HARVARD UNIVERSITY Ernst Mayr Library of the Museum of Comparative Zoology •t s? I Volume 122y Number 1, March 2010 Published by the Wilson Ornithological Society THE WILSON ORNITHOLOGICAL SOCIETY FOUNDED 3 DECEMBER 1888 Named after ALEXANDER WILSON, the first American ornithologist. President E. Dale Kennedy, Biology Department, Albion College, Albion, MI 49224, USA; e-mail: dkennedy@albion.edu First Vice-President— Robert C. Beason, P. O. Box 737, Sandusky, OH 44871, USA; e-mail: Robert.C.Beason@gmail.com Second Vice-President — Robert L. Curry, Department of Biology, Villanova University, 800 Lancaster Avenue, Villanova, PA 19085, USA; e-mail: robert.curry@villanova.edu Editor — Clait E. Braun, 5572 North Ventana Vista Road, Tucson, AZ 85750, USA; e-mail: TWILSONJO@ comcast.net Secretary — John A. Smallwood, Department of Biology and Molecular Biology, Montclair State University, Montclair, NJ 07043, USA; e-mail: smallwoodj@montclair.edu Treasurer— Melinda M. Clark, 52684 HighlandDrive, South Bend, IN 46635, USA; e-mail: MClark@tcservices.biz Elected Council Members— Robert S. Mulvihill, Timothy J. O’Connell, and Mia R. Revels (terms expire 2010); Jameson F. Chace, Sara R. Morris, and Margaret A. Voss (terms expire 2011); Mary Bomberger Brown, Mary Garvin, and Mark S. Woodrey (terms expire 2012). Living Past-Presidents — Pershing B. Hofslund, Douglas A. James, Jerome A. Jackson, Clait E. Braun, Mary H. Clench, Richard C. Banks, Richard N. Conner, Keith L. Bildstein, Edward H. Burtt Jr., John C. Kricher, William E. Davis Jr., Charles R. Blem, Doris J. W’att, and James D. Rising. Membership dues per calendar year are: Regular, $40.00; Student, $20.00; Family, $50.00; Sustaining, $100.00; Life memberships, $1,000.00 (payable in four installments). The Wilson Journal of Ornithology is sent to all members not in arrears for dues. THE WILSON JOURNAL OF ORNITHOLOGY (formerly The Wilson Bulletin) THE WILSON JOURNAL OF ORNITHOLOGY (ISSN 1559-4491) is published quarterly in March, June, September, and December by the Wilson Ornithological Society, 810 East 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 OSNA, 5400 Bosque Boulevard, Suite 680, Waco, TX 76710. All articles and communications for publication should be addressed to the Editor. Exchanges should be addressed to The Josselyn Van Tyne Memorial Library, Mu.seum of Zoology, Ann Arbor, Ml 48109, USA. Subscriptions, changes of address, and claims for undelivered copies should be sent to OSNA, 5400 Bosque Boulevard, Suite 680, Waco, TX 76710, USA. Phone: (254) 399-9636; e-mail: business@osnabirds.org. Back issues or single copies are available for $ 12.00 each. Most back issues of the Journal are available and may be ordered from OSNA. Special prices will be quoted for quantity orders. All issues of the journal published before 2000 are accessible on a free web site at the University of New Mexico library (http://elibrary.unm.edu/sora/). The site is fully searchable, and full-text reproductions of all papers (including illustrations) are available as either PDF or DjVu files. © Copyright 2010 by the Wilson Ornithological Society Printed by Allen Press Inc., Lawrence, KS 66044, USA. COVER: Wilson’s Snipe (Gallinago clelicata). Illustration by Scott Rashid. @ This paper meets the requirements of ANSI/NISO Z39.48-1992 (Permanence of Paper). MCZ LIBRARY JUL 1 3 2010 HARVARD UNIVERSITY FRONTISPIECE. Evennann's Rock Ptarmigan {Lagopiis nnita evcrnianni), endemic to the Near Islands in the western Aleutians, were reintroduced to Agattu from Attu to reestablish a breeding population. Translocations during 2003-2006 were successful as translocated birds survived, nested, and produced young which survived until recruitment in successive breeding seasons. Photograph of a male during the breeding period by Steve E. Ebbert. Wilson Journal of Ornithology Published by the Wilson Oniitbologiccil Society VOL. 122, NO. 1 March 2010 PAGES 1-206 The Wilson Journal of Ornithology 122(1): 1-14, 2010 DEMOGRAPHY OF A REINTRODUCED POPULATION OF EVERMANN’S ROCK PTARMIGAN IN THE ALEUTIAN ISLANDS ROBB S. A. KALER,' STEVE E. EBBERT," CLAIT E. BRAUN,-" AND BRETT K. SANDERCOCK'-^ ABSTRACT. — We report results of a 4-year translocation effort to reestablish a breeding population of Evermann's Rock Ptarmigan (Lagopus niuta evermanni) in the Near Islands group of the western Aleutian Archipelago. Habitat restoration was completed by eradication of introduced foxes from Agattu Island by 1979. We captuied and moved 75 ptarmigan from Attu Island to Agattu Island during 2003-2006, and monitored 29 radio-marked females in the last 2 years of the study. We compared the demography of newly translocated birds (n = 13) with resident birds established from translocations in previous years (/? = 16). Mortality risk was increased by translocation and 15% of females died within 2 weeks of release at Agattu Island. All surviving females attempted to nest but initiated clutches 8 days later in the breeding season and laid 1.5 fewer eggs per clutch than resident females. Probability of nest survival (x ± SE) was good tor both translocated (0.72 ± 0.17) and resident females (0.50 ± 0.16), and renests were rare. Probability of brood survival was higher among translocated (0.85 ± 0.14) than resident females (0.25 ± 0.12), partly as a result of inclement weather in 2006. Fecundity, estimated as female fledglings per breeding female, was relatively low for both translocated (0.9 ± 0.3) and resident females (0.3 ± 0.2). No mortalities occurred among radio-marked female ptarmigan during the 10- week breeding season, and the probability of annual survival for females in 2005-2006 was between 0.38 and 0.75. Translocations were successful because females survived, successfully nested, and recruited offspring during the establishment stage. Post-release monitoring provided useful demographic data in this study and should be a key component of translocation programs for wildlife restoration. Future population surveys and additional translocations may be recjuiied to ensure long-term viability of the reintroduced population of ptarmigan at Agattu Island. Received 31 July 2008. Accepted 16 July 2009. Reintroductions and other translocations are an important tool for restoring and enhancing plant and animal populations within their native 'Division of Biology, 116 Ackert Hall, Kansas .Stale University, Manhattan, KS 66506, USA. -U.S. Fish and Wildlife Service, Alaska Maritime National Wildlife Refuge, 95 Sterling Highway. Suite 1. Homer, AK 99603, USA. ■’Grouse Inc., 5572 North Ventana Vista Road, Tucson. AZ 85750, USA. ■‘Corresponding author: e-mail: bsanderc@ksu.edu geographic range (Griffith et al. 1989, Snyder et al. 1999, Ewen and Armstrong 2007). Greater effort may be allocated to moving animals if resources are limited, rather than to post-release monitoring needed to evaluate project success. However, post-release monitoring is critical because it allows estimation of the demographic parameters needed to evaluate status of reintro- duced populations. Population models can then be used to evaluate the effects of removals on source populations, and management strategies and 7 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 1, March 2010 environmental factors which may affect the eventual success or failure of newly established populations (SaiTazin and Legendre 2000, Arm- strong et al. 2007, Dimond and Armstrong 2007, Reynolds et al. 2008). Translocations have been widely used with continental populations of upland gamebirds. Release of wild-caught birds has been used to establish new breeding populations of White- tailed Ptarmigan {Lugopus leiiciira) (Braun et al. 1978, Hoffman and Giesen 1983, Clarke and Johnson 1990) and Ruffed Grouse (Bonasa umhelliis) (Moran and Palmer 1963, Kurzejeski and Root 1988). Reintroductions have been used to bolster declining or isolated populations of prairie grouse {Tympanuchiis spp.) (Snyder et al. 1999, Coates et al. 2006) and Greater Sage- Grouse (Centrocercus urophasianus) (Reese and Connelly 1997, Baxter et al. 2008). Translocations have also been an essential component of applied con.servation for populations of birds on oceanic islands, including waterfowl (Reynolds et al. 2008), flightless rails and parrots (Clout and Craig 1995, Jamieson and Wilson 2003), and songbirds (Komdeur 1994, Armstrong et al. 2002, Robertson et al. 2006, Ewen and Armstrong 2007). Restoration of insular populations poses particular challenges because life-history traits of island and mainland vertebrates can be highly divergent with differences in demography, behav- ior, and morphology (Stamps and Buechner 1985, Blondel 2000). For example, island populations of waterfowl, grouse, and .songbirds tend to breed later, lay smaller clutches of larger eggs, and have higher adult survival than mainland populations (Mercer 1967, Blondel 1985, Rohwer 1988, Atwood et al. 1990, Blondel et al. 1992, Wiggins et al. 1998). The Aleutian Archipelago of western Alaska consists of >450 isolated islands with large populations of breeding seabirds. Historically, the Aleutian Islands had no native terrestrial mammals west of Umnak Island (Murie 1959). Island populations of birds were negatively impacted by deliberate introductions of arctic fox (Alopex lagopus) by fur trappers between 1750 and 1940 (Maron et al. 2006). Depredation of eggs, young, and breeding birds led to steep population declines and local extirpation of waterfowl, seabirds, and terrestrial birds on islands where foxes were present (Bailey 1993). The U.S. Fish and Wildlife Service (USFWS) began systematic removal of foxes in 1949, as part of recovery efforts for Aleutian Cackling Geese {Branta hutchinsii leucopareia), as well as seabirds of conservation concern (Springer et al. 1977, Byrd et al. 1997, Williams et al. 2003). Rock Ptarmigan {Lagopus miita) are arctic- breeding grouse with a Holarctic distribution (Holder and Montgomerie 1993); eight subspecies have been described from the Aleutian Islands based on allopatric distributions and phenotypic differences in plumage coloration (Holder et al. 2000, 2004). Evermann’s Rock Ptarmigan (L. m. evermanni) is an endemic subspecies confined to the Near Islands group in the western Aleutian Archipelago. It is the only subspecies where males have a black nuptial plumage (Frontispiece), and has unique population genetics that distinguish it from all other subspecies of Rock Ptarmigan in the Aleutians and elsewhere (Holder et al. 2000). Evermann’s Rock Ptarmigan were historically present on all islands in the Near Islands group but introduced foxes reduced their range to Attu Island and an estimated population size of ~ 1 ,000 birds (Ebbert and Byrd 2002). Arctic fox are still present on the islands of Nizki-Alaid and Shemya but were eradicated from Agattu Island and Attu Island by 1979 and 1999, respectively (Bailey 1993, Ebbert and Byrd 2002). Evermann’s Rock Ptarmigan, due to their limited geographic range and small population size, have been designated birds of special management concern by the USFWS. We report results from a field study of a newly established island population of Evermann’s Rock Ptarmigan. Prior to our project, comprehensive surveys by the Alaska Maritime National Wildlife Refuge indicated that ptarmigan and foxes were absent on Agattu Island. Ptarmigan may disperse across marine waters (Zimmerman et al. 2005), but natural recolonization across the 28 km strait between the islands of Attu and Agattu had not occurred in the 25-year period since foxes were eradicated. Thus, we translocated wild-caught ptarmigan over a 4-year period and conducted intensive post-release monitoring at Agattu Island in the last 2 years of the project. Ecological studies of Rock Ptarmigan in the Aleutians have been limited to studies of molt and diet of L. m. gahrielsoni at Amchitka Island (Jacobsen et al. 1983, Emison and White 1988), and our project is the first demographic study of an island popula- tion of ptarmigan in Alaska. Our main goal was to compare the demographic rates of translocated and resident ptarmigan, and to evaluate the Killer et ill. • DEMOGRAPHY OF ISLAND PTARMICiAN 3 efficacy of translocations for restoration of insular populations of landbirds. A secondary goal was to compare life history traits between insular and mainland populations of Rock Ptarmigan that might influence translocation success for subspe- cies endemic to the Aleutian Islands. METHODS Study Area.— Anu (52.85° N, 173.19° E; 89,279 ha) and Agattu (52.43° N, 173.60° E; 22,474 ha) are in the Near Islands group in the western range of the Aleutian archipelago and part of the Alaska Maritime National Wildlife Refuge. Attu is a mountainous island with steep hillsides rising from sea level to elevations >600 m (highest point: 861 m). Most of the land mass of Agattu Island is <230 m in elevation but a mountain range composed of seven major sub- massifs lies along the north side and extends from Armeria Bay eastward to Krugloi Point (highest point: 693 m). The dominant plant community in the Near Islands is maritime tundra because the climate is consistently cool, wet, and windy (Maron et al. 2006). Mean minimum and maximum temperatures are 6.5 and 9.2° C, total precipitation is 6.8 cm per month, and wind velocities average 42 kph in the 3-month period from June to August (climate data from Shemya Island, 30 km northeast of Agattu Island). Glau- cous-winged Gulls (Larus glaucescens) and Com- mon Ravens (Corvus cora.x) were potential predators of eggs and chicks, whereas Peregrine Falcons (Falco peregrinus) and Snowy Owls (Nyctea scandiaca) were a threat to adult ptarmigan (Gibson and Byrd 2007). Agattu Island does not have introduced species of rats {Rattus spp.), which are a problem elsewhere in the Aleutian Islands (Major et al. 2006). Translocations. — Five translocations from Attu Island were conducted to reintroduce EvermaniTs Rock Ptarmigan to Agattu Island: four during May to June, 2003-2006 and one during Septem- ber 2003. Translocations could not be conducted at other times of year in this remote area because logistical support was not available and environ- mental conditions were hazardous. Ptarmigan were live-captured at Attu Island using noose poles, noose carpets, and ground nets. Birds were held for up to 48 hrs in fiberglass transport containers and provided small pieces of melon as a source of food and water. The 28-km strait between the islands of Attu and Agattu required 3 hrs to traverse on the research ves.sel M/V Tiglax. Ptarmigan were immediately released upon arrival at Agattu Island at one ol three sites: McDonald Cove on the east coast (2003), Karab Cove on the south coast (2()()4-2()()5), or Binnacle Bay on the north coast (2006). Loss of body mass of translocated ptarmigan was measured by subtracting mass at release from mass at capture. We adjusted the difference in body mass by egg mass (21 g per egg. Holder and Montgomerie 1993) in ca.ses where females laid eggs during the holding period. Necropsies of birds that died during handling or after release were conducted shortly after recovery of the carcass by a wildlife veterinarian (W. P. Taylor, Alaska Fish and Game Department). Field Methods. — We conducted post-release monitoring at Agattu Island from late May to mid-August in 2005 and 2006 to estimate the fecundity and survival of female Evermann’s Rock Ptarmigan (Kaler 2007). Our study sample included newly translocated birds and resident ptarmigan encountered at Agattu Island. The resident population included marked birds surviv- ing from previous translocations and their un- marked offspring. All ptarmigan released were individually banded and color banded with batch marks (2003-2004) or individual band combina- tions (2005-2006) at first capture. Yearlings and adults were distinguished by patterns of pigmen- tation and shape of the wing primaries (Weeden and Watson 1967). Females were fitted with radio transmitters attached with either a bib (2005: 15 g. Telemetry Solutions, Concord, CA, USA) or a necklace harness (2006: 6 g, Holohil Ltd., Caip, ON, Canada). Radio transmitters had a battery life of 12-18 months and were equipped with mortality switches to facilitate detection of dropped transmitters and mortality events. Ra- dio-marked birds were tracked with a portable radio receiver (R2000, Advanced Telemetry Systems, Isanti, MN, USA) and 3-element Yagi antenna. We started radio-tracking immediately after release and relocated females daily by radio triangulation until nests were located. All points were recorded in Universal Transverse Mercator (UTM) coordinates using a hand-held Global Positioning System receiver (Garmin GPSmap 76; Garmin International, Olathe, KS, USA). Nests of translocated ptarmigan were located by radiotelemetry whereas nests of resident ptarmigan were located by searching suitable nest habitat around roosts of males, which often acted as sentinels (Holder and Montgomerie 1993). 4 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. I, March 2010 Unmarked females were captured with noose poles on territories or with dip nets at nest sites, and were marked with radio transmitters. We did not examine the eggs if a nest was discovered during egg-laying and returned after the clutch had been completed. We counted the eggs, if a nest was discovered during incubation, measured (± 0.1 mm) length (L) and width (W) of all eggs, and floated the eggs in a small cup of lukewarm water to estimate stage of incubation (Westerskov 1950). Buoyancy of eggs was related to stage of incubation as; flat on the bottom of the container = 0 days, 45° = 5 days, 90° = 10 days, floating at surface = 1 3 days, and a ~ 1 8 mm diameter circle protruding above the water surface = 18 days. The onset of hatching was detected by days 19, 20, and 21 of incubation by tapping, raised star- shaped pips in the egg shell, and hole-pipped eggs, respectively. Female nest attendance was checked every 3^ days during incubation by triangulating the radio signal at distances >30 m from the nest site. Nests were visited every 1-2 days around the predicted hatching date to capture young and ascertain nest fate. We considered nests to be abandoned if a female ceased nest attendance during egg-laying or incubation, depredated if the eggs were destroyed or removed before the expected hatch date, and successful if egg shells with detached membranes were left in the ne,st bowl after the chicks had departed or if we captured at least one chick at or near the nest. We counted offspring in broods < 15 days of age with binoculars from >50 m and flushed broods 2-3 times at 15-25 days of age to calculate survival of young after hatching. Demographic Parameters. — Demographic per- formance of ptarmigan is, at times, affected by female age or annual conditions (Watson 1965, Steen and Unander 1985, Gardarsson 1988, Novoa et al. 2008). Our sample sizes were small and we pooled data across age clas.ses and years to compare the demography of translocated and resident Rock Ptarmigan. We calculated 10 demographic parameters for breeding females (after Sandercock et al. 2005). (I) Date of clutch initiation was calculated by backdating from stage of egg-laying, stage of incubation, or the date of hatching. We assumed that Rock Ptarmigan laid one egg/ day. started incubation on the penultimate egg. and had an average incubation period of 2 1 days (Holder and Montgomerie 1993). (2) Total clutch laid (TCL) was the total number of eggs laid in the clutch, recorded during the first visit to the nest during incubation. (3) Egg volume (U) was estimated from linear measurements of ptarmigan eggs using V = ALW", where A = 0.49 for Willow Ptarmi- gan (L. lagopus), L = egg length, and W = egg width (Sandercock and Pedersen 1994). We calculated average egg volume per clutch because eggs within a nest were not independent observations. (4) Nest survival until hatching (NEST) was the probability that at least one egg hatched and produced a chick that departed the nest. Values of NEST < 1 included total clutch losses due to abandonment and predation. (5) Renesting (RENEST) was the probability that a female laid a replacement clutch, conditional upon loss of her first clutch. (6) Chicks/egg laid (C/E) was the percentage of eggs laid that eventually hatched and produced chicks that left the nest. C/E was calculated only for nests that survived incubation and hatched at least one egg. Values of C/E < 1 included partial clutch losses due to eggs that disappeared during incubation and eggs that survived incuba- tion but failed to hatch. (7) Brood survival until fledging (FLED) was the probability that at least one chick survived from hatching until fledging at 15-25 days of age. Eledglings completed growth of Juvenal feathers by 2 weeks and began to evade observers with short flights instead of running and hiding. Values of FLED < 1 were due to total brood failure, which was readily ascertained by the behavior and movements of females. If females lost their brood, they flushed during approach by an observer or Joined small groups of failed breeders. (8) Fledglings/chick hatched (F/C) was the percentage of hatched chicks that left the nest and survived until fledging at 15-25 days. F/C was calculated only for broods that survived the brood-rearing period and fledged at least one chick. (9) Seasonal survival of females was calculated by monitoiing radio-marked females over the 10-week breeding season. We left-censored a 2-week acclimation period to control for potential effects of higher mortality rates immediately following stress of capture. Kalcr et al. • DEMOGRAPHY OF ISLAND PTARMIGAN 5 transport, release in a new environment, and attachment of the radio transmitter. (10) Annual survival of females was calculated from the return rates of radio-marked birds from 2005 to 2006. We conducted intensive searches of territories for all radio-marked females in 2006. Birds relocated alive in 2006 were known survivors whereas radio transmitters recovered near scattered feath- ers or carcass remains were considered known mortalities. Female fate was un- known if birds were not relocated or if radio transmitters were recovered without accom- panying remains. Estimation of Fecundity. — ^We calculated a synthetic estimate of fecundity (F) as the expected number of female fledglings produced per breed- ing female per year. We controlled for possible nest losses before discovery by an observer, and differences in renesting rates, by first calculating the number of eggs surviving until hatching for each breeding female (F): F= (TCLi xNEST) + ([1 -NEST] X RENEST x TCL2 x NEST) where TCL„ = total clutch laid, subscripts 1 and 2 are values for first nests and renests, NEST = probability of nest survival, and RENEST = probability of renesting after clutch loss. We calculated fecundity (F) as: F = F X C/E X FLED x F/C x 0.5 where F = the number of surviving eggs, C/E = percentage of chicks hatched/egg laid, ELED = probability of brood survival until fledging, E/C = percentage of fledglings produced/chick leaving the nest, and 0.5 is the proportion of females based on a 1:1 sex ratio. Statistical Analysis. — Comparisons of demo- graphic parameters between translocated and resident females were conducted using statistical procedures in Program SAS (Version 8.1, SAS Institute Inc., Cary, NC, USA). Sample sizes differed among some components of fecundity because we lacked complete information for a few nests and losses to predators caused attrition over the breeding season. Continuous data were compared with two sample r-tests and frequency data were compared with contingency tests. Post hoc comparisons of survival rates were conducted with Program CONTRAST (USGS, Patuxent Wildlife Research Center, Laurel, MD, USA). All tests were two-tailed and considered signili- cant at a < 0.05. Means ± SE are presented. We estimated daily survival rates of nests and broods during the incubation and brood-rearing periods with the nest survival procedure of Program MARK (Version 4.1, G. C. White, Colorado State University, Fort Collins, CO, USA). We collected data for nests for a 63-day exposure period from 1 June until 2 August, and tor broods for a 38-day exposure period from 28 June until 5 August. We created encounter histories for each nest and brood suiwival analyses with five types of information (Dinsmore et al. 2002): ( 1 ) the day the nest or brood was found {k), (2) the last day it was known to be active (/), (3) the last day it was checked (/??), (4) the fate (/) where 0 = successful and 1 = depredated, and (5) the number of nests or broods with the same encounter history (/?). All assumptions of the nest survival model were met in this study (Kaler 2007). We considered five candidate models in our analyses of nest and brood survival. We expected that translocated females might have lower nest survival rates than resident females, and daily survival rates might vary during the breeding season. Survival rates of ground nests of precocial birds might be expected to vary seasonally if vegetative cover changes, poorly concealed nests are selectively destroyed, or predator activity is affected by nest densities and availability of alternative prey (Klett and Johnson 1982, Dins- more et al. 2002, Wilson et al. 2007). Similarly, brood survival rates might be lowest after nest departure and improve as chicks begin to thermoregulate and fly. Thus, our global model included the effects of female group (translocated vs. resident), a linear effect of time, and the interaction of these factors (Sgroup x linear)* We also fit reduced models with main effects (“^group + linear)’ sillglc faCtOrS (Sgroup’ '^tinie)' tUld a model that constrained daily survival to be a constant (Sconstam)- Global models are saturated in the nest survival model, and adjustments for overdispersion are not possible because the variance inflation factor (r) is not identifiable (Dinsmore et al. 2002). All models were fit to the data using design matrices and the logit-link function in Program MARK. Model selection was based on an information theoretic approach (Burnham and Anderson 1998). The model with the lowest Akaike's 6 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. I. March 2010 Information Criterion corrected for small sample sizes (AIC(.) value was the best fit and models with AAIC,. < 2 were equally parsimonious. We based inference on differences in deviance in cases where the difference in number of param- eters was one (AA' = 1 ). We used ratios of Akaike weights to quantify the extent of support among models (m’,/m7). We calculated parameter estimates with uncon- ditional variances with the model-averaging procedure of Program MARK to examine season- al variation in daily survival rates. We extrapo- lated daily survival rates using A"''"', where S = the daily survival rate of nests or broods, and chir — the duration of the nesting or brood-rearing stages to obtain stage-specific estimates of the probabilities of nest and brood survival. For example, the duration of an eight-egg nesting attempt was 28 days, calculated as the duration of the egg-laying (8 days) and incubation periods (21 days) with incubation starting after laying of the penultimate egg (—1 day). The duration of the brood-rearing period was 20 days. Variances of stage-specific estimates were calculated using the delta method, following formulae and methods described by Powell (2007). We used parametric bootstrapping to obtain confidence intervals for our synthetic estimates of fecundity {F) (Gotelli and Ellison 2004). We generated bootstrap distributions using tools of Program MATLAB (Version 6.5, MathWorks, Natick, MA, USA). Estimates of means and variances were taken directly from our parameter estimates with the exception of the probability of renesting where the variance was estimated as Var(/?) = p{ \ — p)ln, where p is the probability and // is the total sample size. Total clutch laid was modeled as draws from a nonnal distribution whereas probabilities were modeled as draws from beta distributions to bound draws within the range of 0 and 1. We drew a set of seven demographic rates at random for each bootstrap replicate, combined the estimates to calculate fecundity, and repeated these steps for 1,()()() bootstrap iterations. Bootstrapping of fecundity was con- ducted .separately for translocated, resident, and pooled females. We generated a di.stribution of differences between random pairs of bootstrap values for translocated and resident females to compare groups, and calculated a P-value by recording the number of boot.strap differences that were greater than the ob.served difference between means (Gotelli and Elli.son 2004). TABLE 1. Number of Evermann’s Rock Ptarmigan translocated from Attu Island to Agattu Island, Aleutian Archipelago, Alaska, in May to June 2003-2006 and September 2003. Year Males Females Juveniles Totals 2003 1 1 13 2 26 2004 1 1 16 0 27 2005 4 10 0 14 2006 5 3 0 8 Totals 31 42 2 75 RESULTS We translocated 75 Evermann’s Rock Ptarmi- gan from Attu Island to Agattu Island over a 4- year period (2003-2006, Table 1). We monitored reproduction and survival rates of 29 radio- marked female ptarmigan (translocated: n — 13; resident: n = 16) from late May to mid-August during a 2-year radiotelemetry study (2005- 2006). The proportion of females that were adults was higher among translocated (92%, n = 13) than resident birds (56%, n = 16, Eisher’s Exact test, P — 0.044), but age-classes were pooled for analyses because sample sizes were small. Our study sample included 28 nests (translocated: n — 10; resident: n = 18, including 2 renests) and 21 broods (translocated: n = 8; resident: // = 13). Translocations. — ^Female ptarmigan translocated from Attu to Agattu Island in 2005-2006 {n = 13) were in breeding condition: 30.8% had partial brood patches and 46.2% laid an egg in the holding containers during transport. Female mass at capture averaged 568 ± 16 g (/? = 31) and translocated females lost an average of 9% of their total body mass (x ± SE loss: -57 ± 15 g, /; = 12). Only one mortality occuired during the holding and transport periods in the 4-year period. A juvenile male died during holding but we were unable to assign cause of death at necropsy. Two mortalities occuired at Agattu Island during post-relea.se monitoring in the last 2 years of the .study. In both cases, radio- marked females died within 2 weeks of release. One female was recovered 2 days after release 390 m from the release site and necropsy indicated that she drowned in a small tundra pool. A second female was recovered 10 days after relea.se 914 m from the release site; evidence at the recovery site indicated she was killed by a raptor. Release site had little effect on translocation success because surviving birds quickly dispersed to settle in alpine habitats at the northern part of Agattu Island. Kalcr cl al. • DEM(X}RA1M 1 Y OF ISLAND PTARMIGAN 7 TABLE 2. Demographic parameters of translocated and resident Evermann's Rock Ptarmigan at Agattu Island. Aleutian Archipelago. Alaska. 2()()5-2006. The first three parameters were calculated lor lirst nests only. Data are presented as means ± SE {n). Parameter Translocated Resident Pooled Statistic P £ Date of clutch initiation (days) 16 June ± 1.0 (10) 8 June ± 1 .3 ( 16) 11 June ± 1.2 (26) u It O.OOl Total clutch laid (eggs) 6.8 ± 0.3 (10) 8.3 ± 0.2 (15) 7.7 ± 0.2 (25) InJ 11 0.001 Mean egg volume (cm’) 23.1 ± 0.6 (8) 23.3 ± 0.2 (16) 23.2 ± 0.2 (24) = 0.41 0.68 Probability of nest survival 0.724 ± 0.166 (10) 0.500 ± 0.155 (18) 0.588 ± 0.1 18 (28) = !•« 0.18 Probability of renesting 0 (2) 0.50 (4) 0.33 (6) Fisher’s Exact 0.47 Chicks/egg laid (%) 85.7 ± 4.6 (8) 82.3 ± 5.1 (13) 83.6 ± 3.5 (21) My = -0.5 0.66 Probability of brood survival 0.849 ± 0.139 (8) 0.251 ± 0.123 (13) 0.470 ± 0.1 19 (21) f 1 = 10.4 ().()()2 Fledglings/chick hatched (%) 48.5 ± LI (7) 53.9 ± 3.8 (4) 50.4 ± 0.7 (1 1) ry = 0.4 0.73 Fecundity (female fledglings/ breeding female/year) 0.9 ± 0.3 0.3 ± 0.2 0.5 ± 0.2 Bootstrap 0.48 Timing of Clutch Initiation, Total Clutch Laid, and Egg Volume. — ^All translocated females that survived the 2-week acclimation period initiated clutches on Agattu Island. The interval between release and onset of egg-laying was 13.1 ± 0.9 days for translocated females (n = 10). The average date of clutch initiation was 16 June for translocated females, which was 8 days later than an average date of 8 June for resident hens (Table 2). Total clutch laid in first nests by translocated females (6.8 eggs) was 1.5 eggs smaller than the average clutch size of resident hens (8.3 eggs). Mean egg volume did not differ between resident and translocated females, and averaged 23.2 cm^ overall. Nest Survival, Renesting, and Chicks! Egg. — ^We used encounter histories from 28 nests (10 translo- cated and 18 resident, 2 females included in both years) to estimate daily survival rates for nests. Nests were monitored over a 63-day exposure period (day 1 = 1 June). The minimum AIC^. model contained a linear effect of time over the breeding season (5|inear, Table 3). The minimum AIC,. model had 1. 3-2.4 times greater support than the other candidate models but all five models were equally parsimonious (AAIQ. < 1.8). The slope coefficient for the linear effect indicated that nest survival declined over the breeding season (from model 5iinear: P —0.07 ± 0.04, logit scale), particularly in the latter part of TABLE 3. Model selection results for daily survival rates (S) of nests and broods of Evermann’s Rock Ptarmigan at Agattu Island, Aleutian Archipelago, Alaska, 2005-2006. Model fit was described by the number of parameters [K). model deviance (Dev), Akaike’s Information Criterion corrected for small sample size (AlC^), deviations from the minimum-AlCc model (AAIQ.), and Akaike weights (w,). Models for daily survival (S) included comparisons of resident vs. translocated birds (group), a seasonal effect of time (linear), factorial (X) and main effects models (+), and a reduced model with no effects (constant). Exposure periods were 63 days for nesting (day 1 = 1 June) and 38 days for brood-reaiing (day 1 — 28 June). Model K Dev AlC,. AAtC, W, Nests 2 51.5 55.5 0.00 0.300 ^linear ‘^constant 1 54.0 56.0 0.54 0.229 3 50.2 56.3 0.76 0.205 ■-^group + linear c 4 48.9 57.0 1.52 0.140 "-^group X linear C ^ group 2 53.2 57.2 1.74 0. 1 26 Broods e 2 43.2 47.2 0.00 0.443 •-'group •^group X linear ‘^group + linear C ‘-'constant •^linear 4 39.7 47.9 0.68 0.316 3 1 43.1 49.3 2.02 0.162 49.7 51.8 4.52 0.046 2 48.4 52.4 5.20 0.033 8 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122. No. 1, March 2010 1.0 lO 2. 0 cc 03 > > 0.6 (A) Nests N \ \ \ \ Translocated Resident \ J 3 1.0 - (/) 0.9 ■ 03 Q 0.8 ■ 0.7 ■ 0.6 ■ 10 20 30 40 50 60 (B) Broods 1 1 \ \ \ \ 0 5 10 15 20 25 30 35 Day of exposure period LIG. 1. Daily survival rates for nests and broods of translocated and resident female Rock Ptarmigan at Agattu Island, Alaska, 2005-2006 (n = 27 nests and 21 broods). Estimates were calculated by model-averaging in Program MARK. Exposure periods were 63 days for nesting (day 1 = 1 June) and 38 days for brood-rearing (day 1 = 28 June). the season (Fig. lA). Estimates of daily survival rates (from model 5group) revealed that nest survival was similar for translocated {S = 0.988 ± 0.008) and resident females (5 = 0.976 ± ().01 1 ), and was high overall (from model 5con,stant: S = 0.981 ± 0.007). Extrapolation of daily survival rates over the duration of a nesting attempt (27 to 28 days for 7- to 8-egg clutches) yielded estimates of the probability of nest survival that were 0.724 ± 0.166 for translocated females, 0.500 ± 0. 1 55 for resident females, and 0.588 ± 0.1 18 for all females combined. Six first nesting attempts failed because of nest predation or abandonment. Two resident females that renested lost their first nests 2 days before on.set of incubation and at 12 days of incubation, and initiated renests after intervals of 13 and 1 1 days, respectively. Clutch size of the two renests was 5 and 6 eggs. Translocated and resident females that did not renest lost their nests at more advanced stages of incubation (12-23 days, n = 4). The percentage of viable eggs that hatched and produced 1-day old chicks did not differ between resident and translocated females, if the nest survived incubation, and was 83.6% overall. Brood Survival and Fledglings per Chick. — We used encounter histories from 21 radio-marked females that successfully hatched young (8 translocated and 13 resident, 1 female included in both years) to estimate daily survival rates for broods. Broods were monitored over a 34-day exposure period (day 1 = 28 June). The minimum AIC^. model included an effect of group with a difference between translocated and resident birds (5group' Table 3). Two additional models with interactive and additive effects between group and a linear effect of time (‘Sgroup X Unean ‘^group + linear) were equally parsimonious or nearly so (AAIC^. < 2.02), but the minimum AIC^. model had 1.4-2. 7 times greater support. The slope coefficient for the effect of group (from model 5group x linear: P = 6.87 ± 6.71, logit scale) was larger than the linear effect of time (P = 0.53 ± 0.51) or the interaction of the two factors (p = —0.57 ± 0.51). Brood survival of translocated females was initially low and in- creased over the exposure period, whereas resident birds had moderate brood survival (Eig. IB). Overall estimates of daily survival rates (from model Sgroup) indicated brood survival was higher among translocated females [S = 0.992 ± 0.008) than resident hens {S = 0.933 ± 0.022). Daily survival rates of broods for all females combined (from model ^con.stant: ^ = 0.963 ± 0.012) were not different from the daily survival rates of nests [S = 0.98 1 ± 0.007, x? = 1.71, P = 0.19). Extrapolation of daily survival rates over the duration of brood-rearing (20 days) yielded estimates of the probability of brood survival that were 0.849 ± 0.139 for translocated females, 0.251 ± 0.123 for resident females, and 0.470 ± 0.1 19 for all females. The percentage of chicks that successfully fledged if the brood survived the brood-rearing stage did not differ between resident and translocated females, and was 50.4% overall. Fecundity. — We combined our estimates of demographic parameters to calculate a synthetic estimate of fecundity that controlled for nest exposure before nest discovery and productivity from renests. Fecundity tended to be higher among translocated birds {F = 0.9 ± 0.3 female fledgling.s/female) than residents [F = 0.3 ± 0.2), but the difference was not significant (Table 2). Kaler ct al. • DEMOGRAPHY Ol' ISLAND PTARMIGAN 9 Survival of Females. — We monitored seasonal survival during the nesting and brood-rearing period for 26 female ptarmigan (10 translocated and 16 resident). All females survived the 10- week breeding season and no mortalities were observed during either 2005 or 2006. We searched during the 2006 breeding season for 16 females that had been radiomarked in 2005 to assess overwinter female survival. Of the 16 females, 37.5% were detected alive, 31.3% were dead recoveries, and 31.3% were not relocated. The frequency of known mortalities was equivalent in translocated (30%, /; = 10) and resident females (33%, n = 6, Fisher’s Exact test, P = 1.0). Surviving females had strong site fidelity and distance between nest locations in consecutive years averaged 123 ± 34 m (/; = 5). We intensively searched female teixitories and our minimum estimate of annual survival would be 0.38 if all missing females had died. However, it missing females survived to disperse elsewhere on the island, annual survival of females could be as high as 0.75. DISCUSSION Our field study is one of the first investigations of the effects of translocation on an insular population of an upland gamebird, and we provide the first estimates of demographic parameters for any of the subspecies of Rock Ptarmigan that are endemic to the Aleutian Archipelago. Our major conclusions were; ( 1 ) translocations had negative impacts on female performance with short-term increases in female mortality, delays in the timing of clutch initiation, and reductions in clutch size; (2) settlement was rapid and unaffected by location of the release sites; (3) translocated and resident females had similar rates of fecundity and survival during the period of population establishment; and (4) island and mainland populations appear to have similar demography, which may have contributed to the success of our reintroductions. Effects of Translocation on Fecundity. — ^The negative impacts of translocation from Attu Island to Agattu Island were transitory and had little net effect on productivity. Logistics constrained us to hold birds for up to 48 hrs, but mortality rates were low during holding (1.4%) and comparable to translocations of other grouse (0%, Hot! man and Giesen 1983; 2.1%, Baxter et al. 2008) and waterfowl (0%, Reynolds et al. 2008). Rock Ptarmigan lost an average of 9% of body mass during transport, and similar rates of loss have been reported for translocated Rulfed Grouse (10%, Kurzejeski and Root 1988) and Laysan Ducks (Anas lay.sanensLs) (6%, Reynolds et al. 2008). The only mortalities among radio-marked Rock Ptarmigan were two females that died shortly after release. Mortality rates ol wildlife are often higher immediately after translocation, because of the physiological stress of captivity or elevated predation risk due to lack of familiarity with a new environment or attachment of a radio transmitter (Musil et al. 1993, Clout and Craig 1995, Armstrong et al. 1999, Baxter et al. 2008). Ptarmigan settled quickly into alpine habitats during the establishment stage and all surviving females initiated clutches shortly after release. Future translocations of ptarmigan in the Aleutian Islands can select release sites based on safety of coastal landings without concerns about proximity to suitable habitats. In other species, translocated birds often range widely after release in search of suitable habitat or conspecifics (Musil et al. 1993, Amistrong et al. 1999, Coates et al. 2006), and long distance movements can contribute to higher mortality rates during the acclimation period (Kurzejeski and Root 1988). Our releases may have had greater success because post-release movements were constrained by the limited area of alpine habitat on Agattu Island and habitat quality was restored to historical conditions by eradication of introduced predators. Transloca- tions of birds in breeding condition during the nesting season facilitated rapid settlement and early reproduction, similar to previous releases of ptarmigan (Braun et al. 1978). Translocations of juveniles during the nonbreeding season might have less impact on a source population but could delay establishment because juveniles must sur- vive to the next breeding season and, at times, have lower success as first time breeders (Kurze- jeski and Root 1988, Sarrazin and Legendre 2000, Robertson et al. 2006, Baxter et al. 2008, Reynolds et al. 2008). Translocated females delayed clutch initiation and laid clutches that were 1 .5 eggs smaller than nests of resident Evermann’s Rock Ptarmigan. Reduced clutch size could have been due to interruptions in egg-laying if translocated females were determinate egg-layers that failed to replace missing eggs laid in clutches at Attu Island or in the handling crates during transit (Sandercock 1993). Alternatively, smaller clutches could have been due to the physiological stress of transloca- tion. White-tailed Ptarmigan breeding in years of 10 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 1. March 2010 harsh environmental conditions delay egg-laying and take more incubation breaks to maintain body condition (Wiebe and Martin 2000). Only resident females renested after nest failure in our study. However, females that failed to renest lost their clutches at more advanced stages of incubation when the probability of renesting was presumably low (Robb et al. 1992). Evermann’s Rock Ptarmigan laid fewer eggs in renests than first nests, similar to other populations of ptarmigan (Brodsky 1986, Cotter 1999, Sandercock et al. 2005) . Translocated and resident ptarmigan differed primarily in demographic parameters that were recorded at the start of the nesting cycle. The potential demographic impacts of differences in timing of laying and clutch size were largely offset by reproductive losses. Nest and brood survival are expected to improve over the season if vegetative cover increases or predators locate offspring in vulnerable sites earlier in the season (Klett and Johnson 1982, Wilson et al. 2007). Survival of ptarmigan broods at Agattu Island increa.sed during the breeding season, but survival of nests was lowest during the latter part of the exposure period. The lowest rates of daily survival for ptarmigan nests and broods coincided with the hatching period of Aleutian Cackling Geese and increased foraging activity by Glaucous-winged Gulls at inland sites. Gulls primarily hunted for goose eggs and goslings, but incidental predation could have reduced the fecundity of ptarmigan. Predation could potentially be a limiting factor for ptarmigan populations becau.se gull populations are maintained at high levels in the Aleutian Islands by access to refu.se from fish processing facilities (Gib.son and Byrd 2007). Ptarmigan young have limited ability to ther- moregulate or evade predators in the week after hatching (Scherini et al. 2003, Hannon and Martin 2006) . Increa.ses in late .season survival of broods could have been due to development of homeo- thermy and mobility among ptarmigan young or to more favorable climatic conditions during the latter part of the exposure period (Aulie 1976, Steen and Unander 1985). Hatching and fledging success were high, if ptarmigan young survived incubation or brood-rearing, and there was no evidence of deleterious effects caused by inbreeding depression (Briskie and Macintosh 2004). Synthesis of the major components of repro- ductive effort showed that fecundity of Ever- mann's Rock Ptarmigan was relatively low at <1 fledgling per breeding female, and that translo- cated females did not have lower fecundity than resident females in our reintroduced population. If any difference occurred, it was that translocated birds tended to have higher productivity, similar to previous reports for Takahe (Porphyria hoch- stetteri) (Jamieson and Wilson 2003). Some variation in productivity could have been ex- plained by female age class and annual variation in demographic rates. Our sample of translocated birds included a high proportion of adults, which at times, have higher reproductive success than yearling Rock Ptarmigan (Steen and Unander 1985, but see Cotter 1999, Scherini et al. 2003). Similarly, annual fecundity of Rock Ptarmigan can be variable and influenced by local environ- mental factors, including food, predation, and climate (Watson 1965, Gardarsson 1988, Novoaet al. 2008). If conditions are favorable for several consecutive years, ptarmigan numbers should exhibit the exponential growth often reported for populations established by island translocations (Moran and Palmer 1963, Klein 1968, Komdeur 1994, Armstrong et al. 2002). Survival Rates. — Seasonal survival of female Evemiann’s Rock Ptarmigan during the breeding period was 100% as no mortalities were observed. Mortality rates during the breeding season are higher in populations of Rock Ptarmigan where mammalian predators and Gyrfalcons (Falco riisticohis) are a threat (23%, Gardarsson 1988; 14%, Cotter et al. 1992; 12%, Scherini et al. 2003). Annual survival of female Evermann’s Rock Ptamiigan was between 0.38 and 0.75 for one interval (2005-2006), and all losses occuired during the fall or winter after we had departed from Agattu Island. Raptors capable of killing adult ptarmigan occun'ed at relatively low densities during the breeding season (Kaler 2007). Seasonal mortality rates of mainland populations of Willow Ptarmigan are usually highest in autumn during raptor migration (Smith and Willebrand 1999, Hannon et al. 2003). Seasonal mortality rates of Evermann’s Rock Ptarmigan may have been higher outside of the breeding .season if Snowy Owls and Peregrine Ealcons relied on coastal seabirds during summer but spent more time foraging at inland sites during fall and winter (Gibson and Byrd 2007). Our con.servative estimate of annual sur- vival of 0.38 for Evermann’s Rock Ptarmigan is comparable to estimates of 0.37 to 0.43 for arctic and alpine populations of Willow Ptarmigan (Sandercock et al. 2005). Kidcr a al. • DEMO(:}RAI’I I Y OF ISLAND P FARMICiAN 1 1 TABLE 4. Estimates of five demographie parameters for island (1) and mainland (M) populations of Rock Ptarmigan. Location Pop. t.al Dale of clinch initiation Clutch si/e Hrobabilily of nest survival Probability of brood survival Probability of adult survival Source Alaska. USA I 5.1° N 8 June 8.3 0.59 0.47 0.38-0.75 This study Iceland 1 66 N >10.4 >0.90 0.26-0.82 Gardars.son 1988 Norway 1 78 N 7.5 0.44 Steen and Unander 1985 Italy M 46 N 10 June 6.8 0.50 0.18-0.67 Scherini et al. 2003 Scotland, UK M 57° N Mid June 6.6 0.90 0.62 Watson 1965 Alaska, USA M 65° N 7.0 0.67 0.81 0.26-0.71 Weeden 1965 NWT, Canada M 68 N 1 1 June 8.8 Brodsky 1986, 1988 NWT, Canada M 69 N 9 June 8.7 0.64 0.53 0.43-0.82 Cotter 1991, 1999 Life-history Traits of Ptarmigan. — Evermann’s Rock Ptarmigan are endemic to oceanic islands and are a unique insular form with a black nuptial plumage and distinct population genetics (Holder et al. 2000, 2004). We found little evidence that selection on demography has contributed to development of an insular syndrome in the life- history traits of this subspecies, despite distinct differences in morphology. Compared to main- land and island populations of Rock Ptarmigan elsewhere in their Holarctic range, Evermann’s Rock Ptarmigan had similar dates of clutch initiation in mid-June, intermediate values for clutch size, and comparable estimates of the probabilities of survival for nests, broods, and adults (Table 4). Evolution of an insular syn- drome among the life-traits of vertebrates is frequently attributed to differences in predation and climate regimes between island and mainland sites (Mercer 1967, Stamps and Buechner 1985, Blondel 2000). Isolation reduces the diversity of predator communities and climatic conditions can be less variable due to proximity to coastal waters. These two ecological factors may have had less influence on Evermann’s Rock Ptarmigan because the Near Island group supports several species of avian predators and has extreme climatic condi- tions associated with an exposed location in the northern Pacific Ocean (Maron et al. 2006, Gibson and Byrd 2007). Most of the geographic variation in life-history traits of Rock Ptarmigan and congeneric species appears to be driven by latitudinal and altitudinal gradients in ecological conditions in arctic and alpine environments (Sandercock et al. 2005, Novoa et al. 2008). CONSERVATION IMPLICATIONS Ten features of our project have been identified as factors predicting translocation success for teirestrial vertebrates (Griffith et al. 1989, Snyder et al. 1999, Ewen and Armstrong 2007). Ever- mann’s Rock Ptarmigan are a herbivorous, native gamebird with high reproductive potential that breed as yearlings. We translocated multiple batches of wild-caught birds over a 4-year period, and birds were released into high quality habitat in the core of their historical range which had been restored by eradication of introduced predators. High survival during the breeding season, suc- cessful nesting, and recruitment of young birds indicated successful reestablishment of an island population. Small initial populations can risk extinction because of demographic stochasticity, loss of genetic diversity, and inbreeding depres- sion. However, annual surveys in June using playbacks of male vocalizations recorded 26-27 territorial male Evermann’s Rock Ptarmigan each year at Agattu Island in 2008-2009, following completion of our field study. An adaptive management strategy based on regular surveys and supplementary translocations will ensure the long-term population viability of Evermann’s Rock Ptarmigan at Agattu Island. ACKNOWLEDGMENTS R. B. Benter, J. K. Nooker. M. A. Schroeder, and W. P. Taylor assisted with capture and translocation of ptarmigan from Attii Island to Agattu Island. L. A. Kenney and G. T. Wann were dedicated research assistants for two field seasons at Agattu Island. Captain K. D. Bell and the crew of the M/V Tiglax provided us with safe passage to the outer Aleutians, and G. V. Byrd coordinated logistical support for our field project. The LORAN Station ot the U.S. Coast Guard at Attu Island provided housing, meals and other support during our capture effort. D. P. Armstrong and L. N. Ellison offered constructive reviews of the manuscript. Funding for field work for this project was provided by the USFWS, Alaska Maritime National Wildlife Refuge: and the U.S. Missile Defen.se Agency. The Division of Biology at Kansas State University provided financial support to R. 12 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. I, March 2010 S. A. Kaler and B. K. Sandercock. Capture and handling of birds was conducted under protocols approved by the Institutional Animal Care and Use Committee at Kansas State University, and capture and transport permits from the State of Alaska. LITERATURE CITED Armstrong. D. P., I. Castro, and R. Griffiths. 2007. Using adaptive management to determine require- ments of re-introduced populations: the case of the New Zealand Hihi. Journal of Applied Ecology 44:953-962. Armstrong, D. P., I. Castro, J. C. Alley, B. Feenstra, AND J. K, Perrott. 1999. Mortality and behaviour of Hihi, an endangered New Zealand honeyeater, in the establishment phase following translocation. Biologi- cal Conservation 89:329-339. Armstrong, D. P., R. S. Davidson, W. J. Dimond, J. K. Perrott, I. Castro, J. G. Ewen, R. Griffiths, and J. Taylor. 2002. Population dynamics of reintroduced forest birds on New Zealand islands. Journal of Biogeography 29:609-621. Atw'ood, J. L., M. J. Elpers, and C. T. Collins. 1990. Survival of breeders in Santa Cruz Island and mainland California Scrub Jay populations. Condor 92:783-788. Aulie, A. 1976. The pectoral muscles and the development of thermoregulation in chicks of Willow Ptarmigan (Lagopus lagopiis). Comparative Biochemistry and Physiology 53:343-346. Bailey. E. 1993. Introduction of foxes to Alaskan Islands: history, effects on avifauna, and eradication. U.S. Fish and Wildlife Service Resource Publication 191. Anchorage. Alaska, USA. Baxter, R. J., J. T. Flinders, and D. L. Mitchell. 2008. Survival, movements, and reproduction of translocated Greater Sage-Grouse in Strawberry Valley, Utah. Journal of Wildlife Management 72:179-186. Blondel. j. 1985. Breeding strategies of the Blue Tit and Coal Tit iPariis) in mainland and island Mediterranean habitats: a comparison. Journal of Animal Ecology 54:531-556. Blondel, J. 2000. Evolution and ecology of birds on islands: trends and prospects. Vie et Milieu-Life and Environment 50:205-220. Blondel, J., R. Pradel, and J.-D. Lebreton. 1992. Low fecundity insular Blue Tits do not survive better as adults than high fecundity mainland ones. Journal of Animal Ecology 61:205-213. Braun, C. E.. D. H. Nish, and K. M. Giesen. 1978. Release and establishment of White-tailed Ptarmigan in Utah. Southwest Naturalist 23:661-668. Briskie. j. V. and M. Mackintosh. 2004. Hatching failure increases with .severity of population bottlenecks in birds. Proceedings of the National Academy of .Science of the United States of America 101 :558-561 . Brodsky, L. M. 1986. Correlates and consequences of the mating tactics of male Rock Ptarmigan (Lagopm niiilii.s). Dissertation. Queen's University, Kingston. Ontario. Canada. Brodsky, L. M. 1988. Mating taetics of male Rock Ptarmigan Lagopus minus', a conditional mating strategy. Animal Behaviour 36:335-342. Burnham, K. P. and D. R. Anderson. 1998. Model selection and inference: a practical information- theoretic approach. Springer- Verlag, New York, USA. Byrd, G. V., E. P. Bailey, and W. Stahl. 1997. Restoration of island populations of Black Oyster- catchers and Pigeon Guillemots by removing intro- dueed foxes. Colonial Waterbirds 20:253-260. Clarke, J. A. and R. E. Johnson. 1990. Biogeography of White-tailed Ptarmigan (Lagopus leucurus): impliea- tions from an introdueed population in the Sierra Nevada. Journal of Biogeography 17:649-656. Clout, M. N. and J. L. Craig. 1995. The conservation of critically endangered flightless birds in New Zealand. Ibis 137 (Supplement): S181-S190. Coates, P. S., S. J. Stiver, and D, J. Delehanty. 2006. Using Sharp-tailed Grouse movement patterns to guide release-site selection. Wildlife Society Bulletin 34:1376-1382. Cotter, R. C. 1991. Population attributes and reproductive biology of Rock Ptarmigan (Lagopus minus) in the central Canadian arctic. Thesis. University of Alberta. Edmonton, Canada. Cotter, R. C. 1999. The reproductive biology of Rock Ptarmigan (Lagopus minus) in the central Canadian arctic. Arctic 52:23-32. Cotter, R. C.. D. A. Boag. and C. C. Shank. 1992. Raptor predation on Rock Ptarmigan (Lagopus minus) in the eentral Canadian arctic. Journal of Raptor Research 26:146-151. Dimond, W. J. and D. P. Armstrong. 2007. Adaptive harvesting of souree populations for transloeation: a case study with New Zealand Robins. Conservation Biology 21:1 14-124. Dinsmore, S. j. G. C. White, and F. L. Knopf. 2002. Advanced techniques for modeling avian nest survival. Ecology 83:3476-3488. Ebbert, S. E. and G. V. Byrd. 2002. Eradications of invasive species to restore natural biological diversity on Alaska Maritime National Wildlife Refuge. Pages 102-109 ill Turning the tide: the eradication of invasive species (C. R. Veitch and M. N. Clout. Editors). lUCN Invasive Speeies Speeialist Group, Gland, Switzerland and Cambridge, United Kingdom. Emison, W. B. and C. M. White. 1988. Foods and weights of the Roek Ptarmigan on Amchilka. Aleutian Islands. Alaska. Great Basin Naturalist 48:533-540. Ewen, J. G. and D. P. Armstrong. 2007. Strategic monitoring of reintroductions in ecological restoration programmes. Ecoscience 14:401^09. Gardarsson, a. 1988. Cyclic population changes and some related events in Rock Ptarmigan in Iceland. Pages 300-329 in Adaptive strategies and population ecology of northern grouse (A. T. Bergerud and M. W. Gratson, Editors). University of Minnesota Press, Minneapolis, USA. Gibson. D. D. and G. V. Byrd. 2007. Birds of the Aleutian Islands, Alaska. Nuttall Ornithological Club and The Kalei- Cl at. • DEMOGRAPHY OF ISLAND PTARMIGAN 13 American Ornithologists' Union. Washington, D.C., USA. Gotelli, N. J. and A. M. Ei.lison. 2004. A primer of ecological statistics. Sinauer Associates, Sundeiiancl, Massachusetts, USA. Griffith. B., J. M. Scott, J. W. Carpenter, and C. Reed. 1989. Translocation as a species conservation tool: status and strategy. Science 245:477-480. Hannon, S. J. and K. Martin. 2006. Ecology of juvenile grouse during the transition to adulthood. Journal of Zoology 269:422^33. Hannon, S. J., R. C. Gruys, and J. O. Schieck. 2003. Differential seasonal mortality of the sexes in Willow Ptarmigan Lagopiis lagopiis in northern British Columbia, Canada. Wildlife Biology 9:317-326. Hoffman, R. W. and K. M. Giesen. 1983. Demography of an introduced population of White-tailed Ptarmigan. Canadian Journal of Zoology 61:1758-1764. Holder, K. and R. Montgomerie. 1993. Rock Ptarmigan (Lagopiis miitus). The birds of North America. Number 5 1 . Holder, K., R. Montgomerie, and V. L. Friesen. 2000. Glacial vicariance and historical biogeography of Rock Ptarmigan (Lagopiis minus) in the Bering region. Molecular Ecology 9:1265-1278. Holder, K., R. Montgomerie, and 'V. L. Friesen. 2004. Genetic diversity and management of Nearctic Rock Ptarmigan (Lagopiis miitiis). Canadian Journal ot Zoology 82:564-575. Jacobsen Jr., E. E., C. M. White, and W. B. Emison. 1983. Molting adaptations of Rock Ptarmigan on Amchitka Island, Alaska. Condor 85:420-426. Jamieson, I. G. and G. C, Wilson, 2003. Immediate and long-term effects of translocations on breeding success in Takahe Porpliyrio lioclisterteri. Bird Conservation International 13:299-306. Kaler, R. S. a. 2007. Demographic measures, habitat use, and movement of a recently reintroduced island population of Evermann’s Rock Ptarmigan. Thesis. Kansas State University, Manhattan, USA, Klein, D. R. 1968. The introduction, increase, and crash of reindeer on St. Matthew Island. Journal of Wildlife Management 32:350-367. Klett, a. T. and D. H. Johnson, 1982. 'Variability in nest survival rates and implications to nesting studies. Auk 99:77-87. Komdeur, j. 1994. Conserving the Seychelles Warbler Acrocephuliis sechelleiisis by translocation from Cous- in Island to the islands of Aride and Cousine. Biological Conservation 67:143-152. Kurzejeski, E. W. and B. G. Root. 1988. Survival of reintroduced Ruffed Grouse in north Missouri. Journal of Wildlife Management 52:248-252. Major, H. L., I. L. Jone.s, G. V. Byrd, and J. C. Williams. 2006. As.sessing the effects of introduced Norway rats (Ralliis norvegiciis) on survival and productivity of Least Anklets (Aeiliia piisilla). Auk 123:681-694. Maron, J. L., J. A. Estes, D. A. Croll, E. M. Danner, S. C, Elmendorf, and S. L, Buckelew. 2006, An introduced predator alters Aleutian island plant communities by thwarting nutrient subsidies. Ecolog- ical Monographs 76:3-24. Mercer, W. E. 1967. Ecology of an island population ol Newfoundland Willow Ptarmigan. Technical Bulletin Number 2. Newfoundland and Labrador Wildlile Service, St. John's, Canada. Moran, R. J. and W. L. Palmer. 1963. Ruffed Grouse introductions and population trends on Michigan islands. Journal of Wildlife Management 27:606-614. Murie, O. j. 1959. Fauna of the Aleutian Islands and Alaska Peninsula. North American Fauna 61:1-406. Musil, D. D., j. W. Connelly, and K. P. Ree.se. 1993. Movements, survival, and reproduction of Sage Grouse translocated into central Idaho. Journal ol Wildlife Management 57:85-91. Novoa, C., A. Besnard, j. F. Brenot, and L. N. Ellison. 2008. Effect of weather on the reproductive rate of Rock Ptarmigan Lagopiis miita in the eastern Pyrenees. Ibis 150:270-278. Powell, L. A. 2007. Approximating variance of demo- graphic parameters using the delta method: a reference for avian biologists. Condor 109:949-954. Reese, K. P. and J. W. Connelly. 1997. Translocation of Sage Grouse Centrocerciis urophasianus in North America. Wildlife Biology 3:235-241. Reynolds, M. H., N. E. Seavy, M. S. Vekasy, J. L. Klavitter, and L. P. Laniawe. 2008. Translocation and early post-release demography of endangered Laysan Teal. Animal Conservation 11:160-168. Robb, L. A., K. Martin, and S. J. Hannon. 1992. Spring body condition, fecundity and survival in female Willow Ptarmigan. Journal of Animal Ecology 61:215-233. Robertson, H. A., 1. Karika, and E. K. Saul. 2006. Translocation of Rarotonga Monarchs Pomarea dinii- diata within the southern Cook Islands. Bird Conser- vation International 16:197-215. Rohwer, E. C. 1988. Inter- and intraspecific relationships between egg size and clutch size in waterfowl. Auk 105:161-176. Sandercock, B. K. 1993. Free-living Willow Ptarmigan are determinate egg-layers. Condor 95:554-558. Sandercock, B. K. and H. C. Pedersen. 1994. The effects of renesting ability and nesting attempts on egg-size variation in Willow Ptarmigan. Canadian Journal of Zoology 72:2252-2255. Sandercock, B. K., K. Martin, and S. J. Hannon. 2005. Life history strategies in extreme environments: comparative demography of arctic and alpine ptarmi- gan. Ecology 86:2176-2186. Sarrazin, F. and S. Legendre. 2000. Demographic- approach to releasing adults versus young in reintro- ductions. Conservation Biology 14:488-500. SCHERINI, G. C., G. Tosi, and L. A. Walters. 2003. Social behaviour, reproductive biology and breeding success of alpine Rock Ptarmigan Lagopiis miitiis lielvericiis in northern Italy. Ardea 91:1 1-23. Smith. A. and T. Willebrand. 1999. Mortality causes and survival rates of hunted and unhunted Willow Grouse. Journal of Wildlife Management 63:722-730. Snyder, J. W., E. C. Pelren, and J. A. Crawford. 1999. 14 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122. No. I, March 2010 Translocation histories of prairie grouse in the United States. Wildlife Society Bulletin 27:428-432. Springer, P. F., G. V. Byrd, and D. W. Woolington. 1977. Reestablishing Aleutian Canada Geese. Pages 331-338 in Endangered birds: management techniques for preserving threatened species (S. A. Temple, Editor). University of Wisconsin Press, Madison, USA. Stamps, J. A. and M. Buechner. 1985. The territorial defense hypothesis and the ecology of insular vertebrates. Quarterly Review of Biology 60:155-181. Steen, J. B. and S. Unander. 1985. Breeding biology of the Svalbard Rock Ptarmigan (Lagopits mutus hyper- horeiis). Ornis Scandinavica 23:366-370. Watson. A. 1965. A population study of ptarmigan [Lagopus mutus) in Scotland. Journal of Animal Ecology 34:135-172. Weeden, R. B. 1965. Breeding density, reproductive success, and mortality of Rock Ptarmigan at Eagle Creek, central Alaska, from 1960 to 1964. Transac- tions of the North American Wildlife and Natural Resources Conference 30:336-348. Weeden, R. B. and A. Watson. 1967. Determining the age of Rock Ptarmigan in Alaska and Scotland. Journal of Wildlife Management 31:825-826. Westerskov, K. 1950. Methods for determining the age of game bird eggs. Journal of Wildlife Management 14:56-67. WiEBE. K. L. and K. Martin. 2000. The use of incubation behavior to adjust avian reproductive costs after egg laying. Behavioral Ecology and Sociobiology 48:463- 470. Wiggins, D. A., A. P. Mgller, M. E. L. Sgrenson, and L. A. Brand. 1998. Island biogeography and the reproductive ecology of Great Tits Pants major. Oecologia 115:478^82. Williams, J. C., G. V. Byrd, and N. B. Konyukhov. 2003. Whiskered Auklets Aethia pygmaea, foxes, humans and how to right a wrong. Marine Ornithology 31:175-180. Wilson, S., K. Martin, and S. J. Hannon. 2007. Nest survival patterns in Willow Ptarmigan: influence of time, nesting stage, and female characteristics. Condor 109:377-388. Zimmerman, C. E., N. Hillgruber, S. E. Burril, M. A. St. Peters, and J. D. Wetzel. 2005. Offshore marine observation of Willow Ptarmigan, including water landings, Kuskokwim Bay, Alaska. Wilson Bulletin 117:12-14. The Wilson Journal of Ornilhology 1 22( 1 ): I 5-22. 20 1 0 APPARENT SURVIVAL OF BREEDING WESTERN SANDPIPERS ON THE YUKON-KUSKOKWIM RIVER DELTA, ALASKA MATTHEW JOHNSON,'-" DANIEL R. RUTHRAUFF,^ BRIAN J. McCAFFERY,' SUSAN M. HAIG,' AND JEFFREY R. WALTERS^ ABSTRACT. — We used 8 years of live recapture data ( 1998-2005) to estimate apparent annual survival lor male (/) = 237) and female (/; = 296) Western Sandpipers (Calidris maitri) breeding on a 36-ha plot on the Yukon-Kuskokwim River Delta, western Alaska. Apparent annual survival (O) is the product of true survival and site fidelity, and estimates ol d) were corrected for the probability of encounter. Overall return rates (individual returned to the study site in a subsequent season) were lower for females (40%) than males (65%), as was th (± SE, females = 0.65 ± 0.05, males = 0.78 ± 0.03), and encounter rate (females = 0.51 ± 0.07, males = 0.74 ± 0.04). Results differed from previous estimates of O for this species as our estimates of O were higher for both males and females compared to estimates from another breeding site and two nonbreeding locations. Disparity among O estimates from breeding and nonbreeding areas highlights the need to delineate site-specific factors throughout the annual cycle that inlluence population dynamics of the Western Sandpiper. Received 30 May 2009. Accepted 9 October 2009. Twenty-one percent of the world’s 135 shore- bird species are listed as species of conservation concern by Birdlife International (Piersma et al. 1997, Johnson and Oring 2002, Sandercock 2003). Several long-term studies have revealed popula- tion declines among migratory shorebirds (Char- adriiformes) in Europe (Tucker and Heath 1994, Hagemeijer and Blair 1997), North America (Howe et al. 1989, Momson et al. 2001, Haig et al. 2005), and Asia (Rose and Scott 1997). The Western Sandpiper (Calidris mauri) is considered a species of high conservation concern despite being one of the most abundant shorebirds in the Western Hemisphere (Bishop et al. 2000, Morri- son et al. 2000, Brown et al. 2001). Declining numbers during spring (Butler and Lemon 2001) and fall migration surveys (Neil 1992, Butler and Lemon 2001) at staging areas in British Columbia and Texas coupled with a limited breeding distribution and threats to nonbreeding habitat have led to this conservation assessment (Brown et al. 2001, Fernandez et al. 2006). However, reported declines in numbers at staging sites may be caused by shifts in migration patterns and not ' U.S. Geological Survey, Forest and Rangeland Ecosys- tem Science Center, 3200 Southwest .lefferson Way, Corvallis, OR 97331, USA. ^U.S. Geological Survey, Alaska Science Center, 4210 University Drive, Anchorage, AK 99508, USA, ^U.S. Fish and Wildlife Service, Yukon Delta National Wildlife Refuge, P. O. Box 346, Bethel, AK 99559, USA. ■* Department of Biological Sciences, Virginia Polytech- nic Institute and State University, Blacksburg, VA 24061, USA, ■‘’Corresponding author; e-mail: matthewJohnson@usgs.gov impacts of demographic rates on population growth (Ydenberg et al. 2004). The life history of migratory shorebirds is generally characterized by delayed maturity, low productivity, and relatively high adult survivor- ship (Evans and Pienkowski 1984, Piersma and Baker 2000). Robust estimation of demographic parameters is central to understanding population dynamics of avian species (Eberhardt 1985, Lande 1988, Clobert and Lebreton 1991). Estimation of survival rate is particularly important because the persistence of any animal population is the result of a balance between recruitment of new breeders into the population via reproduction and immi- gration, and losses due to the emigration and mortality of established breeders (Lebreton et al. 1993, Lieske et al. 2000). Estimation of these parameters and identification of ecological factors that affect avian populations requires long-term studies of marked individuals (Lebreton and North 1993). A better understanding of shorebird demogra- phy would facilitate conservation efforts; howev- er, population models have been developed for only a few shorebird species. Sensitivity analyses have found that adult survival often has the highest elasticity value and potentially the greatest impact on population growth rates (Hill and Carter 1991, Hitchcock and Gratto-Trevor 1997, Reed et al. 1998, Plissner and Haig 2000, Larson et al. 2002, Sandercock 2003). Effective conser- vation of migratory shorebirds requires reliable estimates of annual survival rates, but estimates are available for <10% of shorebird species worldwide (Sandercock 2003). 15 16 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. I. March 2010 Western Sandpipers breed along the coasts of western Alaska and eastern Siberia and are distributed over a large geographic area in nonbreeding areas ranging on the Pacific Coast from California to Peru, and on the Atlantic Coast from New Jersey to Suriname (Wilson 1994). Male Western Sandpipers spend the nonbreeding season at more northerly latitudes compared to females (Page et al. 1972, Harrington and Haase 1994, Buenrostro et al. 1999, Nebel et al. 2002), and males generally precede females north during spring migration (Holmes 1971, Senner et al. 1981, Butler et al. 1987) with the opposite pattern during fall migration (Butler et al. 1987, Yden- berg et al. 2005). Comparative studies within their nonbreeding distribution have revealed divergent life history strategies (O’Hara et al. 2005). Juveniles and females from the southern portion of the nonbreeding range are more likely to remain at nonbreeding sites during their first breeding sea.son compared to adults and males at northerly sites. This suggests Western Sandpiper life history varies with first-year females favoring increased survival and first-year males emphasiz- ing initial breeding opportunities (O’Hara et al. 2005). Given the species’ expansive geographic distribution and age-and gender-specific variation in life history strategies, it is likely Western Sandpipers likewise exhibit region-, age-, and gender-specific variation in survival. Our objective was to produce estimates of apparent annual survival at a representative breeding site in the middle of the breeding range of Western Sandpipers in western Alaska. We used 8 years of live recapture data to examine whether apparent annual survival rate varied between males and females at our study site, and whether apparent survival varied significantly during the course of this study. We compare our estimates with those from a more northerly breeding site and two nonbreeding sites. METHODS Study Area and Field Methods. — We studied breeding Western Sandpipers at the Yukon Delta National Wildlife Refuge’s Kanaryarmiut Field Station, Yukon-Kuskokwim River Delta, Alaska (61 22' N, 165 07' W). We u.sed a 36-ha study plot in dry upland tundra habitat along the Kuyungsik River (lowland moist low .scrub com- munity, Jorgenson and Ely 2001) to address our objectives. The study site supports an average of 80 breeding pairs annually (Johnson et al. 2005). Two to four observers searched this study plot daily from early May through late July from 1998 to 2005 for banded birds, nests, and broods. We banded each adult at the nest with a U.S. Geological Survey identification band and unique color combinations (3 UV-stable color bands/bird). We recaptured all birds when band loss or discoloration occumed (2% of all banded birds), indentified the individual by its U.S. Geological Survey identification band, and re-banded the bird with new color bands. Adults were classified as male or female by culmen length during banding (>93% of all birds; Page and Fearis 1971, Carter 1984). Birds were classified as males or females via behavior (courtship displays, copulation posi- tion) and comparison with their mate (females > males) when culmen length was inconclusive (24- 25 mm). Locations of nests and banded individuals were mapped using geographic information system layers in ArcGIS (ESRI 2007) and nests were monitored through hatch, predation, or abandon- ment. It is possible to differentiate first-year Western Sandpipers that hatched in the previous breeding season from birds that are >2 years old based on plumage characteri.stics (Sandercock et al. 1999). However, our sample of first-year birds was limited (24% of females and 17% of males banded were first-year birds), and previous attempts to estimate apparent survival in this population revealed unacceptable overdispersion when fitting age-specific models to the data set (r > 8; M. Johnson, unpubl. data). We chose not to estimate apparent survival by age class (i.e., first-year and >2 year old birds), and instead estimated apparent survival for all breeding adults. Statistical Analyses. — ^We used Cormack-Jolly- Seber mark-recapture models to estimate apparent annual survival (4)) coirected for probability of encounter (p). Apparent survival rate was the probability that a bird alive in breeding season / survived until the following breeding season (/ -t- 1 ) and returned to our study plot. Encounter probability was the probability that we detected a bird given that it was alive. Losses from the breeding population may result from permanent emigration or mortality, and are the complement ol apparent survival. We conducted mark-recap- ture analysis following methodologies outlined by Burnham and Anderson (1998) using Program MARK (Version 4.2, Gooch and White 2004). We incorporated gender (,vc.v) and time-depen- dence (/) as variables in our global model of O and p because both parameters are known to vary Johnson cl ul. • WESTHRN SANDPIPER SEIRVIVAI. 17 between males and females and aeross years in many shorebird speeies (Sandereock 2003). We also modeled lemporal variation as a linear (7 ) and quadratie (TI) trend across years. Several nonexckisive factors may lead to lower apparent survival rates during the first interval after initial capture in many bird populations (Sandereock and Jaramillo 2002). Thus, we modeled (1> for the interval following initial capture (' v,.,v /2«-a- AQAIC,. = 1 .6; (I)', 18 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122. No. I. March 2010 TABLE 1. Mark-recapture models for annual estimates ot apparent survival for adult Western Sandpipers breeding near Kanaryarmiut Field Station, Yukon-Kuskokwim River Delta. Alaska. 1998-2005. Model statistics'* Model structure^ K Deviance AQAIC,. W, Psex 4 122.6 0.0 0.32 Psex 5 121.9 1.3 0.16 ^ sex ^ sex' Psex 6 120.2 1.6 0.14 + TT- Psex 6 120.8 2.2 0.1 1 4)'.ve.v Psex 5 123.1 2.5 0.09 Psex 3 127.6 3.0 0.07 4>/ . Psex 4 127.2 4.6 0.03 4’7T- Psex 5 126.5 5.9 0.02 Pc 3 130.8 6.2 0.01 ■st'A + / ^ sex + /' Pse.\ , 15 110.9 10.7 0.002 Psex*t 26 105.3 28.1 < 0.001 sex*i ^ sex*h Psex"^! 37 97.2 43.5 < 0.001 We described model fit by the number of parameters {K). deviance, and the difference in quasi-Akaike's Information Criterion (AQAIC, ) from the best-fit model. QAIC, values were calculated using a variance infiation factor of 2.9. We present models with moderate support (Akaike weight w, ^ 0.01) in order of relative fit to the best-fit model (i.e., AQAIC, = 0). followed by a two age- cla.ss (time-since-marking) model with additive effects of sex and time, a standard Cormack-Jolly-Seber model that did not control for time since marking, and the starting global model. Model factors included; O' = apparent survival during the first-year post banding. O*' = apparent .survival during subsequent years. O = annual apparent .suiwival in models lacking age structure, c = constant, = gender, t = time or annual variation. T — linear time trend. 7T = quadratic lime trend. * = a factorial model, and -i- = an additive model. males = 0.72 ± 0.06, females = 0.57 ± 0.08; males = 0.81 ± 0.04, females = 0.69 ± 0.06). However, the best-fit model p^ex) had >2 times the support of both of these models. Model selection did not support annual variation (f) in apparent survival rates (AQAIC,. = 7.3); however, we examined the point estimates of annual apparent survival from our global model ^^^'^scx*h Psex*i) because time dependence is often not supported in sparse data .sets. Female apparent survival was lower in 2001 compared to other years in the interval after first capture (d)') and later intervals Fig. 1). Inter-annual dispersal distances varied among site-faithful females, males, and reuniting pairs {X~2 ~ 54.2, P < 0.001). Inter-annual dispersal distances were smallest among reuniting pairs (56.0 ± 12.3 m, ii = 47 pairs). Inter-annual dispersal distances were greater among females (220.9 ± 27.7 m, n = 66) compared to males (86.9 ± 8.5 m, n = 174) for site-faithful individuals that did not reunite with their prior mate. We surveyed 29 plots (8.3-ha/plot) annually within a 29 km^ area surrounding our study site (15-31 May, 2()()4-2()()5). We did not detect any LJJ w B ro 7c > w c 0) CO Q. Q. CO To 3 c c < 1.0 1 0.8 - 0.6 ■ 0.4 - 0.2 - 0.0 -■ Subsequent intervals T I "V I T" • " ■ • ' "» 1998-1999 1999-2000 2000-2001 2001-2002 2002-2003 2003-2004 Transition period (year) FIG. 1. Annual apparent survival rates (mean ± SE) of adult Western Sandpipers (296 females, 237 males) breeding near Kanaryarmiut Field Station, Yukon-Kuskok- wim River Delta, Alaska (1998-2005) in the interval after first capture and subsequent intervals (males = squares connected with solid lines, females = circles connected with broken lines). Annual estimates were calculated via model averaging, and estimates for interval 2004-2005 were not identifiable because of the time dependence in apparent annual survival and the probability of encounter. color-marked birds during 50 hrs of observation in the sun'ounding region. No banded birds that went undetected in our surveys were encountered by concurrent research projects being conducted in the surrounding area (2003-2005). DISCUSSION We used Cormack-Jolly-Seber mark-recapture models to estimate apparent annual survival of breeding Western Sandpipers on the Yukon- Kuskokwim River Delta, Alaska. These estimates complement those from other parts of the breeding and nonbreeding range by identifying potential temporal and spatial patterns in survival rates. Apparent survival of female Western Sandpipers was lower than males, and hierarchical model selection results indicated that apparent survival did not vary annually. Previous estimates ot apparent survival for breeding Western Sand- pipers also were higher for males than females (Sandercock et al. 2000), hut our estimates were higher for both males and females (Table 2). Two studies also have used mark-recapture techniques to estimate the apparent annual survival of Western Sandpipers in nonbreeding areas. Juve- nile and adult males in Baja California had relatively low apparent annual survival (Fernan- .Johnson cl a/. • WESTERN SANDPIPER SURVIVAL. 19 TABLE 2. Annual estimates of apparent survival for Western sandpipers from studies that used mark-recapture statistics for live (CuUdris nuniri) encounter data. and Sent i|)al mated (C. semipalmuiiis) Species Range-' Age" GendeL Apparent .survival'' a birds # yeans'’ Reference Western Sandpiper B A F/M 0.57 236 4 Sandercock et al. 2000 NB J M 0.49 139 4 Fernandez et al. 2003 A M 0.45 1 17 NB J/A F 0.62 1 ,990 3 Fernandez et al. 2004 M 0.54 1,383 B A F 0.65 296 8 This study M 0.78 237 Semipalmated B A F 0.56 249 8 Sandercock and Gratto-Trevor Sandpiper M 0.61 237 1997 B A F 0.59 122 4 Sandercock et al. 2000 M 0.73 108 NB A F/M 0.65 215 4 Rice et al. 2007 J F/M 0.55 319 4 “ Range: B = breeding area, NB = nonbreeding area. Age; J = juvenile.s SI year of age. A = adulls >1 year of age. '■ Gender: F = females, M = males. ^ .Apparent survival: estimate from best-fit model, or average of annual estimates (i.e.. Sandercock and Gratto-Trevor 1997). ^ # years = duration of study in years. dez et al. 2003, Table 2), and females had higher rates of apparent annual survival than males in Chitre, Panama (Fernandez et al. 2004, Table 2). Our results are comparable to those from breeding areas for a sympatric Calidris species with similar life history characteristics. Semipal- mated Sandpipers (C. pusilla) are small-bodied shorebirds that breed in the arctic and sub-arctic that also undertake long distance migrations and distribute themselves over a relatively large nonbreeding range (Gratto-Trevor 1992). Male Semipalmated Sandpipers near Nome, Alaska had higher apparent survival rates than females (Ta- ble 2) and both rates were similar to our estimates for Western Sandpipers. Male Semipalmated Sandpipers breeding at La Perouse Bay, Manitoba, Canada had higher apparent survival rates than females (Table 2). These estimates were lower than ours for Western Sandpipers; however, there was considerable annual variation in that study (1980-1987, males 0.53-0.74, females 0.43-0.71; Sandercock and Gratto-Trevor 1997). Differences between estimates of apparent survival for Western Sandpipers in this study and those from Nome may reflect temporal and/or spatial variation in survival rates. Sandercock et al. (2000) conducted their study from 1993 to 1996 at a site near Nome, Alaska. It is ditlicult to assess whether differences are due .solely to temporal variation because the two studies were not contemporary. Survival rates vary over time in other species (e.g.. Lesser Snow Goose \Anser caerulescens], Francis et al. 1992; Mallard {Anas platyrhynchos], Franklin et al. 2002; Barn Owl [Tyto alba], Altwegg et al. 2003), and this phenomenon may occur with Western Sandpipers. However, model selection results did not indicate apparent annual survival varied during the course of our study (Table 1). Other researchers have demonstrated that demographic parameters including survival rates often vary spatially (e.g., Nichols et al. 1990, Kanyamibwa et al. 1993, Franklin et al. 2002, Pearce et al. 2005), and higher estimates of apparent survival at our study site may reflect site-specific conditions that promote higher sur- vival compared to Nome, which is ~365 km and 4° latitude further north than our study site. Western Sandpipers breeding at Nome experience a shorter breeding season with longer periods of inclement weather compared to birds at our study site (Johnson and Walters 2008). A more severe climate coupled with longer annual migration distances may predispose birds at Nome to decreased survival. Annual variation in apparent survival did not occur in our study population, even though it has been reported in other studies of Western Sandpipers that were of considerably shorter duration (4 years, Sandercock et al. 2000; 3 years, Fernandez et al. 2003). Examination of point estimates from our global model revealed that female apparent survival was lower in 2001 compared to other years (Fig. 1 ). We are not able 20 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. I, March 2010 to account for this variation, but lower female survival in 2001 may partially explain why overall estimates of apparent survival were lower for temales compared to males. Whether lower apparent annual survival among temale Western Sandpipers represents variation in true survival or sexual variation in breeding site fidelity is unknown because we are not able to differentiate the relative influence of mortality versus permanent emigration in our estimation of apparent annual survival. Males generally exhibit greater breeding site fidelity than females among monogamous sandpipers (Oring and Lank 1982, Piersma et al. 1996). We cannot conclude that temale annual survivorship is lower than for males in our study population because females may exhibit lower breeding site fidelity than males. However, average inter-annual dispersal distances were relatively small in relation to size of our study site, and similar to those reported near Nome (median inter-annual dispersal dis- tance near Nome: reuniting pairs = 38 m, males = 48 m, females = 157 m). Further, no banded birds that went undetected in our surveys were encountered in the surrounding area. We conclude that local permanent emigration was not common at our study site. Greater inter-annual dispersal distances among site-faithful females may indicate that females also emigrate from our study site at a higher rate than males. However, local movement patterns may or may not be indicative of larger scale dispersal patterns. Encounter probabilities were lower for females during our study, which may indicate temporary emigration (i.e., attempting reproduc- tion at another location for I or more intervening years) is more common among females, or there are behavioral differences that result in lower detection rates for females. Males are more conspicuous early in the breeding sea.son becau.se they engage in conspicuous epigamic displays daily, whereas females are more cryptic spending greater amounts of time in concealed locations, such as wet meadows (Lanctot et al. 2()()()). Further, female encounter rates may be lower if only a portion of tho.se individuals that permanently emigrate still pa.ss through our study site before settling at other locales, or if .some individuals pa.ss through quickly and escape detection. Western Sandpipers are a species of high conservation concern yet the underlying causes warranting this classification are unclear. Disparity between and among apparent survival estimates from breeding and nonbreeding areas highlights the need to delineate site-specific factors throughout the annual cycle that influence population dynam- ics of Western Sandpipers. The closely related Semipalmated Sandpiper is likely to have compa- rable population demographics to the Western Sandpiper (Fernandez et al. 2003). Our estimates of apparent annual survival for Western Sandpipers are similar to true annual survival rates required to sustain a breeding population of Semipalmated Sandpipers (—0.80, Hitchcock and Gratto-Trevor 1997). However, observed population declines may be misleading. Reported declines in numbers at staging sites may be caused by a shift in migration patterns (Ydenberg et al. 2004) and not by a decreasing population, further underscoring the necessity of additional research to elucidate these questions. Effective modeling and predicting population trends requires accurate estimation of vital rates encompassing the range of natural variability. Our estimates of adult survival of Western Sandpipers augment and complement previous estimates by providing current estimates from a centrally located site within the species’ breeding range, and serve to better inform future con.servation efforts. ACKNOWLEDGMENTS We thank the .staff of the Yukon Delta National Wildlife Reluge for supporting this research. Financial support was received from the U.S. Fish and Wildlife Service (Yukon Delta National Wildlife Refuge), the Harold F. Bailey Fund at Virginia Tech, and the U.S. Geological Service (Forest and Rangeland Ecosystem Science Center). We thank T. L. Booms, J. R. Conklin, Z. M. Fairbanks, C. E. Fitzpatrick. L. F. Hamblin. P. N. Laver, B. L. Johnson, Silke Nebel, A. C. Niehaus. L. W. Oring, D. J. Rizzolo, and M. K. Spies for assistance in the tield. G. J. Fernandez, D. B. Lank. C. A. Nicolai, C. M. Taylor, B. K. Sandercock, and two anonymous reviewers provided assistance and comments that greatly improved this manuscript. Any use of trade, or tirm names is tor descriptive purposes only and does not imply endorsement by the U.S. Government. LITERATURE CITED Altwegg, R., a. Rouun, M. Ke.stkniiolz, and L. Jenni. 2003. Variation and covariation in survival, dispersal, and population size in Barn Owls Tylo alha. Journal of Animal Ecology 72:391-399. Bishop, M. A.. P. M. Meyers, and P. E. Mc Neley. 2()()(). A method to estimate migrant shorebird numbers on the Copper River Delta, Alaska. Journal of Field Ornithology 71:627-637. Brown, S., C. Hickey, B. Harrington, and R. Giee. 2001. flic U.S. Shorebird Conservation Plan. Second Johnson d al. • WESTERN SANDPIPER SURVIVAE 21 Edition. Manomcl Center for Conservation Seienees, Manomet, Massachusetts. USA. Buenorostro, M. a., N. Warnock, and H. dela Clieva. 1999. Wintering Western Sandpipers Caliciris nuntri at Estero de Pnnta Banda, Baja Calilornia, Mexico. Wader Study Group Bulletin 88:59-63. Burnham, K, P. and D. R. Anderson. 1998. Model selection and inl'erence: a practical information- theoretic approach. Springer-Verlag, New York, USA. Butler. R. W. and M. J. E. Lemon. 2001. Trends in abundance of Western and Least sandpipers migrating through southern British Columbia. Bird Trends 8:36- 38. Butler, R. W., G. W. Kaiser, and G. E. J. Smith. 1987. Migration ehronology, length of stay, sex ratio, and weight of Western Sandpipers (Calidris niauri) on the south coast of British Columbia. Journal of Field Ornithology 58:103-1 1 1. Carter, R. V. 1984. A morphometric comparison of Western and Semipalmated sandpipers. Wilson Bulle- tin 96:277-286. Clobert, j. and J.-D. Lebreton. 1991. Estimation of demographic parameters in bird populations. Pages 75-104 in Bird population studies (C. M. Perrins, J.-D. Lebreton, and G. J. M. Hirons, Editors). Oxford University Press, Oxford, United Kingdom. COOCH, E. AND G. White, 2004. Program MARK: a gentle introduction, http://www.phidot.org/software/mark/ docs/book/ Eberhardt, L. L. 1985. Assessing the dynamics of wild populations. Journal of Wildlife Management 49:997- 1012. ESRI Inc. 2007. ArcGIS. Version 9.2. Earth Systems Research Institute Inc., Redlands, California, USA. Evans, P. R. and M. W. Pienkowski, 1984. Population dynamics of shorebirds. Pages 83-123 in Behaviour of marine animals (J. Burger and B. L. Olla, Editors). Volume 5. Plenum Press, New York, USA. FernAndez, G.. P. D. O’Hara, and D. B. Lank. 2004. Tropical and subtropical Western Sandpipers {Calidris mauri) differ in life history strategies. Ornitologia Neotropical 15:385-394. Fernandez, G., H. de la Cueva, N. Warnock, and D. B. Lank. 2003. Apparent survival rates of Western Sandpipers {Calidris mauri) wintering in northwest Baja California, Mexico. Auk 120:55-61. Fernandez, G., N. Warnock, D. L. Lank, and J. B. Buchanan. 2006. Version 1.1. Conservation Plan for the Western Sandpiper. Manomet Center for Conser- vation Science, Manomet, Massachusetts, USA. Francis, C. M., M. H. Richards, F. Cooke, and R. F. Rockwell. 1992. Long-term changes in survival rates of Lesser Snow Geese. Ecology 73:1346-1362. Franklin, A. B., D. R. Anderson, and K. P. Burnham. 2002. Estimation of long-term trends and variation in avian survival probabilities using random effects models. Journal of Applied Statistics 29:267-287. Gratto-Trevor, C. L. 1992. Semipalmated Sandpiper {Calidris pnsilla). The birds of North America. Number 6. Haig, S. M., C. L. Ferland, F. J. Cuthbert, .1. Dingledine, j. P. Goossen, A. Hechi, and N. McPhillirs. 2005. A complete species census and evidence for regional declines in Piping Plovers. Journal of Wildlife Management 69:160-173. Hagemeijer. E. j. M. and M. J. Blair. 1997. The EBCC Atlas of European breeding birds: their distribution and abundance. T. & A. D. Poyser, London, United Kingdom. Harrington, B. and B. Haase. 1994. Latitudinal differ- ences in sex ratios among nonbreeding Western Sandpipers in Puerto Rico and Ecuador. Southwestern Naturalist 39:188-189. Hill, D. and N. Carter. 1991. An empirical simulation of an Avocet Recurvirostra avosetta population. Ornis Scandinavica 22:65-72. Hitchcock, C. and C. L. Gratto-Trevor. 1997. Diagno.s- ing a shorebird local population decline with a stage- structured population model. Ecology 78:522-534. Holmes, R. T. 1971. Density, habitat and the mating system of the Western Sandpiper {Calidris mauri). Oecologia 7:191-208. Howe, M. A., P. H. Geissler, and B. A. Harrington. 1989. Population trends of North Ameriean shorebirds based on the International Shorebird Survey. Biolog- ical Conservation 49:185-199. Johnson, M. and L. W. Oring. 2002. Are nest exclosures an effective tool in plover conservation? Waterbirds 25:184-190. Johnson, M. and J. R. Walters. 2008. Effects of mate and site fidelity on nest survival of Western Sandpipers. Auk 125:76-86. Johnson, M., S. E. Jamieson, and B. J. McCaffery. 2005. Arctic breeding conditions locality report: Kanaryar- miut Field Station, Yukon Delta National Wildlife Refuge. Arctic Birds 7:26-28. Jorgenson, T. and C. Ely. 2001 . Topography and Hooding of coastal ecosystems on the Yukon-Kuskokwim Delta, Alaska: implications for sea-level rise. Journal of Coastal Research 17:124-136. Kanyamibwa, S., F. Bairlein, and a. Schierer. 1993. Comparison of survival rates between populations of the White Stork Ciconia ciconia in central Europe. Ornis Scandinavica 24:297-302. Kruskal, W. H. and W. a. Wallis. 1952. Use of ranks in one-criterion analysis of variance. Journal of the American Statistical Association 47:583-621. Lanctot, R. B., B. K. Sandercock, and B. Kempenaers. 2000. Do male breeding displays function to attract mates or defend territories? The explanatory role of mate and site fidelity. Waterbirds 23:15.5-164. Lande, R. 1988. Demographic models of the Northern Spotted Owl (Stri.s occidenlalis canrina). Oecologia 75:601-607. Larson, M. A., M. R. Ryan, and R. K. Murphy. 2002. Population viability of Piping Plovers: effects of predator exclusion. Journal of Wildlife Management 66:361-371. Lieske, D. j., I. G. Warkentin, P. C. James, L. W. Oliphant, and R. H. M. Espie. 2000. Effects of population density on survival in Merlins. Auk 1 17:184-193. THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 1, March 2010 22 Lebreton, J.-D. and P. M. North. 1993. Marked individuals in the study of bird populations. Birkhauser Verlag, Basel, Switzerland. Lebreton, J.-D., R, Pradel, and J. Clobert. 1993. The statistical analysis of animal populations. Trends in Ecology and Evolution 8:91-95. Lebreton, J.-D., K. P. Burnham, J. Clobert, and D. R. Anderson. 1992. Modeling survival and testing biological hypotheses using marked animals: a unified approach with case studies. Ecological Monographs 62:67-118. Morrison, R. 1. G, R. E. Gill Jr., B. A. Harrington, S. Skagen, G. W. Page, C. L. Gratto-Trevor, and S. M. Haig. 2000. Population estimates of Nearctic shorebirds. Waterbirds 23:337-352. Morrison, R. I. G., Y. Aubry, R. W. Butler, G. W. Beyersbergen, G. M. Donaldson, C. L. Gratto- Trevor, P. W. Hicklin, V. H. Johnston, and R. K. Ross. 2001. Declines in North American shorebird populations. Wader Study Group Bulletin 94: 34-38. Nebel, S., D. B. Lank, P. D. O’Hara, G. Fernandez, B. Haase, F. Delgado, F. A. Estela, L. J. Evans Ogden, B. Harrington, B. E. Kus, J. E. Lyons, F. Mercier, B. Ortego, J. Y. Takekawa, N. Warnock, AND S. E. Warnock. 2002. Western Sandpipers [Catidris maiiri) during the nonbreeding season: spatial segregations on a hemispheric scale. Auk 1 19:922-928. Neil, R. L. 1992. Recent trends in shorebird migration for north-central Texas. Southwestern Naturalist 37:87-88. Nichols, J. D., M. Williams, and T. Caithness. 1990. Survival and band recovery rates of Mallards in New Zealand. Journal of Wildlife Management 54:629-636. O'Hara, P. D., G. FernAndez, F. Becerril, H. de al CuEVA, AND D. B. Lank. 2005. Life history varies with migratory distance in Western Sandpipers Calidris nuiuri. Journal of Avian Biology 36:191-202. Oring, L. W. and D. B. Lank. 1982. Sexual selection, arrival times, philopatry, and site fidelity in the polyandrous Spotted Sandpiper. Behavioral Ecology and Sociobiology 10:185-191. P.vge, G. and B. Fearis. 1971. Sexing Western Sandpipers by bill length. Bird-Banding 42:297-298. Page, G., B. Fearis, and R. M. Jurek. 1972. Age and .sex composition of Western Sandpipers on Bolinas Lagoon. California Birds 3:79-86. Pearce, J. M., J. A. Reed, and P. L. Flint. 2005. Geographic variation in survival and migratory tendency among North American Common Mergan- sers. Journal of Field Ornithology 76:109-1 18. PlERSMA, T. AND A. J. BAKER. 2000. Life history character- istics and the con.servation of migratory shorebirds. Pages 105-124 in Behaviour and conservation (L. M. Gosling and W. J. Sutherland. Editors). Cambridge University Press. Cambridge, United Kingdom. PlERSMA. T.. J. VAN GiLS. AND P. WiERSMA. 1996. Family Scolopacidae (Sandpipers. Snipes and Phalaropes). Pages 444-487 in Handbook of the birds of the world. Volume 3, Hoatzin to Auks (J. del Hoyo, A. Elliott, and J. Sargatal, Editors). Lynx Edicions, Barcelona, Spain. PtERSMA, T., P. WiERSMA, AND J. VAN GiLS. 1997. The many unknowns about plovers and sandpipers of the world: introduction to a wealth of research opportu- nities highly relevant for .shorebird conservation. Wader Study Group Bulletin 82:23-33. Plissner, J. H. AND S. M. Haig. 2000. Viability of Piping Plover Charadrius melodus metapopulations. Biolog- ical Conservation 92:163-173. Reed, J, M., C. S. ELPHtCK, and L. W. Oring. 1998. Life- history and viability analysis of the endangered Hawaiian Stilt. Biological Conservation 84:35-45. Rice, S. M., J. A. Collazo, M. W. Alldredge, B. A. Harrington, and A. R. Lewis. 2007. Local annual survival and seasonal residency rates of Semipalmated Sandpipers {Calidris piisilla) in Puerto Rico. Auk 124:1397-1406. Rose, P. M. and D. A. Scott. 1997. Waterl’owl population estimates. Second Edition. Wetlands International Publication 44. Wageningen, The Netherlands. Sandercock, B. K. 2003. Estimation of survival rates for wader populations: a review of mark-recapture methods. Wader Study Group Bulletin 100:163-174. Sandercock, B. K. and C. L. Gratto-Trevor. 1997. Local survival in Semipalmated Sandpipers Calidris piisilla breeding at La Perouse Bay, Canada. Ibis 139:305-312. Sandercock, B. K. and A. Jaramillo. 2002. Annual survival rates of wintering sparrows: assessing demo- graphic consequences of migration. Auk 119:149- 165. Sandercock, B. K., D. Lank, and F. Cooke. 1999. Seasonal declines in the fecundity of two arctic- breeding sandpipers: different tactics in two species with an invariant clutch size. Journal of Avian Biology 30:460-468. Sandercock, B. K., D. B. Lank, R. B. Lanctot, B. Kempenaers, and F. Cooke. 2000. Ecological correlates of mate fidelity in two arctic-breeding sandpipers. Canadian Journal of Zoology 78:1948- 1958. Senner, S. E., D. W. Norton, and G. C. West. 1981. The spring migration of Western Sandpipers and Dunlins in southcentral Alaska: numbers, timing, and sex ratios. Journal of Field Ornithology 52:271-284. Tucker, G. M. and M. Heath. 1994. Birds in Europe: their conservation status. Birdlife Conservation Series Number 3. Birdlife International. Cambridge, United Kingdom. Wilson, W. H. 1994. Western Sandpiper {Calidris niaiiri). The birds of North America. Number 90. Ydenberg, R. C., a. C. Niehaus, and D. B. Lank. 2005. Interannual differences in the relative timing of southward migration of male and female Western Sandpipers {Calidris nuutri). Naturwissen.schaften 92: 332-335. Ydenberg, R. C., R. W. Butler, D. B. Lank, B. D. Smith, AND J. Ireland. 2004. Western Sandpipers have altered migration tactics as Peregrine Falcon popula- tions have recovered. Proceedings of the Royal Society of London. Series B 271:1263-1269. The Wilson Joitnial of Oniilhology 1 22( 1 ):23 -28, 2010 NATAL PHILOPATRY AND APPARENT SURVIVAL OF JUVENILE SEMIPALMATED PLOVERS ERICA NOLT^ SIMONE WILLIAMS,' AND BRETT K. SANDERCOCK^ ABSTRACT. — Natal philopatry is rare in long-distance migrant shorebirds and requires long-term population studies to detect. We report on the rate of natal philopatry from a 1 8-year study of Semipalmated Plovers (Charadrius seniipalniatus) marked as hatchlings to an arctic breeding site near Churchill, Manitoba. About 2% (27/1271) of banded hatchlings returned to the Churchill area to breed. There was no male/female bias in rates of philopatry: 17 male and 10 female hatchlings recruited into the local breeding population. The annual rate of recruitment of hatchlings varied between 0 and 10.7%. Age of first encounter on breeding areas ranged from 1 to 8 years (median age 4) suggesting either unusually delayed age at first breeding, or low detection rates for philopatric hatchlings. The maximum age of a recruited (known- age) hatchling was 9 years. Natal dispersal distances did not differ between males and females, and averaged 5 km between hatching and breeding locations. We used a time-since-marking mark-recapture model to calculate apparent survival of hatchlings. Apparent survival in the interval after first capture was (j)| = 0.0475 (95% Cl: 0.030-0.075), whereas apparent survival (cj)2+) of birds during subsequent intervals was 0.866 (95% Cl: 0.764-0.927). Low rates of natal philopatry suggest little advantage to site familiarity for juveniles, and agree with theoretical predictions for migratory species with widespread habitat availability. Received 15 June 2009. Accepted 7 October 2009. Natal philopatry is rare in long-distance mi- grant birds and requires either long-term popula- tion studies or large sample sizes to detect (Thompson and Hale 1989, Renner and McCaff- ery 2008). Using known-age birds in survival analyses avoids the complications of assuming age of first breeding in population viability analyses and, with sufficiently large sample sizes, allows precise estimates of age-specific survival (Sidhu et al. 2007), probability of breeding (Sedinger et al. 2001, 2008), and life-expectancy (Lavers et al. 2007). Analyzing natal philopatry can also help evaluate hypotheses about the adaptive significance of gender bias in natal dispersal rates and distances (Greenwood and Harvey 1982, Weatherhead and Forbes 1994, Sutherland et al. 2000). Studies of arctic-breeding birds are disadvan- taged in documenting natal philopatry and dispersal becau.se of inaccessibility of breeding areas and low encounter probabilities (Renner and McCaffery 2008). Possibly as a con.sequence, studies of temperate breeding shorebirds typically report higher rates of encounter of breeding juveniles, especially if the species is of con.serva- tion concern (e.g., Charadrius alexandrimts and C. melodus', Larson et al. 2000, Colwell et al. 2007b, Stenzel et al. 2007). ‘Biology Department, Trent University, Peterborough. ON, K9J 7B8. Canada. ^Division of Biology. Kansas Slate University. Manhat- tan, KS 66506, USA. ^Corresponding author; e-mail: enol@lrentu.ca The Semipalmated Plover (C. semipalmatus) has been the subject of an 18-year population study near Churchill, Manitoba. This small shorebird breeds in open gravel, tundra, beaches, and riverbeds throughout the sub-arctic regions of North America (Nol and Blanken 1999). Semi- palmated Plovers migrate long distances from sub-arctic breeding areas to a large, dispersed wintering range that spans coastal areas of the southern United States, Mexico, Central America, and northern and central South America. Unlike many Charadrius plovers, there is currently no evidence to suggest this species is declining (Morrison et al. 2006, Bart et al. 2007). An earlier study of this species suggested low rates of natal philopatry and no gender bias in these rates (Flynn et al. 1999). Our objectives are to: (1) update knowledge of these factors, (2) evaluate natal di,spersal distances, and (3) estimate apparent juvenile and adult survival based on a known-age sample. We also estimate the earliest age of first breeding and report on a longevity record based on a known-age hatchling. METHODS Study Sites and Species. — Field work was conducted each year from the beginning of June to mid-August in 18 years including 1988 and 1992 to 2008. We regularly searched 11 major breeding sites in the area .south and east of Churchill, Manitoba. Canada (58° 45' N, 95° 04' W), spanning ~ 20 km of coastal beaches and two inland sites in an area of 34.2 km' (Fig. 1). Each field .season, researchers walked 23 24 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 1, March 2010 FIG. I. Churchill, Manitoba area showing the 1 1 major study sites (★). potential nesting areas in search of Semipalmated Plovers. The nests of Semipalmated Plovers in 1988 and from 1992 to 2000 were assigned identification (ID) numbers sequentially as they were found within the 1 1 areas, and locations of nests were drawn on field maps and aerial photos. Nests of Semipalmated Plovers were assigned ID numbers from 2001 to 2008 and the locations were recorded using a Global Positioning System (GPS). In early June, observers searched for previously banded Semipalmated Plovers and recorded the band combinations of color-marked birds using 20-40 X spotting .scopes. Any nesting Semipalmated Plovers that were unhanded, or needed clarification of identification (e.g., faded color bands) were captured using circular walk-in traps with a keyhole entrance, made of hardware cloth, that were placed over the nest. Once adults were captured, they were banded with an aluminum band and given a unique color combination of bands or faded color and alumi- num bands were replaced. Gender of adults was assigned using the number of brown feathers in the crown and supcrcilium with females having more brown feathers than males (Teather and Nol 1997). Gender assignments were confirmed by behavioral postures with males using display flights or mounting females during copulations. Chicks were banded in the nest with unique aluminum bands and, depending on the year of study, with either brood-specific or individual color band combinations. Statistical Analysis. — ^We assigned the location of the natal site as the UTM of the center of one of the 1 1 nesting areas for older nest and banding records without reliable GPS locations (in years prior to 2000) (Fig. 1 ), and the location of the first breeding site as the UTM of the center of the site at which we first encountered them breeding. We used a binomial test to examine whether there was a bias in the gender of returning hatchlings, and t- tests to compare natal dispersal distances of males and females. We used a two-stage ‘time-since-marking’ model for live encounter data using Program MARK to calculate annual probabilities of apparent survival (cj)) and encounter (p). The time-since-marking model separates apparent survival in the first interval after capture ((t)|) Irom apparent survival in subsequent intervals No! et ill. • SllRVIVAl. OI’ JUVCNILH SEM I I’ALM ATEI) IM.OVERS 25 ((j)2+) (Larson et al. 2(){)()). We modeled apparent survival as a fLinetion of year of marking and gender beeanse these factors are important in other plovers (Larson et al. 2()()()), and treated the encounter rate as a nuisance parameter. We tested for possible overdispersion before testing reduced models by using the median c procedure to calculate a variance inflation factor for the global starting model (Cooch and White 2004). RESULTS Twenty-seven of 1,271 (2.1%) banded hatch- lings from 1988 to 2008 returned to breed in the Churchill area. Rates of return were low in all years, but reached 10% in 2000 and 2001 (Table 1). Ten local females and 17 males recruited into the breeding population, a ratio that was not significantly different than LI (Binomial test, P = 0.09). The age (±SE) at which birds were first encountered breeding ranged widely from 1 to 8 years (median 4 years) and did not differ between males and females (males: 3.9 ± 0.4 years, females: 4.2 ± 0.6 years; t = 0.38, P = 0.65). The maximum age of a recruited (known- age) hatchling was 9 years. The distance (±SE) between natal sites and nesting sites where we first encountered philopatric individuals did not differ between males and females (males: 5.8 ± 1.4 km, median = 4.2 km, range: 0.7 to 12.6 km, n = 11; females: 3.7 ± 1.8 km, median = 0.8 km, range: 0.3 to 12.7 km, n = l\t = 0.93, P = 0.37). Our estimate of c for the global model indicated there was little evidence of overdispersion (c < 1 ), and we set the variance inflation factor to ? = 1 and used AlCc for model selection. The minimum AICc model tor the survival analysis from Program MARK included two capture periods for apparent survival (cj)) and a constant encounter rate (p). Apparent survival in the first interval after capture ((t)i) was 0.048 (95% Cl: 0.030-0.075), whereas survival ((t)2+) fo*' birds during subsequent capture periods was 0.866 (95% CL 0.764-0.927). Encounter probabilities (p) were relatively low, 0.174 (95% Cl: 0.113- 0.230). The next best model included a time effect for encounter probabilities but was 28.8 AlCr units from the top model and received no weight relative to the top model (u’l < 0.01 ). DISCUSSION The frequency of returning Semipalmated Plovers marked as hatchlings is among the lowest reported for shorebirds (Thompson and Hale I'ABLH 1. Number ol' l^lover chicks per year and IVom that year’s cohort. newly banded Semipalmated number ol' new adult recruits Year Newly banded chiek.s Min. niiinber of recruits front cohort (%) 1988 57 1 (1.75) 1992 31 0 (0) 1 993 87 3 (3.44) 1994 92 1 (1.08) 1 995 81 1 (1.23) 1996 106 2 (1.89) 1997 75 2 (2.67) 1998 58 2 (3.45) 1 999 25 0 (0) 20(){) 28 3 (10.70) 2001 38 4 (10.50) 2002 44 1 (2.27) 2003 94 1 (1.06) 2004 50 2 (4.00) 2005 105 2 (1.90) 2006 103 0 (0) 2007 122 0 (0) 2008 75 0 (0) 1989, Flynn et al. 1999). Our estimate was similar to an earlier estimate from the same study area using observation data for returning juveniles from 1988 to 1996 (1.57%, n = 455, Flynn et al. 1999) despite a sample size that has nearly tripled. Our hatchlings were banded within 48 hrs of hatching, and low rates may be due to high mortality in the period after hatching when young are unable to thermoregulate or fly, and are vulnerable to inclement weather conditions and predators (Thompson and Hale 1989, Flynn et al. 1999, Sandercock et al. 2005, Colwell et al. 20()7a). Our survival estimates obtained for the first sample period were also low compared to that reported for the temperate congeners: Piping Plover (C. melodiis) and Snowy/Kentish Plover (C. a. alexandrinus) (survival from hatching to the following year, ()) = 0.15, Sandercock et al. 2005; (j) = 0. 1 8 for true survival of C. a. nivosus, Stenzel et al. 2007; (j) = 0.32 for C. melodiis banded as juveniles >16 days of age, Larson et al. 2000). Predator control can reduce mortality of hatch- lings for both temperate North American species (Stenzel et al. 2007). However, large differences between these rates and tho.se of Semipalmated Plovers may also be due to greater permanent emigration (through greater natal dispersal) from the Churchill study area, rather than lower survival, because this population of Semipalmated Plovers appeared to be stable over the study 26 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122. No. 1. March 2010 period (E. Nol, unpiibl. data) and otherwise could probably not sustain mortality of this magnitude. The difference between these survival rates among congeners also mirrors that seen in comparing natal philopatry and dispersal rates between migratory and resident passerines (Weatherhead and Forbes 1994, Sutherland et al. 2000). Lower rates of natal philopatry in long- distance migrants like Semipalmated Plovers may be due to dispersal costs accrued when they migrate south and relinquish familiarity with their natal area. It may only be important to find similar, but not necessarily familiar habitat when they return to breed (Weatherhead and Forbes 1994). There was substantial variability in number of recruits into the breeding season among hatchling cohorts, but a model of survival with time- dependence was not supported. A similar absence of time effects was also found for Kentish Plover in .southern Turkey (Sandercock et al. 2005) and, in both cases, may have been due to sparse data with low numbers of philopatric young. Alterna- tively, variation in the number of recruiting individuals may be more common when resources for breeding birds fluctuate widely (e.g.. Black- throated Blue Warbler, Dendroica caerulescens\ Rodenhouse et al. 2003) or where interference competition results in exclusion of potential recruits (Arcese et al. 1992). Low fecundity of arctic-breeding shorebirds may keep populations below the carrying capacity of breeding areas (Piersma and Baker 2000), but as yet we know little about the possible role of density-depen- dence in this group of birds. Territorial shorebirds are often assumed to have male-biased natal philopatry (reviewed in Colwell et al. 2007b) because of a re.source-defen.se mating system (Greenwood 1980). However, bias to- wards philopatric recruiting juvenile males, where it occurs, is usually small (Gratto et al. 1985, Reed and Oring 1993, Colwell et al. 2007b, Stenzel et al. 2007; but see Redmond and Jenni 1982). The advantages of familiarity with the natal site should be approximately equal for males and females of monogamous, bi-parental incubating species. Fe- male vSnowy Plovers share in incubation duties but typically de.sert the brood and, at times, re-nest with other males (Page et al. 1995). In this species, where females are presumed to have a relatively low attachment to the territory, the return rate of juvenile females to the local study area to breed is only 12% lower than for juvenile males, but still 64% (Stenzel el al. 2007). This difference reversed in winter with juvenile males more likely to disperse (54%) from wintering areas than juvenile females (44%). Thus, a re- evaluation of the theoretical basis for gender- bia.sed natal dispersal in birds is probably needed (Redmond and Jenni 1982, Sutherland et al. 2000). Most Semipalmated Plovers appear to delay breeding beyond 1 year although, contrary to the results from the previous study on this species (Flynn et al. 1999), at least one male bred in its first year. Delayed age at maturity differs from other studies of migratory shorebirds, where most individuals attempt to breed in their first year (reviewed in Sandercock et al. 2005). This suggests that a large proportion of the Churchill population may over-summer on wintering areas (e.g., McNeil et al. 1994), similar to reports for three species of calidridine sandpipers in Africa and Central America (Summers et al. 1995, O’Hara et al. 2002). The precise age of first breeding is difficult to ascertain because numbers of philopatric young were low and re-encounter probabilities of banded hatchlings are low (Sandercock et al. 2005). Age- specific variation in breeding propensity with higher rates of philopatry can be estimated with robust design models (Sedinger et al. 2001, 2008), but our data were not adequate for this approach. Most hatchlings were banded with an aluminum band and a single color band. Band loss and color fading after 3-4 years reduces the probability of observing these individuals if breeding in the Churchill study area. That two hatchlings were not detected in the local breeding population until the age of 8 years strongly suggests that some birds are temporarily leaving the population and then returning. In another study of this species on Akimiski Island, Nunavut, a breeding male underwent a long-distance (30 km) dispersal event. This also suggests, that despite generally greater nest site tenacity by males than females (Flynn et al. 1999, Fishman et al. in prep), males can and will disperse between breeding attempts in the same or different years. These difficult to document movements will continue to result in lower estimates of adult and juvenile survival for shorebirds breeding in mostly inaccessible habi- tats across the North American arctic, and reaffirm the importance of scale of the study area for documenting dispersal events (Greenwood and Harvey 1982, Jackson 1994, Sandercock and No! et cil. • SURVIVAl. OF’ JUVENll.E SEM 1 1'Al.M ATEF) F^L()VEF 0.6). Females alone incubated the eggs, and the incubation period was 14-15 days (mean ± SE = 14.2 ± 0.1 days; Eig. 2B). Hatching was gener- ally synchronous, and hatching success of eggs was high; of 160 eggs from 55 nests that survived through the end of incubation, 150 (93.8%) hatched. Nestling and Fledgling Periods. — Nestlings averaged 1.4 ± 0.1 g (mean ± SE) at hatching (n = 3 nests) and reached mean adult body mass of 9.5 g ~ day 7-8 after hatching (Fig. 3). Both males and females provisioned nestlings, and female brooding of young nestlings ended when nestlings reached 5-6 days of age. Overall provisioning rates at nests with nestlings of age 5-9 days was 20.3 ± 1.0 feedings/hr (n = 24 nests); females provisioned young at a signifi- cantly higher rate (11.7 ± 0.8 feedings/hr) than males (8.3 ± 0.5 feedings/hr; two-tailed paired t test, t = 4.23, df = 23, P < 0.001). Age at fledging was 1 1.3 ± 0.2 days (mean ± SE; Eig. 2C). Juveniles began to forage on their own by day 30, but parents continued to feed juveniles for at least 4 weeks after fledging; the latest record of adults feeding dependent young was of 40-day-old Juveniles that had fledged 29 days earlier. Number of nests Number of clutches Number of clutches 32 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. 