  a dead  oak  in  southeastern 
Massachusetts  (Bent  1938).  I have  found  no 
other  records  of  Barred  Owls  using  buildings 
for  nest  sites,  although  they  use  artificial  nest 
boxes  in  Wisconsin  and  Minnesota  (Johnson 
and  Follen  1984).  In  Europe,  other  species  of 


FIG.  1.  Author  banding  young  Barred  Owls  in  the 
attic  of  a shed.  Photograph  by  D.  G.  Miller. 


SHORT  COMMUNICATIONS 


273 


Strix  use  buildings,  especially  bai'ns,  as  nest 
sites.  The  Tawny  Owl  {Strix  aluco)  uses  build- 
ings as  nest  sites  15%  of  the  time,  similarly, 
the  Ural  Owl  {Strix  iiralensis)  nests  in  build- 
ings 2-4%  of  the  time  (Mikkola  1983). 

ACKNOWLEDGMENTS 

I thank  M.  Belcher,  J.  S.  Marks,  J.  F.  Roy,  and  four 
anonymous  reviewers  for  constructive  criticism. 

LITERATURE  CITED 

Bent,  A.  C.  1938.  Life  histories  of  North  American 
birds  of  prey.  U.S.  Nat.  Mus.  Bull.  170:182—201. 


Houston,  C.  S.  1959.  First  records  of  the  Barred  Owl 
in  Saskatchewan.  Blue  Jay  17:94. 

Hou.ston,  C.  S.  1961.  First  Saskatchewan  nest  of 
Barred  Owl.  Blue  Jay  19:1  14-1  15. 

Johnson,  D.  H.  and  D.  G.  Folehn.  1984.  Barred  owls 
and  nest  boxes.  Raptor  Res.  18:34-35. 

Ma/.ur,  K.  M.,  P.  C.  James,  and  S.  D.  Frith.  1997. 
Barred  Owl  (Strix  varia)  nest  site  characteristics 
in  the  boreal  forest  of  Saskatchewan,  Canada.  U.S. 
Forest  Service  General  Technical  Report  NC-190: 
267-27 1 . 

Mikkola,  H.  1983.  Owls  of  Europe.  Buteo  Books, 
Vermillion,  South  Dakota. 


Wilson  Bull.,  111(2),  1999,  pp.  273-216 


Double  Broo(iing  in  the  Long-eareii  Owl 

Jeffrey  S.  Marks'-^  and  Alison  E.  H.  Perkins' 


ABSTRACT. — Owls  in  the  family  Strigidae  typi- 
cally raise  no  more  than  one  brood  per  year.  We  doc- 
umented what  apparently  is  the  first  unequivocal  case 
of  double  brooding  in  Long-eared  Owls  (Asia  otiis).  A 
banded  female  raised  12  young  in  two  nesting  attempts 
compared  with  a mean  of  5.3  young  for  three  single- 
brooded  females  that  nested  in  the  same  grove.  Two 
factors  may  have  influenced  the  occurrence  of  double 
brooding:  the  first  nest  was  initiated  unusually  early  in 
the  year  (mid-February)  and  food  availability  (in  the 
form  of  voles)  was  high.  The  rare  description  of  double 
brooding  in  Long-eared  Owls  may  be  due  to  the  dif- 
ficulty of  detecting  it.  Alternatively,  double  brooding 
may  be  uncommon  because  it  is  seldom  an  economi- 
cally viable  strategy.  Factors  that  would  select  against 
double  brooding  include  low  probability  of  recruitment 
of  the  first-brood  young,  and  reduced  survival  and  fe- 
cundity of  the  adults.  Received  17  Sept.  1998,  accepted 
29  Dec.  1998. 


The  number  of  young  raised  per  year  is  an 
important  component  of  an  individual’s  life- 
time reproductive  success.  Double  brooding, 
in  which  a second  brood  is  attempted  after  a 
successful  first  attempt,  is  a viable  strategy  if 
the  increase  in  fitness  that  results  outweighs 
the  cost  of  any  reduction  in  future  survival  or 


' Montana  Cooperative  Wildlife  Research  Unit, 
Univ.  of  Montana,  Missoula,  MT  59812. 

'Corresponding  author; 

E-mail:  jmarks(3> selway.umt.edu 


fecundity  of  the  adults  and  their  young  from 
the  first  brood.  The  occunence  of  double 
brooding  may  be  influenced  by  factors  such 
as  length  of  the  breeding  season,  food  avail- 
ability, growth  rates  of  the  young,  and  the  du- 
ration and  quality  of  parental  care  (e.g.,  Drent 
and  Daan  1980,  Askenmo  and  Unger  1986, 
Tinbergen  and  van  Balen  1988). 

Double  brooding  is  relatively  rare  in  rap- 
tors, presumably  because  the  length  of  the 
breeding  cycle  and  extended  postfledging  care 
preclude  its  occunence  (Newton  1979,  Mor- 
rison 1998).  Among  nocturnal  raptors,  double 
brooding  occurs  regularly  in  Bani  Owls  {Tyto 
alba-,  Marti  1992,  1997)  and  occasionally  in 
Florida  Bunowing  Owls  {Athene  cimicularia 
floridana-,  Millsap  and  Bear  1990)  and  Boreal 
Owls  {Aegoliiis  funereiis-,  Kellomaki  et  al. 
1977,  Solheim  1983).  During  a study  of 
breeding  Long-eared  Owls  {Asia  otiis),  we 
documented  a female  that  raised  two  broods 
during  the  same  nesting  season.  Here,  we  de- 
scribe the  event  and  discuss  factors  that  may 
have  influenced  its  occuirence. 

STUDY  AREA  AND  METHODS 

The  study  area  is  a small  grove  (ca  2 ha)  of  quaking 
aspens  (Populiis  trenudoides)  and  black  hawthorns 
(Crataegus  douglasii)  located  about  16  km  west  of 
Poison.  Lake  County,  Montana  (47°  40'  N,  114°  20' 
W).  The  elevation  at  the  site  is  888  m,  and  the  nesting 


274 


THE  WILSON  BULLETIN 


Vol.  Ill,  No.  2,  June  1999 


grove  is  surrounded  by  grasslands  and  agricultural 
fields  (mostly  hay).  The  1 1 Long-eared  Owl  nests  that 
occurred  in  the  grove  in  1997  and  1998  were  in  old 
nests  of  Black-billed  Magpies  (Pica  pica)  and  Amer- 
ican Crows  (Conns  brachyrhynchos). 

Adults  were  captured  at  night  in  mist  nests  placed 
near  the  nest  or  at  dusk  with  the  aid  of  a plastic  decoy 
of  a Great  Horned  Owl  (Bubo  virginianus).  Captured 
adults  were  classified  as  after  hatching  year  (AHY)  or 
after  second  year  (ASY)  based  on  the  absence  or  pres- 
ence of  two  generations  of  secondaries,  respectively 
(see  Pyle  1997).  During  the  breeding  season,  the  sex 
of  most  adults  can  be  determined  in  the  field  by  dif- 
ferences in  plumage  coloration  and  in  the  hand  by 
presence  or  absence  of  an  incubation  patch  (Marks  et 
al.  1994). 

RESULTS 

On  16  February  1998,  we  observed  a fe- 
male Long-eared  Owl  incubating  at  a nest 
(PSNII)  about  25  m north  of  a nest  that  had 
produced  young  the  previous  year.  The  male 
was  roosting  nearby,  but  we  could  not  deter- 
mine whether  he  was  banded.  Three  of  the 
five  Long-eared  Owl  nests  that  occuired  in  the 
grove  in  1998  were  initiated  in  February; 
PSNII  appeared  to  be  the  earliest  of  these 
nests.  On  2 April,  JSM  captured  the  adult 
(AHY)  female  (band  no.  951)  by  hand  at 
PSNII  as  she  was  brooding  seven  young  that 
ranged  in  age  from  about  1 to  3 weeks  old. 
Based  on  the  estimated  age  of  the  chicks  and 
an  incubation  period  of  28  days  (Marks  et  al. 
1994),  female  951  would  have  initiated  egg 
laying  on  12  February,  and  the  oldest  chick 
would  have  hatched  on  12  March.  Other  du- 
ties prevented  us  from  catching  the  mate  of 
female  951  during  this  nesting  attempt. 

During  a visit  to  the  nesting  grove  on  25 
June,  we  found  a new  Long-eared  Owl  nest  at 
the  northern  edge  of  the  grove  28  m from  nest 
PSNII.  A female  was  brooding  small  chicks 
that  appeared  to  be  about  2 weeks  old,  and  a 
male  was  flushed  from  the  same  roost  site  typ- 
ically used  by  the  PSNII  male  earlier  that 
spring.  We  returned  to  the  nest  on  the  evening 
of  30  June  and  captured  the  female  in  a mist 
net  placed  directly  in  front  of  the  nest.  She 
proved  to  be  951,  the  same  female  that  had 
fledged  seven  chicks  earlier  in  the  spring.  The 
next  morning  we  banded  her  five  chicks, 
which  were  about  2 to  3 weeks  old.  Female 
951  was  very  aggressive  as  we  handled  her 
chicks,  diving  and  perching  within  1 m of  us 
and  enabling  us  to  observe  her  band  and  to 


note  her  pattern  of  flight-feather  molt  that  we 
had  confirmed  in  the  hand  the  previous  night 
(4  primaries  and  2 secondaries  growing  on 
each  wing).  Also  at  this  time,  we  noticed  that 
her  mate  was  banded.  After  several  attempts, 
we  succeeded  in  capturing  the  male  on  the 
evening  of  13  July,  at  which  time  the  oldest 
chicks  were  capable  of  short  flights  from  tree 
to  tree.  The  male  proved  to  be  no.  914,  the 
same  male  that  had  nested  at  this  site  in  1997. 
Female  95 1 was  still  present,  and  both  adults 
presumably  were  provisioning  their  fledglings. 
Interestingly,  male  914  had  not  staited  flight- 
feather  molt. 

Female  951  fledged  12  young  (defining 
“fledging”  as  capable  of  sustained  flight; 
Marks  1986)  in  two  nesting  attempts  com- 
pared with  a mean  of  5.3  young  (range  5 to 
6)  produced  by  the  other  three  pairs  that  nest- 
ed in  the  grove  in  1998.  The  estimated  time 
between  the  initiation  of  95 1 ’s  two  nesting  at- 
tempts was  90  days  (i.e.,  12  February  and  12 
May).  At  the  time  951  initiated  her  second 
clutch,  the  oldest  offspring  from  her  first 
brood  would  have  been  about  6 weeks  old. 

DISCUSSION 

Several  records  of  double  brooding  by 
Long-eared  Owls  have  been  reported  in  Eu- 
rope (Reinsch  and  Warncke  1968,  Rinne  1981, 
Scott  1997),  but  in  each  case  the  evidence  was 
circumstantial.  To  our  knowledge,  ours  is  the 
first  report  of  double  brooding  in  Long-eared 
Owls  based  on  a banded  individual. 

We  suspect  that  weather  and  food  avail- 
ability played  a major  role  in  this  case  of  dou- 
ble brooding.  The  winter  of  1998  was  unusu- 
ally mild  in  western  Montana.  The  ground  at 
the  study  area  was  virtually  free  of  snow  from 
January  onward  (pers.  obs.),  and  the  mean 
temperature  in  February  was  2.1°  C above 
normal  at  the  Keir  Dam  weather  station  1 3 km 
from  the  study  area  (data  obtained  from  the 
National  Climatic  Data  Center).  In  addition, 
voles  (Microtiis  spp.)  were  abundant  in  winter 
and  spring;  we  saw  many  during  the  day,  and 
other  vole-eating  raptors  [i.e..  Northern  Har- 
rier (Circus  cyaneus).  Rough-legged  Hawk 
(Buteo  lagopus),  and  Short-eared  Owl  (Asia 
flammeiis)]  were  numerous  in  the  study  area. 
The  mild  weather  and  abundant  food  probably 
induced  Long-eared  Owls  to  nest  in  Febmary, 
which  is  very  early  for  this  species  (see  Marks 


SHORT  COMMUNICATIONS 


275 


et  al.  1994).  The  continued  high  numbers  of 
voles  in  summer  provided  an  opportunity  for 
double  brooding,  at  least  for  one  of  the  three 
pairs  that  began  nesting  in  February. 

In  general,  the  incidence  of  second  nesting 
attempts  in  facultatively  double-brooded  spe- 
cies is  negatively  conelated  with  the  laying 
date  of  the  first  clutch  (e.g..  Smith  et  al.  1987, 
Geupel  and  DeSante  1990,  Monison  1998). 
Our  case  agrees  with  this  finding,  but  it  raises 
the  question  of  why  the  other  two  Long-eaied 
Owl  pairs  that  nested  early  did  not  raise  a sec- 
ond brood.  One  possibility  is  that  the  pheno- 
typic quality  of  the  double-brooded  pair  was 
high  relative  to  the  other  pairs  (see  Verboven 
and  Verhulst  1996).  Although  we  have  no  ob- 
jective measure  of  phenotypic  quality  in  the 
Long-eared  Owls  we  studied,  we  note  that  the 
male  of  the  second  nesting  attempt  had  bred 
successfully  at  the  site  in  the  previous  year 
(the  other  males  did  not  breed  there  in  1997), 
and  the  female  that  nested  twice  was  in  good 
physical  condition  and  was  unusually  aggres- 
sive. Indeed,  during  her  first  attempt  she  at- 
tacked JSM  when  he  entered  the  nest.  More- 
over, the  ratio  of  her  body  mass  (g)  to  wing 
length  (mm)  at  first  capture  (1.33)  was  higher 
than  that  of  all  but  one  of  the  other  eight  fe- 
males captured  late  in  the  brooding-rearing 
period  in  1997  and  1998  (Jc  = 1.11  ± 0.13 
SD,  range  0.98-1.34).  The  local  experience  of 
the  male  and  the  physical  condition  and  ag- 
gressiveness of  the  female  are  consistent  with 
the  notion  that  they  were  high-quality  individ- 
uals relative  to  the  other  early  nesters  in  the 
grove. 

Double  brooding  in  Long-eared  Owls  may 
be  more  common  than  previously  thought.  Al- 
ternatively, it  may  indeed  be  rare  because  it  is 
seldom  an  economically  viable  strategy.  For 
instance,  the  fitness  gain  from  double  brood- 
ing would  be  marginal  if  the  probability  of 
recruitment  of  first-brood  young  is  low  (i.e., 
because  of  reduced  care  from  parents  that  di- 
rect their  efforts  to  a new  brood),  or  if  future 
survival  and  fecundity  of  the  adults  are  re- 
duced. Female  Long-eared  Owls  in  Idaho  de- 
serted their  broods  when  the  young  were  6.5 
to  8 weeks  old,  and  males  continued  to  care 
for  the  young  until  they  were  8.5  to  1 1 weeks 
old  (Ulmschneider  1990).  If  parental  care  of 
this  duration  is  typical  in  Long-eared  Owls, 
then  the  first  of  the  two  broods  would  have 


received  a normal  amount  of  parental  care  (the 
second  clutch  was  started  when  the  oldest 
chicks  from  the  first  nest  were  6 weeks  old) 
only  if  the  female  changed  mates  between 
nesting  attempts  (and  the  male  continued  to 
care  for  the  young),  or  if  one  or  both  parents 
continued  to  provision  the  first  brood  while 
starting  the  second  (an  unlikely  occurrence 
given  that  the  male  must  provide  food  to  the 
incubating  female).  Moreover,  the  timing  of 
second  broods  could  interfere  with  the  molt 
schedule  of  adults.  Long-eared  Owls  generally 
begin  molting  in  early  June  soon  after  breed- 
ing (Marks  et  al.  1994).  The  male  attending 
the  second  brood  had  not  started  flight-feather 
molt  in  mid-July,  suggesting  that  his  molt  was 
delayed  because  of  the  late  breeding  effort. 
Delayed  molt  potentially  could  influence  sur- 
vivorship and  fecundity  (see  Pietiainen  et  al. 
1984,  Kjellen  1994). 

In  conclusion,  double  brooding  appears  to 
be  rare  in  Long-eared  Owls,  and  it  probably 
occurs  only  when  first  nests  are  initiated  early 
and  food  availability  is  high.  Nothing  is 
known  about  how  double  brooding  affects  re- 
cruitment of  young  from  first  versus  second 
broods,  or  whether  it  affects  the  survivorship 
and  future  fecundity  of  the  pai'ents.  Whether 
double  brooding  is  a viable  strategy  in  Long- 
eared Owls  remains  to  be  determined  from  ad- 
ditional research. 

ACKNOWLEDGMENTS 

We  thank  R and  D.  Smith  for  granting  access  to  their 
land;  C.  Olson.  T Dial,  and  R.  Petty  for  help  in  catch- 
ing owls;  C.  Marti  for  comments  on  the  manuscript; 
and  H.  Wright  for  translating  two  papers.  This  research 
was  supported  by  a grant  from  the  University  of  Mon- 
tana Research  and  Creativity  Committee. 

LITERATURE  CITED 

Askknmo,  C.  AND  U.  Ungkr.  1986.  How  to  be  double- 
brooded:  trends  and  liming  of  breeding  perfor- 
mance in  the  Rock  Pipit.  Ornis  Scand.  17:237- 
244. 

Drknt,  R.  H.  and  S.  Daan.  1980.  The  prudent  parent: 
energetic  adjustments  in  avian  breeding.  Ardea 
68:223-252. 

Gkupkl,  G.  R.  and  D.  F.  DhSantk.  1990.  Incidence 
and  determinants  of  double  brooding  in  Wrentits. 
Condor  92:67-75. 

Khllomaki,  E.,  E.  Heinonen,  and  H.  Tiainen.  1977. 
Two  successful  nestings  of  Tengmalm’s  Owl  in 
one  summer.  Ornis  Fenn.  54:134-135. 


276 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


Kjkllen,  N.  1994.  Moult  in  relation  to  migration  in 
birds:  a review.  Ornis  Svecica  4:1-24. 

Marks,  J.  S.  1986.  Nest  site  characteristics  and  re- 
productive success  of  Long-eared  Owls  in  south- 
western Idaho.  Wilson  Bull.  98:547-560. 

Marks,  J.  S.,  D.  L.  Evans,  and  D.  W.  Holt.  1994. 
Long-eared  Owl  {Asia  otiis).  In  The  birds  of  North 
America.  No.  133  (A.  Poole  and  L.  Gill,  Eds.). 
Academy  of  Natural  Sciences,  Philadelphia,  Penn- 
sylvania; American  Ornithologists’  Union,  Wash- 
ington, D.C. 

Marti,  C.  D.  1992.  Barn  Owl  (Tyto  alba).  In  The 
birds  of  North  America,  no.  1 (A.  Poole,  P.  Stet- 
tenheim,  and  P.  Gill,  Eds.).  Academy  of  Natural 
Sciences,  Philadelphia,  Pennsylvania;  American 
Ornithologists’  Union,  Washington,  D.C. 

Marti,  C.  D.  1997.  Lifetime  reproductive  success  in 
Barn  Owls  near  the  limit  of  the  species’  range. 
Auk  1 14:581-592. 

Millsap,  B.  a.  and  C.  Bear.  1990.  Double-brooding 
by  Plorida  Burrowing  Owls.  Wilson  Bull.  102: 
313-317. 

Morrison,  J.  L.  1998.  Effects  of  double  brooding  on 
productivity  of  Crested  Caracaras.  Auk  1 15:979- 
987. 

Newton,  I.  1979.  Population  ecology  of  raptors.  Bu- 
teo  Books,  Vermillion,  South  Dakota. 

PlETIAINEN,  H.,  P.  SaUROLA,  AND  H.  KOLUNEN.  1984. 
The  reproductive  constraints  on  moult  in  the  Ural 


Owl  Stri.K  uralensis.  Ann.  Zool.  Penn.  21:277- 
281. 

Pyle,  P.  1997.  Plight-feather  molt  patterns  and  age  in 
North  American  owls.  Am.  Birding  Assoc.  Mon- 
ogr.  Pield  Ornithol.  2:1-32. 

Reinsch,  a.  and  K.  Warncke.  1968.  Zweibruten  bei 
der  Waldohreule.  Anz.  Ornithol.  Ges.  Bayern  8: 
400-401. 

Rinne,  U.  1981.  Beitrag  zur  Brutbiologie  der  Waldoh- 
reule {Asia  otus).  Ornithol.  Mitt.  33:62—65. 

Scott,  D.  1997.  The  Long-eared  Owl.  The  Hawk  and 
Owl  Trust,  London,  U.K. 

Smith,  H.  G.,  H.  Kallander,  and  J.-A.  Nilsson. 
1987.  Effect  of  experimentally  altered  brood  size 
on  frequency  and  timing  of  second  clutches  in  the 
Great  Tit.  Auk  104:700-706. 

SoLHEiM,  R.  1983.  Bigyny  and  biandry  in  the  Teng- 
malm’s  Owl  Aegolius  funereu.s.  Ornis  Scand.  14: 
51-57. 

Tinbergen,  J.  M.  and  J.  H.  van  Balen.  1988.  Pood 
and  multiple  breeding.  Acta  Congr.  Int.  Ornithol. 
19:380-391. 

Ulm.schneider,  H.  1990.  Post-nesting  ecology  of  the 
Long-eared  Owl  (Asio  otus)  in  southwestern  Ida- 
ho. M.S.  thesis,  Boise  State  Univ.,  Boise,  Idaho. 

Verboven,  N.  and  S.  Verhulst.  1996.  Seasonal  var- 
iation in  the  incidence  of  double  broods:  the  date 
hypothesis  fits  better  than  the  quality  hypothesis. 
J.  Anim.  Ecol.  65:264—273. 


Wilson  Bull,  1 1 1(2),  1999,  pp.  276-278 


Planning  to  Facilitate  Caching: 
Possible  Suet  Cutting  by  a Common  Raven 


Bernd  Heinrich' 


ABSTRACT — Many  species  of  birds  feed  on  suet 
in  winter.  As  far  as  is  known,  they  all  take  bite-sized 
chunks  by  pecking  into  this  food  randomly  and/or  they 
tear  off  protruding  pieces.  I compared  the  peck-marks 
left  on  suet  by  Blue  Jays  (Cyanocitta  cristata)  and 
American  Crows  (Connis  hrachyrhyncho.s)  with  those 
left  by  Common  Ravens  (Corvus  corax).  Although 
most  ravens  feed  like  jays  and  crows,  at  least  one  in- 
dividual made  distinct  grooves,  aligning  dozens  of 
consecutive  pecks,  apparently  to  cut  transportable 
chunks  off  large  suet  blocks.  Received  2H  Aug.  1998, 
accepted  7 Jan.  1999. 


' Dept,  of  Biology,  Univ.  of  Vermont,  Burlington, 
VT  05405. 


The  Common  Raven,  Corvus  corax,  is  a 
feeding  generalist  (Bent  1946,  Ratcliffe  1997). 
Ravens  feed  on  carrion  (Ewins  et  al.  1986), 
fruit,  grain,  eggs,  and  “garbage”  (Nelson 
1934,  Marquiss  and  Booth  1986,  Engel  and 
Young  1989).  Ravens  also  capture  insects, 
reptiles,  amphibians,  fish,  small  mammals, 
and  other  birds  (Man’  and  Knight  1982,  Camp 
et  al.  1993).  I here  describe  a raveri  removing 
fat  from  a chunk  of  suet  in  an  unusual  or  ab- 
enant  way  that  differs  markedly  from  the 
method  used  by  jays,  crows,  and  most  other 
ravens. 

Chunks  of  beef  suet  that  were  either  of  suf- 
ficient size  so  that  they  could  not  be  earned 
otl  or  that  were  nailed  onto  the  frozen  ground 


SHORT  COMMUNICATIONS 


277 


FIG.  I.  Two  top  photographs  show  grooves  left  by  ravens  in  suet  when  they  were  interrupted  while  feeding 
in  the  wild.  A:  Deep  (1-2  cm)  groove  in  a clear  block  of  beef  suet.  B:  Two  grooves  cut  into  suet  adhering  to 
ribs.  C:  Typical  pecking  pattern  in  suet  by  American  Crow.  D:  Typical  pecking  pattern  in  suet  by  a Blue  Jay. 
(Right  two  pictures  show  heads  of  spikes  used  to  secure  the  suet  onto  frozen  ground  so  that  it  could  not  be 
carried  off.) 


were  provided  in  the  forest  near  Hinesburg, 
Vermont.  For  many  years  the  following  spe- 
cies fed  on  suet  at  this  site:  American  Crows 
{Corvus  brachyrhynchos).  Blue  Jays  {Cyano- 
citta  cristata).  Hairy  (Picoides  villosus)  and 
Downy  woodpeckers  (Picoides  pubescens). 
Black-capped  Chickadees  (Poecile  atricapil- 
lus)  and  White-breasted  Nuthatches  (Sitta  car- 
olinensis).  All  of  these  birds  fed  on  suet  by 
picking  into  it  and/or  taking  the  most  promi- 
nently protruding  pieces. 

On  one  occasion  a raven  flew  away  from 
the  feeding  station  as  I approached,  and  the 
pattern  of  pecks  the  raven  had  left  on  the  suet 
was  distinctive  (Fig.  lA).  The  raven  had 
carved  a 7.5  cm  long  (and  1—2  cm  deep) 
groove.  I presume  that  I had  inteiTupted  the 
bird  just  before  it  had  finished  the  task  of  cut- 
ting a smaller  chunk  of  suet  off  the  larger.  The 
same  chunk  of  suet  had  what  appeared  to  be 
a pattern  of  a previous  parallel  cut  (2-3  cm 


anterior  to  the  unfinished  cut)  where  the  bird 
had  already  removed  one  slice  of  fat.  The 
readily  available  small  pieces  of  frozen  suet 
chips  (2-3  mm)  that  had  been  loosened  were 
next  to  the  groove  (Fig.  lA). 

When  I inteiTupted  two  ravens  as  they  were 
feeding  on  the  suet  at  the  same  site  on  three 
later  occasions,  I found  similar  grooves  and 
only  raven  tracks  around  this  suet.  When  I left 
suet  that  was  firmly  attached  to  ribs,  the  ra- 
vens cut  grooves  through  the  fat  down  to  the 
bone  underlying  the  fat.  No  suet  chunks  could 
be  removed,  but  the  bird(s)  then  cut  another, 
parallel  groove  (Fig.  IB).  Crows  and  Blue 
Jays  whose  feeding  patterns  were  observed 
routinely  at  the  same  site  always  left  only 
peck-marks,  never  grooves  (Fig.  1C,  D).  At 
no  time  have  I observed  groove  cutting  in 
caged  ravens  that  I watched  routinely.  Neither 
have  I seen  such  behavior  in  hundreds  of 


278 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


hours  of  watching  groups  of  ravens  feeding 
on  frozen  muscle  meat  in  the  wild. 

What  accounts  for  the  ravens’  unique  feed- 
ing patterns  on  suet?  The  peck  marks  were 
clearly  aligned  in  rows.  Each  peck  could  have 
provided  only  a small  immediate  reward,  but 
it  made  possible  the  removal  of  a laige  chunk 
of  fat  and  hence  a large  reward  later.  Since 
loose  “crumbs”  of  fat  were  left  (Fig.  lA),  the 
delayed  (greater)  reward  was  apparently  of 
more  importance  to  the  birds  than  the  small 
proximate  reward.  Ravens  exhibit  similar  ap- 
parent foresight  during  some  aspects  of  their 
caching  behavior  (Heinrich  and  Pepper  1998). 

Food  access  behavior  is  of  interest  because 
it  has  traditionally  provided  a tool  for  exam- 
ining cognition.  Examples  with  corvids  in- 
clude studies  of  memory  (Baida  and  Kamil 
1989,  Kamil  and  Baida  1985,  Bednekoff  et. 
al.  1997),  tool  use  (Hunt  1996),  optimal  for- 
aging (Zach  1979,  Waite  and  Ydenberg  1994), 
and  insight  learning  (Heinrich  1995). 

The  fat  cutting  behavior  is  probably  very 
rare  in  ravens  and  it  is  not  likely  an  innate  or 
hard-wired  response.  Therefore,  the  raven 
could  have  had  insight  of  how  to  remove  a 
large  chunk  of  food  for  storage  and/or  later 
consumption. 

LITERATURE  CITED 

Balda,  R.  P.  and  a.  C.  Kamil.  1989.  A comparative 
study  of  cache  recovery  by  three  corvid  species. 
Anim.  Behav.  38:486-495. 

Bednkkohh,  P.  a.,  a.  C.  Kamil,  and  R.  P.  Balda. 
1997.  Clark’s  Nutcracker  (Aves:  Corvidae)  spatial 


memory:  interference  effects  on  cache  recovery 
performance?  Ethology  103:554-565. 

Bent,  A.  C.  1946.  Life  histories  of  North  American 
jays,  crows,  and  titmice.  Part  1.  Dover,  New  York. 

Camp,  R.  J.,  R.  L.  Knight,  H.  A.  L.  Knight,  M.  W. 
Sherman,  and  T.  Y.  Kawashima.  1993.  Food 
habits  of  the  nesting  Common  Ravens  in  the  east- 
ern Mojave  Desert.  Southwest.  Nat.  38:163-165. 

Engel,  K.  A.  and  L.  S.  Young.  1989.  Spatial  and 
temporal  patterns  in  the  diet  of  Common  Ravens 
in  southwestern  Idaho.  Condor  91:371—378. 

Ewins,  P.  J.,  J.  N.  Dymond,  and  M.  Marquiss.  1986. 
The  distribution,  breeding  and  diet  of  ravens,  Cor- 
vus  corax,  in  Shetlands.  Bird  Study  33:110-116. 

Heinrich,  B.  1995.  An  experimental  investigation  of 
insight  in  Common  Ravens  (Corvus  corax).  Auk 
1 12:994-1003. 

Heinrich,  B.  and  J.  W.  Pepper.  1998.  Influence  of 
competitors  on  caching  behavior  in  the  Common 
Raven  Connis  corax.  Anim.  Behav.  56:1083- 
1090. 

Hunt,  G.  1996.  Manufacture  and  use  of  hook-tools 
by  New  Caledonia  Crows.  Nature  379:249-251. 

Kamil,  A.  C.  and  R.  P.  Balda.  1985.  Cache  recovery 
and  spatial  memory  in  Clark’s  Nutcrackers  (No- 
cifraga  Columbiana).  J.  Exp.  Psychol.  1 1:95-1 1 1. 

Marquiss,  M.  and  C.  J.  Booth.  1986.  The  diet  of 
ravens,  Corx’us  corax,  in  Orkney.  Bird  Study  33: 
190-195. 

Marr,  V.  and  R.  L.  Knight.  1982.  Raven  predation 
of  feral  Rock  Dove  eggs.  Murrelet  63:25. 

Nelson,  A.  L.  1934.  Some  early  summer  food  pref- 
erences of  the  American  Raven  in  southwestern 
Oregon.  Condor  36:10-15. 

Ratcliffe,  D.  1997.  The  Raven.  Academic  Press  Inc., 
San  Diego,  California. 

Waite,  T.  A.  and  R.  C.  Ydenberg.  1994.  What  cur- 
rency do  scatter-hoarding  jays  maximize?  Behav. 
Ecol.  Sociobiol.  34:43—49. 

Zach,  R.  1979.  Shell-dropping:  decision  making  and 
optimal  foraging  in  Northwestern  Crows.  Behav- 
iour 68: 106-1 17. 


SHORT  COMMUNICATIONS 


279 


Wilson  Bull..  1 1 1(2),  1999,  pp.  279-281 


Pairing  Success  of  Wood  Thrushes  in  a Fragmented 
Agricultural  Landscape 

Lyle  E.  Friesen,'  - Valerie  E.  Wyatt,'  and  Michael  D.  Cadman' 


ABSTRACT — Habitat  fragmentation  has  been  as- 
sociated with  low  pairing  success  of  some  Neotropical 
migrant  songbirds  occupying  forest  fragments.  From 
1996  to  1998,  we  conducted  a nest  study  of  Wood 
Thrushes  (Hylocichia  miistelina)  in  21  woodlots  rang- 
ing in  size  from  3-12  ha  in  a highly  fragmented  ag- 
ricultural landscape  in  southwestern  Ontario.  We  found 
active  nests  for  46  of  48  singing  Wood  Thrushes  that 
we  detected  in  the  forest  fragments.  Our  results  suggest 
that  in  at  least  some  highly  fragmented  agricultural 
landscapes,  most  singing  Wood  Thrushes  in  small 
woodlots  are  successfully  paired.  Received  17  Sept. 
1998,  accepted  18  Jan.  1999. 


Habitat  fragmentation  has  been  associated 
with  low  pairing  success  of  some  Neotropical 
migrant  songbirds  residing  in  forest  fragments 
and  along  forest  edges.  For  example,  fewer 
territorial  male  Ovenbirds  (Seiurus  aurocap- 
illus)  were  paired  in  small  forests  than  in  large 
ones  in  Missouri  (Gibbs  and  Faaborg  1990, 
Van  Horn  et  al.  1995),  New  Jersey  (Wander 
1985),  Ontario  (Burke  and  Nol  1998),  and 
Quebec  (Villard  et  al.  1993).  Ziehmer  (1993, 
cited  in  Faaborg  et  al.  1995)  documented  low- 
er pairing  success  for  Red-eyed  Vireos  (Vireo 
olivaceus)  and  Wood  Thrushes  {Hylocichia 
mustelina)  around  clearings  in  large,  selec- 
tively logged  forests  in  Missouri. 

With  respect  to  Wood  Thrushes,  diminished 
pairing  success  may  not  apply  generally 
across  all  fragmented  landscapes.  We  report 
on  high  pairing  success  of  Wood  Thrushes  in 
forest  fragments  in  Waterloo  Region,  an  in- 
tensively farmed  landscape  with  14%  forest 
cover  and  where  the  mean  patch  size  of  wood- 
lots  was  12.8  ha  (±  18.3  SD).  (See  Friesen  et 
al.,  1999  for  a fuller  description  of  the  region- 
al landscape.) 

From  1996  to  1998,  as  part  of  a lai'ger  re- 


‘ Canadian  Wildlife  Service,  75  Farquhar  Street, 
Guelph,  ON  NIH  3N4,  Canada. 

- Corresponding  author; 

E-mail:  lyle. friesen @sympatico.ca 


gional  study  on  nesting  success  and  produc- 
tivity of  several  species  of  forest  birds  (Frie- 
sen et  al.  1999),  nest  searches  were  conducted 
in  21  woodlots  known  to  hold  Wood  Thrush- 
es; woodlot  size  ranged  from  3-12  ha  (x  = 
8.2  ± 3.3).  The  canopy  at  all  sites,  which  av- 
eraged 24  m in  height,  was  dominated  by  sug- 
ar maple  {Acer  sacchariim)  and  smaller 
amounts  of  white  ash  {Fraxinus  americana) 
and  American  beech  {Fagus  grandifolia).  Ma- 
ple and  ash  saplings,  alternate-leaved  dog- 
wood {Cornus  alternifolia),  and  red-benied 
elder  {Sambucus  pubens)  predominated  in  the 
openings  created  by  ongoing  and  recent  selec- 
tive logging  at  all  sites. 

Searches  for  singing  males  involved  four  to 
eight  early  morning  visits  to  each  site  begin- 
ning the  last  week  of  May  and  continuing  to 
20  June.  We  located  and  mapped  the  location 
of  all  singing  birds  at  each  woodlot  by  walk- 
ing parallel  transects  100  m apart  and  using 
taped  song  playbacks  (Yahner  and  Ross 
1995).  We  attempted  to  find  nests  for  all  sing- 
ing birds  using  four-  and  five-person  teams. 
Nest  searching,  conducted  within  a 200  m ra- 
dius of  the  singing  bird,  was  discontinued  if 
an  active  nest  (i.e.,  containing  eggs  or  young) 
was  not  found  within  15  person  hours.  Sing- 
ing birds  were  assumed  to  be  paired  if  an  ac- 
tive nest  was  found  in  their  vicinity. 

Although  many  singing  birds  were  detected 
after  20  June,  less  effort  was  expended  in  find- 
ing their  nests  because  of  the  time  constraints 
involved  in  monitoring  the  nests  found  pre- 
viously. In  addition,  first  broods  in  southern 
Ontario  generally  fledge  aiound  20  June  (Frie- 
sen, Wyatt,  and  Cadman,  unpubl.  data)  and 
singing  birds  encountered  thereafter  could 
have  been  unpaired  males  moving  about  be- 
tween nests  or  they  could  have  become  un- 
paired following  the  failure  of  eailier  nests 
(Roth  and  Johnson  1993).  We  were  less  con- 
fident of  locating  all  temtories  later  in  the 
breeding  season  because  birds  were  then  less 


280 


THE  WILSON  BULLETIN  • Vol.  HI.  No.  2.  June  1999 


likely  to  respond  to  taped  playbacks  (Friesen, 
Wyatt,  and  Cadman,  pers.  obs.)-  Consequent- 
ly, data  collected  after  20  June  were  not  in- 
cluded in  this  analysis. 

One  to  three  singing  males  were  found  in 
each  woodlot,  with  48  territorial  birds  detect- 
ed overall.  Active  nests  were  found  for  46 
(96%)  of  the  singing  Wood  Thrushes,  with  2.0 
(±  1.7)  person  hours  expended  on  average  to 
find  each  nest  following  the  detection  of  a 
singing  bird.  Nest  height  averaged  3.1  m (± 
1.5  m,  range  1.2-6. 4 m).  Our  estimate  of  pair- 
ing success  may  be  conservative  because  a 
new  but  empty  nest  was  found  in  the  vicinity 
of  one  of  the  two  “unpaired”  singing  males 
(suggesting  recent  predation)  and  it  is  possible 
that  we  missed  finding  the  nest  of  the  other 
“unpaired”  male. 

Although  Neotropical  migrants  can  experi- 
ence poor  pairing  success  in  fragmented  hab- 
itats, sensitivity  in  this  regard  likely  varies 
among  species.  Based  on  their  pairing  success 
in  small  fragments,  Ovenbirds  seem  to  be  par- 
ticularly sensitive  to  fragmentation  effects 
while  Wood  Thrushes  are  less  so.  Burke  and 
Nol  (1998)  speculated  that  the  lower  pairing 
success  of  Ovenbirds  in  small  forests  might 
be  attributed  to  the  absence  of  females  who 
avoid  these  areas  of  lower  food  abundance  in 
favor  of  larger  forests.  Wood  Thrushes,  but 
not  Ovenbirds,  commonly  inhabit  small  naral 
woodlots  in  our  study  area  (Friesen  et  al. 
1995).  As  both  species  are  ground  foragers 
dependent  upon  similar  types  of  prey  (Kauf- 
man 1996),  it  may  be  that  factors  other  than 
food  availability  or  perhaps  subtle  differences 
in  food  preferences  are  limiting  the  distribu- 
tion of  Ovenbirds  in  our  region. 

It  may  be,  too,  that  a species’  pairing  suc- 
cess varies  across  regions,  perhaps  in  response 
to  factors  such  as  differences  in  landscape 
configuration,  forest  structure,  disturbance  re- 
gimes, and  population  density.  In  Missouri, 
where  Wood  Thrushes  are  “forest  interior” 
species  that  only  occasionally  occupy  small 
woodlots  (Jacobs  and  Wilson  1997),  lower 
pairing  success  occurred  as  formerly  contin- 
uous forest  became  fragmented  by  logging  ac- 
tivities (Ziehmer  1993,  cited  by  Faaborg  et  al. 
1995).  In  our  highly  fragmented  landscape. 
Wood  Thrushes  exhibit  a high  level  of  pairing 
success.  Weinberg  and  Roth  (1998)  did  not 
explicitly  comment  on  the  pairing  success  of 


Wood  Thrushes  in  their  study  in  Delaware; 
however,  based  on  the  large  number  of  active 
nests  (120  over  two  years)  they  found  in  14 
forest  fragments  ranging  in  size  from  0.2— 2.1 
ha,  it  appears  that  Wood  Thrush  experience  a 
high  degree  of  pairing  success  in  other  frag- 
mented landscapes  as  well. 

Our  findings  on  pairing  success  do  not  nec- 
essarily imply  that  Wood  Thrushes  in  south- 
western Ontaiio  are  immune  to  fragmentation 
pressures.  For  example,  almost  half  of  the 
Wood  Thrush  nests  found  in  Waterloo  Region 
in  1996  and  1997  were  parasitized  by  Brown- 
headed Cowbirds  (Molothrus  ater)  resulting  in 
a significant  decline  in  host  productivity  (Frie- 
sen et  al.  1999).  With  I'espect  to  mating  status, 
however,  our  data  strongly  suggest  that  most 
singing  males  detected  in  forest  fragments 
early  in  the  breeding  season  are  likely  to  be 
paired. 

ACKNOWLEDGMENTS 

Parts  of  this  project  were  supported  by  the  Ontario 
Region  of  Environment  Canada’s  Canadian  Wildlife 
Service,  Human  Resources  Development  Canada,  En- 
vironmental Youth  Corps-Ontario,  Long  Point  Bird 
Observatory,  and  the  Regional  Municipality  of  Water- 
loo. 

LITERATURE  CITED 

Burke,  D.  M.  and  E.  Not.  1998.  Influence  of  food 
abundance,  nest-site  habitat,  and  forest  fragmen- 
tation on  breeding  Ovenbirds.  Auk  1 15:96—104. 
Faaborg,  J.,  M.  C.  Brittingham,  T.  M.  Donovan,  and 
J.  Blake:.  1995.  Habitat  fragmentation  in  the  tem- 
perate zone.  Pp.  357-380  in  Ecology  and  man- 
agement of  neotropical  migratory  birds  (T.  E. 
Martin  and  D.  M.  Finch,  Eds.).  Oxford  Univ. 
Press,  New  York. 

Friesen,  L.  E.,  M.  D.  Cadman,  and  R.  J.  Mackay. 
1999.  Nesting  success  of  neotropical  migrant 
songbirds  in  a highly  fragmented  landscape.  Con- 
serv.  Biol.  13: 1-9. 

Friesen,  L.  E.,  P.  F.  J.  Eagles,  and  R.  J.  Mackay. 
1995.  Effects  of  residential  development  on  for- 
est-dwelling neotropical  migrant  songbirds.  Con- 
serv.  Biol.  9:1408-1414. 

Gibbs,  J.  P.  and  J.  Faaborg.  1990.  Estimating  the 
viability  of  Ovenbird  and  Kentucky  Warbler  pop- 
ulations in  forest  fragments.  Conserv.  Biol.  4:193- 
196. 

Jacobs,  B.  and  J.  D.  Wilson.  1997.  Missouri  breed- 
ing bird  atlas.  Natural  History  Series.  No.  6.  Mis- 
souri Dcpartmenl  ol  Conservation,  Jefferson  City. 
Kaui-man,  K.  1996.  Lives  of  North  American  birds. 
Houghton  Mifflin  Company,  Boston,  Massasschu- 
.setts. 


SHORT  COMMUNICATIONS 


281 


Roth.  R.  R.  and  R.  K.  Johnson.  1993.  Loiig-ierm 
dynamics  of  a Wood  Thrush  population  breeding 
in  a forest  fragment.  Auk  1 10:37-48. 

Van  Horn,  M.  A.,  R.  M.  Gentry,  and  J.  Faaborg. 
1995.  Patterns  of  pairing  success  of  the  Ovenbird 
{Seiiiriis  aurocapillus)  within  Missouri  forest 
tracts.  Auk  1 12:98-106. 

ViLLARD,  M.,  P.  R.  Martin,  and  C.  G.  Drummond. 
1993.  Habitat  fragmentation  and  pairing  success 
in  the  Ovenbird.  Auk  1 10:759-768. 

Wander,  S.  A.  1985.  Comparative  breeding  biology 
of  the  Ovenbird  in  large  vs.  fragmented  forests: 


implications  for  the  conservation  of  Neotropical 
migrant  birds.  Pb.D.  diss.,  Rutgers  Univ.,  New 
Brunswick,  New  Jersey. 

Weinberg,  H.  J.  and  R.  R.  Roth.  1998.  Forest  area 
and  habitat  quality  for  nesting  Wood  Thrushes. 
Auk.  115:879-889. 

Yahner,  R.  H.  and  B.  D.  Ross.  1995.  Seasonal  re- 
sponse of  Wood  Thrushes  to  taped-playback 
songs.  Wilson  Bull.  107:738-741. 

Ziehmer,  R.  L.  1993.  Effects  of  uneven-aged  timber 
management  on  forest  bird  communities.  M.Sc. 
thesis,  Univ.  of  Missouri,  Columbia. 


Wilson  Bull.,  111(2),  1999,  pp.  281-282 

Connecticut  Warbler,  a North  American  Migrant  New  to  Ecuador 

Olaf  Jahn,'  - Maria  Eugenia  Jara  Viteri,'*  and  Karl-L.  Schuchmann-  '* 


ABSTRACT. — We  present  the  first  record  of  the 
Connecticut  Warbler  (Oporornis  agilis)  for  Ecuador. 
The  bird  was  mist-netted  and  photographed  on  21  No- 
vember 1996  at  Playa  de  Oro,  Rio  Santiago,  Esmer- 
aldas  Province,  northwestern  Ecuador.  Received  15 
Sept.  1998,  accepted  14  Nov.  1998. 


Between  August  1995  and  December  1996, 
we  studied  understory  bird  communities  near 
the  village  Playa  de  Oro  (00°  52'  N,  78°  47' 
W)  situated  at  about  50  m above  sea-level  on 
the  Rio  Santiago,  Esmeraldas  Province,  north- 
western Ecuador.  Since  February  1997  we 
have  carried  out  line  transect  censuses  in  dif- 
ferent habitat  types  from  50  m to  400  m in 
the  community  of  Playa  de  Oro,  close  to  the 
border  of  the  Cotocachi-Cayapas  ecological 
reserve.  On  the  morning  of  21  November 
1996,  a warbler  of  the  genus  Oporornis  was 
mist-netted  in  secondary  vegetation  near  a 
natural  backwater  pond  on  the  outskirts  of 
Playa  de  Oro.  The  bird  had  a complete,  pale. 


' Fundacion  Ecuatoriana  de  Estudios  Ecologicos, 
EcoCiencia,  Isla  San  Cristobal  1523  e Isla  Seymour, 
P.O.  Box  17-12-257,  Quito,  Ecuador. 

- Alexander  Koenig  Research  Institute  and  Museum 
of  Zoology,  Research  Group  “Biology  and  Phylogeny 
of  Tropical  Birds,”  Adenauerallee  160,  D-53  I 13  Bonn, 
Germany. 

^ Conocoto,  Olmedo  No  501  y Montarlo  Valle  de 
los  Chillos,  Quito,  Ecuador. 

■'  Corresponding  author. 


whitish-yellow  eye  ring,  very  long  undertail 
coverts,  a dull  brown  breast  band  encircling  a 
paler  throat,  a grayish  tinged  hood  and  grayish 
legs  (Fig.  1).  Referring  to  the  description  in 
Curson  and  coworkers  (1994),  we  determined 
that  the  bird  was  a first  winter  Connecticut 
Warbler  {Oporornis  agilis).  Its  measurements 
were:  total  length  (non-stretched)  125  mm, 
wing  65  mm,  tail  43  mm,  wing  — tail  22  mm, 
bill  length  11.95  mm,  tai'sus  19.65  mm,  body 
mass  12.5  g.  P7  and  p8  were  emarginated.  Af- 
ter the  bird  was  photographed  in  different  po- 
sitions (uppeiparts,  undeiparts,  laterally,  and 
with  opened  wing)  it  was  released.  The  pho- 
tographs of  the  wing  formula  (primaiies) 
show  p9  longer  than  p6.  The  difference  was 
not  measured  exactly  in  the  field.  Because  a 
few  female  and  immature  O.  Philadelphia  also 
have  complete  eye  rings,  it  was  necessary  to 
verify  the  identification  using  wing  and  tail 
measurements.  According  to  Lanyon  and  Bull 
(1967),  O.  agilis  can  usually  be  separated 
from  O.  Philadelphia  by  a wing  - tail  value 
equal  to  19  mm  or  more.  The  bird  we  captured 
showed  a wing  — tail  value  of  22  mm,  hence 
its  identification  as  O.  agilis  is  virtually  cer- 
tain. Photos  have  been  deposited  at  VIREO, 
where  the  identity  as  O.  agilis  has  been  con- 
firmed by  L.  Bevier  and  R.  S.  Ridgely  (pers. 
comm.).  Academy  of  Natural  Sciences,  Phil- 
adelphia. 

Oporornis  agilis  is  a rare  to  locally  uncom- 


282 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2.  June  1999 


FIG.  1.  Connecticut  Warbler,  Oporornis  agilis,  captured  at  Playa  de  Oro,  Esmeraldas  Province,  northwestern 
Ecuador. 


mon  boreal  winter  resident  south  and  east  of 
the  Andes,  from  eastern  Colombia  and  Vene- 
zuela south  to  eastern  Peru,  northern  Bolivia 
and  west  central  Brazil  (Curson  et  al.  1994, 
Ridgely  and  Tudor  1989).  Recent  evidence  (R. 
S.  Ridgley,  pers.  comm.)  indicates  that  the 
species  occurs  in  the  northern  part  of  South 
America  mainly  as  a transient,  but  not  during 
the  northern  mid-winter  months.  The  bird 
trapped  in  Playa  de  Oro  apparently  represents 
not  only  the  first  record  of  O.  agilis  west  of 
the  Andes  but  also  the  first  record  from  any- 
where in  Ecuador  (R.  S.  Ridgely,  pers. 
comm.),  although  it  had  been  expected  in  that 
country. 

ACKNOWLEDGMENTS 

This  work  is  a result  of  the  project  ‘Birds  as  Indi- 
cators for  Human-influenced  Tropical  Habitats’  and  is 
part  of  the  ‘Tropical  Ecology  Support  Program’  (TOB) 
of  the  Deutsche  Ge.sellschaft  fiir  Technische  Zusam- 
menarbeit  (GTZ),  Germany.  The  study  would  have 
been  impossible  without  the  financial  support  of  the 
Brehm-Funds  for  International  Bird  Conservation, 


Germany.  The  project  was  carried  out  in  cooperation 
with  the  Fundacion  para  el  Estudio  e Investigacion  de 
los  Colibries  Ecuatorianos  (FEICE),  Quito,  as  a na- 
tional partner,  with  the  Fundacion  Ecuatoriana  de  Es- 
tudios  Ecologicos  (EcoCiencia),  Quito,  as  a logistical 
partner,  and  with  the  logistic  help,  for  which  we  are 
most  grateful,  of  the  SUBIR  Project,  CARE/Ecuador, 
and  the  Deutscher  Entwicklungsdienst  (DED),  Ger- 
many. We  would  like  to  thank  the  Institute  Ecuatoriano 
Forestal  de  Areas  Naturales  y Vida  Silvestre  (INE- 
FAN),  Quito,  for  permission  to  work  in  Ecuador,  and 
the  community  of  Playa  de  Oro  for  permission  to  work 
on  their  land.  We  thank  L.  Bevier  and  R.  S.  Ridgely 
for  the  confirmation  of  the  identity  of  our  Oporornis 
agilis  photos  and  for  critical  comments  on  the  manu- 
script. 

LITERATURE  CITED 

CuR.soN,  J.,  D.  Quinn,  and  D.  Beadle.  1994.  New 
World  warblers.  Christopher  Helm,  London,  U.K. 
Lanyon,  W.  E.  and  J.  Bull.  1967.  Identification  of 
Connecticut,  Mourning,  and  MacGillivray’s  war- 
blers. Bird-Banding  38:187-194. 

Ridgely,  R.  S.  and  G.  Tudor.  1989.  The  birds  of 
South  America.  Vol.  1.  The  oscine  passerines.  Ox- 
ford Univ.  Press.  Oxford,  U.K.  . • 


SHORT  COMMUNICATIONS 


283 


Wilson  Bull..  1 I 1(2),  1999.  pp.  283-286 

Parental  Behavior  of  a Bigamous  Male  Northern  Cariiinal 

Randall  Breitwisch,'-^  Amy  J.  Schilling,'  - and  Joshua  B.  Banks' 


ABSTRACT. — Parental  behavior  of  a bigamous 
male  Northern  Cardinal  (Canlinalis  cardintiUs)  in 
southwestern  Ohio  in  1997  is  described.  The  male  was 
neither  brighter  in  plumage  nor  larger  than  average. 
Nesting  periods  of  the  two  females  overlapped.  The 
male  provisioned  the  primary  female  during  incubation 
but  not  the  secondary  female.  The  male  delayed  pro- 
visioning the  secondary  female’s  nestlings  until  two 
days  after  they  hatched  but  then  fed  both  sets  of  nest- 
lings at  rates  typical  of  monogamous  males.  Despite 
initially  reduced  paternal  care,  the  brood  of  the  sec- 
ondary female  fledged  successfully.  Received  12  Mar. 
1998.  accepted  15  Nov.  1998. 


The  majority  of  bird  species  are  socially 
monogamous  (Lack  1968),  the  hypothesis  be- 
ing that  ecological  constraints  explain  the  rel- 
ative infrequency  of  polygyny  in  birds  (Emlen 
and  Oring  1977).  At  the  same  time,  males  in 
most  species  of  socially  monogamous  birds 
may  have  the  behavioral  capability  of  becom- 
ing polygynous  should  ecological  conditions 
allow  multiple  mates  (Smith  et  al.  1982, 
Wingfield  1984).  The  description  of  infre- 
quent cases  of  bigamy  in  socially  monoga- 
mous species  is  relevant  to  any  discussion  of 
intraspecific  variability  in  mating  arrange- 
ments. The  behavior  of  bigamous  males  to- 
ward two  females  and  their  offspring  may  pro- 
vide information  on  the  costs  to  females  in- 
volved in  such  mating  arrangements. 

Northei’n  Cardinals  {Cardinali.s  cardinalis) 
are  socially  monogamous  and  sexually  dichro- 
matic; the  parental  behavior  of  bigamous 
males  has  not  been  previously  described  in  de- 
tail. Here,  we  document  bigamy  and  paternal 
care  by  a male  Northern  Cardinal  observed 
during  20  h over  a 10  day  period  in  early  June 
1997. 

The  three  cardinals,  all  of  unknown  age. 


' Dept,  of  Biology,  Univ.  of  Dayton.  Dayton,  OH 
45469-2320. 

- Present  address:  Dept,  of  Anthropology.  Univ.  of 
California,  Davis,  CA  95616-8522. 

’’  Corresponding  author; 

E-mail:  breit@neelix.udayton.edu 


were  members  of  a color-banded  population 
located  at  Aullwood  Audubon  Center,  15  km 
northwest  of  Dayton,  Ohio  (39°  52'  N,  84°  16' 
W)  and  under  continuous  observation  since 
1991.  The  80  ha  property  is  a mixture  of  de- 
ciduous woodlands,  meadows,  and  prairies 
where  cardinals  are  abundant.  The  male  that 
became  bigamous  in  1997  was  banded  in  the 
spring  of  1996.  That  year  he  was  successful 
over  several  others  competing  for  a territory 
that  had  been  occupied  for  several  years  by  a 
male  that  disappeared  over  the  1995-1996 
winter.  The  territory  was  one  of  the  largest  in 
the  study  area  and  among  those  with  the  most 
plant  cover,  a variable  that  might  provide  an 
advantage  to  nesting  success  (Conner  et  al. 
1986,  Wolfenbarger  1996;  however,  see  Fillia- 
ter  et  al.  1994).  The  male  enlarged  this  terri- 
tory in  1997  and  it  became  the  site  of  the  big- 
amous mating. 

One  of  the  two  females  paired  with  this 
male  in  1997  (9  650)  was  banded  in  1996  on 
a tenitory  adjacent  to  the  one  he  occupied  in 

1996.  This  female’s  mate  disappeared  in  the 
non-breeding  season  1996-1997,  as  did  the 
mate  of  the  [bigamous]  male.  Female  650  re- 
mained on  the  same  territory  in  1997  and  the 
bigamous  male  expanded  his  1996  territory  to 
include  the  area  occupied  by  female  650.  In 

1997,  this  female  was  treated  differently  by 
the  male  from  the  manner  in  which  females 
of  monogamous  males  are  treated  by  their 
mates  (described  below),  and  for  these  reasons 
we  consider  her  the  “secondary”  female  of 
the  bigamous  male.  The  other  female  (9  555) 
had  been  banded  two  years  previously  in  an 
area  two  territories  away  from  the  bigamous 
male’s  1996-1997  territory.  She  mated  in 
1995  but  was  not  found  in  1996;  her  1995 
mate  retained  the  same  territory  in  1996  and 
mated  with  another  female.  Female  555  then 
reappeared  in  1997  on  the  territory  from 
which  the  bigamous  male’s  1996  mate  had 
disappeared.  The  bigamous  male  treated  fe- 
male 555  similarly  to  how  monogamous 


284 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


males  treat  their  mates,  and  we  consider  her 
the  “primary”  mate  of  the  bigamous  male. 
We  do  not  know  which  female  first  associated 
with  the  bigamous  male  in  1997. 

Male  cardinals  provision  their  mates  during 
courtship,  egg-laying,  and  incubation  (Lemon 
1968;  Vemer  and  Willson  1969;  Breitwisch, 
Banks,  Donahoo,  LeClair,  and  Schilling,  un- 
publ.  data);  this  is  considered  a form  of  indi- 
rect parental  care  (Lyon  and  Montgomerie 
1985).  We  sampled  provisioning  behavior  of 
the  bigamous  male  toward  both  females  dur- 
ing six  1-h  observations  on  six  days  of  the 
12-d  incubation  period.  The  nests,  located  ap- 
proximately 60  m apart  and  out  of  sight  of 
each  other,  were  monitored  simultaneously 
during  these  observation  periods,  which  in- 
cluded both  mornings  and  afternoons.  Neither 
age  of  eggs  nor  time  of  day  influences  rate  of 
mate  provisioning  by  male  cardinals  (Breit- 
wisch, Banks,  Donahoo,  LeClair,  and  Schil- 
ling, unpubl.  data).  Primary  female  555  began 
nesting  a few  days  before  secondary  female 
650,  and  the  bigamous  male  provisioned  fe- 
male 555  at  a similar  rate  (x  = 1 .00  feeding/ 
h)  to  the  mean  value  for  the  monogamous 
population  [x  = 1.05  ± 0.48  (SD)  feeding/h; 
Breitwisch,  Banks,  Donahoo,  LeClair,  and 
Schilling,  unpubl.  data].  However,  female  650 
received  no  food  from  the  bigamous  male. 
This  is  in  marked  contrast  to  a sample  of  18 
monogamous  males  in  the  population,  all  of 
which  provisioned  their  mates  during  incu- 
bation (Breitwisch,  Banks,  Donahoo,  LeClair, 
and  Schilling,  unpubl.  data).  Although  the  big- 
amous male  did  not  provision  female  650,  he 
remained  active  throughout  his  tenitory  and 
did  not  appear  to  favor  the  area  near  the  nest 
of  female  555. 

Male  cardinals  feed  nestlings  at  high  rates, 
frequently  surpassing  the  rate  at  which  fe- 
males feed  nestlings  (Filliater  and  Breitwisch 
1997,  Linville  et  al.  1998).  We  sampled  nest- 
ling feedings  by  the  bigamous  male  and  the 
two  females  during  l-h  samples  on  seven  days 
at  female  555’s  nest  and  six  days  at  female 
65()’s  nest.  Sampling  periods  on  four  of  the 
days  were  coincident  (see  below).  Observa- 
tion periods  included  mornings  and  after- 
noons, although  Filliater-Lee  (1992)  showed 
that  neither  male  nor  female  feeding  rate  is 
related  to  time  of  day.  The  eggs  of  female  555 
hatched  three  days  before  the  eggs  of  female 


650.  The  male  fed  the  two  nestlings  of  female 
555  at  a mean  rate  of  1.3  feeding/nestling/h, 
similar  to  the  feeding  rate  by  monogamous 
males  (T  = 1.1  ± 0.53  feeding/nestling/h;  Fil- 
liater and  Breitwisch  1997).  Female  555  fed 
her  nestlings  at  a mean  rate  of  0.57  feeding/ 
nestling/h,  similar  to  the  feeding  rate  by  fe- 
males mated  to  monogamous  males  (x  = 0.87 
± 0.38;  Filliater  and  Breitwisch  1997).  How- 
ever, the  male  did  not  begin  feeding  the  two 
nestlings  of  female  650  until  two  days  after 
they  hatched.  Female  650  fed  her  nestlings  at 
a mean  rate  of  1.0  feeding/nestling/h,  similar 
to  that  of  females  mated  to  monogamous 
males.  When  the  male  began  to  feed  female 
650’s  nestlings,  he  fed  them  at  a mean  rate  of 
1.0  feeding/nestling/h,  similar  to  the  rate  of 
feeding  by  monogamous  males.  During  four 
days,  nestlings  were  present  in  both  nests  and 
the  male  fed  nestlings  at  both,  roughly  alter- 
nating his  deliveries  to  the  two  sets  of  nest- 
lings. 

The  fates  of  these  two  nests  differed.  The 
nestlings  of  the  primary  female  were  preyed 
upon  a few  days  before  they  would  have 
fledged,  but  the  nestlings  of  the  secondary  fe- 
male fledged  successfully.  We  do  not  know 
whether  the  three  adults  maintained  the  biga- 
mous relationship  throughout  the  season. 

We  determined  that  the  bigamous  male  was 
neither  exceptionally  ornamented  nor  notably 
large  in  body  size.  Using  a technique  de- 
scribed by  Linville  and  coworkers  (1998),  we 
measured  the  brightness  of  the  red  breast 
plumage  of  the  male  and  of  the  red  underwing 
plumage  of  the  two  females.  The  bigamous 
male  and  one  other  male  were  tied  as  the  dull- 
est in  a sample  of  14  males  in  1997.  The  big- 
amous male  was  also  of  average  body  size,  as 
measured  by  both  tarsus  and  flattened  wing 
arc  (R.  Breitwisch  and  S.U.  Linville,  unpubl. 
data).  The  females  were  both  found  to  be  at 
least  equal  to  the  median  plumage  brightness 
of  15  females  in  1997.  Primary  female  555 
was  one  score  lower  in  brightness  than  sec- 
ondary female  650.  We  lack  size  measure- 
ments of  the  two  females. 

Our  observations  suggest  that  there  are  at 
least  potential  costs  for  a secondaiy  female 
mated  to  a bigamous  male  cardinal.  Most  dra- 
matically, the  bigamous  male  failed  to  provi- 
sion the  secondary  female  during  incubation. 
A monogamous  male  typically  provides  ap- 


SHORT  COMMUNICATIONS 


285 


proxiniately  150  feedings  at  the  nest  during 
the  12-d  incubation  period  and  probably  sup- 
plies the  female  with  a significant  amount  of 
food  away  from  the  nest  (Breitwisch,  Banks, 
Donahoo,  LeClair,  and  Schilling,  unpubl. 
data).  The  high  rate  of  mate  provisioning  in- 
dicates that  the  amount  of  food  provided  may 
be  important  to  the  female’s  nutritional  state, 
especially  when  considering  the  three  or  more 
clutches  of  eggs  laid  by  a typical  female  in 
this  population  during  a breeding  season  (Fil- 
liater  et  al.  1994).  The  bigamous  male’s  be- 
havior toward  the  secondary  female’s  nest- 
lings was  not  typical  of  monogamous  males 
in  this  population.  Although  the  male  even- 
tually began  to  feed  the  nestlings  and  did  so 
at  a rate  typical  for  monogamous  males,  he 
delayed  two  days  after  these  nestlings  hatched 
before  beginning  to  feed  them. 

The  primary  contribution  of  male  cardinals 
to  raising  young  appears  to  be  pi'ovisioning 
the  female,  nestlings,  and  fledglings.  Guarding 
and  active  defense  against  predators  are  of  mi- 
nor importance  and  effectiveness  (Filliater  et 
al.  1994,  Nealen  and  Breitwisch  1997).  Thus, 
we  think  it  unlikely  that  any  reduced  level  of 
these  components  of  paternal  care  were  a sig- 
nificant additional  cost  of  bigamy  to  either  fe- 
male. 

Bigamy  in  cardinals  appears  to  be  quite  rare 
(see  Linville  and  Halkin,  in  press).  Lemon 
(1968)  observed  two  cases  in  which  he  noted 
that  the  females  “less  tended’’  by  the  biga- 
mous males  eventually  left  and  were  probably 
unsuccessful  (R.  E.  Lemon,  pers.  comm.).  In 
our  own  studies,  the  instance  of  bigamy  de- 
scribed here  is  the  first  witnessed  in  seven 
years  of  monitoring  mating  relationships  in 
this  population.  Each  of  the  last  six  years,  we 
have  observed  an  average  of  about  20  terri- 
tories, suggesting  that  the  incidence  of  bigamy 
is  probably  less  than  5%  [Verner  and  Will- 
son’s (1969)  criterion  for  monogamy]  and 
may  be  even  less  than  1%.  Two  other  re- 
searchers have  not  observed  bigamy  in  multi- 
year studies  with  a combined  sample  size  of 
more  than  50  pairs  (G.  Ritchison,  pers. 
comm.;  L.  L.  Wolfenbarger,  pers.  comm.).  D. 
M.  Scott  (pers.  comm.)  and  R.  E.  Lemon 
(1957,  1968,  pers.  comm.)  have  records  of  at 
least  three  bigamous  males  in  multi-year  stud- 
ies of  more  than  50  pairs  of  cardinals,  al- 


though Scott  (pers.  comm.)  agrees  with  the 
above  estimate  of  less  than  a 5%  incidence. 

It  has  been  hypothesized  that  staggered  tim- 
ing of  nesting  by  two  females  mated  to  a big- 
amous male  may  be  critical  to  reducing  the 
cost  of  bigamy  to  the  females  (Verner  1964, 
Breitwisch  et  al.  1986,  Deirickson  1989).  Big- 
amous males  should  be  able  to  apportion  care 
more  easily  when  nests  do  not  overlap  in  time. 
Obviously,  we  cannot  know  if  the  bigamous 
cardinal  would  have  provisioned  the  second- 
ary female  if  her  incubation  period  had  not 
overlapped  with  that  of  the  primary  female. 
Second,  with  staggered  nesting,  a female  oc- 
cupied with  caring  for  eggs  or  nestlings  might 
display  reduced  aggression  toward  a second 
female  attempting  to  nest  (Derrickson  1989). 
Although  female  cardinals  can  be  very  ag- 
gressive toward  other  females  (R.  Breitwisch, 
pers.  obs.),  we  did  not  witness  aggression  be- 
tween the  two  females  we  observed. 

In  any  case,  the  secondary  female  cardinal 
was  successful  in  producing  fledglings  despite 
limited  paternal  care.  Richmond  (1978)  re- 
moved male  cardinals  from  nesting  pairs  and 
also  found  that  females  were  able  to  raise 
young  by  themselves.  We  speculate,  as  did 
Richmond,  that  neglected  females  may  still 
pay  a cost  in  future  survival  from  such  high 
parental  effort. 

The  question  that  remains  is  whether  a sec- 
ondary female  in  a bigamous  relationship  is 
making  the  coirect  decision  at  the  time  of 
pairing  with  a male  or  committing  an  enor.  It 
seems  likely  that  there  is  no  single  answer  to 
this  question.  In  some  monogamous  species, 
secondary  females  may  be  able  to  “predict” 
that  their  young  will  receive  paternal  provi- 
sioning [e.g..  Northern  Shrikes,  Lanius  exciib- 
itor,  and  Loggerhead  Shrikes,  L.  ludovicianiis, 
(Yosef  1992)].  In  others,  lack  of  paternal  pro- 
visioning may  be  equally  predictable  [e.g.. 
Song  SpaiTOWS,  Melospiza  melodia  (Smith  et 
al.  1982),  Florida  Scrub  Jays,  Aphelocoma  c. 
coendescens  (Woolfenden  1976)],  or  paternal 
provisioning  may  depend  on  degree  of  overlap 
in  nesting  [e.g..  Northern  Mockingbirds,  Mi- 
mas polyglottos  (Logan  and  Rulli  1981,  Breit- 
wisch et  al.  1986)].  Moreover,  there  are  other 
factors  that  may  be  involved  in  determining 
level  of  paternal  care  at  nests  of  secondary 
females,  both  in  species  that  are  opportunis- 
tically bigamous  and  those  that  ai'e  typically 


286 


THE  WILSON  BULLETIN 


Vol.  Ill,  No.  2,  June  1999 


more  polygynous.  These  include  the  degree  to 
which  an  aspect  of  paternal  care  is  shareable, 
the  age  and  number  of  nestlings,  and  the 
male’s  confidence  of  paternity  (Searcy  and 
Yasukawa  1995).  Explanations  of  such  varied 
patterns  will  await  additional  reports  on  big- 
amous relationships  in  socially  monogamous 
birds. 

ACKNOWLEDGMENTS 

We  thank  P.  Donahoo  and  J.  LeClair  for  assistance 
in  the  field  and  D.  Scott  and  R.  Lemon  for  providing 
unpublished  data.  C.  Logan,  J.  Jawor,  and  an  anony- 
mous reviewer  offered  useful  suggestions  on  the  man- 
uscript. C.  Krueger,  Director  at  Aullwood  Audubon 
Center  and  Farm,  and  J.  Ritzenthaler,  Head  of  Re- 
search, granted  permission  to  conduct  this  study  on  the 
Aullwood  property.  This  study  was  conducted  under 
USFWS  Banding  Permit  No.  22351  and  ODNR  Band- 
ing Permit  No.  5-57-04,  both  issued  to  RB. 

LITERATURE  CITED 

Brkitwisch,  R.,  R.  C.  Ritter,  and  J.  Zaias.  1986. 
Parental  behavior  of  a bigamous  male  Northern 
Mockingbird.  Auk  103:424—427. 

Conner,  R.  N.,  M.  E.  Anderson,  and  J.  G.  Dickson. 
1986.  Relationships  among  territory  size,  habitat, 
song,  and  nesting  success  of  Northern  Cardinals. 
Auk  103:23-31. 

Derrickson,  K.  C.  1989.  Bigamy  in  Northern  Mock- 
ingbirds: circumventing  female-female  aggres- 
sion. Condor  91:728-732. 

Emlen,  S.  T.  and  L.  W.  Oring.  1977.  Ecology,  sexual 
selection,  and  the  evolution  of  mating  systems. 
Science  197:215—223. 

Filliater,  T.  S.  and  R.  Breitwisch.  1997.  Nestling 
provisioning  by  the  extremely  dichromatic  North- 
ern Cardinal.  Wilson  Bull.  109:145-153. 
Filliater,  T S.,  R.  Breitwlsch,  and  P.  M.  Nealen. 
1994.  Predation  on  Northern  Cardinal  nests:  does 
choice  of  nest  site  matter?  Condor  96:761-768. 
Filliater-Lee,  T.  S.  1992.  Parental  roles  in  feeding 
nestlings,  and  nest  sites  and  nest  success  in  North- 
ern Cardinals  (Cardinalis  cardinalis).  M.Sc.  the- 
sis, Univ.  of  Dayton,  Dayton,  Ohio. 

Lack.  D.  1968.  Ecological  adaptations  for  breeding 
in  birds.  Methuen,  London,  U.K. 


Lemon,  R.  E.  1957.  A study  of  nesting  cardinals 
(Richmondena  cardinalis)  at  London,  Canada. 
M.Sc.  thesis,  Univ.  of  Western  Ontario,  London. 

Lemon,  R.  E.  1968.  The  displays  and  call  notes  of 
cardinals.  Can.  J.  Zool.  46:141—151. 

Linville,  S.  U.,  R.  Breitwisch,  and  A.  J.  Schilling. 
1998.  Plumage  brightness  as  an  indicator  of  pa- 
rental care  in  northern  cardinals.  Anim.  Behav.  55: 

1 19-127. 

Linville,  S.  U.  and  S.  L.  Halkin.  In  press.  Northern 
Cardinal  {Cardinalis  cardinalis).  In  The  birds  of 
North  America  (A.  Poole  and  F.  Gill,  Eds.).  The 
Academy  of  Natural  Sciences,  Philadelphia,  Penn- 
sylvania; The  American  Ornithologists’  Union, 
Washington,  D.C. 

Lyon,  B.  E.  and  R.  D.  Montgomerie.  1985.  Incuba- 
tion feeding  in  snow  buntings:  female  manipula- 
tion or  indirect  male  parental  care?  Behav.  Ecol. 
Sociobiol.  17:279-284. 

Nealen,  P.  M.  and  R.  Breitwisch.  1997.  Northern 
Cardinal  sexes  defend  nests  equally.  Wilson  Bull. 
109:269-278. 

Richmond,  A.  W.  1978.  An  experimental  study  of  ad- 
vantages of  monogamy  in  the  cardinal.  Ph.D. 
diss.,  Indiana  Univ.,  Bloomington. 

Searcy,  W.  A.  and  K.  Yasukawa.  1995.  Polygyny 
and  sexual  selection  in  Red-winged  Blackbirds. 
Princeton  Univ.  Press,  Princeton,  New  Jersey. 

Smith,  J.  N.  M.,  Y.  Yom-Tov,  and  R.  Moses.  1982. 
Polygyny,  male  parental  care,  and  sex  ratio  in 
Song  Sparrows:  an  experimental  study.  Auk  99: 
555-564. 

Verner,  j.  1964.  Evolution  of  polygyny  in  the  Long- 
billed Marsh  Wren.  Evolution  18:252-261. 

Verner,  J.  and  M.  E Willson.  1969.  Mating  systems, 
sexual  dimorphism,  and  the  role  of  male  North 
American  passerine  birds  in  the  nesting  cycle.  Or- 
nithol.  Monogr.  9:1-76. 

Wingfield,  J.  C.  1984.  Androgens  and  mating  sys- 
tems: testosterone-induced  polygyny  in  normally 
monogamous  birds.  Auk  101:665-671. 

Wolfenbarger,  L.  L.  1996.  Fitness  effects  associated 
with  red  coloration  of  male  Northern  Cardinals 
(Cardinalis  cardinalis).  Ph.D.  diss.,  Cornell  Univ., 
Ithaca,  New  York. 

Woolfenden,  G.  E.  1976.  A case  of  bigamy  in  the 
Florida  Scrub  Jay.  Auk  93:443-450. 

Yosef,  R.  1992.  Behavior  of  polygynous  and  monog- 
amous Loggerhead  Shrikes  and  a comparison  with 
Northern  Shrikes.  Wilson  Bull.  104:747—749. 


Wilson  Bull.,  111(2),  1999,  pp.  287-293 


Special  Report 

A SURVEY  OF  UNDERGRADUATE  ORNITHOLOGY  COURSES  IN 

NORTH  AMERICA 

EDWARD  H.  BURTT,  JR.'-^  AND  W.  HERBERT  WILSON,  JR.^ 


ABSTRACT. — The  Committee  on  Undergraduate  Education  of  the  Wilson  Ornithological  Society  conducted 
a survey  of  ornithology  courses  in  North  America  as  a service  for  teachers  of  ornithology.  Our  survey  of  26 
responses  uncovered  26  creative  approaches  to  teaching  ornithology.  Nonetheless,  a number  of  commonalities 
exist.  Courses  at  small  colleges  and  large  universities  include  both  lecture  and  laboratory  components  and  usually 
extend  into  the  spring.  Most  courses  emphasize  anatomy  and  physiology,  nesting,  evolution  of  birds,  ecology, 
and  flight,  with  other  topics  receiving  few  or  no  lectures.  Almost  60%  of  the  courses  include  student  dissection 
or  faculty  demonstration.  Some  courses  use  preserved  birds,  others  use  birds  that  died  accidentally,  and  one  uses 
roasted  chickens  that  are  eaten  as  part  of  the  skeleto-muscular  dissection.  Laboratory  sessions  emphasize  tax- 
onomy and  identification  of  local  and,  often,  world  birds.  Most  schools  have  at  least  a small  collection  of 
specimens  available  for  student  use.  Courses  usually  include  an  extensive  project  and  written  work.  We  hope 
the  results  of  the  survey  will  stimulate  discussion  among  teachers  of  ornithology  as  we  seek  to  develop  new 
ideas  for  our  courses.  Received  29  Sept.  1997,  accepted  8 Jan.  1999. 


The  Wilson  Ornithological  Society’s  Com- 
mittee on  Undergraduate  Education  seeks  to 
increase  the  quality  of  teaching  of  ornithology 
at  the  undergraduate  level  and  to  foster  com- 
munication among  ornithology  teachers  about 
successful  and  unsuccessful  aspects  of  their 
courses.  With  these  goals  in  mind,  we  pre- 
pared a questionnaire  that  was  sent  to  all  or- 
nithology faculty  who  responded  to  a request 
printed  in  the  Ornithological  Societies  of 
North  America  newsletter.  The  following  is  a 
synthesis  of  the  information  provided  by  the 
26  ornithologists  who  completed  the  question- 
naire in  1993  and  1994.  Some  respondents  left 
one  or  more  questions  unanswered,  thus  our 
analysis  of  some  questions  is  based  on  fewer 
than  26  responses. 

The  questionnaire  included  demographic 
and  course  content  questions.  Copies  of  all 
completed  responses  are  available  from  the 
Van  Tyne  Library  at  the  University  of  Mich- 
igan. We  first  describe  the  demographics  of 
our  sample,  then  summarize  the  quantitative 
data,  and  close  with  a discussion  of  successful 
and  unsuccessful  aspects  of  the  courses. 


' Dept,  of  Zoology,  Ohio  Wesleyan  Univ.,  Delaware, 
OH  43015. 

^ Dept,  of  Biology,  Colby  College,  Waterville,  ME 
04901. 

^ Corresponding  author; 

E-mail:  ehburtt@cc.owu.edu 


DEMOGRAPHICS  OF  THE 
RESPONDENTS’  INSTITUTIONS 

Our  small  sample  is  not  amenable  to  mul- 
tivariate analysis.  Furthermore,  because  it  is 
based  on  only  26  respondents,  our  survey  may 
be  biased.  Our  intent  is  to  document  the  di- 
versity of  approaches  and  stimulate  discus- 
sion. 

The  26  responses  to  the  survey  came  from 
ornithology  teachers  in  18  states  and  1 Ca- 
nadian province.  Thirteen  of  the  respondents 
teach  at  schools  in  the  Eastern  Time  zone,  1 1 
at  schools  in  the  Central  Time  zone,  and  2 in 
schools  in  the  Mountain  Time  zone.  We  re- 
ceived no  data  from  faculty  teaching  at 
schools  in  the  Western  Time  zone.  Of  the  24 
schools  in  the  Eastern  and  Central  Time 
zones,  5 are  in  southern  states.  Fifteen  of  the 
colleges  and  universities  aie  located  in  small 
towns,  whereas  1 1 have  suburban  or  urban 
campuses.  Sixteen  of  the  respondents  teach  at 
state-supported,  public  institutions;  7 teach  at 
privately  supported,  non-denominational  col- 
leges or  universities;  and  3 teach  at  church- 
affiliated  colleges.  Twelve  of  the  schools  offer 
the  Ph.D.,  4 the  M.Sc.  as  their  highest  degree, 
and  the  remaining  10  offer  only  bachelors’  de- 
grees. Twelve  schools  have  more  than  10,000 
undergraduates,  hereafter  referred  to  as  large 
schools,  and  14  schools  have  fewer  than 
10,000  (small  schools);  7 of  these  have  2,000 
or  fewer. 


287 


288 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


TABLE  1.  Quantitative  comparison  of  enrollment  as  affected  by  prerequisites  tor  ornithology  courses  taught 
at  large  and  small  schools. 


Prerequisites 

Large  schools 
(>10,000) 

Small  schools 
(£10,000) 

No.  of 
schools 

Mean  class  size  {±SD) 
(1988-1993) 

No.  of  schools 

Mean  class  size  (±SD) 
1988-1993) 

None 

1 

50.8  ± 6.8 

2 

21.3  ± 2.7 

1 semester  biology 

7 

30.9  ± 20.0 

8 

12.6  ± 4.5 

2+  semesters  biology 

4 

26.6  ± 25.7 

4 

17.0  ± 12.5 

Overall  class  size 

12 

31.4  ± 21.7 

14 

14.9  ± 8.1 

THE  COURSE 

Goals. — The  generally  stated  goal  of  the 
classroom  portion  of  the  course  was  to  pro- 
vide students  with  a broad  overview  of  orni- 
thology, and  to  use  birds  as  examples  of  fun- 
damental concepts  of  biology  and,  to  a lesser 
extent,  cognate  disciplines.  A second  goal  was 
to  use  birds  to  illustrate  the  scientific  process, 
which  includes  hypothesis  testing  and  stimu- 
lation of  new  ideas  through  debate  between 
scientists  with  different  interpretations  of  the 
data.  A third  goal,  emphasized  by  three  re- 
spondents, was  to  build  a genuine  admiration 
of  the  many  adaptations  of  birds,  thereby  pro- 


TABLE 2.  Quantitative  comparison  of  some  char- 
acteristics of  ornithology  courses  taught  at  large  and 
small  schools  (based  on  number  of  students  enrolled). 

Characleristics 

Large 

schools 

(>10,000) 

Small 

schools 

(£10,000) 

Lrequency 

Annually 

10 

5 

Biannually 

2 

8 

2-3  yr  interval 

0 

1 

Duration 

14-16  wks 

10 

11 

13  wks 

1 

1 

10  wks 

0 

1 

7-8  wks 

1 

1 

Lecture  h/wk 

4 

0 

1 

3 

3 

10 

2 

7 

3 

1 

1 

0 

Use  of  primary  literature 

Assigned  to  students 

6 

10 

Not  assigned  to  students 

6 

4 

Labs/wk/student 

0 

1 

0 

1 

6 

1 1 

2 

4 

3 

3 

1 

0 

moting  the  life-long  study  and  enjoyment  of 
birds. 

Goals  for  the  laboratory  portion  of  the 
course  were  more  varied.  Most  faculty  saw 
field  identification  by  sight  and  song  as  a pri- 
mary goal.  Family  and  order  names  were  con- 
sidered part  of  identification  by  most  respon- 
dents. Providing  students  with  a working 
knowledge  of  topography  and  anatomy,  par- 
ticularly of  feathers,  was  another  common 
goal.  Six  respondents  sought  to  instill  appre- 
ciation of  behavior  and  ecology  through  field 
experiences.  Another  six  respondents  indicat- 
ed that  a goal  of  the  laboratory  was  to  intro- 
duce students  to  field  (e.g.,  banding,  census- 
ing,  recording  of  vocalizations,  etc.)  or  mu- 
seum (e.g.,  preparation  and  measurement  of 
study  skins)  techniques. 

Structure. — One  semester  of  introductory 
biology  or  zoology  was  a common  prerequi- 
site although  no  prerequisite  and  two  or  more 
prerequisites  also  occun'ed.  Class  size  varied 
significantly  with  the  number  of  prerequisites 
at  both  large  (Fj  53  = 4.71,  P < 0.05)  and 
small  (Fj  49  = 4.05,  P < 0.05)  colleges  and 
universities  (Table  1).  Ornithology  classes 
with  a single  prerequisite  had  smaller  enroll- 
ments in  schools  of  both  sizes  (large:  t = 3.78, 
df  = 38,  P < 0.001;  small:  t = 3.74,  df  = 35, 
P < 0.001)  than  courses  with  no  prerequisite. 
Enrollment  did  not  decline  further  with  a sec- 
ond or  third  prerequisite  (Table  1).  Overall, 
class  size  at  small  schools  was  significantly 
less  it  = -5.67,  df  = 106,  P < 0.001)  than 
at  large  schools  (Table  1). 

Faculty  at  large  schools  were  more  likely 
(X^  = 6.09,  df  = 1,  P < 0.05)  to  offer  orni- 
thology annually  than  those  at  small  schools, 
but  the  duration  of  the  courses  varied  similarly 
among  schools  of  different  sizes  (Table  2).  All 
but  three  of  the  courses  were  taught  in  the 


liurtt  anil  Wilson  • SURVEY  OF  ORNITHOLOCJY  COURSES 


289 


Flight 


Systematics 


Nesting 


Ecology 


Migration, 

Orientation 


Conservation 


Anatomy, 

Physiology 


Vocalization 


Evolution 
Biogeography 

FIG.  I . The  proportion  of  the  “consensus”  ornithology  course  devoted  to  the  subjects  indicated. 


spring  semester  beginning  in  January  when 
students  learn  to  identify  waterfowl,  raptors, 
and  relatively  few  winter  residents.  Later  in 
the  course,  as  their  field  skills  improve,  stu- 
dents are  exposed  to  an  increasing  diversity  of 
spring  migrants  and  summer  residents. 

Twenty-four  courses  included  both  lecture- 
discussion  and  laboratory.  Lecture-discussion 
sections  met  twice  weekly  for  75  min/meeting 
in  12  courses,  three  times/week  for  50  min/ 
meeting  in  8 courses,  and  as  1 three-hour  sem- 
inar in  3 courses.  Faculty  at  small  schools  pro- 
vided more  hours  of  lecture-discussion/week 
(X"  = 6.75,  df  = 1,  P < 0.01,  Table  2)  than 
those  at  large  schools.  Faculty  expressed  sat- 
isfaction with  the  longer  class  period  and  with 
occasions  when  lecture  and  laboratory  could 
be  integrated. 

The  typical  class  period  was  what  one  re- 
spondent characterized  as  a “loose  lecture,”  a 
mix  of  lecture  and  discussion,  illustrated  with 
specimens,  slides  and  video  tapes,  and  punc- 
tuated with  questions  from  the  teacher.  Sev- 
eral faculty  indicated  plans  to  incorporate 
software  in  the  future,  but  none  were  using 
computers  in  the  classroom  in  1993  and  1994 
when  the  survey  was  completed. 

Content. — Two  courses  used  only  the  pri- 
mary literature,  24  courses  required  texts,  9 of 
these  required  two  texts,  and  1 required  three. 
Gill’s  (1990;  the  survey  was  completed  just 
before  the  second  edition)  Ornithology  was 
the  preferred  text  by  a wide  margin  (Appen- 
dix), but  other  texts  were  used.  In  addition  to 


a text,  15  respondents  assigned  their  students 
readings  from  the  primary  literature.  No  dif- 
ference in  use  of  the  primary  liteiature  was 
evident  among  schools  of  different  sizes  (Ta- 
ble 2). 

Each  respondent  was  asked  to  provide  a 
syllabus  of  his  or  her  course.  We  assigned  the 
lectures  to  1 1 broadly  defined  topics.  The 
mean  proportion  devoted  to  each  topic  by  all 
respondents  is  shown  in  Fig.  1.  Some  eiTor 
was  unavoidable  as  we  tried  to  categorize  lec- 
tures into  the  eleven  topics.  Nevertheless,  this 
figure  represents  the  “consensus”  course  of 
the  surveyed  teachers. 

Most  courses  had  one  laboratory  session/ 
student/week  (Table  2)  and  in  most  courses  it 
was  a mix  of  indoor  and  outdoor  sessions. 
Laboratory  schedules  were  similai'  at  large 
and  small  schools  (Table  2). 

Nineteen  courses  devoted  one  or  more  lab- 
oratories to  dissection  of  birds.  Of  these,  two 
courses  had  demonstration  dissections  by  fac- 
ulty only.  Eight  faculty  provided  their  own 
dissection  guide  (Appendix),  but  others  relied 
on  Pettingill  (1990)  or  Faaborg  and  Chaplin 
(1988b).  Preserved  pigeons  {Colwnba  livia) 
were  used  for  dissection  by  1 3 of  1 9 respon- 
dents. One  person  used  fresh  pigeons.  Chick- 
ens (Gallus  gallus),  Japanese  Quail  {Coturnix 
coturnix),  European  Starlings  {Sturniis  vulgar- 
is), House  SpaiTOWs  (Passer  domesticiis),  and 
birds  killed  in  accidents  were  used  in  the  re- 
maining courses  or  for  compainson  with  pi- 
geons. One  ornithologist  brought  a roasted 


290 


THE  WILSON  BULLETIN 


Vol.  Ill,  No.  2,  June  1999 


chicken  to  laboratory  for  dissection  and  sub- 
sequent consumption. 

Most  instructors  required  students  to  own  a 
field  guide.  The  appropriate  Peterson  guide 
[eastern  (1980)  or  western  (1984)  North 
America]  was  the  most  popular  choice  (Ap- 
pendix). 

Twenty-three  of  25  respondents  who  taught 
a course  with  a laboratory  component  required 
students  to  learn  to  identify  species  of  birds 
by  sight,  usually  of  the  local  avifauna.  The 
number  of  birds  students  had  to  learn  varied 
from  fewer  than  50  to  over  200,  with  101—150 
being  typical. 

Seventeen  of  the  23  also  required  their  stu- 
dents to  learn  to  identify  some  birds  by  song. 
The  number  of  species  each  student  had  to 
learn  ranged  from  21  to  100  with  41-60  being 
typical.  To  help  students  learn  vocalizations, 
respondents  identified  the  Peterson  tapes,  the 
Birding  by  Ear  tapes,  and  the  National  Geo- 
graphic Society  tapes  as  particularly  useful. 
One  respondent  had  prepared  an  audiotape 
specific  to  the  birds  that  students  had  to  learn 
in  the  course.  A few  respondents  taped  songs 
with  students  and  had  the  students  analyze  the 
songs  themselves.  This  not  only  taught  stu- 
dents recording  and  analytical  techniques,  but 
also  gave  them  a thorough  knowledge  of  the 
characteristics  of  the  songs  they  recorded.  To 
test  students’  abilities  to  identify  birds  by 
sight  and  sound,  87%  of  the  respondents  gave 
laboratory  examinations  and  35%  gave  ex- 
aminations in  the  field. 

The  amount  of  taxonomy  students  had  to 
learn  varied.  Twenty  of  25  instructors  required 
students  to  learn  order  names  and  know  the 
distinguishing  features  of  each  order.  Family 
names  were  required  for  students  in  16  of  the 
courses.  Few  instructors  required  that  genera 
(two  courses)  and  species  (one  course)  names 
be  learned. 

Most  courses  required  a long  written  report, 
and  a few  also  required  one  or  more  short 
written  assignments.  Long  written  assign- 
ments included  the  following: 

• detailed  field  journal  based  on  20  hours  or 
more  of  fieldwork  in  addition  to  the  regular 
laboratory  field  trips; 

• term  paper  based  on  original  field  or  labo- 
ratory research  or  a literature  review  cov- 
ering .some  aspect  of  avian  biology; 

• Joint  paper  by  several  students  working  on 


limited  and  local  research  topics.  The  teach- 
er did  the  literature  search.  The  students 
added  their  own  data  and  synthesized  the 
material; 

• paper  based  on  observations  of  a bluebird 
box  on  campus.  Students  monitored  the  as- 
signed box  from  late  March  until  the  young 
fledged; 

• paper  based  on  the  social  behavior  of  a par- 
ticular' species  with  monitoring  of  the  spe- 
cies over  the  course  of  the  semester; 

• research  paper  that  usually  involved  field 
research,  data  analysis,  and  literature  re- 
view in  which  the  teacher  and  classmates 
reviewed  a rough  draft  before  the  final  draft 
was  submitted; 

• paper  based  on  field  research  on  the  behav- 
ior, ecology,  or  migration  of  a local  bird 
species.  One  insti-uctor  disallowed  referenc- 
es to  encourage  the  creativity  and  the  ob- 
servational and  analytical  skills  of  each  stu- 
dent; 

• paper  based  on  a census  of  the  birds  of  a 
site  that  has  been  censused  annually  since 
1971; 

• an  account  of  a local  bird  in  the  style  used 
by  the  Birds  of  North  America; 

• paper  based  on  the  analysis  of  a large  sam- 
ple of  banding  data  for  Yellow  Warblers 
(Dendroica  petechia).  The  students  could 
analyze  site  fidelity  as  a function  of  age  and 
sex,  ai'rival  dates  as  a function  of  age  and 
sex,  etc.; 

• analysis  of  the  population  dynamics  of  a 
species  based  on  Christmas  Bird  Count 
data. 

Short  written  assignments  included  the  fol- 
lowing: 

• one  page  summary  of  a published  article; 

• computer  spreadsheet  assignments  on  en- 
ergetics of  flight  and  thermoregulation; 

• three  critiques  of  a set  of  three  or  four 
papers  with  contradictory  views  on  a pai- 
ticular  issue.  Each  student  summarized  each 
paper  and  then  offered  critical  comment  on 
each,  taking  a position  on  one  side  of  the 
controversy; 

• weekly  5—10  minute  essays  written  in  class 
on  specific  ornithological  questions; 

• two  critiques  of  recent  ornithological  arti- 
cles written  in  the  format  of  the  Recent  Lit- 
erature section  of  the  Journal  of  Field  Or- 


Ihtnt  and  Wilson  • SURVEY  OF  ORNITHOLOGY  COURSES 


291 


nithology.  The  critiques  had  to  be  rewritten 
until  they  reach  “A”  quality. 

SUPPORT  FOR  UNDERGRADUATE 
ORNTIHOLOGY 

Avian  specimens  were  equally  available  to 
respondents  at  large  and  small  schools.  Fac- 
ulty used  collections  to  illustrate  taxonomic 
principles,  avian  systematics,  and  less  often  to 
illustrate  morphological,  ecological,  and  be- 
havioral adaptations.  In  five  courses,  students 
were  required  to  prepare  one  or  more  study 
skins.  Skin  prepaiation  was  optional  in  six 
other  courses. 

Collections  available  at  small  and  large 
schools  were  similar  in  size  with  9 of  23  col- 
lections having  fewer  than  1,000  specimens, 
some  with  fewer  than  300.  Eighteen  of  the  23 
collections  had  a regional  focus,  two  had 
broad  North  American  representation,  and 
three  had  large  collections  representing  birds 
of  the  world.  All  collections  included  study 
skins,  most  included  mounts  and  skeletons, 
and  some  included  eggs,  nests,  and  alcoholic 
specimens. 

Fifteen  schools  owned  some  of  the  natural 
areas  visited  by  the  class.  Use  of  these  areas 
varied  from  caiefully  scheduled,  multi-year 
censusing  of  an  arboretum  managed  by  the 
university,  to  intermittent  visits  to  unmanaged 
areas  for  “birding.”  Here  as  with  use  of  mu- 
seum collections,  faculty  might  benefit  from 
sharing  ideas  on  how  university-owned  natu- 
ral areas  could  be  used  in  conjunction  with  an 
ornithology  course. 

MOST  SUCCESSFUL  PARTS  OF  THE 
COURSE 

Respondents  were  asked  to  describe  the 
most  successful  parts  of  their  courses.  Fifteen 
listed  some  aspect  of  field  trips  as  the  most 
successful  portion  of  the  course.  Interestingly, 
one  respondent  found  that  some  students 
loved  the  laboratory/field  portion  of  the 
course  while  others  hated  it.  Listed  below  are 
the  teaching  aids  and  activities  instructors 
found  most  successful: 

• audiotapes  and  CDs  of  bird  songs  to  facil- 
itate vocal  identification  of  birds; 

• breakfast  with  the  class  before  or  after 
morning  field  trips; 

• color  slides,  whether  the  instructor’s  own  or 
supplemented  from  VIREO.  One  instructor 


provided  detailed  notes  on  each  slide  so 
that  students  could  devote  full  attention  to 
the  slides; 

• demonstration  or  experiment  that  gets  stu- 
dents involved  in  active  learning; 

• field  trips  to  build  enthusiasm  for  learning 
species  identification  and  understanding  the 
biology  of  birds; 

• laboratory  and  lecture  sessions  on  the  same 
day  to  encourage  integration  of  the  mate- 
rial; 

• lectures  on  ecology  and  behavior; 

• list  of  mnemonic  devices  generated  by  stu- 
dents for  learning  vocalizations; 

• lecture  demonstrations,  for  example  use  of 
parachutes,  gliders,  ornithopters,  and 
mounted  wings  in  a wind  tunnel  to  illustrate 
principles  of  flight; 

• mist-netting  and  bird-banding  to  excite  stu- 
dents’ interest,  particularly  early  in  the 
course  or  in  conjunction  with  ongoing  re- 
search in  which  the  students  could  partici- 
pate; 

• morphological,  ecological,  and  behavioral 
adaptations  of  birds; 

• study  specimens  before  field  trips; 

• videotapes,  especially  those  from  the  Na- 
ture series  on  Public  Broadcasting:  for  ex- 
ample Marathon  Bird,  Rulers  of  the  Wind, 
Master  Builders,  the  Bee  Team  (on  social 
behavior  in  White-fronted  Bee-eaters),  and 
Jewels  (hummingbirds); 

• use  of  the  Macintosh  softwaie  SoundEdit 
(Mac Recorder).  Annotated  vocalizations  of 
40  species  were  provided  on  departmental 
hard  disks.  Students  could  play  vocaliza- 
tions of  species  they  found  confusing.  The 
software  also  allowed  students  to  make 
sonograms  and  spectrograms  of  vocaliza- 
tions; 

• laminated  color  photographs  of  birds  to 
shaipen  identification  skills,  most  useful  for 
institutions  with  a limited  teaching  collec- 
tion; 

• students  were  provided  with  essay  ques- 
tions a week  in  advance  of  the  test  and 
could  return  outlines  of  their  answers  at 
least  48  hours  before  the  test  for  comments 
by  the  teacher; 

• use  of  the  “Gone  Birding’’  game  to  intro- 
duce students  to  identification.  The  game 
helped  promote  the  goals  of  enjoyment  and 
group  learning  as  well  as  improving  iden- 


292 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  2,  June  1999 


tification  skills  and  knowledge  of  species- 
habitat  associations. 

LEAST  SUCCESSFUL  PARTS  OF  THE 
COURSE 

Each  respondent  was  asked  to  identify  the 
least  successful  portions  of  his  or  her  course. 
Fellow  ornithology  teachers  can  offer  little 
help  with  three  common  complaints:  insuffi- 
cient time,  cold  weather,  and  conclusion  of  the 
semester  before  the  arrival  of  many  spring  mi- 
grants. Lecture  was  most  often  listed  as  the 
least  successful  part  of  the  course.  The  least 
successful  lecture  topics  included  systematics 
and  physiology  and  anatomy,  despite  the  fact 
that  each  occupied  a substantial  part  of  the 
“consensus”  course  (Fig.  1 ).  Additional  areas 
that  some  faculty  listed  as  least  successful  in- 
cluded using  study  skins  to  teach  identifica- 
tion and  taxonomy,  using  tapes  to  leai'n  bird 
vocalizations,  and  teaching  students  field  tech- 
niques, especially  how  to  quantify  behavior  in 
the  field.  We  hope  that  this  list  can  stimulate 
ideas  that  will  improve  these  portions  of  the 
courses. 

TEACHING  MATERIAL  NOT  PRESENTLY 
AVAILABLE 

When  asked  to  identify  teaching  aids  that 
do  not  appear  to  be  commercially  available, 
most  respondents  indicated  a preference  for 
interactive  software.  Most  would  use  such 
software  to  help  students  learn  identification 
of  birds.  Specific  needs  are  listed  below.  If 
any  readers  know  of  such  aids,  please  notify 
us  or  a member  of  the  Wilson  Ornithological 
Society  Committee  on  Undergraduate  Educa- 
tion. 

• good  video,  laser  disk,  or  CD-ROM  that 
deals  only  with  ordinal  chai'acteristics; 

• good  video,  laser  disk,  or  CD-ROM  that 
deals  only  with  familial  characteristics; 

• software  that  illustrates  the  principles  of 
taxonomy; 

• video  or  computer  disks  illustrating  field 
characteristics  of  birds  and  their  vocaliza- 
tions accompanied  by  sonograms.  Such  ma- 
terial would  allow  students  to  learn  at  their 
own  pace; 

• software  simulations  of  population  dynam- 
ics and  evolution; 

• simple,  user-friendly  manual  on  field  tech- 


niques in  ornithology,  directed  toward  stu- 
dents. 

CONCLUSION 

Ornithology  is  part  of  the  curriculum  at 
large  and  small,  public  and  private,  graduate 
and  undergraduate  schools  throughout  North 
America.  The  responses  to  our  survey  con- 
firmed our  sense  that  ornithology  is  an  excit- 
ing, interactive  subject  taught  by  men  and 
women  with  a genuine  enjoyment  of  birds. 
The  desire  to  engender  that  same  life-long  en- 
joyment in  others  was  a common  theme  run- 
ning throughout  all  the  responses.  Beyond  this 
common  theme  we  were  impressed  with  the 
diversity  of  creative  ideas  contained  within 
the  courses.  We  hope  that  our  respondents’ 
ideas  as  we  have  presented  them  will  stimu- 
late others  to  introduce  new  ideas  into  their 
ornithology  courses  and  to  bring  those  ideas 
and  their  reception  by  students  to  ornitholog- 
ical meetings  for  discussion  by  all  who  teach 
the  biology  of  birds.  Finally  the  Wilson  Or- 
nithological Society’s  Committee  on  Under- 
graduate Education  would  be  glad  to  work 
with  any  ornithologist  who  has  ideas  or  sug- 
gestions related  to  the  teaching  of  ornithology. 
The  Committee  is  eager  to  promote  increased 
public  discussion  of  teaching  philosophies  and 
ideas.  We  hope  this  paper  marks  the  beginning 
of  that  public  discussion. 

ACKNOWLEDGMENTS 

We  thank  the  26  faculty  who  took  the  time  to  re- 
spond to  our  questionnaire.  We  hope  that  we  have  rep- 
resented their  responses  accurately.  We  thank  the 
members  of  the  Wilson  Ornithological  Society’s  Com- 
mittee on  Undergraduate  Education  for  their  sugges- 
tions throughout  the  development  of  the  questionnaire 
and  its  analysis,  most  particularly  E.  J.  Willoughby  and 
L.  Moseley.  The  manuscript  benefited  from  the  many 
helpful  comments  of  K.  L.  Bildstein,  A.  J.  Gatz,  D.  C. 
Radabaugh,  and  an  anonymous  referee. 

LITERATURE  CITED 

Faaborg,  j.  and  S.  B.  Chaplin.  1988a.  Ornithology: 
an  ecological  approach.  Prentice  Half  Englewood 
Cliffs,  New  Jersey. 

Faaborg,  J.  and  S.  B.  CiiAPt.iN.  1988b.  Ornithology: 
an  ecological  approach.  Laboratory  manual  and 
field  exercises.  Prentice  Hall,  Englewood  Clilfs, 
New  Jersey. 

Gii.l.,  E B.  1990.  Ornithology.  W.  H.  Freeman  and 
Company,  New  York. 

Mock,  D.  W.  1991.  Behavior  and  evolution  of  birds. 
W.  H.  Freeman  and  Company,  New  York. 


Burn  and  Wilson  • SURVEY  OF  ORNITHOLOGY  COURSES 


293 


National  Gkographic  Socikty.  1983.  Field  guide  to 
the  birds  of  North  America.  National  Geographic 
Society,  Washington,  D.C. 

Pkrrins,  C.  a.  and  a.  L.  A.  Middleton.  1985.  The 
encyclopedia  of  birds.  Facts  on  File,  New  York. 

Peterson,  R.  T.  1980.  A field  guide  to  the  birds. 
Houghton  Mifflin  Company,  Boston,  Massachu- 
setts. 

Peterson,  R.  T.  1984.  A field  guide  to  the  western 
birds.  Houghton  Mifflin  Company,  Boston,  Mas- 
sachusetts. 

Pettingill,  Jr.,  O.  S.  1990.  Ornithology  in  laboratory 
and  field.  Academic  Press,  Inc.,  New  York. 

Proctor,  N.  S.  and  P.  J.  Lynch.  1993.  Manual  of 
ornithology:  avian  structure  and  function.  Yale 
Univ.  Press,  New  Haven,  Connecticut. 

Robbins,  C.  S.,  B.  Brunn,  and  H.  S.  Zim.  1983.  Birds 
of  North  America.  Golden  Press,  New  York. 

Welty,  J.  C.  and  L.  F.  Baptista.  1988.  The  life  of 


birds.  Saunders  College  Publishing,  Philadelphia, 
Pennsylvania. 

APPENDIX 

Texts  used  for  lecture,  dissection,  and  field- 
work are  listed  alphabetically  by  author  (num- 
ber of  courses  using  the  text  are  in  parenthe- 
ses). 

Primary  lecture  text:  Faaborg  and  Chaplin 
1988a  (2),  Gill  1990  (18),  Pettingill  1990  (5), 
Welty  and  Baptista  1988  (1); 

Supplementary  lecture  text:  Mock  1991  (2), 
Perrins  and  Middleton  1985  (1); 

Dissection  text:  Faaborg  and  Chaplin  1988b 
(1),  instructor’s  own  (8),  Pettingill  1990  (5); 

Field  guide:  National  Geographic  1983  (3), 
Peterson  1980,  1984  (11),  Pettingill  1990  (3), 
Robbins  et  al.  field  guide  1983  (4). 


Wilson  Bull.,  111(2),  1999,  pp.  294-301 


Ornithological  Literature 

Edited  by  William  E.  Davis,  Jr. 


SWALLOW  SUMMER.  By  Charles  R. 
Brown.  Univ.  Nebraska  Press,  Lincoln,  Ne- 
braska. 1998:  xiii  + 371  pp.,  black-and-white 
photographs.  $16.95  (paper). — For  the  past  15 
years,  Charles  R.  Brown  and  his  wife  Mary 
have  studied  Cliff  Swallows  {Petrochelidon 
pyrrhonota)  at  Cedar  Point  Biological  Station 
in  the  Sand  Hills  of  western  Nebraska.  The 
Browns’  research  was  (and  continues  to  be) 
highly  productive,  partly  because  the  swal- 
lows proved  to  be  excellent  research  material. 
Cliff  Swallows  make  mud  nests  in  colonies 
that  may  range  from  a few  birds  to  thousands. 
The  swallows  are  relatively  easy  to  capture 
(but  provide  the  basis  for  many  adventures) 
and  are  tolerant  of  extensive  handling.  Their 
natural  history  is  relevant  to  numerous  issues 
in  the  basic  ecology  of  birds  and  to  the  un- 
derstanding group  behavior  of  animals. 

This  book  is  a non-scientific  account  of 
how  the  Browns  came  to  work  on  this  bird 
and  the  trials  and  tribulations  of  one  season 
of  their  studies.  Each  year  they  arrive  in  Ne- 
braska in  May  in  advance  of  the  first  swal- 
lows. They  capture,  band,  weigh,  and  measure 
birds  until  the  swallows  stop  breeding  in  late 
July.  With  the  help  of  numerous  assistants, 
they  have  obtained  data  from  thousands  of 
swallows.  In  fact,  their  total  sample  size  for 
some  measurements  must  be  in  six  figures! 

The  writing  is  clear,  funny,  insightful,  in- 
teresting, and  informative.  If  you  have  more 
than  a passing  interest  in  swallows,  group  be- 
havior, or  basic  avian  ecology,  you  should  also 
read  the  scientific  account,  “Coloniality  in  the 
Cliff  Swallow:  the  effect  of  group  size  on  so- 
cial behavior”  (co-authored  with  Mary 
Brown;  1996,  Univ.  Chicago  Press),  if  you 
have  not  already  done  so. 

Prairie  thunderstorms,  clouds  of  birds,  the 
mysteries  of  bird  behavior,  quotes  from  west- 
ern movies  (especially  from  “Lonesome 
Dove”!),  the  deep  pleasures  of  spending  the 
summer  at  a field  station,  the  behavior  of  field 
assistants,  the  dirt/frustration/exhaustion  of 
long  days  in  the  field,  the  satisfaction  of  dis- 
covery— these  are  just  a few  of  the  rich 


threads  in  the  texture  of  Brown’s  account.  If 
you  have  worked  for  a long  time  on  one  spe- 
cies, spent  a summer  with  students  at  a field 
station,  love  western  Nebraska,  have  struggled 
to  fund  and  run  a field  research  project,  or 
simply  would  enjoy  a good  account  of  how 
field  ornithologists  see  the  world,  you  will  ap- 
preciate this  book. — CHARLES  R.  BLEM. 


ATLAS  OF  BREEDING  BIRDS  OF  IN- 
DIANA. By  John  S.  Castrale,  Edward  M. 
Hopkins,  and  Charles  E.  Keller.  Available 
from:  Indiana  Department  of  Natural  Resourc- 
es, Customer  Service  Center,  402  W.  Wash- 
ington St.,  Rm.  W160,  Indianapolis,  IN 
46204.  1998:  388  pp.,  14  numbered  text  figs., 
7 tables,  158  range  maps.  $20  plus  $3.50  s&h, 
$1  sales  tax  for  Indiana  residents  (cloth). — 
The  Indiana  Nongame  and  Endangered  Wild- 
life Program  of  the  Indiana  Department  of 
Natural  Resources  sponsored  this  atlas  pro- 
ject, coordinating  the  nearly  600  volunteers 
and  paid  “block  busters”  during  the  fieldwork 
conducted  from  1985-1990.  Workers  targeted 
647  “priority  blocks,”  west-central  of  the  six 
blocks  of  each  U.S.  Geological  Survey  IV2 
topographic  map  of  the  state.  A series  of  maps 
depict  counties,  public  lands,  rivers,  urban  ar- 
eas, natural  regions,  forested  ai'eas  (with  sep- 
arate maps  for  evergreen-deciduous  and 
shrubland-early  successional  woodlands),  ag- 
ricultural row-crops,  pastureland,  and  marshes 
and  open  water.  In  an  attempt  to  provide  some 
indices  of  abundance,  atlas  accounts  used 
Breeding  Bird  Survey  (BBS)  and  Summer 
Bird  Count  (SBC)  data  for  the  1985-1990  pe- 
riod. The  SBC  uses  the  county  as  the  sampling 
area,  and  counts  are  conducted  on  multiple 
days  during  June.  A table  lists  in  rank  order 
the  percentage  of  blocks  in  which  a species 
was  detected  (e.g.,  American  Robin,  Turdus 
rnigratorius,  100%,  rank  = 1),  abundance  val- 
ues and  rank  for  BBS  routes,  and  birds/party 
hour  and  rank  for  SBCs. 

A biogeographic  analysis  by  J.  Dan  Web- 


294 


ORNITHOLOGICAL  LITERATURE 


295 


ster  includes  a table  that  lists  species  extinct 
or  extiipated  by  1929,  species  extiipated  be- 
tween 1929  and  1979,  possible  or  sporadic 
nesters.  Twentieth  Century  additions  as  breed- 
ers, and  confirmed  breeding  species  since 
1990.  The  analysis  of  bird  distributions  does 
not  correspond  well  with  physiographic  pat- 
terns, vegetation,  distribution  of  other  groups 
of  organisms,  or  “natural  regions.”  The  au- 
thor suggests  birds  are  poor  indicators  of  bio- 
geography in  a small,  flat  state  like  Indiana. 
Major  changes  this  century  that  have  had  a 
major  impact  on  nesting  species  include  the 
virtual  disappearance  of  the  prairies,  the  drain- 
age and  pollution  of  wetlands,  and  the  frag- 
mentation of  forests.  The  extirpation  of  prai- 
rie, wetlands,  and  forest  interior  species,  to- 
gether with  a uniform  intrusion  of  alien  spe- 
cies has  resulted  in  a more  uniform  avifauna 
in  the  state. 

The  bulk  of  the  book  is  devoted  to  species 
accounts.  Full  species  accounts  accompany 
the  158  species  that  were  confirmed  breeders, 
supplemented  by  shorter  accounts  for  46  spe- 
cies not  confirmed  as  breeders  during  the  at- 
lasing  period  (including  extirpated  breeders 
and  those  confirmed  as  breeding  since  1990). 
Each  map  occupies  a full  page,  with  the  spe- 
cies account  on  the  facing  page.  The  large  size 
of  the  maps  makes  them  very  easy  to  read. 
The  species  accounts  give  very  brief  natural 
histories  synopses,  and  the  bulk  of  the  ac- 
counts are  concerned  with  historical  distribu- 
tion comparisons,  analysis  of  the  atlas  results, 
and  comparisons  with  bird  distributions  in  the 
surrounding  states  of  Ohio,  Michigan,  Illinois, 
and  Kentucky.  Each  species  account  is  accom- 
panied by  a table  summarizing  the  atlas,  BBS, 
and  SBC  data  for  north,  central,  and  southern 
regions  of  Indiana,  as  well  as  statewide. 

This  is  a well-done  atlas  that  is  a bargain 
at  $20 — the  more  than  350  references,  many 
to  local  publications,  alone  are  worth  that.  It 
should  be  of  interest  to  those  concerned  with 
bird  distribution.— WILLIAM  E.  DAVIS,  JR. 


A GUIDE  TO  THE  NESTS,  EGGS,  AND 
NESTLINGS  OF  NORTH  AMERICAN 
BIRDS,  SECOND  EDITION.  By  Paul  J.  Bai- 
cich  and  Colin  J.  O.  Harrison.  Academic 
Press.  1997:  347  pp.,  64  color  plates,  and  103 


black  and  white  figures.  $22.95  (paper). — 
This  is  an  updated  version  of  the  1978  edition, 
and  this  new  edition  is  a must  for  the  library 
of  anyone  interested  in  the  nesting  period  of 
North  American  birds.  The  guide  begins  with 
an  introduction  that  describes  a variety  of  as- 
pects of  breeding  biology  including  how  and 
where  nests  are  built;  egg  shape,  color,  and 
size;  and  clutch  size,  incubation,  hatching,  and 
the  nestling  period.  Keys  to  nests,  eggs,  and 
nestlings  are  also  provided  in  the  introductory 
pages.  The  introductory  section  is  followed  by 
a series  of  individual  species  accounts.  Each 
species  account  includes  breeding  habitat;  lo- 
cation, description,  size,  and  materials  of  the 
nests;  number,  shape,  size,  and  color  of  eggs; 
breeding  season;  length  of  incubation;  de- 
scription of  nestlings;  length  and  description 
of  nestling  period;  and  roles  of  both  sexes  in 
these  activities. 

For  many  users,  the  most  valuable  aspect  of 
this  book  is  the  color  plates  that  provide  pho- 
tographs of  the  eggs  of  597  species  of  North 
American  breeders  as  well  as  147  color  draw- 
ings of  nestlings.  While  most  species’  eggs 
are  represented  by  a single  photograph,  sev- 
eral with  paiticularly  variable  eggs  are  repre- 
sented by  multiple  photographs;  for  example, 
the  authors  provide  six  different  photographs 
of  Sandwich  Tern  {Sterna  sandvicensis)  eggs. 
Additional  information  about  variation  within 
a species  is  described  in  plate  legends  as  well 
as  individual  species  accounts.  Black-and- 
white  drawings  ai'e  scattered  throughout  the 
text  to  illustrate  additional  nestlings  and  a va- 
riety of  nests. 

The  second  edition  of  this  guide  includes 
substantial  information,  photographs,  and 
drawings  from  the  first  edition;  however,  it 
also  includes  updates  and  new  information 
that  was  not  in  the  first  edition.  Species  names 
(both  common  and  scientific)  and  taxonomic 
affinities  have  been  updated  to  reflect  changes 
since  the  first  edition.  Numerous  species  ac- 
counts have  been  augmented  with  information 
about  breeding  biology,  numbers  of  eggs,  and 
incubation  that  was  missing  or  not  known  in 
the  first  edition.  Species  accounts  have  been 
added  for  a number  of  species  that  were  split 
from  existing  species  (e.g.,  Bicknell’s  Thrush, 
Catharus  bicknellr,  California  Gnatcatcher, 
Polioptila  californica;  and  Island  Scrub-Jay, 
Aphelocoma  insularis)  or  have  begun  to  breed 


296 


THE  WILSON  BULLETIN 


Vol.  Ill,  No.  2,  June  1999 


regularly  in  North  America  (e.g.,  Lesser 
Black-backed  Gull,  Lams  fuscus\  Buff-col- 
lared Nightjar,  Caprimulgus  ridgwayr,  and 
Shiny  Cowbird,  Molothrus  bonariensis) . The 
plates  have  been  updated  to  provide  pictures 
of  the  eggs  of  the  new  species  covered  in  the 
text  and  to  reflect  changes  in  taxonomic  order, 
species  names,  and  family  groups.  Another 
useful  change  in  the  revised  edition  is  the 
grouping  of  all  the  plates  together  in  the  cen- 
ter of  the  guide  rather  than  having  the  plates 
of  nestlings  scattered  throughout  the  text  pag- 
es. This  change  in  format  makes  the  guide 
easier  to  use  than  the  first  edition.  The  guide 
also  now  includes  a selected  bibliography  of 
important  works  used  in  the  revision  and  in- 
formation on  how  to  contact  the  author  to  ob- 
tain specific  references  for  each  of  the  ac- 
counts. 

Professional  and  amateur  ornithologists  will 
find  that  this  guide  provides  extensive  infor- 
mation about  nests  and  eggs  in  an  easy-to-use 
format.  I would  highly  recommend  The  Guide 
to  Nests,  Eggs,  and  Nestlings  of  North  Amer- 
ican Birds  to  anyone  interested  in  identifying 
birds’  nests  and  eggs  or  learning  more  about 
the  nest  and  nestling  stage  of  North  American 
birds.— SARA  R.  MORRIS. 


BREEDING  BIRDS  OF  WASHINGTON 
STATE:  LOCATION  DATA  AND  PREDICT- 
ED DISTRIBUTIONS.  By  Michael  R.  Smith, 
Philip  W.  Mattocks,  Jr.,  and  Kelly  M.  Cassidy. 
Available  from  Seattle  Audubon  Society, 
8050-35th  Avenue  NE,  Seattle,  WA  98115. 
1997:  538  pp.,  10  numbered  text  figs.,  6 ta- 
bles, 244  range  maps.  $30  plus  $3  s&h  (pa- 
per).— The  authors  attempt  to  fill  two  roles 
with  this  book:  compilation  of  the  Seattle  Au- 
dubon Society’s  breeding  bird  atlas  project, 
and  a part  of  the  final  report  of  the  Washing- 
ton State  Gap  Analysis  project  (administered 
by  the  National  Biological  Service).  The  goals 
of  this  aspect  of  the  Gap  Project  were  to  map 
existing  land  cover,  and  model  the  breeding 
distributions  of  birds.  The  book  is  divided  into 
an  introductory  chapter  (22  pp.)  and  258  spe- 
cies accounts  that  include  244  maps  (489  pp.). 

The  introductory  chapter  traces  the  history 
of  the  Atlas  project  and  summarizes  the  data 
collected  and  analyzed.  Over  600  volunteers 


contributed  data  over  the  decade  beginning  in 
1985.  The  sampling  unit  was  a block  of  nine 
square  miles  that  constituted  one  quarter  of  a 
township,  for  a total  of  7912  blocks.  Data  pro- 
vided by  the  Washington  Department  of  Fish 
and  Wildlife,  particularly  from  their  shrub- 
steppe  bird  study  and  National  Heritage  da- 
tabase, were  included.  Coverage  of  blocks 
was  incomplete,  with  56%  containing  at  least 
one  “possible”  species  record,  51%  a “prob- 
able” record,  and  44%  a “confirmed”  record. 
More  than  1 0 species  were  confirmed  breeders 
in  6%  of  the  blocks  (the  highest  number  of 
confirmed  species  in  a block  was  67).  The  au- 
thors discuss  the  biases  in  the  data,  including 
the  concentration  of  records  from  more  de- 
veloped areas  of  the  state.  They  also  point  out 
that  records  do  not  reflect  abundance,  and  give 
the  delightful  example:  “a  confirmed  breeding 
record  of  a Ruby-crowned  Kinglet  in  the  Pon- 
derosa  Pine  zone  in  the  Blue  Mountains 
(where  the  species  is  uncommon,  but  can  usu- 
ally be  found)  occupies  the  same  area  on  the 
map  as  a confirmed  record  derived  from  the 
Subalpine  Fir  zone,  where  Ruby-crowned 
Kinglets  are  best  measured  by  the  ton.” 

For  most  species  models  predicting  breed- 
ing distribution  (extent  of  a species’  breeding 
habitat)  were  prepared  and  appeared  on  the 
species’  distribution  map  as  “Habitats  in  core 
zones”  with,  in  some  cases,  an  additional 
“Habitats  in  peripheral  zones”  presented  in  a 
lighter  shade  of  gray.  Habitats  were  selected 
from  a satellite  image  landcover  map  of  the 
state.  Bird-habitat  associations  were  devel- 
oped from  literature  reviews,  location  of 
breeding  records,  and  consultation  with  ex- 
perts. The  introductory  chapter  detailed  the 
modeling  process  and  mapping  processes,  al- 
beit with  a touch  of  jargonese  (e.g.,  “Each 
scene  was  spectrally  clustered  into  120  to  256 
spectral  classes  using  bands  1,  2,  3,  4,  5,  and 
7”  or  “the  presumption  that  error  in  a cover 
is  equal  to  the  product  of  the  errors  in  each 
layer  comprising  that  cover  is  an  over-simpli- 
fication. The  effect  of  an  error  in  a source  lay- 
er upon  a derived  cover  depends  on  what  is 
being  derived.”).  Ecoregions,  vegetation,  and 
vegetation  zones  are  detailed  in  tables,  a 
black-and-white  map,  and  in  two  color  plates. 

Each  species  account  is  divided  into  three 
parts,  the  first  “Breeding  Status  and  Distri- 
bution,” the  second  “Model”  gives  the  char- 


ORNITHOLOGICAI,  LITERATURK 


297 


acteristics  of  the  model  predicting  the  breed- 
ing distribution,  and  the  third  “comments,” 
provides  aspects  of  the  species’  biology.  The 
latter  contains  such  information  as  maximum 
abundance  numbers  for  island  nesting  species, 
comments  on  taxonomy,  e.g.,  American  (Cor- 
vus  brachyrhynochos)  and  Northwestern  (C. 
caurinus)  crows,  and  an  historical  perspective 
on  the  species.  I found  the  “comments”  sec- 
tions pailicularly  informative. 

This  is  an  interesting  book.  It  is  a hybrid 
volume,  and  thus  provides  a great  deal  of  in- 
formation on  habitat  and  computer  modeling 
not  usually  found  in  breeding  bird  atlases,  but 
is  also  somewhat  jaixing  because  of  the  con- 
trast of  sophisticated  jargon  associated  with 
the  Gap  Analysis  and  the  more  usual  distri- 
bution descriptions  associated  with  atlas  pro- 
jects. The  atlas  coverage  seems  sparse  and  I 
wonder  why  Breeding  Bird  Survey  data 
weren’t  used  to  provide  a better  measure  of 
bird  abundance.  As  is  usually  the  case,  the 
more  than  200  references,  many  from  local 
journals  and  unpublished  sources,  are  a gold 
mine  of  regional  information.  Certainly  this 
book  should  be  part  of  the  library  of  every 
serious  student  of  bird  distribution,  and  all  ac- 
ademic libraries.— WILLIAM  E.  DAVIS,  JR. 


MADE  FOR  EACH  OTHER:  A SYMBIOSIS 
OF  BIRDS  AND  PINES.  By  Ronald  M.  Fan- 
ner, Oxford  Univ.  Press,  New  York,  New 
York.  1996:  160  pp.,  14  chapters,  24  figures, 
8 tables,  and  15  color  photographs  on  4 plates. 
$35.00  (cloth). — This  slim  book  interweaves 
a description  of  symbiosis  on  a “grand  scale” 
between  some  corvids  and  over  twenty  spe- 
cies of  pines  across  “vast  tracts  of  North 
American  and  Asian  wildland”  with  the  co- 
evolutionary story  of  the  Clark’s  Nutcracker 
(Nucifraga  Columbiana)  and  whitebark  pine 
(Pinus  albicaulis)  of  the  western  United 
States.  The  author  makes  the  case  that  the  nut- 
cracker-pine  relationship  is  a strong  mutual- 
ism because  nutcrackers  are  the  most  impor- 
tant dispersal  agents  of  the  pine  seeds  and  the 
pine  seeds  are  a nutrient-rich  food  essential 
for  nutcracker  survival.  The  critical  message, 
delivered  in  the  final  chapter,  “Is  the  Keystone 
Slipping?,”  is  an  alert  that  serious  threats  to 
the  whitebark  pine  may  diminish  populations 


of  the  nutcracker  and  other  animal  species  that 
depend  on  the  pine  seeds. 

This  is  a book  for  pine  lovers.  To  support 
his  argument  for  the  mutualism  between  cor- 
vids and  pines.  Fanner  examines  the  phytog- 
eny, comparative  morphology,  and  ecology  of 
pines,  particularly  the  35  species  of  “soft 
pines”  in  the  subgenus  Strobus.  This  group  of 
species  has  a high  proportion  of  species  with 
wingless  seeds:  a critical  adaptation  to  dis- 
persal by  corvids.  The  five  species  of  stone 
pines — Eurasian  and  one  North  American — 
that  are  close  mutualists  with  the  Eurasian 
{Nucifraga  caryocatactes)  and  Clark’s  nut- 
crackers receive  the  most  attention.  Charac- 
teristics of  stone  pine  cone  fertilization  and 
development,  and  seed  germination  and  nutri- 
tional content,  are  described  in  detail.  Some 
of  these  descriptions  are  less  relevant  to  the 
main  thesis  of  the  book  than  others,  but  they 
provide  a rich  natural  history. 

A review  of  the  family  Corvidae  quickly 
focuses  on  the  “pine  birds”:  the  nutcrackers 
and  the  Pinyon  Jay  {Gymnorhinus  cyanoce- 
phalus).  “Pine  bird”  adaptations  to  a diet  of 
pine  seeds  are  long  bills,  the  ability  to  carry 
numbers  of  pine  seeds,  and  well-developed 
spatial  memory  for  retrieving  cached  seeds. 
The  foundation  for  the  pine-corvid  mutualism, 
including  experimental  evidence  that  Clark’s 
Nutcrackers  use  fixed  objects  as  visual  cues 
to  find  food  caches,  is  provided  primarily  by 
the  research  of  Diana  Tomback,  Stephen  Van- 
der  Wall,  Russell  Baida,  and  Fanner’s  own 
work  on  seed  dispersal  by  birds.  Fanner  is 
careful  to  point  out  that  nutcrackers  are  not 
completely  dependent  on  a single  species  of 
pine.  He  notes  that  nutcrackers  use  a variety 
of  other  pine  seeds  and  foods  and  migrate  (ir- 
rupt) when  the  pine  seed  crop  fails.  However, 
nestlings  are  fed  pine  nuts  that  were  cached 
almost  exclusively,  and  cached  foods  aie  used 
extensively  during  fall,  winter,  and  spring 
when  other  foods  are  scarce. 

In  Fanner’s  view  winged  pine  seeds  are  an- 
cestral to  wingless  seeds,  and  corvids  provide 
the  selection  pressure  to  make  this  transition. 
He  presents  a scenario  for  evolution  of  wing- 
less seeds  in  two  groups  of  closely  related 
pine  species:  the  P.  ayachuite-strobifomis- 
flexilis  complex  in  Mexico  and  the  western 
United  States  and  the  P.  pannflora  complex 
in  eastern  Asia.  Pine  seeds  cached  by  corvids 


298 


THE  WILSON  BULLETIN 


Vol.  Ill,  No.  2.  June  1999 


are  less  likely  to  dessicate  and  more  likely  to 
germinate  than  seeds  dispersed  by  the  wind. 
Therefore,  in  drier  climates,  characteristics  in 
pines  that  enhance  dispersal  by  corvids — ver- 
tical fruiting  branches,  sessile  cones,  non- 
opening “breakaway”  scales,  seed-retaining 
cone  cores,  and  large,  wingless  seeds — would 
be  advantageous. 

Lanner’s  two  main  theses — whitebark  pine 
depends  on  Clark’s  Nutcracker  for  effective 
seed  dispersal,  and  the  whitebark  pine  is  a 
keystone  species — superficially  seem  difficult 
to  reconcile.  Many  animal  species  (birds, 
squirrels,  and  bears)  extract,  move,  and  eat 
whitebark  pine  seeds  (the  keystone  concept), 
but  the  nutcracker  is  essentially  the  only  agent 
of  seed  dispersal  (a  tight  mutualism).  The  final 
chapter  (“Is  the  Keystone  Slipping?”)  pre- 
sents the  case  that  whitebark  pine  is  a species 
upon  which  Clark’s  Nutcracker,  red  squirrel 
(Tamiasciunis  hudsonicus),  and  grizzly  beai' 
{Ursus  arctos)  depend  to  varying  degrees.  The 
whitebark  pine  is  seriously  threatened  by 
“competition  from  more  shade-tolerant  trees 
due  to  fire  exclusion;  heightened  bark  beetle 
attacks,  also  engendered  by  fire  exclusion; 
loss  of  habitat  through  global  warming;  and 
quick  death  from  a parasitic  fungus.”  If  these 
threats  severely  diminish  populations  of 
whitebark  pine,  over  the  long  term  and  in  a 
diffuse  ways,  the  nutcracker,  and  to  a lesser 
extent,  the  red  squinel  and  grizzly  bear  will 
suffer. 

The  writing  is  clear  and  the  pace  brisk.  Fig- 
ures and  tables  are  used  judiciously  and  the 
plates  are  excellent.  Some  readers  may  object 
to  instances  of  anthropomorphism  (“White- 
bark pine  takes  its  seeds  very  seriously  indeed 
when  it  comes  to  distributing  the  tree’s  re- 
sources.”) and  unabashedly  adaptationist  in- 
terpretations, but  this  is  perhaps  expected  in  a 
book  directed  to  a general  readership.  Details 
of  pine  phylogeny  and  the  uses  of  pine  nuts 
by  people  help  to  flesh  out  the  story.  I heartily 
recommend  this  book  to  all  interested  in  good 
natural  history  writing  and  forest  ecology  in 
particular. — R.  TODD  ENGSTROM. 


SKUAS  AND  JAEGERS:  A GUIDE  TO 
THE  SKUAS  AND  JAEGERS  OF  THE 
WORLD.  Klaus  Mailing  Olsen  and  Hans 


Larsson.  Yale  University  Press,  New  Haven 
and  London.  1997:  190  p.  12  color  and  1 
black-and-white  plate,  156  photographs,  20  of 
these  in  color  and  7 maps.  $35.00  (cloth). — 
This  is  an  excellent  field  guide  to  the  skuas 
and  jaegers,  and  the  only  current  reference  I 
know  that  correctly  illustrates  and  describes 
the  juvenal  and  winter  plumages  of  all  seven 
species.  For  this  reason,  the  book  is  an  essen- 
tial reference  for  all  interested  in  seabird  iden- 
tification. The  basic,  alternate,  and  juvenal 
plumages  of  three  jaeger  and  four  skua  species 
are  all  illustrated  with  attractive  color  paint- 
ings by  Larsson;  most  or  all  of  these  are  de- 
picted in  clear  photographs  as  well.  A painting 
of  adults  of  the  four  species  of  skuas  is  oddly 
reproduced  in  black-and-white;  presumably 
this  was  done  to  save  costs — it  slightly  mars 
the  otherwise  excellent  graphic  presentation. 

The  book  is  organized  into  an  Introductory 
section  of  28  pages  that  includes  such  topics 
as  “Breeding  Behavior”,  “Skuas  and  Man”, 
and  “Observing  Skuas  in  the  Field”;  followed 
by  the  species  accounts  which  form  the  main 
body  of  the  text.  The  species  accounts  contain 
sections  on  identification,  geographical  vaiia- 
tion,  food,  and  range  during  migration  and 
winter  (including  maps).  Compared  to  the  ob- 
vious precision  and  attention  to  detail  char- 
acteristic of  the  species  accounts,  I found  the 
Introductory  sections  to  be  somewhat  cursory 
and  containing  a number  of  questionable 
statements.  For  example,  on  page  9 it  is  stated 
that  “Unlike  gulls,  skuas  have  supraorbital 
salt  glands  . . . .” — gulls  certainly  have  these 
as  well,  as  anyone  who  has  watched  “runny- 
nosed”  gulls  at  the  seashore  probably  knows. 
The  dogma  about  clockwise  migrations 
around  the  North  Pacific  and  North  Atlantic 
Oceans  by  South  Polar  Skuas  is  peipetuated, 
despite  the  lack  of  evidence  to  support  the  no- 
tion that  any  individual  bird  follows  such  a 
path.  My  final  quibble  is  about  a point  of 
long-standing  confusion  about  the  occunence 
of  Brown  Skuas  (Catharacta  skua)  in  the  Ca- 
ribbean. While  the  confusion  is  obviously  no 
fault  of  the  authors,  this  book  would  have 
been  an  ideal  forum  for  the  settling  of  this 
issue.  I leave  it  to  the  reader  to  puzzle  out 
whether  the  recovered  banded  skuas  from 
“Guadeloupe”  and  the  “Lesser  Antilles”  are 
one  or  two  individuals,  and  how  likely  it  is 
that  either  were  misidentified  South  Polar 


ORNITHOLOGICAL  LITERATURE 


299 


Skuas  (C.  macconnicki).  Recent  sightings 
from  the  North  Atlantic  suggest  that  Brown 
Skuas  may  be  transequatorial  migrants. 

To  me,  the  most  novel  and  intei'esting  in- 
formation in  this  book  pertains  to  recent  ob- 
servations of  migrating  skuas  and  jaegers, 
both  from  land  and  from  ships  at  sea.  Not  only 
have  some  remarkable  numbers  of  these  birds 
been  observed,  but  compilations  of  such  ob- 
servations suggest  they  could  be  used  as  es- 
timates of  reproductive  success  the  previous 
season.  This  is  due  to  conespondence  within 
year's  of  the  proportions  of  juvenile  birds  seen. 
In  all,  the  authors  are  to  be  congratulated.  This 
is  a valuable  and  long  overdue  work,  skillfully 
and  artfully  executed. — RICHARD  R.  VEIT. 


JOHN  ABBOT’S  BIRDS  OF  GEORGIA: 
SELECTED  DRAWINGS  FROM  THE 
HOUGHTON  LIBRARY,  HARVARD  UNI- 
VERSITY. Introduction  and  commentary  by 
Vivian  Rogers-Price.  Beehive  Press,  Savan- 
nah, Georgia.  1997:  vii-xlii,  25  color  plates 
and  facing-page  commentary,  unnumbered 
and  unpaginated.  $125  (cloth  with  linen  slip- 
case). — This  publication  presents  the  first  col- 
or reproductions,  in  book  form,  of  bird  paint- 
ings of  John  Abbot  (1751-1840),  an  English- 
born  artist-naturalist  who  spent  most  of  his 
long  adult  life  in  Georgia.  He  completed  more 
than  5000  watercolors  of  natural  history  sub- 
jects including  more  than  a thousand  of  birds, 
most  of  which  are  extant,  and  another  thou- 
sand of  insects,  their  life  cycles  and  food 
plants.  The  remainder  includes  everything 
from  mites  and  ticks  to  crabs  and  millipedes. 
Abbot’s  father  had  encouraged  his  early  nat- 
ural history  illustration  predilection,  providing 
him  with  an  abundance  of  fine  bird  books  and 
professional  instruction  from  teacher  Jacob 
Bonneau.  At  age  22  Abbot  sailed  to  America 
and  landed  in  Virginia  where  he  remained  un- 
til the  rumblings  preceding  the  American  Rev- 
olution prodded  him  into  moving  to  the  then 
less  militant  Georgia.  He  remained  there  for 
the  rest  of  his  life  where  he  supported  himself 
largely  through  the  sale  of  natural  history  col- 
lections and  watercolor  paintings  that  were  ea- 
gerly sought  by  Europeans  from  his  agent  in 
London,  John  Francillon.  Abbot  published  lit- 
tle himself,  but  in  England  several  volumes 


about  butterflies,  moths,  and  other  insects 
were  illustrated  by  his  work.  John  Latham 
used  Abbot’s  drawings  and  specimens  in  his 
book  General  History  of  Birds  (1821-1824). 
Abbot  met  and  aided  Alexander  Wilson  in  the 
compilation  of  his  American  Ornithology,  and 
aided  George  Ord  in  the  completion  of  this 
major  work  following  Wilson’s  untimely 
death. 

Probably  at  least  partially  because  his  work 
was  published  by  others  and  because  of  his 
own  parochialism,  John  Abbot’s  bird  work 
was  largely  eclipsed  by  Wilson  and  Audubon. 
However,  Abbot’s  style  of  presentation  was 
apparently  modeled  after  George  Edwards, 
and  similar  to  William  Bartram’s,  with  styl- 
ized foregrounds  and  birds  perched  on 
dwarfed  trees.  Abbot  did  not  progress  artisti- 
cally beyond  the  limitation  of  this  approach, 
which  Wilson  and  Audubon  did,  and  by  the 
time  of  his  death  in  1840,  his  bird  portraits 
appear'  rather  archaic.  Nevertheless,  I find  it 
remarkable  that  nearly  two  centuries  elapsed 
before  a selection  of  this  important  Ar'nerican 
bird  artist’s  watercolor  paintings  of  birds  were 
published. 

This  large-format  (27  X 31  cm)  book  be- 
gins with  a 17-page  introduction  that  provides 
a biographical  sketch  of  Abbot’s  life,  high- 
lighting his  natural  history  collections  and 
paintings  and  his  European  colleagues  and  pa- 
trons. It  is  a scholarly  work,  with  115  end- 
notes  that,  in  small  type-face,  at  19  pages 
much  exceed  the  length  of  the  introduction. 
The  heart  of  the  book  is  the  25  watercolors  of 
birds  that  are  reproduced  at  the  same  size  as 
the  originals.  Each  of  the  paintings  is  accoi-n- 
panied  on  a facing  page  by  a brief  commen- 
tary and  any  of  Abbot’s  notes  that  relate  to 
the  species.  The  book  concludes  with  the  his- 
tory and  final  disposition  of  12  collections  of 
Abbot’s  original  paintings,  including  the  col- 
lection of  181  bird  portraits,  painted  from 
1801-1810,  from  which  the  paintings  repro- 
duced in  this  book  were  selected. 

I compared  the  plates  with  the  originals  at 
the  Houghton  Library  at  Harvard  University. 
The  quality  of  reproduction  was  variable  but 
generally  good  (the  American  Oystercatcher, 
Haematopus  palliatus,  was  faded  in  appear- 
ance but  the  Mourning  Dove,  Zenaida  ma- 
croura,  was  as  crisp  as  the  original).  The 
backgrounds  in  the  reproductions  were  darker 


300  THE  WILSON  BULLETIN  • VoL  III.  No.  2,  June  1999 


and  huffier  than  the  originals  and  made  the 
reproductions  warmer  than  the  originals,  but 
muted  the  colors,  especially  the  greens  some- 
what; and  the  contrast  was  not  as  crisp.  This 
was  particularly  a problem  for  the  light-col- 
ored egret  and  heron  paintings.  The  selection 
of  paintings  includes  a hummingbird,  which 
was  the  only  bird  Abbot  ever  painted  in  flight, 
a signed  and  dated  Bald  Eagle  (Haliaeetus 
leucocephalus),  a vulture,  an  owl,  three  wood- 
peckers, one  sparrow,  one  wren,  a nighthawk, 
a dove,  a crane,  seven  herons  and  ibises,  an 
oystercatcher,  a tern,  and  four  ducks.  I found 
the  heavy  emphasis  on  herons,  ibises,  and 
ducks  somewhat  perplexing,  although  many, 
like  the  preening  Wood  Duck  (Aix  sponsa), 
illustrated  interesting  poses  or  behaviors. 

I have  a strong  bias  against  footnotes  and 
endnotes,  and  found  flipping  back  and  forth 
from  the  introductory  text  to  the  endnotes  an- 
noying. The  commentary  facing  each  plate 
could  have  been  expanded — there  is  a lot  of 
empty  space.  Minor  problems  aside,  this  is  an 
important  work — thoroughly  researched  and 
attractively  presented.  I commend  the  Beehive 
Foundation  for  making  Abott’s  important 
work  accessible.  I recommend  the  book  to 
those  with  particular  interest  in  the  history  of 
ornithology  or  bird  art.  The  price  may  limit 
sales,  but  every  academic  library  should  have 
a copy.— WILLIAM  E.  DAVIS,  JR. 


OISEAUX  DE  LA  REUNION.  By  Nicolas 
Barre,  Armand  Barau,  and  Christian  Jouanin, 
illustrated  by  Nicolas  Barre.  Second  edition, 
revised  and  coirected  by  Nicolas  Barre  and 
Christian  Jouanin.  Les  Editions  du  Pacifique, 
62  me  du  Couedic,  75014  Paris.  1996:  207 
pp.,  10  color  plates,  numerous  color  and 
black-and-white  text  illustrations,  bibliogra- 
phy, indexes.  ISBN  2-87868-027-8.  Cloth.  No 
price  given. — Although  only  ten  pages  longer, 
the  second  edition  of  “Oiseaux  de  la  Re- 
union” is  an  improved  and  slightly  larger- 
sized  volume  (21  X 15  instead  of  20  X 12.5 
cm)  than  the  first  (which  I reviewed  in  1983, 
Auk  100:541-543).  The  disadvantage  of  a 
slightly  larger  size,  of  course,  is  that  the  new 
book  fits  less  easily  in  a pocket  than  the  orig- 
inal edition  did.  As  the  first  edition  had  been 
out  of  print  since  1990,  the  need  for  a second 


edition  had  been  felt  for  quite  some  time.  The 
problem  was  that  the  two  original  authors 
were  no  longer  available  for  this  job.  Sadly, 
one  of  them,  Armand  Barau,  who  was  an 
agronomist,  had  died  in  1987,  and  the  second 
author  and  illustrator,  Nicolas  Barre,  a veter- 
inarian, had  left  Reunion  in  1982,  the  year  the 
first  edition  was  published.  In  the  preface  to 
the  second  edition,  Christian  Jouanin,  Asso- 
ciate at  the  Museum  national  d’histoire  natu- 
relle  in  Paris,  who  brought  the  project  of  a 
second  edition  to  fmition,  explains  how  Ar- 
mand Barau’s  widow  and  Nicolas  Barre  per- 
suaded him  to  undertake  the  revision.  We  ai'e 
fortunate  that  he  accepted  this  task.  This  im- 
portant guide  (which,  when  still  in  print,  was 
difficult  to  get,  as  one  had  to  write  the  authors 
in  Reunion  in  order  to  obtain  copies)  has  now 
been  handsomely  produced  by  the  Editions  du 
Pacifique  in  Paris. 

Although  the  text  of  the  second  edition  is 
quite  similar-  to  that  of  the  first,  Jouanin  has 
brought  the  species  accounts  up  to  date  and 
has  incorporated  much  new  information,  both 
published  and  unpublished,  that  has  been 
gathered  on  the  status  and  distribution  of  the 
birds  of  Reunion  by  a number  of  workers 
since  the  early  1980s.  Eight  of  the  ten  color- 
plates  in  the  second  edition  are  the  same  as 
the  eight  of  the  first,  except  for  their  number- 
ing: Plates  VI,  VIII,  and  X were  numbered  as 
V,  VII,  and  VIII,  respectively,  in  the  first  edi- 
tion. Two  plates  are  new:  numbers  VII  (in- 
cluding seven  species  of  kites,  frigate-birds, 
plovers,  terns,  and  ducks;  pages  148—149)  and 
IX  (including  4 morphs  of  the  endemic  white- 
eye  Zosterops  borbonicus  borbonicus  and 
several  plumage  variations  in  the  endemic 
chat  Saxicola  tectes\  pages  178-179;  a very 
welcome  addition).  Interestingly,  the  eight 
original  plates  are  better  reproduced  in  the 
second  than  in  the  first  edition  (at  least  in  my 
copies).  They  are  fresher,  their  colors  are 
crisper,  and  their  slightly  larger  size  mean  that 
the  birds  on  each  plate  are  slightly  laiger,  an 
improvement  in  my  opinion.  As  in  the  first 
edition,  the  second  has  numerous  black-and- 
white  text  illustrations.  Not  all  original  draw- 
ings aie  included,  however,  and  some  have 
been  redrawn.  For  example,  the  attractive  and 
evocative  drawings  of  several  species  of  sea- 
birds and  a fishing  boat  (first  edition,  page  89) 
and  of  shorebirds  on  a mudflat  (first  edition. 


ORNITHOLOGICAL  LITERATURFi 


301 


page  131)  have  been  omitted  from  the  second 
edition,  a pity.  The  attitudes  of  plovers,  fresh 
from  a field  sketchbook  on  page  138  of  the 
first  edition,  has  been  redrawn  for  the  second 
edition  (page  132),  but  is  now  stiff  and  artis- 
tically much  less  interesting.  A novel  feature 
is  the  inclusion  of  a color  portrait  of  each  spe- 
cies, taken  from  the  plates,  next  to  the  name 
of  each  species  in  the  species  accounts.  As  in 
the  first  edition,  the  introductory  sections  are 
illustrated  with  color  photographs  of  habitats 
of  Reunion.  Whereas  there  were  6 such  pho- 
tographs in  the  first  edition,  the  second  has  12, 
thus  giving  the  potential  visitor  a better  over- 
view of  these  landscapes. 

In  addition  to  the  species  accounts  and  the 
plates,  this  volume  includes  an  excellent  de- 
scription of  Reunion  (pages  14-23);  a thor- 
ough and  fascinating  (if  sad)  review  of  the 
past  avifauna  (pages  26-52,  illustrated  with 
nicely  reproduced  color  plates  of  extinct  spe- 
cies, borrowed  from  older  publications);  a de- 


tailed presentation  of  the  modern  avifauna 
(pages  53-76,  including  hints  about  bird 
watching  and  information  about  conserva- 
tion); and  suggestions  about  how  to  use  the 
guide.  The  book  ends  with  a bibliography 
(pages  194-199)  and  four  indices  (French, 
Latin,  English,  and  Creole  names).  One  regret: 
this  book  does  not  have  a map  showing  the 
position  of  Reunion  in  the  Indian  Ocean  (this 
was  true  of  the  first  edition  also). 

Beautifully  produced,  full  of  carefully  re- 
searched information,  easy  to  use  thanks  to  its 
user-friendly  typography;  I strongly  recom- 
mend this  book  to  all  students  of  insular  avi- 
faunas. No  one  visiting  Reunion  can  be  with- 
out it.  As  I wrote  in  my  review  of  the  first 
edition,  this  book  “should  be  mandatory  read- 
ing for  all  school  children  of  Reunion  taking 
courses  on  the  geography  of  their  magnificent 
island.”  My  thanks  go  to  Christian  Jouanin 
for  having  seen  this  second  edition  through 
the  press.— FRANCOIS  VUILLEUMIER. 


This  issue  of  The  Wilson  Bulletin  was  published  on  10  May  1999. 


THE  WILSON  BULLETIN 


Editor  ROBERT  C.  REASON 


Editorial  Board  KATHY  G.  BEAL 


Department  of  Biology 

State  University  of  New  York 

1 College  Circle 

Geneseo,  NY  14454 

E-mail:  WilsonBull@geneseo.edu 


CLAIT  E.  BRAUN 
RICHARD  N.  CONNER 


Review  Editor  WILLIAM  E.  DAVIS,  JR. 


127  East  Street 

Foxboro,  Massachusetts  02035 


Editorial  Assistants  TARA  BAIDEME 


JOHN  LAMAR 
DANTE  THOMAS 
DORIS  WATT 


Index  Editor  KATHY  G.  BEAL 
616  Xenia  Avenue 


Yellow  Springs,  Ohio  45387 


SUGGESTIONS  TO  AUTHORS 

See  Wilson  Bulletin,  110:152-154,  1998  for  more  detailed  “Instructions  to  Authors.” 
http://www.ummz.lsa.umich.edu/birds/wilsonbull.html 


Submit  four  copies  of  manuscripts  intended  for  publication  in  The  Wilson  Bulletin,  neatly  typewritten, 
double-spaced,  with  at  least  3 cm  margins,  and  on  one  side  only  of  good  quality  white  paper.  Do  not 
submit  xerographic  copies  that  are  made  on  slick,  heavy  paper.  Tables  should  be  typed  on  separate  sheets, 
and  should  be  narrow  and  deep  rather  than  wide  and  shallow.  Follow  the  AOU  Check-list  (Seventh  Edition, 
1998)  insofar  as  scientific  names  of  U.S.,  Canadian,  Mexican,  Central  American,  and  West  Indian  birds 
are  concerned.  Abstracts  should  be  brief  but  quotable.  Where  fewer  than  5 papers  are  cited,  the  citations 
may  be  included  in  the  text.  Follow  carefully  the  style  used  in  this  issue  in  listing  the  literature  cited; 
otherwise,  follow  the  “CBE  Scientific  Style  and  Format  Manual”  (AIBS  1994).  Photographs  for  illustra- 
tions should  have  good  contrast  and  be  on  glossy  paper.  Submit  prints  unmounted  and  provide  a brief  but 
adequate  legend  for  each  figure  with  all  captions  on  a single  page.  Do  not  write  heavily  on  the  backs  of 
photographs.  Diagrams  and  line  drawings  should  be  in  black  ink  and  their  lettering  large  enough  to  permit 
reduction.  Original  figures  or  photographs  submitted  must  be  smaller  than  22  X 28  cm.  Alterations  in 
copy  after  the  type  has  been  set  must  be  charged  to  the  author. 


NOTICE  OF  CHANGE  OF  ADDRESS 


If  your  address  changes,  notify  the  Society  immediately.  Send  your  complete  new  address  to  Ornitho- 
logical Societies  of  North  America,  P.O.  Box  1897,  Lawrence,  KS  66044-8897. 

The  permanent  mailing  address  of  the  Wilson  Ornithological  Society  is:  do  The  Museum  of  Zoology, 
The  University  of  Michigan,  Ann  Arbor,  Michigan  48109.  Persons  having  business  with  any  of  the  officers 
may  address  them  at  their  various  addresses  given  on  the  back  of  the  front  cover,  and  all  matters  pertaining 
to  the  Bulletin  should  be  sent  directly  to  the  Editor. 


MEMBERSHIP  INQUIRIES 


Membership  inquiries  should  be  sent  to  Laurie  J.  Goodrich,  Route  2 Box  301  A,  New  Ringgold.  PA 
17960-9445;  E-mail:  goodrich@haukmountain.org. 


CONTENTS 


MAJOR  PAPERS 

A NEW  SPECIES  IN  THE  MYRMOTHERULA  HAEMATONOTA  SUPERSPECIES  (AVES;  THAM- 
NOPHILIDAE)  EROM  THE  WESTERN  AMAZONIAN  LOWLANDS  OF  ECUADOR  AND  PERU 

Niels  Krahhe,  Morton  L.  Isler.  Phyllis  R.  Isler,  Bret  M.  Whitney, 

Jose  Alvarez  A.,  and  Paul  J.  Greenfield 

COMPARATIVE  SPRING  HABITAT  AND  FOOD  USE  BY  TWO  ARCTIC  NESTING  GEESE  

Suzanne  Carriere,  Robert  G.  Bromley,  and  Gilles  Gauthier 

A TEST  OF  THE  CONDITION-BIAS  HYPOTHESIS  YIELDS  DIFFERENT  RESULTS  FOR  TWO 

SPECIES  OF  SPARROWHAWKS  (ACCIPITER)  - -- - - 

Edna  Gorney,  William  S.  Clark,  and  Yoram  Yom-Tov 

THE  DEVELOPMENT  OF  A VOCAL  THERMOREGULATORY  RESPONSE  TO  TEMPERATURE  IN 
EMBRYOS  OF  THE  DOMESTIC  CHICKEN  ..  Shawn  C.  Bugden  and  Roger  M.  Evans 

BEHAVIOR  AND  VOCALIZATIONS  OF  THE  CAURA  AND  THE  YAPACANA  ANTBIRDS  

Kevin  J.  Zimmer 

HABITAT  PATCH  SIZE  AND  NESTING  SUCCESS  OF  YELLOW- BREASTED  CHATS  

Dirk  E.  Burhans  and  Erank  R.  Thompson,  III 

AVIFAUNA  OF  A PARAGUAYAN  CERRADO  LOCALITY:  PARQUE  NACIONAL  SERRANIA  SAN 
LUIS,  DEPTO.  CONCEPCION  Mark  B.  Robbins,  Rob.  C.  Faucett,  and  Nathan  H.  Rice 

NOTES  ON  THE  AVIFAUNA  OF  TABASCO  — - 

Kevin  Winker,  Stefan  Arriaga  Weiss,  Juana  Lourdes  Trejo  P.,  and  Patricia  Escalante  P. 

PREDATION  OF  SMALL  EGGS  IN  ARTIFICIAL  NESTS:  EFFECTS  OF  NEST  POSITION.  EDGE. 

AND  POTENTIAL  PREDATOR  ABUNDANCE  IN  EXTENSIVE  FOREST  

Richard  M.  DeGraaf,  Thomas  J.  Maier,  and  Todd  K.  Fuller 

BIRD  USE  OF  BURNED  AND  UNBURNED  CONIFEROUS  FORESTS  DURING  WINTER  ^ 

Karen  J.  Kreisel  and  Steven  J.  Stein 

NEST  PREDATORS  OF  OPEN  AND  CAVITY  NESTING  BIRDS  IN  OAK  WOODLANDS  

Kathryn  L.  Purcell  and  Jared  Venter 

SHORT  COMMUNICATIONS 

JUVENILE  MARBLED  MURRELET  NURSERIES  AND  THE  PRODUCTIVITY  INDEX  

Katherine  J.  Kuletz  and  John  F.  Piatt 

"SNORKELING"  BY  THE  CHICKS  OF  THE  WATTLED  JACANA  

Carlos  Bosque  and  Emilio  A.  Herrera 

RAPID  LONG-DISTANCE  COLONIZATION  OF  LAKE  GATUN.  PANAMA,  BY  SNAIL  KITES 
George  R.  Angehr 

THE  "SIGNIFICANT  OTHERS”  OF  AMERICAN  KESTRELS:  COHABITATION  WITH  AR- 
THROPODS   - - --  Jeffrey  P.  Neubig  and  John  A.  Smallwood 

BARRED  OWL  NEST  IN  ATTIC  OF  SHED  C Stuart  Hou.ston 

DOUBLE  BROODING  IN  THE  LONG-EARED  OWl 

. Jeffrey  S.  Marks  and  Alison  E.  H.  Perkins 

PLANNING  TO  FACILITATE  CACHING:  POSSIBLE  SUET  CUTTING  BY  A COMMON  RA- 
VEN   Bernd  Heinrich 

PAIRING  SUCCESS  OF  WOOD  THRUSHES  IN  A FRAGMENTED  AGRICULTURAL  LAND- 
SCAPE   Lyle  E.  Friesen,  Valerie  E.  Wyatt,  and  Michael  D.  Cadman 

CONNECTICUT  WARBLER,  A NORTH  AMERICAN  MIGRANT  NEW  TO  ECUADOR  

..  OlafJahn.  Maria  Eugenia  Jara  Viteri,  and  Karl-L.  Schuchmann 

PARENTAL  BEHAVIOR  OF  A BIGAMOUS  MALE  NORTHERN  CARDINAL  

Randall  Breitwisch,  Amy  J.  Schilling,  and  Joshua  B.  Banks 

SPECIAL  REPORT 

A SURVEY  OF  UNDERGRADUATE  ORNITHOLOGY  COURSES  IN  NORTH  AMERICA  

Edward  H.  Burtt,  Jr.  and  W.  Herbert  Wilson,  Jr. 

ORNITHOLOGICAL  LITERATURE 


157 

166 

181 

188 

195 

210 

216 

229 

236 

243 

251 

257 

262 

265 

269 

272 

273 
276 
279 
281 
283 

287 

294 


TKc  Wilson  Bulletin 

PUBLISHED  BY  THE  WILSON  ORNITHOLOGICAL  SOCIETY 


VOL.  Ill,  NO.  3 SEPTEMBER  1999  PAGES  303-456 

(ISSN  (K)43-S643) 


THE  WILSON  ORNITHOLOGICAL  SOCIETY 
FOUNDED  DECEMBER  3,  1888 

Named  after  ALEXANDER  WILSON,  the  first  American  Ornithologist. 

President — John  C.  Kricher,  Biology  Department,  Wheaton  College,  Norton,  Massachusetts  02766; 
E-mail:  JKricher@wheatonma.edu. 

First  Vice-President — William  E.  Davis,  Jr.,  College  of  General  Studies,  871  Commonwealth  Ave.,  Boston 
University,  Boston,  Massachusetts  02215;  E-mail:  WEDavis@bu.edu. 

Second  Vice-President — Charles  R.  Blem,  Dept,  of  Biology,  816  Park  Ave.,  Virginia  Commonwealth 
Univ.,  Richmond,  Virginia  23284;  E-mail:  cblem@saturn.vcu.edu. 

Editor — Robert  C.  Reason,  Department  of  Biology,  State  University  of  New  York,  1 College  Circle, 
Geneseo,  New  York  14454;  E-mail:  WilsonBull@geneseo.edu. 

Secretary — John  A.  Smallwood,  Department  of  Biology,  Montclair  State  University,  Upper  Montclair, 
New  Jersey  07043;  E-mail:  Smallwood@saturn.montclair.edu. 

Treasurer — Doris  J.  Watt,  Department  of  Biology,  Saint  Mary’s  College,  Notre  Dame,  Indiana  46556; 
E-mail:  DWatt@saintmarys.edu. 

Elected  Council  Members — Charles  E Thompson  and  Sara  R.  Morris  (terms  expire  2000),  Jonathan  L. 
Atwood  and  James  L.  Ingold  (terms  expire  2001),  Robert  A.  Askins  and  Jeffrey  R.  Walters  (terms 
expire  2002). 

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  Bulletin  is  sent  to  all  members  not  in  arrears  for  dues. 

THE  JOSSELYN  VAN  TYNE  MEMORIAL  LIBRARY 
The  Josselyn  Van  Tyne  Memorial  Library  of  the  Wilson  Ornithological  Society,  housed  in  the  University 
of  Michigan  Museum  of  Zoology,  was  established  in  concurrence  with  the  University  of  Michigan  in 
1930.  Until  1947  the  Library  was  maintained  entirely  by  gifts  and  bequests  of  books,  reprints,  and  orni- 
thological 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  administered  by  the  Library  Committee.  William  A.  Lunk,  University  of 
Michigan  is  Chairman  of  the  Committee.  The  Library  currently  receives  over  200  periodicals  as  gifts  and 
in  exchange  for  The  Wilson  Bulletin.  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  University 
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 
University  of  Michigan,  1 109  Geddes  Ave.,  Ann  Arbor,  MI  48109-1079  USA.  Contributions  to  the  New 
Book  Fund  should  be  sent  to  the  Treasurer. 


THE  WILSON  BULLETIN 
(ISSN  0043-5643) 

THE  WIL.SON  BULLETIN  (ISSN  0043-5643)  is  published  quarterly  in  March,  June,  September,  and  December  by 
the  Wilson  Ornithological  Society,  810  East  10th  Street,  Lawrence,  KS  66044-8897.  The  subscription  price,  both  in  the 
United  States  and  elsewhere,  is  S40.00  per  year.  Periodicals  postage  paid  at  Lawrence,  KS.  POSTMASTER:  Send  address 
changes  to  THE  WILSON  BULLETIN,  P.O.  Box  1897,  Lawrence,  KS  66044-8897. 

Back  issues  or  single  copies  are  available  for  .$12.00  each.  Most  back  issues  of  the  Bulletin  are  available  and  may  be 
ordered  from  the  Treasurer.  Special  prices  will  be  quoted  for  quantity  orders. 

All  articles  and  communications  for  publications,  books  and  publications  for  reviews  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  the  OSNA,  P.O.  Box  1897, 
Lawrence.  K.S  66044-8897.  Phone;  (785)  843-1221;  FAX:  (785)  843-1274. 

© Copyright  1999  by  the  Wilson  Ornithological  Society 
Printed  by  Allen  Press,  Inc.,  Lawrence,  Kansas  66044,  U.S.A. 


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


THE  WILSON  BULLETIN 

A QUARTERLY  JOURNAL  OL  ORNITHOLOGY 
Published  by  the  Wilson  Ornithological  Society 

VOL.  Ill,  NO.  3 SEPTEMBER  1999  PAGES  303-456 


Wilson  Bull..  111(3),  1999,  pp.  303-313 


ANTILLEAN  SHORT-EARED  OWLS  INVADE  SOUTHERN  LLORIDA 

WAYNE  HOFFMAN,'  GLEN  E.  WOOLFENDEN,^  AND  P.  WILLIAM  SMITH^ 


ABSTRACT. — Recently,  Short-eared  Owls  {Asio  flammeus)  have  invaded  extreme  southern  Florida  during 
spring  and  summer,  most  appear  to  be  post-fledging  dispersers.  Morphological  and  plumage  characteristics 
identify  the  specimens  as  coming  from  the  Antilles,  most  likely  from  Cuba,  where  numbers  and  range  have 
expanded  greatly  in  recent  years.  This  dispersal  continues  a trend  that  began  in  other  bird  species  more  than 
half  a century  ago.  Since  1932  about  one  landbird  species  per  decade  has  colonized  southern  Florida  from  the 
Antilles.  Received  27  May  1998,  accepted  30  Nov.  1998. 


During  the  last  two  decades.  Short-eared 
Owls  {Asio  flammeus)  have  occurred  with  in- 
creasing frequency  in  extreme  southern  Flor- 
ida. especially  from  March  through  Septem- 
ber. In  the  absence  of  specimens,  these  reports 
were  assumed  to  represent  individuals  of  the 
nominate  race  A.  flammeus  flammeus,  which 
was  the  only  form  of  this  polytypic  species 
known  from  the  North  American  continent. 
Asio  f flammeus  is  Holarctic;  in  the  western 
hemisphere  it  breeds  in  northern  North  Amer- 
ica and  migrates  south  as  far  as  southern  Unit- 
ed States,  Mexico,  and  rarely  the  West  Indies 
(American  Ornithologists’  Union  1957,  1998). 
During  the  1990s  we  obtained  several  speci- 
mens of  Short-eared  Owls  from  extreme 
southern  Florida.  Here  we  summarize  all  re- 
cent records  and  reports  of  Short-eared  Owls 
from  southern  Florida,  describe  the  character- 


' 260  SE  97"’  Ct.,  South  Beach,  OR  97366. 

^ Archbold  Biological  Station,  Venus,  FL  33960. 
3p.O.  Box  1992,  Ocean  Shores,  WA  98569. 

* Corresponding  author; 

E-mail:  gwoolfenden@archbold-station.org 


istics  of  Holarctic  and  Antillean  Short-eared 
Owls  and  conclude  that  most  of  the  recent 
spring-summer  records  are  from  an  Antillean 
population,  and  briefly  review  their  nomencla- 
ture. Finally,  we  discuss  possible  causes  for 
the  dispersal  of  these  owls  into  Florida  and 
their  potential  for  colonizing  the  North  Amer- 
ican continent. 

SPECIMENS,  PHOTOGRAPHS  AND 
REPORTS 

Specimens. — Between  July  1990  and  March 
1998,  we  obtained  eight  dead  Short-eared 
Owls  from  the  Florida  Keys,  Monroe  County, 
Florida.  Three  of  the  six  males  showed  no 
molt,  while  three  showed  light,  scattered  body 
molt.  All  six  males  had  small  testes  and  all 
had  extensive  black  feathering  around  the 
eyes,  which  in  the  nominate  race  typifies  Ju- 
venal plumage  (Holt  and  Leisure  1993).  One 
female  (GEW  5902,  25  May  1996,  ovary  15 
X 4 mm,  largest  ovum  1 mm,  substantial  body 
molt)  also  appeared  to  be  juvenile,  but  the  oth- 
er (GEW  5889,  late  April  1994,  ovary  17  X 


<— 


FRONTISPIECE.  Antillean  Short-eared  Owl  photographed  in  Ft.  Zachary  Taylor  State  Park,  Key  West, 
Monroe  County,  Florida  on  4 April  1994  by  Wayne  Hoffman. 


303 


304 


THE  WILSON  BULLETIN  • Vol.  IN,  No.  3,  September  1999 


4,  largest  ovum  2 mm)  had  worn  plumage  and 
possibly  a regressing  brood  patch,  and  might 
have  been  an  adult. 

Comparative  material  included  five  Short- 
eared Owl  specimens  from  the  Antilles,  four 
from  the  southern  Florida  mainland,  and  46 
from  elsewhere  in  North  America.  In  June 
1995  Orlando  Garrido  kindly  loaned  us  an  un- 
sexed  adult  specimen  (MNHN-1595)  collect- 
ed in  Sancti  Spiritus  Province,  Cuba.  In  1996 
Garrido  donated  to  the  Archbold  Biological 
Station  collections  a male  specimen  collected 
near  Havana  (GEW  5925).  We  also  examined 
two  nestlings  collected  in  the  Dominican  Re- 
public in  November  1963  (LSUMZ  142354 
and  142355;  Schwartz  and  Klinikowski  1965). 
The  four  mainland  specimens,  all  from  Dade 
County,  were  two  from  Everglades  National 
Park  (EVER  5035,  collected  5 January  1971; 
GEW  5890,  collected  9 December  1990)  and 
two  in  the  University  of  Miami  Research  Col- 
lections (UMRC  948,  collected  7 February 
1956;  UMRC  5387,  collected  8 November 
1966),  which  now  are  in  the  collections  at 
Archbold  Biological  Station. 

In  April  1997,  Hoffman  measured  six  spec- 
imens of  the  Holarctic  Asia  flammeus  flam- 
meus  in  the  Pennsylvania  State  University  col- 
lections and  40  specimens  in  the  Carnegie 
Museum  collections.  All  were  taken  in  North 
America  throughout  the  year.  The  few  speci- 
mens with  missing  or  incompletely  grown  pri- 
maries or  central  rectrices  were  excluded  from 
analysis.  Hoffman  also  examined  and  mea- 
sured the  single  Puerto  Rican  specimen  in  the 
Carnegie  Museum. 

Measurements  taken  include  lengths  of 
wing  (flattened),  tail,  tarsus,  and  culmen  from 
the  cere.  Measurements  were  taken  as  de- 
scribed by  Palmer  (1962)  and  Cramp  and  co- 
workers (1977).  Tarsi  are  difficult  to  measure 
on  the  feathered  feet  of  owls,  so  a dissecting 
probe  was  used  to  assist  in  locating  the  mea- 
suring points  on  the  posterior  of  the  intertarsal 
joint  and  the  anterior  of  the  middle  toe  artic- 
ulation with  the  tarsometatarsus. 

Photographs  and  reports. — We  examined 
photographs  of  six  individuals  that  did  not  be- 
come specimens:  three  from  the  Dry  Tortugas, 
two  from  other  of  the  Florida  Keys,  and  one 
from  the  Gulf  of  Mexico  off  Hernando  Coun- 
ty. We  reviewed  reports  of  Short-eared  Owls 
from  Florida  published  in  Audubon  Field 


Notes  and  American  Birds  (Loftin  et  al.  1991) 
and  evaluated  other  written  and  verbal  reports 
of  these  owls  in  southern  Florida  since  1978. 

RESULTS 

Current  status  of  Short-eared  Owls  in 
southern  Florida. — Our  review  of  citations  in 
Audubon  Field  Notes  and  American  Birds 
yielded  about  68  reports  of  Short-eared  Owls 
in  Florida  before  1978.  With  one  exception,  a 
bird  seen  14  June  1963  at  Lakeport,  Glades 
County,  all  pre-1978  reports  were  of  occur- 
rences between  early  October  and  late  March. 
We  obtained  information  on  30  occurrences  of 
33-37  Short-eared  Owls  in  Florida  since 
1978,  including  16  records  (specimens  and 
photographs)  and  14  reports  (no  tangible  ev- 
idence). Twenty-three  of  the  30  appeared  in 
spring  and  summer,  outside  the  early  October- 
late  March  dates  dominating  the  earlier  peri- 
od. During  spring— summer  1994  the  influx 
seemed  particularly  heavy.  The  first  bird  was 
located  at  Ft.  Taylor  State  Recreation  Area  in 
Key  West,  25  March  (photographed  on  4 
April).  It  was  joined  by  a second  bird  in  mid- 
April;  both  disappeared  by  mid-May.  Mean- 
while a birding  tour  located  three  birds  at  the 
Dry  Tortugas  on  8 April  (Woolfenden,  pers. 
obs.).  One  was  picked  up  there  in  weakened 
condition  on  18  April  and  died  while  in  transit 
to  Key  West  for  treatment  (GEW  5889).  On  8 
June,  W.  B.  Robertson,  Jr.  flushed  a group  of 
four  owls  on  Long  Key,  Dry  Tortugas.  Thus, 
a minimum  of  five,  possibly  as  many  as  nine, 
owls  were  found  in  southern  Florida  in  spring 
1994.  Another  record  that  summer  was  of  an 
individual  plucked  from  the  water  offshore  of 
Hernando  Co.,  in  June  (Table  1).  The  dates  of 
occurrence  of  these  records  and  reports  sug- 
gest a source  other  than  the  Holarctic.  The 
specimens  we  have  obtained  allowed  us  to  test 
this  hypothesis  using  geographic  variation  de- 
scribed for  Short-eared  Owls. 

Structural  and  plumage  differences  between 
the  Antillean  and  Holarctic  Short-eared 
Owls. — Based  on  Ridgway  (1914)  and  Wet- 
more  (1928),  and  the  specimens  we  examined, 
Antillean  Short-eared  Owls  differ  from  Hol- 
arctic Short-eared  Owls  in  size,  proportions, 
and  plumage.  These  differences  appear  ade- 
quate to  distinguish  all  specimens  in  the  hand, 
and  to  allow  identification  of  birds  observed 
closely  in  the  field.  They  also  appear  sufficient 


Hoffman  el  al.  • SHORT-EARED  OWLS  IN  SOUTHERN  FLORIDA 


00 

r-* 

ON 


■n 

'u 

O 

Ph 

c 

u. 

<3J 


a 

(U 


V3 

“O 

0 

o 

OJ 


O 

“O 

o 


00 


UU 

CQ 

< 

H 


c/3 

(U 


c 

< 


c 

< 


C^* 

c/3  C/3 

c/3 

c/3 

0 

c/3 

c/3 

c/3 

c/3 

c/3 

c/3 

c/3 

c/3 

c/3 

c/3 

c/3 

c/5 

c/3 

c/3 

0 

D <U 

c c 

OJ 

c 

0 

E 

u 

0 

-3 

c 

£ 

c 

£ 

1 £ 

-E 

c 

£ 

1 3 1 

.E 

1 c 

-E 

c 

.E 

c 

-E 

E 

.E 

E 

-E 

c 

E 

_E 

E 

£ 

c 

0 

U 

-2 

0 

< < 

< 

< 

d:  < 

< 

1 < 

< 

< 1 

1 < 

< 

< 

< 

< 

< 

< 

< 

< 

n: 

s: 

a 

eg 

u 

OJ) 

O 

O 

J= 

D- 


x: 

Cl 

eg 

u 

OjO 

O 

O 

-C 

Cl 

TC 

O 

-C 


-O 

3 

0, 


c/3 

a 

eg 

Un 

Oi) 

o 

o 

-C 

D- 


0, 

Z 

w 

"o' 

> 

eg 

c/3 

W) 


— O ri  NO 
NO  ON  NO  00 
00  00  00  00 
lo  '/o  lo  wo 

^ ^ ^ ^ 
U PJ  UJ  UJ 
O O O O 

00  00  oo"  oo" 
QQ  CQ  CQ  03 
< < < < 


ej 

<D 

a - 

c/^  : 


c 

(U 

E 

'o 

O 

D, 


C 

O 

E 

o 

o 

a 


00  00  00  00 


On 

ri 

0 

rJ 

0 

00 

ON 

0 

0 

00 

00 

On 

ON 

c 

ON 

wo 

wo 

W^J 

WO 

eg 

WO 

c 

T3 

w 

w 

w 

w 

w 

a 

□ 

a 

0 

u 

P^ 

0 

00 

c/f 

00 

00' 

n 

00 

oa 

pa 

PQ 

pa 

PQ 

c/3 

< 

c/3 

< 

< 

< 

c/3 

< 

X 

X 

X 

a 

eg 

C 

o; 

CL 

eg 

C 

0 

d 

<u 

d 

0 

a 

eg 

d 

OJ 

00 

E 

00 

P 

E 

E 

bo 

(— 

1“ 

0 

0 

0 

. M 

CJ 

CJ 

u 

0 

CJ 

0 

(U 

0 

0 

<1/ 

1) 

0 

<1) 

X 

X 

1 X 1 

1 CL 

a 

CL 

1 

a 

Cl 

00 

1 Cl  I 

1 01 

c/5 

00 

1 PL, 

00 

Cl- 

eg 

u 

0/J 

o . 


NJ  7:  SJ 

(U  O 4J 

a *5  c 

c/3  Cl  c/: 


00  sO 


O 

ej 


ON 


O 

o 


OJ 

a 

>c 

OiJ 

eg 

c 

eg 


C 

O 

CJ 


<u 

Cl 


r- 

On 

__ 

0 

0 

eg 

ON 

— 

wo 

3 

0 

X 

On 

X 

(N 

U 

c 

X 

On 

— 

CP 

0 

Ig 

P 

66 

ON 

(O 

c5 

PP 

c/3 

bo 

CQ 

rn 

m 

c 

u 

0 

0 

c/3 

c/3 

c/3 

c/3 

1/3 

c/3 

eg 

X 

c/3 

E 

-o 

s 

"O 

u 

■p 

XJ 

u 

■a 

u 

■0 

E 

0 

0 

Cid 

TD 

u 

P 

X 

-a 

m 

5 

£ 

5 

£ 

£ 

c 

eg 


< q;  < 


< < < < 


oa 

q; 


pa 


6 

E 

o 

O 


u 

o 

a 

ob 

c 

PP 


CL  (D 
. CL 


D (1) 

^ -C 


o 

OC 


u 


E 

E 

o 

0 

c/) 

u 

a s' 

J s 

S 

o E 

*5  ^3"" 
ctC  — 


E 

o 

o 

c/5 

u 

D 

a 


c 

c 


c 

< 


CQ  ^ W 
tc  d 


oa  u 

cd  ^ 


u 

•—> 

s 

0 c/3 

c 

CJ  u 

0 

0 

c/3 

C/3  O- 

c 

eg 

t: 

0 

X 

0 

0 c 

^ £ 

i: 

0 

DP 

(U 

X 

X 

£ 

3 'u 

u ^ 

c 

eg 

P 

L— I 

c 

X 


00  Q 


>> 

^ eg 

^ CQ 

eg 

“O 


ojj 

I 

Ui 

Q 


>» 

o 


-C 

ri 


c/3  c/3  c/3  PU 

eg  eg  eg 
0/j  00  wj  >; 


>N 

<u  cu 


p >> 


eg  Uri 
00 

3 .E 

u eg 

£ 

>U-ClEQQQm5£QQS 


CL  0- 


C/5 

eg 

00 

3 

tc 

f2 


Xi 

(N 


c/3  C/3 

13  “O 


X)  X 


>>  >.  P 

u u o 


C 

(U  O 

rt 

aj  ^ 

^ s 


a (N 
Ofl  ^ 

3 « 

U (U 

t2  ^ 


c/3  c/3  O 

eg  eg  -rt 
00  00  c 


(U 


00 

I 

f^, 


eg 


00 

C 

iL 

ri 

c 

On 

a 

a 

a 

1 

*— > 

> 

u< 

< 

CJ 

< 

< 

u 

eg 

O/J 

c, 

CJ 

iL 

ul 

rj 

0 

0 

ri 

V 

ON 

— 

3 

Cl 

— 

•u 

a 

a 

C^J 

1 

Z 

E 

7 

Q 

n 

1 

ri 

1 

OJ 

< 

00 

< 

3 

•—i 

G 

< 

< 

0 

r- 

On 

E 

'C 

00 

0 

0 

00 

r^i 

rj 

— 

no 

ri 

— 

(N 

< 

WO 

n 

u 1)  u u ^ 

Q Q Q o 


eg 


eg 

<u 

I 

iL 

eg 

S 

wo 


c 
o 
£ 

?3  P 

>N  eg 


c/3 

eg 

00  - . 

*-r  u.  QJ  — . 

c/o  Q O S 


eg 


c/3  c/3 

eg  eg 

00  n 50 

i2  j f2 

^ 

u 1)  u 

Q Q 


C/3 

•o 


C 

3 


C ^ CL 

3 ^ (D 

^ u 00 
P 

00  <u  — 


> 

o 

z 


— . 

b 'nP 
D.  « a 

< 5 < 

_ - . , oo(Nr- 

(N  (N  — n — 


3 X 
3 


eg  X 

3 


s ^ s 

— vO  IT) 


r^t^oooooooo(X)ooooooooooovONONONONO'ONaNOsaNONONOsONONONO'ON 

c^o^a^a^o^CT^^a\o^o^o^a^o^o^a^c^o^o^o^o^c^O'o^a^c^o^a^o^o^o^ 


305 


306 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


TABLE  2.  Measurements  of  Short-eared  Owl  specimens.  All  measurements  (in  mm)  taken  by  Hoffman.  Sta- 
tistical comparisons  of  means  are  by  Student’s  t-test  (t),  unpaired.  Probability  (P)  of  equal  means  is  two-tailed. 


/4.  / fiammeus 

West  Indian 

Measurement 

Sex 

n 

Mean  (Range) 

n 

Mean  (Range) 

Wing 

M 

25 

313.8 

(306-323) 

8 

289.8 

(286-291) 

13.77 

<0.001 

Wing 

E 

24 

316.3 

(308-325) 

3 

294.7 

(292-299) 

8.30 

<0.001 

Wing 

both 

49 

315.0 

(306-325) 

11 

291.0 

(286-299) 

16.05 

<0.001 

Tail 

M 

25 

142.0 

(133-148) 

8 

132.8 

(127-141) 

5.51 

<0.001 

Tail 

F 

24 

146.1 

(134-153) 

3 

140.7 

(136-146) 

1.68 

>0.05 

Tail 

both 

49 

144.0 

(133-153) 

11 

134.9 

(127-146) 

5.25 

<0.001 

Tarsus 

M 

25 

42.7 

(38-46) 

8 

52.2 

(50-53.5) 

10.28 

<0.001 

Tarsus 

F 

24 

43.7 

(39-46) 

3 

50.7 

(50-54) 

5.55 

<0.001 

Tarsus 

both 

49 

43.2 

(38-46) 

1 1 

51.8 

(50-54) 

11.67 

<0.000 

Culmen 

M 

24 

26.3 

(24.0-28.0) 

8 

28.9 

(27.8-29.6) 

5.21 

<0.001 

Culmen 

F 

24 

26.2 

(24.0-28.3) 

3 

29.8 

(29.5-30.1) 

4.51 

<0.013 

Culmen 

both 

48 

26.3 

(24.0-28.3) 

11 

29.1 

(27.8-30.1) 

6.94 

<0.001 

to  distinguish  Antillean  specimens  from  all 
other  members  of  Asia  flarnmeus. 

Compared  to  Holarctic  Asia  f flarnmeus, 
Antillean  Short-eared  Owls  have  shorter 
wings  and  tails,  longer  tarsi,  and  slightly  larg- 
er bills  (Table  2).  Male  and  female  A.  f.  flam- 
meus  overlapped  broadly  for  all  measure- 
ments, and  the  Antillean  male  and  female 
specimens  overlapped  broadly  in  tail  and  tar- 
sus measurements.  No  overlap  existed  be- 
tween A.  f.  flarnmeus  and  the  Antillean  spec- 
imens in  length  of  the  wing,  tarsus,  and  cul- 
men.  The  two  populations  overlapped  sub- 
stantially only  in  tail  length.  Because  of  the 
broad  overlap  between  the  sexes  for  most 
measurements,  tests  of  significant  mensural 
differences  between  the  groups  were  run  for 
both  sexes  combined  as  well  as  for  males  and 
females  separately.  Differences  between  the 
groups  were  highly  significant  for  all  compar- 
isons except  tail  length  among  females  (Table 
2).  Hoffman’s  measurements  of  A./,  flarnmeus 
specimens  are  similar  to  those  published  in 
Cramp  and  coworkers  (1985),  except  that  the 
culmen  measurements  averaged  3 mm  shorter, 
suggesting  a methodological  difference.  Ridg- 
way  (1914),  Wetmore  (1928),  and  Garrido 
(1984)  have  published  wing  measurements  for 
Antillean  owls,  and  Marshall  (pers.  comm.) 
provided  measurements  on  six  specimens  at 
the  US  National  Museum  (Table  3).  Together, 
these  sources  provided  measurements  of  15 
specimens  (five  males,  three  females,  seven 
unsexed);  wing  lengths  ranged  from  274  to 


300  mm,  which  are  similar  to  ours  (286-299 
mm). 

The  most  notable  plumage  differences  in- 
volve coloration  of  the  upper  back  and  the  un- 
derparts (Fig.  1).  On  Holarctic  birds  the  con- 
tour feathers  of  the  upper  back,  between  and 
anterior  to  the  scapulars,  are  mostly  tawny 
with  a dark  brown  central  stripe.  On  Antillean 
birds  these  back  feathers  are  mostly  dark 
brown  with  tawny  edgings.  As  a result  the  up- 
per back  of  Holarctic  birds  appears  distinctly 
striped,  whereas  the  backs  of  Antillean  birds 
appear  overall  dark  brown,  or  dark  brown 
with  obscure  tawny  mottling.  The  underparts 
of  A.  f.  flarnmeus  are  heavily  streaked  with 
dark  brown  on  a pale  tawny  to  whitish  back- 
ground (Fig.  1).  Streaks  are  broadest  and  most 
dense  on  the  upper  breast,  and  gradually  be- 
come narrower  and  more  sparse  posteriorly. 
The  streaked  feathers  most  posterior  are  lo- 
cated on  or  near  the  knee  joints.  The  under- 
parts of  the  Antillean  owls  are  much  more 
buffy  overall.  The  streaking  is  similar  on  the 
upper  breast,  but  abruptly  becomes  much  nar- 
rower and  more  sparse  at  mid-breast.  The  low- 
er breast  and  belly  are  mostly  unstreaked  with 
a few  narrow  streaks  (less  than  2 mm  wide) 
on  the  flanks.  The  feathering  on  and  around 
the  knee  joints  is  unstreaked.  The  difference 
in  pattern  and  coloration  of  the  underparts  was 
noted  as  early  as  1770  (Buffon  in  Wetmore 
and  Swales  1931).  Other  plumage  differences 
include  the  color  of  the  upper  tail  coverts 
(fairly  dark  brown  in  Antillean  birds  versus 


TABLE  3.  Measurements  (in  mni)  of  Short-eared  Owls,  from  the  literature. 


Hoffman  el  al.  • SHORT-EARED  OWLS  IN  SOUTHERN  FLORIDA 


307 


CJ 

o 

3 

O 

(/5 


C 

a> 


£ 


CO 

c 


3: 


X 

K 

C/5 


oc 

00 

ON 

ON 

ON 

00 

ON 

r-- 

on 

ON 

dj 

D 

r-* 

o 

u 

m 

u, 

ro 

O 

X 

c 

a 

0 

ON 

0 

00 

(N 

0 

00 

r- 

00 

o 

n 

iO 

o 

(N 

lO 

(N 

— 

Os 

cn 

ON 

00 

00 

ON 

ON 

*0 

r- 

Q 

■a 

lo 

ON 

c 

CZ 

(U 

c 

cZ 

(U 

Cu 

O 

X 

X 

Cl 

kn 

0 

X 

0 

*T^ 

k« 

0 

C3 

> 

c 

P 7 <5 

c 

P 7 

c 

> 

u ^ 


X 

z 


2 ^ 
U ^ 


fc 

a 

O 


■o 


00 

ON 

On 

n 

1 

(N 

1 

1 

in 

1 

CN 

in 

o 

00 

On 

CN 

C4 

s 

— 

NO 

00 

1 

1 

00 

1 

in 

04 

in 

CN 

NO 

nT 

Tj- 

NO 

in 

in 

ro 

1 

— 

1 

1 

r-- 

o 

1 

NO 

ro 

m 

04 

ro 

NO 

in 

o 

NO 

n 

00 

in 

ro 

o 

00 

00 

ro 

ri 

ro 

ro 

1 

1 

n 

1 

1 

1 

00 

1 

o 

04 

ON 

00 

in 

in 

o 

On 

00 

e 

o 

On 

00 

00 

r- 

0- 

fO 

n 

o 

00 

w 

w 

in, 

ro 

ri 

ON 

0^1 

0- 

00 

— H 

00 

— - 

ON 

m 

m 

ro 

n 

rj 

■D 

yj 

y) 

C/5 

c/5 

dJ 

(U 

a 

o; 

'u  — O')  ’C  — 'C  'u 

r3  c3  c3  c3 

> > > > 


‘S:  ^ 


oc 


^ c 
o a 
t-g 


^ c 

5 > X) 
e e 


>1 


S S 

s $ 


c^U  cx.c:l. 


(U 


a 

B 

V 

u. 


"O 

(U 

X 

Oj 

yj 

c 

D 


— ^5 

“ o I 

=:  c 

<u  , . 

1/5  P ^ 

3 £ 

ii:  3 ^ 


,o 
o , 


s:  v2 


c 

£ = 


- o 


E 

E 2 


C 00  -t 

s I § 


E > 


£ >S  — O 


r<-,  <v}  ^ 

: II  II  II  -3^ 

c s:  s:  1:2 


tawny  yellow  in  A.  f jiammeus).  The  sides  of 
the  head  just  behind  the  lateral  edges  of  the 
facial  disk  have  an  unmarked  brown  patch  not 
found  in  A.  f.  flammeus.  The  dark  bars  on  the 
rectrices  also  tend  to  be  more  complete  and 
have  straighter  margins  in  the  Antillean  birds. 

The  feathering  of  the  feet  also  differs  be- 
tween Holarctic  and  Antillean  birds.  In  Asia  f. 
flammeus  the  tarsi  and  the  upper  surfaces  and 
sides  of  the  toes  are  densely  covered  by  fine 
tawny  feathers.  On  Antillean  birds,  the  feath- 
ering of  the  toes  is  restricted  to  the  dorsal  sur- 
faces and  terminates  3-6  mm  from  the  talons. 
The  featherless  skin  of  the  under  surfaces  ex- 
tends up  the  sides  of  the  toes  to  near  their 
tops.  Feathering  on  the  toes  is  also  more 
sparse,  with  exposed  skin  showing  between 
individual  feathers  on  museum  specimens. 
The  feathers  on  the  tarsi  and  toes  are  shorter 
in  the  Antillean  birds  (ca  6 mm  versus  16  mm 
on  the  posterior  surface  midway  along  the  tar- 
sus and  ca  3 mm  versus  6 mm  long  on  the 
proximal  phalanx  of  the  middle  toe). 

Photographs  of  the  Antillean  owls  appear  to 
show  more  prominent  “ear”  tufts  than  do  A. 
f.  flammeus,  but  an  attempt  to  measure  the 
tufts  on  the  specimens  at  ABS  failed  to  show 
differences  (tuft  length  19-29  mm  for  10  An- 
tillean owls;  20—27  mm  for  3 flammeus).  The 
surrounding  head  feathers  appeared  shorter  on 
the  Antillean  birds  so  the  tufts  may  in  fact 
protrude  farther,  and  possibly  the  Antillean 
owls  are  more  likely  to  erect  the  tufts  when 
confronted  by  photographers. 

The  Antillean  birds  appear  to  weigh  sub- 
stantially less  than  northern  birds.  Three 
seemingly  non-emaciated  males  of  our  Florida 
specimens  weighed  260  g,  280  g,  and  299  g, 
and  the  two  non-emaciated  females  274  g and 
288  g.  In  comparison,  mean  weights  of  A.  f 
flammeus  are  315  g for  males,  and  378  g and 
411  g for  two  samples  of  females  (Holt  and 
Leasure  1993). 

Comparison  of  southern  Florida  specimens 
with  a series  from  the  Greater  Antilles. — We 
sent  the  first  two  specimens  we  obtained 
(GEW  5861  and  GEW  5862)  to  Joe  T.  Mar- 
shall at  the  National  Museum  of  Natural  His- 
tory (NMNH)  who  compared  them  with  a se- 
ries of  eight  specimens  from  the  Greater  An- 
tilles (six  from  Puerto  Rico  and  two  from 
Cuba).  He  informed  us  (pers.  comm.)  that  the 
Florida  specimens  were  similar  to  these  An- 


308 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


LIG.  I.  Specimens  of  Asia  f.  domin^ensis  and  Asia  f.  fiammeu.s.  A.  Dorsal  view.  Lett  row  from  top;  Asia, 
f.  dominf’ensis-,  1 unsexed  adult  from  Cuba,  1 female  from  Llorida.  and  4 males  from  Llorida.  Right  row  from 
top:  Asia.  f.  flcimmeus',  3 females  from  mainland  southern  Llorida,  1 female  from  Grand  Turk.  B.  Ventral  view. 
Left  row  from  top:  Asio.  f.  domin^en.sis',  1 unsexed  adult  from  Cuba,  1 female  from  Llorida,  and  4 males  fiom 
Llorida.  Right  row  from  top:  A.sio.  f flammeu.f,  3 females  from  mainland  .southern  Llorida,  1 female  from  Grand 
Turk.  Photograph  by  Reed  Bowman. 


Hoffmim  el  al.  • SHORT-EARED  OWLS  IN  SOUTHERN  FH.ORIDA 


309 


tillean  birds  in  wing  length  and  back  color  and 
pattern,  but  had  paler  underparts  and  smaller 
bills.  He  did  not  assign  them  unequivocally  to 
any  named  population. 

Comparison  of  southern  Florida  specimens 
with  Cuban  specimens. — Five  of  the  eight 
Florida  owls  not  referable  to  A.  f flammeus 
were  compared  with  the  unsexed  Cuban  spec- 
imen (MNHN  1595),  and  all  eight  were  com- 
paied  to  the  second  Cuban  specimen  (GEW 
5925).  The  first  four  males  from  Rorida  had 
somewhat  paler  underparts  than  the  first  Cuban 
specimen  (MNHN  1595),  but  the  Florida  fe- 
male (GEW  5889)  had  shghtly  darker  under- 
parts. As  is  true  for  A.  f flammeus  (Holt  and 
Leasure  1993),  females  of  the  Antillean  form 
may  have  darker  underparts  than  males.  The 
second  Cuban  specimen  (GEW  5925)  has 
some  darker  markings  dorsally  than  most  of  the 
Rorida  specimens  but  otherwise  appeared  sim- 
ilar. The  Rorida  specimens  also  agreed  in  de- 
tail with  the  Cuban  specimens  in  the  foot-feath- 
ering characteristics.  We  conclude  that  these 
eight  specimens  from  southern  Rorida  are  va- 
grants from  the  Greater  Antilles,  and  probably 
originate  from  Cuba.  We  suspect  the  slightly 
paler  underparts  and  smaller  bill  size  noted  by 
Marshall  are  characteristics  of  age  or  sex. 

Analysis  of  photographs. — The  photo- 
graphs of  six  owls  from  Florida  agree  in 
plumage  with  the  eight  specimens  described 
above,  so  we  consider  them  Antillean  (Table 
1).  The  first  of  these  was  discovered  by  Hoff- 
man on  Bush  Key,  Dry  Tortugas,  on  21  June 
1978  and  photographed  in  the  hand  by  Bar- 
bara Kittleson  (Hoffman  et  al.  1979).  When 
the  1990  specimen  was  recognized  as  resem- 
bling Antillean  representatives  of  the  Asia 
flammeus  species-group,  we  re-examined  the 
photographs  and  found  them  to  show  the  dark 
back  and  finely  streaked  buffy  underparts  of 
Antillean  birds.  The  second  owl  was  photo- 
graphed by  Howard  P.  Langridge  in  April 
1985  (Kale  1985).  It  shows  extensive  black 
surrounding  the  eyes,  and  very  fine  streaking 
on  the  breast,  indicating  it  is  likely  a juvenile 
and  of  Antillean  origin.  The  third  owl  was 
photographed  by  Paul  Cavanagh  on  Bottle 
Key,  in  northeast  Florida  Bay,  on  24  August 
1987.  His  photographs  show  the  characteristic 
unstreaked  back  and  lightly  streaked  belly  of 
Antillean  birds.  The  fourth  owl,  photographed 
by  Hoffman  at  Ft.  Zachary  Taylor,  Key  West 


on  4 April  1994,  also  shows  the  unstreaked 
back,  lightly  streaked  belly,  and  the  sparsely 
feathered  feet  of  Antillean  birds,  as  well  as  the 
black  feathering  around  the  eyes  seen  on  the 
five  presumed  juvenile  male  specimens.  The 
fifth  owl  is  the  first  of  seeming  Antillean  or- 
igin found  north  of  southernmost  Florida.  The 
bird  was  rescued  from  the  surface  of  the  Gulf 
of  Mexico  about  110  km  west  of  Hernando 
Co  (northwest  of  St.  Petersburg)  in  early  June 
1994.  It  was  photographed,  rehabilitated,  and 
eventually  released  by  the  Birds  of  Prey  Cen- 
ter of  the  Florida  Audubon  Society.  The  sixth 
owl  was  photographed  29  April  1997  on  Gar- 
den Key,  Dry  Tortugas,  by  Darlene  Friedman 
and  the  photograph  was  forwarded  to  us  by 
Paul  Lehman;  it  also  appears  to  be  Antillean. 
We  suspect  that  most  of  the  other  recent  re- 
ports from  the  Florida  Keys  and  Dry  Tortugas 
also  are  of  Antillean  birds.  Descriptions  from 
the  observers  generally  support  this  conten- 
tion. Table  1 includes  one  A.  f.  flammeus  spec- 
imen (collected  10  December  1992)  and  two 
sightings  that  may  belong  to  this  race  (24  No- 
vember 1979,  winter  1983-84). 

DISCUSSION 

We  document  the  occurrence  in  southern 
Florida  of  representatives  of  a population  of 
the  Short-eared  Owl  previously  unrecorded 
from  the  North  American  continent.  The  doc- 
umentation includes  eight  specimens  and  nu- 
merous photographs  of  several  birds.  The  only 
taxon  of  this  species  previously  known  to  oc- 
cur on  the  continent  is  the  Holarctic  nominate 
race  Asia  flammeus  flammeus. 

Based  on  the  characteristics  of  the  speci- 
mens now  available,  we  are  confident  that 
these  owls  are  from  the  Greater  Antilles.  On 
geographical  and  historical  grounds  they  most 
likely  come  from  Cuba.  Seven  of  our  speci- 
mens, and  perhaps  most  of  the  other  southern 
Florida  birds,  appear  to  be  juveniles  and  to 
have  arrived  during  post-fledging  dispersal. 
Breeding  by  Short-eared  Owls  is  known  to  oc- 
cur in  Hispaniola  and  Cuba  during  northern 
hemisphere  winter.  Albert  Schwartz  collected 
nestlings  in  the  Dominican  Republic  in  No- 
vember (LSUMZ  142354  and  142355),  and 
Garrido  (1984)  reported  a nest  with  eggs  in 
Cuba  on  8 December.  Based  on  seasonality  of 
reports,  this  pattern  of  dispersal  to  southern 


310 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


Florida  appears  to  be  quite  recent,  beginning 
in  the  late  1970s. 

Nomenclature  of  Antillean  Short-eared 
Owls. — Assigning  a scientific  name  to  the 
owls  invading  southern  Florida  is  a problem 
because  the  nomenclature  of  the  Antillean 
Short-eared  Owl  is  confused  and  poorly  doc- 
umented. Briefly  the  history  is  as  follows. 
Muller  (1776,  reviewed  in  Wetmore  1928)  de- 
scribed an  owl  from  Hispaniola,  based  on 
Buffon  (1770),  as  Strix  domingensis.  Subse- 
quent authors  ignored  this  taxon,  or  assumed 
it  was  based  on  the  Burrowing  Owl  {Athene 
cunicularia)  until  Wetmore  (1928)  examined 
one  short-eared  owl  each  from  the  Dominican 
Republic  and  Haiti,  and  concluded  these  were 
examples  of  the  subject  of  Buffon’s  illustra- 
tion and  account.  However,  prior  to  Wetmore’s 
work,  Ridgway  (1882)  described  similar  owls 
from  Puerto  Rico  as  Asia  portoricensis.  Wet- 
more (1928)  compared  his  two  specimens 
from  Hispaniola  to  five  from  Puerto  Rico,  and 
concluded  that  those  from  Hispaniola  were 
only  subspecifically  distinguishable  from  the 
Puerto  Rican  specimens.  He  used  the  names 
Asia  domingensis  domingensis  and  A.  d.  por- 
toricensis, respectively.  We  find  no  publica- 
tion that  proposes  and  explains  the  merger  of 
these  taxa  into  Asio  flammeus.  Although  Wet- 
more and  Lincoln  (1933)  treated  domingensis 
as  a distinct  species,  only  three  years  later 
Bond  (1936),  without  explanation,  listed  these 
Antillean  owls  as  subspecies  of  A.  flammeus. 
This  treatment  seems  to  have  been  followed 
by  most  subsequent  authors  including  Peters 
(1940).  The  situation  is  further  confused  be- 
cause some  recent  authors  have  grouped  the 
Hispaniolan  {domingensis)  and  Puerto  Rican 
{portoricensis)  populations  together,  also 
without  comment.  Voous  (1988)  and  Holt  and 
Leasure  (1993),  for  example,  referred  to  all 
Antillean  Short-eared  Owls  as  the  race  por- 
toricensis of  A.  flammeus  despite  the  fact  that 
the  name  domingensis  seemingly  has  priority 
over  portoricensis  (Wetmore  1928).  The  ex- 
istence of  a Cuban  breeding  population  has 
been  recognized  only  since  1981,  and  no  for- 
mal determinations  of  its  taxonomic  status 
have  been  published.  Pending  further  study, 
we  recommend  using  the  single  epithet  dom- 
ingensis for  all  the  Antillean  populations; 
Wetmore’s  (1928)  justification  for  maintaining 


portoricensis  separate  from  domingensis 
seems  insufficient  given  his  sample  sizes. 

We  feel  that  the  systematics  and  nomencla- 
ture of  the  Short-eared  Owl  are  in  need  of 
revision.  The  validity  of  several  races  is  in- 
adequately established,  in  South  America  as 
well  as  in  the  West  Indies.  We  also  suspect 
that  Asio  flammeus  may  deserve  splitting  into 
two  or  more  species.  Ideally,  such  a revision 
would  include  detailed  analyses  of  vocaliza- 
tions, as  well  as  studies  of  molecular  genetic 
differences.  The  West  Indian  birds  are  among 
the  most  distinctive  in  plumage  and  structure, 
but  the  Galapagos  race  and  the  South  Amer- 
ican populations  could  plausibly  deserve  spe- 
cies status  as  well. 

Biogeographic  considerations. — The  recent 
occurrence  and  increasing  frequency  of  these 
owls  in  Florida  raises  the  possibility  of  a new, 
northward  colonization  from  the  Antilles. 
Owls  that  reach  mainland  Florida  may  find 
habitat  suitable  for  nesting.  Nesting  habitat  in 
Cuba  apparently  includes  pasturelands,  rice 
fields  (Garrido  1984),  and  sugar  cane  planta- 
tions (Garrido,  pers.  comm.).  Habitats  similar 
to  all  of  these  occur  extensively  in  southern 
Florida. 

If  these  Antillean  owls  colonize  southern 
Florida,  they  will  be  part  of  an  ongoing  wave 
of  colonizations  from  the  West  Indies.  When 
Howell’s  Florida  Bird  Life  was  published 
(Howell  1932),  the  breeding  landbird  fauna  of 
southern  Florida  contained  only  five  species’ 
populations  clearly  of  West  Indian  origin: 
White-crowned  Pigeon  {Columba  leucocepha- 
la),  (Cuban)  Mourning  Dove  {Zenaida  m.  rna- 
croura).  Mangrove  Cuckoo  {Coccyzus  minor). 
Gray  Kingbird  {Tyrannus  dominicensis),  and 
Black-whiskered  Vireo  {Vireo  altiloquus). 
Robertson  and  Kushlan  (1984),  in  their  in- 
sightful analysis  of  the  southern  Florida  avi- 
fauna, considered  all  these  to  be  quite  recent 
immigrants,  in  part  because  none  showed  geo- 
graphic variation  in  Florida.  A sixth  and  sev- 
enth species,  Zenaida  Dove  {Zenaida  aurita) 
and  Key  West  Quail-Dove  {Geotrygon  chry- 
sia),  were  reported  breeding  in  the  Florida 
Keys  prior  to  1850,  but  both  now  occur  only 
as  vagrants. 

Since  1932  southern  Florida  has  experi- 
enced an  average  of  about  one  natural  land- 
bird  invasion  per  decade  from  the  West  Indies. 
These  recent  immigrants  are  Smooth-billed 


Hoffman  el  at.  • SHORT-EARED  OWLS  IN  SOUTHERN  ET.ORIDA 


Ani  {Crotophaga  ani;  Sprunt,  A.  Jr.  1939, 
1954),  Cuban  Yellow  Warbler  {Deudroica  pe- 
techia gundlachi;  Greene  1942),  Antillean 
Nighthawk  {Chordeiles  gundlachii;  Greene 
1943),  Fulvous  Whistling-Duck  {Dendrocyg- 
na  bicolor;  reviewed  by  Palmer  1976,  Turn- 
bull  et  al.  1989),  Cave  Swallow  [Hirundo  fid- 
va  cavicola  (=  H.  f.  Jidva?);  Smith  et  al. 
1988j,  and  Shiny  Cowbird  {Molothrus  bon- 
airiensis;  Smith  and  Sprunt  1987). 

Most  of  these  invading  species  had  been 
known  as  vagrants  to  southern  Florida  for 
some  time,  and  reports  became  increasingly 
frequent  before  breeding  in  Florida  was  doc- 
umented. This  fits  a general,  but  often  over- 
looked rule:  range  expansion  tends  to  be  driv- 
en by  population  dynamics  (often  population 
increases)  in  the  source  areas,  rather  than  by 
habitat  changes  or  initial  reproductive  success 
in  the  colonized  areas. 

Several  West  Indian  birds,  in  addition  to 
these  owls,  seem  poised  to  invade  Florida 
from  the  south  or  east  (Bahamas).  La  Sagra’s 
Flycatcher  (Myiarchus  sagrae;  Smith  and  Ev- 
ered  1992),  Bahama  Mockingbird  (Mimus 
gundlachii),  and  Thick-billed  Vireo  (Vireo 
crassirostris;  Smith  et  al.  1990)  seem  to  be 
increasing  in  frequency  as  vagrants.  Pearly- 
eyed  Thrasher  (Margarops  fuscatus)  has  not 
been  documented  in  North  America  as  of  this 
writing,  but  it  has  been  extending  its  range 
northward  in  the  Bahamas  in  recent  years  and 
could  begin  appearing  in  Florida  in  the  near 
future.  The  Cuban  subspecies  of  American 
Kestrel  (Falco  sparverius  sparveroides)  also 
has  been  expanding  its  range  in  the  Bahamas, 
and  recently  was  photographed  by  WH  at  Key 
West.  Robertson  and  Kushlan  (1984:226) 
speculated  on  potential  immigrant  West  Indian 
species,  naming  “the  Masked  Duck,  a hum- 
mingbird, Bahama  [Tachycineta  cyaneoviri- 
dis),  and  Cave  swallows,  Bananaquit  [Coereba 
flaveola].  Stripe-headed  Tanager  [Spindalis 
zena]  and  Black-faced  Grassquit  [Tiaris  bi- 
color]”  most  likely.  Of  these,  the  Cave  Swal- 
low and  possibly  the  Masked  Duck  (Bowman 
1995)  already  have  colonized. 

The  rate  of  immigration  in  recent  decades, 
then,  must  be  much  higher  than  the  overall 
post-Pleistocene  rate,  unless  prehistoric  ex- 
tinction rates  for  immigrant  populations  were 
extremely  high.  The  Zenaida  Dove  and  Key 
West  Quail  Dove,  once  reported  to  breed  in 


31 1 

the  Florida  Keys,  no  longer  do  so.  These  pos- 
sible extirpations  of  breeding  populations 
most  likely  resulted  from  hunting  and  habitat 
destruction  (Robertson  1978a,  1978b;  Robert- 
son and  Woolfenden  1992). 

The  apparent  increase  in  immigration  rates 
from  the  West  Indies  to  Florida  may  have  re- 
sulted from  anthropogenic  changes  in  the  en- 
vironment, both  in  the  West  Indies  and  in 
Florida.  Flabitat  changes,  associated  with  for- 
est clearing  for  grazing,  cultivation,  and  urban 
development,  and  global  climate  changes  are 
two  nonexclusive  anthropogenic  changes  to 
the  regional  environments  that  could  drive 
these  colonizations.  All  the  known  Cave 
Swallow  colonies  in  Florida,  for  example,  are 
located  on  concrete  bridges  and  overpasses 
along  highways  (Smith  et  al.  1988),  so  the  in- 
crease in  numbers  of  vagrants  prior  to  colony 
establishment  (Robertson  and  Woolfenden 
1992)  must  have  reflected  population  or  hab- 
itat changes  in  Cuba. 

Explanations  for  northward  colonization  in- 
volving global  climate  change  potentially  can 
explain  both  the  recent  wave  of  colonizations 
and  the  dearth  of  West  Indian  birds  in  the 
southern  Florida  avifauna  prior  to  this  centu- 
ry. A global  warming  trend  has  been  under- 
way since  the  end  of  the  “Little  Ice  Age”  in 
about  1870  (Pielou  1991),  a trend  that  has  ac- 
celerated because  of  increases  in  atmospheric 
carbon  dioxide,  methane,  and  other  green- 
house gasses  (Maul  1989).  Some  West  Indian 
birds  may  have  colonized  the  Florida  penin- 
sula previously  during  the  Holocene,  but 
stopped  during  the  “Little  Ice  Age”.  Al- 
though records  are  inadequate  or  nonexistent, 
the  lower  average  temperatures  of  the  “Little 
Ice  Age”  likely  were  manifested  in  southern 
Florida  by  more  frequent  and  more  severe  ep- 
isodes of  cold  winter  weather  rather  than  by 
cooler  summer  weather.  These  hypothetical 
cold  episodes  could  have  been  particularly 
damaging  to  West  Indian  bird  populations  at- 
tempting to  persist  in  southern  Florida.  The 
current  warming  trend  also  may  be  fueling 
population  increases  in  the  West  Indies  for  the 
species  that  have  recently  colonized,  or  are 
appearing  more  frequently  as  vagrants.  This 
warming  trend  also  may  be  making  southern 
Florida  habitats  subtly  more  suitable  for  these 
birds  (Robertson  and  Kushlan  1984). 


312 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


ACKNOWLEDGMENTS 

Much  of  the  preparation  of  this  paper  took  place  at 
Archbold  Biological  Station.  We  thank  the  staff  for 
providing  us  the  facilities  and  opportunity  to  work 
there.  L.  Quinn  of  the  Llorida  Keys  Wild  Bird  Center 
in  Key  Largo,  B.  Arnold  of  Wildlife  Rescue  of  the 
Llorida  Keys  in  Key  West,  and  K.  Grinter  of  Marathon 
Wild  Bird  Rescue  have  saved  numerous  bird  speci- 
mens for  us,  including  all  eight  of  the  Antillean  owls. 
R.  Collins  provided  information  on  the  owl  rehabili- 
tated by  Llorida  Audubon’s  Birds  of  Prey  Center.  S. 
Cardiff  and  W.  Meshaka  assisted  us  in  borrowing  spec- 
imens from  the  Museum  of  Natural  Science,  Louisiana 
State  University  and  Everglades  National  Park,  respec- 
tively. O.  Garrido  graciously  transported  the  Cuban 
specimens  to  us,  and  J.  Marshall  examined  certain  of 
our  recent  specimens.  H.  Gonzalez  Alonso,  Director  of 
the  Museo  Nacional  de  Historia  Natural,  Havana,  do- 
nated to  Archbold  Biological  Station  specimen  GEW 
5925  through  the  request  of  O.  Garrido.  P.  Lehman  sent 
to  us  the  photograph  documenting  the  1997  record  at 
the  Dry  Tortugas.  D.  Steadman,  R.  Browning,  and  R. 
Banks  provided  certain  references.  J.  Litzpatrick,  D. 
Holt,  J.  Marshall,  D.  Steadman,  and  K.  Voous  re- 
viewed various  drafts  of  the  manuscript  and  made  nu- 
merous helpful  comments.  We  thank  all  of  these  per- 
sons for  their  assistance. 

LITERATURE  CITED 

American  Ornithologists’  Union.  1957.  Check-list 
of  North  American  birds,  hfth  ed.  American  Or- 
nithologists’ Union,  Washington,  D.C. 

American  Ornithologists’  Union.  1998.  Check-list 
of  North  American  birds,  seventh  ed.  American 
Ornithologists’  Union,  Washington,  D.C. 

Bond,  J.  1936.  Birds  of  the  West  Indies.  The  Academy 
of  Natural  Sciences  of  Philadelphia,  Philadelphia, 
Pennsylvania. 

Bowman,  M.  1995.  Sighting  of  Masked  Duck  duck- 
lings in  Llorida.  Lla.  Lield  Nat.  23:35. 

Buffon,  G.  L.  L.  1770.  Histoire  Naturelle  des  Oiseaux, 
vol.  1.  Paris,  Prance. 

Cramp,  S.,  K.  E.  L.  Simmons,  I.  J.  Perguson-Lees,  R. 
Gillmor,  P.  a.  D.  Hollom,  R.  Hudson,  E.  M. 
Nicholson,  M.  A.  Ogilvie,  P.  J.  S.  Olney,  K.  H. 
Voous,  AND  J.  Wattel.  1977.  Handbook  of  the 
birds  of  Europe,  the  Middle  East,  and  North  Af- 
rica: the  birds  of  the  Western  Palearctic.  Vol.  1. 
Oxford  University  Press,  Oxford,  U.K. 

Cramp,  S.,  D.  J.  Brooks,  E.  Dunn,  R.  Gillmor,  P.  A. 
D.  Hollom,  R.  Hudson,  E.  M.  Nicholson,  M.  A. 
Ogilvie,  P.  J.  S.  Olney,  C.  S.  Roselaar,  K.  E.  L. 
Simmons,  K.  H.  Voous,  D.  1.  M.  Wallace,  J. 
Wattel,  and  M.  G.  Wilson.  1985.  Handbook  of 
the  birds  of  Europe,  the  Middle  East,  and  North 
Africa:  the  birds  of  the  western  Palearctic.  Vol.  4. 
Oxford  University  Press,  Oxford,  U.K. 

Garrido,  O.  H.  1984.  A.sio  fiammeu.s  (Aves:  Strigidae) 
nesting  in  Cuba.  Caribb.  J.  Sci.  20:67—68. 


Greene,  E.  R.  1942.  Golden  Warbler  nesting  in  lower 
Florida  Keys.  Auk  59:114. 

Greene,  E.  R.  1943.  Cuban  Nighthawk  nesting  in  low- 
er Florida  Keys.  Auk  60:105. 

Hoffman,  W.,  W.  B.  Robertson,  Jr.,  and  P.  Patty. 
1979.  Short-eared  Owl  on  Bush  Key,  Dry  Tortu- 
gas, Florida.  Fla.  Field  Nat.  7:29—30. 

Holt,  D.  W.  and  M.  S.  Leasure.  1993.  Short-eared 
Owl  {Asia  flammeus).  In  The  birds  of  North 
America,  no.  62  (A.  Poole  and  F.  Gill,  Eds.).  The 
Academy  of  Natural  Sciences,  Philadelphia,  Penn- 
sylvania; The  American  Ornithologists’  Union, 
Washington,  D.C. 

Howell,  A.  H.  1932.  Florida  bird  life.  Coward-Mc- 
Cann,  Inc.  New  York. 

Kale,  H.  W.,  II.  1985.  Florida  region.  Am.  Birds  39: 
288-291. 

Loftin,  R.  W.,  G.  E.  Woolfenden,  and  J.  A Wool- 
fenden.  1991.  Florida  bird  records  in  American 
Birds  and  Audubon  Field  Notes  (1947—1989): 
species  index  and  county  gazetter.  Fla.  Ornithol. 
Soc.  Sp.  Pub.  4:1-99. 

Maul,  G.  1989.  Implications  of  climatic  changes  in 
the  wider  Caribbean  region.  CEP  Tech.  Rept.  3: 
1-23. 

Palmer,  R.  S.  (Ed.).  1962.  Handbook  of  North  Amer- 
ican birds.  Vol.  1.  Yale  Univ.  Press,  New  Haven, 
Connecticut. 

Palmer,  R.  S.  (Ed.).  1976.  Handbook  of  North  Amer- 
ican birds.  Vol.  2.  Yale  Univ.  Press,  New  Haven, 
Connecticut. 

Peters,  J.  L.  1940.  Check-list  of  birds  of  the  world. 
Vol.  4.  Harvard  Univ.  Press,  Cambridge,  Massa- 
chusetts. 

PiELOU,  E.  1991.  After  the  Ice  Age:  the  return  of  life 
to  glaciated  North  America.  Univ.  of  Chicago 
Press,  Chicago,  Illinois. 

Ridgway,  R.  1882.  Description  of  a new  owl  from 
Porto  Rico.  Proc.  U.S.  Natl.  Mus.  4:366-371. 

Ridgway,  R.  1914.  Birds  of  North  and  Middle  Amer- 
ica. Bull.  U.S.  Natl.  Mus.  50(6):  1-882. 

Robertson,  W.  B.,  Jr.  1978a.  Key  West  Quail-Dove. 
P.  119  in  Rare  and  endangered  biota  of  Florida. 
Vol.  2.  Birds  (H.  W.  Kale,  Ed.).  Univ.  Presses  of 
Florida,  Gainesville. 

Robertson,  W.  B.,  Jr.  1978b.  Zenaida  Dove.  P.  120  in 
Rare  and  endangered  biota  of  Florida.  Vol.  2. 
Birds  (H.  W.  Kale,  Ed.).  Univ.  Presses  of  Florida, 
Gainesville. 

Robertson,  W.  B.,  Jr.  and  J.  A.  Kushlan.  1984.  The 
southern  Florida  avifauna.  Pp.  219-257  in  Envi- 
ronments of  southern  Florida:  past  and  present  II 
(P.  J.  Gleason,  Ed.).  Miami  Geological  Society, 
Coral  Gable,  Florida. 

Robertson,  W.  B.,  Jr.  and  G.  E.  Woolfenden.  1992. 
Florida  bird  species:  an  annotated  list.  Fla.  Orni- 
thol. Soc.  Sp.  Publ.  6:1-260. 

Schwartz,  A.  and  R.  F.  Klinikowski.  1965.  Addition- 
al observations  on  West  Indian  birds.  Not.  Nat. 
376:1-16. 


Hoffman  et  al.  • SHORT-EARED  OWLS  IN  SOUTHERN  FLORIDA 


313 


Smith,  R W.  and  D.  S.  Evered.  1992.  Photo  note — La 
Sagra’s  Flycatcher.  Birding  24:294-297. 

Smith,  P.  W.,  D.  S.  Evered,  L.  R.  Messick,  and  M.  C. 
Wheeler.  1990.  Eirst  veritiable  record  of  the 
Thick-billed  Vireo  from  the  United  States.  Am. 
Birds  44:372-376. 

Smith,  P.  W.,  W.  B.  Robertson,  Jr.,  and  H.  M.  Ste- 
venson. 1988.  West  Indian  Cave  Swallows  nest- 
ing in  Florida,  with  comments  on  the  taxonomy 
of  Hirundo  fulva.  Fla.  Field  Nat.  16:86—90. 
Smith,  P.  W.  and  A.  S.  Sprunt,  IV.  1987.  The  Shiny 
Cowbird  reaches  the  United  States.  Am.  Birds  41: 
370-371. 

Sprunt,  A.  S.  Jr.  1939.  Smooth-billed  Ani  nesting  in 
Florida.  Auk  56:335. 

Sprunt,  A.  S.  Jr.  1954.  Florida  bird  life.  Coward- 
McCann,  Inc.,  New  York. 


Turnbull,  R.  E.,  E A.  John.son,  and  D.  A.  Brokhage. 
1989.  Status,  distribution,  and  foods  of  Eulvous 
Whistling-Ducks  in  southern  Florida.  J.  Wildl. 
Manage.  53:1046-1051. 

Voous,  K.  H.  1988.  Owls  of  the  northern  hemisphere. 

The  MIT  Press,  Cambridge,  Massachusetts. 
Wetmore,  a.  1928.  The  Short-eared  Owls  of  Porto 
Rico  and  Hispaniola.  Proc.  Biol.  Soc.  Wash.  41: 
165-166. 

Wetmore,  A.  and  F.  C.  Lincoln.  1933.  Additional 
notes  on  the  birds  of  Haiti  and  the  Dominican  Re- 
public. Proc.  U.S.  Natl.  Mus.  82:1-68. 

Wetmore,  A.  and  B.  H.  Swales.  1931.  The  birds  of 
Haiti  and  the  Dominican  Republic.  U.S.  Natl. 
Mus.  Bull.  155:1-483. 


Wilson  Bull.,  111(3),  1999,  pp.  314-320 


WITHIN-  AND  BETWEEN- YEAR  DISPERSAL  OF  AMERICAN 
AVOCETS  AMONG  MULTIPLE  WESTERN  GREAT 

BASIN  WETLANDS 

JONATHAN  H.  PLISSNER,'  SUSAN  M.  HAIG,'  ^ AND  LEWIS  W.  ORING^ 


ABSTRACT. — Connectivity  of  discrete  habitat  patches  may  be  described  in  terms  of  the  movements  of 
individual  organisms  among  such  patches.  To  examine  connectivity  of  widely  dispersed  alkali  lake  systems,  we 
recorded  post-breeding  and  subsequent  breeding  locations  of  color-banded  American  Avocets  (Recun’irosira 
americana)  in  the  western  U.S.  Great  Basin,  from  1995—1997.  Among  individuals  observed  during  the  post- 
breeding/premigratory  season,  over  half  of  the  188  breeding  adults  were  observed  at  lakes  other  than  their 
breeding  locations,  whereas  70%  of  125  post-fledged  young  were  observed  only  at  their  natal  lake  systems.  Of 
46  breeding  adults  observed  in  consecutive  years,  only  eight  (17%)  dispersed  between  different  lake  systems. 
Only  8%  of  chicks  were  observed  after  their  first  year,  and  only  1.3%  returned  to  the  natal  area  in  subsequent 
breeding  sea.sons.  Adult  and  recently  fledged  birds  from  the  southernmost  breeding  site  were  regularly  observed 
in  post-breeding  aggregations  at  lakes  several  hundred  kilometers  to  the  north,  suggesting  seasonal  differences 
in  habitat  quality  at  the  lake  systems  studied.  These  results  indicate  the  importance  of  maintaining  habitat  for 
post-breeding  movements.  Received  10  Dec.  1998,  accepted  3 April  1999. 


Concurrent  with  the  recognition  that  habitat 
fragmentation  is  a key  threat  to  regional  bio- 
diversity, the  role  of  dispersal  in  maintaining 
connectivity  between  populations  and  subpop- 
ulations has  become  a major  focus  in  assess- 
ing extinction  risks  and  other  dynamics  of 
populations  and  communities  (Saunders  et  al. 
1991,  McPeek  and  Holt  1992,  Taylor  et  al. 
1993,  Dunning  et  al.  1995,  With  et  al.  1997, 
Haig  et  al.  1998).  For  vertebrate  populations, 
dispersal  studies  generally  have  focused  upon 
movements  within  contiguous  habitats  or 
among  adjacent  patches.  Although  appropriate 
for  some  species,  this  approach  fails  to  ac- 
count for  a significant  proportion  of  dispersal 
events  among  more  mobile  species,  especially 
those  that  inhabit  discrete,  widely-dispersed 
patches  of  habitat.  In  addition,  connectivity 
measures  require  monitoring  of  multiple  sites 
as  both  potential  sources  and  recipients  of  dis- 
persers. Such  interchange  among  patches  has 
been  incorporated  to  varying  degrees  into  nu- 
merous spatial  models  of  populations  (see 
Doebeli  and  Ruxton  1997,  Ims  and  Yoccoz 
1997,  Wiens  1997),  but  empirical  measures  of 
such  rates  of  exchange  are  often  difficult  to 
determine  for  vertebrate  metapopulations. 


' U.S.  Geological  Survey,  Forest  and  Rangeland 
Ecosystem  Science  Center,  3200  SW  Jefferson  Way, 
Corvallis,  OR  97331. 

^ Dept,  of  Environmental  and  Resource  Sciences, 
Univ.  of  Nevada,  1000  Valley  Rd.,  Reno,  NV  89512. 

’ Corresponding  author;  E-mail:  haig.s@fsl.orst.edu 


Specifically,  studies  of  avian  dispersal  are 
largely  limited  by  the  spatial  scale  that  re- 
searchers are  able  to  effectively  monitor  dis- 
persing individuals  and  by  a traditional  focus 
on  return  rates  rather  than  on  broader  dispersal 
patterns.  Typically,  such  studies  focus  on  a 
population  inhabiting  a single  site  or  a few 
neighboring  areas,  with  the  probabilities  of 
detection  for  dispersers  decreasing  geometri- 
cally with  distance  from  the  point  of  origin 
(Barrowclough  1978,  Cunningham  1986). 

To  determine  large-scale  connectivity  pat- 
terns among  discrete  wetlands  of  the  western 
Great  Basin,  we  examined  movements  of 
banded  American  Avocets  (Recurvirostra 
americanci)  among  major  lake  systems  of  an 
otherwise  arid  region.  In  addition  to  monitor- 
ing dispersal  in  relation  to  breeding  sites,  we 
also  examined  premigratory  movements  of  in- 
dividuals, an  often  neglected  aspect  of  indi- 
vidual life  histories  that  may  be  a critical  con- 
nective element  of  patchy  landscapes  (Haig  et 
al.  1998). 

METHODS 

From  1995  to  1997,  we  color-banded  and  observed 
American  Avocets  at  four  major  alkali  lake  systems  in 
high  desert  regions  of  the  western  Great  Basin:  Sum- 
mer Lake  and  Lake  Abert  in  Lake  County,  Oregon; 
Goose  Lake  in  Lake  County,  Oregon  and  Modoc 
County,  California;  and  Honey  Lake  in  Lassen  County, 
California  (Fig.  1).  Interlake  distances  range  from  45- 
315  km.  Avocet  breeding  locations  in  the  region,  other 
than  our  study  areas,  were  scarce  during  the  years  of 
the  study;  although  additional  local  breeding  popula- 


314 


Flissner  et  al.  • AVOCET  DISPERSAL 


315 


tions  occur  sporadically  in  response  to  suitable  water 
conditions  (Neel  and  Henry  1996;  Oring  and  Reed 
1996;  L.  W.  O.,  pers.  observ.)-  During  breeding  peri- 
ods, our  efforts  at  Summer  Lake  and  Honey  Lake  were 
focused  on  managed  wetlands  adjacent  to  the  main 
lake  bodies:  the  Summer  Lake  Wildlife  Area  (SLWA) 
and  Jay  Dow,  Sr.  Wetlands  (JDW),  respectively,  al- 
though we  also  conducted  regular  surveys  along  the 
entire  lake  shorelines. 

During  the  three  breeding  seasons,  339  incubating 
adults  were  captured  at  nests  and  given  unique  com- 
binations of  color  bands.  Recently  hatched  chicks  (n 
= 457)  and  flightless  fledglings  {n  = 19)  were  captured 
and  individually  marked  opportunistically  near  nesting 
areas.  In  1995,  an  additional  61  individuals  in  post- 
breeding flocks  were  captured  and  banded  at  Summer 
Lake.  All  birds  were  given  U.S.D.I.  numbered  alumi- 
num and  celluloid  or  Darvic®  plastic  leg  bands.  In 
1996  and  1997,  radio  transmitters  were  attached  to  alu- 
minum leg  bands  and  placed  on  185  adult  avocets, 
distributed  among  the  four  lake  systems  (Plissner  et  al. 
in  press).  Hatch-year  birds  were  given  brood-specific 
band  combinations,  with  individuals  identified  by  col- 
ored plastic  tape  wrapped  around  the  numbered  alu- 
minum band.  Sexes  of  adults  were  distinguished  by 
relative  bill  curvature  (Hamilton  1975)  whenever  pos- 
sible but  were  deferred  to  judgments  during  banding 
in  cases  of  conflict  with  observations  of  resighted  in- 
dividuals. 

Banded  avocets  were  resighted  during  the  breeding 
and  post-breeding  periods  at  each  lake,  from  April 
through  September  of  each  year.  The  Summer  Lake 
WA  and  Jay  Dow,  Sr.  Wetlands  were  surveyed  weekly 
throughout  the  field  season.  Because  of  difficulties  in 
accessing  most  of  the  lakeshores,  the  four  main  lakes 
were  surveyed  less  frequently  by  foot  and  from  vehi- 
cles and  hovercraft.  In  1996  and  1997,  complete  sur- 
veys of  the  three  Oregon  lakes  were  conducted  weekly 
during  the  breeding  and  post-breeding  periods.  Aerial 
surveys  and  previous  monitoring  efforts  at  Honey 
Lake  indicated  relatively  little  use  of  the  main  lake 
body  by  avocets  throughout  the  breeding  and  post- 
breeding periods  (L.W.O.,  unpubl.  data).  Nevertheless, 
focal  lake  sites  in  proximity  to  Jay  Dow.  Sr.  Wetlands 


were  regularly  surveyed.  Observers  recorded  band 
combinations  observed  with  20-60X  telescopes. 
Breeding  status  was  assigned  to  birds  observed  at  nests 
or  in  the  presence  of  young  chicks.  As  a conservative 
measure,  only  birds  observed  after  1 August  were  as- 
sumed to  be  post-breeding.  We  use  the  term  “dispers- 
al” to  refer  to  movements  between  different  lake  sys- 
tems; whereas  we  refer  to  birds  returning  to  the  same 
lake  system  as  “philopatric”.  We  also  report  anecdotal 
observations  of  banded  avocets  from  wintering  areas 
in  California. 

Resightings  were  subsequently  screened  against  lists 
of  known  band  combinations.  Because  of  occasional 
band  loss  and  the  fact  that  some  pairs  of  band  colors 
became  difficult  to  distinguish  as  they  faded  over  time 
(Robinson  and  Oring  1997a),  alternatives  were  consid- 
ered for  those  observations  that  did  not  correspond  to 
known  combinations.  If  only  a single  alternative  com- 
bination existed,  the  sighting  was  retained  in  the  data- 
set. Other  observations  were  discarded.  Observations 
of  individuals  breeding  in  different  years  (bird-years) 
were  considered  independent  for  all  summaries  and  a- 
nalyses.  Frequency  data  were  analyzed  using  G-tests 
with  Williams  correction  (Excel  macro  based  on  Sokal 
and  Rohlf  1981). 

RESULTS 

Within  year  movements. — During  the  study 
period,  we  monitored  postfledging/postbreed- 
ing  movements  for  476  hatch-year  and  339 
breeding  adult  American  Avocets  (151  fe- 
males, 170  males,  18  unknown  gender)  band- 
ed in  the  western  Great  Basin.  One  bird,  orig- 
inally banded  as  a chick,  was  subsequently  re- 
captured and  rebanded  as  a nesting  adult. 
Eight  individuals  were  observed  at  nests  dur- 
ing two  breeding  seasons.  Of  the  total  moni- 
tored, 125  chicks  (26%)  and  188  adults  (55%) 
observed  during  nesting  periods  were  resight- 
ed during  the  subsequent  post-fledging/post- 
breeding  periods  (i.e.,  after  1 August).  Among 


316 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


TABLE  1.  Post-breeding  locations  of  adult  American  Avocets  in  the  western  Great  Basin,  1995-1997. 


Breeding  location 

Post-breeding  location^ 

Adults  monitored^ 

Honey  Lake 

Goose  Lake 

Lake  Abert 

Summer  Lake 

Honey  Lake 

105 

12 

2 

10 

9 

Goose  Lake 

57 

0 

17 

11 

12 

Lake  Abert 

50 

0 

5 

16 

12 

Summer  Lake 

133 

0 

4 

42 

62 

“ Includes  seven  individuals  observed  at  nests  during  two  breeding  seasons. 
Multiple  post-breeding  locations  included  for  30  individuals. 


adults  resighted  during  the  post-breeding  pe- 
riod, 53%  were  observed  at  lakes  other  than 
where  they  had  bred  (Table  1).  There  was  no 
significant  difference  in  proportions  of  males 
(53.5%,  n = 99)  and  females  (52.6%,  n = 76) 
that  dispersed  from  breeding  locations  to  other 
lake  systems  (G^^,  = 0.014,  P > 0.05).  Fur- 
thermore, of  observed  post-fledged  young, 
30%  (38  of  125)  were  resighted  away  from 
their  natal  lake  systems  (Table  2),  a significant 
difference  from  the  adult  rates  of  post-breed- 
ing dispersal  (G^,,  = 15.59,  P < 0.05).  Five 
chicks  and  16  adult  breeders  dispersed  follow- 
ing fledging  or  breeding  from  Honey  Lake  to 
post-breeding  locations  among  the  northern 
three  lakes,  while  no  birds  from  the  northern 
three  lakes  were  observed  postbreeding/post- 
fledging  at  Honey  Lake.  Four  chicks  and  three 
adults  were  reported  from  wintering  areas 
along  the  northern  California  coast  (Humboldt 
Bay  and  San  Francisco  Bay)  after  leaving 
Goose  Lake  and  Honey  Lake  natal/breeding 
areas.  Prior  to  being  reported  on  the  wintering 
grounds,  all  seven  birds  were  observed  on  at 
least  one  of  the  three  Oregon  lakes  during  the 
post-breeding  period. 

Between-year  movements. — We  monitored 
between-year  dispersal  locations  for  454 
chicks  and  259  adult  American  Avocets  (137 
females,  122  males).  Altogether,  197  birds 
were  observed  in  the  region  in  multiple  years, 
including  51%  of  all  banded  adults  and  8%  of 


chicks.  Of  46  adults  that  were  observed  in  the 
region  in  consecutive  breeding  seasons,  38 
(83%)  returned  to  the  same  lake.  Eight  addi- 
tional adults  that  bred  at  Honey  Lake  in  1995 
were  not  observed  the  following  year  but  re- 
turned to  breed  in  1997.  Overall,  return  rates 
for  chicks  were  0.4%  for  birds  observed  in 
their  first  post-natal  breeding  period  (i.e.,  at 
age  1)  and  1.3%  for  birds  observed  during 
their  first  two  post-natal  breeding  seasons  (i.e., 
ages  1 or  2).  Of  eleven  banded  chicks  resight- 
ed during  subsequent  breeding  seasons,  seven 
(64%)  returned  to  the  natal  lake  system.  The 
other  four  chicks,  all  banded  at  Honey  Lake, 
dispersed  to  Oregon  lakes  in  subsequent 
breeding  seasons.  No  adults  dispersed  from 
Honey  Lake  to  the  northern  lakes  between 
years,  and  no  birds  of  any  age  class  from 
Goose,  Abert,  or  Summer  Lakes  dispersed 
south  to  Honey  Lake  during  subsequent 
breeding  periods.  Four  birds  (three  males  and 
one  female)  were  observed  during  the  breed- 
ing period  of  all  three  study  years.  Three  of 
the  four  were  observed  at  Honey  Lake  during 
all  three  years.  The  other  individual  bred  at 
Summer  Lake  in  1995  and  1996  and  was  ob- 
served at  Lake  Abert  during  the  breeding  pe- 
riod in  1997.  Only  four  individuals  that  bred 
in  1996  and  1997  were  observed  during  the 
post-breeding  period  of  the  previous  year.  One 
of  the  four  nested  in  1997  at  Summer  Lake 
after  nesting  and  spending  the  post-breeding 


TABLE  2.  Post-Hedging  locations  of  American  Avocets  in  the  we.stern  Great  Basin,  1995-1997. 


Natal  location 

Po.st-fledging  location 

Chicks  monitored 

Honey  Lake 

Goose  Lake 

Lake  Abert 

Summer  Lake 

Honey  Lake 

201 

21 

0 

6 

2 

Goose  Lake 

169 

0 

45 

12 

14 

Lake  Abert 

10 

0 

0 

5 

1 

Summer  Lake 

97 

0 

1 

2 

16 

Flissiier  el  al.  • AVOCET  DISPERSAL 


317 


period  at  Lake  Abert  in  1996.  The  other  three 
bred  and  spent  the  post-breeding  period  at  the 
same  lake  in  both  years. 

DISCUSSION 

Our  results  demonstrate  within  and  between 
year  patterns  of  dispersal  at  a regional  scale. 
Earlier  studies  of  movement  patterns  of  avo- 
cets  and  other  recurvirostrids  focused  on  rates 
of  philopatry  and  dispersal  of  individuals 
within  a single  wetland  (Cadbury  and  Olney 
1978,  Watier  and  Fournier  1980,  Sordahl 
1984,  Cadbury  et  al.  1989,  James  1995,  Rob- 
inson and  Oring  1997b)  or  relied  upon  occa- 
sional surveys  and  anecdotal  reports  of 
marked  individuals  from  areas  away  from  the 
study  area  (Robinson  and  Oring  1996).  By 
regularly  monitoring  birds  at  multiple,  distant 
lake  systems,  we  were  better  able  to  define  the 
movement  patterns  and  extent  of  movements 
at  a scale  more  appropriate  to  the  life  history 
of  the  species.  Still,  we  recognize  the  fact  that, 
in  order  to  gain  an  objective  measure  of  dis- 
persal, monitoring  of  more  distant  sites  in 
multiple  directions  from  breeding  areas  would 
be  required. 

Return  rates  of  individuals  we  observed  dif- 
fered somewhat  from  those  described  for  av- 
ocets  in  an  earlier  Great  Basin  study  at  Honey 
Lake,  California.  Robinson  and  Oring  (1997b) 
estimated  that  21-25%  of  avocets  chicks  that 
were  known  to  have  survived  to  breeding  age 
returned  to  the  natal  site  to  breed,  whereas  we 
observed  64%  of  known  survivors  (i.e.,  indi- 
viduals observed  after  their  hatch-year)  re- 
turning to  the  natal  lake  system.  In  relation  to 
the  total  number  of  chicks  banded,  however, 
our  return  rates  were  lower  (4.2%,  Robinson 
and  Oring  1997b;  1.3%,  this  study),  suggest- 
ing that  survivorship  of  chicks  was  likely  the 
distinguishing  variable.  Return  rates  of  breed- 
ing adults  were  significantly  different  in  the 
two  studies  (24%,  Robinson  and  Oring  1997b; 
18%,  this  study;  = 4.24,  P < 0.05).  Rel- 
ative to  other  species  of  Charadriiformes  (Or- 
ing and  Lank  1984),  natal  and  breeding  phil- 
opatry rates  of  avocets  in  the  western  Great 
Basin  were  low,  perhaps  resulting  from  the 
extensive  annual  variability  in  breeding  habi- 
tat suitability. 

The  small  number  of  interlake  breeding  and 
natal  dispersal  events  during  the  three  years 
of  the  study  suggests  limited,  but  adequate 


(Wright  1951),  gene  fiow  among  Great  Basin 
breeding  populations.  Furthermore,  the  pattern 
of  both  post-breeding  and  between  year  dis- 
persal suggests  an  unbalanced  pattern  of  con- 
nectivity, perhaps  indicative  of  source-sink 
metapopulation  dynamics  (Pulliam  1988), 
with  Honey  Lake/Jay  Dow,  Sr.  Wetlands  serv- 
ing as  a source  population  in  relation  to  the 
three  Oregon  lakes.  This  hypothesis  is  further 
supported  by  observations  that  productivity  at 
Jay  Dow,  Sr.  Wetlands  has  generally  been 
higher  than  at  the  other  study  areas  (S.M.H. 
and  L.W.O.,  unpubl.  data).  Historically,  in  a 
system  with  high  interannual  variability  in 
habitat  patch  quality  resulting  from  fluctua- 
tions in  precipitation  (Engilis  and  Reid  1997, 
Robinson  and  Warnock  1997),  sources  and 
sinks  may  have  shifted  periodically.  Such 
shifts  would  characterize  a rescue  effect  me- 
tapopulation (Brown  and  Kodric-Brown 
1977),  in  which  local  populations,  in  danger 
of  extinction  if  isolated,  nevertheless  persist 
as  their  numbers  are  buffered  by  immigration 
from  populations  with  more  favorable  breed- 
ing conditions  (Stacey  et  al.  1997).  As  a re- 
cently developed,  managed  wetland,  the  Jay 
Dow,  Sr.  Wetlands  may  in  fact  now  be  a stable 
source  for  avocet  populations  at  other  wetland 
systems  throughout  the  western  Great  Basin. 

Post-breeding  movements  also  suggested 
seasonal  differences  in  the  use  of  various  wet- 
lands within  the  region.  Primary  breeding  ar- 
eas such  as  the  managed  wetlands  (Jay  Dow, 
Sr.  Wetlands  and  Summer  Lake  WA)  were  rel- 
atively less  important  as  post-breeding  habi- 
tats, whereas  the  major  lake  bodies,  particu- 
larly the  three  northern  waterbodies  in  our 
study  (Goose  Lake,  Lake  Abert,  and  Summer 
Lake),  supported  much  higher  densities  of 
birds  in  late  summer  and  early  fall  than  during 
the  breeding  season  (Warnock  et  al.  1998). 
Our  observations  of  post-breeding  dispersal 
from  Honey  Lake  to  northern  lakes  may  re- 
flect differential  habitat  suitability  of  the  lake 
systems  during  this  time  period.  Avocets  from 
Honey  Lake  also  have  been  reported  in  post- 
breeding flocks  at  water  bodies  east  and  south 
of  the  breeding/natal  lake  system  (Robinson 
and  Oring  1996,  1997b;  Plissner  et  al.  in 
press);  indicating  a multidirectional  exodus 
from  this  particular  breeding  area  in  late  sum- 
mer and  early  fall.  As  an  alternative  to  hy- 
potheses of  seasonal  resource  tracking,  other 


318 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


studies  have  suggested  that  premigratory 
movements  of  birds  are  exploratory  or  other- 
wise associated  with  identification  of  future 
breeding  areas  (Morton  et  al.  1991,  Reed  and 
Oring  1992,  Reed  et  al.  in  press).  Our  results, 
however,  do  not  provide  evidence  for  associ- 
ations between  post-breeding  locations  and 
subsequent  nest  sites,  as  would  be  expected 
for  support  of  this  hypothesis. 

The  observed  patterns  of  dispersal  move- 
ments were  confirmed  by  radio  telemetry  data 
(Plissner  et  al.  in  press),  which  also  demon- 
strated a northerly  trend  in  post-breeding 
movements  of  avocets  between  Carson  Lake, 
Churchill  County,  Nevada  (Fig.  1)  and  the 
three  Oregon  lakes  of  this  study,  without  ev- 
idence of  southward  post-breeding  move- 
ments by  Oregon  breeders.  Telemetry  data 
further  indicated  that  a very  high  proportion 
of  post-breeding  adults  (71%)  frequented  wet- 
lands other  than  their  breeding  locations.  In 
addition,  radio-tagged  individuals  frequently 
were  not  detected  at  any  of  the  primary  lakes 
during  some  telemetry  surveys,  suggesting 
even  greater  rates  of  movements  away  from 
breeding  locations  during  the  time  period. 
Therefore  it  is  likely  that  visual  resightings  of 
individuals  represent  a conservative  estimate 
of  the  extent  of  post-breeding  movements  in 
the  region.  Nevertheless,  these  data  provided 
information  on  movements  across  multiple 
years  and  for  first-year  birds,  which  can  not 
be  obtained  effectively  using  current  teleme- 
try methodologies. 

Post-breeding  locations  may  also  be  asso- 
ciated with  specific  migration  routes  and/or 
wintering  areas  for  these  populations.  All 
birds  reported  during  the  fall/winter  along  the 
northern  California  coast,  including  those 
originating  from  Honey  Lake,  were  last  ob- 
served at  one  of  the  three  northern  lakes.  Pre- 
vious reports  suggest  that  birds  originating 
from  Honey  Lake  primarily  migrate  to  win- 
tering areas  in  California’s  Central  Valley  and 
coastal  areas  from  San  Francisco  Bay  south  to 
the  central  coast  of  Sinaloa,  Mexico  (Robin- 
son and  Oring  1996,  Robinson  et  al.  1997). 
Numbers  of  avocets  at  the  northern  extreme 
of  the  species’  winter  range  (Humboldt  Bay, 
California),  however,  have  been  increasing 
since  I960  (Evans  and  Harris  1994),  and  it  is 
clear  that  many  of  these  birds  originated  from 
breeding  areas  throughout  the  western  Great 


Basin.  Although  further  data  are  needed  from 
other  wintering  areas,  it  appears  possible  that 
birds  that  aggregate  at  post-breeding  areas  in 
the  northwestern  Great  Basin  may  overwinter 
along  the  northern  California  coast,  whereas 
others,  even  from  the  same  breeding  area,  may 
migrate  directly  south  and  west  to  more  south- 
erly post-breeding  and  wintering  sites. 

Our  studies  of  American  Avocet  move- 
ments in  the  western  Great  Basin  indicate  that 
a dispersal-based  evaluation  of  habitat  con- 
nectivity requires  an  understanding  of  move- 
ments at  multiple  temporal  and  spatial  scales. 
Based  solely  upon  movements  of  individuals 
between  breeding  sites,  connectivity  among 
the  different  wetlands  of  the  region  would  ap- 
pear to  be  weak.  Rates  of  post-breeding  move- 
ments among  different  lake  systems,  however, 
was  substantially  higher,  providing  evidence 
of  a strong  link  between  the  different  systems. 
An  apparent  northward  trend  in  long-distance 
post-breeding  dispersal  suggests  that  the 
northern  lake  systems  may  provide  better  re- 
sources for  avocets  during  this  time  period, 
while  the  same  trend  in  between-year  move- 
ments may  be  simply  a geographic  artifact  of 
the  location  of  a source  population  relative  to 
other  study  populations  with  lower  productiv- 
ity. This  hypothesis  also  is  supported  by  ob- 
servations of  avocets  from  Jay  Dow,  Sr.  Wet- 
land at  stopover  areas  farther  south  (Robinson 
and  Oring  1996),  suggesting  a multi-direction- 
al exodus  of  post-breeding  birds  from  Honey 
Lake  breeding  sites.  Thus,  avocets  use  a large 
array  of  Great  Basin  wetlands  within  and 
among  years,  suggesting  that  conservation  ef- 
forts should  consider  this  complexity  in  defin- 
ing appropriate  habitat  conservation  strate- 
gies. 