1, March 2010 Clutch size Incubation period (days) Nestling period (days) Nestling age (days) FIG. 3. Growth in body mass of nestling Slate-throated Whitestarts in 1 1 nests of known hatch date in Monteverde, Costa Rica, 2000-2003. Each point repre.sents the mean ma.ss of nestlings from a single nest with either two or three young. Daily Nest Survival and Nest Success. — ^Esti- mated nest daily survival rate (DSR) for the 128 nests used in the Program MARK analysis was 0.968 with a 95% confidence interval (Cl) of 0.958-0.975. Estimated nest success, based on an average nest exposure period of 27.6 days (2.1 days between laying of the first egg and clutch completion, 14.2 days incubation, and 11.3 days during the nestling period), was 40.3% (95% Cl, 30.5^9.9%). An average (± SE) of 2.61 ± 0.07 young fledged from 74 successful nests. The MARK model of nest DSR that best fit the data included interaction terms between covari- ates year and date (Table 1). This model was 2.3 1 AIC(. units better than the second best model, which included the same two covariates in an additive model. These results suggest (1) there is considerable annual and seasonal variation in nest DSR, but (2) the precise nature of the seasonal effect of date is year-dependent. Estimated DSR varied annually from a low of 0.953 (95% Cl, 0.927-0.970) in 2004 to a high of 0.989 (95% Cl, 0.965-0.996) in 2003, coiresponding to estimated nesting success of 26.3 and 72.7%, respectively (Fig. 4A). Nest DSR generally declined over the course of the nesting season, but the nature of the FIG. 2. Variation in clutch size (A), length of the incubation period (B), and nestling period (C) in Slate- throated Whitestarts in Monteverde, Costa Rica. Mwuliw liREEDlNCJ BIOLCXJY ()!• MYIOHORUS MINIA'IVS 33 TABLE 1. Models of nest survival for 128 nests ol Slate-throated Whitestarts in Monteverde, Costa Riea, 2()()0-2()()4, based on analysis with Program MARK. Models are ranked from best to least supported aeeording to values of AAIC,, the differenee in Akaike's information eriterion for small samples between the listed model and the top model. The eovariates used in eaeh model are shown in parentheses. Other eolumns show the number of parameters estimated in eaeh model (K), model devianee, AAIC,, and Akaike weight (u,), a measure of the relative support for eaeh model. Model K Deviance AAIC, S(year X date) - interaction 10 333.48 O.OOO 0.606 Styear + date) - additive 6 343.88 2.3 1 2 0.191 S(year) 5 347.89 4.307 0.070 S(date) 2 354. 1 1 4.495 0.064 S(.) - constant survival 1 357.23 5.613 0.037 Stnest habitat type) 3 354.73 7.126 0.017 Stnest age) 2 356.94 7.326 0.016 seasonal decline varied from year to year (Fig. 4B). Neither nest site habitat type (road banks, trail banks, or natural slopes) nor nest age appeared to have a major influence on DSR; models incorporating these covariates received little support and were ranked lower than the null model of constant DSR (Table 1). Thus, DSR appears to be largely independent of both nest site characteristics and nest stage. Apparent predation of the entire clutch (/? = 19) or brood (/; = 26) was the leading cause of nest failure, accounting for 83.3% of all observed nest failures (/? = 54). Abandonment of intact nests with eggs ill = 3) or young in = 2) accounted for 9.3% of nest failures. An additional three nests (5.6% of failures) were abandoned after apparent partial nest predation reduced the nest contents to either one egg in = 2) or one nestling in = 1 ). I also observed a single case of nest failure that may have been a result of male egg removal; at one nest where the banded male had disappeared and been replaced by an unbanded immigrant, the three eggs were found intact underneath the empty nest. The female was observed building a new nest with her new mate 6 days later. DISCUSSION Breeding’ Biology.— My results are generally comparable to those ol other studies of the breeding biology of the Slate-throated Whitestart and other Mviohorns species. The March-June breeding season coinciding with the end ot the dry season and onset of the rainy season is consistent with the seasonality reported lor Slate-throated Whitestarts in Costa Rica (Skutch 1954, Shopland 1985, Stiles and Skutch 1989) and Venezuela (Collins and Ryan 1994), and the Collared Whitestart iM. torqnutns) at higher elevations in Monteverde (Shopland 1985). The Brown-capped Whitestart (M. hriinnkeps) in subtropical Argen- tina (26° S) is also strongly seasonal with a September-December nesting season coinciding with a rainy season that begins in November (Auer et al. 2007). However, recent work on the Spectacled Whitestart iM. melanocephcdns) sug- gests this species has a much more extended May- December breeding season in Ecuador (00° 36' S) that occurs during relatively dry periods (Greeney et al. 2008). Placement of Slate-throated Whitestart nests in banks along roads and trails, and on open steep slopes is consistent with other reports (Skutch 1954, Shopland 1985, Collins and Ryan 1994). Skutch (1954) reported the male participated in nest building at one of four nests where he observed nest construction. In contrast, I observed nest building at 31 nests where at least one member of the pair was color banded and gender was known, and males did not carry nest material or assist in nest construction. Skutch’ s report, however, is based on observations ot unbanded birds, and he did not see both members of the pair arrive at the nest simultaneously with nest material. He wrote “1 had proof that both birds worked at the nest, since I saw the second fly up to it with material in its bill, before the first, which had already added its contribution, had left the clearing” (Skutch 1954:360). However, the am- biguities in his description, and the possibility the bird building the nest had slipped away unseen by Skutch before returning with more nest material, suggest the possibility of nest construction by males of this species should be viewed skeptically until more persuasive evidence is presented. Nest construction in other species of Myioborns is either by the female alone or, in cases where gender could not be ascertained, a single adult (Marshall and Baida 1974. Barber et al. 2000, Greeney et al. 2008). The average clutch size of 2.9 eggs is similar to that reported by others in Costa Rica (Skutch 1954, 2.8 eggs, = 1 1 nests; Shopland 1985, 2.8 eggs, n = 33 nests), but slightly higher than the mean clutch size of 2.4 eggs in = 1 nests) reported for Slate-throated Whitestarts at 34 THE WILSON JOURNAL OF ORNITHOLOGY • Vul. 122, No. 1. March 2010 H Estimated daily survival rate H Estimated nest success 2000 2001 2002 2003 2004 All years Year combined FIG. 4. Annual and .seasonal variation in daily nest survival rate and overall nesting success of Slate-throated Whitestarts from Monteverde, Costa Rica, 2000-2004. (A). Annual variation in estimated daily survival rate and estimated nesting success. Mean it 95% confidence intervals are shown. (B) The effect of date on predicted daily nest survival rates for 2000-2004, based on the most strongly supported model of nest survival, the S(year X date) interaction model (Table 1). Predicted daily survival rates are shown only tor dates when active nests were being monitored. 10° 21' N, 67° 41' W in Venezuela (Collins and Ryan 1994). The clutch size of the congeneric Painted Whitestart (M. pictiis) in the Chiricahua Mountains of Arizona is larger, averaging 3.1 eggs (Barber et al. 2000). Mean egg size, 16.7 X 13.4 mm, was similar to that reported by Skutch ( 1954) (17.5 X 13.4 mm) and Collins and Ryan (1994) (16.3 X 12.3 mm). Egg size was not related to laying sequence, unlike many species of birds (e.g., Slagsvold et al. 1984, Litjeld et al. 2005). Incubation periods of 13- 15 days have been reported for Slate-throated Whitestarts in both Costa Rica (Skutch 1954) and Venezuela (Collins and Ryan 1994), but 1 observed incubation periods of 14-15 days. However, the mean incubation period 1 recorded in Monteverde, 14.2 days, is nearly identical to that reported by Skutch (1954) (14.4 day.s, // = 5 nests) and by Collins and Ryan (1994) (14.3 days, /; = 4 nests). The percentage of eggs surviving through the end of incubation that hatched successfully was 93.8%, relatively high for a tropical bird (Koenig Munmw BREEDING BIOLOCJY OB MY/OBORUS MINIAl'US 35 TABLE 2. Ijl'e-hislory traits of the Slate-throated Whitestart and three related north temperate zone warblers, the Red- faeed Warbler, Canada Warbler, and Wilson's Warbler. The three related species were selected based on the phylogeny of Lovette and Bermingham (2002). Data for Slate-throated Whitestarts based on this study. Data for other species derived from Martin and Barber (1995), Conway (1999) and Ammon and Gilbert (1999). Life-hi.sU)r>’ trail Slate-tliroaled Whitestart Red-l'aced Warbler Canada Warbler Wilson’s Warbler Breeding range Tropical montane forest Subtropical montane coniferous forest Northern boreal forest Northern boreal and humid Pacific forest Nest site Ground Ground Ground Usually ground, .some low in vegetation Mean clutch size 2.9 4.2 4.3 3.9 (Pacific coast) 4.4 (NE US) 5.3 (Alaska) Incubation period 14-15 days 13 days 11-12 days 11-12 days Nestling period 10-12 days (mean 1 1.3 days) 11-13 days (mean 12 days) 10-11 days 10-11 days Post-Hedging dependence 29 days after Hedging No data No data 24-25 days after Hedging 1982). The low incidence of hatching failure at Monteverde, a montane cloud-forest site with cool ambient temperatures during the nesting season (Clark et al. 2000), is consistent with the hypothesis that one of the major cau.ses of loss of egg viability and hatching failure is exposure to high ambient temperatures during the laying period (Stoleson and Beissinger 1999, Beissinger et al. 2005, Cooper et al. 2006). Duration of the nestling period averaged 1 1 .3 days, somewhat shorter than reported by Skutch (1954) (12.4 days, /? = 4 nests) elsewhere in Costa Rica, but similar to that found for Monteverde by Shopland (1985) (1 1.7 days, n = 16 nests). The relatively greater role of females in feeding nestlings was also noted by Skutch (1954) at a nest of this species in Guatemala. Nestling growth appears to be slightly more rapid than that of Venezuelan nestlings (Collins and Ryan 1994), and the average provisioning rate I observed at nests (20.3/hr) is about twice the rate at which food is delivered to Brown-capped Whitestart nests in subtropical Argentina (Auer et al. 2007). Few comparative data are available on duration of post-fledgling care of young, but both Skutch (1954) and Chipley (1976) noted the apparently long period of juvenile dependence by Slate- throated Whitestarts. My observation of adults feeding 40-day-old juveniles, 29 days after Hedg- ing is comparable to Shopland’s (1985) ob.serva- tion of Collared Whitestarts feeding young 28 days after Hedging. Nesting Success. — Daily survival rate (DSR) of Slate-throated Whitestart nests in Monteverde was 0.968, yielding an estimated overall nesting success of 40.3%, comparable to the 45% nest success reported by Skutch (1954) for 11 nests. Data on overall nest DSR are not available for other species of Myiohorus, but Auer et al. (2007) report the daily nest predation rate for Brown- capped Whitestarts in subtropical Argentina was 0.023. I found that —85% of nest failure was due to predation, and the comparable figure of daily nest predation rate for Monteverde Slate-throated Whitestarts would be 0.028. Two other studies of passerines in Monteverde have reported data on nest DSR and nesting success. Sargent (1993) found that nest DSR for Yellow-throated Euphonias {Euphonia hiritndina- cea), a species that nests in many of the same road banks favored by Slate-throated Whitestarts, was 0.972, yielding an estimated nest success of 33.6%. Nest DSR and estimated nesting success for cavity-nesting House Wrens {Troglodytes aedon) in Monteverde were considerably higher, 0.991 and 72.8%, respectively (Young 1994). My observed nest DSR of 0.968 in the Costa Rican highlands is greater than found in many studies of lowland neotropical birds (e.g., Robinson et al. 2000, Ryder et al. 2008) but comparable to an average nest DSR of 0.977 at 400 m elevation in Braulio Carrillo National Park, Costa Rica (Young et al. 2008). Comparison to North Temperate Zone War- blers.— Recent phylogenetic analyses of the Par- Lilidae and Myiohorus (Lovette and Bermingham 2002, Perez-Eman 2005, Rabosky and Lovette 2008) indicate the three North Temperate Zone warblers most closely related to the Myiohorus whitestarts are Red-faced Warbler {Cardellina 36 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. I. March 20/0 riihrifrons), Wilson’s Warbler (Wilsonia pitsilla), and Canada Warbler (W. canadensis). Molecular evidence suggests the Hooded Warbler (W. citrina) is not part of the clade that includes Myiohonis and the other Wilsonia (Rabosky and Lovette 2008). The Slate-throated Whitestart in Monteverde, in contrast to its relatives in the North Temperate Zone, displays the standard suite of life-history traits — small clutch size, prolonged incubation, and perhaps prolonged nestling and post-fledging care (Table 2) — that characterize tropical birds (Martin 2004, Russell et al. 2004). Differences in duration of incubation are particularly striking, as the incubation period is 1-3 days longer for Slate- throated Whitestarts than for Cardellina or Wilsonia (Table 2). Recent studies have shown that longer incubation periods of tropical birds are at least partially a result of slower inherent developmental rate of eggs (Tieleman et al. 2004, Robinson et al. 2008), although proximate differences in nest attentiveness and incubation temperature may have a role as well (Martin et al. 2007). The nestling period of Slate-throated Whitestarts in Monteverde is only slightly longer than for Wilsonia and apparently shorter than for Cardellina (Table 2). It also appears to be slightly shorter than for other Myiohorus whitestarts (Barber et al. 2000, Auer et al. 2007, Greeney et al. 2008). The relatively short nestling period in Monteverde merits further study, but may be affected by the relatively high rate at which food is delivered to nests (20.3 feedings/hr), the relatively rapid rate of nestling development, and the moderately high daily nest predation rate (0.028) (Martin et al. 2000). Several studies have indicated the post-Oedging period of juvenile dependence on parents is typically greater in tropical birds than in related temperate species (Skutch 1949, Russell et al. 2004, Tieleman et al. 2006). Data from Slate- throated Whitestarts are consistent with this hypothesis and suggest a slightly longer period of post-Oedging care than for Wifson’s Warbler (Table 2). However, additional data are needed to reach a firm conclusion about the relative duration of post-Hedging parental care in Myiohorus versus related temperate-zone warblers. ACKNOWLEDGMENTS I thank Pawcl Cygan. Mark Galatowilsch, Piotr .lablohski, Lydia Pctcll. Sarah Sargent. Tadck Stawarezyk, Ursula Valdez, and Scott Wissinger for field assistance. Marvin Hidalgo of the Estacion Biologica Monteverde (EBM) allowed me to work on EBM properties, and many other Monteverde landowners, including James Eorrest, Steven and Betsy Kendall, Sharon Kinsman, Alan and Karen Masters, Jim and Phoebe Richards, Bruce Young, and Willow Zuchowski provided me with acce.ss to their land. Alicia Mora and Erancisco Campos of the San Jose office of the Organization for Tropical Studies provided generous assistance with permits. Jack Eitniear and tw'o anonymous reviewers provided helpful criticisms of an earlier version of the manuscript. Einancial support was provided by grant 7194-02 from the Committee for Research and Exploration of the National Geographic Society, and by the Allegheny College Academic Support Committee. LITERATURE CITED American Ornithologists’ Union (AOU). 1998. Check- list of North American birds. Seventh Edition. American Ornithologi.sts’ Union, Washington, D.C., USA. Ammon, E. M. and W. M. Gilbert. 1999. Wilson’s Warbler (Wilsonia pusilla). The birds of North America. Number 478. Auer, S. K., R. D. Bassar, J. J. Eontaine, and T. E. Martin. 2007. Breeding biology of passerines in a subtropical forest in northwestern Argentina. Condor 109:321-333. Barber, D. R., P. M. Barber, and P. G. Jabonski. 2000. Painted Redstart (Myiohonis pictii.f). The birds of North America. Number 528. Beissinger, S. R., M. I. Cook, and W. J. Arendt. 2005. The shelf life of bird eggs: testing egg viability using a tropical climate gradient. Ecology 86:2164-2175. Chipley, R. M. 1976. The impact of wintering migrant wood warblers on resident insectivorous passerines in a subtropical Colombian oak woods. Living Bird 15:1 19-141. Clark, K. L., R. O. Lawton, and P. R. Butler. 2000. The physical environment. Pages 15-38 in Monteverde: ecology and conservation of a tropical cloud forest (N. M. Nadkarni and N. T. Wheelwright. Editors). Oxford University Press, New York, USA. Collins. C. T. and T. P. Ryan. 1994. Notes on the breeding biology of the Slate-throated Redstart (Myiohorus minialn.s) in Venezuela. Ornitologia Neo- tropical 5:125-128. Conway, C. J. 1999. Canada Warbler (Wilsonia canaden- sis). The birds of North America. Number 421. Cooper, C. B., W. M. Hocuachka, T. B. Phillips, and A. A. Dhondt. 2006. Geographical and seasonal gradi- ents in hatching failure in Eastern Bluebirds Sialia sialis reinforce clutch size trends. Ibis 148:221-230. CUR.SON, J., D. Quinn, and D. Beadle. 1994. Warblers of the Americas: an identification guide. Houghton MifOin, Bo.ston, Ma.ssachusctls, USA. Digiacomo, a. G. 1995. Two new species for Argentine avifauna. Hornero 14:77-78. Gill, E.. M. Wright, and D. Donsker. 2009. IOC World bird names (Version 2.0). www.worldbirdnames.org/ (accessed 10 February 2009). Mummc- BREEDING BIOLOCJY OB MYlOliORUS MINIA'IVS 37 Greeney, H. F., P. R. Martin, R. C. Dobhs. R. A. Gelis, A. D. L. Bucker, and H. Montag. 2008. Nesting ecology of the Spectacled WhitestaiT in Ecuador. Ornitologia Neotropical 19:335-344. Haber, W. A. 2000. Plants and vegetation. Pages 39-94 in Monteverde: ecology and conservation of a tropical cloud forest (N. M. Nadkarni and N. T. Wheelwright, Editors). Oxford University Press, New York, USA. jABtONSKi, P. G. 1999. A rare predator exploits prey escape behavior: the role of tail-fanning and plumage contrast in foraging of the Painted Redstart (Myiohorns pictns). Behavioral Ecology 10:7-14. Koenig, W. D. 1982. Ecological and social factors affecting hatchability of eggs. Auk 99:526-536. Lifjeld, J. T., a. Johnsen, and T. Petitguyot. 2005. Egg- size variation in the Bluethroat (Liiscinki s. svecica): constraints and adaptation. Journal fiir Ornithologie 146:249-256. Lovette, 1. J. AND E. Bermingham. 2002. What is a wood warbler? Molecular characterization of a monophyletic Parulidae. Auk 119:695-714. Marshall, J. and R. P. Balda. 1974. The breeding ecology of the Painted Redstart. Condor 76:89-101. Martin, T. E. 2002. A new view for avian life history evolution tested on an incubation paradox. Proceed- ings of the Royal Society of London, Series B 269:309-316. Martin, T. E. 2004. Avian life-history evolution has an eminent past: does it have a bright future? Auk 121:289-301. Martin, T. E. and P. M. Barber. 1995. Red-faced Warbler iCardellina nihrifrons). The birds of North America. Number 152. Martin, T. E., P. R. Martin, C. R. Olson, B. J. Heidinger, and j. J. Fontaine. 2000. Parental care and clutch sizes in North and South American birds. Science 287:1482-1485. Martin, T. E., S. K. Auer, R. D. Bassar, A. M. Niklison, and P. Lloyd. 2007. Geographic variation in avian incubation periods and parental influences on embry- onic temperature. Evolution 61:2558—2569. McCormack, J. E., G. Ca,staneda Guyasamin, B. MilA, and F. Heredia Pinada. 2005. Slate-throated Red- starts (Myiohorus miniatus) breeding in Maderas del Carmen, Coahuila, Mexico. Southwestern Naturalist 50:501-503. Mumme, R. L. 2002. Scare tactics in a neotropical warbler: white tail feathers enhance flush pursuit foraging performance in the Slate-throated Redstart (Myiohorus minialus). Auk 119:1024—1035. Mumme, R. L., M. L. Galatowitsch, P. G. Jabionski, T. M. Stawarczyk, and J. P. Cygan. 2006. Evolution- ary significance of geographic variation in a plumage- based foraging adaptation: an experimental test in the Slate-throated Redstart (Myiohorus minialus). Evolu- tion 60: 1086-1097. Nadkarni, N. M. and N. T. Wheelwright. 2000. Monteverde: ecology and conservation of a tropical cloud forest. Oxford University Press, New York, USA. Perez-Eman, j. L. 2005. Molecular phylogenetics and biogeography ol ihe neotropical redstarts (Myiohoi us, Aves, Parulinae). Molecular Phylogenetics and Evo- lution 37:51 1-528. Rabosky, D. L. and I. J. Ldveite. 2008. Dcnsity- dependenl diversification in North American wood warblers. Proceedings of the Royal Society ol London, Series B 275:2363-2371. Ricklefs, R. E. and M. Wikedski. 2002. The physiology/ life history nexus. Trends in Ecology and Evolution 17:462^68. Ridgely, R. S. and G. Tudor. 1989. The birds of South America. Volume I. The o.scine passerines. University of Texas Press, Austin. USA. Robinson, W. D., T. R. Robinson, S. K. Robinson, and J. D. Brawn. 2000. Nesting success of understory forest birds in central Panama. Journal of Avian Biology 31:151-164. Robinson, W. D., J. D. Styrsky, B. J. Payne, R. G. Harper, and C. F. Thompson. 2008. Why are incubation periods longer in the tropics? A common- garden experiment with House Wrens reveals it is all in the egg. American Naturalist 171:532-535. Rotella, j. 2008. Nest survival models. In Program MARK: a gentle introduction, Seventh Edition (E. Gooch and G. White, Editors), www.phidot.org/ software/mark/docs/book/ (accessed 10 February 2009). Russell, E. M., Y. Yom-Tov, and E. Geffen. 2004. Extended parental care and delayed dispersal: north- ern, tropical and southern passerines compared. Behavioral Ecology 15:831-838. Ryder, T. B., R. Duraes, W. P, Tori, J. R. Hidalgo, B. A. Loiselle, and j. G. Blake. 2008. Nest survival for two species of manakins (Pipridae) in lowland Ecuador. Journal of Avian Biology 39:355-358. Sargent, S. 1993. Nesting biology of the Yellow-throated Euphonia: large clutch size in a neotropical frugivore. Wilson Bulletin 105:285-300. Shopland, j. M. 1985. Facultative following of mixed- species flocks by two species of neotropical warbler. Dissertation. University of Chicago, Illinois, USA. Skutch, a. F. 1949. Do tropical birds rear as many young as they can nourish? Ibis 91:430-455. Skutch, A. F. 1954. Life histories of Central American birds: families Fringillidae, Thraupidae, Icteridae, Parulidae and Coerebidae. Pacific Coast Avifauna. Cooper Ornithological Society. Berkeley, California, USA. Slagsvold, T., j. Sandvik, G. Rofstad, O Lorentsen, AND M. Husby. 1984. On the adaptive value of intraclutch egg-size variation in birds. Auk 101:685- 697 .Stiles, F. G. and A. F. Skutch. 1989. A guide to the birds of Costa Rica. Cornell University Press. Ithaca. New York, USA. .Stole,son. S. H. and S. R. Beissinger. 1999. Egg viability as a constraint on hatching synchrony at high ambient temperatures. Journal of Animal Ecology 68:951—962. Tieleman, B. 1., .1. B. Williams, and R. E. Ricklefs. 2004. Nest attentiveness and egg temperature do not 38 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. I, March 2010 explain the variation in incubation period in tropical birds. Functional Ecology 18:571-577. Tieleman, B. I., T. H. Dukstra, J. R. Lasky, R. A. Mauck, G. H. Visser, and J. B. Williams. 2006. Physiological and behavioral correlates of life-history variation: a comparison between tropical and temper- ate zone House Wrens. Functional Ecology 20:491- 499. White, G. C. and K. P. Burnham. 1999. Program MARK: survival estimation from populations of marked animals. Bird Study 46 (Supplement): 120-138. Young, B. E. 1994. The effects of food, nest predation and weather on the timing of breeding in tropical House Wrens. Condor 96:341-353. Young, B. E., T. W. Sherry, B. J. Sigel, and S. WoLTMANN. 2008. Nesting success of Costa Rican lowland rain forest birds in response to edge and isolation effects. Biotropica 40:615-622. The Wilson Joiinial of Ornithology 122(1):39 45, 2010 BREEDING BIOEOGY OF KEEP GUEES ON THE BRAZIEIAN COAST GISELE PIRES de MENDON^A DANTAS'-^ AND JOAO STENGHEL MORGANTE' ABSTRACT. — We studied clutch size, hatching and fledging success, and time necessary lor chick Kelp Gulls {Lams clominicanns) to leave the nest throughout two breeding seasons (2004 and 2005) on Guararitama Island, Sao Paulo, Brazil. We followed 93 nests in 2004 and 97 nests in 2005. The average (± SD) clutch size was 2.09 ± 0.64 in 2004 and 1 .93 ± 0.59 in 2005. Hatching success was 74% in 2004 and 53% in 2005, and fledging success was 54% in 2004 and 58% in 2005. Chicks grew quickly, following the linear equation y(t) = 61g + 17.03g X age (in days), and began to fly at 40 days. Received 1 1 August 2008. Accepted 28 August 2009. The Kelp Gull (Larits clominicamis) (Family Laridae) is widely distributed in the Southern Hemisphere, occumng in South America, South Africa, Australia, New Zealand, the sub-Antarctic islands, and the Antarctic Peninsula (Burger and Gochfeld 1996). It is a resident and common species in Brazil, occurring from Rio de Janeiro to the coast of Rio Grande do Sul (Sick 1997). This species usually establishes colonies with few to thousands of breeding pairs on islets and islands near the continent and on continental locations (Burger and Gochfeld 1996). Studies have report- ed a large increase in Kelp Gull populations in recent decades (Steele and Hockey 1989, Quin- tana and Yorio 1998). The Kelp Gull population at Peninsula Valdez, Argentina, increased from 3,200 to 6,500 pairs over a 10-year period (Quintana and Yorio 1998). This population growth has been attributed to enhanced post- fledging survival of juveniles due to the provision of supplementary food through anthropomorphic activities (Steele and Hockey 1989, 1991; Bertel- lotti and Yorio 1999). Some authors believe gull population growth has been involved in displacement of other seabird species from breeding sites (Thomas 1972; Burger 1979, 1985; Quintana and Yorio 1998). Kelp Gulls have been observed to prey on eggs and chicks of Magellanic Penguin (Sphenis- cus magellanicus). Imperial Shag (Leitcocarho alriceps). Royal Tern (Sterna maxima), Cayenne Tern (S. eiirygnatha) and Southern Giant Petrel (Macronectes giganteus) (Yorio et al. 1998). Effects of gulls on coastal wildlife are not confined to other seabirds. Thomas (1988) and ' Institute de Biociencias, Universidade de .Sao Paulo, Brazil, Rua do Matao, 277, Sao Paulo, SP, Brazil, Cep. 05508-090. ^Corresponding author; e-mail: giseledantas@mail.icav.up.pt, dantasgpm@gmail.com Rowntree et al. (1998) observed Kelp Gulls injuring Right whales (Enhalaena australis) by picking off skin and fat when the whales emerge at the surface to breathe. These authors argue that intense harassment by gulls induces Right whales to abandon breeding areas before their young are sufficiently strong for the open sea. Little is known about the breeding biology of the Kelp Gull despite its wide distribution and its great impact on other seabirds and mammals. (Crawford et al. 1982, Soares and Fonseca Schielfler 1995, Yorio et al. 1996, Yorio and Borboroglu 2002). Breeding biology studies are essential to understand the dynamics of a population, because changes in population size may result from processes of reproduction, survival, and migration (Doherty and Grubb 2002). Reproductive success is associated with several factors, including ecological conditions (intra- and inter-specific interactions, food avail- ability, breeding site, etc.), and weather conditions (humidity, temperature, and wind) (Paynter 1949, Borboroglu and Yorio 2004). Studies that seek to understand long term survival and mortality rates in a population may be efficient for monitoring environmental change. An estimate of reproduc- tive success is also essential for evaluating population growth. This information is needed to create future conservation programs for seabird communities. We studied the breeding biology of the Kelp Gull on the south Brazilian coast. Our objectives were to: (1) measure the incubation period, (2) ascertain clutch size, (3) measure predation rate, (4) identify predators, (5) calculate the length ot time required for chicks to become independent, and (6) measure the growth rate of chicks. METHODS Study Area. — Data were collected on Guarar- itama Islet, Sao Paulo, Brazil (24° 23' S, 39 40 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. I, March 2010 46°59'W). This islet is 1.9 km from the continent and has an area of 1 .2 ha with a rocky substrate, and shrub and grass-herb vegetation. Campos et al. (2004) estimated the Lams dominicanus population on Guararitama to be 100 breeding pairs. Field Procedures. — We visited the islet every 5 to 7 days during 2004 and 2005 from the end of June to mid-November We recorded egg-laying dates, clutch sizes, egg measurements (length, width, and mass), chick measurements (mass, wing length, tarsus length, and beak length) and hatching dates. We measured eggs and chicks to the nearest 0.1 mm using calipers and to the nearest 1.0 g using a Pesola scale. We marked all Kelp Gull nests on Guararitama Islet with satin tape. We marked eggs with a waterproof pen and chicks with satin tape, until they reached an appropriate size to receive a metal ring provided by CEMAVE/IBAMA (Instituto Brasileiro do Meio Ambiente e Recursos Naturais Renovaveis). Nests were kept under surveillance throughout the breeding season to estimate survival, predation, and abandonment rates. A nest was considered active when it was well formed and contained eggs. It was considered to have been predated when broken eggs were found inside or near the nests, or when eggs disappeared without sign of hatching (e.g., excrement of the chicks). Nests were considered abandoned when eggs remained for 40 days without hatching. We defined hatching success as the number of chicks per eggs laid and fledging success as the number of chicks alive after 28 days per eggs laid (Mougeot and Bretagnolle 2006). Statistical Analysis. — ^Egg volume was calculat- ed following Hoyt (1979): V = length X width^ X 0.51. We estimated the number of females by the number of breeding pairs (bx), assuming that each nest consisted of one pair. The estimated number of females was used to calculate fecundity (fx) and fertility (mx) (Pianka 2()()()): bx = number of nests, fx = number of eggs laid/number of females, and mx = number of surviving eggs/number of females. Two distinct peaks were observed in egg laying in 2004 and 2005. The 2004 breeding sea.son was subdivided into an early and a later period, and the 2005 breeding season was divided into early, intermediate, and later periods, because the first peak was later in 2005. Differences within and between breeding seasons in egg volume and clutch size were evaluated with a t-test. Survival rates of offspring, hatching success, and fledging success were compared between breeding seasons using a yj test. Chicks captured on the day of hatching and at least one additional time during the breeding season (/? = 17) were u.sed to estimate the growth equation. We fitted the von Bertalanffy, Gom- pertz, and logistic three-parameter growth equa- tions (Ricklefs 1967) to the data on chick mass using least squares procedures to analyze chick growth patterns. We analyzed the variance of the residuals distribution to identify which models best fit the data, and chose the model with low variance. We estimated the age of all other chicks captured only once during the breeding season based on the chosen growth equation. Growth curves that best fit the body size measurements (beak, tarsus, and wing length) were calculated using von Bertalanffy, Gompertz, and logistic three-parameter functions. RESULTS The Kelp Gull breeding season on Guararitama Islet began in mid-June and ended in early November. The first eggs were found on 24 June in 2004 and on 30 June in 2005. Egg laying continued until October with the last eggs being laid on 31 October in 2004 and 28 October in 2005. We recorded 195 eggs in 93 nests in 2004 and 172 eggs in 97 nests in 2005. Egg laying in both years showed two peaks during the breeding sea.son. The first peak in 2004 occuiTed in July (93 eggs) and the second in September (102 eggs). The first peak in 2005 was in August (53 eggs) and the second in October (40 eggs) (Fig. 1 ). Mean (± SD) clutch size was 2.09 ± 0.64 (/; = 93) in 2004 and 1.93 ± 0.59 (/? = 97) in 2005. The difference between the years was not significant (/ = 1.77, P = 0.07). The average (± SD) incubation length was 24.12 ± 4.48 days in 2004 and 22.78 ± 3.27 days in 2005. The number of breeding females (bx) was 93 in 2004 and 97 in 2005. Fecundity (fx) was 2.096 in 2004 and 1.773 in 2005, and fertility (mx) was 1.387 in 2004 and 0.422 in 2005. Estimated egg volume differed between the two peaks of each breeding season (2004: t = 4.34, P < O.OOl; 2005: / = 2.24, P = 0.026). This difference cannot be attributed to the first peak consisting mainly of the first eggs laid by each Number of eggs laid Dantas unci Mor^ranlc • BREEDINCi BIOEOGY Ol KllLP CiUl.ES 41 80 FIG. E Number of eggs laid throughout the Kelp Gull breeding season in 2004 (gray bar) and 2005 (black bar), on Guararitama Islet, Sao Paulo, Brazil. female, and the second peak consisting of the second and third eggs laid, because all eggs from the first peak hatched during the period of time between the peaks. The average volume of eggs laid during the first peak in 2004 was 88.11 ± 8.28 cm^ while that of eggs laid during the second peak was 83.80 ± 12.68 cm^ The average egg volume during the early period in 2005 was 87.01 ± 8.12 cml The average egg volume during the first peak (intermediate period) was 87.18 ± 9.01 cm-\ and the average egg volume during the second peak (later period) was 82.43 ± 12.97 cnr\ The difference in egg volume between 2004 and 2005 was significant (t = 1.99, P = 0.046) with a mean (± SD) of 86.32 ± 10.37 cm^ in 2004 and 84.13 ± 10.75 cm^ in 2005 (Fig. 2). The first peak of the season produced greater hatching success than the second peak in both years (2004: f = 10.09, P = 0.006; 2005: f = 7 IP P = 0.03) (Table 1). Hatching succe.ss was greater in 2004 than in 2005 {yj = 18.13, F < 0.01) (Fig. 3). Predation was the principal cause of egg loss (25.6% in 2004 and 69.9% in 2005) and death of chicks (34.9% and 5.8%). The principal predators of Kelp Gulls were Black Vultures (Coragyps atratus) on eggs and adult Kelp Gulls on chicks. Heat stress also affected loss of eggs and chicks. Air temperature can reach 33° C during the late breeding season in Guararitama and, on hot days, the chicks appear lethargic and had difficulty breathing. Thus, they become easy prey for vultures and other raptors. Environmental data from Iguape Meteorological Station, Cptec/INPE, (www.cptec.inpe.br) tor the breeding period indicated that 2005 was hotter than 2004, but the years did not differ in precipitation (temper- 120,0 86.7 - 53,3 - A 20,0 120.0 86,7 - 53,3 20.0 B Early Laler 2004 Early Intermediate Later 2005 120.0 86,7 - 53,3 - 20,0 2004 2005 FIG. 2. Kelp Gull egg volume (x ± SD. cmb through- out the 2004(A) and 2005 (B) breeding seasons ami tor both years (C), on Guararitama islet, Sao Paulo. Brazil. ature: / = 8.78, P = O.OOO; precipitation: t = 0.53, P = 0.59). Field observations indicated that chicks grew rapidly, remaining in the nest until they were 5 days of age. When chicks were 5 to 1 5 days of age, they moved around the island and become more visible to predators. Chicks were able to avoid predators after 30 days of age. They reached 42 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. I. March 2010 TABLE 1, Hatching success rates throughout the breeding season of Lams dominicaniis in 2004 and 2005, at Guararitama Islet, Sao Paulo, Brazil. Year Period Successful eggs Predated eggs Abandoned eggs Totals 2004 Early 65 (69.9%) 16 (17.2%) 12 (12.9%) 93 Later 64 (62.7%) 34 (33.3%) 4 (3.9%) 102 Totals 129 (66.1%) 50 (25.6%) 16 (8.2%) 195 2005 Early 14 (35.9%) 23 (58.9%) 2 (5.1%) 39 Intermediate 18 (20.0%) 69 (76.7%) 3 (3.3%) 90 Later 9 (20.5%) 29 (65.9%) 6 (13.6%) 44 Totals 41 (23. 7%) 121 (69.9%) 1 1 (6.4%) 173 an average mass and body size that made them difficult to catch by predators at this time. Chicks were flying at 40 days of age. The survival rate to fledging was 54% in 2004 and 58% in 2005 (X- = 1.79; P = 0.1798). Three possible growth functions were com- pared (i.e., von Bertalanffy, Gompertz, and logistic functions). The variance of the residual distributions indicated the models did not differ significantly, and we present results only from the logistic equation. The logistic equation for body weight of 17 Kelp Gull chicks is weight (mass) = A/( l-hB*EXP{ — C*(days)}) (parameter values are in Table 2). We were able to predict the ages of chicks from their measured body mass using this curve. When chick mass increased linearly between 5 and 35 days of age, we derived a simplified formula: y = yo + cit, where yo is birth mass (mean mass on day of hatching = 61 g); a is the estimated growth rate of 17.03 g/day, and t is age in days (Fig. 4). We ob.served rapid growth in all of the measured traits (beak, tarsus, and wing length) (Fig. 5). These measurements were fitted to growth models, and there were no significant differences among models (parameters estimated for the logistic three-parameter model. Table 2). FIG. .1. Hatching and fledging succes.s rate.s (%) of Kelp Gulls during the 2004 and 200.‘i breeding seasons, on Guararitama Islet. .Sao Paulo. Brazil. DISCUSSION The Kelp Gull breeding season on Guararitama Islet was similar in both years, starting in mid- June and ending in early November. This same pattern was reported for the Santa Catarina coast in southern Brazil (Soares and Fonseca Schielfler 1995, Branco 2003), although Prellvitz et al. (2009) reported egg laying began on I June on Moleques do Sul, Santa Catarina. The Kelp Gull in Argentina reproduces from November to January in central and northern Patagonia (Bor- boroglu and Yorio 2004) and in mid-October in the Valdez Peninsula (Yorio and Borboroglu 2002). Yorio and Borboroglu (2002) report the Kelp Gull breeding season shows a direct relationship with latitude with gulls breeding earlier in more northern colonies. Laying stages may be related to latitude and to climatic conditions of the breeding localities (Fordham 1964). Clutch size of Kelp Gulls on the Brazilian coast varies from one to three eggs. This clutch size has been reported for Kelp Gulls in Brazil, Argentina, and South Africa (Saliva and Burger 1989, Soares and Fonseca Schielfler 1995, Branco 2003, Borboroglu and Yorio 2004, Prellvitz et al. 2009), and tor other species oi Lams in temperate regions (Paynter 1949, Kadlec and Drury 1968). The average clutch size in our study (2.09 in 2005 TABLE 2. Logistic growth model parameters. A = asymptotic growth rate; B = constant integration growth; C = growth rate of Kelp Gull chicks (/; = 17), at Guararitama Islet. .Sao Paulo, Brazil. A B c Body mass, g 1,165.34 14.25 0.1 1 Wing length, mm 1.35.. 30 4.86 9.41 Tarsus length, mm 77.88 1.78 0.06 Beak length, mm 47.63 1.41 5.69 Daniels (111(1 Morf^antc • l}REE[)INCj BIOLOGY OL KELP GULLS 43 FIG. 4. Growth (mass) of Kelp Gull chicks on Guararitama Islet. Sao Paulo, Brazil. and 1.95 in 2005) is comparable to that of other studies in Patagonia, Argentina (2.3-2.5 eggs, Yorio and Borboroglu 2002), southern Brazil (2.3 eggs, Soares and Fonseca Schielfler 1995; 1.98- 2.26 eggs, Branco 2003; 1.97 eggs, Prellvitz et al. 2009), and the Antarctic Peninsula (2.5 eggs, Fraser 1989). There were two peaks of egg laying at Guararitama Islet in both years. However, Yorio and Borboroglu (2002) in Argentina and Saliva and Burger (1989) in South Africa observed only one laying peak for Kelp Gulls. We believe the bimodal egg laying at Guararitama Islet may be a consequence of environmental conditions, including temperature and humidity. Prellvitz et al. (2008) reported Kelp Gull colonies are smaller in warmer and less productive waters, suggesting that environmental conditions can influ- ence size and reproductive success of colonies. Egg size is related to the ability of the female to obtain food resources throughout the laying period (Herbert and Barclay 1988). We observed significant differences in egg volume between the two laying peaks during each breeding season (Fig. 2). Females produced larger eggs during the early period than during the later period. Skorka et al. (2005) have shown that food availability is reduced late in the breeding sea.son. Thus, older pairs breeding early in the season are expected to have better success/survival than younger pairs due to greater experience and provisioning skills. Additionally, smaller average egg volume was observed in 2005 than in 2004 (t = 1.99; P = 0.046). These differences may reflect environ- mental changes that inlluence chick survival. The optimal temperature for bird embryo development is 37-38° C; temperatures above FIG. Growth (tarsus, wing, beak length) of Kelp Gull chieks on Guararitama Islet, Sao Paulo. Brazil, during the 2004 and 200.5 breeding seasons. 40.5° C are deadly (Gill 1994). The air temper- ature during the 2004 breeding season ranged from 17° to 36° C, and during 2005 from 17.5° to 40° C (Iguape Meteorological Station). Borbor- oglu and Yorio (2004) reported the breeding season of Kelp Gulls at Isla Vernacci, southwest- 44 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122. No. /, March 2010 ern Argentina, ends in the middle of January, when the temperature reaches 40° C. By that time, the chicks are already flying. Salzman (1982) reported that temperatures above 32° C are harmful to chicks of Lams occidentalis, often causing death. Salzman ( 1982) reported that stress caused by excessive heat is one of the main causes for egg loss and chick mortality. The high mortality observed in 2005 may have been due to high temperatures recorded through- out the breeding season, which left nests exposed to predators. The principal cause of egg loss (25.6% in 2004, 69.9% in 2005) at Guararitama was predation by vultures, while the principal cau.se of chick mortality (34.9% in 2004, 25.0% in 2005) was predation by adult Kelp Gulls. Borboroglu and Yorio (2004) reported egg loss of 28% and chick morality of 50%. These authors also reported chick predation by adults of the same species. They believe that chick predation by adults of the same species occurs when chicks invade breeding territories, searching for shade and protection under vegetation. Predation can be reduced by parental care and by selection of a more hidden site for nest building (Borboroglu and Yorio 2004). Other predators observed to prey on eggs and chicks of Kelp Gulls includes Subantartic Skuas (Stercorariiis antarcticiis) and Southern Crested Caracara (Caracara plancus) (Yorio et al. 1995, Sick 1997, Yorio and Borboroglu 2002). Howev- er, Prellvitz et al. (2009) observed parental negligence and starvation to be the main mortality factor for eggs and chick Kelp Gulls, respectively, on Deserta Island. The number of eggs abandoned and chick death by starvation was small (<5%) in our study. Pierotti and Bellrose (1986) reported that high los.ses from parental negligence or egg fertility could indicate lower food availability or inferior food quality. Prellvitz et al. (2009) reported 51% hatching success on Deserta Island, similar to that in 2005 on Guararitama Islet. However, hatching success on Guararitama in 2004 was 74%, similar to that reported by Yorio and Borboroglu (2002) at Peninsula Valdez, Argentina. These differences may be related to environmental variability in temperature and humidity. Stable years can have great hatching and fledging success and hotter years can result in great mortality from predation or starvation. Kelp Gull chicks grow rapidly and this species displays parental care during incubation and feeding of chicks. Borboroglu and Yorio (2004) reported that parental care gradually decreases over the breeding season, increasing predation risk. Kelp Gull chicks start to move around the island after 5 days of age. This movement, when they are only a few days of age, seems to be a behavioral characteristic (Yorio et al. 1996). Chicks weigh 800 g at 35 days and are already flying. Capture by predators at this phase is difficult. Yorio and Borboroglu (2002) also observed rapid growth of chick Kelp Gulls in colonies on the coast of Argentina. Prellvitz et al. (2009) reported that most chick mortality on Deserta Island, Brazil, occurred in the first week of life. This rapid growth rate may represent an evolutionary adaptation for escape from predators. Kelp Gulls breed on the Brazilian coast during the winter, probably to avoid higher summer temperatures and to seek better conditions for embryo development. Individuals that breed early in the season have an advantage over pairs breeding later in the .season. The main predators of eggs and chick Kelp Gulls are Black Vultures. who.se impact is greatest at the egg stage, and adult Kelp Gulls, which prey on young chicks. Chicks have rapid growth, probably as a means of avoiding predators. Despite the high predation rate observed in 2005 (a hotter year). Kelp Gulls had a high survival rate (54% in 2004 and 58% in 2005). If these chicks survive to the reproductive stage, this suggests an incremental population growth rate of 50% each year. This population growth may pose a serious problem for the conservation of other seabirds. ACKNOWLEDGMENTS This work was funded hy grants from the Funda^ao de Amparo a Pesquisa de Sao Paulo (Fapesp; processo O.'i/ 354.J8-4) and a Ph.D. fellowship grant to Gi.sele Dantas (processo 03/01599-1). Wc thank the Instituto Florestal de .Sao Paulo, Esta9ao Ecologica .lurei-ltatins for logistical support (783/2003), and CEMAVE/IBAMA for metal rings and for the license to perform the study (permit 1060). We also thank several friends for help with fieldwork, and M. A. Aires, J. O. Braneo, C. Y. Miyiaki, M, C, Arias, F. A. Santos, and anonymous reviewers for helpful comments on earlier drafts of the manuscript. LITERATURE CITED BERTKi.i.orri, M. and P. Yorio. 1999. Spatial and temporal patterns in the diet of Kelp Gull in Patagonia. Condor 101:790-798. Borboroglu, P. G. and P. Yorio. 2004. Flabitat requirements and selection by Kelp Gulls (Lants dontinicanus) in central and northern Patagonia, Argentina. Auk 121:243-252. D3 fledglings). These categories were used to de- crease variability while maintaining the structure of the relationship between the outcome and the predictor variables (Hosmer and Lemeshow 2000). We used multinomial logistic regression to evaluate variables as predictors of Buirowing Owl fledging success categories for each habitat type. Data within habitat types (grassland vs. urban) were combined for the 2 years, and regression analysis was conducted separately for each habitat type. We developed a set of candidate a priori models for each habitat type that we believed would be the most biologically meaningful, 16 for 54 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 1, March 2010 TABLE I. Habitat variables (mean ± SE) and t statistics for BuiTowing Owl nest sites in urban (Las Cruces) and grassland (Armendaris Ranch) habitats pooled for 2000 and 2001 in southern New Mexico. Sites Nest site variables Urban vs. grassland Grassland by year Urban Grassland 041 P Nearest-neighbor distance, m 159.1 ± 11.7 49.7 ± 6.3 7.66 <0.001 # Nests within 75 m 0.8 ± 0.1 3.6 ± 0.3 -8.99 <0.001 Nest density/ha 0.13 ± 0.01 1.58 ± 0.10 -14.12 <0.001 # Satellite burrows 10.7 ± 1.3 44.3 ± 2.4 -13.14 <0.001 Perch distance, m 184.0 ± 56.3 27.9 ± 3.4 2.46 0.015 2000 2001 ^61 p Nearest-neighbor distance, m 45.0 ± 7.6 52.2 ± 9.1 -0.54 0.589 # Nests within 75 m 2.8 ± 0.5 4.0 ± 0.4 1.81 0.093 Nest density/ha 1.6 ± 0.1 1.9 ± 0.1 -3.78 <0.001 # Satellite burrows 38.2 ± 4.4 47.6 ± 2.6 -1.93 0.058 Perch distance, m 31.4 ± 7.6 26.0 ± 3.2 0.76 0.451 gra.ssland and 18 for urban habitats. We used a balanced combination of the same five variables for each habitat type (year, number of nests within a 75-m radius of each nest, fledging success of the nearest neighboring nest, number of satellite burrows within a 50-m radius of each nest, and distance to nearest perch) but added one addition- al variable, category of urbanization, to the urban data set. The city of Las Cruces makes multiple distinctions concerning land use, but to increase the number of nests within a land-use group, we limited land use to two categories: open space (including agricultural fields, parks, arroyos, and vacant lots) and developed land (including residential, commercial, and community lands). We calculated Akaike’s Information Criterion (AIC) corrected for small sample size (AIC^.) for each model using SYSTAT 12.0 (Systat 2007). All models were ranked according to AAIC^. and models with a AAIC^. 2 were considered competing models (Burnham and Andersen 2002). Model weights were calculated for each model and cumulative weights for all variables were calculated. All variables were evaluated for pairwise collinearity using Pearson Correlation. None of the variables included in either sets of models was considered collinear (variables corre- lated < 0.70). A Chi-.square homogeneity test was u.sed to evaluate whether a burrow was more likely to be re-used if the nest had been successful the previous year. A Mann Whitney U test was u.sed to compare fledgling success between new nest burrows and burrows that were re-used as nests from the previous year. All statistical analyses were performed using SYSTAT, Version 12.0 (Systat 2007). RESULTS Sixty-six nests were located and monitored in 2000, 43 within 19 urban sites and 23 within six prairie dog colonies. Seventy-eight nests were located in 2001, 37 within 15 urban sites and 41 within seven prairie dog colonies, including one newly established colony. There was no differ- ence in the mean number of nests per site in urban areas between years {P = 0.470), but the number of nests established within the original six prairie dog colonies in grassland habitats was greater in 2001 (Hi = —3.72, P = 0.003). In urban areas, 15% of nests were associated with rock squiiTel systems. All monitored nests in grassland areas on the Armendaris Ranch were associated with black-tailed prairie dog colonies. Site characteristics and owl population dynam- ics differed between the two habitats (Table 1). Mean nearest-neighbor nest distances in grass- lands were significantly closer, and the mean number of nests within 75 m and nest densities were both significantly greater compared to urban sites (Table I). The mean number of satellite burrows within a 50 m-radius of the nest was significantly higher, whereas mean perch distance of the adult male from the nest burrow was significantly less in grassland compared to urban habitats (Table 1 ). Grassland site characteristics exhibited differ- ences between years. The number of nests in prairie dog colonies almost doubled the second year and, predictably, nest density was greater in Bcrardelli ct al. • BURROWING OWL REPRODUCTIVE SUCCESS 55 TABLE 2. Burrowing Owl nesi success in urban and grassland habitats in southern New Mexico pooled for 2000 and 2001. Percent reproductive effort within three Hedging success categories (failure, low success, high success) was examined using a Chi-square test for homogeneity (X- = 5.59, df = 2, P = 0.061). Fledging data were not normally distributed. Productivity piirameters Urban Grassland % Failure'’ (/;) .32.5 (26) 18.8 (12) % Low succes.s*’ (/;) 16.5 (13) 28.0 (18) % High success'-' {n) 51.0 (41) 53.0 (34) Totals 100 (80) 100 (64) Successful nests'* {% ± SE) 54 (67.5 ± 5.6) 52 (81.3 ± 4.8) Young fledged 208 160 Fledgling.s/attempt (X ± SE) 2.6 ± 0.25 2.5 ± 0.22 Fledglings/successful nest (X ± SE) 3.9 ± 0.22 3.1 ± 0.20 ■* Failure = 0 fledglings. Low success = 1-2 fledglings. High success = 3+ fledglings. Successful nests = si chick. 2001 (Table 1). However, no difference was detected in the mean number of nests within a 75-m radius of a nest, mean nearest-neighbor distance, or mean perch distance from the nest between years (Table 1). The mean number of satellite burrows within a 50-m radius of a nest in 2001 was higher (Table 1). No differences {P > 0.05) in site characteristics were ob.served be- tween years in urban landscapes. No differences were detected among the three reproductive success categories between habitat types {X- = 5.59, df = 2, P = 0.061); however, there was a tendency for higher nest failure in urban (32.5%) compared to grassland (18.8%) habitats, but more nests in the low-success category for grassland compared to urban sites (28 vs. 16.5%; Table 2). Sixty-eight percent of Burrowing Owl pairs ne.sting in urban areas compared to 81% in grassland habitats success- fully fledged at least one chick (Table 2). The mean number of young fledged per successful nest was 3.85 and 3.07 for urban and grassland sites, respectively. On average, nests in urban and grassland areas fledged 2.60 and 2.50 young per nest attempt, respectively (Table 2). Differences were detected in variables influ- encing fledging success between the two habitats. The global model containing all five variables in grassland habitats had overwhelming support (Table 3). This was the only competing model for grassland sites with a model weight of 0.99. Higher success was associated with nests having fewer satellite burrows within a 50-m radius (40 vs. 57), greater perch distance from the nest (34 vs. 19 m), and lower fledging success of the nearest Burrowing Owl nest (2.40 vs. 3.60 fledglings), as compared with failure. Owls with low success had a greater number of owl nests within a 75-m radius (5.60 compared to 2.70) when compared with high success owls. A year effect was also observed, as significantly more nests were in the low success category the second year (Table 3). Cumulative model weights sup- ported the importance of all five variables (Table 4). The two top models in urban habitats were competing models; variables included in the top model were number of nests within a 75-m radius, perch distance, and number of satellite buiTOWs TABLE 3. Top three multinomial logistic regression models for each habitat type ranked in order of increasing AAIC, influencing Burrowing Owl fledging success in urban and grassland habitats in southern New Mexico in 2000 and 2001. -i- (or — ) on a predictor variable indicates increased (or decreased) probability ol high-success vs. lailure or high success vs. low success as the predictor variable increases for the two multinomial submodels. Habitat Models High v.s. failure High vs. low succe.ss K Ate, AAIC, U’, Grassland 1. NEST, PERCH, NNFS, SAT, YR -fNEST, -perch, -t-NNES, +SAT, -i-YR -hNEST, -PERCH, -NNFS, -SAT, -YR 5 98.94 0 0.99 2. NEST, PERCH, SAT +NEST, -PERCH, +SAT +NE,ST, -PERCH, -SAT 3 107.86 8.92 0.01 3. NEST, NNES, SAT -fNEST, -fSAT, -fNNFS -fNEST, -sat, -NNFS 3 112.26 13.32 0 Urban 1. NEST, PERCH, SAT -NEST, -PERCH, -t-SAT -t-NEST, -PERCH, -hSAT 3 157.03 0 0.44 2. NEST, SAT, URB -NEST, +SAT, -t-URB -t-NEST, -SAT, +URB 3 158.63 1.60 0.20 3. NEST -NEST -hNEST 3 160.36 3.33 0.08 56 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. I, March 2010 TABLE 4. Cumutative model weights of variables in a priori multinomial logistic regression models used to identify habitat characteristics that predicted Burrowing Owl fledging success in grassland and urban environments in southern New Mexico for 2000 and 2001. Habitat Variable (j) W + (j) Grassland Year 0.99 Nests 1.00 NNFS 0.99 Satellites 1.00 Perch distance 0.99 Urban Year 0.18 Nests 0.89 NNFS 0.17 Satellites 0.73 Perch distance 0.61 Urbanization 0.36 within 50-111 of each nest. The second ranking model contained the number of nests within a 75- m radius, number of satellite burrows within a 50- m radius, and measure of urbanization (Table 3). Nests in the high success category, as compared to failed nests, had fewer nests within a 75-m radius (0.40 vs. 0.80) and nests in open space had higher success than nests in developed areas (Table 3). The number of satellite burrows was retained in both models with a higher number associated with high versus low success. Perch distance was retained in one of the top two models with greater perch distance associated with high success compared to low success and failure. Cumulative model weights supported the number of nests within a 75-m radius, number of satellite burrows within a 50-m radius, and perch distance as strongly inlluencing fledging success; number of nests negatively, and number of satellite bun'ows and perch distance positively influenced produc- tivity (Table 4). Owls in urban areas in 2001 re-used 51.4% of nest burrows from the previous year whereas in prairie dog colonies 22.0% of the nest burrows were re-used. There was no significant relation- ship between burrow re-use and previous nest success (P > 0.05). There was also no relation- ship between burrow re-use and nest success in urban environments (P = 0.373), but there was a tendency for owls in grasslands occupying a burrow used the previous year to produce more fledglings than new nest burrows {V = 90, P = o.oni). DISCUSSION BuiTowing Owl nest densities were approxi- mately one order of magnitude greater in prairie dog colonies than in urban areas. The majority of nests in urban areas were associated with rock squin'el bun'ows, which resemble prairie dog colonies as both provide multiple nesting oppor- tunities. However, rock squirrel colonies are smaller, more ephemeral, and may be in open rocky areas, flat ten'ain, imgation canals, em- bankments, or among rock piles (Findley 1987). Prairie dog colonies are larger, provide more nesting opportunities, have higher burrow densi- ties, occur in grasslands, and are more stable (King 1955). Prairie dog colonies had a greater number and density of buirows, covered substan- tially more area compared to rock squiirel colonies, and were surrounded by more foraging habitat. Owls nesting in urban environments maintained nearest neighbor distances (X ± SE = 159 ± 12 m) similar to other studies in urban and agricultural systems (Rosenberg and Haley 2004) whereas spacing of nests in prairie dog colonies in our study was closer (A" ± SE = 50 ± 6 m) than commonly observed in this system (Desmond et al. 1995, Griebel and Savidge 2007). However, available data supports closer spacing in prairie dog colonies for southern sites (Ross 1974, McNicoll 2005, Teaschner 2005), possibly attributable to fewer and smaller colonies result- ing in fewer suitable nesting sites. This is supported by the rapid occupancy of small colonies (A ± SE = 4.85 ± 1.33 ha) on our study area by Burrowing Owls immediately following establishment (T. E. Waddell, pers. comm.). Eledging success did not differ between the two habitat types as predicted but was low in comparison to average clutch sizes for Burrowing Owls (7-9 eggs; Haug et al. 1993) and reproduc- tive performance from several other studies (CoLilombe 1971; Martin 1973; Botelho and Arrowood 1998; Lutz and Plumpton 1999; MJD, unpubl. data). Different mechanisms appeared to be influencing the low productivity at these sites. Sixty-eight percent of all nest failures in our study occurred in urban sites; nests in these sites tended to either fail or be highly successful. In compar- ison, grassland sites had lower rates of failure but produced fewer young per successful nest. Higher nest failure in urban areas suggests higher rates of Bcrardclli cl al. • BURROWING OWL REPRODUCTIVE SUCCESS 57 nest predation and/or abandonment in this envi- ronment, whereas fewer lledglings per successful nest in grasslands suggest increased competition for resources. We were not, however, able to differentiate between predation and nest abandon- ment. Owls in west Texas experienced higher rates of disturbance in urban compared to rural environments (Chipman 2006), but it was not known if this influenced productivity. Foraging data from this study suggested that owls in urban environments may have higher foraging rates possibly compensating for increased disturbance. Studies in Canada (Wellicome 2005) and Califor- nia (Haley 2002), using food supplementation, found that Burrowing Owls were food-limited during the nestling period. York et al. (2002) attributed low reproductive success of Burrowing Owls in the Imperial Valley of California to the lower amounts of vertebrate prey in their diets compared to other studies. Fledging rates in grassland habitats appeared to be negatively impacted by the high density of owls in these small, recently established colonies. This was most apparent in 2001 when number of nests in prairie dog colonies almo.st doubled, resulting in a higher density of nesting owls compared to the previous year. Measures of BuiTowing Owl density appeared to strongly influence productivity at both sites but for different reasons. Productivity in our grassland habitats was strongly negatively impacted by measures of nest/owl density resulting in fewer fledglings per successful nest. This suggests that competition for resources and/or aggressive inter- actions were factors in this environment. Griebel and Savidge (2007) in South Dakota reported higher productivity for nests that were further from the nearest neighboring nest. The higher success of nests with fewer satellite burrows in our study was associated with nests along colony edges, suggesting that reproductive success was higher for edge nests. Several studies have suggested that Burrowing Owls prefer to nest near colony edges (Eckstein 1999, Desmond et al. 2000, Orth and Kennedy 2001). Possible advan- tages include reduced competition related to fewer surrounding nests and easier access to off-colony foraging. Perching distance is an indication of the territory size suiTOunding the nest burrow and sentinel owls closer to the actual nest likely had smaller territories and increased competitive interactions. Owl proximity to the nest may also increase nest vulnerability to predation; however. predation rates in the incubation and early nestling stage for this habitat type were low (81% ol nests were successful). Nest failure rates were higher in urban habitats where owls perched lurther from the nest. The combination of increased owl densities and below normal precipitation in grassland habitats in 2001 likely contributed to increased competition for fewer resources on those sites. Our results from grassland habitats support the need for larger prairie dog colonies in the southern plains that have experienced greater fragmentation compared to northern populations (Miller et al. 1994). Desmond et al. (1995) found that BuiTOwing Owls in prairie dog colonies exhibited clustered distributions with greater inter-nest distance when colonies were ^35 ha, and were randomly distributed with smaller inter- nest distances in colonies <35 ha. These data suggest owls require larger colonies to exhibit preferred nest spacing, and our results indicate that closer nest spacing negatively impacts productivity. Nest density in urban areas also appeared to have the strongest impact on reproduction. Nests in the high success category were associated with lower densities of nesting owls compared to the low success category, suggesting competition for resources also influenced reproduction at these sites. However, failed nests had the lowest nest density and tended to be solitary, suggesting the presence of other nesting pairs may increase productivity, possibly through increased predator detection, as long as sufficient space is available to reduce competition for nearby resources. Most unsuccessful nests failed during the incubation or early nestling .stages, suggesting nest abandon- ment or predation. Increased development in Florida resulted in reduced owl densities despite higher prey populations in these environments (Millsap and Bear 2000). The loss of suitable burrows and space for multiple nesting pairs may increase predation risk for owls nesting in developed areas. Productivity in urban environments in our study was also associated with land use practices with high nest success associated with open space compared to developed areas. Open space is associated with larger habitat patches that provide more nesting opportunities and may have lower predation pressure and less disturbance (vehicular or pedestrian traffic). Density of some of the major urban predators, including domestic cats {Fclis catus) and common raccoons (Procyon lotor) are 58 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122. No. I, March 2010 likely higher in residential areas; we observed owls mobbing domestic cats at urban sites on several occasions. As Burrowing Owls become increasingly associated with urban environments, city planning offices should work with local and state biologists to identify and preserve areas important to Burrowing Owls. This should include identifying areas free from disturbance with sufficient space for multiple nesting pairs, pres- ervation and management of fossorial rodent populations associated with Buirowing Owls, and installing and maintaining artificial nest burrows when necessary. Owls in urban habitats re-used significantly more nest burrows than those in grassland sites. Two factors likely contributed: burrow availabil- ity and resident owls. Fewer suitable bun'ows were available to owls in urban compared to grassland habitats. Holmes et al. (2003) reported re-use rates in Oregon ranging from 57 to 87% where owls primarily nest in American badger {Taxidea taxus) burrows. They attributed high re- use rates to a shortage of suitable bun'ows, especially in sandy soils. Fewer burrows are available in rock squirrel colonies, and rock squirrel populations in urban areas are often eradicated due to conflict with human activities. A portion of our urban owl population was resident year-round (27% of urban owls in 2001 were present the first week in Feb). Bun'owing Owls in a mainly resident population in California had 85% nest site fidelity (Rosenberg and Haley 2004). The combination of a partially year-round resident population and fewer suitable nest burrows in urban compared to grassland habitats likely contributed to a higher re-use rate in urban environments. ACKNOWLEDGMENTS T. E. Waddell, J. C. Truett. and T. A. Ostemier of Ihc Turner Foundation and the Turner Endangered Speeies Fund granted access to the Armendaris Ranch and graciously provided logistical support. T. S. Schrader of New Mexico State University and T. J. Hunteman ot the city of Las Cruces assisted with nest mapping and AreView. Field a.ssistance was provided by J. S. Enriquez. L. A. Murphy, I. R. Murray. E. J. Quintana, and J. P. Waring. This manuscript was improved by comments from P. C. Arrowood. C. E. Braun, E. L. Fredrickson. S. R. Sheffield, B. C. Thompson, and an anonymous reviewer. Funding was provided by Threatened and Endangered Species Inc. (T&E Inc.) and New Mexico State University. This is a New Mexico Agricultural Experiment Station publication, sup- ported by state funds and the U.S. Hatch Act. This publication is dedicated to the memory of T. O. Wooten. a staunch advocate for research and conservation in the Chihuahuan Desert. LITERATURE CITED Botelho, E. S. and P. C. Arrowood. 1998. The effect of burrow site use on the reproductive success of a partially migratory population of western Burrowing Owls. Journal of Raptor Research 32:233-240. Burnham, K. P. and D. R. Anderson. 2002. Model selection and multimodel inference: a practical information-theoretic approach. Springer-Verlag, New York, USA. Chipman, E. D. 2006. Behavioral ecology of western Burrowing Owls {Athene ciinicidaria hypiigaea) in northwestern Texas. Thesis. Texas Tech University, Lubbock, USA. COULOMBE, H. N. 1971. Behavior and population ecology of the Burrowing Owl {Speotyto cunicularia) in the Imperial Valley of California. Condor 73:162-176. Conway, C., V. Garcia, M. D. Smith. L. A. Ellis, and J. L. Whitney. 2006. Comparative demography of BuiTowing Owls in agricultural and urban landscapes in southeastern Washington. Journal of Field Orni- thology 77:280-290. Desmond, M. J. and J. A. Savidge. 1999. Satellite burrow use by Burrowing Owl chicks and its influence on nest fate. Studies in Avian Biology 19:128-130. Desmond, M. J., J. A. Savidge, and T. F. Seibert. 1995. Spatial patterns of Burrowing Owl (Speotyto cunicularia) ne.sts in black-tailed prairie dog (Cynomys hulovicianus) towns. Canadian Journal of Zoology 73:1375-1379. Desmond, M. J., J. A. Savidge, and K. E. Eskridge. 2000. CoiTelations between Burrowing Owl and black-tailed prairie dog declines: a 7-year analysis. Journal of Wildlife Management 64:1067-1075. Desmond, M. J., T. J. Parsons, T. O. Powers, and J. A. Savidge. 2001. An initial examination of mitochon- drial DNA structure in Burrowing Owl populations. Journal of Raptor Research 35:274-281. Desante, D. F., E. D. Ruhlen, and D. K. Rosenberg. 2004. Density and abundance of Burrowing Owls in the agrieultural matrix of the Imperial Valley, California. Studies in Avian Biology 27:1 16-1 19. Eckstein, R. T. 1999. Local and landscape factors affecting nest site selection and nest success of Burrowing Owls in western Nebraska. Thesis. University of Nebraska. Lincoln, USA. Findley, J. S. 1987. The natural history of New Mexico mammals. University of New Mexieo Press, Albu- querque. USA. Garcia. V. and C. J. Conway. 2009. What constitutes a nesting attempt? Variation in criteria causes bias and hinders comparisons across studies. Auk 126:31^0. Gorman, L. R., D. K. Rosenberg, N. A. Ronan, K. L. Haley, J. A. Gervais, and V. Franke. 2003. Estimation of reproductive rales of Burrowing Owls. Journal of Wildlife Management 67:493-500. Griebel, R. L. and j. a. Savidge. 2007. Factors influencing Burrowing Owl reproductive performance Bcrardi’lli et al. • BURROWING OWL REPRODUCTIVE SUCCESS 59 in unfragniented shoitgrass prairie. Journal of Raptor Research 41:212-221. H.aley, K. L. 2002. The role of food limitation and predation on reproductive success of Burrowing Owls in southern California. Thesis. Oregon State Univer- sity, Corvallis, USA. Hartsough, M. J. 2002. Dietary preferences of black-tailed prairie dogs reintroduced into a Chihuahuan Desert grassland. Thesis. New Mexico State University, Las Cruces, USA. Haug, E. a. and L. W. Oliphant. 1990. Movements, activity patterns, and habitat use of Burrowing Owls in Saskatchewan. Journal of Wildlife Management 54:27-35. Haug, E. A., B. A. Millsap, and M. S. Martell. 1993. Burrowing Owl. The birds of North America. Number 61. Holmes, A. L., G. A. Green, R. L. Morgan, and K. B. Livezey. 2003. Burrowing Owl nest success and burrow longevity in north central Oregon. Western North American Naturalist 63:244-250. Holroyd, G. L., R. Rodriguez-Estrella, and S. R. Sheffield. 2001. Conservation of the Burrowing Owl in western North America: issues, challenges, and recommendations. Journal of Raptor Research 35:399- 407. Hosmer, D. W. and S. Lemeshow. 2000. Applied logistic regression. John Wiley and Sons, New York, USA. King, J. A. 1955. Social behavior, social organization, and population dynamics in a black-tailed prairie dog town in the Black Hills of South Dakota. Contributions of the Laboratory of Vertebrate Biology, Number 67. University of Michigan, Ann Arbor, USA. Lantz, S. J., C. Conway, and S. H. Anderson. 2006. Multiscale habitat selection by Burrowing Owls in black-tailed prairie dog colonies. Journal of Wildlife 'Management 71:2664—2672. Lutz, R. S. and D. L. Plumpton. 1999. Philopatry and nest site reuse by Burrowing Owls: implications for productivity. Journal of Raptor Research 33:149-153. Martin, D. J. 1973. Selected aspects of Burrowing Owl ecology and behavior in central New Mexico. Condor 75:446-456. McNicoll, j. L. 2005. Burrowing Owl {Athene cunicii- laria) nest site selection in relation to prairie dog colony characteristics and surrounding land-use prac- tices in Janos, Chihuahua, Mexico. Thesis. New Mexico State University, Las Cruces, USA. Miller. B., G. Ceballos, and R. Reading. 1994. The prairie dog and biotic diversity. Conservation Biology 8:677-681. Millsap. B. A. and C. Bear. 2000. Density and reproduction of Burrowing Owls along an urban development gradient. Journal of Wildlife Manage- ment 64:33-41. Mrykalo, R. j., M. M. Grigione, and R. J. Sarno. 2009. A comparison of available prey and diet ol Florida BuiTOwing Owls in urban and rural environments: a first study. Condor 111: 556—559. Oakes. C. L. 2000. History and consequences of keystone mammal eradication in the desert grasslands: the Arizona black-tailed prairie dog (Cynontys luclovicia- niis arizonensis). Dissertation. Texas A&M University, College Station, USA. Orth, P. B. and P. L. Ki;nnedy. 2001. Do land-u,se patterns inlluence nest-site selection by Burrowing Owls {Athene ciinicniarici hypugueci) in northeastern Colorado? Canadian Journal of Zoology 79:1038- 1045. PouLtN, R. G., L. D. Todd. K. M. Dohms, R. M. Brigham, AND T. 1. Wellicome. 2005. Factors associated with nest- and roost-burrow selection by Burrowing Owls {Athene ciinicniaria] on the Canadian prairies. Cana- dian Journal of Zoology 83:1373-1380. Rosenberg, D. K. and K. L, Haley. 2004. The ecology of Burrowing Owls in the agroecosystem of the Imperial Valley, CA. Studies in Avian Biology 27:120-135. Ross, P. V. 1974. Ecology and behavior of a dense colony of Burrowing Owls in the Texas Panhandle. Thesis. West Texas State University, Canyon, USA. Sauer, J. R., J. E. Hines, and J. Fallon. 2007. The North American Breeding Bird Survey, Results and Analysis 1966-2006. USGS, Patuxent Wildlife Research Cen- ter, Laurel, Maryland, USA. Sheffield, S. R. 1997. Current status, distribution, and conservation of the Burrowing Owl {Speotyto cunicii- laria) in midwestem and western North America. Pages 399^07 in Biology and conservation of owls of the northern hemisphere: Second International Sym- posium (J. R. Duncan, D. H. Johnson, and T. H. Nicholls, Editors). USDA, Forest Service, General Technical Report NC-190. Northcentral Forest Exper- iment Station, St. Paul, Minnesota, USA. Systat. 2007. Systat for Windows: Systat Version 12.0. Systat Software Inc., San Jose, California, USA. Teaschner, a. 2005. Burrowing Owl nest site use and productivity on prairie dog colonies in the southern high plains of Texas. Thesis. Texas Tech University, Lubbock, USA. U.S. Census Bureau. 2000. Population, housing units, area, and density. http://factfinder.census.gov/Itome/ en/datanotes/expsfl u.htm (accessed 5 December 2002). Wellicome, T. 1. 2005. Hatching asynchrony in Burrowing Owls is inlluenced by clutch size and hatching success but not by food. Oecologia 142:326-334. WE.SEMANN, T. AND M. RowE. 1987. Factors influencing the distribution and abundance of Burrowing Owls in Cape Coral, Florida. Pages 129-137 in Proceedings of the national symposium on urban wildlife (L. W. Adams and D. L. Leedy, Editors). National Institute for Urban Wildlife, Columbia, Maryland. USA. WE.STERN Regional Climate Center, Desert Research Institute. 2002. Monthly climate summary, southern. NM 1959-2005. Reno. Nevada. USA. htlp://www. wrcc.dri.edu/cgi-bin/cliMAlN.pl7nmstat (accessed 19 September 2002). York, M. M., D. K. Rosenberg, and K. K. Sturn. 2002. Diet and food breadth of Burrowing Owls {Athene cnniciil(irici) in the Imperial Valley, California. Western North American Naturalist 62:280-287. The Wilson Journal of Ornithology 122( 1 );60-67, 2010 DIETARY TRENDS OE BARN OWLS IN AN AGRICULTURAL ECOSYSTEM IN NORTHERN UTAH CARL D. MARTI' ABSTRACT. — Bam Owl {Tyto alha) diets were studied for 15 years in Utah. Ninety-eight percent of 1 1 1,016 prey items were mammals, heavily dominated by voles (Microtus spp.). Food-niche breadth (FNB) was 3.33 for the entire sample and varied gradually but significantly among the 15 years and among seasons. Frequency of prey in the diet did not vary significantly from year to year or among seasons. Mean daily temperatures did not vary significantly among years but annual precipitation totals and days when deep snow covered the ground varied significantly among years. Irrigation for agriculture may have partially mitigated annual precipitation fluctuations. Flay, one of the most important crops on the study area, increased over the study period and other crops decreased slightly in the amount planted. Flectares of hay planted, hectares of corn planted, and hectares of barley planted were the variables that combined to best predict annual FNB. Received 9 February 2009. Accepted 31 July 2009. Long-term studies can reveal how variation in prey populations, climatic factors, and changes in the landscape might influence the diet of Barn Owls (Tyto alha). Love et al. (2000), for example, found significant changes in diets of Barn Owls in the United Kingdom that they attributed primarily to changing agricultural practices. Taylor (1994) examined Barn Owl diets over 12 continuous years in a Scottish agricultural area and reported fluctuations in prey consumed related to cyclic changes in small mammal populations. Barn Owls prey mainly on a wide variety of mammalian species over their cosmopolitan range (Taylor 1994, Bruce 1999, Marti et al. 2005); they eat mostly small, nocturnal mammals (Herrera and Jaksic 1980, Jaksic et al. 1982), only occasionally capturing birds at high frequencies (Brosset 1956, Corner 1978). Barn Owl diets have been studied most intensively in the United States and Europe (Campbell et al. 1987, Taylor 1994, Marti et al. 2005), but also in South .America (Jaksic et al. 1982, Bellocq 2000), Africa (Vernon 1972, Coodman 1986), Asia (Lenton 1984, Mahmood-Ul-Hassan et al. 2007), and Australia (Morton and Martin 1979, Palmer 2001). The broad geographic distribution and the large quantity of published dietary information make the Barn Owl a good candidate for understanding factors that affect diet variation. Collectively, the large number of published studies provide an overview of the diet of Barn Owls, including types of prey taken (Taylor 1994, Bruce 1999, Marti et al. 2005), size of prey (Dickman et al. 1991, llle 1991, Marti et al. 1993), habitats used (Begall 2005, Bond et al. 2004, ' Raptor Research Center. Boi.sc State University, Boi.se, 11) 83712, USA; e-mail: cmarti@q.eom Askew et al. 2007), effect of season on diet (Baudvin 1983, Campbell et al. 1987), and the influence of climate on diet (Avery 1999, Jaksic and Lazo 1999). However, most studies of Barn Owl diets have been short-term and at risk of being biased because data were often collected during a single season or year and, at times, during atypical conditions, e.g., abnormal weather or in areas with unusual prey populations (Hey- wood and Pavey 2002, Sahores and Trejo 2004, Altwegg et al. 2006). Short-term studies, thus, cannot provide insight into the dynamics of predation. 1 studied Barn Owls in a north-temperate agricultural area with the objectives to: ( I ) describe the diet over a 15-year period, (2) quantify the effects of weather variation on the diet, and (3) quantify the impact of vegetation changes on the diet. METHODS Stiuly Area. — I studied Barn Owls in an area of ~ 1,000 km' in Box Elder, Davis, and Weber counties, Utah. The area was in a long, naiTow, and essentially flat valley at an elevation of 1,300 m between the Wasatch Mountains and the Creat Salt Lake (4I°0I' N, 112° II' W). The area was formerly shrub-steppe desert, but the natural plant community has been supplanted by irrigated agriculture and urban development. Primary crops were hay, corn, wheat, barley, and some livestock pasture. Uncultivated land was limited to narrow strips along roads, water courses, borders of agricultural fields, and adja- cent to the Great Salt Lake. Annual precipitation averaged 35 cm, and mean temperatures for January and July were —3.5° C and 23.9° C, 60 A/15 cm) to interfere with prey detection and capture by Barn Owls also varied among years (F14J5 = 3.4, P = 0.0003). The frequency of voles in the diet of Barn Owls during winters having the greatest number of days with deep snow was less than in milder winters, but the relationship was not significant (F|,i4 = 0.16, r — 0.01, P = 0.69). Winters when the ground was covered longest with deep snow were followed by XJ CD 0 0 -C o 'c *6 o o LL Winter Season FIG. 2. Barn Owl food-niche breadth in northern Utah by season, years 1977-1991 combined. • TRENDS IN BARN OWL DIETS IN NOR THERN UTAH 63 Year FIG. 3. Number of days with deep snow (>15 cm) on ground (bars) in northern Utah and frequency of voles in the diet (line) of Barn Owls during spring. Dietary Variation. — Frequencies ol' different prey items consumed did not vary among years (5^2 = 64.3, df ^ 10, P = 0.67; Fig. 5), among seasons for all years combined ix~ = 6.4, dl =15, p = 0.97), or among seasons within years. The percentage of voles (both species combined) in the diet of Barn Owls varied by an average of only 2.9% from year to year (x" = 4.6, df =14, Z' = 0.99). Variation in Food-niche Breadth.— Tht model that best explained variation in FNB among years included the variables: hectares of hay planted, hectares of corn planted, and hectares of barley planted (Table 2). DISCUSSION a significant increase in the frequency of voles in spring Barn Owl diets (Ti 14 = 7.79, r = 0.32, P = 0.01; Fig. 3). Summer diets following winters with deep snow cover also contained a higher frequency of voles (C1J4 = 4.6, r = 0.26, P = 0.05). The same relationship continued into autumn but the trend was not significant (F,j4 = 1.5, /- = 0.1, P = 0.25). The number of hectares planted in different crops varied among years with the number of hectares of hay increasing during the study and the area devoted to other crops decreasing (Fig. 4). Crop diversity did not vary among years (6-11.35 = 1-0, r = 0.32, P = 0.45). Bam Owls exhibited little variation in diet over a 15-year period in northern Utah where habitat changes were minimal and variation in precipita- tion was buffered by crop irrigation, minimizing the effects of year-to-year variation of precipita- tion on vegetation. In contrast to my findings, Taylor (1994) found pronounced fluctuation in Barn Owl diets corresponding to vole population cycles during a long-term study in Scotland. The proportion of voles in the diet of Barn Owls in my study, unlike several studies in Europe (Bohnsack 1966, de Bruijn 1979, Taylor 1994), did not show cyclic fluctuations, possibly because irrigated agriculture produced a consistent, high-quality Area (ha) planted in crops in northem Utah (data for hay were not available for 1977-1979). FIG. 4. 64 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. I. March 2010 '■o o c CD CD c 0 o 0 CL 90 80 70 60 50 40 30 20 10 0 Sorex Mus Peromyscus Reithrodontomys ■■ Microtus i i £l i El M M rfiJ EJ ifl J1 1977 1979 1981 1983 1985 1987 1989 1991 Year FIG. 5. Frequency of major prey in Bam Owl diets in northern Utah by year. habitat that damped vole population fluctuations (Negus et al. 1986). Changes in Barn Owl diets sampled in two periods separated by a 19-year interval (1956-1974 and 1993-1997) in the United Kingdom were mostly attributable to a change from mixed-crop farms to more homog- enous farms (Love et al. 2000). The mode of farming did not change during my study, but changes in proportion of crops planted did and those changes were related to the small changes in TABLE 2. Models of variation in food-niche breadth among years using .seven explanatory variables: annual precipitation (AP), number of days with deep snow on ground (DOS), hectares of corn planted (HC), hectares of hay planted (FIH), hectares of barley planted (HB), and hectares of wheat planted (HW).“ Model A," »v' Adj. R- HH -b HC -b HB 0 0.962 0.99 HC -b DOS -b AP 6.65 0.035 0.99 HH -b HC -b HW 12.78 0.002 0.96 HH -b HB -b AP 14.1 1 0.001 0.95 HH -b HC -b AP 14.91 0.001 0.94 “ Akaikc’.s information crilcrion (AIC) for Ihe best model was -42.8,1. A, = difference between the AIC value of a given model and that of the best-fitting model. ‘ w, = AIC model weight. the diet — the more hay planted, the more voles in the diet. Voles are known to be abundant in forage crops like hay (Getz 1985), and Barn Owls in Utah may have concentrated their foraging in the vegetation that served as best vole habitat as they did in the United Kingdom (Askew et al. 2007). Voles were the most common prey in the diet of Barn Owls in my study, and have been the most numerous prey of these owls in many other studies conducted in the north-temperate zone (Campbell et al. 1987, Taylor 1994, Bruce 1999). In studies where an index of small mammal population size was available. Barn Owls captured voles in greater proportion than expected based on relative abundance (Marti 1974, Colvin 1984, Askew et al. 2007). The high proportion of voles in the diet of Barn Owls in my study also suggests that voles were taken .selectively. I found strong evidence of selective predation by Barn Owls on voles inhabiting an island in the Great Salt Lake 10 km from the present study area, probably due to the owls hunting preferentially in vole habitat (Marti 1986); Barn Owl diets on that island contained 81% voles by number despite vole habitat being extremely limited. Voles were not known to occur on Antelope Island prior to my finding them in the Barn Owl diet (Bowers 1982, A/r/r// • TRENDS IN BARN OWL DIETS IN NORTHERN UI'AII 65 Marti 1986). A similar pattern was seen in southwestern Idaho where frequency of voles was higher in Barn Owl diets at nests surrounded by iirigated agricultural land (Marti 1988). Barn Owls appeared to concentrate their foraging in habitats where voles were most numerous both on Antelope Island and in southwestern Idaho. Fast and Ambrose (1976) and Derting and Cranford (1989) found that captive Barn Owls selected Microtiis over Peromyscus when given a simultaneous choice. Prey size may be an important factor in prey selection — Microtiis are about twice the mass of Peromyscus and Mus, the next largest among the most commonly taken prey in my study. Captive Barn Owls, given a simultaneous choice between two sizes of the same species, selected larger prey up to the point where prey became hard to subdue (Ille 1991). It seems likely that Barn Owls may not discriminate between the two vole species be- cause both are found in the same habitats (Getz 1985), and are similar in appearance and mass. A sample of voles trapped near my study area (University of Utah Museum collection; H. J. Egoscue, unpubl. data) revealed no difference in mass between the species (t = 1.53, df = 63, F = 0.13). Despite these similarities, I separated the two voles in my analyses because I did not know whether Bam Owls can or do distinguish between them. Tores et al. (2005) concluded that Barn Owls prefer the kind of prey that voles exemplify, but readily switch to different prey if the prefemed species declines below a certain level. Barn Owls in Utah ate all of the small mammal species known to occur in the area (Newey 1951, Frost 1970). Some prey were available or vulnerable only at certain times of the year; seasonal variation in uncommon prey was most noticeable for northern pocket gophers (Thomomys tal- poides), which were numerous in the diet in May when young animals disperse above ground (Nowak 1991). Birds were a minor component of the diet, and the frequency of .some species in the Bam Owl diet could be attributed to migration. Sora (Porzana Carolina), for example, occurred in the diet only in spring and early summer. Others, like European Starlings (Sturnus vulgaris), were not migratory and Barn Owls preyed on them throughout the year. However, starlings reached their highest numbers in the Barn Owl diet in winter when communal roosting may have made them more vulnerable. 1 observed the same pattern ol seasonal variation in the diet ol Barn Owls in Utah that Taylor (1994) reported in Scotland — frequency of voles in the diet was lowest in spring, increased into summer, and peaked in late lall and early winter. During winters with more days of deep snow cover. Bam Owls in my study preyed on fewer voles. Winter fluctuations were likely explained by voles under deep snow cover having reduced vulnerability to predators (Canova 1989, Halonen et al. 2007). The number of Barn Owls breeding in the study area varied 12-fold over the study period (Marti 1994, 1997), primarily due to differences in severity of winter weather (Marti and Wagner 1985). However, owl population density did not influence the rate of predation on voles; owl density as indexed by the number of nest attempts per year was not correlated with the frequency of voles in the diet (ly = —0.25, P — 0.37). Diets of Barn Owls inhabiting this agricultural ecosystem exhibited remarkable uniformity over the long term. The measures I applied — FNB variation and prey frequencie.s — showed the diet of Barn Owls in my study exhibited little variation. The owls preyed primarily on voles at all collection sites and during all seasons and years, and the small changes that I found from year to year likely were the result of an increase in the amount of hay grown and a concomitant increase in vole numbers. The reverse — changing from grass-oriented agriculture to corn and soybeans — has been associated with declines of Barn Owl populations in Ohio (Colvin 1985). Variation in annual precipitation in my study area was mitigated by crop irrigation providing consistent moisture levels each year, and, in turn, consistent prey habitat. The only weather variable strongly associated with diet variation ot Barn Owls in Utah was the number of days the ground was covered with deep snow which interfered with owls ability to prey on voles. ACKNOWLEDGMENTS I ihank Weber State University for financial support through Research and Professional Growth grants, and equipment to conduct much of this research. P. W. Wagner helped with data collection during the first 6 years of this study. 1 thank B. C. Harvey for advice on using Akaike's information criterion. I also thank Karen Steenhof, M. N. Kochert, Motti Charter, Marco Restani, and an anonymous reviewer for their comments which helped refine this paper. 66 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. /, March 2010 LITERATURE CITED Altwegg, R., a. Roulin, M. Kestenholz, and L. Jenni. 2006. Demographic effects of extreme winter weather in the Barn Owl. Oecologia I49;44-51. Askew, N. P., J. B. Searle, and N. P. Moore. 2007. Agri- environment schemes and foraging of Barn Owls Tyto alha. Agriculture, Ecosystems and Environment 1 18:109-1 14. Avery, D. M. 1999. A preliminary assessment of the relationship between trophic variability in southern African Barn Owls Tyto alha and climate. Ostrich 70:179-186. Baudvin, H. 1983. La regime alimentaire de la Chouette Effraie (Tyto alha). Le Jean-le-Blanc 22:1-108. Begall, S. 2005. The relationship of foraging habitat to the diet ot Barn Owls (Tyto alha) from central Chile. Journal of Raptor Re.search 39:97-101. Bellocq, M. I. 2000. A review of the trophic ecology of the Barn Owl in Argentina. Journal of Raptor Research 34:108-119. Bohnsack. P. 1966. Uber die Ernahrung der Schleiereule, Tyto alha, insbesondore auBerhalb der Brutzeit, in einem wesholsteinischen Massenwechselgebiet der Feldmaus, Microtiis arvalis. Corax 1:162-172. Bond, G., N. G. Burnside, D. J. Metcalfe, D. M. Scott, and j. Blamire. 2004. The effects of land-use and landscape structure on Barn Owl (Tyto alha) breeding success in southern England, U.K. Landscape Ecology 20:555-566. Bowers, M. A. 1982. Insular biogeography of mammals in the Great Salt Lake. Great Basin Naturalist 42:589- 596. Brosset, a. 1956. Le regime alimentaire de FEffraye Tyto alha au Maroc oriental. Alauda 24:303-305. Bruce, M. D. 1999. Family Tytonidae (Barn-owls). Pages 34-75 in Handbook of the birds of the world. Volume 5. Barn-owls to hummingbirds (J. del Hoyo. A. J. Elliott, and J. Sargatal, Editors). Lynx Edicions, Barcelona. Spain. Burnham, K. P. and D. R. Anderson. 2002. Model selection and multimodel inference: a practical information-theoretic approach. Second Edition. Springer-Verlag, New York, USA. Campbell, R. W., D. A. Manuwal. and A. S. Harestad. 1987. Food habits of the Common Barn-Owl in British Columbia. Canadian Journal of Zoology 65:578-586. Canova, L. 1989. Influence of snow cover on prey selection by Long-eared Owls (A.sio otits). Ethology, Ecology and Evolution 1 :367-372. Colvin, B. A. 1984. Barn Owl foraging behavior and .secondary poisoning hazard from rodenticide use on farms. Dissertation. Bowling Green State University, Bowling Green. Ohio, USA. Coi.viN. B. A. 1985. Common Barn-Owl population decline in Ohio and the relationship to agricultural trends. Journal of Field Ornithology 56:224—235. DE Bruijn. O. 1979. Voedseloccologie van ide Kerkuil Tyto alha in Nederland. Limosa 52:91-154. Derting, R. L. and j. a. Cranford. 1989. Physical and behavioral correlates of prey vulnerability to Barn Owl (Tyto alha) predation. American Midland Naturalist 121:11-20. Dickman, C. R., M. Predavec, and a. J. Lynam. 1991. Differential predation of size and sex classes of mice by the Barn Owl, Tyto alha. Oikos 62:67-76. Fast, S. J. and H. W. Ambrose III. 1976. Prey preference and hunting habitat selection in the Barn Owl. American Midland Naturalist 96:503-507. Frost, J. L. 1970. The mammals of Davis County, Utah. Thesis. University of Utah, Salt Lake City, USA. Getz, L. L. 1985. Habitats. Pages 286-309 in Biology of New World Microtus. (R. H. Tamarin, Editor). American Society of Mammalogists Special Publica- tion Number 12. Allen Press Inc., Lawrence. Kansas, USA. Goodman, S. M. 1986. The prey of Barn Owls (Tyto alha) inhabiting the ancient temple complex of Karnak, Egypt. Ostrich 57:109-1 12. Gorner, M. 1978. Schleiereule, Tyto alha, als Vogeljiiger. Beitraege zur Vogelkunde 24:273-275. Halonen, M., T. Mappes, T. Meri, and J. Suhonen. 2007. Influence of snow cover on food hoarding in Pygmy Owls Glaiicidiiini passehnum. Ornis Fennica 84:105- 111. Herrer.x, C. M. and F. M. Jaksic. 1980. Feeding ecology of the Barn Owl in central Chile and southern Spain: a comparative study. Auk 97:760-767. Heywood, M. R. and C. R. Pavey. 2002. Relative importance of plague rodents and dasyurids as prey of Barn Owls in central Australia. Wildlife Research 29:203-207. Ille, R. 1991. Preference of prey size and profitability in Barn Owls Tyto alha guttata. Behaviour 1 16:180-189. Jaksic, F. M. and I. Lazo. 1999. Response of a bird assemblage in semiarid Chile to the 1997-1998 El Nino. Wilson Bulletin 1 1 1:527-535. Jaksic, F. M., R. L Seib, and C. M. Herrera. 1982. Predation by the Barn Owl (Tyto alha) in Mediterra- nean habitats of Chile, Spain and California: a comparative approach. American Midland Naturalist 107:151-162. Lenton, G. M. 1984. The feeding and breeding ecology of Barn Owls Tyto alha in peninsular Malaysia. Ibis 126: 551-575. Levins, R. 1968. Evolution in changing environments: some theoretical explorations. Princeton University Press, Princeton. New Jersey, USA. Love. R. A.. C. Webbon, D. E. Glue, and S. Harris. 2000. Changes in the food of British Barn Owls (Tvto alha) between 1974 and 1997. Mammal Review 30:107-129. Mahmood-Ul-Hassan, M., M. A. Beg. and M. Mushtaq- Ul-Hassan. 2007. Locality related changes in the diet of the Barn Owl (Tyto alha stertens) in agroecosystems in central Punjab, Pakistan. Wilson Journal of Ornithology 119:479^83. Marii. C. D. 1974. Feeding ecology of four sympatric owls. Condor 76:45-61. Marti, C. D. 1986. Barn Owl diet includes mammal species new to the island fauna of the Great Salt Lake. Great Basin Naturalist 46:307-309. A/(//7/ • I RENDS IN BARN OWL DIETS IN NORTHERN UTAH 67 Marti, C. D. 1‘')88. A long-ierm study of food-niche dynamics in the Common Barn-Owl: comparisons within and between populations. Canadian Journal of Zoology 66:1803-1812. Marti, C. D. 1994. Barn Owl reproduction: patterns and variation near the limit of the species' distribution. Condor 96:468—484. Marti, C. D. 1997. A 20-year study of Barn Owl {Tylo alba) reproduction in northern Utah. Page 261 in Biology and conservation of owls of the Northern Hemisphere. (J. R. Duncan, D. H. Johnson, and T. H. Nicholls, Editors). USDA, Forest Service, General Technical Report NC-190. Northcentral Research Station, St. Paul, Minnesota, USA. Marti, C. D. and P. W. Wagner. 1985. Winter mortality in Common Barn-Owls and its effect on population density and reproduction. Condor 87:1 1 1-1 15. Marti, C. D., M. Bechard, and F. M. Jaksic. 2007. Food habits. Pages 129-149 in Raptor research and management techniques (K. L. Bildstein and D. M. Bird, Editors). Second Edition. Raptor Research Foundation, Hancock House Publishers. Blaine, Washington. USA. Marti, C. D., A. F. Poole, and L. R. Bevier. 2005. Barn Owl {Tylo alba). The birds of North America online (A. Poole, Editor). bna.birds.cornell.edu/BNA/ account/Bam Owl (accessed 1 1 December 2008). Marti. C. D., P. W. Wagner, and K. W. Denne. 1979. Nest boxes for the management of Barn Owls. Wildlife Society Bulletin 7:145-148. Marti, C, D., K. Steenhof. M. N. Kochert, and J. S. Marks. 1993. Community trophic structure; the roles of diet, body size, and activity time in vertebrate predators. Oikos 67:6-18. Morton, S. R. and A. A. Martin. 1979. Feeding ecology of the Barn Owl, Tylo alba, in arid southern Australia. Australian Wildlife Research 6:191-204. Negus, N. C., P. J. Berger, and B. W. Brown. 1986. Microtine population dynamics in a predictable environment. Canadian Journal of Zoology 64:785— 792. Newey, j. P. 1951. Mammals of Weber County, Utah. Thesis. University of Utah. Salt Lake City, USA. Nowak, R. M. 1991. Walker’s mammals of the world. Johns Hopkins Press, Baltimore, Maryland, USA. Pai.mer, R. a. 2001 . Notes on the diet of the Barn Owl Tylo alba from Mulyungarie Station in north-eastern South Australia. South Australian Ornithologist 33:137-138. Sahores, M. and a. Trejo. 2004. Diet shift of Barn Owls (Tylo alba) after natural fires in Patagonia, Argentina. Journal of Raptor Research 38:174-177. SAS Institute. 2000. SAS/STAT software: changes and enhancements. Release 8.1. SAS Institute Inc., Cary, North Carolina, USA. Taylor, I. 1994. Barn Owls: predator-prey relationships and conservation. Cambridge University Press. Cam- bridge, United Kingdom. Tores, M.. Y. Motro, U. Motro, and Y. Yom-Tov. 2005. The Barn Owl — a selective opportunist predator. Israel Journal of Zoology 51:349-360. Vernon, C. J. 1972. An analysis of owl pellets collected in southern Africa. Ostrich 43:109-124. The Wilson Journal of Ornithology 122( l);68-74, 2010 REGIONAL VARIATION IN DIETS OF BREEDING RED-SHOULDERED HAWKS BRAD N. STROBEL'-^ AND CLINT W. BOAL^ ABSTRACT. — We collected data on breeding season diet composition of Red-shouldered Hawks (Biiteo lineatus) in south Texas and compared these data, and those reported from studies elsewhere to examine large scale spatial variation in prey use in eastern North America. Red-shouldered Hawk diets aligned into two significantly different groups, which appear to correlate with latitude. The diets of Red-shouldered Hawks in group 1, which are of more northern latitudes, had significantly more mammalian prey and significantly less amphibian prey than those in group 2, which are at more southerly latitudes. Our meta-analysis demonstrated the dietary flexibility of Red-.shouldered Hawks, which likely accounts tor their broad distribution by exploiting regional variations in taxon-specific prey availability. Received 27 April 2009. Accepted 28 August 2009. Red-shouldered Hawks (Buteo lineatus\ hereaf- ter RSHA) were once considered the most abundant raptor in moist woodlands of North America (Bent 1937), and are thought to prefer mature contiguous forests, especially those asso- ciated with wetlands (Craighead and Craighead 1956, Bednarz and Dinsmore 1981, Dykstra et al. 2008). Breeding Bird Survey data indicate RSHA populations have declined in portions of the midwestern and northeastern United States, while populations in other areas appear to be relatively stable or increasing (Sauer et al. 2008). Population declines have heightened concern and protection for RSHA by several state agencies (Jacobs and Jacobs 2002). Conversion of forests to agriculture is one factor thought to have contributed to RSHA population declines (Bednarz and Dinsmore 1981, Gehring 2003, Dykstra et al. 2008). Greater amounts of forest edge and smaller forest stand size caused by agricultural and silvicultural practices may provide a competitive advantage for habitat generalist raptors such as Red-tailed Hawks (Buteo jamaecensis) (Bednarz and Dins- more 1981). Wetland degradation and drainage may also have reduced the quality and availability of foraging areas important to RSHAs (Jacobs and Jacobs 2002). Other studies have documented apparently sustainable RSHA populations in highly human altered landscapes (e.g.. Bloom et ' Texa.s Cooperative Fish and Wildlife Researeh Unit, Department of Natural Re.sources Management, Texas Tech University, laibbock, TX 79409, USA. - U.S. Geological Survey, Texas Cooperative Fish and Wildlife Research Unit, Department of Natural Resources Management. Texas Tech University, Lubbock, TX 79409. USA. 'Corresponding author; e-mail: bradley. .strobel@ttu.edu al. 1993, Rottenborn 2000, Dykstra et al. 2001), suggesting that factors other than habitat alter- ation may be directly or indirectly contributing to RSHA population trends. For example. Bloom (1989) suggested that habitat use by RSHAs is strongly influenced by prey availability. The abundance and diversity of prey available to breeding RSHAs is likely associated with, and influenced by, landscape characteristics within RSHA foraging habitat. Previous studies of breeding season diets indicate RSHAs use a variety of prey items across their range. Diets documented for RSHAs nesting in Missouri were primarily amphibians and reptiles (Parker 1986), while nearly 75% of the diets of RSHAs nesting in Maryland were mammals (Portnoy and Dodge 1979). The diets of breeding RSHAs in Arkansas (Townsend 2006) and Texas (Strobel 2007), in contrast, were largely invertebrates. Bednarz and Dinsmore (1985) documented RSHA dietary differences between two breeding seasons in Iowa, apparently associated with annual differences in precipitation and riparian flooding. Thus, it appears RSHAs adapt to some natural and anthropogenic land- scape changes, likely facilitated by their ability to exploit a diverse suite of prey. Regional variation in prey use by RSHAs is likely caused by large-scale ecological differences such as precipitation patterns, temperature varia- tion, and composition of vegetation communities influencing prey communities. Anthropogenic changes to the landscape have also influenced population trends of RSHAs in North America (Brown 1971, Henny et al. 1973, Peterson and Crocoll 1992, Jacobs and Jacobs 2002). Under- standing the coarse scale patterns in prey u.se, and the breadth and plasticity of RSHA diets may 68 Sirohel am! Boa! • RED-SIIOULDERED HAWK DIET 69 offer insights into productivity and ecology of RSHA populations that would facilitate develop- ment of sound conservation strategies for the species. We used breeding season food habits information from our study in south Texas and published studies conducted on breeding RSHA populations east of the Continental Divide to examine large scale spatial variation in prey use by eastern populations of Red-shouldered Hawks in North America. METHODS Study Area and Field Procedures. — 'We located active Red-shouldered Hawk nests on the Rob and Bessie Welder Wildlife Refuge and Twin Oaks Hunting Resort in San Patricio and Refugio counties (respectively), Texas. Nests of breeding RSHA pairs were located using broadcast survey methods as described by McLeod and Andersen (1998) and intensive systematic nest searching between April and July 2005 and 2006. We installed color video surveillance cameras (Mod- els OC-225 and C3320AX, Clover Electronics®, Cerritos, CA, USA) after eggs hatched and young were ~1 week of age. We recorded all diurnal activities using time-lapse VCRs (Piczel video security products® and Security Labs®, Nobles- ville, IN, USA) until each nesting attempt was completed. We based our analysis on prey items delivered to nestlings prior to fledging to reduce bias caused by off camera prey deliveries. We identified prey items to the most specific taxa possible using regional field guides, museum specimens, and Welder Wildlife Refuge staff expertise (Strobel 2007). Literature Search. — 'We conducted a literature search for published studies (peer reviewed publications, books, and theses) of RSHA breed- ing season diet. Methods used in other studies included: direct visual observation, video surveil- lance, identification of uneaten prey remains, composition of regurgitated pellets, and crop contents analysis. We classified methods of the publications as either direct observation (DO) or prey remains analysis (PR). Inconsistent methods of data collection are a common concern in meta-analysis and the poten- tial biases of different techniques to assess diets of raptors have been noted previously (Redpath et al. 2001, Rogers et al. 2005, Marti et al. 2007). Results from studies using prey remains analysis can be skewed by higher persistence rates of heavy boned prey species, such as mammals, and falsely indicate higher frequency of occurrence of those species (Marti et al. 2007). Correct identi- fication of prey items in direct observation studies often depends on prey item size and its distance from the observer (Marti et al. 2007). We attempted to reduce these potential confounding effects by pooling prey types into five taxonomic categories: mammals, birds, reptiles, amphibians, and invertebrates. When studies used multiple methods (i.e., DO and PR) to collect data and presented those data separately, we only used those collected through direct observation and excluded unknown or unidentified prey items (Table 1). Some studies presented yearly data separately and, to remain consistent in our analyses, we pooled all data from multiple years within those studies. Statistical Analyses. — ^We calculated the Pear- son’s correlation coefficient, based on the pro- portion of prey types reported in each study to compare RSHA diets among studies (Krebs 1999). Values of Pearson’s correlation coefficient range from 0 to 1 indicating no dietary overlap (0) to complete dietary overlap (1). We used the Pearson’s correlation coefficient as the distance function within an unweighted pair-group method using arithmetic means (UPGMA) cluster analysis to calculate and compare the similarity of diets reported for breeding RSHAs. We assessed the significance of clusters using the approximately unbiased significance level {P ~ \-au) calculated from 10,000 multiscale bootstrap iterations of the distance matrix using the PVCLUST package (Suzuki and Shimodaira 2006). We used Ho- telling’s T- test, in the ICSNP package, on the arcsine transformed proportions of each prey category to identify which prey categories con- tributed to the formation of significant clusters. Reported values represent means ± SD. Analyses were conducted using program R version 2.8.0 (R Development Core Team 2008). RESULTS We monitored prey use at 10 different Red- shouldered Hawk nests in south Texas in 2005 and 2006, and collected 1,457 hrs of video footage yielding 1,315 prey items categorized into the five taxonomic categories (Table 1). We located 1 1 other studies that reported the diets of breeding RSHAs in different states in the eastern United States (Table 1). Eew multi-year studies identified prey delivered to individual nests or by individual RSHAs. Thus, the independence of 70 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. I, March 2010 TABLE 1. Composition of diets ot breeding Red-shouldered Hawks as reported in 12 studies across the eastern United States. Methods of data collection included direct observation (DO) and pellet and/or prey remains examination (PR). Shading indicates significantly similar groups identified though cluster analyses. Adapted from Dykstra et al. (2008). Mammal Bird Reptile Amphib. Invert. State Year # of nests # Prey items # % # % # % # % # % Method Source lA 1977-1978 7 115 41 36 3 3 3 3 38 33 30 26 DO Bednarz and Dinsmore 1985 IN 1999-2(X)0 8 195 82 42 17 9 48 25 8 4 40 21 PR Gehring 2003 MA 1974 4 46 33 72 2 4 2 4 9 20 0 0 DO Portnoy and Dodge 1979 MD 1978-1979 17 30 23 77 1 3 2 7 4 13 0 0 DO Janik and Moser 1982 MI 1942/1948 29 1,428 583 41 228 16 311 22 68 5 238 17 PR Craighead and Craighead 1956 NY 1939-1942 16 — — 58 — 8 — 3 — 18 — 11 DO/PR Ernst 1945“'’ OH 1997-2001 21 198 64 32 14 7 46 23 36 18 38 19 DO Dykstra et al. 2003 WI 1973-1979 9 97 57 59 1 1 14 14 12 12 13 13 DO Welch 1987 AR 2004-2005 20 1,139 69 6 4 0 171 15 523 46 372 33 DO Townsend 2006 GA 1994 8 181 34 19 15 8 43 24 46 25 43 24 DO Howell and Chapman 1998 MO 1982-1983 3 — — 17 — 0 — 26 — 46 — 1 1 DO Parker 1986“ TX 2005-2006 10 1315 197 15 63 5 412 31 107 8 536 41 DO Strobel 2007 ^ Only reported the relative proportion.s of prey items to total diet. Invertebrates approximate because of small unknown proportion (<3%) of crayfi.sh categorized as miscellaneous. data within each study is unknown; however, we assumed that natural and stochastic processes (pair and territory fidelity, nest accessibility, etc.) were similar among study populations, which facilitated comparison among studies. These data, in combination with our data, describe the diets of three {B. 1. alleni, B. I. lineatus, B. 1. texanus) of the five currently recognized subspecies of RSHA. The cluster analysis of the Pearson’s correlation coefficient for the 12 study areas identified several significant clusters {P < 0.05) with a primary division creating two distinct groups (Fig. 1 ). Our results indicate significant diver- gences at several levels of correlation; however, we were most interested in the dichotomous nature of the results and focus further analyses on the two most distinct groups. We arbitrarily titled the groups. Group 1 was comprised of data from studies in Indiana, Iowa. Maryland, Massa- chu.setts, Michigan. New York, Ohio, and Wis- consin. Group 2 was comprised of data collected in Arkansas, Georgia, Missouri, and Texas. Although grouped by similarity in breeding .season diet, the latitude of each study site appears to correlate with group membership with studies in group 1 occurring in states of higher latitudes than studies in group 2. Analyses of those data from the northern and southern groups indicated substantial differences existed between diets in the two primary clusters (^5,7 = 8.1, P < 0.01). Red-shouldered Hawk diets in group 1 had more mammalian prey than those in group 2 (group 1. .v ± SD = 53.2 ± 20.3%, group 2 = 17.9 ± 9.6%; t\\ = 3.8, P < 0.01, Fig. 2). Diets of RSHAs in group 2 had a higher proportion of amphibians (group 1 = 1 1.2 ± 6.2%, group 2 = 32.9 ± 20.3%; tu = -2.7, P = 0.02) than those in group 1 (Fig. 2). Avian, reptilian, and invertebrate prey items comprise similar proportions of RSHA diets in both groups. All studies using prey remains analysis were assigned to Group 1, suggesting the suspected biases previously discussed may have skewed their results; however, the latitudinal pattern still exists in direct observational studies (Table 1). DISCUSSION Our analysis indicated there are two groups of Red-shouldered Hawks distinguishable by their breeding sea.son diet. Breeding pairs in group 1 preyed more frequently on mammals while those in group 2 preyed more frequently on amphibians. In addition to diet similarities, there also appears to be a coarse spatial relationship within groups. Most studies conducted at northern latitudes Strobe! and Baal - RED-SHOULDERED HAWK DIET 71 Study site/year FIG. 1. Dendrogram illustrating similarity of diets of Red-shouldered Elawks reported by studies in 12 states across the eastern USA. ait values indicate approximately unbiased significance level of each specified grouping {P ~ \-aii) based on 10,000 multiscale bootstrap iterations of the Pearson’s correlation coefficient topology matrix of an UPGMA cluster analysis. Dashed boxes identify the primary significant (P ^ 0.01) groups. (>38° N) were classified into group 1 and most southern latitude studies into group 2. We speculate the pattern between latitude and diet composition is ultimately caused by climate, vegetation, and precipitation patterns influencing prey type availability differently across North America. Results from Bednarz and Dinsmore’s (1985) study in Iowa and Strobel’s (2007) in Texas, support this hypothesis and show diets of RSHAs were primarily mammalian prey during drier conditions, but were primarily amphibian and invertebrate prey under conditions with abundant moisture. This may also explain why Bednarz and Dinsmore’s (1985) and Strobel s (2007) data were the furthest diverged from other studies in their respective groups (Fig. 1). The effects of latitude, temperature, and moisture regimes on species’ distributions are largely accepted (MacArthur 1972), and clearly influence prey availability. These effects would presumably influence productivity and survival of RSHAs breeding in different areas of North America. For example, nutritional value varies widely among prey types and likely influences reproductive rates of RSHAs breeding in different regions of their range. Four studies in our analysis reported estimates of prey biomass and estimated mammalian prey items taken by RSHAs averaged 30% heavier than amphibian prey items (Howell and Chapman 1998, Dykstra et al. 2003, Town- send 2006, Strobel 2007). Strobel (2007) estimat- ed that RSHAs breeding in south Texas deliver -6,000 g of prey to a nest to fledge one nestling. Common Kestrels {Falco tinniinculns, Wiehn and Korpimaki 1997), European Starlings (Stiinnis yiilgciris, Wright et al. 1998), and RSHAs (Strobel 2007) increased prey delivery rates to meet the increased nutritional demands of larger broods or older nestlings. Red-shouldered Hawks breeding in areas with larger and more nutrient rich prey may be able to rear larger broods, fledge higher quality young, or allocate more time to activities 72 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. I. March 2010 FIG. 2. Composition of Red-shouldered Flawk diets among studies in two significantly different groups (Group I = northern studies; Group 2 = .southern studies) identified by 10,000 multiscale bootstrap iterations of the Pearson’s correlation coefficient. Spreads indicate ± 1 SD from the mean percent contributed by each of the five prey types. such as nestling defense, and increase their productivity. A post hoc examination of produc- tivity rates (average number of nestlings produced by successful nests) indicated little difference in productivity of RSHAs in group 1 or group 2 (2.42 ± 0.51, 2.38 ± 0.67; respectively). Bednarz and Dinsmore (1985) noted similar RSHA productivity on their study site between years when RSHAs had markedly different diets. This suggests RSHAs have similar productivity despite using different prey types. This also suggests different prey types result in differences in foraging effort, prey delivery rates, and energetic demands upon adults. Recognizing patterns in prey use by RSHAs throughout North America may provide valuable insight for conservation planning. Breeding pop- ulations of RSHAs feeding primarily on mamma- lian prey items may not be as greatly inHuenced by wetland degradation as populations feeding primarily on amphibians, but may be more susceptible to fluctuations in forest harvest regimes or natural fluctuations in small mammal populations. It is unclear how factors like bioaccumulation of toxins and juvenile survival rates vary between populations with different breeding season diets. Few studies have examined regional specific survival patterns or lifetime reproductive success in widespread raptor popu- lations, both of which are likely influenced by diet composition and prey availability. Evaluating the importance of different types of prey for RSHAs throughout North America would be greatly enhanced if future studies provided accurate estimates of local prey biomass and availability. ACKNOWLEDGMENTS We thank the Rob and Bes.sie Welder Wildlife Foundation for indispensable financial and logistic support as well as the U.S. Geological Survey, Texas Cooperative Fish and Wildlife Research Unit and Texas Tech Univer- sity. Additional funding was provided by the Houston Safari Club. We appreciate Terry Blankenship’s assistance with prey identification and Carey Haralson for field and technical assistance. We also thank Jeff Rooke and the staff Sirohcl (iihI Boa/ - RED-SI lOUlJ^ERED HAWK DIET 73 of the Twin Oaks Hunting Resort. The use of trade, product, industry, or firm names or products is for informative purposes only and does not constitute an endorsement by the U.S. Government, U.S. Geological Survey, or Texas Tech University. This is Welder Wildlife Foundation Contribution Number 688. LITERATURE CITED Bednarz, J. C. and J. J. Dinsmore. 1981. Status, habitat use, and management of Red-shouldered Hawks in Iowa. Journal of Wildlife Management 45:236-241. Bednarz, J. C. and J. J. Dinsmore. 1985. Flexible dietary response and feeding ecology of the Red-shouldered Hawk, Buteo lineatiis, in Iowa. Canadian Field- Naturalist 99:262-264. Bent, A. C. 1937. Life histories of American birds of prey. U.S. National Museum Bulletin 167. Washington, D.C., USA. Bloom. P. H. 1989. Red-shouldered Hawk home range and habitat use in southern California. Thesis. California State University, Long Beach, USA. Bloom, P. H., M. D. McCrary, and M. J. Gibson. 1993. Red-shouldered Hawk home-range and habitat use in southem California. Journal of Wildlife Management 57:258-265. Brown. W. H. 1971. Winter population trends in the Red- shouldered Hawk. American Birds 25:813-817. Craighead, J. J. and F. C. Craighead Jr. 1956. Hawks, owls and wildlife. Stackpole Company, Harrisburg, Pennsylvania and Wildlife Management Institute, Washington, D.C., USA. Dykstra, C. R.. j. L. Hays, and S. T. Crocoll. 2008. Red-shouldered Hawk {Buleo lineatiis). The birds of North America Online (A. Poole, Editor). Cornell Laboratory of Omithology, Ithaca, New York, USA. http://bna.birds.cornell.edu/bna/species/l07doi: 10. 2173/bna. 107 (accessed 16 December 2007). Dykstra, C. R., J. L. Hays, F. B. Daniel, and M. M. Simon. 2001. Home range and habitat use of suburban Red-shouldered Hawks in southwestern Ohio. Wilson Bulletin 113:308-316. Dykstra, C. R., J. L. Hays, M. M. Simon, and F. B. Daniel. 2003. Behavior and prey of nesting Red- shouldered Hawks in southwestern Ohio. Journal of Raptor Research 37:177-187. Ernst, S. G. 1945. The food of the Red-shouldered Hawk in New York State. Auk 62:452-453. Gehring, j. L. 2003. The ecology of Red-tailed Hawks and Red-shouldered Hawks in forested landscapes and in landscapes fragmented by agriculture. Dissertation. Purdue University, West Lalayette, Indiana, USA. Henny, C. j., F. C. Schmidt, E. M. Martin, and L. L. Hood. 1973. Territorial behavior, pesticides, and the population ecology of Red-shouldered Hawks in central Maryland. 1943-1971. Ecology 54:545-554. Howell, D. L. and B. R. Chapman. 1998. Prey brought to Red-shouldered Hawk nests in the Georgia piedmont. Journal of Raptor Research 32:257-260. Jacobs, J. and E. J. Jacobs. 2002. Con.servation assess- ment for Red-shouldered Hawk (Buteo hneatus) in national forests of the North Central .States. U.SDA, Forest Service, Eastern Region, Milwaukee, Wiscon- sin, USA. Janik. C. a. and j. a. Mosher. 1982. Breeding biology of raptors in the central Appalachians. Journal ol Raptor Re.search 1 6: 1 8-24. Krebs, C. J. 1999. Ecological methodology. Second Edition. Addison-Wesley Educational Publishers Inc., New York, USA. MacArthur, R. H. 1972. Geographical ecology: patterns in the distribution of species. Princeton University Press, Princeton, New Jersey, USA. Marti, C. D., M. Bechard, and F. M. Jasik. 2007. Food habits. Pages 129-149 in Raptor research and management techniques (D. L. Bird and K. L. Bildstein. Editors). Hancock House Publishers, Blaine, Washington, USA. McLeod, M. A. and D. E. Andersen. 1998. Red- shouldered Hawk broadcast surveys: factors affecting detection of responses and population trends. Journal of Wildlife Management 62:1385-1397. Parker, M. A. 1986. The foraging behavior and habitat use of breeding Red-shouldered Hawks Buteo Hneatus in southeastern Missouri. Thesis. University of Missouri, Columbia, USA. Peterson, J. M. C. and S. T. Crocoll. 1992. Red- shouldered Hawk, Buteo Hneatus. Pages 333-351 in Migratory nongame birds of management concern in the Northeast (K. J. Schneider and D. M. Pence, Editors). USDl, Fish and Wildlife Service, Newton Corner, Massachusetts, USA. Portnoy, J. W. and W. E. Dodge. 1979. Red-shouldered Hawk nesting ecology and behavior. Wilson Bulletin 91:104-1 17. R Development Core Team. 2008. R: a language and environment for statistical computing. R Foundation for Statistical Computing. Vienna, Austria. Redpath, S. M., R. Clarke. M. Madders, and S. J. Thirgood. 2001. Assessing raptor diet: comparing pellets, prey remains, and observational data at Hen Harrier nests. Condor 103:184-188. Rogers, A. S., S. Destefano, and M. F. Ingraldi. 2005. Quantifying Northern Goshawk diets using remote cameras and observations from blinds. Journal of Raptor Research 39:303-309. Rottenborn, S. C. 2000. Nest-site selection and repro- ductive success of urban Red-shouldered Hawks in central California. Journal of Raptor Research 34:18- 25. Sauer. J. R., J. E. Hines, and J. Fallon. 2008. The North American Breeding Bird Survey, results and analysis 1966—2007. Version 5.15.2008. USGS. Pa- tuxent Wildlife Research Center. Laurel. Maryland. USA. .Strobel. B. N. 2007. Nest site selection and nestling diet of the Texas Red-shouldered Hawk Buteo Hneatus te.xanus in south Texas. Thesis. Texas Tech University. Lubbock. USA. Suzuki, R. and H. Shimodaira. 2006. PVCLUST: hierarchical clustering with P-values via multiscale bootstrap resampling. R package Version 1.2-0. http:// 74 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. I. March 2010 www.is.titech.ac.jp/~shimo/prog/pvclust/ (accessed 16 December 2008). Townsend, K. A. L. 2006. Nesting ecology and sibling behavior of Red-shouldered Hawks at the St. Francis Sunken Lands Wildlife Management Area in north- eastern Arkansas. Thesis. Arkansas State University, State University, USA. WiEHN, J. AND E. Korpimaki. 1997. Food limitation on brood size: experimental evidence in Eurasian Kestrel. Ecology 78:2043-2050. Welch, R. J. 1987. Food habits of the Red-shouldered Hawk in Wisconsin. Pas.senger Pigeon 49:81-91. Wright, J., C. Both, P. A. Cotton, and D. Bryant. 1998. Quality vs. quantity: energetic and nutritional trade- offs in parental provisioning strategies. Journal of Animal Ecology 67:620-634. The Wilson Joiinnil oj Oniitholoi’y 1 22( 1 ):75 -81 , 2010 EFFECTS OF BROWN-HEADED COWBIRD PARASITISM ON PROVISIONING RATES OF SWAINSON’S WARBLERS SARA PAPPAS,' THOMAS J. BENSON,' -^ AND JAMES C. BEDNARZ' ABSTRACT. — We studied the effects of brood parasitism by Brown-headed Cowbirds (Mololhrus ater) and other factors on food provisioning rates of Swainson’s Warblers (Lunnothlypis swainsoiiii), a secretive and poorly understood species of conservation concern. We used time-lapse video systems to collect provisioning data at 25 nests, nine of which were parasitized by Brown-headed Cowbirds. We found strong relationships between feeding rate and brood size with increases from 2.0 feeding visits/hr for nests with a single Swainson’s Warbler nestling to 3.0/hr for broods of four. We also found an effect of nestling age with 1.9 visits/hr early and 3.2 visits/hr late in the nestling period. The relationships between cowbird parasitism and provisioning were complex. Nests with brood size of one or three that contained a single cowbird nestling had greater provisioning rates (2.5 and 3.3 visits/hr, respectively) than non-parasitized nests (2.0 and 2.6 visits/hr, respectively). Nests with two cowbirds and no Swainson's Warbler young had greater provisioning rates (3.4 visits/hr) than those with two warbler young (2.4 visits/hr), or one warbler and one cowbird (2.6 visits/hr). The increase m provisioning rate with nestling age was more pronounced with two cowbird nestlings present than in nests with zero or one cowbird nestling. These results suggest that parasitized nests, especially those with multiple cowbird nestlings, can impose greatei energetic demands on parents. Food limitation may constrain the ability of Swainson s Warblers to adequately care for both their own nestlings and cowbird young. Received 10 April 2009. Accepted 2 September 2009. Swainson’s Warbler {Limnothlypis swainsonii) is a medium-sized wood warbler that occurs in dense understory vegetation in bottomland hard- wood forests of the southeastern United States. Nondescript olive-brown coloration and secretive behavior make it elusive and difficult to locate and study. Swainson’s Warblers have specialized breeding habitat requirements, preferring forest stands with dense understory vegetation, often of giant cane {Arunclinaria gigantea) (Brewster 1885, Meanley 1971, Brown et al. 2009). Conversion of higher elevation bottomland hard- wood forests and loss of canebrakes has reduced the availability of breeding habitat for Swainson’s Warblers (Brown et al. 2009), likely contributing to the status of this species as among those of greatest conservation concern nationally (Hunter et al. 1993, Rich et al. 2004). Fragmentation of bottomland hardwood forests has exposed some Swainson’s Warbler popula- tions to high frequency of brood parasitism by Brown-headed Cowbirds (Mololhrus ater) (Ben- son et al. 2010a). Thirty-six percent of Swainson’s Warbler nests in eastern Arkansas were parasit- ized (/? = 135 nests), and 10% received multiple cowbird eggs (Benson et al. 2010a). Brood parasitism may be a major contributing factor ' Department of Biological Sciences, Environmental Sciences Program, Arkansas State University, P. O. Box 599, Jonesboro, AR 72467. USA. -Current address; Illinois Natural Flistory Survey, 1816 South Oak Street, Champaign, IL 61820, USA. ■'Corresponding author; e-mail: tjbenson@gmail.com for population declines of some songbirds (Schmidt and Whelan 1999). Cowbirds reduce the productivity of host species by removing or damaging eggs, and through competition between host and cowbird young (Trine et al. 1998, Hoover 2003). This competition may lead to starvation of host nestlings in parasitized nests, especially for smaller host species such as Swainson’s Warblers (Benson et al. 2010a). Parasitism of smaller hosts may also reduce the growth rate of surviving nestlings (Dearborn et al. 1998, Burhans et al. 2000), leading to higher post- fledgling mortality. Parasitism of Swainson’s Warblers in eastern Arkansas decreased produc- tivity of successful nests from 2.75 in non- parasitized to 0.60 Swainson’s Warbler young in parasitized nests (Benson et al. 2010a); parasitized nests with multiple Brown-headed Cowbird young produced no Swainson’s Warbler young (Benson 2008). Beyond the direct effects of parasitism, in- creased food provisioning resulting from parasit- ism also may have harmful effects on adults. Presence of cowbird young in nests of relatively small hosts may lead to increased prey delivery rates by the host parents (Hoover and Reetz 2006) and increase food demand from adults (Kilpatrick 2002). This increased parental investment may lower future reproductive investment (Gustafsson and Sutherland 1988, Linden and Mpller 1989). Increased activity at the nest may attract predators and lead to decreased survival of parasitized nests (Haskell 1994, Leech and Leonard 1997, Dear- 75 76 THE WILSON JOURNAL OL ORNITHOLOGY • Vul. 122, No. I. March 2010 born et al. 1998, Dearborn 1999, Tewksbury et al. 2002). The stress imposed on the host parents by increased provisioning rates may ultimately lead to lower survival or increased emigration rates of individuals with parasitized nests (Hoover and Reetz 2006). We investigated food provisioning rates at parasitized and non-parasitized nests of Swain- son’s Warblers to understand why brood parasit- ism leads to decreased productivity and potential implications for con.servation. We expected the food-delivery rate of adults to increase with parasitism due to the large size of young cowbirds relative to Swainson’s Warblers. We also predict- ed multiple parasitism to have a greater impact than single parasitism, and for larger and older broods to receive more food deliveries than smaller and younger broods. METHODS Study Area. — We studied Swainson’s Warbler breeding biology in 2006 and 2007 at White River National Wildlife Refuge (Brown et al. 2009). This >60,000-ha refuge is among the largest remaining areas of bottomland hardwood forest in the Mississippi Alluvial Valley (Twedt and Loesch 1999). Areas with Swainson’s Warblers at White River National Wildlife Refuge were dominated by sugarberry {Celtis laevigata), sweet- gum {Liciuidamhar styraciflua), box elder {Acer negiindo), elm (Ulmiis spp.), oak (Qiiercus spp.), sycamore (Plataniis occidentalis), and hickory (Carya spp.). Dominant plants in the forest understory included greenbrier {Srnila.x spp.), Virginia creeper (Parthenocissus quinquefolia), peppervine (Ampelopsis arhorea), grape (Vitis spp.), spicebush (Lindera benzoin), box elder, and dense thickets of giant cane. Nest Searching and Video Monitoring. — We .searched for Swainson’s Warbler nests from late April to early August systematically by searching known territories with one to six observers, and opportunistically while engaged in other activities (Benson et al. 2()l()a). We recorded GPS (Global Positioning System) coordinates once nests were located, and returned at I- to 4-day intervals to ascertain the status and fate of each nest (Benson et al. 2()l()a). We installed time-lapse video systems at a subset of nests as part ol an effort to identify nest predators for this species (Benson et al. 20 1 Ob). These systems were a small color video camera (models PC5()61R and PCI68-IR, Supercircuits, Austin, TX, USA) mounted on 2.2- cm diameter wooden dowels; the camera was connected to a time-lapse VCR with two deep- cycle batteries >10 m from the nests in plastic weatheiproof containers. Camera systems were camouflaged to blend in with suiToundings; the presence of cameras appeared to not affect nest survival (Benson et al. 2010b). We returned to video-monitored nests at 1- to 4-day intervals to change batteries and tapes until nests failed or fledged young. Video Processing. — We collected data on food provisioning rates of adult Swainson’s Warblers at 25 nests, nine of which were parasitized by Brown-headed Cowbirds. We recorded the time of each feeding event for each nest. Data Analyses. — We assessed the effect of Brown-headed Cowbird parasitism on Swainson’s Warbler provisioning rates (feedings/hr) using repeated-measures ANOVA (SAS PROC MIXED) (Littell et al. 2006). We examined differences among nests containing no, one, and two cowbird nestlings to document the effect of parasitism. Other factors, including brood size and age of nestlings, may affect provisioning rates, and we tested for the influence of brood size, nest period (1^, 5-8, and 9-12 days), and interactions among parasitism and these two factors. We set signifi- cance at a = 0.05, and followed significant main effects or interactions with pair-wise contrasts. We considered nest, hour within nest, and day within nest as random effects to account for potential non- independence of observations, and used the Ken- ward-Roger approximation for denominator de- grees of freedom (Littell et al. 2006). We incoiporated brood size as a time-specific factor; con.sequently, nests were often observed with multiple brood sizes and fit into more than one category. We inspected the distribution of residuals and plots of residual versus predicted values to confirm our data conformed to the assumptions of normality and homogeneity of variances. RESULTS The mean brood size for the 16 non-parasitized nests was 2.86 (SE = 0.22; range = 1^), and the mean brood size for parasitized nests (including cowbirds) was 2.67 (SE = 0.33; range = 1^). The parasitized nests consisted of seven with one cowbird nestling and two with two cowbird nestlings. We recorded 6,320 feeding visits (.v = 252.8/ nest, SE = 24.1, range = 28-465) between 17 May and 25 July 2006 and 2007, and data were Pappas cl al. • SWAINSON’S WARBLER PROVISIONING 77 TABLE 1. Repeated-measures ANOVA examining the influence of number of Brown-headed Cowbird (BHCO) nestlings, brood size, nest period (1-4, 5-8, and 9-12 days = nestling age), and interactions among these factors on Swainson's Warbler provisioning rates at White River National Wildlife Refuge, Arkansas, 2006 and 2007. Effect df F p BHCO 2, 34 4.08 0.026 Brood size 3, 307 9.89 <0.001 Nest period 2, 291 35.92 <0.001 BHCO X brood size 5, 276 2.53 0.029 BHCO X nest period 4, 235 5.36 <0.001 Brood size X nest period BHCO X brood size X nest 6, 326 0.72 0.64 period 5, 297 0.64 0.67 collected from 2 to 12 days for each nest (.v = 8.4, SE = 0.6) representing 2,495 total nest hours {x = 99.8, SE = 7.7, range = 18-160). There was no significant three-way interaction among number of cowbird nestlings, brood size, and ne.st period, nor was there a significant interaction among brood size and nest period (Table 1 ). There were significant interactions between number of cow- bird nestlings and brood size, and number of cowbird nestlings and nest period (Table 1). Provisioning rate increased with brood size for non-parasitized nests (r > 1.88, P ^ 0.06; Fig. 1). The provisioning rate for nests with one cowbird nestling was similar to that for nests with a cowbird and a warbler nestling {t = —0.49, P = 0.63), but the rates for both nest types differed from nests with a single cowbird and two or three warbler nestlings {t 2: 3.09, P ^ 0.002). Nests with two cowbird nestlings and zero or one warbler nestling had similar provisioning rates {t = 1.27, P = 0.21). There was a marginally significant difference in provisioning rate between nests with a single warbler and a single cowbird nestling (r = -1.83, P = 0.07), and a significant difference between nests with three warbler 4.0 - 3.5 - •& 3.0 c "a CD CD 2.5 2.0 - • One nestling o Two nestlings T Three nestlings A Four nestlings (5) I I (10) ~r (12)1 (13)X (4) (2) II (6) (5) (2) io) 1.5 ~r" 0 Number of cowbird nestlings FIG. I . Effects of Brown-headed Cowbird parasitism on provisioning rale (teedmgs/hr ± SE) at Swainson s Warbler nests for brood sizes of one through four (including cowbird young). Data were collected from I6 non-parasitized and nine parasitized nests at White River National Wildlife Refuge, Arkansas, 2006 and 2007. The sample size of nests for each category is in parentheses; most nests were observed at multiple states and do not equal sums of the total number of nests observed. 78 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. I. March 2010 FIG. 2. Effects of Brown-headed Cowbird parasitism on provisioning rate (feedings/hr ± SE) at Swainson’s Warbler nests during three different periods of the nestling stage. Values were calculated at the mean for brood size. Data were collected from 16 non-parasitized and nine parasitized nests at White River National Wildlife Refuge, Arkansas, 2006 and 2007. The sample size of nests for each category is in parentheses; most nests were observed at multiple states and do not equal sums of the total number of nests observed. nestlings and those with two warblers and a cowbird nestling (/ = —2.76, P = 0.008). There were no differences between nests with two warblers and tho.se with a warbler and a cowbird nestling (t = —0.87, P = 0.39), or those with four warbler nestlings and those with three warblers and a cowbird nestling (t = —0.35, P = 0.73). Nests with two warbler nestlings or a warbler and a cowbird nestling had lower provisioning rates than those with two cowbird nestlings (/ = -3.75, P < 0.001, and t — —2.87, P = 0.007, respectively). Nests with three warbler nestlings or two warblers and a cowbird nestling had similar provisioning rates to a nest with one warbler and two cowbird nestlings it = — 1.19, = 0.24, and t = 0.60, P = 0.55, respectively). The provisioning rate in nests that were not parasitized increased with nestling age (/ ^ 6.65, P < 0.001; Fig. 2). There was a difference in provisioning rates between days 1 to 4 and 5 to 8 for nests with one cowbird nestling, (f = — 1 .98, P = 0.050), but no other combinations of two periods (/ < 0.92, P > 0.36). A nest with two cowbird nestlings had a greater provisioning rate from days 5 to 12 than for days 1 through 4 (r > 5.56, P < 0.001 ), but rates for days 5 to 8 and 9 to 12 were similar (/ = — 1.30, P = 0.19). Nests with a single cowbird nestling during the early nestling period had greater provisioning rates than non- parasitized nests {t = 2.88, P = 0.007), but there were no differences between rates for nests with two versus no cowbirds {t = 1.02, P — 0.32), or for nests with one versus two cowbird nestlings during this stage (/ = 0.69, P = 0.49). The frequency of parental feedings from days 5 through 8 was similar for non-parasitized nests and those with a single cowbird nestling (/ = -1.34, P = 0.19), but both differed from a nest with two cowbirds {t — —5.12, P < 0.001, and t = —4.17, P < 0.001, respectively). The rate of provisioning during the late nestling period was marginally greater in non-parasitized when com- Pappas ct al. • SWAINSON’S WARBLER PROVISIONING 79 pared to single-parasitized nests {t = 1.85, P = 0.07), but both differed from a double-parasitized nest (t = -2.97, P = 0.004, and t ^ -3.65, P < 0.001, respeetively). DISCUSSION The effect of cowbird parasitism on the provisioning rate of Swainson’s Warblers was variable and depended on brood size and age of nestlings. Our observed increases in provisioning rates with brood size for non-parasitized nests were consistent with trends observed for other species (e.g., Johnson and Best 1982, Haggerty 1992, Rosa and Muiphy 1994), as were the results for increased provisioning to older nestlings (e.g., Johnson and Best 1982, Knapton 1984, Goodbred and Holmes 1996). These results were expected given the increasing energetic demand of each nestling as they grow (Williams and Prints 1986). There were effects of parasitism on provision- ing rates at nests with one to three nestlings, and for all three nestling-age categories. Differences were generally more pronounced for nests with two cowbird young. However, these results must be viewed with caution due to small sample sizes. We collected provisioning data from only two multiply parasitized nests, only one of those nests also contained a warbler nestling, and only one of the nests survived into the middle and late nestling periods. Similar effects of single and multiple parasitism have been observed for other species (Dearborn et al. 1998, Hoover and Reetz 2006). However, there appeared to be a limit to the provisioning rate for Swainson’s Warblers. The mean provisioning rate of non-parasitized nests with four nestlings was similar to the rates for single-parasitized nests with three or four total nestlings (4 warblers vs. 2 warblers and 1 cowbird: /6o — 1.08, P = 0.28) and double- parasitized nests with two or three total nestlings (t3o = 1.40, P = 0.17; t57 = 0.08, P -= 0.94, respectively). This similarity of mean rates among these brood sizes, given the larger size and begging intensity of cowbird nestlings (Dearborn 1998), suggests food may be a limiting factor (the mass of a cowbird at fledging is approximately twice that of an adult Swainson’s Warbler; Lowther 1993, Brown and Dickson 1994, Hoover 2003). The maximum observed clutch size for Swain- son’s Warblers at our study area was tour eggs (Benson et al. 2010a). Our provisioning data support the hypothesis that this maximum is at least partly affected by food limitation (Lack 1947, 1954). The difference in productivity for Swainson’s Warblers between successful non- parasitized (2.75 young/successful nest) and parasitized nests (0.60 young/successtul nest) is among the more-extreme reductions in similar- sized passerines (Payne 1998). This reduction for Swainson’s Warblers is mostly due to starvation of host nestlings (Benson et al. 2010a). The similar-sized Prothonotary Warbler {Protonotaria citrea), which breeds in lower elevation bottom- land hardwood forests, is often able to fledge multiple cowbird and warbler young (Hoover 2003). The observed provisioning rates for Prothonotary Warblers (.v = 9.54/hr for non- parasitized, X — 16.96/hr for parasitized nests; Hoover and Reetz 2006) far exceed the rates we observed for Swainson’s Warblers (.v < 5/hr). Further evidence of food limitation of Swainson’s Warblers in our study area includes food abun- dance which appears to influence habitat use (Brown 2008). These birds have larger home ranges than similar-sized passerines (,v = 9.4 ha; Anich et al. 2009), suggesting they require large areas to acquire necessary resources. The increased sound and activity at parasitized nests has been observed to increase nest predation in some cases (Haskell 1994, Leech and Leonard 1997, Dearborn 1999, Tewksbui^ et al. 2002). Nest predation rates in our study population appeared to be unaffected by brood size or parasitism, possibly because predation rates were already relatively high (Benson et al. 2010a, b). However, the complex relationships among brood size, nestling age, parasitism status, and provisioning rate may lead to increased predation risk only in limited circumstances and not necessarily as a simple function of brood size or parasitism status. We observed effects of parasitism on prey delivery rates by Swainson’s Warblers, but several questions remain. For example, the size or type of prey that adults bring to the nest may differ among brood sizes or with nestling age in some species (Biermann and Sealy 1982), and the relative contributions of males and females to provisioning young may also vary (Nolan 1978, Carey 1990); these issues merit further study. The inlluence of habitat quality on provisioning rate and interaction between habitat quality and brood parasitism effects also remain unknown. Habitat loss and fragmentation present signif- icant conservation challenges for Swainson’s Warbler populations. This fragmentation may 80 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 1. March 2010 lead to low productivity due to high predation and brood parasitism rates in some populations (Benson et al. 2010a). Efforts should be made to conserve and improve existing habitats by pro- viding forested buffers between quality Swain- son’s Warbler habitat and agricultural areas to reduce the negative effects of cowbird parasitism in this and other populations. Cowbird removal programs may be beneficial in limited circum- stances, but are not cost-effective, or a permanent solution, and should be of secondary importance to habitat restoration and landscape-level man- agement (Stutchbury 1997, Rothstein and Peer 2005). ACKNOWLEDGMENTS Funding was provided by the Arkansas Game and Fish Commission and the U.S. Fish and Wildlife Service (USFWS) through a State Wildlife Grant and a separate cost-share grant from the USFWS. We are grateful to Catherine Rideout, Laurel Barnhill, Richard Hines, W. C. Hunter, and Steve Reagan for their support of this research. We thank the National Science Foundation which provided funding for Sara Pappas through the Research Experiences for Undergraduate (REU) program (NSF-DBI 0552608 Dowling) and Carolyn Dowling, the head of Arkansas State University’s REU program. Sara Pappas thanks Daniel Albrecht-Mallinger for his support. We are especially grateful to Jeremy Brown, Eric Huskinson, Nick Anich, Carolina Roa, Jason Sardell, Mary-Beth Albrechtsen, Christy McCarroll, and Jessica O’Connell for providing valuable field a.ssistance and Nick Anich, C. E. Braun, and two anonymous reviewers for providing comments on the manuscript. LITERATURE CITED Anich, N. M., T. J. Benson, and J. C. Bednarz. 2009. Estimating territory and home-range sizes: do singing locations alone provide an accurate estimate of space use? Auk 126:626-634. Benson, T. J. 2008. Habitat use and demography of Swainson’s Warblers in eastern Arkansas. Disserta- tion. Arkansas State University, Jonesboro, USA. Benson, T. J., J. D. Brown, and J. C. Bednarz. 2010b. Identifying predators clarifies predictors of nest success in a temperate passerine. Journal of Animal Ecology 79:225-234. Benson. T. J., N. M. Anich, J. D. Brown, and J. C. Bednarz. 2010a. Habitat and landscape effects on brood parasitism, nest survival, and fledgling produc- tion in Swain.son's Warblers. Journal of Wildlife Management 74:81-93. Biermann. G. C. and S. G. Sealy. 1982. Parental feeding of ne.stling Yellow Warblers in relation to brood size and prey availability. Auk 99:332-341. BREW.STRR, W. 1885. Swainson’s Warbler. Auk 2:65-80. Brown, J. D. 2008. Arthropod communities and habitat characteristics of sites occupied by Swain.son’s Warblers in the White River National Wildlife Refuge, Arkansas. Thesis. Arkansas State University, Jones- boro, USA. Brown, J. D., T. J. Benson, and J. C. Bednarz. 2009. Vegetation characteristics of habitat occupied by Swainson’s Warblers at the White River National Wildlife Refuge, Arkansas. Wetlands 29:586-597. Brown, R. E. and J. G. Dickson. 1994. Swainson’s Warbler {Limnothlypis swainsonii). The birds of North America. Number 126. Burhans, D. E., F. R. Thompson III, and J. Faaborg. 2000. Costs of parasitism incurred by two songbird species and their quality as cowbird hosts. Condor 102:364-373. Carey, M. 1990. Effects of brood size and nestling age on parental care by male Field Sparrows {Spizella pusilla). Auk 107:580-586. Dearborn, D. C. 1998. Begging behavior and food acquisition by Brown-headed Cowbird nestlings. Behavioral Ecology and Sociobiology 43:259-270. Dearborn, D. C. 1999. Brown-headed Cowbird nestling vocalizations and risk of nest predation. Auk 1 1 6:448- 457. Dearborn, D, C., A. D. Anders, F. R. Thompson III, and J. Faaborg. 1998. Effects of cowbird parasitism on parental provisioning and nestling food acquisition and growth. Condor 100:326-334. Goodbred, C. and R. T. Holmes. 1996. Factors affecting food provisioning of nestling Black-throated Blue Warblers. Wilson Bulletin 108:467^79. Gustafsson, L. and W. j. Sutherland. 1988. The costs of reproduction in the Collared Flycatcher Ficediila alhicollis. Nature 335:813-815. Haggerty, T. M. 1992. Effects of nestling age and brood size on nestling care in Bachman’s Sparrow {Aimo- phila aestivalis). American Midland Naturalist 128:1 15-125. Haskell, D. 1994 Experimental evidence that nestling begging behaviour incurs a cost due to nest predation. Proceedings of the Royal Society of London, Series B 257:161-164. Hoover, J. P. 2003. Multiple effects of brood parasitism reduce the reproductive success of Prothonotary Warblers, Protonotaria citrea. Animal Behaviour 65:923-934 Hoover, J. P. and M. J. Reetz. 2006. Brood parasitism increases provisioning rate, and reduces offspring recruitment and adult return rates, in a cowbird host. Oecologia 149:165-173. Hunter, W. C., D. N. Pashley, and R. E. F. E.scano. 1993. Neotropical migratory landbird species and their habitat of .special concern within the Southeast Region. Pages 159-171 in Status and management of neotrop- ical migratory birds (D. M. Finch and S. W. Stangel, Editors). USDA, Forest Service, General Technical Report RM-229. Rocky Mountain Research Station, Fort Collins, Colorado, USA. Johnson, E. J. and L. B. Best. 1982. Factors affecting feeding and brooding of Gray Catbird nestlings. Auk 99:148-156. Kilpatrick, M. A. 2002. Variation in growth of Brown- Pappas cl al. • SWAINSON’S WARBLER PROVISIONING 81 headed Cowbird (Molothrus ater) nestlings and energetic impacts on their host parents. Canadian Journal of Zoology 80; 145-153. Knapton, R. W. 1984. Parental feeding of nestling Nashville Warblers: the effects of food type, brood-size, nestling age, and time of day. Wilson Bulletin 96:594-602. Lack, D. 1947. The significance of clutch size. Ibis 89:302-352. Lack, D. 1954. The natural regulation of animal numbers. Clarendon Press, Oxford, United Kingdom. Leech, S. M. and M. L. Leonard. 1997. Begging and the risk of predation in nestling birds. Behavioral Ecology 8:644-646. Linden, M. and A. P. Moller. 1989. Cost of reproduction and covariation of life history traits in birds. Trends in Ecology and Evolution 4:367-371. Littell, R. C., G. a. Milliken, W. W. Stroup, R. D. WOLFINGER, AND O. SCHABENBERGER. 2006. SAS for mixed models. SAS Institute Inc., Cary, North Carolina, USA. Lowther, P. E. 1993. Brown-headed Cowbird {Molothrus ater). The birds of North America. Number 47. Meanley, B. 1971. Natural history of the Swainson’s Warbler. North American Fauna Number 69. U.S. Department of the Interior, Washington, D.C., USA. Nolan Jr., V. 1978. The ecology and behavior of the Prairie Warbler Dendroica discolor. Ornithological Monographs 26:1-595. Payne, R. B. 1998. Brood parasitism in birds; strangers in the nest. BioScience 48:377-386. Rich, T. D., C. J. Beardmore, H. Berlanga, P. J. Blancher, M. S. W. Bradstreet, G. S. Butcher, D. W. Demarest, E. H. Dunn, W. C. Hunter, E. E. Inigo-Elias, j. a. Kennedy, A. M. Martell, A. O. Panjabi, D. N. Pashley, K. V. Rosenberg, C. M. Rustay, j. S. Wendt, and T. C. Will. 2004. Partners in Flight North American landbird conservation plan. Cornell Laboratory of Ornithology, Ithaca, New York, USA. Rosa, S. M. and M. T. Murphy. 1994. Trade-offs and constraints on Eastern Kingbird parental care. Wilson Bulletin 106:668-678. Rothstein, S. I. AND B. D. PEER. 2005. Conservation solutions for threatened and endangered cowbird {Molothrus spp.) hosts: separating fact from fiction. Ornithological Monographs 57:98-114. Schmidt, K. A. and C. J. Whelan. 1999. The relative impacts of nest predation and brood parasitism on seasonal fecundity in songbirds. Conservation Biology 13:46-57. Stutchbury, B. j. M. 1997. Effects of female cowbird removal on reproductive success of Hooded Warblers. Wilson Bulletin 109:74-81. Tewksbury, J. J., T. E. Martin, S. J. Hejl, M. J. Kuehn, AND J. W. Jenkins. 2002. Parental care of a cowbird host: caught between the costs of egg-removal and nest predation. Proceedings of the Royal Society of London, Series B 269:423^29. Trine, C. L., W. D. Robinson, and S. K. Robinson. 1998. Consequences of Brown-headed Cowbird brood par- asitism for host population dynamics. Pages 273-295 in Parasitic birds and their hosts (S. I. Rothstein. and S. K. Robinson, Editors). Oxford University Press, New York, USA. Twedt, D. j. and C. R. Loesch. 1999. Forest area and distribution in the Mississippi Alluvial Valley: impli- cations for breeding bird conservation. Journal of Biogeography 26:1215-1224. Williams, J. B. and A. Prints. 1986. Energetics of growth in nestling Savannah Sparrows: a comparison of doubly labeled water and laboratory estimates. Condor 88:74-83. The Wilson Journal of Ornithology 122( 1 ):82-87, 2010 EFFECTS OF FAT AND FEAN BODY MASS ON MIGRATORY FANDBIRD STOPOVER DURATION CHAD L. SEEWAGEN' 2 ' AND CHRISTOPHER G. GUGLIELMO' ABSTRACT. — We used quantitative magnetic resonance body composition analysis and radiotelemetry to examine whether fat and lean body mass affected stopover durations of 1 1 birds captured during autumn migration in New York City, USA. Two Swainson’s Thrushes (Catharus ustulatus), two Hermit Thaishes (C. giittatus), and seven Ovenbirds {Seiurus aurocapilla) were used in the study. Ovenbird stopover duration was significantly and negatively related to fat mass but unrelated to lean body mass. The same relationships were found when data from all three species were combined to increase sample sizes. Birds that departed within I day had fat stores upon capture that represented at least 1 1 % of their total body mass whereas those with fat content <6% of total body mass remained for no fewer than 4 days. Arrival fat mass clearly influenced time birds spent at the site but lean body mass did not. Conditions for increasing or maintaining fat stores provided by urban stopover sites may affect the migration timing of birds. Received 26 May 2009. Accepted 16 September 2009. Migrating landbirds require several stopovers for rest and refueling, as most are incapable of completing their migration in a single flight. Time and energy spent during stopover periods greatly exceeds that spent aloft (Wikelski et al. 2003, Bowlin et al. 2005), and behavior of birds at stopover sites can greatly influence their overall migration success. Factors that govern a bird’s decision to terminate a stopover and begin the next flight have received much attention, partic- ularly through development of theoretical models of optimal migration strategies (Alerstam and Hedenstrdm 1998, Houston 1998). There is evidence that fuel stores, refueling rate, distance from final destination, date, predation risk, and weather conditions can individually, or in some combination influence a bird’s length of stay at a given stopover site (Wang and Moore 1997, Akesson and Hedenstrom 2000, Danhardt and Lindstrdm 2001, Dierschke and Delingat 2001, Schaub et al. 2008). Field studies of the effects of intrinsic and extrinsic factors on stopover duration are ham- pered by the difficulty of knowing when birds arrive at, and depart from, the site of interest. Several studies have measured stopover duration using mark-recapture data, while assuming birds were marked upon arrival and final recapture occurred on their actual departure date (e.g.. ' Department of Biology, University of Western Ontario, 1151 Richmond .Street North, London, ON N6A .5K7, Canada. ^Department of Ornithology, Wildlife Con.servation Society. 2300 Southern Boulevard, Bronx, NY 10460, USA. 'Corresponding author; e-mail: cseewagcn@wc.s.org Cherry 1982, Wang and Moore 1997, Morris and Glasgow 2001). The latter assumption can be especially tenuous, as migrants may learn to avoid mist nets (Mac Arthur and MacArthur 1974) or move beyond the coverage area within the stopover site (Chernetsov and Mukhin 2006, Tsvey et al. 2007). This method also relies on condition and behavior of the small proportion of birds that are recaptured to represent the majority, when recaptures may be biased towards individ- uals in poor condition that remain at the site for a relatively long time (Guglielmo et al. 2005, Morris et al. 2005, Hays 2008). More recently, survival analyses and other probabilistic models have been used to relax these assumptions and improve accuracy of stopover duration estimates ba.sed on mark-recapture data (Schaub et al. 2001, 2008), although large sample sizes are generally needed (Schaub 2006). Radio tracking offers another approach for measuring stopover duration that does not rely on recaptures and allows knowing departure date with near certainty (Chernetsov and Mukhin 2006, Tsvey et al. 2007). Drawbacks of radiotelemetry include high equipment costs, possible effects of transmitters on bird behavior, and the assumption that birds are captured upon arrival. Examining the effect of energetic condition on stopover duration is complicated by the need for accurate, non-lethal measurements of body com- position. Commonly used condition indices (e.g., body mass, size-corrected body mass, visual fat scores) can potentially provide reliable estimates of fat content (Conway et al. 1994, Spengler et al. 1995, Seewagen 2008). However, their accuracy may be species-specific (Skagen et al. 1993, Spengler et al. 1995, Seewagen 2008) or weak- 82 Secn'cigen ami Guglielmo • BODY COMPOSITION AND STOPOVER DURATION 83 ened by inter-observer variation (Krementz and Pendleton 1990), and they do not provide separate, direct measures of fat and lean body mass. Quantitative magnetic resonance analysis (QMR; Taicher et al. 2003, Tinsley et al. 2004) in contrast provides accurate, objective, and direct measures of the fat and lean body mass of small birds (CGG, unpubl. data). We attached radio transmitters on 1 1 migrant songbirds at our study site in New York City during autumn 2008 in a pilot study to test the feasibility of radio tracking birds in an urban setting. We used a QMR body composition analyzer on all captured migrants for a concurrent study of body composition dynamics during stopover refueling. The stopover duration and body composition data obtained by radiotelemetry and QMR, respectively, afforded an additional opportunity to examine relationships between these variables. We coupled telemetry and QMR data to examine if arrival fat and lean body mass affected the stopover durations of these 1 1 migrants. METHODS Study Site and Animal Capture. — We captured birds during autumn 2008 on the grounds of the Bronx Zoo as part of an ongoing study of urban stopover ecology of migrant songbirds. The Bronx Zoo is a 107-ha park in Bronx County, New York, USA. Our study site was a 4.9-ha fragment of riparian forest on the eastern edge of the zoo that does not contain exhibits and is not open to visitors (40° 85' N, 73° 87' W). Red oak {Quer- cus rubra), sweet gum (Lic/uidambar styraciflua), swamp dogwood (Cornus foemina), and willows (SaUx spp.) are the dominant tree species in the area. Previous research has suggested this site offers suitable refueling conditions for landbird migrants (Seewagen and Slayton 2008; CLS and CGG, unpubl. data). Birds were captured in mist nets from sunrise until -1200 hrs during 4 September to 22 October. Captured birds were marked with a U.S. Geological Survey aluminum leg band, measured (unflattened wing length to 1 mm), and weighed on a digital balance to 0.1 g. Approximately 75 pi of blood was collected by brachial veinipuncture for a separate study. Magnetic Resonance Analysis. — Conscious birds were scanned within 45 min of capture and banding in a QMR body composition analyzer (Echo-MRI, Echo Medical Systems, Houston, TX, USA) housed in a customized mobile laboratory (Glendale Recreational Vehicles, Strathroy, ON, Canada) at the study site. Birds were scanned in duplicate on the “small bird’’ and “two-accumu- lation” settings of the Echo-MRI software, yielding measures of fat mass and wet lean body mass to 0.001 g. Total scanning time was —4 min/individual and the average coefficient of variation for individual birds was 1 .1% for fat and 0.4% for wet lean mass. Raw QMR scan data were transformed using calibration equations for small birds developed with House Sparrows (Passer domesticus) and Zebra Finches (Taenio- pygia guttata) in a laboratory validation study (CGG, unpubl. data). Validation indicated the relative error for predictions of fat and wet lean mass were ± 11% and ± 1.5%, respectively. Radiotelemetry. — ^We radiomarked each Swain- son’s Thrush (Catharus ustulatus). Hermit Thrush (C. guttatus), and Ovenbird (Seiurus aurocapilla) captured beginning 3 October until all 1 1 transmitters we had were deployed. These species were used because they are relatively common migrants in the area, body masses are amenable to radiotelemetry, and they are the focus of concur- rent studies in progress at the site. The 0.5-g transmitters (A2415, Advanced Telemetry Sys- tems, Isanti, MN, USA) were affixed after QMR scanning with eyelash adhesive directly to a cleared area of skin in the interscapular region (adapted from Raim 1978). Birds were held in bags for - 15 min to allow the adhesive to dry and released within 50 m of where originally cap- tured. We could not be certain birds were captured and marked upon arrival. Eight of 1 1 birds were marked on days when migrant capture rates were high (-double) relative to the 2 previous days at this site and at two affiliated banding stations elsewhere in New York City. This increased our confidence these individuals were new amvals and marked during their first day at the stopover site (sensu Tsvey et al. 2007). However, we use the term “minimum stopover duration” hereafter because of this uncertainty. We checked for presence/absence daily at sunrise, noon, and sunset using a handheld Yagi antenna and receiver (R45()0S, Advanced Telem- etry Systems, Isanti, MN, USA) from within the study site. We searched for birds from other points throughout the zoo including roofs of two three- story buildings if we initially failed to detect a signal from within the study site. We assumed 84 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 1, March 2010 TABLE 1. Means ± SD and ranges of total body mass, fat mass, lean body mass, and minimum stopover durations of three species of migratory landbirds radiotracked during autumn stopovers in New York City, USA. Species n Total body mass (g) Fat mass (g) Lean body mass (g) Stopover duration (days) Swainson’s Thrush 2 27.3, 30.1 1.700, 2.410 22.760,26.119 4,5 Hermit Thrush 2 30.2, 34.6 1.570, 4.000 27.103,29.089 2,7 Ovenbird 7 20.0 ± 1.3 18.3-21.7 2.461 ± 1.069 1.068-3.911 17.009 ± 0.795 4.3 ± 5.0 15.713-18.145 1-14 birds had left the area permanently and resumed migration if they were undetected for at least 4 consecutive days. We considered a stopover of 1 day as the initial capture of a bird at the site in the morning and departure any time that night. Statistical Analyses. — ^We used standardized residuals from linear regressions of lean body mass and wing length to correct lean body mass for body size variation. We could not adjust fat mass in the same manner because fat mass was not significantly related to wing length. We instead divided fat mass by total body mass and arcsine-root transformed the proportions (Zar 1999, Gotelli and Ellison 2004). We used backwards selection multiple regres- sion with adjusted fat and lean body mass as predictor variables, and number of days of stopover as the response variable to examine the effect of body composition on stopover duration. Collinearity did not occur, as fat and lean body mass were not significantly correlated. We performed the analyses on Ovenbird data alone and on data from all three species combined. Age and gender were not considered because sample sizes were small and not all birds could be confidently assigned to age or gender classes. All variables met normality assumptions. We con- ducted tests with SPSS 16.0 and interpreted results as significant when P < 0.05. Mean (± SD) values are reported. RESULTS We radiomarked two Swainson’s Thrushes, two Hermit Thrushes, and seven Ovenbirds. Swainson's Thrushes were marked on 2 and 6 October, and Hermit Thrushes were marked on 10 and 11 October. Four Ovenbirds were marked on 3 October and three were marked on 6 October. Minimum stopover durations ranged Irom 1 to 14 days (Table 1). Three individuals departed within 1 day and the remainder stayed for at least 2 days. Initial fat content ranged from 6 to 8% of total body mass in Swainson’s Thrushes, from 5 to 12% in Hermit Thrushes, and from 6 to 18% in Ovenbirds. An Ovenbird was recaptured on the eighth day of its 14 day stopover; total body mass (measured by digital balance) increased 4.2 g (23% of original), and fat and lean body mass (measured by QMR) increased 2.767 g (259% of original) and 1.884 g (11% of original), respec- tively. The initial fat loads of birds that stayed at the site for only 1-2 days (/? = 5) averaged 14 ± 3% of total body mass, whereas the initial fat loads of birds that stayed for the longest periods (7, 8, and 14 days; n = 3) averaged 6 ± 0.5% of total body mass. Ovenbird stopover duration was unrelated to lean body mass (r = 0.02, Fja = 0.36, P = 0.58), and significantly and negatively related to fat mass {r = 0.74, F,,5 = 13.95, P = 0.013). The same relationships were found when data from all three species were combined to increase sample sizes (lean: r = 0.03, F2,8 = 0.00, P = 0.98; fat: r = 0.56, F,,9 =11.31, F = 0.008; Fig. 1). The relationship between fat and stopover duration appeared non-linear, and we explored the fit of an exponential model to the data a posteriori. The exponential model was also highly significant (r = 0.71, F,.9 = 22.08, P = 0.001; Fig. 1). DISCUSSION This is the first combination of magnetic resonance technology and radiotelemetry to examine the influence of fat and lean body mass on stopover duration to our knowledge. We believe this is also the first description of autumn stopover durations of migrants within a major urban center. Fat mass upon presumed arrival date appeared to strongly influence how long birds remained at the stopover site. This is consistent with other findings that lean birds stop over longer than fatter birds (Cherry 1982, Loria and Moore 1990, Wang and Moore 1997, Matthews 2008; but see Safriel and Lavee 1988, Salewski and Schaub 2007). The Seewageii and Guglielnio • BODY COMPOSITION AND STOPOVER DURATION 85 FIG. 1. Relationship between arrival fat content (arcsine-root transformed g fat/g total body mass) and minimum stopover duration of Swainson’s Thrushes (squares), Hermit Thrushes (triangles), and Ovenbirds (circles) captured during autumn migration in New York City, USA. Solid and dashed lines represent linear (r = 0.56, T,,9 = 1 1.31, P = 0.008) and exponential (r = 0.71, P,, 9 = 22.08, P = 0.001) models, respectively. arrival fat content of individuals that stayed for only 1 day was at least 1 1 % of total body mass, whereas all individuals with fat content <6% of total body mass remained for no fewer than 4 days, suggesting a possible threshold fat level for departure. However, how much additional fat birds acquired between capture and departure is unknown. We did not find any relationship between lean body mass and minimum stopover duration. Lean tissues (e.g., flight muscles, digestive organs) may contribute significantly to total body mass dy- namics during migration. This is particularly apparent in shorebirds that routinely make excep- tionally long non-stop flights and in passerines when they must cross ecological barriers (Biebach 1998, Karasov and Pinshow 1998, Battley et al. 2000, Guglielmo and Williams 2003). No formi- dable ecological barriers exist immediately north of New York City, and it is possible most arriving autumn passerine migrants have not recently metabolized substantial non-fat tissue, and re- building of lean mass is not a significant component of stopover refueling. This is not supported by the lean mass gained by the recaptured Ovenbird which accounted for ~40% of its total body mass increase. Four of eight other (non-radio-marked) birds that we recaptured during the season also had gains in lean mass (CLS and CGG, unpubl. data). It is unknown from our small sample sizes of recaptured and radio- marked birds whether migrants commonly deposit lean tissue at this site and if lean body mass affects departure timing. Fusani et al. (2009) did not find lean mass (muscle scores) to influence migratory restlessness in two of three species examined. This also suggests lean mass does not influence departure decisions. Further study on the relationship between lean body mass and stopover duration is needed. Fuel load is not the only variable that can influence stopover duration; weather, predation risk, and date are also factors that possibly govern a bird’s decision to depart a stopover site. Our small sample size of 1 1 birds prohibited including additional predictor variables in the analyses and we cannot assess any effect they may have had on stopover duration. However, the birds were marked on 5 days over a span of only 9 days; thus, they probably experienced similar predation 86 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. I. March 2010 risk and temporal pressure, and were exposed to comparable weather conditions. Arrival body composition likely differed among the individuals we studied more so than any of these potential extrinsic influences. The recaptured Ovenbird may illustrate how multiple variables can affect stopover duration. This bird appeared to have stored sufficient fat to resume migration by its eighth stopover day, yet it remained at the site for 6 additional days. Overnight wind direction following recapture was primarily from the south until the night the bird departed when winds came from the north, the preferred direction of southbound passerines (Gauthreaux 1991). It is possible this bird was energetically prepared for departure by the eighth day but waited 6 additional days for more favorable overnight flying conditions. Six of the other birds similarly departed on nights with northern winds; the remaining four departed during eastern winds. Our sample size was small, but arrival fat mass had a clear and strong effect on time birds remained at the stopover site. Similar relation- ships between stopover duration and some measure of energetic condition have been docu- mented previously (Cherry 1982, Loria and Moore 1990, Wang and Moore 1997, Matthews 2008), but the contributions of fat and lean mass were not addressed individually. Quantitative magnetic resonance scanning allowed us to .separately examine how each tissue affected stopover duration. Arrival fat mass at our study site affected migrants’ decisions to leave whereas lean body mass did not. Thus, the conditions for increasing or maintaining fat stores provided by this site and possibly other similar urban habitats can affect the migration timing of birds using them. Our results demonstrate quantitative mag- netic re.sonance analysis can be useful under field conditions, and that combining it with telemetry and other approaches will improve our ability to understand the stopover biology of birds. ACKNOWLEDGMENTS This research would not have been possible without the hard work and contributions of Rafael Campos. Nancy Clum. Gary Del ‘Abate. Robert Haupt, Liam McGuire. Mark .Shaw, Christine Sheppard. Eric Slayton, and especially Quentin Hays. Funding was provided by the Canadian Foundation for Innovation. Ontario Ministry of Re.search and Innovation, and an NSFRC Discovery Grant to CGG. Part of this research was undertaken as part of an Environmental Benefit Project funded through the re.solu- tion of an enforcement action for violations of the Environmental Conservation Law of New York State and its implementing regulations. Helpful comments on earlier drafts of this manuscript were provided by David Cerasale, Robert DeCandido, and Liam McGuire. Research protocols were approved by the University of Western Ontario’s Council on Animal Care and the Wildlife Conservation Society’s Institutional Animal Care and Use Committee. LITERATURE CITED Akesson. S. and A, Hedenstrom. 2000. Wind selectivity of migratory flight departures in birds. Behavioral Ecology and Sociobiology 47:140-144. Alerstam, T. and A. Hedenstrom. 1998. The develop- ment of bird migration theory. Journal of Avian Biology 29:343-369. Battley P. F.. T. Piersma. M. W. Dietz. S. Tang. A. Dekinga. and K. Hulsman. 2000. Empirical evidence for differential organ reductions during trans-oceanic bird flight. Proceedings of the Royal Society of London, Series B 267:191-195. Biebach. H. 1998. Phenotypic organ flexibility in Garden Warblers (Sylvia horin) during long-distance migra- tion. Journal of Avian Biology 29:529-535. Bowlin, M. S., W. W. Cochran, and M. C. Wikelski. 2005. Biotelemetry of New World thrushes during migration: physiology, energetics and orientation in the wild. Integrative and Comparative Biology 45:295-304. Chernetsov, N. and A. Mukhin. 2006. Spatial behavior of European Robins during migratory stopovers: a telemetiy study. Wilson Bulletin 118:364-373. Cherry, J. D. 1982. Fat deposition and length of stopover of migrant White-crowned Sparrows. Auk 99:725- 732. Conway, C. J., W. R. Eddleman, and K. L. Simpson. 1994. Evaluation of lipid indices of the Wood Thrush. Condor 96:783-790. Danhardt, j. and A. Lindstrom. 2001. Optimal departure decisions of songbirds from an experimental stopover site and the significance of weather. Animal Behavior 62:235-243. Dierschke, V. AND J. Delingat. 2001. Stopover behaviour and departure decision of Northern Wheatears, Oc- nanlhe oenantlic. facing different onward non-stop flight distances. Behavioral Ecology and Sociobiology 50:53.5-545. Fusani, L., M. Cardinale, C. Carere, and W. Goymann. 2009. Stopover decision during migration: physiolog- ical conditions predict nocturnal restlessness in wild passerines. Biology Letters 5:302-305. Gauthreaux Jr., S, A. 1991. The flight behavior of migrating birds in changing wind fields: radar and visual analyses. American Zoologist 31:187-204. Gotelli, N. j. and A. M. Ellison. 2004. A primer of ecological statistics. Sinauer A.ssociates Inc., Sunder- land, Massachu.setts, USA. Guglielmo, C. G. and T. D. Williams. 2003. Phenotypic flexibility of body composition in relation to migratory state, age, and .sex in the Western Sandpiper (Calidris Scewagcn ami Guglielnio • BODY COMPOSITION AND STOPOVER DURATION 87 maitri). Physiological and Biochemical Zoology 76:84-98. Guglielmo, C. G., D. J. Cerasale, and C. Eldermire. 2005. A field validation of plasma metabolite profiling to assess refueling performance of migratory birds. Physiological and Biochemical Zoology 78:116-125. Hays, Q. R. 2008. Stopover behavior and refueling physiology in relation to migration distance in Wilson's Warblers {Wilsonia piisilla). Thesis. Univer- sity of Western Ontario, London, Canada. Houston, A, 1. 1998. Models of avian optimal migration: state, time and predation. Journal of Avian Biology 29:395-404. Kar.asov, W. H. and B. Pinshow. 1998. Changes in lean mass and in organs of nutrient assimilation in a long- distance passerine migrant at a springtime stopover site. Physiological Zoology 71:435^48. Krementz, D. G. and G. W. Pendleton. 1990. Fat scoring: sources of variability. Condor 92:500-507. Loria, D. E. and F. R. Moore. 1990. Energy demands of migration on Red-eyed Vireos, Vireo olivaceus. Behavioral Ecology 1:24—35. MacArthur, R. H. and a. T. MacArthur. 1974. On the use of mist nets for population studies of birds. Proceedings of the National Academy of Sciences of the USA 71:3230-3233. Matthews, S. N. 2008. Stopover habitat utilization by migratory landbirds within urbanizing landscapes of central Ohio. Dissertation. Ohio State University, Columbus, USA. Morris, S. R. and J. L. Glasgow. 2001. Comparison of spring and fall migration of American Redstarts on Appledore Island, Maine. Wilson Bulletin 113:202- 210. Morris, S. R., D. A. Liebner, A. M. Larracuente, E. M. Escamilla, and H. D. Sheets. 2005. Multiple- day constancy as an alternative to pooling for estimating mark-recapture stopover length in Nearc- tic-neotropical migrant landbirds. Auk 122:319- 328. Raim, A. 1978. A radio transmitter attachment for small passerine birds. Bird-Banding 49:326-332. Safriel, U. N. and D. Lavee. 1988. Weight changes of cross-desert migrants at an oasis — do energetic considerations alone determine the length of stopover? Oecologia 76:61 1-619. Salewski, V. AND M. SCHAUB. 2007. Stopover duration of Palearctic passerine migrants in the western Sahara - independent of fat stores? Ibis 149:223—236. ScHAUB, M. 2006. How to study departure decisions of migrants from stopover sites using capture-recapture data. Acta Zoologica Sinica 52:602-605. SCHAUB. M., L. Jenni and F. Bairlein. 2008. Fuel stores, fuel accumulation, and the decision to depart from a migration stopover site. Behavioral Ecology 19:657- 666. SCHAUB, M.. R. Pradel, L. Jenni, and J. Lebreton. 2001. Migrating birds stop over longer than usually thought: an improved capture-recapture analysis. Ecology 82:852-859. Seewagen, C. L. 2008. An evaluation of condition indices and predictive models for non-invasive estimates of lipid mass of migrating Common Yellowthroats. Ovenbirds and Swainson’s Thrushes. Journal of Field Ornithology 79:80-86. Seewagen, C. L. and E. J. Slayton. 2008. Mass changes of migratory landbirds during stopovers in a New York City park. Wilson Journal of Ornithology 120:296- 303. Skagen, S. K., F. L. Knopf, and B. S. Cade. 1993. Estimation of lipids and lean mass of migrating sandpipers. Condor 95:944-956. Spengler, T. j., P. L. Leberg, and W. C. Barrow Jr. 1995. Comparison of condition indices in migratory passerines at a stopover site in coastal Louisiana. Condor 97:438—444. Taicher, G. Z., F. C. Tinsley, A. Reiderman, and M. L. Heiman. 2003. Quantitative magnetic resonance (QMR) method for bone and whole-body-composition analysis. Analytical and Bioanalytical Chemistry 377:990-1002. Tinsley, F. C., G. Z. Taicher, and M. L. Heiman. 2004. Evaluation of a quantitative magnetic resonance method for mouse whole body composition analysis. Obesity Research 12:150-160. Tsvey, A., V. N. Bulyuk, and V. Kosarev. 2007. Influence of body condition and weather on departures of first-year European Robins. Erithaciis mhecula. from an autumn migratory stopover site. Behavioral Ecology and Sociobiology 61:1665-1674. Wang, Y. and F. R. Moore. 1997. Spring stopover of intercontinental migratory thrushes along the north- ern coast of the Gulf of Mexico. Auk 114: 263-278. WiKELSKi. M., E. M. Tarlow, A. Raim, R. H. Diehl, R. P. Larkin, and G. H, Visser. 2003. Costs of migration in free-llying songbirds. Nature 423:704. Zar, j. H. 1999. Biostatistical analysis. Fourth Edition. Prentice Hall. New Jersey. USA. The Wilson Journal of Ornithology 122( l);88-94, 2010 SEASONAL FLUCTUATION OF THE ORANGE- WINGED AMAZON AT A ROOSTING SITE IN AMAZONIA LEILIANY NEGRAO de MOURAT^ JACQUES M. E. VIELLIARDr AND MARIA LUISA da SILVA' STRACT.— We recorded fluctuations in a population of Orange-winged Amazon (Amazona amazonica) during 1 year at a roostmg site on an island near Belem, Para, Brazil. Parrots were counted from a boat by a minimum of three n ^ observers, each team oriented in different directions. Orange-winged Amazons were observed flying alone n xcr "’’- a" U5.7%), and small numbers in family groups (pairs with young) of three (8.7%), four (1.2%), or five . o) individuals. The larger number of groups of three compared with groups of four and five individuals reflects the low Qnn generally only one surviving offspring per brood. The total number of parrots increased from pnl (J,899) to July (8,539), and began to decrease in August (5,351). This decrease was presumably due to onset of the reeding season, when paired individuals leave the roost in search of a nest, where they breed, nest, and rear youn« until the nestlings can fly. Received 24 January 2009. Accepted 28 August 2009. Temporal and spatial variations in estimates of wild animal populations are important indicators of underlying demographic processes and ecolog- ical interactions (Nunes and Betini 2002). Popu- lation estimates are particularly important in the case of parrots (Aves, Psittacidae) because this family has a disproportionate number of threat- ened species (Collar and Juniper 1992). Censusing over significant periods of time is an efficient tool to distinguish between natural and anthropogenic fluctuations (Bibby et al. 1992), which is essential for future conservation of species. There are different methods to estimate popu- lation size of birds. In the case of parrots, most methods are based on counts of individuals by visual and/or auditory contacts (Collar and Juniper 1992). They are registered in high number early in the morning and late afternoon (Pizo et al. 1997), and counts in the middle of the day should be avoided because activity of parrots decreases abruptly at this time (Marsden 1999). Roost counts may permit targeted assessments of parrot populations in small areas and on islands where parrots roost communally (Snyder et al. 1987, Matuzak and Brightsmith 2007). The Orange-winged Amazon (Amazona ama- zonica) gathers overnight in communal roosts (Sick 1997, Juniper and Parr 1998). Individuals meet just before or shortly after sunset (Sick 1997), making it relatively easy to obtain ' Laboratory of Ornithology and Bioacoustics, Institute of Biological Sciences, Federal University of Para, 66075-1 10 Belem, PA, Brazil. "Laboratory of Bioacoustics, Institute of Biology, State University of Campinas, 13083-970 Campinas, SP. Brazil. ’Corresponding author; e-mail: leiliany(®ufpa.br estimates of the minimum number of individuals (Nunes and Betini 2002). Our objective was to conduct population estimates of Orange-winged Amazons for all months of the year at a roosting site on the Ilha dos Papagaios (Parrot’s Island), near Belem (Para State, Brazil), discriminating family groups (single birds, pairs, and groups of three, four, or five individuals), and measuring seasonal fluctuation throughout the year. METHODS Study Area. — Ilha dos Papagaios is a roosting site for a significant Orange-winged Amazon popula- tion, a common species in the region. It is in Guajara Bay, south ofBelem, Brazil (01° 31' 37" S, 48° 30' 22" W), and has an area of 7.4 ha, eight houses, and about 36 inhabitants. The island is low and subject to periodic tidal flooding (Novaes and Lima 1998). The climate is tropical humid and the source of the seasonality is the precipitation, although there is no pronounced diy season. There are two seasons: a rainy season and a less rainy season. The rainy season begins in December, peaks in March, and subsides in May. January, February, and March are the wettest months with average monthly rainfalls exceeding 400 mm. The less rainy sea.son usually corresponds to the period from June to November, and October is the driest month with an average precipitation of 86 mm (annual precip- itation = 2,800 mm; SECT AM 1994). The vegetation is mostly dense alluvial forest with emergent and uniform canopy in areas of lowland floodplain. Flooded forests, and sandy alluvial deposits (beaches) covered by sparse vegetation, characteristic of Guajara Bay, are also present (Novaes and Lima 1998). 88 de Moiira el cd. • ORANGE-WINGED AMAZON ROOST IN AMAZONIA 89 Surveys. — Counts of A. amazonica were con- ducted at the roost on llha dos Papagaios. We conducted 96 surveys (54 in the afternoon and 42 in the morning) from September 2004 through September 2005. Pan'ots were counted from a single medium-sized boat (12 m long), positioned in a point (about 100 m southeast of the island) where it was possible to clearly observe all the main flight routes of the parrots’ amval and departure to and from the roost. Two visits were made per week, one between 1700 and 1900 hrs and one the following day between 0500 and 0800 hrs, before sunset and sunrise, respectively. During counts, single birds, pairs, and groups of three, four, or five individuals were noted when possible with larger groups being counted or estimated when it was not possible to separate individuals. Counts were made by three teams each with at least one observer and one auxiliary. Observers counted all individuals passing in their field of view, arriving or departing to or from previously selected directions in relation to the roost; (1) to or from the north and northeast of the island; (2) to or from the southeast and south of the island; and (3) parrots going to or arriving from the other side of the island. The teams were careful to not duplicate individuals already counted by them- selves or by another team, eventually announc- ing loudly their counts to the near-by team. Each observer announced to his auxiliary the number and group composition of the parrots seen with the aid of 7 X 50 binoculars; the auxiliaries recorded these data on standardized forms, separating the numbers of birds seen per minute. Statistical Analyses. — Data were analyzed with Statistica 7.1 software package (Stasoft Inc. 2005) and are expressed as means ± SD. Variables were examined for normality using Kolmogorov-Smir- nov tests. All data followed a normal distribution and parametric tests were used (ANOVA and paired r-test). Differences between means were considered statistically significant at P < 0.05. We calculated the percentages of the parrot flocks flying to or from the roost (single birds, pairs and groups of three, four, or five individuals) per visit and averaged the data by month. Cluster analysis was conducted using the City-block (Manhattan) distance method and weighted pair-group aveiage to test for seasonal differences (Sneath and Sokal 1973). RESULTS Patterns of Flight Behavior. — Single individu- als (probably unpaired adults or young) accounted for 14.2 ± 6.7% of the observed groups, pairs (sexually mature adults) 75.7 ± 6.6%, and family groups of three (presumably pairs with 1 young) 8.7 ± 3.3%, groups of four (presumed pairs with 2 young) 1.2 ± 1.6%, and groups of five (presumed pairs with 3 young) 0.3 ± 0.4%. Total Number of Individuals and Seasonal Fluctuation. — The average number of individuals per visit varied by month (Eig. 1). June (7,084 ± 1,708) and July (8,539 ± 704) had the highest number of pan'ots counted at the roosting site (max = 9,603 on 2 July 2005). Total numbers decreased in August (5,351 ± 1,534) and again in September (3,819 ± 462 in 2005 but only 2,040 ± 383 in 2004). The lowest counts were recorded from October (1,545 ± 245) through December (1,364 ± 161), followed by a slightly higher level in January (1,895 ± 223), February (1,778 ± 356), and March (2,361 ± 453), after which the number increased steadily (3,899 ± 547 in April and 4,484 ± 814 in May) until the July peak. Size of Flocks and Seasonal Fluctuation.— Iht number of single individuals, pairs, and groups of three, four, and five individuals varied over the year (Fig. 2). These patterns were similar to those for total numbers, except for groups of four and five individuals, which were essentially absent from September until December. In January 2005 they began to increase, reaching a level of —3.5 and —2.3%, respectively between March and July 2005, and then decreased in August and September 2005. Annual Cycle. — ^Monthly values separated into three major groups (Fig. 3). The first period included September 2004 to March 2005, the second period was April, May, August, and September 2005, and the third period was June and July 2005. There was a significant difference between these periods in total number of paiTots across all Hock sizes (ANOVA E2.92totai ~ 339.96, P < 0.01; /^2.92.single ~ 37.15, P < 0.01; E2,92pairs ~ 204.51, P < 0.01; F2,92ihree = 152.55, P < 0.01; p2.92rour = 36.58, P < 0.01; P2.92f.ve = >8.80, P < 0.01) (Table 1). All comparisons of total number and number of single individuals, pairs, and groups of three, were significant; the averages for period 1 were smaller than those for period 2, which were smaller than those for period 3. Total numbers of parrots and 90 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122. No. I. March 2010 Tf VY WD IT) o o o o O o o o O O o o o o o o o o o o o o o o o o a D 00 Ci o O > o Z u 0.05 >0.05 de Maura ct al. • ORANGE-WINGED AMAZON ROOST IN AMAZONIA 93 1998). The higher number of groups of three (presumably pairs with one young), compared with groups of four and five individuals (pre- sumed pairs with 2 and 3 young, respectively), reflects the characteristically low survival rate of nestling parrots (Gnam and Rockwell 1991, Munn 1992, Guedes 1993, Lindsey et al. 1994, Carrillo et al. 2002, Seixas and Mourao 2002). We observed clutches of 2-4 eggs in five nests, but only one nestling survived in each of the two nests that were not visited for the illegal pet trade (Moura et al. 2008) when we studied reproductive biology of the Orange-winged Amazon during the 2006/2007 breeding season in Santa Barbara (Para State, 41 km from the roosting site). Orange-winged Amazons have a wide distribu- tion and are usually abundant and, at the moment, are not at risk (BirdLife International 2008). However, this species is largely captured from the wild for the international trade in Brazil, being fourth on the list of most apprehended species of parrots by the Brazilian Institute of Environment and Renewable Natural Resources (IB AM A) (Oliveira and Caparroz 2007). It is a popular, although illegal, pet in the region and poaching is intensive. Thus, it is important to monitor Orange- winged Amazon populations for effective conser- vation actions in the future. llha dos Papagaios should be considered a priority area for conservation as it is an important refuge for the parrots. It is close to an urban center (8 km from Belem, population of 2-million inhabitants) and is a convenient site for environ- mental education programs, where people can see a great number of free parrots and become aware that the strong culture of having a wild bird as pet is a crime against nature. ACKNOWLEDGMENTS We thank Ana Paula Assump^ao, Camilo Araujo, Dnilson Ferraz, Eliane Reis, Isabela Brcko, Karine Pereira, Jose Leonardo Magalhaes, Paulo Costa, Renata Emin-Lima and Vitor Lima for assistance with field work; Dr. Jonathan Widdicombe for the revision of an earlier draft; C. E. Braun and the reviewers for providing constructive comments and help in improving the contents of this paper; and PARATUR and CAPES for financial support of this project. LITERATURE CITED Berg, K. S. and R. R. Angel. 2006. Seasonal roosts of Red-lored Amazons in Ecuador provide information about population size and structure. Journal of Field Ornithology 77:95-103. Bibby, C. J., D. N. Burgess, and D. A. Hill. 1992. Bird census techniques. Academic Press, London, United Kingdom. Birdlife International. 2008. Amazona amazonica. Version 2009. 1. lUCN Red List of Threatened Species, www.iucnredlist.org (accessed 5 August 2009). Cannon, C. E. 1984. Flock size of feeding Eastern and Pale-headed rosellas (Aves: Psittaciformes). Austra- lian Wildlife Research 1 1:349-355. Carrillo, A. C., E. A. B. Sipinski, M. L. Cavalheiro, and K. L. Oliveira. 2002. Conserva9ao do Papagaio- de-cara-roxa (Amazona hrasiliensis) no estado do Parana. Pages 193-213 in Ecologia e conservafao de psitacideos no Brasil (M. Galetti and M. A. Pizo, Editors). Melopsittacus Publica96es Cientificas, Belo Horizonte, Brazil. Casagrande, D. G. and S. R. Beissinger. 1997. Evaluation of four methods for estimating parrot population size. Condor 99:445-457. Chapman, C. A., L. J. Chapman, and L. Lefebvre. 1989. Variability in parrot flock size: possible functions of communal roosts. Condor 91:842-847. Collar, N. J. and A. T. Juniper. 1992. Dimensions and causes of the parrot conservation crisis. Pages 1-24 in New World parrots in crisis (S. R. Beissinger and N. F. R. Snyder, Editors). Smithsonian Institution Press, Washington, D.C., USA. Costa, P. C. R. 2006. Comportamento alimentar e dinamica populacional do Periquito— de— asa— branca Brotogeris versicolurus (Aves - Psittacidae) na cidade de Belem, PA. Monograph. Universidade Federal do Pari Belem. Brazil. COUGILL, S. and S. j. Marsden. 2004. Variability in roost size in an Amazona parrot: implications for roost monitoring. Journal of Field Ornithology 75:67—73. Gilardi, j. D. and C. Munn. 1998. Patterns of activity, flocking, and habitat use in parrots of the Peruvian Amazon. Condor 100:641-653. Gnam, R. S. and R. F. Rockwell. 1991. Reproductive potential and output of the Bahama Parrot Amazona leiicocephala hahamensis. Ibis 133:400^05. Guedes, N. M. R. 1993. Biologia reprodutiva de Arara-azul (Anodorhynchus liyacinihinus) no Pantanal-MS. Brasil. Dissertation. Universidade de Sao Paulo. Piracicaba. Brazil. Guedes, N. M. R. and G. H. F. Seixas. 2002. M^todos para estudos de reprodu9ao de Psitacideos, Pages 123-139 in Ecologia e conserva9ao de psitacideos no Brasil (M. Galetti and M. A, Pizo. Editors). Melopsittacus Publica96cs Cientificas, Belo Horizonte. Brazil. Harms, K. E. and J. R. Eberhard. 2003. Roosting behavior of the Brown-throated Parakeet (Aratinga pertinax) and roost locations on four southern Caribbean islands. Ornitologia Neotropical 14:79-89. Juniper, T. and M. Parr. 1998. Parrots: a guide to the parrots of the world. Pica Press, London, United Kingdom. Lindsey, G. D., W. J Arendt, and J. Kalina. 1994. Survival and causes of mortality in juvenile Puerto Rican Parrots. Journal of Field Ornithology 65:76-82. 94 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 1, March 2010 Mabb, K. T. 1997. Roosting behavior of naturalized parrots in the San Gabriel Valley, California. Western Birds 28:202-208. Marsden, S. J. 1999. Estimation of parrot and hornbill densities using a point count distance sampling method. Ibis 141:377-390. Martinez, J. and N. P. Prestes. 2002. Ecologia e conservagao do Papagaio-charao Aniazona pretrei. Pages 173-192 in Ecologia e conservafao de psitaci- deos no Brasil (M. Galetti and M. A. Pizo, Editors). Melopsittacus Publica96es CientiTicas, Belo Hori- zonte, Brazil. Martuscelli, P. 1995. Ecology and conservation of the Red-tailed Amazon (Aniazona hrasiliensis) in south- eastern Brazil. Bird Conservation International 5:405- 420. Matuzak, G. D. and D. j. Brightsmith. 2007. Roosting of Yellow-naped Parrots in Costa Rica: estimating the size and recruitment of threatened populations. Journal of Field Ornithology 78:159-169. Moura, L. N., j. M. E. Vielliard, and M. L. Silva. 2008. Flutuafao populacional e comportamento reprodutivo do Papagaio-do-mangue (Aniazona aniazonica). Pages 223-238 in Biologia da Conserva9ao: estudo de caso com o Papagaio-charao e outros papagaios brasileiros (J. Martinez and N. P. Prestes, Editors). UPF Editora, Passo Fundo, Brazil. Munn, C. a. 1992. Macaw biology and ecotourism, or “When a bird in the bush is worth two in the hand?” Pages 47-72 in New World parrots in crisis (S. R. Beissinger and N. F. R. Snyder, Editors). Smith.sonian Institution Press. Washington, D.C., USA. Novaes, F. C. and M. F. C. Lima. 1998. Aves da Grande Belem: Municfpios de Belem e Ananindeua. Museu Paraense Emilio Goeldi, Belem, Brazil. Nunes, M. F. C. and G. S. Betini. 2002. Metodos de estimativa de abundancia de psitacideos. Pages 99-1 12 in Ecologia e conserva9ao de psitacideos no Brasil (M. Galetti and M. A. Pizo, Editors). Melopsittacus Publica96es CientiTicas, Belo Horizonte, Brazil. Oliveira, T. G. and R. Caparroz. 2007. Analise do impacto do trafico ilegal sobre as popula96es de Psittacidae no Brasil. Page 215 in Annals of the XV Brazilian Ornithological Congress. Porto Alegre. Brazil. Olmos, F. 1997. Parrots of the Caatinga of Piaui, northeastern Brazil. Papageienkunde 1:173-182. Pizo, M. A., 1. Simao, and M. Galetti. 1997. Daily variation in activity and flock size of two parakeet species from southeastern Brazil. Wilson Bulletin 109:343-348. Rocha, C. E. D., H. G. Bergallo, and S. Siciliano. 1988. Migra9ao circadiana em cinco especies de psitacideos em Parintins-AM. Acta Amazonica 18:371-374. SECTAM. 1994. Parque ambiental de Belem: Plano de manejo. Sofrelec Engenharia, Rio de Janeiro, Brazil. Seixas, G. H. F. and G. M. Mourao. 2002. Biologia reprodutiva do Papagaio-verdadeiro (Aniazona aestiva) no Pantanal sul-mato-grossense, Brasil. Pages 157- 171 in Ecologia e conserva9ao de psitacideos no Brasil (M. Galetti and M. A. Pizo, Editors). Melopsittacus Publica96es Cienti'ficas, Belo Horizonte, Brazil. Sick, H. 1997. Ornitologia Brasileira. Nova Fronteira, Rio de Janeiro, Brazil. Sneath, P. H. and R. R. Sokal. 1973. Numerical taxonomy: the principles and practice of numerical classification. W. H. Freeman, San Francisco, Cali- fornia, USA. Snyder, N. F. R.. J. W. Wiley, and C. B. Kepler. 1987. The parrots of Luquilo: natural history and conserva- tion of the Puerto Rican Pan'ot. Western Foundation of Vertebrate Zoology, Los Angeles. California, USA. STATSOFT iNC. 2005. Statistica software. Version 7.1. StatSoft Inc., Tulsa. Oklahoma, USA. The Wilson Joiinial oj' Ornithology 122( 1 ):95-l()l, 2010 COSTS AND BENEFITS OF FORAGING AFONE OR IN MIXED- SPECIES AGGREGATIONS FOR FORSTER’S TERNS LISA SCHREFFLER,' - JOHN K. LEISER,- AND TERRY L. MASTER'"'’ ABSTRACT.— Forster's Terns (Sterna forsteri) often forage alone or within mixed- species toraging aggregations. We investigated the relative costs and benefits of these two foraging strategies. Metrics used to assess diflerences in strategies include^d the number of attempts for prey (both successful and missed), capture efficiency, and capture rate. Forster s Terns foraoinc socially cue on presence and behavior of other aggregating species, especially Snowy Egrets (E^retta tinila). JoinTng^iggregations significantly increased their capture rate, but not their capture efficiency, compared to solitarily foraaing terns. However, total aggregation size, influenced primarily by the two most abundant species, Laughing Gu Is (Laras \itrieilla) and Snowy Egrets, affected tern behavior. For example, diving terns often collided with Laughing Gu s, causing frequent missed attempts for socially foraging individuals. Thus, large aggregation sizes had a negative effect on foraaina success, perhaps because of interference and/or prey depletion. In contrast, solitarily foraging terns dove less often for prey but also missed less often. This may be why terns were observed foraging solitarily as aggregation sizes and conaestion increased. Received 20 February 2009. Accepted 17 July 2009. Important among an individual’s behavioral decisions are those regarding food acquisition. Each individual must decide when and where to forage (Chamov 1976), on what prey item(s) to focus (Schoener 1971, Collier et al. 1991), and what capture strategy to use (Ydenberg et al. 1994). These decisions are constrained by the necessity of energy acquisition and will ultimately influence an animal’s fitness. One option animals can choose to make food acquisition decisions easier is to join a group in a process called local enhancement. Foraging in groups has been shown to provide benefits, including reduction of an individual’s energy intake variance (Caraco 1981), increased prey capture rate (Pulliam and Millikan 1982), reduced energy expenditure (Master et al. 1993), increased vigilance (Caldwell 1986), reduced predation risk from “many-eyes” (Lima 1995) or dilution effects (Hamilton 1971), and the imparting of information about novel foraging techniques (Webster and Lefebvre 2001). Foraging in a group also has disadvantages, including increased conspicuousness to predators (Hamilton 1971, Bertram 1978, Caldwell 1986) and increased energy expenditure on aggressive interactions with competing foragers (Grand and Dill 1999). Some disadvantages of feeding within a group may be assuaged if individuals join mixed rathei than single-species groups to forage. The com- ' Department of Biological Sciences, East Stroudsburg University, East Stroudsburg, PA 18301, USA. ^Biology Department, Northampton Community Col- lege, Monroe Campus. Tannersville, PA 18372, USA. ^Corresponding author; e-mail: tmaster(®po-box. esu.edu petitive costs of food acquisition are, at times, reduced by species-specific preferences for prey items in mixed-species feeding aggregations among wading birds (Moreno et al. 2005). Competitive costs are also moderated by differ- ences in preferred feeding locations (Moreno et al. 2005) and by reduced interference due to variation in prey catching techniques (Latta and Wunderle 1996). Few studies of Forster’s Tern (Sterna forsteri) foraging behavior have been conducted (but see Salt and Willard 1971, Reed 1985). We investi- gated tern feeding behavior within mixed-species aggregations of wading birds and while foraging solitarily in a salt marsh in southern New Jersey, USA. Our goals were to: ( 1 ) identify the role of terns in formation and structure of mixed-species aggregations that progress from an initial for- mation phase” with few species and individuals to an active phase with more species and individuals, and finally to a dispersal phase (Master et al. 1993); and (2) investigate foraging tactics used by terns under changing social conditions by comparing benefits accrued versus costs incurred by feeding within aggregations and solitarily. METHODS Study Area. — ^This study was conducted over two summer seasons (2004—2005) in a 2,083-ha salt marsh in southeastern New Jersey, USA (~39° 3' 74° 46' W). This marsh provided an unusually large number of relatively shallow (~3_60 cm) pools in which terns forage (Master 1992). The majority of fish prey within the pools were mummichogs {Funduhis heteroclitus) and 95 96 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 1, March 2010 sheepshead minnows {Cyprinodon variegatus); prey density averaged (± SD) 147.15 ± 57.92 fish/nr in the pools (Master 1992). The popula- tion of Forster’s Terns in southern New Jersey was estimated at 403 individuals in the late 1970s and early 1980s (Spendelow and Patton 1988). At least eight separate colonies were within Cape May County more recently although numbers of birds were not reported (Walsh et al. 1999). F oraging Observations. — ^Terns were observed trom June until mid-August during each field season. Both solitary and aggregated terns were observed using binoculars (Nikon 10 X 40) or a spotting scope (KOWA TSN 821, 20-60X zoom), depending on the subject’s proximity to the observer. We defined a mixed-species aggregation as a group of at least four birds from at least two different species foraging within close proximity (~1 m) to one another at a single pool. The species most often composing aggregations, in addition to terns, were Great Egret {Ardea alba). Snowy Egret (Egretta thida). Glossy Ibis {Plega- dis folcinellus). Great Blue Heron (Ardea lier- odias). Laughing Gull (Larus atricilla), and Least Tern (Sterna antillarinm). Aggregations ranged in size from four to 430 individuals (mean aggrega- tion size ± SE: 93.74 ± 4.70, n = 87). Observations began each morning at sunrise as aggregations were forming and continued until — 1130 hrs EDT, when nearly all aggregations had dispersed. An aggregation was chosen randomly at sunrise and the total number of species represented and number of individuals of each species were counted. Numbers of individuals were summed to calculate total aggregation size. A single Eorster’s Tern was selected for observation, after the initial count, using the focal animal sampling technique (Schmitz and Baldassarre 1992). The focal tern was observed for 7 min (Balph and Romesburg 1986, Rooney and Sleeman 1998) during which the number of successful and missed foraging attempts was recorded. Capture rate and capture efficiency were calculated from these observations by dividing the number of successful attempts by the observa- tion time of 7 min, and by dividing the number of successful attempts by the total number of attempts, both successful and missed, respectively. A 10-min intermission followed observation of a focal individual prior to the next 7-min ob.servation to allow terns to re-distribute them- selves throughout the aggregation. After the intermission, a .second count of the participants in the aggregation was recorded, another tern was selected, and a 7-min focal observation was begun. Our protocol continued in this manner for 58 min, providing four separate focal tern observations and species counts per aggregation (3 10-min intermissions between 4 7-min obser- vation periods). Additional aggregations were located for continued observation until the — 1130 hrs dispersion time. The Schmitz and Baldassame (1992) sampling procedure does not ensure an individual tern is observed only once, but this procedure is well accepted when studying aggregating animals and is resistant to potential sampling biases (Weinstein 1995). There were 38 aggregations studied in 2004 and 49 in 2005. Observations were conducted on 275 focal terns (126 in 2004 and 149 in 2005). Observations were also conducted on terns foraging solitarily. These observations followed a similar protocol with two exceptions. Eirst, aggregation counts were unnecessary. Second, solitary terns were observed for as long as they were visible. Nineteen solitary terns were ob- served with a mean (± SD) observation time of 7.00 ± 1.09 min. Total attempts, capture rates, and capture efficiencies were calculated on a per 7-min basis for comparison with aggregating terns. Solitary terns were observed during the same morning period as aggregations (sunrise to -1130 hrs). Data Analyses. — ^Each focal tern and the coire- sponding aggregation survey were treated as independent samples throughout the data analyses, (n — 275), rather than considering each aggrega- tion with four individual terns as an independent sample (n = 87). This methodology was justified as intermission periods between observations allowed terns to redistribute themselves randomly throughout the space above a pool providing for Linbia.sed observation of foraging birds. Observa- tions of focal individuals can be considered independent it the time lag between observations is sullicient for an animal to traverse its range (Swihart and Slade 1985, Minta 1992). The 10-min interval between tern observations was sufficient tor a tern to traverse the space above a pool. This also provided us with a preci.se time of day and aggregation size for each focal bird. Analyses using each tern and corresponding suiwey as independent samples were compared to analyses in which one tern and one survey from each aggregation were randomly .selected for analysis. The results of these random selections were similar in all cases with those using all four focal individuals from a single Schreffler cl at. • AGGREGATIONS OF FORSTER'S TERNS 97 FIG. 1. Numbers of Laughing Gulls (open pyramids, dotted line). Snowy Egrets (shaded diamonds, dashed li.^e) and Forster's Terns (dark boxes, solid line) by aggregation size. The number ot gulls (slope - 0.48) and egrets (slope 0.17) increased faster than the number of terns (slope = 0.06) as group size increased. aggregation (Pearson product-moment coirelation, y = 0.99). Observations were separated into four time intervals: dawn until 0630, 0631-0730, 0731- 0830, and 0831 hrs EDT until dispersal. Group sizes were divided into four categories: solitai^ terns; and aggregations with 4 to 50, 5 1 to 100, and >100 birds. The data were then analyzed using two-factor completely randomized ANOVAs. RESULTS Aggregation Characteristics. — ^Eorster’s Terns were often observed foraging within aggregations. Laughing Gulls (44% of the total birds within a group). Snowy Egrets (15%), and terns (12%) were the three most abundant species, respective- ly, in these groups. Total aggregation size was most significantly influenced by number of Laughing Gulls ((3 = 0.60), followed by Snowy Egrets (P = 0.38), and Forster’s Terns (P = 0.20; stepwise multiple regression: T3.272 “ 1693.95, P < 0.0001, r = 0.95; Fig. 1 ). The number of terns in aggregations increa.sed with total aggregation size, but number of terns in these groups increased more slowly compared to the number ot gulls and egrets. Thus, the proportion of total aggregation size comprised by terns decreased with increasing group size (/^i,274 “ 33.93, P < 0.01 , r = 0. 1 1, P = -0.33). The number of terns participating in feeding aggregations was influenced not only by total aggregation size, but by time of day (one-way ANOVA on tern number across the 4 time intervals: F3,272 = 2.964, P = 0.03). Terns were most numerous in aggregations during the early morning periods from sunrise until 0730 his (mean number of terns in aggregations ± SE: 10.58 ± 1.17 from sunrise to 0630, 10.11 ± 1.01 from 0631-0730) with mean numbers declining significantly throughout the remainder of the morning to 7.72 ± 0.75 during the 0731- 0830 period, and to 5.73 ± 0.99 during the 0831 -dispersal time period (ad hoc linear comparison within the ANOVA: Fi,272 “ 7.798, P < 0.01). The contribution that terns made to aggrega- tions was most substantial during the formative period of the aggregation but then declined (one- way ANOVA: ^3,272 = 2.55, P = 0.05). Forster's Terns comprised a high percentage of aggrega- tions initially (18.43 ± 2.48%), followed by a decline for the rest of the aggregation's existence (12.44 ± 1.44% for 0631-0730; 12.43 ± 1.20% for 0731-0830; 13.66 ± 0.38% for 083 1 -dispers- al). Rather than a linear decline (ad hoc linear comparison within the ANOVA: F\,2ii ~ 188, P = 0.17), the percentage of terns in aggregations was initially high and then decreased to a relatively stable lower level until aggregation dissolution (comparison of initial vs. other 3 intervals: F| 272 6.38, P = 0.01). 98 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. I. March 2010 Dawn - 0630 0631 - 0730 0731 - 0830 0831 - Dispersal Time of day FIG. 2. Mean (± SE) number of (A) missed and (B) successful attempts over time by Forster's Terns foraging solitarily (dark diamonds, .solid line) and within aggregations of: 4 to 50 birds (open boxes, dashed line), 51 to 100 birds (open pyramids, dotted line), and >100 birds (open diamonds, variably dashed line). Foraging Strategies. — Both aggregated and solitary terns frequently dove into pools to capture prey; however, many of their attempts mis.sed. Solitary birds tended to make fewer dives overall than aggregated birds and therefore missed prey less often than aggregated birds (aggregation size effect within the two-way ANOVA: = 4.078, P = 0.007; Fig. 2A). The number of missed attempts did not change significantly for all birds over time (time of day effect within the two-way ANOVA: T3.273 = 0.500, P = 0.687), but there was a significant interaction between aggregation size and time of day (F3 273 = 1.928, P = 0.048). Solitary birds tended to miss capture of prey less than aggregated birds. Terns in small (4 to 50 total individuals) to moderate-sized (5 1 to 100) groups mis.sed more often than solitary birds but less often than birds in large aggregations (>100 birds) and missed more often early in the morning than later (linear trend comparison within the ANOVA: F1.273 = 8.562, P = 0.004). Terns in large congested aggregations frequently missed prey continuing throughout the morning. Occasional collisions with competitors, Schrcfficr cl al. • AGGREGATIONS OF FORSTER’S TERNS 99 especially Laughing Gulls, may have been al least partially responsible for misses by terns in aggregations. The number of times that terns were successful in capturing prey during dives was also influenced by group size (aggregation size effect within the two-way ANOVA: ^3273 ~ 3.239, P = 0.023; Fig. 2B). Solitary terns overall dove relatively infrequently and had the fewest successful dives. Terns in small or moderate-sized aggregations were most successful, while terns in large aggregations made comparatively few successful dives (quadratic trend comparison of group size within the ANOVA: F] 273 ^ 4.673, P — 0.031). The number of successful dives by terns in aggregations appeared to decline throughout the morning, but the trend was not significant (time of day effect within the two-way ANOVA: F3273 ~ 1.234, P = 0.298), nor was the interaction between time of day and group size {Fg.m “ 0.430, P = 0.918). Terns in aggregations, as a result of diving more frequently, had higher capture rates than solitary terns (aggregation size effect within the two-way ANOVA: ^3,273 = 3.197, P = 0.024; Fig. 3A). However, aggregated terns had fewer successful dives later in the day, causing their capture rates to decline throughout the morning (linear trend comparison within the ANOVA: = 9.341, P = 0.002), while capture rates for solitary birds did not (time of day effect within the two-way ANOVA: ^3 273 1.463, P = 0.225; interaction between time of day and group size: F9.273 = 0.338, P = 0.962). Capture rates differed for the two strategies, but capture efficiency was similar for aggregated and solitarily foraging terns (aggregation size effect within the two-way ANOVA: F3 273 ~ 1.414, f* = 0.239; Fig. 3B). The capture efficiency for solitary birds appeared more variable than for aggregated birds throughout the morning, but there were no significant differences in either lime of day for all birds (F3 273 = 2.225, P = 0.086) or the interaction between group size and time (F9.273 = 1.478, P = 0.156). DISCUSSION Forster’s Terns were frequent participants within aggregations, contributing significantly to the overall size of these mixed-species groups. Terns were less numerous in aggregations than either Snowy Egrets or Laughing Gulls and, as mornings progressed and more ol these hi ids joined aggregations, the terns left. These observations likely reflect the foraging strategy of terns relative to wading biids that comprised these large, congested aggregations. The pre.sence of birds at a pool during the early morning lormative period indicated to seaiching terns that it was a worthwhile patch to exploit (Armstrong 1970, Master 1992). Foraging maneuvers of wading birds, especially neai the shoreline of pools, serve to flush prey into open water (Kushlan 1978). This maneuvering, particularly by the actively foraging Snowy Egrets, has been shown to enhance capture opportunities for other wading birds in the group (Russell 1978), and apparently for Forster’s Terns. Terns used a “hover-hunt” technique over the center of the pool, after prey had been flushed. In this behavior, terns would actively beat their wings to remain stationary in the air ~7 to 10 m above the center of the pool’s surface. A tern would then dive into the water to capture a fish and, if successful, would fly across the pool while swallowing its prey until returning to its hovering position. If unsuccessful, the bird would return to its hovering position. Hover hunting was most successful for terns early in the morning, when both their number of successful dive attempts and rates of prey capture were highest. These results seem intuitive as, during this time of the morning, prey has been beaten from cover but are still abundant while competing foragers are still low in number. The total number and proportion of terns in aggregations increased during the lormative period reflecting good foraging conditions. This technique became less successful as time pro- gressed and aggregations grew in size, prey density declined and other birds became obstacles during diving, reducing both the terns’ number ot successful attempts and capture rates. Previous studies have shown that foraging activity of aggregations does impact pool fish populations by reducing density by 80% on average (Master 1992). Solitary terns, in contrast to aggregated terns, used a “circle hunt” behavior consisting of flying a circle around the perimeter of the pool al a relatively low height of 1 to 3 m. The tern would dive into the water after spotting a prey item. Whether successful or not, the bird would rise from the water and continue its circling flight until diving again or leaving the pool. This behavior likely flushed some prey items, but 100 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 1. March 2010 Time of day FIG. 3. Mean ( ± SE) (A) capture rate and (B) capture efficiency over time for Forster’s Terns foraging solitarily (dark diamonds, solid line) and within aggregations of: 4 to 50 birds (open boxes, dashed line), 51 to 100 birds (open pyramids, dotted line), and >100 birds (open diamonds, variably dashed line). probably not a.s many as the combined effects of multiple egrets and gulls and, thus, solitary terns dove less often for prey items, which ultimately reduced their capture rates. However, solitary birds also missed less frequently, likely due to lack of interference from waders and gulls, and their capture efficiency was similar to that of aggregated terns. Ultimately, the tactics of Forster’s Terns reflect the apparent advantages and disadvantages of aggregated versus solitary foraging over the course of the morning. Terns Hew over the salt marsh. encountering multiple pools that either were or were not occupied by an aggregation. Terns could easily join a group or forage alone,, or switch between the two tactics. Incentives for Joining a group were high as long as the aggregation was not so large that competition was high and prey items had been depleted, conditions that seemed to exist during the early morning. Terns foraging alone had fewer opportunities to capture prey, but they were equally likely to catch fish throughout the day as their prey had not been over-exploited and they had far le.ss competition. Schrcfflcr ei at. • AGGREGATIONS OF FORSTER’S TERNS 101 ACKNOWLEDGMENTS We thank Roger Wood and The Wetlands Institute, Stone Harbor, New Jersey for generously providing logistical support throughout this project, the Biology Departments ot East Stroudsburg University of Pennsylvania and North- ampton Community College for providing technical support, the East Stroudsburg University Foundation for providing travel funds, and two anonymous reviewers for helpful suggestions which greatly improved the manuscript. LITERATURE CITED Armstrong, E. A. 1970. Social signaling and white plumage. Ibis 113:534. B.^LPH, D. F. AND H. C. Romesburg. 1986. The possible impact of observer bias on some avian research. Auk 103:831. Bertram, B. C. R. 1978. Living in groups: predators and prey. Pages 64-96 in Behavioural ecology: an evolutionary approach (J. R. Krebs and N. B. Davies, Editors). Blackwell Scientific Publishers, Oxford, United Kingdom; and Boston, Massachusetts, USA. Caldwell, G. S. 1986. Predation as a selective force on foraging herons: effects of plumage color and flocking. Auk 103:494-505. Caraco, T. 1981. Risk-sensitivity and foraging groups. Ecology 62:527-531. Charnov, E. L. 1976. Optimal foraging: the marginal value theorem. Theoretical Population Biology 9:129-136. Collier, G., D. F. Johnson, and J. Berman. 1991. Patch choice as a function of procurement cost and encounter rate. Journal of the Experimental Analysis of Behav- iour 69:5-16. Grand, T. C. and L. M. Dill. 1999. The effect of group size on the foraging behaviour of Juvenile coho salmon: reduction of predation risk or increased competition? Animal Behaviour 58:443-451. Hamilton, W. D. 1971. Geometry for the selfish herd. Journal of Theoretical Biology 31:295-311. Kushlan, j. A. 1978. Feeding ecology of wading birds. Pages 249-297 in Wading birds (A. Sprunt IV, J. C. Ogden, and S. Winckler, Editors). Research Report Number 7. National Audubon Society, New York, USA. Latta, S. C. and J. M. WUNDERLE. 1996. The composition and foraging ecology ot mixed-species flocks in pine forests of Hispaniola. Condor 98:595-607. Lim.a, S. L. 1995. Back to the basics of anti-predatory vigilance: the group-size effect. Animal Behaviour 49:11-20. Master, T. L. 1992. Composition, structure, and dynamics of mixed-species foraging aggregations in a southern New Jersey salt marsh. Waterbirds 15:66-74. Master, T. L., M. Frankel, and M. Russell. 1993. Benefits of foraging in mixed-species wader aggrega- tions in a southern New Jersey salt marsh. Waterbirds 16:149-157. Minta, S. 1992. Tests of spatial and temporal interaction among animals. Ecological Applications 2:178—188. Moreno, A. B., A. R. Lagos, and M. A. Alves. 2005. Water depth selection during foraging and efficiency in prey capture by the egrets Casmerodiiis alhits and Egrelta tiuilci (Aves, Ardeidae) in an urban lagoon in Rio de Janeiro State, Brazil. Iheringia. Serie Zoologia 95:107-109. Pulliam, H. R. and G. C. Millikan. 1982. Social organization in the non-reproductive season. Pages 1 17-132 in Avian biology. Volume 6 (D. S. Earner and J. R. King, Editors). Academic Press, New York, USA. Reed, J. M. 1985. Relative energetics of two foraging behaviors of Forster’s Terns. Waterbirds 8:79-82. Rooney, M. B. and J. Sleeman. 1998. Effects of selected behavioral enrichment devices on behavior of western lowland gorillas (Gorilla gorilla gorilla). Journal of Applied Animal Welfare Science 1:339-351. Russell, J. 1978. Effects of interspecific dominance among egrets commensally following Roseate Spoonbills. Auk 95:608-610. Salt, G. W. and D. E. Willard. 1971. The hunting behavior and success of Forster’s Tem. Ecology 52:989-998. Schmitz, R. A. and G. A. Baldassarre. 1992. Correlates of flock size and behavior of foraging American Flamingos following Hurricane Gilbert in Yucatan. Mexico. Condor 94:260-264. Schoener, T. W. 1971. Theory of feeding strategies. Annual Review of Ecology and Systematics 2:368- 404. Spendelow, j. a. and S. R. Patton. 1988. National atlas of coastal waterbird colonies in the contiguous United States: 1976-1982. U.S. Fish and Wildlife Service Biological Report 88 (5). Washington, D.C., USA. SwiHART, R. K. AND N. A. Slade. 1985. Testing for independence of observations in animal movements. Ecology 66: 1176-1184. Walsh. J., V. Elia. R. Kane, and T. Halliwell. 1999. Birds of New Jersey. New Jersey Audubon Society. Barnardsville, USA. Webster, S. J. and L. Lefebvre. 2001. Problem solving and neophobia in a columbiform-passeriform assem- blage in Barbados. Animal Behaviour 62:23-32. Weinstein. R. B. 1995. Locomotor behavior of nocturnal ghost crabs on the beach: focal animal sampling and instantaneous velocity from three-dimensional motion analysis. Journal of Experimental Biology 198:989- 999. Ydenberg. R. C., C. V. J. Welham, R. Schmid-Hempel, P. Schmid-Hempel, and G. Beauchamp. 1994. Time and energy constraints and the relationships between currencies in foraging theory. Behavioral Ecology 5:28-34. The IVilson Journal of OmUhology 122( 1 ): 102-1 1 5, 2010 ABUNDANCE AND DISTRIBUTION OF WATERBIRDS IN THE LLANOS OF VENEZUELA FRANCISCO J. VILELLA'-^ AND GUY A. BALDASSARRE- ABSTRACT. — The Llanos is a significant waterbird site in the Western Hemisphere, but abundance and distribution of waterbirds across this vast region are poorly known, which hampers conservation initiatives. We used point counts along road routes in the Llanos region of Venezuela to examine abundance and distribution of waterbirds during 2000-2002 within five ecoregions across the Llanos. We detected 69 species of waterbirds and recorded 283,566 individuals, of which 10 species accounted for 80% of our observations. Wading birds (Ciconiifomies) represented the largest guild both in numbers of species (26) and individuals (55%), followed by waterfowl (26%), and shorebirds (11%). Five species comprised 62% of all individuals: Cattle Egret {Biihiilciis ibis). White-faced Whistling Duck (Dendrocygna vidiiata). Black- bellied Whistling Duck (D. ainiimnalis). Great Egret (Ardea alba), and Wattled Jacana (Jacana jacana). Wading birds were particularly ubiquitous with at least 21 ot 26 species recorded in each ot the ecoregions. Species richness (66), proportion of waterbirds detected (54%), and mean number of birds per route (1,459) were highest in the Banco-Bajio-Estero savanna ecoregion. Our study provides the most comprehensive data set available on waterbirds in the Llanos of Venezuela and highlights regions of special conservation concern. Received 20 April 2009. Accepted 7 October 2009. The Llanos is the second largest savanna ecosystem in South America after the Cerrado of Brazil, covering some 451,474 km^ of which two- thirds occur in Venezuela and the remainder in Colombia (Bibby et al. 1992, Mittermeier et al. 1998). Wetlands in the Llanos are concentrated in southwest Venezuela and eastern Colombia and can cover an estimated 107,530 km^ which makes the Llanos the second largest wetland complex in South America after the Brazilian Pantanal (Hamilton et al. 2002). This vast expan.se of wetlands, along with poor soils for agriculture and low human population density, combine to make the Llanos one of the most significant waterbird habitats in the world (Scott and Carbonell 1986, Roca et al. 1997, Mittermeier et al. 1998). Over 80 species of waterbirds have been recorded in Llanos wet- lands, of which the major groups are wading birds, sandpipers and plovers, waterfowl, and rails (Hilty 2003). Waterbird occurrence in the Llanos has been documented and general range maps are available in field guides (Hilty 2003), but detailed information on abundance and distribution is lacking and often dated, sparse, or limited in scope and duration (e.g., Ramo and Busto 1984, Frederick and Bildstein 1992, Tamisier and Dehorter 2000). ' U.S. Geological .Survey. Cooperative Re.search Unit, Department of Wildlife and Fisheries, Mississippi State University. Mississippi State. MS 39762. USA. LSUNY College of Environmental .Science and Fore.stry, Syracuse. NY 13210, USA. ’Corresponding author: e-mail: fvilella@cfr.msstate.edu Little is known about abundance and distribu- tion patterns of waterbirds in the Llanos despite its huge size and significance as waterbird habitat (Hilty 2003). This lack of information is note- worthy because the Llanos, after many years of relative isolation, is experiencing increased hu- man population growth and accompanying distur- bance including road construction, water control activities, and hunting, all of which can affect waterbird populations (Bisbal 1988, Mittermeier et al. 1998). Lack of information on waterbird populations in the Llanos seriously hampers informed conservation decisions. Our objective was to conduct a geographically extensive survey to examine the abundance and distribution of waterbirds across this vast wetland savanna landscape. METHODS Study Area. — ^The Llanos is a vast sedimentary plain bordering the Orinoco River and its tributar- ies. Average annual temperature exceeds 24° C and rainfall varies from 500 to 4,000 mm/year (Sarmiento 1983, Mittermeier et al. 1998, Sar- miento and Pinillos 2001). Precipitation is highly sea.sonal (May-Sep) with 23% of the Llanos floodplain susceptible to .sea.sonal Hooding at maximum inundation. In contrast, only 2,080 knr of open water remain during the dry season (Nov- Mar), principally in rivers and lakes (Hamilton et al. 2002). This interaction of .seasonal rainfall, soil quality, topography, high temperatures, and fire promotes savanna vegetation across much of the Llanos, regardless of seasonal inundation. Moist gallery forest, dry tropical forest, permanent 102 Vilcllci and Baldassarrc • LLANOS WA TIiRBIRDS 103 wetlands, and agriculture occur throughout the Llanos and contribute to a complex mosaic ot habitats across the landscape (Sarmiento 1983, Huber and Alarcon 1988). Vegetation and Ecoregions. — Several studies have delimited Llanos ecoregions based on vegetation, flooding, and soil type, but these have not produced universally accepted detinitions or boundaries (Blydenstein 1967, Ramia 1967, Sarmiento 1983, Velasco and Ayarzaguena 1985, Huber and Alarcon 1988, Bulla et al. 1990). We reviewed the existing classifications and selected three well-recognized ecoregions containing most of the waterbird biodiversity in the Llanos: Alluvial Overflow Plains, Western Llanos, and Central Llanos. The Alluvial Overflow Plains include 7.45 mil- lion ha of flat savanna grassland with gallery forest along permanent rivers (Ramia 1967, Velasco and Ayarzaguena 1985). The region is subdivided into two areas based on topography and vegetation: (1) Paspalum Savanna, and (2) Banco-Bajio-Estero Savanna (Ramia 1967, Bulla et al. 1990). Paspalum Savanna occurs in low terrain subject to extensive flooding by rivers and is dominated by Paspalum fasciculatum. Banco- Bajio-Estero Savanna (hereafter, BBE) borders Paspalum Savanna (hereafter, Paspalum), but topographical variations of 1—2 m create consis- tent differences in inundation patterns. This savanna is recognized as “banco-bajio-estero”; literally bank, slope, and marsh. Esteros are the lowest areas, bancos are higher areas of former riverbanks, and bajios are intermediate sites. Esteros comprise —80% of the region (Sarmiento 1983, Bulla et al. 1990). The BBE includes an area of about 1 million ha of man-made impound- ments (modulos). These impoundments increase maximum flood level of the savannas an average of 50 cm and prolong inundation by nearly 3 months (Bulla et al. 1990). The Western Llanos ecoregion (5.15 million ha) is characterized by fertile soils and partial flooding (Huber and Alarcon 1988, Ramo and Busto 1988). Historically this region was deciduous tropical forest interspersed by savan- na; the remaining forest fragments are now largely in reserves (Veillon 1976). Irrigation has allowed this area to become the largest rice- producing region in Venezuela (Ministerio de Agricultura y Cna 1998). We divided our sampling in this ecoregion into (1) Ricefields and (2) Western Llanos. The Central Llanos ecoregion (7.15 million ha) consists ot rolling grasslands dominated by Traclivpogon spp. This ecoregion includes areas of chaparral dominated by the tree species Curatella ainericana, Bowdichia virgilioides, and Byrsonima crassifolia. Gallery lorest, Moriche palm {Maiuitia flexuosa) swamps, and patches of the Orinoco palm (Copernicia tectornm) occur where standing water persists; patches ol tropical deciduous forest occurs at higher elevation (Blydenstein 1962, Ramia 1993, Huber and Alarcon 1988, Ramo and Busto 1988). Rice fields have expanded into the Central Llanos (Ministerio de Agricultura y Cria 1998). Waterbird Surveys. — We used point counts along roadsides (Eig. 1 ) to survey waterbirds because roadside surveys can efficiently assess community composition and habitat associations of birds in open habitats and habitat mosaics across large landscapes (Thiollay 1978, Ellis et al. 1990, Hanowski and Niemi 1995). Detection of waterbird species in open wetlands may actually be greater from roads than off-road (Hanowski and Niemi 1995). Conway and Simon (2003) also noted that point counts along transects increase detection of less conspicuous species. Disadvantages of road count methods are recognized (Buckland et al. 2004), but we believe these limitations were not significant in our study because wetland habitats in the Llanos occur continuously along both roadsides and roadless areas. We sampled each ecoregion based on rainfall patterns that occur in four distinct seasons, each about 3 months in duration: (1) Late Wet, July- September; (2) Early Dry, October-December; (3) Late Dry, January-March; and (4) Early Wet. April-June (Bulla et al. 1990). The hydrological extremes of these seasons were the Late Wet. which is characterized by heavy raintall and maximum annual inundation, whereas the Late Dry is characterized by drought conditions and minimal inundation. Ecoregions were visited in the same order each season (Western. Ricefields. Central, Paspalum, and BBE), and routes were surveyed within 7 days of the calendar date surveyed each year. An initial sample of 30 routes established in July 2000 was increased to 54 by year's end as our familiarity with the Llanos improved (Table 1). Each route was 22.5 km with 16 point-count locations spaced at L5-km intervals. A road was considered for route placement if at least 30 km was accessible year-round, and it was separated 104 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 1. March 2010 Caribbean Sea Q River •oo s Kilometers FIG. 1 . Area of the Llanos of Venezuela surveyed for waterbirds. State boundaries, major cities, and route locations as referenced by GPS are included. from another route by at least 5 km. The first point-count location was randomly selected with- in the first 5 km of a ,selected road. We used a Garmin 12 Global Positioning System (GPS) unit to identify our first point location. Road traffic was light to none and we did not believe traffic affected bird detection. Our point count methodology was based on the North American Breeding Bird Survey (Robbins et al. 1986, Droege 1990, Peterjohn 1994) as modified by local conditions. Counts began shortly after sunrise and were usually completed by noon. A two-person team recorded all waterbirds seen or heard for 3 min within a 500- TABLE 1. Number of route.s sampled per season in each ecoregion of the Llanos, Venezuela. Ecoregion Fieldwork periods Sea.son Ceniral Weslem Llanos Alluvial Overllow Plains Totals Central Ricefields Western Paspalum BBE 14 Aug-3() Sep 2000 Late Wet 3 6 6 8 7 30 1 Nov- 15 Dec 2000 Early Dry 3 6 7 10 7 33 14 Feb-31 Mar 2001 Late Dry 4 6 7 12 14 43 14 May-30 Jun 2001 Early Wet 4 8 14 14 14 54 14 Aug-30 Sep 2001 Wet 4 8 14 14 14 54 1 Nov- 15 Dec 2001 Early Dry 4 8 14 14 14 54 14 Feb-31 Mar 2002 Dry 4 8 14 14 14 54 14 May-30 Jun 2002 Early Wet 4 8 14 14 14 54 I 'ilclla and Balda.s.sarre • LLANOS WAIERBIRDS 105 m radius on each side of the road (6 min/point) as measured by rangefinder binoculars; we included birds in flight but tried to minimize double- counting. We selected the observation duration based on a January 2000 pilot study, and all observers were trained in species identification before surveys began. Observations were made using 10 X 42 binoculars and 20X spotting scopes. Common and scientific names follow Gill et al. (2009). Data Analysis. — We selected 69 waterbird species to survey and classified these into six guilds based on feeding niche, habitat preference, and taxonomy: (1) wading birds, (2) waterfowl, (3) shorebirds, (4) open water, (5) forest, and (6) passerines. The Open Water guild comprised 13 piscivorous species requiring deep water for feeding: Least Grebe (Tachybaptiis dominicus). Neotropic Cormorant {Phalacrocorax hrasilia- nus), Anhinga {Anhinga anhinga), Sungrebe (Heliovnis fulica), terns (Laridae, 4 species); and kingfishers (Alcedinidae, 5 species). The Wading Birds guild (26 species) comprised long-legged species that occur in or close to water and forage on fish, amphibians, and/or invertebrates. These included herons, egrets, and bitterns (Ardeidae, 14 species), ibises (Threskiornithidae, 8 species), storks (Ciconiidae, 3 species), and Limpkin (Aramus guaraima). The Waterfowl guild (10 species) included ducks (Anatidae, 8 species) and gallinules (Rallidae, 2 species). The Shorebird guild (12 species) included sandpipers (Scolopa- cidae, 8 species) as well as Double-striped Thick- knee (Burhinus histriatus). Collared Plover (Char- adriiis collaris). Black-necked Stilt (Himantopus mexicanii.s), and Wattled Jacana (Jacana jacaiia). The Forest guild (four species) included Hoatzin {Opisthocomus hoazin), Grey-necked Wood Rail {Aramides cajanea), Sunbittern (Eurypyga helias), and Homed Screamer (Anhima corniita). The Passerine guild included species associated with water: Pied Water Tyrant (Fluvicola pica). White- headed Marsh Tyrant (Arundinicoia ieitcoce- phala). Black-capped Donacobius {Donacohius atricapilia), and Red-capped Cardinal (Paroaria gidaris). We assessed general patterns of waterbird abundance by summarizing the total numbers of birds recorded by species and guild, ranking each in order of abundance, and calculating percent relative abundance for each species. We also expressed abundance as the percentage of routes and points where we detected each species. We TABLE 2. Number of species and total number of birds recorded for each guild in the Llanos, Venezuela. Guild # .species % lolal species Total.s % total Wading birds 26 37.7 154,450 54.5 Waterfowl 10 14.5 74,335 26.2 Shorebirds 12 17.4 31,277 11.0 Open water 13 18.8 17,286 6.1 Forest 4 5.8 3,280 1.2 Passerines 4 5.8 2,938 1.0 Totals 69 100.0 283,566 100.0 standardized for sampling effort in each ecoregion by calculating the mean number of each species per route per field season. We averaged mean values per sample period to derive a grand mean per route for each species or guild in each ecoregion. We used a species accumulation curve to evaluate the relationship between sampling effort and species detection over the survey period (Colwell 2006). We grouped data from both years and used t- tests to compare species abundance in each ecoregion and season between years; only 3.2% of the tests were significant and differences were distributed among all seasons and ecoregions. We calculated the mean number of individuals per route to combine data between years when routes were sampled in both fieldwork periods, and retained those routes added in the second period or near the end of the first. This combination standardized sample effort in each ecoregion. We compared seasonal differences in number of birds by guild in each ecoregion using analysis of variance. We used SPSS 11.0.0 for all statistical analysis, and accepted significance at the 0.05 level (SPSS 2001). RESULTS General Patterns. — We recorded 283,566 indi- viduals from 69 species of waterbirds in 23 families (Appendix). The waterbird community was dom- inated by wading birds, which comprised 54.5% of the individuals and 37.7% oi the species (Table 2). Waterfowl were second in number of individuals (26.2%), followed by shorebirds (1 1.0%). Howev- er, waterfowl only represented 14.5% of all species. The Open Water guild ranked second in number of species ( 1 8.8%) but fourth in abundance (6.1%). The Passerine and Forest guilds contained the fewest species (4) and individuals (1.0-1. 