ACKNOWLEDGMENTS 

Funding  for  this  project  was  provided  by  the  Bio- 
logical Resources  Division  of  the  U.S.  Geological  Sur- 
vey, the  USGS  Forest  and  Rangeland  Ecosystem  Sci- 
ence Center,  a USDA  Hatch  grant  to  L.W.O.  via  the 
University  of  Nevada-Reno,  NSF  grant  DEB  9424375 
to  L.W.O.,  and  USDA  grant  95-37101-2026  to  L.W.O. 
and  M.  Rubega.  Over  30  field  assistants  were  indis- 
pensible  in  gathering  the  data  during  the  years  of  the 
study.  We  also  thank  the  12  private  landowners  who 
provided  access  to  properties  adjoining  many  of  the 
wetlands.  M.  Colwell  and  C.  Gratto-Trevor  provided 
reports  of  banded  birds  observed  along  the  California 
coast  during  the  non-breeding  season.  Additional  sup- 


Flissiier  et  al.  • AVOCET  DISPERSAL 


319 


port  was  provided  by  M.  St.  Louis,  O.  Tat't,  Oregon 

Department  of  Eish  and  Wildlife,  the  Jay  Dow,  Sr. 

Wetlands,  and  the  University  of  Nevada-Reno.  We 

thank  C.  Gratto-Trevor  and  two  anonymous  reviewers 

for  their  helpful  comments. 

LITERATURE  CITED 

Barrowclough,  G.  E 1978.  Sampling  bias  in  dispersal 
studies  based  on  finite  area.  Bird-Banding  49:333- 
341. 

Brown,  J.  H.  and  A.  Kodric-Brown.  1977.  Turnover 
rates  in  insular  biogeography:  effect  of  immigra- 
tion on  extinction.  Ecology  58:445-449. 

Cadbury,  C.  J.,  D.  Hill,  J.  Partridge,  and  J.  Sor- 
ensen. 1989.  The  history  of  the  Avocet  population 
and  its  management  in  England  since  recolonisa- 
tion. R.  Soc.  Prot.  Birds  Conserv.  Rev.  3:9-13. 

C.xdbury,  C.  j.  and  P.  j.  S.  Olney.  1978.  Avocet  pop- 
ulation dynamics  in  England.  British  Birds  71: 
102-121. 

Cunningham,  M.  A.  1986.  Dispersal  in  White-crowned 
Sparrows:  a computer  simulation  of  study-area  ef- 
fects on  estimates  of  local  recruitment.  Auk  103: 
79-85. 

Doebeli,  M.  and  G.  D.  Ruxton.  1997.  Evolution  of 
dispersal  rates  in  metapopulation  models:  branch- 
ing and  cyclic  dynamics  in  phenotype  space.  Evo- 
lution 51:1730-1741. 

Dunning,  J.  B.,  Jr.,  D.  J.  Stewart,  B.  J.  Danielson, 
B.  R.  Noon,  T.  L.  Root,  R.  H.  Lamberson,  and 
E.  E.  Stevens.  1998.  Spatially  explicit  population 
models:  current  forms  and  future  uses.  Ecol.  Appl. 
5:3-1 1. 

Engilis,  a.,  Jr.  and  E A.  Reid.  1997.  Challenges  in 
wetland  restoration  of  the  western  Great  Basin. 
Int.  Wader  Stud.  9:71-79. 

Evans,  T.  J.  and  S.  W.  Harris.  1994.  Status  and  hab- 
itat use  by  American  Avocets  wintering  at  Hum- 
boldt Bay,  California.  Condor  96:178-189. 

Haig,  S.  M.,  D.  W.  Mehlman,  and  L.  W.  Oring.  1998. 
Avian  movements  and  wetland  connectivity  in 
landscape  conservation.  Conserv.  Biol.  12:749- 
758. 

Hamilton,  R.  B.  1975.  Comparative  behavior  of  the 
American  Avocet  and  the  Black-necked  Stilt  (Re- 
curvirostridae).  Ornithol.  Monogr.  17:1-98. 

IMS,  R.  A.  AND  N.  G.  Yoccoz.  1997.  Studying  transfer 
processes  in  metapopulations:  emigration,  migra- 
tion, and  colonization.  Pp.  247-265  in  Metapopu- 
lation biology:  ecology,  genetics,  and  evolution  (I. 
A.  Hanski  and  M.  E.  Gilpin,  Ed.)  Academic  Press, 
San  Diego,  California. 

James,  R.  A.,  Jr.  1995.  Natal  philopatry,  site  tenacity, 
and  age  of  first  breeding  of  the  Black-necked  Stilt. 
J.  Field  Ornithol.  66:107-1 1 I. 

McPeek,  M.  a.  and  R.  D.  Holt.  1992.  The  evolution 
of  dispersal  in  spatially  and  temporally  varying 
environments.  Am.  Nat.  140:1010-1027. 

Morton,  M.  L.,  M.  W.  Wakamatsu,  M.  E.  Pereyra, 
AND  G.  A.  Morton.  1991.  Postfledging  dispersal. 


habitat  imprinting,  and  philopatry  in  a montane, 
migratory  sparrow.  Ornis  Scand.  22:98-106. 

Neel,  L.  A.  and  W.  G.  Henry.  1997.  Shorebirds  of 
the  Lahontan  Valley,  Nevada,  USA:  a case  history 
of  western  Great  Basin  shorebirds.  Int.  Wader 
Stud.  9:15-19. 

Oring,  L.  W.  and  D.  B.  Lank.  1984.  Breeding  area 
fidelity,  natal  philopatry,  and  the  social  systems  of 
sandpipers.  Pp.  125-147  in  Shorebirds:  breeding 
behavior  and  populations  (J.  Burger  and  B.  L. 
Olla,  Ed.).  Plenum  Publishing,  New  York. 

Oring,  L.  W.  and  J.  M.  Reed.  1997.  Shorebirds  of  the 
western  Great  Basin  of  North  America:  overview 
and  importance  to  continental  populations.  Int. 
Wader  Stud.  9:6-12. 

Plissner,  j.  R,  S.  M.  Haig,  and  L.  W.  Oring.  In  press. 
Post-breeding  movements  of  American  Avocets 
and  wetland  connectivity  in  the  western  great  ba- 
sin. Auk. 

Pulliam,  H.  R.  1988.  Sources,  sinks,  and  population 
regulation.  Am.  Nat.  132:652-661. 

Reed,  J.  M.,  T.  Boulinier,  E.  Danchin,  and  L.  W. 
Oring.  In  press.  Informed  dispersal:  prospecting 
by  birds  for  breeding  sites.  Curr.  Ornithol.  15:  In 
press. 

Reed,  J.  M.  and  L.  W.  Oring.  1992.  Reconnaissance 
for  future  breeding  sites  by  Spotted  Sandpipers. 
Behav.  Ecol.  3:310-317. 

Robinson,  J.  A.  and  L.  W.  Oring.  1996.  Long-distance 
movements  by  American  Avocets  and  Black- 
necked Stilts.  J.  Field  Ornithol.  67:307-320. 

Robinson,  J.  A.  and  L.  W.  Oring.  1997a.  Fading  of 
UV-stable  coloured  bands  on  shorebirds.  Wader 
Stud.  Group  Bull.  84:45—46. 

Robinson,  J.  A.  and  L.  W.  Oring.  1997b.  Natal  and 
breeding  dispersal  in  American  Avocets.  Auk  1 14: 
416-430. 

Robinson,  J.  A.,  L.  W.  Oring,  J.  P.  Skorupa,  and  R. 
Boettcher.  1997.  American  Avocet  {Recun’iros- 
tra  americana).  In  The  birds  of  North  America, 
no.  275  (A.  Poole  and  E Gill,  Ed.).  The  Academy 
of  Natural  Sciences,  Philadelphia,  Pennsylvania; 
The  American  Ornithologists’  Union,  Washington, 
D.C. 

Robinson,  J.  A.  and  S.  E.  Warnock.  1997.  The  stag- 
ing paradigm  and  wetland  conservation  in  arid  en- 
vironments; shorebirds  and  wetlands  of  the  North 
American  Great  Basin.  Int.  Wader  Stud.  9:37-44. 

Saunders,  D.  A.,  R.  J.  Hobbs,  and  C.  R.  Margules. 
1991.  Biological  con.sequcnces  of  ecosystem  frag- 
mentation: a review.  Con.serv.  Biol.  5:18-32. 

SoKAL,  R.  R.  AND  E J.  Rohlf.  1981.  Biometry.  W.  H. 
Freeman  and  Company.  San  Francisco,  California. 

SoRDAHL,  T.  A.  1984.  Observations  on  breeding  site 
fidelity  and  pair  formation  in  American  Avocets 
and  Black-necked  Stilts.  N.  Am.  Bird  Bander  9: 
8-11 

Stacey,  P.  B.,  V.  A.  John.son,  and  M.  L.  Taper.  1997. 
Migration  within  metapopulations:  the  impact 
upon  local  population  dynamics.  Pp.  267-291  in 
Metapopulation  biology:  ecology,  genetics,  and 


320 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


evolution  (1.  A.  Hanski  and  M.  E.  Gilpin,  Ed.). 
Academic  Press,  San  Diego,  California. 

Taylor,  P.  D.,  L.  Fahrig,  K.  Henein,  and  G.  Merri- 
AM.  1993.  Connectivity  is  a vital  element  of  land- 
scape structure.  Oikos  68:571-573. 

Warnock,  N.,  S.  M.  Haig,  and  L.  W.  Oring.  1998. 
Monitoring  species  richness  and  abundance  of 
shorebirds  in  the  western  Great  Basin.  Condor 
100:589-600. 

Waxier,  J.-M.  and  O.  Fournier.  1980.  Elements  de 
demographie  de  la  population  d’Avocettes  (Re- 
curvirostra  avosetta)  de  la  cote  atlantique  fran- 


(jaise.  LOiseau  Rev.  Fran^aise  Ornithol.  50:307- 
321. 

Wiens,  J.  A.  1997.  Metapopulation  dynamics  and  land- 
scape ecology.  Pp.  43—62  in  Metapopulation  bi- 
ology: ecology,  genetics,  and  evolution  (I.  A. 
Hanski  and  M.  E.  Gilpin,  Ed.).  Academic  Press, 
San  Diego,  California. 

With,  K.  A.,  R.  H.  Gardiner,  and  M.  G.  Turner. 
1997.  Landscape  connectivity  and  population  dis- 
tributions in  heterogeneous  environments.  Oikos 
78:151-169. 

Wright,  S.  1951.  The  genetical  structure  of  popula- 
tions. Ann.  Eugenics  15:323-354. 


Wilson  Bull.,  111(3),  1999,  pp.  321-329 


HIGH  MORTALITY  OF  PIPING  PLOVERS  ON  BEACHES  WITH 
ABUNDANT  GHOST  CRABS:  CORRELATION,  NOT  CAUSATION 

DONNA  L.  WOLCOTT*  2 AND  THOMAS  G.  WOLCOTT* 


ABSTRACT. — Ghost  crabs  (Ocypode  quad  rata)  have  been  implicated  in  mortality  of  eggs  and  chicks  of  the 
beach-nesting  Piping  Plover  (Charadrius  melodus)  whose  Atlantic  Coast  populations  are  listed  as  threatened. 
Through  observation  and  experimentation,  we  investigated  the  interactions  between  ghost  crahs  and  plovers  on 
Wild  Beach,  a Piping  Plover  nesting  area  on  Assateague  Island,  Virginia.  This  site  has  a high  abundance  of 
ghost  crabs  and  historically  low  fledging  success  compared  to  adjacent  areas  with  fewer  crabs.  We  observed 
encounters  of  crabs  with  plover  eggs,  chicks,  and  adults  in  the  field,  but  never  predation.  In  staged  encounters 
of  crabs  with  eggs  and  chicks  (using  hatchery  reared  quail  as  plover  surrogates),  we  were  unable  to  elicit 
predatory  behavior  either  on  the  beach  or  in  the  lab.  We  conclude  that  although  instances  of  ghost  crab  predation 
on  Piping  Plover  eggs  and  chicks  occur,  they  are  rare  and  cannot  account  for  the  high  mortality  frequently 
reported  on  beaches  where  ghost  crabs  are  abundant.  Adult  plovers  behave  toward  crabs  as  if  they  were  dan- 
gerous to  eggs  and  chicks,  and  their  young  hroods  in  the  study  area  did  not  forage  along  the  foreshore.  Hence, 
ghost  crabs  may  increase  mortality  indirectly.  Frequent  responding  to  crabs  by  parents  may  attract  more  deadly 
brood  predators.  Brood  nutrition  may  suffer  as  adult  plovers  direct  chicks  away  from  areas  where  forage  is 
reportedly  richer  but  crabs  are  abundant,  such  as  the  foreshore.  Nutrient  intake  may  be  further  reduced  on  more 
southerly  breeding  grounds  where  high  temperatures  on  backshores  force  chicks  to  stop  foraging  and  take  shelter 
during  mid-day.  Although  high  mortality  cannot  be  attributed  directly  to  predation  by  crabs,  it  may  be  due  to 
factors  that  covary  with  crab  abundance,  such  as  high  temperature,  behavioral  responses  of  adult  birds,  and  poor 


forage.  Received  28  April  1998,  accepted  7 Feb.  1999. 

Anecdotal  and  published  reports  of  ghost 
crab  predation  on  Piping  Plovers  {Charadrius 
melodus;  Loegering  et  al.  1995,  Watts  and 
Bradshaw  1995)  have  led  to  concern  that  crab 
predation  may  hamper  recovery  of  plovers  on 
the  Atlantic  Coast,  where  the  species  is  listed 
as  threatened  (Loegering  and  Fraser  1995; 
U.S.  Fish  and  Wildlife  Service  1993,  1996). 
To  assess  the  extent  of  crab-caused  mortality, 
we  investigated  interactions  between  ghost 
crabs  {Ocypode  quadrata)  and  Piping  Plovers 
during  incubation  and  chick  rearing  on  Wild 
Beach  on  Assateague  Island,  Virginia,  within 
the  Chincoteague  National  Wildlife  Refuge. 
Compared  to  other  portions  of  this  barrier  is- 
land that  are  used  for  nesting  areas  by  the  Pip- 
ing Plover,  Wild  Beach  has  higher  abundances 
of  ghost  crabs  (Britton  1979)  and  lower  rates 
of  fledging  success  (U.S.  Fish  and  Wildlife 
Service  1994). 

Piping  Plovers  breed  from  eastern  Canada 
to  North  Carolina,  as  well  as  in  the  Great 
Lakes  region  and  the  Great  Plains  of  Canada 
and  the  U.S.  (Haig  1992).  On  the  Atlantic 
Coast,  plovers  typically  lay  four  eggs  in  a 


' Dept,  of  Marine,  Earth  and  Atmospheric  Sciences, 
Box  8208,  North  Carolina  State  Univ.,  Raleigh,  NC 
27695-8208. 

^ Corresponding  author;  E-mail:  dwolcott@ncsu.edu 


shallow  scrape  in  the  sand,  usually  well  be- 
yond the  high-tide  mark,  or  in  shelly  storm- 
flattened  areas  (washouts)  between  and  behind 
the  primary  dunes.  Chicks  are  precocial  and 
forage  in  moist  backshore  areas  where  avail- 
able, or  on  the  foreshore  (the  area  between  the 
tides;  Loegering  and  Fraser  1995). 

Ghost  crabs,  named  for  their  cryptic  col- 
oration, range  along  the  Atlantic  Coast  from 
Rhode  Island  to  Brazil  and  throughout  the  Ca- 
ribbean (Chace  and  Hobbs  1969).  They  are 
among  the  fastest  terrestrial  invertebrates  (Full 
and  Weinstein  1992)  and  formidable  predators 
with  acute  sensory  receptors  for  vision,  vibra- 
tion, taste,  and  smell  (Cowles  1908,  Wellins 
et  al.  1989).  They  are  most  abundant  on  high 
energy  beaches,  where  they  obtain  over  90% 
of  their  diet  preying  on  intertidal  invertebrates 
(Wolcott  1978).  They  are  extremely  flexible 
foragers,  also  scavenging,  deposit  feeding, 
consuming  seeds  and  insects,  and  are  docu- 
mented predators  of  turtle  hatchlings  (Arndt 
1994,  Robertson  and  Pfeiffer  1982).  At  dusk 
crabs  move  from  their  burrows  on  the  back- 
shore  and  among  the  dunes  to  feed  in  the 
swash  zone.  Crabs  seek  out  and  take  refuge  in 
burrows  as  dawn  approaches  (Wolcott  1978). 

Poor  fledging  success  on  beaches  where 
crabs  are  abundant,  coupled  with  extensive 


321 


322 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


seasonal  and  spatial  overlap  between  plovers 
and  the  predatory  ghost  crabs,  led  to  the  hy- 
pothesis that  ghost  crab  predation  is  a com- 
mon source  of  mortality  for  plover  eggs  and 
chicks  (U.S.  Fish  and  Wildlife  1993,  Loeger- 
ing  and  Fraser  1995).  To  test  this  hypothesis, 
we  documented  and  quantified  natural  en- 
counters between  crabs  and  Piping  Plover 
eggs  and  chicks,  and  staged  encounters  be- 
tween crabs  and  the  eggs  and  chicks  of  non- 
threatened  species. 

METHODS 

We  conducted  our  study  on  a 6.4  km  stretch  of  the 
Chincoteague  National  Wildlife  Refuge’s  (NWR)  Wild 
Beach,  Virginia,  in  June  and  July,  1994.  Wild  Beach 
has  a steeply-sloped  foreshore  rising  from  a high  en- 
ergy surf  zone  to  the  berm  and  was  characterized  in 
1994  by  one  or  two  wave-cut  scarps  less  than  1 m 
high.  At  the  berm  the  slope  decreased  abruptly  and 
there  was  a narrow  (20-50  m),  flat  backshore  region. 
Inland  of  the  backshore  are  low  vegetated  foredunes, 
fronting  2-4  m high  artificially  stabilized  dunes  veg- 
etated with  beach  grass  (Ammophila  breviligulala)  and 
seaside  goldenrod  (Solidago  semperviren.s). 

We  observed  four  natural  plover  nests  through  the 
last  days  of  incubation  and  past  hatching  using  contin- 
uous video  monitoring  to  document  crab  predation  on 
eggs  and  newly  hatched  chicks.  When  the  third  egg 
was  laid  in  each  nest,  staff  from  the  refuge  surrounded 
the  nest  with  a 4 m diameter,  1.5  m high  exclosure 
constructed  of  5 X 10  cm  welded  mesh  wire  supported 
by  reinforcing  bar  and  covered  with  plastic  netting.  A 
dummy  camera  was  mounted  at  the  top  of  an  exclosure 
support  bar  on  the  southwest  side,  facing  toward  the 
nest  and  away  from  the  prevailing  wind.  At  the  time 
nest  observation  began,  the  dummy  camera  was  re- 
placed with  a similar  appearing  video  surveillance 
camera  fitted  with  infrared  light-emitting  diodes  for 
nocturnal  illumination,  and  an  infrared  filter  to  prevent 
saturation  in  strong  sunlight.  The  camera  imaged  an 
area  approximately  2 X 3 m centered  on  the  nest.  In- 
stalling the  camera  in  place  of  the  dummy  took  3-12 
min,  and  birds  returned  to  the  exclosure  1-9  min  there- 
after (mean  = 4.25  min.,  n = 4).  Coaxial  and  power 
cables  (150  m)  led  to  a TV/VCR  (Magnavox  model 
CCR095)  powered  by  a deep-cycle  lead-acid  battery. 
These  were  housed  in  a tent  behind  nearby  dunes  to 
reduce  disturbance  to  the  plovers.  Tapes  were  changed 
at  six-hour  intervals  and  the  battery  every  18  hours. 
According  to  the  video  record,  the  incubating  adult 
plovers  typically  did  not  flush  from  the  nest  at  those 
times.  Video  tapes  were  reviewed  for  the  following 
occurrences  and  the  times  logged:  adult  plover  brood- 
ing, alarming  and  flushing  from  the  nest  (rapid  rising 
off  eggs,  standing  near  nest  with  wing  raises,  rapid  exit 
from  the  field  of  view),  parental  exchanges  of  incu- 
bating/brooding duty  (second  parent  appears  on  screen 
and  changes  places  with  the  bird  on  the  nest),  eggs 


hatching,  activities  of  chicks  (including  walking,  for- 
aging and  being  brooded),  and  any  activities  of  ghost 
crabs. 

To  explore  the  role  of  crab  predation  after  chicks 
left  the  nest,  we  observed  encounters  between  crabs 
and  adult  plovers  and  chicks  for  eight  days,  concen- 
trating on  periods  of  peak  crab  activity.  Observations 
of  hatched  broods  were  made  from  a vehicle  at  least 
30  m away,  using  binoculars  during  daylight  hours, 
and  an  image-intensifying  scope  (Varo  Noctron  IV)  at 
night  fitted  with  an  infrared  diode  laser  to  enhance 
illumination.  Plover  chicks  in  two  broods  were  marked 
on  each  thigh  with  10  mm  diameter  disks  of  Scotch- 
lite®  (3M)  reflectorized  tape  glued  to  the  surface  of 
their  down  to  make  them  visible  with  night  vision 
equipment.  Two  of  the  marked  chicks  were  the  only 
chicks  to  fledge  in  the  study  area;  thus  the  treatment 
did  not  appear  to  increase  mortality.  Data  collected 
during  each  observation  period  included  the  location 
of  the  brood,  the  place  and  time  spent  foraging,  and 
the  location  of  brooding. 

To  further  assess  the  probability  of  ghost  crabs  prey- 
ing on  eggs,  we  presented  crabs  with  surrogate  eggs 
that  were  similar  in  size  and  shape  to  those  of  Char- 
adrius  melodus  (see  Macivor  et.  al.  1990).  Japanese 
Quail  (Coturni.x  japonicu.s)  eggs  were  obtained  from  a 
local  bird  breeder,  and  Northern  Bobwhite  (Colinus 
virginianu.s)  eggs  were  obtained  from  Seven  Oaks 
Game  Earm  and  Supply,  Wilmington,  North  Carolina. 
On  three  nights,  we  constructed  four  scrapes  near  the 
berm  and  placed  four  Cotiimix  eggs  in  each  just  prior 
to  the  time  that  crabs  emerge  from  their  burrows.  Dur- 
ing the  nocturnal  peak  in  crab  activity  (19:30-22:00 
EST),  the  artificial  nests  were  observed  for  any  ghost 
crab  encounters  using  the  video  camera  unit  fixed  on 
a tripod. 

To  further  explore  the  vulnerability  of  eggs  to  ghost 
crab  predation,  we  placed  four  opaque  gray  plastic  bins 
(38L  X 18W  X 1 ID  cm)  in  a rectangular  array  in  the 
laboratory  and  filled  them  to  a depth  of  about  3 cm 
with  damp  sand  from  the  foreshore.  In  each  we  placed 
a ghost  crab  [average  weight  49.0  ± 5.4  (SD)  g,  with 
a carapace  width  of  43.2  ± 1.6  mm,  n = 8]  freshly- 
collected  from  the  foreshore  of  Wild  Beach,  two  Co- 
turni.x  eggs,  and  the  bins  were  covered  with  chicken 
wire.  Crab  activity  was  recorded  using  the  infrared 
camera  and  VCR  from  the  initiation  of  the  experiment 
at  23:15  until  05:15,  and  condition  of  the  eggs  was 
assessed  at  10:00.  The  experiment  was  repeated  using 
dry  beach  sand  and  Colinu.s  eggs  on  another  night,  and 
condition  of  the  eggs  assessed  after  I 1 hours  of  ex- 
posure to  the  crabs.  From  the  video  tapes  we  noted  the 
length  of  time  crabs  spent  in  contact  with  eggs,  their 
behavior,  and  the  condition  of  the  eggs  after  manipu- 
lation by  the  crabs. 

To  elucidate  potential  predatory  interactions  be- 
tween crabs  and  chicks,  we  used  chicks  of  Northern 
Bobwhite  as  surrogates  for  Piping  Plover  chicks.  Bob- 
white  chicks  are  similar  in  size  and  behavior  to  Piping 
Plover  chicks,  but  darker  and  hence  less  cryptic  on 
beach  sand.  Two-  and  three-day-old  chicks  (u  = 14) 


Wolcotl  ami  Wolcott  • PIPING  PLOVERS  AND  GHOST  CRABS 


323 


were  released  near  the  beach  berm  into  areas  of  high 
crab  activity  to  maximize  encounters  with  crabs  and 
were  observed  with  binoculars  as  they  wandered  on 
the  beach.  In  four  cases,  chicks  were  deployed  from 
the  parked  vehicle  (to  which  crabs  showed  little  re- 
sponse) into  the  immediate  vicinity  of  active  crabs.  To 
verify  that  crabs  were  motivated  to  forage,  we  placed 
dead  chicks,  freshly  cracked  oysters,  and  pieces  of 
thawed  chicken  upwind  near  the  mouths  of  occupied 
burrows  at  the  same  time  that  live  chicks  were  released 
onto  the  beach.  Encounters  between  crabs  and  chicks 
and  dead  prey  were  documented. 

To  determine  temporal  overlap  in  activity  between 
crabs  and  plovers,  nocturnal  and  diurnal  observations 
of  crab  activity  and  behavior  were  conducted  from  a 
vehicle.  The  onset  of  migration  of  crabs  to  the  surf 
zone  for  nocturnal  foraging  was  determined  by  noting 
the  time  of  appearance  of  the  first  crab  moving  down 
the  beach  each  night. 

Statistics  are  reported  as  mean  plus  or  minus  one 
standard  deviation. 

RESULTS 

Predation  by  crabs  on  eggs. — Ghost  crabs 
did  not  prey  on  eggs  in  either  natural  nests  or 
in  experimental  trials,  and  crabs  showed  com- 
parable behavior  toward  eggs  in  all  settings. 
In  147  hours  of  video  observations  of  incu- 
bation by  Piping  Plovers,  there  were  seven  ap- 
pearances of  ghost  crabs  in  the  video  field. 
These  appearances  generally  were  confined  to 
the  periods  of  crepuscular  movement  by  the 
crabs  (dawn,  04:16  ± 11.3  min,  n = 2 and 
dusk,  20:01  ± 4.6  min,  n = 3).  Two  excep- 
tions occurred  on  one  rainy  day  when  high 
humidity  and  reduced  insolation  resulted  in 
diurnal  crab  activity.  In  two  cases,  crabs  di- 
rectly contacted  and  manipulated  the  eggs  in 
the  nest  cup  after  the  incubating  adult  had 
flushed  from  the  nest.  Crabs  appeared  to  be 
testing  the  eggs  as  potential  prey,  using  chelae 
and  mouthparts  which  contain  dense  arrays  of 
chemoreceptors  to  “taste”  the  eggs.  The  du- 
rations of  manipulations  were  13  s and  23  s. 

Crabs  that  contacted  Coturnix  eggs  in  pseu- 
do-nests in  the  field  either  continued  slowly 
toward  the  foreshore  or  stopped  to  manipulate 
and  taste  the  eggs.  Crabs  spent  more  time  in 
contact  with  eggs  when  manipulating  them 
(17.3  ± 3.06  s,  n = 3)  than  when  they  were 
not  manipulating  them  (1 1.4  ± 8.4  s,  n — 5). 
In  one  instance,  a crab  spent  6 minutes  and 
12  seconds  at  a pseudo-nest  of  Coturnix  eggs 
on  the  beach,  repeatedly  tasting  the  eggs  and 
rolling  one  egg  several  cm  prior  to  moving  on 
to  the  surf  zone.  We  interpret  this  intense  in- 


terest as  a response  to  the  bird  feces  present 
on  that  particular  egg. 

Crabs  confined  with  Colinus  eggs  in  the 
laboratory  explored  the  eggs  as  potential  prey 
at  least  once,  not  necessarily  on  the  first  con- 
tact, but  simply  walked  over  the  eggs  in  other 
encounters.  Four  crabs  spent  an  average  of 
12.2  s (±  8.2  s,  n = 11),  10.6  s (±  11.2  s,  n 
= 8),  6.9  s (±  6.7  s,  n = 10),  and  7.3  s (± 
5.9  s,  n = 6)  on  eggs  per  encounter.  Average 
time  of  encounters  in  which  manipulation  oc- 
curred (all  crabs)  was  13.7  s (±  8.4  s,  n = 7). 

Potential  for  predation  by  crabs  on 
chicks. — Three  different  Piping  Plover  broods 
were  observed  for  a total  of  26.2  h.  A chick 
was  seen  to  pass  near  a crab  only  once,  and 
in  that  instance,  there  was  no  response  from 
the  crab. 

After  leaving  the  nest  and  exclosure,  chicks 
were  almost  always  found  foraging  in  vege- 
tated areas,  principally  on  the  low  foredunes 
(ji  = 18).  Chicks  from  Nest  Five  were  found 
on  high  dunes  because  low  foredunes  were 
uncommon  in  their  territory.  This  brood  was 
seen  once  on  the  foreshore  of  Wild  Beach  on 
the  day  after  it  left  the  exclosure.  Surviving 
chicks  from  all  broods  were  taking  5-10  m 
forays  away  from  the  attending  parent  by  1.5 
days,  often  in  divergent  directions.  The  at- 
tending adult  was  often  seen  following  after  a 
rapidly  moving  chick.  Typically  the  attending 
adult  stayed  near  the  chicks,  outside  the  veg- 
etation, while  the  other  adult  stationed  itself 
near  the  berm,  standing  sentinel  and  frequent- 
ly feeding  on  the  foreshore.  During  midday 
chicks  alternated  between  foraging  and  resting 
in  the  shade  of  vegetation.  The  parent  typi- 
cally moved  the  chicks  back  to  the  same  gen- 
eral area  each  evening  to  brood.  Typical 
brooding  areas  were  in  shell  hash  (broken 
shell  fragments)  on  the  backshore  near  fore- 
dunes, a few  meters  from  dune  vegetation. 

Activity  periods  of  birds  and  crabs  showed 
little  overlap  once  the  chicks  left  the  nest.  Par- 
ent birds  began  brooding  chicks  as  each  even- 
ing became  cool,  from  about  18:30  to  shortly 
before  20:00  (19:05  ± 28  min,  n = 8),  until 
chicks  ceased  making  forays  altogether,  typi- 
cally between  19:00  and  20:30  (20:00  ± 38 
min,  n = 8).  First  sightings  of  crabs  moving 
to  the  foreshore  occurred  about  the  same  time 
that  chick  activity  fully  ceased  (19:51  ± 9 
min,  n = 7 nights).  Crab  activity  increased 


324 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


during  the  following  hour,  and  chicks  were 
brooded  during  the  peak  of  activity.  In  the 
morning,  crabs  migrated  from  the  foreshore  to 
their  burrows  before  dawn  (04:16  ± 1 1.3  min, 
n = 2 mornings),  and  before  broods  were  seen 
to  become  active.  Some  crab  activity  persisted 
throughout  the  day,  generally  restricted  to  the 
vicinity  of  burrows  whose  densities  increased 
from  the  dunes  toward  the  berm. 

No  instances  of  crab  predation  on  bobwhite 
chicks  occurred,  although  more  than  30  en- 
counters were  observed  in  12  hours  of  diurnal 
and  nocturnal  observation.  In  the  one  case  in 
which  a crab  seized  a quail  chick  that  was 
precipitously  deposited  next  to  its  burrow,  the 
crab  promptly  released  the  chick  unharmed 
and  retreated  to  its  burrow.  Most  commonly, 
crabs  showed  no  response  to  chicks  that  wan- 
dered nearby  within  visual  range.  Crabs  were 
attracted  to,  and  fed  readily  on,  cracked  oyster 
and  chicken  during  the  same  intervals  in 
which  chicks  were  presented,  demonstrating 
the  crabs’  willingness  to  forage  and  feed. 
However,  crabs  routinely  ignored  dead  quail 
chicks,  even  when  they  physically  contacted 
them  in  the  course  of  foraging. 

Crab  interactions  with  adult  Piping  Plo- 
vers.— Interactions  between  incubating  adult 
Piping  Plovers  and  approaching  crabs  were 
variable.  Plovers  either  remained  on  the  nest 
{n  = 4)  or  flushed  in  = 4).  Flushing  typically 
occurred  while  the  crabs  were  further  than  0.5 
m from  the  nest  cup  (n  = 3),  but  once  not 
until  the  crab  approached  within  10  cm.  De- 
fensive encounters  initiated  by  the  plovers 
could  involve  both  parents  (n  — 2).  In  the  four 
instances  in  which  birds  responded  to  crabs, 
the  minimum  length  of  engagement  in  the  vid- 
eo field  was  2 min  (120  ± 49  s,  n = 4).  How- 
ever, adults  left  the  nest  cup  unattended  for 
about  5 min  (307  ± 197  s,  /?  = 4),  presumably 
continuing  the  defense  out  of  the  camera’s 
view. 

Aggressive  displays  by  incubating  adult 
plovers  against  approaching  crabs  were  large- 
ly ineffectual.  Crabs  generally  remained  mo- 
tionless or  maintained  course  (n  = 3)  when 
confronted  by  adult  plovers  advancing  slowly 
with  uplifted  wings,  but  sometimes  they  ran 
when  charged  by  a displaying  bird  (n  = 2). 
In  one  instance,  a crab  whose  course  would 
have  bypassed  the  nest  was  deflected  onto  the 
nest  while  veering  from  the  displaying  parent. 


Thirteen  encounters  between  crabs  and 
adult  plovers  with  unfledged  chicks  were  ob- 
served, with  variable  behavior  by  the  plovers. 
Once,  an  adult  passed  within  10  cm  of  a crab 
with  no  apparent  response.  Where  interactions 
occurred  (n  — 12),  plovers  always  initiated 
them,  although  the  crab  was  8-50  m from  the 
brood.  Plovers  would  approach  and  display 
within  10-20  cm  of  the  crab.  Birds  were  seen 
feigning  and  leading  crabs  toward  the  fore- 
shore (max  distance  = 10  m,  max  time  = 4 
min,  n = 3).  Three  encounters  involved  both 
parents. 

Adult  plovers  appear  to  associate  the  pres- 
ence of  burrows  with  ghost  crabs.  A burrow 
that  was  near  a nest  was  ignored  for  the  first 
85  hours  of  video  observation,  but  was  closely 
inspected  by  the  adult  plovers  on  1 1 separate 
occasions  in  the  final  1 1 hours  of  observation 
after  a crab  had  approached  the  nest  from  that 
direction.  Burrows  were  also  investigated  on 
the  beach  by  parents  with  hatched  broods  (n 
= 2). 

Plover  behavior  and  mortality. — Although 
we  observed  no  instances  of  Piping  Plover 
mortality  directly  attributable  to  predation  by 
ghost  crabs,  we  documented  other  factors  that 
might  directly  or  indirectly  contribute  to  the 
low  fledging  success  on  Wild  Beach.  Hatching 
asynchrony  (substantial  time,  e.g.,  >24  hrs 
between  hatching  of  the  first  and  final  eggs  in 
a given  clutch)  was  responsible  for  the  only 
mortality  for  which  a cause  could  be  estab- 
lished, and  the  only  mortality  to  occur  prior 
to  hatching.  Hatching  asynchrony  showed  a 
strong  seasonal  correlation  in  this  study.  Six 
clutches  of  eggs  were  laid  on  Wild  Beach  in 
1994  (Table  1).  For  the  two  monitored  broods 
in  June,  hatching  was  highly  synchronous. 
The  time  between  the  hatching  of  the  first  egg 
and  the  last  egg  in  an  entire  clutch  averaged 
104  min.  In  the  three  July  broods,  duration  of 
hatching  averaged  at  least  1680  min.  The 
asynchrony 'Contributed  directly  to  the  aban- 
donment of  an  egg,  which  was  determined  to 
be  viable  (Refuge  Staff  pers.  comm.). 

Of  23  chicks  that  hatched  on  Wild  Beach 
in  1994,  two  fledged.  Half  the  mortality  oc- 
curred in  the  first  two  days,  and  75%  by  day 
five  (Fig.  1 ).  This  pattern  of  chick  loss  soon 
after  hatching  is  typical  for  the  species,  but 
even  more  pronounced  than  reported  in  pre- 
vious studies  (most  mortality  in  the  first  10 


Wolcott  and  Wolcott  • PIPING  PLOVERS  AND  GHOST  CRABS 


325 


TABLE  1.  Data  tor  broods  of  Piping  Plover  {Cliaradrins  melodns)  hatching  on  Wild  Beach,  Assateague 
Island,  1994.  Based  on  six  clutches  laid  on  Wild  Beach;  nests  2,  3,  5 and  6 were  monitored  via  video  camera. 


Nest  1 

Nest  2 

Nest  3 

Nest  4 

Nest  .S 

Nest  6 

Hatch  date 

June  10 

June  10 

June  16 

July  4- 

6 July  6/7 

July  8/9 

Hatch  duration 

9 

20  min 

3 hr  08  min 

>36  hr 

>24  hr 

>24  hr 

Interval  to  walk 

0 

1 hr 

1-2.5  hr 

9 

6 hr 

5 hr 

Time  nest  abandoned 

1 ? 

18:49 

16:10 

9 

16:37 

15:44-18:45 

Survival,  days  (time 

of  death  where  known") 

chick  A 

0.5  (n) 

0.5  (n) 

1.5 

9 

1-2  (d) 

0.5  (n) 

chick  B 

0.5  (n) 

0.5  (n) 

1.5 

7 

3-4  (d) 

0.5-1. 5 (n) 

chick  C 

7 

1 (d) 

fledged 

9 

4 (d) 

1-2 

chick  D 

16 

1.5 

fledged 

13 

never  hatched 

4-5 

“ Where  frequency  of  observation  permits,  the  time  of  disappearance  of  chicks  is  given  as  day  (d)  or  night  (n). 


days;  Macivor  1990;  Patterson  1988;  U.S. 
Fish  and  Wildlife  Service  1993).  Mortality  of 
chicks  in  all  the  June  broods  was  highly  con- 
centrated in  the  first  48  hours  (67%),  with 
25%  occurring  during  the  first  night.  In  the 
monitored  July  broods,  mortality  did  not  occur 
during  the  first  night,  but  was  spread  rather 
uniformly  over  the  first  week. 

We  could  not  document  the  cause  of  mor- 
tality and  the  fate  of  “disappeared”  chicks. 
All  mortality  on  Wild  Beach  occurred  after 
broods  had  left  the  nest.  Where  confirmed 
sightings  allowed  the  time  of  disappearance  of 
chicks  to  be  established  with  some  certainty, 
chicks  were  as  apt  to  vanish  during  the  night 
as  during  the  day  (Table  1).  Five  chicks  dis- 
appeared between  sunset  and  09:00,  and  five 
between  08:00  and  19:00.  Signs  of  direct  pre- 
dation were  never  found. 

Differences  in  diel  activity  patterns  between 
June  and  July  broods  were  documented 


through  video  observation,  and  indicated  dif- 
ferences in  potential  for  prey  acquisition  and 
for  practicing  locomotor  and  feeding  behav- 
iors during  the  first  day  of  life.  Hatching  was 
highly  synchronous  in  June,  but  asynchronous 
in  July  (Table  1).  Chicks  emerged  from  the 
nest  cup  for  brief  excursions  as  soon  as  1 hour 
after  hatching  in  the  June  broods,  but  after  5- 
6 hours  in  the  July  broods  (Table  1).  For  June 
broods  (2  and  3),  chicks  spent  extended  pe- 
riods in  the  shade  of  vegetation  near  the  nest, 
or  out  of  the  camera’s  field  of  view,  before 
returning  to  the  brooding  parent,  and  were  ac- 
tive throughout  the  day.  July  chicks  spent  little 
time  away  from  the  nest  and  were  continually 
brooded  during  the  hot  midday. 

Newly  hatched  chicks  in  all  broods  left  the 
exclosure  between  16:00  and  19:00  (Table  1). 
The  chicks  ranged  from  4 to  12.5  hours  in  age, 
and  exhibited  obvious  age-related  differences 
in  coordination  for  walking  and  running,  both 
within  and  between  broods. 

Once  the  nest  was  abandoned,  it  was  no 
longer  possible  to  observe  broods  by  using 
video.  However,  in  one  of  the  asynchronous 
broods,  detailed  observation  of  brooding  and 
foraging  of  day-old  chicks  was  made  by  video 
on  three  chicks  that  spent  their  first  day  post- 
hatch (July  7)  in  the  area  of  the  nest  while  the 
fourth  egg  was  still  being  incubated.  There 
was  an  initial  peak  in  activity  of  the  chicks 
(50%  of  each  hour  spent  moving  about)  be- 
tween 05:00  and  07:00  and  another  (30%)  af- 
ter 14:00.  Activity  remained  below  10%  be- 
tween 09:00-13:00,  and  fell  to  zero  at  12:00 
when  chicks  were  continuously  brooded.  For- 
aging, as  seconds  per  hour  that  a chick  spent 


326 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


in  hunt  and  peck  behaviors,  was  congruent 
with  the  activity  pattern  (300  s/hr  05:(X)-07:00, 
<10  s/hr  11:00-12:00). 

DISCUSSION 

Crab  predation  on  Piping  Plover  chicks. — 
Ghost  crabs  do  not  appear  to  be  directly  re- 
sponsible for  the  poor  breeding  success  of 
Piping  Plovers  on  Wild  Beach  on  Assateague 
Island.  Ghost  crabs  showed  no  predatory  re- 
sponse to  nearby  chicks  of  Piping  Plover  or 
Northern  Bobwhite,  and  occasionally  even  ran 
away  from  chicks.  Since  crabs  that  were  ac- 
tive during  daylight  hours  readily  fed  on  other 
prey,  we  conclude  that  chicks  (at  least  North- 
ern Bobwhite)  are  not  preferred  prey. 

Beyond  the  crabs’  apparent  lack  of  interest 
in  chicks,  there  is  temporal  and  spatial  sepa- 
ration in  crab  and  chick  activity  that  further 
reduces,  but  does  not  eliminate  the  likelihood 
of  interactions.  Ghost  crabs  are  principally 
nocturnal;  they  become  active  after  sunset  and 
migrate  to  the  foreshore  to  prey  on  macroin- 
vertebrates. Crabs  leave  the  foreshore  and  re- 
turn to  burrows  by  dawn.  During  the  early 
morning  hours,  they  may  be  active  near  their 
burrows,  spending  time  on  burrow  repair,  de- 
fense, and  intraspecific  aggression,  until  they 
are  confined  below  the  surface  by  microcli- 
matic conditions.  Crab  activity  is  extended  on 
days  with  low  desiccation  risk  (pers.  obs.).  Al- 
though breeding  adult  plovers  may  forage  ex- 
tensively on  the  foreshore  at  night  (Staine  and 
Burger  1994),  unfledged  chicks  on  Wild 
Beach  are  almost  entirely  diurnal.  During  the 
periods  when  crabs  are  most  active,  chicks  are 
being  continuously  brooded,  at  least  through 
the  first  week  post-hatching  when  most  chick 
mortality  occurs.  In  addition  to  temporal  sep- 
aration, spatial  separation  of  chicks  and  crabs 
was  also  evident.  Only  once  was  a brood  seen 
foraging  in  the  intertidal  prior  to  two  weeks 
post-hatching;  all  other  foraging  occurred 
within  5 m of  the  dunes  onto  the  backshore, 
and  hence  was  well  inland  of  the  berm  and 
the  densest  aggregation  of  crab  burrows. 

The  possibility  of  occasional  predation  on 
foraging  chicks  by  crabs  cannot  be  dismissed. 
Only  a single  documented  event  (Loegering  et 
al.  1995)  has  occurred  in  over  1 16  h of  direct 
observation  in  this  and  other  studies  (Patter- 
.son  1988,  44  h;  Loegering  1992,  46  h;  this 
study,  26  h).  Based  on  our  observations,  pre- 


dation is  most  apt  to  occur  when  a chick  star- 
tles a crab,  especially  one  that  has  been  re- 
cently defending  its  burrow.  Aggression  be- 
tween crabs  peaks  during  the  early  morning 
hours  as  they  compete  for  burrows.  Land 
crabs  of  several  species  will  jump,  either  on 
potential  prey,  or  as  part  of  aggressive  en- 
counters (Herreid  1963,  Evans  et  al.  1976, 
pers.  obs.)  On  video  we  observed  an  hours- 
old  plover  chick  careening  head  first  into  a 
crab  burrow.  Had  the  bunow  been  occupied, 
the  chick  might  have  been  killed.  Neverthe- 
less, our  data  indicate  that  stalking  and  killing 
of  chicks  by  crabs  is  highly  unlikely. 

Crab  predation  on  eggs. — Video  observa- 
tions in  our  study  show  that  ghost  crabs  that 
make  contact  with  Piping  Plover  eggs  inves- 
tigate the  eggs  as  potential  prey  items,  using 
stereotypical  tasting  behaviors.  Crabs  showed 
the  same  behavior  toward  surrogate  eggs  of 
Coturnix  japonica  and  Colinus  virginianus  on 
the  beach  and  in  the  lab.  However,  the  claws 
of  the  largest  crabs  on  Wild  Beach  do  not  have 
a gape  large  enough  to  directly  crush  a plover 
egg,  and  no  predation  was  observed. 

Nevertheless,  ghost  crabs  are  confirmed, 
though  infrequent,  predators  of  Piping  Plover 
eggs  on  barrier  islands  in  Virginia  (Watts  and 
Bradshaw  1995;  Refuge  staff,  this  study)  and 
North  Carolina  (S.  Philhower,  pers.  comm.). 
Viable  eggs  have  been  found  in  crab  burrows 
(S.  Philhower,  pers.  comm.),  but  from  known 
crab  behavior  and  observations  during  this 
study,  we  conclude  that  ghost  crabs  are  most 
likely  to  attack  or  manipulate  eggs  that  are 
rotting,  cracked,  or  dirty.  Crabs  use  dactyls 
and  claws  for  contact  chemoreception,  and 
distance  olfaction  to  track  odor  plumes  to  the 
source  of  a smell  (Wellins  et  al.  1989).  Ghost 
crabs  that  encounter  a large  food  parcel  (e.g., 
a dead  fish)  typically  dig  a burrow  immedi- 
ately adjacent  to  it,  which  provides  security 
for  extended  scavenging.  Crabs  that  burrow 
next  to  a nest  may  do  so  because  they  identify 
plover  eggs  as  potential  food.  For  instance,  a 
ghost  crab  burrow  was  found  immediately  ad- 
jacent to  a Piping  Plover  nest  on  a beach  10 
km  south  of  Wild  Beach  when  refuge  staff 
inspected  the  nest  after  it  was  abandoned  by 
the  adults  after  35  days  of  continuous  incu- 
bation (normal  development  time  is  27  d).  The 
eggs  were  missing.  Excavation  of  the  crab 
burrow  yielded  three  of  the  four  eggs,  one  of 


Wolcott  ami  Wolcott  • PIPING  PLOVERS  AND  GHOST  CRABS 


327 


which  was  emitting  a powerful  smell.  The 
missing  egg  was  assumed  to  have  been  con- 
sumed by  the  crab.  A crab  that  has  experi- 
enced eggs  as  prey  may  subsequently  recog- 
nize intact  and  odor-free  eggs  as  food  through 
non-associative  learning  (Evans  et  al.  1976). 
It  is  unclear  whether  broken  eggs  that  have 
been  found  in  crab  burrows  were  already 
cracked,  and  hence  emitting  an  attractive  odor, 
or  were  cracked  from  hitting  each  other  in  the 
burrow.  Whether  instances  of  nest  predation 
by  ghost  crabs  were  initiated  as  scavenging  or 
as  predation,  the  end  result  is  that  some  viable 
eggs  are  lost  to  crabs.  Nevertheless,  we  con- 
clude that  egg  predation  by  ghost  crabs  cannot 
account  for  poor  breeding  success  of  Piping 
Plovers  on  Wild  Beach. 

Piping  Plover  response  to  crabs. — What- 
ever the  actual  threat  from  ghost  crabs,  adult 
Piping  Plovers  treat  them  as  potential  preda- 
tors. We  observed  16  instances  in  which  one 
or  both  adults  engaged  in  extensive  displays 
against  crabs,  and  in  5 cases,  the  defense  left 
hatched  broods  unguarded.  Further  underscor- 
ing the  perceived  threat  from  crabs,  adult  plo- 
vers seem  able  to  connect  the  presence  of  bur- 
rows with  ghost  crabs,  and  invest  time  and 
energy  in  investigating  burrows. 

The  presence  of  abundant  ghost  crabs  may 
create  indirect  problems  for  plovers  by  several 
mechanisms  related  to  the  adults’  perception 
of  crabs  as  potential  predators.  First,  obvious 
responses  to  ghost  crabs  may  alert  truly  dan- 
gerous predators,  both  avian  and  mammalian, 
to  the  location  of  the  brood  at  the  very  time 
the  parents  are  busy  elsewhere  and  leave  it 
undefended.  Second,  more  frequent  alarm  and 
defense  behaviors  carry  an  energetic  cost  (re- 
viewed in  Walters  1984).  Finally,  it  is  possible 
that  the  abundance  of  crabs  on  the  backshore 
induces  the  adult  Piping  Plovers  to  shepherd 
their  broods  away  from  the  foreshore,  where 
forage  might  be  more  abundant  and  have  a 
higher  water  content.  Broods  elsewhere  on 
Assateague  Island  are  routinely  taken  to  the 
foreshore  (refuge  staff,  pers.  comm.). 

Given  the  minimal  direct  threat  posed  by 
ghost  crabs,  and  the  potential  negative  con- 
sequences of  frequent  display  and  restricted 
foraging,  it  seems  maladaptive  on  the  part  of 
the  adult  Piping  Plovers  to  treat  crabs  as  dan- 
gerous predators.  Natural  selection  acting  on 
adult  defensive  behavior  should  have  elimi- 


nated the  behavior  if  ghost  crabs  are  not  sig- 
nihcant  predators  and  if  engaging  in  defense 
towards  ghost  crabs  increases  the  likelihood 
of  predation  by  other  predators.  However,  lack 
of  sufficient  genetic  isolation  between  birds 
breeding  in  areas  with  and  without  ghost  crabs 
would  preclude  such  selection.  Instead,  some 
Charadriidae  appear  to  recognize  several  ani- 
mal categories,  including  avian,  mammalian 
(with  a subset  of  ungulate),  reptilian  and  “oth- 
er”. They  have  evolved  unique  displays  to 
each  group  (reviewed  in  Gochfield  1984). 
They  generally  distinguish  potential  predators 
from  non-predators,  especially  among  birds, 
thereby  minimizing  false  alarms  (Walters 
1990).  However,  they  seem  less  discriminat- 
ing about  other  intruders,  lumping  disparate 
taxonomic  groups  into  a category  of  “poten- 
tial threat”  in  an  “urgency  of  response 
scheme”  (Walters  1990).  Perhaps  Piping  Plo- 
vers indiscriminately  categorize  anything  ter- 
restrial but  “neither  a large  mammal  nor  a 
snake”,  and  moving  near  a nest  or  near 
chicks,  as  requiring  immediate  alarm.  This 
group  might  include  dangerous  predators  such 
as  rats  or  mustelids  to  which  immediate  alarm 
would  be  adaptive.  If  one  postulates  that  or- 
ganisms such  as  crabs  and  turtles,  which  do 
not  pose  a significant  threat  but  which  do  elic- 
it alarm  responses  from  parenting  plovers  (ref- 
uge staff,  pers.  obs.),  are  lumped  into  the  same 
“dangerous  predator”  category  in  the  alarm 
response  hierarchy,  the  apparently  maladap- 
tive alarming  by  Piping  Plover  parents  could 
be  explained.  Given  the  large  geographic 
range  and  variety  in  breeding  habitat,  with 
concomitant  and  unpredictable  variation  in  the 
suite  of  predators,  mounting  a defensive  dis- 
play against  anything  novel  in  the  area  of  the 
brood  might  have  at  least  neutral  if  not  ben- 
eficial effects  on  fitness. 

Correlations  betM’een  low  fledging  success 
and  high  ghost  crab  abundance. — We  hypoth- 
esize that  high  ghost  crab  abundance  and  low 
fledging  success  of  Piping  Plovers  have  a cor- 
relative, not  causative,  relationship.  Three  fac- 
tors contribute  to  the  correlation:  beach  and 
dune  morphology,  climate,  and  parental  be- 
havior. 

Ghost  crabs  are  most  abundant  on  high  en- 
ergy beaches  backed  by  high  dunes.  The  high 
dunes  provide  overwintering  habitat  in  which 
crabs  are  able  to  burrow  below  their  lethal  iso- 


328 


THE  WILSON  BULLETIN  • VoL  III,  No.  3.  September  1999 


therm  (6—8°  C)  before  being  blocked  by  the 
water  table  (T.G.W.,  unpubl.  data),  while  high- 
energy  beaches  provide  habitat  for  the  crabs’ 
preferred  prey  (Wolcott  1978).  However, 
beaches  that  are  backed  by  dunes  and  that  lack 
low-lying  moist  habitat  away  from  the  surf, 
may  be  poorer  habitat  for  raising  Piping  Plo- 
ver broods,  even  in  the  absence  of  crabs.  On 
Assateague  Island,  slower  growth  with  con- 
comitant reduced  survival  has  been  docu- 
mented for  chicks  reared  on  an  ocean  beach, 
compared  with  chicks  from  other  areas  with 
low-lying  moist  habitat  for  foraging  (Loeger- 
ing  and  Fraser  1995).  Prey  abundance  (mea- 
sured in  the  wrack  zone)  and  foraging  rates 
were  lower  on  the  ocean  beach  as  well.  Wild 
Beach  lacks  the  low-lying  moist  areas  that  are 
the  major  foraging  habitat  for  young  plover 
broods  in  the  more  productive  breeding  areas 
on  Assateague.  Even  the  steep  intertidal  with 
its  sharp  escarpments  may  pose  a physical 
barrier  to  young  broods  moving  to  the  fore- 
shore. 

During  the  same  time  as  our  study,  over 
90%  of  chicks  successfully  fledged  on  a site 
approximately  400  m inland  from  our  site, 
with  extensive  moist  low-lying  forage  but  no 
access  to  the  beach.  Similarly  91%  fledging 
success  occurred  on  a low  energy  beach  at  the 
southern  tip  of  Assateague  Island,  with  exten- 
sive backshore  foraging  areas.  Meanwhile  less 
than  10%  of  chicks  fledged  on  Wild  Beach,  a 
percentage  similar  to  its  long-term  average 
(1988-1994,  U.S.  Fish  and  Wildlife  1994). 
Poor  forage  is  more  likely  to  contribute  to  the 
unusually  high  chick  mortality  on  Wild  Beach 
than  direct  ghost  crab  predation. 

Climate  and  latitude  probably  play  a role  in 
the  plover-crab  relationship.  From  their  north- 
ern limit  in  New  Jersey  through  the  southern 
limit  of  plover  breeding  in  North  Carolina, 
ghost  crabs  increase  in  size  and  abundance 
(U.S.  Fish  and  Wildlife  Service  1996).  Pre- 
sumably, at  lower  latitudes  more  adult  crabs 
are  able  to  successfully  overwinter,  emerge 
earlier  in  the  spring,  have  a longer  active  sea- 
son, and  grow  to  a larger  adult  size.  Higher 
abundance  of  large  crabs  leads  to  more  en- 
counters between  birds  and  crabs,  with  an  in- 
crease in  adverse  indirect  effects  on  Piping 
Plovers. 

The  hotter  summer  temperatures  associated 
with  lower  latitudes  may  directly  affect  brood 


survival  of  Piping  Plovers.  High  daytime  tem- 
peratures may  speed  the  rate  of  embryo  de- 
velopment and  lead  to  greater  hatching  asyn- 
chrony (reviewed  in  Magrath  1990,  Shields 
1998).  Asynchrony  may  contribute  to  mortal- 
ity directly  by  causing  abandonment  of  viable, 
late  eggs,  and  indirectly,  by  reducing  the  co- 
operative attendance  by  adults  of  the  hatched 
young.  In  our  study,  severe  asynchrony  also 
resulted  in  broods  having  chicks  of  very  dif- 
ferent locomotor  capabilities,  which  could  in- 
crease the  likelihood  that  chicks  will  become 
separated  from  one  another  and  lost. 

Desiccation  poses  an  even  more  immediate 
danger  to  the  chicks.  Piping  Plovers  acquire 
water  from  their  food  and  thermoregulate  by 
panting  (Haig  1992).  During  the  critical  first 
day,  broods  hatching  in  hot  weather  during 
our  study  showed  a reduction  in  foraging 
time  and  activity  compared  to  broods  hatch- 
ing in  cooler  weather.  Should  heat  cause  a 
persistent  shift  in  activity,  it  implies  that  dur- 
ing hot  weather  intake  of  food  and  its  in- 
cluded water  is  lowered  at  the  same  time  that 
evaporative  losses  become  greater.  Under- 
standing the  relationship  between  elevated 
temperatures,  asynchrony,  and  brood  survival 
is  critical  to  informed  management  decisions 
at  the  southern  end  of  the  Piping  Plover’s 
breeding  range. 

The  behavior  of  the  adult  plovers  to  the  per- 
ceived threat  of  the  crabs  compounds  the 
problem  of  desiccation  because  the  adults  ap- 
parently restrict  their  broods  to  the  duneline, 
where  forage  is  sparser  and  drier.  On  longer 
time  scales,  restricted  feeding  times  and  poor 
forage  will  result  in  slow  growth  and  in- 
creased mortality  (Loegering  and  Fraser 
1995).  The  threat  perceived  by  adult  plovers 
of  ghost  crabs  on  the  Wild  Beach  may  act  as 
a barrier  between  broods  and  the  richer  food 
resources  of  the  foreshore. 

ACKNOWLEDGMENTS 

Financial  support  for  the  study  came  from  the  U.S. 
Biological  Survey.  Logistical  support  was  provided  by 
Chincoteague  National  Wildlife  Refuge  and  Virginia 
Department  of  Game  and  Inland  Fisheries.  Special 
thanks  to  A.  Hecht,  J.  Schrocr  and  his  staff,  R.  Cross, 
K.  Terwilliger  for  help  with  Scotch-liting®  chicks,  D. 
Turner  for  donating  quail  eggs,  and  J.  Walters  for  help- 
ful comments  on  the  manuscript. 


Wolcoit  and  Wolcott  • PIPING  PLOVERS  AND  GHOST  CRABS 


329 


LITERATURE  CITED 

Arndt,  R.  G.  1994.  Predation  on  hatchling  diamond- 
back  temipin,  Mulaclemys  terrapin  (Schoepff),  by 
the  ghost  crab.  Ocypode  qiiadrata  (Fabriciiis)  II. 
Fla.  Sci.  57:1-5. 

Britton,  E.  E.  1979.  Evaluation  of  public  use  impacts 
upon  nesting  shorebirds  and  the  beach  habitat  on 
Chincoteague  National  Wildlife  Refuge.  U.S.  Fish 
and  Wildlife  Service,  Chincoteague  National 
Wildlife  Refuge,  Chincoteague,  Virginia. 

Chace,  F.  a.  and  H.  H.  Hobbs,  Jr.  1969.  The  fresh- 
water and  terrestrial  crabs  of  the  West  Indies  with 
special  reference  to  Dominica.  U.S.  Natl.  Mus. 
Nat.  Hist.  Bull.  292:1-258. 

Cowles,  R.  P.  1908.  Habits,  reactions,  and  associations 
in  Ocypode  arenaria.  Pap.  Carnegie  Inst.  Wash. 
2:3-41. 

Evans,  P.  R.  and  M.  W.  Pienkowski.  1984.  Population 
dynamics  of  shorebirds.  Behav.  Mar.  Anim.  5:91. 

Evans,  S.  M.,  A.  Cram,  K.  Eaton,  R.  Torrance,  and 
V.  Wood.  1976.  Foraging  and  agonistic  behavior 
in  the  ghost  crab  Ocypode  kuhlii  de  Haan.  Mar. 
Behav.  Physiol.  4:121-135. 

Full,  R.  J.  and  R.  B.  Weinstein.  1992.  Integrating  the 
physiology,  mechanics  and  behavior  of  rapid  run- 
ning ghost  crabs:  slow  and  steady  doesn’t  always 
win  the  race.  Am.  Zool.  32:382-395. 

Gochfield,  M.  1984.  Antipredator  behavior:  aggres- 
sive and  distraction  displays  of  shorebirds.  Behav. 
Mar.  Anim.  5:289-377. 

Haig,  S.  M.  1992.  Piping  Plover.  In  The  birds  of  North 
America,  no.  2 (A.  Poole,  P.  Stettenheim,  and  F. 
Gill,  Eds.).  The  Academy  of  Natural  Sciences, 
Philadelphia,  Pennsylvania;  The  American  Orni- 
thologists’ Union,  Washington,  D.C. 

Herreid,  C.  F.  1963.  Observations  on  the  feeding  be- 
havior of  Cardisoma  guanhumi  (Latreille)  in 
southern  Florida.  Crustaceana  5:176-180. 

Loegering,  J.  P.  1992.  Piping  Plover  breeding  biology, 
foraging  ecology  and  behavior  on  Assateague  Is- 
land National  Seashore,  Maryland.  M.S.  thesis, 
Virginia  Polytechnic  Institute  and  State  Univ., 
Blacksburg. 

Loegering,  J.  P.  and  J.  D.  Fraser.  1995.  Factors  af- 
fecting Piping  Plover  chick  survival  in  different 
brood-rearing  habitats.  J.  Wildl.  Manage.  59:646- 
655. 

Loegering,  J.  R,  J.  D.  Fraser,  and  L.  L.  Loegering. 
1995.  Ghost  crab  preys  on  a Piping  Plover  chick. 
Wilson  Bull.  10:768-769. 

MacIvor,  L.  H.  1990.  Population  dynamics,  breeding 
ecology,  and  management  of  Piping  Plovers  on 


outer  Cape  Cod,  Massachu.setts.  M.S.  thesis,  Univ. 
of  Massachusetts,  Amher.st. 

MacIvor,  L.  H.,  S.  M.  Melvin,  and  C.  R.  Griffin. 
1990.  Etlects  of  research  activity  on  Piping  Plover 
nest  predation.  J.  Wildl.  Manage.  54:443-447. 

Magrath,  R.  D.  1990.  Hatching  asynchrony  in  altri- 
cial  birds.  Biol.  Rev.  65:587-622. 

Patterson,  M.  E.  1988.  Piping  Plover  breeding  biol- 
ogy and  reproductive  success  on  A.s.sateague  Is- 
land. M.S.  thesis,  Virginia  Polytechnic  Institute 
and  State  Univ.,  Blacksburg. 

Robertson,  J.  R.  and  W.  J.  Pfeiffer.  1982.  Deposit 
feeding  by  the  ghost  crab  Ocypode  quad  rata  (Fa- 
bricius).  J.  Exp.  Mar.  Biol.  Ecol.  56:165-177. 

Shields,  M.  A.  1998.  Proximate  and  ultimate  causes 
of  hatching  asynchrony  in  the  Brown  Pelican. 
Ph.D.  diss..  North  Carolina  State  Univ.,  Raleigh. 

Staine,  K.  j.  and  j.  Burger.  1994.  Nocturnal  foraging 
behavior  of  breeding  Piping  Plovers  (Charadrius 
melodus)  in  New  Jersey.  Auk  111:579-587. 

U.S.  Fish  and  Wildlife  Service.  1993.  Piping  Plover 
monitoring  and  management,  summer  1993.  U.S. 
Fish  and  Wildlife  Service,  Chincoteague  NWR, 
Chincoteague,  Virginia. 

U.S.  Fish  and  Wildlife  Service.  1994.  Piping  Plover 
monitoring  and  management-summer  1994.  U.S. 
Fish  and  Wildlife  Service,  Chincoteague  NWR, 
Chincoteague,  Virginia. 

U.S.  Fish  and  Wildlife  Service.  1996.  Piping  Plover 
{Charadrius  melodus)  Atlantic  coast  population 
revised  recovery  plan.  U.S.  Fish  and  Wildlife  Ser- 
vice, Hadley,  Virginia. 

Walters,  J.  R.  1982.  Parental  behavior  in  Lapwings 
(Charadriidae)  and  its  relationships  with  clutch 
sizes  and  mating  systems.  Evolution  36:1030- 
1040. 

Walters,  J.  R.  1984.  The  evolution  of  parental  be- 
havior and  clutch  size  in  shorebirds.  Behav.  Mar. 
Anim.  5:234-287. 

Walters,  J.  R.  1990.  Antipredatory  behavior  of  Lap- 
wings: field  evidence  of  discriminative  abilities. 
Wilson  Bull.  102:49-70 

Watts,  B.  D.  and  D.  S.  Bradshaw.  1995.  Ghost  crab 
preys  on  Piping  Plover  eggs.  Wilson  Bull.  107: 
767-768. 

Wellins,  C.  a.,  D.  Rittschof,  and  M.  Wachowiak. 
1989.  Location  of  volatile  odor  sources  by  ghost 
crab  Ocypode  quadrata  (Fabriciiis).  J.  Chem. 
Ecol.  15:1161-1169. 

Wolcott,  T.  G.  1978.  Ecological  role  of  ghost  crabs, 
Ocypode  quadrata  (Fab.)  on  an  ocean  beach: 
scavengers  or  predators?  J.  Exp.  Mar.  Biol.  Ecol. 
3 1 :67-82. 


Wilson  Bull.,  111(3),  1999,  pp.  330-339 


A TAXONOMIC  STUDY  OF  CRESTED  CARACARAS 

(FALCONIDAE) 

CARLA  J.  DOVE'  2^  AND  RICHARD  C.  BANKS^ 


ABSTRACT. — The  taxonomic  status  of  the  crested  caracaras  (Caracara  spp.,  Falconidae)  has  been  unsettled 
for  many  years.  Current  sources  such  as  the  AOU  Check-list  recognize  a single  species  that  includes  three  taxa 
formerly  considered  distinct,  citing  observations  by  Hellmayr  and  Conover  (1949)  on  two  specimens  considered 
to  be  intermediate.  We  studied  plumage  characters  and  measurements  of  over  392  museum  specimens  and  found 
no  evidence  of  clinal  change  between  the  northern  and  southern  continental  populations.  Sixteen  specimens  from 
localities  near  the  Amazon  River  where  these  two  populations  sporadically  meet  exhibit  a mosaic  of  plumage 
elements  from  both  forms.  Measurements  of  wing  chord,  bill  length,  and  bill  depth  indicate  that  size  is  positively 
correlated  with  latitude  north  and  south  of  the  equator  and  that  females  are  larger  than  males  in  the  northern 
population.  These  populations  do  not  meet  in  western  South  America.  We  conclude  that  three  biological  species 
can  be  identified  in  the  crested  caracaras;  the  insular  Guadalupe  Caracara  {Caracara  hitosus)-,  and  two  continental 
species.  Northern  (C.  clieriway)  and  Southern  caracara  (C.  plancus),  neither  of  which  shows  subspecific  variation. 
Received  6 Oct.  1998,  accepted  16  Feb.  1999. 


The  Florida  population  of  the  Crested  Ca- 
racara (Caracara  plancus  audubonii)  is  con- 
sidered threatened  by  the  U.S.  Fish  and  Wild- 
life Service  (1987),  a recognition  that  subjects 
that  population  to  strict  permit  regulations  and 
consideration  for  conservation  efforts  (Mor- 
rison 1996).  Although  the  generic  name  given 
in  the  Fish  and  Wildlife  Service  listing  is  Po- 
lyborus.  Banks  and  Dove  (1992)  have  shown 
that  the  generic  name  should  be  Caracara. 
The  threatened  status,  together  with  the  un- 
certainty reflected  by  various  taxonomic  treat- 
ments at  the  species  and  subspecies  levels,  has 
led  us  to  complement  Morrison’s  (1996)  de- 
tailed report  by  examining  the  taxonomy  of 
crested  caracaras  throughout  their  range. 

Ridgway  (1876)  treated  the  crested  caraca- 
ras as  three  species  in  the  genus  Polyborus 
(now  Caracara)'.  tharus  Molina,  1782  of 
southern  South  America;  chertway  Jacquin, 
1784  from  northern  South  America  to  south- 
ern North  America;  and  lutosus  Ridgway, 
1876  of  Guadalupe  Island,  Mexico.  Except  for 
the  replacement  of  the  specific  name  tharus 
with  the  earlier  plancus  Miller,  1777  by  Bra- 


' Dept,  of  Biology,  George  Mason  Univ.,  Fairfax, 
VA  22030. 

2 Present  address:  Dept,  of  Vertebrate  Zoology,  Na- 
tional Museum  of  Natural  History,  Smithsonian  Insti- 
tution, Washington,  DC  20560-01  16. 

^ U.S.  Geological  Survey,  Patuxent  Wildlife  Re- 
search Center,  National  Museum  of  Natural  History, 
Washington.  DC  20560-01  1 1. 

* Corresponding  author; 

E-mail;  dovec@nmnh.si.edu 


bourne  and  Chubb  (1912),  an  action  followed 
by  Swann  (1925)  and  all  subsequent  authors, 
this  treatment  remained  unchanged  for  three- 
quarters  of  a century.  Swann  (1925)  thought 
that  birds  from  the  northern  part  of  the  range 
of  C.  plancus  in  Brazil  were  “more  or  less 
intermediate  in  appearance”  between  more 
southerly  C.  plancus  and  C.  clieriway,  but 
gave  no  details.  Hellmayr  and  Conover 
(1949),  stating  that  C.  clieriway  “appears  to 
us  nothing  else  but  a well-marked  race  of  the 
Southern  Caracara,”  were  the  first  to  unite  the 
two  continental  populations  into  a single  spe- 
cies, C.  plancus.  The  chief  distinguishing 
characters  of  blacker  coloration  and  reduction 
of  white  barring  both  on  rump  and  chest  were 
considered  by  Hellmayr  and  Conover  (1949) 
as  “merely  differences  of  degree.”  They  also 
noted  that  the  apparent  gap  in  measurements 
between  extreme  southern  C.  plancus  (Strait 
of  Magellan)  and  C.  clieriway  was  bridged  by 
specimens  from  intermediate  localities.  Hell- 
mayr and  Conover  (1949)  admitted  that  they 
had  seen  C.  clieriway  only  from  north  of  the 
Amazon  River  but  cited  records  of  that  form 
from  Santarem  and  Rhomes,  south  of  the  riv- 
er. At  the  same  time,  they  reported  that  spec- 
imens from  Marajo  and  Mexiana  Islands,  in 
the  mouth  of  the  Amazon,  belonged  to  C. 
plancus.  The  only  actual  evidence  of  inter- 
mediacy mentioned  by  Hellmayr  and  Conover 
(1949:283-284)  was  based  on  two  adult  birds 
from  Obidos,  Brazil,  classified  as  C.  clieriway, 
“which,  by  more  heavily  barred  lateral  upper 


330 


Dove  and  Banks  • TAXONOMY  OF  CRESTED  CARACARAS 


33] 


TABLE 

1.  Plumage  characters  of  Ca 

racara  cheriway  and  C.  plancus. 

Character — area 

C.  cheriway 

C.  plancus 

1 Breast 

dark  spots  or  wedge-shaped 
bars,  heavier  posteriorly 

dark  and  light  bars  over  entire  breast  area 

2 Vent  area 

pale  patch  between  thighs 

dark  feathers  between  thighs 

3 Upper  back/Scapulars 

pale  wedge-shaped  patch 
with  broad  black  bars, 
scapulars  always  black 

finely  barred,  no  wedge-shaped  patch,  scapu- 
lars usually  heavily  barred,  only  lightly  in 
n.  part  of  range 

4 Lower  back 

black 

barring  continuous  with  upper  back  and  tail 

5 Upper  tail  coverts 

white  or  faintly  barred 

barred 

tail  coverts,  mark  a decided  step  in  the  direc- 
tion of  plancus.” 

Friedmann  and  coworkers  (1950)  and  the 
American  Ornithologists’  Union  (AOU  1957) 
retained  C.  cheriway  as  a species.  Wetmore 
(1965)  paraphrased  Swann  (1925)  in  men- 
tioning intergradation  south  of  the  Amazon 
and  accepted  “the  present-day  tendency  to 
unite”  the  two  mainland  forms.  Brown  and 
Amadon  (1968)  considered  C.  plancus  and  C. 
cheriway  conspecific  as  did  Blake  (1977), 
who  briefly  noted  “evidence  of  intergrada- 
tion in  Brazil.”  Vuilleumier  (1970)  was  the 
first  to  treat  the  insular  C.  lutosus  as  a well- 
marked  subspecies  of  the  combined  C.  plan- 
cus, apparently  following  the  suggestion  of 
Brown  and  Amadon  (1968:736)  that  if  cher- 
iway and  plancus  were  combined,  then  “it 
would  be  no  great  extension  to  include  lu- 
tosus” Stresemann  and  Amadon  (1979)  and 
Sibley  and  Monroe  (1990)  merged  the  con- 
tinental forms  but  retained  C.  lutosus  as  an 
allospecies.  The  AOU  (1983,  1998)  merged 
all  the  taxa,  noting  the  report  of  intergrada- 
tion “near  the  mouth  of  the  Amazon.”  Thus, 
the  merger  of  C.  plancus  and  C.  cheriway  is 
based  mainly  on  Hellmayr’s  and  Conover’s 
(1949:283)  statement  that  color  characters 
were  “merely  differences  of  degree”,  and  on 
the  two  specimens  of  C.  cheriway  from  Ob- 
idos,  Brazil,  that  they  stated  showed  a 
“marked  step”  in  the  direction  of  C.  plancus. 
The  degree  of  intergradation  or  variation  has 
never  been  thoroughly  examined  in  the  lim- 
ited geographical  area  where  the  northern  and 
southern  populations  meet,  and  the  distinc- 
tiveness of  C.  lutosus  has  not  been  reviewed 
since  Friedmann  (1950). 


METHODS 

To  reevaluate  the  taxonomic  relationships  of  the 
crested  caracaras  at  the  species  level,  we  compared  five 
plumage  characters  and  three  body  measurements  of 
specimens  from  all  continental  geographic  popula- 
tions. All  plumage  color  comparisons  were  made  under 
museum  Examolites®  or  in  daylight.  We  took  mea- 
surements on  more  than  392  specimens  of  wing  chord, 
bill  depth  (at  the  cere),  bill  length  (from  the  base  of 
the  bill  to  the  tip),  and  tarsus  of  sexed  adult  birds  from 
all  parts  of  the  species’  range.  Tarsal  measurements 
were  extremely  variable  within  populations  and  were 
deemed  not  useful  for  comparison.  Measurements  of 
males  and  females  were  analyzed  separately  to  deter- 
mine sexual  variation. 

We  used  an  early  version  of  Table  1 to  categorize 
plumage  characters  in  specimens  examined,  and  we 
asked  colleagues  at  some  other  museums  (see  ac- 
knowledgments) to  use  that  and  xeroxed  photographs 
of  Fig.  1 to  evaluate  specimens  for  us,  thereby  avoid- 
ing the  need  of  extensive  loans  of  these  large  birds. 
Voice  recordings  and  tissue  samples  of  this  species  are 
insufficient  for  accurate  analysis  at  this  time. 

Specimens  from  the  northern  populations  were  com- 
pared with  those  from  Bolivia  and  southern  Brazil 
south  to  Tierra  del  Fuego.  Twenty-one  specimens  from 
localities  in  Brazil  along  the  Amazon  River  and  the 
northeastern  coast  of  Brazil  are  from  the  contact  zone 
of  the  two  populations  and  were  examined  separately 
for  plumage  patterns.  These  were  compared  with 
northern  and  southern  populations  to  determine  the  ex- 
tent of  variation  within  this  region.  Because  this  spe- 
cies is  not  sexually  dichromatic  in  plumage  color  or 
pattern,  adult  specimens  (definitive  plumage)  were  not 
separated  for  plumage  compari.sons. 

At  the  subspecies  level  we  examined  the  purported 
characters  of  recently  recognized  taxa  (Hellmayr  and 
Conover  1949,  Peters  1931),  all  of  which  were  de- 
scribed long  ago  from  the  periphery  of  the  range  of 
the  species. 

South  American  collecting  localities  were  confirmed 
using  Stevens  and  Traylor  (1983),  Paynter  (1988),  and 
Paynter  and  Traylor  (1991).  Descriptive  statistics,  AN- 


332 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


FIG.  1 Typical  adult  plumage,  ventral  (above)  and  dorsal  (below)  views  of  (top)  Cciracara  cheriway  (USNM 
I 32101 . Sonora,  Mexico),  (middle)  specimen  from  zone  ol  contact  (USNM  276906,  Maica,  Brazil),  and  (bottom), 
C.  pUmcus  (USNM  284790,  Argentina). 


Dove  ami  Banks  • TAXONOMY  OF  CRESTED  CARACARAS 


333 


OVAs,  and  /-tests  were  done  with  SYSTAT  (1992) 
version  5.0  for  Windows. 

RESULTS 

Specific  variation. — Typical  cidult  crested 
caracaras  from  south  (C.  plancus)  and  north 
(C.  cheriway)  of  the  Amazon  have  clearly  dis- 
tinct color  patterns  (Fig.  1 ) with  major  differ- 
ences in  all  five  plumage  characters.  The 
plumage  patterns  of  specimens  from  the 
northern  and  southern  portions  of  the  range  do 
not  overlap,  and  specimens  from  these  popu- 
lations can  be  identified  and  allocated  un- 
equivocally to  either  C.  cheriway  or  C.  plan- 
cus on  that  basis.  All  adult  birds  from  northern 
Brazil  (Roraima),  Venezuela,  Colombia,  Ec- 
uador, and  Peru  (Rio  Chinchipe,  Pacasmayo, 
and  Catacaos)  north  to  the  United  States  (ex- 
cept Guadalupe  Island)  are  of  the  C.  cheriway 
type.  Birds  from  southern  Brazil  (Bahia)  and 
Bolivia  to  Tierra  del  Fuego  are  of  the  C plan- 
cus type.  In  addition  to  the  plumage  differ- 
ences evaluated  (Table  1),  the  undertail  co- 
verts of  C.  cheriway  are  either  all  white,  or 
faintly  or  incompletely  barred,  giving  the  base 
of  the  tail  a whitish  appearance,  whereas  the 
undertail  coverts  of  C.  plancus  are  finely 
barred  with  distinct  wide  white  and  narrow 
dark  bars  that  extend  completely  across  the 
feather.  The  base  of  the  tail  is  more  white  than 
barred  on  C.  cheriway  specimens,  but  this 
character  does  not  always  hold  true  because 
some  C.  cheriway  specimens  have  barring 
near  the  base  of  the  tail. 

Plumage  patterns  of  specimens  from  the 
zone  of  contact. — The  apparent  zone  of  sec- 
ondary contact  for  the  two  populations  ex- 
tends from  the  mouth  of  the  Amazon  River 
westward  along  the  river  and  its  southern  trib- 
utaries to  the  Rio  Tapajos  (Fig.  2).  It  may  ex- 
tend west  as  far  as  the  Rio  Purus  (Canutama), 
where  two  immature  specimens  (Field  Muse- 
um of  Natural  History;  FMNH  100805, 
100806)  having  mosaic  plumage  tendencies 
were  collected.  A bird  from  farther  west  on 
the  Rio  Jurua  (Museum  of  Comparative  Zo- 
ology; MCZ  173161)  is  pure  C.  plancus  and 
was  not  considered  to  represent  the  contact 
zone.  The  zone  extends  to  the  southeast  from 
the  mouth  of  the  Amazon  to  approximately 
Morros  de  Mariana.  The  latest  specimen  col- 
lected from  the  contact  zone  that  we  found 
was  dated  1937. 


In  two  adult  specimens  (American  Museum 
of  Natural  History;  AMNH  241501,  241502) 
from  well  to  the  south  at  Remanso,  on  the  Rio 
Sao  Francisco,  the  plumage  on  the  upper 
back/scapular  area  is  similar  to  C.  cheriway, 
but  the  lower  back  area  is  intermediate  in  col- 
or between  C.  cheriway  and  C.  plancus.  The 
breast  of  a bird  (Naturhistorisches  Museum, 
Vienna;  NMWZ  39885)  from  Juazeiro,  Bahia, 
was  scored  as  C.  cheriway.  Specimens  from 
other  nearby  Bahia  localities  (AMNH  163138, 
Salvador;  The  Natural  History  Museum; 
BMNH  73.3.19.4,  Ilha  de  Itaparica;  NMWZ 
39884,  Lago  de  Pamagua;  NMWZ  39886, 
Barra)  all  are  pure  C.  plancus.  The  appearance 
of  these  few  c/zer/way- like  characters  well  to 
the  south  may  indicate  a much  wider  zone  of 
intergradation  than  we  recognize,  but  the  lack 
of  available  specimens  from  the  intervening 
500  km  makes  this  conclusion  problematical. 
This  observation  may  merely  be  a reflection 
of  the  generally  more  variable  plumage  in  the 
southern  population.  Gyldenstolpe  (1951)  also 
has  commented  on  the  lack  of  comparative 
material  from  parts  of  the  range  of  this  spe- 
cies. 

Hellmayr  and  Conover  (1949)  mentioned 
birds  from  two  other  localities  (Santarem  and 
Rhomes)  in  the  zone  and  we  accepted  their 
determination  of  those  birds  as  cheriway  (Ta- 
ble 2).  We  found  no  gradual  intergradation  in 
plumage  characters  but  instead  an  abrupt  shift 
in  plumage  type.  There  was  no  consistent  pat- 
tern of  intermediacy  but  rather  a mosaic  of 
plumage  combinations.  A summary  of  the 
plumage  data  shows  that  54%  of  the  115  in- 
dividual character  states  were  like  C.  plancus, 
40%  were  like  C.  cheriway,  and  6%  were  in- 
termediate. The  lower  back  (character  4)  re- 
ceived 4 of  the  6 intermediate  ratings.  Speci- 
mens rated  intermediate  in  this  character  have 
dark  backs  with  white-tipped,  not  barred, 
feathers  in  the  mid-  to  lower  back  region.  No 
specimen  was  intermediate  in  more  than  one 
character.  Seven  of  the  birds  from  the  contact 
zone  were  rated  as  pure  for  one  or  the  other 
species. 

The  two  specimens  from  Obidos  that  were 
the  basis  for  the  original  merger  of  the  species 
do  not  agree  in  plumage  patterns.  FMNH 
101538  is  typical  C.  cheriway,  whereas 
FMNH  101539  has  the  breast  and  heavily 
barred  tail  pattern  typical  of  C.  plancus  and 


ATLANTIC  OCEAN 


334 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


OJ 


FIG.  2.  Map  of  localities  in  zone  of  contact  with  icons  corresponding  to  character  states  listed  in  Table  2. 


Dove  ami  Banks  • TAXONOMY  OF  CRESTED  CARACARAS 


335 


TABLE  2.  Summary  of  plumage  characters  (numbers  from  Table  I ) of  specimens  from  the  zone  ol'  contact 
between  northern  anti  southern  continental  populations  of  crested  caracaras.  Localities  without  specimen  numbers 
are  included  on  the  basis  of  records  mentioned  by  Hellmayr  and  Conover  (1949).  C = C.  cheriwav,  P = C. 
planciis.  I = intermediate,  ? = unscored. 


Character 


Locality 

Museum  tt 

1 

2 

3 

4 

5 

(Dbidos 

FMNH  101538 

C 

c 

c 

c 

c 

FMNH  101539 

P 

p 

c 

c 

p 

BMNH  1908.8.21.17 

P 

c 

c 

c 

c 

Igarpe  Arriba 

FMNH  101424 

P 

p 

c 

p 

p 

FMNH  101425 

P 

p 

c 

p 

p 

FMNH  101427 

P 

p 

c 

1 

p 

Diamantina  Parintins 

USNM  121077 

I 

p 

c 

p 

p 

AMNH  276706 

P 

p 

c 

1 

p 

AMNH  277572 

P 

c 

c 

1 

p 

AMNH  277573 

P 

c 

c 

1 

p 

Maica  Ituqui 

USNM  276906 

P 

p 

c 

c 

p 

FMNH  101157 

c 

p 

c 

p 

p 

FMNH  101158 

c 

c 

c 

c 

c 

Agarape  Brabo  (Rio  Tapajos) 

AMNH  285747 

p 

p 

c 

c 

c 

Humberto  do  Campos  (Maranho) 

FMNH  100401 

p 

p 

p 

p 

p 

Santo  Amaro  Maranhao 

MCZ  92682 

I 

7 

p 

p 

p 

Morros  de  Mariana 

AMNH  241499 

p 

c 

p 

p 

p 

AMNH  241500 

p 

c 

p 

p 

p 

Ilha  Mexiana 

BMNH  73.3.19.4 

p 

p 

p 

p 

p 

Ilha  de  Marajo 

MCZ  22996 

p 

I 

c 

c 

p 

Ilha  Caviana 

UM  7504 

p 

p 

p 

p 

p 

Santarem 

c 

c 

c 

c 

c 

Rhomes 

c 

c 

c 

c 

c 

the  back  pattern  of  C.  cheriway  with  the 
wedge-shaped  barring  on  the  upper  back  and 
completely  black  mid-  and  lower  back.  An  ad- 
ditional specimen  from  Obidos  (BMNH 
1908.8.21.17)  is  typical  of  C.  cheriway  except 
for  the  barring  on  the  breast.  Our  efforts  to 
locate  the  two  other  specimens  from  Obidos 
mentioned  by  Hellmayr  and  Conover  (1949: 
284)  were  unsuccessful. 

Size  variation. — We  divided  mensural  data 
from  the  continental  populations  into  four 
geographic  sets  as  indicated  in  Table  3.  A one- 
way analysis  of  variance  shows  significant 
differences  among  the  four  continental  groups 
in  wing  length  (F  = 27.8,  P < 0.01,  n = 336), 
bill  length  (F  = 3.4,  P < 0.02,  n = 376),  and 
bill  depth  (F  = 4.2,  P < 0.007,  n = 391),  but 
body  size,  as  estimated  by  wing  chord  mea- 
surements, was  positively  correlated  with  lat- 
itude (Fig.  3).  Although  individual  variation 
in  body  size  is  extensive  throughout  the  range, 
specimens  from  the  extreme  southern  parts  of 
South  America  are  the  largest.  Northern  and 


southern  populations  were  analyzed  separately 
for  sexual  size  dimorphism,  f-tests  of  northern 
and  southern  populations  revealed  a signifi- 
cant sexual  size  difference  in  the  northern 
population,  with  females  being  larger  (wing 
length,  t = —5.50,  n = 243;  bill  length,  t = 
-5.96,  n = 21 U bill  depth,  t = -6.93,  n = 
277;  P < 0.001  in  all  measurements)  but  no 
difference  in  the  southern  populations  (wing 
length,  t = —1.37,  n = 77;  bill  length,  t = 
-0.65,  n = 84;  bill  depth,  t = -0.84,  n = 92; 
P > 0.05  in  all  measurements).  That  the  os- 
tensible gap  in  measurements  between  C. 
cheriway  and  extreme  southern  birds  was 
bridged  by  specimens  from  intermediate  lo- 
calities, was  used  by  Hellmayr  and  Conover 
(1949)  as  an  argument  to  justify  conspecific 
treatment.  Our  data  (Table  3,  Fig.  3)  show  that 
these  caracaras  vary  clinally  within  each 
hemisphere,  and  become  smaller  toward  the 
equator.  This  is  true  for  wing  chord,  bill 
length,  and  bill  depth. 


336 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


TABLE  3.  Descriptive  statistics  for  four  geographic 
dalupe  Island  caracara. 

groups  of  continental  crested  caracaras  and  the  Gua- 

Wing 

(mm) 

Bill  length 
(mm) 

Bill  depth 
(mm) 

US,  Cuba,  Mexico 

d /?“  = 85 

90 

92 

Rb  = 343-409 

28.6-36.7 

16.2-19.9 

x‘^  = 382.51 

32.65 

17.70 

SD'*  = 13.67 

1.47 

0.75 

9 n = 19 

84 

86 

R = 363-415 

29.8-38.7 

16.5-22.0 

X = 391.27 

33.60 

18.3 

SD  = 1 1.35 

1.44 

0.88 

Central  Am.  & 

II 

51 

51 

Northern  South  Am. 

R = 353-409 

29.5-35.0 

14.8-19.2 

X = 374.17 

32.13 

17.24 

SD  = 12.14 

1.27 

0.84 

$ n = 37 

46 

48 

R = 358-407 

31.1-36.4 

15.7-20.0 

X = 383.57 

33.28 

18.05 

SD  = 11.64 

1.31 

0.81 

Contact  Zone 

S n = 10 

13 

14 

R = 351-401 

30.2-34.7 

16.3-18.9 

X = 379.5 

32.32 

17.31 

SD  = 16.15 

1.41 

0.88 

9/7  = 6 

8 

8 

R = 359-406 

29.2-33.8 

17.3-18.8 

X = 386.00 

31.7 

18.04 

SD  = 16.54 

1.55 

0.48 

Southern  South  Am. 

6 n = 35 

40 

43 

Inch  S.  Brazil 

R = 358-438 

26.5-37.7 

14.9-21.4 

X = 397.34 

32.49 

17.92 

SD  = 22.46 

2.37 

1.47 

CM 

11 

0+ 

44 

49 

R = 361-455 

28.2-37.3 

15.7-21.4 

X = 404.45 

32.82 

18.18 

SD  = 22.84 

2.28 

1.44 

Guadalupe  Island 

(5/7=1 

406 

32.2 

17.8 

9/7  = 2 

401.5 

31.0 

18.2 

“ n = number  of  specimens  measured. 
R = range. 

X = mean. 

SD  = standard  deviation. 


DISCUSSION 

Species  limits. — We  conclude  that  there  are 
3 species  of  crested  caracaras:  Caracara  piem- 
ens, C.  cheriwciy,  and  C lutosus.  Our  exami- 
nation of  nearly  400  specimens  from  the  con- 
tinental range  of  the  crested  caracaras  (Cara- 
cara) revealed  a mix  in  plumage  characters  of 
the  northern  and  southern  populations  only  in 
the  limited  zone  of  contact  near  the  Amazon 
River  in  Brazil.  We  consider  this  limited  char- 
acter sharing  a result  of  secondary  contact, 
first  suggested  by  Vuilleumier  (1970).  The 


specimens  indicate  an  abrupt  shift  in  appear- 
ance from  the  northern  to  the  southern  plum- 
age pattern,  and  most  of  the  intermediate 
specimens  exhibit  a non-consistent  mosaic  of 
characters  (back  pattern,  breast).  Juvenile 
birds  from  the  contact  zone  exhibit  the  same 
tendencies,  but  were  not  studied  as  extensive- 
ly as  adults. 

The  essential  reproductive  isolation  of  these 
populations  is  expressed  by  the  low  number 
of  intermediate  characters  in  specimens  from 
the  contact  zone,  and  the  relatively  narrow 


Dove  am!  Banks  • TAXONOMY  OF  CRESTED  CARACARAS 


337 


Wing  Chord 

FIG.  3.  Linear  regression  model  of  size  (indicated 
by  wing  chord)  and  latitude  of  196  specimen  records. 
Triangles  represent  specimens  from  south  latitude  and 
circles  represent  specimens  form  north  latitude. 

area  of  overlap.  Crested  caracaras  are  not 
widespread  in  Amazonia  east  of  the  Andes, 
occurring  only  as  wanderers  in  isolated  patch- 
es of  savanna  in  Amazonian  Colombia,  Ec- 
uador, and  Bolivia.  Individuals  of  the  two 
forms  probably  meet  only  infrequently  when 
they  wander  into  areas  of  sufficiently  open 
habitat  along  the  Amazon  River.  There  is  no 
record  of  contact  of  the  populations  in  western 
South  America.  Localities  of  South  American 
specimens  reveal  a distributional  gap  from 
northern  Peru  (excluding  three  cheriway  spec- 
imens taken  from  coastal  Peru  and  the  Peru- 
Ecuador  border)  south  to  the  middle  of  Chile. 
Parker  and  coworkers  (1982)  indicate  that 
crested  caracaras  are  uncommon  or  rare  in 
southern  Peru,  and  Johnson  (1965)  reports 
they  are  very  scarce  and  confined  to  the  sea 
coast  in  the  two  northernmost  provinces  of 
Chile  (see  map  77  in  Brown  and  Amadon 
1968). 

Review  of  the  Guadalupe  Island  species. — 
Morphologically,  adults  of  the  extinct  Gua- 
dalupe Island  population  differ  from  those  of 
North  and  South  American  mainland  popula- 
tions more  than  the  latter  two  differ  from  one 
another.  The  crest  of  C.  lutosus  is  brown  rath- 
er than  black,  and  the  crest  feathers  are  longer 
than  those  of  mainland  specimens.  The  throat 


is  buff  to  pale  brown,  not  white.  Ventral  and 
dorsal  surfaces  of  the  rest  of  the  body  are  en- 
tirely banded  with  brown  and  white  or  buff, 
and  there  is  no  solid  abdominal  patch  as  in 
the  mainland  populations.  Remiges  and  some 
scapulars  are  solid  brown  rather  than  black; 
most  upper-wing  coverts  are  bordered  narrow- 
ly with  darker  brown.  Sexed  adult  represen- 
tatives of  extinct  C.  lutosus  are  rare  in  collec- 
tions, and  only  three  specimens  were  mea- 
sured (see  Table  3).  They  do  not  differ  in  size 
from  mainland  birds. 

A few  specimens  from  southern  South 
America  (e.g.,  U.S.  National  Museum,  USNM 
13926,  Patagonia)  show  some  resemblance  to 
C.  lutosus.  These  South  American  mainland 
birds  are  primarily  brown  rather  than  black, 
and  most  of  the  body,  except  for  the  thighs 
and  a small  abdominal  patch,  is  banded.  How- 
ever, the  banding  is  finer  on  the  mainland 
birds  than  in  lutosus.  It  was  undoubtedly  birds 
like  this  that  Ridgway  (1876:460)  alluded  to 
when  he  stated  “This  species  resembles  the  P. 
tharus  [=  C.  plancus]  much  more  than  P. 
cheriway.  ...”  An  explanation  for  the  mor- 
phological similarity  of  birds  at  opposite  ends 
of  the  range,  with  birds  of  quite  different  ap- 
pearance occupying  the  intervening  continent, 
would  be  speculative  at  this  stage. 

We  agree  with  Ridgway  (1876)  that  the 
Guadalupe  Island  birds  are  specifically  dis- 
tinct. Recognition  of  that  distinct  and  isolated 
population  follows  logically  from  our  sepa- 
ration of  C.  cheriway  from  C.  plancus,  but  we 
suggest  that  the  extinct  C.  lutosus  should  be 
recognized  at  the  species  level  regardless  of 
the  treatment  of  the  mainland  populations. 

Intraspecific  variation. — Both  southern  (C. 
plancus)  and  northern  (C.  cheriway)  crested 
caracaras  have  been  variously  divided  at  the 
subspecific  level  by  different  authors.  Popu- 
lations of  the  Southern  Caracara  (C.  plancus) 
from  northern  Paraguay  to  the  Amazon  have 
frequently  been  separated  under  the  trinomial 
C.  p.  brasiliensis  (Gmelin,  1788)  on  the  basis 
of  being  darker  and  smaller  (Swann  1925, 
Wetmore  1926,  Peters  1931).  The  name  bras- 
iliensis cannot  be  applied  to  any  caracara 
(Banks  and  Dove  1992),  and  Gyldenstolpe 
(1951)  has  indicated  that  the  subspecific  name 
caracara  Spix,  1824  must  be  used  if  two 
forms  of  birds  south  of  the  Amazon  are  rec- 
ognized. However,  we  are  unable  to  document 


338 


THE  WILSON  BULLETIN  • Vol.  1 1 1,  No.  3.  September  1999 


any  consistent  difference  in  plumage  pattern, 
color,  or  size  among  these  birds,  and  so  we 
follow  most  recent  authors  (Hellmayr  and 
Conover  1949,  Gyldenstolpe  1951,  Blake 
1977,  Stresemann  and  Amadon  1979)  in  rec- 
ognizing the  populations  south  of  the  Amazon 
as  being  monotypic  Caracara  plancus. 

Populations  of  the  northern  mainland  spe- 
cies, C.  cheriway,  have  been  recognized  by  up 
to  four  subspecific  names  (cheriway  Jacquin, 
1784;  audubonii  Cassin,  1865;  pallidus  Nel- 
son, 1898;  arnmophilus  van  Rossem,  1939). 
Divisions  in  this  species  have  been  based  on 
the  amount  and  intensity  of  black  as  opposed 
to  brown  on  the  wings  and  back,  and  on  size. 
The  amount  of  black  is  subject  to  individual 
variation,  and  depends  on  stage  of  molt  and 
extensive  fading  related  to  the  open  habitat 
occupied  by  the  species.  Contrary  to  Griscom 
(1932),  we  do  not  believe  that  fading  from 
black  to  brown  occurs  post  mortem.  Size 
varies  less  than  some  earlier  writers  have  sug- 
gested and  increases  clinally  from  the  Ama- 
zon to  the  north  (Fig.  3),  so  that  the  birds  in 
the  southern  United  States  tend  to  be  among 
the  largest.  There  are  no  major  changes  in 
these  characters  to  warrant  the  recognition  of 
van  Rossem’s  (1939)  arnmophilus  of  Sonora, 
Mexico  with  a supposedly  smaller  bill  and 
feet  and  more  prominently  barred  tail,  or,  de- 
spite the  present  disjunction  of  range,  the  tri- 
nomial audubonii  that  Cassin  (1865)  based  on 
Florida’s  larger  birds.  We  agree  with  Strese- 
mann and  Amadon  (1979)  in  considering 
those  names  synonyms  of  C.  cheriway. 

Nelson  (1898)  separated  birds  from  the  Tres 
Marias  Islands  off  western  Mexico  under  the 
subspecific  name  pallidus,  on  the  basis  of  pal- 
er or  lighter  brown  coloration  and  slightly 
smaller  size.  Grant  (1965)  agreed  that  the  is- 
land birds  are  generally  paler  brown  than 
those  on  the  mainland  but  attributed  the  dif- 
ference to  greater  fading  of  the  island  birds. 
We  agree  that  the  color  is  not  diagnostic,  and 
note  that  new  feathers  or  those  generally  hid- 
den from  solar  radiation  are  no  paler  on  the 
island  birds  than  on  those  from  elsewhere  in 
the  range.  Furthermore,  AMNH  471349,  taken 
by  Nelson  at  the  Tres  Marias,  is  dark  both 
dorsally  and  ventrally.  Grant  (1965)  consid- 
ered a shorter  terminal  tail  band  and  a shorter 
wing  and  tarsus  in  males  to  be  sufficient  for 
recognition  of  pallidus,  even  though  compa- 


rable differences  could  not  be  demonstrated  in 
females.  However,  tarsus  length  is  too  indi- 
vidually variable  to  be  a useful  taxonomic 
character.  Wing  and  culmen  lengths  given  by 
Nelson  (1898)  and  Grant  (1965)  fit  well  in  the 
range  of  variation  of  the  much  larger  mainland 
sample  we  measured  (Table  3).  We  believe 
that  the  minor  difference  in  the  length  of  the 
terminal  tail  band  may  be  related  to  wear  of 
the  rectrices  and  that  other  differences  of  tail 
band  characters  mentioned  by  Grant  (1965) 
are  not  sufficient  to  warrant  separation  of  the 
Tres  Marias  population.  Some  specimens  from 
the  Tres  Marias  exhibit  extensive  barring  on 
the  abdomen,  but  this  is  within  the  range  of 
variation  of  the  mainland  specimens.  Thus,  we 
synonymize  pallidus  with  C.  cheriway. 

ACKNOWLEDGMENTS 

Part  of  this  study  was  supported  by  a Collections 
Study  Grant  to  CJD  from  the  American  Museum  of 
Natural  History  (AMNH).  The  study  was  initiated  as 
a graduate  course  by  CJD  at  George  Mason  University, 
supervised  by  RCB  and  J.  Shaffer.  The  work  was  con- 
ducted at  the  U.S.  National  Museum  of  Natural  His- 
tory (USNM).  We  thank  J.  V.  Remsen,  S.  W.  Cardiff, 
L.  L.  Kiff,  K.  C.  Parkes,  and  D.  Willard  for  scoring 
specimens  in  their  care  and  gratefully  acknowledge 
The  Lield  Museum  of  Natural  History  (LMNH)  for  the 
loan  of  specimens.  This  study  greatly  benefited  from 
help  by  G.  C.  Banks  in  measuring  and  scoring  many 
specimens  at  the  following  collections;  Lield  Museum 
of  Natural  History  (LMNH),  The  Natural  History  Mu- 
seum (BMNH),  Louisiana  State  University  Museum  of 
Natural  Science,  California  Academy  of  Sciences,  Cin- 
cinnati Natural  History  Museum,  Canadian  Museum  of 
Nature,  University  of  Michigan  Museum  of  Zoology 
(UM),  Museum  of  Vertebrate  Zoology  (MVZ),  Muse- 
um of  Comparative  Zoology  (MCZ),  University  of 
Kansas  Museum  of  Zoology,  Yale  Peabody  Museum, 
Royal  Ontario  Museum,  Museum  National  d’Histoire 
Naturelle,  Paris,  and  Carnegie  Museum  of  Natural  His- 
tory. CJD  measured  specimens  at  AMNH,  USNM  and 
Naturhistorisches  Museum,  Vienna  (NMWZ).  A.  R. 
Phillips  provided  sub.specific  comments;  M.  R.  Brown- 
ing and  J.  V.  Remsen  reviewed  early  drafts  of  this  pa- 
per. We  thank  M.  Isler  for  assistance  with  the  map  and 
V.  Krantz  (USNM)  for  the  photographs. 

LITERATURE  CITED 

Amkrican  Ornithologists'  Union.  1957.  Check-list 
of  North  American  birds.  Lifth  ed.  .Vmerican  Or- 
nithologi.sts’  Union,  Baltimore,  Maryland. 
American  Ornithologists’  Union.  1983.  Check-li.st 
of  North  American  birds.  Sixth  ed.  American  Or- 
nithologists’ Union,  Washington,  D.C. 

American  Ornithologlsts’  Union.  1998.  Check-list 


Dove  and  Banks  • TAXONOMY  OF  CRESTED  CARACARAS 


339 


of  North  American  birds.  Seventh  ed.  American 
Ornithologists’  Union,  Washington,  D.C. 

Banks,  R.  C.  and  C.  J.  Dove.  1992.  The  generic  name 
for  Crested  Caracaras  (Aves:  Falconidae).  Proc. 
Biol.  Soc.  Wash.  105:420-425. 

Blake,  E.  R.  1977.  Manual  of  Neotropical  birds.  Vol. 

1.  Univ.  of  Chicago  Press,  Chicago,  Illinois. 
Brabourne,  L.  and  C.  Chubb.  1912.  The  birds  of 

South  America.  Vol.  1.  Taylor  & Francis,  London, 
U.K. 

Brown,  L.  and  D.  Amadon.  1968.  Eagles,  hawks  and 
falcons  of  the  world.  Vol.  2,  Country  Life  Books, 
London,  U.K. 

Cassin,  J.  1865.  Notes  on  some  new  and  little  known 
rapacious  birds.  Proc.  Acad.  Nat.  Sci.  Phila.  17: 
2-3. 

Friedmann,  H.  1950.  The  birds  of  North  and  Middle 
America.  Part  XL  U.S.  Nat.  Mus.  Bull.  50:1-793. 
Friedmann,  H.,  L.  Griscom,  and  R.  T.  Moore.  1950. 
Distributional  check-list  of  the  birds  of  Mexico. 
Part  1.  Pac.  Coast  Avif.  29:1—202. 

Grant,  P.  R.  1965.  A systematic  study  of  the  terrestrial 
birds  of  the  Tres  Marfas  Islands,  Mexico.  Postilla 
60:1-106. 

Griscom,  L.  1932.  The  distribution  of  bird-life  in  Gua- 
temala. Bull.  Amer.  Mus.  Nat.  Hist.  64: 1-439. 
Gyldenstolpe,  N.  1951.  The  ornithology  of  the  Rio 
Purus  region  in  western  Brazil.  Arkiv  Zook,  ser. 

2,  2:1-320. 

Hellmayr,  C.  E.  and  B.  Conover.  1949.  Catalogue 
of  birds  of  the  Americas  and  adjacent  islands. 
Zook  Sen,  Field  Mus.  Nat.  Hist.  13,  Part  1 (4):1- 
358. 

Johnson,  A.  W.  1965.  The  birds  of  Chile  and  adjacent 
regions  of  Argentina,  Bolivia  and  Peru.  Platt  Es- 
tablecimientos  Graficos,  S.A.,  Buenos  Aires,  Ar- 
gentina. 

Morrison,  J.  L.  1996.  Crested  Caracara  (Caracara 
plancus).  In  The  birds  of  North  America,  no.  249 
(A.  Poole  and  F.  Gill,  Eds.).  The  Academy  of  Nat- 
ural Sciences,  Philadelphia,  Pennsylvania;  The 
American  Ornithologists’  Union,  Washington, 
DC. 

Nelson,  E.  W.  1898.  New  birds  from  western  Mexico. 
Proc.  Biol.  Soc.  Wash.  12:5-11. 


Parker,  T.  A.,  Ill,  S.  A.  Parker,  and  M.  A.  Plenge. 
1982.  An  annotated  checklist  of  Peruvian  birds. 
Buteo  Books,  Vermillion,  South  Dakota. 

Paynter,  R.  a.,  Jr.  1988.  Ornithological  gazetteer  of 
Chile.  Harvard  Univ.,  Cambridge,  Massachusetts. 

Paynter,  R.  A.,  Jr.  and  M.  A.  Traylor,  Jr.  1991. 
Ornithological  gazetteer  of  Brazil.  Harvard  Univ., 
Cambridge,  Massachuetts. 

Peters,  J.  L.  1931.  Check-list  of  birds  of  the  world. 
Vol.  1.  Harvard  Univ.  Press,  Cambridge,  Massa- 
chusetts. 

Ridgway,  R.  1876.  Studies  of  the  American  Falconi- 
dae. Monograph  of  the  Polybori.  Bulk  U.S.  Geok 
Geogr.  Surv.  Territories  1:451-473. 

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

Stevens,  L.  and  M.  Traylor,  Jr.  1983.  Ornithological 
gazetteer  of  Peru.  Harvard  Univ.,  Cambridge, 
Massachusetts. 

Stresemann,  E.  and  D.  Amadon.  1979.  Order  Ealcon- 
iformes.  Pp.  271—425  in  Check-list  of  birds  of  the 
world,  Vol.  1,  second  ed.  (E.  Mayr  and  G.  W. 
Cottrell,  Eds.).  Museum  of  Comparative  Zoology, 
Cambridge,  Massachusetts. 

Swann,  H.  K.  1925.  A monograph  of  the  birds  of  prey 
(order  Accipitres).  Part  Ik  Wheldon  & Wesley, 
London,  U.K. 

Systat.  1992.  Version  5 ed.  Systat,  Inc.  Evanston,  Il- 
linois. 

U.S.  Eish  and  Wildlife  Service.  1987.  Endangered 
and  threatened  wildlife  and  plants;  threatened  sta- 
tus for  the  Florida  population  of  the  Audubon’s 
Crested  Caracara.  Federal  Register  52:25229- 
25231. 

VAN  Rossem,  a.  j.  1939.  Some  new  races  of  birds  from 
Mexico.  Ann.  Mag.  Nat.  Hist.  ser.  11,  4:439-443. 

VuiLLEUMiER,  F.  1970.  Generic  relations  and  speciation 
patterns  in  the  caracaras  (Aves:  Falconidae).  Bre- 
viora  355: 1-29. 

Wetmore,  a.  W.  1926.  Birds  of  Argentina,  Paraguay, 
Uruguay,  and  Chile.  U.S.  Nat.  Mus.  Bulk  133:95- 
98. 

Wetmore,  A.  W.  1965.  Birds  of  Panama.  Vol.  1. 
Smithson.  Misc.  Coll.  150:1-483. 


Wilson  Bull.,  111(3),  1999,  pp.  340-345 


VISUAL  COMMUNICATION  AND  SEXUAL  SELECTION  IN  A 
NOCTURNAL  BIRD  SPECIES,  CAPRIMULGUS  RUFICOLLIS,  A 
BALANCE  BETWEEN  CRYPSIS  AND  CONSPICUOUSNESS 

JUAN  ARAGONES,  ' LUIS  ARIAS  DE  REYNA,'  AND  PILAR  RECUERDA' 


ABSTRACT. — Cryptic  protective  mechanisms  and  the  conspicuousness  required  to  communicate  result  in  a 
conflict  of  opposing  selection.  In  the  Red-necked  Nightjar  (Caprimulgus  nificollis)  a nocturnal  bird,  the  use  of 
a restricted  signaling  strategy  provides  an  appropriate  balance  between  these  two  selection  forces.  Conspicuous 
white  wing  and  tail  bands  may  have  been  favored  by  sexual  selection  in  this  species.  We  studied  the  variation 
of  visual  signals  and  found  conspicuousness  to  be  closely  related  to  sex  and  age,  being  much  higher  in  males 
and  adults.  This  variation  allows  an  individual  to  identify  the  reproductive  status  of  conspecifics,  providing 
sexual  selection  a basis  to  select  these  visual  signals  in  this  and  other  nocturnal  bird  species.  We  believe  that  a 
relationship  between  restricted  signaling  strategy  and  sexually  selected  visual  signals  may  occur  in  nocturnal 
species  that  use  visual  comunication.  Received  24  April  1998,  accepted  5 Feb.  1999. 


Crypsis  is  one  of  the  more  effective  anti- 
predator mechanisms  (Baker  and  Parker 
1979).  Because  it  relies  on  inconspicuousness, 
its  use  can  conflict  with  the  conspicuousness 
required  for  communication.  Selection  pres- 
sure drives  populations  to  address  signals 
among  conspecifics  (conspicuousness  increas- 
es at  close  range)  and  to  conceal  them  from 
predators  (inconspicuouness  or  long  range 
crypsis;  Butcher  and  Rohwer  1989).  An  ap- 
propriate balance  between  crypsis  and  com- 
munication is  achieved  by  the  use  of  restricted 
signaling,  for  example,  the  presence  of  hidden 
conspicuous  signals  in  highly  specific  body 
zones  that  are  only  exhibited  in  some  situa- 
tions (see  Butcher  and  Rohwer  1989). 

In  strictly  nocturnal  birds,  which  account 
for  less  than  3%  of  all  bird  species  (Martin 
1990),  the  visual  channel  is  assumed  to  play 
an  insignificant  role  in  communication  com- 
pared to  diurnal  birds.  However,  this  is  only 
an  assumption  and  nocturnal  birds  may  use 
visual  signals  to  communicate  more  widely 
than  is  thought.  The  Red-necked  Nightjar  {Ca- 
primulgus ruficolUs)  is  a nightbird  of  open  and 
semi-open  habitats  that  uses  sight  to  capture 
flying  insects.  The  plumage  of  caprimulgids  is 
highly  cryptic,  but  includes  white  spots  on 
wings,  tail,  and  throat.  These  markings,  be- 
cause of  their  anatomical  location,  are  only 


' Dpto.  Biologia  Animal,  Univ.  de  Cordoba,  Avda. 
San  Alberto  Magno,  sn.  14004  Cordoba,  Spain. 

2 Present  address:  Avda.  Cadiz  n"  5,  I4009-C6rdoba. 
Spain. 

’Corresponding  author;  E-mail:  ba2biani@Lieo.es 


visible  in  some  contexts  (agonistic,  antipred- 
ator, and  sexual  displays)  when  the  sender  in- 
tentionally shows  its  markings  to  the  receiver 
(Bent  1940,  Mengel  1972,  Bruce  1973,  Cramp 
1985,  Fry  et  al.  1988,  Cleere  1998,  Aragones 
et  al.  in  press).  Some  caprimulgid  species 
have  been  reported  to  show  individual  varia- 
tion in  spot  size  and  color  (Common  Night- 
hawk,  Chordeiles  minor,  Selander  1954;  Red- 
necked Nightjar,  Beven  1973,  Soares  1973; 
Blackish  Nightjar,  C.  nigrescens,  Ingels  and 
Ribot  1982).  This  variation  in  spots  suggests 
that  signal  conspicuousness  might  vary  with 
age  and  sex. 

The  objective  of  this  paper  is  to  describe 
the  variation  of  spots  in  signal  conspicuous- 
ness in  the  Red-necked  Nightjar  and  to  ex- 
amine whether  this  variation  is  related  to  age 
and/or  sex.  We  hypothesize  that  if  males,  es- 
specially  adult  males,  show  greater  conspicu- 
ousness in  wing  and  tail  bands,  then  these 
plumage  characters  may  be  sexually  selected. 

METHODS 

We  used  data  from  170  specimens  obtained  from 
bird  collections  at  Estacion  Biologica  de  Donana, 
CSIC  (/?  = 22)  and  from  road  casualties  (/t  = 148)  in 
Cordoba  (southern  Spain)  during  the  1986  to  1994 
breeding  seasons.  From  this  sample  we  measured  the 
following  variables  related  to  conspicuousness.  (1) 
Band  size.  The  wing-band  consisted  of  3 to  4 spots 
and  the  tail-band  of  2 to  3 spots.  Using  a digital  caliper, 
the  longer  and  shorter  axis  of  each  spot  was  measured 
(±  0.1  mm)  and  the  area  for  each  was  calculated.  The 
surface  areas  of  the  spots  present  in  the  outer  primaries 
of  the  right  wing  were  summed  and  recorded  as  wing- 
band  size  and  those  on  right  rectrices  were  summed 
and  recorded  as  tail-band  size.  The  area  of  each  spot 


340 


Arai>(mes  et  al.  • CRYPSIS  AND  CONSPICUOUSNESS 


341 


FIG.  1.  Wing-band  (WBC)  and  tail-band  contrast 
(TBC),  calculated  from  color  and  sharpness  values  of 
spots  forming  the  bands  ( 1 = maximum  contrast  value, 
white  spot  with  sharp  contrasting  edges,  0 = minimum 
contrast  value,  dark  spots  without  shaip  contrasting 
edges). 

was  indexed  by  the  product  of  the  long  and  the  short 
axis.  (2)  Band  contrast.  Contrast  was  measured  from 
the  color  and  sharpness  of  the  spots  that  formed  the 
wing  and  tail  bands.  The  contrast  gradient  was  estab- 
lished from  a reference  series  formed  by  four  types  of 
spots  that  were  assigned  the  values  0,  0.33,  0.66  and 
1;  where  0 denoted  the  spot  with  the  darkest  colour 
and  least  sharp  contrasted  edges,  and  1 that  with  the 
lightest  color  and  sharpest  edges.  A given  individual 
can  possess  spots  of  different  types  (depending  on  col- 
or and  sharpness  variations),  so  each  spot  was  assigned 
a value,  and  a mean  index  was  used  for  the  contrast 
analysis  in  wing-band  contrast  and  tail-band  contrast 
which  included  all  spots  of  the  wing  and  tail  (Fig.  1). 
(3)  Uniformity  of  color  spots  within  a band.  Some  in- 
dividuals showed  bands  consisting  of  a single  type  of 
spot,  whereas  others  had  one  or  two  different  types; 
the  former  were  designated  “uniform”  and  the  latter 
“non-uniform”.  (4)  Spot  number  in  wing  and  tail 
bands.  (5)  Band  conspicuousness.  Band  size  and  con- 
trast values  were  first  log  transformed,  normalizing  the 
data  for  the  application  of  parametric  statistics  and 
used  to  calculate  a conspicuousness  index  (C)  from  the 
following  expression: 

C = WBS  X log(WBC+l)  + TBS  X log(TBC-M)/ 
WBS  + TBS 

where  WBS  = wing-band  size,  TBS  = tail-band  size, 
WBC  = wing-band  contrast,  and  TBC  = tail-band 
contrast.  This  index  varied  from  0 to  I,  with  1 corre- 
sponding to  maximum  conspicuousness.  Sex  was  de- 
termined by  gonadal  dissection.  Differences  in  band 
size  and  conspicuouness  were  analysed  by  ANOVA 
and  the  Tukey  test,  and  differences  in  the  number  of 
spots  and  their  uniformity  by  means  of  the  G-test.  Con- 
trast differences  were  identified  by  multiple  logistic  re- 
gression, which  allows  one  or  more  categorical  vari- 
ables (contrast  in  our  case)  to  be  analysed  and  related 
to  a dependent  variable.  This  method  is  especially  suit- 


able tor  data  that  are  not  normally  distributed  and  is 
more  eftective  than  other  classifying  methods  such  as 
discriminant  analysis  (Press  and  Wifson  1978;  see  also 
Hanell  1986,  Schlinger  1990,  Fancy  et  al.  1993).  Con- 
spicuousness data  were  transformed  logarithmically 
for  parametric  analysis.  All  analyses  were  performed 
using  JMP  software  (version  2,  Macintosh  computer). 

RESULTS 

There  were  significant  differences  in  wing- 
band  size  {F  = 31.47,  df  = 3,  90,  P < 0.001) 
and  tail-band  size  (F  = 41.61,  df  = 3,  90,  P 
< 0.001)  between  sex-age  categories.  Wing 
and  tail  bands  were  significantly  larger  in 
males  than  in  females  and  larger  in  adults  rel- 
ative to  young.  There  was  a clear  trend  for 
band  size  to  increase  in  an  age-sex  sequence, 
both  in  wing-band  (mean  young  females  = 
662.3  ± 37.87  SD,  n = 29;  mean  adult  fe- 
males = 803.1  ± 52.66,  A?  = 15;  mean  young 
males  = 986.3  ± 37.24,  n = 30;  mean  adult 
males  = 1210.5  ± 45.61,  n = 20)  and  in  tail 
band  (mean  young  females  = 743.4  ± 67.06, 
n = 29;  mean  adult  females  = 1046.4  ± 
93.24,  n = 15;  mean  young  males  = 1428.7 
± 65.93,  n = 30;  mean  adult  males  = 1852.6 
± 80.75,  n = 20).  Both  variables  were  found 
to  be  positively  correlated  (r  = 0.67,  n = 93, 
P < 0.001). 

The  multiple  logistic  regression  discrimi- 
nated between  the  four  age-sex  classes  in 
terms  of  wing-band  or  tail-band  contrast  (Ta- 
ble 1).  There  were  significant  differences  in 
sex  and  age  for  wing-band  contrast  (x^3  = 
86.33,  P < 0.001,  n = 94)  and  tail-band  con- 
trast X"3  = 94.21,  P < 0.001,  n = 94).  Males 
and  adults  always  had  the  largest  contrast  in- 
dex both  for  wing-band  and  tail-band  (Fig.  2). 
Sharp  edges  were  detected  only  in  males  with 
wing-band  or  tail-band  contrast  index  close  to 
one  because  only  such  spots  had  sharp  edge 
contrasts.  Males  showed  greater  uniformity  in 
wing-band  contrast  than  females,  and  adults 
were  more  uniform  than  young  in  this  respect. 
All  groups  showed  uniform  tail  bands  (Table 
2). 

The  presence  of  four  spots  on  the  wing  is 
an  exclusive  feature  of  adult  males  (Table  2); 
over  45%  (/?  = 16)  of  adult  males  have  the 
fourth  spot.  About  80%  of  adult  males  also 
tended  to  possess  three  caudal  spots  (/?  = 28; 
Table  2).  Additional  spots  on  the  tail  was  not 
significantly  different  among  the  other  groups 
(Table  2).  The  additional  wing  spot  increased 


342 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  3.  September  1999 


TABLE  1.  Parameter  estimates  of  logistic  functions  derived  from  multiple  logistic  regression. 


Logistic  function  parameters 

Estimate 

Wald  test 

p 

(Young  female  vs  adult  female) 

Intercept  I 

9.47 

17.62 

<0.001 

Wing-band  contrast 

-13.22 

21.22 

<0.001 

(Adult  female  vs  young  male) 

Intercept  2 

7.52 

1 1.51 

<0.001 

Wing-band  contrast 

-10.17 

14.43 

<0.001 

(Young  male  vs  adult  male) 

Intercept  3 

1.63 

0.56 

>0.05 

Wing-band  contrast 

-1.34 

0.32 

>0.05 

(Young  female  vs  adult  female) 

Intercept  1 

10.05 

23.21 

<0.001 

Tail-band  contrast 

-16.03 

31.09 

<0.001 

(Adult  female  vs  young  female) 

Intercept  2 

8.45 

16.36 

<0.001 

Tail-band  contrast 

-12.29 

20.59 

<0.001 

(Young  male  vs  adult  male) 

Intercept  3 

6.09 

10.54 

<0.001 

Tail-band  contrast 

-6.90 

10.20 

<0.001 

the  wing-band  size  of  adult  males  an  average 
of  6%  and  the  tail-band  size  of  adult  males  an 
average  of  12%.  The  presence  of  additional 
spots  on  wing  and  tail  increased  the  overall 
band  area  by  9.6%  and  presumably  increased 
signal  conspicuousness.  There  were  signifi- 
cant differences  in  conspicuousness  (Table  2) 
with  regard  to  age-sex  classes,  with  greater 
conspicuousness  in  males  and  adults. 

DISCUSSION 

We  detected  a marked  trend  for  signal  con- 
spicuousness to  increase  with  age  and  in 
males  which  was  reflected  in  an  increase  of 
contrast,  number  of  spots,  wing-  and  tail- 
bands  area,  and  uniformity  of  contrast.  The 
differences  in  conspicuousness  in  wing-  and 
tail-bands  were  related  to  the  age  and  sex  class 
of  the  individual.  Potentially,  these  signals  let 
conspecifics  evaluate  one  another  in  aggres- 
sive or  reproductive  contexts.  Signal  conspic- 
uousness seemingly  involves  little  energy  ex- 
penditure but  is  combined  with  expensive  dis- 
plays; the  combination  results  in  an  efficient 
signal  that  conveys  a much  greater  amount  of 
information  (Hasson  1991).  We  found  that 
spot  number  was  related  to  sex  and  age,  but 
Forero  and  coworkers  (1995)  found  no  statis- 
tically significant  differences  in  the  average 


FIG.  2.  Heavy  line  indicates  probability  of  belong- 
ing to  a sex  cla.ss  as  a function  of  wing-band  contra.st 
(WBC)  or  tail-band  contra.st  (TBC),  thin  lines  indicates 
probability  of  belonging  to  an  age  class. 


Aragones  cl  al.  • CRYPSIS  AND  C(3NSPICUOUSNESS 


343 


TABLE  2.  Frequency  distribution  of 
index  for  age-sex  groups. 

contrast  uniformity. 

additional  spots  in 

bands  and 

conspicuousness 

Female 

Male 

Young 

Adult 

Young 

Adult 

p 

Contrast  uniformity  (n  = 94) 

n = 29 

/7  = 15 

n = 30 

/;  = 20 

WBC  (G,  = 27.29) 

37% 

66% 

66% 

95% 

<0.001 

TBO'  (Gj  = 1 1.09) 

88% 

87% 

91% 

89% 

>0.05 

Additional  spots  (n  = 170) 

n = 50 

n = 31 

n = 54 

n = 35 

In  wing  (G,  = 40.33) 

0% 

0% 

0% 

55% 

<0.001 

In  tail  (G,  = 39.85) 

6% 

6% 

6% 

80% 

<0.001 

Conspicuousness  (/;  = 94) 

n = 29 

/;  = 15 

n = 30 

n = 20 

(F  = 71.76,  n = 3,  90) 

0.23 

0.43 

0.43 

0.91 

<0.001 

“ WBC  = wing-band  contra.si,  TBC  = 

tail-band  contrast 

number  of  spots  between  groups  of  the  same 
species.  We  believe  the  addition  of  new  spots 
has  a strong  biological  significance  because  it 
occurs  only  in  adult  males  as  reported  by  For- 
ero  and  co workers  (1995),  but  their  method, 
failed  to  demonstrate  a significant  difference. 


In  the  Red-necked  Nightjar  increased  signal 
conspicuousness  may  increase  the  likelihood 
of  mating  (Andersson  1982,  Andersson  1992, 
Saetre  et  al.  1994),  and  conspicuous  males 
may  have  a higher  mating  success  (Payne 
1982,  Price  1984,  Jarvi  et  al.  1987,  Harvey 


FIG.  3.  For  restricted  signaling  strategy  to  be  effective,  conspicuous  plumage  zones  (black  and  white  areas 
in  the  drawing)  must  be  concealed  by  cryptic  plumage  zones  (gray  areas  in  the  drawing)  while  at  rest.  A 
hypothetical  model  based  on  the  genus  Caprimulf’us  (I)  represents  the  maximum  possible  conspicuousness  for 
restricted  signaling  strategy;  a larger  conspicuous  area  would  defeat  the  defensive  system  based  on  crypsis  and 
hence  cancel  restricted  signaling  strategy.  The  Nacunda  Nighthawk  (2),  the  Sand-Colored  Nighthawk  (3)  and 
the  White-Winged  Nigthjar  (4)  use  three  restricted  signaling  strategies  that  are  very  similar  to  that  of  the  model, 
esspecially  the  Sand-Colored  Nighthawk. 


344 


THE  WILSON  BULLETIN  • VoL  III,  No.  3,  September  1999 


and  Bradbury  1991,  Sundberg  and  Larsson 
1994).  The  fact  that  increased  signal  conspic- 
uousness was  associated  with  adult  males  sug- 
gests that  the  signal  could  be  modified  by  sex- 
ual selection  (Hbglund  1993).  For  adult  males, 
this  signal  indicates  that  those  individuals 
have  survived  at  least  two  reproductive  sea- 
sons and,  therefore,  may  reflect  a high  repro- 
ductive quality.  In  another  nocturnal  bird  spe- 
cies, the  Great  Snipe  {Gallinago  media)  males 
have  white  spots  on  their  tails,  and  females 
choose  the  males  with  the  whitest  signals 
(Hbglund  et  al.  1992).  This  suggests  that  sig- 
nal conspicuousness  in  nocturnal  birds  could 
be  an  effective  way  of  distinguishing  between 
potential  mates. 

Our  results  show  that  in  Red-necked  Night- 
jar both  sexes  possess  ornaments,  and  that  in 
both  sexes  spots  tend  to  increase  in  conspic- 
uousness with  age.  Forero  and  co workers 
(1995)  claimed  that  visual  signals  increased 
only  in  males.  We  believe  that  their  results  are 
due  to  a less  precise  method  of  measuring  spot 
size. 

The  restricted  signaling  strategy  allows  two 
scarcely  compatible  mechanisms  to  be  com- 
bined and  might  occur  in  many  cryptic  species 
of  the  genera  Burhinus,  Charadrius  and  Gal- 
linago, although  this  has  not  yet  been  inves- 
tigated. For  a nocturnal  signal,  contrast  is 
more  important  than  color,  as  in  the  capri- 
mulgids  and  other  species  that  use  restricted 
signaling  strategy  with  white  spots  (i.e.,  Gal- 
linago media,  Hbglund  et  al.  1992;  Burhinus 
spp.,  Martin  1990,  Hayman  et  al.  1986). 
Therefore,  increasing  the  amount  of  white  col- 
or increases  contrast  and  presumably  maxi- 
mizes the  signaling  ability.  Similar  trends  of 
increasing  conspicuousness  with  age  have 
been  detected  in  other  caprimulgids  (Little 
Nightjar,  C.  parvulus  heterurus,  Schwartz 
1968;  Chuck-WiU’s-Widow,  C.  vociferus, 
Rohwer  1971;  Blackish  Nightjar,  Ingels  and 
Ribot  1982;  Scrub  Nightjar,  C.  anthonyi,  Rob- 
bins and  Ridgeley  1994;  Nacunda  Nighthawk, 
Podager  nacunda,  Aragones  unpubl.  data; 
Pauraque,  Nyctidromus  albicollis,  Aragones 
1997a,  b). 

From  the  distribution  of  visual  signals  in 
Red-necked  Nightjar  (throat,  wings,  and  tail- 
bands),  we  developed  a model  plumage  pat- 
tern for  a nightjar  that  represents  the  optimum/ 
maximum  distribution  of  conspicuous  areas  in 


the  body  (Fig.  3).  We  subsequently  found  that 
the  Nacunda  Nighthawk  and  the  White- 
Winged  Nightjar  {Caprimulgus  candicans) 
have  patterns  that  closely  resemble  our  model, 
and  the  Sand-colored  Nighthawk’s  (Chordei- 
les  rupestris)  visual  signals  patterns  are  iden- 
tical with  that  of  the  model.  All  three  are  gre- 
garious South  American  species  of  open  or 
semiopen  habitats  with  sexual  variations  in 
wing-  and  tail-bands  (Ffrench  1986,  Hilty  and 
Brown  1986,  Sick  1993).  It  is  interesting  to 
note  that  the  Sand-colored  Nighthawk  and  the 
Yellow-billed  Tern  {Sterna  superciliaris),  two 
Amazonic  species  that  use  the  same  fluvial 
habitat,  show  almost  identical  visual  signal 
patterns  and  general  appearance  (see  Sick 
1993).  This  convergent  evolution  stresses  the 
significance  of  open  and  semiopen  habitats  for 
restricted  signaling  strategy  evolution.  Such 
habitats  are  also  characterized  by  other  species 
that  employ  this  strategy,  Burhinus  spp.  and 
the  Great  Snipe  (Hayman  et  al.  1986).  We  be- 
lieve that  a relationship  between  restricted  sig- 
naling strategy  and  sexually  selected  visual 
signals  may  occur  in  nocturnal  species  that 
uses  visual  comunication. 

ACKNOWLEDGMENTS 

Valuable  assistance  was  provided  at  various  stages 
of  this  study  by  many  friends.  We  thank  M.  C.  Casaut. 
A.  Leiva,  M.  A.  Nunez  and  R.  Pulido  for  help  with 
the  collection  of  traffic  casualties,  and  E M.  Coco  for 
help  with  statistical  analysis.  We  are  grateful  to  M.  A. 
Nunez,  E M.  Coco,  R.  Reques,  S.  Carpintero  and  J. 
Marin  for  practical  support,  discussion,  and  comments. 
Two  anonymous  referees  provided  valuable  sugges- 
tions on  an  early  version  of  this  manuscript. 

LITERATURE  CITED 

Andersson,  M.  1982.  Female  choice  selects  for  ex- 
treme tail  lengh  in  a widowbird.  Nature  299:818- 
820. 

Andersson,  S.  1992.  Female  preference  for  long  tails 
in  lekking  Jackson's  Widowbirds:  experimental 
evidence.  Anim.  Behav.  43:379-388. 

Aragones,  J.  1997a.  Influencia  de  la  cripsis  sobre  el 
comportamiento  del  Chotacabras  pardo,  Capri- 
mulf>u.s  ruficoUis.  Tesis  Doctoral.  Univ.  de  Cor- 
doba. Cordoba,  Spain. 

Aragones,  J.  1997b.  Risk  taking  and  flushing  distance: 
a way  of  parental  investment  in  the  Pauraque 
(Nyctidromu.s  albicollis).  Etologia  5:83—90. 
Aragones,  J.,  L.  Arias  de  Reyna,  and  P.  Recuerda. 
1998.  Antipredator  strategies  and  mimicry  in  Red- 
necked Nightjar  Caprimulgus  ruficollis.  Etologia 
6:53-59. 


Aragones  el  al.  • CRYPSIS  AND  CONSPICUOUSNESS 


345 


Baker,  R.  E.  and  G.  A.  Parker.  1979.  The  evolution 
of  bird  coloration.  Phil.  Tran.s.  R.  Soc.  Lond.  B 
287:63-130. 

Bent,  A.  C.  1940.  Life  histories  of  North  American 
cuckoos,  goatsuckers,  hummingbirds  and  their  al- 
lies. Bull.  U.S.  Nat.  Mus.  176:147-254. 

Beven,  G.  1973.  Studies  of  less  familiar  birds.  Red- 
necked Nightjar.  Brit.  Birds  66:390-396. 

Bruce,  A.  1973.  Tail  flashing  display  in  the  Whip- 
poor-will.  Auk  90:682. 

Butcher,  G.  S.  and  S.  Rohwer.  1989.  The  evolution 
of  conspicuous  and  distinctive  coloration  for  com- 
munication in  birds.  Curr.  Ornithol.  6:51-108. 

Cleere,  N.  1998.  Nightjars.  A guide  to  nightjars  and 
related  nightbirds.  Pica  Press.  Sussex,  U.K. 

Cramp,  S.  1985.  The  birds  of  the  western  Palearctic. 
Vol  4.  Oxford  Univ.  Press,  Oxford,  U.K. 

Fancy,  S.  G.,  T.  K.  Pratt,  G.  D.  Lindsey,  C.  K.  Har- 
ada,  a.  H.  Parent,  and  J.  D.  Jacobi.  1993.  Iden- 
tifyng  sex  and  age  of  Apapane  and  liwi  on  Ha- 
waii. J.  Field  Ornithol.  64:262—269. 

Ffrench,  R.  1986.  A guide  to  the  birds  of  Trinidad  and 
Tobago.  Harrowood  Books, Valley  Forge,  Penn- 
sylvania. 

Forero,  M.  G.,  j.  L.  Tella,  and  L.  Garcia.  1995.  Age 
related  evolution  of  sexual  dimorphism  in  the 
Red-necked  Nightjar  Caprirnulgus  ruficolUs.  J. 
Ornithol.  136:447-451. 

Fry,  C.  H.,  S.  Keith,  and  E.  K.  Urban.  1988.  The 
birds  of  Africa.  Vol  3.  Academic  Press,  London, 
U.K. 

Harvey,  P.  H.  and  J.  W.  Bradbury.  1991.  Sexual  se- 
lection. Pp.  203—233  in  Behavioural  ecology.  An 
evolutionary  approach  (J.  R.  Krebs  and  N.  B.  Da- 
vies, Eds.).  Blackwell  Scientific  Publications,  Ox- 
ford, U.K. 

Harrell,  F.  E.  1986.  The  logistic  procedure.  Pp.  269— 
293  in  SUGI  supplemental  library  user’s  guide, 
version  5 (R.  P.  Hastings,  Ed.).  SAS  Institute  Inc., 
Cary,  North  Carolina. 

Hasson,  O.  1991.  Sexual  displays  as  amplifiers:  prac- 
tical examples  with  emphasis  on  feather  decora- 
tions. Behav.  Ecol.  2:189-197. 

Hayman,  R,  P.  Marchant,  and  T.  Prater.  1986. 
Shorebirds:  an  identification  guide  to  the  waders 
of  the  world.  Croom  Helm,  London,  U.K. 

Hilty,  S.  L.  and  W.  L.  Brown.  1986.  The  birds  of 
Colombia.  Princeton  Univ.  Press,  Princeton,  New 
Jersey. 

Hoglund,  j.  1993.  Cost  of  mate  advertisement.  Eto- 
logia  3:143-150. 

Hoglund,  J.,  M.  Eriksson,  and  L.  Lindell.  1992.  Fe- 
males of  the  lek  breeding  Great  Snipe,  GoUinago 


media,  prefer  males  with  white  tails.  Anim.  Be- 
hav. 40:23-32. 

Ingels,  J.  and  J.  H.  Ribot.  1982.  Variations  in  the 
white  markings  of  the  Blackish  Nightjar  Capri- 
iniilgus  nigrescens.  Bull.  Brit.  Ornithol.  Club  102: 
1 19-122. 

Jarvi,  T,  E.  Roskaft,  M.  Bakken,  and  B.  Zum.steg. 
1987.  Evolution  of  variation  in  male  .secondary 
sexual  characteristics:  a test  of  eight  hypothesis 
applied  to  Pied  Flycatchers.  Behav.  Ecol.  Socio- 
biol.  20:161-169. 

Martin,  G.  1990.  Birds  by  night.  T.  & A.  D.  Poyser, 
London,  U.K. 

Mengel,  R.  M.  1972.  Wing  clapping  in  territorial  and 
courtship  behavior  of  the  Chuck-will’s-widow  and 
Poor- Will  (Caprimulgidae).  Auk  89:441-444. 

Payne,  R.  B.  1982.  Ecological  consequences  of  song 
matching:  breeding  success  and  intraspecific  song 
mimicry  in  Indigo  Buntings.  Ecology  63:401- 
411. 

Press,  S.  J.  and  S.  Wilson.  1978.  Choosing  between 
logistic  regression  and  discriminant  analysis.  J. 
Am.  Stat.  Assoc.  73:699-705. 

Price,  T.  D.  1984.  Sexual  selection  on  body  size,  ter- 
ritory and  plumage  variation  in  a population  of 
Darwin’s  finches.  Evolution  38:327-341. 

Robbins,  M.  B.  and  R.  S.  Ridgeley.  1994.  Voice, 
plumage  and  natural  history  of  Anthony’s  Nightjar 
Caprirnulgus  anthonyi.  Condor  96:224-228. 

Rohwer,  S.  1971.  Molt  and  the  annual  cycle  of  the 
Chuck-will’s-widow,  Caprirnulgus  carolinensis. 
Auk  88:485-519. 

Saetre,  G.  R,  S.  Dale,  and  T.  Slagsvold.  1994.  Fe- 
male Ried  Flycatchers  prefer  brightly  coloured 
males.  Anim.  Behav.  48:1407-1416. 

ScHLiNGER,  B.  A.  1990.  A non-parametric  aid  in  iden- 
tifying sex  of  cryptically  dimorphic  birds.  Wilson 
Bull.  102:545-550. 

Schwartz,  R.  1968.  Notes  on  two  Neotropical  night- 
jars, Caprirnulgus  anthonyi  and  C.  pan’ulus.  Con- 
dor 70:223-227. 

Selander,  R.  K.  1954.  A systematic  review  of  the 
booming  nighthawks  of  we.stern  North  America. 
Condor  56:57-82. 

Sick,  H.  1993.  Birds  in  Brazil.  A natural  history. 
Rrinceton  Univ.  Rress,  Rrinceton,  New  Jersey. 

Soares,  A.  A.  1973.  Sobre  Caprirnulgus  europaeus  e 
C.  ruficollis  em  Rortugal  (Aves,  Caprimulgifor- 
mes).  Arquivos  Museu  Bocage  4:215-233. 

SuNDBERG,  J.  AND  C.  Larsson.  1994.  Male  coloration 
as  an  indicator  of  parental  quality  in  the  Yellow- 
hammer,  Emheriza  citrinella.  Anim.  Behav.  48: 
885-892. 


Wilson  Bull.,  111(3),  1999,  pp.  346—353 


INTERSPECIFIC  INTERACTIONS  WITH  FORAGING 
RED-COCKADED  WOODPECKERS  IN  SOUTH-CENTRAL  FLORIDA 

REED  BOWMAN,'  5 DAVID  L.  LEONARD,  JR.,'  ^ LESLIE  K.  BACKUS,'  ^ AND 

ALLISON  R.  MAINS'^ 


ABSTRACT. — Inter.specific  competition  for  Red-cockaded  Woodpecker  (Picoicles  borealis)  cavities  has  been 
well  documented  and  may  be  one  factor  contributing  to  the  species’  decline.  Other  forms  of  interspecific  inter- 
actions have  rarely  been  documented  over  most  of  the  species’  range  and  have  received  little  attention.  During 
806  hours  of  Red-cockaded  Woodpecker  foraging  observations  in  south-central  Florida  we  documented  306 
interspecific  interactions  with  19  species.  We  observed  fewer  non-foraging  interactions  (98)  than  foraging  inter- 
actions (208).  Red-cockaded  Woodpeckers  lost  70  (71%)  of  the  non-foraging  interactions  and  177  (85%)  of  the 
foraging  interactions.  Most  non-foraging  interactions  (64%)  were  with  non-woodpecker  species,  several  of  which 
frequently  and  consistently  dominated  Red-cockaded  Woodpeckers.  Together,  Eastern  Kingbirds  (Tyrannus  tyr- 
annies), Great  Crested  Flycatchers  (Myiarclnis  crinitus).  Eastern  Bluebirds  (Sialia  sialis),  and  Pine  Warblers 
(Dendroica  pinus)  won  45  of  their  48  (94%)  non-foraging  interactions  with  Red-cockaded  Woodpeckers.  Most 
foraging  interactions  (97%)  were  with  other  woodpecker  species.  Red-bellied  Woodpeckers  (Melanerpes  caro- 
linits)  were  involved  in  172  (85%)  of  these  interactions,  of  which  they  won  168  (98%).  We  found  no  relationship 
between  the  rate  of  interactions  and  the  habitats  or  the  local  landscape  in  which  these  interactions  occuiTed. 
Red-cockaded  Woodpeckers  did  not  appear  to  move  to  different  and  possibly  less  productive  foraging  sites  after 
being  usurped.  In  south-central  Florida,  where  hardwood  basal  areas  are  relatively  low  in  Red-cockaded  Wood- 


pecker habitat,  the  foraging  niche  of  these  two  species 
range.  Received  20  July  1998,  accepted  5 Feb.  1999. 

The  Red-cockaded  Woodpecker  (Picoides 
borealis)  is  a cooperative  breeder  restricted  to 
the  old  growth  pine  forests  of  the  southeastern 
United  States  (Jackson  1971).  Despite  nearly 
30  years  of  Federal  protection,  Red-cockaded 
Woodpecker  populations  have  continued  to 
decline  (James  1991).  Habitat  loss  and  frag- 
mentation have  ultimately  been  responsible 
for  the  species’  decline  (Lennartz  et  al.  1983, 
Conner  and  Rudolph  1991).  Interspecific  com- 
petition for  Red-cockaded  Woodpecker  nest 
and  roost  cavities  has  been  well  documented 
(Jackson  1978,  Harlow  and  Lennartz  1983, 
Kappes  and  Harris  1995)  and  may  be  one 
proximate  factor  contributing  to  the  species’ 
decline  (U.S.  Fish  and  Wildlife  Service  1985). 

Interspecific  interactions,  other  than  those 
involving  cavities,  have  rarely  been  reported 
over  most  of  the  Red-cockaded  Woodpecker’s 


‘ Archbokl  Biological  Station,  PO  Box  2057,  Lake 
Placid,  FL  33862. 

j.'^  Dept,  of  Wildlife  Ecology  and  Conservation,  Univ. 
of  Florida,  PO  Box  1 1043,  Gainesville,  FL  3261  1. 

’James  San  Jacinto  Mountains  Re.serve,  Univ.  of 
California,  PO  Box  1775,  Idyllwild,  CA  92549. 

’ Dept,  of  Forestry,  Wildlife  and  Fisheries,  Univ.  of 
Tennessee,  Knoxville,  TN  37996. 

’ Corresponding  author; 

E-mail:  rbowman@archbold-station.org 


may  overlap  to  a greater  extent  than  elsewhere  in  then- 


range  (Morse  1970,  Nesbitt  et  al.  1978). 
Hooper  and  Lennartz  (1981)  observed  forag- 
ing Red-cockaded  Woodpeckers  from  May  to 
March  in  South  Carolina  and  documented  21 
interspecific  interactions  between  Red-cock- 
aded Woodpeckers  and  one  of  four  wood- 
pecker species  or  the  Brown-headed  Nuthatch 
{Sitta  pusilla).  Only  three  interactions  were  re- 
lated to  foraging.  Ligon  (1970)  reported  six 
interactions  between  Red-cockaded  Wood- 
peckers and  Downy  {Picoides  pubescens)  and 
Hairy  woodpeckers  {P.  villosus)  during  240 
hours  of  observations  from  May  to  December 
in  north-central  Florida.  In  contrast,  Nesbitt 
and  coworkers  (1981)  documented  149  inter- 
specific interactions  between  Red-cockaded 
Woodpeckers  and  five  woodpecker  species 
during  221  hours  of  observations  from  July  to 
October  in  southwestern  Florida.  Most  inter- 
actions involved  Red-bellied  Woodpeckers 
{Melanerpes  carolinus)  that  often  usurped 
Red-cockaded  Woodpeckers  from  foraging 
sites.  These  interactions  may  have  reduced  the 
caloric  intake  of  foraging  Red-cockaded 
Woodpeckers  (Nesbitt  et  al.  1981). 

Geographic  variation  in  interactions  be- 
tween species  is  common  (Travis  1996).  Ex- 
plaining this  variation  may  lead  to  a better  un- 
derstanding of  geographical  differences  in  be- 


346 


Bowman  et  al.  • RED-COCKADED  WOODPECKER  EORAGING  INTERACTIONS 


347 


havior,  demography,  and  habitat  selection  of 
potentially  interacting  species.  In  this  paper 
we  report  on  interspecific  interactions  with 
Red-cockaded  Woodpeckers  in  a small  popu- 
lation in  south-central  Florida. 

METHODS 

The  Avon  Park  Air  Force  Range  (APR)  i.s  a 42,900 
ha,  multiple-use,  active  military  training  installation  in 
Polk  and  Highlands  counties,  Florida.  Dominant  native 
pine  communities  consist  of  longleaf  (Finns  paliistris) 
and  south  Florida  slash  pine  (P.  elliotlii  var.  densa) 
and  approximately  9,000  ha  planted  in  north  Florida 
slash  pine  (P.  elliottii  var.  elliottii).  The  pine  habitats 
are  interspersed  with  other  communities  typical  of  this 
region  such  as  oak  scrub  and  fresh  water  marshes.  The 
natural  pine  habitats  support  the  characteristic  bird 
community  for  this  region  (Engstrom  1993),  including 
21  groups  of  Red-cockaded  Woodpeckers. 

To  determine  the  foraging  preferences  of  Red-cock- 
aded Woodpeckers  at  Avon  Park  AFR,  we  observed 
individuals  from  12  groups  once  a month  from  April 
1995  to  March  1996.  Red-cockaded  Woodpeckers 
were  observed  from  dawn  to  dusk  whenever  possible; 
observations  that  ended  prior  to  13:00  EST  were  re- 
peated. During  a foraging  observation  period,  we  re- 
corded the  location  of  the  focal  individual,  its  foraging 
maneuver,  and  substrate  use  at  10  minute  intervals. 
Locations  were  entered  into  a Geographical  Informa- 
tion System  (ArcView  GIS  Version  3.0).  We  deter- 
mined home  range  boundaries  and  overlaid  these 
boundaries  with  existing  habitat  type  coverages.  From 
these  maps,  we  calculated  the  area  of  each  habitat  type 
(13  categories)  in  each  home  range  and  linked  indi- 
vidual foraging  locations  to  specific  habitat  types. 
These  habitat  types  included  pine  flatwoods,  scrubby 
flatwoods,  oak  scrub,  sand  pine  scrub,  pine  plantation, 
mixed  natural  pine  and  plantation,  pine  swamp,  oak 
hammock,  hardwood  forests,  cypress,  marsh,  lake,  and 
human  disturbed. 

During  a subset  (806  hours)  of  the  total  observation 
time  (1168  hours),  we  documented  all  interspecific  in- 
teractions. We  recorded  the  species,  sex  (if  determin- 
able), type  of  aggression  (aerial  chase,  tree  chase, 
lunge,  usurp,  etc.),  and  the  outcome  (winner/loser).  In- 
dividuals that  retreated  without  retaliation  were  clas- 
sified as  losing.  We  categorized  interactions  as  forag- 
ing or  non-foraging  interactions.  Interactions  where  the 
winner  examined,  or  foraged  at,  the  usurped  site  were 
categorized  as  foraging  related.  All  other  interactions 
were  categorized  as  non-foraging.  Monthly  observa- 
tion periods  varied  as  did  the  number  of  individuals  in 
each  group.  To  avoid  observation  time  and  group  size 
biases,  we  used  only  those  interactions  that  involved 
the  breeding  pair  in  each  cluster  and  converted  those 
interactions  to  a rate  per  hour  for  all  analyses.  We  u.sed 
the  number  of  interactions  between  all  individuals  to 
describe  the  species  involved  in  interspecific  interac- 
tions with  Red-cockaded  Woodpeckers,  and  the  num- 
ber, type  and  outcome  of  those  interactions. 


Because  most  interactions  were  instantaneous  or  no 
longer  than  15-60  s (aerial  cha.ses),  we  assumed  inter- 
actions between  the  same  individuals  were  indepen- 
dent if  they  occuired  more  than  1 5 min  apart.  For  in- 
teractions that  occuiTed  less  than  15  min  apart,  we  ex- 
cluded all  but  the  first  interaction  as  long  as  the  type 
of  interaction  (foraging  or  non-foraging)  and  habitat 
were  the  same.  When  the  type  of  interaction  or  habitat 
differed,  we  excluded  all  the  interactions,  since  they 
could  not  be  aggregated  into  a single  type  of  interac- 
tion. However,  if  we  suspected  two  different  individ- 
uals of  the  other  species  (e.g.,  one  male  and  one  fe- 
male) were  involved  in  sequential  interactions  less 
than  15  min  apart  then  both  observations  were  consid- 
ered independent. 

To  examine  whether  the  frequency  of  interspecific 
interactions  was  habitat  specific,  we  compared  the  fre- 
quency of  interactions  per  habitat  type  to  the  expected 
frequency  based  on  the  proportion  of  time  Red-cock- 
aded Woodpeckers  foraged  in  each  habitat  type.  To 
determine  if  the  frequency  of  interspecific  interactions 
was  related  to  the  local  landscape,  we  compared  the 
frequency  of  interactions  per  group  to  the  mean  basal 
area  of  pines  and  hardwoods  in  each  Red-cockaded 
Woodpecker’s  home  range.  Given  the  frequency  of  for- 
aging interactions  between  Red-cockaded  and  Red-bel- 
lied woodpeckers,  we  repeated  the  above  analyses  for 
those  interactions. 

To  determine  whether  usurpations  had  a measurable 
effect  on  Red-cockaded  Woodpecker  foraging  patterns 
we  performed  two  analyses.  First,  we  compared  the 
habitats  used  by  Red-cockaded  Woodpeckers  before 
and  after  interactions  with  Red-bellied  Woodpeckers 
to  determine  whether  the  former  moved  to  a different 
and  potentially  less  productive  habitat  after  an  inter- 
action. Second,  we  compared  foraging  tree  character- 
istics [dbh  (diameter  at  breast  height)  and  height]  and 
Red-cockaded  Woodpecker  foraging  height  before  and 
after  usurpations  by  Red-bellied  Woodpeckers  to  de- 
termine whether  they  moved  to  different  micro-sites 
after  interactions.  Male  and  female  Red-cockaded 
Woodpeckers  forage  at  different  locations  (Ligon 
1970),  therefore  we  analyzed  each  sex  separately.  All 
statistical  tests  were  nonparametric  and  were  per- 
formed in  the  Microsoft  Windows  95  operating  system 
using  SPSS  (version  8.0). 

RESULTS 

We  observed  306  independent  interspecific 
interactions  between  45  color-banded  Red- 
cockaded  Woodpeckers  and  19  other  bird  spe- 
cies (Table  1 ).  Interactions  involved  26  breed- 
ing adult  Red-cockaded  Woodpeckers  (13  <3, 
13  $),  10  hatch-year  birds  (6  d,  4 9),  6 older 
helpers  (all  c3),  and  2 floaters  (both  6).  Of  the 
306  interactions  observed,  203  occurred  with 
the  breeding  Red-cockaded  Woodpeckers,  50 
with  hatch-year  birds,  16  with  older  helpers, 
and  6 with  floaters.  Red-cockaded  Woodpeck- 


348 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  3,  September  1999 


TABLE  1.  Species  observed  interacting  with  Red-cockaded  Woodpeckers,  the  outcome  (loser  or  winner), 
and  the  type  (foraging  or  non-foraging)  of  interaction  during  806  hours  of  foraging  observations  at  the  Avon 
Park  Air  Force  Range,  1995-1996. 


Red-cockaded  Woodpecker 

Loser 

Red-cockaded  Woodpecker 

Winner 

Species 

Foraging 

Non-foraging 

Foraging 

Non-foraging 

Red-shouldered  Hawk 



1 



1 

Red-bellied  Woodpecker 

168 

6 

4 

8 

Red-headed  Woodpecker 

— 

3 

— 

— 

Northern  Flicker 

1 

— 

1 

2 

Yellow-bellied  Sapsucker 

— 

— 

— 

2 

Downy  Woodpecker 

2 

— 

23 

11 

Hairy  Woodpecker 

3 

— 

— 

3 

Eastern  Kingbird 

1 

14 

— 

— 

Great  Crested  Flycatcher 

— 

1 1 

— 

— 

Florida  Scrub-Jay 

— 

1 

— 

— 

Blue  Jay 

2 

— 

— 

— 

Brown-headed  Nuthatch 

— 

1 

1 

— 

Eastern  Bluebird 

— 

8 

— 

1 

Loggerhead  Shrike 

— 

3 

— 

— 

Northern  Mockingbird 

— 

3 

— 

— 

Pine  Warbler 

— 

12 

2 

— 

Eastern  Towhee 

— 

2 

— 

— 

Red-winged  Blackbird 

— 

4 

— 

— 

Summer  Tanager 

— 

1 

— 

— 

Total  # of  Interactions 

177 

70 

31 

28 

Total  # of  Species 

6 

14 

5 

7 

ers  lost  247  (81%)  interactions  to  18  species 
and  won  59  (19%)  interactions  with  9 species 
(Table  1).  Windoss  ratios  for  Red-cockaded 
Woodpeckers  did  not  differ  between  life  his- 
tory stage  (x^  = 1.23,  df  = 3,  P > 0.05)  or 
sex  (x^  = 1.49,  df  = 1,  P > 0.05). 

Interactions  between  Red-cockaded  Wood- 
peckers and  other  woodpecker  species  were 
most  frequent,  accounting  for  237  (77%)  of 
all  interactions.  Excluding  species  with  fewer 
than  five  observed  interactions,  five  species 
won  more  than  85%  of  their  interactions  with 
Red-cockaded  Woodpeckers  [Eastern  King- 
bird (Tyrannus  tyrannus),  100%;  Great  Crest- 
ed Flycatcher  {Myiarchus  crinitus),  100%; 
Red-bellied  Woodpecker,  94%;  Eastern  Blue- 
bird {Sicilia  sialis),  89%;  and  Pine  Warbler 
(Dendroica  pinus),  86%;  Table  1].  The 
Downy  Woodpecker  was  the  only  species 
Red-cockaded  Woodpeckers  consistently 
dominated  (34  of  36  encounters). 

The  rate  of  interspecific  interactions  with 
breeding  Red-cockaded  Woodpeckers  was 
greatest  in  June  and  July.  Although  interac- 
tions varied  between  months  from  0.10  (± 
0.03  SE)  to  0.40  (±  0.31)  interactions  per 


hour  (Fig.  la),  these  differences  were  not  sig- 
nificant (Kruskal-Wallis  One  Way  ANOVA: 
X“  = 16.7,  df  = 11,  P > 0.05). 

Non-foraging  interactions. — Red-cockaded 
Woodpeckers  had  98  non-foraging  interac- 
tions with  18  species;  however,  they  had  only 
non-foraging  interactions  with  12  of  those 
species  (Table  1).  Interactions  with  Red-head- 
ed Woodpeckers  {Melanerpes  erythocephal- 
us).  Eastern  Kingbirds,  Great  Crested  Fly- 
catchers, Northern  Mockingbirds  {Mirnus  po- 
lyglottos),  and  Pine  Warblers  often  involved 
aerial  chases  that  lasted  from  15-60  s.  Few  of 
these  interactions  were  initiated  by  Red-cock- 
aded Woodpeckers.  However,  in  one  instance 
a group  of  Red-cockaded  Woodpeckers 
mobbed  and  successfully  evicted  a Red-shoul- 
dered Hawk  {Buteo  lineatus).  Of  the  98  inter- 
actions, Red-cockaded  Woodpeckers  won  28 
and  lost  70. 

Non-foraging  interactions  were  most  fre- 
quent during  June  and  July  (Fig.  lb)  and 
monthly  differences  were  statistically  different 
(Kruskal-Wallis  One  Way  ANOVA:  x^  ~ 
23.9,  df  = 11,  P = 0.013). 

Foraging  interactions. — Red-cockaded 


Bowman  el  at.  • RED-COCKADED  WOODPECKER  FORAGING  INTERACTIONS 


349 


Month 


Month 


FIG.  1.  Interspecific  interactions  (T  ± 1 SE  for  each  month)  observed  per  hour  with  breeding  adult  Red- 
cockaded  Woodpeckers  at  Avon  Park  Air  Force  Range,  in  south-central  Florida,  1995-1996;  (a)  all  interactions, 
(b)  non-foraging  interactions,  (c)  foraging  interactions,  and  (d)  all  interactions  with  Red-bellied  Woodpeckers. 


Woodpeckers  had  208  foraging  interactions 
with  eight  species.  With  six  of  these  species 
they  had  both  foraging  and  non-foraging  in- 
teractions, but  foraging  interactions  were 
more  frequent  than  non-foraging  interactions 
(Table  1 ).  Most  foraging  interactions  were  be- 
tween Red-cockaded  Woodpeckers  and  other 
woodpeckers  (202  of  208,  97%),  but  interac- 
tions with  Blue  Jays  (Cyanocitta  cristata). 
Eastern  Kingbirds,  Brown-headed  Nuthatches, 
and  Pine  Warblers  also  were  observed.  Red- 
cockaded  Woodpeckers  lost  most  (177  of  208, 
85.1%)  foraging  interactions;  however,  74% 
of  the  31  interactions  they  won  were  with 
Downy  Woodpeckers.  Red-cockaded  Wood- 
peckers lost  a greater  percentage  of  foraging 
interactions  (85.1%)  than  they  did  non-forag- 
ing interactions  (71.4%;  = 7.14,  df  = 1,  P 

= 0.008).  None  of  the  foraging  interactions 
were  initiated  by  Red-cockaded  Woodpeckers, 
except  for  those  with  Downy  Woodpeckers. 


Downy  Woodpeckers  frequently  foraged  near 
Red-cockaded  Woodpeckers  and  often  were 
aggressively  chased  and  their  foraging  loca- 
tions usurped.  The  rate  of  foraging  interac- 
tions with  breeding  Red-cockaded  Woodpeck- 
ers did  not  vary  monthly  (Kruskal-Wallis  One 
Way  ANOVA:  x"  = 14.2,  df  = 1 1,  P = 0.22; 
Fig.  Ic). 

Red-bellied  Woodpeckers. — Most  interspe- 
cific interactions  occurred  between  Red-cock- 
aded and  Red-bellied  woodpeckers  (186  of 
306,  61%).  Of  the  186  interactions  between 
Red-cockaded  and  Red-bellied  woodpeckers, 
the  latter  won  174  (94%).  Red-bellied  Wood- 
peckers successfully  usurped  foraging  Red- 
cockaded  Woodpeckers  in  all  but  4 of  168  for- 
aging interactions.  Red-bellied  Woodpeckers 
frequently  foraged  within  sight  of  Red-cock- 
aded Woodpeckers  but  usurped  them  only  af- 
ter the  Red-cockaded  Woodpecker  had  found 
food.  We  also  observed  Red-bellied  Wood- 


350 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


TABLE  2.  Variation  in  habitat  characteristics  and  the  rate  (x  ± 1 SE)  of  all  interspecific  interactions  (per 
hour  of  observation),  foraging  and  non-foraging  interactions,  and  interactions  with  Red-bellied  Woodpeckers, 
with  breeding  Red-cockaded  Woodpeckers  from  different  groups  at  the  Avon  Park  Air  Eorce  Range,  1995-1996. 


Hardwood 


RCW  group 

Pine  basal  area 
(m-/ha) 

basal  area 
( m-/ha) 

All  interactions 

Foraging 

interactions 

Non-foraging 

interactions 

Red-bellied  Woodpecker 
interactions 

1 

8.13 

0.0 

0.16 

-h 

0.09 

0.11 

4- 

0.07 

0.05  ± 

0.04 

0.10 

4- 

0.06 

2 

12.44 

0.0 

0.19 

-h 

0.05 

0.15 

4- 

0.05 

0.04  ± 

0.04 

0.11 

4- 

0.04 

3 

8.97 

0.0 

0.20 

-F 

0.06 

0.17 

4- 

0.01 

0.04  ± 

0.02 

0.14 

4- 

0.06 

5 

6.6 

0.28 

0.20 

4- 

0.06 

0.15 

4- 

0.01 

0.06  ± 

0.03 

0.14 

4- 

0.06 

7 

7.49 

0.07 

0.10 

4- 

0.05 

0.06 

4- 

0.03 

0.04  ± 

0.04 

0.04 

4- 

0.02 

8 

9.28 

0.04 

0.10 

4- 

0.06 

0.04 

4- 

0.02 

0.06  ± 

0.06 

0.04 

4- 

0.02 

15 

1 1.34 

0.03 

0.42 

4- 

0.09 

0.27 

4- 

0.09 

0.15  ± 

0.05 

0.22 

4- 

0.09 

19 

9.18 

0.46 

0.22 

4- 

0.07 

0.18 

4- 

0.07 

0.04  ± 

0.02 

0.15 

4- 

0.05 

21 

10.25 

0.44 

0.17 

4- 

0.11 

0.11 

4- 

0.07 

0.05  ± 

0.04 

0.11 

4- 

0.07 

23 

6.55 

0.01 

0.42 

4- 

0.15 

0.34 

4- 

0.13 

0.08  ± 

0.03 

0.35 

4- 

0.14 

31 

12.15 

0.86 

0.22 

4- 

0.09 

0.16 

4- 

0.08 

0.07  ± 

0.03 

0.13 

4- 

0.07 

33 

12.33 

0.0 

0.34 

4- 

0.09 

0.24 

4- 

0.07 

0.09  ± 

0.04 

0.22 

4- 

0.02 

peckers  following  Red-cockaded  Woodpeck- 
ers as  they  foraged  between  different  pine 
stands.  Red-cockaded  Woodpeckers  won  only 
12  interactions  with  Red-bellied  Woodpeck- 
ers: 4 foraging  interactions  and  10  non-for- 
aging interactions. 

The  rate  of  interactions  between  Red-bel- 
lied Woodpeckers  and  breeding  Red-cockaded 
Woodpeckers  varied  monthly  (Kruskal-Wallis 
One  Way  ANOVA:  x'  = 19.3,  df  = 11,  P = 
0.055;  Fig.  Id);  however,  no  consistent  pattern 
was  evident.  The  rate  of  interactions  did  not 
differ  between  the  breeding  and  non-breeding 
season  (Mann-Whitney  [/-test:  Z = -0.59,  P 
> 0.05).  No  sex-related  difference  existed  in 
the  rate  of  interactions  between  Red-cockaded 
and  Red-bellied  woodpeckers  (x^  = 5.13,  df 
= 1,  P > 0.05). 

Habitat  relationships. — Red-cockaded 
Woodpeckers  foraged  predominately  (93.8% 
of  observation  time)  in  pine  flatwood,  scrubby 
flatwood,  and  pine  plantation  habitats  (Bow- 
man et  al.  1998,  unpubl.  data).  The  frequency 
of  foraging  and  non-foraging  interactions  did 
not  differ  from  the  relative  frequency  of  hab- 
itats used  by  foraging  Red-cockaded  Wood- 
peckers (x^  — 4.74  and  2.18,  df  = 4 and  2, 
respectively,  P > 0.05).  We  also  found  no  sig- 
nificant correlations  between  the  rate  of  inter- 
actions and  the  area  of  any  of  the  13  habitat 
types  in  Red-cockaded  Woodpecker  home 
ranges  (Pearson  correlations:  all  P > 0.05). 
Pine  basal  area  in  home  ranges  varied  from 


6.6  to  12.4  m^  per  ha,  and  hardwood  basal 
area  varied  from  0.0  to  0.44  m^  per  ha  (Table 
2);  however,  neither  the  total  number  of  inter- 
actions nor  foraging  or  non-foraging  interac- 
tions were  correlated  with  the  basal  area  of 
pines  or  hardwoods  within  each  home  range. 

Interactions  with  Red-bellied  Woodpeckers 
in  different  habitat  types  did  not  differ  from 
the  relative  frequency  of  habitats  used  by  Red- 
cockaded  Woodpeckers  (x^  = 0.18,  df  = 3,  P 
> 0.05).  However,  the  rate  of  these  interac- 
tions was  positively  correlated  with  the  per- 
centage of  each  home  range  comprised  of  pine 
plantation  (Spearman  rank  correlation:  r = 
0.62,  P < 0.05).  No  significant  correlations 
existed  between  the  rate  of  interactions  and 
the  area  of  any  of  the  other  12  habitat  types 
or  the  pine  or  hardwood  basal  area  in  Red- 
cockaded  Woodpecker  home  ranges  (Spear- 
man rank  correlation:  all  P > 0.05). 

Red-cockaded  Woodpeckers  moved  to  a 
new  habitat  type  following  only  5 of  120  (4%) 
usurpations  by  Red-bellied  Woodpeckers  for 
which  we  had  data.  No  significant  difference 
existed  in  the  dbh  or  height  of  the  trees  used 
by  foraging  Red-cockaded  Woodpeckers 
(male  or  female)  before  and  after  usurpation 
(Kruskal-Wallis  One  Way  ANOVAs:  all  P > 
0.05),  nor  did  any  differences  exist  in  the 
height  at  which  Red-cockaded  Woodpeckers 
(male  or  female)  foraged  before  and  after 
usurpation  (Kruskal-Wallis  One  Way  ANO- 
VAs: all  P > 0.05). 


liowimin  et  al.  • RED-COCKADED  WOODPECKER  FORAGING  INTERACTIONS 


351 


DISCUSSION 

Thirty-two  percent  of  interspecific  interac- 
tions were  not  related  to  foraging.  Non-for- 
aging  interactions  were  highly  seasonal,  oc- 
cuiTing  during  the  breeding  season  for  most 
species.  Many  of  these  interactions  may  have 
been  related  to  nest  and/or  fledgling  defense 
as  many  occurred  near  nests  or  young  of  the 
species  interacting  with  Red-cockaded  Wood- 
peckers. Although  these  interactions  were  sea- 
sonal and  relatively  infrequent  in  our  popu- 
lation, other  forms  of  non-foraging  interac- 
tions (e.g.,  cavity  competition)  could  play  an 
important  role  in  the  dynamics  of  Red-cock- 
aded Woodpecker  populations  (Kappes  and 
Harris  1995). 

Most  interspecific  interactions  were  related 
to  foraging  and  occurred  between  Red-cock- 
aded and  Red-bellied  woodpeckers.  Red-cock- 
aded Woodpeckers  lost  virtually  all  foraging 
interactions  with  Red-bellied  Woodpeckers. 
Red-cockaded  Woodpeckers  interacted  fre- 
quently with  Downy  Woodpeckers,  winning 
most  encounters.  Therefore,  the  latter  inter- 
actions likely  had  no  deleterious  impacts  on 
Red-cockaded  Woodpeckers. 

Habitat  use,  foraging  behavior,  and  diet  of 
Red-cockaded  and  Red-bellied  woodpeckers 
appear  to  be  dissimilar.  Red-bellied  Wood- 
peckers use  most  habitats  occurring  within 
their  range  (Sprunt  1954,  Breitwisch  1977  and 
references  within)  but  may  prefer  hardwood 
habitats  (Short  1982,  Root  1988).  In  Florida, 
their  use  of  tree  species  for  foraging  is  diverse 
and  varies  by  habitat  type  (Breitwisch  1977). 
Red-bellied  Woodpeckers  spend  20-69%  of 
their  foraging  time  on  dead  trees  (Williams 
1975,  Breitwisch  1977,  Williams  and  Batzli 
1979).  In  contrast,  Red-cockaded  Woodpeck- 
ers forage  almost  exclusively  on  living  pines 
(Hooper  and  Lennartz  1981 ) in  relatively  open 
pine  forests. 

In  south  Florida  pine  habitat,  Breitwisch 
(1977)  observed  foraging  Red-bellied  Wood- 
peckers gleaning  and  probing  (80%)  but  rarely 
excavating  (10%).  At  Avon  Park  AFR,  for- 
aging Red-cockaded  Woodpeckers  used  sur- 
face probes  (54%)  most  frequently,  excavated 
frequently  (40%),  and  rarely  gleaned  (4%; 
Bowman  et  al.l998,  unpubl.  data). 

Little  dietary  overlap  appears  to  exist  be- 
tween Red-cockaded  and  Red-bellied  wood- 


peckers (Beal  191 1).  Red-bellied  Woodpecker 
stomachs  (n  — 271)  contained  3 1 % animal 
matter,  of  which  6%  was  ants;  Red-cockaded 
Woodpecker  stomachs  (n  = 76)  contained 
81%  animal  matter,  of  which  56%  was  ants. 
Both  species  consumed  a similar  percentage 
of  beetles  (~10%);  however,  little  overlap  ex- 
isted in  the  remaining  fraction  of  animal  mat- 
ter. 

Niche  overlap  between  these  two  species 
appears  to  be  low,  even  in  south  Florida,  yet 
interactions  between  Red-bellied  and  Red- 
cockaded  woodpeckers  appear  to  be  higher 
here  than  elsewhere  in  their  ranges.  It  is  pos- 
sible that  these  interactions  are  simply  over- 
looked elsewhere,  especially  if  they  are  more 
frequent  outside  of  the  breeding  season.  If  so, 
and  these  interactions  have  deleterious  im- 
pacts on  Red-cockaded  Woodpeckers,  then 
they  should  be  examined  more  closely  else- 
where. However,  geographical  variation  in  in- 
terspecific competition  may  be  real  and  be 
caused  by  variation  in  population  densities  of 
the  species  (Thompson  1988),  indirect  effects 
as  species  assemblages  change,  the  productiv- 
ity or  vegetation  composition  of  habitats 
(Travis  1 996)  or  some  interaction  of  these  fac- 
tors. 

Data  on  the  density  of  Red-bellied  Wood- 
peckers across  their  range  are  not  available;  in 
general  they  appear  as  abundant  in  Florida  as 
elsewhere  in  the  southeastern  coastal  plain 
(Bock  and  Lepthien  1975,  Root  1988,  Price  et 
al.  1995).  At  Avon  Park  AFR,  the  density  of 
Red-cockaded  Woodpeckers  is  low  compared 
to  populations  outside  of  peninsular  Florida 
(Bowman  et  al.  1998,  unpubl.  data).  Data  on 
the  regional  variation  in  density  of  both  Red- 
cockaded  and  Red-bellied  woodpeckers  are 
needed  to  determine  whether  differences  in 
density  contribute  to  variations  in  interspecific 
interactions. 

Indirect  effects  related  to  the  presence  of 
other  species  may  have  contributed  to  the  high 
rate  of  observed  interactions.  At  Avon  Park 
AFR,  five  species  of  woodpeckers  and  the 
Brown-headed  Nuthatch  are  sympatric  with 
Red-cockaded  Woodpeckers;  however,  many 
of  these  species  are  sympatric  in  pine  habitats 
outside  of  peninsular  Florida.  The  abundance 
and  diversity  of  species  utilizing  similar  re- 
sources in  different  habitats  may  contribute  to 


352 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


variation  in  the  rate  of  interspecific  interac- 
tions. 

Differences  in  pine  forests  between  south- 
central  Florida  and  more  temperate  forests 
may  have  contributed  to  the  relatively  high 
rates  of  interactions  with  Red-bellied  Wood- 
peckers. In  southern  Florida,  most  Red-cock- 
aded  Woodpeckers  occur  in  mesic  and  hydric 
flatwoods.  These  habitats  have  lower  hard- 
wood basal  area  than  do  more  temperate  pine 
communities  (Beever  and  Dryden  1992,  pers. 
obs.).  Elsewhere,  Red-bellied  Woodpeckers’ 
preference  for  hardwoods  may  minimize  their 
foraging  overlap  with  Red-cockaded  Wood- 
peckers, but  we  know  little  about  habitat-spe- 
cific foraging  strategies  of  either  species.  Al- 
though hardwood  basal  area  varied  among  the 
12  Red-cockaded  Woodpecker  home  ranges, 
overall,  basal  area  was  low  and  was  not  cor- 
related with  the  frequency  of  Red-bellied 
Woodpecker  interactions. 

All  Red-cockaded  Woodpecker  populations 
in  peninsular  Florida  support  fewer  than  50 
groups  (Cox  et  al.  1995).  In  peninsular  Flor- 
ida, Red-cockaded  Woodpeckers  have  larger 
home  ranges  (Nesbitt  et  al.  1981;  DeLotelle 
et  al.  1983;  Bowman  et  al.  1998,  unpubl.  data) 
and  produce  fewer  fledglings  (Jansen  and  Pat- 
terson 1983;  DeLotelle  and  Epting  1992; 
Bowman  et  al.  1998,  unpubl.  data)  than  other 
populations.  These  characteristics  suggest  that 
these  populations  may  occupy  relatively  poor 
quality  habitat;  however,  few  correlations  ex- 
ist between  various  measures  of  Red-cockad- 
ed Woodpecker  demography  and  habitat  char- 
acteristics (Beyer  et  al.  1996;  Bowman  et  al. 
1998,  unpubl.  data).  Although  these  results  do 
not  suggest  a deleterious  effect  of  interspecific 
competition,  the  relatively  high  rates  we  doc- 
umented bear  further  investigation,  especially 
where  these  interactions  have  not  been  re- 
ported. Aggressive  interaction  between  spe- 
cies is  not  sufficient  to  demonstrate  competi- 
tion, but  interspecific  competition  may  con- 
tribute to  variation  in  the  abundance  and  re- 
productive potential  of  species.  It  is  possible 
that  some  synergistic  interaction  of  habitat 
and  community  structure,  such  as  competition, 
may  be  related  to  regional  differences  in  Red- 
cockaded  Woodpecker  demography. 

ACKNOWLEDGMENTS 

Financial  support  was  provided  by  the  United  States 
Department  of  Defense.  We  thank  P.  Ebersbach,  B. 


Progulske,  P Walsh,  and  K.  Olson  for  cooperation 
throughout  the  course  of  this  study.  T.  Dean,  D.  Barber, 
P.  Barber,  and  D.  Swan  provided  very  helpful  com- 
ments on  early  drafts  of  this  manuscript  and  R.  Conner, 
E James,  C.  Rudolph  and  J.  Thomlinson  provided  con- 
structive reviews  of  a later  version. 

LITERATURE  CITED 

Beal,  F.  E.  L.  1911.  Food  of  the  woodpeckers  of  the 
United  States.  U.S.  Dept.  Agric.  Biol.  Surv.  Bull., 
37:1-64. 

Beever,  J.  W.  and  K.  A.  Dryden.  1992.  Red-cockaded 
Woodpeckers  and  hydric  slash  pine  flatwoods. 
Trans.  N.  Am.  Wildl.  Nat.  Res.  Conf.  57:693-700. 
Beyer,  D.  E.,  R.  Costa,  R.  G.  Hooper,  and  C.  A. 
Hess.  1996.  Habitat  quality  and  reproduction  of 
Red-cockaded  Woodpecker  groups  in  Florida.  J. 
Wildl.  Manage.  60:826—835. 

Bock,  C.  E.  and  L.  W.  Lepthien.  1975.  A Christmas 
count  analysis  of  woodpecker  abundance  in  the 
United  States.  Wilson  Bull.  87:355-366. 

Bowman,  R.,  D.  L.  Leonard,  Jr.,  L.  K.  Backus,  P.  M. 
Barber,  A.  R.  Mains,  L.  M.  Richman,  and  D. 
Swan.  1998.  Demography  and  habitat  character- 
istics of  the  Red-cockaded  Woodpecker  (Picoides 
borealis)  at  the  Avon  Park  Air  Force  Range:  Final 
Report,  1994-1998.  Department  of  Defense, 
Avon  Park,  Florida. 

Breitwisch,  R.  j.  1977.  The  ecology  and  behavior  of 
the  Red-bellied  Woodpecker,  Centurus  carolinus 
(Linnaeus)  (Aves:  Picidae),  in  south  Florida. 
M.Sc.  thesis,  Univ.  of  Miami,  Miami,  Florida. 
Conner,  R.  N.  and  D.  C.  Rudolph.  1991.  Forest  hab- 
itat loss,  fragmentation,  and  Red-cockaded  Wood- 
pecker populations.  Wilson  Bull.  103:446—457. 
Cox,  J.,  W.  W.  Baker,  and  D.  Wood.  1995.  Status, 
distribution,  and  conservation  of  the  Red-cock- 
aded Woodpecker  in  Florida:  a 1992  update.  Pp. 
475-464  in  Red-cockaded  Woodpecker:  recovery, 
ecology,  and  management  (D.  L.  Kulhavey,  R.  G. 
Hooper,  and  R.  Costa,  Eds.).  College  of  Forestry, 
Stephen  F.  Austin  State  Univ.,  Nacogdoches,  Tex- 
as. 

DeLotelle,  R.  S.  and  R.  J.  Epting.  1992.  Reproduc- 
tion of  the  Red-cockaded  Woodpecker  in  central 
Florida.  Wil.son  Bull.  104:285—294. 

DeLotelle,  R.  S.,  J.  R.  Newman,  and  A.  E.  Jerauld. 
1983.  Habitat  use  by  Red-cockaded  Woodpeckers 
in  central  Florida.  Pp.  59-67  in  Proceedings  of 
the  Red-cockaded  Woodpecker  symposium  11 
(D.A.  Wood,  Ed.).  Florida  Game  and  Fresh  Water 
Fish  Comm.,  Tallahassee. 

Engstrom,  R.  T.  1993.  Characteristic  mammals  and 
birds  of  longleaf  pine  forests.  Proc.  Tall  Timbers 
Fire  Ecol.  Conf.  18:127-138. 

Harlow,  R.  F.  and  M.  R.  Lennartz.  1983.  Interspe- 
cific competition  for  Red-cockaded  Woodpecker 
cavities  during  the  nesting  season  in  South  Caro- 
lina. Pp.  41-43  in  Red-cockaded  Woodpecker 
symposium  II  (D.  A.  Wood,  Ed.).  Florida  Game 
and  Fresh  Water  Fish  Comm.,  Tallahassee. 


Bowman  et  al.  • RED-COCKADED  WOODPECKER  EORAGING  INTERACTIONS 


353 


Hooper,  R.  G.  and  M.  R.  Lennartz.  1981.  Eoraging 
behavior  of  the  Red-eockaded  Woodpecker  in 
South  Carolina.  Auk  98:321-334. 

Jackson,  J.  A.  1971.  The  evolution,  taxonomy,  distri- 
bution, past  populations  and  current  status  of  the 
Red-cockaded  Woodpecker.  Pp.  4-29  in  The  ecol- 
ogy and  management  of  the  Red-cockaded  Wood- 
pecker (R.  L.  Thompson,  Ed.).  Bur.  Sport  Eish. 
Wildl.,  U.S.  Dept,  of  Interior,  and  Tall  Timbers 
Research  Station,  Tallahassee,  Florida. 

Jackson,  J.  A.  1978.  Competition  for  cavities  and  Red- 
cockaded  Woodpecker  management.  Pp.  103-112 
in  Endangered  birds:  management  techniques  for 
threatened  species  (S.  A.  Temple,  Ed.).  Univ.  of 
Wisconsin  Press,  Madison. 

James,  E C.  1991.  Signs  of  trouble  in  the  largest  re- 
maining population  of  Red-cockaded  Woodpeck- 
ers. Auk  108:419-423. 

Jansen,  D.  K.  and  G.  A.  Patterson.  1983.  Late  nest- 
ing and  low  reproductive  success  of  Red-cockad- 
ed Woodpeckers  in  the  Big  Cypress  National  Pre- 
serve, Florida,  in  1983.  Pp.  99  in  Red-cockaded 
Woodpecker  symposium  II  (D.  A.  Wood,  Ed.). 
Florida  Game  and  Fresh  Water  Fish  Comm.,  Tal- 
lahassee. 

Kappes,  j.,  Jr.  and  L.  D.  Harris.  1995.  Interspecific 
competition  for  Red-cockaded  Woodpecker  cavi- 
ties in  Apalachicola  National  Forest.  Pp.  389-393 
in  Red-cockaded  Woodpecker:  recovery,  ecology, 
and  management  (D.  L.  Kulhavey,  R.  G.  Hooper, 
and  R.  Costa,  Eds.).  College  of  Forestry,  Stephen 
F.  Austin  State  Univ.,  Nacogdoches,  Texas. 

Lennartz,  M.  R.,  P.  H.  Geissler,  R.  F.  Harlow,  R.  C. 
Long,  K.  M.  Chitwood,  and  J.  A.  Jackson.  1983. 
Status  of  the  Red-cockaded  Woodpecker  on  fed- 
eral lands  in  the  south.  Pp.  7—12  in  Red-cockaded 
Woodpecker  Symposium  II  (D.  A.  Wood,  Ed.). 


Florida  Game  Fresh  Water  Eish  Comm.,  Tallahas- 
see. 

Ligon,  j.  D.  1970.  Behavior  and  breeding  biology  of 
the  Red-cockaded  Woodpecker.  Auk  87:255-278. 

Morse,  D.  H.  1970.  Ecological  aspects  ol'some  mixed- 
species  foraging  flocks  of  birds.  Ecology  40: 1 19- 
168. 

Nesbitt,  S.  A.,  D.  T.  Gilbert,  and  D.  B.  Barbour. 
1978.  Red-cockaded  Woodpecker  fall  movements 
in  a Florida  flatwood  community.  Auk  95:145- 
151. 

Nesbitt,  S.  A.,  B.  A.  Harris,  A.  E.  Jerauld,  and 
C.  B.  Brownsmith.  1981.  Report  of  the  investi- 
gation of  Red-cockaded  Woodpeckers  in  Charlotte 
County.  Florida  Game  and  Fresh  Water  Fish 
Comm.,  Tallahassee. 

Price.  J.,  S.  Droege,  and  A.  Price.  1995.  The  summer 
atlas  of  North  American  birds.  Academic  Press. 
San  Diego,  California. 

Root,  T.  1988.  Atlas  of  wintering  North  American 
birds.  Univ.  of  Chicago  Press,  Chicago,  Illinois. 

Short,  L.  L.  1982.  Woodpeckers  of  the  world.  Dela- 
ware Museum  of  Natural  History,  Greenville. 

Sprunt,  a.,  Jr.  1954.  Florida  bird  life.  Coward-Mc- 
Cann,  Inc.,  New  York. 

Thompson,  J.  N.  1988.  Variation  in  interspecific  inter- 
actions. Annu.  Rev.  Ecol.  Syst.  19:65-87. 

Travis,  J.  1996.  The  significance  of  geographic  vari- 
ation in  species  interactions.  Am.  Nat.  148:S1-S8. 

U.S.  Fish  and  Wildlife  Service.  1985.  Red-cockaded 
Woodpecker  recovery  plan.  U.S.  Fish  and  Wildlife 
Service,  Atlanta,  Georgia. 

Williams,  J.  B.  1975.  Habitat  utilization  by  four  spe- 
cies of  woodpeckers  in  a central  Illinois  wood- 
land. Am.  Midi.  Nat.  93:354-367. 

Williams,  J.  B.  and  G.  O.  Batzli.  1979.  Interference 
competition  and  niche  shifts  in  the  bark-foraging 
guild  in  central  Illinois.  Wilson  Bull.  91 :400-41 1. 


Wilson  Bull..  111(3),  1999,  pp.  354-362 


SPATIAL  AND  TEMPORAL  DYNAMICS  OF  A PURPLE  MARTIN 

PRE-MIGRATORY  ROOST 

KEVIN  R.  RUSSELL'  23  AND  SIDNEY  A.  GAUTHREAUX,  JR.‘ 


ABSTRACT. — We  used  simultaneous  WSR-88D  radar  (NEXRAD)  and  direct  visual  observations  to  investi- 
gate the  spatial  and  temporal  dynamics  of  a Purple  Martin  (Prague  suhis)  pre-migratory  roost  in  South  Carolina. 
The  timing  of  mass  flights  of  martins  from  and  to  the  roost  was  related  to  levels  of  ambient  light.  Each  morning, 
the  birds  first  departed  approximately  40  min  before  sunrise  independent  of  date,  with  peak  departures  occurring 
about  10  min  before  sunrise.  The  time  of  evening  flights  was  more  variable,  but  peak  movement  of  birds  into 
the  roost  consistently  occurred  at  sunset.  Purple  Martins  exited  the  roost  in  organized,  annular  departures  (360°) 
that  were  visible  on  radar  up  to  100  km  away  from  the  roost,  but  returned  to  the  roost  over  an  extended  period 
in  scattered  flocks.  During  morning  departures  we  recorded  flight  speeds  up  to  13.4  m/s.  Radar  echoes  corre- 
sponding to  martin  flights  were  recorded  farther  from  the  roost,  and  flights  from  and  to  the  roost  occuiTed  later 
and  earlier,  respectively,  in  response  to  increased  cloud  cover.  The  departures  of  birds  from  the  roost  appeared 
to  be  displaced  by  winds  aloft.  At  the  peak  of  the  roosting  season  in  late  July,  the  total  roost  population  was 
estimated  to  be  at  least  700,000  birds.  Received  18  Aiig.  1998,  accepted  5 Feb.  1999. 


Purple  Martins  (Progne  siibis)  are  neotrop- 
ical migratory  swallows  that  breed  across 
North  America  (Brown  1997,  AOU  1998).  In 
eastern  North  America,  Purple  Martins  are 
conspicuous  colonial  nesters  that  almost  ex- 
clusively are  dependent  on  man-made  nesting 
houses.  As  a result,  martin  breeding  biology 
and  behavior  have  been  the  focus  of  consid- 
erable study  (see  Brown  1997).  After  the 
fledging  period  eastern  populations  of  Purple 
Martins  often  congregate  in  distinctive  noc- 
turnal roosts  that  may  reach  concentrations  of 
100,000  or  more  as  a prelude  to  fall  migration 
(Brown  1997).  From  late  June  through  August 
or  early  September  these  assemblages  engage 
in  two  mass  movements  daily;  a morning  ex- 
odus from  the  roost  for  aerial  foraging  and  an 
evening  return  (Allen  and  Nice  1952).  Al- 
though the  existence  of  these  large  pre-migra- 
tory roosts  is  well  documented  (e.g..  Bent 
1942,  Allen  and  Nice  1952,  Anderson  1965, 
Brown  and  Wolfe  1978,  Rogillio  1989,  Rus- 
sell et  al.  1998),  little  quantitative  data  are 
available  concerning  their  spatial  and  tempo- 
ral dynamics. 

Studies  of  communal  roosts  may  be  limited 
by  the  inability  to  collect  data  at  appropriate 
spatial  and  temporal  scales  (Caccamise  et  al. 


' Dept,  of  Biological  Science.s.  Clemson  Univ., 
Clem.son.  SC  29634. 

2 Current  addre.ss:  Willamette  Industries,  Inc.,  P.O. 
Box  488.  Dallas.  OR  97338-0488. 

’ Corresponding  author; 

E-mail;  krussell@wii.com 


1983,  Russell  et  al.  1998).  Locating  roosts  of- 
ten is  labor-intensive  (see  Caccamise  and 
Fischl  1985,  Komar  1997).  Even  when  roost 
locations  are  known,  visual  surveys  alone  pro- 
vide only  limited  data  on  the  spatial  extent 
and  direction  of  roosting  flights  (Brown  and 
Wolfe  1978),  and  the  time  of  movements  (e.g., 
pre-dawn  flights)  may  limit  visual  observa- 
tions (Russell  and  Gauthreaux  1998).  Recent- 
ly, we  developed  methods  for  using  the  Na- 
tional Weather  Service’s  new  doppler  weather 
surveillance  radar,  the  WSR-88D  or  NEX- 
RAD to  locate  and  study  communal  roosting 
assemblages  (Russell  and  Gauthreaux  1998, 
Russell  et  al.  1998).  In  this  study,  we  provide 
quantitative  data  on  the  spatial  and  temporal 
dynamics  of  a pre-migratory  roost  of  Purple 
Martins  in  South  Carolina  using  simultaneous 
WSR-88D  radar  and  direct  visual  observa- 
tions. Our  specific  objectives  were  to  docu- 
ment the  daily  timing  and  spatial  pattern  of 
roosting  flights,  flight  directions  and  speeds, 
the  influence  of  weather  conditions  on  roost- 
ing flights,  and  seasonal  changes  in  roost  pop- 
ulation size. 

METHODS 

Study  area. — Our  study  site  was  a Purple  Martin 
roost  on  Lunch  fsland  (34°  03'  N.  81°  18'  W),  a 5-ha 
island  located  in  Lake  Murray,  near  Columbia,  South 
Carolina  (Russell  and  Gauthreaux  1998:  fig.  2).  We 
conducted  visual  surveys  from  a peninsula  on  the 
shoreline  of  the  lake,  approximately  3 km  south  of  the 
roost.  Radar  data  were  collected  at  the  National  Weath- 
er Service  Office,  Columbia  Metropolitan  Airport,  28 
km  southeast  of  the  roost  site.  A detailed  de.scription 


354 


Russell  and  Gauthreaux  • PURPLE  MARTIN  ROOSTING  DYNAMICS 


355 


of  the  study  area  is  provided  in  Russell  (1996)  and 
Russell  and  Gauthreaux  (1998). 

Sun’ey  methods. — We  conducted  timed  visual  sur- 
veys of  roosting  flights  5 days  per  week  from  30  June 
to  27  August  1995.  We  originally  conducted  morning 
surveys  from  05:50  to  06:50  EST,  but  extended  them 
to  07:30  when  the  duration  of  departures  lengthened 
in  late  July.  Evening  surveys  were  conducted  from  19: 
15  to  21:15  throughout  the  study.  During  each  visual 
survey,  we  recorded  the  numbers  and  flight  directions 
of  individual  martins  as  they  passed  over  a natural  cir- 
cular opening  (24  m diameter;  48°  angle  of  observa- 
tion) in  the  forest  canopy,  using  methods  described  by 
Lowery  and  Newman  ( 1963)  and  applied  to  WSR-88D 
radar  by  Russell  and  Gauthreaux  ( 1998).  We  identified 
martins  in  flight  by  their  distinctive  profile,  behavior, 
and  vocalizations  (Brown  1997).  The  presence  of  other 
roosting  species  was  negligible  during  the  study  (Rus- 
sell 1996).  When  both  numbers  and  flight  directions 
could  not  be  accurately  recorded,  we  ( I ) assigned  di- 
rections to  flocks,  or  (2)  only  made  counts  for  many 
of  the  birds.  During  surveys  we  also  made  incidental 
observations  of  the  roost  site  with  a 30X  spotting 
scope  and  recorded  general  weather  conditions.  We  lat- 
er obtained  detailed  local  climatological  data  (LCD) 
from  the  National  Weather  Service  Office  at  the  Co- 
lumbia Metropolitan  Airport. 

Immediately  after  each  visual  survey  we  visited  the 
National  Weather  Service  Office  and  acquired  WSR- 
88D  radar  images  that  coincided  with  our  surveys,  ex- 
cept for  eight  mornings  when  weather  conditions  ob- 
scured martin  flights  on  radar.  Evening  flights  of  Pur- 
ple Martins  typically  failed  to  produce  recognizable 
patterns  on  the  WSR-88D  (Russell  and  Gauthreaux 
1998);  thus  we  present  no  radar  data  on  evening  arriv- 
als. We  collected  base  reflectivity  images  to  monitor 
flights  of  martins  in  the  proximity  of  the  roost  and 
composite  reflectivity  images  to  track  flights  over  a 
wider  geographic  area  (Russell  and  Gauthreaux  1998). 
We  also  collected  several  radial  velocity  images  to  de- 
termine directions  and  speeds  of  each  mass  flight.  We 
recorded  radar  images  by  taking  a 35-mm  color  slide 
exposure  of  each  image  as  it  was  displayed  on  a mon- 
itor at  the  radar  site.  Detailed  descriptions  of  the  WSR- 
88D,  radar  images,  and  our  survey  methods  are  pro- 
vided in  Russell  and  Gauthreaux  (1998). 

Analyses. — We  examined  the  timing  of  roosting 
flights  by  recording  the  beginning,  end,  duration,  and 
peak  of  each  mass  flight  from  both  the  radar  and  visual 
data  (Russell  and  Gauthreaux  1998).  We  also  exam- 
ined temporal  changes  in  the  numbers  of  departing  and 
arriving  birds  by  pooling  our  visual  counts  into  15-min 
totals  ba.sed  on  our  timed  surveys  (Russell  1996).  We 
characterized  the  spatial  patterns  of  roosting  flights  by 
examining  the  origin,  timing,  intensity,  number,  and 
spatial  distribution  of  echoes  from  each  radar  image  of 
the  roost  area  (Russell  and  Gauthreaux  1998).  We  mea- 
sured the  maximum  spatial  extent  of  each  mass  flight 
by  relating  echoes  created  from  the  leading  edge  of  the 
flight  to  landmarks  depicted  on  the  radar  (e.g.,  county 
boundaries).  We  excluded  mornings  when  regional 


precipitation  or  spurious  ground  echoes  obscured  the 
maximum  extent  of  flights. 

We  pooled  visual  flight  tracks  of  individual  birds  so 
that  peak  movement  from  or  to  the  roost  was  not  di- 
vided artificially  (Russell  1996);  then  examined  tem- 
poral changes  in  flight  directions  using  circular  statis- 
tics (Zar  1984).  We  scored  directions  and  speeds  of 
each  mass  flight  directly  from  the  radar  images  (Ru.s- 
sell  and  Gauthreaux  1998). 

We  used  univariate  correlation  and  stepwise  multi- 
ple regression  to  examine  the  influence  of  weather  con- 
ditions on  roosting  flights.  Daily  climatological  data 
(mean,  04;00,  07;00,  19:00)  included  temperature,  dew 
point,  total  precipitation,  atmospheric  pressure,  surface 
wind  speed,  visibility,  and  cloud  cover.  Variables  ini- 
tially entered  or  left  the  stepwise  models  at  P = 0. 1 . 
We  examined  the  influence  of  winds  aloft  on  morning 
departures  by  scoring  the  modal  azimuth  of  annular 
displacement  (relative  to  the  roost  site;  Eastwood  et  al. 
1962)  and  regressing  these  values  against  the  corre- 
sponding direction  of  geostrophic  winds  aloft  (1000  m 
above  the  ground  or  900  mb;  Riehl  1972)  at  07:00. 

The  total  roost  population  (TRP)  represents  the  es- 
timated daily  sum  of  all  birds  at  the  roost  (Caccamise 
et  al.  1983).  Based  on  a preliminary  examination  of 
the  radar  data,  we  assumed  that  departures  and  arrivals 
of  Purple  Martins  were  omnidirectional  at  approxi- 
mately equal  densities.  Our  visual  sample  area  was 
0.47°  or  1/766  of  the  circumference  around  the  roost 
at  a distance  of  3 km.  Thus,  we  calculated  separate 
morning  and  evening  TRPs  by  multiplying  the  total 
count  for  each  movement  by  766  (Russell  1996)  and 
examined  changes  in  roost  size.  We  used  univariate 
correlation  to  investigate  relationships  between  our 
morning  and  evening  counts,  and  patterns  of  flight 
with  time  and  date.  We  used  JMP  3.1  (SAS  Institute 
1995)  on  an  IBM  computer  for  all  analyses  with  prob- 
ability values  P < 0.05  recognized  as  significant  and 
means  reported  ± 1 standard  deviation  (SD)  unless 
otherwise  noted.  We  standardized  all  survey  times  by 
converting  to  minutes  before  or  after  sunrise  or  sunset 
(as  given  by  the  National  Weather  Service  Office  in 
Columbia,  South  Carolina). 

RESULTS 

Timing  of  flights. — We  detected  the  initia- 
tion of  mass  departures  as  a single  radar  echo 
corresponding  to  the  location  of  the  roost. 
Time  of  initial  departure  from  the  roost  aver- 
aged 41.4  ± 4.0  (SD)  min  before  sunrise 
(range  31—48  min)  and  was  independent  of 
date  {r-  = 0.02,  P > 0.05,  n = 32;  Russell 
and  Gauthreaux  1998).  Time  of  peak  depar- 
ture also  was  independent  of  date  (r-'  = 0.06, 
P > 0.05,  n = 32).  Mean  duration  of  morning 
departures  (interval  from  first  radar  echo  to 
last  bird  observed  visually)  was  67.9  ± 12.2 
min  (range  49-101  min)  and  was  positively 


356 


THE  WILSON  BULLETIN  • Vol.  ///,  No.  3,  September  1999 
A B 


TIME  BEFORE  (-)/AFI  ER  (+)  SUNRISE  (MIN)  TIME  BEFORE  (+)/AFI'ER  (-)  SUNSET  (MIN) 

FIG  I . Changing  numbers  of  roosting  Purple  Martins  during  visual  counts  of  (A)  morning  departures  and 
(B)  evening  arrivals,  by  time  before  and  after  sunrise  or  sunset. 


related  to  the  number  of  birds  counted  exiting 
the  roost  (r^  = 0.60,  P < 0.001,  n = 32). 

Changes  in  the  numbers  of  Purple  Martins 
observed  departing  the  roost  were  related  to 
the  time  of  sunrise.  Because  some  birds  first 
departed  in  darkness,  changes  in  the  percent- 
age of  birds  we  observed  40—30  min  before 
sunrise  reflected  their  increased  visibility  in 
the  morning  sky  rather  than  actual  numbers 
aloft  (Fig.  lA).  Corresponding  radar  data  in- 
dicated that,  on  average,  the  movement  of 
most  birds  from  the  roost  began  20.4  ± 4.6 
min  before  sunrise,  when  sufficient  light  was 
available  for  accurate  visual  counts  (Russell 
and  Gauthreaux  1998).  The  number  of  birds 
departing  from  the  roost  consistently  peaked 
about  10  min  before  sunrise,  then  steadily  de- 
clined over  the  next  40  min  (Fig.  lA).  We 
occasionally  observed  a small,  secondary  in- 
crease in  the  number  of  birds  departing  from 
the  roost  as  late  as  30-40  min  after  sunrise 
(Fig.  lA). 

Purple  Martins  first  returned  to  the  roost  an 
average  of  59.8  ± 13.5  min  before  sunset 
(range  35-84  min),  and  the  last  birds  entered 
the  roost  2-24  min  after  sunset  (T  = 15.5  ± 
5.9  min).  We  found  a weak  negative  relation- 
ship between  the  time  that  the  last  birds  were 
observed  entering  the  roost  and  date  (r  = 
0.23,  P = 0.0033,  n = 36),  but  not  for  the 
time  of  peak  movement  into  the  roost  (r-’  = 
0.07,  P > 0.05,  n = 36).  Mean  duration  of 
observed  evening  arrivals  at  the  roost  (71.4  ± 


15.2  min,  range  43-103  min)  was  not  related 
to  our  visual  counts  (r-’  = 0.05,  P > 0.05,  n 
= 36),  but  we  detected  a negative  relationship 
between  the  duration  of  evening  flights  and 
date  = 0.71,  P < 0.001,  n = 36). 

Numbers  of  Purple  Martins  returning  to  the 
roost  increased  50-30  min  before  sunset  (Fig. 
IB).  At  about  30  min  before  sunset,  the  num- 
ber of  arrivals  increased  rapidly  until  peaking 
at  sunset  (Fig.  IB).  After  sunset,  numbers 
aloft  declined  rapidly  and  ceased  by  24  min 
after  sunset,  and  always  before  complete  dark- 
ness. 

Spatial  pattern  of  flights. — Each  morning 
until  the  peak  of  the  roosting  season.  Purple 
Martins  departed  en  masse  from  the  roost  in 
all  directions  (360°).  These  omnidirectional 
departures  first  were  visible  on  radar  as  a 
roughly  circular  mass  of  echoes  that  extended 
across  Lake  Murray  (Russell  and  Gauthreaux 
1998:  fig.  3).  As  departures  continued,  an  ex- 
panding ring  or  annulus  was  seen  on  radar  as 
the  departing  birds  extended  across  the  land- 
scape directly  away  from  the  roost  (Russell 
and  Gauthreaux  1998:  fig.  4).  The  maximum 
daily  extent  of  flights  we  could  detect  on  radar 
averaged  77.9  ± 14.1  km  (range  50-100  km, 
n = 20;  Russell  and  Gauthreaux  1998);  We 
found  no  relationship  between  the  maximum 
distance  of  flights  and  date  (r^  = 0.02,  P > 
0.05,  n = 20)  or  our  morning  counts  = 
0.01,  P > 0.05,  n = 20).  Beginning  12  August 
we  observed  on  radar  that  more  birds  departed 


Russell  (uul  Gauthreaux  • PURPLE  MARTIN  ROOSTING  DYNAMICS 


357 


south  than  other  directions;  by  18  August 
birds  only  departed  to  the  south  and  after  26 
August  we  no  longer  observed  departures  on 
radar. 

In  the  evening.  Purple  Martins  arrived  at 
the  roost  in  sporadic,  loosely  organized  flocks. 
The  birds  flew  over  the  observation  point  at 
tree-top  height  (ca  18-25  m)  and  then  flew 
down  to  just  above  the  lake’s  surface,  usually 
below  the  radar  beam  (Russell  and  Gauth- 
reaux  1998).  After  reaching  the  roost  the  birds 
remained  aloft  in  a growing  mass  that  circled 
high  over  the  island  counter-clockwise  until  a 
final,  spiraling  entry  into  the  roost  at  sunset. 

Flight  directions  and  speeds. — During 
morning  departures,  the  Purple  Martins  we 
observed  visually  were  consistently  and  sig- 
nificantly {P  < 0.001)  oriented  to  the  south. 
The  mean  direction  (±95%  confidence  inter- 
val, vector  length,  angular  deviation)  of  mar- 
tins at  40-21  min  before  sunrise  was  180°  (2°, 
0.953,  18°);  at  20-1  min  before  sunrise  was 
184°  (3°,  0.922,  23°);  at  0-20  min  after  sunrise 
was  183°  (4°,  0.884,  28°);  at  21-40  min  after 
sunrise  was  182°  (4°,  0.885,  28°);  and  at  41- 
60  min  after  sunrise  was  190°  (3°,  0.931,  21°). 
The  radar  images  also  showed  the  birds  flying 
in  a southerly  direction  over  the  observation 
point,  consistent  with  an  omnidirectional  de- 
parture from  the  roost.  Radar  images  of  ve- 
locity that  we  acquired  on  32  mornings 
showed  that  flight  speeds  of  the  departing 
birds  ranged  from  10.8-13.4  m/s  (see  figure  5 
in  Russell  and  Gauthreaux  1998). 

We  observed  distinct  changes  in  the  orien- 
tation of  Purple  Martins  during  evening  flights 
to  the  roost.  The  mean  direction  of  martins  at 
90-71  min  before  sunset  was  141°  (90°,  0.364, 
65°)  and  not  significant  (P  > 0.05).  During 
this  time  we  often  observed  birds  feeding  in 
small  flocks  and  flying  away  from  the  roost. 
As  the  time  of  sunset  approached,  however, 
the  mean  direction  of  martins  at  the  observa- 
tion point  became  significant  (P  < 0.001)  and 
increasingly  oriented  towards  the  northeast, 
but  still  east  of  the  roost  (azimuth  = 10°  from 
observation  point).  The  mean  direction  of 
martins  at  70-51  min  before  sunset  was  93° 
(18°,  0.318,  67°);  at  50-31  min  before  sunset 
was  73°  (14°,  0.399,  63°);  at  30-1 1 min  before 
sunset  was  57°  (9°,  0.590,  52°);  at  10  min  be- 
fore to  9 min  after  sunset  was  54°  (5°,  0.836, 


33°);  and  at  10-29  min  after  sunset  was  53° 
(4°,  0.843,  32°). 

Weather  conditions  and  roosting  flights. — 
Purple  Martins  first  departed  the  roost  earlier 
relative  to  sunrise  with  higher  atmospheric 
pressure,  surface  wind  speed,  and  relative  hu- 
midity, but  later  relative  to  sunrise  with  in- 
creased cloud  cover  (stepwise  multiple  regres- 
sion: = 0.66,  P < 0.001,  n = 32).  Of  these 

variables,  atmospheric  pressure  was  the  most 
important  and  the  only  significant  univariate 
relationship  {r-  = 0.20,  P = 0.0068,  n = 32). 
We  also  found  a significant  relationship  be- 
tween the  duration  of  morning  departures  and 
atmospheric  pressure  (stepwise  multiple  re- 
gression: R~  = 0.29,  P < 0.001,  n = 32). 

Purple  Martins  were  detected  on  radar  far- 
ther from  the  roost  on  days  with  increased 
cloud  cover  or  decreased  visibility  at  07:00 
(stepwise  multiple  regression:  R^  = 0.61,  P < 
0.001,  n = 20),  although  univariate  analysis 
indicated  a significant  relationship  only  for 
cloud  cover  (ri  = 0.35,  P = 0.0063,  n = 20). 
Mass  departures  from  the  roost  also  appeared 
to  be  displaced  by  winds  aloft.  Azimuthal 
wind  direction  was  strongly  related  to  the 
modal  direction  of  annulus  displacement  (r-  = 
0.76,  P < 0.001,  n = 20). 

Arrival  of  the  first  birds  over  the  observa- 
tion point  was  not  related  to  any  of  the  weath- 
er variables,  but  arrival  of  the  last  martins  at 
the  roost  occurred  earlier  relative  to  sunset  on 
days  with  more  cloud  cover  and  later  in  the 
season  (stepwise  multiple  regression:  R~  = 
0.40,  P < 0.001,  n = 36).  Univariate  analysis 
also  indicated  a significant  relationship  for 
cloud  cover:  (ri  = 0. 13,  P = 0.0275,  n = 36). 
Maximum  temperature,  atmospheric  pressure, 
and  surface  wind  speed  at  19:00  were  identi- 
fied by  stepwise  regression  as  significant  pre- 
dictors of  the  duration  of  evening  flights  {R~ 
= 0.89,  P < 0.001,  n = 36),  although  no  var- 
iable was  significant  when  subjected  to  uni- 
variate analysis. 

Seasonal  changes  in  total  roost  popula- 
tion.— Despite  fluctuations  in  the  daily  number 
of  departing  (x  = 368.3  ± 201.9  SD,  range  29- 
916,  n — 14,732  birds)  and  returning  (x  = 
755.1  ± 548.9  SD,  range  108-2,531,  n = 
27,182  birds)  martins,  the  roosting  population 
exhibited  a seasonal  pattern  of  growth  and  de- 
cline (Fig.  2).  On  the  first  morning  census  of 
3 July  we  counted  only  172  birds  over  the  ob- 


358 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


MORNING  EXODUS 


EVENING  RETURN 


FIG  2.  Seasonal  changes  in  visual  counts  of  roost- 
ing Purple  Martins  during  (A)  morning  departures  and 
(B)  evening  arrivals. 


servation  point  yielding  an  estimated  total  roost 
population  (TRP)  of  131,752  birds,  but  on  31 
July  we  counted  916  birds,  yielding  a peak 
morning  TRP  of  701,656  birds.  On  the  last 
morning  census  (27  August),  we  counted  only 
29  birds  (TRP  = 22,214).  Although  evening 
counts  usually  were  higher  than  morning 
counts,  numbers  of  birds  returning  to  the  roost 
also  followed  a pattern  of  increase  and  decline 
(Fig.  2).  The  first  (30  June),  peak  (17  July), 
and  last  (27  August)  evening  counts  yielded 
TRPs  of  301,804  birds,  1,938,746  birds,  and  0 
birds,  respectively.  However,  the  peak  morning 
count  occurred  two  weeks  after  the  peak  even- 
ing count  and  morning  counts  were  not  related 
to  evening  counts  from  the  previous  night  (r" 
= 0.001,  P > 0.05,  n = 32)  or  the  same  night 
(r  = 0.01,  P > 0.05,  n = 32). 

DISCUSSION 

Timing  of  flights. — Our  observations  indi- 
cate that  flights  of  Purple  Martins  from  and  to 
pre-migratory  roosts  are  related  to  levels  of 


ambient  light.  Purple  Martins  (Brown  and 
Wolfe  1978,  Oren  1980,  Hill  1988),  Crag 
Martins  (Ptyonoprogne  rupestris',  Elkins  and 
Etheridge  1974),  and  other  species  of  swal- 
lows (Rudebeck  1955,  Loske  1984,  Skutch 
1989,  Komar  1997)  have  been  shown  previ- 
ously to  enter  and  exit  roosts  in  response  to 
changing  intensities  of  light.  Although  endog- 
enous factors  likely  are  a significant  influence 
on  the  timing  of  departures  and  arrivals  of 
birds  at  roosts  (Aschoff  1967),  cueing  on 
changing  light  levels  has  been  suggested  as  a 
selective  advantage  for  maximizing  time 
available  for  feeding  (Eastwood  et  al.  1962, 
Summers  and  Eeare  1995). 

The  timing  of  evening  flights  was  less  pre- 
dictable than  morning  departures  and  ap- 
peared to  be  related  to  date.  Across  a season 
the  departure  or  arrival  times  of  the  first  and 
last  birds  at  roosts  may  show  greater  variation 
than  peaks  of  departure  or  arrival  (Summers 
and  Eeare  1995).  In  our  study,  more  variation 
was  associated  with  the  times  martins  first  ar- 
rived in  the  vicinity  of  the  roost  than  when 
flights  were  terminated,  as  was  reported  pre- 
viously for  other  roosting  species  (Eastwood 
et  al.  1962,  Meanley  1965,  Bunning  1973, 
Krantz  and  Gauthreaux  1975).  We  suggest 
that  relationships  between  date  and  the  timing 
of  evening  arrivals  of  martins  are  related 
mostly  to  the  large  variation  associated  with 
these  flights. 

Spatial  pattern  of  flights. — As  with  other 
roosting  species  (Harper  1959,  Eastwood  et  al. 
1962),  the  omnidirectional  departures  of  Pur- 
ple Martins  that  we  observed  on  radar  were 
associated  with  daily  feeding  flights  (Russell 
1996).  In  contrast,  directionally  biased  pat- 
terns of  flight  on  radar  are  often  associated 
with  migration  departures  (Harper  1959,  Rich- 
ardson and  Haight  1970).  The  strong  southerly 
bias  in  the  annulus  and  its  eventual  disap- 
pearance during  the  latter  part  of  August,  in 
conjunction  with  our  declining  visual  counts, 
marked  increasing  fall  migration  departures 
and  eventual  abandonment  of  the  roost. 

Among  our  more  surprising  results  was  the 
long  distances  Purple  Martins  flew  from  the 
roost.  Prior  to  our  study,  it  was  believed  that 
the  birds  foraged  within  10-15  km  of  the 
Lake  Murray  roost  (J.  Cely,  pers.  comm.). 
Brown  and  Wolfe  (1978)  suggested  that  Pur- 
ple Martins  may  travel  as  far  as  48  km  from 


Russell  and  Gauthreaux  • PURPLE  MARTIN  ROOSTING  DYNAMICS 


359 


pre-niigratory  roosts  during  the  day  while 
feeding.  Our  radar  data  clearly  showed  that 
birds  regularly  flew  almost  80  km  and  occa- 
sionally as  far  as  100  km  from  the  roost.  If 
the  large  population  size  of  the  roost  and  thus 
foraging  competition  (Caccamise  et  al.  1983, 
Sunmiers  and  Feare  1995)  was  responsible  for 
the  flight  distances  we  observed,  a significant 
relationship  should  have  emerged  between  the 
extent  of  departures  and  changes  in  TRP,  but 
it  did  not.  In  fact,  some  of  the  longest  flights 
occurred  early  in  the  season,  when  TRP  was 
relatively  small.  Because  we  did  not  track  in- 
dividuals, we  do  not  know  the  ultimate  des- 
tination of  the  birds  or  whether  all  departing 
individuals  returned  the  same  evening. 

Flight  directions  and  speeds. — In  marked 
contrast  to  the  uniform  departures  of  Purple 
Martins,  we  observed  high  variability  in  the 
flight  tracks  of  birds  arriving  at  the  roost. 
Even  during  the  peak  of  arrivals,  mean  flight 
direction  of  the  birds  was  east  of  the  roost. 
This  directional  bias  and  our  own  incidental 
observations  indicate  that  Purple  Martins  re- 
turn to  the  roost  via  specific  flight  corridors. 
Use  of  preferred  flight  lines  often  is  associated 
with  evening  movements  to  roosts  (Eastwood 
et  al.  1962,  Meanley  1965,  Skutch  1989).  Af- 
ter arriving  at  the  roost  the  birds  assembled  in 
a high,  circling  mass  before  a final  entry  at 
sunset.  This  pattern  is  very  similar  to  the 
phases  of  “staging”  or  pre-settling  behavior 
(initial  aimless  flight,  formation  of  a tight 
flock  near  the  roost,  final  descent;  Loske 
1984)  previously  described  for  Prague  spp. 
(Brown  and  Wolfe  1978,  Oren  1980,  Hill 
1988)  and  other  genera  of  roosting  swallows 
(Bent  1942;  Loske  1984,  1986;  Skutch  1989, 
Komar  1997).  Flight  speeds  of  departing  Pur- 
ple Martins  we  recorded  on  radar  (10.8-13.4 
m/s)  were  similar  to  those  reported  by  Evans 
and  Drickamer  (8.45-11.09  m/s;  1994)  and 
Southern  (12.1  m/s;  1959). 

Weather  conditions  and  roosting  flights. — 
Because  Purple  Martins  apparently  respond  to 
some  threshold  level  of  ambient  light  for  the 
cueing  of  flights  from  and  to  the  roost,  the 
timing  of  these  movements  should  vary  with 
daily  differences  in  light  levels  caused  by 
changing  cloud  cover  (Richardson  1978,  Elk- 
ins 1983).  In  our  study,  cloud  cover  and  at- 
mospheric pressure  were  the  most  important 
weather  conditions  explaining  variation  in  the 


timing  of  roosting  flights.  Increasing  or  de- 
creasing light  levels  reach  a given  intensity 
later  or  earlier,  respectively,  on  days  with 
overcast  conditions  compared  to  clear  days. 
Atmospheric  pressure  is  often  inversely  cor- 
related with  the  amount  of  cloud  cover  (Riehl 
1972).  The  inclusion  of  other  variables  into 
the  stepwise  models  also  may  have  resulted 
from  intercorrelations  with  cloud  cover,  at- 
mospheric pressure,  or  date  (Richardson 
1978).  Because  these  weather  variables  were 
highly  intercorrelated,  the  most  that  can  be  as- 
sumed is  that  the  birds  responded  to  some  as- 
pect of  the  weather  that  was  interrelated  with 
the  significant  variables  (Richardson  1978). 
Thus,  we  view  our  results  as  a preliminary 
explanation  of  variation  in  the  flight  behaviors 
of  roosting  Purple  Martins,  rather  than  repre- 
senting causal  mechanisms. 

Finlay  (1976)  observed  the  influence  of 
cloud  cover  on  Purple  Martins  during  the  nest- 
ing season;  the  birds  departed  from  nests  ear- 
lier on  clear  days  than  on  cloudy  days.  Also, 
Elkins  and  Etheridge  (1974)  found  that  Crag 
Martins  returned  to  roosts  up  to  2.5  h earlier 
on  overcast  or  cloudy  days.  The  influence  of 
cloud  cover  on  the  extent  of  martin  roosting 
flights  may  in  part  be  related  to  the  aerial  for- 
aging habits  of  the  birds  (Elkins  1983).  In  two 
studies  of  Purple  Martin  food  habits  (Spice 
1972,  Walsh  1978),  the  intake  of  airborne  in- 
sects was  negatively  correlated  with  cloud 
cover.  Cloudy  weather  also  was  shown  to  re- 
duce the  amount  and  type  of  airborne  insects 
taken  by  Brown-chested  Martins  {Phaeoprog- 
ne  tapera)  in  Venezuela  (Turner  1984).  Ad- 
ditionally, Finlay  (1976)  reported  that  during 
overcast  conditions  martins  appeared  to  spend 
more  time  away  from  nests,  presumably 
searching  for  food.  It  is  possible  that  cloudy 
conditions  forced  roosting  Purple  Martins  to 
travel  longer  distances  in  search  of  prey  (Elk- 
ins 1983). 

Our  radar  observations  that  winds  aloft  dis- 
placed the  annular  departures  of  Purple  Mar- 
tins indicate  that  the  birds  continued  to  fly 
with  constant  headings  and  speeds,  apparently 
making  no  correction  for  drift  in  spite  of  their 
presumed  knowledge  of  the  local  topography. 
Similar  radar  observations  were  made  by 
Eastwood  et  al.  (1962)  of  wind  displaced  de- 
partures of  European  Starlings  {Sturnus  vul- 
garis) from  roosts. 


360 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3.  September  1999 


Seasonal  changes  in  total  roost  popula- 
tion.— The  overall  increase  in  TRP  through 
the  end  of  July  likely  reflects  recruitment  of 
local  and  regional  populations  after  the  fledg- 
ing period  (Brown  1997).  Although  Purple 
Martins  are  colonial  breeders  they  nest  asyn- 
chronously (Brown  1997),  and  in  South  Car- 
olina clutches  have  been  observed  as  early  as 
1 1 April  and  as  late  as  19  June  (Post  and 
Gauthreaux  1989).  Thus,  it  is  not  surprising 
that  increases  in  TRP  during  July  occurred 
gradually.  The  influx  of  birds  at  the  roost  also 
occurred  too  early  in  the  season  to  be  signif- 
icantly influenced  by  migration.  However, 
subsequent  declines  in  TRP  and  changes  in 
flight  patterns  evident  on  radar  throughout 
August  corresponded  with  fall  migration  de- 
partures (Hamel  1992,  Brown  1997).  In  the 
southeastern  United  States,  Purple  Martins  of- 
ten depart  for  South  America  in  early  August 
(Hamel  1992). 

Although  the  roost  exhibited  a seasonal  pat- 
tern of  growth  and  decline  consistent  with  pre- 
migrator y assembly,  large  daily  fluctuations  in 
TRP  were  evident.  The  accuracy  of  our  TRP 
estimates  depended  on  the  following  assump- 
tions: (1)  departures  and  arrivals  were  equally 
distributed  around  the  roost,  (2)  each  bird  was 
counted  only  once  during  a survey,  and  (3) 
birds  did  not  switch  among  roosts.  The  an- 
nular patterns  from  radar  and  uniform  flight 
tracks  over  the  observation  point  indicate  that 
the  first  two  assumptions  probably  were  met 
for  most  morning  departures,  although  some 
daily  variation  was  likely.  In  contrast,  the 
large  variability  and  directional  bias  of  even- 
ing flights  and  the  lack  of  relationships  be- 
tween morning  and  evening  counts  indicate 
that  the  birds  did  not  return  to  the  roost  in  a 
uniform  manner,  and  at  least  early  in  the  even- 
ing some  birds  were  counted  multiple  times. 
Thus,  our  evening  counts  likely  did  not  pro- 
vide an  accurate  estimate  of  the  number  of 
martins  at  the  roost  on  a daily  basis. 

Another  potential  source  of  variation  in 
TRP  is  movement  of  birds  among  roosts.  Dur- 
ing the  present  and  subsequent  (Russell  et  al. 
1998)  studies,  we  discovered  another  pre-mi- 
gratory  roost  of  martins  100  km  west  in  Geor- 
gia and  others  within  230  km.  Fluctuations  in 
TRP  through  late  July  may  reflect  some 
switching  of  birds  among  roosts  on  a regional 
basis,  while  temporary  increases  in  TRP  late 


in  the  season  also  could  result  from  birds  en- 
countering the  roost  during  migration  from 
more  northerly  breeding  or  roosting  sites. 
Likewise,  birds  migrating  from  Lake  Murray 
probably  encounter  roosts  along  the  Gulf 
coast  (Russell  et  al.  1998).  Although  we  be- 
lieve our  morning  counts  provide  a good  ini- 
tial assessment  of  daily  changes  in  TRP  and 
maximum  roost  size,  more  accurate  estimates 
will  require  monitoring  roost  populations  on  a 
regional  basis.  In  part  this  may  require  rigor- 
ous, labor  intensive  ground  surveys  at  multi- 
ple roost  sites  (e.g.,  Caccamise  et  al.  1983). 
However,  with  further  refinements  in  meth- 
odology, WSR-88D  radar  may  provide  the  po- 
tential to  remotely  monitor  seasonal  and  an- 
nual changes  in  roost  populations  over  large 
geographical  areas  (Russell  and  Gauthreaux 
1998,  Russell  et  al.  1998). 

Why  do  Purple  Martins  assemble  in  pre- 
migratory  roosts? — Large  pre-migratory 
roosts  of  Purple  Martins  are  neither  a recent 
nor  an  isolated  phenomenon.  Wayne  (1910), 
Stone  (1937),  and  Anderson  (1965)  observed 
roosts  reaching  concentrations  of  at  least 
100,000,  and  a well-established  roost  at  Lake 
Pontchartrain,  Louisiana  may  support  200,000 
birds  (Rogillio  1989).  Some  roosts,  including 
Lake  Murray  and  a site  in  southern  Oklahoma 
(Brown  and  Wolfe  1978)  have  been  used  by 
martins  for  20  years  or  more  (Russell  et  al. 
1998).  At  least  30  additional  major  pre-migra- 
tory roosts  are  known  to  exist  in  the  eastern 
United  States  (Russell  et  al.  1998).  Enormous 
concentrations  of  roosting  Purple  Martins  also 
have  been  documented  on  their  wintering 
grounds  (Oren  1980;  Hill  1988,  1993). 

Several  selective  advantages  have  been  pro- 
posed for  communal  roosting  behavior:  pre- 
migratory  assembly  (Allen  and  Nice  1952, 
Michael  and  Chao  1973,  Skutch  1989),  re- 
duced risk  of  predation  (Lack  1968),  more  ef- 
ficient thermoregulation  (Williams  et  al. 
1991),  enhanced  foraging  ability  through  in- 
formation exchange  (Ward  and  Zahavi  1973, 
Brown  and  Brown  1996),  and  association  with 
super-abundant  food  supplies  or  other  diurnal 
activity  centers  (Caccamise  and  Morfison 
1986,  Caccamise  1993).  Although  the  late- 
summer  roosting  habits  of  Purple  Martins  of- 
ten are  attributed  to  the  need  for  pre-migratory 
assembly  (Allen  and  Nice  1952),  this  seasonal 
pattern  of  roosting  does  not  explain  the  ben- 


Russell  and  Gaullireau.x  • PURPLE  MARTIN  ROOSTING  DYNAMICS 


361 


efits  gained  by  gathering  in  a large  communal 
assemblage  for  several  weeks  prior  to  migra- 
tion (Caccamise  et  al.  1983,  Brown  1997).  Be- 
cause the  Lake  Murray  roost  is  on  an  island 
and  at  least  3 km  from  the  lakeshore  the  birds 
likely  have  little  threat  from  most  predators  or 
other  disturbances;  other  large  pre-migratory 
roosts  of  Purple  Martins  also  are  associated 
with  bodies  of  water  (Russell  et  al.  1998).  Ad- 
ditionally, the  anti-predator  benefits  of  roost- 
ing should  reach  their  maximum  value  at  rel- 
atively small  population  sizes  (e.g.,  < 1000; 
Pulliam  1973).  Although  roost  advertising  is 
evident  in  the  pre-settling  flight  behavior  of 
Purple  Martins,  the  omnidirectional  departures 
of  the  birds  and  their  absence  until  evening 
make  it  unlikely  that  Lunch  Island  functions 
as  an  information  center  about  local  changes 
in  food  supply  (Skutch  1989,  Brown  and 
Brown  1996,  Brown  1997). 

Our  study  has  quantified  aspects  of  the  pre- 
migratory  roosting  behavior  olF  Purple  Martins 
that  previously  were  known  only  from  anec- 
dotal accounts  (see  Brown  1997)  or  brief  com- 
ments made  during  studies  of  nesting  or  im- 
mediate post-fledging  activities  (Allen  and 
Nice  1952,  Finlay  1971,  Brown  1978).  Future 
studies  should  focus  on  the  costs  and  benefits 
of  large  pre-migratory  roosts  to  Purple  Mar- 
tins and  be  conducted  at  scales  sufficiently 
large  to  monitor  spatial  and  temporal  patterns 
of  roosting  on  a regional  basis  (Caccamise  et 
al.  1983,  Russell  et  al.  1998).  Studies  employ- 
ing radiotelemetry  or  other  tracking  methods 
also  are  needed  to  determine  the  daily  and 
seasonal  movements  of  individual  martins,  in- 
cluding whether  individuals  switch  among 
roosts,  or  commute  between  roosts  and  diurnal 
activity  centers  (e.g.,  Caccamise  and  Morrison 
1986,  Caccamise  1993). 

ACKNOWLEDGMENTS 

Funding  was  provided  by  the  South  Carolina  Elec- 
tric & Gas  Company  (SCE&G)  and  the  Riverbanks 
Zoological  Park  and  Botanical  Garden  (Columbia, 
South  Carolina).  The  Department  of  Defense  (DoD), 
Legacy  Resource  Management  Program  provided  ad- 
ditional funding.  We  thank  B.  Palmer  and  the  staff  of 
the  National  Weather  Service  Office  (Columbia,  South 
Carolina)  for  use  of  the  WSR-88D  and  assistance  with 
its  operation.  Logistical  support  from  SCE&G  and  P. 
Krantz  of  Riverbanks  was  greatly  appreciated.  We  also 
gratefully  acknowledge  the  helpful  contributions  of  M. 
A.  Russell  and  C.  G.  Belser.  J.  A.  Waldvogel,  J.  D. 


Lanham,  C.  R.  Brown,  and  anonymous  reviewers  pro- 
vided critical  ct)mments  that  significantly  improved  the 

manuscript. 

LITERATURE  CITED 

Allen,  R.  W.  and  M.  M.  Nice.  1952.  A study  of  the 
breeding  biology  of  the  Purple  Martin  (Progne 
suhis).  Am.  Midi.  Nat.  47:606-665. 

American  Ornithologists’  Union.  1998.  Checklist  of 
North  American  birds,  seventh  ed.  American  Or- 
nithologists’ Union,  Washington,  D.C. 

Anderson,  D.  1965.  A preliminary  study  of  a Purple 
Martin  roost.  Bluebird  32:3-4. 

Aschoff,  J.  1967.  Circadian  rhythms  in  birds.  Proc. 
Int.  Ornithol.  Congr.  14:81-105. 

Bent,  A.  C.  1942.  Life  histories  of  North  American 
flycatchers,  larks,  swallows,  and  their  allies.  U.S. 
Natl.  Mus.  Bull.  179:1-555. 

Brown,  C.  R.  1978.  Post-fledging  behavior  of  Purple 
Martins.  Wilson  Bull.  90:376-385. 

Brown,  C.  R.  1997.  Puiple  Martin  (Progne  suhis).  In 
The  birds  of  North  America,  no.  287  (A.  Poole 
and  E Gill,  Eds.).  The  Academy  of  Natural  Sci- 
enees,  Philadelphia,  Pennsylvania;  The  American 
Ornithologists’  Union,  Washington,  D.C. 

Brown,  C.  R.  and  M.  B.  Brown.  1996.  Coloniality  in 
the  Cliff  Swallow:  the  effect  of  group  size  on  so- 
cial behavior.  Univ.  of  Chicago  Press,  Chicago, 
Illinois. 

Brown,  C.  R.  and  S.  D.  Wolfe.  1978.  Post-breeding 
movements  of  Purple  Martins  in  the  Lake  Texoma 
area.  Bull.  Tex.  Ornithol.  Soc.  11:22-23. 

Bunning,  E.  1973.  The  physiological  clock.  Springer- 
Verlag,  New  York. 

Caccamise,  D.  F.  1993.  The  “patch-sitting”  hypothe- 
sis: a parsimonious  view  of  communal  roosting 
behavior.  Wilson  Bull.  105:372-378. 

Caccamise,  D.  F and  J.  Fischl.  1985.  Patterns  of  as- 
sociation of  secondary  species  in  roosts  of  Euro- 
pean Starlings  and  Common  Graekles.  Wilson 
Bull.  97:173-182. 

Caccamise,  D.  F and  D.  W.  Morrison.  1986.  Avian 
communal  roosting:  implications  of  diurnal  activ- 
ity centers.  Am.  Nat.  128:191-198. 

Caccamise,  D.  F,  L.  A.  Lyon,  and  J.  Fischl.  1983. 
Seasonal  patterns  in  roosting  flocks  of  starlings 
and  Common  Crackles.  Condor  85:474-481. 

EA.STWOOD,  E.,  G.  A.  ISTED,  AND  G.  C.  Rider.  1962. 
Radar  ring  angels  and  the  roosting  movements  of 
starlings.  Philos.  Trans.  R.  Soc.  Lond.  B 156:242- 
267. 

Elkins,  N.  1983.  Weather  and  bird  behaviour.  T & A 
D Poyser,  Staffordshire,  U.K. 

Elkins,  N.  and  B.  Etheridge.  1974.  The  Crag  Martin 
in  winter  quarters  at  Gibraltar.  Brit.  Birds  72:376- 
387. 

Evans,  T.  R.  and  L.  C.  Drickamer.  1994.  Flight 
speeds  of  birds  determined  using  Doppler  radar. 
Wilson  Bull.  106:154-156. 

Finlay,  J.  C.  1971.  Post-breeding  nest  cavity  defense 
in  Pui-ple  Martins.  Condor  73:381-382. 


362 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3.  September  1999 


Finlay.  J.  C.  1976.  Some  effects  of  weather  on  Puiple 
Martin  activity.  Auk  93:231-244. 

Hamel,  P.  B.  1992.  Land  manager's  guide  to  the  birds 
of  the  South.  The  Nature  Conservancy,  Chapel 
Hill,  North  Carolina. 

Harper,  W.  G.  1959.  Roosting  movements  of  birds  and 
migration  departures  as  seen  by  radar.  Ibis  101: 
201-208. 

Hill,  J.  R.,  III.  1988.  A tale  of  two  cities:  the  Purple 
Martin  on  its  Brazilian  wintering  grounds.  Puiple 
Martin  Update  1(4):  1-5. 

Hill,  J.  R..  III.  1993.  A giant  martin  roost  on  the  Am- 
azon River.  Purple  Martin  Update  4(4):5— 7. 

Komar,  O.  1997.  Communal  roosting  behavior  of  the 
Cave  Swallow  in  El  Salvador.  Wilson  Bull.  109: 
332-337. 

Krantz,  P.  E.  and  S.  a.  Gauthreaux,  Jr.  1975.  Solar 
radiation,  light  intensity,  and  roosting  behavior  in 
birds.  Wilson  Bull.  87:91-95. 

Lack,  D.  1968.  Ecological  adaptations  for  breeding  in 
birds.  Methuen,  London,  U.K. 

Loske,  K.  H.  1984.  Observations  on  the  swallow  Hi- 
riiiulo  riistica  at  mass  roosting  places  in  the  mid- 
dle of  North  Rhine,  Westphalia,  Germany.  Vogel- 
welt  105:51-60. 

Loske,  K.  H.  1986.  On  the  behavior  of  swallows  Hi- 
rundo  riistica  at  mass  roosting  places  in  Namibia, 
Southwest  Africa.  Beitr.  Vogelkd.  32:273-280. 

Lowery,  G.  H.,  Jr.  and  R.  J.  Newman.  1963.  Studying 
bird  migration  with  a telescope.  Museum  of  Zo- 
ology, Louisiana  State  Univ.,  Baton  Rouge. 

Meanley,  B.  1965.  The  roosting  behavior  of  the  Red- 
winged Blackbird  in  the  southeastern  United 
States.  Wilson  Bull.  77:217-228. 

Michael,  E.  D.  and  W.  H.  Chao.  1973.  Migration  and 
roosting  of  Chimney  Swifts  in  east  Texas.  Auk  90: 
100-105. 

Oren,  D.  C.  1980.  Enormous  concentration  of  martins 
(Prague  spp.)  in  Iquitos,  Peru.  Condor  82:344- 
345. 

Post,  W.  and  S.  A.  Gauthreaux,  Jr.  1989.  Status  and 
distribution  ot  South  Carolina  birds.  Charleston 
Museum,  Charleston,  South  Carolina. 

Pulliam,  H.  R.  1973.  On  the  advantages  of  flocking. 
J.  Theor.  Biol.  38:419-422. 

Richardson,  W.  J.  1978.  Timing  and  amount  ot  bird 
migration  in  relation  to  weather:  a review.  Oikos 
30:224-272. 

Richardson,  W.  J.  and  M.  E.  Haight.  1970.  Migration 
departures  from  starling  roosts.  Can.  J.  Zool.  48: 
31-39. 


Riehl,  H.  1972.  Introduction  to  the  atmosphere,  second 
ed.  McGraw-Hill,  New  York. 

Rogillio,  C.  1989.  Plans  for  a Purple  Martin  sanctuary 
take  wing.  Purple  Martin  Update  2(l):6-7. 

Rudebeck,  G.  1955.  Some  observations  at  a roost  of 
European  swallows  and  other  birds  in  south-east- 
ern Transvaal.  Ibis  97:572—580. 

Russell,  K.  R.  1996.  Spatial  and  temporal  patterns  of 
a Purple  Martin  {Prague  siihis)  roost:  a radar  and 
direct  visual  study.  M.S.  thesis,  Clemson  Univ., 
Clemson,  South  Carolina. 

Russell,  K.  R.  and  S.  A.  Gauthreaux,  Jr.  1998.  Use 
of  weather  radar  to  characterize  movements  of 
roosting  Purple  Martins.  Wildl.  Soc.  Bull.  26:5— 
16. 

Russell,  K.  R.,  D.  S.  Mizrahi,  and  S.  A.  Gauth- 
reaux, Jr.  1998.  Large-scale  mapping  of  Purple 
Martin  pre-migratory  roosts  using  WSR-88D 
weather  surveillance  radar.  J.  Field  Ornithol.  69: 
316-325. 

SAS  Institute.  1995.  JMP  statistics  and  graphics 
guide.  Vers.  3.1.  SAS  Inst.  Inc.,  Cary,  North  Car- 
olina. 

Skutch,  a.  F.  1989.  Birds  asleep.  Univ.  of  Texas  Press, 
Austin. 

Southern,  W.  E.  1959.  Homing  of  Puiple  Martins. 
Wilson  Bull.  71:254—261. 

Spice,  H.  1972.  Food  habits  of  nestling  Purple  Martins. 
M.S.  thesis,  Univ.  of  Alberta,  Edmonton. 

Stone,  W.  1937.  Bird  studies  at  Old  Cape  May,  vol. 
2.  Delaware  Valley  Ornithol.  Club,  Academy  of 
Natural  Sciences  of  Philadelphia,  Philadelphia, 
Pennsylvania. 

Summers,  R.  W.  and  C.  J.  Feare.  1995.  Roost  depar- 
ture by  European  Starlings  Stiinnis  vulgaris:  ef- 
fects of  competition  and  choice  of  feeding  site.  J. 
Avian  Biol.  26:193-202. 

Turner,  A.  K.  1984.  Nesting  and  feeding  habits  of 
Brown-chested  Martins  in  relation  to  weather  con- 
ditions. Condor  86:30-35. 

Walsh,  H.  1978.  Food  of  nestling  Puiple  Martins.  Wil- 
son Bull.  90:248-260. 

Ward,  P.  and  A.  Zahavi.  1973.  The  importance  of 
certain  assemblages  of  birds  as  “information  cen- 
tres” for  food-finding.  Ibis  115:517-534. 

Wayne,  A.  T.  1910.  Birds  of  South  Carolina.  Charles- 
ton Museum,  Charleston,  South  Carolina. 

Williams,  J.  B.,  M.  A.  du  Plessis,  and  W.  R.  Sieg- 
fried. 1991.  Green  Woodhoopoes  (Phaenicidits 
piirpureus)  and  obligate  cavity  roosting  provide  a 
test  of  the  thermoregulatory  insufficiency  hypoth- 
esis. Auk  108:285-293. 

Zar,  j.  H.  1984.  Biostatistical  analysis,  second  ed. 
Prentice-Hall,  Englewood  Cliffs,  New  Jersey. 


Wilson  Bull.,  111(3),  1999,  pp.  363-367 


AGGRESSIVE  RESPONSE  OE  CHICKADEES  TOWARDS 
BLACK-CAPPED  AND  CAROLINA  CHICKADEE  CALLS  IN 

CENTRAL  ILLINOIS 

ERIC  L.  KERSHNER'  23  AND  ERIC  K.  BOLLINGER* 


ABSTRACT. — Aggressive  responses  of  Black-capped  (Poecile  atricapillus)  and  Carolina  chickadees  (Poecile 
carolinensis)  to  heterospecitic  and  conspecific  vocalization  playbacks  were  measured  across  a historic  contact 
zone  in  east-central  Illinois  to  determine  the  magnitude  of  interspecific  aggression.  Within  the  traditional  Carolina 
Chickadee  range,  chickadees  responded  more  aggressively  towards  Carolina  Chickadee  calls  than  Black-capped 
Chickadee  calls.  Within  the  traditional  Black-capped  Chickadee  range,  chickadees  did  not  respond  to  either 
vocalization  significantly  more  than  the  other.  The  aggressive  response  towards  presumed  heterospecific  vocal- 
izations for  all  chickadees  was  marginally  more  aggressive  closer  to  the  contact  zone.  Thus,  we  conclude  that 
interspecific  aggression  may  not  act  as  a gap  producing  mechanism  between  chickadee  ranges.  Received  24  Nov. 
1998,  accepted  3!  March  1999. 


Black-capped  {Poecile  atricapillus)  and 
Carolina  {Poecile  carolinensis)  chickadees  are 
extremely  similar  in  appearance,  behavior,  and 
ecology  (Brewer  1963,  Johnston  1971,  Merritt 
1978).  They  occupy  largely  parapatric  breed- 
ing ranges,  although  there  are  a few  narrow 
zones  of  overlap.  Interspecific  encounters  and 
recognition  of  heterospecifics  are  common  in 
these  overlap  areas  (Ward  and  Ward  1974). 
Interspecific  territoriality  in  chickadees  is  rare 
except  with  other  chickadees  in  contact  zones 
(Brewer  1963,  Smith  1993).  Interspecific  ter- 
ritoriality may  arise  through  the  competition 
for  limited  resources,  which  may  contirbute  to 
a competitively-induced  gap  between  Black- 
capped  and  Carolina  chickadee  ranges  (Tanner 
1952,  Slade  and  Robertson  1977). 

We  measured  the  level  of  aggression  exhib- 
ited by  chickadees  across  a historic  contact 
zone  in  Illinois  defined  by  Brewer  (1963)  to 
test  the  role  of  interspecific  aggression  as  a 
range  segregating  mechanism.  We  hypothe- 
sized that  responses  towards  the  conspecific 
vocalization  would  be  more  aggressive  than 
those  to  heterospecific  vocalizations.  We  also 
hypothesized  that  levels  of  aggression  to  het- 
erospecific vocalizations  would  be  greatest 
closer  to  the  contact  zone  and  weaker  away 
from  the  contact  zone.  We  expected  this  re- 


' Dept,  of  Zoology,  Eastern  Illinois  Univ.,  Charles- 
ton, IL  61920. 

^ Present  address:  Dept,  of  Natural  Resources  and 
Environmental  Sciences,  350  Burnsides  Research  Lab- 
oratory, 1208  W.  Pennsylvania  Ave.,  Univ.  of  Illinois, 
Urbana,  IL  61801. 

^ Corresponding  author;  E-mail:  kershner@uiuc.edu 


sponse  because  closer  to  the  zone  of  overlap 
chickadees  would  have  more  encounters  with 
congeners  and  therefore  should  exhibit  more 
aggressive  territorial  defense  if  interspecific 
aggression  is  used  to  maintain  range  bound- 
aries. 

METHODS 

Study  area. — Our  5 study  sites  formed  a transect 
across  the  historical  contact  zone  for  chickadees  in 
east-central  Illinois  (Fig.  1).  The  distance  from  the 
midline  of  the  contact  zone  varied  for  each  study  site 
(Shelby ville  10  km,  Douglas  Hart  22  km.  Fox  Ridge 
40  km,  Lincoln  Trail  58  km,  Sangchris  65  km). 
Sangchris  State  Park  and  Shelbyville  State  Park  were 
located  in  traditional  Black-capped  Chickadee  range. 
Fox  Ridge  State  Park,  Douglas  Hart  Nature  Center,  and 
Lincoln  Trail  State  Park  were  located  in  traditional 
Carolina  Chickadee  range.  Study  sites  were  visited 
weekly. 

Playback  e.xperiments. — Two  chickadee  calls  and 
one  White-breasted  Nuthatch  (Sitta  carolinensis)  vo- 
calization were  used  in  this  experiment.  A playback 
tape  (Maxell  UDII  60  minutes)  was  made  for  each 
vocalization  on  a Magnavox  FA9403  dual  recording 
stereo  system.  Each  tape  had  a call  rate  of  18  calls/ 
minute.  Vocalizations  of  both  chickadee  species  and 
the  White-breasted  Nuthatches  were  taken  from  the 
Peterson  Guide  to  Bird  Songs®  (Peterson  1983).  One 
call  of  each  species  was  used  for  all  trials.  This  seemed 
reasonable  given  that  differences  between  species  calls 
are  much  greater  than  variation  of  calls  within  a spe- 
cies. Vocalizations  were  played  to  subjects  on  a Pan- 
asonic FW18  dual  speaker  cassette  recorder. 

Playbacks  were  used  to  test  the  abilities  of  chicka- 
dees to  discriminate  between  conspecific  and  hetero- 
specific vocalizations  (Emlen  et  al.  1975).  Several  re- 
searchers have  shown  that  chickadees  respond  to  con- 
specific songs  more  than  to  heterospecific  songs  (Hill 
and  Lein  1989,  Menitt  1978,  Robbins  et  al.  1986b, 


363 


364 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  3,  September  1999 


EIG.  I.  Map  of  the  study  sites.  Dark  line  indicates 
the  historic  contact  zone  across  east-central  Illinois 
(from  Brewer,  1963).  North  of  the  contact  zone  rep- 
resents Black-capped  Chickadee  range  and  south  of  the 
contact  zone  represents  Carolina  Chickadee  range. 
Study  sites  are  (1)  Sangchris  State  Park,  (2)  Shelby- 
ville  State  Park,  (3)  Douglas  Hart  Nature  Center,  (4) 
Pox  Ridge  State  Park,  and  (5)  Lincoln  Trail  State  Park. 

Ward  and  Ward  1974).  The  chick-a-dee  call  was  used 
in  this  study  because  it  contains  sufficient  information 
that  can  potentially  be  used  by  chickadees  for  individ- 
ual recognition  (Mamman  and  Nowicki  1981,  Smith 
1991).  These  calls  tend  to  be  short,  less  musical,  less 
variable,  and  relatively  specialized  for  particular  lunc- 
tions  such  as  alarm  or  changing  the  spacing  between 
individuals  (Smith  1991).  The  playback  of  chick-a-dee 
calls  may  elicit  elevated  calling  rates  (Nowicki  1983) 
by  simulating  intruding  males  (Hill  and  Lein  1989, 
Shackleton  et  al.  1992),  and  therefore  would  measure 
any  differences  in  aggression  levels  between  vocali- 
zation type. 

Playback  experiments  were  conducted  08:00-1 2:00 
CST,  1 May-31  July  1995.  Two  different  paired  play- 
back trials  were  conducted.  One  trial  type  consisted  of 
broadcasting  Black-capped  and  Carolina  chick-a-dee 
call  to  chickadees  in  all  study  sites.  The  .second  trial 


type  consisted  of  playing  both  a randomly-selected 
chickadee  call  (Black-capped  or  Carolina)  and  a 
White-breasted  Nuthatch  vocalization.  This  second  tri- 
al type  was  played  every  third  trial  at  each  study  site. 
Nuthatches  were  used  because  chickadees  should  be 
familiar  with  this  vocalization  through  winter  flock  as- 
sociation. Therefore,  the  responses  to  the  nuthatch  vo- 
calizations gave  us  a baseline  level  of  aggression  to 
compare  to  the  responses  to  chickadee  vocalizations. 

Playback  trials  were  conducted  by  slowly  walking 
around  the  study  site  until  chickadees  were  detected. 
Trials  were  conducted  by  approaching  a single  bird  as 
close  as  possible  without  visibly  agitating  it.  Only  data 
that  fit  the  following  criteria  were  used:  (1)  the  focal 
bird  could  be  approached  within  15  m,  and  (2)  weather 
conditions  matched  those  required  for  the  Breeding 
Bird  Survey  (Robbins  et  al.  1986a).  Each  bird  was 
exposed  to  only  one  trial  (either  a two-chickadee  trial 
or  a chickadee-nuthatch  trial),  and  specific  areas  within 
each  study  site  were  used  for  only  one  trial  to  avoid 
influencing  neighboring  chickadees.  The  minimum 
distance  between  trial  locations  was  500  m and  was 
usually  over  1000  m.  Vocalizations  were  played  for  2 
minutes  with  a 5 minute  silent  period  between  the  two 
sets  of  vocalizations.  The  silent  period  allowed  the  fo- 
cal bird  to  return  to  normal  activity  after  being  exposed 
to  the  first  vocalization.  The  order  of  the  two  vocali- 
zations were  alternated  to  reduce  bias  (Kroodsma 
1989,  Lampe  and  Baker  1994,  Ward  and  Ward  1974). 

Statistical  analyses. — The  degree  of  aggression  by 
the  focal  bird  was  quantified  based  on  its  behavior  dur- 
ing the  two-minute  trial  period  (Table  1).  The  identity 
of  species  at  each  study  site  was  assumed  to  be  that 
of  historic  record,  although  the  possibility  of  hybrid- 
ization may  render  this  assumption  invalid.  However, 
all  analyses  were  conducted  without  regard  to  species 
identity.  Response  scores  were  analyzed  by  paired  t- 
tests.  To  compare  all  chickadee  calls  combined  to  nut- 
hatch vocalizations,  one  trial  per  site  was  used  and  all 
sites  were  pooled  together.  The  relationships  between 
mean  aggressive  response  and  both  date  and  distance 
to  the  contact  zone  were  analyzed  by  Pearson  Cone- 
lation  Analysis  (SAS  Institute  1994). 

RESULTS 

Chickadees  responded  more  aggressively  to 
Black-capped  and  Carolina  chickadee  calls 
combined  than  to  songs  of  White-breasted 
Nuthatches  (t  = 4.6,  df  = 9,  P < 0.001;  Fig. 
2).  Chickadees  also  responded  more  aggres- 
sively to  each  chickadee  species  call  separate- 
ly than  to  White-breasted  Nuthatch  vocaliza- 
tions (Black-capped:  t = 8.9,  df  = 4,  P < 
0.001;  Carolina:  t = 5.65,  df  = 4,  P < 0.005; 
Fig.  2). 

There  was  no  significant  difference  in  ag- 
gressive response  to  Carolina  calls  versus 
Black-capped  chickadee  calls  (t  = 1.06,  df  = 
33,  P > 0.05;  Fig.  3).  Chickadees  within  tra- 


Kershner  ami  Bollinf>er  • PLAYBACK  RESPONSES  OF  CHICKADEES 


365 


TABLE  1 . Categories  of  behavioral  responses  for 
Black-capped  Chickadees  (Poecile  ciiricapillus)  and 
Carolina  Chickadees  (Poecile  carolinensis)  to  play- 
backs of  conspecihc  and  heterospecific  calls.  Catego- 
ries were  derived  by  combining  information  from 
Brindley  (1991),  Censky  and  Ficken  (1982),  Ficken 
and  Wiese  (1984),  Popp  and  coworkers  (1990), 
Schroeder  and  Wiley  (1983),  and  Shackleton  and  co- 
workers (1992).  The  categories  run  on  a 0—10  scale 
with  10  being  the  most  aggressive. 


Category 

Behavior 

10 

Flights  <2  m from  tape  player,  wing  twit- 
tering 

9 

Flights  2-5  m from  tape  player 

8 

Flights  <5  m from  tape  player,  >2  gargles 

7 

Flights  <5  m from  tape  player,  <2  gargles 

6 

Flights  5-10  m from  tape  player,  >4  calls 
made 

5 

Flights  5-10  m from  tape  player,  <4  calls 
made 

4 

Flights  5—10  m from  tape  player,  songs 
elicited 

3 

Flights  >10  m from  tape  player,  calls  elic- 
ited, some  gargling 

2 

Flights  >10  m from  tape  player,  songs 
elicited 

1 

minimal  interest  shown  in  recording,  mov- 
ing away  from  area 

0 

no  interest,  left  area  during  trial 

ditional  Black-capped  Chickadee  range  did 
not  respond  significantly  more  aggressively  to 
Black-capped  calls  than  to  Carolina  calls  (/  = 
1.35,  df  = 12,  F > 0.05;  Fig.  3).  However, 
chickadees  within  traditional  Carolina  Chick- 
adee range  were  more  aggressive  towards  Car- 
olina calls  than  Black-capped  calls  (r  = 2.75, 
df  = 20,  F < 0.01;  Fig.  3). 

When  sites  were  analyzed  separately,  only 
chickadees  at  Fox  Ridge  State  Park  responded 
significantly  more  aggressively  towards  a giv- 
en chickadee  vocalization.  Chickadees  at  Fox 
Ridge  responded  more  aggresssively  to  Car- 
olina calls  than  Black-capped  chickadee  calls 
{t  = -2.74,  df  = 8,  F < 0.03;  Fig.  3). 

Aggressive  responses  towards  both  Black- 
capped  (r  = —0.71,  n = 5)  and  Carolina 
chickadee  vocalizations  (r  = —0.29,  n = 5) 
increased  at  decreasing  distances  to  the  con- 
tact zone,  although  neither  was  statistically 
significant  (F  > 0.05).  With  both  species  com- 
bined, the  relationship  (r  = —0.49,  F = 0.08, 
n = 10)  was  still  not  significant  (F  > 0.05). 
Futhermore,  chickadees  did  not  show  a sea- 


7-, 


Combined  Black-capped  Carolina 

Vocalization 


FIG.  2.  Mean  aggressive  response  of  chickadees 
in  paired  Black-capped  Chickadee-nuthatch  (n  = 5), 
Carolina  Chickadee-nuthatch  (n  = 5),  and  for  all  trials 
combined  (n  = 10).  Dark  bars  represent  the  response 
to  chickadee  calls  and  light  bars  represent  the  response 
to  nuthatch  vocalizations.  **  p < 0.01,  ***  P < 0.001. 


Study  Site 

FIG.  3.  Mean  aggressive  response  of  chickadees 
to  Black-capped  and  Carolina  chick-a-dee  vocaliza- 
tions at  all  study  sites.  Dark  bars  represent  the  response 
to  Black-capped  Chickadee  calls  and  light  bars  repre- 
sent the  response  to  Carolina  Chickadee  calls.  The 
study  sites  are  arranged  from  West  to  East  and  the 
contact  zone  is  noted  between  Shelbyville  and  Douglas 
Hart.  (Sample  sizes  were  6,  7,  5,  9,  7 respectively;  * 
* Z’  < 0.05.) 


366 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3.  September  1999 


sonal  difference  in  response,  when  all  trials 
were  combined  (r  = —0.12,  P > 0.05,  n = 
36).  Separately,  there  was  no  relationship  be- 
tween mean  aggressive  response  to  either 
Black-capped  or  Carolina  Chickadee  vocali- 
zations and  date  (r  = —0.28,  P > 0.05,  n — 
18  and  r = 0.05,  P > 0.05,  n = 18,  respec- 
tively). 

DISCUSSION 

The  fact  that  chickadees  responded  more 
aggressively  to  chickadee  calls  than  nuthatch 
vocalizations  indicated  that  chickadees  were 
able  to  discriminate  between  congeneric  and 
heterogeneric  vocalizations  (Fig.  2).  These  re- 
sults are  consistent  with  other  studies  that  sug- 
gest that  interspecific  territoriality  in  chicka- 
dees is  rare  except  with  other  chickadees  in 
contact  zones  (Brewer  1963,  Smith  1993). 

We  found  that  within  the  traditional  Caro- 
lina Chickadee  range,  chickadees  responded 
more  aggressively  to  presumed  conspecific 
calls.  This  is  consistent  with  the  results  of 
studies  on  buntings  in  allopatric  populations 
(Emlen  et  al.  1975)  and  in  tropical  birds  in 
Peru  (Robinson  and  Terborgh  1995).  Several 
researchers  found  that  chickadees  responded 
more  aggressively  to  their  own  song  type  than 
to  songs  of  other  chickadee  species  except  in 
the  contact  zone  where  they  responded  ag- 
gressively to  both  conspecific  and  heterospe- 
cific song  types  (Ratcliffe  and  Weisman  1986, 
Robbins  et  al.  1986b).  This  differs  from  our 
results  at  one  site.  Chickadees  at  Fox  Ridge 
State  Park  showed  significantly  more  aggres- 
sion to  presumed  conspecific  vocalizations 
than  towards  heterospecifics.  Overall,  chicka- 
dees clearly  responded  more  aggressively  to 
chickadee  calls  than  nuthatch  calls  with  little 
difference  between  presumed  hetero-  and  con- 
specific chickadee  calls.  This  suggests  that 
chickadees  may  either  not  perceive  nuthatches 
as  a competitive  threat  or  that  chickadees  near 
contact  zones  may  not  distinguish  between 
chickadee  species  calls. 

We  did  not  get  the  predicted  increase  in  ag- 
gression towards  heterospecifics  closer  to  the 
contact  zone.  It  is  possibile  that  the  maximum 
distance  from  the  contact  zone  used  in  this 
study  was  not  far  enough  to  detect  any  sig- 
nificant differences  in  aggression.  This  would 
suggest  that  chickadees  across  the  area  are  fa- 
miliar with  congeners.  Merritt  (1981)  sug- 


gested that  individual  chickadees  expand  and 
contract  their  ranges  seasonally.  As  the  ranges 
of  these  species  approach  each  other,  cogni- 
zance of  the  heterospecific  vocalization  should 
increase  (Ward  and  Ward  1974).  At  increasing 
distances  from  the  contact  zone,  there  could 
be  a point  where  the  mean  aggressive  respons- 
es would  be  significantly  lower  than  closer  to 
the  contact  zone. 

Two  other  possible  explanations  exist  for 
the  aggression  shown  towards  a presumed  het- 
erospecific call  within  species’  ranges:  mis- 
directed aggression  and  the  presence  of  hy- 
brids. Misdirected  aggression  could  arise  from 
mistaken  identity  (Murray  1971,  1981).  This 
is  possible  because  the  vocalizations  of  both 
species  are  similar  and  variable  between  in- 
dividuals, and  we  used  only  one  example  of 
each  vocalization  type  in  this  experiment 
(Mammen  and  Nowicki  1981,  Smith  1991). 
However,  this  aggression  could  be  intentional, 
because  response  to  heterospecific  calls  may 
promote  recognition  and  facilitate  heterospe- 
cific spacing  (Emlen  et  al.  1975,  Merritt  1981, 
Robinson  and  Terborgh  1995).  Hybridization 
may  be  more  common  in  contact  zones  than 
previously  thought  (Brewer  1963,  Johnston 
1971,  Rising  1968,  Robbins  et  al.  1986b, 
Ward  and  Ward  1974).  Thus,  the  lack  of  spe- 
cies specific  aggressive  responses  could  be  the 
result  of  the  presence  of  hybrids  that  are  fa- 
miliar with  and  respond  similarly  to  calls  of 
both  species.  If  hybridization  is  the  cause  for 
the  observed  interspecific  aggression,  it  is 
likely  that  interspecific  territoriality  may  not 
act  as  a gap  producing  mechanism.  Gaps  be- 
tween Black-capped  and  Carolina  chickadee 
ranges  may  occur  if  hybrids  within  these  gaps 
had  severely  reduced  fitness  (Brewer  1963). 
However,  other  factors  may  cause  gaps.  One 
such  example  is  the  lack  of  suitable  habitat  in 
gaps  areas  (Grubb  et  al.  1994). 

ACKNOWLEDGMENTS 

We  would  like  to  thank  Eastern  Illinois  University 
Council  for  Faculty  Research  for  partially  funding  this 
project.  Thanks  to  T.  Grubb  Jr.,  M.  Morrison,  and  3 
anonymous  reviwers  for  comments  on  this  manuscript. 

LITERATURE  CITED 

Baker,  M.  C.  1991.  Response  of  male  Indigo  and  Laz- 
uli buntings  and  their  hybrids  to  song  playback  in 
allopatric  and  sympatric  populations.  Behaviour 
119:225-242. 


Kershner  am!  Bolliiif’er  • PLAYBACK  RESPONSES  OE  CHICKADEES 


367 


Brewer,  R.  1963.  Ecological  and  reproductive  rela- 
tionships of  Black-capped  and  Carolina  chicka- 
dees. Auk  80:9-47. 

Brindley,  E.  L.  1991.  Response  of  European  Robins 
to  playback  of  song:  neighbor  recognition  and 
overlapping.  Anim.  Behav.  41:503-512. 

Censky,  E.  J.  and  M.  S.  Ficken.  1982.  Responses  of 
Black-capped  Chickadees  to  miiTors.  Wilson  Bull. 
94:590-593. 

Emlen,  S.  T,  J.  D.  Rising,  and  W.  L.  Thompson.  1975. 
A behavioral  and  morphological  study  of  sympat- 
ry  in  the  Indigo  and  Lazuli  buntings  of  the  Great 
Plains.  Wilson  Bull.  87:145-177. 

Ficken,  M.  S.  and  C.  M.  Weise.  1984.  A complex  call 
of  the  Black-capped  Chickadee  (Pants  atricapil- 
lus).  1.  Microgeographic  variation.  Auk  101:349- 
360. 

Grubb,  T.  C.,  Jr.,  R.  A.  Mauck,  and  S.  L.  Earnst. 
1994.  On  no-chickadee  zones  in  midwestern 
North  America:  evidence  from  the  Ohio  Breeding 
Bird  Atlas  and  the  North  American  Breeding  Bird 
Survey.  Auk  111:191-197. 

Hill,  B.  G.  and  M.  R.  Lein.  1989.  Natural  and  sim- 
ulated encounters  between  sympatric  Black- 
capped  Chickadees  and  Mountain  Chickadees. 
Auk  106:645-652. 

Johnston,  D.  W.  1971.  Ecological  aspects  of  hybrid- 
izing chickadees  in  Virginia.  Am.  Midi.  Nat.  85: 
124-134. 

Kroodsma,  D.  E.  1989.  Suggested  experimental  de- 
sign for  song  playbacks.  Anim.  Behav.  37:600- 
609. 

Lampe,  H.  M.  and  M.  C.  Baker.  1994.  Behavioral 
response  to  song  playback  by  male  and  female 
White-crowned  Sparrows  of  two  subspecies.  Bio- 
acoustics 5:171—185. 

Mammen,  D.  L.  and  S.  Nowicki.  1981.  Individual  dif- 
ferences and  within-flock  convergence  in  chicka- 
dee calls.  Behav.  Ecol.  Sociobiol.  9:179-186. 

Merritt,  P.  G.  1978.  Characteristics  of  Black-capped 
and  Carolina  chickadees  at  the  range  interface  in 
Northern  Indiana.  Jack-Pine  Warbler  56:170—179. 

Merritt,  P.  G.  1981.  Narrowly  disjunct  allopatry  be- 
tween Black-capped  and  Carolina  chickadees  in 
Northern  Indiana.  Wilson  Bull.  93:54-66. 

Murray,  B.  G.  1971.  The  ecological  consequences  of 
interspecific  territorial  behavior  in  birds.  Ecology 
52:414-423. 

Murray,  B.  G.  1981.  The  origins  of  adaptive  inter- 
specific tenitorialism.  Biol.  Rev.  56:1-22. 


Nowk  ki,  S.  1983.  Flock-specific  recognition  of  chick- 
adee calls.  Behav.  Ecol.  Sociobiol.  12:317-320. 

Peterson,  R.  T.  1983.  A held  guide  to  bird  songs  of 
eastern  and  central  North  America.  Houghton 
Mifflin  Co.  Boston,  Massachusetts. 

Popp,  J.  W.,  M.  S.  Ficken,  and  C.  M.  Weise.  1990. 
How  are  agonistic  encounters  among  Black- 
capped  Chickadees  resolved?  Anim.  Behav.  39: 
980-986. 

Ratcliffe,  L.  M.  and  R.  G.  Weisman.  1986.  Song 
sequence  discrimination  in  the  Black-capped 
Chickadee.  J.  Comp.  Psychol.  100:361-367. 

Rising,  J.  D.  1968.  A multivariate  assessment  of  inter- 
breeding between  the  chickadees.  Pants  atricap- 
illtis  and  P.  carolittensis.  Syst.  Zool.  17:160-169. 

Robbins,  C.  S.,  D.  Bystrak,  and  P.  H.  Geissler. 
1986a.  The  breeding  bird  survey:  its  hrst  hfteen 
years,  1965-1979.  U.S.  Fish  Wildl.  Serv.  Resour. 
Publ.  157:1-196. 

Robbins,  M.  B.,  M.  J.  Braun,  and  E.  A.  Tobey. 
1986b.  Moiphological  and  vocal  variation  across 
a contact  zone  between  the  chickadees  Pants  atri- 
capillus  and  P.  carolittensis.  Auk  103:655-666. 

Robinson,  S.  K.  and  J.  Terborgh.  1995.  Interspecihc 
aggression  and  habitat  selection  by  Amazonian 
birds.  J.  Anim.  Ecol.  64:1-1 1. 

SAS  Institute.  1994.  SAS/STAT  user’s  guide.  SAS 
Institute,  Inc.,  Cary,  North  Carolina. 

Schroeder,.D.  j.  and  R.  H.  Wiley.  1983.  Communi- 
cation with  shared  song  themes  in  Tufted  Titmice. 
Auk  100:414-424. 

Shackleton,  S.  a.,  L.  Ratcliffe,  and  D.  M.  Weary. 
1992.  Relative  frequency  parameters  and  song 
recognition  in  Black-capped  Chickadees.  Condor 
94:782-785. 

Slade,  N.  A.  and  P.  B.  Robertson.  1977.  Comments 
on  competitively-induced  disjunct  allopatry.  Occ. 
Pap.  Mus.  Nat.  His.  65:1-8. 

Smith,  S.  M.  1991.  The  Black-capped  Chickadee:  be- 
havioral ecology  and  natural  history.  Cornell 
Univ.  Press,  Ithaca,  New  York. 

Smith,  S.  M.  1993.  Black-capped  Chickadee  in  The 
birds  of  North  America,  no.  39.  (A.  Poole.  P.  Stet- 
tenheim,  and  E Gill,  Eds).  The  Academy  of  Nat- 
ural Sciences,  Philadelphia,  Pennsylvania;  The 
American  Ornithological  Union,  Washington  D.C. 

Tanner,  J.  T.  1952.  Black-capped  and  Carolina  chick- 
adees in  the  southern  Appalachian  Mountains. 
Auk  69:407-424. 

Ward,  R.  and  D.  A.  Ward.  1974.  Songs  in  contiguous 
populations  of  Black-cappcd  and  Carolina  chick- 
adees in  Pennsylvania.  Wilson  Bull.  86:344-356. 


Wilson  Bull.,  111(3),  1999,  pp.  368-375 


USE  OF  SONG  TYPES  BY  MOUNTAIN  CHICKADEES 

{POECILE  GAMBELI) 

MYRA  O.  WIEBE'-  AND  M.  ROSS  LEIN'  ^ 


ABSTRACT. — We  investigated  the  composition  and  function  of  individual  song  repertoires  in  Mountain 
Chickadees  (PoecUe  gambeli)  in  Alberta,  Canada.  Individual  males  had  repertoires  of  4—7  song  types,  but  three 
types  made  up  90%  of  all  songs.  We  tested  and  rejected  the  hypothesis  that  all  song  types  convey  the  same 
behavioral  messages.  Different  song  types  were  associated  with  different  behavioral  situations.  Males  used  3- 
note  songs  predominantly  during  undisturbed  singing  and  2-note  songs  predominantly  during  non-aggressive 
activity.  Three-note  songs  with  each  successive  note  lower  pitched  were  associated  with  male— male  interactions. 
We  suggest  that  different  song  types  convey  messages  indicating  different  levels  of  aggression  by  the  singer. 
The  function  of  individual  repertoires  in  Mountain  Chickadees  appears  to  be  similar  to  that  of  other  North 
American  chickadees  and  titmice,  with  different  song  types  having  different  communicative  functions.  Received 
7 Aug.  1998,  accepted  II  Feb.  1999. 


Although  individuals  of  some  species  of 
songbirds  sing  only  one  type  of  song  (Searcy 
1983),  males  of  many  species  possess  reper- 
toires comprising  a number  of  discrete  cate- 
gories of  songs,  or  song  types  (Dodson  and 
Lemon  1975).  The  significance  of  song  rep- 
ertoires has  been  the  subject  of  much  research 
and  speculation  (reviewed  by  Krebs  and 
Kroodsma  1980,  Kroodsma  1982,  Kroodsma 
and  Byers  1991,  MacDougall-Shackleton 
1998). 

Most  species  of  the  genera  Parus,  Poecile, 
and  Baeolophus  have  individual  song  reper- 
toires (Hailman  1989),  although  the  role  of 
repertoires  appears  to  differ  among  species 
[these  closely-related  genera  were  formerly 
lumped  as  Pcirus  (American  Ornithologists’ 
Union  1997)].  Individual  male  Great  Tits  {Pa- 
rus major)  may  use  different  song  types  in 
sequence  with  no  apparent  change  in  the  ex- 
ternal situation  (Hinde  1952),  suggesting  that 
all  song  types  convey  the  same  messages. 
Consequently,  Great  Tits  have  been  used  to 
test  many  of  the  hypotheses  that  suggest  that 
overall  repertoire  size  is  important  (e.g.,  Krebs 
1976,  1977;  McGregor  et  al.  1981;  Baker  et 
al.  1986;  Lambrechts  and  Dhondt  1988).  In 


' Divi.sion  of  Ecology  (Behavioral  Ecology  Group), 
Dept,  of  Biological  Sciences,  Univ.  of  Calgary,  Cal- 
gary, Alberta  T2N  1N4,  Canada; 

E-mail;  mrlein@ucalgary.ca 

2 Present  address:  Canadian  Wildlife  Service,  Suite 
301,  5204  50th  Avenue,  Yellowknife,  Northwest  Ter- 
ritories XI  A 1E2,  Canada; 

E-mail;  myra.robertson@ec.gc.ca. 

’ Corresponding  author. 


contrast,  other  species,  such  as  the  Blue  Tit 
{Parus  caeruleus),  the  North  American  tit- 
mice {Baeolophus),  and  some  of  the  North 
American  chickadees  {Poecile),  seem  to  pos- 
sess song  types  that  convey  different  messag- 
es (Smith  1972,  Dixon  and  Martin  1979,  Gad- 
dis 1983,  Schroeder  and  Wiley  1983,  Bijnens 
and  Dhondt  1984,  Johnson  1987). 

We  examine  the  role  of  individual  reper- 
toires of  Mountain  Chickadees  {Poecile  gam- 
beli). Mountain  Chickadees  sing  relatively 
simple  songs  consisting  of  2-6  whistled  notes, 
with  any  number  of  these  notes  shifted  to  fre- 
quencies lower  than  the  others  (Hill  1987). 
Song  types  are  defined  easily  by  variation  in 
number  and  pitch  of  notes;  individual  reper- 
toires consist  of  3-5  song  types  (Hill  and  Lein 
1989).  Although  no  studies  have  investigated 
whether  Mountain  Chickadees  can  differenti- 
ate among  song  types,  this  seems  probable  be- 
cause closely  related  Black-capped  Chicka- 
dees {Poecile  atricapillus)  distinguish  among 
songs  varying  in  note  number  and  pitch  (Rat- 
cliffe  and  Weisman  1986;  Weisman  and  Rat- 
cliffe  1989). 

We  documented  song  variation  and  singing 
behavior  of  male  Mountain  Chickadees  during 
the  breeding  season.  Our  null  hypothesis  was 
that  all  song  types  of  the  Mountain  Chickadee 
convey  the  same  behavioral  messages,  and 
thus  the  song  type  that  a male  sings  would  be 
independent  of  the  situation  in  which  it  is 
used.  The  alternate  hypothesis  was  that  song 
types  convey  different  messages  and  thus  cer- 
tain song  types  would  have  a higher  proba- 
bility of  being  sung  in  specific  situations. 


368 


Wiehe  and  Lean  • SONG  TYPES  OF  MOUNTAIN  CHICKADEES 


369 


METHODS 

The  study  was  conducted  at  the  Barrier  Lake  site 
(51°  00'  N,  I 15°  00'  W)  of  the  University  of  Calgary’s 
Kananaskis  Field  Stations  in  the  Kananaskis  Valley  of 
the  Rocky  Mountains  of  southwestern  Alberta.  The 
forest  inhabited  by  Mountain  Chickadees  is  dominated 
by  trembling  aspen  (Populus  tremuloides),  white 
spruce  (Picea  glaiica),  and  lodgepole  pine  (Pinus  con- 
torta).  Mountain  Chickadees  are  secondary  cavity- 
nesters,  using  pre-existing  cavities  such  as  natural 
crevices  or  deserted  nests  of  other  cavity-nesting  birds 
(Hill  1987,  pers.  obs.).  In  southwestern  Alberta,  Moun- 
tain Chickadees  start  to  search  for  suitable  nesting  cav- 
ities during  April  (Hill  1987,  pers.  obs.).  Nest-building 
occurs  in  early  May.  Incubation  starts  near  the  end  of 
May  and  lasts  about  14  days.  The  nestling  period  gen- 
erally lasts  18-21  days  (Dahlsten  and  Copper  1979) 
but  may  be  as  short  as  14  days  (pers.  obs.).  Nestlings 
fledge  during  the  last  week  in  June  and  early  July. 

Mountain  Chickadee  songs,  and  information  on  the 
situations  of  song  use,  were  recorded  from  April  to 
July  1993.  Data  were  collected  from  1 1 males,  of 
which  7 were  marked  with  a unique  combination  of 
colored  plastic  leg  bands.  We  could  identify  unbanded 
males  because  males  in  adjacent  territories  were  band- 
ed; we  also  used  territorial  location,  use  of  favorite 
singing  sites,  or  association  with  a particular  nest  cav- 
ity to  identify  unbanded  males.  All  males  except  one 
foraged  frequently  with  one  other  Mountain  Chickadee 
and  therefore  were  presumed  to  be  mated.  Because 
Mountain  Chickadees  defend  their  entire  home  range 
(Hill  1987),  we  determined  the  extent  of  a male’s  ter- 
ritory by  noting  areas  in  which  he  was  found  regularly. 
Observations  of  aggressive  interactions  between  males 
helped  to  confirm  the  locations  of  territorial  boundar- 
ies. 

It  was  impossible  to  sample  songs  of  all  individuals 
in  one  day,  but  usually  each  individual  was  observed 
at  least  once  in  a three-day  period.  Observations  oc- 
curred between  04:00  and  12:00  (MST),  the  period 
when  chickadees  are  most  active.  An  observation  pe- 
riod began  when  a male  started  to  sing  and  lasted  from 
a few  minutes  to  over  an  hour,  depending  on  how  long 
the  individual  sang.  Songs  and  verbal  descriptions  of 
the  different  elements  of  the  situation  during  singing 
were  recorded  onto  Sony  C-90HF  cassette  tapes  using 
AKG  D190E  or  Sony  ECM-33P  microphones,  Sony 
PBR-330  parabolic  reflectors,  and  Sony  TCM-5000  or 
Sony  TCM-5000EV  tape  recorders. 

All  recordings  were  made  by  either  MOW  or  a field 
assistant.  Before  working  independently,  the  field  as- 
sistant accompanied  MOW  during  15  observation  pe- 
riods to  ensure  that  both  observers  were  making  com- 
parable observations.  We  did  not  notice  chickadees  en- 
gaging in  behaviors  directed  at  the  observer  during  any 
of  the  observation  periods,  so  we  believe  that  the  effect 
of  our  pre.sence  was  minimal. 

We  recorded  three  elements  of  the  situation  while 
males  were  singing.  (1)  Behavior  of  the  singer:  a.  un- 
disturbed (i.e.,  singing  while  stationary  and  not  en- 


gaged in  other  activities);  b.  engaged  in  non-aggressive 
activities  (e.g.,  foraging,  preening,  feeding  nestlings, 
etc.);  c.  engaged  in  aggre.ssive  activities  (i.e.  counter- 
singing with  or  chasing  territorial  neighbors).  (2)  Po- 
sition on  territory:  a.  within  50  m of  the  nest  site;  b. 
away  from  the  ne.st  site.  (3)  Location  of  mate:  a.  pres- 
ent in  the  vicinity  of  the  singer;  b.  absent  from  the 
vicinity.  Each  song  was  assigned  to  a particular  situ- 
ation. If  the  situation  changed  while  the  male  was  sing- 
ing, then  songs  recorded  before  the  change  were  as- 
signed to  the  first  situation,  and  songs  recorded  after 
the  change  assigned  to  the  second  situation.  Nests  of 
six  pairs  were  found,  allowing  us  to  also  assign  songs 
for  these  males  to  particular  breeding  stages  (nest 
search/build,  egg-laying,  incubation,  or  nestling). 

Mountain  Chickadees  often  were  out  of  sight  when 
in  the  tops  of  coniferous  trees.  We  assumed  that  songs 
given  during  such  intervals  were  in  the  same  situation 
as  the  last  song  given  prior  to  disappearance.  However, 
if  a subject  was  out  of  sight  for  more  than  5 minutes, 
we  recorded  its  behavior  for  that  interval  as  unknown. 

We  categorized  different  song  types  by  variation  in 
number  of  notes  and  relative  pitch  of  notes  within  a 
song.  These  two  features  showed  the  most  obvious 
variation  among  songs  and  were  relatively  easy  to  dis- 
tinguish by  ear.  We  confirmed  these  classifications  by 
examining  audiospectrographs  of  many  songs  using 
SIGNAL  bioacoustical  analysis  software  (Engineering 
Design,  Belmont,  Massachusetts).  Transcriptions  of  re- 
cordings were  made  using  OBSER'VER  software  (Nol- 
dus  Information  Technology,  Wageningen,  The  Neth- 
erlands). 

We  examined  separately  the  relative  importance  of 
number  of  notes  in  a song  and  the  pitch  of  notes  within 
a song.  To  determine  the  influence  of  number  of  notes, 
we  combined  all  song  types  with  the  same  number  of 
notes  regardless  of  the  pitch  of  the  notes  within  the 
song.  Statistical  tests  were  performed  only  for  2-note 
and  3-note  songs  because  sample  sizes  for  songs  with 
1 note  and  4 notes  or  more  were  too  small.  To  examine 
the  influence  of  pitch,  statistical  tests  were  performed 
only  for  “common  3-note  songs’’  and  “descending  3- 
note  songs”  (see  descriptions  in  Results).  Sample  sizes 
of  other  song  types  with  variations  in  pitch  were  too 
small  for  statistical  testing. 

If  different  song  types  are  not  used  in  different  sit- 
uations, one  would  predict  that  a particular  song  type 
should  occur  in  a specific  situation  at  the  frequency 
expected  if  song  types  are  used  at  random.  To  examine 
the  influence  of  note  number  in  songs,  we  calculated 
X'  values  for  2X2  contingency  tables  of  either  the 
relative  number  of  2-note  songs  and  all  other  song 
types  compared  between  specific  situations  of  use,  or 
the  relative  number  of  3-note  songs  and  all  other  song 
types  compared  between  specific  situations  of  use.  To 
examine  the  influence  of  pitch  in  songs,  we  calculated 

values  for  2 X 2 contingency  tables  of  the  number 
of  common  3-note  songs  and  descending  3-note  songs 
compared  between  specific  situations  of  use.  The  in- 
dependent event  in  all  contingency  tables  was  a single 
song. 


370 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  3,  September  1999 


We  conducted  a separate  contingency  analysis  for 
each  individual  for  every  situation  of  use.  We  were  not 
able  to  record  all  individuals  in  different  situations. 
Consequently,  the  number  of  individuals  used  in  each 
comparison  varied  from  four  to  seven.  We  combined 
the  results  of  individual  contingency  analyses  using  a 
test  described  by  Cochran  (1971:151).  This  test  ac- 
counts for  differences  in  direction  of  response  among 
individuals  and  can  be  used  even  if  there  is  a wide 
range  of  sample  sizes  and  probabilities  among  individ- 
uals. The  test  calculates  a z-value  that  can  be  compared 
to  the  normal  distribution  to  determine  significance, 
with  non-significant  results  indicating  that  the  song 
types  are  independent  of  the  situation  of  use.  The  sign 
of  the  z-value  indicates  the  direction  of  deviation  from 
the  expected. 

Most  statistical  tests  were  performed  using  STATIS- 
TIX  3.5  for  Windows  (Analytical  Software,  St.  Paul, 
Minnesota).  Differences  at  an  a < 0.05  were  consid- 
ered to  be  significant. 

RESULTS 

Composition  of  song  repertoires. — We  re- 
corded an  average  of  1,501  songs  (±  435  SE; 
range  = 222-4,372)  from  each  of  the  1 1 focal 
males,  with  an  individual  male  being  recorded 
an  average  of  15.1  days  (±  2.2;  range  = 6- 
28).  Average  total  repertoire  size  was  7.4  song 
types  (±  0.3;  range  = 6-9).  However,  song 
types  used  rarely  by  an  individual  may  have 
been  “accidental”  productions  rather  than 
regular  elements  in  the  repertoire.  Excluding 
song  types  that  represented  less  than  1%  of 
the  total  songs  for  an  individual,  the  average 
repertoire  size  was  5.1  song  types  (±  0.3; 
range  = 4-7).  There  was  no  relationship  be- 
tween the  number  of  songs  recorded  from  an 
individual  and  the  estimated  size  of  his  rep- 
ertoire (all  song  types:  Spearman  r = 0.36,  P 
> 0.05,  n = 11;  song  types  > 1%:  Spearman 
r = -0.27,  P > 0.05,  n = 11),  indicating  that 
all  repertoires  were  sampled  adequately. 

“Common  3-note  songs”  (3-note  songs 
with  the  last  two  notes  lower  in  pitch  than  the 
first  note.  Fig.  lA),  “common  2-note  songs” 
(2-note  songs  with  the  second  note  lower- 
pitched,  Fig.  IB),  and  “descending  3-note 
songs”  (3-note  songs  with  each  successive 
note  at  least  200  Hz  lower  in  pitch  than  the 
previous  note.  Fig.  1C)  were  the  prevalent 
song  types.  The  most-frequent  song  type  was 
common  2-note  song  for  5 of  the  1 1 focal 
males,  common  3-note  song  for  4 males,  and 
descending  3-note  song  for  2 males.  All  males 
sang  these  three  song  types,  but  for  three  in- 


6- 

4- 

A 

B 

. 2-1 

N 

X 

^ 6- 

■ 

o 

1 -T  1 1 1 H 

C 

1 

D 

C 4- 
0) 

3 

O" 

O’  O. 

k 2- 

U- 

6- 

E 

F 

4- 

2- 

1 1 1 1 j 

m m m 

1 1 ’"'I 1 1 

Time  (seconds) 

PIG.  1.  Audiospectrograms  of  song  types  most 
commonly  used  by  Mountain  Chickadees  on  the  study 
area.  A.  Common  3-note  song.  B.  Common  2-note 
song.  C.  Descending  3-note  song.  D.  2-note  song  with 
both  notes  the  same  pitch.  E.  Three-note  song  with  all 
notes  the  same  pitch.  F.  Typical  4-note  song.  All  high- 
quality  recordings  of  songs  had  the  short,  upswept  note 
at  the  beginning  of  the  song  that  is  present  in  the  com- 
mon 3-note,  the  common  2-note  song,  and  the  de- 
scending 3-note  song  shown  in  this  figure. 

dividuals  < 4%  of  total  songs  recorded  were 
descending  3-note  songs.  All  other  variants 
had  frequencies  of  < 12%  of  an  individual’s 
repertoire,  including  2-note  and  3-note  songs 
with  all  notes  of  the  same  pitch  (i.e.,  within 
200  Hz  of  one  another;  Figs.  ID,  IE)  and 
songs  with  four  or  more  notes  (Fig.  IF).  Wie- 
be  (1995)  gives  details  of  frequencies  of  use 
of  song  types  by  each  focal  male  and  the  use 
of  2-note  songs,  common  3-note  songs  and 
descending  3-note  songs,  in  different  situa- 
tions by  individual  males.  These  data  are  sum- 
marized in  the  following  sections. 

Variation  in  song  use  among  breeding  stag- 
es.— Two-note  and  3-note  songs  were  not 
used  at  random  during  many  breeding  stages, 
although  there  was  much  individual  variation 
in  the  relative  use  of  these  song  types  within 
all  stages  (Tables  1,  2).  Likewise,  common 
and  descending  3-note  songs  were  not  used  at 
random  during  different  stages  and  there  was 
also  much  individual  variation  (Table  3). 


Wiehe  and  Lean  • SONG  TYPES  OF  MOUNTAIN  CHICKADEES 


371 


TABLE  I.  The  use  of  2-note  songs  by  male  Mountain  Chickadees  in  various  situations,  compared  to  all 
other  song  types.  Columns  headed  “AM  males”  refer  to  combined  analysis  of  contingency  tables  for  individual 
males.  Columns  headed  “Individual  males”  indicate  numbers  of  contingency  tables  for  individual  males  with 
significant  departures  from  expectations. 


Situation 

All  males 

Individual  males 

Preferred  song  type 

n 

Preferred  song^ 

c-value 

p 

2-note 

Other 

None 

Nest  searching/building  stage 

5 

Other 

-6.77 

<0.01 

2 

3 

0 

Egg-laying  stage 

6 

+ 1.86 

0.06 

3 

1 

2 

Incubation  stage 

6 

2-note 

+ 5.99 

<0.01 

3 

1 

2 

Nestling  stage 

4 

Other 

-2.30 

0.02 

1 

1 

2 

Near  nest 

6 

2-note 

+3.94 

<0.01 

4 

0 

2 

Mate  present 

7 

Other 

-2.45 

0.01 

1 

3 

3 

Undisturbed  singing 

7 

Other 

-9.1 1 

<0.01 

0 

5 

2 

Male-male  interaction 

6 

-0.22 

0.83 

0 

0 

6 

“ Indicate.s  that  2-note  songs  were  used  significantly  more  frequently  (2-note)  or  less  frequently  (Other)  than  expected  in  that  situation.  A blank  indicates 
no  significant  deviation  from  expected  frequencies. 


Variation  in  song  use  among  behavioral  sit- 
uations.— When  singing  near  the  nest  site, 
males  used  2-note  songs  significantly  more 
frequently  (Table  1)  and  3-note  songs  signifi- 
cantly less  frequently  (Table  2)  than  expected. 
This  pattern  also  is  reflected  in  the  song  types 
used  preferentially  by  individuals.  Common 
3-note  songs  were  less  frequent  than  expected 
(and  descending  3-note  songs  more  frequent) 
when  males  were  singing  near  the  nest,  al- 
though there  was  much  individual  variation 
(Table  3). 

Males  sang  2-note  songs  significantly  less 
frequently  than  expected  in  the  presence  of 
their  mates,  as  indicated  by  the  combined 
analysis  and  by  the  preferred  song  types  of 
individuals  (Table  1).  Common  3-note  songs 
were  used  significantly  more  frequently  than 
expected  in  this  situation,  with  five  of  seven 


individuals  using  common  3-note  songs  pref- 
erentially (Table  3). 

When  engaged  in  undisturbed  singing,  in 
comparison  to  singing  while  engaged  in  an- 
other non-aggressive  activity  such  as  foraging, 
2-note  songs  were  used  significantly  less  fre- 
quently (Table  1),  and  3-note  songs  signifi- 
cantly more  frequently  (Table  2)  than  expect- 
ed. Song  type  preferences  of  individuals  are 
entirely  concordant  with  this  overall  pattern. 
Frequencies  of  common  and  descending  3- 
note  songs  did  not  differ  from  expectations 
during  undisturbed  singing  (Table  3). 

During  male— male  interactions,  common  3- 
note  songs  were  used  significantly  less  fre- 
quently, and  descending  3-note  songs  signifi- 
cantly more  frequently,  than  expected  (Table 
3).  Three  of  five  males  showed  significant 
preferential  use  of  descending  3-note  songs 


TABLE  2.  The  use  of  3-note  songs  by  male  Mountain  Chickadees  in  various  situations,  compared  to  all 
other  song  types.  See  Table  I for  an  explanation  of  the  format. 


Situation 

All  males 

Individual  males 

Preferred  song  type 

n 

Preferred  song“ 

z-value 

p 

3-note 

Other 

None 

Nest  searching/building  stage 

5 

3-note 

+ 5.71 

<0.01 

3 

2 

0 

Egg-laying  stage 

6 

-0.76 

0.45 

1 

3 

2 

Incubation  stage 

6 

Other 

-5.55 

<0.01 

0 

3 

3 

Nestling  stage 

4 

+ 1.55 

0.12 

1 

1 

2 

Near  nest 

6 

Other 

-4.82 

<0.01 

0 

4 

2 

Mate  present 

7 

+ 1.48 

0.14 

3 

2 

2 

Undisturbed  singing 

7 

3-note 

+ 10.34 

<0.01 

5 

0 

2 

Male-male  interaction 

6 

+0.59 

0.56 

0 

1 

5 

^ Indicates  that  3-note  songs  were  u.sed  significantly  more  frequently  (3-note)  or  less  frequently  (Other)  than  expected  in  that  situation.  A blank  indicates 
no  significant  deviation  from  expected  frequencies. 


372 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


TABLE  3.  The  use  of  common  3-note  songs  by  male  Mountain  Chickadees  compared  to  the  use  of  descend- 
ing 3-note  songs  in  various  situations.  See  Table  1 for  an  explanation  of  the  format. 

Individual  males 
Preferred  song  type 


All  males 

Descend- 

- Common 

ing 

Situation 

n 

Preferred  song“ 

z-value 

p 

3-note 

3-note 

None 

Nest  searching/building  stage 

4 

Common 

+ 8.73 

<0.01 

3 

1 

0 

EgR-laying  stage 

6 

Common 

+ 8.36 

<0.01 

3 

2 

1 

Incubation  stage 

5 

Descending 

-10.73 

<0.01 

2 

3 

0 

Nestling  stage 

4 

Descending 

-5.65 

<0.01 

1 

2 

1 

Near  nest 

6 

Descending 

-4.35 

<0.01 

3 

2 

1 

Mate  present 

7 

Common 

-6.56 

<0.01 

5 

1 

1 

Undisturbed  singing 

4 

+0.94 

0.35 

2 

1 

1 

Male-male  interaction 

5 

Descending 

-6.72 

<0.01 

0 

3 

2 

^ Indicates  the  3-note  song  type  that  was  used  significantly  more  frequently  than  expected  in  that  situation.  A blank  indicates  no  significant  deviation 
from  expected  frequencies. 


during  this  situation  whereas  none  used  com- 
mon 3-note  songs  preferentially. 

DISCUSSION 

Individual  song  repertoires. — Excluding  in- 
frequent song  types,  males  had  individual  rep- 
ertoires of  4-7  song  types.  Hill  and  Lein 
(1989)  estimated  a slightly  smaller  repertoire 
size  of  3-5  song  types.  However,  they  sam- 
pled songs  less  intensively  and  may  not  have 
recorded  enough  songs  to  obtain  complete 
song  repertoires  for  all  individuals. 

Other  researchers  imply  that  the  most  com- 
mon song  type  of  Mountain  Chickadees  is  a 
3-note  song  with  all  notes  of  the  same  pitch 
(Gaddis  1985,  Hailman  1989,  Hill  and  Lein 
1989).  In  contrast,  we  found  that  males  in  our 
study  rarely  sang  songs  of  this  type.  We  are 
unable  to  explain  this  difference,  but  it  may 
support  Gaddis’  (1985)  suggestion  of  geo- 
graphical variation  in  Mountain  Chickadee 
songs. 

Use  of  .song  types  in  different  situations. — 
Our  analyses  of  song  length  grouped  different 
song  types  with  the  same  numbers  of  notes 
into  a single  category,  possibly  obscuring 
some  patterns  in  the  use  of  different  song 
types.  However,  because  almost  all  2-note 
songs  recorded  were  common  2-note  songs, 
the  results  from  analyses  of  all  2-note  songs 
would  be  almost  identical  to  results  using  only 
common  2-note  songs.  Almost  all  3-note 
songs  recorded  were  either  common  or  de- 
scending 3-note  songs.  Any  differences  be- 
tween these  song  types  that  were  obscured  in 


the  song  length  analyses  should  be  revealed 
in  the  song  pitch  analyses. 

Different  song  types  had  significant  asso- 
ciations with  different  breeding  stages.  How- 
ever, there  was  also  much  individual  variation 
in  all  stages,  with  different  males  using  dif- 
ferent songs  preferentially  during  the  same 
breeding  stage.  This  suggests  that  although  the 
z-values  were  significant  statistically,  the  re- 
sults may  not  be  meaningful  biologically.  Dif- 
ferent song  types  are  probably  not  signaling 
messages  about  the  breeding  stage  of  the  sing- 
er. Instead,  different  breeding  stages  may  be 
associated  with  other  situations  that  are  more 
relevant  biologically  to  the  messages  of  the 
song  type. 

The  number  of  notes  in  songs  of  Mountain 
Chickadees  was  associated  significantly  with 
location  relative  to  the  singer’s  nest.  All  in- 
dividuals showing  significant  preferences  ei- 
ther used  more  2-note  songs  near  the  nest  than 
away  from  it,  or  more  3-notes  away  from  the 
nest  than  near  it.  Some  other  studies  on  song- 
birds also  have  found  that  males  sang  different 
song  types  depending  on  their  territorial  lo- 
cation (e.g.,  Lein  1978,  Weary  et  al.  1994). 
Although  pitch  of  the  last  note  in  3-note  songs 
was  associated  with  location  relative  to  the 
nest,  this  association  is  weak  because  of  the 
large  degree  of  individual  variation. 

In  Mountain  Chickadees,  the  tendency  to 
sing  2-note  songs  near  the  nest  and  3-note 
songs  away  from  the  nest  may  be  influenced 
by  factors  other  than  just  the  singer’s  location. 
Males  were  more  likely  to  engage  in  non-ag- 


Wiehe  am!  Lein  • SONG  TYPES  OF  MOUNTAIN  CHICKADEES 


373 


gressive  activities  while  singing  near  the  nest 
than  when  singing  away  from  the  nest,  sug- 
gesting that  location  relative  to  the  nest  and 
male  behavior  were  related.  Although  note 
number  may  communicate  some  information 
about  the  singer’s  location,  these  factors  may 
be  related  only  indirectly  and  the  association 
between  male  behavior  and  note  number  may 
be  more  important  biologically. 

Different  song  types  were  associated  with 
the  presence  or  absence  of  the  singer’s  mate. 
Other  researchers  (e.g.,  Temrin  1986,  Staicer 
1989)  have  claimed  that  song  types  used  pref- 
erentially in  the  presence  of  females  have  a 
greater  intersexual  function.  One  methodolog- 
ical problem  in  our  study  is  that  a female  may 
have  been  recorded  as  absent  when  she  actu- 
ally was  nearby  in  the  nest  hole,  but  not  vis- 
ible to  the  observer.  Two-note  songs  and  de- 
scending 3-note  songs,  which  were  positively 
associated  with  the  absence  of  the  female, 
were  also  positively  associated  with  the  nest 
site.  Furthermore,  in  most  cases  singers  did 
not  seem  to  be  directing  song  specifically  at 
females.  Males  sometimes  sang  low-volume, 
“quiet”  songs,  usually  of  1 or  2 notes,  when 
approaching  nests  to  feed  incubating  mates. 
Although  such  songs  may  be  directed  specif- 
ically at  females,  we  never  observed  normal 
volume  songs  used  in  this  manner.  Therefore, 
we  are  hesitant,  without  further  experimental 
study,  to  suggest  that  common  3-note  songs 
have  a greater  intersexual  function  than  do  2- 
note  songs  and  descending  3-note  songs. 

The  behavior  of  singing  males  was  associ- 
ated with  the  number  of  notes  in  the  song,  but 
not  with  pitch.  Two-note  songs  were  positive- 
ly associated  with  singing  while  engaged  in 
non-aggressive  activity  and  3-note  songs  were 
positively  associated  with  undisturbed  sing- 
ing. Although  not  all  males  had  individual  re- 
sults that  were  significant,  all  males  showed 
this  trend. 

We  suggest  that  3-note  songs  signal  a high- 
er motivation  level  of  the  singer  to  sing  than 
do  2-note  songs.  If  so,  3-note  songs  might  in- 
dicate that  the  singer  is  more  willing  to  en- 
gage in  some  of  the  agonistic  actions  associ- 
ated with  singing  in  Mountain  Chickadees, 
such  as  countersinging  bouts  or  interacting 
with  other  males  at  the  edge  of  the  territory 
to  confirm  boundaries.  Thus,  3-note  songs 


could  convey  more  aggressive  messages  than 
2-note  songs. 

Male-male  interactions  in  Mountain  Chick- 
adees were  associated  with  changes  in  pitch 
of  notes  in  the  song.  Because  most  males  were 
more  likely  to  sing  common  3-note  songs 
when  apparently  unprovoked  by  another  bird’s 
activities  than  during  interactions  with  rival 
males,  this  song  type  may  function  in  spon- 
taneous advertisement  of  the  territory.  De- 
scending 3-note  songs  were  associated  with 
male-male  interactions  and  it  is  probable  that 
lowering  the  pitch  of  the  last  note  in  3-note 
songs  may  convey  some  message  to  the  rival. 
Other  researchers  have  suggested  that  song 
types  associated  with  male-male  interactions 
probably  convey  more  aggressive  messages 
than  do  other  song  types  (e.g..  Nelson  and 
Croner  1991).  Interactions  between  males  are 
situations  of  high  levels  of  agonistic  stimula- 
tion for  Mountain  Chickadees  and  so  they 
may  be  more  likely  to  use  songs  that  convey 
stronger  aggressive  tendencies  at  this  time. 
Thus,  descending  3-note  songs  may  convey 
more  aggressive  messages  than  common  3- 
note  songs. 

Comparison  to  closely-related  species. — 
Our  findings  suggest  that  the  function  of  in- 
dividual repertoires  in  Mountain  Chickadees 
is  similar  to  that  of  other  North  American 
chickadees  and  titmice,  with  different  song 
types  used  in  different  situations  and  appear- 
ing to  have  different  communicative  func- 
tions. Three  species  of  North  American  tit- 
mice have  certain  song  types  that  are  used 
predominantly  in  male-male  interactions 
(Gaddis  1983,  Schroeder  and  Wiley  1983, 
Johnson  1987).  These  are  probably  similar  in 
function  to  the  descending  3-note  song  of  the 
Mountain  Chickadee,  which  is  also  used  in 
male-male  interactions.  The  Bridled  Titmouse 
(Baeolophus  wollweheri)  has  one  song  type 
used  predominantly  in  spontaneous  advertise- 
ment of  territory  (Gaddis  1983).  and  we  found 
that  the  common  3-note  song  of  the  Mountain 
Chickadee  is  used  predominantly  in  undis- 
turbed singing.  Schroeder  and  Wiley  (1983) 
suggested  that  different  song  types  of  the  Tuft- 
ed Titmouse  (B.  bicolor)  convey  different  lev- 
els of  aggression  by  the  singer.  This  corre- 
sponds to  our  suggestion  that,  in  Mountain 
Chickadees,  descending  3-note  songs,  com- 
mon 3-note  songs,  and  2-note  songs  indicate 


374 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


high,  intermediate,  and  low  levels  of  aggres- 
sion, respectively. 

There  are  also  some  differences  in  the  man- 
ner in  which  titmice  and  Mountain  Chicka- 
dees use  songs.  For  instance,  the  Bridled  Tit- 
mouse has  a song  type  used  predominantly  in 
long-distance  countersinging  (Gaddis  1983). 
We  did  not  find  any  song  type  in  Mountain 
Chickadees  that  was  used  in  this  manner,  al- 
though it  is  possible  that  some  songs  that  we 
categorized  as  “undisturbed”  singing  may 
have  actually  been  in  response  to  far  away 
song  that  was  inaudible  to  the  observer.  John- 
son (1987)  noted  that  the  Plain  Titmouse 
{Baeolophus  inornatus)  was  more  likely  to 
use  some  song  types  in  situations  related  to 
nesting  activities.  Although  2-note  songs  were 
associated  with  close  proximity  to  the  nest, 
there  was  no  indication  that  Mountain  Chick- 
adees were  using  these  songs  in  any  way  that 
was  related  specifically  to  nesting  activity. 

Carolina  Chickadees  (Poecile  carolinensis) 
have  one  song  type  associated  with  counter- 
singing that  is  thought  to  be  a more  aggressive 
song  type  (Smith  1972).  This  song  type  could 
be  similar  in  function  to  descending  3-note 
songs  given  by  Mountain  Chickadees  during 
male— male  interactions.  However,  Smith 
(1972)  also  observed  that  Carolina  Chicka- 
dees were  more  likely  to  use  this  aggressive 
song  type  while  patrolling  territorial  bound- 
aries whereas  we  did  not  note  any  strong  as- 
sociation between  descending  3-note  songs 
and  territorial  boundaries. 

Pitch  may  be  an  important  cue  in  coding 
information  in  the  songs  of  both  Mountain 
and  Black-capped  chickadees,  but  the  two 
species  differ  in  how  they  vary  the  pitch  of 
their  songs.  Unlike  Black-capped  Chickadees, 
which  shift  the  entire  song  downward  in  pitch. 
Mountain  Chickadees  shift  individual  notes  in 
a song  to  a lower  pitch.  Black-capped  Chick- 
adee songs  shifted  downward  in  pitch  were 
observed  during  countersinging  between 
males  in  the  field  and  in  response  to  playback 
in  both  wild  and  captive  birds  (Ratcliffe  and 
Weisman  1985,  Hill  and  Lein  1987).  Moun- 
tain Chickadees  3-note  songs,  with  the  last 
note  lower  in  pitch  (descending  3-note  songs), 
were  associated  with  male-male  interactions 
whereas  3-note  songs  with  the  last  two  notes 
of  the  same  pitch  (common  3-note  songs) 
were  associated  with  the  less-aggressive  situ- 


ation of  territorial  advertisement.  In  both  spe- 
cies, pitch  seems  to  be  lowered  during  more 
aggressive  situations.  Morton  (1977)  suggest- 
ed that  calls  of  low  pitch  indicate  higher  ag- 
gressiveness by  the  signaler  than  do  calls  of 
higher  pitch.  This  idea  seems  to  be  applicable 
to  song  in  Mountain  Chickadees  and  Black- 
capped  chickadees. 

ACKNOWLEDGMENTS 

S.  R.  M.  Shima  provided  able  assistance  in  collec- 
tion of  the  data.  R Rodriguez  de  la  Vega,  M.  R.  Evans, 

J.  Bolstad,  J.  Goddard,  and  K.  Olson  also  assisted  with 
field  work.  We  thank  G.  Chilton  for  his  help  and  ad- 
vice during  various  stages  of  the  project.  The  staff  at 
the  Barrier  Lake  site  of  the  University  of  Calgary’s 
Kananaskis  Field  Stations  provided  support  and  ac- 
commodation during  the  field  season.  R.  M.  R.  Bar- 
clay, A.  P.  Russell,  P.  A.squith,  and  J.  P.  Hailman  gave 
constructive  comments  on  earlier  drafts  of  the  manu- 
script. This  research  was  supported  by  a graduate 
scholarship  from  Ellis  Bird  Farm  Limited  and  research 
assistantships  from  the  University  of  Calgary  to  MOW, 
and  a research  grant  from  the  Natural  Sciences  and 
Engineering  Research  Council  of  Canada  to  MRL. 

LITERATURE  CITED 

American  Ornithologists’  Union.  1997.  Forty-first 
supplement  to  the  American  Ornithologists’ 
Union  Check-list  of  North  American  Birds.  Auk 
1 14:542-552. 

Baker,  M.  C.,  T H.  Bjerke,  H.  Lampe,  and  Y.  Esp- 
MARK.  1986.  Sexual  response  of  female  Great  Tits 
to  variation  in  size  of  males'  song  repertoires.  Am. 
Nat.  128:491-498. 

Bijnens,  L.  and  a.  a.  Dhondt.  1984.  Vocalizations  in 
a Belgian  Blue  Tit,  Pants  c.  caeruleus,  popula- 
tion. Le  Gerfaut  74:243-269. 

Cochran,  W.  G.  1971.  Some  methods  for  strengthen- 
ing the  common  x’  test.  Pp.  132—156  in  Readings 
in  statistics  for  the  behavioral  scientist  (J.  A.  Ste- 
ger,  Ed.).  Holt,  Rinehart,  and  Winston,  New  York. 
Dahlsten,  D.  L.  and  W.  A.  Copper.  1979.  The  use  of 
nesting  boxes  to  study  the  biology  of  the  Moun- 
tain Chickadee  (Pants  gamheli)  and  its  impact  on 
selected  forest  insects.  Pp.  217-260  in  The  role  of 
insectivorous  birds  in  forest  ecosystems  (J.  G. 
Dickson,  R.  N.  Conner,  R.  R.  Fleet,  J.  C.  Kroll, 
and  J.  A.  Jackson,  Eds.).  Academic  Press,  New 
York. 

Dixon,  K.  L.  and  D.  J.  Martin.  1979.  Notes  on  the 
vocalization  of  the  Mexican  Chickadee.  Condor 
81:421-423.  . - 

Dodson,  D.  W.  and  R.  E.  Lemon.  1975.  Re-examina- 
tion of  monotony  threshold  hypothesis  in  bird 
song.  Nature  257:126-128. 

Gaddis,  P.  K.  1983.  Differential  usage  of  song  types 
by  Plain,  Bridled  and  Tufted  titmice.  Ornis  Scand. 
14:16-23. 


Wiehe  ami  Lein  • SONG  TYPES  OF  MOUNTAIN  CHICKADEES 


375 


Gaddis,  P.  K,  1985.  Structure  and  variability  in  the 
vocal  repertoire  of  the  Mountain  Chickadee.  Wil- 
son Bull.  97:30-46. 

H.ailman,  J.  P.  1989.  The  organization  of  major  vocal- 
izations in  the  Paridae.  Wilson  Bull.  101:305-343. 

Hill,  B.  G.  1987.  Tenitorial  behaviour  and  ecological 
relations  of  sympatric  Black-capped  (Pams  alri- 
capiUus)  and  Mountain  chickadees  (Pams  gam- 
beli)  in  southwestern  Alberta.  M.Sc.  thesis,  Univ. 
of  Calgary,  Calgary,  Alberta. 

Hill,  B.  G.  and  M.  R.  Lein.  1987.  Function  of  fre- 
quency-shifted songs  of  Black-capped  Chicka- 
dees. Condor  89:914-915. 

Hill,  B.  G.  and  M.  R.  Lein.  1989.  Natural  and  sim- 
ulated encounters  between  sympatric  Black- 
capped  Chickadees  and  Mountain  chickadees. 
Auk  106:645-652. 

Hinde,  R.  a.  1952.  The  behaviour  of  the  Great  Tit 
(Pams  major)  and  some  related  species.  Behav- 
iour, Suppl.  2:1-201. 

Johnson,  L.  S.  1987.  Pattern  of  song  types  use  for 
territorial  defense  in  the  Plain  Titmouse  Parus  in- 
ornatus.  Ornis  Scand.  18:24—32. 

Krebs,  J.  R.  1976.  Habituation  and  song  repertoires  in 
the  Great  Tit.  Behav.  Ecol.  Sociobiol.  1:297-303. 

Krebs,  J.  R.  1977.  The  significance  of  song  reper- 
toires: the  Beau  Geste  hypothesis.  Anim.  Behav. 
25:475-478. 

Krebs,  J.  R.  and  D.  E.  Kroodsma.  1980.  Repertoires 
and  geographical  variation  in  bird  song.  Adv. 
Study  Behav.  11:143-177. 

Kroodsma,  D.  E.  1982.  Song  repertoires:  problems  in 
their  definition  and  use.  Pp.  125-146  in  Acoustic 
communication  in  birds.  Vol.  2.  Song  learning  and 
its  consequences  (D.  E.  Kroodsma  and  E.  H.  Mill- 
er, Eds.).  Academic  Press,  New  York. 

Kroodsma,  D.  E.  and  B.  E.  Byers.  1991.  The  func- 
tionfs)  of  bird  song.  Am.  Zool.  31:318-328. 

Lambrechts,  M.  and  a.  A.  Dhondt.  1988.  The  anti- 
exhaustion hypothesis:  a new  hypothesis  to  ex- 
plain .song  performance  and  song  switching  in  the 
Great  Tit.  Anim.  Behav.  36:327-334. 

Lein,  M.  R.  1978.  Song  variation  in  a population  of 
Chestnut-sided  Warblers  (Dendroica  pensylvani- 
ca):  its  nature  and  suggested  significance.  Can.  J. 
Zool.  56:1266-1283. 


MacDougali.-Shackleton,  S.  a.  1998.  Sexual  selec- 
tion and  the  evolution  of  song  repertoires.  Curr. 
Ornithol.  14:81-124. 

McGregor,  P.  K.,  J.  R.  Krebs,  and  C.  M.  Perrins. 
1981.  Song  repertoires  and  lifetime  reproductive 
success  in  the  Great  Tit  (Pams  major).  Am.  Nat. 
188:149-159. 

Morton,  E.  S.  1977.  On  the  occurrence  and  signifi- 
cance of  motivation-structural  rules  in  some  bird 
and  mammal  .sounds.  Am.  Nat.  11  1:855-869. 

Nelson,  D.  A.  and  L.  J.  Croner.  1991.  Song  catego- 
ries and  their  functions  in  the  Field  Sparrow  (Spi- 
zella  piisUla).  Auk  108:42-52. 

Ratcliffe,  L.  and  R.  G.  Weisman.  1985.  Frequency 
shift  in  the  fee  bee  song  of  the  Black-capped 
Chickadee  (Pams  alricapillus).  Condor  87:555— 
556. 

Ratcliffe,  L.  and  R.  G.  Weisman.  1986.  Song  se- 
quence discrimination  in  the  Black-capped  Chick- 
adee (Pams  atricapillus).  J.  Comp.  Psychol.  100: 
361-367. 

Schroeder,  D.  j.  and  R.  H.  Wiley.  1983.  Communi- 
cation with  repertoires  of  song  themes  in  Tufted 
Titmice.  Anim.  Behav.  31:1128-1138. 

Searcy,  W.  A.  1983.  Response  to  multiple  song  types 
in  male  Song  Sparrows  and  Field  Sparrows. 
Anim.  Behav.  31:948-949. 

Smith,  S.  T.  1972.  Communication  and  other  social 
behavior  in  Pams  carolinensis.  Publ.  Nuttall  Or- 
nithol. Club  11:1-125. 

Staicer,  C.  a.  1989.  Characteristics,  use,  and  signifi- 
cance of  two  singing  behaviors  in  Grace’s  Warbler 
(Dendroica  graciae).  Auk  106:49-63. 

Temrin,  H.  1986.  Singing  behaviour  in  relation  to  po- 
lyterritorial polygyny  in  the  Wood  Warbler  (Phyl- 
loscopus  sihilatrix).  Anim.  Behav.  34:146-152. 

Weary,  D.  M.,  R.  E.  Lemon,  and  S.  Perreault.  1994. 
Male  Yellow  Warblers  vary  use  of  song  types  de- 
pending on  pairing  status  and  distance  from  the 
nest.  Auk  1 1 1 :727-729. 

Weisman,  R.  G.  and  L.  Ratcliffe.  1989.  Absolute  and 
relative  pitch  processing  in  Black-capped  Chick- 
adees, Pams  atricapillus.  Anim.  Behav.  38:685- 
692. 

WiEBE,  M.  O.  1995.  The  function  of  song  types  in  the 
Mountain  Chickadee  (Pams  gamheli).  M.Sc.  the- 
sis, Univ.  of  Calgary,  Calgary,  Alberta. 


Wilson  Bull..  111(3).  1999,  pp.  376-380 


SURVIVAL  AND  LONGEVITY  OF  THE  PUERTO  RICAN  VIREO 

BETHANY  L.  WOODWORTH,'  JOHN  FAABORG,^  AND  WAYNE  J.  ARENDU 


ABSTRACT. — The  Puerto  Rican  Vireo  {Vireo  latimeri),  a Puerto  Rican  endemic,  is  declining  in  at  least  one 
forest  reserve  as  the  result  of  pressures  from  introduced  nest  predators  and  an  introduced  brood  parasite.  We 
collected  data  on  adult  survival,  adult  longevity,  and  juvenile  survival  from  a long-term  mist  netting  study 
(1973—1999)  and  a demographic  study  of  color-marked  birds  (1990—1993)  in  Guanica  Forest,  Puerto  Rico.  Of 
the  adult  birds  banded  in  the  first  three  years  of  the  demographic  study,  24  of  32  males  (75%)  and  6 of  7 females 
(86%)  were  known  to  survive  until  June  of  the  year  following  their  banding.  Model-based  estimates  of  adult 
survival  rate  from  capture/resighting  of  65  color-marked  birds  was  0.74  (±  0.05  SE);  for  51  adult  males  analyzed 
separately,  survival  rate  was  0.74  (±0.06;  data  were  insufficient  to  estimate  survival  rate  of  females).  We 
recorded  a new  longevity  record  for  the  Puerto  Rican  Vireo  of  13  years,  2 months.  Juvenile  survival  was 
estimated  by  enumeration  to  be  0.40  (±0.15).  Juveniles  spent  prolonged  periods  on  their  natal  teiTitory,  which 
might  increase  their  probability  of  surviving  to  first  breeding.  Puerto  Rican  Vireos  have  lelatively  high  survival 
rates  despite  the  presence  of  numerous  introduced  predators  in  their  habitat,  a highly  seasonal  environment,  and 
the  stress  of  renesting  as  many  as  6 times  in  a season.  Received  19  Oct.  1998,  accepted  25  Feb.  1999. 


Survival  rate  is  an  important  component  of 
life  history  models,  and  it  is  a central  parameter 
examined  in  comparative  demographic  studies 
(e.g.,  Ricklefs  1982,  Martin  1995).  From  a con- 
servation perspective,  precise  estimation  of 
adult  and  Juvenile  survival  rates  is  critical  be- 
cause population  dynamics  often  show  great 
sensitivity  to  variation  in  these  parameters 
(e.g.,  Lande  1988,  Ryan  et  al.  1993).  Despite 
their  importance,  relatively  few  data  are  avail- 
able regarding  the  survival  rates  and  longevity 
of  tropical  birds  in  general,  and  insular  species 
in  particular  (but  see  Karr  et  al.  1990,  Faaborg 
and  Arendt  1995,  Johnston  et  al.  1997). 

The  Puerto  Rican  Vireo  (Vireo  latimeri),  is 
a small  (11-12  g)  passerine  restricted  to  the 
island  of  Puerto  Rico  (Wetmore  1916).  The 
population  of  vireos  in  Guanica  Forest,  Puerto 
Rico’s  largest  dry  forest  reserve,  has  declined 
steadily  over  the  past  20  years  as  a result  of 
parasitism  by  an  exotic  avian  brood  parasite, 
the  Shiny  Cowbird  (Molothrus  bonarien.sis), 
and  nest  predation  by  introduced  mammals 


' Dept,  of  Ecology,  Evolution  and  Behavior,  Univ. 
of  Minnesota,  1987  Upper  Buford  Circle.  St.  Paul,  MN 
55108. 

- Present  address:  Pacific  Island  Ecosystems  Re- 
search Center,  Biological  Resources  Division, 
U.S.G.S.,  Kilauea  Field  Station,  PO.  Box  44,  Hawaii 
National  Park,  HI  96718. 

•’  Division  of  Biological  Sciences,  Univ.  of  Missouri, 
Columbia,  MO  6521  1. 

■'  USDA  Forest  Service,  International  In.stitute  of 
Tropical  Forestry,  Sabana  Field  Research  Station,  P.O. 
Box  490,  Palmer.  PR  00721. 

^ Corresponding  author; 

E-mail:  Bethany-WoodworthC® usgs.gov 


[rats  (Rattus  spp.),  mongoose  (Herpestes  au- 
ropunctatus),  and  feral  cats  (Felis  catus)-,  Faa- 
borg et  al.  1997;  Woodworth  1997,  1999].  A 
population  dynamics  analysis  indicated  that  es- 
timates of  the  Puerto  Rican  Vireo’s  population 
growth  rate  were  very  sensitive  to  adult  sur- 
vival rate;  consequently  precise  estimates  of 
this  parameter  are  crucial  for  useful  population 
modehng  (Woodworth  1999).  In  this  paper  we 
present  data  on  adult  and  juvenile  recoveries, 
from  which  we  estimate  survival  rates  and  lon- 
gevity of  the  Puerto  Rican  Vireo,  compare 
them  to  temperate  mainland  congeners,  and 
comment  on  their  implications  for  persistence 
of  this  single-island  endemic. 

METHODS 

We  .studied  the  Puerto  Rican  Vireo  population  in 
Guanica  Forest  Reserve  (17°  58'  N,  66°  52'  W)  along 
the  southwestern  coast  of  Puerto  Rico.  The  reserve 
comprises  4,0 1 5 ha  of  mature  dry  subtropical  forest 
over  shallow  limestone  soils.  Rainfall  averages  860 
mm  annually,  almost  all  of  which  tails  between  April 
and  November  (Murphy  and  Lugo  1986),  and  Puerto 
Rican  Vireos  generally  breed  from  April  through  July 
(Woodworth  1997).  Guanica  is  the  site  of  a long-term 
constant  effort  mist  netting  study  of  wintering  and  res- 
ident landbirds  (see  Faaborg  and  Arendt  1989a),  dur- 
ing which  135  Puerto  Rican  Vireos  were  marked  with 
aluminum  bands  from  1973-1996. 

From  1990-1993.  B.L.W.  conducted  a demographic 
study  of  color-marked  vireos  in  four  50-ha  study  areas 
(Woodworth  1997).  Resident  vireos  were  captufed  by 
playing  recorded  vireo  songs  to  lure  territorial  males 
or  pairs  into  mist  nets.  Because  males  were  more  ag- 
gressive than  females  toward  intruders,  most  known 
sex  birds  we  captured  were  males  (78%,  n = 65;  1 
bird  was  of  unknown  sex).  Individuals  were  resighted 
by  revisiting  all  territories  within  the  study  areas  and 


376 


Woodworth  el  at.  • PUERTO  RICAN  VIREO 


377 


areas  within  300  ni  (about  two  territory  widths)  of  their 
borders,  and  by  broadcasting  Puerto  Rican  Vireo  song. 
Color-marked  birds  were  relocated  every  few  days 
(range:  1-9)  throughout  the  breeding  season  as  part  of 
a study  of  their  seasonal  reproductive  success  (Wood- 
worth  1997).  In  order  to  approximate  the  general  as- 
sumption of  capture-recapture  models  that  all  sampling 
is  instantaneous,  "capture  periods”  were  defined  so  as 
to  be  short  in  relation  to  interval  length  (breeding  sea- 
son; Smith  and  Anderson  1987).  Thus  we  defined  two 
sample  (capture)  periods  each  breeding  season  (1990, 
1991,  and  1993),  one  consisting  of  the  first  two  weeks 
after  arrival  on  the  study  area,  and  the  second  includ- 
ing the  two  weeks  immediately  preceding  the  end  of 
the  field  season.  Birds  banded  at  other  times  were  in- 
cluded only  if  they  were  resighted  during  one  of  these 
sampling  periods  and  were  treated  as  if  they  were  orig- 
inally banded  during  that  sampling  period.  The  June 
1992  and  January  1993  capture  periods  consisted  of 
15  and  11  day  visits  to  the  study  area,  respectively. 

We  estimated  adult  survival  rate  from  capture/resight- 
ing data  on  66  color-marked,  territorial  adults  over  4 
years  and  7 capture  intervals  [average  interval  length  = 
0.45  ± 0.31  (SE)  years].  Five  of  the  birds  used  in  this 
analysis  were  originally  banded  by  J.F.  and  W.J.A.  prior 
to  1990,  and  so  were  included  in  a survival  analysis  by 
Faaborg  and  Arendt  (1995),  but  time  periods  of  the  two 
survival  datasets  did  not  overlap.  We  used  the  program 
JOLLY  (Pollock  et  al.  1990)  to  produce  estimates  of 
survival  rate  under  five  different  capture-recapture  mod- 
els which  vary  in  their  assumptions  about  capture  and 
survival  probabilities.  These  models  and  their  assump- 
tions have  been  presented  in  detail  elsewhere  (Pollock 
et  al.  1990  and  references  therein).  In  general,  the  cap- 
ture/resighting field  methods  used  here  and  the  more 
widely  u.sed  constant  effort  mist  netting  methods  meet 
(or  not)  the  assumptions  of  the  Jolly-Seber  models  to 
similar  degrees,  with  a few  exceptions:  (1)  although 
fixed  placement  of  nets  in  relation  to  territory  bound- 
aries may  result  in  heterogeneous  capture  probabilities 
in  mistnetting  .studies  (Pollock  et  al.  1990),  we  were  able 
to  search  entire  territories  for  marked  individuals;  (2) 
trap  response  (net  shyness)  was  not  a concern  in  this 
study  because  we  did  not  need  to  catch  a bird  in  a net 
in  order  to  resight  it;  (3)  we  were  able  to  exclude  tran- 
sients from  the  study  (the  presence  of  transients  in  a 
sample  may  bias  survival  rate  estimates  if  special  mod- 
els are  not  employed;  Pradel  et  al.  1997);  (4)  we  were 
able  to  rule  out  temporary  emigration.  Becau.se  the  cap- 
ture probability  in  this  study  was  very  high  (0.92),  we 
expect  the  model-based  estimators  to  provide  reasonably 
unbiased  estimates  of  survival  rate  despite  the  relatively 
short  time  span  of  the  study  and  moderate  sample  size 
(Gilbert  1973). 

The  program  JOLLY  provides  goodness-of-fit  te.sts 
to  assess  the  fit  of  a model  to  a given  data  set.  Where 
several  models  fit  the  data,  likelihood  ratio  tests  were 
used  to  test  among  models,  with  the  simplest  adequate 
model  preferred.  For  statistical  comparisons  among 
survival  rates  we  used  the  X“  statistic  proposed  by 
Sauer  and  Williams  (1989). 


Model-based  estimators  account  for  the  possibility 
that  a bird  is  alive  and  in  the  study  area,  but  is  not 
resighted  in  a particular  sample  period.  To  facilitate 
comparison  with  other  studies,  we  also  present  surviv- 
al rate  as  the  number  of  birds  banded  in  the  first  three 
years  of  the  study  that  were  known  to  be  alive  in  the 
June  following  their  banding  (i.e.,  enumeration). 

Because  of  the  small  sample  size  of  fledglings,  we 
could  not  use  model-based  estimators  of  juvenile  sur- 
vival. Thus,  juvenile  survival  was  calculated  as  the  pro- 
portion of  birds  originally  banded  as  fledglings  that  were 
recaptured  or  resighted  in  any  subsequent  year,  and  var- 
iance was  estimated  assuming  binomial  sampling. 

As  is  true  in  all  capture-recapture  studies  of  open 
populations,  dispersal  outside  of  the  study  area  could 
not  be  distinguished  from  mortality.  However,  typical 
dispersal  rates  and  distances  for  the  Puerto  Rican  Vireo 
are  small  (Woodworth  et  al.  1998)  so  the  effect  should 
be  relatively  minor  in  this  study. 

During  the  demographic  study,  B.L.W.  recaptured  7 
birds  that  had  been  originally  banded  prior  to  1990. 
Estimated  maximum  longevity  of  the  recaptured  birds 
was  calculated  as  the  time  from  initial  banding  to  the 
last  recapture,  plus  the  time  from  initial  banding  to  the 
previous  June  1,  assuming  that  all  birds  were  hatched 
on  that  date  (following  Klimkiewicz  et  al.  1983). 

RESULTS 

We  color-banded  51  males,  14  females,  13 
fledglings,  and  10  birds  of  unknown  sex  (win- 
ter captures).  Of  the  adult  birds  banded  in  the 
first  three  years  of  the  study,  24  of  32  males 
(75%)  and  6 of  7 females  (86%)  were  known 
to  survive  (i.e.,  were  alive  and  present  on  the 
study  area)  until  the  June  of  the  year  follow- 
ing their  banding  (Table  1).  Over  the  four 
years,  there  were  59  opportunities  for  males 
to  survive  between  breeding  seasons,  and  the 
males  survived  in  at  least  43  of  these  cases 
(73%).  Females  were  documented  to  survive 
in  11  of  13  opportunities  (85%). 

Territorial  adult  Puerto  Rican  Vireos  during 
this  study  had  an  estimated  annual  survival  rate 
of  0.74  (±  0.05).  JOLLY  model  D,  which  is 
based  on  constant  survival  and  capture  proba- 
bilites  throughout  the  study,  provided  the  best 
fit  to  the  data  (overall  ~ 9.7,  df  = 8,  P = 
0.29).  Our  capture/resighting  methodology  re- 
sulted in  a very  high  annual  capture  probability 
(0.92  ± 0.03).  Male  annual  survival  rate  esti- 
mated for  51  males  was  0.74  (±  0.06;  Model 
D,  capture  probability  = 0.95  ± 0.02).  Data 
for  14  females  were  insufficient  to  fit  a model 
describing  female  survival  because  most  were 
banded  in  the  last  year  of  the  study.  Juvenile 


378 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


TABLE  1.  Bandings  and  resightings  of  male  Puerto  Rican  Vireos  known  to  be  at  least  one  year  old  when 
banded  (demographic  study  only).  The  8 sample  periods,  each  2 weeks  long,  took  place  in  April  1990,  July 
1990,  March  1991,  August  1991,  June  1992,  January  1993,  March  1993,  and  August  1993,  resulting  in  7 
intervals.  No  new'  birds  were  banded  during  sample  periods  in  July  1990,  June  1992,  or  August  1993. 


Sample  period 
when  banded 

Returns  in 

sample  period  following  banding 

Number 

banded 

One 

Two 

Three 

Four 

Five 

Six 

Seven 

Apr  1990 

10 

8 

7 

7 

4 

2 

1 

1 

Mar  1991 

7 

5 

6 

5 

6 

6 

— 

— 

Aug  1991 

13 

10 

1 1 

8 

6 

— 

— 

— 

Jan  1993 

3 

3 

2 

— 

— 

— 

— 

— 

Mar  1993 

18 

17 

— 

— 

— 

— 

— 

— 

Total 

51 

43 

26 

20 

16 

8 

1 

1 

survival  rate  from  fledging  to  first  breeding 
was  estimated  at  0.40  (±  0.15;  n = 10). 

The  oldest  Puerto  Rican  Vireo  we  recap- 
tured was  at  least  13  years,  2 months  (13-02) 
old,  exceeding  the  previous  longevity  record 
for  the  Puerto  Rican  Vireo  by  nearly  4 years 
(09-04;  Faaborg  and  Arendt  1989b).  Three 
other  birds  that  nearly  matched  the  previous 
record  were  also  recaptured  (09-01,  09-02, 
and  09-02).  All  were  color-banded  territorial 
males  that  we  observed  over  1-2  complete 
breeding  seasons  before  they  disappeared  or 
the  study  ended.  None  had  dispersed  more 
than  500  m in  the  decade  since  they  were  orig- 
inally banded  (see  Woodworth  et  al.  1998  for 
details  of  dispersal  behavior). 

DISCUSSION 

Faaborg  and  Arendt  (1995)  estimated  an 
adult  annual  survival  rate  of  0.68  (±  0.08)  for 
the  Puerto  Rican  Vireo  population  in  Guanica, 
based  on  their  long-term  mist  netting  study 
(Jolly-Seber  model  D,  19  individuals  over  18 
years  and  15  capture  intervals).  Although  their 
mean  survival  value  is  slightly  less  than  that 
presented  here  (probably  because  of  the  inclu- 
sion of  a higher  proportion  of  female  birds, 
along  with  non-territorial  individuals),  the  two 
estimates  are  not  significantly  different  (y^  = 
0.40,  df  = 1,  P > 0.05). 

Our  Puerto  Rican  Vireo  survival  rate  esti- 
mate is  high  relative  to  survival  and  recovery 
rates  reported  for  temperate  vireos.  Recovery 
percentages  of  White-eyed  Vireos  {Vireo  gri- 
.seus)  studied  over  9 years  on  their  breeding 
grounds  were  48%  for  males,  and  50%  for 
females  (Hopp  et  al.  1999).  Return  rates  of 
adult  male  Black-capped  Vireos  {V.  atricap- 
illiis)  to  breeding  territories  in  central  Texas 


were  55-75%  (Grzybowski  1991).  The  sur- 
vival rate  of  Red-eyed  Vireos  {V.  olivaceous) 
based  on  returns  to  breeding  grounds  in  Mar- 
yland, was  estimated  at  59%  (Jolly-Seber 
model  A;  Nichols  et  al.  1981).  Return  rates  of 
adult  Gray  Vireos  (V.  vicinior)  to  wintering 
territories  in  Mexico  were  46-71%  (Bates 
1992),  and  for  Bell’s  Vireo  {V.  bellii)  return- 
ing to  breeding  territories  in  California,  47% 
(Salata  1983).  Interestingly,  compared  to  sur- 
vival rate  estimates  for  other  tropical  island 
passerines  studied  to  date,  the  survival  rate  of 
the  Puerto  Rican  Vireo  is  not  unusually  high 
[e.g.,  average  68%  (51—79%)  for  7 Puerto  Ri- 
can species,  Faaborg  and  Arendt  1995;  65.3% 
(45-85%)  for  17  Trinidadian  species,  John- 
ston et  al.  1997;  and  76%  (55-88%)  for  5 Ha- 
waiian species,  van  Riper  1987,  Lepson  and 
Freed  1995,  Ralph  and  Fancy  1995,  Wood- 
worth  et  al.  in  press]. 

Likewise,  the  longevity  record  for  the 
Puerto  Rican  Vireo  of  13  years,  2 months  is 
long  relative  to  most  of  its  temperate  conge- 
ners. Records  for  six  other  temperate  Vireo 
species  range  from  6 years,  1 month  to  10 
years  (Davis  1995,  Kennard  1975,  Klimkiew- 
icz  et  al.  1983,  Rodewald  and  James  1996). 
The  record  for  the  Warbling  Vireo  (Vireo  gil- 
vis),  a neotropical  migrant,  is  very  similar  (13- 
01,  Klimkiewicz  et  al.  1983)  to  the  Puerto  Ri- 
can Vireo.  Such  records  are  complicated  by 
many  factors  (Krementz  et  al.  1989).  Al- 
though it  is  surprising  that  such  a seemingly 
long  longevity  record  from  a nonmigratory 
species  would  be  equaled  by  a long  distance 
migrant,  it  is  worth  noting  that  over  15,000 
Warbling  Vireos  have  been  banded  (Kli- 
mkiewicz et  al.  1983),  but  only  a few  hundred 


Woodworth  et  cil.  • PUERTO  RICAN  VIREO 


379 


Puerto  Rican  Vireos  have  been  bandeci  and 
many  of  these  have  been  long  lived. 

The  relatively  high  survival  and  longevity 
of  this  insular  species  is  remarkable  in  light 
of  the  presence  of  numerous  introduced  pred- 
ators in  its  habitat,  high  rates  of  nest  failure 
causing  females  to  renest  up  to  6 times  in  a 
season  (Woodworth  1997),  and  the  stresses  of 
a highly  seasonal  environment  (almost  no  rain 
falls  from  December  to  March,  and  Guanica 
loses  up  to  50%  of  its  leaf  area  in  winter; 
Murphy  and  Lugo  1986).  Puerto  Rican  Vireos 
have  a small  clutch  size  relative  to  temperate 
vireos  (Woodworth  1995),  which,  when  cou- 
pled with  the  generally  observed  trade-off  be- 
tween fecundity  and  survival  (Martin  1995), 
might  allow  birds  to  survive  through  more 
breeding  seasons  (Cody  1966).  In  addition,  a 
non-migratory  insular  species  might  outlive 
its  migratory  counterparts  because  it  does  not 
pay  the  price  of  annual  migration. 

Juvenile  survival  rates  of  passerines  are 
poorly  known,  especially  for  tropical  birds. 
Return  rates  of  juvenile  Bell’s  Vireos  and 
Black-capped  Vireos  were  measured  as  24%, 
although  actual  survival  rate  is  likely  to  be 
higher  (Salata  1983,  Grzybowski  1991;  re- 
spectively). Survival  of  juvenile  Wood  Tlirush 
{Hylocichla  mustelina)  in  their  first  12  weeks 
is  only  0.42  (Anders  et  al.  1997).  Juvenile  sur- 
vival may  be  enhanced  if  young  are  allowed 
to  remain  in  their  natal  territory  for  an  ex- 
tended period  (discussed  in  Karr  et  al.  1990) 
as  Puerto  Rican  Vireo  fledglings  have  been 
observed  to  do  (at  least  80-98  days  post- 
fledging;  Woodworth  1995). 

A population  dynamics  model  of  this  pop- 
ulation showed  that,  as  is  common  in  many 
population  models,  the  vireo ’s  predicted  pop- 
ulation growth  rate  was  greatly  dependent  upon 
the  value  of  adult  survivorship  used  in  the 
model  (Woodworth  1999).  Therefore,  a precise 
and  accurate  estimate  of  adult  survivorship  is 
critical  to  evaluating  the  long-term  prospects 
for  survival  of  this  population.  The  close  agree- 
ment between  two  independent  estimates  (Faa- 
borg  and  Arendt  1995  and  this  study)  of  adult 
survival  rate  for  this  population  improves  con- 
fidence in  the  predictions  of  a model  using 
these  estimates,  although  additional  data  on  fe- 
male and  juvenile  survivorship  is  needed. 

The  relatively  high  adult  and  juvenile  sur- 
vival rates  we  documented  would  seem  to  bode 


well  for  the  persistence  of  Puerto  Rican  Vireos 
in  Guanica  Forest.  However,  other  work  on  this 
population  has  shown  that  the  vireos  suffer  ex- 
tremely high  nest  losses  to  native  and  intro- 
duced predators,  and  to  parasitism  by  the  ex- 
otic Shiny  Cowbird  (Woodworth  1997).  De- 
spite as  many  as  6 nest  attempts  in  a single 
season,  females  succeed  in  fledging  young 
from  only  0.41-0.67  nests  per  year  (Wood- 
worth  1997).  These  factors  result  in  an  overall 
negative  population  growth  rate  for  the  vireo 
over  the  range  of  “reasonable”  survival  rate 
values  (the  95%  confidence  limits  of  the  esti- 
mates; Woodworth  1999).  Thus,  the  declines 
observed  over  the  previous  decade  (Faaborg  et 
al.  1997)  are  likely  to  continue  unless  active 
management  is  undertaken  to  reduce  predation 
and/or  brood  parasitism  in  the  forest. 

ACKNOWLEDGMENTS 

J.  Colon  kindly  provided  information  on  two  birds 
he  originally  banded.  We  thank  the  dozens  of  volun- 
teers who  have  participated  in  these  projects  over  the 
years.  The  Puerto  Rico  Departamento  de  Recursos  Na- 
turales  Ambiente  and  M.  Canals  kindly  gave  permis- 
sion to  work  in  Guanica.  B.L.W.  was  supported  by  the 
International  Council  for  Bird  Preservation-U.S.  Sec- 
tion; Frank  M.  Chapman  Fund  of  the  American  Mu- 
seum of  Natural  History;  Sigma  Xi  Grant-in-Aid  of 
Research;  Dayton  Natural  History  Fund  and  Wilkie 
Fund  for  Natural  History  Research,  Bell  Museum  of 
Natural  History;  Eastern  Bird  Banding  Association; 
Paul  A.  Stewart  Award  of  the  Wilson  Ornithological 
Society;  fellowships  and  assistantships  from  the  De- 
partment of  Ecology,  Evolution  and  Behavior,  Univer- 
sity of  Minnesota;  and  Grants  for  Research  Abroad  and 
a Doctoral  Dissertation  Fellowship  from  the  Graduate 
School  of  the  University  of  Minnesota.  Support  for  J.F 
and  W.J.A.  was  provided  by  the  Frank  M.  Chapman 
Fund  of  the  American  Museum  of  Natural  Hi.story,  Na- 
tional Sciences  Foundation  Doctoral  Dissertation  Im- 
provement Fund,  University  of  Missouri-Columbia 
(Research  Council  of  the  Graduate  School),  USD  A 
Forest  Service  (International  Institute  of  Tropical  For- 
estry), Biological  Resources  Division  of  the  U.S.  Geo- 
logical Survey,  and  the  U.S.  Fish  and  Wildlife  Service. 
K.  Dugger,  S.  Hopp,  J.  Bates,  and  two  anonymous  re- 
viewers provided  helpful  comments  on  earlier  drafts  of 
this  manuscript. 

LITERATURE  CITED 

Anders,  A.  D.,  D.  C.  Dearborn,  J.  Faaborg,  and  F. 
R.  Thompson,  III.  1997.  Juvenile  survival  in  a 
population  of  Neotropical  migrant  birds.  Conserv. 
Biol.  11:698-707. 

Bates,  J.  M.  1992.  Winter  tenitorial  behavior  of  Gray 
Vireos.  Wilson  Bull.  104:425-433. 


380 


THE  WILSON  BULLETIN  • VoL  III.  No.  3.  September  1999 


Cody,  M.  L.  1966.  A general  theory  of  clutch-size. 
Evolution  20:174-184. 

Davis,  J.  N.  1995.  Hutton's  Vireo  (Vireo  hiittoni).  In 
The  birds  of  North  America,  no.  189  (A.  Poole 
and  L Gill,  Eds.).  The  Academy  of  Natural  Sci- 
ences, Philadelphia,  Pennsylvania;  The  American 
Ornithologists'  Union,  Washington,  D.C. 

Faaborg,  J.  and  W.  J.  Arendt.  1989a.  Long-term  de- 
clines in  winter  resident  warblers  in  a Puerto  Ri- 
can dry  forest.  Am.  Birds  43:1226-1230. 

Faaborg,  J.  and  W.  J.  Arendt.  1989b.  Longevity  es- 
timates of  Puerto  Rican  birds.  N.  Am.  Bird  Bander 
14:1  1-13. 

Faaborg,  J.  and  W.  J.  Arendt.  1995.  Survival  rates 
of  Puerto  Rican  birds:  are  islands  really  that  dif- 
ferent? Auk  1 12:503-507. 

Faaborg,  J.,  K.  M.  Dugger,  W.  J.  Arendt,  B.  L. 
Woodworth,  and  M.  E.  Baltz.  1997.  Population 
declines  of  the  Puerto  Rican  Vireo  {Vireo  lalimeri) 
in  Guanica  Forest.  Wilson  Bull.  109:195-202. 

Gilbert,  R.  O.  1973.  Approximations  of  the  bias  in 
the  Jolly-Seber  capture-recapture  model.  Biomet- 
rics 29:501-526. 

Grzybowski,  j.  a.  1991.  Survivorship,  dispersal  and 
population  structure  of  Black-capped  Vireos  at  the 
Kerr  Wildlife  Management  Area,  Texas.  Resour. 
Protection  Div.,  Texas  Parks  Wildl.  Dept.,  Austin. 

Hope,  S.  L.,  A.  Kirby,  and  C.  A.  Boone.  1999.  Band- 
ing returns,  arrival  pattern,  and  site-hdelity  of 
White-eyed  Vireos.  Wilson  Bull.  111:46—55. 

Johnston,  J.  P,  W.  J.  Peach,  R.  D.  Gregory,  and  S. 
A.  White.  1997.  Survival  rates  of  tropical  and 
temperate  passerines:  a Trinidadian  perspective. 
Am.  Nat.  150:771-789. 

Karr,  J.  R.,  J.  D.  Nichols,  M.  K.  Klimkiewicz,  and 
j.  D.  Brawn.  1990.  Survival  rates  of  birds  of  trop- 
ical and  temperate  forests:  will  the  dogma  sur- 
vive? Am.  Nat.  136:277—291. 

Kennard,  j.  H.  1975.  Longevity  records  of  North 
American  birds.  Bird-Banding  46:55—57. 

Klimkiewicz,  M.  K.,  R.  B.  Clapp,  and  A.  G.  Futcher. 
1983.  Longevity  records  of  North  American  birds: 
Remizidae  through  Parulinae.  J.  Field  Ornithol. 
54:287-294. 

Krementz,  D.  G.,  j.  R.  Sauer,  and  J.  D.  Nichols. 
1989.  Model-based  estimates  of  annual  survival 
rate  are  preferable  to  observed  maximum  lifespan 
statistics  for  use  in  comparative  life-history  stud- 
ies. Oikos  56:20.3-208. 

Lande,  R.  1988.  Demographic  models  of  the  NoHhern 
Spotted  Owl  (Sfri.x  occidenlali.s  caiirina).  Oecol- 
ogia  (Berlin)  75:601-607. 

Lep.son,  j.  K.  and  L.  a.  Freed.  1995.  Variation  in  male 
plumage  and  behavior  ot  the  Hawaii  Akepa.  Auk 
1 12:402-414. 

Martin,  T.  E.  1995.  Avian  life  history  evolution  in 
relation  to  nest  sites,  nest  predation,  and  food. 
Ecol.  Monogr.  65:101-127. 

Murphy,  P.  G.  and  A.  E.  Liigo.  1986.  .Structure  and 
biomass  of  a subtropical  dry  fore.st  in  Puerto  Rico. 
Biotropica  18:89-96. 


Nichols,  J.  D.,  B.  R.  Noon,  S.  L.  Stokes,  and  J.  E. 
Hines.  1981.  Remarks  on  the  ii.se  of  mark-recap- 
ture methodology  in  estimating  avian  population 
size.  Stud.  Avian  Biol.  6:121-136. 

Pollock,  K.  H.,  J.  D.  Nichols,  C.  Brownie,  and  J.  E. 
Hines.  1990.  Statistical  inference  for  capture-re- 
capture experiments.  Wildl.  Monogr.  107:1—97. 

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. 

Ralph,  C.  J.  and  S.  G.  Fancy.  1995.  Demography  and 
movements  of  'Apapane  and  'I'iwi  in  Hawai'i. 
Condor  97:729-742. 

Ricklefs,  R.  E.  1982.  Comparative  avian  demography. 
Cliit.  Ornithol.  1:1-32. 

Rodewald,  P.  G.  and  R.  D.  James.  1996.  Yellow- 
throated  Vireo  ( Vireo  jiavifrons).  In  The  birds  of 
North  America,  no.  247  (A.  Poole  and  F.  Gill, 
Eds.).  The  Academy  of  Natural  Sciences,  Phila- 
delphia. Pennsylvania;  The  American  Ornitholo- 
gists' Union,  Washington,  D.C. 

Ryan,  M.  R.,  B.  G.  Root,  and  P.  M.  Mayer.  1993. 
Status  of  Piping  Plovers  in  the  Great  Plains  ot 
North  America:  a demographic  simulation  model. 
Conserv.  Biol.  7:581—585. 

Salata,  L.  R.  1983.  Status  of  the  Least  Bell's  Vireo 
on  Camp  Pendleton,  California:  research  done  in 
1983.  U.S.  Fish  and  Wildl.  Serv.,  Laguna  Niguel, 
California. 

Sauer,  J.  R.  and  B.  K.  Williams.  1989.  Generalized 
procedures  for  testing  hypotheses  about  survival 
or  recovery  rates.  J.  Wildl.  Manage.  53:137—142. 

Smith,  D.  R.  and  D.  R.  Anderson.  1987.  Effects  of 
lengthy  ringing  periods  on  estimates  ot  annual 
survival.  Acta  Ornithol.  23:69—76. 

van  Riper,  C..  III.  1987.  Breeding  ecology  of  the  Ha- 
waii Common  Amakihi.  Condor  89:85—102. 

Wetmore,  a.  1916.  Birds  of  Porto  Rico.  U.S.  Dept. 
Agric.  Bull.  326:1-140. 

Woodworth,  B.  L.  1995.  Ecology  of  the  Puerto  Rican 
Vireo  and  the  Shiny  Cowbird  in  Guanica  Forest, 
Puerto  Rico.  Ph.D.  diss.,  Univ.  of  Minnesota,  St. 
Paul. 

Woodworth,  B.  L.  1997.  Brood  parasitism,  nest  pre- 
dation. and  season-long  reproductive  success  of  a 
tropical  island  endemic.  Condor  99:605-612. 

Woodworth,  B.  L.  1999.  Modeling  the  population  dy- 
namics of  a songbird  exposed  to  parasitism  and 
predation  and  evaluating  management  options. 
Conserv.  Biol.  13:67—76. 

Woodworth,  B.  L.,  J.  Faaborg,  and  W.  J.  Arendt. 
1998.  Breeding  and  natal  dispersal  in  the  Puerto 
Rican  Vireo.  J.  Field  Ornithol.  69:1-7. 

Woodworth.  B.  L.,  J.  T.  Nel.son,  E.  J.  TVeed,  S.  G. 
Fancy,  M.  P.  Moore,  E.  B.  Cohen,  and  M.  S. 
Collins.  In  press.  Breeding  productivity  and  sur- 
vival of  the  endangered  Hawai'i  Creeper  in  a wet 
forest  refuge  on  Manna  Kea.  Hawai'i.  In  The  sta- 
tus, ecology,  and  conservation  of  the  Hawaiian 
avifauna  (.1.  M.  Scott,  S.  Conant,  and  L.  A.  Freed. 
Eds.).  Stud.  Avian  Biol. 


Wilson  Bull..  111(3),  1999,  pp.  381-388 


EFFECTS  OF  PRIOR  RESIDENCE  AND  AGE  ON  BREEDING 
PERFORMANCE  IN  YELLOW  WARBLERS 

G.  A.  LOZANO'  - ' AND  R.  E.  LEMON' 


ABSTRACT. — Age-related  increases  in  reproductive  success  could  be  the  result  of  better  survival  by  suc- 
cessful breeders  (survival  hypothesis),  greater  dispersal  by  unsuccessful  breeders  (dispersal  hypothesis),  and/or 
age-related  differences  in  the  ability  to  compete  for  breeding  opportunities  (constraint  hypothesis).  We  used 
banding  and  nesting  data  horn  four  consecutive  breeding  seasons  to  examine  the  effects  of  prior  residency  on 
several  indices  of  breeding  performance  in  Yellow  Warblers  {Dendroica  petechia).  We  compared  the  breeding 
performance  of  returning  birds  with  that  of  new  arrivals,  and  of  individuals  between  successive  breeding  seasons. 
There  were  no  differences  in  clutch  size  between  new  anivals  and  returning  individuals,  nor  within  individuals 
between  successive  breeding  seasons.  Among  males,  prior  residence  had  no  effect  on  whether  a clutch  was 
started,  but  among  females  the  number  of  prior  residents  that  initiated  a clutch  was  higher  than  expected,  and 
the  number  of  new  arrivals  that  did  not  was  lower  than  expected.  In  contrast,  there  were  no  differences  in  laying 
or  hatching  date  between  new  arrivals  and  returning  individuals,  but  within-individual  comparisons  showed  that 
males  bred  earlier  in  successive  breeding  seasons.  Previous  reproduction  increased  subsequent  return  rates  only 
1 out  of  3 years  in  both  sexes.  Returning  males  were  larger  than  new  arrivals,  but  there  were  no  differences  in 
females.  Within-individual  size  increases  between  successive  breeding  seasons  occurred  in  both  sexes.  These 
results  are  consistent  with  the  constraint  hypothesis,  but  the  proximate  mechanisms  by  which  these  differences 
arise  remain  to  be  determined.  Received  1 Oct.  1998,  accepted  20  Feb.  1999. 


Age-related  increases  in  reproductive  suc- 
cess have  been  documented  in  many  bird  spe- 
cies (reviewed  by  Saether  1990,  Forslund  and 
Part  1995).  Several  hypotheses  have  been  pro- 
posed to  explain  this  phenomenon.  Forslund 
and  Part  (1995)  divided  these  hypotheses  into 
three  groups  depending  on  whether  they  are 
based  on  (1)  the  gradual  appearance  or  dis- 
appearance of  certain  phenotypes,  (2)  the  life- 
time optimization  of  reproductive  effort,  or  (3) 
age-related  improvements  in  competence. 
These  three  groups  of  hypotheses  are  not  mu- 
tually exclusive. 

In  the  first  group  of  hypotheses,  differences 
in  reproductive  success  among  age  classes  are 
not  viewed  as  the  result  of  individuals  increas- 
ing their  reproductive  success  with  age,  but 
rather  as  a result  of  the  elimination  of  some 
phenotypes  from  the  population.  The  proba- 
bility of  breeding  may  be  positively  correlated 
with  the  likelihood  of  survival  (survival  hy- 
pothesis) or  negatively  correlated  with  the 
likelihood  of  dispersal  (dispersal  hypothesis; 
e.g..  Smith  1981,  Nol  and  Smith  1987,  Wheel- 
wright and  Schultz  1994);  either  mechanism 


' Dept,  of  Biology,  McGill  Univ.,  120.3  Dr.  Pentield 
Ave.,  Montreal,  Quebec,  Canada  H3A  IBI. 

^ Present  address:  Dept,  of  Biological  Sciences,  Si- 
mon Fraser  Univ.,  Burnaby,  British  Columbia,  Canada, 
V5A  IS6;  E-mail:  LOZANO (®SFU.CA 
^ Corresponding  author. 


would  yield  a positive  correlation  of  age  with 
reproductive  success.  These  hypotheses  are 
clearly  not  applicable  to  species  that  show 
within-individual  increases  in  reproductive 
success  between  successive  reproductive 
bouts  (Pyle  et  al.  1991,  Smith  1993). 

The  second  group  of  hypotheses  asserts  that 
older  individuals  allocate  more  effort  to  re- 
production. The  restraint  hypothesis  states  that 
this  is  because  residual  reproductive  value  de- 
creases with  age,  so  as  individuals  become 
older  the  value  of  current  reproduction  in- 
creases relative  to  the  value  of  future  repro- 
duction. This  idea  is  based  on  the  theoretical 
trade-off  between  current  and  future  reproduc- 
tion (Williams  1966,  Pianka  and  Parker  1975) 
and  only  applies  if  the  probability  of  survival 
decreases  with  age.  Reproductive  effort  may 
also  increase  with  age  if  age-specific  improve- 
ments in  breeding  experience  make  each  suc- 
cessive reproductive  bout  relatively  less  ex- 
pensive (Reid  1988).  Accordingly,  individuals 
may  allocate  the  same  relative  effort  into  re- 
production as  they  age,  but  increase  it  in  ab- 
solute terms.  These  hypotheses  are  based  on 
life  history,  in  which  age  is  the  explanation. 

In  contrast,  hypotheses  in  the  third  group, 
collectively  known  as  the  constraint  hypoth- 
esis (Curio  1983),  are  based  on  ecological 
considerations.  These  hypotheses  posit  that 
younger  individuals  are  unable  to  reproduce 


381 


382 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


as  well  as  older  ones  because  the  former  are 
less  capable  of  competing  for  breeding  op- 
portunities. Age-related  variance  in  reproduc- 
tive success  is  considered  to  be  a result  of 
differences  in  the  ability  to  obtain  territories, 
forage  efficiently,  resist  competitors,  avoid 
predators,  attract  prospective  mates,  and/or 
raise  offspring.  These  factors  may,  in  turn,  be 
affected  by  the  familiarity  of  individuals  with 
the  breeding  area.  This  effect,  known  as  “lo- 
cal experience”  or  “local  familiarity”,  may 
allow  prior  residents  to  exploit  the  resources 
of  an  area  more  efficiently  than  new  arrivals 
(Hinde  1956,  Greenwood  1980). 

Here  we  used  banding  and  nesting  data 
from  four  consecutive  breeding  seasons  to  ex- 
amine the  effects  of  prior  residency  on  several 
indices  of  breeding  performance  in  Yellow 
Warblers  (Dendroica  petechia).  The  Yellow 
Warbler  is  a socially  monogamous  10  g pas- 
serine found  from  Alaska  and  northern  Can- 
ada to  the  coasts  of  Peru  and  Venezuela,  in- 
cluding the  Caribbean  and  Galapagos  islands. 
Formerly,  several  species,  or  subspecies  were 
recognized  (Aldrich  1942,  Bent  1963),  but 
currently  these  are  considered  “groups”  of 
one  species  (Sibley  and  Monroe  1990).  South- 
ern, non-migratory  subspecies  are  easily  dis- 
tinguished from  North  American  migratory 
populations  by  morphological  differences  (Al- 
drich 1942,  Wiedenfeld  1991,  Curson  et  al. 
1994)  and  by  phylogenetic  analyses  based  on 
mitochondrial  DNA  (Klein  and  Brown  1994). 
The  migratory  group  {Dendroica  petechia 
aestiva)  used  in  this  study,  breeds  in  the  Unit- 
ed States  and  Canada  and  winters  in  Central 
America,  northern  South  America,  and  the 
Caribbean. 

Using  morphological  and  nesting  data  from 
four  consecutive  breeding  seasons,  we  first  de- 
termined whether  prior  residence  increased 
the  likelihood  or  timing  of  breeding  in  Yellow 
Warblers.  We  then  tested  the  survival  and  dis- 
persal hypotheses  by  examining  whether  in- 
dividuals were  more  likely  to  breed  or  to  do 
so  earlier  in  successive  seasons,  and  we  dealt 
with  the  possible  disappearance  of  phenotypes 
from  the  population  by  relating  morphology 
and  breeding  performance  to  subsequent  re- 
turn rates.  Finally,  we  explored  morphological 
differences  between  new  arrivals  and  return- 
ing birds,  and  individual  morphological 
changes  between  successive  breeding  seasons. 


METHODS 

Data  were  collected  from  1992  to  1995  at  a 5 ha 
(approximately  500  m X 100  m)  area  between  Pointe 
a Lourneau  and  Pointe  dii  Moulin  on  lie  Perrot,  Que- 
bec, Canada  (45°  22'  N,  73°51'  W).  Yellow  Warblers 
began  arriving  to  our  study  area  during  the  second 
week  of  May  in  all  four  years.  Typically  one  or  two 
males  arrived  first;  two  or  three  days  later  large  num- 
bers of  males  arrived,  along  with  the  first  females. 
Nests  are  built  exclusively  by  females,  and  nest-build- 
ing  takes  about  four  days.  Eggs  are  usually  laid  on 
consecutive  days,  at  a rate  of  one  per  day.  Clutches 
normally  contain  four  or  five  eggs.  Incubation  begins 
when  the  last  egg  is  laid,  and  the  first  egg  hatches  10 
days  later.  During  the  nestling  period  only  females 
brood,  but  both  parents  feed  their  nestlings.  Pledging 
normally  occurrs  approximately  10  days  after  hatch- 
ing. 

Adult  birds  were  captured  using  mist  nets;  each  bird 
captured  was  banded  with  a U.S.  Fish  and  Wildlife 
Service  (USFWS)  aluminum  band  and  three  colored 
plastic  bands.  Mist  nets  were  operated  throughout  the 
study  area  on  a daily  basis  from  May  to  early  June, 
starting  at  dawn  and  ending  at  11:00-13:00,  weather- 
permitting.  Three  to  five  nets  were  used  simultaneous- 
ly in  close  proximity  to  one  another,  and  these  were 
moved  such  that  the  entire  study  area  was  sampled 
every  7-10  days.  After  3 or  4 such  cycles,  we  focused 
our  effort  for  1-2  weeks  on  unbanded  birds  whose 
nests  had  already  been  found.  Nestlings  were  banded 
at  the  nest  when  they  were  six  days  old,  but  only  with 
a USFWS  band.  In  any  given  year,  each  adult  bird 
captured  was  classified  as  a returning  bird  if  it  had 
been  banded  as  an  adult  in  previous  years;  otherwise 
it  was  classified  as  a new  arrival.  It  is  impossible  to 
know  with  certainty  whether  every  bird  was  captured 
every  year,  so  a few  returning  birds  may  have  been 
misclassified  as  being  new  to  the  site;  all  tests  must 
therefore  be  considered  conservative. 

Several  standard  morphological  measurements  were 
taken  from  each  bird:  the  lengths  of  the  flattened  wing 
chord,  the  ninth  primary  feather,  the  outermost  rectrix, 
and  the  tarsus.  The  weight  of  each  bird  was  recorded 
to  the  nearest  0.1  g using  a calibrated  Pesola®  spring 
balance. 

From  early  May  to  the  middle  of  June  the  area  was 
thoroughly  searched  daily  for  nests,  concentrating  in 
areas  where  birds  had  been  banded  and  using  behav- 
ioral cues  from  females.  Except  during  days  of  contin- 
uous rain,  every  nest  was  visited  daily  to  document  the 
chronology  of  nest  building  and  egg  laying.  Many 
nests  failed  as  a result  of  harsh  weather  or  predators; 
renesting  attempts  usually  followed.  Two-  categorical 
measures  were  used  to  determine  the  breeding  stage 
reached:  whether  at  least  one  egg  was  laid  at  a given 
nest  (clutch  initiation)  and  whether  at  least  one  egg 
hatched.  Clutch  initiation  and  hatching  dates  were  used 
to  indicate  the  timing  of  reproduction;  replacement 
nests  were  not  used  in  these  comparisons.  Laying  date 
was  defined  as  the  date  the  first  egg  was  laid,  and 


Lozano  and  Lemon  • BREEDING  SUCCESS  IN  YELLOW  WARBLERS 


383 


TABLE 

were  new 

test.s;  male 

1.  Frequency  of  Yellow  Warblers  that  reached  the  "first  egg”  stage  depending  on 
to  the  area  or  returning  birds,  by  year  and  sex  (expected  frequencies  in  parentheses), 
s:  X-  = 0.006,  df  = 2,  P > 0.05;  females:  X'  = 0.706,  df  = 2,  P > 0.05. 

whether  they 
Heterogeneity 

Did  not  reach  the 

■•first  egg"  stage 

Reached  the  "first 

egg"  stage 

n 

New 

Return 

New 

Return 

*) 

X' 

e 

Males 

1993 

47 

13  (13.3) 

13  (12.7) 

1 1 (10.7) 

10  (10.3) 

0.02 

»0.()5 

1994 

71 

30  (31.1) 

15  ( 13.9) 

19  (17.9) 

7 (8.1) 

0.09 

»0.05 

1995 

53 

19  (19.3) 

12  (11.7) 

14  (13.7) 

8 (8.3) 

0.01 

»0.05 

Total 

171 

62  (63.2) 

40  (38.8) 

44  (42.8) 

25  (26.4) 

0.06 

»0.05 

Females 

1993 

46 

21  (16) 

2 (7) 

1 1 (16) 

12  (7) 

8.32 

0.004 

1994 

66 

23  (21.1) 

8 (9.9) 

22  (23.9) 

13  (11.1) 

0.52 

»0.05 

1995 

45 

15  ( 1 1.4) 

4 (7.6) 

12  (15.6) 

14  (10.4) 

3.65 

0.056 

Total 

157 

59  (48.3) 

14  (24.7) 

45  (55.6) 

39  (28.3) 

1 1.8 

0.001 

similarly,  hatching  date  as  the  date  the  first  egg 
hatched.  As  used  here,  the  terms  “clutch  initiation”, 
"first-egg  stage",  “laying  date”,  and  “hatching  date” 
refer  to  features  of  the  nest,  and  therefore  apply  to  both 
the  male  and  the  female  associated  with  the  nest.  In 
several  other  species  clutch  size  often  decreases  as  the 
breeding  sea.son  progresses  (Erikstad  et  al.  1985,  Mur- 
phy 1986,  Perrins  and  McCleery  1989),  so  only  initial 
nesting  attempts  were  used  in  comparisons  of  clutch 
size.  Other  estimates  of  breeding  performance  were  not 
used  in  these  analyses  because  intrusive  experiments, 
beginning  at  the  time  of  hatching  but  random  with  re- 
spect to  the  variables  examined  here,  were  carried  out 
every  year  (Lozano  and  Lemon  1995,  1996,  1998). 

Although  the  date  on  which  birds  began  to  arrive 
during  the  four  years  of  the  study  varied  by  up  to  two 
weeks,  the  first  day  on  which  an  egg  was  laid  in  the 
population  was  fairly  consistent.  Erom  1992  to  1995, 
these  first  egg  dates  were  May  26  and  24,  June  2,  and 
May  25  respectively.  These  yearly  first  egg  dates  were 
used  to  account  for  variation  in  laying  date  among 
years;  relative  laying  and  hatching  dates  of  individual 
nests  are  defined  as  the  number  of  days  after  the  re- 
spective yearly  first  egg  dates,  added  to  the  mean  first 
egg  date  from  all  four  years. 

Relationships  between  categorical  variables  were 
examined  using  Fisher’s  exact  tests  for  n < 20,  tests 
for  40  > n > 20,  and  x“  with  Yates’  adjustment  for 
continuity  for  n > 40  (Cochran  1954,  Fienberg  1980, 
Everitt  1992).  Because  the  mobility  of  the  two  sexes 
may  have  differed,  e.specially  early  in  the  breeding  .sea- 
son when  the  birds  were  banded,  the  data  were  ana- 
lyzed separately  for  males  and  females.  Within-indi- 
vidual  increases  in  size  and  breeding  performace  be- 
tween con.secutive  seasons  were  te.sted  with  one-tailed 
paired  /-tests  or  Wilcoxon  matched-pairs  signed  ranks 
te.sts;  if  an  individual  was  sampled  repeatedly  in  sev- 
eral years,  data  from  only  the  first  2 years  were  used 
in  these  analyses.  Size  differences  between  groups 
were  also  as.sessed  using  the  four  morphological  vari- 
ables in  a MANOVA,  with  group  and  year  as  the  in- 


dependent variables.  All  statistical  analyses  were  per- 
formed or  verified  using  Statistica  (5.1  and  98,  under 
Windows  3.1 1 and  95).  Statistical  significance  was  ac- 
cepted at  P < 0.05. 

RESULTS 

Among  males  there  was  no  relation  be- 
tween prior  residence  and  whether  a clutch 
was  initiated  in  their  nests  (Table  1).  In  con- 
trast, among  females  the  number  of  prior  res- 
idents that  initiated  a clutch  was  higher  than 
expected,  as  was  the  number  of  birds  new  to 
the  area  that  failed  to  initiate  a clutch.  The 
pattern  was  the  same  in  all  three  years,  sig- 
nificant in  2 of  the  3 years  and  in  the  analysis 
with  data  pooled  from  all  years  (Table  1).  Fi- 
nally, there  were  no  significant  differences  in 
first  egg  date  (F,  ,5,  = 0.65,  P > 0.05),  clutch 
size  (F,  ,47  = 0.17,  P > 0.05),  or  hatching  date 
~ 0.21,  P > 0.05)  between  birds  new 
to  the  area  and  returning  individuals. 

Differences  between  groups  do  not  neces- 
sarily imply  changes  within  individuals.  Com- 
parisons within  individuals  in  successive 
breeding  seasons  show  no  increases  in  clutch 
size  in  either  sex  (Wilcoxon  matched-pairs 
signed-ranks  tests:  males  Z = 0.36,  //  = 18,  F 
> 0.05;  females  Z = 0.82,  n = 21,  F > 0.05). 
However,  for  males  laying  date  was  signifi- 
cantly earlier  in  successive  breeding  seasons, 
on  average  2.8  days  earlier  (Wilcoxon 
matched-pairs  signed-ranks  test:  Z = 1.78,  n 
~ 18,  F = 0.04;  Fig.  la).  This  was  largely 
due  to  late-nesting  males  nesting  markedly 
earlier  in  the  following  year,  as  there  was  little 
difference  among  males  that  nested  in  late 


Relative  Clutch  Initiaion  Date,  Year  (n+1 ) Relative  Clutch  Initiation  Date,  Year  (n+1 ) 


384 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  3,  September  1999 


25-May  30-May  4-Jun  9-Jun  14-Jun  19-Jun  24-Jun  29-Jun 
Relative  Clutch  Initiation  Date,  Year  n 


29- Jun 

24-Jun 

19-Jun 

14-Jun 

9-Jun 

4-Jun 

30- May 


25-Ma' 


• 

• 

o 

Females 

O 

O 

o 

O 

0 

o 

0 

• * O 0 

. o o 

o 

o 

o o 

May  30-May  4-Jun  9-Jun  14-Jun  19-Jun  24-Jun  29-Jun 
Relative  Clutch  Initiation  Date,  Year  n 


FIG.  1.  Clutch  initiation  dates  by  individuals  in 
consecutive  breeding  seasons,  corrected  for  differences 
between  years  (see  methods).  The  dashed  lines  have  a 
slope  of  1,  which  would  result  if  clutch  initiation  oc- 
curred on  the  same  relative  date  in  successive  breeding 
seasons.  The  two  solid  circles  in  the  lower  graph  were 
categorized  as  outliers  (see  results). 


May— early  June  (Fig.  la).  In  contrast,  the  dif- 
ference in  females  was  only  1.3  days  earlier, 
and  was  not  significant  (Wilcoxon  matched- 
pairs  signed-ranks  test  Z = 0.93,  n = 2\,  P = 
0.18;  Fig.  lb).  However,  the  relationship  is 
heavily  influenced  by  2 nests  that  may  have 
been  re-nesting  attempts.  Without  these  out- 
liers the  margin  becomes  4.0  days  earlier  and 
statistically  significant  (Wilcoxon  matched- 
pairs  signed-ranks  test:  Z = 1.85,  n = 19,  P 
= 0.032;  Fig.  lb).  The  removal  of  these  out- 
liers did  not  lead  to  categorically  different 
conclusions  in  all  other  analyses. 

A significant  relationship  between  breeding 
performance  and  subsequent  return  occurred 
in  only  one  of  the  three  years.  Both  males  and 
females  that  had  at  least  started  a clutch  in 
1992  were  more  likely  to  return  the  following 
year  (Table  2),  but  there  was  no  evidence  of 
this  relationship  in  the  other  2 years.  Although 
the  results  were  also  significant  when  using 
pooled  data,  this  was  due  solely  to  the  1992 
results  (Heterogeneity  t^sts:  males  P = 
0.014;  females  P = 0.065;  Table  2).  Similar 
results  were  obtained  if  return  frequency  was 
compared  to  having  hatched  at  least  one  off- 
spring. Among  birds  that  nested,  there  were 
no  significant  differences  in  first  egg  date 
(F,,i66  = 0-60,  P > 0.05),  clutch  size  (F,.i66  = 
0.21,  P > 0.05),  or  hatching  date  (F,  = 

1.10,  P > 0.05)  between  birds  that  subse- 
quently returned  and  those  that  did  not.  There- 
fore, the  effect  of  breeding  performance  on 
subsequent  return,  as  measured  here,  was  cer- 
tainly not  consistent,  if  present  at  all. 


TABLE  ^ Relationship  between  having  reached  the  “first  egg"  stage  in  one  breeding  season  and  returning 

in  the  following  year.  Continuity  adjusted  or  Fisher's  exael  tests  (•)  were  used  dependin_g  on  the  sample  siee 

and  its  distribution  (expected  frequencies  in  parentheses).  Heterogeneity  tests;  males,  x 8.51.  dt 

0.014;  females:  X'  = 5.461,  df  = 2,  F = 0.065. 


Did  not  reach  the 

■•first  egg"  .stage 

Reached  the 

“Hrst  egg"  stage 

/f 

Did  not  return 

Returned 

Did  not  return 

Returned 

X" 

p 

Males 

1992 

67 

29  (20.5) 

3 (11.7) 

14  (22.5) 

21  (12.5) 

16.5 

<0.001 

1995 

47 

15  (14.9) 

1 1 (11.1) 

12  (12.1) 

9 (8.9) 

0.07 

»0.05 

1994 

71 

53  (32.3) 

12  (12.7) 

18  (18.7) 

8 (7.3) 

0.01 

»0.05 

Total 

185 

77  (67.4) 

26  (35.6) 

44  (53.6) 

38  (28.4) 

8.07 

0.005 

Females 

1 99'> 

57 

8 (4.1) 

1 (4.9) 

9 (12.9) 

19  (15.1) 

* 

0.005 

1995 

46 

17  (14.5) 

6 (8.5) 

12  (14.5) 

1 1 (8.5) 

1.49 

»0.05 

1 994 

66 

25  (26.3) 

6 (4.7) 

31  (29.7) 

4 (5.3) 

0.31 

»0.05 

Total 

149 

50  (43.1 ) 

13  (19.9) 

52  (58.9) 

34  (27.1) 

5.17 

0.023 

Lozano  and  Lemon  • BREEDING  SUCCESS  IN  YELLOW  WARBLERS 


385 


Wing  Chord  9th  Primary  Tail  Tarsus 


70 


Wing  Chord  9th  Primary  Tail  Tarsus 


FIG.  2.  Mean  size  ( + SE)  of  Yellow  Warblers  ver- 
sus residency  history.  MANOVAs  were  carried  out  us- 
ing all  variables  (Males:  Wilks’  \ = 0.901,  /^4i5i  = 
4.14,  P = 0.003;  Females;  Wilks’  X = 0.976,  C4  ,33  = 
0.803,  P ^ 0.05).  Asterisks  indicate  significant  differ- 
ences {P  < 0.05)  resulting  from  univariate  ANOVAs. 

Returning  males  were  significantly  larger 
than  new  arrivals  (Wilks’  \ = 0.901,  ^4  ,5,  = 
4.14,  P = 0.003;  Fig.  2)  but  returning  females 
were  not  larger  than  new  arrivals  (Wilks’  X.  = 
0.976,  ^4  ,33  = 0.803,  P > 0.05;  Fig.  2).  In 


contrast,  there  were  significant  within-individ- 
ual  increases  in  size  between  successive 
breeding  seasons  in  both  sexes  (Table  3). 

Finally,  there  were  no  significant  differenc- 
es in  morphology  between  nesting  birds  that 
subsequently  returned  and  nesting  birds  that 
did  not  return,  in  either  sex  (MANOVA: 
Males  Wilks’  \ = 0.989,  F4  ,67  = 0.448,  P > 
0.05;  Females  Wilks’  X = 0.9477,  F4  ,,5  = 
1.724,  P > 0.05). 

DISCUSSION 

The  effects  of  prior  residency  on  breeding 
performance  differed  substantially  between 
the  sexes.  Comparisons  of  new  arrivals  with 
returning  birds  showed  that  the  likelihood  of 
breeding  increased  with  prior  residency  in  fe- 
males, but  not  in  males.  Among  birds  that 
nested,  there  were  no  significant  differences 
in  the  timing  of  breeding  and  clutch  size  be- 
tween new  arrivals  and  returning  birds.  How- 
ever, within-individual  comparisons  showed 
that  males  began  breeding  significantly  earlier 
in  subsequent  breeding  seasons.  Therefore,  the 
effect  of  prior  residency  was  greater  for  fe- 
males than  for  males;  in  females  local  expe- 
rience affected  the  prospect  of  breeding,  and 
in  males  it  only  affected  the  timing  of  breed- 
ing. 

There  can  be  several  advantages  to  breed- 
ing earlier.  Early  breeders  often  have  larger 
clutches  (Erikstad  et  al.  1985,  Murphy  1986, 
Perrins  and  McCleery  1989).  In  the  popula- 
tion we  studied  the  median  clutch  size  de- 
creased from  five  eggs  in  nests  started  before 
June  15,  to  four  eggs  in  nests  initiated  there- 
after. Fledglings  from  earlier  nests  also  have 


TABLE  3.  Within-individual  increa.se.s 
results  from  one-tailed  paired  /-tests. 

in  size  between 

successive 

breeding  seasons 

in  Yellow  Warblers  and 

Variable 

n 

Difference  (mm) 

Year^ , i - Year, 

SE 

t 

p 

Males 

Tail 

48 

0.50 

0.32 

1 .55 

0.06 

Wing  chord 

49 

0.46 

0.21 

2.14 

0.02 

Ninth  primary 

49 

0.97 

0.28 

3.53 

<0.001 

Tarsus 

49 

0.20 

0.14 

1.49 

0.07 

Females 

Tail 

31 

0.27 

0.31 

0.89 

>0.05 

Wing  chord 

31 

0.74 

0.34 

2.18 

0.02 

iNinth  primary 

31 

0.74 

0.32 

2.32 

0.01 

Tarsus 

31 

0.31 

0.14 

2.28 

0.02 

386 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3.  September  1999 


more  time  to  develop  before  migration,  which 
in  our  area  begins  in  the  latter  half  of  July. 
Perhaps  most  important,  nest  losses,  whether 
they  are  due  to  inclement  weather  or  preda- 
tion, are  very  common  in  open  nesting  birds 
(Nice  1957,  Martin  and  Li  1992).  An  earlier 
start  provides  a longer  available  breeding  sea- 
son, which  enhances  the  probability  of  re- 
nesting following  nest  losses  (Lozano  et  al. 
1996). 

Birds  that  arrive  first  may  also  benefit  by 
obtaining  preferred  territories  (Wooller  and 
Coulson  1977,  Newton  1988,  Morris  and 
Lemon  1989,  Lemon  et  al.  1996).  On  the  other 
hand,  early  arrivals  also  risk  death  from  ex- 
posure to  cold  weather  and  lack  of  food  early 
in  the  spring  (Anderson  1965,  Whitmore  et  al. 
1977).  Several  other  researchers  have  shown 
that  earlier  arrivals  are  larger  and  in  better 
condition  than  later  arrivals  (e.g.,  Arvidsson 
and  Neergaard  1991,  Lozano  1994,  Stolt  and 
Fransson  1995).  Size  is  important  because  ar- 
rival time  depends  partially  on  the  ability  to 
withstand  adverse  conditions  early  in  the 
breeding  season.  In  our  population,  however, 
it  is  unlikely  that  size  affected  arrival  time, 
because  there  was  little  variance  in  arrival 
dates,  but  even  given  similar  arrival  times, 
large  size  may  still  be  advantageous  in  intra- 
sexual  competition  during  territory  establish- 
ment and  defense  (Arcese  1987,  Hogstad 
1989). 

Individual  males  were  larger  in  subsequent 
breeding  seasons,  and,  as  a group,  returning 
males  were  also  larger  than  males  new  to  the 
area.  In  contrast,  returning  females  were  not 
significantly  larger  than  females  new  to  the 
area,  but  individual  females  recaptured  in  con- 
secutive breeding  seasons  were  larger  from 
one  breeding  season  to  the  next.  At  first 
glance,  the  latter  results  may  appear  contra- 
dictory, but  the  two  analyses  are  not  equiva- 
lent. Females  may  increase  in  size  between 
consecutive  breeding  seasons  without  neces- 
sarily leading  to  returning  females  being  larg- 
er than  new  arrivals.  The  differences  between 
the  sexes  may  result,  for  instance,  if  the  breed- 
ing dispersal  or  size  variance  of  females  is 
greater  than  that  of  males. 

Size  may  also  play  a role  in  mate  choice 
later  in  the  season.  In  Yellow  Warblers  extra- 
pair paternity  is  widespread  (Yezerinac  et  al. 
1995),  and  within-pair  paternity  increases 


with  male  size  (Yezerinac  and  Weatherhead 
1997).  Female  preference  for  larger  males 
may  occur  if  size  is  an  honest  indicator  of 
phenotypic  or  genotypic  quality,  or  if  size  is 
being  used  as  an  indicator  of  age.  If  there  is 
a genetic  component  of  survival  ability,  age, 
in  itself,  would  be  a measure  of  genetic  qual- 
ity, and  females  would  benefit  by  mating  with 
older  males  (Manning  1985). 

We  do  not  know  the  exact  ages  of  all  adult 
birds  studied,  but  we  assume  the  mean  age  of 
new  arrivals  is  less  than  that  of  returning 
birds.  This  is  because  in  Yellow  Warblers 
breeding  dispersal  is  limited  (Yezerinac  and 
Weatherhead  1997),  but  natal  dispersal  is 
common,  as  it  is  in  birds  in  general  (Green- 
wood 1980,  Greenwood  and  Harvey  1982, 
Clarke  et  al.  1997).  We  have  no  measures  of 
dispersal  exclusive  of  mortality,  but  only  4.5% 
of  all  nestlings  returned,  compared  to  an  over- 
all yearly  return  of  36%  for  adults.  Although 
we  cannot  be  certain  that  every  individual 
classified  as  a new  arrival  was  only  one  year 
old,  returning  birds  were,  by  necessity,  at  least 
two  years  old.  Therefore,  when  compared  as 
two  separate  groups,  not  as  individuals,  it  is 
safe  to  conclude  that  returning  birds  were  old- 
er than  new  arrivals  (see  also  Yezerinac  and 
Weatherhead  1997). 

Age-dependent  increases  in  reproductive 
success  have  been  well  documented  in  birds 
(reviewed  by  Ssether  1990,  Forslund  and  Part 
1995),  but  the  proximate  mechanisms  respon- 
sible have  been  difficult  to  determine  because 
potential  factors  are  often  correlated.  Apparent 
age-dependent  increases  in  reproductive  suc- 
cess can  be  a statistical  consequence  of  the 
gradual  disappearance  of  poor  breeders  from 
a population,  which  would  result  if  the  phe- 
notypic or  genotypic  quality  of  an  individual 
affects  both  its  chances  of  breeding  and  sur- 
viving (Curio  1983).  Alternatively,  individuals 
may  be  more  likely  to  disperse  to  other  areas 
in  subsequent  breeding  seasons  after  a failed 
breeding  attempt  (Harvey  et  al.  1979,  Bensch 
and  Hasselquist  1991).  Either  mechanism 
would  yield  a difference  in  reproductive  suc- 
cess between  age  classes,  but  this  did  not  oc- 
cur in  our  study.  Except  for  one  year,  breeding 
performance  did  not  affect  subsequent  return 
for  either  sex.  Furthermore,  there  were  no 
morphological  differences  between  birds  that 
subsequently  returned  and  those  that  did  not. 


Lozano  and  Lemon  • BREEDING  SUCCESS  IN  YELLOW  WARBLERS 


387 


These  results  also  confirm  that  non-breeders 
were  not  actually  transients,  captured  while  on 
route  to  their  final  destinations.  Had  this  been 
the  case  they  would  have  been  less  likely  to 
return  than  breeders. 

The  restraint  hypothesis  proposes  that 
young  individuals  deliberately  withhold  repro- 
ductive effort.  Life-history  theory  suggests 
that  age  of  first  reproduction  is  an  important 
component  of  lifetime  reproductive  success 
(Charlesworth  1980,  Clutton-Brock  1988). 
Our  records  show  that  64%  of  all  adult  Yellow 
Warblers  fail  to  return,  which  makes  it  very 
unlikely  that  individuals  would  purposely 
forego  the  opportunity  to  reproduce,  especial- 
ly after  having  already  migrated  to  the  breed- 
ing grounds.  Moreover,  the  decision  not  to  re- 
produce would  be  optimal  only  if  there  is  a 
large  cost  to  reproduction,  but  we  found  that 
breeding  was  not  related  to  subsequent  return 
rates.  Therefore,  the  restraint  hypothesis  is 
probably  not  appropriate  to  explain  age-de- 
pendent  increases  in  reproductive  success  in 
Yellow  Warblers  or  other  short-lived  migra- 
tory species  (Wheelwright  and  Schultz  1994). 

Our  results  are  consistent  with  the  con- 
straint hypothesis,  which  predicts  that  younger 
birds  are  disadvantaged  when  competing  for 
breeding  opportunities.  However,  it  is  difficult 
to  know  to  what  extent  these  results  are 
caused  by  differences  in  local  experience  or 
by  age  in  itself.  Experimental  work  will  be 
required  to  determine  the  ecological  and  prox- 
imate mechanisms  responsible  for  these  dif- 
ferences (Martin  1995). 

ACKNOWLEDGMENTS 

We  thank  C.  Daniel,  G.  Goggin-Michaud,  D.  Lafleiir 
and  C.  Riley  for  their  help  with  the  field  work.  Finan- 
cial support  was  provided  by  NSERC,  FCAR,  and  the 
John  K.  Cooper  Foundation.  We  appreciate  the  com- 
ments of  J.  Grant.  D.  Kramer,  E.  Nol,  G.  Pollack,  R. 
Titman,  H.  Reiswig,  N.  Wheelwright,  and  D.  A.  Wie- 
denfeld. 

LITERATURE  CITED 

Aldrich,  J.  W.  1942.  Specific  relationships  of  the 
Golden  and  Yellow  warblers.  Auk  59:47-449. 
Anderson,  D.  W.  1965.  Spring  mortality  in  insectiv- 
orous birds.  Loon  37:134-135. 

Arcese,  P.  1987.  Age,  intrusion  pressure  and  defense 
against  floaters  by  temtorial  male  Song  Sparrows. 
Anim.  Behav.  35:773-784. 

Arvidsson,  B.  L.  and  R.  Neergaard.  1991.  Mate 


choice  in  the  Willow  Warbler — a field  experiment. 
Behav.  Ecol.  Sociobiol.  29:225-229. 

Bensch,  a.  and  D.  Hasselquist.  1991.  Territory  infi- 
delity in  the  polygynous  Great  Reed  Warbler  Ac- 
rocephaliis  ariindinacetis:  the  effect  of  variation 
in  tenitory  attractiveness.  J.  Anim.  Ecol.  60:875- 
871. 

Bent,  A.  C.  1963.  Life  histories  of  North  American 
wood  warblers,  part  one.  Dover  Publications  Inc., 
New  York. 

Charlesworth,  B.  1980.  Evolution  in  age-structured 
populations.  Cambridge  Univ.  Press,  Cambridge, 
U.K. 

Clarke,  A.  L.,  B-E.  SasTHER,  and  E.  Rpskaft.  1997. 
Sex  bia.ses  in  avian  dispersal:  a reappraisal.  Oikos 
79:429-438. 

Clutton-Brock,  T.  H.  1988.  Reproductive  success. 
Univ.  of  Chicago  Press,  Chicago,  Illinois. 

Cochran,  W.  G.  1954.  Some  methods  for  strengthen- 
ing the  common  tests.  Biometrics  10:417-451. 

Curio,  E.  1983.  Why  do  young  birds  reproduce  less 
well?  Ibis  125:400-404. 

CuRSON,  J.,  D.  Quinn,  and  D.  Beadle.  1994.  Warblers 
of  the  Americas.  Houghton  Miffin  Company,  Bos- 
ton, Massachusetts. 

Erikstad,  K.  E.,  H.  C.  Pedersen,  and  J.  B.  Steen. 
1985.  Clutch  size  and  egg  size  variation  in  Willow 
Grouse  Lagopus  /.  lagopus.  Ornis  Scand.  16:88- 
94. 

Everitt,  B.  S.  1992.  The  analysis  of  contingency  ta- 
bles, second  edition.  Chapman  & Hall,  London, 
U.K. 

Fienberg,  S.  E.  1980.  The  analysis  of  cross-classified 
categorical  data,  second  ed.  MIT  Press,  Cam- 
bridge, Massachusetts. 

Forslund,  P.  and  T.  Part.  1995.  Age  and  reproduction 
in  birds-hypotheses  and  tests.  Trends  Ecol.  Evol. 
10:374-378. 

Greenwood,  P.  J.  1980.  Mating  systems,  philopatry 
and  dispersal  in  birds  and  mammals.  Anim.  Be- 
hav. 28:1140-1 162. 

Greenwood,  P.  J.  and  P.  H.  Harvey.  1982.  The  natal 
and  breeding  dispersal  of  birds.  Annu.  Rev.  Ecol. 
Syst.  13:1-21. 

Harvey,  P.  H.,  P.  J.  Greenwood,  and  C.  M.  Perrins. 
1979.  Breeding  area  fidelity  of  the  Great  Tit  (Pa- 
ms major).  J.  Anim.  Ecol.  48:305-313. 

Hinde,  R.  a.  1956.  The  biological  significance  of  the 
ten'itories  of  birds.  Ibis  98:340-369. 

Hogstad,  O.  1989.  The  presence  of  non-territorial 
males  in  Willow  Warbler  Phyllo.uopiis  trochilus 
populations — a removal  .study.  Ibis  131:263-267. 

Klein,  N.  K.  and  W.  M.  Brown.  1994.  Intraspecific 
molecular  phylogeny  in  the  Yellow  Warbler  (Den- 
droica  petechia)  and  implications  for  avian  bio- 
geography of  the  We.st  Indies.  Evolution  48: 1914- 
1932. 

Lemon,  R.  E.,  S.  Perreault,  and  G.  A.  Lozano.  1996. 
Breeding  di.spersions  and  site  fidelity  of  American 
Redstarts  (Setophaga  mticilla).  Can.  J.  Zool.  74: 
2238-2247. 


388 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  Sepleinher  1999 


Lozano,  G.  A.  1994.  Size,  condition,  and  territory 
ownership  in  male  Tree  Swallows  (Tachycineta 
bicolor).  Can.  J.  Zool.  72:330-333. 

Lozano.  G.  A.  and  R.  E.  Lemon.  1995.  Food  abun- 
dance and  parental  care  in  Yellow  Warblers  (Den- 
droica  petechia).  Behav.  Ecol.  Sociobiol.  37:45— 
50. 

Lozano,  G.  A.  and  R.  E.  Lemon.  1996.  Male  plumage, 
paternal  care  and  reproductive  success  in  Yellow 
Warblers,  Dendroica  petechia.  Anim.  Behav.  51: 
265-272. 

Lozano,  G.  A.  and  R.  E.  Lemon.  1998.  Parental  care 
responses  by  Yellow  Warblers  (Dendroica  pete- 
chia) to  simultaneous  manipulations  of  food  abun- 
dance and  brood  size.  Can.  J.  Zool.  76:916—924. 

Lozano,  G.  A.,  S.  Perreault,  and  R.  E.  Lemon.  1996. 
Age,  arrival  date  and  reproductive  success  of  male 
American  Redstarts  Setophaga  nilicilla.  J.  Avian 
Biol.  27:164-170. 

Manning,  J.  T.  1985.  Choosy  females  and  correlates 
of  male  age.  J.  Theor.  Biol.  1 16:349-354. 

Martin,  K.  1995.  Patterns  and  mechanisms  for  age- 
dependent  reproduction  and  survival  in  birds.  Am. 
Zool.  35:340-348. 

Martin,  T.  E.  and  P.  Li.  1992.  Life  history  traits  of 
open-  vs.  cavity-nesting  birds.  Ecology  73:579- 
592. 

Morris,  M.  M.  J.  and  R.  E.  Lemon.  1989.  Mate  choice 
in  American  Redstarts:  by  territory  quality?  Can. 
J.  Zool.  66:2255-2261. 

Murphy,  T.  M.  1986.  Temporal  components  of  repro- 
ductive variability  in  Eastern  Kingbirds  (Tyranini.s 
tyrannu.s).  Ecology  67:1483—1492. 

Newton,  I.  1988.  Age  and  reproduction  in  the  Spar- 
rowhawk.  Pp.  201-219  in  Reproductive  success 
(T.  H.  Clutton-Brock,  Ed.).  Univ.  of  Chicago 
Press,  Chicago,  Illinois. 

Nice,  M.  M.  1957.  Nesting  success  in  altricial  birds. 
Auk  74:305-321. 

Nol,  E.  and  j.  N.  M.  Smith.  1987.  Effects  of  age  and 
breeding  experience  on  seasonal  reproductive  suc- 
cess in  the  Song  Sparrow.  J.  Anim.  Ecol.  56:301- 
313. 

Perrins,  C.  M.  and  R.  H.  McCleery.  1989.  Laying 


date  and  clutch  size  in  Great  Tits.  Wilson  Bull. 
101:236-253. 

PiANKA,  E.  R.  and  W.  S.  Parker.  1975.  Age-specific 
reproductive  tactics.  Am.  Nat.  109:453-464. 

Pyle,  R,  L.  B.  Spear,  W.  J.  Sydeman,  and  D.  G.  Ain- 
LEY.  1991.  The  effects  of  experience  and  age  on 
the  breeding  performance  of  Western  Gulls.  Auk 
108:25-33. 

Reid,  W.  V.  1988.  Age-specific  patterns  of  reproduc- 
tion in  the  Glaucous-winged  Gull:  increased  effort 
with  age?  Ecology  69:1454-1465. 

S/ETHER,  B-E.  1990.  Age-specific  variation  in  repro- 
ductive performance  of  birds.  Curr.  Ornithol.  7: 
251-283. 

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

Smith,  J.  N.  M.  1981.  Does  high  fecundity  reduce  sur- 
vival in  Song  SpaiTOWS?  Evolution  35: 1 142-1 148. 

Smith,  H.  G.  1993.  Parental  age  and  reproduction  in 
the  Marsh  Tit  Parit.s  pahistris.  Ibis  135:196-201. 

Stolt,  B.  O.  and  T.  Fransson.  1995.  Body  mass,  wing 
length  and  spring  arrival  of  the  Ortolan  Bunting 
Emheriza  hortidana.  Ornis  Fennica  72:14-18. 

Wheelwright,  N.  T.  and  C.  B.  Schultz.  1994.  Age 
and  reproduction  in  Savannah  Span'ows  and  Tree 
Swallows.  J.  Anim.  Ecol.  63:686-702. 

Whitmore,  R.  C.,  J.  A.  Mosher,  and  H.  H.  Frost. 
1977.  Spring  migrant  mortality  during  unseason- 
able weather.  Auk  94:778-781. 

WiEDENFELD,  D.  A.  1991.  Geographical  moiphology  of 
male  Yellow  Warblers.  Condor  93:712—723. 

Williams,  G.  C.  1966.  Natural  selection,  the  cost  of 
reproduction  and  a refinement  of  Lack's  principle. 
Am.  Nat.  100:687—690. 

WOOLER,  R.  D.  AND  J.  C.  CouLSON.  1977.  Factors  af- 
fecting the  age  of  first  breeding  of  the  Kittiwake. 
Ris.sa  tridactyki.  Ibis  119:339-349. 

Yezerinac,  S.  M.  and  P.  J.  Weatherhead.  1997.  Ex- 
tra-pair mating,  male  plumage  coloration,  and  sex- 
ual selection  in  Yellow  Warblers  (Dendroica  pe- 
techia). Proc.  R.  Soc.  Lond.  B 264:527-532. 

Yezerinac,  S.  M.,  P.  J.  Weatherhead,  and  P.  T.  Boag. 
1995.  Extra-pair  paternity  and  the  opportunity  for 
sexual  selection  in  a socially  monogamous  bird 
(Dendroica  petechia).  Behav.  Ecol.  Sociobiol.  37: 
179-188. 


Wilson  Bull..  1 I 1(3),  1999,  pp.  389-396 


DISTRIBUTION  AND  HABITAT  ASSOCIATIONS  OF  THREE 
ENDEMIC  GRASSLAND  SONGBIRDS  IN 
SOUTHERN  SASKATCHEWAN 

S.  K.  DAVIS, D.  C.  DUNCAN,'^  AND  M.  SKEEL'^ 

ABSTRACT. — We  conducted  1675  point  counts  on  93  survey  routes  to  determine  the  distribution  and  habitat 
associations  of  three  endemic  grassland  songbirds  across  the  four  prairie  ecoregions  of  southern  Saskatchewan, 
Canada.  Within  the  four  habitat  types  surveyed,  Sprague’s  Pipits  (Anthiis  spragueii)  and  Chestnut-collared  Long- 
spurs  (Calcariits  ornatu.s)  occuned  more  frequently  in  native  and  seeded  pastures  than  in  hayland  and  cropland, 
whereas  Baird's  Sparrows  (Ammodramus  hairdii)  occurred  as  frequently  in  hayland  as  in  native  and  seeded 
pastures.  The  occuiTcnce  of  Baird's  Spanows  and  Chestnut-collared  Longspurs  did  not  differ  significantly  be- 
tween lightly,  moderately,  and  heavily  grazed  native  pastures,  whereas  Sprague’s  Pipits  occurred  less  frequently 
in  heavily  grazed  pastures.  Sprague’s  Pipits  and  Chestnut-collared  Longspurs  occurred  more  often  in  the  drier 
prairies  of  the  southern  portion  of  the  province,  but  Chestnut-collared  Longspurs  were  virtually  absent  from  the 
cypress  upland  ecoregion.  In  contrast,  Baird’s  Sparrows  occurred  most  frequently  in  the  .semi-arid  grasslands  of 
the  moist-mixed  grassland  ecoregion.  In  native  pastures.  Chestnut-collared  Longspurs  were  associated  with  a 
lower  density  of  short  grasses  and  lesser  amounts  of  litter  whereas  Baird’s  Sparrows  were  associated  with  a 
higher  density  of  taller  grasses  and  sparse  shrub  cover.  Our  results  suggest  that  conservation  programs  that 
convert  annually  tilled  cropland  to  perennial  forage  could  provide  additional  habitat  for  endemic  grassland  birds. 
Received  2 Nov.  1998,  accepted  2 April  1999. 

Land  settlement  and  agriculture  have  great- 
ly altered  the  landscape  of  southern  Saskatch- 
ewan. Only  17%  of  the  province’s  original  na- 
tive prairie  remains  (Samson  and  Knopf  1994) 
and  is  currently  threatened  by  cultivation, 
over-grazing  by  livestock,  invasion  by  exotic 
plant  species,  and  urban  development.  In  areas 
of  Saskatchewan  where  soils  and  landscapes 
are  particularly  suited  for  crop  production, 
less  than  0.1%  of  the  original  native  prairie 
remains  (Riemer  et  al.  1997). 

Despite  the  loss  of  native  grassland,  Sas- 
katchewan supports  a diverse  grassland  avi- 
fauna (Smith  1996),  including  1 1 of  the  12 
primary  endemic,  and  17  of  the  25  secondary 
endemic  grassland  bird  species  as  outlined  by 
Knopf  (1994).  Primary  endemic  species  gen- 
erally have  more  restricted  breeding  ranges 
(Sauer  et  al.  1997)  and  are  less  flexible  in  their 
habitat  requirements  than  more  broadly  dis- 
tributed species  (Owens  and  Myres  1973, 

Knopf  1996,  Davis  and  Duncan  in  press). 

Consequently,  continental  populations  of  7 of 


' Saskatchewan  Wetland  Conservation  Corporation. 
202-2050  Cornwall  St.,  Regina  SK,  Canada,  S4P  2K5. 

-Present  address:  Canadian  Wildlife  Service,  200- 
4999-98"-  Ave.,  Edmonton  AB,  Canada,  T6B  2X3. 

^Present  address:  Nature  Saskatchewan,  206-1860 
Lome  St.,  Regina  SK,  Canada,  S4P  2L7. 

^ Corresponding  author;  E-mail:  .sdavis@wetland.sk. ca 


the  1 1 primary  endemic  species  are  currently 
in  decline  (Sauer  et  al.  1997),  possibly  as  a 
result  of  loss  and  degradation  of  native  prairie 
habitat. 

Information  on  the  habitat  associations  of 
endemic  songbirds  can  provide  insight  into 
population  declines  by  identifying  habitat  fea- 
tures that  correlate  with  their  occurrence.  This 
information  may  be  used  in  making  manage- 
ment decisions  and  formulating  land-use  pol- 
icies. Although  many  researchers  have  ex- 
amined habitat  selection  by  grassland  song- 
birds (Cody  1968,  Owens  and  Myres  1973, 
Whitmore  1979,  Rotenberry  and  Wiens  1980, 
Johnson  and  Temple  1986,  Mahon  1995),  lit- 
tle research  has  been  focused  on  birds  of  the 
northern  mixed-grass  prairie.  Furthermore, 
most  studies  in  the  mixed-grass  prairie  region 
have  been  conducted  on  intensively  managed 
sites  (Dale  1983,  Winter  1994,  Madden  1996, 
Dale  et  al.  1997)  or  were  located  within  a 
small  geographic  area  (Arnold  and  Higgins 
1986,  Sutter  1996;  but  see  Johnson  and 
Schwartz  1993,  Davis  and  Duncan  in  press). 

We  examined  habitat  associations  and  the 
distribution  of  three  endemic  songbird  species 
of  the  northern  mixed-grass  prairie  across  the 
entire  Prairie  Ecozone  of  southern  Saskatch- 
ewan. The  objectives  of  the  study  were  (1)  to 
determine  whether  the  frequencies  of  occur- 


389 


390 


THE  WILSON  BULLETIN  • Vol.  III.  No.  3.  September  1999 


rence  of  Sprague’s  Pipit  {Anthus  spragueii), 
Baird’s  Sparrow  {Ammodramus  bairdii),  and 
Chestnut-collared  Longspur  (Calcarius  orna- 
tus)  differ  among  native  pasture,  seeded  pas- 
ture, hayland,  and  cropland;  (2)  to  determine 
whether  various  levels  of  grazing  intensity  in- 
fluence the  occurrence  of  these  species;  (3)  to 
determine  the  distribution  of  the  three  species 
relative  to  the  prairie  ecoregions  of  Saskatch- 
ewan; and  (4)  to  identify  structural  compo- 
nents of  native  prairie  vegetation  important  in 
predicting  songbird  occurrence. 

STUDY  AREA  AND  METHODS 

Study  area. — We  conducted  grassland  bird  surveys 
throughout  the  Prairie  Ecozone  of  southern  Saskatch- 
ewan. The  Prairie  Ecozone  covers  24,103,000  ha  and 
comprises  four  ecoregions;  cypress  upland,  mixed 
grassland,  moist-mixed  grassland,  and  aspen  parkland 
(Fig.  1;  Ecological  Stratification  Working  Group 
1995).  The  cypress  upland  in  the  extreme  southwestern 
portion  of  the  province  rises  400—500  m above  the 
prairie  landscape.  This  region  is  characterized  by  slop- 
ing escarpments,  valleys,  and  coulees.  Wheatgrass 
(Agropyron  spp.)  and  speargrass  {Stipa  spp.)  dominate 
the  dark  brown  soils  of  the  lower  elevations,  whereas 
fescue  (Festuca  spp.)  prairie  predominates  on  the 
slopes  and  at  higher  elevations.  The  mixed  grassland 
is  the  driest  region  of  Saskatchewan  and  is  character- 
ized by  wheatgrass,  speargrass,  and  blue  grama  grass 
(Bouteloua  gracilis).  Because  of  the  lack  of  moisture, 
trees  and  wetlands  are  scarce;  shrubs  are  restricted  to 
mesic  areas.  The  moist-mixed  grassland  represents  the 
northern  extent  of  the  open  grasslands  in  Saskatche- 
wan. This  region  is  characterized  by  semiarid  condi- 
tions and  dark  brown  soils.  Speargrass,  wheatgrass, 
and  deciduous  shrubs  predominate.  The  aspen  park- 
land ecoregion  is  characterized  by  trembling  aspen 
{Populus  tremuloide.s)  groves  and  fescue  grasslands, 
although  the  latter  habitat  is  now  rare  (Sask.  Wetland 
Conserv.  Corp.,  unpubl.  data). 

Bird  surveys  were  conducted  in  native  pasture,  seed- 
ed pasture,  hayland,  and  cropland.  Survey  routes  were 
designed  to  sample  mostly  grassland  habitat,  although 
each  of  the  four  habitat  types  was  sampled  on  most 
routes.  Native  prairie  was  characterized  by  Stipa  spp., 
June  grass  (Koeleria  cristata).  northern  wheatgrass 
(Agropyron  dasystachyum),  western  wheatgrass  (A. 
smithii).  blue  grama  grass,  Care.x  spp.,  club  moss  (Se- 
laginella  densa).  pasture  sage  (Artemisia  frigida),  and 
various  forbs.  The  most  common  shrubs  were  western 
snowberry  (Symphoricarpos  occidentalis),  rose  (Rosa 
spp.),  and  wolf  willow  (Eleagmis  commutata).  Seeded 
pasture  was  defined  as  land  that  had  been  broken  and 
seeded  with  exotic  perennial  gras.ses  tor  grazing,  most 
commonly  crested  wheatgrass  (Agropyron  cri.statum) 
or  brome  grass  (Bronnts  spp.),  with  altalta  (Medi(  ago 
spp.)  or  sweet  clover  (Melilotus  spp.)  sometimes  pre- 
sent. Hayland  was  defined  as  cultivated  land  that  had 


been  seeded  to  perennial  crops  for  haying.  Vegetation 
on  hayland  ranged  from  100%  alfalfa,  to  mixes  of  al- 
falfa, sweet  clover  and  introduced  grasses  such  as 
brome  grass,  crested  wheatgrass,  or  bluegrass  (Poa 
spp.).  Cropland  was  cultivated  land  that  was  seeded  to 
annual  crops,  most  commonly  wheat  (Triticum  aesti- 
vum)  or  canola  (Brassica  spp.). 

Route  selection. — We  numbered  (1-42)  townships 
from  the  U.S.  border  to  the  northern  extent  of  the  as- 
pen parkland,  and  renumbered  (1-64)  range  locations 
from  the  Manitoba  border  west  to  the  Alberta  border. 
The  starting  points  of  76  routes  were  then  located  by 
selecting  township-range  numbers  from  a random 
number  table.  Because  the  target  species  (Baird’s  Spar- 
row) was  believed  to  be  a grassland  specialist  (Cart- 
wright et  al.  1937,  Owens  and  Myres  1973),  we  only 
included  townships  where  most  point  counts  (>80%) 
could  be  located  in  grassland  habitat.  Thus  more  routes 
were  established  in  the  mixed  grassland  ecoregion  than 
the  other  ecoregions  because  most  of  the  grassland  in 
Saskatchewan  exists  in  this  region  (Saskatchewan  Dig- 
ital Land  Cover  Project,  unpubl.  data).  Because  Baird’s 
Sparrow  was  a threatened  species  in  1994  (Goossen  et 
al.  1993),  we  assumed  they  were  uncommon  and  es- 
tablished an  additional  19  non-random  routes  in  areas 
where  the  species  was  thought  to  occur  regularly.  The 
proportion  of  point  counts  in  grassland  habitat  was 
nearly  identical  for  random  and  non-random  routes  (79 
and  80%,  re.spectively). 

Bird  occurrence. — Six  surveyors  recorded  the  num- 
ber of  singing  males  using  five-minute,  100-m  fixed 
radius  point  counts  (Ralph  et  al.  1993)  conducted  from 
roads  and  trails  at  approximately  0.8  km  intervals 
along  each  route.  Each  point  count  was  subdivided  into 
half-circle  counts  by  recording  birds  detected  on  the 
left  or  right  side  of  the  road  or  trail  because  habitat 
types  and  grazing  intensity  often  differed  on  each  side. 
Although  roadside  sampling  may  confound  species- 
habitat  as.sociations  because  of  the  influence  of  road- 
side vegetation,  the  three  species  in  this  study  are  not 
attracted  to  roadside  habitat  (Sutter  et  al.,  in  press). 
Thus  the  occurrence  of  these  species  in  the  four  habitat 
types  is  likely  not  confounded  by  roadside  vegetation. 
Surveyors  attended  a training  session  prior  to  the  study 
to  standardize  protocol  and  reduce  surveyor  bias.  Sur- 
veys were  conducted  between  4 June  and  2 July,  1994, 
commenced  30  minutes  before  sunrise,  on  days  with 
no  precipitation,  and  winds  less  than  20  km/hr.  Each 
route  had  20-25  stops. 

Habitat  as.sociations. — Bird  surveyors  quantified 
eight  measures  of  vegetation  structure  at  every  second 
native  pasture  point  count  location  on  the  same  day  as 
the  survey.  Vegetation  was  assessed  on  only  native 
pastures  because  of  logistic  constraints  in  sampling  all 
habitats  over  a large  area  and  because  of  our  increased 
sampling  of  native  pasture.  Each  surveyor  laid  a meter 
stick  on  the  ground  at  least  35  m from  the  road/trail 
at  a random  location  within  each  portion  of  the  half- 
circle. A 6 mm  diameter  metal  rod  was  passed  verti- 
cally through  the  vegetation  at  each  end  of  the  meter 
stick  and  the  number  of  contacts  by  different  vegeta- 


Davis  el  al.  • ENDEMIC  GRASSLAND  SONGBIRDS 


391 


FIG.  1.  Location  of  random  (closed  circles)  and  non-random  (open  circles)  survey  routes  and  ecoregions 
within  the  Prairie  Ecozone  of  southern  Saskatchewan. 


tive  life  forms  (e.g.,  standing  dead  vegetation,  narrow- 
leaf  grass,  broad-leaf  grass,  forb,  shrubs  >15  cm,  and 
dwarf  shrubs  <15  cm  high)  counted  in  each  successive 
1 decimeter  (dm)  height  interval  (Rotcnbeny  and 
Wiens  1980).  The  occuiTence  of  each  life  form  was 
lumped  into  two  height  categories;  1 dm  (number  of 
contacts  <1  dm)  and  2+  dm  (number  of  contacts  >1 
dm)  height  intervals  because  vegetation  contacts  in  na- 
tive pastures  rarely  occurred  in  the  higher  levels.  Litter 
depth  was  measured  from  the  surface  of  the  ground  to 


the  top  of  the  litter,  and  distance  to  the  nearest  shrub 
was  visually  e.stimated. 

Grazing  intensity  was  estimated  in  June  and  early 
July  by  a range  ecologist  as  idle,  lightly,  moderately, 
or  heavily  grazed.  Idle  prairie  was  defined  as  native 
vegetation  that  had  not  been  grazed  for  at  least  two 
years.  Heavily  grazed  pa.stures  were  characterized  by 
virtually  all  plant  material  and  litter  removed,  greater 
than  20%  bare  .soil,  greater  than  40%  club  moss,  and 
small  plants  with  poor  vigor.  Lightly  grazed  pasture 


392 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


TABLE  1 . Influence  of  habitat  type, 
singing  males  in  southern  Saskatchewan. 

ecoregion,  and  grazing  intensity  on  the  occurrence  (%  half  circles)  of 
For  each  species,  values  followed  by  the  same  letter  do  not  differ  (P 

> 0.05)  from  one  another. 

Sprague's  Pipit 

Baird's  Sparrow 

Chestnut-collared 

Longspur 

Habitat 

Native  pasture  (n  = 1158) 

Seeded  pasture  (/;  = 192) 

Hayland  (/?  = 1 16) 

Cropland  {n  = 209) 

18.5  A 

14.1  A 

2.6  B 

0.5  B 

30.5  A 
31.2  A 
25.0  A 

5.3  B 

21.4  A 

17.7  A 

1.7  B 

0.5  B 

Ecoregion 

Aspen  parkland  (;?  = 238) 

Cypress  upland  (n  = 149) 

Mixed  grassland  {n  = 805) 

Moist-mixed  grassland  (n  = 483) 

8.4  A 

24.2  B 

16.9  B 

6.2  A 

15.9  A 

16.8  A 

25.8  B 
37.7  C 

6.7  A 

2.0  B 

24.4  C 

14.5  D 

Lightly  grazed  (/?  = 112) 

Moderately  grazed  (n  = 137) 

Heavily  grazed  (/?  = 81 ) 

Grazing  intensity 

19.6  A 

22.6  A 

9.9  B 

34.8 

37.9 

25.9 

23.2 

24.1 

22.2 

was  defined  as  having  little  evidence  of  grazing,  abun- 
dant litter  and  plant  material,  less  than  10%  bare  soil, 
less  than  10%  club  moss,  and  robust  and  vigorous 
plants.  Moderately  grazed  pastures  exhibited  charac- 
teristics intermediate  between  the  previous  two  types. 

Statistical  analyses. — We  used  songbird  occurrence 
(presence/absence)  within  half-circle  point  counts  for 
all  analyses  because  more  than  one  individual  was  re- 
corded in  only  9%,  8%,  and  2%  of  the  half-circle 
counts  for  Baird's  Sparrow,  Chestnut-collared  Long- 
spur  and  Sprague’s  Pipit,  respectively.  Because  the  two 
half-circle  counts  at  any  given  stop  do  not  represent 
independent  observations,  we  randomly  chose  only 
one  half-circle  count  on  either  side  of  the  road  or  trail 
for  inclusion  in  all  subsequent  analyses.  x“  contingency 
analyses  were  used  to  determine  whether  the  frequency 
of  occurrence  of  songbirds  inside  halt-circle  point 
counts  was  influenced  by  land-use,  grazing  intensity, 
and  ecoregion.  Pair-wise  comparisons  were  pertormed 
only  for  those  species  where  the  overall  significance 
level  was  P < 0.05.  Although  most  of  the  native  and 
seeded  pastures  were  grazed  by  cattle,  5%  of  the  sam- 
ple points  had  been  idle  for  at  least  two  years  and  were 
thus  omitted  from  all  analyses.  A multivariate  assess- 
ment of  songbird  occurrence  in  native  pasture  was  con- 
ducted using  step-wise  logistic  regression  on  half-cir- 
cle point  counts  using  vegetative  structure,  grazing  in- 
tensity, and  ecoregions.  None  of  these  variables  were 
highly  correlated  with  each  other  (all  comparisons  r’ 
< 0.42,  P > 0.001 ) except  for  the  number  of  contacts 
of  broad-leaf  grass  in  the  first,  and  2+  decimeter  cat- 
egories (/-’  = 0.71.  P < 0.001);  thus  only  broad-leaf 
grass  contacts  in  the  first  decimeter  were  used.  Level 
of  significance  for  variable  inclusion  in  the  models  was 
set  at  0.05.  All  analyses  were  performed  using  SAS 
statistical  software  v.  6.12  (.SAS  Institute  Inc.  1989). 


RESULTS 

Land  use. — The  occurrence  of  each  of  the 
three  endemic  species  differed  significantly 
among  habitat  types  (x^  = 59.2-75.6,  df  = 3, 

P < 0.001).  Sprague’s  Pipit  and  Chestnut-col- 
lared Longspur  occurred  more  frequently  in 
native  and  seeded  pasture  than  in  hay  land  or 
cropland,  whereas  Baird’s  Sparrows  occurred 
as  frequently  in  native  and  seeded  pastures  as 
in  hayland,  but  occurred  least  frequently  in 
cropland  (Table  1). 

Distribution. — The  occurrence  of  each  spe- 
cies differed  among  ecoregions  (Table  1 ; x"  = 
51.1-74.4,  df  = 3,  P < 0.001).  Furthermore, 
ecoregion  type  was  a significant  predictor  of 
occurrence  in  each  of  the  three  logistic  re- 
gression models  (Table  2).  Sprague  s Pipits 
were  recorded  in  relatively  low  abundance 
throughout  the  study  area,  occurring  most  fre- 
quently in  the  cypress  upland  and  mixed 
grassland  ecoregions  (Table  1).  Chestnut-col- 
lared Longspurs  were  primarily  restricted  to 
the  extreme  southern  portion  of  the  province, 
particularly  within  the  mixed  grassland  ecore- 
gion (Tables  1,  2).  Baird’s  Sparrows  Were 
abundant  throughout  much  of  the  study  area, 
but  were  recorded  most  frequently  in  the 
moist-mixed  grassland  (Tables  1,  2). 

Habitat  associations. — Grazing  intensity 
had  little  influence  on  the  occurrence  of 


Davis  el  al.  • ENDEMIC  GRASSLAND  SONGBIRDS 


393 


TABLE  2.  Results  summary  ot  step-wise  logistic  regression  analyses  of  grassland  songbird  occurrence  in 
native  pasture.  Variables  are  presented  in  the  order  they  were  entered  into  the  model. 

Purunieter 

Species 

Variable 

estimate 

Wald 

r 

R2 

Sprague’s  Pipit 

Intercept 

-0.894 

19.922 

<0.001 

Moist-mixed  grassland 

-1.920 

12.377 

<0.001 

0.092 

Heavily  grazed 

-0.981 

4.261 

0.039 

Baird's  Sparrow 

Intercept 

- 1 .346 

26.653 

<0.00 1 

Narrow-leaf  grass  2'  dm 

0.315 

9.480 

0.002 

0.075 

Moist-mixed  grassland 

0.825 

8.209 

0.004 

Shrub  distance 

0.008 

3.941 

0.002 

Chestnut-collared  Longspur 

Intercept 

-0.605 

3.664 

0.055 

Litter  depth 

-0.035 

4.525 

0.033 

0.088 

Narrow-leaf  grass  1 dm 

-0.278 

6.053 

0.014 

Mixed  grassland 

0.675 

4.535 

0.033 

Baird’s  Sparrow  and  Chestnut-collared  Long- 
spur  in  native  pasture  (Table  1;  x“  = 3.4,  df 
= 2,  P > 0.05  and  = 0.1,  df  = 2,  P > 
0.05,  respectively).  The  occurrence  of 
Sprague’s  Pipits  in  native  pasture  differed  sig- 
nificantly among  grazing  intensity  levels  (x^ 
= 6.2,  df  = 2,  P = 0.045).  Sprague’s  Pipits 
were  negatively  associated  with  heavy  grazing 
(Table  2),  occurring  twice  as  often  in  lightly 
and  moderately  grazed  pastures  as  in  heavily 
grazed  pastures  (Table  1).  Although  grazing 
intensity  did  not  significantly  influence  the  oc- 
currence of  Baird’s  Sparrows,  the  species  was 
associated  with  pastures  having  greater  cov- 
erage of  grasses  over  10  cm  and  were  attracted 
to  pastures  with  sparse  shrub  cover  (Table  2). 
In  contrast.  Chestnut-collared  Longspurs  ap- 
peared to  be  associated  with  pastures  that 
were  characterized  by  less  dense  vegetative 
cover.  The  species  was  negatively  associated 
with  depth  of  the  litter  and  the  density  of  nar- 
row-leaf grasses  in  the  first  decimeter  (Table 
2).  However,  the  amount  of  variation  ex- 
plained by  each  model  was  extremely  poor 
(Table  2). 

DISCUSSION 

Land-use. — Sprague’s  Pipits  and  Chestnut- 
collared  Longspurs  were  mostly  restricted  to 
grassland  habitat  in  southern  Saskatchewan. 
Although  we  detected  no  differences  in  the 
frequency  of  occurrence  of  these  species  in 
native  and  seeded  pastures,  others  have  re- 
ported Sprague’s  Pipits  (Owens  and  Myres 
1973,  Hartley  1994,  Madden  1996,  Dale  et  al. 
1997)  and  Chestnut-collared  Longspurs 
(Stewart  and  Kantrud  1972,  Owens  and  Myres 


1973)  to  prefer  native  prairie  over  a number 
of  other  habitat  types.  The  attractiveness  of 
seeded  pastures  in  this  study  may  have  been 
influenced  by  the  age  and  structural  compo- 
sition of  the  seeded  pastures  such  that  a num- 
ber of  these  pastures  may  have  been  similar 
to  native  pastures.  Indeed,  surveyors  consult- 
ed with  the  range  ecologist  on  several  occa- 
sions to  confirm  whether  pastures  were  native 
or  had  been  cultivated  in  the  past.  More  re- 
cently seeded  pastures,  or  those  dominated  by 
certain  exotic  plant  species  may  be  less  suit- 
able for  Sprague’s  Pipits  and  Chestnut-col- 
lared Longspurs.  In  Saskatchewan,  both  spe- 
cies occur  more  frequently  and  in  higher 
abundance  in  native  pastures  than  in  pastures 
dominated  by  crested  wheatgrass  (Sutter 
1996,  Davis  et  al.  1996,  Davis  and  Duncan  in 
press).  Siinilarly,  fields  comprised  predomi- 
nantly of  smooth  brome  grass  {Bromus  iner- 
mis)  are  unsuitable  for  these  species  (Wilson 
and  Belcher  1989). 

In  our  study,  Baird’s  Sparrows  exhibited 
more  flexibility  in  their  habitat  use  than  either 
Chestnut-collared  Longspurs  or  Sprague’s 
Pipits.  The  occurrence  of  Baird’s  Sparrows  in 
habitats  other  than  native  prairie  has  been  well 
documented  (reviewed  in  Davis  et  al.  1996) 
despite  earlier  studies  that  suggested  the  spar- 
row was  a native  prairie  specialist  (Cartwright 
et  al.  1937,  Owens  and  Myres  1973).  Al- 
though we  frequently  recorded  Baird’s  Spar- 
rows in  hayfields,  this  habitat  may  act  as  a 
population  sink  (Pulliam  1988).  Using  a pro- 
ductivity index.  Dale  and  coworkers  (1997) 
found  significantly  fewer  signs  of  productive 


394 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


behavior  following  mowing  of  hayfields  in 
southern  Saskatchewan.  Mowing  destroys  ap- 
proximately 50%  of  ground  nests  and  the  pro- 
ductivity of  breeding  birds  in  hayfields  is  of- 
ten below  that  required  to  maintain  a stable 
population  (Frawley  1989,  Bollinger  et  al. 
1990). 

Distribution  and  habitat  associations. — We 
found  that  the  distribution  of  Sprague’s  Pipit, 
Baird’s  Sparrow,  and  Chestnut-collared  Long- 
spur  closely  resembled  that  outlined  by  the 
Saskatchewan  breeding  bird  atlas  (Smith 
1996)  and  the  Breeding  Bird  Survey  (BBS; 
Sauer  et  al.  1997).  Additionally,  our  results 
suggest  that  these  species  are  associated  with 
certain  regions  within  the  Prairie  Ecozone. 
Sprague’s  Pipits  and  Chestnut-collared  Long- 
spurs  occurred  mostly  in  the  drier  prairies  of 
the  southern  portion  of  the  province,  but 
Chestnut-collared  Longspurs  were  virtually 
absent  from  the  cypress  upland  ecoregion  de- 
spite the  presence  of  large  tracts  of  native 
grassland.  Grassland  habitat  in  this  region  is 
taller  and  denser  than  the  surrounding  grass- 
lands (W.  Harris,  pers.  comm.)  and  thus  may 
not  provide  Chestnut-collared  Longspurs  with 
the  short,  sparse  vegetative  cover  they  require 
(Harris  1944,  Owens  and  Myres  1973,  Dale 
1983).  In  contrast  to  Sprague’s  Pipits  and 
Chestnut-collared  Longspurs,  Baird’s  Spar- 
rows occurred  most  frequently  in  the  more 
mesic  grasslands  of  the  moist-mixed  grassland 
ecoregion,  consistent  with  their  overall  pref- 
erence for  taller  and  denser  vegetation  (Dale 
1983,  Winter  1994,  Madden  1996,  Sutter  and 
Brigham  1998,  Davis  and  Duncan  in  press). 
Furthermore,  our  study  indicates  that  the  Mis- 
souri Coteau,  which  borders  the  mixed  and 
moist-mixed  grassland  ecoregions,  is  an  im- 
portant landform  for  Baird’s  Sparrows  in  Sas- 
katchewan (McMaster  and  Davis,  upubl. 
data).  This  landform  has  also  been  identified 
as  a significant  region  for  Baird  s Sparrows  in 
North  Dakota  (Stewart  1975). 

Grazing  by  livestock  can  have  a profound 
influence  on  the  structure  of  rangeland  vege- 
tation (Wiens  and  Dyer  1975,  Ryder  1980). 
Because  habitat  selection  by  grassland  birds  is 
likely  influenced  by  habitat  structure  (Wiens 
1969),  grazing  intensity  should  affect  the  oc- 
currence of  the  three  songbird  species  exam- 
ined in  this  study.  Sprague’s  Pipits,  and  to  a 
lesser  degree,  Baird’s  Sparrows,  were  both  in- 


fluenced by  grazing  intensity.  Light  to  mod- 
erately grazed  pasture  has  been  described  as 
preferred  habitat  for  both  Sprague  s Pipits  and 
Baird’s  Sparrows  (Kantrud  1981;  Kantrud  and 
Kologiski  1982,  1983;  Knopf  1996)  although 
others  have  suggested  that  these  species  prefer 
idle  prairie  (Maher  1973,  Owens  and  Myres 
1973,  Dale  1984).  The  latter  studies,  however, 
examined  only  a small  number  of  grazed  and 
ungrazed  sites  and  did  not  discriminate  be- 
tween lightly  or  heavily  grazed  pastures.  We 
could  not  assess  the  suitability  of  ungrazed 
prairie  to  either  species  in  our  study  because 
this  habitat  type  is  uncommon  in  Saskatche- 
wan. Despite  the  lack  of  information  on  the 
response  of  Sprague’s  Pipits  and  Baird  s Spar- 
rows to  low  intensity  grazing  relative  to  un- 
grazed native  prairie,  our  results  suggest  that 
low  levels  of  grazing  intensity  are  tolerated  by 
these  species. 

We  found  no  difference  in  the  response  of 
Chestnut-collared  Longspurs  to  grazing  inten- 
sity in  native  pasture.  Numerous  studies  have 
reported  that  Chestnut-collared  Longspurs  re- 
spond positively  to  grazing  (Maher  1973,  Dale 
1983,  Kantrud  and  Kologiski  1983,  Renken 
and  Dinsmore  1987,  Bock  et  al.  1993).  Graz- 
ing may  not  have  influenced  Chestnut-collared 
Longspurs  in  our  study  because  the  structure 
of  the  vegetation  under  all  grazing  intensities 
may  have  fallen  within  an  acceptable  range. 
Descriptions  of  Chestnut-collared  Longspur 
breeding  habitats  have  ranged  from  over- 
grazed  pastures  with  sparse  vegetation  to  sit- 
uations where  the  “.  . .thicker  and  taller  grass- 
es afford  adequate  concealment”  (DuBois 
1935:70).  Chestnut-collared  Longspurs  in  our 
study  were  generally  associated  with  sparsely 
vegetated  native  pastures  with  low  plant  litter 
depths  (see  also  Harris  1944,  Owens  and  My- 
res 1973,  Dale  1983,  Johnson  and  Schwartz 
1993). 

Conservation. — While  continental  popula- 
tions of  Baird’s  Sparrows  and  Chestnut-col- 
lared Longspurs  appear  to  be  relatively  stable 
(Sauer  et  al.  1997),  Sprague’s  Pipits  are  cur- 
rently undergoing  significant  population  de- 
clines of  4.7%  per  year,  one  of  the  steepest 
declines  recorded  for  grassland  songbirds  in 
North  America.  The  conversion  of  native 
grassland  to  annually  cropped  land,  and  the 
pattern  of  habitat  loss  (i.e.,  habitat  fragmen- 
tation) have  likely  played  significant  roles  in 


Dcivis  ei  al.  • ENDEMIC  GRASSLAND  SONGBIRDS 


395 


these  declines  (Davis,  unpubl.  data).  For  ex- 
ample, Sprague’s  Pipits  reach  their  highest 
densities  in  southeastern  Alberta  and  south- 
western Saskatchewan  (Sauer  et  al.  1997),  ar- 
eas characterized  by  large  tracts  of  contiguous 
native  grassland  (South  Digital  Land  Cover 
Project,  unpubl.  data).  While  land-use  pro- 
grams that  convert  annually  tilled  cropland  to 
perennial  cover  will  likely  provide  additional 
habitat  for  endemic  grassland  birds  (Johnson 
and  Schwartz  1993,  Reynolds  et  al.  1994,  Sut- 
ter and  Brigham  1998,  Davis  and  Duncan  in 
press),  it  is  imperative  that  the  reproductive 
consequences  of  selecting  alternative  nesting 
habitats  be  determined  to  accurately  assess 
habitat  quality  (Johnson  and  Temple  1986, 
Van  Horne  1983,  Vickery  et  al.  1992). 

ACKNOWLEDGMENTS 

We  thank  G.  Butcher,  M.  Hartley,  B.  Dale,  D.  Hjer- 
taas,  D.  H.  Johnson,  and  D.  McKinnon  for  their  helpful 
advice  on  the  initial  design  of  the  study.  We  are  grate- 
ful to  our  held  crew,  L.  Banman,  C.  Bjorklund,  R Hjer- 
taas,  R.  Kreba,  J.  Pollock,  and  T.  Troupe.  Thanks  to  J. 
Keith  (Saskatchewan  Conservation  Data  Center)  for 
assistance  in  the  production  of  maps  and  T.  Hairison 
(Saskatchewan  Wetland  Conservation  Corporation)  for 
his  expertise  in  rangeland  ecology.  Comments  by  S.  L. 
Jones,  D.  G.  McMaster,  G.  C.  Sutter,  and  an  anony- 
mous reviewer  greatly  improved  the  manuscript.  This 
study  would  not  have  been  possible  without  the  co- 
operation of  private  landowners  and  the  staff  and  pa- 
trons of  the  Prairie  Farm  Rehabilitation  Administration 
pastures.  Provincial  Community  pastures,  and  Provin- 
cial Co-op  pa.sture.s.  Financial  support  for  this  study 
was  provided  by  the  National  Fish  and  Wildlife  Foun- 
dation (U.S.),  Endangered  Species  Recovery  Fund 
(World  Wildlife  Fund  Canada  and  Canadian  Wildlife 
Service  of  Environment  Canada),  and  Saskatchewan 
Endangered  Species  Fund  (Saskatchewan  Environment 
and  Resource  Management). 

LITERATURE  CITED 

Arnold,  T.  W.  and  K.  F.  Higgins.  1986.  Effects  of 
shrub  coverages  on  birds  of  North  Dakota  mixed- 
grass  prairies.  Can.  Field-Nat.  100:10-14. 

Bock,  C.  E.,  V.  A.  Saab,  T,  D.  Rich,  and  D.  S.  Dob- 
kin.  1993.  Effects  of  livestock  grazing  on  Neo- 
tropical migratory  landbirds  in  western  North 
America.  Pp.  296—309  in  Status  and  management 
of  Neotropical  migratory  birds  (D.  M.  Finch  and 
P.  W.  Stangel,  Eds.).  U.S.  For.  Serv.  Gen.  Tech. 
Rep.  RM-229,  Fort  Collins,  Colorado. 

Bollinger,  E.  K.,  P.  B.  Bollinger,  and  T.  A.  Gavin. 
1990.  Effects  of  hay-cropping  on  ea.stern  popula- 
tions of  the  Bobolink.  Wildl.  Soc.  Bull.  18:142- 
150. 


Cartwrigh'i,  B.  W.,  T.  M.  Siiorit,  and  R.  D.  Harris. 
1937.  Baird’s  Sparrow.  Trans.  R.  Can.  Inst.  46: 
153-198. 

Cody,  M.  L.  1968,  On  the  methods  of  resource  divi- 
sion in  grassland  bird  communities.  Am.  Nat.  102: 
107-147. 

Dale,  B.  C.  1983.  Habitat  relationships  of  seven  spe- 
cies of  passerine  birds  at  Last  Mountain  Lake, 
Saskatchewan.  M.Sc.  thesis,  Univ.  of  Regina,  Re- 
gina, Saskatchewan. 

Dale,  B.  C.  1984.  Birds  of  grazed  and  ungrazed  grass- 
lands in  Saskatchewan.  Blue  Jay  42:102-105. 

Dale,  B.  C.,  P.  A.  Martin,  and  P.  S.  Taylor.  1997. 
Effects  of  hay  management  on  grassland  song- 
birds in  Saskatchewan.  Wildl.  Soc.  Bull.  25:616- 
626. 

Davis,  S.  K.,  D.  C.  Duncan,  and  M.  Skeel.  1996.  The 
Baird’s  SpaiTow:  status  re.solved.  Blue  Jay  54: 
185-191. 

Davis,  S.  K.  and  D.  C.  Duncan.  In  press.  Grassland 
songbird  occurrence  in  native  and  crested  wheat- 
grass  pastures  of  southern  Saskatchewan.  Stud. 
Avian  Biol. 

Dubois,  A.  D.  1935.  Nests  of  Horned  Larks  and  long- 
spurs  on  a Montana  prairie.  Condor  37:56-72. 

Ecological  Stratification  Working  Group.  1995.  A 
national  ecological  framework  for  Canada.  Agri- 
culture and  Agri-Food  Canada  and  Environment 
Canada,  Ottawa. 

Fairfield,  G.  M.  1968.  Chestnut-collared  Longspun 
U.S.  Nat.  Mus.  Bull.  237:1635-1652. 

Frawley,  B.  j.  1989.  The  dynamics  of  nongame  bird 
breeding  ecology  in  Iowa  alfalfa  fields.  M.S.  the- 
sis, Iowa  State  Univ.,  Ames. 

Goossen,  j.  R,  S.  Brechtel,  K.  D.  De  Smet,  D.  Hjer- 
TAAS,  AND  C.  Wershler.  1993.  Canadian  Baird’s 
Sparrow  recovery  plan.  Recovery  of  Nationally 
Endangered  Wildlife  Rep.  No.  3.  Canadian  Wild- 
life Federation,  Ottawa. 

Harris,  R.  D.  1944.  The  Chestnut-collared  Longspur 
in  Manitoba.  Wilson  Bull.  56:105-115. 

Hartley,  M.  J.  1994.  Passerine  abundance  and  pro- 
ductivity indices  in  grasslands  managed  for  wa- 
terfowl nesting  cover.  Trans.  Am.  Wildl.  Nat.  Re- 
sour. Conf.  59:322-327. 

Johnson,  D.  H.  and  M.  D.  Schwartz.  1993.  The  con- 
servation reserve  program:  habitat  for  grassland 
birds.  Great  Plains  Res.  3:273-295. 

Johnson,  R.  G.  and  S.  A.  Temple.  1986.  Assessing 
habitat  quality  for  birds  nesting  in  fragmented  tail- 
grass  prairie.  Pp.  245-249  in  Wildlife  2000.  Mod- 
elling habitat  relationships  of  tenestrial  verte- 
brates (J.  'Verner,  M.  L.  Morrison,  and  C.  J.  Ralph, 
Eds.).  Univ.  of  Wisconsin  Press,  Madison. 

Kantrud,  H.  a.  1981.  Grazing  intensity  effects  on  the 
breeding  avifauna  of  North  Dakota  native  grass- 
lands. Can.  Field-Nat.  95:404-417. 

Kantrud,  H.  A.  and  R.  L.  Kologiski.  1982.  Effects 
of  soils  and  grazing  on  breeding  birds  of  uncul- 
tivated upland  grasslands  of  the  northern  great 


396 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3.  September  1999 


plains.  U.S.  Eish  and  Wildlife  Service.  Washing- 
ton, D.C. 

Kantrud,  H.  a.  and  R.  L.  Kologiski.  1983.  Avian 
associations  of  the  northern  Great  Plains  grass- 
lands. J.  Biogeogr.  10:331—350. 

Knopf,  E L.  1994.  Avian  assemblages  on  altered  grass- 
lands. Stud.  Avian  Biol.  15:247-257. 

Knopf,  E L.  1996.  Prairie  legacies-birds.  Pp.  135—148 
in  Prairie  conservation.  Preserving  North  Ameri- 
ca's most  endangered  ecosystem  (E  B.  Samson 
and  E L.  Knopf,  Eds.).  Island  Press,  Washington, 
D.C. 

Madden,  E.  M.  1996.  Passerine  communities  and  bird- 
habitat  relationships  on  prescribed-burned,  mixed- 
grass  prairie  in  North  Dakota.  M.S.  thesis,  Mon- 
tana State  Univ.,  Bozeman. 

Maher,  W.  J.  1973.  Birds:  I.  Population  dynamics. 
Matador  project  technical  report  No.  34.  Canadian 
Committee  for  the  International  Biological  Pro- 
gramme, Saskatoon,  Saskatchewan. 

Mahon,  C.  L.  1995.  Habitat  selection  and  detectability 
of  Baird's  Sparrows  in  southwestern  Alberta. 
M.Sc.  thesis,  Univ.  of  Alberta,  Edmonton. 

Owens,  R.  A.  and  M.  T.  Myres.  1973.  Effects  of  ag- 
riculture upon  populations  of  native  passerine 
birds  of  Alberta  fescue  grassland.  Can.  J.  Zool. 
51:697-713. 

Pulliam,  H.  R.  1988.  Sources,  sinks,  and  population 
regulation.  Am.  Nat.  132:652—661. 

Ralph,  C.  J.,  G.  R.  Geupel,  P.  Pyle,  T.  E.  Martin, 
and  D.  E Desante.  1993.  Handbook  of  field 
methods  for  monitoring  landbirds.  Gen.  Tech. 
Rep.  PSW-GTR-144.  Pacific  Southwest  Research 
Station,  Albany,  California. 

Renken,  R.  B.  and  j.  j.  Dinsmore.  1987.  Nongame 
bird  communities  on  managed  grasslands  in  North 
Dakota.  Can.  Field-Nat.  101:551—557. 

Reynolds,  R.  E.,  T.  L.  Shaffer,  J.  R.  Sauer,  and  B. 
G.  Peterjohn.  1994.  Conservation  Reserve  Pro- 
gram: benefit  for  grassland  birds  in  the  noithern 
plains.  Trans.  Am.  Wildl.  Nat.  Resour.  Conf.  59: 
328-336. 

Riemer,  G.,  T.  Harrison,  L.  H.all,  and  N.  Lynn.  1997. 
The  native  prairie  stewardship  program.  Pp.  111- 
1 16  in  Caring  for  the  home  place:  protected  areas 
and  landscape  ecology  (P.  Jonker,  J.  Vandall,  L. 
Ba.schak,  and  D.  Gauthier.  Eds.).  Univ.  Extension 
Press.  Univ.  of  Saskatchewan,  Saskatoon. 

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. 

Ryder,  R.  A.  1980.  Effects  of  grazing  on  bird  habitat. 
Pp  51-66  in  Management  of  western  forests  and 
grasslands  for  nongame  birds  (R.  M.  DeGrafl  and 
N.  G.  Tilghman.  Eds.).  U.S.  For.  Serv.  Gen.  Tech. 
Rep.,  Ogden,  Utah. 


Samson,  E B.  and  E L.  Knopf.  1994.  Prairie  conser- 
vation in  North  America.  Bioscience  44:418-421. 

SAS  Institute  Inc.  1989.  SAS/STAT  user's  guide.  Re- 
lease 6.12  ed.  SAS  Inst.,  Inc.,  Cary,  North  Caro- 
lina. 

Sauer,  J.  R.,  J.  E.  Hines,  G.  Gough,  I.  Thomas,  and 
B.  G.  Peterjohn.  1997.  The  North  American 
breeding  bird  survey  results  and  analysis.  Version 
96.4.  Patuxent  Wildlife  Research  Center,  Laurel, 
Maryland.  URL  = http://www.mbr.nbs.gov/bbs/ 
bbs.html 

Smith,  A.  R.  1996.  Atlas  of  Saskatchewan  birds.  Sas- 
katchewan Natural  History  Society,  Regina,  Sas- 
katchewan. 

Stewart,  R.  E.  1975.  Breeding  birds  of  North  Dakota. 
Tri-college  Center  for  Environmental  Studies,  Far- 
go, North  Dakota. 

Stewart,  R.  E.  and  H.  A.  Kantrud.  1972.  Population 
estimates  of  breeding  birds  in  North  Dakota.  Auk 
89:766-788. 

Sutter,  G.  C.  1996.  Habitat  selection  and  prairie 
drought  in  relation  to  grassland  bird  community 
structure  and  the  nesting  ecology  of  Sprague  s 
Pipit,  Anthii.s  spragueii.  Ph.D.  diss.,  Univ.  of  Re- 
gina, Regina,  Saskatchewan. 

Sutter,  G.  C.  and  M.  Brigham.  1998.  Avifaunal  and 
habitat  changes  resulting  from  conversion  of  na- 
tive prairie  to  crested  wheat  grass:  patterns  at 
songbird  community  and  species  levels.  Can.  J. 
Zool.  76:869-875. 

Sutter,  G.  C.,  S.  K.  Davis,  and  D.  C.  Duncan.  In 
press.  Grassland  songbird  abundance  along  roads 
and  trails  in  southern  Saskatchewan.  J.  Field  Or- 
nithol. 

Van  Horne,  B.  1983.  Density  as  a misleading  indi- 
cator of  habitat  quality.  J.  Wildl.  Manage.  47:893- 
901. 

Vickery,  P.  D.,  M.  L.  Hunter,  Jr.,  and  J.  V.  Wells. 
1992.  Use  of  a new  reproductive  index  to  evaluate 
relationship  between  habitat  quality  and  breeding 
success.  Auk  109:697—705. 

Whitmore.  R.  C.  1979.  Short-term  changes  in  vege- 
tation structure  and  its  effects  on  Grasshopper 
SpaiTows  in  West  Virginia.  Auk  96:621—625. 

Wiens,  J.  A.  1969.  An  approach  to  the  study  of  eco- 
logical relationships  among  grassland  birds.  Or- 
nithol.  Monogr.  8:1-93. 

Wiens,  J.  A.  and  M.  I.  Dyer.  1975.  Rangeland  avi- 
faunas: their  composition,  energetics,  and  role  in 
the  ecosystem.  Pp.  146—182  in  Symposium  on 
management  of  forest  and  range  habitats  for  non- 
game birds  (D.  R.  Smith,  Ed.).  U.S.  For.  Serv. 
Gen.  Tech.  Rep.  WO-L,  Tucson,  Arizona. 

Wilson.  S.  D.  and  J.  W.  Belcher.  1989.  Plant  and 
bird  communities  of  native  prairie  and  introduced 
Eurasian  vegetation  in  Manitoba,  Canada.  Con- 
serv.  Biol.  3:39—44. 

Winter,  M.  1994.  Habitat  selection  of  Baird's  Spar- 
rows in  northern  mixed-grass  prairie.  Diplomar- 
beit,  Univ.  of  Tubingen,  Tubingen,  Germany. 


Wilson  Bull..  1 1 1(3),  1999,  pp.  397-414 


BIRD  COMMUNITIES  IN  NATURAL  FOREST  PATCHES  IN 

SOUTHERN  BRAZIL 

LUIZ  DOS  ANJOS'  3 AND  ROBERTO  BO^ON^ 


ABSTRACT. — Avifaimal  composition  was  evaluated  for  natural  (not  artificial)  patches  of  mixed  temperate 
rain  forest  in  the  Campos  Gerais  region,  Parana  State,  southern  Brazil.  A large  patch  (840  ha)  and  1 1 smaller 
patches  (0.5-40  ha)  were  censused  from  September  to  December  of  1995  (five  hours  per  month,  each  site).  The 
total  species  number  was  strongly  conelated  with  patch  size  (r  = 0.92,  P < 0.001).  However,  the  number  of 
edge  species  increased  with  decreasing  patch  area;  the  opposite  happened  with  forest  species.  Thus,  the  ratio  of 
edge  to  forest  species  increased  with  decreasing  patch  area.  The  number  of  leaf  insectivore  species  decreased 
the  most  with  a decrease  in  area.  The  mean  Simpson  similarity  index  was  73.8%  among  forest  patches  of  similar 
size.  Smaller  forest  patches  linked  to  the  840  ha  patch  were  more  similar  to  this  larger  patch  than  isolated 
patches.  Point  counts  from  January  to  December  1991  in  four  patches  (72  points  each  area)  showed  that  several 
species,  specially  trunk  (and  twig)  insectivores  and  omnivores,  increa.sed  in  relative  abundance  with  decreases 
in  area  of  the  patch  (density  compensation).  The  “habitat  appropriation”  hypothesis,  the  expansion  of  niches  to 
include  slightly  different  habitats,  could  explain  the  increased  relative  abundance  of  two  trunk  (and  twig)  insec- 
tivores: Cranioleuca  ohsoleta  and  Cranioleuco  pallida.  Received  15  Oct.  1998,  accepted  19  Feb.  1999. 


Forest  fragmentation  in  the  Neotropical  re- 
gion has  been  considered  an  important  force 
in  the  loss  of  biodiversity  (Bierregaard  and 
Lovejoy  1989).  Decreases  in  the  number  of 
bird  species  and  changing  avifaunal  compo- 
sition have  been  documented  by  many  work- 
ers (Willis  1974,  1979;  Karr  1982;  Bierre- 
gaard 1990;  Anjos  1992;  Aleixo  and  Vielliard 
1995). 

Density  compensation  (increased  relative 
abundance  of  species  in  biologically  isolated 
habitats)  is  another  frequent  feature  in  habitat 
fragments  (Mac Arthur  et  al.  1972,  Wright 
1980).  Reduced  competition,  habitat  differ- 
ences, and  differences  in  colonization  (Rick- 
lefs  and  Cox  1978,  Blondel  1991)  also  have 
been  suggested  as  possible  reasons  for  density 
compensation  in  habitat  fragments.  Blondel 
and  coworkers  (1988)  presented  the  “habitat 
appropriation”  hypothesis  in  which  popula- 
tions increase  on  islands  because  of  expansion 
into  additional  habitats. 

Studies  of  birds  in  Neotropical  forest  frag- 
ments have  been  carried  out  after  fragmenta- 
tion took  place  through  logging.  The  state  of 
Parana,  southern  Brazil,  has  a grassland  re- 


‘ Univ.  Estadual  de  Londrina,  Depto.  de  Biol.  Ani- 
mal e Vegetal,  Caixa  Postal  6001,  Londrina  86051- 
970,  Parana,  Brazil;  E-mail:  llanjo.s@sercomtel.com.br 
^ Socied.  de  Pesquisa  em  Vida  Selvagem  e Educagao 
Ambiental,  Rua  Gutemberg  345,  Curitiba  80420-030, 
Parana,  Brazil. 

Corresponding  author. 


gion  called  “Campos  Gerais”  with  isolated 
forest  patches  of  0.1-100  ha  (Fig.  1).  This 
landscape  is  natural;  it  was  not  deforested  by 
humans.  The  forest  patches  appear  in  areas  of 
suitable  soil  conditions  surrounded  by  grass- 
land and  are  slowly  increasing  in  area  through 
a natural  ecological  succession  (Klein  1960, 
1972).  The  forest  structure  among  different 
patch  series  is  similar  (Klein  1960,  Maack 
1981,  Klein  and  Hatschbach  1971).  Because 
fragmentation  has  been  natural,  the  biological 
processes  may  not  have  been  affected  by  hu- 
man activity.  Our  goal  in  this  study  was  to 
analyze  the  differences  in  the  composition  of 
the  bird  community  among  various-sized  for- 
est patches  in  Campos  Gerais  region. 

STUDY  AREA  AND  METHODS 

Study  sites. — The  study  areas  are  the  Fazenda  Santa 
Rita  (25°  15'  S,  49°  48'  W)  and  Vila  Velha  State  Park 
(25°  15'  S,  49°  55'  W),  Parana  State,  southern  Brazil, 
in  a region  called  Campos  Gerais  (Fig.  2).  The  patches 
of  forest  are  mixed  temperate  rainforest.  The  dominant 
trees  are:  Araucaria  angustifolia  ( Araucariaceae),  Po- 
docarpus  lainhertii  (Podocaipaceae),  Sehastiana  coni- 
tnersoniana  (Euphorbiaceae),  Ocotea  porosa,  and  Nec- 
tandra  grandijiora  (both  Lauraceae;  Klein  and  Hatsch- 
bach 1971).  The  average  annual  temperature  is 
17.3°  C,  varying  from  20.9°  C during  summer  (Decem- 
ber-February)  and  14.I°C  during  winter  (June-Au- 
gust). The  range  of  average  annual  precipitation  is 
l5()0-200()  mm.  Elevations  range  from  950-1100  m 
a.s.l.  (Maack  1981). 

Eleven  forest  patches  (called  B,  C,  D,  E,  F,  G,  H,  I, 
J,  L and  M)  from  0.5-40  ha  in  size  and  one  large  forest 
patch  of  840  ha  (called  A)  were  censused.  All  these 


397 


398 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


LIG.  1.  Landscape  in  the  “Campos  Gerais”  region,  southern  Brazil;  grassland  with  naturally  isolated  forest 
patches. 


sites  are  covered  by  similar  forest  structure.  Some 
patches  are  linked  (patches  B,  C,  D,  and  E)  to  the 
largest  forest  patch  (A)  by  forest  corridors  while  others 
are  completely  isolated  by  grassland  (Table  1 ). 

Species  composition. — Bird  species  in  each  patch, 
micro-habitat,  and  feeding  habit  were  recorded  during 
five-hour  monthly  visits  in  the  morning  (20  h total) 
September-December,  1995.  Each  species  was  classi- 
fied as  to  its  typical  habitat  and  feeding  habit  through 
field  observations  and  references  (Eitzpatrick  1980; 
Belton  1984,  1985;  Sick  1997;  Ridgely  and  Tudor 
1989,  1994).  The  habitats  were  categorized  as  open 
area,  forest  border  (up  to  5 m from  the  edge),  and 
forest  interior.  We  divided  the  forest  into  three  levels; 
(1)  understory  (below  2 m),  (2)  mid-levels  (2-7  m) 
and  (3)  canopy  (above  7 m).  Feeding  habits  were  cat- 
egorized as;  omnivore,  insectivore,  Irugivore,  carni- 
vore or  nectarivore.  For  insectivores,  the  area  where 
the  insect  or  its  larvae  was  most  often  captured  was 
also  recorded.  Three  classes  of  capture  sites  were  de- 
fined: (1)  trunk  (and  twigs),  (2)  leaves,  and  (3)  gen- 
eralized insectivores. 

Relative  abundance. — Relative  abundance  was  de- 
termined by  monthly  point  counts  of  unlimited  dis- 
tance (Blondel  et  al.  1970)  January-December  in 
1991.  Fifty-three  points  were  used  to  calculate  relative 
abundance:  24  in  A,  12  in  B,  10  in  C.  and  7 in  G. 
Each  point  was  100  m from  another  and  at  least  50  m 
from  the  edge  of  the  forest.  Each  month  six  points  in 
each  site  were  chosen  at  random  to  be  sampled  giving 


a total  of  72  counts  per  site  during  the  year  (288  point 
counts  total  in  the  four  sites).  The  relative  abundance 
of  each  species  was  determined  by  dividing  the  total 
number  of  species  contacts  by  the  total  number  of 
points  (72)  sampled  in  each  patch  (Blondel  et  al. 
1970).  Sampling  began  at  dawn  at  the  first  randomly 
selected  point  and  finished  about  3.5  h later  at  the  sixth 
point.  The  time  for  sampling  at  each  point  was  20  min. 
Snecies  were  identified  primarily  by  sound  (99%).  The 
same  observer  (LdA)  performed  all  counts.  Each  pair 
or  flock  of  each  species  was  counted  once  (one  con- 
tact) while  vocalizing.  Precautions  were  taken  not  to 
count  the  same  individual  or  group  more  than  once  (a 
form  was  used  in  order  to  locate  the  counted  individ- 
uals), especially  highly  mobile  species.  Bird  recordings 
(1601  recordings  of  414  species,  deposited  in  the  Bio- 
acoustic Laboratory  of  the  Universidade  Estadual  de 
Londrina)  were  used  to  aid  identification.  The  observer 
had  six  years  (1984-1990)  of  field  experience  in  spe- 
cies identification  in  the  region  (Anjos  1992,  Anjos 
and  Graf  1993,  Anjos  et  al.  1997). 

Atudxses. — The  birds  were  identified  mostly  using 
the  taxonomy  of  Meyer  de  Schauensee  (1982)  and 
Sick  (1997).  Similarities  in  bird  species  composition 
between  forest  patches  were  determined  using  the 
Simpson  Index.  3 diversity  (Whittaker  1960)  was  used 
to  measure  the  degree  of  turnover  in  species  compo- 
sition along  the  six  points  sampled  monthly  (in  A and 
G).  This  index  is  a measure  of  how  different  the  sam- 
ples were,  in  terms  of  the  variety  of  species.  Differ- 


Anjos  ami  Bo<^on  • BIRD  COMMUNITIES  IN  FOREST  PATCHES 


399 


FIG.  2.  Study  area  in  “Campos  Gerais”  region,  Parana  State,  southern  Brazil,  indicated  by  hatching. 


ences  in  numbers  of  species  or  numbers  of  contacts  of 
species  between  the  sites  were  tested  with  X"  analysis 
at  a < 0.05.  Analysis  of  Variance  (one  way  ANOVA) 
and  the  Tukey  multiple  range  test  (P  < 0.05)  were  used 
to  evaluate  the  average  number  of  species  and  contacts 
recorded  per  sampled  point  in  A,  B,  C and  G.  The 
relationship  between  number  of  bird  species  and  size 
of  fragment  was  expressed  through  a transformation 
following  Preston  (1962):  log^  = log^  + zlogA  where 


TABLE  1.  Size  (ha)  and  on  the  distance  (m)  from 
a continuous  forest  of  each  study  site. 

Site 

Size 

Type-Di.stance 

A 

840  ha 

continuous  forest 

B 

40  ha 

linked  by  forest  corridor- 100  m 

C 

20  ha 

linked  by  forest  corridor-200  m 

D 

12  ha 

linked  by  forest  corridor-600  m 

E 

10.5  ha 

linked  by  forest  corridor-4()0  m 

F 

10  ha 

isolated-3000  m 

G 

9 ha 

isolated-2000  m 

H 

8.5  ha 

isolated-800  m 

I 

6.5  ha 

isolated-2500  m 

J 

4 ha 

isolated- 1500  m 

L 

1.5  ha 

isolated-500  m 

M 

0.5  ha 

isolated- 1000  m 

S is  the  number  of  species  and  A is  the  size  of  the  area. 
SAS/STAT  (version  6.11,  IBM,  mainframe)  was  used 
to  calculate  most  of  the  above  values. 

RESULTS 

Habitat  distribution. — Birds  in  all  patches 
totaled  189  species,  including  13  open  area 
species,  51  edge  species,  and  125  forest  spe- 
cies (Appendix).  The  open  area  birds  used  the 
forest  only  for  roosting  and/or  for  nesting  but 
not  for  foraging. 

There  were  13  edge  species  in  the  largest 
patch  (A)  and  32  species  in  the  6.5  ha  frag- 
ment (1).  By  contrast,  there  were  93  forest  spe- 
cies in  the  largest  patch  and  only  13  species 
in  the  smallest  fragment  (M;  Table  2).  These 
results  indicate  a general  decrease  in  the  ratio 
of  forest  to  edge  species  with  a decrease  in 
area  (Fig.  3).  The  exceptions  are  the  patches 
E (34/21  species),  G (41/28  species),  and  L 
(32/19  species).  The  number  of  understory 
forest  species  decreased  faster  {y}  = 15.37,  df 
= 2,  P < 0.05)  with  decrease  in  area  of  forest 
(33  species  in  A,  20  species  in  B,  and  8 spe- 


400 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


TABLE  2.  Number  of  bird  species  in  the  study  sites  accordin; 

g their  habitats  and  feeding 

habits. 

Sites 

A 

B 

c 

D 

E 

F 

G 

H 

I 

j 

L 

M 

Habitats 

Open  area 

2 

2 

3 

3 

6 

2 

4 

4 

3 

3 

1 

1 

Edae 

13 

22 

26 

29 

21 

27 

28 

29 

32 

30 

19 

30 

Forest/understory 

33 

20 

8 

10 

5 

8 

8 

4 

10 

4 

7 

3 

Forest/mid-levels 

32 

29 

27 

22 

20 

17 

23 

18 

16 

15 

17 

6 

Forest/canopy 

28 

19 

16 

12 

9 

5 

10 

7 

7 

7 

8 

4 

Feeding  Habits 

Omnivores 

29 

32 

28 

31 

22 

19 

27 

28 

23 

24 

21 

16 

Insectivores/trunk 

18 

14 

12 

10 

10 

7 

10 

6 

3 

4 

6 

3 

Insectivores/leaves 

30 

16 

16 

13 

9 

1 1 

15 

10 

1 1 

10 

8 

7 

Insectivores/general 

1 1 

12 

8 

10 

9 

12 

12 

7 

16 

10 

8 

1 1 

Frugivores 

15 

12 

12 

7 

8 

6 

5 

6 

8 

8 

8 

5 

1 

Carnivores 

3 

3 

2 

4 

3 

3 

3 

4 

4 

1 

Nectarivores 

2 

3 

2 

1 

0 

1 

1 

1 

3 

1 

0 

Total  Species 

108 

92 

80 

76 

61 

59 

73 

62 

68 

59 

52 

45 

cies  in  C)  than  mid-level  forest  species  (32 
species  in  A,  29  species  in  B,  and  27  species 
in  C;  X-  = 0-43,  df  = 2,  F > 0.05);  canopy 
forest  species  were  intermediate  in  the  ratio  of 
decreasing  (28  species  in  A,  19  species  in  B, 
and  16  species  in  C;  “ 3.71,  df  = 2,  P > 
0.05;  Table  2). 

Similarity  between  sites. — The  similarity 
index  (Simpson)  between  the  largest  site  (A: 
108  species)  and  the  smallest  (M;  45  species) 
was  40%  (17  species  common  to  both  sites); 
however,  the  mean  was  73.8  ± 2.7%  (SE) 
among  forest  patches  of  similar  size. 

The  number  of  species  was  correlated  with 
area  (Fig.  4;  r = 0.92,  df  = 10,  P < 0.001). 
Species  living  in  open  areas  were  not  consid- 


FIG. 3.  Ratio  of  forest  (solid)  and  edge  (hatched) 
species  in  the  study  sites.  Sites  B-E  linked  to  A by 
forest  corridors. 


ered  in  this  analysis;  edge  species  were  in- 
cluded with  forest  species  because  the  major- 
ity of  them  were  commonly  found  within  the 
forest.  The  number  of  species  can  be  predicted 
according  to  the  Preston  (1962)  model;  \ogS 
= 3.81  + 0.17  logA,  where  S is  the  number 
of  species  and  A is  the  fragmented  area  in  ha. 

Feeding  habits. — Of  the  189  species  re- 
corded, 85  were  insectivores,  58  omnivores, 
28  frugivores,  12  carnivores,  and  6 nectari- 
vores.  Among  the  insectivores,  33  were  leaf 
insectivores,  3 1 generalized  insectivores  (cap- 
ture insects  in  various  ways),  and  21  trunk 
insectivores. 

Insectivores  were  less  abundant  in  B (40  ha; 


FIG.  4.  Correlation  (log/log)  between  number  of 
species  (S)  and  area  (A)  of  fragment.  Dotted  lines  in- 
dicate 95%  confidence  interval. 


Anjos  am!  lio^on  • BIRD  COMMUNITIES  IN  FOREST  PATCHES 


401 


42  species)  than  in  A (840  ha;  59  species) 
although  not  significantly  so  (x"  = 2.86,  df  = 
I,  P > 0.05).  In  patches  smaller  than  40  ha 
the  rate  of  decrease  slowed  as  patch  size  di- 
minished (36  species/20  ha,  33  species/12  ha, 
28  species/10.5  ha);  in  patches  smaller  than 
10.5  ha,  there  was  no  clear  pattern  (Table  2). 
The  greatest  reduction  in  species  occurred  in 
leaf  insectivores,  from  30  species  in  the  forest 
(A)  to  16  species  in  the  largest  patch  (B;  x“ 
= 4.26,  df  = 1,  P < 0.05);  the  loss  of  trunk 
insectivores  was  not  so  great  (18  species  in  A 
and  14  species  in  B;  x‘  = 0.5,  df  = 1,  P > 
0.05)  while  generalized  insectivores  gained 
one  species  (11  species  in  A and  12  species 
in  B). 

Omnivores  were  represented  by  two  more 
species  in  B (32  species)  than  in  A (30  spe- 
cies). Their  numbers  remained  relatively  con- 
stant down  through  the  12  ha  patch  (D;  31 
species)  but  decreased  in  smaller  patches  (Ta- 
ble 2). 

Frugivores  decreased  slightly  in  number  (x^ 
= 0.46,  df  = 2,  P < 0.05)  from  A (15  species) 
to  B and  C (both  with  12  species).  The  num- 
ber of  frugivores  stayed  between  5 and  8 spe- 
cies in  patches  smaller  than  12  ha  (Table  2). 

Carnivores  and  nectarivores  were  repre- 
sented by  few  species  in  all  sites  (1-4  species; 
Table  2). 

Number  of  species  and  contacts. — In  point 
counts  conducted  in  1991  we  recorded  138 
species  in  A,  125  in  B,  103  in  C,  and  91  in 
G.  The  average  numbers  of  species  per  sam- 
pled point  were;  23.6  in  A,  16.5  in  B,  18.5  in 
C,  and  21.4  in  G.  The  number  of  species  per 
sampled  point  was  not  significantly  different 
(Tukey  test:  P > 0.05)  between  A and  G al- 
though they  are  different  in  the  size  and  total 
number  of  species  recorded.  The  intermediate- 
sized patches  B (40  ha)  and  C (20  ha)  did  not 
show  significant  differences  (Tukey  test:  P > 
0.05)  among  the  numbers  of  species  per  sam- 
pled point,  but  both  were  significantly  differ- 
ent from  A and  G (Tukey  test:  P < 0.05). 

The  average  numbers  of  contacts  per  sam- 
pled point  were:  27.0  in  A,  19.1  in  B,  22.1  in 
C,  and  25.8  in  G.  The  numbers  of  contacts  per 
point  were  not  significantly  different  between 
A and  G nor  between  B and  C;  but  they  were 
significantly  different  between  A-B,  A-C,  G- 
B and  G-C  (Tukey  test:  P < 0.05). 

There  were  variations  in  the  average  num- 


45 

40 

35 

30 

25 

20 

15 
10 

5 

0 

FIG.  5.  Average  number  of  contacts  in  the  studied 
sites  during  the  year.  Square  = A,  triangle  = B,  circle 
= C,  diamond  = G. 


ber  of  contacts  during  the  year  in  the  four 
study  sites.  The  number  of  contacts  was  high- 
est in  September/October— January  than  Feb- 
ruary-August (Fig.  5). 

Relative  abundance. — The  highest  relative 
abundance  (number  of  contacts  divided  by  the 
number  of  sampled  points)  in  this  study  was 
Basileuterus  leucoblepharus  (1.2)  followed  by 
Turdus  rufiventris  (0.85),  Lepidocolaptes 
squamatus  (0.68),  Basileuterus  culicivorus 
(0.68),  Cyclarhis  gujanensis  (0.67),  and  Cran- 
ioleuca  obsoleta  (0.61). 

The  majority  of  species  recorded  in  the 
point  counts  presented  significant  differences 
in  the  values  of  relative  abundance  when  oc- 
curring in  three  or  four  sampled  sites  (indi- 
cated with  “s”  in  Appendix;  df  = 2 or  3,  P 

< 0.05).  Some  of  these  species  increased  in 
abundance  with  the  decrease  in  area  (density 
compensation)  such  as  Veniliornis  spilogaster, 
Cranioleuca  obsoleta,  Turdus  amaurochali- 
nus,  Cyclarhis  gujanensis.  Panda  pitiayumi, 
Thraupis  sayaca,  and  Stephanophorus  diade- 
matus.  The  distribution  of  the  abundances  in 
a rank  order  showed  that  the  12  species  with 
highest  abundances  (indicated  with  an  * in 
Appendix)  in  the  smallest  patch  analyzed  (G) 
comprised  27%  of  total  contacts  but  they  com- 
prised only  20%  in  the  continuous  forest  (A). 

Relative  abundance  increased  with  decreas- 
ing area  for  12  omnivorous  species  (Appen- 
dix); Penelope  obscura  (x“  = 9.12,  df  = 3,  P 

< 0.05),  Pachyramphus  polychopterus  (x^  = 
73.87,  df  = 3,  P < 0.001),  Pitangus  sulphur- 
atus  (x'  = 106.39,  df  = 3,  P < 0.001),  Elaen- 


402 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


ia  mesoleuca  (x“  = 23.38,  df  = 3,  P < 0.001), 
Cyanocorax  caeruleus  (x"  = 21.08,  df  = 3,  P 

< 0.001),  Turdus  rufiventris  (x‘  = 8.52,  df  = 

3,  P < 0.05),  Turdus  amaurochalinus  (x'  = 
23.9,  df  = 3,  P < 0.001),  Cyclarhis  gujanen- 
sis  (x-  = 36.19,  df  = 3,  P < 0.001),  Stephan- 
ophorus  diadematus  (x“  — 12.92,  df  = 3,  P < 
0.01),  Thraupis  sayaca  (x“  = 19.8,  df  = 3,  P 

< 0.001),  Zonotrichia  capensis  (x^  = 33.64, 

df  = 3,  P < 0.001),  and  Carduelis  magellan- 
icus  (x"  = 14.07,  df  = 3,  P < 0.01).  The  sum 
of  relative  abundances  of  the  omnivorous  spe- 
cies increased  (x“  28.41,  df  = 2,  P < 0.001) 

from  A (6.56)  to  the  patches  C (7.52)  and  G 
(9.0). 

The  sum  of  relative  abundances  of  leaf  in- 
sectivorous species  was  highest  (x^  = 77.15, 
df  = 3,  P < 0.001)  in  A (8.42),  followed  by 
G (6.18),  C (5.56),  and  B (5.14).  Parula  pi- 
tiayumi  (x“  = 12.86,  df  = 3,  P < 0.01)  in- 
creased in  relative  abundance  with  decreases 
in  area  (Appendix). 

The  sum  of  relative  abundances  of  species 
of  trunk  insectivores  increased  (x“  = 10.8,  df 
= 1,  P < 0.01)  from  A (4.06)  to  G (5.24)  as 
the  result  of  a substantial  increase  in  relative 
abundance  of  two  species:  Cranioleuca  ob- 
soleta  (0.21  in  A and  1.54  in  G)  and  C.  pallida 
(0.14  in  A and  0.93  in  G). 

Relative  abundances  of  Furnarius  rufus  (x^ 
= 49.89,  df  = 2,  P < 0.001),  Tyrannus  rne- 
lancholicus  (x^  = 23,  df  = 3,  P < O.OOL),  and 
Phyllomyias  fasciatus  (x^  = 39.72,  df  = 3,  P 
< 0.001)  generalized  insectivores,  increased 
with  decreases  in  area  (Appendix).  The  sum 
of  the  relative  abundances  of  this  group  was 
higher  (x^  = 8.43,  df  = 1,  P < 0.01)  in  G 
(3.82)  than  A (2.93). 

The  sum  of  the  relative  abundances  of  fru- 
givores  decreased  quickly  (x'  “ 176.93,  df  — 
3^  p < 0.001)  with  decreases  in  area:  A,  4.18; 
B,  2.89;  C,  1.76,  and  G,  0.92.  Only  Leptotila 
verreauxi  (x^  = 20.43,  df  = 3,  P < 0.001 ) and 
L.  rufaxilla  (x^  ~ 11.96,  df  = 3,  P < 0.01) 
increased  in  relative  abundance. 

DISCUSSION 

Number  of  species. — Based  on  censuses 
from  a few  visits  to  many  islands  in  the  spe- 
cies/area studies,  Haila  and  Jiirvinen  (1981: 
561)  suggested  that  the  “gain  by  increasing 
the  number  of  visits  to  an  island  is  the  im- 
proved accuracy  in  the  recording  of  rare  spe- 


cies.” Almost  90%  of  the  species  in  a site  are 
recorded  during  a single  visit  (Haila  and  Ku- 
usela  1982).  Taking  into  account  the  point 
counts  of  one  year  carried  out  in  the  present 
study,  the  combined  samplings  in  A,  B,  C,  and 
G from  September  to  December  (four  visits) 
revealed  73—78%  of  the  total  species  in  these 
sites;  this  is  less  than  the  90%  predicted  by 
Haila  and  Kuusela  (1982).  The  fewer  species 
we  recorded  in  four  visits  is  problably  due  to 
a higher  proportion  of  rare  species  in  our 
study  sites.  Therefore  researchers  in  the  Neo- 
tropical region,  especially  in  forests,  should 
make  several  visits  to  a site  instead  of  a single 
visit. 

There  was  a strong  correlation  (r  = 0.92, 
df  = 10,  P < 0.001)  between  the  number  of 
bird  species  and  the  size  of  the  forest  patches 
in  the  Campos  Gerais  region.  Other  effects, 
such  as  habitat  heterogeneity,  might  be  more 
important  than  area  in  predicting  the  number 
of  species  found  on  an  island  (Martin  et  al. 
1995).  In  more  homogeneous  sites  it  is  ex- 
pected that  the  effect  of  area  would  increase 
(Connor  and  McCoy  1979).  Galli  and  co- 
workers (1976)  studied  woodlots  in  New  Jer- 
sey with  similar  foliage  height  diversity  and 
found  a strong  correlation  between  the  num- 
ber of  species  and  size  of  the  area.  Martin 
(1981)  studied  homogeneous  shelterbelts  in 
South  Dakota  and  concluded  that  area  was  the 
most  important  factor  in  determining  the  num- 
ber of  species.  The  similarity  in  vegetation 
structure  between  Campos  Gerais  forest 
patches  is  due  to  their  being  at  approximately 
the  same  stage  of  ecological  succession  (Klein 
1960,  1972;  Klein  and  Hatschbach  1971; 
Maack  1981),  and  probably  explains  the  high 
species/area  correlation. 

Based  on  the  study  by  Tomialojc  and  co- 
workers (1984)  in  Poland,  Blondel  (1986) 
suggested  that  the  habitats  in  large  forests  ap- 
pear to  be  broken  into  a mosaic  of  “sink”  and 
“source”  because  a large  forest  is  represented 
by  a mosaic  of  habitat  patches.  This  results  in 
a potential  heterogeneity  of  the  distribution  of 
birds  in  large  forests.  In  order  to  test  the 
source/sink  (Blondel  1986)  role  of  larger  for- 
ests, we  calculated  the  ^-diversity  (Whittaker 
1960)  among  the  six  points  sampled  monthly, 
in  the  9 ha  patch  (G)  and  in  the  largest  patch 
(A,  840  ha).  The  mean  (3-diversity  was  higher 
in  A (1.68)  than  in  G (1.27;  Mann-Whitney 


Anjos  am!  lioi-on  • BIRD  COMMUNI  TIES  IN  FOREST  PATCHES 


403 


t/-test:  U = 12.5,  P < 0.05),  which  suggests 
a more  homogeneous  bird  distribution  in  the 
smaller  area.  The  average  number  of  species 
per  point  sampled  was  not  significantly  dif- 
ferent (Tukey  test:  P > 0.05)  between  G and 
A,  which  suggests  that  the  higher  number  of 
total  species  in  the  largest  patch  is  mainly  due 
to  the  heterogeneity  of  species’  distributions 
compared  to  smaller  patch. 

It  was  expected  that  there  would  be  more 
species  in  A (138  species,  840  ha)  than  in  G 
(91  species,  9 ha)  because  of  the  difference  in 
size  of  the  fragments.  The  small  difference 
can  be  explained  by  the  heterogeneity  of  spe- 
cies distribution  in  A;  the  area  sampled  in  A 
(24  sampled  points,  which  represents  around 
80  ha)  corresponds  to  about  10%  of  the  total 
site,  but  the  total  area  of  G (9  ha)  was  sam- 
pled. Two  hundred  and  seven  species  were  re- 
corded in  A in  a parallel  study  (Anjos  1992). 
Thus,  the  heterogeneity  in  bird  distribution  re- 
sulted in  Just  67%  of  the  total  avifauna  in  A 
being  recorded  in  this  study. 

Composition  of  the  bird  communities. — In 
the  Neotropics,  as  in  the  rest  of  the  world, 
losses  of  species  and  changes  in  avian  com- 
munity composition  occur  with  fragmentation 
of  forest  areas.  In  the  Campos  Gerais,  the 
scarcity  of  forest  species  are,  in  part,  balanced 
by  gains  in  edge  species  (Fig.  3).  Willis 
(1979)  pointed  out  that  the  woodlots  of  Sao 
Paulo,  southeastern  Brazil,  have  become  more 
like  temperate  zone  forests  with  an  emphasis 
on  oscine  birds  characteristic  of  canopy  and 
edge.  Bierregaard  and  Lovejoy  (1988)  noted 
some  edge  species  {Troglodytes  aedon  and 
Ramphocelus  carbo)  wandering  into  small 
forest  patches  (1  ha)  after  experimental  defor- 
estation in  Amazonia.  In  the  Campos  Gerais 
region  a gradual  increase  in  the  ratio  of  edge 
species  to  forest  species  was  obvious  with  a 
decrease  in  fragment  area.  The  majority  of 
these  edge  species  are  oscines,  some  of  them 
living  in  the  canopy. 

The  proportion  of  forest/edge  species  does 
not  seem  to  be  related  only  to  the  size  of  the 
area.  MacArthur  and  Wilson  (1963,  1967) 
demonstrated  that  isolation  is  also  an  impor- 
tant factor.  Among  the  isolated  patches,  L (1 .5 
ha)  possessed  a higher  proportion  of  forest 
species  than  F (10  ha),  H (8.5  ha),  I (6.5  ha), 
or  J (4  ha)  probably  because  it  was  closer  to 
the  largest  patch  (Fig.  3,  Table  1). 


Among  lorest  birds,  the  principal  decrea.se 
in  number  ol  species  in  the  Campos  Gerais 
patches  was  associated  with  the  loss  of  un- 
derstory species.  Canopy  species  probably  fly 
to  a nearby  patch  when  there  is  a shortage  of 
resources.  Bierregaard  and  Lovejoy  (1989) 
verified  a decrease  in  the  number  of  species 
also  occurring  among  understory  birds  in 
Amazonia. 

Aleixo  and  Vielliard  (1995)  pointed  out  that 
leaf  insectivores  of  the  understory  were  those 
most  likely  to  be  absent  in  a woodlot  of  251 
ha  in  Sao  Paulo.  We  also  found  that  most  of 
the  understory  species  that  were  absent  from 
the  small  patches  in  the  Campos  Gerais  region 
were  leaf  insectivores.  Trunk  insectivores 
were  rarely  found  in  small  woodlots  by  Willis 
(1979);  these  were  mostly  small  birds,  such  as 
Picumnus  spp.  (12  g)  and  Veniliornis  spilo- 
gaster  (43  g).  Large  trunk  insectivores,  such 
as  Phloeoceastes  robustus  (263  g)  and  Dry- 
ocopus  lineatus  (246  g),  were  recorded  only 
in  the  forest  of  our  study.  Only  small  or  mid- 
size species  such  as  Cranio  I euca  obsoleta  (14 
g),  C.  pallida  (12  g),  Lepidocolaptes  squa- 
matus  (28  g),  and  Veniliornis  spilogaster  were 
recorded  in  the  smallest  patches  (Appendix). 

Frugivores  were  rare  in  small  patches  of 
forest  of  Amazonia  (Bierregaard  and  Lovejoy 
1989)  and  Sao  Paulo  (Willis  1979,  Aleixo  and 
Vielliard  1995).  Willis  (1979)  thought  that  be- 
cause frugivores  can  easily  travel  to  other 
woodlots  to  look  for  available  resources  they 
probably  disappear  from  the  small  woodlots; 
they  depend  on  scattered  trees  of  different 
species  at  different  seasons  or  years,  and  only 
large  woodlots  have  enough  tree  diversity  to 
support  them.  Frugivores  occurred  in  all  the 
patches  we  studied  in  the  Campos  Gerais  re- 
gion. This  ability  to  easily  move  between 
patches  may  have  prevented  their  numbers 
from  declining  (although  the  relative  abun- 
dance of  frugivores  in  smaller  patches  was 
low,  as  shown  above). 

Omnivores,  which  can  switch  from  fruit  to 
insects  or  vice  versa,  may  benefit  from  small 
woodlots  of  Sao  Paulo  (Willis  1979).  This 
group  was  represented  by  the  greatest  number 
of  species  in  the  Campos  Gerais  and  tended 
to  increase  its  contribution  to  the  avifaunal 
composition  with  decreases  in  area;  the  sum 
of  relative  abundance  increased  from  A (6.56) 
to  G (9.0). 


404 


THE  WILSON  BULLETIN  • Vol.  1 1 1,  No.  3.  September  1999 


The  present  composition  of  the  avifauna  in 
the  patches  at  the  Campos  Gerais  region  is  the 
result  of  a long  period  during  which,  presum- 
ably, many  factors  have  operated.  Forest 
patches  resulting  from  human  disturbance  of 
a continuous  forest  are  isolated  more  rapidly. 
The  effects  of  isolation  on  natural  and  artifi- 
cial forest  fragments  may  be  different.  Ac- 
cording to  Willis  (1979),  frugivores  seem  to 
be  more  sensitive  to  isolation  in  woodlots  in 
Sao  Paulo  than  in  those  of  the  Campos  Gerais. 
But  the  woodlots  studied  by  Willis  (1979)  are 
larger  and  more  isolated  than  the  patches  stud- 
ied in  Campos  Gerais.  Thus,  a study  carried 
out  under  similar  conditions  is  necessary  for 
a more  precise  comparative  analysis. 

Relative  abundance. — Vielliard  and  Silva 
(1990)  and  Aleixo  and  Vielliard  (1995),  using 
the  same  census  method  of  point  counts  in 
Sao  Paulo,  Brazil,  obtained  a similar  pattern 
of  monthly  variation  in  number  of  contacts  as 
we  did  (Fig.  5).  This  is  probably  because  Sep- 
tember-December  is  the  main  period  when  all 
the  species  are  more  obvious  when  they  breed 
and  vocalize  (Vielliard  and  Silva  1990,  Aleixo 
and  Vielliard  1995).  There  are  also  summer 
transients  of  some  species  in  the  study  area 
from  September  to  March  (Anjos  and  Graf 
1993).  Therefore,  counts  of  relative  abun- 
dance are  influenced  by  vocalization  and  mi- 
gratory behavior.  Because  censuses  were  per- 
formed monthly  in  each  patch,  seasonal  vari- 
ation was  assumed  to  be  the  same  for  all  sites, 
making  comparative  analysis  possible  be- 
tween the  patches.  In  addition,  because  habitat 
physiognomy  was  similar  for  all  of  our  patch- 
es, bias  between  the  sites  resulting  from  dif- 
ferences in  detection  of  vocalizations  (Schieck 
1997)  was  unlikely. 

Density  compensation  was  detected  in  the 
present  study  for  several  species.  For  edge 
species  the  greater  relative  abundance  is  prob- 
ably due  to  the  relatively  greater  habitat  area 
in  the  smaller  patches.  Edge  species  may  also 
be  better  colonizers  of  isolated  patches  than 
forest  species. 

Decreased  competition  in  smaller  patches 
could  explain  density  compensation  (Ricklefs 
and  Cox  1978).  This  seems  to  be  the  case  for 
two  leaf  insectivores  {Thamnophilus  caerules- 
cen.s  and  Parula  pitiayurni)  and  some  trunk 
insectivores  {Veniliornis  spUoga.ster,  Lepido- 
colaptes  .squarnatu.s,  Cranioleuca  ohsoleta. 


and  C.  pallida)  in  the  present  study.  These 
groups  decreased  in  number  of  species  as  area 
decreased,  which  could  mean  fewer  competi- 
tors. However,  documenting  that  competition 
occurs  among  species  is  not  easy  (Wiens 
1989). 

The  “habitat  appropriation”  hypothesis  of 
Blondel  and  coworkers  (1988;  a population 
size  increases  in  an  island  because  of  an  ex- 
pansion of  habitat  occupation)  was  examined 
with  two  trunk  and  twig  insectivores.  Crani- 
oleuca obsoleta  (x"  “ 162.72,  df  = 3,  F < 
0.05)  and  C.  pallida  (x^  = 91.41,  df  = 3,  F 
< 0.05)  increased  in  relative  abundance  with 
decreases  in  area  (Appendix).  Points  were  se- 
lected in  A and  G with  either  dense  or  open 
understory.  The  relative  abundance  of  the  two 
species  was  calculated  in  these  two  habitats 
(dense  and  open  understory).  Cranioleuca  ob- 
soleta had  a relative  abundance  of  0.33  in 
open  understory  and  0.08  in  dense  understory 
areas  of  the  840  ha  patch  (A)  but  2.07  and 
1.06,  respectively,  in  the  9 ha  patch  (G).  Cran- 
ioleuca pallida  was  not  observed  in  dense  un- 
derstory and  had  a relative  abundance  of  0.67 
in  open  understory  areas  of  A,  and  0.43  and 
1.29  respectively  in  G.  This  suggests  that  both 
species  increase  the  habitats  they  use  (they 
tended  to  be  more  abundant  in  areas  of  dense 
understory  in  G).  These  results  were  consis- 
tent with  what  could  be  interpreted  as  the 
“habitat  appropriation”  hypothesis.  This 
should  be  considered  as  a complementary  ex- 
planation for  density  compensation.  Crani- 
oleuca obsoleta  also  was  common  with  high 
relative  abundance  in  another  natural  patch  of 
forest  in  Curitiba  city,  Parana  (Anjos  and  Lar- 
oca  1989). 

The  causes  of  density  compensation  are  not 
clear  but  they  may  be  different  for  each  spe- 
cies. 

ACKNOWLEDGMENTS 

We  acknowledge  Conselho  Nacional  de  Pesquisa 
(CNPq-Brasilia)  for  research  grants  (LdA-350054/95- 
9).  Financial  support  for  the  field  study  was  provided 
by  the  Universidade  Estadual  de  Londrina,  Londrina, 
the  Consorcio  Intermunicipal  para  Prote9ao  Antbiental 
do  rio  Tibagi.  Londrina,  and  the  Klabin  Fabricadora 
de  Papel  e Celulose,  Tclemaco  Borba.  We  thank  T. 
Matsuo  for  statistical  analysis  assistance  and  M.  M.  F. 
Costa  and  A.  H.  F.  de  Toledo  for  English  revision.  M. 
C.  Dias  and  M.  Torrezan  helped  us  with  the  vegetation 
description. We  appreciated  the  assistance  of  E.  A. 


Anjos  ami  Boi^on  • BIRD  COMMUNITIES  IN  FOREST  PATCHES 


405 


Lima,  H.  D.  Garcia,  and  V.  L.  Ogassawara  using  SAS/ 

STAT  J.  C.  V.  Lopes  kindly  permitted  us  to  work  at 

the  Santa  Rita  farm.  Finally  we  sincerely  thank  W. 

Belton,  J.  Bates,  E.  O.  Willis  and  Y.  Oniki  for  com- 
ments and  criticisms  that  improved  the  manuscript  and 

J.  M.  E.  Vielliard  for  suggestions  during  the  held  work. 

LITERATURE  CITED 

Aleixo,  a.  and  j.  M.  E.  Vielliard.  1995.  Composi^ao 
e dinamica  da  avifauna  da  mata  de  Santa  Genebra, 
Campinas,  Sao  Paulo,  Brasil.  Rev.  Bras.  Zool.  12; 
493-5 1 1 . 

Anjos,  L.  dos.  1992.  Riqueza  e abundancia  de  aves 
em  “ilhas”  de  floresta  de  araucaria.  Ph.D.  diss., 
Univ.  Federal  do  Parana,  Curitiba,  Brazil. 

Anjos,  L.  dos  and  S.  Laroca.  1989.  Abundancia  re- 
lativa  e diversidade  especifica  em  duas  comuni- 
dades  urbanas  de  aves  de  Curitiba  (sul  do  Brasil). 
Arq.  Biol.  Tecnol.  (Curitiba)  32:637-643. 

Anjos,  L.  dos  and  V.  Graf.  1993.  Riqueza  de  aves  da 
fazenda  Santa  Rita,  regiao  dos  Campos  Gerais, 
Palmeira,  Parana,  Brasil.  Rev.  Bras.  Zool.  10:673— 
693. 

Anjos,  L.  dos,  K.-L.  Schuchmann,  and  R.  Berndt. 
1997.  Avifaunal  composition,  species  richness, 
and  status  in  the  Tibagi  River  basin,  Parana  State, 
southern  Brazil.  Omitol.  Neotrop.  8:145—173. 

Belton,  W.  1984.  Birds  of  Rio  Grande  do  Sul,  Brazil. 
Part  1.  Bull.  Amer.  Mus.  Nat.  Hist.  178:369-636. 

Belton,  W.  1985.  Birds  of  Rio  Grande  do  Sul,  Brazil. 
Part  2.  Bull.  Amer.  Mus.  Nat.  Hist.  180:1-241. 

Bierregaard,  R.  O.,  Jr.  1990.  Avian  communities  in 
the  understory  of  the  Amazonian  forest  fragments. 
Pp.  333-343  in  Biogeography  and  ecology  of  for- 
est bird  communities  (A.  Keast,  Ed.).  SPB  Aca- 
demic. The  Hague,  Netherlands. 

Bierregaard,  R.  O.,  Jr.  and  T.  E.  Lovejoy.  1988. 
Birds  in  Amazonian  forest  fragments:  effects  of 
insularization.  Proc.  Ornithol.  Congr.  19:1564- 
1579. 

Bierregaard,  R.  O.,  Jr.  and  T.  E.  Lovejoy.  1989.  Ef- 
fects of  forest  fragmentation  on  Amazonian  un- 
derstory bird  communities.  Acta  Amazonica  19: 
215-241. 

Blondel,  j.  1986.  Biogeographie  evolutive.  Masson, 
Paris,  France. 

Blondel,  J.  1991.  Birds  in  biological  isolates.  Pp.  45- 
72  in  Bird  population  studies  (C.  M.  Perrins,  J.- 
D.  Lebreton,  and  G.  J.  M.  Hirons,  Eds.).  Oxford 
Univ.  Press,  Oxford,  U.K. 

Blondel,  J.,  C.  Ferry,  and  B.  Frochot.  1970.  La 
methode  des  indices  ponctuels  d’abondance 
(I. PA.)  ou  des  releves  d’avifaune  par  “stations 
d'ecoute”.  Alauda  38:55-71. 

Blondel,  J.,  D.  Chessel,  and  B.  Frochot.  1988.  Birds 
impoverishment,  niche  expansion,  and  density  in- 
flation in  mediterranean  island  habitats.  Ecology 
69:1899-1917. 

Connor,  E.  F.  and  E.  D.  McCoy.  1979.  The  statistics 
and  biology  of  the  species-area  relationship.  Am. 
Nat.  113:791-833. 


Fitzpatrick,  J.  W.  1980.  Foraging  behavior  of  Neo- 
tropical tyrant  flycatchers.  Condor  82:43—57. 

Galli,  a.  E.,  C.  F.  Lecj,  and  R.  T.  Forman.  1976. 
Avian  distribution  patterns  within  sized  forest  is- 
lands in  central  New  Jersey.  Auk  93:356-365. 

Haila,  Y.  and  O.  Jarvinen.  1981.  The  underexploited 
potential  of  bird  censuses  in  insular  ecology.  Stud. 
Avian  Biol.  6:559-565. 

Haila,  Y.  and  S.  Kuusela.  1982.  Efficiency  of  one 
visit  censuses  of  birds  breeding  on  small  islands. 
Ornis  Scand.  13:17-24. 

Karr,  J.  R.  1982.  Avian  extiction  on  Barro  Colorado 
Island,  Panama;  a reassessment.  Am.  Nat.  119: 
220-239. 

Klein,  R.  M.  1960.  O aspecto  dinamico  do  pinheiro 
brasileiro.  Sellowia  12:17-44. 

Klein,  R.  M.  1972.  As  florestas  da  America  do  Sul. 
Univ.  de  Brasilia,  Brasilia,  Brazil. 

Klein,  R.  M.  and  G.  Hatschback.  1971.  Fitofision- 
omia  e notas  complementares  sobre  a mapa  fito- 
geogrifico  de  Quero-quero  (Parana).  Bol.  Par. 
Geoc.  28: 159-188. 

Leck,  C.  F.  1979.  Avian  extinctions  in  an  isolated  trop- 
ical wet-forest  preserve,  Ecuador.  Auk  96:343- 
352. 

Maack,  R.  1981.  Geografia  ffsica  do  Estado  do  Para- 
na. Livraria  Jose  Olympio,  Rio  de  Janeiro,  Brasil. 

MacArthur,  R.  and  E.  O.  Wilson.  1963.  An  equilib- 
rium theory  of  insular  zoogeography.  Evolution 
17:373-387. 

MacArthur,  R.  and  E.  O.  Wilson.  1967.  The  theory 
of  island  bigeography.  Princeton  Univ.  Press, 
Princeton,  New  Jersey. 

MacArthur,  R.,  J.  R.  Kaar,  and  J.  M.  Diamond. 
1972.  Density  compensation  in  island  faunas. 
Ecology  53:330-342. 

Martin,  J.-L.,  A.  J.  Gaston,  and  S.  Hitier.  1995.  The 
effect  of  island  size  and  isolation  and  old  growth 
forest  habitat  and  bird  diversity  in  Gwaii  Haanas 
(Queen  Charlotte  Islands,  Canada).  Oikos  72:1 15- 
131. 

Martin,  T.  E.  1981.  Limitation  in  small  habitat  islands; 
chance  or  competition.  Auk  98:715-734. 

Meyer  de  Schauensee,  R.  1982.  A guide  to  the  birds 
of  South  America.  Academy  Natural  Science, 
Philadelphia,  Pennsylvania. 

Preston,  E W.  1962.  The  canonical  distribution  of 
commoness  and  rarity.  Ecology  43:185-215. 

Ricklefs,  R.  E.  and  G.  W.  Cox.  1978.  Stage  of  taxon 
cycle,  habitat  distribution  and  population  density 
in  the  avifauna  of  the  West  Indies.  Am.  Nat.  1 12: 
875-895. 

Ridgely,  R.  S.  and  G.  Tudor.  1989.  The  birds  of 
South  America.  Vol.  1.  Oxford  Univ.  Press,  Ox- 
ford, U.K. 

Ridgely,  R.  S.  and  G.  Tudor.  1994.  The  birds  of 
South  America.  Vol.  2.  Oxford  Univ.  Press,  Ox- 
ford, U.K. 

SAS  Institute,  Inc.  1989.  SAS/STAT®  user's  guide, 
version  6.1 1,  fourth  ed.  SAS  Institute,  Inc.,  Carey, 
North  Carolina. 


406 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


SCHIECK,  J.  1997.  Biased  detection  of  bird  vocaliza- 
tions affects  comparisons  of  bird  abundance 
among  forested  habitats.  Condor  99:179-190. 

Sick,  H.  1997.  Omitologia  brasileira.  Nova  Lronteira 
ed.,  Rio  de  Janeiro,  Brazil. 

Tomialojc,  L.,  T.  Wesolowski,  and  W.  Walankiew- 
icz.  1984.  Breeding  bird  community  of  a primae- 
val temperate  forest  (Bialowieza  National  Park, 
Poland).  Acta  Omitol.  20:241-310. 

ViELLiARD,  J.  E.  M.  AND  W.  R.  SiLVA.  1990.  Nova  me- 
todologia  de  levantamento  quantitativo  da  avifau- 
na e primeiros  resultados  no  interior  do  Estado  de 
Sao  Paulo,  Brasil.  Pp.  117-151  in  Anais  do  IV 
Encontro  de  Anilhadores  de  Aves  (S.  Mendes, 


Ed.).  Univ.  Federal  Rural  de  Pernambuco  Press, 
Recife,  Brazil. 

Whittaker,  R.  H.  1960.  Vegetation  of  the  Siskiyou 
Mountains,  Oregon  and  California.  Ecol.  Monogr. 
30:279-338. 

Wiens,  J.  A.  1989.  The  ecology  of  bird  communities. 

Vol.  2.  Cambridge  Univ.  Press,  Cambridge,  U.K. 
Willis,  E.  O.  1974.  Populations  and  local  extinctions 
of  birds  on  Barro  Colorado  Island,  Panama.  Ecol. 
Monogr.  44:153-169. 

Willis,  E.  O.  1979.  The  composition  of  avian  com- 
munities in  remanescent  woodlots  in  southern 
Brazil.  Papeis  Avulsos  Zoologia  33:1-25. 
Wright,  S.  J.  1980.  Density  compensation  in  island 
avifaunas.  Oecologia  45:385—389. 