2%). The species accumulation curve indicated sam- pling effort over the 2-year survey period was 106 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. I, March 2010 FIG. 2. Species accumulation curve for waterbird species detected during 2000-2002 surveys in the Llanos of Venezuela. adequate for detecting most species in the study area (Fig. 2). Approximately 10 additional routes, surveyed seasonally for 2 years, would be needed to detect one additional species. Waterbird abundance in the Llanos was dom- inated by a small number of species; most species occurred infrequently. Only five species com- prised 62% of all individuals recorded: Cattle Egret (Buhulcus ihis). White-faced Whistling Duck (Dencirocygiia viditata). Black-bellied Whistling Duck (D. autumnalis). Great Egret {Arclea alha), and Wattled Jacana. Ten species comprised 80% of all individuals recorded, and 17 species comprised 90%. The Cattle Egret was the most abundant species (23.7%) and was at 97.5% of transect sample points. The second most abundant species was the White-faced Whistling Duck (14.7%), followed by the Black-bellied Whistling Duck (10.1%). The three most abun- dant species collectively accounted for 48.5% of all detections. Great Egret (8.0%) and Wattled Jacana (5.9%) were much less common. We recorded <50 individuals for 14 of the 69 species detected (20.3%), including four species repre- sented by a single observation; Sungrebe, Green- and-rufous Kingfisher (Chloroceryle iiu/a), Amer- ican Pygmy Kingfisher (C. acnea), and Least Bittern (l.xohrycluis e.xilis). Eorty-one (59%) of the 69 species occurred in every ecoregion and on >50% of routes; 20 species (29%) occurred in 3^ ecoregions, and 8 (12%) occuiTed in only 1-2 ecoregions. Seven- teen species occurred on 90-100% of all routes, but 15 occurred on <20% of routes. Wading birds, including the Double-striped Thick-knee, were especially ubiquitous with 17 of 26 species (65.4%) occLiiTing on >70% of all routes. Routes in the BBE had the greatest water cover, ranging from 15 to 19% in the Late Dry season to 54-55% in the Late Wet. The Western Llanos had the lowest water cover, ranging from 4 to 5% in the Late Dry season to 1 1-12% in the Late Wet (Table 3). The BBE contained the highest proportion of sample points among ecoregions with small, excavated ponds along the roadside (prestamos, 63%), followed by Paspalum (45%), Western (24%), Central (34%), and Ricefields (7%). Species Richness and Abundance by Ecore- gion.— Most waterbirds were recorded in the BBE (54%), followed by Ricefields (17%), whereas the Central had the fewest (3%, Table 4). The mean/ route provides a less biased comparison among ecoregions because it standardized our sampling effort for an uneven number of routes per ecoregion (Table 5). The BBE had the highest mean number TABLE 3. Mean (%) water cover at each point count during each fieldwork season and in each ecoregion of the Llanos. Water coverage was estimated within 200 m of each point count. No data were collected during the Late Wet season 2001. Ficorogion l,alc Wci (2000) Ciirly Dry (2000) I.ale Dry (2001) tuirly Wel (2001) Late Wcl (2001) Early Drv 12001)' (.ate I2ry (2002) Early Wet (2002) Central 1 1.7 15.0 9.9 9.0 1 1.7 7.4 12.5 Ricefields 17.4 27.8 12.6 13.8 14.5 9.7 12.9 Western 1 1.1 14.0 5.4 7.8 9.4 4.0 1 1.8 Paspalum 46.9 44.1 1 1.4 22.4 21.2 10.5 25.8 BBE 54.6 44.0 14.8 29.4 30.9 18.5 54.3 MleUa (111(1 liaUlassarrc • LLANOS WAT ERBIROS 107 TABLE 4. Number of wulerbirds within guilds in surveyed over two field seasons. each eeoregion sampled in the Llanos, Venezuela, based on all routes Guild Cenlnil RiceHelds WL-slern Paspaluni BBE Toials Wadins birds 6,1 15 36,330 23,944 29,554 58,507 154,450 Waterfowl 378 755 2,280 3,574 67,348 74,335 Shorebirds 449 10,085 1,518 6,045 1 ,005 31,227 Open water 44 703 225 4,153 12,161 17,286 Forest 152 23 591 1,1 10 1,401 3,280 Passerines 211 281 536 905 1,005 2,938 Totals 7,349 48,177 29,094 45,341 153,605 283.566 Percent total 2.6 17.0 10.3 16.0 54.2 TABLE 5. Abundance of waterbirds expressed as mean (± SE) number of individuals per route in each guild for each eeoregion. Relative percentage (Rel. %) of each mean for each guild among (Rel. %-A) and within (Rel. %-W) ecoregions and species richness are included. Guild Central Ricefietds Western Paspalum BBE Wading birds Mean ± SE 200.6 ± 29.0 617.2 ± 78.4 258.9 ± 25.8 289.8 ± 10.2 574.2 ± 50.9 Rel. %-A 10.3 31.8 13.3 14.9 29.6 Rel. %-W 82.4 76.9 82.1 65.9 39.4 Richness 23 21 25 25 26 Waterfowl Mean ± SE 13.6 ± 4.6 1 1.9 ± 3.1 24.9 ± 4.2 33.4 ± 8.5 611.0 ± 240.6 Rel. %-A 2.0 1.7 3.6 4.8 87.9 Rel. %-W 5.6 1.5 7.9 7.6 41.9 Richness 6 7 7 7 10 Shorebirds Mean ± SE 14.9 ± 4.6 157.7 ± 48.1 15.9 ± 1.0 57.5 ± 5.6 126.6 ± 14.9 Rel. %-A 4.0 42.3 4.3 15.4 34.0 Rel. %-W 6.1 19.6 5.0 13.1 8.7 Richness 9 9 7 10 12 Open water Mean ± SE 1.6 ± 0.3 11.1 ±7.0 2.5 ± 0.1 39.2 ± 1.6 122.5 ± 13.8 Rel. %-A 0.9 6.3 1.4 22.2 69.2 Rel. %-W 0.7 1.4 0.8 8.9 8.4 Richness 5 6 10 1 1 12 Forest Mean ± SE 5.8 ± 1.6 0.4 ± 0.1 7.6 ± 1 .2 10.8 ± 1.9 13.8 ± 4.7 Rel. %-A 15.1 1.0 19.8 28.2 36.0 Rel. %-W 2.4 0.1 2.4 2.5 0.9 Richness 5 6 10 1 1 10 Passerines Mean ± SE 6.8 ± 1.9 4.9 ± 0.6 5.5 ± 0.7 9.0 ± 1.2 10.5 ± 1.3 Rel. %-A 18.5 13.4 15.0 24.5 28.6 Rel. %-W 2.8 0.6 1.7 2.1 0.7 Richness 4 4 4 4 4 All guilds Mean ± SE 243.3 803.2 315.3 439.7 1,458.6 Richness 51 49 57 61 66 108 THE WILSON JOURNAL OL ORNITHOLOGY • Voi 122, No. I, March 2010 of waterbirds per route (1,459/route), and the Central contained the lowest (243/route). Wading birds were the most abundant guild per route in all ecoregions (201-617/route) except the BBE, where waterfowl were most abundant (61 1/route). Wading birds comprised 39-82% of the mean/route in each guild. Shorebirds were most abundant in Ricefields (158/route) and BBE (127/route) but comprised <20% of all waterbirds detected within an ecoregion. Open Water species were concentrated in the BBE (123/route) and Paspalum (39/route), but were < 1 1/route in other ecoregions and comprised <9% of all waterbirds within a guild. Eorest and Passerine species were not abundant (< 14/route) or proportionally substantial (<3%) in any ecoregion. Overall species richness was greatest in the BBE (66) and least in the Central and Ricefields (Table 5). Wading birds were well represented in all ecoregions, but only the BBE included all 26 species. The fewest wading birds (21) were recorded in Ricefields. Species richness of the Open Water, Waterfowl, and Shorebird guilds showed greater variation among ecoregions than the Wading Bird guild. Eor example, only six of 13 species in the Open Water guild occuired in the Central and Ricefields ecoregions. Waterhird Guilds (Wading Birds). — ^We record- ed 154,450 wading birds with this guild dominat- ed by herons (14 species), followed by ibises (8), and storks (3). The herons included two (Cattle Egret, Rank 1; and Great Egret, Rank 4) of the 10 most abundant species. The herons also included 4 species ranked in the top 20: Snowy Egret (Egretta thida), Cocoi Heron (Ardea cocoi). Little Blue Heron (£. caeruleu), and Whistling Heron (Syrigma sihilaOix). Three species were rarely recorded: Least Bittern, Green Heron (Biitorides virescens), and Pinnated Bittern (Botaiirus pinna- tus). Overall, herons were widespread, as 1 1 species occurred on >50% of routes, 1 2 species occurred in all five ecoregions, and eight occiured on >40% of tran.sect sample points. The ibis group was also widespread and frequently encountered. Three species ranked in the top 10 by abundance: Bare-faced Ibis (Phinwsiis infuscatiis, Rank 6), Glossy Ibis (Plegadis falciiudliis. Rank 7), and Scarlet Ibis (Eudocimus ruhei\ Rank 9). Ibi.ses occurred on 57-100% of all routes and in all ecoregions with the exception of the Sharp-tailed Ibis (Ccrcihis o.xycerca), which was absent from the Ricefields. The Wood Stork (Mycteria americana) was most abundant; it represented only 2.6% of the wading bird guild but 74.6% of all storks and occLin'ed on 70% of all routes. The Wood Stork also ranked 12th in overall abundance. The Maguari Stork (Ciconia maguari) occurred on only 43% of routes, and the Jabiru (Jahini mycteria) occuiTed on only 35%. Waterhird Guilds (Waterfowl). — We grouped 10 species, including the two species of gallinules into the waterfowl guild. However, 95.1% of the 74,076 individuals recorded consisted of two species: White-faced Whistling Duck and Black- bellied Whistling Duck. Only the Purple Gallinule (Porphyrula martinica) and Brazilian Teal (Ama- zonetta hrasiliensis) of the remaining eight species exceeded 1% of the guild total. Blue-winged Teal (Anas discors), the only migratory species of waterfowl to reach the Llanos from North America, were rarely recorded (196 individuals). The waterfowl guild generally was not widely distributed. Only the whistling ducks and galli- nules were recorded in all ecoregions and on >50% of routes. Lour species were concentrated in only one ecoregion (the Alluvial Overflow Plains): Orinoco Goose (Neochen juhata), Brazi- lian Teal, White-faced Whistling Duck, and Black-bellied Whistling Duck. Waterhird Guilds (Shorehirds). — ^The Wattled Jacana was the most abundant of the 12 shorebirds, accounting for 53.6% of the 31,277 birds recorded in the guild and ranked 5th in overall abundance; the Black-necked Stilt was the second most abundant shorebird (#12). Two species occurred on >90% of routes (Wattled Jacana and Double-striped Thick-knee), but most shorebirds were more patchily distributed (8 of 12 occurred on <50% of routes), and not abundant. The South American Snipe (Gallinago para- guaiae) was the least frequently encountered shorebird with only eight individuals. Waterhird Guilds (Open Water). — We recorded 13 species (17,286 individuals) in this guild, of which the Neotropic Cormorant was the most frequently recorded species (61.1%), followed by the Large-billed Tern (Pliaetusa simple.v, 18.8,%). The Ringed Kingfisher (Megaceryle torcpiata) was the most abundant kingfisher (3.7%). Swimming species (Neotropic Cormorant, Anhinga, Least Grebe, Sungrebe) were more than twice as common as terns and kingfishers combined. Open water species were primarily recorded in the BBE (69.2%); 22.2% of all observations occurred in Paspalum. The Neolropic Cormorant Vilelhi and BaUlassarre • LLANOS WATERBIRDS 109 accounted for >60% of observations, but all guild species were more abundant (>75% of observa- tions) in these two ecoregions with the exception of the Least Grebe, which was recorded on only five routes, four of which were in the Western Llanos. Five species were recorded <100 times: Sungrebe, Green-and-rufous Kingfisher, Ameri- can Pygmy Kingfisher, Least Grebe, and Green Kingfisher {Chloroceryle americana). We only encountered one Green-and-rufous Kingfisher and one American Pygmy Kingfisher. Waterbird Guilds (Forest). — ^The Hoatzin was the most abundant species in this guild, repre- senting 94.2% of the 3,280 individuals recorded. All species were observed in at least four ecoregions, but distribution was patchy (<50% of routes) and lowest in the Ricefields ecoregion. Waterbird Guilds (Passerines). — We recorded only 2,938 individuals for the four species in this guild, of which the Pied Water Tyrant was most abundant (46.6%). All species were widespread (>80% of routes) and occurred in all ecoregions. Temporal Patterns Within Ecoregions. — We recorded fewer waterbirds in the first fieldwork period (101,146) than the second (182,422), but combined data by seasons because seasonal differences between fieldwork periods generally were not significant (Table 6). Overall, we recorded more species and individual birds in the Late Dry season of the second fieldwork period. The proportions of waterbirds also dif- fered between years for the Late Wet season in the BBE (P = 0.031) and the Early Dry season in the Central {P = 0.015). Wading birds were the most abundant guild, representing >50% of all detections in all ecoregions and all seasons, except for the BBE in the Late Dry season of the second fieldwork period. There were seasonal changes in mean/ route in three ecoregions (F = 2.84; df = 4, 49; P = 0.056), primarily caused by a decrease in Cattle Egrets. Wading bird numbers in the Western ecoregion were lowest in the Late Wet season (230/route) before increasing in the Early Wet season (F = 8.13; df = 4, 49; P < 0.001). Wading bird numbers in the BBE were highest in the Late Dry season (846/route) and lowest in the Early Wet season (383/route; F = 3.85; df = 4, 49; P = 0.015). This seasonal .shift in wading bird distribution between the Western and BBE ecoregion was most obvious for five ibis species: Scarlet, American White Ibis (Eudocinnis albus). Sharp-tailed, Glossy, and Bare-faced Ibis. Waterfowl were most abundant in the BBE (F = 22.01; df = 4, 49; P < 0.001) where we recorded >90% of all individuals, but detections in each ecoregion showed significant seasonal patterns. Eor example, detections in the BBE in the Eate Dry season increased to nearly 2,000/ route (F = 8.06; df = 3, 52; P < 0.001), because of seasonal aggregations of White-faced Whis- tling Ducks and Black-bellied Whistling Ducks. Waterfowl detections were lowest (P < 0.003) outside the BBE in the Late Dry season and increased in the Early Wet season. The BBE was most important for shorebird abundance (F = 9.45; df = 4, 49; P < 0.001 ) with counts exceeding 70/route in all seasons. Howev- er, Ricefields exceeded this average in the Late Wet season, when mean route totals were 352/ route with the airival of Nearctic migrants (mostly Calidris and Tringa species). Off-route observa- tions of Calidris species exceeded 2,000 individ- uals, but large flocks were infrequently encoun- tered along routes. Migrants accounted for 87% of all shorebirds recorded during the Late Wet season, but this proportion decreased to 40% in the Early Dry season, as migrants dispersed. The seasonal dynamic of shorebird abundance in the Alluvial Overflow Plains was driven mainly by detections of the Wattled Jacana. Eor example, numbers of Wattled Jacanas recorded in the Ricefields accounted for <5% of all shorebirds, but increased to >70% in the Western, Paspalum, and BBE. Wattled Jacanas (> 100/route) were recorded in the BBE during the dry season, although the mean number of detections for shorebirds increased to > 200/route in the Eate Dry season with amval of Black-necked Stilts. Seasonal variation in the mean count per route for Open Water birds was not significant in any ecoregion except the Ricefields (F = 3.20; df = 3. 28; P — 0.039), where numbers peaked at 41 birds/route in the Late Wet season. This differ- ence potentially resulted from flooding of the Rio Portuguesa, which was at the southern limit of the agricultural region and sampled with a single route. Over 500 Neotropic Cormorants were attracted to these temporary conditions, which raised the mean route count for the season to almost lOO/route. Excluding this point reduced the average abundance of the Open Water guild in the Ricefields to <1 /route in any season. Several species within the Open Water guild exhibited distinct seasonal trends. Most notably, the Yel- low-billed Tern (Sternula superciliaris). Gull- 110 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122. No. I. March 2010 TABLE 6. Waterbirds (mean/route ± SE) detected across seasons for each guild and ecoregion. Means were calculated trom combined data over both fieldwork seasons. Means for each guild within ecoregions that differed across seasons (P < 0.05) are denoted by lower case letters (one-way ANOVA and Waller-Duncan post-hoc comparisons). Central Llanos ecoregion (« = 4) Late Wet Early Dry Late Dry Early Wet Guild Wading birds 148.5 ± 26.9 269.5 ± 130.6 156.9 ± 54.9 227.5 ± 99.0 Waterfowl 21.6 ± 12.6'’ 1.6 ± 0.6“ 2.0 ± 1.7“ 29.3 ± 9.8" Shorebirds 28.1 ± 7.2'’ 9.4 ± 2.7“-'’ 14.5 ± 5.5“-'’ 8.0 ± 2.4" Open water 2.1 ± 0.8 1.3 ± 0.6 1.8 ± 0.7 1.0 ± 0.5 Forest 4.5 ± 4.2 5.0 ± 4.0 10.5 ± 10.3 3.3 ± 3.3 Passerines 10.3 ± 2.5'’ 4.1 ± 1.3“ 3.0 ± 0.5“ 9.8 ± 1.7" Ricefield.s ecoregion (n = 8) Late Wet Early Dry Late Dry Early Wei Guild Wading birds 321.3 ± 1 10.7“ 858.7 ± 243.9" 635.0 ± 187.6“'’ 653.9 ± 158.0“ Waterfowl 17.8 ± 5.1'’ 20.0 ± l.T 0.4 ± 0.3“ 9.6 ± 4.8" Shorebirds 351.9 ± 239.5 147.9 ± 62.3 86.4 ± 37.5 44.7 ± 12.3 Open water 41.0 ± 38.2'’ 2.4 ± 1 .5“-'’ 0.4 ± 0.2“ 0.6 ± 0.2“ Forest 0.2 ± 0.2 0.3 ± 0.3 0.5 ± 0.4 0.5 ± 0.4 Passerines 6.3 ± 1.5'’ 3.6 ± 1.0“ 2.6 ± 1.4“ 7.1 ± 1.9" Western Llanos ecoregion (/? = 14) Late Wet Early Dry Late Dry Early Wet Guild Wading birds 230.2 ± 24.4'’ 238.3 ± 34.5'’ 170.6 ± 38.4“ 396.4 ±41.2“ Waterfowl 34.5 ± 10.6'’ 8.7 ± 2.7“ 14.4 ± 10.9“ 41.8 ± 6.4" Shorebirds 13.2 ± 1.9“'’ 12.4 ± 3.3“ 20.5 ± 3.8“ 17.8 ± 3.5"-“ Open water 3. 1 ± 0.7 2.8 ± 0.7 1 .9 ± 0.4 2.2 ± 0.5 Forest 5.1 ± 2.8 4.7 ± 2.1 13.9 ± 7.9 6.9 ± 3.2 Passerines 7.0 ± 0.9'’ 3.5 ± 0.6“ 3.4 ± 0.7“ 8.3 ± 1.3" Paspalum ecoregion (« = 14) Late Wet Early Dry Late Dry Early Wet Guild Wading birds 240.5 ± 38.7 301.8 ± 58.7 332.2 ± 1 18.3 284.8 ± 53.5 Waterfowl 36.7 ± 13.3'-’ 14.2 ± 6.0'’ 5.5 ± 4.6“ 77.0 ± 32.3“ Shorebirds 41.2 ± 10.7“-'’ 38.0 ± 1 1 .5“ 76.4 ± 14.0“ 74.6 ± 15.0"“ Open water 36.2 ± 10.5 42.5 ± 12.2 45.4 ± 17.3 32.6 ± 10.6 Forest 4.1 ± 1.4 6.3 ± 1.9 20.8 ± 1 1.0 12.0 ± 4.1 Passerines 12.9 ± 1.3'’ 4.8 ± 1 .0“ 5.4 ± 0.9“ 13.0 ± 1.2" BBE (/I = 14) Late Wet Early Dry Late Dry Early Wet Guild Wading birds 582.5 ± 149.6“'’ 640.4 ± 87.8'’'^' 845.8 ± 143.6“ 382.6 ± 46.9“ ' Waterfowl 165.5 ± 51.0^ 96.9 ± 43.4“ 1.958.7 ± 1.254.2“ 223.0 ± 42.7"-“ Shorebirds 71.8 ± 13.2“ 133.0 ± 33.9“'’ 201.2 ± 52.4'’ 100.5 ± 18.8“ Open water 93.5 ± 39.7 187.9 ± 89.2 137.6 ± 69.7 71.2 ± 26.6 Forest 2.2 ± 1.2 7.8 ± 3.9 39.6 ± 2 1 .6 5.5 ± 2.9 Passerines 15.3 ± 2.2'’ 5.4 ± 0.9“ 7.3 ± 1.3“ 13.8 ± 1.9" I 'ilella ciihI Bcilclassanc • LLANOS WATER BIRDS billed Tern {Gelochelickm nilotica), and Black Skimmer (Rynchops niger) were dry season visitors to the BBE. Anhinga and Ringed Kingfisher also became less abundant in the Late Dry season (P < 0.05). The mean per route for the Forest guild was greatest in all ecoregions at the end of the dry season, peaking at 40/route in the BBE as flocks of Hoatzin became more common. Counts during the dry season also increased for the Sunbittern but seasonal differences in abundance were not signif- icant in any ecoregion (P > 0.05). Passerines were most abundant in the BBE and Paspalum (P = 6.00; df = 1, 4; P < 0.001) but seasonal abundance was similar among ecoregions. The fewest passerines were detected in all ecoregions during the Early and Late Dry seasons (P < 0.052). DISCUSSION The Llanos support a diverse and abundant community of wetland dependent birds. We emphasize three major findings: (1) the waterbird community is dominated by wading birds, (2) waterbirds are not homogeneously distributed across the Llanos ecosystem, and (3) compara- tively few species comprised the majority of individuals recorded. The savanna wetlands of the Pantanal of Brazil are similarly dominated by wading birds (Figueira et al. 2006). Like the Pantanal, waterbird communities of the Llanos of Venezuela were characterized by habitat general- ists. Specialization is a function of the number of habitats used (Rotenberry and Wiens 1980). The waterbirds of the Llanos, both residents as well as seasonal species, generally occurred in a variety of wetland types within and between ecoregions. This may reflect the ecological flexibility and/or similarities among these habitats regarding struc- ture and food availability. This flexibility may reflect the spatiotemporal habitat heterogeneity of the Llanos and may be fundamental to maintain- ing local diversity of waterbirds. We found significant spatial and temporal heterogeneity across this vast savanna ecosystem. Specifically, 70% of all waterbirds recorded occurred in the Alluvial Overflow Plains (BBE and Paspalum); Ricefields ranked a distant second at 17%. The BBE, within the Alluvial Overflow Plains, contained the most waterbird detections per route (1,459) and was especially important during the dry seasons. The wettest ecoregion during all seasons was the Alluvial Overflow Plains averaging 1 1-55% water cover at each 1 1 1 sample point during each field season. The BBE was especially wet during the dry seasons (11- 44%), even exceeding water cover of the Rice- fields (10-27%). Overall, 63% of all half-points sampled in the BBE and 45% in the Paspalum contained permanent water in the lorm ol prestamos along the roadside versus 7% in Ricefields, 24% in the Western Llanos, and 34% in the Central. Many species, especially in the dry seasons, shift distributions in response to these hydrological patterns, concentrating in the Allu- vial Overflow Plains. The BBE contained the highest relative percentage of each waterbird guild, while the Paspalum usually ranked second or third (Table 5). Only a few of the 69 waterbird species dominate the abundance of waterbirds we en- countered in the Llanos. Nearctic migrants had marked seasonal patterns of abundance. We recorded only 196 Blue-winged Teal during our 2-year study, despite Venezuela’s recognition as an important wintering site for this species (Botero and Rusch 1988). Most migrant water- birds travel and winter along the Caribbean coast of Venezuela with smaller numbers in the inland wetlands (Rodner 2006). Recent surveys have also highlighted the importance of the northern Orinoco Delta of Venezuela as migratory shore- bird habitat (Rodner 2006). The BBE was the most important ecoregion for migrant shorebirds in the Llanos (Thomas 1987). Numbers increased in the Ricefields during the Late Wet season. Waterfowl were mainly concentrated in the BBE and Ricefields ecoregions with decreasing detections in all other ecoregions sampled. Only Black-bellied and White-faced whistling ducks were recorded in all ecoregions. Waterfowl exhibited a distinct relationship between seasonal patterns of abundance in different ecoregions. which generally reflected the transition from the Late Dry season to the Early Wet season. Waterfowl hunting is permitted during the breed- ing season for some species, particularly whistling ducks, which are considered serious rice crop pests (Sanz-Agreda 1998). Our surveys recorded dramatically reduced numbers of whistling ducks in the Llanos compared to estimates reported by Gomez-Dallmeier and Cringan (1989). CONSERVATION IMPLICATIONS Waterbirds tended to spread across the eco- regions during wet seasons but contracted to the Alluvial Overflow Plains in dry seasons. We 112 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. I. March 2010 believe this ecoregion should be a focus of conservation efforts, because it contains extensive water coverage during the dry season, and supports the greatest number and diversity of Llanos waterbirds. The dominant land use in the Llanos is cattle ranching, which at current levels is compatible with sustaining waterbird popula- tions and habitats (Rios-Uzcategui 1994). How- ever, cattle ranching and agriculture can require deforestation. Widespread deforestation has oc- curred in parts of the Western Llanos (Hamilton et al. 1976, Veillon 1976). Extensive natural grass- lands are the dominant habitat in the Alluvial Overflow Plains. Major modifications to the environment, such as dams, dikes, and deforesta- tion have not yet occurred on most of these vast plains, and the low human density over most of the area prevents other significant impacts. Additional studies are needed on the waterbird communities in the Llanos, particularly in the Alluvial Overflow Plains where the role of impoundments as waterbird habitats should be evaluated. Information on population dynamics and species-habitat relationships will help better understand ecological factors involved in regulat- ing waterbird community structure and population levels. Additional studies should also survey for secretive marsh birds such as rails which, although common in the Llanos, were not detected during our study. About 250,000 ha of private property on large ranches (hatos) operate ecotourism enterprises that attract tourists from around the world (Lara 1999). These arrangements in turn provide a strong incentive to promote wildlife con.servation. Com- parisons of waterbird abundance and diversity between private hatos (where wildlife is protected) versus areas where hunting is likely to occur may provide an indication of the severity of hunting pressure (Perez and OJasti 1996) and the role of hatos in conservation of Llanos waterbirds. Long- term conservation benefits for waterbird communi- ties on private lands of the Venezuelan Llanos will require the continuing support of government and conservation agencies as well as outreach to landowners as a major approach toward protecting the rich waterbird heritage of this region. ACKNOWLEDGMENTS This paper is dedicated to the memory of our friend and collaborator. Prof. Gilberto Ri'os-llzcategui of the Uni- versidad de los Llanos Occidentales in Venezuela. Funding for this research was provided by Ducks Unlimited, the International Programs Office of the USDA Forest Service, and the Neotropical Migrant Bird Conservation Program of the U.S. Fish and Wildlife Service. We especially thank Bruce Batt of Ducks Unlimited and Doug Ryan of the U.S. Fish and Wildlife Service for logistic and administrative support. We are indebted to Mark Gregory and Alexis Araujo for field assistance. We are especially grateful to the caretakers and landowners of the following ranches for access to their properties and use of facilities; Hato El Frio, Hato El Cedral, Hato Pinero, Hato Eemando Corrales, and Mantecal. The manuscript was greatly improved by comments from two anonymous referees. LITERATURE CITED Bibby, C. J., N. J. Collar, M. J. Crosby, M. E. Heath, C. H. IMBODEN, T. H. Johnson, A. J. Long, A. J. Sattersfield, and S. j. Thirgood. 1992. Putting biodiversity on the map: priority areas for global conservation. International Council for Bird Preserva- tion, Cambridge, United Kingdom. Bisbal, E. j. 1988. Impacto humano sobre los habitats de Venezuela. Interciencia 13:226-232. Blydenstein, j. 1962. La sabana de Traebypogon del alto llano. Boleti'n de la Sociedad Venezolana de Ciencias Naturales 103:223-238. Blydenstein, J. 1967. Tropical savanna vegetation of the Llanos of Colombia. Ecology 48:1-15. Botero, j. E. and D. H. Rusch. 1988. Recoveries of North American waterfowl in the Neotropics. Pages 469^82 in Waterfowl in winter (M. W. Weller, Editor). University of Minnesota Press, Minneapolis, USA. Buckland, S. T., D. R. Anderson, K. P. Burnham, J. L. Laake, D. L. Borchers, and L. Thomas. 2004. Advanced distance sampling: estimating abundance of biological populations. Oxford University Press Inc., New York, USA. Bulla, L., J. Pacheco, and G. Morales. 1990. Seasonally flooded neotropical savanna closed by dikes. Pages 172-210 in Ecosystems of the world: managed gra,sslands (E. Breymeyer, Editor). Elsevier Sciences, New York, USA. Colwell, R. K. 2006. Estimates: statistical estimation of species richness and shared species from samples. Version 8.0. http://viceroy.eeb.unconn.edu/EstimateS (accessed 26 October 2006). Conway, C. J. and J. C. Simon. 2003. Comparison of detection probability as.sociated with Burrowing Owl survey methods. Journal of Wildlife Management 67:501-51 1. Droege, S. 1 990. The North American Breeding Bird Survey. Pages 1-4 in Survey designs and statistical methods for the estimation of avian population trends (J. R. Sauer and S. Droege, Editors). USDl, Eish and Wildlife Service, Biological Report 90. Washington, D.C., USA. Eli.i,s, D. H., R. L. Glin.ski, and D. G. Smith. 1990. Raptor road surveys in South America. Journal of Raptor Re.search 24:98-106. Eigueira, j. E., R. Cintra, L. R. Viana, and C. Yamashita. 2006. Spatial and temporal patterns of bird .species diversity in the Pantanal of Mato Grosso. Vilclla and Baldassarrc • LL.ANOS WATERIilRDS 113 Brazil; implications for conservation. Brazilian Journal of Biology 66:393—104. Frederick. P. C. and K, L. Bildstein. 1992. Foraging ecology of seven species of neotropical ibises (Thrcskiornithidae) tliiring the dry season in the Llanos of Venezuela. Wilson Bulletin 104:1-21. Gill, F., M. Wright, and D. Donsker. 2009, IOC World bird names (Version 2.1). http://www.worldbirdname.s. org/ (accessed 20 August 2009). Gomez-Dallmeier, F. and a. T. Cringan. 1989. Water- fowl in Venezuela. Editorial Ex Libris, Caracas, Venezuela. Hamilton, L. S., J. A. Steyermark, .1. P., Veillon, and E. Mondolfi. 1976. Conservacidn de los Bosques Humedos de Venezuela. Sierra Club, Consejo de Bienestar Rural, Caracas, Venezuela. Hamilton, S. K., S. J. Sippel, and J. M. Melack. 2002. Comparison of inundation patterns in South American floodplains. Journal of Geophysical Research-Atmo- spheres 107: Article 8038. Hanowski. j. M. and G. j. Niemi. 1995. A comparison of on- and off-road bird counts: do you need to go off road to count birds accurately? Journal of Field Ornithology 66:469^83. Hilty, S. L. 2003. Birds of Venezuela. Second Edition. Princeton University Press, Princeton. New Jersey, USA. Huber, O. and C. Alarcon. 1988. Mapa de vegetacion de Venezuela. 1:2,000,000. Ministerio del Ambiente y de los Recursos Naturales Renovables, Caracas, Venezuela. Lara, L. C. 1999. El ecoturismo: /,una opcion para las reservas forestales? Caso: unidad experimental de Caparo. Revista Forestal Venezolana 43:69-78. Ministerio de Agricultura y Cria. 1998. Anuario estadfstico agropecuario 1996. Publicaciones del Ministerio de Agricultura y Cn'a, Caracas, Venezuela. Mittermeier, R. a., N. Myers, J. B. Thompsen, G. A. B. DA Fonseca, and S. Olivieri. 1998. Global biodiver- sity hotspots and major tropical wilderness areas. Conservation Biology 12:516-520. Peterjohn, B. G. 1994. The North American Breeding Bird Survey. Birding 26:386-398. Perez, E. M. and J. Ojasti. 1996. La utilizacion de la fauna silvestre en la America Tropical y recomenda- ciones para su manejo sustentable en las sabanas. Ecotropicos 9:71-82. Ramia, M. 1967. Tipos de sabanas en los llanos de Venezuela. Boletm de la Socicdad Venezolana de Ciencias Naturales 27:264—288. Ramia, M. 1993. Ecologi'a de las sabanas del Estado Cojedes: relaciones vegetacion-suelo en sabanas secas. Fundacion La Salle de Ciencias Naturales, Serie Ciencia y Tecnologia Number 4. Caracas, Venezuela. Ramo, C. and B. Busto. 1984. Censo aerco de corocoros (Eiulocinns riihher) y otras aves acuaticas en Vene- zuela. Boletm Sociedad Venezolana de Ciencias Naturales 39:65-88. Ramo, C. and B. Busto. 1988. Status of the nesting population of the Scarlet Ibis (Endocimus rither) in the Venezuelan Llanos. Colonial Waterbirds 1 1:311-314. Rios-UzCATEGUi, G. A. 1994. Role of private landowners in wildlife conservation; Venezuelan case. Thesis. Uni- versity of Maryland, College Park, USA. Robbins, C. S., D. Bystrack, and P. H. Geissler. 1986. The Breeding Bird Survey: its first 15 years, 1965- 1979. USDl, Fish and Wildlife Service Resource Publication 157. Washington D.C., USA. Roca, R., L. Adkins, M. C. Wurschy, and K. Skerl. 1997. Transboundary conservation: an ecoregional approach to protect neotropical migratory birds in South America. Environmental Management 21:477— 481. Rodner, C. 2006. Waterbirds in Venezuela. Sociedad Conservacionista Audubon de Venezuela, Caracas. 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. Sanz-Agreda, a. 1998. Evaluacion del subprograma manejo de anatidos en Venezuela; caso Dendrocygna. Direccion de Manejo de Fauna Silvestre PROFAUNA, Ministerio del Ambiente y de los Recursos Naturales Renovables, Caracas, Venezuela. Sarmiento, G. 1983. The savannas of tropical America. Pages 245-288 in Tropical savannas (F. Bourliere. Editor). Volume 13. Elsevier Sciences, New York, USA. Sarmiento, G. and M. Pinillos. 2001. Patterns and processes in a seasonally flooded tropical plain: the Apure Llanos, Venezuela. Journal of Biogeography 28:985-996. Scott, D. A. and M. Carbonell. 1986. Directory of neotropical wetlands. lUCN Intemational Waterfowl and Wetland Research Bureau, Slimbridge, United Kingdom. SPSS Inc. 2001. SPSS 11.0 for Windows. SPSS Inc., Chicago, Illinois, USA. Tamisier, a. and O. Dehorter. 2000. Eauna of the Llanos and Pantanal; ecological basis for the sustainable management of tropical Hooded ecosystems. Case studies in the Llanos (Venezuela) and the Pantanal (Brazil). Centre d’Ecologie Eonctionelle et Evolutive CNRS, Montpellier. Erance. Thiollay, j. M. 1978. Population structure and seasonal tluctuations of the Falconiformes in Uganda Na- tional parks. East African Wildlife Journal 16:145- 151. Tiioma.s, B. 1987. Spring shorebird migration through central Venezuela. Wilson Bulletin 99:571-578. Veillon, J. P. 1976. Las deforcstaciones de los Llanos Occidcntales de Venezuela desde 1950 hasta 1975. Pages 67-112 in Conservacion de los bosques humedos de Venezuela (L. S. Hamilton. J. Steyermark. .1. P. Veillon, y E. Mondolfi, Editores). Ministerio del Ambiente y de los Recursos Naturales Renovables, Caracas, Venezuela. Velasco, A. and J. Ayarzaguena. 1985. Situacion actual de las poblaciones de baba (Caiman crocodi- Ins) sometidas a aprovechamiento comercial en los Llanos Venezolanos. Publicaciones de la Asociacion de Amigos de Donana Number 5 Caracas, Vene- zuela. 114 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. I. March 2010 c o •— (U O v; c. c C. E = h; a ■a O ^ 3 .E CJ c - ^ . #. ^ ■a 5 w -a • E o =5 D. ? E! ^ O (D 1) c/: C C3 CJ (U o. i: ^ c « O (U L. c/5 4J .2 CJ & OJ Q. c/5 C/5 C ^ .2 o '=‘^' Q- 5 O J2 — 8S CJ 2 E ^ c3 Q o C . X E 5 z z w CL D- ^5 in Tt f<5 lAi i/--, rY lAi pj- — r- (n C500C — ocNoccNsccnoopfr-vOcxcnin^ O CN Tt — r- d d so oC ^ ■5: ^ C S 2 W ^ w C C3 Y o c oo O •r- O ^ CL- o ^ Sj >, c o u CJ X X 00 . 'S Cj s: '5' § 1 1 5^"§ CJ X X5 ^ — ^ "O ■a ji ^ c "5 -2 ? ^ Oc 'Ll CL ^ ^ Cj :§, s E" .2 ^ c = kj c o o CL ^ cn s* -- s s Ti “§ “O ^ CJ (A •S s s3 22 d d 2: ^ CJ 5 E 4^ L. o -o w 3 c c/3 X '5 1 q CJ CJ CJ CJ c uL >> c/3 C/3 X c CJ CJ 3 CJ 3 1 B- X CJ *»««» s •c 3 3 Oi) cn TD o *H cJ5 X o c U "cj >- u 3 X 4— 3 X O a CJ u- a c/3 O CC 3 CJ c/5 3 X cn X X 3 3 S c cd CJ u CJ 00 •o CJ E o *§ § **» 3 2 00 r^ S ^ • Q , ^ --- D Q Q ^ ^ o £ =3 d Q c/3 1 I -o CJ J= 2 ^ CJ q m -li o J I m X 2 2 D, CL Oc & ^ Q L<2 -02 Q Q OXj go C C 2 2 S "§ -= d CJ ~o d c ^ D 5 S U ^ CJ U m 3 O d Q r o 8 8 cn c S o CJ 3 -g CJ 'o 3 3 -■E o ^ -o 8 X CJ o ^ E- S U _3 “5 3 o x: c X 'qj •£ w -»- 3 d Cl Cu < CJ 3 -o E o X c/; CJ u X h- CJ 3 i£ ’E o CJ U CJ 3 •u E X 3 C < yUc'lla anil Baldassarre • LLANOS WATERBIRDS 115 ■o o p p ^ r^ mi O m m os p ^ m* c ’ r- m r- mi o r- m in ms in r- OJ ooosOS'^Opinp ■ — — Tt o o6 — 00 os O 00 00 00 in — r- m o Os ooop — inp — o — in o o — o — ni — mosTfinoso — oomoor-miO oooo'^o — oo — ooom'^irm^ — — ; 00000000 — 0 — 000000000 os'sor-'^ — os'^mmino — m* m m so sD — m^ in — ^ m r- ^ m* o mi nO so r-sosDso — r-inoo'^rj-m'?fmm — mm^ — in — '^msosom^mm'^'^ mooo — os — somm — Os"^m — oommnJinoo r:}- oj p \o — m^ r-^ — — in so" oo os r- O in m in m in oooommm'^m — osD — mlnmJmsO'^lno^r-m m m m — o m m o so r- in so rj- r- m so ^ so so m r- mi m — m m m 00 m o s ^ ^ O O u- x)"0'0’0'a’0"a"opppp’p o o -c x: cn c/5 o -C cn 00 o o x: xz cn o o x: xz 00 00 o o H) c fU a. OO 00 OO 00 00 oo Ui Ln u. Lh L-. L- O 0) a (U 0) o cd 3 td 3 cd 0) o a > C ^:3 UJ < o) 'r- 6f} 1/5 -C; (U ^0 -ii c OJ C3 c/2 w 1) czi • — W- (U C o ^ u -2 ^ 5^ ^ 'c i .1 c rv “'J -a .y c c ■- ■r n jj I c3 3 O O Cl 00 C/D S Cl w O. 6T 1/5 ^sj Sc 5 ^ 2 B t'l 6c O Cl ■ «>3 ■r: a ' o- O c i_ (U S G G fe H H (U ^ -C L, 5/2 (3J i:: j= ^ Si "2 B G G k. L. a o x: xz c/2 C/5 6JJ 61j C C 2 5 C5 Sj . TO c 3 00 E oo 3 3 £ H 00 “3 — -E O 3 5 'oij 3 3 L. -o 00 TO c OO 00 00 I) iE x $ 3 3 O 5 3 3 3 X) a Q ’n N 3 3 “2 jd o __ 6Ij 3 3 OJ 0/ o Cl 'E cl Jd ~B u. 3 "E O £ 1) L. :n < a o D 3 p o 'C 3 p c/2 O fU C3 ~a “O C3 u. C3 X (U C3 *0 'E C6 CQ U QJi cj c6 D- O O O OO (U 63 *o G o o o 63 "O 3 O 3 3 p w '*5 c/2 ^ ^ B c w Li r3 13 1- 6c C H 5 i ^ — c/2 U. — is C f3 c/2 ^ ^ ^ -O 00 ^ a C L- TO "S ^ OJ oxj -Q • — .E ^ Cl > o 63 ;p 'E c 3 H -c ^ C2 ^ s ^ 2 ^ ^ Ci c/) *3 -O ^ o _ c .E o p Q S “O ^ o ^ Q- o ^ & V 3 V (J p i3 ^ cQ ac o 3 (D TO c3 w “O ■;:3 00 .i:i o i- — a OU Xj o c LLl The I'Vilsun Journal of Ornithology 122(1):1 16-125, 2010 PHENOLOGY OF SIX MIGRATORY COASTAL BIRDS IN RELATION TO CLIMATE CHANGE CHARLES R. FOSTER, ANTHONY F. AMOS,' AND LEE A. FUIMAN' ABSTRACT.— The migration phenology of six species of coastal birds on Mustang Island, Texas, USA was examined tor a 27-year period ( 1978-2005). First arrival date, last date of departure, and duration of stay were quantified for three winter and three summer residents. These three variables were analyzed for changes over time and correlation with local, regional, and global temperature indices. Mean local summer temperature increased 0,03° C/year (0.74° C overall), while mean local winter temperature increased 0.10° C/year (2.76° C overall). The three winter residents had a trend for increasingly later arrival, increasingly earlier departure, and decrea.sed duration of stay over the 27-year period. These trends reflect a shortening of the winter season for these birds and are consistent with expected responses due to warming temperatures. The three terns representing summer residents had less homogeneity in temporal trends than the three winter residents. Correlations of local temperature with arrival and departure dates, and duration of stay yielded few significant results and no overall pattern. Only Double-crested Cormorant (Phalacrocorax aiiritus; a winter resident) and Least Tern (Sterna antillarum; a summer resident) had significant correlations between arrival date and arrival temperature. Received 27 March 2009. Accepted 3 September 2009. The Earth’s climate has warmed by —0.6° C over the last 100 years according to the Intergov- ernmental Panel on Climate Change (IPCC 2001 ). There have been two distinct periods of warming during this time; one from 1910 to 1945 and the other from 1976 continuing to the present (IPCC 2001). The rate of warming during the latter period has been approximately twice the magni- tude of the first and greater than any other time during the past 1,000 years (Walther et al. 2002). Regional changes in temperature are more significant when examining ecological changes and, in some cases, the.se climates have exhibited greater increa.ses than the mean global tempera- ture over the past 30 years. The largest regional increases have been over the middle and high latitudes of the northern hemisphere (IPCC 2001). There has been a recent paradigm shift from focusing on climatic consequences of global climate change to collecting existing evidence of current alterations in the biological activity of organisms as the result of warming temperatures (Wuethrich 2000, McCarty 2001, Walther et al. 2002, Parmesan and Yohe 2003). Climate change has altered the activities of a broad range of organisms in a variety of ways (McCarty 2001. Walther et al. 2002). Organisms worldwide are experiencing shifts in range. ' Marine .Science Instilulc, University of Texas at Austin, 750 Channel View Drive, Port Aransas, TX 78373, USA. -Current address: Gulf Coa.st Research Laboratory, University of Southern Mississippi, 703 Ea.sl Beach Drive, Ocean Springs, MS 39564. USA. 'Corresponding author; e-mail; charles.foster@usm.edu breeding and migration phenology, and commu- nity structure. It should be expected that organ- isms in the northern hemisphere will shift their ranges northward or to higher elevations with increasing temperatures to remain within their thermal preferences. Parmesan et al. (1999) reported significant range shifts consistent with regional climate change for 22 species of non- migratory European butterflies over the past century. GrabheiT et al. (1994) measured signif- icant upward shifts in elevation in nine plant species over the previous 90 years in the Austrian Alps. Phenology, or the sea.sonal timing of biological events, can also be an important tool for assessing effects of climate change. An organism may shift its range and/or adjust aspects of its phenology to cope with a changing climate. Phenological changes have been observed for a wide range of species, including woody and herbaceous plants, birds, insects and other invertebrates, amphibians, and fishes (Beebee 1995, Alias 1999, Bradley et al. 1999, Roy and Sparks 2()()(), Cotton 2003). A review of phenologies for 677 species by Parmesan and Yohe (2003) found the shifts in phenology were overwhelmingly in the direction expected from climate change (87% of all cases studied). Phenological events can occur through- out the year, but most studies of phenology have focLi.sed on the spring sea, son. This is because many biologically relevant events occur in spring that are likely to indicate an imprint of climate change since it is a transitional period. Examples of biological events that occur in spring are blooming of flowers, thawing of ice on frozen Foster el al. • COASTAL BIRD PHENOLOGY AND CLIMATE CHANGE 117 lakes, breeding of a variety of organisms, and migrations of butterflies and birds. Several studies have shown that arrival dates of numerous species of migratory birds to breeding or stopover areas have been advancing over the past century (Bradley et al. 1999, Cotton 2003, Jonzen et al. 2006). In addition, the advancement of airival has been linked to recent increases in global and regional temperatures. These findings have been documented by several studies of bird migrations (Sparks 1999, Sparks et al. 2005, Gordo and Sanz 2006). Nearly all studies of avian migration phenology have focused on passerine species, leaving other bird groups virtually unexamined. We examined migration phenologies for six bird species on the Texas coast with diverse life histories. Our objectives were to investigate the migration phenology for winter and summer residents separately by testing the following hypotheses: (1) that phenological traits have changed over the past 27 years, and (2) that phenological traits are not independent of local, regional, and global temperature indices. METHODS Survey Methods. — Surveys (BEACHobs) of Mustang Island, Texas, USA began on 14 April 1978 and continue to the present. Surveys were conducted along a 1 1.7-km stretch of barrier island shoreline between beach Access Road 1 (27° 47.6' N, 97° 05.1' W) and Access Road 2 (27° 42.2' N, 97° 09.0' W), south of Port Aransas, Texas (Eig. 1). The primary objective of BEA- CHobs was to census bird species that use the beach. Other data recorded on the surveys included numbers of people, dogs, and motor vehicles, as well as atmospheric and oceanographic parameters, beach and dune morphology, tidal stage, and occurrence of natural and anthropogenic debris. The survey protocol was established from results of several tests conducted during initial development of the survey to optimize surveying techniques with respect to direction of travel, survey frequency, and diurnal distribution of birds. The survey was then designed and con- ducted every other day, starting at sunrise (0600 to 0740 hrs CST) traveling from north to south. This sampling frequency ensured that bird pres- ence was consistently represented in the data set. All surveys were conducted by one of us (AEA) from a four-wheel-drive vehicle while driving slowly southward between the two access roads. The surveyor, who was also the driver, had a clear view of birds, the great majority of which were feeding, roosting, or loafing. Vehicular traffic is permitted along many Texas barrier island beaches, and the beach is considered a roadway where all traffic laws pertain. This made con- ducting the survey possible. Surveys were done under all weather conditions and took from 40 min to more than 3 hrs to complete, depending on beach conditions, abun- dance and diversity of bird species, and number of people. The full survey distance was not com- pleted on rare occasions due to dangerous weather events, high storm tides, or unsafe beach surfaces. There were some gaps in the survey especially in the winter months of some years when the surveyor was unavailable (Table 1). Most birds were seen in the swash zone (averaging 15 m wide) between the shoreline and the most recent high tide line. Birds in flight were also counted and occasionally there were small roosts of shorebirds between the high tide line and the foredune, which were also counted. Birds were usually counted individually, but block counting was used when abundance was high. Statistical Analyses. — ^Data for complete calen- dar years from 1979 through 2004 were analyzed for the current study. Partial year data for 1978 and 2005 were used for applicable species. Statistical analyses were computed using SYSTAT 11.0 (SYSTAT Software Inc. 2004). Migration phenologies of six species (three winter and three summer residents) were examined. These six species were selected because they exhibited a clear annual pattern of occurrence (all were present one part of the year but completely absent for the other part). This allowed identifi- cation of date of anival and date of departure for each species each year. The three winter residents were Double-crested Cormorant {Phalacrocorax auritus). Eared Grebe (Podiceps nigricolli.'i). and Herring Gull (Larus argentatiis). The three summer residents were Gull-billed Tern (Sterna nilotica). Least Tern (S. antillarum). and Sand- wich Tern (S. sandvicensi.s). Phenology was described by: first arrival date (PAD), last date of departure (LDD), and duration of stay (DUR). PAD was the first day of each year that a species was observed on Mustang Island. LDD was the day that a species was last seen each year. Duration of stay was calculated as the difference between LDD and FAD. If a significant (>1 week) gap in the 118 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 1, March 2010 97'30'0"W 97’0'0''W 96"30'0"W COPANi ORPUS CHRll BAY fAFPIN -28’0'0"N ■27'30'0"N •27"0'0"N •26°30'0"N -28”30’0"N 26’30'0"NH 28'30’0''N' Legend National Seashore Boundary Stream — I — 97-0'0"W Corpus Chrlsti Bay Gun of Mexic USTANG ISLAND JOSE ISLAND 28’0'0"N' GULF OF MEXICO N 97’30'0''W 1— 96’30'0”W Access Road # 1 Access Road U 2 Mustang Island Beach Study Area W-^ Acc ! USTAn ss Road # 1 G ISLAND Access Rc ad 2 27’30'0"N' 27‘0'0"N' FIG. I. Beach survey area on Mustang Island, Texas. surveys existed during the period of arrival or departure, that year was omitted from the analysis. Temporal trends in phenological traits were de.scribed for each species using linear regression and the slope of each regression line was used to estimate the trend. The hypothesis on temporal change in a phenological trait for winter or summer residents was tested with Student’s t statistic ba.sed on the trend value for each species (i.e., /? = 3 trends). Local temperature data bases were used to as.sess changing temperature and possible effects on avian migration phenology. These data bases included the temperature record taken during the Mustang Island beach surveys. Temperature records from Corpus Christi, Texas (—30 km Foster et al. • COASTAL BIRD PHENOLOGY AND CLIMATE CHANGE 1 19 TABLE 1. Number of surveys completed by month on Mustang Island, Texas, 1978-2005. Month Number of surveys Jan 208 Feb 227 Mar 323 Apr 322 May 347 Jun 320 Jul 327 Aug 335 Sep 313 Oct 303 Nov 332 Dec 266 from the survey area) were also used. These data, obtained from the NASA Goddard Institute for Space Studies (http://data.giss.nasa.gov/cgi-bin/ csci/), served as an inland location less influenced by the Gulf of Mexico. Two additional temperature data bases were used to assess the importance of large-scale climatological changes for the six species exam- ined. Global and northern hemispheric tempera- ture anomaly records were obtained from the Climatic Research Unit at the University of East Anglia, Norwich, UK (http;//www.cru. uea.ac.uk/ cru/data/temperature). These temperature records are those used by the Intergovernmental Panel on Climate Change (IPCC). The anomaly values are measures of temperature relative to the average temperature during the period of best coverage (1961-1990) and have become the standard for most studies of climate change. These global and hemispheric averages are accurate to — h/— 0.05° C (2 SEs) for the period since 1951. All four temperature indices were analyzed using linear regression to test for a significant change over time, and for correlations between temperature and migration phenology. FAD was compared with mean temperature for the month(s) in which that particular species arrived on Mustang Island (arrival temperature). FDD was compared with the mean temperature of the month(s) during which that species left Mustang Island (departure temperature). For instance. Sandwich Terns arrived in March and departed in November or December. Arrival temperature used for the Sandwich Tern was the mean March temperature, while the departure temperature was the average of the monthly mean temperatures for November and December. RESULTS Local air temperature on Mustang Island, Texas increased, on average, by 0.04° C/year during 1978-2005. This overall increase of 1.08° C did not represent a significant linear trend (P = 0.17; Fig. 2A). However, a highly significant trend of similar magnitude (0.05° C/year) was present in the data for Corpus Christi, Texas over the same period {P < 0.001; Fig. 2B). All three species that winter in the study area had positive slopes for airival date, negative slopes for departure date, and negative slopes for duration of stay (Table 2). The mean slopes for FAD and FDD were not significantly different from zero. However, the mean slope for DUR was —0.8 days/year and demonstrated a significant departure from the null model of no change over time {t — —5.188, n = 3, P = 0.035; Table 2). This mean slope equates to an average decrease in duration of stay for winter residents of 21.9 days over the 27-year study period. The three terns had less homogeneity in the temporal pattern of the phenological traits than the three winter residents. Sandwich and Gull-billed terns had a non-significant negative slope for FAD, while Least Terns had an opposite, significant trend (Table 3). All three species had negative slopes for FDD and DUR (Sandwich Tern was non-significant for both). Sandwich and Gull-billed terns had a trend toward earlier aiTival. but the departure date for these two species was also earlier. The mean slope of DUR for all three species combined was not significantly different from zero (t = — 1 .750, n = 3. P = 0.22; Table 2). Local mean winter (Dec-Feb) and summer temperatures (Jun-Aug) were calculated separate- ly to examine the relationship between trends in phenology for winter and summer residents. Mean summer temperature increased by 0.03° C/year over 1978-2005, an overall significant increase of 0.74° C over the past 27 years {P = 0.013; Fig. 3). Mean winter temperature increased sig- nificantly by 0.10° C/year or 2.76° C over the same period {P < 0.001; Fig. 3). Comparisons of local temperatures with arrival and departure dates, and duration of stay yielded few significant results and no consistent pattern. Only the Double-crested Cormorant, a winter resident, and Least Tern, a summer resident, had a significant con-elation between amval date and aiTival temperatures and both were positive trends 120 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 1, March 2010 (A) (B) FIG. 2. Mean annual air temperature for Port Aransas, Texas (A) from 1979 to 2004 (trend not significant; P = 0.17), and for Corpus Christi, Texas (B) from 1978 to 2005 (trend is significant, P < 0.001). Foster et al. • COASTAL BIRD PHENOLOGY AND CLIMATE CHANGE 121 TABLE 2. Slopes of regression lines and results ot one- sample student's r-test for first arrival date (FAD), last date of departure (LDD), and duration of stay for both winter and summer species (individually significant values are in bold). Species FAD Year vs. LDD Duration Winter Double-crested Cormorant (Phalacrocorax auritus) 0.409 -0.416 -0.812 Eared Grebe (Podiceps nigrieolis) 1.262 -0.171 -0.489 Herring Gull (Larits argentalits) 0.141 -0.644 -0.993 Mean 0.604 -0.410 -0.765 SE 0.585 0.237 0.255 t 1.787 -3.005 -5.188 P 0.216 0.095 0.035 Summer Gull-billed Tern (Sterna nilotica) -0.404 -1.738 -2.087 Least Tern (S. antillanim) 0.462 -0.393 -0.825 Sandwich Tern (S. sandvicensis) -0.542 -0.436 -0.115 Mean -0.161 -0.856 - 1 .009 SE 0.544 0.764 0.999 t -0.513 -1.939 -1.750 P 0.659 0.192 0.222 (Fig. 4). The Eared Grebe, a winter resident, had a significant (negative) relationship between depar- ture date and local departure temperature (Fig. 5). None of the six species had a significant relationship between duration of stay and local temperature. Similar results were found when temperature data from Corpus Christi, Texas were used. The temperature anomaly data sets used to represent the global and northern hemispheric temperatures over the past 27 years yielded only two significant results of a possible 36 (Gull- billed Tern and Least Tern exhibited significant relationships between DUR and global tempera- ture) even though significant warming occurred. DISCUSSION The general reduction in time spent by six species of birds on Mustang Island over the past 27 years, along with confirmation that local temperatures have increased more in winter than in summer, suggests that warming temperatures may be influencing the migration phenology of three winter residents. However, when phenology variables were related directly to measures of C O ON ri SO 00 OS Cj Cl so r- 1— ' r- Q. — to o p 00 odd d d d Cl re O r*-, O tCj ' — ' O ^ O' Cl ^ O 00 O O c- 53 UJ c/3 d d d d Ui + 1 +1 +1 +i +1 +1 -fl o TD C Cl as rc r- IT) ic) (U 3 O — 00 OS 00 Cl — "o H 00 "rt OS 0 00 — c o d d d ci d d > 1 1 1 1 1 1 o. « 3 "O Q Tt cn c- Os tC) Os c r- OO c- sO C3 Q (Os tn sO to p r c/3 d o6 ci d rc tu Cl d Tt ^ Cl CO +1 C — Tf 00 r- m 4J — O 00 vO (U CI — Cl d — Cl c/3 c 00 c~ O Os O rc Cl Tf OS \0 OX) re so o — ^ cO d d d odd C o ^ 00 00 r- so c/3 OS to O O Os 00 . U- rv W 4^ CJ o o CJ h 1 < O Os os ^ o tn d ^ c* 00 — 00 3 Q OS p Os so OS r- c/3 c/3 d ci ci 00 tc — "O — (^ rc C3 u CJ c > c 3 0) w > O O 3 o z < 2 < s S r- OS (^i 00 vn cS O C4 Cl d o — c/3 u __ c C3 C3 O L. ‘oX) o o E o >> o c CJ -C CJ o T3 U E c Cu '35 Ci d) u o 43 t/5 :rest( re be Gull Um cj ^ n ^ ua JD Cu i, O 60 — (2 - CO < C c/3 CJ C/3 S *o .£ u- ■= OJ h o ^ h w C3 CJ •_ 'i _ 1 122 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. I. March 2010 FIG. 3. Average summer (top; P = 0.01 3) and winter (bottom; P < 0.001 ) temperatures for Corpus Christi, Texas from 1978 to 2005. local temperature, few significant correlations were found. It is possible that temperature does not have a direct effect on migration phenology at this location, but may affect other places/times along the migration route. Our results suggest that migration is triggered by conditions in breeding areas or by changes in the availability of prey items, both possible products of climate change. Double-crested Cormorants, Eared Grebes, and Herring Gulls are all widely distributed across North America. Thus, identifying whether or not climate change is affecting the.se birds, and where on their migration route the effects are greatest, is difficult. A lack of banding studies of these birds prohibits inferences on where the birds that winter on Mustang Island actually spend the rest of the year. However, it is known that climate change affects higher latitudes at a more rapid pace than equatorial locations (IPCC 2001). The Herring Gull has the northernmost breeding range of the three winter species examined and had the strongest trend (but not significant) toward contraction of the winter season on Mustang Island. The three species of terns representing summer residents did not exhibit a consistent pattern in aiTival date, departure date, or duration of stay. This inconsistency may arise from' a smaller temperature increa.se during summer over the prtst 27 years. All three species winter from Mexico to South America. If long-term changes in pheno- logical traits are more common in the “winter portion” of the migration cycle and temperature increase is greater in winter than in summer, terns, which winter in tropical latitudes or move to a southern hemisphere summer, might not be inlluenced by a long-term temperature increase. Their migratory pattern may keep them in Foster et ai • COASTAL BIRD PHENOLOGY AND CLIMATE CHANGE 123 ■o Mean September-October temperature (°C) "<5 > 'C CQ of) FIG. 4. Mean arrival temperature (° C) v.s. first arrival date (Julian days) on Mustang Island, Texas for Double-crested Cormorant (top; P < 0.001) and Least Tern (bottom; P < 0.05) from 1978 to 2005. “summer” temperatures year-round. All three summer residents exhibited a decrease in duration of stay on Mustang Island, although the only significant trend was tor the Least Tern. Contrac- tion of the summer season is opposite of what would be expected in a warming climate where summers should become warmer and longer. However, variability in migration phenologies has been observed in previous studies, where some species exhibit the opposite trend (Parmesan and Yohe 2003). All three terns are known to breed locally, and a decrease in amount of time on Mustang Island may translate to a shortened breeding season. Ramifications of an abridged breeding sea.son for these species are unknown and warrant further investigation. Studies on 124 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. I, March 2010 FIG. 5. Mean April (departure) temperature in Port Aransas, Texas vs. the last date of departure for Eared Grebe from Mustang Island, Texas from 1978 to 2005 (P < 0.01 ). breeding success, clutch size, and food resources should be initiated, especially for the Least Tern which is listed as either threatened or endangered throughout its range. A major concern in conducting phonological studies is the effect that small population size may have on the ability to detect arrival and departure dates (Sparks 1 999, Tryjanowski and Sparks 2001 ). Conclusions on whether or not population size influences detection ability have been mixed and focused on whether a change in population size affects detection of the first anlval date. The variability in airival and departure dates (repre- sented by the standard deviation around the 27-year mean; Table 3) in our study was compared to the mean daily abundance of each species. Mean daily abundance among these six species ranged from 2.77 birds/day for Gull-billed Terns to 81 .84 birds/ day for Least Terns. No significant correlations were found between mean daily abundance and arrival date (P — 0.28, n — 6) or departure date {P — 0.24, n = 6). This suggests that, at least for this study, population size (mean daily abundance) did not cau.se variability in arrival and departure dates, and did not inlluence the ability of the observer to detect migration phenologies for the.se species. Most studies of avian migration phenology have occurred in Britain (Sparks 1999, Cotton 2003) or elsewhere in Europe (Gordo and Sanz 2006, Jonzen et al. 2006) and have focused on passerine species. Information concerning breeding or migration phenology of North American birds is sparse (Brown et al. 1999, Butler 2003). Efforts aie needed to as,sess changes in breeding and migration phenology of North American bird species to help quantify the effects of climate change. These data will allow compari.sons with European studies to examine if climate change is similarly or dispropor- tionately affecting birds on different continents. Studies of coastal birds are also needed to allow comparisons across a greater breadth of avian fauna. ACKNOWLEDGMENTS The authors thank K. H. Dunton, G. .1. Holt, and A. F. Ojanguren for helpful suggestions and critical review of earlier drafts of this manuscript. An earlier draft of this manuscript was also greatly improved through suggestions by C. E. Braun and two anonymous reviewers. CRF thanks his wife and family for their support and encouragement. Funding for analysis of the data was provided by the Sibyl Ranfranz and Kenneth F. Wells Endowed Excellence Fund in Marine Science. This is eonlribution 1510 of The University of Texas Marine Science Institute. Foster et al. • COASTAL BIRD PHENOLOGY AND CLIMATE CHANGE 125 LITERATURE CITED Ahas, R. 1999. Long-term phyto-, ornitho-, and ichthyo- phenologicul tinie-serie.s analyses in Estonia. Interna- tional Journal of Biometeorology 42:1 19-123. Beebee, T. J. C. 1995. Amphibian breeding and climate. Nature 374:219-220. Bradley, N. L., A. C. Leopold, J. Ross, and W. Huffaker. 1999. Phenological changes reflect climate change in Wisconsin. Proceedings of the National Academy of Sciences of the USA 96:9701-9704. Brown, J. L., S. Li, and N. Bhagabati. 1999. Long-term trend toward earlier breeding in an American bird: a response to global warming? Proceedings of the National Academy of Sciences of the USA 96:5565- 5569. Butler, C. J. 2003. The disproportionate effect of global warming on the arrival dates of short-distance migratory birds in North America. Ibis 145:484^95. Cotton, P. 2003. Avian migration phenology and global climate change. Proceedings of the National Academy of Sciences of the USA 100:12219-12222. Gordo, O. and J. J. Sanz. 2006. Climate change and bird phenology: a long-term study in the Iberian Peninsula. Global Change Biology 12:1993-2004. Grabherr, G., M. Gottfried, and H. Paull. 1994. Climate effects on mountain plants. Nature 369:448. IPCC. 2001. Climate Change 2001: third assessment of the Intergovernmental Panel on Climate Change. Cam- bridge University Press, Cambridge, United Kingdom. JoNZEN, N., A. Linden, T. Ergon, E. Knudsen, J. O. Vik, D. Rubolini, D. Piacentini, C. Brinch, F. Spina, L. Karlsson, M. Stervander, a. Andersson, j. Waldenstrom, A. Lehikoinen, E. Edvardsen, R. SoLVANG, AND N. C. Stenseth. 2006. Rapid advance of spring arrival dates in long-distance migratory birds. Science 312:1959-1961. McCarty, J. P. 2001. Ecological consequences of recent climate change. Conservation Biology 15:320-331. Parmesan, C. and G. Yohe. 2003. A global ly-coherent footprint of climate change across natural systems. Nature 421 :37^2. Parmesan, C., N. Ryrholm, C. Stefanescus, J. K. Hill, C. D. Thomas, H. Descimon, B. Huntley, L. Kaila, j. Killberg, T. Tammaru, W. J. Tennent, J. A. Thomas, and M. Warren. 1999. Poleward shifts in geographic ranges of butterfly species associated with regional warming. Nature 399:579-583. Roy, D. B. and T. H. Sparks. 2000. Phenology of British butterflies and climate change. Global Change Biology 6:407^16. Sparks, T. H. 1999. Phenology and the changing pattern of bird migration in Britain. International Journal of Biometeorology 42:134-138. Sparks, T. H., F. Bairlein, J. G. Bojarinova, O. Huppop, E. A. Lehikoinen, K. Rainios, L. V. Sokolov, and D. Walker. 2005. Examining the total arrival distribution of migratory birds. Global Change Biol- ogy 1 1 :22-30. SYSTAT Software Inc. 2004. SYSTAT Version 11.0. SYSTAT Software Inc., Richmond, California, USA. Tryjanowski, P. and T. H. Sparks. 2001. Is the detection of the first arrival date of migratory birds influenced by population size? A case study of the Red-backed Shrike Lanins collwio. International Journal of Biometeorology 45:217-219. Walther, G. R., E. Post, P. Convey, A. Menzel, C. Parmesan, T. J. C. Beebee, J.-M. Fromentin, O. Heogh-Guldberg, and F. Bairlein. 2002. Ecological responses to recent climate change. Nature 416:389- 385. WuETHRiCH, B. 2000. How climate change alters the rhythms of the wild. Science 287:793-795. The Wilson Journal of Ornithology 122( 1 ): 126-1 34, 2010 NEST-SITE LIMITATION OF SECONDARY CAVITY-NESTING BIRDS IN EVEN-AGE SOUTHERN PINE FORESTS KARL E. M1LLER‘- ABSTRACT. — 1 used a randomized, replicated, controlled study design to test the hypothesis that nest-site availability limits breeding densities ot secondary cavity nesting species in even-age southern pine (Pinus spp.) forests. Breeding densities of secondary cavity nesters increased significantly on treated plots after nest boxes were introduced. Total number of nesting attempts also increased several told post-treatment. These data indicate nest-site availability was a limiting factor tor breeding densities of secondary cavity nesting species. The response ot individual species to nest boxes ranged from moderately high (Great Crested Flycatcher [Myuirchus crinitiis]), to low (Tufted Titmouse [Baeolophus hicolor]. Eastern Bluebird [Siaha sialis]) to no response (Carolina Wren \Thryothonis htdovicianus]). At least three factors accounted for interspecific ditferences in this study: different levels of reliance on cavities excavated in snags, different body sizes, and differences in local population densities. The large number of unoccupied nest boxes (only 9% were used for nesting) suggests secondary cavity nesters were not limited solely by cavity availability but also by habitat quality. Prescribed burning appeared to facilitate discovery and use of nest boxes by birds in this study, consistent with the hypothesis that nest- site limitation is mitigated by habitat structure. Received 4 September 2007 . Accepted 3 September 2009. Nest sites are often assumed to be the primary factor limiting densities of secondary (i.e., non- excavating) cavity-nesting birds (e.g., von Haart- man 1957, 1971; Thomas et al. 1979; Cody 1985). However, tests of the nest-site limitation hypoth- esis by manipulative experiments, which include knowledge of initial conditions, controls, and replication (Hairston 1989), are still rare. Newton (1994) reviewed the evidence for nest-site limita- tion in cavity-nesting birds and found most to be circumstantial. Newton (1994) found only six studies that added or subtracted nest sites for cavity-nesting birds where controls and pre- and post-manipulation data were included. Only two of these studies (Enemar and Sjbstrand 1972, Brawn and Baida 1988) had replicate study plots. Other common bia.ses in tests of the nest-site limitation hypothesis include lack of spatial independence of experimental and control plots (e.g.. East and Peirins 1988, Beis.senger and Bucher 1992), and inadequate controls because of placement of study plots in dissimilar plant communities (e.g., Biaish 1983). Inadequate experimental controls can leave apparent experimental effects open to alternative explanations such as temporal and spatial variation in habitat quality. Most nest-site limitation studies have been conducted on a small number of species inhabit- ' Department of Wildlife Ecology and Conservation, University of Florida. Gainesville, FI. 3261 I, USA. ’Current address: Fish and Wildlife Research Institute, Florida Fish and Wildlife Conservation Commission, I lO,*) Southwest Williston Road, Gainesville, Ft. 32601, USA; e-mail: karl . m i I lcr@ myfwc.com ing simplified European forests, especially Great Tit (Pants major) and European Pied Flycatcher (Ficedula hypoleuca), and often focusing on only one species at a time (Wesolowski 2007). Controlled nest box experiments have not been conducted with cavity-nesting birds in pine (Pinus spp.) forests of the southeastern United States. Research from other forest systems indicates that cavity density is positively con'elated with tree density and tree age (van Balen et al. 1982), and natural cavities formed from decay are more abundant in hardwood trees than in conifers (Waters et al. 1990). Thus, secondary cavity nesters should be more likely to be nest-site limited in younger forests and in forests dominat- ed by conifers than in older, more structurally complex forests with abundant natural cavities (Brawn and Baida 1988, Waters et al. 1990, Wesolowski 2007). I predicted that secondary cavity nesting species would be nest-site limited in pine forests of the southeastern United States, where silvicul- tural practices are considered to limit the availability of dead trees that provide nest sites for cavity nesters (McComb et al. 1986, Jackson 1988). The objective of my study was to test the hypothesis that nest-site availability limits breed- ing densities of secondary cavity nesting species in even-age managed pine forests, which are the predominant forest type in the southeastern coastal plain. METHODS Study Plots and Experimental Design. — Field- work was conducted on 12 study plots in 35- 126 Miller - NEST-SITE LIMITATION IN NON-EXCAVATORS 127 40 year-old even-age slash pine (Finns elliottii) plantations at Camp Blanding Training Site, Clay County, Florida. I distributed study plots evenly between mesic “tlatwoods” sites and drier “sandhill” sites to control for variation in site productivity. Flatwoods plantations were charac- terized by moderately to poorly drained soils and a dense shrub layer of gallberry (Ilex glabra), saw palmetto (Serenoa repens), and ericaceous shrubs (Abrahamson and Hartnett 1990). Sandhill plan- tations occLiiTed on sites originally dominated by longleaf pine (P. palustris) (Myers 1990) and were characterized by well-drained sandy soils and a more open shrub layer. I measured live vegetation within each study plot at point-count stations (Martin et al. 1997) to develop stand profiles. Flatwoods plantations were densely stocked with a basal area (T ± SE) of 31.0 ± 2.2 m%a. Sandhill plantations were moderately stocked with a basal area of 24.5 ± 2.5 mVha. All plantations had been unburned for >3 years. I counted snags >10 cm diameter at breast height (dbh) on two 50 X 400 m transects within each stand. Snags were abundant (17.7 ± 2.0 [x ± SE] snags/ha on flatwoods plots, 14.8 ± 2.1 snags/ha on sandhill plots) relative to the average of 6.2 snags/ha reported for pine plantations on public land throughout Elorida (McComb et al. 1986). Most snags were too small for excavation by woodpeckers; <2% were ^25 cm dbh (Miller 2000). More than 97.5% of all snags counted were slash pines with the remainder comprising turkey oak (Quercus laevis) and longleaf pine. Study plots were of equal size (10 ha) and dimensions (250 X 400 m) and as far as possible (typically >75 m) from dirt roads and other openings. Study plots were broadly dispersed to encourage spatial independence; the northernmost study plot was 13.1 km distant from the south- ernmost plot, and the westernmost study plot was 9.2 km distant from the easternmost plot. Nearest- neighbor distances between study plots averaged 1.2 km (range: 0.5— 2.6 km). Dense wetland forest separated the two study plots, that were only 0.5 km apart. I applied treatments randomly in a balanced design (Table 1 ) after 2 years of breeding bird surveys to document initial conditions (1995- 1996). Prescribed burns were conducted on two flatwoods plots and two sandhill plots during January-February 1997. Winter bums consumed leaf litter and pruned or killed shrubs and saplings, but did not cause tree mortality. I TABLE I. Experimental design used for test of nest- site limitation. Treatments (O = no manipulation; X = nest boxes added; Y = nest boxes added after control burn) were randomly applied within forest type. Clay County, Florida. 1995-1998. Plot # 1995 1996 1997 1998 Flatwoods 1 o o 0 o 2 0 0 0 0 3 o o X X 4 0 0 X X 5 o o Y Y 6 o 0 Y Y Sandhills 7 o o o O 8 o 0 0 O 9 o 0 X X 10 o 0 X X 11 o 0 Y Y 12 o 0 Y Y installed nest boxes during the first week of March 1997 on all burned plots and on four of the unburned plots, leaving four unburned plots without nest boxes as controls (Table 1). Two years of breeding bird surveys were conducted to assess the effects of treatments (1997-1998). Nest Boxes. — constructed 320 wood nest boxes with dimensions that would accommodate all species of secondary cavity nesters that breed in Florida pinelands, except raptors and water- fowl. Nest boxes differed in inside floor dimen- sions (10.2 X 8.9 cm or 14.0 X 14.0 cm), diameter of the entrance hole (3.5 or 5.1 cm), and in height above ground (1.9 or 4.8 m). I installed 40 nest boxes (4/ha) on each nest-box plot, at 50-m intervals in a standardized an’ay. The nest box array was a balanced, three-factor design: box size (small vs. large) X hole size (naiTOw vs. wide) X box height (low vs. high). 1 standardized the orientation of nest boxes, placing them with the entrance hole oriented east by southeast, because cavity openings facing east, southeast, .south, or southwest are typical for many cavity nesting species (Conner 1975, Pinkowski 1976, McEllin 1979, Jackson 1994, Rendell and Ro- bertson 1994). Nest Searches and Nest Monitoring. — used standard methods (Martin and Geupel 1993) to search for nests in tree cavities on study plots from early April through early July 1996-1998. Each nest located was monitored regularly at 3- 4 day intervals to assess nesting status (Martin 128 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. I. March 2010 and Geupel 1993, Ralph et al. 1993). I used an aluminum extension ladder to inspect the contents ot each nest box once every 10-14 days during April-May and once every 14-21 days during June-early July 1997-1998. Nests in nest boxes also were monitored regularly at 3^ day inter- vals. A nesting attempt was defined as a nest where >1 egg was laid. Southern flying squirrels {Glaucomys volans), snakes, and other non-avian occupants were not discouraged from using nest boxes, because I wanted to study cavities as a limiting factor for birds under natural conditions. Squirrel “roost sites” were defined as those nest boxes in which flying squirrels were observed on >2 occasions within a season. Estimation of Bird Densities. — I used the intensive point count method (Ralph et al. 1993, Wilson et al. 1995) to estimate densities of breeding birds in study plots. Intensive point counts differ from extensive point counts in that points are established relatively close together within a nest search plot or mist net plot, and count data within each plot are pooled for analyses (Ralph et al. 1993). Point-count survey methods followed Ralph et al. (1993): counts were conducted within 3 hrs post-sunrise; no counts were made during rain or strong winds; birds counted within a fixed-width circle were not recounted if they were observed moving to an adjacent fixed-width circle. 1 conducted all point- count surveys to eliminate inter-observer bias. 1 .selected a 9-min count period because pilot studies indicated that 84% of the species in this system were detected within the first 9 min (unpubl. data); a longer count period would increase the likelihood of double counting (Scott and Ramsey 1981). Data from pilot studies demonstrated the effective detection distance was >35m for most species, including all cavity-nesting species. 1 established six 35-m radius point-count stations spaced 125 m apart within each plot (Hutto et al. 1986). Each point- count station was sampled three times per year al intervals of 2.5 weeks between 20 April and the second week of June. Analyses of Experimental Effects. — I pooled fixed-radius point count data during each visit to a study plot because each plot was considered an experimental unit for analyses (Ralph et al. 1993). I used the maximum pooled count for each species within each plot for compari.sons among plots, treatments, and years. Count data were not transformed for analyses because of homogeneity of variances across treatments and years. I analyzed experimental effects with a mixed model analysis-of-variance (ANOVA) using generalized least squares estimation (PROC MIXED, SAS Institute Inc. 1997). The model included four factors: treatment (control vs. nest box vs. burned with nest box), forest (flatwoods vs. sandhills), year, and a random plot effect. 1 modeled the differences in count data between pre- (1995, 1996) and post-treatment years (1997, 1998) using an average of the 2 pre-treatment years as a covariate (sensitivity analysis indicated that pre- treatment years did not differ). Post-treatment years were treated separately in the ANOVA because I wanted to examine if responses to treatments changed over time. Goodness-of-fit criterion (Schwarz’s Bayesian Criterion) indicated a model including interaction terms (treatment X forest, treatment X year) best described the data for each species. Pairwise differences of least- squares means (PROC MIXED) identified the relationships between variables for significant factors in the model, and P-values < 0.10 were considered significant. RESULTS Response of Cavity-nesting Bird Populations to Nest Box Treatments. — I detected 14 species of cavity nesters during point-count surveys. Three species of woodpeckers (Red-bellied Woodpecker [Melanerpes carolinus]. Downy Woodpecker [Pi- coides puhescens]. Northern Elicker [Colaptes auratus]), two species of weak excavators (Car- olina Chickadee [Poecile carolinensis], and Brown-headed Nuthatch \Sitta pusilla\), and five species of non-excavators (Eastern Screech-Owl \Megascops asio\. Great Crested Elycatcher \Myiarchus crinitus]. Tufted Titmouse [Baeolo- plius hicolor], Carolina Wren \Thryotiioriis liidovi- f77//?//.v|, and Eastern Bluebird {Sialia sialis]) were documented nesting on study plots. 1 categorized Carolina Chickadees and Brown-headed Nuthatch- es as primary (i.e., excavating) cavity nesters because of their strong affinity for cavity excava- tion in this locality. The other four species detected (Barred Owl \Strix varia]. Red-headed Woodpeck- er [M. erythroceplialiis], Red-cockaded Wood- pecker \P. borealis], and Pileated Woodpecker \Dryocopus pileatus]) were rare (recorded on < 1 % of surveys) and did not nest on study plots. Counts of primary cavity nesting species did not vary among treatments, either for all species Miller - NEST-SITE LIMITATION IN NON-EXCAVATORS 129 combined (P = 0.23) or for any individual species {P > 0.10). No primary cavity nester used nest boxes for nesting. Red-bellied and Downy wood- peckers accounted for 93% of all woodpeckers counted within 35-m radius point counts. Counts of secondary cavity nesting species varied among treatments. The mixed model ANOVA for all species combined indicated that treatment was the only significant factor (F — 7.57, df = 2,5; P = 0.03) explaining the change in counts between pre-treatment (1995, 1996) and post-treatment years (1997, 1998); secondary cavity nesters increased on treated plots but decreased slightly on control plots (Fig. 1). The treatment X year interaction term approached statistical significance (F = 2.89, df = 2,9; P = 0.11), reflecting the different rates of response on unburned and burned plots. Secondary cavity nesters did not increase on unburned plots with nest boxes (NB) until 1998 but increased nearly twofold on burned plots with nest boxes (BNB) in 1997 (Fig. 1). Differences of least-squares means, post-treatment, indicated that counts {x ± SE) were greater (F = 0.01) in NB plots (5.5 ± 0.7) than in control plots (2.9 ± 0.2) and greater (F = 0.02) in BNB plots (5.0 ± 0.4) than in control plots. Counts did not differ between NB plots and BNB plots. Interspecific Variation in Response to Nest Box Treatments. — ^Treatment was the only signif- icant factor in the mixed model (F = 4.42, df = 2,5; F = 0.07) for Great Crested Flycatchers. This species increased on treated plots after treatments were applied and decreased slightly on control plots (Fig. 2). Counts in NB plots (2.6 ± 0.4) post-treatment were greater (F = 0.02) than counts in control plots (1.3 ± 0.3), and counts in BNB plots (2.9 ± 0.4) were greater (F = 0.03) than counts in control plots. Great Crested Flycatcher numbers reached their peak during the study in the BNB plots during the first year after boxes were introduced (Fig. 2). Tufted Titmou.se response was similar to that of Great Crested Flycatchers; counts increased three- fold on treated plots after treatments were applied (Fig. 2). Tufted Titmice were not recorded in fixed-radius point counts before nest boxes were introduced on any BNB plots in 1995 and on only one BNB plot in 1996 (Fig. 2). Treatment was not significant (F = 0.37, df = 2,5; F = 0.70), in part because of the small samples involved. Eastern Bluebirds were not recorded on study plots until after nest boxes were introduced Year FIG. 1. .Mean number of secondary cavity-nesting (SCN) birds on study plots before (1995-1996) and after (1997- 1998) treatments (C = control, NB = unburned plots with nest boxes, BNB = burned plots with nest boxes). Bars equal one SE. (Fig. 2), which suggests they responded to the nest boxes. No factors were significant, but the statistical analysis had low power because of small samples. Carolina Wrens had no response to treatments (Fig. 2). The Carolina Wren was the only secondary cavity nester species that did not exhibit an immediate response to nest boxes in BNB plots in 1997. Forest type was the only significant factor (F = 9.62, df = 1,9; F = 0.01) explaining the differences in wren counts; wrens were more common in flatwoods plots than in sandhill plots, regardless of treatment. Eastern Screech-Owls were not monitored effectively by point-count surveys and sample sizes were insufficient for statistical analyses. Patterns of Nest Box Occupancy. — recorded 57 nesting attempts in nest boxes during 1997- 1998, including 33 (58%) Great Crested Flycatch- er nests, 13 (23%) Tufted Titmouse nests, six (11%) Eastern Bluebird nests, three (5%) Car- olina Wren nests, and two (3%) Eastern Screech- Owl nests. Screech-Owls were only able to use those nest boxes that had entrance holes enlarged by squirrels. Red-bellied Woodpeckers also were documented roosting in 19 nest boxes; these nest boxes contained fecal debris and typically were near (<50 m) an active Red-bellied Woodpecker nest in a pine snag. Overall avian occupancy of ne.st boxes, including woodpecker roost sites, increased from the first year (8%) to the second year (16%). Flying squirrels used 70 (22%) and 56 (18%) nest boxes as roost sites in 1997 and 1998, respectively. The total number of secondary cavity nests on study plots increa.sed twofold in 1997 (28 total nests) and threefold in 1998 (42 total nests) 130 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 1. March 2010 A B 1995 1996 1997 1998 c CO < LU 4 3 2 1 0 1995 1996 1997 1998 D « 4 -1 1995 1996 1997 1998 FIG. 2. Mean number of (A) Great Crested Flycatchers (GCFLs), (B) Tufted Titmice (TUTls), (C) Eastern Bluebirds (EABLs), and (D) Carolina Wrens (CAWRs) on study plots before (1995-1996) and after (1997-1998) treatments (C = control, NB = unburncd plots with nest boxes, BNB = burned plots with nest boxes). Bars equal one SE. Miller - NEST-SITE LIMITATION IN NON-EXCAVATORS 131 because of the addition of nest boxes (Table 2). Nesting activity on NB plots lagged behind that on BNB plots; cavity nesters made little use of nest boxes on NB plots until the second year after the nest boxes were in place (Table 2). DISCUSSION Response of Cavity-nesting Bird Populations to Nest Box Treatments. — Numbers of secondary cavity nesters recorded during the breeding season increased significantly on treated plots after nest boxes were introduced. The number of nesting attempts on treated plots also increased several fold. These data indicate that nest-site availability was limiting breeding densities of secondary cavity nesters. It is unlikely that birds moved from control to treated plots because interplot distances were many times larger than the territory sizes of these species (Grubb and Pravosudov 1994, Lanyon 1997, Gowaty and Plissner 1998). Even if a few individuals moved between plots, the slight decline in number of cavity nests on control plots (from 4 in 1996 to 3 in each of the following years; Table 2) cannot account for the large increase in the number of total nests on treated plots during 1997-1998. Point-count surveys indicated the bird commu- nity was lacking in large-bodied woodpeckers (i.e.. Northern Flicker, Pileated Woodpecker) that excavate cavities used by secondary cavity nesters. The Northern Flicker is the primary creator of large cavities used by other species in longleaf pine forest (Blanc and Walters 2008), especially larger species such as American Kestrel (Falco sparverius), which was absent from my study site. TABLE 2. Nests of secondary cavity-nesting birds (Eastern Screech-Owl, Great Crested Flycatcher, Tufted Titmouse, Carolina Wren, Eastern Bluebird) in tree cavities and nest boxes, before (1996) and after (1997-1998) treatments were applied. Clay County, Florida. Treatment Pre- treatmeni Post-treatment years 19% 1997 1998 Cavity Box Cavity Box Cavity Box Control plot.s in = 4) 4 3 a 3 Nest box plots [n = 4) 6 3 5 2 16 Burned nest box plots (n = 4) 4 1 16 1 20 Totals 14 7 21 6 36 Nesl boxes not available. Interspecific Variation in Nest-site Limitation. — Population response to nest boxes by individual species ranged from moderately high (Great Crested Flycatcher) to low (Tufted Titmouse, Eastern Bluebird) to no response (Carolina Wren). At least three factors accounted for interspecific differences in this study; different levels of reliance on cavities excavated in snags, different body sizes, and differences in local population densities. Great Crested Flycatchers responded more favorably to addition of nest boxes in pine plantations than Tufted Titmice and Eastern Bluebirds, indicating they were more nest-site limited. Similarly, Ash-throated Flycatchers {Myiarchus cinerascens) increased from 0 to three breeding pairs after 20 nest boxes were added to a 16-ha plot of riparian forest that was devoid of cavities (Brush 1983). Myiarchus nest-site selec- tion is sufficiently generalized to allow use of hollow branches and other natural cavities if they are of sufficient size (Lanyon 1997), but natural cavities available in 35--40 year-old slash pine trees apparently were too small for this species. Flycatchers relied heavily on woodpecker-exca- vated cavities; only one (6%) of 18 flycatcher cavity nests monitored was in a natural cavity in a living pine tree (Miller 2000). Tufted Titmice increased in numbers following introduction of nest boxes but not as dramatically as Great Crested Flycatchers. Nest monitoring indicated that three of six (50%) titmouse nests in slash pine trees were in shallow slit cavities formed between forked pine trunks (Miller 2000). The smaller body size of the Tufted Titmouse may have allowed it to be more opportunistic in its use of natural cavities than the Great Crested Flycatcher. Eastern Bluebirds showed only slight increases on treated plots after nest boxes were introduced. However, Eastern Bluebirds were rare on the study plots and it was difficult to assess whether they were nest-site limited. Local or regional population densities for rare species may be insufficient to provide surplus or floater individ- uals that would breed if more nest sites were available (Newton 1998). Carolina Wrens had no response to treatments and, despite their abundance, rarely used nest boxes. Nest searches indicated the Carolina Wren was a facultative cavity nester at this locality, typically building its nests in crevices in brush piles, stumps, and shrubs or weaving them into 132 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. I. March 2010 palmetto fronds (Miller 2000). Only one wren nest was found in a tree cavity during 1996-1998, a ‘chimney’ nest {sensu Wesolowski 2007) built in a low hollow stump. Ejf’ect of Habitat on Nest-site Limitation. — ^The large number of unoccupied nest boxes (only 9% were used for nesting) suggests that secondary cavity nesters were not limited solely by cavity availability but also by habitat quality (e.g., unsuitable habitat structure, low food availabili- ty). Studies in western forests have demonstrated a greater response to nest boxes. For example, bird use of nest boxes in thinned ponderosa pine (P. ponderosa) forest in Arizona increased from 10% in the first year of their availability to 58% in the fourth year (Brawn and Baida 1988). Nest box occupancy rates in mixed oak-pine wood- lands in California increased from 25% in the first year to 68% in the sixth year (Purcell et al. 1997). Extensive tree cutting during 1999-2000 curtailed my experiment, but I continued monitoring three plots unaffected by forestry operations and nest box occupancy rates during 1999-2000 did not exceed 13%. Thus, low nest box occupancy rates persisted for several years, indicating that habitat quality may .set an upper limit on breeding bird density in these young forests. Populations of several cavity nester species, especially Eastern Bluebird, were low in these densely-stocked slash pine plantations, possibly because the forest may not match the niche structure {sensu Grinellian niche, James et al. 1984) prefen'ed by these species. Eor example, the Eastern Bluebird prefers forest gaps, forest edges, and other plant communities that have few trees and sparse ground cover (Gowaty and Plissner 1998). Birds in flatwoods plantations and sandhill plantations responded similarly to treatments despite modest site differences in basal area and snag availability. Regardless of forest type, the respon.se to nest boxes was most rapid on burned plots, probably becau.se birds were attracted to areas where winter burns had opened up the shrub layer and iinderstory. Populations of many secondary cavity nesting species respond favor- ably to open microhabitats created by burning (e.g., Bock and Lynch 1970, Pinkowski 1976, Hutto 1995, Krei.sel and Stein 1999, Provencheret al. 2002, Allen et al. 2006). Other Factors Potentially Limiting, Cavity- nestOip, Bird Populations. — Breeding populations of cavity-nesting birds potentially can be limited by other factors including territoriality and interspecific competition for nest sites (Newton 1979, 1998). However, ten'itorial behavior at present bird densities is unlikely to have prevent- ed individuals from breeding. Eor example, territory mapping on study plots (unpubl. data) demonstrated that Great Crested Elycatcher terri- tories frequently were not contiguous. The large number of nest boxes unused by birds or flying squirrels ( —two-thirds of the total each year) also suggested little potential for interference compe- tition among species. 1 rarely observed any evidence of inter- or intra-specific usuipation of nest boxes, although flying squiirels did depredate some flycatcher nests (Miller 2002). Interference competition for nest sites was not observed among secondary cavity nesting species at natural nests in tree cavities, although 1 did observe several aggressive interactions between Red-bellied Woodpeckers and Great Crested Elycatchers at Red-bellied Woodpecker nest cavities. Nest-site limitation should not be assumed for all cavity-nesting bird species, given that some species are more opportunistic in their choice of nest sites than others. Reliance on cavities excavated in dead wood by primary cavity-nesters and larger body size were factors associated with nest-site limitation. Supporting evidence comes from studies in ponderosa pine forests, where secondary cavity nesting species that nested almost exclusively in cavities excavated by primary cavity nesters (i.e., Violet-green Swallow [Tachycineta thalassina]\ Pygmy Nuthatch [Sitta pygmaea]-. Western Bluebird [Sialia me.xicana)) exhibited greater population responses to nest boxes than species that nested primarily in natural cavities and crevices (Brawn and Baida 1988). Nest-site limitation cannot be assumed for all plant communities and locales. Tree cavities are most abundant in older forest systems, which tend to be structurally more complex (Aitken and Martin 2007, Wesolowski 2007). Results of my study provide additional support for the hypoth- esis that secondary cavity-nesting birds are most nest-site limited in younger forests and in forests dominated by conifers (Brawn and Baida 1988, Waters et al. 1990, Wesolowski 2007). Pre.scribed burning appeared to facilitate dis- covery and Li.se of nest boxes by birds in my study, which is consistent with the hypothesis that nest- site limitation is mitigated by habitat structure (Brawn and Baida 1988). If increasing cavity- nesting bird populations is a management goal in Miller - NEST-SITE LIMITATION IN NON-EXCAVATORS 133 t'ire-mainlained habitats, such as southern pine forests, managers may find that addition of nest boxes to newly burned sites will result in the highest local densities of secondary cavity-nesting birds, at least in the short-term. ACKNOWLEDGMENTS My research was suppoited by funding from the Florida Fish and Wildlife Conservation Commission, Florida Army National Guard. University of Florida. North American Bluebird Society, and a Student Research Grant from Sandpiper Technologies Inc. I thank M. P. Moulton for advice and encouragement throughout this project. G. A. Jones, M. I. Williams, S. A. Johnson, and Annemarie van Doom assisted in the field. K. M. Portier assisted with statistical analy.ses. L. A. Blanc, D. L. Leonard, D. J. Levey, M. P. Moulton, K. E. Sieving, G. L. Slater, and an anonymous reviewer made constructive comments on earlier versions of the manuscript. I dedicate this paper to my father, J. H. Miller, who taught me the value of hard work. LITERATURE CITED Abrahamson, W. G. and D. C. FIartnett. 1990. Pine flatwoods and dry prairies. Pages 103-149 in Ecosys- tems of Florida (R. L. Meyers and J. J. Ewel, Editors). University of Central Elorida Press, Orlando, USA. Aitren, K. E. H. and K. Martin. 2007. The importance of excavators in hole-nesting communities: availability and use of natural tree holes in old mixed forests of western Canada. Journal of Ornithology 148 (Supple- ment 2):S425^34. Allen. J. C., S. M. Krieger, J. R. Walters, and J. A. Collazo. 2006. As.sociations of breeding birds with fire-influenced and riparian-upland gradients in a longleaf pine ecosystem. Auk 123:1 1 10-1 128. Beissinger, S. R. and E. FI. Bucher. 1992. Sustainable harvesting of parrots for conservation. Pages 73-1 15 in New World parrots in crisis: .solutions in conserva- tion biology (S. R. Beissinger and N. F. R. Snyder, Editors). Smithsonian Institution Press, Washington, D.C., USA. Blanc. L. A. and J. R. Walters. 2008. Cavity-nest webs in a longleaf pine ecosystem. Condor 1 10:80-92. Bock, C. E. and J. F. Lynch. 1970. Breeding bird populations of burned and unburned conifer forests in the Sierra Nevada, Condor 72:182-189. Brawn. J. D. and R. P. Bai.da. 1988. Population biology of cavity nesters in northern Arizona: do nest sites limit breeding densities? Condor 90:61-71. Brush. T. 1983. Cavity use by secondary cavity-nesting birds and response to manipulations. Condor 85:461-466. Cody, M. L. 1985. An introduction to habitat selection in birds. Pages 3-56 in Habitat .selection in birds (M. L. Cody, Editor). Academic Press. Orlando. Florida, USA. Conner, R. N. 1975. Orientation of entrances to wood- pecker nest cavities. Auk 92:371-374. East, M. L. and C. M. Perrins. 1988. The effect of nestboxes on breeding populations of birds in broad- leaved temperate woodlands. Ibis 130:393-401. Enemar, a. and B. Sjostrand. 1972. Effects of the introduction ol Pied Elycatchers Ficedulu liypolenca on the composition of a passerine bird community. Ornis Scandinavica 3:79-89. Gowaty, P. a. and j. H. Plissner. 1998. Eastern Bluebird (Sialia sialis). The birds of North America. Number 381. Grubb, T. C. and V. V. Pravosudov. 1994. Tufted Titmou.se (Baeolophus hicolor). The birds of North America. Number 86. Hairston Sr., N. G. 1989. Ecological experiments: purpose, design, and execution. Cambridge University Press, Cambridge, United Kingdom. Hutto, R. L. 1995. Composition of bird communities following stand replacement fires in northern Rocky Mountain conifer forests. Conservation Biology 9:1041-1058. Hutto, R. L., S. M. Pletschet, and P. Hendricks. 1986. A fixed-radius point count method for nonbreeding and breeding season use. Auk 103:593-602. Jackson, J. A. 1988. The southeastern pine forest ecosystem and its birds: past, present, and future. Pages 119-159 in Bird conservation 3 (J. A. Jack.son, Editor). University of Wisconsin Press, Madison, USA. Jackson, J. A. 1994. Red-cockaded Woodpecker (Picoides borealis). The birds of North America. Number 85. James, F. C., R. F. Johnston, N. O. Wamer, G. J. Niemi, and W. j. Boecklen. 1984. The Grinnellian niche of the Wood Thrush. American Naturalist 124:17-30. Kreisel, K, j, and S. j. Stein. 1999. Bird use of burned and unburned coniferous forests during winter. Wilson Bulletin 1 1 1 :243-250. Lanyon, W. E. 1997. Great Crested Flycatcher (Myiarchus crinitm). The birds of North America. Number 300. Martin, T. E. and G. R. Geupel. 1993. Nest-monitoring plots: methods for locating nests and monitoring success. Journal of Eield Ornithology 64:507-519. Martin, T. E., C. R. Paine, C. J. Conway, W. M. Hochachka, P. Allen, and W. Jenkins. 1997. BBIRD field protocol. Montana Cooperative Wildlife Research Unit. University of Montana, Mis.soula. USA. McComb, W. C., S. A. Bonney, R. M. Sheffield, and N. D. Cost. 1986. Snag resources in Florida — are they sufficient for average populations of primary cavity- nesters? Wildlilc Society Bulletin 14:40^8. McEllin, S. M. 1979. Nest sites and population demog- raphics of White-breasted and Pigmy nuthatches. Condor 8 1 :348-352. Miller, K. E. 2000. Nest-site limitation, nest predation, and nest-site selection in a cavity-nesting bird community. Dis.sertation. University of Florida, Gaines- ville. USA. Miller, K. E. 2002. Nesting success of the Great Crested Flycatcher in nest boxes and in tree cavities: are nest boxes safer from nest predation? Wilson Bulletin 1 14:179-185. Myers, R. L. 1990. Scrub and high pine. Pages 150-193 in Ecosystems of Elorida (R. L. Meyers and J. J. Ewel, 134 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 1, March 2010 Editors). University of Central Florida Press, Orlando, USA. Newton, I. 1979. Population ecology of raptors. T. and A. D. Poyser, London, United Kingdom. Newton, I. 1994. The role of nest sites in limiting the numbers of hole-nesting birds; a review. Biological Conservation 70:265-276. Newton, I. 1998. Population limitation in birds. Academic Press, San Diego, California, USA. PlNKOWSKl, B. C. 1976. The use of tree cavities by nesting Eastern Bluebirds. Journal of Wildlife Management 40:556-563. Provencher, L., N. M. Gobris, L. A, Brennan, D, R. Gordon, and J. L. Hardesty. 2002. Breeding bird response to midstory hardwood reduction in Florida sandhill longleaf pine forests. Journal of Wildlife Management 66:641-661. Purcell, K. L., J. Verner, and L. W. Oring. 1997. A comparison of the breeding ecology of birds nesting in boxes and tree cavities. Auk 1 14:646-656. Ralph, C. J., G. R. Geupel, P. Pyle, T. E. Martin, and D. F. Desante. 1993. Handbook of field methods for monitoring landbirds. USDA, Forest Service, General Technical Report PSW-GTR-144. Pacific Southwest Re.search Station, Albany, California, USA. Rendell, W. B. and R. j. Robertson. 1994. Cavity- entrance orientation and nest-site use by secondary hole- nesting birds. Journal of Field Ornithology 65:27-35. SAS Institute Inc. 1997. SAS/STAT guide for personal computers. Version 6.12. SAS Institute Inc., Cary, North Carolina, USA. Scott, J. M. and F. L. Ramsey. 1981. Length of count period as a possible source of bias in estimating bird densities. Studies in Avian Biology 6:409- 413. Thomas, J. W., R. G. Anderson, C. Maser, and E. L. Bull. 1979. Snags. Pages 60-67 in Wildlife habitats in managed forest — the Blue Mountains of Oregon and Washington (J. W, Thomas, Editor). USDA, Forest Service, Agriculture Handbook Number 553. VAN Balen, j. H., C. j. H. Booy, j. A. Franeker, and E. R. OsiECK. 1982. Studies on hole-nesting birds in natural nest sites. 1. Availability and occupation of natural nest sites. Ardea 70:1-24. VON Haartman, L. 1957. Adaptation in hole-nesting birds. Evolution 11:339-347. VON Haartman, L. 1971. Population dynamics. Pages 391-459 in Avian biology (D. S. Earner and J. R. King, Editors). Volume 1. Academic Press, New York, USA. Waters, J. R., B. R. Noon, and J. Verner. 1990. Lack of nest-site limitation in a cavity-nesting bird community. Journal of Wildlife Management 54:239-245. Wesolowski, T. 2007. Lessons from long-term hole-nester studies in a primeval temperate forest. Journal of Ornithology 148 (Supplement 2):S395-S405. Wilson, C. W., R. E. Masters, and G. A. Buken- HOFER. 1995. Breeding bird response to pine- grassland community restoration for Red-cockaded Woodpeckers. Journal of Wildlife Management 59: 56-67. The Wilson Journal of Ornithology 1 22( I ): 1 35- 1 38, 2010 HOUSE SPARROWS ASSOCIATED WITH REDUCED CLIEE SWALLOW NESTING SUCCESS DOUGLAS R. LEASURE,' RAGUPATHY KANNAN,' AND DOUGLAS A. JAMES^' ABSTRACT. — We quantified the impact of nesting and roosting House Spanows (Passer domesticus) on nesting success of Cliff Swallows (Petroehelidon pyrrhonota) in colonies in western Arkansas in 2007 and 2008. Two sections of a large swallow colony under a bridge with House Sparrows were compared in 2007 to two sections with little House Sparrow usage. Nesting success of Cliff Swallows (percent of nests yielding at least 1 chick) was 61% in sections with low House Sparrow activity, significantly higher than the 30% in sections with high House Sparrow activity. House Sparrows defended a broad zone surrounding their nests from Cliff Swallow nesting attempts. We compared the proportion of nests used, clutch sizes, and brood sizes of Cliff Swallows in two colonies in 2008, one with and one without House Sparrow activity. In the colony without House Sparrow activity, 48% of old and new nests were used by swallows versus only 8% in the colony with House Sparrows. Swallow clutch sizes were similar in the two colonies, but swallow brood sizes in the colony with no House Sparrows were significantly higher, mean = 2.3 nestlings per nest (mode = 2; 75th percentile = 3) compared to 0.8 nestlings (mode = 0; 75th percentile = 1) in the colony with House Sparrows. This suggests Cliff Swallows are less successful when House Spamows are present in colonies. Received 3 April 2009. Accepted 27 September 2009. Cliff Swallow (Petroehelidon pyrrhonota) nest- ing colonies can contain up to 3,500 gourd-shaped mud nests adhering to rock overhangs, cliff faces, and concrete structures such as bridges and culverts (Brown and Brown 1995). Their breeding range was historically limited to mountain ranges in western North America but urbanization produced a myriad of suitable Cliff Swallow nesting sites, which allowed range expansion eastward to include most of North America. House Sparrows (Passer domesticus) were intro- duced from Europe into North America in the 1850s and 1880s (Lowther and Cink 1992). Their preference for human altered environments has enabled them to spread throughout North and Central America. House Sparrows in North America are known to have interfered with nesting of several species of native birds (Barrows 1889, Estabrook 1907, Jackson and Tate 1974, Weisheit 1989) including Cliff Swallows (Brewster 1906, Burleigh 1930, Grinnell 1937, Stoner 1939, Bent 1942, Buss 1942, Samuel 1969, Brown and Brown 1995). Previous studies did not quantify the impact of House Sparrows on Cliff Swallow nesting effort and reproduction, or compare colonies with and without House Sparrows. Our objective was to quantify nesting effort and reproduction of Cliff Swallows in the Arkansas River valley of western ' Department ot Biology, University of Arkansas-Fort Smith, Fort Smith, AR 72913, USA. “Department of Biological Sciences, University ot Arkansas, Fayetteville, AR 72701, USA. ^CoiTesponding author; e-mail: djames@uark.edu Arkansas. We compared nesting of Cliff Swal- lows in several portions of a colony, each with different levels of House Sparrow activity, and also between two separate colonies, one with House Sparrows and one without. The three colonies were in urban or semi urban areas in or around Fort Smith, Arkansas. METHODS We observed an active Cliff Swallow colony in 2007 under the state highway 59 bridge spanning the Arkansas River near Barling, Arkansas in Crawford County. The bridge height made examination of nest contents impossible. Four clusters each spanning 6 m were chosen for study from over 2,100 nests: clusters Ah and Bh with high House Span'ow activity had 125 and 46 old swallow nests, clusters Cl and Dl with low House Sparrow activity had 158 and 77 old swallow nests, respectively. Subscript “L” denotes low sparrow activity while subscript “H” denotes high span-ow activity. These clusters were chosen .so that Ah and Cl as well as Bh and Dl had similar density of nest aggregations, and were on opposite sides of the bridge adjacent to one another. Observations began prior to swallow nesting. Nesting in all clusters progressed simi- larly due to breeding synchrony of Cliff Swallows (Brown and Brown 1996). The colony was visited twice each week. At least one cluster was observed for at least 1 hr during each visit. Extra observations were made when nestlings began to peer from nests to ensure documentation of all nests producing at least one nestling. Observations were recorded on photo- 135 136 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. I. March 2010 graphs with each nest numbered. Nest observa- tions included identity of bird species entering nest, any nest building activity, nest defense behaviors, provisioning trips, and direct nestling observations. These data allowed quantification of nesting success of House Sparrows and Cliff Swallows in each cluster measured by proportion of available nests that yielded at least one nestling. Distance to the nearest successful Cliff Swallow nest was measured for all House Sparrow, Clift Swallow, and European Starling (Sturnus vulgaris) nests. Distances between nests were measured from the center of each nest opening to the nearest 0.5 mm on photographs and scaled by measuring a known distance on each photograph. Two colonies near Fort Smith, Sebastian Coun- ty, Arkansas were studied in 2008, one with no House Sparrow activity (colony El; 252 old swallow nests prior to nesting) and one with high House Sparrow activity (colony Fh; 196 old swallow nests prior to nesting). Both colonies were accessible by ladder and were on the underside of highway overpasses over drainage canals. The level of House Sparrow activity at each colony was as.sessed prior to swallow nesting by observing their pre-roost activity, recording House Sparrows entering nests, and counting nests that were stuffed with House Sparrow nesting material. Nest contents were observed several times throughout the nesting cycle (May-Jun) in colonies El and Fh using an extension ladder, dental mirror, and flashlight following Brown and Brown (1996). Nest contents were observed during the appropriate period of nesting at respective colonies to ensure documentation of eggs and nestlings. One hundred and five (41.7%) of 252 nests in colony El, and 1 1 8 (60.2%) of 196 nests in Colony Fh were investigated. A ladder was randomly positioned throughout the colony and all nests within reach of each ladder position were sampled. Nests were numbered with chalk on the concrete bridge substrate and monitored during visits. We recorded presence of nest material, eggshells, living and dead nestlings, and number of eggs for each nest. This informa- tion was u.sed to quantify clutch size, brood size, and the proportion of Cliff Swallow eggs yielding successful nestlings. Active House Sparrow nests were recorded, as were clutch and brood sizes, whenever possible. Nest observations ceased before nestlings were 10 days of age to prevent premature fledging. Chi-square contingency table analyses were peifomied to examine the relationship between Cliff Swallow nesting success and House Sparrow activity. Mann-Whitney f/-tests were used to examine the difference between nesting performance (clutch size and brood size) in colonies El and Fh- We used SPSS (2007) in all .statistical analy.ses. RESULTS House Sparrows roosting in clusters A through D gathered every evening on shrubby vegetation 3 m from the bridge for over an hour before entering their roost nests. The highest number recorded in these pre-roost congregations was 58 on 9 March 2007. Swallow nests used for roosting or nesting by sparrows were concentrated on the side of the bridge that faced nearby shrubby vegetation, which included clusters Ah and Bh. Clusters Ah and Bh had high House Sparrow activity and were grouped together for analyses. We recorded Hou.se Sparrows entering or defend- ing 38 nests and successfully nesting in seven nests in clusters Ah and Bn. Sparrows nesting outside our sample still defended nests within sampled clusters. These clusters also supported two European Starling ne.sts. Cliff Swallows nested successfully (produced at least 1 nestling) in 52 of 171 available nest cavities (30.4%). Clusters Ci^ and Dl were grouped together for analy.ses because of low House Sparrow activity. These clusters faced away from nearby shrubs. We recorded House SpaiTows entering or defending eight ne.sts and successfully nesting in only one nest. These clusters supported a single starling nest. Cliff Swallows nested successfully in 144 of 235 available nest cavities (61.3%). Cliff Swallow nesting success was significantly higher in clusters with low sparrow activity when compared to those with high .sparrow activity (x" = 37.767, df ^ \,P < 0.001, Table 1). Mean (± SD) distance to the nearest successful Cliff Swallow nest in clusters A- D (Table 2) was 33.2 ± 28.6 cm for ne.sts ofHou.se TABLE 1, ClitI Swallow (CS) ne.sling .success (producing > 1 neslling) among high or low House Sparrow activity. Data from 2007 arc presented first and 2008 data are in parentheses. House Sparrow activity High Low Nests containing > 1 CS nestling 52 (.21) 144 (41) Empty ne.sts 119(112) 91 (56) Lcasiire et at. • HOUSE SPARROWS AND CLIFF SWALLOW REPROi:)UCTION 137 Sparrows, 16.6 ± 9.2 cm for starlings, and 1 1 .0 ± 4.7 cm for Cliff Swallows. Colonies Ei, and Fh were observed in 2008. Colony El was devoid of House Sparrows except for one observation of a brief sparrow visit after most swallow eggs had hatched. An Eastern Bluebird (Sialia sialis) began nesting in an old Cliff Swallow nest but abandoned the nest soon after Cliff Swallows airived on 1 April 2008. This colony had no shrubby vegetation nearby. Fifty (47.6%) of 105 sampled nests were used by Cliff Swallows for nesting. The mean (± SD) clutch size was 2.64 ± 0.12 eggs (mode = 2; 75"’ percentile = 3, Fig. 1) and mean brood size was 2.31 ± 0.12 nestlings (mode = 2; 75"’ percentile = 3, Fig. 1). At least 74% of eggs yielded nestlings in colony El. Colony Fh had a high level of sparrow activity. This colony was near shrubs that House Sparrows used as pre-roosts throughout nesting. Thirty-one nests were stuffed with House SpaiTOW nesting material, but only 12 active House Sparrow nests were recorded. The 12 active nests produced at least 37 House Sparrow eggs. Only 10 of 118 sampled nests were active Cliff Swallow nests (8.5%). Cliff Swallows nesting in this colony (producing > 1 nestling) had lower success than in colony El iy} — 50.319, df = \, P < 0.001, Table 1). The mean (± SD) clutch size among Cliff Swallows in Fh was 2.60 ± 0.27 eggs (mode = 3; 75'" percentile = 3, Fig. 1) but the mean brood size was only 0.83 ± 0.48 nestlings (mode = 0; 75'" percentile = 1, Fig. 1). Only 20% of Cliff Swallow eggs yielded nestlings and several Cliff Swallow clutches were abandoned. Clutch size of Cliff Swallows did not differ between colonies E[ and Fh (C = 248, P = 0.966) but brood size was significantly lower in colony Fh than in colony El (C = 39, P = 0.005). DISCUSSION Cliff Swallow nesting success was low in both years when House Sparrows were nesting within the swallow colony. There was a 3 1 % reduction in TABLE 2. Distance to the nearest Clid' Swallow (CS) nest for House Sparrows, European Starlings, and ClitI Swallows nesting in Clusters A-D in 2007. Mean ± SD distance (cm) to nearest CS nest n House Sparrow 33.2 ± 28.6 8 European Starling 16.6 ± 9.2 3 Cliff Swallow 1 1.0 ± 4.7 195 Cliff Swallow clutch size FIG. 1. Cliff Swallow clutch and brood sizes from two colonies in 2008. Colony F had high sparrow activity (black) while colony E (white) had low sparrow activity. Nests with brood size of zero are those in which eggs were documented but no nestlings were produced. Cliff Swallow nests producing at least one nestling in 2007 in sections of a colony with House Sparrows, and a 54% reduction in proportion of Cliff Swallow eggs producing nestlings in 2008 in the colony with House Sparrows compared to sections without House Sparrows. House Sparrow activity was greater in nests facing and in close proximity to shrubs used as pre-roosts in all colonies. The nesting season of House SpaiTows in the southern United States begins in March (Lowther and Cink 1992), 1 month before that of Cliff Swallows (Brown and Brown 1995). House Sparrows in the present study roosted in Cliff Swallow colonies throughout winter. Our earliest observation of Cliff Swallows was 1 April 2008. House Sparrows had established themselves in old Cliff Swallow nests at this point and some were incubating eggs, similar to reports by Buss (1942) and Samuel (1969). House SpaiTOws defended a broad zone around their nests often preventing Cliff Swallows from nesting nearby (Table 2), 138 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 1, March 2010 which could explain the significant impact to Cliff Swallow nesting associated with even a modest number of House Sparrows. Samuel (1969) reported that 33.6% of Cliff Swallow nests were lost to House Sparrows with most of these nests occupied before swallows began nesting, but some were taken despite Cliff Swallow eggs or nestlings being present. House Sparrows have been observed pecking eggs and nestlings of Barn Swallows (Hinmdo rustica) (Weisheit 1989), but the enclosed nests of Cliff Swallows made direct observation of this behavior difficult. Our data suggests this occurred in Cliff Swallow colonies as there was signifi- cantly reduced Cliff Swallow production despite similar clutch sizes in a colony with House Sparrows compared to a colony without House Sparrows. We observed a House Sparrow entering an active Cliff Swallow nest despite aggressive protest from the incubating swallow and we observed dead Cliff Swallow nestlings and un- hatched eggs on the ground beneath colonies with House Sparrows; these observations are insuffi- cient to implicate House Sparrows directly. Buss ( 1942) reported successful House Sparrow management that included shooting House Spar- rows and removing Cliff Swallow nests in winter to prevent sparrow roosting. We suggest that removal of nearby shrubbery could discourage House Sparrow roosting and nesting in Cliff Swallow colonies as we observed more Hou.se Sparrow activity at swallow nests facing and in close proximity to shrubby vegetation used for sparrow pre-roost congregations. Our finding that House Sparrow activity was negatively related to the number of active Cliff Swallow nests, nesting success, brood size, and the proportion of eggs yielding nestlings indicates that Hou.se SpaiTows can significantly reduce Cliff Swallow reproductive success in western Arkansas. ACKNOWLEDGMENTS This project was funded by a grant from the Student Undergraduate Research Fellowship (SURF) of the Arkan- sas Department of Higher Education. Peter Lowther provided insightful review and suggested the nearest neighbor analysis. Field work could not have been completed without the assistance of Colby Marshall and Mike Fielder. Charles R. Brown was a valuable source of advice during development of the project. LITERATURE CITED Barrows, W, B. 1889. The English Sparrow in North America, especially in its relations to agriculture. U.S. Department of Agriculture, Division of Economic Ornithology and Mammals Bulletin 1:1^05. Bent, A. C. 1942. Northern Cliff Swallows. U.S. National Museum Bulletin 179:468. Brewster, W. 1906. The birds of the Cambridge region of Massachusetts. Memoirs of the Nuttall Ornithological Club 4:1. Brown, C. R. and M. B. Brown. 