Aujos  and  Bo^on  • BIRD  COMMUNITIES  IN  FOREST  PATCHES 


(U  w 

fc 

3 
U 
CJ 
O 


-C 


c 

(u  •£ 

(U 

-O  o 

v? 


X V 

w CJ 


OJ 

o 

c 

o 


c 

■o 
c 

3 

,(U  -O 


c 

3 

CJ 


c O 

OJD 

•—  c/2 

c/2  (U 

c/: 

O 3 
3 >■ 
O ^ 
L3  c/2 
^ (D 
C -3 

c/2  .3 


O'^ 


u 

U D. 

c/2 

03  ^ 

u 

c/2  CC 

OJ 

.3  O 

C4  ^ 

^ c/2 

■£  ^ 

C O 


CJ 

CJ  . 

a 


<u 

j= 


■o 

c 

3 

<72 


3 

C 

“O 

C .3 
o c/2 

H >- 

OJ 

- •£ 
Z o 


£ 

TO 

CJ 

3 

O 

c 


“ o 
c 1/ 
c a. 


uo 

< 

Q 


*5 

-c 

C' 


< 

O 


< 

o 


W 

< 

Q 

5 

Q 


S Cj  ftC  CO 


H 


W 

9 S 
£ ^ 
^ 3 

7 C 

Cj 

g S 

2 = 

in  -2 

QC  r3 

X ^ 
H 


<<  <<<  <<S2<-  S 

uu  uuu  uuoouo  o 


<<  <Q<  UUQQ<<  S 

OO  OUO  uih-WWOO  if 


pq 

< 

Q 

H 

aC 

< 

X 

u 


o o 
ci.  ^ 
^ C 
2f  3 


U U 


w 

<: 

Q 

5 

H 

s 

u 

u 

< 


■f 

3 

Cj 


:2 

s 

o 

.2 

f2 


W 

< 

Q 

Z 

O 

U 

J 

ii 


^ 5 

Rc-  S 

^ 5 


•3  o 

"S  ^ 

c I 

'3  "S. 


3 

Oc 


- 

-- 

- 

- 

- 

- 

— 

- 

— 

- 

— 

— 

- 

- 

- 

- 

— 

- 



-- 

- 

— 

- 

— 

- 

- 

00 

<N 

00  (N  ro 

VO  00 

>o 

Os 

r- 

CN  CN  00 

>0  (O 

(N 

(N 

O 

O (N  O 

o — 

d 

d 

d 

d d d 

d d 

d 

00 

m 

fo 

(N 

ID 

m 

ON 

r^j 

^T) 

o 

>o 

ON  CD 

VD 

o 

O 

o 

o — 

O 

o 

d 

d 

d 

d 

d 

d d 

d 

d 

ro 

r- 

'O 

VD  VD  (N 

(N 

00 

00 

a^ 

o 

00 

— . — c ro 

m 

o 

o 

O O O 

O 

c 

o 

d 

d 

d 

d 

d d d 

d 

d 

d 

n 

(N 

— CO 

00 

'O 

— ■ 00 

CN 

CN 

o 

O 

— o 

o 

c 

d 

d 

d 

d 

d d 

d 

d 

d 

>. 

5/ 


c C 


5 i:  i;  << 


w 

< 

9 

u 

< 

ac 

CJ 


w o 

9 2 

z S- 

< 3 

CO  5 

< ^ 

cu 


W 

< 

9 

j 

X 

< 

ai 


407 


APPENDIX.  CONTINUED. 


408 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


UQUQUQQPDD 

mutLLija,tLiwti.u-p- 


-1-  n 

— — r)- 


o o o 


\D 

un  00 

o o 
d d 


Qi  Di  Oi  Qi  Oi  Qi  Qi 
u-  U,  U-  U-  U-  tin  u- 


u Q u u u u u 

tu  UJ  tU  Uh  pH  tl.  tt, 


Q 

— 

r- 

o^ 

'd- 

r- 

(N 

o 

sO 

so 

00 

O 

o 

o 

O 

d 

d 

d 

d 

d 

d 

0^  00  ^ 
m <N  — 
— o o 
d d d 


OO  ^ 00 

— cN  m — ' n 

o o o p p 

d d d d d 


C 

E CL 


-c 


PJ 

< 

Q 

5 

S 

-j 

o 

u 


'S. 


ii  3 ~ 

.S-  § ^ 

s S:  ^ 

^ S' 

s ^ ^ 


^ s 


UPO'oONO-JhjO 


g 

C/1 

oc 

b d 5 c 

2i  -H  3 3 

s 

w 

2 

c 

S-  .2  -i;  ^ d 

-C  d Cl.  .5 

s 

C 

n 

C:  - „ i-  ~ 

~ .to  U 3i 

1-^ 

s 

i;  d -C:  C 2 

oc 

U 

< 

Ik 

^ ^ ^ 5 
s c ^ 2 

H 

H 

iK 

>. 

^ ^ S c 5 

^ k.  -S  -S  s 

oo 

Ok 

k,  CQ  a,  d.  rt; 

Cl, 

S S S S 
o o o o 


S S q q 

Ph  tu  PJ  PJ 


< 

u 


Q 

PJ 


< < < 
u u u 


u Q Q 

P-  PJ  w 


o 

d 


o 

d 


00 

o 


r<'i 

o o 

r- 

os 

q 

o 

O' 

n 

00 

o 

r*^ 

o 

os 

CN 

SO 

O 

lO 

O 

SO 

q 

(N 

m 

q 

o 

d d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

^ — 

tt 

so 

o 

* 

■tt 

(N 

n 

o 

00 

os 

so 

so 

m 

. 

o — 

o 

o 

so 

o 

rJ 

— 

o 

o 

q 

d 

d d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

a 


Q 

pj 


00 

tN 

O 


\D 

LO 

O 

o 

d 

d 

so 

ON 

yn 

NO 

o 

o 

d 

d 

1h 


PJ 

< 

Q 

J 

D 

U 

D 

U 


00 


5 t 


U ^ ^ 


PJ 

< 

Q 

Z 

O 

H 

H 


PJ 

c < 

I s 

= 2 

^ H 


-c 


■ s: 
c 


<3 

-c  r;; 

H.  -s: 

It 


<■  .c 

Q 

5 s 

D 


5 'C 
O 


a: 

Cu 

< 

u 


APPENDIX.  CONTINUED. 


Anjos  and  Bo^on  • BIRD  COMMUNITIES  IN  FOREST  PATCHES 


409 


PQUJWUPJW  SS  (V 

Z2ZZZZ  OO  £ 


QQQpS  SS  U 

upjwUhE  Euh  [i, 


QuSuuQSu 

WPuEPhPuPJUhPh 


H H H H H H H 


S S S 
u.  u. 


UJ 


* 

r- 

CN 

00 

00 

o 

00 

CN 

r- 

NO 

m 

so 

ON 

.— ( 

(N 

(N 

o 

CN 

ON 

ON 

o 

00 

o 

r-H 

(N 

in 

o 

00 

o 

CN 

o 

00 

o 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

vO 

'Cf 

r- 

r- 

00 

CN 

CN 

NO 

cn 

On 

On 

* 

On 

NO 

IT) 

ON 

ON 

CN 

o 

NO 

O 

00 

.— • 

00 

in 

o 

o 

O 

o 

o 

O 

O 

O 

m 

o 

CN 

m 

o 

00 

o 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

C^J 

vO 

CN 

vO 

NO 

(N 

, 

CN 

NO 

CN 

VO 

NO 

* 

o 

* 

in 

NO 

m 

— - 

00 

<n 

OO 

— H 

CN 

m 

00 

m 

cn 

CN 

r- 

NO 

ON 

00 

m 

CN 

o 

o 

o 

O 

o 

o 

CN 

o 

o 

o 

o 

o 

o 

CN 

00 

o 

CN 

o 

CN 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

00 

CN 

VO 

o 

r- 

NO 

NO 

ON 

Tj- 

On 

r- 

00 

00 

* 

CN 

NO 

* 

CN 

NO 

* 

ON 

m 

C^l 

>ri 

in 

ON 

in 

in 

NO 

m 

CN 

CN 

r- 

m 

m 

(n 

O 

o 

O 

O 

o 

C^l 

o 

o 

o 

o 

o 

o 

1-H 

cn 

o 

o 

ON 

o 

in 

o 

NO 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

rz  o 

eu 

a. 

00 


w 

< 

Q 


X 

U 

0 

01 
H 


':3  ^ 


O 


K 

C 

-C5 


"S 

-s:  o 
0.  U 


c 


a 
oo  ^ 

•§l 

5 


^ -S  S 

Co  G F.  ^ 


W 

< 

Q 

Z 

O 

O 

O 

cd 

H 


s:  c 
Oc  Cic 

c c 

ti;  f; 


w 

< 

Q 

H 

00 

< 

X 

a- 

s 

< 

Ct' 


3 

i*. 

•5 

*0 


g 


w 

< 

Q 

U 

E 


-2  ■§ 

"S  S 

-O  K 

S:  *>M 

>-  ^ 
3 3 

S E 

3 3 

;j  ;j 


c 


^-2 


>1 
3 


3 


3 
>) 


S3  >. 

3 I 

i2  2 


3 : 
3 


So  So 


3 3 -E 
U U O. 


3^3 
3 3 0 
3 3 0 
O ? 

Q 5 ^ 


s:  c 

a. 


w 

< 

Q 

H 
Co 
< 

J 
O 
U 
O 
DC 
Q 

W Q 
Q 


3 

3 

s 


,2 

G 

a 

V 

03 

c 

<4) 


c\  C t.“ 

a -ii  'c 
■2  :§  a. 

•C  ^ 2 

^ 2 ia, 

S--S 

s ^ a 

C C Cj 

2 i g 

a S 

.s:^G 

CO  ><  Q 


3 
Oc  s 


S'  cS 

CO  c*-. 


Cn  Co  to 
^ ^ 

S-  S'  ^ 
2 -2  “2 
c "2  "5 


2i  ?=  2s 

^ ^ ^ 


APPENDIX.  CONTINUED. 


410 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3.  September  1999 


U 


SU 

o. 


PL,[UP-U,t)HpHU,U.P-UpHtl-P-U-pHtUP-pH 


s s 

O O J J J 


j J o o o o 


pdSSSdddppp 

uuP-Puu-ti-P-.P-P-u,ti-u. 


* 

ON 


o 

d 


ON  NO 
O fN 

d d 


* * 
— ON 
— ^ 

\0  — 00 
d d d 


O'!  NO  00 

o o o 
d d d 


(N 

O 

d 


o 

d 


* * 
r)  — ' 


NO 


00 

■Nt  ro  NO  >£N 
iri  ON  (N 


o o o 


o 

d 


— 00  Nt  NO 
00  ol  ^ ro 

— IT)  (N 

d d d d 


00 

o 


■*■ 

ON 


d 


— ri 

00  cn 


(N  in  (N  00 

— (N  On)  (N 

^ O — (N  P 

d d d d d 


oominincNcSNOoo 


m m 
O 00 
^ o 
d d 


* 

(N 
NO  ^ 
NO  O 

d d 


o 

d 


in  NO 


m 

d 


o — 
d d 


(n  tn 

o o 


o o c 


'll- 

o 


NO 

o 

d 


— 't:}-  — csr^-^oocn 
t^t^Nomr'NO'^rNi 
;pp 

dddddddd 


00 

n 

o 

d 


fN 

It 

O 

d 


* 

O On 
00  m 

NO  — 

d d 


m 00  On 
mom 
mm—' 
d d d 


— -minoomON^m 
minmoooNOONNO 


m (N 
d d d 


d o 
d d d 


^ 00  (N  in  — 


00 


* 

It 


* 

m ON  m 


NO  in  m m 
m NO  — ■ m 


o o o o o o o 


C: 


u 

< 

9 

5 

< 

z 

a; 

u. 


c;  No  ^ 

^ tjs  -> 
^ ^ i; 
d 2 

F 5 5 
*»■  —* 

:S  ^ t. 

C d 


No 


c 

lo 

<3 

Nj 


£3. 


5 c 


c c 


!3. 

-J 


cn 


to 


cn 


^ S lo 
2 
i; 

= c 
-cs  >■. 


No 


u 

< 

Q 

2 

< 

u 


^ ^c.0 

^ ::: 

5 § 


PS- 


2 


^ s ^ 

'nJ  1 

2 i 

'nJ  ^ 

5 Nc  3; 
a ? £3 


C 33 


C 
No  j:; 


QC 

O 

Ph 


£3  £3 
K?  li 

^ £u  k; 
53  3:^ 

K C Cn. 


NnJ  ^ t*"*  •>*  -.C 

^ £3  S d S .2f 

•<S!;5;i<l<£n>5  QCQ^iSf^QQft-GCdO 


APPENDIX.  CONTINUED. 


Atijos  and  Bo(;(m  • BIRD  COMMUNITIES  IN  FOREST  PATCHES 


411 


S S S 2 QJ  05 

o o o o £ £ 


o 


c « 
c o. 


w 

< 

Q 

O 

Z 

H 

O 

U 


2 


Cl 

S 

S 

>, 

Q 

C. 


>•  ^ 
CL,  C 

^ Ei. 

5 S 
-c: 

&•  &• 

I I 
^ &■ 
"5  "C 

a <3 

Q,  Q, 


to 

^ I 

a £:■ 


o 

u 

2 ^ 

2 2 

;>> 

3!  ^ 
EJ  Q 


o:  oi 

U,  Ph 


oaooaooaooojoojaooaoj 


^U^<<QQQQQQuuQQSppQ<p 

OiiOOOtuwujWpQpjupppwpjEEEwOEI- 


* 

* 

'tt 

so 

ON 

00 

CN 

m 

On 

t» 

OS 

so 

m 

NO 

o 

00 

— H 

CN 

(D 

o 

o 

(N 

o 

q 

ID 

CD 

o 

m 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

* 

* 

'O 

c:^ 

00 

(N 

00 

so 

r- 

o 

CN 

00 

•o 

00 

1-^ 

CN 

CN 

ID 

>-H 

o 

1— 

On 

tN 

CO 

o 

(N 

o 

o 

(N 

o 

o 

OS 

ID 

o 

CN 

O 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

* 

On 

'O 

os 

o 

ID  m 

so 

so 

NO 

00 

(D 

r- 

o 

lO 

so 

n 

so  O 

OS 

Os 

os 

NO 

— H 

sO 

o 

(N 

ID 

o 

o 

o 

o 

O 

(N 

o 

o 

d 

d 

d 

d 

d d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

SO 

CD 

r- 

■Nt 

(N 

ID 

On 

00 

ID 

— 

00 

CD 

ID 

00 

00 

00 

Tf 

CN 

CD 

CN 

(N 

so 

(N 

o 

o 

o 

o 

-H 

ID 

so 

CD 

o 

o 

O 

o 

CD 

o 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

Co 

C 

to 

<L> 


. ^ s 

Q-  Q, 


W 

< 

Q 

s 

CP 


■S-  g 

II 

U 


to 

W § I 

I -~ 

Z C 
< .5  -S 

^ ? r2 

>H  ><  tJ 

H 


3 

to  S 
-3 

C V 
-C  tp 

cx  Oc 

^ '5 

to  to 
3 3 
oc  Oc 

-2i  -Si 
"c  "S 
A .£. 

Xo  5^ 


.to 

p 

S; 

3 

■S  a 
=:  "s 
"3  ^ 

po.  J) 
to  3 

=3  b 
S>c  EX 

tJJ 


.£. 


■ I-  ^ 


^ § 


Cp  -2 

s :: 

^ .'o 

■S  S 
^ i 

IX.  tp 

5 "5 

CO 


3 

O 


-s; 


3 c .:i 

5 5 ^ 


oc 

2 

'S. 

to 

"o 

3 

6C  ■ 


- 

5 5 


Co  •:2 
2 § 
3 to 

b .*3 

-5  I 

A ^ 

3 

to  to 


C 3 
2 -p 

£ 5 


3 to 
S!  2i  t3 
EU  5 P 
.C  Eu 


3 P;  3 

» eS  -2 
2 § - 

2 I-  S 
,0  S c 
U UJ  u 


3 3 
tL>  ^ 
3 'J 

■5b  S 

g 5 


35  S 

3 -C 
C 'J 
;3  c 


Cp  c E- 
.J  3 
.b  3! 

^ £ 


APPENDIX.  CONTINUED. 


412 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


J J J 




JJhJJJJOOOOhJOOJJPh 


UDDDDSQSQQQQQUUUSPD 


u 


5 o- 


2 

s 

-c  s: 


•2  2 

~ 


t 

3 

"2 

s:  ^ 

•S-  o 
'c 

5 =s 

^ C3 


3 

-Ci 

3 


5 


c 

Ci.  o 

o -3 


K K 


i: 


3 

O 

3 

O 

Oc 


0,  O, 


3 

C/5 

3 

O 

■Si 

k 

3 

O 

-3 

a. 


3 

3 

k 

-3 

3 

3 

00 

3 

-3 

CX 

O 

Co 


-2  -2 
kj  kj 


0 

c 

3 

3 

•Si 

3 

k 

3 

3 

to 

_k 

-S 

k» 

-3 

0 

0 

•S 

■S 

.3 

5 

5 

S 

2 

3 

S.  "s 
■§ 


c 

3 

3 


3 3 

kO  kj 


3 

5 
o 

oo  ^ 

<3  2 

•2  s 


3 5 

_3  2 

C -Cj 

^ 5 

S §■ 
■1,'^ 
5 ^ 

i- 

^ k 
•3  3 

a,  < 


O O 


U 2 

kn  P- 


Q 

PQ 


S 2 S S S 
o o o o o 


S 2 S S P 

Uh  Ph  Ph  Ph  P-. 


Tf 

CN 

so 

CNl  ^ 

00 

0 

m 

so 

On 

a\ 

ON  — 

04 

tn 

»n 

04 

0 

(N  0 

0 

CN 

m 

d 

d 

d 

d 

d d 

d 

d 

d 

d 

n 

* 

so 

NO 

os 

r4 

04 

os 

00 

Tf 

os 

m 

10 

NO 

so 

r4 

Os 

so 

0 

0 

SO 

0 

0 

(N 

0 

(N 

(N 

0 

04 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

(N 

00 

0 

in 

CN 

m 

so 

so 

■O 

ro 

so 

os 

NO 

os 

0 

0 

0 

<N 

0 

0 

0 

0 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

0 

<N 

00 

rt 

<N 

so 

04 

r- 

m 

QS 

04 

m 

m 

os  — 

If 

10  »n 

so 

0 

(N 

0 

0 

0 

0 0 

0 

0 — ' 

0 

0 

d 

d 

d 

d 

d 

d 

d 

d d 

d 

d d 

d 

d 

d 

d 

d 

o 

d 


VD 


ro 

00 

O 


r'  (N  ON  ko 
On  ■■t  — ' 00 

O o m -r]- 

d P d d 


* 

-H  Tt 

O 00 

d d 


ON  NO 

— (N 

d d 


00 

■■t 

o 


* 

r'  ro 

00  — ' U1 

00  — ' ro 

d d d 


* 

-H  Tf  NO 

^ — I 00 

NO  O ■■t 

d d d 


lo 

3 

"3 

td 
^ 2 
U 2* 

3 § 

3 

3 ? 

2 S 

'2  3 

c -s 

3 FX 


3 

-Si 

"3 

k 


to 

to 

3 -3 

3 3 


W 

< 

Q 


3 

■3 

1) 

3 


iO 

- 3 
3.  2 

P a. 


w 

< 

Q 

> 

Di 

O 

U 


i 

3 

3 

3 

3 

3 


U O 


> s 

Q S. 

3 I 

8 I 

Di  K 

H 


W 

< 

Q 

5 

at. 

P 

H 


•S- 

S 

3 

3 

■— .j 

-3 

•2 

3 

d 

3 


a.  L ^ 

^33 

•2^3 

.2?^  2 :2 

3 k 3 ■ 

to  >1  t<3 

3 3 3- 

^ 

^ s 


APPENDIX.  CONTINUED. 


Anjos  and  Bo^on  • BIRD  COMMUNITIES  IN  FOREST  PATCHES 


413 


o 


>>.2 
rz  o 

I ^ 

c/5 


2 « S 
o o o 


C 

C ^ 
^ 3 

^ Cj 

5; 


O 

Ci- 

to 

3 


S S 2 S 
o o o o 


Q U U Q 
PJ  tin  Uh  W 


J O J J 


CC 

u. 


Q 

PJ 


EEEotSoooooooo 


u u u 

P>  PP  Ph 


U Q U Q P 

pH  U P-,  U Pu 


.-H 

* 

* 

* 

* 

* 

r' 

00 

Os 

m 

00 

CN 

00 

CO 

00 

«-H 

c^ 

(N 

CO 

r- 

r- 

r- 

Os 

CO 

r- 

OS 

CO 

o 

CO 

r-- 

Os 

CN 

O 

(N 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

* 

* 

* 

Tf 

<— 1 

Os 

so 

fO 

to 

r- 

so 

00 

o 

so 

CO 

in 

so 

os 

so 

00 

Os 

m 

CO 

(O 

SO 

00 

o 

o 

04 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

* 

* 

* 

* 

* 

* 

os 

o 

Os 

CO 

(N 

so 

m 

CO 

Os 

to 

00 

CN 

1-H 

r- 

C4 

(N 

00 

— H 

o 

SO 

1-H 

o 

SO 

hH 

CO 

00 

CO 

00 

00 

Tf 

lo 

O 

fN 

o 

CO 

Tf 

o 

o 

O 

o 

(N 

o 

o 

o 

o 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

* 

* 

so 

o 

Os 

Oi 

00 

so 

'tt 

so 

00 

CO 

os 

*— ( 

00 

SO 

o 

fO 

Ol 

(N 

CO 

so 

w—( 

CO 

o 

CO 

so 

04 

CO 

<N 

CO 

o 

(N 

o 

o 

o 

o 

o 

(N 

t-H 

o 

»— ( 

o 

d 

d 

d 

d 

d 

d 

— i 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

d 

— a\ 
o o 
d d 


W 
< 
Q 

Z 

o 

> 


.to 

3 

c 

3 

C 

o 

to 

•g 

o 


to 

3 

O 

■5 

O 


§ ^ 

-3 

to  t/5 

3 3 


to 

3 . 

I ^ 


^ Cj 

^ O 
^ C 


k. 

§■ 


w 
< 

Q 

s 

w 

H 

P 5 G O P 


PP 

< 

Q 

P 

P 

Qi 

< 

Cu 


S 

s 

>, 

a 


s -= 
Q.  a 

^ t^ 

.a  -2  S 

Cp  £X  ^ 
r*  •^‘  S 

-S  ^ 

3 ■ 

O,  c:)  cq 


to 

3 

k 

3 

-3 

-C5 

O 

Cj 

3 

'HH 

a 

>H 

Si 

a 


W 

< 

Q 


?3 


5u  t3 


Vl 

a „ 
03  W 


Cei  .^u 


W 

< 

Q 

Cu 

P 

< 

OC 

X 

H 


a 

to 

3 

s 


3 

-Ci 

'■**.* 

a 

-a 

a 


■a 

a 


a a a 


■a  a; 
Op  a, 
a a 
kj  kj 


§ 

a ^ 
a Si 

a •a- 
kj  a. 


.2  S 

^ a 
a a; 

ct  ^ 

s S 

05  2 
ac  -j- 
a Ea 
,a  Si 

Ki  Co 


d s 

a 

a '3 

I § .2 

tr:  -a 

to  a 
a.  5. 

5 S 5 

-a  a:  ^ 

p p 5: 


a a 
a -2 
a tp 

^ -2 
to  S 

2 ^ 
a Q 

a c 
a a 

^ 2 
■to 
eS  tt; 


to  Ck 

a.  a 
ST  a 
a a 
a a 
^ a 

a iH 


a a 
a h 
a -£ 
a -2 
a a 

-S 

a,  a: 


414 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


'3 


X 


ooooooooo 


2 

o 


3 

3 

« 

I 

n 


qsqdqqdqqq 

[UtuUPuUtOU-WUU 


I 


Q 

uu 

D 

Z 

H 

Z 

O 

U 


Q 


>< 

Q 

Z 

tu 

CL. 

Cu 

< 


o 


u 


m 


< 


w-H 

- 

-- 

- 

- 

-- 

(N 

'O 

o^ 

o^ 

1-^ 

00 

00 

m 

o 

o 

fO 

d 

d 

d 

d 

d 

d 

* 

o 

00 

r4 

(N 

00 

00 

(N 

o 

CN 

(N 

O 

o 

lO 

(N 

o 

d 

d 

d 

d 

d 

d 

* 

00 

as 

o 

(N 

yn 

•— 

— H 

ro 

(N 

(N 

O 

d 

d 

d 

d 

00 

(N 

00 

r-- 

o 

Tj- 

(N 

(N 

O 

o 

ro 

o 

o 

o 

d 

d 

d 

d 

d 

d 

d 

£ o- 

i2"’ 


c 

o 


PJ 

< 

Q 

J 

5 

z 

5 

u, 


§ 'C 

■ij 

t:? 

^ 5 

3 'c: 
to  ^ ^ 


2 

^3 

tj 


C ^3  ^ 

■a  N 

ji  ^ a 

^ ^ -ii 

to  LO 


Cs  t 

I 2 

a -2 
-c 

&•  •- 

z E 

"S.  I 


3 

1?  -2 

w s; 

•a  ^ 

5 I, 

5u  3 

3 ^a 
•2  ^ 
3 3 
t3 
C >- 
C 3 
0,  O 


Habitat  (OA  = open  area;  ED  = edge;  FU  = forest  understory;  FM  = forest  mid-levels;  FC  = forest  canopy). 

Feeding  habit  (OM  = omnivore;  FR  = frugivore;  NE  = nectarivore;  CA  = carnivore;  GI  = generalized  insectivore;  LI  = leaf  insectivore;  T1  = trunk  insectivore)  is  shown  for  each  species. 


Wilson  Bull.,  111(3),  1999,  pp.  415-420 


DO  MAMMALIAN  NEST  PREDATORS  FOLLOW  HUMAN  SCENT 
TRAILS  IN  THE  SHORTGRASS  PRAIRIE? 

SUSAN  K.  SKAGEN,'  2 THOMAS  R.  STANLEY,'  AND  M.  BETH  DILLON' 


ABSTRACT. — Nest  predation,  the  major  cause  of  nest  failure  in  passerines,  has  exerted  a strong  influence  on 
the  evolution  of  life  history  traits  of  birds.  Because  human  disturbance  during  nest  monitoring  may  alter  predation 
rates,  we  investigated  whether  human  scent  affected  the  survival  of  artificial  ground  nests  in  shortgrass  prairie. 
Our  experiment  consisted  of  two  treatments,  one  in  which  there  was  no  attempt  to  mask  human  scent  along 
travel  routes  between  artificial  nests,  and  one  in  which  we  masked  human  scent  with  cow  manure,  a scent 
familiar  to  mammalian  predators  in  the  study  area.  We  found  no  evidence  that  human  scent  influenced  predation 
rates,  nor  that  mammalian  predators  followed  human  trails  between  nests.  We  conclude  that  scent  trails  made 
by  investigators  do  not  result  in  lower  nesting  success  of  passerines  of  the  shortgrass  prairie  where  vegetation 
trampling  is  minimal,  mammalian  predators  predominate,  and  avian  predators  are  rare.  Received  9 Nov  1998, 
accepted  10  Feb.  1999. 


Predation  has  exerted  a strong  influence  on 
the  evolution  of  habitat  selection  and  life  his- 
tory traits  for  many  avian  species  (Martin 
1993b).  Research  on  a broad  array  of  ecolog- 
ical topics  requires  estimates  of  avian  fecun- 
dity. Because  nest  predation  is  the  major  cause 
of  nest  failure  in  passerines  (Ricklefs  1969; 
Martin  1992,  1993a,  b),  researchers  have  fre- 
quently expressed  concerns  that  monitoring 
might  artificially  increase  predation  rates 
(Mayfield  1975,  Major  1990,  Gotmark  1992). 

Predators  might  be  attracted  to  nests  by  vi- 
sual cues,  such  as  the  presence  of  researchers, 
trampling  of  vegetation,  increased  activity  of 
parent  birds,  and  by  olfactory  cues.  Mamma- 
lian predators  are  thought  to  follow  tracks  in 
the  vegetation  and  to  respond  to  human  scent 
along  the  trails  or  at  the  nests  (Creighton 
1971,  Wilson  1976,  Nol  and  Brooks  1982, 
Gotmark  1992,  Whelan  et  al.  1994).  In  a re- 
view paper  on  investigator  bias,  Gotmark 
(1992)  concluded  there  was  little  or  no  evi- 
dence that  researcher  disturbance  increased 
mammalian  predation  rates.  Of  three  studies 
that  have  directly  addressed  whether  human 
scent  increases  mammalian  predation  rates 
(Keith  1961,  Macivor  et  al.  1990,  Whelan  et 
al.  1994),  one  (Whelan  et  al.  1994)  supported 
the  hypothesis.  Even  though  evidence  is  scant, 
the  use  of  rubber  boots  and  gloves  is  widely 


' U.S.  Geological  Survey,  Biological  Resources  Di- 
vision, Midcontinent  Ecological  Science  Center,  Fort 
Collins,  CO  80525-3400. 

^ Corresponding  author; 

E-mail;  susan_skagen(g> usgs.gov 


recommended  to  alleviate  the  potential  prob- 
lem of  human  scent  leading  to  bird  nests  (Nol 
and  Brooks  1982,  Yahner  et  al.l993.  Major 
and  Kendal  1996). 

Artificial  bird  nests  have  been  widely  used 
in  predation  studies  (e.g.,  Gottfried  and 
Thompson  1978,  Yahner  and  Wright  1985, 
Yahner  et  al.  1993).  Despite  problems  with  in- 
terpretation of  results  (Major  and  Kendal 
1996),  they  remain  a useful  tool  for  testing 
predation  theories.  We  conducted  an  experi- 
ment using  artificial  ground  nests  in  a short- 
grass prairie  where  the  primary  nest  predators 
are  mammals  and  human  presence  is  rare.  Our 
objective  was  to  test  if  human  scent  increased 
the  rates  of  predation  on  shortgrass  prairie 
ground  nesting  birds  by  comparing  two  meth- 
ods of  experimenter  travel  between  nests. 

The  purpose  of  our  study  was  to  determine 
the  most  expedient  technique  for  ongoing 
breeding  bird  studies  in  the  shortgrass  prairie. 
We  do  not  intend  to  make  inferences  from  this 
study  to  other  ecosystems  and  predator  com- 
munities. Because  breeding  systems  vary  in 
predator  communities,  predator  behavior,  ex- 
posure to  human  presence,  vegetation  struc- 
ture, and  nest  position,  many  systems  need  to 
be  evaluated  before  we  can  fully  understand 
the  effect  of  human  scent  on  predation  rates. 

METHODS 

We  conducted  this  experiment  in  July  1997  on  Paw- 
nee National  Grassland.  7 km  northwest  of  Briggsdale, 
Weld  County,  Colorado  (40°  41'  N,  104°  24'  W).  The 
259  ha  tract  of  grazed  shortgrass  prairie  is  character- 
ized by  short  and  mid-grasses,  cacti  {Opunlia  sp.). 


415 


416 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


forbs,  and  patchy  areas  of  yucca  {Yucca  glauca).  Com- 
mon ground  nesting  passerines  in  the  vicinity  include 
Homed  Lark  {Eremophila  alpestris).  Lark  Bunting 
(Calamospiza  melancorys),  McCown’s  Longspur  (Cal- 
cariu.s  mccownii).  Chestnut-collared  Longspur  {Cal- 
carius  omatus),  and  Western  Meadowlark  {Sturnella 
neglecta).  Potential  predators  of  ground-nesting  birds 
include  thirteen-lined  ground  squirrel  {Spermophilus 
tridecemlineatu.s),  deer  mouse  (Peromyscus  manicu- 
latu.s),  northern  grasshopper  mouse  {Onychomys  leu- 
cogaster),  coyote  (Canis  latrans),  swift  fox  (Vulpes  ve- 
lox),  raccoon  (Procyon  lotor),  long-tailed  weasel  {Mus- 
tela  frenata),  badger  {Taxidea  taxus),  striped  skunk 
{Mephitis  mephitis),  and  several  snake  species. 

We  placed  100  artificial  nests  along  two  transects 
established  800  m apart  in  similar  habitat.  Each  tran- 
sect contained  25  lines  perpendicular  to  the  transect 
and  alternating  in  opposite  directions  at  50  m intervals 
so  that  adjacent  lines  were  100  m apart.  Each  line  con- 
tained two  nests;  one  at  100  m (Nest  A)  and  the  second 
at  200  m (Nest  B)  from  the  transect.  From  Nest  B we 
walked  an  additional  100  m so  that  both  nests  on  the 
line  were  treated  equally.  To  aid  in  relocating  nests  we 
noted  any  distinguishing  features  around  the  nest  and 
tied  surveyor’s  tape  to  low-growing  vegetation  10  m 
from  each  nest,  a distance  not  associated  with  in- 
creased predation  rates  (Major  and  Kendal  1996). 

To  test  if  human  scent  trails  influenced  predation 
rates,  we  subjected  artificial  nests  to  two  treatments. 
In  Treatment  1 (human  scent),  we  wore  leather  boots 
and  made  no  attempts  to  mask  human  scent  while 
walking  between  nests.  In  Treatment  2 (masked  scent), 
we  masked  human  scent  with  a scent  familiar  to  po- 
tential predators  in  the  study  area  by  wearing  rubber 
boots  that  were  sprayed  with  a cow  manure  tea  (fresh 
cow  manure  steeped  in  water,  in  a 1:3  mixture  for  at 
least  12  hours)  at  the  beginning  of  each  line.  Treatment 
types  were  randomly  assigned  to  the  50  lines  (25  lines 
per  transect);  both  nests  on  a line  received  the  same 
treatment.  Because  we  specifically  wanted  to  deter- 
mine an  effect  of  scent  trails,  we  wore  vinyl  craft 
gloves  (standard  field  practice)  while  handling  eggs  in 
both  treatments  to  minimize  human  scent  on  the  eggs. 

Nests  consisted  of  a scrape  on  the  ground  and  con- 
tained two  fresh  Japanese  quail  (Cnturnix  japonica) 
eggs  (mean  length  X width,  3.3  X 2.6  cm,  n = 20) 
and  one  clay  egg  (2.2  X 1.5  cm,  n = 20).  Scrapes 
were  created  using  the  broad  end  of  a large  wooden 
tongue  depressor.  While  wearing  rubber  gloves  we 
constructed  clay  eggs  out  of  soft  modeling  compound 
(Sculpey  III  brand)  to  approximate  the  size  of  Lark 
Bunting  eggs.  Clay  eggs  aided  in  the  identification  of 
nest  predators  and  enabled  us  to  record  predation  by 
predators  too  small  to  handle  quail  eggs  (i.e.,  small 
rodents;  Major  and  Kendal  1996). 

Nests  were  set  out  on  9 July  1997  and  checked  three 
days  later,  a time  interval  during  which  we  expected 
50%  of  the  ne.sts  to  survive  ba.sed  on  preliminary  re- 
sults of  trials  using  artificial  nests  constructed  in  the 
same  manner.  Although  .several  studies  used  longer  tri- 
al intervals,  we  expected  that  our  ability  to  detect  dif- 


TABLE  1.  Predation  outcomes  for  25  lines  receiv- 
ing the  human  scent  treatment  and  25  lines  receiving 
the  masked  scent  treatment. 


Predation  outcome 

( I = depredated.  Number  of  lines 

0 = survived)  with  outcome 


i 

Nest  A 

Nest  B 

Human  scent 
(«,) 

Masked  scent 
("!/) 

1 

0 

0 

6 

5 

2 

1 

0 

5 

8 

3 

0 

1 

9 

7 

4 

1 

1 

5 

5 

ferences  would  be  diminished  if  nearly  all  nests  were 
depredated.  Nests  were  classified  as  intact  or  disturbed 
based  on  signs  of  disturbance  to  either  quail  or  clay 
eggs.  Nests  were  considered  disturbed  if  quail  eggs 
were  missing,  broken,  or  moved,  or  if  clay  eggs  were 
missing,  moved,  or  had  tooth  impressions.  We  collect- 
ed extant  clay  eggs  for  examination  and  identification 
of  any  diagnostic  marks.  We  classified  markings  on  the 
clay  eggs  as  rodent,  non-rodent,  insect,  or  unknown  by 
comparing  them  with  known  tooth  impressions  made 
from  skulls  in  the  zoology  collection  at  Colorado  State 
University,  Fort  Collins,  Colorado.  In  the  absence  of 
other  signs  of  disturbance,  nests  containing  clay  eggs 
with  only  insect  marks  were  considered  intact. 

The  data  from  this  experiment  are  counts  and  can 
be  arranged  into  an  / X j contingency  table  (Table  1), 
where  i denotes  the  predation  outcome  and  j denotes 
the  treatment  (i.e.,  human  or  masked  scent).  While  the 
cell  probabilities  for  such  tables  are  commonly  mod- 
eled and  estimated  using  standard  loglinear  models 
(e.g.,  Agresti  1990),  reparameterization  of  the  under- 
lying multinomial  model  can  lead  to  loglinear  models 
that  are  difficult  to  construct  or  difficult  to  interpret. 
In  this  study  we  reparameterize  the  underlying  multi- 
nomial model  to  address  the  following  specific  ques- 
tions: (1)  do  predation  probabilities  differ  for  nests  on 
a line  because  of  differences  in  their  proximity  to  the 
transect,  (2)  do  predation  probabilities  for  nests  differ 
because  of  differences  in  human  and  masked  scent 
treatments,  and  (3)  is  there  evidence  that  predators  fol- 
lowed the  investigator’s  trail  between  nests  on  a line. 
Hence,  instead  of  using  a loglinear  modeling  approach, 
we  derived  parameter  estimates  and  constructed  hy- 
pothesis tests  using  classical  maximum  likelihood 
methods  (e.g.,  Lar.sen  and  Marx  1986:261).  The  gen- 
eral procedure  was  to  ( 1 ) construct  the  appropriate 
likelihood  function  for  the  data,  (2)  derive  estimators 
and  compute  estimates  for  parameters  under  the  model, 
(3)  evaluate  the  likelihood  function  at  the  maximum 
likelihood  parameter  estimates  to  obtain  the  deviance 
(here  we  omit  the  term  for  the  saturated  model  and 
define  deviance  as  -2  X (log-likelihood),  and  then  (4) 
test  specific  hypotheses  using  likelihood  ratio  tests  for 
nested  models  (Agresti  1990:21  1).  The  models  used  in 
this  study  are  presented  in  the  Appendix.  SAS  statis- 


Skcificn  el  al.  • MAMMALIAN  NEST  PREDATORS 


417 


TABLE  2.  Three  candidate  models  for  estimating 
predation  probabilities  constructed  using  (A.l). 

Model 

Pariinielers  and  Con.straint.s 

Deviance 

1 

Pa>  Pb, 

137.182 

2 

II 

II 

138.549 

3 

Pa  Pb  Q Cg 

138.589 

tical  software  (version  6.12  on  an  IBM-eompatible  mi- 
crocomputer; SAS  Institute  Inc.  1990)  was  used  for  all 
computations.  Values  reported  are  means  (±SE). 

RESULTS 

During  the  trial,  49  of  100  nests  were  dis- 
turbed, 24  from  the  human  scent  treatment 
and  25  from  the  masked  scent  treatment.  Dis- 
turbance to  quail  eggs  was  apparent  in  45  of 
49  (92%)  nests;  eggs  were  missing  from  12 
nests,  broken  in  an  additional  7 nests,  and 
moved  in  an  additional  26  nests.  In  four  nests, 
the  quail  eggs  were  undisturbed,  yet  clay  eggs 
were  either  moved  or  had  tooth  impressions. 
Clay  eggs  were  undisturbed  in  only  two  nests 
with  disturbed  (broken)  quail  eggs.  Rodent 
tooth  impressions  were  identified  on  22  clay 
eggs,  and  non-rodent  impressions  on  one  clay 
egg.  Clay  eggs  were  missing  from  20  nests. 
Quail  eggs  and  clay  eggs  that  were  moved 
were  displaced  an  average  of  31.3  cm  (±  9.10, 
median  1 cm,  range  0.5-330  cm,  n = 41)  and 
43.2  cm  (±  15.07,  median  20  cm,  range  0.5- 
250  cm,  n = 20)  from  their  original  positions, 
respectively. 

Predation  outcomes  for  the  two  treatments 
are  summarized  in  Table  1.  In  general,  few 
differences  between  the  treatments  were  evi- 
dent. In  Table  2,  the  three  candidate  models 
constructed  under  A.l  are  presented,  along 
with  their  deviance.  The  likelihood  ratio  test 
between  model  2 and  model  1,  which  tests  for 
differences  in  predation  rates  between  nest  A 
and  nest  B caused  by  their  proximity  to  the 
transect,  had  a P-value  of  0.505  (y^  = 1.37, 
df  = 2).  Hence,  there  appears  to  be  no  effect 
as  a result  of  proximity  to  the  transect.  The 
likelihood  ratio  test  between  model  3 and 
model  2,  which  tests  for  differences  in  pre- 
dation rates  caused  by  differences  in  the  hu- 
man and  masked  scent  treatments,  had  a P- 
value  of  0.841  (x^  = 0.04,  df  = 1).  Hence,  we 
conclude  there  was  no  treatment  effect. 

For  the  two-parameter  model  in  A. 2 (i.e.,  p, 


p'),  which  allows  unconditional  and  condi- 
tional predation  probabilities  for  nests  on  a 
line  to  differ,  we  get  a deviance  of  97.094. 
When  we  impose  the  constraint  p = p'  (giving 
us  a one-parameter  model  that  is  equivalent  to 
model  3 in  Table  2),  we  get  a deviance  of 
98.387.  The  likelihood  ratio  test  for  these 
models,  which  tests  whether  predators  were 
following  the  investigator’s  trail  between  nests 
on  a line,  had  a P-value  of  0.256  (x^  = 1.29, 
df  = 1).  Consequently,  we  conclude  predators 
did  not  follow  the  investigator’s  trail  between 
nests  on  a line. 

The  one  parameter  models  from  A.l  and 
A. 2 are  mathematically  equivalent  and,  based 
on  the  likelihood  ratio  tests,  are  the  appropri- 
ate models  to  use  for  parameter  estimation. 
Hence,  the  estimated  three-day  predation 
probability  for  nests  in  this  study  was  0.49 
(±0.050),  which  gives  an  estimated  daily  sur- 
vival probability  of  0.80  (±0.026). 

DISCUSSION 

We  found  no  evidence  that  human  scent 
trails  to  nests  altered  predation  rates  on  arti- 
ficial nests  in  grasslands  where  the  main  pred- 
ators are  small  mammals,  nor  did  we  find  ev- 
idence that  predators  were  more  likely  to  dep- 
redate nests  on  the  same  trail.  A learned  as- 
sociation of  human  scent  with  food  is  unlikely 
because  human  presence  is  rare  throughout 
much  of  our  study  area.  Rather,  the  scent 
would  be  novel  to  small  mammals  of  the  re- 
gion. We  found  no  evidence  that  novel  scent 
was  an  attractant  to  predators  in  our  region. 

Whether  human  scent  is  an  attractant  or  de- 
terrent to  predators  has  been  a topic  of  spec- 
ulation (Creighton  1971,  Mayfield  1975,  Wil- 
son 1976,  Gotmark  1992)  that  has  been  di- 
rectly tested  in  only  three  other  studies.  Re- 
sults differ  between  studies.  Keith  (1961) 
reported  no  effect  of  human  scent  on  survival 
of  artificial  duck  nests  in  wetlands  with  pre- 
dominately mammalian  predators.  Macivor 
and  coworkers  (1990)  found  that  red  fox  {Vid- 
pes  vulpes)  avoided  human  scent  associated 
with  experimental  plover  nests  along  a beach. 
In  contrast,  Whelan  and  coworkers  (1994)  re- 
ported raccoons  in  a forested  system  preying 
on  nests  with  human  scent  and  novel  scent 
more  frequently  than  nests  with  no  scent  or 
familiar  scent.  In  another  study  evaluating  the 
influence  of  familiar  and  novel  scents,  Clark 


418 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


and  Wobeser  (1997)  determined  that  a novel 
odor  (lemon  juice  and  ground  ginger  root)  did 
not  affect  survival  of  artificial  waterfowl 
nests.  Collectively,  these  studies  represent  a 
broad  variability  in  predator  community  com- 
position, predator  behavior,  exposure  to  hu- 
man presence,  vegetation  structure,  and  nest 
placement. 

Evidence  of  other  aspects  of  investigator 
bias  on  predation  rates  is  also  equivocal.  Fre- 
quency of  nest  visits  had  no  effect  on  preda- 
tion rates  of  artificial  nests  in  several  studies 
(Bowen  et  al.  1976,  Gottfried  and  Thompson 
1978,  Erikstad  et  al.  1982,  Maclvor  et  al. 
1990,  Gotmark  1992,  Mankin  and  Warner 
1992),  but  did  in  two  studies  (Major  1990, 
Esler  and  Grand  1993),  presumably  because 
of  vegetation  trampling.  In  our  study,  we  did 
not  evaluate  the  effects  of  frequency  of  visi- 
tation nor  vegetation  trampling.  In  the  short- 
grass  prairie,  the  one  visit  to  artificial  nests 
during  construction  resulted  in  minimal  veg- 
etation trampling. 

One  criticism  of  artificial  nests  is  that  they 
often  contain  only  quail  eggs  and  that  small 
predators  unable  to  handle  the  quail  eggs  may 
be  under-represented  (Major  and  Kendal 
1996).  We  addressed  this  problem  by  consid- 
ering nests  disturbed  when  eggs  were  moved 
as  well  as  broken  or  removed,  and  by  using 
smaller  clay  eggs  in  addition  to  quail  eggs.  We 
found,  however,  that  only  two  nests  would 
have  been  misclassified  as  undisturbed  if  only 
quail  eggs  had  been  used. 

We  conclude  that  the  procedures  we  used 
while  visiting  nests  are  unlikely  to  contribute 
to  reduced  nesting  success  of  passerines  of  the 
shortgrass  prairie  where  vegetation  trampling 
is  minimal,  mammalian  predators  predomi- 
nate, and  avian  predators  are  rare.  Our  con- 
clusion is  consistent  with  Gotmark  (1992) 
who  surmises  that  passerines  are  less  sensitive 
to  investigator  disturbance  than  other  groups 
of  birds,  scent  having  less  effect  than  vege- 
tation trampling,  and  increases  in  predation  in 
response  to  human  cues  more  common  for 
avian  than  mammalian  or  reptilian  predators. 
We  recommend  that  investigators  continue  to 
evaluate  whether  human  scent  alters  predation 
rates  in  avian  breeding  systems  and  not  make 
inappropriate  inferences  across  systems.  Hu- 
man scent  studies  that  identify  and  describe 
the  predator  communities,  habitat  structure. 


and  human  influence  will  ultimately  contrib- 
ute to  better  understanding  of  observer  bias  in 
research. 

ACKNOWLEDGMENTS 

We  thank  A.  Yackel  Adams  and  R.  Adams  for  field 
assistance,  J.  Miller  for  suggestions  on  experimental 
design,  B.  A.  Wunder  for  access  to  the  zoology  col- 
lection at  Colorado  State  University,  and  D.  T.  Arm- 
strong for  information  on  mammals  of  the  study  area. 
J.  Bradley  and  M.  Laubhan  reviewed  a draft  of  the 
manuscript.  This  study  was  funded  by  the  Biological 
Resources  Division  of  the  U.  S.  Geological  Survey. 

LITERATURE  CITED 

Agresti,  a.  1990.  Categorical  data  analysis.  John  Wi- 
ley and  Sons,  New  York. 

Bowen,  D.  E.,  R.  J.  Robel,  and  R G.  Watt.  1976. 
Habitat  and  investigators  influence  artificial 
ground  nest  losses.  Trans.  Kans.  Acad.  Sci.  79; 
141-148. 

Clark,  R.  G.  and  B.  K.  Wobeser.  1997.  Making  sense 
of  scents:  effects  of  odour  on  survival  of  simulat- 
ed duck  nests.  J.  Avian  Biol.  28:31—37. 
Creighton,  P.  D.  1971.  Nesting  of  the  Lark  Bunting 
in  northcentral  Colorado.  U.S.I.B.P.  Grassland  Bi- 
ome  Tech.  Rep.  68.  Colorado  State  Univ.,  Port 
Collins. 

Erikstad,  K.  E.,  R.  Blom,  and  S.  Myrberget.  1982. 
Territorial  Hooded  Crows  as  predators  on  Willow 
Ptarmigan  nests.  J.  Wildl.  Manage.  46:109-114. 
Esler,  D.  and  J.  B.  Grand.  1993.  Factors  influencing 
depredation  of  artificial  duck  nests.  J.  Wildl.  Man- 
age. 57:244-248. 

Gotmark,  E 1992.  The  effects  of  investigator  distur- 
bance on  nesting  birds.  Cuix  Ornithol.  9:63-104. 
Gottfried,  B.  M.  and  C.  F.  Thompson.  1978.  Exper- 
imental analysis  of  nest  predation  in  an  old-field 
habitat.  Auk  95:304-312. 

Keith,  L.  B.  1961.  A study  of  waterfowl  ecology  on 
small  impoundments  in  southeastern  Alberta. 
Wildl.  Monogr.  6:1-86. 

Larsen,  R.  J.  and  M.  L.  Marx.  1986.  An  introduction 
to  mathematical  statistics  and  its  applications,  sec- 
ond ed.  Prentice-Hall,  Englewood  Cliffs,  New  Jer- 
sey. 

MacIvor,  L.  H.,  S.  M.  Melvin,  and  C.  R.  Griffin. 
1990.  Effects  of  research  activity  on  Piping  Plover 
nest  predation.  J.  Wildl.  Manage.  54:443-447. 
Major,  R.  E.  1990.  The  effect  of  human  observers  on 
the  intensity  of  nest  predation.  Ibis  132:608-612. 
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,  R C.  and  R.  W.  Warner.  1992.  Vulnerability 
of  ground  nests  to  predation  on  an  agricultural 
habitat  island  in  east-central  Illinois.  Am.  Midi. 
Nat.  128:281-291. 

Martin,  T.  E.  1992.  Breeding  productivity  consider- 


Skai>en  el  al.  • MAMMALIAN  NEST  PREDATORS 


419 


ations:  what  are  tlie  appropriate  habitat  features 
for  management?  Pp.  455-473  in  Ecology  and 
conservation  of  Neotropical  migrant  land  birds  (J. 
M.  Hagan  and  D.  W.  Johnston,  Eds.).  Smithsonian 
Institution  Press,  Washington,  D.C. 

Martin,  T.  E.  1993a.  Nest  predation  among  vegetation 
layers  and  habitats:  revising  the  dogmas.  Am.  Nat. 
141:897-913. 

Martin,  T.  E.  1993b.  Nest  predation  and  nest  sites: 
new  perspectives  and  old  patterns.  BioScience  43: 
523-532. 

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

Nol,  E.  and  R.  J.  Brooks.  1982.  Effects  of  predator 
exclosures  on  nesting  outcome  of  Killdeer.  J.  Field 
Ornithol.  53:263—268. 

Ricklefs,  R.  E.  1969.  An  analysis  of  nesting  mortality 
in  birds.  Smithson.  Contrib.  Zool.  9:1—48. 

SAS  Institute  Inc.  1990.  SAS  language  reference, 
version  6,  first  ed.  SAS  Institute,  Inc.,  Cary,  North 
Carolina. 

Whelan,  C.  J.,  M.  L.  Kilger,  D.  Robson,  N.  Hallyn, 
AND  S.  Dilger.  1994.  Effects  of  olfactory  cues  on 
artificial-nest  experiments.  Auk  111:945-952. 
Wilson,  J.  K.  1976.  Nesting  success  of  the  Lark  Bun- 
ting near  the  periphery  of  its  breeding  range. 
Kans.  Ornithol.  Soc.  Bull.  27:13-22. 

Yahner,  R.  H.  and  a.  W.  Wright.  1985.  Depredation 
on  artificial  ground  nests:  effects  of  edge  and  plot 
age.  J.  Wildl.  Manage.  49:158-161. 

Yahner,  R.  H.,  C.  G.  Mahan,  and  C.  A.  Delong. 
1993.  Dynamics  of  depredation  on  artificial 
ground  nests  in  habitat  managed  for  Ruffed 
Grouse.  Wilson  Bull.  105:172-179. 

APPENDIX 

For  a particular  line  receiving  either  the  hu- 
man scent  or  the  masked  scent  treatment,  four 
outcomes  are  possible  (Table  1).  If  we  denote 
these  outcomes  by  / (/  = 1,  . . . , 4),  let  the 
probability  of  the  /-th  outcome  be  tt,  and  y, 
(respectively)  for  human  and  masked  scent 
treatments,  and  let  n,  and  m,  (respectively)  be 
the  number  of  lines  for  which  the  /-th  outcome 
was  observed  for  human  and  masked  scent 
treatments,  then  the  probability  of  the  ob- 
served data  is  the  product  of  two  multinomi- 
als: 

4 4 

;=1  1=1 

where  C,  and  C2  are  multinomial  coefficients, 
2,77,  = 2,y,-  = 1,  and  2,/2,  = 2,m,  = 25.  Under 
the  assumption  that  lines  and  nests  on  a line 
are  independent  (the  latter  assumption  is  test- 
ed below  using  A. 2),  we  can  reparameterize 
this  model  in  terms  of  the  probability  nest  A 


and  nest  B were  depredated  for  human  scent 
treatments  and  p/d  and  the  probability  nest 
A and  nest  B were  depredated  for  masked 
scent  treatments  (c^  and  Cn),  to  obtain  a model 
with  likelihood  function  proportional  to: 

L(1  - Pa){\  -Pb)]'''[Pa(1  -Pfi)''4(l  - Pa)Pb]''^ 

X - Q)(1  - Cb)Y"[ca{\  - Cb)]"'2 

X [(1  - Q)cB]"'3[c^Ca]'"L  (A.l) 

We  derived  estimators  for  p^,  pg,  c^,  and  Cg 
using  standard  maximum  likelihood  methods 
(Larsen  and  Marx  1986:261).  Differences  in 
predation  probabilities  between  nests  A and  B 
due  to  differences  in  proximity  to  the  transect, 
and  differences  in  predation  probabilities  due 
to  differences  in  treatments,  were  tested  by 
constraining  parameters  in  A.l  to  obtain  the 
appropriate  submodels,  and  then  performing 
likelihood  ratio  tests.  In  the  first  submodel  pa- 
rameters were  constrained  so  that,  within  a 
treatment,  predation  probabilities  for  nests  A 
and  B were  constant  (i.e.,  p^  = pg  and  = 
Cg).  In  the  second  submodel  parameters  were 
constrained  so  that  predation  probabilities 
were  constant  between  nests  A and  B and 
across  treatments  (i.e..  Pa  — Pb  ~ = Cg,  see 

Table  1). 

In  an  effort  to  determine  whether  predators 
were  following  the  human  trail  between  nests, 
one  additional  model  was  constructed.  This 
model  assumed  that  predation  probabilities 
among  nests  on  a line  and  among  treatments 
did  not  differ,  but  allowed  the  unconditional 
and  conditional  predation  probabilities  of 
nests  on  a line  to  differ.  Here,  the  conditional 
predation  probability  is  the  probability  nest  A 
would  be  depredated  given  nest  B had  already 
been  depredated,  or  the  converse.  If  we  denote 
the  unconditional  predation  probability  by  p 
and  the  conditional  predation  probability  by 
p' , then  the  probability  neither  nest  on  a line 
is  depredated  is  given  by  (1  — p)(l  — p)  and 
the  probability  both  nests  on  a line  are  dep- 
redated is  given  by  pp' . To  obtain  the  proba- 
bility that  only  one  nest  on  a line  is  depre- 
dated, we  exploit  the  fact  that  the  cell  proba- 
bilities for  a multinomial  must  sum  to  one. 
Hence,  the  probability  that  only  one  nest  on  a 
line  is  depredated  is  given  by  1 — (1  — p)(l 
— p)  ~ pp'  which,  after  some  algebraic  ma- 
nipulation, yields  the  intuitively  reasonable 


420 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


p{\  — p')  + (1  — p)p.  Thus,  the  resulting 
model  has  likelihood  function  proportional  to: 

[(1  - p)ii  - p)Y'’r\w  - p')  + 

(1  - (A.2) 


where  Uj  and  ntj  (i  = 1,  . . . , 4)  are  as  defined 
above.  A test  for  Hq:  p — p'  versus  H^:  p ^ 
p'  was  constructed  using  a likelihood  ratio 
test. 


Short  Communications 


Wilson  Bull.,  111(3),  1999,  pp.  421-422 


Christmas  Shearwater  Egg  Dimensions  and  Shell  Characteristics  on 
Laysan  Island,  Northwestern  Hawaiian  Islands 

G.  C.  Whittow^'^  and  M.  B.  Naughton^ 


ABSTRACT. — The  mean  fresh  egg  mass  of  Christ- 
mas Shearwaters  (Puffinus  nativitatis)  on  Laysan  Is- 
land, in  the  Northwestern  Hawaiian  Islands,  was  44.9 
± 3.4  (SD)  g,  and  the  mean  egg  volume  was  42.3  ± 
2.9  cm-’.  The  measured  length  and  breadth  of  the  eggs, 
the  shell  mass,  shell  thickness,  and  number  of  pores  in 
the  shell  were  within  10%  of  predictions  for  procel- 
lariiform  birds,  based  on  fresh  egg  mass  or  on  both 
fresh  egg  mass  and  incubation  period.  These  data  con- 
form with  evidence  that  there  are  few  allometric  dif- 
ferences between  the  eggs  of  tropical  Procellariiformes 
and  those  of  Procellariiformes  from  higher  latimdes. 
Received  19  Oct.  1998,  accepted  25  Feh.  1999. 


The  Christmas  Shearwater  (Puffinus  nativ- 
itatis) is  a tropical  procellariiform  seabird  that 
breeds  on  islands  in  the  central  North  and 
South  Pacific  oceans  (Warham  1990).  It  has 
been  little  studied,  perhaps  because  the  spe- 
cies is  not  abundant  anywhere  (Shallenberger 
1984).  The  purpose  of  the  present  note  is  to 
report  data  for  Christmas  Shearwater  eggs  col- 
lected on  Laysan  Island  (28°  12'  N;  177°  20' 
W)  in  the  northwestern  Hawaiian  Islands. 

We  measured  egg  volume  by  weighing  the 
egg  in  air  and  again  submerged  in  water.  From 
the  difference  between  the  mass  in  air  and  in 
water,  and  the  density  of  water,  egg  volume  can 
be  determined  in  accordance  with  Archimedes’ 
Principle  (Rahn  et  al.  1976).  We  determined 
fresh  egg  mass  by  weighing  the  egg  after  fiUing 
the  airceU  with  distilled  water  (the  mass  loss  of 
the  egg  during  incubation  being  entirely  the  re- 
sult of  the  loss  of  water  vapor;  Grant  et  al. 
1982),  and  we  measured  both  the  length  and 
breadth  of  the  egg  with  a dial  caliper.  We  ob- 
tained the  shell  mass,  its  thickness,  and  we 


■ Dept,  of  Physiology,  John  A.  Bums  School  of  Med- 
icine, Univ.  of  Hawaii,  Honolulu,  HI  96822;  E-mail: 
whittowg@jabsom.biomed.hawaii.edu 
^ Corresponding  author. 

’ U.S.  Fish  & Wildlife  Service,  Prince  Kuhio  Fed- 
eral Building,  Honolulu,  HI  96850. 


counted  the  number  of  pores  in  the  shell  of  ran- 
domly-selected sub-samples  of  eggs  after  drying 
the  shells  in  a desiccator.  We  measured  pore 
density  using  the  procedure  described  by  Tyler 
(1953)  and  Roudybush  and  coworkers  (1980). 

The  mean  fresh  egg  mass  of  18  Christmas 
Shearwater  eggs  was  44.9  ± 3.4  (SD)  g. 
Knowledge  of  the  fresh  egg  mass  provides  an 
opportunity  to  compare  some  of  the  other  mea- 
sured values  (Table  1)  with  predictions  for  pro- 
cellariiform eggs,  based  on  the  mass  of  their 
freshly  laid  eggs  (Rahn  and  Whittow  1988). 
There  are  no  predictive  equations  for  the  vol- 
ume of  the  eggs  of  Procellariiformes,  but  mea- 
sured egg  lengths  and  breadths  were  similar 
(100.3%  and  96.0%,  respectively)  to  predicted 
values  (Table  1).  Measured  shell  mass  and  shell 
thickness  were  also  similar  (94.7%  and 
105.6%,  respectively)  to  predictions  (Table  1). 
Rahn  and  Whittow  (1988)  presented  two  pre- 
dictive equations  for  the  total  number  of  pores 
in  the  eggshell.  Both  require  the  calculation  of 
the  surface  area  of  the  egg  from  its  mass  (Tul- 
lett  and  Board  1977),  which  is  then  multiplied 
by  the  measured  pore  density  (Table  1).  The 
resulting  estimated  total  number  of  pores  in  the 
shell  of  a Christmas  Shearwater’s  egg  is  3103 
pores.  This  value  falls  between  the  two  pre- 
dicted values  (2963  and  3584).  The  predicted 
values  were  both  based  on  the  incubation  pe- 
riod as  well  as  on  the  fresh  egg  mass;  for  this 
purpose,  we  used  an  incubation  period  of  53 
days  (Byrd  et  al.  1983,  Naughton  1983)  and  a 
fresh  egg  mass  of  44.9  g. 

The  measured  values  for  the  eggs  and  egg- 
shells of  the  tropical  Christmas  Shearwater  are 
close  to  predictions  for  Procellariiformes  in 
general.  This  finding  supports  evidence  that 
there  are  few  differences  in  the  allometric  re- 
lationships of  eggs  and  eggshells  of  Procel- 
lariiformes between  tropical  and  non-tropical 
species.  In  contrast,  there  are  substantial  dif- 
ferences between  Procellariiformes  and  other 


421 


422 


THE  WILSON  BULLETIN  • Vol.  HI.  No.  3,  September  1999 


TABLE  1.  Measured  and  predicted  values  for  the  eggs  and  eggshells  of  Christmas  Shearwaters  on  Laysan 
Island.  The  mean  measured  values  (±1  SD)  are  shown;  n = the  numbers  of  observations.  Predicted  values  are 
calculated  following  Rahn  and  Whittow  (1988). 


Measured  {x  ± SD) 


n 


Predicted 


Volume  (cm^) 

42.3  ± 2.9 

19 

Length  (cm) 

55.6  ± 2.0 

22 

55.5 

Breadth  (cm) 

38.3  ± 1.3 

22 

39.9 

Eggshell 

Mass  (g) 

2.9  ± 0.2 

8 

3.0 

Thickness  (mm) 

0.3  ± 0.0 

11 

0.3 

Pore  density  [pores  (cm0“'] 

59.8  ± 4.7 

6 

orders  of  seabirds  in  this  regard  (Whittow 
1984,  Ar  and  Rahn  1985). 

ACKNOWLEDGMENTS 

We  thank  the  U.S.  Fish  and  Wildlife  Service  for 
granting  permission  to  collect  eggs  on  Laysan  Island. 

LITERATURE  CITED 

Ar,  a.  and  H.  Rahn.  1985.  Pores  in  avian  eggshells: 
gas  conductance,  gas  exchange  and  embryonic 
growth  rate.  Resp.  Physiol.  61:1-20. 

Byrd,  G.  V.,  D.  I.  Moriarty,  and  B.  G.  Brady.  1983. 
Breeding  biolgy  of  Wedge-tailed  Shearwaters  at 
Kilauea  Point,  Hawaii.  Condor  85:292—296. 
Grant,  G.  S.,  C.  V.  Paganelli,  T.  N.  Pettit,  and  G. 
C.  Whittow.  1982.  Determination  of  fresh  egg 
mass  during  incubation.  Condor  84:121-122. 
Naughton,  M.  1982.  Breeding  biology  of  the  Christ- 
mas Shearwater  (Puffinus  nativitatis)  on  Laysan 
Island,  Hawaii.  Pac.  Seabird  Group  Bull.  9:71-72. 
Rahn,  H.  and  G.  C.  Whittow.  1988.  Adaptations  to  a 


pelagic  life:  eggs  of  the  albatross,  shearwater  and 
petrel.  Comp.  Biochem.  Physiol.  91A:415— 423. 
Rahn,  H.,  C.  V.  Paganelli,  I.  C.  T.  Nisbet,  and  G.  C. 
Whittow.  1976.  Regulation  of  incubation  water 
loss  in  eggs  of  seven  species  of  terns.  Physiol. 
Zool.  49:245-259. 

Roudybush,  T,  L.  Hoffman,  and  H.  Rahn.  1980. 
Conductance,  pore  geometry,  and  water  loss  of 
eggs  of  Cassin’s  Auklet.  Condor  82:105-106. 
Shallenberger,  R.  J.  1984.  Fulmars,  shearwaters  and 
gadfly  petrels.  Pp.  42—56  in  Seabirds  of  eastern 
North  Pacific  and  Arctic  waters  (D.  Haley,  Ed.). 
Pacific  Search  Press,  Seattle,  Washington. 
Tullett,  S.  G.  and  R.  G.  Board.  1977.  Determina- 
tions of  avian  eggshell  porosity.  J.  Zool.  (Lond.) 
183:203-21 1. 

Tyler,  C.  1953.  Studies  on  egg  shells.  II:  method  for 
marking  and  counting  pores.  J.  Sci.  Food  Agric. 
4:266-272. 

Warham,  j.  1990.  The  petrels:  their  ecology  and 
breeding  systems.  Academic  Press,  London,  U.K. 
Whittow,  G.  C.  1984.  Physiological  ecology  of  incuba- 
tion in  tropical  seabirds.  Stud.  Avian  Biol.  8:47-72. 


Wilson  Bull.,  111(3),  1999,  pp.  422-424 

The  Paint-billed  Crake  Breeding  in  Costa  Rica 

David  M.  Watson'-^  and  Brett  W.  Benz' 


ABSTRACT. — We  report  a recent  observation  from 
southern  Costa  Rica  of  the  Paint-billed  Crake  {Neocre.x 
erythrops),  a little  known  species  from  eastern  and 
northern  South  America.  An  adult  and  recently 
hatched  chick  were  observed  at  close  range  in  wet 


‘ Natural  History  Museum  & Biodiversity  Research 
Center,  and  Dept,  of  Ecology  and  Evolutionary  Biol- 
ogy, The  Univ.  of  Kansas,  Lawrence,  KS  66045. 

2 Corresponding  author;  E-mail:  vergil@ukans.edu 


grassy  second-growth.  This  observation  constitutes  the 
first  record  of  the  young  of  this  species  and  represents 
the  only  breeding  record  for  Central  America.  Re- 
ceived 12  Nov.  1998,  accepted  12  Feb.  1999. 


On  5 June  1998,  at  16:30  an  adult  Paint- 
billed Crake  (Neocrex  erythrops)  was  ob- 
served, accompanied  by  a chick,  near  the 


SHORT  COMMUNICATIONS 


423 


town  of  Golfito  on  the  Pacific  coast  of  Costa 
Rica,  close  to  the  Panama  border  (8°  37'  N, 
83°  1 r W).  Observations  were  made  by 
both  authors  and  K.  Cohoon,  while  walking 
slowly  beside  the  Golfito  airstrip  amidst 
grassy  second  growth  interspersed  with  a 
row  of  large  Ficus  trees.  Behind  this  vege- 
tation was  a slow  moving  stream  with  thick- 
ets of  tall  grass  along  its  banks.  We  saw  the 
birds  from  approximately  15  m and  watched 
them  for  25  seconds  using  binoculars.  The 
adult  paused  in  the  middle  of  the  path,  even- 
tually returning  to  the  wet  grassy  second 
growth  from  which  it  had  walked.  The 
bright  yellow  bill  with  a scarlet  base,  and 
black  and  white  barred  flanks  were  clearly 
visible,  clearly  distinguishing  it  from  the 
congeneric  Colombian  Crake  (Neocrex  col- 
ombianus).  The  chick  was  covered  uniform- 
ly in  black  natal  down  and  the  tarsi  were 
dark  grey  or  horn.  Further  soft-part  colors 
were  not  noted  because  it  quickly  ran  away 
from  the  adult,  across  the  path  into  thick  un- 
dergrowth beside  the  airstrip.  Despite  sub- 
sequent visits  to  this  locality  for  several 
weeks,  we  made  no  further  observations  of 
this  species. 

Neocrex  erythrops  is  a little  known  species 
that  ranges  widely  in  eastern  and  central  South 
America;  N.  e.  olivascens  is  known  east  of  the 
Andes  from  Colombia  and  Venezuela  south  to 
Paraguay  and  Argentina,  and  N.  e.  erythrops 
from  west  of  the  Andes  in  coastal  Peru  and 
the  Galapagos  Islands  (Ripley  1977).  There 
have  been  reports  of  vagrants  within  South 
America  (Osborne  and  Beissinger  1979),  from 
suburban  areas  and  up  to  3375  m elevation 
(Remsen  and  Traylor  1983),  and  some  recent 
reports  (Tostain  et  al  1992,  Haverschmidt  and 
Mees  1994)  suggest  that  the  species  may  be 
resident  in  Surinam  and  French  Guyana.  As 
with  many  other  species  of  rail,  the  chicks  of 
this  species  are  undescribed  (Ripley  1977,  del 
Hoyo  et  al  1996).  The  uniform  black  down 
and  dark  tarsi  are  similar  to  the  young  of  other 
neotropical  rails  in  the  genera  Loterallus  and 
Porzana. 

The  status  of  this  species  in  Central  Amer- 
ica is  unclear.  There  is  only  one  definite  record 
from  Costa  Rica,  from  the  Sarapiqui  lowlands 
in  the  northeast  by  Stiles  and  Rosselli  on  22 
August  1987  (Stiles  and  Skutch  1989).  There 
is  an  additional  record  of  either  this  species  or 


the  similar  N.  colombianus  from  southern 
Costa  Rica,  near  Hitoy  Cerere  in  March  1985 
(Pratt  et  al.,  reported  in  Stiles  and  Skutch 
1989).  Two  specimens  collected  in  the  coastal 
lowlands  of  Bocas  del  Toro,  Panama  on  10 
November  1981,  were  later  identified  by  Rip- 
ley as  the  wide-ranging  N.  erythrops  olivas- 
cens-, several  individuals  were  seen  at  Tocu- 
men  Marsh  in  eastern  Panama  by  Behrstock 
(1983).  All  of  these  records  are  from  the  Ca- 
ribbean lowlands,  thus  the  record  reported 
herein  constitutes  the  first  for  the  Pacific  slope 
of  Central  America. 

There  are  two  records  from  North  Ameri- 
ca: from  east  central  Texas  on  17  February 
1972  (Arnold  1978),  and  Virginia  on  15  De- 
cember 1978  (Blem  1980).  Both  these  re- 
cords were  probably  wandering  individuals, 
a pattern  seen  in  many  other  species  of  rail 
(Remsen  and  Parker  1990).  In  contrast  to 
these  winter  records,  the  two  records  from 
Costa  Rica  are  from  August  and  March. 
Based  on  the  June  15  record  we  report,  and 
the  clear  evidence  of  breeding,  we  suggest 
that  N.  erythrops  has  a breeding  population 
in  southern  Costa  Rica. 

ACKNOWLEDGMENTS 

We  thank  D.  Levey,  J.  Eberhard,  and  two  anony- 
mous reviewers  for  helpful  and  constuctive  comments 
on  the  manuscript. 

LITERATURE  CITED 

Arnold,  K.  A.  1978.  First  United  States  record  of 
Paint-billed  Crake  (Neocrex  etythrops).  Auk  95: 
745-746. 

Behrstock,  R.  A.  1983.  Colombian  Crake  (Neocrex 
colombianus)  and  Paint-billed  Crake  (N.  eiy- 
throps):  hrst  breeding  records  for  Central  Ameri- 
ca. Am.  Birds  37:956-957. 

Blem,  C.  R.  1980.  A Paint-billed  Crake  in  Virginia. 
Wilson  Bull.  92:393-394. 

DEL  Hoyo,  J.,  A.  Elliot,  and  J.  Sargatal.  1996. 
Handbook  of  the  birds  of  the  world.  Vol.  3:  Hoa- 
tzin  to  auks.  Lynx  Edicions,  Barcelona,  Spain. 
Haverschmidt,  E and  G.  E Mees.  1994.  Birds  of  Su- 
riname. Vaco  Press,  Paramaribo,  Suriname. 
Osborne,  D.  R.  and  S.  R.  Beissinger.  1979.  The 
Paint-billed  Crake  in  Guyana.  Auk  96:425. 
Remsen,  J.  V.,  Jr.  and  M.  A.  Traylor,  Jr.  1983.  Ad- 
ditions to  the  avifauna  of  Bolivia,  part  2.  Condor 
85:95-98. 

Remsen,  J.  V.,  Jr.  and  T.  A.  Parker,  III.  1990.  Sea- 
sonal distribution  of  the  Azure  Gallinule  (Porphy- 


424 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  3,  September  1999 


nila  flavirostris),  with  comments  on  vagrancy  in 
rails  and  gallinules.  Wilson  Bull.  102:380-399. 
Ripley,  S.  D.  1977.  Rails  of  the  world.  David  R.  God- 
ine.  Pub.,  Boston,  Massachusetts. 

Stiles,  E G.  and  A.  E Skutch.  1989.  A guide  to  the 


birds  of  Costa  Rica.  Cornell  Univ.  Press,  Ithaca, 
New  York. 

Tostain,  O.,  J.  L.  Dujardin,  C.  Erard,  and  J.  M. 
Thiollay.  1992.  Oiseaux  de  Guyana.  Societe 
d’Etudes  Ornithologiques,  Brunoy,  Guyana. 


Wilson  Bull.,  11  1(3),  1999,  pp.  424-426 


Additional  Records  of  Fall  and  Winter  Nesting  by  Killdeer  in  Southern 

United  States 

Kimberly  G.  Smith,*’^  W.  Marvin  Davis,-  Thomas  E.  Kienzle,^  William  Post,^  and 

Robert  W.  Chinn^ 


ABSTRACT. — We  report  on  successful  nesting  at- 
tempts in  fall  by  Killdeer  (Cliaradrius  vociferii.s)  in 
southern  Mississippi  in  November,  1987  and  in  central 
Arkansas  in  October,  1998,  and  a winter  nesting  at- 
tempt in  South  Carolina  in  December,  1998.  The  first 
nest  was  found  1 year  before  previously  reported  fall 
nestings  in  the  Southeast  and  1 month  earlier  in  the 
season.  The  second  is  the  most  northern  and  western 
fall  nesting  site  in  the  South,  and  the  third  is  the  latest 
reported  nesting  attempt  in  the  southern  United  States. 
Taken  together  with  3 other  reported  successful  fall 
nests  in  Mississippi  and  South  Carolina,  Killdeer 
would  appear  to  be  the  only  fall  breeding  shorebird  in 
North  America  and,  based  on  those  6 widely-scattered 
observations  over  the  last  1 1 years,  should  now  be 
considered  a rare  fall  and  winter  breeder  across  the 
southern  United  States.  Received  24  Nov.  1998,  ac- 
cepted 31  March  1999. 


Although  an  anomalous  report  of  breeding 
in  November  exists  from  Michigan  in  1982 
(Tessen  1983),  Jackson  and  coworkers  (1995) 
were  the  first  to  document  fall  and  winter 
breeding  by  Killdeer  {Charcidrius  vociferus) 


' Dept,  of  Biological  Sciences,  Univ.  of  Arkansas, 
Fayetteville,  AR  72701;  E-mail: 
kgsmith@comp.uark.edu 

2 The  School  of  Pharmacy,  Univ.  of  Mississippi, 
University,  MS  38677. 

' Dept,  of  Microbiology  & Immunology,  Univ.  of 
Arkan.sas  for  Medical  Sciences,  Little  Rock,  AR 
72205. 

■’Charleston  Museum,  360  Meeting  St.,  Charleston, 
SC  29403. 

■’  7666  Chippendale  Rd.,  North  Charleston,  SC 
29420. 

’’  Corresponding  author. 


in  the  southeastern  United  States,  reporting  1 
set  of  chicks  and  adults  on  16  November  and 
another  set  on  1 1 December  1988  in  Okibbeha 
Co.,  Mississippi.  Subsequently,  Post  (1996) 
reported  3 downy  young,  apparently  1-2  days 
old,  taken  to  a veterinarian  in  Berkeley  Co., 
South  Carolina  on  13  November  1995.  Here 
we  report  on  two  more  successful  fall  nesting 
attempts  by  Killdeer  in  the  south:  one  from 
Mississippi  that  is  earlier  than  observations  by 
Jackson  and  coworkers  (1995),  and  one  from 
central  Arkansas,  the  most  northern  and  west- 
ern fall  nesting  site  yet  reported  in  the  South. 
We  also  document  a mid-December  winter 
nesting  attempt  in  South  Carolina,  which  is 
the  latest  nesting  activity  yet  reported. 

On  7 November  1987,  W.M.D.  found  and 
photographed  a pair  of  adults  with  one  chick, 
which  appeared  to  be  several  days  old,  at  the 
wastewater  treatment  plant  lagoon  in  Wave- 
land,  Hancock  Co.,  coastal  Mississippi.  The 
race-track  shaped  lagoon  was  surrounded  by 
a 4. 5-9. 2 m raised  strip  of  excavated  soil, 
which  varied  from  well-grassed  to  almost  bare 
areas,  one  of  which  was  evidently  chosen  for 
nesting  by  the  Killdeer. 

On  5 October  1998,  T.E.K.  and  his  wife  dis- 
covered a nest  with  four  eggs  located  in  a 
stone  area  on  the  barrier  of  the  parking  lot  at 
the  Veterans  Administration  Hospital  in  Little 
Rock,  Pulaski  Co.,  Arkansas.  During  daily  ob- 
servations, two  birds  were  usually  present  and 
the  female  was  observed  incubating  during 
the  day.  They  found  two  chicks  on  26  October 
and  a third  on  28  October.  The  nest  was  aban- 


SHORT  COMMUNICATIONS 


425 


doned  with  one  egg  remaining  on  29  October. 
Subsequent  analysis  determined  that  the  egg 
was  fertile,  but  did  not  hatch. 

On  5 December  1998,  R.W.C.  found  a Kill- 
deer  nest  containing  four  eggs  in  North 
Charleston,  Charleston  Co.,  South  Carolina. 
The  nest  was  a shallow  depression  located  on 
an  approximately  0.2  ha  lawn  covered  with 
short  (2-3  cm)  grass.  The  nest  was  about  10 
m from  a frequently  traveled  road  in  a U.S. 
Post  Office  complex.  He  made  repeated  visits 
to  the  site,  and  found  adults  incubating  daily 
during  5-15  December.  On  16  December,  af- 
ter arrival  of  a cold  front  on  the  coast,  no  Kill- 
deer  were  seen  in  the  area,  and  Chinn  con- 
cluded that  the  adults  had  abandoned  the  nest. 
On  17  December  at  15:00  EST,  W.P.  checked 
the  area,  and  finding  no  Killdeer,  collected  the 
four  eggs  (ChM  # 1998.11.50a-d),  which 
were  intact  but  cold.  The  heaviest  egg 
weighed  11.7  g;  the  lightest,  10.8  g.  Dimen- 
sions of  the  longest  egg  were  40.2  X 25.6 
mm;  the  shortest,  37.9  X 26.5  mm.  No  egg 
had  a discernible  embryo,  although  the  con- 
tents of  all  appeared  to  be  fresh. 

Killdeer  would  appear  to  be  the  only  fall 
nesting  shorebird  in  North  America.  There  are 
apparently  no  records  of  Killdeer  breeding  in 
Mexico  (P.  Escalante,  A.  T.  Peterson,  pers. 
comm.),  and  no  evidence  of  breeding  later 
than  July  in  southern  California  based  on 
clutches  in  the  collection  at  the  Western  Eoun- 
dation  of  Vertebrate  Zoology  (M.  Marin,  pers. 
comm.).  However,  based  on  breeding  records 
from  October,  Eebruary,  May,  and  August, 
Robertson  (1962)  concluded  that  Killdeer 
breed  throughout  the  year  in  the  Caribbean. 
Schardien  (1981)  determined  that  Killdeer 
have  year-round  territories  in  Mississippi  and 
copulations  have  been  observed  during  winter 
months  (Jackson  and  Jackson,  in  press).  Thus, 
those  six  widely  scattered  records  of  breeding 
over  the  last  1 1 years  would  suggest  that  the 
Killdeer  should  now  be  considered  a rare  fall 
and  winter  breeder  in  southern  United  States. 

Incidental  fall  breeding  in  temperate  re- 
gions has  been  documented  for  a wide  vari- 
ety of  birds  (e.g.,  Orians  1960),  but  repeated 
fall  (or  winter)  breeding  seems  to  be  trig- 
gered by  either  an  appropriate  stimulus  ap- 
pearing naturally,  e.g.,  green  cones  stimulat- 
ing breeding  in  Pinyon  Jays  (Gymnorhinus 
cyanocephalus;  Ligon  1978),  or  an  appropri- 


ate stimulus  occurring  during  the  wrong  sea- 
son, e.g..  Tricolored  Blackbirds  (Agelaius  tri- 
color) breeding  in  fall  in  response  to  flooding 
of  rice  fields  (Orians  1960).  Those  recent  ob- 
servations of  fall  breeding  by  Killdeer  could 
be  due  to  a combination  of  events,  including 
more  birdwatchers  being  active  during  fall 
and  early  winter  in  areas  of  the  South  than 
in  previous  times.  However,  the  most  likely 
explanation  is  the  unusually  warm  years  that 
have  occurred  over  the  last  decade  or  so 
(Mann  et  al.  1998).  Jackson  and  coworkers 
(1995)  noted  that  their  observations  followed 
a summer  drought  and  mild  fall  weather, 
which  is  similar  to  the  situation  in  Arkansas 
during  1998.  The  period  from  May  through 
September  of  1998  was  the  hottest  on  record 
at  Little  Rock,  with  18  days  above  37.8°  C, 
59  days  above  35°  C,  and  111  days  above 
32.2°  C and  below  average  rainfall  (National 
Weather  Service,  North  Little  Rock,  Arkan- 
sas). Warm  fall  weather  may  be  extending  the 
breeding  season  into  fall,  and  even  winter, 
months,  but  there  are  apparently  few,  if  any, 
breeding  attempts  reported  for  late  August 
and  September.  More  likely,  extended  warm 
fall  weather  stimulates  Killdeer  to  resume 
breeding,  as  they  typically  are  very  early 
spring  breeders  in  the  south  [e.g.,  nests  found 
in  Eebruary  of  1999  in  Louisiana  (W.  M. 
D.)]. 

ACKNOWLEDGMENTS 

J.  and  B.  J.  Jackson  supplied  unpublished  informa- 
tion on  Killdeer;  A.  Jobes,  H.  Parker.  R.  Payne,  and 
R.  E.  Ricklefs  supplied  important  references;  and  C.  E 
Bailey  examined  the  Arkansas  egg.  E.  Nol  and  P.  W. 
Bergstrom  made  helpful  suggestions  as  reviewers.  This 
is  Contribution  No.  2 from  the  Arkansas  Breeding  Bird 
Atlas. 

LITERATURE  CITED 

Jackson,  B.  J.  S.  and  J.  A.  Jackson.  In  press.  Killdeer 
(Cluiradrius  vociferus).  In  The  birds  of  North 
America  (A.  Poole  and  E Gill.  Eds.).  The  Acad- 
emy of  Natural  Sciences.  Philadelphia,  Pennsyl- 
vania; The  American  Ornithologists'  Union, 
Washington,  D.C. 

Jackson,  J.  A.,  M.  E Hodoes,  D.  J.  Ingold,  and  B.  J. 
S.  Jackson.  1995.  Fall  nesting  of  Killdeers  in  Mis- 
sissippi. Miss.  Kite  25:16—17. 

Eicon,  J.  D.  1978.  Reproductive  interdependence  of 
Pifion  Jays  and  pinon  pines.  Ecol.  Monogr.  48: 
I I 1-126. 

Mann,  M.  E.,  R.  S.  Bradley,  and  M.  K.  Hughes. 


426 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


1998.  Global-scale  temperature  patterns  and  cli- 
mate forcing  over  the  past  six  centuries.  Nature 
392:779-787. 

Orians,  G.  H.  1960.  Autumnal  breeding  in  the  Tricol- 
ored Blackbird.  Auk  77:379-398. 

Post,  W.  1996.  Late  autumnal  breeding  by  Killdeer. 
Ela.  Eield  Nat.  24:109. 


Robertson,  W.  B.,  Jr.  1962.  Observations  of  the  birds 
of  St.  John,  Virgin  Islands.  Auk  79:44-76. 
ScHARDiEN,  B.  J.  1981.  Behavioral  ecology  of  a south- 
ern population  of  Killdeer.  Ph.D.  diss.,  Mississippi 
State  Univ.,  Mississippi  State. 

Tessen,  D.  D.  1983.  Western  Great  Lakes  region.  Am. 
Birds  37:182-185. 


Wilson  Bull..  1 1 1(3),  1999,  pp.  426-427 


Wild  Turkeys  {Meleagris  gallopavo)  Renest  After  Successful  Hatch 

Craig  A.  Harper'  --^  and  Jay  H.  Exurn^"* 


ABSTRACT. — Wild  Turkey  [Meleagris  gallopavo) 
hens  frequently  renest  after  disturbance  on  the  nest, 
especially  while  laying  or  early  during  the  incubation 
period.  However,  no  record  exists  of  Wild  Turkey  hens 
renesting  after  a successful  hatch.  We  document  three 
Wild  Turkey  hens  that  renested  after  having  hatched  a 
brood.  None  of  the  renests  were  successful.  Received 
2 Dec.  1998,  accepted  6 March  1999. 


Nesting  success  of  Wild  Turkeys  {Melea- 
gris gallopavo)  varies  widely  across  their 
range  and  is  influenced  by  many  factors  (Van- 
glider  1992).  A successful  nest  generally  is 
defined  as  one  in  which  at  least  one  poult 
hatches.  Researchers  have  documented  many 
instances  of  Wild  Turkey  hens  renesting  after 
their  initial  nest  was  disturbed  or  depredated 
(Everette  et  al.  1980,  Williams  et  al.  1980, 
Vangilder  et  al.  1987),  most  often  while  laying 
or  during  early  incubation  (Williams  et  al. 
1976).  Causes  for  nest  failure  and  subsequent 
renesting  include  nest  destruction  by  predators 
(Speake  1980,  Vander-Haegen  et  al.  1988),  se- 
vere weather  (Roberts  and  Porter  1998,  Kim- 
mel  and  Zwank  1985),  and  disturbance  by  re- 
searchers (Still  and  Baumann  1990).  Renest- 
ing after  a successful  nest  was  not  thought  to 


' Dept,  of  Eorcst  Resources,  Clcmson  Univ.,  Clem- 
son.  SC  29634;  E-mail:  charpei@utk.edu 

- Pre.sent  address:  Dept,  of  Forestry,  Wildlife  and 
Fisheries.  P.O.  Box  1071,  Univ.  of  Tennessee,  Knox- 
ville. TN  37901. 

' Dept,  of  Zoology-Entomology,  Auburn  Univ.,  Au- 
burn, AL  36849. 

■'  Present  address:  Glatting,  Jackson,  Kcrcher,  An- 
glin. Lopez  and  Rinehart,  Orlando,  EL  32801. 

■'  Corresponding  author. 


occur.  In  his  Book  of  the  Wild  Turkey,  Wil- 
liams (1981:53)  stated,  “No  example  is 
known  of  a hen  nesting  again  in  the  same  year 
after  her  brood  hatched,  and  there  has  been  no 
reported  case  of  a turkey  hatching  two  broods 
in  one  year.”  Below,  we  document  three  cases 
of  hens  renesting  after  having  successfully 
nested  but  with  early  loss  of  broods. 

In  1983,  while  working  in  southern  Ala- 
bama, J.H.E.  found  a Wild  Turkey  hen  that 
renested  three  times  after  hatching  a brood 
that  did  not  survive  more  than  two  days.  This 
particular  hen  hatched  all  1 1 eggs  in  her  initial 
clutch  after  a normal  incubation  period  (28 
days).  On  the  day  after  hatching,  five  of  the 
poults  were  found  dead  in  the  nest  from  un- 
known causes.  The  hen  had  no  poults  with  her 
two  days  after  hatching,  and  the  fate  of  the 
remaining  six  poults  was  never  determined. 
Twenty-five  days  after  hatching,  the  hen  re- 
nested. This  nest  was  disturbed  by  investiga- 
tors, which  prompted  the  hen  to  abandon  that 
nest  and  eventually  renest  two  more  times. 
None  of  the  renests  was  successful. 

In  the  southern  Appalachians  of  North  Car- 
olina, C.A.H.  monitored  two  wild  turkey  hens 
that  renested  after  hatching  clutches  of  1 1 and 
14  eggs.  These  nests  were  initiated  in  early 
April  of  1996,  and  incubated  29  and  27  days. 
Both  broods  were  killed  within  five  days.  One 
hen  initiated  a second  nest  17  days  after  her 
initial  clutch  hatched.  This  renest  contained 
nine  eggs  that  were  incubated  65  days,  37 
days  beyond  the  normal  28-day  incubation  pe- 
riod. Subsequently,  this  nest  was  abandoned, 
and  the  eggs  were  determined  to  be  infertile. 


SHORT  COMMUNICATIONS 


427 


The  second  hen  began  incubating  another  nest 
27  days  after  her  initial  clutch  hatched.  This 
renest  was  incubated  for  1 1 days  and  aban- 
doned for  unknown  reasons. 

None  of  the  renests  we  documented  was 
successful,  either  because  of  infertility  or  nest 
abandonment.  Ultimately,  infertility  could  be 
a primary  factor  limiting  success  of  late  re- 
nests. Sperm  may  be  contained  in  a hen’s  ovi- 
duct up  to  56  days  after  copulation  (Blanken- 
ship 1992)  and  remain  viable  for  a ‘normal’ 
renesting  attempt.  However,  the  renests  we 
documented  were  initiated  considerably  later 
than  most  renests  because  they  occurred  after 
an  entire  incubation  period  plus  some  addi- 
tional days.  Thus,  a hen  might  need  to  copu- 
late again  in  order  to  lay  fertile  eggs  two 
months  after  the  primary  mating  season  (i.e., 
April-May).  However,  copulation  in  July 
would  be  exceptional  because  the  urge  to 
breed,  which  is  associated  with  the  rise  in  tes- 
tosterone (for  males)  and  prolactin  (for  fe- 
males) levels,  is  regressing  (Blankenship 
1992). 

The  ability  of  Wild  Turkey  hens  to  renest 
after  having  hatched  a brood  did  not  contrib- 
ute to  productivity  in  these  cases.  However, 
we  now  know  it  is  possible  for  hens  to  renest 
after  a successful  nesting  attempt  and  it  is 
conceivable  that  such  nesting  could  be  suc- 
cessful. 

LITERATURE  CITED 

Blankenship,  L.  H.  1992.  Physiology.  Pp.  84-100  in 
The  Wild  Turkey:  biology  and  management  (J.  G. 


Dickson,  Ed.).  Stackpole  Books,  Harrisburg, 
Pennsylvania. 

Everetth,  D.  D.,  D.  W.  Speake,  and  W.  K.  Maddox. 
1980.  Natality  and  mortality  of  a north  Alabama 
Wild  Turkey  population.  Proc.  Natl.  Wild  Turkey 
Symp.  4:1  17-126. 

Kimmel,  E G.  and  P.  J.  Zwank.  1985.  Habitat  selection 
and  nesting  responses  to  spring  flooding  by  east- 
ern Wild  Turkey  hens  in  Louisiana.  Proc.  Natl. 
Wild  Turkey  Symp.  5:155-171. 

Roberts,  S.  D.  and  W.  E Porter.  1998.  Relation  be- 
tween weather  and  survival  of  Wild  Turkey  nests. 
J.  Wildl.  Manage.  62:1492-1498. 

Speake,  D.  W.  1980.  Predation  on  Wild  Turkeys  in 
Alabama.  Proc.  Natl.  Wild  Turkey  Symp.  4:86- 
101. 

Still,  H.  R.,  Jr.  and  D.  P.  Baumann,  Jr.  1990.  Wild 
Turkey  nesting  ecology  on  the  Erancis  Marion  Na- 
tional Forest.  Proc.  Natl.  Wild  Turkey  Symp.  6: 
13-17. 

Vander-Haegen,  W.  M.,  W.  E.  Dodge,  and  M.  W. 
Sayre.  1988.  Factors  affecting  productivity  in  a 
northern  Wild  Turkey  population.  J.  Wildl.  Man- 
age. 52:127-133. 

Vangilder,  L.  D.  1992.  Population  dynamics.  Pp. 
144-164  in  The  Wild  Turkey:  biology  and  man- 
agement. (J.  G.  Dickson,  Ed.).  Stackpole  Books, 
Harrisburg,  Pennsylvania. 

Vangilder,  L.  D.,  E.  W.  Kurzejeski,  V.  L.  Kimmel- 
Truitt,  and  j.  B.  Lewis.  1987.  Reproductive  pa- 
rameters of  Wild  Turkey  hens  in  north  Missouri. 
J.  Wildl.  Manage.  51:535-540. 

Williams,  L.  E.,  Jr.  1981.  The  book  of  the  Wild  Tur- 
key. Winchester  Press,  Tulsa,  Oklahoma. 

Williams,  L.  E.,  Jr.,  D.  H.  Austin,  and  T.  E.  Peoples. 
1976.  The  breeding  potential  of  the  Wild  Turkey 
hen.  Proc.  Annu.  Conf.  Southeast.  Assoc.  Fish 
Wildl.  Agenc.  30:371-376. 

Williams,  L.  E.,  Jr.,  D.  H.  Austin,  and  T.  E.  Peoples. 
1980.  Turkey  nesting  success  on  a Florida  study 
area.  Proc.  Natl.  Wild  Turkey  Symp.  4:102-107. 


428 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


Wilson  Bull.,  111(3),  1999,  pp.  428-432 


Post-migration  Weight  Gain  of  Swainson’s  Hawks  in  Argentina 

Michael  I.  Goldstein,'  Peter  H.  Bloom,-  Jose  H.  Sarasola,^  and  Thomas  E.  Lacher' 


ABSTRACT. — Swainson’s  Hawks  (Biiteo  swain- 
soni\  Aguilucho  Langostero)  were  captured  and  band- 
ed in  La  Pampa,  Argentina  from  28  November  1996 
through  25  January  1997.  We  collected  morphometric 
measurements  to  determine  if  hawks  gained  weight  on 
the  austral  (non-breeding)  grounds.  Hawks  in  appar- 
ently healthy  condition  weighed  819.7  ± 1 1.1  g (mean 
± SE,  n = 127,  range  540-1090  g).  Weight  increased 
significantly  over  the  length  of  the  non-breeding  sea- 
son in  Argentina  (P  = 0.0059),  but  wingspan  (124.9 
± 0.5  cm,  n = 127,  range  105.7-137.1  cm)  and  tail 
(19.8  ± 0.1  cm,  n = 127,  range  16.5-23.1  cm)  did 
not.  When  separated  by  age,  weight  increased  signifi- 
cantly for  juveniles  (P  = 0.0083)  but  was  marginally 
non-significant  for  adults  (P  = 0.0555).  Received  15 
Oct.  1998,  accepted  15  Feb.  1999. 


Swainson’s  Hawks  (Buteo  swainsoni)  are 
long-distance  migrants  that  travel  between  the 
plains,  shrublands,  and  pampas  (grasslands)  of 
North  and  South  America.  During  the  south- 
ward migration,  some  hawks  temporarily  stop 
in  Central  America  or  in  the  agricultural  zones 
of  Mexico  (Ridgely  and  Gwynne  1989,  Stiles 
and  Skutch  1989,  Howell  and  Webb  1995). 
Occasionally  some  hawks  are  found  in  Florida 
(Robertson  and  Woolfenden  1992)  and  others 
winter  in  California  (Herzog  1996).  The  pam- 
pas of  Argentina  are  a major  non-breeding 
(austral)  destination,  supporting  much  of  this 
species’  population  from  mid-November 
through  early  March.  Two  satellite  radio- 
tagged  hawks  trapped  by  Brian  Woodbridge 
in  California  in  1994  were  tracked  to  La  Pam- 
pa province,  Argentina  (Woodbridge  et  al. 

1995).  Large  numbers  of  hawks  were  subse- 
quently found  in  La  Pampa,  Buenos  Aires, 


' Texas  A&M  Univ..  Dept,  of  Wildlife  and  Fisheries 
Scienees,  210  Nagle  Hall,  College  Station,  TX  77843- 
2258. 

2 Western  Foundation  of  Vertebrate  Zoology,  439 
Calle  San  Pablo,  Camarillo,  CA  93010. 

^ Univ.  Naeional  de  La  Pampa,  Facultad  de  Cieneias 
Exactos  y Naturales,  Uruguay  151,  (6300)  Santa  Rosa, 
La  Pampa,  Argentina. 

■’  Corresponding  author; 

E-mail;  mgold.stein@tamu.edu 


and  Cordoba  provinces  (Goldstein  et  al. 

1996) . 

Adults  feed  primarily  vertebrates  to  nes- 
tlings (England  et  al.  1997).  Summer  flocks  of 
non-breeding  birds  eat  a more  varied  diet,  in- 
cluding insects  (Johnson  et  al.  1987).  On  the 
non-breeding  grounds,  hawks  are  generally 
seen  in  large  aggregations:  roosting,  foraging, 
and  traveling  together.  In  the  pampas,  the 
hawks  primarily  eat  invertebrates  (White  et  al. 
1989,  Jaramillo  1993,  Woodbridge  et  al.  1995, 
Goldstein  et  al.  1996).  The  dietary  shift  allows 
for  more  birds  to  be  supported  per  unit  area. 
In  La  Pampa,  flocks  of  birds  following  grass- 
hopper (Orthoptera)  outbreaks  have  been  re- 
ported (Rudolph  and  Fisher  1993).  We  found 
flocks  as  large  as  12,000  birds  (Goldstein 

1997) . 

Long-distance  migration  using  only  stored 
fat  has  been  suspected  but  not  documented 
(Smith  et  al.  1986).  Whether  hawks  forage  or 
fast  en  route  to  Argentina,  or  whether  specific 
stopover  habitats  are  regularly  used  is  not 
known  (Goldstein  and  Smith  1991,  Kirkley 
1991).  The  extent  of  predation  on  airbom 
dragonflies  and  other  flying  insects  during  mi- 
gration is  also  unknown.  Nevertheless,  if  mi- 
grating hawks  used  only  stored  fat  they  might 
arrive  in  the  pampas  in  poor  condition  (Smith 
et  al.  1986).  Hawks  have  been  reported  arriv- 
ing in  Argentina  in  such  weak  condition  that 
they  were  picked  up  by  hand  (C.  C.  Olrog  in 
Smith  1980).  None  of  these  birds,  however, 
were  checked  for  contaminants. 

Swainson’s  Hawks  are  found  in  Argentina 
from  early  November  through  mid-February, 
although  later  arrivals  and  earlier  departures 
have  been  documented  (England  et  al.  1997). 
Substantial  periods  of  fasting  during  migration 
would  result  in  a substantial  loss  of  weight, 
weakened  condition,  and  a subsequent  in- 
crease in  weight  when  on  the  wintering 
grounds.  From  November  1996  through  Jan- 
uary 1997,  we  captured  and  measured  Swain- 
son’s Hawks  on  the  non-breeding  grounds.  We 


SHORT  COMMUNICATIONS 


429 


present  these  data  and  observations  regarding 
Swainson’s  Hawk  austral  weight  gain. 

METHODS 

The  research  area  in  northern  La  Pampa  (35°  14'  S, 
63°  57'  W,  149  m ASL)  is  a flat  grassland  dominated 
by  row  crop  agriculture.  One  hundred  and  twenty-eight 
Swainson's  Hawks  were  captured  and  sampled  in  La 
Pampa  from  28  November  1996  through  25  January 
1997.  Most  hawks  were  trapped  by  bal  chatri  traps  in 
open  fields  or  on  dirt  roads  between  foraging  fields 
(Bloom  1987).  Nine  of  the  128  hawks  were  captured 
by  hand  after  they  were  grounded  from  their  nighttime 
Eucalyptus  viminalis  roost  by  a thunderstorm.  All 
hawks  were  captured  in  the  morning  between  05:15 
and  I 1 :00  local  time. 

Hawks  were  banded  and  weighed  to  the  nearest 
gram  with  a 1500  g Pesola  scale.  Wingspan  and  tail 
length  were  recorded  to  the  nearest  mm.  Hawks  were 
classified  as  juveniles  or  adults  based  on  plumage,  with 
immatures  grouped  as  juveniles  (Wheeler  and  Clark 
1995).  A regression  analysis  was  used  to  analyze 
changes  in  weight,  wingspan,  and  tail  length  for  the 
population  and  for  each  age  category.  Time  was  ex- 
pressed in  number  of  trap  days  from  Day  1 (28  No- 
vember 1996).  Because  sexes  could  not  be  distin- 
guished morphologically,  we  did  not  analyze  the  data 
by  sex  (Wheeler  and  Clark  1995).  Statistical  analyses 
were  performed  using  SPSS  8.0  for  Windows  (SPSS 
Inc.,  Chicago,  Illinois). 

RESULTS 

We  collected  complete  data  sets  for  127 
healthy  hawks.  One  individual  that  might  have 
been  exposed  to  pesticides  was  excluded 
(Goldstein  1997).  Weight  averaged  819.7  g [± 
11.1  (SE),  range  540-1090  gj.  Wingspan 
measured  124.9  cm  (±  0.5,  range  105.7-137.1 
cm),  and  tail  length  measured  19.8  cm  (±  0.1, 
range  16.5-23.1  cm).  Regression  analysis  in- 
dicated that  weight  increased  significantly 
over  time  (R^  = 0.0705,  P = 0.0059),  but  nei- 
ther wingspan  (R^  = 0.0001,  P > 0.05)  nor 
tail  length  (R^  = 0.0078,  P > 0.05)  increased 
during  the  season.  This  pattern  suggested  an 
increase  in  mass  from  late  November  until  late 
January  without  structural  growth. 

There  were  65  adult,  58  juvenile,  and  4 un- 
known aged  birds.  Adults  weighed  836.6  g (± 
16.5,  range  560-1090  g).  Juveniles  weighed 
794.2  g (±  15.9,  range  540-1080  g).  Weight 
significantly  increased  over  time  for  Juveniles 
{P  = 0.0083,  df  = 57,  R2  = 0.1 179;  Fig.  1 A), 
but  was  marginally  non-significant  for  adults 
(P  = 0.0555,  df  = 64,  R^  = 0.0570;  Fig.  IB). 
Although  both  age  categories  showed  weight 


gains,  the  larger  slope  for  juveniles  indicated 
they  gained  weight  at  a faster  rate  than  adults. 
When  examined  by  age,  neither  wingspan  nor 
tail  length  changed  over  time  (P  > 0.05  for 
all). 

Wingspan  was  significantly  and  positively 
correlated  with  tail  length  (r  = 0.509;  P < 
0.05)  and  with  weight  (r  = 0.342;  P < 0.05). 
Weight  and  tail  length  showed  no  correlation 
(r  = -0.048;  P > 0.05). 

DISCUSSION 

Swainson’s  Hawks  remain  in  Argentina 
from  mid-November  through  mid-February 
(Houston  and  Schmutz  1995,  England  et  al. 
1997).  When  we  began  following  large  flocks 
on  20  November  1996  we  saw  no  signs  of 
birds  in  poor  condition.  Whether  birds  arrived 
in  Argentina  in  poor  condition  is  uncertain, 
but  neither  our  observations  nor  conversations 
with  farmers  supported  that  conjecture.  Al- 
though hawks  were  not  in  poor  condition 
when  captured  in  late  November,  they  were 
somewhat  lighter  than  those  measured  on  the 
breeding  grounds.  Breeding  males  weighed 
808  g (range  693-936  g,  n = 69)  and  females 
weighed  1109  g (range  937-1367  g,  n = 50), 
for  an  overall  mean  of  934  g (J.  K.  Schmutz 
in  England  et  al.  1997).  Although  hawks  cap- 
tured in  Argentina  averaged  819.7  g,  birds 
captured  on  28  November- 17  December  1996 
were  notably  lighter  (779.1  g)  than  the  mean. 
Hawks  captured  from  Days  21-59  weighed 

860.5  g.  Both  adults  and  juveniles,  when  ex- 
amined separately,  show  similar  patterns  of 
weight  for  Days  1-20  (adults  814.8  g,  juve- 
niles 737.0  g)  and  Days  21-59  (adults  872.5 
g,  juveniles  847.7  g).  This  pattern  suggests 
that  both  adult  and  juvenile  hawks  lost  weight 
during  migration  and  then  regained  it  during 
their  stay  in  Argentina,  although  the  trend  for 
juveniles  was  stronger.  However,  if  the  hawks 
foraged  in  the  northern  pampas  for  several 
days  or  weeks  prior  to  their  arrival  near  the 
La  Pampa  field  sites,  our  measure  of  weight 
loss  is  underestimated. 

Rectrix  lengths  were  consistent  with  those 
reported  from  North  America  (England  et  al. 
1997).  Alberta  males  averaged  18.4  cm  (range 
17.0-19.8  cm,  /?  = 61)  and  females  averaged 

20.5  cm  (range  19.3-22.1  cm,  n = 43)  for  an 
overall  length  of  19.3  cm  (J.  K.  Schmutz  in 
England  et  al.  1997).  We  found  no  wingspan 


430 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


A. 


1200 

1000 


800 

s 

% 600 
O 
§ 

400 

200 

0 

0 10  20  30  40  50  60  70 

Time  (trapday) 

FIG.  1.  Swainson’s  Hawk  juvenile  (A)  and  adult  (B)  austral  weights  from  28  November  1996  through  25 
January  1997  as  measured  in  northern  La  Pampa,  Argentina.  Juvenile  weight  gain  was  significant  (Y  = 2.3972x 
+ 736.65,  = 0.1  179,  n = 58,  P = 0.0083).  Adult  weight  gain  was  marginally  non-significant  (Y  = L8196x 
+ 806.89,  R2  = 0.0570,  n = 65,  P = 0.0555). 


data  for  North  American  hawks  in  the  litera- 
ture. In  addition,  hawks  were  molting 
throughout  the  non-breeding  season  in  Argen- 
tina, a phenomenon  that  has  not  been  studied. 

We  captured  only  one  hawk  that  was  light 
weight  and  appeared  weak.  The  hawk 
weighed  560  g when  trapped  on  25  November 
1996.  It  gained  100  g in  5 days  of  captivity 
and  was  subsequently  released.  Whether  this 
bird  was  weak  from  migration  or  had  been 


exposed  to  toxins  is  unknown.  Although  no 
chemical  residues  were  found  on  either  foot- 
wash  or  feather  residue  samples,  cholinester- 
ase measurements  from  blood  plasma  samples 
taken  at  the  time  of  capture  were  30%  below 
those  taken  at  the  time  of  release  (Goldstein 
1997).  Exposure  to  organophosphate  insecti- 
cides decreases  plasma  cholinesterase  activity 
levels  (Hill  and  Fleming  1982). 

Large  flocks  foraging  on  the  ground  during 


SHORT  COMMUNICATIONS 


431 


migration  were  thought  unlikely  in  the  past 
(Smith  et  al.  1986),  but  such  behavior  may  not 
be  uncommon  (see  England  et  al.  1997,  Gold- 
stein 1997).  Insect  outbreaks  in  agricultural 
grasslands  may  provide  terrestrial  foraging 
opportunities  along  the  migration  corridor.  Re- 
cently discovered  populations  wintering  in 
southwestern  Mexico  confirm  the  need  for 
further  study  of  stopover  habitats.  In  addition, 
hawks  are  known  to  forage  aerially  for  grass- 
hoppers and  dragonflies  across  the  pampas  of 
Argentina  (White  et  al.  1989,  Jaramillo  1993, 
Rudolph  and  Fisher  1993,  Woodbridge  et  al. 
1995,  Goldstein  1997).  We  have  no  idea  what 
altitudes  are  reached  while  traveling  on  ther- 
mal air  currents  across  the  insect-rich  tropical 
and  sub-tropical  rainforests,  nor  whether  any 
appreciable  distance  is  ever  achieved  without 
aid  of  thermals.  We  do  not  know  whether  ae- 
rial insects  are  in  adequate  supply  at  the 
heights  of  travel.  The  question  of  the  duration 
and  extent  of  fasting  of  Swainson’s  Hawks 
during  migration  needs  further  study  although 
lengthy  periods  of  fasting  seem  unlikely  given 
the  weights  of  November  birds  captured  in 
Argentina. 

ACKNOWLEDGMENTS 

M.  I.  Goldstein  wishes  to  express  thanks  to  the 
Graduate  Program  Enhancement  Fund  of  the  Depart- 
ment of  Wildlife  and  Fisheries  Sciences  of  Texas 
A&M  University  for  financial  support,  enabling  pre- 
sentation of  this  work  at  the  annual  meeting  of  the 
Raptor  Research  Foundation  in  Ogden,  Utah  in  Sep- 
tember 1998.  We  wish  also  to  thank:  A.  Lanusse  and 
S.  Salva  who  hosted  our  stay  at  Estancia  Chanilao  in 
La  Pampa;  B.  Woodbridge  and  M.  Bechard  who  pro- 
vided financial  assistance  and  technical  guidance;  M. 
Parker,  M.  Bechard,  M.  Kochert,  S.  Canavelli,  S.  Bak- 
er, S.  Weidensaul,  and  the  Lanusse  family  for  trapping 
hawks;  and  M.  E.  Zaccagnini  and  J.  L.  Panigatti  for 
logistical  support.  This  project  was  funded  in  part  by 
the  Hawk  Mountain-Zeiss  Raptor  Research  Award, 
the  Frank  M.  Chapman  Memorial  Fund,  U.S.  Fish  and 
Wildlife  Service  (Office  of  International  Affairs),  No- 
vartis Crop  Protection,  and  the  Archbold  Tropical  Re- 
search Center.  P.  H.  Bloom  thanks  B.  Woodbridge  and 
the  U.S.  Forest  Service  for  facilitating  travel  to  Ar- 
gentina. A.  Palka,  C.  S.  Houston,  A.  Jaramillo,  and  an 
anonymous  reviewer  provided  critical  reviews  of  ear- 
lier versions  of  the  manuscript. 

LITERATURE  CITED 

Bloom,  P.  H.  1987.  Capturing  and  handling  raptors. 
Pp.  99—123  in  Raptor  management  techniques 
manual  (B.  A.  G.  Pendleton,  B.  A.  Milsap,  K.  W. 


Cline,  and  D.  M.  Bird,  Eds.).  National  Wildlife 
Federation.  Wa.shington,  D.C. 

England,  A.  S.,  M.  J.  Bechard,  and  C.  S.  Hou,ston. 
1997.  Swain.son's  Hawk  (liiileo  swainsoni).  In  The 
birds  of  North  America,  no.  265  (A.  Poole  and  F. 
Gill  Eds.).  The  Academy  of  Natural  Sciences, 
Philadelphia,  Pennsylvania;  The  American  Orni- 
thologists’ Union,  Washington,  D.C. 

Goldstein,  D.  L.  and  N.  G.  Smith.  1991.  Response  to 
Kirkley.  J.  Raptor  Res.  25:87-88. 

Goldstein,  M.  I.  1997.  Toxicological  assessment  of  a 
neotropical  migrant  on  its  non-breeding  grounds: 
case  study  of  the  Swainson’s  Hawk  in  Argentina. 
M.Sc.  thesis,  Clemson  Univ.,  Clemson,  South 
Carolina. 

Goldstein,  M.  I.,  B.  Woodbridge,  M.  E.  Zaccagnini, 
S.  B.  Canavelli,  and  A.  Lanusse.  1996.  An  as- 
sessment of  mortality  of  Swainson’s  Hawks  on 
wintering  grounds  in  Argentina.  J.  Raptor  Res.  30: 
106-107. 

Herzog,  S.  K.  1996.  Wintering  Swainson’s  Hawks  in 
California’s  Sacramento-San  Joaquin  River  Delta. 
Condor  98:876-879. 

Hill,  E.  F.  and  W.  J.  Fleming.  1982.  Anticholinester- 
ase poisoning  of  birds:  field  monitoring  and  di- 
agnosis of  acute  poisoning.  Fnv.  Toxicol.  Chem. 
1:27-38. 

Houston,  C.  S.  and  J.  K.  Schmutz.  1995.  Swainson’s 
Hawk  banding  in  North  America  to  1992.  N.  Am. 
Bird  Bander  20:120-127. 

Howell.  S.  N.  G.  and  S.  Webb.  1995.  A guide  to  the 
birds  of  Mexico  and  northern  Central  America. 
Oxford  Univ.  Press,  New  York. 

Jaramillo,  A.  P.  1993.  Wintering  Swainson’s  Hawks 
in  Argentina:  food  and  age  segregation.  Condor 
95:475-479. 

Johnson,  C.  G.,  L.  A.  Nickerson,  and  M.  J.  Bechard. 
1987.  Grasshopper  consumption  and  summer 
flocks  of  non-breeding  Swainson’s  Hawks.  Con- 
dor 89:676-678. 

Kirkley,  J.  S.  1991.  Do  migrant  Swainson’s  Hawks 
fast  en  roitte  to  Argentina?  J.  Raptor  Res.  25:82- 
86. 

Ridgely,  R.  S.  and  j.  a.  Gwynne.  Jr.  1989.  A guide 
to  the  birds  of  Panama  with  Costa  Rica,  Nicara- 
gua, and  Honduras,  second  ed.  Princeton  Univ. 
Press,  Princeton,  New  Jersey. 

Robertson.  W.  D.,  Jr.  and  G.  E.  Woolfenden.  1992. 
Florida  bird  species:  an  annotated  list.  Florida  Or- 
nithological Society,  Cocoa.  Florida. 

Rlidolph,  D.  C.  and  C.  D.  Fisher.  1993.  Swainson’s 
Hawk  predation  on  dragonflies  in  Argentina.  Wil- 
son Bull.  105:365-366. 

Smith.  N.  G.  1980.  Hawk  and  vulture  migrations  in 
the  Neotropics.  Pp.  51-65  in  Migrant  birds  in  the 
Neotropics:  ecology,  behavior,  distribution,  and 
conservation  (A.  Keast  and  F.  S.  Morton.  Eds.). 
Smithsonian  Institution  Press,  Washington,  D.C. 

Smith,  N.  G.,  D.  L.  Goldstein,  and  G.  A.  Barthol- 
omew. 1986.  Is  long-distance  migration  possible 
using  only  stored  fat?  Auk  103:607-61  1. 


432 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  3,  September  1999 


Stiles,  E G.  and  A.  E Skutch.  1989.  A guide  to  the 
birds  of  Costa  Rica.  Cornell  Univ.  Press,  Ithaca, 
New  York. 

Wheeler,  B.  K.  and  W.  S.  Clark.  1995.  A photo- 
graphic guide  to  North  American  raptors.  Aca- 
demic Press,  San  Diego,  California. 

White,  C.  M.,  D.  A.  Boyce,  and  R.  Straneck.  1989. 
Observations  on  Buteo  .swainsoni  in  Argentina, 


1984,  with  comments  on  food,  habitat  alteration 
and  agricultural  chemicals.  Pp.  79—87  in  Raptors 
in  the  modern  world  (B.  U.  Meyburg,  and  R.  D. 
Chancellor,  Eds.).  World  Wildlife  Group  on  Birds 
of  Prey,  London,  U.K. 

WOODBRIDGE,  B.,  K.  K.  PiNLEY,  AND  S.  T.  SEAGER. 
1995.  An  investigation  of  the  Swainson’s  Hawk 
in  Argentina.  J.  Raptor  Res.  29:202-204. 


Wihon  Bull.,  111(3),  1999,  pp.  432-436 

Siblicide  at  Northern  Goshawk  Nests:  Does  Food  Play  a Role? 

Wendy  A,  Estes,'  - Sarah  R,  Dewey, and  Patricia  L.  Kennedy' 


ABSTRACT. — Siblicide  as  a mechanism  for  brood 
reduction  has  been  reported  in  a number  of  asynchro- 
nously hatching  bird  species.  Although  researchers 
have  documented  the  occurrence  of  facultative  sibli- 
cide in  several  raptor  species,  its  cause  is  still  debated. 
Most  hypotheses  relate  incidences  of  siblicide  to  food 
availability.  The  food-amount  hypothesis  predicts  a 
negative  relationship  between  the  amount  of  food 
available  and  nestling  aggression.  While  the  food- 
amount  hypothesis  has  received  much  attention,  few 
studies  show  more  than  correlational  support  for  this 
activity  in  raptors.  Our  observation  of  a siblicide  event 
at  a Northern  Goshawk  (Accipiter  gentili.s)  nest  used 
as  a control  in  a supplemental  feeding  experiment,  and 
a similar  incident  where  a nestling  goshawk's  death 
can  be  attributed  to  siblicide  provide  support  for  the 
negative  correlation  between  food  amount  and  sibling 
aggression.  These  observations  and  the  lack  of  any  re- 
ported sibling  aggression  at  seven  supplementally  fed 
nests  showing  extreme  hatching  asynchrony  also  in- 
dicate a relationship  between  food  resources  and  brood 
reduction.  Our  observations  are  consistent  with  the 
idea  that  goshawks  exhibit  facultative  siblicide,  and 
that  resource  levels  as  predicted  by  the  food-amount 
hypothesis  directly  influence  it.  Received  7 Oct.  1998, 
accepted  16  Feb.  1999. 


Hatching  asynchrony  in  birds  facilitates 
brood  reduction  because  the  last  hatched  nest- 


' Dept,  of  Eishery  and  Wildlife  Biology,  Colorado 
State  Univ.,  Eort  Collins,  CO  80523. 

2 Pre.scnt  address:  104  Biological  Sciences  East, 
Univ.  of  Arizona,  Tuc.son,  AZ  85721. 

’ Present  address:  USDA  Forest  .Service,  Ashley  Na- 
tional Forest,  Vernal  Ranger  District,  355  N.  Vernal 
Ave.,  Vernal,  UT  84078. 

•'  Corresponding  author; 

E-mail:  patk@cnr.colostate.edu 


ling  is  at  a competitive  disadvantage  if  re- 
sources provided  by  the  parents  prove  inade- 
quate. In  asynchronously  hatched  broods,  the 
youngest  nestling  occasionally  dies  from  ag- 
gressive sibling  behavior  including  pecking, 
exclusion  during  feeding  bouts,  or  eviction 
from  the  nest  (Lack  1954,  Mock  et  al.  1990, 
Creighton  and  Schnell  1996).  Asynchronous 
hatching  results  in  adapting  a brood  size  to  an 
unpredictable  food  supply  by  allowing  all 
young  to  survive  when  food  is  plentiful,  but 
ensuring  brood  reduction  to  match  parental 
provisioning  capabilities  when  prey  levels  are 
meager  (Lack  1954,  Newton  1979,  Bryant  and 
Tatner  1990,  Heeb  1994).  Species  in  which 
the  frequency  of  siblicide  events  are  variable 
are  termed  facultative,  while  those  in  which 
siblicide  occurs  in  nearly  all  nest  attempts  are 
called  obligate  (Edwards  and  Collopy  1983). 
Although  the  occurrence  of  obligate  siblicide 
appears  to  be  largely  innate  (Mock  et  al.  1990, 
Gerhardt  et  al.  1997),  the  causes  of  facultative 
siblicide  are  still  debated  (Forbes  and  Mock 
1994). 

Fatal  sibling  aggression  has  been  docu- 
mented in  a range  of  avian  species  (Stinson 
1979,  Braun  and  Hunt  1983,  Anderson  1989, 
Drummond  and  Garcia  Chavelas  1989,  Bryant 
and  Tatner  1990,  Mock  et  al.  1990;  Mock  and 
Lamey  1991,  Heinsohn  1995,  Reynolds 
1996).  However,  an  understanding  of  the 
proximate  factors  that  influence  the  occur- 
rence of  facultative  siblicide  remains  elusive 
because  such  events  are  rare  and  unpredict- 
able. Most  similar  hypotheses  attempt  to  ex- 
plain facultative  siblicide  in  relation  to  food. 


SHORT  COMMUNICATIONS 


433 


The  food-amount  hypothesis  predicts  a nega- 
tive relationship  between  the  amount  of  food 
available  and  nestling  aggression  (Mock  et  al. 
1987,  Creighton  and  Schnell  1996).  By  killing 
its  sibling  when  food  is  scarce,  a nestling  may 
increase  its  chance  of  survival  by  increasing 
its  share  of  food  delivered  to  the  nest.  Mock 
and  coworkers  (1990)  found  that  smaller  food 
morsels  can  be  monopolized  through  combat 
and,  therefore,  reward  sibling  aggression. 
Higher  rates  of  aggression  were  observed  in 
larger  broods  of  Cattle  Egrets  (Bubulcus  ibis) 
where  individual  food  portions  are  expected 
to  be  smaller  (Mock  et  al.  1987). 

Because  siblicide  events  are  uncommon, 
and  tend  to  go  unwitnessed  unless  nests  are 
under  constant  watch,  few  studies  have  estab- 
lished a causal  link  between  food  resources 
and  fatal  sibling  aggression  (but  see  Mock  et 
al.  1987).  Facultative  siblicide  has  been  doc- 
umented in  several  raptor  species,  but  its 
cause  has  not  been  fully  investigated  (Schnell 
1958,  Pilz  and  Seibert  1978,  Newton  1979, 
Bechard  1983,  Zachel  1985,  Bortolotti  et  al. 
1991,  Boal  and  Bacorn  1994).  Although  the 
hypothesis  that  food  supplies  influence  sibling 
aggression  is  intuitively  appealing,  few  stud- 
ies, with  the  exception  of  Wiebe  and  Borto- 
lotti (1995)  and  Wellicome  (1997),  have  pro- 
vided more  than  correlational  support  for  this 
activity  in  raptors. 

Northern  Goshawks  (Accipiter  gentilis) 
hatch  asynchronously  and  exhibit  siblicide 
(Newton  1979,  Stinson  1979).  Observational 
accounts  of  siblicide  in  goshawks  are  rare 
(Schnell  1958,  Zachel  1985,  Boal  and  Bacorn 
1994),  and  its  occurrence  is  thought  to  be  lim- 
ited to  times  when  food  is  in  very  low  supply 
(Newton  1979).  Experimental  data  linking 
food  resources  and  incidence  of  siblicide  in 
this  species  are  nonexistent.  In  this  paper,  we 
report  the  occurrence  of  a siblicide  event  at  a 
goshawk  nest  in  northeastern  Utah.  We  also 
describe  another  incident  in  which  a nestling’s 
death  was  likely  the  result  of  siblicide  and  we 
provide  experimental  evidence  that  the  inci- 
dence of  sibling  aggression  may  be  related  to 
food  supplies.  Our  observations  are  consistent 
with  the  hypothesis  that  goshawks  exhibit  fac- 
ultative siblicide  directly  influenced  by  food 
resources. 

Our  siblicide  observations  occurred  at  nests 
that  were  part  of  a study  on  the  influence  of 


food  provisioning  on  female  nest  attendance 
and  nestling  begging  vocalizations.  This  in- 
vestigation was  part  of  a larger  experiment  ex- 
amining the  influence  of  supplemental  food  on 
parental  care  strategies  and  juvenile  survival 
(Dewey  1999).  In  1997,  14  nests  were  includ- 
ed in  the  food  supplementation  experiment 
(experimental  design  similar  to  that  of  Ward 
and  Kennedy  1996).  Seven  of  these  nests  were 
randomly  assigned  as  treatments  and  were 
provided  Japanese  Quail  (Coturnix  coturnix) 
from  hatching  through  the  fledgling  depen- 
dency period.  We  visited  treatment  nests  every 
two  to  three  days  and  provided  sufficient  food 
to  meet  the  energy  requirements  of  the  female 
and  young  until  the  next  scheduled  visit  (see 
Dewey  1999  for  details).  Control  nests  were 
visited  at  the  same  interval  and  for  the  same 
amount  of  time,  but  were  not  given  food.  The 
nest  attendance/vocalization  study  was  con- 
ducted from  mid-June  to  mid-July  1997  and 
consisted  of  a subset  of  the  nests  used  in  the 
food  supplementation  experiment.  Each  nest 
was  observed  for  a 3 h period  on  three  dif- 
ferent occasions  from  a portable  blind  located 
approximately  30  m from  the  nest.  Observa- 
tion times  were  rotated  to  include  both  morn- 
ings and  afternoons. 

The  first  event  was  witnessed  during  a 3 h 
observation  period  on  the  afternoon  of  10  July 
at  control  nest  LGD.  The  adult  female  had  not 
been  observed  in  the  nest  stand  since  8 July, 
and  likely  had  deserted  the  nest  or  died.  Al- 
though we  were  unable  to  trap  the  female  to 
verify  her  age  prior  to  her  disappearance,  her 
unusually  dark  maroon  colored  eyes,  behavior, 
and  degree  of  scarring  above  and  around  her 
right  eye  indicated  old  age.  The  role  of  the 
male  in  caring  for  the  nestlings  after  the  fe- 
male’s disappearance  is  unknown  because  he 
was  never  observed  visiting  the  nest. 

Shortly  after  observations  had  begun,  two 
nestlings  (21  and  22  days  old)  were  begging 
periodically.  The  older  nestling  (N 1 ) then  be- 
gan flapping  its  wings  and  pecking  at  the  head 
of  nestling  2 (N2).  Nestling  2 initially  retali- 
ated by  flapping  its  wings  and  pecking  at  the 
head  of  N 1 , but  soon  turned  its  back  to  N 1 
and  assumed  the  defense  stance  described  by 
Schnell  (1958),  with  its  head  lowered  and  its 
rump  elevated.  Nestling  1 responded  by  in- 
creasing its  intensity  of  pecking  and  then  be- 
gan pulling  down  out  of  N2’s  thighs  and 


434 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3.  September  1999 


rump.  Nestling  2 uttered  a high-pitched  call 
and  moved  toward  the  edge  of  the  nest.  This 
behavior  continued  as  N2  was  forced  out  of 
the  west  side  of  the  nest  and  onto  the  nest 
branch.  Nestling  1 perched  on  the  edge  of  the 
nest  and  continued  pulling  down  from  N2’s 
rump  until  N2  moved  out  of  reach,  at  which 
point  N 1 walked  to  the  center  of  the  nest  and 
rested. 

After  approximately  15  min,  during  which 
several  strong  gusts  of  wind  nearly  blew  N2 
off  the  branch,  N2  moved  back  into  the  nest; 
N 1 resumed  aggressive  attacks  within  5 min. 
Nestling  2 again  assumed  a defense  stance, 
and  N1  began  tearing  down  from  N2’s  back- 
side, forcing  N2  onto  the  south  edge  of  the 
nest.  Nestling  1 then  began  rushing  at  N2  and 
colliding  with  N2’s  hind  end.  This  behavior 
continued  while  N2  called  and  flapped  its 
wings  in  an  attempt  to  maintain  balance  as  it 
clung  to  the  rim  of  the  nest.  Nestling  2 then 
turned  quickly  and  climbed  over  N1  and  into 
the  center  of  the  nest.  Nestling  1 pursued  N2 
to  the  east  edge  of  the  nest  next  to  the  tree 
trunk,  where  N2  again  took  a defense  stance. 
Nestling  1 resumed  ramming  and  tearing 
down  from  N2.  Nestling  2 was  knocked  out 
of  the  nest  but  caught  its  wing  on  a branch. 
Nestling  1 leaned  out  of  the  nest  and  contin- 
ued to  rip  down  from  N2  while  N2  screamed. 
Nestling  1 then  backed  off,  uttered  an  adult- 
like alarm  call,  and  returned  to  the  center  of 
the  nest.  Nestling  2 climbed  back  into  the  nest 
and  remained  in  the  nest  for  approximately  10 
min  until  N1  again  chased  N2  out  of  the  nest 
and  onto  the  nest  branch.  Nestling  2 was  not 
allowed  back  onto  the  nest  for  the  rest  of  the 
observation  period. 

Two  days  later  we  found  N2  dead  on  the 
ground  under  the  nest.  Nestling  1 directed 
loud  alarm  and  begging  calls  at  us  throughout 
the  visit  to  the  nest  stand.  Nestling  1 was 
found  dead  in  the  nest  on  14  July  1997.  Bod- 
ies of  both  nestlings  were  sent  to  the  Colorado 
Veterinary  Diagnostic  Laboratory  at  the  Col- 
lege of  Veterinary  Medicine,  Colorado  State 
University,  where  necropsies  were  performed. 
Nestling  2 was  mildly  emaciated,  had  two 
fractured  ribs,  and  pulmonary  hemorrhaging, 
presumably  incurred  during  its  fall  from  the 
nest  tree.  The  exact  cause  of  death  for  N1  was 
unknown;  however,  the  necropsy  showed  this 
bird  suffered  from  advanced  emaciation  re- 


sulting in  pectoral  muscle  atrophy,  which 
strongly  suggests  starvation. 

The  other  probable  case  of  siblicide  oc- 
curred at  the  control  nest  SNK  on  2 July  1997. 
When  we  entered  the  nest  stand  the  female 
was  not  in  the  immediate  vicinity.  Two  nest- 
lings (20  and  22  days  old)  were  in  the  nest; 
one  was  obviously  dead  with  blood  around  its 
head.  The  adult  female  returned  shortly  there- 
after, poked  at  the  dead  nestling  briefly,  and 
then  carried  the  body  away  from  the  nest.  She 
returned  within  several  minutes  without  the 
dead  nestling.  Although  we  did  not  witness 
aggression  between  the  siblings,  the  fact  that 
the  dead  nestling  was  still  in  the  nest  suggests 
that  a predator  did  not  kill  it.  Because  of  the 
obvious  head  injury,  we  believe  the  nestling’s 
death  resulted  from  siblicide  and  not  merely 
starvation.  Although,  it  is  possible  that  the  re- 
maining nestling  attempted  cannibalism  after 
its  sibling  had  died,  we  did  not  observe  the 
nestling  trying  to  feed  on  its  dead  sibling.  The 
SNK  nestling’s  death  could  also  have  been  the 
result  of  filial  infanticide  if  parental  behavior 
(e.g.,  nest  desertion,  favoritism,  or  aggression) 
contributed  to  its  death  (Mock  and  Parker 
1997).  However,  O’Connor’s  (1978)  brood  re- 
duction model  predicts  that  conditions  favor- 
able to  siblicide  will  occur  more  often  than 
those  favorable  to  filial  infanticide.  To  our 
knowledge,  filial  infanticide  resulting  from  fa- 
tal parental  aggression  has  not  been  docu- 
mented in  goshawks;  nevertheless,  we  cannot 
rule  it  out  as  a possibility. 

Several  details  of  these  observations  dif- 
fered from  those  of  similar  events  observed  in 
goshawks  and  other  raptors.  Cannibalism  was 
documented  to  have  followed  siblicide  in  a 
Swainson’s  Hawk  (Buteo  swainsoni)  nest  (Pilz 
and  Seibert  1978),  three  Burrowing  Owl 
{Athene  cuniciilaria)  nests  (Wellicome  1997), 
four  American  Kestrel  {Falco  spar\>erius) 
nests  (Bortolotti  et  al.  1991),  and  three  gos- 
hawk nests  (Schnell  1958,  Zachel  1985,  Boal 
and  Bacorn  1994).  However,  cannibalism  was 
not  observed  at  either  nest  in  this  study.  At 
the  LCD  nest  cannibalism  might  have  oc- 
curred if  N2  had  not  fallen  to  its  death;  but 
NTs  behavior  gave  no  indication  that  it  was 
attacking  N2  for  the  purpose  of  consumption. 
Nestling  1 seemed  intent  on  expelling  N2 
from  the  nest  and  Nl’s  aggression  stopped 
once  N2  was  out  of  the  nest.  If  N1  was  at- 


SHORT  COMMUNICATIONS 


435 


tempting  to  kill  N2  for  consumption  we  would 
have  expected  the  aggression  to  continue  until 
N2  was  dead.  Cannibalism  was  also  not  ob- 
served at  the  SNK  nest  but  this  may  have  been 
due  to  the  presence  of  the  adults  that  were 
providing  food  to  the  remaining  nestling,  or 
to  the  removal  of  the  dead  nestling  before  it 
could  be  cannibalized. 

A second  disparity  between  our  observa- 
tions and  those  in  the  literature  is  the  potential 
function  of  the  submissive  posture  of  the  de- 
fense stance.  According  to  Schnell’s  (1958) 
observations  of  nestling  aggression,  the  ag- 
gressor terminated  attacks  when  its  sibling  as- 
sumed the  defense  stance.  In  our  observations, 
N1  continued  aggressive  attacks  after  N2  as- 
sumed the  defense  stance.  Nestling  I’s  behav- 
ior also  differed  from  aggressor  behavior  in 
other  documented  siblicide  events  in  that  N I’s 
attacks  were  aimed  primarily  at  N2’s  rump 
and  thighs  instead  of  at  its  sibling’s  head 
(Schnell  1958,  Pilz  and  Seibert  1978,  Boal 
and  Bacom  1994). 

Although  our  LGD  siblicide  observation 
differs  in  the  aforementioned  ways  from  those 
previously  reported  by  Schnell  (1958),  Zachel 
(1985),  and  Boal  and  Bacom  (1994),  our 
event  is  similar  in  that  it  occurred  during  a 
period  of  apparent  low  food  supply.  We  did 
not  measure  food  availability  in  our  study 
area,  but  provided  half  of  our  experimental 
goshawks  with  supplemental  food.  Including 
the  LGD  nest  failure  and  the  SNK  mortality, 
we  documented  brood  reductions  at  four  of 
the  seven  control  nests  in  1997  and  no  nest- 
ling deaths  at  any  of  the  treatment  nests.  Three 
of  the  seven  (43%)  control  nests  failed  (i.e., 
fledged  no  young).  In  addition,  the  youngest 
nestling  at  one  supplemented  nest  hatched  10 
days  after  its  closest  sibling  (mean  age  differ- 
ence between  oldest  sibling  and  each  of  the 
younger  siblings  = 2.12  days)  and  was  no- 
ticeably smaller  than  its  two  nest  mates,  yet 
survived  to  fledging  age  with  little  aggression 
between  siblings.  Ward  and  Kennedy  (1996, 
unpubk  data)  documented  similar  results  in 
their  experiment,  where  a nest  with  supple- 
mental food  successfully  fledged  four  young 
including  a nestling  7-10  days  younger  than 
its  closest  sibling.  Because  nestlings  that  hatch 
significantly  later  than  their  siblings  in  asyn- 
chronous broods  often  die  unless  enough  food 
is  provided  (Bryant  and  Tatner  1990,  Wiebe 


and  Bortolotti  1995),  we  attribute  the  higher 
survival  of  these  treatment  nestlings  to  the 
high  food  abundance. 

Although  our  study  was  not  designed  to  in- 
vestigate the  role  of  food  in  sibling  aggression 
in  goshawks,  our  finding  of  higher  survival 
for  supplementally  fed  nestlings,  coupled  with 
the  siblicide  observations  provided  us  with  the 
opportunity  to  consider  this  relationship.  Lack 
(1954)  hypothesized  that  asynchronous  hatch- 
ing in  avian  species  occurs  to  facilitate  brood 
size  reduction  to  match  available  levels  of  re- 
sources provided  by  parents.  If  Lack’s  hy- 
pothesis is  correct,  occurrences  of  siblicide 
should  be  influenced  by  levels  of  prey  abun- 
dance. Forbes  and  Mock  (1994)  differentiate 
two  types  of  facultative  siblicide:  one  in 
which  aggression  is  triggered  by  food  short- 
age and  the  other  where  it  is  not.  Mock  and 
coworkers  (1987)  observed  that  the  occur- 
rence of  fatal  sibling  aggression  in  some  spe- 
cies was  only  indirectly  influenced  by  food. 
They  observed  aggressive  behavior  between 
siblings  regardless  of  food  levels,  but  mortal- 
ity from  aggression  was  lower  if  food  was 
abundant  because  the  younger  siblings  were 
sufficiently  strong  to  withstand  the  attacks. 
Our  observations  are  consistent  with  the  form 
of  facultative  siblicide  directly  influenced  by 
resource  levels  and  provides  evidence  for  the 
hypothesis  that  low  food  supplies  trigger  sib- 
ling aggression  in  goshawks.  Additional  em- 
pirical research,  coupled  with  measurement  of 
background  resource  levels  is  needed  to  fur- 
ther substantiate  this  assertion  and  clarify  the 
nature  of  the  relationship. 

ACKNOWLEDGMENTS 

These  observations  were  made  while  conducting  re- 
search funded  by  the  USDA  Forest  .Service,  Ashley 
National  Forest.  We  thank  C.  W.  Boal.  R.  R Gerhardt. 
J.  S.  Marks,  D.  W.  Mock  and  two  anonymous  referees 
for  their  helpful  reviews  of  the  manuscript.  We  would 
also  like  to  thank  J.  Cropley,  S.  Lewis,  C.  Neilson,  M. 
Painter,  and  S.  Rayroux  for  assistance  in  the  Held,  K. 
Paulin  for  her  support  of  this  research,  and  C.  Sigurd- 
son  for  generously  volunteering  her  time  and  expertise 
to  conduct  necropsies. 

LITERATURE  CITED 

Anderson,  D.  J.  1989.  The  role  of  hatching  asynchro- 
ny in  siblicidal  brood  reduction  of  two  booby  spe- 
cies. Behav.  Ecol.  Sociobiol.  25:363—368. 
Bechard,  M.  j.  1983.  Food  supply  and  the  occurrence 


436 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3.  September  1999 


of  brood  reduction  in  Swainson’s  Hawk.  Wilson 
Bull.  95:233-242. 

Boal,  C.  W.  and  J.  E.  B acorn.  1994.  Siblicide  and 
cannibalism  at  Northern  Goshawk  nests.  Auk  111: 
748-750. 

BoRTOLorn,  G.  R.,  K.  L.  Wiebe,  and  W.  M.  Iko.  1991. 
Cannibalism  of  nestling  American  Kestrels  by  their 
parents  and  sibhngs.  Can.  J.  Zool.  69:1447-1453. 

Braun,  B.  M.  and  G.  L.  Hunt,  Jr.  1983.  Brood  re- 
duction in  Black-legged  Kitti wakes.  Auk  100: 
469-476. 

Bryant,  D.  M.  and  R Tatner.  1990.  Hatching  asyn- 
chrony, sibling  competition  and  siblicide  in  nest- 
ling birds:  studies  of  swiftlets  and  bee-eaters. 
Anim.  Behav.  39:657—671. 

Creighton,  J.  C.  and  G.  D.  Schnell.  1996.  Proximate 
control  of  siblicide  in  Cattle  Egrets:  a test  of  the 
food-amount  hypothesis.  Behav.  Ecol.  Sociobiol. 
38:371-377. 

Dewey,  S.  R,  1999.  Effects  of  supplemental  food  on 
parental  care  strategies  and  juvenile  survival  in 
Northern  Goshawks.  M.Sc.  thesis,  Colorado  State 
University,  Fort  Collins. 

Drummond,  H.  and  C.  Garcia  Chavelas.  1989.  Food 
shortage  influences  sibling  aggression  in  the  Blue- 
footed Booby.  Anim.  Behav.  37:806—819. 

Edwards,  T.  C.,  Jr.  and  M.  W.  Collopy.  1983.  Ob- 
ligate and  facultative  brood  reduction  in  eagles: 
an  examination  of  factors  that  influence  fratricide. 
Auk  100:630-635. 

Forbes,  L.  S.  and  D.  W.  Mock.  1994.  Proximate  and 
ultimate  determinants  of  avian  brood  reduction. 
Pp.  237-256  in  Infanticide  and  parental  care  (S. 
Parmigiani  and  F.  S.  vom  Saal,  Eds.).  Hardwood 
Academic  Publ.,  Chur,  Switzerland. 

Gerhardt,  R.  P,  D.  M.  Gerhardt,  and  M.  A.  Vas- 
QUEZ.  1997.  Siblicide  in  Swallow-tailed  Kites. 
Wilson  Bull.  109:1 12-120. 

Heeb,  P.  1994.  Intraclutch  egg-mass  variation  and 
hatching  asynchrony  in  the  Jackdaw  Con'us  mo- 
nedula.  Ardea  82:287—297. 

Heinsohn,  R.  G.  1995.  Hatching  a.synchrony  and  brood 
reduction  in  cooperatively  breeding  White-winged 
Choughs  Corcora.x  melanorhamphos.  Emu  95: 
252-258. 


Lack,  D.  1954.  The  natural  regulation  of  animal  num- 
bers. Oxford  Univ.  Press,  Oxford,  U.K. 

Mock,  D.  W.  and  T.  C.  Lamey.  1991.  The  role  of 
brood  size  in  regulating  egret  sibling  aggression. 
Am.  Nat.  138:1015-1026. 

Mock,  D.  W.  and  G.  A.  Parker.  1997.  The  evolution 
of  sibling  rivalry.  Oxford  Univ.  Press,  Oxford, 
U.K. 

Mock,  D.  W.,  T.  C.  Lamey,  and  B.  J.  Ploger.  1987. 
Proximate  and  ultimate  roles  of  food  amount  in 
regulating  egret  sibling  aggression.  Ecology  68: 
1760-1772. 

Mock,  D.  W.,  H.  Drummond,  and  C.  H.  Stinson. 
1990.  Avian  siblicide.  Am.  Sci.  78:438-449. 

Newton,  I.  1979.  Population  ecology  of  raptors.  Buteo 
Books,  Vermillion,  South  Dakota. 

O’Connor,  R.  J.  1978.  Brood  reduction  in  birds:  se- 
lection for  fratricide,  infanticide  and  suicide. 
Anim.  Behav.  26:79-96. 

PiLZ,  W.  R.  AND  L.  K.  Seibert.  1978.  Fratricide  and 
cannibalism  in  Swainson’s  Hawk.  Auk  95:584- 
585. 

Reynolds,  P.  S.  1996.  Brood  reduction  and  siblicide 
in  Black-billed  Magpies  (Pica  pica).  Auk  113: 
189-199. 

Schnell,  J.  H.  1958.  Nesting  behavior  and  food  habits 
of  goshawks  in  the  Sierra  Nevada  of  California. 
Condor  60:377-403. 

Stinson,  C.  H.  1979.  On  the  selective  advantage  of 
fratricide  in  raptors.  Evolution  33:1219-1225. 

Ward,  J.  M.  and  P.  L.  Kennedy.  1996.  Effects  of  sup- 
plemental food  on  growth  and  survival  of  juvenile 
Northern  Goshawks.  Auk  113:200-208. 

Wellicome,  T.  I.  1997.  Reproductive  performance  of 
Burrowing  Owls  (Speotyto  cimicularia):  effects  of 
supplemental  food.  J.  Raptor  Res.  9:68-73. 

Wiebe,  K.  L.  and  G.  R.  Bortolotti.  1995.  Food-de- 
pendent  benefits  of  hatching  asynchrony  in  Amer- 
ican Kestrels  Falco  .xpan’eriu.K.  Behav.  Ecol.  So- 
ciobiol. 36:49-57. 

Zachel,  C.  R.  1985.  Food  habits,  hunting  activity,  and 
post-fledging  behavior  of  Northern  Goshawks 
(Accipiter  gentili.x)  in  interior  Alaska.  M.Sc.  the- 
sis, Univ.  of  Alaska,  Fairbanks. 


SHORT  COMMUNICATIONS 


437 


Wilson  Bull.,  I 1 1(3),  1999,  pp.  437-439 


Cooperative  Foraging  in  the  Mountain  Caracara  in  Peru 

Jason  Jones'  - 


ABSTRACT. — Cooperative  foraging  behavior  is 
rarely  observed  in  ground-walking  birds.  I report  on 
observations  of  cooperative  foraging  behavior  by 
Mountain  Caracaras  (Phalcoboenits  megaloptenis)  in 
the  puna  region  of  Peru  in  September  of  1995.  On 
several  occasions,  three  individuals  (two  adults  and 
one  immature)  were  observed  working  together  to  turn 
over  large  rocks  to  obtain  prey  from  beneath.  These 
cooperative  foraging  events  are  notable  in  that,  unlike 
cooperative  foraging  behavior  observed  in  other 
ground-walking  birds,  only  one  individual  obtained 
prey  from  a given  cooperative  effort.  The  presence  of 
the  immature  individual  may  be  indicative  of  delayed 
dispersal,  a behavior  not  previously  described  for  this 
poorly  known  species.  Received  14  Dec.  1998,  ac- 
cepted 28  March  1999. 


Cooperative  foraging  involves  two  or  more 
individual  organisms  assisting  one  another  in 
obtaining  a food  item.  Among  vertebrates, 
this  behavior  is  well  described  in  many  social 
mammals  (Macdonald  1983,  Serfass  1995) 
and  some  fish  (Dugatkin  and  Mesterton-Gib- 
bons  1996)  but  is  relatively  uncommon  in 
birds  (Sullivan  1984).  Among  birds,  cooper- 
ative hunting  appears  most  frequently  in  sea- 
birds (Parasitic  Jaegers,  Stercorarius  parasi- 
ticus, Pruett- Jones  1980;  Brown  and  Ameri- 
can White  pelicans,  Pelecanus  occidentalis 
and  P.  erythrorhynchos,  J.  Jones,  unpub. 
data).  Examples  from  land  birds  include  the 
cooperative  hunting  behavior  exhibited  by 
Harris’  Hawks  (Parabuteo  unicinctus;  Mader 
1979,  Bednarz  1988),  Golden  Eagles  (Aqiiila 
chrysaetos;  Collopy  1983)  and  Crested  Ca- 
racaras {Caracara  plancus\  Morrison  1996). 
Most  of  these  instances  of  cooperative  hunt- 
ing involve  highly  mobile  prey  items;  coop- 
erative foraging  for  less  mobile  organisms  is 
uncommon  (Sullivan  1984).  In  this  report,  I 


' 4974  Lakeshore  Rd.,  Kelowna,  British  Columbia, 
Canada,  VI Y 7R3. 

- Present  address:  Dept,  of  Biology,  Queen’s  Univ., 
Kingston,  Ontario,  Canada,  K7L  3N6; 

E-mail:  jonesja® biology. queensu.ca 


detail  observations  of  cooperative  foraging 
behavior  in  the  Mountain  Caracara  (Phalco- 
boenus  megalopterus).  This  species  is  adept 
at  ground  foraging  and  in  non-urban  areas 
feeds  on  large  arthropods,  rodents,  and  birds 
(Brown  and  Amadon  1968).  Breeding  usually 
occurs  between  October  and  December  with 
two,  rarely  three,  eggs  laid  (Brown  and  Ama- 
don 1968). 

STUDY  AREA 

The  observations  were  made  in  the  Peruvian  puna 
zone  on  the  road  between  the  towns  of  Quillabamba 
and  Ollantaytambo,  Department  of  Cuzco  (1 3°  9'  S, 
72°  14'  W;  3750  m elevation).  This  region  is  charac- 
terized by  dry  grasslands,  dominated  by  genera  such 
as  Calamagrostis  and  Festuca  with  interspersed  shrubs 
of  the  genera  Astragalus,  Berberis,  and  Lupiniis  (Park- 
er et  al.  1982).  Mountain  Caracaras  are  common  in  this 
region  (Parker  et  al.  1982)  and  are  often  found  near 
towns  where  they  feed  on  refuse  and  carrion  (White 
and  Boyce  1987). 

RESULTS 

I observed  cooperative  rock-turning  on  four 
occasions  from  1-6  September  1995,  as  I 
watched  three  individuals  (two  adults  and  one 
immature)  foraging  together  on  the  puna 
grassland.  The  immature  bird  was  easily  dis- 
tinguished by  its  plumage.  On  each  occasion, 
one  of  the  adults  approached  a large  rock, 
walked  around  it,  uttered  a high-pitched 
kieeer,  and  then  stood  by  the  rock.  Apparently 
responding  to  the  vocalization,  the  other  in- 
dividuals joined  the  first  at  the  rock  and  pro- 
ceeded to  work  together  to  flip  the  rock  from 
its  resting  place,  with  each  bird  using  one  of 
its  talons.  The  bird  that  made  the  call  partic- 
ipated in  the  the  turning  but  also  appeared  to 
act  as  a “watcher”  and  was  the  individual  re- 
sponsible for  prey  capture.  On  one  occasion, 
the  item  was  captured  by  an  adult  which  then 
gave  it  to  the  immature  bird.  No  begging  vo- 
calizations were  uttered  by  the  younger  bird 
nor  did  it  adopt  any  unusual  posture.  Each  lift- 
ing event  took  approximately  30  min  from  call 


438 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3.  September  1999 


to  prey  capture.  This  species,  although  a 
ground-foraging  specialist,  has  weak  legs 
(Brown  and  Amadon  1968).  After  examining 
the  rocks  (approximate  dimensons  30  X 20  X 
10  cm),  1 do  not  believe  that  one  individual 
could  have  turned  over  any  of  the  rocks  by 
itself. 

At  no  time  during  the  days  of  observation 
was  the  immature  bird  more  than  100  m from 
one  or  the  other  of  the  adults,  although  the 
adults  were  often  separated  by  distances  up  to 
500  m.  In  a series  of  12  one-hour  watches  {n 
= 4 for  each  individual),  I determined  that  the 
two  adults  appeared  to  spend  more  time  for- 
aging than  did  the  immature  bird  (adult  = 
68.6%,  imm.  = 52.3%)  although  the  differ- 
ence was  not  statistically  significant  (x“  = 
3.334,  df  = 1,  P = 0.067). 

Subsequent  investigation  revealed  that 
worms  or  arthropods  could  be  found  under 
most  (19  of  24)  of  the  rocks  in  the  vicinity; 
several  of  the  rocks  (5  of  24)  also  hid  rodent 
runways.  Examination  of  the  surrounding 
grassland  showed  that  prey  items  of  a similar 
size  but  different  taxa  (e.g.,  grasshoppers  rath- 
er than  millipedes)  were  available  without 
rock  lifting. 

DISCUSSION 

Unlike  most  land  bird  species.  Mountain 
Caracaras  were  not  using  cooperative  foraging 
behavior  to  pursue  and  subdue  large,  highly 
mobile  prey  items.  Rather,  they  required  co- 
operation to  obtain  access  to  otherwise  inac- 
cessible prey  items.  While  the  turning  over  of 
small  rocks  was  one  of  the  main  foraging  be- 
haviours exhibited  by  these  individuals,  the 
cooperative  rock-turning  events  did  not  occur 
within  the  set  watches  and  did  not  seem  to 
represent  a major  foraging  strategy  for  these 
individuals.  The  main  difference  between  the 
cooperative  behavior  observed  in  Mountain 
Caracaras  and  that  observed  in  other  cooper- 
atively foraging  birds  is  that  only  one  individ- 
ual obtained  food  from  a given  foraging  event; 
Harris’  Hawks,  for  example,  share  large  prey 
that  are  cooperatively  caught  (Bednarz  1988). 
This  disparity  in  obtaining  a food  reward  may 
even  out  over  time  (e.g.,  one  of  the  four  prey 
items  was  given  to  the  immature  bird)  but 
there  apparently  is  often  no  immediate  reward 
for  some  of  the  individuals  participating.  That 
individuals  are  willing  to  help  without  a re- 


ward is  perhaps  indicative  of  the  length  of 
time  these  birds  remain  together  as  a foraging 
unit;  that  is,  an  individual  is  willing  to  help 
today  because  its  turn  will  come  eventually 
(see  Trivers  1971  for  discussion  of  reciprocal 
altruism). 

Congeners  of  the  Mountain  Caracara  hatch 
their  eggs  in  December  and  fledglings  are  usu- 
ally independent  by  March  (Brown  and  Ama- 
don 1968).  As  my  observations  took  place  in 
September,  the  immature  member  of  the  trio 
was  probably  a chick  from  a previous  breed- 
ing effort  and  its  presence,  therefore,  may  rep- 
resent delayed  dispersal.  Delayed  dispersal  is 
fairly  common  in  Neotropical  raptors  (Mader 
1981).  How  common  delayed  dispersal  is  in 
Mountain  Caracaras  and  how  it  may  affect  the 
incidence  of  cooperative  foraging,  is  uncer- 
tain. 

ACKNOWLEDGMENTS 

I thank  J.  Barg,  E.  Canuthers,  F.  Chavez-Ramirez, 
D.  Jone.s,  W.  Rendell  and  four  anonymous  reviewers 
for  providing  helpful  comments  on  earlier  drafts  of  this 
manuscript. 

LITERATURE  CITED 

Bednarz,  J.  C.  1988.  Cooperative  hunting  in  Hanis’ 
Hawks  {Parahuteo  unicinctus).  Science  239: 
1525-1527. 

Brown,  L.  and  D.  Amadon.  1968.  Eagles,  hawks  and 
falcons  of  the  world.  Vol.  2.  McGraw-Hill,  New 
York. 

COLLOPY,  M.  W.  1983.  Foraging  behavior  and  success 
of  Golden  Eagles.  Auk  100:747-749. 

Dugatkin,  L.  a.  and  M.  Mesterton-Gibbons.  1996. 
Cooperation  among  unrelated  individuals:  recip- 
rocal altruism,  by-product  mutualism  and  group 
selection  in  tishes.  Biosystems  37:19-30. 
Macdonald,  D.  W.  1983.  The  ecology  of  carnivore 
social  behaviour.  Nature  301:379-384. 

Mader.  W.  J.  1979.  Breeding  behavior  of  a polyan- 
drous  trio  of  Harris'  Hawks  in  southern  Arizona. 
Auk  96:776-788. 

Mader,  W.  J.  1981.  Notes  on  nesting  raptors  in  the 
llanos  of  Venezuela.  Condor  83:48-51. 

Morri.son,  j.  L.  1996.  Crested  Caracara  (Ccinieara 
plancus).  In  The  birds  of  North  America,  no.  249 
(A.  Poole  and  E Gill,  Eds.).  The  Academy  of  Nat- 
ural Sciences,  Philadelphia,  Pennsylvania;  .The 
American  Ornithologists'  Union,  Washington,  D.C. 
Parker,  T.  A.,  Ill,  S.  A.  Parker,  and  M.  A.  Plenge, 
1982.  An  annotated  checklist  of  Peruvian  birds. 
Buteo  Books,  Vermillion,  South  Dakota. 

Pruett -Jones,  S.  G.  1980.  Team-hunting  and  food 
sharing  in  Parasitic  Jaegers.  Wilson  Bull.  92:524- 
526. 


SHORT  COMMUNICATIONS 


439 


Serfass,  T.  L.  1995.  Cooperative  toragiiig  by  North 
American  river  ottens.  Lutra  caiuuk'iisis.  Can. 
Field-Nat.  109:458-459. 

Sullivan,  K.  A.  1984.  Cooperative  foraging  and  court- 
ship feeding  in  the  Laughing  Gull.  Wilson  Bull. 
96:710-71  1. 


Trivf:r.s,  R.  L.  1971.  The  evolution  of  rceiprtteal  altru- 
ism. Q.  Rev.  Biol.  46:35-57. 

White,  C.  M.  and  D.  A.  Boyce.  1987.  Notes  on  the 
Mountain  Caracara  ( kluilcohoenu.s  inef>cilople- 
rus)  in  the  Argentine  puna.  Wilson  Bull.  99: 
283-284. 


Wilson  Bull..  1 1 1(3),  1999,  pp.  439-440 


Predation  by  Rufous  Motmot  on  Black-and-Green  Poison  Dart  Frog 

Terry  L.  Master' 


ABSTRACT. — I observed  a Rufous  Motmot  (Bar- 
yphthengus  martii)  feeding  a black-and-green  poison 
dart  frog  (Denclrobates  aurotiis)  to  another  motmot  in 
the  Caribbean  Slope  lowland  rainforest  of  northeastern 
Costa  Rica.  Neither  individual  appeared  to  suffer  any 
ill  effects  from  what  was  probably  courtship  feeding. 
Small  vertebrates  are  typical  prey  for  the  larger  species 
of  motmots.  Blue-crowned  Motmots  (Momotus  moni- 
ota)  have  been  observed  consuming  several  species  of 
poison  dart  frogs  raised  in  captivity  but  captive  reared 
frogs  either  do  not  contain,  or  have  reduced  levels  of, 
the  toxins  that  native  frogs  produce.  Relatively  little  is 
known  about  the  effects  of  poison  dart  frog  toxins  on 
predators.  Presumably,  the  digestive  system  of  the  Ru- 
fous Motmot  is  capable  of  neutralizing  the  potentially 
toxic  effects  of  such  prey.  Received  15  Sept.  1998, 
accepted  15  Feb.  1998. 


Poison  dart  frogs  have  long  been  known  to 
possess  toxic  skin  secretions,  and,  because  of 
their  bright  coloration,  are  thought  to  be  apo- 
sematic  to  visually  hunting  predators  such  as 
Rufous  Motmots  {Baryphthengus  martii) 
which  presumably  have  excellent  color  vision 
(Brodie  and  Tumbarello  1977).  Smith  (1975) 
demonstrated  that  hand-reared  Torquoise- 
browed  Motmots  (Eumomota  superciliosa) 
showed  an  innate  avoidance  of  snake-shaped 
models  with  patterns  simulating  those  of  coral 
snakes.  All  other  snake  models  were  readily 
attacked  implying  that  aposematic  coloration 
is  a deterrent  to  this  species.  Observations  in- 
dicate that  Blue-crowned  Motmots  {Momotu.s 
momota)  at  the  National  Aquarium  consume 
several  species  of  poison  dart  frogs  including 


' Dept,  of  Biological  Sciences,  East  Stroudsburg 
Univ.,  East  Stroudsburg,  PA  18301; 

E-mail:  tmaster@esu.edu 


the  black-and-green  poison  dart  frog  (Dendro- 
bates  auratus)  and  phantasmal  poison  dart 
frog  (Dendrobates  tricolor).  However,  these 
frogs  were  raised  in  captivity  and  either  do  not 
produce  or  have  relatively  low  levels  of  the 
characteristic  skin  toxins  (Kricher  1997;  C. 
Rowsom,  pers.  comm.). 

At  approximately  9:30  CST  on  26  March 
1995,  an  adult  Rufous  Motmot  was  observed 
in  secondary  lowland  tropical  forest  from  a 
hiking  trail  located  at  Estacion  Biologica  La 
Suerte,  near  Cariari,  Limon  Province,  north- 
eastern Costa  Rica  (10°  26'  N,  83°  46'  W). 
The  bird  landed  25  m from  the  trail  on  an 
exposed  perch  3 m above  the  ground  and  was 
easily  observed  for  approximately  4 min.  Af- 
ter 4 min  another  individual  landed  on  the 
same  branch  next  to  the  first  individual.  The 
newly  arrived  motmot  was  carrying  a black- 
and-green  poison  dart  frog  in  its  beak  which 
it  fed  immediately  to  the  first  individual.  It  is 
not  possible  to  distinguish  between  sexes  in 
Rufous  Motmots;  however,  this  behavior  was 
interpreted  as  a male  who  was  feeding  the  fe- 
male as  a courtship  gesture.  Both  individuals 
had  diagnostic  black  breast  marks  and  raquet- 
tails  indicative  of  adult  birds,  suggesting  that 
this  was  probably  not  a fledgling  being  fed. 
The  pair  continued  sitting  on  the  branch  for 
approximately  30  min  after  which  they  flew 
off  together  into  the  forest.  Neither  individual 
appeared  to  suffer  any  ill  effects  from  either 
grasping  or  consuming  the  poison  dart  frog. 

The  typical  diet  of  motmots  varies  somewhat 
in  conjunction  with  body  size.  Smaller  species 
prefer  insects  while  larger  species  consume  in- 
sects along  with  other  invertebrates,  small  ver- 


440 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  3,  September  1999 


tebrates,  and  fmit  (Orejuela  1980,  Remsen  et 
al.  1993).  The  Rufous  Motmot  consumes  ar- 
thropods, other  invertebrates  including  crabs, 
small  vertebrates  including  fish,  lizards  and 
birds,  as  well  as  fruit  (Remsen  et  al.  1993). 
Frogs  have  been  reported  as  a dietary  compo- 
nent of  the  Rufous,  Broad-billed  {Electron  pla- 
tyrhynchum)  and  Torquoise-browed  motmots 
(Remsen  et  al.  1993),  and  Blue-crowned  Mot- 
mots  in  captivity  (C.  Rowsom,  pers.  comm.). 

The  effect  of  poison  dait  frog  toxins  on  var- 
ious potential  predators  has  received  relatively 
little  attention.  Brodie  and  Tumbarello  (1977) 
tested  the  response  of  garter  snakes  {Thamno- 
phis  sirtalis)  to  D.  auratus  offered  as  prey. 
Snakes  readily  mouthed,  or  in  some  cases  con- 
sumed the  frogs  but  all  exhibited  head  shaking, 
mouth  opening,  convulsions,  and  loss  of  equi- 
librium. Only  one  snake  actually  died  and  that 
was  after  consuming  its  third  frog.  These 
snakes  do  not  possess  color  vision  and  might 
not  be  influenced  by  the  aposematic  coloration 
to  the  extent  that  an  organism  with  color  vision 
would  be  (Brodie  and  Tumbarello  1977). 

While  motmots  in  general  may  be  warned 
by  aposematic  coloration,  the  Rufous  Motmot 
at  least  is  capable  of  handling  and  consuming 
this  particular  species  of  poison  dart  frog. 
Dendrobates  auratus  reaches  densities  of  1 in- 
dividual/180  m^  in  one  locality  at  La  Suerte 


known  to  be  frequented  by  Rufous  Motmots 
(B.  Graves,  pers.  comm.)  One  pair  was  ob- 
served on  the  ground  rummaging  through  leaf 
litter  where  they  would  undoubtedly  encoun- 
ter D.  auratus  (B.  Graves,  pers.  comm.).  The 
level  of  toxins  in  the  frogs  of  this  area,  how 
the  motmots  physiologically  handle  the  tox- 
ins, and  the  frequency  with  which  they  con- 
sume D.  auratus  remain  unknown. 

ACKNOWLEDGMENTS 

I would  like  to  thank  Dr.  T LaDuke  of  East  Strouds- 
burg University,  Dr.  B.  Graves  of  Northern  Michigan 
University,  and  C.  Row.som  of  the  National  Aquarium 
in  Baltimore  for  encouragement  and  assistance  with 
this  manuscript. 

LITERATURE  CITED 

Brodie,  E.  D.  and  M.  S.  Tumbarello.  1977.  The  an- 
tipredator functions  of  Dendrobates  auratus  (Am- 
phibia, Anura,  Dendrobatidae)  skin  secretion  in 
regard  to  a snake  predator.  J.  Herp.  12:264—265. 
Kricher,  J.  1977.  A Neotropical  companion.  Princeton 
Univ.  Press.,  Princeton,  New  Jersey. 

Orejuela,  J.  E.  1980.  Niche  relationships  between 
Torquoise-browed  and  Blue-crowned  motmots. 
Wilson  Bull.  92:229-244. 

Remsen,  J.  V.,  M.  A.  Hyde,  and  A.  Chapman.  1993. 
The  diets  of  Neotropical  trogons,  motmots,  bar- 
bets  and  toucans.  Condor  95:178-192. 

Smith,  S.  M.  1975.  Innate  recognition  of  coral  snake 
pattern  by  a possible  avian  predator.  Nature 
(Lond.)  187:759-760. 


Wil.son  Bull..  111(3),  1999,  pp.  440-442 


Evidence  Of  Egg  Ejection  In  Mountain  Bluebirds 

Percy  N.  Hebert' 


ABSTRACT. — When  the  last  two  eggs  of  Mountain 
Bluebird  {Sialia  currucoides)  clutches  were  replaced 
with  another  bluebird  egg  and  one  House  Sparrow 
{Passer  domesticus)  egg,  20%  (3/15)  of  the  spaiTOW 
eggs  were  removed  within  24  hr.  None  of  the  surrogate 
bluebird  eggs  was  removed.  This  is  the  first  recorded 
instance  of  interspecific  egg  ejection  in  a bluebird  spe- 
cies, and  hole-nesters  in  general.  Received  2 Nov. 
I99S,  accepted  IH  Feb.  1999. 


' Dept,  of  Zoology,  Univ.  of  Manitoba.  Winnipeg, 
Manitoba.  Canada,  R3T  2N2; 

E-mail:  phcberKScc.umanitoba.ca 


Of  the  approximately  140  biological  hosts 
of  the  Brown-headed  Cowbird  (Molothrus 
ater),  fewer  than  7%  have  been  classified  as 
rejectors  (Friedmann  and  Kiff  1985,  Ortega 
1998).  Rejectors  typically  remove  cowbird 
eggs  from  the  nest  within  24  hr  of  introduc- 
tion (Rothstein  1982).  Ejection  is  accom.- 
plished  either  by  grasping  the  cowbird  egg. be- 
tween the  mandibles  or  by  puncturing  the  egg 
with  the  beak  and  then  lifting  the  egg  out  of 
the  nest  (Sealy  1996).  Acceptors,  by  contrast, 
do  not  remove  cowbird  eggs  and  in  most  cases 
provision  the  cowbird  nestling(s)  (see  Petit 
1991,  Sealy  1996). 


SHORT  COMMUNICATIONS 


441 


Unlike  the  Shiny  Cowbird  {M.  honarensis), 
the  Brown-headed  Cowbird  infrequently  par- 
asitizes hole-nesters  (Ortega  1998;  but  see 
Petit  1991).  Bluebirds  (Sicilia  spp.)  are  para- 
sitized infrequently  by  Brown-headed  Cow- 
birds  (Friedmann  and  Kiff  1985).  Cowbird 
eggs  have  been  found  in  0.2-2. 6%  of  Eastern 
Bluebird  (S.  sialia)  nests,  but  there  are  only  4 
records  of  parasitism  on  Mountain  Bluebirds 
(S.  currucoides)  and  none  for  the  Western 
Bluebird  (S.  mexicanus;  Friedmann  and  Kiff 
1985).  These  low  frequencies  of  parasitism 
may  be  due  to  aggression  by  adult  bluebirds 
towards  female  cowbirds  (Gowaty  and  Wag- 
ner 1988).  Furthermore,  the  cowbird  parasit- 
izes smaller  hosts  than  itself  (Friedmann  et  al. 
1977),  thus  female  cowbirds  may  be  too  large 
to  squeeze  through  bluebird  cavity  entrances 
(Friedmann  et  al.  1977,  Pribil  and  Pieman 
1997). 

Given  such  low  frequencies  of  parasitism 
by  Brown-headed  Cowbirds  on  hole-nesters  in 
general  (Friedmann  and  Kiff  1985),  apparent- 
ly there  has  been  little  selection  pressure  fa- 
voring the  evolution  of  rejection  behavior 
(Davies  and  Brooke  1989).  In  fact,  there  is 
only  one  published  record  of  interspecific  egg 
ejection  in  hole  nesting  species  (Moksnes  et 
al.  1990).  Here  I present  data  that  indicate  that 
Mountain  Bluebirds  apparently  cannot  distin- 
guish between  conspecific  eggs,  whereas  they 
can  recognize  interspecific  eggs  as  different 
from  their  own,  and  that  these  eggs  are  some- 
times removed  from  the  nest. 

METHODS 

I collected  the  data  between  May  and  July,  in  1995 
and  1996,  on  a population  of  Mountain  Bluebirds  nest- 
ing in  boxes  near  Virden,  Manitoba  (49°  51'  N, 
100°  55'  W).  Nest-boxes  were  visited  every  2-3  days 
during  ne.st-building  and  daily  during  laying.  Eggs 
were  measured  and  weighed  within  24  hr  of  laying, 
and  numbered  on  the  blunt  end  using  a non-toxic  felt 
marker.  Once  the  clutch  was  complete,  the  penultimate 
and  ultimate  eggs  were  removed  for  24  and  48  hr,  re- 
spectively. To  minimize  the  risk  of  abandonment,  some 
of  these  clutches  received  one  bluebird  egg  from  failed 
clutches  and  one  House  Sparrow  [Passer  clomeslicus) 
egg.  The  presence  or  absence  of  these  replacement 
eggs  was  then  recorded  24  and  48  hours  later  when 
the  original  eggs  were  returned  to  their  clutches. 

Because  House  Sparrow  eggs  are  very  similar  to 
Brown-headed  Cowbird  eggs  (see  Lowther  1993, 
Lowther  and  Cink  1992),  I expected  bluebirds  to  re- 
spond to  a sparrow  egg  the  same  way  they  would  re- 
spond to  a cowbird  egg  (see  also  Rothstein  1977). 


RESULTS 

Fifteen  nests  received  a bluebird  egg  and  a 
sparrow  egg,  and  none  of  these  nests  was 
abandoned.  None  of  the  replacement  bluebird 
eggs  was  removed  from  the  nest  within  48  hr. 
By  contrast,  3/15  (20%)  of  the  sparrow  eggs 
were  removed  from  the  nest,  all  within  24  hr 
of  introduction.  In  2 of  the  3 ejections,  the 
sparrow  egg  was  removed  from  the  nest-box, 
whereas  in  the  third  instance  the  undamaged 
egg  ended  up  on  the  rim  of  the  nest.  For  both 
years  combined,  12  bluebird  eggs  were  known 
to  have  been  cracked  or  dented  during  mea- 
suring. Of  these,  one  was  found  on  the  rim  of 
the  nest  cup  the  following  day,  7 were  gone 
the  following  day,  and  4 remained  in  the  nest. 

DISCUSSION 

The  results  of  this  study  indicate  that 
Mountain  Bluebirds  are  capable  of  egg  ejec- 
tion. Mountain  Bluebirds  possess  several  traits 
that  Rothstein  (1975)  identified  as  pre-adap- 
tations for  the  evolution  of  ejection  behavior. 
For  instance,  Rothstein  (1975)  suggested  that 
the  evolution  of  ejection  behavior  would  be 
facilitated  if  the  hosts’  eggs  differed  from 
those  of  the  cowbird  in  at  least  two  respects: 
base  color,  maculation,  and  size.  Mountain 
Bluebird  eggs  differ  from  sparrow  and  cow- 
bird eggs  in  color  and  maculation  (see  Lowth- 
er 1993,  Lowther  and  Cink  1992,  Power  and 
Lombardo  1996).  Mountain  Bluebirds  can  re- 
move their  damaged  eggs  from  the  nest;  their 
eggs  are  similar  in  size  to  those  of  the  cow- 
bird. Thus  it  can  be  assumed  that  bluebirds 
would  be  capable  of  removing  cowbird  eggs 
from  their  nests. 

Given  that  Mountain  Bluebirds  are  sympat- 
ric  with  Brown-headed  Cowbirds  (see  Lowth- 
er 1993,  Power  and  Lombardo  1996),  and  that 
their  eggs  are  sufficiently  different  to  facilitate 
recognition  and  ejection  of  a cowbird  egg,  a 
rejection  rate  of  only  20%  would  appear  to  be 
low.  However,  the  low  rate  of  ejection  I ob- 
served is  likely  an  underestimate  of  the  fre- 
quency of  ejection  behavior  in  Mountain 
Bluebirds.  For  example,  Rothstein  (1982)  ob- 
served that  American  Robins  (Turdus  migra- 
torius)  are  less  likely  to  eject  cowbird  eggs 
that  are  introduced  into  the  nest  after  laying. 
As  I introduced  sparrow  eggs  at  clutch  com- 
pletion, it  is  thus  possible  that  bluebirds  were 


442 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


less  likely  to  eject  them.  Furthermore,  the 
ejected  sparrow  eggs  were  removed  within  24 
hr  of  placement  in  bluebird  nests.  Such  a 
quick  response  is  typical  of  most  rejector  spe- 
cies (Sealy  1996)  and  suggests  at  least  a mod- 
erate level  of  intolerance  (sensu  Rothstein 
1982)  to  cowbird  parasitism  in  Mountain 
Bluebirds. 

ACKNOWLEDGMENTS 

I thank  E Asselin,  H.  Khan,  and  H.  McElmoyle  for 
their  assistance  in  the  field.  J.  Hare,  A.  Smith  and  the 
Canadian  Wildlife  Service  provided  valuable  logistical 
support.  I also  thank  S.  Rondeau  for  providing  some 
of  the  House  Sparrow  eggs.  S.  G.  Sealy,  and  3 anon- 
ymous reveiwers  made  helpful  comments  on  a previ- 
ous draft  of  the  manuscript.  This  study  was  supported 
by  a research  grant  from  the  Natural  Sciences  and  En- 
gineering Research  Council  of  Canada. 

LITERATURE  CITED 

Davies,  N.  B.  and  M.  de  L.  Brooke.  1989.  An  ex- 
perimental study  of  co-evolution  between  the 
Cuckoo,  Cuculus  canoriis,  and  its  hosts.  II.  Host 
egg  markings,  chick  discrimination  and  general 
discussion.  J.  Anim.  Ecol.  58:225-236. 
Friedmann,  H.,  L.  E Kief,  and  S.  I.  Rothstein.  1977. 
A further  contribution  to  knowledge  of  the  host 
relations  of  the  parasitic  cowbirds.  Smithson. 
Contr.  Zool.  235:1-75. 

Friedmann,  H.  and  L.  E Kiff.  1985.  The  parasitic 
cowbirds  and  their  hosts.  Proc.  West.  Found.  Vert. 
Zool.  2:226-302. 

Gowaty,  P.  a.  and  S.  j.  Wagner.  1988.  Breeding  sea- 
son aggression  of  female  and  male  Eastern  Blue- 
birds (Sialia  .sialis)  to  models  of  potential  conspe- 
cific  and  interspecihc  egg  dumpers.  Ethology  78: 
238-250. 

Lowther,  P.  E.  1993.  Brown-headed  Cowbird  (Mol- 
othrii.s  ater).  In  The  birds  of  North  America,  no. 


47  (A.  Poole  and  F.  Gill,  Eds.).  The  Academy  of 
Natural  Sciences,  Philadelphia,  Pennsylvania;  The 
American  Ornithologists’  Union,  Washington, 
DC. 

Lowther,  P.  E.  and  C.  L.  Cink.  1992.  House  Sparrow. 
In  The  birds  of  North  America,  no.  12  (A.  Poole, 
P.  Stettenheim,  and  E Gill,  Eds.).  The  Academy 
of  Natural  Sciences,  Philadelphia,  Pennsylvania; 
The  American  Ornithologists’  Union,  Washington, 
DC. 

MOKSNES,  a.,  E.  R0SKAFT,  A.  T.  Braa,  L.  Korsnes,  H. 
M.  Lampe,  and  H.  C.  Pedersen.  1990.  Behav- 
ioural responses  of  potential  hosts  towards  artih- 
cial  cuckoo  eggs  and  dummies.  Behaviour  116: 
64-89. 

Ortega,  C.  P.  1998.  Cowbirds  and  other  brood  para- 
sites. Univ.  of  Arizona  Press,  Tucson. 

Petit,  L.  J.  1991.  Adaptive  tolerance  of  cowbird  par- 
asitism by  Prothonotary  Warblers:  a consequence 
of  nest-site  limitation.  Anim.  Behav.  41:425-432. 

Power,  H.  W.  and  M.  P.  Lombardo.  1996.  Mountain 
Bluebird.  In  The  birds  of  North  America,  no.  222 
(A.  Poole  and  E Gill,  Eds.).  The  Academy  of  Nat- 
ural Sciences,  Philadelphia,  Pennsylvania;  The 
American  Ornithologists’  Union,  Washington, 
DC. 

Pribil,  S.  and  j.  Picman.  1997.  Parasitism  of  House 
Wren  nests  by  Brown-headed  Cowbirds:  why  is  it 
so  rare?  Can.  J.  Zool.  75:302-307. 

Rothstein,  S.  I.  1975.  An  experimental  and  teleonom- 
ic  investigation  of  avian  brood  parasitism.  Condor 
77:250-271. 

Rothstein,  S.  I.  1977.  Cowbird  parasitism  and  egg 
recognition  of  the  Northern  Oriole.  Wilson  Bull. 
89:21-32. 

Rothstein,  S.  I.  1982.  Mechanisms  of  avian  egg  rec- 
ognition: which  egg  parameters  elicit  responses  by 
rejecter  species?  Behav.  Ecol.  Sociobiol.  11:229- 
239. 

Sealy,  S.  G.  1996.  Evolution  of  host  defenses  against 
brood  parasitism:  implications  of  puncture-ejec- 
tion by  a small  passerine.  Auk  1 13:346-355. 


SHORT  COMMUNICATIONS 


443 


Wilson  Bull..  111(3),  1999,  pp.  443-444 


Foraging  Ovenbird  Follows  Armadillo 

Douglas  J.  Levey' 


ABSTRACT — I report  an  observation  of  a foraging 
Ovenbird  (Seiurus  aurocapillus)  following  a nine- 
banded  armadillo  (Dasypits  novemcinctus),  near 
Gainesville,  Florida.  Close  attendance  only  while  the 
armadillo  was  moving  and  disturbing  leaf  litter  sug- 
gests the  Ovenbird  was  taking  advantage  of  increased 
prey  availability  caused  by  the  armadillo's  flushing  of 
insects.  Received  27  May  1998,  accepted  / Oct.  1998. 


Many  species  of  birds  forage  in  association 
with  groups  of  other  species.  The  two  most 
common  explanations  of  such  interspecific 
groups  relate  to  decreased  risk  of  predation 
and  increased  foraging  efficiency  (Bertram 
1978).  Foraging  efficiency  can  be  increased 
via  several  mechanisms  (Morse  1970).  Per- 
haps the  least-well  studied  mechanism  occurs 
when  one  species  follows  another  species  and 
captures  prey  incidentally  flushed  by  the  sec- 
ond species.  Such  a relationship  between  “fol- 
lowers” and  “beaters”  has  been  reported  for 
groups  of  birds  and  groups  of  cattle  (Scot 
1984),  dolphins  (Evans  1987),  primates  (Ter- 
borgh  1983,  Boinski  and  Scott  1988),  wolves 
(Silveira  et  al.  1997),  ants  (Willis  and  Oniki 
1978),  and  other  birds  (Bennetts  and  Dreitz 
1997).  Here  I report  an  observation  of  an  Ov- 
enbird {Seiurus  aurocapillus)  following  a 
nine-banded  armadillo  (Dasypus  novemcinc- 
tus). 

On  6 March  1998  at  10:35  EST  along  the 
rim  of  Paynes  Prairie  (Alachua  County,  Flor- 
ida), I flushed  an  Ovenbird  from  the  ground. 
After  perching  for  1-2  min,  it  flew  to  the 
ground  within  3 m of  an  armadillo.  As  the 
armadillo  started  to  move  forward,  the  Ov- 
enbird flew  directly  to  it,  landing  approxi- 
mately 30  cm  from  its  tail  and  maintaining 
that  distance  of  separation  as  the  armadillo 
walked.  I followed  them  for  17  min  at  a dis- 
tance of  10-12  m.  The  armadillo  often 


' Dept,  of  Zoology,  PO  Box  1 18525,  Univ.  of  Flor- 
ida, Gainesville,  FL  3261  1-8525; 

E-mail:  DLEVEY(§)zoo. ufl.edu 


Stopped  briefly  (<  10  s)  to  dig  or  push  its 
snout  into  the  leaf  litter.  The  Ovenbird  did  not 
approach  the  armadillo’s  head  during  these 
times  but  rather  remained  by  its  tail.  On  two 
occasions  the  armadillo  stopped  for  1-2  min 
and  the  Ovenbird  walked  2-3  m away  from 
it.  When  the  armadillo  started  to  move  again, 
the  Ovenbird  immediately  resumed  following 
it,  once  returning  by  flight.  It  frequently 
pecked  at  the  leaf  litter.  I was  unable  to  de- 
termine if  these  presumed  foraging  attempts 
were  successful.  They  did  not  appear  more 
frequent  when  the  Ovenbird  was  following 
close  behind  the  armadillo  than  when  it  tem- 
porarily foraged  by  itself.  I did  not  notice  any 
insects  being  flushed  by  the  armadillo,  but 
there  was  heavy  shade  and  I lacked  binocu- 
lars. I stopped  my  observation  when  the  ar- 
madillo walked  into  a clearing  and  the  Ov- 
enbird did  not  follow. 

The  behavior  of  the  Ovenbird  suggests  its 
association  with  the  armadillo  was  not  due  to 
both  animals  being  attracted  to  an  area  of  high 
prey  abundance.  In  particular,  its  close  prox- 
imity to  the  armadillo  only  when  the  armadillo 
was  moving  suggests  it  was  using  the  arma- 
dillo as  a beater.  A similar  pattern  of  atten- 
dance has  been  noted  for  Double-toothed 
Kites  {Harpagus  didenlatus),  Gray-headed 
Tanagers  {Eucornetis  pencillata),  and  Tawny- 
winged  Woodcreepers  (Dendrocincla  anaba- 
tina)  following  Squirrel  Monkey  {Sainiiri  oer- 
stedi)  troops;  attendance  frequencies  of  these 
species  were  higher  when  the  Squirrel  Mon- 
key troops  were  moving  and  foraging  than 
when  they  were  immobile  (Boinski  and  Scott 
1988). 

Despite  high  levels  of  disturbance  created 
by  armadillos  while  foraging,  I am  unaware 
of  any  other  published  accounts  of  birds  fol- 
lowing them.  Further  observations  are  re- 
quired before  it  can  be  concluded  that  the  bird 
I watched  was  not  idiosyncratic  and  that  Ov- 
enbirds  benefit  from  following  armadillos. 


444 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


ACKNOWLEDGMENTS 

Thanks  to  E.  Bollinger,  S.  Duncan,  G.  Pryor,  K. 

Smith  and  an  anonymous  reviewer  for  constructive 

comments  on  the  manuscript. 

LITERATURE  CITED 

Bennetts,  R.  E.  and  V.  Dreitz.  1997.  Possible  use  of 
wading  birds  as  beaters  by  Snail  Kites,  Boat-tailed 
Crackles,  and  Limpkins.  Wilson  Bull.  109:169- 
173. 

Bertram,  B.  C.  R.  1978.  Living  in  groups:  predators 
and  prey.  Pp.  64-96  in  Behavioural  ecology  (J.  R. 
Krebs  and  N.  B.  Davies,  Eds.).  Sinauer  Press, 
Sunderland,  Massachusetts. 

Boinski,  S.  and  P.  E.  Scott.  1988.  Association  of  birds 
with  monkeys  in  Costa  Rica.  Biotropica  20:136- 
143. 


Evans,  D.  L.  1987.  Dolphins  as  beaters  for  gulls?  Bird 
Behav.  7:47-48. 

Morse,  D.  H.  1970.  Ecological  aspects  of  some  mixed- 
species  foraging  flocks  of  birds.  Ecol.  Monogr.  40: 
119-168. 

Scot,  D.  1984.  The  feeding  success  of  Cattle  Egrets 
in  flocks.  Anim.  Behav.  32:1089-1100. 

Silveira,  L.,  a.  T.  a.  Jacomo,  E H.  G.  Rodrigues, 
and  P.  G.  Crawshaw.  1997.  Hunting  association 
between  the  Aplomado  Falcon  (Falco  femoralis) 
and  the  Maned  Wolf  {Chry.socyon  bracbyurus)  in 
Emas  National  Park,  central  Brazil.  Condor  99: 
201-202. 

Terborgh,  J.  1983.  Five  New  World  primates.  Prince- 
ton Univ.  Press,  Princeton,  New  Jersey. 

Willis,  E.  O.  and  Y.  Oniki.  1978.  Birds  and  army  ants. 
Annu.  Rev.  Ecol.  Syst.  9:243-263. 


Wilson  Bull..  1 1 1(3),  1999,  pp.  445-456 


Ornithological  Literature 

Edited  by  William  E.  Davis,  Jr. 


THE  GREAT  BLUE  HERON:  A NATU- 
RAL HISTORY  AND  ECOLOGY  OF  A 
SEASHORE  SENTINEL.  By  Robert  W.  But- 
ler. UBC  Press,  Vancouver,  British  Columbia. 
1997:  167  pp.,  30  black-and-white  photos,  24 
color  plates,  17  numbered  text  figs.,  18  tables. 
$39.95  (Canadian)  (cloth). — This  is  a very 
nice  book  about  a very  interesting  bird — the 
Great  Blue  Heron  {Ardea  herodias) — with  a 
particular  emphasis  on  the  subspecies  that  fre- 
quents the  northwest  coast  of  North  America, 
A.  h.  fannini.  However,  frequent  reference  to 
herons  elsewhere  broadens  the  scope  and  per- 
spective of  the  book.  It  is  more  than  just  the 
study  of  a heron,  however,  it  is  the  story  of 
the  ecology  and  conservation  of  a region. 
Robert  Butler  is  certainly  well  qualified  to  tell 
this  story — 24  of  the  more  than  1 50  references 
cited  bear  his  name  as  an  author.  In  the  brief 
introductory  chapter  Butler  outlines  the  aims 
of  the  book  that  include  detailing  the  natural 
history  of  the  Great  Blue  Heron,  “a  worthy 
symbol  of  the  conservation  of  coastal  habi- 
tats,” along  the  27,000  kms  of  fragmented 
British  Columbia  coastline.  Chapters  that  fol- 
low consider  in  detail  the  habitat  of  the  heron 
(the  shores  of  temperate  rainforest),  the  her- 
on’s food  web,  the  sites  where  the  heron  has 
been  studied  year  around,  all  with  an  histori- 
cal perspective  on  interactions  with  man  wo- 
ven through  the  narrative.  There  are  chapters 
on  foraging,  food  and  diet,  social  and  territo- 
rial behavior,  colonial  nesting,  habitat  selec- 
tion, population  dynamics,  and  finally  a chap- 
ter on  the  conservation  of  Great  Blue  Herons 
and  the  Strait  of  Georgia  ecosystem.  An  epi- 
logue concludes  with  the  optimistic  thought 
that  many  people  are  beginning  to  recognize 
that  the  environment  provides  more  than  just 
resources,  and  that  restoring  environments 
will  yield  great  future  dividends.  Appendices 
report  records  of  Great  Blue  Heron  colonies, 
length-mass  regression  equations  for  fish,  ef- 
fects of  increased  disturbance  on  heron  pop- 
ulations, and  lists  of  scientific  names  of  plants 
and  animals. 

The  book  is  very  well  written.  It  is  clear 


that  the  author  thoroughly  enjoys  his  research 
on  herons  (including  night-vision  telescopic 
sleuthing),  and  his  descriptions  have  an  almost 
poetic  touch:  “1  look  back  on  the  long  hours 
spent  watching  herons  catch  fish  as  an  enjoy- 
able period  of  my  life.  Perched  in  the  shade 
on  a prominent  location  overlooking  spectac- 
ular scenery,  the  smell  of  the  sea  and  arbutus 
leaves  carried  on  a warm  breeze,  gulls  and 
shorebirds  busily  feeding  along  the  mudflat — 
it  was  hard  to  beat,”  or  “On  calm  nights  I 
often  slept  on  the  beach  beneath  the  stars.  I 
welcomed  the  silence  of  the  night  after  a day 
in  the  colony,  though  the  quiet  was  periodi- 
cally disturbed  by  landing  calls  ringing  from 
the  forest.”  Even  his  descriptions  of  natural 
history  phenomena  make  for  pleasant  reading: 
“The  delicate  choreographed  displays  of  a 
threatening  heron  are  exquisite.” 

I thoroughly  enjoyed  reading  this  book.  It 
contained  a great  deal  of  interesting  and  im- 
portant scientific  information  about  Great 
Blue  Herons,  and  a well  articulated  conser- 
vation perspective.  I recommend  it  to  anyone 
interested  in  avian  biology  or  conservation. — 
WILLIAM  E.  DAVIS,  JR. 


A GUIDE  TO  THE  IDENTIFICATION 
AND  NATURAL  HISTORY  OF  THE  SPAR- 
ROWS OF  THE  UNITED  STATES.  By 
James  D.  Rising.  Illustrated  by  David  D.  Bea- 
dle. Academic  Press,  New  York.  1996:  365 
pp.,  27  color  plates  with  captions,  53  color 
range  maps,  $42.00  (cloth). — Until  the  last 
three  years,  those  of  us  fascinated  by  details 
of  the  natural  history  of  sparrows  in  North 
America  had  only  Bent’s  life  histories  for  a 
reference.  With  the  publication  of  Sparrows 
and  Buntings — A Guide  to  the  Sparrows  and 
Buntings  of  North  American  and  the  World  in 
1995,  and  now  this  guide,  we  have  substan- 
tially more  information  of  a recent  vintage  at 
our  finger  tips.  This  is  a good  book.  Because 
of  its  less  ambitious  geographic  coverage. 


445 


446 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


there  is  more  information  and  greater  detail 
for  each  species  in  this  book. 

Sixty-two  species,  from  Olive  Sparrow 
(Embernagra  rufivirgata)  through  McKay’s 
Bunting  (Plectrophenax  hyperboreus)  are 
treated.  Each  species  account  includes  infor- 
mation about  identification,  similar  species, 
details  of  plumage  color,  voice,  habits,  habitat, 
breeding  behavior,  range,  history,  geographic 
variation,  measurements,  and  technical  refer- 
ences (for  further  reading).  A clearly  drawn, 
easily  interpreted,  color  range  map  shows 
breeding,  winter,  and  “all  year”  ranges  for 
each  species.  Including  state  and  provincial 
boundaries  on  the  range  maps  facilitates  their 
interpretation  substantially.  Where  sufficient 
data  exist,  an  additional  color  map,  showing 
relative  abundance,  based  on  Breeding  Bird 
Survey  information,  also  is  included  as  part  of 
the  species  account.  Though  somewhat  useful, 
the  shading  of  the  relative  abundance  maps  is 
such  that  it  is  difficult  to  distinguish  the 
shades  chosen  for  the  two  middle  classes  of 
the  four  relative  abundance  categories 
mapped.  In  addition,  the  two  middle  catego- 
ries of  5-20  and  20-50  individuals  appear  to 
overlap.  One  wonders  how  the  computer  map- 
ping software  dealt  with  abundances  of  20  in- 
dividuals and  how  one  really  should  interpret 
the  relative  abundance  maps. 

The  book  is  comprehensive  in  its  coverage. 
Among  the  62  “species”  descriptions  in  the 
book  are  a few  forms  not  yet  elevated  to  full 
species  status  by  the  AOU  Committee  on 
Classification  and  Nomenclature  in  its  1998 
check-list.  These  include  Bell’s  Sparrow  [con- 
sidered a “group”  of  Sage  Sparrow  (Amphis- 
pizcibelli)  by  AOUJ  and  Red,  Sooty,  and  Slate- 
colored  Fox  sparrows  [considered  “groups” 
of  Fox  Sparrow  (Passerella  iliaca)  by  AOUJ. 
There  also  is  substantial  treatment  of  a num- 
ber of  “Alaskan  rarities,”  including  Pine  {Em- 
berizci  leucocephaliis).  Little  (E.  pu.silla).  Rus- 
tic (E.  ru.stica).  Yellow-breasted  {E.  aureola). 
Gray  (E.  variablis),  Pallas’s  (E.  pallasi),  and 
Reed  buntings  {E.  schoeniclus).  Two  other 
rare  North  American  species.  Yellow-faced 
(Tiaris  oUvacea)  and  Black-faced  {T.  bicolor) 
grassquits,  reported  occasionally  from  the 
southern  U.S.,  also  are  described.  I also  ap- 
preciated the  concise  comments  and  descrip- 
tions of  subspecies  for  Savannah  Sparrow 
( Passe rculu.s  sandwichensis).  Song  Sparrow 


{Melospiza  melodia),  and  Dark-eyed  Junco 
(Junco  hyemalis),  including  tables  of  mea- 
surements illustrating  geographic  variation  in 
size.  Special  introductory  sections  for  most  of 
the  genera  described  in  the  book  provide  help- 
ful additional  information  about  intergeneric 
relationships  among  the  sparrows  and  identify 
areas  where  more  research  is  needed  to  clarify 
relationships  among  species  and  genera. 

The  illustrations  by  David  Beadle  add  sig- 
nificantly to  the  book’s  appeal.  In  addition  to 
excellent  color  plates,  each  species  account  is 
illustrated  by  a black-and-white  drawing  of 
the  species.  The  artist  has  done  an  excellent 
job  of  capturing  the  subtle  beauty  of  plumage 
colors  for  this  generally  somber-colored  as- 
semblage of  species. 

In  general  this  is  a good,  solid,  user-friendly 
reference  book  for  a challenging  group  of 
North  American  species.  It  is  small  enough  to 
be  carried  in  the  field,  though  not  necessarily 
a “field  guide”  in  the  traditional  sense.  I rec- 
ommend it  to  anyone  as  an  exceptionally  well 
illustrated,  very  readable  introduction  to  North 
American  sparrows. — CHARLES  R.  SMITH. 


NATURAL  HISTORY  OF  THE  WATER- 
FOWL.  By  Frank  S.  Todd.  Ibis  Publishing 
Co.,  Vista,  California.  1996:  490  pp.,  more 
than  750  color  photographs  with  captions,  164 
range  maps,  appendix.  $80.00  (cloth). — This 
visually  stunning  and  informative  book  effec- 
tively portrays  the  diversity  of  behavior,  ecol- 
ogy, and  plumage  of  the  more  than  160  spe- 
cies of  waterfowl.  Its  large  format  {XQVi  by 
13")  allows  the  author  to  display  to  best  ad- 
vantage the  many  fine  photos  that  he  has  taken 
in  pursuit  of  waterfowl  and  other  birds,  al- 
though even  the  heaviest  binoculars  will  seem 
like  featherweights  after  lifting  this  hefty 
tome!  The  book  begins  with  introductions  to 
natural  history  (Chapter  1)  and  taxonomy 
(Chapter  2),  then  proceeds  to  cover  primarily 
typical  waterfowl  groups  in  the  next  15  chap- 
ters (including  chapters  on  “Waterfowl  Odd- 
ities” and  “Whitewater  Ducks”).  The  chapter 
on  waterfowl  natural  history  includes  a nice 
overview  of  this  group  and  its  adaptations, 
and  many  photos  are  included  to  illustrate  typ- 
ical comfort  movements  and  postures.  The 
photo  legends  throughout  the  book  add  infor- 


ORNITHOLOGICAL  LITERATURE 


447 


mation  that  is  sometimes  not  present  in  the 
text,  including  general  information  on  birds. 
The  level  of  detail  varies  for  examples  given 
in  this  chapter,  creating  some  redundancy  with 
future  chapters,  and  many  of  the  examples  are 
from  work  on  geese  and  swans.  The  inclusion 
of  the  term  “gang  rape”  (p.  41)  is  inappro- 
priate and  unnecessary  since  the  preferable 
terminology  (forced  copulation)  is  also  used 
in  the  text,  and  references  to  “hyperactive” 
and  “hot-blooded”  males  in  photo  legends  de- 
picting forced  copulation  attempts  are  regret- 
table and  misleading.  The  taxonomy  chapter 
includes  a very  brief  coverage  of  past  and  cur- 
rent thoughts  on  waterfowl  taxonomy.  The  au- 
thor recognizes  the  need  for  further  corrobo- 
ration of  more  recent  taxonomic  treatments  of 
this  group  (some  of  which  have  not  been  sup- 
ported by  recent  findings),  and,  for  consisten- 
cy with  other  sources,  he  uses  more  traditional 
groups  for  subsequent  chapters  with  a few  ex- 
ceptions. 

Chapters  3-17  each  begin  with  an  overview 
of  the  similarities  within  the  group  being  cov- 
ered, including  movements  on  land  and  in  the 
air,  basic  calls,  feeding  habits,  nesting  and 
egg-laying,  flocking,  mating  and  family  be- 
havior, and  migration.  Individual  species  ac- 
counts follow,  including  descriptions  of  ex- 
tinct species  and  separate  accounts  for  each 
race.  The  description  of  each  species  is  ac- 
companied by  a small  color  range  map  (on 
which  race  distributions  are  not  delineated), 
and  usually  several  excellent  photos  of  the 
species  in  the  wild.  Valuable  information  on 
conservation,  captive  propagation,  and  human 
uses/conflicts  is  presented  in  addition  to  more 
detailed  information  on  topics  outlined  in  the 
chapter  overview.  Incubation  period,  time  to 
fledging,  conditions  under  which  dump  nest- 
ing data  were  collected,  and  nesting  density 
are  presented  inconsistently  within  the  species 
accounts,  although  further  details  on  weight, 
egg  and  clutch  size,  incubation  period,  and 
fledging  period  are  presented  for  each  species 
(and  race  if  appropriate)  in  the  Appendix.  In- 
sights from  the  close  observation  of  captives 
complement  information  from  extensive  field 
experience  in  many  species  accounts. 
Throughout  these  accounts,  the  author  has  in- 
serted personal  experiences  and  origins  of 
some  scientific  names  that  add  extra  interest 
for  the  reader.  I especially  found  interesting 


the  accounts  of  endangered  species  and  races 
within  various  waterfowl  groups. 

The  accounts  of  the  northern  geese  (Chap- 
ter 5),  swans  (Chapter  6),  and  eiders  (Chapter 
14)  are  some  of  the  most  complete  of  the  vol- 
ume, demonstrating  the  author’s  experience 
with  these  groups  and  also  information  avail- 
able from  the  scientific  studies  of  The  Wild- 
fowl and  Wetlands  Trust  in  England.  These 
chapters  are  full  of  photos,  including  all  of  the 
Canada  Goose  {Branta  canadensis)  races,  and 
details  on  the  Nene  (Hawaiian  Goose;  Branta 
sandvicensis)  reintroduction  project  and  suc- 
cessful Barnacle  Goose  {Branta  leucopsis) 
conservation.  Nesting  densities  for  the  eiders 
are  well-documented,  as  is  the  solving  of  the 
mystery  of  the  location  of  Spectacled  Eider 
{Somateria  fischeri)  wintering  areas.  Atypical 
waterfowl,  such  as  the  Magpie  Goose  {Anser- 
anas  semipalmata\  Chapter  3);  Cape  Barren 
Goose  {Cereopsis  novaehollandiae).  Freckled 
Duck  (Stictonetta  naevosa).  Spur-winged 
Goose  (Plectropterus  gambensis).  Comb 
Duck  {Sarkidiornis  melanotos),  and  Pink- 
eared Duck  {Malacorhynchus  membranaceus) 
(grouped  as  “waterfowl  oddities”  in  Chapter 
7);  and  screamers  (Chapter  17)  are  covered 
quite  extensively  in  accounts  that  highlight  the 
unusual  traits  of  these  species  and  their  con- 
tinued puzzling  taxonomy.  The  author  groups 
together  and  covers  well  the  white-water 
ducks  (Blue  Duck,  Hyrnenolairnus  malacor- 
hynchos'.  Torrent  Duck,  Merganetta  armata; 
and  Salvadori’s  Duck,  Salvadorina  waigiuen- 
sis)  in  Chapter  10,  describing  adaptations  to 
this  demanding  environment  such  as  their  sed- 
entary lifestyle,  territorial  behavior,  and  diet 
of  benthic  invertebrates. 

The  sheldgeese  and  shelducks,  sharing 
strong  pairbonds  and  an  aggressive  disposi- 
tion, are  described  in  Chapter  8,  including  es- 
pecially good  information  on  the  ongoing  con- 
flicts with  humans  when  sheldgeese  use  crop 
lands.  Presented  in  Chapter  9 is  another  very 
pugnacious  group  of  waterfowl  that  includes 
three  flightless  species,  the  steamerducks.  The 
chapters  on  the  pochards  (Chapter  13),  sea- 
ducks  (Chapter  15),  and  stiff-tailed  ducks 
(Chapter  16)  present  many  useful  photos  and 
information  on  feeding  and  diving  adapta- 
tions. Also  included  in  these  chapters  are  fas- 
cinating accounts  of  the  (probably)  extinct 
Pink-headed  Duck  {Rhodonessa  caryophylla- 


448 


THE  WILSON  BULLETIN  • Vol.  HI.  No.  3,  September  1999 


cea),  the  extinct  Labrador  Duck  {Camptorhyn- 
chus  labradorius)  and  Auckland  Islands  Mer- 
ganser (Mergus  australis),  and  the  only  wa- 
terfowl obligate  brood  parasite,  the  extant 
Black-headed  Duck  (Heteronetta  atricapilla). 
Accounts  of  the  whistling  ducks  (Chapter  4), 
including  the  White-backed  Duck  (Thalassor- 
nis  leuconotus),  are  informative  but  generally 
short  because  of  a lack  of  detailed  information 
for  many  species.  This  general  lack  of  infor- 
mation available  for  southern  hemisphere 
ducks  is  especially  evident  in  short  accounts 
for  these  species  in  Chapters  1 1 (perching 
ducks)  and  12  (dabbling  ducks).  Chapter  11 
covers  many  of  the  traditional  perching  duck 
group  members  while  recognizing  that  the 
taxonomic  organization  of  these  ducks  is  still 
changing.  The  account  of  a well-studied  spe- 
cies, the  (North  American)  Wood  Duck  {Aix 
sponsa),  was  disappointing  in  its  omission  of 
available  information  on  dump-nesting  and 
other  aspects  of  its  natural  history.  In  general, 
members  of  the  large  dabbling  duck  group  re- 
ceive more  complete  coverage,  including 
more  personal  observations  and  photos  by  the 
author,  although  I found  some  aspects  of  the 
Andean  Teal  (Anas  flavirostris  andiurn)  and 
Puna  Teal  (Anas  puna)  accounts  at  odds  with 
my  own  observations. 

The  Epilogue  presents  a balanced  treatment 
of  threats  to  waterfowl  populations  and  ben- 
efits greatly  from  the  author’s  own  experience 
with  the  continued  conservation  challenges 
that  face  this  group.  Overall,  the  book  is  writ- 
ten in  a style  that  is  easy  to  read;  a glossary 
is  included  to  aid  the  lay  reader  and  the  Index 
at  the  end  of  the  book  facilitates  finding  in- 
formation on  particular  species.  Scientific 
names  are  not  presented  in  the  book  except 
for  those  of  waterfowl  species.  There  are  few 
citations  in  the  text  and  the  bibliography  in- 
cludes mostly  books,  limiting  the  use  of  this 
volume  as  a scientific  or  research  reference. 
Despite  a few  reoccurring  grammatical  prob- 
lems, the  writing  style  conveys  well  the  thrill 
of  viewing  waterfowl  and  the  author’s  enthu- 
siasm about  his  experiences  with  this  group  of 
birds.  This  book  is  a treat  to  the  eyes  for  any- 
one that  appreciates  birds,  and  in  addition,  is 
at  the  least  a good  overview  of  waterfowl  nat- 
ural history.  Highly  recommended  as  a visual, 
general  reference  book  for  anyone  interested 
in  waterfowl. — GWENDA  L.  BREWER. 


FAIRY-WRENS  AND  GRASSWRENS 
MALURIDAE.  By  Ian  Rowley  and  Eleanor 
Russell,  illus.  by  Peter  Marsack.  Oxford  Uni- 
versity Press,  Oxford.  1997:  274  pp.,  9 color 
plates,  numerous  maps,  tables  and  black-and- 
white  line  drawings.  $75  (cloth). — This  fas- 
cinating family  of  birds  is  found  only  in  Aus- 
tralia and  New  Guinea  and  consists  of  five 
genera:  emu-wrens  (Stipiturus)  and  grass- 
wrens  (Amytornis)  found  only  in  Australia, 
tree- wrens  (Sipodotus)  and  russet- wrens  (C/v- 
tomyias)  of  New  Guinea,  and  the  largest  ge- 
nus, fairy-wrens  (Malurus)  found  in  both. 
Several  species  of  fairy-wrens  have  been  in- 
tensively studied  using  color-banded  popula- 
tions and  biochemical  analyses  and  the  results, 
particularly  those  relating  to  breeding  biology, 
are  intriguing  and  make  for  fascinating  read- 
ing. 

This  fourth  volume  in  Oxford  University 
Press’  series  on  bird  families  of  the  world  pro- 
vides a thorough  review  of  this  interesting 
family  of  birds  by  authors  who  have  done 
much  of  the  primary  research  on  several  of 
the  species  considered.  The  monograph  is  di- 
vided into  two  parts:  the  first  consists  of  eight 
chapters  dealing  with  various  aspects  of  the 
biology,  behavior,  ecology,  evolution,  and 
conservation  of  the  Maluridae;  and  the  second 
consists  of  accounts  of  the  5 genera  and  25 
species  that  constitute  the  family.  Chapter  2 
discusses  two  centuries  of  the  rather  confused 
taxonomy  of  the  Maluridae,  including  more 
recent  biochemical  studies.  Chapter  3 deals 
with  the  environment,  biogeography,  and  evo- 
lution, including  plate  tectonics  and  Gond- 
wana  breakup,  past  and  present  climates  and 
vegetation,  refugia  and  speciation  during  the 
past  two  million  years,  changes  since  human 
settlement,  and  the  evolution  of  the  Maluridae 
including  the  five  main  lineages  of  fairy- 
wrens  and  the  grass  wrens.  Chapter  4 deals 
with  morphology,  locomotion,  and  feeding  be- 
havior of  these  largely  insectivorous,  ground, 
and  shrub-dwelling  birds.  Chapter  5 discusses 
vocal  communication  and  social  organization, 
and  includes  a number  of  sonagrams,  and  a 
thorough  analysis  of  courtship  displays.  The 
chapter  also  details  the  remarkable  findings 
from  electrophoresis  and  DNA  fingerprinting 
studies  focused  on  reproductive  biology  in 
fairy-wrens.  Although  monogamous  and  mat- 
ed for  life,  extra-pair  copulations  outside  of 


ORNITHOLOGICAL  LITERATURE 


449 


the  territorial  family  group  in  one  study,  ac- 
counted for  more  than  three-quarters  of  the 
young!  Chapter  6 is  devoted  to  co-operative 
breeding  and  an  analysis  of  helpers  at  the  nest 
(mostly  surviving  young  from  earlier  years). 
Chapter  7 contains  a generalized  life  history 
study  of  the  Maluridae,  based  mostly  on  long- 
term studies  of  fairy-wrens,  and  includes  sec- 
tions on  nests  and  nest  building,  eggs,  clutch 
size,  number  of  broods,  reproductive  success, 
parasitism  and  predation,  dispersal,  and  sur- 
vival rates.  Chapter  8,  on  conservation,  traces 
the  clearance  and  fragmentation  of  vegetation 
for  agriculture,  forestry,  and  grazing,  and  ac- 
companying habitat  degradation,  introduction 
of  alien  plants,  draining  of  wetlands,  and 
changes  introduced  into  fire  regimes  by  Eu- 
ropeans. The  author’s  analysis  suggests  that  5 
species  and  5 additional  subspecies  are  cur- 
rently threatened. 

In  Part  II,  brief  accounts  of  each  genus  are 
followed  by  detailed  accounts  of  each  species. 
These  accounts  begin  with  descriptions  of 
adult  males  and  females  in  breeding  and 
eclipsed  plumages,  of  immatures  and  moult 
sequences,  and  typically  continue  with  history 
of  taxonomy,  weights  and  measurements,  field 
characters,  voice,  range  and  status,  habitat, 
displays,  breeding  behavior,  and  life  cycle. 
Each  account  is  accompanied  by  a range  map 
and  sonagram(s).  Additional  brief  accounts  of 
groupings  of  species  within  genera  are  present 
where  needed,  e.g.,  a section  on  the  four 
chestnut-shouldered  fairy-wren  species.  The 
eight  color  plates  are  excellent.  They  include, 
where  appropriate,  depictions  of  adult  males 
and  females,  immatures,  eclipsed  males,  and 
subspecies.  The  color  of  the  plates  is  excel- 
lent, although  the  breast  of  the  Blue-breasted 
Fairy- Wren  (Maluras  piilcherrimus)  appears 
bluer  than  I remember  it  in  the  field.  A color 
figure,  grouped  with  the  plates,  contains  six 
photographs  of  typical  Australian  habitats.  A 
glossary  helps  with  terms  like  “samphire,” 
“spinifex,”  or  “billabong”  that  might  not  be 
familiar  to  everyone.  The  bibliography  in- 
cludes more  than  500  references. 

This  is  a well-written,  thoroughly  re- 
searched, monograph.  I looked  through  the 
Acknowledgments  for  people  who  I know 
have  done  work  on  malurids — they  were  all 
there.  This  is  a comprehensive  book,  easy  to 
read,  and  loaded  with  interesting  information 


about  a fascinating  and  lovely  family  of  birds. 
Anyone  working  on  Australasian  birds  or  with 
interest  in  avian  breeding  biology  should  have 
this  book.— WILLIAM  E.  DAVIS,  JR. 


AVIAN  CONSERVATION.  By  John  M. 
Marziuff  and  Rex  Sallabanks  (Eds.).  Island 
Press,  Washington,  D.C.  1998;  563pp.  (no 
price  given) — Avian  Conservation  is  a collec- 
tion of  chapters  of  which  about  one  half  were 
presented  in  a symposium  of  the  American 
Ornithologist’s  Union  and  Raptor  Research 
Foundation  meeting  in  1996.  The  remaining 
chapters  were  solicited  by  the  editors  to  fill 
gaps  and  to  provide  a land  manager’s  per- 
spective on  relevant  research  for  avian  con- 
servation. The  book  is  divided  into  7 parts  and 
31  chapters.  Part  1 introduces  the  reader  to  the 
past  and  present  approaches  taken  in  conser- 
vation. Part  2 reviews  a variety  of  techniques 
applied  in  conservation  research  including  ge- 
netics, spatial  modeling,  indicator  species  con- 
cept, and  monitoring  landbirds.  Part  3 pro- 
vides examples  of  approaches  used  to  con- 
serve endangered  and  sensitive  species.  Part  4 
deals  with  conservation  of  forested  landscapes 
and  Part  5 covers  non-forested  and  urban 
landscapes.  Part  6 examines  conservation  of 
birds  outside  North  America  and  includes  the 
European  agricultural  environment,  research 
needs  and  applications  for  Neotropical  birds, 
and  conservation  in  Israel,  Russia,  the  Mari- 
ana Islands  and  Australia.  Part  7 is  a discus- 
sion on  making  conservation  research  relevant 
to  land  managers.  The  aims  of  the  authors  are 
to  review  current  research  and  identify  infor- 
mation. 

Marziuff  and  Sallabanks  should  be  com- 
mended for  assembling  concise  reviews  of 
many  important  topics  for  conservationists 
and  applications  for  land  managers.  Anyone 
interested  in  the  latest  information  and  status 
of  projects  aimed  at  conserving  the  Northern 
Spotted  Owl  (Strix  occidentalis).  Northern 
Goshawk  (Accipter  gentiiis),  Hawaiian  Goose 
(Anser  sandvicensis),  Red-cockaded  Wood- 
pecker {Picoides  borialis)  and  others  will  find 
this  book  very  useful.  Also  anyone  wishing  to 
apply  techniques  such  as  GIS  modeling  of 
populations,  understanding  threats  to  seabirds, 
invasions  by  exotics  or  affects  of  urban  en- 


450 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


vironments  on  birds  should  read  this  book.  Is- 
land Press  has  published  an  attractive  book  in 
an  easy-to-read  format. 

This  book  contains  a wealth  of  information 
and  is  an  excellent  review  of  the  topic  al- 
though it  is  biased  in  favor  of  North  American 
issues — 6 of  the  31  chapters  were  devoted  to 
areas  of  the  world  outside  the  USA,  2 of  those 
were  written  by  Americans,  and  50  of  the  58 
contributors  were  from  American  institutions. 
Consequently,  the  conclusions  were  strongly 
biased  towards  problems  perceived  by  Amer- 
icans. The  topic  of  bird  conservation  is  im- 
mense and  the  book  would  have  been  more 
successful  if  it  had  dealt  only  with  birds  that 
use  the  USA.  A minor  annoyance  is  the  as- 
sumption by  many  American  authors  that  all 
readers  know  that  place  names  such  as  Pacific 
Northwest,  Midwestern  States,  the  Snake  Riv- 
er, and  issues  such  as  the  enactment  of  the 
Endangered  Species  Act  are  in  the  USA  when 
they  are  writing  for  a world-wide  audience. 
Once  again,  if  the  book  had  been  focused  on 
birds  that  use  the  USA,  these  terms  would 
have  been  appropriate.  A more  important 
oversight  is  the  small  amount  of  attention  de- 
voted to  existing  conservation  programs  in  the 
USA  and  abroad.  For  example,  the  North 
American  Waterfowl  Management  Plan 
(NAWMP)  is  the  largest  and  most  ambitious 
avian  conservation  program  undertaken  in 
North  America.  It  has  overcome  political  bar- 
riers, raised  billions  of  dollars  and  set  aside 
1000s  of  hectares  of  wetlands  in  Canada,  the 
USA,  and  Mexico.  The  success  of  NAWMP 
has  prompted  other  programs  such  as  the 
Western  Hemisphere  Shorebird  Reserve  Net- 
work (WHSRN),  Partners  in  Flight,  and  the 
Seaduck  Joint  Venture.  There  is  no  or  very 
little  mention  of  these  and  many  other  con- 
servation programs  in  Avian  Conservation. 
Both  NAWMP  and  WHSRN  have  been  in 
place  for  many  years,  with  well  established 
newsletters  and  web  sites  but  only  one  paper 
addressed  concerns  about  waterfowl  and  it  fo- 
cused on  the  Hawaiian  Goose;  there  were  no 
papers  on  shorebirds.  A few  chapters  on  the 
status  and  trends  of  these  and  other  birds 
would  have  strengthened  the  book  consider- 
ably. There  was  no  mention  of  Birdlife  Inter- 
national’s Important  Bird  Areas  program  in 
Europe,  the  Middle  East,  Canada,  and  Mexi- 
co, and  a similar  program  by  National  Au- 


dubon Society  and  the  American  Bird  Con- 
servancy in  the  US  was  also  overlooked.  A 
review  of  the  many  approaches,  their  success- 
es and  failures,  and  the  research  questions 
they  require  would  have  greatly  strengthened 
Avian  Conservation.  These  oversights  suggest 
that  a wide  gap  remains  between  the  two  sol- 
itudes of  research  and  conservation  manage- 
ment, at  least  in  the  USA.  The  five  chapters 
on  land  management  were  written  by  Ameri- 
cans and  for  Americans.  The  superb  reviews 
in  Avian  Conservation  will  appeal  to  conser- 
vationists world-wide  and  should  be  on  their 
shelves  as  an  up-to-date  summary  of  field  and 
a reference  source.  However,  its  applicability 
is  limited  largely  to  a North  American  audi- 
ence.—ROBERT  W.  BUTLER. 


RUDDY  DUCKS  AND  OTHER  STIFF- 
TAILS:  THEIR  BEHAVIOR  AND  BIOLO- 
GY. By  Paul  A.  Johnsgard  and  Montserrat 
Carbonell.  University  of  Oklahoma  Press, 
Norman,  Oklahoma.  1996:  291  pp.,  16  color 
photos  with  captions,  33  numbered  text  fig- 
ures including  line  drawings  and  range  maps, 
19  tables,  13  black-and-white  illustrations. 
$49.95  (cloth). — The  collaboration  of  these 
two  authors  brings  together  a wealth  of  ex- 
perience with  wild  and  captive  stifftails,  and 
with  studies  of  waterfowl  natural  history  and 
behavior  in  general.  The  result  is  a fine  book 
on  a fascinating  group  of  ducks  that  includes 
excellent  illustrations  and  much  detailed  in- 
formation. Although  there  are  few  color  plates 
(photos  of  both  sexes  of  all  species  covered 
would  be  helpful),  there  are  nice  illustrations 
of  each  species  and  excellent  drawings  of  dis- 
play behavior  by  the  first  author,  many  of 
which  are  tracings  from  films  or  photos. 

The  first  section  of  the  book  covers  general 
characteristics  of  the  stifftails,  effectively  in- 
troducing a number  of  interesting  features  of 
this  group.  The  summary  chapters  that  follow 
make  especially  good  use  of  the  second  au- 
thor’s thesis  work  on  a variety  of  captive  stiff- 
tails  at  The  Wildfowl  and  Wetlands  Trust  in 
England.  Covered  as  the  stifftail  group  are 
members  of  the  genera  Heteronetta,  Nomonyx, 
Oxyura,  and  Bizinra.  Although  the  White- 
backed  Duck  (Thassalornis  leuconotus)  and 
Freckled  Duck  {Stictonetta  naevosa)  are  no 


ORNI  FHOLOGICAL  LITERATURE 


451 


longer  considered  to  be  stifftails,  some  com- 
ments are  still  included  where  relevant  (es- 
pecially display  behavior).  Chapter  1 begins 
with  a detailed  historical  treatment  of  the  tax- 
onomy of  the  stifftails,  and  concludes  with  a 
synthesis  of  available  information  and  a dis- 
cussion of  remaining  questions  (some  of 
which  have  recently  been  tackled  through  ge- 
netic analyses).  Chapter  2 presents  a detailed 
summary  of  stifftail  morphology  and  anato- 
my, including  a discussion  of  the  adaptations 
for  diving  that  these  species  possess  and  com- 
parisons between  them.  Molts  and  plumages 
are  also  treated  in  this  chapter,  but  in  a very 
general  way,  with  detailed  accounts  appearing 
in  species  chapters  when  possible.  I believe 
that  it  would  have  been  helpful  to  include  a 
clear  summary  of  at  least  the  most  common 
molt  patterns  of  stifftails  here  as  an  overview. 
Chapter  3,  General  Behavior  and  Ecology,  at- 
tempts to  summarize  the  postures  and  loco- 
motion, comfort  movements,  time  budgets, 
feeding  behavior,  habitats,  dispersal,  migra- 
tion, and  important  interspecific  interactions 
of  the  stifftails.  Accounts  of  comfort  move- 
ments are  unusually  complete  due  to  a com- 
bination of  information  obtained  by  the  au- 
thors and  previous  work  by  Frank  McKinney. 
Time  budget  data  and  dive  durations  are  pri- 
marily available  only  from  captives,  and  diet 
data  are  presented  from  previous  studies  on 
wild  birds.  Especially  rare  for  waterfowl  are 
the  data  on  dive  durations  and  activity  budgets 
for  ducklings. 

Sex  ratios,  pairbonds,  ritualized  display  be- 
havior, aggression  and  territoriality,  and  con- 
tributions of  display  behavior  to  taxonomic  re- 
lationships are  presented  in  Chapter  4 (Com- 
parative Social  and  Sexual  Behavior).  This 
chapter  also  is  based  on  a mixture  of  data 
from  captive  and  wild  birds.  Unfortunately, 
some  of  the  sample  sizes  from  captives  are 
small,  and  there  is  a general  lack  of  data  on 
number  of  individuals  and  variability,  making 
it  difficult  for  the  reader  to  interpret  the  level 
of  support  for  the  authors’  statements.  Also, 
the  display  information  focuses  almost  com- 
pletely on  male  displays  in  this  chapter.  A ta- 
ble of  male  stifftail  structures  and  displays 
makes  for  an  easy  comparison  of  behavioral 
similarities,  and  hints  at  some  of  the  recent 
findings  on  the  taxonomy  of  this  group  using 
genetic  characters.  Reproductive  and  popula- 


tion biology  are  summarized  in  Chapter  5,  in- 
cluding a plea  for  more  studies  on  species  oth- 
er than  the  well-known  North  American  Rud- 
dy Duck  {Oxyura  jamaicensis).  To  underscore 
this  point,  the  timing  of  pairbonding,  breed- 
ing, and  nesting,  and  hatching  success,  brood 
behavior,  renesting,  and  annual  recruitment  in 
this  chapter  are  only  available  in  any  detail 
for  the  North  American  Ruddy  Duck.  Much 
of  this  information  was  drawn  from  the  same 
three  studies.  Information  from  more  species, 
including  data  from  captives,  is  summarized 
for  nest  site  characteristics,  eggs  and  laying 
behavior,  clutch  sizes,  and  duckling  weights, 
often  in  tables  that  allow  comparisons  to  be 
made  between  the  stifftails.  Nest  parasitism  is 
discussed  in  some  detail,  including  an  indi- 
cation of  reproductive  success  relative  to  host 
nests  for  North  American  Ruddy  Duck  and 
Black-headed  Duck  (Heteronetta  atricapilla), 
an  obligate  nest  parasite. 

The  second  section  of  the  book.  Chapters 
6-13,  presents  species  accounts  that  include 
vernacular  names,  range  of  species  and  races, 
measurements  and  plumage  descriptions, 
identification  cues,  ecology,  annual  cycle,  so- 
cial and  sexual  behavior,  nesting  and  parental 
behavior,  and  reproductive  success  and  status. 
Range  maps  have  been  updated  and  improved 
from  Johnsgard’s  Ducks,  Geese,  and  Swans  of 
the  World  (1978).  In  general,  the  species  ac- 
counts present  a good  degree  of  specific  data, 
and  sample  sizes  are  given  for  measurements 
and  some  other  data.  Plumage  and  soft  tissue 
descriptions  are  supplemented  by  references 
to  the  location  of  photographs  or  illustrations 
in  the  literature,  and  for  each  species  there  are 
detailed  drawings  or  tracings  from  film  of 
courtship  display  postures  or  sequences  in  ad- 
dition to  written  accounts.  Although  phonetic 
descriptions  of  calls  are  given,  sonagrams 
would  have  been  useful.  Identification  of  birds 
in  the  hand  and  in  the  field  is  noted,  and  an 
Appendix  features  a dichotomous  key  to  in- 
hand identification  with  drawings  of  adults 
and  ducklings. 

The  section  on  ecology  for  each  species  in- 
cludes habitat,  density,  foods,  foraging,  and  a 
short  coverage  of  competitors,  predators,  and 
symbionts.  Data  on  densities,  foods  con- 
sumed, and  predators  are  unavailable  for 
many  species,  and  the  discussions  of  compet- 
itors are  mostly  speculations  about  diet  over- 


452 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  3,  September  1999 


laps  with  various  waterbirds  and  other  stiff- 
tails.  Symbionts  apparently  refers  to  parasitic 
egg-laying  interactions,  and  these  data  would 
appear  to  have  been  more  appropriate  in  the 
section  on  nest  choice.  Movements  and  mi- 
grations, molts  and  plumages,  and  the  breed- 
ing cycle  are  discussed  as  parts  of  the  annual 
cycle,  with  limited  information  on  wild  birds 
for  the  lesser-known  species.  Descriptions  of 
the  mating  system,  territoriality,  courtship  and 
pairbonding,  and  copulatory  behavior  are  in- 
cluded under  social  and  sexual  behavior. 
Again,  information  is  limited  for  wild  birds 
for  a number  of  species,  and  captive  studies 
have  provided  the  majority  of  the  detailed  de- 
scriptions of  courtship  displays  and  copula- 
tion. The  use  of  “rape  behavior”  is  outdated 
and  even  in  the  book’s  Glossary  the  reader  is 
instructed  to  “See  forced  copulation”.  The 
nesting  and  parental  behavior  subsection  in- 
cludes nest  choice  and  egg  laying,  and  hatch- 
ing and  brood-related  behavior.  A mixture  of 
data  from  wild  and  captive  birds  expands  the 
coverage  of  these  topics  appreciably,  although 
in  general,  little  information  on  duckling  be- 
havior is  known  for  any  of  the  ducks.  Aspects 
of  reproductive  success  and  status  for  each 
species  were  summarized  in  tables  in  the  first 
section  of  the  book,  so  in  the  species  accounts, 
the  focus  is  primarily  on  population  estimates 
and  conservation  challenges  (although  cover- 
age is  a little  disappointing  on  this  topic). 
Brought  to  light  here  is  the  especially  alarm- 
ing predicament  of  the  White-headed  Duck 
iOxyura  leucocephala),  which  is  threatened 
by  overhunting,  habitat  destruction,  and  hy- 
bridization and  interactions  with  the  intro- 
duced North  American  Ruddy  Duck.  The  sta- 
tus of  most  of  the  .southern  hemisphere  species 
has  been  difficult  to  determine  because  of  a 
lack  of  focused  studies,  but  as  human  activi- 
ties continue  to  increase  pressure  on  wetland 
habitats,  we  cannot  afford  our  ignorance  if 
these  species,  and  many  others,  are  to  survive. 

The  book  is  peppered  with  a number  of  ty- 
pographical errors,  but  it  is  quite  readable  and 
a glossary  is  included  to  aid  the  lay  reader.  In 
general,  the  references  throughout  both  sec- 
tions of  the  book  do  not  seem  to  be  as  up-to- 
date  as  they  could  have  been,  probably  as  a 
result  of  delays  between  preparation  and  pub- 
lishing. Some  unpublished  material  supple- 
ments the  text,  and  the  references  section  in- 


cludes sources  not  cited  in  the  text.  This  book 
goes  a long  way  towards  identifying  where 
our  gaps  in  knowledge  lie  and  what  future 
studies  are  needed  on  this  interesting  group  of 
birds.  1 recommend  it  as  a valuable  and  quite 
easy  to  use  resource  for  researchers  and  stu- 
dents of  waterfowl  behavior  and  ecology  or 
others  particularly  interested  in  the  stiff- 
tails.— GWENDA  L.  BREWER. 


THE  NUTHATCHES.  By  Erik  Matthysen, 
illus.  by  David  Quinn.  T & A D Poyser,  Lon- 
don, U.K.  U.S.  edition  published  by  Academic 
Press,  San  Diego,  CA.  1998:  xx-l-315  pp.,  one 
color  plate,  many  black  & white  drawings,  17 
black  & white  photos,  103  figs.,  29  tables 
$39.95. — Books  devoted  to  single  bird  fami- 
lies are  in  vogue  nowadays  and  several  series 
are  being  published.  The  series  coming  from 
the  Poyser  company  of  England  differs  from 
most  in  that  the  books  do  not  attempt  to  be  a 
field  guide  and  they  have  no  extensive  color 
plates.  Instead  most  of  them  are  thorough 
studies  of  the  natural  history  of  the  species 
involved. 

The  nuthatch  volume  is  of  that  nature. 
While  treating  all  24  species  of  the  genus  Sit- 
ta,  approximately  60%  of  the  space  is  devoted 
to  the  Eurasian  Nuthatch  {S.  europaea).  The 
author  has  spent  a major  portion  of  his  pro- 
fessional career  studying  this  species  starting 
with  an  undergraduate  thesis  in  1982.  The 
species  is  treated  in  9 chapters:  Taxonomy, 
Morphology  and  Moult;  Habitat  and  Popula- 
tion Density;  Foraging,  Food  and  Hoarding; 
The  Pair  and  Its  Territory;  Breeding  Biology; 
Finding  a Territory;  Dispersal  and  Migration; 
Population  Dynamics;  and  Nuthatches  in  For- 
est Fragments.  Each  chapter  bristles  with  data, 
often  of  a sort  not  usually  found  in  avian  life 
history  studies.  Tables  and  graphs  abound.  Be- 
sides his  own  data  from  Belgium,  the  author 
gathered  data  from  the  literature  from 
throughout  the  range  of  the  species.'  The  result 
is  an  exemplary  life  history  account. 

The  other  23  species  are  treated  in  four 
chapters.  In  so  far  as  possible  the  species  are 
treated  by  the  same  topics  as  above,  but  un- 
derstandably very  little  is  known  about  some 
aspects  of  the  biology  of  some  species. 

The  Mediterranean  Nuthatches  are  three 


ORNITHOLOGICAL  LITERATURE 


453 


Red-breasted  Nuthatch  (5.  canadensis)  look- 
alikes:  Algerian  (5.  ledanti),  Corsican  (S.  whi- 
teheadi),  and  Krueper’s  (5.  krueperi)  nut- 
hatches. All  these  have  very  limited  distribu- 
tion and  the  Algerian  Nuthatch  was  discov- 
ered as  late  as  1975.  The  story  of  the 
discovery  of  this  species  is  told  in  detail.  The 
original  population  estimate  was  12  pairs  in  a 
very  restricted  area  but  more  recently  the  es- 
timate is  somewhere  between  500  and  a few 
thousand  pairs  as  the  result  of  the  discovery 
of  additional  populations.  All  three  of  these 
species  are  susceptible  to  threats  due  to  loss 
of  habitat. 

The  two  Rock  Nuthatches  (5.  tephronota 
and  S.  neumayer)  have  abandoned  the  forest 
habitat  of  the  rest  of  the  family  and  are  found 
on  rocky  slopes  and  cliffs  in  dry  regions  from 
Yugoslavia  to  Pakistan.  The  two  overlap  in 
Iran  and  eastern  Turkey,  and  it  was  not  real- 
ized until  1911  that  they  were  two  species. 

Fourteen  species  are  covered  under  the  col- 
lective heading.  Oriental  Nuthatches.  These 
range  from  the  Himalayas  to  the  Philippines, 
and  as  many  as  7 species  can  be  found  to- 
gether in  some  parts  of  southeastern  Asia.  As 
might  be  expected  many  of  these  are  not  well 
known  and  the  treatment  is  less  detailed  than 
for  other  species.  Included  in  this  group  are 
the  195  mm  Giant  Nuthatch  (5.  magna),  the 
Blue  Nuthatch  (5.  azurea),  and  the  Beautiful 
Nuthatch  (S.  formosa)  both  of  which  depart 
from  the  drab  coloration  of  most  of  the  family, 
as  well  as  two  more  Red-breasted  look-alikes. 

The  final  chapter  discusses  the  4 New 
World  Nuthatches:  White-breasted  {S.  caroli- 
nensis).  Red-breasted  (5.  canadensis),  Brown- 
headed (S.  pusilla),  and  Pygmy  (5.  pygmaea). 
Despite  the  abundance  of  these  species  none 
of  them  has  been  studied  to  the  extent  that  S. 
europaea  has.  The  Red-breasted  is  the  only 
migrating  nuthatch.  The  “Dwarf”  nuthatches, 
S.  pusilla  and  S.  pygmaea,  are  unique  in  the 
family  in  having  small,  often  overlapping,  ter- 
ritories, communal  roosts,  and  extra  male 
helpers  at  the  nest. 

The  evolutionary  history  of  the  S.  canaden- 
sis superspecies  which  consists  of  one  species 
in  northern  North  America,  three  species  in 
the  Mediterranean  region,  and  two  geograph- 
ically separated  species  in  China  presents  an 
interesting  puzzle. 

David  Quinn’s  black-and-white  sketches. 


many  illustrating  behaviors,  enhance  the  book 
and  the  single  color  plate  illustrates  four  spe- 
cies that  will  be  unfamiliar  to  British  and 
American  readers.  As  of  present  knowledge 
this  appears  to  be  the  definitive  work  on  the 
family  Sittidae. — GEORGE  A.  HALL. 


WORKING  EOR  WILDLIFE:  THE  BE- 
GINNING OF  PRESERVATION  IN  CANA- 
DA. By  Janet  Foster,  with  a foreword  and  an 
afterword  by  Lome  Hammond.  University  of 
Toronto  Press,  Toronto.  Second  ed.  1998:  297 
pp.,  38  black  and  white  photographs,  5 maps. 
$21.95  (paper). — In  1904,  Howard  Douglas, 
Superintendent  of  Canada’s  Rocky  Mountains 
Park,  learned  that  Michel  Pablo  wanted  to  sell 
a large  herd  of  bison  he  kept  in  western  Mon- 
tana. Although  as  many  as  20  to  30  million 
of  the  animals  had  once  roamed  the  North 
American  continent,  by  the  end  of  the  Nine- 
teenth Century  their  numbers  had  been 
thinned  to  fewer  than  a thousand.  Pablo’s 
herd,  which  he  estimated  at  around  360  indi- 
viduals, represented  the  largest  surviving  ag- 
gregation of  a species  that  seemed  to  be  rush- 
ing headlong  into  oblivion.  Although  Ameri- 
can conservationists  hoped  to  keep  Pablo’s  bi- 
son in  the  United  States,  they  failed  to  secure 
the  necessary  funds.  At  Douglas’s  urging,  the 
Canadian  government  purchased  the  herd  and 
shipped  it  by  rail  to  Rocky  Mountains  Park. 

This  story  is  one  of  many  fascinating  epi- 
sodes that  Janet  Foster  recounts  in  her  study 
of  the  origins  of  Canadian  wildlife  conserva- 
tion. According  to  Foster,  the  first  significant 
efforts  to  address  wildlife  decline  in  Canada 
came  at  the  end  of  the  nineteenth  century, 
when  a handful  of  senior  federal  civil  servants 
began  using  their  position  and  influence  to 
push  a protectionist  agenda.  Howard  Douglas 
and  his  dedicated  colleagues — Robert  Camp- 
bell, Director  of  the  Forestry  Branch,  Depart- 
ment of  Interior;  James  Harkin,  Commissioner 
of  Dominion  Parks;  Maxwell  Graham,  Chief 
of  the  Parks  Branch  Animal  Division;  and 
Gordon  Hewitt,  Division  Entomologist  with 
the  Department  of  Agriculture — are  the  main 
protagonists  in  Eoster’s  account,  which  begins 
in  the  mid  1880s  and  ends  in  the  early  1920s. 

None  of  this  small  group  was  a particularly 
prominent  public  figure,  and  only  one,  Harkin, 


454 


THE  WILSON  BULLETIN  • VoL  III,  No.  3,  September  1999 


had  any  formal  training  in  the  biological  sci- 
ences. Yet,  working  together  with  provincial 
officials,  other  interested  citizens,  and  their 
counterparts  in  the  United  States,  they  were 
remarkably  successful  in  transforming  their 
personal  commitment  to  wildlife  into  federal 
policy.  Fighting  a pervasive  belief  in  the  su- 
perabundance of  nature,  a national  agenda  that 
emphasized  settlement  and  development,  a 
public  that  seemed  largely  indifferent  to  the 
desperate  plight  of  wildlife,  and  a tradition 
that  left  resource  management  in  the  hands  of 
provincial  governments,  these  federal  officials 
established  national  parks,  created  wildlife 
preserves,  rallied  public  support  for  native 
species,  and  pushed  through  protective  legis- 
lation and  regulations.  Aiding  this  quintet  of 
federal  civil  servants  was  a larger  cast  of  char- 
acters who  receive  much  less  attention  in  this 
book,  including  the  ornithologists  Hoyes 
Lloyd,  Percy  Tavener,  James  Fleming,  and 
others. 

Foster’s  account  of  the  Migratory  Bird 
Treaty,  negotiated  in  the  years  around  World 
War  I,  will  be  of  particular  interest  to  readers 
of  this  journal.  After  more  than  a decade  of 
lobbying,  in  1913  wildlife  advocates  in  the 
United  States  finally  secured  a federal  law 
protecting  migratory  birds.  Fearing  that  the 
new  legislation  might  be  struck  down  on  con- 
stitutional grounds,  the  bill’s  supporters  then 
moved  to  have  its  provisions  introduced  into 
a treaty  with  Great  Britain.  Negotiations  soon 
bogged  down,  however,  when  officials  from 
the  Maritime  Provinces  balked  at  the  idea  of 
eliminating  spring  shooting.  Foster  demon- 
strates Hewitt’s  central  role  in  garnering  Ca- 
nadian support  for  this  landmark  treaty,  which 
remains  in  effect  to  this  day. 

This  book  is  a second  edition  of  a work  first 
published  two  decades  ago.  The  environmen- 
tal historian  Lome  Hammond  has  contributed 
the  only  significant  additions:  a new  foreword, 
which  briefly  describes  the  larger  context  of 
Foster’s  book,  and  a new  afterword,  which  re- 
views the  literature  on  Canadian  wildlife  con- 
servation published  since  the  first  edition.  If 
Foster  were  to  write  her  book  today,  undoubt- 
edly she  would  pity  more  attention  to  the  con- 
tributions of  sportsmen,  naturalists,  humani- 
tarians, and  provincial  wildlife  officials — all 
of  whom  have  received  much  scholarly  atten- 
tion since  the  first  edition.  Yet,  because  her 


book  is  so  well  written  and  based  on  solid 
archival  research,  it  remains  a useful  starting 
point  for  anyone  interested  in  the  early  history 
of  wildlife  conservation  in  Canada. — MARK 
V.  BARROW,  JR. 


HABITATS  FOR  BIRDS  IN  EUROPE:  A 
CONSERVATION  STRATEGY  FOR  THE 
WIDER  ENVIRONMENT.  Compiled  by  Gra- 
ham M.  Tucker  and  Michael  I.  Evans. 
BirdLife  Conservation  Series  No.  6,  BirdLife 
International,  Cambridge,  U.K.  1997:  464  pp., 
6 appendices.  $45.00  (paper)  (in  North  Amer- 
ica, contact  via  email:  BTUCKER® 
SIPRESS.SI.EDU). — This  ambitious  compi- 
lation from  8 habitat  working  groups  marks 
the  third  and  final  leg  of  a decadal  marathon 
sponsored  by  BirdLife  to  promote  the  conser- 
vation of  Europe’s  birds  (the  first  two  culmi- 
nated in  Grimmet  and  Jones’  1989  Important 
Bird  Areas  in  Europe  and  number  3 in  the 
BirdLife  series.  Tucker  and  Heath’s  1994 
Birds  in  Europe:  their  conserx’ation  status). 
Thirteen  workshops  were  held  across  Europe 
bringing  experts  together  to  prepare  conser- 
vation strategies  for  each  of  the  following  ma- 
jor habitat  types:  marine  habitats;  coastal  hab- 
itats; inland  wetlands;  tundra,  mires  (bogs), 
and  moorlands;  lowland  Atlantic  heathland; 
boreal  and  temperate  forests;  Mediterranean 
forest,  shrubland,  and  rocky  habitats;  and  ag- 
ricultural and  grassland  habitats.  For  each 
habitat  type,  information  is  provided  on  cur- 
rent distribution  (with  maps)  and  trends,  its 
history,  physical  and  biological  processes,  and 
its  dominant  flora  and  fauna.  In  addition,  the 
chief  values  of  habitat  to  humans  is  given,  and 
the  major  threats  to  the  habitat  quality  and 
quantity  of  priority  bird  species  are  identified 
in  both  text  and  tables.  The  last  section  of 
each  habitat  chapter  then  lists  conservation 
opportunities  such  as  legislation,  financial  in- 
centives, and  policy  initiatives,  then  broad 
conservation  recommendations  are  given. 

The  rationale  for  priority  bird  rankings  is 
explained  early  in  the  book,  providing  a very 
useful  model  for  other  large-scale  bird  (or  oth- 
er fauna)  conservation  efforts  around  the 
globe.  In  this  scheme,  5 classes  are  established 
of  “Species  of  European  Conservation  Con- 
cern” (or  SPECs);  SPEC  1 species  are  of 


ORNITHOLOGICAL  LITERATURE 


455 


global  concern  (rare,  endangered,  or  declining 
populations),  while  SPEC  4 and  5 species 
have  favorable  conservation  status  in  Europe. 
Next,  priority  categories  (from  A to  D)  for 
bird  species  in  each  habitat  are  established  us- 
ing a matrix  of  SPEC  category  X habitat  im- 
portance (percent  of  European  population  us- 
ing that  habitat).  Thus,  Priority  A species  in 
any  particular  habitat  are  those  most  vulner- 
able to  further  losses,  while  Priority  D species 
are  those  that  are  more  stable  and  widespread, 
with  less  dependence  upon  that  particular  hab- 
itat. 

Next,  the  principles  and  strategies  for  broad 
conservation  initiatives  in  Europe  are  out- 
lined, spanning  across  international  treaties 
and  conventions,  economic  instruments,  and 
policy  doctrine.  A dazzling  litany  of  some  25 
legislative  instruments  are  reviewed,  from 
global  to  more  local  European  perspectives, 
ranging  from  specifics  (agricultural  nitrate 
control  policy)  to  broad  measures  of  biodi- 
versity (so-called  Rio  Convention  of  1992). 
Then,  numerous  economic  instruments  are  re- 
viewed (e.g.,  various  European  Union,  EU, 
and  World  Bank  funds)  followed  by  other 
broad  initiatives  (e.g..  Birds  and  Habitats  Di- 
rectives in  EU).  It  would  seem  that  with  this 
bewildering  array  of  conventions  and  plans 
cutting  across  landscapes  and  political  bound- 
aries, that  Europe’s  conservation  needs  would 
all  be  well  taken  care  of!  But  alas,  as  with 
most  large  Plans,  “the  devil  is  in  the  de- 
tails”— developing  consensus  for  habitat  pro- 
tection, harvest  criteria,  or  emission  standards 
among  an  array  of  nations  with  vastly  differ- 
ent ideologies  and  histories  in  human-nature 
interactions  is  difficult. 

A survey  across  the  habitat  chapters  reveals 
the  following  order  (from  most  to  least)  from 
the  perspective  of  priority  species:  (1)  Agri- 
cultural and  grassland  species — 173  species 
(ca  70%  unfavorable  conservation  status), 
with  6 Priority  A species  (4  are  SPEC  1);  (2) 
Boreal  and  temperate  forests — 114  species 
(40%  unfavorable  status),  with  2 Priority  A 
species;  (3)  Inland  wetlands — 102  species 
(55%  unfavorable),  with  all  8 Priority  A spe- 
cies being  SPEC  1;  (4)  Mediterranean  forest, 
shrubland,  and  rocky  habitats — 100  species 
(65%  unfavorable),  with  10  Priority  A species 
(1  SPEC  1);  (5)  Coastal  habitats — 75  species 
(70%  unfavorable),  with  13  Priority  A species 


(5  are  SPEC  1);  (6)  Tundra,  mires,  and  moor- 
land— 73  priority  species  (37%  unfavorable), 
with  only  2 Priority  A species  (1  SPEC  1);  (7) 
Marine  habitats — 62  species  (45%  unfavor- 
able), with  6 Priority  A species  (1  SPEC  1); 
(8)  Lowland  Atlantic  heathland — 16  species 
(all  small  populations),  no  Priority  A or  B spe- 
cies. 

Some  recurrent  themes  are  the  need  to  in- 
tegrate habitat  conservation  planning  with 
other  sectors  and  programs  using  mechanisms 
such  as  Environmental  Impact  Assessment 
and  Strategic  Environmental  Assessment,  to 
modify  established  policies  within  the  EU 
such  as  Common  Agricultural  Policy  and 
Common  Eisheries  Policy  to  work  toward  sus- 
taining biodiversity  rather  using  the  more  my- 
opic traditional  focus,  to  remove  economic  in- 
centives that  destroy  habitat  and  reduce  di- 
versity (e.g.,  non-indigenous  tree  plantations), 
to  better  educate  the  public  about  the  benefits 
and  ecological  services  of  such  natural  habi- 
tats, and  to  work  to  develop  standards  and  cri- 
teria for  sustainability  of  habitats  across  na- 
tional boundaries.  The  fact  that  nearly  40%  of 
Europe’s  bird  species  show  an  unfavorable 
conservation  status  is  largely  due  to  intensi- 
fication. That  is,  intensification  of  farming  and 
silvicultural  practices  on  land,  aquaculture  and 
open  sea  fishing  in  coastal  and  marine  habi- 
tats, and  of  coastal  development,  recreation, 
and  tourism,  much  of  the  latter  especially  in 
Mediterranean  Europe. 

This  book  represents  an  enormous  effort  by 
many  experts.  The  details  are  displayed  in  the 
extensive  appendices,  the  figures  and  tables 
are  very  useful  in  summarizing  the  vast 
amounts  of  species,  habitat,  and  threat  infor- 
mation, and  the  chapters  are  neatly  and  co- 
herently packaged  so  that  they  can  nearly 
stand  alone  for  those  with  more  specific  in- 
terests. The  approaches  taken  in  prioritizing 
the  species  and  their  threats,  and  the  conser- 
vation recommendations  made  by  each  habitat 
working  group  provide  an  extremely  valuable 
reference  for  bird  conservationists  in  any  con- 
tinent, not  simply  Europe.  The  challenges  now 
lie  in  translating  all  of  those  directives,  con- 
ventions, and  biodiversity  initiatives  into  real 
Action  Plans  that  can  survive  the  turbulent  po- 
litical and  economic  seas,  especially  as  the 
new  Euro  currency  takes  hold.  This  book 
should  be  a library  requisite  for  conservation- 


456 


THE  WILSON  BULLETIN  • VoL  HI,  No.  3,  September  1999 


ists,  land  managers,  and  environmental  policy 
professionals  because  it  transcends  issues  of 
bird  conservation  and  prompts  us  to  consider 
true  integration  not  just  of  land-  and  ocean- 


scapes  and  their  intersections,  but  also  of  ecol- 
ogy, agriculture,  forestry,  fisheries,  econom- 
ics, and  of  course,  politics. — R.  MICHAEL 
ERWIN. 


Wilson  Bull.,  1 1 1(3),  1999,  pp.  456 


Announcement 


The  Lincoln  Park  Zoo  Scott  Neotropic  and 
Africa/Asia  Funds  support  field  research  in 
conservation  biology  around  the  world.  The 
Scott  Neotropic  fund  focuses  on  projects  un- 
dertaken in  Latin  America  and  the  Caribbean. 
The  fund  emphasizes  the  support  of  graduate 
students  and  other  young  researchers,  partic- 
ularly those  from  Latin  America.  Since  1986, 
the  fund  has  awarded  over  126  grants  in  19 
countries.  The  Africa/Asia  fund,  lauched  in 
1997,  focuses  on  projects  throughout  Africa, 
Asia,  and  the  Pacific.  Each  fund  supports 
projects  of  young  conservation  biologists  and 
between  5 and  15  projects  for  each  fund  are 


supported  each  year.  The  fund  awards  are  sel- 
dom greater  than  US$7500,  and  most  awards 
fall  in  the  range  of  $3000-$6000.  Initial  sup- 
port is  for  up  to  12  months  from  the  date  of 
award,  and  the  maximum  duration  of  support 
is  two  years.  The  current  deadline  for  receipt 
of  Scott  Neotropic  proposals  is  1 September, 
and  Africa/Asia  proposals  have  no  deadline 
for  1999.  For  additional  information  and  ap- 
plication procedures  go  to  www.lpzoo.com, 
email  steveedC® ix.netcom.com,  or  write  to: 
LINCOLN  PARK  ZOO  SNF/AA  FUNDS, 
% Director  of  Conservation  and  Science, 
Lincoln  Park  Zoo,  Chicago,  IL  60614. 


This  issue  of  The  Wilson  Bulletin  was  published  on  10  August  1999. 


THE  WILSON  BULLETIN 


Editor  ROBERT  C.  REASON 


Editorial  Board  KATHY  G.  BEAL 


Department  of  Biology 

State  University  of  New  York 

1 College  Circle 

Geneseo,  NY  14454 

E-mail:  WilsonBull@geneseo.edu 


Review  Editor  WILLIAM  E.  DAVIS,  JR. 


127  East  Street 

Foxboro,  Massachusetts  02035 


CLAIT  E.  BRAUN 
RICHARD  N.  CONNER 


Editorial  Assistants  TARA  BAIDEME 


Index  Editor  KATHY  G.  BEAL 


JOHN  LAMAR 
DANTE  THOMAS 
DORIS  WATT 


616  Xenia  Avenue 
Yellow  Springs,  Ohio  45387 


SUGGESTIONS  TO  AUTHORS 

See  Wilson  Bulletin,  110:152-154,  1998  for  more  detailed  “Instructions  to  Authors.” 
http://www.ummz.lsa.umich.edu/birds/wilsonbull.html 
Submit  four  copies  of  manuscripts  intended  for  publication  in  The  Wilson  Bulletin,  neatly  typewritten, 
double-spaced,  with  at  least  3 cm  margins,  and  on  one  side  only  of  good  quality  white  paper.  Do  not 
submit  xerographic  copies  that  are  made  on  slick,  heavy  paper.  Tables  should  be  typed  on  separate  sheets, 
and  should  be  narrow  and  deep  rather  than  wide  and  shallow.  Follow  the  AOU  Check-list  (Seventh  Edition, 
1998)  insofar  as  scientific  names  of  U.S.,  Canadian,  Mexican,  Central  American,  and  West  Indian  birds 
are  concerned.  Abstracts  should  be  brief  but  quotable.  Where  fewer  than  5 papers  are  cited,  the  citations 
may  be  included  in  the  text.  Follow  carefully  the  style  used  in  this  issue  in  listing  the  literature  cited; 
otherwise,  follow  the  “CBE  Scientific  Style  and  Format  Manual”  (AIBS  1994).  Photographs  for  illustra- 
tions should  have  good  contrast  and  be  on  glossy  paper.  Submit  prints  unmounted  and  provide  a brief  but 
adequate  legend  for  each  figure  with  all  captions  on  a single  page.  Do  not  write  heavily  on  the  backs  of 
photographs.  Diagrams  and  line  drawings  should  be  in  black  ink  and  their  lettering  large  enough  to  permit 
reduction.  Original  figures  or  photographs  submitted  must  be  smaller  than  22  X 28  cm.  Alterations  in 
copy  after  the  type  has  been  set  must  be  charged  to  the  author. 


NOTICE  OF  CHANGE  OF  ADDRESS 


If  your  address  changes,  notify  the  Society  immediately.  Send  your  complete  new  address  to  Ornitho- 
logical Societies  of  North  America,  P.O.  Box  1897,  Lawrence,  KS  66044-8897. 

The  permanent  mailing  address  of  the  Wilson  Ornithological  Society  is:  c/o  The  Museum  of  Zoology, 
The  University  of  Michigan,  Ann  Arbor,  Michigan  48109.  Persons  having  business  with  any  of  the  officers 
may  address  them  at  their  various  addresses  given  on  the  back  of  the  front  cover,  and  all  matters  pertaining 
to  the  Bulletin  should  be  sent  directly  to  the  Editor. 


MEMBERSHIP  INQUIRIES 


Membership  inquiries  should  be  sent  to  Laurie  J.  Goodrich,  Route  2 Box  301  A,  New  Ringgold,  PA 
17960-9445;  E-mail:  goodrich@haukmountain.org. 


CONTENTS 


ANTILLEAN  SHORT-EARED  OWLS  INVADE  SOUTHERN  FLORIDA  

- Wayne  Hoffman,  Glen  E.  Woolfencien,  and  P.  William  Smith  303 

WITHIN-  AND  BETWEEN-YEAR  DISPERSAL  OF  AMERICAN  AVOCETS  AMONG  MULTIPLE 

WESTERN  GREAT  BASIN  WETLANDS 

Jonathan  H.  Plissner,  Susan  M.  Hai^,  and  Lewis  W.  Orin^  3 1 4 

HIGH  MORTALITY  OF  PIPING  PLOVERS  ON  BEACHES  WITH  ABUNDANT  GHOST  CRABS: 

CORRELATION,  NOT  CAUSATION Donna  L.  Wolcott  and  Thomas  G.  Wolcott  321 

A TAXONOMIC  STUDY  OF  CRESTED  CARACARAS  (FALCONIDAE) 

- - — Carla  J.  Dove  and  Richard  C.  Banks  330 

VISUAL  COMMUNICATION  AND  SEXUAL  SELECTION  IN  A NOCTURNAL  BIRD  SPECIES,  CA- 
PRIMULGUS  RUF/COLUS,  A BALANCE  BETWEEN  CRYPSIS  AND  CONSPICUOUSNESS  ... 

Juan  Aragones,  Luis  Arias  De  Reyna,  and  Pilar  Recuerda  340 

INTERSPECIFIC  INTERACTIONS  WITH  FORAGING  RED-COCKADED  WOODPECKERS  IN 

SOUTH-CENTRAL  FLORIDA  

Reed  Bowman,  David  L.  Leonard,  Jr.,  Leslie  K.  Backus,  and  Allison  R.  Mains  346 

SPATIAL  AND  TEMPORAL  DYNAMICS  OF  A PURPLE  MARTIN  PRE-MIGRATORY  ROOST  ...... 

- - Kevin  R.  Russell  and  Sidney  A.  Gauthreau.x,  Jr.  354 

AGGRESSIVE  RESPONSE  OF  CHICKADEES  TOWARDS  BLACK-CAPPED  AND  CAROLINA 

CHICKADEE  CALLS  IN  CENTRAL  ILLINOIS  Eric  L.  Kershner  and  Eric  K.  Bollinger  363 

USE  OF  SONG  TYPES  BY  MOUNTAIN  CHICKADEES  (POECILE  GAMBELL)  

Myra  O.  Wiehe  and  M.  Ross  Lein  368 

SURVIVAL  AND  LONGEVITY  OF  THE  PUERTO  RICAN  VIREO  

- - Bethany  L.  Woodworth,  John  Faahorg.  and  Wayne  J.  Arendt  376 

EFFECTS  OF  PRIOR  RESIDENCE  AND  AGE  ON  BREEDING  PERFORMANCE  IN  YELLOW  WAR- 
BLERS   G.  A.  Lozano  and  R.  E.  Lemon  381 

DISTRIBUTION  AND  HABITAT  ASSOCIATIONS  OF  THREE  ENDEMIC  GRASSLAND  SONG- 
BIRDS IN  SOUTHERN  SASKATCHEWAN  S.  K.  Davis,  D.  C.  Duncan,  and  M.  Skeel  389 

BIRD  COMMUNITIES  IN  NATURAL  FOREST  PATCHES  IN  SOUTHERN  BRAZIL  

Luiz  Dos  Anjos  and  Roberto  Bo^  on  397 

DO  MAMMALIAN  NEST  PREDATORS  FOLLOW  HUMAN  SCENT  TRAILS  IN  THE  SHORTGRASS 

PRAIRIE?  Susan  K.  Skagen,  Thomas  R.  Stanley,  and  M.  Beth  Dillon  415 

SHORT  COMMUNICATIONS 

CHRISTMAS  SHEARWATER  EGG  DIMENSIONS  AND  SHELL  CHARACTERISTICS  ON  LAY- 

SAN  ISLAND,  NORTHWESTERN  HAWAIIAN  ISLANDS  

- G.  C.  Whittow  and  M.  B.  Naughton  42 1 

THE  PAINT-BILLED  CRAKE  BREEDING  IN  COSTA  RICA  

David  M.  Watson  and  Brett  W.  Benz  422 

ADDITIONAL  RECORDS  OF  FALL  AND  WINTER  NESTING  BY  KILLDEER  IN  SOUTHERN 

UNITED  STATES  Kimberly  G.  Smith,  W.  ManJn  Davis,  Thomas  E.  Kienzle, 

William  Post,  and  Robert  W.  Chinn  424 

WILD  TURKEYS  (MELEAGRIS  GALLOPAVO)  RENEST  AFTER  SUCCESSFUL  HATCH  

— Craig  A.  Harper  and  Jay  H.  E.xum  426 

POST-MIGRATION  WEIGHT  GAIN  OF  SWAINSON’S  HAWKS  IN  ARGENTINA  

Michael  I.  Goldstein,  Peter  H.  Bloom.  Jose  H.  Sarasola,  and  Thomas  E.  Lacher  428 

SIBLICIDE  AT  NORTHERN  GOSHAWK  NESTS:  DOES  FOOD  PLAY  A ROLE?  

. Wendy  A.  Estes,  Sarah  R.  Dewey,  and  Patricia  L.  Kennedy  432 

COOPERATIVE  FORAGING  IN  THE  MOUNTAIN  CARACARA  IN  PERU  Ja.son  Jones  437 

PREDATION  BY  RUFOUS  MOTMOT  ON  BLACK-AND-GREEN  POISON  DART  FROG  


Terry  L.  Master  439 

EVIDENCE  OF  EGG  EJECTION  IN  MOUNTAIN  BLUEBIRDS  Percy  N.  Hebert  440 

FORAGING  OVENBIRD  FOLLOWS  ARMADILLO  ..  Douglas  J.  Uvey  ,443 

ORNITHOLOGICAL  LITERATURE  ...._ 445 


The  Wilson  Bulletin 

PUBLISHED  BY  THE  WILSON  ORNITHOLOGICAL  SOCIETY 


VOL.  Ill,  NO.  4 DECEMBER  1999  PAGES  457-630 

(ISSN  0043-5643) 


THE  WILSON  ORNITHOLOGICAL  SOCIETY 
FOUNDED  DECEMBER  3,  1888 

Named  after  ALEXANDER  WILSON,  the  first  American  Ornithologist. 

President  John  C.  Kricher,  Biology  Department,  Wheaton  College,  Norton,  Massachusetts  02766; 
E-mail;  JKricher@wheatonma.edu. 

First  Vice-President — William  E.  Davis,  Jr.,  College  of  General  Studies,  871  Commonwealth  Ave.,  Boston 
University,  Boston,  Massachusetts  02215;  E-mail;  WEDavis@bu.edu. 

Second  Vice-President — Charles  R.  Blem,  Dept,  of  Biology,  816  Park  Ave.,  Virginia  Commonwealth 
Univ.,  Richmond,  Virginia  23284;  E-mail;  cblem@saturn.vcu.edu. 

Editor — Robert  C.  Beason,  Department  of  Biology,  State  University  of  New  York,  1 College  Circle, 
Geneseo,  New  York  14454;  E-mail;  WilsonBull@geneseo.edu. 

Secretary — John  A.  Smallwood,  Department  of  Biology,  Montclair  State  University,  Upper  Montclair, 
New  Jersey  07043;  E-mail;  Smallwood@mail.montclair.edu. 

Treasurer — Doris  J.  Watt,  Department  of  Biology,  Saint  Mary’s  College,  Notre  Dame,  Indiana  46556; 
E-mail;  DWatt@saintmarys.edu. 

Elected  Council  Members — Charles  F.  Thompson  and  Sara  R.  Morris  (terms  expire  2000),  Jonathan  L. 
Atwood  and  James  L.  Ingold  (terms  expire  2001),  Robert  A.  Askins  and  Jeffrey  R.  Walters  (terms 
expire  2002). 

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  Bulletin  is  sent  to  all  members  not  in  arrears  for  dues. 

THE  JOSSELYN  VAN  TYNE  MEMORIAL  LIBRARY 
The  Josselyn  Van  Tyne  Memorial  Library  of  the  Wilson  Ornithological  Society,  housed  in  the  University 
of  Michigan  Museum  of  Zoology,  was  established  in  concurrence  with  the  University  of  Michigan  in 
1930.  Until  1947  the  Library  was  maintained  entirely  by  gifts  and  bequests  of  books,  reprints,  and  orni- 
thological 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  administered  by  the  Library  Committee.  William  A.  Lunk,  University  of 
Michigan  is  Chairman  of  the  Committee.  The  Library  currently  receives  over  200  periodicals  as  gifts  and 
in  exchange  for  The  Wilson  Bulletin.  For  information  on  the  library  and  our  holdings,  see  the  Society’s 
web  page  at  http;//www.ummz. Isa. 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  University 
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 
University  of  Michigan,  1 109  Geddes  Ave.,  Ann  Arbor,  MI  48109-1079  USA.  Contributions  to  the  New 
Book  Fund  should  be  sent  to  the  Treasurer. 


THE  WILSON  BULLETIN 
(ISSN  0043-5643) 

THE  WILSON  BULLETIN  (ISSN  0043-5643)  is  published  quarterly  in  March,  June,  September,  and  December  by 
the  Wilson  Ornithological  Society,  810  East  1 0th  Street,  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  THE  WILSON  BULLETIN,  P.O.  Box  1897,  Lawrence,  KS  66044-8897. 

Back  issues  or  single  copies  are  available  for  $12.00  each.  Most  back  issues  of  the  Bulletin  are  available  and  may  be 
ordered  from  the  Treasurer.  Special  prices  will  be  quoted  for  quantity  orders. 

All  articles  and  communications  for  publications,  books  and  publications  for  reviews  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  the  OSNA,  P.O.  Box  1897, 
Lawrence,  KS  66044-8897.  Phone:  (785)  843-1221;  FAX:  (785)  843-1274. 

© Copyright  1999  by  the  Wilson  Ornithological  Society 
Printed  by  Allen  Press,  Inc.,  Lawrence,  Kansas  66044,  U.S.A. 


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


MCZ 

LIDRA.^Y 

btU  22  1^99 


HA'^V'AR'O 

UlMlVZi^JirY 


THE  WILSON  BULLETIN 

A QUARTERLY  JOURNAL  OF  ORNITHOLOGY 
Published  by  the  Wilson  Ornithological  Society 

VOL.  Ill,  NO.  4 DECEMBER  1999  PAGES  457-630 

Wilson  Bull.,  111(4),  1999,  pp.  457-464 

A NEW  SPECIES  OF  HAWK-OWL  NINOX  FROM  NORTH 

SULAWESI,  INDONESIA 

PAMELA  C.  RASMUSSEN* 


ABSTRACT. — A distinctive  new  species  of  hawk-owl,  Ninox  ios,  is  described  from  a specimen  collected  in 
1985  in  forest  at  1120  m in  Bogani  Nani  Wartabone  (then  Dumoga-Bone)  National  Park,  North  Sulawesi, 
Indonesia.  It  was  previously  identified  as  a rufous  morph  of  the  Ochre-bellied  Hawk-Owl,  N.  ochracea.  Ninox 
ios  is  small,  predominantly  bright  chestnut,  and  lacks  facial  patterning;  it  has  pink  orbital  skin,  yellow  irides, 
triangular  whitish  scapular  spots,  a finely  banded  and  relatively  long  tail,  unusually  short,  slender  tarsi  that  are 
feathered  for  most  of  their  length,  and  weak  claws.  Its  relationships  within  the  genus  Ninox  are  unclear;  it  differs 
in  several  morphological  characters  from  all  other  species.  Because  Ninox  ios  is  only  known  from  one  specimen, 
its  distribution  and  conservation  status  are  unknown;  nothing  is  known  of  its  ecology,  but  it  probably  occurs 
primarily  at  higher  elevations  than  N.  ochracea.  Received  14  Dec.  1998,  accepted  5 May  1999. 


For  many  years  two  endemic  species  of  the 
genus  Ninox  were  thought  to  occur  on  the  cen- 
tral Indonesian  island  of  Sulawesi.  Of  these, 
the  Speckled  Hawk-Owl  (Ninox  punctulata) 
primarily  inhabits  disturbed  lowland  habitats 
throughout  the  island  (White  and  Bruce  1986), 
and  is  morphologically  quite  different  from 
other  endemic  Indonesian  Ninox.  The  poorly 
known  Ochre-bellied  Hawk-Owl  [A.  ochracea 
(=  perversa)]  of  the  lowland  rainforests  in 
North  and  Central  Sulawesi  (White  and  Bruce 
1986)  is  a small,  fairly  typical  member  of  its 
genus  (Frontispiece).  Because  there  had  been 
no  indication  that  a third  species  might  occur, 
it  was  a surprise  when  in  1985  F.  G.  Rozen- 


' NHB  336  MRC  1 14,  Smithsonian  Institution, 
Washington,  DC  20560,  and  Michigan  State  University 
Museum,  East  Lansing,  MI  48824-0590; 

E-mail:  rasmussen.pamela@nmnh.si.edu 


daal  netted  an  almost  entirely  bright  rufous 
Ninox  (Frontispiece)  in  Bogani  Nani  Warta- 
bone (then  Dumoga-Bone)  National  Park, 
North  Sulawesi,  Indonesia  (Fig.  1).  He  con- 
cluded that  this  individual  represented  “a  pre- 
viously undescribed  rufous  phase”  of  N. 
ochracea  (Rozendaal  and  Dekker  1989),  and 
this  treatment  was  followed  by  Coates  and 
Bishop  (1997). 

While  working  on  small  owls  at  the  Na- 
tional Museum  of  Natural  History/Naturalis, 
Leiden  (NNM,  formerly  Rijksmuseum  van 
Natuurlijke  Historie,  RMNH)  in  June  and  Oc- 
tober 1998,  I chanced  to  see  the  rufous  Sula- 
wesi specimen,  which  had  been  registered  as 
RMNH  84701  but  had  not  yet  been  incorpo- 
rated into  the  main  collection  following  its 
purchase  by  NNM.  On  the  second  occasion  I 
noted  that  it  differed  in  several  morphological 
features  from  Ninox  ochracea,  in  addition  to 


FRONTISPIECE.  Cinnabar  Hawk-Owl  (Ninox  ios,  upper  two)  compared  with  Oehre-bellied  Hawk-Owl  (TV. 
ochracea,  lower  left),  and  Bum  race  of  Moluccan  Hawk-Owl  (TV.  squamipila  hantu,  lower  right).  Original 
watercolor  painting  by  Ian  Lewington. 


457 


458 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


FIG.  1.  Map  of  Wallacea  showing  collection  locality  of  the  holotype  of  the  Cinnabar  Hawk-Owl  (Ninox 
ios),  other  locations  mentioned  in  the  text,  and  approximate  ranges  within  the  region  of  other  species  of  Ninox 
and  subspecies  of  W.  scutulata,  which  occurs  throughout  the  region. 


the  obvious  color  differences.  Subsequent 
mensural  analyses  of  series  of  all  species  of 
Ninox  have  confirmed  the  distinctness  of  the 
rufous  Sulawesi  specimen  (an  adult  in  good 
condition)  in  many  characters.  Although  sev- 
eral Ninox  species  from  other  areas  are  typi- 
cally rufous,  morphism  (and  thus  true  rufous 
morphs)  appears  to  be  unknown  for  any  Ni- 
nox, and  in  any  case  most  of  the  differences 
are  structural  and  thus  would  not  be  related  to 
morph.  Despite  the  fact  that  only  one  rufous 
specimen  is  known  thus  far  from  Sulawesi, 
there  is  no  reason  to  believe  that  any  of  its 
several  novel  character  states  are  aberrant,  and 
there  can  be  no  reasonable  doubt  that  it  rep- 
resents a new  species. 


CINNABAR  HAWK-OWL  Ninox  ios  sp.  nov. 

Holotype. — RMNH  84701,  adult  male 
(Frontispiece),  according  to  the  label  collected 
in  a forested  valley  at  1 120  m at  Clark’s  camp 
(Hill  1440),  east-central  Bogani  Nani  Warta- 
bone  National  Park,  North  Sulawesi,  Indone- 
sia (ca  0°40'  N,  123°  0'  E)  by  F.  G.  and  C. 
M.  Rozendaal  the  night  of  5-6  April  1985  (the 
label  date  of  7 April  presumably  indicates  date 
of  death).  Label  data:  “Completely  ossified 
skull”,  “weight  78  g”. 

Diagnosis. — A small,  lightly  built,  nearly 
uniformly  rich  chestnut  hawk-owl  with  a rel- 
atively long  tail  and  narrow  pointed  wings,  lax 
feathering,  no  facial  pattern,  mostly  feathered 


Rasmussen  • A NEW  SPECIES  OF  HAWK-OWL 


459 


short  slender  tarsi,  and  rufous,  narrowly  dark- 
barred  wings  and  tail. 

Compared  with  all  flying  states  of  Ninox 
ochracea  [n  = 20  (three  of  which  are  fully 
grown  juveniles);  6 males,  4 females,  10  un- 
sexed],  N.  ios  is  much  smaller  in  most  dimen- 
sions (Table  1,  Fig.  2),  but  has  a relatively 
longer  tail  and  rictal  bristles.  Its  wing,  while 
shorter  than  that  of  N.  ochracea,  is  narrower 
and  more  pointed  (Fig.  3).  Ninox  ios  has  a 
much  shorter,  shallower  bill  and  smaller  nares 
than  N.  ochracea.  It  has  short,  slender  tarsi 
that  are  mostly  feathered  on  both  surfaces, 
whereas  N.  ochracea  has  longer,  stout  tarsi 
that  are  largely  unfeathered  on  the  anterior 
(acrotarsal)  side  and  are  virtually  unfeathered 
on  the  posterior  (plantar)  side,  with  numerous 
stiff  bristles  over  the  unfeathered  areas.  The 
new  species  has  relatively  sparse,  fine  rufous 
bristles  on  the  extreme  lower  tarsi  and  on  its 
slender  toes  (although  the  bristles  are  heavier 
and  longer  on  the  hallux),  while  N.  ochracea 
has  more  profuse,  heavier,  mostly  pale  bristles 
(which  are  usually  longer  but  sometimes  worn 
down  to  stubs)  on  the  tops  and  sides  of  its 
stouter  toes.  Ninox  ios  has  much  smaller,  more 
slender  claws  that  are  dark  for  most  of  their 
length  (vs  large  and  mostly  pale  in  N.  ochra- 
cea). The  holotype  of  TV.  ios  had  pink  orbital 
skin  (vs  blackish  in  TV.  ochracea)  and  yellow 
eyes,  as  does  TV.  ochracea  according  to  Strese- 
mann  (1940),  who  based  this  statement  on  G. 
Heinrich’s  specimens  [although  Meyer  and 
Wiglesworth  (1898)  mentioned  a brown-eyed 
TV.  ochracea].  The  base  of  the  bill  and  the  cere 
of  TV.  ios  appear  entirely  pale  (vs  the  basal 
two-thirds  conspicuously  dark  in  specimens  of 
TV.  ochracea). 

In  plumage,  TV.  ios  differs  conspicuously 
from  both  adults  and  juveniles  of  TV.  ochracea 
in  its  overall  bright  rufous  coloration  (vs  dark 
brown  and  yellow-ocher).  Unlike  all  flying 
stages  of  TV.  ochracea,  it  lacks  facial  pattern- 
ing, including  the  whitish  supercilia  typical  of 
most  of  its  relatives,  and  also  lacks  white 
markings  in  the  wing  coverts  and  flight  feath- 
ers. Less  obvious  distinctions  from  TV.  ochra- 
cea include  its  more  triangular  (vs  squarer 
tipped)  whitish  scapular  spots,  its  mainly  ru- 
fescent  rictal  bristles  (vs  blackish  with  white 
bases),  its  more  narrowly  barred  rectrices,  its 
vaguely  dark-scalloped  lower  underparts  (vs 
plain  ocher  or  somewhat  brown- streaked),  and 


the  patterning  of  its  breast  feathers,  which 
have  a light  rufous  (vs  dark  brown)  area  sur- 
rounding the  whitish  shafts. 

The  Philippine  Hawk-Owl  {Ninox  philip- 
pensis)  superspecies  (sensu  Dickinson  et  al. 
1991,  but  see  Collar  and  Rasmussen  1998)  is 
composed  of  several  dark  brown  to  brown- 
and-ocher  forms  that  are  either  barred  or 
streaked  below.  None  of  the  taxa  included  in 
TV.  philippensis  can  be  described  as  warmer- 
toned  than  rufescent  brown.  All  have  much 
heavier  claws  and  relatively  shorter  tails  (Fig. 
2A)  than  TV.  ios,  from  which  they  also  differ 
in  wing  shape  (Fig.  3).  One  form,  TV.  [philip- 
pensis] mindorensis  (see  Frontispiece),  is 
somewhat  similar  in  overall  size  and  tarsal 
feathering  to  TV.  ios  than  is  any  other  taxon, 
including  TV.  ochracea  (Fig.  2C),  but  not  in 
plumage  or  the  above-mentioned  shape  char- 
acters. 

All  taxa  of  the  paraphyletic  Moluccan 
Hawk-Owl  {Ninox  squamipila\  split  provi- 
sionally into  at  least  three  species  by  Norman 
et  al.  1998)  are  considerably  larger  and  heavi- 
er-legged than  TV.  ios,  and  all  differ  from  it 
additionally  in  having  whitish-barred  under- 
parts and  scapulars.  Despite  the  above  differ- 
ences, Ninox  s.  hantu  (Frontispiece)  of  Bum 
superficially  resembles  the  much  smaller  TV. 
ios  because  of  its  overall  mfescence  and  re- 
duced barring  below,  as  well  as  its  obscure 
facial  pattern  and  finely  barred  tail.  The  Sum- 
ba  Hawk-Owl  (TV.  rudolfi)  is  large  and  strik- 
ingly different,  with  a heavily  spotted  crown, 
barred  underparts,  and  broadly  banded  and 
speckled  upperparts.  The  widespread  and  var- 
iable Brown  Hawk-Owl  {Ninox  scutulata)  is 
also  a much  larger  species,  with  a broadly 
banded  tail  and  large,  heavily  feathered  tarsi. 
It  is  dark  brown  above  with  the  underparts 
heavily  streaked,  or  nearly  solid  dark  brown 
in  TV.  s.  obscura  of  the  Andamans.  The  nom- 
inate race  of  the  Andaman  Hawk-Owl  (TV.  a. 
affinis)  is  smaller  than  TV.  scutulata,  to  which 
it  is  otherwise  quite  similar,  while  the  larger 
Nicobar  race  (TV.  a.  isolata)  is  even  more  like 
some  races  of  TV.  scutulata. 

The  highly  varied  subspecies  (including  a 
new  one  described  from  Roti  Island,  south- 
west of  Timor,  Lesser  Sundas;  Johnstone  and 
Darnell  1997)  usually  grouped  in  the  Southern 
Boobook  {Ninox  novae seelandiae)  as  well  as 
the  Manus  Hawk-Owl  (TV.  meeki)  are  also 


460 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


in 

f \ r" 


a 


(j  V 
CL  ^ 


c/5 


00 

c 


c 

<u 

C ! 

o • 


s 

<u 

s '5 

> *5 

u 


- ^ 


-*  ^'.SJ 

I d 

^ s 

u 

<£  ° M 

_ c 

- "5  ^ 


Q 

(/) 


C -O 
o u 


C/5 

C 

S3  ^ 
1)  ^ 


+1-^2 
+1  - 


__  SJ 

ULl  V 

J i 
CQ  .S 

S? 

,o 


o 

U« 

a 

B 


c/5 


c/5 

<D  .ti 

u 

c/5 

2 ofl 

d) 

x: 

TD 

O c^ 

o 

(N 

o 

00 

04 

ro 

ro  fN  St  O 

o 

On 

ON 

OS 

NO 

o 

m 

o 

NO 

d) 

D.  0 

m 

ro 

CO 

fO  ro  fO  fO 

ro 

04 

04 

04 

CO 

CO 

C^5 

u 

" (N 

C3 

a 

d)  1 

-C  — 

*§ 

;:; 

3.9 

CN 

(N 

ro 

sd 

0^ 

3.5 

sd  O;  ; 

04  d d 04 

q 

rd 

q 

04 

oi 

QO 

04 

04 

in 

q 

cn 

d 

00 

d 

04 

04 

c 

Ci. 

■ 

O CJ 

<9j 

L< 

+1 

+1 

+1 

+1 

+1 

+1 

+1  +1  +1  +1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

c 

OX) 

Cs 

NO 

CL  C 

(N 

On 

SO 

tN  NO  St 

o 

ON 

00 

CO 

00 

00 

03 

rn 

d 

<N 

d 

04 

NO  d d 1/5 

in 

d 

rd 

oi 

NO 

d 

cd 

—H 

NO 

ob 

CN 

CN 

(N 

r- 

in 

04 

04 

ro 

I*-* 

04 

— - 

<u 

Ui 

c3 

^ -= 

3 

c/5 

d) 

E ji 

'u 

OJJ 

C 

3 

to  ^ 

DC 

V) 

« 

'O 

On 

00 

o 

r- 

in 

04  — OS  O' 

fO 

o 

o 

os 

in 

o 

00 

in 

•o 

d) 

r- 

SO 

NO 

r- 

NO 

NO 

SO  so  m »n 

m 

in 

m 

O' 

O' 

NO 

O' 

sO 

“O 

3 ^ 

k. 

r4 

o 

O' 

00 

ON  On  — 

fO 

00 

04 

00 

m 

04 

CO 

0- 

CO 

_3 

o 2 

a. 

d 

04 

so 

rd 

d 

— d d 04 

04 

04 

cd 

cd 

oi 

d 

d 

04 

O 

c 

^ 2 

s/3 

<<1 

+1 

+1 

+1 

+1 

+1 

+1 

+1  +1  +1  +1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

p o 

NO 

m 

in 

so 

00 

O 3-  St  NO 

— 

NO 

os 

Os 

00 

0- 

NO 

04 

c 

d 

§; 

NO 

,-H 

sd  d d in 

in 

in 

cd 

d 

d 

d 

cd 

in 

d) 

,(D 

(N 

r4 

04 

0- 

in 

04 

04 

CO 

00 

04 

" 

C3 

(U  Nm 

3 

X 

£ 3 

a 

2 o 

■<— * 

C/5  lU 

53 

^ 3 

ou 

y s 

NO 

a 

-5  3 

in 

SO 

0- 

so 

sO 

in  so  so  so 

NO 

m 

CO 

SO 

in 

in 

CN 

CN 

04 

04 

04 

04 

04  04  04  04 

04 

04 

04 

04 

04 

04 

04 

04 

04 

x: 

OX)  ^ 

o 

3 

2 13 

_o 

p 

q 

in 

d 

O) 

5.2 

04 

1.7 

1.2 

0.8 

2.1 

2.6 

2.7 

2.8 

CO 

cd 

3.9 

q 

oi 

04 

d 

0- 

d 

q 

— ■£ 

+1 

+1 

+1 

+1 

+1 

+1 

+1  +1  +1  +1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+1 

+ 1 

c-^ 

2 « 

■l 

in 

OS 

o 

— ' -t  ro  NO 

04 

NO 

ON 

04 

NO 

04 

O 

d 

cn 

d 

•d 

O'! 

04 

— d 2 

rd 

cd 

00 

oi 

Os 

cd 

cd 

CN 

(N 

NO 

m 

04 

04 

CO 

CO 

00 

04 

»— 1 

o 

C/5  d) 

_ x: 

OJ 

N 

&-  d:; 

* 

- o 

1 

C 

c a 

. 

o - 

/-'V 

r-s 

d) 

c/5 

n 

S', 

OS 

0^ 

q 

M x; 

IT) 

in 

C4 

m 

00 

o 

04 

sp 

04 

"o 

Wi  -w 

CS 

(N 

04 

Os 

o 

<o 

04 

m 

1 

d 

in 

CO 

1 

04 

1 

> E 

O 

1 

O 

m 

o 

1 

so 

04 

1 

ON  ro  — . 

1 1 1 1 

04 

1 

q 

1 

in 

1 

q 

in 

1 

CO 

Np 

d 

d 

d 

O' 

r- 

in 

in 

00 

00 

CO 

00 

NO 

d 

cd 

rN 

04  O O loi 

04 

CO 

00 

04 

cd 

r-  “O 

w w w w 

3 QJ 

r- 

On 

NO 

On 

00 

m 

r-  m 

NO 

SO 

so 

NO 

00 

NO 

NO 

ON 

d 

0 E 

u ^ 

? 

— 

— ' 

"" 

^ ^ ^ 

" 

<u 

M s 

§ 2 


ir)'r'--00O000N00^O'O'Ot^0N— I'sTroiriO 
— r-iOu-istr<S— 'O— -(NCNricNCNVO— "OOrO 
+1  +1  +1  +1  +1  +1  +1  +1  +1  +1  +1  +1  +1  +1  +1  +1  +1  +1  +1 
m — ^cNO(Nmr<^staNOt^t^— 'f^'Ooor^oo 

d — 


ro  — O'  lO  O 
fs  n — 00  lo  (N 


0\  — ■ On  00  On  m 

'sj'  On  OJ 


On  NO  O; 

r'CNror'imvocNO 
— 04  (N  r'  in  — 


NO  O 00  NO 
04— ' — (SOONl^t^fNldON— ' 
— <04rOrON^ONr4  — 


3 

CO 

E 

o 


a 

jj 

c 

E 

3 

U 


wj 

s:  c 

w d) 
&0  — 
C d) 

c/5 


D 

E « 
^ 2 
< Qi 


— 3: 


222222222 

tcct^dctcdtt: 

ooooooooc 

^C/5C/5C/5C/5C/5C/5C/5C/5 

-^rl^^i^»r)'Or-ooc^ 

cucuCua<(xa<cu(XCL 


o 

-C 

c/5 

o 

a! 


Clij 

c 

-E 

c/5 

D 

c/5 


■S 

c/5 

3 

c/5 

Ui 

s 

3 

s 


c/5 

£ S3 
00 
c 

(U 


d) 


d> 


s s D 


Rasmussen  • A NEW  SPECIES  OF  HAWK-OWL 


461 


Culmen  from  skull  (mm) 


Tarsus  length  (mm) 

FIG.  2.  Bivariate  scatter  plots  (measurements  in  mm)  for  Ninox  ios  (filled  circles),  N.  ochracea,  N.  “squam- 
ipila”  (sensu  White  and  Bruce  1986),  and  N.  philippensis  (sensu  Dickinson  et  al.  1991):  A.  culmen  vs  tail 
length;  B.  auricular  vs  rictal  bristle  length;  and  C.  tarsus  length  vs  extent  of  unfeathered  tarsus.  For  A,  the  main 
Philippine  taxon  groups  are  treated  separately,  while  for  B and  C they  are  combined. 


462 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


60- 


FIG.  3.  Shortfalls  from  wingpoint  of  each  primary 
(PI  = outer  primary;  shortfalls  are  distance  of  tip  of 
each  primary  from  longest  primary  in  folded  wing)  for 
Ninox  ios,  N.  ochracea,  N.  [philippensis]  spilocephala, 
and  N.  [/?.]  spilonota.  Ninox  p.  philippensis  and  similar 
races  are  virtually  identical  in  pattern  of  primary  short- 
falls to  N.  [p.]  spilocephala  and  thus  are  not  shown 
separately,  while  all  bar-bellied  populations  in  the  N. 
philippensis  superspecies  are  similar  to  race  N.  p.  spi- 
lonota. 


larger  than  N.  ios  and  are  streaked  or  heavily 
blotched  below.  None  of  the  remaining  Aus- 
tralasian taxa  (Papuan  Boobook  Owl,  N.  theo- 
macha;  Rufous  Owl,  N.  rufa'.  Powerful  Owl, 
N.  strenua-,  or  Barking  Owl,  N.  connivens)  ap- 
proach N.  ios  more  closely  than  the  above. 
The  other  Sulawesi  endemic.  Speckled  Hawk- 
Owl  (N.  punctulata),  and  some  Melanesian 
taxa  (Bismarck  Hawk-Owl,  N.  variegata-.  New 
Britain  Hawk-Owl,  N.  odiosa;  and  Solomons 
Hawk-Owl,  N.  jacquinoti)  are  strikingly  dif- 
ferent in  plumage  and  morphology,  with  short 
tails,  very  heavy  tarsi,  and  Athene plum- 
age pattern  and  toe  bristles;  in  fact  some  had 
been  placed  in  that  genus  (among  others)  in 
the  past.  The  White-browed  Owl  (Ninox  su- 
perciliaris)  of  Madagascar  is  very  different 
from  other  Ninox  (H.  F.  James,  pers.  comm.) 
as  would  be  predicted  by  its  distribution. 

Distribution. — To  date  Ninox  ios  is  known 
only  from  the  type  locality  in  North  Sulawesi, 
Indonesia.  It  might  occur  at  similar  elevations 
elsewhere  in  the  Minahasa  Peninsula  of  North 
Sulawesi. 

Description  of  the  holotype. — Color  match- 
ing was  done  under  natural  light  using  Mun- 


sell  (1977)  notation,  in  which  the  first  number 
and  letters  refer  to  the  hue,  the  number  pre- 
ceding the  slash  is  the  value  or  lightness,  and 
the  last  number  is  the  chroma  or  saturation. 
The  holotype  was  directly  compared  with  11 
specimens  of  N.  ochracea  (including  the  ho- 
lotype) at  NNM,  and  a series  of  photographs 
of  it  was  compared  to  specimens  from  other 
museums. 

Front  of  head  from  base  of  bill  through  cen- 
ter of  forecrown  and  including  supercilia,  uni- 
form rich  chestnut  (5YR  5/8);  center  of  crown 
through  mantle  slightly  darker  (close  to  5YR 
4/6);  rictal  bristles  fairly  long  (maximum  24 
mm),  profuse,  and  dark  chestnut,  somewhat 
blackish  near  tips;  auriculars  with  fairly  long 
distally  extended  barbs  (total  length  of  longest 
feather  23  mm)  that  are  paler  basally  (5YR  5/ 
10)  and  grade  to  black  near  the  tips;  chin  and 
throat  paler  chestnut  (5YR  6/8)  than  forehead. 

Sides  of  neck  and  breast,  back,  rump,  and 
uppertail  coverts  are  all  approximately  the 
same  rich  dark  chestnut  (5YR  4/8).  The  un- 
derparts appear  very  lightly  dappled,  slightly 
paler  chestnut  (5YR  5/8)  than  upperparts. 
Most  breast  feathers  have  pale  shaft  streaks 
(5YR  7/8)  and  pale  rufous  surrounding  areas, 
some  with  darker  dappling  at  sides,  and  feath- 
ers of  lower  underparts  are  mostly  pale  rufous 
with  vague  darker  scalloping  (2.5YR  5/8);  un- 
dertail coverts  rufescent  whitish  with  the  tips 
scalloped  rufous  (5YR  6/8). 

The  scapulars  have  large  mostly  triangular 
whitish  spots  with  broad  dark  chevron-shaped 
tips  (5YR  4/4).  The  upper  secondary  coverts 
are  almost  uniformly  rufous  (5YR  6/8)  and  the 
upper  primary  coverts  are  darker  (5YR  4/2). 
The  remiges  are  faded,  pale,  and  worn,  in 
striking  contrast  to  the  fresh,  richly  colored 
scapulars.  The  inner  webs  of  the  primaries  and 
narrow  vague  dark  bands  of  the  outer  webs 
are  dark  grayish  brown  (5YR  4/4);  only  the 
outer  webs  have  broader  light  bands  (5YR  7/ 
6).  The  base  color  of  the  secondaries  is  dull 
rufescent  ochraceous  (7.5YR  6/8),  with  fine 
dark  dusky  brown  bars  (7.5YR  4/4).  The  inner 
webs  of  the  undersurfaces  of  the  inner  pri- 
maries and  secondaries  are  basally  pale  rufous 
(7.5  YR  8/6),  as  are  the  uppersurfaces  of  the 
inner  webs  of  the  inner  secondaries,  which 
contrast  strongly  with  the  dark  bands.  The  un- 
derwing coverts  are  solid  pale  rufous  (7.5YR 
7/8).  The  uppertail  surface  has  pale  bands  of 


Rasinussen  • A NEW  SPECIES  OF  HAWK-OWL 


463 


dull  rufous  (SYR  5/6)  that  are  narrow  basally 
and  wider  distally,  and  about  12  narrow  very 
dark  brown  bands  (SYR  3/2)  that  fade  out  to- 
ward the  tip.  There  are  no  definite  bands  for 
the  terminal  20  mm.  The  rectrices  are  heavily 
worn  and  faded. 

The  short,  slender  tarsi  are  completely 
feathered  with  short  pale  cinnamon  (7. SYR  7/ 
6)  pennaceous  feathers  to  about  12  mm  ante- 
riorly (measured  from  joint  of  digits  1-2  of 
middle  toe)  and  posteriorly  to  about  6 mm 
(measured  from  base  of  hallux).  The  toes  ap- 
pear to  have  been  slender,  with  sparse,  short 
rufous  bristles  on  the  tops  and  sides  of  each 
toe.  The  claws  are  small,  delicate,  and  mostly 
blackish  but  with  pale  bases. 

The  soft  part  colors  recorded  on  the  original 
label  are:  eyes  “bright  yellow;  pink  orbital 
skin”,  bill  “ivory”,  feet  “pale  whitish-yel- 
low”. 

Measurements  of  the  holotype  (by  au- 
thor).— Culmen  (from  skull)  17.9  mm;  cul- 
men  (from  distal  edge  of  cere)  10.7  mm;  tar- 
sus 22.6  mm;  wing  172  mm;  tail  97  mm.  Total 
length  of  prepared  specimen  220  mm.  See  Ta- 
ble 1 for  measurements  of  other  characters  of 
the  holotype  and  those  of  other  species. 

Etymology. — This  new  species  is  named 
Ninox  ios  (Greek  for  rust)  for  its  striking  over- 
all coloration.  The  specific  epithet  is  here  used 
as  a noun  in  apposition  to  Ninox,  which,  al- 
though usually  treated  as  feminine,  is  a port- 
manteau combining  Nisus  and  Noctua.  The 
common  name  “cinnabar”  also  refers  to  its 
predominant  color,  which  is  similar  to  that  of 
mercuric  sulfide  before  prolonged  exposure  to 
light. 

DISCUSSION 

Voice. — Not  definitely  known.  Rozendaal 
(Rozendaal  and  Dekker  1989:97)  mentioned 
“disyllabic  calls  ascribed  to  [N.  ochracea]  re- 
corded at  Clark’s  camp  and  on  the  summit  of 
G.[unung]  Muajat  during  April  1985.”  Ek- 
strom  and  co workers  (1998:39)  reported  “an 
unknown  owl  Ninox  sp.”  giving  a series  of 
dry  hoots  rising  and  falling  in  pitch  in  dense 
evergreen  valley  forest  near  the  eastern 
boundary  of  Lore  Lindu  National  Park,  at 
about  1300  m,  in  the  northern  part  of  central 
Sulawesi  (J.  Tobias,  pers.  comm.).  Either  of 
these  reports  might  refer  to  Ninox  ios  but  con- 
firming field  data  are  required. 


Habitat  and  elevation. — Most  researchers 
have  considered  N.  ochracea  to  be  restricted 
to  the  lowlands  below  800  m (Stresemann 
1939,  White  and  Bruce  1986,  Stattersfield  et 
al.  1998).  More  recently,  Coates  and  Bishop 
(1997)  gave  the  elevational  range  of  N.  ochra- 
cea as  up  to  1780  m,  but  this  was  probably 
based  on  the  questionable  vocal  records  men- 
tioned in  Rozendaal  and  Dekker  (1989)  and 
the  collection  of  the  type  of  N.  ios  at  1120  m. 
All  montane  records  of  N.  ochracea  therefore 
require  review  in  light  of  this  new  species. 
Ninox  ios  clearly  occurs  in  sympatry  with,  al- 
though very  likely  at  higher  elevations  than, 
N.  ochracea. 

Molt,  breeding,  and  ecology. — The  holo- 
type of  N.  ios  clearly  had  recently  molted  its 
scapulars,  which  were  bright  and  fresh  and 
contrasted  strikingly  with  the  relatively  dull 
tertials  and  other  flight  feathers.  The  feathers 
on  the  crown  appeared  to  be  worn,  while  those 
of  the  back  appeared  fresh.  Only  10  rectrices 
were  present.  Active  molt  of  the  flight  feathers 
was  not  detected,  but  avoid  damaging  the 
unique  specimen  a thorough  examination  was 
not  attempted.  The  size  of  the  label  drawing 
of  the  largest  testis  (which  measures  6X4 
mm)  suggests  a bird  not  completely  reproduc- 
tively  quiescent.  Because  nothing  is  known  of 
the  habits  of  N.  ios,  it  is  possible  only  to  spec- 
ulate that  its  morphology  (which  recalls  that 
of  owlet-nightjars  Aegothelidae)  suggests  the 
likelihood  of  its  preying  largely  upon  soft- 
bodied  invertebrates  caught  in  flight. 

Systematics. — The  affinities  of  Ninox  ios 
are  unclear;  it  shows  many  morphological  dif- 
ferences from  all  other  species,  particularly  in 
its  small  size,  relatively  long  tail,  narrow 
pointed  wing,  and  weak  tarsi  and  claws.  Al- 
though membership  in  the  polytypic  N.  phi- 
lippensis  superspecies  might  ‘^eem  geograph- 
ically plausible,  the  pattern  ^f  primary  feather 
lengths  shown  by  TV.  ios  is  closer  to  that  of  TV. 
ochracea  than  to  any  form  of  TV.  philippensis. 
Phylogenetic  analyses  will  be  required  to  un- 
derstand the  relationships  of  TV.  ios. 

Conservation. — As  only  one  specimen  is 
known,  it  appears  likely  that  Ninox  ios  has  a 
limited  range  and/or  is  rare.  However,  noctur- 
nal birds  are  frequently  overlooked.  Also, 
most  scientific  bird  collecting  in  Sulawesi 
took  place  before  mist-nets  were  widely  avail- 
able, and  at  lower  elevations.  Ascertaining  its 


464 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


vocalizations  and  calling  periods  will  be  a pre- 
requisite to  carrying  out  effective  surveys, 
which  will  be  essential  to  establish  the  degree 
of  risk  faced  by  this  unique  new  species. 

The  only  other  bird  species  thought  to  be 
restricted  to  North  Sulawesi  is  the  poorly 
known  Matinan  Flycatcher  (Muscicapa  san- 
fordi),  which  has  been  found  only  in  the  Du- 
moga-Bone  and  Tentolo-Matinan  mountains 
between  1400  and  1780  m.  The  fact  that  a 
species  as  distinctive  as  Ninox  ios  could  have 
escaped  description  until  now  clearly  under- 
scores the  fact  that  our  knowledge  of  the  avi- 
fauna of  Sulawesi  is  still  in  a rudimentary 
state. 

ACKNOWLEDGMENTS 

Special  thanks  are  due  R.  W.  R.  J.  Dekker,  M.  S. 
Hoogmoed,  and  H.  van  Grouw,  NNM/Naturalis,  Lei- 
den, and  to  E G.  Rozendaal.  Specimens  of  comparative 
taxa  were  examined  in  museums  too  numerous  to  enu- 
merate individually,  but  chiefly  the  following,  for 
which  I thank  G.  E Barrowclough  and  R Sweet,  Amer- 
ican Museum  of  Natural  History,  New  York;  G.  Hess, 
Delaware  Museum  of  Natural  History,  Wilmington;  D. 
Willard,  Field  Museum  of  Natural  History,  Chicago; 
R.  S.  Kennedy,  Museum  of  Natural  History,  Cincin- 
nati; S.  Prijono  and  Darjono,  Museum  Zoologicum  Bo- 
goriense,  Cibinong,  West  Java;  S.  L.  Olson,  G.  R. 
Graves,  and  B.  M.  McPhelim,  National  Museum  of 
Natural  History,  Smithsonian  Institution,  Washington, 
D.C.;  M.  P.  Walters  and  R.  P.  Prys-Jones,  The  Natural 
History  Museum,  Tring,  U.K.;  G.  Boenigk,  Staatliches 
Naturhistorische  Museum,  Braunschweig,  Germany; 
and  E C.  Sibley,  Yale  Peabody  Museum,  New  Haven. 
The  manuscript  was  improved  by  the  comments  of  R. 
W.  R.  J.  Dekker,  N.  J.  Collar,  B.  M.  Beehler,  and  R. 
Hill;  J.  Tobias  provided  information;  and  the  frontis- 
piece was  painted  by  1.  Lewington. 


LITERATURE  CITED 

Coates,  B.  J.  and  K.  D.  Bishop.  1997.  A guide  to  the 
birds  of  Wallacea.  Alderley,  Queensland,  Austra- 
lia. 

Collar,  N.  J.  and  P.  C.  Rasmussen.  1998.  Species 
limits  in  the  Ninox  philippensis  complex.  Ostrich 
69(3-4);398. 

Dickinson,  E.  C.,  R.  S.  Kennedy,  and  K.  C.  Parkes. 
1991.  The  birds  of  the  Philippines.  British  Orni- 
thologists’ Union,  Tring,  U.K. 

Ekstrom,  j.,  j.  Tobias,  and  J.  Robinson-Dean.  1998. 
Forests  at  the  edge  of  Lore  Lindu  National  Park, 
central  Sulawesi.  Oriental  Bird  Club  Bull.  28:36— 
39. 

Johnstone,  R.  and  J.  Darnell.  1997.  Description  of 
a new  subspecies  of  Boobook  Owl  Ninox  novae- 
seelandiae  (Gmelin)  from  Roti  Island,  Indonesia. 
West.  Aust.  Nat.  21:161—173. 

Meyer,  A.  B.  and  L.  W.  Wiglesworth.  1898.  The 
birds  of  Celebes  and  the  neighbouring  islands. 
Vol.  1.  R.  Friedlander  & Sohn,  Berlin,  Germany. 

Munsell.  1977.  Munsell  color  charts  for  plant  tissues. 
Gretagmacbeth,  New  Windsor,  New  York. 

Norman,  J.  A.,  L.  Christidis,  M.  Westerman,  and  E 
A.  R.  Hill.  1998.  Molecular  data  confirms  the 
specific  status  of  the  Christmas  Island  Hawk-Owl 
Ninox  natalis.  Emu  98:197—208. 

Rozendaal,  E G.  and  R.  W.  R.  J.  Dekker.  1989.  An- 
notated checklist  of  the  birds  of  the  Dumoga-Bone 
National  Park,  North  Sulawesi.  Kukila  4:85-109. 

Stattersfield,  a.  j.,  M.  J.  Crosby,  A.  J.  Long,  and 
D.  C.  Wege.  1998.  Endemic  bird  areas  of  the 
world.  BirdLife  International,  Cambridge,  U.K. 

Stresemann,  E.  1939.  Die  Vogel  von  Celebes.  Part  1. 
J.  Ornithol.  87:299-425. 

Stresemann,  E.  1940.  Die  Vogel  von  Celebes.  Part  2. 
J.  Ornithol.  88:389-487. 

White,  C.  M.  N.  and  M.  D.  Bruce.  1986.  The  birds 
of  Wallacea.  British  Ornithologists’  Union,  Lon- 
don, U.K. 


Wilson  Hull.,  1 1 1(4),  1999.  pp.  465-471 


PATTERNS  OF  VARIATION  IN  SIZE  AND  COMPOSITION  OF 
GREATER  SCAUP  EGGS:  ARE  THEY  RELATED? 

PAUL  L.  FLINT'  3 AND  J.  BARRY  GRAND'  ^ 


ABSTRACT. — We  studied  egg  size  variation  of  Greater  Scaup  (Aythya  mcirila)  nesting  on  the  Yukon-Kus- 
kokwini  Delta,  Alaska  from  1991  — 1996.  Mean  egg  size  was  64.36  ± 0.03  (SE)  ml.  Egg  size  did  not  vary  with 
clutch  size  or  serve  as  an  index  of  body  size.  There  was  less  than  2%  overlap  in  total  clutch  volumes  for  clutches 
of  different  sizes  indicating  that  phenotypic  clutch  size-egg  size  trade-offs  are  not  occurring  among  individuals. 
At  the  population  level.  Greater  Scaup  have  less  variation  in  egg  size  than  other  species  of  waterfowl.  The 
proportion  of  variation  in  egg  size  caused  by  differences  among  females  was  0.20,  caused  by  differences  within 
females  among  years  was  0.25,  and  caused  by  differences  within  females  and  years  (i.e.,  clutches)  was  0.56. 
The  proportion  of  egg  lipid  decreased  with  increasing  egg  size  while  the  proportion  of  egg  protein  increased 
with  egg  size.  Thus,  Greater  Scaup  appear  to  trade-off  lipid  for  protein  as  egg  size  increases.  The  proportion  of 
variation  that  was  due  to  differences  among  females  in  total  egg  protein  was  0.79  and  in  total  egg  lipid  was 
0.49.  We  conclude  that  in  the  absence  of  a fitness  trade-off  between  clutch  size  and  egg  size,  selection  has 
reduced  among-individual  variation  in  egg  size.  Received  16  April  1999,  accepted  4 August  1999. 


Lack  (1967)  suggested  that  trade-offs  may 
occur  between  the  number  and  size  of  eggs 
produced  by  waterfowl  species.  Rohwer 
(1988)  argued  that  the  same  trade-off  between 
number  and  size  of  eggs  should  occur  within 
species  as  well.  The  high  repeatability  of  egg 
size  (i.e.,  volume  or  weight)  generally  found 
in  waterfowl  suggests  that  such  trade-offs 
likely  occur  among  rather  than  within  individ- 
uals (Lessells  et  al.  1989,  Larsson  and  Fors- 
lund  1992).  Thus,  the  concept  of  a clutch 
size-egg  size  trade-off  implies  that  females 
laying  small  clutches  of  large  eggs  and  fe- 
males laying  large  clutches  of  small  eggs  have 
equal  fitness.  Accordingly,  for  a trade-off  to 
exist  both  clutch  size  and  egg  size  must  be 
positively  related  to  fitness. 

Waterfowl  laying  larger  clutches  may  be 
more  fit  because  they  tend  to  fledge  more 
young  (Lessells  1986,  Rockwell  et  al.  1987, 
Flint  1993).  For  example,  Rockwell  and  co- 
workers (1987)  demonstrated  that  female 
Lesser  Snow  Geese  (Chen  caerulescens  cae- 
rulescens)  laying  larger  clutches  recruited 
more  young  than  females  laying  smaller 
clutches.  Fitness  may  also  be  related  to  egg 
size  in  .some  species  of  waterfowl  because 


' Alaska  Biological  Science  Center,  U..S.  Geolog- 
ical Survey,  1011  East  Tudor  Rd.,  Anchorage,  AK 
99503. 

- Present  address:  Alabama  Cooperative  Fisheries 
and  Wildlife  Research  Unit,  331  Funchess  Hall,  Au- 
burn University,  Auburn,  AL  36849. 

^ Corresponding  author;  E-mail;  pauLflint(§>usgs.gov 


young  from  larger  eggs  are  better  able  to  sur- 
vive extreme  conditions  (Ankney  1980, 
Rhymer  1988,  Thomas  and  Brown  1988).  For 
example,  Dawson  and  Clark  (1996)  found  that 
Lesser  Scaup  {Aythya  affinis)  ducklings  from 
large  eggs  survived  better  than  those  from 
small  eggs  under  natural  conditions  (but  see 
Williams  et  al.  1993).  The  mechanism  by 
which  larger  eggs  yield  higher  juvenile  sur- 
vival may  be  related  to  egg  composition 
(Dawson  and  Clark  1996).  Egg  composition, 
in  terms  of  lipid  and  protein,  typically  varies 
isometrically  with  egg  weight  for  waterfowl 
species  (Ankney  1980,  Hepp  et  al.  1987, 
Owen  and  West  1988,  Hill  1995,  Slattery  and 
Alisauskas  1995);  however,  some  species 
show  a proportional  increase  in  lipid  with  in- 
creasing egg  size  (Birkhead  1984,  1985;  Ali- 
sauskas 1986;  Rohwer  1986;  Williams  1994). 
In  either  case,  young  hatching  from  large  eggs 
tend  to  be  larger  at  hatching  and  have  abso- 
lutely larger  reserves  than  young  hatching 
from  smaller  eggs  (Ankney  1980,  Slattery  and 
Alisauskas  1995,  Erikstad  et  al.  1998). 

Fitness  trade-offs  are  potential  mechanisms 
maintaining  heritable  variation  in  both  egg 
size  and  composition  at  the  population  level 
(Falconer  1989).  Egg  size  has  been  shown  to 
be,  at  least  partially,  under  genetic  control, 
and  heritable  genetic  variation  has  been  found 
in  several  species  of  waterfowl  (Batt  and 
Prince  1978,  Lessells  et  al.  1989,  Larsson  and 
Forslund  1992).  Thus,  in  the  absence  of  fitness 
trade-offs,  mean  egg  size  for  a population 


465 


466 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


would  be  expected  to  respond  to  directional 
selection.  Therefore,  we  predict  that  patterns 
of  egg  size  variation  within  and  among  pop- 
ulations are  linked  to  clutch  size-egg  size 
trade-offs.  Additionally,  if  clutch  size-egg 
size  trade-offs  exist,  then  the  corresponding 
variation  in  egg  composition  can  be  used  to 
infer  the  required  relationship  between  egg 
size  and  fitness.  It  follows  then,  that  patterns 
of  variation  in  egg  composition  within  popu- 
lations are  also  influenced  by  clutch  size-egg 
size  trade-offs. 

Our  goal  in  this  study  was  to  describe  egg 
size  variation  in  Greater  Scaup  {Aythya  mar- 
ilci)  with  regard  to  female  body  size  and  clutch 
size.  Further,  we  looked  for  evidence  of  clutch 
size-egg  size  trade-offs  among  females.  We 
examined  variation  in  egg  composition  within 
and  among  females  and  in  relation  to  egg  size 
and  compared  these  results  to  other  species. 

METHODS 

This  study  was  conducted  along  the  lower  Kashunuk 
River  drainage  (61°  20'  N,  165°  35'  W)  on  the  outer 
coastal  fringe  of  the  Yukon  Delta  National  Wildlife 
Refuge,  Alaska.  This  study  was  conducted  under  ap- 
propriate Federal  (PRT-692350),  state  (93-69),  and  ref- 
uge special  use  collection  permits.  The  study  area  con- 
sists of  relatively  flat  sedge  meadows  and  numerous 
ponds  (Flint  and  Grand  1996a,  Grand  et  al.  1997).  Nest 
searches  were  conducted  from  mid-May  through  mid- 
July  (see  Flint  and  Grand  1996a).  When  nests  were 
discovered,  the  number  of  eggs,  nest  location,  and  date 
were  recorded.  Eggs  were  individually  numbered  and 
candled  to  determine  the  stage  of  embryonic  devel- 
opment (Flint  and  Grand  1996a).  Nests  were  revisited 
at  7 day  intervals  and  the  number  of  eggs  and  stage 
of  development  of  embryos  were  recorded.  Maximum 
lengths  and  breadths  of  all  eggs  were  measured  to  the 
nearest  0. 1 mm.  Clutch  size  was  defined  as  the  number 
of  eggs  known  to  have  been  laid  into  a nest  for  nests 
found  during  egg  laying  and  as  the  number  of  eggs  in 
the  nest  at  time  of  discovery  for  nests  found  during 
incubation. 

In  1994-1996  a sample  of  hens  was  captured  on  the 
nest  at  hatching  using  bow  traps  (Flint  and  Grand 
1996b).  Hens  were  weighed  to  the  nearest  10  grams 
and  culmen  and  total  tarsus  lengths  were  measured  to 
the  nearest  0.1  mm.  Females  were  marked  with  alu- 
minum U.S.  Fi.sh  and  Wildlife  Service  tarsal  bands. 

In  1996,  we  collected  a sample  of  30  complete 
clutches  at  4 days  of  incubation  in  conjunction  with  a 
separate  study  of  renesting  ecology.  These  eggs  were 
weighed,  measured  (length  and  breadth)  with  calipers, 
and  we  measured  the  external  volume  by  submerging 
the  egg  and  measuring  the  displacement  ol  water  to 
the  nearest  0.5  ml.  Variables  in  the  relationship  be- 
tween volume  and  linear  measures  were  estimated  us- 


ing analysis  of  covariance  with  volume  as  the  depen- 
dent variable,  females  as  a factor,  and  length  X 
breadth^  as  the  single  covariate  (Hoyt  1979,  Flint  and 
Sedinger  1992,  Flint  and  Grand  1996b).  We  included 
the  interaction  between  females  and  length  X breadth^ 
to  compare  the  slope  of  this  relationship  among  fe- 
males. The  fitted  relationship  was  used  to  predict  egg 
volume  from  linear  measures  for  eggs  where  volume 
was  not  measured  directly  (i.e.,  eggs  measured  in  the 
field). 

In  1993,  a sample  of  nests  was  visited  every  other 
day  after  7 eggs  had  been  laid  to  determine  the  date 
of  clutch  completion.  We  collected  a sample  of  88  eggs 
from  1 1 clutches  as  soon  as  laying  was  complete. 
These  eggs  were  boiled  for  15  min  and  frozen.  Each 
egg  was  later  thawed  and  weighed  whole.  Because  en- 
tire clutches  were  collected  at  the  end  of  laying  and 
some  incubation  occurred  during  laying,  vasculariza- 
tion or  embryonic  development  was  noted  as  being 
present  during  dissection.  The  egg  was  separated  into 
shell,  albumen,  and  yolk  and  these  components  were 
dried  to  a constant  weight  at  approximately  80°  C.  Fat 
was  extracted  from  the  yolk  using  petroleum  ether 
(Dobush  et  al.  1985)  in  a Soxhlet  apparatus.  Total  yolk 
protein  was  estimated  as  whole  yolk  dry  weight  minus 
yolk  fat.  We  assumed  that  dried  albumen  weight  was 
composed  entirely  of  protein  (Montevecchi  et  al.  1983) 
and  estimated  total  egg  protein  as  total  yolk  protein 
plus  albumen  dry  weight. 

We  examined  variation  in  egg  size  with  clutch  size 
using  a nested  ANOVA  with  clutches  nested  within 
clutch  sizes.  We  used  the  mean  square  error  among 
clutches  with  clutch  sizes  as  the  denominator  in  the  F- 
test  of  clutch  size  effects.  Total  clutch  volume  was 
calculated  as  the  sum  of  the  individual  egg  volumes 
within  a clutch  for  the  sample  of  nests  that  survived 
to  incubation.  We  used  the  sum  of  the  log  transformed 
measurements  of  culmen  and  tarsus  as  an  index  of 
structural  size  and  examined  variation  in  mean  egg  size 
in  relation  to  this  index  for  the  sample  of  nests  from 
which  we  captured  females  using  linear  regression. 
The  proportion  of  variation  in  egg  size  caused  by  dif- 
ferences within  clutches,  within  females  among  years, 
and  among  females  (i.e.,  repeatability)  were  calculated 
using  a nested  ANOVA  and  modifying  the  methods  of 
Lessells  and  Boag  (1987)  for  a nested  design  (Sokal 
and  Rohlf  1981). 

We  examined  variation  in  both  total  egg  lipid  and 
protein  (separately)  using  ANCOVA  with  female  (i.e., 
clutch)  and  egg  development  as  factors  and  egg  size 
as  a covariate.  We  also  included  an  interaction  between 
female  and  egg  size.  To  examine  allometric  relation- 
ships of  egg  components  we  used  log|„-logio  regression 
of  egg  components  against  egg  weight;  isometry  was 
concluded  if  the  slope  of  these  relationships  was  not 
different  from  1.  Repeatability  of  egg  components  was 
e.stimated  using  a nested  design  with  eggs  nested  with- 
in females.  All  analyses  were  conducted  using  SAS 
version  6.12  (SAS  Institute  1990). 


Flint  iind  Grand  • GREATER  SCAUP  EGG  SIZE  AND  COMPOSITION 


467 


(f) 

<u 

o 

_3 

o 


(U 

E 

D 


20 

15 

10 

5 

0 

250  300  350  400  450  500  550  600  650  700  750 


Total  dutch  volume  (ml) 


FIG.  1.  Total  clutch  volume  for  Greater  Scaup  clutches  of  different  sizes  from  the  Yukon-Kuskokwim  Delta 
1991-1996.  Clutch  volumes  were  rounded  to  the  nearest  5 ml.  Overall  only  3 of  235  (1.3%)  clutch  volumes 
(i.e.,  two  9 egg  clutches  and  one  10  egg  clutch)  occurred  within  the  volume  distribution  of  another  clutch  size. 
Other  stacked  bars  represent  tied  values  resulting  from  rounding  of  total  clutch  volumes  where  actual  estimates 
of  total  clutch  volume  did  not  overlap.  For  example,  a 7 egg  clutch  with  a volume  of  467  ml  is  stacked  with 
an  8 egg  clutch  of  468  ml  because  rounding  is  required  to  develop  the  distributions. 


RESULTS 

We  measured  length,  breadth,  weight,  and 
volume  of  271  eggs.  The  equation: 

Volume  = 31.84  + 0.2729  X length 

X breadth^,  ( 1 ) 

described  the  relationship  between  displace- 
ment (ml)  and  linear  egg  measurements  (cm) 
(r2  = 0.44,  F,  269  = 209.10,  P < 0.001).  To 
compare  variation  in  predicted  egg  volumes 
with  variation  in  measured  volumes  we  cal- 
culated the  repeatability  of  measured  egg  vol- 
umes (33.4%)  and  found  it  similar  to  the  re- 
peatability of  estimated  egg  volumes  (36.4%). 

We  measured  a total  of  3937  eggs  in  the 
field.  Mean  egg  length  was  6.352  ± 0.003 
(SE)  cm  and  egg  breadth  was  4.328  ± 0.002 
cm.  Mean  estimated  egg  size  was  64.36  ± 
0.03  ml  (CV  = 0.03).  We  had  no  data  on  var- 
iation in  egg  size  with  laying  sequence,  but 
the  average  range  of  egg  size  within  clutches 
was  3.3  ml. 

Egg  size  did  not  vary  with  clutch  size 
(^12.470  “ 0.74,  P > 0.05).  For  the  analyses  of 
the  relationship  between  egg  size  and  body 
size  we  used  93  observations  of  individual  fe- 
males captured  on  nests.  Average  egg  size  per 
clutch  was  not  related  to  our  index  of  body 
size  (F|  9o  = 0.35,  P > 0.05).  Overall  only  3 


of  235  (1.3%)  clutch  volumes  (i.e.,  two  9-egg 
clutches  and  one  10-egg  clutch)  occurred 
within  the  volume  distribution  of  another 
clutch  size  (Fig.  1). 

For  estimates  of  repeatability  of  egg  size  we 
captured  20  females  41  times  (only  1 female 
was  captured  3 times).  The  proportion  of  var- 
iation in  egg  size  attributed  to  differences 
among  females  (i.e.,  repeatability)  was  0.20, 
proportion  of  variation  attributed  to  differenc- 
es within  females  among  years  was  0.25,  and 
proportion  of  variation  attributed  to  differenc- 
es within  females  and  years  (i.e.,  clutches) 
was  0.56.  Using  a standard  approach  of  cal- 
culating the  mean  egg  size  per  clutch  and 
treating  clutches  as  individual  observations 
within  females  (e.g.,  Flint  and  Grand  1996b) 
yields  a repeatability  of  mean  egg  size  of  0.36. 

Total  egg  protein  varied  with  egg  size  m 
= 67.96,  P < 0.001),  but  not  among  females 
(^9,64  ~ 1.55,  P > 0.05).  The  relationship  be- 
tween egg  protein  and  size  also  did  not  vary 
among  females  (i.e,  no  interaction,  ^9^,4  = 
1.59,  P > 0.05).  Total  egg  protein  was  not 
significantly  influenced  by  early  embryonic 
development  ^ = 3.23,  P > 0.05).  Simi- 
larly, total  egg  lipid  varied  with  egg  size  (F,  (,4 
= 65.80,  P < 0.001),  but  not  among  females 
(^9.64  ~ 1.98,  P = 0.055).  However,  the  rela- 


468 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


TABLE  1.  Allometric  relationships  between  total  eg 

;g  lipid,  protein  and  water 

in  relation  to  egg  weight. 

Allometric  relationship 

r~ 

Slope 

ECU' 

UCL-’ 

r'' 

p 

Weight  and  egg  lipid'-' 

0.62 

0.85 

0.71 

0.99 

2.09 

0.04 

Weight  and  egg  protein‘s 

0.84 

1.11 

1.01 

1.21 

2.16 

0.03 

Weight  and  egg  water 

0.95 

1.01 

0.96 

1.06 

0.52 

>0.05 

“ Upper  and  lower  95%  confidence  limits  for  slope. 
'’Test  of  null  hypothesis  that  slope  = 1.0. 

Egg  lipid  and  protein  are  in  dry  weight. 


tionship  between  total  egg  lipid  and  egg  size 
varied  among  females  (i.e.,  interaction;  = 
2.04,  P = 0.049).  Total  egg  lipid  was  not  in- 
fluenced by  early  embryonic  development 
(F,  ,4  = 0.17,  P > 0.05). 

The  proportion  of  lipid  in  an  egg  declined 
with  increasing  egg  weight,  whereas  the  pro- 
portion of  protein  increased  with  increasing 
egg  weight  (Table  1).  Among-female  differ- 
ences explained  most  of  the  variation  in  total 
egg  protein;  whereas  within-female  (i.e.,  with- 
in clutch)  differences  explained  most  of  the 
variation  in  egg  lipid  components  (Table  2). 
Even  with  the  variation  in  egg  composition 
within  and  among  females,  total  clutch  vol- 
ume still  predicted  most  of  the  variation  in 
total  protein  {F  — 0.86,  F,  ^ = 48.71,  P < 
0.001)  and  lipid  (r-  = 0.93,  F,  g = 100.05,  P 
< 0.001)  investment  in  a clutch. 

DISCUSSION 

Egg  size. — The  relationship  between  length 
X width^  and  measured  volumes  fit  poorly  (F 
= 0.44)  for  Greater  Scaup.  For  other  species 
of  waterfowl  these  relationships  have  shown 
a high  F (>90%;  Flint  and  Sedinger  1992, 
Flint  and  Grand  1996b,  Slattery  and  Alisaus- 


TABLE  2.  Average  egg  components  dry  weight 
and  proportion  of  variation  due  to  differences  among 
females. 


Among 

female 

variance 

component 

Variable 

X 

SE 

(%) 

Albumen  (g) 

4.97 

0.046 

87.9 

Shell  (g) 

5.23 

0.032 

73.2 

Yolk  (g) 

14.84 

0. 1 09 

52.4 

Total  egg  lipid  (g) 

9.97 

0.069 

48.8 

Total  egg  protein  (g) 

9.89 

0.077 

79.1 

Lipid  per  mi 

0.15 

0.001 

36.2 

Protein  per  ml 

0. 1 5 

0.001 

79.9 

Density  (g/ml) 

0.91 

0.003 

67.9 

kas  1995).  We  believe  this  relatively  poor  fit 
is  the  result  of  the  low  coefficient  of  variation 
in  egg  size  (i.e.,  lack  of  variation  in  the  y- 
axis).  Additionally,  we  likely  detected  varia- 
tion in  slope  among  females  because  we  sam- 
pled entire  clutches  and  thus  had  relatively 
large  samples  within  females.  Other  studies 
have  failed  to  find  variation  in  slope  of  pre- 
dictive relationships  between  egg  measures 
and  egg  size  among  females  but  only  sampled 
a few  eggs  from  each  clutch  (Flint  and  Sedin- 
ger 1992,  Flint  and  Grand  1996b).  We  found 
little  difference  between  estimates  of  repeat- 
ability of  predicted  volumes  and  measured 
volumes.  The  comparability  of  these  two  re- 
sults indicates  that  error  associated  with  pre- 
dicting egg  volumes  is  random  within  and 
among  females. 

We  found  less  variation  in  egg  size  at  the 
population  level  than  has  been  reported  for 
other  species  of  waterfowl.  The  coefficient  of 
variation  of  average  egg  size  for  Greater 
Scaup  (3%)  was  substantially  less  than  for 
Northern  Pintails  {Anas  acuta,  11%;  Flint  and 
Grand  1996b),  Blue- winged  Teal  {Anas  dis- 
cors,  9%;  Rohwer  1986),  Canada  Geese 
{Branta  canadensis,  7—9%;  Cooper  1978, 
Leblanc  1989),  Black  Brant  {Branta  bernicla 
nigricans,  8%;  Flint  and  Sedinger  1992),  or 
Lesser  Snow  Geese  (8%;  Cooke  et  al.  1995). 
Thus,  it  appears  selection  has  reduced  the  var- 
iation in  egg  size  at  the  population  level  com- 
pared to  other  species. 

The  repeatability  of  egg  size  that  we  mea- 
sured among  females  is  substantially  lower 
than  reported  for  other  waterfowl  (Batt  and 
Prince  1978,  Duncan  1987,  Lessells  et  al. 
1989,  Flint  and  Sedinger  1992,  Flint  and 
Grand  1996b).  When  we  partitioned  variance 
using  a nested  ANOVA  design,  including  var- 
iation within  clutches,  we  found  the  among 
female  component  of  the  variance  was  even 
lower.  Using  an  approach  similar  to  ours. 


Flint  and  Grand  • GREATER  SCAUP  EGG  SIZE  AND  COMPOSITION 


469 


Leblanc  (1989)  reported  that  the  proportions 
of  variation  among  female  Canada  Geese, 
within  females-among  years,  and  within 
clutches  were  0.62,  0.05,  and  0.33,  respec- 
tively. Thus,  the  patterns  of  egg  size  variation 
within  and  among  female  Greater  Scaup  are 
substantially  different  than  those  reported  for 
other  waterfowl.  The  average  range  of  egg 
sizes  within  clutches  was  similar  to  the  range 
of  egg  sizes  reported  for  other  waterfowl 
(Leblanc  1987,  Owen  and  West  1988,  Flint 
and  Sedinger  1992,  Robertson  and  Cooke 
1993,  Cooke  et  al.  1995).  Thus,  the  low  re- 
peatability we  report  is  not  the  result  of  rela- 
tively high  variation  within  females  (i.e., 
clutches),  but  is  strongly  influenced  by  the 
lack  of  variation  in  egg  size  among  females. 
Relative  to  other  waterfowl  species,  it  appears 
selection  has  resulted  in  reduced  egg  size  var- 
iation among  female  Greater  Scaup  and  main- 
tained variation  within  clutches  that  may  be 
related  to  sequence  (Flint  and  Sedinger  1992, 
Cooke  et  al.  1995). 

Mean  egg  size  did  not  vary  with  clutch  size 
in  this  study.  This  is  consistent  with  findings 
for  other  waterfowl  (Duncan  1987,  Rohwer 
1988,  Rohwer  and  Eisenhauer  1989,  Flint  and 
Sedinger  1992,  Flint  and  Grand  1996b). 
Therefore,  we  did  not  detect  a trade-off  be- 
tween clutch  size  and  egg  size  predicted  under 
the  nutrient  limitation  hypothesis  and  a simple 
model  of  nutrient  allocation  to  eggs  (Flint  et 
al.  1996).  Additionally,  mean  egg  size  was  not 
related  to  body  size  and  the  alternative  allo- 
cation model  described  by  Flint  and  cowork- 
ers (1996)  cannot  explain  the  failure  to  detect 
a negative  relationship  between  clutch  size 
and  egg  size.  Finally,  the  lack  of  significant 
overlap  in  total  clutch  volumes  for  clutches  of 
different  size,  indicated  phenotypic  trade-offs 
between  clutch  size  and  egg  size  among  fe- 
males with  equal  investments  in  their  clutches 
did  not  occur  (Flint  and  Sedinger  1992,  Flint 
and  Grand  1996b).  Thus,  we  find  no  evidence 
of  phenotypic  clutch  size-egg  size  trade-offs 
among  individuals  for  Greater  Scaup. 

Ankney  and  Bissett  (1976)  proposed  that 
egg  size  variation  in  a population  was  main- 
tained by  annual  variation  in  environmental 
conditions  that  caused  annual  variation  in  op- 
timal egg  size.  While  not  explicitly  stated,  the 
concept  of  a clutch  size-egg  size  trade-off  is 
inherent  in  this  hypothesis.  In  their  example. 


Ankney  and  Bissett  (1976)  state  that  the  ad- 
vantage to  females  laying  small  eggs  is  that 
they  can  produce  more  eggs  from  finite  re- 
serves than  females  laying  large  eggs.  We 
found  considerably  less  egg  size  variation  at 
the  population  level  than  has  been  reported  for 
other  species.  Correspondingly,  we  also  found 
no  evidence  of  a clutch  size— egg  size  trade- 
off whereas  studies  of  other  species  have 
shown  some  evidence  of  a trade-off  among 
individuals  with  equal  investments  in  their 
clutches  (Ankney  and  Bissett  1976,  Flint  and 
Sedinger  1992,  Flint  and  Grand  1996b).  Fur- 
ther, repeatability  sets  the  upper  limit  to  her- 
itability  (Falconer  1989);  thus,  the  low  re- 
peatability we  measured  implies  little  genetic 
variability  for  egg  size  in  our  study  popula- 
tion. Traits  influenced  by  selection  are  ex- 
pected to  approach  fixation  and  therefore  have 
low  heritabilities  (Falconer  1989).  We  suggest 
for  Greater  Scaup  that  clutch  size-egg  size 
trade-offs  do  not  occur  and  selection  has  re- 
duced variability  in  egg  size. 

Egg  composition. — Because  female  water- 
fowl  commonly  begin  incubation  before  egg 
laying  is  complete  (Flint  et  al.  1994),  and  we 
did  not  collect  eggs  until  egg  laying  was  ter- 
minated, some  eggs  had  slight  embryonic  de- 
velopment at  the  time  of  collection.  The  pres- 
ence of  egg  development  was  not  related  to 
either  total  egg  lipid  or  protein  after  control- 
ling for  variation  among  females,  egg  volume, 
and  an  interaction  between  females  and  egg 
volume.  Thus,  we  do  not  believe  that  the  early 
development  observed  in  some  of  our  col- 
lected eggs  influenced  our  results. 

The  relationship  between  egg  lipid  and  egg 
size  for  Greater  Scaup  varied  among  females. 
This  is  similar  to  what  Rohwer  (1986)  found 
for  Blue-winged  Teal.  Birkhead  (1985)  re- 
ported greater  variation  in  egg  composition 
among  females  than  within  clutches  for  Mal- 
lards {Anas  platyrhynchos).  Further,  Alisaus- 
kas  (1986)  found  that  egg  lipid  was  highly 
variable  among  female  American  Coots  {Fu- 
Uca  americana)  and  varied  within  clutches  in 
relation  to  egg  sequence.  Hepp  and  coworkers 
(1987)  found  that  about  half  of  the  variation 
in  egg  lipid  in  Wood  Ducks  {Aix  sponsa)  was 
due  to  differences  within  females,  similar  to 
our  results.  Varying  egg  composition  may  al- 
low females  to  slightly  adjust  the  total  invest- 
ment in  the  clutch  without  altering  clutch  size 


470 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  4,  December  1999 


or  egg  size  (Owen  and  West  1988).  The  con- 
sistent finding  that  egg  lipid  varied  within  fe- 
males suggests  that  there  may  be  some  adap- 
tive partitioning  of  resources  within  clutches, 
perhaps  related  to  egg  laying  sequence.  How- 
ever, even  with  the  variation  in  egg  compo- 
sition described  above,  total  clutch  volume  ex- 
plained more  than  85%  of  the  variation  in  to- 
tal lipid  and  protein  investment  in  a clutch. 
Thus,  contrary  to  the  results  of  Flint  and 
Grand  (1996b)  for  Northern  Pintails,  clutch 
size  is  a good  predictor  of  nutrient  investment 
in  Greater  Scaup  clutches. 

Heavier  eggs  had  proportionally  more  pro- 
tein and  less  lipid  than  lighter  eggs.  Thus,  it 
appears  that  Greater  Scaup  trade  off  lipid  for 
protein  as  egg  weight  increases.  These  results 
differ  from  studies  of  waterfowl  that  show  a 
proportional  increase  in  lipid  with  egg  size 
(Birkhead  1984,  1985;  Alisauskas  1986;  Roh- 
wer  1986;  Williams  1994),  and  differ  from 
studies  of  other  precocial  species  where  egg 
protein  and  lipid  both  increase  proportionately 
with  egg  weight  (Ankney  1980,  Hepp  et  al. 
1987,  Owen  and  West  1988,  Hill  1995,  Slat- 
tery and  Alisauskas  1995).  Ankney  and  Bis- 
sett  (1976)  argued  that  because  egg  yolk,  and 
hence  egg  lipid,  increased  with  egg  size, 
young  hatching  from  larger  eggs  survived  bet- 
ter because  they  had  absolutely  larger  lipid  re- 
serves. However,  young  from  larger  eggs  tend 
to  be  larger  at  hatch  (Ankney  1980,  Slattery 
and  Alisauskas  1995,  Erikstad  et  al.  1998)  and 
thus  will  also  have  absolutely  higher  energy 
requirements  (Rhymer  1988).  For  Greater 
Scaup,  the  proportion  of  egg  lipid  decreased 
with  increasing  egg  size.  Therefore,  we  would 
not  expect  large  egg  size  to  confer  a survival 
advantage  to  offspring  in  our  study  popula- 
tion. 

Given  our  conclusion  that  clutch  size-egg 
size  trade-offs  do  not  occur,  and  egg  size  has 
been  optimized  by  selection  for  our  popula- 
tion, we  would  expect  no  relationship  between 
egg  size  and  fitness  for  our  study  population. 
The  proportional  decline  in  egg  lipid  with  in- 
creasing egg  size  fits  this  prediction.  There- 
fore, we  conclude  that  clutch  size-egg  size 
trade-offs  likely  influence  both  egg  size  vari- 
ation and  patterns  of  egg  composition  within 
species.  If  this  hypothesis  is  correct,  we  would 
predict  that  species  showing  evidence  ol  phe- 
notypic trade-offs  between  clutch  size  and  egg 


size  (e.g..  Northern  Pintails;  Flint  and  Grand 
1996b)  would  also  have  patterns  of  variation 
in  egg  composition  consistent  with  the  expec- 
tation that  offspring  from  larger  eggs  have  a 
survival  advantage. 

ACKNOWLEDGMENTS 

This  project  was  funded  by  U.S.  Fish  and  Wildlife 
Service,  Office  of  Migratory  Bird  Management,  Re- 
gion 7,  and  the  Alaska  Biological  Science  Center.  We 
thank  D.  Derksen,  R.  Oates,  and  R.  Leedy  for  their 
support,  and  J.  Morgart  and  the  staff  of  the  Yukon 
Delta  National  Wildlife  Refuge  for  logistical  support. 
We  acknowledge  the  many  people  who  helped  search 
for  nests  and  measure  eggs  over  the  years.  We  thank 
S.  Lee  for  conducting  nutrient  composition  analysis  of 
eggs  and  R.  Rockwell  and  C.  Babcock  for  helping  us 
understand  repeatability  analysis.  We  thank  T.  Arnold, 
D.  Derksen,  K.  Lessells,  M.  MacCluskie,  and  J.  Se- 
dinger  for  reviewing  earlier  drafts  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. 

Ankney,  C.  D.  1980.  Egg  weight,  survival,  and  growth 
of  Lesser  Snow  Goose  goslings.  J.  Wildl.  Manage. 
44:174-182. 

Ankney,  C.  D.  and  A.  R.  Bissett.  1976.  An  expla- 
nation of  egg-weight  variation  in  the  Lesser  Snow 
Goose.  J.  Wildl.  Manage.  40:729-734. 

Batt,  B.  D.  j.  and  H.  H.  Prince.  1978.  Some  repro- 
ductive parameters  of  Mallards  in  relation  to  age, 
captivity,  and  geographic  origin.  J.  Wildl.  Man- 
age. 42:834-842. 

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. 

Cooke,  E,  R.  E Rockwell,  and  D.  B.  Lank.  1995. 
The  Snow  Geese  of  La  Perouse  Bay:  natural  se- 
lection in  the  wild.  Oxford  Univ.  Press,  New  York. 
Cooper,  J.  A.  1978.  The  history  and  breeding  biology 
of  the  Canada  Goose  of  Marshy  Point,  Manitoba. 
Wildl.  Monogr.  61:1-87. 

DAW.SON,  R.  D.  AND  R.  G.  Clark.  1996.  Effects  of 
variation  in  egg  size  and  hatching  date  on  survival 
of  Lesser  Scaup  Aythya  qffinis  ducklings.  Ibis  1 38: 
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. 
Can.  J.  Zool.  63:1917-1920. 

Duncan,  D.  C.  1987.  Variation  and  heritability  in  egg 
size  of  Northern  Pintails.  Can.  J.  Zool.  65:992- 
996. 

Erikstad,  K.  E.,  T.  Tveraa,  and  J.  O.  Bustnes.  1998. 


Flint  and  Grand  • GREATER  SCAUP  EGG  SIZE  AND  COMPOSITION 


471 


Significance  ot  intraclutch  egg-size  variation  in 
Common  Eider:  the  role  of  egg  size  and  quality 
of  ducklings.  J.  Avian  Biol.  29:3-9. 

Falconer,  D.  S.  1989.  Introduction  to  quantitative  ge- 
netics. Third  ed.  John  Wiley  and  Sons,  Inc.,  New 
York. 

Flint,  P.  L.  and  J.  S.  Sedinger.  1992.  Reproductive 
implications  of  egg-size  variation  in  the  Black 
Brant.  Auk  109:896-903. 

Flint,  P.  L.  1993.  Pre fledging  survival  and  reproduc- 
tive survival  in  Black  Brant.  Ph.D.  diss.,  Univ.  of 
Alaska,  Fairbanks. 

Flint,  P.  L.,  M.  S.  Lindberg,  M.  C.  MacCluskie,  and 
J.  S.  Sedinger.  1994.  The  adaptive  significance  of 
hatching  synchrony  of  waterfowl  eggs.  Wildfowl 
45:248-254. 

Flint,  P.  L.  and  J.  B.  Grand.  1996a.  Nesting  success 
of  Northern  Pintails  on  the  Coastal  Yukon-Kus- 
kokwim  Delta.  Condor  98:54-60. 

Flint,  P.  L.  and  J.  B.  Grand.  1996b.  Variation  in  egg 
size  of  the  Northern  Pintail.  Condor  98:162-165. 

Flint,  P.  L.,  J.  B.  Grand,  and  J.  S.  Sedinger.  1996. 
Allocation  of  limited  reserves  to  a clutch:  a model 
explaining  the  lack  of  a relationship  between 
clutch  size  and  egg  size.  Auk  113:939-942. 

Grand,  J.  B.,  P.  L.  Flint,  and  P.  J.  Heglund.  1997. 
Habitat  use  by  breeding  Northern  Pintails  on  the 
Yukon-Kuskokwim  Delta,  Alaska.  J.  Wildl.  Man- 
age. 61:1 199-1207. 

Hepp,  G.  R.,  D.  j.  Strangohr,  L.  A.  Baker,  and  R. 
A.  Kennamer.  1987.  Factors  affecting  variation  in 
the  egg  and  duckling  components  of  Wood  Ducks. 
Auk  104:435-443. 

Hill,  W.  L.  1995.  Intraspecific  variation  in  egg  com- 
position. Wilson  Bull.  107:382-387. 

Hoyt,  D.  F.  1979.  Practical  methods  of  estimating  vol- 
ume and  fresh  weight  of  bird  eggs.  Auk  96:73- 
77. 

Lack,  D.  1967.  The  significance  of  clutch-size  in  wa- 
terfowl. Wildfowl  18:125-128. 

Larsson,  K.  and  P.  Forslund.  1992.  Genetic  and  so- 
cial inheritance  of  body  and  egg  size  in  the  Bar- 
nacle Goose  (Branta  leucopsis).  Evolution  46: 
235-244. 

Leblanc,  Y.  1987.  Intraclutch  variation  in  egg  size  of 
Canada  Geese.  Can.  J.  Zool.  65:3044-3047. 

Leblanc,  Y.  1989.  Variation  in  size  of  egg  of  captive 
and  wild  Canada  Geese.  Ornis  Scand.  20:93-98. 

Lessells,  C.  M.  1986.  Brood  size  in  Canada  Geese:  a 
manipulation  experiment.  J.  Anim.  Ecol.  55:669- 
689. 

Lessells,  C.  M.  and  P.  T.  Boag.  1987.  Unrepeatable 


repeatabilities:  a common  mistake.  Auk  104:1  16- 
121. 

Lessells,  C.  M.,  F.  Cooke,  and  R.  F Rockwell.  1989. 
Is  there  a trade-off  between  egg  weight  and  clutch 
size  in  Lesser  Snow  Geese  (Anser  c.  caerides- 
cens)l  J.  Evol.  Biol.  2:457-472. 

Montevecchi,  W.  a.,  1.  R.  Kirkman,  D.  D.  Roby,  and 
K.  L.  Brink.  1983.  Size,  organic  composition,  and 
energy  content  of  Leach’s  Storm  Petrel  (Ocean- 
odronia  leuohoa)  eggs  with  reference  to  position 
in  the  precocial-altricial  spectrum  and  breeding 
ecology.  Can.  J.  Zool.  61:1457-1463. 

Owen,  M.  and  J.  West.  1988.  Variation  in  egg  com- 
position in  semi-captive  Barnacle  Gee.se.  Ornis 
Scand.  19:58-62. 

Rhymer,  J.  M.  1988.  The  effect  of  egg  size  variability 
on  thermoregulation  of  Mallard  (Anas  platyrhyn- 
chos)  offspring  and  its  implications  for  survival. 
Oecologia  75:20-24. 

Robertson,  G.  J.  and  F.  Cooke.  1993.  Intraclutch  egg 
size  variation  and  hatching  success  in  the  Com- 
mon Eider.  Can.  J.  Zool.  71:544-549. 

Rockwell,  R.  F,  C.  S.  Findlay,  and  F Cooke.  1987. 
Is  there  an  optimal  clutch  size  in  Snow  Geese? 
Am.  Nat.  130:839-863. 

Rohwer,  F.  C.  1986.  Composition  of  Blue-winged  Teal 
eggs  in  relation  to  egg  size,  clutch  size,  and  the 
timing  of  laying.  Condor  88:513-519. 

Rohwer,  F.  C.  1988.  Inter-  and  intraspecific  relation- 
ships between  egg  size  and  clutch  size  in  water- 
fowl.  Auk  105:161-176. 

Rohwer,  F.  C.  and  D.  I.  Eisenhauer.  1989.  Egg  mass 
and  clutch  size  relationships  in  geese,  eiders,  and 
swans.  Ornis  Scand.  20:43-48. 

SAS  Institute.  1990.  SAS/STAT  user’s  guide.  Version 
6.  SAS  Institute,  Inc.  Cary,  North  Carolina. 

Slattery,  S.  M.  and  R.  T.  Alisauskas.  1995.  Egg 
characteristics  and  body  reserves  of  neonate  Ross’ 
and  Lesser  Snow  geese.  Condor  97:970-984. 

SoKAL,  R.  R.  AND  F J.  Rohlf.  1981.  Biometry,  second 
ed.  W.  H.  Freeman  and  Co.,  New  York. 

Thomas,  V.  G.  and  H.  C.  P.  Brown.  1988.  Relation- 
ships among  egg  size,  energy  reserves,  growth 
rate,  and  fasting  resistance  of  Canada  Goose  gos- 
lings from  .southern  Ontario.  Can.  J.  Zool.  66:957- 
964. 

Williams,  T.  D.  1994.  Intraspecific  variation  in  egg 
size  and  egg  composition  in  birds:  effects  on  off- 
spring fitness.  Biol.  Rev.  68:35-59. 

Williams,  T.  D.,  D.  B.  Lank,  F Cooke,  and  R.  F. 
Rockwell.  1993.  Fitness  consequences  of  egg- 
size  variation  in  the  Lesser  Snow  Goose.  Oecol- 
ogia 96:331-338. 


Wilson  Bull.,  111(4),  1999,  pp.  472-477 


POSTFLEDGING  BEHAVIOR  OF  GOLDEN  EAGLES 

LAURA  T.  O’TOOLE,'  2 PATRICIA  L.  KENNEDY,'  ^ RICHARD  L.  KNIGHT,'  AND 

LOWELL  C.  MCEWEN' 


ABSTRACT. — We  predicted  that  extended  parental  care,  asynchronous  hatching,  and  incidences  of  siblicide 
in  Golden  Eagles  (Aquila  chrysaeto.s)  could  increase  the  chances  for  conflict  between  siblings,  and  between 
parents  and  offspring  as  juveniles  aged.  This  conflict  could  motivate  independence  and  dispersal  in  this  species. 
To  test  our  predictions,  during  the  1993  and  1994  breeding  seasons  we  examined  post-fledging  behavior  in 
Golden  Eagles  from  the  Little  Missouri  National  Grassland  and  contiguous  areas  of  western  North  Dakota.  We 
collected  observations  of  28  radio-tagged  juveniles  to  determine  whether  predispersal  movements  were  correlated 
with  age  and  with  the  presence  of  a sibling  or  parent  during  the  first  6—10  weeks  after  fledging.  We  also  recorded 
juvenile  vocalization  rates  to  determine  if  they  changed  with  age  or  the  presence  of  a parent.  We  found  that 
distance  from  the  natal  nest  increased  with  time  since  fledging.  This  was  attributed  to  an  increased  proficiency 
in  flight  and  gradual  development  of  independence  from  parental  care.  We  found  that  calling  rate  and  distance 
between  individuals  of  sibling  pairs  did  not  change  with  time  after  fledging  but  was  highly  variable.  Calling 
rates  of  fledglings  in  the  presence  of  parents  were  higher  than  when  parents  were  absent.  Increased  calling  may 
facilitate  juvenile  location  or  inform  the  parents  of  offspring  nutritional  status.  Parents  were  not  visible  for  most 
observation  periods  and  we  did  not  observe  any  aggression  by  parents  directed  toward  offspring.  Siblings 
engaged  in  “pl^y”  activity  and  we  did  not  detect  any  signs  of  aggression  between  siblings.  Our  data  do  not 
support  the  predictions  that  an  increase  in  parental  and/or  sibling  aggression  is  associated  with  independence  m 
this  population  of  Golden  Eagles.  Received  15  Oct.  1998,  accepted  20  April  1999. 


Factors  influencing  the  timing  and  duration 
of  dispersal  may  be  extrinsic  (environmental), 
endogenous,  or  some  combination  of  the  two 
(Howard  1960,  Ritchison  et  al.  1992,  Belthoff 
and  Duffy  1995).  Extrinsic  factors  include  pa- 
rental aggression  toward  young  (Alonso  et  al. 
1987,  Hiraldo  et  al.  1989,  Wiggett  and  Boag 
1993),  sibling  aggression  (Holleback  1974,  De 
Laet  1985,  Strickland  1991),  ectoparasitism 
(Brown  and  Brown  1992),  increased  predation 
risk  (Harfenist  and  Ydenberg  1995),  and  de- 
clining food  availability  within  the  natal  area 
(Messier  1985,  Kenward  et  al.  1993,  Busta- 
mante 1994b).  In  raptors,  the  role  of  sibling 
and  parental  aggression  in  family  break-up  is 
disputed.  In  some  species,  dispersal  may  oc- 
cur after  a period  of  parent-offspring  conflict, 
the  parents  being  the  ones  who  promote  the 
independence  of  juveniles  by  gradually  reduc- 
ing the  food  supply  or  increasing  aggression 
towards  the  juveniles  (Alonso  et  al.  1987,  De- 
lannoy  and  Cruz  1988,  Hiraldo  et  al.  1989). 
Family  break-up  may  also  take  place  without 
apparent  conflict  (Bustamante  and  Hiraldo 


' Dept,  of  Fishery  and  Wildlife  Biology,  Colorado 
State  Univ.,  Ft.  Collins.  CO  80523. 

2 Current  address:  1815  W.  Grange  Avenue,  Mil- 
waukee, WI  53221 . 

’ Ct^rresponding  author; 

E-mail : patk@cnr.colostate.edu 


1990,  Bustamante  1994b).  Although  Watson 
(1997)  thought  that  there  was  little  evidence 
of  aggression  between  adult  Golden  Eagles 
and  their  young.  Walker  (1987)  observed 
some  aggression  by  parents  toward  offspring 
during  the  fledgling  dependency  period.  There 
is  little  information  on  the  factors  that  affect 
timing  of  dispersal  in  Golden  Eagles  because 
few  studies  have  been  conducted  on  the  be- 
havior of  juvenile  eagles  during  this  stage 
(Watson  1997).  We  predicted  that  the  extend- 
ed parental  care,  asynchronous  hatching  (Wat- 
son 1997),  and  high  incidence  of  siblicide 
(Edwards  and  Collopy  1983,  Edwards  et  al. 
1988)  in  this  species  would  increase  the  prob- 
ability of  predispersal  conflict. 

METHODS 

Study  area. — We  conducted  this  study  from  May  to 
November  1993-1994  in  the  Little  Missouri  National 
Grassland  (46°  00-48°  07'  N,  102°  50'-104°  00'  W) 
and  contiguous  areas  in  western  North  Dakota.  The 
dominant  habitat  was  mixed  prairie  with  patches  of 
shrubs  managed  primarily  for  livestock  grazing  (Hop- 
kins et  al.  1986,  Fowler  et  al.  1991).  Woodlands  (1.8% 
of  total  vegetative  cover)  were  found  in  areas  of  higher 
soil  moisture,  such  as  valley  bottoms,  lower  valley 
slopes,  and  along  stream  banks  and  floodplains  of  the 
Little  Missouri  River  (Girard  et  al.  1989).  Large  buttes, 
easily  eroded  sandstone,  and  clay  badlands  character- 
ized the  topography.  The  climate  is  semi-arid  conti- 
nental with  wide  daily  fluctuations  in  temperature  and 


472 


O'Toole  ei  al.  • FLEDGLING  GOLDEN  EAGLE  BEHAVIOR 


473 


variable  precipitation.  Annual  precipitation  averages 
33-41  cm.  and  the  average  temperature  in  July  is 
21°C  (Jensen  1972).  We  found  Golden  Eagle  nests  on 
siltstonc  and  clay  buttes,  and  in  trees  along  rivers  and 
streams. 

Data  collection. — We  located  active  Golden  Eagle 
nests  by  aerial  and  ground  surveys.  We  aged  chicks 
based  on  behavior  and  plumage  (Ellis  1979),  and  sexed 
them  based  on  bill  depth  and  head,  hallux  claw,  and 
tail  length  (A.  Harmata,  pers.  comm.).  We  entered  each 
of  the  20  eagle  nests  (/;  = 10  in  1993;  /?  = 10  in  1994) 
when  nestlings  were  8-10  weeks  of  age  (near  fledg- 
ing). We  selected  this  age  because  harassment  can 
cause  prolonged  ab.sence  of  eagle  adults  which  could 
result  in  nestlings  being  exposed  to  direct  sunlight  or 
missed  feedings — both  fatal  to  younger  chicks.  Be- 
cause young  eagles  will  fledge  prematurely  if  dis- 
turbed, we  waited  until  nestlings  were  fully  feathered 
and  could  fly  well  enough  to  avoid  injury  if  they 
fledged  in  response  to  our  presence  (Fyfe  and  Olen- 
dorft  1976).  Twenty-eight  (12  male,  16  female)  chicks 
(/!  = 12  in  1993;  n = 16  in  1994)  were  weighed,  mea- 
sured, fitted  with  a 25-g  backpack-style  radio  trans- 
mitter (L.  L.  Electronics,  Mahomet,  Illinois),  and  band- 
ed with  an  aluminum  U.S.  Fish  and  Wildlife  Service 
leg  band.  After  we  instrumented  and  measured  each 
eagle,  we  placed  it  back  in  the  nest  where  it  remained 
until  it  fledged.  We  spent  an  average  of  51.3  (±  3.9 
SE)  min  from  eagle  capture  to  replacement  in  the  nest. 

We  considered  a juvenile  as  fledged  after  its  first 
flight  from  the  nest.  After  fledging,  we  located  birds 
using  ground  searches  every  2-4  days  for  behavioral 
observations.  We  recorded  over  416  hr  of  direct  be- 
havioral observations  from  14  July  to  20  August  1993, 
and  15  July  to  22  August  1994.  We  recorded  move- 
ments of  radio-telemetered  birds  until  they  dispersed 
from  their  natal  territories,  or  we  found  them  dead. 
During  the  first  five  weeks,  we  were  usually  able  to 
locate  eagles  visually  (94.1%  of  attempts).  We  pri- 
marily u.sed  triangulation  (Kenward  1987)  by  four  ob- 
servers to  find  juveniles  in  the  later  stages  of  the  post- 
fledging  period. 

We  observed  eagles  during  daylight  hours  (06:00- 
21:00  MST)  using  I4-45X  zoom  spotting  scopes  from 
vehicles  or  at  locations  at  lea.st  300  m from  the  bird. 
We  determined  target  observation  times  by  randomly 
selecting  a 4-hr  block  of  daylight.  We  covered  the  oth- 
er time  periods  during  later  observations  using  the 
same  method.  During  each  observation  period,  wc 
continually  recorded  data  until  the  predetermined  time 
elapsed,  or  until  we  could  no  longer  observe  the  bird. 
We  collected  a mean  of  15.4  (±  1.2)  hr  of  observation 
per  bird  over  4.3  (±  0.3)  observation  periods.  We  re- 
corded eagle  locations  and  activities  on  a data  check- 
sheet.  We  did  not  record  sibling  behaviors  simulta- 
neously. We  recorded  the  following  data  on  adults:  lo- 
cation in  relation  to  focal  bird,  incidences  of  food  de- 
liveries by  an  adult,  stooping  and  physical  aggression 
directed  toward  the  focal  bird.  We  observed  few  prey 
deliveries  (/?  = 9),  so  we  could  not  analyze  temporal 
changes  in  food  provisioning  rates. 


We  attempted  to  sample  behavior  equally  among 
birds  tor  diltcrent  ages  and  times  of  day.  We  calculated 
distance  from  local  bird  to  the  nest  and  the  distance 
between  siblings  using  simple  geometry  and  UTM  co- 
ordinates. 

Statistical  analy.si.s. — Using  Repeated  Measures 
Analysis  of  Variance  (ANOVA;  Proc  GLM;  SAS  Inst. 
Inc.  1987;  PC  ver.  6.10)  we  tested  our  predictions  on 
the  pooled  1993—1994  data  because  there  were  too  few 
data  to  examine  a year  effect  using  repeated  measures 
analysis.  To  determine  the  proper  sampling  unit  for 
analysis  (individual  bird  vs  nest)  we  used  the  ratio  of 
Type  III  Sum  of  Squares  (Proc  GLM).  A relatively 
large  ratio  (>2)  would  indicate  that  individual  birds 
should  be  combined  into  one  “nest”  unit  because  sep- 
arate consideration  results  in  unexplained  error  (Sum 
of  Squared  Errors;  P.  Chapman,  pers.  comm.).  There 
was  no  evidence  of  a strong  sibling  effect  (average 
ratio  = 1.3),  and  because  all  behavioral  observations 
were  made  after  fledging,  we  used  individuals  as  the 
unit  of  analysis. 

We  compared  hourly  calling  rate  between  three  age 
groups:  0-14,  15-28,  and  29-42  days  post-fledging; 
and  a fledgling’s  distance  from  the  natal  nest  over  six 
age  periods  (0-15,  16-28,  29-52,  53-74,  75-97,  and 
98-137  days  post-fledging).  We  tested  whether  the  dis- 
tance between  siblings  increased  over  three  post-fledg- 
ing age  periods  (0-37,  38-74,  and  75-121  days).  For 
analysis,  we  grouped  fledgling  ages  differently  because 
the  data  collected  for  calling,  distance  from  the  nest, 
and  distance  between  siblings  varied  in  quantity.  We 
were  able  to  collect  data  on  distance  from  nest  longer 
than  any  other  measure.  Also,  calling  rates  were  re- 
corded hourly,  unlike  both  distance  measurements.  The 
selected  age  groups  fulfilled  minimum  sample  size  re- 
quirements for  the  most  suitable  statistical  tests  and  to 
achieve  a stratified  distribution  of  data  points  (P.  Chap- 
man, pers.  comm.).  All  data  sets  used  for  these  anal- 
yses were  log-transformed  to  equalize  the  variance  of 
errors  (SAS  Inst.  Inc.  1987;  PC  ver.  6.10)  but  the  raw 
data  are  reported  in  this  paper.  The  a level  for  all  sta- 
tistical tests  was  0.05. 

To  analyze  data  on  parent-offspring  interactions,  we 
used  ANOVA  (Proc  GLM;  SAS  Inst.  Inc.  1987;  PC 
ver.  6.10)  on  log-transformed  data  to  test  whether  call- 
ing rates  or  fledgling  distance  from  the  nest  changed 
when  parents  were  present.  We  also  examined  whether 
fledgling  distance  from  the  nest  depended  upon  an  in- 
teraction between  parents'  presence  and  fledging  age. 

RESULTS 

Hourly  calling  rate. — The  number  of  calls 
per  hour  did  not  change  with  time  after  fledg- 
ing (Wilks’  Lambda  ^24,2  = 0.06,  P > 0.05). 
The  mean  values  for  each  of  the  three  groups 
were  similar,  with  an  overall  mean  of  1 1.4  (± 
0.01)  calls  hr  ' (Fig.  1).  Calling  rate  did  not 
differ  between  the  sexes  (F,  4,3  = 0.50,  P > 
0.05). 


474 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


s 

o 

JC. 


u 

o 


u 

£ 


E 


3 

Z 


25 


20  - 


15  - 


10 


5 


0 


0-14 


15-28  29-42 

Days  postlledging 


FIG.  1.  Mean  number  of  calls  ISEI  per  hour  for 
Golden  Eagles  in  three  post-hedging  time  periods  (0- 
14,  15-28,  29-42  d)  in  the  Little  Missouri  National 
Grassland  of  western  North  Dakota,  1993  and  1994. 


Distance  from  nest. — The  distance  of  fledg- 
lings from  their  nests  increased  significantly 
with  time  (Wilks’  Lambda  F4  ,o  = 11.53,  P = 
0.001;  Fig.  2)  and  eventually  resulted  in  loss 
of  radio  contact  within  the  study  area  (27,500 
km^)  as  the  birds  dispersed.  There  was  no  dif- 
ference between  sexes  in  the  distance  individ- 
uals dispersed  (F,  ,3  = 0.06,  P > 0.05). 

Sibling  interactions  and  juvenile  behav- 
ior.— We  collected  data  from  six  sibling  pairs 
to  determine  the  mean  distance  between  sib- 
lings with  time  since  fledging.  There  was  no 
significant  (Wilks’  Lambda  F2.3  = 0.34,  P > 
0.05)  change  over  the  time  period  of  our  study 
(Fig.  3). 

Fledgling  pairs  frequently  exhibited  non- 
aggressive  social  behavior.  Siblings  often  fol- 
lowed one  another  or  flew  together.  We  ob- 
served mutual  preening  or  “nibbling  (Ellis 
1979)  when  the  fledglings  perched  together  (n 
= 288),  as  they  frequently  did.  We  observed 
three  different  sibling  pairs  display  “play 
catching  and  plucking  of  prey  together.  We 
recorded  no  overtly  aggressive  social  behavior 
between  sibling  pairs,  and  this  social  activity 
did  not  appear  to  change  over  time. 

The  mean  age  at  first  flight  was  10.1  (± 
0.08)  weeks.  We  observed  1 1 attempts  of  prey 
capture  by  juveniles.  Two  of  the  observed  at- 


FIG.  2.  Mean  distance  from  the  nest  (m)  ISEI  for 
Golden  Eagles  in  six  post-fledging  time  periods  (0-15, 
16-28,  29-52,  53-74,  75-97,  98-137  d)  in  the  Little 
Missouri  National  Grassland  of  western  North  Dakota, 
1993  and  1994. 


500 

450 

400 

350 


300 


s 250 

o 

i: 

e 200 

3 


Q 


150 

100 

50 

0 


0-37  38-74 


i 


75-121 


Days  postfledging 


FIG.  3.  Mean  distance  ISEI  between  Golden  Eagle 
siblings  (6  pairs)  in  three  post-fledging  time  periods 
(0-37,  38-74,  75-121  d)  in  the  Little  Missouri  Na- 
tional Grassland  of  western  North  Dakota,  1993  and 
1994. 


O'Toole  el  al.  • FLEDGLING  GOLDEN  EAGLE  BEHAVIOR 


475 


tempts  were  successful,  with  prey  items  in- 
cluding an  unidentified  snake  and  a rodent. 

Parental  interactions. — We  observed  par- 
ents near  the  focal  bird  relatively  infrequently 
(29.6%)  during  observation  periods.  Calling 
rates  of  the  young  were  higher  in  the  presence 
of  parents  than  when  parents  were  not  visible 
(^1.41.1  ~ 14.39,  P — 0.001).  The  rate  of  calling 
nearly  doubled  for  juveniles  with  a parent 
present  (x  ± SE  = 21.8  ± 3.5  calls  hr  ‘)  ver- 
sus parents  absent  (11.5  ± 1.7  calls  hr”‘). 
Fledglings  were  closer  to  the  nest  when  par- 
ents were  present  (F,  ,3  = 10.81,  P = 0.001), 
but  distance  from  the  nest  was  not  signifi- 
cantly correlated  with  the  interaction  of  fledg- 
ling age  and  parental  presence  (F4  10  = 0.96, 
P > 0.05).  We  did  not  observe  any  aggressive 
behavior  by  the  parent  toward  the  offspring. 

DISCUSSION 

Calling  rate  did  not  change  as  the  birds 
aged  after  fledging,  although  we  observed  a 
high  degree  of  individual  variation.  A similar 
finding  was  reported  in  a Japanese  population 
of  Black  Kites  (Milvus  migrans;  Koga  and 
Shiraishi  1994).  Increased  calling  with  age 
was  reported  in  Spanish  populations  of  Black 
Kites  (Bustamante  and  Hiraldo  1990),  the 
Egyptian  Vulture  {Neophron  percnopterus; 
Ceballos  and  Donazar  1990),  and  the  Spanish 
Imperial  Eagle  (Aquila  adalberti;  Alonso  et  al. 
1987).  The  increased  calling  rates  in  these 
studies  were  attributed  to  a decrease  in  pro- 
visioning by  parents  as  the  young  aged.  Our 
results  do  not  corroborate  findings  of  in- 
creased calling  with  age.  It  is  possible  that 
food  provisioning  did  not  decrease  over  time 
in  our  study,  resulting  in  no  change  in  calling 
rates.  It  is  also  possible  that  food  provisioning 
decreased,  but  calling  did  not  increase  because 
the  young  were  becoming  independent  and 
beginning  to  hunt  on  their  own.  We  cannot 
evaluate  these  hypotheses  because  we  did  not 
have  enough  data  to  analyze  food  provision- 
ing rates. 

Throughout  our  study,  the  appearance  of  a 
parent  resulted  in  increased  juvenile  calling 
compared  with  when  parents  were  absent.  In- 
creased calling  in  the  presence  of  a parent  is 
commonly  observed  in  raptors  (Alonso  et  al. 
1987,  Ikeda  1987,  Hiraldo  et  al.  1989,  Bus- 
tamante 1994a).  Calling  by  juveniles  may  fa- 
cilitate juvenile  location  by  parents  (Ikeda 


1987).  Calling  also  informs  the  parents  of  the 
nutritional  status  of  their  offspring  (Trivers 
1974). 

Distance  from  the  nest  increased  as  the  ju- 
veniles aged,  probably  in  part  due  to  increased 
flying  proficiency.  As  independence  nears, 
movements  may  not  represent  a linear  dis- 
persal, but  may  resemble  “wanderings”  out- 
side the  natal  area.  Similar  observations  have 
been  reported  for  other  species  (Boeker  and 
Ray  1971,  Beecham  and  Kochert  1975,  Steen- 
hof  et  al.  1984,  Walker  1987,  Bahat  1992). 

Siblings  tended  to  move  together  after  leav- 
ing the  nest,  and  the  distance  between  them 
remained  relatively  constant.  This  contradicts 
Newton’s  (1979)  idea  that  fledgling  raptors 
tend  to  perch  apart.  Other  studies  also  re- 
vealed increased  distance  between  siblings 
with  age  (Hiraldo  et  al.  1989,  Bustamante  and 
Hiraldo  1990,  Ceballos  and  Donazar  1990). 
Increased  sibling  distance  has  been  attributed 
to  “tension”  between  juveniles  or  increased 
flight  proficiency.  Bustamante  (1994a)  report- 
ed that  Kestrels  {Falco  tinnunculus)  perched 
close  together  and  engaged  in  social  behavior 
similar  to  the  young  Golden  Eagles  we  ob- 
served. Bustamante  (1993)  also  found  siblings 
of  Black-shouldered  Kites  often  perched  to- 
gether. Kenward  and  coworkers  (1993)  saw 
the  same  tendency  with  Northern  Goshawk 
{Accipiter  gentilis)  siblings. 

If  parents  and  offspring  were  in  conflict,  re- 
sulting in  the  adults  “driving-off”  the  juve- 
niles, or  if  parents  used  aggression  to  evaluate 
the  young’s  flight  proficiency  (Ferrer  1992), 
we  should  have  observed  agonistic  interac- 
tions during  the  times  when  the  parents  were 
seen  with  the  offspring,  but  we  did  not.  Al- 
though parental  aggression  has  been  observed 
in  raptor  species  (Robertson  1985,  Alonso  et 
al.  1987,  Walker  1987,  Hiraldo  et  al.  1989),  it 
is  often  the  case  that  the  parents  feed  their 
offspring  long  after  they  have  fledged  (Ikeda 
1987,  Walker  1987,  Bustamante  and  Hiraldo 
1990,  Ceballos  and  Donazar  1990,  Bahat 
1992). 

We  did  not  find  an  increase  in  aggression 
between  siblings,  as  we  predicted.  Although 
juveniles  moved  away  from  their  parental 
home  range,  they  apparently  remained  togeth- 
er, indicative  of  sibling  attraction  rather  than 
aversion.  Our  data  do  not  support  the  predic- 
tions that  an  increase  in  parental  and/or  sibling 


476 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


agression  is  associated  with  independence  in 
this  population. 

ACKNOWLEDGMENTS 

We  would  like  to  acknowledge  the  following  for 
field  assistance;  M.  Melendez,  L.  Masten,  P.  O’Brien, 
R.  Brunotte,  C.  Stubblefield,  and  B.  O'Toole.  Many 
thanks  go  to  C.  Althouse  and  B.  Peterson  for  technical 
assistance;  T.  Tiara  and  M.  Hooper  for  blood  chemical 
analysis;  B.  Tokaruk  for  recovery  of  one  eagle  in  Sas- 
katchewan; and  D.  Getzy  for  necropsy.  We  thank  A. 
Harmata,  G.  Fob,  R.  Kreil,  K.  Sanchez,  L.  Brand,  K. 
Winks,  D.  Hirsch,  W.  Cushing,  and  J.  Jeske  for  other 
help.  We  thank  J.  Smith  and  two  anonymous  reviewers 
for  their  helpful  reviews  of  the  manuscript.  Support  for 
this  research  project  was  provided  by  the  United  States 
Department  of  Agriculture,  Animal  and  Plant  Health 
Inspection  Service,  PPQ  Grasshopper  IPM  Project;  the 
U.S.  Fish  and  Wildlife  Service,  U.S.  Forest  Service 
LMNG;  and  the  North  Dakota  Game  and  Fish  De- 
partment. The  Animal  Care  and  Use  Committee  of 
Colorado  State  University  approved  the  methodologies 
used  in  this  project. 

LITERATURE  CITED 

Alonso,  J.  C.,  L.  M.  Gonzalez,  B.  Heredia,  and  J. 
L.  Gonzalez.  1987.  Parental  care  and  the  transi- 
tion to  independence  of  Spanish  Imperial  Eagle 
Aqiiila  heliaca  in  Donana  National  Park,  South- 
west Spain.  Ibis  129:212-224. 

Bahat,  O.  1992.  Post-fledging  movements  of  Golden 
Eagles  (Aquila  chry.saetos  homeryeri)  in  the  Ne- 
gev Desert,  Isreal,  as  determined  by  radio-telem- 
etry. Pp.  612-621  in  Wildlife  telemetry:  remote 
monitoring  and  tracking  of  animals  (LG.  Preide 
and  S.M.  Swift,  Eds.).  Ellis  Horwood  Ltd,  West 
Sussex,  U.K. 

Beecham,  j.  j.  and  M.  N.  Kochert.  1975.  Breeding 
biology  of  the  Golden  Eagle  in  .southwestern  Ida- 
ho. Wilson  Bull.  87:506—513. 

Belthoff,  j.  R.  and  a.  M.  Duffy,  Jr.  1995.  Loco- 
motor activity  levels  and  the  dispersal  of  Western 
Screech-Owl,  Otus  kennicottii.  Anim.  Behav.  50: 
558-561. 

Boeker,  E.  L.  and  T.  D.  Ray.  1971.  Golden  Eagle 
population  studies  in  the  Southwest.  Condor  73: 
463-467. 

Brown,  C.  R.  and  M.  B.  Brown.  1992.  Ectoparasitism 
as  a cause  of  natal  dispersal  in  Cliff  Swallows. 
Ecology  73:1718—1723. 

Bu,stamante,  j.  1993.  The  post-fledging  dependence 
period  of  the  Black-Shouldered  Kite  {Elanns  cae- 
ruleiis).  J.  Raptor  Res.  27:185-190. 

BU.STAMANTE,  J.  1994a.  Behavior  of  colonial  Common 
Kestrels  (Falco  tinminculus)  during  the  post- 
fledging  dependence  period  in  southwestern 
Spain.  J.  Raptor  Res.  28:79-83. 

Bustamante,  J.  1994b.  Family  break-up  in  Black  and 
Red  kites  Milvu.s  mif>ran.s  and  M.  milvii.s:  is  time 


of  independence  an  offspring  decision?  Ibis  136: 
176-184. 

Bustamante,  J.  and  F.  Hiraldo.  1990.  Factors  influ- 
encing family  rupture  and  parent-offspring  con- 
flict in  the  Black  Kite  (Milvu.s  mii’rans).  Ibis  132; 
58-67. 

Ceballos,  O.  and  j.  a.  Donazar.  1990.  Parent-off- 
spring  conflict  during  the  post-fledging  period  in 
the  Egyptian  Vulture  Neophron  perenopterus 
(Aves,  Accipitridae).  Ethology  85:225-235. 

DeLaet,  J.V.  1985.  Dominance  and  aggression  in  ju- 
venile Great  Tits,  Panes'  major  major,  in  relation 
to  dispersal.  Pp.  375-380  in  Behavioral  ecology: 
ecological  consequences  of  adaptive  behavior 
(Sibley,  R.  M.  and  R.  H.  Smith,  Eds.).  Blackwell 
Scientific  Publ.,  London,  U.K. 

Delannoy,  C.  a.  and  a.  Cruz.  1988.  Breeding  biol- 
ogy of  the  Puerto  Rican  Sharp-shinned  Hawk  (Ac- 
cipiter  striatus  Venator).  Auk  105:649—662. 

Edwards,  T.  C.,  Jr.  and  M.  W.  Collopy.  1983.  Ob- 
ligate and  facultative  brood  reduction  in  eagles; 
an  examination  of  factors  that  influence  fratricide. 
Auk  100:630-635. 

Edwards,  T.  C.,  Jr.,  M.  W.  Collopy,  K.  Steenhof, 
AND  M.  N.  Kochert.  1988.  Sex  ratios  of  fledgling 
Golden  Eagles.  Auk  l(J5;793-796. 

Ellis,  D.  H.  1979.  Development  of  behavior  in  the 
Golden  Eagle.  Wildl.  Monogr.  43:1-94. 

Ferrer,  M.  1992.  Regulation  of  the  period  of  post- 
fledging  dependence  in  the  Spanish  Imperial  Ea- 
gle/\r/n//u  Ibis  134:128—133. 

Fowler,  A.  C.,  R.  L.  Knight,  T.  L.  George,  and  L. 
C.  McEwen.  1991.  Effects  of  avian  predation  on 
grasshopper  populations  in  North  Dakota  grass- 
lands. Ecology  72:1775—1781. 

Fyfe,  R.  W.  and  R.  R.  Olendorff.  1976.  Minimizing 
the  dangers  of  nesting  studies  to  raptors  and  other 
sensitive  species.  Can.  Wildl.  Serv.  Occas.  Pap. 
23:1-17. 

Girard,  M.  M.,  H.  Goetz,  and  A.  J.  Bjugstad.  1989. 
Native  woodland  habitat  types  of  southwestern 
North  Dakota.  USDA  For.  Serv.  Res.  Pap.  RM- 
281:1-36. 

Harfenist,  a.  and  R.  C.  Ydenberg.  1995.  Parental 
provisioning  and  predation  risk  in  Rhinoceros 
Auklets  (Cerorhinca  monocerala):  effects  on  nest- 
ling growth  and  fledging.  Behav.  Ecol.  6:82-86. 

Hiraldo,  E,  M.  Delibes,  and  R.  R.  E.strella.  1989. 
Observations  of  a Zone-tailed  Hawk  family  during 
the  post-fledging  period.  J.  Raptor  Res.  23:103— 

1 06. 

Holleback,  M.  1974.  Behavioral  interactions  and  the 
dispersal  of  the  family  in  Black-capped  Chicka- 
dees. Wilson  Bull.  76:28-36. 

Hopkins.  R.  B.,  J.  F.  Cassel,  and  A.  J.  Bjugstad. 
1986.  Relationships  between  breeding  birds  and 
vegetation  in  four  woodland  types  of  the  Little 
Missouri  National  Grasslands.  USDA  For.  Serv. 
Res.  Pap.  MR-27():I-I2. 

Howard,  W.  E.  1960.  Innate  and  environmental  dis- 


O'Toole  et  al.  • FLEDGLING  GOLDEN  EAGLE  BEHAVIOR 


477 


persal  of  individual  vertebrates.  Am.  Midi.  Nat. 
63:152-161. 

Ikeda,  Y.  1987.  Ecological  research  on  the  Japanese 
Golden  Eagle  Aquila  chrysaetos  japonica  during 
the  post-fledging  period  in  the  Hakusan  Range.  J. 
Raptor  Res.  21:79-80. 

Jensen,  R.  R.  1972.  Climate  of  North  Dakota.  North 
Dakota  State  Univ.,  Eargo. 

Kenward,  R.  E.  1987.  Wildlife  radio  tagging:  equip- 
ment, field  techniques  and  data  analysis.  Academ- 
ic Press,  London,  U.K. 

Kenward,  R.  E.,  V.  Marcstrom,  and  M.  Karlblom. 
1993.  Post-nestling  behavior  of  Goshawks,  Accip- 
iter  genfilis.  1.  The  causes  of  dispersal.  Anim.  Be- 
hav.  46:65-70. 

Koga,  K.  and  S.  Shiraishi.  1994.  Parent-offspring  re- 
lations during  the  post-fledging  dependency  peri- 
od in  the  Black  Kite  {Milvus  migrans)  in  Japan. 
J.  Raptor  Res.  28:171-177. 

Meissier,  E 1985.  Solitary  living  and  extraterritorial 
movements  of  wolves  in  relation  to  social  status 
and  prey  abundance.  Can.  J.  Zool.  63:239-245. 

Newton,  I.  1979.  Population  ecology  of  raptors.  Buteo 
Books,  Vermillion,  South  Dakota. 

Ritchison,  a.  j.,  j.  R.  Belthoff,  and  E.  J.  Sparks. 


1992.  Di.spersal  restlessness:  evidence  for  innate 
dispersal  by  juvenile  Eastern  Screech  Owls? 
Anim.  Behav.  43:57-65. 

Robertson,  A.  S.  1985.  Observations  on  the  post- 
fledging  dependence  period  of  Cape  Vultures.  Os- 
trich 56:58-66. 

SAS  Institute,  Inc.  1987.  SAS  stati.stical  guide  for 
personal  computers,  version  6.10.  SAS  Inst.  Inc., 
Cary,  North  Carolina. 

Steenhof,  K.,  M.  N.  Kochert,  and  M.  Q.  Moritsch. 
1984.  Dispersal  and  migration  of  southwestern 
Idaho  raptors.  J.  Field  Ornithol.  55:357-368. 

Strickland,  D.  1991.  Juvenile  dispersal  in  Gray  Jays: 
dominant  brood  member  expels  siblings  from  na- 
tal teiTitory.  Can.  J.  Zool.  69:2935-2945. 

Trivers,  R.  L.  1974.  Parent-offspring  conflict.  Am. 
Zool.  14:249-264. 

Walker,  D.  G.  1987.  Observations  on  the  post-fledg- 
ing period  of  the  Golden  Eagle  Aquila  chrysaetos 
in  England.  Ibis  129:92-96. 

Watson,  J.  1997.  The  Golden  Eagle.  T & A D Poyser, 
London,  U.K. 

WiGGETT,  D.  R.  and  D.  a.  Boag.  1993.  The  proximate 
cause  of  male-biased  natal  emigration  in  Colum- 
bian ground  squirrels.  Can.  J.  Zool.  71:201-218. 


Wilson  Bull.,  111(4),  1999,  pp.  478-487 


GROWTH  PATTERNS  OF  HAWAIIAN  STILT  CHICKS 


J.  MICHAEL  REED,'  ELIZABETH  M.  GRAY,^^  DIANNE  LEWIS,^ 
LEWIS  W.  ORING,3  RICHARD  COLEMAN, ^ TIMOTHY  BURR,^  AND 

PETER  LUSCOMB^ 


ABSTRACT. — We  studied  chick  growth  and  plumage  patterns  in  the  endangered  Hawaiian  Stilt  (Himantopus 
mexictmus  knudseni).  Body  mass  of  captive  chicks  closely  fit  a Gompertz  growth  curve,  revealing  a growth 
coefficient  {K)  of  0.065  day  ' and  point  of  inflection  (T)  of  17  days.  When  chicks  fledged  about  28  days  after 
hatching,  they  weighed  only  60%  of  adult  body  mass;  at  42  d,  birds  still  were  only  75%  of  adult  mass;  culmen, 
tarsus,  and  wing  chord  at  fledging  also  were  less  than  adult  size.  This  trend  of  continued  growth  to  adult  size 
after  fledging  is  typical  for  most  shorebirds.  After  hatching,  captive  chicks  grew  more  rapidly  than  wild  chicks, 
probably  because  of  an  unlimited  food  supply.  We  found  no  evidence  for  adverse  effects  of  weather  on  the 
growth  of  wild  chicks.  As  with  other  shorebirds,  the  tarsus  started  relatively  long,  with  culmen  and  then  wing 
chord  growing  more  rapidly  in  later  development.  Tarsal  and  wing  chord  growth  were  sigmoidal,  whereas  culmen 
growth  was  linear.  We  describe  plumage  characteristics  of  weekly  age  classes  of  chicks  to  help  researchers  age 
birds  in  the  wild.  Received  28  Dec.  1998,  accepted  20  April  1999. 


Avian  growth  patterns  have  been  studied 
primarily  because  of  their  relationships  to  the 
ecology  and  evolutionary  history  of  different 
species  (Ricklefs  1968,  1973,  1983;  O’Connor 
1984;  Anthony  et  al.  1991),  and  to  maximize 
food  yields  of  domestic  animals  (e.g.,  An- 
thony et  al.  1991).  Although  there  is  selection 
for  rapid  independence  of  chicks,  which 
should  reduce  variance  in  growth  rates,  intra- 
specific growth  patterns  can  be  variable  and 
flexible  because  of  environmental  variability 
and  competing  selective  pressures  (Cooch  et 
al.  1991,  Emlen  et  al.  1991).  In  studies  of  wild 
birds,  altricial  species  have  been  studied  more 
often  than  precocial  species,  at  least  in  part 
because  the  former  remain  in  the  nest  from 
hatching  until  fledging. 

In  this  paper  we  present  information  on 
chick  growth  patterns  of  the  Hawaiian  Stilt 


' Biological  Resources  Research  Center,  Univ.  of 
Nevada,  Reno,  NV  89557. 

2 Current  address;  Dept,  of  Biology,  Tufts  Univ., 
Medford,  MA  02155;  E-mail;  mreed@tufts.edu 

’Environmental  and  Resource  Sciences/ 1 86,  1000 
Valley  Rd.,  Univ.  of  Nevada,  Reno,  NV  89512-0013. 

Current  address;  USG.S-BRD-PIERC,  RO.  Box  44, 
Hawaii  National  Park,  HI  96718. 

’ U.S.  Fi.sh  and  Wildlife  Service,  91  1 NE  1 1th  Ave., 
Eastside  Federal  Complex,  Portland.  OR  97232. 

^ Southwest  Division,  Naval  Facilities  Engineering 
Command,  1220  Pacific  Hwy.,  San  Diego,  CA  92132- 
5190. 

' 151  Kapahulu  Ave.,  Honolulu  Zoo,  Honolulu,  HI 
96815. 

**  Corresponding  author. 


{Himantopus  mexicanus  knudseni),  a precocial 
bird  that  is  an  endangered  subspecies  of  the 
Black-necked  Stilt.  Like  all  shorebirds,  stilts 
are  precocial  and  nidifugeous.  Hawaiian  Stilts 
are  significantly  larger  than  the  nominate  race 
(Coleman  1981)  and  differ  somewhat  in  adult 
plumage  characteristics  (Wilson  and  Evans 
1893,  Coleman  1981).  Stilts  are  found  on  all 
five  major  islands  in  Hawaii,  breed  exclusive- 
ly in  shallow,  lowland  wetlands  (USFWS 
1985),  and  statewide  population  counts  indi- 
cate a steady  increase  in  population  size  (Reed 
and  Oring  1993).  Our  specific  objectives  were 
to  ( 1 ) describe  patterns  of  Hawaiian  Stilt  chick 
growth  from  captive  and  wild  birds  and  com- 
pare them  to  other  shorebirds,  and  (2)  provide 
a method  for  aging  chicks  in  the  field.  The 
last  objective  was  designed  for  studying  pre- 
adult mortality  patterns  by  providing  aging 
criteria  that  do  not  requiring  capturing  the 
bird. 

METHODS 

Captive  birds. — Growth  data  for  captive  birds  came 
from  15  individuals  raised  from  eggs  in  1980  in  the 
Honolulu  Zoo.  Because  chicks  were  kept  in  a common 
enclosure,  some  competition  for  food  might  have  oc- 
curred, although  food  was  provided  ad  libitum.  Be- 
cause all  birds  were  subject  to  the  same  feeding  and 
environmental  conditions,  inter-individual  variability 
in  growth  should  be  minimized.  All  birds  were 
weighed  daily  for  42  days  to  the  nearest  0.1  g.  Ha- 
waiian Stilts  fledge  approximately  28  d after  hatching 
(Coleman  1981). 

One  of  the  15  birds  was  used  only  for  the  first  13  d 
because  a bill  deformity  developed  at  this  time,  caus- 


478 


Reed  et  al.  • HAWAIIAN  STILT  CHICK  GROWTH 


479 


ing  the  individual  to  lose  mass  quickly.  A sixteenth 
bird  was  not  included  in  the  analysis  because  of  ab- 
eiTant  fluctuations  in  growth.  Its  mass  at  hatch  was 
over  5 standard  deviations  above  the  mean,  and  it 
gained  mass  rapidly  for  1 1 days.  Between  days  12-17, 
however,  it  lost  25%  of  its  body  mass,  dropping  well 
below  the  mean  (ca  2 standard  deviations);  on  day  18 
it  began  to  grow  rapidly  again,  reaching  mean  mass 
for  the  group  24  d later. 

Other  variables  (culmen,  tarsus,  and  wing  chord) 
were  measured  less  regularly.  Measurements  were 
made  every  2-4  d after  hatching  and  became  less  fre- 
quent (every  4—10  d)  after  fledging.  Some  individuals 
were  measured  more  often  than  others.  Despite  this 
variation,  we  were  able  to  derive  useful  growth  pat- 
terns for  these  body  measurements.  Mass  was  mea- 
sured by  one  person  and  lengths  by  another. 

A growth  curve  for  body  mass  was  fit  to  a Gompertz 
equation  (r^  = 0.99;  SPSS,  Inc.  1995,  NONLIN  pro- 
cedure) because  it  is  used  most  often  for  shorebirds 
(e.g.,  Beintema  and  Visser  1989a)  and  we  wanted  to 
allow  interspecific  comparisons  to  be  made  (O’Connor 
1984).  The  fit  was  made  on  average  values  for  each 
day  from  12—15  individuals.  The  Gompertz  equation 
has  the  form 

W = A X 

where  W is  body  mass  (g),  A is  asymptotic  (adult)  mass 
(g),  K is  the  growth  coefficient  (day  ‘),  t is  age  (d), 
and  e is  the  base  for  natural  logarithms.  Adult  mass 
came  from  43  adult  males  and  42  adult  females  (Cole- 
man 1981).  Although  adult  females  weigh  slightly 
more  than  males  (mean  difference  = 7.0  g),  the  dif- 
ference is  a small  percentage  (<4%)  of  total  body 
mass,  consequently  A was  averaged  across  sexes 
(202.5  g). 

Wild  birds. — Wild  chicks  were  captured  by  hand  on 
the  islands  of  Oahu,  Maui,  and  Kauai  in  1978-1980 
and  1993.  During  1978-1980,  we  captured  chicks  with 
known  hatching  dates  142  times.  Because  chicks  from 
the  same  clutch  were  not  considered  to  be  independent, 
they  were  averaged  within  each  clutch  (maximum  of 
four  chicks  averaged  per  clutch).  This  resulted  in  33 
measurements  of  chicks  less  than  24  h old  (designated 
day  0;  « = 64  chicks).  Chicks  were  remeasured  every 
time  they  were  encountered  and  captured.  This  resulted 
in  43  measurements  of  birds  from  2-32  d old  in  = 78 
chicks).  We  measured  mass  to  the  nearest  1 .0  g,  cul- 
men and  tarsus  lengths  to  the  nearest  0.1  mm,  and 
wing  chord  to  the  nearest  1.0  mm.  In  1993,  we  took 
measurements  on  55  birds  ranging  in  age  from  hatch- 
ing to  fledging  using  the  above  methods.  During  1993 
we  rarely  knew  the  exact  age  of  each  chick,  so  these 
measurements  were  used  only  to  determine  the  rela- 
tionships among  body  measurements.  Tarsus  and  wing 
chord  measurements  were  made  on  the  right  side  of 
the  chick  and  the  same  person  made  all  measurements 
in  1993.  We  also  noted  the  presence  or  absence  of  an 
egg  tooth.  Field  measurements  from  1978-1980  were 
made  by  one  person,  and  in  1993  by  another,  so  values 
were  not  compared. 


Fluinage. — We  considered  only  tho.se  plumage  char- 
acteristics that  were  visible  in  the  field:  fuzzy  appear- 
ance associated  with  down,  brown  versus  black  cast, 
presence  of  an  eye  ring,  etc.  We  used  the  above  char- 
acteristics to  describe  plumage  of  weekly  age  classes. 
In  several  cases,  plumage  descriptions  for  weekly  age 
classes  were  incomplete  (e.g.,  lacking  description  of 
wing  coloration  for  week  3).  Because  plumage  is  es- 
sentially the  same  for  chicks  of  both  Hawaiian  and 
Black-necked  stilts  (Coleman  1981),  we  supplemented 
our  descriptions  of  Hawaiian  Stilts  with  plumage  ob- 
servations of  wild,  known-aged  Black-necked  Stilt 
chicks  at  Honey  Lake,  California  in  1997.  Plumage  of 
adult  Hawaiian  Stilts  is  different  from  fledglings  (Rob- 
inson et  al.,  in  press). 

Analyses. — Statistical  analyses  were  conducted  us- 
ing version  7.0  of  SPSS  (SPSS,  Inc.  1995).  One  as- 
sumption in  comparing  body  measurements  between 
captive  and  wild  birds  is  that  initial  body  sizes  are 
equal.  To  test  this,  we  used  multiple  analysis  of  vari- 
ance (MANOVA)  to  compare  mass  and  culmen  length, 
tarsus  and  wing  chord  measurements  between  known- 
aged  captive  and  wild  hatch  day  (day  0)  birds.  For  ages 
after  day  0,  we  determined  whether  or  not  mean  values 
for  wild  birds  fell  within  95%  confidence  intervals  for 
mean  values  of  captive  birds.  All  statistical  tests  were 
two-tailed.  Values  presented  are  means  ± SD. 

RESULTS 

Growth  in  captivity. — Growth  parameters 
for  the  Gompertz  equation  indicated  a growth 
coefficient  {K)  of  0.065  and  time  to  inflection 
point  (T)  of  17  days.  Although  chick  mass 
varied  little  among  the  1 1 individuals  on  day 
of  hatch  (15.7  ± 0.6  g),  variability  in  mass 
among  individuals  increased  greatly  over  the 
first  two  weeks  (60.4  ± 9.2  g),  and  remained 
high  up  to  fledging  at  day  28  (122.5  ± 10.6 
g).  In  general,  differences  among  chick  mass 
at  day  14  are  consistent  until  fledging,  indi- 
cating that  chicks  that  gain  relatively  more 
mass  in  the  first  two  weeks  after  hatching  tend 
to  fledge  at  a heavier  mass  than  chicks  that 
gain  less  mass  their  first  two  weeks.  Captive 
individuals  did  not  experience  a significant 
mass  loss  between  day  0 (hatch  day)  and  day 
1 (paired  t-test:  t = —0.432,  df  = 10,  P > 
0.05). 

At  fledging,  chicks  had  not  attained  adult 
body  mass  or  body  measurements.  Mass  at 
fledging  was  60%  of  adult  mass,  culmen 
length  was  67%  of  adult  length,  tarsus  length 
was  at  66%,  and  wing  chord  length  was  at 
55%  (adult  measurements  from  Coleman 
1981). 

Growth  in  the  wild. — There  was  no  differ- 


480 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


LIG.  1.  Captive  and  wild  Hawaiian  Stilt  chick  mass  and  percentage  of  adult  body  mass  (202.5  g)  as  a 
function  of  age.  Values  are  means  ± SE. 


ence  between  mean  mass  of  captive  (x  = 15.7 
± 0.6  g)  and  wild  (x  = 15.6  ± 1.1  g)  chicks 
at  hatch  {t  = 0.551,  df  = 40,  P > 0.05).  The 
apparent  decrease  in  mass  between  day  0 
(hatch  day)  and  day  1 for  wild  chicks  was  not 
significant  (paired  f-test:  t = 0.585,  df  = 12, 
P > 0.05).  From  days  1 to  17,  masses  of 
same-aged  wild  birds  typically  fell  within  the 
95%  confidence  interval  of  captive  birds, 
though  below  the  mean.  In  three  comparisons 
(day  9,  14,  15),  the  mass  of  wild  birds  fell 
below  the  95%  confidence  interval  for  captive 
mass.  Mass  gain  with  age  generally  followed 
a sigmoidal  pattern,  with  individuals  not 
reaching  an  asymptote  until  after  42  days  of 
age  (Fig.  1).  Similarly,  from  days  1 to  17, 
mean  wing  chord  of  same-aged  wild  birds  fell 
within  the  95%  confidence  interval  of  captive 
birds,  with  the  exception  of  days  9,  14  and  15, 
when  mean  wing  chord  measurements  for 
wild  birds  fell  below  the  95%  confidence  in- 
terval. Growth  of  the  wing  chord  also  fol- 
lowed a sigmoidal  pattern,  although  the  slope 
of  the  curve  was  less  steep  for  wing  chord 
growth  than  it  was  for  mass  gain  (Fig.  2). 
Mean  culmen  length  and  mean  tarsus  length 
did  not  differ  between  wild  and  captive  birds 
from  days  1 to  17.  Mean  culmen  growth  for 
both  wild  and  captive  chicks  was  relatively 
linear  with  increasing  age  (Fig.  2). 


Relative  growth  rates. — Relative  growth 
rates  among  different  parts  of  the  body  can  be 
assessed  without  reference  to  age.  We  found 
tarsus  length  to  be  long  in  early  development 
relative  to  culmen  and  wing  chord,  and  it  con- 
tinued to  grow  at  a faster  rate  than  the  culmen 
throughout  development.  Culmen  and  wing 
chord  grew  at  approximately  the  same  rate  in 
early  development  until  wing  chord  reached 
about  40  mm;  as  wing  chord  continued  to 
grow,  culmen  length  growth  rate  slowed  con- 
siderably. Changes  in  wing  chord  and  body 
mass  were  similar  throughout  the  growth  pe- 
riod observed  (Fig.  3).  Changes  in  tarsus 
length  and  body  mass  also  were  similar  until 
individuals  reached  approximately  80  g,  when 
tarsus  growth  slowed. 

Plumage. — Using  field  data  from  known- 
aged  chicks,  we  constructed  a table  of  weekly 
plumage  characteristics  for  Hawaiian  Stilt 
chicks  (Table  1).  The  presence  or  loss  of 
down,  as  well  as  overall  body  color,  appear  to 
be  the  two  best  indicators  of  chick  age  in  the 
wild  for  weeks  1-3.  Aging  during  this  time  is 
more  precise  if  one  can  determine  the  pres- 
ence and  condition  of  primary  sheaths;  this 
cannot  be  done,  however,  without  chicks  in 
hand.  Specifically,  in  week  1 chicks  are  en- 
tirely covered  with  down,  and  primary  sheaths 
are  absent.  The  dorsal  surface  of  the  body  in- 


Reed  el  al.  • HAWAIIAN  STILT  CHICK  GROWTH 


481 


FIG.  2.  Growth  patterns  for  wing  chord,  culmen 
length,  and  tarsus  length  of  captive  and  wild  Hawaiian 
Stilt  chicks. 


eluding  head,  neck,  back,  and  wing  is  mottled 
black,  golden  brown,  and  white;  the  ventral 
surface  is  creamy  white.  In  week  2,  the  head 
begins  to  turn  brown  and  is  distinctly  lighter 
than  the  rest  of  the  body.  Mottling  on  the  neck 
changes  to  a more  solid  pattern  of  gray  and 
tan.  Most  importantly,  primary  sheaths 
emerge  on  day  12.  During  week  3,  down  be- 
gins to  disappear,  giving  chicks  a sleeker  ap- 
pearance. Overall  body  coloration  changes 
from  mottled  black,  golden  brown,  and  white 
to  plain  gray  and  white,  and  primary  sheaths 
are  broken  about  day  16.  In  all  cases  where 
we  had  information  on  both  subspecies,  plum- 
age descriptions  of  known-aged  Black-necked 
Stilt  chicks  matched  exactly  the  plumage  de- 
scriptions of  known-aged  Hawaiian  Stilt 
chicks  up  to  and  including  six  weeks  of  age. 


E 

E 


M 

3 

(A 


E 

E_ 

c 

(U 

E 

3 

o 


160i 


120 


80 


40- 


a □ 

□ □ oa° 

^>13  ° 


Q f □ 


a„Eb  a 


0 

160- 

120- 

80- 

40 


4 0 


8 0 


1 20 


1 60 


f □ % 


P5 


4 0 


8 0 


120  160 


E 

E 

160  - 

■a 

k. 

o 

120  - 

o 

80  - 

O) 

c 

40  - 

% 

□ B 


pa 


40  80  120 

Mass  (g) 


160 


FIG.  3.  Relative  growth  rates  of  three  body  mea- 
surements of  wild  Hawaiian  Stilt  chicks  compared  to 
body  mass. 


From  4-6  weeks,  age  classes  can  be  differ- 
entiated by  the  presence  of  tail  feathers,  the 
ability  to  fly,  and  the  presence  of  an  eye  patch 
and  eye  ring.  In  week  4,  tail  feathers  emerge 
and  the  eye  patch  and  eye  ring  become  visible. 
During  week  5,  all  down  is  lost,  wing  feathers 
are  fully  developed  enabling  short  distance 
flight  (up  to  1.5  m),  and  the  eye  patch  is  dis- 
tinct. Finally  in  week  6,  chicks  are  capable  of 
prolonged  flight. 

As  with  other  shorebirds  (Clark  1961),  the 
egg  tooth  typically  was  lost  after  the  first  day 
and  always  was  gone  after  48  h. 

DISCUSSION 

Because  shorebird  chicks  feed  themselves, 
they  hatch  with  well  developed  legs  and  a 


482 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4.  December  1999 


(U 

cc 


D 


W) 

c 


o 

o 

o 

00 

O 

S 

CJ 


CO 


a 

a: 

2 

'5 


o 

« 

u 

-C 

u 

(U 

W) 

C3 

s 


UJ 

CQ 

< 

f- 


CQ 


■o  ^ 

I ^ 

E o 


c 

* <u 

^ 2 

o 

Oi) 


o •- 

C«  £ 

s ^ 


c 

CJ 

2 15 

M § 

^ (U 

o 

cd  • '' 

<u 


.ti 


j=  ^ 
^ 2 


^ u 

S ^ 


cd 

(U 


c 

1) 

o i- 

OJ 

^ > 

0 

cd  • »• 

3 

1 ”1 
s 

- i 

'5  o £1 


c 

o 


c 

V 

2 

O 

2' 

o 

cd 


c 

CJ 

2 

“o 

W) 

o 

cd 


■o 

o 


T3 

V 


o ^ 
B c 

2 o 

c/3  U 

O 

Q 


c 

<u 

2 13 
o h 
wj  c 

^ (U 

> 
u ^ 
cd  • 

— (U 

n w 
■=  U 
"O  ■§  .ti 
u 5 x: 

s ^ * 

ffi" 

c3  o 


C 

:> 


2 


c 
o 

i:;  Ui 

o ^ 
2 


6 

cd 

(U 


cd 

u 

C 

:> 


- TD 

^ X> 

S ^ 

2 o 

C to 

Ui 

3 w 

!l!  c 

O cd 
4= 
W)  ^ 
2 

c -2 
.3  Wj 
13 

1) 

CQ 


c 

(U 

2 13 

o h 

00  C 

r (U 

> 

o 

cd  • ^ 
— (u 


cd 

t-i 

w 

c 

:> 

u 

e 2 

s 

o ^ 

Q 


•o 

OJ 


cd 


JC 

3 

O 


>> 

B 

cd 

V 

u 

CJ 

O 

>. 

cd 


i::  X)  W) 


0) 

o 


•o 

e 

o 


X 

cd 

u 

o 


3 

dj 

2 

2 

00 

ca 

o 

3 


TJ 

<U 


3 

o 

zz  u- 

O ^ 

2 


>> 

S 

cd 

OJ 


3 

;> 


cd 

cd 

cd 

u, 

u 

Uh 

g 

00 

g 

^ a 

00 

cd 

<D 

u 

CJ 

2 

c/3 

U> 

O 

cd 

<D 

u 

o 

X 
c/3  *“ 

0 ^ 

2 

c/3 

o 

Q 

Q 

Q 

X 


<u 

X 


a 

o 


3 

O 


4^  O 
O ^ 
Cd  X 

S 


0) 

o 

3 

3 


s 

o 


X 

CJ 

2 

QQ 


3 

OJ 

2 

2 

00 

O 

cd 


■X 

)U 


3 
O 

X '- 

O x: 

2 


X 

2 

u 

3 

3 

00 

2 

c/3 


o 

Q 


u 

(U 

X 

3 

o 

U 

3 


U 

2 

s 


-u 

§ ^ 

>> 

3 

X 

3 

dJ 

V <€. 

o o 

O 

X 

>> 

>» 

CO  <u 

X 

3 

o 

c ^ 

O 

X 

2 Un 

. ^ 

o 

o 

c/3  ^ 

3 

00 

o ^ 

X 

X 

3 ^ 

3 

0)  X 

(U 

u 

(U 

o 3 

O 

3 

2 

3 

c/3  cd 

o 

4— > 

3 

(U 

c/3 

Ui 

3 

0) 

c/3 

00  T3 
c ^ 
.3  (U 
3 X 

c 

X 

u. 

dj 

Cl 

E 

dJ 

Ui 

dJ 

c 

o 

<D 

3 ^ 

X x 

(U 

(U 

(U 

•3  C 

Cl 

w w 

> 

> 

W)  o 

3 

3 

3 3 

3 

O 

0 

<u 

O 

O 

U 

U 

PQ 

CO 

OI 

m 

»rt 

M 

(U 

dJ 

dJ 

dJ 

<U 

JU 

X 

dj 

X 

X 

Week  6 None  Black  & white  Mostly  black  w/brown  flecks  Dorsal  gray;  Ventral  white  Mottled  black  & golden 

brown 


Reed  el  al.  • HAWAIIAN  STILT  CHICK  GROWTH 


483 


Q 

W 

Q 

Z 

w 

H 

Z 

w 


w 


00 
>>  C 

w-c 


(U  O 

c ca 

<l> 

Ih 

c 

C 

c 

c 

c 

C 2 

3 

Cl 

X 

o 

o 

u 

o 

o 

u 

o 

u 

o 

u 

Z 

m 

CQ 

DQ 

c 

c 

c 

(U 

lU 

CJ 

c/3 

C/) 

c/3 

(U 

CJ 

CJ 

u 

u 

u 

CL 

a 

a 

j_, 

O 

o 

o 

Z 

Z 

Z 

^ . 

c 


O 

Z 


^ o 

— 

5 


c 

(U 

"o 

Oi) 

o 

a 


<D 

5 

6 
*U 
fU 

a 

<u 

£ 


T3  > 

:=  O 

Ui 

O ^ 

S 


0) 


c 

(U 

(/) 

u 

(X 


VO 

>> 
ca 
•O 

/ 

CJ  c/3 

cd  y 


c 

o 


CJ 

cd 


O 

^ 5::^ 
iH 

S 


(U 

= P 

<U  C 
T3  'C 
<u 
Cl 


^ -5 


O 

00 

^ B 
^ ■ 
CJ 

a 


TD 

V 


o 

.2 

S 


c 

(U 

•o 


oa 

CJ 

03 


TD 

1) 


ji:  ^ 
w o 
cd  cii 
o 

pH 


u 

03 


C 

<U 


u 

cd 


c 

a> 


c 

c 

c 

CJ 

CJ 

CJ 

CJ 

2 

c/3 

c/3 

c/3 

• f 

CJ 

CJ 

CJ 

‘w  2 

r; 

w 

u 

u 

u 

CJ  W 

CJ 

w 

a 

CL 

CL 

> X 

.S  X 

c 

X 

o 

O 

O 

to  ^ 

c/3 

Z 

Z 

Z 

Q 

5 

c 

c 

c 

X 

13  O 

w ^ 

00 

O X) 

o 

o 

o 

U) 

CQ 

OP 

u 

CQ 

c/3 

oo 

0 

(N 

X 

X 

CJ 

— 

cd 

"S 

w 

m 

>» 

>> 

CJ 

CJ 

3 

>» 

>. 

V) 

JC 

C 

Cd 

’O 

cd 

-o 

X 

C/) 

X 

c/3 

6 

cd 

c 

_E 

w 

0/ 

CJ 

c/3 

C 

CJ 

C 

CJ 

CJ 

c/3 

CL 

s/5 

CJ 

u 

a 

w 

c 

CJ 

c 

CJ 

X 

o 

u 

X 

0 

u 

CJ 

CJ 

CJ 

3 

CJ 

i~> 

CL 

o 

Z 

c/3 

CJ 

u 

X 

o 

CQ 

X 

c 

X 

c 

[X 

CL 

O 

Z 

C 

CJ 

2 

CZI 

c: 

CJ 

"O  c/3 

c 

CJ 

^3  c/3 

C dj 
dJ  J- 
C/3  C 

3 

CJ 

u> 

u/ 

2 

3 

-S? 

CZl 

C 

CJ 

2 

c/3 

3 

.9- 

o .B- 

3 .& 

CJ 

CJ 

.9- 

o 

.9- 

_ 

00 

■*— ' 

M 

00 

CL  w 

00 

■*— ' 

Tai 

c 

^ I 

^ I 

c/3 

C/3  »H 

L-  CX 

(1^  ^ 

Lh 

X 

CJ 

c 

c 

« O 

^3;  Lh  o j-i 

CQ  X)  X3 

S 


00 

c 

00 

c 

CJ 

9 

3 

2 

3 

cd 

3 

Cd 

JJ 

00 

2 

>% 

>> 

cd 

cd 

u 

00 

X 

CJ 

(h 

^0 

2 

X 

c 

c 

U "u 

o 

00 

0 

o 

cd 

3 

C .. 

c ^ 

03  ^ 


T3 

CJ 


(U 

O 


dJ 

(U  T3 
00 


^s-- 

o i3 
u Z 5 
> -a  ^ 
"'  „ o 

i- 


o -c 


o -C)  <u 


^ N-  SJ 

o -D  lU 

s 


— i..  o =$ 
■a  Z'  x;  ^ 


C/3 


(N 

r<i 

X 

X 

X 

CJ 

CJ 

CJ 

X 

CJ 

»o 

!U 

I 


00 

c 

*>. 

E 

X 

QJ 


484 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4.  December  1999 


partly  developed  bill;  wing  development  be- 
gins later  and  is  rapid  once  started  (Galbraith 
1988,  Thompson  et  al.  1990).  Growth  patterns 
are  variable  among  species  (Holland  and  Yal- 
den  1991;  Starck  and  Ricklefs  1998a,  b).  For 
example,  body  mass  at  fledging  as  a percent- 
age of  asymptotic  adult  body  mass  varies  in 
shorebirds  (Scolopacidae  and  Charadriidae) 
from  53%  to  91%  (Beintema  and  Visser 
1989a).  In  addition,  it  has  been  suggested  that 
shorebirds  have  a higher  growth  coefficient 
{K)  than  other  terrestrial,  precocial  birds 
(Beintema  and  Visser  1989b).  Of  15  species 
reviewed  by  Beintema  and  Visser  (1989a),  K 
ranged  from  0.051  to  0.158,  and  the  inflection 
point  {T)  ranged  from  5.5  to  23.8  d after 
hatch.  Not  surprisingly,  body  size  is  positively 
correlated  with  the  inflection  point  and  nega- 
tively correlated  with  the  growth  coefficient. 
That  is,  larger  species  reach  the  half-way  point 
in  growth  at  a relatively  larger  size,  and  grow 
at  a slower  rate  in  proportion  to  their  adult 
body  size,  than  do  smaller  species.  Hawaiian 
Stilts  conform  to  these  patterns. 

Shape  of  growth  curves. — Captive  Hawai- 
ian Stilt  chicks  grew  from  approximately  15 
g at  hatching  to  125  g at  fledging,  attaining 
only  60%  of  adult  body  mass  when  they 
fledged.  Culmen,  tarsus,  and  wing  chord  also 
were  still  growing  at  fledging,  well  below 
adult  sizes,  and  did  not  reach  adult  values  un- 
til after  42  days  after  hatching.  Culmen  and 
tarsus  sizes  increased  rapidly  between  hatch- 
ing and  fledging,  with  culmen  growth  gener- 
ally following  a linear  trajectory  and  tarsus 
following  a slightly  sigmoidal  pattern.  Wing 
chord  growth  was  sigmoidal,  with  slow 
growth  from  hatch  day  to  day  12  followed  by 
a substantial  increase  in  growth  rate  when 
chicks  reached  13—15  days  old. 

Mass  loss  in  the  first  24-48  h after  hatch 
has  been  reported  in  some  shorebird  species 
(e.g..  Lapwing,  Vcinellus  vanellus;  Galbraith 
1988)  and  is  attributed  to  movement  away 
from  the  nest  cup  soon  after  hatching.  Al- 
though Hawaiian  Stilts  also  leave  their  nest 
cup  within  a day  of  hatching,  we  found  no 
significant  mass  loss  for  captive  or  wild  chicks 
from  day  of  hatch  to  day  1.  Differences  in  the 
distance  traveled  and  the  amount  of  food 
available  in  the  first  24  h may  explain  inter- 
specific and  intraspecific  differences  in  shore- 
bird  mass  loss  immediately  after  hatching. 


Reasons  for  variation  in  shorebird  post-hatch- 
ing mass  loss  require  further  investigation. 

Comparison  of  captive  and  wild  chick 
growth. — Captive  and  wild  chick  masses  did 
not  differ  significantly  for  most  ages;  when 
they  differed,  wild  birds  were  lighter  than  cap- 
tive birds.  By  the  end  of  week  1 captive 
chicks  generally  were  growing  at  a faster  rate 
than  wild  chicks  for  all  growth  parameters 
measured.  This  trend  mirrors  results  from  oth- 
er studies  of  precocial  birds  (Beintema  and 
Visser  1989a).  In  most  cases,  captive  and  wild 
chicks  have  similar  growth  curves,  with  more 
variation  in  the  growth  of  wild  chicks  (Visser 
and  Ricklefs  1993).  Faster  growth  in  captivity 
could  be  due  to  an  unlimited  food  supply, 
while  slower  growth  in  the  wild  could  be  at- 
tributed to  colder  weather,  which  increases  the 
costs  of  thermoregulation  and  reduces  the 
amount  of  time  that  chicks  can  spend  forag- 
ing. A study  of  time  budgets  in  the  field  of 
three  precocial  charadriiform  species  revealed 
that  during  adverse  weather,  young  chicks 
were  brooded  for  75%  of  the  daytime,  and  as 
a result,  they  could  not  obtain  enough  food  to 
satisfy  their  energy  requirements  (Beintema 
and  Visser  1989a).  In  contrast,  during  good 
weather  conditions,  chicks  foraged  almost 
continuously  once  they  were  able  to  thermo- 
regulate. 

Beintema  and  Visser  (1989a,  b)  hypothe- 
sized that  for  shorebird  species,  cold  temper- 
atures and  cold  with  rain  are  the  main  causes 
of  slower  chick  growth  in  the  wild.  Specifi- 
cally, temperatures  dropping  below  15°  C 
slowed  chick  growth.  In  Hawaii,  temperatures 
in  coastal  wetlands  where  Hawaiian  Stilts 
breed  rarely  fall  below  21°  C,  and  there  are  no 
records  of  temperatures  as  low  as  15°  C.  In 
addition,  rains  at  coastal  areas  typically  are 
short-lived.  The  fact  that  growth  was  slower 
in  wild  chicks  despite  temperatures  above 
15°  C suggests  that  temperature  itself  is  not 
the  main  factor  affecting  slower  Hawaiian 
Stilt  chick  growth  in  the  field.  At  warmer  tem- 
peratures, Pierce  (1986)  observed  faster 
growth  in  other  stilt  species.  Either  a different 
threshold  applies  to  Hawaiian  Stilts  or  differ- 
ences were  due  to  food  availability  (Beintema 
1994). 

Comparison  to  other  species. — Hawaiian 
Stilts  grow  slowly  in  comparison  to  other 
shorebirds.  Of  the  42  growth  coefficients 


Reed  et  al.  • HAWAIIAN  STILT  CHICK  GROWTH 


485 


Starck  and  Ricklefs  (1998a)  reported  for  27 
species  of  shorebird,  only  5 were  lower  than 
what  we  calculated  for  Hawaiian  Stilts,  and 
all  came  from  heavier  species.  The  only  pub- 
lished estimates  of  Himantopus  growth  coef- 
ficients are  Starck’s  and  Ricklefs’  (1998a)  cal- 
culations from  Pierce’s  (1986)  data  on  Pied 
{Himantopus  himantopus  leucocephalus)  and 
Black  {H.  novaezealandiae)  stilts.  These  spe- 
cies have  lower  adult  masses  (129  g and  130 
g,  respectively)  than  do  Hawaiian  Stilts,  but 
do  not  fledge  until  a later  age.  Hawaiian  Stilts 
fledge  approximately  28  days  after  hatching; 
Pied  Stilt  chicks  do  not  fledge  until  they  are 
34  d,  and  the  Black  Stilt  fledges  even  later  (at 
46  d;  Pierce  1986).  Similar  to  the  Hawaiian 
Stilt,  both  species  continue  to  grow  after 
fledging.  However,  based  on  data  presented  by 
Pierce  (1986:  fig.  6),  Pied  and  Black  stilts 
fledge  at  a higher  percent  of  their  adult  body 
mass.  Consequently,  despite  the  longer  time 
to  fledging.  Pied  and  Black  stilt  growth  co- 
efficients are  consistent  with  expectations 
based  on  their  adult  size  {K  = 0.175  and  0.129 
respectively;  Starck  and  Ricklefs  1998a).  A K 
of  0.074  would  be  expected  for  the  202.5  g 
Hawaiian  Stilt  (Beintema  and  Visser  1989a), 
but  we  observed  K = 0.065  for  Hawaiian 
Stilts  in  captivity  (and  possibly  lower  in  the 
field;  Fig.  1).  Starck  and  Ricklefs  (1998a)  also 
reported  faster  growth  coefficients  for  the  Eu- 
ropean Avocet  (Recurvirostra  avosetta\  K = 
0.213  and  0.171  from  two  different  studies), 
which  is  similar  in  mass  to  Hawaiian  stilts 
(168  g and  250  g,  respectively).  Although  the 
relationship  between  body  m.ass  and  K in 
Charadriiformes,  is  poor  {H  = 8%,  n = 75 
species;  Starck  and  Ricklefs  1998b),  these 
data  demonstrate  that  the  slow  growth  rate  ob- 
served in  Hawaiian  Stilts  is  not  a character- 
istic of  the  Recurvirostridae. 

We  do  not  know  why  Hawaiian  Stilts  have 
slow  growth.  The  two  obvious  hypotheses  do 
not  provide  satisfactory  explanations.  First, 
growth  rate  could  be  correlated  with  latitude. 
Tropical  environments  provide  a longer  breed- 
ing season,  and  growth  rates  of  tropical  altri- 
cial  species  are  lower  than  are  those  of  taxo- 
nomically  related  temperate  species  (Ricklefs 
1976,  Oniki  and  Ricklefs  1981).  The  Hawai- 
ian Stilt  breeding  season  lasts  six  months 
(Coleman  1981).  Despite  this,  neither  the  in- 
cubation nor  fledging  period  is  prolonged. 


Worldwide,  stilts  average  22-26  days  of  in- 
cubation (Johnsgard  1981),  which  incorpo- 
rates the  Hawaiian  Stilt’s  incubation  length  of 
25  days  (Colemen  1981).  As  noted  above,  the 
fledging  time  is  shorter  in  this  species  than  in 
others  of  its  genus  (Johnsgard  1981,  Pierce 
1986)  so  there  is  no  extended  time  as  a chick. 
There  are  no  studies  of  which  we  are  aware 
comparing  growth  rates  of  precocial  species 
across  a latitudinal  gradient,  but  it  would  be 
an  interesting  assessment. 

Second,  the  lower  growth  rate  could  be  a 
consequence  of  evolving  in  an  island  environ- 
ment where  predation  rates  might  have  been 
relatively  low  before  human  occupancy,  and 
selection  for  rapid  growth  might  have  been 
relaxed.  Most  recorded  mortality  of  adult  Ha- 
waiian Stilts  is  attributed  to  introduced  species 
(Woodside  1979).  However,  one  would  expect 
slower  growth  to  be  associated  with  an  older 
age  at  fledging,  which  does  not  occur.  In  con- 
trast, the  Hawaiian  Stilt  fledges  at  a smaller 
percent  of  adult  body  mass  than  do  other  stilts, 
resulting  in  an  extended  post-fledging  growth 
period. 

Estimating  age. — Ideally,  estimates  of 
chick  age  would  be  based  on  a trait  that 
changes  rapidly  and  monotonically  throughout 
growth.  One  problem  with  this  method  is  that 
often  no  one  trait  is  ideal  throughout  the  entire 
growth  period.  Rather,  traits  differ  in  their  ac- 
curacy for  aging  as  chicks  become  older.  For 
example,  measurements  of  tarsus  and  wing 
chord  for  Hawaiian  Stilts  are  not  useful  for 
aging  chicks  at  early  and  late  ages  because  of 
their  sigmoidal  growth  patterns.  Using  mass 
as  an  indicator  of  chick  age  is  problematic  be- 
cause it  fluctuates  rapidly,  depending  on  en- 
vironmental conditions  and  when  chicks  are 
weighed  in  relation  to  their  last  feeding.  For 
Hawaiian  Stilts,  culmen  length  may  be  the 
most  useful  parameter  for  aging  chicks  be- 
cause its  growth  trajectory  is  fairly  linear.  Be- 
cause it  typically  has  a constant  growth  rate 
throughout  the  chick  stage,  culmen  length  has 
been  used  to  age  chicks  of  other  shorebird 
species  in  the  wild  (Beintema  and  Visser 
1989a).  However,  even  for  traits  that  tend  to 
vary  linearly  and  monotonically  throughout 
development,  there  is  a tremendous  amount  of 
individual  variation  in  daily  growth.  Unfor- 
tunately, this  individual  variation  is  magnified 
by  measurement  error  when  all  measurements 


486 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


are  not  made  by  the  same  person.  Thus,  de- 
termination of  chick  age  using  body  measure- 
ments and  mass,  regardless  of  the  species, 
should  be  viewed  as  approximate  at  best. 

As  a result,  we  decided  to  describe  general 
plumage  patterns  for  Hawaiian  Stilt  chicks  of 
known  age  in  the  field  to  set  up  criteria  for 
establishing  weekly  age  classes  for  chicks,  de- 
fined by  the  presence  or  absence  of  specific 
plumage  characteristics.  The  ability  to  identify 
approximate  chick  age  in  the  field  without 
capturing  young  of  this  endangered  species 
could  aid  in  management  by  helping  to  iden- 
tify the  age  at  which  chicks  disappear.  To  this 
end,  we  found  definitive  differences  between 
plumage  characteristics  of  specific  age  classes 
of  Hawaiian  Stilts.  This  should  increase  abil- 
ities to  assess  survival,  the  least  understood 
life-history  component  of  this  species  (Reed 
et  al.  1998). 

We  found  plumage  characteristics  to  be  use- 
ful for  identifying  weekly  age  classes  of  Ha- 
waiian Stilts.  Because  culmen  length  is  similar 
for  captive  and  wild  chicks  and  has  a linear 
positive  relationship  with  age  throughout  de- 
velopment, a combination  of  culmen  length 
and  plumage  description  may  be  the  most  ac- 
curate way  to  age  wild  Hawaiian  Stilt  chicks. 
Relying  primarily  on  plumage  characteristics, 
specifically  because  they  are  non-invasive, 
and  supplementing  these  observations  with 
culmen  lengths  if  chicks  are  captured,  will 
help  minimize  interference  in  this  endangered 
species  while  providing  managers  with  a tool 
for  monitoring  reproductive  success  and  pop- 
ulation numbers.  Because  adults  and  fledg- 
lings differ  in  plumage  patterns,  it  also  will 
allow  accurate  monitoring  of  reproductive 
success  before  molt. 

ACKNOWLEDGMENTS 

We  thank  the  United  States  Fish  and  Wildlife  Ser- 
vice’s Pacific  Refuges  Office,  the  Hawaii  State  Divi- 
sion of  Forestry  and  Wildlife,  and  the  Marine  Corps 
Base  Hawaii,  Kaneohe  Bay  for  support  and  use  of  their 
facilities,  and  M.  Silbernagle  and  C.  Terry  for  logistic 
support.  This  manuscript  benefited  from  reviews  by  C. 
Elphick,  N.  Warnock,  R.  Ricklefs,  and  two  anonymous 
reviewers.  Research  was  supported  by  a grant  from  the 
North  DakoUi/National  Science  Foundation  EPSCoR 
to  LWO,  by  National  Science  Foundation  grants  DEB 
9322733  fto  J.M.R.  and  L.W.O.)  and  DEB  9424375 
(to  L.W.O.),  and  by  the  U.S.  Dept,  of  Agriculture. 


LITERATURE  CITED 

Anthony,  N.  B.,  D.  A.  Emmerson,  K.  E.  Nestor,  W. 
L.  Bacon,  P.  B.  StECEL,  and  E.  A.  Dunnington. 
1991.  Comparison  of  growth  curves  of  weight  se- 
lected populations  of  turkey,  quail,  and  chickens. 
Poultry  Sci.  70:13-19. 

Beintema,  a.  J.  1994.  Condition  indices  for  wader 
chicks  derived  from  body-weight  and  bill-length. 
Bird  Study  41:68-75. 

Beintema,  A.  J.  and  G.  H.  Visser.  1989a.  Growth  pa- 
rameters on  chicks  of  charadriiform  birds.  Ardea 
77:169-180. 

Beintema,  A.  J.  and  G.  H.  Visser.  1989b.  The  effect 
of  weather  on  time  budgets  and  development  of 
chicks  of  meadow  birds.  Ardea  77:181-192. 

Clark,  G.  A.,  Jr.  1961.  Occurrence  and  timing  of  egg 
teeth  in  birds.  Wilson  Bull.  73:268—278. 

Coleman,  R.  A.  1981.  The  reproductive  biology  of  the 
Hawaiian  subspecies  of  the  Black-necked  Stilt, 
Himantopus  mexicanu.s  knudseni.  Ph.D.  diss., 
Pennsylvania  State  Univ.,  State  College. 

CoocH,  E.  G.,  D.  B.  Lank,  A.  Dzubin,  R.  E Rockwell, 
AND  E Cooke.  1991.  Body  size  variation  in  Lesser 
Snow  Geese:  environmental  plasticity  in  gosling 
growth  rates.  Ecology  72:503-512. 

Emlen,  S.  T,  P.  H.  Wrege,  N.  J.  Demong,  and  R.  E. 
Hegner.  1991.  Flexible  growth  rates  in  nestling 
White-fronted  Bee-eaters:  a possible  adaptation  to 
short-term  food  shortage.  Condor  93:591-597. 

Galbraith,  H.  1988.  Adaptation  and  constraint  in  the 
growth  pattern  of  Lapwing  Vanellus  vanellu.s 
chicks.  J.  Zool.  Lond.  215:537—548. 

Holland,  P.  K.  and  D.  W.  Yalden.  1991.  Growth  of 
Common  Sandpiper  chicks.  Wader  Study  Group 
Bull.  62:13-15. 

JoHNSGARD,  P.  A.  1981.  The  plovers,  sandpipers,  and 
snipes  of  the  world.  Univ.  of  Nebraska  Press,  Lin- 
coln. 

O’Connor,  R.  J.  1984.  Growth  and  development  of 
birds.  Wiley,  New  York. 

Oniki,  Y.  and  R.  E.  Ricklefs.  1981.  More  growth  rates 
of  birds  in  the  humid  New  World  tropics.  Ibis  123: 
349-354. 

Pierce,  R.  J.  1986.  Differences  in  susceptibility  to  pre- 
dation during  nesting  between  Pied  and  Black 
stilts  (Himantopus  spp.).  Auk  103:273-280. 

Reed,  J.  M.,  C.  S.  Elphick,  and  L.  W.  Oring.  1998. 
Life-history  and  viability  analysis  of  the  endan- 
gered Hawaiian  Stilt.  Biol.  Conserv.  84:34—45. 

Reed,  J.  M.  and  L.  W.  Oring.  1993.  Long-term  pop- 
ulation trends  of  the  endangered  Ae’o  (Hawaiian 
Stilt,  Himantopus  mexicanus  knudseni).  Trans. 
West.  Wild.  Soc.  29:54-60. 

Ricklefs,  R.  E.  1968.  Patterns  of  growth  in  birds.  Ibis 
1 10:419-451. 

Ricklefs,  R.  E.  1973.  Patterns  of  growth  in  birds.  11. 
Growth  rate  and  mode  of  development.  Ibis  1 15: 
177-200. 

Ricklefs,  R.  E.  1976.  Growth  rate  of  birds  in  the  hu- 
mid New  World  tropics.  Ibis  118:176-207. 


Reed  el  al.  • HAWAIIAN  STILT  CHICK  GROWTH 


487 


Ricklefs,  R.  E.  1983.  Avian  po.stnatal  development. 
Avian  Biol.  7: 1-83. 

Robinson,  J.  A.,  J.  M.  Reed,  J.  R Skorupa,  and  L.  W. 
Oring.  In  press.  Black-necked  Stilt  (Himantopus 
mexicanus).  In  The  birds  of  North  America  (A. 
Poole  and  E Gill,  Eds.).  The  Academy  of  Natural 
Sciences,  Philadelphia,  Pennsylvania;  The  Amer- 
ican Ornithologists’  Union,  Washington,  D.C. 

SPSS,  Inc.  1995.  SPSS  for  Windows.  Base  system 
user’s  guide.  Release  7.0  ed.  SPSS  Inc.,  Chicago, 
Illinois. 

Starck,  j.  M.  and  R.  E.  Ricklefs.  1998a.  Avian 
growth  rate  data  set.  Pp.  381-423  in  Avian  growth 
and  development:  evolution  within  the  altricial- 
precocial  spectrum  (J.  M.  Starck  and  R.  E.  Rick- 
lefs, Eds.).  Oxford  Univ.  Press,  New  York. 

Starck,  J.  M.  and  R.  E.  Ricklefs.  1998b.  Variation, 
constraint,  and  phylogeny:  comparative  analysis 
of  variation  in  growth.  Pp.  247-265  in  Avian 


growth  and  development:  evolution  within  the  al- 
tricial-precocial  spectrum  (J.  M.  Starck  and  R.  E. 
Ricklefs,  Eds.).  Oxford  Univ.  Press,  New  York. 

Thompson,  P.  S.,  C.  McCarty,  and  W.  G.  Hale.  1990. 
Growth  and  development  of  Redshank  Trinf^a  lo- 
taniis  chicks  on  the  Ribble  saltmarshes,  N.W.  Eng- 
land. Ringing  & Migr.  11:57-64. 

U.S.F.W.S.  1985.  Hawaiian  waterbirds  recovery  plan. 
U.S.  Fish  and  Wild!.  Serv.,  Portland,  Oregon. 

VissER,  G.  H.  AND  R.  E.  Ricklefs.  1993.  Development 
of  temperature  regulation  in  shorebirds.  Physiol. 
Zool.  66:771-782. 

Wilson,  S.  B.  and  A.  H.  Evans.  1893.  Aves  Ha- 
waiienses:  birds  of  the  Sandwich  Islands.  Arno 
Press  (1974),  New  York. 

WooDSiDE,  D.  H.  1979.  Limited  study  of  nesting  by 
stilt  on  the  islands  of  Maui,  Oahu  and  Kauai.  Pro- 
gress Report,  Job  No.  R-III-C,  Project  No.  W-18- 
R-4.  Division  of  Forestry  and  Wildlife,  Honolulu, 
Hawaii. 


Wilson  Bull.,  111(4),  1999,  pp.  488-493 


NESTING  BEHAVIOR  OF  THE  LILAC-CROWNED  PARROT 

KATHERINE  RENTON'  ^ AND  ALEJANDRO  SALINAS -MELGOZA^ 


ABSTRACT. — Nesting  behavior  of  the  Lilac-crowned  Parrot  (Amazona  finschi)  was  observed  over  a three 
year  period  at  24  nests  in  the  tropical  dry  forest  of  the  Chamela-Cuixmala  Biosphere  Reserve,  western  Mexico. 
Nest  site  characteristics  and  the  pattern  of  parental  care  throughout  the  nesting  cycle  are  described  for  this 
mainland  Amazon  parrot  and  compared  with  that  reported  for  other  Amazon  parrot  species.  Nest  sites  were 
located  in  natural  cavities  of  large  mature  trees  characteristic  of  semi-deciduous  forest.  Nest  sites  were  similar 
to  one  another  in  tree  species,  tree  size,  cavity  height,  and  entrance  width,  indicating  that  Lilac-crowned  Parrots 
may  select  nest  sites  based  on  these  characteristics.  Unlike  most  parrot  species.  Lilac-crowned  Parrots  showed 
low  nest  site  reuse  and  high  synchrony  of  nest  initiation.  Throughout  the  nesting  cycle,  females  and  nestlings 
were  fed  only  twice  a day  on  average.  Nest  attendance  during  feeding  visits  was  short.  The  infrequent  feeding 
visits  and  short  nest  attendance  exhibited  by  Lilac-crowned  Parrots  corresponds  with  that  found  for  other  main- 
land Amazon  parrots  in  northeastern  Mexico,  but  contrasts  with  the  multiple  feedings  and  longer  nest  attendance 
observed  for  island  Amazon  species.  The  distinct  aspects  of  Lilac-crowned  Parrot  nesting  behavior  may  be 
related  to  predation  rate  and  food  resource  availability  during  the  extreme  dry  season.  Received  10  March  1999, 
accepted  15  July  1999. 


Most  studies  on  the  ecology  of  Neotropical 
Amazon  parrots  have  been  conducted  on  spe- 
cies inhabiting  the  Caribbean  Islands  (Snyder 
et  al.  1987,  Gnam  1991,  Gnam  and  Rockwell 
1991,  Wilson  et  al.  1995).  A comparative 
study  of  three  species  of  mainland  Amazon 
parrot  in  north-eastern  Mexico  found  distinct 
differences  compared  with  island  species  in 
some  aspects  of  reproductive  behavior  and 
productivity  (Enkerlin-Hoeflich  1995).  Fe- 
males and  nestlings  of  three  mainland  Ama- 
zon parrot  species  were  fed  only  twice  a day 
(Enkerlin-Hoeflich  1995),  in  comparison  with 
the  multiple  feedings  observed  for  Caribbean 
Amazons  (Snyder  et  al.  1987,  Gnam  1991, 
Wilson  et  al.  1995)  and  other  Neotropical  par- 
rots (Fanning  1991,  Waltman  and  Beissinger 
1992).  Additional  data  on  mainland  Amazon 
parrots  are  needed  to  determine  whether  con- 
clusions from  studies  on  island  species  are  ap- 
plicable to  mainland  species.  The  Amazon 
parrot  species  of  Mexico  are  particularly  suit- 
ed for  comparison  with  the  Amazona  species 
of  the  Greater  Antilles  in  the  Caribbean  be- 
cause of  their  close  evolutionary  relationship 
(Snyder  et  al.  1987,  Forshaw  1989).  The  Li- 
lac-crowned  Parrot  (Amazona  finschi)  is  en- 


' Durrell  Institute  of  Con.servation  & Ecology,  The 
Univ.  of  Kent  at  Canterbury,  Kent,  CT2  7NJ,  U.K. 

^ Fundacion  Ecologica  de  Cuixmala,  Apartado  Post- 
al 161 , .San  Patricio-Melaque,  Jalisco,  CP  48980,  Mex- 
ico. 

’ Corresponding  author; 

E-mail;  fundacion_ecologica@gdl.icanet.net.mx 


demic  to  western  Mexico  and  has  a restricted 
distribution  from  southeastern  Sonora  to  Oa- 
xaca (Forshaw  1989).  There  have  been  no 
studies  on  the  ecology  of  the  Lilac-crowned 
Parrot,  and  little  is  known  of  its  breeding  bi- 
ology (Forshaw  1989).  Anecdotal  reports 
from  captive  breeding  give  an  incubation  pe- 
riod of  28  days,  with  the  young  chick  leaving 
the  nest  after  60  days  (Mann  and  Mann  1978). 
In  this  paper  we  present  observations  on  the 
nest  site  requirements  and  nesting  behavior  of 
the  Lilac-crowned  Parrot  in  the  wild,  and 
compare  them  to  observations  for  other  island 
and  mainland  Amazon  parrots. 

STUDY  AREA  AND  METHODS 

Studies  on  the  breeding  biology  of  the  Lilac- 
crowned  Parrot  were  conducted  at  the  13,142  ha  Cha- 
mela-Cuixmala Biosphere  Reserve  (19°  22'  N, 
104°  56'  W to  19°  35'  N,  105°  03'  W)  on  the  Pacific 
coast  of  Mexico.  The  study  site  has  a dry  tropical 
climate  exhibiting  a marked  seasonality  in  precipita- 
tion, with  80%  of  the  748  mm  average  annual  rainfall 
occurring  June  to  November,  and  a prolonged  drought 
from  mid-February  to  late  May  (Bullock  1986).  The 
reserve  has  a hilly  topography  varying  in  elevation 
from  20-520  m above  sea  level.  The  dominant  veg- 
etation type  on  the  slopes  is  tropical  dry  deciduous 
forest,  with  semi-deciduous  forest  in  the  larger  drain- 
ages and  more  humid  valleys  (Lott  et  al.  1987,  Lott 
1993).  Monospecific  forests  of  Celaenodendron  mex- 
icaniun  also  occur  as  discontinuous  patches  within 
the  tropical  deciduous  forest  mosaic  (Martijena  and 
Bullock  1994). 

Observations  on  the  nesting  behavior  of  Lilac- 
crowned  Parrots  were  conducted  from  January  to  June 
in  1996-1998.  Nest  searches  were  earned  out  in  Feb- 


488 


Renum  and  Salinas-Melf>oza  • LILAC-CROWNED  PARROT  NESTING  BEHAVIOR 


489 


TABLE  1. 

Cavity  dimensions  for  26  Lilac-crowned  Parrot  nests. 

Nest  character 

Mean 

Standard  deviation 

Range 

Coefficient  of  variation 
(s/m)  X 100 

Tree  diameter  at  brea.st  height  (cm) 

43.1 

1 1.4 

27.7-66.3 

26.3% 

Cavity  height  from  ground  (m) 

9.7 

1.7 

7.4-14.7 

17.8% 

Entrance  width  (cm) 

10.0 

2.3 

6.4-14.0 

22.5% 

Entrance  length  (cm) 

21.0 

16.0 

7.5-71.6 

76.2% 

Internal  diameter  (cm) 

19.9 

6.8 

10.5-35.0 

34.3% 

Cavity  depth  (cm) 

66.2 

51.7 

24-260 

78.1% 

Circumference  at  entrance  (cm) 

86.7 

30.0 

26-135 

34.5% 

Neare.st  active  nest  (m) 

948.9 

707.7 

25-2419 

74.6% 

ruary  during  the  nest  prospecting  and  early  incubation 
phases  of  the  parrot  breeding  cycle.  No  additional 
nests  were  located  later  in  the  nesting  cycle  because 
the  behavior  of  breeding  pairs  made  detection  of  nest 
sites  difficult.  A cavity  was  considered  a potential  nest 
site  if  one  or  both  of  the  adult  panots  were  observed 
entering  it.  The  cavity  was  considered  an  active  nest 
site  if  one  of  the  adult  parrots  remained  within  the 
cavity  for  longer  than  30  min.  Nest  site  reuse  was  de- 
termined from  the  frequency  of  cavity  occupancy  be- 
tween years. 

Access  to  nest  cavities  was  achieved  using  both  sin- 
gle-rope ascending  (Perry  1978,  Perry  and  Williams 
1981)  and  a tree  bole  climbing  technique  (Donahue 
and  Wood  1995).  Nesting  requirements  of  the  Lilac- 
crowned  Parrot  were  determined  by  measurement  of 
nest  cavity  dimensions:  tree  species,  diameter  at  breast 
height  (DBH)  of  the  tree,  height  above  ground  of  the 
entrance,  width  and  length  of  entrance,  cavity  depth, 
internal  diameter,  and  circumference  of  the  tree  at  en- 
trance (Saunders  1979,  Saunders  et  al.  1982).  The  lo- 
cation of  each  nest  site  and  where  possible  the  tree 
used  by  the  nesting  pair  for  the  transfer  of  food  from 
the  male  to  the  female  were  obtained  using  a geo- 
graphic positioning  system.  The  coefficient  of  varia- 
tion was  determined  for  the  mean  cavity  dimensions 
to  evaluate  the  variability  of  characteristics  between 
nest  sites. 

Behavior  of  breeding  pairs  was  determined  by  ob- 
servations of  parrot  nests  from  covered  blinds  using 
10  X 40  binoculars.  Continuous  dawn  to  dusk  obser- 
vations were  conducted  on  30  man-days  (360  hours) 
at  8 nests.  No  activity  was  observed  at  nests  during 
mid-day;  therefore  additional  observations  were  re- 
stricted to  the  first  four  hours  after  sunrise  and  the  last 
three  hours  prior  to  sunset  giving  an  additional  299 
hours  of  observation  at  16  nests.  Parental  care  and  in- 
vestment was  evaluated  from  the  number  of  feeding 
visits  to  the  nest,  arrival  time,  duration  of  feeding  visit, 
time  spent  in  the  nest  cavity,  and  time  spent  in  the  nest 
area  (defined  as  within  100  m of  the  nest).  Descriptive 
statistics  are  presented  with  means,  ranges,  and  stan- 
dard deviations. 


RESULTS 

Nest  site  characteristics. — A total  of  29 
nest  sites  were  located  in  1995-1998,  all  of 
which  occurred  in  natural  cavities.  Nest  cav- 
ities were  located  in  live  trees  of  Celaenoden- 
dron  rnexicanum,  local  name  Guayabillo 
(51.7%,  n = 15),  and  Astronium  graveolens, 
local  name  Culebro  (31.0%,  n = 9).  Of  the 
remaining  5 cavities,  2 were  located  in  a Ta- 
bebua  species,  1 was  located  in  a dead  tree, 
and  2 were  located  in  unidentified  trees.  Nest 
site  reuse  was  low,  with  only  3 (10.3%)  of  the 
29  nests  sites  located  between  1995  and  1998 
being  used  by  nesting  pairs  over  more  than 
one  breeding  season.  One  cavity  was  used  in 
three  of  the  four  years;  1995,  1997,  and  1998. 
Two  other  cavities  were  reused  once  after  a 
vacancy  of  one  year. 

Mean  cavity  dimensions  for  26  active  nest 
sites  are  presented  in  Table  1;  three  cavities 
could  not  be  accessed  for  safety  reasons.  The 
cavity  dimensions  with  the  least  variation 
were  height  of  entrance  from  the  ground  and 
width  of  entrance.  Diameter  of  tree  at  breast 
height  was  relatively  consistent  between  nest 
sites  and  reflects  the  fact  that  parrot  nests  were 
located  in  large,  mature  trees  characteristic  of 
semi-deciduous  forest.  The  greatest  variability 
was  found  in  depth  of  cavity  and  length  of 
entrance. 

Egg-laying  and  incubation. — Timing  of 
egg-laying  was  highly  synchronized  between 
nests  with  most  pairs  commencing  incubation 
within  14  days  of  the  first  nest  being  initiat- 
ed. Mean  nest  initiation  date  was  6 February 
± 4.6  (SD)  days  in  1996  (range:  30  January- 
13  February,  n — 8),  and  15  February  ±5.3 


490 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


days  in  1997  (range:  10-23  February,  n = 
6). 

Nest  attendance  by  the  female  was  high 
during  incubation,  with  the  female  leaving  the 
nest  only  once  in  the  morning  and  once  in  the 
afternoon  to  be  fed  by  the  male.  The  female 
spent  a mean  of  39.4  ± 26.5  min  per  day  out 
of  the  nest  over  both  morning  and  afternoon 
feeding  sessions  (range:  15-95  min  per  day,  n 
= 20).  Conversely,  the  male  was  rarely  ob- 
served entering  the  nest  or  perching  on  the 
nest  rim.  Daily  activity  periods  were  consis- 
tent between  nests  with  the  male  making  an 
average  2.1  ± 0.3  nest  visits  per  day  (range 
2-3  visits,  n — 35)  to  feed  the  female.  Mean 
arrival  times  for  the  morning  and  afternoon 
activity  periods  were  respectively  08:24  EST 
± 46  min  (range:  06:07-10:08,  n — 45)  and 
18:14  EST  ± 26  min  (range:  17:15—19:03,  n 
= 40).  Each  feeding  visit  by  a male  lasted  an 
average  of  33.9  ± 25.6  min  (range:  5-113 
min,  n = 80). 

The  male  usually  vocalized  loudly  on  his 
approach  to  the  nest  area  and  perched  in  a tree 
adjacent  to  the  nest  cavity  making  low  contact 
vocalizations  until  the  female  emerged.  The 
nesting  pair  gave  a characteristic  take-off 
squawk,  or  bugle,  as  the  female  flew  from  the 
nest  cavity  to  join  the  male.  Food  transfer 
from  the  male  to  the  female  took  place  in  a 
regular  perch  tree  located  an  average  423  ± 
228  m (range:  149-983  m,  n = 11)  from  the 
nest  cavity.  The  food  transfer  session  was  the 
only  time  during  the  incubation  phase  when 
both  adults  were  away  from  the  nest  area  and 
was  short  in  duration  (average  11.5  ± 10.4 
min,  range:  2-48  min,  n = 78). 

Parental  care. — Eggs  hatched  asynchro- 
nously, and  females  continued  to  brood  nest- 
lings during  the  day  until  the  oldest  nestling 
was  19.6  ± 2.7  days  old  (range:  15-23  days, 
n = 9).  Females  ceased  roosting  in  nests  over- 
night when  the  youngest  chick  was  in  its  third 
week.  During  this  early  nestling  phase,  the 
male  continued  to  feed  the  female  twice  a day 
and  was  occasionally  observed  to  enter  the 
nest  for  a mean  1.7  ± 2.8  min  (range:  0—8 
min,  n — 9)  per  feeding  visit.  Later  in  the 
nesting  cycle  when  the  chicks  were  larger, 
both  parents  entered  the  nest  to  feed  the 
young. 

The  behavior  of  nesting  pairs  altered  once 
the  female  began  to  forage  with  the  male. 


Nesting  pairs  became  more  secretive  around 
nests,  arriving  and  departing  silently.  Pairs 
used  low,  almost  inaudible  vocalizations  when 
in  the  nest  area.  Pairs  were  cautious  about  ap- 
proaching the  actual  nest,  and  would  not  do 
so  if  they  detected  an  observer  or  another  dis- 
turbance. The  nesting  pair  made  an  average 
2.6  ± 0.9  visits  per  day  (range:  2-4  visits,  n 
= 25)  to  feed  the  nestlings.  Average  duration 
of  feeding  visits  during  the  nestling  phase  was 
72.3  ± 42.3  min  (range:  12-171  min,  n = 31). 
However,  the  nesting  pair  spent  the  majority 
of  this  time  perched  in  trees  around  the  nest 
area.  Attendance  at  the  actual  nest  cavity  was 
short,  lasting  an  average  of  10.6  ± 11.2  min 
[range:  1.0-27.2  min,  n = 30  (total  time  either 
adult  in  nest  cavity  or  at  entrance)],  with  a 
mean  of  5.4  ± 4.5  min  per  visit  (range:  0-17 
min,  n — 30)  spent  within  the  nest  cavity,  and 
a mean  of  6.0  ± 12  min  (range:  0-20.6  min, 
n = 30)  perched  at  the  nest  rim.  Each  adult 
spent  a mean  of  only  4.0  ± 3.5  min  (range: 
0.42-17  min,  n = 41)  inside  the  nest  cavity 
per  feeding  visit.  This  was  sufficient  time  to 
feed  the  young;  however,  there  was  no  indi- 
cation that  parent  birds  spent  any  other  time 
in  the  nest  with  the  young  except  when  feed- 
ing them. 

Prior  to  fledging,  nestlings  began  to  climb 
to  the  nest  entrance,  and  were  fed  at  the  nest 
rim.  During  this  stage,  nesting  pairs  spent 
more  time  perched  near  the  nest  entrance 
making  low  contact  vocalizations  to  the 
young.  Mean  age  at  fledging  was  63.7  ± 3.2 
days  (range:  56-68  days,  n = 22).  Nestlings 
fledged  asynchronously,  and  all  nests  fledged 
young  within  a 2-3  week  period.  All  nestlings 
fledged  within  12  days  in  1996  (mean  fledge 
date  = 10  May  ± 4.34  days,  range:  6-18  May, 
« = 8),  17  days  in  1997  (mean  fledge  date  = 
18  May  ± 6.68  days,  range:  11-28  May,  n = 
7),  and  13  days  in  1998  (mean  fledge  date  = 
8 May  ± 5.16  days,  range:  2-15  May,  n = 
7). 

DISCUSSION 

The  low  variability  between  nest  sites  in 
tree  species,  size,  cavity  height,  and  entrance 
width  suggests  that  Lilac-crowned  Parrots 
may  select  nest  sites  based  on  these  charac- 
teristics. Predation  rates  decrease  with  increas- 
ing height  of  nest  sites  from  the  ground  (Nils- 
son 1984,  Wilcove  1985),  while  the  increased 


Renton  and  Salinas-Mel}>oza  • LILAC-CROWNED  PARROT  NESTING  BEHAVIOR 


491 


size  of  nest  entrance  required  by  large  birds 
may  pose  greater  risks  from  predation,  leading 
to  specific  requirements  for  entrance  dimen- 
sions (Christman  and  Dhondt  1997).  Amazon 
parrots  in  northeastern  Mexico  appear  to  se- 
lect cavities  based  on  tree  species,  cavity 
height,  and  entrance  length  (Enkerlin-Hoeflich 
1995).  Australian  cockatoos  also  demonstrate 
species  specific  requirements  related  to  body 
size  for  entrance  dimensions  and  internal  di- 
ameter of  nest  hollows  (Saunders  et  al.  1982). 
Enkerlin-Hoeflich  (1995)  suggested  that  vari- 
ability in  several  cavity  characteristics  com- 
bined with  narrow  criteria  for  a few  key  char- 
acters may  provide  parrots  with  the  flexibility 
to  exploit  a wide  range  of  available  cavities 
while  limiting  predation  and  competition 
threats.  In  addition,  low  nest  site  reuse  by  Li- 
lac-crowned  Parrots  is  contrary  to  the  30- 
40%  cavity  reuse  observed  for  most  other  par- 
rot species  (Saunders  1982,  Snyder  et  al. 
1987,  Rowley  and  Chapman  1991,  Smith 
1991,  Enkerlin-Hoeflich  1995).  Natural  pre- 
dation is  the  main  cause  of  nest  failure  for 
Lilac-crowned  Parrots  (Renton  1998),  hence 
infrequent  cavity  reuse  may  help  to  prevent 
predators  from  learning  nest  site  locations 
(Sonerud  1985,  1989). 

The  Lilac-crowned  Parrot  is  notably  differ- 
ent from  other  parrot  species  in  its  high  syn- 
chrony in  nest  initiation,  with  all  nests  com- 
mencing within  two  weeks  in  each  season, 
and  the  general  nest  initiation  period  compris- 
ing the  first  three  weeks  in  February.  There  is 
no  evidence  that  breeding  pairs  of  the  Lilac- 
crowned  Parrot  relay  after  a nest  failure, 
which  would  also  lengthen  the  nesting  period. 
Most  Amazona  species  have  a 3-5  week  nest 
initiation  period  each  breeding  season  (Snyder 
et  al.  1987,  Gnam  1991,  Enkerlin-Hoeflich 
1995).  The  Monk  Parakeet  (Myiopsitta  mon- 
achus)  in  Argentina  extends  egg-laying  over 
a nine  week  period  (Navarro  et  al.  1992).  Aus- 
tralian cockatoos  have  a similar  broad  egg- 
laying  period  of  5-8  weeks  (Saunders  1982, 
Smith  and  Saunders  1986,  Rowley  and  Chap- 
man 1991,  Smith  1991).  The  nesting  season 
of  the  Lilac-crowned  Parrot  may  be  so  sharply 
defined  by  the  extreme  climatic  seasonality  in 
tropical  deciduous  forest  and  food  resource 
availability.  Nesting  pairs  may  need  to  fledge 
young  before  the  end  of  the  long  dry  season 
in  late  May-June  when  food  abundance  de- 


clines (Renton  1998).  Delaying  nest  initiation 
may  result  in  breeding  pairs  having  to  conduct 
energetically  demanding  activities  of  raising 
young  during  this  environmentally  difficult 
period. 

The  infrequent  feeding  visits  to  the  nest  by 
breeding  pairs  of  the  Lilac-crowned  Parrot 
contrasts  with  the  multiple  daily  feedings  not- 
ed for  island  Amazona  species  (Snyder  et  al. 
1987,  Gnam  1991)  and  other  Neotropical  par- 
rots (Lanning  and  Shiflett  1983;  Lanning 
1991;  Waltman  and  Beissinger  1992;  K.R., 
pers.  obs.),  but  is  consistent  with  the  two  nest 
visits  per  day  observed  for  three  mainland 
Amazon  parrots  in  northeastern  Mexico  (En- 
kerlin-Hoeflich 1995).  Morning  and  afternoon 
arrival  times,  approximately  one  hour  after 
sunrise  and  one  hour  before  sunset,  for  nesting 
pairs  of  the  Lilac-crowned  Parrot  were  similar 
to  the  three  Amazona  species  in  northeastern 
Mexico  (Enkerlin-Hoeflich  1995).  Large 
cockatoos  in  dry  areas  of  Australia  also  re- 
strict nest  visitation  activity  to  the  early  morn- 
ing and  late  afternoon,  spending  the  hot,  mid- 
day periods  resting  under  the  shade  of  leafy 
trees  (Saunders  1982).  Lilac-crowned  Parrots 
at  the  study  site  have  been  noted  to  demon- 
strate signs  of  heat  stress  during  the  mid-day 
hours  of  12:00—14:00  by  holding  wings  away 
from  their  bodies  and  panting  with  beaks  open 
(K.R.,  pers.  obs.).  Therefore,  restricting  feed- 
ing activity  to  the  early  morning  and  late  af- 
ternoon may  enable  parrots  to  conserve  en- 
ergy during  high  mid-day  temperatures,  par- 
ticularly in  dry  habitats. 

In  addition  to  being  infrequent,  nest  atten- 
dance by  Lilac-crowned  Parrots  during  feed- 
ing visits  was  brief.  Most  Lilac-crowned  Par- 
rot activity  was  conducted  away  from  the  nest 
area.  Nesting  pairs  were  never  observed  to 
forage  near  the  nest,  and  food  transfers  from 
the  male  to  the  female  took  place  an  average 
423  m from  the  nest.  Island  Amazon  parrots, 
by  comparison,  may  spend  longer  periods  in 
the  nest  cavity  brooding  and  preening  young 
(Snyder  et  al.  1987),  as  well  as  conducting 
food  transfers  and  foraging  activities  near  the 
nest  (Snyder  et  al.  1987,  Gnam  1991).  Infre- 
quent visits,  short  nest  attendance,  and  feeding 
away  from  the  nest  by  Lilac-crowned  Parrots 
may  serve  to  limit  the  amount  of  activity  in 
the  nest  area,  and  reduce  the  risk  of  attracting 
predators  to  the  nest. 


492 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


Breeding  birds  are  restricted  by  competing 
demands  to  maintain  their  own  energetic  re- 
quirements and  to  protect  and  nourish  the 
young  (Martin  1987).  While  infrequent,  short 
nest  visits  may  reduce  both  mid-day  energetic 
expenditures  by  foraging  adults  and  the  risks 
of  nest  predation,  there  is  a cost  in  having  to 
meet  the  energy  demands  of  the  young  in  a 
shorter  time.  Hence,  variations  in  nesting  be- 
havior suggest  that  island  and  mainland  Am- 
azon parrots  may  be  employing  differing  strat- 
egies to  meet  time  constraints  in  caring  for  the 
young. 

ACKNOWLEDGMENTS 

We  are  grateful  to  the  Cuixmala  Ecological  Eoun- 
dation  for  financial  and  logistical  support  during  the 
study,  and  to  the  Mexican  government  department  of 
SEMARNAP  who  granted  permits  for  the  research. 
K.R.  received  a grant  from  the  Durrell  Institute  of  Con- 
servation and  Ecology  and  A.S.M.  was  supported  by 
the  Cuixmala  Ecological  Foundation  and  the  Denver 
Zoological  Foundation.  S.  Berman,  R.  A.  Griffiths,  D. 
V.  Fanning,  B.  Miller,  N.ER.  Snyder,  and  an  anony- 
mous reviewer  provided  constructive  comments  on  the 
manuscript. 

LITERATURE  CITED 

Bullock,  S.  H.  1986.  Climate  of  Chamela,  Jalisco,  and 
trends  in  the  south  coastal  region  of  Mexico.  Arch. 
Met.  Geoph.  BiocL,  Sen  B 36:297-316. 
Chrlstman,  B.  J.  and  A.  A.  Dhondt.  1997.  Nest  pre- 
dation in  Black-capped  Chickadees:  how  safe  are 
cavity  nests?  Auk  1 14:769-773. 

Donahue,  P.  K.  and  T.  M.  Wood.  1995.  A safe,  flex- 
ible, and  non-injurious  technique  for  climbing  tall 
trees.  Selbyana  16:196—200. 

Enkerlin-Hoeflich,  E.  C.  1995.  Comparative  ecology 
and  reproductive  biology  of  three  species  of  Ama- 
zonci  parrots  in  northeastern  Mexico.  Ph.D.  diss., 
Texas  A & M Univ.,  Kingsville. 

Forshaw,  J.  M.  1989.  Parrots  of  the  world,  third  (re- 
vised) ed.  Lansdowne  Editions,  Melbourne,  Aus- 
tralia. 

Gnam,  R.  S.  1991.  Nesting  behaviour  of  the  Bahama 
Parrot  Atnazona  leucocephala  bahamensis  on 
Abaco  Island,  Bahamas.  Acta  Congr.  Int.  Ornithol. 
20:673-680. 

Gnam,  R.  S.  and  R.  F.  Rockwell.  1991.  Reproductive 
potential  and  output  of  the  Bahama  Parrot  Ama- 
zona  leucocephala  bahamen.sLs.  Ibis  133:400- 
405. 

Fanning,  D.V.  1991.  Distribution  and  breeding  biol- 
ogy of  the  Red-fronted  Macaw.  Wilson  Bull.  103: 
357-365. 

Fanning,  D.  V.  and  J.  T.  Shielett.  1983.  Nesting  ecol- 
ogy of  Thick-hilled  Parrots.  Condor  85:66-73. 
Lorr,  E.  J.  1993.  Annotated  checklist  of  the  vascular 


flora  of  the  Chamela  Bay  region,  Jalisco,  Mexico. 
Occas.  Pap.  Calif.  Acad.  Sci.  148:1-60. 

Lott,  E.  J.,  S.  H.  Bullock,  and  J.  A.  Solis-Magal- 
LANES.  1987.  Floristic  diversity  and  structure  of 
upland  and  arroyo  forests  of  coastal  Jalisco.  Bio- 
tropica  19:228-235. 

Mann,  R.  E.  H.  and  P.  D.  Mann.  1978.  Breeding 
Finch’s  Amazon  Parrot  (Atnazona  finschi).  Avicul. 
Mag.  84:187-189. 

Martijena,  N.  E.  and  S.  H.  Bullock.  1994.  Mono- 
specific  dominance  of  a tropical  deciduous  forest 
in  Mexico.  J.  Biogeogr.  21:63-74. 

Martin,  T.  E.  1987.  Food  as  a limit  on  breeding  birds: 
a life-history  perspective.  Annu.  Rev.  Ecol.  Syst. 
18:453-487. 

Navarro,  J.  L.,  M.  B.  Martella,  and  E.  H.  Bucher. 
1992.  Breeding  season  and  productivity  of  Monk 
Parakeets  in  Cordoba,  Argentina.  Wilson  Bull. 
104:413-424. 

Nilsson,  S.  G.  1984.  The  evolution  of  nest-site  selec- 
tion among  hole-nesting  birds:  the  importance  of 
nest  predation  and  competition.  Ornis  Scand.  15: 
167-175. 

Perry,  D.  R.  1978.  A method  of  access  into  the 
crowns  of  emergent  and  canopy  trees.  Biotropica 
10:155-157. 

Perry,  D.  R.  and  J.  Williams.  1981.  The  tropical  rain 
forest  canopy:  a method  providing  total  access. 
Biotropica  13:283-285. 

Renton,  K.  1998.  Reproductive  ecology  and  conser- 
vation of  the  Lilac-crowned  Parrot  (Atnazona  fin- 
schi) in  Jalisco,  Mexico.  Ph.D.  diss.,  Univ.  of 
Kent,  Canterbury,  U.K. 

Rowley,  I.  and  G.  Chapman.  1991.  The  breeding  bi- 
ology, food,  social  organization,  demography  and 
conservation  of  the  Major  Mitchell  or  Pink  Cock- 
atoo, Cacatua  leadbeateri,  on  the  margin  of  the 
Western  Australian  wheatbelt.  Aust.  J.  Zool.  39: 
21 1-261. 

Saunders,  D.  A.  1979.  The  availability  of  tree  hollows 
for  use  as  nest  sites  by  White-tailed  Black  Cock- 
atoos. Aust.  Wild.  Res.  6:205—216. 

Saunders,  D.  A.  1982.  The  breeding  behaviour  and 
biology  the  short-billed  form  of  the  White-tailed 
Black  Cockatoo  Calyptorhynchus  fiinereits.  Ibis 
124:422-455. 

Saunders,  D.  A,  G.  T.  Smith,  and  I.  Rowley.  1982. 
The  availability  and  dimensions  of  tree  hollows 
that  provide  nest  sites  for  cockatoos  (Psittacifonnes) 
in  Western  Australia.  Aust.  Wild.  Res.  9:541—556. 

Smith,  G.  T 1991.  Breeding  ecology  of  the  Western 
Long-billed  Corella,  Cacatua  pastihator  paslina- 
/or.  Wild.  Res.  18:91-1  10. 

Smith,  G.  T.  and  D.  A.  Saunders.  1986.  Clutch  size 
and  productivity  in  three  sympatric  species  of 
cockatoo  (Psittacifonnes)  in  the  south-west  of 
Western  Australia.  Aust.  Wild.  Res.  13:275-285. 

Snyder,  N.  E R.,  J.  W.  Wiley,  and  C.  B.  Kepler. 
1987.  The  parrots  of  Luquillo:  natural  history  and 
conservation  of  the  Puerto  Rican  Parrot.  Western 


Renton  and  Salinas-Melgoza  • LILAC-CROWNED  PARROT  NESTING  BEHAVIOR 


493 


Foundation  of  Vertebrate  Zoology,  Los  Angeles, 
California. 

SoNERUD,  G.  A.  1985.  Nest  hole  shift  in  Tengmalm’s 
Owl  Aeo}>olius  funereiis  as  defence  against  nest 
predation  involving  long-term  memory  in  the 
predator.  J.  Anim.  Ecol.  54:179-192. 

SoNERUD,  G.  A.  1989.  Reduced  predation  by  pine  mar- 
tins on  nests  of  Tengmalm’s  Owl  in  relocated  box- 
es. Anim.  Behav.  37:332-334. 


Waltman,  J.  R.  and  S.  R.  Beissinger.  1992.  Breeding 
behavior  of  the  Green-rumped  Parrotlet.  Wilson 
Bull.  104:65-84. 

WiLCOVE,  D.  S.  1985.  Nest  predation  in  forest  tracts 
and  the  decline  of  migratory  songbirds.  Ecology 
66:121 1-1214. 

Wilson,  K.  A.,  R.  Field,  and  M.  H.  Wilson.  1995. 
Successful  nesting  behavior  of  Puerto  Rican  Par- 
rots. Wilson  Bull.  107:518-529. 


Wilson  Bull.,  111(4),  1999,  pp.  494-498 


RELATIONSHIPS  AMONG  RED-COCKADED  WOODPECKER 
GROUP  DENSITY,  NESTLING  PROVISIONING 
RATES,  AND  HABITAT 

RICHARD  N.  CONNER,'  3 D.  CRAIG  RUDOLPH,'  RICHARD  R.  SCHAEFER,' 
DANIEL  SAENZ,'  AND  CLIFFORD  E.  SHACKELFORD^ 


ABSTRACT. — We  examined  Red-cockaded  Woodpecker  (Picoides  borealis)  food  provisioning  rates  of  nest- 
lings during  the  1992  and  1993  breeding  seasons  on  the  Vernon  Ranger  District  of  the  Kisatchie  National  Forest 
in  Louisiana.  Provisioning  rates  were  monitored  at  nest  trees  in  moderate  (9.8  groups/2  km  radius,  n = 10)  and 
low  (5.9  groups/2  km  radius,  n = 10)  density  populations.  Habitat  around  each  cluster  was  measured  within 
three  radii  (100  m,  400  m,  and  800  m)  to  evaluate  the  possible  influence  of  habitat  quality  on  group  density 
and  nestling  provisioning  rates.  We  tested  the  null  hypothesis  that  habitat  quality  and  provisioning  rates  would 
be  similar  in  areas  with  different  densities  of  woodpecker  groups.  We  failed  to  detect  differences  in  nestling 
provisioning  rates  between  woodpecker  groups  in  moderate  versus  low  group  densities.  Woodpecker  groups 
from  areas  where  group  densities  were  moderate  attempted  to  nest  significantly  more  often  than  woodpecker 
groups  occurring  in  low  densities.  Hardwood  midstory  vegetation  was  more  abundant  in  areas  with  low  wood- 
pecker group  density.  Old-growth  pines,  which  are  known  to  be  important  for  cavity  excavation,  were  present 
in  habitat  around  cavity-tree  clusters  of  moderate-density  groups,  but  generally  absent  in  areas  where  group 
density  was  low.  Woodpecker  group  density  may  be  related  to  hardwood  midstory  conditions  and  the  abundance 
and  spatial  distribution  of  remnant  old  pines.  Received  18  August  1998,  accepted  9 August  1999. 


The  Red-cockaded  Woodpecker  {Picoides 
borealis)  is  a cooperatively  breeding  species 
closely  associated  with  older-growth  pine  for- 
ests of  the  southeastern  United  States  (U.S. 
Fish  and  Wildlife  Service  1985,  Walters  et  al. 
1988).  A single  tree,  or  aggregation  of  cavity 
trees,  termed  the  cluster,  is  inhabited  by  a 
group  of  woodpeckers  that  includes  a single 
breeding  pair.  Other  adults  present  in  the 
group  are  typically  male  offspring  from  pre- 
vious breeding  seasons  (Ligon  1970,  Lennartz 
et  al.  1987). 

Considerable  information  is  known  about 
the  woodpecker’s  cavity  tree  requirements. 
The  Red-cockaded  Woodpecker  requires  old 
living  pines  for  its  cavities  (Conner  and 
O’Halloran  1987).  The  presence  of  fungal  de- 
cay within  the  heartwood  of  pines  increases 
in  frequency  as  pines  age  and  significantly 
shortens  the  time  required  for  woodpeckers  to 
excavate  nest  and  roost  cavities  (Hooper  et  al. 


' Wildlife  Habitat  and  Silviculture  Laboratory 
(maintained  in  cooperation  with  the  College  of  For- 
estry, Stephen  F.  Austin  State  Univ.),  Southern  Re- 
.search  Station,  U.S.D.A.  Forest  Service,  Nacogdoches, 
TX  75962. 

2 Texas  Partners  in  Flight,  Texas  Parks  and  Wildlife 
Department,  Austin,  TX  78704. 

’ Corresponding  author; 

E-mail:  c_connerrn@ titan.sfasu.edu 


1991,  Conner  et  al.  1994,  Rudolph  et  al. 
1995).  A sufficient  diameter  of  heartwood, 
which  increases  with  pine  age,  also  is  required 
to  provide  adequate  space  for  cavity  excava- 
tion (Conner  et  al.  1994).  Pines  with  suitable 
cavities  are  known  to  be  a critical  resource  for 
Red-cockaded  Woodpeckers  (Walters  et  al. 
1992a,  Conner  and  Rudolph  1995).  Thus,  lo- 
cation of  existing  cavities  and  the  number  and 
distribution  of  old-growth  pines  that  can  be 
excavated  for  cavities  likely  has  a strong  in- 
fluence on  the  density  and  spatial  distribution 
of  woodpecker  groups. 

The  quality  of  foraging  habitat  across  the 
forest  landscape  also  may  influence  nesting 
success  and  the  density  of  Red-cockaded 
Woodpecker  populations.  Male  and  female 
Red-cockaded  Woodpeckers  forage  at  differ- 
ent locations  on  trees  and  use  different  meth- 
ods to  exploit  arthropod  prey  (Ligon  1968, 
Hooper  and  Lennartz  1981).  Hooper  and  Har- 
low (1986)  observed  that  foraging  Red-cock- 
aded Woodpeckers  showed  little  preference 
among  pine  stands  that  were  more  than  30 
years  old,  and  concluded  that  once  pine  stands 
reach  30  years  old  their  quality  as  a foraging 
substrate  does  not  improve  with  further  aging. 
More  recent  studies  suggest  that  old-growth 
pines  provide  an  increased  foraging  benefit 
(Zwicker  1995,  Jones  and  Hunt  1996,  Eng- 


494 


Conner  el  al.  • RED-COCKADED  NESTLING  PROVISIONING 


495 


Strom  and  Sanders  1997).  In  this  study  we 
evaluate  relationships  among  forest  habitat, 
adult  woodpecker  provisioning  rates  of  nes- 
tlings, and  woodpecker  fledging  success  of 
woodpecker  groups  where  group  density  is 
moderate  and  where  group  density  is  low.  We 
ask  if  foraging  habitat  characteristics  are  re- 
lated to  woodpecker  group  density,  nestling 
provisioning  rates,  and  nesting  success.  If 
woodpecker  group  density  is  a function  of  for- 
aging habitat  quality,  provisioning  rates  of 
nestlings  where  woodpecker  group  density  is 
moderate  might  be  expected  to  exceed  those 
of  groups  living  in  lower  densities. 

STUDY  AREA  AND  METHODS 

The  Vernon  Ranger  District  of  the  Kisatchie  Na- 
tional Eorest  (31°  01'  N,  93°  02'  W)  is  located  in  west- 
central  Louisiana.  Longleaf  pines  (Finns  palustris) 
compose  the  bulk  of  the  overstory  with  grasses  and 
forbs  as  the  primary  ground  cover.  Hardwood  midstory 
vegetation  is  typically  minimal  on  the  Vernon  Ranger 
District  but  does  occur  in  some  areas  where  the  effec- 
tiveness of  prescribed  fire  has  been  reduced.  We  se- 
lected 10  Red-cockaded  Woodpecker  groups  from  ar- 
eas of  the  national  forest  where  group  density  was 
moderate  (x  = 9.8  active  cavity-tree  clusters  per  2-km 
radius)  and  10  groups  from  portions  of  the  Ranger  Dis- 
trict where  group  density  was  low  [x  = 5.9  active  cav- 
ity-tree clusters  per  2-km  radius;  moderate  vs  low 
group  density:  = 5.01,  P < 0.001;  see  Hooper  and 

Lennartz  (1995)  for  classification  (moderate  vs  low)  of 
woodpecker  group  densities].  Prior  to  the  1992  breed- 
ing season,  all  adult  woodpeckers  roosting  within  these 
20  clusters  were  captured  at  their  roost  cavities, 
weighed,  and  banded  with  U.S.  Fish  and  Wildlife  Ser- 
vice metal  bands  and  combinations  of  color  bands  to 
facilitate  individual  recognition. 

At  the  onset  of  the  nesting  seasons  during  1992  and 
1993  we  climbed  cavity  trees  using  Swedish  climbing 
ladders  to  determine  the  location  of  the  nest  tree,  the 
number  of  eggs  in  the  clutch,  and  the  initial  number 
of  nestlings.  During  July  and  August  of  both  years  we 
determined  the  number  of  post-fledging  survivors  for 
each  woodpecker  group.  We  quantified  adult  provi- 
sioning rates  of  nestlings  for  each  woodpecker  group 
when  nestlings  were  8,  20,  and  23  days  old  (see  Schae- 
fer 1996).  The  total  number  of  provisioning  trips  made 
by  adults  to  feed  nestlings  was  counted  during  the  3- 
hour  period  following  the  breeding  male’s  initial  de- 
parture from  the  nest  cavity  in  the  morning.  We  also 
identified  which  adult  brought  food  to  nestlings  during 
each  provisioning  trip  and  made  an  estimate  of  prey 
size  (cm)  using  the  adult  woodpecker’s  bill  as  a size 
scale.  Nest  trees  were  climbed  after  each  3-h  provi- 
sioning sampling  period  to  verify  the  number  of  nest- 
lings present  in  each  nest  cavity.  An  adjusted  provi- 
sioning rate  was  calculated  for  each  nest  by  dividing 
the  total  number  of  feeding  trips  by  the  number  of 


adult  woodpeckers  in  the  group,  yielding  the  number 
of  provisioning  trips  per  adult.  Provisioning  rates  were 
also  adjusted  to  simultaneously  account  for  different 
numbers  of  nestlings  in  cavities  and  group  size. 

Red-cockaded  Woodpecker  use  of  forest  stands  for 
toraging  depends  in  part  on  the  distance  ol'  the  stand 
from  cavity  trees  (DeLotelle  et  al.  1987).  We  measured 
vegetational  characteristics  around  the  geometric  cen- 
ter of  each  woodpecker  cavity-tree  cluster  within  three 
radii:  0-100  m,  101—400  m,  and  401-800  m.  Forest 
compartment  stand  maps  were  obtained  from  the  Kis- 
atchie National  Forest  supervisor’s  office  for  those 
compartments  falling  within  800  m of  each  cluster 
studied.  Each  compartment  is  comprised  of  forest 
stands  of  varying  size  and  tree  age.  Five  dominant  or 
codominant  pine  trees  were  randomly  selected  as  cen- 
tral points  in  each  forest  stand,  and  habitat  character- 
istics for  each  stand  were  gathered  around  each  of 
these  five  trees.  For  each  stand,  means  were  calculated 
for  each  habitat  measurement  taken  around  the  five 
central  trees. 

Stand  age  was  determined  by  coring  each  central 
tree  at  breast  height  (1.3  m)  with  an  increment  borer 
and  counting  the  growth  rings  of  the  cores.  Five  years 
were  added  to  the  growth  ring  counts  for  longleaf  pine 
to  adjust  for  the  minimum  years  spent  as  a seedling 
(Conner  and  O’Halloran  1987).  Stands  were  divided 
into  five  age  classes:  0-29,  30-49,  50-69,  70-89,  and 
>90  years  old.  The  diameters  of  each  central  tree  and 
of  all  live  stems  larger  than  2 cm  within  1 1.3  m of  the 
central  tree  were  measured  at  breast  height  with  cali- 
pers. Stands  were  divided  into  two  diameter  classes 
based  on  average  diameters  of  pines:  30-40  cm  and 
40.1-50  cm  diameter  at  breast  height  (dbh).  Pines  in 
smaller  diameter  classes  were  excluded  because  they 
were  rarely  encountered.  A one-factor  metric  basal 
area  prism  was  used  to  measure  pine  overstory  basal 
area  (m°/ha).  Stands  were  divided  into  three  basal  area 
classes  based  on  average  basal  areas  of  overstory 
pines:  0—3,  3.1-12,  and  12.1-21  m^/ha. 

Hardwood  midstory  density  was  visually  estimated 
and  placed  into  one  of  five  categories:  none,  sparse, 
moderate,  dense,  and  very  dense.  The  effects  of  mid- 
story height  (measured  with  a clinometer)  and  mids- 
tory density  may  not  be  obvious  when  considered  in- 
dependently. For  example,  tall,  dense  midstory  con- 
ditions may  have  a different  impact  on  the  woodpeck- 
ers than  would  tall,  sparse  midstory  conditions. 
Therefore,  both  midstory  height  and  midstory  density 
were  considered  together  to  obtain  measures  of  suit- 
able and  unsuitable  midstory  conditions.  Midstory 
conditions  were  considered  suitable  if  the  height  was 
less  than  3 m regardless  of  the  density,  or  if  the  density 
was  none  to  sparse  regardless  of  the  height.  Midstory 
conditions  were  considered  unsuitable  if  the  height 
was  more  than  3 m and  the  density  was  moderate  to 
very  dense. 

The  area  (ha)  of  each  forest  stand  within  distance 
zones  of  100  m,  400  m.  and  800  m from  each  nest  tree 
was  measured  with  a digitizer  using  Sigma-Scan*.  The 
area  of  each  stand  was  summed  for  each  habitat  vari- 


496 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


TABLE  1.  Comparisons  of  mean  (±  SD)  nestling  provisioning  (feeding)  and  nest  productivity  rates  between 
Red-cockaded  Woodpecker  groups  in  areas  of  moderate  (n  = 10)  and  low  {n  = 10)  woodpecker  group  density 
on  the  Vernon  Ranger  District,  Kisatchie  National  Lorest,  Louisiana,  during  the  1992  and  1993  breeding  seasons. 

Group  density 

Variable 

Moderate 

Low 

t 

p 

No.  feeding  trips  to  8-day-old  nestlings 

37.0  (17.3) 

37.6  (15.5) 

0.61 

>0.05 

No.  feeding  trips  to  20-day-old  nestlings 

42.1  (17.5) 

43.9  (29.2) 

0.19 

>0.05 

No.  feeding  trips  to  23-day-old  nestlings 

42.6  (18.4) 

46.1  (19.6) 

0.06 

>0.05 

Adjusted  8-day  feeding  rate 

12.9  (7.4) 

13.3  (4.0) 

0.13 

>0.05 

Adjusted  20-day  feeding  rate 

14.3  (5.9) 

15.5  (9.5) 

0.38 

>0.05 

Adjusted  23 -day  feeding  rate 

14.8  (6.3) 

16.2  (5.1) 

0.57 

>0.05 

No.  adult  woodpeckers  in  group 

3.1  (0.9) 

2.9  (0.8) 

0.48 

>0.05 

No.  of  eggs  in  clutch  from  successful  groups 

3.1  (0.9) 

2.8  (1.2) 

0.80 

>0.05 

No.  of  nestlings  fledged  from  successful  groups 

1.9  (0.9) 

1.7  (0.9) 

0.67 

>0.05 

able  by  age,  dbh,  and  basal  area  classes  as  described 
above  for  the  100  m,  400  m,  and  800  m concentric 
distance  zones  around  each  woodpecker  cluster  ex- 
amined and  converted  to  a percentage  for  each  distance 
zone. 

Data  were  analyzed  using  SAS  (version  6.12;  SAS 
Inst.  Inc.  1988)  on  an  IBM  compatible  computer.  Per- 
centage data  for  stand  area  in  various  habitat  classes 
were  transformed  with  an  arcsin  transformation,  and 
count  data  for  nesting  variables  were  transformed  with 
a square  root  function  to  approximate  normality  in  all 
parametric  statistical  tests.  A stepwise  discriminant 
analysis  was  used  as  a data  reduction  technique  to 
compare  habitat  variables  between  groups  occurring  in 
moderate  and  low  densities.  In  order  to  test  the  hy- 
pothesis that  habitat  quality  influences  provisioning 
rates  and  woodpecker  group  density,  we  calculated 
two-tailed  r-tests  (adjusted  for  unequal  variances)  to 
test  for  differences  in  nest  productivity  and  provision- 
ing effort.  Lrequencies  of  nest  success,  nest  attempts, 
and  nest  failures  for  moderate  and  low  group  densities 
were  compared  with  analysis  using  2X2  contin- 
gency tables.  Logistic  regression  (stepwise)  was  used 
to  evaluate  nest  success  (successful  or  not)  and  nest 
attempts  (attempted  to  nest  or  not)  as  a function  of 
habitat  characteristics. 

RESULTS 

Of  all  habitat  variables  measured  only  the 
percentage  area  with  suitable  midstory  con- 
ditions within  100  m of  the  nest  tree  (Wilks’ 
Lambda  0.78,  P = 0.0356,  1,  18  df)  entered 
the  stepwise  discriminant  function  (75%  clas- 
sification accuracy)  comparing  groups  in  mod- 
erate [96.3%  ± 6.5  (SE)  area  with  suitable 
midstory]  and  low  (67.8%  ± 38.5)  densities, 
suggesting  considerable  homogeneity  of  hab- 
itat throughout  the  Vernon  Ranger  District. 
Stands  with  old-growth  pines  (pines  >90 
years  old)  within  100  m and  400  m of  clusters 
were  present  around  groups  occurring  in  mod- 


erate densities  (1.7%  ± 5.2  and  2.4%  ±5.7 
of  the  area,  respectively)  but  were  totally  ab- 
sent in  areas  with  low  woodpecker  group  den- 
sities. 

Differences  in  woodpecker  group  density 
did  not  appear  to  be  a function  of  foraging 
habitat  quality.  We  detected  no  significant  dif- 
ferences between  moderate  and  low  wood- 
pecker group  density  in  the  unadjusted  and  ad- 
justed rates  that  adults  fed  nestlings  at  8,  20, 
or  23  days  post  hatching,  woodpecker  group 
size,  clutch  size,  and  the  number  of  young 
successfully  fledged  (Table  1).  However,  the 
power  of  our  ability  to  detect  a difference  is 
low  (5-9%)  because  of  the  relatively  small 
sample  size  (n  = 20).  As  with  provisioning 
rates,  we  failed  to  detect  a difference  between 
moderate  and  low  woodpecker  group  density 
in  the  size  of  prey  that  adults  fed  to  nestlings 
(t  = 0.45,  P > 0.05,  df  = 24).  The  average 
weight  of  breeding  males  was  identical  in  ar- 
eas of  moderate  group  density  (48.3  g)  and 
low  group  density  (48.3  g).  Breeding  female 
woodpeckers  differed  by  only  0.1  g,  46.6  g 
and  46.7  g,  respectively. 

A significantly  higher  proportion  of  wood- 
pecker groups  attempted  to  nest  in  moderate 
group  densities  (19  of  20  nest  years)  than  in 
low  group  densities  (10  of  20  nest  years; 

= 10.2,  P < 0.001,  df  = 1).  Also,  a signifi- 
cantly higher  proportion  of  woodpecker 
groups  nested  successfully  in  moderate  group 
densities  (17  of  20  nest  years)  than  in  low 
group  densities  (9  of  20  nest  years;  = 7.03, 
P - 0.008,  df  = 1).  The  three  groups  that 
failed  to  produce  fledglings  in  the  areas  with 
moderate  group  density  were  the  result  of  two 


Conner  el  al.  • RED-COCK ADED  NESTLING  PROVISIONING 


497 


nesting  attempts  that  failed  and  one  instance 
where  the  woodpeckers  did  not  attempt  to  nest 
because  the  breeding  female  abandoned  the 
cluster  (or  died)  immediately  prior  to  the 
breeding  season.  The  1 1 groups  that  failed  to 
produce  fledglings  in  the  low-density  groups 
were  the  result  of  one  attempt  (eggs  laid)  that 
failed,  five  instances  where  the  breeding  fe- 
male disappeared  immediately  prior  to  the 
breeding  season,  and  five  instances  where  a 
pair  was  present  but  did  not  attempt  to  nest. 
Two  (10.5%)  of  19  nesting  attempts  in  areas 
with  moderate  group  density  failed,  whereas 
1 (10.0%)  of  10  nesting  attempts  in  areas  with 
low  group  density  failed  (x“  = 0.002,  P > 
0.05,  df  = 1).  Cluster  abandonment  by  fe- 
males prior  to  the  breeding  season  was  mar- 
ginally higher  in  low  group  densities  (5  of  20 
nest  years)  than  in  moderate  group  densities 
(1  of  20  nest  years;  = 3.14,  P = 0.08,  df 
= 1). 

Attempts  to  evaluate  nest  success  (success- 
ful or  not)  and  nest  attempts  (attempted  to  nest 
or  not)  as  a function  of  habitat  characteristics 
through  logistic  regression  failed  as  no  mea- 
sured habitat  characteristic  had  a sufficient  re- 
lationship to  enter  the  analyses. 

DISCUSSION 

The  observed  lower  rate  of  nesting  attempts 
in  the  low-density  groups  relative  to  moder- 
ate-density groups  could  be  related  to  at  least 
several  factors  singly  or  in  combination:  (1) 
failure  to  nest  because  of  a foraging  habitat 
insufficiency,  (2)  demographic  dysfunction  re- 
sulting from  increased  isolation  of  low-density 
groups  relative  to  moderate-density  groups, 
and  (3)  an  inadequate  number  of  older- growth 
pines  suitable  for  nest  trees.  Inadequacy  of 
foraging  habitat  appears  to  be  an  unlikely  ex- 
planation. The  lack  of  older-growth  pines  in 
the  low  group  density  area  may  have  reduced 
the  number  of  sites  available  for  cavity-tree 
clusters,  and  because  of  increased  group  iso- 
lation, may  have  had  a negative  effect  on  pop- 
ulation demographics.  Unfortunately,  the  his- 
toric demographics  of  groups  we  studied  was 
not  known.  It  is  also  possible  that  prey  avail- 
ability, as  provided  by  these  older-growth 
pines,  had  an  influence  on  cluster  abandon- 
ment by  females  and  whether  groups  attempt- 
ed to  nest  or  not.  Limiting  factors  appear  to 


prevent  nest  initiation  rather  than  decreasing 
the  success  rate  of  nesting  attempts. 

Recent  research  suggests  that  Red-cockad- 
ed  Woodpeckers  have  a preference  for  older 
pines.  Zwicker  (1995),  Engstrom  and  Sanders 
(1997),  and  Jones  and  Hunt  (1996)  observed 
that  Red-cockaded  Woodpeckers  used  larger, 
older-growth  pines  at  much  higher  rates  than 
would  be  expected  based  on  availability. 
However,  using  logistic  regression  we  failed 
to  detect  a relationship  between  availability  of 
old  pines  and  nesting  attempts. 

The  age  and  experience  of  the  breeders  oc- 
cupying moderate  and  low  density  group  areas 
also  may  have  influenced  the  observed  differ- 
ences in  nesting  productivity.  Older  and  more 
experienced  woodpeckers  might  preferentially 
capture  and  occupy  habitat  with  higher  den- 
sities of  older-growth  pines  because  such  hab- 
itat is  viewed  as  better  quality  than  habitat 
lacking  older-growth  pines.  The  number  of 
young  fledged  by  Red-cockaded  Woodpeckers 
is  known  to  increase  with  the  age  and  expe- 
rience of  breeders  (Walters  et  al.  1992b).  We 
did  not  know  the  ages  of  the  woodpeckers  in 
the  groups  we  studied. 

The  suitability  of  hardwood  midstory  con- 
ditions within  100  m of  the  center  of  cavity- 
tree  clusters  was  significantly  greater  for  mod- 
erate-density groups,  but  it  was  not  related  to 
nest  productivity  or  the  propensity  of  groups 
to  nest.  The  greater  presence  of  unsuitable 
midstory  conditions  in  areas  of  low  group 
density  than  areas  of  moderate  group  density 
suggests  that  cluster  abandonment  could  also 
have  influenced  the  observed  differences  in 
group  density.  The  Red-cockaded  Woodpeck- 
er’s requirement  for  open  pine  stands  relative- 
ly devoid  of  hardwood  midstory  is  well 
known  (Conner  and  Rudolph  1989,  Loeb  et 
al.  1992). 

ACKNOWLEDGMENTS 

We  thank  U.S.  Forest  Service  personnel  from  the 
Vernon  Ranger  District  of  the  Kisatchie  National  For- 
est for  logistical  assistance  throughout  the  study  and 
the  Southern  Region  Office  R-8  for  financial  support. 
We  thank  K.  G.  Beal.  N.  R.  Canie,  R.  T.  Engstrom.  S. 
Forbes,  E C.  James,  R.  G.  Hooper,  N.  E.  Koerth,  J.  D. 
Ligon,  and  J.  R.  Walters  for  constructive  comments 
leading  to  the  improvement  of  the  manuscript. 

LITERATURE  CITED 

Conner,  R.  N.  and  K.  A.  O'Halloran.  1987.  Cavity- 
tree  selection  by  Red-cockaded  Woodpeckers  as 


498 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


related  to  growth  dynamics  of  southern  pines. 
Wilson  Bull.  99:398-412. 

Conner,  R.  N.  and  D.  C.  Rudolph.  1989.  Red-cock- 
aded  Woodpecker  colony  status  and  trends  on  the 
Angelina,  Davy  Crockett  and  Sabine  National 
Forests.  U.S.D.A.  For.  Serv.,  Res.  Pap.  SO-250:1- 
15. 

Conner,  R.  N.  and  D.  C.  Rudolph.  1995.  Excavation 
dynamics  and  use  patterns  of  Red-cockaded 
Woodpecker  cavities:  relationships  with  coopera- 
tive breeding.  Pp.  343-352  in  Red-cockaded 
Woodpecker  recovery,  ecology  and  management 
(D.  L.  Kulhavy,  R.  G.  Hooper,  and  R.  Costa, 
Eds.).  College  of  Forestry,  Stephen  F.  Austin  State 
Univ.,  Nacogdoches,  Texas. 

Conner,  R.  N.,  D.  C.  Rudolph,  D.  Saenz,  and  R.  R. 
Schaefer.  1994.  Heartwood,  sap  wood,  and  fungal 
decay  associated  with  Red-cockaded  Woodpecker 
cavity  trees.  J.  Wildl.  Manage.  58:728-734. 

DeLotelle,  R.  S.,  R.  J.  Epting,  and  J.  R.  Newman. 
1987.  Habitat  use  and  territory  characteristics  of 
Red-cockaded  Woodpeckers  in  central  Florida. 
Wilson  Bull.  99:202-217. 

Engstrom,  R.  T.  and  F J.  Sanders.  1997.  Red-cock- 
aded Woodpecker  foraging  ecology  in  an  old- 
growth  longleaf  pine  forest.  Wilson  Bull.  109: 
203-217. 

Hooper,  R.  G.  and  R.  F.  Harlow.  1986.  Foraging 
stands  selected  by  foraging  Red-cockaded  Wood- 
peckers. U.S.D.A.  For.  Serv.  Res.  Pap.  SE-259: 1- 
10. 

Hooper,  R.  G.  and  M.  R.  Lennartz.  1981.  Foraging 
behavior  of  the  Red-cockaded  Woodpecker  in 
South  Carolina.  Auk  98:321—334. 

Hooper,  R.  G.  and  M.  R.  Lennartz.  1995.  Short-term 
response  of  a high  density  Red-cockaded  Wood- 
pecker population  to  loss  of  foraging  habitat.  Pp. 
283-289  in  Red-cockaded  Woodpecker  recovery, 
ecology  and  management  (D.  L.  Kulhavy,  R.  G. 
Hooper,  and  R.  Costa,  Eds.).  College  of  Forestry, 
Stephen  F.  Austin  State  Univ.,  Nacogdoches,  Tex- 
as. 

Hooper,  R.  G.,  M.  R.  Lennartz,  and  H.  D.  Muse. 
1991.  Heart  rot  and  cavity  tree  selection  by  Red- 
cockaded  Woodpeckers.  J.  Wildl.  Manage.  55: 
323-327. 

Jones,  C.  M.  and  H.  E.  Hunt.  1996.  Foraging  habitat 
of  the  Red-cockaded  Woodpecker  on  the 


D’Arbonne  National  Wildlife  Refuge,  Louisiana. 
J.  Field  Ornithol.  67:511-518. 

Lennartz,  M.  R.,  R.  G.  Hooper,  and  R.  F.  Harlow. 

1987.  Sociality  and  cooperative  breeding  of  Red- 
cockaded  Woodpeckers,  {Picoides  borealis).  Be- 
hav.  Fcol.  Sociobiol.  20:77-88. 

Ligon,  j.  D.  1968.  Sexual  differences  in  foraging  be- 
havior in  two  species  of  Dendrocopos  woodpeck- 
ers. Auk  85:203-215. 

Ligon,  J.  D.  1970.  Behavior  and  breeding  biology  of 
the  Red-cockaded  Woodpecker.  Auk  87:255-278. 
Loeb,  S.  C.,  W.  D.  Pepper,  and  A.  T.  Doyle.  1992. 
Habitat  characteristics  of  active  and  abandoned 
Red-cockaded  Woodpecker  colonies.  So.  J.  Appl. 
For.  16:120-125. 

Rudolph,  D.  C.,  R.  N.  Conner,  and  R.  R.  Schaefer. 
1995.  Red-cockaded  Woodpecker  detection  of  red 
heart  infection.  Pp.  338-342  in  Red-cockaded 
Woodpecker  recovery,  ecology  and  management 
(D.  L.  Kulhavy,  R.  G.  Hooper,  and  R.  Costa, 
Eds.).  College  of  Forestry,  Stephen  F.  Austin  State 
Univ.,  Nacogdoches,  Texas. 

SAS  Inst.,  Inc.  1988.  SAS/STAT  user’s  guide.  Release 
6.03.  SAS  Institute,  Inc.,  Cary,  North  Carolina. 
Schaefer,  R.  R.  1996.  Red-cockaded  Woodpecker  re- 
production and  provisioning  of  nestlings  in  rela- 
tion to  habitat.  M.Sc.  thesis,  Stephen  F.  Austin 
State  Univ.,  Nacogdoches,  Texas. 

U.S.  Fish  and  Wildlife  Service.  1985.  Red-cockaded 
Woodpecker  recovery  plan.  U.S.  Fish  and  Wildlife 
Service,  Atlanta,  Georgia. 

Walters,  J.  R.,  C.  K.  Copeyon,  and  J.  H.  Carter,  III. 
1992a.  Test  of  the  ecological  basis  of  cooperative 
breeding  in  Red-cockaded  Woodpeckers.  Auk 
109:90-97. 

Walters,  J.  R.,  P.  D.  Doerr,  and  J.  H.  Carter,  III. 
1992b.  Delayed  dispersal  and  reproduction  as  a 
life-history  tactic  in  cooperative  breeders:  fitness 
calculations  from  Red-cockaded  Woodpeckers. 
Am.  Nat.  139:623-643. 

Walters,  J.  R.,  P.  D.  Doerr,  and  J.  H.  Carter,  III. 

1988.  The  cooperative  breeding  system  of  the 
Red-cockaded  Woodpecker.  Ethology  78:275- 
305. 

ZwiCKER,  S.  M.  1995.  Selection  of  pines  for  foraging 
and  cavity  excavation  by  Red-cockaded  Wood- 
peckers. M.Sc.  thesis.  North  Carolina  State  Univ., 
Raleigh. 