1995. Cliff Swallow {Hinmdo pyrrhonota). The birds of North America. Number 149. Brown, C. R. and M. B. Brown. 1996. Coloniality in the Cliff Swallow: the effect of group size on social behavior. University of Chicago Press, Chicago, Illinois, USA. Burleigh, T. D. 1930. Notes on the birdlife of northwest- ern Washington. Auk 47:48. Buss, I. O. 1942. A managed Cliff Swallow colony in southern Wisconsin. Wilson Bulletin 54:153-161. Estabrook, a. H. 1907. The present status of the English Sparrow problem in the United States. Auk 24:129- 134. GRtNNELL, J. 1937. The swallows of the Life Sciences Building. Condor 39:206-210. Jackson, J. A. and J. Tate Jr. 1974. An analysis of nest box use by Purple Martins, House Sparrows and starlings in eastern North America. Wilson Bulletin 86:435^49. Lowther, P. E. and C. L. Cink. 1992. House Sparrow {Passer domesticus). The birds of North America. Number 12. Samuel, D. E. 1969. House Sparrow occupancy of Cliff Swallow nests. Wilson Bulletin 81:103-104. SPSS Institute Inc. 2007. SPSS 16.0.1 for Windows. SPSS Institute Inc., Chicago, Illinois, USA. Stoner, D. 1939. Parasitism of the English Sparrow on the Northern Cliff Swallow. Wilson Bulletin 51:221-222. Weisheit, A. S. 1989. Interference by House Sparrows in nesting activities of Barn Swallows. Journal of Eield Ornithology 60:323-328. The li ilson Joiiniiil oj Oniiihology 122( 1 ): 1 39-145, 2010 HOME-RANGE SIZE AND SITE TENACITY OE OVERWINTERING EE CONTE’S SPARROWS IN A EIRE MANAGED PRAIRIE HEATHER Q. BALDWIN,' CLINTON W. JESKE,- MELISSA A. POWELL,^ PAUL C. CHADWICK,' AND WYLIE C. BARROW JR.' ABSTRACT. — We evaluated home-range size and site tenacity of Le Conte’s Sparrows {Ammodramus lecontii) during winter 2002-2003 at Brazoria National Wildlife Refuge, Texas. Twenty-six wintering Le Conte’s Sparrows were radiomarked in 1- and 2-year post-burn units, and monitored for ~10 days. Additionally, 1-ha plots on each 1-, 2- and 3- year (n = 15) post-burn units were flush-netted once monthly. Telemetry results indicated Le Conte’s Sparrows were sedentary during winter with a 50% probability mean home-range of 2.41 ha (72% < 1 ha) and a 95% probability mean home range of 10.31 ha (44% < 1 ha and 55% < 1.5 ha). Home-range size did not differ between post-bum year 1 and 2 (P = 0.227). Le Conte’s Sparrows appeared to exhibit a behavioral response to flush-netting {P < 0.001) with estimated capture probability of 0.462 and recapture probability of 0.056. Our findings suggest Le Conte’s Sparrows remain fairly sedentary throughout the winter. Received 28 March 2008. Accepted 7 August 2009. Site tenacity has been reported for many overwintering grassland birds (Plentovich et al. 1998, Gordon 2000); familiarity with an area may increase an individual’s ability to acquire resourc- es and avoid predators. Optimal management of habitats in wintering areas for high-priority species that exhibit strong site tenacity may increase overwinter survival, and enhance breed- ing (Pulliam and Enders 1971, Fretwell 1972, Wiens 1974, Raitt and Pimm 1976, Grzybowski 1982). Le Conte’s Sparrows {Ammodramus lecontii) have been listed as a “high priority species in need of conservation attention” by Partners in Flight (Carter et al. 2000, Shackelford et al. 2001). They are fairly uncommon over much of their breeding range with the highest densities in the western portions (Lowther 1996). Le Conte’s Sparrows are poorly covered by the North American Breeding Bird Survey, in part due to their secretive behavior (Igl and Johnson 1999). Similarly, during the non-breeding season, the species may be underestimated or undetected on many Christmas Bird Counts. Concern for this species is compounded by severe habitat loss in ' Louisiana Slate University, 227 Renewable Natural Resources, Baton Rouge, LA 70803, USA. ^U.S. Geological Survey, National Wetlands Research Center, 700 Caiundome Boulevard, Lafayette, LA 70506, USA. ^National Park Service. 2282 South West Resource Boulevard, Moab, UT 84532, USA. ''Current address: lAP World Services at USGS. National Wetlands Research Center, 700 Cajundome Boulevard. Lafayette, LA 70506, USA. ■’Corresponding author; e-mail; heather_baldwin@usgs.gov wintering areas. Wintering areas, primarily coast- al prairie, still used by this species and other grassland-dependent birds are degraded largely due to fire suppression, agricultural practices and, more recently, invasive plant species (Grace 1998, Igl and Ballard 1999, Allain and Grace 2001). Knowledge of Le Conte’s Sparrow ecology and space use during winter can facilitate efforts to provide optimal habitat during the non-breeding season. Little is known about the general ecology of Le Conte’s SpaiTOw (Igl and Johnson 1999), espe- cially during winter. They reluctantly flush from observers and spend much of their time on or near the ground in dense vegetation. Most literature related to Le Conte’s Sparrow refers to their ecology and behavior during the breeding season (Madden et al. 1999, Johnson and Igl 2001, Winter et al. 2005). Few studies have examined Le Conte’s Sparrow winter ecology. Igl and Ballard (1999) evaluated habitat associations of migrating and overwintering grassland birds in southern Texas. Their study showed significant differences in densities of Le Conte’s Sparrows among habitats during both winter and spring but not in the fall. Baldwin et al. (2007) examined habitat relations in coastal prairie and found Le Conte’s Sparrows are most common in areas burned ^2 years and characterized by medium herbaceous density and sparse shrubs. Reynolds and Krausman (1998) examined relative abun- dance between unburned and recently burned (I0 days), and periodically move to other areas with suitable habitat. Le Conte’s Sparrow may exhibit site fidelity in wintering areas. Preliminary banding of Le Conte’s Sparrows was conducted in January 2002 on one burn unit. A mist net was randomly placed in the same unit in January 2003, and one of the 84 Le Conte’s Sparrows banded the previous year was recaptured less than 90 m from its original capture site. ACKNOWLEDGMENTS Funding wa.s provided by the U. S. Geologieal Survey (USGS), in a cooperative effort with Regions 2 and 4 of the U.S. Fish and Wildlife Service. We thank the field cre\v: M. A. Powell, S. W. Stuart. J. A. Umniel, and M. L. Keprta. We e.specially thank employees and volunteers associated with the Texas Mid-Coast National Wildlife Refuge Complex, Gulf Coast Bird Observatory, USGS National Wetlands Research Center, and Texas Parks and Wildlife Department who devoted their time to this project. We thank J. B. Grace, L. K. Allain, W. G. Vermilion, T. C. Michot. F. C. Rohwer, P. E. Lowther, L. D. Igl, and C. E. Braun for editorial comments and advice. Any use of trade, product, or firm names is for descriptive purposes only and does not imply endorsement by the U.S. Government. Bakhvin et al. • HOME-RANGE SIZE AND FIDELITY OF EE CONTE'S SPARROW 145 LITERATURE CITED Allain, L. and J. B. Grace. 2001. Changes in density and height of the shrub Baccharis halimifoUa following burning in coastal tallgrass prairie. Proceedings of the North American Prairie Conference 17:66-72. Baldwin, H. Q., J. B. Grace, W. C. Barrow Jr., and F. C. Rohwer. 2007. Habitat relationships of birds over wintering in a managed coastal prairie. Wilson Journal of Ornithology 119:189-197. Beeman, j. W., N. Bower, S. Juhnke, L. Dingmon, M. VAN den Tillaart, AND T. Thomas. 2007. Effects of antenna length and material on output power and detection of miniature radio transmitters. Hydrobiolo- gia 582:221-229. Carter, M. F., W. C. Hunter, D. N. Pashley, and K. V. Rosenberg. 2000. Setting conservation priorities for landbirds in the United States: the Partners In Flight approach. Auk 117:541-548. Cooper, S. 1984. Habitat and size of the Le Conte’s Sparrow’s territory. Loon 56:162-165. Environmental Systems Research Institute (ESRI). 2002. ARCVIEW 'Version 3.3. Environmental Systems Research Institute Inc., Redlands, California, USA. Fretwell, S. 1972. Populations in a seasonal environment. Princeton University Press, Princeton, New Jersey, USA. Gordon, C. E. 2000. Movement patterns of wintering grassland sparrows in Arizona. Auk 117:748-759. Grace, J. B. 1998. Can prescribed fire save the endangered coastal prairie ecosystem from Chinese tallow inva- sion? Endangered Species Update 15:70-76. Grzybowski, J. A. 1982. Population structure in grassland bird communities during winter. Condor 84:137-152. Grzybowski, J. A. 1983. Patterns of space use in grassland bird communities during winter. Wilson Bulletin 95:591-602. Helms, C. W. and W. H. Drury Jr. 1960. Winter and migratory weight and fat field studies on some North American buntings. Bird Banding 31:1-40 IGL, L. D. AND B. M. Ballard. 1999. Habitat associations of migrating and overwintering grassland birds in southern Texas. Condor 101: 771-782. iGL. L. D. AND D. H. Johnson. 1999. Le Conte’s Sparrows breeding in Conservation Reserve Program fields: precipitation and patterns of population change. Studies in Avian Biology 19:178-186. Johnson, D. H. and L. D. Igl. 2001. Area requirements of grassland birds: a regional perspective. Auk 1 18:24—34. Krementz, D. G. and G. W. Pendleton. 1990. Fat scoring: sources of variability. Condor 92:500—507. Lorenz, S. 2007. Site fidelity of wintering Le Conte’s Sparrows in northeast Texas. North American Bird Bander 32: 153-157. Lowther, P. E. 1996. Le Conte’s Sparrow (Ammodranms leconleii). The birds of North America. Number 224. Madden, E. M., A. J. Hansen, and R. K. Murphy. 1999. Influence of prescribed fire history on habitat and abundance of passerine birds in northern mixed— grass prairie. Canadian Field-Naturalist 113:627-640. National Audubon Society. 2002. The Chrisimas Bird Count historical results [Online]. http://www.audubon. org/bird/cbc (accessed 1 August 2008) Otis, D. L., K. P. Burnham, G. C. White, and D. R. Anderson. 1978. Statistical inference from capture data on closed animal populations. Wildlife Monographs 62. Plentovich, S. M., N. R. Holler, and G. E. Hill. 1998. Site fidelity of wintering Henslow’s Sparrows. Journal of Field Ornithology 69:486^90. Pulliam, H. R. and F. Enders. 1971. The feeding ecology of five sympatric finch species. Ecology 52:557-566. Pulliam, H. R. and G. S. Mills. 1977. The use of space by wintering sparrows. Ecology 58:1393-1399. Rappole, j. H. and A. R. Tipton. 1991. New harness design for attachment of radio transmitters to small passerines. Journal of Field Ornithology 62:335-337. Raitt, R. j. and S. L. Pimm. 1976. Dynamics of bird communities in the Chihuahuan Desert, New Mexico. Condor 78:427-442. Rexstad, E. and K. Burnham. 1991. User’s guide for interactive program CAPTURE. Colorado Cooperative Fish and Wildlife Research Unit, Fort Collins, USA. Reynolds, M. C. and P. R. Krausman. 1998. Effects of winter burning on birds in mesquite grassland. Wildlife Society Bulletin 26:867-876. SAS Institute Inc. 1999. SAS/STAT. "Version 8. SAS Institute Inc., Cary, North Carolina, USA. Shackelford, C. E., N. R. Carrie, C. M. Riley, and D. K. Carrie. 2001. Project prairie birds: a citizen science project for wintering grassland birds. Second Edition. PWD BK W7000-485 (1/01). Texas Parks and Wildlife, Austin. USA. Sinclair, A. R. E. 1984. The function of distance movements in vertebrates. Pages 240-258 in The ecology ot animal movement (I. R. Swingland and P. J. Greenwood. Editors). Clarendon Press, Oxford, United Kingdom. Tucker Jr., J. W. and W. D. Robinson. 2003. Influence of season and frequency of fire on Henslow’s Sparrows (Ammodrainus henslowii) wintering on Gull Coast pitcher plant bogs. Auk 120:96-106. White, G. C., K. P. Burnham, D. L. Otis, and D. R. Anderson. 1978. User’s manual for Program CAP- TURE. Utah State University Press. Logan. USA. Wiens, J. A. 1974. Climatic instability and the "ecological saturation’’ of grassland bird communities in North American grasslands. Condor 76:385-400. Winter. M., J. A. Shaffer. D. H. Johnson. T. M. Donovan, W. D. Svedarsky, P. W. Jones, and B. R. EULISS. 2005. Habitat and nesting of Le Conte’s SpaiTows in the northern tallgrass prairie. Journal of Field Ornithology 76:61-71. Worton, B. j. 1989. Kernel methods for estimating the utilization distribution in home-range studies. Ecology 70:164-168. ZiPPiN, C. 1956. An evaluation of the removal method of estimating animal populations. Biometrics 12:163- 189. ZiPPiN, C. 1958. The removal method of population estimation. Journal of Wildlife Management 22:82-90. The Wilson Journal of Ornithology 122( 1):146-152, 2010 PATERNAL SONG COMPLEXITY PREDICTS OFFSPRING SEX RATIOS CLOSE TO FLEDGING, BUT NOT HATCHING, IN SONG SPARROWS DOMINIQUE A. POTVIN' --^ AND ELIZABETH A. MacDOUGALL-SHACKLETON’ ABSTRACT. — Sex allocation theory predicts that population sex ratios should be generally stable and close to unity, but individuals may benefit by adjusting the sex ratio of their offspring. For example, females paired with attractive males may benefit by overproducing sons relative to daughters, as sons inherit their fathers' attractive ornaments (“sexy son” hypothesis). Similarly, if compatible gene effects on fitness are more pronounced in males than females, genetically dissimilar mated pairs may enhance fitness by overproducing sons (“outbred son” hypothesis). We tested these hypotheses in Song Sparrows (Melospiza melodia) by examining offspring sex ratios of 64 complete families shortly after hatching (“early-stage”) and again shortly before fledging ("late-stage”) in relation to paternal song complexity and the genetic similarity of social mates. Neither early nor late-stage offspring sex ratio was related to parental genetic similarity. Nests of males with larger song repertoires contained more male-biased broods by the late-stage nestling period, but not in the early- stage nestling period. These findings suggest that attractive males may be better able to successfully raise male-biased broods, but not that females adaptively adjust primary sex ratios in response to their social mate’s attractiveness. Received 20 April 2009. Accepted 24 July 2009. Sex allocation theory predicts that population sex ratios tend to be stable and close to 1 : 1 (Fisher 1930), but variation among individuals in circum- stances and quality may favor parents facultative- ly adjusting the sex ratios of their offspring (Trivers and Willard 1973). The “sexy son” hypothesis (Weatherhead and Robertson 1979), originally postulated to explain female tolerance for polygyny, also predicts that attractive males and their mates should produce predominantly sons, as sons may inherit elaborate ornaments. Consistent with this hypothesis, paternal attrac- tiveness has been linked to offspring sex ratios in some birds (Ellegren et al. 1996) although not in others (Westerdahl et al. 1997; see analyses by Ewen et al. 2004, and also Cassey et al. 2006). These mixed results may indicate mechanistic constraints preventing some species from manip- ulating sex ratios, and/or a relatively weak selective advantage to doing so (Komdeur and Pen 2002, Fawcett et al. 2007). The question of how paternal attractiveness should affect offspring sex ratios in the case of sexually .selected traits that are learned and/or highly environment- dependent, such as bird song, remains almost entirely unanswered (but see Leitner et al. 2006). Good genes models of sexual selection have recently expanded to consider nonadditive, or ' Department of Biology, University of Western Ontario, London. ON N6A 5B7. Canada. 'Current address: Department of Zoology, University of Melbourne. Melbourne, VIC 3010, Australia. 'Corresponding author; e-mail: d.potvin@pgrad.unimelb.edu.au compatible-gene effects on fitness (Neff and Pitcher 2005) with particular emphasis on hetero- zygote advantage. Compatible-gene effects are often observed in the expression of secondary sexual traits (Marshall et al. 2003, Reid et al. 2005), raising the possibility that sons may profit more than daughters from parental genetic compatibility. These effects are also commonly implicated in immunocompetence (e.g., Reid et al. 2005); this trait is expressed by male and females, but may affect male fitness disproportionately due to the immunosuppressive effects of high levels of circulating testosterone (Folstad and Karter 1992). If compatible-gene effects on fitness are stronger in males, optimal offspring sex ratios may depend in part upon parental genetic similarity. We refer to this possibility as the “outbred son” hypoth- esis. Our objective was to examine if offspring sex ratios in Song Sparrows {Melospiza melodia) are associated with paternal attractiveness based on a learned trait (song repertoire size) and/or with parental genetic similarity. The mechanisms by which birds might manipulate offspring sex ratios comprise two major classes: before hatching,- for example through .selectively developing Z- or W- bearing ova; and after hatching, for example through preferentially feeding either males or females (Pike and Petrie 2003). We investigated sex ratios at two different points in the nesting cycle, once shortly after hatching and again shortly before fledging. Female Song Sparrows prefer males with more complex song repertoires (Searcy 1984, Reid et al. 2004), and song repertoire size has been correlated with immuno- 146 Polvili and MacDoiigall-Shackleton • SONG COMPLEXITY AND SEX RATIOS 147 competence, teiritory defense, and other measures of male fitness (Reid et al. 2005, Pfaff et al. 2007). Thus, the "sexy son" hypothesis predicts that males with larger repertoires (more attractive males) should produce male-biased broods. The "outbred son” hypothesis predicts that genetical- ly dissimilar mated pairs should also produce male-biased broods, because individual genetic diversity is related both to immunocompetence and to repertoire size in this species (Reid et al. 2005). METHODS Study Subjects and General Field Methods. — We examined offspring sex ratios in 64 families of Song Sparrows breeding near Newboro, Ontario, Canada (44° 38' N, 76° 19' W). We captured adults in seed-baited treadle traps in April and May 2006, 2007, and 2008, outfitted each with a unique combination of colored leg bands, and collected a small (<25 pL) blood sample for genetic analysis. We identified mated pairs and located nests by behavioral observa- tions. Each adult was included only once in the analysis; for birds that bred in multiple years we randomly selected which year’s nest to include in the analysis. We included only first nesting attempts to reduce potential seasonal effects on sex ratio. We collected a small (15 mL) blood sample from each nestling via femoral venipuncture 2 days after eggs hatched for subsequent genetic identification as male or female. We used felt-tip markers to mark nestlings’ toes to track subse- quent survival, weighed individuals to the nearest 0.1 g, and collected any unhatched eggs for genetic identification of inviable embryos as male or female. We returned to the nest 4 days later (6 days after hatching), noted which individuals were still present, and weighed each individual to the nearest 0.5 g. Day 6 was used to avoid artificially inducing early fledging, which nor- mally occurs 8-10 days after hatching (pers. obs.). Song Analysis. — ^We recorded songs of adult males early in the mating season onto Marantz Professional PMD 671 solid state recorders using Telinga Twin Science Pro parabolic microphones. We considered a repertoire to have been sampled in full after recording 300 consecutive or 450 non- consecutive songs following the guidelines estab- lished for this population by Pfaff et al. (2007). We generated spectrograms in SYRINX (John Burt, www.syrinxpc.com), and classilied song types by visual inspection and sorting to identify each male’s repertoire size. This was done blind with respect to offspring sex ratios. Genetic Analysis. — ^We genotyped all adults at seven microsatellite loci: Escm 1 (Hanotte et al. 1994), Mme 2 and 7 (Jeffrey et al. 2001), Pdom 5 (Griffith et al. 1999), and Sosp 3, 13, and 14 (L. E. Keller, pers. comm.). We tested for deviations from Hardy- Weinberg expectations and from linkage equilibrium using GENEPOP Version 3.3 (Raymond and Rousset 1995), and found no evidence for either. We used microsatellite profiles to calculate Wang’s (2002) coefficient of genetic similarity for each mated pair using MARK (Kermit Ritland, http://genetics.forestry. ubc.ca/ritland/programs.html). We identified 208 nestlings and unhatched embryos from 55 nests (33 of which had survived to 6 days after hatching) as males or females over 3 years of study. We used primers P2 and P8 (Griffiths et al. 1998) to amplify portions of the CHD-W and CHD-Z gene, located on avian sex chromosomes. Each gel included control amplifi- cations from a known male and a known female. Statistical Analysis. — ^We investigated the rela- tionships between offspring sex ratios and pater- nal repertoire size and/or parental genetic simi- larity using generalized linear models (GEM; PROG GENMOD) (SAS 2004) with a logit-link function and binomial error distribution. Embryos collected from unhatched eggs were included in the early-stage (2 days after hatching) analyses, while late-stage analyses included only nestlings that survived until 6 days after hatching. We constructed two statistical models for both early and late-stage sex ratios. Initial models included paternal repertoire size, parental genetic similarity (using Wang’s [2002] coefficient), year, and relative hatching date (number of days before or after the mean hatching date for all first-brood nests that year) as predictor variables. Einal models excluded any clearly non-informative predictor variables {y- < 1). Number of sons within a nest for all models was the response variable and number of offspring (sons plus daughters) was the binomial denominator. All predictor variables were normally distributed (Kolmogorov-Smirnov test) and all statistical tests were two-tailed. We tested whether the use of color bands might confound sex ratios in our study population, because colored leg bands may affect attractive- 148 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 1, March 2010 ness and offspring sex ratios in some species (Burley 1986; but see Rutstein et al. 2005). We investigated the relationship between offspring sex ratio and the number of bands of that color on each adult male for each of the eight band colors using Pearson’s correlations with the appropriate Bonferroni corrections. We did not observe any relationship between color bands and either hatching or fledging sex ratios (all P > 0.05), suggesting our use of colored leg bands was unlikely to be an important confounding issue in this study. RESULTS Fifty-four (53.8) percent (112 of 208) of nestlings analyzed 2 days after hatching were male. Sixty (55%) of the 109 nestlings that survived until at least 6 days after hatching were male. Our initial GLM identified a significant effect of year on early-stage offspring sex ratios (/■ = 6.01, P = 0.014) with sex ratios somewhat more male-biased in 2007 than other years. None of the other predictor variables included in this model explained a significant amount of variation in early-stage offspring sex ratios (repertoire size: /• = 0.34, P = 0.56; parental genetic similarity: /- — 0.25, P = 0.56; relative hatching date: = 0. 13, P = 0.72). Our initial GLM at 6 days after hatching identified a near-significant association between paternal repertoire size and offspring sex ratios (x~ — 3.28, P = 0.070), and no significant effects of other predictor variables on late-stage sex ratios (parental genetic similarity: x' = 0.77, P = 0.38; relative hatching date: x' ~ P = 0.38; year: x" = P = 0.27). Our final GLMs included only repertoire size and year as predictor variables, and indicated a significant positive relationship between repertoire size and late-stage iX^ = 4.66, P - 0.030) but not early-stage (x’ = 0.52, P = 0.47) offspring sex ratios (Fig. 1). Differences in .sex ratios at day 6 could be explained by one of two possibilities: either there was increased mortality of female nestlings in attractive males’ nests (creating a male bias), or there was increased mortality of male nestlings in less attractive males’ nests (creating a female bias). Twelve of the thirty-three nests (36%) that had at least one nestling survive to day 6 (surviving nests) had partial brood loss. We used a /-test to investigate differences between the social fathers’ repertoire sizes for nests in which male and female nestlings were dying, and found that daughters were consistently dying in nests of males with larger repertoire sizes (/ = —3.78, df = 64, P < 0.001). The total number of offspring surviving to day 6 was not significantly coirelated with social father’s repertoire size {r = 0.003, df = 34, P = 0.763). However, when analyzing predictors of number of nestlings lost in surviving nests, social father’s repertoire size had a significant effect (x" = 20, P = 0.002). We also compared the average growth rates of brothers to the average growth rates of sisters in the same nest to help explain differences in mortality. Brothers consistently grew faster than sisters (/ = —2.88, df — 2Q, P — 0.009). The data indicate, although there was not sufficient statis- tical power to compare nestling mortality rates between years, that female mortality in surviving nests was higher than male mortality in surviving nests for 2 of the 3 years (Tables 1-2). DISCUSSION Offspring sex ratios near the end of the nestling period were significantly related to paternal song repertoire size (Fig. IB), consistent with the predictions of the “.sexy son’’ hypothesis. How- ever, this relationship did not appear to be mediated through manipulation of primary sex ratios, as repertoire size did not predict sex ratios earlier in the nestling period (Fig. lA). It is unlikely the lack of a relationship between song repertoire size and early-stage sex ratios was due to low statistical power, as we were able to detect an effect of repertoire on late-stage sex ratios despite a lower sample size of surviving nests (38 vs. 66). Adjusting primary sex ratios is presum- ably less costly in terms of fitness than altering sex ratios after hatching, but the apparent absence of primary sex-ratio manipulation may reflect an inability to do so. The positive relationship between paternal song repertoire size and brood sex ratio late in the nestling period may or may not represent adaptive post-hatching male/female allocation. It remains to be learned whether adult Song Sparrows preferentially feed male or female offspring, much less how such biases might relate to paternal attractiveness. A growing body of evidence points to the importance of early developmental condi- tion in shaping song-learning ability (e.g., Bu- chanan et al. 2003), suggesting that complex singers and their mates may benefit from oveiproducing, or preferentially feeding sons, if sons stand to gain more than daughters from direct benefits via increased paternal care. Secondary Potvin and MacDougall-Sluickleton • SONG COMPLEXITY AND SEX RATIOS 149 O Male repertoire size FIG. 1. Relationship between paternal repertoire size and olTspring sex ratio, measured as proportion ot sons, (A) 2 days and (B) 6 days after hatch over 3 years (filled circles TABLE 1. Population-wide nestling survival of Song Sparrows near Newboro, Ontario, Canada. 2006 2007 2008 Males Females Males Female.s Males Females Day 2 59 35 26 26 27 35 Day 6 35 22 15 18 10 8 Survival rate 59% 63% 58% 69% 37% 23% 2006, open circles = 2007, and inverted triangles = 2008). sexual traits in many species honestly advertise paternal ability or effort (Buchanan and Catchpole 2000, Voltura et al. 2002, Siefferman and Hill 2003). Thus, more attractive males would provision at a higher rate, resulting in a more noticeable difference between male and female sibling growth rates in nests of more attractive males. We observed a difference in growth rates between brother and sister nestlings, consistent with biased feeding. 150 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 1, March 2010 TABLE 2. Population-wide nestling survival of Song Sparrows near Newboro, Ontario, Canada, including only nests in which at least one nestling survived to day 6. 2006 2007 2008 Males Females Males Females Males Females Day 2 36 26 15 18 15 11 Day 6 35 22 15 18 10 8 Survival rate 97% 85% 100% 100% 80% 72% Comparing mortality rates between female and male nestlings indicates the observed relationship between paternal repertoire size and sex ratios near fledging may be due to disproportionate mortality of daughters in the nests of attractive males. The reason for these female-biased deaths is not completely clear. One possibility is that male nestlings that are fed more often, perhaps by an attractive male, may be better able to OLitcompete their sisters. There may be a positive feedback effect whereby an increase in the size of a male nestling may enhance his begging ability, resulting in further biased feeding from parents. Consistent with this possibility, female Song Sparrow nestlings in nests parasitized by Brown- headed Cowbirds (Molothrus ater) are more susceptible to being outcompeted by the faster- growing cowbird nestling (Zanette et al. 2005). This could be interpreted as an adaptive sex ratio bias in that sons may be preferentially fed; however, the deaths of female nestlings could be an “unintended” result of this feedback effect. We cannot rule out effects of disease and partial predation, but starvation resulting from differen- tial provisioning by parents is likely the main cause of partial brood reduction, as theorized by Ricklefs (1969), and supported by experiments by Arce.se and Smith (1988) and observations by Hochachka and Smith (1991). Nestling growth rate in this population is correlated with parental visits, providing further support to the theory that nestling success is heavily influenced by parental care (Potvin and MacDougall-Shackleton 2009). We observed no relationship between parental genetic similarity and offspring sex ratio, contrary to the predictions of the “outbred son” hypoth- esis. This may reOect low statistical power, an inability to detect genetic compatibility, and/or a relatively weak selective advantage to adjusting sex ratios in response to this variable. Some traits exhibiting heterozygote advantage are expressed mainly by males, such as song-learning ability (Marshall et al. 2003, Reid et al. 2005), but genetic diversity can also enhance female fitness, for example through fecundity (Ortego et al. 2007) and hatching success (Mair et al. 2006). There may be little, if any, selective advantage to manipulating primary or secondary brood sex ratios in response to expected offspring heterozy- gosity. Adaptive adjustments to primary sex ratios appear taxonomically widespread among birds, but our study adds to a growing body of evidence that the pattern is by no means universal (Komdeur and Pen 2002, Ewen et al. 2004, Cassey et al. 2006). Moreover, apparently adap- tive patterns of variation in secondary sex ratios may represent a byproduct of environmental conditions and the quality of care provided by attractive versus less attractive males, rather than a facultative response by females to the perceived attractiveness of their mates. ACKNOWLEDGMENTS We thank Scott MacDougall-Shackleton for helpful comments; Steve Russell, Megan Barclay, Casey Hoggard, Janet Lapierre, Laura Dindia, Yanina Sarquis-Adamson, Vishalla Singh, and Erica Tennenhouse for field and laboratory assistance; Jeremy Pfaff and Kathryn Stewart for recordings; Lucas Keller for primer sequences; the Queen’s University Biological Station for logistic support; and NSERC Canada and the Canada Foundation for Innovation for funding. LITERATURE CITED Arcese, P. and J. N. M. Smith. 1988. Effects of population density and supplemental food on reproduction in Song Sparrows. Journal of Animal Ecology 57;119- 136. Buchanan, K. L. and C. L. Catchpole. 2000. Song as an indicator of male parental effort in the Sedge Warbler. Proceedings of the Royal Society of London, Series B 267;32l-326. Buchanan, K. L., K. A. Spencer, R. A. Goldsmith, and C. K. Catchpole. 2003. Song as an honest signal of past developmental stress in the European Starling. Proceedings of the Royal Society of London, Series B 270:1 149-1 156. Burley, N. 1986. Sex-ratio manipulation in color-banded populations of Zebra Finches. Evolution 40:1191- 1206. Cassey, P., J. G. Ewen, and A. P. M0ller. 2006. Revised evidence for facultative sex ratio adjustment in birds: a correction. Proceedings of the Royal Society of London, Series B 273:3129-3130 Ellegren, H., L. Gustafs.son, and B. C. Sheldon. 1996. Sex ratio adjustment in relation to paternal attractive- ness in a wild bird population. Proceedings of the Polvin aiul A/cicDoiigcill-Slicicklelon • SONG COMPLEXITY AND SEX RATIOS 151 National Academy of Sciences of the USA 93; 1 1723- 1 1728. Ewen. J. G.. P. Cassey, and A. P. Mdller. 2004. Facultative primary sex ratio variation: a lack of evidence in birds? Proceedings of the Royal Society of London, Series B 27 1 ; 1277-1282 F.awcett, T. W., B. Kuuper, I. Pen, and F. J. Weissing. 2007. Should attractive males have more sons? Behavioral Ecology 18:71-80. Fisher, R. A. 1930. The genetical theory ot natural selection. Oxford University Press, Oxtoid, United Kingdom. Folstad, I. and a. j. Karter. 1992. Parasites, bright males, and the immunocompetence handicap. Amer- ican Naturalist 139:603-622. Griffith, S. C., I. R. K. Stewart, D. A. Dawson, I. P. F. Owens, and T. Burke. 1999. Contrasting levels of extra-pair paternity in mainland and island populations of the House Sparrow (Passer domesticus): is there an “island effect”? Biological Journal of the Linnean Society 68:303-316. Griffiths, R., M. C. Double, K. Orr, and R. J. G. Dawson. 1998. A DNA test to sex most birds. Molecular Ecology 7:1071-1075. H.^NOTTE, O., C. Zanon, a. Pugh, C. Grieg, A. Dixon, AND T. Burke. 1994. Isolation and characterization of microsatellite loci in a passerine bird; the Reed Bunting Emheriza schoenidus. Molecular Ecology 3:529-530. Hochachka, W. and j. N. M. Smith. 1991. Determinants and consequences of nestling condition in Song Sparrows. Journal of Animal Ecology 60:995-1008. Jeffery, K. J., L. F. Keller, P. Arcese, and M. W. Bruford. 2001. The development of microsatellite loci in the Song Sparrow, Melospiza melodia (Aves) and genotyping errors associated with good quality DNA. Molecular Ecology Notes 1:11-13. KOMDEUR, J. AND 1. PEN. 2002. Adaptive sex allocation in birds: the complexities ot linking theory and practice. Philosophical Transactions of the Royal Society of London, Series B 357:373-380. Leitner, S., R. C. Mar.shall, B. Leisler, and C. K. Catchpole. 2006. Male song quality, egg size and offspring sex in captive Canaries (Serinus camiria). Ethology 112:554-563. Marr, A. B., P. Arcese, W. M. Hochachka, J. M. Reid, AND L. F. Keller. 2006. Interactive effects of environmental stress and inbreeding on reproductive traits in a wild bird population. Journal ot Animal Ecology 75:1406-1415. Marshall, R. C., K. L. Buchanan, and C. K. Catchpole. 2003. Sexual selection and individual genetic diversity in a songbird. Proceedings of the Royal Society ot London, Series B 270:S248-S250. Neff, B. D. and T. E. Pitcher. 2005. Genetic quality and sexual selection; an integrated framework for good genes and compatible genes. Molecular Ecology 14:19-38. Ortego, j., G. Calabuig, P. J. Cordero, and J. M. Aparicio. 2007. Egg production and individual genetic diversity in Lesser Kestrels. Molecular Ecol- ogy 16:2383-2392. Pfaff, j., L. Zanette, S. A. MacDougall-Shackleton, AND E. A. MacDougall-Shackleton. 2007. Song repertoire size varies with HVC volume and is indicative of male quality in Song Sparrows (Melos- piza melodia). Proceedings of the Royal Society of London, Series B 274:2035-2040. Pike, T. W. and M. Petrie. 2003. Potential mechanisms of avian sex manipulation. Biological Reviews 78:553- 574. PoTViN, D. A. AND E. A. MacDougall-Shackleton. 2009. Parental investment amplifies effects of genetic complementarity on growth rates in Song Sparrows (Melospiza melodia). Animal Behaviour 78:943-948. Raymond, M. and F. Rousset. 1995. GENEPOP (Version 1.2): a population genetics software for exact test and ecumenicism. Journal of Heredity 86:248-249. Reid, J. M., P. Arcese, A. Cassidy, A. Marr, J. Smith, AND L. Keller. 2005. Hamilton and Zuk meet heterozygosity? Song repertoire size indicates inbreed- ing and immunity in Song Sparrows (Melospiza melodia). Proceedings of the Royal Society of London, Series B 272:481-487. REtD, J. M., P. Arcese, A. CASStDY, S. M. Hiebert, J. N. M. Smith, P. K. Stoddard, A. B. Marr. and L. F. Keller. 2004. Song repertoire size predicts initial mating success in male Song Sparrows, Melospiza melodia. Animal Behavior 68. 1055— 1063. Ricklefs, R. E. 1969. An analysis of nestling mortality in birds. Smithsonian Institution Press, Washington. D.C.. USA. Rutstein. A. N., H. E. Gorman, K. E. Arnold. L. Gilbert, K. J. Orr, A. Adam, R. Nager. and J. Graves. 2005. Sex allocation in response to paternal attractiveness in the Zebra Finch. Behavioral Ecology 16:763-769. SAS. 2004. Version 9.1 u.ser's guide. SAS Institute Inc.. Cary, North Carolina. USA. Searcy, W. A. 1984. Song repertoire size and female preferences in Song Sparrows. Behavioral Ecology and Sociobiology 14:281-228 Siefferman, L. and G. E. Hill. 2003. Structural and melanin coloration indicate parental effort and repro- ductive success in male Eastern Bluebirds. Behavioral Ecology 14:855-861 Trivers, R. L. and D. E. Willard. 1973. Natural selection of parental ability to vary the sex ratio of offspring. Science 1 79:90-92. VOLTURA. K. M., P. L. SCHWAGMEYER, AND D. W. MOCK. 2002. Parental feeding rates in the Hou.se Sparrow, Passer domesticus'. are larger-badged males better fathers? Ethology 108:1011-1022. Wang, J. 2002. An estimator for pairwise relatedness using molecular markers. Genetics 160:1203-1215. Weatherhead, P. and R. Robertson. 1979. Offspring quality and the polygyny threshold: “the .sexy son hypothesis.” American Naturalist 113:201-208. 152 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. I, March 2010 Westerdahl, H., S. Bensch, B. Hansson. D. Hassel- QUIST, AND T. VON SCHANTZ. 1997. Sex ratio variation among broods of Great Reed Warblers Acrocephaliis aritndinaceus. Molecular Ecology 6:543-548. Zanette, L., E. Clinchy, and J Cowbirds skew 86:815-820. . MacDougall-Shackleton, M. N. M. Smith. 2005. Brown-headed host offspring sex ratios. Ecology The ITilson Joiinuil of Oniilliology 1 22( 1 ); 1 53-1 59, 2010 INFLUENCE OF AGE AND SEASON ON MORPHOMETRIC MEASUREMENTS OF THE BISCUTATE SWIFT {STREPTOPROCNE BISCUTATA) MAURO PICHORIM' ABSTRACT —Little emphasis has been placed on the influence of season and age on morphological measurements of swifts which makes comparative mensural studies difficult in the Apodidae. This study provides information about the imporlance of moiphological variation in measurements to the biology of the Biscutate Swift (Streptoprocne h,sculata)A studied individuals captured at four colonies in southern Brazil. Mass varied with age, time of day, and season. Birds captured at dawn weighed 1 15.5 ± 9.9 g {n = 1,320). Mass was greater in fall and spring and lower m summer^and winter Subadults weighed less than adults during summer, fall, and winter. The wing measured 206.3 ±5.2 mrn in „), an the tail measured 69.9 ± 4.6 mm (n = 802); both showed seasonal variation in length. The tarsus measured 25.9 _ 0.6 mni („ = 695) and the exposed culmen measured 9.94 ± 0.36 mm (n = 658). The Biscutate Swift has significant seasonal variation in several morphological measurements. This variation should be considered m studies comparing populations of this species as well as those of other apodids. Received 6 April 2009. Accepted 4 September 2009. The morphology of the Apodidae is conserva- tive and they are remarkably uniform in many respects, such that differences among them are subtle and may even be undetectable in prepared skins. Some species of the genera Cypseloides, Aeroclramus, Chaetiira, and Apiis are not easily distinguished even when observed in the hand (Wetmore 1957; Sims 1961; Medway 1966; Chantler and Driessens 1995; Chantler 1995, 1999) Swifts do not vary greatly in size interspecifically, and only minor differences are apparent in the shape of the beak, tail, wing, and tarsus. Morphological measurements, in some cases, contribute to diagnoses of species and subspecies (e.g.. Sick 1991, Mann and Stiles 1992, Browning 1993, Parkes 1993, Chantler and Driessens 1995). Clinal variations, at times, lead to high numbers of subspecies, although not all variations justify taxonomic separation (Chantler and Driessens 1995). The body mass of many swifts also depends upon age and season, and variations are much greater than those in passerine birds of the same size (Lack and Lack 1951, Naik and Naik 1966). Subadults in some species weigh less than adults during part of the first year of life, and this feature is used to estimate the age of individuals (Coffey 1958, Collins 1968, Telia et al. 1995). However, several studies that made morphological compar- isons among species did not consider possible variation in measurements caused by environmen- ' Departamento de Botanica, Ecologia e Zoologia. Univer.sidade Federal do Rio Grande do Norte, Campus Universitario, Lagoa Nova, Natal, RN, Brazil, 59072-970, e-mail: mauropichorim@yahoo.com.br tal and age-related factors. Sick (1991) based the diagnosis of Streptoprocne hisciitata seridoensis on differences in body mass and wing and tail length, but did not consider seasonal variation. Some descriptions of new Cypseloides species also ignored variation in measurements through- out the year (Eisenmann and Lehmann 1962, Collins 1972, NavaiTO et al. 1992). Taxonomists have used the data available and, considering the scarcity of museum material and the difficulty of studying swifts, those surveys have helped us to understand species relationships. However, little emphasis has been placed on the seasonal and age dependency of some measurements. This tenden- cy can make comparative population studies difficult. 1 investigated age-related and seasonal variation in moiphological measurements at four colonies of the Biscutate Swift (Streptoprocne hisciitata) in southern Brazil, and 1 discuss the results with regard to the use of those parameters in taxonomic studies. METHODS Four Biscutate Swift colonies were studied in the eastern part of the State of Parana, southern Brazil (Anhangava colony, municipality of Quatro Barras - 25° 22' S, 48° 58' W, 1,250 m asl; Seira do Capivari colony, municipality of Campina Grande do Sul - 25° 1 1' S, 48° 51' W, 1,050 m asl; and Arenitos and Furna 1 colonies, munici- pality of Ponta Grossa - 25° 15' S, 50° 00' W, -1,000 m asl; Fig. 1). Only measurements from captured individuals were used. Mist nets and other nets were used to capture swifts near the most frequently used exit of each cave or roost 153 154 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. I. March 2010 FIG. 1. Biscutate Swift colonies studied in southern Brazil. (A) Anhangava, municipality of Quatro Barras. (B) Capivari, municipality of Campina Grande do Sul, (C) Arenitos, and (D) Furna 1, both in the Vila Velha State Park, municipality of Ponta Grossa. site (Pichorim and Monteiro-Filho 2010). Metallic numbered bands, supplied by CEMAVE/ICMBio, (4.5 mm inner diam) were used to mark captured swifts. The following measurements of each captured individual were taken: body mass, wing length (tlattened and straightened), tail length (measuring with a metal ruler inserted between the central pair of rectrices), tarsus length, expo.sed culmen length (measured from the beak tip in a straight line to the edge of the feathers in the central part of the forehead), and length of beak from nostril (hereafter nostril beak tip; measured from the anterior end of the maxilla in a straight line to the anterior edge of the nostril). Pesola spring scales (300 g capacity; estimated to the nearest Ig), calipers (15 cm), and rulers (30 cm) were used to measure swifts. All birds were relea.sed near the place of capture. Analysis of variance (ANOVA) and Mests (alpha = 0.05) were u.sed to test the hypothesis ot the existence of only one group of means for measurements taken during different seasons, age classes, and colonies following Zar (1984). Normality of distribution was tested for each comparison, and variances were considered anal- ogous if one did not exceed double of the other. Measurement sets with samples of similar sizes were compared by ANOVA and tested for normal distribution and homogeneity of the variances using Bartlett, Brown-Eorsythe, and Levene tests (JMP 4.0.4, SAS 2001). Data are presented as means ± SD. Individuals younger than 1 year of age were considered subadults. Age could be ascertained for marked nestlings captured after leaving the nest, and for individuals captured without molt of remiges and rectrices between January and March, the months in which all adults molt (pers. obs.). RESULTS Adult swifts weighed 115.7 ± 9.9 g (n = 1,340); considering only birds captured at dawn, the mass was 115.5 ± 9.9 g (u = 1,320). The heaviest individual (153 g) was captured in Arenitos on the afternoon of 2 September 2001. This individual had been captured the previous year at dawn on 14 October 2000 and weighed 125 g. The lightest individuals (87 g, n = 2) were captured in Arenitos on 2 December 2000. The plumage of these birds did not allow estimation of their age. However, subadults weighed less than adults captured in the same period. Adults in summer (Jan-Mar) weighed 110.9 ± 6.8 g (n = 292), while subadults weighed 102.5 ± 9.3 g (// = 8) (t — 3.43, df = 298, P < 0.001). Adults in fall (Apr-Jun) weighed 1 18.9 ± 6.7 g (/? = 128), and subadults weighed 1 1 1.0 ± 10.4 g (n = 5) (A = 2.54, df = 131, P = 0.012). Adults in winter (Jul- Sep) weighed 116.1 ± 10.5 g (// = 456), and subadults weighed 97.7 ± 1 1 .5 g (/; = 6) (r = 4.28, df = 460, P < 0.001). Adults in spring (Oct-Dec) weighed 116.9 ± 10.4 g (/; = 444), and subadults weighed 108.8 ± 4.2 g (n = 6) (/.= 1.88, df = 448, P — 0.061). The lightest confirmed subadult (86.0 g) was captured at dawn on 9 July 2002 in Arenitos. No differences in mass were apparent between males and females, as measurement of adults by season had a normal distribution. The mass of birds captured at dawn varied during the year; birds weighed more in fall and spring and less in summer and winter. Analyzing random samples of the same size from each season confirmed these Piclwrim . MORPHOMETRICS OF THE BISCUTATE SWIFT 155 differences (ANOVA F,,39(, = 19.8, P < 0.001). Adults were heavier in September and lighter in February (124.0 ± 8.7 g, n = 105 and 108.0 ± 5.2 g, n = 35; Table 1). Seasonal variation in mass was noted in some individuals captured repeatedly. An individual weighing 139 g was captured in the Anhangava colony on 1 1 October 1994, and the same individual weighed 104 g on 24 February 1995. This variation coiTesponds to a loss in mass of 25% in a 4-month period. Two other individuals, both captured in Arenitos, had mass changes of 30 g in a few months. The first individual weighed 138 g on 13 October 2000 and 108 g on 12 December (a 22% decrease). The second individual weighed 101 g on 12 December 2000 and 131 g on 3 September 2001 (a 30% increase). These values were not related to time of capture, as all individuals were weighed at dawn as they were leaving the night roost. Variations in mass were also observed in individuals captured at day’s end. The maximum variation was 15 g, for an individual from the Capivari colony, captured on 24 November 2000 weighing 131 g and recaptured the following day weighing 1 16 g (a 12% decrease). Six birds captured at day’s end had mass decrease of 9- 13 g (a 7-10% decrease). Sixteen of 25 birds captured at day’s end had mass decreases, six had mass increases (not more than 5 g), and three had no variation. An individual captured in Capivari at 0615 hrs on 27 October 2001 weighed 1 1 1 g, and 109 g at 1000 hrs the same day, a 2 g decrease during a period of 3 hrs and 45 min. A nocturnal decrease of 17 g (13%) was observed for an individual captured on 24 November 2000 at Capivari colony; it weighed 131 g in late afternoon and 114 g the following morning. The mean annual mass of adults captured at dawn varied among the colonies (Capivari = 120.2 ± 9.0 g, n = 169; Arenitos = 115.9 ± 10.1 g, n = 610; Anhangava = 113.7 ± 8.7 g, n = 360; and Furna 1 = 1 13.2 ± 7.8 g, /? = 181). However, these data are not totally comparable because of seasonal variations of individuals captured in each colony. Considering only March, April, August, October, and November, when I obtained random samples of the same size from each colony, the Furna 1 average was smaller than TABLE 1. Measurements of body mass, at four colonies in southern Brazil (Mean ± wing SD, length, and tail length in adult Streptoprocne hisciitata captured at dawn n; minimum - median - maximum). .Month Mass (g) Flattened wing (mm) Tail (mm) Jan 108.1 ± 5.4, 134 203.6 ± 5.2, 134 66.6 ± 3.8, 99 93.0-108.0-120.0 186.0-204.0-215.0 51.0-67.0-78.0 Feb 108.0 ± 5.2, 35 203.1 ± 6.9, 34 72.7 ± 2.5, 35-'' 98.0-109.0-117.0 189.0-204.0-218.0 66.0-73.0-77.0 Mar 1 15.0 ± 6.6, 119 207.3 ± 7.3, 20 75.0 ± 2.5. 16 98.0-1 15.0-132.0 184.0-209.0-217.0 70,0-75.0-79.0 Apr 1 18.5 ± 5.4, 78 200.7 ± 5.6, 59” 75.7 ± 2.82, 52 106.0-1 19.0-134.0 185.0-201.0-214.0 68.0-76.0-83.0 May 122.8 ± 2.6, 5 207.8 ± 3.0, 5'-' 74.8 ± 1 .9. 5 120.0-122.0-126.0 205.0-207.0-21 1.0 72.0-75.0-77.0 Jun 122.3 ± 8.3, 96 207.8 ±4.1. 96 76.0 ± 3.5, 96 101.0-122.5-139.0 199.0-208,0-219.0 62.0-73.5-80.0 Jul 109.9 ±8.1, 233 208.7 ± 3.9, 1 15 72.0 ± 3.0, 115 90.0-1 10.0-135.0 198.0-209.0-218.0 64.0-72.0-80.0 Aug 1 19,4 ± 8.1. 62 208.2 ± 4.4, 62 70.3 ± 3.9. 62 99.0-120.0-134.0 198.0-209.0-217.0 60.0-71.0-77.0 Sep 124.0 ± 8.7, 105 207.2 ± 4.0. 71 69.6 ± 3.6. 72 107.0-125.0-141.0 197.0-208.0-218.0 60.0-70.0-78.0 Oct 120.7 ± 10.2, 217 206.7 ± 4.9. 132 67.6 ± 4.0, 135 96.0-121.0-144.5 194.0-207.0-219.0 54.0-67.5-78.0 Nov 120.2 ± 7.4. 75 206.9 ± 4.8, 50 68.3 ± 2.6. 52 100.0-122.0-134.0 192.0-207.0-214.0 62.0-68.0-73.0 Dec 109.6 ± 8.0, 161 204.2 ± 5.2, 98 67.7 ± 4.7. 98 87.0-1 10.0-130.0 192.0-205.0-214.0 57.0-68.0-76.0 “ Individual.s molting rectrices (27 in the 1st and 2nd pair. 6 in the l.st pair, and 2 in the 2nd pair). ” Individuals molting primaries (33 in the 9th and lOlh pair, 14 in the 9th pair, and 12 in the lOth pair). Individuals molting the lOlh pair ol primaries. 156 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. I, March 2010 those of the other colonies (ANOVA ^3,3 14 = 5.6, P < 0.001; meancap.vari = 120.8 ± 8.3 g, SE = 0.92, /; = 80; meanArenitos = 1 19.9 ± 7.7 g, SE = 0.86, 11 = 80; meanAnhangava = 1 19.5 ± 8.6 g, SE = 0.96, n = 80; meanpuma = 1 15.9 ± 7.7 g, SE = 0.87, n = 78; variances were homogeneous according to Bartlett, Brown-Eorsythe, and Le- vene tests). Of individuals captured that were not molting primary remiges 9 and 10, wing length was 206.3 ± 5.2 mm (/? = 812). The longest wings measured 219 mm (bird captured in Anhangava colony on 1 1 October 1994 and another captured in Arenitos colony on 17 June 2001). The shortest wing measured 184 mm for a bird captured in Capivari on 2 March 2002 (this wing was severely worn). Generally, individuals without molt of the outer primary feathers, captured from January to March, showed wear on the tip of the primaries. As a result, the mean was smaller than during the rest of the year (203.9 ± 5.9 mm, « = 188). Wing length was 207.1 ± 4.7 mm (n = 624) between June and December. The largest monthly mean was in July (208.7 ± 3.9 mm, n = 115), and the smallest was in Eebruary (203.1 ± 6.9 mm; Table 1). Subadults did not have the same differences in means between the January-March period and the rest of the year (meanjan-Mar — 203.2 ± 4.4 mm, n = 9; meanApr-oec “ 201.6 ± 3.7 mm, n = 18) (/ = 0.99, df = 25, P = 0.33). Wings of subadults measured 202.1 ± 4.0 mm (range 196.0-209.0 mm, n = 27). This value was not different from that for adult wings at the beginning of the year (Jan-Mar) (/ = 1.48, df = 213, P = 0.14), but differed from the June- December period (t = 5.33, df = 649, P < 0.00 1 ). Mean wing length of adults in the June-December period had little variation among colonies (Capi- vari = 207.5 ± 4.8 mm, n — 82; Arenitos = 206.6 ± 4.7 mm, n = 242; Anhangava = 208.6 ± 4.1 mm, n - 167; Furna 1 = 205.7 ± 4.9 mm, n — 133). Considering similar random samples of the same period and size from each colony, wing length of birds from Furna 1 was smaller than in the other colonies (ANOVA = 1 1.44, P < 0.001; variances were homogeneous according to Bartlett, Brown-Eorsythe, and Levene tests). Tail length of adults not molting the innermost feathers (first and second pairs of rectrices) was 69.9 ± 4.6 mm (/; = 802). An individual captured at Furna 1 on 8 April 2001 had the longest tail (83.0 mm), while an individual captured in Anhangava on 19 January 1995 had the shortest tail (51.0 mm) with a worn tip and molt of rectrices 5-6 (outermost feathers). Individuals without molt of the first rectrices, captured from January to March, exhibited worn tail tips. The tail in this period was shorter than during the rest of the year (meanjan-Mar = 67.8 ± 4.7 mm, n = 1 16; meanApr-Dec = 70.3 ± 4.5 mm, n = 685) {t = 5.38, df = 799, P < 0.001). The largest monthly mean (76.0 ± 3.5 mm, n = 96) was in June and the smallest (66.6 ± 3.8 mm, >1 = 99) was in January (Table 1 ). Tail lengths of subadults varied between 55.0 and 71.0 m,m throughout the year, resulting in a mean of 65.4 ± 4.0 mm (n = 27). The birds had longer tails during January-March than during the rest of the year (meanjan-Mar = 68.3 ± 2.5 mm, n = 9; meanApr-Dec = 64.0 ± 3.8 mm, n = 18) (r = 3.13, df = 25, P = 0.005). Tails of adults in the second half of the year were longer than those of subadults (t = 5.89, df = 701, P < 0.001). Length of tails of adults in the period without molt of the rectrices (Apr-Dee) varied among colonies (Cap- ivari = 68.0 ± 4.0 mm, n = 95; Furna 1 = 68.9 ± 5.1 mm, n = 144; Anhangava = 70.5 ± 3.4 mm, /? = 183; Arenitos = 71.7 ± 4.4 mm, n = 263). Considering similar random samples of the same period and size from each colony, tail length of birds from Furna 1 was smaller than in the other colonies and larger than birds from Arenitos (ANOVA F3218 = 28.79, P < 0.001). The tarsus of adults measured 25.9 ± 0.6 mm (/? = 695). An individual captured in Anhangava had the longest tarsus (27.5 mm), and two individuals from this colony and two from Furna 1 had the shortest tarsi (24.3 mm). The tarsus of subadults was 25.8 ± 0.5 mm (24.7-26.7 mm, n = 24), which did not differ from measurements of the adults (l = 0.22, df = 717, P = 0.83). The mean length of the tarsus of adults was similar among colonies (Furna 1 = 25.8 ± 0.6 mm, n = 155; Capivari = 25.8 ± 0.6 mm, 11 = 108; Anhangava = 25.8 ± 0.7 mm, /? = 114; Arenitos = 25.9 ± 0.6 mm, n = 318) (ANOVA F3.69i.= 1.42, P = 0.24; variances were homogeneous according to Bartlett, Brown-Forsythe, and Le- vene tests). The exposed culmen of adults measured 9.94 ± 0.36 mm (/; = 658). An individual captured in Arenitos had the longest exposed culmen (11.1 mm), and at least one specimen from each colony had the shortest expo.sed culmen (9.0 mm). Individuals with head feather molt had longer exposed culmens (meanwith molt = 9.98 ± Pichorim • MORPHOMETRICS OF THE BISCUTATE SWIFT 157 mm, f\ 376, mctin^id-^oui mou 9.88 — 0.37 mm, n = 282; t = 3.40, df = 656, P < 0.001). The nostril beak, tip (mean = 6.27 ± 0.29 mm, n = 658) did not vai'y with molt (meanwini mou = 6.29 it 0.28 mni, /? 376, mean^ij).(Qi^ii niou 6.24 — 0.32 mm, n = 282) (/ = 1.95, df = 656, P = 0.052). Siibadults had exposed culmens similar to adults without molt (meaUsubaduits = ± 0.29 mm, n = 15; / = 1.90, df = 295, P = 0.058), but the nostril beak tip was shorter (meansubaduit = 6.07 ± 0.31 mm, n = 15; r = 2.09, df = 295, P = 0.037). Considering the nostril beak tip measurement, which did not vary with molt, and random samples of similar size from each colony, individuals from Furna 1 had the shortest beak (Furna 1 = 6.19 ± 0.34 mm, n — 100; Capivari = 6.28 ± 0.26 mm, n = 100; Arenitos = 6.31 ± 0.27 mm, n = 100; Anhangava = 6.31 ± 0.30 mm, /? = 97) (ANOVA ^3,393 = 3.66, P = 0.013). DISCUSSION Body mass of Biscutate Swifts depended on time of day, age, and season. There was no evidence of differences in mass between males and females, a valid comparison as seasonal samples had normal distributions. Feeding was presumably the cause of daily mass variations, as individuals captured in late afternoon were heavier due to their digestive systems being full. Mass of subadults was lower than adults during the greatest part of the first year of life. The difference in mass was observed in summer, fall, and winter. Other subadult swifts have a similar tendency (e.g.. Chimney Swift [Chaetura pela- gica]. Chapman’s Swift [C. chapmani], and Alpine Swift [Tachymarptis melha]) (Coffey 1958, Collins 1968, Telia et al. 1995). Fat storage in nestlings increases their chances of survival after leaving the nest. However, young swifts probably experience .some difficulty obtaining food during the first months of independent life. It is known that Biscutate Swifts decrea.se in mass between the nestling pha.se and the first months as subadults (Pichorim 2002). However, the new data demonstrate that mass of subadults is below that of adults for longer than previously known. It is likely that subadult survival during the first year of life is lower than for adults, as it is for some Apodidae (Chantler and Driessens 1995, Chantler 1999). Mass equalization between age groups occurs during the spring at the end of the first year of life. Reserves stored during the nestling phase appear to be indispensable for the survival of subadults. Control of brood size in the Biscutate Swift is achieved by egg ejection, probably during unfavorable conditions (Pichorim and Monteiro- Filho 2008). This behavior allows for heavier fledglings and, consequently, higher subadult survival during the first year. Seasonal variation in the mass of adults may be related to prey availability, reproductive period, and molt with mass peaking in fall and spring, as observed for Chimney Swift (Coffey 1958, Johnston 1958) and Common Swift {Apiis apus) (Lack and Lack 1951). The first period coincides with the end of molt when demand for energy is reduced. The new plumage may also act to increase mass, as it improves flight efficiency and facilitates capture of prey. It is also possible that storing of reserves before winter indicates an adaptation for low food availability. Increases in mass occur in spring due to increased resource availability and breeding. Mass starts decreasing by December, which is certainly related to the effort of caring for nestlings. The decrease in mass in summer is probably caused by the increased metabolic demand of molt. The maxi- mum gain and loss of mass for the same individual in a year (30 and 25%, respectively) were close to rates for the Common Swift (31.3 and 13.2%, respectively (Lack and Lack 1951). The 13% overnight decrease in mass detected is the highest rate reported for Apodidae. The highest value previously reported for the Common Swift was 10% (Lack and Lack 1951). Mass decrease of individuals captured in morning and at day’s end could be linked to stress of capture. However, captured individuals were set free only after other individuals left the overnight shelter and were subsequently isolated from other individuals of the colony. Thus, the mass decrease could be an indication the species u.ses a technique of hunting for food in flocks and. when an individual is apart from the Bock, it forages less efficiently. Schoener (1971) reports that grouping benefits animals when food avail- ability fluctuates over time. This technique enlarges the monitored area and increases the chances of finding prey when resources are restricted or infrequent. Individuals may experi- ence more difficulty in feeding when flying alone or in small flocks. Seasonal variation of wing and tail lengths was mainly due to natural wear. Wings were shorter at the start of molt because the feathers were worn 158 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 1, March 2010 from use. Mean lengths were greater after molt with the new plumage. Wings and tails of subadults were shorter than those of the adults, as subadults do not molt during the first year of life and have older feathers. The difference in tail lengths of subadults between summer and other periods of the year also appeared to be caused by wear. There was no difference in tarsus length of adults and subadults. The tarsus of S. hisciitata grows quickly during the nestling stage, reaching maximum development at the 19th day of life; this serves the nestlings well for hanging onto their nest (Pichorim 2002). Tarburton (1986) reported the tarsi of White-rumped Swiftlets (Aerodramus spocliopydiiis assimilis) in Fiji reached adult size on day 14, whereas adult weight was not reached until day 23 and wing- length sometime after fledging. Marin (1997) also showed the tarsus of Cypseloides uiger reached adult length in about 12 days whereas mass took about 20 days and wing reached adult length post fledging. The tarsus is a good moiphological measurement to compare populations due to its stability and quick development. Molt of head feathers influenced the exposed CLilmen, which implied a certain limitation in using it as a comparison between populations that were not sampled during the same period. No plausible cause points to age-related variations of the nostril beak tip length, but the differences could be caused by increased cranial ossification or by the development of the beak. Means of body mass and lengths of wings, tails, and nostril beak tip were significantly smaller in Furna 1. These data suggest that individuals from this colony experience special conditions, which may be related to competition for better places for resting, overnight roosting, and nesting with White-collared Swifts (Streptopvocne zonaris) (a large number of which inhabit the same shelter). This study demonstrates that several morpho- logical parameters of the Biscutate Swift have significant seasonal variations. There is evidence that similar variation occurs in other species of the Family Apodidae. Thus, studies to compare Apodidae populations must consider the depen- dence of certain morphological measurements on season and age. The existence of morphological differences over a latitudinal gradient is probable in some swifts. However, comparison of samples from different seasons could accentuate the differences between populations and lead to misinterpretation. AC KNOWLEDGMENTS The author thanks the Conselho Nacional de Desenvol- vimento CientiTico e Tecnologico (CNPq) for financial support, the Post-Graduate Program in Zoology at the Federal University of Parana for institutional support, and the Instituto Ambiental do Parana (lAP) for access to Vila Velha State Park. 1 also thank Alexandre Lorenzetto, Aline DaPMaso Ferreira, Andressa C. Bonat, Arthur Angelo Bispo de Oliveira, Bianca Luiza Reinert, Deborah Afon.so Cornelio, Cristina Imaguire, Cristiani Hagar Flores, Diter Liepsch, Douglas Kajiwara, Juliano Ribeiro, Marcos Ricardo Bornschein, Marlon Carlos Borba, Nilton Carlos Caceres, Olivia Isfer, Paula Elbl, Paulo Eduardo de Souza, Priscila Barbosa, Reginaldo Assencio Machado, and Ricardo Voiles for help in the field. I appreciate the improvements in English usage made by James Armacost through the Association of Eield Ornithologists’ program of editorial assistance. The study complies with the current laws of Brazil in which it was performed. LITERATURE CITED Browning, M. R. 1993. Species limits of the Cave Swiftlets {Collocalia) in Micronesia. Avocetta 17:101-106. Chantler, P. 1995. Identification of three Chaetura swifts; Band-rumped Swift Chaetura spinicauda, Grey- rumped Swift Chaetura ciuereiventris and Pale- rumped Swift Chaetura egregia. Cotinga 4:44-5 1 . Chantler, P. 1999. Family Apodidae (Swifts). Pages 388- 457 in Handbook of the birds of the world. Volume 5. Barn-owls to hummingbirds (J. del Hoyo, A. Elliott, and J. Sargatal, Editors). Lynx Edicions, Barcelona. Spain. Chantler, P. and G. Driessens. 1995. Swifts, a guide to the swifts and treeswifts of the world. Pica Press. East Sussex. United Kingdom. Coffey, L. C. 1958. Weights of .some Chimney Swifts at Memphis. Bird Banding 29:98-104. Collins, C. T. 1968. Notes on the biology of Chapman’s Swift Chaetura chapmani (Aves: Apodidae). Ameri- can Museum Novitates 2320:1-15. Collins, C. T. 1972. A new species of swift of the genus Cypseloides from northeastern South America. Con- tributions in Science of the Natural History Museum of Los Angeles County 229:1-9. Eisenmann. E. and F. C. Lehmann. 1962. A new species of swift of the genus Cypseloides from Colombia. American Museum Novitates 21 17:1-16. Johnston, D. W. 1958. Sex and age characters and salivary glands of the Chimney Swift. Condor 60:73-84. Lack. D. and E. Lack. 1951. The breeding biology of the swift Apus apus. Ibis 93:501-546. Marin, A. M. and F. G. .Stiles. 1992. On the biology of five species of swifts (Apodidae, Cypseloidinae) in Costa Rica. Western Foundation of Vertebrate Zool- ogy 4:287-351. Marin, M. 1997. Some aspects of the breeding biology of the Black Swift. Wilson Bulletin 109:290-306 Medway, F. L. S. 1966. Field characters as a guide to the specific relations of swiftlets. Proceedings of the Linnean Society of London 177:151-172. Pichorini • MORPHOMETRICS OF THE BISCUTATE SWIFT 159 Naik, S. and R. M. Naik. 1966. Studies on the House Swift, Apiis iijfhiis (G. E. Gray), body weight. Pavo 4:84-91. N.avarro. a. G. S., a. T. Peterson, B. P. P. E.scalante, and H. D. BenItez. 1992. Cypseloides sloreri, a new species of swift from Mexico. Wilson Bulletin 104:55- 64. Parkes, K. C. 1993. Taxonomic notes on the White- collared Swift (Streptoprocne zonaris). Avocetta 17:95-100. PiCHORiM, M. 2002. The breeding biology of the Biscutate- swift, Streptoprocne hiscutata (Apodidae) in southern Brazil. Ornitologia Neotropical 13:61-84. PiCHORiM, M. AND E. L. A. MoNTEtRO-FiLHO. 2008. Brood size and its importance for nestling growth in the Biscutate Swift (Streptoprocne hiscutata, Aves: Apod- idae). Brazilian Journal of Biology 68:851-857. PtCHORIM, M. AND E. L. A. MONTEIRO-FlLHO. 2010. Population size, survival, longevity, and movements of the Biscutate Swift in southern Brazil. Annales Zoologici Fennici 47: In press. SAS. 2001. Program JUMP. SAS In.stitute Inc, Cary, North Carolina, USA. Schoener, T. W. 1971. Theory of feeding strategies. Annual Review of Ecology and Systematics 2:369- 404. Sick, H. 1991. Distribution and subspeciation of the Biscutate Swift Streptoprocne hiscutata. Bulletin of the British Ornithologists’ Club 3:38^0. Sims, R. W. 1961. The identification of Malaysian species of swiftlets Coliocaiia. Ibis 103:205-209. Tarburton, M. K. 1986. Breeding of the White-rumped Swiftlet in Fiji. Emu 86: 214-227. Tella, J. L., C. Gortazar, R. Lopez, and U. Osacar. 1995. Age related differences in biometrics and body condition in a Spanish population of Alpine Swift (Apus meiha). Journal ftir Ornithologie 136:77-79. Wetmore, a. 1957. Species limitation in certain groups of the swift genus Chaetura. Auk 74:383-385. Zar, J. H. 1984. Biostatistical analysis. Second Edition. Prentice-Hall, Englewood Cliffs, New Jersey, USA. Short Communications The Wilson Journal of Ornithology 122(1): 160-162, 2010 Cooperative Breeding by Red-Headed Woodpeckers Megan R. Atterberry-Jones' and Brian D. Peer'-^ ABSTRACT. — Red-headed Woodpeckers (Mela- nerpes eiythrocephalus) are one of the most easily recognizable bird species, but little is known about their social system and breeding activity due to their lack of sexual dimorphism. We observed a dense population of Red-headed Woodpeckers and found evidence of cooperative breeding. The habitat at our site may promote cooperative breeding because it has a high density of utility poles that woodpeckers use for nesting, roosting, and caching food. Received 7 April 2009. Accepted 30 July 2009. Cooperative and communal breeding in which some individuals forgo breeding and assist other breeding pairs have been described for > 300 species of birds (Stacey and Koenig 1990, Cockburn 2003, Koenig and Dickinson 2004). Seven of 21 members of Melanerpes woodpeck- ers, including the Acorn Woodpecker (M. for- micivoriis) and Hispaniolan Woodpecker (M. striatus), are known to breed either cooperatively or communally (Koenig and Mumme 1987, Ligon and Burt 2004). Relatively little is known of the breeding behavior and social interactions of the Red-headed Woodpecker (M. erythrocephalus) as they are sexually monomorphic and difficult to capture (Smith et al. 2000). Anecdotal observa- tions suggest Red-headed Woodpeckers may not be the monogamous loners as portrayed. They have been observed migrating in groups of 5-6 individuals suggesting they may remain in family groups for extended periods of time (Hall 1983), and three adults have been observed attending the same nest (Short 1982). These observations raise the possibility that some populations of Red- headed Woodpeckers may be cooperative breed- ers. The objectives of our study were to: ( 1 ) observe a locally dense population of Red-headed Wood- peckers in west-central Illinois to better under- stand their social behaviors during the breeding .season, and (2) examine if the den.se population ' Department of Biological Sciences, Western Illinois University, Macomb, IL 6145.'i, USA. ^Corresponding author; e-mail: BD-Peer@wiu.cdu was a result of cooperative breeding or non- breeding floaters in the area. METHODS The study site was a 6.75-ha campground at Spring Lake Park (40° 50' N, 90° 71' W) in Macomb, Illinois, USA. The area was relatively open with the dominant vegetation being oak {Quercus spp.) and hickory [Carya spp.) trees with snags present throughout the area, in addition to 31 utility poles. Individual wood- peckers were captured in mist nets using call backs and mounted specimens, and banded with uses bands and a unique combination of color bands. Fifteen Red-headed Woodpeckers were captured and banded. Red-headed Woodpeckers were observed during the 2007 and 2008 breeding seasons (late Mar through early Aug) to identify: (1) number of territories, (2) territory boundaries and whether territories overlapped, (3) color band combinations of marked individuals in each territory, (4) nest site locations, (5) number of birds and the identity of individuals tending a nest, and (6) number of birds and the identity of individuals defending territories. Territories were identified by observing a mated pair’s movements and then mapped onto a satellite image of the campground. Each territory was observed for a 2-hr time interval between 0800 and 1200 hrs CDT once a week. Nests in territories became the main focus of observations, and the identities of individuals coming and going to the nest were recorded. If a nest was not located, we followed the individuals in that territory to locate potential future nest sites. Additional ob.servations were made if more than two individuals were found in a territory to learn if the extra birds were floaters impinging on the mated pair’s territory, or if they were helpers assisting the mated pair. Helpers were classified as individuals that helped defend the territory, but did not care for the young, or as individuals observed at the nest caring for the young and defending the territory. 160 SHORT COMMUNICATIONS 161 RESULTS Red-headed Woodpeckers occurred in the campground year around and at least some individuals maintained year-around territories. Eleven breeding territories were located during the 2007 breeding season. Ten territories had two individuals, while one had three individuals. Three floaters were observed in at least one census during the breeding season. Boundaries of territories overlapped with at least one other breeding territory. Two breeding pairs used utility poles as their nesting site and the other nine used trees. There were 10 breeding territories in the 2008 breeding season. One contained three individuals while there were two each in the other nine territories. All territories in 2008 had boundaries that overlapped with at least one other breeding territory. One floater was observed during one census in the 2008 breeding season. The same two utility poles were used as nesting sites for two breeding pairs with the remainder of the breeding pairs using trees. Fifty-eight percent of the utility poles (« = 31) were used for nesting, roosting, and/or caching food. We observed cooperative breeding in both years of study. The cooperative group in 2007 consisted of an individual with a white band and two unbanded individuals that nested in a white oak (Qiiercus alba) on the east side of the campground. All three individuals were seen simultaneously on the tree that contained the nest and responded with chatter calls when they landed near another territory member. All three also responded aggressively with nest attentiveness and alarm calls to a Red-headed Woodpecker call played in front of the tree that contained the nest. All three birds defended the territory, but only the white-banded individual and one of the unbanded individuals that had a distinctive stripe on its beak were observed tending the nest. The cooperative breeding group during the 2008 breeding season was on the west side of the campground nesting in a utility pole. The group consisted of the same white banded individual, an unbanded bird which retained juvenile coloration around the eyes, and a third unbanded bird. All three birds defended the territory from invaders and tended the nest. All three individuals were seen on the utility pole with the nest and all flew to the utility pole when one bird emerged from the nest, giving nest attentiveness and chatter calls to one another in the process. All three were observed on several occasions taking turns incubating eggs or brooding young. All three responded aggressively with alarm calls when a Red-headed Woodpecker call was played near the utility pole. DISCUSSION The Red-headed Woodpecker population we studied in west-central Illinois appeared to occasionally have cooperative breeding. One breeding territory in 2007 had three individuals, all of which defended the territory, but only two were observed tending the nest. It is possible the third bird also tended the nest unobserved. One breeding territory in 2008 contained three birds, all of which defended the territory and all three were observed tending the nest. The relationship between helpers and the breeding pair was unknown because we were unable to band all birds and did not document genetic relationships. The presence of the after hatching year bird in 2008 suggests it could have been one of “white’s” young from 2007. It is unclear whether these were two distinct cases of cooperative breeding or whether it was the same white-banded bird and its mate with a different helper in both years because all of the birds were not banded. This also may be an example of irregular cooperative breeding (Ligon and Burt 2004), although the breeding behavior of these birds requires further study. The cooperative breeding we observed in addition to Short’s (1982) observations of three adult Red-headed Woodpeckers tending a nest suggests a low level of cooperative breeding occurs within this species. What factors are driving cooperative breeding in our population? First, most birds in our population maintained year-around territories which can result in greater adult survival and increased population densities, and more breeding adults than teiritories (Arnold and Owens 1999). Second, the open oak habitat in the park appears ideal for Red-headed Wood- peckers (Smith et al. 2000). The Red-headed Woodpecker population density in the park was greater than average for Illinois. The average number of breeding individuals in similar habitats in Illinois was 0.68/ha (Graber et al. 1977), but at our site there was an average of 1.6 breeding individuals/ha. The park is typical of many midwestern parks and campgrounds in that it is relatively open and dominated by oaks and 162 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 1, March 2010 hickories (pers. obs.). Parks within 16 km of our site were similar structurally, but have typical numbers of Red-headed Woodpeckers (unpubl. data). The most obvious feature that distinguished our site was the large number of utility poles, the majority of which were used by woodpeckers. It is possible the utility poles make this habitat productive by providing additional nesting and food storage sites. ACKNOWLEDGMENTS We thank S. H. Abott and Raymond Peterson for allowing us to study the Red-headed Woodpecker population at Spring Lake Park, J. D. Jones for helping with field work, and two anonymous reviewers who provided insightful comments on the manuscript. This research was supported by grants from the Illinois Ornithology Society and the School of Graduate Studies at Western Illinois University. LITERATURE CITED Arnold, K. E. and I. P. F. Owens. 1999. Cooperative breeding in birds: the role of ecology. Behavioral Ecology 5:465-471. Cockburn, A. 2003. Cooperative breeding in oscine passerines: does sociality inhibit speciation? Proceed- ings of the Royal Society of London, Series B 270: 2207-2214. Graber, J. W., R. R. Graber, and E. L. Kirk. 1977. Illinois birds: Picidae. Biological Notes Number 102. Illinois Natural History Survey, Urbana, USA. Hall, G. A. 1983. West Virginia birds. Special Publica- tions Number 7. Carnegie Museum of Natural History, Pittsburgh, Pennsylvania, USA. Koenig, W. D. and J. L. Dickinson. 2004. Ecology and evolution of cooperative breeding in birds. Cambridge University Press, Cambridge, United Kingdom. Koenig, W. D. and R. L. Mumme. 1987. Population ecology of the cooperatively breeding Acorn Wood- pecker. Monographs in Population Biology Number 24. Princeton University Press, Princeton, New Jersey, USA. Ligon, j. D. and D. B. Burt. 2004. Evolutionary origins. Pages 5-34 in Ecology and evolution of cooperative breeding in birds (W. D. Koenig and J. L. Dickinson, Editors). Cambridge University Press, Cambridge, United Kingdom. Short, L. L. 1982. Woodpeckers of the world. Monograph Series Number 4. Delaware Museum of Natural History, Wilmington, USA. Smith, K. G., J. H. Withgoti', and P. G, Rodewald. 2000. Red-headed Woodpecker (Melanerpes erythrocepha- lus). The birds of North America. Number 518. Stacey, P. B. and W. D. Koenig. 1990. Cooperative breeding in birds: long-term studies of ecology and behavior. Cambridge University Press, Cambridge, United Kingdom. The Wilson Journal of Ornithology 122( 1 ): 162-165, 2010 Observations on the Breeding Biology of the Collared Crescentchest (Melanopareia torquata) Mieko F. Kanegae,'-^ Marina Telles,“ Severino A. Lucena,^ and Jose Carlos Motta-Junior' ABSTRACT. — The Collared Crescentchest (Melano- pareia torquata) is an endemic bird of the Cerrado (Family: Melanopareiidae), and is listed in the State ol Sao Paulo, Brazil as “endangered". We studied the breeding biology of Collared Crescentchest at two nests in the State of Sao Paulo, southeast Brazil. Males were identified genetically and equipped with radio-transmit- ' Laboratorio de Ecologia de Aves, Departamento de Ecologia, Instituto de Biociencias da Univcrsidade de Sao Paulo, 05508-900 Sao Paulo, SP, Brazil. - Pos-Graduayao cm Ecologia e Recursos Nalurais pela Univcrsidade Federal de Sao Carlos, Brazil. 'Laboratorio de Bioqui'mica de Insetos, lOC-FlOCRUZ, Brazil. ■‘Corresponding author: e-mail: mieko. kanegae@gmail. com. miekok@hotmail.com ters. The incubation period was 12-16 days and the nestling period was 12-14 days. Nestling body mass was measured every second day for the first 10 days. Males participated in incubation and helped with nesting care. Measurements of eggs and nests are compared to those from the single previously known nest. These data are the first for any member of the Family Melanopareiidae. Received 27 March 2009. Accepted 28 August 2009. Little is known about the life history and ecology of tropical species, particularly with regard to endangered and threatened species of the Melanopareiidae and Rhinocryptidae (Krabbe and Schulenberg 2003). Recently, the nest, eggs. SHORT COMMUNICATIONS 163 and young of the Collared Crescentchest (Mela- nopareia torqiiata) were first described (Gressler and Marini 2007). The Collared Crescentchest is an endemic bird of the Cerrado region (Silva and Bates 2002). Willis and Oniki (2003) indicated the population is declining in the State of Sao Paulo, Brazil. The species is classified as “Endangered” in the List of Animals Threatened with Extinction in the State of Sao Paulo (Decree 53.494, 2 Oct 2008). Conservation of biodiversity requires a refined knowledge of species and their biology, especially in the tropics, where life histories of birds have been little investigated (Stutchbury and Morton 2001). Our study was directed at adding new natural history information about the Collared Crescentchest, especially reproductive and behav- ioral aspects with a conservation focus. We asked the following questions: (1) are both parents involved with incubation and nestling care, (2) what is the nest visitation frequency of adults, and (3) how long do nestlings stay in the nest? METHODS Study Area. — ^This study was conducted in the 2,300-ha Esta9ao Ecologica de Itirapina (EEI) in the municipalities of Itirapina and Brotas (22° 15' S, 47° 49' W) in Sao Paulo State, Brazil. This area is one of the last remnants of natural grasslands and grassland savannahs in Sao Paulo State (Gianotti 1988). The Cerrado has a complex of physiognomies from open grasslands (campo limpo) to forest (cerradao), including intermediate types “campo sujo”, “campo cerrado”, and “cerrado sensu stricto" (Ribeiro and Walter 1998). Climate in this region, according to Kdppen’s (1948) classification, is defined as Cwa type, receiving on average 1,376 mm of precipitation annually. During the dry season (Apr-Sep), rainfall ranges from 32 to 88 mm per month and, during the rainy season (Oct- Mar), from 117 to 257 mm per month (DAEE Post D4-014, Itirapina, SP). Field Procedures. — We captured individual Collared Crescentchest, after hearing them, using a single mist net (12 X 2 m) and, at times using playback calls. Males and females were identified genetically by taking blood samples (0.1 ml) from the jugular vein with analysis at the Laboratorio de genetica e evolu^ao de aves in the Universidade de Sao Paulo following Griffiths et al. (1998). We attached LD-2 radio transmitters (Holohil Systems Ltd, Carp, ON, Canada) using a back- pack harness on 10 individuals. We obtained radio-telemetry data between October and No- vember 2007 during the morning (0600-1200 hrs) and evening (1600—1800 hrs) in areas of “campo sujo” and “campo cerrado”. We located nests by following these individuals and observing their behavior. Nests were monitored every 2 days and the behavior of parents was observed during each visit for 30 min at a distance of ~ 15 m from the nest. Observations were conducted 3 hrs after sunrise, 3 hrs around noon, and 3 hrs before sunset. This procedure allowed us to record frequency and duration of visits to the nest. We measured body mass using a spring scale (Pesola, 50 g, accurate to 1 g), and size of eggs and nestlings was measured with a digital calliper (Mitutoyo, 0.01 mm precision). RESULTS We found two nests of M. torquata in 2007. Both were woven of dry grass leaves and a few dry leaves of trees or shrubs. One of the nests was half covered by grass leaves of Loudetiopsis chrysothrix (Poaceae), a grass commonly ob- served in the area. Average egg mass was 1.55 g, and average egg width and length was 19.05 and 14.95 mm, respectively (Table 1). Eggs were ovoid-shaped, unspotted with a greenish-blue coloration. The incubation period was 12-16 days. The first nest, with two nestlings, was found in an area of “campo sujo” on 12 November. Both nestlings left the nest 12 days later on 24 November. Considering developmental stage of the nestlings at the moment of nest discovery, the birds probably fledged 13-14 days after the eggs hatched. The second nest, with two eggs, was found on 19 November in an area of “campo cerrado”. One egg hatched on 29 November and the offspring fledged 12 days later. However, this nestling may have fledged 1-2 days prematurely due to our visits. We could still detect the parents around the nest 2 days after the young fledged, where they gave a “pirrr” call. The other egg did not hatch and was collected (deposited in the Museu de Zoologia da Universidade de Sao Paulo). The growth rates of three nestlings were similar (Fig. 1). Both males and females shared incubation and nestling care. The male made few visits to the nest, which were brief (1-2 min) and without periodicity. We recorded two visits during the nestling period in midday and one during 164 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 1, March 2010 TABLE 1. Measurements of three nests of the Collared Crescentchest obtained by Gressler and Marini (2007) in Distrito Federal and average measurements of two nests during the present study, in Sao Paulo, Brazil. Characteristics Gressler and Marini (2007) Present study Width of entry, cm 4.2 6.0 Height of entry, cm 3.7 5.25 Nest length, cm 10 14.75 Nest depth, cm 10.5 14.75 Height from nest’s base to the ground, cm 13.5 15.5 Egg mass, g 1.55 Egg width, mm 21.2/19.5 19.05 Egg length, mm 15.7/15.4 14.95 Incubation period, days 15-18 12-16 Nestling period, days 12-14 afternoon. A male was in the nest during incubation twice in the morning. It is probable the males knew of our presence during monitor- ing, which may have affected their behavior. DISCUSSION The proportions of the two Collared Cres- centchest nests we found were similar, but larger (Table 1) than those of the nests described by Gressler and Marini (2007). They also observed that dry leaves in the nests were used only as support. Egg color was similar to that described by Gressler and Marini (2007), but differed from the descriptions of Sick (1997) for coloration of Olive-crowned Crescentchest (Melanopareia maximiliani) eggs by the lack of any speckles. Eggs in this species were whitish with large dark gray speckles near the large end (Di Giacomo 2005). Periodicity of nest visits was low, which may reflect a strategy against predators, thus reducing chances of discovery of nests and eggs or nestlings (Skutch 1976). We also noted a different “pirr” call from adults near nests and suggest that one of the functions of this call is to keep the young near the nest, a behavior that may increase their survival (Martin 1996, Stutchbury and Morton 2001). Our study complements the known life history information of the Collared Crescentchest, a species less studied, and contributes to the knowledge of the behavior and reproductive ecology of tropical birds. — F1 -m— F2 -A- F3 FIG. I . Growth rates (body ma.ss) of three (FI , F2, F3) Collared Crescentchest nestlings measured every 2 days. FI and F2 (Nest I ) were monitored between 1 2 and 24 November 2007. F3 (Nest 2) was monitored from hatching on 29 November to I I December 2007. SHORT COMMUNICATIONS 165 ACKNOWLEDGMENTS We thank CEMAVE and the Institute Florestal do Estado de Sao Paulo (IF) for authorization to capture birds; CAPES for the doctoral scholarship; Laboratorio de Genetica e Evolugao de Aves do Instituto de Biociencias da Uni- versidade de Sao Paulo for bird gender identification; Instituto de Biociencias da Universidade de Sao Paulo; Idea Wild; the E. Alexander Bergstrom Memorial Research Award; Neotropical Grassland Conservancy; W. J. and Virsinia W. Moorhouse Memorial; Pamela and Alexander F. Skutch; and Birders Exchange for the financial support and equipment donation. LITERATURE CITED Di Giacomo, A. G. 2005. Aves de la Reserva El Bagual. Pages 201^65 in Historia natural y paisaje de la Reserva El Bagual, Provincia de Formosa, Argentina. Inventario de la fauna de vertebrados y de la flora vascular de un area protegida del Chaco Hiimedo. (A. G. Di Giacomo and S. F. Krapovickas, Editors). Aves Argentinas. Temas de Naturaleza y Conservacion 4. Asociacion Ornitologica del Plata, Buenos Aires, Argentina. Gianotti, E. 1988. Composi§ao florfstica e estrutura fitossociologica da vegeta9ao de Cerrado e mata ciliar da Esta9ao Ecologica de Itirapina (SP). Disserta9ao. Campinas, Universidade Estadual de Campinas, Sao Paulo, Brazil. Gressler, D. R. and M. A. Marini. 2007. Nest, eggs and nestling of the Collared Crescentchest Melanopareia torquata in the Cerrado region, Brazil. Revista Brasileira de Ornitologia 15:574-576. Griffiths, R., M. Double, K. C. Y. Orr, and R. J. G. Dawson. 1998. A DNA test to sex most birds. Molecular Ecology 7:1071-1075. Koppen, W. 1948. Climatologia: con un estudio de los climas de la Tierra. Fondo de Cultura Economica, Mexico. Krabbe, N. and T. Schulenberg. 2003. Family Rhino- cryptidae (Tapaculos). Pages 748-787 in Handbook of the birds of the world. Volume 8, Broadbills to Tapaculos. (J. del Hoyo, A. Elliott, and D. Christie, Editors). Lynx Edicions, Barcelona, Spain. Martin, T. E. 1996. Life history evolution in tropical and south temperate birds: what do we really know? Journal of Avian Biology 27:263-271. Ribeiro, j. F. and B. M. T. Walter. 1998. Fitofisionomias do Bioma Ceirado. Cerrado: Ambiente e Flora. Pages 89-166 in Embrapa-CPAC (S. M. Sano and S. P. Almeida, Editors). Planaltina, DF, Brazil. Silva, J. M. C. and J. M. Bates. 2002. Biogeographic patterns and conservation in the South American Cerrado: a tropical savanna hotspot. BioScience 52:225-233. Skutch, A. F. 1976. Parent birds and their young. University of Texas Press, Austin, USA. Stutchbury, B. j. M. and E. S. Morton. 2001. Behavioral ecology of tropical birds. Academic Press, San Diego, California, USA. Willis, E. O. and Y. Oniki. 2003. Aves do Estado de Sao Paulo. Divisa, Rio Claro, Brazil. The Wilson Journal of Ornithology 122(1): 165-168, 2010 Effects of a Flood on Foraging Ecology and Population Dynamics of Swainson’s Warblers Nicholas M. Anich'-^ and Bryan M. Reiley'-^ ABSTRACT.— Swainson’s Warblers (Limnothlypis swainsonii) have been described as terrestrial leaf litter foraging specialists, and are thought to vacate flooded areas entirely. We report Swainson’s Warblers occupy- ing a flooded area in Arkansas and foraging on unusual substrates. We observed Swainson’s Warblers feeding in novel ways in the tree canopy and on floating debris. This is the first report of Swainson’s Warblers occupying and foraging in flooded habitat, and suggests this species may have more short-term flexibility in ' Department of Biological Sciences, Arkansas State University, P. O. Box 599, Jonesboro, AR 72467, USA. ^Current address: 2414 Fellman Circle, Ashland, WI 54806, USA. ^Corresponding author; e-mail: bryan.reiley@gmail.com foraging substrate than previously thought. Return rates to flooded territories were reduced compared to years without floods and several marked individuals switched to an unflooded area protected by man-made levees. The altered hydrology of rivers constrained by levees may reduce habitat quality for this species. Received 19 February 2009. Accepted 14 September 2009. The Swainson’s Warbler {Limnothlypis swain- soniv. Parulidae) is a medium-sized wood-warbler that breeds in the southeastern United States, winters in the Caribbean Basin, and inhabits areas with dense understory vegetation (Brown and Dickson 1994). Pairs are socially monogamous. 166 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 1. March 2010 and males exhibit breeding site fidelity and defend their territory against conspecific males (Brown and Dickson 1994). Swainson’s Warblers are uncommon and local throughout their breeding range, and are ranked among the highest-priority species for conservation (Partners in Flight 2007). Swainson’s Warblers forage for arthropods in terrestrial dead leaf litter, primarily by flipping dead leaves and poking their beaks into litter on the ground (Meanley 1970, 1971; Barrow 1990; Graves 1998). Barrow (1990) classified 82% of foraging substrates {n = 41) as dead leaves, ground litter, or fallen debris, and classified the remainder as live leaves or herbs. Most (99%) observations of foraging Swainson’s Warblers {n = 399) by Graves (1998) involved lifting dead leaves from the ground. More rarely, they have been reported feeding at shrub level and hawking insects (Meanley 1970, 1971; Barrow 1990), although Graves (1998) attributed arboreal forag- ing observations to instances of birds disturbed by humans or audio playback. Swainson’s Warblers at times sing from branches in the mid-story (Meanley 1968), but they are considered primarily ground foragers (Brown and Dickson 1994, Graves 1998). Meanley (1966) reported that none of four territorial males returned to an area they had used the previous year after a spring flood in Georgia, and characterized the effect of flooding on Swainson’s Warblers as “devastating.” Benson and Bednarz (2010) examined Swainson’s War- bler habitat in Arkansas several years after Hooding and found only nine of 42 previously occupied locations were still occupied by Swain- son’s Warblers. Graves (2001) and Klaus (2004) also reported abandonment of flooded territories. It is not unusual for Swainson’s Warblers to occur adjacent to water or to inhabit areas that can be seasonally flooded, but flooding during the breeding season is thought to preclude site occupancy (Graves 2001, Brown et al. 2009). Our objective is to report on response of Swainson’s Warblers to a catastrophic Hood event. We hypothesized that flooding would make the leaf litter and other ground debris unavailable for foraging, and therefore cause individuals to vacate an inundated area, possibly even resulting in local population extirpation. OBSERVATIONS Our observations were made in the Alligator Lake area of White River National Wildlife Refuge (WRNWR), Arkansas (34° 2' N, 91° 5' W) from 24 April to 6 May 2008. This site was being studied as part of an ongoing project: ( I ) it had a continuous Swainson’s Warbler population since at least 2000, (2) had been intensively studied since 2004, and (3) most temtorial males were color-banded. The site is several meters higher in elevation than most of the refuge, which protects it from flooding in most years allowing development of dense understory vegetation preferred by Swainson’s Warblers. Heavy rains in spring 2008 caused the White River to rise, flooding WRNWR, which is within earthen levees that contain the river. The floodwaters that would have naturally inundated the much larger flood- plain were confined within the boundaries of the levees. High water in the Mississippi River also prevented normal drainage of the White River. Additionally, an Army Corps of Engineers pumping plant was pumping water into WRNWR from agricultural areas bordering the refuge. The site became completely inundated in late March and the water started to recede in late April, disappearing (except for isolated pools) by 6 May 2008. The first Swainson’s Warblers returned to this site on 8 or 9 April in 2005 through 2007 (years with no flooding). We surveyed all previously known Swainson’s Warbler temtories in the study area beginning 24 April 2008 via canoe by listening for their songs and using audio play- backs. We encountered Swainson’s Warblers singing and foraging in trees in areas that were inundated with 0.91-2.75 m of water. We observed four Swainson’s Warblers forag- ing in atypical ways. One unbanded male sally- hovered (terminology follows Remsen and Rob- inson 1990) above a pile of floating debris and appeared to attempt to pick an insect from it. This bird then perched on a vine [Smila.x spp.) 2 cm above the water and picked an insect from the flotsam. A banded male was observed hopping along horizontal branches of a tree ~ 10 m above the water level, using typical leaf-lifting motions to search for arthropods under loose pieces of bark. This bird methodically hopped along one horizontal branch, looking for arthropods under loose bark, then gleaned insects from several other large horizontal branches, until it flew to the canopy of another tree. We observed a different banded male and one banded female on separate occasions searching for in.sects by lifting leaves from hanging dead leaf clusters lodged in tree SHORT COMMUNICATIONS 167 limbs. We observed some of these activities after we had judiciously used audio playback, and in other instances, we had not used playback. We observed 10 different males and four females in flooded habitat, including four instanc- es of pairs traveling together. Seven of the 10 males were banded and three were unbanded. One male that settled in flooded habitat had been banded at WRNWR as a nestling in 2007 (T. J. Benson and NMA, unpubl. data). We detected one male in 2006 and 2007 holding a territory outside the levee bordering the refuge. This area remained dry during the flood, and the number of males detected holding territories outside the levee increased to six in 2008, including four coloi- banded males that held territories inside the levee in previous years. DISCUSSION Some Swainson’s Warblers survived the period of flooding and eventually settled on territories when the water receded, but we found reduced return rates of color-banded individuals and reduced site occupancy in 2008. We detected 25 males in 2008, compared to higher numbers in previous years (40 in 2005; 43 in 2006; 47 in 2007; Benson 2008, Anich et al. 2009). The return rate for color-banded males from 2007 to 2008 was 0.36 (17 of 47), compared to higher proportions in previous years (0.53 from 2005 to 2006; 0.58 from 2006 to 2007; Benson 2008, Anich et al. 2009). These are the first observations of Swainson’s Warblers foraging in the tree canopy. Tree branches and floating debris were the only substrate available on which birds could feed during the prolonged flooding period. We ob- served birds using previously unreported foraging substrates, but these birds were using typical foraging maneuvers supporting Graves (1998) observation that Swainson’s Warblers use a limited repertoire of foraging maneuvers. Graves (1998) classified 99% {n = 399) of foraging maneuvers as “leaf lifting,’’ which occuired with associated gleaning, probing, and gaping motions, and two instances of flush-pursuit of moths (Lepidoptera) from the leaf litter. Strong (2000) found 80% of foraging maneuvers in wintering areas consisted of leaf lifting. Barrow (1990) reported that 63% of foraging maneuvers (/; — 41 ) were flaking (analogous to leaf lifting), 32% gleaning, and 5% (2 instances) sally-gleaning (includes sally-strike and sally-hover-, Remsen and Robinson 1990). The literature, to our knowledge, does not report Swainson’s Warblers occupying flooded landscapes or feeding above shrub-level height. Thus, we were suiprised to document their return to breeding areas during the flood and their use of novel foraging substrates. It is perhaps not surprising that Swainson’s Warblers have evolved the flexibility to survive temporary flooding, as they often inhabit bottomland forests. However, anthropogenic changes to natural flooding regimes appear to have increased the duration and depth of the high water at our site, and at other sites occupied by Swainson’s Warblers (Graves 2001, Thompson 2005). Some Swainson’s Warblers persisted at the site thi'ough the flood and settled on ten-itories when the flooding subsided. Howev- er, lower return rates of color-banded individuals and reduced site occupancy in 2008 suggest that other individuals either vacated their territories when confronted with flooding or perished. Several authors have noted abandonment of territories after flooding (Meanley 1966, Graves 2001, Klaus 2004). Previously we assumed the mechanism by which flooding prevented Swain- son’s Warbler occupancy was through direct removal of the primary foraging substrate. Our observations suggest that abandonment may result from the initial effects of the flood (i.e., elimination of ground for foraging) but also longer-term changes that reduce habitat quality for Swainson’s Warblers. Benson and Bednarz (2010) examined Swainson’s Warbler habitat at WRNWR and three other sites within 3 years of flooding and found only 21% of sites occupied after the flood. They also found that locations exhibiting more flood-related habitat changes, such as reduced leaf litter and shrub cover, were more likely to be unoccupied. We observed that Swainson’s Warblers have the flexibility to survive short-term flooding, but flooded and recently flooded habitats appear to be suboptimal for this species. ACKNOWLEDGMENTS Funding for this research was provided hy the U.S. Fish and Wildlife Service and Arkansas Slate University. T. J. Benson provided unpublished banding records. J. C. Bednarz. T. J. Benson. S. K. Robinson. P. A. Spaeth, and an anonymous reviewer provided insightful comments on this manuscript. We thank J. C. Bednarz. T. J. Benson, Carolina Roa, and Alex Zachary for assistance in the field. THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 1. March 2010 LITERATURE CITED Swainson’s Warbler along a hydrological gradient in Anich, N. M., T. J. Benson, and J. C. Bednarz. 2009. Effect of radio transmitters on return rates of Swainson’s Warblers. Journal of Field Ornithology 80:206-211. Barrow Jr., W. C. 1990. Ecology of small insectivorous birds in a bottomland hardwood forest. Dissertation. Louisiana State University, Baton Rouge, USA. Benson, T. J. 2008. Habitat use and demography of Swainson’s Warblers in eastern Arkansas. Disserta- tion. Arkansas State University, Jonesboro, USA. Benson, T. J. and J. C. Bednarz. 2010. Short-term effects of flooding on understory habitat and presence of Swainson’s Warblers. Wetlands 30: In press. Brown, J. D., T. J. Benson, and J. C. Bednarz. 2009. Vegetation characteristics of Swainson’s Warbler habitat at the White River National Wildlife Refuge, Arkansas. Wetlands 29:586-597. Brown, R. E. and J. G. Dickson. 1994. Swainson’s Warbler (Limnothlypis swainsonii). The birds of North America. Number 126. Graves, G. R. 1998. Stereotyped foraging behavior of the Swainson’s Warbler. Journal of Field Ornithology 69:121-127. Graves, G. R. 2001. Factors governing the distribution of Great Dismal Swamp. Auk 118:650-664. Klaus, N. A. 2004. Swainson’s Warblers may shift territories in response to spring flooding. Oriole 69: 19. Meanley, B. 1966. Some observations on habitats of the Swainson’s Warbler. Living Bird 5:151-165. Meanley, B. 1968. Singing behavior of the Swainson’s Warbler. Wilson Bulletin 80:72-77. Meanley, B. 1970. Method of searching for food by the Swainson’s Warbler. Wilson Bulletin 82:228. Meanley, B. 1971. Natural history of the Swainson’s Warbler. North American Fauna Number 69. Partners in Flight. 2007. Partners in Flight species assessment database, http://www.rmbo.org/pif/pifdb. html (accessed 10 November 2008). Remsen Jr., j. V. and S. K. Robinson. 1990. A classification scheme for foraging behavior of birds in terrestrial habitats. Studies in Avian Biology 13:144-160. Strong, A. M. 2000. Divergent foraging strategies of two neotropical migrant warblers: implications for winter habitat use. Auk 117:381-392. Thompson, J. L. 2005. Breeding biology of Swainson’s Warblers in a managed South Carolina bottomland forest. Dissertation. North Carolina State University, Raleigh, USA. The Wilson Journal of Ornithology 122(1): 1 68-1 73, 2010 Preiiator Vocalizations Affect Foraging Trade-offs of Northern Cardinals Mark T. Stanback'-^ and Emily M. Powell' ABSTRACT. — We investigated the roles of food quality and predator avoidance in shaping foraging behavior of Northern Cardinals (Cardinalis cardinalis). Cardinals visited feeders near cover and feeders with sunflower seed significantly more than they visited those in the open and those with commercial wild bird seed mix. Cardinals with a choice between their preferred food and preferred location visited the sunflower seed feeder in the open significantly more than the mixed seed feeder near cover, despite their prior preference for feeding near cover. We recorded visits to these feeders over a 2-day period during which we played Cooper’s Hawk (Accipiter cooperi) calls every 2 hrs. The overall visitation rate did not decrea.se significantly, nor did the visitation rate to feeders near cover increase significantly. However, the decrease in the visitation rate to feeders in the open approached significance. Overall, we observed a significant shift in the proportion of visits to feeders near cover (containing ' Department of Biology, David.son College, Davidson, NC 28035, USA. ^Corresponding author; c-mail: ma.stanback@davidson.edu less preferred seed). We conclude Northern Cardinals are sensitive to both food quality and predation risk, and adjust foraging to prevailing perceptions of risk. Received 23 March 2009. Accepted 23 July 2009. Natural selection should favor organisms that make optimal foraging decisions when presented with conflicting options (Lima and Dill 1990, Krebs and Kacelnik 1991, Cuthill and Houston 1997). The efficiency of energy/nutrient intake and avoidance of predation are particularly important factors affecting foraging animals (Dill and Fraser 1984, Lima 1985, Lima and Valone 1986, Valone and Lima 1987). Many passerines prefer to feed near cover due to higher risk of predation in the open by Accipiter hawks (Morse 1973, Barnard 1979). Feeding in open areas increases a bird’s visibility and the time it takes to reach safety (Pulliam and Mills 1977); small birds spend less time scanning for predators when SHORT COMMUNICATIONS 169 near cover (Caraco et al. 1980). Given a particular probability of the arrival of a predator, birds should be able to perceive the relative danger of visiting feeders that differ in distance to cover (Lima 1985). Granivorous birds present an excellent system for investigating trade-offs. Not only can the quality of presented food be both quantified and controlled, but researchers can also manipulate predation risk by varying the distance to the nearest cover. Predation risk may vary seasonally or even daily (Fenn and Macdonald 1995, Lucas et al. 1996, Lima and Bednikoff 1999). Moreover, prey species can use visual, auditory, and/or chemical cues to recognize the presence and threat level of predators (Lima and Dill 1990, McLean and Rhodes 1991). Consequently, researchers have used a variety of techniques to manipulate the perception of predation risk (Blumstein et al. 2008, Schmidt et al. 2008), including model predators (Lima 1985), playbacks of conspecific and heterospecific alarm calls (Peake 2005), and recordings of predator sounds (McLean and Rhodes 1991). Our objectives were to: (1) assess how Northern Cardinals (Cardinalis cardinalis), a common feeder visitor, respond to variation in food quality and distance to cover when subject to predation by both migrant and resident Cooper’s Hawks {Accipiter cooperi), and (2) examine it playing calls of a Cooper’s Hawk near our test site would alter the foraging decisions of Northern Cardinals. METHODS Field Procedures. — We installed four squirrel (Sciurus spp.)