Wilson  Bull.,  111(4),  1999,  pp.  499-504 


RESPONSES  OF  BELL’S  VIREOS  TO  BROOD  PARASITISM  BY  THE 
BROWN-HEADED  COWBIRD  IN  KANSAS 

TIMOTHY  H.  PARKER'  2 


ABSTRACT. — I studied  patterns  of  cowbird  parasitism  and  responses  to  this  parasitism  by  Bell’s  Vireos  (Vireo 
hellii)  in  Kansas.  Bell’s  Vireos  abandoned  parasitized  nests  at  a significantly  higher  rate  than  unparasitized  nests. 
Lower  probability  of  brood  parasitism  later  in  the  season  may  help  make  abandonment  followed  by  renesting 
beneficial.  Burial  of  cowbird  eggs  by  vireos  was  also  observed  in  several  cases.  I did  not  detect  a strong 
relationship  between  nest  site  vegetation  characteristics  and  the  probability  of  brood  parasitism.  Received  9 Nov. 
1998.  accepted  27  May  1999. 


Bell’s  Vireo  (Vireo  bellii)  is  a well  known 
host  of  the  brood  parasitic  Brown-headed 
Cowbird  (Molothrus  ater\  Barlow  1962, 
Mayfield  1965,  Franzreb  1987,  Brown  1993). 
The  arrival  of  the  cowbird  in  California  in 
this  century  (Laymon  1987)  has  been  cited 
as  a major  factor  causing  the  severe  range 
restriction  and  endangerment  of  the  Least 
Bell’s  Vireo  (V.  b.  pusillus;  Franzreb  1987, 
1989;  Laymon  1987,  Brown  1993).  In  his  re- 
view Brown  (1993)  reported  that  between 
one  third  to  over  one  half  of  all  Bell’s  Vireo 
nests  monitored  in  California  were  parasit- 
ized by  cowbirds.  High  rates  of  parasitism 
were  also  reported  in  the  Great  Plains  race 
(V.  b.  bellii',  Barlow  1962,  Brown  1993).  Al- 
though declines  in  Bell’s  Vireo  population 
have  been  detected  in  some  areas  of  the  Great 
Plains  by  the  Breeding  Bird  Survey  (Brown 
1993),  this  species  is  still  at  least  locally 
common.  The  long-term  data  set  (1981- 
1997)  from  my  study  site  shows  no  decline 
[Konza  Prairie  Long  Term  Ecological  Re- 
search (LTER)  Site,  data  set  CBPOl].  Sur- 
veys elsewhere  in  the  region  also  have  de- 
tected higher  Bell’s  Vireo  densities  than  near- 
by Breeding  Bird  Survey  routes  (Robbins  et 
al.  1992,  1993;  M.  B.  Robbins,  pers.  comm.). 
This  suggests  that  cowbird  parasitism,  de- 
spite its  frequency,  may  not  be  causing  a rap- 
id decline  in  Bell’s  Vireo  on  the  Great  Plains. 

It  is  important  to  study  nest  success  be- 
cause local  population  numbers  may  not  re- 
flect local  reproduction  (because  of  source- 


' Division  of  Biology,  Kansas  State  Univ.,  Manhat- 
tan, KS  66506. 

^ Present  address:  Dept,  of  Biology,  Univ.  of  New 
Mexico,  Albuquerque,  NM  87131; 

E-mail:  tparker@unm.edu 


sink  dynamics;  Brawn  and  Robinson  1996). 
If  the  Bell’s  Vireo  is  not  declining  rapidly  on 
the  Great  Plains,  we  might  expect  this  pop- 
ulation to  possess  traits  that  would  allow  its 
persistence  in  the  face  of  cowbird  parasitism. 
Vireos  could  try  to  avoid  parasitism  altogeth- 
er, they  could  attempt  to  salvage  nesting  at- 
tempts after  parasitism  has  occurred,  or  they 
could  simply  abandon  parasitized  nests  and 
renest  (Clark  and  Robertson  1981,  Hill  and 
Sealy  1994).  Avoidance  measures  could  in- 
clude cryptic  nest  placement,  secretive  be- 
havior around  the  nest  (Uyehara  and  Narins 
1995),  and/or  aggressive  nest  defense  (Neu- 
dorf  and  Sealy  1994,  Robertson  and  Norman 
1977).  Two  means  of  salvaging  a parasitized 
nest  include  removal  of  cowbird  eggs  (Roth- 
stein  1975)  or  burial  of  cowbird  eggs  with 
nesting  material  (Clark  and  Robertson  1981, 
Sealy  1996). 

In  this  paper,  I consider  the  potential  roles 
for  avoidance  of  cowbirds  and  salvaging  or 
abandoning  parasitized  nests  by  Bell’s  Vireos 
in  Kansas.  Analysis  of  nest  site  vegetation 
coupled  with  observations  of  nest  contents  al- 
lowed exploration  of  cryptic  nest  placement, 
burial  of  cowbird  eggs,  and  nest  abandonment 
followed  by  renesting. 

METHODS 

From  May  through  Augu.st  of  1996  I inve.stigated 
cowbird  parasitism  and  nest  success  of  Bell's  Vireos 
in  the  Flint  Hills  of  northeastern  Kansas.  My  study  site 
was  located  on  a portion  of  the  Nature  Con.servancy's 
Konza  Prairie  Research  Natural  Area  (in  Riley  and 
Geary  counties).  The  site  consisted  of  tallgrass  prairie 
interspersed  with  deciduous  shrub  vegetation  concen- 
trated around  ephemeral  streams  and  limestone  out- 
croppings. Vireos  arrive  at  this  site  beginning  in  mid- 
May  and  initiate  nest  building  in  late  May,  but  re- 
nesting attempts  continue  into  early  July.  Nests  are 


499 


500 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


placed  in  low  deciduous  shrubs  usually  within  1.5  m 
of  the  ground  (unpubl.  data). 

Nest  building  by  Bell’s  Vireos  usually  takes  3 to  4 
days  and  egg  laying  follows  1 or  2 days  after  nest 
completion.  Males  aid  in  nest  site  selection,  nest  con- 
struction, incubation,  and  feeding  of  young.  Typically 
4 eggs  are  laid  (Brown  1993).  On  the  Great  Plains, 
two  broods  are  sometimes  reared  in  one  season  (Bar- 
low  1962,  Brown  1993). 

I searched  for  and  monitored  nests  throughout  the 
season  using  the  locations  of  singing  males  to  narrow 
my  search.  I located  most  nests  before  the  onset  of 
incubation.  The  entire  study  area  (110  ha)  was 
searched  every  3-4  days.  I also  visited  active  nests 
once  every  3—4  days  to  record  the  number  of  eggs  and/ 
or  nestlings  present  (including  cowbird)  and  to  look 
for  any  buried  cowbird  eggs.  If  no  adult  vireo  was 
active  in  the  vicinity  of  the  nest,  I felt  the  eggs  for 
warmth  to  determine  if  they  were  being  incubated.  I 
ceased  visiting  a nest  and  concluded  it  had  failed  after 
two  successive  visits  where  I observed  cold  eggs  and 
no  parental  activity.  If  the  entire  contents  of  a nest 
were  removed,  or  if  during  incubation  or  the  nestling 
period  most  of  the  contents  were  removed  and  the  par- 
ents ceased  attending  the  nest,  I concluded  that  the  nest 
was  depredated.  I did  not  visit  nests  if  cowbirds  or 
predators  were  in  the  vicinity  (Martin  and  Geupel 
1993). 

To  minimize  nest  disturbance,  I waited  until  the  nest 
was  no  longer  in  use  before  assessing  vegetation  near 
the  nest.  I measured  height  of  nest,  height  of  the  nest 
shrub,  and  depth  of  leaf  litter.  I counted  the  number  of 
woody  stems  within  a 50  cm  horizontal  radius  of  the 
nest  shrub  and  estimated  nest  concealment  (percentage 
of  nest  hidden  from  the  observer  by  vegetation  at  dis- 
tance of  1 m,  BBIRD  protocol)  above,  below,  and  in 
the  four  cardinal  directions  around  the  nest.  I estimated 
ground  cover  (a  proxy  for  vegetation  density;  defined 
as  the  percent  of  ground  covered  by  a given  vegetation 
type  within  5 m of  the  nest)  for  nest  substrate  (the  plant 
species  in  which  nest  placed),  large  shrubs  (the  size 
class  used  by  vireos  for  nesting),  all  woody  vegetation, 
woody  clumps  (closed  canopy  continuous  woody  veg- 
etation), sparse  woody  clumps  (open  canopy  continu- 
ous woody  vegetation),  grass,  and  the  three  most  com- 
mon woody  species  within  5 m.  Also  within  the  5 m, 
1 estimated  the  median  height  of  the  woody  canopy, 
measured  the  height  of  the  tallest  woody  stem,  and 
counted  the  number  of  dead  woody  stems  and  the 
numbers  of  live  woody  stems  under  2.5  cm  diameter 
and  over  2.5  cm  diameter.  To  assess  the  area  of  ground 
covered  (both  within  and  outside  of  the  5 m radius) 
by  the  woody  clump  in  which  the  nest  was  placed,  1 
measured  the  maximum  width  of  the  clump  and  the 
width  perpendicular  to  the  maximum.  The  distance 
from  the  nest  to  the  nearest  corridor  of  woody  vege- 
tation along  a stream  bed  was  recorded  as  well. 

By  noting  changes  in  nest  contents  and  whether  or 
not  nests  were  active,  I assessed  vireo  re.sponses  to 
cowbird  parasitism:  nest  material  placed  over  cowbird 
eggs  or  nest  abandonment  subsequent  to  the  laying  of 


TABLE  1.  Fates  of  Bell’s  Vireo  nests  parasitized 
and  not  parasitized  by  Brown-headed  Cowbirds.  All 

nests  included  were  completed  and 
found  during  building  or  laying. 

active 

and  were 

Para.sitized 

Unparasitized 

nests 

nests 

Abandoned 

32“ 

8*’ 

Depredated 

Fledged  (cowbird  if  parasitized. 

8^- 

4 

vireos  if  not) 

3 

8“= 

TotaE 

44 

19 

“ Includes  two  nests  in  which  cowbird  eggs  were  buried  and  other  cowbird 
eggs  were  later  laid. 

^ Includes  one  nest  in  which  a cowbird  egg  was  buried. 

Total  offspring  fledged  = 24. 

One  nest  counted  as  parasitized  here  contained  a buried  cowbird  egg. 
Without  further  cowbird  eggs  being  laid,  the  nest  was  later  abandoned  and 
so  was  considered  unparasitized  when  abandoned. 


cowbird  eggs.  A nest  was  considered  abandoned  (as 
opposed  to  depredated)  if  parental  activity  ceased  and 
either  the  number  of  vireo  eggs  had  not  declined  be- 
tween visits  or  the  number  of  vireo  eggs  had  decreased 
but  the  number  of  cowbird  eggs  had  inereased  between 
visits.  Using  a test  (Sokal  and  Rohlf  1987),  I com- 
pared the  proportion  of  parasitized  nests  that  were 
abandoned  to  the  proportion  of  unparasitized  nests 
which  were  abandoned.  All  completed  (nest  lining 
complete),  active  (adults  defending  nest)  nests  located 
before  or  during  the  laying  stage  were  included  in  the 
analysis  (/?  = 63). 

To  identify  the  factors  associated  with  nest  aban- 
donment, I compared  parasitized  nests  (n  = 43;  does 
not  include  1 nest  abandoned  after  vireos  had  buried 
a cowbird  egg;  Table  1)  that  were  abandoned  to  those 
that  were  not  abandoned  based  on  the  numbers  of  vireo 
eggs  and  cowbird  eggs  in  the  nests.  Numbers  of  vireo 
and  cowbird  eggs  were  considered  separately  in  two  t- 
te.sts  (using  f-test  assuming  equal  variances,  Microsoft 
Excel  7.0). 

Although  the  vireos  were  not  banded,  I conserva- 
tively estimated  renesting  attempts  by  comparing  nest 
locations  with  dates  of  nest  use  for  all  nests  (n  = 63) 
included  in  this  study.  I considered  a nesting  attempt 
to  be  a renesting  event  if  it  occurred  within  7 days  of 
the  cessation  of  use  of  a nearby  nest.  If  a nesting  at- 
tempt was  begun  after  a longer  period,  I considered  it 
a possible  renesting  attempt  (presumably  in  some  of 
these  cases  I may  have  missed  an  intervening  nesting 
attempt).  Furthermore,  a nest  could  be  considered  part 
of  a given  .series  of  renestings  (or  possible  renestings) 
only  if  the  location  of  the  nest  did  not  overlap  with 
the  locations  of  a different  series  of  apparent  rene.st- 
ings.  Becau.se  Bell’s  Vireos  are  territorial  (Brown 
1993),  I made  the  conservative  as.sumption  that  terri- 
tories (i.c.,  series  of  nesting  attempts)  did  not  overlap 
and  were  consistent  throughout  the  season  to  avoid 
overestimating  renesting.  1 located  a number  of  i.solat- 
ed  nests  which,  based  on  their  late  dates  of  initiation. 


Parker  • BROOD  PARASITISM  OF  BELL’S  VIREOS 


501 


were  probably  renesting  attempts;  however,  I did  not 
count  these  as  renestings  because  I could  not  identify 
any  previous  nests. 

A reasonable  estimate  of  the  proportion  of  pairs  pro- 
ducing offspring  was  not  possible  because  for  25  of 
the  estimated  33  vireo  pairs,  only  one  or  two  nesting 
attempts  were  observed  for  each  pair.  Therefore  I 
could  not  rule  out  the  possibility  that  other,  possibly 
successful,  nesting  attempts  were  not  detected. 

To  assess  the  timing  of  nest  initiation  on  nest  suc- 
cess, I conducted  the  following  analyses.  I compared 
vireo  nest  initiation  dates  (Julian  dates)  for  both  par- 
asitized and  unparasitized  nests  {n  = 56)  using  a 
Mann- Whitney  f/-test.  This  analysis  included  all  com- 
pleted, active  nests  found  during  building  or  laying 
except  for  those  unparasitized  nests  that  were  aban- 
doned early  in  the  nesting  cycle  (/;  = 7).  For  the  nests 
1 excluded  from  the  analyses,  I could  not  rule  out  the 
possibility  that  parasitism  might  have  occurred  had  the 
nest  remained  in  use.  I also  compared  the  nest  initia- 
tion dates  for  both  depredated  and  fledged  (fledged  ei- 
ther cowbird  or  vireo  young)  nests  {n  = 23)  using  a 
Mann- Whitney  t/-test.  Included  in  this  analysis  were 
completed,  active  nests  found  during  building  or  lay- 
ing that  were  either  depredated  or  fledged.  Finally,  us- 
ing a Mann-Whitney  t/-test,  I compared  vireo  nest  ini- 
tiation dates  for  successful  (fledged  vireos)  and  unsuc- 
cessful (all  other  fates)  nests.  All  63  complete,  active 
nests  found  during  building  or  laying  were  included. 
t/-test  F-values  were  obtained  from  Sokal  and  Rohlf 
(1987). 

I included  28  variables  describing  vegetation  sur- 
rounding nests  in  a step-wise  discriminant  function 
analysis  (using  PROC  STEPDISC,  SAS  6.12,  for  a 
UNIX  operating  system)  to  compare  parasitized  to  un- 
parasitized nests.  I set  the  critical  P-value  for  entering 
and  remaining  in  the  model  at  0.05.  I included  nests 
found  at  all  stages  of  the  nesting  cycle  for  which  I had 
measured  vegetation  (unparasitized  n = 15,  parasitized 
n = 50)  except  for  those  unparasitized  nests  that  were 
abandoned  early  in  the  nesting  cycle.  For  these  aban- 
doned nests,  I could  not  rule  out  the  possibility  that 
parasitism  might  have  occurred  had  the  nest  remained 
in  use. 

RESULTS 

Of  the  63  completed  and  active  Bell’s  Vireo 
nests  found  during  nest  building  and  laying, 
44  (70%)  were  parasitized  by  at  least  one 
cowbird  egg  but  only  3 of  these  fledged  a 
cowbird  young  (Table  1).  None  of  the  para- 
sitized nests  fledged  any  vireo  young.  A mean 
of  1.5  cowbird  eggs  were  laid  in  each  para- 
sitized nest,  and  a mean  of  1.5  vireo  eggs  were 
present  in  each  such  nest  after  cowbird  activ- 
ity (possibly  egg  removal;  Brown  1993). 

Of  the  44  parasitized  Bell’s  Vireo  nests  in- 
cluded in  this  analysis,  in  only  4 (9%)  did  the 
vireo  parents  use  additional  nest  material  to 


cover  one  or  more  cowbird  eggs  laid  in  their 
nests.  All  nests  with  buried  eggs  subsequently 
failed  for  a variety  of  reasons  (Table  1).  In 
none  of  the  nests  with  buried  eggs  could  1 rule 
out  the  possibility  that  cowbird  eggs  had  been 
buried  during  the  process  of  nest  building  be- 
cause they  were  laid  in  nests  under  construc- 
tion. 

Nest  abandonment  following  cowbird  par- 
asitism in  my  study  was  frequent.  Of  the  43 
parasitized  nests  (does  not  include  1 nest 
abandoned  after  vireos  had  buried  cowbird 
egg,  see  Table  1),  32  were  abandoned.  This  is 
a significantly  higher  proportion  of  abandon- 
ment than  that  expected  based  on  the  frequen- 
cy of  abandonment  for  unparasitized  nests  (8 
of  20;  = 21.22,  P < 0.001). 

Abandoned  nests  had  significantly  fewer 
host  eggs  than  non-abandoned  nests  [aban- 
doned: X = 0.9  ±0.1  (SE);  non-abandoned:  x 
= 3.3  ± 0.1;  t = -7.04,  P < 0.001].  Aban- 
donment was  not  significantly  related  to  the 
number  of  cowbird  eggs  laid  (abandoned:  x = 
1.6  ± 0.1;  non-abandoned:  x = 1.2  ± 0.1;  f 
= 1.45,  P > 0.05). 

Of  63  nests,  1 estimated  20  (32%)  were  re- 
nesting attempts  and  10  (16%)  were  probable 
renesting  attempts.  Of  the  8 nests  that  fledged 
vireo  young,  6 appeared  to  have  been  renest- 
ing attempts. 

Unparasitized  nests  («  =12)  were  initiated 
significantly  later  {U  — 378,  P < 0.05)  than 
parasitized  nests  {n  = 44;  Fig.  1).  No  differ- 
ence in  initiation  date  was  found  between  dep- 
redated {n  = 12)  and  fledged  (/?  = 11)  nests 
in  initiation  date  ((/  = 91,  P > 0.05;  Fig.  1). 
Successful  nests  (/?  = 8)  did  not  differ  from 
failed  nests  (n  = 55)  in  date  of  initiation  {U 
= 235.5,  P > 0.5;  Fig.  1). 

The  nest  substrate  species  was  selected  by 
step-wise  discriminant  analysis  as  a significant 
predictor  of  cowbird  parasitism  (F  = 5.29,  P 
= 0.0248,  F = 0.08).  Unparasitized  nests 
were  surrounded  within  5 m by  more  of  the 
plant  species  in  which  the  nest  was  placed 
than  were  parasitized  nests.  No  other  vegeta- 
tion variables  distinguished  parasitized  from 
unparasitized  nests. 

DISCUSSION 

During  my  one  season  of  study,  abandon- 
ment (and  apparent  renesting)  was  the  most 
common  response  of  Bell’s  Vireos  to  brood 


number  of  nests  number  of  nests  number  of  nests 


502 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


18 

15 

12 

9 

6 

3 

0 


25 

20 

15 

10 

5 

0 


nest  initiation  dates 


failed  nests 


\///A  successful  nests 


1c 


J7A ^ 


GO 

CN 

>N 

CO 


O) 

CN 


h- 

CN 

0) 

c 

3 


0) 

c 

3 

”5 

00 

CN 


3 


3 

“D 

00 


LIG.  1.  Dates  of  Bell’s  Vireo  nest  initiation  (in  10 
day  intervals;  May-July  1996)  for  (1  A)  parasitized  and 
unparasitized  nests,  (IB)  depredated  and  fledged  (ei- 
ther cowbird  or  vireo  young)  nests,  and  ( 1C)  failed  and 
successful  (Bell’s  Vireo  fledged)  nests. 


parasitism  by  cowbirds.  Abandoned  nests  had 
fewer  host  eggs  than  non-abandoned  nests. 
This  result  is  consistent  with  other  findings 
that  egg  removal  by  cowbirds,  rather  than  the 
presence  of  cowbird  eggs  in  the  nest,  is  the 
stimulus  that  leads  to  nest  abandonment  (Bar- 
low  1962,  Hill  and  Sealy  1994,  Woodworth 
1997).  It  is  also  consistent  with  the  hypothesis 
that  nest  abandonment  is  a generalized  re- 
sponse to  egg  loss  as  opposed  to  a specific 
response  to  parasitism  (Rothstein  1975). 

The  seasonal  activity  of  Brown-headed 
Cowbirds  could  be  a factor  favoring  nest 
abandonment  by  the  Bell’s  Vireo.  Unparasit- 
ized vireo  nests  were  initiated  significantly 
later  in  the  season.  Egg  laying  by  cowbirds 


declined  more  quickly  than  vireo  nest  initia- 
tion over  the  breeding  season  at  my  study  site; 
a finding  similar  to  other  studies  (Scott  and 
Ankney  1980,  Hill  and  Sealy  1994).  Those 
Bell’s  Vireos  that  nest  later  therefore  are  less 
likely  to  be  parasitized.  This  suggests  that  nest 
abandonment  followed  by  renesting  is  bene- 
ficial for  the  vireos.  No  costs  to  later  nesting 
were  detected;  neither  depredated  nests  nor 
nests  that  failed  from  all  causes  differed  in 
initiation  date  from  successful  nests.  However, 
post-fledging  success  was  not  followed. 

Nest  abandonment  may  be  a complemen- 
tary tactic  to  egg  burial.  Abandonment  may 
be  effective  at  high  rates  of  parasitism  while 
egg  burying  may  be  effective  at  lower  rates. 
In  this  study,  burying  of  cowbird  eggs  was 
rare  and  was  not  a successful  tactic,  partially 
because  of  subsequent  cowbird  parasitism. 
However,  when  rates  of  parasitism  are  lower 
(i.e.,  with  a lower  probability  of  subsequent 
cowbird  eggs  being  laid)  this  behavior  might 
be  beneficial.  Burial  is  probably  less  expen- 
sive energetically  than  constructing  a new  nest 
(Clark  and  Robertson  1981).  Frequency  of 
parasitism  on  my  study  site  may  be  unusually 
high  in  comparison  to  the  Great  Plains  as  a 
whole  (29%  of  nests  parasitized,  Friedmann  et 
al.  1977;  13-69%  of  nests  parasitized.  Brown 
1993).  If  this  is  so,  then  this  study  may  un- 
derestimate the  importance  of  egg  burying  in 
allowing  Great  Plains  Bell’s  Vireos  to  persist 
in  the  presence  of  cowbirds. 

Egg  burying  has  not  been  reported  for 
Bell’s  Vireos  in  California  (Salata  1983,  Gray 
and  Greaves  1984,  Franzreb  1989).  Cowbirds 
have  occupied  most  of  California  only  in  the 
past  century  (Laymon  1987),  so  their  hosts 
there  may  not  have  had  enough  time  to  evolve 
adaptive  responses  to  brood  parasitism  (May- 
field  1965). 

I did  not  attempt  to  compare  cowbird  in- 
duced nest  abandonment  rates  in  my  study  to 
those  reported  from  California  because  of  two 
potentially  confounding  factors.  Unlike  my 
study,  in  California  cowbird  eggs  were  re- 
moved by  researchers  (Salata  1983,  Gray  and 
Greaves  1984,  Franzreb  1989).  Therefore,  the 
observed  rate  of  abandonment  in  California 
may  be  reduced  because  not  all  vireos  that 
abandon  parasitized  nests  do  so  immediately 
upon  receiving  a cowbird  egg  (pers.  obs.). 
Secondly,  usually  only  one  cowbird  egg  was 


Parker  • BROOD  PARASIl'ISM  OF  BELL’S  VIREOS 


503 


laid  per  nest  in  California  (Salata  1983,  Gray 
and  Greaves  1984,  Franzreb  1989),  possibly 
coinciding  with  removal  of  only  one  host  egg 
by  the  cowbirds  (Lowther  1993).  The  lower 
intensity  of  parasitism  in  California  than  in 
Kansas  could  mean  a less  intense  proximate 
cue  for  vireos  to  abandon  in  California.  This 
could  lead  to  the  observation  of  different 
abandonment  tendencies  in  these  two  vireo 
populations  regardless  of  the  presence  or  ab- 
sence of  any  evolved  differences  between 
them. 

To  better  understand  cowbird  behavior  and 
the  possibility  for  cryptic  nest  placement  by 
vireos,  I considered  the  relationship  between 
nest-site  vegetation  and  parasitism.  I found 
that  unparasitized  nests  had  more  ground  cov- 
ered (within  5 m of  the  nest)  by  the  plant  spe- 
cies in  which  a given  nest  was  placed  (nest 
substrate).  However,  this  finding  does  not  nec- 
essarily support  the  idea  that  an  increased  den- 
sity of  vegetation  generally  hinders  searching 
by  cowbirds  because  no  other  measures  of 
vegetation  density  were  associated  with  brood 
parasitism.  Although  cowbird  parasitism 
seems  to  be  affected  by  vegetation  structure 
in  forests  (Brittingham  and  Temple  1996), 
such  effects  were  not  apparent  in  this  study. 
The  predictive  value  of  the  variable  ‘nest  sub- 
strate’ was  low  (r^  = 0.08).  With  such  a weak 
relationship  between  nest  placement  and 
brood  parasitism,  cowbirds  may  be  a negligi- 
ble selective  pressure  further  refining  nest 
placement  in  the  Bell’s  Vireo. 

ACKNOWLEDGMENTS 

I would  like  to  thank  J.  L.  Zimmerman  for  his  ideas 
as  I conducted  this  research.  I also  wish  to  thank  J.  E 
Cavitt  for  his  help  during  data  collection.  R.  Kimball 
and  C.  Fellows  provided  useful  suggestions  during 
manuscript  preparation.  Anonymous  reviewers  made 
helpful  comments  on  earlier  versions  of  this  manu- 
script. I conducted  this  study  at  the  Konza  Prairie  Re- 
search Natural  Area  Long  Term  Ecological  Research 
Site  and  I utilized  resources  provided  by  Konza  LTER. 
Nest  site  measurement  protocol  and  data  entry  pro- 
gram were  provided  by  BBIRD. 

LITERATURE  CITED 

Barlow,  J.  C.  1962.  Natural  history  of  the  Bell  Vireo, 
Vireo  hellii  Audubon.  Univ.  Kans.  Publ.  Mus. 
Nat.  Hist.  12:241-296. 

Brawn,  J.  D.  and  S.  K.  Robinson.  1996.  Source-sink 
population  dynamics  may  complicate  the  interpre- 
tation of  long  term  census  data.  Ecology  77:3-12. 


Brittingham,  M.  C.  and  S.  A.  Templk.  1996.  Vege- 
tation around  parasitized  and  non-parasitized  nests 
within  deciduous  forest.  J.  Field  Ornithol.  67:406- 
413. 

Brown,  B.  T.  1993.  Bell’s  Vireo  (Vireo  hellii).  In  The 
birds  of  North  America,  no.  35  (A.  Poole,  P.  Stet- 
tenheim,  and  F Gill,  Eds.).  The  Academy  of  Nat- 
ural Sciences,  Philadelphia,  Pennsylvania;  The 
American  Ornithologists’  Union,  Washington, 
D.C. 

Clark,  K.  L.  and  R.  J.  Robertson.  1981.  Cowbird 
parasitism  and  evolution  of  anti-parasite  strategies 
in  the  Yellow  Warbler.  Wilson  Bull.  93:249-258. 

Franzreb,  K.  E.  1987.  Endangered  status  and  strate- 
gies for  conservation  of  the  Least  Bell’s  Vireo 
(Vireo  hellii  pusillu.s)  in  California.  West.  Birds 
18:43-49. 

Franzreb,  K.  E.  1989.  Ecology  and  conservation  of 
the  endangered  Least  Bell’s  Vireo.  Biological  Re- 
port 89(1).  U.S.  Fish  and  Wildlife  Service,  Wash- 
ington, D.C. 

Friedmann,  H.,  L.  E Kief,  and  S.  I.  Rothstein.  1977. 
A further  contribution  to  knowledge  of  the  host 
relations  of  the  parasitic  cowbirds.  Smithson.  Con- 
trib.  Zool.  235:1—75. 

Gray,  M.  V.  and  J.  M.  Greaves.  1984.  Riparian  for- 
ests as  habitat  for  the  Least  Bell’s  Vireo.  In  Cal- 
ifornia riparian  systems:  ecology,  conservation, 
and  productive  management  (R.  E.  Warner  and  K. 
M.  Hendrix,  Eds.).  Univ.  of  California  Press, 
Berkeley. 

Hill,  D.  P.  and  S.  G.  Sealy.  1994.  Desertion  of  nests 
parasitized  by  cowbirds:  have  Clay-colored  Spar- 
rows evolved  an  anti-parasite  defense?  Anini.  Be- 
hav.  48:1063-1070. 

Laymon,  S.  a.  1987.  Brown-headed  Cowbirds  in  Cal- 
ifornia: historical  perspectives  and  management 
opportunities  in  riparian  habitats.  West.  Birds  18: 
63-70. 

Lowther,  P.  E.  1993.  Brown-headed  Cowbird  (Mol- 
othru.s  ater).  In  The  birds  of  North  America,  no. 
47  (A.  Poole,  P.  Stettenheim,  and  F.  Gill,  Eds.). 
The  Academy  of  Natural  Sciences,  Philadelphia, 
Pennsylvania;  The  American  Ornithologists’ 
Union,  Washington,  D.C. 

Martin,  T.  E.  and  G.  R.  Geupel.  1993.  Nest-monitor- 
ing  plots:  methods  for  locating  nests  and  moni- 
toring success.  J.  Field  Ornithol.  64:507-519. 

Mayfield,  H.  1965.  The  Brown-headed  Cowbird,  with 
old  and  new  hosts.  Living  Bird  4:12-28. 

Neudorf.  D.  L.  and  S.  G.  Sealy.  1994.  Sunrise  ne.st 
attentiveness  in  cowbird  hosts.  Condor  96:162— 
169. 

Robbins,  M.  B.,  D.  A.  Easterla,  and  D.  Mead.  1992. 
Avian  census  of  the  Nodaway  River,  northwestern 
Missouri.  Bluebird.  59:105-107. 

Robbins,  M.  B.,  D.  A.  Ea.sterla,  and  D.  Mead.  1993. 
1993  avian  census  of  the  Nodaway  River,  north- 
western Missouri.  Bluebird.  60:110-1  1 1. 

Robertson,  R.  J.  and  R.  E Norman.  1977.  The  func- 
tion and  evolution  of  aggressive  host  behavior  to- 


504 


THE  WILSON  BULLETIN  ♦ Vol.  Ill,  No.  4,  December  1999 


wards  the  Brown-headed  Cowbird  (Molothrus 
ater).  Can.  J.  Zool.  55:508-518. 

Rothstein,  S.  I.  1975.  An  experimental  and  teleonom- 
ic  investigation  of  avian  brood  parasitism.  Condor 
77:250-271. 

Salata,  L.  R.  1983.  Status  of  the  Least  Bell’s  Vireo 
on  Camp  Pendleton,  California.  U.S.  Fish  and 
Wildlife  Service,  Laguna  Niguel,  California. 

Scott,  D.  M.  and  C.  D.  Ankney.  1980.  Fecundity  of 
the  Brown-headed  Cowbird  in  southern  Ontario. 
Auk  97:677-687. 


Sealy,  S.  G.  1996.  Evolution  of  host  defenses  against 
brood  parasitism:  implication  of  puncture-ejection 
by  a small  passerine.  Auk  1 13:346-355. 

SOKAL,  R.  R.  AND  E J.  Rohle.  1987.  Introduction  to 
biostatistics,  second  ed.  W.  H.  Freeman  and  Com- 
pany, New  York. 

Uyehara,  j.  C.  and  P.  M.  Narins.  1995.  Nest  defense 
by  Willow  Flycatchers  to  brood-parasitic  intrud- 
ers. Condor  97:361—368. 

Woodworth,  B.  L.  1997.  Brood  parasitism,  nest  pre- 
dation, and  season-long  reproductive  success  of  a 
tropical  island  endemic.  Condor  99:606-621. 


Wilson  Bull.,  I 1 1(4),  1999,  pp.  505-514 


THE  TYPE  B SONG  OF  THE  NORTHERN  PARULA:  STRUCTURE 
AND  GEOGRAPHIC  VARIATION  ALONG  PROPOSED 
SUB-SPECIES  BOUNDARIES 

MICHAEL  D.  BAY'  2 


ABSTRACT. — The  type  B song  of  the  Northern  Parula  (Parula  aniericana)  was  described  from  120  males 
recorded  throughout  much  of  the  species  range  in  North  America.  Most  songs  were  structured  with  a series  of 
complex  syllables,  followed  by  simple  syllables,  trill  syllables,  and  a terminal  simple  syllable.  Some  birds  sang 
songs  that  contained  2 phrases  per  song  with  syllables  that  varied  in  structure  and  number  between  individuals. 
Analysis  ot  song  variables  revealed  variation  at  the  macrogeographic  level  with  songs  from  western  populations 
differing  significantly  from  eastern  populations  in  song  duration,  frequency,  number  of  trill  syllables,  and  simple 
syllables.  In  addition,  variation  was  evident  between  eastern  and  western  populations  in  the  structuring  of  phrase 
patterns.  Received  8 March  J999.  accepted  3 August  1999. 


Several  investigations  into  wood-warbler 
(Parulidae)  song  behavior  have  shown  that 
some  species  sing  two  song  types  (Ficken  and 
Ficken  1967,  Morse  1967,  Highsmith  1989) 
while  other  species  have  a repertoire  of  sev- 
eral songs  classified  as  first  and  second  cate- 
gory songs  (Lein  1978,  Staicer  1989,  Byers 
1995).  The  two  song  types  have  been  referred 
to  as  types  A and  B (Morse  1967,  Nolan 
1978),  types  I and  II  (Lanyon  and  Gill  1964, 
Gill  and  Murray  1972,  Morrisson  and  Hardy 
1983),  or  accented  and  unaccented  ending 
songs  (Morse  1966,  Lein  1978).  Studies  on 
the  function  of  song  suggest  that  type  A,  I,  or 
accented  ending  songs  [or  the  type  B in 
Black-throated  Green  Warblers  (Dendroica  vi- 
rens)  and  Blackburnian  Warblers  (D.  fusca)] 
are  used  as  intersexual  signals  and  are  more 
stereotyped;  while  B,  II,  or  unaccented  ending 
songs  (or  the  type  A in  Black- throated  Green 
and  Blackburnian  warblers)  are  used  intrasex- 
ually  and  tend  to  be  more  variable  (e.g., 
Kroodsma  1981,  reviewed  by  Spector  1992). 

Although  detailed  descriptions  of  parulid 
song  types  can  be  found  for  several  species,  a 
few,  like  the  Northern  Parula  {Parula  ameri- 
cana),  are  less  well  studied.  Moldenhauer 
(1992)  presented  a detailed  account  of  the 
type  A song,  but  the  type  B song  has  yet  to 
be  described  in  detail  spectrographically.  In 
this  study,  I present  a description  of  B songs 


' Dept,  of  Biological  Sciences.  Sam  Houston  State 
Univ.,  Huntsville,  TX  77341. 

^ Present  address:  Dept,  of  Biology,  East  Central 
Univ.,  Ada,  OK  74820; 

E-mail : mbay @ mailclerk.ecok.edu 


recorded  from  several  males  located  through- 
out the  species’  breeding  range  (eastern  Unit- 
ed States  and  southeastern  Canada). 

METHODS 

Type  B songs  of  29  male  Parula  Warblers  were  re- 
corded from  20  localities  in  Texas,  Alabama,  Missis- 
sippi, Louisiana,  Georgia,  and  Tennessee  (Appendix) 
from  15  May  to  10  June  1986,  during  the  morning 
hours  of  07:00-10:00  (CST).  Although  no  birds  were 
color  marked,  only  1—3  individuals  were  recorded  per 
locality,  with  two  investigators  recording  different  in- 
dividuals, often  at  the  same  time.  Type  B songs  were, 
in  most  instances,  elicited  by  song  playback.  These 
recordings  (46  songs)  were  made  using  a Uher  4000 
Report  IC  recorder  at  a tape  speed  of  19  cm/sec  and  a 
Dan  Gibson  P650  parabolic  microphone.  Spectrograms 
of  recorded  songs  were  produced  with  a Kay  Eleme- 
trics  606 IB  sonagraph  with  a wideband  filter.  Addi- 
tional song  recordings  were  obtained  from  the  Texas 
Bird  Sound  Library  (Department  of  Biological  Scienc- 
es, Sam  Houston  State  University;  33  birds,  50  songs), 
Cornell  Library  of  Natural  Sounds  (15  birds,  27 
songs),  and  the  Borror  Laboratory  of  Bioacoustics  of 
The  Ohio  State  University  (43  birds,  58  songs).  These 
songs  were  recorded  using  Nagra  III  (38  cm/sec)  or 
Magnemite  (38  cm/sec)  recorders.  An  AKG  micro- 
phone (or  an  unknown  type)  was  used  with  a parabola 
to  record. 

The  following  variables  were  measured  for  each 
song:  (1)  duration  of  song  (sec),  (2)  total  number  of 
syllables,  (3)  number  of  syllable  types,  (4)  minimum 
frequency,  and  (5)  maximum  frequency.  For  individ- 
uals with  multiple  recordings,  1 computed  a within-bird 
mean.  For  the  sake  of  comparison  with  variation  as 
reported  in  the  type  A song  (Moldenhauer  1992),  I 
used  analysis  of  variance  (ANOVA;  SAS  Institute 
1985;  a = 0.05)  to  test  for  significant  differences  be- 
tween song  populations  from  three  geographic  areas. 
Moldenhauer  (1992)  primarily  reported  on  the  differ- 
ences between  songs  in  the  western  versus  eastern  re- 
gions of  the  species’  breeding  range.  I followed  this 


505 


506 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


M if  i 

^ 0 30  see  ^ 

FIG.  1.  Syllable  classes  of  Northern  Parula  type  B 
songs.  (A)  complex  syllable  (B)  simple  syllable  (C) 
trill  syllable. 


division  with  the  exception  that  I divide  the  eastern 
population  into  northeast  and  southeast  (see  Appen- 
dix). These  geographical  ranges  correspond  to  the  sub- 
species (western  race,  P.  a.  hidoviciancc,  southeastern 
race,  P.  a.  americaner,  northeastern  race,  P.  a.  pusillci) 
proposed  by  Oberholser  (1974). 

1 classified  song  elements  based  on  morphological 
structure  (Baptista  1974,  Weins  1982,  Bradley  and 
Bradley  1983)  using  terminology  partly  adopted  from 
Baptista  (1974)  and  Staicer  (1989).  A note  was  any 
short  tracing  on  the  spectrogram,  and  syllables  were 
represented  by  repeated  notes  or  series  of  notes  form- 
ing a coherent  unit.  Three  classes  of  syllables  were 
identified:  (1)  simple  syllables  (SS),  those  containing 
1 or  2 simple  notes;  (2)  complex  syllables  (CS),  syl- 
lables with  more  than  2 notes  forming  a coherent  unit, 
or  in  rare  instances  1-2  wavy  and  continuous  notes; 
(3)  trill  syllables  (TS),  high  frequency  slurs  near  the 
end  of  a song  (Fig.  1).  To  distinguish  between  the 
number  of  different  variations  or  types  within  each  syl- 
lable class,  a subscript  number  was  added  to  a sylla- 
ble’s abbreviation  (e.g.,  CS14). 

A phrase  was  defined  as  a series  of  repeated  sylla- 
bles forming  the  following  4 phrase  classes  of  a B 
song:  (1)  complex  (C)  phra.se,  composed  of  complex 
syllables  of  one  type;  (2)  mixed  (M)  phra.se,  composed 
of  a mixture  of  repeated  complex  and  simple  syllables 
of  one  type  each;  (3)  trill  (T)  phrase,  composed  of  trill 
syllables  of  one  type;  (4)  repeated  trill  (R)  phrase, 
composed  of  trill  syllables  of  2 types  (Fig.  2).  A phrase 
pattern  was  the  entire  sequence  of  phrases  in  a song 
and  was  symbolized  by  the  letter  codes  for  each  phrase 
class  (e.g.,  C-T-SS  = a song  composed  of  a C phrase, 
T phrase,  and  ending  with  a simple  syllable;  Fig.  3). 

RESULTS 

Song  structure. — Thirty-nine  syllable  types 
from  3 syllable  classes  were  discovered  in  the 
type  B song,  including  21  complex  syllables 
(Fig.  4),  1 1 simple  syllables,  and  7 trill  syl- 
lables (Fig.  5).  It  appeared  that  no  one  indi- 
vidual contained  more  than  one  complex  syl- 
lable type  and  no  more  than  2 types  of  simple 
and  trill  syllables. 

Each  of  the  syllable  classes  (complex,  sim- 


81  A 


8 D 


\mm 

3J 

6 05  L5 

Time  (sec) 

FIG.  2.  Phrase  classes  of  Northern  Parula  type  B 
songs.  (A)  C phrase  (B)  M phrase  (C)  T phrase  (D)  R 
phrase. 


pie,  and  trill)  was  used  to  construct  song 
phrases  that  typically  began  with  an  introduc- 
tory series  of  complex  syllables  {x  = 3.6  syl- 
lables per  song,  SD  = 1.4,  n = 181  songs) 
followed  by  a high  frequency  trill  (T  phrase) 
(x  = 4.6  syllables  per  song,  SD  = 2.7,  n = 
181  songs)  and  ending  in  one  to  several  sim- 
ple syllables  (SS;  x = 0.88  syllables  per  song, 
SD  = 2.6,  n = 181  songs).  Some  birds  sang 
songs  that  lacked  a few  of  these  syllable  class- 
es (e.g.,  only  the  introductory  complex  sylla- 
bles were  present)  and  therefore  were  shorter 
and  usually  more  difficult  to  hear  (low  ampli- 
tude; e.g.,  somewhat  like  muted  songs;  Morse 
1967).  The  number  of  syllables  within  a song 
ranged  from  2 to  19.5  (x  = 10.7),  while  most 
(54%)  were  composed  of  10  to  15  syllables. 
Few  songs  contained  more  than  15  syllables 
(12.6%).  Some  birds  began  their  song  with 
chip  notes  (see  Morse  1967)  but  this  was  rare- 
ly recorded  and  was  not  considered  in  this 
analysis. 

The  most  frequent  phrase  pattern,  C-T-SS 
(complex-trill-simple  syllable),  accounted  for 
most  of  the  songs  (67%).  Nine  different  pat- 
terns were  found,  the  five  most  common  are 
illustrated  in  Fig.  6.  It  is  likely  that  some  ex- 
tremely rare  patterns  may  represent  scrambled 


Bay  • NORTHERN  PARULA  TYPE  B SONG 


507 


i- ^ f 1 ^ ^ ^ 1 ^ ] ^ 1 hHf--U-Hh-]hHfH  h-H 

CSi3  syllables  TSs  syllables  SS9 

[ - - Song  Duration - - -I 

EIG.  3.  Measurements  and  coding  of  a type  B song.  MxF  = maximum  frequency,  MnF  = minimum  fre- 
quency, CS  = complex  syllable,  TS  = trill  syllable,  SS  = simple  syllable. 


(e.g.,  T-M-T-SS  occurred  in  0.4%  of  songs)  or 
incomplete  song  phrases  (e.g.,  M occurred  in 
1.4%  of  songs  and  M-T  occurred  in  1.5%  of 
songs). 

Geographic  variation. — Eight  phrase  pat- 
terns (except  for  the  phrase  C-T-SS)  occurred 
in  populations  that  occupied  the  western  half 
of  the  species’  breeding  range  and  correspond 
to  those  birds  that  sang  the  western  type  A 
song  (see  Moldenhauer  1992).  Birds  occupy- 
ing the  eastern  range  and  corresponding  to 
those  that  sing  the  eastern  type  A song,  sang 
B songs  that  were  constructed  of  only  three 
phrase  patterns  (C-T-SS,  C-T,  C;  Fig.  6).  The 
additional  variation  in  phrase  patterns  ob- 
served for  the  western  population  is  attributed 
to  the  addition  of  simple  and  complex  sylla- 
bles to  the  introductory  portion  of  the  song  to 
form  M phrases.  This  phrase  difference  ap- 
pears to  change  or  overlap  between  the  Mis- 
sissippi and  Alabama  boundary  much  as  oc- 
curs with  the  type  A song  (Moldenhauer 
1992).  However,  more  data  are  needed  to  de- 
termine the  distribution  of  type  B song  phrase 
patterns  within  the  east  to  west  type  A song 
overlap  zone. 

In  addition  to  phrase  variation,  songs  varied 
significantly  between  geographic  areas  in 
mean  duration,  mean  maximum  frequency, 
and  in  the  mean  number  of  simple  and  trill 
syllables  (Table  1).  Songs  of  western  birds  av- 
eraged 0.10  second  longer  than  songs  of 
southeastern  birds  and  0.20  second  longer 
than  songs  of  northeastern  birds  {F  = 10.90, 
df  = 2,  134,  P < 0.001).  Birds  from  the  latter 
region  averaged  0.5  kHz  less  than  the  average 


maximum  frequency  of  western  and  south- 
eastern populations  (F  = 4.31,  df  = 2,  134,  P 

< 0.025). 

Comparing  the  usage  of  syllable  types  with- 
in the  3 syllable  classes  (complex,  simple,  and 
trill)  among  birds  in  the  three  geographic  ar- 
eas, the  most  frequent  syllable  for  western 
birds  was  CSjj  (56.6%;  Fig.  4,  no.  13)  fol- 
lowed by  Tj  (43.9%;  Fig.  5,  no.  8)  and  SS^ 
(29.2%;  Fig.  5,  no.  8).  For  birds  in  the  south- 
east, songs  most  frequently  contained  SS^ 
(47.3%;  Fig.  5,  no.  8)  followed  by  T,  and  T, 
(both  at  39.4%;  Fig.  5,  no.  1 and  2,  respec- 
tively). The  most  frequently  used  complex 
syllables  for  southeastern  songs  were  CS,3 
(18.4%;  Fig.  4,  no.  13)  and  CS4  (15.7%;  Fig. 

4,  no.  4).  Birds  from  the  northeastern  region 
primarily  used  T,  (51.2%;  Fig.  5,  no.  3),  CS,9 
(36.5%;  Fig.  4,  no.  19),  and  SS9  (34.1%;  Fig. 

5,  no.  9)  to  construct  their  songs.  Of  the  three 
syllable  types  used  to  construct  the  type  B 
song,  significant  differences  were  evident  be- 
tween the  three  geographical  areas  in  the 
mean  number  of  syllables  used  per  song  (Ta- 
ble 1).  For  instance,  western  singers  used  sig- 
nificantly more  simple  syllables  to  construct 
the  type  B song  (F  = 16.3,  df  = 2,  134,  P < 
0.005)  but  fewer  trill  syllables  in  comparison 
to  eastern  singers  (F  = 23.9,  df  = 2,  134,  P 

< 0.005). 

Only  within  specific  localities  did  some 
birds  sing  identical  songs  (i.e.,  same  phrase 
patterns  and/or  syllables),  while  others  sang 
songs  using  a different  type  of  one  or  more 
syllables  within  the  three  syllable  classes  (e.g., 
one  individual  might  use  08,3  and  T^,,  while 


Frequency  (IcHz) 


508 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


Complex  Syllables 


6- 

5- 

4- 

3- 

2- 

1- 

0- 


^ 0.30  sec  ^ 

FIG.  4.  Complex  syllables  of  Northern  Parula  type  B songs.  Syllable  types  1,  3,  5,  10,  11,  and  20  from 
eastern  birds  only,  while  variants  14,  16,  17,  and  18  from  we.stern  birds  only. 


another  used  CS4  and  TSO-  It  is  unknown 
whether  individuals  with  the  same  song  (i.e., 
same  phra.se  and  syllables)  within  a specific 
locality  were  neighbors  because  most  of  these 
songs  were  not  recorded  by  the  author.  It  is 
likely  that  some  of  these  individuals  were 
members  of  the  same  local  population  because 


many  of  these  songs  were  recorded  within  the 
same  year.  Some  morphological  differences 
were  evident  in  specific  syllable  types  be- 
tween individuals  (e.g.,  CS,,  might  differ 
slightly  between  two  birds;  Fig.  7).  Morpho- 
logical differences  in  simple  syllables  were 
evident  between  those  used  in  the  middle  of 


Frequency  (IdHz) 


Bay  • NORTHERN  PARULA  TYPE  B SONG 


509 


Simple  Syllables 


6- 

5- 

4- 

3- 

2- 

1- 

0- 


\/  7 X!  y 


7 


Trill  Syllables 


6- 

5- 

4- 

3- 

2- 

I- 

0- 


0.30  sec 

FIG.  5.  Simple  and  trill  syllable.s  of  Northern  Parula  type  B songs.  Simple  syllable  types  1,  3,  and  5 and 
trill  syllable  type  6 from  western  birds  only. 


Frequency  (IcHz) 


510 


THE  WILSON  BULLETIN  • Vol.  HI.  No.  4.  December  1999 


8 


3^ 

8i 


3^ 

8i 


3 

8 


3^ 

8i 


A 

I- 


• I 


I 


/ 


-C  ph  rase I I--T  phrase--I  simple  syllable 


8 


I M phrase I I — T phrase — I simple  syllable 


' / / A,/  f A'  / W 

. 1 ' 1 t 1 Ml' 

c 

I C phrase I I--T  phrase— I 


\i 


Ml  ^ 


fi  ‘M 


mm 


E 

I C phrase I 


6 0.5  EO  E6  ?0 

Time  (sec) 

FIG.  6.  The  five  most  common  song  or  phrase  patterns  observed  in  Northern  Panda  type  B songs.  (A) 
pattern  C-T-SS  [67%  of  all  .songs  (west  = 0.0%,  southeast  = 92%,  northeast  = 100%)],  (B)  pattern  M-T-SS 
1 13.1%  of  all  songs  (west  = 45.0%,  southeast  = 0.0%,  northeast  = 0.0%)],  (C)  pattern  C-T  [6.2%  of  all  songs 
(west  = 17%.  .southeast  = 2.6%,  northeast  = 0.0%)],  (D)  pattern  M-R-SS  [3.6%  of  all  songs  (west  = 12.1%, 
southeast  = 0.0%,  northeast  = ().()%)],  (E)  pattern  C [2.1%  of  all  .songs  (west  = 2.4%,  southeast  = 4.1%, 
northeast  = 0.0%)]. 


Hay  • NORTHERN  PARULA  TYPE  B SONG 


511 


TABLE  1.  A comparison  of  song  variables  from  western,  southeastern,  and  northeastern  type  B songs  of 
the  Northern  Panda  (values  are  mean  ± SD).  Number  of  individuals  in  which  songs  were  analyzed:  West  = 41, 
Southeast  = 38,  Northeast  = 41. 


Song  population 


ANOVA 


Variable 

West 

Southeast 

Northeast 

/■'-value'-' 

Song  duration  (s)“ 

1.5 

-h 

0.2 

1.4 

-h 

0.2 

1.3 

-i- 

0.1 

10.90** 

Number  of  syllables 

11.9 

~h 

2.9 

11.3 

-h 

3.1 

11.4 

-h 

2.7 

0.53 

Syllable  types  in  a song 

4.0 

-h 

1.3 

3.7 

-h 

1.0 

3.6 

-h 

0.5 

1.26 

Maximum  frequency  (k.Hz)'“ 

7.0 

-h 

0.4 

7.2 

~h 

0.4 

6.6 

-h 

0.4 

4.31* 

Minimum  frequency  (kHz) 

4.4 

-h 

0.5 

4.4 

-h 

0.5 

4.2 

0.4 

2.10 

Complex  syllables''^’ 

3.7 

-h 

1.6 

3.5 

-h 

1.3 

3.4 

H- 

0.9 

0.71 

Simple  syllables"*’ 

4.6 

-h 

3.0 

1.1 

-h 

1.0 

1.0 

-F 

0.3 

16.3*** 

Trill  syllables"*’ 

2.9 

1.5 

5.1 

-+■ 

2.6 

6.5 

H- 

2.7 

23  9*** 

^ Multiple  Comparison  (Bonferroni  Correction;  a = 0.05):  Song  Duration  (W  # Se  = Ne),  Maximum  Frequency  (W  = Se  Ne).  Simple  Syllables  (W 
# Se  = Ne).  Trill  Syllables  (W  # Se  Ne). 

^ Number  per  song. 

>=*  P < 0.025.  **  P < 0.001.  ***  P < 0.0005. 


the  song  (to  form  M phrases)  versus  the  single 
simple  syllable  terminating  the  song  (as  usu- 
ally occurs  in  most  parula  B songs). 

DISCUSSION 

Song  structure. — Northern  Parula  B songs 
are  complex  and  show  much  intraspecific  var- 
iation. Song  complexity  occurs  because  most 
songs  contain  multiple  phrases  (usually  two) 
that  vary  in  syllable  types  between  individu- 
als. A similar  arrangement  occurs  in  B songs 
of  the  Grace’s  Warbler  (Dendroica  graciae; 
Staicer  1989)  and  the  type  II  song  of  the  Blue 
and  Golden-winged  warblers  (Gill  and  Murray 
1972,  Highsmith  1989).  Some  evidence  sug- 
gests that  complexity  in  B songs  may  be  a 
result  of  intrasexual  usage  (e.g.,  territorial 
clashes  between  males),  while  A songs,  used 
for  mate  attraction,  are  more  stereotyped 
(Kroodsma  1981). 

Geographic  variation. — Northern  Parula 
type  A songs  are  very  similar  in  most  vari- 
ables (except  trill  rate,  frequency,  and  song 
length),  but  differ  significantly  between  east- 
ern and  western  populations  in  the  type  of  ter- 
minal syllable  (Moldenhauer  1992).  Males 
from  each  population  recognize  and  respond 
differentially  to  the  two  types  of  A songs  (Re- 
gelski  and  Moldenhauer  1996).  Similarly,  the 
type  B song  exhibited  differences  between 
eastern  and  western  populations  in  song 
length  and  maximum  frequency,  but  primarily 
differed  in  the  number  of  specific  syllable 
types  and  phrase  patterns  composing  songs. 
For  instance,  western  birds  frequently  used 


more  simple  syllables  per  song  and  often  used 
these  syllables  to  construct  the  M phrase  pat- 
tern, which  was  absent  from  eastern  singers. 

Moldenhauer  (1992)  argued  for  subspecific 
recognition  of  P.  americana  (P.  a.  americana 
for  the  east  and  P.  a.  ludoviciana  for  the  west) 
based  upon  the  terminal  note  difference  in  the 
type  A songs.  Results  from  my  study  are  con- 
sistent with  this  division,  based  upon  B song 
phrase  patterning  between  eastern  and  western 
populations.  The  terminal  note  in  the  type  A 
song  is  readily  identifiable  both  audibly  and 
visually  by  sonogram.  Likewise,  B songs  with 
M phrases  (western  population)  may  be  au- 
dibly distinguished  from  songs  without  M 
phrases  (eastern  population);  they  are  longer 
(composed  of  more  syllables)  and  more  buz- 
zy.  These  differences  are  easily  viewed  by 
comparing  sonograms  (Fig.  6B,  D vs  Fig.  6 A, 
C).  Whether  these  structural  differences  in 
eastern  and  western  type  A and  B songs  are 
influenced  by  learning  and/or  have  a genetic 
basis  has  yet  to  be  determined. 

Populations  from  each  geographic  region 
contained  a repertoire  of  unshared  syllables 
(15.3%  confined  to  the  eastern  population, 
20.5%  confined  to  the  western  population); 
however,  many  (57%)  were  shared.  Syllable 
confinement  within  specific  macrogeographi- 
cal  areas  reported  in  my  study  might  be 
viewed  with  some  speculation.  For  instance, 
some  individual  songs  contained  more  than 
one  syllable  type  within  a syllable  class,  par- 
ticularly trill  and  simple  syllables.  No  individ- 


Frequency  (IcHz) 


512 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


A 


C 

6- 
5- 
4- 
3- 
2- 
I- 
0- 

' 0 30  see  * 

FIG.  7.  Examples  of  morphological  variations  of  specific  syllable  types  between  individuals  (A)  CSivIcom- 
plex  syllable),  (B)  SS,  (simple  syllable),  (OTSj  (trill  syllable). 


Bay  • NORTHERN  PARULA  TYPE  B SONG 


513 


uals  for  which  multiple  songs  were  examined 
sang  more  than  1 complex  syllable  type.  How- 
ever, because  of  the  small  sample  size  (only  5 
individuals  recorded  with  at  least  5 songs 
from  each),  these  results  should  be  viewed 
with  caution.  Within  specific  microgeographic 
areas,  syllable  sharing  could  be  variable  even 
though  a few  individuals  sang  some  or  all  of 
the  same  syllables. 

In  other  warbler  species  that  have  a song 
similar  to  the  type  B of  the  Northern  Parula, 
local  dialect  patterns  are  usually  evident, 
which  suggests  that  young  males  learn  B 
songs  from  neighboring  conspecifics  (Kroods- 
ma  1981).  Thus  an  individual  from  an  area 
that  contains  a song  similar  to  his  neighbors, 
will  be  more  effective  during  countersinging 
bouts  (Kroodsma  1981).  Playback  experi- 
ments show  that  Blue-winged  Warbler  males 
respond  more  intensely  to  their  local  type  II 
(B)  songs  but  do  not  differentiate  among  type 
I (A)  songs  of  different  localities  (Kroodsma 
et  al.  1984).  Although  Northern  Parula  B 
songs  tend  to  be  complex  and  differ  between 
locations  (in  either  phrase  pattern  or  syllable 
types  used),  Bay  (1987)  found  no  discemable 
patterns  or  dialects  in  songs  from  the  two  best 
sampled  areas  (representing  10  and  15  indi- 
viduals) in  Texas. 

Future  studies  should  concentrate  on  gath- 
ering information  concerning  site  fidelity  in 
successive  breeding  years  to  allow  a better  un- 
derstanding of  type  B song  structure  at  the 
microgeographic  level.  Such  data  might  also 
reveal  whether  young  males  learn  the  same 
types  of  syllables  and  phrase  patterns  in  suc- 
cessive years  or  different  ones  as  a result  of 
the  exposure  to  new  individuals.  In  addition, 
researchers  should  determine  what  role,  if  any, 
the  disjunct  winter  distribution  of  the  North- 
ern Parula  has  on  song  learning  and  the  geo- 
graphical differences  in  the  type  A and  B 
songs  on  the  breeding  grounds 

ACKNOWLEDGMENTS 

Many  thanks  to  R.  Moldenhauer  for  suggesting  this 
research  and  for  his  advice  throughout  the  study.  D. 
Spector,  C.  Staicer,  and  anonymous  reviewers  made 
valuable  suggestions  for  improving  the  manuscript.  A. 
Dewees  and  M.  Duggan  gave  helpful  advice  on  statis- 
tics. Part  of  this  study  was  caixied  out  to  fulfill  require- 
ments for  the  M.A.  degree  at  the  Dept,  of  Biological 
Sciences,  Sam  Houston  State  University. 


LITERATURE  CITED 

Baptista,  L.  E 1974.  The  effects  of  songs  of  wintering 
White-crowned  Spanows  on  .song  development  in 
.sedentary  populations  of  the  .species.  Z.  Tierpsy- 
chol.  34:147-171. 

Bay,  M.  D.  1987.  Singing  behavior  and  geographic 
variation  in  the  type  B song  of  the  Northern  Parula 
(Parula  americana).  M.A.  thesis,  Sam  Houston 
State  Univ.,  Huntsville,  Texas. 

Bradley,  D.  and  R.  Bradley.  1983.  Application  of 
sequence  comparison  to  the  study  of  bird  songs. 
In  Time  warps,  string  edits  and  macromolecules: 
the  theory  and  practice  of  sequence  comparison 
(D.  Sankoff  and  J.  Kruskal,  Eds.).  Proceedings  of 
the  Conference  on  Sequence  Comparison,  Mon- 
treal, Canada. 

Byers,  B.  E.  1995.  Song  types,  repertoires  and  song 
variability  in  a population  of  Chestnut-sided  War- 
blers. Condor  97:390-401. 

Eicken,  M.  S.  and  R.  W.  Eicken.  1967.  Singing  be- 
havior of  the  Blue-winged  Warblers  and  Golden- 
winged Warblers  and  their  hybrids.  Behaviour  28: 
149-181. 

Gill,  E B.  and  B.  G.  Murray.  1972.  Song  variation 
in  sympatric  Blue-winged  and  Golden-winged 
warblers.  Auk  89:625-643. 

Highsmith,  R.  T.  1989.  The  singing  behavior  of  Gold- 
en-winged Warblers.  Wilson  Bull.  101:36-50. 

Kroodsma,  D.  E.  1981.  Geographical  variation  and 
functions  of  song  types  in  warblers  (Parulidae). 
Auk  98:743-751. 

Kroodsma,  D.  E.,  W.  R.  Meservey,  A.  L.  Whitlock, 
AND  W.  M.  Vanderhaegen.  1984.  Blue-winged 
Warblers  (Vermivora  pinus)  “recognize”  dialects 
in  type  II  but  not  type  1 songs.  Behav.  Ecol.  So- 
ciobiol.  15:127-131. 

Lanyon,  W.  E.  and  E B.  Gill.  1964.  Spectrographic 
analysis  of  variation  in  the  songs  of  a population 
of  Blue-winged  Warblers  {Vermivora  pinus).  Am. 
Mus.  Novitates,  No.  2176:1  — 18. 

Lein,  M.  R.  1978.  Song  variation  in  a population  of 
Chestnut-sided  Warblers  (Dendroica  pensylvani- 
ca):  its  nature  and  suggested  significance.  Can.  J. 
Zool.  56:1266-1283. 

Moldenhauer,  R.  R.  1992.  Two  song  populations  of 
the  Northern  Parula.  Auk  109:215-222. 

Morrison,  M.  M.  and  J.  W.  Hardy.  1983.  Vocaliza- 
tions of  the  Black-throated  Gray  Warbler.  Wilson 
Bull.  95:640-643. 

Morse,  D.  H.  1966.  The  context  of  .songs  in  the  Yel- 
low Warbler.  Wilson  Bull.  78:444—455. 

Morse,  D.  H.  1967.  The  context  of  songs  in  the  Black- 
throated  Green  and  Blackburnian  warblers.  Wil- 
.son  Bull.  79:64-74. 

Nolan,  V.,  Jr.  1978.  The  ecology  and  behavior  of  the 
Prairie  Warbler,  Dendroica  discolor.  Ornithol. 
Monogr.  26:1-595. 

Oberholser,  H.  C.  1974.  The  bird  life  of  Texas,  2 
vols.  Univ.  of  Texas  Press,  Austin. 

Regelski,  D.  j.  and  R.  R.  Moldenhauer.  1996.  Dis- 


514 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4.  December  1999 


crimination  between  regional  song  forms  in  the 
Northern  Parula.  Wilson  Bull.  108:335-341. 

SAS  Institute.  1985.  U.ser's  guide:  statistics.  SAS  In- 
stitute Inc.,  Cary,  North  Carolina. 

Spector,  D.  a.  1992.  Wood  warbler  song  systems:  a 
review  of  paruline  singing  behaviors.  Curr.  Orni- 
thol.  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. 

Weins,  J.  a.  1982.  Song  pattern  variation  in  the  Sage 
Sparrow  (Amphispiza  belli):  dialects  or  epiphe- 
nomena?  Auk  99:208-229. 

APPENDIX 

Recording  localities  for  Northern  Parula 
type  B songs.  Localities  are  listed  by  geo- 
graphic region  and  then  alphabetically  by  state 
(or  country)  and  county,  parish,  or  province. 


Number  of  individuals  per  locale  are  indicated 
in  parenthesis. 

Southeast  region. — Alabama:  Dale  (1), 
Dallas  (1),  Clark  (1),  Covington  (2),  Monroe 

(2) ,  Wilcox  (4).  Florida:  Alatchua  (1),  Dade 

(3) ,  Flager  (1),  Jefferson  (2),  Polk  (1),  Wak- 
ulla (1).  Georgia:  Bulloch  (3),  Chatham  (4), 
Jenkins  (6),  Richmond  (1).  North  Carolina: 
New  Hanover  (3).  South  Carolina:  Richland 
(8).  Tennessee:  Dickson  (1),  Sevier  (1). 

Northeast  region. — Canada:  Gaspe  (1). 
Maine:  Lincoln  (44).  Ohio:  Franklin  (2),  Sen- 
eca (1).  West  Virginia:  Monogalia  (1). 

West  region. — Lousiana:  St.  Tammany  (2). 
Mississippi:  George  (3),  Hancock  (4).  Texas: 
Hardin  (2),  Montgomery  (2),  Trinity  (2), 
Walker  (25). 


Wilson  Bull.,  1 I 1(4),  1999,  pp.  515-527 


NESTING  BIOLOGY  OF  DICKClSSELS  AND  HENSLOW’S 
SPARROWS  IN  SOUTHWESTERN  MISSOURI 
PRAIRIE  FRAGMENTS 

MAIKEN  WINTER'  2 


ABSTRACT. — According  to  data  from  the  North  American  Breeding  Bird  Survey,  populations  of  Dickcissel 
(Spizci  americana)  and  Henslow’s  SpaiTow  {Ammodrannis  henslowii)  have  declined  severely  during  the  last  30 
years.  The  reasons  for  their  population  declines  seem  to  differ;  habitat  fragmentation  on  the  breeding  grounds 
has  been  suggested  to  have  little  negative  impact  on  Dickcissels,  but  appears  to  be  a major  reason  for  Henslow’s 
Sparrow  declines.  Previous  reports  on  the  status  of  Dickcissels  and  Henslow’s  Sparrows  largely  were  based  on 
density  estimates  without  considering  the  nesting  biology  of  the  two  species.  My  comparison  of  the  nesting 
biology  of  Dickcissel  and  Henslow’s  Spanow  provides  some  insight  into  potential  factors  that  might  contribute 
to  their  population  declines.  During  1995-1997,  I studied  the  nesting  biology  of  Dickcissels  and  Henslow’s 
Sparrows  in  fragments  of  native  tallgrass  prairie  in  southwestern  Missouri.  Both  species  had  similar  clutch  sizes, 
rates  of  hatching  success,  and  numbers  of  young  fledged  per  successful  nest.  Dickcissels  tended  to  have  lower 
rates  of  nesting  success  and  higher  rates  of  brood  parasitism  by  Brown-headed  Cowbirds  (Molothrus  ater)  than 
Henslow’s  Sparrows.  Although  several  vegetation  characteristics  at  the  nest  differed  between  successful  and 
depredated  nests  in  Dickcissels,  no  differences  were  found  between  successful  and  depredated  Henslow’s  Spar- 
row nests  or  between  parasitized  and  unparasitized  Dickcissel  nests.  My  results  indicate  that  Dickcissels  might 
reproduce  less  successfully  than  Henslow’s  Sparrows  in  southwestern  Missouri,  and  might  therefore  be  of  higher 
conservation  concern  on  the  breeding  ground  than  previously  thought.  Received  27  January  1999,  accepted  3 
June  1999. 


Data  from  the  North  American  Breeding 
Bird  Survey  indicated  that  populations  of 
Dickcissel  (Spiza  americana)  and  Henslow’s 
Sparrows  (Ammodramus  henslowii)  have  de- 
clined by  about  39%  and  91%,  respectively, 
during  the  last  30  years  (Peterjohn  et  al. 
1994).  The  reasons  for  the  declines  are 
thought  to  differ  between  the  two  species: 
Dickcissels  are  assumed  to  have  declined 
mainly  because  of  poisoning  on  their  South 
American  wintering  grounds  (Basili  and  Tem- 
ple 1995,  Basili  1997),  and  are  thought  to  be 
little  affected  by  breeding  habitat  loss  or  frag- 
mentation (Herkert  et  al.  1993).  In  contrast, 
the  population  decline  of  Henslow’s  Sparrows 
seems  to  be  mainly  caused  by  loss  and  frag- 
mentation of  suitable  grassland  habitat  on 
their  breeding  grounds  (Herkert  1994).  How- 
ever, status  assessments  of  Dickcissels  and 
Henslow’s  Sparrows  are  based  largely  on  es- 
timates of  density  or  relative  abundance,  with- 
out considering  the  breeding  ecology  of  the 
two  species.  A comparison  of  the  breeding 
ecology  of  Dickcissels  and  Henslow’s  Spar- 


' Division  of  Biological  Sciences,  Univ.  of  Missouri, 
Columbia,  MO  65211. 

^Current  address;  611  Winston  Ct,  Apt  4,  Ithaca, 
NY  14850;  E-mail;  mwinte02@syr.edu 


rows  might  provide  information  on  factors 
that  could  cause  differential  reproductive  suc- 
cess in  the  two  species.  Such  factors  might 
include  clutch  sizes,  rates  of  brood  parasitism 
by  Brown-headed  Cowbirds  {Molothrus  ater), 
rates  of  nest  predation,  and  hatching  and 
fledging  rates.  Vegetation  characteristics  at  the 
nest  site  might  differ  between  the  species  and 
cause  one  species  to  be  more  susceptible  to 
nest  predation  or  cowbird  parasitism. 

In  southwestern  Missouri,  little  information 
has  been  collected  on  the  nesting  success  of 
passerines  breeding  in  tallgrass  prairie  frag- 
ments. In  this  study  I describe  and  compare 
nesting  characteristics  of  the  Dickcissel  and 
the  Henslow’s  Sparrow  in  fragments  of  native 
tallgrass  prairie  in  southwestern  Missouri  be- 
tween 1995  and  1997.  Detailed  analyses  on 
the  effect  of  fragment  size,  proximity  to  hab- 
itat edge,  management  practices,  and  land- 
scape structure  on  density  and  nesting  success 
of  these  species  are  described  elsewhere  (Win- 
ter 1998,  Winter  and  Faaborg  in  press). 

Dickcissel. — Dickcissels  are  grassland  hab- 
itat generalists;  they  can  be  found  breeding  in 
a wide  variety  of  grassland  vegetation  (Bent 
1968).  Because  males  often  sing  from  elevat- 
ed perches  and  females  often  place  their  nests 
above  the  ground,  they  tolerate  a relatively 


515 


516 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4.  December  1999 


large  number  of  shrubs  and  trees.  As  with 
most  polygynous  species,  the  female  generally 
tends  both  eggs  and  young  alone. 

In  spite  of  their  ability  to  use  secondary 
habitats  such  as  non-native  grasslands  or  road 
right-of-way,  Dickcissel  populations  have  de- 
clined by  about  39%  since  1966  (Peterjohn  et 
al.  1994).  Population  trends  in  Dickcissels  are 
difficult  to  estimate  because  their  abundance 
and  distribution  fluctuate  considerably  among 
years  (Fretwell  1986),  but  habitat  fragmenta- 
tion did  not  seem  to  be  responsible  for  their 
population  declines  (Herkert  et  al.  1993).  In- 
stead, poisoning  of  tens  of  thousands  of  birds 
on  the  wintering  grounds  has  been  suggested 
as  cause  for  its  population  decline  (Basili  and 
Temple  1995,  Basili  1997).  Recent  evidence 
indicates  that  habitat  fragmentation  on  the 
breeding  grounds  might  also  have  a negative 
impact  on  this  species  (Winter  1998,  Winter 
and  Faaborg  in  press). 

Henslow's  Sparrow. — Little  is  known 
about  the  nesting  behavior  of  this  inconspic- 
uous species  because  of  its  furtive  behavior 
and  its  tendency  to  spend  most  of  its  time  on 
the  ground  (Bent  1968).  Its  highest  densities 
occur  in  grasslands  with  tall,  dense  vegetation 
and  a well-developed  layer  of  litter  (Wiens 
1969,  Robins  1971,  Zimmerman  1988,  Her- 
kert 1994,  Mazur  1996,  Winter  1998).  Based 
on  the  few  existing  nesting  studies  (Hyde 
1939,  Bent  1968,  Robins  1971,  Schulenberg 
et  al.  1994,  Rohrbaugh  et  al.  in  press),  we 
know  that  this  monogamous  species  generally 
nests  close  to  the  ground  in  tall  dense  vege- 
tation, preferably  within  large  clumps  of  litter. 

With  the  destruction  of  tallgrass  prairie  and 
similar  grassland  habitats,  the  breeding  range 
of  Henslow’s  Sparrows  has  contracted  consid- 
erably during  the  last  30  years,  mainly  in  the 
northeastern,  eastern,  and  northwestern  parts 
of  its  range  (Pruitt  1996).  Although  Henslow’s 
Sparrows  also  nest  in  secondary  habitats  such 
as  hayfields  and  reclaimed  surface  mines  (see 
review  in  Swanson  1996),  it  has  shown  a con- 
sistent population  decline  (Peterjohn  et  al. 
1994,  Herkert  1997).  Analysis  of  Christmas 
Bird  Count  data  in  the  southeastern  United 
States  also  indicates  population  declines  on 
the  wintering  grounds  (Butcher  and  Lowe 
1990).  The  major  reason  for  the  large  popu- 
lation decline  of  Henslow’s  Spanows  has  been 
suggested  to  be  loss  and  fragmentation  of  hab- 


itat on  the  breeding  grounds  (Herkert  1994). 
However,  studies  in  Missouri,  Kansas,  and 
Ohio  indicated  that  Henslow’s  Sparrows  can 
occurr  in  even  small  fragments  (see  Winter 
1998),  and  since  1988  its  populations  have 
been  steadily  increasing  in  Illinois  (J.  R.  Her- 
kert, pers.  comm.). 

STUDY  SITES  AND  METHODS 

Study  area. — Between  1995  and  1997,  I studied  the 
nesting  biology  of  Dickcissels  and  Henslow’s  Spar- 
rows in  13  fragments  of  native  tallgrass  prairie  in 
southwestern  Missouri  (approx.  37°  30'  N,  93°  30'  W; 
Winter  1998).  Dominant  grasses  in  the  study  area  in- 
cluded big  bluestem  (Andropogon  gerardii),  little  blue- 
stem  (Schizocliyrium  scoparium),  and  Indian  grass 
{Sorghastrum  nutans).  Dominant  forbs  included  sun- 
flower (Heliantlius  spp.),  milkweed  (Asclepias  spp.), 
blazing  star  (Liatris  spicata),  and  sensitive  briar 
(Schrankia  nuttallii).  Prairies  were  owned  by  the  Mis- 
souri Department  of  Conservation,  the  Missouri  Prairie 
Foundation,  The  Nature  Conservancy,  and  the  Missou- 
ri Department  of  Natural  Resources  and  were  actively 
managed  by  prescribed  burning  and  haying  (see  Winter 
1998). 

Nest  searching  and  monitoring. — Throughout  each 
field  season  (early  May  to  end  of  July)  my  field  assis- 
tants and  I located  and  monitored  nests  of  all  grassland 
species  that  we  found,  but  focused  our  nest  searching 
efforts  on  finding  nests  of  Henslow's  Sparrows  and 
Dickcissels.  Nests  were  found  by  walking  across  the 
study  sites  and  adjacent  areas  of  similar  vegetation, 
while  paying  close  attention  to  behavior  and  vocali- 
zations of  nearby  adult  birds.  Most  nests  were  found 
by  observing  adults  (Dickcissels:  80%;  Henslow's 
Sparrows:  56%).  Behavioral  patterns  of  adults  that  we 
used  as  clues  that  nests  might  be  nearby  were  chipping, 
flying  short  distances  away  or  around  the  observer, 
flushing  close  to  the  observer  followed  by  a short 
flight,  and  canying  nest  material,  fecal  sacs,  or  food. 
The  location  of  a potential  nest  site  was  marked  with 
a short  length  of  flagging  tape  at  three  locations  within 
I m of  a potential  nest  site  forming  a triangle.  We  then 
retreated  10-30  m and  tried  to  locate  the  nest  when 
the  bird  returned.  Nests  also  were  located  by  flushing 
birds  while  randomly  walking  across  the  prairie  (Dick- 
cissel: 10%;  Henslow's  Sparrow:  30%).  The  remaining 
10%  and  14%  of  all  nests,  respectively,  were  found 
fortuitously;  flushing  birds  while  doing  other  research 
activities  such  as  vegetation  measurements  or  census- 
ing.  Because  rope-dragging  and  systematic  search 
were  ineffective  methods  for  nest  finding  in  1995,  1 
did  not  use  those  methods  in  the  following  years. 

We  did  not  search  for  or  monitor  nests  when  vege- 
tation was  wet  (after  rain  or  heavy  dew  immediately 
after  sunrise)  to  minimize  disturbance  of  vegetation 
surrounding  nests.  Each  nest  was  marked  with  a flag 
5 m to  the  north  and  a small  ribbon  was  placed  about 
30  cm  south  of  those  nests  that  were  hard  to  find.  Ev- 
ery 3-4  days  nest  fate  was  checked  by  walking  past 


Winter  - HENSLOW’S  SPARROW  AND  DICKCISSEL  NESTING 


517 


the  nest  to  avoid  creating  ‘‘dead  ends”  that  might  lead 
nest  predators  to  the  nest.  During  each  nest  check  we 
recorded  the  number  of  host  and  cowbird  eggs  and 
young,  presence  or  absence  of  adults,  and  the  state  of 
the  nest  if  the  nest  was  found  empty.  An  empty  nest 
was  considered  successful  if  one  or  more  of  the  fol- 
lowing cues  were  observed;  feces  in  the  nest,  feather 
sheaths  in  the  nest,  nest  rim  flattened,  adults  carrying 
food  or  chipping,  or  fledgling  close  to  nest. 

Nest  vegetation. — Nest  vegetation  was  characterized 
within  one  week  after  activity  at  a nest  had  ceased. 
Vegetation  was  measured  at  five  locations  around  the 
nest  site:  directly  at  the  nest  and  0.5  m from  the  nest 
in  each  cardinal  direction.  At  each  of  the  five  points  I 
measured  vegetation  cover  (Daubenmire  1959),  the 
number  of  woody  stems  within  each  Daubenmire 
frame,  vegetation  height,  litter  depth,  and  visual  ob- 
struction (Robel  et  al.  1970;  for  a more  detailed  de- 
scription see  Winter  1998).  For  each  nest  I calculated 
the  mean  for  each  of  the  five  measuring  points,  and 
used  the  mean  of  those  five  data  points  for  further 
analysis. 

Estimates  of  nesting  success. — When  calculating 
rates  of  nesting  success,  I excluded  nests  for  which  it 
was  not  possible  to  determine  if  predation  happened 
before  or  after  a nest  was  abandoned.  This  was  true 
for  nests  that  had  small  clutch  sizes  ( 1-2  eggs)  and 
were  depredated  the  next  time  the  nest  was  checked. 
Those  nests,  however,  were  included  for  estimating 
rates  of  cowbird  parasitism.  For  each  year  I estimated 
species  specific  probabilities  of  daily  nest  survival 
(Mayfield  1975)  separately  for  incubation  and  nestling 
stages,  and  for  the  total  nesting  period.  The  total  prob- 
ability of  nest  survival  was  defined  as  the  probability 
that  a nest  successfully  survived  incubation  and  nest- 
ling periods  and  fledged  at  least  one  young  of  the  pa- 
rental species.  In  the  two  species  that  I investigated, 
incubation  begins  with  the  laying  of  the  last  egg  (Bent 
1968).  I used  the  following  exponents  to  estimate  the 
probability  of  nesting  success  over  the  entire  nesting 
period;  Dickcissel:  21  days  (12  incubation  days  plus  9 
nestling  days),  and  Henslow’s  Sparrow:  20  days  (11 
incubation  days  plus  9 nestling  days;  Ehrlich  et  al. 
1988).  Standard  errors  for  daily  nest  survival  rates 
were  calculated  by  using  the  formula  for  binomial  dis- 
tributions (Zar  1996).  1 used  means  and  confidence  in- 
tervals (Johnson  1979)  to  compare  rates  of  nesting  suc- 
cess among  years  and  between  incubation  and  nestling 
stages  in  each  year.  To  allow  comparison  with  other 
studies  that  did  not  use  Mayfield  estimates,  1 also  pre- 
sent the  apparent  proportion  of  successful  nests. 

Statistical  analyses. — Logistic  regression  was  used 
to  investigate  if  nesting  success  was  related  to  the  date 
in  the  breeding  season.  I calculated  mean  clutch  size 
for  each  week  in  the  breeding  season  and  used  linear 
regression  analysis  to  investigate  if  clutch  size  varied 
during  the  nesting  season.  Because  the  number  of  nests 
found  varied  among  weeks,  I weighted  the  mean  week- 
ly clutch  size  by  its  standard  error.  For  this  analysis  I 
used  all  unparasitized  nests  of  the  three  years  of  the 
study. 


Vegetation  characteri.stics  of  depredated  and  suc- 
cessful Dickcissel  and  Henslow’s  Sparrow  nests  were 
compared  with  a two-tailed  /-test.  T’-values  were  com- 
pared to  the  T’-values  obtained  from  a sequential  Bon- 
ferroni  adjustment  (Rice  1989).  The  same  analysis  was 
used  to  compare  characteristics  between  Dickcissel 
and  Henslow’s  Sparrow  nests.  Dickcissel  was  the  only 
species  with  enough  parasitized  nests  to  allow  for  sta- 
tistical analysis.  For  this  species,  host  clutch  size,  num- 
ber of  host  fledglings,  and  nest  characteristics  of  par- 
asitized and  unparasitized  nests  were  compared  with 
/-tests.  Nesting  success  of  unparasitized  and  parasitized 
Dickcissel  nests  were  compared  by  using  means  and 
confidence  intervals  (Johnson  1979).  Logistic  regres- 
sion was  used  to  investigate  if  cowbird  parasitism  was 
related  to  the  date  in  the  breeding  season.  All  data 
were  analyzed  with  SAS  (SAS  6.03  for  PC;  SAS  In- 
stitute, Inc.  1988)  and  are  presented  as  means  and  stan- 
dard errors;  the  level  of  significance  was  set  at  0.05. 

RESULTS 

Nesting  biology. — Henslow’s  Sparrows  typ- 
ically arrived  in  the  study  area  in  early  May, 
about  1-2  weeks  earlier  than  most  Dickcis- 
sels,  and  stopped  nesting  by  the  end  of  July, 
also  about  1-2  weeks  earlier  than  most  Dick- 
cissels  (Fig.  1).  The  latest  observed  initiation 
of  incubation  in  Henslow’s  Sparrows  was  16 
July.  In  contrast  to  Henslow’s  Sparrows,  some 
Dickcissels  were  observed  carrying  nesting 
material  in  early  August.  Although  most  Dick- 
cissels seem  to  have  completed  their  nesting 
activity  by  the  end  of  July,  some  might  nest 
until  the  end  of  August.  The  peak  of  Dickcis- 
sel nest  initiation  did  not  occur  until  early 
June,  and  they  continued  to  nest  throughout 
June  and  early  July  (Fig.  lA).  In  contrast  to 
Dickcissels,  Henslow’s  Sparrows  had  two 
peaks  of  nest  initiation,  one  in  the  second  and 
third  week  of  May,  and  one  in  the  middle  of 
June  (Fig.  IB). 

Dickcissels  and  Henslow’s  Sparrows  had 
almost  identical  clutch  sizes,  hatching  and 
fledging  rates,  and  lengths  of  nestling  stages 
(Table  1).  Clutch  size  of  Dickcissels  and  Hen- 
slow’s Sparrows  tended  to  decline  with  date 
in  the  breeding  season  but  not  significantly 
(Dickcissel:  F = 4.8,  F = 0.35,  df  = 9,10,  P 
= 0.06,  slope  = —0.05  ± 0.02;  Henslow’s 
Sparrow:  F = 5.0,  F = 0.38,  df  = 8,9,  P = 
0.06,  slope  = -0.05  ± 0.02). 

Nesting  success. — The  main  cause  of  nest 
failure  was  nest  predation;  86%  of  all  failed 
Henslow’s  Sparrow  nests  and  84%  of  all 
failed  Dickcissel  nests  were  depredated  (Table 


518 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4.  December  1999 


a)  Dickcissel 


0.20  - 


1-7  8-14  15-21  22-28  29-4  5-11  12-18  19-25  26-2  3-9  10-16  17-23 


May  June  July 


b)  Henslow's  Sparrow 


FIG.  1.  Frequency  distribution  of  clutch  initiation  dates  in  (A)  Dickcissels  and  (B)  Henslow’s  SpaiTOWS  in 
southwestern  Missouri,  1995-1997. 


1 ).  Nest  failure  from  unknown  causes  (prob- 
ably weather)  or  nest  abandonment  were  min- 
imal (Table  1).  None  of  the  nests  of  either 
species  failed  as  a result  of  cowbird  parasit- 


ism. Dickcissel  nesting  success  was  lower 
during  incubation  than  during  the  nestling 
stage  in  1996,  whereas  nesting  success  during 
incubation  and  nestling  stages  did  not  differ 


Winter  • HENSLOW’S  SPARROW  AND  DICKCISSEL  NESTING 


519 


TABLE  1.  General  nesting  data  (.v  ± SE)  of  Dickcissel 
prairie  fragments,  1995-1997. 

and  Henslow’s  Sparrow 

in  southwestern  Missouri 

Dickcissel 
(II  = 242  f 

Henslow’s  Sparrow 
(n  = 59) 

General  nesting  data: 

Successful  nests  {n) 

112 

34 

Depredated  nests  (n) 

128 

25 

Unknown  loss  (n) 

2 

0 

Abandoned  nests  («)*’ 

a)  during  nest  building 

13 

0 

b)  with  eggs 

2 

1 

c)  parasitized 

3 

0 

Mowed  nests  (nY 

4 

3 

Nesting  success: 

Mayfield  nesting  success  (%)‘* 

29.7 

39.5 

Apparent  nesting  success^ 

46.3 

57.6 

Nesting  biology: 

Clutch  size  (n)^ 

3.9  ± 0.05  (227) 

3.8 

± 0.10  (56) 

Incubation  days  («)« 

11.45  ± 0.08  (11) 

12.0 

± 0 (1) 

Nestling  days  (n)*’ 

8.7  ± 0.02  (52) 

9.1 

± 0.08  (9) 

Hatching  success  («)' 

92.9  (69) 

93.2  (12) 

Young  fledged/nest 

1.7 

2.0 

Young  fledged/successful  nest 

3.6 

3.5 

Broad  parasitism: 

Parasitized  nests  (%>' 

8.8  ± 0.005  (21) 

5.3 

± 0.006  (3) 

Young  fledged  from  successful  unparasitized  nests 

3.7  ± 0.13  (105) 

3.6 

± 0.25  (33) 

Young  fledged  from  successful  parasitized  nests 

2.3  ± 0.26  (6) 

2.0 

± 0 (1) 

^ Total  number  of  nests  found  excluding  those  that  were  abandoned  and  mowed, 

^ Not  included  in  the  total  number  of  nests. 

Not  included  in  the  total  number  of  nests. 

After  Mayfield  (1975). 

' Percent  of  successful  nests  from  all  nests  found. 
fQnly  unparasitized  nests  were  used. 

8 n = number  of  nests  that  could  be  followed  from  nest  building  until  hatching. 

n = number  of  nests  that  could  be  followed  from  hatching  until  fledging. 

‘ Percent  hatched  eggs  from  all  eggs  for  which  the  clutch  size  was  known  with  certainty  (see  Methods). 
J Percent  of  parasitized  nests  out  of  all  nests  found. 


significantly  in  any  other  year  or  for  Hen- 
slow’s  Sparrows  (Table  2).  Nesting  success 
did  not  vary  significantly  with  the  date  in  the 
breeding  season  for  either  Dickcissels  (Wald- 
= 0.22,  P > 0.05,  n — 240)  or  Henslow’s 
Sparrows  (Wald-x^  = 2.66,  P > 0.05,  n = 59). 
Dickcissel  nesting  success  was  higher  in  1997 
than  in  1996,  whereas  nesting  success  of  Hen- 
slow’s Sparrows  did  not  vary  significantly 
among  years  (Table  2).  Mayfield  nesting  suc- 
cess tended  to  be  higher  in  Henslow’s  Spar- 
rows (40%)  than  in  Dickcissels  (30%;  Tables 
1,  2);  however,  the  95%  confidence  intervals 
for  the  estimates  of  nesting  success  in  these 
two  species  overlapped. 

Cowbird  parasitism. — The  rate  of  brood 
parasitism  by  Brown-headed  Cowbirds  was 


low,  but  slightly  higher  in  Dickcissels  (9.6%) 
than  in  Henslow’s  Sparrows  (5.3%;  Table  1). 
Dickcissel  nests  were  parasitized  throughout 
the  nesting  season  except  for  the  first  and  third 
week  of  May  and  the  last  week  of  July  (Fig. 
2). 

Parasitized  nests  generally  had  smaller 
clutches,  fewer  fledglings,  and  lower  nesting 
success  than  unparasitized  nests  (Table  3).  On 
average,  cowbirds  laid  1.4  eggs  per  parasit- 
ized Dickcissel  nest.  None  of  the  three  para- 
sitized Henslow’s  Sparrow  nests  had  more 
than  one  cowbird  egg.  Host  clutch  size  in  both 
species  was  reduced  by  about  0.9  eggs  per 
parasitized  nest  (Table  3).  The  reduction  in 
clutch  size  was  significant  in  Dickcissels  {t  = 
4.07,  df  = 23,  P < 0.001),  with  fewer  host 


520 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


TABLE  2. 
1995-1997. 

Nesting  success  of  Dickcissel  and  Henslow’s  Sparrow  in  southwestern  Missouri  prairie  fragments. 

Year 

Nest  cycle 
interval 

Depredated 

(«)•> 

Exposure 

days'^ 

Survival  ± SE‘* 

Success 

(%)' 

Dickcissel 

1995 

Incubation 

9 

4 

69.0 

0.94  ± 0.03 

21.6  < 48.8  < 100 

Nestling 

12 

4 

74.5 

0.95  ± 0.03 

35.0  < 60.8  < 100 

Totab 

18 

8 

143.5 

0.94  ± 0.02 

10.9  < 30.0  < 65.0 

1996 

Incubation 

82 

44 

511.5 

0.91  ± 0.01 

24.7  < 34.0  < 44.58 

Nestling 

61 

19 

388.5 

0.95  ± 0.01 

52.0  < 63.7  < 76.08 

Total 

113 

67 

900.0 

0.92  ± 0.01 

12.0  < 19.7  < 24.98 

1997 

Incubation 

90 

32 

517.0 

0.94  ± 0.01 

36.8  < 46.4  < 61.3 

Nestling 

98 

30 

618.0 

0.95  ± 0.01 

53.0  < 63.9  < 74.6 

Total 

117 

53 

1235.0 

0.96  ± 0.01 

32.6  < 39.8  < 55.18 

Alb 

Incubation 

181 

80 

1097.5 

0.93  ± 0.01 

34.0  < 40.3  <51.4 

Nestling 

171 

53 

1081.0 

0.95  ± 0.01 

46.4  < 63.6  < 70.6 

Total 

248 

128 

2278.5 

0.94  ± 0.00 

21.8  < 29.7  < 34.0 

Henslow’s  Sparrow 

1995 

Incubation 

6 

0 

34.5 

1.0  ± 0.00 

100.0 

Nestling 

7 

2 

59.0 

0.97  ± 0.02 

52.0  < 73.3  < 100 

Total 

7 

2 

83.5 

0.98  ± 0.02 

29.0  < 66.8  < 100 

1996 

Incubation 

1 1 

5 

103.5 

0.95  ± 0.02 

35.4  < 58.0  < 89.5 

Nestling 

16 

8 

146.0 

0.94  ± 0.02 

38.7  < 60.2  < 83.4 

Total 

21 

13 

249.5 

0.95  ± 0.01 

23.4  < 34.3  < 54.4 

1997 

Incubation 

14 

4 

84.5 

0.95  ± 0.02 

35.4  < 58.7  < 89.5 

Nestling 

26 

6 

133.0 

0.95  ± 0.02 

42.8  < 66.0  <91.0 

Total 

31 

10 

217.5 

0.95  ± 0.01 

23.4  < 39.0  < 54.0 

All 

Incubation 

31 

9 

222.5 

0.96  ± 0.01 

50.6  < 63.5  < 80.1 

Nestling 

49 

16 

334.0 

0.95  ± 0.01 

52.0  < 64.3  < 76.0 

Total 

59 

25 

550.5 

0.95  ± 0.01 

24.4  < 39.5  < 52.2 

“ Total  number  of  nests  monitored  in  a specific  nesting  interval  during  incubation  and  nestling  stages.  Because  nests  were  mostly  monitored  during  parts 
of  both  nesting  stages,  the  sum  of  nests  in  each  interval  is  higher  than  the  total  number  of  nests  found. 

Total  number  of  depredated  nests. 

Total  number  of  exposure  days  (Mayfield  1975). 

Probability  of  daily  Mayfield  nesting  succe.ss  (Day)  = - (#  depredated  nest.s/#  Mayfield  days)  + 1 SE  = sqrt  ((Day  » (#  depredated  nests/#  Mayfield 
days))/#  Mayfield  days). 

' Probability  of  nesting  succe.ss  over  the  entire  interval  = Day‘"'"''“'  shown  are  means  and  lower  and  upper  95%  confidence  intervals  (Johnson 

1979). 

f Both  nesting  stages  combined. 

? Intervals  with  the  same  letter  do  not  overlap. 

All  years  combined. 


young  fledged  from  successful  parasitized 
nests  (t  = 4.10,  df  = 34,  P < 0.001).  The 
reduction  in  nesting  success  was  not  caused 
by  competition  with  cowbird  young,  but  rather 
by  a higher  predation  rate  on  parasitized  nests; 
alt  successful  parasitized  nests  fledged  young 
of  both  host  and  cowbird. 

Nest  characteristics. — Compared  to  Hen- 
slow’s  Sparrows,  Dickcissels  chose  a variety 
of  nest  sites.  Most  (45%)  nests  were  placed  in 
forbs,  especially  leadplant  (Amorpha  canes- 
cens)  and  ashy  sunflower  {Helianthus  mollis), 
but  shrubs  (29%),  grass  (16%),  and  litter 
(10%)  also  were  used  as  nesting  substrates. 
Nests  were  typically  woven  in  the  stems  of 
forbs  or  woody  plants.  Because  nest  searches 


were  restricted  to  grassland  habitat,  nests  were 
not  found  within  shrubby  edge  habitats.  How- 
ever, many  Dickcissels  were  observed  breed- 
ing in  such  edge  habitats  (Winter  1998).  Suc- 
cessful nests  had  taller  vegetation,  greater  vi- 
sual obstruction  values,  greater  coverage  by 
grass,  and  smaller  areas  of  bare  soil  than  un- 
successful nests  (Table  4).  None  of  the  vege- 
tative characteristics  that  1 measured  at  Dick- 
cissel  nest  sites  differed  between  parasitized 
and  unparasitized  nests  (Table  4). 

Henslow’s  Sparrows  typically  placed  their 
nests  among  layers  of  thick  litter  (82%  of  all 
nests).  Compared  to  Dickcissel  nests,  the  veg- 
etation surrounding  Henslow’s  Sparrow  nests 
had  deeper  litter  (3.5  ± 0.27  vs  1.9  ± 0.23 


Winler  • HENSLOW’S  SPARROW  AND  DICKCISSEL  NESTING 


521 


May  June  July 


FIG.  2.  Frequency  distribution  of  rates  of  cowbird  parasitism  on  Dickcissel  nests  in  southwestern  Missouri, 
1995-1997. 


cm;  t = 3.3,  df  = 296,  P = 0.001),  lower 
vegetation  (42.0  ± 1.12  vs  46.1  ± 0.67  cm;  t 
= —2.8,  df  = 296,  P = 0.005),  greater  cover 
by  litter  (29.6  ± 2.0  vs  11.6  ± 0.71%;  t = 
10.4,  df  = 296,  P < 0.001),  and  less  cover  by 
forbs  (17.9  ± 1.03  vs  26.0  ± 0.83%;  t = 
—4.6,  df  = 296,  P < 0.001),  woody  plants 
(1.5  ± 0.12  vs  5.2  ± 0.62%;  t = -2.9,  df  = 
296,  P = 0.004),  and  soil  (0.4  ± 0.12  vs  4.8 
± 0.40%;  t = -5.5,  df  = 296,  P < 0.001). 
Henslow’s  Sparrow  nests  also  had  a higher 
percentage  of  nest  cover  (90.3  ± 2.68  vs  67.7 


± 1.79%;  t = 5.9,  df  = 296,  P < 0.001),  and 
were  located  closer  to  the  ground  (7.2  ± 0.53 
vs  18.4  ± 1.04  cm;  t = —5.4,  df  = 134,  P < 
0.001).  All  significant  P-values  remained  sig- 
nificant after  a sequential  Bonferroni  adjust- 
ment. Henslow’s  Sparrows  were  never  ob- 
served to  place  their  nest  within  or  in  imme- 
diate proximity  to  woody  vegetation.  In  con- 
trast to  Dickcissels,  Henslow’s  Sparrows  did 
not  weave  their  nests  into  the  surrounding 
vegetation,  but  placed  them  loosely  among  the 
surrounding  stems  of  grass  and  dead  vegeta- 


TABLE  3.  Clutch  size  (.v  ± SE)  and  nesting  success  of  unparasitized  and  parasitized  nests  in  Dickcissel  and 
Henslow’s  Sparrow  in  southwestern  Missouri,  1995-1997. 


Unparasitized  ne.sts 

Parasitized  nests 

Year 

n 

Clutch  size 

Exposure 

days 

Failed 

nests 

Success” 

(%) 

n 

Host  clutch 

Cowbird  clutch 

Exposure 

days 

Failed 

nests 

Succe.ss 

(%) 

Dickcissel 

1995 

17 

3.9 

-+* 

0.13 

143.5 

7 

35.0 

1 

3.0  ± 0 

2.0  ± 0 

12.0 

1 

16.1 

1996 

104 

3.9 

4- 

0.07 

931.0 

62 

23.5 

5 

3.4  ± 0.40 

2.0  ± 0.55 

26.0 

5 

1.1 

1997 

98 

3.9 

-+- 

0.07 

1014.5 

44 

39.4 

15 

3.0  ± 0.26 

1.2  ± 0.14 

152.0 

9 

27.7 

All 

220 

3.9 

H- 

0.05 

2089.0 

1 13 

31.2 

21 

3.0  ± 0.22 

1.4  ± 0.18 

190.0 

15 

17.8 

Henslow’s 

Sparrow 

1995 

6 

4.3 

-h 

0.21 

79.0 

2 

59.9 

1 

3.0  ± 0 

1.0  ± 0 

15.0 

0 

100 

1996 

21 

3.7 

-h 

0.17 

138.0 

13 

13.8 

0 

N/A 

0 

N/A 

N/A 

N/A 

1997 

29 

3.7 

-h 

0.15 

201.5 

8 

44.5 

2 

3.0  ± 0 

1.0  ± 0 

16.5 

2 

7.5 

All 

56 

3.8 

4- 

0.10 

418.5 

23 

32.3 

3 

3.0  ± 0 

1.0  ± 0 

31.5 

2 

26.9 

“Probability  that  a nest  survived  both  incubation  and  nesting  periods,  estimated  after  Mayfield  (1975). 


522 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


'u 

'a 

Urn 

a. 

— m — ootNoo 

a. 

ddddddddddd 

5 

0 

C/D 

C/5 

S 

fNr^l  fN(N(N(N  (N(N 

c 

u 

2 

c/5 

o 

00  — re)  oOinsOOOCN— 'fN 

-C 

*- 

d — ddd  — -^-^d-^d 

3 

1 III 

o 

c/5 

■y. 

c/5 

C/5 

a. 

— vori^  — 

•eb  — p On  ■'t  ^ Os  (^j  rn  VD 

ij 

z 

^ac 

doi4d— 'oireirN  — dcN 

"O 

0/ 

7.  II 

+1  +1  +1  +1  +1  +1  +1  +1  +1  +1  +1 

N 

2 

cz  ^ 

p p p 00  (N  ^ lo  p 

Q.  ^ 

— (ddddoOONU'rtrritC 

r3 

CS  — -eb  (N  — 

C3 

CL 

•o 

S 

i/^r^lONt^rNONOsrriTtt^m 

73 

— t^NOOt^OOOONO-'tON  — 

•o 

a> 

.3  ^ 

ddddddddd  — 

N 

ri 

+1  +1  +1  +1  +1  +1  +1  +1  +1  +1  +1 

b 11 

X 

a. 

00  00  00  p p p oj  p b'  (N  (p 

u 

5 

— \or~-d  — ^re)d'rt4Ndod 

ri-  rN  — (^1  no  — 

O. 

o F 

~o  S 

— 00  Oi  ON 

C X 

— OOOO^OON  — Oro  — 

TZ  D 

— OO  — ooO'rt-u-iOrNp 

X ^ 

a,. 

ddddddddddd 

w a 

X 

V V 

V -p 

= o 

? fc 

r3  ''P 

re)0N>/-)rr)ini/^>/^O’rrt''00 

F 

— OOnO  — — — — 'P  — On 

r-i  n — M ri  ri  oi  (N  — (N 

^ CQ 

D. 

1)  a 

u 

"O  3 

e3  f3  n rt  a 

C '**- 

*- 

— ^rnr-i— -dreiddr-i  — d 

3 c 

^ C3 

■yi 

•j'. 

III  1 II 

X (J 

U 

(U  ^ 

H 'c 

O 

u 

3 

3 .SJJ 
X 'x 

y 

iJ 

ONce,  re,  — onOnOnonO^On 

z 

— .O  — — ^OnO  — ^ONNDi/^'e)- 

c 73 

•—  u 

73 

o — — dd— ■— -ddrN  — 

^ E 

3 O 
73  — 

L" 

+1  +1  +1  +1  +1  4-1  4-1  +\  4-1  4-1  4-1 
>/',^0'ND'e)-r''  — — poop 

+1 

Q — 

— ^et->6d— •dUiEiri4U 

X 

^ OJ 

X 3 

Tj-  ri  — in  (N  no  — 

X ^ 

m m o 00  oi  no  NO  00  m On  O' 

T a. 

oi  00  00  O On  rN|  m p p p 

E 

II 

d d d d d — — d d ONi  — 

O c 
tZ  rt 

y,  — 

U 

4-1  4-1  4-1  4-1  -H  -H  4-1  -H  4-1  4-1  4-1 

u 3 
i: 

U r- 

a II 

3 - 
X 3. 

ooNNOitoiopppoip 
oiodoid— ■NOin'rfmONOo 

Urn 

OXJ 
X • — 

X 

-rf  ri  — m O'!  NO  — 

C 

"S  <! 

X 

y— s 

!-■ 

E S 

d 3 

n 1 

^ r: 

1— 1 iTi 

. O' 
^ — 

UJ  X 
J c 
CC  y 

Variable 

3'Ci 

r«px 

r“  U-  <— * 

•<  E 

H 

Litt 

Veg 

Visi 

No. 

Litt 

Gra 

For 

Wo 

Soi 

Nc! 

Ne: 

•I) 

O 

c 

rs 


tion.  In  late  June  and  throughout  July,  18%  of 
all  nests  were  found  in  areas  that  had  been 
burned  the  same  spring  and  therefore  lacked 
any  litter.  In  these  areas,  Henslow’s  Sparrows 
placed  their  nests  within  large  clumps  of  grass 
(mostly  big  bluestem  and  Indian  grass)  close 
to  the  ground.  Successful  and  depredated  Hen- 
slow’s  Sparrow  nests  did  not  differ  in  any  nest 
characteristic  (Table  5). 

DISCUSSION 

In  southwestern  Missouri,  Dickcissels  and 
Henslow’s  Sparrows  had  nearly  identical 
clutch  size,  hatching  success,  length  of  incu- 
bation and  nestling  stages,  and  number  of 
young  fledged  per  successful  nest.  These  var- 
iables were  similar  to  previous  reports  for 
Dickcissels  (Bent  1968,  Harmeson  1974,  Zim- 
merman 1982,  Fretwell  1986,  Patterson  and 
Best  1996).  Fewer  studies  have  monitored 
Henslow’s  Sparrow  nests,  because  their  nests 
are  difficult  to  locate  (Bent  1968;  Robins 
1971;  Schulenberg  et  al.  1994;  D.  Reinking, 
pers.  comm.).  Clutch  size  of  Dickcissels  and 
Henslow’s  Sparrows  tended  to  decrease  with 
date  in  the  breeding  season  in  southwestern 
Missouri.  For  Dickcissels,  Harmeson  (1974) 
described  a peak  in  clutch  size  in  the  middle 
of  the  nesting  season,  whereas  changes  of 
Henslow’s  Sparrow  clutch  size  over  time  had 
not  yet  been  described. 

Although  Dickcissels  and  Henslow’s  Spar- 
rows had  almost  identical  nesting  variables, 
their  nesting  phenologies  seemed  to  differ. 
Generally,  nesting  success  is  relatively  low  in 
most  grassland  nesting  birds,  varying  from 
25-50%  (Wiens  1969,  Vickery  et  al.  1992, 
Martin  1995).  Grassland  birds  often  compen- 
sate for  low  nesting  success  by  several  re- 
nesting attempts  throughout  the  breeding  sea- 
son. Consequently,  most  grassland  nesting 
species  raise  an  average  of  1.5-2  broods  per 
female  per  year  (Wiens  1969,  Martin  1995). 
The  single  nesting  peak  of  Dickcissels  in  my 
study  seemed  to  indicate  that  Dickcissels 
raised  only  one  brood  in  my  study  area,  as  has 
been  described  by  Zimmerman  (1982,  1984). 
However,  E.  Bollinger  (pers.  comm.)  observed 
a second  brood  in  one  color-banded  Dickcissel 
female,  indicating  that  Dickcissels  can  be  dou- 
ble-brooded. Because  Dickcissels  appear  to 
frequently  move  within  one  breeding  season 
(Fretwell  1986),  possibly  because  of  displace- 


Winler  • HENSLOW’S  SPARROW  AND  DICKCISSEL  NESTING 


523 


TABLE  5.  Henslow’s  Sparrow  nest  characteristics  (.f  ± SE)  at 
western  Missouri  prairie  fragments,  1993-1997. 

successful  and  depredated  ne.sts 

in  south- 

Variable 

Successful 
(n  = 35) 

Depredated 
(«  = 25) 

t 

df 

H 

Litter  depth  (cm) 

3.3  ± 0.37 

3.7  ± 0.41 

0.72 

58 

OAl 

Vegetation  height  (cm) 

43.0  ± 1.48 

40.7  ±1.71 

- 1 .00 

58 

0.32 

Visual  obstruction  (dm) 

25.4  ± 1.89 

24.9  ± 2.09 

-0.20 

57.7 

0.84 

No.  woody  stems 

0.43  ± 0.14 

0.26  ±0.11 

-0.94 

57.7 

0.35 

Litter  cover  (%) 

27.1  ± 2.70 

33.1  ± 2.85 

1.50 

58 

0.14 

Grass  cover  (%) 

51.4  ± 2.50 

49.4  ± 2.43 

-0.57 

58 

0.57 

Eorb  cover  (%) 

19.1  ± 1.35 

16.2  ± 1.58 

-1.42 

58 

0.16 

Woody  cover  (%) 

1.82  ± 0.68 

1.14  ± 0.45 

-0.98 

55.2 

0.40 

Soil  cover  (%) 

0.55  ± 0.12 

0.22  ±0.11 

-1.47 

50.4 

0.15 

Nest  cover  (%) 

89.6  ± 4.02 

91.7  ± 3.03 

0.41 

57.7 

0.68 

Nest  height  (cm) 

7.0  ± 0.63 

7.9  ± 1.02 

0.74 

26.8 

0.47 

ment  from  hayfields  after  mowing  (Igl  1991, 
Frawley  and  Best  1991),  females  might  renest 
or  raise  a second  brood  in  another  area.  None 
of  the  Dickcissels  in  my  study  area  were  col- 
or-banded; therefore,  I could  not  determine  if 
late  nesting  females  (as  also  described  by  Har- 
meson  1974)  had  arrived  from  other  areas,  or 
if  they  had  started  a second  brood  or  a re- 
nesting attempt  in  the  same  area. 

Henslow’s  Sparrows  seemed  to  be  more 
likely  to  be  double-brooded  in  southwestern 
Missouri  than  Dickcissels  because  they  clearly 
exhibited  two  peaks  of  nest  initiation.  How- 
ever, as  with  Dickcissels,  individual  birds 
were  not  color-banded,  making  it  impossible 
to  determine  if  the  second  nesting  peak  was 
caused  by  females  on  their  second  brood,  by 
renesting  attempts,  by  newly  arriving  females, 
or  if  it  was  an  artifact  of  small  sample  size. 
This  lack  of  adequate  information  is  also  true 
for  all  other  studies  that  describe  this  species 
as  double-brooded  (Hyde  1939,  Bent  1968, 
Robins  1971). 

Nest  predation  was  the  main  reason  for  nest 
failure,  as  has  been  described  for  many  other 
bird  species  (Martin  1993,  Patterson  and  Best 
1996).  Mean  Mayfield  nesting  success  of 
Dickcissels  was  similar  to  that  reported  from 
Kansas  (Zimmerman  1984)  and  Missouri 
Crop  Reserve  Program  fields  (McCoy  1996), 
but  lower  rates  of  nesting  success  were  re- 
ported from  Iowa  (Bryan  and  Best  1994,  Pat- 
terson and  Best  1996),  Kansas  (Hill  1976), 
and  Oklahoma  (Rohrbaugh  et  al.  in  press). 
Robins  (1971)  reported  that  6 of  1 1 Henslow’s 
Sparrow  nests  found  in  Michigan  successfully 
fledged  young.  This  apparent  success  rate 


(54.5%)  is  comparable  to  the  apparent  success 
rate  in  my  study  (57.6%).  However,  the  num- 
ber of  young  fledged  per  nest  in  Michigan 
(0.37)  and  the  number  of  young  fledged  per 
successful  nest  (2.8)  were  lower  than  in  Mis- 
souri. In  Oklahoma,  40.9%  of  22  Henslow’s 
Sparrow  nests  were  successful  (D.  Reinking, 
pers.  comm.),  which  was  about  17%  lower 
than  in  Missouri.  The  number  of  young 
fledged  per  unparasitized  Henslow’s  Sparrow 
nest  was  also  slightly  lower  in  Oklahoma  than 
in  Missouri  (3.3  vs  3.6  young  fledged  per  nest; 
Reinking,  pers.  comm.).  Southwestern  Mis- 
souri thus  seems  to  be  a relatively  productive 
breeding  area  for  Henslow’s  Sparrows. 

Daily  Mayfield  nesting  success  in  Dickcis- 
sels was  lower  during  incubation  than  during 
nestling  stages  in  1996,  and  tended  to  be  low- 
er in  1995  and  1997.  Higher  nesting  success 
during  the  nestling  stage  was  also  reported  by 
Bryan  and  Best  (1994)  and  by  Harmeson 
(1974),  and  generally  is  the  most  frequently 
observed  pattern  of  nest  survival  (Nice  1957; 
but  see  Patterson  and  Best  1996).  Nesting  suc- 
cess could  be  lower  during  incubation  because 
poorly  concealed  nests  are  the  first  to  be  found 
by  nest  predators,  or  because  visually  hunting 
nest  predators  find  nests  with  eggs  more  eas- 
ily. Shorter  and  sparser  vegetation  at  depre- 
dated Dickcissel  nests  indicated  that  these 
nests  were  in  fact  less  well  concealed  than 
successful  nests.  High  incidence  of  nest  pre- 
dation by  mammals  (see  Winter  1998),  which 
hunt  based  on  visual  and  olfactory  cues,  might 
explain  the  tendency  for  slightly  lower  nesting 
success  during  incubation  in  southwestern 
Missouri.  However,  the  only  nest  predators 


524 


THE  WILSON  BULLETIN  • Vo/.  Ill,  No.  4.  December  1999 


that  I observed  at  Dickcissel  nests  were  two 
snakes,  one  eastern  yellowbellied  racer  {Col- 
uber constrictor  flaviventris)  and  one  prairie 
kingsnake  {Lximpropeltis  calligaster  calligas- 
ter).  Because  rates  of  nesting  success  in  Hen- 
slow’s  Sparrows  were  nearly  identical  during 
incubation  and  nestling  stages,  and  because 
their  nests  were  extremely  well  concealed,  it 
seems  that  visually  hunting  nest  predators 
rarely  destroy  its  nests.  Instead  of  visually 
hunting  predators,  snakes  are  possibly  the 
main  nest  predators  of  Henslow’s  Sparrow 
nests.  This  could  also  be  the  reason  why  nest 
vegetation  did  not  differ  between  successful 
and  depredated  nests. 

Rates  of  brood  parasitism  by  Brown-headed 
Cowbirds  in  southwestern  Missouri  were  rel- 
atively low  compared  to  parasitism  rates  de- 
scribed in  other  studies  on  grassland-nesting 
birds  (Hergenrader  1962;  Zimmerman  1966, 
1983;  Hill  1976;  Elliott  1978;  Patterson  and 
Best  1996;  Koford  et  al.  in  press).  Because 
cowbirds  did  not  cause  direct  mortality  to  any 
Dickcissel  young  in  my  study,  brood  parasit- 
ism by  itself  did  not  directly  decrease  nesting 
success.  However,  the  reduction  in  clutch  size 
decreased  the  number  of  host  fledglings  by 
about  one  young  per  parasitized  nest,  as  also 
was  reported  by  Hill  (1976). 

In  Henslow’s  Sparrows,  brood  parasitism 
by  Brown-headed  Cowbirds  was  slightly  low- 
er than  in  Dickcissels;  only  5%  of  all  Hen- 
slow’s Sparrow  nests  were  parasitized  in 
southwestern  Missouri.  The  only  other  records 
of  parasitized  Henslow’s  Sparrow  nests  are 
from  Oklahoma  (Reinking,  pers.  comm.)  and 
Kansas  (Schulenberg  et  al.  1994).  In 
Oklahoma,  2 out  of  22  Henslow’s  Sparrow 
nests  were  parasitized.  Only  1 of  the  parasit- 
ized nests  successfully  fledged  both  host  and 
cowbird  young,  whereas  the  other  nest  was 
depredated.  The  one  Henslow’s  Sparrow  nest 
that  was  found  by  Schulenberg  and  coauthors 
(1994)  in  Kansas  contained  two  cowbird  eggs 
and  was  abandoned  during  incubation.  Cow- 
bird parasitism  is  probably  low  in  Henslow’s 
Sparrows  because  their  nests  are  well  con- 
cealed. Low  parasitism  rates  in  Henslow’s 
Sparrows  were  previously  noted  by  Bent 
(1968:786),  who  mentioned  that  this  species 
“appears  to  escape  heavy  parasitism,  possibly 
because  the  nests  are  so  well  hidden.’’ 

Nest  placement  differed  significantly  be- 


tween Dickcissels  and  Henslow’s  Sparrows. 
Dickcissels  chose  a variety  of  nesting  habitats 
(Bent  1968);  they  preferred  forbs  and  shrubs, 
and  did  not  avoid  edge  habitats.  Henslow’s 
Sparrow  nests,  on  the  other  hand,  were  never 
found  in  either  of  the  nest  substrates  preferred 
by  Dickcissels  or  within  shrubby  edge  habitat. 
Instead,  this  species  built  its  nest  lower  to  the 
ground,  mainly  within  large  clumps  of  litter 
where  it  was  almost  100%  covered  by  vege- 
tation (Hyde  1939,  Robin  1971,  Schulenberg 
et  al.  1994).  Several  researchers  that  described 
the  relationship  between  Henslow’s  Sparrow 
breeding  densities  and  vegetation  parameters 
also  noted  the  species’  preference  for  tall 
grass  and  litter  cover  (Wiens  1969,  Skinner  et 
al.  1984,  Herkert  1994,  Swanson  1996).  1 
found  that  Henslow’s  Sparrows  were  able  to 
build  nests  in  recently  burned  areas  that 
lacked  litter,  as  did  Zimmerman  (pers.  comm, 
in  Schulenberg  et  al.  1994)  in  Kansas.  Be- 
cause Dickcissel  nests  were  more  conspicuous 
than  Henslow’s  Sparrow  nests,  they  were 
probably  more  easily  detected  by  visually 
hunting  nest  predators  and  Brown-headed 
Cowbirds,  resulting  in  slightly  higher  rates  of 
nest  predation  and  nest  parasitism  in  this  spe- 
cies. 

General  nesting  data  indicated  that  Dickcis- 
sels tended  to  be  less  productive  in  south- 
western Missouri  than  Henslow’s  Sparrows. 
These  findings  are  in  contrast  to  the  general 
notion  that  Dickcissels  are  of  little  conserva- 
tion concern  on  the  breeding  grounds  (Herkert 
et  al.  1993,  Swanson  1996).  The  discrepancy 
may  be  because  previous  reports  on  Dickcis- 
sels and  Henslow’s  Sparrows  were  based  only 
on  breeding  density  estimates.  This  study 
showed  that  basic  data  on  the  general  nesting 
ecology  of  a species  are  necessary  for  a better 
understanding  of  the  factors  that  might  influ- 
ence a species  in  a given  area. 

ACKNOWLEDGMENTS 

I thank  J.  Faaborg  for  hi.s  support  throughout  my 
study,  and  J.  Faaborg,  E.  Fink,  W.  Hochachkil,  L.  Igl, 
D.  Johnson,  R.  Koford,  M.  Ryan,  and  R Vickery  for 
many  comments  that  greatly  improved  the  manuscript. 
I also  thank  my  field  assi.stants  J.  Bernier,  M.  Harry, 
P.  Kohn,  S.  Panken,  D.  Rutka,  and  K.  Warren.  This 
research  was  supported  by  the  U.S.  Geological  Survey 
(Northern  Prairie  Wildlife  Research  Center),  the  Mis- 
souri Department  of  Conservation,  the  U.S.  Fish  and 


Winter  • HENSLOW’S  SPARROW  AND  DICKCISSEL  NESTING 


525 


Wildlife  Service,  the  Missouri  Prairie  Foundation,  and 

Sigma  Xi. 

LITERATURE  CITED 

Basili,  G.  D.  1997.  Continental-scale  ecology  and  con- 
servation of  Dickcissels.  Ph.D.  diss.,  Univ.  of 
Wisconsin,  Madison. 

Basili,  G.  and  S.  A.  Temple.  1995.  A perilous  migra- 
tion. Nat.  Hist.  104:40-46. 

Bent,  A.  C.  1968.  Life  histories  of  North  American 
cardinals,  grosbeaks,  buntings,  towhees,  finches, 
sparrows,  and  allies.  Bull.  U.S.  Natl.  Mus.  237:1- 
1889. 

Bryan,  G.  G.  and  L.  B.  Best.  1994.  Avian  nest  den- 
sity and  success  in  grassed  waterways  in  Iowa 
rowcrop  fields.  Wildl.  Soc.  Bull.  22:583-592. 

Butcher,  G.  and  Lowe.  1990.  Population  trends  of 
twenty  species  of  migratory  birds  as  revealed  by 
Christmas  Bird  Counts,  1963-87.  Final  report. 
Cooperative  Agreement  No.  14-16-0009-88-941, 
U.S.  Fish  and  Wildl.  Serv.  Office  of  Migratory 
Bird  Manage.,  Washington,  D.C. 

Daubenmire,  R.  1959.  A canopy-coverage  method  of 
vegetational  analysis.  Northwest  Sci.  33:43—64. 

Ehrlich,  P.  R.,  D.  S.  Dobkin,  and  D.  Wheye.  1988. 
The  birder’s  handbook.  Simon  and  Schuster  Inc., 
New  York. 

Elliott,  P.  E 1978.  Cowbird  parasitism  in  the  Kansas 
tallgrass  prairie.  Auk  95:161-167. 

Frawley,  B.  J.  and  L.  B.  Best.  1991.  Effects  of  mow- 
ing on  breeding  bird  abundance  and  species  com- 
position in  alfalfa  fields.  Wildl.  Soc.  Bull.  19:135- 
142. 

Fretwell,  S.  1986.  Distribution  and  abundance  of  the 
Dickcissel.  Current  Ornithol.  4:211-242. 

Harmeson,  J.  P.  1974.  Breeding  ecology  of  the  Dick- 
cissel. Auk  91:348-359. 

Hergenrader,  G.  L.  1962.  The  incidence  of  nest  par- 
asitism by  the  Brown-headed  Cowbird  (Molothrus 
ater)  on  roadside  nesting  birds  in  Nebraska.  Auk 
79:85-88. 

Herkert,  j.  R.  1994.  Status  and  habitat  selection  of 
the  Henslow’s  Sparrow  in  Illinois.  Wilson  Bull. 
106:35-45. 

Herkert,  J.  R.  1997.  Population  trends  of  the  Hen- 
slow’s Sparrow  in  relation  to  the  Con.servation  Re- 
serve Program  in  Illinois,  1975-1995.  J.  Field  Or- 
nithol. 68:235-244. 

Herkert,  J.  R.,  R.  S.  Szafone,  V.  M.  Kleen,  and  J. 
E.  ScHWEGMAN.  1993.  Habitat  establishment,  en- 
hancement, and  management  for  forest  and  grass- 
land birds  in  Illinois.  Division  of  Natural  Heritage, 
Illinois  Department  of  Conservation,  Springfield. 

Hill,  R.  A.  1976.  Host-parasite  relation.ships  of  the 
Brown-headed  Cowbird  in  a prairie  habitat  of 
west-central  Kansas.  Wilson  Bull.  33:555-565. 

Hyde,  A.  S.  1939.  The  life  history  of  Henslow’s  Spar- 
row, Passerherbulus  henslowii  (Audubon).  Univ. 
Mich.  Misc.  Publ.  41:1-72. 

Igl,  L.  D.  1991.  The  role  of  climate  and  mowing  on 
Dickcissel  (Spiza  americana)  movements,  distri- 


bution, and  abundance.  M..Sc.  thesis,  Iowa  State 
Univ.,  Ames. 

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

Koford  R.  R.,  B.  S.  Bowen,  J.  T.  Lokemoen,  and  A. 
D.  Kruse.  In  press.  Cowbird  parasitism  in  grass- 
land and  cropland  in  North  Dakota.  In  The  ecol- 
ogy and  management  of  cowbirds  (J.  N.  M.  Smith, 

T.  Cook,  S.  K.  Robinson,  S.  I.  Rothstein,  and  S. 
G.  Sealy,  Eds.).  University  of  Texas  Press,  Austin. 

Martin,  T.  E.  1993.  Nest  predation  among  vegetation 
layers  and  habitat  types:  revising  the  dogmas.  Am. 
Nat.  141:897-913. 

Martin,  T.  E.  1995.  Avian  life  history  evolution  in 
relation  to  nest  sites,  nest  predation,  and  food. 
Ecol.  Monogr.  65:101-127. 

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

Mazur,  R.  1996.  Implications  of  field  management  for 
Henslow’s  Sparrow  habitat  at  Saratoga  National 
Historical  Park,  New  York.  M.Sc.  thesis.  State 
Univ.  of  New  York,  Syracuse. 

McCoy,  T.  D.  1996.  Avian  abundance,  composition, 
and  reproductive  success  on  Conservation  Re- 
serve Program  fields  in  northern  Missouri.  M.Sc. 
thesis,  Univ.  of  Missouri,  Columbia. 

Nice,  M.  M.  1957.  Nest  success  in  altricial  birds.  Auk 
74:305-321. 

Patterson,  M.  P.  and  L.  B.  Best.  1996.  Bird  abun- 
dance and  nest  success  in  Iowa  CRP  fields:  the 
importance  of  vegetation  structure  and  composi- 
tion. Am.  Midi.  Nat.  135:153-167. 

Peterjohn,  B.  G.,  j.  R.  Sauer,  and  W.  A.  Link.  1994. 
The  1992  and  1993  summary  of  the  North  Amer- 
ican breeding  bird  survey.  Bird  Populations  2:46- 
61. 

Pruitt,  L.  1996.  Henslow's  Sparrow  status  assessment. 

U. S.  Fish  and  Wildlife  Service,  Minneapolis.  Min- 
nesota. 

Rice,  W.  R.  1989.  Analyzing  tables  of  statistical  tests. 
Evolution  43:223—225. 

Robel,  R.  j.,  j.  N.  Briggs,  A.  D.  Dayton,  and  L.  C. 
Hulbert.  1970.  Relationship  between  visual  ob- 
struction measurements  and  weight  of  grassland 
vegetation.  J.  Range  Manage.  23:295-297. 

Robins,  J.  D.  1971.  A study  on  Henslow's  SpaiTows 
in  Michigan.  Wilson  Bull.  83:39-48. 

Rohrbaugh,  R.  W.,  D.  L.  Reinking,  D.  H.  Wolfe,  S. 
K.  Sherrod,  and  M.  A.  Jenkins.  In  pre.ss.  Effects 
of  prescribed  burning  and  grazing  on  nesting  and 
reproductive  success  of  three  grassland  passerine 
species  in  tallgrass  prairie.  Stud.  Avian  Biol. 

SAS  Institute,  Inc.  1988.  SAS/STAT  user's  guide,  re- 
lease 6.03  ed.  SAS  Inst.,  Cary,  North  Carolina. 

ScHULENBERG,  J.  H.,  G.  L.  Horak,  M.  D.  Schwilling, 
and  E.  j.  Finck.  1994.  Nesting  of  Henslow's  Spar- 
rows in  Osage  County,  Kansas.  Kans.  Ornith.  Soc. 
Bull.  44:1-4. 

Skinner,  R.  M.,  T.  M.  Baskett,  and  M.  D.  Blenden. 


526 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


1984.  Bird  habitat  on  Missouri  prairies.  Missouri 
Department  of  Conservation,  Jefferson  City. 
Swanson,  D.  A.  1996.  Nesting  ecology  and  nesting 
habitat  requirements  of  Ohio's  grassland-nesting 
birds;  a literature  review.  Ohio  Fish  Wildl.  Serv. 
Rep.  13:1-60. 

Vickery,  P.  D.,  M.  L.  Hunter,  and  J.  V.  Wells.  1992. 
Evidence  of  incidental  nest  predation  and  its  effect 
on  nests  of  threatened  grassland  birds.  Oikos  63; 
281-288. 

Wiens,  J.  A.  1969.  An  approach  to  the  study  of  eco- 
logical relationships  among  grassland  birds.  Or- 
nithol.  Monogr.  8;  1-93. 

Winter,  M.  1998.  Effects  of  habitat  fragmentation  on 
grassland-nesting  birds  in  southwestern  Missouri. 
Ph.D.  diss.,  Univ.  of  Missouri,  Columbia. 
Winter,  M.  and  J.  Faaborg.  In  press.  Patterns  of  area- 


sensitivity  in  grassland-nesting  birds.  Conserv. 
Biol. 

Zar,  j.  H.  1996.  Biostatistical  analysis,  third  ed.  Pren- 
tice-Hall, Englewood  Cliffs,  New  Jersey. 

Zimmerman,  J.  L.  1966.  Polygyny  in  the  Dickcissel. 
Auk  83:534-546. 

Zimmerman,  J.  L.  1982.  Nesting  success  of  Dickcissels 
{Spiza  americana)  in  preferred  and  less  preferred 
habitats.  Auk  99:292-298. 

Zimmerman,  J.  L.  1983.  Cowbird  parasitism  of  Dick- 
cissels in  different  habitats  and  at  different  nest 
densities.  Wilson  Bull.  95:7—22. 

Zimmerman,  J.  L.  1984.  Nest  predation  and  its  rela- 
tionship to  habitat  and  nest  density  in  Dickcissels. 
Condor  86:68-72. 

Zimmerman,  J.  L.  1988.  Breeding  season  habitat  se- 
lection by  the  Henslow’s  Sparrow  (Ammodramus 
henslowii)  in  Kansas.  Wilson  Bull.  100:17—24. 


Wilson  Bull.,  1 1 1(4),  1999,  pp.  527-535 


RESPONSE  OF  A BIRD  ASSEMBLAGE  IN  SEMIARID  CHILE  TO 

THE  1997-1998  EL  NINO 

FABIAN  M.  JAKSIC  2 AND  IVAN  LAZO' 


ABSTRACT. — The  semiarid  region  of  Chile  is  influenced  by  El  Nino  Southern  Oscillation.  Its  absence  causes 
droughts  and  its  presence  causes  wet  years,  which  in  turn  result  in  variations  in  resource  levels  for  avian 
assemblages.  We  show  that  bird  species  richness  and  density  follow  some  of  these  pulses  closely.  Sixty-one  bird 
species,  32  of  which  were  Passeriformes,  were  sighted  during  five  years  in  Las  Chinchillas  National  Reserve 
(300  km  N of  Santiago).  Overall,  30  species  (49%)  were  residents  and  31  (51%)  were  migratory.  The  most 
speciose  trophic  groups  were  insectivores  (34%),  carnivores  (28%),  and  granivores  (25%).  Bird  species  richness 
and  density  declined  from  43  species  and  45-50  individuals/ha  in  spring  1993,  to  29  species  and  15-20  indi- 
viduals/ha in  autumn  1996.  Increases  were  observed  with  the  onset  of  El  Nino,  reaching  totals  of  42  species  (a 
45%  increase  from  29)  and  densities  of  55—60  bird.s/ha  in  summer  1997.  Similar  trends  were  observed  in  one 
of  two  major  food  resources  measured:  small  mammals.  Positive  correlations  were  found  between  raptor  species 
richness  and  density  and  small  mammal  density,  but  not  between  insectivorous  bird  species  richness  or  density 
and  terrestrial  arthropod  abundance.  Because  the  climate  was  very  dry  during  most  of  the  time  of  our  study,  we 
may  have  witnessed  the  lowest  boundary  for  species  richness  and  bird  density.  Whether  the  1997-1998  El  Nino 
brought  the  maximum  bird  species  richness  and  density  for  the  site  is  yet  to  be  seen.  Received  20  Jan.  1999, 
accepted  2 June  1999. 


Bird  assemblages  vary  through  time  in  spe- 
cies composition  and  absolute  and  relative 
abundances,  in  the  short  term  (seasonal,  Av- 
ery and  van  Riper  1989),  medium  term  (be- 
tween years,  Wiens  1990a,  b),  or  both.  Most 
authors  agree  that  this  variation  reflects 
changes  in  the  resource  base,  mainly  food 
(Feinsinger  et  al.  1985,  Wiens  1993).  Three 
factors  are  generally  proposed  to  account  for 
how  birds  use  resources  (Pearson  1991):  com- 
petition (Martin  1987,  Pulliam  and  Dunning 
1987),  predation  (Lima  1987),  and  physical  or 
abiotic  stresses  (Karr  and  Freemark  1983). 
The  degree  to  which  these  factors  determine 
bird  assemblages  is  controversial  (Loiselle 
and  Blake  1991).  Although  they  are  not  mu- 
tually exclusive,  these  three  factors  are  usually 
assessed  individually,  rarely  two  simulta- 
neously (Martin  1985,  Kotler  and  Holt  1989). 
Most  attention  has  concentrated  on  the  role  of 
species  interactions  rather  than  on  physical 
factors  (Pearson  1991),  although  arguably  the 
latter  set  the  stage  for  biotic  interactions. 

The  periodic  intrusion  of  El  Nino  Southern 
Oscillation  along  the  western  coasts  of  the 
Americas  constitutes  a major  physical  distur- 
bance that  brings  warm  water  to  the  shores 


‘ Dept,  de  Ecologi'a,  Pontificia  Univ.  Catolica  de 
Chile,  Casilla  1 14-D,  Santiago,  Chile; 

E-mail:  fjak.sic @ genes. bio. puc. cl 
^ Corresponding  author. 


and  increased  precipitation  to  the  adjacent 
land  masses.  Although  initially  studied  in  its 
oceanographic  and  climatic  aspects,  increased 
awareness  of  the  multiple  effects  of  El  Nino 
is  shifting  the  focus  to  the  effects  of  this  phe- 
nomenon on  birds  (Barber  and  Chavez  1983, 
Schreiber  and  Schreiber  1984,  Gibbs  and 
Grant  1987,  Grant  and  Grant  1987,  Hall  et  al. 
1988,  Miskelly  1990,  Massey  et  al.  1992, 
Lindsey  et  al.  1997).  Previous  studies  dealt 
with  the  putative  El  Nino  effects  on  seabird 
colonies  or  on  terrestrial  island  birds.  Effects 
on  inland  birds  have  been  little  studied. 

The  semiarid  areas  of  northcentral  Chile 
(27-32°  S),  apart  from  seasonal  fluctuations  in 
weather  and  food  resources,  are  characterized 
by  medium  term  fluctuations  in  rainfall  (Pu- 
entes et  al.  1988).  Accordingly,  plant  cover, 
amount  of  herbage  production,  and  size  of 
seed  bank  vary  markedly  among  years  (Gu- 
tierrez et  al.  1993).  Small  mammals  track 
these  food  resources  closely  (Jimenez  et  al. 
1992)  and  perhaps  arthropods  do  also  (Puen- 
tes and  Campusano  1985).  It  has  become  in- 
creasingly clear  that  the  unusual  rainfall 
brought  by  El  Nino  events  to  semiarid  Chile 
are  responsible  for  increased  primary  produc- 
tivity, which  in  turn  leads  to  population  out- 
breaks of  small  mammals  and  to  local  increas- 
es in  the  populations  of  carnivorous  birds  that 
prey  on  them  (Meserve  et  al.  1995;  Jaksic  et 
al.  1996,  1997). 


527 


528 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4.  December  1999 


Interestingly,  no  multi-year  studies  have 
been  conducted  on  the  response  of  an  entire 
bird  assemblage  in  such  semiarid  areas  to 
changes  in  precipitation,  or  in  primary  (veg- 
etation) or  secondary  production  (mammals, 
arthropods).  Here  we  report  upon  the  effect  of 
variation  in  such  secondary  production  on 
bird  species  richness  and  density. 

STUDY  AREA  AND  METHODS 

Study  area. — The  study  site  is  located  inside  Las 
Chinchillas  National  Reserve  (simply  called  the  re- 
serve hereafter),  at  31°  30'  S,  71°  06'  W,  about  300  km 
N of  Santiago,  Chile,  and  60  km  E of  the  Pacific  coast. 
The  reserve  spans  400-1700  m elevation,  has  a rugged 
topography,  and  is  dominated  physiographically  by  an 
alternation  of  ridges  and  ravines,  with  few  flat  areas 
between.  The  climate  is  classified  as  semiarid  by  di 
Castri  and  Hajek  (1976),  with  sporadic  precipitation 
concentrated  during  Austral  winter  months  (June 
through  August).  Mean  annual  rainfall  1986-1997  was 
170  mm,  but  with  marked  increases  in  1987  (513  mm), 
1992  (307  mm),  and  1997  (367  mm)  associated  with 
the  respective  El  Nino  events  ( 1986-1987,  1991-1992, 
1997-1998).  Two  dominant  landscape  types  character- 
ize the  study  site;  ravines  and  slopes.  Ravines  are  more 
vegetated  (70.3%  shrub  cover)  than  slopes  (52.7% 
cover). 

Vegetation. — The  vegetation  is  a thornscrub  com- 
posed mainly  of  spiny  dicots,  bromeliads,  and  cacti 
(details  in  Gajardo  1978).  Dominant  species  in  ravines 
are  Stevia  sp.  (15.5%  cover),  Colliguaya  odorifera 
(12.5%),  Pleocarphu.s  revolutu.s  (6.6%),  Bacchari.s 
paniculata  (5.4%),  Proustia  cimeifolia  (5.4%),  and 
Maytenus  hoaria  (5.2%).  Dominant  species  in  slopes 
are  Bahia  amhrosioide.s  (16.9%),  Proustia  cuneifolia 
(19.3%),  Baccharis  paniculata  (8.6%),  Porlieria  chi- 
lensis  (8.0%),  Lobelia  polyphylla  (4.0%),  and  Puya 
berteroniana  (3.0%). 

Census  techniques. — During  the  four  calendar  sea- 
sons of  every  year  from  July  (winter)  1993  to  January 
1998  (summer),  we  conducted  fixed  band  transects  to 
census  birds  (Burnham  et  al.  1980,  Conner  and  Dick- 
.son  1980,  Bibby  et  al.  1993).  One  transect  was  in  a 
ravine  (1500  X 20  m = 3 ha.  time  spent  = 45-55  min/ 
transect)  and  another  in  a southern  exposure  mid-slope 
(500  X 40  m = 2 ha.  20-30  min/transect)  of  each  El 
Cobre  and  El  Grillo  creek  beds  (sampling  effort  = 4 
transects/season,  the  two  creek  beds  combined).  The 
two  ravine  tran.sects  were  surveyed  throughout  the 
study  period,  but  the  two  on  slopes  were  terminated  in 
the  summer  of  1995.  The  two  creek  beds  were  dry 
except  for  the  wet  winter  of  1997,  but  are  hereafter 
called  creeks  nonetheless.  Transects  were  started  1 h 
after  the  sun's  rising  above  the  top  of  ridges  east  of 
the  two  creeks.  Diurnal  raptors  (Falconiformes)  were 
censused  opportunistically  in  a 2000  ha  area  centered 
around  the  two  creeks.  The  abundance  of  the  four  noc- 
turnal raptor  species  (Strigiformes)  at  the  site  was  as- 
sessed based  on  responses  to  playbacks  of  their  calls 


(Johnson  et  al.  1981,  Haug  and  Didiuk  1993)  0.5  h 
after  the  first  star  was  spotted  in  the  sky.  Every  play- 
back was  broadcast  for  1 min  for  each  species  sequen- 
tially, at  four  fixed  stations  located  in  the  ravine  of  the 
El  Cobre  creek,  at  two  stations  in  the  ravine  of  El 
Grillo  creek,  and  at  two  stations  in  the  bottom  of  the 
cliffs  that  border  the  Auco  stream.  For  the  Strigiformes, 
we  considered  the  area  sampled  to  be  about  2000  ha, 
and  all  estimates  of  abundance  refer  to  the  minimum 
number  of  individuals  detected.  All  densities  are  stan- 
dardized as  the  number  of  individuals  per  species  per 
hectare. 

Bird  categorizations. — Bird  nomenclature  follows 
Meyer  de  Schauensee  (1982).  We  categorized  species 
according  to  their  reproductive  status,  diet,  residence 
status,  and  habitat.  Reproduction:  birds  were  classified 
as  either  nesting  or  non-nesting  at  the  site,  depending 
on  whether  they  were  observed  actually  nesting.  Some 
cryptic  nesters  may  have  escaped  our  detection  and 
thus  the  number  of  nesting  species  may  be  underesti- 
mated. Diet;  we  followed  Jaksic  and  Feinsinger  (1991) 
in  establi.shing  the  following  primary  diets:  carnivores, 
insectivores,  granivores,  nectarivores,  frugivores,  foli- 
vores,  and  omnivores.  These  categorizations  come 
from  the  literature  or  from  direct  observation.  Resi- 
dence: we  considered  a species  to  be  resident  at  the 
study  site  if  it  was  observed  on  13  (ca  70%)  of  the  19 
visits  to  the  site.  A species  was  categorized  as  migrant, 
either  if  it  stayed  at  the  site  only  during  spring  and 
summer  (summer  visitor)  or  if  it  stayed  only  during 
autumn  and  winter  (winter  visitor).  Habitat;  the  follow- 
ing landscape  units  were  considered  as  “habitats”;  ra- 
vines (bottom  of  creeks),  slopes  (sides  of  creeks),  flat 
areas,  and  water  edges  (around  permanent  streams). 

Food  availability. — -Small  mammal  density  was  as- 
sessed during  five  days/four  nights  trapping  bouts  in 
each  season  on  opposite  slopes  of  El  Cobre  creek. 
Each  of  the  two  grids  consisted  of  a 7 X 7 arrangement 
of  stations  at  15  m intervals  equipped  with  one  Sher- 
man live  trap  at  each  station.  Total  trapping  area  was 
2.2  ha.  Jimenez  and  coworkers  ( 1992)  provided  details 
of  this  trapping  scheme.  Teirestrial  arthropods  were 
sampled  with  Barber  pitfall  traps  (Southwood  1978) 
consisting  of  a 200  ml  plastic  vial  with  its  rim  at 
ground  level,  inside  of  which  a 100  ml  plastic  vial  was 
tightly  fitted  and  filled  with  water  and  biodegradable 
detergent.  One  hundred  such  traps  were  placed  at  El 
Cobre  creek  along  four  tran.sects  with  stations  at  10  m 
intervals.  Two  transects  were  on  opposite  slopes  and 
had  20  traps  each;  two  transects  (with  30  traps  per 
transect)  were  in  the  ravine.  Specimens  were  collected 
at  24  h intervals  during  two  consecutive  days  and  their 
total  abundance  expressed  as  individuals  per  trap  per 
24  h.  This  total  abundance  was  weighted  by  the  pro- 
portion of  traps  placed  in  each  habitat  type  (2:3, 
slopes : ravine).  Unlike  the  case  for  mammals,  this 
method  provides  information  only  on  relative  densities 
of  large  terrestrial  arthropods  through  time. 

Statistical  analyses. — To  determine  whether  there 
were  as.sociations  between  habitat  and  bird  densities 
through  time,  we  compared  data  obtained  in  the  two 


Jaksic  ami  Lazo  • BIRD  RESPONSES  TO  EL  NINO  IN  CHILE 


529 


slopes  with  those  from  the  two  ravines  because  these 
were  the  two  physiognomically  most  salient  landscape 
features  in  the  reserve.  The  degree  of  association  be- 
tween bird  densities  and  two  food  resources  (mammals 
and  arthropods)  was  determined  separately  for  carniv- 
orous (Falconiformes  and  Strigiformes)  and  insectiv- 
orous birds  by  Spearman  correlation  coefficients  (So- 
kal  and  Rohlf  1981).  Because  raptor  sightings  could 
not  be  assigned  unequivocally  to  either  slopes  or  ra- 
vines (as  they  apparently  hunted  in  both  habitat  types), 
these  data  were  pooled.  In  the  case  of  insectivorous 
birds,  we  could  assign  sightings  unequivocally  to  one 
or  the  other  of  the  two  habitats,  and  thus  we  were  able 
to  make  separate  comparisons  with  arthropod  abun- 
dance in  either  slopes  or  ravines.  Bird  and  arthropod 
densities  were  compared  by  means  of  Wilcoxon 
matched-pairs  tests  (Sokal  and  Rohlf  1981).  The  re- 
sults from  Barber  traps  may  be  considered  only  as  ap- 
proximations of  terrestrial  arthropod  density.  We  did 
not  sample  aerial  or  foliage  arthropods.  All  statistical 
analyses  were  performed  with  software  Systat  version 
7.0  for  Windows  95. 

RESULTS 

Characterization  of  the  bird  assemblage. — 
We  sighted  61  species  during  our  5 year  study 
(Table  1),  32  (53%)  of  which  were  Passeri- 
formes. Thirty  species  (49%)  were  residents 
(but  only  13  species  were  sighted  in  all  19 
visits).  Transient  species  were  equally  divided 
between  winter  visitors  (10%)  and  summer 
visitors  (10%).  Another  31%  of  the  species 
were  sighted  too  few  times  to  enable  us  to 
categorize  them,  save  as  accidentals  or  occa- 
sional. Two  species  were  not  observed  but 
known  to  be  present.  The  White-tailed  Kite 
(Elanus  leucurus)  had  been  sighted  previously 
(Jaksic  et  al.  1996)  and  thus  we  categorized  it 
as  a migrant.  We  considered  the  Great  Homed 
Owl  Bubo  virginianus  to  be  resident,  despite 
having  detected  it  during  only  five  visits  be- 
cause we  collected  freshly  regurgitated  pellets 
at  each  visit  (Jaksic  et  al.  1996). 

In  terms  of  species  numbers,  the  best  rep- 
resented trophic  groups  were  insectivores 
(34%),  carnivores  (28%),  and  granivores 
(25%),  accounting  for  87%  (53  species)  of  the 
local  assemblage  (Table  1).  Thirty-eight 
(62%)  of  the  61  species  nested  in  the  reserve, 
including  three  summer  visitors:  Aplomado 
Falcon  (Falco  femoralis).  Giant  Hummingbird 
{Patagona  gigas),  and  White-tufted  Tyrant 
{Elaenia  albiceps;  Table  1).  Seventy-seven 
percent  of  the  species  were  sighted  in  ravines, 
54%  on  slopes,  11%  on  flat  areas,  and  7% 
near  small  streams  and  ponds.  (These  per- 


centages add  to  more  than  100%  because 
some  species  visit  more  than  one  habitat  type; 
Table  1.)  Bird  densities  were  not  different,  ei- 
ther between  the  two  ravines  (Wilcoxon 
matched-pairs  test;  Z = 0.283;  n = 19,  P > 
0.05)  or  between  the  two  slopes  (Z  = 0.891; 
n = 11;  P > 0.05).  Thus,  we  felt  justified  to 
analyze  our  data  from  the  two  creeks  by  hab- 
itat type  only. 

Multi-year  trends. — Drought  conditions 
prevailed  during  the  first  four  years  of  our 
study  (1993-1996);  rainfall  ranged  40—106 
mm  compared  to  a mean  of  170  mm  for  1986- 
1997  (Fig.  1).  By  contrast,  1997  had  over 
twice  the  mean  annual  precipitation  recorded 
for  the  study  site  (Fig.  1).  Concomitantly, 
there  was  a declining  trend  in  bird  species 
richness  from  43  species  in  winter  1993  to  29 
at  the  end  of  the  drought  in  autumn  1997  (a 
33%  decrease;  Fig.  1).  Twenty-nine  species 
may  well  represent  the  minimum  number  of 
bird  species  present  in  the  reserve  at  any  time. 
As  soon  as  El  Nino  driven  precipitation  re- 
appeared at  the  site  (the  previous  two  occur- 
rences were  1987  and  1992),  there  was  an  in- 
crease from  29  species  in  autumn  1997  to  42 
species  in  summer  1997  (a  45%  increase;  Fig. 
1).  Whether  this  is  close  to  the  maximum  bird 
species  richness  that  the  site  can  accommo- 
date has  not  been  determined. 

Bird  density  also  tracked  precipitation  pat- 
terns (Fig.  2).  Bird  numbers  in  ravines  de- 
clined from  about  45-50  individuals/ha 
(1993)  to  about  10/ha  in  summer  1994  (the 
driest  year  in  the  series;  Fig.  1),  and  started  a 
slow  recovery  through  1995  and  1996,  reach- 
ing densities  of  55-60  birds/ha  during  sum- 
mer 1997.  Bird  densities  on  slopes  were  ap- 
parently less  than  those  in  ravines  (Fig.  2)  but 
paralleled  the  same  trends  until  summer  1995, 
when  we  terminated  censuses  on  slopes. 

Similar  trends  were  observed  among  small 
mammals  but  not  among  terrestrial  arthropods 
(Fig.  3).  Small  mammals  declined  markedly 
through  1993  and  1994,  remained  at  very  low 
levels  1995-1996,  and  recovered  after  the 
spring  1997  (Fig.  3).  Terrestrial  arthropods  did 
not  display  such  marked  fluctuation  (Fig.  3). 
There  were  significant  positive  correlations 
between  small  mammal  density  and  both  rap- 
tor species  richness  {r^  = 0.77,  df  = 17,  P < 
0.001)  and  raptor  density  (r^  = 0.76,  df  = 17, 
P < 0.001).  Although  positive,  there  were  no 


530 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


TABLE  1.  Birds  of  Las  Chinchillas  National  Reserve  (northcentral  Chile)  and  their  categorization  by  repro- 


diictive  status,  dietary  category,  residence  status  (in  parentheses, 
sighted),  and  habitat  type  (landscape  aspect). 

, number  of  calendar 

seasons  the  species  was 

Specie.s 

Reproduction 

Diet 

Residence 

Habitat 

Nothoprocta  perdicaria 

Nesting 

Granivore 

Resident  (18) 

Ravine/Slope 

Ca.s  me  radius  albus 

Non-nesting 

Carnivore 

Migrant  (6)“ 

Water  edge 

Egretta  thula 

Nesting 

Carnivore 

Migrant  (8) 

Water  edge 

Nycticorax  nycticorax 

Non-nesting 

Carnivore 

Migrant  (1) 

Water  edge 

Vultur  gryphus 

Non-nesting 

Carnivore 

Migrant  (5) 

Slope 

Elanus  leucurus 

Non-nesting 

Carnivore 

Migrant  (0)'’ 

Elat  areas 

Geranoaetus  melanoleucus 

Nesting 

Carnivore 

Resident  (16) 

Ravine/Slope 

Buteo  polyosoma 

Nesting 

Carnivore 

Resident  (14) 

Ravine 

Parabuteo  unicinctus 

Non-nesting 

Carnivore 

Migrant  (8) 

Flat  areas 

Milvago  chimango 

Nesting 

Insectivore 

Resident  (14) 

Ravine/Flat  areas 

Ealco  peregrinus 

Non-nesting 

Carnivore 

Summer  visitor  (5) 

Ravine/Slope 

Falco  femoralis 

Nesting 

Carnivore 

Summer  visitor  (4) 

Slope 

Ealco  sparverius 

Nesting 

Insectivore 

Resident  (13) 

Ravine/Flat  areas 

Callipepla  californica 

Nesting 

Granivore 

Resident  (19) 

Ravine/Slope 

Rallus  sanguinolentus 

Nesting 

Carnivore 

Migrant  (8) 

Water  edge 

Vanellus  chilensis 

Nesting 

Carnivore 

Resident  (14) 

Flat  areas 

Columba  araucaria 

Non-nesting 

Granivore 

Winter  visitor  (3) 

Ravine 

Zenaida  auriculata 

Nesting 

Granivore 

Resident  (14) 

Ravine/Slope 

Metriopelia  melanoptera 

Unknown 

Granivore 

Summer  visitor  (11) 

Ravine/Slope 

Cyanoliseus  patagonus 

Unknown 

Erugivore 

Resident  (15) 

Ravine 

Tyto  alba 

Nesting 

Carnivore 

Migrant  (11) 

Ravine/Slope 

Bubo  virginianus 

Nesting 

Carnivore 

Resident  (6)'’ 

Ravine/Slope 

Glaucidium  nanum 

Nesting 

Carnivore 

Resident  (13) 

Ravine/Slope 

Speotvto  cunicularia 

Nesting 

Carnivore 

Resident  (16) 

Ravine/Slope 

Caprirnulgus  longirostris 

Nesting 

Insectivore 

Resident  (19) 

Ravine 

Patagona  gigas 

Nesting 

Nectarivore 

Summer  visitor  ( 12) 

Ravine/Slope 

Seplianoides  galeritus 

Non-nesting 

Nectarivore 

Winter  visitor  (10) 

Ravine/Slope 

Colaptes  pitius 

Non-nesting 

Insectivore 

Migrant  (8) 

Slope 

Picoides  lignarius 

Nesting 

Insectivore 

Resident  (19) 

Ravine/Slope 

Geositta  rufipennis 

Non-nesting 

Granivore 

Winter  visitor  (6) 

Ravine/Slope 

Upucerthia  dumetaria 

Non-nesting 

Insectivore 

Migrant  (10) 

Ravine 

Upucerthia  ruficauda 

Non-nesting 

Insectivore 

Summer  visitor  (5) 

Ravine 

Chilia  melanura 

Nesting 

Insectivore 

Resident  (19) 

Ravine/Slope 

Leptasthenura  aegithaloides 

Nesting 

Insectivore 

Resident  (19) 

Ravine/Slope 

Asthenes  modes ta 

Nesting 

Insectivore 

Resident  (16) 

Ravine/Slope 

Pteroptochos  megapodius 

Nesting 

Insectivore 

Resident  (19) 

Ravine/Slope 

Scelorchilus  albicollis 

Nesting 

Insectivore 

Resident  (19) 

Ravine/Slope 

Scytalopus  magellanicus 

Nesting 

Insectivore 

Resident  (16) 

Ravine 

Agriornis  livida 

Non-nesting 

Carnivore 

Migrant  (7) 

Slope 

Pyrope  pyrope 

Nesting 

Insectivore 

Resident  (18) 

Ravine/Slope 

Muscisaxicola  macloviana 

Non-nesting 

Insectivore 

Winter  visitor  (2) 

Ravine/Slope 

Elaenia  albiceps 

Nesting 

Insectivore 

Summer  visitor  (7) 

Ravine 

Anairetes  parulus 

Nesting 

Insectivore 

Resident  (18) 

Ravine/Slope 

Colorhamphus  parvirostris 

Non-nesting 

Insectivore 

Winter  visitor  (8) 

Ravine 

Tachycineta  leucopyga 

Nesting 

Insectivore 

Resident  (19) 

Flat  areas 

Pygochelidon  cyanoleuca 

Non-nesting 

Insectivore 

Migrant  (3) 

Flat  areas 

Troglodytes  aedon 

Nesting 

Insectivore 

Resident  ( 19) 

Ravine/Slope 

Phytotoma  rara 

Non-nesting 

Eolivore 

Migrant  ( 1 ) 

Ravine 

Turdus  falcklandii 

Nesting 

Omnivore 

Migrant  (9) 

Ravine 

Mimus  thenca 

Ne.sting 

Omnivore 

Resident  ( 19) 

Ravine/Slope 

Sicalis  sp. 

Non-nesting 

Granivore 

Migrant  (3) 

Ravine 

Zonotrichia  capensis 

Nesting 

Granivore 

Resident  ( 19) 

Ravine/Slope 

Sturnella  loyca 

Nesting 

Omnivore 

Resident  (19) 

Ravine 

Curaeus  curaeus 

Nesting 

Omnivore 

Resident  ( 16) 

Ravine/Slope 

Phrygilus  gayi 

Non-nesting 

Granivore 

Winter  visitor  (8) 

Ravine/Slope 

Jciksic  and  Uizo  • BIRD  RESPONSES  TO  EL  NINO  IN  CHILE 


531 


TABLE  1.  CONTINUED. 

Species 

Reproduction 

Diet 

Residence 

Habiial 

Phiygilus  fniticeti 

Nesting 

Granivore 

Resident  (18) 

Ravine/Slope 

Phry’gilus  aluudinus 

Nesting 

Granivore 

Migrant  (8) 

Ravine 

Diuca  diuca 

Nesting 

Granivore 

Resident  ( 19) 

Ravine/Slope 

Carduelis  uropygialis 

Non-nesting 

Granivore 

Migrant  (2) 

Ravine 

Corduelis  harhatus 

Non-nesting 

Granivore 

Migrant  (2) 

Slope 

Passer  domesticus 

Nesting 

Granivore 

Migrant  (12) 

Ravine 

^ Species  categorized  simply  as  migrants  did  not  yield  enough  data  to  determine  whether  they  are  summer  or  winter  visitors,  occasionals,  or  accidentals. 
^ Not  sighted  during  our  work,  but  present  in  the  area.  See  text  for  details. 


significant  correlations  between  terrestrial  ar- 
thropod abundance  and  insectivorous  bird 
species  richness  (r^  = 0.06,  df  = 17,  P > 
0.05),  density  on  slopes  (r^  = 0.49,  df  = 9,  P 
> 0.05)  nor  density  on  ravines  (r^  = 0.37,  df 
= 17,  P > 0.05). 

During  the  transition  from  drought  (1993- 
1996)  to  wet  year  (1997),  two  previously  un- 
recorded species  arrived  at  the  reserve,  the  fo- 
livorous  Rufous-tailed  Plantcutter  {Phytotoma 


rara)  and  the  fish  and  amphibian  eating 
Brown  Heron  (Nycticorax  nycticorax). 

DISCUSSION 

The  percentage  (51%)  of  species  at  our 
study  site  that  were  migrants,  is  similar  (48% 
of  88  species)  to  that  reported  by  Marone 
(1992a,b)  in  the  Monte  scrubland  across  the 
Andes  at  similar  latitudes  in  Argentina.  As  in 
Chile,  the  high  percentage  of  migrants  in  the 


Calendar  seasons 

EIG.  1.  Seasonal  variation  in  rainfall  and  number  of  bird  species  at  Las  Chinchillas  National  Reserve 
throughout  the  study  period  (19  calendar  .seasons  = 5 years).  W = Winter  (June-August),  Sp  = Spring  (Sep- 
tember-November),  Su  = Summer  (December-Eebruary),  A = Autumn  (March-May).  The  segmented  line 
represents  the  mean  annual  rainfall  of  the  last  12  years. 


532 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


Calendar  seasons 

FIG.  2.  Bird  density  (mean  ± SD)  at  the  reserve  throughout  the  study  period,  by  habitat  type:  ravines 
and  slopes.  W = Winter,  Sp  = Spring,  Su  = Summer,  A = Autumn.  Censuses  on  the  slopes  were  discontinued 
after  summer  1995. 


Calendar  seasons 

FIG.  3.  Abundance  of  terrestrial  arthropods  (mean  ± SD)  and  of  small  mammals  (mean  ± SD)  at  the 
reserve  throughout  the  study  period.  Abundance  of  arthropods  is  the  mean  of  four  samples  (two  from  opposite 
facing  slopes  and  two  from  a ravine),  that  of  mammals  is  an  average  for  the  north  and  south  facing  slopes.  W 
= Winter,  Sp  = Spring,  Su  = Summer,  A = Autumn. 


Jciksic  and  Uizo  • BIRD  RESPONSES  TO  EL  NINO  IN  CHILE 


533 


Argentine  Monte  may  be  associated  with  the 
extreme  fluctuations  in  precipitation  that  char- 
acterize arid  and  semiarid  regions  of  South 
America.  Varying  precipitation  levels  in  Chile 
are  associated  with  changes  in  primary  and 
secondary  production  (Fuentes  and  Campu- 
sano  1985;  Jaksic  et  al.  1996,  1997).  Migrant 
birds  may  exploit  these  unpredictable  resourc- 
es by  moving  north  from  the  more  mesic 
Mediterranean  region  of  central  Chile  or  down 
from  the  Coastal  Range  to  the  west  or  the  An- 
des to  the  east  of  our  study  area. 

Marone  (1992a,b)  observed  that  insecti- 
vores  (52%),  carnivores  (19%),  and  grani- 
vores  (15%)  dominated  the  Argentine  Monte, 
accounting  for  86%  of  the  local  assemblage, 
similar  to  the  87%  we  recorded  in  Chile.  Nev- 
ertheless, there  were  more  insectivores  and 
fewer  carnivores  and  granivores  in  the  Argen- 
tine site  than  in  our  Chilean  site.  This  suggests 
that  there  may  be  differences  in  resource  lev- 
els between  Chilean  and  Argentine  semiarid 
sites.  More  data  are  needed  from  the  Monte. 

We  detected  positive  correlations  between 
number  and/or  density  of  bird  species  and  spe- 
cific resource  levels  (i.e.,  with  regard  to  small 
mammals  but  not  to  terrestrial  arthropods). 
Gutierrez  and  coworkers  (pers.  comm.)  mea- 
sured the  seed  bank  of  the  reserve  before 
(1996)  and  during  El  Nino  (1997).  We  found 
a positive  association  between  this  resource 
and  richness/density  of  granivorous  birds  dur- 
ing the  respective  years.  In  central  Chile,  Lo- 
pez-Calleja  (1995)  found  that  the  granivorous 
Diuca  Finch  (Diuca  diuca)  and  Rufous-col- 
lared  Sparrow  (Zonotrichia  capensis),  two 
species  that  were  also  present  at  our  site, 
changed  their  seed  preferences  in  response  to 
short-term  fluctuations  in  seed  abundance.  A 
detailed  study  of  the  abundance  of  these  two 
species  and  their  respective  diets  in  relation  to 
the  seed  bank  in  the  reserve  is  needed. 

Why  were  birds  in  our  study  site  more 
abundant  in  ravines  than  in  slopes?  Perhaps 
food  levels  are  higher  in  ravines  than  on 
slopes.  Also,  there  is  more  vegetated  ground 
in  ravines  (70%)  than  on  slopes  (53%).  Struc- 
tural aspects  of  the  former  habitat  may  render 
it  more  attractive  to  birds.  The  presence  of  a 
tree  layer  (e.g.,  Maytenus  boaria,  Quillaja  sa- 
ponaria)  in  ravines,  which  is  absent  from 
slopes,  may  favor  higher  bird  densities  be- 
cause it  provides  more  shelter,  roosting  and 


nesting  sites,  and  perhaps  greater  food  diver- 
sity (e.g.,  foliage  insects). 

The  bird  assemblage  in  the  reserve  showed 
both  short  (seasonal)  and  multi-year  variation 
in  its  composition  and  density.  Although  bird 
species  richness  and  density  in  our  study  site 
were  lower  during  drought  years  and  higher 
during  wet  years,  it  should  be  noted  that 
droughts  are  more  frequent  and  last  longer 
than  El  Nino  events  (eight  dry  years  versus 
four  wet  ones  in  1986-1997).  In  a semiarid 
climate,  perhaps  dry  years  set  the  baseline  in 
species  richness  and  density  for  the  bird  as- 
semblage, and  both  increase  during  wet  years 
because  of  the  immigration  of  opportunistic 
species  from  elsewhere.  Thus,  El  Nino  driven 
rains  impose  a strong  abiotic  influence,  which 
cascades  from  increased  primary  and  second- 
ary productivity  to  bird  species  richness  and 
density. 

ACKNOWLEDGMENTS 

B.  Saavedra,  E.  Silva,  R.  Soto,  and  S.  Tellier  helped 
us  with  their  respective  expertises.  R Feinsinger  helped 
us  in  many  ways.  C.  F.  Estades.  F.  Hertel,  M.  Marini, 
F Vuilleumier  and  two  anonymous  reviewers  made  co- 
gent criticisms  that  helped  us  improve  our  presentation 
of  data.  The  study  was  funded  by  grants  Fondecyt  193- 
0639  and  196-0319,  NSF-INT  92-14085,  the  Mellon 
Fund.  Lazo  acknowledges  the  support  of  the  Mellon 
Foundation  toward  postgraduate  training.  Jaksic  is  sup- 
ported by  a Presidential  Chair  in  Science,  Republic  of 
Chile. 

LITERATURE  CITED 

Avery,  M.  L.  and  C.  van  Riper,  III.  1989.  Seasonal 
changes  in  bird  communities  of  the  chapanal  and 
blue-oak  woodland  in  central  California.  Condor 
91:288-295. 

Barber,  R.  T.  and  F P.  Chavez.  1983.  Biological  con- 
sequences of  El  Nino.  Science  222:1203-1210. 
Bibby,  C.  J..  N.  D.  Burgess,  and  D.  A.  Hill.  1993. 
Bird  census  techniques.  Academic  Press,  London, 
U.K. 

Burnham,  K.  R,  D.  R.  Anderson,  and  J.  L.  Laake. 
1980.  Estimation  of  density  from  line  transect 
sampling  of  biological  populations.  Wildl.  Mon- 
ogr.  72:1-202. 

Conner,  R.  N.  and  J.  G.  Dickson.  1980.  Strip  transect 
sampling  and  analysis  for  avian  habitat  studies. 
Wildl.  ,Soc.  Bull.  8:4-10. 

di  Castri,  F.  and  E.  Hajek.  1976.  Bioclimatologi'a  de 
Chile.  Ediciones  Univ.  Catolica  de  Chile,  Santia- 
go, Chile. 

Feinsinger  R,  L.  A.  Swarm,  and  J.  A.  Wolfe.  1985. 
Nectar-feeding  birds  on  Trinidad  and  Tobago: 


534 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


comparison  of  diverse  and  depauperate  guilds. 
Ecol.  Monogr.  55:1-28. 

Fuentes,  E.  R.  and  C.  Campusano.  1985.  Pest  out- 
breaks and  rainfall  in  the  semi-arid  region  of 
Chile.  J.  Arid  Envir.  8:67-72. 

Fuentes,  E.  R.,  E.  R.  Hajek,  and  G.  Espinoza.  1988. 
Some  consequences  of  rainfall  variability  for 
Mediterranean  type  ecosystems  in  Chile.  Pp.  347- 
360  in  Time  scales  and  water  stress  (F.  di  Castri, 
C.  Floret,  S.  Rambal,  and  J.  Roy,  Eds.).  Interna- 
tional Union  of  Biological  Sciences,  Paris,  France. 

Gajardo,  R.  1978.  Antecedentes  para  la  determinacion 
de  las  comunidades  vegetales  de  Auco.  (Illapel, 
IV  Region).  Cs.  For.  Chile  1:19-27. 

Gibbs,  H.  L.  and  P.  R.  Grant.  1987.  Ecological  con- 
sequences of  an  exceptionally  strong  El  Nino 
event  on  Darwin’s  finches.  Ecology  68:1735- 
1746. 

Grant,  P.  R.  and  B.  R.  Grant.  1987.  The  extraordi- 
nary El  Nino  event  of  1982-83:  effects  on  Dar- 
win’s finches  on  Isla  Genovesa,  Galapagos.  Oikos 
49:55-66. 

Gutierrez,  J.  R.,  P.  L.  Meserve,  E M.  Jaksic,  L.  C. 
Contreras,  S.  Herrera,  and  H.  Vasquez.  1993. 
Structure  and  dynamics  of  vegetation  in  a Chilean 
arid  thornscrub  community.  Acta  Oecol.  14:271  — 
285. 

Hall,  G.  A.,  H.  L.  Gibbs,  P.  R.  Grant,  L.  W.  Bots- 
FORD,  AND  G.  S.  Butcher.  1988.  Effects  of  El 
Nino-Southem  Oscillation  (ENSO)  on  terrestrial 
birds.  Proc.  Int.  Ornithol.  Congr.  19:1759—1775. 

Haug,  E.  a.  and  a.  B.  Didiuk.  1993.  Use  of  recorded 
calls  to  detect  burrowing  owls.  J.  Field  Ornithol. 
64:188-194. 

Jaksic,  E M.  and  P.  Feinsinger.  1991.  Bird  assem- 
blages in  temperate  forests  of  North  and  South 
America:  a compari.son  of  diversity,  dynamics, 
guild  structure,  and  resource  use.  Rev.  Chilena 
Hist.  Nat.  64:491-510. 

Jaksic,  E M.,  P.  Feinsinger,  and  J.  E.  Jimenez.  1996. 
Ecological  redundancy  and  long-term  dynamics  of 
vertebrate  predators  in  semiarid  Chile.  Conserv. 
Biol.  10:252-262. 

Jaksic,  F.  M.,  S.  I.  Silva,  P.  L.  Meserve,  and  J.  R. 
Gutierrez.  1997.  A long-term  study  of  vertebrate 
predator  responses  to  an  El  Nino  (ENSO)  distur- 
bance in  western  South  America.  Oikos  78:341- 
354. 

Jimenez,  J.  E.,  P.  Feinsinger,  and  E M.  Jaksic.  1992. 
Spatiotemporal  patterns  of  an  imiption  and  de- 
cline of  small  mammals  in  north-central  Chile.  J. 
Mamm.  73:356-364. 

Johnson,  R.,  R.  Brown,  L.  T.  Haight,  and  J.  M.  Simp- 
son. 1981.  Playback  recording  as  a special  avian 
censusing  technique.  Stud.  Avian  Biol.  6:68-75. 

Karr,  J.  R.  and  K.  E.  Freemark.  1983.  Habitat  selec- 
tion and  environmental  gradients:  dynamics  in  the 
“stable”  tropics.  Ecology  64:1481-1494. 

Koti.er,  B.  P.  and  R.  D.  Holt.  1989.  Predation  and 
competition:  the  interaction  of  two  types  ot  spe- 
cies interactions.  Oikos  54:256-260. 


Lima,  S.  L.  1987.  Clutch  size  in  birds:  a predation 
perspective.  Ecology  68:1062-1070. 

Lindsey,  G.  D.,  T.  K.  Pratt,  M.  H.  Reynolds,  and  J. 
D.  Jacobi.  1997.  Response  of  six  species  of  Ha- 
waiian forest  birds  to  a 1991—1992  El  Nino 
drought.  Wilson  Bull.  109:339—343. 

Loiselle,  B.  a.  and  j.  G.  Blake.  1991.  Temporal  var- 
iation in  birds  and  fruits  along  an  elevational  gra- 
dient in  Costa  Rica.  Ecology  72:180-193. 

Lopez-Calleja,  M.  V.  1995.  Dieta  de  Zonotrichia  ca- 
pensis  (Emberizidae)  y Diuca  cliuca  (Fringillidae): 
efecto  de  la  variacion  estacional  de  los  recursos 
troficos  y la  riqueza  de  aves  granivoras  en  Chile 
central.  Rev.  Chil.  Hist.  Nat.  68:321—331. 

Marone,  L.  1992a.  Seasonal  and  year-to-year  fluctu- 
ations of  bird  populations  and  guilds  in  the  Monte 
Desert,  Argentina.  J.  Field  Ornithol.  63:264—308. 

Marone,  L.  1992b.  Estatus  de  residencia  y categori- 
zacion  trofica  de  las  especies  de  aves  en  la  Re- 
serva  de  la  Biosfera  de  Nacunan,  Mendoza.  Hor- 
nero  13:207-210. 

Martin,  T.  E.  1985.  Resource  selection  by  tropical  fru- 
givorous  birds:  integrating  multiple  interactions. 
Oecologia  66:563-573. 

Martin,  T.  E.  1987.  Food  as  a limit  on  breeding  birds: 
a life  history  perspective.  Annu.  Rev.  Ecol.  Syst. 
18:453-487. 

Massey,  B.  W.,  D.  W.  Bradley,  and  J.  L.  Atwood. 
1992.  Demography  of  a California  Least  Tern  col- 
ony including  effects  of  the  1982—83  El  Nino. 
Condor  94:976-983. 

Meserve,  P.  L.,  J.  A.  Yunger,  J.  R.  Gutierrez,  L.  C. 
Contreras,  W.  B.  Milstead,  B.  K.  Lang,  K.  L. 
Cramer,  S.  Herrera,  V.  O.  Lagos,  S.  I.  Silva,  E. 
L.  Tabilo,  M.  a.  Torrealba,  and  E M.  Jaksic. 
1995.  Heterogeneous  responses  of  small  mammals 
to  an  El  Nino  Southern  Oscillation  event  in  north- 
central  semiarid  Chile  and  the  importance  of  eco- 
logical scale.  J.  Mamm.  76:580-595. 

Meyer  de  Schauensee,  R.  1982.  A guide  to  the  birds 
of  South  America.  International  Council  for  Bird 
Preservation,  Philadelphia,  Pennsylvania. 

Miskelly,  C.  M.  1990.  Effects  of  the  1982—83  El  Nino 
event  on  two  endemic  landbirds  on  the  Snares  Is- 
lands, New  Zealand.  Emu  90:24—27. 

Pearson,  D.  L.  1991.  A basis  for  developing  broader 
generalizations  of  bird  community  structure  and 
species  co-occurrence.  Proc.  Int.  Ornithol.  Congr. 
20:1462-1469. 

Pulliam,  H.  R.  and  J.  B.  Dunning.  1987.  The  influ- 
ence of  food  supply  on  local  density  and  diversity 
of  sparrows.  Ecology  68:1009-1014. 

ScHREiBER,  R.  W.  AND  E.  A.  ScHREiBER.  1984.  Central 
Pacific  seabirds  and  the  El  Nino  Southern  Oscil- 
lation: 1982  to  1983  perspectives.  Science  225: 
713-716. 

SoKAL,  R.  R.  AND  E J.  Rohlf.  1981.  Biometry:  the 
principles  and  practice  of  statistics  in  biological 
re.search.  Second  ed.  W.  H.  Freeman  & Co.,  San 
Francisco,  California. 

SouTHWOOD,  T.  R.  E.  1978.  Ecological  methods  with 


Jciksic  and  Lazo  • BIRD  RESPONSES  TO  EL  NINO  IN  CHILE 


535 


particular  reference  to  the  study  of  insect  popu- 
lations, second  ed.  Chapman  and  Hall,  London, 
U.K. 

Wiens,  J.  A.  1990a.  The  ecology  of  bird  communities. 
Vol.  1.  Cambridge  Univ.  Press,  Cambridge,  U.K. 


Wiens,  J.  A.  1990b.  The  ecology  of  bird  communities. 

Vol.  2.  Cambridge  Univ.  Pre.ss,  Cambridge,  U.K. 
Wiens,  J.  A.  1993.  Pat  times,  lean  times  and  compe- 
tition among  predators.  Trends  Ecol.  Evol.  8:348- 
349. 


Wilson  Bull.,  111(4),  1999,  pp.  536-540 


POTENTIAL  FOR  PREDATOR  LEARNING  OF  ARTIFICIAL 
ARBOREAL  NEST  LOCATIONS 

RICHARD  H.  YAHNER"  AND  CAROLYN  G.  MAHAN'  - 


ABSTRACT. — We  examined  the  potential  for  predators  to  learn  the  location  of  artificial  arboreal  (1.5  m above 
ground)  nests  in  a managed  forested  landscape  of  central  Pennsylvania  from  June— July  1995.  We  tested  the 
hypothesis  that  predators  do  not  learn  the  location  of  artificial  arboreal  nests  placed  repeatedly  at  the  same  sites 
(fixed  nests)  versus  those  placed  at  random  sites  in  three  habitats  created  by  clearcutting  (forested  patches, 
forested  corridors,  contiguous  forest).  Sixty-nine  (23%)  of  299  total  nests  in  five  combined  trials  were  disturbed 
by  predators;  11  (16%)  of  these  disturbances  were  attributed  to  corvids.  Predation  rates  were  greater  on  nests 
placed  at  random  (28%)  compared  to  fixed  sites  (18%,  P < 0.05),  indicating  predators  did  not  learn  or  return 
to  the  location  of  arboreal  nests  during  our  study.  Predation  rates  varied  significantly  (P  < 0.001)  among  habitats, 
with  49%  of  the  nests  disturbed  in  the  forested-patch  habitat  versus  only  7%  and  13%  in  forested-corridor  and 
contiguous-forest  habitats,  respectively.  We  propose  that  predation  was  higher  in  forested  patches  than  in  the 
other  two  habitats  because  the  former  had  greater  amounts  of  edge.  Received  12  Nov.  1998,  accepted  10  May 
1999. 


Artificial  nest  studies  have  been  useful  in 
examining  the  relationships  between  avian 
nesting  success  and  landscape  patterns  (e.g., 
Paton  1994,  Bayne  and  Hobson  1997).  Sev- 
eral investigators  have  indicated  that  depre- 
dation of  artificial  and  natural  avian  nests  in 
managed  forests  varies  with  landscape  pat- 
terns created  by  clearcutting  (Yahner  and  Ross 
1995,  Vander  Haegen  and  DeGraaf  1996, 
Yahner  and  Mahan  1996a).  However,  if  pre- 
dation rates  on  artificial  nests  are  used  as  an 
indicator  of  temporal  or  spatial  trends  in  avian 
nesting  success  (Yahner  1996,  Sargent  et  al. 
1998,  Wilson  et  al.  1998),  then  the  potential 
effect  of  the  ability  of  predators  to  learn  the 
locations  of  artificial  nests  needs  to  be  deter- 
mined. For  example,  as  a consequence  of 
clearcutting  in  a localized  area,  the  availability 
of  suitable  nest  sites  may  decline,  thereby  en- 
abling predators  to  find  nests  located  in  the 
remaining  uncut  forested  tracts  (patches  or 
conidors). 

Forest  clearcutting  for  Ruffed  Grouse  {Bon- 
asa  umbellus)  at  the  Barrens  Grouse  Habitat 
Management  Area  (GHMA)  in  central  Penn- 
sylvania provided  us  with  an  ideal  opportunity 
to  test  the  hypothesis  that  predation  rates  did 
not  vary  between  artificial  arboreal  (1.5  m 


' .School  of  Forest  Resources,  The  Pennsylvania 
.State  Univ.,  University  Park,  PA  16802-4300;  E-mail: 
rhy@psu.edu 

2 Current  address:  Dept,  of  Biology,  Penn  State  Al- 
toona, Altoona,  PA  16601. 

’ Corresponding  author. 


above  ground)  nests  placed  at  sites  used  re- 
peatedly (fixed  nests)  versus  random  sites  in 
a managed  forested  landscape.  To  our  knowl- 
edge, predation  rates  on  artificial  nests  at  fixed 
vs  random  sites  has  been  examined  only  with 
ground  nests  (Yahner  and  Mahan  1996a). 

STUDY  AREA  AND  METHODS 

Our  study  was  conducted  on  a 1166-ha  Barrens 
GHMA,  State  Game  Lands  176,  Centre  County,  Penn- 
sylvania, where  a series  of  experimental  studies  deal- 
ing with  depredation  of  artificial  and  actual  nests  have 
been  conducted  (e.g.,  Yahner  and  Wright  1985,  Yahner 
1991,  Yahner  and  Ross  1995,  Yahner  and  Mahan 
1996a).  The  Barrens  GHMA  includes  reference  (con- 
tiguous forest  habitat)  and  treated  (forested-patch  and 
forested-corridor  habitats)  sectors  of  similar  size  (Fig. 
1).  The  treated  sector  is  divided  into  136  contiguous 
4-ha  blocks,  and  each  block  is  partitioned  into  four  1 
ha  ( 100  X 100  m)  plots  arranged  in  a elockwise  pattern 
(plots  A-D).  At  the  first  cutting  cycle  (winter  1976- 
1977),  plot  A was  clearcut  in  each  block.  At  the  second 
cycle  (winter  1980-1981),  plot  B was  clearcut  in  each 
block  of  the  forested-patch  habitat.  At  the  third  and 
last  cycle  (winters  1985—1986  and  1986—1987),  plot  B 
in  each  block  of  the  forested-conidor  habitat  and  plot 
C in  each  block  of  the  forested-patch  habitat  were 
clearcut.  The  remaining  uncut  plots  in  the  treated  sec- 
tor and  forest  in  the  reference  sector  have  not  been 
clearcut  for  75-80  years.  As  a result  of  these  three 
cutting  cycles,  a mosaic  of  uncut  plots  (plot  D)  entirely 
surrounded  by  clearcut  plots  of  three  age  classes  (plots 
A-C)  occuned  in  the  forested-patch  habitat,  whereas 
100  m wide  corridors  of  uncut  plots  (plots  C— D)  re- 
mained in  the  forested-corridor  habitat  (Fig.  1). 

We  placed  artificial  arboreal  ( 1 .5  m above  ground) 
nests  during  five  time  periods  (trials)  from  early  June 
through  July  1995  (Yahner  and  Mahan  1996a).  A trial 
was  6 days  in  length,  with  8 days  between  trials.  At 


536 


Yahner  and  Mahan  • PREDATOR  LEARNING 


537 


REFERENCE  TREATED 

SECTOR 


Tt  976-77  r> 

’777777Z 

31985-87T 

rm 

1 Uncut  1 

1 Uncut  \ 

c 

D 

UNIMPROVED  ROAD 
I I CONTIGUOUS-FOREST  HABITAT 
V777\  FORESTED-CORRIDOR  HABITAT 
FORESTED-PATCH  HABITAT 


FIG.  I.  Schematic  of  reference  and  treated  sectors  at  the  Barrens  GHMA,  Centre  County,  Pennsylvania. 
Dates  of  cutting  cycles  are  given  in  plots  A and  B of  the  76  blocks  in  the  forested-corridor  habitat  of  forest 
clearcutting  and  in  plots  A— C of  the  60  blocks  in  the  forested-patch  habitat  of  clearcutting.  Forest  in  the 
contiguous-forested  habitat  of  clearcutting  (reference  sector),  in  plots  C and  D of  the  forested-corridor  habitat, 
and  in  plot  D of  the  forested-patch  habitat. 


the  beginning  of  the  study,  10  uncut  plots  (plot  D) 
were  chosen  randomly  in  both  forested-patch  and  for- 
ested-corridor habitats  and  10  sites  were  randomly  se- 
lected in  the  contiguous  forest.  These  30  sites  were 
designated  as  fixed  nests  and  were  used  in  all  trials  (1— 
5)  for  ne.st  placement.  For  each  trial,  we  randomly 
chose  10  additional  uncut  plots  (plot  D)  each  in  both 
forested-patch  and  forested-corridor  habitats  and  10 
sites  in  the  contiguous  forest;  these  additional  30  sites 
were  termed  random  nests.  This  resulted  in  60  nests/ 
trial,  with  20  nests/habitat  (forested  patch,  forested  cor- 
ridor, and  contiguous  forest)  and  30  nests/nest-site  type 
(fixed  and  random). 

Artificial  nests  (10  cm  diam  and  10  cm  deep)  were 
constructed  of  chicken  wire  painted  flat  black  to  reduce 
glare  and  lined  with  leaf  litter;  nests  were  attached  to 
the  nearest  woody  stem  (1-5  cm  dbh)  with  green  wire 
(Yahner  and  Scott  1988).  Two  fresh,  brown  chicken 
eggs  were  placed  in  each  nest  and  sunk  slightly  below 
the  rim  of  the  nest  to  minimize  detection.  We  chose 
large  brown  chicken  eggs  in  this  study  because  they 
allowed  us  to  directly  compare  our  results  with  those 
obtained  in  other  studies  at  the  study  site,  including 


artificial  ground  and  arboreal  nest  studies  conducted 
before  the  third  cutting  cycle  (e.g.,  Yahner  and  Wright 
1985,  Yahner  and  Scott  1988),  an  artificial  ground  nest 
study  conducted  after  the  third  cutting  cycle  (Yahner 
and  Mahan  1996a),  and  a study  of  Wood  Thmsh  nest- 
ing success  after  the  third  cutting  cycle  (Yahner  and 
Ross  1995).  One  nest  was  established  at  each  site.  In 
forested-patch  and  forested-corridor  habitats,  nests 
were  located  50  m from  an  edge  in  the  center  of  plot 
D;  in  the  contiguous  forest,  nests  were  placed  at  least 
50  m from  an  edge  (e.g.,  logging  road).  Rubber  gloves 
and  boots  were  worn  when  placing  nests  to  reduce 
human  scent  (Nol  and  Brooks  1982). 

We  determined  the  fates  of  nests  (e.g..  undisturbed, 
di.sturbed  by  an  avian  predator,  disturbed  by  a nona- 
vian  predator)  at  the  end  of  each  trial  (Yahner  and 
Mahan  1996a).  Nest  predators  were  classified  by  mode 
of  disturbance  and  general  nest  appearance;  eggs  with 
peck  holes  were  categorized  as  preyed  upon  by  birds, 
and  nests  without  eggs  or  with  crushed  eggs  were  clas- 
sified as  preyed  upon  by  nonavian  predators  (Rearden 
1951,  Yahner  and  Scott  1988,  Hernandez  et  al.  1997). 
Eggs  and  egg  fragments  were  removed  from  nests  at 


538 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


TABLE  1.  Pate  of  artificial  arboreal  nests  in  relation  to  type  of  nest  site,  habitat,  and  trial  in  a 
forested  landscape  at  the  Barrens  GHMA,  Centre  County,  Pennsylvania,  June— July  1995. 

managed 

Nest  fate 

Type  of 
nest-site 

Trial 

Habitat 

1 

2 

3 

4 

5 

Total 

Undisturbed 

Pixed 

Porested  patch 

8 

7 

5 

4 

6 

30 

Porested  corridor 

8 

9 

8 

10 

10 

45 

Contiguous  forest 

9 

9 

9 

10 

10 

47 

25 

25 

22 

24 

26 

122 

Random 

Porested  patch 

7 

3 

4 

3 

4 

21 

Forested  corridor 

10 

10 

8 

10 

9 

47 

Contiguous  forest 

7 

9 

6 

10 

8 

40 

24 

22 

18 

23 

21 

108 

Disturbed 

Pixed 

Forested  patch 

2 

3 

5 

6 

4 

20 

Forested  corridor 

1 

1 

2 

0 

0 

4 

Contiguous  forest 

1 

1 

1 

0 

0 

3 

4 

5 

8 

6 

4 

27 

Random 

Forested  patch 

3 

7 

6 

7 

6 

29 

Forested  corridor 

0 

0 

2 

0 

1 

3 

Contiguous  forest 

3 

1 

4 

0 

2 

10 

6 

8 

12 

7 

9 

42 

the  end  of  each  trial.  The  location  of  one  nest  in  the 
forested-corridor  habitat  was  not  found  after  nest 
placement  during  trial  1 . 

Common  bird  species  nesting  in  uncut  forest  within 
2 m of  ground  level  at  the  Barrens  Grouse  HMA  were 
Wood  Thrush  (Hylocichla  mustelina)  and  Eastern 
Towhee  (Pipilo  erythrophlhaImu.s\  Yahner  1991).  Po- 
tential predators  on  artificial  arboreal  nests  were  Amer- 
ican Crow  (Corvus  hrachyrhynchos).  Blue  Jay  (Cyan- 
ocitta  cristata),  and  raccoon  (Procyon  lotor,  Yahner 
and  Scott  1988,  Yahner  and  Morrell  1991).  Smaller 
mammalian  predators,  e.g.,  eastern  chipmunk  (Tamia.s 
striatu.s)  and  white-footed  mice  (Peromy.scu.s  leuco- 
pu.s),  probably  had  minimal  effect  on  our  artificial 
nests  because  of  the  relatively  large  egg  size  (see  Rop- 
er 1992,  Haskell  1995,  DeGraaf  and  Maier  1996,  Yah- 
ner and  Mahan  1996b). 

We  examined  dependency  of  nest  fate  (undisturbed 
and  disturbed)  on  nest-site  type  (fixed  versus  random), 
habitat  (forested  patch,  forested  corridor,  and  contigu- 
ous forest),  and  trial  (1-5)  using  a four-way  test-of- 
independence  (BMDP4L  Log-Linear  Model;  Dixon 
1990).  Likelihood  ratios  (G’)  were  used  to  determine 
interactions  of  nest  fate  with  the  three  other  variables 
using  log-linear  models  (Dixon  1990.  Sokal  and  Rohlf 
1995).  If  nest  fate  was  dependent  on  a variable  with 
more  than  two  levels,  we  used  2X2  G-tests-of-inde- 
pendence  about  the  cell(s)  of  interest. 

RESULTS 

Sixty-nine  (23%)  of  the  299  artificial  ar- 
boreal nests  were  disturbed  during  the  five  tri- 
als combined  (Table  1);  one  nest  location  was 


not  found  in  trial  1.  We  attributed  11  (16%) 
of  the  disturbed  nests  to  avian  predators.  Nest 
fate  was  dependent  on  nest  type,  with  fewer 
arboreal  nests  disturbed  at  fixed  than  at  ran- 
dom sites  (18%  vs  28%,  respectively;  G = 
4.0,  df  = I,  P < 0.05). 

Nest  fate  varied  with  habitat  (G  = 55.8,  df 
= 2,  P < 0.001).  Rate  of  nest  disturbance  was 
higher  in  the  forested-patch  habitat  (49%) 
compared  to  either  forested-corridor  (7%)  or 
contiguous-forested  habitats  (13%;  G ^ 22.3, 
df  = 1,  P < 0.001).  The  number  of  disturbed 
nests  in  the  forested-corridor  habitat,  however, 
was  similar  to  that  in  the  contiguous-forest 
habitat  (P  > 0.05).  In  contrast,  nest  fate  was 
not  associated  with  trial  or  with  interactions 
of  two  or  more  variables  (P  > 0.05). 

DISCUSSION 

We  believe  that  predators  did  not  learn  the 
location  of  arboreal  nests  in  our  study  (Eibl- 
Eibesfeldt  1970,  Krebs  1978,  Yahner  and 
Wright  1985)  because  disturbance  rates  were 
higher  at  random  than  at  fixed  sites  and  be- 
cause rates  did  not  vary  among  trials.  In  an- 
other study  of  artificial  nests,  both  avian  and 
mammalian  predators  preyed  upon  nests  ran- 
domly and  did  not  learn  the  location  of  ex- 
perimental nests  (Angelstam  1986).  In  con- 


Yalwer  ami  Mahan  - PREDATOR  LEARNING 


539 


trast,  previous  work  at  the  Barrens  GHMA 
showed  that  predators  probably  learned  the  lo- 
cation of  ground  nests  at  fixed  nests  in  the 
forested-patch  sector,  particularly  as  the  study 
progressed  (trials  4 and  5;  Yahner  and  Mahan 
1996a). 

Because  artificial  nests  pose  potential  biases 
and  the  debate  on  their  usefulness  in  assessing 
success  of  natural  nests  continues,  caution 
should  be  used  in  interpreting  the  results  ob- 
tained from  artificial  nest  studies  in  making 
management  decisions  (e.g.,  Yahner  1996,  Or- 
tega et  al.  1998,  Wilson  et  al.  1998).  Care 
should  be  used  when  extrapolating  results  ob- 
tained from  artificial  nest  studies  compared  to 
naturally  occurring  nests  because  predation 
rates  on  the  two  types  of  nests  may  vary  and 
predation  rates  may  differ  among  years  (Sto- 
raas  1988).  For  example,  predators  may  use 
behavioral  cues  from  nesting  birds  to  locate 
naturally  occurring  nests.  Well  designed  stud- 
ies using  artificial  nests  remain  a useful  ap- 
proach to  making  inferences  about  factors  af- 
fecting avian  nesting  success,  especially  when 
comparisons  are  made  between  local  habitats, 
among  nests  in  a given  locality,  at  the  same 
locality  over  several  years,  or  in  detecting 
trends  in  rates  of  predation  (Roper  1992,  Yah- 
ner and  Mahan  1996a,  Wilson  et  al.  1998). 

Our  study  and  others  provided  evidence 
that  uncut  wooded  corridors,  which  are  at 
least  100  m wide  in  a forested  landscape  af- 
fected by  clearcutting,  may  provide  consider- 
ably more  secure  nesting  habitat  for  breeding 
birds  than  small  uncut  forest  stands.  For  ex- 
ample, Yahner  and  Ross  (1995)  found  lower 
predation  on  Wood  Thrush  nests  in  the  for- 
ested-corridor  habitat  (50%)  than  in  the  con- 
tiguous forest  (61%)  or  forested-patch  habitats 
(100%).  Based  on  a study  of  nest  predation 
along  uncut  buffer  strips  retained  after  clear- 
cutting  near  streams  in  Maine,  Vander  Haegen 
and  DeGraaf  (1996)  provided  evidence  that 
relatively  wide  (>150  m)  strips  enhanced 
nesting  success  of  forest  birds.  Their  study  in- 
cluded artificial  ground  and  arboreal  nests 
containing  Japanese  Quail  (Coturnix  coturnix) 
eggs.  Despite  conflicting  evidence  for  preda- 
tor learning  of  the  location  of  artificial  arbo- 
real versus  ground  nests,  we  recommend  that 
investigators  using  artificial  nests  in  frag- 
mented forested  landscapes  carefully  random- 
ize nest  placement  in  order  to  mitigate  detec- 


tion of  nests  by  predators  (see  Yahner  and  Ma- 
han 1996a). 

ACKNOWLEDGMENTS 

Our  study  was  funded  by  the  Pennsylvania  Agricul- 
tural Experiment  Station  and  the  Max  McGraw  Wild- 
life Eoundation.  We  thank  B.  Niccolai,  B.  Ross,  and 
C.  Stem  for  field  assistance. 

LITERATURE  CITED 

Angelstam,  P.  1986.  Predation  on  ground-nesting 
birds’  nests  in  relation  to  predator  densities  and 
habitat  edge.  Oikos  47:365-373. 

Bayne,  E.  M.  and  K.  A.  Hobson.  1997.  Comparing 
the  effects  of  landscape  fragmentation  by  forestry 
and  agriculture  on  predation  of  artificial  nests. 
Conserv.  Biol.  11:1418-1429. 

DeGraaf,  R.  M.  and  T.  J.  Maier.  1996.  Effect  of  egg 
size  on  predation  by  white-footed  mice.  Wilson 
Bull.  108:535-539. 

Dixon,  W.  J.  (Chief  Ed.).  1990.  BMDP  statistical  soft- 
ware manual.  Univ.  of  California  Press,  Berkeley. 
Eibl-Eibesfeldt,  I.  1970.  Ethology:  the  biology  of  be- 
havior. Holt,  Rinhart,  and  Winston,  New  York. 
Haskell,  D.  G.  1995.  Reevaluation  of  the  effects  of 
forest  fragmentation  on  rates  of  bird-nest  preda- 
tion. Conserv.  Biol.  9:1316—1318. 

Hernandez,  E,  D.  Rollins,  and  R.  Cantu.  1997.  An 
evaluation  of  Trialmaster*  camera  systems  for 
identifying  ground-nest  predators.  Wildl.  Soc. 
Bull.  25:848-853. 

Krebs,  J.  R.  1978.  Optimal  foraging:  decision  rules  for 
predators.  Pp.  23-63  in  Behavioural  ecology  (J. 
R.  Krebs  and  N.  B.  Davies,  Eds.).  Sinauer  Asso- 
ciates, Inc.,  Sunderland,  Massachusetts. 

Nol,  E.  and  R.  j.  Brooks.  1982.  Effects  of  predator 
exclosures  on  nesting  success  of  Killdeer.  J.  Field 
Ornithol.  53:263-268. 

Ortega,  C.  P,  J.  C.  Ortega,  C.  A.  Rapp,  and  S.  A. 
Backensto.  1998.  Validating  the  use  of  artificial 
nests  in  predation  experiments.  J.  Wildl.  Manage. 
62:925-932. 

Baton,  W.  C.  1994.  The  effect  of  edge  on  avian  nest- 
ing success:  how  strong  is  the  evidence?  Conserv. 
Biol.  8:17-26. 

PiCMAN,  J.  1988.  Experimental  study  of  predation  on 
eggs  of  ground-nesting  birds:  effects  of  habitat 
and  nest  distribution.  Condor  90:124-131. 
Rearden,  j.  D.  1951.  Identification  of  waterfowl  nest 
predators.  J.  Wildl.  Manage.  15:386-395. 

Roper,  J.  J.  1992.  Nest  predation  experiments  with 
quail  eggs:  too  much  to  swallow?  Oikos  65:528- 
530. 

Sargeant,  R.  a.,  j.  C.  Kilgo,  B.  R.  Chapman,  and  K. 
V.  Miller.  1998.  Predation  of  artificial  nests  in 
hardwood  fragments  enclosed  by  pine  and  agri- 
cultural habitats.  J.  Wildl.  Manage.  62: 1438-1442. 
SoKAL,  R.  R.  AND  E J.  Rohlf.  1995.  Biometry,  third 
ed.  W.  H.  Freeman  and  Company,  San  Francisco, 
California. 


540 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  4,  December  1999 


Storaas,  T 1988.  A comparison  of  losses  in  artificial 
and  naturally  occurring  Capercaillie  nests.  J. 
Wildl.  Manage.  52:123-126. 

Vander  Haegen,  W.  M.  and  R.  M.  DeGraaf.  1996. 
Predation  on  artificial  nests  in  forested  riparian 
buffer  strips.  J.  Wildl.  Manage.  60:542-550. 
Wilson,  G.  R.,  M.  C.  Brittingham,  and  L.  J.  Good- 
rich. 1998.  How  well  do  artificial  nests  estimate 
success  of  real  nests?  Condor  100:357-364. 
Yahner,  R.  H.  1991.  Avian  nesting  ecology  in  small 
even-aged  aspen  stands.  J.  Wildl.  Manage.  55: 
155-159. 

Yahner,  R.  H.  1996.  Forest  fragmentation,  artificial 
nest  studies,  and  predator  abundance.  Conserv. 
Biol.  10:672-673. 

Yahner,  R.  H.  and  C.  G.  Mahan.  1996a.  Depredation 


of  artificial  ground  nests  in  a managed  forested 
landscape.  Conserv.  Biol.  10:285-288. 

Yahner,  R.  H.  and  C.  G.  Mahan.  1996b.  Effects  of 
egg  type  on  depredation  of  artificial  ground  nests. 
Wilson  Bull.  108:129-136. 

Yahner,  R.  H.  and  T.  E.  Morrell.  1991.  Depredation 
of  artificial  nests  in  irrigated  forests.  Wilson  Bull. 
103:113-117. 

Yahner,  R.  H.  and  B.  D.  Ross.  1995.  Distribution  and 
success  of  Wood  Thrush  nests  in  a managed  for- 
ested landscape.  Northeast  Wildl.  52:1-9. 

Yahner,  R.  H.  and  D.  P.  Scott.  1988.  Effects  of  forest 
fragmentation  on  depredation  of  artificial  avian 
nests.  J.  Wildl.  Manage.  52:158-161. 

Yahner,  R.  H.  and  A.  L.  Wright.  1985.  Depredation 
on  artificial  nests:  effects  of  edge  and  plot  age.  J. 
Wildl.  Manage.  49:508-513. 


Wilson  Bull.,  111(4),  1999,  pp.  541-549 


PREDATION  ON  ARTIFICIAL  NESTS  ALONG  THREE  EDGE  TYPES 
IN  A NORTH  CAROLINA  BOTTOMLAND  HARDWOOD  FOREST 

JAMES  F.  SARACCO'^  AND  JAIME  A.  COLLAZO' 


ABSTRACT. — Many  researchers  have  reported  high  rates  of  nest  predation  near  forest  edges.  However,  edges 
may  be  of  various  types  (e.g.,  interior  or  exterior,  abrupt  or  gradual),  which  may  not  always  result  in  elevated 
predation.  We  compared  predation  rates  on  artificial  arboreal  nests  along  three  types  of  edges  in  a bottomland 
forest  in  North  Carolina  during  the  1996  breeding  season.  Edge  types  were  forest— farm,  forest— river,  and  the 
transition  zone  between  the  two  dominant  forest  types  in  the  floodplain  (cypress-gum  swamps-natural  levees). 
We  tested  for  differences  in  predation  rates  using  two  egg  types:  Northern  Bobwhite  (Colinus  virginianus)  and 
clay  eggs.  Predation  rates  were  higher  (P  < 0.05)  along  forest-farm  edges  than  along  the  other  two  edges. 
Predation  rates  did  not  differ  between  forest-river  and  transition  zone  edges.  Patterns  of  predation  on  the  two 
egg  types  and  higher  avian  predator  abundance  on  forest-farm  edges  suggested  that  avian  predators  may  have 
exerted  more  predation  pressure  along  these  edges.  These  results  are  consistent  with  other  studies,  which  suggest 
that  encroachment  by  agriculture  into  forested  landscapes  may  negatively  affect  breeding  birds.  Our  findings 
also  suggest  that  not  all  edge  types  are  equivalent  in  terms  of  predation  rates.  This  is  important  in  assessing  the 
conservation  value  of  bottomland  forests,  which  may  contain  various  edge  types  resulting  from  natural  processes 
(e.g.,  hydrodynamics).  Received  19  Feh.  1999,  accepted  6 July  1999. 


Predation  is  the  primary  cause  of  nest  loss 
for  a wide  range  of  passerine  birds  (Martin 
1992)  and  may  be  the  most  important  factor 
affecting  their  population  dynamics  (Temple 
and  Cary  1988).  Forest  birds  nesting  in  highly 
fragmented  landscapes  or  near  edges  may  ex- 
perience higher  rates  of  nest  predation  than 
birds  nesting  in  contiguous  forests  (Paton 
1994,  Andren  1995,  Robinson  et  al.  1995). 
However,  forest  edges  occur  in  a variety  of 
contexts  which  may  not  always  lead  to  in- 
creased predation  levels.  For  example,  edges 
may  be  in  the  interior  (e.g.,  clearcuts  within 
contiguous  forest)  or  along  the  exterior  (e.g., 
agricultural  encroachment  from  outside)  of 
forests  and  they  exhibit  varying  degrees  of 
contrast  from  subtle  to  abrupt  (Ratti  and  Reese 
1988,  Yahner  et  al.  1989,  Hawrot  and  Niemi 
1996,  Fenske-Crawford  and  Niemi  1997, 
Suarez  et  al.  1997).  Most  researchers  reporting 
high  predation  rates  near  edges  have  exam- 
ined abrupt  exterior  edges  (reviewed  by  An- 
dren 1995).  Those  that  have  considered  inte- 
rior and  more  subtle  edges  have  reported  less 
consistent  results  (e.g.,  Ratti  and  Reese  1988, 
Yahner  et  al.  1989,  Fenske-Crawford  and  Nie- 


‘ North  Carolina  Cooperative  Fish  and  Wildlife  Re- 
search Unit,  Biological  Resources  Division,  U.S.  Geo- 
logical Survey,  Dept,  of  Zoology,  North  Carolina  State 
Univ.,  Raleigh,  NC  27695-7617. 

^Corresponding  author;  E-mail:  jfsaracc(§>unity.ncsu. 
edu 


mi  1997,  Suarez  et  al.  1997).  Further  inves- 
tigation into  the  characteristics  of  edges  that 
influence  levels  of  predation  is  clearly  needed. 
Such  information  could  be  used  to  assess  the 
conservation  value  of  complex  landscapes, 
such  as  bottomland  hardwood  forests  that  sup- 
port diverse  breeding  bird  communities  (e.g., 
Wharton  et  al.  1981,  Mitchell  and  Lancia 
1990,  Mitchell  et  al.  1991,  Pashley  and  Bar- 
row  1992).  These  forested  wetlands  may  con- 
tain a variety  of  edge  types  that  result  from 
the  patchwork  of  plant  communities  whose  ar- 
rangement is  influenced  by  site-specific  hy- 
drodynamics and  sediment  deposition  rates 
along  floodplains  (Wharton  et  al.  1982). 

We  compared  predation  rates  on  artificial 
arboreal  nests  among  three  edge  types  in  bot- 
tomland hardwood  forests  along  the  Roanoke 
River  in  North  Carolina.  The  three  edge  types 
were;  (1)  forest-farm  edge  (an  abrupt  exterior 
edge),  (2)  forest-river  edge  (an  abrupt  interior 
edge),  and  (3)  levee-swamp  edge  (a  gradual 
interior  edge  where  the  two  dominant  plant 
communities  in  the  floodplain  meet).  We  used 
artificial  nests  primarily  because  of  the  logis- 
tic and  experimental  advantages  afforded  by 
their  use.  We  do  not  claim  that  predation  rates 
on  artificial  nests  represent  those  experienced 
by  natural  nests,  only  that  the  pattern  of  pre- 
dation among  edge  types  are  likely  to  be  sim- 
ilar for  the  two.  For  example,  the  few  studies 
that  have  compared  patterns  of  predation 
among  habitats  using  both  artificial  and  nat- 


541 


542 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


ural  nests  as  well  as  studies  comparing  similar 
habitats  using  either  of  these  methods  have 
typically  found  a close  match  in  predation  pat- 
terns for  the  two  nest  types  (see  Andren 
1995).  To  our  knowledge,  this  is  the  first  study 
to  examine  predation  rates  at  an  edge  between 
two  relatively  undisturbed  forested  plant  com- 
munities and  only  the  third  to  examine  a forest 
edge  abutting  water  (Bollinger  and  Peak  1995, 
Vander  Hagen  and  DeGraff  1996).  Differenc- 
es in  predator  communities  among  the  three 
edge  types  were  assessed  by  comparing  pat- 
terns of  predation  on  two  egg  types  and  the 
abundance  of  likely  avian  nest  predators. 

STUDY  AREA  AND  METHODS 

This  study  was  conducted  within  a contiguous  forest 
corridor  along  the  lower  Roanoke  River  between  the 
towns  of  Palmyra  and  Jamesville,  North  Carolina 
(36°  9'  N to  35°  50'  N,  77°  20'  W to  76°  53'  W).  The 
forested  areas  we  studied  have  been  undisturbed  for 
more  than  60  years.  Loss  and  alteration  of  forests  in 
the  floodplain  have  come  primarily  from  crop  (e.g., 
peanuts,  cotton,  wheat)  and  timber  production.  The 
lower  Roanoke  ecosystem  is  comprised  of  20  vegeta- 
tive community  types  (Schafale  and  Weakley  1990),  2 
of  which  are  clearly  dominant:  cypress-gum  swamp 
and  coastal  plain  levee  forests  (hereafter  swamps  and 
levees,  respectively).  Swamps  are  flooded  for  extended 
periods  throughout  the  year.  The  dominant  canopy  spe- 
cies are  water  tupelo  (Nyssa  aquatica)  and  bald  cy- 
press (Taxodium  distichum)-,  Carolina  ash  (Fraxinu.s 
caroliniana)  is  common  in  the  understory  (Lynch  et 
al.  1994).  Levees  occur  at  slightly  higher  elevations 
and  are  comprised  of  a diverse  mixture  of  canopy  spe- 
cies including  American  elm  (Ulmu.s  americana), 
green  ash  {Frcixinu.s  penn.sylvanica),  hackberry  (Celtis 
laevigata),  boxelder  (Acer  negundo),  water  hickory 
(Carya  aquatica),  and  sweetgum  (Liquidamhar  .styra- 
ciflua).  The  understory  of  levees  is  characterized  by 
pawpaw  (A.simina  triloba),  ironwood  (Carpinu.s  caro- 
liniana), and  various  vines  (Lynch  et  al.  1994).  Al- 
though the  sizes  and  shapes  of  patches  of  the  two  for- 
est types  are  variable,  levees  generally  occur  as  linear 
patches  close  to  the  river  channel  formed  by  the  de- 
position of  sediment  following  flooding  events.  Farther 
from  the  river  channel,  these  forests  grade  into 
swamps.  Levee-swamp  edges  are  comprised  of  a mix- 
ture of  species  typical  of  the  two  forest  types.  Forest- 
river  edges  are  compri.sed  of  species  typical  of  levees. 
Forest-farm  edges  are  dominated  by  swamp  trees,  with 
red  maple  (Acer  rubrum)  also  a dominant  species. 

Artificial  nests  were  placed  along  two  1.5  km  tran- 
sects established  within  each  edge  type.  Survey  tape 
was  u.sed  to  mark  59  25-m  intervals  (nest  site  locations 
1-60)  along  each  transect.  Nests  were  placed  at  50  m 
intervals  beginning  at  the  first  survey  flag  (nest  site  1) 
during  trials  one  and  three  and  beginning  at  the  second 
survey  flag  (nest  site  2)  during  trial  two.  Thus,  30  nests 


were  placed  along  each  transect  during  each  trial.  All 
nests  were  placed  on  a suitable  substrate  within  15  m 
of  the  survey  flag.  Transects  along  forest-river  and  for- 
est-farm edges  ran  parallel  to  the  river  and  fields,  re- 
spectively, and  were  approximately  15  m inside  the 
forest.  Levee-swamp  edge  transects  ran  along  the  es- 
timated center  of  the  levee-swamp  transition  zone. 
Transition  zones  were  characterized  by  the  presence  of 
bald  cypress  and  water  tupelo,  wetter  soils  (often  with 
standing  water),  and  a noticeable  opening  of  the  un- 
derstory. This  design  resulted  in  all  nests  being  within 
30  m of  a habitat  boundary.  Paton  (1994)  found  that 
edge  effects  on  nest  predation  are  typically  found  with- 
in 50  m of  a habitat  boundary;  the  30  m distance  cutoff 
we  used  was  well  within  this  range.  All  transects  were 
separated  by  at  least  2 km  and  were  at  least  100  m 
from  any  other  edge  type. 

Because  predators  may  respond  to  artificial  nests 
differently  than  to  natural  nests  (Major  and  Kendal 
1996),  we  attempted  to  mimic  as  closely  as  possible 
the  size,  color,  and  locations  of  nests  of  Acadian  Fly- 
catchers (Empidonax  virescens),  a common  breeding 
species  in  the  floodplain.  Several  other  common  breed- 
ing species  place  their  nests  in  similar  locations  (Lynch 
et  al.  1994).  Artificial  nests  were  constructed  from 
commercially  available  miniature  grape  vine  wreaths 
(approximately  8 cm  outside  and  5 cm  inside  diame- 
ters) with  bottoms  of  dried  grass  or  leaves  lining  wire 
mesh  frames  (approximately  4 cm  deep).  Nests  were 
attached  with  wire  to  the  fork  of  a low  hanging  tree 
branch,  sapling,  or  shrub  at  a height  of  approximately 
2.5  m. 

Three  15  day  trials  were  run  over  the  course  of  the 
1996  nesting  season  (30  May-24  July).  Fifteen  days 
approximates  a typical  incubation  period  for  open- 
nesting passerines  in  the  area.  Two  egg  types  were 
placed  in  each  nest:  one  Northern  Bobwhite  egg  (Col- 
inu.s  virginianus)  and  one  smaller  white  clay  egg 
(“Plastalina”,  Van  Aken  International;  approximately 
20  X 10  mm)  to  account  for  potential  biases  associated 
with  egg  type  (Roper  1992;  Haskell  1995a,  b;  Major 
and  Kendal  1996).  Eggs  were  placed  in  each  nest  3-5 
days  after  nests  were  placed  in  the  field.  This  was  in- 
tended to  mimic  the  interval  between  nest  building  and 
egg  laying  (Marini  et  al.  1995).  We  minimized  human 
scent  at  nest  sites  by  wearing  rubber  boots  and  gloves 
while  placing  nests  and  eggs,  and  while  checking  nests 
(Nol  and  Brooks  1982).  Nests  were  checked  for  signs 
of  predation  on  three  occasions  during  each  trial  (day 
5,  10,  and  15).  We  considered  a nest  to  be  depredated 
if  either  egg  was  damaged  or  missing.  Predation  was 
attributed  to  a bird  if  the  clay  egg  was  found  with  bill 
imprints  and/or  the  bobwhite  egg  was  found  with 
punctures  suggestive  of  a bill  (e.g.,  as  described  for 
crows  by  Rearden  1951).  We  considered  a nest  to  be 
depredated  by  a large  mouthed  mammal  if  bobwhite 
eggs  were  found  half  eaten  from  one  end  [suggesting 
raccoon,  Procyon  lotor  (Rearden  1951),  or  gray  squir- 
rel, Sciurus  carolinen.sis  (C.  J.  Whelan,  pers.  comm.)], 
if  chewed  up  clay  eggs  were  found,  or  if  nests  were 
destroyed  (e.g.,  nest  ring  gone  or  pulled  apart;  Best 


Saracco  and  Collazo  • ARTIFICIAL  NEST  PREDATION  AT  EDGES 


543 


and  Stauffer  1980).  Nests  for  which  tooth  imprints  or 
scratches  were  found  on  clay  eggs,  or  for  which  both 
eggs  were  found  still  in  the  nest  or  in  the  immediate 
vicinity  and  the  clay  egg  was  scratched,  were  consid- 
ered to  have  been  depredated  by  small-mouthed  mam- 
mals (e.g.,  Peromysciis  mice;  Major  1991,  Haskell 
1995b).  Although  snakes  may  have  also  contributed  to 
predation,  we  were  unable  to  attribute  predation  events 
to  snakes  based  on  evidence  at  nest  sites. 

Birds  were  censu.sed  at  10  count  stations  located  at 
150  m intervals  along  each  transect.  All  birds  seen  or 
heard  within  a 50  m radius  and  more  than  50  m but 
within  the  area  of  interest,  over  a 10  min  interval  were 
recorded  (Hutto  et  al.  1986).  One  census  was  con- 
ducted along  each  transect  during  the  morning  hours 
(06:30-09:45  EST)  between  20  May  and  6 June. 
Abundance  of  species  likely  to  depredate  nests  [Amer- 
ican Crow  (Con>us  hrachyrhynchos).  Fish  Crow  (Cor- 
vus  ossfragus).  Blue  Jay  (Cyanocilta  chstata)  and 
Common  Crackle  (Quiscalus  quiscula)]  was  expressed 
as  total  detections  per  point  (i.e.,  all  detections,  un- 
bounded radius).  Although  some  independence  among 
sampling  stations  may  have  been  sacrificed  by  using 
detections  at  all  distances,  each  of  these  were  “high- 
detection-ratio”  species  (i.e.,  each  had  a high  propor- 
tion of  the  total  detections  recorded  outside  of  the  50 
m radius  count  circle),  suggesting  that  detections  at  all 
distances  within  the  edge  and  immediately  adjacent 
habitats  were  more  appropriate  for  comparisons  (Hutto 
et  al.  1986). 

Univariate  repeated  measures  ANOVA  was  used  to 
test  for  differences  in  predation  rates  among  edge  types 
(Proc  GEM,  SAS  Institute  1990;  Winer  et  al.  1991). 
The  response  variable  was  the  proportion  of  the  30 
nests  depredated  on  each  transect.  The  independent 
variable  was  edge  type  (forest— farm,  levee-swamp, 
and  forest-river);  trial  (1,  2,  and  3)  and  days  of  ex- 
posure (5,  10,  and  15)  were  repeated  measures.  Prior 
to  analyses  the  response  variable  was  square  root-arc- 
sine  transformed  to  meet  homogeneity  of  variance  as- 
sumption (Levene’s  test:  P > 0.05;  JMP,  SAS  Institute 
1994).  In  order  to  test  for  potential  biases  associated 
with  egg  type,  we  used  McNemar’s  tests  conducted 
separately  for  each  edge  type  (Proc  FREQ,  SAS  Insti- 
tute 1990).  We  tested  for  differences  in  selected  avian 
predator  abundance  among  edge  types  using  nested- 
ANOVA  (Proc  NESTED,  SAS  Institute  1990).  The  re- 
sponse variable  was  the  number  of  detections  of  se- 
lected avian  predators  per  point.  Model  terms  were 
edge  type  and  transect  [edge  type].  Data  met  homo- 
geneity of  variance  assumption  (Levene’s  test;  P > 
0.05;  JMP,  SAS  Institute  1994).  Differences  in  abun- 
dance of  individual  species  of  avian  predators  among 
edge  types  were  assessed  using  Kruskal-Wallis  tests 
(Proc  NPARIWAY,  SAS  Institute  1990).  For  species 
where  a significant  edge  effect  was  found,  a posteriori 
contrasts  were  computed  using  the  nonparametric  all- 
treatments multiple  contrast  test  described  in  Hollander 
and  Wolfe  (1999).  An  a < 0.05  was  used  for  all  an- 
alyses and  values  presented  are  means  ± SE.  Statistical 


analyses  were  performed  with  JMP  (version  3.2.2)  and 
SAS  (version  7.0)  for  Windows. 

RESULTS 

Predation  rates  differed  significantly  among 
edge  types  (F  = 11.33,  df  = 2,  3;  P = 0.04) 
and  were  higher  along  the  agricultural  field- 
forest  edges  than  along  the  other  two  edge 
types  (F  — 22.31,  df  = 1,  3;  P = 0.01;  Fig. 
1).  Predation  rates  did  not  differ  between  for- 
est-river and  levee-swamp  edges  (F  = 0.35, 
df  = 1,  3;  P > 0.05).  There  was  no  difference 
in  predation  rate  among  trials  (P  = 0.05,  df 
= 2,  6;  P > 0.05).  Within  trials,  predation  rate 
increased  with  day  of  exposure  (F  = 94.54, 
df  = 2,  6;  P < 0.001).  Interaction  between 
day  of  exposure  and  edge  type  was  nearly  sig- 
nificant (P  = 3.55,  df  = 4,  6;  P = 0.08).  This 
nearly  significant  interaction  was  likely 
caused  by  differences  in  response  pattern 
(slope)  between  levee-swamp  and  forest-river 
edges  from  5-10  days  of  exposure  (Fig.  1). 
The  difference  among  these  two  edges  at  5 
days  of  exposure  was  not  significant  (P  = 
5.23,  df  = 1,3;  P > 0.05).  Predation  rates 
were  highest  along  forest-farm  edges  regard- 
less of  exposure  time. 

The  number  of  nests  for  which  the  bob- 
white  egg  was  damaged  or  missing  was  high- 
est on  forest-farm  edges,  while  the  number  of 
nests  for  which  only  the  clay  egg  was  depre- 
dated was  similar  among  edge  types  (Fig.  2A). 
Bobwhite  eggs  were  preyed  upon  more  fre- 
quently on  forest-farm  edges  than  on  the  other 
two  edge  types  (Fig.  2B).  Conversely,  the  per- 
centage of  depredated  nests  in  which  only  the 
clay  egg  was  preyed  upon  was  lowest  on  for- 
est-farm edges  and  highest  on  levee-swamp 
edges.  For  each  edge  type,  the  clay  egg  was 
depredated  significantly  more  often  than  the 
bobwhite  egg  in  nests  where  only  one  egg  was 
depredated  (Forest-farm:  = 9.49,  df  = 1; 

P < 0.01;  Forest-river:  G,,jj  = 19.15,  df  = 1; 
P < 0.001;  Levee-swamp:  G,,jjj  = 45.83,  df  = 
1;  P < 0.001).  Despite  this  egg  type  bias,  the 
pattern  of  predation,  higher  on  forest-farm 
edges  than  on  the  other  two  edge  types,  was 
the  same  regardless  of  whether  predation  was 
on  bobwhite  or  clay  eggs  (Fig.  2A). 

We  identified  nest  predators  for  30%  of 
depredated  nests  (114/368).  Of  these,  69% 
(79)  were  birds,  22%  (25)  were  smaller 
mouthed  mammals,  and  9%  (10)  were  larger 


544 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


100 
90 
80 
70 

c 60 

o 

TO 

"S  50 

k_ 

Q. 

^ 40 

30 
20 
10 
0 

LIG.  1.  Predation  rates  on  artificial  nests  at  three  edge  types  in  bottomland  hardwood  forest  along  the 

Roanoke  River  floodplain.  North  Carolina  during  the  1996  breeding  season.  Predation  rates  were  significantly 

higher  along  the  forest-farm  edges  than  on  the  other  two  edges  {P  < 0.05). 


5 10  15 

Days  of  exposure 


mouthed  mammals.  Although  measurements 
of  bill  imprints  in  clay  eggs  were  not  taken, 
the  size  and  shape  of  these  imprints  suggested 
that  crows,  Blue  Jays,  and  Common  Crackles 
were  among  the  avian  predators.  The  abun- 
dance of  these  nest  predators  differed  by  edge 
type  (F  = 36.84,  df  = 2,  3;  F < 0.01)  and 
was  higher  on  forest-farm  edges  than  along 
forest-river  and  levee-swamp  edges  (F  = 
65.79,  df  = 1,  3;  F < 0.01;  Fig.  3).  Avian 
predator  abundance  did  not  differ  significantly 
between  forest-river  and  levee-swamp  edges 
but  tended  to  be  higher  along  the  forest-river 
edge  (F  = 7.89,  df  = 1,  3;  F = 0.07;  Fig.  3). 
Considered  individually,  the  four  predator 
species  were  not  consistent  in  their  responses 
to  edge  type  (Fig.  3).  The  numbers  of  Amer- 
ican and  Fish  Crows  detected  differed  signif- 
icantly among  edge  types  (American  Crow; 

= 1 1.21,  df  = 2,  F < 0.01;  Fish  Crow:  = 

12.27,  df  = 2;  F < 0.01),  and  both  of  these 
species  were  significantly  more  abundant 
along  forest-farm  edges  than  along  levee- 
swamp  edges  (American  Crow  q = 4.28,  Fish 


Crow  q = 4.42;  F < 0.01).  American  Crows 
were  also  significantly  more  abundant  on  for- 
est-farm edges  than  along  forest-river  edges 
{q  — 3.65;  F < 0.05).  Fish  Crows  tended  to 
be  more  abundant  along  forest-farm  edges 
than  along  forest-river  edges,  although  this 
difference  was  not  significant  {q  = 3.06;  0.05 
< F < 0.10).  In  contrast.  Blue  Jays,  which 
also  showed  a significant  edge  effect  (x^  = 
6.05,  df  = 2;  F < 0.05),  were  more  abundant 
along  forest-river  edges  than  on  levee-swamp 
edges  {q  = 3.33;  F < 0.05).  Contrasts  be- 
tween forest-farm  edges  and  the  other  two 
edge  types  for  this  species  were  not  significant 
(F  > 0.05).  Common  Crackle  abundance  did 
not  differ  significantly  among  edge  types  (x^ 
= 3.74,  df  = 2;  F > 0.05). 

DISCUSSION 

Our  findings  are  consistent  with  most  pre- 
vious studies  that  have  reported  high  rates  of 
nest  predation  along  abrupt  exterior  edges 
(our  forest-farm  edge  type;  see  Andren  1995). 
The  presence  of  more  avian  predators  along 


% of  depreciated  nests  Number  depredated 


Saracco  and  Collazo  • ARTIFICIAL  NEST  PREDATION  AT  EDGES 


545 


30 


25 


20 


15 

10 


5 

0 

100  T 

90 
80 
70  - 
60 
50 
40 
30 
20 
10 
0 

Forest-farm  Forest-river  Swamp-levee 

FIG.  2.  (A)  Absolute  number  of  depredated  nests  and  (B)  the  percentage  of  depredated  nests  for  which  the 

bobwhite  egg,  clay  egg,  or  only  the  clay  egg  was  preyed  upon.  Re.sults  of  McNemar's  tests  conducted  for  each 
edge  type  (/?  = 180  nests)  .suggested  a significant  egg  type  bias  (P  < 0.01  for  each  edge  type);  however,  the 
same  pattern  of  predation  (highest  on  forest-farm  edges)  was  apparent  regardless  of  whether  bobwhite  or  clay 
eggs  were  considered. 


Forest-farm  Forest-river  Swamp-levee 


546 


THE  W ILSON  BULLETIN  • Vol.  III.  No.  4.  December  1994 


3 


2.5 


2 


o '•■5 
O 


1 


0.5 


0 

FIG.  3.  Mean  number  of  detections  per  point  for  selected  avian  nest  predators  at  three  edge  types  in  a 
bottomland  hardwood  forest  along  the  Roanoke  River  floodplain.  North  Carolina.  The  pooled  abundance  for  all 
species  was  significantly  higher  on  forest-farm  edges  than  on  the  other  two  edge  types  (F  < 0.01);  the  responses 
of  individual  species  were  not  consistent  among  edge  types. 


IIIIIIIIP 


■ All  Species 
e American  Crow 
ID  Fish  Crow 
^ Blue  Jay 
0 Common  Crackle 


m 


m. 


Forest-farm 


Forest-river 


Swamp-levee 


forest-farm  edges  may  have  contributed  to  the 
higher  predation  rates  we  observed  along 
these  edges.  Marini  and  coworkers  (1995) 
found  a positive  correlation  between  avian 
predator  abundance  (American  Crows.  Blue 
Jays,  and  Common  Crackles)  and  predation 
levels  on  artificial  nests  in  forest  saplings,  as 
well  as  significantly  higher  predation  rates  on 
these  nests  at  forest-fimn  edges.  Other  re- 
searchers have  also  related  avian  nest  predator 
abundance  (e.g..  corvids)  to  forest-farm  edges 
or  shown  predation  on  artificial  nests  by  these 
predators  to  be  higher  near  such  edges  (e.g.. 
Whitcomb  et  al.  1981.  Angelstam  1986.  An- 
dren  1992.  Nour  et  al.  1993). 

Our  finding  that  nests  at  forest-river  edges 
experienced  lower  predation  rates  than  forest- 
farm  edges  is  in  accordance  w ith  Vander  Hae- 
gen  and  DeGraff  ( 1 986)  w ho  found  no  effect 
of  distance  from  a river  edge  on  predation 
rate.  In  contrast.  Bollinger  and  Peak  (1995) 
found  predation  rates  to  be  uniformly  high  on 
artificial  ground  nests  along  a forest  edge  bor- 


dering water  and  a forest-fami  edge  in  one 
forest  fragment  in  an  agricultural  setting. 
Small  forest  fragments  in  agricultural  land- 
scapes such  as  this  may  become  inundated 
w ith  certain  mammalian  predator  species  (e.g.. 
raccoons;  gray  squirrels.  Sciiirus  carolinensis: 
and  opossums.  Didelphis  marsupiali.s',  Bider 
1968.  Matthiae  and  Steams  1981).  The  rela- 
tively w ide  and  heavily  forested  river  corridor 
in  our  study  may  have  alleviated  any  such 
packing  effects  by  predators. 

The  levee-swamp  edges  we  studied  are 
unique  in  that  they  ivre  naturally  occurring 
boundaries  between  plant  communities  rather 
than  edges  resulting  from  human  activities 
(e.g..  agriculture,  forestry  practices).  As  such, 
they  may  not  be  perceived  as  edges  by  some 
predators  that  may  move  freely  between  le- 
vees and  swamps  rather  than  concentrating  ac- 
tivities along  the  edge  or  using  it  as  a travel 
lane  (Bider  1968.  Chasko  and  Gates  1982). 
This  could  explain  the  relatively  low  preda- 
tion rates  we  observed  at  these  edges. 


Sarcuco  ami  Collazo  • AR'I  Il-ICIAL  NEtST  RKHDA  I'ION  A I lilKiliS 


547 


It  is  difficult  to  determine  the  relative  im- 
pact of  different  predators  at  the  three  edge 
types  because  predators  were  only  identified 
for  30%  of  depredated  nests.  The  greater 
abundance  of  avian  predators  at  forest-farm 
edges  may  have  contributed  to  the  higher  pre- 
dation rates  there;  however,  some  mammalian 
predators  and  snakes  might  also  be  abundant 
and  concentrate  their  activities  or  travel  along 
abrupt  edges  (Bider  1968,  Chasko  and  Gates 
1982,  Durner  and  Gates  1993;  but  see  Heske 
1995).  Unfortunately,  we  were  unable  to  as- 
sess the  relative  abundance  of  non-avian  pred- 
ators or  their  relative  contribution  to  predation 
on  artificial  nests.  Nonetheless,  there  was 
some  indication  that  the  predators  responsible 
for  depredating  nests  may  have  differed 
among  edge  types.  For  example,  because 
small  mouthed  predators  may  have  been  un- 
able to  damage  the  Northern  Bobwhite  eggs 
[as  has  been  reported  for  Japanese  Quail  {Co- 
turnix  coturnix)  eggs;  Roper  1992,  Haskell 
1995aJ,  our  finding  that  the  proportion  of 
nests  for  which  only  the  clay  egg  was  depre- 
dated was  higher  at  the  forest  interior  edges 
suggests  that  small  mouthed  predators  (e.g., 
mice)  may  have  been  more  important  at  these 
edges.  In  contrast,  both  the  proportion  and  ab- 
solute number  of  depredated  nests  for  which 
the  bobwhite  egg  was  preyed  upon  was  high- 
est at  forest-farm  edges.  This  supports  the 
contention  advanced  by  Haskell  (1995b)  and 
Nour  and  coworkers  (1993)  that  avian  and 
larger  mammalian  predators  increase  in  im- 
portance in  small  forest  patches  or  at  the  edg- 
es of  forests.  Smaller  mouthed  predators,  al- 
though possibly  more  frequent  at  the  interior 
edges,  appear  to  have  depredated  similar  pro- 
portions of  nests  along  the  three  edge  types. 
Finally,  differential  predation  rates  at  different 
types  of  edges  could  also  be  influenced  by 
factors  other  than  the  types  of  predators  in- 
volved and  their  abundance.  Future  studies 
should  be  designed  to  consider  factors  influ- 
encing nest  site  selection  (e.g.,  number  of  po- 
tential nest  sites)  and  nest  densities  of  avian 
community  members  (Martin  1993). 

Bottomland  hardwood  forests  of  the  south- 
eastern U.S.  are  being  destroyed  and  frag- 
mented at  high  rates  (Turner  et  al.  1981,  Ab- 
ernathy and  Turner  1987).  These  areas  provide 
important  breeding  habitats  for  many  migra- 
tory and  resident  birds  (Wharton  et  al.  1981, 


Mitchell  and  Lancia  1990,  Mitchell  et  al. 
1991,  Pashley  and  Barrow  1992).  Understand- 
ing how  edges  resulting  from  natural  process- 
es (e.g.,  hydrodynamics),  as  well  as  from  an- 
thropogenic modifications,  affect  breeding 
bird  communities  is  important  to  their  conser- 
vation and  management.  Our  results  suggest 
that  encroachment  by  agriculture  may  nega- 
tively affect  breeding  birds  through  higher 
predation  rates  along  forest-farm  edges.  Nat- 
ural edges  between  adjacent  plant  communi- 
ties and  at  the  forest-river  interface  may  not 
affect  breeding  birds  in  the  same  way. 

ACKNOWLEDGMENTS 

We  thank  L.  Peoples  and  E.  Sandlin  for  assistance 
in  the  field.  We  also  thank  The  Nature  Conservancy, 
the  North  Carolina  Wildlife  Resources  Commission, 
U.S.  Fi.sh  and  Wildlife  Service,  and  private  land  own- 
ers for  granting  permission  to  work  on  and  facilitating 
access  to  their  land.  C.  Parkhurst  provided  Northern 
Bobwhite  eggs.  K.  Pollock  provided  advise  regarding 
the  design  of  the  experiment,  and  G.  Brown  provided 
assistance  with  statistical  analyses.  This  project  was 
supported  by  the  North  Carolina  Chapter  of  The  Na- 
ture Conservancy,  North  Carolina  State  University, 
and  Biological  Resources  Division,  U.S.  Geological 
Survey.  We  thank  M.  Groom,  D.  Haskell,  J.  Lyons,  T. 
Simons,  J.  Walters,  C.  Whelan,  and  two  anonymous 
reviewers  for  comments  on  earlier  versions  of  this 
manuscript. 

LITERATURE  CITED 

Abernathy,  Y.  and  R.  Turner.  1987.  U.S.  forested 
wetlands:  status  and  changes  1940-1980.  Bio- 
Science  37:721—727. 

Andren,  H.  1992.  Corvid  density  and  nest  predation 
in  relation  to  forest  fragmentation:  a landscape 
perspective.  Ecology  73:794—804. 

Andren,  H.  1995.  Effects  of  landscape  composition  on 
predation  rates  at  habitat  edges.  Pp.  225-255  in 
Mosaic  landscapes  and  ecological  processes  (L. 
Hansson,  L.  Fahrig,  and  G.  Merriam.  Eds.).  Chap- 
man and  Hall,  London,  U.K. 

Angelstam,  P.  1986.  Predation  on  ground-nesting 
birds  nests  in  relation  to  predator  densities  and 
habitat  edge.  Oikos  47:365-373. 

Be.st,  L.  B.  and  D.  E Stauffer.  1980.  Factors  affect- 
ing nesting  success  in  riparian  bird  communities. 
Condor  82:149-158. 

Bider,  J.  R.  1968.  Animal  activity  in  uncontrolled  ter- 
restrial communities  as  determined  by  a sand  tran- 
sect technique.  Ecol.  Monogr.  38:269-308. 
Bollinger,  E.  K.  and  R.  G.  Peak.  1995.  Depredation 
of  artificial  avian  nests:  a comparison  of  forest- 
field  and  forest-lake  edges.  Am.  Midi.  Nat.  134: 
200-203. 

Chasko,  G.  G.  and  J.  E.  Gates.  1982.  Avian  habitat 


548 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  4,  December  1999 


suitability  along  a transmission-line  corridor  in  an 
oak-hickory  forest  region.  Wildl.  Monogr.  82:1- 
41. 

Durner,  G.  M.  and  J.  E.  Gates.  1993.  Spatial  ecology 
of  black  rat  snakes  on  Remington  Farms,  Mary- 
land. J.  Wildl.  Manage.  57:812-826. 

Fenske-Crawford,  T.  J.  and  G.  J.  Niemi.  1997.  Pre- 
dation of  artificial  ground  nests  at  two  types  of 
edges  in  a forest-dominated  landscape.  Condor  99; 
14-24. 

Haskell,  D.  G.  1995a.  Forest  fragmentation  and  nest 
predation;  are  experiments  with  Japanese  Quail 
eggs  misleading?  Auk  112:767-770. 

Haskell,  D.  G.  1995b.  A reevaluation  of  the  effects 
of  forest  fragmentation  on  rates  of  bird-nest  pre- 
dation. Conserv.  Biol.  9:1316-1318. 

Hawrot,  R.  Y.  and  G.  j.  Niemi.  1996.  Effects  of  edge 
type  and  patch  shape  on  avian  communities  in  a 
mixed  conifer-hardwood  forest.  Auk  113:586- 
598. 

Heske,  E.  j.  1995.  Mammalian  abundances  on  forest- 
farm  edges  versus  forest  interiors  in  southern  Il- 
linois; is  there  an  edge  effect?  J.  Mamm.  76:562- 
568. 

Hollander,  M.  and  D.  A.  Wolfe.  1999.  Nonpara- 
metric  statistical  methods.  Second  ed.  John  Wiley 
and  Sons,  Inc.,  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. 

Lynch,  J.  M.,  J.  A.  Collazo,  and  J.  R.  Walters. 
1994.  Breeding  birds  of  the  lower  Roanoke  River 
floodplain — A land  managers  guide.  The  Nature 
Conservancy,  Durham,  North  Carolina. 

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. 

Marini,  M.  A.,  S.  K.  Robinson,  and  E.  J.  Heske.  1995. 
Edge  effects  on  nest  predation  in  the  Shawnee  Na- 
tional Forest,  Southern  Illinois.  Biol.  Conserv.  74: 
203-213. 

Martin,  T.  E.  1992.  Breeding  productivity  consider- 
ations; what  are  the  appropriate  habitat  features 
for  management?  Pp.  455—473  in  Ecology  and 
conservation  of  Neotropical  migrant  landbirds  (J. 
M.  Hagan  and  D.  W.  Johnston,  Eds.).  Smithsonian 
Institution  Press,  Washington,  D.C. 

Martin,  T.  E.  1993.  Nest  predation  and  nest  sites:  new 
perspectives  on  old  patterns.  BioScience  43:523- 
532. 

Matthiae,  P.  E.  and  E Stearns.  1981.  Mammals  in 
forest  islands  in  .southeastern  Wisconsin.  Pp.  55- 
66  in  Forest  island  dynamics  in  man-dominated 
landscapes  (R.  L.  Burgess  and  D.  M.  Sharpe. 
Eds.).  Springer- Verlag,  New  York. 

Mitchell,  L.  J.  and  R.  A.  Lancia.  1990.  Breeding  bird 


community  changes  in  bald-cypress-tupelo  wet- 
land following  timber  harvesting.  Proc.  Annu. 
Conf.  Southeast.  Assoc.  Fish  Wildl.  Agencies  44: 
189-201. 

Mitchell,  L.  J.,  R.  A.  Lancia,  R.  Lea,  and  S.  A. 
Gauthreaux,  Jr.  1991.  Effects  of  clearcutting  and 
natural  regeneration  of  breeding  bird  communities 
of  bald-cypress-tupelo  wetlands.  Pp.  155—161  in 
Proc.  Symp.  Wetlands  River  Corridor  Manage.  (J. 
Kusler  and  S.  Daly,  Eds.).  Association  of  Wetland 
Managers,  Inc.,  Berne,  New  York. 

Nol,  E.  and  R.  j.  Brooks.  1982.  Effects  of  predator 
exclosures  on  nesting  success  of  Killdeer.  J.  Field 
Ornithol.  53:263-268. 

Nour,  N.,  E.  Matthysen,  and  A.  A.  Dhondt.  1993. 
Artificial  nest  predation  and  habitat  fragmentation: 
different  trends  in  bird  and  mammal  predators. 
Ecography  16:111-116. 

Pashley,  D.  N.  and  W.  C.  Barrow.  1992.  Effects  of 
land  use  practices  on  Neotropical  migratory  birds 
in  bottomland  hardwood  forests.  Pp.  315-320  in 
Status  and  management  of  Neotropical  migratory 
birds  (D.  M.  Finch  and  P.  W.  Stangel,  Eds.). 
USDA  Forest  Service,  Rocky  Mountain  Forest 
and  Range  Experimental  Station.  Gen.  Tech.  Rep. 
RM-229.  Fort  Collins,  Colorado. 

Paton,  P.  W.  C.  1994.  The  effect  of  edge  on  avian  nest 
success:  how  strong  is  the  evidence?  Conserv. 
Biol.  8:17-26. 

Ratti,  j.  T.  and  Reese,  K.  P.  1988.  Preliminary  test  of 
the  ecological  trap  hypothesis.  J.  Wildl.  Manage. 
52:484-491. 

Rearden,  j.  D.  1951.  Identification  of  waterfowl  nest 
predators.  J.  Wildl.  Manage.  15:386-395. 

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. 

Roper,  J.  J.  1992.  Nest  predation  experiments  with 
quail  eggs;  too  much  to  swallow?  Oikos  65:528- 
530. 

SAS  Institute.  1990.  SAS/STAT  users  guide.  Version 
6,  fourth  ed.  SAS  Institute  Inc.,  Cary,  North  Car- 
olina. 

SAS  Institute.  1994.  JMP  statistics  and  graphics 
guide.  Version  3.0.  SAS  Institute  Inc.,  Cary,  North 
Carolina. 

SCHAFALE,  M.  P.  AND  A.  S.  WEAKLEY.  1990.  Classifi- 
cation of  the  natural  communities  of  North  Caro- 
lina: third  approximation.  North  Carolina  Natural 
Heritage  Program,  Raleigh,  North  Carolina. 

Small,  M.  E and  M.  L.  Hunter.  1988..  Forest  frag- 
mentation and  avian  nest  predation  in  forested 
landscapes.  Oecologia  76:62-64. 

Suarez,  A.  V.,  K.  S.  Pfennig,  and  S.  K.  Robinson. 
1997.  Nesting  success  of  a disturbance-dependent 
songbird  on  different  kinds  of  edges.  Conserv. 
Biol.  11:928-935. 

Temple,  S.  A.  and  J.  R.  Cary.  1988.  Modeling  dy- 
namics of  habitat-interior  bird  populations  in  frag- 
mented landscapes.  Conserv.  Biol.  2:340-347. 


Saracco  and  Collazo  • ARTIFICIAL  NEST  PREDATION  AT  EDGES 


549 


Turner,  R.  E.,  S.  Forsythe,  and  N.  Craig.  1981.  Bot- 
tomland hardwood  forest  land  resources  of  the 
southeastern  U.S.  Pp.  13-18  in  Wetlands  of  bot- 
tomland hardwood  forest  (J.  R.  Clark  and  J.  Ben- 
forado,  Eds.).  Elsevier  Scientific  Publishing  Co., 
New  York. 

Vander  Haegen,  W.  M.  and  R.  M.  DeGraaf.  1996. 
Predation  on  artificial  nests  in  forested  riparian 
buffer  strips.  J.  Wildl.  Manage.  60:542-550. 

Wharton,  C.  H.,  W.  M.  Kitchens,  F.  C.  Pendleton, 
AND  T.  W.  SiPE.  1982.  The  ecology  of  bottomland 
hardwood  swamps  of  the  Southeast:  a community 
profile.  FWS/OBS-81/37.  U.S.  Fish  and  Wildlife 
Service,  Biological  Services  Program,  Washing- 
ton, D.C. 

Wharton,  C.  H.,  V.  W.  Lambour,  J.  Newson,  P.  V. 
Winger,  L.  L.  Gaddy,  and  R.  Mancke.  1981.  The 
fauna  of  bottomland  hardwoods  of  the  southeast- 


ern United  States.  Pp.  87-160  in  Wetlands  of  bot- 
tomland hardwood  forest  (J.  R.  Clark  and  J.  Ben- 
forado,  Eds.).  Elsevier  Scientific  Publishing  Co., 
New  York. 

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.  Pp.  125-205  in 
Forest  island  dynamics  in  man-dominated  land- 
scapes (R.  L.  Burgess  and  D.  M.  Sharpe,  Eds.). 
Springer- Verlag,  New  York. 

Winer,  B.  J.,  D.  R.  Brown,  and  K.  M.  Michaels. 
1991.  Statistical  principles  in  experimental  design. 
McGraw-Hill,  New  York. 

Yahner,  R.  H.,  T.  E.  Morrell,  and  E.  S.  Rachael. 
1989.  Effects  of  edge  contrast  on  depredation  of 
artificial  avian  nests.  J.  Wildl.  Manage.  49:508- 
513. 


Wilson  Bull.,  111(4),  1999,  pp.  550-558 


THE  RESPONSE  OF  A KANSAS  WINTER  BIRD  COMMUNITY  TO 
WEATHER,  PHOTOPERIOD,  AND  YEAR 

MARTIN  A.  STAPANIAN,'  CHRISTOPHER  C.  SMITH, AND  ELMER  J.  FINCK^ 

ABSTRACT. — We  conducted  a bird  census  along  the  same  route  nearly  each  week  for  14  winters  (194 
censuses),  and  compared  the  mean  number  of  species  per  station  and  the  total  number  of  species  recorded  on 
the  census  with  the  length  of  photoperiod  and  weather  variables.  We  found  significant  differences  among  winters 
for  both  indicators  of  species  richness.  This  result  is  consistent  with  previous  studies  in  which  abundance  of 
food  was  measured  in  the  same  general  area.  Both  indicators  of  species  richness  were  negatively  associated  with 
the  number  of  days  after  1 November.  This  result  is  consistent  with  the  hypothesis  that  wintering  species 
dependent  on  nonrenewed  food  resources  lose  individuals  to  mortality  or  emigration.  Further,  there  was  a positive 
relationship  between  photoperiod  and  both  indicators  of  species  richness.  This  result  is  consistent  with  the 
hypothesis  that  the  detection  of  individuals  in  the  early  morning  hours  increases  with  the  amount  of  daylight 
they  have  available  for  foraging  and  social  behaviors.  Wind  speed  and  temperature  had  negative  and  positive 
relationships,  respectively,  to  species  richness.  The  number  of  species  per  station  was  greatest  on  days  when  the 
ground  was  covered  with  dew  and  least  on  days  when  snow  depth  was  more  than  15  cm.  When  the  “winters” 
were  divided  into  four  30-day  “quarters”,  most  of  the  61  species  were  recorded  with  equal  frequency  in  each 
quarter.  Eight  species  were  detected  less  frequently  at  the  end  of  winter  than  in  the  beginning.  Four  species 
exhibited  the  reverse  pattern.  Two  species  were  recorded  more  frequently  at  the  beginning  and  at  the  end  of  the 
winter  than  during  the  middle.  Temperature,  wind,  photoperiod,  successive  winter  day,  year,  and  species-specific 
evolutionary  history  all  affect  winter  bird  species  richness.  Received  1 Oct.  1998,  accepted  5 August  1999. 


Winter  is  a stressful  season  of  the  year  for 
endotherms  at  mid-  and  high  latitudes.  Severe 
cold,  short  photoperiod,  and  a mostly  nonre- 
newed food  supply  make  it  a challenge  to 
maintain  a constant  body  temperature.  Many 
bird  species  migrate  to  more  hospitable  cli- 
mates. For  those  species  that  overwinter  at 
higher  latitudes,  weather  conditions  have  been 
shown  to  affect  the  amount  of  body  fat  stored 
(White  and  West  1977,  Dawson  and  Marsh 
1986,  Peach  et  al.  1992,  Waite  1992,  Houston 
and  McNamara  1993,  Rogers  et  al.  1994,  Pi- 
lastro  et  al.  1995).  Collins  (1989)  provided  a 
short  review  on  some  of  the  major  physiolog- 
ical adaptations  in  birds  for  surviving  the  win- 
ter. Robbins  (1972,  1981a)  and  Altman  (1983) 
discussed  the  importance  of  weather  condi- 
tions on  winter  bird  populations. 

Although  detailed,  long-term  winter  studies 
exist  for  specific  species  (Loery  and  Nichols 


' Ohio  Cooperative  Fish  and  Wildlife  Re.search  Unit, 
The  Ohio  State  Univ.,  1735  Neil  Ave.,  Columbus,  OF! 
43210. 

2 Division  of  Biology,  Ackert  Hall,  Kansas  State 
Univ..  Manhattan,  KS  66506. 

’ Division  of  Biological  Sciences,  Box  4050,  Em- 
poria State  Univ..  Emporia,  KS  66801. 

■*  Present  address:  USGS/BRD.  Lake  Erie  Biological 
Station,  6100  Columbus  Ave.,  Sandusky,  OH  44870. 

' Corresponding  author;  E-mail:  cccsmith@ksu.edu 


1985),  data  for  studies  of  overwintering  bird 
communities  often  are  collected  only  for  a few 
days  per  year  (cf.,  Erskine  1992).  For  exam- 
ple, three  counts  per  year  are  made  for  the 
Finnish  winter  bird  census  routes  (Hilden 
1987),  and  the  Christmas  Bird  Count  is  an  an- 
nual one-day  count  of  an  area.  The  daily  ef- 
fects of  weather  components  and  the  annual 
effects  of  available  food  resources  (e.g.,  mast 
crop  failure)  on  the  number  of  species  in  an 
area  are  often  difficult  to  determine  or  are  sta- 
tistically confounded.  Instructions  for  the 
Winter  Bird-Population  Study  (Robbins 
1981b)  call  for  a minimum  of  six  visits  per 
site  per  year.  However,  daily  weather  data  for 
the  study  sites  and  analyses  of  the  effects  of 
weather  components  on  species  richness  are 
typically  lacking  (Robbins  1981a).  Further, 
depending  on  species-specific  responses  to 
abiotic  factors  and  food  abundance,  bird  spe- 
cies may  differ  in  their  detectability  during  the 
course  of  winter. 

We  analyzed  data  from  bird  censuses  con- 
ducted at  nearly  weekly  intervals  for  14  years 
(194  censuses)  along  the  same  route.  We  in- 
clude in  our  analyses  weather  data  collected 
from  a permanent  station  approximately  10 
km  from  the  route.  Our  objectives  were  to  (1) 
quantify  the  effects  of  weather  components, 
photoperiod,  and  the  cumulative  number  of 


550 


Suipcinian  el  al.  • WEATHER  EFFECTS  ON  WINTER  BIRDS 


551 


winter  days  on  species  richness;  and  (2)  to 
determine  if  the  frequency  of  detection  of  in- 
dividual species  changed  over  the  course  of 
the  winter.  To  accomplish  the  first  objective, 
we  used  a statistical  procedure  that  accounts 
for  the  correlation  structure  (i.e.,  time-depen- 
dency) between  censuses  taken  within  each 
winter.  We  accomplished  the  second  objective 
by  testing  the  null  hypothesis  that  individual 
species  were  recorded  with  equal  frequency 
within  each  of  four  30-day  intervals  during 
the  winter. 

METHODS 

Census  route  and  field  method. — Our  study,  like 
others  based  on  seeing  and/or  hearing  birds  to  count 
their  presence,  measured  the  visual  and  auditory  de- 
tectability of  birds.  Birds  were  counted  with  a modified 
Breeding  Bird  Survey  procedure  (Robbins  et  al.  1986) 
along  a regular  census  route  across  the  border  between 
Riley  and  Pottawatomie  counties,  Kansas  (Stapanian 
1982,  Stapanian  et  al.  1994).  The  route  consisted  of 
16  stations;  unlike  the  Breeding  Bird  Survey  routes, 
the  stations  were  not  separated  by  regular  0.81  km  in- 
tervals. Instead,  stations  were  selected  to  represent  typ- 
ical upland  and  riparian  forest  habitats  with  some  tree 
species  bearing  fleshy,  bird-dispersed  fruit  in  propor- 
tion to  their  presence  in  the  Kansas  Flint  Hills.  Nine 
stations  were  along  one  road  and  seven  were  along 
another.  There  were  eight  convenient  sequences  in 
which  the  16  stations  could  be  visited.  Each  of  the 
eight  sequences  of  stations  was  used  during  eight  con- 
secutive censuses.  Therefore,  there  was  no  consistent 
pattern  in  the  time  after  official  sunrise  that  each  sta- 
tion was  visited.  Distances  between  stations  on  the 
same  road  ranged  from  0.3  to  1.6  km  (mean  = 1.0). 
The  nearest  stations  on  the  two  roads  were  separated 
by  13  km.  Because  our  goal  was  to  quantify  the  effects 
of  weather,  photoperiod,  and  cumulative  number  of 
winter  days  on  species  richness  in  the  entire  area,  data 
were  pooled  for  all  stations.  Birds  were  identified  to 
species,  and  the  number  of  individuals  was  counted 
for  3 min  at  each  station.  Birds  flying  overhead  were 
included  in  the  analysis.  Censuses  were  conducted  at 
approximately  weekly  intervals  November-February, 
1982-1996.  Each  census  began  within  1 h after  sunrise 
and  required  approximately  2 h to  complete.  In  accor- 
dance with  instructions  for  Breeding  Bird  Surveys 
(Robbins  1981b),  no  censuses  were  conducted  in  fog, 
steady  drizzle,  prolonged  rain,  or  winds  stronger  than 
Beaufort  3 (13-19  km/h). 

In  selecting  stations  for  the  census,  the  original  cri- 
terion was  a wooded  area  with  concentrations  of  trees 
of  Juniperus  virginiana,  Monts  rubra,  or  Celtis  occi- 
dentalis  that  would  attract  frugivorous  birds  (Stapanian 
1982).  The  two  roads  along  which  the  stations  were 
spaced  held  a variety  of  habitats  (Table  1 ),  which  af- 
fected our  bird  censuses.  At  each  station  we  visualized 
a line  perpendicular  to  the  road  and  classified  each  of 


TABLE  1 . Habitat  type  for  census  stops  by  the 
number  of  stops  at  which  the  habitat  was  represented 
and  by  the  number  of  90°  arcs  at  the  16  stops  that  were 
predominantly  composed  of  that  habitat. 


Habitat 

Number  of 
census 
slops 

Number  of 
90^"  arcs 

Native  prairie 

1 

1 

Cj  grass  pastures 

5 

6 

Row  crops 

7 

13 

Residential  and  farm  buildings 

7 

9 

Dense  shrub 

2 

4 

Juniper  forest 

4 

4 

Young  mixed  forest 

7 

12 

Mature  mixed  forest  with  oak 

3 

6 

Mature  mixed  forest  without  oak 

2 

6 

Riparian  margin  forest 

2 

3 

Forest  beyond  crops 

5 

— 

the  four  90°  sections  thus  formed  as  being  predomi- 
nantly in  one  category  for  Table  1.  Thus,  there  are  a 
total  of  64  sections  for  the  16  stations  that  form  Table 
1 . The  Flint  Hills  area  of  Kansas  held  almost  no  forests 
before  European  settlement  (Axelrod  1985).  Only 
about  16  species  of  native  trees  have  spread  into  the 
area  from  the  eastern  deciduous  forests  after  the  con- 
trol of  prairie  fires.  Two  stations  were  completely  sur- 
rounded by  forest,  but  14  stations  had  at  least  one  90° 
section  of  forest  holding  one  of  the  three  tree  species 
producing  fleshy  fruit  and  the  other  two  stations  had 
fence  rows  with  M.  rubra.  The  mature  forests  are  sep- 
arated into  those  with  and  without  bur  oaks  {Quercus 
macrocarpa)  because  this  tree  species  must  have  a 
large  acorn  crop  in  order  for  Red-headed  Woodpeckers 
(Melanerpes  erynhrocephalus)  to  winter  in  the  area. 
Some  of  the  residences  near  stations  on  the  census 
were  homes  with  lawns  while  others  had  corrals  for 
livestock.  At  five  stations  birds  could  be  heard  calling 
from  mature  forests  beyond  extensive  fields  of  row 
crops  (Table  1 ). 

Although  our  survey  has  been  conducted  along  the 
same  route  nearly  every  week  since  1978  (Stapanian 
et  al.  1994),  because  weather  data  are  not  available 
before  1982,  we  only  analyzed  data  from  November 
1982  through  February  1996.  We  .selected  the  period 
between  1 November  through  28  February  because  it 
represents  a time  interval  in  the  study  area  during 
which  (1)  food  sources  are  not  renewed  and  (2)  Neo- 
tropical migrants  are  rarely  present.  We  divided  this 
interval  into  four  3()-day  periods  (quarters)  for  analysis 
of  the  presence  of  individual  species.  Our  censuses 
were  designed  to  monitor  populations  of  upland  birds 
(Stapanian  1982,  Stapanian  et  al.  1994).  Aquatic  and 
nocturnal  species  were  eliminated  from  the  present 
analysis.  For  each  census,  we  calculated  the  mean 
number  of  species  recorded  per  station  and  the  total 
number  of  species  recorded  from  all  stations. 

Our  procedures  differed  from  Breeding  Bird  Sur- 
veys in  three  ways.  First,  when  no  birds  were  evident 


552 


THE  WILSON  BULLETIN  • Vol.  111.  No.  4.  December  1999 


at  a station,  we  spished  to  attract  them.  Second,  when 
we  were  unable  to  find  new  birds  where  the  car  was 
parked,  we  walked  along  the  road  in  search  of  birds. 
Third,  we  had  more  than  one  observer  on  58%  of  the 
censuses.  Neither  of  the  first  two  differences  biased  the 
data.  Making  noise  and  walking  along  the  road  when 
no  birds  were  evident  would  tend  to  overestimate  the 
number  of  bird  species  and  individuals  when  they  were 
lowest.  Thus,  any  conclusions  we  would  make  about 
which  factors  decreased  bird  activity  and  the  number 
of  species  would  be  conservative.  The  number  of  ob- 
servers ranged  from  one  (82  censuses,  42. 1 %)  to  four 
(3  censuses,  1.5%).  The  number  of  censuses  in  which 
there  were  two  and  three  observers  were  81  (41.5%) 
and  29  (14.9%),  respectively.  In  exploratory  analyses, 
we  found  that  the  number  of  species  recorded  was 
greater  when  more  than  one  observer  participated  in 
the  census.  Therefore,  we  adjusted  mean  species  per 
station  and  total  species  per  census  for  the  number  of 
observers.  In  controlled  experiments  performed  during 
winter  on  this  route  (C.  C.  Smith,  unpubl.  data),  we 
found  that  the  mean  number  of  species  per  station  and 
total  species  per  census  increased  on  average  by  fac- 
tors of  1.32  and  1.08,  respectively,  for  multiple  ob- 
servers over  those  values  found  by  one  observer.  Thus, 
when  the  number  of  observers  was  greater  than  1,  we 
divided  mean  species  per  station  and  total  species  per 
census  by  1.32  and  1.08,  respectively.  Further,  C.C.S. 
participated  in  all  censuses  and  his  hearing  still  allows 
him  to  detect  a Brown  Creeper  (Certhia  americana)  at 
30  m.  E.J.E  participated  in  almost  all  censuses  from 
1982  through  February  1989.  J.  Cavitt,  S.  Hansen,  S. 
Hull,  C.  Pacey,  G.  Radke,  and  C.  Rebar  participated 
in  at  least  four  censuses  each. 

Weather  data. — Weather  data  for  each  census  were 
collected  automatically  from  a permanent  station  at  the 
Konza  Prairie  Research  Natural  Area,  located  within 
18  km  from  all  our  census  stations.  The  weather  station 
measured  wind  speed  at  hourly  intervals  on  the  hour. 
We,  therefore,  selected  weather  data  recorded  at  07:00 
on  each  census  day.  Because  each  cen.sus  began  within 
1 h after  official  sunrise,  07:00  does  not  represent  a 
standard  time  relative  to  sunrise  for  all  censuses.  How- 
ever, we  were  confident  that  the  data  were  represen- 
tative of  the  weather  during  our  censuses. 

We  use  a standard  weather  service  formula  to  con- 
vert temperature  and  wind  speed  to  a wind  chill  tem- 
perature. Wind  chill  temperature  exceeded  air  temper- 
ature only  for  wind  speeds  greater  than  6.7  km/h, 
which  occurred  on  only  six  cen.suses.  In  exploratory 
analyses  of  variance,  we  found  that  of  the  weather  var- 
iables recorded,  only  temperature  and  wind  speed  ac- 
counted for  a significant  proportion  of  the  variance  in 
our  .statistical  models. 

We  ranked  ground  conditions  from  1 though  6 ac- 
cording to  what  we  perceived  as  increased  difficulty 
for  birds  in  finding  food  on  the  ground:  (1)  dry,  (2) 
dew,  (3)  frost,  (4)  wet  from  rain  or  melting  snow,  (5) 
snow  15  cm  or  le.ss  deep,  and  (6)  snow  more  than  15 
cm  deep.  Ground  condition  was  recorded  at  the  first 
station  we  visited  on  all  but  six  censuses.  All  stations 


were  then  assigned  the  same  weather  data  and  ground 
condition  class  as  the  first  station  for  the  census. 

Statistical  analyses. — “Winter  day”  was  designated 
as  the  number  of  days  after  31  October  for  each  cen- 
sus. Photoperiod  was  calculated  from  published  tables 
(U.S.  Naval  Observatory  1945)  as  the  number  of  min- 
utes between  official  sunrise  and  official  sunset  on 
each  census  date,  and  ranged  from  565  to  679  min. 

We  tested  for  the  effects  of  winter  day,  photoperiod, 
and  weather  components  on  our  two  indicators  of  spe- 
cies richness.  In  exploratory  analyses  and  previous 
studies  (Stapanian  et  al.  1994)  we  found  considerable 
variance  in  the  species  richness  among  winters.  Fur- 
ther, we  found  significant  time  dependency  among  the 
successive  censuses  within  winters.  Therefore,  we  used 
a mixed  models  procedure  (Crowder  and  Hand  1990, 
SAS  Institute  1992,  Littell  et  al.  1996)  in  which  each 
winter  was  treated  as  a random  effect  (i.e.,  whole  plot), 
and  the  remaining  variables  were  treated  as  fixed  co- 
variates (i.e.,  subplots)  to  account  for  the  correlation 
structure  among  the  censuses  within  winters.  The  de- 
grees of  freedom  and  mean  squares  were  adjusted  for 
time  dependency  based  on  the  covariance  structure  and 
inference  space.  Ecologically,  this  meant  that  we  re- 
moved winters  as  a random  effect  from  the  statistical 
model  tests  for  the  effects  of  the  fixed  covariates  based 
on  an  average  set  of  conditions  at  the  beginning  of 
winter.  The  resulting  model  was  general,  not  winter- 
specific. 

We  evaluated  the  covariance  structure  in  three  ways: 
(1)  uniform  correlation  (compound  symmetry),  (2)  ex- 
ponentially decaying,  and  (3)  Markov  chain.  In  ex- 
ploratory analyses,  we  found  that  the  uniform  corre- 
lation method  best  represented  the  covariance  structure 
of  the  data  set.  Further,  we  found  that  none  of  the  two- 
way  interactions  between  the  fixed  covariates  contrib- 
uted significantly  to  the  models  {P  > 0.05  in  all  cases). 
Thus,  we  performed  the  mixed  models  analyses  only 
on  the  main  effects  of  the  fixed  covariates.  We  per- 
formed Tukey’s  tests  for  a posteriori  testing  on  the 
effects  of  specific  ground  condition  classes  on  diver- 
sity. 

We  defined  ordinal  year  as  the  ordinal  number  of  a 
census  year  (i.e.,  year  1 = 1982-1983,  year  2 = 1983- 
1984,  . . . , year  14  = 1995-1996).  We  performed  stan- 
dard Pearson  correlations  between  ordinal  year,  species 
richness,  and  our  weather  variables.  In  this  manner,  we 
were  able  to  test  for  overall  temporal  trends  in  weather 
and  diversity  on  census  days  on  our  census  route. 

For  each  species,  we  calculated  the  proportion  of  the 
censuses  conducted  in  each  quarter  (30-day  interval) 
of  each  year  in  which  that  species  was  recorded  (Ap- 
pendix). These  quarterly  proportions  were  then  pooled 
across  all  14  winters  for  each  species.  Using  analysis 
of  variance,  we  then  tested  the  null  hypothesis  that 
each  species  was  recorded  in  equal  proportions  in  all 
four  quarters.  Tukey’s  pair-wi.se  comparisons  were 
used  for  all  a posteriori  testing.  SAS  for  Personal 
Computers,  version  6.12  for  Windows  was  used  for 
statistical  computations. 


Stapauian  el  cil.  • WEATHER  EFFECTS  ON  WINTER  BIRDS 


553 


TABLE  2.  Summary  statistics  for  each  winter  for  the  total  number  of  species  recorded  on  each  census  and 
the  mean  number  of  species  per  station  on  each  census.  These  variables  were  adjusted  for  number  of  observers. 
Years  that  share  a grouping  letter  were  not  significantly  different  (Tukey’s  pair-wise  comparisons,  P > 0.05)  for 
that  indicator  of  diversity. 

Species  per  census 

Species  per  station 

Year 

IV' 

Mean 

SD*’ 

Mean 

SD*’ 

1982-1983 

12 

24.18 

2.29  B,  C,  D 

5.00 

0.74  A,  B,  C 

1983-1984 

14 

18.21 

4.97  E 

3.52 

1.54  D,  E 

1984-1985 

14 

19.48 

2.65  D,  E 

3.99 

1.02  C,  D 

1985-1986 

13 

23.40 

3.32  B,  C 

5.04 

1.44  A,  B,  C 

1986-1987 

1 1 

22.09 

3.67  B,  C,  D,  E 

4.89 

1.02  A,  B,  C,  D 

1987-1988 

12 

23.73 

2.59  B,  C 

5.50 

1.18  A,  B 

1988-1989 

14 

24.45 

2.07  B,  C 

5.37 

0.91  A,  B 

1989-1990 

15 

23.75 

4.16  B,  C 

4.98 

0.59  A,  B,  C 

1990-1991 

14 

25.93 

2.76  B 

5.86 

1.09  A 

1991-1992 

15 

18.39 

3.26  E 

3.65 

1.25  D 

1992-1993 

15 

23.71 

2.99  B,  C 

4.97 

1.07  A,  B,  C 

1993-1994 

16 

29.81 

3.67  A 

5.36 

1.12  A,  B 

1994-1995 

13 

24.00 

2.16  B,  C 

5.34 

1.36  A,  B 

1995-1996 

16 

22.32 

1.37  C,  D 

4.63 

0.95  B,  C,  D 

^ n = number  of  censuses. 
^ SD  = standard  deviation. 


RESULTS 

There  were  significant  differences  among 
winters  for  the  annual  means  of  both  total  spe- 
cies per  census  and  species  per  station  (F,3  ,go 
= 13.87  and  5.46,  respectively,  P < 0.001  in 
both  cases;  Table  2).  Consequently,  we  treated 
winters  as  random  effects  in  our  mixed  model 
analysis.  The  results  from  the  mixed  model 
procedure  (Table  3)  suggested  significant  ef- 
fects from  winter  day,  photoperiod,  tempera- 
ture, and  wind  speed  for  the  number  of  species 
per  census  and  species  per  station.  Tempera- 
ture and  photoperiod  were  positively  related 
to  both  indicators  of  species  richness  when  the 
covariance  structure  was  taken  into  account 
(slopes  in  Table  3).  On  the  other  hand,  wind 
speed  and  winter  day  had  negative  effects  on 
both  indicators  of  diversity  (slopes  in  Table  3). 


On  average,  a change  of  1°  C in  temperature 
or  1 km/h  in  wind  speed  had  a greater  effect 
on  species  richness  than  did  either  a change 
of  1 min  in  photoperiod  or  1 day  further  into 
winter. 

Ground  condition  had  a significant  effect  on 
species  per  station,  but  not  on  species  per  cen- 
sus in  the  mixed  model  analyses  (Table  3). 
Values  of  species  per  station  were  lowest 
when  there  was  more  than  15  cm  of  snow  on 
the  ground  and  greatest  when  the  ground  was 
covered  with  dew  (Table  4). 

Both  of  our  indicators  of  species  richness 
increased  over  the  course  of  our  study.  There 
was  weak  but  positive  correlation  between 
species  per  census  and  ordinal  year  (r  = 0.27, 
df  = 12,  P < 0.001)  and  between  species  per 
station  and  ordinal  year  (r  — 0. 15,  df  = 12, 


TABLE  3.  Results  of  the  mixed  models  ANOVA  procedure.  Two-way  interactions  were  not  found  to  be 
significant  in  exploratory  analyses  (P  > 0.05).  Slopes  and  standard  errors  (SE)  of  the  slopes  are  not  reported 
for  ground  condition  because  it  was  not  a continuous  variable. 

Specie.s  per  census  Species  per  station 


Source  tlf  P > F Slope  SE  /•'  P > F Slope  SE 


Winter  day  (days) 

1 20.35 

0.001 

Photoperiod  (minutes) 

1 11.51 

0.0009 

Temperature  (°C) 

1 7.65 

0.0063 

Wind  speed  (km/h) 

1 26.36 

0.0001 

Ground  condition 

5 1.10 

0.3636 

0.033 

0.007 

21.89 

0.0001 

-0.01 1 

0.002 

0.029 

0.008 

40.03 

0.0001 

0.017 

0.003 

0.1 1 1 

0.040 

30.53 

0.0001 

0.070 

0.013 

0.634 

0.123 

45.86 

2.54 

0.0001 

0.0301 

-0.264 

0.039 

554 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


TABLE  4.  A posteriori  tests  (Tukey’s  pair-wise 
comparisons)  on  the  effects  of  ground  condition  on  the 
mean  values  of  the  mean  number  of  species  per  station 
on  a census.  Means  of  ground  condition  classes  with 
at  least  one  letter  in  common  are  not  signihcantly  dif- 
ferent (P  > 0.05). 


Ground 

condition 

Class 

Mean  of  mean 
species  per 
station 

Clear,  dry 

1 

53 

4.93  B 

Dew 

2 

10 

6.12  A 

Erost 

3 

39 

5.14  B 

Wet 

4 

31 

4.75  B 

Snow  < 15  cm 

5 

50 

4.47  B 

Snow  > 15  cm 

6 

5 

3.56  C 

^ n = number  of  censuses. 

P = 0.035). 

Further, 

there  was 

a weak  but 

negative  correlation  between  wind  speed  and 
ordinal  year  (r  = —0.16,  df  = 12,  P = 0.022), 
which  suggested  that  wind  speed  on  census 
trips  decreased  over  the  course  of  this  study. 
We  made  no  conscious  change  in  our  policy 
of  when  to  hold  censuses  during  our  study  that 
would  have  resulted  in  lower  wind  speeds  dur- 
ing censuses.  Neither  temperature  nor  ground 
condition  class  was  significantly  correlated 
with  ordinal  year  (P  > 0.05  in  both  cases). 

Sixty-one  species  were  recorded  for  our 
study  (Appendix).  Only  two  species,  Ameri- 
can Crow  (Corvus  brachyrhynchos)  and 
Black-capped  Chickadee  (Poecile  atricapil- 
lus),  were  recorded  on  all  194  censuses. 
Twenty-two  species  (36.1%  of  the  total  spe- 
cies) were  recorded  on  at  least  50%  of  the 
censuses.  Thirteen  species  (21.3%  of  the  total 
species)  were  seen  on  less  than  5%  of  the  cen- 
suses. There  was  no  consistent  pattern  to  the 
temporal  occurrences  of  individual  species 
(Appendix).  For  most  species,  the  proportion 
of  the  censuses  in  which  they  were  recorded 
was  the  same  for  each  quarter.  Eight  species. 
Northern  Flicker  (Colaptes  ciurcitus).  Golden- 
crowned  Kinglet  (Regulus  satrapa).  Northern 
Bobwhite  (Colinus  virginianus).  White- 
crowned  Sparrow  (Zonotrichio  leucophrys). 
Common  Crackle  (Quiscalus  quiscula).  Field 
Sparrow  (Spizella  piisiUa),  White-throated 
Sparrow  (Zonotrichia  alhicollis),  and  Lin- 
coln’s Sparrow  (Melospiz.o  lincolnii),  occurred 
more  frequently  in  early  winter  than  in  late 
winter.  The  reverse  trend  was  exhibited  by 
four  species,  American  Tree  Sparrow  (Spizella 


arborea).  Tufted  Titmouse  (Baeolophus  bicol- 
or), Western  Meadowlark  (Sturnella  neglec- 
ta),  and  Eastern  Meadowlark  (Sturnella  mag- 
na).  Two  species.  Red-winged  Blackbird 
(Agelaius  phoeniceus),  and  Ring-necked 
Pheasant  (Phasianus  colchicus),  were  record- 
ed less  frequently  in  mid-winter  (i.e.,  in  the 
second  and  third  quarters)  than  at  the  begin- 
ning or  end.  Species  observed  on  fewer  than 
5%  of  the  censuses  were  not  considered  com- 
mon enough  to  test  for  patterns  of  occurrence 
by  winter  quarter  (Appendix). 

DISCUSSION 

Our  study  is  the  first  of  which  we  are  aware 
that  demonstrated  significant  effects  of  pho- 
toperiod, cumulative  number  of  winter  days, 
and  weather  components  on  bird  species  rich- 
ness of  upland  and  riparian  forest  birds  in 
winter.  The  results  appear  to  differ  consider- 
ably from  those  of  Robbins  (1981a).  He  de- 
tected no  effects  of  weather  conditions  on  the 
numbers  of  selected  species  or  families  of 
birds  from  repeated  coverage  of  a Winter  Bird 
Survey  route.  Robbins’  8-km  route  in  Mary- 
land was  covered  at  least  three  times  per  year 
for  five  consecutive  years  in  late  December  or 
early  January.  He  also  analyzed  data  from 
eight  years  of  Audubon  Winter  Bird-Popula- 
tion Studies  on  two  forest  plots  in  Maryland. 
There  were  no  significant  effects  of  tempera- 
ture on  the  number  of  species  he  recorded. 
The  differences  between  our  results  and  those 
from  Robbins  may  be  due  to  (1)  our  larger 
sample  size,  (2)  a longer  season  (i.e.,  Novem- 
ber through  February)  in  our  study,  (3)  the 
fact  that  Robbins’  (1981a)  analyses  were  re- 
stricted to  selected  species  and  families,  or  (4) 
differences  in  location  and  climate.  Most  im- 
portantly, Robbins  (1981a)  selected  for  calm, 
dry  mornings  in  both  studies.  Therefore,  it  is 
not  surprising  that  he  reported  no  weather  ef- 
fects on  the  number  of  species  recorded. 

Although  ours  is  a long-term  study,  there 
were  too  few  censuses  to  analyze,  the  effects 
of  number  of  winter  days,  photoperiod,  and 
weather  on  species  richness  for  specific  win- 
ters. The  data  strongly  suggest  that  differences 
in  the  detection  of  bird  species  occurred 
among  winters.  Previously,  we  (Stapanian  et 
al.  1994)  estimated  extremely  low  seed  crops 
for  weeds,  herbs,  grasses,  and  bur  oak  for  the 
winters  of  1983-1984  and  1984-1985.  Simi- 


Stapanum  et  cil.  • WEATHER  EEFECTS  ON  WINTER  BIRDS 


555 


lady,  we  estimated  extremely  low  weed  seed 
and  wild  fleshy  fruit  crops  for  the  winter  of 
1991—1992.  These  low  food  supplies  may 
partly  explain  why  the  fewest  species  were  re- 
corded in  those  years.  Large  crops  of  herb  and 
grass  seeds  were  estimated  for  the  winter  of 
1982-1983,  which  had  relatively  high  values 
for  species  per  station.  Similarly,  there  were 
large  crops  of  acorns  and  fleshy  fruits  for 
1988-1989  and  1990-1991.  In  both  winters, 
species  richness  was  relatively  high.  These 
trends  support  the  importance  of  the  size  of 
unrenewed  food  supplies  in  determining  the 
detection  of  winter  bird  populations. 

By  treating  the  large  winter  differences  as 
random  effects,  the  statistical  analysis  dem- 
onstrated that  photoperiod  affects  the  morning 
activity  of  birds.  The  influences  of  photope- 
riodism  on  the  physiology  and  activity  of 
birds  are  well  documented  (Bissonette  1932, 
1937;  Bartholomew  1949;  Welty  and  Baptista 
1988  and  references  therein;  Ball  1993;  Hau 
et  al.  1998).  Perhaps  when  less  time  is  avail- 
able for  feeding,  as  in  mid-winter,  birds  spend 
less  time  in  easily  detected  behaviors.  The  be- 
havioral effect  of  reduced  feeding  time  rela- 
tive to  energy  needs  is  likely  to  be  larger  flock 
size  (Caraco  1979,  Sullivan  1988)  and  a lower 
probability  of  seeing  birds  at  the  average  sta- 
tion. 

Temperature  and  ground  condition  classes 
were  significantly  and  negatively  correlated  (r 
—0.37,  P < 0.001,  n = 187).  Thus,  what 
we  perceived  as  difficult  foraging  conditions 
might  have  been  simply  a consequence  of  low 
temperature.  Dew  (ground  condition  class  2) 
and  ground  wet  from  rain  or  snowmelt  (class 
4)  required  that  the  air  temperature  exceed 
0°  C,  while  the  temperature  may  be  below 
0°  C for  dry  ground  (class  1)  and  will  be  for 
frost  (class  3).  When  we  switched  the  number 
class  of  frost  to  4 and  rain  or  snowmelt  to  3, 
the  correlation  coefficient  between  tempera- 
ture and  ground  condition  class  increased  in 
absolute  magnitude  (r  = —0.508,  P < 0.001). 
Ground  condition  may  have  little  effect  in- 
dependent of  temperature. 

We  are  not  sure  how  to  interpret  the  nega- 
tive effects  of  wind  speed  on  bird  species  per 
census  or  bird  species  per  station.  Wind  speed 
had  a negative  effect  on  both  indicators  of 
species  richness  even  when  we  considered 
only  those  censuses  in  which  wind  speeds 


were  less  than  6.7  km/h,  the  speed  above 
which  wind  chill  temperature  is  less  than  air 
temperature.  The  effects  of  wind  speed  on 
species  richness  appear  to  be  due  to  neither  a 
decrease  in  our  ability  to  hear  birds,  nor  ap- 
parent additional  thermoregulatory  stress  for 
the  birds.  However,  wind  speed  typically  in- 
creases after  sunrise,  and  the  wind  speed  at 
the  end  of  a census  may  be  greater  than  at  the 
beginning.  Small  differences  in  wind  speed  at 
07:00  may  be  magnified  later  in  the  census. 
There  is  evidence  that  some  species,  particu- 
larly those  with  small  body  sizes,  can  reduce 
metabolic  demands  in  winter  by  selecting  mi- 
crohabitats  that  are  sheltered  from  the  wind 
and  exposed  to  solar  radiation  (Wolf  and 
Walsberg  1996). 

Similarly,  we  are  unsure  why  both  indica- 
tors of  species  richness  increased  in  later 
years.  The  same  principal  observer  (C.C.S.) 
was  present  for  all  censuses  in  our  study.  Eye- 
sight and  hearing  typically  deteriorate  over 
time  (Cyr  1981),  but  these  effects  can  be 
countered  by  individual  experience  with  a 
specific  route.  The  increases  in  species  rich- 
ness were  not  due  to  changes  in  mean  annual 
temperature,  because  temperature  and  ordinal 
census  year  were  not  significantly  correlated 
(r  = 0.109,  df  = 12,  P > 0.05).  Wind  speed 
on  the  census  trips  was  negatively  correlated 
with  year  and  with  both  indicators  of  species 
richness.  Thus,  a decrease  in  wind  speed  on 
census  days  may  partially  explain  the  increas- 
es in  species  per  station  and  species  per  cen- 
sus over  the  census  years.  Species  composi- 
tion on  the  census  route  changed  over  time. 
For  example,  the  population  of  Carolina  Wren 
(Thryothorus  ludovicicvms),  a sedentary  bird 
species,  increased  steadily  during  the  14  years 
of  the  census  after  a time  when  it  was  at  low 
levels  in  the  Manhattan  Christmas  bird  counts. 
Wild  Turkeys  (Meleagris  gallopavo)  were  re- 
introduced in  the  area  in  the  early  1980s  and 
their  populations  have  increased  since.  These 
changes  may  be  due  to  milder  winters  during 
the  study  period.  There  may  have  been  an  in- 
crease in  canopy  closure  or  structural  diversity 
of  the  habitat  on  the  route  over  the  14  years 
of  the  study,  but  that  was  not  measured. 

White-crowned  Sparrows  and  White-throat- 
ed Sparrows  feed  in  large  mixed  flocks  of 
sparrows  in  late  winter  and  in  smaller  groups 
in  late  fall.  These  species  were  recorded  on  a 


556 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


greater  proportion  of  the  censuses  in  the  first 
quarter  than  in  any  other.  November  (first 
quarter)  typically  has  the  mildest  average 
weather  of  any  quarter  and  food  abundance 
should  be  greatest  at  this  time.  Therefore,  the 
selective  pressure  on  birds  for  being  in  large 
flocks  should  be  least  at  this  time  (Caraco 
1979,  Sullivan  1988).  If  members  of  a species 
were  more  widely  distributed  geographically, 
an  observer  would  be  more  likely  to  record 
that  bird  species  at  least  once  on  a census. 
Field  Sparrows,  Lincoln’s  Sparrows,  and 
Common  Crackles  were  also  seen  with  greater 
frequency  in  November  than  in  late  winter. 
The  Kansas  Flint  Hills  are  along  the  northern- 
most edge  of  the  winter  range  of  these  three 
species  (Thompson  and  Ely  1992).  Although 
most  members  of  these  species  migrate  south 
out  of  the  census  area,  a few  overwinter  (Zim- 
merman 1993).  The  decrease  in  numbers  for 
these  species  during  the  winter  could  be  a re- 
sult of  continued  migration  south  or  mortality 
in  a marginal  range. 

The  proportions  of  censuses  on  which 
Northern  Bobwhite  (a  resident  species).  Gold- 
en-crowned Kinglet  (a  winter  migrant),  and 
Northern  Flicker  (winter  resident  and  winter 
migrant;  Thompson  and  Ely  1989,  1992;  Zim- 
merman 1993)  were  recorded  dropped  steadily 
from  the  second  through  the  fourth  quarter. 
This  decline  may  have  been  due  to  mortality 
to  the  wintering  populations  of  those  species. 
The  results  from  our  study  for  Golden- 
crowned  Kinglet  agreed  with  those  of  Zim- 
merman (1993)  who  hypothesized  that  the 
variation  in  departure  of  Golden-crowned 
Kinglets  from  the  area  was  related  to  avail- 
ability of  food. 

Eastern  Meadowlark,  Western  Meadowlark, 
and  Tufted  Titmouse  were  recorded  most  fre- 
quently on  censuses  in  the  last  quarter.  These 
are  resident  species  (Thompson  and  Ely  1992, 
Zimmerman  1993)  that  breed  early  in  spring 
and  begin  establishing  territories  and/or  ob- 
taining mates  in  late  winter.  The  American 
Tree  Sparrow,  a winter  migrant  to  the  area 
(Zimmerman  1993),  was  recorded  least  fre- 
quently in  the  first  quarter.  This  agrees  with 
Finck  (1986),  who  found  this  species  to  be 
most  numerous  from  December  through  Feb- 
ruary, suggesting  a late  migratory  arrival. 
Red-winged  Blackbird  and  Ring-necked 
Pheasant  were  most  often  recorded  in  Novem- 


ber and  February.  These  are  resident  or  partly 
resident  species  that  flock  in  severe  weather 
in  mid-winter.  However,  they  begin  prepara- 
tion for  breeding  in  late  winter  (Zimmerman 
1993).  Zimmerman  (1993)  found  Red- winged 
Blackbird  to  be  “occasional”  during  the  win- 
ter months  of  most  years  in  upland  habitats 
until  the  migrants  returned  in  late  winter. 

Our  results  are  consistent  with  at  least  five 
hypotheses:  (1)  species  are  lost  by  mortality 
resulting  from  nonrenewed  resources  over  the 
course  of  winter,  (2)  resident  species  move  in 
and  out  of  detection  distance  in  the  census 
area,  (3)  selective  pressures  for  flock  sizes 
change  with  weather  conditions  and  food 
abundance,  (4)  some  species  are  more  easily 
detected  in  late  winter  because  of  early  court- 
ship behavior,  and  (5)  species  richness  in  the 
census  area  changes  as  a result  of  the  arrival 
and  departure  of  seasonally  migrant  species. 
The  results  suggest  a complex  relationship 
among  weather  components,  photoperiod, 
abundance  of  resources,  and  species-specific 
evolutionary  histories  on  winter  bird  species 
richness.  We  suggest  further  studies  to  analyze 
responses  by  individual  species  to  resource 
abundance  and  abiotic  factors  in  winter. 

ACKNOWLEDGMENTS 

R.  Leighty  provided  assistance  with  statistical  ana- 
lyses. C.  Robbins,  T Grubb,  and  J.  Harder  commented 
on  earlier  drafts  of  the  paper.  Weather  data  were  col- 
lected as  part  of  the  NSF  Long  Term  Ecological  Re- 
search Program  at  Konza  Prairie  Research  Natural 
Area.  The  USGS/BRD  Ohio  Cooperative  Fish  and 
Wildlife  Research  Unit  provided  support  to  work  on 
this  manuscript. 

LITERATURE  CITED 

Altman,  R.  1983.  The  effects  of  time  of  day  on  a 
winter  bird  survey.  Ky.  Warbler  59:23-26. 
Axelrod,  D.  I.  1985.  Rise  of  the  grassland  biome. 
Central  North  America.  Bot.  Rev.  51:163-201. 
Ball,  G.  E 1993.  The  neural  integration  of  environ- 
mental information  by  seasonally  breeding  birds. 
Am.  Zool.  33:185-199. 

Bartholomew,  G.  A.  1949.  The  effect  of  light  inten- 
sity and  day  length  on  reproduction  in  the  English 
Sparrow.  Bull.  Mus.  Comp.  Zool.  101:433-476. 
Blssonnette,  T.  H.  1932.  Light  and  diet  as  factors  in 
relation  to  sexual  photoperiodicity.  Nature  (Lond.) 
129:612. 

Bissonnette,  T.  H.  1937.  Photoperiodicity  in  birds. 
Wilson  Bull.  49:241-270. 

Caraco,  T.  1979.  Time  budgeting  and  group  size:  a 
test  of  theory.  Ecology  60:618-627. 


StapanUm  et  al.  • WEATHER  EFFECTS  ON  WINTER  BIRDS 


557 


Collins,  R T.  1989.  Surviving  the  winter;  the  physi- 
ology of  thermoregulation  in  birds.  Passenger  Pi- 
geon 51:315-320. 

Crowder,  M.  J.  and  D.  J.  Hand.  1990.  Analysis  of 
repeated  measures.  Monogr.  Stat.  Appl.  Probabil- 
ity 41:1-257. 

Cyr,  a.  1981.  Limitation  and  variability  in  hearing 
ability  in  censusing  birds.  Stud.  Avian  Biol.  6: 
327-333. 

Dawson,  W.  R.  and  R.  L.  Marsh.  1986.  Winter  fat- 
tening in  the  American  Goldfinch  and  the  possible 
role  of  temperature  in  its  regulation.  Physiol.  Zool. 
59:357-368. 

Erskine,  a.  j.  1992.  A ten-year  urban  winter  bird 
count  in  Sackville,  New  Brunswick.  Can.  Field- 
Nat.  106:499-506. 

Finck,  E.  j.  1986.  Birds  wintering  on  the  Konza  Prairie 
Research  Natural  Area.  Proc.  N.  Am.  Prairie  Conf. 
9:91-94. 

HaU,  M.,  M.  WlKELSKI,  AND  J.  C.  WiNGFlELD.  1998.  A 
Neotropical  bird  can  measure  the  slight  changes 
in  tropical  photoperiod.  Proc.  R.  Soc.  London  B 
265:89-95. 

Hidden,  O.  1987.  Finnish  winter  bird  censuses:  long- 
term trends  in  1956-1984.  Acta  Oecol.  8:157- 
168. 

Houston,  A.  I.  and  J.  M.  McNamara.  1993.  A theo- 
retical investigation  of  the  fat  reserves  and  mor- 
tality levels  of  small  birds  in  winter.  Ornis  Scand. 
24:205-219. 

Littell,  R.  C.,  G.  a.  Milliken,  W.  W.  Stroup,  and 
R.  D.  WoLFiNGER.  1996.  SAS  systems  for  mixed 
models.  SAS  Institute,  Inc.,  Cary,  North  Carolina. 

Loery,  G.  AND  J.  D.  Nichols.  1985.  Dynamics  of  a 
Black-capped  Chickadee  population,  1958-1983. 
Ecology  66:1195-1203. 

Peach,  W.  J.,  D.  P.  Hodson,  and  J.  A.  Fowler.  1992. 
Variation  in  the  winter  body  mass  of  starlings 
Sturnus  vulgaris.  Bird  Study  39:89-95. 

PiLASTRO,  A.,  G.  BERTORELLE,  AND  G.  Marin.  1995. 
Winter  fattening  strategies  of  two  passerine  spe- 
cies: environmental  and  social  influences.  J.  Avian 
Biol.  26:25-32. 

Robbins,  C.  S.  1972.  An  appraisal  of  the  winter  bird- 
population  study  technique.  Am.  Birds  26:1-4. 

Robbins,  C.  S.  1981a.  Bird  activity  levels  related  to 
weather.  Stud.  Avian  Biol.  6:301-310. 


Robbins,  C.  S.  1981b.  Reappraisal  of  the  winter  bird- 
population  study  technique.  Stud.  Avian  Biol.  6: 
52-57. 

Robbins,  C.  S.,  D.  Bystrak,  and  P.  H.  Geissler.  1986. 
The  breeding  bird  survey;  its  first  fifteen  years, 
1965-1979.  Spec.  Sci.  Rep.-Wildl.  157:1-196. 

Rogers,  C.  M.,  V.  Nolan,  Jr.,  and  E.  D.  Ketterson. 
1994.  Winter  fattening  in  the  Dark-eyed  Junco: 
plasticity  and  possible  interaction  with  migration 
trade-offs.  Oecologia  97:526-532. 

SAS  Institute,  Inc.  1992.  SAS/STAT  software: 
changes  and  enhancements.  Release  6.07.  SAS  In- 
stitute, Inc.,  Cary,  North  Carolina. 

Stapanian,  M.  a.  1982.  Evolution  of  fruiting  strate- 
gies among  fleshy-fruited  plant  species  in  eastern 
Kansas.  Ecology  63:1422-1431. 

Stapanian,  M.  A.,  C.  C.  Smith,  and  E.  J.  Finck.  1994. 
Population  variabilities  of  bird  guilds  in  Kansas 
during  fall  and  winter:  weekly  censuses  versus 
Christmas  bird  counts.  Condor  96:58-69. 

Sullivan,  K.  A.  1988.  Ontogeny  of  time  budgets  of 
Yellow-eyed  Juncos:  adaptation  to  ecological  con- 
straints. Ecology  69:118-124. 

Thompson,  M.  C.  and  C.  Ely.  1989.  Birds  in  Kansas; 
vol.  I.  Univ.  of  Kansas,  Univ.  Press  of  Kansas, 
Lawrence. 

Thompson,  M.  C.  and  C.  Ely.  1992.  Birds  in  Kansas: 
vol.  II.  Univ.  of  Kansas,  Univ.  Press  of  Kansas, 
Lawrence. 

U.S.  Naval  Observatory.  1945.  Supplement  to  the 
American  Ephemeris,  1946.  Tables  of  sunrise, 
sunset,  and  twilight.  U.S.  Government  Printing 
Office,  Washington,  D.C. 

Waite,  T.  A.  1992.  Winter  fattening  in  Gray  Jays:  sea- 
sonal, diurnal,  and  climatic  correlates.  Ornis 
Scand.  23:499-503. 

Welty,  j.  C.  and  L.  Baptista.  1988.  The  life  of  birds, 
fourth  ed.  Saunders  College  Publishing,  New 
York. 

White,  C.  M.  and  G.  C.  West.  1977.  The  annual  lipid 
cycle  and  feeding  behavior  of  Alaskan  Redpolls. 
Oecologia  27:227-238. 

Wolf,  B.  O.  and  G.  E.  Walsberg.  1996.  Thermal  ef- 
fects of  radiation  and  wind  on  a small  bird  and 
implications  for  microsite  selection.  Ecology  77: 
2228-2236. 

Zimmerman,  J.  L.  1993.  The  birds  of  Konza.  Univ. 
Press  of  Kansas,  Lawrence. 


558 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


APPENDIX 

The  number  of  all  censuses  (All)  and  the  mean  proportion  of  censuses  in  30-day  intervals  over  14  winters  in 
which  species  were  observed.  For  each  species,  we  tested  the  null  hypothesis  that  it  was  recorded  in  equal 
proportions  in  all  four  quarters.  A posteriori  tests  (Tukey’s  pair-wise  comparisons)  were  performed  on  those 
species  for  which  the  null  hypothesis  was  rejected.  Quarterly  means  for  a species  having  the  same  letter  were 
not  significantly  (P  < 0.05)  different.  Species  listed  at  the  bottom  lacked  significant  quarterly  differences. 
Abbreviations:  All  = all  censuses  combined,  n = number  of  censuses. 


Common  name 

Scientific  name 

All 

n = 194 

1-30  Nov 
n = 52 

1-30  Dec 
n = 42 

31  Dec-29  Jan 
n = 46 

30  Jan-28  Feb 
n = 54 

Northern  Flicker 

Colaptes  aiiratiis 

171 

0.962  A 

0.986  A 

0.873  AB 

0.768  B 

American  Tree  Sparrow 

Spizella  arborea 

157 

0.562  B 

0.958  A 

0.962  A 

0.857  A 

Tufted  Titmouse 

Baeolophus  bicolor 

153 

0.673  B 

0.642  B 

0.902  A 

0.929  A 

Red-winged  Blackbird 

Agelaiiis  phoeniceus 

143 

0.926  A 

0.626  BC 

0.579  C 

0.831  AB 

Western  Meadowlark 

Sturnella  neglecta 

99 

0.576  AB 

0.318  C 

0.396  BC 

0.737  A 

Golden-crowned  Kinglet 

Regulus  satrapa 

84 

0.604  A 

0.669  A 

0.321  B 

0.127  B 

Northern  Bobwhite 

Colinus  virginianus 

66 

0.382  AB 

0.461  A 

0.364  AB 

0.211  B 

White-crowned  Sparrow 

Zonotrichia  leucophrys 

36 

0.385  A 

0.095  B 

0.143  B 

0.157  B 

Ring-necked  Pheasant 

Phasianus  colchicus 

34 

0.240  A 

0.183  AB 

0.014  B 

0.294  A 

Common  Crackle 

Quiscalus  quiscula 

24 

0.308  A 

0.060  B 

0.024  B 

0.071  B 

Eastern  Meadowlark 

Sturnella  magna 

20 

0.070  B 

0.018  B 

0.089  B 

0.205  A 

Field  Sparrow 

Spizella  pusilla 

18 

0.270  A 

0.071  B 

0.056  B 

0.032  B 

White-throated  Sparrow 

Zonotrichia  albicollis 

14 

0.173  A 

0.060  AB 

0.018  B 

0.018  B 

Lincoln’s  Sparrow 

Melospiza  lincolnii 

13 

0.130  A 

0.065  AB 

0.000  B 

0.050  AB 

American  Crow  (Corvus  brachyrhynchos)  194,  Black-capped  Chickadee  (Poeciie  alricapillus)  194,  Dark-eyed  Junco  (Junco  hyemalis)  193,  Northern 
Cardinal  (Cardinalis  cardinalis)  192,  House  Sparrow  (Passer  dome sticus)  192,  Blue  Jay  (Cyanocilla  cristala)  190,  American  Goldfinch  (Carduelis  iristis) 
187,  White-breasted  Nuthatch  (Silla  carolinensis)  187,  Red-bellied  Woodpecker  (Melanerpes  carolinus)  184,  Downy  Woodpecker  (Picoides  pubescens) 
170,  European  Starling  (Slurnus  vulgaris)  169,  American  Robin  (Tiirdus  migralorius)  167,  Red-tailed  Hawk  (Buteo  jamaicensis)  157,  Hairy  Woodpecker 
(Picoides  villosus)  143,  Harris’s  Sparrow  (Zonotrichia  querula)  136,  Eastern  Bluebird  (Sialia  sialis)  128,  Carolina  Wren  (Thryothorus  ludovicianus)  97, 
Red-headed  Woodpecker  (Melanerpes  erythrocephalus)  92,  Rock  Dove  (Columba  livia)  81,  Song  Sparrow  (Melospiza  melodia)  68,  American  Kestrel 
(Falco  sparverius)  49,  Brown  Creeper  (Cerlhia  americana)  47,  Northern  Harrier  (Circus  cyaneus)  41,  Cedar  Waxwing  (Bombycilla  cedrorum)  40,  Bewick’s 
Wren  (Thrvomanes  bewickii)  36,  Spotted  Towhee  (Pipilo  maculalus)  33,  Wild  Turkey  (Meleagris  gallopavo)  29,  Mourning  Dove  (Zenaida  macroura)  27, 
Pine  Siskin  (Carduelis  pinus)  23,  Brown-headed  Cowbird  (Molothrus  ater)  20,  Yellow-bellied  Sapsucker  (Sphyrapicus  varius)  16,  Winter  Wren  (Troglodytes 
iroglodvles)  16,  Sharp-shinned  Hawk  (Accipiter  sirialus)  14.  Yellow-rumped  Warbler  (Dendroica  coronata)  1 1,  Loggerhead  Shrike  (Lanius  ludovicianus) 
7.  Rough-legged  Hawk  (Buteo  lagoputs)  6,  Horned  Lark  (Eremophila  alpestris)  6,  Rusty  Blackbird  (Euphagus  carolinu.s)  6,  Cooper’s  Hawk  (Accipiter 
cooperii)  4.  Purple  Finch  (Carpodacus  purpureus)  3,  Prairie  Falcon  (Falco  mexicanus)  3,  Fox  Sparrow  (Passerella  iliaca)  3,  Sedge  Wren  (Cistothorus 
platensis)  2,  Red-breasted  Nuthatch  (Sitta  canadensis)  2,  Brown  Thrasher  (Toxostoma  rufum)  2.  Hermit  Thrush  (Catharus  guttatus)  1.  Yellow-headed 
Blackbird  (Xanthocephalus  .xanthocephalus)  1. 


Short  Communications 


Wilson  Bull.,  1 1 1(4),  1999,  pp.  559-560 


Possible  Winter  Quarters  of  the  Aleutian  Tern? 

Norman  P.  Hill'  and  K.  David  Bishop-^ 


ABSTRACT. — Recent  observations  of  the  Aleutian 
Tern  {Sterna  aleutica)  in  the  coastal  waters  around 
Hong  Kong  in  spring  and  fall,  and  Singapore  and  the 
Indonesian  islands  of  Karimun  and  Bintan  between 
October  and  April  indicate  that  at  least  part  of  the  pop- 
ulation of  this  species  migrates  through  and  winters  in 
these  areas.  Our  observations  during  December  1997, 
suggest  that  the  coastal  waters  of  Java,  Bali  and  Su- 
lawesi may  form  an  additional  part  of  the  winter  range 
of  this  species.  Received  3 August  1998,  accepted  17 
June  1999. 


The  Aleutian  Tern  {Sterna  aleutica)  breeds 
along  the  western  coast  of  Alaska  (USA)  and 
in  Asia  on  the  east  coast  of  Kamchatka  and 
Sakhalin  (American  Ornithologists’  Union 
1998).  In  Alaska,  birds  return  to  colonies  dur- 
ing early  May  and  then  disperse  during  Au- 
gust and  September  after  breeding,  (Harrison 
1983).  Both  the  AOU  Checklist  (1998)  and 
Harrison  (1983)  state  that  this  species’  winter 
range  is  “unknown”. 

Daring  the  last  decade  or  so  there  has  been 
a steady  accumulation  of  records  for  this  spe- 
cies outside  its  breeding  range  from  Southeast 
Asia.  Lee  (1992)  reported  six  specimens  col- 
lected in  May  1984  in  the  Mindanao  Sea  off 
Bohol,  Philippines.  Brazil  (1991)  noted  ap- 
proximately ten  records  from  Honshu  and 
Hokkaido,  Japan  including  one  instance  of 
probable  breeding.  One  exceptional  record  in- 
volves the  occurrence  of  a single  vagrant  bird 
on  the  Fame  Islands  off  the  coast  of  northeast 
England  during  May  1979  (Dixey  et  al.  1981). 
More  recently,  during  August  and  September 
1992  as  many  as  190  birds  were  observed  off 
the  southern  and  southeastern  coast  of  Hong 
Kong  (Kennerley  et  al.  1993). 

Initially  birds  were  observed  in  breeding 


' 38  North  Main  Street,  Assonet,  MA  02702. 

2 ‘Semioptera’,  RO.  Box  6068,  Kincumbcr,  NSW 
2251,  Australia. 

3 Corresponding  author;  E-mail:  kdbishop@ozemail. 
com.au 


(alternate)  plumage  in  late  August  but  most 
then  molted  into  non-breeding  (basic)  plum- 
age during  September.  Details  of  the  latter, 
poorly  known  plumage,  can  be  found  in  Lee 
(1992),  Kennerley  and  coworkers  (1993)  and 
Kennerley  and  Ollington  (1998).  Subsequent- 
ly, small  numbers  have  been  observed  annu- 
ally in  Hong  Kong  waters  including  individ- 
uals in  breeding  plumage  during  April-June. 
It  is  now  established  there  as  a regular  and 
fairly  common  migrant  in  varying  numbers. 
Several  hundred  migrants  have  also  been  re- 
corded annually  in  August  and  September, 
with  occasional  birds  recorded  in  October 
(Leven  et  al.  1994;  Carey  et  al.  1995,  1996). 

From  September  to  October  1994  this  spe- 
cies was  common  within  the  Riau  Archipela- 
go, Indonesia,  and  a single  bird  was  also  re- 
corded there  in  March  1996  (Rajathurai  1996). 
Kennerley  and  Ollington  (1998)  observed 
small  numbers  between  18  September  and  25 
April  in  the  Straits  of  Malacca  and  in  the  seas 
around  the  island  of  Grand  Karimun  with  a 
maximum  number  of  15  positively  identified 
birds  on  19  January  1996,  with  more  than  100 
distant  terns  also  present. 

Small  numbers  of  Aleutian  Terns  were  also 
observed  off  the  east  coast  of  Singapore  in 
September  and  October  1994  (Kennerley  and 
Ollington  1998).  No  records  were  reported 
during  1995  but  a concerted  effort  to  locate 
this  species  in  1996  resulted  in  at  least  15 
birds  sighted  13  October  in  the  Straits  of  Sin- 
gapore between  Jurong  and  the  Horsburgh 
Light  (Kennerley  and  Ollington  1998). 

We  report  here  the  first  probable  records 
from  Java,  Bali,  and  Sulawesi  in  the  Republic 
of  Indonesia  and  identify  a likely  proportion 
of  this  species’  wintering  range. 

On  30  November  1997  between  Labuan 
and  the  north  coast  of  Ujung  Kulon  National 
Park,  at  the  western  tip  of  Java,  from  a boat 
we  observed,  about  20  probable  adult  Aleu- 
tian Terns  in  groups  of  3-6,  all  in  non-breed- 


559 


560 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


ing  plumage.  We  obtained  good  views  of  these 
birds  in  flight  and  perched  as  they  fished  to- 
gether with  other  seabirds  among  and  from  the 
numerous  fish-trap  platforms.  Similarly,  on  4 
December  1997  we  also  observed,  from  a boat 
a flock  of  approximately  35  probable  Aleutian 
Terns  during  the  late  afternoon.  They  were 
feeding  just  off  the  west  coast  of  Bali,  within 
the  maritime  boundary  of  Bali  Barat  National 
Park.  On  13  December  1997  we  observed  a 
group  of  15  probable  Aleutian  Terns  and 
Crested  Terns  {Sterna  bergii)  fishing  at  the 
mouth  of  the  Bone  River,  Gorontalo,  Sula- 
wesi. At  the  same  time,  about  200  unidentified 
terns  were  feeding  offshore  along  the  interface 
between  the  muddy  river  water  and  the  clear 
sea. 

We  observed  the  flocks  closely  for  10-30 
minutes,  as  two  of  the  flocks  of  the  birds  often 
fed  close  to  our  boat.  N.P.H.  is  familiar  with 
the  species  on  its  breeding  grounds  and  has 
observed  it  twice  in  Alaska  and  once  in  Si- 
beria. Both  of  us  are  familiar  with  the  black- 
billed race  of  Common  Tern  {Sterna  hirundo 
longipennis),  which  the  Aleutian  Tern  most 
resembles.  All  the  birds  we  saw  appeared  to 
be  adults  in  winter  or  near  winter  plumage 
with  a sharply  defined  pattern  of  extensive 
white  forehead,  with  white  extending  onto  the 
crown  of  some  individuals  creating  a bald- 
headed  appearance.  Some  individuals  had  a 
black  nape  with  whitish  streaking  in  the  upper 
edge  of  the  black  nape  (for  an  example  see 
photographs  6 and  10  in  Kennerley  and  Ol- 
lington  1998).  The  wings  and  back  were  dark 
grey  with  a contrasting  pale  rump  and  tail. 
The  bill  was  black  on  all  the  birds  as  were  the 
legs  of  perched  birds  off  the  west  coast  of 
Java.  Flight  appeared  to  be  slower  than  that 
of  a Common  Tern  and  feeding  was  accom- 
plished more  by  dipping,  similar  to  Chlidon- 
ias  terns  or  the  tropical  Sooty  and  Bridled 
terns  {S.  fuscata  and  S.  anaethetus).  Unfortu- 
nately we  did  not  take  note  of  the  underwing 
pattern  which  Kennerley  and  Ollington  (1998) 
and  Kennerley  (pers.  comm.)  demonstrate  to 
be  a diagnostic  field  identification  character 
for  the  Aleutian  Tern  in  winter  plumage. 


While  our  records  of  the  Aleutian  Tern  in 
Java,  Bali  and  Sulawesi  are  not  conclusive,  we 
have  presented  our  records  here  in  order  to 
draw  attention  to  a possible  wintering  area  for 
this  species  and  to  encourage  field  workers 
visiting  Southeast  Asia  and  Wallacea  to  be 
more  diligent  in  their  examination  of  tern 
flocks,  especially  from  September  to  April. 
Our  records  of  a probable  Aleutian  Tern  in 
Indonesia  together  with  previously  published 
observations  of  this  species  elsewhere  in 
Southeast  Asia  suggest  that  a significant  pro- 
portion of  the  western  Pacific  population  of 
Aleutian  Tern  migrates  along  the  southern 
coast  of  China  and  Southeast  Asia  to  winter 
in  the  islands  of  Indonesia  and  possibly  the 
Philippines. 

ACKNOWLEDGMENTS 

We  are  most  grateful  to  P.  Kennerley  for  his  invalu- 
able comments  on  our  manuscript. 

LITERATURE  CITED 

American  Ornithologists’  Union.  1998.  Checklist  of 
North  American  birds,  seventh  ed.  American  Or- 
nithologists’ Union,  Washington,  D.C. 

Brazil,  M.  A.  1991.  The  birds  of  Japan.  Christopher 
Helm.  London,  U.K. 

Carey,  G.  J.,  D.  A.  Diskin,  P.  J.  Leader,  M.  R.  Leven, 
R.  W.  Lewthwaite,  M.  L.  Chalmers,  P.  R.  Ken- 
nerley, AND  V.  B.  PiCKEN.  1996.  Systematic  list. 
Hong  Kong  Bird  Report  1995:45. 

Carey,  G.  J.,  D.  A.  Diskin,  V.  B.  Picken,  and  P.  J. 
Leader.  1995.  Systematic  list.  Hong  Kong  Bird 
Report  1994:53. 

Dixey,  a.  E.,  a.  Ferguson,  R.  Heywood,  and  A.  R. 
Taylor.  1981.  Aleutian  Tern  on  the  Fame  Islands. 
Brit.  Birds  74:  41  1-416. 

Harrison,  P.  1983.  Seabirds  an  identification  guide. 

Croom  Helm  Ltd.,  Beckenham,  Kent,  U.K. 
Kennerley,  P.  R.,  P.  J.  Leader,  and  M.  R.  Leven. 
1993.  Aleutian  Tern:  the  first  records  for  Hong 
Kong.  Hong  Kong  Bird  Report  1992:107-113. 
Kennerley,  P.  and  R.  Ollington.  1998.  Aleutian 
Terns  in  South-East  Asia.  Oriental  Bird  Club  Bull. 
27:34-41. 

Lee,  D.  S.  1992.  Specimen  records  of  Aleutian  Terns 
from  the  Philippines.  Condor  94:276-279. 

Leven,  M.  R.,  G.  J.  Carey,  and  V.  B.  Picken.  1994. 

Systematic  list.  Hong  Kong  Bird  Report  1993:50. 
Rajathurai,  S.  1996.  The  birds  of  Batam  and  Bintan 
Islands,  Riau  Archipelago.  Kukila  8:86-113. 


SHORT  COMMUNICATIONS 


561 


Wilson  Bull.,  111(4),  1999,  pp.  561-564 


Arthropods  and  Predation  of  Artificial  Nests  in  the  Bahamas: 
Implications  for  Subtropical  Avifauna 

Nancy  L.  Staus'  --^  and  Paul  M.  Mayer^'^ 


ABSTRACT. — Little  is  known  of  nest  predation 
patterns  in  the  dry  subtropics.  We  used  artificial  nests 
to  examine  patterns  of  nest  predation  and  to  identify 
possible  nest  predators  in  the  Bahamas.  Unlike  pre- 
dation patterns  in  temperate  areas,  we  found  no  rela- 
tionship between  predation  rates  and  nest  cover  or  dis- 
tance to  the  road.  Instead,  the  rate  of  nest  predation 
depended  on  distance  to  ocean.  This  result  and  a pho- 
tograph taken  at  a disturbed  nest  implicated  the  giant 
white  land  crab  {Cardisoma  guanhumi)  as  a possible 
nest  predator.  Because  land  crabs  are  prevalent 
throughout  the  subtropics  and  could  potentially  influ- 
ence nesting  behavior,  we  advise  researchers  to  con- 
sider variables  associated  with  land  crabs  when  ex- 
amining nest  predation  in  the  subtropics.  Received  14 
July  1998,  accepted  15  April  1999. 


Nest  predation  studies  are  abundant  in  the 
literature;  most  have  been  conducted  in  north- 
ern, temperate  areas  (reviewed  by  Paton  1994, 
Major  and  Kendal  1996,  Hartley  and  Hunter 
1998).  Although  a few  similar  studies  have 
taken  place  in  the  tropics  (e.g.,  Gibbs  1991, 
Laurance  et  al.  1993)  and  wet  subtropics  (Lat- 
ta  et  al.  1995),  no  such  study  has  been  con- 
ducted in  dry,  subtropical  habitat  where  pred- 
ator species  assemblages  may  be  quite  differ- 
ent. Patterns  of  nest  predation  might  differ  in 
the  dry  subtropics  as  a result  of  differences  in 
numbers  and  species  of  egg  predators. 

Long  Island,  an  outer  island  in  the  southern 
Bahamas  archipelago,  is  characterized  by  dry. 


' James  Ford  Bell  Museum  of  Natural  History  and 
Graduate  Program  in  Conservation  Biology,  100  Ecol- 
ogy Building,  Univ.  of  Minnesota,  St.  Paul,  MN 
55108. 

^ Conservation  Biology  Institute,  800  NW  Starker 
Ave.,  Suite  3 1C,  Corvallis,  OR  97330. 

^Graduate  Program  in  Conservation  Biology,  100 
Ecology  Building,  Univ.  of  Minnesota,  St.  Paul,  MN 
55108. 

U.S.  Environmental  Protection  Agency,  Kerr  Re- 
search Center,  P.O.  Box  1 198,  Ada,  OK  74820. 

^ Corresponding  author;  E-mail; 
nstaus@earthdesign.com 


scrubby  vegetation  and  a relatively  depauper- 
ate fauna.  There  has  been  no  prior  study  to 
examine  avian  nest  predation  on  any  of  the 
Bahama  islands  or  to  determine  which  egg 
predators  are  present.  We  used  artificial  nests 
to  determine  factors  influencing  nest  survival 
of  ground-nesting  birds  and  to  identify  im- 
portant nest  predators  on  Long  Island  and  Hog 
Cay,  Bahamas.  Artificial  nests  are  frequently 
used  in  predation  experiments  where  it  is  as- 
sumed that  they  provide  a reasonable  assess- 
ment of  the  impact  of  predators  on  real  nests 
(Burger  et  al.  1994,  but  see  Major  and  Kendal 
1996).  In  temperate  zone  studies,  nest  preda- 
tion rates  often  varied  with  nest  visibility  (Ma- 
jor and  Kendal  1996)  and  distance  from  edge 
(Paton  1994).  We  conducted  an  experiment  to 
determine  whether  patterns  of  nest  predation 
in  the  Bahamas  were  similar  to  those  observed 
elsewhere  and  to  identify  possible  nest  pred- 
ators. 

STUDY  AREA  AND  METHODS 

We  conducted  our  study  on  the  northern  20  km  of 
Long  Island  and  on  Hog  Cay,  Bahamas.  Long  Island, 
one  of  the  outer  islands  of  the  Bahamas  archipelago, 
is  128  km  long  and  6.4  km  wide  at  its  widest  point. 
Hog  Cay  is  a small  (100  ha),  privately  owned  island 
located  off  the  northern  tip  of  Long  Island.  Both  is- 
lands are  covered  with  dry,  scrubby  vegetation.  Man- 
groves (Rhn.ophora  mangle  and  Avicennia  genninans) 
grow  along  the  coasts  of  both  islands. 

We  observed  7 ground  nesting  bird  species  on  Long 
Island  and  Hog  Cay,  Bahamas.  The  largest  included 
the  West  Indian  Whistling-duck  (Dendrocygna  arho- 
rea)  and  White-cheeked  Pintail  (Anas  hahatnensis  ha- 
hamensis).  Smaller  species  included  Antillean  Night- 
hawks  (Chordeiles  minor).  Common  Ground-Doves 
(Calumhigallina  passerina).  Snowy  and  Wilson's  plo- 
vers (Cliaradrius  ale.xandrinus  and  C.  wilsonia),  and 
Willets  (Catoptrophorus  semipalmatu.s). 

Potential  terrestrial  nest  predators  included  intro- 
duced rats  (Rattus  spp.),  domestic  dogs  (Canis  fami- 
liaris),  and  native  giant  white  land  crabs  (Cardisoma 
guanhumi).  Possible  avian  egg  predators  included 
Laughing  Gulls  (Larus  atricilla).  Yellow-crowned 


562 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


Night-Herons  {Nyctanassa  violacea),  and  Smooth- 
billed Anis  (Crotophaga  ani). 

We  utilized  artificial  ground  nests  in  two  experi- 
mental trials  lasting  from  mid-May  through  July  1995. 
Each  trial  was  30  days,  the  approximate  incubation 
time  for  several  ground  nesting  species  in  the  Bahamas 
(Paterson  1972).  For  each  trial,  two  plots  were  selected 
on  Long  Island  and  one  plot  was  selected  in  similar 
habitat  on  Hog  Cay.  Trial  1 took  place  May  17-June 
19  and  consisted  of  25  and  16  nests  on  Long  Island, 
and  30  nests  on  Hog  Cay.  Trial  2 took  place  June  20- 
July  21  and  consisted  of  24  and  16  nests  on  Long 
Island,  and  30  nests  on  Hog  Cay.  Artificial  nests  were 
placed  in  a grid  pattern  within  each  of  the  six  sites 
such  that  one  edge  of  the  grid  was  located  parallel  to 
and  50  m from  a road.  All  nests  within  the  grid  were 
placed  25  m apart  and  were  randomly  assigned  as 
“hidden”  or  “open.”  Hidden  nests  were  completely 
covered  by  vegetation,  whereas  open  nests  could  be 
seen  from  within  1 m. 

Nests  consisted  of  a shallow  scrape  containing  five 
domestic  chicken  eggs  located  under  thatch  palm 
(Thrinax  microcarpa)  or  a dense  bush.  Although  some 
studies  have  detected  effects  of  egg  size  because  small 
predators  (e.g.,  mice)  were  unable  to  break  larger  eggs 
(Pieman  1988,  Haskell  1995),  we  believe  that  all  po- 
tential predators  in  our  study  sites  were  large  enough 
to  handle  chicken  eggs.  Nests  were  examined  for  sur- 
vivorship at  days  6,  12,  18,  and  24,  and  were  consid- 
ered depredated  if  one  or  more  eggs  was  missing  or 
damaged. 

To  identify  specific  nest  predators,  we  placed  three 
automatic  cameras  with  flash  capability  at  one  nest  on 
each  study  site  during  each  trial.  Cameras  were  trig- 
gered by  a motion-sensitive  mercury  switch  glued  to 
the  bottom  of  the  eggs.  Because  cameras  were  con- 
spicuous, they  were  placed  at  previously  depredated 
nests,  which  were  then  rebaited  with  chicken  eggs. 
Cameras  were  rotated  among  nests  within  study  sites 
on  a weekly  basis. 

We  also  noted  the  remnants  of  eggs  at  the  first  pre- 
dation event  for  each  nest.  Predators  can  sometimes  be 
identified  by  the  type  of  egg  remains  they  leave  behind 
(Reardon  1951;  but  see  Trevor  et  al.  1991 ).  Depredated 
eggs  were  classified  as  missing  or  broken  (portion  of 
an  egg  remaining  in  nest),  and  appearance  of  broken 
eggs  was  also  noted  (e.g.,  many  small  fragments,  half 
shell  remaining). 

We  developed  a logistic  regression  model  to  exam- 
ine the  dependency  of  ne.st  fate  on  nest  type  (i.e.,  hid- 
den or  open),  distance  to  road  (to  examine  edge  ef- 
fects), and  distance  to  the  ocean  (a  variable  as.sociated 
with  land  crab  presence).  Logistic  regression  models 
have  been  used  to  analyze  factors  affecting  the  success 
of  both  natural  (Thomas  ct  al.  1996)  and  artificial  nests 
(Burger  et  al.  1994,  Vander  Haegen  and  DeGraaf 
1996)  and  are  appropriate  when  response  variables  are 
binary  (e.g.,  nest  succc.ss  or  failure)  and  factors  are 
continuous  (e.g.,  distance  to  road/occan;  Hosmer  and 
Lemeshow  1989).  We  determined  the  suitability  of  the 
model  by  using  the  Hosmer  and  Lemeshow  Goodness- 


of-Fit  Test  and  associated  statistic  with  a significance 
level  of  P < 0.05.  Individual  variables  within  the  mod- 
el were  tested  with  the  Wald  statistic.  Analyses  were 
conducted  with  SAS  (Windows  version  6.12;  SAS  In- 
stitute Inc.,  Cary,  North  Carolina). 

RESULTS 

Of  141  artificial  nests,  99  (70%)  were  dep- 
redated during  the  two  trials  combined.  Our 
overall  regression  model  fit  our  data  (Hosmer 
and  Lemeshow  Goodness-of-Fit  statistic  = 
8.57,  df  = 8,  P = 0.38)  and  was  significant 
(score  statistic  = 21.73,  df  = 3,  P < 0.001). 
Nest  fate  depended  primarily  on  distance  to 
the  ocean  (score  statistic  = 16.4,  df  = 1,  P < 
0.001).  Nests  ranged  from  100-1500  m from 
the  ocean;  the  probability  of  nest  success  in- 
creased with  distance  from  the  ocean.  Nests 
located  farthest  from  the  ocean  (1500  m)  had 
the  greatest  success  (71%),  whereas  those  lo- 
cated 100-325  m from  the  ocean  and  had  an 
average  success  rate  of  23%  (range  7-37%). 

Nest  fate  was  independent  of  nest  type  (hid- 
den or  open;  x“  = 2.56,  df  = 1,  P > 0.05); 
44  (63%)  and  54  (76%)  of  the  nests  were  dis- 
turbed at  hidden  and  open  nests,  respectively. 
In  addition,  nest  fate  was  not  associated  with 
distance  to  roads  (x^  = 0.18,  df  = 1,  P > 
0.05). 

One  camera  successfully  captured  activity 
near  a nest.  A photograph  was  taken  of  a giant 
white  land  crab  near  two  damaged  eggs  in  a 
nest  on  Hog  Cay.  It  was  not  clear  whether  the 
crab  broke  the  eggs,  or  found  them  after  they 
had  been  broken. 

Of  the  190  eggs  from  92  nests  on  which 
data  were  collected,  73  (38%)  were  missing 
and  117  (62%)  were  broken  at  the  first  nest 
check  after  predation.  Thirteen  (12.5%)  of  the 
broken  eggs  were  attributed  to  rat  predation 
(Flack  and  Lloyd  1976,  pers.  obs.). 

DISCUSSION 

Although  the  results  of  artificial  nest  ex- 
periments conducted  in  the  temperate  zone  are 
often  inconsistent,  a few  common  patterns 
have  emerged.  In  general,  predation  rates  -are 
higher  in  nests  that  are  more  visible  and  in 
habitats  with  little  understory  cover  (Major 
and  Kendal  1996,  Hartley  and  Hunter  1998). 
In  addition,  Paton  (1994)  found  a negative  re- 
lationship between  nest  predation  rates  and 
distance  from  habitat  edge  in  most  of  the  14 


SHORT  COMMUNICATIONS 


563 


artificial  nest  studies  he  re-analyzed.  We 
found  no  such  patterns  in  the  Bahamas. 

In  our  study,  the  only  environmental  vari- 
able that  successfully  predicted  nest  fate  was 
distance  to  the  ocean,  suggesting  that  nests 
were  depredated  by  a species  residing  in  or 
near  the  water.  Our  photograph  of  a giant 
white  land  crab  at  a nest  suggests  that  land 
crabs  were  depredating  artificial  nests. 

Although  land  crabs  are  terrestrial  and  do 
not  rely  on  the  ocean  directly  on  a daily  basis, 
they  do  need  some  source  of  water  nearby  to 
survive  and  the  females  migrate  to  the  ocean 
for  reproduction  to  release  larvae  (Wolcott 

1988) .  Given  their  reliance  on  water  for  re- 
production and  oxygen  exchange,  land  crabs 
are  generally  limited  to  low-lying  areas  near 
mangroves,  swamps,  and  streams,  and  are 
rarely  more  than  a few  kilometers  from  the 
sea  (Wolcott  1988). 

Giant  white  land  crabs  were  abundant  on 
both  Hog  Cay  and  Long  Island  but  did  not 
appear  to  be  associated  with  edge  habitats. 
Their  mostly  vegetarian  diets  and  ground  for- 
aging habits  could  bring  them  into  contact 
with  ground  nests  regardless  of  whether  the 
nests  were  hidden  or  open.  Experiments  in- 
volving captive  land  crabs  revealed  that  crabs 
were  able  to  crack  and  consume  eggs  of  var- 
ious sizes  corresponding  to  the  egg  sizes  of 
ground  nesting  birds  in  the  Bahamas  (Staus 
and  Barnwell  1996).  This  study  also  indicated 
that  crabs  could  be  responsible  for  both  bro- 
ken and  missing  eggs. 

Although  some  studies  document  chick  pre- 
dation by  giant  white  land  crabs  (Gnam  1991), 
ours  is  the  first  study  to  implicate  Cardisoma 
crabs  as  egg  predators.  Egg  eating  behavior 
has  been  documented  in  several  Gecarcinus 
spp.,  land  crabs  in  the  same  family  as  C. 
guanhiimi  (Rockwell  1932,  Atkinson  1985, 
Burger  and  Gochfeld  1988,  Burger  et  al. 

1989) .  Other  egg  eating  species  include  hermit 
crabs  (Coenobita  rugoscr,  Atkinson  1985,  Bur- 
ger et  al.  1989),  coconut  crabs  (Birgus  latro\ 
Atkinson  1985),  and  ghost  crabs  (Ocypode 
quadrata;  Watts  and  Bradshaw  1995). 

It  has  been  suggested  that  land  crabs  play 
an  ecological  role  similar  to  that  of  rats,  and 
that  crabs  may  have  exerted  considerable  in- 
fluence on  tropical  island  avifuanas  (Atkinson 
1985).  For  example,  after  examining  the  fossil 
record,  Olson  (1981)  hypothesized  that  Ge- 


carcinus  land  crabs  in  the  South  Atlantic  may 
have  prevented  the  colonization  of  some  is- 
lands by  burrowing  and  ground-nesting  pe- 
trels. Burger  and  Gochfeld  (1988)  provided 
evidence  that  Roseate  Terns  {Sterna  dougallii) 
in  Puerto  Rico  chose  nest  sites  far  from  suit- 
able land  crab  (G.  ruricola)  habitat.  The  wide- 
ly distributed  giant  white  land  crab  might  have 
a similar  effect  on  bird  populations  within  its 
range. 

Our  results  suggest  that  egg  predation  pat- 
terns and  predator  species  assemblages  in  the 
dry  subtropics  may  be  different  than  those  in 
northern  temperate  areas.  Specifically,  land 
crabs  may  play  a significant  role  as  egg  pred- 
ators. In  the  future,  we  urge  researchers  to 
consider  environmental  variables  associated 
with  the  presence  of  land  crabs  (e.g.,  density 
of  crab  burrows,  altitude,  distance  from  ocean) 
when  examining  nest  predation  in  the  sub- 
tropics. 

ACKNOWLEDGMENTS 

This  research  was  funded  by  grants  from  Ducks  Un- 
limited (Bahamas),  Birdlife  International,  Dayton  and 
Wilkie  Funds  for  Natural  History  and  Behavior,  As- 
sociation of  Field  Ornithologists,  and  World  Nature 
Association.  Permission  to  work  in  the  Bahamas  was 
granted  by  M.  Isaacs  of  the  Ministry  of  Agriculture. 
Mr.  and  Mrs.  P.  Graham  kindly  allowed  us  to  work  on 
Hog  Cay.  We  thank  N.  Graham  for  excellent  field  as- 
sistance during  this  project.  Thanks  also  to  L.  Gape  of 
the  Bahamas  National  Trust,  P.  Maillis,  M.  Lightbourn, 
C.  Crissey,  and  E.  Wilson  for  their  friendship  and  lo- 
gistical support  throughout  this  project.  We  gratefully 
acknowledge  K.  Beal  for  her  assistance  with  the  sta- 
tistical analysis.  A special  thanks  to  D.  F.  McKinney 
for  guidance  and  support  throughout  the  project,  and 
helpful  comments  on  earlier  drafts  of  this  manuscript. 

LITERATURE  CITED 

Atkinson,  I.  A.  E.  1985.  The  spread  of  commensal 
species  of  Rattus  to  oceanic  islands  and  their  ef- 
fects on  island  avifaunas.  Pp.  35-81  in  Conser- 
vation of  island  birds  (P  J.  Moors,  Ed.).  ICBP 
Technical  Publication  No.  3,  Page  Bros.  Ltd.,  Nor- 
wich, England. 

Burger,  J.  and  M.  Gochfeld.  1988.  Nest-site  selec- 
tion by  Roseate  Terns  in  two  tropical  colonies  on 
Culebra,  Puerto  Rico.  Condor  90:843—851. 
Burger,  J.,  M.  Gochfeld.  J.  E.  Saliva,  D.  Gochfeld, 
AND  H.  Morales.  1989.  Antipredator  behaviour  in 
nesting  Zenaida  Doves  (Zenaidci  auhta):  Parental 
investment  or  offspring  vulnerability.  Behaviour 
111:129-143. 

Burger,  L.  D.,  L.  W.  Burger,  Jr.,  and  J.  Faaborg. 
1994.  Effects  of  prairie  fragmentation  on  preda- 


564 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


tion  of  artificial  nests.  J.  Wild!.  Manage.  58:249- 
254. 

Flack,  J.  A.  D.  and  B.  D.  Lloyd.  1976.  The  effect  of 
rodents  on  the  breeding  success  of  the  South  Is- 
land Robin.  Pp.  59—66  in  The  ecology  and  control 
of  rodents  in  New  Zealand  nature  reserves  (P.  R. 
Dingwall,  I.  A.  E.  Atkinson,  and  C.  Hay,  Eds.). 
Departments  of  Land  and  Survey,  Wellington, 
New  Zealand. 

Gibbs,  J.  P.  1991.  Avian  nest  predation  in  tropical  wet 
forest;  an  experimental  study.  Oikos  60:155-161. 

Gnam,  R.  S.  1991.  Natural  history  of  the  Bahama  Par- 
rot Amazona  leucocephalo  hahamensis  on  Abaco 
Island.  Proc.  Symp.  Nat.  Hist.  Bahamas  4:67-74. 

Hartley,  M.  J.  and  M.  L.  Hunter,  Jr.  1998.  A meta- 
analysis of  forest  cover,  edge  effects,  and  artificial 
nest  predation  rates.  Conserv.  Biol.  12:465-469. 

Haskell,  D.  G.  1995.  A reevaluation  of  the  effects  of 
forest  fragmentation  on  rates  of  bird-nest  preda- 
tion. Conserv.  Biol.  5:1316-1318. 

Hosmer,  D.  W.  and  S.  Lemeshow.  1989.  Applied  lo- 
gistic regression.  John  Wiley  and  Sons,  New  York. 

Latta,  S.  C.,  j.  M.  Wunderle,  Jr.,  E.  Terranova, 
and  M.  Pagan.  1995.  An  experimental  study  of 
nest  predation  in  a subtropical  wet  forest  follow- 
ing hurricane  disturbance.  Wilson  Bull.  107:590— 
602. 

Laurance,  W.  E,  j.  Garesche,  and  C.  W.  Payne. 
1993.  Avian  nest  predation  in  modified  and  nat- 
ural habitats  in  tropical  Queensland:  an  experi- 
mental study.  Wildl.  Res.  20:711—723. 

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. 


Olson,  S.  L.  1981.  Natural  history  of  vertebrates  on 
the  Brazilian  islands  of  the  mid  south  Atlantic. 
Nat.  Geog.  Soc.  Res.  Rep.  13:481-492. 

Paterson,  A.  1972.  Birds  of  the  Bahamas.  Durell  Pub- 
lications, Kennebunkport,  Maine. 

Paton,  P.  W.  C.  1994.  The  effect  of  edge  on  avian  nest 
success:  how  strong  is  the  evidence?  Conserv. 
Biol.  8:17-26. 

PiCMAN,  J.  1988.  Experimental  study  of  predation  on 
eggs  of  ground-nesting  birds:  effects  of  habitat 
and  nest  distribution.  Condor  90:124-131. 

Reardon,  J.  D.  1951.  Identification  of  waterfowl  nest 
predators.  J.  Wildl.  Manage.  15:386—395. 

Rockwell,  R.  H.  1932.  Southward  through  the  dol- 
drums. Nat.  Hist.  32:424—436. 

Staus,  N.  L.  and  E H.  Barnwell.  1996.  The  tropical 
Atlantic  land  crab  Cardi.soma  guanhumi  feeds  on 
bird  eggs.  Am.  Zool.  36:80A. 

Thomas,  B.  G.,  E.  P.  Wiggers,  and  R.  L.  Clawson. 
1996.  Habitat  selection  and  breeding  status  of 
Swainson’s  Warblers  in  southern  Missouri.  J. 
Wildl.  Manage.  60:611-616. 

Trevor,  J.  T,  R.  W.  Seabloom,  and  R.  D.  Sayler. 
1991.  Identification  of  mammalian  predators  at  ar- 
tificial waterfowl  nests.  Prairie  Nat.  23:93-99. 

Vander  Haegen,  W.  M.  and  R.  M.  DeGraaf.  1996. 
Predation  on  artificial  nests  in  forested  riparian 
buffer  strips.  J.  Wildl.  Manage.  60:542-550. 

Watts,  B.  D.  and  D.  S.  Bradshaw.  1995.  Ghost  crab 
preys  on  Piping  Plover  eggs.  Wilson  Bull.  107: 
767-768. 

Wolcott,  T.  G.  1988.  Ecology.  Pp.  55-96  in  Biology 
of  the  land  crabs  (W.  W.  Burggren  and  B.  R.  Mc- 
Mahon, Eds.).  Cambridge  Univ.  Press,  Cam- 
bridge, U.K. 


Wilson  Bull:  111(4),  1999,  pp.  564-569 

Notes  About  the  Distribution  of  Pauxi  pauxi  and 
Aburria  aburri  in  Venezuela 

Jose  L.  Silva'  - 


ABSTRACT. — In  this  paper  I review  the  current 
distribution  of  the  Northern  Helmeted  Curassow 
(Baicxi  pciuxi)  and  the  Wattled  Guan  {Aburria  aburri) 
in  Venezuela.  The  historical  range  of  P.  pau.xi  was  re- 
duced as  a result  of  human  population  growth  and  hab- 
itat perturbations.  The  current  distribution  corresponds 


' NYZS  The  Wildlife  Con.servation  Society,  185th 
& Southern  Blvd.,  Bronx,  New  York  10460-1099. 

2 Current  address:  Univ.  of  Florida,  Wildlife  Ecolo- 
gy and  Con.servation  Dept.,  RO.  Box  I 10430,  303 
Newins — Ziegler  Hall,  Gainesville,  FL  32611-0430; 
E-mail:  jo.selugo@ufl.edu 


principally  with  18  national  parks  located  from  the 
northern  coastal  mountains  of  central  Venezuela  to  the 
Andes  Cordillera  and  Sierra  de  Perija.  Pau.xi  pau.xi  was 
recorded  only  in  three  localities  outside  national  parks 
and  may  have  expanded  from  its  historical  distribution 
in  the  eastern  part  of  the  country.  Aburria  aburri  was 
recorded  in  Sierra  de  Perija  and  western  Merida  to 
southern  Tachira,  including  four  new  localities;  three 
in  national  parks.  Both  species  are  endangered  in  Ven- 
ezuela and  their  survival  will  depend  on  environmental 
education  programs  and  enforcement  of  the  law.  Re- 
ceived 9 Feb.  199H,  accepted  20  .July  1999. 


SHORT  COMMUNICATIONS 


565 


The  Wildlife  Conservation  Society  of  the 
New  York  Zoological  Society  funded  a study 
on  human  impacts  on  game  species  in  pro- 
tected areas  of  Venezuela  from  1985  to  1990 
(Silva  and  Strahl  1991,  1994,  1996,  1997). 
During  1985-1996,  censuses  and  interviews 
were  conducted  and  new  data  about  the  dis- 
tribution of  Pauxi  pauxi  (Northern  Helmeted 
Curassow)  and  Aburria  aburri  (Wattled  Guan) 
were  collected.  My  objective  in  this  paper  is 
to  present  these  data  and  review  the  status  of 
P.  pauxi  and  A.  aburri  in  Venezuela. 

In  Venezuela  Pauxi  pauxi  ranges  from  the 
northern  coastal  mountains  of  central  Vene- 
zuela to  the  Andes  Cordillera  and  Sierra  de 
Perija  in  rain  forest  and  cloud  forest.  Most  au- 
thors (Wetmore  and  Phelps  1943;  Phelps  and 
Phelps  1958,  1962;  Delacour  and  Amadon 
1973;  Meyer  de  Schauensee  and  Phelps  1978; 
Collar  et  al.  1992;  Rodriguez  and  Rojas-Suar- 
ez  1995)  cited  the  states  and  localities  of  the 
historical  range  as  follows:  P.  p.  pauxi:  south- 
ern Miranda  state  in  Cerro  Negro  (Guatopo 
National  Park);  north  Caracas  in  El  Calvario; 
Distrito  Federal;  coastal  mountains  in  Aragua 
state  (Henri  Pittier  National  Park);  Carabobo 
state  in  Valencia,  San  Esteban,  and  Montal- 
ban;  east  Falcon  to  Yaracuy  state  in  Tucacas, 
Nirgua,  mountains  inland  from  Aroa,  and  La- 
gunita  de  Aroa;  Lara  state  in  Cubiro,  and  Ya- 
cambu  National  Park;  from  northern  Merida 
to  southern  Tachira  state  in  Montana  de  Li- 
mones.  La  Azulita,  and  Burgua.  Pauxi  p.  gil- 
liardi:  Zulia  state  from  southern  Sierra  de  Per- 
ija (Sierra  de  los  Motilones)  to  southern  Rio 
Tucuco,  in  Fila  Macoita,  Campamento  Avispa, 
Cerro  Yin-taina,  upper  Rio  Negro,  and  La  Sa- 
bana.  A continuous  distribution  in  the  histor- 
ical range  was  assumed  (Fig.  lA)  because  of 
historical  records,  and  Central  Cordillera  and 
Los  Andes  Cordillera  were  almost  a continu- 
ous forest  in  the  past. 

The  habitat  available  for  P.  pauxi  has  been 
greatly  reduced  as  a result  of  deforestation, 
fragmentation,  and  habitat  alteration.  Almost 
all  the  remnant  forest  available  in  northern 
Venezuela  was  decreed  as  national  parks  by 
the  Venezuelan  government.  Consequently, 
these  national  parks  are  isolated.  1 found  that 
the  current  distribution  of  P.  pauxi  mainly  co- 
incided with  the  distribution  of  national  parks 
situated  in  its  historical  range  (Fig.  IB),  as 
well  as  new  localities  such  as  Sierra  de  San 


Luis,  Cueva  Quebrada  del  Toro,  and  Tirgua 
National  Parks.  Although  it  was  reported  in 
Morrocoy  National  Park  (Collar  et  al.  1992, 
Wege  and  Long  1993),  according  to  the  rang- 
ers, it  was  no  longer  present  in  the  park  in 
1996. 

Pauxi  pauxi  is  rare  in  national  parks  and 
almost  extinct  outside  national  parks  because 
hunting  pressure  is  highest  outside  the  parks 
(Silva  and  Strahl  1991,  1996,  1997).  The  few 
locations  where  P.  pauxi  was  found  outside 
national  parks  included  the  Sanchon  River 
Hydraulic  Reserve  (10°  24'  N,  68°  09'  W),  the 
Cojedes  River  Protectoral  High  Basin  (10°  24' 
N,  68°  15'  W)  and  Finca  El  Jaguar  (10°  26' 
N,  68°  59'  W). 

From  interviews  with  hunters  I found  that 
P.  pauxi  probably  existed  or  may  still  live  in 
eastern  Venezuela.  A hunter  in  Teresen  (Mon- 
agas  State)  narrated  the  size,  color  pattern,  and 
helmeted  color  of  this  species,  and  imitated  its 
booming  song.  He  recognized  the  bird  from  a 
set  of  cracid  pictures.  Pauxi  pauxi  was  seen 
in  La  Hormiga  (9°  54'  N,  62°  58'  W),  Cano 
Payanuco,  Guarapiche  Forest  Reserve  (Sucre 
and  Monagas  States)  between  1968  and  1973. 
Because  only  1 of  25  interviewed  hunters  in 
Teresen  saw  a P.  pauxi,  and  saw  it  only  once, 
this  should  not  be  interpreted  as  range  exten- 
sion. The  nearest  locality  of  the  historical  dis- 
tribution (Guatopo  National  Park)  is  approxi- 
mately 405  km  from  Guarapiche  Forest  Re- 
serve, and  this  separation  is  settled  with  towns 
and  cities.  More  likely  the  former  distribution 
record  was  incomplete.  Pauxi  pauxi  is  very 
likely  to  be  extinct  in  Guarapiche  because  of 
high  hunting  pressure. 

An  interesting  characteristic  of  P.  pauxi  is 
the  brown  phenotype  that  sometimes  occurs 
in  females.  Males  and  females  are  typically 
black  with  a white  belly.  Hunters  call  the 
brown  morph  “Canaguey”  or  “Pauji  Amaril- 
lo.” It  was  reported  in  the  Sierra  de  Perija, 
where  two  specimens  were  collected  between 
1941  and  1957  (Delacour  and  Amadon  1973). 
Here  I report  26  new  localities  of  the  brown 
morph  seen  between  1949  and  1993  (Table  1). 
Of  the  34  birds  sighted,  a single  brown  phe- 
notype was  seen  with  one  black  phenotype  at 
Fila  Real  (1975),  one  with  two  black  pheno- 
types at  Casa  de  Tejas  (1980),  and  one  with 
seven  black  phenotypes  at  El  Corazon  (1988). 
Two  brown  phenotypes  were  seen  with  two 


566 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


FIG.  I . A.  Historical  range  of  Pait.xi  pau.xi  according  to  Delacour  and  Amadon  { 1973),  Meyer  de  Schauensee 
and  Phelps  (1978),  Collar  and  coworkers  (1992),  and  Rodriguez  and  Rojas-Suarez  (1995).  B.  Cunent  range  of 
PaiLxi  pau.xi  in  Venezuela.  Boundaries  of  national  parks  are  shown,  but  the  species  may  not  be  distributed 
through  the  entire  park.  Numbers  are  national  parks,  Roman  numerals  are  other  localities,  and  letters  are  states 
with  the  exception  of  Caracas  (Ca),  the  capital  of  Venezuela.  Code;  1 = Guatopo  (122,464  ha).  2 = El  Avila 
(81,800  ha).  3 = Macarao  (15,000  ha).  4 = Henri  Pittier  (107,800  ha).  5 = San  Esteban  (43,500  ha).  6 = 
Morrocoy  (32,090  ha).  7 = Siena  de  San  Luis  (2(),()()()  ha).  8 = Cueva  Quebrada  del  Toro  (4,885  ha).  9 = 
Yurubi  (23,670  ha).  10  = Tirgua.  1 1 = Terepaima  (18,650  ha).  12  = Yacambii  (14,580  ha).  13  = Dinira  (42,000 
ha).  14  = Guaramacal  (21.000  ha).  15  = Sierra  de  La  Culata  (200,400  ha).  16  = Siena  Nevada  (276,446  ha). 
17  = Tapo — Caparo.  18  = El  Tama  (109,000  ha).  19  = Siena  de  Perija  (295,288  ha).  I = Sanchon  River 
Hidraulic  Re.serve  (8,100  ha).  11  = Cojedes  River  Protectoral  High  Basin  (276,000  ha).  Ill  = Finca  El  Jaguar 
(I6,()()()  ha).  IV  = Guarapiche  Forest  Reserve  (576,500  ha).  Mo  = Monagas.  S = Sucre.  G = Guarico.  Mi  = 
Miranda.  Ca  = Caracas.  A = Aragua.  C = Carabobo.  Y = Yaracuy.  F = Falcon.  L = Lara.  Tr  = Trujillo.  Me 
= Merida.  T = Tachira.  Z = Zulia. 


SHORT  COMMUNICATIONS 


567 


TABLE  1. 

Brown  phenotype  of  Pauxi  pauxi  recorded 

in  Venezuela. 

Locality 

Date 

Location 

Coordinates 

n 

Henri  Pittier 

1975 

La  Glorieta 

10°  28'N  67°45'W 

1 

Henri  Pittier 

03/1984 

La  Regresiva 

I0°22'N  67°44'W 

1 

Henri  Pittier 

1993 

El  Saltico 

a 

2 

Henri  Pittier 

1993 

Los  Riitos 

— 

2 

San  Esteban 

1955 

Burro  Sin  Cabezas 

— 

1 

San  Esteban 

1960 

Buito  Sin  Cabezas 

— 

1 

San  Esteban 

1986-87 

Flor  Amarillo 

— 

1 

San  Esteban 

1987 

Flor  Amarillo 

— 

1 

San  Esteban 

07/1989 

El  Tanque 

— 

1 

San  Esteban 

1991 

San  Felipe 

— 

1 

San  Esteban 

7 

El  Dique 

10°  18'N  67°59'W 

1 

San  Esteban 

7 

La  Panta  (Qda.  Yaguas) 

— 

1 

San  Esteban 

7 

Ranchitos 

— 

2 

San  Esteban 

7 

La  Manguera 

— 

2 

Terepaima 

1949 

Los  Portones 

9°  52'N  69°  20'W 

1 

Terepaima 

1975 

Fila  Real 

9°  55'N  69°  16'W 

1 

Terepaima 

04/1992 

Fila  Real 

9°53'N  69°  17'W 

1 

Yacambu 

1979 

El  Blanquito,  Qda.  La  Toma 

9°42'N  69°  34'W 

1 

Yacambu 

1983 

Barro  Amarillo 

— 

1 

Yacambu 

1988 

El  Corazdn 

— 

1 

Yacambu 

1988 

La  Canada 

— 

1 

Yacambu 

06/1992 

El  Blanquito 

9°  42'N  69°  34'W 

1 

Yacambu 

7 

La  Escalera 

9°  42'N  69°  30'W 

1 

Yacambu 

7 

La  Postora 

9°41'N  69°37'W 

1 

Yacambu 

7 

El  Blanquito 

9°  42'N  69°  34'W 

1 

Yacambu 

7 

Cerro  Blanco 

9°  37'N  69°  30'W 

1 

Sierra  Nevada 

7 

Alto  de  la  Aguada 

8°  37'N  70°40'W 

1 

Sierra  Nevada 

7 

San  Benito 

8°  40'N  70°  37'W 

1 

Rio  Sanchon 

1980 

La  Cumbre  del  Cacho 

— 

1 

Rio  Sanchon 

11/1980 

Casa  de  Tejas 

— 

1 

“ Name  of  localities  do  not  appear  on  maps  because  they  are  local  names  used  by  hunters  and  the  exact  locations  are  unknown. 


black  phenotypes  at  El  Saltico  and  Los  Riitos 
in  1993. 

Aburria  aburri  was  mainly  recorded  in  the 
western  part  of  Venezuela  (Fig.  2).  The  his- 
torical range  was  reported  to  be  in  the  Sierra 
de  Perija  and  west  Merida  to  southern  Tachira 
in  rain  and  cloud  forest  (Delacour  and  Ama- 
don  1973,  Meyer  de  Schauensee  and  Phelps 
1978,  Rodriguez  and  Rojas-Suarez  1995).  The 
current  distribution  of  A.  aburri  indicates  that 
the  record  of  the  historical  range  may  have 
been  incomplete.  Aburria  aburri  was  recorded 
in  Sierra  Nevada  National  Park  (54  inter- 
viewed hunters),  in  Terepaima  National  Park 
(observed),  and  in  Yacambu  National  Park 
and  the  basin  of  Yacambu  River  (S.  Boher, 
pers.  comm.,  and  338  interviewed  hunters). 
These  were  new  distribution  records,  but  they 
did  not  suggest  an  extension  of  the  historical 
range  because  hunters  over  60  years  old  hunt- 


ed A.  aburri  since  they  were  young.  Perhaps, 
the  historical  range  was  continuous.  Although 
Rodriguez  and  Rojas-Suarez  (1995)  stated  that 
A.  aburri  probably  was  found  in  the  eastern 
part  of  Costa  Cordillera,  they  did  not  mention 
the  source  of  their  information. 

The  present  status  of  P.  pauxi  and  A.  aburri 
is  worrisome.  According  to  the  population 
censuses  (Silva  and  Strahl  1991,  1997)  and 
the  interviews  (Silva  and  Strahl  1996),  P. 
pauxi  and  A.  aburri  have  very  low  densities 
with  A.  aburri  being  more  rare  than  P.  pauxi. 
Both  species  were  considered  Endangered  by 
the  Cracid  Specialist  Group  (Strahl  et  al. 

1994)  and  by  researchers  of  a recent  study  in 
Venezuela  (Rodriguez  and  Rojas-Suarez 

1995) .  Habitat  destruction  and  illegal  hunting 
are  the  principal  causes  of  the  decline  in  pop- 
ulation of  both  species,  and  their  conservation 
will  rely  on  hunter  education  (Silva  and  Pel- 


568 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


(B) 


FIG.  2.  Historical  (A)  and  current  distribution  (B)  of  Ahurria  aburri  in  Venezuela.  Abbreviations  are  the 
same  as  Fig.  1. 


legrini  1996,  Silva  1997)  and  enforcement  of 
the  law. 

ACKNOWLEDGMENTS 

NYZS  The  Wildlife  Conservation  Society  funded 
the  main  projects.  I thank  hunters  who  collaborated 
with  the  interviews.  A.  Grigss  reviewed  the  manu- 
.script. 

LITERATURE  CITED 

Collar.  N.  J.,  L.  P.  Gonzaga,  N.  Krabbe,  A.  Mad- 
rono Nieto,  L.  G.  Naranjo,  T.  A.  Parker,  III, 
AND  D.  C.  Wege.  1992.  Threatened  birds  of  the 
Americas:  the  ICBP/IUCN  red  data  book  2.  Inter- 
national Council  for  Bird  Preservation,  Cam- 
bridge, U.K. 

Delacour,  J.  and  D.  Amadon.  1973.  Curassows  and 
related  birds.  American  Museum  of  Natural  His- 
tory, New  York. 

Meyer  de  Schauensee,  R.  and  W.  H.  Phelps,  Jr. 
1978.  A guide  to  the  birds  of  Venezuela.  Princeton 
Univ.  Press,  Princeton,  New  Jersey. 

Phelps,  W.  H.  and  W.  H.  Phelps,  Jr.  1958.  Lista  de 
Aves  de  Venezuela  con  su  dislribucion  I:  no  pas- 
ser! formes.  Bol.  Soc.  Venezolana  Cienc.  Nat.  19: 
1-317. 

Phelps,  W.  H.  and  W.  H.  Phelps,  Jr.  1962.  Two  new 


subspecies  of  birds  from  Venezuela,  the  rufous 
phase  of  Pauxi  paiixi  and  other  notes.  Proc.  Biol. 
Soc.  Wash.  75:199-204. 

Rodriguez.  J.  P.  and  F.  Rojas-Suarez.  1995.  Libro 
Rojo  de  la  Fauna  Venezolana.  ExLibris,  Caracas, 
Venezuela. 

Silva,  J.  L.  1997.  Proyecto  de  Educacion  para  la  Con- 
servacion  de  la  Familia  Cracidae  en  los  Parques 
Nacionales  al  Norte  de  Venezuela.  Pp.  424-436 
in  The  Cracidae:  their  biology  and  conservation 
(S.  Strahl,  S.  Beaujon,  D.  M.  Brooks,  A.  J.  Be- 
gazo,  G.  Sedaghatkish,  and  F.  Olmos,  Eds.).  Han- 
cock House,  Hong  Kong,  China. 

Silva,  J.  L.  and  N.  Pellegrini.  1996.  Proyecto  de  Ed- 
ucacion para  Cazadores  Furtivos  en  Parques  Na- 
cionales. Natura  106:59-62. 

Silva,  J.  L.  and  S.  Strahl.  1991.  Human  impact  on 
populations  of  guan.  curassow  and  chachalacas  in 
Venezuela.  Pp.  36-52  in  Neotropical  wildlife  use 
and  conservation  (J.  G.  Robinson  and  K.  H.  Red- 
ford,  Eds.).  Chicago  Univ.  Press,  Chicago,  Illinois. 

Silva,  J.  L.  and  S.  Strahl.  1994.  Usos  Folcloricos  de 
la  Fauna  Silvestre  en  Nueve  Parques  Nacionales 
al  Norte  de  Venezuela.  Vida  Silvestre  Neotropical 
3:100-107. 

Silva,  J.  L.  and  S.  Strahl.  1996.  La  Caza  Furtiva  en 
los  Parques  Nacionales  al  Norte  de  Venezuela. 
Vida  Silvestre  Neotropical  5:126-139. 


SHORT  COMMUNICATIONS 


569 


Silva,  J.  L.  and  S.  Strahl.  1997.  Presion  de  Caza 
sobre  Poblaciones  de  Cracidos  en  los  Parques  Na- 
cionales  al  Norte  de  Venezuela.  Pp.  437—448  in 
The  Cracidae:  their  biology  and  conservation  (S. 
Strahl,  S.  Beaujon,  D.  M.  Brooks,  A.  J.  Begazo, 
G.  Sedaghatkish,  and  F.  Olmos,  Eds.).  Hancock 
House,  Hong  Kong,  China. 

Strahl,  S.,  S.  Ellis,  O.  Byers,  and  C.  Plasse.  1994. 
Conservation  assessment  and  management  plan 


for  Neotropical  guans,  curassows,  and  chachala- 
cas.  lUCN/SSC  Conservation  Breeding  Specialist 
Group,  Apple  Valley,  Minnesota. 

Wege,  D.  C.  and  a.  Long.  1993.  Bird  conservation 
priorities;  country  profile,  Venezuela.  BirdLife 
Conserv.  Series  3:1-85. 

Wetmore,  a.  and  W.  H.  Phelps.  1943.  Description  of 
a third  form  of  curassow  of  the  Genus  Pauxi.  J. 
Washington  Acad.  Sci.  33:142-146. 


Wilson  Bull.,  111(4),  1999,  pp.  569-571 

Western  Burrowing  Owls  in  California  Produce  Second  Broods  of  Chicks 

Jennifer  A.  Gervais'  - and  Daniel  K.  Rosenberg' 


ABSTRACT. — We  present  the  first  evidence  that 
western  Burrowing  Owls  are  capable  of  raising  a sec- 
ond brood  of  chicks  within  a nesting  season  once  their 
first  brood  successfully  fledges.  Two  pairs  of  owls  in 
central  California  known  to  have  successfully  fledged 
chicks  from  a first  brood  renested  in  1998,  with  one 
pair  producing  five  additional  fledglings.  Received  29 
March  1999,  accepted  15  July  1999. 


Western  Burrowing  Owls  {Athene  cunicu- 
laria)  are  thought  to  be  declining  throughout 
much  of  their  range  (DeSante  et  al.  1997, 
James  and  Espie  1997).  The  potential  causes 
of  these  declines  vary  with  location,  but  likely 
include  large-scale  habitat  destruction  from 
farming  or  development,  reductions  in  species 
such  as  ground  squirrels  that  create  the  bur- 
rows that  the  owls  use,  and  agricultural  chem- 
icals (James  and  Espie  1997,  Gervais  et  al.  in 
press).  Because  of  the  perceived  threat  to  the 
viability  of  Burrowing  Owl  populations,  the 
species  has  been  listed  as  endangered,  threat- 
ened, or  of  special  management  concern  in  a 
number  of  North  American  states  and  prov- 
inces (Haug  et  al.  1993). 

Effective  conservation  at  the  species  level 
requires  understanding  the  population  dynam- 
ics of  the  species  in  question,  which  in  turn 
means  accurate  estimation  of  demographic  pa- 


' Oregon  Cooperative  Fish  & Wildlife  Research 
Unit,  Dept,  of  Fisheries  and  Wildlife,  Oregon  State 
Univ.,  Corvallis,  OR  97331. 

- Corre.sponding  author;  E-mail: 
gervaisj  @ ucs.orst.edii 


rameters  such  as  survival  and  reproductive 
rates.  These  can  be  used  in  simplified  models 
that  allow  the  examination  of  the  effects  of 
possible  management  actions  or  environmen- 
tal perturbations  on  population  persistence. 
Such  an  approach  has  recently  been  used  for 
the  northern  Spotted  Owl  (Strix  occidentalis\ 
Noon  and  Biles  1990),  and  for  predicting  the 
effects  of  pesticide  exposure  on  wildlife  pop- 
ulations (Caswell  1996,  Calow  et  al.  1997). 

Simulations  of  generalized  life  history  strat- 
egies have  shown  that  for  a species  with  rel- 
atively low  adult  survivorship  and  a short  life 
span,  reproductive  success  may  be  most  influ- 
ential in  maintaining  population  viability  (Em- 
len  and  Pikitch  1989).  This  is  likely  to  be  gen- 
erally true  for  Burrowing  Owls.  They  are  ca- 
pable of  producing  up  to  12  eggs  in  a clutch 
(Haug  et  al.  1993),  and  we  have  observed  up 
to  10  young  fledged  per  nest  in  good  repro- 
ductive years.  In  addition.  Burrowing  Owl  an- 
nual adult  survivorship  appears  to  be  quite 
low,  with  between-year  return  rates  ranging 
from  33-58%  (Haug  et  al.  1993),  and  a lon- 
gevity record  for  a wild  banded  owl  of  8 years 
and  8 months  (Kennard  1975).  If  sensitivity 
analyses  prove  that  the  Burrowing  Owl  fits  the 
predictions  of  the  Emlen-Pikitch  model  (Em- 
len  and  Pikitch  1989)  for  a small,  relatively 
short-lived  species,  then  accurate  assessment 
of  reproductive  potential  of  Burrowing  Owls 
is  essential  to  evaluating  population  processes. 

Only  Florida  Burrowing  Owls  have  been 
known  to  produce  second  broods  within  a sea- 
son (Millsap  and  Bear  1990).  We  report  two 


570 


THE  WILSON  BULLETIN  • Vol.  1 1 1,  No.  4,  December  1999 


instances  of  western  Burrowing  Owl  pairs  at- 
tempting second  broods  after  the  first  brood 
had  successfully  been  fledged  during  the  1998 
breeding  season.  To  our  knowledge,  this  is  the 
first  time  the  production  of  more  than  a single 
brood  per  season  has  been  verified  in  western 
Burrowing  Owls. 

METHODS  AND  RESULTS 

We  have  conducted  demographic  research 
since  1997  on  a population  of  Burrowing 
Owls  at  Naval  Air  Station  Lemoore  (36°  20' 
N,  119°  57'  W),  50  km  southwest  of  Fresno, 
Califoma.  The  naval  air  station  supports  ap- 
proximately 65  breeding  pairs  of  owls,  which 
appear  to  be  winter  residents  (Gervais  and  Ro- 
senberg, unpubl.  data).  Nesting  habitat  on  the 
station  is  primarily  small  patches  of  exotic  an- 
nual grasses  along  runway  easements  and  in 
wildlife  areas  surrounded  by  agricultural 
fields.  Wildlife  areas  are  fallow  fields  that  are 
composed  of  exotic  annual  grasses  and  weeds, 
although  they  are  burned  annually.  Approxi- 
mately 75%  of  the  adult  resident  population 
of  owls  is  now  banded  with  U.S.  Fish  and 
Wildlife  Service  bands  and  unique  alpha-nu- 
meric rivet  bands  (Acraft  Bird  Bands,  Edmon- 
ton, Alberta,  Canada). 

Early  in  the  season,  we  collected  eggs  for 
use  in  an  ongoing  toxicology  study  (Gervais 
et  al.  in  press).  One  sampled  burrow  contained 
at  least  four  eggs  and  the  incubating  female 
on  19  April  1998.  At  that  time,  we  identified 
the  female  from  her  bands  and  we  removed 
one  egg.  Her  mate  also  was  previously  banded 
and  was  identified  by  resighting  his  bands  ear- 
ly in  the  nesting  season.  This  nest  successfully 
fledged  two  chicks  in  early  June.  We  recap- 
tured the  female  owl  on  14  June  using  a 
mouse  baited  spring  net.  At  capture,  she 
weighed  198  g,  well  above  the  150  g average 
for  this  species  (Haug  et  al.  1993),  and  was 
noticeably  swollen  in  the  lower  abdomen.  Her 
brood  patch  was  well  developed  and  vascu- 
larized, suggesting  nesting  activity. 

To  verify  that  this  female  was  indeed  relay- 
ing, we  used  an  infrared  burrow  probe  (Chris- 
tensen Designs,  Manteca,  California)  to  ex- 
amine the  burrow  on  16  June.  We  observed 
the  two  fledged  chicks  in  the  entrance  to  the 
nesting  chamber,  but  were  unable  to  see  be- 
yond them.  The  burrow  entrance  had  fresh 
decorations  of  coyote  dung  and  the  nest  tunnel 


was  lined  with  similar  debris.  The  adults  were 
observed  at  the  burrow  entrance  throughout 
the  next  few  weeks;  individual  identity  was 
confirmed  using  their  color  bands. 

We  examined  the  nest  again  on  27  June, 
and  observed  four  eggs  in  the  nest  chamber 
after  the  female  flushed  from  the  burrow  en- 
trance. We  removed  two  eggs  through  an  ac- 
cess hole  originally  dug  for  the  toxicological 
study  egg  sampling  (Gervais  et  al.,  in  press). 
The  eggs  were  cool,  but  the  shells  were  very 
clean  and  candling  revealed  clear  egg  contents 
with  no  visible  development.  The  eggs  were 
returned  to  the  nest  after  inspection  and  the 
access  holes  covered  again  with  dirt  and 
boards.  We  do  not  believe  these  eggs  were 
from  the  previous  nesting  attempt  because  of 
their  clear  contents  and  clean  shells.  Eggs  that 
sit  in  the  burrow  for  eight  weeks  would  have 
dark  contents  as  they  began  to  rot  and  shells 
would  be  covered  with  dirt  and  fecal  matter 
from  the  chicks. 

When  we  examined  the  burrow  on  14  July, 
the  eggs  were  gone.  No  owls  were  present  at 
the  burrow  during  that  visit,  although  both 
adults  continued  to  be  sighted  in  the  area 
through  July. 

A second  double  nesting  attempt  also  oc- 
curred in  1998.  We  observed  with  the  infrared 
burrow  probe  a banded  female  owl  in  her  bur- 
row with  nine  eggs  on  16  April;  she  raised  one 
chick  to  fledging  after  the  disappearance  of  her 
mate.  We  observed  this  same  female  at  the 
same  burrow  entrance  in  early  September  with 
five  buffy  breasted  chicks.  These  chicks  clearly 
had  recently  emerged  from  their  burrows,  be- 
cause juvenile  owls  fledged  during  the  main 
breeding  attempt  at  this  site  have  typically  un- 
dergone a body  molt  by  this  time  and  their 
breasts  are  heavily  streaked  in  the  manner  of 
adult  birds.  No  other  nests  within  the  area  still 
contained  chicks  at  this  time.  The  five  chicks 
were  frequently  seen  at  the  burrow  entrance 
through  the  middle  of  September  when  field- 
work was  discontinued.  This  is.  typical  of 
young  owls  still  fully  dependent  on  their  par- 
ents for  food;  owls  fledged  earlier  in  the  season 
had  dispersed  from  their  natal  burrows  by  early 
August  as  indicated  by  radio  telemetry  (Ger- 
vais and  Rosenberg,  unpubl.  data).  The  owl’s 
mate  for  this  second  attempt  was  also  banded, 
but  his  bands  were  consistently  too  muddy  to 
read  and  he  was  never  identified. 


SHORT  COMMUNICATIONS 


571 


Our  fieldwork  did  not  include  detailed  ob- 
servations of  all  nests  in  our  study  area 
throughout  the  summer  and  early  fall,  but  we 
did  not  find  any  other  evidence  in  support  of 
double  brooding  attempts.  These  attempts 
may  be  quite  rare  and  only  occur  in  excep- 
tional years.  The  1998  breeding  season  was 
marked  by  very  late  rains,  resulting  in  a high 
proportion  of  renesting  efforts  by  the  owls 
(Rosenberg  and  Gervais,  unpubl.  data).  The 
prolonged  growing  season  that  followed  the 
wet  spring  may  have  led  to  conditions  con- 
ducive to  late-season  breeding  attempts,  such 
as  greater  food  availability. 

Nevertheless,  it  is  clear  that  at  least  in  some 
conditions  western  Burrowing  Owls  can  raise 
two  broods  of  chicks,  thus  increasing  their  re- 
productive output.  This  may  be  important  for 
individuals  whose  first  broods  were  small  be- 
cause of  predation  or  the  loss  of  a mate,  or 
for  populations  recovering  from  environmen- 
tal damage  such  as  pesticides  or  burrowing 
rodent  control.  This  information  is  also  im- 
portant for  use  in  sensitivity  modeling  such  as 
that  done  by  Emlen  and  Pikitch  (1989)  or 
Noon  and  Biles  (1990),  because  accurate  de- 
mographic parameter  estimation  is  essential  to 
determining  life  history  strategies  and  evalu- 
ating demographic  risks  to  populations. 

ACKNOWLEDGMENTS 

This  research  was  funded  by  the  U.S.  Navy  Engi- 
neering Field  Activity  West,  the  Bureau  of  Land  Man- 
agement, Bakersfield,  California  Field  Office,  the  U.S. 
Fish  and  Wildlife  Service  Partners  in  Wildlife  Act 
Fund,  California  Department  of  Fish  and  Game,  and 
the  National  Fish  and  Wildlife  Foundation.  We  thank 
T.  Lanman  and  J.  Podulka  for  assistance  with  field  ob- 
servations; E.S.  Botelho,  G.A.  Green,  and  an  anony- 
mous reviewer  for  comments  on  the  manuscript;  and 
J.  Crane  for  logistical  support  at  NAS  Lemoore  that 
made  this  research  possible.  Cooperators  ot  the  Oregon 
Cooperative  Fish  and  Wildlife  Research  Unit  include 
the  U.S.  Fish  and  Wildlife  Service,  Oregon  State  Uni- 
versity, Oregon  Department  of  Fish  and  Wildlife,  the 


Wildlife  Management  Institute,  and  the  Biological  Re- 
sources Division  of  the  U.S.  Geological  Survey.  This 
work  was  conducted  in  conjunction  with  The  Institute 
for  Bird  Populations,  and  is  The  Institute  for  Bird  Pop- 
ulations Publication  No.  175. 

LITERATURE  CITED 

Calow,  P,  M.  Sibley,  and  V.  Forbes.  1997.  Risk  as- 
sessment on  the  basis  of  simplified  life-history 
scenarios.  Environ.  Toxicol.  Chem.  16:1983- 
1989. 

Caswell,  H.  1996.  Demography  meets  ecotoxicology: 
untangling  the  population  level  effects  of  toxic 
substances.  Pp.  255-288  in  Ecotoxicology:  a hi- 
erarchical approach  (M.  C.  Newman  and  C.  H. 
Jagoe,  Eds.).  Lewis  Publishers,  Boca  Raton,  Flor- 
ida. 

DeSante,  D.  E,  E.  D.  Ruhlen,  S.  L.  Adamany,  K.  M. 
Burton,  and  S.  Amin.  1997.  A census  of  Burrow- 
ing Owls  in  central  California  in  1991.  Pp.  38-48 
in  The  Burrowing  Owl:  its  biology  and  manage- 
ment (J.  L.  Lincer  and  K.  Steenhof,  Eds.).  Raptor 
Research  Foundation,  Inc.,  Boise,  Idaho. 

Emlen,  J.  M.  and  E.  K.  Pikitch.  1989.  Animal  popu- 
lation dynamics:  identification  of  critical  compo- 
nents. Ecol.  Model.  44:253—273. 

Gervais,  J.  A.,  D.  K.  Rosenberg,  D.  M.  Fry,  L.  Tru- 
Lio,  AND  K.  K.  Sturm.  In  press.  Burrowing  Owls 
and  agricultural  pesticides:  evaluation  of  residues 
and  risks  for  three  populations  in  California.  En- 
viron. Toxicol.  Chem. 

Haug,  E.  a.,  B.  a.  Millsap,  and  M.  S.  Martell. 
1993.  Burrowing  Owl  {Speotyto  cunicularia).  In 
The  birds  of  North  America,  no.  61  (A.  Poole  and 
F Gill,  Eds.).  The  Academy  of  Natural  Sciences, 
Philadelphia,  Pennsylvania;  The  American  Orni- 
thologists’ Union,  Washington,  D.C. 

James,  P.  C.  and  R.  H.  M.  Espie.  1997.  Cunent  status 
of  the  Burrowing  Owl  in  North  America:  an  agen- 
cy survey.  Pp.  3—5  in  The  Burrowing  Owl:  its  bi- 
ology and  management  (J.  L.  Lincer  and  K.  Steen- 
hof, Eds.).  Raptor  Research  Foundation,  Inc.,  Boi- 
se, Idaho. 

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

Millsap,  B.  A.  and  C.  Bear.  1990.  Double  brooding 
by  Florida  Burrowing  Owls.  Wilson  Bull.  102: 
313-317. 

Noon,  B.  R.  and  C.  Biles.  1990.  Mathematical  demog- 
raphy of  Spotted  Owls  in  the  Pacific  Northwest.  J. 
Wildl.  Manage.  54:18-27. 


572 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


Wilson  Bull.,  111(4),  1999,  pp.  572-573 


Continuous  Nesting  of  Barn  Owls  in  Illinois 

Jeffery  W.  Walk,'-^  Terry  L.  Esker,-  and  Scott  A.  Simpson- 


ABSTRACT. — Bam  Owls  (Tyto  alba)  typically  be- 
gin nesting  in  temperate  zones  in  early  spring.  The 
species  has  high  reproductive  output  (large  clutch  size, 
occasionally  double-brooded)  and  high  mortality  for  a 
member  of  the  Strigiformes.  We  report  on  Barn  Owls 
in  one  nest  box  that  hatched  hve  clutches  and  fledged 
young  from  four  clutches  within  23  months.  Laying, 
incubation,  or  brood-rearing  was  attempted  in  every 
month  of  the  year.  In  the  only  clutch  mortality  we  wit- 
nessed, three  young  apparently  died  of  exposure  dur- 
ing a period  of  cold  weather  (temperatures  as  low  as 
— 15°C).  Received  24  February  1999,  accepted  27 
May  1999. 


Bam  Owls  {Tyto  alba)  have  a high  repro- 
ductive rate  and  relatively  short  life  span,  fit- 
ting an  r-selected  life  history  strategy  (Colvin 
et  al.  1984,  Marti  1997).  Average  clutch  sizes 
in  North  America  range  from  4.2— 7.1  eggs 
(mostly  5-6  eggs;  Hands  et  al.  1989,  Marti 
1992)  and  the  average  Bam  Owl  life  span  is 
less  than  2 years  (1.4  years,  Stewart  1952;  1.7 
years,  Keran  1981).  Most  nest  initiations  in 
temperate  areas  occur  from  February  to  June, 
with  the  peak  probably  occurring  March-May 
(Hands  et  al.  1989,  Marti  1992).  Nestlings 
have  been  banded  in  every  month  except  Feb- 
ruary in  the  northern  United  States  (Stewart 
1952),  suggesting  that  nesting  may  occur  in 
any  season.  Second  broods  are  uncommon  in 
temperate  climates.  For  example,  only  about 
5%  of  pairs  in  Utah  produced  a second  brood 
(Marti  1992).  In  contrast,  nesting  in  tropical 
areas  occurs  year  around  with  double  broods 
being  common  (e.g.,  Lenton  1984).  In  captiv- 
ity, Maestrelli  (1973)  reported  a pair  fledging 
six  broods  in  22  months. 

A nest  box  (enclosed  design  from  Colvin 
1983)  was  erected  4.6  m above  ground  in  an 
empty  barn  in  1986.  This  location  was  within 


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

^ Illinois  Dept,  of  Natural  Resources,  Prairie  Ridge 
State  Natural  Area,  Newton,  IL  62448. 

' Corresponding  author;  E-mail:  j-walk@uiuc.edu 


a 64  ha  grassland  tract  of  Prairie  Ridge  State 
Natural  Area,  Marion  County,  Illinois  (38°  45' 
N,  88°  51'  W).  Prairie  Ridge  State  Natural 
Area  grasslands  are  a combination  of  restored 
native  grasses  and  introduced  cool-season 
grasses  managed  by  the  Illinois  Department  of 
Natural  Resources  for  grassland  wildlife. 
Within  2 km  of  the  nest  site,  land  use  was 
about  70%  rowerop  agriculture,  15%  grass- 
lands (96  ha  of  Prairie  Ridge  State  Natural 
Area,  98  ha  of  Conservation  Reserve  Program 
grasslands),  8%  small  grains,  and  small 
amounts  of  woodland,  pasture  and  farmsteads. 

The  nest  box  was  checked  periodically  and 
Rock  Dove  {Columba  livia)  nests  were  re- 
moved. We  first  observed  one  adult  Barn  Owl 
perched  upon  the  nest  box  on  20  September 
1993.  On  1 November  1993,  two  adults  and  a 
clutch  of  six  eggs  were  noted.  We  saw  three 
chicks  on  21  December  1993  and  found  them 
dead  in  the  nest  box  on  6 January  1994.  The 
young  apparently  died  of  exposure  to  harsh 
weather  between  23  and  31  December.  Mean 
temperature  during  this  period  was  — 5.5°C 
(—15°  to  3°  C)  with  7 cm  snowfall  on  25  De- 
cember and  northeasterly  winds  20-25  km/hr 
blowing  into  the  nest  box  on  28  and  29  De- 
cember (weather  data  from  Midwestern  Cli- 
mate Center,  Champaign,  Illinois).  The  second 
nest  attempt  began  about  1 1 March  1994  (one 
egg  in  nest  box),  with  five  young  seen  through 
May  1994.  Fledging  occurred  in  early  July 
(four  grown  birds  observed  14  July).  An  adult 
and  the  third  clutch  of  five  eggs  were  noted 
23  August  1994.  Four  chicks  from  this  clutch 
fledged  between  25  and  30  October.  The 
fourth  nest  attempt  was  apparently  initiated  in 
early  to  mid-February  1995.  A clutch  of  seven 
eggs  was  observed  13  March,  and  five  yolmg 
about  to  fledge  were  seen  9 May  1995.  A fifth 
clutch  consisted  of  three  eggs  recorded  on  7 
July  1995,  three  young  (estimated  two  weeks 
old)  on  23  August  and  two  fledglings  on  19 
October  1995.  Single  Barn  Owls  were  ob- 
served only  sporadically  after  this  date. 


SHORT  COMMUNICATIONS 


573 


The  five  clutches  averaged  5.2  eggs  per 
clutch  (range  3-7).  Of  the  four  successful 
clutches,  3.75  young  fledged  per  clutch  (range 
2-5).  We  estimate  an  average  of  160  days 
(range  130-190  days)  for  each  successful  nest 
cycle  (estimated  time  between  the  start  of  one 
attempt  and  the  start  of  the  next).  While  these 
birds  were  not  banded  or  marked,  their  con- 
stant presence  at  the  nesting  site  from  Septem- 
ber 1993  to  October  of  1995  suggests  the 
same  adults  were  involved  in  all  nest  attempts. 

Excluding  the  first  nest  attempt,  the  tem- 
poral pattern  of  the  four  successful  nests  better 
fits  the  typical  early  spring/late  summer  pat- 
tern of  double-brooded  Barn  Owls  (Taylor 
1994),  although  the  1995  nests  were  about  one 
month  earlier  than  their  1994  counterparts. 
Taken  as  a whole,  this  nest  site  was  used  near- 
ly continuously  for  two  years  with  an  addi- 
tional delay  of  about  30  days  between  nest 
attempts  during  the  coldest  time  of  year. 

This  is  only  the  fourth  report  of  Bam  Owls 
using  a nest  box  in  Illinois  (Illinois  Biological 
and  Conservation  Data  System,  unpubl.  data). 
Once  this  nest  site  was  discovered,  breeding 
activity  was  brief,  but  quite  productive. 

ACKNOWLEDGMENTS 

We  thank  R.  Day,  R.  Edgin,  B.  Griffith  and  R.  Jan- 
sen for  their  assistance  with  monitoring  the  Barn  Owl 
nest  box.  We  gratefully  acknowledge  C.  Becker,  D. 
Cooper,  J.  Herkert,  E.  Kershner,  G.  Kruse,  C.  Marti, 
D.  Olson,  D.  Smith,  G.  Therres,  and  R.  Warner  for 
reviewing  this  manuscript.  We  would  also  like  to  thank 
the  Illinois  Endangered  Species  Protection  Board,  the 


Illinois  Natural  History  Survey,  the  Illinois  Nature  Pre- 
serves Commission  and  The  Nature  Conservancy  for 

their  support  of  Prairie  Ridge  State  Natural  Area. 

LITERATURE  CITED 

Colvin,  B.  A.  1983.  Nest  boxes  for  Barn  Owls.  Publ. 
346(183).  Ohio  Dept,  of  Nat.  Res.,  Columbus. 

Colvin,  B.  A.,  P.  L.  Hegdal,  and  W.  B.  Jackson. 
1984.  A comprehensive  approach  to  research  and 
management  of  Common  Barn-Owl  populations. 
Pp.  270-282  in  Proceedings  of  workshop  man- 
agement of  nongame  species  and  ecological  com- 
munities (W.  Comb,  Ed.).  Univ.  of  Kentucky,  Lex- 
ington. 

Hands,  H.  M.,  R.  D.  Drobney,  and  M.  R.  Ryan.  1989. 
Status  of  the  Common  Barn-Owl  in  the  northcen- 
tral  United  States.  Missouri  Cooperative  Fish  and 
Wildlife  Research  Unit,  Columbia. 

Reran,  D.  1981.  The  incidence  of  man-caused  and 
natural  mortalities  to  raptors.  Raptor  Res.  15:108— 
112. 

Lenton,  G.  M.  1984.  The  feeding  and  breeding  ecol- 
ogy of  Barn  Owls  Tyto  alba  in  peninsular  Malay- 
sia. Ibis  126:551-575. 

Maestrelli,  J.  R.  1973.  Propagation  of  Barn  Owls  in 
captivity.  Auk  90:426-428. 

Marti,  C.  D.  1992.  Barn  Owl.  In  The  birds  of  North 
America,  no.  1 (A.  Poole,  P.  Stettenheim,  and  E 
Gill,  Eds.).  The  Academy  of  National  Sciences, 
Philadelphia,  Pennsylvania;  The  American  Orni- 
thologists’ Union,  Washington,  D.C. 

Marti,  C.  D.  1997.  Lifetime  reproductive  success  in 
Barn  Owls  near  the  limit  of  the  species’  range. 
Auk  114:581-592. 

Stewart,  P.  A.  1952.  Dispersal,  breeding  behavior  and 
longevity  of  banded  Barn  Owls  in  North  America. 
Auk  69:227-245. 

Taylor,  I.  1994.  Barn  Owls:  predator-prey  relation- 
ships and  conservation.  Cambridge  Univ.  Press, 
Cambridge,  U.K. 


574 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


Wilson  Bull.,  111(4),  1999,  pp.  574-575 


Unusual  Nest  Sites  for  Southwestern  Willow  Flycatchers 

Scott  H.  Stoleson''^  and  Deborah  M.  Finch' 


ABSTRACT — The  endangered  southwestern  Wil- 
low Elycatcher  (Empidona.x  traillii  extimus)  is  an  ob- 
ligate riparian  species  that  typically  nests  in  willow 
(Salix  spp.)  thickets  or  other  dense,  shrubby  habitats. 
We  report  on  the  first  nests  in  Arizona  sycamore  (Pla- 
tanus  wrightii)  and  in  a climbing  rose  vine  (Rosa  mul- 
tiflora). Although  these  nests  were  located  in  novel 
substrates,  they  were  typical  for  the  species  in  being 
supported  by  multiple  small  stems  and  in  having  a 
dense  canopy  cover.  We  suggest  that  nest  substrate 
preferences  of  Willow  Elycatchers  in  the  Southwest 
may  be  broader  than  generally  considered.  Received 
17  Nov.  1998,  accepted  6 March  1999. 


The  southwestern  Willow  Flycatcher  (Empi- 
donax  trcdllii  extimus),  a federally-listed  endan- 
gered species,  is  an  obligate  riparian  specialist 
that  breeds  in  dense  vegetation  associated  with 
watercourses  (U.S.  Fish  and  Wildlife  Service 
1995).  Most  studies  of  habitat  preferences  in 
Willow  Flycatchers  have  shown  a strong  asso- 
ciation with  willow  (Salix  spp.)  thickets  or  other 
shrubby  habitats  (McCabe  1991,  Sedgwick  and 
Knopf  1992).  In  the  Southwest,  nests  of  this 
subspecies  have  been  found  most  commonly  in 
willows,  salt  cedar  (Tamarix  spp.),  and  locally 
in  forbs  such  as  stinging  nettles  (Urtica  dioica 
holosericea)  or  trees  such  as  boxelder  (Acer  ne- 
gundo),  alder  (Alnus  spp.),  Russian  olive 
(Eleagnus  angustifolia),  and  young  Fremont 
cottonwoods  (Populus  fremontii;  Sferra  et  al. 
1997;  Sogge  et  al.  1997).  In  this  paper  we  report 
on  the  first  recorded  incidence  of  Willow  Fly- 
catchers nesting  in  Arizona  sycamore  (Platanus 
wrightii)  and  in  a nonnative  climbing  rose  (Rosa 
multiflora). 

These  observations  were  made  as  part  of  a 
study  of  southwestern  Willow  Flycatchers  in 
the  Gila  River  valley  near  the  towns  of  Cliff 
and  Gila.  Grant  County,  New  Mexico  (32°  57' 
N,  108°  35'  W).  The  study  area  consists  of 


' USDA  Forest  Service,  Rocky  Mountain  Re.search 
Station,  2205  Columbia  SE,  Albuquerque,  NM  87106. 

^Corresponding  author;  E-mail: 
sstolcso/rmrs_albq@fs.fcd.us 


patches  of  riparian  woodland  along  the  river 
and  earthen  irrigation  ditches  at  elevations  av- 
eraging 1400  m.  Most  of  the  valley  bottom  is 
used  for  ranching  and  farming.  Woodland 
patches  are  composed  primarily  of  Fremont 
cottonwood,  Goodding’s  willow  (Salix  good- 
dingii),  boxelder,  Arizona  sycamore  and  Ari- 
zona walnut  (Juglans  major),  with  an  under- 
story of  shrubs,  forbs,  and  grasses.  This  valley 
supports  the  largest  known  population  of  Wil- 
low Flycatchers  in  the  Southwest,  with  an  es- 
timated 230  pairs  in  1998  (Stoleson  and 
Finch,  unpubl.  data;  P.  Boucher,  pers.  comm.). 

Nesting  habits  at  this  site  differed  from 
what  has  been  reported  elsewhere  in  the 
Southwest.  Of  257  nests  located  in  1997- 
1998,  76.5%  were  placed  in  boxelder,  8.6%  in 
willows,  6.3%  in  Russian  olive,  and  the  re- 
mainder (<5%  each)  in  Arizona  alder  (Alnus 
oblongifolia),  seepwillow  (Baccharis  glutino- 
sa),  Fremont  cottonwood,  salt  cedar,  Arizona 
sycamore,  and  rose. 

One  pair  of  Willow  Flycatchers  was  found 
building  a nest  in  a sycamore  on  8 June  at  the 
Fort  West  Ditch  site  on  the  Gila  National  For- 
est (FS-4).  This  nest  was  too  high  for  its  con- 
tents to  be  visible  (nest  characteristics  in  Table 
1).  Parents  were  observed  carrying  food  to  the 
nest  on  1 and  6 July.  On  13  July,  the  parents 
were  observed  feeding  at  least  two  fledglings 
in  the  surrounding  trees.  This  nest  was  in  a 
cluster  of  five  vertical  twigs  on  a small 
branch.  Although  the  nest  tree  was  very  open, 
the  nest  itself  was  immediately  beneath  a 
dense  layer  of  foliage. 

A second  sycamore  nest  (GRP-7)  was  lo- 
cated in  The  Nature  Conservancy  of  New 
Mexico’s  Gila  Riparian  Preserve  on  23  July. 
The  female  was  on  the  nest  incubating  or 
brooding.  On  31  July  the  nest  was  found  errip- 
ty  and  disheveled,  presumably  the  result  of 
predation.  This  nest  was  located  in  a cluster 
of  about  1 2 vertical  twigs  at  the  end  of  a short, 
broken  branch.  Like  the  previous  nest,  the  nest 
was  visible  from  the  sides  but  covered  from 
above  by  dense  foliage. 


SHORT  COMMUNICATIONS 


575 


TABLE  1.  Characteristics  and  outcomes  of  southwestern  Willow  Flycatcher  nests  in  rose  and  sycamore  in 
the  Gila  River  valley,  southwestern  New  Mexico,  1998. 

Nesl  characierisiics 

Nesl  plant 


Ne.st 

Sub.striue 

Nest  ht. 
(m) 

Canopy  ht. 
at  nest  (m) 

Canopy 
cover  (%)^ 

diameter 

(cm)*' 

Distance  to 
water  (cm)*^ 

Distance  to 
edge  (m)‘* 

Outcome 

SEl-19 

rose 

3.5 

13.9 

94 

0.7 

8 

33 

fledged  2 

FS-4 

sycamore 

13.6 

17.9 

94 

33.0 

31 

10 

fledged  ^2 

GRP-7 

sycamore 

8.0 

12.2 

93 

33.2 

25 

12 

depredated 

^ Average  percent  canopy  cover  measured  at  base  of  nest  plant  and  at  points  4 and  8 m from  base  in  tour  cardinal  directions,  measured  using  densiomeiers. 
^Measured  at  1.7  m above  ground. 

^ Horizontal  distance  between  base  of  nest  plant  and  nearest  perennial  water. 

^ Defined  here  as  horizontal  distance  between  nest  site  and  nearest  area  with  no  tree  cover  (i.e.,  shrubs,  sand,  or  pasture). 


On  18  June,  an  incubating  bird  was  flushed 
from  a nest  in  a rose  vine  climbing  a large 
boxelder  tree  on  the  U-Bar  Ranch  (SEl-19). 
The  nest  contained  two  eggs  at  that  time.  On 
13  July,  two  almost  fully  feathered  fledglings 
were  observed  being  fed  in  the  undergrowth 
near  the  nest.  The  nest  was  placed  at  the  junc- 
tion of  four  stems  of  the  nonnative  Rosa  mul- 
tiflora, hanging  from  and  about  a meter  below 
a leaning  trunk  of  boxelder. 

Willow  Flycatcher  nests  have  been  found 
only  rarely  in  native  shrubby  Rosa  species  in 
the  Southwest,  in  California  (W.  Haas,  pers. 
comm.)  and  at  high  elevations  in  Arizona 
(McCarthey  et  al.  1998).  In  the  Palouse  Hills 
of  Washington,  where  Willow  Flycatchers  are 
not  restricted  to  riparian  habitats,  rose  was  the 
most  frequent  nest  substrate  (King  1955). 
Similarly,  56%  of  nests  in  the  interior  of  Brit- 
ish Columbia  were  in  rose  (Campbell  et  al. 
1997).  Nests  have  been  reported  in  rose  else- 
where as  well  (Walkinshaw  1966;  McCabe 
1991;  J.  Sedgwick,  pers.  comm.). 

Our  observations  emphasize  that  Willow  Fly- 
catchers are  opportunistic  in  their  choice  of  nest- 
ing substrates,  apparently  requiring  only  dense 
fohage  and  a suitable  twig  structure  to  support 
their  nests  (McCabe  1991,  Sogge  et  al.  1997). 
Although  the  three  nests  reported  here  were  un- 
usual in  terms  of  substrate  species,  they  were 
very  typical  of  flycatcher  nests  with  respect  to 
fohage  density  and  twig  structure  (Table  1). 

ACKNOWLEDGMENTS 

We  thank  G.  Bodner,  K.  Brodhead,  P.  Chan.  J.  Gar- 
cia, B.  Gibbons,  D.  Hawk.sworth,  and  H.  Walker  for 
field  assistance;  P.  Boucher,  J.  Monzingo,  and  R.  Pope 
of  the  Gila  National  Forest  and  T.  Bays,  T.  Shelley, 
and  C.  Rose  of  Phelps  Dodge  for  logistical  support;  D. 


Parker  for  sharing  his  expertise;  and  T and  D.  Ogilvie 
for  their  hospitality.  Funding  was  provided  by  the  Gila 
National  Forest,  Phelps  Dodge  Corporation,  and  The 
Nature  Conservancy.  Comments  by  T.  Bays,  D.  Mei- 
dinger,  C.  Rose,  J.  Sedgwick.  M.  Whitfield  and  two 
anonymous  reviewers  improved  the  manuscript. 

LITERATURE  CITED 

Campbell,  R.  W,  N.  K.  Da  we,  I.  McTaggart-Cowan, 
G.  E.  Smith,  and  J.  M.  Cooper.  1997.  The  birds  of 
Briti.sh  Columbia;  passerines:  flycatchers  through 
vireos.  Univ.  of  British  Columbia  Press,  Vancouver. 
King,  J.  R.  1955.  Notes  on  the  life  history  of  Traill’s 
Flycatcher  {Empidonax  traillii)  in  southeastern 
Washington.  Auk  72:148-173. 

McCabe,  R.  A.  1991.  The  little  green  bird.  Palmer 
Publications,  Inc.,  Amherst,  Wisconsin. 
McCarthey,  T.  D.,  C.  E.  Paradzick,  J.  W.  Rourke, 
M.  W.  Sumner,  and  R.  F.  Davidson.  1998.  Ari- 
zona Partners  in  Flight  southwestern  Willow  Fly- 
catcher 1997  survey  and  nest  monitoring  report. 
Nongame  and  Endangered  Wildlife  Program 
Technical  Report  130.  Arizona  Game  and  Fish 
Department,  Phoenix. 

Sedgwick,  J.  A.  and  F.  L.  Knopf.  1992.  Describing 
Willow  Flycatcher  habitats:  scale  perspectives  and 
gender  differences.  Condor  94:720-733. 

Sferra,  S.  j.,  T.  E.  Corman,  C.  E.  Paradzick,  J.  W. 
Rourke,  J.  A.  Spencer,  and  M.  W.  Sumner.  1997. 
Arizona  Partners  in  Flight  southwestern  Willow 
Flycatcher  survey;  1993-1996  summary  report. 
Nongame  and  Endangered  Wildlife  Program 
Technical  Report  113.  Arizona  Game  and  Fish 
Department,  Phoenix. 

Sogge,  M.  K.,  R.  M.  Marshall,  S.  J.  Sferra,  and  T. 
J.  Tibbitts.  1997.  A southwestern  Willow  Fly- 
catcher natural  history  summary  and  survey  pro- 
tocol. National  Park  Service  Technical  Report 
NPS/NAUCPRS/NRTR-97/12.  Flagstaff,  Arizona. 
U.S.  Fish  and  Wildlife  Service.  1995.  Final  rule  deter- 
mining endangered  status  for  the  southwestern  Wil- 
low Flycatcher.  Federal  Register  60:10694-10715. 
Walkinshaw,  L.  H.  1966.  Summer  biology  of  Traill’s 
Flycatcher.  Wilson  Bull.  78:31—46. 


576 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


Wilson  Bull..  1 1 1(4),  1999,  pp.  576-577 


Mortality  of  an  Adult  Veery  Incurred  During  the  Defense  of  Nestlings 

David  I.  King' 


ABSTRACT — Cost-benefit  analyses  of  the  adaptive 
significance  of  nest  defense  in  birds  are  based  on  the 
assumption  that  parent  birds  attacking  nest  predators 
risk  serious  injury  or  death.  However,  there  are  few 
published  records  of  adult  birds  dying  during  nest  de- 
fense. I found  an  adult  male  Veery  (Catharus  fiisce- 
sens)  dead  in  circumstances  indicating  that  the  bird 
died  while  defending  his  nest.  This  observation  sup- 
ports speculation  that  adult  birds  risk  injury  or  death 
in  the  course  of  nest  defense,  and  lends  support  to 
explanations  of  variability  in  nest-defense  behavior 
that  are  presented  in  terms  of  cost-benefit  analysis  and 
the  optimization  of  fitness.  Received  11  Jan.  1999,  ac- 
cepted 27  May  1999. 


On  27  June  1997  I located  a Veery  {Catha- 
rus fusee  sens)  nest  containing  3 eggs  1.6  m 
from  the  ground  in  a 9 year-old  clearcut  in  the 
White  Mountain  National  Forest  (43°  58'  N- 
70°  97'  W)  in  north-central  New  Hampshire. 
On  1 1 July  I approached  the  nest  and  ob- 
served an  adult  on  the  rim  of  the  nest  with  its 
tail  tilted  up  in  an  unusual  posture.  The  adult 
was  dead,  the  body  cold.  Autopsy  revealed 
that  the  adult  was  a male  and  that  the  body 
had  numerous  puncture  wounds,  one  on  the 
central-posterior  region  of  the  right  pectoralis 
muscle  and  four  more  on  the  back  in  the  vi- 
cinity of  the  synsacrum.  The  four  puncture 
wounds  on  the  back  were  symmetrically  ori- 
ented on  either  side  of  the  spine,  1. 5-2.0  cm 
apart  and  up  to  7 mm  deep,  and  probably 
caused  the  death  of  the  bird.  The  depth  and 
spacing  of  the  wounds  suggest  that  the  pred- 
ator may  have  been  a Sharp-shinned  Hawk 
(Accipiter  striatus),  a species  known  to  attack 
Veery  nests  (Day  1953). 

The  adult  was  missing  all  of  its  secondary 
feathers  and  all  but  three  rectrices.  The  miss- 
ing feathers  were  scattered  over  the  nest  and 
the  immediate  vicinity  of  the  nest  bush.  An- 
other adult,  presumably  the  female,  was  heard 


‘ Dept,  of  Forestry  and  Wildlife  Management,  Univ. 
of  Massachusetts  Amherst,  Amherst,  MA  01003; 
E-mail:  daveking@forwild.umass.edu 


alarm  calling  nearby.  The  nest  cup  was  tom 
down  and  had  come  to  rest  on  some  branches 
of  the  nest  bush  below  the  original  level  of 
the  nest.  Two  nestlings  were  found  prone  on 
ground  beneath  the  nest.  The  nestlings  were 
cold,  and  gaped  weakly.  Based  on  previous 
experience  with  the  nesting  phenology  of  this 
species,  I estimated  the  nestlings  to  be  at  least 
three  days  short  of  fledging.  The  original 
clutch  size  was  three  eggs.  It  is  unknown 
whether  the  third  egg  hatched.  No  sign  of  a 
third  nestling  or  eggshells  was  found  in  the 
vicinity. 

On  the  basis  of  several  lines  of  evidence,  I 
conclude  the  male  Veery  was  killed  in  the 
course  of  nest  defense.  First,  based  on  the 
spacing  of  the  puncture  marks,  the  predator 
was  evidently  large  enough  to  consume  a Vee- 
ry (K.  Doyle,  pers.  comm.).  Under  these  cir- 
cumstances, it  is  difficult  to  conceive  of  a sce- 
nario in  which  a predator  would  kill  the  adult, 
leave  it  on  the  nest,  attempt  to  depredate  the 
nest,  and  subsequently  leave.  In  contrast,  the 
disposition  of  the  male,  and  of  the  nest  and  its 
contents  are  all  consistent  with  the  hypothesis 
that  the  adult  was  killed  in  association  with 
nest  defense.  The  adult  was  killed  yet  not  con- 
sumed, indicating  that  it  was  not  the  original 
target  of  the  predator.  Furthermore,  the  nes- 
tlings were  not  consumed,  suggesting  that  the 
predator  was  interrupted  during  the  predation 
event.  Veerys  have  been  observed  successful- 
ly defending  nests  by  striking  the  predator 
with  their  wings  (Nice  1962;  Pettingill  1976, 
pers.  obs.),  which  would  account  for  the  sym- 
metrical loss  of  wing  feathers  I observed.  It 
could  be  argued  that  the  male  was  the  original 
target  of  the  predator,  was  wounded  else- 
where, and  returned  to  the  nest  seeking  a se- 
cure hiding  place,  were  it  not  for  the  fact  that 
the  nest  had  been  attacked. 

Cost-benefit  analyses  of  the  adaptive  sig- 
nificance of  nest  defense  in  birds  are  based  on 
the  assumption  that  birds  attacking  predators 
during  the  course  of  nest  defense  are  at  some 


SHORT  COMMUNICATIONS 


577 


risk  of  injury  or  death  (Montgomerie  and 
Weatherhead  1988).  However,  observations  of 
nest  predation  events  under  natural  conditions 
are  rare  (Pettingill  1976)  and  observations  of 
attacks  on  parent  birds  by  predators  during  the 
course  of  nest  predation  are  even  more  scarce. 
Brunton  (1986)  observed  a Killdeer  {Char- 
adrius  vociferous)  killed  by  a red  fox  (Vulpes 
vulpes)  while  performing  a distraction  display. 
This  observation  of  a Veery  confronting  a 
predator  at  the  cost  of  its  own  life  during  ac- 
tive defense  of  the  nest  is  to  my  knowledge, 
unprecedented.  This  observation  supports 
speculation  that  adult  birds  assume  risk  of  in- 
jury or  death  in  the  course  of  nest  defense 
(Curio  and  Regelman  1985),  and  lends  sup- 
port to  explanations  of  variability  in  nest-de- 
fense  behavior  that  are  couched  in  terms  of 
cost-benefit  analysis  and  the  optimization  of 
fitness  (Montgomerie  and  Weatherhead  1988). 


ACKNOWLEDGMENTS 

I thank  K.  Doyle  of  the  Vertebrate  Museum  at  the 
University  of  Massachusetts  Amherst  for  assistance  in 
conducting  the  autopsy,  and  D.  Albano,  R.  DeGraaf, 
and  C.  Griffin  for  commenting  on  the  manuscript. 

LITERATURE  CITED 

Brunton,  D.  H.  1986.  Fatal  antipredator  behavior  of 
a Killdeer.  Wilson  Bull.  98:605-607. 

Curio,  E.  and  K.  Regelman.  1985.  The  behavior  and 
dynamics  of  Great  Tits  (Purus  major)  approaching 
a predator.  Z.  Tierpsychol.  69:3-18. 

Day,  K.  C.  1953.  The  home  life  of  the  Veery.  Bird- 
Banding  24:100-106. 

Montgomerie,  R.  D.  and  R J.  Weatherhead.  1988. 
Risks  and  rewards  of  nest  defense  by  parent  birds. 
Q.  Rev.  Biol.  63:167-187. 

Nice,  M.  M.  1962.  Observations  on  breeding  behavior 
of  Veeries  in  Michigan.  Bird-Banding  33:114. 
Pettingill,  O.  S.,  Jr.  1976.  Observed  acts  of  predation 
on  birds  in  northern  lower  Michigan.  Living  Bird 
15:33-41. 


Wilson  Bull.,  111(4),  1999,  pp.  577—581 

Relationships  of  Clutch  Size  and  Hatching  Success  to  Age  of  Female 

Prothonotary  Warblers 


Charles  R.  Blem,'-  Leann  B.  Blem,'  and  Claudia  I.  Barrientos' 


ABSTRACT. — We  obtained  1033  clutch  sizes  from 
281  known-age  female  Prothonotary  Warblers  (Pro- 
tonotaria  citrea)  nesting  in  nest  boxes  at  Pre.squile  Na- 
tional Wildlife  Refuge  in  eastern  Virginia  from  1987 
through  1998.  Prothonotary  Warblers  typically  nested 
twice  during  each  breeding  season;  first  clutches  of  all 
birds  averaged  1.01  eggs  greater  than  second  clutches 
[4.96  ± 0.72  (SD)  vs  3.94  ± 0.55].  Clutch  size  was 
significantly  smaller  in  first  nests  of  one-year-old  war- 
blers (4.64  ± 0.48)  than  in  first  clutches  of  females 
two  to  eight  years  old  (5.05  ± 0.62).  First  clutches  did 
not  differ  among  age  classes  of  birds  older  than  one 
year.  The  mean  size  of  second  clutches  was  not  sig- 
nificantly different  among  any  of  the  age  classes.  One 
year  old  birds  initiated  laying  significantly  later  than 
older  birds  (125.0  ± 6.4  vs  121.5  ± 7.7;  Julian  dates). 
The  average  number  of  infertile  eggs  in  first  clutches 
was  larger  in  one  year  old  females  and  differed  sig- 
nificantly from  that  of  older  females  (1.01  ± 0.90  vs 


' Dept,  of  Biology,  Virginia  Commonwealth  Univ., 
816  Park  Ave.,  Richmond,  VA  23284-2012. 

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


0.63  ± 0.87).  The  number  of  infertile  eggs  in  second 
clutches  did  not  differ  significantly  with  female  age. 
Significantly  fewer  eggs  hatched  in  first  nests  of  one 
year  old  birds  than  in  those  of  older  birds  (3.75  ± 0.89 
vs  4.33  ± 1.09).  Received  2 Dec.  1998,  accepted  2 
May  1999. 


The  Prothonotary  Warbler  {Protonotaria  ci- 
trea) is  unusual  among  wood  warblers  (Pa- 
rulidae)  because  it  nests  in  secondary  cavities. 
It  shares  this  trait  with  only  one  other  member 
of  the  116  members  of  the  subfamily  Paruli- 
nae  [Lucy’s  Warbler  {Vermivora  luciae):  Cur- 
son  et  al.  1994],  It  is  likewise  noteworthy 
among  the  birds  of  the  eastern  United  States 
in  that  it  migrates  farther  than  the  other  small 
passerines  nesting  in  secondary  cavities.  De- 
terminants of  clutch  size  of  Prothonotary  War- 
blers therefore  may  be  of  interest  for  compar- 
ison with  other  cavity-nesting  passerines  and 
with  other  Neotropical  migrants.  Several  re- 


578 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


searchers  have  documented  changes  in  repro- 
ductive success  associated  with  age  in  cavity- 
nesting, passerine  birds  (Bryant  1988,  Mc- 
Cleery  and  Perrins  1989,  Sternberg  1989, 
Seether  1990),  but  to  date  there  have  been  few 
such  data  for  a warbler. 

Prothonotary  Warblers  are  declining  in  the 
United  States  (Sauer  et  al.  1997),  thus  any 
knowledge  of  their  demography  may  be  useful 
in  conservation  of  the  species.  In  this  paper 
we  provide  an  analysis  of  a large  set  of  mea- 
surements of  reproductive  performance  of  the 
Prothonotary  Warbler  that  we  have  accumu- 
lated over  the  past  12  years.  Specifically,  we 
examine  clutch  size  and  infertility  in  this  spe- 
cies and  ask  the  question:  Does  clutch  size 
and/or  infertility  of  eggs  change  with  age  of 
females? 

STUDY  AREA  AND  METHODS 

Our  data  were  obtained  from  1987-1998  in  swamp 
forest  along  the  James  River  near  and  on  Presquile 
National  Wildlife  Refuge,  Hopewell,  Virginia  (37°  20' 
N,  77°  15'  W).  The  habitat  of  the  study  area  is  tidal 
swamp  in  which  the  dominant  tree  species  are  black 
gum  (Vv.v.vfl  .'iylvatica),  red  maple  (Acer  ruhrum),  and 
ash  {Fraxinu.s  sp.).  These  swamps  have  a relatively 
harsh  environment  where  tree-surface  temperatures 
regularly  exceed  45°  C and  tidal  amplitude  in  the 
swamp  during  spring  tides  often  exceeds  1 m.  Begin- 
ning in  March  1987,  we  placed  nest  boxes  made  of 
salt-treated  pine  or  red  cedar  at  100  m intervals  along 
the  creek  banks.  Box  dimensions  were  28  L X 9 W X 
6 D cm  and  the  entrance  hole  was  3.8  cm  in  diameter 
(see  Blem  and  Blem  1991,  1992,  1994,  for  details). 
We  gradually  increased  the  number  of  nest  boxes  in 
the  study  from  141  in  1987  to  300  in  1993—1998. 

We  checked  the  contents  of  boxes  8-15  times  during 
each  breeding  season.  Old  nest  material  was  removed 
from  the  boxes  in  late  winter.  Eggs  that  failed  to  hatch 
were  opened  to  determine  fertility  and  degree  of  de- 
velopment. The  present  paper  includes  only  those 
clutches  that  were  incubated  by  females  and  only  those 
eggs  that  failed  to  hatch  because  of  infertility.  Clutches 
that  failed  because  nests  were  abandoned  were  not  in- 
cluded in  the  analyses.  Parasitism  by  Brown-headed 
Cowbirds  (Molotlirus  ater)  is  relatively  uncommon  at 
our  .study  site  (<5%  of  all  clutches),  but  nests  con- 
taining cowbird  eggs  were  excluded  from  our  analyses. 
We  recorded  dates  of  hrst  eggs  and  clutch  sizes  only 
for  those  nests  visited  often  enough  that  we  could  be 
certain  of  laying  dates.  In  some  instances,  we  deter- 
mined clutch  size  but  ne.sts  were  subsequently  taken 
by  predators  and  we  were  unable  to  determine  fertility 
of  the  eggs.  Sample  sizes  therefore  vary  among  various 
subsets  of  the  data.  Because  Prothonotary  Warblers 
typically  produce  two  clutches  ctich  sea.son  (Petit  1989, 
Blem  and  Blem  1992),  we  divided  nests  with  eggs  into 


two  groups:  first  clutches  in  which  first  eggs  were  laid 
from  25  April  through  20  May  and  second  clutches  in 
which  first  eggs  were  laid  after  20  May.  Recaptures  of 
banded  birds  indicated  that  this  division  was  accurate 
for  this  data  set.  Nest  boxes  were  originally  attached 
to  trees.  We  moved  them  to  metal  poles  in  1995,  al- 
most completely  eliminating  predation  on  nests.  Since 
then  many  females  have  been  recaptured  during  sec- 
ond broods  in  the  same  nest  box.  We  captured  adults 
by  hand-netting  them  as  they  emerged  from  boxes  and 
banded  all  birds  with  aluminum  USFWS  bands.  In 
1998,  we  used  the  criteria  in  Pyle  (1997)  to  age  adults, 
but  older  adults  and  many  birds  captured  before  1998 
could  only  be  aged  relative  to  previous  captures.  We 
designated  such  birds  with  a + (i.e.,  3 + ) to  indicate 
minimal  age,  and  analyzed  age  classes  accordingly. 

All  data  are  reported  as  means  ± SD.  Differences 
among  groups  were  analyzed  using  nonparametric 
Kruskal-Wallis  tests  (x'  approximation;  Zar  1984,  SAS 
Institute  Inc.  1990,  Proc  NPARIWAY).  In  all  statisti- 
cal tests,  a probability  of  0.05  or  less  was  accepted  as 
significant  (P  < 0.05).  All  analyses  were  performed 
using  SAS  (Ver.  6;  SAS  Institute,  Inc.  1990)  on  an 
IBM  mainframe  computer  (VM  operating  system). 

RESULTS 

Sample  size. — Over  the  12  years  we  banded 
2968  nestlings  and  482  adult  females.  Birds 
were  recaptured  opportunistically,  therefore 
sample  sizes  varied  from  year  to  year  and  in- 
dividual age  classes  came  from  various  years. 
We  recovered  487  adults  and  103  birds  banded 
as  nestlings.  Recaptures  during  the  same  clutch 
were  counted  only  once.  Of  all  females  banded 
as  adults,  47.9%  were  recaptured  at  least  once 
in  subsequent  years  {n  = 231).  Only  1.7%  of 
all  nestlings  were  recaptured  (n  = 50).  Some 
females  {n  = 112)  were  captured  over  several 
years  and,  therefore,  are  represented  in  several 
age  classes  in  Tables  1 and  2. 

Clutch  size.- — We  obtained  1033  clutch  siz- 
es from  281  female  Prothonotary  Warblers  of 
known  age  (Table  1).  First  clutch  sizes  dif- 
fered significantly  among  age  classes  (x^  = 
22.4,  P = 0.002,  df  = 7),  but  there  was  no 
difference  in  second  clutches  (x^  = 5.0,  P > 
0.05,  df  = 7).  One  year  old  female  Protho- 
notary Warblers  laid  an  average  of  0.4  fewer 
eggs  in  first  clutches  (4.64  ± 0.48;  n = 42) 
than  did  older  females  (5.03  ± 0.73;  x^'~ 
17.7,  P < 0.001,  df  - 1;  Table  1).  First  clutch- 
es did  not  differ  among  age  classes  of  birds 
older  than  one  year  (x^  = 4.6,  P > 0.05,  df  = 
6).  First  clutches  of  all  birds  averaged  1.01 
eggs  more  than  second  clutches  (4.96  ± 0.72 
vs  3.94  ± 0.55;  x^  = 356.8,  P < 0.001,  df  = 


SHORT  COMMUNICATIONS 


579 


TABLE  1.  Clutch 
Refuge,  Virginia. 

size  of  known-age 

female  Prothonotary  Warblers  captured 

at  Presquile  National  Wildlife 

Age  (years)'* 

First  clutchesh 

Second  clutches^ 

Julian  date  of  first  egg*’ 

1 

4.64  ± 0.48  (42) 

3.81  ± 0.60  (11) 

125.0  ± 6.4  (40) 

1 + 

4.91  ± 0.72  (395) 

3.86  ± 0.62  (128) 

124.1  ± 5.5  (381) 

2 

4.94  ± 0.79  (33) 

3.90  ± 0.32  (10) 

123.2  ± 5.6  (32) 

2+ 

5.04  ± 0.79  (211) 

4.06  ± 0.46  (65) 

121.4  ± 5.8  (205) 

3 

5.20  ± 0.42  (10) 

4.25  ± 0.50  (4) 

120.3  ± 2.5  (10) 

3 + 

4.97  ± 0.59  (58) 

4.12  ± 0.33  (17) 

121.3  ± 6.2  (55) 

4 

4.50  ± 0.71  (2) 

— (0) 

121.0  ± 4.2  (2) 

5-8 

5.18  ± 0.53  (33) 

4.00  ± 0.39  (14) 

121.0  ± 6.1  (32) 

Totals 

4.96  ± 0.72  (784) 

3.94  ± 0.55  (249) 

123.0  ± 5.8  (757) 

^ Plus  signs  indicate  that  females  were  that  age  or  older. 

^ Numbers  in  parentheses  are  sample  sizes.  Values  are  means  ± SD. 


1;  Table  1).  The  mean  of  second  clutches  of 
one  year  old  birds  was  0.24  eggs  fewer  than 
that  of  females  2-8  years  old  (3.81  ± 0.60  vs 
4.05  ± 0.54),  but  did  not  differ  significantly 
among  the  age  classes  (x^  = 9.1,  P > 0.05, 
df  = 5;  Table  1).  We  found  no  change  in 
clutch  size  over  consecutive  years  for  82  of 
122  individuals;  in  24  cases  clutch  size  in- 
creased by  one  egg,  in  14  cases  clutch  size 
decreased  by  one  egg,  and  in  two  cases  clutch 
size  decreased  by  two  eggs. 

Nest  initiation  dates. — Nest  initiation  dates 
(Julian)  differed  significantly  among  age  clas- 
ses in  the  whole  data  set  (x^  = 54.7,  P < 
0.001,  df  = 7),  but  not  among  age  classes  of 
females  2 years  old  or  older  (x^  = 6.2,  P > 
0.05,  df  = 5).  One  year  old  birds  initiated  lay- 
ing of  first  clutches  significantly  later  than 
older  birds  (125.0  ± 6.4  vs  121.5  ± 5.8;  (x^ 
= 13.9,  P < 0.001,  df  = 1).  The  mean  date 
of  nest  initiation  was  remarkably  stable 


among  females  greater  than  two  years  old, 
varying  very  little  from  1 May  (Julian  date  = 
121). 

Infertility  rate. — The  number  of  infertile 
eggs  in  first  clutches  was  significantly  larger 
in  one  year  old  females  than  in  older  birds  (x^ 
= 3.9,  P < 0.05,  df  = 1;  Table  2),  but  the 
number  of  infertile  eggs  in  second  clutches 
did  not  differ  significantly  with  female  age  (x“ 
= 8.7,  P > 0.05,  df  = 7).  Infertile  eggs  were 
more  frequent  in  first  clutches  than  in  second 
(X^  = 8.8,  P < 0.01,  df  = 1).  We  found  no 
effect  of  age  on  frequency  of  clutches  in 
which  all  eggs  hatched,  regardless  of  clutch 
size  (x'  = 0.04,  P > 0.05,  df  = 1).  One  year 
old  birds  hatched  all  eggs  in  36.0%  (18/50)  of 
their  clutches.  Older  birds  hatched  all  eggs  in 
35.6%  (235/660)  of  their  clutches  and  there 
was  no  significant  difference  between  the  two 
groups  (x^  = 0.005,  P > 0.05,  df  = 1). 


TABLE  2.  Number  of  infertile  eggs  and  nestlings  per  clutch  of  known-age  Prothonotary  Warblers  captured 
1987-1998  at  Presquile  National  Wildlife  Refuge,  Virginia. 

Number  of  infertile  eggs 

Number  of  nestlings 

Age  (years)” 

First  clutches*’ 

Second  clutches*’ 

First 

clutches*’ 

Second  clutches*’ 

1 

1.00  ± 0.90  (28) 

0.43  ± 0.53  (7) 

3.70  ± 

0.89  (28) 

3.43  ± 0.79  (7) 

1 + 

0.55  ± 0.75  (295) 

0.42  ± 0.61  (67) 

4.41  ± 

0.92  (295) 

3.58  ± 0.69  (67) 

2 

0.92  ±1.41  (26) 

0.50  ± 0.68  (10) 

4.04  ± 

1.64  (26) 

3.40  ± 0.71  (10) 

2-h 

0.66  ± 0.88  (179) 

0.47  ± 0.79  (49) 

4.36  ± 

0.94  (179) 

3.61  ± 0.86  (49) 

3 

0.90  ± 1.20  (10) 

0.50  ± 0.71  (2) 

4.30  ± 

1.25  (10) 

3.50  ± 0.71  (2) 

3 + 

0.71  ± 1.01  (45) 

0.21  ± 0.43  (14) 

4.31  ± 

1.18  (45) 

3.86  ± 0.36  (14) 

4 

0.50  ± 0.71  (2) 

— (0) 

4.00  ± 

0.00  (2) 

— (0) 

5-8 

0.73  ± 0.87  (26) 

0.33  ± 0.71  (9) 

4.46  ± 

1.17  (26) 

3.55  ±1.01  (9) 

Totals 

0.64  ± 0.87  (611) 

0.42  ± 0.69  (158) 

4.34  ± 

1.01  (611) 

3.59  ± 0.77  (158) 

“ Plus  signs  indicate  that  females  were  that  age  or  older, 
h Numbers  in  parentheses  are  sample  sizes.  Values  are  means  ± SD. 


580 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


DISCUSSION 

In  eastern  Virginia,  Prothonotary  Warblers 
lay  2-8  eggs  per  clutch  and  clutch  size  varies 
significantly  among  years  (Blem  and  Blem 
1992).  In  1986,  we  initiated  a study  to  identify 
those  factors  that  were  responsible  for  those 
yearly  differences.  To  lend  explanatory 
strength  to  our  analyses,  we  attempted  to  iden- 
tify potential  sources  of  variation.  For  exam- 
ple, date  of  clutch  initiation  is  significantly  re- 
lated to  clutch  size  (Blem  and  Blem  1992), 
while  nest  cavity  volume,  position  of  the  nest, 
presence  of  old  nest  materials,  and  female 
mass  are  not  (Blem  and  Blem  1991;  Blem  et 
al.  1999,  unpubl.  data).  Female  age  is  related 
significantly  to  Prothonotary  Warbler  clutch 
size  in  a manner  similar  to  that  found  in  sev- 
eral other  cavity-nesting  species  (Klomp 
1970,  Saether  1990),  particularly  the  European 
Starling  {Sturnus  vulgaris;  Kluijver  1935), 
House  Martin  {Delichon  urbica;  Bryant 
1988),  Pied  Flycatcher  (Ficedula  hypoleuca; 
Sternberg  1989)  and  Great  Tit  (Parus  major; 
McCleery  and  Perrins  1989).  In  all  of  these 
species,  older  females  laid  0.4— 1.0  more  eggs 
per  clutch  than  birds  producing  their  first 
clutch.  The  increase  in  mean  size  of  first 
clutches  with  age  may  be  due  to  several  fac- 
tors including  proximate  factors  such  as  in- 
creased development  of  reproductive  tracts  or 
enhanced  ability  to  collect  and  process  energy, 
thus  allowing  females  to  produce  more  eggs. 

The  effects  of  senility  on  clutch  size  are  less 
well  known.  There  seems  to  be  no  document- 
ed decreases  of  clutch  size  accompanying  lon- 
gevity in  passerine  birds.  This  might  be  due 
to  the  difficulty  in  obtaining  clutch  sizes  from 
the  rare  individuals  that  reach  more  than  a few 
years  of  age.  In  the  present  study  we  obtained 
clutch  measurements  from  several  birds  more 
than  four  years  old,  including  one  each  from 
six,  seven,  and  eight  year  old  birds.  One  fe- 
male originally  caught  as  an  adult  in  1990  was 
recaptured  a total  of  8 times  from  1990—1997. 
Her  first  clutch  sizes  were  4 (1990),  5 (1992), 
5 (1993),  5 (1995),  5 (1996),  and  5 (1997).  In 
1997  the  bird  was  at  least  eight  years  of  age. 
The  previous  longevity  record  for  Prothono- 
tary Warblers  was  5 years,  1 1 months  (Ken- 
nard  1975). 

Conservation  measures,  including  intensive 
use  of  predator-proof  nest  boxes,  have  been 


successful  in  increasing  local  abundance  of 
Prothonotary  Warblers  (Blem  and  Blem 
1992).  However,  elimination  of  predation  at 
nest  boxes  could  skew  age  structures  of  war- 
bler populations  either  by  increasing  produc- 
tion of  young  or  by  decreasing  mortality  of 
adult  females  nesting  in  boxes.  Skewed  age 
structures  could  then  affect  clutch  size  and  in- 
fertility rates,  this  factor  must  be  taken  into 
account  in  any  analysis  of  annual  variations 
in  clutch  size.  For  conservation  of  the  species, 
reduced  clutch  size  and  greater  infertility  of 
young  birds  seem  to  have  only  a modest  im- 
pact on  reproductive  performance  of  Protho- 
notary Warblers.  Furthermore,  effects  of  se- 
nility were  not  obvious  even  in  relatively  old 
warblers. 

ACKNOWLEDGMENTS 

We  thank  D.  Brehmer,  C.  Cosgrove,  S.  Horne,  B. 
Monroe,  J.  Reilly,  R.  Reilly,  A.  Seidenberg,  K.  Sei- 
denberg,  and  T.  Thorp  for  help  in  monitoring  nest  box- 
es. We  are  grateful  to  B.  Brady,  refuge  manager  of 
Presquile  National  Wildlife  Refuge,  for  his  continued 
cooperation  in  this  research.  The  North  American 
Bluebird  Society  provided  funds  to  place  nest  boxes 
on  metal  pipes  to  reduce  predation  of  warblers.  The 
comments  of  two  anonymous  reviewers  significantly 
improved  an  earlier  version  of  this  manuscript. 

LITERATURE  CITED 

Blem,  C.  R.  and  L.  B.  Blem.  1991.  Nest  box  selection 
by  Prothonotary  Warblers.  J.  Field  Ornithol.  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 
1 1 1:197-200. 

Blem,  C.  R.,  L.  B.  Blem,  and  L.  S.  Berlinghoff. 
1999.  Old  nests  in  Prothonotary  Warbler  nest  box- 
es: effects  on  reproductive  performance.  J.  Field 
Ornithol.  70:95-100. 

Bryant,  D.  M.  1988.  Lifetime  reproductive  success  of 
House  Martins.  Pp.  173-188  in  Reproductive  suc- 
cess (T.  H.  Clutton-Brock,  Ed.).  Univ.  of  Chicago 
Press,  Chicago,  Illinois. 

CuRSON,  J.,  D.  Quinn,  and  D.  Bendle.  1994.  Warblers 
of  the  Americas.  Houghton-Mifflin  Co.,  Boston, 
Massachusetts. 

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

Klomp,  H.  1970.  The  determination  of  clutch-size  in 
birds:  a review.  Ardea  58:1-124. 

Kluijver,  H.  N.  1935.  Waarnemingen  over  de  leven- 
swijze  van  den  Spreeuw  (Sturnus  v.  vulf^ciris  L.)  met 
behulp  van  geringde  individuen.  Ardea  24:133-166. 


SHORT  COMMUNICATIONS 


581 


McCleery,  R.  H.  and  C.  M.  Perrins.  1989.  Great  Tit. 
Pp.  35-53  in  Lifetime  reproduction  in  birds  (I. 
Newton,  Ed.).  Academic  Press,  New  York. 

Petit,  L.  S.  1989.  Breeding  biology  of  Protlionotary 
Warblers  in  riverine  habitat  in  Tennessee.  Wilson 
Bull.  101:51-61. 

Pyle,  P.  1997.  Identification  guide  to  North  American 
birds.  Part  I.  Columbidae  to  Ploceidae.  Slate 
Creek  Press,  Bolinas,  California. 

S^THER,  B.  1990.  Age-specific  variation  in  reproduc- 
tive performance  of  birds.  Current  Ornithol.  7; 
251-283. 


SAS  Institute.  1990.  SA.SASTAT®  user's  guide,  ver- 
sion 6,  fourth  ed.  Vol.  2.  SAS  Institute,  Inc.,  Cary, 
North  Carolina. 

Sauer,  J.  R.,  J.  E.  Hines,  G.  Gough,  1.  Thomas,  and 
B.  G.  Peterjohn.  1997.  The  North  American 
breeding  bird  survey  results  and  analysis.  Patux- 
ent Wildlife  Research  Center,  Laurel,  Maryland. 

Sternberg,  H.  1989.  Pied  Flycatcher.  Pp.  55-74  in 
Lifetime  reproduction  in  birds  (1.  Newton,  Ed.). 
Academic  Press,  New  York. 

Zar,  j.  H.  1984.  Biostatistical  analysis.  Second  ed. 
Prentice-Hall,  Inc.,  Englewood  Cliffs,  New  Jersey. 


Wilson  Bull.,  111(4),  1999,  pp.  581-584 

Hybridization  Between  Clay-colored  Sparrow  and  Field  Sparrow  in 

Northern  Vermont 

David  J.  Hoag' 


ABSTRACT. — A male  sparrow  showing  hybrid 
characteristics  between  Clay-colored  {Spizella  pallida) 
and  Field  sparrows  (Spizella  pusilla)  was  first  observed 
in  Grand  Isle,  Vermont,  in  1997.  In  1998,  the  same 
hybrid  defended  a territory  and  mated  with  a female 
Field  Sparrow.  The  pair  produced  one  fledgling.  The 
hybrid's  signature  song  was  composed  of  the  buzzy 
notes  of  a Clay-colored  Sparrow  rising  to  a final  trill 
as  if  copying  a Field  Sparrow's  accelerating  clear  whis- 
tles. Received  18  Dec.  1998,  accepted  9 May  1999. 


I have  found  only  two  previous  records  of 
Clay-colored  Sparrows  (Spizella  pallida)  and 
Field  Sparrows  (Spizella  pusilla)  cooperating 
at  a nest.  Finch  and  Smart  (1974)  mention, 
without  further  details,  a Clay-colored  Spar- 
row found  breeding  with  a Field  Sparrow  at 
Rockefeller  Institute,  Dutchess  County,  New 
York;  “young  were  taken  for  study.”  The  one 
example  of  hybridization  presented  by  Knap- 
ton  (1994)  is  the  account  by  Brooks  (1980)  of 
a trio  of  adults,  a male  Clay-colored  Sparrow 
and  a pair  of  Field  Sparrows,  at  a nest  near 
Millbrook,  Dutchess  County,  New  York;  how- 
ever, “the  fledged  young  appeared  identical  to 
young  Field  Sparrows.”  Carey  and  coworkers 
(1994)  refer  to  the  same  report  as  “possible” 
hybridization.  Hybridization  between  these 


' 173  West  Shore  Rd.,  Grand  Isle,  VT  05458; 
E-mail:  sr71blbrd@aol.com 


two  species  is  not  unexpected  because  of  their 
close  phylogenetic  relationship  (Patten  and 
Fugate  1998).  Examples  exist  of  apparent 
crossbreeding  between  Clay-colored  Sparrows 
and  other  Spizella  species,  and  between  Chip- 
ping Sparrow  (Spizella  passerina)  and  Brew- 
er’s Sparrow  (Spizella  breweri;  Knapton  1994, 
Pyle  and  Howell  1996). 

Clay-colored  Sparrows  are  rarely  reported 
in  Vermont  (Faccio  et  al.  1997,  1998).  In  con- 
trast, Field  Sparrows  may  be  abundant  in 
proper  habitat  such  as  the  abandoned  over- 
grown fields  and  pastures  of  Grand  Isle,  a 
town  on  Lake  Champlain  in  northwestern  Ver- 
mont. There,  on  29  May  1997,  I identified  a 
Clay-colored  Sparrow  by  its  song  which  con- 
sisted of  two  long  buzzes.  At  09:30,  11:00, 
and  16:30  EST,  for  a total  of  30  minutes,  I 
listened  to  the  sparrow  sing  from  elevated 
perches  in  a grassy  clearing  surrounded  by  red 
cedar  (Juniperus  virginiana),  staghorn  sumac 
(Rhus  typhina),  and  common  barberry  (Ber- 
heris  vulgaris).  This  sparrow  was  relocated 
650  m north  on  2 June,  and  last  heard  on  5 
June. 

From  23  July  through  14  August,  1 ob- 
served and  recorded  a second  Clay-colored 
Sparrow  in  a similar  clearing  300  m southwest 
of  the  original  location.  I recorded  the  songs 
on  a microcassette  recorder  and  transferred  the 
songs  to  a computer  using  either  Creative 


582 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


3 kHz 


2 kHz 


Time  1 s 


5 kHz 


4 kHz  - 


3 kHz 


2 kHz 


B 


5 kHz 


4 kHz  - 


3 kHz 


2 kHz 


Time  1 s 


Time  1 s 


D 


Time  1 s 


FIG.  1.  Three  sonograms  illustrate  the  buzzy  songs  of  the  probable  Clay-colored/Field  Sparrow  hybrid.  A. 
Trilled  version  of  the  hybrid  Clay-colored  Sparrow’s  song  recorded  July  1997.  B.  Basic  song  of  the  same  hybrid 
Clay-colored  Sparrow  recorded  May  1998.  C.  Trilled  version  of  the  same  hybrid  sparrow’s  song  recorded  May 
1998.  The  two  trilled  songs  have  a very  similar  pattern  to  a Field  Sparrow’s  clear  whistles  as  illustrated  by  D. 
D.  A Field  Sparrow’s  basic  song  recorded  May  1998. 


Technology’s  Sound  Blaster  Pro  Voice  Editor 
version  2.08,  or  Microsoft  Windows  Sound 
Recorder  version  3.1.  The  sonograms  were 
created  in  Windows  3.1  with  Spectrogram 
(version  2.3;  Horne  1995). 

I assumed  this  second  bird  was  a different 
Clay-colored  Sparrow  because  its  song,  5-8 
short  buzzes,  was  extraordinarily  different 
from  the  first  bird’s  song  of  two  long  buzzes. 
The  second  sparrow’s  song  frequently  ended 
with  a buzzy  trill  that  varied  in  duration  (Fig. 
lA).  With  the  added  trill,  the  song  resembled 
a Field  Sparrow’s  basic  song  (Fig.  ID).  The 
sonograms  show  the  similarity  of  the  song 
pattern  to  a Field  Sparrow  song,  and  show  the 
dissimilarity  of  the  buzzy  notes  to  the  clear 
whistles  of  a Field  Sparrow.  The  mimicry  ei- 
ther was  learned  from  Field  Sparrows  in  the 
surrounding  area  (Knapton  1994)  or  perhaps 
was  a product  of  hybridization. 

Although  the  face  markings  of  the  second 


sparrow  seemed  somewhat  indistinct,  the 
plumage  pattern  was  generally  compatible 
with  Clay-colored  Sparrow.  The  wide  central 
stripe  and  the  streaking  on  the  crown  were 
typical  of  Clay-colored  Sparrow.  The  shapes 
of  the  bill  and  the  tail  were  also  representative 
of  Clay-colored  Sparrow.  However,  the  red- 
dish tint  of  its  plumage  and  the  pink  color  of 
its  entire  bill  caused  me  and  other  observers 
to  accept  it  as  a probable  hybrid  between  a 
Clay-colored  Sparrow  and  a Field  Sparrow. 
Two  observers  thought  the  flank  color  at  the 
bend  of  the  wing  was  indicative  of  Field  Spar- 
row parentage.  The  Vermont  Bird  Records 
Committee  agreed  with  the  hybrid  designation 
at  its  November  1998  meeting  (Nicholson, 
pers.  comm.). 

From  7—14  April  1998,  Field  Sparrows  re- 
turned to  the  area.  On  28  April,  on  the  same 
territory  that  had  been  occupied  by  the  hybrid 
in  July  and  August  of  1997,  I found  a bird 


SHORT  COMMUNICATIONS 


583 


FIG.  2.  A Clay-colored  Sparrow’s  mimicry  of  a Field  Sparrow’s  complex  song  as  illustrated  by  two  sono- 
grams. A.  The  new  Clay-colored  Sparrow’s  four-part  song  recorded  8 July  1998.  B.  A Field  Spanow’s  three- 
part  complex  song  recorded  July  1998. 


whose  coloration  matched  that  of  the  hybrid  I 
had  seen  in  1997.  Also,  the  bird’s  song  was 
the  same  distinct  vocalization  of  about  eight 
buzzes  (Fig.  IB)  with  trilled  notes  occasion- 
ally added  (Fig.  1C),  as  heard  and  recorded  in 
1997  (Fig.  lA).  According  to  Knapton  (1994), 
Clay-colored  Sparrows  retain  their  song  type 
from  one  year  to  the  next. 

On  23-24  July  1997,  the  hybrid  sang  rapid 
eight  buzz  songs  interspersed  with  seven  buzz 
songs  having  the  added  trill.  The  length  of  the 
trill  varied;  twice  it  was  very  short  with  only 
three  notes.  The  repertoire  from  25  July  to  3 
August  consisted  of  five  buzzes  heard  14 
times,  seven  buzzes  heard  9 times,  seven 
buzzes  with  a trill  heard  17  times,  eight  buzz- 
es heard  over  50  times,  and  nine  buzzes  heard 
twice.  Errors  in  judging  sounds  may  have  af- 
fected the  true  syllable  count.  The  sonograms 
verify  the  softness  of  the  initial  buzz.  In  two 
variant  seven  buzz  songs,  the  fifth  syllable 
was  abbreviated.  One  long  song  of  faster 
buzzing  was  heard  on  7 August.  In  1998,  the 


hybrid’s  songs  were  less  variable.  I heard  few 
five  buzz  songs,  and  seven  buzz  songs  without 
a trill  were  very  scarce.  Approximately  25% 
of  the  songs  were  seven  buzzes  with  a trill; 
75%  were  eight  buzzes. 

By  3 May  1998,  the  hybrid  had  moved 
about  100  m north.  Singing  occurred  less  fre- 
quently during  the  second  half  of  May  when 
the  hybrid  acquired  a Field  Sparrow  as  a mate. 

I made  frequent  observations  throughout  May 
and  June  and  found  no  extra  Field  Sparrows 
within  the  territory.  I did  not  observe  any  ex- 
tra-pair copulations. 

Nest  inspections  at  09:00—10:30  revealed 
no  egg  on  4 June,  one  egg  on  5 June,  two 
eggs  on  7 June,  and  three  eggs  on  8 June,  the 
first  day  that  the  Field  Sparrow  was  brooding. 
On  19-23  June,  the  female  spaiTow  sat  on  the 
nest,  blocking  viewing  of  the  nest  contents. 
Both  adult  sparrows  carried  food  to  the  nest 
21-26  June.  On  28  June,  the  abandoned  nest 
contained  two  infertile  eggs.  Both  adults  chap- 
eroned me  and  chipped  continuously  as  I at- 


584 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


tempted  to  view  the  single  fledgling  at  07:45 
and  13:35.  The  hybrid  carried  food  on  2 and 
5 July  and  continued  to  scold  me  throughout 
July.  On  26  July,  the  hybrid  guarded  the  ju- 
venile. The  two  adults  and  the  juvenile  were 
still  in  close  association. 

The  female  Field  Sparrow’s  reaction  during 
my  unsuccessful  search  for  a second  nest  on 
26  July  indicated  that  a second  brood  existed. 
I saw  the  hybrid  carrying  food  again  on  5 Au- 
gust, a late  date  for  assisting  the  first  brood 
juvenile  hatched  approximately  40  days  ear- 
lier. In  December  1998  I found  the  second 
nest  which  was  obscured  by  grass  within  a 
thicket  of  barberry  centrally  located  in  the  0.3 
ha  territory.  This  second  nest  contained  one 
Field  Sparrow  egg.  Since  all  1997  Field  Spar- 
row nests  in  the  surrounding  area  were  de- 
stroyed by  ice  accumulation  during  three  days 
of  freezing  rain  in  January  1998,  this  undam- 
aged nest  provided  additional  evidence  that 
the  hybrid  and  its  mate  raised  a second  1998 
brood. 

I did  not  hear  the  hybird  singing  from  mid- 
June  until  5 July,  the  day  a new  Clay-colored 
Sparrow  began  a four  day  encroachment  upon 
the  territory.  The  new  Clay-colored  Sparrow’s 
song  (Fig.  2A)  closely  resembled  a Field 
Sparrow  song  (Fig.  2B).  All  songs  of  this  new 
Clay-colored  Sparrow  were  identical  except 
for  minor  variations  in  length.  The  hybrid  re- 
sponded with  its  own  songs  (Fig.  lA-C) 
through  12  July. 

The  ability  of  some  emberizid  sparrows  to 
learn  other  species’  songs  (Tasker  1955,  Bap- 
tista  et  al.  1981),  possibly  useful  in  defense  of 
territory,  may  also  attract  mates  from  closely 
related  species.  Albrecht  and  Oring  (1995)  in- 
dicated that  the  primary  function  of  song  for 
Chipping  Sparrows  is  mate  attraction  rather 
than  territorial  defense.  The  song  mimesis 
may  result  in  occasional  interspecific  pairing 
of  Spizella  species.  Unpublished  accounts  of 
Spizella  mimesis  include  Chipping  Sparrows 
and  Clay-colored  Sparrows  singing  each  oth- 
er’s .song  (Bailey,  pers.  comm.)  and  a Field 
Sparrow  singing  a Chipping  Sparrow  song 
(unpubl.  data). 

The  probable  hybrid’s  unusual  plumage,  its 
atypical  song,  its  pairing  with  a female  Field 
Sparrow,  and  its  intensive  parental  activity  at 
the  nest  imply  that  crossbreeding  between 


Clay-colored  Sparrows  and  Field  Sparrows 
occurs.  A DNA  study  of  these  individuals 
might  verify  hybridization. 

Note  added  in  proof:  The  hybrid  sparrow 
returned  5 May  1999  and  eventually  paired 
with  a female  Field  Sparrow  9 July  to  2 Au- 
gust. His  songs  in  1999  initially  were  identical 
to  those  in  1997  and  1998,  but  stopped  in- 
cluding the  trill  during  the  last  half  of  the 
breeding  season. 

ACKNOWLEDGMENTS 

I thank  D.  Capen  for  reviewing  and  assisting  with 
the  manuscript.  I thank  the  referees  for  their  assistance 
and  suggestions  for  improving  the  manuscript. 

LITERATURE  CITED 

Albrecht,  D.  J.  and  L.  W.  Oring.  1995.  Song  in  Chip- 
ping Sparrows,  Spizella  pa.sserina:  structure  and 
function.  Anim.  Behav.  50:1233-1241. 

Baptista,  L.  E,  M.  L.  Morton,  and  M.  E.  Pereyra. 
1981.  Interspecific  song  mimesis  by  a Lincoln 
Sparrow.  Wilson  Bull.  93:265-267. 

Brooks,  E.  W.  1980.  Interspecific  nesting  of  Clay-col- 
ored and  Field  sparrows.  Wilson  Bull.  92:264- 
265. 

Carey,  M.,  D.  E.  Burhans,  and  D.  A.  Nelson.  1994. 
Field  Sparrow  (Spizella  pu.silla).  In  The  birds  of 
North  America,  no.  103  (A.  Poole  and  F.  Gill, 
Eds.).  The  Academy  of  Natural  Sciences,  Phila- 
delphia, Pennsylvania;  The  American  Ornitholo- 
gists' Union,  Washington,  D.C. 

Faccio,  S.,  J.  Nicholson,  J.  Peterson,  and  C.  Rimmer. 
1997.  Vermont  bird  records  committee  report 
1995.  Vermont  Institute  of  Natural  Science, 
Woodstock. 

Faccio,  S.,  J.  Nicholson,  J.  Peterson,  B.  Pfeiffer, 
AND  C.  Rimmer.  1998.  Vermont  bird  records  com- 
mittee report  1996.  Vermont  Institute  of  Natural 
Science,  Woodstock. 

Finch,  D.  W.  and  R.  Smart.  1974.  Clay-colored  Spar- 
row. Kingbird  24:211. 

Horne,  R.  S.  1995.  Spectrogram,  version  2.3.  URL  = 
www.intranet.csupomona.edu/~biology/ 
gram.html 

Knapton,  R.  W.  1994.  Clay-colored  Sparrow  (Spizella 
pallida).  In  The  birds  of  North  America,  no.  120 
(A.  Poole  and  F.  Gill,  Eds.).  The  Academy  of  Nat- 
ural Sciences,  Philadelphia,  Pennsylvania;  The 
American  Ornithologists’  Union,  Washington, 
D.C. 

Patten,  M.  A.  and  M.  Fugate.  1998.  Systematic,  re- 
lationships among  the  emberizid  sparrows.  Auk 
1 15:412-424. 

Pyle,  P.  and  S.  N.  G.  Howell.  1996.  Spizella  spar- 
rows: intraspecific  variation  and  identification. 
Birding  28:374-387. 

Tasker,  R.  R.  1955.  Chipping  Sparrow  with  song  of 
Clay-colored  Sparrow  at  Toronto.  Auk  72:303. 


Wilson  Bull.,  I 1 1(4),  1999.  pp.  585-588 


Commentary 


A CRITIQUE  OF  WANG  YONG  AND  FINCH’S 
FIEFD-IDENTIFICATIONS  OF  WIFFOW  FFYCATCHER 
SUBSPECIES  IN  NEW  MEXICO 

John  P.  Hubbard' 


In  a recent  paper  in  the  Wilson  Bulletin, 
Wang  Yong  and  Finch  (1997;  henceforth 
Y&F)  reported  that  they  subspecifically  iden- 
tified 83  of  84  Willow  Flycatchers  {Ernpidon- 
ax  traillii)  captured,  banded,  and  released  in 
central  New  Mexico  in  spring  and  autumn 
1994  and  1995.  Given  the  nature  of  these  sub- 
species and  the  means  by  which  Y&F  appar- 
ently identified  them,  I am  extremely  doubtful 
about  the  reliability  of  their  determinations 
and  thus  the  validity  of  these  as  scientific  data. 
The  fact  is  that  identifying  these  taxa  is  quite 
difficult,  even  for  trained  taxonomists  working 
in  the  laboratory  under  the  best  protocols  and 
conditions.  This  difficulty  stems  from  a num- 
ber of  factors,  the  major  one  being  the  per- 
vasive subtlety  of  the  plumage-color  charac- 
ters by  which  these  subspecies  mainly  differ. 
Not  surprisingly,  these  differences  are  difficult 
to  describe  in  words,  which  is  exacerbated  by 
the  fact  that  none  of  the  available  classifica- 
tion systems  accurately  portrays  the  range  of 
plumage  coloration  observed  in  this  flycatcher 
(e.g..  Browning  1993).  This  means  that  this 
species’  plumage-color  characters  are  best  ob- 
served in  specimens  (i.e.,  study  or  flat  skins), 
which  also  provide  the  best  avenue  for  iden- 
tifying subspecies.  To  do  this,  one  must  first 
assemble  series  of  skins  representing  all  rel- 
evant taxa,  as  well  as  such  important  subcat- 
egories as  age  classes  (e.g.,  adult  vs  immature) 
and  seasonal  groupings  (e.g.,  spring  vs  au- 
tumn). Then  one  sorts  “unknowns”  (which 
could  include  live  birds)  into  subcategories 
and  compares  them  to  the  taxa  therein,  which 
should  produce  at  least  tentative  subspecific 
identifications.  In  fact,  this  is  the  standard  lab- 
oratory approach  for  identifying  color-based 
subspecies,  and  it  is  the  only  means  proven 


' Route  5,  Box  431,  Espanola,  New  Mexico  87532. 


reliable  for  this  purpose  in  the  Willow  Fly- 
catcher. 

As  my  earlier  comments  suggest,  I do  not 
believe  Y&F  used  the  approach  described 
above  in  their  attempts  to  identify  subspecies 
in  the  Willow  Flycatcher.  In  other  words,  they 
did  not  take  synoptic  series  of  study  skins  into 
the  field,  against  which  the  birds  they  captured 
were  compared  to  determine  subspecific  iden- 
tities. However,  I cannot  be  100%  certain 
about  this  because  the  methods  section  in  their 
paper  is  so  incomplete  and  otherwise  deficient 
one  can  only  guess  at  many  aspects  of  their 
approach.  Nonetheless,  it  seems  logical  that  if 
they  had  used  skins  as  the  basis  for  their  iden- 
tifications, they  would  have  said  so.  Given  this 
assumption,  if  they  did  not  use  skins,  how  did 
they  go  about  identifying  their  birds  to  sub- 
species? On  this  matter  Y&F  are  at  best 
vague,  providing  a few  clues  but  no  definitive 
explanations  of  their  identification  methodol- 
ogy. For  example,  we  are  told  that  they 
“.  . . adopted  the  four-subspecies  classification 
system  of  Hubbard  (1987)  and  Unitt  (1987), 
in  which  “subspecies  identity  ...  is  based  [in 
part]  on  . . . coloration  of  the  head  [=  crown] 
and  neck  [=  forenape]  and  its  contrast  with 
the  back,  and  the  contrast  between  the  breast- 
band  and  the  throat  (see  Phillips  1948,  Hub- 
bard 1987,  Unitt  1987,  Browning  1993).” 
Based  on  this,  I assume  that  Y&F  chose  lit- 
erature descriptions  (as  opposed  to  specimen 
comparisons)  as  the  basis  for  their  identifica- 
tion of  Willow  Flycatcher  subspecies.  In  ad- 
dition, 1 also  suspect  they  converted  these  de- 
scriptions into  the  color  values  of  Smithe 
(1975),  as  this  is  the  system  they  used  to  clas- 
sify coloration  in  birds  captured  in  the  field. 
Beyond  this,  one  could  also  speculate  on  such 
matters  as  (a)  how  converted  values  were  ac- 
tually used  to  identify  birds,  e.g.,  whether  in 
a dichotomous  key,  probability  table,  or  other 


585 


586 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


framework;  or  (b)  what  Y&F’s  perceptions 
were  of  color  characters  in  various  races,  giv- 
en that  no  such  descriptions  were  offered  by 
them.  However,  I see  no  purpose  in  further 
speculation  concerning  these  or  other  aspects 
of  their  methodology.  This  is  because  if  they 
did  base  their  identifications  on  the  literature 
rather  than  specimens,  1 believe  the  process 
became  so  flawed  that  the  details  are  irrele- 
vant— like  rearranging  deck  chairs  on  the 
sinking  Titantic! 

The  message  here  is  that  the  literature  is  no 
substitute  for  specimen  comparisons  for  any- 
one attempting  to  identify  Willow  Flycatcher 
subspecies,  at  least  if  attaining  the  most  reli- 
able scientific  data  is  the  goal.  Furthermore, 
given  logistical  and  other  problems,  1 doubt 
even  specimen  comparisons  would  consistent- 
ly yield  reliable  identifications  of  live  birds 
under  field  conditions.  Not  only  would  it  be 
unwieldy  to  take  and  use  museum  skins  in  the 
field,  but  setting  up  and  maintaining  constant 
conditions  (e.g.,  lighting)  would  also  be  dif- 
ficult. In  addition,  except  for  recaptures,  only 
one  opportunity  would  be  available  to  identify 
each  live  bird  in  the  field.  This  means  that  one 
could  not  reassess  identifications  at  a later 
time,  which  is  both  frequent  and  necessary 
when  studying  specimens  in  the  laboratory.  In 
this  regard,  photographs  and  certainly  color 
readings  (e.g.,  from  Smithe  1975)  would  not 
be  adequate  for  such  reexaminations  because 
these  do  not  exactly  duplicate  colors  observed 
in  the  birds  or  specimens  themselves.  Given 
these  considerations,  I believe  that  identifying 
subspecies  in  the  Willow  Flycatcher  is  best 
done  in  the  laboratory,  using  study  skins  ex- 
amined under  proper  protocols  and  procedures 
by  people  trained  in  the  process.  In  other 
words,  this  is  a task  that  should  be  left  to  an 
alpha-taxonomic  approach,  which  is  appropri- 
ate when  one  considers  that  subspecies  arose 
and  largely  remain  as  products  of  that  realm. 

Even  when  approached  as  outlined  above, 
the  reality  is  that  not  every  specimen  or  even 
population  of  this  flycatcher  can  be  reliably 
assigned  to  subspecies.  Intergradation  and 
overlap  occur  in  all  characters  that  distinguish 
these  taxa,  so  birds  exhibiting  such  character- 
istics may  be  un-  or  misidentified  as  a result. 
In  addition,  characteristics  in  some  popula- 
tions remain  poorly  known,  mainly  because  of 
the  paucity  of  specimens  from  these  areas.  For 


example,  in  the  latest  revision  of  the  species. 
Browning  (1993)  could  only  assemble  270 
specimens  of  breeding  season  adults — includ- 
ing fewer  than  20  of  the  endangered  subspe- 
cies E.  t.  extirnus  of  the  Southwest.  As  a con- 
sequence, it  is  not  surprising  that  he  ques- 
tioned boundaries  between  four  of  the  five 
subspecies  recognized  in  his  paper.  Even 
when  populational  characteristics  are  better 
known,  opinions  may  differ  as  regards  their 
taxonomic  treatment.  Thus,  Browning  (1993) 
recognized  two  subspecies  (i.e.,  E.  t.  trail  Hi 
and  E.  t.  campestris)  as  breeding  in  the  region 
east  of  the  Rocky  Mountains,  whereas  Unitt 
(1987)  merged  the  latter  with  the  nominate 
form.  Differences  in  opinion  also  exist  on  a 
broader  scale,  such  as  concerning  the  overall 
number  of  subspecies  recognizable  in  the  Wil- 
low Flycatcher.  For  example,  some  taxono- 
mists maintain  that  none  should  be  recognized 
(e.g.,  Mayr  and  Short  1970,  Traylor  1979), 
while  others  accept  four  to  six  as  valid  (e.g., 
Phillips  1948,  Aldrich  1951,  Wetmore  1972, 
Oberholser  1974,  Unitt  1987,  Browning 
1993).  Thus,  although  specimen  comparisons 
provide  our  only  reliable  means  for  identify- 
ing subspecies  in  this  flycatcher,  this  approach 
must  be  used  with  the  clear  recognition  that  it 
is  just  the  first  step  in  this  very  difficult  en- 
deavor. 

Incidentally,  the  above  differences  in  taxo- 
nomic opinion  present  a problem  for  those 
that  rely  largely  or  entirely  on  the  literature 
for  their  knowledge  of  geographic  variation  in 
this  species.  That  is,  how  does  one  choose 
which  authorities  to  follow  and  thus  which 
viewpoints  to  accept  on  this  subject?  Among 
others,  one  way  around  this  would  be  to  ad- 
here strictly  to  a single  point  of  view,  such  as 
the  recent  revision  of  this  flycatcher  by 
Browning  (1993).  However,  Y&F  chose  not 
do  this,  instead  electing  to  cobble  their  con- 
cept of  variation  from  a variety  of  sources 
(e.g.,  Phillips  1948,  Hubbard  1987,  Unitt 
1987,  Browning  1993).  Given  the  lack  of  con- 
sensus among  these  sources,  this  was  a ques- 
tionable decision.  In  fact,  it  would  be  a chal- 
lenge even  for  people  with  firsthand  experi- 
ence with  geographic  variation  in  this  species, 
as  seen  from  the  variety  of  opinions  cited 
above.  As  a consequence,  it  is  not  surprising 
that  I would  quibble  with  Y&F’s  choices,  in- 
cluding that  of  which  authorities  to  follow. 


Huhhard  • COMMENTARY 


587 


For  example,  as  indicated  earlier,  they  cited 
my  unpublished  paper  (Hubbard  1987)  as  a 
basis  for  the  “four-subspecies  classification 
system”  adopted  in  their  study.  However,  that 
so-called  system  was  actually  a cobbling  job 
itself,  my  aim  being  to  summarize  color  char- 
acters of  various  subspecies  from  the  treat- 
ments of  Phillips  (1948),  Aldrich  (1951),  Wet- 
more  (1972),  and  Oberholser  (1974).  As  such, 
it  was  not  meant  either  to  provide  definitive 
descriptions  of  these  subspecies  or  to  recom- 
mend which  should  be  recognized  as  valid. 
For  it  to  have  been  otherwise  used  by  Y&F 
may  seem  flattering,  but  it  certainly  was  not  a 
sound  decision  from  a taxonomic  viewpoint. 

Given  the  flawed  nature  of  their  approach, 
it  is  no  surprise  that  Y&F’s  findings  on  Wil- 
low Flycatcher  subspecies  would  also  be  open 
to  question.  For  example,  when  compared 
with  what  is  known  from  specimens  (e.g., 
Hubbard  1987),  significant  differences  emerge 
on  the  New  Mexico  status  of  three  of  the  four 
taxa  recognized  in  that  study.  (In  light  of  the 
relative  scientific  standing  of  the  two  sources, 
I would  obviously  accept  the  specimen  ver- 
sion over  that  of  Y&F  in  every  case.)  The 
most  significant  difference  occurs  in  the  sub- 
species E.  t.  brewsteri  (sensu  stricto),  which 
breeds  along  the  Pacific  slope  of  North  Amer- 
ica. Although  occurring  regularly  in  migration 
eastward  to  Arizona  (Monson  and  Phillips 
1981),  this  form  has  rarely  been  collected  east 
and  north  of  that  state,  e.g.,  in  Utah  (Behle 
1985),  Colorado  (Bailey  and  Niedrach  1965), 
Oklahoma  (Sutton  1967),  and  Texas  (Ober- 
holser 1967).  Hard  data  from  New  Mexico 
clearly  conform  to  this  pattern,  with  only  two 
(4.7%)  of  the  43  specimens  so  attributed  in 
Hubbard  (1987)  and  even  these  were  some- 
what equivocal.  By  contrast,  Y&F  identified 
33  (39.8%)  of  their  83  birds  as  E.  t.  brewsteri, 
which  is  about  8.5  times  more  frequent  than 
reported  by  Hubbard.  Another  notable  depar- 
ture involves  the  subspecies  E.  t.  traillii  (in 
which  Y&F  include  E.  t.  campestris),  which 
breeds  from  the  Great  Plains  to  the  northeast- 
ern Atlantic  Coast.  In  the  Southwest,  E.  t. 
traillii/campestris  occurs  regularly  in  the 
plains  of  eastern  Colorado  (Bailey  and  Nied- 
rach 1965)  and  New  Mexico  (Hubbard  1987), 
but  it  has  not  been  collected  as  far  west  as 
Arizona  (Monson  and  Phillips  1981).  Yet 
Y&F  reported  that  8.4%  of  their  birds  were 


this  form,  even  though  the  the  middle  Rio 
Grande  Valley  lies  some  200  miles  west  of  the 
nearest  specimen  localities  in  New  Mexico. 
Finally  is  the  race  E.  t.  adastus,  which  breeds 
widely  in  the  interior  U.S.  north  of  the  .south- 
western states,  through  which  it  passes  in  both 
spring  and  autumn.  In  New  Mexico,  it  com- 
prised 25.6%  of  the  specimens  reported  by 
Hubbard  (1987),  compared  to  10.8%  in  Y&F’s 
sample. 

As  for  the  fourth  subspecies  {E.  t.  extimus), 
Y&F  identified  34  (41.1%)  of  their  birds  as 
this  form,  compared  to  the  48.8%  from 
throughout  New  Mexico  by  Hubbard  (1987). 
Thus,  on  the  face  of  it,  their  findings  would 
seem  not  to  differ  significantly  from  what  is 
known  from  specimens  of  this  taxon.  How- 
ever, the  number  of  questionable  literature  re- 
cords of  this  subspecies  suggests  it  may  be 
more  subject  to  misidentification  than  certain 
other  forms,  such  E.  t.  brewsteri  and  E.  t. 
traillii  (both  sensu  lato).  Birds  that  might  be 
mistaken  for  E.  t.  extimus  could  include  sun- 
bleached  or  worn  individuals  of  other  races, 
as  well  as  pale  variants  of  E.  t.  adastus,  in- 
tergrades between  the  latter  and  E.  t.  extimus, 
and  carelessly-examined  E.  t.  campestris.  If  so 
misidentified,  such  instances  could  help  ex- 
plain records  of  E.  t.  extimus  from  areas  out- 
side its  known  breeding  range,  such  as  the 
northern  two-thirds  of  Colorado  (Bailey  and 
Niedrach  1965)  and  Texas  east  of  the  Trans- 
Pecos  region  (Oberholser  1974).  As  for  New 
Mexico,  1 am  dubious  of  E.  t.  extimus  records 
from  the  eastern  plains,  such  as  two  speci- 
mens reported  in  Hubbard  (1987)  from  Roo- 
sevelt County.  In  addition,  1 have  definitely 
reidentified  two  of  the  purported  E.  t.  extimus 
from  that  report,  one  from  San  Juan  County 
(=  E.  t.  adastus  > extimus)  and  another  from 
Socorro  County  (=  E.  t.  extimus  > adastus). 
Of  course,  as  mentioned  earlier,  we  do  not 
have  the  luxury  of  reexamining  E.  t.  e.xtimus 
(or  other  subspecies)  reported  by  Y&F,  so 
their  identifications  cannot  be  reassessed  in 
light  of  potential  sources  of  misidentification. 
Given  this  and  their  flawed  methodology,  1 see 
no  reason  to  regard  their  findings  on  this  form 
as  any  more  acceptable  than  those  on  the  other 
races  reported  in  their  paper.  As  a final  point, 
Y&F  make  no  mention  of  the  differences  be- 
tween their  findings  on  the  various  subspecies 
and  the  specimen  record  as  discussed  above. 


588 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


While  the  need  for  this  would  not  have  been 
obvious  as  regards  E.  t.  extimus  and  perhaps 
even  E.  t.  adastus,  this  could  hardly  have  been 
the  case  with  E.  t.  traillii  and  especially  E.  t. 
brewsteri. 

To  summarize,  geographic  variation  in  the 
Willow  Flycatcher  mainly  involves  subtle  dif- 
ferences in  plumage  coloration,  concerning 
which  taxonomists  disagree  in  terms  of  the 
number  of  subspecies  that  should  be  recog- 
nized. Anyone  contemplating  identifying 
these  subspecies  should  do  so  with  these  ca- 
veats in  mind,  as  well  as  by  approaching  the 
process  through  the  use  of  specimen  compar- 
isons— preferably  in  the  laboratory  under  con- 
trolled conditions  and  with  proper  training  in 
alpha-taxonomic  procedures.  Given  that 
Y&F’s  approach  appears  to  have  been  other- 
wise, I submit  that  their  field  identification  of 
these  subspecies  cannot  be  regarded  as  a bona 
fide  assessment  of  this  parameter  in  the  birds 
they  processed  in  New  Mexico  in  1994  and 
1995.  Furthermore,  for  those  that  would  use 
their  subspecific  findings,  I urge  them  to  do 
so  with  extreme  caution  to  say  the  least.  Be- 
yond this,  I would  like  to  state  that  as  an  al- 
pha-taxonomist, I am  dismayed  that  a study 
with  such  a flawed  approach  to  subspecies 
identification  could  make  its  way  into  print  in 
a major  ornithological  journal.  To  wit,  orni- 
thology has  come  to  rely  almost  entirely  on 
non-specimen  data  for  monitoring  the  distri- 
bution and  status  of  birds  on  this  planet.  While 
not  necessarily  a bad  thing,  sometimes  we 
may  fail  to  recognize  the  very  real  limitations 
of  such  data.  No  better  example  of  this  exists 
than  as  regards  the  identification  of  difficult 
taxa,  of  which  subspecies  in  Empidonax  trail- 
lii provide  a perfect  case  in  point. 


LITERATURE  CITED 

Aldrich,  J.  W.  1951.  A review  of  the  races  of  Traill’s 
Elycatcher.  Wilson  Bull.  63:192-197. 

Bailey,  A.  M.  and  R.  J.  Niedrach.  1965.  Birds  of 
Colorado,  vol.  2.  Denver  Museum  of  Natural  His- 
tory, Denver,  Colorado. 

Behle,  W.  C.  1985.  Utah  birds:  geographic  variation 
and  systematics.  Utah  Mus.  Nat.  Hist.  Occ.  Pub. 
5:1-147. 

Browning,  M.  R.  1993.  Comments  on  the  taxonomy 
of  Empidonax  traillii  West.  Birds  24:241-257. 

Hubbard,  J.  P.  1987.  The  status  of  the  Willow  Fly- 
catcher in  New  Mexico.  Unpublished  report.  New 
Mexico  Dept,  of  Game  and  Fish,  Santa  Fe. 

Mayr,  E.  and  L.  L.  Short,  Jr.  1970.  Species  taxa  of 
North  American  birds.  Nuttall  Ornithithological 
Club,  Cambridge,  Massachusetts. 

Monson,  G.  and  a.  R.  Phillips.  1981.  Annotated 
checklist  of  the  birds  of  Arizona.  Univ.  of  Arizona 
Press,  Tucson. 

Oberholser,  H.  C.  1974.  The  birdlife  of  Texas,  vol. 
2.  Univ.  of  Texas  Press,  Austin. 

Phillips,  A.  R.  1948.  Geographic  variation  in  Etnpi- 
donax  traillii.  Auk  65:507-514. 

Phillips,  A.  R.,  J.  Marshall,  and  G.  Monson.  1964. 
The  birds  of  Arizona.  Univ.  Arizona  Press,  Tuc- 
son. 

Smithe,  F.  B.  1975.  Naturalist’s  color  guide.  American 
Museum  of  Natural  History,  New  York. 

Sutton,  G.  M.  1967.  Oklahoma  birds.  Univ.  of 
Oklahoma  Press,  Norman. 

Traylor,  M.  A.,  JR.  1979.  Check-list  of  birds  of  the 
world,  vol.  8.  Harvard  Univ.,  Cambridge,  Massa- 
chusetts. 

Unitt,  P.  1987.  Empidonax  traillii  extimus'.  an  endan- 
gered subspecies.  West.  Birds  18:137-162. 

Yong,  W.  and  D.  M.  Finch.  1997.  Migration  of  the 
Willow  Flycatcher  along  the  Middle  Rio  Grande. 
Wilson  Bull.  109:253-268. 

Wetmore,  a.  1972.  The  birds  of  the  Republic  of  Pan- 
ama, part  3.  Smithsonian  Misc.  Coll.  150(3):  1- 
631. 


Vonf;  and  Finch  • RESPONSE 


589 


Wilson  Bull.,  111(4),  1999,  pp.  589-592 


RESPONSE 

Wang  Yong'  and  Deborah  M.  Finch*  - 


Hubbard  (1999)  criticizes  our  paper  Migra- 
tion of  the  Willow  Flycatcher  along  the  middle 
Rio  Grande  (Yong  and  Finch  1997),  where  we 
reported  aspects  of  stopover  ecology  of  the 
species  including  timing,  abundance,  fat 
stores,  stopover  length,  and  habitat  use.  Hub- 
bard questions  our  identification  of  subspecies 
of  the  Willow  Flycatcher  {Empidonax  traillii) 
and  the  methods  we  used  to  identify  them.  He 
also  attempts  to  evaluate  the  accuracy  of  our 
results  of  subspecies  composition  by  compar- 
ing them  with  data  from  other  researchers.  We 
welcome  and  applaud  this  scrutiny  in  the  hope 
that  this  interchange  will  stimulate  greater  in- 
terest, research,  and  capability  to  distinguish 
the  phenotypic  characteristics  of  subspecies  of 
the  Willow  Flycatcher.  Given  that  the  south- 
western race  {E.  t.  extimus)  of  the  Willow  Fly- 
catcher is  federally  listed  as  Endangered,  re- 
liable methods  for  identifying  this  subspecies 
need  to  be  developed  to  more  effectively  con- 
serve and  recover  its  populations. 

We  are  aware  that  the  subspecific  taxonomy 
of  the  Willow  Flycatcher  is  inconsistent 
among  taxonomists  as  are  the  techniques  to 
identify  subspecies.  Consequently,  reliable 
identification  of  subspecies  is  difficult,  espe- 
cially in  field  situations.  We  acknowledge  that 
issues  of  taxonomic  status,  population  distri- 
butions, and  identification  methods  of  subspe- 
cies of  the  Willow  Flycatcher  should  be  ex- 
plored further.  However,  Hubbard’s  criticisms 
of  our  paper  are  generally  based  on  erroneous 
information  as  well  as  incorrect  assumptions 
about  our  methods,  and  they  do  not  alter  our 
conclusions  about  Willow  Flycatcher  stopover 
ecology  at  the  species  level. 

Hubbard’s  first  criticism  focuses  on  the 
methods  we  used  for  identifying  the  subspe- 
cies. Rather  than  using  an  assemblage  of  sub- 


' USDA  Forest  Service,  Rocky  Mountain  Forest  Re- 
search Station,  2205  Columbia,  SE,  Albuquerque,  NM 
87106. 

^ Corresponding  author;  E-mail:  Finch_Deborah_M/ 
rmrs_albq@fs.fed.us 


species  skins  as  advocated  by  Hubbard  to 
identify  Willow  Flycatcher  subspecies  in  the 
field,  we  relied  on  descriptions  and  records  of 
coloration  and  morphology  published  in  the 
available  literature  by  taxonomists.  Contrary 
to  what  Hubbard  speculates,  we  did  not  con- 
vert color  descriptions  into  Smithe’s  (1975) 
color  code  values.  We  based  our  identification 
of  back  plumage  color  on  the  most  recent  re- 
search by  Unitt  (1987)  and  Browning  (1993). 
Using  Smithe’s  color  codes  to  describe  back 
plumage,  Unitt  (1987)  writes:  “In  brewsteri 
the  green  is  in  the  direction  of  olive  green 
(color  48),  in  adastus  in  the  direction  of 
greenish  olive  (color  49),  and  in  extimus  and 
traillii  in  the  direction  of  grayish  olive  (color 
43).  That  is,  brewsteri  is  a dark  brownish  ol- 
ive, adastus  a dark  grayish  green,  and  extimus 
and  traillii  a pale  grayish  green.  . .’’  Browning 
(1993)  suggested  that  Smithe’s  color  system 
is  problematical  because  the  color  swatches 
generally  are  not  identical  matches  for  actual 
colors.  Hence,  he  used  Munsell  Color  Charts 
(1990)  to  describe  the  crown  and  back  con- 
trast for  his  specimens.  During  our  fieldwork, 
we  consulted  both  Unitt’s  (1987)  color  codes 
for  subspecies’  back  color  and  Browning’s 
color  contrast  scores  between  crown  and  back. 
Although  Hubbard  suggests  that  live  speci- 
mens have  some  disadvantages,  we  counter 
that  the  plumage  coloration  of  live  birds  is 
more  likely  to  be  true  to  type  than  skin  spec- 
imen plumage  that  may  have  faded.  If  our  hy- 
pothesis that  the  coloration  of  fresh  plumage 
differs  from  that  of  faded  plumage  is  correct, 
then  data  collected  from  live  specimens  may 
be  more  reliable,  or  at  least  not  less  reliable, 
than  results  obtained  from  study  skins.  Birds 
occasionally  called  or  sang  in  our  study  after 
being  released.  Information  about  song  and 
call  characteristics  were  also  recorded  when 
possible.  Such  data  are  available  from  living 
flycatchers  but  not  from  skins.  Sedgewick’s 
(pers.  comm.)  preliminary  analyses  of  Willow 
Flycatcher  song  and  call  signatures  collected 


590 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


in  different  regions  suggest  that  E.  t.  extimus 
song  structure  can  be  distinguished  from  that 
of  its  northern  conspecifics  and  we  used  this 
kind  of  data  to  aid  identification  also. 

We  did  not  rely  solely  on  coloration  for 
subspecies  identification,  contrary  to  Hub- 
bard’s second  assumption.  Unitt  (1987)  sug- 
gested that  wing  formula  (relative  length  of 
primary  feather  length)  can  be  used  to  assist 
subspecies  identification.  Of  the  305  speci- 
mens that  Unitt  (1987)  examined,  wing  for- 
mula distinguished  93%  of  the  E.  t.  extimus 
and  E.  t.  traillii,  88%  of  the  E.  t.  adastus  and 
E.  t.  traillii,  and  89%  of  the  E.  t.  brewsteri 
and  E.  t.  traillii.  Browning  (1993)  also  applied 
wing  formula  to  assess  variation  in  subspecific 
characteristics,  and  his  results  also  demon- 
strated that  wing  formula  may  be  useful  for 
distinguishing  some  subspecies  although  his 
sample  size  was  smaller  than  Unitt’s.  Hubbard 
himself  (1987)  noted  that  E.  t.  brewsteri  was 
smaller  than  other  described  forms.  In  the 
field,  we  relied  partly  on  non-overlapping  ex- 
treme wing  measurements  to  assist  in  the 
identification  of  this  subspecies.  In  addition, 
we  measured  and  recorded  more  than  30  var- 
iables from  each  individual.  Following  Unitt 
(1987),  we  used  wing  formula  to  aid  in  iden- 
tifying subspecies. 

Thirdly,  Hubbard  (1999)  comments  that 
“even  when  characteristics  of  populations  are 
better  known,  opinions  may  differ  as  regards 
their  taxonomic  treatment”  because  of  limited 
sample  sizes,  interbreeding  among  populations, 
and  differences  in  taxonomists’  methods, 
views,  and  findings.  Although  we  agree  that 
taxonomists  have  been  inconsistent  in  their 
treatment  of  subspecific  taxonomy,  we  consid- 
er this  to  be  an  incentive  for  finding  areas  of 
common  ground  among  researchers,  rather 
than  a justification  for  concluding  that  reliable 
identification  of  subspecies  is  impossible. 
Hubbard  states  that  we  should  have  strictly 
adhered  to  a single  view  of  subspecies  tax- 
onomy. We  followed  a single  view  of  subspe- 
cies treatment,  but  we  did  not  credit  this  single 
view  to  a single  researcher.  We  made  it  clear 
that  we  adopted  the  “four  subspecies  classi- 
fication system  of  Hubbard  (1987)  and  Unitt 
(1987).”  We  warned  readers  in  our  Methods 
section  that:  “Given  morphological  overlap 
and  hybridization  among  subspecies,  complete 
accuracy  in  identifying  subspecies  is  not 


achievable.”  Although  taxonomists  disagree 
in  their  interpretations  of  within-species  vari- 
ation and  subspecies  recognition,  there  is  un- 
mistakable agreement  about  use  of  a four  sub- 
species classification  among  recent  research 
papers  (Hubbard  1987,  Unitt  1987,  Browning 
1993).  Hubbard  (1987)  clearly  advocates  ac- 
ceptance of  the  four  subspecies  classification 
in  his  report  by  stating  that:  “Given  the  degree 
of  agreement  among  recent  workers,  I believe 
the  most  prudent  course  is  to  accept  all  of  the 
above  subspecies  [i.e.,  E.  t.  extimus,  brewsteri, 
and  adastus^  and  traillii  as  valid — at  least  un- 
til more  definitive  studies  are  available.”  Al- 
though in  his  commentary  Hubbard  declares 
his  own  report  to  be  a “cobbling  job”,  its 
quality  is  deemed  sound  by  other  authorities. 
Indeed,  it  has  been  widely  distributed  and  cit- 
ed both  unofficially  and  officially  by  the  En- 
dangered Species  Programs  of  U.S.  Fish  and 
Wildlife  Service  regions,  by  state  Game  and 
Fish  Departments,  and  by  other  agencies  and 
ornithologists  in  the  western  United  States,  es- 
pecially in  the  Southwest.  Given  Hubbard’s 
background  as  a competent  taxonomist  in 
New  Mexico  and  as  an  officer  of  the  state  en- 
dangered species  branch,  his  paper  is  judged 
as  an  authoritative  source  on  the  species.  For 
example,  in  the  process  use.d  for  listing  the 
southwestern  Willow  Flycatcher  as  a federally 
endangered  subspecies,  Hubbard’s  paper  was 
one  of  the  most  heavily  cited  reports  by  the 
U.S.  Fish  and  Wildlife  Service  (1995). 

Unitt  (1987)  also  states  that  the  four  races 
of  E.  traillii  are  valid  and  may  be  distin- 
guished from  each  other  by  “color,  wing  for- 
mula, or  both”.  Browning  (1993)  further  sep- 
arated subspecies  E.  t.  traillii  into  two  popu- 
lations: E.  t.  campestris  of  the  Great  Plains 
and  Great  Lakes  regions,  and  E.  t.  traillii  to 
the  southeast  of  E.  t.  campestris.  We  recently 
became  aware,  that  Unitt  has  conducted  fur- 
ther research  on  the  same  specimens  and  may 
soon  be  updating  his  taxonomic  treatment  (P. 
Unitt,  pers.  com.  through  J.  E.  Cartron).  These 
different  authors  describe  subspecies  distri- 
butions that  are  very  similar  although  popu- 
lation boundaries  are  not  exactly  the  same. 
U.S.  Fish  and  Wildlife  Service  relied  partly 
on  these  studies  to  conclude  that  listing  the 
southwestern  Willow  Flycatcher  as  an  endan- 
gered subspecies  was  appropriate. 

Fourthly,  Hubbard  evaluates  our  results  by 


Y(>nf>  and  Finch  • RESPONSE 


591 


comparing  our  subspecies  composition  data 
with  subspecies  data  from  his  own  and  other 
reports  and  sources.  While  such  comparisons 
may  be  valid  for  the  purpose  of  exploring  po- 
tential sources  of  viuriation,  the  conclusions 
that  Hubbard  draws  are  incorrect  because  of 
spatial  and  temporal  differences  among  stud- 
ies. Species,  subspecies,  and  population  com- 
position of  migratory  birds  captured  at  spe- 
cific stopover  sites  in  fall  or  spring  can  dra- 
matically differ  from  what  is  observed  at  the 
same  location  during  the  breeding  season  at 
the  same  location  or  from  other  locations  dur- 
ing migration.  For  example,  the  overall  spe- 
cies composition  we  detected  indicated  that 
the  majority  of  individuals  captured  were  not 
local  breeders  and  many  did  not  even  breed 
in  New  Mexico  (Finch  and  Yong  1999).  While 
we  used  a standardized,  systematic  procedure 
to  sample  throughout  the  entire  migration  sea- 
sons of  spring  and  fall,  1994  and  1995,  other 
studies  that  Hubbard  (1999)  cites  and  com- 
pares to  ours  were  not  conducted  during  mi- 
gration seasons  and/or  did  not  use  standard- 
ized procedures.  In  addition,  source  studies 
cited  by  Hubbard  are  heterogeneous  in  rela- 
tion to  study  goals,  year  of  study,  number  of 
years,  geographical  location,  sampling  design, 
sampling  season,  and  quality  of  data,  leading 
to  uncontrolled  and  unknown  factors  that  in- 
validate comparisons  with  our  data  set.  Our 
data  are  restricted  to  two  sites  during  two 
years  in  the  middle  Rio  Grande  valley  of  New 
Mexico,  and  thus  are  only  truly  comparable  to 
other  data  from  the  same  vicinity,  year,  and 
sampling  design.  Given  that  different  studies, 
especially  earlier  ones,  used  controversial  cri- 
teria for  classifying  and  counting  their  speci- 
mens, Hubbard’s  argument  that  our  results  are 
inaccurate  because  they  are  not  completely 
consistent  with  other  studies  that,  when  com- 
pared, also  yielded  dissimilar  results  is  circu- 
lar. In  our  manuscript,  we  did  not  make  such 
comparisons  for  at  least  two  reasons:  (1)  our 
research  focus  was  on  the  stopover  biology  of 
the  species,  not  on  the  taxonomic  status  of  the 
subspecies,  and  (2)  other  data  sources  were 
not  homogeneous  or  similar  enough  to  draw 
comparisons. 

Our  data  and  conclusions  about  the  fly- 
catcher’s stopover  ecology  are  not  dependent 
on  the  validity  or  accuracy  of  its  subspecies 
status  or  on  the  methods  used  to  identify  sub- 


species. Because  E.  t.  extimus  is  endangered, 
U.S.  Fish  and  Wildlife  permits  for  collecting 
voucher  specimens  during  migration  are  not 
issued  in  the  Southwest,  eliminating  the  pos- 
sibility of  having  an  alpha-taxonomist  identify 
locally  caught  specimens  to  subspecies  for  the 
purpose  of  setting  standards.  Because  most 
current  research  studies  and  conservation  ef- 
forts pertaining  to  the  Willow  Flycatcher  have 
focused  on  its  breeding  grounds,  the  impor- 
tance of  our  research  centers  on  when,  where, 
and  how  migration  stopover  sites  in  riparian 
woodlands  along  the  middle  Rio  Grande  are 
used  for  resting  and  fat  depositions  by  the  spe- 
cies. Without  understanding  the  migration 
strategy  of  the  species  and  without  justifying 
efforts  to  conserve  the  stopover  habitat  that 
the  species  uses,  the  Willow  Flycatcher’s  fate 
in  the  Southwest  will  be  jeopardized  regard- 
less of  how  perfect  or  imperfect  our  ability  in 
identifying  subspecies  is. 

Throughout  ornithological  history,  subspe- 
cies classification  and  identification  have  tra- 
ditionally been  a “problematic”  area,  partic- 
ularly within  the  genus  Empidonax.  Uncer- 
tainties about  subspecies  or  even  species  sta- 
tus do  not  negate  the  value  of  our  migration 
research  or  refute  our  results  about  Willow 
Flycatcher  stopover  ecology  or  intraspecific 
variation  in  migration  patterns.  We  assert  that 
increased  knowledge  of  the  stopover  behavior 
and  energetic  condition  of  the  Willow  Fly- 
catcher is  important  for  understanding  the  bi- 
ology of  the  species  as  a whole  and  that  in- 
formation about  within-species  variation  is 
valuable  in  conserving  the  endangered  south- 
western subspecies. 

Our  paper  and  Hubbard’s  (1999)  critique 
have  opened  up  the  opportunity  to  develop 
and  expand  discussion  and  evaluation  of  the 
different  subspecies,  the  subspecies  concept  as 
a whole,  and  whether  subspecies  should  be 
recognized  for  the  Willow  Flycatcher  given 
the  disagreement  about  their  identification  and 
the  difficulty  in  identifying  birds  in  hand.  We 
invite  and  challenge  others  to  contribute  ideas 
and  knowledge  to  this  controversy  in  the  hope 
that  new  or  better  techniques  for  identifying 
willow  flycatcher  subspecies  may  result.  Such 
discussion  or  results  would  certify  beyond  a 
doubt  the  worthwhile  contribution  of  our  pa- 
per. Subjecting  any  paper  to  a critical  com- 
mentary, however,  automatically  attracts  the 


592 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


notice  of  additional  readers.  We  are  pleased 
with  the  extra  attention  in  the  hope  that  further 
research,  understanding,  and  conservation  ef- 
forts will  be  directed  toward  the  endangered 
southwestern  Willow  Flycatcher  and  its  dis- 
appearing habitat. 

ACKNOWLEDGMENT 

We  thank  R.  Banks  for  his  review  of  the  manuscript. 

LITERATURE  CITED 

Browning,  M.  R.  1993.  Comments  on  the  taxonomy 
of  Empidonax  traillii  (Willow  Elycatcher).  West. 
Birds  24:241-257. 

Einch,  D.  M.  and  W.  Yong.  1999.  Landbird  migration 
in  riparian  habitats  of  the  middle  Rio  Grande. 
Stud.  Avian  Biol.  In  press. 

Hubbard,  J.  R 1987.  The  status  of  the  Willow  Fly- 


catcher in  New  Mexico.  New  Mexico  Dept,  of 
Game  and  Fish,  Endangered  Species  Program, 
Santa  Fe. 

Hubbard,  J.  P.  1999.  A critique  of  Wang  Yong  and 
Finch’s  Field-identification  of  Willow  Flycatcher 
subspecies  in  New  Mexico.  Wilson  Bull.  Ill: 
585-588. 

Munsell  Color  Charts.  1990.  Munsell  Color  Com- 
pany, Inc,  Baltimore,  Maryland. 

Smithe,  F.  B.  1975.  Naturalist’s  color  guide.  American 
Museum  of  Natural  History,  New  York. 

Unitt,  P.  1987.  Empidonax  traillii  extiinu.s:  an  endan- 
gered subspecies.  West.  Birds  18:137-162. 

U.S.  Fish  and  Wildlife  Service.  1995.  Final  mle  de- 
termining endangered  status  for  the  Southwestern 
Willow  Flycatcher.  Federal  Register  60:10694— 
10715. 

Yong,  W.  and  D.  M.  Finch.  1997.  Migration  of  the 
Willow  Flycatcher  along  the  middle  Rio  Grande. 
Wilson  Bulletin  109:253-268. 


Wilson  Bull.,  1 1 1(4),  19W,  pp.  593-599 


Ornithological  Literature 

Edited  by  William  E.  Davis,  Jr. 


THE  AUKS.  By  Anthony  J.  Gaston  and 
Ian  Jones,  illustrated  by  Ian  Lewington,  line 
drawings  by  Ian  Lewinton  and  Ian  Jones.  Ox- 
ford University  Press,  New  York,  New  York. 
1998:  XX  plus  349  pp.,  8 color  plates  with 
caption  figs.,  41  figures,  32  maps,  43  tables 
29  photographs.  ISBN  0-19-8540320-9.  $75 
(Cloth). — This  book  summarizes  the  biology 
of  these  northern  seabirds.  The  authors  have 
spent  much  time  studying  many  species  of 
auks  and  are  well  prepared  to  summarize  the 
biology  of  these  northern  seabirds.  As  the  au- 
thors point  out,  these  birds  are  primarily  ma- 
rine organisms  but  most  studies  have  been 
conducted  at  their  breeding  sites  on  land. 

The  book  includes  three  sections:  Plan  of 
the  book.  General  chapters,  and  Species  ac- 
counts. The  Plan  of  the  book  is  comparable  to 
the  introduction  in  most  books:  it  explains  the 
intent  of  the  authors  and  the  layout  of  the 
book.  This  section  includes  an  important  table 
presenting  common  nomenclature  between 
Europe  and  North  America.  The  plates  of  the 
species  are  excellent.  Literature  coverage  is 
extensive.  The  book  is  written  in  British  En- 
glish rather  than  American  English. 

The  General  chapters  include  7 chapters: 
Auks  and  their  world,  Systematics  and  evo- 
lution, Distribution  and  biogeography.  Auks 
and  ecosystems.  Social  behavior.  Chick  de- 
velopment and  the  transition  from  land  to 
sea,  and  Populations  and  conservation. 
These  chapters  include  the  authors’  under- 
standing of  the  family,  comparative  analyses, 
and  syntheses.  These  are  worth  the  price  of 
the  book.  These  chapters  are  strong  in  that 
they  deal  with  auks  at  sea  as  well  as  the 
breeding  on  land.  Sections  in  these  chapters 
are  usually  one  to  a few  pages  in  length.  As 
summaries,  these  cover  the  material  ade- 
quately but  each  could  be  expanded  into  a 
more  thorough  monograph.  Our  understand- 
ing of  these  birds  has  developed  slowly  be- 
cause these  are  marine  birds  breeding  in  re- 
mote areas.  These  summary  chapters  include 
an  important  historical  perspective  when 


covery  of  auks  and  systematics.  These  sec- 
tions allow  the  reader  to  put  the  literature 
into  an  historical  perspective.  We  are  only 
now  increasingly  understanding  their  marine 
biology,  while  often  information  in  the  lit- 
erature is  based  on  early  articles  based  on 
early  assumptions. 

The  authors  discuss  the  biology  of  these 
birds  at  sea  and  during  breeding  at  land. 
Some  comparisons  are  among  most  auks  and 
related  species;  some  comparisons  are  among 
select  auk  species.  These  comparisons  are 
important  in  keeping  these  birds  in  perspec- 
tive with  other  seabirds,  and  for  understand- 
ing how  they  are  adapted  to  their  marine  en- 
vironment. The  chapter  on  populations  and 
conservation  is  very  timely.  It  includes  sec- 
tions on  changes  in  populations,  species  of 
concern,  as  well  as  fisheries  impacts.  While 
this  is  a good  review,  it  should  have  included 
more  timely  information  on  immediate  is- 
sues. The  Species  accounts  are  well  written. 
They  include  sections  on  Description,  Range 
and  status.  Habitat,  food  and  feeding  behav- 
ior, Displays  and  breeding  behaviour.  Breed- 
ing and  life  cycle,  and  Population  dynamics. 
It  would  have  been  nice  to  include  a sum- 
mary of  the  conservation  status. 

This  is  a fine  book  with  a great  deal  of  in- 
formation. However,  confusion  may  result 
from  inconsistencies  among,  and  in  some  cas- 
es poor  organization  of,  tables,  figures,  and 
maps.  Caption  formats  are  different  between 
the  General  chapters  and  the  Species  accounts. 
Timing  of  activity  at  colonies  is  dealt  with 
inconsistently  among  the  species  accounts  but 
is  summarized  in  a table  in  the  chapter  on  So- 
cial behaviour  under  the  confusing  heading  of 
Activity  timing.  Furthermore,  in  the  tables 
that  include  measurements  in  the  Species  ac- 
counts, data  are  summarized  according  to  ref- 
erence numbers  with  no  indication  as  to  where 
to  find  the  references,  leaving  it  to  the  reader 
to  guess  where  in  the  text  these  reference 
numbers  are  indicated.  Maps  in  the  first  seven 
chapters  have  a figure  number  (without  men- 
tion of  Fig.);  among  the  species  accounts. 


dealing  with  auks  and  people,  scientific  dis- 

593 


594 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


maps  do  not  contain  a figure  number  or  a title, 
leaving  it  to  the  reader  to  assume  that  these 
are  for  the  species  accounts  in  which  they  are 
included.  Finally,  among  range  maps,  the 
summer  range  coloration  is  often  difficult  to 
distinguish  from  the  background  map  colora- 
tion. These  problems  will  lead  to  confusion 
and  detract  from  an  otherwise  fine  and  infor- 
mative book. 

The  seven  summary  chapters  present  the 
author’s  family  analyses  and  strongly  based 
understanding  of  these  birds.  This  information 
is  well  worth  the  book.  The  species  accounts 
and  difficulty  in  extracting  information  are 
disappointing. — MALCOLM  C.  COULTER. 


THE  ECOTRAVELLERS’  WILDLIFE 
GUIDE  TO  COSTA  RICA.  By  Les  Beletsky. 
Academic  Press,  San  Diego,  California.  1998: 
426  pp.,  80  color  plates,  19  habitat  photos. 
$27.95. 


THE  ECOTRAVELLERS’  WILDLIFE 
GUIDE  TO  BELIZE  AND  NORTHERN 
GUATEMALA.  By  Les  Beletsky.  Academic 
Press,  San  Diego,  California,  1998:  488  pp. 
104  color  plates,  20  habitat  photos.  $27.95. 


THE  ECOTRAVELLERS’  WILDLIFE 
GUIDE  TO  TROPICAL  MEXICO.  By  Les 
Beletsky,  Academic  Press,  San  Diego,  Cali- 
fornia, 1999:  497  pp.  104  color  plates,  21 
habitat  photos.  $27.95. — Each  of  these  vol- 
umes is  part  of  a series  recently  released  un- 
der the  sponsorship  of  the  Wildlife  Conser- 
vation Society  and  authored  by  Les  Beletsky. 
As  all  of  the  volumes  are  quite  similar  they 
are  reviewed  together.  The  books  are  intend- 
ed as  ecological  introductions.  There  is  a 
brief  opening  section  about  ecotourism  that 
is  essentially  identical  in  each  of  the  books, 
as  is  the  structure  of  each  book.  There  is  an 
ecological  overview  of  the  country  followed 
by  chapters  on  how  to  use  the  book  and  on 
conservation  issues.  The  major  parts  of  each 
book  are  chapters  on  amphibians,  reptiles, 
birds,  and  mammals,  and,  in  the  Belize  book, 
on  marine  life.  Unfortunately  terrestrial  in- 


vertebrates are  totally  ignored.  There  is  noth- 
ing about  butterflies  or  other  prominent  in- 
verts in  either  book.  It  is  disappointing  to 
look  in  the  index  and  see  reference  to  ant- 
birds,  anteaters,  antshrikes,  ant-tanagers, 
antthrushes,  but  not  to  ants! 

Birds  are  featured  prominently  in  these 
books,  the  bird  chapter  being  89  pages  long 
in  Costa  Rica,  80  pages  long  in  Belize,  83 
pages  long  in  Tropical  Mexico  (with  50  color 
plates  on  birds  in  each  book).  Much  of  the  text 
and  plates  is  duplicated  among  the  three  vol- 
umes. Chapters  on  taxa  discuss  natural  histo- 
ry, breeding,  ecological  interactions,  lore  and 
notes,  and  conservation  issues.  “Profiled” 
species  are  illustrated  on  the  plates.  Illustra- 
tions are  large  format  and  are  confined  to  (ar- 
guably) the  most  common  species.  For  most 
readers  of  The  Wilson  Bulletm  these  guides 
will  not  be  useful  in  field  identification  of 
birds  because  too  many  species  are  omitted. 
However,  these  books  will  help  bird  watchers 
and  ornithologists  to  identify  amphibians  and 
reptiles,  many  of  which  are  not  illustrated  in 
convenient  field-sized  books.  As  with  birds, 
better  and  more  complete  guides  exist  for 
mammals  and  fish.  It  is  regretable  that  Belet- 
sky selected  tropical  Mexico,  much  of  which 
duplicates  what  he  includes  in  his  Belize 
book,  rather  than  western  Mexico,  where  there 
are  major  avifaunal  as  well  as  other  taxonomic 
distinctions. 

Beletsky’s  text  draws  heavily  from  the 
published  literature  and  does  a credible  job 
of  presenting  sound  overviews  of  the  subject 
matter.  The  text  ranges  widely,  from  sum- 
maries of  why  some  birds  have  evolved  to 
become  frugivores  to  Mayan  legends  about 
how  hummingbirds  became  so  bright.  Much 
information,  though  charming  in  a way,  is  su- 
perfluous to  a utilitarian  field  guide.  The  lore 
and  notes  sections  abound  with  such  quaint 
insights  as  the  belief  that  cows  belonging  to 
a farmer  who  has  destroyed  a swallow’s  nest 
will  give  bloody  milk.  We  are  told  (in  each 
volume)  that  the  Common  Raven  (Corvus 
corax)  is  the  largest  passerine,  though  the  Au- 
thor readily  acknowledges  that  they  do  not 
occur  in  Costa  Rica,  Belize,  northern  Gua- 
temala, or  the  Mexican  Yucatan.  I always 
thought  the  Superb  Lyrebird  {Menura  novae- 
hollandiae)  was  actually  the  largest  passerine 
(neither  does  it  occur  in  the  Neotropics). 


ORINTHOLOGICAL  LITERATURE 


595 


Nonetheless,  visitors  to  the  countries  profiled 
should  find  the  appropriate  book  useful  in 
gaining  a better  understanding  (and  a fair 
dose  of  trivia)  about  the  local  vertebrate 
wildlife.— JOHN  KRICHER. 


A BIRD-FINDING  GUIDE  TO  MEXICO. 
By  Steve  N.  G.  Howell,  illus.  by  Sophie 
Webb.  Cornell  University  Press,  Ithaca,  New 
York.  1999;  365  pp.,  54  locality  maps.  $20 
softcover. — This  is  an  exciting  addition  to  the 
libraries  of  couch-birders  and  those  eager  to 
bird  or  investigate  new  corners  of  Mexico. 
This  compact  and  well-organized  guide  is  6" 
X 9"  and  appears  to  be  solidly  bound.  Steve, 
Sophie,  and  their  occasional  fellow  travelers 
have  spent  hundreds  of  days  and  nights  in  fine 
resorts,  dives,  and  camping  sites  to  gather  lo- 
cality lists  for  III  sites  from  Baja  to  the  Yu- 
catan. Every  locality  map  is  extremely  useful 
to  any  visitor,  although  indicated  habitat  may 
be  altered  in  time. 