-proof, hopper-style feeders in a semi-secluded location on the campus ot David- son College in Davidson, North Carolina, USA. Birds were acclimated to these feeders for 1 month prior to the beginning of the experiment. Two feeders were positioned 1.5 m apart, 15 m from cover, in the middle of a rectangular clearing. The other two feeders were also 1.5 m apart, and placed at the edge of the clearing 1.5 m from den.se vegetation. Northern Cardinals were by tar the most common visitor to these feeders. Their large body size and aggressive behavior also minimized the potentially confounding effects of interspecific competition (Alatalo et al. 1987). The extent to which Accipiter hawks concentrate their hunting efforts near bird feeders is disputed (Roth and Lima 2007), but Cooper’s Hawks are present in Davidson year-round and have been ob.served hunting at our feeders. In our first experiment, adapted from Mumme (2003), we filled one feeder in each pair with black-oil sunflower seed and the other with ACE® brand wild bird seed mix (consisting primarily of millet, but also with milo and black-oil sunflower seed). We allowed birds several days to acclimate to the feeder contents before beginning observa- tions. We recorded the number of feeding visits by cardinals at each feeder for four 1-hr observation periods per day. Observations were made from a chair ~ 50 m from the feeders. All observations occurred between 1000 and 1800 hrs EDT. A feeding visit was one in which the bird was observed to land and consume at least one food item. We did not quantify the amount of food consumed or the length of each feeding visit. Cardinals were not individually marked and it is probable that we re-sampled many of the same individuals, violating strict independence. How- ever, several factors convinced us to pursue the study. First, the nutritional state of any bird (as well as the danger associated with visiting a particular feeder) can change over the course of a day. Second, the number of cardinals in the area was large (up to 10 were observed in the vicinity of the feeders at any given time). Third, capturing and color-banding all birds present prior to conducting this undergraduate experiment was not feasible. We recorded feeding visits on 12 and 13 September 2007. We switched the food type in each feeder pair at 1 800 hrs on 1 3 September to control for any locality or feeder preferences within the pair. We allowed birds to acclimate to the change in seed location on 14 September and resumed data collection on 15 and 16 September. We again changed the contents of the feeders at 1800 hrs on 16 September, placing the favored seed type from 12 to 16 September into the two feeders in the center of the clearing and the other seed type into feeders near cover. We allowed birds to acclimate to the new locations on 17 September and resumed data collection on 18 September, continuing through 21 September. We continued to use two feeders at both locations to minimize the chance that subordinate individuals would be forced to visit undesirable feeders. We repeated this trial on 30-31 January 2008 with preferred seed in the open and the non- preferred seed near cover. We collected data from 170 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. I, March 2010 (/) > c g ■c o Q. o Q_ FIG. 1. Northern Cardinal visitation in winter 2008 to exposed feeders with preferred food (sunflower seed) vs. visits to less preferred food near cover on days without (/? = 668 visits) and with (/? = 455 visits) Cooper’s Hawk call playback (** = P < 0.001). this trial in three 1-hr sessions per day (1 hr between 0800 and 1 100 hrs EST, another between 1 100 and 1400 hrs EST, and again between 1400 and 1700 hrs EST). We played calls of a Cooper’s Hawk (Neville 2004) on 1 and 2 Eebruary at 0800, 1000, 1200, 1400, and 1600 hrs. We recorded feeder visits by cardinals on these days as we had on 30-3 1 January. Playbacks consisted of a 5-sec call played 12 times in succession. Calls were played at a natural volume from a hidden position 80 m from the feeders. We did not have an explicit control playback because of the indirect nature of the hawk playback and the richness of the acoustic environment (corvids, dogs, car horns, trucks). Statistical Analyses. — ^We used JMP 7.0 (SAS 2007) to analyze our data. We performed X- tests to compare relative use of feeders containing different seeds, feeders in different locations, and feeders in the presence and absence of Cooper’s Hawk calls. We used Wilcoxon/Kruskal-Wallis tests to assess absolute changes in feeder visita- tion rates resulting from playback of hawk calls. Data are pre.sented as means ± SE. RESULTS Northern Cardinals visited feeders containing sunlJower seed significantly more often than those with mixed seed (Y"/ = 22.2, P < 0.001) when both foods were provided at both locations. Sixty- four percent of 532 visits were to feeders containing sunflower seeds. Cardinals visited feeders near cover significantly more often than those far from cover (X‘/ = 36.6, P < 0.001). Sixty-eight percent of 532 visits were to feeders near cover. We provided cardinals with a choice between their prefeired food type (sunflower seed) and preferred location (near cover) in the next phase of the experiment. Cardinals visited the exposed sunflower seed significantly more often than mixed seed near cover {X- / = 34.3, P < 0.001). Sixty-seven percent of 571 visits were to the exposed feeders containing sunflower seed. We repeated the trial to ensure the cardinals continued to prefer higher quality seed in a less safe location, even in winter. Cardinals visited the exposed feeders significantly more often (X- / = 12.4, P < 0.001), as we found in the fall. Sixty percent of 668 visits were to the exposed feeders containing sunflower seed (Eig. 1). We then moved on to our playback experiment. The decrease in the mean visitation rate of cardinals as a result of exposure to hawk calls (47.7 ± 8.6/ hr to 37.8 ± 4.9/hr) was not significant (Wil- coxon/Kruskal-Wallis Z = -0.54, P = 0.59). The increase in use of feeders near cover (38.6 ± 5.3/ hr to 48.3 ± 5.3/hr) was not significant (Wil- coxon/Kruskal-Wallis Z = —1.14, P = 0.25), but the decrease in the visitation rate to exposed SHORT COMMUNICATIONS 171 feeders (56.9 ± 16.3/hr to 27.3 ± 5.6/hr) approached significance (Wilcoxon/Kruskal- Wallis Z = — 1.79, P = 0.074). Overall, exposure to Cooper's Hawk calls resulted in a dramatic and significant shift in feeder use (X-/ = 60.0, P < 0.001; Fig. 1). Cardinals, instead of preferring the higher quality seed in the exposed location, as they had in the first 2 days, had a significant preference for the lower quality seed in the safer location on playback days (X"/ = 18.1, f < 0.001). Only 36% of 454 visits were to exposed feeders containing sunflower seed (Fig. 1). DISCUSSION A basic tenet of foraging theory is that organisms seek to maximize energy acquisition (Krebs and Kacelnik 1991, Cuthill and Houston 1997). One means by which this can be achieved is by foraging selectively, concentrating on higher quality foods when opportunities present them- selves. We presented cardinals with black-oil sunflower seeds (with shells) versus commercial wild bird seed mix in our experiment. We were not surprised that, when presented with both foods at both locations, cardinals visited feeders with sunflower seed significantly more often; relative to commercial wild bird seed, black-oil sunflower seed has a high energetic content and nutritional value (Willson 1971, Willson and Harmeson 1973). It is possible that cardinals varied the length of their stay (and quantity of seed ingested) at feeders, but our results suggest that visitation rates alone provide a meaningful comparison. Foraging organisms are also under pressure to minimize risk of death from predation (Lima and Dill 1990, Lima 1998). Cardinals should be sensitive to risks associated with different feeding locations and, all things being equal, should spend more time in safer foraging locations (those near cover) and less time in dangerous locations (those in the open). We were not surprised to find that, when presented with both foods at both locations, cardinals visited feeders near cover significantly more often. Optimizing foraging requires making decisions when all things are not equal (Sih 1980, Lima and Dill 1990, Cuthill and Houston 1997). In the first trial, the strength of the preference for sunflower seed was ostensibly similar to the preference for foraging near cover. However, when cardinals had to choose between their prefen'ed food and their preferred location, they showed a clear bias toward higher quality food (sunflower seed) — even though it was available only in the dangerous (exposed) location. This suggests that cardinals can recognize a risk of feeding in exposed locations, but willingly incur that risk to obtain high quality food. Our last experiment demonstrated the cardinals we observed were not applying a single assess- ment of risk to foraging activities. Most cardinals under normal conditions perceived a threat in foraging at the exposed feeders, but not suffi- ciently great to forego sunflower seeds when available only at exposed feeders. However, upon hearing hawk calls in the vicinity throughout the day, cardinals adjusted their perception of risk. Neither the decline in the visitation rate to all feeders nor the increase in the visitation rate to feeders near cover was significant, but the decline in the visitation rate to exposed feeders ap- proached significance. We observed a significant change in the relative use of exposed and protected feeders. Most cardinals were unwilling to risk predation to obtain higher quality food when exposed to the calls of a Cooper’s Hawk. Our experiment lacked an explicit control play- back, but the strength of the foraging response suggests the trends we observed were real. This plasticity is presumably adaptive, allowing indi- viduals to maximize their fitness (Lima and Bednikoff 1999). Birds recognize, assess, and respond to danger using a variety of cues including visual, auditoiT, and even olfactory (Roth et al. 2008). The cardinals we tested altered their foraging strategy dramatically to the playback of hawk calls, but these results are not always obtained. Schmidt et al. (2008) observed that eastern chipmunks {Tamias striatus) altered their foraging in re- sponse to the alarm calls of Tufted Titmice (Baeolophus hicolor), but not to the calls of an actual predator, the Broad-winged Hawk (Buteo platypterus). Schmidt et al. (2008) hypothesized the hawk call provided information upon which chipmunks could make a foraging decision, whereas the titmouse call simply alerted them to an eminent danger for which additional vigilance and information would be required. Similarly, Van der Veen (2002) found that Yellowhammers (Emberizu citronella) exposed to an accipter model resumed foraging sooner than individuals with less reliable information (those which did not see the model but did hear the alarm calls of those exposed to the model). Our experiment is the first to demonstrate that Northern Cardinals weigh the 172 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. 1, March 2010 importance of food quality differently in response to auditory cues of the presence of a particularly dangerous predator. The plasticity we observed suggests the out- comes obtained were particular to the circum- stances of our experiments. Our results may have been different had we varied certain aspects of our design (e.g., types of food used in the feeders). Moreover, had our exposed feeders been placed in a large field 50 m from the forest edge, rather than in the center of a small meadow, the perceived risk involved in visiting them may have differed. Our findings suggest that Northern Cardinals in our population tended to respond in an adaptive way to variation in food quality and differing levels of predation risk. ACKNOWLEDGMENTS We thank the students of BIO 1 12 (David Baker, Austin Bond, Imani Bryan, Kyri Bye-Nagel, Elizabeth Cannon, Fareed Cheema, Lauren Childs, Jesse Dimmock, Andrew Dunn, Lauren Felkel, Francisco Fiallo, Joel Fineman, Robert Flowers, Rieti Gengo, Sarah Griffith, Courtney Hart, Jon Huggins, Harriet King, Scott Lester, Nina Mace, Natasha Meyer, Shane Purvis, Allison Ruhe, Chris Ryan, Kyle Sanders, Meena Sangar, Jackie Tan, Ellen Thomas, Amber Townsend, Alysen Wallace, Katie Wallace) and BIO 322 (Elizabeth Arellano, Mickey Belcher, Laura Bergner, Lindsay Brownell, Jay Buckingham, Middleton Chang, Esther Cline, Kara Earle, Emma Garren, Nate Geigle, Christa Goeke, MacLean Hall, Chris Hampton, Courtney Hart, Andrew Johnson, Alex Kim, Deanna Lomax, Peggy McKay, Rachel Miranda, Shannon Pittman, Erevan Rankin, Philip Ruzycki, Sam Sheline, Ellen Thomas, Brent Winship) for their help in conducting this experiment. The manuscript was improved by the com- ments of Howell Burke. Esther Cline, Austin Mercadante, Ron Mumme, C. E. Braun, and two anonymous reviewers. LITERATURE CITED Alatalo. R. V., D. Erik.s,son, L. Gustafsson, and K. Larsson. 1987. Exploitation competition influences the u.se of foraging sites by tits: experimental evidence. Ecology 68:284-290. Barnard. C. J. 1979. Interaction between House Sparrows and Sparrowhawks. British Birds 72:569-.573. Blumstein. D. T.. L. Cooley, J. Winternitz, and J. C. Daniel. 2008. Do yellow-bellied marmots respond to predator vocalizations? Behavioral Ecology and So- ciobiology 62:457-468. Caraco. T., S. Martindale, and H. R. Pulliam. 1980. Avian time budgets and distance to cover. Auk 97:872-875. CUTHILL, I. C. AND A. 1. HOUSTON. 1997. Managing time and energy. Pages 97-120 in Behavioural ecology: an evolutionary approach (J. R. Krebs and N. B. Davies, Editors). Fourth Edition. Blackwell Scientific Publi- cations, Oxford, United Kingdom. Dill, L. M. and A. H. G. Fraser. 1984. Risk of predation and the feeding behavior of juvenile coho salmon {Oncorhynchiis kisiitch). Behavioral Ecology and Sociobiology 16:65-71. Fenn, M. G. P. and D. W. MacDonald. 1995. U.se of middens by red foxes: risk reverses rhythms of rats. Journal of Mammalogy 76:130-136. Krebs, J. R. and A. Kacelnik. 1991. Decision-making. Pages 105-136 in Behavioural ecology: an evolution- ary approach (J. R. Krebs and N. B. Davies, Editors). Fourth Edition. Blackwell Scientific Publications, Oxford, United Kingdom. Lima, S. L. 1985. Maximizing feeding efficiency and minimizing time exposed to predators: a trade-off in the Black-capped Chickadee. Oecologia 66:60-67. Lima, S. L. 1998. Stress and decision-making under the risk of predation: recent developments from behavioral, reproductive, and ecological perspectives. Advances in the Study of Behavior 27:215-290. Lima, S. L. and P. A. Bednikoff. 1999. Temporal variation in danger drives antipredator behavior: the predation risk allocation hypothesis. American Naturalist 153:649-659. Lima, S. L. and L. M. Dill. 1990. Behavioral decisions made under the risk of predation: a review and prospectus. Canadian Journal of Zoology 68:619-640. Lima, S. L. and T. J. Valone. 1986. IntJuence of predation risk on diet selection: a simple example in the grey squirrel. Animal Behaviour 33:155-165. Lucas, J. R., R. D. Howard, and J. G. Palmer. 1996. Callers and satellites: chorus behaviour in anurans as a stochastic dynamic game. Animal Behaviour 51:501- 518. McLean, I. G. and R. Rhodes. 1991. Enemy recognition and response in birds. Current Ornithology 8: 173-21 1. Morse, D. H. 1973. Interaction between tit flocks and Sparrowhawks Accipiter nisiis. Ibis 1 15:591-593. Mumme, R. L. 2003. Economic decisions and foraging trade-offs in chickadees. Pages 231-237 in Exploring animal behavior in laboratory and field (B. J. Ploger and K. Yasukawa, Editors). Academic Press, New York, USA. Neville, J. 2004. Bird songs — Western Boreal Forest. Neville Recording. Salt Spring Island. British Colum- bia. Canada. Peake, T. M. 2005. Eavesdropping in communication networks. Pages 13-37 in Animal communication networks (P. McGregor, Editor). Cambridge Univer- sity Press, Cambridge, United Kingdom. Pulliam, H. R. and G. S. Mills. 1977. The use of space" by wintering sparrows. Ecology 58:1393-1399. Roth, T. C. and S. L. Lima. 2007. Use of predictable prey hotspots by an avian predator: purpo.seful randomness? American Naturalist 169:264-273. Roth, T. C., J. G. Cox, and S. L. Lima. 2008. Can foraging birds asse,ss predation risk by scent? Animal Behav- iour 76:2021-2027. SAS. 2007. JMP 7.0. SAS Institute Inc., Cary, North Carolina, USA. SHORT COMMUNICATIONS 173 Schmidt, K. A., E. Lee, R. S. Ostfeld, and K. Sieving. 2008. Eastern chipmunks increase their perception ot predation risk in response to titmouse alarm calls. Behavioral Ecology 19:759-763. SiH, A. 1980. Optimal behavior: can foragers balance two conflicting demands? Science 210:1041-1043. Valone, T. J. and S. L. Lima. 1987. Carrying food items to cover for consumption: the behavior of ten bird species feeding under the risk of predation. Oecologia 7 1 :286-294. Van der Veen, 1. T. 2002. Seeing is believing: infor- mation about predators influences Yellowhammer behavior. Behavioral Ecology and Sociobiology 51: 466-47 1 . Willson, M. F. 1971. Seed selection in .some North American finches. Condor 73:415-429. Willson, M. E. and J. C. Harmeson. 1973. Seed preferences and digestive efficiency of Cardinals and Song Sparrows. Condor 75:225-234. The Wilson Journal of Ornithology 122( 1 ): 173-177, 2010 Observations and Predation of a Coral-billed Ground Cuckoo {Carpococcyx renauldi) Nest in Northeastern Thailand Korakoch Pobprasert' and Andrew J. Pierce ABSTRACT. — We found the first documented wild nest of a Coral-billed Ground Cuckoo {Carpococcyx renauldi) at Khao Yai National Park, Thailand in June 2007. The large stick nest was monitored for 24 days including 1095 hrs of video footage; it contained two eggs and was in dense vegetation 4.85 m above the ground. Nest attentiveness of adults was almost constant with both birds taking turns to incubate or brood. Food delivered to nestlings included lizards, nestlings, a snake, frogs, earthworms, and other invertebrates. Nest defense was observed against several known nest predators but the nest ultimately failed due to predation by a pig-tailed macaque (Macaca nemistrina). These observations provide insight into the breeding ecology of the only two congeners, C. viridis and C. radiceus, for which little is known of their ecology and both are endangered. Received 27 March 2009. Accepted 18 September 2009. There is an extreme paucity of studies on the breeding ecology of tropical Asian birds (Sodhi and Brook 2006) and they should not be assumed to have the same life histories as the more widely studied neotropical birds. It is essential that studies are undertaken for the long-term manage- ment of bird communities as forest cover and forest quality continues to decline in Asia (FAO 2009). There are 1 1 species of non-parasitic cuckoos on mainland Southeast Asia including: six mal- kohas (Phaenicophaeus), four coucals {Centro- ‘ Conservation Ecology Program, King Mongkut’s University of Technology Thonburi, 83 Moo 8 Thakham, Bangkhuntien, Bangkok 10150, Thailand. ^Corresponding author; e-mail: andyp67@gmail.com pus), and one ground cuckoo (Carpococcyx). There are two other ground cuckoos which inhabit the forests of Sumatra and Borneo. The Sumatran Ground Cuckoo (Carpococcyx viridis) has only recently been rediscovered (Zetra et al. 2002) and nothing is known of its breeding ecology; it is Critically Endangered (lUCN 2008). The Bornean Ground Cuckoo (C. radiceus) is Near Threatened (lUCN 2008) and what is known of its breeding ecology is vague and ambiguous derived from a few captive birds and questionable museum specimens (Long and Collar 2002). Most wide- spread of the three species is the Coral-billed Ground Cuckoo (C. renauldi) (CBGC), that occurs in Thailand, Cambodia, Laos, and Tonkin and Annam (Vietnam). It is an uncommon resident of broad-leaved evergreen forest and secondary growth up to 1,000 m asl in Thailand (Robson 2000); its conservation status is of Least Concern (lUCN 2008). The only documented nesting attempts are of captive pairs. A pair at Metro Toronto Zoo produced 13 fledglings and clutches of up to five eggs were laid between early April and mid-August from 1976 to 1979 (Atkinson 1982). Robilleret al. (1992) and Rinke ( 1999) both document nests of captive pairs with clutches of 3-4 eggs and nestlings reared by hand. We present details, obtained through video recordings, of the first documented wild nest of a Coral-billed Ground Cuckoo, including its preda- tion by pig-tailed macaques (Macaca nemistrina leonina) in Khao Yai National Park, northeastern Thailand in 2007. 174 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. I. March 2010 METHODS We initiated a 3-year study of nest predation in 2006 on birds in the Mo-Singto area of Khao Yai National Park (14° 26' N, 101° 22' E), Nakhon Nayok Province, Thailand. Small waterproof Eujiko™ cameras (65 X 65 X 85 mm) with infrared diodes for 24-hr surveillance were deployed 3-5 m from nests of ground and understory nesting birds. The cameras were attached by cables to a digital video recorder (DVR) placed 20-25 m from the nest whereby the power source and DVR could be changed without disturbing the nesting birds (Pierce and Pobprasert 2007). A small microphone was also fitted within the surrounding case of the camera. Data from the DVR were stored on DVDs for later analysis. Cameras have been shown to have little effect on nesting success when placed off the ground (King et al. 2001, Staller et al. 2005, Pierce and Pobprasert 2007). Measurements of the nest were taken after it had failed. Male Coral-billed Ground Cuckoos are slightly larger than females, but they are monomorphic in plumage and males and females are indistinguishable in the field. RESULTS An adult Coral-billed Ground Cuckoo was observed to Ily down from a tree within a dense tangle of woody climbers on 13 June 2007. Closer inspection on 17 June revealed a nest containing two white eggs. A camera was placed in an adjacent tree ~4 m from, and slightly below, the nest which still contained two eggs at 1400 hrs on 18 June. The sitting bird flushed quickly and silently as we approached the nest area on each of these visits despite our attempts to avoid disturb- ing it. An adult returned to incubate within an hour of us leaving the area after initial setting of the camera. Apart from a mechanical failure for 8 hrs on I day during incubation, the cameras recorded continuously until the nest was predated on the morning of 7 July. The adults were only disturbed once following placement of the camera when a person inadvertently pas.sed close to the nest. Ne.st Placement. — ^The nest was in seasonally wet evergreen forest at 725 m asl. The canopy was generally 20-30 m high with emergent trees up to —60 m, but the area within 20 m of the nest included an old tree fall with most trees less than 10 m in height and covered with masses of woody climbers. The nest was an untidy pile of dead twigs in the shape of a shallow cup similar to that of a large raptor. It was lined with a few fresh small twigs that still had green leaves. It was placed at the base of a branch forking from the leaning trunk of a tree with a diameter at breast height of ~10cm, and further supported by woody climbers. The rim of the nest was 4.85 m above ground and vegetation, in the form of small branches and woody climbers, extended 50- 60 cm above the nest. The outer diameter of the nest was 33 cm, and the total height was —25 cm. Parental Care and Food Provisioning. — Both adults shared incubation duties and one adult was on or close to the nest at all times with few exceptions. There may have been some contact calls from the returning bird as the sitting bird became aware of the incoming bird shortly before it arrived at the nest but, if so, they were too soft to be picked up by the microphones. Both adults usually brought a small twig complete with fresh green leaves and placed it in the nest during the egg stage before commencing incubation. The adults changed brooding only twice per day with changeover times quite variable; the first between 0921 and 1338 hrs and the second between 1612 and 1754 hrs. Changeovers typical- ly took <30 sec with two of exceptionally longer duration of 2 min 32 sec and 8 min 32 sec. Excluding these apparent anomalies, changeovers, from when the sitting bird left the nest until the other bird resumed incubation, averaged (± SD) 20 ± II sec (/; = 24). The bird incubating during the day averaged (± SD) 5 hrs 36 ± 52 min {n = 12) on the nest while the bird that incubated overnight averaged 18 hrs 43 ± 50 min (// = 12) on the nest. We were unable to identify gender of the adults; however, for 9 days where the data are continuous, and one or both adults were in view at all times, the same adult incubated over night as on the previous night. It is likely therefore that only one male or one female incubated at night and the other during the day for the entire period. Gaps in attentiveness after the eggs hatched meant we could not ascertain the proportion of time spent brooding by each bird during the nestling stage. The first egg hatched in the early morning of 1 July when the adult started pecking at something beneath it.self and ate what was presumably eggshell (no eggshells were removed and only small fragments of eggshell were found in the nest after it had failed). The second egg hatched on the morning of 2 July when the same sequence of SHORT COMMUNICATIONS 175 events occun-ed. Changeovers followed the same pattern during hatching as in the incubation period and it was not until 1 147 hrs on 2 July that the first food item was brought to the nest. Attentive- ness at the nest continued to be almost constant throughout the nestling period although visitation rates increased to provision the nestlings. There were two visits with food on 2 July followed by 16, 24. 33, and 29 visits on the sub.sequent 4 days and a further seven visits in the morning of 7 July. The incoming bird during the nestling stage usually perched on the rim of the nest for a few seconds before the sitting bird stood up and left. The incoming bird would then hop into the nest and feed the nestlings before brooding them. Nestling Diet. — Only a limited number of prey items could be positively identified. These included lizards (3), nestlings of small birds (2), frogs (2), a blind snake (Typhlops or Rampho- typhlops), Orthoptera (2), and earthwomis (2). Prey was too large for the nestlings to swallow on two occasions and it was either taken away or eaten by the adult. Other prey items tentatively identified were worms, caterpillars, arthropods, and slugs or snails but many of the small food items appeared to be regurgitated or were out of view of the camera. Nest Defense. — Variable squirrels {Calloscinrus finlaysonii) and northern tree shrews (Tupaia helangeri) visited the nest nine times. The adults usually turned to face these relatively small predators with crown feathers depressed and tail raised and the intruder quickly withdrew. Two visits to the nest by small snakes (—50-100 cm long; probably Boiga spp.) were the only potential nocturnal nest predators, and both soon retreated rapidly after reaching the sitting bird which did not react. Pig-tailed macaques passed close by twice on the afternoon of 4 July. The adult adopted a threat posture with this no doubt more serious threat to the nest and, on these occasions, the macaques left the nest alone. A troop of macaques could be heard passing the nest at 0854 hrs on 7 July and the sitting CBGC became alert raising its body and spreading its tail. At exactly 0858 hrs it raised its crown feathers and turned to face a macaque with bill open, wings apart and tail spread over its back. This macaque dropped down quickly to the nest 17 sec later and the CGBC flushed to nearby branches. It stayed within —I m of the nest maintaining its threat posture and making short darting movements at the macaque. However, the macaque was not deterred and grabbed the nestlings one at a time, put them quickly into its mouth and left. The whole episode took 3 sec from when the CBGC left the nest until the macaque took the nestlings and departed. The CBGC returned to the rim of the nest; when another (smaller) macaque ap- proached, <1 min after the first attack, it flew at the macaque causing the latter to retreat; the CBGC flew on away from the nest. Two small macaques came to investigate the nest but left on finding it empty. The CBGC returned cautiously at 0918 hrs from below the nest and stood on the nest for 6 min before leaving. A CBGC entered the nest carrying a large insect at 0957 hrs but left again after nearly 1 min. After another visit at 1014 hrs with a different insect there were no further visits by 1500 hrs the following day when the camera was removed. DISCUSSION Coral-billed Ground Cuckoos have been re- corded nesting in captivity but ours is the first observation of a wild nest and provides more appropriate information about their nesting be- havior. The CBGC is under no immediate threat but there is an almost complete lack of informa- tion on the breeding ecology of the globally threatened congeners, Sumatran and Bornean ground cuckoos, the latter of which is known to be similar to CBGC (Long and Collar 2002). The CBGC nests at the Toronto Zoo were built by both adults and were comparable in dimen- sions to the nest we describe (Atkinson 1982) but. otherwise quantifying these nests is not appropri- ate due to necessary restrictions of captive birds. We found no fecal sacs in or near the nest after it failed and it is possible the adults ate them; Robiller et al. (1992) observed the nestlings defecated on the rim of their nest, but they were being hand fed with no adult birds present. Atkinson (1982:170) states “an incubation period of 28 days seems the normal condition”. However, the data are ambiguous and could be interpreted as being anywhere between 17 to 28 days while Payne (2005) reports the eggs hatch in 18-19 days. Eggs in our nest hatched within 15 days of seeing the full clutch and was probably in the early stages of incubation when found. The clutch of two eggs is at the lower end of clutch sizes reported from captive birds. Atkinson ( 1982) details 10 nests with clutches of 1-5 eggs (mean 3.1, median 4). Rinke (1999) details two nests with four eggs and Robiller et al. (1992) records 176 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. I, March 2010 two nests, each with three eggs but also states the clutch size to be 3-4 eggs. Ground cuckoos are notoriously difficult to observe in the wild and, apart from a few observations of birds coming to food waste at restaurants adjacent to the forest (Payne 2005; AP, pers. obs.), there are few published data on CBGC diet or behavior. Delacour and Jabouille (1931) reported small birds and mammals in their diet. Based on the stomach contents of a recently dead adult, found within a kilometer of this nest in 2006, they feed on a wide range of invertebrates; the stomach included the remains of several hundred individual prey items, especially Orthop- tera and Isoptera but also Coleoptera, Lepidop- tera, Hymenoptera, and a Chilapoda (centipede) (John Milne, in prep.) We did not definitively record any of these but it is possible they were among the many unidentified prey items fed to the nestlings. There is no evidence that they (or Bornean Ground Cuckoos [Long and Collar 2002]) eat fruit despite the generic name Carpo- coccy.x meaning fruit-cuckoo. Nest success in tropical forests is low; ~8-30% for most species with predation being the main cause of nest failure (Martin 1995, Robinson et al. 2000). Our research using video surveillance at nests in Khao Yai National Park (Pierce and Pobprasert 2007, and unpubl. data) identified the main nest predators as pig-tailed macaques, rodents (Muridae and Sciuridae), tree-shrews (Tupoia), birds, and snakes. Several of these were recorded visiting the CBGC nest. We recorded similar size snakes disturbing sitting Puff-throated Bulbuls (Alophoxius palUdiis) and trogons (Har- pactes spp.) and, being able to take the eggs or nestlings (Pierce and Popbrasert 2007). Rinke ( 1999) shows a CBGC on a nest in captivity in a defense posture similar that recorded at our nest. The CBGC in our study was able to deter macaques on two occasions by adopting this stance, although ultimately unsuccessful against a large macaque. A high level of nest attentiveness has been recorded for other non-parasitic cuckoos in the region (Payne 2005), and it has been suggested this may have developed in response to high predation pressures observed in the tropics (Auer et al. 2007). The con.stant vigilance of the adult CBGC at their nest no doubt stopped many potential predators from taking eggs or nestlings. However, adults were highly .sensitive to distur- bance, Hushing from the nest and not returning for an hour or more afterwards, even when we approached quietly knowing the nest location. These results highlight the potential increased predation risk caused by human disturbance of nesting birds that might otherwise be able to defend their nest. ACKNOWLEDGMENTS We thank Will Duckworth. Kim McConkey, and two anonymous reviewers for valuable comments on an earlier draft. John Milne kindly provided details of the CBGC stomach contents and Matthias Fehlow found and translated the two German references. We are grateful to Khao Yai National Park and the Department of National Parks, Wildlife and Plants Conservation for facilitating our research. We thank our fellow researchers for assistance and support. This project was supported by King Mongkut’s University of Technology Thonburi, Bangkok with grant BID-BD-RD-48-02 KMUTT from the National Research Council of Thailand and the TRF/BIOTEC Special Program for Biodiversity Research and Training Grant BRT 350002. LITERATURE CITED Atkinson, R. W. 1982. Breeding the Renauld’s or Coral- billed Ground Cuckoo Carpococcyx renauldi at the Metro Toronto Zoo. International Zoo Yearbook 22:168-171. Auer, S. K.. R. D. Bassar, J. J. Fonttaine, and T. E. Martin. 2007. Breeding biology of passerines in a subtropical montane forest in northwestern Argentina. Condor 109:321-333. Delacour. J. and P. J.abouille. 1931. Les oiseaux de rindochine fran^aise. 14. Exposition Coloniale Internationale. Paris, France. Food and Agriculture Organization of the United Nations (FAO). 2007. State of the world's forests 2009. FAO. Rome, Italy, www.fao.org (accessed 8 September 2009). lUCN. 2008. The lUCN Red List of threatened species. http://www.iucnredlist.org/ (accessed 8 September 2009). King, D. 1., R. M. DeGraaf, P. J. Champlin, and T. B. Champlin. 2001. A new method for wireless video monitoring of bird nests. Wildlife Society Bulletin 29: 349-353. ^ Long, A. J. and N. J. Collar. 2002. Distribution, status and natural history of the Bornean Ground Cuckoo Carpococcyx radiatiis. Forktail 18:101-119. Martin, T. E. 1995. Avian life history evolution in relation to nest sites, nest predation, and food. Ecological Monographs 65:101-127. Payne, R. B. 2005. The cuckoos. Oxford University Press, Oxford. United Kingdom. Pierce. A. J. and K. Pobprasert. 2007. A portable system for continuous monitoring of bird nests using digital video recorders. Journal of Field Ornithology 78:322- 328. Rinke, D. 1999. Zucht des Renauldkuckticks (Carpococcyx renaiddi) im Vogelpark. Walsrode 22:212-214. Robiller, F., H. Michi, and M. Michi. 1992. Uber den SHORT COMMUNICATIONS 177 Renauld kuckuck (Carpococcyx rcnaiiUli Oustalet, 1896) und seine Zucht in der Forschnngsstalion Ornis Mallorca. Tropische Vogel 13:92-97. Robinson, W. D., T. R. Robinson, S. K. Robinson, and J. D. Brawn. 2000. Nesting success of understory forest birds in central Panama. Journal of Avian Biology 31:31-164. Robson, C. 2000. A field guide to the birds of Thailand and South-East Asia. Asia Book Co. Ltd., Bangkok, Thailand. SoDHi. N. S. AND B. W. Brook. 2006. Southeast Asian biodiversity in crisis, Cambridge University Press, Cambridge, United Kingdom. Staller, E. L., W. E. Palmer, .1. P. Carroll, R. P. Thornton, and D. C. Sisson. 2005. Identifying predators at Northern Bobwhite nests. Journal ot Wildlife Management 69:124-132. Zetra, B., A. Rafiastanto, W. M. Rombang, and C. R. Trainor. 2002. Rediscovery of the critically endan- gered Sumatran Ground Cuckoo Carpococcyx viridis. Forktail 18:63-65. The Wilson Journal of Ornithology 122( 1 ): 177-1 80, 2010 Diet of Chinese Grouse {Tetrastes sewerzowi) during Preincubation Wang Jie,' - Yang Chen,' - Lu Nan,' - Fang Yun,' and Sun Yue-Hua'-^ ABSTRACT. — The endemic Chinese Grouse (Tetra- stes sewerzowi) inhabits subalpine coniferous forests. Little is known about its diet in the breeding season, especially while birds feed on the ground. Analysis of crop contents indicated that willow (Salix spp.) was the primary food of males (>98% wet weight, n — 4), whereas Dragon spruce (Picea asperata) seeds and willow were the primary foods of females (both >40% wet weight, n = 2). Dragon spruce seeds, invertebrates (mainly ants), and forbs were frequently consumed by females, but seldom by males, possibly to meet the nutrient constraints of egg formation. We suggest the different diets of males and females of this monoga- mous species may be the result of females allocating more time to searching for scarce, nutritious food, whereas males spend more time in vigilance behavior. Received 2 April 2009. Accepted 5 September 2009. The Chinese Grouse (Tetrastes sewerzowi), endemic to the coniferous forests of Qinghai- Tibet Plateau, is one of the rarest grouse species in the world (Storch 2000) and is legally considered an endangered species in China (Zheng and Wang 1998). The small body size of Chinese Grouse (Sun et al. 2005) may reduce its ability to meet nutritional requirements when feeding on plants because gut capacity scales linearly with body mass, whereas mass-specific nutritional require- ments increase with decreasing body mass (Dem- ' Key Laboratory of Animal Ecology and Conservation Biology, Institute of Zoology, Chinese Academy ol Sciences, Beijing 100101, PR China. ’Graduate School of Chinese Academy of Sciences, Beijing 100049, PR China. ^Corresponding author; e-mail: sunyh@ioz.ac.cn ment and Van Soest 1985). Small body size also precludes storage of energy and nutrients in the body for egg formation (Swenson et al. 1994). Chinese Grouse have the greatest nutritional investment in clutches among tetraonids (Sun et al. 2005) and longer egg laying intervals (48 hrs) than other grouse (Sun et al. 2002). Thus, nutrient constraints possibly exist for egg formation for this species and female Chinese Grouse may rely heavily on nutritionally rich food during the prelaying and laying period to gain mass and produce eggs when food may not yet be plentiful (Nager 2006). The diet of Chinese Grouse during the breeding season is poorly known and difficult to clarity by observation because of dense vegetation and birds not habituated to observers. We present informa- tion on the diet of this monogamous grouse from analysis of crop contents to help understand its foraging strategy to meet nutrient constraints of egg formation. METHODS Study Area. — Our study area was in the Lianhuashan Natural Reserve, southern Gansu Province, China (34 57' N, 103° 46' E). Eorests occur on north-facing, and some northeast and northwest-facing slopes at elevations of 2,600- 3,200 m; only shrubs and grasses grow on south- facing slopes. The main cover types are: (1) coniferous forest dominated by Dragon spruce (Picea asperata) and Earges fir (Abies fargesii). (2) coniferous-deciduous forest, including spruce, fir, Himalayan birch (Betula utilis), and willow 178 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 1. March 2010 TABLE 1. Loods in crops of six Chinese Grouse at Lianhuashan, Gansu, China (Collection dates; Male # 1, 25 Apr; Male # 2, 9 May; Male # 3, 3 May; Male # 4, 17 May; Lemale #1,16 Apr; Female # 2, 14 May). Portion of wet weight (%) Food Types of food Male #1 Male #2 Male #3 Male #4 Female #1 Female #2 Invertebrate.s (ant, midge)“ 0 0 0 0 3.0 9.9 Willow Twigs, buds, catkins, leaves 100 100 98.3 100 56.0 43.3 Dragon spruce Seeds 0 0 1.7 0 40.8 43.3 Lorbs^ Leaves 0 0 0 0 0.2 3.5 Total wet mass, g 2.3 0.3 3.2 0.3 4.9 1.5 ^ Most invertebrates were ants (Hymenoptera; Formicidae); only one midge (Diptera) occurred in the crop of Female # 2. Included Fraf>aria orienuilis, Gcttm atcppicum. and Triosteiim piiuuilifulum. (Salix spp.), and (3) shrublands, including willow, sea buckthorn {Hippophae rhamnoides), and barberry (Berheris spp.). The study area has been described by Sun et al. (2003). Field and Analysis Procedures. — We defined the preincubation period as 20 April-30 May, which represents 2 weeks before egg laying to when the last egg is laid. We obtained crops of four males and two females that died from predation or during capture while attaching radio transmitters. One female which died on 16 April was also included in our analysis. Identification of gender of grouse carcasses was based on PCR amplification of CHD genes using P2/P8 primers (Griffiths et al. 1998) and sexl/sex2 primers (Wang and Zhang 2009), or based on the presence of the black chin patch for males (Bergmann et al. 1996). Age (adult or yearling) was not assigned, as there is no known reliable method (Swenson et al. 1996). Crops were removed and contents were identified, sorted, and weighed to the nearest 0.001 g. Plants were identified to species and invertebrates were identified to Order. We sampled 1 3 plots (40 X 40 cnr) where grouse pecked and counted the surface seeds to assess their availability. Seeds from cones were collected, dried to constant weight at 50° C and homogenized to measure nitrogen (Foss Kjeltec™ 2100, Hilleroed, Denmark), lipid (Soxhlet), fiber (ANKOM 200/220 Fiber Analyzer, Fairport, NY, USA), tannin (Phosphomolybdium Tungstic Acid colorimetric method, China Entry/Exit Inspection and Quarantine Association 1999), calcium and phosphorus (AOAC 1995), ash (combustion at 550 C), and energy (PARR 1281 bomb calorim- eter, Moline, IE, USA). SPSS 13.0 (SPSS Inc., Chicago. IE. USA) was used for all statistical analysis. Data are expressed as x ± SE. RESULTS Willow, including catkins, leaves, twigs, and buds occun'ed in all crops and comprised 99.6 ± 0.4% and 49.7 ± 6.4% of the wet biomass for males (/? = 4) and females (n = 2), respectively (Table 1). Dragon spruce seeds were consumed intensively by females but seldom by males (42. 1 ± 1.3 vs. 0-1.7% of the wet biomass). Inverte- brates (ants and midges) and forbs (strawberry [Fragaria orientalis], yellow avens [Geum alep- piciim], and feverwort [Triosteum pinnatifidiim]) ranged from 0.2 to 9.9% of the wet biomass in crops of females, but were not found in crops of males. Earges fir seeds were not found in crops of either males or females. Compared to Dragon spruce seeds, Earges fir seeds occuired in low densities on the ground (35 ± 67 vs. 296 ± 367 seeds/m", paired t test, P = 0.04, n = 13), and had lower protein and ash content, and higher fiber, tannin, phosphorus, and calcium content, whereas their lipid and energy contents were similar (Fig. 1). DISCUSSION Willow was the primary food of male grouse (>98% of wet biomass), while Dragon spruce seeds as well as willow were the primary food of females. Females ate invertebrates and forbs, whereas males did not. These findings indicate the diet of female grouse differed from males prior to incubation. This is consistent with our observations that both males and females foraged in willows intensively at dawn and dusk, but females allocated more time than their mates to foraging on the ground during the day (60.8 vs. 15.1%) (Sun 2004). Egg formation is a demanding process in terms of energy and nutrient requirements (Nager 2006). SHORT COMMUNICATIONS 179 S’ O) E S o S’ TO E Food species FIG. 1. Content (x ± SE) of protein, fiber, tannin, lipid, ash, phosphorous, calcium, and energy in samples of seeds of Dragon spruce (Picea asperata) (PA) {n = 3) and Farges fir (Abies fargesii) (AF) (/? = 3) collected from the ground. Some female grouse shift dietary patterns during spring to select high-nutrient foods, including invertebrates and forbs, which are high in crude protein, calcium, and phosphorus (Swenson 1991, Barnett and Crawford 1994, Gregg et al. 2008). Female Chinese Grouse in our study similarly shifted their diet from willow to selected ground seeds, invertebrates, and forbs during the prelay- ing period. Female grouse in smaller species with fewer energy and nutrients reserves are more dependent on exogenous sources (i.e., nutrition- ally rich food) for egg production (Moss et al. 1975, Brittas 1988, Swenson et al. 1994). Green leaves typically contain lower concentrations of nutrients and are less digestible than animal tissue (Robbins 1993). Butterfield and Coulson (1975) estimated that consumption of 8% insect material in the food of Red Grouse (Lagopiis lagopiis scoticiis) could increa.se nitrogen and phosphorus intake by 67 and 51%, respectively. Dragon spruce seeds had higher protein, energy, and lipid, and lower fiber and tannin than willow buds (unpubl. data). Thus, Dragon spruce seeds, invertebrates, and forbs may be important foods for Chinese Grouse hens during preincubation. Dragon spruce seeds rather than Farges fir seeds were consumed by hens possibly because the latter are less available on the ground and contain higher tannin and lower protein contents. Nutritionally rich foods, including inverte- brates, forbs, and Dragon spruce seeds are not highly available for prelaying hens, as it is still cold (average <10° C) and some areas are covered by snow until mid-May in the subalpine forest. These foods are less available in years of late snowmelt and could affect the breeding performance of Chinese Grouse. Clutch size was 6.2 ± 0.3 (/? = 9) and number of chicks hatching was 2.1 ± 0.9 (/; =11) in a year that snowmelt was delayed by —10 days, whereas the corre- sponding figures were 7.0 ± 0.3 (/? = 13) and 5.6 ± 0.8 (n = 13) in a good year (unpubl. data). Our sample size was small, but we found females commonly ate nutritionally rich loods whereas males seldom foraged on them. Male Chinese Grouse in this monogamous species (Sun 2004), spend more time (60.8 vs. 15.1%) in vigilance behavior than females, allowing their mates more time to search for nutritious food. The Chinese Grouse is listed in Category 1 of 180 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 1. March 2010 nationally protected animals and for which hunting is forbidden (Sun 2000). The crop samples and diet analysis reported here are valuable and add to the knowledge of foods used by Chinese Grouse. This information is also helpful in understanding the social behavior of small monogamous birds. ACKNOWLEDGMENTS We are grateful to J. E. Swenson. S. J. Hannon, Yi Tao, and J. F. Bendell for helpful comments, and the staff of Lianhuashan Natural Reserve for assistance in the field. We appreciate the constructive comments of M. A. Gregg, C. E. Braun, and an anonymous reviewer on earlier drafts of the manuscript. This study was supported by the National Natural Science Foundation of China (Project 30620- 130110). Licenses to collect the birds were issued by the Chinese Directorate for Nature Management. LITERATURE CITED Association of Analytical Chemists (AOAC). 1995. Official methods of analysis. Sixteenth Edition. Associ- ation of Analytical Chemists, Arlington, Virginia, USA. Barnett, J. K. and J. A. Crawford. 1994. Pre-laying nutrition of Sage Grouse hens in Oregon. Journal of Range Management 47:114-1 18. Bergmann, H.-H., S. Klaus, F. Muller, W. Scherzinger, J. E. Swenson, and J. Wiesner. 1996. Die haselhuh- ner, Bonasa honasia und B. sewerzowi. Die Neue Brehm-Biicherei. Westarp Wissenschaften. Magdeburg, Germany. Brittas, R. 1988. Nutrition and reproduction of the Willow Grouse Lagopiis lagopus in central Sweden. Ornis Scandinavica 19:49-57. Butterfield, J. and J. C. Coulson. 1975. Insect food of adult Red Grouse Lagopus lagopus scoticus (Lath.). Journal of Animal Ecology 44:601-608. China Entry/Exit Inspection and Quarantine Associ- ation. 1999. Cereals and feedstuff for import and export — Method for the determination of tannin content [SN/T0800.9-1999]. China Entry/Exit Inspec- tion and Quarantine Association, Beijing. Demment, M. W. and P. j. van Soe.st. 1985. A nutritional explanation for body-size patterns of ruminant and nonruminant herbivores. American Naturalist 125: 641-672. Gregg, M. A., J. K. Barnett, and J. A. Crawford. 2008. Temporal variation in diet and nutrition of preincu- bating Greater Sage-Grouse. Rangeland Ecology and Management 61:535-542. Griffiths R, M. C. Double, K. Orr, and R. J. G. D.awson. 1998. A DNA test to sex most birds. Molecular Ecology 7:1071-1075. Moss. R., A. Watson, and R. Parr. 1975. Maternal nutrition and breeding success in Red Grouse (Lagopus lagopus scolicus). Journal of Animal Ecology 44:233-244. Nager, R. G. 2006. The challenges of making eggs. Ardea 94:323-346. Robbins, C. T. 1993. Wildlife feeding and nutrition. Academic Press, San Diego, California, USA. Storch, I. 2000. Grouse status survey and conservation action plan 2000-2004. WPA/BirdLife/SSC Grouse Specialist Group, lUCN, Gland, Switzerland. Sun. Y.-H. 2000. Distribution and status of the Chinese Grouse Bonasa sewerzowi. Wildlife Biology 6:271- 275. Sun, Y.-H. 2004. Distribution, reproduction strategy and population biology of the Chinese Grouse (Bonasa sewerzowi). Dissertation. Beijing Normal University, Beijing, China. Sun, Y.-H., Y. Fang, S. Klaus, C.-X. Jia, and G.-M. Zheng. 2002. The application of data logger technique to the study of incubation rhythms of the Chinese Grouse. Journal of Beijing Normal University (Natural Science) 38:260-265. Sun, Y.-H., Y. Fang, J. E. Swenson, S. Klaus, and G.-M. Zheng. 2005. Morphometries of the Chinese Grouse Bonasa sewerzowi. Journal of Ornithology 146:24-26. Sun, Y.-H., J. E. Swenson, Y. Fang, S. Klaus, and W. Scherzinger. 2003. Population ecology of the Chi- nese Grouse, Bonasa sewerzowi, in a fragmented landscape. Biological Conservation 110:177-184. Swenson, J. E. 1991. Social organization of Hazel Grouse and ecological factors influencing it. Dissertation. University of Alberta, Edmonton, Canada. Swenson, J. E., L. Saari, and Z. Bonczar. 1994. Effects of weather on Hazel Grouse reproduction: an allome- tric perspective. Journal of Avian Biology 25:8-14. Swenson, J. E., Y.-H. Sun, and N. Liu. 1996. A potential method for age determination of the Chinese Grouse Bonasa sewerzowi. Journal fur Ornithologie 137:255- 258. Wang, N. and Z.-W. Zhang. 2009. The novel primers for sex identification in the Brown-eared Pheasant and their application to other species. Molecular Ecology Resources 9:186-188. Zheng, G.-M. and Q.-S. Wang. 1998. China Red Data Book of Endangered animals, Aves. Science Press, Beijing. SHORT COMMUNICATIONS 181 The Wilson Joiirncil of Ornithology 122(1): 18 1- 182, 2010 Observations of a Possible Foraging Tool used by Common Ravens Paul F. Frame ABSTRACT. — Common Ravens [Coitus cora.x) are opportunistic generalist foragers. Ravens during winter, in some portions of their range, rely heavily on meat scavenged from carcasses. Ravens use a variety of strategies to find carcasses on the landscape, including keying on audible cues that suggest the presence of a food source. I documented Common Ravens investi- gating a simulated animal distress call on 13 of 17 trials, suggesting that investigating animal distress vocaliza- tions may be one tool in the suite of foraging strategies used by ravens. Received 6 June 2009. Accepted 6 September 2009. The Common Raven [Coitus cora.x) is an opportunistic generalist forager and uses a variety of food sources including carrion, rodents, insects, nestlings, eggs, seeds, fruit, and garbage (Bent 1946, Heinrich 1988, Engel and Young 1989, Kelly et al. 2005). Many ravens at northern latitudes rely heavily on meat from carcasses of ungulates during winter (Temple 1974, Heinrich 1988, Hayes et al. 2000, Vustich et al. 2004). These carcasses are an unpredictable food source that is often difficult to locate on the landscape (Heinrich 1988, Stabler et al. 2002, Vustich et al. 2004). Stabler et al. (2002) summarized strategies that carrion feeders, such as ravens, use when foraging for carcasses as: flying the landscape and finding them visually by chance, foraging where conspecifics or other species are seen foraging, responding to vocalizations from other carcass feeders, following other ravens that have previ- ously located food from nocturnal roost sites, and associating with predators that make carcasses available. White (2005) found that ravens were attracted to gunshots as a strategy for finding ungulate gut piles discarded by successful hunters. Ravens also cue in on appeasement calls of conspecifics (Heinrich et al. 1993) and likely the howling of wolves (Harrington 1978) as a means ' Department of Fisheries, Wildlife, and Conservation Biology, University of Minnesota. St Paul, MN 55108, USA. ‘Current address: Carcross Tagish First Nation, P. O. Box 130, Carcross. YT YOB I BO, Canada; e-mail: pframe@ ualberta.net of finding food. These examples indicate ravens respond to audible stimuli as a foraging strategy. 1 present observations made during winter of ravens responding to imitation prey distress calls and suggest this response is another tool in the suite of foraging strategies used by ravens. OBSERVATIONS I made these observations from 28 December through 12 January during winter 2000-2001 on the Superior National Eorest (SNE) in northeastern Minnesota (47° 50' N, 91° 25' W) while attempting to capture gray wolves [Canis lupus) for radio collai'ing. I located radio-marked wolves from fixed-wing aircraft and then homed in on the location with hand-held telemetry immediately after landing. I used a manual prey distress predator call to attract wolves sufficiently close to dart them with anesthesia for capture and radio collaring. I called wolves at kill sites, non-kill sites, and sites where kill information was unknown. I approached wolves from downwind to a distance they would be able to hear and respond to the predator call. I concealed myself so that wolves would not immediately identify me when they approached, usually by backing into spruce [Picea spp.) or balsam fir [Abies balsamea) trees, which also concealed me from above. The predator call was then blown to imitate a dying rabbit or young deer. Notes were kept of the change of signal strength and other occumences including ravens flying over. The actual distance between my location and wolves was unknown and varied from site to site. Typically 6-25 ravens have been reported at wolf kills (summarized in Vustich et al. 2004) with as many as 80 being present in some cases (Carbyn et al. 1993); between 0 and 16 ravens have been reported in association with wolves away from kills (Stabler et al. 2002). I observed between one and five ravens responding to the predator call on 13 of 17 trials (76%) by flying within 100 m or directly over my location. Those that flew directly over me approached in a manner that suggested directed Bight. Those that came within 100 m Bew directly toward me and then changed course when they appeared to detect me. Ravens 182 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. I, March 2010 responded during all trials near known kill sites (/? = 8), once when it was known there was no kill present {n = 3), and four of six times when kill information was unknown. DISCUSSION I observed ravens investigating a sound that simulates the distress cry of a dying animal. Ravens are known to use audible cues as a foraging technique (Heinrich et al. 1993, White 2005), but this behavior has not previously been reported regarding animal distress calls. Ravens congregate at wolf kill sites (Carbyn et all 993, Stabler et al. 2002) and it can be assumed the eight trials at known kills were conducted in the proximity of ravens. The response rate of 100% (// = 8) to trials in the proximity of ravens suggests this may be an established foraging technique. Alternatively, ra- vens may have been investigating the simulated distre.ss call out of curiosity. My observations are reported to pre.sent the possibility that ravens key on prey distress calls as one tool in a suite of foraging strategies. However, a more standardized study with controls is required to test for an effect. White (2005), in an experimental study, fired a rifle, blew an air horn, a whistle, and did nothing in an attempt to learn if ravens responded to gunshots. He found that ravens were attracted to the gunshot, but not the air horn or whistle. My observations were made incidental to another project and I did not have a control. White’s (2005) results suggest the raven behavior I observed was in response to the distress call and not a response to a novel loud noise. White (2005) suggested that ravens might respond to gun.shots because they associate the sound with other sounds that indicate potential food. It is possible prey distress calls are one such sound. A carcass for ravens is an unpredictable food source on the landscape (Heinrich 1988, Stabler et al. 2002, Vustich et al. 2004). Using prey distress calls to find a carcass would likely increase a raven’s resource intake and better insure an individual’s survival. ACKNOWLEDGMENTS These observations were made during a project funded by the Undergraduate Research Opportunities Program in the College of Natural Re.sources at the University of Minnesota. C. E. Braun, L. D. Mech, D. R. Stabler, Crow White, and an anonymous reviewer made useful comments on earlier drafts of this manuscript. M. E. Nelson provided field support and many idea generating conversations over the years. LITERATURE CITED Bent, A. C. 1946. Life histories of North American jays, crows, and titmice. Part 1. U.S. National Museum Bulletin Number 191. Carbyn, L. N., S. M. Oosenbrug, and D. W. Anions. 1993. Wolves, bison, and the dynamics related to the Peace-Athabasca Delta in Canada’s Wood Buffalo National Park. University of Alberta, Edmonton, Canada. Engel, K. A. and L. S. Young. 1989. Spatial and temporal patterns in the diet of Common Ravens in southwest- ern Idaho. Condor 91:372-378. Harrington, E. H. 1978. Ravens attracted to wolf howling. Condor 80:236-237. Hayes, R. D., A. M. Baer, U. Wotschikowsky, and A. S. Harestad. 2000. Kill rate by wolves on moose in the Yukon. Canadian Journal of Zoology 78:49-59. Heinrich, B. 1988. Winter foraging at carcasses by three sympatric corvids with emphasis on recruitment by the Raven, Corviis corax. Behavioral Ecology and Socio- biology 23: 141-156. Heinrich, B., J. M. Marzluff, and C. S. Marzluff. 1993. Common Ravens are attracted by appeasement calls of food discoverers when attacked. Auk 110:247-254. Kelly, J. P., K. L. Etienne, and J. E. Roth. 2005. Factors influencing the nest predatory behaviors of Common Ravens in heronries. Condor 107:402-415. Stahler, D., B. Heinrich, and D. Smith. 2002. Common Ravens, Corvus corax. preferentially associate with grey wolves, Canis lupus, as a foraging strategy in winter. Animal Behaviour 64:283-290. Temple, S. A. 1974. Winter food habits of raven on the Arctic Slope of Alaska. Arctic 27:41^6. Vustich, J. A., R. O. Peterson, and T. A. Waite. Raven scavenging favours group foraging in wolves. Animal Behaviour 67:1 17-1 126. White, C. 2005. Hunters ring dinner bell for ravens: experimental evidence of a unique foraging strategy. Ecology 86:1057-1060. SHORT COMMUNICATIONS 183 The Wilson Jonrmtl of Ornithology 122( 1 ): 1K3-185, 2010 Hanging Behavior of the Hooded Crow {Corviis coi nix) Mario Melletti' -^ and Marzia Mirabile" ABSTRACT. — We report observations of Hooded Crows (Corvns corni.x) hanging upside down in the wild, for both playing and harvesting acorns. This behavior was recorded in six different events at the same site (Villa Chigi’s urban park in Rome, Italy) and, in two cases, involved two individuals at the same time. Hanging behavior has been observed mainly in captive Northern Ravens (Conns cora.x) and in few cases in wild corvids. Our observations indicate hanging behav- ior can be used to obtain food. These observations confirm that corvids have enormous plasticity that can be adapted to obtain food. Received 5 March 2009. Accepted 19 July 2009. Play by birds is difficult to define and is a behavior that apparently has no purpose (Bekoff 1984, Heinrich 1999). Ficken (1977) noted that play is much more difficult to define in birds than in mammals. Birds may play socially, but only a few species exhibit the full range of play behaviors, from play chases to reciprocal object play (Fagen 1981, Ortega and Bekoff 1987). Some birds, such as corvids and parrots, exhibit more extensive social play than others (Fagen 1981, Iwaniuk and Pellis 2001). Corvids, among the Passeriformes, show the most complex play behavior including play with objects (e.g., objects carried into the air, dropped, and then caught with the beak many times), flight play, bathing play, vocal play, hanging (upside down posture from a branch holding with 1 foot, 2 feet or the beak), allospecific interactions, sliding, and ‘snowromp- ing’ (e.g., a raven lay on its breast and slid head forwards downhill in the snow) (Ficken 1977, Moffett 1984, Heinrich 1990, Ratcliffe 1997, Heinrich and Smolker 1998, Brazil 2002). Corvids may use cognitive tools in a way similar to apes, as reported in a study comparing mentality in these different animal groups (Emery and Clayton 2004). In fact, corvids have a brain ' Estacion Biologica De Donana, C.S.l.C. Depaitment ot Conservation Biology c/Americo Vespucio s/n. 41092 Seville, Spain. ^High Institute for Environmental Protection and Re- search, Via Brancati 48. 00144 Rome, Italy. -’Corresponding author; e-mail: mario.melletti@yahoo.it significantly larger than in other birds and, in proportion, their brain has similar size as that of the chimpanzee (Pan spp.). Emery and Clayton (2004) concluded that cognition in corvids and apes must have evolved through a process of divergent brain evolution with convergent mental evolution. Heinrich and Bugnyar (2005) studied the acquisition ol problem-solving behavioi in captive ravens. Their results support the idea that behavior of ravens in accessing meat on a string is not only a result of rapid learning, but may involve the understanding of cause-effect relation between string and food. Corvids and parrots can perform upside down hanging behavior, which can be considered as play (Elliot 1977, Ortega and Bekoff 1987, Heinrich and Smolker 1998, Diamond and Bond 2003). Hanging behavior has been observed in captivity many times in Northern Ravens (Corvus corax) (e.g., Gwinner 1966, Coombs 1978, Heinrich and Smolker 1998), but only a few records have been reported for corvids in the wild (McIntyre 1953, Gwinner 1966, Elliot 1977, Heinrich and Smolker 1998, Melletti 1999). Hanging behavior is primarily a type of play performed by highly social birds, such as corvids and parrots, which have the largest brain volume (Heinrich and Smolker 1998, Diamond and Bond 2003, Emery and Clayton 2004). Hanging behav- ior can be adopted for purposes other than play, but most reasons for this behavior remain unknown (Coombs 1978). Gwinner ( 1966) report- ed ravens performing hanging behavior to store food in an aviary and Melletti (1999) recorded a case of hanging behavior, lasting up to 5 min, in a wild raven for five consecutively times without apparent reason. Corvids have the ability to access food that is difficult to reach, in particular ravens in captivity (Heinrich 1995, Heinrich and Bugnyar 2()()5). Our observations show that Hooded Crows (Corvus cornix) are able to obtain acorns otherwise difficult to access from an upright posture. Use of hanging behavior by Hooded Crows demon- .strates the ability to solve a problem. To our knowledge this is the first observation reported of 184 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 1, March 2010 FIG. 1. Hanging behavior of Hooded Crows while feeding. Photograph by Mario Melletti. hanging behavior in wild Hooded Crows with the intent to obtain food. Little is still known about this behavior. OBSERVATIONS We made observations of hanging behavior by Hooded Crows from October 2005 through October 2006 in the urban park of Villa Chigi in Rome, Italy. The park is 5 ha in size and it is characterized by Stone pine (Finns pinea). Holm- oak (Querciis ilex), and open meadows. Data were collected weekly for 6 hrs/week divided between morning and afternoon (total number of hrs = 288). We conducted the study using lOX binoculars and a digital camera. Hooded Crows were observed for 96 of 288 hrs during which hanging in upside down posture was observed in six different events; two on 7 November 2005 (2 individuals), one on 14 January 2006 (1 individual), one on 6 October 2006 ( 1 individual), and two on 26 October 2006 (2 different individ- uals). Hanging behavior lasted from a few seconds up to 1 min during which birds held the wings pressed to the body and the head was held both horizontally and vertically (Fig. I ). DISCUSSION Hanging behavior was observed by McIntyre (1953) for wild Camon Crows (Corviis corone). Gwinner (1966) and Heinrich and Smolker ( 1998) for captive Northern Ravens, and Elliot (1977) and Melletti (1999) for wild Northern Ravens. Hooded Crows would hang upside down with one foot and then land upright on the ground. We observed two Hooded Crows that performed this behavior alternating one and both feet while hanging upside down. A Hooded Crow in hanging posture was imitated by another bird on two occasions, which performed the same posture at the same time. Gwinner (1966) reported similar cases of imitation in captive ravens. We observed hanging behavior by Hooded Crows in three events that .seemed as play, as described by other authors (Gwinner 1966, Heinrich and Smolker 1998), while in three other occasions the goal of the crows was to reach acorns of Holm-oaks, which were eaten or stored in the ground. Acorns were located on small twigs and upside down posture might seem to be the only way for crows to obtain them. The upright posture probably was not the best way to pick SHORT COMMUNICATIONS 185 acorns, because the twigs were not sufficiently strong to support the weight of crows in that position. By hanging upside down, crows were able to obtain acorns with their beak, before coming back to the ground. They remained upside down for few seconds. These observations indicate that hanging behavior is not only a type of play, but can be used to reach food, as also reported by Gwinner (1966) for captive ravens. Elliot (1977) suggested the hanging behavior of ravens was for display by courting males. These observations confirm that corvids have huge plasticity that can be adapted to solve different situations. ACKNOWLEDGMENTS We thank in particular Maria del Mar Delgado, Marc Bekoff, and Igor Krugersberg for useful comments and criticism. LITERATURE CITED Bekoff, M. 1984. Social play behavior. BioScience 34:228-233. Brazil, M. 2002. Common Raven Conns corcix at play; records from Japan. Ornithological Science 1 : 150—152. Coombs, F. 1978. The Crows. Batsford, London, United Kingdom. Diamond, J. and A. B. Bond. 2003. A comparative analysis of social play in birds. Behaviour 140:1091-1 1 15. Elliot, R. D. 1977. Flanging behavior in Common Ravens. Auk 94:777-778. Emery, N. J. and N. S. Clayton. 2004. The mentality of crows: convergent evolution of intelligence in corvids and apes. Science 306:1903-1907. Fagen, R. 1981. Animal play behavior. Oxford University Press, New York, USA. Ficken, M. S. 1977. Avian play. Auk 94:573-582. □winner, E. 1966. U ber einige Bewegungsspiele des Kolkraben (Conns corax L.). Zeitschrifl fur Tiei-psy- chologie 23:28-36. Heinrich, B. 1990. Ravens in winter. Banie and Jenkins, London, United Kingdom. Heinrich, B. 1995. An experimental investigation ot insight in Common Ravens (Corvus corax). Auk 1 12:994-1003. Heinrich. B. 1999. Mind of the raven: investigating and adventures with wolf-birds. Harper Collins, New York. USA. Heinrich, B. and T. Bugnyar. 2005. Testing problem solving in ravens: string-pulling to reach food. Ethology 111:962-976. Heinrich, B. and R. Smolker. 1998. Play in Common Ravens (Corvus cora.x). Pages 27-44 in Animal play, evolutionary, comparative and ecological perspectives (M. Bekoff and J. A. Byers, Editors). Cambridge University Press, Cambridge, United Kingdom. IWANIUK, A. N. AND S. M. Pellis. 2001. Do big-brained animals play more? Comparative analysis of play and relative brain size in mammals. Journal of Compara- tive Psychology 115:29-41. McIntyre, N. 1953. Curious behaviour of Canion Crow. British Birds 46:377-378. Melletti, M. 1999. Un caso di “hanging in un corvo imperiale Conns corax. Picus 1:112. Moffett, A. T. 1984. Ravens sliding in snow. British Birds 77:321-322. Ortega, J. C. and M. Bekoff. 1987. Avian play: comparative evolutionary and developmental trends. Auk 104:338-341. Ratcliffe, D. 1997. The Raven. T. & A. D. Poyser, London, United Kingdom. 186 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 1. March 2010 The Wilson Journal of Ornithology 122(1): 186- 187, 2010 Conspecific Brood Parasitism by the Dickcissel Brian D. Peer' ABSTRACT. — I report the first observation of conspecific brood parasitism in the Dickcissel (Spiia americana). I monitored 302 nests during a study of the interactions between parasitic Brown-headed Cowbirds (Molothrus ater) and Dickcissels. This and a possible second case of conspecific brood parasitism may have resulted from females laying in nests after their original nests were destroyed in nearby hayfields. Received 9 June 2009. Accepted 4 September 2009. Avian brood parasites lay their eggs in nests of other individuals thereby avoiding costs associat- ed with parental care. Obligate brood parasites such as the Brown-headed Cowbird (Molothrus ater) lay their eggs in nests of other species (Peer et al. 2005), whereas conspecific brood parasites lay in nests of individuals of the same species (Yom-Tov 1980). Conspecific brood parasitism (CBP) has been reported in 236 species (Yom-Tov 1980, 2001) and tends to be most common in precocial species such as waterfowl, colonial species, and species with specialized nest sites such as cavities (Yom-Tov 1980, 2001; Rohwer and Freeman 1989). Females may practice CBP to increase their fecundity by laying extra eggs in nests of other individuals, in addition to tending their own clutches (Soren.son 1991, Lyon 1993). Alterna- tively, females that do not have a nest may attempt to make the best of a bad situation and lay their eggs in nests of other females or, if their nests are destroyed during the laying cycle, they may lay eggs in nests of nearby neighbors (Hamilton and Orians 1965, Feare 1991; but see Rothstein 1993, Peer and Sealy 2000). 1 report the first known instance of CBP in the Dickcis.sel (Spiza americana). METHODS I monitored Dickcissel nests at a reclaimed mine from 2006 to 2008 in southern McDonough County, Illinois, USA during a study of the ' Department of Biological Sciences, Western Illinois University. Macomb. IL 61455, USA; e-mail: BD-Peer@ wiu.edu interactions between Brown-headed Cowbirds and their Dickcissel hosts. Dickcissel nests were inspected for presence of cowbird eggs and to record the Dickcissel’ s response to parasitism. Most nests were checked daily, and each egg was numbered. Nests were checked beginning at least 3 hrs following sunrise, after the time of day Dickcissels typically lay (unpubl. data). I checked nests sufficiently late to avoid missing an egg being laid 1 day and returning the second day and finding two additional eggs; this could lead to the eiToneous conclusion that a second female laid in the nest. Dickcissels nested in grasslands planted as part of the mine reclamation and were suiTounded by agricultural fields that included hay in which the Dickcissels also nested. Nests in hayfields were not monitored because the hay was mowed once a month destroying the nests. RESULTS A total of 302 Dickcissel nests was monitored. Modal clutch size was four eggs with a range of three to seven eggs. Nest #41 was found in the afternoon of 24 June 2008 with four eggs and I numbered all eggs at this time. Seven eggs were in this nest on 26 June. Because birds lay no more than one egg per day (Johnson 1999), at least one additional female laid an egg in this nest. The hayfields immediately adjacent to this nest were mowed the afternoon of 24 June. The nest contents remained the same until 3 July when five eggs had been removed, and two eggs remained. The nest was subsequently deserted. DISCUSSION This is the first reported observation of Conspecific Brood Parasitism (CBP) in Dickcis- sels. The rarity of CBP in this species is not surprising because Dickcissels do not demonstrate traits associated with conspecific brood parasites. What caused the parasitism that I observed'? Dickcissels at my site nested in grasslands and in adjacent hayfields. The hayfields were mowed regularly during the breeding sea.son and in 2008 the hay adjacent to nest #41 was mowed in the days prior to the parasitic event. It is possible the SHORT COMMUNICATIONS 187 parasitic female was nesting in the hayfield and her nest was destroyed by the mowing, after which she laid one of her remaining eggs in nest #41. However, I did not monitor nests in hayfields and I did not ascertain the genetic identities of the eggs. Thus it is unclear whether nest destruction precipitated this case of CBP. Birds may make the best of a bad situation by laying an egg in another nest if their own nests are destroyed during the laying stage (Hamilton and Orians 1965). Con- specific brood parasites such as European Star- lings {Stiiniis vulgaris) and Wood Ducks (A/.v sponsa) lay in nests of conspecifics after their nests have been destroyed or experimentally removed during the laying cycle (Haramis et al. 1983, Feare 1991). Species not known to practice CBP (Red-winged Blackbird [Agelaius phoeni- ceus], Great-tailed Crackle [Quiscalus mexica- nus]) do not lay parasitically after their nests have been experimentally removed (Rothstein 1993, Peer and Sealy 2000). Illinois was mostly grassland prior to European colonization. Today <1% of the original grassland habitat remains (Iverson 1988) and there is little suitable nesting habitat for Dickcissels and other grasslands birds (e.g., Herkert 2003). The density of Dickcissels at my site was high (upubl. data), possibly due to lack of suitable habitat in the western Illinois region. Dickcissels may nest in suboptimal habitat such as hayfields that are mowed causing nest failure during laying. These females may have responded by laying in nests of nearby females, which is a better strategy than dumping physio- logically committed eggs indiscriminately in the environment. The seven-egg clutch at this nest is the largest reported clutch for the Dickcissel, one egg more than what had been reported previously (Temple 2002). I also observed a second nest with a seven- egg clutch. On 26 June 2006 nest #81 was found with four eggs. On 28 June it contained six eggs, on 29 June seven eggs, and the contents remained the same until the eggs hatched. There were no laying irregularities during the time I monitored the nest that would suggest CBP, but a second female could have laid an egg in the nest before I found it on 26 June. This is only the second time a seven-egg clutch has been reported tor this species and the other was a result ol CBP. ACKNOWLEDGMENTS 1 thank G. K. Arnett anti the hreeman Mining Company lor allowing me to conduet my studies on their propeity. M. D. Benson and C. G. Peer assisted with the field work. LITERATURE CITED Feare, C. J. 1991. Inlraspecinc ne.st parasitism in Starlings Sliinuis vulgaris: effects of disturbance on laying females. Ibis 133:75-79. Hamilton III, W. J. and G. H. Orians. 1965. Evolution of brood parasitism in altricial birds. Condor 67:361-382. Haramis, G. M., W. G. Alliston, and M. E. Richmond. 1983. Dump nesting in the Wood Duck traced by tetracycline. Auk 100:729-730. Herkert, J. R., D. L. Reinking, D. A. Wiedenfield, M. Winter, J. L. Zimmerman, W. E. Jensen, E. J. Finck, R, R. Koford, D. H. Wolfe, S. K. Sherrod, M. A. Jenkins, J. Faaborg, and S. K. Robinson. 2003. Effects of prairie fragmentation on the nest success of breeding birds in the midcontinental United States. Conservation Biology 17:587-594. Iverson, L. R. 1988. Land-use changes in Illinois, USA: the influence of landscape attributes on current and historic land use. Landscape Ecology 2:45-61. Johnson, A. L. 1999. Reproduction in the female. Pages 569-596 in Sturkie’s avian physiology (G. C. Whittow, Editor). Fifth Edition. Academic Press, New York, USA. Lyon, B. E. 1993. Conspecific brood parasitism as a flexible female reproductive tactic in American Coots. Animal Behaviour 46:91 1-928. Peer, B. D. and S. G. Sealy. 2000. Conspecific brood parasitism and egg rejection in Great-tailed Grackles. Journal of Avian Biology 31:271-277. Peer, B. D., S. I. Rothstein, M. J. Kuehn, and R. C. Fleischer. 2005. Host defenses against cowbird (Molothrus spp.) parasitism: implications for cow'bird management. Ornithological Monographs 57:84—97. Rohwer, F. C. and S. Freeman. 1989. The distribution of conspecific nest parasitism in birds. Canadian Journal of Zoology 67:239-253. Rothstein, S. 1. 1993. An experimental test of the Hamilton-Orians hypothesis for the origin of avian brood parasitism. Condor 95:1000-1005. Sorenson, M. D. 1991. The functional-significance of parasitic egg-laying and typical nesting in Redhead ducks: an analysis of individual behavior. Animal Behaviour 42:771-796. Temple, S. A. 2002. Dickcissel (Spiza amcricana). The birds of North America. Number 703. Yom-Tov, Y. 1980. Intraspecific nest parasitism in birds. Biological Reviews 55:93-108. Yom-Tov, Y. 2001. An updated list and some comments on the occLiiTence of intraspccific nest parasitism in birds. Ibis 143:133-143. 188 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 1. March 2010 The Wilson Journal of Ornithology 122( 1 ): 1 88-1 89. 2010 First Report of Olrog’s Gull Depredation by Sympatric Kelp Gulls Luciano La Sala' " and Sergio Martorelli' ABSTRACT. — We observed adult Kelp Gulls {Larns clominicanus) capture and eat Olrog’s Gull (L. atlanti- CHs) chicks in the Bahia Blanca estuary, Argentina. This estuary holds the largest breeding colony of Olrog's Gulls. There are no previously published reports of Kelp Gulls capturing and eating Olrog’s Gull chicks. Our data support suggestions made by other authors about the possible existence of conflicts in colonies where both species breed in close proximity. Received 13 March 2009. Accepted 7 October 2009. Olrog’s Gull {Lams atkmticus) (Olrog 1958) is endemic to the coastal wetlands of the Atlantic Coast in Argentina, Uruguay, and southern Brazil (Yorio et al. 2005), and is listed as a Vulnerable species (Birdlife International 2008). Factors that led to this classification include the limited geographical range of the species, small estimated population size (4,000-5,000 pairs), its special- ized diet, and high susceptibility to anthropogenic changes in the environment. Few Olrog’s Gull breeding sites have been identified, and all are in Argentina (Yorio et al. 2005). Over 80% of the total breeding population is concentrated in the Bahia Blanca estuary with the largest colony (—3,600 pairs) on Isla del Puerto (38° 48' S, 62° 15' W) surrounded by a colony of Kelp Gulls (L. dominicaniis) (Delhey et al. 2001). The estimated size of the later was —5,000 pairs in 2003 but appears to have grown considerably since then (P. F. Petracci, pers. comm.). There is only one previous report of depreda- tion attempts by Kelp Gulls on Olrog’s Gull eggs and chicks (Petracci et al. 2004), and no reports of successful depredation were found during a literature search. This is the first report of Olrog’s Gull chicks as prey of adult Kelp Gulls. ' Centro cic E.stuciios Parasitologicos y Vectores (CON- ICET-UNLP). Calle 2 N 384 CEPAVE, 1900 La Plata. Buenos Aires, Argentina. ’Corresponding author; e-mail: lucianolasala@yahoo. com.ar OBSERVATIONS We monitored depredation of Olrog’s Gull eggs and chicks in the Isla del Puerto breeding colony in 2005, 2006, and 2007. We studied the colony in 2005 during part of the egg-laying phase (29 Sep- 7 Oct) and part of the hatching and chick-rearing periods (27 Oct- 15 Nov) and, in 2006 and 2007, during the early hatching and chick-rearing periods (17 Oct-9 Nov and 1 Oct-1 Nov, respectively). We spent —8 hrs per day in the colony between 0800 and 1200 hrs ST, and between 1500 and 1900 hrs. We did not observe direct egg depredation in any year but did observe depredation of Olrog’s Gull chicks in 2006 and 2007. We observed two depredation events in 2006, one on 20 October (1015 hrs) and the other on 1 November (1945 hrs). We observed six depredation events in 2007; 21 October (1800 hrs), 22 October (1905 hrs), 27 October (1715 hrs), 28 October (1045 hrs and 1910 hrs), and 30 October (2018 hrs). Adult Kelp Gulls usually made fast, low-flying passes over the Olrog’s Gull colony and used their beaks to snatch chicks from or around their nests. We observed one Kelp Gull drop the Olrog’s Gull chick while hovering above the colony, causing the death of the chick shortly after its impact with the ground. We calculated all depredated chicks were —10 days of age based on body size. We strongly suspect the same Kelp Gull was respon- sible for four of the six events observed in 2007. This Kelp Gull returned to the same nest —4 m from the Olrog’s Gull colony while eating the chicks. Adult Olrog’s Gulls did not display defensive behavior during the attacks, nor while Kelp Gulls fed on chicks a short distance from the Olrog’s Gull colony. DISCUSSION Kelp Gull populations are expanding through- out most of their geographical range (Yorio et al. 2005), and their generalistic/opportunistic feeding habits are thought to have an important role in this expansion (Giccardi el al. 1997). The negative effects of expanding Kelp Gull populations on SHORT COMMUNICATIONS 189 other seabird species has been associated with increased artificial food sources (McGehee and Eitniear 2007). One study conducted at six colonies, where both Kelp and Olrog’s gulls breed, indicated these species share many micro- habitat characteristics, and suggested the presence of spatial conflicts between them (Garcia Borbor- oglu and Yorio 2007a, b). Isla del Puerto is within an international deepwater port and 3.2 km from the city of Bahia Blanca (-450,000 inhabitants). Discarded fish from a small-scale fishing fleet and presence of open-sky refuse tips, slaughter houses, and pork and poultry production systems near these colo- nies provide readily available and abundant food sources throughout the year, and might sustain growth of this colony (Petracci et al. 2004). This could lead to increased predatory pressure by Kelp Gulls on Olrog’s Gulls. Individual specialization as predators of seabird chicks and eggs has been reported for a number of marine birds including Kelp Gulls (Yorio and Quintana 1997, Quintana and Yorio 1998), other gull species (Harris 1965, Hatch 1970, Parsons 1971, Fuchs 1977, Hulsman 1977), and Great Skuas {Stercorahiis skua) (Votier et al. 2004, 2007). In our study, it is likely the series of consecutive predatory attacks observed in 2007 was by a single specialist Kelp Gull. Our observation of eight chicks depredated over a 3-year period would not suggest substantial pressure on this Olrog’s Gull colony. However, we provide sufficient evidence to suggest further studies that accurately quantify the impact of depredation and other possible conflicts between these species. ACKNOWLEDGMENTS We thank Club Nautico Bahia Blanca, all members of the Fernandez family and park rangers Martin Sotelo and Lucrecia Diaz for their support. We also thank Joaquin Cereguetti and Nicolas Acosta for field assistance. This study was partly funded by Agenda Nacional de Promocion CientiTica y Tecnologica (PICT 34412/05). LITERATURE CITED BirdLife International. 2008. Species fact sheet: Lants atlanticus. http://www.birdlife.org (accessed 20 May 2008). Delhey J. K.. V. P. F. Petracci, and C. M. Grassini. 2001. Hallazgo de una nueva colonia de Gaviota de Olrog (Lams atlanticus) en la ria de Bahia Blanca. Argentina. Hornero 16:39-42. Fuchs, E, 1977. Predation and anti-predator behaviour in a mixed eolony ol terns Sterna sp. and Blaek-headed Gulls Lams mdihunchis with speeial rclerenee to the Sandwieh Tern Sterna sandvieencis. Ornis Seandina- vica 8: 17-32. GarcIa Borboroglu, P. and P. Yorio, 20()7a. Breeding habitat requirements and selection by Olrog’s Gull (Lams atlanticus), a threatened species. Auk 124: 1201-1212. GarcIa-Borboroglu, P. and P. Yorio. 2007b. Compar- ative habitat use by syntopic Kelp Gulls (Lams donunicanus) and Olrog’s Gulls (L. atlanticus) in coastal Patagonia. Emu 107:321-326. Giccardi, M., P. Yorio, and M. E. Lizurme. 1997. Patrones estacionales de abundancia de la Gaviota Cocinera (Lams donunicanus) en un basural patago- nico y si relacion con el manejo de residues urbanos y pesqueros. Ornitologia Neotropical 8:77-84. Harris, M. P. 1965. The food of some Lams gulls. Ibis 107:43-53. Hatch, J. J. 1970. Predation and piracy by gulls at a ternery in Maine. Auk 87:244-254. Hulsman, K. 1977. Breeding success and mortality of terns at One Tree Island, Great Barrier Reef. Emu 77:49-60. McGehee, S. M. and J. C, Eitniear. 2007. Kleptoparasit- ism of Magellanic flightless steamer-ducks (Tachyeres pteneres) by Kelp Gulls (Lams donunicanus). Boletin SAO 14:143-146. Olrog, C. C. 1958. Notas ornitologicas sobre la coleccion del Instituto Muguel Lillo, Tucuman. Acta Zoologica Lilloana 15:5-18. Parsons, J. 1971. Cannibalism in Herring Gulls. British Birds 64:528-537. Petracci, P., L. La Sala, G. Aguerre, C. Perez, N. Acosta, M. Sotelo, and C. Pamparana. 2004. Dieta de la gaviota cocinera (Larus donunicanus) durante el periodo reproductivo en el estuario de Bahia Blanca. Buenos Aires, Argentina. Hornero 19:23-28. Quintana, F. and P. Yorio, 1998. Kelp Gull Lams donunicanus predation on an Imperial Cormorant Phcdacrocora-x atriceps colony in Patagonia. Marine Ornithology 26:84-85. Votier, S. C., S. Bearhop, N. R.atcliffe, and R. W. Furness. 2004. Reproductive consequences for Great Skuas specializing as seabird predators. Condor 106:275-287. Votier, S. C.. S. Bearhop. J. E. Crane, J. M. Arcos, and R. W. Furness. 2007. Seabird predation by Great Skuas Stercorarius skua: intra-specific competition for food? Journal of Avian Biology 38:234-246. Yorio. P. and F. Quintana. 1997. Predation by Kelp Gulls Larus donunicanus at a mixed-speeies colony of Royal and Cayanne terns Sterna ma.xima and S. euryf’natha in Patagonia. Ibis 139:536-541, Yorio, P., M. Bertellotti. and P. Garcia Borboroglu. 2005. Estado poblacional y de conscrvacion de gaviotas quo se reproducen en el litoral maritimo Argentino. Hornero 20:5.3-74. 190 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. I, March 2010 The Wilson Journal of Ornithology 122(1): 190- 193, 2010 Second and Third Records of Snares Penguins {Eudyptes rohustus) in the Falkland Islands Laurent Demongin, ' Maud Poisbleau,’ “ Georgina Strange, '' and Ian J. Strange^ ABSTRACT. — The Snares Penguin (Eudyptes rohus- tus) breeds only on the Snares Islands, New Zealand, and is vagrant throughout the New Zealand region and southeast Australia. The only previous record outside this area was one in the Falkland Islands in 1988. We report the unusual occurrence of two Snares Penguins in the same colony in the Falkland Islands in 2008, and discuss identification issues. Vagrant penguins demon- strate the incredible dispersal ability of these flightless birds. Received 9 March 2009. Accepted 19 July 2009. Snares Penguins (Eudyptes rohustus) are en- demic to New Zealand, breeding only on the Snares Islands (48° 00' S, 166° 33' E), and straying to surrounding islands, south-east Aus- tralia, and Tasmania (Marchant and Higgins 1990, Williams 1995). They are absent from their breeding areas during winter, when they are probably pelagic and migratory. Snares Penguins are rarely observed away from the Snares Islands and their movements are poorly known (Marchant and Higgins 1990). The occurrence of at least five Snares Penguins in the Chatham Islands in January and February 2003 was considered exceptional (Miskelly and Bell 2004, Miskelly et al. 2006). The only previous record of this species beyond Australasia was an adult in the Settlement Rookery on New Island (51° 43' S, 61° 17' W) in the Falkland Islands on 10 December 1988 (Lamey 1990, Martinez 1992). We discuss the unusual occurrence of two Snares Penguins in the same colony in the Falkland Islands 20 years later. ' Max Planck Institute for Ornithology, Vogelwarte Radolfzell, Schlossallee 2, 7831.‘i Radolfzell, Germany. ^University of Antwerp, Campus Drie Eiken, Depart- ment Biology-Ethology, Universiteitsplein 1 , 2610 Antwerp (Wilrijk), Belgium. ’New Island Conservation Trust, New Island, FIQQ IZZ, Falkland Islands. ■“Corresponding author; e-mail: laurentdemongin® gmail.corn OBSERVATIONS We visited daily the Settlement Rookery on New Island during the entire breeding season during the course of scientific studies of Western Rockhopper Penguins (Eudyptes chrysocome) (Poisbleau et al. 2008). An adult Snares Penguin was found on 30 November and 1 December 2008 (Fig. 1). The rookery was visited every day until the middle of January 2009, but the bird was not seen again. However, a second bird was observed on 24 and 25 December 2008 (Fig. 2). We captured both birds to take standard measure- ments (Table 1). The appearance of the birds was sufficiently different to be certain they were not the same individual; the second bird was larger, the black color of its head was duller, it had a partially gray chin, its crests were very short, the pink color of its gape and of the skin around its beak was less obvious, and its eyes were brownish (not red) even under strong light. DISCUSSION The identification of extra-Iimital Snares Pen- guins requires care. The following characteristics allow separation from Western Rockhopper Pen- guins (Shirihai 2007): a wider yellow stripe above the eye starting close to the beak, an obvious pink gape and skin along the edge of the lower mandible, a larger and more bulbous beak, the absence of a black crest on the rear crown, a shorter yellow crest less pendulous behind the eye, a bold black tip on the underside of the nipper, and an overall larger size. The large beak and overall size, and the pink gape and skin along the beak could also resemble a hybrid between a Rockhopper Penguin and a Macaroni Penguin (Eudyptes ehrysolophus). Several instances of such hybridization have been documented in the Falkland Islands (White and Clausen 2002). However, the shape of the yellow stripe and crest of the birds that we found were typical of Snares Penguin and completely different to those of Rockhopper or Macaroni penguins or postulated hybrids between them. SHORT COMMUNICATIONS 191 FIG. 1. Adult Snares Penguin (right) next to a Western Rockhopper Penguin on New Island, Falkland Islands, 30 November 2008. Photograph by Laurent Demongin. A Snares Penguin could also be mistaken for an Erect-crested Penguin (Eiidyptes sclatevi), a spe- cies that breeds on the Antipodes and Bounty islands southeast of New Zealand (Marchant and Higgins 1990) and that occasionally reaches the Falkland Islands (Strange 1982, Summers 2005), including one that bred with Rockhopper Pen- guins on Pebble Island ( 120 km northeast of New Island) between January 1997 and January 2008 (Morrison et al. 2005, Arnold 2008). Erect-crested Penguins have crests that are erectile and nearly parallel when seen from the front (Marchant and Higgins 1990) while Snares Penguins have crests that droop behind the eye, and forms a V when seen from the front. Erect-crested Penguins also have a different head shape with a domed crown, and larger chin reaching closer to the beak tip (Shirihai 2007). Cases of hybridization between Erect-crested and Rockhopper penguins have been reported at the Falkland Islands (Napier 1968) but the descendants were not described. An apparent Erect-crested x Rockhopper hybrid seen next to the Erect-crested Penguin on Pebble Island in January 2006 looked similar to the birds that we saw, but differed in having the extended chin profile of an Erect-crested Penguin, and black feathers in the crest (a Rockhopper characteristic). Other details of the superciliary stripe and bare skin at the gape also matched Western Rockhop- per or Erect-crested penguins rather than Snares Penguins (C. M. Miskelly, pers. comm.). The first bird we found, based on measurements and plumage patterns, was probably an adult female whereas the second was probably an immature male. The beak of the second bird was much longer and was not as deep. The beak shapes were quite different, the first one being bulbous with parallel edges while the second was less robust and more slender. The latter beak shape is typical of imma- tures; the more massive beak of adults is achieved during the third or later years (Stonehouse 1971). The weight of crested penguins varies season- ally with the heaviest birds being recorded before the molt (Warham 1975). The weight of the 192 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 1. March 2010 FIG. 2. Immature Snare.s Penguin (right) among Western Rockhopper Penguins on New Island, Falkland Islands, 24 December 2008. Photograph by Laurent Demongin. second bird we found exceeded the range of breeding Snares Penguins (Stonehouse 1971) but fit well with pre-molting immature males (War- ham 1974a). Most vagrant crested penguins are sub-adults recorded before or during their molt (Woehler 1992, Miskelly and Bell 2004), like the bird we observed on 24-25 December. The two other Falkland Islands records of Snares Penguin (including our first bird) were adults during the breeding season. However, observers rarely visit the rookeries after the end of the breeding season, TABLE 1 . Mea.surements (following Warham 1975) of the two Snares Penguins found in the Falkland Islands in 2008, and comparison with x, (n), and (range] data from (a) Stonehouse (1971) and (b) Warham (1974b). Characicristic Bird 1 Adult male Adult female Bird 2 (a) (b) (a) (b) Flipper, mm 181 191 183 (61) 184.0 (1 14) 177.3 (47) 178.8 (82) [170-193] (167-187] Beak length, mm 54.2 57,1 59.2 (68) 59.1 (35) 52.5 (58) 52.0 (12) (54-691 (49-61] ' Beak depth, mm 23.3 22.6 28.2 (35) 24.2 (12) Head and beak length. 122.6 127.9 mm Mass, g 3.300 4,360 3,320 (41) 2,780 (32) ] 2,450-4,300] [2,300-3,400] SHORT COMMUNICATIONS 193 and vagrant birds coming ashore to molt could easily be missed. The Chatham Islands are only ~ 1 ,400 km from the Snares Islands, and yet the presence of several Snares Penguins in 2003 was considered anoma- lous, possibly reflecting a changed at-sea distri- bution (Miskelly and Bell 2004). It is difficult to assess the reasons for the presence of this species in the Falkland Islands, —8,000 km from the Snares Islands. However, the occurrence of penguins endemic to the New Zealand region in the Falkland Islands demonstrates their incredible dispersal ability. The opposite is also true with Western Rockhopper Penguins reported from the Snares Islands south of New Zealand (Tennyson and Miskelly 1989, Miskelly et al. 2001). ACKNOWLEDGMENTS We are grateful to the New Island Conservation Trust for permission to work on the island and for providing accommodations. We thank Maria Strange, Dan Birch, Rafael Matias, and Petra Quillfeldt for their support and advice during the field season. The manuscript benefited greatly from critical comments by Maria Strange. C. M. Miskelly, C. E. Braun, and an anonymous referee. LITERATURE CITED Arnold, N. 2008. Falkland Islands. 10 to 23 January 2008. Trip report. The Travelling Naturalist, Limosa Holi- days. www.naturalist.co.uk/reports2008/falklands08. pdf (accessed 13 July 2009). Lamey, T. C. 1990. Snares Crested Penguin in the Falkland Islands. Notomis 37:78. Marchant, S. and P. J. Higgins. 1990. Handbook of Australian. New Zealand and Antarctic birds. Volume 1 . Ratites to ducks. Part A. Ratites to petrels. Oxford University Press, Melbourne, Australia. Martinez, I. 1992. Family Sphenicidae (Penguins). Pages 140-160 in Handbook of the birds of the world. Volume 1. Ostrich to ducks (J. del Hoyo, A. Elliott, and J. Sargatal, Editors). Lynx Edicions, Barcelona. Spain. Miskelly, C. M. and M. Bell. 2004. An unusual influx of Snares Crested Penguins (Eiidyptes robnstus) on the Chatham Islands, with a review of other crested penguin records from the islands. Notomis 5 1 :235-237. Miskelly, C. M., A. J. Bester. and M. Bell. 2006. Additions to the Chatham Islands’ bird list, with further records of vagrant and colonising bird species. Notomis .63:213-228. Miskelly, C. M., P. M. Sagar, A. J. D. Tenny.son, and R. Scofield. 2001. Birds of the Snares Islands, New Zealand. Notomis 48:1-40. MoRRtsoN, M., A. Henry, and R. Woods. 2005. Rare and vagrant birds in the Falkland Islands 2005. Falk- lands Conservation, London, United Kingdom, www. falklandsconservation.com/wildlife/birds/Fl-RareBirds- 05.pdf (acces.sed 13 July 2009). Napier, R. B. 1968. Erect-crested and Rockhopper penguins interbreeding in the Falkland Islands. British Antarctic Survey Bulletin 16:71-72. PoisBLEAU, M., L. Demongin, I. J. Strange, H. Otley, AND P. Quillfeldt. 2008. Aspects of the breeding biology of the Southern Rockhopper Penguin Endyptes c. chrvsocome and new consideration on the intrinsic capacity of the A-egg. Polar Biology 31:925-932. Shirihai, H. 2007. A complete guide to Antarctic wildlife. The birds and marine mammals of the Antarctic continent and the Southern Ocean. Second Edition. A&C Black Publishers, London, United Kingdom. Stonehouse, B. 1971. The Snares Islands Penguin Endyptes robnstus. Ibis 113:1-7. Strange, I. J. 1982. Breeding ecology of the Rockhopper Penguin (Endyptes crestatus) in the Falkland Islands. Le Gerfaut 72:137-188. Summers, D. 2005. A visitor’s guide to Falkland Islands. Second Edition. Falklands Conservation, London. United Kingdom. Tennyson, A. J. D. and C. M. Miskelly. 1989. “Dark- faced” Rockhopper Penguins at the Snares Islands. Notomis 36:183-189. Warham, j. 1974a. The breeding biology and behaviour of the Snares Crested Penguin. Journal of the Royal Society of New Zealand 4:63-108. Warham, J. 1974b. The Fiordland Crested Penguin Endyptes pachyrhynclins. Ibis 1 16:1-27. Warham, J. 1975. The crested penguins. Pages 189-269 in The biology of penguins (B. Stonehouse, Editor). Macmillan Press, London, United Kingdom. White, R. W. and A. P. Clausen. 2002. Rockhopper Endyptes chrysocome chrysocome x Macaroni E. chrysolopbns penguin hybrids apparently breeding in the Falkland Islands. Marine Ornithology 30:40^2. Williams, T. D. 1995. The penguins. Oxford University Press, Oxford, United Kingdom. WOEHLER. E. J. 1992. Records of vagrant penguins from Tasmania. Marine Ornithology 20:61-73. 194 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122. No. 1, March 2010 The Wilson Journal of Ornithology 122( 1 ): 194-195, 2010 A New Bird Species for Costa Rica; Sapphire-throated Hummingbird (Lepidopyga coerideogidahs) Esteban Biamonte* ABSTRACT. — I present the first confirmed record of the Sapphire-throated Hummingbird (Lepidopyga coer- uleogularis) at Punta Banco, southwestern Costa Rica. An adult male was seen and photographed perched on a twig at the border of a second growth forest tract. This observation extends the geographic distribution of this .species from northern Colombia to southwestern Costa Rica. Received 28 Fehriiarv 2009. Accepted 27 August 2009. The Sapphire-throated Hummingbird {Lepi- dopyga coenileogidaris) is known to occur from western Panama to northern Colombia, along most of the Pacific slope and over the eastern region of the Caribbean slope of Panama into Colombia. This hummingbird is more abundant in coastal forests (Ridgely and Gwynne 1989), occupying primarily secondary forests and scrubby clearings, and in lower frequency, around mangrove patches. It is usually observed foraging for both nectar and insects at low strata (Ridgely and Gwynne 1989, Stiles et al. 1989). This species is known in Costa Rica from an old and possibly mislabelled specimen and from an unconfirmed record from 1962 (Stiles et al. 1989). Since then, this species has not been reported for Costa Rica (Obando-Calderon et al. 2007). OBSERVATIONS An adult male Sapphire-Throated Humming- bird was observed foraging and photographed on 2 April 2008 at Tiskita Jungle Lodge landing strip at Punta Banco, Costa Rica (8° 21' N, 83° 06' W) (Fig. 1). This area is —10 km from the Panamanian border. Possibly the same male was observed the previous day, foraging at the same spot. The hummingbird was feeding on flowers ' E.scucla (Je Biologi'a, Univensidad de Costa Rica. San Pedro de Montes dc Oca. Costa Rica; e-mail: cstebanbiamontc@ yahcKi.com of a vine of the Rubiaceae family growing near the lodge’s airstrip. The area is an open field surrounded by patches of secondary forest and early second growth vegetation, numerous orna- mental plants in the lodge’s gardens, and an extensive living fence of flowering Ihiscus rosa- sinensis, which attracts several other species of hummingbirds. On 23 December 2008 a new report of this species was made by Kevin Easley even further into Costa Rica, —35 km from the Panamanian border at la Gamba near Golfito. This region is comprised of extensive pastures with living fences and isolated trees suiTounded by mature protected forest on the slopes that surround the valley. The species was seen visiting flowers of garden plants and living fences. DISCUSSION This is one of several bird species, typical of open areas that have been recently reported in Costa Rica, possibly coming from Panama (Stiles et al. 1989; Sanchez et al. 1998; EB, unpubl. data). The range expansion of these species may be caused by the increasing destruction of the natural mature forests in the region during the last decades (Sanchez-Azofeifa et al. 2001, 2003). Mo,st of the original forest has been eliminated in the lowlands of Central America and replaced by pastures or agricultural fields, allowing some birds to expand their original range. The Sap- phire-throated Hummingbird is expected to soon become more common along the Costa Rican Pacific Coast, given its use of open areas and behavior. ACKNOWLEDGMENTS I thank .Saul Bocian, for being in the right place at the right time to take the great picture.s and for sharing them, everyone at Tiskita .lungle Lodge, Onik Morrison, and Jay Morrison for revising the English, and Luis Sandoval and Gilbert Barrantes for checking the manuscript. SHORT COMMUNICATIONS 195 FIG. 1. Male Sapphire-throated Hummingbird in Costa Rica’s Pacific Coast (photograph by Saul Bocian). LITERATURE CITED Obando-Calderon, G., L. Sandoval, J. Chaves-Cam- pos, J. Villareal-Orias, and W. Alfaro-Cer- vantes. 2007. Lista Oficial de aves de Costa Rica. Zeledonia. Numero Especial 1 1:1-76. Ridgely, R. S. and j. a. Gwynne. 1989. A guide to the birds of Panama with Costa Rica. Nicaragua and Honduras. Princeton University Press. Princeton. New Jersey. USA. Sanchez-Azofeifa, G. a., R. C. Harrlss, and D. L. Skole. 2001. Deforestation in Costa Rica: a quanti- tative analysis using remote sensing imagery. Biotrop- ica 33:378-384. Sanchez-Azofeifa, G. A., G. C. Daily, A. S. P. Pfaff, AND C. Busch. 2003. Integrity and Isolation of Costa Rica's national parks and biological reserves: exam- ining the dynamics of land-cover change. Biological Conservation 109:123-135. Sanchez. J. E., K. Naoki. and J. Zook. 1998. New information about Costa Rican birds. Ornitologia Neotropical 9:99-102. Stiles. G. F., A. F. Skutch. and D. Gardner. 1989. A guide to the birds of Costa Rica. Cornell University Press. Ithaca. New York. USA. The Wilson Journal of Ornithology 122( 1): 196-205, 2010 Ornithological Literature Robert B. Payne, Review Editor THE CURSE OF THE LABRADOR DUCK. MY OBSESSIVE QUEST TO THE EDGE OF EXTINCTION. By Glen Chilton. Simon & Schuster, New York, USA. 2009: x + 305 pages, 10 black and white figures, 1 map. ISBN: 978-1- 4391-0247-3; 978-1-4391-2499-4 (ebook). $25.00 (hardcover). — For those whose interest in birds extends beyond another tick on their life list, extinct birds often provide a source of intense fascination. Each species has its own allure, but one that is also coupled with an air of unattain- ability, like a photograph of a beautiful woman on the obituary page. The Labrador Duck {Campto- rhynciius lahradorius), the first endemic North American bird to disappear historically, about 1875, remains one of the most perversely enigmatic of extinct birds. Although known from a fair number of specimens (at least 55 still exist), few have reliable data and virtually nothing is known about the breeding of the species or its summer distribution. It has escaped the attention of monographers such as have labored over the Great Auk (Pingiiinus impennis). Passenger Pi- geon (Ectopistes migratoriiis), and Carolina Par- akeet (Coniiropsis carolinensis). Enter Glen Chilton, who debuted in ornithology on the bandwagon of studies of song dialects of the White-crowned Sparrow, which led to his being an author on the account of Zonotrichia leiicophrys in the Birds of North America series. Lured by the prospect of a gratis set of that costly work, he chose to write another species account and what could possibly have been easier than the Labrador Duck, about which so little was known? With the account of the Labrador Duck under his belt, Chilton became its de facto “authority,” thus setting the stage for his “obsessive” quest to examine every specimen of the species, perhaps not unlike the person who patronized every Starbucks coffee shop in the world to set a record. This book is about Chilton’s visits to museums scattered through Europe and North America to view specimens of Labrador Ducks. The opposite of a work of .scholarship, it is very much a travelogue, and a tedious one at that. From the author’s web site we learn that Chilton wants people to believe that he is a bon vivant engaged in “outrageous adventures,” which in this case means going to a distant city, measuring a “stuffed duck,” and then wandering about town in search of alcohol and vegetarian food (to me, the thought of eating tofu in Paris and Vienna is more depressing than the fate of the Labrador Duck). The title, of course, is a gimmick, the only curse being the book itself. Nor does the subtitle make much sense considering that the Labrador Duck is long past the edge of extinction. Chilton’s background and attitude did little to prepare him for museum work, as he plainly knows next to nothing about collecting or preparing birds. The specimens he looked at met their ends by being “blasted,” “blown to kingdom come,” or shot full of “bullets,” which, if true, would have left little for him to examine. Although he includes brief instructions on prep- aration techniques taken from a museum pam- phlet, he seems not to understand the process and refers more than once to the incision through which the “guts were pulled out.” Specimens are invariably “stuffed,” a word generally avoided by museum ornithologists but apparently deliber- ately overused by Chilton to convey his poorly disguised disdain for museums as places full of “stuffed” dead things. Not being knowledgeable of the history of museum ornithology, Chilton is bound to have overlooked facts that would be evident to someone with more experience. What else could one expect from one for whom the name Verreaux carried no significance? We also read (page 304) that: “In November of 1844, Colonel Nicolas Pike shot a drake Labrador Duck at the mouth of the Ipswich River at the south end of Plum Island, New York. History doesn’t say whether or not the duck was doing anything to provoke the colonel. Perhaps Pike just really, really hated ducks”. Not only is this gratuitously stupid, but Chilton shows no evidence of recog- nizing the identity of Nicolas Pike nor of Pike’s connection with other extinct birds [Subtropical Rambles in the Land of the Aphanapteryx, 1873). Chilton provides a brief description, usually only a paragraph, mostly detailing defects, of the various specimens of Labrador Duck. In a few instances he appears to have unearthed some new 196 ORNITHOLOGICAL LG ERATLJRi-: 197 information on the history of a given specimen, which is commendable. But most of the book is simply an ego-indulging procession of irrelevan- cies. Why do we need to know that he was a breech birth, a nervous and obsessive child, and that on at least four occasions in his duck travels he shared a room, or even a bed with various young women who were not his wife'.^ Chiiton seems to regard himself as quite a humorist, but his humor is so tiresomely puerile as tei make George Gobel and Dave Barry seem like genuine wits. Pick any pejorative adjective that describes annoying prose and it will apply somewhere in this book; arrogant, facetious, facile, flippant, glib, precious, sopho- moric, and snotty. Throughout, Chilton manages to be condescending or even downright insulting to individuals, institutions, and entire countries. I daresay that following the appearance of this book he will be unwelcome in many places where he was once treated with courtesy and respect. Errors large and small throughout the book call into question the accuracy of almost anything Chilton writes. For example, the Portuguese word for “cultivator” is lavrador, not llavrador (incidentally, Chilton seems to revel in his ignorance of languages). Some older bird speci- mens were preserved with arsenic, not cyanide. There are no sea lions, large or small, around Grand Manan Island or anywhere else in the North Atlantic. The national mall in Washington does not run just from the Capitol to the Washington Monument, which latter he has apparently confused with the Lincoln Memorial. It seems intended that the book not be consulted again after its first reading, as there is no table of contents, no index, and no bibliography. As partial recompense for inflicting this fulsome blot on the literature of extinct birds, Chilton owes ornithology and the staffs of all the museums he visited an exceptionally scholarly and insightful monograph on the Labrador Duck. Considering that he was ill equipped to undertake such an endeavor in the first place, I doubt that he will ever be able to make satisfactory repayment of his debt. — STORRS L. OLSON, Curator, Division of Birds, National Museum of Natural History, Smithsonian Institution, Washington, D.C. 20560, USA; e-mail: olsons@si.edu BIRD BANDING IN NORTH AMERICA: THE FIRST HUNDRED YEARS. Edited by Jerome A. Jackson, William E. Davis Jr., and John Tautin. Nuttall Ornithological Club, Cam- bridge, Massachusetts, USA, 2008; ix and 280 pages, 62 figures and photographs. ISBN: 1- 877973-45-9. $40.00 (cloth).— This volume brings together contributions presented at a symposium which occurred at the North American Ornitho- logical Conference in New Orleans in 2002. The symposium commemorated 100 years of scientific bird banding in North America, the birth of which is designated by the banding of Black-crowned Night-herons (Nycticorax nycticorax) by the Smith- sonian Institution’s Paul Bartsch in 1902. The 12 chapters trace the trajectory of bird marking from modest but enthusiastic beginnings through the wide variety of scientific pursuits that have used this important ornithological tool. The first three chapters provide historical perspective. Chapter 1 gives the early history of bird banding in North America. It starts with a bit of background on marking birds in the 1 6th century in the Old World, and quickly settles in North America, relating the well-known story of John James Audubon’s 1804 experiment of marking Eastern Phoebes (Sayornis phoehe) with silver wire. This is followed by an interesting overview of the factors that led to the popularity of nature study in general and bird marking in particular in the early 20"’ century - including such disparate elements as increased leisure time due to the availability of electricity and the sudden accessibility of aluminum. The chapter goes on to the more specific underpinnings: the genesis of organized banding, founding of the American Bird Banding Association and the first banding stations, and finally the federal govern- ment’s role in recognizing and controlling band- ing in the United States Chapter 2 discusses the role of banding organizations in education and training of band- ers; the distribution of data, techniques, and materials; and in encouraging regional coopera- tion. The organizations mentioned include nation- al (including the present North American Banding Council) and regional groups, as well as major bird observatories and banding station networks. They are arranged by type with a brief summary of each group's history and function. Chapter 3 focuses on the history of the Bird Banding Laboratory (BBL). The BBL wasn t formally designated as such until 1961, but this chapter begins in 1920, when Frederick Lincoln was charged by the USDA Bureau ot Biological Survey with organizing their “banding office.” 198 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. /, March 2010 He remained in this position until 1946 and is noted as the true founder of the modern bird banding program. The chapter proceeds to tell the history of the BBL by decade, ending in 2002, which coincided with the retirement of the BBL’s sixth Chief and chapter author John Tautin. The rest of the book provides overviews of the role of bird banding in various research areas, focusing on particular instances where banding has made important scientific contributions. These include population monitoring and ecology, rare species recovery, migratory game bird and waterbird con.servation, avian ecotoxicology and disease, habitat use, and behavioral research. The chapters overall make a fine case for the value of bird banding, even though, as several authors point out, the original intent was focused primarily on bird movements. Most chapters conclude with recent developments and future directions in each area, and how banding might fit into these new horizons. The chapters in this compendium vary in form from chronologies and narratives to something more resembling typical scientific style. Every chapter ended with literature cited, a convention I prefer over paging to the end of the entire volume to look up a reference. I wasn’t able to discern whether there was any particular order to the chapters past the first three, which dealt with history. The volume might have flowed a little better if the chapters hacj progressed from broad research themes to more specific studies, or perhaps followed some chronological order. I found only one major editing gaff. In Chapter 5, a concluding paragraph was printed twice, once at the end of a series of species accounts, and again three paragraphs later at the end of the chapter. There are 62 black-and-white figures in the book, mostly photographs. Many of them are “mug shots,’’ generally of historical figures or researchers whose work is featured in the chapter. I often don’t notice the.se things, but in this volume I could not help being struck that over half of the photos featured white guys, mostly middle- aged. There were a few women, and no people of color. Are bird banding and ornithology really that devoid of ethnic, racial, and gender diversity? Yikes. We have something to work toward in the next hundred years. — JULIE A. CRAVES, Rouge River Bird Observatory, University of Michigan-Dearborn, Dearborn, MI 48128, USA; e-mail: )craves@umd. umich.edu INGESTION OE LEAD FROM SPENT AM- MUNITION; IMPLICATIONS FOR WILDLIFE AND HUMANS. By Richard T. Watson, Mark Fuller, Mark Pokras, and W. Grainger Hunt. The Peregrine Fund, Boise, Idaho, USA. 2009: 383 pages. ISBN: 0-9619839-5-7. $ None shown, (paper). — As a rule, conference proceedings are uninspiring tomes. The ugly stepchildren of the academic publication family; their contents are often limited to progress reports of ongoing research, literature reviews, and the results of short-term Master’s Degree research projects. There are some conferences, however, for which a published proceedings is not only appropriate, but also provides an exceptionally useful resource for a wide audience. The proceedings of the conference Ingestion of Lead from Spent Ammu- nition: Implications for Wildlife and Humans, held 12-15 May 2008 at Boise State University is one of these. The book fully addresses the ubiquitous nature of lead toxicity by including voices from science, medicine, policy, and business. Whereas it would be difficult to find one scientific joui'nal that would accept this breadth of viewpoints, the proceedings format allows for the cross-disciplin- ai'y compilation of knowledge necessary to put the lead issue in context. Everyone knows something about the toxic effects of lead but few know everything about it. For example, most Americans know that in the 1970s, awai'eness of the effects of lead on human health I'esulted in its removal fr'om house paint and gasoline. Hunter's know that in 1991, documented adverse effects of lead ingestion on eagles and water-fowl species resulted in r'egula- tions outlawing lead shot for waterfowl hunting. Ingestion of Lead from Spent Ammunition pr'o- vides infor'ination of which most people ar'e not aware, such as: how lead is toxic (in a fascinating paper by Pokras and Kneeland); effects of low- dose exposure of lead on humans who consume hunted game; and the magnitude of the pr'oblem in ter'rns of the amount of lead released into the environment by outdoor spor'ts (6-10 thousand rnetr'ic tons annually in the U.S.) or the number of species known to have incidences of high lead levels (over 130 species of ver'tebr'ates). The unique strength of this book is it’s “con.servation medicine’’ appr'oach which Pokras and Kneeland (page 7) define as “examin[ing] the linkages among the health of people, animals, and the envir'onment’’. Imagine the value to wildlife r'ehabilitation veter'inar'ians when they tap into the ORNITHOLOGICAL LITERATURE 199 rich field of lead exposure in human development, and vice versa; or the public health official who becomes familiar with the work on retention ol lead fragments in hunted wildlife. Proof of this book’s integrated approach to lead is the diversity of author affiliations: wildlife research centers, zoos, environmental agencies and non-protits, human medicine, and state conservation offices, to name a few. The book is composed of 53 papers (25 peer- reviewed, 10 not peer-reviewed, and 18 extended abstracts) organized in four main sections: review of lead uptake and toxicosis in humans and wildlife; lead exposure in humans from spent ammunition; lead exposure, sources, and toxicosis in wildlife; and, remediation of lead exposure from spent ammunition. There are also nine expert commentaries transcribed from the meeting and an introduction and conference summary by the esteemed raptor biologist Ian Newton. Birds, especially raptors, are the predominant taxonomic group in the research. Not surprisingly, work on the California Condor {Gymnogyps californianus) is well represented. Other bird species addressed include Golden Eagles {Aquila chrysaetos) and fish eagles (Haliaeetus spp.), waterfowl, doves, and other terrestrial birds. Geographically, most papers concern the Unit- ed States, but there is good representation from Europe. In the one paper from South America, Saggese et al. review lead toxicosis in raptors from Argentina. In discussing sources of lead in the environment, they cite the recent and growing popularity of dove and pigeon hunting in central Argentina where, absent any limits on the number of birds killed, hunters regularly discharge over 1,000 cartridges per day, killing or injuring as many birds. The barbarism of these hunts is, therefore, compounded by the 1,600 metric tons of lead released into the local environment each year and the fact that crippled and dead birds are left for scavengers (including humans) to con- sume, lead shot and all. There are only two minor flaws with the book. The first is the repetition of basic information in many of the papers. One would only notice this if the book was read cover-to-cover, a method tor which the book was not designed. The second flaw concerns the 18 entries that are extended abstracts. Some of these are quite extensive with figures and tables, etc. while others aren’t ‘extended’ at all. Either way. no one likes to cite abstracts and when I searched for full papers that these abstracts should have evolved into by now I wasn’t able to find any. In his closing comments for the meeting, Ian Newton noted the conference showed that lead from spent ammunition poses a bigger human health problem than previously recognized. In- deed, after reading this book, one cannot help but be alarmed at the widespread and insidious nature of lead ingestion by humans and wildlife. These proceedings are a call for action and, unlike many environmental problems that seem insurmount- able; the removal of lead from outdoor sporting equipment is attainable. The book is published by the non-profit organi- zation The Peregrine Eund and the entire proceed- ings are available at http://www.peregrinefund.org/ Lead_conference/. Other than a few color photos and figures online that are represented in grayscale in the print edition, there are no substantial differences between the book and the online version. The reader’s personal preference is the best guide in deciding which version to acquire. Either way, acquire this book and spread the word! I strongly recommend this book to anyone who wants to be more knowledgeable about the threats to wildlife, humans, and the environment posed by the release of lead into our fields and wetlands. — JOHN CURNUTT, Regional Wildlife Ecologist, USDA, Forest Service, Eastern Region, 626 East Wisconsin Avenue, Milwaukee, WI 53203, USA; e-mail; jcurnutt(§>fs.fed.us THE BIRDS OE THE REPUBLIC OE PANA- MA. PART 5. GAZETTEER AND BIBLIOG- RAPHY. By Deborah C. Siegel and Stoms L. Olson. Buteo Books, Shipman, Virginia, USA. 2008: 516 pages plus 1 inset map. ISBN: 978-0- 931130-17-5. $45.00 (hardcover). — Even in these modern days of hand-held GPS receivers and Google Earth, the final volume of Alexander Wetmore’s magnum opus: The Birds of the Republic of Panama is a welcome addition to the library of any serious student of neotropical birds who will want to make room for it along side the previous four volumes. It is outstanding in its primary role, as a 2()th century ornithological gazetteer, but most modern readers will find it wanting a few 21st century trappings that were left out and we can hope will be included in some future format. The authors, Deborah Siegel and Stoms Olson, have done meticulous work in providing geo- 200 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. I, March 2010 graphic precision to ambiguous place names, which are the norm in much of Latin America. Those who have done field work in these parts are tamiliar with the simultaneous tendency for multiple local names for a single place, the application of a general name to a region too large to be considered a single biogeographic point, and the duplication of common names across distant districts and provinces. For exam- ple, this volume lists nine unique locations with the moniker “San Jose” which the authors dutifully sort out. Siegel and Olson have also been mindful of the tendency of place names to change, which is especially problematic for the areas surrounding the Panama Canal; in many cases, a Spanish-language place name replaced earlier English-language names used during the U.S. administration of the Canal Zone. I recently reviewed collecting localities for the niglaris subspecies of Myrmeciza e.xsul mentioned in The Birds of Panama. Volume 3. Not surprisingly I was able to find precise locations and geograph- ical coordinates for all, but 1 should also note that 1 found the same information on Google; the value of this volume is in the annotations for each location. For example, many of the place name descriptions follow a hyperlink format, whereby other place names of interest relative to the location in question are referenced. It was in this way that 1 was able to make sense of the confusion surrounding Cerros Colorado, Santiago and Flores; an important endemic bird area in the Serrania de Tabasca in western Panama. It should also be noted the authors were careful with resolving the many unintentional variants of place names caused by collectors (including me) whose command of Spanish was less than perfect. Equally valuable is the detailed annotated bibliography of Panamanian ornithology. The span of this bibliography is vast. Here, topics as diverse as systematics, paleozoography, and natural history comingle from both recent (up until ~ 2005) and historical sources. I am especially appreciative for the effort the authors undertook to include works not typically found in searchable data ba.ses such as museum mono- graphs and Latin American regional journals published in Spanish. This volume should foster the inclusion of the.se works into the cited literature of future studies. Particularly of interest is the inclusion of the reference for a multitude of named taxa (genus, species or subspecies) with a Panamanian type location; perhaps only the authors know how close to being an exhaustive list this may be. Given this attention to detail, I wonder how much greater an impact this final volume - in what is largely a 20th century work - could have had if only more 21st century informatics were used. As an active collector of birds in Panama, one of my favorite features of the book is a map highlighting every collecting location (as well as the geographic gaps in our collective efforts). However, the map appears to be an afterthought, as it is a separate inset to the book. There is no way to go from a dot on the map to that place name in the text to learn more about who collected there and when. Similarly, there is no way to find all of the place names in any geographic region, such as the Pearl Islands or the Darien Province, despite how useful that would be. I imagine that many of my colleagues from the gene Jockey cohort of ornithologists will appreciate the bibliography’s detailed listing of 19'*’ century taxonomic literature when it comes time to give proper names to resurrected lineages. However, without a .searchable index for scientific names, 1 wonder if this resource will be used to its fullest capability. Providing a digital PDF of the text would allow for searching on strings such as province name or latitude and longitude ranges. Alternatively, perhaps the basics of the place names (and the point map) could be created as an internet-based resource; readers could be directed to the text for the details pertaining to specific locations. To be fair, other gazetteers of Latin American ornithology have these problems, but the fact that this volume was published in 2008 makes their omission more apparent. It is worth noting the book begins with a short bibliography of Dr. Wetmore that includes his portrait and a detailed timeline of his field expeditions in Panama. This is fitting for the final volume of Wetmore’s unparalleled work. Panama has had a much larger role in the development of neotropical ornithology than could be predicted by its geographic size. This is in no small part due to the generations of ornithologists that refer to well-thumbed pages of The Birds of Panama to learn detail after detail about the distribution and natural history of that country’s birds. We are all indebted for his effort. — MATTHEW J. MIL- LER, Postdoctoral Fellow in Molecular Evolu- tion, Smithsonian Tropical Research Institute, Apartado Postal 0843-03092 Panama, Repiib- lica de Panama; e-mail: millerma@si.edu ORNITHOLOGICAL LITERATURE 201 THE MIGRATION OF BIRDS: SEASONS ON THE WING. By Janice M. Hughes. Firefly Books, Buffalo, New York, USA. 2009: 208 pages, 76 color photographs, 27 maps and figures. ISBN: 13-978-1-55407-432-7. $40.00 (hardcover with jacket). — The biological study of migration, specifically among birds, is an extremely diverse field of endeavor. Over the past hundred years the literature is filled with diverse views and studies, each attempting to explain one particular facet of migration. The author of this work attempts to unify the great number of studies contributing to the understanding of migration. The book is divided into six broad chapters, each with its own strengths and shortfalls-as to be expected in any attempt to cover such a complex topic. Ten “Profiles” are presented, two per chapter. These vignettes deal with specific species or groups of species that exhibit similar migration patterns or experience similar migration difficul- ties. The opening chapter, “Bird Migration through Human History,” was perhaps my personal favorite as it covered a great diversity of bird lore starting with mythological references, passing through classical references (such as Aristotle and Pliny the Elder) on to the treatise on falconry by the Holy Roman Emperor Frederick II in the 13th Century. At this point we start to see the origins of a more scientific study of birds that progresses to the present century. The second chapter, “The Five Ws of Avian Migration,” is the fuel for this amble through migration. It is here that the author introduces the reader to the varied aspects of bird migration. Researchers are occasionally and appropriately mentioned but there is a complete lack of reference citations. While this is undoubtedly a style preference for this type of book, citations would have made this a powerful chapter, instead of just an interesting one. The third and fourth chapters, “The Phenom- enon of Flight” and “Fueling the Migration,” provide interesting aspects regarding the anatomy and energetic needs of birds. Again, citations would have been appreciated. “Finding the Way,” the fifth chapter meanders through the various ways that birds are believed to navigate. Here the reader is exposed to some classic experiments and studies in navigation. It is here that the anecdotal story of Cheri Ami should hold the reader on edge heading towards the ultimate chapter of the book. This World War I carrier pigeon (Columha livia) earned “the Croi.x de Guerre, one of the country’s | France] highest honors for distinguished acts ol heroism during combat.” The final chapter, “Migratory Birds in Peril,” is really a one page overview followed by three birds with contemporary problems, a possibly extinct species (Eskimo Curlew [Niinienius bor- ealis]) and “An Indiscriminate Killer,” which discusses the role of lighthouses and radio towers as man-made migratory hazards. A glossary is a welcomed addition to any book of a highly specific nature. The one in this book covers the basic terms with which most non- scientists would need help. While it is quite brief, it is well done and should be of use to the casual reader. Following the glossary is a section entitled “Further Reading.” This two page list of 33 books is inadequate for the topic, especially when considering three of the texts are reprints of Aristotle, Frederick II, and Pliny. The writing throughout is well done and makes for enjoyable reading. The photographs are first- rate. This attractive combination draws the reader interested in migration deeper in, looking for more. But here it stops. The complete lack of references to the great number of studies that must have been consulted for this work will frustrate the reader in efforts to look elsewhere for greater detail. While the author has attempted to cover the topic of migration thoroughly, I found it disturb- ing that Pliny the Elder garnered a citation in the index, whereas world-renowned migratory sites such as Point Pelee (Ontario), Whitefish Point (Michigan), Cape May (New Jersey), Malmo (Sweden), and Istanbul (Turkey), among others were lacking. This book is a cursory overview of a complex topic. The appeal of this book will not be to the ornithological community due to the lack of referencing or detailed accounts of species or migration sites. It should, however, be in public libraries where it will have general audience appeal. Additionally, it will make a fine coffee- table book for the casual bird-watcher. Overall this book is best suited to middle school and high school libraries where young students will find it an interesting read, providing enough detail to stimulate interest. With adequate assistance, a student should be able to wade through the “Further Reading” section and continue on in studying the various aspects of migration. — MICHAEL A. KIELB, Visiting Lecturer, 202 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122. No. I, March 2010 Eastern Michigan University, Department of Biology, 316 Mark Jefferson, Ypsilanti, MI 48197, USA; e-mail; mkielb@emich.edu THE EASTERN SCREECH OWL: LIEE HISTORY, ECOLOGY, AND BEHAVIOR IN THE SUBURBS AND COUNTRYSIDE. By Frederick R. Gehlbach. Texas A&M University Press, College Station, Texas. USA. First Printing (1994). Second Printing (2008): 320 pages, 2 color, 36 black and white photos, 23 figures, and 27 tables. ISBN: (1994): 13-978-I-60344-I21-6). ISBN (2008): 10-1-60344-121-2. $24.95 (paper- back).— I highly recommend reading this study if there is anything at all you want to know or think you might want to know about screech-owls. There is little here that I can think of that might be gleaned from natural history observations that might have been left out. The text contains 10 chapters, 10 appendices, 297 notes, an index, and about 315 references. This is the story of serious field research and intimate contact with individual screech-owls, and with their populations in central Texas. It is a study spanning 41 years, that included 1,453 banded owls. The author involved homeowners, birders, school children, and teach- ers. It is a comparison of suburban versus nearby rural populations that were studied concuirently. The bulk of the study, reported in the first printing of this book (in 1994) was completed between 1976 and 1987. The (2008) printing contains new findings. Some things had changed: eggs are now laid 5 days earlier in the season than reported earlier. Wooded suburban habitat has decreased 26% due to suburban sprawl in the study plots, and nesting pairs have decreased 21% although their nesting success is still high. In Chapter 1, the Introduction, Gehlbach tells us how, as a youngster, he was first introduced to screech-owls. He was riding his bicycle in a city park after a softball game when he saw three IJedglings perched out of reach on a branch above him. He “knocked each bird off’ (page 3) with a softball. Years later the mystery of the owls he had discovered that day remained in his memory as he encountered a pair nesting in a “squirrel box." (page 3). He .soon made detailed ob.serva- tions and ended up putting up nest boxes and monitoring owls in several states (Ohio, New York, Michigan, New Mexico, Arizona, and also in Mexico). His close watch of the central Texas population that he reports started later, in 1967. In the Introduction, he outlines his study approach and methods of his statistical analyses. In Chapter 2, he delineates the climate and weather, and the vegetation of his study sites with correlates of the owls’ choice of nest sites, as derived from nest boxes, as well as seasonal choice of roost sites. Chapters 3 and 4 examine food supplies and predation, and body weight and molt, respective- ly. The next three chapters concern details of nesting from eggs and chicks and their care to roles of males versus females during nesting. All of the data culminate in the last three chapters that concern lifetime reproduction and population structure, and comparisons of rural versus urban populations. The wealth of detail is at times overwhelming, but each chapter has a summaiY page or two of ‘The Essentials’. One can thus skim through the book to areas of individual interest and refer back to the detailed quantifications and statistics in each chapter. The coverage is far too wide and too detailed for me to reiterate in a review, and perhaps it should not be done, because eveiyone will find their own field of interest represented. As in any long-teim study, the goal is to try to find repeating patterns. This is the exciting game of natural histoiy, but there are also always unanticipated SLiiprises that go beyond the original goals. Although the primary stated goal of this study was to compare rural versus suburban populations- and a number of significant differences were docLimented-for me the most fascinating discover- ies were the unintended. To cheiry-pick a few perhaps arbitrary examples, I was most intrigued by documentation of a 9-year population cycle, which coincides with a lunar periodicity, and occuiTed in both rural and suburban owls. It has apparently also been documented for diurnal birds. Other interest- ing nuggets were that life-long monogamy was the rule, but sequential and concurrent polygyny occurred when food was plentiful and nesting density high. The rufous phase declined more than the gray pha.se in hard winters but increased following rainy years. Birds were the primary vertebrate prey of these sit-and-wait predators, but insects and retiles were supplemental .items. Blind snakes were taken into the nest as prey items but some survived in nests to live in a novel commensal association with the owls. Gehlbach’s three major conclusions, (pages 192-193) are: (I) the .screech-owl is “especially well-suited to coexistence with humans”, (2) “modern suburbia is quite munificent toward this ORNITHOLOGICAL LITERATURE 203 species... and |it] can utilize its resources”, and (3) "the amenity-pre-adaptation connections cer- tainly modify some ideas about avian ecology.” Screech-owls are rather small owls and pre- adapted to live in the suburban environment as long as there are cavities and food that will satisfy a catholic diet. That said, the devil is, as almost always, in the details. And this book has plenty of them. As Gehlbach mentions himself (page 198): ”1 have learned much and perhaps even more from Eastern Screech Owls [Megascops but long-term events like population cycles need more time...” He continues to monitor the suburban population and writes (page 200); “Field exper- iments are possible now that I know the rules of the game” so the final phase of the investigation begins. He proposed that in the next few years he would use the owls in another study, in another city around a school curriculum to educate children (some of the owls were so habituated they would land on persons and one could reach under them to take eggs and chicks to examine and measure without disturbing the birds). These owls are an excellent study animal for children and the public to get familiar with wildlife, and through his birds he hopes to continue to promote an urban blend of nature and culture. Gehlbach’ s practical directions for future comparative studies will raise awareness and help promote the provision of habitat and nesting sites for these charismatic birds, which are a barometer of some aspects of the urban environment. — BERND HEINRICH, Department of Biology, University of Vermont, Burlington, VT 05401 , USA; e-mail: bemd.heinrich@uvm.edu BIRDS OF EAST ASIA: CHINA, TAIWAN, KOREA, JAPAN, AND RUSSIA. By Mark Brazil. Princeton University Press, Princeton, New Jersey, USA. 2009: 528 pages, 236 color plates, 2 pages line drawings, 2 maps, 950 distribution maps, family key, 2 appendices, index. ISBN: 978-0-691-13926-5. $39.95 (paper); ISBN: 978-0-691-13925-8. $79.50 (cloth).— When I first visited Japan, Korea, and Hong Kong in 1960 aboard the USS Helena (a heavy cruiser) as a freshly minted ensign in the U.S. Navy, there were no field guides covering any part of Asia. My only guide to the area was Keisuke Kobayashi’s ""Birds of Japan in Natural Colors."' Although helpful because of its excellent illustrations, it was difficult to use as the text was in Japanese. Over the ensuing years, numerous books on the birds of eastern Asia were published, including guides to the birds ol China, Taiwan, Japan, and Korea (but none lor northeastern Russia). All of these books are useful and expand the knowledge of the birds in those areas. However, each has their drawbacks, one ol the main ones being that they treat only the birds already known to have been found in the area covered. The gap that remained was a modern field guide that tied all of these areas together in a broad overview. This new guide fills that void very well. It covers all the birds known to have occun'ed in northeastern Asia roughly ea.st of longitude 116° E, thus all the coastal Chinese provinces north from Fujian (opposite Taiwan) to the Arctic Ocean and east to the Bering Sea, including northeastern Russia, Japan, Korea, Taiwan, and their satellite islands. Further, it is a competently executed field identification guide lavishly illus- trated in the modem fashion with the text for each species facing its plate. The use of “East” Asia in the title is both misleading and inaccurate as the guide is concerned only with “northeastern” Asia. None of southeastern Asia is covered. The brief introduction of 8 pages contains: Aims of this guide. Geographical scope. Taxonomy, Nomen- clature, Bird identification. Bird habitats. Migra- tion, Vagrancy, and How to use this book (which includes 2 pages of line drawings illustrating avian topography and terminology). This is followed by a 16-page key to the families with a color thumbnail portrait of a representative species from each family. The brief text for each family discusses general habits of the family members which will help place an unidentified bird in its family. The bibliography lists only 16 titles. Since a work such as this requires use of hundreds of references, listing only a few seems pointless. However, the entire reference list (and other information about the book) can be found at: htlp://sites. google.com/site/birdsofeastasia. Appen- dix 1 is a useful table giving the status in five broad areas (Chinese coastal and northeastern provinces. Taiwan, Korea, Japan, and northeastern Russia) of each of the species recorded in those areas. Appendix 2 is a list of 46 species considered likely to occur eventually in northeastern Asia. All 969 species known to have occurred in northeastern Asia are illustrated in color paintings on 236 color plates. An additional 16 species that are expected to turn up are illustrated in the plates. 204 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 1. March 2010 as well as eight extinct species. Each plate displays 2-8 species with 1-1 1 images/species. The entire text for each species is opposite its plate and consists of a small (2.2 X 2.1 cm) range map in seven colors, length, wingspan and weight, distribution, habitat and habits, identification, bare parts, and vocalizations. Some have taxonomic notes or alternate English names. On the same line with the title ‘English name’ is a useful code indicating in which of the five geographical areas covered the bird has occurred. By using tiny print, a large amount of information was crammed into the text for each species. However, for some complex groups such as large gulls and diurnal birds of prey, the book’s format is crippling as it does not allow a sufficiently long text. The taxonomy throughout is up to date and forward-looking. The color plates were executed by 14 artists, resulting in uneven quality, most in the good to very good level with some in the excellent to superb range. With only a few exceptions, their goal of enabling identification is attained. Fe- males and immatures are illustrated where they differ from the adult male, as are the more different subspecies, offering mostly complete coverage of plumage variation. All individual images are identified by age, male or female, and often subspecies. Where identification character- istics are revealed only or best in flight, inserts of birds in Hight are given. The artists responsible for each plate are listed only in the copyrights section on the backside of the title page. 1 believe the name of the artist responsible for each plate should be displayed prominently either on the plate itself or on its facing page. Some of the plates 1 found most satisfying are: Alan Harris’s diurnal raptors, Brian Small’s bush-warblers through Acrocephaliis warblers, and Ren Hath- away’s large thrushes. 1 am uncomfortable with the brush strokes in many of Dave Nurney’s birds, as well as the odd postures and proportions and too-large heads of some of his birds. Some comments by plate follow. Plate 22. While the text coirectly notes the crown and nape of the adult Pacific Loon (Diver) (Gavia pacifica) is noticeably paler than the Black-throated Loon (Diver) {G. airtica), the plate has this reversed and the difference dimin- ished, likely due to using specimens rather than field experience, as the contrast is accentuated in the brighter light in the field. Plate 41. The underwing pattern of both adult and immature Spot-billed Pelican {Felecamts philippensis) is inaccurate. It should have a rather banded appearance with white greater and median underwing coverts contrasting with vinous (adult) or dusky brown (immature) lesser underwing coverts and dusky brown (looking paler or darker depending on light angle) primaries and second- aries. The outer primaries are blackish. Plate 46. Adult and immature labels are switched on Eastern Osprey (Pandion cristatiis). Plate 48. I had the privilege of seeing an adult of the "dugeP' subspecies (or moiph) of Steller’s Sea Eagle {Haliaeetus pelagicus) in South Korea on 3 January 1962 and aver this entirely black eagle with its huge yellow bill and white tail could not possibly be mistaken for anything else, contra the text. Plate 52. While the crest of the Crested Goshawk (Accipiter oivirgatus) in the wind could possibly look like the plate rendition, it normally lies flat on the nape with the tip slightly above the nape (perhaps 2 mm.) and just barely visible. Plate 54. The complexity of the plumages of the buzzards (Biiteo) can only be hinted at in the space available here, e.g., the individual tail feathers of most Upland Buzzards {B. hemilasiiis) are white in the center and darker on the edges so that the tail looks mostly white when spread and darker when closed. Plate 59. While the flight picture of Ruddy- breasted Crake {Porzana fusca) accurately depicts this bird’s “jizz”, the walking bird is both badly misshapen and too large (should be smaller than Band-bellied Crake [P. paykuUii]). Plate 65. The specific name of European Golden Plover (Pluvicdis apricaria) is misspelled {aricaria rather than apricaria). Plate 69. Neither mentioned in the text nor pictured properly is the fact the white flank feathers of the Greater Painted Snipe (Rostratida henglia- lensis) in flight curve up and over the sides of the base of the tail to give a distinctive Ruff-like pattern. Plate 7 1 . The width of the white band on the tips of the secondaries of the Common Snipe {Gallinago gallinago) and Wilson’s Snipe (G. delicata) are different and incorrectly depicted (switched) in the plate (correct in text). Plate 90. Missing in text is the fact that Relict Gull (Ichthyaetus relictus) often forages along shore like a shorebird. Plate 104. The Marbled Murrelet (Brachyram- plius marmoratus) is pictured and listed as a species that might turn up in northeastern Asia. However, there is a record for Idlidlya Island in ORNITHOLOGICAL LITERATURE 205 KoUichia Bay on the northern Chukotka Peninsula (Dement’ev et al. 1951, Birds of the Soviet Union). Plate 1 15. Eastern Grass Owl {Tyto longimeni- hris) text and plates depict only huffy phase. White phase has white underparts and white upperside of tail with narrow black bars. Plate 120. The text account of Brown Hawk- Owl (Nino.x sentidata) is a bit muddled. My 2002 paper (Bulletin B.O.C. 122: 250-257) on this species complex showed the song of the migratory northern race japonica is the same throughout its range and different from the resident races in southern Asia. I recommended splitting them based on vocalizations as well as some moipho- metric differences. The resident subspecies totogo (southern Japan and Taiwan) is close to japonica both vocally and morphometrically with a weak potential for a split. Brazil chose to split the japonica subspecies into eastern japonica and western florensis and suggests that only japonica might be separate from scutulata, which doesn’t make sense (japonica and florensis have identical vocalizations and similar morphometries, which are consistently distinct from the subspecies of scutulata). The English name I suggested in my paper, Northern Boobook, is mentioned, but the paper is not listed in the online bibliography. Plates 127, 130. The colors of Rufous (Celeus hrachyurus). Pale-headed (Gecinulus grantia), and Bay (Blythipicus pyrrhotis) woodpeckers are very badly off. Plate 133. Heads of Tiger Shrikes (Lanins tigrinus) are much too big. It should be noted that any new work with as broad a sweep as this one, and its consequent massive number of data points, inevitably has short-comings. That this one has so tew is a tribute to the care and expertise applied to the project by the author. It is remarkable indeed that the book is so far along the route to perfection on its first try. This guide covers its area quite thoroughly, enough so, that I would be quite comfortable catrying it as my only guide in its geographic range. Further, it will be quite useful in the Philippines and Indonesia. Birders throughout Eurasia and North America will find it helpful for identifying strays. 1 highly recommend this book to anyone interested in Asian birds and expect to be using it in my Asian jaunts. — BEN F. KING, Ornithology Department, American Museum of Natural History, Central Park West at 79th Street, New York, NY 10024, USA; e-mail; kingbirdtours@earthlink.net THE WILSON JOURNAL OL ORNITHOLOGY Editor CLAIT E. BRAUN 5572 North Ventana Vista Road Tucson, AZ 85750-7204 E-mail: TWILSONJO@comcast.net Editorial NANCY J. K. BRAUN Assistant Editorial Board RICHARD C. BANKS JACK CLINTON EITNIEAR SARA J. OYLER-McCANCE LESLIE A. ROBB Review Editor ROBERT B. PAYNE 1 306 Granger Avenue Ann Arbor. MI 48104, USA E-mail: rbpayne@umich.edu GUIDELINES FOR AUTHORS Please consult the detailed ‘'Guidelines for Authors” found on the Wilson Ornithological Society web site (http://www.wilsonsociety.org). All manuscript submissions and revisions should be sent to Clait E. Braun, Editor, The Wilson Journal of Ornithology, 5572 North Ventana Vista Road, Tucson, AZ 85750-7204, USA. The Wilson Journal of Ornithology office and fax telephone number is (520) 529-0365. The e-mail address is TWilsonJO@comcast.net NOTICE OF CHANGE OF ADDRESS Notify the Society immediately if your address changes. Send your complete new address to Ornithological Societies of North America, 5400 Bosque Boulevard, Suite 680, Waco, TX 76710. The permanent mailing address of the Wilson Ornithological Society is: 7oThe Museum of Zoology, The University of Michigan, Ann Arbor, MI 48109, USA. Persons having business with any of the officers may address them at their various addresses given on the inside of the front cover, and all matters pertaining to the journal should be sent directly to the Editor. MEMBERSHIP INQUIRIES Membership inquiries should be sent to Timothy J. O’Connell, Department of Natural Resources Ecology and Management, Oklahoma State University, 240 AG Hall, Stillwater, OK 74078; e-mail: oconnet@ okstate.edu 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 e.stabli.shed in concurrence with the University of Michigan in 1930. Until 1947 the Library was maintained entirely by gifts and bequests of books, reprints, and ornithological magazines from members and friends of tbe 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 is administered by the Library Committee. Jerome A. Jackson, Florida Gulf Coast Univeristy, is Chairman of the Committee. The Library currently receives over 200 periodicals as gifts and in exchange for The Wilson Journal of Ornithology. For information on the Library and our holdings, see the Society’s web page at http:// www.wilsonsociety.org. 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 pos.sessions, 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, 1109 Geddes Avenue, Ann Arbor, MI 48109-1079, USA. Contributions to the New Book Fund should be sent to the Treasurer. This issue of The Wilson Journal of Ornithology was published on 1 March 2010. Continued from outside back cover 139 Home-range size and site tenacity ol overwintering Le Conte’s Sparrows in a fire managed prairie Heather Q. Baldivin, Clinton W. Jeske, Melissa A. Powell, Paul C. Chadwick, and Wylie C. Barrow Jr. 146 Paternal song complexity predicts offspring sex ratios close to Hedging, but not hatching, in Song Sparrows Dominique A. Potvin and Elizabeth A. MacDougall-Shackleton 153 Influence of age and season on morphometric measurements ol the Biscutate Swift {Streptoprocne biscutata) Mauro Pichorim Short Communications 160 Cooperative breeding by Red-Headed Woodpeckers Megan R. Atterberry-Jones and Brian D. Peer 162 Observations on the breeding biology of the Collared Crescentchest {Melanopareia torquata) Mieko F. Kanegae, Marina Telles, Severino A. Lucena, and Jose Carlos Motta-Junior 165 Effects of a flood on foraging ecology and population dynamics of Swainson’s Warblers Nicholas M. Anich and Bryan M. Reiley 168 Predator vocalizations affect foraging trade-offs of Northern Cardinals Mark T. Stanback and Emily M. Powell 173 Observations and predation of a Coral-billed Ground Cuckoo {Carpococcyx renauldi) nest in northeastern Thailand Korakoch Pobprasert and Andrew J. Pierce 177 Diet of Chinese Grouse {Tetrastes sewerzowi) during preincubation Wangjie, Yang Chen, Lu Nan, Fang Yun, and Sun Yue-Hua 181 Observations of a possible foraging tool used by Common Ravens Paul E Frame 183 Hanging behavior of the Hooded Crow {Corvus cornix) Mario Melletti and Marzia Mirabile 186 Conspecific brood parasitism by the Dickcissel Brian D. Peer 188 First report of Olrog’s Gull depredation by sympatric Kelp Gulls Luciano La Sala and Sergio Martorelli 190 Second and third records of Snares Penguins {Eudyptes robustus) in the Falkland Islands Laurent Demongin, Maud Poisbleau, Georgina Strange, and LanJ. Strange 194 A new bird species for Costa Rica: Sapphire-throated Hummingbird {Lepidopyga coeruleogularis) Esteban Biamonte 196 Ornithological Literature Robert B. Payne, Review Editor The Wilson Journal of Ornithology (formerly The Wilson Bulletin) Volume 122, Number 1 CONTENTS March 2010 Major Articles 1 Demography of a reintroduced population of Evermann’s Rock Ptarmigan in the Aleutian Islands Robb S. A. Kaler, Steve E. Ebbert, Clait E. Braun, and Brett K Sandercock 15 Apparent survival of breeding Western Sandpipers on the Yukon-Kuskokwim River Delta, Alaska Matthew Johnson, Daniel R. Ruthrauff, Brian J. McCajfery, Susan M. Haig, and Jejjrey R. Walters 23 Natal philopatry and apparent survival of juvenile Semipalmated Plovers Erica Nol, Simone Williams, and Brett K Sandercock 29 Breeding biology and nesting success of the Slate-throated Whitestart {Myioborus miniatus) in Monteverde, Costa Rica Ronald L. Mumme 39 Breeding biology of Kelp Gulls on the Brazilian coast Gisele Fires de Mendonga Dantas and Joao Stenghel Morgante 46 Cooperative breeding of the Society Kingfisher {Todiramphus veneratus) Dylan C. Kesler, Thomas Ghestemme, Emmanuelle Fortier, and Anne Gouni 51 Reproductive success of Burrowing Owls in urban and grassland habitats in southern New Mexico Daniele Berardelli, Martha ]. Desmond, and Leigh Murray 60 Dietary trends of Barn Owls in an agricultural ecosystem in northern Utah Carl D. Marti 68 Regional variation in diets of breeding Red-Shouldered Hawks Brad N. Strobel and Clint W. Boal 75 Effects of Brown-Headed Cowbird parasitism on provisioning rates of Swainson’s Warblers Sara Fappas, Thomas J. Benson, and James C. Bednarz 82 Effects of fat and lean body mass on migratory landbird stopover duration Chad L. Seewagen and Christopher G. Guglielmo 88 Seasonal fluctuation of the Orange-winged Amazon at a roosting site in Amazonia Leiliany Negrao de Moura, Jacques M. E. Vielliard, and Maria Luisa da Silva 95 Costs and benefits of foraging alone or in mixed-species aggregations for Forster’s Terns Lisa SchreJJler, John K Leiser, and Terry L. Master 102 Abundance and distribution of waterbirds in the Llanos of Venezuela j Francisco J. Vilella and Guy A. Baldassarre | 1 16 Phenology of six migratory coastal birds in relation to climate change ; Charles R. Foster, Anthony F. Amos, and Lee A. Fuiman . j 126 Nest-site limitation of secondary cavity-nesting birds in even-age southern pine forests Karl E. Miller 1 35 House Sparrows associated with reduced Cliff Swallow nesting success Douglas R. Leasure, Ragupathy Kannan, and Douglas A. James Continued on inside back cover Volume 122, Number 2, June 2010 I Published by the Wilson Ornithological Society THE WILSON ORNITHOLOGICAL SOCIETY FOUNDED 3 DECEMBER 1888 Named after ALEXANDER WILSON, the first American ornithologist. President — E. Dale Kennedy, Biology Department, Albion College, Albion, MI 49224, USA; e-mail: dkennedy@albion.edu First Vice-President — Robert C. Beason, P. O. Box 737, Sandusky, OH 44871, USA; e-mail: Robert.C.Beason@gmail.com Second Vice-President — Robert L. Curry, Department of Biology, Villanova University, 800 Lancaster Avenue, Villanova, PA 19085, USA; e-mail; robert.curry@villanova.edu Editor — Clait E. Braun, 5572 North Ventana Vista Road, Tucson, AZ 85750, USA; e-mail: TWILSONJO@ comcast.net Secretary — John A. Smallwood, Department of Biology and Molecular Biology, Montclair State University, Montclair, NJ 07043, USA; e-mail: smallwoodj@montclair.edu Treasurer — Melinda M. Clark, 52684 Highland Drive, South Bend, IN 46635, USA; e-mail: MClark@tcservices.biz Elected Council Members — Robert S. Mulvihill, Timothy J. O’Connell, and Mia R. Revels (terms expire 20 1 0); Jameson F. Chace, Sara R. Morris, and Margaret A. Voss (terms expire 2011); Mary Bomberger Brown, Mary Garvin, and Mark S. Woodrey (terms expire 2012). Living Past-Presidents — Pershing B. Hofslund, Douglas A. James, Jerome A. Jackson, Clait E. Braun, Mary H. Clench, Richard C. Banks, Richard N. Conner, Keith L. Bildstein, Edward H. Burtt Jr., John C. Kricher, William E. Davis Jr., Charles R. Blem, Doris J. Watt, and James D. Rising. Membership dues per calendar year are: Regular, $40.00; Student, $20.00; Family, $50.00; Sustaining, $100.00; Life memberships, $1,000.00 (payable in four installments). The Wilson Journal of Ornithology is sent to all members not in arrears for dues. THE WILSON JOURNAL OF ORNITHOLOGY (formerly The Wilson Bulletin) THE WILSON JOURNAL OF ORNITHOLOGY (ISSN 1559-4491) is published quarterly in March, June, September, and December by the Wilson Ornithological Society, 810 East 10th 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 OSNA, 5400 Bosque Boulevard, Suite 680, Waco, TX 76710. All articles and communications for publication should be addressed to the Editor. Exchanges should be addressed to The Josselyn Van Tyne Memorial Library, Museum of Zoology, Ann Arbor, Ml 48109, USA. Subscriptions, changes of address, and claims for undelivered copies should be sent to OSNA, 5400 Bosque Boulevard, Suite 680, Waco, TX 76710, USA. Phone: (254) 399-9636; e-mail: business@osnabirds.org. Back issues or single copies are available for $12.00 each. Most back issues of the Journal are available and may be ordered from OSNA. Special prices will be quoted for quantity orders. All issues of the journal published before 2000 are accessible on a free web site at the University of New Mexico library (http://elibrary.unm.edu/sora/). The site is fully searchable, and full-text reproduetions of all papers (including illustrations) are available as either PDF or DjVu files. © Copyright 2010 by the Wilson Ornithological Society Printed by Allen Press Inc., Lawrence, KS 66044, USA. COVER: Wilson’s Storm Petrel {Oceanites oceanicus). Illustration by Don Radovich. @ This paper meets the requirements of ANSI/NISO Z39.48-1992 (Permanence of Paper). MCZ LIBRAf=»Y JUN 0 3 2010 HARVARO UNIVERSITY FRONTISPIECE. We quantified male and female parental care by House Sparrows {Passer domesticiis) to ascertain food provisioning rates by each parent (page 211) during the 2-week nestling period. Adult fitness was shaped primarily by the parents’ ability to deliver enormous insect prey items to offspring, f’arental success (i.e., total recruitment of offspring) was predicted by size and numbers of offspring at banding (day I 1 ). Water color by Don Radovich. Wilson Journal ofO rnithology Published by the Wilson Ornithological Society VOL. 122, NO. 2 June 2010 PAGES 207-416 The Wilson Journal of Ornithology 122(2):207-2 16, 2010 NOT THE NICE SPARROW The 2007 Margaret Morse Nice Lecture DOUGLAS W. MOCK' - AND P. L. SCHWAGMEYER' ABSTRACT. — We began our studies of House Sparrow (Passer domesticiis) biparental care in the mid-1990s by applying the classic Margaret Morse Nice field technique of color-banding individuals. Over the ensuing summers, we slowly accumulated quantitative provisioning records for 100 broods, even as we commenced a series of experimental manipulations. Provisioning data showed parental fitness, as expressed by offspring recruitment into the local breeding population, to be shaped mainly by the adults' ability to deliver enormous insect prey items. It also turned out that production of robust and competitive fledglings routinely involves losing one or more nestlings (brood reduction in 42% of 1.000 multi-chick families). Recruitment success was compensated for the death of an offspring if the subsequent reallocation of food enables surviving nest-mates to gain at least 2 g more before fledging. Video samples showed that parents of day 3 broods favored larger siblings, even though brood reduction typically occurs on —day 4, suggesting that adults participate actively in promoting some offspring over others. The social dynamic affecting how parents work as a team during provisioning does not fit the pattern expected if partners negotiate actively with one another, but points more toward the likelihood of “sealed bids.” Specifically, experimental handicapping of individual parents (tail-weights and hormone implants) indicates partners operate quite independently during brood-rearing. We are now extending our experiments into the incubation phase, where parents are probably better-informed about partner activities, thus potentially able to adjust to fluctuating contributions. Finally, behavioral rules affecting food deliveries seem to differ for females and males. Females normally increase provisioning as the brood ages, but males do not. However, when broods received supplemental nutrients, males matched the female upsurge, accelerating their deliveries by 25%. showing they usually work well below capacity. Received 31 December 2009. Accepted 28 February 2010. Among the many contributions for which bird behaviorists are grateful to Margaret Morse Nice, her pioneering use of color bands is probably the most universal. Color-banding seems utterly fundamental to behavioral ornithologists today — a trifle really — but on 26 March 1928, one pink celluloid ring provided Mrs. Nice with a means of recognizing a particular male Song Sparrow 'Department of Zoology. 730 Van Vleet Oval. Univer- sity of Oklahoma. Norman, OK 73019, USA. “Corresponding author; e-mail: dmock@ou.edu (Metospiza meloclia) near her Columbus, Ohio home. Soon she banded a second male with a green ring, thereby exposing consistent differ- ences between the two in song, aggressiveness, etc. It is hard to overstate what a difference this innovation has made to behavioral field studies. Pausing to think about it. much that we take for granted today was not available to Margaret Nice. For example, well before today's ornithological neophyte affixes any color bands, the step of identifying the local avifauna, or even finding out which birds live nearby, promotes a trip to the 207 208 THE WILSON JOURNAL OF ORNITHOLOGY ♦ Vol. 122, No. 2. June 2010 bookshelf for a field guide where the only dilemma is which one to choose. When Margaret Morse Nice arrived in Oklahoma in 1913 (her husband Blaine had been hired in what is now the Zoology Department where we both work), no bird guides or even checklists existed. Although she had strong academic credentials (including a Master’s degree from Clark University), she never held a position at the University of Oklahoma. Instead, she stayed physically busy with the traditional duties of wife and mother, while fighting boredom and intellectual frustration by reviving her ornithological interests. Eventually she wrote Oklahoma’s first state bird book but, by the time it was published, the family had moved (with Blaine’s career) to Ohio State, where she began to focus on avian behavior. The Oklahoma house that Blaine Nice built (literally!) at 445 College Avenue is just a block from where we now live, but nearly everything else has changed. The streets have been paved, more than one faculty member owns an automo- bile (indeed, it seems as if most OU students have more than one!), and women now comprise a third of our department’s faculty. Meanwhile, Margaret Nice’s fields of natural history and ethology have been infused with theoretical and mathematical fibers to form modem behavioral ecology (Parker 2006). We all inhabit this new world, but should pause from time to time and acknowledge the extraordinary spadework that made our studies possible. That said, many Nice-era roots remain clearly visible, despite being taken for granted. House Sparrows {Passer domesticus) were already abun- dant in Norman when she arrived in 1913, but we have now color-banded .several thousand since 1994. Otherwise, we spend much of our field time sitting quietly and, hopefully, recording what the birds would have done in our absence. The overarching goal of our collaborative research program is to understand what makes biparental care evolutionarily stable. Because cooperation in parenting makes a major contribu- tion to social monogamy, which predominates in Class Aves, we .see it as central to understanding that mating system. House Sparrows are abundant and willing to use nest boxes, so we study House Sparrows. This also is consistent with the Margaret Morse Nice tradition of studying “... what ap- peared common to .so many” (Nice 1979:263). We thank the Wilson Ornithological Society for honoring us with the 2007 Margaret Morse Nice Medal and offer the following sketch of some things we have learned about these quintessen- tially ordinary birds through a series of shared projects. Our goal is not any sort of autobio- graphical detail, except as it relates to the evolution of our sparrow research program, but to provide an overview of several related studies that have led us to our current work. We do not presume to offer this account as a model of how anyone else should proceed, but students of bird research need not regard the whole field process as mysterious. To reduce that mystery, we explain how we try to extract our guiding questions from the general picture, use them to design empirical exercises, and then seek to re-apply the analyzed results back to the bigger issue from whence we start. Along the way, it must be confessed, we often stumble. Space limitation and reader interest preclude a full chronicling of all our mistakes (trap designs that were prone to equipment malfunctions, an attempt to supplement breeding parents with live mealworms that escaped, occasional duplications of color band combina- tions, etc.), but we hope students will forgive themselves for similar false starts as they press forward. We both came of age as scientists more or less as Darwin’s theory of sexual selection was being revived after a century of neglect (Campbell 1972) and as several other major pieces of the evolutionary puzzle, especially inclusive fitness theory (Hamilton 1964a, b) and optimality modeling (reviewed in Parker and Maynard Smith 1990), were coming into focus as the foundations of behavioral ecology. Unlike some previous awardees of the Margaret Morse Nice Medal, we had not grown up as bird-watchers per se, but more generally as animal enthusiasts. Knowing from childhood that we wanted to have “animals” in our lives, we had gravitated naturally toward animal behavior research while in school. Arriv- ing independently 2 years apart and meeting for the first time as University of Oklahoma faculty members, we continued with our separate lines of study for quite a few years focusing on scramble- competition polygyny and .sperm competition in local ground squiirels (e.g., Schwagmeyer 1988, Foltz and Schwagmeyer 1989, Schwagmeyer and Foltz 1990), and fatal sibling rivalry and piirent- off.spring conflict in egrets (e.g.. Mock 1984, 1985; Mock et al. 1987). Realizing that we needed to learn more about mathematical modeling, we visited several British universities over the 1983 Mock and Scliwai'nicyer • NOT THE NIC'E SPARROW 209 Christmas break, pausing to get married in Edinburgh. It was easy to agree on Geoffrey Parker as our theoretical mentor and collaborator, as he had discovered sperm competition (e.g., Parker 1970) was having a major role in the development of Evolutionary Game Theory (e.g., Maynard Smith and Parker 1976, Parker and Stuart 1976, Parker 1978, Parker 1983, Parker and Maynard Smith 1990), and was already deeply involved in modeling both of our research topics (e.g., Parker and Macnair 1978; Parker 1982, 1985). We spent fall semester 1984 with him at the University of Liverpool, taking turns working on models with Geoff and spending the rest of our time either pulling numbers out of squin-el and egret data sets or desperately salvaging our neglected mathematical training. Over the ensuing years, we produced two separate series of publications with Geoff (e.g., Schwagmeyer and Parker 1987, 1990; Parker et al. 1989; Mock and Parker 1986, 1997), then began searching for a field system to explore together. That led first to two field seasons working on Glaucous-winged Gulls {Larus glaiicescens) in Puget Sound, specifically on their pre-hatching acoustic com- munications (Schwagmeyer et al. 1991), before we decided to focus on biparental care in House Sparrows. CHOICE OE PROBLEM Biparental care and monogamy are not super- ficially sexy topics, at least on the surface. Except for feathered vertebrates, both traits are exceedingly rare in nature. Thus, the question naturally arises: what ecological and evolutionary forces generally promote the evolutionary stabil- ity of reproductive cooperation between two adults? Why have birds, in particular, adopted these as the norm, when other taxa have not? Einally, although many human cultures (includ- ing our own) claim both as societal goals, why are they so elusive? It seems logical to explore the underlying problems of shared parental care (seeking what may be broadly general causes) by analyzing the behavioral dynamics with birds. Eurthermore, it makes sense to choose a bird that is common and amenable to (non-destructive) experimental manipulation of the targeted be- havioral components. Biparental investment is most simply ex- plained as the pattern that emerges when neither of the sexual partners has anything better to do with its time. Specifically, if either parent has the realistic option of alternative pursuits that deliver higher average fitness returns, those distractions would likely undermine the joint endeavor (Maynard Smith 1977). Understanding the fitness payoffs available to each “player” for its various alternatives lies at the heart of the matter. In many taxa (e.g., mammals and plants), dependent offspring remain physically anchored to the female, extracting nutrients from her for extend- ed periods. This arrangement “emancipates” the other parent to explore additional potential reproductive opportunities. Similarly, when off- spring are not physically attached, and especially when their basic requirements are lessened (e.g., by precocial development), it is easy to see how uniparental care may suffice across a broader array of taxa. If low-cost offspring are shed soon after fertilization, the uniparent is less rigidly pre-ordained and “sex-role reversal” is more easily evolved (males serving as solo care-givers in many fishes and shorebirds, for example). In most birds, combining of egg-laying with altricial development seems to have created a middle zone in which more than one adult can make valuable contributions to offspring welfare. In short, when each embryo is placed in a neutral arena (nest) soon after it has been fertilized — and especially if a hasty departure by the first male reduces the chance that his sperm will fertilize the next ovum — the male has inducements for sticking around at least through the laying period. Thereafter, if he can further his own selfish interests better by providing additional services to his offspring (directly to the eggs and/or by proxy to his partner), than by seeking additional mating opportunities, he may remain. Once hatching begins, the need for provisioning with vast quantities of exogenous food can extend male participation on the home front (along with the continuing parental tasks of defense against weather and enemies, etc.). Eollowing this argument, ecological and phy- logenetic circumstances may provide plausible reasons why biparental care is evolutionarily stable across —90% of feathered vertebrates (Lack 1968, Glutton-Brock 1991), yet rare in all other life forms. Parental duties .seem sufficiently extensive to require a team of care-givers and, perhaps equally important, to preclude either the option of dumping the whole brood-rearing exercise on the other (i.e., the “cruel bind” of Trivers 1972), lest the abandoned parent simply follow suit and allow the brood to fail. 210 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2. June 2010 To put a face on the overall dynamic, consider House Sparrow reproduction. After ~11 days of incubation (some by the male), four or five hatchlings emerge that will develop from tiny (1-2 g) naked ectotherms into fully feathered, warm-blooded, adult-sized (20-25 g) flying machines within just 2 weeks. During the nestling period, the two adults must find, capture, and deliver (typically one per trip) 3,000 to 4,000 unwilling insects to fuel this growth. At the same time, the parents also provide brooding warmth (at night and on chilly days), resist certain predators (and various other intruders, including hostile conspecifics), and maintain a modicum of nest sanitation by removing fecal sacs. Although most House Sparrows are socially monogamous, some males acquire two partners simultaneously and many females accept gametes from extra-pair males; in our Oklahoma population, ~ 17-20% of young are sired by other males (and 41-45% of broods contain one or more such extra-pair offspring) (Whitekiller et al. 2000, Edly-Wright et al. 2007). Overall, these are pretty ordinary passerines and, thus, a useful model system. CHOICE OE SUBJECTS According to W. L. Dawson (1903:40), “With- out question the most deplorable event in the history of American ornithology was the intro- duction of the English Sparrow.” We may wince over that sweeping indictment, hastily noting that the arrival of European hominids was surely more “deplorable” for native bird species by virtue of myriad associated ecological nightmares (e.g., novel pathogens, deforestation, pollution, and introductions of numerous destructive exotics, especially pets, livestock, and all the birds mentioned in Shakespeare, etc.), but we get Dawson’s point: this is not the most glamorous and beloved bird in the New World. We did not pick it for glamour. As behavioral ecologists, we chose our topic first and then sought the best local bird with which to pursue it. Had bluebirds, chickadees, or wrens piled into our nest boxes, we might well have studied them. But we were expecting sparrows and chose two sparsely occupied (by humans), university-owned tracts near campus for our study areas, partly because many House Sparrows were already nesting in the porous old buildings left from a World War II naval air station. One of these decaying structures is still dignified as the university’s Animal Behavior Laboratory. Our goal was to lure sparrows from those decrepit buildings to fine, new nest boxes mounted where we could reach them conveniently. Surveying possible box de- signs, we learned that standard bluebird houses are annoyingly popular with House SpaiTows and explicit warnings advised us that House Span'ows like having a perch, prefer being near human- occupied buildings, etc. We actively catered to sparrow tastes. In the early seasons, we had problems with a few avian predators, particularly Loggerhead Shrikes {Lanins ludovicianus) that trap-lined our boxes, harvesting sparrow nestlings simply by sticking their heads through the hole and seizing the tallest beggars. To thwart this, we attached a simple 10-cm long corridor of galva- nized hardware cloth to every box. The sparrows did not mind entering through these hall-like structures and mass-predation losses dropped. TRAPPING, MARKING, AND CENSUSING We used the usual trapping methods for building a color-banded population, including walk-in traps baited with seed and bread, mist nets, and even sneaking up on brooding males to catch them in nest boxes. We learned the hard way not to touch desertion-prone females during incubation. We applied three color bands (slitting toy Perler Beads as an inexpensive supply; Hill 1992) and one aluminum band in unique combinations for each bird. After much trial-and-error with a variety of in- box trap designs, we developed a Rube Goldberg contraption involving a modified wire conndor (sporting a drop-down door to prevent exit), some monofilament fishing line, a rat-trap, and remote- control toy car electronics that empowered us to catch uncooperative individual adults during their brief visits to provision chicks (Mock et al. 1997). The general problem was that House Sparrows are smart and wary: catching one member of a pair without disturbing its mate required a trap that was simultaneously invisible, under our instantaneous control from considerable distances, and impervi- ous to strong prairie winds. Once captured, we carried subjects inside an opaque bag to a car, processed it there (weighing, banding, etc.), and then canned it to some stranger’s vehicle for release; simply releasing it from one’s own car was found to lead to scolding of that vehicle during subsequent attempts to observe behavior. BEHAVIORAL SAMPLING Reliable quantification of male and female parental care took advance preparation to opti- Mock and Schwugnwycr • NOT THE NICE SI^ARROW 211 mize the later research elfort. A chief goal was to ascertain food provisioning rates by each parent (Frontispiece) across the 2-week nestling period. We needed a sampling regime that would generate representative estimates. Average figures in quite a few published field studies of avian parental care are based on rather scant samples (at times just 15-30 min observations taken on 1-2 days convenient to the researchers) without any attention given to whether these capture the essence of the parents’ activities across daylight hours, age of the nestlings, shifts in food availability within and between seasons, etc. Because we were going to do a lot of monitoring over several summers, we wanted to know how much sampling was needed and when this should be done. For this purpose, we conducted a preliminary study in which we recorded parental delivery rates intensively at five nests (covering all daylight hours and brood ages), then performed regressions on these over-sized samples to see if some times of day (and/or combinations thereof) captured most of the variation. We learned that a single hour-long sample taken mid-day yielded a good picture of the whole day’s provisioning (between-nest comparison), but one taken in mid- morning did the worst (see Schwagmeyer and Mock 1997 for details). Combining samples delivered 91-99.7% (depending on which hours) of the variation between nests and nearly 80% of within-nest variation. In the seasons that followed, we used these findings to guide our deployments, eventually amassing parenting records for more than 100 broods based on >13 of these selected hours per brood. During these observations, we recorded parents’ arrivals and departures, size of prey being carried (three categories, scaled against adult beak), and other activities performed near the nest (e.g., foraging, singing, defending, etc.) These data provided a good picture of parental care in regular sparrow nesting cycles and in those subjected to experimental manipula- tions. BUILDING SUCCESSFUL OFFSPRING At the heart of any parental care problem lies the question of what is needed to produce valuable offspring. There are many components, including protection from the elements and predators, but finding and delivering sufficient food to the brood (provisioning) and then distributing it optimally once home (allocating) are generally assumed to supersede most other concems, because of the phenomenal growth challenge ol oilspring. Once eggs have hatched, parentally delivered food is the chief factor limiting growth, survival to Hedging, and recruitment to breeding age of House Sparrow offspring; hence it is a major determinant of parental fitness. Provisioning, as the central task shared between the parents, was the initial focus of our exploration into the stability of biparental care. If food is limiting, the simplest prediction is that offspring growth and survival should be density (brood-size) dependent for a given level of parental effort. This is usually assessed over the short span of the nestling period (volant offspring being necessarily harder to count) as one discrete exercise. In many field studies, brood size at hatching is compared with survival and body mass at fledging (typically defined as banding age, just before first flight). This requires only a brief census visit at hatching and measurement of body size during the banding process. Few field studies calibrate those relationships with intensive sam- pling of actual food delivered, presumably because this feature is tedious and time-consum- ing. A separate exercise, seldom performed in tandem with pre-fledging evaluation, extends the matter from that point by comparing how fledging production (e.g., size and numbers of banded chicks) translates into offspring survival to breeding age. This second step has produced far greater documentation (for at least 22 avian species; reviewed by Magrath 1991, Schwag- meyer and Mock 2008) of the commonsense idea that parents producing more and larger fledglings should consequently achieve greater recruitment of breeding offspring. We were committed to doing the necessary observations of provisioning, as an essential part of analyzing biparental division of labor, and took the opportunity to combine the two steps (pre- fledging and post-fledging) on the marked House SpaiTow population. Mean offspring mass on day 1 I, to our mild initial surpri.se, was not predicted by total food deliveries, but was strongly predicted by the rate at which the “enormous” category of insects (>2 cm total body length, hereafter, “e-prey”) were provided (Schwag- meyer and Mock 2008). Once a correction is made for the much greater volume and mass of e- prey, it seems likely these grasshoppers, caterpil- lars, etc. can provide most of the nutrition early in the season, despite being relatively uncommon (—14% of all deliveries). Thus, total amount of 212 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2. June 2UI0 tood is important, as expected, but simple counts ot parental food trips — as might be recorded by automated devices — would have pointed to the opposite conclusion. We used reobservations of 100 banded off- spring from a sample of 1,028 broods to evaluate how chick numbers and mean mass relate to local recaiitment. This measure of parental success was predicted by both size and numbers of offspring at banding (Schwagmeyer and Mock 2008). We found that our data from 99 broods observed for provisioning rates were sufficient for a direct assessment of how food deliveries affect recruitment. This connection had been established before only once, for Long-tailed Bushtits (Aegithalos caiidatus), a cooperatively- breeding British bird (MacColl and Hatchwell 2003). Total delivery counts once again failed to predict survival to breeding age in our sparrows, but e-prey deliveries did (Schwagmeyer and Mock 2008). Overall, it seems clear that e-prey availability is a key determinant of parental success and these food items have an especially strong effect on recruiting for earlier broods. Considering that parents presumably cannot simultaneously maximize both the number and sizes of current brood members, much attention has been given over the years to short-term adjustments they may make during the period of offspring dependency. Early mortality of individ- ual nestlings has the potential to relax competition for insufficient food and the idea that brood reduction might be engineered by parents as part of an adaptive strategy for tracking unpredictable food availability dates to the classic writings of David Lack (1947, 1954). The resolution of acute sibling rivalry in some taxa includes extreme forms of nestling aggressiveness and even spe- cialized weaponry, but in many it results from low-key processes of scramble competition (re- viewed in Mock and Parker 1997). Passerine brood reduction falls overwhelmingly into the latter category and it is hard to learn unambigu- ously whether the loss of any given brood member makes a positive contribution to the inclusive fitness interests of surviving parents and siblings, an overall negative, or a matter of little impor- tance one way or the other (O’Connor 1978). We found that 42% of 1,000 multi-chick House Sparrow broods experienced partial-brood mor- tality between hatching and Hedging (Mock et al. 2009) with greater frequency in larger broods, as one would expect. Most of the.se losses occurred early in the nestling period (median day 4), before parental capacities to provision are likely to be overtaxed. We found parental provisioning (in the closely observed sample) increased from broods of three to broods of five, and per-capita food remained relatively even. Family size was a strong predictor of whether brood reduction occurred (provisioning and family size are both good predictors of the proportion of the brood surviving to fledge). Once a single-chick mortality event had occurred, parents did not trim their provisioning rate, as has been documented for siblicidal Cattle Egrets (Buhiilcus ibis'. Mock and Lamey 1991) and Brown Pelicans (Pelecaniis occidentalis', Ploger 1997); thus, sparrow survi- vors tended to receive a food bonus in the aftermath of brood reduction. We took two further steps to explore whether parents might favor or disfavor brood reduction. We affixed tiny video cameras above false ceilings inside 1 1 nest boxes containing broods of four and recorded how food was distributed when nestlings were sufficiently young (3 days) that allocation could be attributed safely to parental decisions (i.e., not to nest-mates’ Jock- eying effectively for positional advantage). The idea was to see if fully empowered parents show overt favoritism toward the smallest chick(s) in each brood by giving runts more food. These parents skewed allocation more to larger chicks (Mock et al. 2009), a pattern that persists (and often increases) after the stronger chicks usurp increasing control over allocations. Thus, brood reduction victims tend to be those that were smallest initially. Second, we analyzed the trade- off between size and numbers of fledglings on recruitment to estimate how much additional body mass of surviving nest-mates is required to compensate parental fitness for loss of one offspring. This exercise suggested that one chick’s death is fully reimbursed if its surviving siblings fledge about 2 g heavier (e.g., by virtue of the food the victim would have consumed). Beneficial reallocation probably occurs often, apparently when parental control is at its peak (Mock et al. 2009). The foraging task shared between two sparrow parents is something of a malleable abstraction, perhaps roughly estimated (within a range of limits) when clutch size is first known to both partners, but then shifting over the course of events including hatching failure, seasonal ebb and How of food availability (insect hatches. Mock (iiicl Sclnccif'incycr • NOT T1 IE NICE SPARROW 213 periods of inclement weather, etc.), and loss of nestlings. DO PARENTS NEGOTIATE WITH EACH OTHER? The most influential hypothesis for biparental cooperation outlines a behavioral dynamic be- tween two monogamous partners as being the fruit of a negotiation (Chase 1980, Houston and Davies 1985, McNamara et al. 1999). This assumes that each parent has a generally self-serving level at which it is willing to work, but must coordinate with a like-minded partner. The “self-serving” qualifier reflects that participants are not genetic kin, so their confederation is held together by personal fitness interests (cf. inclusive fitness incentives). Each individual seeks the best deal possible, despite needing a sexual partner, within the framework of its joint reproductive enterprise, while prefering that its partner provide somewhat more than half of the work. Both partners cannot realize this preference simultaneously. Just as union representatives and owners engage in a series of offers and counter-offers during a labor dispute (Maynard Smith 1982), male and female parents are envisioned as adjusting and re- adjusting their levels of effort, either in evolu- tionary time or in real time, until each is satisfied that it cannot get a better deal. This final pair of positions is, in the parlance of evolutionary game theory, the ESS (Evolutionarily Stable Strategy) solution. One attractive feature of this Negotiation Model for biparental care is its testable prediction of partial compensation, the optimal response that a parent should make if its partner seems to do less than its share. Eor illustration purposes, imagine that an idealized pair has a clean 50-50 split of the family-raising labor, then the male starts gold-bricking by cutting his share down to 30% of the total, leaving a 20% shortfall. The model reasons the female should not compensate fully (by making up the whole difference), but take up only part of the slack. If the system is working properly, the male’s interest in offspring welfare should prompt him to increase his contributions (whereupon she decreases, etc.) until equilibrium occurs. Eield tests of the Negotiation Model had begun to appear in print as we were starting our sparrow project. The basic experimental approach involves unilateral handicapping, that is imposing a burden on one partner to reduce its contributions and simulating “gold-bricking.” With one parent slowed, the unencumbered partner’s behavior is monitored to see if it provides partial compensa- tion (vs. full compensation or even no compen- sation, etc.). We followed the example of Wright and Cuthill (1989) in crimping four small lead split-shot fishing weights onto the rectrices of the handicapped parent (male or female), loading —5-7% of total body mass away from its center of gravity. After finding that some birds in the first season had lost these weights, we painted them with red nail polish in later years, enabling us to confirm retention through spotting scopes. Limiting our study pairs to those with clutches of either four or five eggs, we rotated among the experimental and control treatments (minimizing season effects). Our routine was to sample pre- treatment provisioning behavior for a total of —5 hrs prior to treatment (typically by day 5 post- hatching) and to collect another 9-10 hrs of data after treatment. Females in control (unmanipulated) pairs typ- ically increased provisioning throughout the nestling period, while males leveled off after about the fifth day (Mock et al. 2005). Our experimental treatments showed unexpected re- sponses. When males were handicapped, they had only a slight decrease in mean delivery rate, but they reduced their e-prey deliveries significantly; their mates showed a non-significant increase. More unexpectedly, at the nests where females received tail-weights, burdened females showed little reduction in their food deliveries, but their partners substantially increased provisioning any- way. Consequently, the broods received consid- erably more food than usual. Thus, if the positive male response was to his mate’s behavior at all. it could not have been because she was doing less. More generally, we found little evidence that parents are responding to each other's provision- ing rate (as required by the Negotiation Model), leading to the conclusion that parental effort is tied more directly to other factors (perhaps brood condition) and is more like a “sealed bid” with independent parental decisions (Schwagmeyer et al. 2002). We also handicapped male provisioning per- formance with a second experimental manipula- tion, this time by surgically inserting small tubes containing time-release testosterone beneath the scapular skin, implanting empty tubes for con- trols. This produced a more dramatic reduction in male provisioning, as well as a reduction in male 214 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2. June 2010 incubation time (62% lower than controls for both measures), but a significant female com- pensatory increase was observed only during incubation. At that stage of the cycle, testoster- one-dosed males appeared to reallocate time and effort to non-parental activities, as they did more singing and behaved more aggressively (both to other males and to their own partners) than control males. These hormone results showed that male and female workloads are strongly inter-related when offspring are still inside eggs, but not after hatching (Schwagmeyer et al. 2005). One interpretation of this difference considers that both parents are likely too busy when seeking and delivering prey items to keep track of each other’s contributions (their independent visits to the brood typically last only 1-5 sec and encounter each other only by chance). In contrast, most incubation changeovers are face- to-face and the other changeovers are likely to carry information about when the partner depart- ed (from the lingering tell-tale warmth of the eggs). We are currently exploring the behavioral dynamics of shared incubation in a new series of experiments. WHEN PARENTING COSTS ARE REDUCED Parents detecting their offspring are developing ahead of schedule might be expected to reduce their costly provisioning activities, either to use the time and effort for immediate reproductive advantage (especially males that can fertilize additional females) or for maintaining higher personal condition to benefit future offspring. The approach typically used for such load- lightening studies is to make food readily available to the experimental pairs (e.g., a feeder near the nest) and simply assume that a nontrivial amount of the extra nutrients is passed along to the offspring. We took a less trusting route by hand-feeding nestlings with commercial baby bird food twice a day. Specifically, we suspended nutrient powder in warm water to form a thin gruel and squirted it directly down the esophagus of each brood member through a modified syringe. Supplemented broods thus received an extra 25-30% of their estimated Daily Energy Budget, over and above whatever their parents supplied. We sampled parental provisioning as before, averaging ~I5 hrs at each of the 23 experimental broods. Parents did not abridge their food deliveries to these supplemented nests (Mock et al. 2005). Instead, females showed their usual provisioning increase after day 5, but this time their males escalated in lock step, outperforming control males by 25%. This male rate increase did not come at the expense of food quality (e.g., the proportion of e-prey in the diet remained con- stant). Subsidized nestlings, despite this extra nutrition, neither gained weight more rapidly nor achieved higher mass by banding age, but they recruited into the local breeding population at a somewhat higher rate (11.8%) than matched controls (4.9%). Thus, it appears males do not automatically divert their attention away from the current brood when parental burdens are lessened, but may respond positively to stronger offspring. Some data on begging call intensities showed that our supplemented broods begged more loudly, sug- gesting a possible proximate cue underlying paternal responses, but we found no evidence these broods differed visibly from normal (i.e., no mass effects). The asymmetrical responsiveness of the parents to subsidized broods may stem primarily from females already operating near their maximum (a ceiling effect), while males have extra reserves that normally remain unspent, or are channeled elsewhere. EULL CIRCLE We close this discussion of our research program where we started with explicit appreci- ation of our debt to Margaret Morse Nice for demonstrating the value of color bands. The huge amount of time and effort we have invested over the years in building and maintaining study populations of recognizable individual span'ows is a direct rellection on its central role in our work. Of course, numbered metal bands opened many research areas, but the advantage of being able to identify individuals without re-capture and from a distance truly revolutionized behavioral ornithology. Every project described would have been either totally impossible or substantially weaker without that preliminary step. One cannot even be certain that a given subject has not been Li.sed already in either the same study (with loss of statistical independence) or in a previous exper- imental manipulation (with the risk of introducing bias from that earlier treatment) without known individuals. One small celluloid band placed on one special sparrow in 1928 gave so much to so many. Mock am! Scinvai'oicycr • NOT THE NICE SI^ARROW 215 ACKNOWLEDGMENTS Our research on sparrows has been funded primarily by the National Science Foundation (1BN-94()8148. IBN- 9982661, and lOS-0843673). LITERATURE CITED Campbell, B. (Editor). 1972. Sexual selection and the descent of man, 1871-1971. Aldine Atherton, Chi- cago, Illinois, USA. Chase, I. D. 1980. Cooperative and noncooperative behavior in animals. American Naturalist 115: 827- 857. Clutton-Brock, T. H. 1991. The evolution of parental care. Princeton University Press, Princeton, New Jersey, USA. Dawson, W. L. 1903. Birds of Ohio. Wheaton Publishing Company, Columbus, Ohio, USA Edly-Wright, C., P. L. Schwagmeyer, P. G. Parker, AND D. W. Mock. 2007. Genetic similarity of mates, offspring health and extrapair fertilization in House Sparrows. Animal Behaviour 73: 367-378. Foltz, D. W. and P. L. Schwagmeyer. 1989. Sperm competition in the thirteen-lined ground squirrel: differential fertilization success under field conditions. American Naturalist 133: 257-265. Hamilton, W. D. 1964a. The genetical evolution of social behaviour. Journal of Theoretical Biology 7:1-16. Hamilton, W. D. 1964b. The genetical evolution of social behaviour. Journal of Theoretical Biology 7:17-52. Hill, G. E. 1992. An inexpensive source of colored leg bands. Journal of Field Ornithology 63: 408-410. Houston, A. I. and N. B. Davies, 1985. The evolution of co-operation and life history in the Dunnock, Prunella modiilaris. Pages 471—487 in Behavioral ecology: ecological consequences of adaptive behaviour (R. M. Sibly and R. H. Smith, Editors). Blackwell, Oxford, United Kingdom. Lack, D. 1947. The significance of clutch-size. Parts 1 and 2. Ibis 89:302-352. Lack, D. 1954. The natural regulation of animal numbers. Clarendon Press, Oxford, United Kingdom. Lack, D. 1968. Ecological adaptations for breeding in birds. Methuen Press, London, United Kingdom. MacColl, A. D. C. AND B. J. Hatchwell. 2003. Determinants of lifetime reproductive success in a cooperative breeder, the Long-tailed Tit Aegithalos caudatus. Journal of Animal Ecology 73: 1 137-1 148. Magrath, R. D. 1991. Nestling weight and juvenile survival in the Blackbird, Tiirdus merula. Journal ol Animal Ecology 60:335-351. Maynard Smith, J. 1977. Parental inve.stment: a prospec- tive analysis. Animal Behaviour 25:1-9. Maynard Smith, J. 1982. Evolution and the theory ot games. Cambridge University Press, Cambridge, United Kingdom. Maynard Smith, J. and G. A. Parker. 1976. The logic ot asymmetric contests. Animal Behaviour 24:159-175. McNamara, J. M., C. E. Ga.sson, and A. I. Houston. 1999. Incorporating rules for responding into evolu- tionary games. Nature 401:368-371. Mock. D. W. 1984. Siblicidal aggression and resource monopolization in birds. Science 225:731-733. Mock, D.W. 1985. Siblicidal brood reduction: the prey-size hypothesis. American Naturalist 125:327-343. Mock, D. W. and T. C. Lamey. 1991. The role of brood size in regulating egret sibling aggression. American Naturalist 138: 1015-1026. Mock, D. W. and G. A. Parker. 1986. Advantages and disadvantages of ardeid brood reduction. Evolution 40: 459-470. Mock, D. W. and G. A. Parker. 1997. The evolution of sibling rivalry. Oxford University Press, Oxford, United Kingdom. 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., P. L. Schwagmeyer, and J. A. Gieg. 1997. A trap design for capturing individual birds at the nest. Journal of Field Ornithology 70:276-282. Mock, D. W., P. L. Schwagmeyer, and G. A. Parker. 2005. Male House Sparrows deliver more food to experimentally subsidized offspring. Animal Behav- iour 70: 225-236. Mock, D. W., P. L. Schwagmeyer, and M. B. Dugas. 2009. Parental provisioning and nestling mortality in House SpaiTows. Animal Behaviour 78:677-684. Nice, M. M. 1979. Research is a passion with me. Consolidated Amethyst Communications Inc., Toron- to, Ontario. Canada. O'Connor. R. J. 1978. Brood reduction in birds: selection for infanticide, fratricide, and suicide? Animal Behav- iour 26: 79-96. Parker, G. A. 1970. Sperm competition and its evolution- ary consequences in the insects. Biological Review 45: 525-567. Parker, G. A. 1978. Selfish genes, evolutionary games, and the adaptiveness of behaviour. Nature 274: 849- 855. Parker, G. A. 1982. Why are there so many tiny sperm? Sperm competition and the maintenance of two sexes. Journal of Theoretical Biology 96: 281-294. Parker, G. A. 1983. Arms races in evolution: an ESS to the opponent-independent costs game. Journal ot Theo- retical Biological 101: 619-648. Parker, G. A. 1985. Models of parent-offspring conHict. V. Effects of the behaviour of the two parents. Animal Behaviour 33: 519-533. Parker, G. A. 2006. Behavioural ecology: the science of natural history. Pages 23-56 in Essays on animal behaviour: celebrating 50 years of animal behaviour (J. R. Lucas, and L. W. Simmons. Editors). Elsevier, Burlington. Massachusetts, USA. Parker, G. A. and M. R. Macnair. 1978. Models of parent-offspring conllict. 1. Monogamy. Animal Be- haviour 26: 97-1 1 1 . Parker, G. A. and .1. Maynard Smith. 1990. Optimality theory in evolutionary biology. Nature 348: 27-33. Parker, G. A. and R. A. Stuart, 1976. Animal behaviour as a strategy optimizer: evolution of resource assess- 216 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2. June 2010 merit strategies and optimal emigration thresholds. American Naturalist 110; 1055-1076. Parker, G. A., D. W. Mock, and T. C. Lamey. 1989. How selfish should stronger sibs be? American Naturalist 133: 846-868. Ploger, B. J. 1997. Does brood reduction provide nestling survivors a food bonus? Animal Behavior 54:1063- 1076. SCHWAGMEYER, P. L. 1988. Scramble competition polyg- yny in a social mammal: male mobility and mating success. American Naturalist 131: 885-892. SCHWAGMEYER, P. L. AND D. W. FoLTZ. 1990. Factors affecting the outcome of sperm competition in thirteen-lined ground squirrels. Animal Behaviour 39: 156-162. SCHWAGMEYER, P. L. AND D. W. MOCK. 1997. How to minimize sample sizes while preserving statistical power. Animal Behaviour 54: 470^74. SCHWAGMEYER, P. L. AND D. W. MocK. 2008. Parental provisioning and offspring fitness: size matters. Animal Behaviour 75: 291-298. SCHWAGMEYER, P. L. AND G. A. PARKER. 1987. Queuing for mates in thirteen-lined ground squirrels. Animal Behaviour 44: 1015-1025. SCHWAGMEYER, P. L. AND G. A. PARKER. 1990. Male mate choice as predicted by sperm competition in thirteen- lined ground squirrels. Nature 348: 62-64. SCHWAGMEYER, P. L., D. W. MOCK, AND G. A. PARKER. 2002. Biparental care in House Sparrows: negotiation or sealed bid? Behavioral Ecology 13: 713-721. SCHWAGMEYER, P. L., H. G. SCHWABL, AND D. W. MOCK. 2005. Dynamics of biparental care in House Sparrows: hormonal manipulations of parental contributions. Animal Behaviour 69: 481^88. SCHWAGMEYER, P. L., D. W. MOCK, T. C. LaMEY, C. S. Lamey, and M. D. Beecher. 1991. Effects of sibling contact on hatch timing in an asynchronously hatching bird. Animal Behaviour 41: 887-894. Trivers, R. L. 1972. Parental investment and sexual selection. Pages 136-179 in Sexual selection and the descent of man, 1871-1971 (B. Campbell, Editor). Aldine Atherton, Chicago, Illinois, USA. Whitekiller, R. R., D. F. Westneat, P. L. Schwagmeyer, and D. W. Mock. 2000. Badge size and extra-pair fertilization in the House Sparrow. Condor 102:342-348 Wright, J. and 1. Cuthill. 1989. Manipulation of sex differences in parental care. Behavioral Ecology and Sociobiology 25: 171-181. The Wilson Journal of Ornithology 122(2):217 -227, 2010 THE POSTBREEDING MIGRATION OE EARED GREBES JOSEPH R. JEHL JR.'-' AND ANNETTE E. HENRY- ABSTRACT. — Eared Grebes {Podiceps nii>ricollis) in autumn make a postbreeding/molt migration from breeding areas in western North America to hypersaline lakes in the Great Basin. We studied their biology in 2001-2006 during this phase of the annual migration near Green River, Wyoming, USA where migrants en route to Great Salt Lake, Utah land on industrial ponds. Most evidently originate in the Prairie Pothole Region of North Dakota. The main movement extends from late July to mid-October. Migrants arrive almost daily with the cumulative percentage of transiting birds increasing by about 1% per day. Adult males and females migrate on the same schedule and precede juveniles by 2-3 weeks. Annual differences in phenology, abundance, age ratio, and wing molt vary with availability of wetland habitats in the main source area. Data on mass, body composition, energetics, and stomach contents indicate a typical flight involves a direct 2-3 day non-feeding migration, which is accomplished at night. Grebes are quiescent during the day and do not resume their migration until 45 min after sunset. We documented two undescribed vocalizations, a short-range contact note and one associated with departure. The possibility that Eared Grebe productivity, as inferred from studies of migrants through Wyoming, can provide insight into the status of waterbirds in the source area is worth further investigation. Received 16 March 2009. Accepted 6 December 2009. The autumnal migration of the Eared Grebe (Podiceps nigricoUis) in North America occurs in two phases. Promptly after the breeding season adults and young undertake a postbreeding migration from nesting areas in the interior to molting/staging sites at hypersaline lakes in the Great Basin (Fig. 1; Storer and Jehl 1985). Nearly the entire North American population stages at Mono Lake, California and Great Salt Lake, Utah from late summer through early winter and then undergoes a poststaging migration to wintering areas in Mexico and southern California (Jehl 1988, Cullen et al. 1999). Major details of grebe biology during the nonbreeding season are reasonably well estab- lished (Cullen et al. 1999; Jehl and McKernan 2002; Jehl et al. 2002, 2003; Ellis and Jehl 2003; Jehl 2007). Most grebes undertake a 800-1,500 km flight, which apparently occurs without feeding, to reach staging lakes (Jehl and Yochem 1986, Boyd et al. 2000). Few can make the trip on a single flight because migration occurs only at night. Migrants passing through western Wyo- ming toward Great Salt Lake must cross large spans of steppe desert, where rest stops are few and unreliable. Near the town of Green River some encounter an apparent oasis in the form ol large (up to 400 ha) and sterile evaporation ponds associated with mining a hydrated form of soda ash (Na2C03) known as trona (Na2C03/ ‘ Smithsonian Ornithology, U.S. National Museum ot Natural History, Washington, D.C. 20560, USA. ^NOAA Fisheries, Southwe.st Fisheries Science Center, 8604 La Jolla Shores Drive, La Jolla, CA 92037, USA. ^Corresponding author; e-mail: grebe5k@verizon.net NaHC03-2H20). Sodium salts in the supersatu- rated ponds can encrust feathers and lead to mortality. The trona industry has developed a program to reduce potential losses by monitoring the migration, and capturing and rehabilitating afflicted birds through the entire migration season. We took advantage of this exceptional sample to document Eared Grebe biology and behavior during a phase of the annual cycle that had been largely ignored. The data also generated basic information on age ratios, timing and pattern of migration, molt patterns, and annual patterns of production that are essential for understanding grebe biology but which, for logistical and sampling reasons, cannot be obtained at the major fall staging areas (Jehl and Johansson 2002). METHODS We studied grebe migration from 2001 through 2006 at the EMC West Vaco and OCI trona ash plants, which are ~15 km apart and 30 km northwest of Green River and 220 km east- northea.st of Great Salt Lake (Fig. 1). Each year we visited the sites early in the migration season and instructed staff on Eared Grebe natural history and procedures for collecting data. The staff inspect the ponds several times each day starting about 0700 hrs over the entire fall migration season, from late June through November, and record all species of waterbirds encountered (captured, not captured, found dead). Encrusted birds, predominantly grebes, are captured as quickly as possible, transported to a nearby (1- 2 km) treatment facility, washed in warm water to remove external salts, and allowed to rest and dry 217 218 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 2. June 2010 FIG. 1. The major breeding distribution of the Eared Grebe ir 1999). Great Salt Lake, Utah (2) and Mono Lake, California, (3) passing through Green River, Wyoming (I) probably originate i Oregon (4) is evidently used by some migrants en route to Natural History. in a large fre.shwater tank. In mid-afternoon they are released locally along the Green River (fresh water). Staff collected data on age, wing molt, and other characteristics as time permitted. Body mass was measured in early afternoon, after the birds had been cleaned of salt deposits and thoroughly dried. Age was recorded as adult (AHY = after hatch year) or Juvenile (HY = hatch year) based on iris color; crimson in adults, brown to orange- brown in juveniles. Birds in their .second year (SY North America (outlined in black; from Cullen et al. hold >1 million staging birds in fall. Most migrants 1 eastern North Dakota (cross-hatched). Abert Lake, staging lakes. Map by Dan Cole, U.S. National Mu.seum of = hatched in the previous year) have irides that are orange or pale red and would be classified as AHY (Storer and Jehl 1985). Adult males and females were separated by bill length: ^23.5 mm = male, <23.5 mm = female, a procedure that is >90% reliable (Jehl et al. 1998). We made ho consistent effort to sex juveniles because bill growth may not be completed by onset of migration. Grebes were weighed on a digital scale (±1.0 g) after they were cleaned and dried, typically 4-6 hrs after capture (1400 hrs). We Jeh! ami Henry • POSTBREEDING MIGRATION OE EARED CJREBES 219 TABLE 1. Eared Grebe migration schedule by age class, 2001-2006, based on birds captured or lound dead during autumn migration in southwestern Wyoming. AdulLs Juveniles Year Migration span Day.s 1S% complete Migration span Days 75% complete 2001 12 Jul-16 Oct 97 1 Sep 14 Aug-6 Nov 85 22 Sep 2002 5 Jul-12 Oct 100 7 Aug 10 Aug- 16 Oct 68 25 Sep 2003 23 Jiin-28 Sep 98 3 Sep 19 Aug-24 Oct 67 10 Oct 2004 6 Jul-24 Sep 81 19 Aug 20 Aug-1 Nov 84 23 Sep 2005 12 Jul-30 Sep 81 6 Sep 10-Aug-ll Nov 94 22 Sep 2006" Combined migrulion Span Days l2Jul-6Nov 118 5Jul-16 0ct 104 23 Jun-24 Oct 124 6Jul-01Nov 119 12Jul-llNov 123 18Jul-15 0ct 90 " No data by age class. used color of the primaries, brown in unmolted grebes, black in newly-molted, to learn whether wing molt had been completed before grebes left breeding locations. We observed grebe behavior on the evaporation ponds and in the treatment facility on >40 days (2001-2005), and rehabilitated birds from release through departure on over 30 occasions. We recorded vocalizations in the laboratory and in the field using a Radio Shack Unidirectional micro- phone (frequency response 0.01-10 kHz), record- ed sounds directly to computer hard drive, and analyzed them using ISHMAEL (Mellinger 2001). Body composition and sex identification of birds found freshly dead (<24 hrs) were ascertained by dissection (Jehl 1997). Fat content of migrants was calculated as: fat (g) = 0.76 body mass (g) ~ 163.54 g (H. 1. Ellis and JRJ, unpubl. data). The source of migrants passing through Green River can only be inferred. Breeding grebes are scarce in eastern Wyoming and across the High Plains. Only a few hundred pairs breed in South Dakota (Drilling 2007), and several thousand in western Minnesota (Boe 1993). Grebes departing Great Salt Lake in early winter fly in a straight line toward wintering areas (JRJ and S. A. Gauthreaux, unpubl. data), and we assume post- breeding migrants behave similarly. Thus, we infer that most Green River migrants originate in the Prairie Pothole Region of North Dakota, where tens of thousands breed (Stewart 1975, Igl et al. 1999). Those originating farther north or south would likely by-pass the trona ponds. The distance from the eastern Dakotas is > 1 ,000 km, requiring a flight of at least 2 days. We examined synoptic weather maps from the northcentral U.S. 2 days prior to 19 large arrivals (36-369 individuals) to investigate if weather influenced the grebes’ decision to migrate. We used a paired t-test to compare the number of ponds in the northcentral U.S. to the number of grebes encountered near Green River. RESULTS Phenology. — ^Eared Grebes usually arrived be- tween dawn and 1000 hrs. This species avoids flowing water (JRJ and AEH, unpubl. data) and only rarely did grebes land on the nearby Green River, which is only 1 km from some ponds (Julie Lutz and Lan-y Cherny, pers. comm.). Phenology was similar at both trona plants in all years and the data were combined. The earliest arrivals appeared in late June (Table 1). Small numbers (1-15) arrived almost daily (Fig. 2A) during the main influx, 1 August-2 October; the cumulative increase averaged 1.3% per day (Fig. 2B). An-ivals slowed to 0.5% per day from 3 to 15 October before ceasing in early November. Yearly patterns were similar despite an occasional pulse of >100 birds or variation in the total encounters, and paralleled those documented in 1992-1995 (Jehl and Johansson 2002). Age and Se.x Ratios. — Adults migrated earlier than juveniles (Fig. 3), 75% by the first days of September (Table 1). Those arriving before 1 August were probably nonbreeders or failed breeders. Juveniles began to appear in mid- August and, in most years, 75% passed through by 25 September. The sex ratio of adults (2001. 2002. 2005; n = 396 males, 380 females) did not differ from 1:1 overall or in any individual year. More juveniles were classilied as females than males (2001, 28:15; 2002, 67:58; no data for other years), probably because growth had not finished, inflating the apparent proportion of the smaller females. Wing Molt. — Data on wing molt were obtained in 2001 and 2002. The pattern was similar in both 220 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2. June 20 JO c QJ O FIG. 2. Phenology of Eared Grebe autumn migration in .southwestern Wyoming, 2001-2005. A. Individual years; 2001-2005, n = total grebes recorded at FMC and OCI plants. B. All years combined. years. The first migrants with new remiges arrived in mid-August and the frequency increased as the season progressed (Fig. 4); by 20 September all passage adults had molted. The frequency of molt in breeding areas was greater in 2001 (21% of 206 adults) than in 2002 (7% of 425). Body Mass, Body Composition, and Stomach Contents. — Mass did not vary from year to year. The mean (± SD) was 297 ± 32 g for adults with males averaging 7.5% heavier than females, and 295 ± 35 g for juveniles (Table 2). There was no consistent or significant pattern by males and females, age, or date in the 3 years for which we have adequate data. Carcasses showed the en- larged flight muscles and reduced digestive organs characteristic of birds that had reconfig- ured body composition prior to migrating. Body and component masses were greatly reduced in comparison to migrants elsewhere known to have been in migration for only one night (Table 3). Gizzards lacked a pyloric plug and were nearly devoid of digestible food; the weight of non- feather debris averaged about 0.4-0. 5 g, most of which was finely ground and unidentifiable. Eight of 10 gizzards examined in detail in 2006 contained unidentified insect material — mostly Jch! and Henry • POSTBREEDING MIGRATION Ol' EARED (IREBES 221 100 80 -I 60 40 - 20 - 0 ♦ AHY=222 oHY=199 ♦ ♦ 2001 U) FIG. 3. Arrival schedule of adult (AHY) and juvenile (HY) Eared Grebes in southwestern Wyoming. 2001-2005. fragments of exoskeleton from surface-dwelling aquatic beetles and alkali flies (Ephydra spp.); two contained brine shrimp {Anemia spp.) eggs, four-miscellaneous seeds, three — fragmentary plant material, three — jaws of nereid worms (/? = 3, 6, and 200-300), and two— small bits of grit or sand. Eight also contained feathers, which were extremely worn in five cases, but freshly ingested (1-2 days) in three others. Behavior Diiriny the Day. — Grebes that landed on the trona ponds remained largely inactive throughout the day, as is characteristic of birds stopping over in natural situations. They did not attempt to drink or feed, nor did they dive, unless pursued for capture. Typical of grebes, they preened continually. If salt crystals began to form, they increased preening efforts, pulling several feathers through the bill and using vigorous head-shaking to flick off adhering crystals. As encrustation increased they moved closer to shore, often attempting to haul out. Unimpaired birds remained on the ponds and evidently departed successfully after dark. There was little variation in behavior of rehabilitated birds upon release. They swam directly to the middle of the Green River and engaged in comfort movements (bathing, stretch- ing, preening, adjusting plumage, drinking) for up to a half-hour; some also made shallow and brief (8-15 sec) dives, evidently to readjust plumage. Subsequently they began to investigate their suiToundings, swimming up- or downstream until they encountered rapids or other bamers and then reversed course. They then moved into secluded sites near shore, out of any cuiTent, and rested or slept until evening. Departure. — Sunset is a key time. Grebes began to become active a few minutes after the sun dipped below the horizon, when they returned to mid-river, and began predeparture activities. These included ritualized preening of the back and scapulars, one or two short practice flights (always upstream), and group diving, submerging, and surfacing in near-unison. As dusk deepened the birds became quiescent for several minutes. Then, 40-45 min after sunset, when it was almost completely dark, they performed a few intention movements, consi.sting of pumping the head up and down and giving soft calls. Then, with no further signals, they flew upstream ~1 m above the river for about 100 m or so before climbing and disappearing westward. Small groups left as a unit but, when relatively large numbers (>15) were present, llighls could be spread over 3- 10 min. Any that did not leave swam back to shallow water and rested without feeding until the next sunset. We noted zugunruhe in captive grebes twice. Six adults captured before noon on 17 August 222 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 2, June 2010 (/) "O 'o FIG. 4. Timing of remige molt of Eared Grebes migrating though southwestern Wyoming, 2001 and 2002. Data presented in weekly intervals. 2001 were retained overnight for metabolic studies. They became agitated at 2100 hrs (46 min after sunset), despite having had no view of the sky or access to natural light for 9 hrs, started to vocalize, and swim actively for about 30 min (H. I. Ellis, pers. comm.). Other birds released earlier that day departed between 2054 and 2112 hrs. In another case, birds in a dark box being transported to a release site in the early evening suddenly became restless and started calling 27 min before sunset. This behavior was not observed when birds were released earlier in the day (Julie Lutz, pers. comm.). Vocalizations. — ^Two vocalizations were record- ed (Fig. 5). Both were compound sounds, made up of more than one unit and within the frequency range (1-5 kHz) of calls associated with breeding, but their structures were distinct. The “A” call consisted of three frequency-modulated down- swept vocalizations in a 1.2 to 1.5 kHz bandwidth. We heard it as grebes were preparing to depart and immediately before the actual takeoff. The “B” call, reminiscent of a soft amphibian-like grunt, consisted of a patterned series of 5-1 1 individual pulses with a consistent inteipulse interval of 0.025 sec. Each series was repeated TABLE 2. Mass (g), by males southwestern Wyoming, 2001-2005. and females, and age of migrating Eared Grebes captured in autumn migration. Adult niale.s Adult females All adults Juveniles Year n Range (.v ± SD) n Range (jf ± SD) n Range (.f ± SD) n Range (.f ± SD) 2001 83 249^09 (305 ± 29) 75 210-345 (277 ± 23) 158 210-409 (292 ± 30) 191 197-388 (299 ± 35) 2002 223 223-389 (305 ± 30) 210 200-394 (285 ± 30) 424 200-394 (296 ± 31) 125 222-372 (287 ± 31) 2003“ 10 228-332 (288 ± 31) 31 218-360 (291 ± 35) 2004“ 29 225-367 (304 ± 33) 25 197-348 (290 ± 35) 2005 68 270-396 (316 ± 28) 79 217-390 (297 ± 32) 147 217-396 (306 ± 32) 162 213-398 (297 ± 36) All years 374 223^09 (307 ± 30) 364 200-394 (286 ± 30) 768 200^09 (297 ± 32) 534 197-398 (295 ± 35) No duty by categories of males and females. Jeh! and Henry • POSTliRIIHDINCJ M ICJRA I'lON ()l liARED GRERES 223 TABLE 3. Body composition (g) August-September 200 1-2006, and of Eared Grebes found freshly dead at the end of a migration: Green River, Wyoming, Iron Mountain, California, .lanuary 1993 (from .lehl 1997: Table 1). Green River, Wyttming Iron Mountain. California Description n Riinge (X ± St)) n R:inge (X ± SD) Mass 25 240-330 (280.2 ± 24.0) 7 312423 (368 ±41.4) Breast 24 16.0-23.2 (19.7 ± 1.9) 7 21.7-27.0 (24.0 ± 1.9) Heart 26 2.7-t.7 (3.6 ± 0.4) 7 4.5-5.4 (5.0 ± 0.3) Intestine 10 7.5-13.6 (9.7 ± 1.7) 7 7.0-12.4 (10.2 ± 1.9) Leg 10 1 1.1-14.4 (13.1 ± 1.2) 7 14.7-20-1 (17.1 ± 2.0) Liver 26 6.1-1 1.8 (8.8 ± 1.2) 7 7.8-13.7 (10.6 ± 1.9) Stomach 26 6.6-15.5 (10.6 ± 2.1) 7 9.4-13.3 (1 1.2 ± 1.3) Stomach contents 26 0. 1-5.6 (1.2 ± 1.2) 7 1. 6-7.0 (2.7 ± 1.8) Estimated time in flight 36-60 hrs 9.3 hrs Fat stores 20.6-86.4 (49.7 ± 17.7) 72.6-151.4 (1 12.9 ± 30.6) five times within a call. The interval between the first and second series was 3.0 sec, while the interval between the 2nd, 3rd, 4th, and 3th series was 2.0 sec. The fundamental peak frequency approximated 1.5 kHz with harmonics. More than 50 calls were recorded in 10 min from a tank holding 28 individuals. Weather and Migration. — ^The main source of migrants at Green River is probably the Prairie Pothole Region of North Dakota. Synoptic weather maps showed calm conditions prevailed there at dusk 2 days prior to all major arrivals at Green River. In all cases, winds were <5 km/hr, generally from the NE-SE; in 15 cases a cold front had passed 1-2 days earlier. DISCUSSION Eared Grebes migrating through Wyoming en route to fall .staging areas exhibited behaviors that had not been documented elsewhere, including energy-saving sluggishness by day, ziigunriihe, strict correspondence of departure with onset ot near-total darkness, and unique vocalizations. Our finding that adults migrated earlier than juveniles had been established in breeding areas and at Mono Lake (Jehl 1988, Cullen et al. 1999), but the pattern lacked detailed quantification. That adult males and females migrate on the same schedule was not unexpected, because adults split the brood, each providing a similar level of care for about 2 weeks, and then leave belore the young can fly (Cullen 1998). Some adults were known to complete wing molt in breeding areas (Storer and Jehl 1985, Jehl 1988). The Wyoming data, supplemented by >5,000 individuals banded cit Mono Lake over two decades (JRJ, unpubl. data), indicate that about 11% of adults molt before migrating, and that frequency of molt increases as the season progresses and is greater in years when grebes remain longer in breeding areas. Eared Grebes are generally silent during migration. The two undescribed calls we recorded were within the frequency range (1-5 kHz) of calls associated with breeding, but their structures were distinct (spectrograms in Cullen et al. 1999). We consider the “B” call a short-range contact note. The “A” call was associated with departure. Eared Grebes migrate in the dark, depart staging areas in flocks, and are susceptible to mid-air collisions (Jehl 1998). Whether vocalizations or other sounds are used to minimize that risk remains to be investigated. From Breeding Areas to Staging Lakes: Evi- dence for a Direct, Non-feeding Flight. — ^Eared Grebes reorganize their body composition before departing staging and breeding areas (Jehl 1997). Departure mass approximates 400-450 g, which decreases to 350-370 g after an overnight flight (Jehl 1993, 1994, 1997, 1998). Wyoming migrants averaged —300 g, >50 g lighter, indicating they had been en route for at least 2 days. Their stomachs and intestines were small and essentially empty, confirming a nonfeeding migration and indicating they had ceased feeding and cleared their gut several days before departing (Jehl 1997). Meteorological data support earlier find- ings (Jehl and Johansson 2002) that grebes depart breeding areas in periods of calm weather. There is no time for a leisurely migration, as south- bound adults begin to pass through Green River as early as late July-the same time that departures are noted in breeding areas and migrants appear at the major staging lakes (Jehl 1988, Cullen et al. 224 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 2, June 2010 ^ I I i I i r 0.1 0,2 0.3 0.4 Time (s) FIG. 5. Vocalizations of migrating Eared Grebes, Green River, Wyoming, 1 5 September 2005. A. Type “A” call (sample rate 20 kHz, 512 FFT). B. Type “B” call (.sample rate 20 kHz, 512 FFT) showing the entire call, which consists of five series of pulses. C. Fig. 5B expanded to show individual pulses. 1999). They do not feed during this movement, and migrants leave stopover locations the evening after they arrive. Thus, weather conditions appear to have little effect once the decision to migrate has been made. Energetic.s. — Migrating across arid lands is a challenge (Jehl 1993, 1994; Cullen et al. 1999; Jehl et al. 1999, 2003). Grebes leaving Green River after rehabilitation averaged about 293 g and would be expected to reach Great Salt Lake weighing 271 g with 41 g of fat remaining (H. I. Ellis and JRJ, unpubl. data). This estimate is based on flight costs of 6 g fat/hr (vs. 7. 1 g/hr for 400^50 g migrants; Jehl 1994) and a flight time of 3.7 hrs at 60 km/hr (Cullen et al. 1999). The minimal weight for healthy adults is ~240 g (Jehl 1988, Anderson et al. 2007; JRJ, unpubl. data), and migrants need to depart Green River at ^262 g, which includes 85% of our sample (Fig. 6). Laying over 1 day would require catabolizing at least an additional 17 g of fat (0.7 g/hr resting rate X 24 hrs; JRJ and AEH, unpubl. data). Those delayed for 2 days would deplete nearly all remaining fat and require an additional stop to rebuild the digestive organs and refuel before continuing (Hume and Biebach 1996). The rate of weight gain in staging grebes is 3-4 g/day (Jehl 1988). Consequently, migrants originating more than a 3-night flight from the staging areas (—1,500-1,600 km) probably re- quire a stopover of several weeks or more to accumulate sufficient fat stores to continue. The only present candidate locality for a regular stopover is saline Lake Abert, Oregon where 5,000-7,000 migrants have been reported in September and early October (Jehl 1988). Annual Differences. — ^Timing and pattern of migration differed from year to year. Adults and young in 2001 passed through steadily in their respective periods; 75% of adults migrated by 1 September, 21% of adults underwent wing molt before migrating, and juveniles comprised 47% of the migrants. In 2002, 75% of adults migrated 3 weeks earlier, only 7% had molted, and juveniles were few (23%) and concentrated in the last half of September. We consider 2001 a year of successful breeding, whereas in 2002 most adults left too early to have bred successfully, and the relatively few young evidently resulted from late nesting. These differences are probably indicative of habitat loss resulting from the eastward shift into the Dakotas of the severe drought that had beset the western U.S. since the late I99()s (U.S. Drought Monitor maps for 29 May 2001 and 28 May 2002, www.drought.unl. edu/dm/archive.html). Our view that North Da- kota is a main source of Green River migrants is supported by the significant relationship (( = 4.33, P < 0.001) between the number of grebes encountered near Green River in autumn and the Jehl and Henry • POSTBREEDING MIGRATION OP EARED GREBES 225 FIG. 6. Mass of Eared Grebes (adults and juveniles combined) captured at Green River, Wyoming, 2001-2005. Arrow indicates estimated minimum mass needed to reach Great Salt Lake without laying over or making an additional stop. number of May ponds in the northcentral U.S. by year (Fig. 7). Patterns from other years are less clear. There was an early exodus of adults in 2003, yet nearly half remained in breeding areas into September; juveniles migrated much later than in other years, but comprised a high percentage (65%) of the population. Evidently conditions were poor early, but breeders able to take advantage of late- summer rains fared well, a pattern also inferred from studies of waterfowl (USDl 2003). Data in 2004 were too few for analysis. There was a steady passage of adults in 2005 when juveniles were plentiful (52%), but concentrated late in the season. Are the Data Representative? — One might question whether a sample from industrial ponds fairly represented the population passing through 3000 2500 C/D “D C 0 -§■ 2000 c CD C/D 1 1500 OD 05 1000 ,Q E 13 500 0 CM CO in CD t^ CXD 2 year-old individuals. These data sets were analyzed separately for differences in distance moved by yearlings and those dispersing from natal areas until they became established on breeding areas. The distance moved by male and female yearlings was combined as individuals could not be identified prior to performing courtship displays. We recorded the location of birds re-encoun- tered using Global Positioning System (GPS) receivers. The distance moved was calculated on 1:50,000 topographic maps as a straight-line between natal sites and first breeding territories for natal dispersers or post-fledging areas for juveniles. We estimated the azimuth from the nest location to the final known location for each dispersing ibis (final direction). We calculated the mean final direction (/i) and the angular deviation (s) around the mean final direction for each cohort, and used Rayleigh’s z to test for directional pattern within cohorts (Zar 1996). All circular analyses were conducted using Oriana for Windows (Version 2.0) (Oriana 2003). Non-parametric statistical treatments (SPSS for WINDOWS, Version 1 1.5) were used for tests if dispersal data was not normally distributed. We report means ± SD unless otherwise noted, and statistical tests were considered significant at F = 0.05. RESULTS Postfledging Dispersal of Young. — ^We followed 36 yearlings of the 304 nestlings banded including three radio-marked individuals from fledging through dispersal to areas used for breeding. All natal home ranges were close to paddy fields near human settlements at the edge of even-aged, mature oak-pine forests characterized by 25- 80 cm dbh, and canopy heights ranging from 15 to 33 m. Stands ranged in age from 25 to 183 years. One movement pattern was used by Crested Ibis family groups during postfledging dispersal in all cases. Movements of family groups prior to dispersal were centered on nest sites, and areas within the nest territory were visited repeatedly up to time of dispersal. The distance between family groups and their natal nest sites was 0.54 ± 0.67 km. Nest territories were occupied for an average of 21.1 ± 10.2 days (n - 36) from time of fledging (38-40 days of age) to dispersal. Family groups, composed of chicks and adults, moved progressively farther (Fig. 2) along the }•;/ ('/ ul. • CRESTED IBIS DISPERSAL 231 tributaries of Hanjiang River in a southerly direction with a mean direction of 189.2 ± 46.5° (/? = 36 ibis). Siblings of each brood increasingly become independent of their parents at the beginning of August and all young, as well as adults from different families, formed wander- ing flocks in post-dispersal areas characterized by a sweep of distant hills with many artificial ponds, reservoirs, and vast farmlands along the Hanjiang River. Mean dispersal distance was 20.3 ± 7.0 km (n = 36). Young took 51.4 ± 8.1 days (range = 40-64 days) to disperse from their natal area to their post-dispersal range, where most remained for nearly 2-3 years until they became sexually mature. Most immature individuals (2^ years of age) spent summers in wintering areas 15.5 ± 4.3 km in = 25) south (191.5 ± 33.6°, n = 36 ibis) of their first breeding sites. The average elevation of the roosting sites (560.0 ± 55.6 m, n = 15) in wintering areas was significantly lower than that in first breeding sites (786.5 ± 56.8 m, n = 25) (/ = 12.3, P = O.OOO). Natal Dispersal Distance. — Seventy-three (36 males and 37 females) of the 304 chicks banded from 1987 to 2002 were followed to their first reproductive attempt (i.e., were recorded with eggs or chicks) at 25 natal sites (Table 1 ). Males and females as first breeders nested at a mean distance of 5.9 ± 6.3 km (// = 36) and 9.6 ± 10.2 km (n — 37) from their natal sites, respectively. Females dispersed significantly farther than males (/ = 2. 