This  is  the  first  detailed  information  on 
many  important  sites  in  Mexico,  many  of 
which  scream  for  recognition  and  protection. 
Very  few  sites  are  protected  in  Mexico.  Lo- 
cality descriptions  are  well  written,  to  the 
point,  and  include  attention  to  directions  with- 
in a tenth  of  a kilometer.  Comments  on  avail- 
ability of  second-class  buses  and  camping 
sites  are  included  for  those  without  their  own 
(or  rental)  vehicles  and  those  on  limited  bud- 
gets. The  species  lists  are  English  common 
names  only,  and  wisely  run  in  4 columns  per 
page  (very  economical).  Relative  abundances 
are  not  given,  except  for  those  rarely  seen. 
Mexican  endemics  are  bolded  in  the  text  and 
lists. 

A variety  of  taxonomic  decisions  and  En- 
glish name  selections  differ  from  Seventh 
Edition  of  the  A.O.U.  Check-List  of  North 
American  Birds  (1998).  Most  involve 
“splits”  as  both  the  world  bird  species  list 
and  the  Dow  Jones  Average  struggle  to  pass 
10,000.  Well-differentiated  subspecies  clus- 
ters are  given  distinct  English  names  in  an- 
ticipation of  these  forms  being  given  full  spe- 
cies status  in  the  future.  Numbers  gathering 
bird  listers  pressure  for  all  splits,  boo  all 
lumps.  One  would  hope  to  see  less  attention 
by  birders  focused  on  one  unit  of  taxonomy 


and  more  attention  to  higher  (genus)  and 
lower  (subspecies)  levels  in  the  future.  As  for 
the  names  selected,  it’s  great  to  see  the  use 
of  whitestart  replacing  redstart  for  members 
of  the  genus  Myioborus.  Redstart  was  created 
for  Old  World  thrushes  with  red  on  the  tail, 
and  is  erroneously  used  for  an  American  pa- 
rulid  which  could  easily  be  renamed  Oran- 
gestart,  Setophaga  ruticilla.  There  is  no  red 
in  the  tail  of  any  Myioborus,  most  have  no 
red  anywhere,  and  they  are  no  longer  placed 
adjacent  to  Setophagal  Guy  Tudor  and  1 
came  up  with  whitestart  back  in  the  late 
1970s  as  a solution.  This  British  author  has 
sold  his  publisher  on  using  grey  in  place  of 
the  American  gray  throughout  in  contradic- 
tion of  the  1998  A.O.U.  Check-List.  I am 
concerned  with  using  subspecific  modifiers  in 
front  of  species  names;  it  can  create  much 
confusion.  Perhaps  it  should  be  Sooty  race  of 
Fox  Sparrow  or  Fox  Sparrow  (Sooty  race)  for 
Passerella  iliaca  unaleschensis,  not  Sooty 
Fox  Sparrow. 

This  book  is  not  designed  for  “lite  birders” 
doing  cruises  of  the  Mexican  Riviera,  the 
whale  lagoons,  or  the  islands  of  the  Gulf  of 
California,  nor  for  those  doing  single  desti- 
nation beach  resort  vacations.  It  will  greatly 
aid  self  sufficient,  street-smart  birders  and  or- 
nithologists with  a taste  for  adventure  and 
great  birding.  Great  job  Steve! — PETER  AL- 
DEN. 


THE  BIRDS  OF  SONORA.  By  Stephen  M. 
Russell  and  Gale  Monson,  illus.  By  Ray 
Harm.  The  University  of  Arizona  Press,  Tuc- 
son, Arizona.  1998:  360  pp.,  2 color  plates,  34 
b+w  figures.  ISBN  0-8165-1635-9.  $75  hard- 
cover.— A long  awaited  book  authored  by  two 
excellent  fieldmen  who  have  lived  just  across 
the  border  in  or  near  Tucson  for  years.  This 
is  a fairly  heavy  book  {^Vi"  X 11")  focused  on 
the  ranges,  habitats,  seasonal  abundances,  his- 
torical records,  and  current  status  of  over  500 
species  of  birds.  It  accomplishes  its  tasks  well, 
especially  with  the  well-researched  range 
maps  for  most  species.  This  work  is  the  first 
update  in  many  decades  for  a province  that 
should  attract  many  more  birders  from  the 
southwestern  states.  Sonora  has  tropical  de- 
ciduous forest  around  the  colonial  hilltown  of 


596 


THE  WILSON  BULLETIN  • Vol.  HI,  No.  4.  December  1999 


Alamos,  great  pine-oak  woodland  in  the  Sierra 
Madre  Occidental,  and  cactus-scapes  of  the 
Sonoran  Desert. 

The  dust  jacket  features  a colorful  White- 
fronted  Parrot  {Arnazona  albifrons),  which  is 
not  reproduced  within  and  lost  to  users  of  li- 
braries that  routinely  toss  dust  jackets.  Nice  to 
see  the  separate  large  maps  of  mountain  rang- 
es, cities,  rivers  and  reservoirs,  and  the  full 
color  vegetation  map  in  the  introduction.  I 
question  the  wisdom  of  using  only  the  metric 
system  to  indicate  elevations  and  distances. 
Outside  of  scientific  circles  the  metric  system 
is  dying  in  the  U.S.,  why  obfuscate  the  ma- 
jority of  the  book’s  users?  Add  American 
equivalents  in  parentheses. 

The  geographical  coverage  excludes  the 
Sonoran  Islands  in  the  Gulf  of  California.  Isla 
Tiburon  and  other  islets  have  no  endemic 
birds  and  this  book  should  have  included  a 
summary  of  known  residents  and  visitors. 
While  habitat  loss  is  discussed  and  lamented, 
a rundown  of  any  protected  areas  and  a focus 
on  areas  most  in  need  of  protection  would 
have  been  welcome. 

The  appendices  cover  plants  named  in  the 
text,  an  exhaustive  gazetteer  useful  to  any  bi- 
ologist, and  literature  cited.  Adding  a Spanish 
common  name  throughout  the  text  and  includ- 
ing them  in  the  index  is  an  outstanding  step. 
One  hopes  this  will  be  made  available  in 
Spanish  for  Sonoran  citizens. 

Wouldn’t  it  be  nice  to  have  similar  books 
state  by  state  throughout  Mexico  being  re- 
searched and  published  in  an  orderly  fashion? 
This  would  be  a good  model. — PETER  AL- 
DEN. 


A FIELD  GUIDE  TO  THE  BIRDS  OF 
MEXICO  AND  ADJACENT  AREAS  (BE- 
LIZE, GUATEMALA,  AND  EL  SALVA- 
DOR), third  edition.  By  Ernest  Preston  Ed- 
wards, principal  illustrator  Edward  Murrell 
Butler.  University  of  Texas  Press,  Austin,  Tex- 
as. 1998:  284  pp.  incl  51  color  plates.  $35 
hardcover,  $17.95  softcover. — Ernest  Edwards 
has  been  a pioneer  in  producing  a series  of 
bird-finding  guides  and  compact  bird  field 
guides  to  Mexico.  This  third  edition  updates 
names  and  taxonomy,  adds  a few  plates,  and 
competes  with  the  Peterson’s  and  Chalif’s 


Field  Guide  to  Mexican  Birds  (Houghton  Mif- 
flin, Boston,  1973).  As  a portable  pocket 
guide  it  is  significantly  wider  than  the  Peter- 
son’s (fitting  fewer  pockets),  the  artwork  a bit 
stiff  and  stylized.  The  text  is  concise  and  the 
book  well  indexed  between  text  and  plates  and 
back.  The  sequence  of  families  and  species  in 
the  color  plates  is  jumbled  and  confusing. 
Neither  book  has  range  maps.  Gray  is  spelled 
gray  not  grey  in  Edwards. 

At  four  times  the  size  and  weight  of  the 
Edwards  or  Peterson/Chalif,  Steve  Howell’s 
and  Sophie  Webb’s  A Guide  to  the  Birds  of 
Mexico  and  Northern  Central  America  (New 
York,  Oxford  Univ.  Press,  1995)  is  the  clear 
choice  for  serious  students  of  Mexican  birdlife 
with  its  exhaustive  text,  superior  plates,  and 
excellent  range  maps.  However,  its  bulk  and 
weight  will  force  many  to  consider  leaving  it 
at  home  or  in  the  car  when  deciding  which  of 
the  two  portable  quick  reference  guides  to 
take  in  the  field.  The  Edwards  book  gives 
much  less  thought  to  taxonomic  changes  and 
English  name  modification  than  does  Steve 
Howell’s  works,  a mixed  blessing. 

The  time  has  come  to  stop  redoing  at- 
tempts to  cover  close  to  1100  species  of  birds 
from  such  disparate  places  in  4-6  countries. 
What’s  needed  is  field  guides  to  Pacific  slope 
birdlife.  Gulf  and  Caribbean  slope  birdlife, 
highland  birdlife,  etc.  No  matter  where  you 
are,  one’s  book  has  over  50%  of  its  species 
totally  inapplicable  to  wherever  you  are.  Far 
more  useful  would  be  a guide  to,  say,  just 
Yucatan,  Belize,  and  Caribbean  Guatemala’s 
birds.  That’s  one  area  that  can  support  such 
a smaller  geographical  focus  book. — PETER 
ALDEN. 


CHECKLIST  OF  THE  BIRDS  OF  EUR- 
ASIA. By  Ben  E King.  Ibis  Publishing,  Vista, 
California.  1997:  105  pp.  $19.95  (paper). — 
Ben  King  has  produced  a concise,  and  func- 
tional birder’s  checklist  to  the  contiguous  con- 
tinents of  Europe  and  Asia,  plus  their  atten- 
dant islands,  ranging  all  the  way  from  Iceland 
to  Novaya  Zemlya  to  Japan,  the  Philippines, 
the  Greater  Sundas,  and  Wallacea,  two-thirds 
of  the  Old  World,  in  fact.  For  purposes  of  dis- 
tributional coding.  King  has  partitioned  this 
vast  expanse  into  ten  regions:  Europe,  the 


ORINTHOLOGICAL  LITERATURE 


597 


Middle  East,  the  former  Soviet  Union,  Japan, 
the  Indian  Subcontinent,  China,  Southeast 
Asia,  the  Greater  Sundas,  the  Philippines,  and 
Wallacea.  The  list  follows  “generally”  that  of 
Peters’  world  list,  and  includes  3062  species. 
Taxonomy  is  reported  to  be  conservative,  al- 
though the  author  notes  that  he  has  adopted 
“some  new  ideas  from  the  literature  as  well 
as  unpublished  field  studies,  especially  where 
they  appear  to  corroborate  my  own  experi- 
ence.” The  checklist  indicates  regional  pres- 
ence or  absence,  but  provides  no  information 
on  status  (e.g.,  breeding,  migrant,  vagrant)  or 
abundance.  The  author  notes  that  introduced 
species  are,  in  many  cases  omitted.  Species 
endemic  to  a single  one  of  King’s  regions  are 
indicated  by  boldface. 

The  main  focus  of  King’s  six  page  intro- 
duction is  the  construction  and  clarity  of  En- 
glish names.  From  King’s  strongly  worded 
statement,  it  is  clear  that  a main  purpose  of 
this  checklist  was  to  provide  a standardized 
and  revised  set  of  English  names  for  the  birds 
of  Eurasia.  King  notes  that  names  that  are  elit- 
ist or  that  are  difficult  to  pronounce  need  to 
be  changed,  as  do  patronyms  memorializing 
westerners.  Conducting  such  a nomenclatural 
“cleansing”  must  be  a difficult  task,  indeed. 
And  it  is  interesting  to  compare  King’s  bird 
names  against  those  used  in  the  array  of  other 
checklists  and  field  guides  for  the  region.  For 
reasons  of  space  I will  focus  only  on  a few 
widespread  Asian  passerine  species  that  hap- 
pen to  also  inhabit  Wallacea,  an  area  familiar 
to  me.  For  this  I refer  to  the  English  names 
used  in  King  and  Dickinson’s  (1975)  Field 
Guide  to  the  Birds  of  South-East  Asia,  White 
and  Bruce’s  (1986)  Birds  of  Wallacea,  Inskipp 
et  al.’s  (1996)  Annotated  Checklist  of  the 
Birds  of  the  Oriental  Region,  and  Coates  and 
Bishop’s  (1997)  Guide  to  the  Birds  of  Wal- 
lacea. At  least  for  these  widespread  Asian 
species,  the  names  King  today  chooses  to  use 
are  all  quite  reasonable,  and,  in  fact,  the  no- 
menclature across  the  various  publications 
produced  over  a 24  year  span  is  surprisingly 
uniform. 

It  is  rare  that  one  can  obtain  consensus  on 
English  names,  even  when  working  in  com- 
mittee. I am  happy  to  report,  however,  that,  at 
least  by  comparison  of  16  widespread  song- 
bird species  that  served  as  my  sample  exhibit 
vary  little  variation  in  name  in  the  six  sampled 


texts.  Of  these  16,  only  4 were  represented  by 
more  than  one  name  (Red  Avadavat/Straw- 
berry  Waxbill,  Chestnut/Black-headed  Munia, 
Hair-crested/Spangled  Drongo,  and  Eurasian 
Tree-Sparrow/Tree  Sparrow).  One  of  the  birds 
with  two  names,  the  drongo,  is  in  fact  a geo- 
graphically variable  taxon  that  may  constitute 
more  than  a single  species.  Thus  it  is  evident 
that  even  for  Southeast  Asian  birds  there  is 
considerable  stability  and  uniformity  of  En- 
glish nomenclature.  Still,  given  that  King 
deals  with  3000  species,  there  will  be  plenty 
of  species  with  multiple  English  names  in  cur- 
rent use.  Thus,  one  of  the  disappointments  of 
the  King  checklist  is  that  alternative  names  are 
not  listed,  probably  because  of  space  limita- 
tions. 

The  checklist  is  completed  by  a compre- 
hensive index  of  English  and  scientific  names, 
which  lists  all  scientific  names  by  species 
(e.g.,  ‘deucophaeus,  Dicrurus'")  and  all  En- 
glish names  by  group-name  (e.g.,  “Drongo, 
Ashy”).  At  the  very  least  I would  have  pre- 
ferred to  see  listing  by  genus,  as  well  (e.g., 
""Dicrurus  leucophaeus").  That  complaint 
aside,  this  is  a compact,  well-produced,  and 
useful  checklist  that  covers  a huge  avifauna. 
This  will  be  a must  buy  for  many  world  bird- 
ers and  ornithogeographers. — BRUCE  M. 
BEEHLER. 


THE  HANDBOOK  OF  BIRD  IDENTIFI- 
CATION FOR  EUROPE  AND  THE  WEST- 
ERN PALEARCTIC.  By  Mark  Beaman  and 
Steve  Madge,  illus.  by  Hilary  Burn,  Martin 
Elliott,  Alan  Harris,  Peter  Hayman,  Laurel 
Tucker,  and  Dan  Zetterstrom.  Princeton  Uni- 
versity Press,  Princeton,  New  Jersey.  1998: 
868  pp.,  291  color  plates  with  captions,  77 
other  color  illustrations  dropped  in  the  text 
elsewhere,  625  color-coded  range  maps, 
$99.50  (cloth). — This  book,  originally  pub- 
lished in  Great  Britain  by  Christopher  Helm, 
is  a monumental  achievement.  It  covers  al- 
most 900  species  known  to  have  occurred  in 
the  Western  Palearctic,  the  area  defined  for  the 
9-volume  Oxford  University  Press  series  ed- 
ited by  S.  Cramp  et  al..  The  Birds  of  the  West- 
ern Palearctic  (1977-1994).  This  area  extends 
from  Franz  Joseph  Land  and  Novaya  Zemlya 
south  to  Kuwait,  west  to  the  southwest  comer 


598 


THE  WILSON  BULLETIN  • VoL  III,  No.  4.  December  1999 


of  Morocco,  and  north  to  Iceland,  Jan  Mayen, 
and  Spitsbergen,  including  the  Azores,  Ma- 
deira, and  Canary  Island  groups. 

All  species  are  illustrated  in  color,  but  that 
is  only  the  beginning.  Many  are  shown  in 
flight,  at  rest,  as  adults  and  juveniles,  from 
above  and  below,  and  in  several  racial  forms 
as  appropriate.  About  600  of  the  species  cov- 
ered breed  in  the  area  and  the  rest  occur  only 
seasonally  or  are  vagrants.  Although  a vagrant 
may  have  occurred  only  once  or  twice,  its 
plumages  are  covered  thoroughly,  usually  in 
an  illustration  dropped  in  the  text,  and  typi- 
cally with  several  images.  Vagrants  that  occur 
with  some  frequency  are  included  on  the 
plates  with  the  local  species.  To  provide  a per- 
spective on  the  breadth  of  coverage,  the  33 
plates  depicting  60  species  of  jaegers,  skuas, 
gulls,  terns,  and  alcids  contain  no  fewer  than 
480  images  and  include  at  least  12  species  that 
are  vagrants  from  North  America  or  Asia. 
Two  plates  and  29  images  are  devoted  to  the 
recently-split  Yellow-legged  (Lams  cachin- 
nans),  Heuglin’s  (L.  heuglini),  and  Armenian 
gulls  (L.  arrnenicus)  alone.  Swainson’s  Hawk 
(Buteo  swainsoni)  has  only  occurred  two  or 
three  times,  but  it  rates  eight  images  including 
both  pale  and  dark  morphs,  perched  and  in 
flight,  and  even  the  rare  rufous  variant.  The 
plates  are  supported  by  detailed  discussion  of 
identification  criteria  and  comparisons  be- 
tween similar  species  in  the  text,  including 
mention  of  racial  populations  where  current 
taxonomy  is  in  doubt  or  where  geographical 
variation  is  significant. 

The  same  lavish  treatment  is  given  every 
group,  including  sandpipers  and  plovers.  A 
high  percentage  of  the  shorebird  species 
known  for  eastern  North  America  is  depicted 
on  the  plates  right  next  to  the  most  similar 
Western  Palearctic  species.  Even  Eskimo  Cur- 
lew is  there  (Niimenius  horeals).  Indeed,  spe- 
cies such  as  Semipalmated  (Colidns  pusilla). 
Least  (C.  minytilla).  Western  (C.  mouri), 
Baird’s  (C.  hairdd),  White-rumped  (C.  fusci- 
collis).  Stilt  (C.  himantopu.s)  and  Pectoral  (C. 
mekmotos)  sandpipers  and  the  two  dowitchers 
are  treated  more  fully  than  in  all  but  special- 
ized shorebird  guides.  If  anything.  Western 
Palearctic  species  are  even  more  fully  treated. 
For  those  with  a virtual  field-identification 
death-wish,  there  are  two  plates  with  32  im- 
ages of  Phylloscopus  warblers  that  give  new 


meaning  to  the  notion  of  confusing  (and  vir- 
tually inseparable)  species,  including  five 
forms  of  Chiffchaff  (P.  collybita)  that  may  be 
made  separate  species  someday. 

Though  the  greatest  wealth  of  detail  is  re- 
served, appropriately,  for  local  species.  North 
American  birders  will  be  impressed  with  the 
description  details  provided  for  North  Amer- 
ican vagrants.  Six  species  of  thrushes,  22 
wood  warblers,  2 tanagers,  15  emberizids,  and 
5 icterids  are  covered  exhaustively,  for  ex- 
ample. In  addition  to  critical  field  marks,  var- 
iations associated  with  age  or  sex,  voice,  and 
preferred  habitat  in  each  species’  natural  range 
are  covered  in  detail,  just  the  same  as  for  local 
birds. 

The  taxonomy  is  relatively  conservative 
and  current  to  about  1995,  a significant 
achievement  itself  when  dealing  with  so  many 
species.  However,  the  authors  have  carefully 
called  attention  to  many  races  that  may  be  el- 
evated to  full  species  rank  in  the  future:  Taiga 
and  Tundra  Bean  Goose  (Anser  fabalis  ssp.), 
two  or  more  forms  of  Brant  (Branta  bernicla). 
Common  and  Black  scoters  (Melanitta  nigra 
ssp.).  Pharaoh  Eagle  Owl  (Bubo  bubo  asca- 
laphus),  two  or  more  Yellow  Wagtails  (Mo- 
tacilla  flava),  Moroccan  Wagtail  (Motacilla 
alba  subpersonata),  Sykes’  Warbler  (Hippo- 
lais  caligata  rama),  the  Chiffchaffs  mentioned 
above,  and  others. 

There  are  brief  but  excellent  introductory 
sections  outlining  the  content  of  the  species 
accounts,  defining  terms,  and  commenting  on 
the  techniques  and  pitfalls  of  field  identifica- 
tion. Full  indices  of  English  and  scientific 
names  are  provided.  There  are  also  appendices 
of  14  recent  additions  with  full  descriptions, 
another  of  50  species  intentionally  omitted, 
and  an  appendix  listing  other  important  dis- 
tribution and  identification  references  for  the 
area  covered.  As  is  inevitable  in  a publication 
of  such  size,  there  are  occasional  typographi- 
cal errors,  but  I noted  only  three  or  four  in 
studying  the  book  for  over  eight  hours.  Each 
of  the  illustrators  has  done  superlative  work, 
and  though  every  plate  can  be  called  “good”, 
in  this  reviewer’s  opinion  there  are  a few 
plates  that  are  less  successful  than  others  at 
capturing  a vibrant,  lifelike  quality  to  the  im- 
ages. 

It  is  easy  to  lapse  into  superlatives  after 
only  a short  acquaintance  with  this  book.  It 


ORINTHOLOGICAL  LITERATURE 


599 


weighs  about  five  pounds,  and  at  two  inches 
thick  it  is  too  large  and  heavy  to  fit  conve- 
niently in  the  pocket  of  a field  jacket.  But 
most  active  North  American  birders  will  want 
to  have  it,  despite  the  high  price,  even  if  they 


do  not  always  carry  it  with  them.  It  sets  a new 
high  standard  for  the  part  of  the  world  it  cov- 
ers, a standard  not  yet  met  or  even  closely 
approached  elsewhere.  Highly  recommend- 
ed.—ALLAN  R.  KEITH. 


Wilson  Bull.,  111(4),  1999,  pp.  600-607 


PROCEEDINGS  OF  THE  EIGHTIETH 
ANNUAL  MEETING 

JOHN  A.  SMALLWOOD,  SECRETARY 


The  eightieth  annual  meeting  of  the  Wilson  Orni- 
thological Society  was  held  Thursday,  10  June, 
through  Sunday,  13  June,  1999,  at  Colby  College  in 
Waterville,  Maine.  W.  Herbert  Wilson  chaired  the  local 
committee;  support  for  the  conference  was  provided 
by  the  Special  Programs  office  of  Colby  College. 

The  Council  met  from  13:07  to  18:27  on  Thursday, 
10  June,  and  again  from  15:35  to  16:30  on  Saturday, 

12  June,  in  Room  335,  Olin  Hall,  Colby  College.  On 
Thursday  evening  there  was  an  informal  reception  in 
Cotter  Union  for  conferees  and  their  guests. 

The  opening  session  on  Friday  convened  in  Room 
101  Keys  Hall,  being  called  to  order  at  08:37  by  WOS 
President  Edward  H.  Burtt,  Jr.  After  several  announce- 
ments were  made  by  Local  Chair  Herb  Wilson  and  by 
Scientific  Program  Chair  Ted  Davis,  President  Burtt 
introduced  Professor  Ed  Yeterian,  Dean  of  Faculty  at 
Colby  College,  who  welcomed  those  in  attendance. 
Welcoming  remarks  were  also  offered  by  Dr.  Miriam 
Bennett,  Professor  Emeritus,  Department  of  Biology, 
Colby  College.  Eollowing  these  opening  remarks. 
President  Burtt  introduced  the  third  annual  Margaret 
Morse  Nice  Plenary  Lecture,  “Intraspecific  variation 
in  the  sizes  and  shapes  of  birds,”  presented  by  Frances 
C.  James.  The  Nice  Lecture  was  followed  by  the  first 
business  meeting  of  the  Wilson  Ornithological  Society. 

The  scientific  program  included  45  presentations  or- 
ganized into  six  paper  sessions  and  one  session  for  the 

13  poster  presentations.  Approximately  half  of  all  pre- 
sentations were  by  students.  In  addition,  there  were 
two  workshops,  one  on  post-baccalaureate  careers  in 
wildlife  and  conservation,  presented  jointly  by  the  Wil- 
son Ornithological  Society  and  the  Ornithological 
Council,  and  the  other  on  bird  skinning,  conducted  by 
WOS  Librarian  Janet  Hinshaw. 

On  Friday  evening  the  conferees  enjoyed  a “down 
East”  lobster  bake  at  Johnson  Pond  on  the  Colby  Col- 
lege campus.  At  the  conclusion  of  this  meal,  the  sat- 
isfied decapodivores  reconvened  for  a social  at  the 
Cotter  Union.  Field  trips  on  Friday  and  Saturday  morn- 
ings included  excursions  to  Sidney  Bog  and  Colby  Ar- 
boretum. Several  longer  field  trips  were  scheduled  for 
Sunday,  including  birding  along  coastal  sites  in  south- 
ern Maine,  white-water  rafting  on  the  Kennebec  River, 
birding  in  the  greater  Waterville  area,  and  a pelagic 
trip  to  Matinicus  rock,  where  those  participating 
viewed  Atlantic  Puffins  (Fratercula  arctica).  Razor- 
bills (Alca  torcla).  Black  Guillemots  {Cepphus  grylle), 
and  several  species  of  procellariiformids. 

A social  hour  preceded  the  annual  banquet,  which 
was  held  in  Dana  Hall.  At  the  conclusion  of  a fine 
dinner  President  Burtt  briefly  addressed  the  conferees, 
commended  Herb  Wilson  and  the  Local  Committee  for 
a pleasant  conference  venue  and  Ted  Davis  and  the 


Scientific  Program  Committee  for  a successful  meet- 
ing, thanked  retiring  Members  of  Council  Peter  Fred- 
erick and  Danny  Ingold  for  their  service  to  the  Society, 
and  recognized  all  student  presenters  for  their  contri- 
butions. The  following  awards  were  presented: 

MARGARET  MORSE  NICE  MEDAL 

(for  the  WOS  plenary  lecture) 

Frances  C.  James,  “Intraspecific  variation  in  the  siz- 
es and  shapes  of  birds.” 

EDWARDS  PRIZE 

(for  the  best  major  article  in  volume  110  of 
The  Wilson  Bulletin) 

Sheila  Conant,  H.  Douglas  Pratt,  and  Robert  J.  Shal- 
lenberger,  1998,  “Reflections  on  a 1975  expedi- 
tion to  the  lost  world  of  the  Alaka’i  and  other 
notes  on  the  natural  history,  .semantics,  and  con- 
servation of  Kaua’i  birds.”  Wilson  Bull.  110:1- 
22. 

LOUIS  AGASSIZ  FUERTES  AWARD 

Kazuya  Naoki,  “Community  evolution  in  the  An- 
dean tanagers  of  the  genus  Tangara.'" 

PAUL  A.  STEWART  AWARDS 

Paul  M.  Brandy,  “Olive-sided  Flycatcher  habitat  use 
in  managed  landscapes.” 

Thomas  V.  Dietsch,  “Ecology  and  conservation  of 
Neotropical  birds  in  coffee  agro-ecosystems  of 
southern  Mexico.” 

Amanda  D.  Rodewald,  “Disturbance  in  forested 
landscape:  influence  of  type  and  magnitude  on 
forest  birds.” 

WILSON  ORNITHOLOGICAL  SOCIETY 
TRAVEL  AWARDS 

Thomas  V.  Dietsch,  “Relating  Neotropical  birds  and 
vegetative  structure  to  certification  criteria  for 
coffee  agrosystems  in  Chiappas,  Mexico.” 

Falk  Huettmann,  “Wintering  Razorbills,  Alca  torcla, 
and  auk  assemblages  in  the  lower  Bay  of  Fundy, 
Canada.  Results  from  two  winter  surveys  1997/ 
98  and  1998/99.” 

Rachael  Z.  Jennings,  “Spatial  and  temporal  Varia- 
tion in  the  distributions  of  Calypte  hummingbirds 
along  an  elevational  tran.sect.  Riverside  County, 
California.” 

Karl  E.  Miller,  “Nesting  .success  of  the  Great  Crest- 
ed Flycatcher  in  natural  nests  and  in  nest  boxes: 
predation  rates  increase  with  nest  box  age.” 

Kimberly  A.  Peters,  “Swain.son’s  Warbler  habitat 


600 


Annum,  rhport 


601 


selection  in  a managed  bottomland  hardwood  for- 
est in  South  Carolina.” 

ALEXANDER  WILSON  PRIZE 

(for  best  student  paper) 

Christopher  M.  Somers,  "Bird  depredation  of  grapes 
in  a Niagara  Vineyard;  do  predictable  trends  ex- 
ist?” 

Selection  committee  for  the  Nice  Medal-Edward  H. 
Biirtt.,  Jr.  (chair),  John  Kricher,  William  E.  Davis,  Jr., 
and  W.  Herbert  Wilson;  for  the  Edwards  Prize-Robert 
Beason  (chair),  Charles  Blem,  and  Doris  Watt;  for  the 
Fuertes  and  Stewart  Awards— Richard  Stiel  (chair), 
James  Sedgwick,  and  Kelli  Stone;  for  the  Wilson  Or- 
nithological Society  Travel  Award— William  E.  Davis, 
Jr.  (chair),  and  John  C.  Kricher;  and  for  the  Wilson 
Prize-William  E.  Davis,  Jr.  (chair),  Jon  Barlow,  Robert 
Beason,  Richard  Conner,  and  Sara  Morris. 

FIRST  BUSINESS  MEETING 

The  first  business  meeting  was  called  to  order  by 
President  Burtt  at  10:02  on  Friday,  1 1 June,  in  Room 
105  Keyes  Hall.  Secretary  Smallwood  then  presented 
to  those  who  had  gathered  a synopsis  of  Thursday’s 
council  meeting,  commenting  on  the  successful  efforts 
of  Herb  Wilson  and  the  local  committee  and  of  Second 
Vice-president  Ted  Davis  and  the  scientific  program 
committee.  He  reviewed  the  awards  offered  by  the  So- 
ciety, the  winners  to  be  announced  at  the  annual  ban- 
quet, and  reported  that  Council  had  enthusiastically  in- 
creased their  monetary  values:  the  prestigious  Louis 
Agassiz  Fuertes  Award  increased  to  $2500,  the  Mar- 
garet Morse  Nice  Award  increased  to  $1000,  and  the 
annual  funding  for  up  to  four  Paul  A.  Stewart  Awards 
was  increased  to  $2000.  Secretary  Smallwood  in- 
formed the  membership  that  to  alleviate  any  confusion 
over  two  separate  WOS  awards  made  in  honor  of  Mar- 
garet Mor.se  Nice,  namely  the  MMN  Award  for  ama- 
teur ornithologists  and  the  MMN  Medal  for  ornithol- 
ogists presenting  plenary  lectures  on  a lifetime  of  or- 
nithological research,  the  MMN  Award  would,  in  the 
spirit  of  the  contribution  of  amateurs  to  the  field  of 
ornithology,  henceforth  be  known  as  the  George  A. 
Hall  and  Harold  E Mayfield  Award.  Council  had  also 
increased  the  amount  of  funding  available  for  student 
travel  awards  to  $5000  annually.  Secretary  Smallwood 
announced  that  Bob  Beason  had  been  elected  to  a third 
year  as  Editor  of  The  Wilson  Bulletin,  noting  that  the 
journal  was  not  only  published  under  budget,  due 
mostly  to  format  changes  and  electronic  submissions, 
but  also  published  slightly  ahead  of  schedule.  The  sec- 
retary then  reviewed  the  Conservation  Committee  re- 
port on  reauthorization  of  the  Endangered  Species  Act. 
The  2000  meeting  will  be  held  27—30  April  at  Hotel 
Galvez  in  Galveston,  at  the  invitation  of  the  Houston 
Audubon  Society  and  the  Gulf  Coast  Bird  Observa- 
tory; Dwight  Peake  will  be  the  local  host.  The  2001 
meeting  will  be  held  jointly  with  the  Arkansas  Audu- 
bon Society,  3-6  May,  at  the  University  of  Arkansas 


Continuing  Education  Center  and  Fayetteville  Hilton; 
Doug  James  will  chair  the  local  committee. 

Although  this  information  was  not  available  at  the 
time  of  the  first  business  meeting,  here  the  Wilson  Or- 
nithological Society  honors  the  memory  of  WOS 
members  who  passed  away  since  the  1998  meeting: 
Robert  E.  Ball  (North  Canton,  OH),  Roger  M.  Evans 
(Winnipeg,  MB),  Richard  R.  Graber  (Golconda,  IL), 
Frances  Hamerstrom  (Plainfield,  WI),  Robert  R. 
Knickmeyer  (Hazelwood,  MO),  H.  Elliot  McClure 
(Camarillo,  CA),  Henri  Ouellet  (Hull,  QC),  Edward  F. 
Rivinus  (Upper  Marlboro,  MD),  and  Charles  G.  Sibley 
(Santa  Rosa,  CA). 

The  treasurer’s  report  was  then  presented  by  Doris 
Watt. 

Bob  Beason  offered  the  editor’s  report. 

President  Burtt  delivered  the  report  of  the  Nominat- 
ing Committee  for  Peter  Stettenheim,  Chair,  who  was 
not  able  to  attend;  other  committee  members  included 
Kenneth  Able,  Patricia  Gowaty,  and  Ellen  Ketterson. 
The  following  slate  of  candidates  was  offered:  Presi- 
dent, John  C.  Kricher;  First  Vice-President,  William  E. 
Davis,  Jr.;  Second  Vice-President,  Charles  R.  Blem; 
Secretary,  John  A.  Smallwood;  Treasurer,  Doris  J. 
Watt,  and  Members  of  Council  for  1999—2002,  Robert 
A.  Askins,  Gary  Ritchison,  Charles  E Thompson,  and 
Jeffrey  R.  Walters. 

The  first  business  meeting  was  adjourned  at  approx- 
imately 10:20 

SECOND  BUSINESS  MEETING 

The  second  business  meeting  was  called  to  order  by 
President  Burtt  at  13:31,  Saturday,  12  June,  in  Room 
105  Keys  Hall,  at  which  time  he  recalled  the  report  of 
the  nominating  committee  to  the  floor.  Calling  for  ad- 
ditional nominations  and  hearing  none.  President  Burtt 
accepted  a motion  from  Dick  Banks,  seconded  by  Sara 
Morris,  that  the  nominations  be  closed.  That  motion 
passed  unanimously.  Jerry  Jackson  moved  that  the 
slate  of  officers  (but  not  councilors)  be  passed  unani- 
mously by  acclimation.  After  a second  from  Sara  Mor- 
ris, it  was.  Because  there  were  four  candidates  for  three 
seats  on  the  Council  (two  Members  of  Council  for 
1999-2002,  and  one  to  complete  Charles  Blem's 
1997-2000  Council  term)  election  was  by  paper  ballot, 
with  the  following  result:  Members  of  Council  for 
1999-2002,  Robert  A.  Askins  and  Jeffrey  R.  Walters; 
Member  of  Council  for  1999-2000.  Charles  E Thomp- 
son. 

Dwight  Peake,  Local  Chair,  updated  those  assem- 
bled on  the  2000  meeting  in  Galveston.  After  a brief 
discussion  about  the  proposed  name  change  for  The 
Wilson  Bulletin,  and  remarks  on  the  electronic  publi- 
cation of  the  Bulletin  and  the  Ornithological  Newslet- 
ter, Sara  Morris  moved  for  adjournment.  Ted  Davis 
and  Peter  Frederick  seconded  the  motion,  and  the 
membership  indicated  with  the  appropriate  body  lan- 
guage that  the  motion  had  indeed  passed.  This  oc- 
curred shortly  before  14:00. 


602 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


REPORT  OF  THE  TREASURER 

1 July  1998  through  30  June  1999 

GENERAL  EUNDS 


Balance  Eorward ^ 

Receipts 

Regular  and  Sustaining  Memberships $ 29,688.00 

Student  Memberships $ 4 278.00 

Family  Memberships  $ 112.00 

Total  Dues j 

Subscriptions $ 24,005.00 

Newsletter $ 181.00 

Back  Issues $ 406.25 

Page  Charges $ 16,911.25 

Total  Income  from  Publications  $ 

Contributions:  Research  Fund  $ 291.00 

Van  Tyne  Library $ 560.00 

General  Endowment $ 30.00 

Life,  Patrons  (Gen  Endowment) $ 1,000.00 

Unrestricted $ 937.00 

Total  Contributions $ 

Royalties $ 1,127.70 

Interest  from  Checking  Account $ 490.52 

Interest  from  Endowments $ 24,573.16 

Dividends  from  Dreyfus  Acct  (reinvested)  $ 999.02 

Subtotal $ 

Miscellaneous  $ 

Net  OSNA  Adjustment $ 

TOTAL  RECEIPTS  $ 

Disbursements 
Bulletin  Publication 

June  1998  $ 17,191.67 

September  1998  $ 16,280.53 

December  1998  $ 18,239.61 

March  1999  $ 16,570.33 

Editor’s  expenses  $ 4,315.07 

Total  Publication  Costs  $ 

OSNA  Expenses  $ 5,921.00 

Officer’s  expenses  $ 253.95 

CPA  (tax  filing) $ 465.00 

Incorporation  fee  $ 5.00 

Editor’s  honorarium  $ 2,000.00 

Editor’s  travel  $ 500.00 

Library $ 1,000.00 

Miscellaneous  $ 81.80 

Total  Operating  Expenses  $ 

Organizational  Awards  $ 1,700.00 

Ornithological  Council  Contribution $ 1,000.00 

ABC  Dues $ 150.00 

Total  Philanthropies  $ 

TOTAL  DISBURSEMENTS $' 

Endinf’  Balance $ 

CASH  ACCOUNTS 

First  Source  Bank  Checking  1 July  99  $ 61,832.05 

Dreyfus  Liquid  Assets  I July  99  $ 22,662. 16 

Total  Cash  on  Hand  $ 

Van  Tyne  Library  Accounts 

Starting  Balance $ 2,531.63 


64,264.33 


34,078.00 


41,503.50 


2,818.00 


27,190.40 

994.35 

(680.41) 

105,903.84 


72,597.21 


10,226.75 


2,850.00 

85,673.96 

84,494.21 


84,494.21 


Annual  rkport 


603 


Receipts  . . . 
Expenses  . . 
Ending  Balance 


SUTTON  DESIGNATED  ACCOUNT 


$ 8,788.79 
$ (3,046.13) 
$ 


Endowment  Principal  as  of  1 July  98 $ 66,956.97 

Increase  in  Principal  Value  98-99*  $ 68,898.72 

1998  Earnings*  $ 1,998.06 

Funds  disbursed  for  Color  Plates  (2)  $ 1,935.00 

Balance $ 


8,274.29 


63.06 


TOTAE  ENDOWMENT  FUNDS 


1996  Market  Value $ 676,029.00 

1997  Market  Value $ 814,315.00 

1998  Market  Value  (1  July  98)  $ 968,838.00 

1999  Market  Value  (1  July  99)  $ 1,096,733.81 


*Based  on  2.9%  yield  on  Mellon  account 

EDITOR’S  REPORT— 1998 

During  1998,  the  editorial  office  of  The  Wilson  Bul- 
letin received  245  manuscripts  including  6 that  were 
provided  by  Jed  Burtt  for  the  Symposium  on  Neotrop- 
ical Ornithology.  This  is  a continued  increase  in  the 
number  of  manuscripts  submitted  over  the  past  several 
years.  The  acceptance  rate  was  40%  for  these  manu- 
scripts, similar  to  that  of  the  past.  The  review  process 
took  an  average  of  13  weeks  and  most  of  manuscripts 
were  returned  to  the  authors  with  reviewers’  comments 
2-4  months  after  receipt.  We  have  been  using  E-mail 
for  much  of  the  correspondence  with  authors  and  re- 
viewers. This  has  resulted  not  only  in  monetary  sav- 
ings from  postage,  but  also  has  resulted  in  faster  turn 
around  for  manuscripts;  in  some  cases  the  author  re- 
ceived a decision  within  a month  of  submission. 

As  a result  of  the  new  format  and  page  size  of  The 
Wilson  Bulletin,  the  journal  has  gone  from  approxi- 
mately 800  pages  per  volume  to  600  pages.  The  cur- 
rent volume  contains  the  same  number  of  manuscript 
pages,  figures,  and  tables  as  previous  issues.  Most  of 
the  difference  in  length  is  because  many  of  the  figures 
and  tables  require  less  space  with  the  2-column  format 
than  they  did  with  the  single  column  format.  The  use 
of  electronic  manuscript  (disk)  submission  has  resulted 
in  fewer  corrections  in  the  galley  stage.  Three  issues 
in  1998  contained  color  frontispieces  and  3 of  the  4 
issues  that  are  out  or  in  the  works  for  1999  contain 
color  frontispieces.  We  have  had  some  problems  with 
the  printer  to  get  them  to  make  the  frontispieces  to 
bleed  to  the  edge  of  the  paper,  but  I think  those  have 
been  resolved. 

I greatly  appreciate  the  assistance  of  the  editorial 
board,  Clait  Braun,  Richard  Conner,  and  Kathy  G.  Beal 
for  their  timely  advise  on  many  manuscripts,  especially 
Kathy  for  her  advice  on  manuscripts  with  difficult  sta- 
tistical problems  and  for  doing  the  index.  Editorial  as- 
sistants Tara  Baideme,  Melanine  Daniels,  John  Lamar, 
Dante  Thomas,  and  Doris  Watt  assisted  in  tracking  and 
checking  the  many  manuscripts.  The  State  University 


Doris  J.  Watt,  Treasurer 

of  New  York  at  Geneseo  and  the  Biology  Department 
continue  to  support  the  editor  and  the  running  of  the 
editorial  office  in  many  ways. 

Robert  C.  Beason,  Editor 
The  reports  of  the  standing  committees  are  as  follows; 

REPORT  OF  THE  MEMBERSHIP 
COMMITTEE 

In  October  1998,  Laurie  Goodrich  took  over  as 
membership  chair  from  John  Smallwood,  WOS  Sec- 
retary. Current  members  of  the  WOS  membership 
committee  are  Laurie  Goodrich,  Chair,  Hawk  Moun- 
tain (PA),  Jim  Ingold  of  Louisiana  State  University 
(LA),  John  Smallwood  of  Montclair  State  University 
(NJ),  Amanda  Rodewald  of  Pennsylvania  State  Uni- 
versity (PA),  Christine  Howell  of  University  of  Mis- 
souri-Columbia  (MO),  and  Daniel  Ingold  of  Musking- 
um College  (OH). 

The  membership  poster  was  displayed  during  the 
last  year  at  the  joint  OSNA  societies  meeting  in  St. 
Louis  (MO),  April  1998,  the  Raptor  Research  Foun- 
dation meeting  in  Salt  Lake  City  (UT),  October  1998, 
and  the  Pennsylvania  Wildlife  Society  meeting  in  Wil- 
liamsport (PA),  March  1999.  Brochures  were  also  dis- 
played at  a few  other  meetings,  including  the  Penn- 
sylvania Society  of  Ornithology  in  Wellsboro,  May 
1998. 

Since  October  1998,  the  chair  has  received  at  least 
ten  inquiries  from  people  who  have  not  received  their 
journal.  Most  inquiries  occurred  between  March  and 
May.  All  inquiries  have  been  forwarded  to  Allen  Press 
and/or  the  OSNA  Director,  Anthony  Bledsoe.  Most 
people  seem  to  be  locating  a contact  person  for  mem- 
bership via  the  web  site.  The  chair  has  not  received 
any  other  inquiries  for  brochures  or  information  during 
her  tenure. 

The  membership  brochure  needs  to  be  reprinted.  Be- 
fore we  undertake  a large  run,  the  committee  would 
like  to  have  Council  review  the  text  and  layout,  etc., 
to  make  any  necessary  changes.  Ideally,  we  need  to 


604 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


make  these  changes  before  the  August  AOU  meeting. 
Please  return  all  suggestions  to  Laurie  Goodrich  or 
John  Smallwood  by  June  30.  If  anyone  needs  a copy 
of  a brochure,  please  contact  Laurie  by  reply  to  this  e- 
mail,  and  she  will  send  one  directly.  The  committee 
would  like  to  request  that  any  Council  Member  at- 
tending a meeting  other  than  the  AOU  consider  help- 
ing out  with  the  display.  The  chair  will  ship  the  poster 
to  the  site,  and  the  member  would  only  have  to  put  it 
up  and  then  repack  and  ship  it  back  at  end  of  the  con- 
ference. To  volunteer  for  a meeting,  contact  Laurie 
Goodrich,  goodrich@hawkmountain.org.  If  unable  to 
put  up  a display,  please  consider  taking  a stack  of  bro- 
chures for  distribution  tables. 

Laurie  J.  Goodrich,  Chair 

REPORT  OF  THE  UNDERGRADUATE 
OUTREACH  COMMITTEE 

Two  people  have  joined  the  committee  since  our  last 
meeting  in  April,  1998.  Thomas  Knight  of  Denver, 
CO,  and  Yves  de  Repentigny,  Departement  de  Scienc- 
es Biologiques,  Universite  de  Montreal. 

The  committee  is  pleased  to  announce  a Wilson  Bul- 
letin paper  stemming  from  the  workshop  on  teaching 
held  at  the  Guelph  meeting  in  1992.  The  citation  is: 
Burtt,  E.  H.,  Jr.,  and  W.  H.  Wilson,  Jr.  1999.  A survey 
of  ornithology  courses  in  North  America.  Wilson  Bull. 

1 1 1:287-293. 

The  Guide  to  Graduate  Programs  in  Ornithology 
that  the  Committee  has  compiled  continues  to  be  a 
valuable  resource  for  undergraduates  seeking  to  pursue 
graduate  research  in  ornithology.  The  guide  can  be 
found  on  the  Wilson  Ornithological  Society  home  page 
(http://www.ummz.lsa.umich.edu/birds/studies.html). 
Some  users  have  found  that  the  information  provided 
for  some  programs  is  outdated.  The  committee  will  be 
working  this  year  on  updating  the  material  in  the  on- 
line guide. 

Jed  Burtt  continues  to  compile  laboratory  exercises 
in  ornithology.  These  exercises  will  be  made  available 
at  a reasonable  cost  for  any  instructor  requesting  them. 

W.  Herbert  Wilson,  Jr.,  Chair 

REPORT  OF  THE  JOSSELYN  VAN 
TYNE  MEMORIAL  LIBRARY 
COMMITTEE 

I am  very  pleased  to  be  the  new  Chair  of  the  Jos- 
selyn  Van  Tyne  Memorial  Library  (JVTML)  Commit- 
tee. As  such  I would  like  to  take  this  opportunity  to 
thank  William  A.  Lunk  for  his  tireless  leadership  of 
the  Library  Committee.  He  has  been  chair  of  this  com- 
mittee for  40  years.  Since  1958  many  changes  have 
occurred  (e.g.,  computers  have  become  a standard  re- 
search tool)  and  under  Dr.  Lunk’s  chairmanship  the 
Library  has  grown  and  modernized  along  with  the 
times.  Dr.  Lunk  deserves  our  heart-felt  thanks  for  help- 
ing to  ensure  that  the  Library  has  been  able  to  provide 
WOS  members  and  others  with  ornithological  infor- 
mation that  they  need.  During  my  tenure  as  chair,  I 


hope  to  do  as  good  a job,  but  I do  not  think  I will  be 
able  to  do  it  for  40  years! 

My  philosophy  for  managing  the  JVTML  has  two 
parts.  First,  the  Library  needs  to  continue  acquiring 
volumes,  thereby  ensuring  that  it  stays  current  and  thus 
as  useful  as  possible.  Second,  it  needs  to  be  as  acces- 
sible as  possible  to  WOS  members  and  others,  while 
maintaining  the  integrity  of  the  Library. 

With  respect  to  acquiring  volumes,  the  following 
has  happened  over  the  past  calendar  year: 

• Loans  of  library  materials  to  members  included 
102  transactions  to  46  people  or  institutions.  These 
loans  included  417  books,  journals  and  photocop- 
ied articles,  with  many  of  the  articles  going  to  au- 
thors of  BNA  accounts. 

• A total  of  201  publications  were  received  from  167 
organizations  or  individuals.  These  included  120 
exchanges,  23  subscriptions,  and  24  gifts. 

• A few  journals  have  been  added,  primarily  from 
Africa,  and  others  dropped  because  they  have  ei- 
ther ceased  or  merged. 

• 758  items  were  donated  by  members  and  friends. 
These  donations  included  17  books,  522  journal 
issues,  and  219  reprints,  reports  and  misc.  items. 

• 17  members  and  friends  donating  materials  includ- 
ed: C.  Braun,  D.  Klem,  Jr.,  J.  Hinshaw,  O.  Komar, 
J.  Marks,  R.  Meek,  R.  Payne,  D.  Pioli,  W.  Post,  W. 
J.  Richardson,  T.  Shane,  J.  Spendelow,  R.  Tashian, 
E.  Walters,  R.  Whiting,  J.  Youngman,  and  H.  Zer- 
nickow. 

• Five  institutions  also  donated  materials:  L.  Birch 
for  the  Edward  Grey  Institute  of  Field  Ornithology, 
J.  Buki  for  the  Hungarian  Institute  of  Ornithology, 
L.  Kiff  for  the  Peregrine  Fund,  E Lohrer  for  Arch- 
bold Biological  Station,  and  G.  Penn  for  Point 
Reyes  Bird  Observatory. 

• New  items  purchased  for  $781.52  from  the  New 
Book  Fund  and  for  the  $742.88  credit  with  Buteo 
Books  included  19  books,  journal  issues,  CD’s  and 
tapes. 

• The  sale  of  209  books,  45  journal  issues  and  5 
color  plates,  all  of  which  were  duplicates,  resulted 
in  $5,082.30  plus  a credit  of  $2900  with  Buteo 
Books. 

• Gifts  to  other  institutions  included  216  journal  is- 
sues to  Hungarian  Institute  of  Ornithology,  38  jour- 
nal issues  to  Walter  Thiede,  and  3 journal  issues  to 
The  Peregrine  Fund. 

As  to  the  accessibility  of  the  materials  to  members 
and  others,  the  following  has  been  accomplished: 

• The  web  site  (http://www.ummz.lsa.umich.edu/ 
birds/wos.html)  continues  to  be  enhanced.  Journals 
currently  received  are  listed  on  the  site  as  well  as 
how  to  access  the  University  of  Michigan's  on-line 
catalogue,  which  can  be  used  to  check  holdings. 

• The  large  number  of  duplicate  books  for  sale  from 
Helen  Lapham’s  bequest  will  soon  be  listed  on  the 
web.  When  it  is  available,  a notice  will  be  posted 
in  the  0,SNA  Newsletter. 


Annual  rkport 


605 


Due  to  lack,  of  space,  we  need  to  reduce  our  stock 
of  back  issues  of  The  Wilson  Bulletin.  We  plan  to  keep 
around  50  to  75  copies  of  each  issue  where  possible. 
We  would  like  to  sell  (give  away,  actually)  surplus 
copies  for  the  cost  of  postage  and  handling.  Thus  far 
we  have  “sold”  over  1300  issues.  Certainly  we  do  not 
want  surplus  issues  to  end  up  in  the  recycle  bin  and 
ideally  we  would  like  these  issues  to  go  to  institutions 
in  lesser  developed  countries.  I do  not  know  the  best 
way  to  determine  which  institutions  could  use  them. 
Any  ideas  to  help  us  solve  this  problem  would  be  most 
appreciated. 

Needs  for  the  1999/2000  academic  year: 

• We  need  help  in  identifying  institutions  in  lesser 
developed  countries  that  could  use  copies  of  back 
issues.  We  currently  have  suiplus  copies  of  most 
issues,  which  will  allow  us  to  provide  nearly  com- 
plete runs. 

• An  additional  part-time  student  is  needed  to  help 
the  Library  work  to  reduce  significantly  its  surplus 
of  back  issues  of  the  Bulletin,  to  get  more  infor- 
mation up  on  the  web,  and  to  send  out  the  large 
numbers  of  orders  we  will  hopefully  generate  (via 
the  web)  from  members  and  friends  wanting  to  buy 
duplicate  copies  from  Helen  Lapham’s  bequest.  I 
do  not  know  the  particulars  of  how  to  obtain  funds 
to  hire  a student.  I assume  that  funds  would  need 
to  be  run  through  the  University  of  Michigan, 
which  would  mean  overhead  would  be  applied. 
The  amount  of  overhead  would  depend  on  what 
the  Society’s  bylaws  say  or  what  is  customary  for 
the  Society  to  pay.  (The  full  indirect-cost  rate  at 
the  University  is  52.5%,  but  can  range  from  that 
to  0%,  depending  on  what  the  funder  requires.)  We 
would  like  to  have  a student  who  works  at  $7.00 
an  hour  for  an  average  of  15  hours  a week  over  1 1 
months,  and  with  8.55%  in  benefits  (i.e.,  FICA). 
Thus  a total  of  $5,000  (direct  cost)  would  be  need- 
ed. Certainly  if  we  can  find  a work/.study  student, 
then  we  would  need  only  around  2/3  of  that  or 
$3,400.  We  anticipate  that  sales  from  Helen  La- 
pham’s bequest  will  provide  that  amount  of  money 
and  more,  but  it  will  not  be  available  ahead  of 
time. 

My  job  has  been  quite  ea.sy  due  to  the  help  of  Janet 
Hinshaw,  who  manages  the  day-to-day  tasks  in  the  Li- 
brary. Indeed,  she  is  the  one  who  provided  me  with 
the  figures  I am  reporting  below. 

In  closing,  I want  to  offer  my  thanks  to  those  people 
who  make  the  JVTML  work  .so  efficiently:  Joann  Con- 
stantinides,  the  secretary  for  the  Bird  Division  in  the 
Museum  of  Zoology,  who  handles  many  of  the  library 
requests;  Kari  Chciuk,  the  library  work/.study  student, 
who  for  2 years  has  handled  most  of  the  day-to-day 
filing  and  routine  jobs,  as  well  as  sorting  thousands  of 
reprints;  and  of  course  to  THE  most  important  person 
connected  with  the  Library,  Janet  Hinshaw,  who  pro- 
vided me  with  the  figures  I reported  above  and  who 
literally  “runs  the  show”  by  making  sure  the  library 
is  kept  up-to-date,  and  as  useful  to  members  and 


friends  as  possible.  All  of  these  people  make  my  job 
quite  easy. 

Terry  L.  Root,  Chair 

The  Committee  on  the  Scientific  Program,  consist- 
ing of  William  E.  Davis,  Jr.,  chair,  and  John  C.  Kricher, 
presented  the  following  program,  assisted  by  session 
moderators  and  workshop  organizers  Edward  H.  Burtt, 
Jr.,  Richard  N.  Conner,  Janet  Hinshaw,  Jerome  A. 
Jackson,  Sara  R.  Monis,  John  A.  Smallwood,  and  Su- 
san M.  Smith. 

PAPER  SESSIONS 

D.  J.  Albano,  Univ.  of  Massachusetts,  Amherst,  MA, 
“Partial  migration  in  the  Belted  Kingfisher  (Cer- 
yle  o Icy  on).” 

J.  C.  Barlow  and  S.  N Leckie,  Royal  Ontario  Museum, 
Toronto,  ON,  Canada,  “Winter  frugivory  in  Gray 
Vireos — do  eastern  populations  eat  berries  too?” 
R.  E.  Brown,  USFWS  Southern  Research  Station,  Nac- 
ogdoches, TX,  “Habitat,  nest  site,  and  nest  box 
selection  by  the  Prothonotory  Warbler  in  eastern 
Texas.” 

R.  N.  Conner,  D.  Saenz,  and  D.  C.  Rudolph,  Southern 
Research  Station,  Nacogdoches,  TX,  “The  value 
of  Red-cockaded  Woodpecker  cavity  trees  for 
studying  initial  attack  of  pines  by  southern  pine 
beetles.” 

R.  L.  Curry  and  J.  Wotanis,  Villanova  Univ.,  Villan- 
ova,  PA,  “Northward  movement  of  the  Carolina/ 
Black-capped  Chickadee  contact  zone  in  south- 
eastern Pennsylvania:  inferences  from  song  pat- 
terns.” 

T.  ’V.  Dietsch  and  A.  H.  Mas,  Univ.  of  Michigan,  Ann 
Arbor,  Ml,  “Relating  Neotropical  birds  and  veg- 
etative structure  to  certification  criteria  for  coffee 
agroecosystems  in  Chiapas,  Mexico.” 

P.  C.  Frederick,  Univ.  of  Florida,  Gainesville,  FL, 
“Community  structure  and  population  dynamics 
of  breeding  wading  birds  in  the  Everglades.” 

M.  J.  Forman,  E.  H.  Burtt,  Jr.,  and  J.  M.  Ichida,  Ohio 
Wesleyan  Univ.,  Delaware,  OH,  “Microorganisms 
in  the  plumage  of  Galliformes.” 

M.  J.  Hartley,  Univ.  of  Maine,  Orono,  ME,  “Effects 
of  partial  cutting  on  avian  community  composi- 
tion.” 

F.  Huettmann,  Atlantic  Cooperative  Wildlife  Ecology 
Research  Network,  Fredericton,  NB,  Canada, 
“Wintering  Razorbills,  AIca  torda.  and  auk  as- 
semblages in  the  lower  Bay  of  Fundy,  Canada. 
Results  from  two  winter  surveys  1997/98  and 
1998/99.” 

J.  M.  Ichida,  J.  M.  Mann,  and  E.  H.  Burtt,  Jr.,  Ohio 
Wesleyan  Univ.,  Delaware,  OH,  “The  nest,  the 
nestling,  and  the  microbe.” 

J.  A.  Jackson,  Whitaker  Center,  Florida  Gulf  Coast 
Univ.,  Ft.  Myers,  FL,  and  W.  E.  Davis,  Jr.,  Boston 
Univ.,  Boston,  MA,  “Great  Egrets,  catfish,  and  the 
economics  of  birds  and  humans.” 

R.  Z.  Jennings,  Univ.  of  Texas.,  Austin,  TX,  “Spatial 


606 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


and  temporal  variation  in  the  distributions  of  Ca- 
lypte  hummingbirds  along  an  elevational  transect. 
Riverside  County,  CA.” 

J.  M.  Jonas  and  J.  A.  Smallwood,  Montclair  State 

Univ.,  Upper  Montclair,  NJ,  “The  influences  of 
weather  and  migration  pressure  on  the  growth  of 
nestling  American  Kestrels  in  northwestern  New 
Jersey.” 

E.  L.  Kershner,  J.  W.  Walk,  and  R.  E.  Warner,  Univ. 
of  Illinois,  Urbana,  IL,  “The  effect  of  predator 
removal  on  grassland  bird  nest  success.” 

D.  L.  King,  R.  M.  DeGraaf,  USDA  Forest  Service 
Northeastern  Forest  Experiment  Station,  and  C.  R. 
Gritfin,  Univ.  of  Massachusetts,  Amherst,  MA, 
“Do  predation  rates  on  artificial  nests  accurately 
reflect  predation  rates  on  natural  bird  nests?” 

P.  A.  McDowell,  J.  M.  Ichida,  and  E.  H.  Burn,  Jr.,  Ohio 
Wesleyan  Univ.,  Delaware,  OH,  "Streptomyces 
and  the  evolution  and  microecology  of  avian 
plumage.” 

K.  E.  Miller,  Univ.  of  Florida,  Gainesville,  FL,  “Nest- 

ing success  of  the  Great  Crested  Flycatcher  in  nat- 
ural nests  and  in  nest  boxes;  predation  rates  in- 
crease with  nest  box  age.” 

J.  J.  Nocera  and  P.  D.  Taylor,  Acadia  Univ.,  Wolfville, 

NS,  Canada,  “Behavior  of  failed  and  non-breed- 
ing Common  Loons  during  the  summer  breeding 
season.” 

K.  A.  Peters,  North  Carolina  State  Univ.,  Raleigh,  NC 

“Swainson’s  Warbler  habitat  selection  in  a man- 
aged botttomland  hardwood  forest  in  South  Car- 
olina.” 

Z.  Richards  and  D.  Capen,  Univ.  of  Vermont,  Burling- 
ton, VT,  “Landscape  effects  on  the  Black-throated 
Blue  Warbler  {Denciroica  caende.scensy.  a com- 
parison of  demographics  in  large  forest  isolates 
and  contiguous  forest.” 

D.  C.  Rudolph,  R.  N.  Conner,  R.  H.  Johnson,  W.  G. 
Ross,  S.  J.  Burgdorf,  R.  R.  Schaefer,  and  D. 
Saenz,  Southern  Research  Station,  Nacogdoches, 
TX,  “Red-cockaded  Woodpecker  cavity  trees, 
southern  pine  beetle  mortality,  and  pine  resin 
characteristics.” 

D.  Saenz,  R.  N.  Conner,  C.  S.  Collins,  and  D.  C.  Ru- 
dolph, Southern  Research  Station,  Nacogdoches, 
TX,  “Long-term  use  of  artificial  cavities  by  Red- 
cockaded  Woodpeckers  in  eastern  Texas.” 

J.  A.  Sedgwick,  Biological  Resources  Div.,  Fort  Col- 
lins, CO,  “Geographic  variation  in  the  song  of  the 
Willow  Flycatcher." 

C.  M.  Somers  and  R.  D.  Morris,  Brock  Univ.,  St.  Cath- 
arines, ON,  Canada,  “Bird  depredation  of  grapes 
in  a Niagara  vineyard:  do  predictable  trends  ex- 
ist?” 

R.  Stiehl  and  A.  Farmer,  Midcontinent  Ecological  Sci- 
ence Center,  Fort  Collins,  CO,  “Richness  of  Pip- 
ing Plover  foraging  habitat  on  Fire  Island,  NY.” 

J.  W.  Walk,  E.  L.  Kershner,  and  R.  E.  Warner,  Univ. 
of  Illinois,  Urbana,  IL,  “Area  and  age  of  habitat 
patches  and  nesting  succe.ss  in  grassland  birds.” 

J.  R.  Walters,  Virginia  Polytechnic  Institute  and  State 


Univ.,  Blacksburg,  VA,  “Experimental  studies  of 
effects  of  cavities  on  territory  quality  in  Red- 
cockaded  Woodpeckers.” 

W.  H.  Wilson,  Jr.,  Colby  College,  Waterville,  ME, 
“Arrival  dates  of  Maine  migratory  breeding  birds: 
a trans-century  comparison.” 

POSTERS 

R.  C.  Banks,  National  Museum  of  Natural  History, 
Washington,  DC,  “Variation  in  Greater  White- 
fronted  Geese  in  the  Central  Fly  way.” 

M.  T.  Bradley  and  S.  R.  Morris,  Canisius  College,  Buf- 
falo, NY,  “Is  tail  feather  shape  a reliable  indicator 
of  age  in  warblers  and  thrushes?” 

E.  H.  Burtt,  Jr.,  and  P.  Y.  Burtt,  Ohio  Wesleyan  Univ., 
Delaware,  OH,  “Ice  damage  to  feathers.” 

M.  Carey  and  J.  Mills,  Univ.  of  Scranton,  Scranton, 
PA,  “Nest-site  selection  and  breeding  biology  of 
Field  Sparrows  in  a rapidly  changing  old  field 
habitat.” 

M.  C.  Herceg,  J.  A.  Weiner,  and  S.  R.  Morris,  Canisius 
College,  Buffalo,  NY,  “A  seasonal  comparison  of 
the  stopover  patterns  and  behavior  of  migratory 
passerines  on  Appledore  Island,  Maine.” 

J.  H.  Hull  and  A.  R.  Buckelew,  Jr.,  Bethany  College, 
Bethany,  WV,  “Group  anting  in  the  Common 
Grackle.” 

J.  L.  Ingold,  Louisiana  State  Univ.,  Shreveport,  LA, 
“Winter  birds  of  small  remnant  prairies  in  the  Pin- 
ey  Woods  of  northern  Louisiana.” 

E.  D.  Kennedy  and  A.  McCauley,  Albion  College,  Al- 
bion, MI,  “Acquisition  by  Albion  College  of 
prints  from  10  original  copper  plate  engravings 
used  in  Alexander  Wilson’s  “American  Ornithol- 
ogy”-” 

G.  J.  Robertson,  G.  Chapdelaine,  and  R.  D.  Elliot,  Ca- 
nadian Wildlife  Service,  A.  W.  Diamond  and  F. 
Huettmann,  Univ.  of  New  Brunswick,  Fredericton, 
NB,  Canada,  “Population  status  and  trends  of  Ra- 
zorbills in  Canada.” 

C.  R.  Smith,  J.  T.  Weber,  and  M.  E.  Richmond,  Cornell 

Univ.,  Ithaca,  NY,  “Using  gap  analysis  informa- 
tion to  guide  planning  for  con.servation  of  birds  in 
New  York  state;  a comparison  of  science-based 
and  expert-opinion  approaches.” 

D.  Westmoreland  and  M.  Moseley,  US  Air  Force 
Academy,  Colorado  Springs,  CO,  “The  cost  of 
bright  egg  coloration  in  American  Robin  nests.” 

D.  W.  White  and  E.  D.  Kennedy,  Albion  College,  Al- 
bion, MI,  “Do  House  Wrens  identify  target  eggs 
by  sight  or  by  feel?” 

J.  B.  Whittier  and  D.  M.  Leslie,  Jr.,  US' Geological 
Survey,  Biological  Resources  Division,  Oklahoma 
Cooperative  Fish  and  Wildlife  Research  Unit, 
Oklahoma  State  Univ.,  Stillwater,  OK,  “The  effect 
of  drought  on  nest  success  in  Least  Terns.” 

ATTENDANCE 

COLORADO:  For!  CoUin.s,  Jim  Sedgwick,  Dick 
Stiehl;  Colorado  Sprinf>.s,  David  Westmoreland. 


Annual  report 


607 


CONNECTICUT;  New  Britain,  Sylvia  Halkin;  Sharon, 
Elyse  Glover. 

DISTRICT  OF  COLUMBIA:  Washington,  Dick 
Banks. 

FLORIDA:  Fort  Myers,  Jerry  Jackson;  Gainesville, 
Mary  Clench,  Peter  Frederick,  Karl  Miller;  Tal- 
lahassee, Fran  James. 

ILLINOIS:  Bushell,  Larry  Hood;  Centralia,  Pricilla 
McDowell;  Collinsville,  Kimberly  Peters;  Grays- 
lake,  Scott  Hickman;  Urbana,  Eric  Kershner,  Jef- 
frey Walk. 

INDIANA;  Notre  Dame,  Doris  Watt. 

KANSAS;  Manhattan,  Dave  Rintoul. 

LOUISIANA:  Shreveport,  James  Ingold. 

MAINE;  Bar  Harbor,  Goodale  Wing;  Belgrade,  Don 
Mairs;  Fairfield,  Miriam  Bennett;  Farmington, 
Sarah  Sloane;  Orono,  Mitschka  Hartley;  Rich- 
mond, Peter  Vickery;  Unity,  Ed  Beals;  Waterville, 
Larkspur  Morton,  Neal  Taylor,  Herb  Wilson;  Wil- 
ton, Wendy  Howes. 

MARYLAND:  Chevy  Chase,  Ellen  Paul;  St.  Mary’s 
City,  Ernie  Willoughby. 

MASSACHUSETTS;  Boston,  Ted  Davis;  South  Had- 
ley, Susan  Smith. 

MICHIGAN:  Albion,  Dale  Kennedy,  Doug  White;  Ann 
Arbor,  Tom  Dietsch;  Chelsea,  Janet  Hinshaw; 
Kalamazoo,  Richard  Brewer. 

NEW  HAMPSHIRE;  Keene,  Jon  Atwood;  West  Swa- 
zey,  Lewis  Kibler. 

NEW  JERSEY:  Belleville,  Jeffrey  Jonas;  Cape  May, 


Tom  Parsons;  Edison,  Robert  Colburn;  Upper 
Montclair,  John  Smallwood. 

NEW  YORK:  Buffalo,  Sara  Morris;  Cheetkowaga,  Je- 
anette Weiner;  Geneseo,  Bob  Beason;  Grand  Is- 
land, Maria  Bradley;  Ithaca,  Charles  Smith;  yo/?n- 
son  City,  Michael  Herceg;  Utica,  Judy  McIntyre. 

NORTH  CAROLINA:  Asheville,  Lou  Weber;  Chape! 
Hill,  Helmut  Mueller,  Nancy  Mueller. 

OHIO:  Columbus,  Sandy  Gaunt,  Toby  Gaunt;  Dela- 
ware, Jed  Burtt,  Martin  Forman,  Jann  Ichida;  Sun- 
bury,  Kathy  Wildman. 

OKLAHOMA;  Stillwater,  Joanna  Whittier. 

PENNSYLVANIA:  Berwyn,  Phil  Street;  Doylestown, 
Jennifer  Niese;  Kempton,  Laurie  Goodrich;  Scran- 
ton, Michael  Carey;  Villanova,  Bob  Curry; 
Swarthmore,  Janet  Williams. 

RHODE  ISLAND;  Wakefield,  Cecil  Kersting. 

TEXAS:  Austin,  Rachel  Jennings;  Belton,  John  Cor- 
nelius; Galveston,  Dwight  Peake;  Nacogdoches, 
Raymond  Brown,  Dick  Conner,  Craig  Rudolph, 
Dan  Saenz. 

VERMONT;  Northfield,  Bill  Barnard. 

VIRGINIA:  Arlington,  Marcus  Koenen;  Blacksburg, 
Jeff  Walters. 

WEST  VIRGINIA:  Bethany,  Jay  Buckelew;  Morgan- 
town, George  Hall. 

CANADA;  Fredericton  New  Brunswick,  Falk  Huett- 
man;  St.  Catherine’s  Ontario,  Christopher  Somer; 
Toronto  Ontario,  Jon  Barlow,  Sheridan  Leckie; 
Wolfville  Nova  Scotia,  Joseph  Nocera. 


Wilson  Bull,  111(4),  1999,  pp.  608-609 


Acknowledgments 


The  following  individuals  graciously  served  as  referees  for  the  volume  of  The  Wilson  Bulletin.  1 am  deeply 
grateful  for  their  assitance  and  advice — Robert  C.  Reason,  Editor. 


C.  S.  Adkisson,  D.  G.  Ainley,  J.  L.  B.  Albu- 
querque, T.  A1  worth,  E.  Ammon,  D.  Ander- 
son, J.  G.  T.  Anderson,  T.  R.  Anderson,  B.  A. 
Andres,  C.  D.  Ankney,  R.  D.  Applegate,  D. 
Arcese,  W.  J.  Arendt,  F.  Arengo,  K.  A.  Arnold, 
R.  A.  Askins,  J.  E.  Austin,  M.  L.  Avery,  M. 
C.  Baker,  N.  E.  Baldaccini,  R.  C.  Banks,  D. 
R.  Barber,  E K.  Barker,  J.  R.  Bart,  J.  M.  Bates, 
E.  Bayne,  K.  G.  Beal,  D.  L.  Beaver,  J.  C.  Bed- 
narz,  B.  M.  Beehler,  S.  R.  Beissinger,  L.  D. 
Beletsky,  J.  R.  Belthoff,  W.  Belton,  T.  M.  Ber- 
gin,  R W.  Bergstrom,  S.  L.  Berman,  L.  B. 
Best,  J.  Bielefeldt,  K.  L.  Bildstein,  M.  A. 
Bishop,  J.  G.  Blake,  C.  R.  Blem,  W.  M.  Block, 
J.  Blondel,  C.  W.  Boal,  C.  E.  Bock,  E.  K.  Bol- 
linger, E.  S Botelho,  R.  Bowman,  G.  T.  Bra- 
den, C.  E.  Braun,  J.  D.  Brawn,  R.  M.  Brigh- 
am, D.  J.  Bright-Smith,  M.  C.  Brittingham,  E. 
W.  Brooks,  B.  T.  Brown,  C.  R.  Brown,  K.  M. 
Brown,  A.  H.  Brush,  T.  Brush,  A.  Burger,  D. 
E.  Burhans,  B.  E.  Byers,  V.  J.  Byre,  D.F.  Cac- 
camise,  L.  D.  Caldwell,  S.  B.  Canale,  R.  A. 
Canterbury,  T.  Caraco,  M.  D.  Carey,  T.  W. 
Carpenter,  N.  R.  Carrie,  E.  L.  Caton,  E.  Cham- 
berlain, B.  R.  Chapman,  M.  K.  Chase,  A.  R. 
Clark,  R.  G.  Clark,  L.  M.  Coburn,  J.  A.  Col- 
lazo, D.  M.  Collister,  R.  N.  Conner,  P.  G.  Con- 
nors, C.  J.  Conway,  E Cooke,  C.  J.  Counard, 
R.  J.  Craig,  D.  A.  Cristol,  J.  E Cully,  R.  L. 
Curry,  D.  R.  Curson,  T.  W.  Custer,  E J.  Cuth- 
bert,  T.  V.  Dailey,  G.  C.  Daily,  S.  J.  Daniels, 
M.  P.  Darveau,  W.  B.  Davison,  D.  K.  Dawson, 
R.  D.  Dawson,  B.  Day,  S.  Debus,  R.  M.  De- 
graaf,  J.  R.  Des  Lauriers,  T.  DeSanto,  T.  L. 
Devault,  A.  W.  Diamond,  V.  M.  Dickison,  D. 

R.  Diefenbach,  R.  H.  Diehl,  T.  V.  Dietsch,  J. 
J.  Dinsmore,  R.  C.  Dobbs,  S.  L.  Dodd,  T.  Don- 
ovan, A.  W.  Doolittle,  S.  Dove,  R.  D.  Drob- 
ney,  D.  C.  Duffy,  J.  R.  Duncan,  J.  B.  Dunning, 
C.  R.  Dykstra,  S.  W.  Eaton,  J.  R.  Eberhard,  L. 

S.  Eberhardt,  M.  E.  Eddins,  W.  R.  Eddleman, 
R.  T.  Engstrom,  A.  J.  Erskine,  R.  M.  Erwin, 

C.  Estades,  M.  R.  Evans,  T.  J.  Evans,  J.  R. 
Faaborg,  S.  Faccio,  G.  H.  Farley,  G.  Farns- 
worth, M.  B.  Fenton,  E.  J.  Finck,  L.  D.  Flake, 

D.  J.  Flaspohler,  P L.  Flint,  T.  B.  Ford,  M.  S. 


Foster,  A.  D.  Fox,  A.  B.  Franklin,  J.  D.  Fraser, 
H.  Freifeld,  L.  E.  Friesen,  R.  W.  Furness,  S. 
W.  Gabrey,  J.  L Ganey,  K.  L.  Garrett,  R.  J. 
Gates,  A.  S.  Gaunt,  D.  G.  Gawlik,  E R.  Gehl- 
bach,  R.  P Gerhardt,  J.  C.  Gering,  J.  A.  Ges- 
saman,  H.  G.  Gilchrist,  C.  E.  Gill,  S.  Gill,  J. 
Glahn,  R.  S.  Gnam,  N.  M.  Gobris,  M.  Goch- 
feld,  C.  B.  Goguen,  T.  Goldsmith,  G.  Good- 
ing, L.  J.  Goodrich,  C.  E.  Gordon,  B.  M.  Gott- 
fried, C.  Graham,  W.  Gram,  J.  G.  Granlund, 

C.  L.  Gratto-Trevor,  G.  A.  Green,  C.  H. 
Greenberg,  J.  S.  Greenlaw,  J.  W.  Grier,  C.  R. 
Griffin,  C.S.  Griffiths,  K.  Groschupf,  T.  C. 
Grubb,  R.  Grundel,  J.  A.  Grzybowski,  K.  J. 
Gutzwiller,  S.  J.  Hackett,  J.  H.  Haffer,  T.  M. 
Haggerty,  S.  M.  Haig,  J.  P Hailman,  E R. 
Hainsworth,  D.  A.  Hall,  M.  J.  Hamas,  P B. 
Hamel,  I.  K.  Hanski,  R.  G.  Harper,  M.  J.  Har- 
tley, D.  Haskell,  E.  W.  Hein,  H.  T.  Hendrick- 
son, C.  J.  Henny,  G.  R.  Hepp,  J.  R.  Herkert, 
F Hertel,  S.  K.  Herzog,  D.  O.  Hill,  S.  L.  Hilty, 
A.  Holmes,  D.  W.  Holt,  P D.  Hooper,  J.  P. 
Hoover,  S.  L.  Hopp,  D.  J.  Horn,  C.  S.  Hous- 
ton, J.  D.  Hubbard,  J.  Hughes,  W.  G.  Hunt,  D. 
J.  T.  Hussell,  R.  L.  Hutto,  L.  D.  Igl,  D.  J. 
Ingold,  M.  L.  Isler,  E M.  Jaksic,  D.  A.  James, 
E C.  James,  R.  A.  James,  A.  P.  Jaramillo,  R. 

L.  Jarvis,  D.  A.  Jenni,  W.  E.  Jensen,  D.  H. 
Johnson,  L.  S.  Johnson,  M.  D.  Johnson,  J.  Jo- 
livette,  I.  L.  Jones,  J.  Jones,  K.  L.  Jones,  S. 
Jones,  D.  G.  Jorde,  S.  M.  Joy,  R.  Kane,  T.  L. 
Kast,  W.  Kendall,  R.  A.  Kennamer,  E.  D.  Ken- 
nedy, R L.  Kennedy,  D.  M.  Keppie,  L.  H.  Ker- 
mott,  H.  Khanna,  A.  A.  Kinsey,  M.  Klaassen, 
P K.  Kleintjes,  J.  T.  Klicka,  D.  S.  Klute,  R. 
W.  Knapton,  M.  T.  Koenen,  W.  D.  Koenig,  R. 
R.  Koford,  O.  Komar,  N.  Krabbe,  A.  W.  Krat- 
ter,  D.  G.  Krementz,  J.  C.  Kricher,  K.  S.  Kritz, 

D.  E.  Krodsma,  J.  D.  Lambert,  R.  B.  Lanctot, 
R.  S.  Lange,  D.  V.  Fanning,  W.  E.  Lanyon,  S. 
Lariviere,  B.  Lauro,  J.  J.  Lawler,  R E.  Lederle, 

M.  R.  Lein,  R.  E.  Lemon,  R.  A.  Lent,  M.  L. 
Leonard,  D.  J.  Levey,  E.  T.  Linder,  B.  D.  Link- 
hart,  C.  D.  Littlefield,  J.  R Loegering,  B.  A. 
Loiselle,  E.  R.  Loos,  P.  E.  Lowther,  R.  C. 
Lundquist,  J.  E.  Lyons,  T.  R.  Mace,  R.  H.  E 


608 


ACKNOWLEDGMENTS 


609 


Macedo,  R.  E.  Maier,  T.  J.  Maier,  B.  P Mar- 
ansky.,  M.  A.  Marin,  M.  A.  Marini,  J.  S. 
Marks,  C.  D.  Marti,  P.  M.  Mayer,  H.  L.  Mays, 
J.  C.  Mazourek,  K.  M.  Mazur,  T.  McCoy,  K. 
P McFarland,  W.  B.  McGillivray,  J.  W.  Mc- 
Intyre, K.  McKay,  M.  R.  McLandress,  D.  B. 
McNair,  R.  McNeil,  S.  M.  Melvin,  B.  D.  Mey- 
burg,  J.  M.  Meyers,  E.  H.  Miller,  K.  E.  Miller, 

B.  A.  Millsap,  M.  Milonoff,  D.  Mock,  A.  P 
M0ller,  R.  Montgomery,  D.  Moore,  E R. 
Moore,  D.  C.  Morimoto,  S.  R.  Morris,  M.  L. 
Morrison,  S.  G.  Mosa,  R.  S.  Mulivihill,  M.  T. 
Murphy,  J.  J.  Negro,  D.  A.  Nelson,  S.  A.  Nes- 
bitt, G.  J.  Niemi,  M.  Nogales,  E.  Nol,  C.  J. 
Norment,  R.  L.  Norton,  J.  P.  O’Neill,  P Olsen, 
L.  W.  Oring,  C.  P.  Ortega,  H.  Ouellet,  R.  B. 
Owen,  K.  C.  Parsons,  D.  N.  Pashley,  C.  A. 
Paszkowski,  P.  W.  Paton,  R.  B.  Payne,  G.  Pen- 
delton,  C.  M.  Perrins,  B.  G.  Peterjohn,  T.  A. 
Peterson,  J.  Pieman,  P.  J.  Pietz,  W.  H.  Piper, 

B.  J.  Ploger,  D.  W.  Pogue,  W.  Post,  B.  Poulin, 
A.  N.  Powell,  H.  W.  Power,  S.  Pribil,  R.  O. 
Prum,  K.  L.  Purcell,  W.  H.  Pyle,  P Radewald, 
S.  A.  Raouf,  J.  H.  Rappole,  S.  E.  Reinert,  D. 

L.  Reinking,  L.  Reitsma,  J.  V.  Remsen,  M.  K. 
Reynolds,  R.  E.  Ricklefs,  R.  S.  Ridgely,  C.  M. 
Riley,  J.  D.  Rising,  G.  Ritchison,  J.  H.  Rivera, 

C.  S.  Robbins,  M.  B.  Robbins,  G.  J.  Robert- 
son, W.  D.  Robinson,  N.  L.  Rodenhouse,  H. 

M.  Rogers,  R.  N.  Rosenfield,  R.  R.  Roth,  S. 
I.  Rothstein,  S.  M.  Russell,  M.  R.  Ryan,  D. 
Saenz,  L.  Santisteban,  R.  A.  Sargent,  G.  Sat- 
tler,  J.  A.  Savidge,  T.  L.  Scarlett,  P E 
Schempf,  A.  Scheuerlein,  S.  R.  Schmidt,  J.  K. 


Schmutz,  G.  D.  Schnell,  E.  A.  Schreiber,  T. 
Schulenberg,  D.  Schulter,  J.  A.  Sedgwick,  G. 
Seutin,  C.  E.  Shackelford,  D.  H.  Shedd,  M.  A. 
Shields,  W.  G.  Shriver,  M.  Siders,  T.  R.  Si- 
mons, C.  G.  Sims,  J.  A.  Smallwood,  C.  R. 
Smith,  D.  G.  Smith,  J.  N.  M.  Smith,  M.  K. 
Sogge,  S.  S.  Soukup,  D.  A.  Spector,  J.  A. 
Spendelow,  S.  E H.  Spofford,  D.  W.  Stead- 
man, K.  Steenhof,  M.  A.  Stem,  R.  B.  Stiehl, 
C.  M.  Stinson,  S.  H.  Stoleson,  D.  E Stotz,  R 

C.  Stouffer,  W.  D.  Svedarsky,  D.  L.  Swanson, 
P W.  Sykes,  D.  A.  Tallman,  W.  K.  Taylor,  D. 
L.  M.  Teixeira,  R.  C.  Telfair,  E.  J.  Temeles,  G. 

D.  Therres,  G.  R.  Thomas,  C.  E Thompson, 
C.  W.  Thompson,  J.  E.  Thompson,  S.  C. 
Thompson,  T.  L.  Tibbits,  M.  E.  Tobin,  T Tra- 
cy, C.  L.  Trine,  J.  D.  Tyler,  T J.  Underwood, 
C.  Van  Riper,  W.  M.  Vander  Hagen,  D.  W. 
Verser,  P.  D.  Vickery,  M.  A.  Villard,  G.  H. 
Visser,  C.  M.  Vleck,  K.  H.  Voous,  E Vuilleu- 
mier,  S.  J.  Wagner,  J.  W.  Walk,  G.  E.  Wals- 
burg,  C.  Ward,  D.  Ward,  I.  G.  Warkentin,  K. 
S.  Warner,  B.  D.  Watts,  L.  M.  Weber,  L.  C. 
Wemmer,  C.  G.  West,  G.  B.  West,  D.  West- 
moreland, C.  J.  Whelan,  D.  Whitacre,  C.  M. 
White,  M.  J.  Whitfield,  R.  C.  Whitmore,  D. 
A.  Wiedenfeld,  E A.  Wilkinson,  D.  E.  Willard, 
J Williams,  L.  E.  Williams,  C.  N.  Willis,  W. 
Wiltschko,  J.  C.  Wingfield,  D.  W.  Winkler,  M. 
Winter,  L.  L.  Wolf,  D.  H.  Wolfe,  R B.  Wood, 
R W Woodward,  B.  L.  Woodworth,  A.  L. 
Wright,  R.  H.  Yahner,  P.  H.  Yaukey,  Y Yom- 
Tov,  R.  Yosef,  S.  Zack,  G.  D.  Zenitsky,  M.  C. 
Zicus 


Wilson  Bull.,  111(4),  1999,  pp.  610-630 


Index  to  Volume  111,  1999 

By  Kathleen  G.  Beal 


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  AOU  Check-list  of  North  American  Birds  (1998). 
Reference  is  made  to  books  reviewed,  and  announcements  as  they  appear  in  the  volume. 


Ahurria  ahurri,  564-569 
abundance 

effect  of  wind  turbines  on  upland  nesting  birds, 
100-104 

of  breeding  birds  in  old-growth  forests,  89-99 
of  Sitta  pusilla  after  pine  plantation  thinning,  56-60 
Accipiter  hicolor,  225 
hrevipes,  181-187 
cooperii,  7—14,  558 
gentilis,  432-436,  475 
ni.sus,  181-187 
stricitus,  7—14,  92,  558,  575 
superciliosus, 

Acrochordopus  burmeisteri,  412 
Actophilornis  africanus,  264 
Aegolius  funereus,  273 
age 

effect  on  breeding  performance  in  Dendroica  pete- 
chia, 381—388 

Agelaius  phoeniceus,  84,  86,  100-104,  105—114,  131, 
254,  348,  554,  555,  558 
tricolor,  425 
Agriornis  Uvula,  530 
Aimophila  aestivalis,  56 
Aix  sponsa,  1-6,  108,  109,  469 
Alden,  Peter,  reviews  by,  595,  595-596,  596 
Alectrurus  tricol.  I'll 
Alouatta  palliata,  125 

Alvarez  A.,  Jose,  see  Krabbe,  Niels,  Morton  L.  Isler, 

Phyllis  R.  Isler,  Bret  M.  Whitney, , and 

Paul  J.  Greenfield 

Alworth,  Tom,  and  Isabella  B.  R.  Scheiber,  An  incident 
of  female-female  aggression  in  the  House 
Wren,  130-132 

Amazon,  Blue-fronted,  see  Amazona  aestiva 
Amazona  aestiva,  225,  408 
finschi,  488-493 
vinacea,  408 

Amhassis  gymnocephalus,  19 
Ammodramus  hairdii,  389—396 
henslowii,  515-527 
humeralis,  228 
leconteii,  1 00—  1 04 
savannarum.  1 00—  1 04 
Anahacerthia  amaurotis,  410 
Anairetes  parulus,  530 
Anas  acuta,  468,  470 

hahamensis  haliamensis,  56 1 
discor.s,  108,  109,  111,  468,  469 


platyrhynchos,  108,  109,  111,469 
ruhripes,  108 
spp.,  17 

Anderson,  James  T,  and  Thomas  C.  Tacha,  Habitat  use 
by  Masked  Ducks  along  the  Gulf  Coast  of  Tex- 
as, 119-121 

Angehr,  George  R.,  Rapid  long-distance  colonization 
of  Lake  Gatun,  Panama,  by  Snail  Kites,  265- 
268 

Ani,  Greater,  see  Crotophaga  major 
Smooth-billed,  see  Crotophaga  ani 
Anjos,  Luiz  Dos,  and  Roberto  Bo^on,  Bird  commu- 
nities in  natural  forest  patches  in  southern  Bra- 
zil, 397-414 
announcement 

Lincoln  Park  Zoo  Scott  Neotropic  and  Africa/Asia 
Funds,  456 

Anodorhynchus  hyacinthinus,  225 
Anser  alhifrons,  167 

albifrons  frontalis,  166—180 
caerulescens,  166,  167,  177 
canagica,  1 66 
rossii,  166 

Ant-Tanager,  Red-crowned,  see  Habia  rubica 
Antbird,  Black-chinned,  see  Hypocnemoides  melano- 
pogon 

Black-throated,  see  Myrmeciza  atrothorax 
Caura,  see  Percnostola  caruensis 
Dusky,  see  Cercomacra  tyrannina 
Jet,  see  Cercomacra  nigricans 
Rio  Branco,  see  Cercomacra  carbonaria 
Silvered,  see  Sclateria  naevia 
Yapacana,  see  Myrmeciza  disjuncta 
Anthus  spragueii,  389—396 
Antilophia  galeata,  222 
Antwren,  Ornate,  see  Myrmotherula  ornatua 
Rusty-backed,  see  Formicivora  rufa 
White-eyed,  see  Myrmotherula  leucophthalma 
Anumhius  annumbi,  410 
Aphanotriccus  capitalils,  124-128 
Aphelocoma  californica,  252 
coerulescens,  285,  348 
Aquila  adalberti,  475 

chrysaetos,  437,  472-477 
Ara  chloropterus,  225 
maracana,  225 

Aragones,  Juan,  Luis  Arias  de  Reyna,  and  Pilar  Re- 
cuerda.  Visual  communication  and  sexual  se- 
lection in  a nocturnal  bird  species,  Caprimul- 


610 


INDEX  TO  VOLUME  I 1 I 


611 


guM  ruficollis,  a balance  between  crypsis  and 
conspicLiousness,  340-345 
Aramides  cajanea,  225 
saracurti,  407 
Aratinga  aurea,  225 
leucophthcdmus,  225 
Archilochus  colubris.  92,  94 
Arenaria  interpres,  263 

Arendt,  Wayne  J.,  see  Woodworth,  Bethany  L.,  John 

Faaborg,  and  

Arius  mudagascariensis,  19 

armadillo,  nine-banded,  see  Dasypus  novemcinclus 
Arremon  aurantiirostris,  234 
flavirostris,  228,  233 
Arremonops  rufivirgatus,  234 
arrival  patterns 

of  Vireo  griseus,  46—55 
Asia  capensis, 

flanitneus  [5rr/x  domingensis,  Asia  portoricensis, 
Asia  domingensis},  274,  303—313  (Frontis- 
piece) 
otus,  273-276 
spp.,  218 

Asthenes  modesta,  530 
Ateles  geojfroyi,  125 

Athene  cunicularia,  273,  310,  434,  569—571 
Attagis  goyi,  264 

Avery,  Michael  L.,  Carrie  L.  Schreiber,  and  David  G. 
Decker,  Fruit  sugar  preferences  of  House 
Finches,  84—88 
avifauna 

at  the  Paraguayan  Cerrado  site,  Parque  Nacional 
Seram'a  San  Luis,  depto.  Concepcion,  216— 
228 

of  Tabasco,  229—235 

Avocet,  American,  see  Recurvirostra  americana 
European,  see  Recurvirostra  avosetta 
awards  and  grants 

Lincoln  Park  Zoo  Scott  Neotropic  and  Africa/Asia 
Funds,  456 
Aythya  affinis,  465 
americana.  108,  109 
marila,  465—471 

Backus,  Leslie  K.,  see  Bowman,  Reed,  David  L.  Leon- 
ard, Jr., , and  Allison  R.  Mains 

badger,  see  Taxidea  tcvca 

Baeolophus  hicolor,  92,  94,  139-143,  373,  554,  555, 
558 

inornatus.  25 1 , 374 
spp.,  368 

wollweheri,  373,  374 

Baicich,  Paul  J.,  and  Colin  J.  O.  Harrison,  A guide  to 
the  nests,  eggs,  and  nestlings  of  North  Ameri- 
can birds,  reviewed,  295—296 
Bananaquit,  see  Coereha  fiaveola 
banding 

returns  of  Vireo  griseus.  46-55 
Banks,  Joshua  B.,  see  Breitwisch,  Randall,  Amy  J. 
Schilling,  and 

Banks,  Richard  C.,  see  Dove,  Carla  J.,  and 


Barau,  Armand,  see  Barre,  Nicolas, , and  Chris- 

tian Jouanin 

Barre,  Nicolas,  Armand  Barau,  and  Christian  Jouanin, 
Oiseaux  de  la  Reunion,  reviewed,  300-301 
Barrientos,  Claudia  I.,  see  Blem,  Charles  R.,  Leann  B. 
Blem,  and 

Barrow,  Mark  V.,  Jr.,  A passion  for  birds,  American 
ornithology  after  Audubon,  reviewed,  152-153 
Barrow,  Mark  V.,  Jr.,  review  by,  453-454 
Bartramia  longicauda,  263 
Baryphthengus  martii,  439 
ruficapillus,  221,  226 

Basileuterus  culicivorus.  216,  220,  401,  413 
flaveolus,  220,  228 
hypoleuciis.  216,  220,  228 
leucoblepharus,  401,  413 
Batara  cinerea,  410 

Bay,  Michael  D.,  The  type  B song  of  the  Northern 
Parula:  structure  and  geographic  variation 
along  proposed  sub-species  boundaries,  505- 
514 

Beaman,  Mark,  and  Steve  Madge,  The  handbook  of 
bird  identification  for  Europe  and  the  western 
Palearactic,  reviewed,  597-599 
Becard,  Gray-collared,  see  Pachyramphus  major 
Beehler,  Bruce  M.,  review  by,  596—597 
behavior 
aggressive 

female-female  in  Troglodytes  aedon,  130—132 
response  of  Poecile  atricapillus  and  Poecile  car- 
olinensis  to  calls,  363—367 
anti-predator 

“snorkeling”  by  Jacana  Jacana  chicks,  262-265 
courtship 

of  Theristicus  caudatus,  1 18—119 
interspecific 

with  foraging  Picoides  borealis.  346-353 
learning 

of  artificial  nest  location  by  predators,  536-540 
nesting 

of  Amazona  finschi,  488-493 
responses  of  Vireo  belli  to  brood  parasitism,  499- 
504 

parental 

of  bigamous  Cardinalis  cardinalis.  283-286 
postfiedging 

in  Aquila  chrysaetos.  472—477 
roosting 

in  pre-migratory  Progne  subis.  354-362 
siblicide 

role  of  food  at  Accipiter  gentilis  nests,  432-436 
Beletsky,  Les,  The  ecotravellers'  wildlife  guide  to  Be- 
lize and  northern  Guatemala,  reviewed,  594— 
595 

Beletsky,  Les,  The  ecotravellers’  wildlife  guide  to  Cos- 
ta Rica,  reviewed,  594-595 
Beletsky,  Les,  The  ecotravellers'  wildlife  guide  to 
tropical  Mexico,  reviewed,  594-595 
Bellbird,  Bare-throated,  see  Procnias  nudicollis 

Benz,  Brett  W.,  see  Watson,  David  M.,  and 

Berkelman,  James,  James  D.  Fraser,  and  Richard  T. 


612 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


Watson,  Madagascar  Fish-Eagle  prey  prefer- 
ence and  foraging  success,  15-21 
Biatas  nigropectiis,  410 

Bielefeldt,  John,  see  Trexel,  Dale  R.,  Robert  N.  Ro- 
senfield, , and  Eugene  A.  Jacobs 

biology 

of  nesting  Spiza  americana  and  Ammodramus  hen- 
slowii,  515-527 

Bishop,  K.  David,  see  Hill,  Norman  R,  and 

Bittern,  American,  see  Botaiirus  lentiginosus 
Least,  see  Ixobrychus  exilis 
Blackbird,  Brewer’s,  see  Euphagu.s  cyanocephalus 
Chopi,  see  Gnorimopsar  chopi 
Red-winged,  see  Agelaiu.s  phoeniceus 
Rusty,  see  Euphagiis  carolinus 
Tricolored,  see  Agelaius  tricolor 
Yellow-headed,  see  Xanthocephalus  .xanthocephalus 

Blair,  M.  J.,  see  Hagemeijer,  W.  J.  M.,  and 

Blarina  brevicauda,  239 

Blem,  Charles  R.,  Leann  B.  Blem,  and  Claudia  I.  Bar- 
rientos, Relationships  of  clutch  size  and  hatch- 
ing success  to  age  of  female  Prothonotary  War- 
blers, 577-580 

Blem,  Charles,  R.,  review  by,  294 

Blem,  Leann  B.,  see  Blem,  Charles  R.,  , and 

Claudia  I.  Barrientos 

Bloom,  Peter  H.,  see  Goldstein,  Michael  I.,  , 

Jose  H.  Sarasola,  and  Thomas  E.  Lacher 
Bluebird,  Eastern,  see  Sicilia  sialis 
Mountain,  see  Sialia  ciirrucoides 
Western,  see  Sialia  mexicana 
Bobolink,  see  Doiichonyx  oryzivoru.s 
Bobwhite,  Northern,  see  Colinus  virginianus 
Bock,  Walter  J.,  review  by,  152—153 

Bo^on,  Roberto,  see  Anjos,  Luiz  Dos,  and 

Bollinger,  Eric  K.,  see  Kershner,  Eric,  L.,  and 

Bombycilla  cedroriim.  84,  92,  94,  558 
Bonasa  iimbeiiu.s,  536 

Boobook,  Southern,  see  Ninox  novaeseelandiae 

Boone,  C.  A.,  see,  Hopp,  S.  L.,  A.  Kirby,  and 

Bosque,  Carlos,  and  Emilio  A.  Herrera,  “Snorkeling” 
by  the  chicks  of  the  Wattled  Jacana,  262-265 
Botauru.x  lentiginosus,  107,  108,  111 
pinnatus. 

Bowman,  Reed,  David  L.  Leonard,  Jr.,  Leslie  K.  Back- 
us, and  Allison  R.  Mains,  Interspecific  inter- 
actions with  foraging  Red-cockaded  Wood- 
peckers in  south-central  Florida,  346-353 
Brachyramphus  marmoratus,  257-261 
Brant,  Black,  see  Branta  bernicia  nigricans 
Branta  bernicia  bernicia.  166 
bernicia  nigricans,  167,  468 
canadensis,  1 08,  1 66—  1 80,  181,  468,  469 
canadensis  occidentalis,  167 
Breitwisch,  Randall,  Amy  J.  Schilling,  and  Joshua  B. 
Banks,  Parental  behavior  of  a bigamous  male 
Northern  Cardinal,  283-286 
Brewer,  Gwenda  L.,  reviews  by,  447—448,  450—452 
Bristlefront,  Slaty,  see  Meridaxis  ater,  199 
brood  parasitism 

responses  of  Vireo  belli,  499—504 


Bromley,  Robert  G.,  Carriere,  Suzanne,  , and 

Gilles  Guathier 
Brotogeris  chiriri,  225 
tirica,  408 

Brown,  Charles  R.,  Swallow  summer,  reviewed,  294 
Bubo  virginianus,  218,  226,  274,  529,  530 
Bubuicus  ibis,  225,  433 

Bugden,  Shawn  C.,  and  Roger  M.  Evans,  The  devel- 
opment of  a vocal  thermoregulatory  response 
to  temperature  in  embryos  of  the  domestic 
chicken,  188-194 

Bunting,  Indigo,  see  Passerina  cyanea 
Lark,  see  Calaniospiza  melanocorys 
Burhans,  Dirk  E.,  and  Frank  R.  Thompson  III,  Habitat 
patch  size  and  nesting  success  of  Yellow- 
breasted Chats,  210-215 
Burhinus  spp.,  344 

Burr,  Timothy,  see  Reed,  J.  Michael,  Elizabeth  M. 
Gray,  Dianne  Lewis,  Lewis  W.  Oring,  Richard 

Coleman, , and  Peter  Luscomb 

Burtt,  Edward  H.,  Jr.,  and  W.  Herbert  Wilson,  Jr.,  A 
survey  of  undergraduate  ornithology  courses  in 
North  America,  287—293 
Buteo  aibicaudatus,  407 
buteo  vuipinus,  182 
jamaicensis,  558 
lagopus,  558 
lineatus,  89-99,  348 
magnirostris,  225,  407 
piatypterus,  230-231,  254 
polyosoma,  530 
spp.,  218 

swainsoni,  428—432,  434 
Buteogallus  meridionaUs,  218,  225 
urubitinga,  218 

Butler,  Robert  W.,  The  Great  Blue  Heron:  a natural 
history  and  ecology  of  a seashore  sentinel,  re- 
viewed, 445 
Butorides  striatus,  225 
Buttonquail,  see  Tumix  sp. 

Buzzard,  Steppe,  see  Buteo  buteo  vuipinus 
Buzzard-Eagle,  Gery,  see  Geranoaetus  meianoieucus 

Cacicus  chrysopteriis,  218,  228,  413 
haemorrhous,  222,  413 
meianicterus,  84 

Cacique,  Golden-winged,  see  Cacicus  chrysopterus 
Red-rumped,  see  Cacicus  haemorrhous 
Yellow-winged,  see  Cacicus  meianicterus 
Cadman,  Michael  D.,  see  Friesen,  Lyle  E.,  Valerie  E. 

Wyatt,  and 

Cairina  moschata,  225 
Calaniospiza  melanocorys,  416 
Calcarius  mccownii,  416 
ornatus,  389-396 
Calidris  fuscicollis,  225,  263 
himantopus,  263 
Callipepla  californica,  530 
Campephilus  melanoleucos,  226 
Camptostoma  ohsoletum,  226,  412 
Campylopterus  curvipennis,  232 


INDEX  TO  VOLUME  I I I 


613 


excellens.  229,  232,  234 
heinileucurits,  232 

Campylorhynchiis  brunneicapillus,  1 29 
Ccinis  familiaris,  56 1 
Caprimulgus  anthonyi,  344 
candiciins.  222,  343,  344 
carolinensis,  344 
longirostris,  530 
nigrescens,  340,  344 
pan’ulus,  218,  226,  344 
rufus.  218,  219,  226 
vociferus,  229,  231 

Caracara,  Chimango,  see  Mitvago  chimango 

Crested,  see  Caracara  plancus  [Caracara  lutosus, 
Caracara  cheriway,  Polyborus  tharus] 
Mountain,  see  Phalcoboenus  megalopterus 
Caracara  plancus,  225,  [Caracara  lutosus,  Caracara 
cheriway,  Polyborus  tharus]  330-339,  437 
Carhonelle,  Montserrat,  see  Johnsgard,  Paul  A.,  and 


Cardinal,  Northern,  see  Cardinalis  cardinalis 
Red-crested,  see  Paroaria  coronata 
Cardinalis  cardinalis,  76-83,  137,  141,  142,  283-286, 
558 

Cardisoma  guanhani,  561—564 
Carduelis  magellanica,  228,  402,  414 
pinus,  558 

tristis,  93,  105,  108,  111,  558 
Cariatna  cristata,  225 
Carpodacus  mexicanus,  84—88 
purpureas,  93,  94,  558 

Carriere,  Suzanne,  Robert  G.  Bromley,  and  Gilles 
Guathier,  Comparative  spring  habitat  and  food 
use  by  two  Arctic  nesting  geese,  166-180 

Carroll,  Adele  E,  see  Gilbert,  William  M.,  and 

Casiornis  rufa,  227 
Casmerodius  albus,  530 

Cassidy,  Kelly  M.,  see  Smith,  Michael  R.,  Philip  W. 
Mattocks,  Jr.,  and 

Castrale,  John  S.,  Edward  M.  Hopkins,  and  Charles  E. 
Keller,  Atlas  of  breeding  birds  of  Indiana,  re- 
viewed, 294-295 
cat,  feral,  see  Felis  catus 
Catbird,  Gray,  see  Dumetella  carolinensis 
Cathartes  aura,  225,  407 
Catharus  fuscescens,  216,  220,  227,  575-576 
guttatus,  92,  94,  558 
ustulatus,  92,  94,  95 
Catherpes  mexicanus,  129,  199 
Catoptrophorus  semipalnuilus,  263,  561 
Cebus  capucinus,  125 
Celeiis  lugubris,  226 
censusing  methods 

temporal  differences  in  point  counts,  139-143 
Cepphus  columba,  257—261 
Cercomacra  carbonaria,  207 
nigricans,  207 
serx’a,  207 
tyrannina,  207 

Certhia  aniericana,  89—99,  245,  246,  558 
Chaetura  meridioncdis,  226 


pelagica,  23  I 
.spp.,  35 
vauxi,  229,  23 1 
Chamaeza  campanisoma,  4 1 0 
ruficauda,  410 

Charadrius  alexandrinus,  56 1 
melodus,  121-123,  321-329 
spp.,  344 

vociferus,  424—426,  576 
wilsonia,  561 

Chat,  Yellow-breasted,  see  Icteria  virens 
Chelydra  serpentina,  115 
Chencaerulescens  caerulescens,  465,  468 
Chickadee,  Black-capped,  see  Poecile  atricapillus 
Carolina,  see  Poecile  carolinensis 
Mountain,  see  Poecile  [Paru,s]  gambeli 
chicken,  see  Callus  gallus 
Chilia  melanura,  530 

Chinn,  Robert  W.,  see  Smith,  Kimberly  G.,  W.  Marvin 
Davis,  Thomas  E.  Kienzle,  William  Post,  and 


chipmunk,  eastern,  see  Tamias  striatus 
Merriam’s,  see  Eutamias  merriami 
Chiroxiphia  caudata,  411 
Chlidonias  niger,  108,  109 
spp.,  560 

Chloroceryle  aenea,  232 
americana,  226 
inda,  226 

Chlorostilbon  aureoventris,  409 
Chordeiles  gundlachii,  311,  561 
minor,  226,  340 
rapes tris,  343,  344 

Chuck-will’s-widow,  see  Caprimulgus  carolinensis 
Circus  cyaneus,  108,  109,  274,  558 
Cistothorus  palustris,  105-1 14,  254 
platensis,  100-104,  105-114,558 
Claravis  pretiosa,  225 

Clark,  William  S.,  see  Gorney,  Edna, , and  Yor- 

an  Yom-Tov 

Clethrionomys  gapperi,  236,  239 
Clibanornis  dendrocolaptoides,  410 
Cnemotriccus  fusctus,  227,  41  1 
Coccyzus  americanus,  141,  408 
melacoryphus,  226 
minor,  231,  310 

Cock-of-Rock,  Guianan,  see  Rupicola  rupicola 
Coereba  flaveola,  3 1 1 

Colaptes  auratus,  92,  108,  110,  141,  142,  348,  554, 
558 

campestris,  226,  409 
melanochloros,  409 
pitius,  530 

Coleman,  Richard,  see  Reed,  J.  Michael,  Elizabeth  M. 

Gray,  Dianne  Lewis,  Lewis  W.  Oring, , 

Timothy  Burr,  and  Peter  Luscomb 
Colibri  serrirostris,  409 

Colinus  virginianus,  236,  326,  541,  542,  545,  554, 
555,  558 

Collazo,  Jaime  A.,  see  Saracco,  James  E,  and 

Colonia  colonus,  4 1 1 


614 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


colonization 

of  Lake  Gatun,  Panama,  by  Rostrhamus  sociabilis, 
265-268 

Colorhamphus  parx’irostris,  530 
Coluber  constrictor  flaviventris,  524 
Cohirnba  araiicana,  530 
cayennensis,  225,  408 
leucocephala,  310 
livia,  86,  289,  558 
maculosa,  408 
picazuro,  408 
plumbea,  408 
speciosa,  408 
Columbina 
passe rina,  561 
picui,  225 
squarnmata,  225 
talpacoti,  225,  408 
communication 

visually  in  Caprimulgus  riificollis,  340—345 
community 

in  old-growth  conifer-hardwood  forest,  89-99 
response  to  1997-1998  El  Nino  in  Chile,  527-535 
response  to  weather,  photoperiod,  and  year,  550-558 
condition 

among  sex/age  classes  in  migrant  Regulus  calen- 
dula. 61-69 

of  bait  trapped  Accipiter  nisus  and  Accipiter  brevi- 
pes,  181-187 

Conirostrum  speciosum,  228 

Conner,  Richard  N.,  D.  Craig  Rudolph,  Richard  R. 
Schaefer,  Daniel  Saenz,  and  Clifford  E.  Shack- 
elford, Relationships  among  Red-cockaded 
Woodpecker  group  density,  nestling  provision- 
ing rates,  and  habitat,  494-498 
Conopophaga  lineata,  410 
conservation 

impact  of  wind  turbines  on  upland  nesting  birds, 
100-104 

Contopus  cinereus,  221,  411 
virens,  92,  108,  110 
Coot,  American,  see  Fulica  americana 
Coragyps  atratus,  225,  407 

Corx’us  brachyrhynchos,  89—99,  276—278,  538,  543, 
544,  546,  554,  558 
cora.x,  86,  92,  94,  1 15,  245,  276-278 
ossifragus,  543,  544 
Coryphospingus  cucullatus,  228 
Corylhopis  delandi,  226 

Coturni.x  japonica,  236—242,  289,  326,  415—420,  433, 
539,  547 

sp.,  236-242,  332,  323 
Coulter,  Malcolm  C.,  review  by,  593-594 
Cowbird,  Bay-winged,  see  Molothrus  badius 
Bronzed,  see  Molothrus  aeneus 
Brown-headed,  see  Molothrus  ater 
Giant,  see  Scaphidura  oryzivora 
Screaming,  see  Molothrus  rufoaxillaris 
Shiny,  see  Molothrus  bonariensis 
Cox,  Robert  R.,  Jr.,  see  Whitt,  Michael  B.,  Harold  H. 
Prince,  and  

crab,  giant  white  land,  see  Cardisoma  guanhani 


Crake,  Columbian,  see  Neocrex  colombianus 
Paint-billed,  see  Neocrex  erythrops 
Ruddy,  see  Laterallus  ruber 
Cranioleuca  obsoleta,  397-414 
pallida,  397-414 
Crax  fasciolata,  218,  225 
Creeper,  Brown,  see  Certhia  americana 
Crescentchest,  Collared,  see  Melanopareia  torquata 
Crossbill,  Red,  see  Loxia  curxnrostra 
Crotophaga  ani,  226,  311,  562 
major,  226,  264 

Crow,  American,  see  Corvus  brachyrhynchos 
Pish,  see  Corvus  ossifragus 
Red-ruffed  Pruit,  see  Pyroderus  scutatus 
Crypturellus  obsoletus,  407 
parvirostris,  225 
undulatus,  225 

Cuckoo,  Mangrove,  see  Coccyzus  minor 
Squirrel,  see  Piaya  cayana 
Yellow-billed,  see  Coccyzus  americanus 
Cuculus  canorus,  213 
Culicivora  caudacuta,  222 
Curaeus  curaeus,  530 
Curassow,  Bare-faced,  see  Crax  fasciolata 
Northern  Helmeted,  see  Pauxi  pauxi 
Curlew,  Long-billed,  Numenius  americanus 
Currasow,  Bare-faced,  see  Crax  fasciolata 
Custer,  Thomas  W.,  review  by, 

Cuthbert,  Prancesca  J.,  Brian  Scholtens,  Lauren  C. 
Wemmer,  and  Robyn  McLain,  Gizzard  contents 
of  Piping  Plover  chicks  in  northern  Michigan, 
121-123 

Cyanocitta  cristata,  92,  94,  96,  139—143,  276—278, 
348,  349,  538,  543,  544,  546,  558 
stelleri,  245 

Cyanocorax  caeruleus,  402,  412 
chrysops,  412 
cristatellus,  217,  227 
cyanomelas,  221 
yncas,  233 

Cyanoliseus  patagonus,  530 
Cyclarhis  gujanensis,  228,  401,  402,  413 
Cyprinus  carpio,  17,  19,  20 
Cypseloides  niger,  30-47 
Cypsnagra  hirundinacea,  228 

Dasypus  novemcinctus,  443-444 
Davis,  S.  K.,  D.  C.  Duncan,  and  M.  Skeel,  Distribution 
and  habitat  associations  of  three  endemic  grass- 
land songbirds  in  southern  Saskatchewan,  389— 
396 

Davis,  William  E.,  Jr.,  reviews  by,  294-295,  296-297, 
299-300,  445,  448-449 

Davis,  W.  Marvin,  see  Smith,  Kimberly  G.,  , 

Thomas  E.  Kienzle,  William  Post,  and  Robert 
W.  Chinn 

de  Reyna,  Luis  Arias,  see  Aragones,  Juan, , and 

Pilar  Recuerda 

Dean,  Kurtis  L.,  see  Swanson,  David  L.,  Eric  T.  Lik- 
nes,  and 


INDEX  TO  VOLUME  I I 1 


615 


Decker,  David  G.,  see  Avery,  Michael  L.,  Canie  L. 
Schreiber,  and 

DeGraat,  Richard  M.,  Thomas  J.  Maier,  and  Todd  K. 
Fuller,  Predation  of  small  eggs  in  artificial 
nests:  effects  of  nest  position,  edge,  and  poten- 
tial predator  abundance  in  extensive  forest, 
236-242 

Delichon  urhica,  579 
Dendragapus  obsciinis,  245 
Dendrohates  auratus,  439-440 
tricolor,  439 

Dendrocincla  anabatina,  443 
tiirdimi,  409 

Dendrocolaptes  platyrostris,  226,  409 
Dendrocygna  arboreo,  561 
bicolor,  3 1 1 

Dendroica  caerulescens,  74,  89-99,  213 
castanea,  93 

coronata,  93,  141,  233,  558 
discolor,  136 
fusca,  93,  94,  96,  505 
graciae,  5 1 1 
magnolia,  93,  94 
occidentalis,  95 
pensylvanica,  93 

petechia,  105-114,  136,  210,  290,  311,  381-388 
pinus,  93,  94,  346,  348,  349 
townsendi,  95 
virens,  93,  94,  505 
density 

of  Picoides  borealis  groups,  494-498 
DeSucre-Medrano,  Atahualpa,  see  Perez-Villafana, 
Monica,  Hector  Gomez  de  Silva  G.,  and 


Dewey,  Sarah  R.,  see  Estes,  Wendy  A.,  and 

Patricia  L.  Kennedy 
Dickcissel,  see  Spiza  americana 
Didelphis  marsupialis,  546 
diet 

fruit  sugar  preference  in  Carpodacus  mexicanus, 
84-88 

gizzard  contents  of  Charadrius  melodus  chicks, 
121-123 

in  spring  in  Anser  albifrons  frontalis  and  Branta 
canadensis,  1 66—  1 80 

in  Strix  occidentalis,  22-29 

of  Cypseloides  niger  in  California,  30—47 
Dillon,  M.  Beth,  see  Skagen,  Susan  K.,  Thomas  R. 

Stanley,  and 

dispersal 

invasion  of  southern  Florida  by  Asio  flammeus  from 
the  Antilles,  303-313 

of  Recun’irostra  americana  among  western  Great 
Basin  wetlands,  314—320 
distribution 

invasion  of  southern  Florida  by  Asio  flammeus  from 
the  Antilles,  303-313 

of  endemic  grassland  songbirds  in  southern  Sas- 
katchewan, 389-396 

of  Pauxi  pauxi  and  Aburria  aburri  in  Venezuela, 
564-569 


Diuca  diuca,  531,  533 
dog,  domestic,  see  Canis  familiaris 
Doliclionyx  oryzivorus,  100-104,  108 
Dove,  Carla  J.,  and  Richard  C.  Banks,  A taxonomic 
study  of  Crested  Caracaras  (Falconidae),  330- 
339 

Dove,  Mourning,  see  Zenaida  macroura 
Rock,  see  Columba  livia 
White-tipped,  see  Leptotila  verreauxi 
Zenaida,  see  Zenaida  aurita 
Drymophila  malura,  410 
Dryocopus  lineatus,  226,  403,  409 
pileatus,  92,  94,  245 

Duck,  American  Black,  see  Anas  rubripes 
Masked,  see  Nomonyx  dominicus 
Wood,  see  Aix  sponsa 

Dugger,  Bruce  D.,  see  Dugger,  Katie  M., , and 

Leigh  H.  Frederickson 

Dugger,  Katie  M.,  Bruce  D.  Dugger,  and  Leigh  H. 
Frederickson,  Annual  survival  rates  of  female 
Hooded  Mergansers  and  Wood  Ducks  in  south- 
eastern Missouri,  1-6 
Dumetella  carolinensis,  108,  254 

Duncan,  D.  C.,  see  Davis,  S.  K., , and  M.  Skeel 

Dysithamnus  mentalis,  222,  410 

Eagle,  Bald,  see  Haliaeetus  leucocephalus 

Crowned  Solitary,  see  Haryphaliaetus  coronatus 
Golden,  see  Aquila  chrysaetos 
Spanish  Imperial,  see  Aquila  adalberti 
White-tailed,  see  Haliaeetus  albicilla 
Edwards,  Ernest  Preston,  A field  guide  to  the  birds  of 
Mexico  and  adjacent  areas  (Belize,  Guatemala, 
and  El  Salvador),  reviewed,  588 

egg 

dimensions  and  shell  characteristics  in  Pufftnus  na- 
tivitatis,  421-422 

ejection  in  Sialia  currucoides,  440-442 
laying  time  of  Molothrus  aeneus,  137—139 
predation  in  artificial  nests,  236—242 
relation  of  age  in  female  Protonotaria  citrea  to 
hatching  success,  577—580 
variation  in  size  and  composition  in  Aytliya  marila. 
465-47 1 

Egret,  Cattle,  see  Bubulcus  ibis 
Egretta  thula,  530 

Eider,  Common,  see  Somateria  mollissima 
Eira  barbarra,  125 
Elaenia  albiceps,  121,  529,  530 
chiriquensis,  4 1 2 
flavoguster,  226 
mesoleuca,  401—402,  412 
obscura,  412 
pan’irostris,  227,  412 
Elaenia,  Gray,  see  Myiopagis  caniceps 
Elanus  caeruleus,  27 
leucurus,  225,  529,  530 
Electron  platyrhynchum,  440 
Emberizoides  herbicola,  228 
Embernagra  platensis,  228 
Empidonax  albigularis,  229,  232 


616 


THE  WILSON  BULLETIN  • VoL  III,  No.  4.  December  1999 


difficilis.  95 
euleri,  4 1 1 
flaviventris,  92,  95 
hammondii,  95 
minimus.  92,  94,  96 
oberholseri,  74 

traillii.  108,  585-588,  589-592 
traillii  e.xtimus,  573-575 
varius,  227 

vire.scens,  92,  94,  95,  542 
Empidonomus  varius,  41 1 
Engstrom,  R.  Todd,  review  by,  297-298 
Eremophila  cdpestris,  416,  558 
Erwin,  R.  Michael,  review  by,  454—456 
Escalante  R,  Patricia,  see  Winker,  Kevin,  Stefan  Ania- 

ga  Weiss,  Juana  Lourdes  Trejo  R,  and 

Esker,  Terry  L.,  see  Walk,  Jeffrey  W.,  , and 

Scott  A.  Simpson 
Esnx  masquinongy,  1 1 5 
niger,  1 15 

Estes,  Wendy  A.,  Sarah  R.  Dewey,  and  Patricia  L. 
Kennedy,  Siblicide  at  Northern  Goshawk  nests: 
does  food  play  a role?,  432-436 
Eucometis  penicillata,  228,  443 
Eumomota  superciliosa,  439,  440 
Euphagus  carolinus,  558 
cyanocephalus,  108 

Euphonia,  Antillean,  see  Euphonia  musica 
Chestnut-bellied,  see  Euphonia  pectoralis 
Euphonia  chalybea,  413 
chlorotica,  228 
musica,  413 
pectoralis,  413 

Euscarthmus  rufomarginatiis , 22 1 
Eutamias  merriami,  252 

Evans,  Michael  I.,  see  Tucker,  Graham  M.,  and 

Evans,  Roger  M.,  see  Bugden,  Shawn  C.,  and 

Exum,  Jay  H.,  see  Harper,  Craig  A.,  and 

Eaaborg,  John,  review  by,  147-148 

Eaaborg,  John,  see  Woodworth,  Bethany  L.,  , 

and  Wayne  J.  Arendt 
Ealco  elenorae,  185 
femoralis,  407,  529,  530 
mexicanus,  558 
peregrinus,  530 

sparx’erius,  225,  269-271,  311,  407,  434,  530,  558 
tinnunculus,  21 , 475 
Falcon,  Aplomado,  see  Ealco  femoralis 
Eleonora's,  see  Ealco  elenorae 
Peregrine,  see  Ealco  peregrinus 
Prairie,  see  Ealco  mexicanus 

Faucett,  Rob.  C.,  see  Robbins,  Mark  B.,  , and 

Nathan  H.  Rice 
Eelis  cams.  376 
Ficedula  hypoleuca.  579 

Finch,  Deborah,  see  Stole.son,  Scott  H.,  and  

Finch,  Deborah,  see  Yong,  Wang,  and  

Finch,  Diuca,  see  Diuca  diuca 

House,  see  Carpodacus  mexicanus 
Glive,  see  Eysurus  castaneiceps 
Purple,  see  Carpodacus  purpureus 


Sooty-faced,  see  Eysurus  crassirostris 
Zebra,  see  Poephila  gullata 
Finck,  Elmer  J.,  see  Stapanian,  Martin  A.,  Christopher 
C.  Smith,  and 

Fire-eye,  White-shouldered,  see  Pyriglena  leucoptera 
Firewood-gatherer,  see  Anumbius  annumbi 
Fish-Eagle,  African,  see  Haliaetus  vicifer 
Madagascar,  see  Haliaeetus  vociferoides 
fisher,  see  Maries  pennanti 
Fleming,  Robert  L.,  review  by  149-150 
Flicker,  Northern,  see  Colaptes  auratus 
Flint,  Paul  L.,  and  J.  Barry  Grand.  Patterns  of  variation 
in  size  and  composition  of  Greater  Scaup  eggs: 
are  they  related?,  465-47 1 
Fluvicolo  leucocephala,  227 
Flycatcher,  Acadian,  see  Empidonax  virescens 

Crowned  Slaty,  see  Griseotyrannus  aurantioatro- 
cristatus 

Dusky,  see  Empidonax  oberholseri 
Great  Crested,  see  Myiarchus  crinitus 
Hammond’s,  see  Empidonax  hammondii 
La  Sagra’s,  see  Myiarchus  sagrae 
Least,  see  Empidonax  minimus 
Matinan,  see  Muscicapa  sanfordi 
Ochre-bellied,  see  Mionectes  oleagineus 
Pacific  Slope,  see  Empidonax  difficilis 
Pied,  see  Ficedula  hypoleuca 
Sepia-capped,  see  Eeptopogon  amaurocephalus 
Southwestern  Willow,  see  Empidonax  traillii  e.xti- 
mus 

Sulphur-rumped,  see  Myiobius  sulphureipygius 
Tawny-chested,  see  Aphanotriccus  capitalis 
White-throated,  see  Empidonax  albigidaris 
Willow,  see  Empidonax  traillii 
Yellow-bellied,  see  Empidonax  flaviventris 
Yellow-olive,  see  Tolmomyias  sulphurescens 
Foliage-gleaner,  Chestnut-capped,  see  Hylocryptus 
rectirostris 

Ochre-breasted,  see  Philydor  lichtensteini 
Planalto  [Russet-mantled],  see  Philydor  dimidiatus 
food 

fRiit  sugar  preference  in  Carpodacus  mexicanus, 
84-88 

role  in  siblicide  at  Accipiter  gentilis  nests,  432—436 
spring  use  by  Anser  albifrons  frontalis  and  Branta 
canadensis,  1 66—  1 80 

foraging 

behavior 

cooperative  effort  in  Phalcoboenus  megalopterus, 
437-439 

of  Cypseloides  niger  in  California,  30-47 
of  Percnostola  caurensis  and  Myrmeciza  di.yunc- 
ta.  195-209 

possible  suet  cutting  by  Corvus  cora.x,  276—278 
Seiurus  aurocapillus  follows  Dasypus  novemcinc- 
tus,  434-444 

ecology 

of  Haliaeetus  vociferoides,  15—21 
Forest-Falcon,  Barred,  see  Micrastur  ruflcollis 
Eormicivora  rufa,  221 
Corpus  .xanthopterygius,  408 


INDEX  TO  VOLUME  1 1 1 


617 


Foster,  Janet,  Working  for  wildlife:  the  beginning  of 
preservation  in  Canada,  reviewed,  453-454 
fox,  red,  see  Vidpes  vulpes 

Franzreb,  Kathleen  E.,  Factors  that  influence  translo- 
cation success  in  the  Red-cockaded  Woodpeck- 
er, 38-45 

Fraser,  James  D.,  see  Berkelman,  James, , and 

Richard  T.  Watson 

Frederickson,  Leigh  H.,  see  Dugger,  Katie  M.,  Bruce 
D.  Dugger,  and 

Friesen,  Lyle  E.,  Valerie  E.  Wyatt,  and  Michael  D. 
Cadman,  Nest  reuse  by  Wood  Thrushes  and 
Rose-breasted  Grosbeaks,  132-133 
Friesen,  Lyle  E.,  Valerie  E.  Wyatt,  and  Michael  D. 
Cadman,  Pairing  success  of  Wood  Thrushes  in 
a fragmented  agricultural  landscape,  279-281 
frog,  black-and-green  poison  dart,  see  Dendrobates  au- 
ra tus 

phantasmal  poison  dart,  see  Dendrobates  tricolor 
Fidica  americana,  108,  469 

Fuller,  Todd  K.,  see  DeGraaf,  Richard  M.,  Thomas  J. 

Maier,  and 

Furnarius  rufus,  226,  402,  410 

Gallinago  gallinago,  231,  263,  344 
media,  344 
paragLiaiae,  225 
spp.,  344 

Gallinida  chloropus,  108 
Gallinule,  Purple,  see  Porphyrida  martinica 
Gallus  gallus,  188-194,  289 
sp.,  236-242 

Gaston,  Anthony  J.,  and  Ian  Jones,  The  auks,  re- 
viewed, 593—594 

Gauthier,  Gilles,  see  Carriere,  Suzanne,  Robert  G. 
Bromley,  and 

Gauthreaux,  Sydney,  Jr.,  see  Russell,  Kevin  R.,  and 


Gavia  immer,  115—116,  116-117 
geographic  variation 

of  type  B song  in  Panda  americana,  505-514 
Geositta  rufipennis,  530 
Geothlypis  aequinoctiaUs,  228,  413 
trichas,  93,  94,  100-104,  105-114 
Geotrygon  chrysia,  310,  311 
montana,  408 

Geranoaetus  melanoleucus,  530 

Gervais,  Jennifer  A.,  and  Daniel  K.  Ro.senberg,  West- 
ern Burrowing  Owls  in  California  produce  sec- 
ond broods  of  chicks,  569—57 1 
Gilbert,  William  M.,  and  Adele  E Carroll,  Singing  in 
a mated  female  Wilson’s  Warbler,  134—137 
Glaucidium  brasilianum,  218,  226,  408 
nanum,  530 

Glaucomys  sabrinus,  21 
sp.,  239 
volans,  38,  269 
Glossogobius  giurus,  19 
Glyphorynchus  spirurus,  232 
Gnorimopsar  cliopi,  220,  221,  228,  413 
Goldfinch,  American,  see  Carduelis  tristis 


Goldstein,  Michael  L,  Peter  H.  Bloom,  Jose  H.  Sara- 
sola,  and  Thomas  E.  Lacher,  Post-migration 
weight  gain  of  Swainson’s  Hawks  in  Argentina, 
428-432 

Gomez  de  Silva  G.,  Hector,  see  Perez- Vallifana,  Mon- 
ica,   , and  Atahualpa  DeSucre-Medrano 

Goose,  Canada,  see  Branta  canadensis 

Greater  White-fronted,  see  Anser  albifrons  frontalis 
Lesser  Snow,  see  Chen  caerulescens  caerulescens 
gopher,  northern  pocket,  see  Thymomas  bottae 
Gorney,  Edna,  William  S.  Clark,  and  Yoran  Yom-Tov, 
A test  of  the  condition-bias  hypothesis  yields 
different  results  for  two  species  of  sparrow- 
hawks  (Accipiter),  181-187 
Goshawk,  Northern,  see  Accipiter  gentilis 
Grackle,  Common,  see  Quiscalus  quiscula 

Grand,  J.  Barry,  see  Flint,  Paul  L.,  and 

grants 

see  awards  and  grants 
Grassquit,  Black-faced,  see  Tiaris  bicolor 

Gray,  Elizabeth  M.,  see  Reed,  J.  Michael.  , 

Dianne  Lewis,  Lewis  W.  Oring,  Richard  Cole- 
man, Timothy  Burr,  and  Peter  Luscomb 
Grebe,  Pied-billed,  see  Podilymbus  podiceps 

Greenberg,  Russell,  see  Schelhas,  John,  and 

Greenfield,  Paul  J.,  see  Krabbe,  Niels,  Morton  L.  Isler, 
Phyllis  R.  Isler,  Bret  M.  Whitney,  Jose  Alvarez 
A.,  and 

Griseotyrannus  aurantioatrocristatus,  218,  227 
Grosbeak,  Rose-breasted,  see  Pheucticus  ludovicianus 
Ground-Dove,  Common,  see  Columbina  passerina 
Ruddy,  see  Columbina  talpacoti 
Grouse,  Blue,  see  Dendragapus  obscurus 
Ruffed,  see  Bonasa  umbellus 
growth 

patterns  in  Himantopus  mexicanus  knudseni,  478- 
487 

Guan,  Wattled,  see  Aburria  aburri 
Gubernetes  yetapa,  227 
Guillemot,  Pigeon,  see  Cepphus  columba 
Guira  guira,  226,  408 
Gull,  Hening,  see  Larus  argentatus 
Laughing,  see  Larus  atricilla 
Ring-billed,  see  Larus  delawarensis 
Gutierrez,  R.  J.,  see  Smith,  Richard  B.,  M.  Zachariah 
Peery, , and  William  S.  Lahaye 

Habia  rubica,  413 
habitat 

association 

of  endemic  grassland  songbirds  in  southern  Sas- 
katchewan, 389-396 

breeding 

association  of  group  density  and  nestling  provi- 
sioning rates  in  Picoides  borealis,  494— 
498 

use  of  Ly thrum  salicaria,  105-114 
patch  size 

in  Icteria  virens,  210—215 


618 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


selection 

by  Setophaga  niticilla  in  aspen-dominated  forest 
fragments,  70-75 
use 

by  Nomonyx  domicus,  1 19-121 
in  spring  by  Anser  albifrons  frontalis  and  Branta 
canadensis,  166—180 

winter 

of  Sterna  aleutica,  559-560 
use  of  burned  and  unburned  coniferous  forests, 
243-250 

Hagemeijer,  W.  J.  M.,  and  M.  J.  Blair,  eds.  The  EBCC 
atlas  of  European  breeding  birds;  their  distri- 
bution and  abundance,  reviewed,  154-156 

Haig,  Susan  M.,  Plissner,  Jonathan  H.,  , and 

Lewis  W.  Oring 
Haliaeetus  albicilla,  15 

leucocephalus,  15-21,  115-116,260 
vocifer,  19 
vociferoides,  1 5—2 1 
Hall,  George  A.,  review  by,  452-453 
Haney,  J.  Christopher,  Hierarchical  comparisons  of 
breeding  birds  in  old-growth  conifer-hardwood 
forest  on  the  Appalachian  plateau,  89-99 
Haplospiza  unicolor,  414 
Harpagus  bidentatus,  443 

Harper,  Craig  A.,  and  Jay  H.  Exum,  Wild  Turkeys  (Me- 
leagris  gallopavo)  renest  after  successful  hatch, 
426-427 

Harpyhaliaetus  coronatus,  218,  225 
Harrier,  Northern,  see  Circus  cyaneus 

Harrison,  Colin  J.  O.,  see  Baicich,  Paul  J.,  and 

Hawk,  Broad-winged,  see  Buteo  platypterus 
Cooper’s,  see  Accipiter  cooperii 
Great  Black,  see  Buteogallus  urubitinga 
Harris’s,  see  Parabuteo  unicinctus 
Red-shouldered,  see  Buteo  lineatus 
Red-tailed,  see  Buteo  jamaicensis 
Roadside,  see  Buteo  magnirostris 
Rough-legged,  see  Buteo  lagopus 
Savannah,  see  Buteogallus  meridionalis 
Sharp-shinned,  see  Accipiter  striatus 
Swainson’s,  see  Buteo  swainsoni 
Hawk-Eagle,  Ornate,  see  Spizaetus  ornatus 
Hawk-Owl,  Bismarck,  see  Ninox  variegata 
Brown,  see  Ninox  scutulata 
Cinnabar,  see  Ninox  ios 
Manus,  Ninox  meeki 
Moluccan,  see  Ninox  squamipila 
New  Britain,  see  Ninox  odiosa 
Ochre-bellied,  see  Ninox  ochracea  [perx’ersa] 
Philippine,  see  Ninox  philippensis 
Solomons,  see  Ninox  jaccpdnoti 
Speckled,  see  Ninox  punctulata 
Sumba,  see  Ninox  rufolfi 

Heinrich,  Bernd,  Planning  to  facilitate  caching;  possi- 
ble suet  cutting  by  a Common  Raven,  276-278 
Heliobletus  contaniinatus,  4 1 0 
Hemithraupis  guira,  228 
ruficapilla.  4 1 3 

Hemitriccus  inargaritaceiventer,  227 


obsoletus,  412 

Hendrickson,  Herbert  T,  reviews  by,  148,  151 
Herbert,  Percy  N.,  Evidence  of  egg  ejection  in  Moun- 
tain Bluebirds,  440-442 
Heron,  Brown,  see  Nycticorax  nycticorax 
Herpetotheres  cachinnans,  225,  407 
Herpsilochmus  atricapillus,  226 

Herrera,  Emilio  A.,  see  Bosque,  Carlos,  and 

Heterospizias  meridionalis,  407 
Heterotis  niloticus,  17,  19 

Higgins,  Kenneth  E,  see  Leddy,  Krecia  L.,  , 

and  David  E.  Naugle 

Hill,  Norman  R,  and  K.  David  Bishop,  Possible  winter 
quarters  of  the  Aleutian  Tern?,  559-560 
Himantopus  himantopus  leucocephalus,  485 
mexicanus,  478,  479 
me.xicanus  knudseni,  478-487 
novaezealandiae,  485 
Hirundinea  ferruginea,  411 
Hirundo  fulva,  3 1 1 

Hoag,  David  J.,  Hybridization  between  Clay-colored 
Sparrow  and  Field  Sparrow  in  northern  Ver- 
mont, 581-584 

Hoffman,  Wayne,  Glen  E.  Woolfenden,  and  P.  William 
Smith,  Antillean  Short-eared  Owls  invade 
southern  Florida,  303-313 

Holmes,  Derek,  and  Karen  Phillipps,  The  birds  of  Su- 
lawesi, reviewed,  151 

Hopkins,  Edward  M.,  see  Castrale,  John  S.,  , 

and  Charles  E.  Keller 

Hopp,  S.  L.,  A.  Kirby,  and  C.  A.  Boone,  Banding  re- 
turns, arrival  pattern,  and  site-fidelity  of  White- 
eyed Vireos,  46—55 

House-Martin,  Common,  see  Delichon  urbica 
Houston,  C.  Stuart,  Barred  Owl  nest  in  attic  of  shed, 
272-273 

Howell,  Steve  N.  G.,  A bird-finding  guide  to  Mexico, 
reviewed,  595 

Hubbard,  John  P.  A critique  of  Wang  Wong  and 
Finch’s  field  identification  of  Willow  Flycatch- 
er subspecies  in  New  Mexico,  585-588 
human  impact 

influence  of  human  scent  on  mammalian  predators, 
415-420 

Hummingbird,  Giant,  see  Patagona  gigas 
Ruby-throated,  see  Archilochus  colubris 
hybridization 

between  Spizella  pallida  and  Spizella  pusilla  in 
northern  Vermont,  581-584 
Hydrophasianus  chirurgus,  264 
Hylocharis  chrysura,  226 
sapphirina,  222 

Hylocichla  mustelina,  92,  94,  132-133,  141,  279-281, 
379,  538,  539 

Hylocryptus  recti rostris,  222 
Hylopezus  nattereri,  410 
Hylophilus  poicilotis,  413 
Hylorchilus  sumichrasti,  1 28- 1 30 
Hymenops  perspicillatus,  227 


INDEX  TO  VOLUME  1 I 1 


619 


Hypocneinoides  melanopo^on,  205 
spp.,  207 

Ibis,  Buff-necked,  see  Theristicus  caudalus 
Icteria  virens,  84,  210—215 
Icterus  hidlockii,  251—256 
cayanensis,  228 
gal  hula,  1 08 
Ictinia  pluiuhea,  225 
Idioptilon  nidipendulum,  412 
Inezia  inornata,  I'll 

Isler,  Morton  L.,  see  Krabbe,  Niels, , Phyllis  R. 

Isler,  Bret  M.  Whitney,  Jose  Alvarez  A.,  and 
Paul  J.  Greenfield 

Isler,  Phyllis  R.,  see  Krabbe,  Niels,  Morton  L.  Isler, 

, Bret  M.  Whitney,  Jose  Alvarez  A.,  and 

Paul  J.  Greenfield 
Ixobrychus  exilis,  105—1  14 
Ixoreus  naevius,  245 

Jacana,  African,  see  Actophilornis  africanus 
Lesser,  see  Microparra  capensis 
Pheasant-tailed,  see  Hydrophasianus  chirurgus 
Wattled,  see  Jacana  jacana 
Jacana jacana,  262—265 

Jacobs,  Eugene  A.,  see  Trexel,  Dale  R.,  Robert  N.  Ro- 

senfield,  John  Bielefeldt,  and 

Jaeger,  Parasitic,  see  Stercorarius  parasiticus 
Jahn,  Olaf,  Maria  Eugenia  Jara  Viteri,  and  Karl-L. 
Schuchmann,  Connecticut  Warbler,  a North 
American  migrant  new  to  Ecuador,  281-282 
Jaksic,  Eabian  M.,  and  Ivan  Lazo,  Response  of  a bird 
assemblage  in  semiarid  Chile  to  the  1997—1998 
El  Nino,  527-535 
Jay,  Blue,  see  Cyanocitta  cristata 

Bushy-crested,  see  Cyanocorax  melanocyaneus 
Curl-crested,  see  Cyanocorax  cristatellus 
Gray,  see  Perisoreus  canadensis 
Green,  see  Cyanocorax  yncas 
Stellar’s,  see  Cyanocitta  stelleri 
Johnsgard,  Paul  A.,  and  Montserrat  Carbonell,  Ruddy 
ducks  and  other  stiff-tails:  their  behavior  and 
biology,  reviewed,  450—452 

Jones,  Ian,  see  Gaston,  Anthony  J.,  and 

Jones,  Jason,  Cooperative  foraging  in  the  Mountain 
Caracara  in  Peru,  437—439 

Jouanin,  Christian,  see  Barre,  Nicolas,  Armand  Barau, 
and 

Junco,  Dark-eyed,  see  Junco  hyemalis 
Junco  hyemalis,  93,  94,  558 

Keehn,  Shannon,  see  Sodhi,  Navjot  S.,  Cynthia  A. 
Paszkowski,  and 

Keith,  Allan  R.,  The  birds  of  St.Lucia,  West  Indies, 
reviewed,  147—148 
Keith,  Allan  R.,  review  by,  597-599 
Keller,  Charles  E.,  see  Castrale,  John  S.,  Edward  M. 
Hopkins,  and 

Kennedy,  Patricia  L.,  see  Estes,  Wendy  A.,  Sarah  R. 
Dewey,  and 


Kennedy,  Patricia  L.,  see  O’Toole,  Laura  T,  , 

Richard  L.  Knight,  and  Lowell  C.  McEwen 
Kershner,  Eric,  L.,  and  Eric  K.  Bollinger,  Aggressive 
response  of  chickadees  towards  Black-capped 
and  Carolina  calls  in  central  Illinois,  363-367 
Kestrel,  American,  see  Falco  spar\>erius 
Common,  see  Falco  tinnunculus 
Kienzle,  Thomas  E.,  see  Smith,  Kimberly  G.,  W.  Mar- 
vin Davis,  , William  Post,  and  Robert 

W.  Chinn 

Killdeer,  see  Charadrius  vociferus 
King,  Ben  E,  Checklist  of  the  birds  of  Eurasia,  re- 
viewed, 596-597 

King,  David  L,  Mortality  of  an  adult  Veery  incurred 
during  the  defense  of  nestlings,  575-576 
Kingbird,  Eastern,  see  Tyrannus  tyrannus 
Gray,  see  Tyrannus  dominicensis 
Tropical,  see  Tyrannus  melancholicus 
Western,  see  Tyrannus  verticalis 
Kingfisher,  Pygmy,  see  Chloroceryle  aenea 
Kinglet,  Golden-crowned,  see  Regulus  satrapa 
Ruby-crowned,  see  Regulus  calendula 
kingsnake,  prairie,  see  Lampropeltis  calligasster  cal- 
ligaster 

Kirby,  A.,  see  Hopp,  S.  L., , and  C.  A.  Boone 

Kiskadee,  Great,  see  Pitangus  sulphuratus 
Kite,  Black,  see  Milvus  migrans 

Black-shouldered,  see  Elanus  caeruleus 
Double-toothed,  see  Harpagus  hidentatus 
Snail,  see  Rostrhamus  sociabilis 
White-tailed,  see  Elanus  leucurus 
Knight,  Richard  L.,  see  O’Toole,  Laura  T,  Patricia  L. 

Kennedy, , and  Lowell  C.  McEwen 

Knipolegus  cyanirostris,  41 1 
lophotes,  41 1 
nigerrimus,  41 1 

Krabbe,  Niels,  Morton  L.  Isler,  Phyllis  R.  Isler,  Bret 
M.  Whitney,  Jose  Alvarez  A.,  and  Paul  J. 
Greenfield,  A new  species  in  the  Myrmotherula 
haematonota  superspecies  (Aves;  Thamnophil- 
idae)  from  the  western  Amazonian  lowlands  of 
Ecuador  and  Peru,  157—165 
Kreisel,  Karen  J.,  and  Steven  J.  Stein,  Bird  use  of 
burned  and  unburned  coniferous  forests  during 
winter,  243-250 

Kricher,  John  C.,  reviews  by,  148-149,  151,  594-595 
Kricher,  John,  A Neotropical  companion:  an  introduc- 
tion to  the  animals,  plants,  and  ecosystems  of 
the  New  World  tropics,  reviewed,  146-147 
Kuletz,  Katherine  J.,  and  John  E Piatt,  Juvenile  Mar- 
bled Murrelet  nurseries  and  the  productivity  in- 
dex, 257-261 

Lacher,  Thomas  E.,  see  Goldstein,  Michael  L,  Peter  H. 

Bloom,  Jose  H.  Sarasola,  and 

Lagopus  lagopus,  274 

Lahaye,  William  S.,  see  Smith,  Richard  B.,  M.  Za- 

chariah  Peery,  R.  J.  Gutierrez,  and 

Lampropeltis  calligaster  calligaster,  524 
Lanins  excubitor,  285 

ludovicianus,  285,  348,  558 


620 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


Lanner,  Ronald  M.,  Made  for  each  other:  a symbiosis 
of  birds  and  pines,  reviewed  297—298 
Lapwing,  Northern,  see  Vanellus  vanellu.s 
Lark,  Homed,  see  Eremophila  alpestri.s 
Larsson,  Hans,  see  Olsen,  Klaus  Mailing  Olsen  and 

Loru.s  argentatus,  115,  192 
atricilla,  561 
dekiwarensis,  192 
Liitercillus  ruber.  231 
spp.,  423 

luithrotricciis  euleri.  Til 

Laso,  Ivan,  see  Jaksic,  Fabian  M.,  and 

leaftosser,  see  Sclerurus  spp. 

Leddy,  Krecia  L.,  Kenneth  E Higgins,  and  David  E. 
Naugle,  Effect  of  wind  turbines  on  upland  nest- 
ing birds  in  conservation  reserve  program 
grasslands,  100-104 
Legatus  leucophaius,  227 

Lein,  M.  Ross,  see  Wiebe,  Myra  O.,  and 

Leistes  siiperciliaris,  228 

Lemon,  R.  E.,  see  Lozano,  G.  A.,  and 

Leonard,  David  L.,  Jr.,  see  Bowman,  Reed,  , 

Leslie  K.  Backus,  and  Allison  R.  Mains 
Lepiclocolaptes  cmgustirostris,  221,  226,  409 
fu.'icatus,  409 
squamatus,  397—414 
Leptastheniira  aegithaloides,  530 
.setaria,  410 
struikcita,  410 
Leptodon  cayanensis,  225 
Leptopogon  cimciurocephalu.<;.  111,  232,  412 
Leptotila  rufaxilla,  225,  408 
verreauxi,  225,  402,  408 
Leucochloris  albicollis,  409 
Leuconerpes  Candidas,  409 

Levey,  Douglas  J.,  Foraging  Ovenbird  follows  arma- 
dillo, 443-444 

Lewis,  Dianne,  see  Reed,  J.  Michael,  Elizabeth  M. 
Gray,  , Lewis  W.  Oring,  Richard  Cole- 

man, Timothy  Burr,  and  Peter  Luscomb 

Liknes,  Eric  T,  see  Swanson,  David  L.,  , and 

Kurtis  L.  Dean 

Limnothlypis  swainsonii,  229,  233,  234 
lizard,  tegu,  see  Tupinambis  sp. 

Lochmias  nematiira,  410 

Longspur,  Chestnut-collared,  see  Calcarius  ornatiis 
McCown’s,  see  Calcarius  mccownii 
Loon,  Common,  see  Gavia  immer 
loosetrife,  purple,  see  Lythrum  salicaria 
Lophodytes  cucullalus,  1—6 
Lowther.  Peter  E.,  review  by,  146 
Loxia  curv’iroslra,  245 

Lozano,  G.  A.,  and  R.  E.  Lemon,  Effects  of  prior  res- 
idence and  age  on  breeding  performance  in 
Yellow  Warblers,  381-388 
l.urocalis  semitorquatus,  218,  226,  408 
Luscomb,  Peter,  see  Reed,  J.  Michael,  Elizabeth  M. 
Gray,  Dianne  Lewis,  Lewis  W.  Oring,  Richard 

Coleman,  Timothy  Burr,  and 

Lysurus  caslaneiceps,  1 26 


crassirostris,  1 24—  1 28 
Lythrus  salicaria,  105—114 

Machetornis  rixosus,  227,  4 1 1 

Mack,  Tara,  see  Paruk,  James  D.,  Dean  Seanfield,  and 


Mackenziaena  leachii,  410 

Madge,  Steve,  see  Beaman,  Mark,  and 

Magpie,  Black-billed,  see  Pica  pica 

Mahan,  Carolyn  G.,  see  Yahner,  Richard  H.,  and 


Maier,  Thomas  J.,  see  DeGraaf,  Richard  M.,  , 

and  Todd  K.  Fuller 

Mains,  Allison  R.,  see  Bowman,  Reed,  David  L.  Leon- 
ard, Jr.,  Leslie  K.  Backus,  and 

Mallard,  see  Anas  platyrhynchos 
management 

response  of  Sitta  pusilla  to  pine  plantation  thinning, 
56-60 

use  of  burned  and  unburned  coniferous  forests  dur- 
ing winter,  243-250 

Manakin,  Helmeted,  see  Antilophia  galeata 
Margarops  fuscatus,  3 1 1 

Marin,  Manuel,  Food,  foraging,  and  timing  of  breeding 
of  the  Black  Swift  in  California,  30-47 
Marks,  Jeffrey  S.,  and  Alison  E.  H.  Perkins,  Double 
brooding  in  the  Long-eared  Owl,  273-276 
Martes  pennanti,  240 

Martin,  Brown-chested,  see  Phaeoprogne  tapera 
Crag,  see  Ptyonoprogne  rupestris 
Purple,  see  Progne  subis 

Marzluff,  John  M.  and  Rex  Sallabanks,  eds..  Avian 
conservation,  reviewed,  449-450 
Master,  Terry  L.,  Predation  by  Rufous  Motmot  on 
black-and-green  poison  dart  frog,  439—440 
mate  choice 

color  preference  in  female  Cardinalis  cardinalis, 
76-83 

Matthynsen,  Erik,  The  nuthatches,  reviewed,  452—453 

Mattocks,  Philip  W.,  Jr.,  see  Smith,  Michael  R., , 

and  Kelly  M.  Cassidy 

Mayer,  Paul  M.,  see  Staus,  Nancy  L.,  and 

McEwen,  Lowell  C.,  see  O’Toole,  Laura  T,  Patricia  L. 

Kennedy,  Richard  L.  Knight,  and 

McLain,  Robyn,  see  Cuthbert,  Francesca  J.,  Brian 

Scholtens,  Lauren  C.  Wemmer,  and 

Meadowlark,  Eastern,  see  Sturnella  magna 
Western,  see  Sturnella  neglecta 
Mearns,  Barbara  and  Richard  Mearns,  The  bird  collec- 
tors, reviewed,  144—146 

Mearns,  Richard,  see  Mearns,  Barbara  and 

Megalops  cyprinoides,  19 
Megarynchus  pitangua.  111,  4 1 1 
Melanerpes  Candidas,  226 

carol  inns,  139-143,  346-353,  558 
erythrocephalus,  141,  142,  348,  551,  558 
flavifrons,  409 
fonnicivorus,  252,  254 
Melanopareia  torquata,  222 
Meleagris  gallopavo,  426-427,  555,  558 
spp.,  17 


INDEX  TO  VOLUME  1 1 I 


621 


Melospiza  i>eori>iana,  105-1 14 
lincolnii.  234,  554,  555,  558 
melodici.  108,  285,  558 
Mephifis  mephitis,  416 

Merganser,  Common,  see  Mergus  merganser 
Hooded,  see  Lophodytes  ciicidlatus 
Mergus  merganser,  92 
Meridaxis  ater,  199 
methods 

critique  of  field-identification  of  Empidonax  trailii 
subspecies,  585-588 
Response,  589-592 

temporal  differences  in  point  counts,  139-143 
Metriopelia  melanoptera,  530 
Micrastur  ruficollis,  225,  407 
semitorquatus,  125 
Microparra  capensis,  264 
Microtiis  pennsylvanicus,  239 
spp.,  274 
migration 

arrival  patterns  of  Vireo  griseiis,  46-55 
differences  in  timing  among  sex/age  classes  of  Reg- 
ains calendula,  61—69 
Milvago  chimachima,  225,  407 
chimango,  530 

Milvus  migrans,  16,  17,  18,  475 
Mimus  gundlachii,  311 
polyglottos,  285,  348 
saturninus,  227 
thenca,  530 
triurus,  228 

Mionectes  oleagineus,  232,  234 
Mniotilta  varia,  93 

Mockingbird,  Bahama,  see  Mimus  gundlachii 
Northern,  see  Mimus  polyglottos 
Molothrus  aeneus,  137-139 

ater,  72,  93,  94,  96,  100-104,  108,  110,  133,  137- 
139,  210,  252,  289,  440,  441,  499,  502,  515- 
527,  558,  577 
badius,  137-139,  220 

honariensis,  137-139,  220,  221,  228,  311,  376,  379, 
413,  441 

rufoaxillaris,  138,  220,  221,  228 
Momotus  momota,  219,  226,  439,  440 
monkey,  see  Cebus  capucinus,  Ateles  geoffroyi,  Al- 
ouatta  palliate 
squirrel,  see  Saimiri  oerstedi 

Monson,  Gale,  see  Russell,  Stephen  M.,  and 

Moorhen,  Common,  see  Gallinula  chloropus 
Morris,  Sara  R.,  review  by,  295-296 
mortality 

of  an  adult  Catharus  fuscescens  during  nestling  de- 
fense, 575-576 

of  Charadrius  melodus  from  Ocypode  quadrata, 
321-329 

Motmot,  Blue-crowned,  see  Momotus  momota 
Broad-billed,  see  Electron  platyrhynchum 
Rufous,  see  Baryphthengus  martii 
Rufous-capped,  see  Baryphthengus  ruficapillus 
Turquoise-browed,  see  Eumomota  superciliaria 
mouse,  see  Peromyscus  spp. 

deer,  see  Peromyscus  maniculatus,  Peromyscus  sp. 


house,  see  Mus  musculus 

northern  grasshopper,  sec  Onychomys  leucogaster 
white-footed,  see  Peromyscus  leucopus 
woodland  jumping,  see  Napaeoz.apus  insignis 
Murrelet,  Marbled,  see  Brachyramphus  marmoratus 
Mus  musculus,  23 
Muscicapa  sanfordi,  463 
Muscipipra  vetula,  4 1 1 
Musci.saxicola  macloviana,  530 
muskellunge,  see  Esox  masquinongy 
Mustela  frenata,  239,  416 
Myadestes  unicolor,  229,  233,  234 
Mycteria  americana,  225 

Myiarchus  crinitus,  91,  92,  108,  110,  141,  232,  269, 
346,  348 
sagrae,  3 1 1 
swainsoni,  227 
tyrannulus,  227 
Myiobius  sulphureipygius,  232 
Myiodynastes  maculatlus,  227,  411 
Myiopagis  caniceps,  226,  412 
viridicata,  226 

Myiophobus  fasciatus,  227,  411 
Myiopsitta  monachus,  491 
Myiornis  auricularis,  221,  226,  412 
Myrmeciza  atrothorax,  205,  207 
disjuncta,  195-209 

Myrmotherula  fjeldsaai,  sp.  nov.,  157—165  (Frontis- 
piece) 

haematonota,  157—165  (Frontispiece) 
leucophthalma,  157-165  (Frontispiece) 
ornata,  157 
sororia,  157-165 

spodionota,  157-165  (Frontispiece) 

Napaeozapus  insignis,  239 

Naughton,  M.  B.,  see  Whittow,  G.  C.,  and 

Naugle,  David  E.,  see  Leddy,  Krecia  L.,  Kenneth  F. 

Higgins,  and 

Nemosia  pileata,  228 
Neocrex  colombianus,  423 
erythrops,  422-424 
Neophron  percnopterus,  475 
Neotoma  fuscipes,  22-29,  252 
nest 

arthropod  fauna  of  Falco  span’erius  boxes,  269—27 1 
influence  of  human  scent  on  mammalian  predators, 
415-420 

of  Strix  varia  in  attic  of  shed,  272-273 
predation  along  three  edge  types,  541-549 
predators  of  open  and  cavity  in  oak  woodlands, 
251-256 

renest  after  successful  hatch  in  Meleagris  gallopavo, 
426-427 

reuse  by  Hylocichla  mustelina  and  Pheucticus  lu- 
dovicianus,  132—133 

nest-site 

habitat  of  Accipiter  striatus  and  Accipiter  cooperii 
in  Wisconsin,  7-14 

unusual  for  Empidonax  trailii  extimus,  573—575 


622 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


nesting 

behavior  of  Amazona  finschi,  488-493 
biology  of  Spiza  americana  and  Ammodramus  hen- 
slowii,  515-527 

continuous  in  Tyto  alba,  571-573 
fall  and  winter  records  for  Charadrius  vociferus  in 
the  southern  United  States,  424-426 
of  four  species  of  the  Caribbean  slope  of  Costa  Rica, 
124-128 

success  of  Icteria  virens,  210-215 
nestling 

mortality  of  an  adult  Catharus  fuscescens  during  de- 
fense of,  575-576 

provisioning  rates  in  Picoides  borealis,  494-498 
Neubig,  Jeffrey  R,  and  John  A.  Smallwood,  The  “sig- 
nificant others”  of  American  Kestrels;  cohabi- 
tation with  arthropods,  269-27 1 
Night-Heron,  Black-crowned,  see  Nycticorax  nycticor- 
a.x 

Yellow-crowned,  see  Nyctanassa  violacea 
Nighthawk,  Antillean,  see  Chordeiles  gundlachii 
Nacunda,  see  Podager  nacimda 
Sand-colored,  see  Chordeiles  rupestris 
Nightjar,  Blackish,  see  Cciprimulgus  nigrescens 
Little,  see  Caprimidgus  parvulus 
Red-necked,  see  Caprimulgus  rujicollis 
Rufous,  see  Caprimulgus  rufus 
Scrub,  see  Caprimulgus  anthonyi 
White-winged,  see  Caprimulgus  candicans 
Nino.x  connivens,  462 

ios,  sp.  nov.,  457-464  (Frontispiece) 

jacquinoti,  462 

meeki,  462 

natal  is,  458 

novaeseelandiae,  462 

ochracea  [jjerver.sa],  457-464  (Frontispiece) 

odiosa,  462 

philipensis,  457-464 

punctulata,  457,  462 

rudolfi,  458,  461 

rufa,  462 

scutulata,  462 

squamipila,  457-464  (Frontispiece) 
strenua,  462 
superciliaris,  462 
theomacha,  458,  462 
variegata,  462 

Nomonyx  dominicus,  1 19—121,  31  I 
Nothoprocta  perdicaria,  530 
Nothura  maculosa.  225 
Nucifraga  columbiana.  245 
Numenius  phaeopus,  263 

Nutcracker,  Clark's,  see  Nucifraga  columbiana 
nuthatch,  see  Sitta  spp. 

Nuthatch,  Brown-headed,  see  Sitta  pusilla 
Red-breasted,  see  Sitta  canadensis 
White-breasted,  see  Sitta  carolinensis 
Nyctibius  grandis,  1 24- 1 28 
griseus,  2 1 8,  226 

Nycticorax  nycticorax,  105,  225,  530,  531 
Nyctidromus  albicollis,  218,  226,  231,  344 
Nyctanassa  violacea,  562 


Nystalus  chacuru,  226 
maculatus,  226 

O’Connor,  Raymond  J.,  review  by,  154—156 
Odontoplwrus  capueira,  407 

Olsen,  Klaus  Mailing  Olsen  and  Hans  Larsson,  Skuas 
and  Jaegers:  a guide  to  the  skuas  and  jaegers 
of  the  world,  reviewed,  298-299 
Onychomys  leucogaster,  416 
Ophicephalus  striatus,  17,  19 
Oporornis  agilis,  281—282 
Philadelphia,  93,  281 

opossum,  Virginia,  see  Didelphis  marsupialis 
Oreochromis  spp.,  17 

Oring,  Lewis  W.,  see  Plissner,  Jonathan  H.,  Susan  M. 
Haig,  and 

Oring,  Lewis  W.,  see  Reed,  J.  Michael,  Elizabeth  M. 
Gray,  Dianne  Lewis,  , Richard  Cole- 

man, Timothy  Burr,  and  Peter  Luscomb 
Oriole,  Baltimore,  see  Icterus  galbula 
Bullock’s,  see  Icterus  bullockii 
O’Toole,  Laura  T,  Patricia  L.  Kennedy,  Richard  L. 
Knight,  and  Lowell  C.  McEwen,  Postfledging 
behavior  of  Golden  Eagles,  472-477 
Otus  asio,  254 

atricapillus,  221,  226 
choliba,  226,  408 

Ovenbird,  see  Seiurus  aurocapillus 
Owl,  Barking,  see  Ninox  connivens 
Barn,  see  Tyto  alba 
Barred,  see  Strix  varia 
Boreal,  see  Aegolius  funereus 
Burrowing,  see  Athene  cunicularia 
Great  Gray,  see  Strix  nebulosa 
Great  Horned,  see  Bubo  virginianus 
Long-eared,  see  Asio  otus 

Northern  Spotted,  see  Strix  occidentalis  caurina 
Papuan  Boobook,  see  Ninox  theomacha 
Powerful,  see  Ninox  strenua 
Rufous,  see  Ninox  rufa 

Short-eared,  see  Asio  fiammeus  [Strix  domingensis, 
Asio  portoricensis,  Asio  domingensis^ 
Spectacled,  see  Pulsatrix  perspicillata 
Spotted,  see  Strix  occidentalis 
Tawny,  see  Strix  aluco 
Ural,  see  Strix  uralensis 
White-browed,  see  Nino.x  superciliaris 
0.xyruncus  cristatus,  227 

Pachyramphus  castaneus,  227,  41  1 
major,  232 

polychopterus,  227,  401,  411 
validus,  227 
y/m/fv.  227,  41 1 

Painted-Snipe,  see  Rost  rat  ula  semicollaris 
Pandion  haliaetus,  19 
Parabuteo  unicinctus,  218,  437,  530 
Parakeet,  Maroon-bellied,  see  Pyrrhura  frontalis 
Monk,  see  Myiopsitta  monachus 
Plain,  see  Brotogeris  tirica 
Reddish-bellied,  see  Pyrrhura  frontalis 


INDEX  TO  VOLUME  1 I 1 


623 


Parker.  Timothy  H.,  Responses  of  Bell’s  Vireos  to 
brood  parasitism  by  the  Brown-headed  Cow- 
bird  in  Kansas,  499-504 
Parkes,  Kenneth  C.,  review  by,  144-146 
Paroaria  coronata,  218,  228 
Parrot,  Lilac-crowned,  see  Amazona  jinschi 
Reddish-bellied,  see  Pyrhura  frontalis 
Paruk,  James  D.,  Dean  Seanfield,  and  Tara  Mack,  Bald 
Eagle  predation  on  Common  Loon  chick,  1 15- 
116 

Paruk,  James  D.,  Territorial  takeover  in  Common 
Loons  (Gavia  immer),  116—117 
Panda  americana,  93,  505—514 
pitiaywni,  228,  397-414 
Parula,  Northern,  see  Panda  americana 
Tropical,  see  Parula  pitiayumi 
Pants  caeruleus,  368 
major,  368,  579 
spp.,  368 

Passer  domesticus,  228,  236—242,  251,  269,  289,  440— 
442,  558 

Passerculus  sandwichensis,  100—104,  108,  234 

Passerella  iliaca,  558 

Passerina  cyanea,  93,  141,  211,  212,  214 

Paszkowski,  Cynthia  A.,  see  Sodhi,  Navjot  S., , 

and  Shannon  Keehn 
Patagona  gigas,  529,  530 
Pauraque,  see  Nyctidromus  alhicollis 
Pauxi  pauxi,  564—569 

Peer,  Brian  D.,  and  Spencer  G.  Sealy,  Laying  time  of 
a Bronzed  Cowbird,  137-139 

Peery,  M.  Zachariah,  see  Smith,  Richard  B.,  , 

R.  J.  Gutierrez,  and  William  S.  Lahaye 
Pelecanus  erythrorhynchos,  437 
occidentalis,  437 

Pelican,  American  White,  see  Pelecanus  erythrorhyn- 
chos 

Brown,  see  Pelecanus  occidentalis 
Penelope  ohsciira,  401,  407 
Percnostola  [Schistocichla]  caurensis,  195-209 
[Schistocichla]  leucostigma,  197,  206 
lophotes,  206 
rufifrons,  206 

[Schistocichla]  schistacea,  206 
Perez-Vallafina,  Monica,  Hector  Gomez  de  Silva  G., 
and  Atahualpa  DeSucre-Medrano,  Sexual  di- 
morphism in  the  song  of  Sumichrast’s  Wren, 
128-130 

Perisoreus  canadensis,  245 

Perkins,  Alison  E.  H.,  see  Marks,  Jeffrey  S.,  and 

Peromyscus  leucopus,  25,  28,  236—242,  269,  538 
rnanicidatus,  236,  416 
spp.,  252,  543 

Phacellodomus  rufifrons,  226 
Phaeomyias  rnurina,  226 
Phaeoprogne  tapera,  227,  359 
Phaethornis  pretrei,  409 
Phalacrocorax  brasilianus,  225 
Phalcohoenus  megalopterus,  437-439 
Phasianus  colchicus,  108,  554,  555,  558 


Pheasant,  Ring-necked,  see  Phasianus  colchicus 
Pheucticus  ludovicianus,  93,  94,  96,  108,  132-133 

Phillipps,  Karen,  see  Holmes,  Derek,  and  

Philydor  dimidiatus,  2 1 6-228 
lichtensteini,  221,  226 
rufus,  226,  410 
Phimosus  infuscatus,  225 
Phloeoceastes  rohustus,  403,  409 
photoperiod 

response  of  a Kansas  winter  bird  community  to, 
550-558 

Phrygilus  alaudinus,  531 
fruticeti,  53 1 
gayi,  530 

Phyllomyias  fasciatus,  402,  412 
reiseri,  222 

Phylloscartes  oustaleti,  412 
ventralis,  412 
Phytotoma  rara,  530,  531 

Piatt,  John  E,  see  Kuletz,  Katherine  J.,  and 

Piaya  cayana,  226,  408 
Pica  pica,  245,  274 
Picoides  arcticus,  243—250 

borealis,  38-45,  56,  346-353,  494-498 
lignarius,  530 
mixtus,  226 

pubescens,  91,  92,  141,  142,  245,  246,  277,  346— 
353,  558 

tridactylus,  243-250 

villosus,  89-99,  244,  245,  246,  277,  346,  348,  558 
Piculus  aurulentus,  409 
chrysochloros,  218,  226 
Picumnus  cirratus,  226 
nebulosus,  409 
spp.,  403 
temminckii,  409 

Pigeon,  White-crowned,  see  Columba  leucocephala 
pike,  northern,  see  Esox  niger 
Pintail,  Northern,  see  Anas  acuta 

White-cheeked,  see  Anas  bahamensis  bahamensis 
Pionopsitta  pileata,  408 
Pionus  maximiliani,  225,  408 
Pipilo  crissalis,  251,  252 
erythrophthalmus.  348,  538 
maculatus.  558 

Pipit,  Sprague’s,  see  Anthus  spragueii 
Pipra  fasciicauda,  227 
Pipraeidea  melanonola,  413 
Piranga  fiava,  228 
leucoptera,  233,  234 
olivacea,  93,  94,  95 
rubra.  141,  348 
Pipromorpha  rufiventris,  4 1 2 
Pitangus  sulphuratus,  227,  401,  411 
Pitylus  fulginosus,  414 

Plains-wanderer,  see  Pedionomus  torquatus 

Plantcutter,  Rufous-tailed,  see  Phytotoma  rara 

Platycichla  flavipes,  412 

Platypsaris  rufus,  4 1 1 

Platyrinchus  mystaceus,  221 , 41 1 

Plissner,  Jonathan  H.,  Susan  M.  Haig,  and  Lewis  W. 


624 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


Oring,  Within-  and  between-year  dispersal  of 
American  Avocets  among  multiple  western 
Great  Basin  wetlands,  314-320 
Plover,  Piping,  see  Charadrius  melodus 
Snowy,  see  Charadrius  alexandrinus 
Wilson’s,  see  Charadrius  wilsonia 
plumage 

color  preference  in  female  Cardirtalis  cardinalis, 
76-83 

Podager  nacunda,  226,  344 
Podilymbus  podiceps,  105-1 14 

Poecile  atricapillus.  92,  94,  245,  277,  363-367,  368, 
374,  554,  558 

carolinensis,  141,  363-367,  374 
gambeli,  245,  368-375 
spp.,  368 

Poephila  guttata,  251 
Pogonotriccus  eximius,  412 
Polioptila  dumicola,  221 
Polyborus  plancus,  407 
Polystictus  pectoralis,  222 
Pooecetes  gramineus,  213 
Poospiza  lateralis,  414 
Porphyrula  flavirostris,  264 
rnartinica,  264 
Porzana  albicollis,  225 
Carolina,  107,  108 
spp.,  423 

Post,  William,  see  Smith,  Kimberly  G.,  W.  Marvin  Da- 
vis, Thomas  E.  Kienzle, , and  Robert  W. 

Chinn 

Potoo,  Common,  see  Nyctibius  griseus 
Great,  see  Nyctibius  grandis 
Pratt,  Jerome  J.,  The  Whooping  Crane;  North  Ameri- 
ca’s symbol  of  conservation,  reviewed,  151 
predation 

by  Baryphthengus  martii  on  Dendrobates  auratus, 
439-440 

influence  of  human  scent  in  shortgrass  prairie,  415- 
420 

of  Haliaeetus  leucocephalus  of  Gavia  immer  chick, 

1 15-1 16 

of  open  and  cavity  nesting  birds  in  oak  woodlands, 
251-256 

of  small  eggs  in  artificial  nests,  236-242 
on  artificial  nests  in  three  edge  types,  541—549 
potential  learning  of  artificial  nest  locations,  536- 
540 

relationship  of  Ocypode  quadrata  abundance  and 
Charadrius  melodus  mortality,  321—329 
“snorkeling”  as  escape  behavior  in  Jacana  jacana, 
262-265 

prey 

preference  in  Haliaeetus  vociferoides,  15-21 

Prince,  Harold  H.,  see  Whitt.  Michael  B., , and 

Robert  R.  Cox,  Jr. 
proceedings 

of  eightieth  annual  meeting,  596-603 
Procnias  nudicollis,  219,  221,  227,  41  I 
Procyon  lotor,  2,  416,  538,  542,  546 


productivity 

juvenile  Brachyrarnphus  marmoratus  nurseries, 
257-261 

Progne  chalybea,  221 
spp.,  359 
subis,  354-362 

Prolonotaria  citrea,  141,  577-580 
Psarocolius  decumanus,  228 
Pseudoleistes  guirahura,  228 
Pteroglossus  castanotis,  226 
Pteroptochos  megapodius,  530 
Ptyonoprogne  rupestris,  358 
Puffinus  nativitatus,  421—422 
Pulsatrix  perspicillata,  218,  226 

Purcell,  Kathryn  L.,  and  Jared  Verner,  Nest  predators 
of  open  and  cavity  nesting  birds  in  oak  wood- 
lands, 251-256 

Pygmy-Owl,  Ferruginous,  see  Glaucidium  brasilianum 
Pygmy-Tyrant,  see  Euscarthmus  rufomarginatus 
Eared,  see  Myiornis  auricularis 
Pygochelidon  cyanoleuca,  530 
Pyrhura  frontalis,  219 
Pyriglena  leucoptera,  410 
Pyrocephalus  rubinus,  221 
Pyroderus  scutatus,  220,  227 
Pyrope  pyrope,  530 
Pyrrhocoma  ruficeps,  413 
Pyrrhura  frontalis,  221,  225 

Quail,  California,  see  Callipepla  californica 
Japanese,  see  Coturnix  japonica 
quail,  see  Coturnix  sp. 

Quail-Dove,  Key  West,  see  Geotrygon  chrysia 
Ruddy,  see  Geotrygon  montana 
Quiscalus  quLscula,  84,  543,  544,  546,  554,  555,  558 

racer,  yellow-bellied,  see  Coluber  constictor  flaventris 
racoon,  see  Procyon  lotor 
Rail,  King,  see  Rallus  elegans 
Virginia,  see  Rallus  limicola 
Rallus  elegans,  107 
limicola,  1 05- 1 1 4 
sanguinolentus,  530 
Ramphastos  discolorus,  409 
toco,  226 

Ramphocelus  carbo,  402 

Rasmussen,  Pamela  C.,  A new  species  of  hawk-owl 
Ninox  from  North  Sulawesi,  Indonesia,  457- 
464 

rat,  see  Rattus  spp. 

Rattus  spp.,  376,  561 

Raven,  Common,  see  Con’us  corax 

record 

first  Oporornis  agilis  for  Ecuador,  281-282 
of  breeding  Neocrex  erythrops  in  Costa  Rica,  422— 
424 

renest  after  successful  hatch  in  Meleagris  gallopavo, 
426-427 

second  brood  after  successful  fledging  in  Athene 
cunicularia,  569—571 

Recuerda,  Pilar,  see  Aragones,  Juan,  Luis  Arias  de 
Reyna,  and 


INDEX  TO  VOLUME  I I 1 


625 


Recurvirostra  avosetta,  485 
americana.  314—320 
Redhead,  see  Aythya  americana 
Redstart,  American,  see  Setophaga  niticilla 
Reed,  J.  Michael,  Elizabeth  M.  Gray,  Dianne  Lewis, 
Lewis  W.  Oring,  Richard  Coleman,  Timothy 
Bun;  and  Peter  Luscomb,  Growth  patterns  of 
Hawaiian  Stilt  chicks,  478-487 
Regiihis  calendula.  61-69 

satrapa.  89-99,  245,  246,  554,  555,  558 
release 

factors  affecting  success  in  Picoides  borealis,  38— 
45 

Renton,  Katherine,  and  Alejandro  Salinas-Melgoza, 
Nesting  behavior  of  the  Lilac-crowned  Panot, 
488-493 

report 

survey  of  undergraduate  ornithology  courses  in 
North  America,  287-293 
reproduction 

double  brooding  in  Asia  otus,  273—276 
double  brooding  in  Athene  cimicidaria,  569-571 
effect  of  prior  residence  and  age  in  Dendroica  pe- 
techia, 381-388 

nesting  biology  of  Spiza  americana  and  Ammodra- 
niLis  henslowii,  515—527 

nurseries  and  the  productivity  index  in  Brachyram- 
phus  marmoratus,  257—261 
pairing  success  of  Hylocichla  mustelina  in  a frag- 
mented agricultural  landscape,  279-281 
relationship  of  clutch  size  and  hatching  success  to 
age  in  female  Protonotaria  citrea,  577—580 
responses  of  Vireo  belli  to  brood  parasitism,  499— 
504 

timing  in  Cypseloides  niger,  30-47 
success  in  Strix  occidentalis,  22—29 
Restall,  Robin,  Munias  and  mannikins,  reviewed,  148 
Rhea  americana,  222,  225 
Rhea,  Greater,  see  Rhea  americana 
Rhynchotus  maculosa,  225 

Rice,  Nathan  H.,  see  Robbins,  Mark  B.,  Rob.  C.  Fau- 
cett,  and 

Rice,  Nathan  H.,  Courtship  behavior  of  the  Buff- 
necked Ibis  (Theristicus  caudatus),  1 18-1 19 
Rising,  James  D.,  A guide  to  the  identification  and 
natural  history  of  the  sparrows  of  the  United 
States,  reviewed,  445—446 
Robbins,  Chandler  S.,  review  by,  146-147 
Robbins,  Mark  B.,  Rob.  C.  Faucett,  and  Nathan  H. 
Rice,  Avifauna  of  a Paraguayan  Cerrado  local- 
ity: Parque  Nacional  Serram'a  San  Luis,  depto. 
Concepcion,  216-228 
Robin,  American,  see  Turdus  migratorius 
Black,  see  Turdus  infuscatus 
White-throated,  see  Turdus  assimilis 
Rogers-Price,  Vivian,  John  Abbot’s  birds  of  Georgia: 
selected  drawings  from  the  Houghton  Library, 
Harvard  University,  reviewed,  299-300 

roost 

pre-migratory  in  Progne  subis,  354-362 


Rosenberg,  Daniel  K.,  .see  Gervais,  Jennifer  A.,  and 


Rosenfield,  Robert  N.,  see  Trexel,  Dale  R.,  , 

John  Bielefeldt,  and  Eugene  A.  Jacobs 
Rostratula  semicollaris,  218,  225 
Rostrhamus  sociabilis,  218,  225,  265—268 
Rowley,  Ian,  and  Eleanor  Russell,  Fairy-wrens  and 
grasswrens  Maluridae,  reviewed,  448-449 

Rudolph,  D.  Craig,  see  Conner,  Richard  N.,  , 

Richard  R.  Schaefer,  Daniel  Saenz,  and  Clif- 
ford E.  Shackelford 
Rupicola  rupicola,  200 

Russell,  Eleanor,  see  Rowley,  Ian,  and 

Russell,  Kevin  R.,  and  Sidney  A.  Gauthreaux,  Jr.,  Spa- 
tial and  temporal  dynamics  of  a Purple  Martin 
pre-migratory  roost,  354—362 
Russell,  Stephen  M.,  and  Gale  Monson,  The  birds  of 
Sonora,  reviewed,  595—596 

Sabrewing,  Long-tailed,  see  Campylopterus  excellens 
Violet,  see  Campylopterus  hemileucurus 
Saenz,  Daniel,  see  Conner,  Richard  N.,  D.  Craig  Ru- 
dolph, Richard  R.  Schaefer,  , and  Clif- 

ford E.  Shackelford 
Saimiri  oerstedi,  443 

Salinas-Melgoza,  Alejandro,  see  Renton,  Katherine, 
and 

Sallabanks,  Rex,  see  Marzluff,  John  M.  and 

Salpinctes  obsoletus,  199 
Saltator  atricollis,  217,  228 
similis,  228,  414 

Saltator,  Black-throated,  see  Saltator  atricollis 
Buff-throated,  see  Saltator  maximus 
Sandpiper,  Stilt,  see  Calidris  himantopus 
Upland,  see  Bartramia  longicauda 
White-rumped,  see  Calidris  fuscicollis 
Sapphire,  Rufous-throated,  see  Hylocharis  sapphirina 
SapsLicker,  Yellow-bellied,  see  Sphyrapicus  varius 
Saracco,  James  E,  and  Jaime  A.  Collazo,  Predation  on 
artificial  nests  along  three  edge  types  in  a North 
Carolina  bottomland  hardwood  forest,  541-549 
Sarasola,  Jose  H.,  see  Goldstein,  Michael  I.,  Peter  H. 

Bloom, , and  Thomas  E.  Lacher 

Sarcoramphus  papa.  225 
Satrapa  icterophrys,  4 1 1 
Scaphidura  oiyzivora.  138,  228 
Scatophagus  tetracanthus,  1 9 
Scaup,  Greater,  see  Aythya  marila 
Scelorchilus  albicollis,  530 

Schaefer,  Richard  R.,  see  Conner,  Richard  N.,  D.  Craig 

Rudolph,  , Daniel  Saenz,  and  Clifford 

E.  Shackelford 

Scheiber,  Isabella  B.  R.,  see  Alworth,  Tom,  and 

Schelhas,  John,  and  Russell  Greenberg, 

eds..  Forest  patches  in  tropical  landscapes,  re- 
viewed, 148-149 
Schijfornis  virescens,  41  1 

Schilling,  Amy  J.,  see  Breitwisch,  Randall,  , 

and  Joshua  B.  Banks 
Schoeniophylax  phryganophila,  226 


626 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


Scholtens,  Brian,  see  Cuthbert,  Erancesca  J., , 

Lauren  C.  Wemmer,  and  Robyn  McLain 

Schreiber,  Carrie  L.,  see  Avery,  Michael  L.,  , 

and  David  G.  Decker 

Schuchmann,  Karl-L.,  see  Jahn,  Olaf,  Maria  Eugenia 

Jara  Viteri,  and 

Sciuru.s  carolinensis,  269,  542,  546 
griseus,  25 
Sclateria  naevia,  204 
Sclerurus  scansor,  410 
spp.,  199,  205 

Screech-Owl,  Eastern,  see  Otiis  asio 
Variable,  see  Otus  atricapilhis 
Scrub-Jay,  Elorida,  see  Aphelocoma  coerulescens 
Western,  see  Aphelocoma  californica 
Scytalopu.s  ma^ellanicu.s,  530 

Sealy,  Spencer  G.,  see  Peer,  Brian  D.,  and 

Seanheld,  Dean,  see  Paruk,  Janies  D.,  , and 

Tara  Mack 

Seedeater,  Dark-throated,  see  Sporophila  ruficoUis 
Seiurus  aurocapilliis,  93,  94,  96,  240,  279,  443-444 
Sephanoide.s  galeritus,  530 

Seriema,  Red-legged,  see  Lurocalis  semitorquatus 
Serpophaga  cinerea,  124—128 
suh.scistata,  4 1 2 

Setophaga  ruticilla,  70-75,  93,  94,  96 
sexual 

dimorphism 

in  song  of  Hylorchilus  surnichrasti,  128—130 
selection 

in  Caprimulgus  ruficoUis,  340—345 
Shackelford,  Clifford  E.,  see  Conner,  Richard  N.,  D. 
Craig  Rudolph,  Richard  R.  Schaefer,  Daniel 
Saenz,  and  

Shearwater,  Christmas,  Puffinus  nativitatus 
shrew,  masked,  see  Sorex  cinereiis 

northern  short-tailed,  see  Blarina  hrevicauda 
smokey,  see  Sorex  fumeus 
Shrike,  Loggerhead,  see  Lxmius  hidovicianus 
Northern,  see  Lanins  excuhitor 
Sialia  currucoides,  440—442 
mexicana,  243,  253,  441 
.sialis,  269,  346,  348,  441,  558 
spp.,  441 

Sicalis  fiaveola,  228,  414 
sp.,  530 

Silva,  Jose,  Notes  about  the  distribution  of  Pauxi  pauxi 
and  Ahurria  ahurri  in  Venezuela,  564-569 
Simpson,  Scott  A.,  see  Walk,  Jeffrey  W.,  Terry  L Es- 

ker,  and 

Sirysles  sihilator,  227 

Siskin,  Hooded,  see  Carduelis  magellanica 
Pine,  see  Carduelis  pinus 
site-fidelity 

in  Vireo  g rise  us.  46—55 
Sitla  canadensis,  89-99,  245,  246,  277,  558 
carolinen.si.s,  92,  94,  245,  246,  363-367,  558 
pu.silla,  56-60,  346-353 
spp.,  204 

Siltasomus  griseicapillus,  22 1 , 226,  409 

.Skagcn,  Susan  K.,  Thomas  R.  Stanley,  and  M.  Beth 


Dillon,  Do  mammalian  nest  predators  follow 
human  scent  trails  in  the  shortgrass  prairie?, 
415-420 

Skeel,  M.,  see  Davis,  S.  K.,  D.  C.  Duncan,  and 

skunk,  striped,  see  Mephitis  mephitis 
Skutch,  Alexander  E,  Life  of  the  flycatcher,  reviewed, 
147 

Smallwood,  John  A.,  see  Neubig,  Jeffrey  P,  and 

Smith,  Charles  R.,  review  by,  445-446 

Smith,  Christopher  C.,  see  Stapanian,  Martin  A., 

, and  Elmer  J.  Einck 

Smith,  Kimberly  G.,  W.  Marvin  Davis,  Thomas  E. 
Kienzle,  William  Post,  and  Robert  W.  Chinn, 
Additional  records  of  fall  and  winter  nesting  by 
Killdeer  in  southern  United  States,  424-426 
Smith,  Michael  R.,  Philip  W.  Mattocks,  Jr.,  and  Kelly 
M.  Cassidy,  Breeding  birds  of  Washington 
state:  location  data  and  predicted  distributions, 
reviewed,  296-297 

Smith,  P.  William,  see  Hoffman,  Wayne,  Glen  E. 
Woolfenden,  and 

Smith,  Richard  B.,  M.  Zachariah  Peery,  R.  J.  Gutier- 
rez, and  William  S.  Lahaye,  The  relationship 
between  Spotted  Owl  diet  and  reproductive 
success  in  the  San  Bernadino  Mountains,  Cal- 
ifornia, 22-29 

Smith,  Winston  Paul,  and  Daniel  J.  Twedt,  Temporal 
differences  in  point  counts  of  bottomland  forest 
landbirds,  139-143 
snake,  garter,  see  Thamnophis  sirtalis 
Snipe,  Common,  see  Gallinago  gallinago 
Sodhi,  Navjot  S.,  Cynthia  A.  Paszkowski,  and  Shan- 
non Keehn,  Scale-dependent  habitat  selection 
by  American  Redstarts  in  aspen-dominated  for- 
est fragments,  70-75 

Solitaire,  Slate-colored,  see  Myadestes  unicolor 
Somateria  mollissima,  1,  3,  4 
Sora,  see  Porzana  Carolina 
Sorex  cinereus,  239 
fumeus,  239 

Sparrow,  American  Tree,  see  Spizella  arhorea 
Bachman’s,  see  Aimophila  aestivalis 
Baird’s,  see  Ammodramus  hairdii 
Brewer’s,  see  Spizella  breweri 
Chipping,  see  Spizella  passerina 
Clay-colored,  see  Spizella  pallida 
Field,  see  Spizella  pusilla 
Fox,  see  Passerella  iliaca 
Grasshopper,  see  Ammodramus  savannarum 
Harris’,  see  Zonotrichia  querula 
Henslow’s,  see  Ammodramus  henslowii  ■ 

House,  see  Passer  domesticus 
Le  Conte’s,  see  Ammodramus  leconteii 
Lincoln’s,  see  Melospiza  lincolnii 
Olive,  see  Arremonops  rufivirgatus 
Orange-billed,  see  Arremon  aurantiirostris 
Rufous-collared,  see  Zonotrichia  capensis 
Savannah,  see  Passerculus  sandwichensis 
Song,  see  Melospiza  melodia 
Swamp,  see  Melospiza  georgiana 


INDEX  TO  VOLUME  1 I 1 


627 


Vesper,  see  Pooecetes  gramineiis 
White-crowned,  see  Zonotrichia  leucophrys 
White-throated,  see  Zonotrichia  alhicollis 
Sparrowhawk,  Eurasian,  see  Accipiter  nisus 
Levant,  see  Accipiter  brevipes 
species  nova 

Myrmotherula  haematonota,  157-165 
Ninox  ios,  457—464 
Speotyto  cunicularia,  226,  530 
Spermophiliis  heecheyi,  252,  253 
tridecemlineatus.  416 
Sphyrapiciis  variiis,  92,  94,  348,  558 
Spindalis  zena,  3 1 1 

Spinetail,  White-lored,  see  Synallaxis  alhilora 
Spiza  americana,  100-104,  515-527 
Spizaetus  ornatiis.  218,  225 
Spizella  arborea,  554,  555,  558 
breweri.  581 

pallida.  100-104,  581-584 
passerina,  93,  94,  581,  584 
pusiUa,  214,  554,  558,  581-584 
Sporophiia  bouvreuil,  228 
caendescens.  228,  414 
collaris,  228 
hypoxantha,  228 
plumbea,  228 
ruficoUis,  218,  222,  228 

squirrel,  California  ground,  see  SpermophUus  beecheyi 
flying,  see  Glaucomys  sp. 
gray,  see  Sciurus  caroiinen.'iis 
northern  flying,  see  Glaucomys  sabrinus 
red,  see  Tamiasciurus  grahamensis  and  Tamiasciu- 
rus  hudsonicus 

southern  flying,  see  Glaucomys  volans 
thirteen-lined  ground,  see  SpermophUus  tridecemli- 
neatus 

western  grey,  see  Sciurus  griseiis 

Stanley,  Thomas  R.,  see  Skagen,  Susan  K.,  , 

and  M.  Beth  Dillon 

Stapanian,  Martin  A.,  Christopher  C.  Smith,  and  Elmer 
J.  Finck,  The  response  of  a Kansas  winter  bird 
community  to  weather,  photoperiod,  and  year, 
550-558 

Starling,  European,  see  Sturnus  vulgaris 
Staus,  Nancy  L.,  and  Paul  M.  Mayer,  Arthropods  and 
predation  of  artificial  nests  in  the  Bahamas:  im- 
plications for  subtropical  avifauna,  561-564 

Stein,  Steven  J.,  see  Kreisel,  Karen  J.,  and 

Stelgidopteryx  ruficoUis,  221 
Stephanophorus  diadematus.  401,  402,  413 
Stephanoxis  lalandi,  409 
Stercorarius  parasiticus,  437 
Sterna  aleutica.  559—560 
anaethetus.  560 
antillarum,  231 
bergii,  560 
dougallii,  563 
forsteri,  108,  109 
fuscata,  560 
hirundo,  560 
sandvicensis,  23 1 


superciliaris,  344 

Stilt,  Black,  see  Himantopus  novaezealandiae 
Black-necked,  see  Himantopus  mexicanus 
Hawaiian,  see  Himantopus  mexicanus  knudseni 
Pied,  see  Himantopus  himantopus  leucocephalus 
Stoleson,  Scott  H.,  and  Deborah  M.  Finch,  Unusual 
nest  sites  for  southwestern  Willow  Flycatchers, 
573-575 

Streptopelia  risoria,  25  1 
Strix  aluco,  273 
hylophila,  408 
nebulosa,  21 
occidentalis,  22—29,  569 
occidentaiis  caurina,  22 
uraiensis,  273 
varia,  89-99,  272-273 
Sturnella  loyca,  530 
magna,  254,  555,  558 
neglecta,  100-104,  416,  554,  555,  558 
Sturnus  vulgaris,  84,  86,  87,  253,  269,  289,  359,  558, 
579 

Sungrebe,  see  Heliornis  fulica 

Suriri  suiriri,  226 

survival 

of  Lophodytes  cucullatus  and  Aix  sponsa  in  south- 
eastern Missouri,  1-6 
of  Vireo  latimeri,  376-380 
Swallow,  Bahama,  see  Tachycineta  cyaneoviridis 
Bam,  see  Hirundo  rustica 
Cave,  see  Hirundo  fulva 
Tree,  see  Tachycineta  bicolor 
Swanson,  David  L.,  Eric  T.  Liknes,  and  Kurtis  L. 
Dean,  Differences  in  migratory  timing  and  en- 
ergetic condition  among  sex/age  classes  in  mi- 
grant Ruby-crowned  Kinglets,  61-69 
Swift,  Black,  see  Cypseioides  niger 
Vaux’s,  see  Chaetura  vauxi 
SyivUagus  spp.,  25 
Synallaxis  albescens,  226 
albilora,  222 
cinerascens.  410 
ruficapilla.  410 
.•ipixi.  410 

Syndactyla  rufosuperciliata,  410 
Svrigma  sibilatrix,  225,  407 

Tacha,  Thomas  C.,  see  Anderson,  James  T,  and 


Tachuri,  Bearded,  see  Polystictus  pectoralis 
Tachycineta  bicolor.  108,  233,  269 
cyaneoviridis.  3 1 1 
leucopyga.  530 
leucorrhoa,  221 
Tachvphonus  coronatus,  413 
rufus,  228 

Taeniopygia  guttata.  237 
Tamias  striatus,  221 , 239,  538 
Tamiasciurus  hudsonicus.  239,  269 
Tanager,  Gray-headed,  see  Eucometis  pencillata 
Rufous-headed,  see  Hemithraupis  ruficapilla 
Sayaca,  see  Thraupis  sayaca 


628 


THE  WILSON  BULLETIN  • Vol.  Ill,  No.  4,  December  1999 


Scarlet,  see  Piranga  olivacea 
Stripe-headed,  see  Spiudcilis  zena 
Summer,  see  Piranga  rubra 
White-winged,  see  Piranga  leucoptera 
Tangara  preciosa,  413 
Tapera  naevia,  226,  408 
Ta.xidea  ta.xa,  416 
taxonomy 

new  species  in  the  Myrmotherida  haematonota  su- 
perspecies, 157-165 
of  Caracara  spp.,  330-339 

of  Percno.stola  caurensis  and  Myrmeciza  disjuncta, 
195-209 

tayra,  see  Eira  barbara 
Teal,  Blue-winged,  see  Ana.s  discor.s 
Tern,  Aleutian,  see  Sterna  aleutica 
Black,  see  Chlidonia.s  niger 
Bridled,  see  Sterna  anaethetus 
Crested,  see  Sterna  bergii 
Common,  see  Sterna  hirundo 
Forster's,  see  Sterna  for.steri 
Least,  see  Sterna  antillarum 
Roseate,  see  Sterna  dougalli 
Sandwich,  see  Sterna  sandvicensis 
Sooty,  see  Sterna  fiiscata 
Yellow-billed,  see  Sterna  superciliari.s 
territory 

takeover  in  Gavia  immer,  1 16—1 17 
Tersina  viridi.x,  413 
Thalurania  furcata,  226 
glaucopis,  409 

Thamnophihis  caerule.xcens,  226,  404,  410 
Thamnophi.x  .xirtalis,  440 
Theri.sticu.s  caudatus.  118-119,  225,  407 
thermoregulation 

vocalization  of  Gallu.s  gallu.s  embryos  in  response 
to  temperature,  188-194 
Thinocorus  spp.,  264 

Thompson,  Frank  R.,  Ill,  see  Burhans,  Dirk  E.,  and 


Thrasher,  Brown,  see  Toxostoma  rufum 
Pearly-eyed,  see  Margarop.s  fu.scatu.s 
Thraiipi.s  bonarien.si.s,  413 
.sayaca,  228,  40 1 , 402,  4 1 3 
Thrush,  Hermit,  see  Catharus  guttatii.x 
Swainson’s,  see  Catharus  ustulatus 
Varied,  see  Ixoreus  naevius 
Wood,  see  Hylocichia  mustelina 
Thryomanes  bewickii,  558 
Thryothorus  leucotis,  205 

ludovicianus.  141,  142,  555,  558 
modest  us.  229,  233 
Thymomas  bottae,  25,  28 
Tiaris  bicolor.  31  1 
fuliginosa.  414 
Tilapia  spp.,  17 
Tit,  Blue,  see  Parus  caerideus 
Great,  see  Parus  major 
Titmouse,  Bridled,  see  Baeolophus  woUweberi 
Plain,  see  Baeolophus  inornatus 
Tufted,  see  Baeolophus  bicolor 


Tityra  cayana.  221 , 41 1 
inquisitor,  227 

Todd,  Frank  S.,  Natural  history  of  the  waterfowl,  re- 
viewed, 446-448 
Todirostrum  plumbeiceps,  412 
Tolmomyias  sulphurescens,  227,  412 
Towhee,  California,  see  Pipilo  crissalis 
Eastern,  see  Pipilo  erythrophthalmus 
Spotted,  see  Pipilo  maculatus 
Toxostoma  rufum.  558 
translocation 

factors  affecting  success  in  Picoides  borealis,  38- 
45 

Trejo  R,  Juana  Lourdes,  see  Winker,  Kevin,  Stefan  Ar- 
riaga Weiss, , and  Patricia  Escalante  P. 

Trexel,  Dale  R.,  Robert  N.  Rosenfield,  John  Bielefeldt, 
and  Eugene  A.  Jacobs,  Comparative  nest  site 
habitats  in  Sharp-shinned  and  Cooper’s  hawks 
in  Wisconsin,  7—14 
Trichothraupis  melanops,  228,  413 
Tringa  soiitaria,  225 

Troglodytes  aedon,  130-132,  227,  253,  254,  403,  412, 
530 

sissonii,  128 

troglodytes,  89-99,  128,  245,  558 
Trogon  curucui,  226 
rufus,  409 
surrucura,  409 

Tucker,  Graham  M.,  and  Michael  I.  Evans,  Habitats 
for  birds  in  Europe:  a conservation  strategy  for 
the  wider  environment,  reviewed,  454-456 
Tupinambis  sp.,  218 
Turdus  albicollis,  412 

amaurochalinus,  227,  401,  402,  412 

assimilis,  233 

faicklandii,  530 

flavipes,  412 

infuscatus,  229,  233,  234 

migratorius,  84,  86,  87,  92,  94,  141,  142,  441,  558 
nigriceps,  412 
rufiventris,  227,  401,  402 
turkey,  see  Meleagris  spp. 

Turkey,  Wild,  see  Meleagris  gallopavo 
Turnix  sp.,  251 

Turnstone,  Ruddy,  see  Arenaria  interpres 
turtle,  snapping,  see  Chelydra  serpentina 
Turtle-Dove,  Ringed,  see  Streptopelia  risoria 

Twedt,  Daniel  J.,  see  Smith,  Winston  Paul,  and 

Tyrannulet,  Reiser’s,  see  Phyllomyias  reiseri 
Torrent,  see  Serpophaga  cinerea 
Tyranniis  dominicensis,  310 
melancholicus,  227,  402,  411 
savana,  221 , 4 1 1 
tyrannus,  108,  346,  348,  349 
vertical  is,  252 

Tyrant,  Cock-tailed,  see  Alecturus  tricolor 
Sharp-tailed,  see  CuUcivora  caudacuta 
White-tufted,  see  Elaenia  albiceps 
Tyto  alba.  27,  273,  408,  530,  571-573 
Upucerthia  dumetaria,  530 
ruficauda,  530 


INDEX  TO  VOLUME  1 1 I 


629 


use 

of  Lythrion  salicaria  dominated  habitat,  105-114 

Vanelhis  chilensis,  225,  530 
vanellus.  131,  484 
Veery,  see  Cathuriis  fuscescens 
Veit,  Richard  R.,  review  by,  298—299 
Veniliornis  passerinus,  226 
spilogaster.  397-414 
Vermivora  luciae,  577 
ruficcipilla.  229,  233 

Verner,  Jared,  see  Purcell,  Kathryn  L.,  and 

Vireo  altiloqiius,  310 

atricapilliis,  46—55,  378,  379 
helm,  51,  52,  378,  379,  499-504 
bellii  pusilliis,  46,  499 
crassirostris,  311 
gilvus,  378 

griseus,  46—55,  141,  378 
latimeri,  376—380 

olivaceus,  52,  91,  93,  94,  141,  228,  279,  378,  413 
pallens,  232 

solitarius,  92,  94,  95,  233 
vicinior,  378 

Vireo,  Bell’s,  see  Vireo  hellii 

Black-capped,  see  Vireo  atricapillus 
Black-whiskered,  see  Vireo  altiloquus 
Blue-headed,  see  Vireo  solitarius 
Gray,  see  Vireo  vicinior 
Least  Bell’s,  see  Vireo  hellii  pusillus 
Mangrove,  see  Vireo  pallens 
Puerto  Rican,  see  Vireo  latimeri 
Red-eyed,  see  Vireo  olivaceus 
Thick-billed,  see  Vireo  crassirostris 
Warbling,  see  Vireo  gilvus 
White-eyed,  see  Vireo  griseus 

Viteri,  Maria  Eugenia  Jara,  see  Jahn,  Olaf, , and 

Karl-L.  Schuchmann 
vocalization 

aggressive  response  of  Poecile  atricapillus  and  Poe- 
cile  carolinensis  to  calls,  363-367 
of  Callus  gallus  embryos  in  response  to  tempera- 
ture, 188—194 

of  Percnostola  caurensis  and  Myrmeciza  disjuncta. 
1 95-209 

sexual  dimorphism  in  song  of  Hylorchilus  sumi- 
chrasti,  128-130 

singing  in  mated  female  Wilsonia  pusilla,  134-137 
type  B song  of  Parula  americana,  505-514 
use  of  song  types  by  Poecile  gamheli,  368-375 
Volatinia  jacarina,  228,  414 
vole,  see  Microtus  spp. 

pine,  see  Microtus  pennsylvanicus 
red-backed,  see  Clethrionornys  gapperi 
Vuilleumier,  Eran9ois,  review  by,  300-301 
Vulpes  vulpes,  417,  576 
Vultur  gryphus,  530 
Vulture,  Black,  see  Coragyps  atratus 
Egyptian,  see  Neophron  percnopterus 
Turkey,  see  Cathartes  aura 


Walk,  Jeffrey  W.,  Ten  y L Eskcr,  and  Scott  A.  Simpson, 
Continuous  nesting  of  Barn  Owls  in  Illinois, 
571-573 

Warbler,  Bay-breasted,  see  Dendroica  castanea 
Black-and-white,  see  Mniotilta  varia 
Black-throated  Blue,  see  Dendroica  caerulescens 
Black-throated  Green,  see  Dendroica  virens 
Blackburnian,  see  Dendroica  fusca 
Chestnut-sided,  see  Dendroica  pensylvanica 
Connecticut,  see  Oporornis  agilis 
Golden-crowned,  see  Basileuterus  culicivorus 
Grace’s,  see  Dendroica  graciae 
Hermit,  see  Dendroica  occidentalis 
Hooded,  see  Wilsonia  citrina 
Magnolia,  see  Dendroica  magnolia 
Mourning,  see  Oporornis  Philadelphia 
Nashville,  see  Vermivora  ruficapilla 
Pine,  see  Dendroica  pinus 
Prairie,  Dendroica  discolor 
Prothonotary,  see  Protonotaria  citrea 
Swainson’s,  see  Limnothlypis  swainsonii 
Townsend’s,  see  Dendroica  townsendi 
White-bellied,  see  Basileuterus  hypoleucus 
Wilson’s,  see  Wilsonia  pusilla 
Yellow,  see  Dendroica  petechia 
Yellow-rumped,  see  Dendroica  coronata 
Watson,  David  M.,  and  Brett  W,  Benz,  The  Paint-billed 
Crake  breeding  in  Costa  Rica,  422-424 
Watson,  Richard  T,  see  Berkelman,  James,  James  D. 
Eraser,  and 

Watts,  Bryan  D.,  see  Wilson,  Michael  D.,  and  

Waxwing,  Cedar,  see  Bomhycilla  cedrorum 
weasel,  long-tailed,  see  Mustela  frenata 
weather 

response  of  a bird  assemblage  to  the  1997-1998  El 
Nino  in  Chile,  527—535 

response  of  a Kansas  bird  community,  550-558 
weight 

post-migration  gain  of  Buteo  swainsoni  in  Argenti- 
na, 428-432 

Weiss,  Stefan  Arriaga,  see  Winker,  Kevin, , Ju- 

ana Lourdes  Trejo  P,  and  Patricia  Escalante  P. 
Wemmer,  Lauren  C.,  see  Cuthbert,  Francesca  J.,  Brian 

Scholtens, , and  Robyn  McLain 

Wheatley,  Nigel,  Where  to  watch  birds  in  Asia,  re- 
viewed, 149-150 

Whimbrel,  see  Numenius  phaeopus 
Whip-poor-will,  see  Caprimulgus  vociferus 
Whistling-Duck,  Fulvous,  see  Dendrocygna  hicolor 
West  Indian,  see  Dendrocygna  arhorea 
Whitney,  Bret  M.,  see  Krabbe,  Niels,  Morton  L.  Isler, 

Phyllis  R.  Isler,  , Jose  Alvarez  A.,  and 

Paul  J.  Greenfield 

Whitt,  Michael  B.,  Harold  H.  Prince,  and  Robert  R. 
Cox,  Jr.,  Avian  use  of  purple  loosestrife  dom- 
inated habitat  relative  to  other  vegetation  types 
in  a Lake  Huron  wetland  complex,  105-114 
Whittow,  G.  C.,  and  M.  B.  Naughton,  Christmas 
Shearwater  egg  dimensions  and  shell  charac- 
teristics on  Laysan  Island,  northwestern  Ha- 
waiian Islands,  421-422 


630 


THE  WILSON  BULLETIN  • Vol.  II],  No.  4,  December  1999 


Wiebe,  Myra  O.,  and  M.  Ross  Lein,  Use  of  song  types 
by  Mountain  Chickadees  {Poecile  gambeli), 
368-375 

Willet,  see  Catoptrophonis  semipabnatus 

Wilson,  Michael  D.,  and  Bryan  D.  Watts,  Response  of 
Brown-headed  Nuthatches  to  thinning  of  pine 
plantations,  56-60 

Wilson,  W.  Herbert,  Jr.,  see  Burtt,  Edward  H.,  Jr.,  and 


Wilsonia  citrina,  93,  94,  95 
pusilla,  134-137 

Winker,  Kevin,  Stefan  Arriaga  Weiss,  Juana  Lourdes 
Trejo  R,  and  Patricia  Escalante  R,  Notes  on  the 
avifauna  of  Tabasco,  229-235 
Winter,  Maiken,  Nesting  biology  of  Dickcissels  and 
Henslow’s  Sparrows  in  southwestern  Missouri 
prairie  fragments,  515-526 

Wolcott,  Donna  L.,  and  Thomas  G.  Wolcott,  High  mor- 
tality of  Piping  Plovers  on  beaches  with  abun- 
dant ghost  crabs:  correlations,  not  causation, 
321-329 

Wolcott,  Thomas  G.,  see  Wolcott,  Donna  L.,  and 

Wolfenbarger,  L.  Lareesa,  Female  mate  choice  in 
Northern  Cardinals:  is  there  a preference  for 
redder  males?,  76-83 

Wood-Pewee,  Eastern,  see  Contopus  virens 
Woodcreeper,  Great  Rufous,  see  Xiphocolaptes  major 
Narrow-billed,  see  Lepidocolapte.s  angu.stirostris 
Olivaceous,  see  Sitta.somu.s  griseicapillii.s 
Tawny-winged,  see  Dendrocincia  anabatina 
Wedge-billed,  see  Glyphorhynchu.s  spirurus 
Woodpecker,  Acorn,  see  Melanerpes  formicivorus 
Black-backed,  see  Picoide.s  arcticus 
Downy,  see  Picoide.s  pube.scen.s 
Golden-green,  see  Piculu.s  chry.sochloro.s 
Hairy,  see  Picoide.s  villo.su.s 
Lineated,  see  Dryocopii.s  lineatii.s 
Pileated,  see  Dryocopii.s  pileatii.s 
Red-bellied,  see  Melanerpe.s  carolinu.s 
Red-cockaded,  see  Picoide.s  boreali.s 
Red-headed,  see  Melanerpe.s  erythrocephahi.s 
Three-toed,  see  Picoide.s  tridactylu.s 
woodrat,  dusky-footed,  see  Neotoma  fu.scipe.s 
Woodworth,  Bethany  L.,  John  Faaborg,  and  Wayne  J. 
Arendt,  Survival  and  longevity  of  the  Puerto 
Rican  Vireo,  376-380 

Woolfenden,  Glen  E.,  see  Hoffman,  Wayne,  , 

and  P.  William  Smith 


Wren,  Bewick’s,  see  Thryomanes  bewickii 
Buff-breasted,  see  Thryothorus  leiicotis 
Cactus,  see  Campylorhynchu.s  bruimeicapillus 
Canyon,  see  Catherpes  mexicamis 
Carolina,  see  Thryothorus  hidovicianus 
House,  see  Troglodytes  aedon 
Marsh,  see  Cistothorus  palustris 
Plain,  see  Thryothorus  modestus 
Rock,  see  Salpinctes  obsoletus 
Sedge,  see  Cistothorus  platensis 
Sorocco,  see  Troglodytes  sissonii 
Sumichrast’s,  see  Hylorchilus  sumichrasti 
Winter,  see  Troglodytes  troglodytes 

Wyatt,  Valerie  E.,  see  Friesen,  Lyle  E.,  , and 

Michael  D.  Cadman 

Xanthocephalus  xanthocephalus,  108,  254,  558 
Xenops  minutus,  410 
rut  Hans,  410 

Xenopsaris  albinucha,  218,  227 
Xenopsaris,  White-naped,  see  Xenopsaris  albinucha 
Xiphocolaptes  albicolis,  409 
major,  221,  226 
Xolmis  cinerea,  227,  411 
velata,  227 

Yahner,  Richard  H.,  and  Carolyn  G.  Mahan,  Potential 
for  predator  learning  of  artificial  arboreal  nest 
locations,  536-540 

Yellowthroat,  Common,  see  Geothlypis  trichas 
Yom-Tov,  Yoram,  see  Gorney,  Edna,  William  S.  Clark, 
and 

Young,  Wang  and  Deborah  M.  Fitch,  Response,  589- 
592 

Young,  Bruce  E.,  and  James  R.  Zook,  Nesting  of  four 
poorly-known  bird  species  on  the  Caribbean 
slope  of  Costa  Rica,  124-128 

Zenaida  auriculata,  408,  530 
aurita,  310,  311 

macroura,  92,  94,  96,  139-143,  310,  558 
Zimmer,  Kevin  J.,  Behavior  and  vocalizations  of  the 
Caura  and  Yapacana  antbirds,  195-209 
Zonotrichia  albicollis,  141,  554,  555,  558 
capensis,  228,  402,  414,  530,  533 
leucophrys,  554,  555,  558 
querula,  213,  558 

Zook,  James  R.,  see  Young,  Bruce  E.,  and 


This  issue  of  The  Wilson  Bulletin  was  published  on  14  December  1999. 


PUBLISHED  BY  THE  WILSON  OKMTHOI-OGIUAI.  SOCIETY 


VOLUME  111  1999  QUAKTEKLY 


EDITOR:  ROBERT  C.  SEASON 
EDITORIAL  BOARD:  KATHY  G.  BEAL 
CLAIT  E.  BRAUN 
RICHARD  N.  CONNER 
INDEX  EDITOR:  KATHY  G.  BEAL 
ASSISTANT  EDITORS:  TARA  BAIDEME 
JOHN  LAMAR 
DANTE  THOMAS 
DORIS  WATT 


The  Wilson  Ornithological  Society 
Founded  December  3,  1888 

Named  after  ALEXANDER  WILSON,  the  first  American  Ornithologist 

President  John  C.  Kricher,  Biology  Department,  Wheaton  College,  Norton,  Massachusetts 
02766;  E-mail:  JKricher@wheatonma.edu. 

First  Vice-President— William  E.  Davis,  Jr.,  College  of  General  Studies,  871  Commonwealth 
Ave.,  Boston  University,  Boston,  Massachusetts  02215;  E-mail:  WEDavis@bu.edu. 

Second  Vice-President — Charles  R.  Blem,  Dept,  of  Biology,  816  Park  Ave.,  Virginia 
Commonwealth  Univ.,  Richmond,  Virginia  23284;  E-mail:  cblem@saturn.vcu.edu. 

Editor — Robert  C.  Beason,  Department  of  Biology,  State  University  of  New  York,  1 College 
Circle,  Geneseo,  New  York  14454;  E-mail:  WilsonBull@geneseo.edu. 

Secretary — John  A.  Smallwood,  Department  of  Biology,  Montclair  State  University,  Upper 
Montclair,  New  Jersey  07043;  E-mail:  Smallwood@mail.montclair.edu. 

Treasurer — Doris  J.  Watt,  Department  of  Biology,  Saint  Mary’s  College,  Notre  Dame, 
Indiana  46556;  E-mail:  DWatt@saintmarys.edu. 

Elected  Council  Members — Charles  F.  Thompson  and  Sara  R.  Morris  (terms  expire  2000), 
Jonathan  L.  Atwood  and  James  L.  Ingold  (terms  expire  2001),  Robert  A.  Askins  and 
Jeffrey  R.  Walters  (terms  expire  2002). 


DATES  OF  ISSUE  OF  VOLUME  1 1 1 
OF  THE  WILSON  BULLETIN 

NO.  1 — I March  1999 
NO.  2 — 10  May  1999 
NO.  3 — 10  August  1999 
NO.  4 — 14  December  1999 


CONTENTS  OF  VOLUME  1 1 1 


NUMBER  I 


MAJOR  PAPERS 

ANNUAL  SURVIVAL  RATES  OF  FEMALE  HOODED  MERGANSERS  AND  WOOD  DUCKS  IN 
SOUTHEASTERN  MISSOURI  Katie  M.  Dugger,  Bruce  D.  Dugger,  and  Leigh  H.  Kredrickson 

COMPARATIVE  NEST  SITE  HABITATS  IN  SHARP-SHINNED  AND  COOPER’S  HAWKS  IN  WIS- 
CONSIN   Dale  R.  Trexel,  Robert  N.  RosenfieUl,  John  Bielefeldt,  and  Eugene  A.  Jacob.s 

MADAGASCAR  FISH-EAGLE  PREY  PREFERENCE  AND  FORAGING  SUCCESS 

James  Berkelman,  James  D.  Fraser,  and  Richard  T.  Watson 

THE  RELATIONSHIP  BETWEEN  SPOTTED  OWL  DIET  AND  REPRODUCTIVE  SUCCESS  IN  THE 

SAN  BERNARDINO  MOUNTAINS,  CALIFORNIA  

Richard  B.  Smith.  M.  Zachariah  Peery,  R.  J.  Gutierrez,  and  William  S.  Lcduiye 

FOOD,  FORAGING,  AND  TIMING  OF  BREEDING  OF  THE  BLACK  SWIFT  IN  CALIFORNIA 

Manuel  Marin 

FACTORS  THAT  INFLUENCE  TRANSLOCATION  SUCCESS  IN  THE  RED-COCKADED  WOOD- 
PECKER   Kathleen  E.  Franzreb 

BANDING  RETURNS,  ARRIVAL  PATTERN,  AND  SITE-FIDELITY  OF  WHITE-EYED  VIREOS  -- 

S.  L.  Hopp,  A.  Kirby,  and  C.  A.  Boone 

RESPONSE  OF  BROWN-HEADED  NUTHATCHES  TO  THINNING  OF  PINE  PLANTATIONS 

Michael  D.  Wilson  and  Bryan  D.  Watts 

DIFFERENCES  IN  MIGRATORY  TIMING  AND  ENERGETIC  CONDITION  AMONG  SEX/AGE 

CLASSES  IN  MIGRANT  RUBY-CROWNED  KINGLETS  

David  L.  Swanson,  Eric  T.  Liknes,  and  Kurtis  L.  Dean 

SCALE-DEPENDENT  HABITAT  SELECTION  BY  AMERICAN  REDSTARTS  IN  ASPEN-DOMINAT- 
ED FOREST  FRAGMENTS  Navjot  S.  Sodhi,  Cynthia  A.  Paszkowski.  and  Shannon  Keehn 

FEMALE  MATE  CHOICE  IN  NORTHERN  CARDINALS:  IS  THERE  A PREFERENCE  FOR  REDDER 
MALES? _ Wolfenbarger 

FRUIT  SUGAR  PREFERENCES  OF  HOUSE  FINCHES  

Michael  L.  Avery,  Carrie  L.  Schreiber,  and  David  G.  Decker 

HIERARCHICAL  COMPARISONS  OF  BREEDING  BIRDS  IN  OLD-GROWTH  CONIFER-HARD- 
WOOD FOREST  ON  THE  APPALACHIAN  PLATEAU  J-  Christopher  Haney 

EFFECTS  OF  WIND  TURBINES  ON  UPLAND  NESTING  BIRDS  IN  CONSERVATION  RESERVE 
PROGRAM  GRASSLANDS  Krecia  L.  Leddy,  Kenneth  F.  Higgins,  and  David  E.  Naugle 

AVIAN  USE  OF  PURPLE  LOOSESTRIFE  DOMINATED  HABITAT  RELATIVE  TO  OTHER  VEGE- 
TATION TYPES  IN  A LAKE  HURON  WETLAND  COMPLEX  

Michael  B.  Whitt.  Harold  H.  Prince,  and  Robert  R.  Cox.  Jr. 


SHORT  COMMUNICATIONS 

BALD  EAGLE  PREDATION  ON  COMMON  LOON  CHICK 

James  D.  Paruk.  Dean  Seanfield.  and  Tara  Mack 

TERRITORIAL  TAKEOVER  IN  COMMON  LOONS  (GAVIA  IMMER) James  D.  Paruk 

COURTSHIP  BEHAVIOR  OF  THE  BUFF-NECKED  IBIS  (THERISTICUS  CAUDATUS) 

Nathan  H.  Rice 

HABITAT  USE  BY  MASKED  DUCKS  ALONG  THE  GULF  COAST  OF  TEXAS  

James  T.  Ander.son  and  Thomas  C.  Tacha 

GIZZARD  CONTENTS  OF  PIPING  PLOVER  CHICKS  IN  NORTHERN  MICHIGAN 

Francesca  J.  Cuthbert.  Brian  Scholtens.  Uiuren  C.  Wemmer.  and  Robyn  McLain 


NESTING  OF  FOUR  POORLY-KNOWN  BIRD  SPECIES  ON  THE  CARIBBEAN  SLOPE  OF 
COSTA  RICA  Bruce  E.  Young  and  James  R.  Zook 

SEXUAL  DIMORPHISM  IN  THE  SONG  OF  SUMICHRAST’S  WREN  — 

Monica  Perez-ViUafana.  Hector  Gomez  de  Silva  G..  and  Atahualpa  De  Sue  re -Medrano 

AN  INCIDENT  OF  FEMALE-FEMALE  AGGRESSION  IN  THE  HOUSE  WREN  — 

Tom  Alworth  and  Isabella  B.R.  Scheiber 

NEST  REUSE  BY  WOOD  THRUSHES  AND  ROSE-BREASTED  GROSBEAKS 


Lyle  E.  Eriesen.  Valerie  E.  Wyatt,  and  Michael  D.  Cadman 


SINGING  IN  A MATED  FEMALE  WILSON’S  WARBLER  

William  M.  Gilbert  and  Adele  F.  Carroll 

LAYING  TIME  OF  THE  BRONZED  COWBIRD  Brian  D.  Peer  and  Spencer  G.  Sealy 

TEMPORAL  DIFFERENCES  IN  POINT  COUNTS  OF  BOTTOMLAND  FOREST  LANDBIRDS 

Winston  Paul  Smith  and  Daniel  J.  Twedt 


ORNITHOLOGICAL  LITERATURE 


I 

7 

15 

22 

30 

38 

46 

56 

61 

70 

76 

84 

89 

100 

105 

115 

1 16 

118 

1 19 

121 

124 

128 

130 

132 

134 

137 

139 

144 


NUMBER  2 


MAJOR  PAPERS 

A NEW  SPECIES  IN  THE  MYRMOTHERULA  HAEMATONOTA  SUPERSPECIES  (AVES;  THAM- 
NOPHILIDAE)  FROM  THE  WESTERN  AMAZONIAN  LOWLANDS  OF  ECUADOR  AND  PERU 

Niels  Krabbe,  Morton  L.  Isler,  Phyllis  R.  Isler,  Bret  M.  Whitney, 

Jose  Alvarez  A.,  and  Paul  J.  Greenfield 

COMPARATIVE  SPRING  HABITAT  AND  FOOD  USE  BY  TWO  ARCTIC  NESTING  GEESE 

Suzanne  Carriere,  Robert  G.  Bromley,  and  Gilles  Gauthier 

A TEST  OF  THE  CONDITION-BIAS  HYPOTHESIS  YIELDS  DIFFERENT  RESULTS  FOR  TWO 
SPECIES  OF  SPARROWHAWKS  {ACCIPITER)  

Edna  Gorney,  William  S.  Clark,  and  Yoram  Yom-Tov 

THE  DEVELOPMENT  OF  A VOCAL  THERMOREGULATORY  RESPONSE  TO  TEMPERATURE  IN 

EMBRYOS  OF  THE  DOMESTIC  CHICKEN  Shawn  C.  Bugden  and  Roger  M.  Evans 

BEHAVIOR  AND  VOCALIZATIONS  OF  THE  CAURA  AND  THE  YAPACANA  ANTBIRDS 

Eevin  J.  Zimmer 

HABITAT  PATCH  SIZE  AND  NESTING  SUCCESS  OF  YELLOW-BREASTED  CHATS  

Dirk  E.  Burhans  and  Erank  R.  Thompson,  III 

AVIFAUNA  OF  A PARAGUAYAN  CERRADO  LOCALITY:  PARQUE  NACIONAL  SERRANIA  SAN 

LUIS,  DEPTO.  CONCEPCION  Mark  B.  Robbins,  Rob.  C.  Eaucett,  and  Nathan  H.  Rice 

NOTES  ON  THE  AVIFAUNA  OF  TABASCO  

Kevin  Winker,  Stefan  Arriaga  Weiss,  Juana  Lourdes  Trejo  P.,  and  Patricia  Escalante  P. 

PREDATION  OF  SMALL  EGGS  IN  ARTIFICIAL  NESTS:  EFFECTS  OF  NEST  POSITION,  EDGE, 

AND  POTENTIAL  PREDATOR  ABUNDANCE  IN  EXTENSIVE  FOREST  ! 

Richard  M.  DeGraaf  Thomas  J.  Maier,  and  Todd  K.  Fuller 

BIRD  USE  OF  BURNED  AND  UNBURNED  CONIFEROUS  FORESTS  DURING  WINTER  

Karen  J.  Kreisel  and  Steven  J.  Stein 

NEST  PREDATORS  OF  OPEN  AND  CAVITY  NESTING  BIRDS  IN  OAK  WOODLANDS  

Kathryn  L.  Purcell  and  Jared  Verner 

SHORT  COMMUNICATIONS 

JUVENILE  MARBLED  MURRELET  NURSERIES  AND  THE  PRODUCTIVITY  INDEX  

Katherine  J.  Kuletz  and  John  E.  Piatt 

“SNORKELING”  BY  THE  CHICKS  OF  THE  WATTLED  JACANA  

Carlos  Bosque  and  Emilio  A.  Herrera 

RAPID  LONG-DISTANCE  COLONIZATION  OF  LAKE  GATUN,  PANAMA,  BY  SNAIL  KITES 

George  R.  Angehr 

THE  “SIGNIFICANT  OTHERS”  OF  AMERICAN  KESTRELS:  COHABITATION  WITH  AR- 
THROPODS   Jeffrey  P.  Neubig  and  John  A.  Smallwood 

BARRED  OWL  NEST  IN  ATTIC  OF  SHED  C.  Stuart  Houston 

DOUBLE  BROODING  IN  THE  LONG-EARED  OWL  

Jeffrey  S.  Marks  and  Alison  E.  H.  Perkins 

PLANNING  TO  FACILITATE  CACHING:  POSSIBLE  SUET  CUTTING  BY  A COMMON  RA- 
VEN   Bernd  Heinrich 

PAIRING  SUCCESS  OF  WOOD  THRUSHES  IN  A FRAGMENTED  AGRICULTURAL  LAND- 
SCAPE   Lyle  E.  Eriesen,  Valerie  E.  Wyatt,  and  Michael  D.  Cadman 

CONNECTICUT  WARBLER,  A NORTH  AMERICAN  MIGRANT  NEW  TO  ECUADOR  

Olaf  Jahn,  Maria  Eugenia  Jara  Viteri,  and  Karl-L.  Schuchmann 

PARENTAL  BEHAVIOR  OF  A BIGAMOUS  MALE  NORTHERN  CARDINAL 

Randall  Breitwisch,  Amy  J.  Schilling,  and  Joshua  B.  Banks 

SPECIAL  REPORT 

A SURVEY  OF  UNDERGRADUATE  ORNITHOLOGY  COURSES  IN  NORTH  AMERICA  ..L . _ 

Edward  H.  Burtt,  Jr.  and  W.  Herbert  Wilson,  Jr. 

ORNITHOLOGICAL  LITERATURE  


157 

166 

181 

188 

195 

210 

216 

229 

236 

243 

251 

257 

262 

265 

269 

272 

273 
276 
279 
281 
283 

2&7 

294 


NUMBER  3 


ANTILLEAN  SHORT-EARED  OWLS  INVADE  SOUTHERN  ELORIDA  - 

Wayne  Hoffman,  Glen  E.  WoolfenJen,  and  P.  William  Smith  303 

WITHIN-  AND  BETWEEN-YEAR  DISPERSAL  OE  AMERICAN  AVOCETS  AMONG  MULTIPLE 

WESTERN  GREAT  BASIN  WETLANDS 

Jonathan  H.  Plissner,  Susan  M.  Haig,  and  Lewis  W.  Oring  3 1 4 

HIGH  MORTALITY  OE  PIPING  PLOVERS  ON  BEACHES  WITH  ABUNDANT  GHOST  CRABS; 

CORRELATION,  NOT  CAUSATION Donna  L Wolcott  and  Thomas  G.  Wolcott  321 

A TAXONOMIC  STUDY  OE  CRESTED  CARACARAS  (EALCONIDAE) 

Carla  J.  Dove  and  Richard  C.  Banks  330 

VISUAL  COMMUNICATION  AND  SEXUAL  SELECTION  IN  A NOCTURNAL  BIRD  SPECIES,  CA- 
PRIMULGUS  RUFICOLLIS,  A BALANCE  BETWEEN  CRYPSIS  AND  CONSPICUOUSNESS  - 
Juan  Aragones,  Luis  Arias  De  Reyna,  and  Pilar  Recuerda  340 

INTERSPECIEIC  INTERACTIONS  WITH  FORAGING  RED-COCKADED  WOODPECKERS  IN 

SOUTH-CENTRAL  FLORIDA  

Reed  Bowman,  David  L.  Leonard,  Jr.,  Leslie  K.  Backus,  and  Allison  R.  Mains  346 

SPATIAL  AND  TEMPORAL  DYNAMICS  OF  A PURPLE  MARTIN  PRE-MIGRATORY  ROOST 

Kevin  R.  Russell  and  Sidney  A.  Gauthreaux,  Jr.  354 

AGGRESSIVE  RESPONSE  OF  CHICKADEES  TOWARDS  BLACK-CAPPED  AND  CAROLINA 

CHICKADEE  CALLS  IN  CENTRAL  ILLINOIS  Eric  L.  Kershner  and  Eric  K.  Bollinger  363 

USE  OF  SONG  TYPES  BY  MOUNTAIN  CHICKADEES  {POECILE  GAMBELL)  

Myra  O.  Wiebe  and  M.  Ross  Lein  368 

SURVIVAL  AND  LONGEVITY  OF  THE  PUERTO  RICAN  VIREO  

Bethany  L.  Woodworth,  John  Faaborg,  and  Wayne  J.  Arendt  2sl() 

EFFECTS  OF  PRIOR  RESIDENCE  AND  AGE  ON  BREEDING  PERFORMANCE  IN  YELLOW  WAR- 
BLERS   Lozano  and  R.  E.  Lemon  381 

DISTRIBUTION  AND  HABITAT  ASSOCIATIONS  OF  THREE  ENDEMIC  GRASSLAND  SONG- 
BIRDS IN  SOUTHERN  SASKATCHEWAN  S.  K.  Davis,  D.  C.  Duncan,  and  M.  Skeel  389 

BIRD  COMMUNITIES  IN  NATURAL  FOREST  PATCHES  IN  SOUTHERN  BRAZIL 

Luiz  Dos  Anjos  and  Roberto  Bogon  397 

DO  MAMMALIAN  NEST  PREDATORS  FOLLOW  HUMAN  SCENT  TRAILS  IN  THE  SHORTGRASS 

PRAIRIE?  Susan  K.  Skagen,  Thomas  R.  Stanley,  and  M.  Beth  Dillon  415 

SHORT  COMMUNICATIONS 

CHRISTMAS  SHEARWATER  EGG  DIMENSIONS  AND  SHELL  CHARACTERISTICS  ON  LAY- 

SAN  ISLAND,  NORTHWESTERN  HAWAIIAN  ISLANDS  

G.  C.  Whittow  and  M.  B.  Naughton  421 

THE  PAINT-BILLED  CRAKE  BREEDING  IN  COSTA  RICA  

_ David  M.  Watson  and  Brett  W.  Benz  422 

ADDITIONAL  RECORDS  OF  FALL  AND  WINTER  NESTING  BY  KILLDEER  IN  SOUTHERN 

UNITED  STATES  Kimberly  G.  Smith,  W.  Marvin  Davis,  Thomas  E.  Kienzle, 

William  Post,  and  Robert  W.  Chinn  424 

WILD  TURKEYS  (MELEAGRIS  GALLOPAVO)  RENEST  AFTER  SUCCESSFUL  HATCH  

Craig  A.  Harper  and  Jay  H.  E.xum  426 

POST-MIGRATION  WEIGHT  GAIN  OF  SWAINSON’S  HAWKS  IN  ARGENTINA  

Michael  I.  GoULstein,  Peter  H.  Bloom,  Jose  H.  Sarasola,  and  Thomas  E.  Lacher  428 

SIBLICIDE  AT  NORTHERN  GOSHAWK  NESTS:  DOES  FOOD  PLAY  A ROLE?  

Wendy  A.  Estes,  Sarah  R.  Dewey,  and  Patricia  L.  Kennedy  432 

COOPERATIVE  FORAGING  IN  THE  MOUNTAIN  CARACARA  IN  PERU  Ja.son  Jones  437 

PREDATION  BY  RUFOUS  MOTMOT  ON  BLACK-AND-GREEN  POISON  DART  FROG  

Terry  L.  Master  439 


EVIDENCE  OF  EGG  EJECTION  IN  MOUNTAIN  BLUEBIRDS  Percy  N.  Hebert  440 

FORAGING  OVENBIRD  FOLLOWS  ARMADILLO  Douglas  J.  Levey  443 

ORNITHOLOGICAL  LITERATURE  445 


NUMBER  4 


A NEW  SPECIES  OF  HAWK-OWL  NINOX  FROM  NORTH  SULAWESI,  INDONESIA  

(j  Rasmussen 

PATTERNS  OF  VARIATION  IN  SIZE  AND  COMPOSITION  OF  GREATER  SCAUP  EGGS:  ARE 

THEY  RELATED?  — Paul  L.  Flint  and  J.  Barry  Grand 

POSTFLEDGING  BEHAVIOR  OF  GOLDEN  EAGLES  

Laura  T.  O'Toole,  Patricia  L.  Kennedy,  Richard  L.  Knight,  and  Lowell  C.  McEwen 

GROWTH  PATTERNS  OF  HAWAIIAN  STILT  CHICKS  

Michael  Reed,  Elizabeth  M.  Gray,  Dianne  Lewis,  Lewis  W.  Oring,  Richard  Coleman, 

Timothy  Burr,  and  Peter  Luscomb 

NESTING  BEHAVIOR  OF  THE  LILAC-CROWNED  PARROT 

- Katherine  Renton  and  Alejandro  Salinas-Melgoza 

RELATIONSHIPS  AMONG  RED-COCKADED  WOODPECKER  GROUP  DENSITY,  NESTLING 

PROVISIONING  RATES,  AND  HABITAT  

Richard  N.  Conner,  D.  Craig  Rudolph,  Richard  R.  Schaefer, 

Daniel  Saenz,  and  Clifford  E.  Shackeljbrd 
RESPONSES  OF  BELL’S  VIREOS  TO  BROOD  PARASITISM  BY  THE  BROWN-HEADED  COWBIRD 

IN  KANSAS Timothy  El.  Parker 

THE  TYPE  B SONG  OF  THE  NORTHERN  PARULA:  STRUCTURE  AND  GEOGRAPHIC  VARIATION 
ALONG  PROPOSED  SUB-SPECIES  BOUNDARIES  Michael  D.  Bay 

NESTING  BIOLOGY  OF  DICKCISSELS  AND  HENSLOW’S  SPARROWS  IN  SOUTHWESTERN 

MISSOURI  PRAIRIE  FRAGMENTS Maiken  Winter 

RESPONSE  OF  A BIRD  ASSEMBLAGE  IN  SEMIARID  CHILE  TO  THE  1997-1998  EL  NINO 

- - - Fabian  M.  Jaksic  and  Ivan  Lazo 

POTENTIAL  FOR  PREDATOR  LEARNING  OF  ARTIFICIAL  ARBOREAL  NEST  LOCATIONS 

---- - - Richard  H.  Yahner  and  Carolyn  G.  Mahan 

PREDATION  ON  ARTIFICIAL  NESTS  ALONG  THREE  EDGE  TYPES  IN  A NORTH  CAROLINA 

BOTTOMLAND  HARDWOOD  FOREST James  F.  Saracco  and  Jaime  A.  Collazo 

THE  RESPONSE  OF  A KANSAS  WINTER  BIRD  COMMUNITY  TO  WEATHER,  PHOTOPERIOD, 
AND  YEAR  Martin  A.  Stapanian,  Christopher  C.  Smith,  and  Elmer  J.  Finck 

SHORT  COMMUNICATIONS 

POSSIBLE  WINTER  QUARTERS  OF  THE  ALEUTIAN  TERN?  

- Norman  P.  Hill  and  K.  David  Bishop 

ARTHROPODS  AND  PREDATION  OF  ARTIFICIAL  NESTS  IN  THE  BAHAMAS:  IMPLICATIONS 

FOR  SUBTROPICAL  AVIFAUNA Nancy  L.  Staus  and  Paul  M.  Maver 

NOTES  ABOUT  THE  DISTRIBUTION  OF  PAUXI  PAUXI  AND  ABURRIA  ABURRI  IN  VENE- 
ZUELA   Jose  L.  Silva 

WESTERN  BURROWING  OWLS  IN  CALIFORNIA  PRODUCE  SECOND  BROODS  OF  CHICKS 

Jennifer  A.  Gervais  and  Daniel  K.  Rosenberg 

CONTINUOUS  NESTING  OF  BARN  OWLS  IN  ILLINOIS  

leffery  W.  Walk,  Terry  L.  Esker,  and  Scott  A.  Simpson 

UNUSUAL  NEST  SITES  FOR  SOUTHWESTERN  WILLOW  FLYCATCHERS  

Scott  H.  Stoleson  and  Deborah  M.  Finch 

MORTALITY  OF  AN  ADULT  VEERY  INCURRED  DURING  THE  DEFENSE  OF  NESTLINGS 

David  /.  King 

RELATIONSHIPS  OF  CLUTCH  SIZE  AND  HATCHING  SUCCESS  TO  AGE  OF  FEMALE  PRO- 

THONOTARY  WARBLERS Charles  R.  Blem,  Leann  B.  Blem,  and  Claudia  /.  Barrientos 

HYBRIDIZATION  BETWEEN  CLAY-COLORED  SPARROW  AND  FIELD  SPARROW  IN 
NORTHERN  VERMONT  David  J.  Hoag 

COMMENTARY 

A CRITIQUE  OF  WANG  YONG  AND  FINCH’S  FIELD-IDENTIFICATIONS  OF  WILLOW  FLY- 
CATCHER SUBSPECIES  IN  NEW  MEXICO  John  P Hubbard 

RESPONSE Wang  Yong  and  Deborah  M.  Finch 

ORNITHOLOGICAL  LITERATURE  

PROCEEDINGS  OF  THE  EIGHTIETH  ANNUAL  MEETING  

ACKNOWLEDGMENTS  

INDEX  TO  VOLUME  I I I 

CONTENTS  TO  VOLUME  I I I 


457 

465 

472 

478 

488 

494 

499 

505 

515 

527 

536 

541 

550 

559 

561 

564 

569 

572 

574 

576 

577 
581 

585 

589 

593 

600 

608 

610 


THE  WILSON  BULLETIN 


Editor  ROBERT  C.  BEASON 


Editorial  Board  KATHY  G.  BEAL 


Department  of  Biology 

State  University  of  New  York 

1 College  Circle 

Geneseo,  NY  14454 

E-mail:  WilsonBull@geneseo.edu 


CLAIT  E.  BRAUN 
RICHARD  N.  CONNER 


Review  Editor  WILLIAM  E.  DAVIS,  JR. 


127  East  Street 

Foxboro,  Massachusetts  02035 


Editorial  Assistants  TARA  BAIDEME 


Index  Editor  KATHY  G.  BEAL 


JOHN  LAMAR 
DANTE  THOMAS 
DORIS  WATT 


616  Xenia  Avenue 
Yellow  Springs,  Ohio  45387 


SUGGESTIONS  TO  AUTHORS 

See  Wilson  Bulletin,  1 10:152-154,  1998  for  more  detailed  “Instructions  to  Authors.” 
http://www.ummz.lsa.umich.edu/birds/wilsonbull.html 
Submit  four  copies  of  manuscripts  intended  for  publication  in  The  Wilson  Bulletin,  neatly  typewritten, 
double-spaced,  with  at  least  3 cm  margins,  and  on  one  side  only  of  good  quality  white  paper.  Do  not 
submit  xerographic  copies  that  are  made  on  slick,  heavy  paper.  Tables  should  be  typed  on  separate  sheets, 
and  should  be  narrow  and  deep  rather  than  wide  and  shallow.  Follow  the  AOU  Check-list  (Seventh  Edition, 
1998)  insofar  as  scientific  names  of  U.S.,  Canadian,  Mexican,  Central  American,  and  West  Indian  birds 
are  concerned.  Abstracts  should  be  brief  but  quotable.  Where  fewer  than  5 papers  are  cited,  the  citations 
may  be  included  in  the  text.  Follow  carefully  the  style  used  in  this  issue  in  listing  the  literature  cited, 
otherwise,  follow  the  “CBE  Scientific  Style  and  Format  Manual”  (AIBS  1994).  Photographs  for  illustra- 
tions should  have  good  contrast  and  be  on  glossy  paper.  Submit  prints  unmounted  and  provide  a brief  but 
adequate  legend  for  each  figure  with  all  captions  on  a single  page.  Do  not  write  heavily  on  the  backs  of 
photographs.  Diagrams  and  line  drawings  should  be  in  black  ink  and  their  lettering  large  enough  to  permit 
reduction.  Original  figures  or  photographs  submitted  must  be  smaller  than  22  X 28  cm.  Alterations  in 
copy  after  the  type  has  been  set  must  be  charged  to  the  author. 


NOTICE  OF  CHANGE  OF  ADDRESS 


If  your  address  changes,  notify  the  Society  immediately.  Send  your  complete  new  address  to  Ornitho- 
logical Societies  of  North  America,  P.O.  Box  1897,  Lawrence,  KS  66044-8897. 

The  permanent  mailing  address  of  the  Wilson  Ornithological  Society  is:  do  The  Museum  of  Zoology, 
The  University  of  Michigan,  Ann  Arbor,  Michigan  48109.  Persons  having  business  with  any  of  the  officers 
may  address  them  at  their  various  addresses  given  on  the  back  of  the  front  cover,  and  all  matters  pertaining 
to  the  Bulletin  should  be  .sent  directly  to  the  Editor. 

MEMBERSHIP  INQUIRIES 

Membership  inquiries  should  be  sent  to  Laurie  J.  Goodrich,  Route  2 Box  301  A,  New  Ringgold,  PA 
17960-9445;  E-mail:  goodrich@hawkmountain.org. 


CONTENTS 


A NEW  SPECIES  OF  HAWK-OWL  A/AQYFROM  NORTH  SULAWESI,  INDONESIA 

— - — - - Pamela  C.  Rasmussen 

PATTERNS  OF  VARIATION  IN  SIZE  AND  COMPOSITION  OF  GREATER  SCAUP  EGGS:  ARE 

THEY  RELATED?  — Paul  L.  Flint  and  J.  Barry  Grand 

POSTFLEDGING  BEHAVIOR  OF  GOLDEN  EAGLES  

— - Laura  T.  O Toole,  Patricia  L.  Kennedy,  Richard  L.  Knight,  and  Lowell  C.  McEwen 

GROWTH  PATTERNS  OF  HAWAIIAN  STILT  CHICKS  

J-  Michael  Reed,  Elizabeth  M.  Gray,  Dianne  Lewis,  Levins  W.  Oring,  Richard  Coleman, 

Timothy  Burr,  and  Peter  Luscomb 

NESTING  BEHAVIOR  OF  THE  LILAC-CROWNED  PARROT 

- - Katherine  Renton  and  Alejandro  Salinas-Melgoza 

RELATIONSHIPS  AMONG  RED-COCKADED  WOODPECKER  GROUP  DENSITY,  NESTLING 

PROVISIONING  RATES,  AND  HABITAT  

Richard  N.  Conner,  D.  Craig  Rudolph,  Richard  R.  Schaefer, 

Daniel  Saenz,  and  Clifford  E.  Shackelford 
RESPONSES  OF  BELL’S  VIREOS  TO  BROOD  PARASITISM  BY  THE  BROWN-HEADED  COWBIRD 

IN  KANSAS  Timothy  H.  Parker 

THE  TYPE  B SONG  OF  THE  NORTHERN  PARULA:  STRUCTURE  AND  GEOGRAPHIC  VARIATION 

ALONG  PROPOSED  SUB-SPECIES  BOUNDARIES Michael  D.  Bay 

NESTING  BIOLOGY  OF  DICKCISSELS  AND  HENSLOW’S  SPARROWS  IN  SOUTHWESTERN 

MISSOURI  PRAIRIE  FRAGMENTS Maiken  Winter 

RESPONSE  OF  A BIRD  ASSEMBLAGE  IN  SEMIARID  CHILE  TO  THE  1997-1998  EL  NINO 

- - - Fabian  M.  Jaksic  and  Ivan  Lazo 

POTENTIAL  FOR  PREDATOR  LEARNING  OF  ARTIFICIAL  ARBOREAL  NEST  LOCATIONS 

Richard  H.  Yahner  and  Carolyn  G.  Mahan 

PREDATION  ON  ARTIFICIAL  NESTS  ALONG  THREE  EDGE  TYPES  IN  A NORTH  CAROLINA 

BOTTOMLAND  HARDWOOD  FOREST  James  F.  Saracco  and  Jaime  A.  Collazo 

THE  RESPONSE  OF  A KANSAS  WINTER  BIRD  COMMUNITY  TO  WEATHER.  PHOTOPERIOD, 
AND  YEAR  Martin  A.  Stapanian,  Christopher  C.  Smith,  and  Elmer  J.  Finck 

SHORT  COMMUNICATIONS 

POSSIBLE  WINTER  QUARTERS  OF  THE  ALEUTIAN  TERN?  

- Norman  P Hill  and  K.  David  Bishop 

ARTHROPODS  AND  PREDATION  OF  ARTIFICIAL  NESTS  IN  THE  BAHAMAS:  IMPLICATIONS 

FOR  SUBTROPICAL  AVIFAUNA Nancy  L.  Staus  and  Paul  M.  Mayer 

NOTES  ABOUT  THE  DISTRIBUTION  OF  PAUXI  PAUXI  AND  ABURRIA  ABURRI  IN  VENE- 
ZUELA   Jose  L.  Silva 

WESTERN  BURROWING  OWLS  IN  CALIFORNIA  PRODUCE  SECOND  BROODS  OF  CHICKS 

Jennifer  A.  Gervais  and  Daniel  K.  Rosenberg 

CONTINUOUS  NESTING  OF  BARN  OWLS  IN  ILLINOIS 

Jeffery  W.  Walk.  Terry  L.  Esker,  attd  Scott  A.  Simpson 

UNUSUAL  NEST  SITES  FOR  SOUTHWESTERN  WILLOW  FLYCATCHERS  

Scott  H.  Stoleson  and  Deborah  M.  Finch 

MORTALITY  OF  AN  ADULT  VEERY  INCURRED  DURING  THE  DEFENSE  OF  NESTLINGS 

- David  1.  King 

RELATIONSHIPS  OF  CLUTCH  SIZE  AND  HATCHING  SUCCESS  TO  AGE  OF  FEMALE  PRO- 

THONOTARY  WARBLERS Charles  R.  Bletn,  Leann  B.  Blern,  and  Claudia  /.  Barrientos 

HYBRIDIZATION  BETWEEN  CLAY-COLORED  SPARROW  AND  FIELD  SPARROW  IN 
NORTHERN  VERMONT  David  J Hoag 

COMMENTARY 

A CRITIQUE  OF  WANG  YONG  AND  FINCH’S  FIELD-IDENTIFICATIONS  OF  WILLOW  FLY- 
CATCHER SUBSPECIES  IN  NEW  MEXICO  John  P.  Hubbard 

RESPONSE Wang  Yong  and  Deborah  M.  Finch 

ORNITHOLOGICAL  LITERATURE  

PROCEEDINGS  OF  THE  EIGHTIETH  ANNUAL  MEETING  

ACKNOWLEDGMENTS  — 

INDEX  TO  VOLUME  I I I 

CONTENTS  TO  VOLUME  I I I 


457 

465 

472 

478 

488 

494 

499 

505 

515 

527 

536 

541 

550 

559 

561 

564 

569 

572 

574 

576 

577 
581 

585 

589 

593 

600 

608 

610 


3 2044