1 4, P = 0.03) and the ranges of their dispersal distances were dissimilar with males dispersing 0.7 to 28.0 km and females 1 .0 to 46.2 km (Fig. 3). Natal dispersal distances of females by age groups were significantly greater than for males (Wil- coxon Test, Z = —2.394, P = 0.003; Table 2). Mean distance of females and males from natal areas to breeding tendtories decreased with age of first reproductive attempt (two-way ANOVA: F = 1 1.381, df = 4, F = 0.019). Direction of Natal Dispersal. — ^There was a significant directional pattern to natal dispersal for males and females, based on directions to final locations for individuals. Mean final direction was not random for both cohorts, and we pooled all cohorts for estimation of overall mean direction (// = 185.6°, s = 86.1°; Fig. 4). Most individuals dispersed southeast from natal sites, suggesting that dispersal movements of individual ibis were significantly concentrated around the mean direc- tion (Rayleigh’s ~; all P < 0.05). Age of First Breeding. — ^There were 73 birds for which age of first breeding was precisely known from observations of courtship displays of marked individuals. The youngest birds participating in breeding were 2 years of age. The oldest bird recorded as a first breeder was 7 years of age. The range of ages of first reproduction was 2 to 6 years for females and 2 to 7 years for males (Fig. 5). The mean age of these birds when first recorded breeding was 3.89 ± 1.40 years (/? = 73). First reproduction by females was not significantly earlier than males (females 3.64 ± 1.36 years, n = 37; males 4.17 ± 1.47 years, n = 36, t = 1.60, F = 0.1 1). 232 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2, June 2U 10 <5 <10 <15 <20 <25 <30 <35 <40 <45 <50 Dutance (km) FIG. 3. Distribution of natal dispersal distances for Crested Ibis (n = 36 males and 37 females). DISCUSSION Measuring dispersal in spatially heterogeneous environments is essential for testing population dynamics models (Kareiva 1990), but obtaining unbiased dispersal data is difficult, because long- distance dispersers often go undetected (Koenig et al. 2000). We were able to track 73 individual ibis until they settled and paired, and our study areas were sufficiently large (—3,000 km-) to allow reasonable quantification of long-distance dis- persal (up to 50 km. Fig. 1). Multi-year obser- vations allowed for large sample sizes for juveniles. Crested Ibises had a skewed dispersal pattern with many individuals dispersing relatively short distances, but with some dispersing much greater distances (Fig. 3). Breeding territories are typically small with a diameter of 50-100 m (Shi and Cao 2001) with several nests on the same tree (Liu et al. 2003). However, all marked individuals dispersed from their natal home ranges (100% dispersed >0.7 km). A similar pattern has been reported for other Ciconii- formes, including Snowy Egrets (Egretta thula). Black-crowned Night Herons {Nycticorax nycti- corax), and Forster’s (Sterna forsteri) and Caspian terns {S. caspia), in contrast to Great Blue Herons (Ardea herodias) which have a tendency to return to natal areas to breed as adults (Gill and Richard 1979). Gender-biased dispersal is common among many bird species with females usually dispersing farthest (Greenwood 1980). Females settled farther from natal sites than males in some Ciconiiformes such as White Storks {Ciconia ciconia) (Chernetsov et al. 2006), but it is most often the female that is more likely to leave the natal area and travel longer distances before settling to breed (Clarke et al. 1997). Our study strengthens the conclusions drawn by Greenwood (1980) regarding this broad, general pattern (Fig. 3). TABLE 2. Natal dispersal distances (km) in relation to age and gender of Crested Ibis. Puramclcr Age of first reproduction (.v ± SD) 2-yr .Vyr 4-yr .S-yr 6-yr 7-yr ' Males 1.3.2 ± 9.7 7.3 ± .3.1 3.9 ± 2.4 3.7 ± 2.3 2.0 ± 0.3 2.6 ± 2.2 (n = 4) (n = 12) (/; = 8) (« = -3) (n = 3) in = 4) Females 14.1 ± 10.2 12.2 ± .3..3 9.4 ± 6.3 7.4 ± 2.7 4.3 ± 1.9 0 (n = 9) in = 9) (n = 9) (/) = 6) (n = 4) (n = 0) Gender combined 14.3 ± 9.9 9.4 ± 10.4 6.6 ± 7.4 .3.9 ± 2.6 2.8 ± 1.6 2.6 ± 2.2 (n = 13) (n = 21) (n = 17) (n = 11) (n = 1) (/; = 4) Yti cl uL • CRESTED IBIS DISPERSAL 233 0 FIG. 4. Graphical depiction showing final direction from the natal site to first breeding territory for 73 banded Crested Ibis. Directions are spatially nonrandom (Ray- leigh’s z: all P < 0.05). Direction of natal dispersal and post-fledging movement was likely in response to habitat factors. Crested Ibis were once commonly seen in towns and cities in agricultural areas, near which they could find large trees for nesting and suitable wetlands for feeding (La Touche 1934). The wild population that survived in Yangxian County persisted in remote mountainous districts at an altitude of 1,200 m, where human exploi- tation was not as severe as in neighboring areas (Shi et al. 1989, Shi and Cao 2001). The species has showed a recent prel'erence to move to lower areas for breeding (Ding 2004). Crested Ibis prior to 1990 usually nested on oaks in mountainous areas above 1,000 m. They have colonized new nesting sites in recent years where they usually nest on pines at lower altitudes (700-900 m), presumably indicating the relict population had become confined to suboptimal upland habitat but is now expanding into more typical lowland areas. More recently, a few birds have been found breeding in patches of secondary forests close to human settlements (Liu et al. 2003). Forty-nine (67.1%) of 73 marked chicks in our study required 2 to 4 years to become territorial (Fig. 5). Females bred no more than 1 year earlier than males on average, similar to Marabou Storks {Leptoptilos cnimeniferus) (Pomeroy 1978), Scar- let Ibis {Eudocimus ruber), and Eurasian Spoon- bill {Platalea leucorodia) (Hancock et al. 1992). These data may indicate the habitat is saturated or reflect other breeding constraints on the popula- tion. Variation in age of first breeding could reflect individual variation in physical condition or foraging skills or, alternatively, individuals selecting different strategies for maximizing reproductive success (Nelson 1977, Osorio-Ber- istain and Drummond 1993). Males are at least 1 year older than females in most pairs (Lu et al. 1997), and males were observed to have a dominant role in territory defense (Shi and Cao 2001). This suggests that greater experience is necessary for males to have a mate and hold a territory. 0 35 2 3 4 5 6 7 Age (yeirs) FIG. 5. Age of first breeding attempt in = 36 males and 37 females). 234 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 2, June 2010 Survival, productivity, and dispersal data are important parameters in population models that highlight stages in a species’ life cycle where conservation actions may be most critical. Lim- ited data on dispersal of both young and yearling ibis led to imprecise survival estimates, and the low success of first breeding attempts for each age class (Yu et al. 2006) suggest low year-to-year survival of juveniles. Our results indicate that young individuals are likely to disperse farther than adults, and females breed nearly 1 year earlier than males. The current distribution is progressively expanding to areas at lower eleva- tion and denser human populations by dispersal of young individuals. These findings have potential- ly important implications for reintroduction pro- grams that are currently being considered (Yu et al. 2009). We strongly recommend that reintro- duction programs be conducted with younger (2- 3 yrs of age) individuals which are more prone to establishing new territories and have low site fidelity. ACKNOWLEDGMENTS We thank Y. M. Zhang, T. Q. Zhai, Y. P. Chen, and Y. J. Wang for assistance in field observations. The State Forestry Administration, Shaanxi Nature Reserve and Wildlife Administration, and Shaanxi Natural Science Foundation provided administrative support and funding. We thank the Yangxian County weather bureau for providing meteorological information. Chris Wood, Daniel White, and Ren Wei significantly improved earlier versions of this manuscript. LITERATURE CITED Archibald, G. W., S. D. FI. Lantis. and I. Munetchika. 1980. Endangered ibises, Threskiornithinae: their future in the wild and in captivity. International Zoo Yearbook 20:6-17. Bull, E. L., M. G. Henjum, and R. S. Rohweder. 1988. Home range and dispersal of Great Gray Owls in northeastern Oregon. Journal of Raptor Research 22:101-106. Chernetsov. N.. W. Chromik, P. T. Dolata. P. Proeu.s, AND P. TRYJANOW.SKI. 2006. Sex-relatcd natal dispers- al of White Storks (Cicouici ciconia) in Poland: how far and where to? Auk 12.1:1 103-1 109. Clarke, A . L.. B. E. S/ether. and E. Ro.skaft. 1997. Sex biases in avian dispersal: a reappraisal. Oikos 79:429- 438. Cl-OBfiRT. J.. E. Danchin. a. a. Diiondt, and j. D. Nichols. 2001. Dispersal. Oxford University Pre.ss, Oxford. United Kingdom. Ding, C. Q. 2004. Research on the Crested Ibis in China. Shanghai Scientific and Technological Publishing House. Shanghai, China. Gill Jr., R. and L. M. Richard. 1979. Dispersal and migratory patterns of San Erancisco Bay produced herons, egrets, and terns. North American Bird Bander 4:4-13. Greenwood, P. J. 1980. Mating systems, philopatry and dispersal in birds and mammals. Animal Behaviour 28:1140-1162. Greenwood, P. J. and P. H. Harvey. 1982. The natal and breeding dispersal of birds. Annual Review of Ecology and Systematics 13:1-21. Hancock, J. A.. J. A. Kushlan, and M. P. Kahl. 1992. Storks, ibises and spoonbills of the world. Academic Press, New York, USA. Hansson. B., B. D. Hasselquist. and D. B. Nielsen. 2002. Restricted dispersal in a long-distance migrant bird with patchy distribution, the Great Reed Warbler. Oecologia 130:536-542. He, L. P., Q. H. Wan, S. G. Fang, and Y. M. Xi. 2006. Development of novel microsatellite loci and assess- ment of genetic diversity in the endangered Crested Ibis. Nipponia nippon. Conservation Genetics 7:157-160. Hobson, K. A., K. P. McFarland, L. 1. Wassenaar, C. C. Rimmer, and j. E. Goetz. 2001. Linking breeding and wintering grounds of Bicknell's Thrushes using stable isotope analyses of feathers. Auk 1 18:16-23. Horn, P. L., J. A. Rafalski. and P. J. Whitehead, 1996. Molecular genetic (RAPD) analysis of breeding Magpie Geese. Auk 1 13:552-557. Kareiva, P. 1990. Population dynamics in spatially complex environments: theory and data. Philosophical Transactions of the Royal Society of London. Series B 330:175-190. Kendall, W. L. and J. D. Nichols. 2004. On the estimation of dispersal and movement of birds. Condor 106:720-731. Koenig, W. D., P. N. Hooge, M. T. Stanback. and J. Haydock. 2000. Natal dispersal in the cooperatively breeding Acorn Woodpecker. Condor 102:492-502. La Touche, J. D. D. 1934. Page 434 in A handbook of the birds of eastern China. 11. Part 5. Taylor and Francis, London, United Kingdom. Lang, J. D., L. A. Powell, D. G. Krementz, and M. J. Conroy. 2002. Wood Thrush movements and habitat use: effects of the forest management for Red- cockaded Woodpeckers. Auk 1 19:109-124. Li, M.. C. Q, Ding, F. W. Wei. S. J. Meng, Y. M. Xi, and B. Z. Lu. 2001 . Sex-related gene and sex identification of Crested Ibis Nipponia nippon (Ciconiiformes: Threskiornithidae), Chinese Science Bulletin 46:669- 671. Li, X. H., Z. .1. Ma. C. Q. Ding, T. Q. Zhai, and D. M. Li. 2002. Relationship between the distribution of Crested Ibis and local farmers. Acta Zoologica Sinica 48:725- 732. . ■ Liu, D. P., C. Q. Ding, and G. Z. Chu. 2003. Home range and habitat utilization of the Crested Ibis in the breeding period. Acta Zoologica Sinica 49:755-763. Liu, Y. Z. 1981. The re-discovery of the Crested Ibis in Qinling Mountain. Acta Zoologica Sinica 27:273. Lu, B. Z., T. Q. Zhai, Y. M. Zhang, and Z. Z. Geng. 1997. Wild populalion slruclure and relationship of the Yn el cil. • CRESTED IBIS DISPERSAL 235 Crested Ibis (Nipponia nippon). Journal of Shaanxi Normal University 25 (Supplement):98-I04. Nelson, J. B. 1977. Some relationships between food and breeding in the marine Pelecaniformes. Pages 77-78 in Evolutionary ecology (B. Stonehouse and C. Perrins, Editors). University Park Press, Baltimore, Maryland, USA. Oriana. 2003. Oriana for Windows, Version 2.0. Kovach Computing Services, Pentraeth, Wales, United King- dom. Osorio-Beristain, M. and H. Drummond. 1993. Natal dispersal and deferred breeding in the Blue-footed Booby. Auk 110: 234-239. Paradis, E., S. R. Baillie, and W. J. Sutherland. 2002. Modeling large-scale dispersal distances. Ecological Modelling 151:279-292. Paradis, E., S. R. Baillie, W. J. Sutherland, and R. D. Gregory. 1998. Patterns of natal and breeding dispersal in birds. Journal of Animal Ecology 67:518-536. Pomeroy, D. 1978. The biology of Marabou Storks in Uganda. II. Breeding biology and general review. Ardea 66: 1-23. Powell, L. A. and L. L. Frasch. 2000. Can nest predation and predator type explain variation in dispersal of adult birds during the breeding season? Behavioral Ecology 11:437^143. Pyle, P. 2001. Age at first breeding and natal dispersal in a declining population of Cassin’s Auklet. Auk 118: 996-1007. Shi, D. C. and Y. H. Cao. 2001. The Crested Ibis in China. China Forestry Publishing House, Beijing. Shi, D. C., X. P. Yu, X. Y. Chang, and B. Z. Lu. 1989. Breeding habits of the Crested Ibis. Zoological Research 10:327-332. Walters, J. R. 2000. Dispersal behavior: an ornithological frontier. Condor 102:479-481. Walls, S. S. and R. E. Kenward. 1998. Movement of radio-tagged Buzzards Buleo hiiteo in early life. Ibis 140:561-568. Wang, X. H. and C. H. Trost. 2001. Dispersal pattern of Black-billed Magpies (Pica hudsonia) measured by molecular genetic (RAPD) analysis. Auk 118:137- 146. Yamashina, Y. 1967. The plight of the Japanese Crested Ibis. Animals 10:275-277. Yu, X. P., N. F. Liu, Y. M. Xi, and B. Z. Lu. 2006. Reproductive success of the Crested Ibis Nipponia nippon. Bird Conservation International 16:325- 343. Yu, X. P., X. Y. Chang, X. Li, W. G. Chen, and L. Shi. 2009. Return of the Crested Ibis Nipponia nippon: a reintroduction programme in Shaanxi Province, China. BirdingASIA 11: 80-82. Zar, j. H. 1996. Biostatistical analysis. Third Edition. Prentice-Hall, Englewood Cliffs, New Jersey, USA. Zhai, T. Q. and Z. Y. Wang. 1994. On the reproduction of two-year-old Crested Ibis. Acta Ecologica Sinica 14:99-101. Zhang, B., S. G. Fang, and Y. M. Xi. 2004. Low genetic diversity in endangered Crested Ibis Nipponia nippon and implications for conservation. Bird Conservation International 14: 183-190. The n'ilson Journal oj Ornithology 122(2):236-243, 2010 MORPHOLOGICAL AND GENETIC VARIATION BETWEEN MIGRATORY AND NON-MIGRATORY TROPICAL KINGBIRDS DURING SPRING MIGRATION IN CENTRAL SOUTH AMERICA ALEX E. JAHNT' DOUGLAS J. LEVEY,' IZENI FIRES FARIAS,- ANA MARIA MAMANI,^ JULIAN QUILLEN VIDOZ,^ AND BEN FREEMAN^ ABSTRACT. — We attempted to distinguish spring passage migrant Tropical Kingbirds (Tyrannus melancholicus) from resident conspecifics where they overlap in South America. Migrant males at our Bolivian study site had significantly less tail feather molt and longer wing chords than resident males. Migrant females had significantly longer wing chords, less flight feather molt, and less flight feather wear than resident females. We found no evidence of genetic population differentiation between migrants and residents. We also compared wing chords of migrants and residents to those of breeding kingbirds in breeding populations further south. Wing chords of migrants were more similar to those of breeders from further south than to those ot breeders at our study site. An ability to distinguish migrant from resident conspecifics will be critical to understanding migrant winter ecology, migratory routes, and connectivity of migratory populations in South America. Received IH May 2009. Accepted 2 December 2009. Almost 70% of neotropical austral migrant species (hereafter “austral migrants”), which migrate wholly within South America, are char- acterized by having both migratory and non- migratory (permanent resident) populations (Stotz et al. 1996). Populations of permanent residents usually occur within or close to tropical latitudes, whereas populations of migratory individuals occur in temperate latitudes during the breeding season and in tropical latitudes during the non- breeding season. Ranges of migratory and resident populations of austral migrants typically overlap during the non-breeding season. Patterns of migration within the range of overlap are rarely documented because distinguishing between mi- gratory and resident individuals is difficult. Thus, for the great majority of austral migrants, we know little about migratory routes, non-breeding season ecology, population connectivity, and intraspecific interactions between residents and migrants. Zimmer (1938) and Lanyon (1978) used .seasonal distribution, morphology, and molt data to study where non-breeding populations of migrants and residents overlap in South America. More recently, Ches.ser (1995, 1997) primarily ' Department of Biology, University of Florida, Gaines- ville. FL 3261 I. U,SA. ^Instituto de Ciencias Biologicas, Departamento de Biologia, Universidade Federal do Amazonas. Manaus. Amazonas, Brazil. 'Museo de llistoria Natural Noel Kempff Mercado, Santa Cruz. Bolivia, ■'8412, 36th Avenue Northeast, Seattle, WA 98 1 1 5, USA. ’Corresponding author; e-mail: alexJahn77@yahoo.com used seasonal distribution data to identify the range of winter overlap of migrant and resident austral migrant populations at the center of the continent. Significant morphological differences in other migratory systems exist between migrant and resident individuals of the same species; those differences have been used to distinguish between co-occLining migrants and residents. For example, migratory European Blackcaps {Sylvia atricapilla) have longer, more pointed wings and smaller bodies than resident conspecifics in Iberia (Tell- eria and Carbonell 1999), allowing accurate classification of migrant and resident Blackcaps in winter (Perez-Tris et al. 1999). Similarly, a migratory subspecies of Common Reed Bunting {Emheriza schoenidus) in Spain has larger and more convex wings than a sedentary subspecies (Copete et al. 1999). Molt schedules may also differ between co-occurring resident and migrant individuals (Lundberg and Eriksson 1984). Range overlap of resident and austral migrant populations in the non-breeding season occurs for Tropical Kingbirds (Tyrannus melancholicus), which are permanent residents in most of tropical South America. Two subspecies (satrapa and despotes) appear to be permanent residents in northern Amazonia (Traylor 1979, Chesser 1995), whereas the subspecies melancholicus is a resident from most of Amazonia south to —18° S. South' of this latitude, melancholicus occurs only in sum- mer, migrating north after breeding into the range of permanent resident populations (Chesser 1995; Fig. 1). We studied a resident population of Tropical Kingbirds (hereafter “kingbirds”) at a site in 236 John ct c/I. • TROPICAL KINGBIRDS 237 Tropical Kingbird {Tyra//niis m. mela/\cholicits) in South America. Gray polygon represents the permanent distribu- tion of the species and the hatched polygon south of 18° S represents the seasonal breeding range. Star represents location of Caparu Biological Station, Bolivia. Adapted from Chesser (1995). eastern Bolivia (~15° S) within the range of the nominate subspecies. We noticed that many kingbirds captured during September and Oc- tober were much fatter than others, and that fewer of those fat individuals were seen again at the site, compared to those individuals that were less fat. We hypothesized that fatter individuals were passage migrants, on their way to breeding areas further south. Our objective was to assess different methods of distinguishing between migrant and resident kingbirds where they co- occur. We compared; ( 1 ) the molt schedule, wing chord, and genetic signature between putative migrants and residents, and (2) the wing chord of migrants and residents at the study site with wing chords of kingbirds breeding lurther south. METHODS Study Site and Field Procedures. — Our study site was Caparu Biological Station (CBS), De- partment of Santa Cruz, eastern Bolivia (14° 49' S, 61° ir W; 170 m asl). The site is near the center of the continent, north of the boundary of migratory kingbirds’ breeding range and south of a large expanse of potential wintering area (i.e.. Amazonia; Fig. 1). The habitat is primarily cerrado grassland, humid forest edge, and cattle pasture. Kingbirds were captured using nylon and polyester mist nets (12 or 18 m X 2.6 m, 36 or 38 mm mesh size) where they were most abundant (e.g., ponds and forest edge). Kingbirds were also banded as nestlings and parents were captured at their nest by placing a stuffed Purplish Jay (Cyanocora.x cyanomelas), a common nest pred- ator, nearby. Kingbirds were banded with uniquely- numbered aluminum bands provided by the Museo de Historia Natural Noel Kempff Mercado (MHNNKM), Santa Cruz, Bolivia as well as with up to three celluloid color bands. Banding was conducted during most months from October 2004 to July 2007. Data recorded for captured kingbirds included reproductive condition (presence/ab- sence of cloacal protuberance or brood patch), fat score (measured on an 8-point scale), body molt (measured on a 5-point scale), primary feather wear (measured on a 6-point scale), length of unflattened wing chord, tail, beak, and tarsus length, and mass (Ralph et al. 1993, Pyle 1997). We classified flight feather and tail molt as present or absent; molt was classified as “pres- ent” if at least one secondary or primary feather or retrix was molting. These measurements were taken by AEJ and AMM, who collected data together on a regular basis to minimize observer error. We classified birds in their first year of life as hatch-year individuals (HY) or after-hatch-year individuals (AHY) for individuals that were of unknown age but at least 1 year of age. Each year began on 1 September because the kingbird breeding season lasts from mid-September to February. Blood was taken for genetic analysis from all color-banded individuals by piercing the brachial vein with a sterile needle. Blood was collected in a 0.5-mL heparinized capillary tube and stored in DNA lysis buffer (Seutin et al. 1991). We assigned gender of live birds based upon primary notch length (Pyle 1997) and, when necessary, through molecular methods using primers 2550F and 271 8R (Fridolfsson and Ellegren 1999). Optimal cycle conditions were 95° C for 5 min, 95° C for 30 sec, an annealing temperature of 50° C for 1 min, and 72° C for 45 sec. repeated 39X. and held at 72° C for 10 min, using 15 pL Mg. Blood samples were deposited at the MHNNKM. We documented seasonal presence/absence of color-banded kingbirds after banding to identify 238 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 2. June 2010 which individuals were residents. We divided the ~ 700-ha study site into 23 sampling plots and methodically searched for banded individuals in each plot at least once/month throughout most of the annual cycle. These censuses were completed from February 2005 to August 2007. We did not visit plots during most of the non-breeding season of 2005 because we were absent during June- September. We also visited plots on 28 January to 12 February, 2-19 March, and 15-27 June 2008. We geo-referenced the location when a color- banded individual was observed using a Global Positioning System receiver (Garmin GPS 76, Olathe, KS, USA), noting the date, time, and color band combination. AEJ measured 120 kingbird skins from Argen- tina and Paraguay housed at Museo Argentine de Ciencias Naturales Bernardino Rivadavia (MACN; Buenos Aires, Argentina) and 79 skins at Louisiana State Museum of Natural Science (LSUMZ; Baton Rouge, USA) to examine whether morphological variation among popula- tions could be useful in identifying breeding locations of passage migrants at CBS. BF also measured 60 study skins from Paraguay and Argentina at the American Museum of Natural History (AMNH; New York City, USA). Un- flattened wing chord was measured as on live kingbirds at CBS, following Ralph et al. (1993). Wing chord was not measured on specimens with excessively worn primary feather tips. We previously standardized our measurements to minimize inter-observer measurement error. We isolated DNA from blood samples using the PUREGENE DNA Purification Kit (Gentra Sys- tems, Minneapolis, MN, USA) and reviewed the literature to identify potential primers of poly- morphic loci developed from related species in Tyrannidae. Six microsatellite primers were successfully optimized from Empidonax (GATA5; Pearson et al. 2006) and Sayornis (SAP22, SAP32, SAP47, SAP50, SAP 156; Watson et al. 2002, Beheler et al. 2007). Optimal PCR cycling conditions for all primers were: 95° C for 4 min followed by 94° C for 1 min, then 55° C for 1 min, and 72° C for 1 min, repeating the last three steps 35X with a final cycle of 72° C for 10 min. We used 12.5 pL of master mix for all primers, 800 pM dNTP’s, 2.5 pM MgCl2, 0.26 pM of each primer, and a concentration of 10-20 ng of sample DNA. We combined males and females for genotyp- ing due to low sample sizes of kingbirds of known gender for which we had DNA available for analysis. We analyzed DNA from 15 migrants and 15 residents. DNA was run on an ABI 3730XL, 96 capillary automated sequencer (Biosystems, Fos- ter City, CA, USA). The program GENEMAP- PER (Softgenetics, State College, PA, USA) was used to score the alleles. Population structure was tested in program STRUCTURE 2.2 (Pritchard et al. 2000) assuming a model of admixture, correlated allele frequencies, and K = 1 to 2 (where K is the number of populations) with 10 independent replicates for each K (length of the run 1,000,000 MCMC re-samples), and 100,000 burn-in. We used program ARLEQUIN 3.01 (Excoffier et al. 2005) to test for departures from Hardy-Weinberg equilibrium, and to calculate the amount of linkage disequilibrium between pairs of loci. Bonferroni correction was applied for multiple comparisons in both tests. There was violation of Hardy-Weinberg equilibrium in one locus (SAP32) for both resident and migrant populations, and in only migrants for another locus (SAP50) but, overall, loci adhered to equilibrium. Population differentiation was tested by performing a hierarchical analysis of molecular variance (AMOVA) to calculate global and pair- wise Fs, in ARLEQUIN 3.01. We only consider adult kingbirds (i.e., those with only trace juvenile plumage or 95% or more of the skull ossified) to avoid potential confound- ing effects of age with wing chord (Low 2006), fat content (e.g., Lundberg 1985), and molt (e.g., Debruyne et al. 2006). After-hatch-year (AHY) kingbirds captured in September and October that had higher subcutaneous lipid scores (i.e., >6 on an 8-point scale, sensu Ralph et al. 1993) were re- observed significantly less often than those with lower lipid scores (i.e., <6 on the same scale; Fig. 2). This was true of both males (x^ = 1 1 .8, P = 0.001) and females = 7 A, P = 0.008). We classified as passage migrants all AHY kingbirds captured in September or October with lipid scores of 6 or higher, that were not re-observed after being color banded, and which had no incubation patch or cloacal protuberance (to exclude the possibility they were breeding at the study site). We classified residents as individuals that were captured in September or October, had subcutaneous lipid scores of 1-5, and re-observed at least once later on the study site. There were insufficient residents for which DNA was avail- able for analyses, and we included as residents for puiposes of DNA analyses individuals with the .lulm cl al. • TROPICAL KINCBIRDS 239 Low □ Females ■ Males Subcutaneous fat score FIG. 2. Relationship between amount of subcutaneous fat and frequency of being re-observed for AHY Tropica) Kingbirds color banded during spring migration at Caparti Biological Station. Bolivia; “Low” = subcutaneous fat score of <6 on an 8-point scale, “High” = subcutaneous fat score >6. same characteristics captured in November, which is well into the breeding season (A. E. Jahn, unpubl. data). RESULTS Resident Tropical Kingbirds at CBS tended to molt during and after the breeding season (Fig. 3). Spring passage migrants had significantly less flight feather molt (Kruskal-Wallis ~ 5.96, P — 0.015, n = 81) and significantly less tail molt (Kruskal- Wallis x^ = 7.66, P = 0.006, n = 79) than re.sidents when data for males and females, and those of unknown gender were combined. However, there was no significant difference between migrants and residents in body molt ( Kruskal-Wallis x^ = 1 .07, P = 0.30, n = 81). Migrants also had less primary feather wear (Kruskal-Wallis x^ = 10.36, P = 0.001, n = 80) and longer wings (/ = 4. 17, P < 0.0001, n - 79) than residents. Female passage migrants had significantly less flight feather molt, less primary feather wear, and longer wings than resident females (Table 1 ). Male residents had significantly more tail molt than male migrants, but there was no signilicant difference in flight feather molt or primary feather wear between migrant and resident males. Male migrants were also characterized by a significant- ly longer wing chord than male residents (Table 1 ). We used discriminant function analysis (DFA) to examine the level at which we could correctly classify migrants and residents at the site during spring migration. We combined data for males and females due to the relatively low sample sizes for each. We first examined which of the five morphological measurements (flight feather molt, tail molt, primary feather wear, body feather molt, and wing chord) provided the most explanatory power within the discriminant function. A test of the equality of group means revealed that body molt contributed the least amount of explanatory power to the discriminant function {F]j2 = 1.09, P = 0.30); we therefore excluded it from further analysis. All four remaining variables significant- ly contributed to the discriminant function and were retained in the full model. The parameter with the least explanatory power was flight feather molt {F\j2 — 4.48, P = 0.038), and the parameter with the highest explanatory power was wing chord (F\j2 = 16.84, P < O.OOOl). The resulting DFA correctly classified 86.0% of migrants and 71.0% of residents with an overall success rate of 79.7%. Wing chord from study skins of males at >18° S in — 146) was significantly longer than that of both resident (/ = —6.93, P < 0.0001, n = 15) and passage migrant males at the study site {t = -2.71, P = 0.007, n = 21; Fig. 4). Wing chord 240 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 2. June 2010 90 n 80 - ■ Body molt May Jun Jul Aug Sep Oct Nov Dec Jan Feb Mar Apr FIG. 3. Molt schedule of AHY color-marked, re-observed resident Tropical Kingbirds at Caparu Biological Station, Bolivia. Data represent the percentage of kingbirds per month with each molt type. Percentages may not sum to 100 because individuals can have more than one molt type. from Study skins of females at >18° S (/? = 83) was significantly longer than that of female residents (r = 6.787, P < 0.0001, n = 11) but not of female passage migrants at the study site {t = 1.055, P = 0.29, n = 16; Fig. 4). We found better support using genetic data for the presence of one population (log likelihood = —515.2) than for the presence of two populations (log likelihood = —516.6), based on Bayesian analyses of population structure, as implemented in program STRUCTURE 2.2. This suggests there is no significant genetic structure between migrant and resident kingbirds in the six loci tested. AMOVA also gave no indication of genetic differentiation between migrants and residents (^st = 0.01576, P = 0.578). There was no indication of departures from Hardy-Weinberg equilibrium or of linkage disequilibrium (Ta- ble 2). We conducted DFA for migrants and residents classified as above, but without considering re- observation data (i.e., using only fat scores) to check whether using fat content alone to distin- guish between migrants and residents is as useful TABLE 1. Morphological and molt comparisons of resident and passage migrant Tropical Kingbirds at Caparu Biological Station, Bolivia. Values for flight feather and tail molt are percent of migrants and residents with molt present. Values for body feather molt and primary feather wear represent mean score on a 5- and a 6-point .scale, respectively (high .scores indicate more molt or wear). Resident.s (n) Migrants in) Test statistic p Males Flight feather molt 12% (17) 0% (21) 2.54“ 0.1 1 1 Tail molt 20% (15) 0% (21) 4.46“ 0.035 Body feather molt 2.44 ± 1.09 (16) 2.24 ± 1.18 (21) 0.28“ 0.595 Primary feather wear 2.06 ± 0.25 (16) 1.86 ± 0.48 (21) 2.42“ 0.120 Wing chord, mm 1 1 1.27 ± 2.87 (15) 1 13.9 ± 2.84 (21) 2.735” . 0.010 Females Flight feather molt 25% (12) 0% (16) 4.06“ 0.044 Tail molt 8% (12) 0% (16) 1.25“ 0.264 Body feather molt 2.82 ± 0.98 (11) 2.47 ± 1.19 (15) 0.50“ 0.481 Primary feather wear 2.58 ± 0.79 (12) 1.80 ± 0.56 (15) 6.92“ 0.009 Wing chord, mm 106.09 ± 1.97 (1 1) 1 10.63 ± 2.63 (16) 4.846” <0.0001 “ Kruskal-Wallis test. ^ Independent samples /-test. Juhn Cl al. • TROPICAL KINGBIRDS 241 H Female lOO-' 1 1 I Resident at Migrant at Southern CBS (26) CBS (37) breeder (229) FIG. 4. Box-and-whisker diagram depicting wing chord of Tropical Kingbird residents, passage migrants, and breeders from seasonal breeding range south of 18° S latitude in South America. Numbers in parentheses on x- axis labels represent sample sizes. Rectangles depict the range of the first to third quartiles and the dark horizontal line within each rectangle depicts the median. Lines from each rectangle extend to the larger and smaller values, and circles outside of rectangles represent outliers. as using fat plus re-observation data. Body molt within the revised DFA had practically no explanatory power (Fijss = 0.01, P = 0.94) and was excluded. The resulting test of equality of group means revealed that flight feather molt had the most explanatory power (F| 159 = 4. 10, P = 0.044) and that wing chord had the least (T1J59 = 0.28, P = 0.60). The DFA correctly classified 75% of migrants and 26.5% of residents with an overall success of 41.0%. DISCUSSION Tropical Kingbirds with high lipid scores and those that were not re-observed after banding (i.e., putative migrants) were characterized by traits typical of migrants in the Northern Hemisphere: advanced molt .schedules (Lundberg and Eriksson 1984) and longer wings (Haberman et al. 1991, Tellerfa and Carbonell 1999, Mila et al. 2008) relative to kingbirds with less fat and which were re-observed at the site (i.e., residents). The differences between the two groups in molt, feather wear, and wing chord from our site at the middle of the continent, where passage migrants would be expected to occur if migrating between low latitudes to the north and temperate breeding areas to the south, strongly suggest some Tropical Kingbirds are long-distance mi- grants that overwinter in central or northern Amazonia. Migrant females at CBS had wing chords similar to breeders from the seasonal breeding range >18° S, whereas male migrants had significantly shorter wings than kingbirds from the seasonal breeding range. Thus, at least some female spring passage migrants were likely returning to the seasonal breeding range, while male migrants may have been returning to the seasonal breeding range, as well as to latitudes closer to the study site. Future research on geographical variation in wing chord and genetic structure south of the study site may reveal geographical variation in traits that could be useful in identifying the breeding location of these migrants. Molt of resident kingbirds at CBS tended to occur primarily just before, during, and after the breeding season (Oct-Dec). Thus, many resident kingbirds at CBS likely had not molted prior to passage of spring migrants. Passage migrant male kingbirds had a more advanced pre-alternate tail molt compared to local resident males, and female migrants had a significantly more advanced flight feather molt than residents at CBS. This indicates migratory kingbirds have more advanced pre- alternate molt schedules than non-migratory individuals at CBS. Lundberg and Eriksson (1984) found the post-juvenile molting period of migratory European Starlings (Sturniis vulgaris) TABLE 2. Genetic diversity for resident and migrant Tropical Kingbirds. Genetic diversity parameters include mean observed heterozygosity {Hq), mean expected heterozygosity (//pj, inbreeding coefficient (F,s) based on the average of six loci, average number of alleles per locus (A), and average number of private alleles {PA). Hii-He Gs Average gene diversity across loci A PA n Residents 0.423-0.478 0.020 (Z' = 0.497) 0.422 ± 0.259 6.67 1.8 15 Passage migrants 0.438—0.611 0.107 (P = 0.088) 0.582 ± 0.339 8.00 3.0 15 242 THE WILSON JOURNAL OL ORNITHOLOGY • Voi 122, No. 2, June 2010 was shortened relative to that of non-migratory individuals with migrants finishing the molt of primary and secondary feathers significantly ear- lier than residents. They postulated the shortened molting period of migrants was constrained by the need to migrate after breeding. Our study focused on spring migration and we do not know if migrants molt their feathers in a shorter span of time compared to residents. That migrant kingbirds finish the pre-alternate molt earlier than resident kingbirds suggests migration can be a constraint on molt timing. It remains unknown whether migra- tion constrains the scheduling of molt at tropical latitudes. Genetic analysis revealed no significant popu- lation subdivision between migrant and resident Tropical Kingbirds. The lack of microsatellite structure in our study may be due to at least three reasons: ( 1 ) too few individuals or loci sampled, (2) relatively high gene flow occurring among populations, and (3) recent colonization of southern latitudes by kingbirds. Our results corroborate previous evidence that populations of Tropical Kingbirds breeding at high southern latitudes in South America (i.e., T. m. melancholiciis) have longer wing chords than populations breeding closer to the equator (i.e., T. m. satrapa and T. m. despotes\ Cory and Hellmayr 1927). Likewise, southern breeding populations of Swainson’s Flycatcher (Myiarchus swainsoni) in South America have longer wing chords than populations breeding at tropical latitudes (Lanyon 1978, Joseph et al. 2003). Distinguishing between migrant and resident conspecifics when they occur together at tropical latitudes is important in advancing our under- standing of the evolution of migratory behavior. However, migrants may not be obviously different in morphology from permanent resident individ- uals in most species of austral migrants because austral migrants that overwinter in South Amer- ica’s tropics migrate an average of 9.2° (±8.5) latitude (Chesser 1994). ACKNOWLEDGMENTS We thank Terry Ches.ser. Scott Robinson. Katie Sieving. C. E. Braun, and an anonymous reviewer for many helpful comments that improved the manuscript. We are grateful to Jaime Rozenman and Guillermo Wei.se for their hospitality and ongoing support of research at Caparu Biological Station, and to Rebecca Kimball, Ginger Clark. Carolyne Bardelebcn, Amanda Behelcr. Christy Watson, and Olin Rhodes for advice and logistical support for molecular analyses. We th;mk collection managers al the AMNH (Paul Sweet), LSUMZ (Steve Cardiff), and MACN (Pablo Tubaro) for access to the specimens under their care. A large number of assistants helped in the field and laboratory, especially Noelia Barrios, Silvia Estevez, Betty Flores, Fabian Hilarion, Jennifer Johnson, Patricia Justin- iano, John Prather, Ana Maria Saavedra, and Jordan Smith. Funding was provided by the American Ornithologists' Union, the National Science Foundation (OISE-03 13429, 0612025), Optics for the Tropics, School of Natural Resources and Environment-University of Florida, NSF Southeast Alliance for Graduate Education and the Professoriate, Western Bird Banding Association, and the Wilson Ornithological Society. The genetic research was supported by a grant to I. P. Farias from CNPq/CT- Amazonia #575603/2008-9. LITERATURE CITED Beheler, a. S., J. A. Fike, and O. E. Rhodes Jr. 2007. Eight new polymorphic microsatellite loci from the Eastern Phoebe (Sayornis pboehe). Conservation Genetics 8:1259-1261. Chesser, R. T. 1994. Migration in South America, an overview of the austral system. Bird Conservation International 4:91-107. Chesser, R. T. 1995. Biogeographic, ecological, and evolutionary aspects of South American austral migration, with special reference to the Family Tyrannidae. Dissertation. Louisiana State University. Baton Rouge, USA. Chesser, R. T. 1 997. Patterns of seasonal and geographical distribution of austral migrant llycatchers (Tyrannidae) in Bolivia. Ornithological Monographs 48:171-204. CoPETE, J. L.. R. Marine, and D. Bigas. 1999. Differences in wing shape between sedentary and migratory Reed Buntings Emheriza schoenichis. Bird Study 46:100-103. Cory, C. B. and C. E. Hellmayr. 1927. Catalogue of birds of the Americas. Field Museum of Natural History Publications, Zoology Series 13, Part 5. Debruyne, C. a., C. M. Hughes, and D. J. T. Hussell. 2006, Age-related timing and patterns of pre-basic molt in wood warblers (Parulidae). Wilson Journal of Ornithology 118:374-379. Excoffier, L., G. Laval, and S. Schneider. 2005. ARLEQUIN Version 3.1: an integrated software package for population genetics data analysis. Evolu- tionary Bioinformatics Online 1:47-50. Fridolfssgn, a. K. and H. Ellegren. 1999. A simple and universal method for molecular sexing of non-ratite birds. Journal of Avian Biology 106:1 16-121. Haberman, K., D. I. Mackenzie, and J. D. Rising. 1991. Geographic variation in the Gray Kingbird. Journal of Field Ornithology 62:1 17-131. Joseph, L., T. Wilke, and D. Alpers. 2003. Independent evolution of migration on the South American landscape in a long-distance temperate-tropical migra- tory bird, Swainson's Flycatcher Myianhus swain.soui. Journal of Biogeography 30:925-937. Lanyon, W. E. 1978. Revision of the Myiarchu.s Hycatch- ers of South America. Bulletin of the American Museum of Natural History 161:429-627. Jahu et al. • TROPICAL KINGBIRDS 243 Low, M. 2006. Sex, age and .season influence morphomet- ries in the New Zealand Stitchbird (or Hihi; Notio- mystis cincta). Emu 106:297-304. Lundberg. P. 1985. Dominance behaviour, body weight and Fat variations, and partial migration in European Blackbirds Turdus meriila. Behavioral Ecology and Sociobiology 17:185-187. Lundberg, P. and L.-O. Eriksson. 1984. Postjuvenile moult in two northern Scandinavian starling Sturniis vulgaris populations-evidence for difference in the circannual time-program. Ornis Scandinavica 15:105-109. Mila, B., R. K. Wayne, and T. B. Smith. 2008. Ecomorphology of migratory and sedentary popula- tions of the Yellow-rumped Warbler (Dendroica coronata). Condor 1 10:335-344. Pearson, T., M. J. Whitfield, T. C. Theimer, and P. Keim. 2006. Polygyny and extra-pair paternity in a population of Southwestern Willow Flycatchers. Condor 108:571-578. Perez-Tris, J., R. Carbonell, and J. L. Telleria. 1999. A method for differentiating between sedentary and migratory Blackcaps Sylvia atricapilla in wintering areas of southern Iberia. Bird Study 46:299-304. Pritch.ard, j. K., M. Stephens, and P. Donnelly. 2000. Inference of population structure using multilocus genotype data. Genetics 155:945-959. PVXE, P. 1997. Identification guide to North American birds. Part I. Slate Creek Press, Bolinas, California, USA. Ralph, C. J., G. R. Guepel, P. Pyle, T. E. Martin, and D. F. Desante. 1993. Handbook of field methods for monitoring landbirds. USDA, Forest Service, General Technical Report PSW-GTR-144. Albany, California, USA. Seutin, G., B. N. White, and P. T. Boag. 1991. Preservation of avian blood and tissue samples for DNA analyses. Canadian Journal of Zoology 69:82- 90. Stotz, D. F., j. W. Fitzpatrick, T. A. Parker III, and D. K. Moskovits. 1996. Neotropical birds: ecology and conservation. University of Chicago Press, Chicago, Illinois, USA. TellerIa, j. L. and R. Carbonell. 1999. Morpho- metric variation of five Iberian Blackcap Sylvia atricapilla populations. Journal of Avian Biology 30:63-71. Traylor Jr., M. A. 1979. Subfamily Tyranninae. Pages 186-229 in Checklist of birds of the world (M. A. Traylor Jr., Editor). Volume 8. Museum of Compar- ative Zoology, Cambridge, Massachusetts, USA. Watson, C. J. W., A. A. Beheler, and O. E. Rhodes Jr. 2002. Development of hypervariable microsatellite loci for use in Eastern Phoebes {Sayornis phoebe) and related tyrannids. Molecular Ecology Notes 2:117- 118. Zimmer, J. T. 1938. Notes on migrations of South American birds. Auk 55:405-410. The Wilson Journal of Ornithology 122(2):244-258, 2010 RED-COCKADED WOODPECKER MALE/EEMALE EORAGING DIEEERENCES IN YOUNG EOREST STANDS KATHLEEN E. FRANZREB' ABSTRACT. The Red-cockaded Woodpecker (Picoides borealis) is an endangered species endemic to pine (Finns spp.) forests ot the southeastern United States. I examined Red-cockaded Woodpecker foraging behavior to learn if there were male/female differences at the Savannah River Site, South Carolina. The study was conducted in largely young forest stands (<50 years of age) in contrast to earlier foraging behavior studies that focused on more mature forest. The Red- cockaded Woodpecker at the Savannah River site is intensively managed including monitoring, translocation, and installation of artificial cavity inserts for roosting and nesting. Over a 3-year period, 6,407 foraging observations covering seven woodpecker family groups were recorded during all seasons of the year and all times of day. The most striking differences occurred in foraging method (males usually scaled [45% of observations] and females mostly probed [47%]), substrate used (females had a stronger preference [93%] for the trunk than males [79%]), and foraging height from the ground (mean ± SE foraging height was higher for males [11.1 ± 0.5 m] than females [9.8 ± 0.5 m[). Niche overlap between males and females was lowest for substrate (85.6%) and foraging height (87.8%), and highest for tree .species (99.0%), tree condition (98.3%), and tree height (96.4%). Both males and females preferred to forage in older, large pine trees. The habitat available at the Savannah River Site was considerably younger than at most other locations, but the pattern of male/female habitat partitioning observed was similar to that documented elsewhere within the range attesting to the species’ ability to adjust behaviorally. Received 24 March 2009. Accepted 9 November 2009. The Red-cockaded Woodpecker (Picoides bo- realis), endemic to the pine (Finns spp.) forests of the southeastern United States, is strongly associ- ated with open, mature longleaf pine (P. palnstris) forest (USFWS 2003). Historically, this preferred habitat has been maintained by fire without which the hardwood midstory may develop to make the area unsuitable for Red-cockaded Woodpeckers. A reduction in the frequency of fire (Jackson 1986, Ligon et al. 1986, Walters 1990), loss of longleaf pine forests (Conner et al. 2001), and a general overall reduction in the age of forest stands resulting from short-rotation ages (Wahlen- berg 1960, Frost 1993) have resulted in modifi- cation or loss of habitat throughout the geographic range of the species. Red-cockaded Woodpeckers are cooperative breeders living in social groups consisting of the mated pair and often one or more helpers, which are usually male and the offspring of one or both members of the breeding pair. They construct their own cavities in mature, living pine trees, which are used for nesting and night roosting (Jackson et al. 1979). This is a fairly long-lived species in which all individuals in the group cooperate in tetritorial defense, excavation of cavities, incubation, and feeding nestlings and fledglings (Hooper and Lennartz 1981 ). There is a long period of juvenile dependency, which Ligon ' U.SDA .Southern Research Station, Department of Forestry, Wildlife, and Fisheries, University of Tennessee, Knoxville, TN 37996, USA; e-mail: fran7.reb@utk.edu (1968) speculated was the result of the young having to learn specialized foraging techniques. The Red-cockaded Woodpecker population at the Savannah River Site (SRS), South Carolina, had declined to a low of four individuals by 1986 at which time an intensive management effort was initiated to prevent extirpation of the species at the site. Results and details of the management program that was implemented are available in Franzreb (1997). Most forest stands on the site had been replanted with pine species when this study was initiated in 1992 and —92% of the pines were <40 years of age (K. E. Franzreb, unpubl. data). Previous Red-cockaded Woodpeck- er studies on foraging and habitat use relied on areas that contained a fairly large proportion of older trees in contrast to the younger age and smaller size-class distribution patterns for trees at SRS where old, large diameter trees were sparse. The Red-cockaded Woodpecker population at SRS is intensively managed including: removal of nest site competitors such as southern flying .squirrels (Gktncomys volans) from cavities; trans- locations of woodpeckers from off-site locations to augment the population; translocations on-site to replace lost mates or establish new pairs; and research to examine habitat requirements, home- range size and configuration, and foraging behavior. All birds were monitored, nestlings were banded, and the reproductive rate was estimated for each group on an annual basis (Franzreb 1997). The Red-cockaded Woodpecker 244 Franzreh • RED-COCKADED WOODPECKER F'ORAGING IN YOUNG FORESTS population increased on the site during the course of this study from 36 to 89 birds, including 21 translocated birds from off-site (Franzreb 1997, 1999). Previous investigations by others were con- ducted in habitat largely consisting of stands with relatively mature trees, offering a more structur- ally complex vegetative environment with greater opportunities to partition the habitat than likely found in much younger forest stands. Further, most studies of Red-cockaded Woodpecker for- aging behavior and habitat use have not differen- tiated observations by males and females; those that did noted differences in foraging behavior. Published studies have noted male/female varia- tion in method of foraging, tree species selection, use of hardwoods, foraging height from the ground, foraging substrate, tree height selection, location, and tree condition (Skorupa 1979, Hooper and Lennartz 1981, Rudolph et al. 2007). However, Zwicker and Walters (1999) found no difference in foraging in relation to tree age, diameter at breast height, or species between male and female Red-cockaded Woodpeckers in the Coastal Plain of North Carolina. Porter and Labisky (1986) suggest that differing foraging requirements in different habitats necessitate habitat-specific forest management guidelines for Red-cockaded Woodpeckers. The objectives of my study were to: (1) examine habitat partitioning between foraging male and female Red-cockaded Woodpeckers in a habitat that largely consisted of stands of trees that were considerably younger than most habitat found elsewhere in the range; (2) compare habitat availability with foraging use for males and females; (3) assess differences among groups in selection and preferences on the SRS; and (4) discuss the possible role younger forest stands may have in the range-wide management and conservation of the species. METHODS Study Area. — ^The Savannah River Site, a National Environmental Research Park, is in the Upper Coastal Plain Physiographic Region in Aiken, Allendale, and Barnwell counties. South Carolina, USA. Most of the land was acquired by the Department of Energy (DOE) in the late 194()s and early 1950s to develop into a nuclear production facility. Most of the site was in agricultural use or recently had been harvested for timber prior to acquisition by DOE. The USDA, Forest Service has managed the natural resources of the site for DOE since 1952. The area managed for woodpeckers during this study contained 31,970 ha of pine forest consisting ot longleaf (37.7% of the pine), loblolly (P. taeda, 45.4%), slash (P. elliottii, 13.4%), and other pines (0.2%), in addition to pine-hardwoods (3.3%) (Glenn Gaines, U.S. Eorest Service, unpubl. data). Most of the existing pine stands are the result of replanting efforts undertaken in the 1950s, although there are some residual older pine trees. Approximately 92% of the pine trees were <40 years of age (K. E. Eranzreb, unpubl. data). Foraging Data Collection. — ^Each Red-cockad- ed Woodpecker on the site was banded with a unique set of colored plastic leg bands for field identification and also with a numbered aluminum leg band provided by the U.S. Geological Survey. I obtained the necessary endangered species permits and banding permits from the U.S. Eish and Wildlife Service, U.S. Geological Survey, and State of South Carolina. I selected seven groups of Red-cockaded Woodpeckers for observation on SRS based largely on their history of successfully breeding, increasing the likelihood the group would be present for the entire length of the study. Each group consisted of a breeding male and breeding female and often at least one helper. There was no turnover in either member of the breeding pair for two groups (groups 3 and 6). The male breeder also remained the same for three of the other groups (groups 1, 4, and 7). The breeding female varied from a group that had no turnover during the study, to changing every breeding season (group 2). Group composition during most years included at least one helper and there was no group that did not have a helper at least one of the years. The Red-cockaded Woodpecker population on the site during this study increased from 36 to 89 birds, including 21 translocated individuals from off-site (Eranzreb 1997, 1999). Foraging observations were obtained from 5 May 1992 to 26 July 1995. Each foraging observation consisted of; individual identification (band colors), date, time, tree species, tree height (m), tree condition (alive or dead), diameter at breast height (dbh) (cm), tree age, foraging method, foraging substrate (trunk, live limb, dead limb, cone), foraging height from the ground (m), timber stand location (compartment and stand number), and weather conditions (percent cloud 246 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 2, June 2010 cover, wind conditions, precipitation). Tree age, height, and dbh are likely to be positively correlated. These three variables are included because forest managers use them in developing management plans and there is a large amount of variation depending on site quality and physio- graphic region. The location of each observation was defined using a global positioning system (GPS) Trimble Pathfinder Professional (use of brand names conveys no recommendation by the U.S. Forest Service). Observations were obtained for all members of the Red-cockaded Woodpecker groups. Data were collected during all seasons of the year and all times during the day. Birds were observed from the time of leaving the roost cavity in the morning and, if possible, until they returned to the cavity at night. Each bird located was followed until it made a foraging “strike” (defined as an actual attempt to collect prey) at which time the first observation was taken. At least 15 min separated sequential observations of the same individual to minimize inter-dependence of subsequent observations (Hejl et al. 1990). Porter et al. (1985) reported that a 15-min interval between sequential observations was sufficient for independence for Red-cockaded Woodpeckers. I accounted for possible correlation among obser- vations of the same individual instead of relying on the informal 15 min rule. Vegetation Data. — I used the Continuous In- ventory of Stand Conditions (CISC) data base for information on stands such as dominant tree vegetation, stand age, and stand condition on the SRS (U.S. Forest Service, unpubl. data). A stand is a contiguous group of trees sufficiently uniform in age-class distribution, composition, and struc- ture growing on a site of sufficiently uniform quality to be a distinguishable unit (Helms 1998). More detailed information was obtained by sampling vegetation in all stands within 800 m of the nest tree of each of the Red-cockaded Woodpecker groups studied. Most foraging ob- servations occurred within 800 m of the nest tree of each group. Hardwood and pine trees in each stand were sampled using 0.04-ha fixed-radius plots established along transect lines at a rate of one plot per 2 ha of stand size following James and Shugart (1970). The beginning of each transect in a stand was established from a random point. Sample stations were separated by 50 m and transects were at least 50 m apart. The following variables were measured in each 0.04- ha plot for every tree: tree species, tree height (m). tree condition (alive, dying, or dead), diameter at breast height (dbh) (cm), age (increment bore reading of 4,945 trees including 1,061 hardwoods and 3,884 pines or CISC for even-aged pine stands), location (compartment/stand/Red-cock- aded Woodpecker group number), and specific plot location. Statistical Analyses. — A randomized block ex- perimental design was used with repeated obser- vations where blocks are the Groups and treat- ment is Gender, which is tested with the Group*Gender interaction as the error term. Observations of each individual bird were repre- sented as BirdID (individual band of a bird) and nested within each Group*Gender. This is equiv- alent to a randomized block repeated measures design where observations of a given bird follow a compound symmetry covariance matrix (common variance and common covariance). There were numerous observations of the same individual during the 3-year study period, often ranging to over 100 observations of certain individuals. I used Proc Mixed (SAS 2004) for analysis that considered nesting or clustering of repeated observations for a given bird. This weighted the observations of individual birds so those with high numbers of observations would not unduly influence the outcome and bias the results (Rao and Scott 1992). This ensured that error terms and variances would be calculated correctly, yielding valid tests of hypotheses. 1 obtained least-squares means (hereafter referred to as means) for all continuous variables (foraging height, tree height, tree dbh, and tree age) for males and females in each group and for all groups combined. I tested differences between means using an F-statistic based on the mixed model using Proc Mixed. Individuals with fewer than five observations over the 3-year period were deleted from the analysis, which affected the degrees of freedom. Groups with the highest number of observations tended to have the highest number of birds with observa- tions of more than five and these groups had the largest degrees of freedom indicated by the denominator term. I examined each group sepa- rately as well as combining data for all males and for all females. The amount and distribution of the habitat described by the variables studied differed within the home ranges of each group. Compar- isons of the data on a per group basis provide a more complete representation of the range of foraging behavioral differences for males and females. Frcinzreh • RED-COCKADEI) WOODPEC KER EORACJINCi IN YOUNCj l-ORESTS 247 I combined all observations of all pine trees regardless of species and then grouped them by age. height, and dbh class interval to allow for comparison of the distribution of observations for the continuous tree variables by gender. Tree height and foraging height data were grouped by 3-m intervals. Age class data were grouped into 10-year intervals. There were few observations in trees <30 years of age and the age class intervals <30 were grouped. Tree dbh data were grouped by 5-cm size class intervals. The distribution of observations of males versus females was com- pared for these variables as well as for the other categorical variables (method, substrate, tree species, and tree condition) using the T-statistic in the Rao-Scott modified likelihood ratio test as obtained from Proc SurveyFreq (SAS 2004) for each group, and for males and females of all groups combined. This considered the nesting of the repeated observations for a bird in much the same way as the mixed model did using Proc Mixed for the other continuous variables. The Rao-Scott modified likelihood ratio test uses the survey design to adjust the typical Pearson Chi- square test by dividing the P-value by the survey design effect. The F-statistic used is an additional modification that provides a better statistic as suggested by SAS (2004). Degrees of freedom varied depending on the number of categories compared. Any cells with fewer than five observations were merged for these tests with the closest adjoining cell until there were at least five observations. The clustering process in the Rao-Scott tests also affected degrees of freedom. I examined tree composition within an 800-m radius circle with a group’s nest tree as its center. The value of 800-m was .selected because it was the same radius used by USFWS in its recom- mended foraging guidelines for this species. I estimated how many pine and hardwood trees of specific ages, heights, and dbh were available within each of these stands or portions thereof using the 0.04-ha plot vegetation information. I then estimated the availability of the resources (tree species, tree heights, dbh, and tree age) within the 800-m radius circle. These availability data were used to derive estimates of how many observations were expected to occur if birds were using the foraging resources randomly. Arcinlo coverages were used to aid in this process. The percent use by each gender in each Red-cockaded Woodpecker group of pine trees by tree species and by tree age, tree height, and dbh class intervals was compared to expected u.se based on availability of the.se trees within the 80()-m radius circle. Males and females in each group, and all males versus all females combined were compared with respect to the distribution of actual observations versus the number ol ob.servations expected. Expected values were based on resource availability within 800 m of the nest tree and these comparisons were used to examine any preference or avoidance of foraging resources. Any cells that contained <5 observations were either deleted or grouped with similar cells so the overall number of observations in each cell was >5. I compared means for tree age and dbh for each pine species for trees that were available in the habitat of each of the groups and overall for all groups combined using Proc GEM and the BonfeiToni adjustment for multiple comparisons (SAS 2004). All F-tests and comparisons of means were performed at the 0.05 significance level. Values reported are means ± SE. I used Levin’s (1968) measure of niche breadth where B = l/Ep", and /;, is the proportion of observations recorded in class I followed Schoener (1970) to estimate niche overlap for each variable: % overlap = 100( 1 — 0.5 Z I F^., — Pyj\) where P^j and F^, are the respective frequencies for males and females in each class for a given type of behavior. An overlap of 100% indicates that males and females acted identically in regard to the type of behavior examined, whereas 0% overlap indicates completely non- overlapping behavior. RESULTS There were 6,407 foraging observations of all seven groups ranging from a low of 794 in group 7 to a high of 1,152 in group 2. Observations of males (/; = 3,769) ranged from 403 in group 7 to a high of 576 in group 6. Observations of females (/? = 2,638) ranged from 269 (group 4) to 502 (group 2). Method of Foraging. — Scaling was the most commonly used foraging method for males in all seven groups, ranging from 40 to 52% of the total observations (// = 3,769; Eig. 1 ), followed closely by probing. However, females in five of the seven groups used probing more commonly than scaling (/? = 2,638; Fig. 1 ). Foraging method significant- ly differed between males and females for all groups combined, primarily because females tended to probe more {F^ 2')i — 2.89, F = 0.04). 248 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 2, June 2010 Group Number 05 c g ro 5 0) o c Q) L 05 CL FIG. 1. Foraging method of male and female Red-cockaded Woodpeckers at the Savannah River Site, South Carolina, 1992-1995. Suhso-ate. — Males strongly preferred the trunk as the foraging substrate with 73 to 87% of observations occurring there (Fig. 2). For all groups combined, 79% of all male foraging was on the trunk surface, followed by 10% on a live limb, and 1 1% on a dead limb (Fig. 2). Females in each group had a stronger preference for foraging on the trunk than males and used that surface a minimum of 87% of the time (Fig. 2). Overall, 93% of all female foraging observations were on the trunk, followed by 4% on a live limb, and 2% on a dead limb. Occasionally males and females were observed foraging on cones (19 instances for males, 7 instances for females), and there was one instance of a female foraging on the ground. There was a significant difference in foraging substrate .selection in every group and for all groups combined between males and females (all F values > 3.79, all P < 0.03). Tree Species Selection. — At least 98% of all foraging by males and females, regardless of group, occurred in pine trees. Overall for males there were 27 (0.7%) instances and up to 1% of observations for groups foraging in hardwood trees. Females were observed foraging in hard- woods on 20 instances (0.8%) and 2% of observations by group. There was a decided preference by both males and females to forage in longleaf pines and this did not .seem to be the result of tree dbh or age. The mean overall dbh values for pine species showed that, on average, longleaf pine trees available at SRS were 1 .0 cm larger than loblolly pines and 2.8 cm larger than slash pine trees. Mean tree age of longleaf pine exceeded that of loblolly pine by 0.6 to 8.5 years but, in some groups, the mean age of the available loblolly pines exceeded that of longleaf by 0.3 to 5.0 years. For example, in group 1, loblolly pines on average were bigger (larger dbh) and older than longleaf pines. The latter species was most frequently used for foraging by males in all groups (range 69 to 90%) and accounted for 79% of all male observations (n = 3,743). Females also preferred to forage in longleaf pine trees, in which 78% of ob.servations occurred (range 62-94% in the 7 groups). Loblolly and slash pine also were used, but to a much less extent. There was a significant difference in tree species selection by males and females in two of the seven groups (F values ^ 4.60, P values < 0.02); however, considering all male and female observations combined, there was no significant difference in tree species selection (F2.196 = 0. 1 1, F = 0.91). Tree species selection by males differed from expected on the basis of tree species availability in all groups and for all males combined (all F > 4.23, all P < 0.04). Females in all but group 2 did not use tree species in the same proportion as available (all F > 4.79, all P < 0.03). Males in all Franzrcb • RED-COCKADEI) WOODPECKER FORAGING IN YOUNG FORESTS 249 Group Number I I Trunk v//\ Live limb I I Dead limb HTTTTTI Cone FIG. 2. Substrates used by foraging male and female Red-cockaded Woodpeckers at the Savannah River Site, South Carolina, 1992-1995. groups and females in five of the seven groups foraged more in longleaf pines than expected based on availability. Foraging Height. — ^Mean foraging height of males ranged from 9.5 ± 0.6 m in group 1 to 12.7 ± 0.4 m in group 5 (Table 1). Mean foraging height of females ranged from 7.5 ± 0.5 m in group 1 to 11.8 ± 0.4 m in group 5 (Table 1). Mean foraging height for males in two groups was higher than for females {F > 6.0, P ^ 0.03; Table 1). Foraging height for males for all groups combined was higher (11.1 ± 0.5 m) than for females (9.8 ± 0.5 m; F\ (, = 17.67, P = 0.006; Table 1). Tree Height. — ^Mean heights of trees selected for foraging by males ranged from 17.6 ± 0.6 m in group 1 to 21.5 ± 0.3 m in group 5. Mean tree height selection by females ranged from 17.1 ± 0.6 to 21.3 ± 0.3 for groups 1 and 5, respectively. Mean tree height selection for all male observa- tions combined (19.6 ± 0.5 m) was not different than for females (19.6 ± 0.5 m; F\fi = 0.02, P = 0.90). There was no difference in mean tree heights between males and females in five of the seven groups (all F 0.42, all P ^ 0.53) or between males and females within any group [F > 4.49, p > 0.05). There was no difference for five of the groups comparing the distribution of observations for male versus female tree height selection (all F < 1.39, all P ^ 0.24). Nor was there a difference between males and females for all groups combined in distribution of observations by tree height (^7.693 = 0.28, P = 0.96). Males and females in every group and for all groups combined did not randomly select trees by height (all F > 10.56, all P < 0.001). Both males and females avoided shorter trees (those less than 12 m), either avoided or used in proportion to availability mid-sized trees (12-18 m), and strongly prefen'ed trees >18 m in height. Tree Condition. — Both males and females heavily foraged on live trees (average of 90% of male and 91% of female foraging observations). There was no difference in selection of trees by condition (alive, dead, or dying) between males and females in five of the groups and for all groups combined (F s 2.33, P ^ 0.12). There was a difference between males and females in two groups (F2,34 = 6.06, P = 0.006; Fj 12 = 53.2. P < 0.0001; groups 2 and 7, respectively) and males in both groups used dying trees more than females. Diameter at Breast Height (dhh). — ^There was a difference in mean diameter at breast height of trees selected by males versus females for two of the seven groups. Females in group 6 (F| 14 = 6.23, P — 0.03) and group 7 (Fi n = 9.22, P = 0.01) selected trees with larger dbh than males. Overall, there was no difference in mean dbh of trees selected for foraging by males and females in all groups combined (Fj ^, = 0.39, P = 0.55). 250 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 2, June 2010 TABLE 1. Foraging height (m) of male and female Red-cockaded Woodpeckers by seven groups and all groups combined at the Savannah River Site, South Carolina, 1992-1995. Percent observations Group 1 Group 2 Group 3 Group 4 Foraging height M F M F M F M F <3 4.9 16.1 5.7 13.0 2.3 9.1 4.7 10.4 >3-6 13.5 23.2 15.1 14.3 8.9 16.5 10.3 17.8 >6-9 25.8 25.7 22.0 16.9 16.9 18.2 20.0 21.6 >9-12 26.4 14.9 21.7 20.5 23.0 16.2 24.7 20.1 >12-15 18.4 13.4 16.6 16.1 27.3 17.6 16.1 15.2 >15-18 7.4 5.5 10.8 1 1.6 16.2 15.6 14.6 7.4 >18-21 2.2 1.0 5.9 6.6 4.6 5.4 6.9 5.6 >21-24 0.8 0.0 1.2 0.2 0.7 1.1 1.7 1.5 >24 0.2 0.0 0.5 0.0 0.0 0.3 0.7 0.4 Ground or no data 0.4 0.3 0.6 0.8 0.0 0.0 0.3 0.0 Sample .size 509 398 646 500 561 .352 589 269 LSMeans ± SE 9.5 ± 0.6 7.5 ± 0.5 10.4 ± 0.3 9.8 ± 0.3 1 1.7 ± 0.7 9.0 ± 0.8 1 1.1 ± 0.5 9.6 ± 0.7 F value” T,.n = 6.54 F 1.16 ” = 1,67 F^.u = 6.00 T|,,4 = = 3.00 P = 0.03 P -- = 0.21 P = 0.03 P -- = 0.10 Group 5 Group 6 Group 7 All groups combined M F M F M F M F 1.9 6.2 7.8 13.4 3.2 12.0 4.5 11.5 3.4 8.5 14.8 22.6 9.2 14.3 1 1.0 16.3 1 1.5 14.7 17.4 14.7 22.3 20.5 19.5 18.8 23.9 23.9 22.7 20.6 26.1 20.0 23.9 19.5 32.1 23.2 20.3 15.1 20.1 15.9 21.3 16.8 18.5 15.6 10.8 9.6 10.7 9.2 12.7 10.9 6.5 5.8 5.4 3.1 6.5 5.9 5.4 5.0 1.1 0.5 0.4 1.0 1.7 2.3 1.1 0.7 0.2 0.0 0.2 0.0 0.3 0.0 0.3 0.1 1.1 1.6 0.4 0.0 0.0 0.0 0.4 0.4 472 428 575 292 402 390 3,754 2,629 12.7 ± 0.4 1 1 .8 ± 0.4 10.5 ± 0.6 9.4 ± 0.8 1 1.9 ± 0.9 10.8 ± 0.84 1 l.l ±0.5 9.8 ± 0.5 T,.,3 ^ = 2.38 F 1.14 = = 1.22 T,,n ^ = 0.78 T,,6 = 17.67 P ■■ = 0.15 P = = 0.28 P = 0.40 P = 0.006 F and df values from Proc Mixed model; Note for dl ot x and y as given in Fy^ y. x represents k — I and _v is calculated by the clustering program There was a difference between males and females in two of the groups when comparing tree selection by distribution of observations by dbh size class with females using trees with larger dbh (group 6, = 3.78, P = O.OOl ; and group 7, T7_s4 = 5.39, P < 0.0001 ). There were no male/ female differences in distribution of observations for dbh size class intervals in the other five groups and for all groups combined (all F values < 1.31, P > 0.26). Males and females did not randomly select trees on the basis of dbh (all F > 9.68, all P < O.OOl ). Both males and females avoided foraging in trees <20 cm dbh and displayed strong preferences for the moderate and larger dbh size classes. Longleat and loblolly pine trees were the two most common pine species in the habitat. The mean dbh for longleaf pine trees available to three groups was higher than for loblolly pine (Ta- ble 2); however, loblolly pine for two groups had higher mean dbh values (all P < 0.001). There was no difference in mean dbh values between longleaf and loblolly pine trees in the remaining two groups (Table 2). The available longleaf pines had a higher mean dbh (longleaf pine •= 17.9 ± 0.2 cm vs. loblolly pine = 16.8 ± 0.2 cm, P = ().()()4; Table 2) for all groups combined. Tfcc Age. — Females in group 5 (mean = 39.9 ± 0.2 years vs. mean = 39.3 ± 0.2 years; F\ = 5.27, P = 0.04) foraged more on older trees than males (Table 3). There was no difference between TABLE 2. Tree species availability (least-squares means ± SE) for tree age (yrs) and diameter at breast height (dbh) (cm) for foraging Red-cockaded Woodpeckers at the Savannah River Site, South Carolina, 1992-1995. Franzreh • RED-COCKADED WOODPECKER FORAGING IN YOUNG FORESTS 251 — u > OO O OS O p lo o cN r- lO Tf so o — — Os — (N d ^ so sd m m o — 00 p — — Tt so r-1 o II V II II II II II V II 11 V II II V II II II II II II II II k. a, s: Ci. a, s: k. a. =: k, k. =: k. k, C k. n F < u CN u p d p "r3 > OO D r3 > D CJJ r3 ■U GJ GJ Cij >% GJ O Ci) £ "o CJ GJ GJ -1^ O GJ CL ■o c "3 GJ *CJ r3 O o 0 1 •a G> c -3 C GJ 3^ I r I o^ UJ -J 6 CO E ^ C U o vO (N O On CN O r<^ CN (N CN (N On 00 OO O O SO m r<) O m* ON NO r- o o r-- ON II p p p p p m q o p d d d II m in mi 00 NO o d o m. NO + 1 II II II II (N rA [y o (N GN d in 1.14 cu vr, r^’, [> r*\ o Jj O d p N<. -2 r3 V a E II p q o ON — o o o — E a. ,GJ a. Tf — O — ^ r-' vO Tf Ov o ON VY — On O -— rj n (N o o o (N +1 ON IT) 00 00 o d r- — o odd — oo oo sD cn r^> 00 c^i c d m O m O rn d iri sOOOsqONr'if^Oir) o d (N — r-’ — d d p On NO NO OO (N — O r- (n 00 II +1 d d d d II o oo II II II II X fNj o rr o d in (N a. in vq 0. d o »n d -2 d (n ti. GJ V C3 V "£ E + 1 E a. cx. nO Os — ITi so ON — 00 sd — * (^1 IT'j ^ o o d d d NO m -s 1 mi — ■ 4 d d d +1 II II II II rj /— N (N -— Q. 3 d d d d NO tri d d d p p o d d d iri r-- r^i NO nO -t — 00 d — r^i o d lO +1 o ON iT'i 00 0^ u. Z1 O Tf O k. c _b' d ^ V E V E 0, k. UJ 00 c C3 a £ > GJ 2 I o o ? t GJ +1 £ C3 E y> C GJ u. ■3 o o o o o -2 Cl c3 GJ C3 CL cn GJ in nO r- 00 J. ON o ON £ C3 S 00 E o CxC. in NO r-- 00 A m J U T3 G^ O D a. X GJ 3 ro — — 00 d lO oS 00 (N rj r-; p p d d — — d r-;rsj(^|pON'^(NOro — dddrsiodrodro ro — (N — (N ro OO d d NO r- rN — 00 ro 00 — (N o *r^o NoddroodcNrod— • m ro pooor-r-(NOO ddddd — lodd c3 iT^ £ ^ G eCJ in o d d o o o d V p O) o d II o o o +1 II O. d p d d d rt II On r- o II 5C r^i d d — d (N ti. o o +1 II d d > _ (u 2 I ^ C cz cz cz (U Q. 2 E CO O -J U 7. ^ a ^ 'Z ■S S S .2 2 e i ji -d d JZ U X) w c CL = X 3 LU Z Eqj X = E o ■£ r3 Qi 3 C«i x -j 2 males and females for six of the seven groups and for all groups combined in tree age classes (all F < 1.79, all P > 0.13, Table 3). Neither males nor females randomly selected trees by age class (all f > 4.25, all P < 0.002). Both males and females used younger trees (<30 years of age) less than expected and preferred trees >30 years of age (Table 3). Mean age was higher for longleaf pine than loblolly pine in three groups {P < 0.0001) within the 800-m radius circles centered on the nest tree of each group with the reverse true in two groups {P < 0.0001; Table 2). There was no difference in mean age between available longleaf and loblolly pines in groups 2 and 7 (Table 2). Mean tree age of longleaf pine (26.2 ± 0.2 years) was higher compared to loblolly pine (23.3 ± 0.3 years, P < 0.0001; Table 2) combining the 800-m circles of all groups. Niche Breadth and Overlap. — ^Foraging vari- ables that had the narrowest niche breadth were the same for males and females: tree species selection (1.54 vs. 1.57), tree condition (1.24 vs. 1.20), and substrate (1.56 vs. 1.15). Both males and females were most generalized in tree dbh (6.25 vs. 6.54), foraging height (5.78 vs. 6.41), tree height (4.85 vs. 4.91), and tree age (5.55 vs. 4.55). Niche overlap between males and females was >90% for foraging method (90.8%), tree species (99.0%), dbh (95.7%), tree height (96.4%), tree condition (98.3%), and tree age (90.9%). The least overlap between males and females was for substrate (85.6%) and foraging height (87.8%). DISCUSSION Results of foraging ecology studies of male and female Red-cockaded Woodpeckers in regard to foraging method have not been consistent (Ramey 1980, Hooper and Lennartz 1981, Rudolph et al. 2007). In my study, males scaled more than females in two groups, whereas females engaged more in probing. Red-cockaded Woodpeckers at SRS did not forage by excavating to the same extent observed at other locations, all of which had more mature pine habitat. It may be that young stands at SRS did not provide the deep, furrowed bark of older trees that would encourage more excavation. Culmen length is significantly longer in males than females (Mengel and Jackson 1977), which may contribute to the differences that have been observed. However, the Red- cockaded Woodpecker is one of the least sexually 254 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2, June 2010 dimorphic species of Picoides and, although difference in beak size could be an adaptation for differential foraging behavior, culmen length would seem to be insufficient to obviate the need to have other means to reduce intraspecific competition. It has been reported that both males and females forage mostly on the trunk (Ligon 1968, Skorupa 1979, Ramey 1980, Hooper and Lennartz 1981). My results showed that males spent more time foraging on limbs than females, whereas females more frequently foraged on trunks. Ramey (1980) found variation in foraging site was the main difference between males and females, and there was substantial foraging in hardwoods, particularly by males. Pizzoni-Arde- mani (1990) reported that male Red-cockaded Woodpeckers have longer beaks, toes, and legs, and shorter tails than females, which may be advantageous when clinging to branches and twigs on limbs. The shorter legs and longer tail of females may be beneficial when foraging on trunks (Pizzonni-Ardemani 1990), the common female foraging substrate. Red-cockaded Woodpeckers forage on living pine trees and rarely u.se non-pine substrates (Ligon 1968, Hooper and Lennartz 1981, De- Lotelle et al. 1983, Porter and Labisky 1986); however, DeLotelle et al. (1983) also found that 10% of foraging occurred in pond cypress (Taxodium ascendens) stands. Both males and females at SRS heavily used pine trees and at least 98% of all observations occurred there, similar to that reported by other studies (Ramey 1980; Zwicker and Walters 1999; Walters et al. 2000, 2002; Rudolph et al. 2007). 1 detected no difference between males and females in four of the groups and in all groups combined in u.se of different pine species. However, based on availability, both males (all groups) and females (5 of 7 groups) foraged in longleaf pine trees more than expected. Previous work on tree species selection produced conflicting results as to whether Red-cockaded Woodpeckers prefer a particular pine species (Nesbitt et al. 1978, Porter and Labisky 1986). Tree species selection can be confounded by tree age and size, presence of a hardwood midstory, and density of trees in the area. All studies that examined foraging height of Red-cockaded Woodpeckers have reported that males forage higher in trees than females (Hooper and Lennartz 1981, Eng.strom and Sanders 1997, Rudolph et al. 2007). However, in my study this was true of only two of the seven groups and all groups combined. It may not have been apparent that for most of the groups there was no difference in foraging height for males and females if only combined data had been analyzed. Habitat partitioning by foraging height could reduce competition for prey and permit specialization by gender so that trees are used more efficiently. Males are dominant to females and may be using the most productive portions of the foraging substrate. Porter and Labisky (1986) reported that when Red-cockaded Woodpeckers foraged in stands that were usually avoided, they selected trees of greatest height and dbh. Ramey (1980) detected no significant difference in tree heights used by males and females. I also found there was little difference in tree height selection between males and females. Both strongly preferred taller and avoided shorter trees, probably reflecting the larger, more deeply fissured bark surface area of the taller trees. Dying and recently dead pine trees, especially those with southern pine beetle {Dendroctomis frontalis), are an important foraging resource for Red-cockaded Woodpeckers, but the reported response by males/females has not been consis- tent. Hooper and Lennartz (1981), Repasky (1984), and this study found that male and female foraging was similar. Five of the groups and all groups combined in my study had no difference in foraging .selection between males and females based on tree condition. Skorupa and MacFarlane (1976), Skorupa (1979), and Repasky (1984) reported that males foraged more frequently than females on dead and dying trees. The differences in the response may reflect the level and occurrence of insect infestations. Hooper and Lennartz (1981) reported niche overlap between males and females was highest for tree condition (99%) and dbh (97%). Most overlap in my study occurred in tree species (99.0%) and tree condition (98.3%). Foraging height overlap was low (56%) in the study by Hooper and Lennartz (1981), but was 87.8% in my study. Higher niche overlap for foraging height and also similarity in foraging height for five of the seven groups in my study reflected the considerably younger forest than that studied by Hooper and Lennartz (1981) where there was less opportunity for vertical divergence in foraging height between males and females. Frcmzreh • RED-COCKADED WOODPECKER FORAGING IN YOUNG FORESTS 255 Hanula et al. (2000) sampled 300 living pine trees at SRS during the same time frame in which my study was conducted and found that arthropod biomass/tree increased with increasing stand age up to —65-70 years, but the arthropod biomass/ha was highest in the youngest stands. Abundance and biomass of arthropods on each tree trunk was positively correlated with bark thickness and tree diameter, and negatively correlated with basal area (mVha). There was no correlation of diversity, abundance, or biomass with arthropods on the tree trunk with site index, numbers of herbaceous plant genera in the understory, number of herbaceous plant stems, or percent ground cover by herbs. The characteristics of the stands, including average bark thickness and dbh, asso- ciated with greater arthropod abundance and biomass on the bark are positively correlated with tree age, but are subject to change depending on how stands are managed (Hanula et al. 2000). Current management guidelines at SRS include prescribed burning at least 2-3 times/ 10 years in active Red-cockaded Woodpecker clusters, mid- story tree/shrub removal, thinning of overstory pines, suppression of pine beetle infestations, and extensive recommendations for regeneration (Ed- wards et al. 2000). Hooper and Lennartz (1981) and Zwicker and Walters (1999) reported Red-cockaded Wood- peckers preferentially selected the largest trees with respect to dbh with no differences between males and females. My results were similar for five of the groups and for all groups combined. Both males and females consistently selected larger dbh trees based on availability, far more than expected on the basis of their presence in the habitat. Larger trees have more arboreal surface area available in which to forage per visit and probably richer resource patches/unit area than smaller trees. All researchers who have investigated individ- ual tree selection have found that large old trees are preferred over smaller, younger ones (De- Lotelle et al. 1987, Jones and Hunt 1996, Hardesty et al. 1997, Walters et al. 2002). Age and size of pine trees are highly correlated until about age 80 (Platt et al. 1988). It is not clear whether tree size (dbh), age, or both is the more important determinant of tree selection by foraging Red- cockaded Woodpeckers. There was a significant difference between males and females in my study in mean tree age selected in four of the groups and in all groups combined; however, it was not always the same gender that had the stronger selection for older trees. Females tended to select trees with higher dbh values. Females spend more time foraging on the trunks than males, and it is likely this difference reflects females taking advantage of the additional foraging surface area provided on larger trees. Both males and females avoided foraging in trees <30 years of age with some groups not using them. There was a strong preference by both males and females to use trees at least 60 years of age in my study. Males and females still preferred to forage on older trees even though a majority of stands at SRS were considered young. There was little difference in dbh size among the pines at the SRS, but longleaf pines on average were strongly preferred as foraging substrate by both males and females. Loblolly, the second most commonly available pine species within the home ranges next to longleaf pine, tended to be selected against. The mean age of longleaf pine in some groups exceeds that of loblolly pine, but in other groups the opposite is true. Males (69.4% of observations) and females (72.7% of observations) strongly selected longleaf pine trees in relation to the predicted use of longleaf pine (46.2%). Red-cockaded Woodpeck- ers evolved in a mixed-age forest containing only a relatively small proportion of young pine trees (Frost 1993) and in a landscape dominated by longleaf pine; thus, they may be more adapted to foraging on surfaces of old, large pine trees. It is logical to assume they are best equipped morpho- logically and behaviorally to use longleaf pine. The strong preference by both males and females to use longleaf pine does not appear to be strictly the result of tree size or age. Habitat segregation observed in foraging Red- cockaded Woodpeckers may be designed to reduce competition between males and females to allow more efficient use of the available foraging habitat. The more socially dominant males may cau.se females to use less productive areas. It is plausible that partitioning by gender in this species may be an efficient mechanism to facilitate cohesion of the group by reducing male/ female competition and aggression. Breeding males at SRS tended to hold their territories until they disappeared (and were presumed dead). Females were more flexible in maintaining their pair bonds. Thus, it would be beneficial to the group to minimize aggressive encounters and to maximize .social cohesion of group members. This 256 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2, June 2010 may be the mechanism that developed differences between males and females as documented at this younger growth site and elsewhere. Red-cockaded Woodpecker mean reproductive success (fledglings/pair) at SRS ranged from 1.4 to 3.4 with the highest rates (2. 8-3. 4) observed for groups with at least 10,000 pine stems >25.4 cm dbh within 800 m of the nest cavity (Franzreb 2004). Reproductive success was significantly related to number of available pine stems 5:25.4 cm dbh (Franzreb 2004). Typical mean reproductive rates in other parts of the range were 1.34-1.67 fledglings/pair (Conner et al. 2001). Higher rates at SRS, rather than due mainly to habitat quality, are possibly the result of the unusually intensive management of the species on the site, including installation of artificial cavity inserts, removal of cavity competitors, translocat- ing birds to re-establish a pair where one member has been lost, frequent monitoring of cavities to detect damage to artificial inserts, and frequent monitoring of all individuals in the population (Franzreb 2004). Subsequent to initiation of these management activities, the population level has increased to — 196 adults and fledglings in 2006 (John Blake, pers. comm.) from a low of four birds in 1986 (Franzreb 1997). Development of the technique to install artificial cavity inserts for the Red- cockaded Woodpecker was pioneered at SRS (Allen 1991) and installation of the.se devices has been a widely used management tool in many parts of the range. Artificial in.serts at SRS were used by 47% of the birds in comparison to 53% that used naturally excavated cavities for the nest location during my study (K. E. Franzreb, unpubl. data). It is likely that Red-cockaded Woodpeckers would not have persisted at SRS without the assistance provided by these management actions. Current federal guidelines for managing Red- cockaded Woodpecker foraging habitat are com- plex and address factors including type of forest regeneration, rotation age, basal area, and home- range size needed per group (USFWS 2003). Management and recovery of the Red-cockaded Woodpecker have focused on providing suitable foraging habitat in more mature and older pine stands than typically found at SRS. The results of my study indicate that younger forest stands, as long as they are at least 30 years of age, may provide acceptable foraging habitat and should not automatically be regarded as unacceptable or sub-optimal. Younger trees (preferably at least 40 cm dbh) in which managers have either installed artificial cavity inserts or drilled cavities are necessary to provide for nesting and night roosting if older trees are insufficient for natural cavity excavation. It is preferable to provide trees of sufficient size so birds can create their own cavities rather than rely on artificial inserts. Allowing pines to age to >60 years, preferably even older, provides better foraging habitat and trees of sufficient size for birds to excavate their own cavities. Maintaining the site in just young pines will not maintain a self-sustaining population without continued in- tensive management actions as there will be insufficient trees for natural cavity excavation nor would there be the larger trees that are preferred for foraging. The need to provide larger, older pines has been recognized in the SRS management plan for the Red-cockaded Wood- pecker, which calls for a number of proactive actions including a rotation age of 120 and 100 years for longleaf pine and loblolly pine stands, respectively, in the main Red-cockaded Woodpecker management area (Edwards et al. 2000). Older pines will be optimal foraging substrate for both male and female Red-cockaded Woodpeckers and will provide the large trees needed for natural cavity excavation. ACKNOWLEDGMENTS This research was funded by the Department of Energy, SRS, through the USDA Forest Service Savannah River under Interagency Agreement DE-IA09-76SR00056 and is gratefully acknowledged. I am especially grateful to Steven Schulze for field support and for generating the data bases to handle a large volume of behavioral and vegetation data. 1 also thank my field staff whose support has been crucial to the success of the Red-cockaded Woodpecker research at the site. I thank C. H. Greenberg, J. A. Jackson, D. C. Rudolph, and John Blake for reviewing the manuscript and Stan Zarnoch for statistical review. The cooperation of the USDA Forest Service Savannah River is gratefully acknowledged (John Irwin, Elizabeth LeMaster, William Jarvis, Glenn Gaine.s, and David Wilson). In particular, 1 thank John Blake, Savannah River Forest Station, for his long-standing support of my research on the site. LITERATURE CITED Allen, D. H. 1991. An insert technique for con.structing artificial Red-cockaded Woodpecker cavities. USDA, Forest Service, General Technical Report SE-73. Southea.stern Forest Experiment Station, Asheville, North Carolina, USA. Conner, R. N., D. C. Rudolph, and J. R. Walters. 2001. The Red-cockaded Woodpecker: surviving in a fire- Frcnizreh • RED-COCKADED WOODPECKER FORAGING IN YOUNG FORESTS 257 maintained ecosystem. University of Texas Press, Austin, USA. DeLotelle, R. S., J. R. Newman, and A. E. Jerauld. 1983. Habitat use by Red-cockaded Woodpeckers in central Florida. Pages 59-67 in Proceedings of the Red-cockaded Woodpecker Symposium, 11 (D. A. Wood, Editor). Florida Game and Fresh Water Fish Commission, Tallahassee, USA. DeLotelle, R. S., R. J. Epting, and J. R. Newman. 1987. Habitat use and territory characteristics of Red- cockaded Woodpeckers in central Florida. Wilson Bulletin 99:202-217. Edwards, J. W., W. M. Smathers Jr., E. T. Lema.ster, AND W. L. Jarvis. 2000. Savannah River Site Red- cockaded Woodpecker management plan. USDA, Forest Service, Unpublished Report. Savannah River Forest Station, New Ellenton, South Carolina, USA. Engstrom, R. T. AND F. J. Sanders. 1997. Red-cockaded Woodpecker foraging ecology in an old-growth longleaf pine forest. Wilson Bulletin 109:203-217. Franzreb, K. E. 1997. Success of intensive management of a critically imperiled population of Red-cockaded Woodpeckers in South Carolina. Journal of Field Ornithology 68:458^70. Franzreb, K. E. 1999. Factors that influence translocation success in the Red-cockaded Woodpecker. Wilson Bulletin 111:38^5. Franzreb, K. E. 2004. Habitat preferences of foraging Red-cockaded Woodpeckers at the Savannah River Site, South Carolina. Pages 553-561 in Red-cockaded Woodpecker: road to recovery (R. Costa and S. J. Daniels, Editors). Hancock House Publishers, Blaine, Washington, USA. Frost, C. C. 1993. Four centuries of changing landscape patterns in the longleaf pine ecosystem. Proceedings of the Tall Timbers Fire Ecology Conference 18:17^4. Hanula, j. L., K. E. Franzreb, and W. D. Pepper. 2000. Longleaf pine characteristics associated with arthro- pods available for Red-cockaded Woodpeckers. Jour- nal of Wildlife Management 64:60-70. Hardesty, J. L., K. E. Gault, and R. P. Percival. 1997. Ecological correlates of the Red-cockaded Woodpeck- er (Picoides borealis) foraging preferences, habitat use, and home range size in northwest Florida (Eglin Air Force Base). Final Report. Research Work Order 99. Florida Cooperative Fish and Wildlife Research Unit. University of Florida, Gainesville, USA. Hejl. S. j., j. Verner, and G. W. Bell. 1990. Sequential versus initial observations in studies in avian foraging. Studies in Avian Biology 13:166-173. Helms, J. A. (Editor). 1998. The dictionary of forestry. S(x;iety of American Foresters, Bethesda, Maryland, USA. Hooper, R. G. and M. L. Lennartz. 1981. Foraging behavior of the Red-cockaded Woodpecker in South Carolina. Auk 98:321-334. Jackson, J. A. 1986. Biopolitics, management of federal lands, and the con.servation of the Red-cockaded Woodpecker. American Birds 40:1 162-1 168. Jackson, J. A., M. R. Lennartz, and R. L. Hooper. 1979. Tree age and cavity initiation by Red-cockaded Woodpeckers. Journal of Forestry 77:102-103. James, F. and H. Shugart. 1970. A quantitative method of habitat description. Audubon Field Notes 24:727-736. Jones, C. M. and H. E. Hunt. 1996. Foraging habitat of the Red-cockaded Woodpecker on the D’Arbonne Nation- al Wildlife Refuge, Louisiana. Journal of Field Ornithology 67:51 1-518. Levins, R. 1968. Evolution in changing environments. Princeton University Press, Princeton, New Jersey, USA. Ligon, j. D. 1968. Sexual differences in foraging behavior in two species of Dendrocopos woodpeckers. Auk 85:203-215. Ligon, J. D., P. B. Stacey, R. N. Conner, C. E. Bock, and C. S. Adkisson. 1986. Report of the American Ornithologists’ Union Committee for the conservation of the Red-cockaded Woodpecker. Auk 103:848-855. Mengel, R. M. and j. Jackson. 1977. Geographic variation of the Red-cockaded Woodpecker. Condor 79:349-355. Nesbitt, S. A., D. T. Gilbert, and D. B. Barbour. 1978. Red-cockaded Woodpecker fall movements in a Florida flatwoods community. Auk 95:145-151. Pizzoni-Ardemani, a. 1990. Sexual dimorphism and geographic variation in the Red-cockaded Woodpeck- er. Thesis. North Carolina State University, Raleigh, USA. Platt, W. J., G. W. Evans, and S. L. Rathburn. 1988. The population dynamics of a long-lived conifer (Pinus palustris). American Naturalist 131:491-525. Porter, M. L. and R. F. Labisky. 1986. Home range and foraging habitat of Red-cockaded Woodpeckers in northern Florida. Journal of Wildlife Management 50:239-247. Porter, M. L., M. W. Collopy, R. F. Labisky, and R. C. Littell. 1985. Foraging behavior of Red-cockaded Woodpeckers: an evaluation of research methodolo- gies. Journal of Wildlife Management 49:505-507. Ramey, P. 1980. Seasonal, sexual, and geographic variation in the foraging ecology of Red-cockaded Woodpeck- ers {Picoides borealis). Thesis. Mississippi State University, Starkville, USA. Rao, j. N. K. and a. j. Scott. 1992. A simple method for the analysis of clustered binary data. Biometrics 48:577-585. Repasky, R. R. 1984. Home range and habitat utilization of the Red-cockaded Woodpecker. Thesis. North Car- olina State University, Raleigh, USA. Rudolph, D. C, R. N. Conner, R. S. Schaefer, and N. E. Koerth. 2007. Red-cockaded Woodpecker foraging behavior. Wilson Journal of Ornithology 1 19:170-180. SAS Institute Inc. 2004. SAS/STAT 9.1. User’s guide. SAS Inslitute Inc., Cary, North Carolina, USA. Schoener, T. W. 1970. Non-synchronous spatial overlap of lizards in patchy environments. Ecology 51:408-418. Skorupa, j. P. 1979. Foraging ecology of the Red- cockaded Woodpecker in South Carolina. Thesis. University of California, Davis, USA. Skorupa, J. P. and R. W. McFarlane. 1976. Seasonal variation in foraging territory of Red-cockaded Woodpeckers. Wilson Bulletin 88:662-665. U.S. Fish AND Wildlife Service (USFWS). 2003. Recovery Plan for the Red-cockaded Woodpecker {Picoides 258 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 2, June 2010 borealis). Second Revision. USDI, Fish and Wildlife Service. Southeast Region, Atlanta, Georgia, USA. Wahlenberg, W. G. 1960. Loblolly pine: its use, ecology, regeneration, protection, growth and management. Duke University School of Forestry, Durham, North Carolina, USA. Walters, J. R. 1990. Red-cockaded Woodpecker: a ‘primitive’ cooperative breeder. Pages 69-101 in Cooperative breeding in birds (P. B. Stacey and W. D. Koenig, Editors). Cambridge University Press, London, United Kingdom. Walters, J. R., S. J. Daniels, J. H. Carter III, and P. D. Doerr. 2002. Defining quality of Red-cockaded Woodpecker foraging habitat based on habitat use and fitness. Journal of Wildlife Management 66:1064- 1082. Walters, J. R., S. J. Daniels, J. H. Carter III, P. D. Doerr, K. Brust, and J. M. Mitchell. 2000. Foraging habitat resources, preferences and fitness of Red-cockaded Woodpeckers in the North Carolina sandhills. Fort Bragg Project Final Report. Virginia Polytechnic Institute and State University, Blacksburg, and North Carolina State University, Raleigh, USA. ZwiCKER, S. M. AND J. R. WALTERS. 1999. Selection of pines for foraging red-cockaded woodpeckers. Journal of Wildlife Management 63:843-852. The Wilsini Joiinicil of Ornithology 1 22(2):259-272, 2010 FORAGING HABITS AND HABITAT USE BY EDIBLE-NEST AND GLOSSY SWIETLETS IN THE ANDAMAN ISLANDS, INDIA SHIRISH S. MANCHl'"' AND RAVI SANKARAN' - ABSTRACT. — Foraging habits and habitats of exclusive aerial insectivores, the Edible-nest Swiftlet (Aerodramus fitciphagus inexpectatiis) and Glossy Swiftlet {Collocaliu escidenta affinis), were studied in Andaman Islands, India. Observations were made during January to June 2004 between 0500 and 1 800 hrs at four locations in the forest and on open paddy land. Edible-nest and Glossy swiftlets, respectively, spent (.v ± SD) 17.2 ± 1 1.4% and 25.8 ± 15.6% of their time foraging with significant temporal variations. Glossy Swiftlets had spatial variations in twist, flutter, and tail-wing-open foraging maneuvers. This species also had diurnal variations in flock size, which were positively correlated with feeding attempts. Both swiftlets shared all microhabitats except Inside Forest Canopy and Inside Stream Bank Canopy. Microhabitat use did not vary significantly in Below Stream Bank Canopy, >10 m Above Forest Canopy, >30 m Above Ground, and Above Forest Canopy for Edible-nest Swiftlets. Inside Forest Canopy and Inside Stream Bank Canopy categories for Glossy Swiftlets were relatively important in descending order. Deforestation near and distant from caves used by swiftlets for breeding in the islands can severely affect the wild population of both species. Received 2 September 2009. Accepted If January 2010. Swiftlets (Apodidae) are exclusively aerial insectivores and land only on their nesting and roosting sites in caves or cave-like conditions (Medway 1962, Charles 1987, Chantler and Driessens 2000, Nguyen et al. 2002). Swiftlets have wings well adapted for long and fast flights (Norberg 1986, Chantler 1999). However, a single observation of a Glossy Swiftlet {Collocalia esculent a) perching on a tree was reported by Spennemann (1928). Detailed information about aerial and foraging habits, and habitat and microhabitat requirements of most swiftlet species is .scarce. Swiftlets have large foraging ranges and travel far from caves used for breeding. Medway (1962) observed Aerodramus maximus and A. salanganus flying >24 km from their breeding caves. Swiftlet abundance has been mostly studied in different habitats or air spaces whereas their behaviors have been briefly studied (Medway 1962, Harrison 1972, Hails and Amirrudin 1981, Waugh and Hails 1983, Tarburton 1986, Charles 1987, Lim and Cranbrook 2002, Nguyen et al. 2002). Most studies have recorded the occurrenc- es of swiftlet species in different habitats and altitudes in relation to breeding seasonality, and weather conditions with information on diet and general foraging habits (Medway 1962, Harrison 1972, Langham 1980, Hails and Amirrudin 1981, Nguyen 1983, Waugh and Hails 1983, Laurie and ' Division of Conservation Ecology, Salim AH Centre for Ornithology and Natural History, Anaikatty P. O., Coim- batore- 641 108, India. ^Deceased 17 January 2009. ^Corresponding author; e-mail: ediblenest@gmail.com Tompkins 2000, Lim and Cranbrook 2002, Nguyen et al. 2002). The breeding biology and breeding habitats of the economically valuable Edible-nest Swiftlet (A. fuciphagus) are well documented in different parts of their ranges (Lim and Cranbrook 2002, Nguyen et al. 2002). India has four species of swiftlets; Indian Swiftlet (A. unicolor), Himalayan Swiftlet (A. hrevirostris). Edible-nest Swiftlet, and Glossy Swiftlet. The later two species are distributed in the Andaman and Nicobar Islands. The Edible- nest Swiftlet {A. fuciphagus inexpectatiis), endem- ic to the Andaman and Nicobar Islands, is known to build highly commercial white edible nests exclusively made up of its saliva (Medway 1963, Lau and Melville 1994, Chantler 1999) inside limestone caves throughout the island arc. The Glossy Swiftlet (C. escidenta affinis), also en- demic to the Andaman and Nicobar Islands, commonly constructs its nest using saliva as the cementing material to bind nesting materials including mosses, leaves, flowers, and Casuarina needles inside caves, suitable houses and build- ings, under jetties, etc. (Medway 1963, Chantler 1999). Uncontrolled exploitation of nests caused a decline of —80% in the Edible-nest Swiftlet population between 1990 and 2000 (Sankaran 2001). In-situ and the e.x-situ conservation pro- grams for the species were initiated at selected sites to aiTest this population decline in the Andaman and Nicobar Islands. The in-situ population of the Edible-nest Swiftlet is being protected during the breeding season. The ability to breed in diverse habitats made the Glossy Swiftlet an important part of the ex-situ conser- 259 260 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2, June 2010 vation program for fostering the Edible-nest Swiftlet (Sankaran and Manchi 2008). The Edible-nest Swiftlet and Glossy Swiftlet of the Andaman and Nicobar Islands have not been studied extensively except for status surveys and brief information on their breeding seasonality (Sankaran 1995, 1998, 2001). Our objectives were to study: (1) aerial and foraging habits, (2) spatial and temporal variations in foraging habits, and (3) habitats with the microhabitat requirements of both species near their breeding sites. METHODS Study Area. — Andaman and Nicobar Islands in the northeastern Indian Ocean are the peaks of a submerged continuous mountain ridge arching from Arakan Yoma in the north to Sumatra in the south, between 06° 45' N and 13° 41' N, and 92° 12' E and 93° 57' E (Manchi and Sankaran 2009). Data were collected near the breeding sites of Edible-nest and Glossy swiftlets at Chalis-ek (a single hillock with a group of inland limestone caves) and Pattilevel (human habitation area adjacent to the Chalis-ek hillock) near Ramnagar, in the southeastern part of North Andaman Island. The Chalis-ek cave complex had 650 breeding pairs of Edible-nest Swiftlets. Eive caves were used by 1 1 7 breeding pairs of the Glossy Swiftlet, along with pairs of Edible-nest Swiftlets. Chota tikree, another small hillock within 0.5 km of Chalis-ek, had 16 breeding pairs of Edible-nest Swiftlets and seven breeding pairs of Glossy Swiftlets in four caves. The Chalis-ek and Pattilevel areas have evergreen, semi-evergreen, and dry deciduous habitats (Sankaran 1998). People inhabiting the area altered part of the forested land into paddy fields. There are two major open paddy fields (within 1-2 km of the Chalis-ek hillock. The remaining forest in the area is under immense pressure of deforestation. Data Collection. — Scan sampling (Altmann 1974) was used to record foraging behavior of Edible-nest and Glossy swiftlets between ()5()() and 1800 hrs (Indian Standard Time). Observa- tions were made at four points every 2 weeks from January to June 2004. Ten-minute observations were made every 15 min. The number of birds was recorded after each observation. There were 2,086 observations in 624 hrs for both species. Activity performed, habitat, microhabitat type, and Hock size were recorded for each encounter. We developed terminology for describing the aerial behavior of swiftlets. Flylglide. — ^The flight with wing beats was recorded as ‘fly’ and flight without wing beat was recorded as ‘glide.’ These two flying patterns are combined as Flylglide. Feeding Attempts (Foraging Maneuvers). — Ob- servations were limited to recording feeding attempts by presuming that different behaviors/ activities were performed to capture different kinds of prey. Five types of behaviors were recorded as Foraging Maneuvers. Twist. — The bird makes a sudden twist while gliding. Fly-pause. — A fast-flying individual occasion- ally makes a sudden pause for a second and then moves on with a slight twist. Roll. — The individual catches prey in its beak and rolls down. Birds were observed rolling down for about 2 sec. Flutter. — A hover performed with a rapid wing beat and a pause in flight. Tail and wing open. — A pre-planned position to capture prey once observed. Wing and tail feathers are stretched while approaching the prey and a small twist or flutter is performed during capture. This event takes at most 2 sec. Call. — ^There are two types of calls: tik-tik-tik and chirk-chirk-chirk made in flight. Follow. — Individuals were observed chasing each other in flight. Preen. — Preening during flight. Defecate. — ^Defecating during flight. Carry Nest Material. — Swiftlets collect and carry nest materials in flight. Major studies related to food preferences (Harrison 1972, Hails and Amin'udin 1981, Laurie and Tompkins 2000, Nguyen et al. 2002) were ba.sed on gut content and bolus collection; these were not attempted in our study to avoid sacrifice of any individual. We presumed that gut content and bolus collection during feeding visits by adults would adversely affect nesting success. The Point-centered Quarter Method (Mitchell 2001) was used to study tree diversity in the forests of Chalis-ek. Habitats were classified following Hails and Amirrudin (1981) with separation of habitats into microhabitats depen- dent on foraging heights in air space. Heights considered for segregating microhabitats differed from those of Hails and Amirrudin ( 1981 ) per the suitability of the study area based on canopy levels. Forest habitat was categorized into four microhabitats: (1) Below Forest Canopy, (2) Inside Forest Canopy, (3) 0-10 m Above Forest Mauchi and Sankaran • SWIITLET FORACHNG IN I'llH ANDAMAN ISLANDS 261 TABLE 1. Activity budgets ( v ± SD, %) of Edible-nest and Glossy swiftlets near breeding caves at Chalis-ek, North Andaman Island, India. Activity Edible-nest Swiftlet n = 719 Glossy Swil'tlet « = l,26.S Mann-Whiiney lesi z r Fly/glide 74.42 ± 16.10 84.60 ± 17.50 -13.609 <0.001 Feeding attempts 14.03 ± 12..30 13.75 ± 17.20 -5.217 <0.001 Call 2.08 ± 5.10 0.67 ± 3. .30 -14.572 <0.001 Follow 9.47 ± 12.10 0.98 ± 4.20 -25.020 <0.001 Canopy, and (4) >10 m Above Forest Canopy. The open paddy lands developed through defor- estation had streams with vegetation along the stream banks. This vegetation was included in the open land habitat which was divided into six microhabitats: (1) 0-5 m Above Ground, (2) 5- 30 m Above Ground, (3) >30 m Above Ground, (4) Below Stream Bank Canopy, (5) Inside Stream Bank Canopy, and (6) 0-10 m Above Stream Bank Canopy. Temperature was recorded using a Zeal thermometer. A rain gauge was placed at the top of the hill in an open canopy area to record the amount of rainfall. Statistical Analyses. — ^Percentage of time spent on each activity was estimated in each set of observations from instantaneous scan samples and the proportion of time spent on different aerial activities was calculated. Behaviors including preening, defecating, and carrying nest material were collectively performed <0.1% of the time by both species, and were not included in the statistical analyses. Variations in activities of Edible-nest and Glossy swiftlets were calculated with Mann-Whitney C-tests (Z). Days were divided into four sets of time: 0500-0815 hrs (early morning), 0815-1130 hrs (late morning), 1130-1445 hrs (afternoon), and 1445-1800 hrs (late afternoon). The data were arcsine trans- formed for one-way ANOVAs. Kruskal-Wallis tests and one-way ANOVAs were performed to examine temporal and spatial variations in activities overall, different behaviors separately, and activities performed in different microhabi- tats. Identification of the differing means was achieved by a Tukey’s test for the detailed activity patterns when significant differences were found. All statistical analyses were also conducted with the data on feeding attempts to derive or understand foraging maneuvering patterns. Pear- son’s correlation test (/) was performed to examine the relation between foraging frequency and foraging flock size. Diurnal variation in size of the foraging flock was also described. An Important Value Index (IVI) of trees was calculated following Mitchell (2001). Chi-.square tests (Crosstab: contingency table) were used to confirm distinctness of the species in terms of microhabitat use. Binary logistic regression anal- ysis was performed to estimate use and relative importance of each microhabitat. All statistical tests were performed using Microsoft Excel 2003 and SPSS software. Version 10.0 (SPSS Institute 1998). RESULTS Edible-nest Swiftlets were encountered 41,959 times in 719 observations with an average (.v ± SD) of 58.4 ±41.6 activities per set. There were 36,906 encounters of Glossy Swiftlets in 1,265 sets of observations with an average of 29.2 ± 36.5 activities per 10-min period. Aerial Activity Budget. — ^Edible-nest and Glossy swiftlets in flight performed fly/glide, feeding attempts, follow, and call in descending order of their frequencies of occuirence. Preen, defecate, and nest material transport were collectively performed for 0.07% of the total activities and were not included in further analyses. The proportions of time by the two species for different activities varied significantly (Table 1). Spatio-temporal Variation in Aerial Activity'. — Edible-nest and Glossy swiftlets had significant temporal and spatial variations in their overall active periods (Eig. 1). The monthly active period varied (/‘ = 105, P < O.OOl) for Edible-nest and Glossy swiftlets (x^ = 23, P < 0.001). The diurnal active period varied for Edible-nest (x' = 55.896, P < O.OOl) and Glossy swiftlets (x' = 8.464, P = 0.037). Spatial variation in active period was also shown by Edible-nest (x* = 58.081, P < 0.001) and Glossy swiftlets (x' = 293.654, P < 0.001). Individual aerial activities of both species, except feeding attempts for Edible-ne.st Swiftlets, had significant monthly 262 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 2, June 2010 Jan Feb Mar Apr May Jun □ Edible-nest Swiftlet I □ Glossy Swiftlet B □ Edible-nest Swiftlet I □ Glossy Swiftlet Indian Standard Time C i □ Edible-nest Swiftlet I □ Glossy Swiftlet Hill base Open land-1 Hill top Open land-ll FIG. i. Spatio-temporal variation (A = monthly, B = diurnal, and C = location; %) in activities of Edible-nest and Glossy .swiftlets at Chalis-ek, North Andaman Island, India. variations (Table 2). All activities had significant diurnal variations except feeding attempts for both species, and call behavior for the Edible-nest Swiftlet (Table 3). All the other activities, unlike OVJ 70 LU 60 (S) CM 4- 50 c 03 S 40 > 30 o < >5 20 10 0 ■ S" ' a ZBI fly/glide and call, had significant spatial varia- tions (Table 4). Foraging Activity Budget. — ^Edible-nest and Glossy swiftlets spent (.v ± SD) 17.2 ± 11.4 and 25.8 ± 15.6% of their time in feeding attempts, respectively. Both species spent sub- stantial time on twist, flutter, tail-wing-open, fly- pause, and roll in descending order. The propor- tion of time spent on different foraging maneuvers varied significantly (Table 5). Spatio-temporal Variations in Foraging Activ- ities.— Both species had significant spatial (Edi- ble-nest Swiftlet: — 61.236, P < 0.001; Glossy Swiftlet: x^ — 84.702, P < 0.001) and monthly (Edible-nest Swiftlet: x~ ~ 19.588, P < 0.001; Glossy Swiftlet: x^ ~ 53.952, P < 0.001) variations in time spent in feeding attempts. However, significant temporal variations across the day were not apparent (Eig. 2). Eoraging maneuvers, except flutter by Edible-nest Swiftlets and roll by Glossy Swiftlets, had significant monthly variations (Table 6), but roll had signif- icant diurnal variations by Glossy Swiftlets (Table 7). Twist alone in both species, and tail- wing-open in Glossy Swiftlets, had significant spatial variations (Table 8). Foraging Flock Size. — ^Edible-nest Swiftlets foraged in significantly larger flocks (T ± SD, 9.4 ± 3.5 individuals; min = 3, max = 17) before returning to roosting caves between 1700 and 1800 hrs (x" = 120.944, P < 0.001). Glossy Swiftlets also formed comparatively large flocks (J ± SD, 5.7 ± 2.4 individuals; min = 2, max = II) between 1700 and 1800 hrs without any significant diurnal variation (Fig. 3). Edible-nest and Glossy swiftlets had significant coirelations (r = 0.400, P < 0.001 and r = 0.307, P < 0.001, respectively), between foraging flock size and frequency of feeding attempts. This correlation did not change when partial con-elation was performed by controlling for month, location, and hours (Edible-nest Swiftlet: r = 0.420, P < 0.001, Glossy Swiftlet: r = 0.370, P < 0.001). Quantification of Vegetation. — Chalis-ek had 63 species of trees in the area. The forest community was dominated by Pterocarpus dalhergioides (IVI = 35.7) followed by Diospyros criimenata (IVl, = 22.4) and Drypetes andamanica (IVl = 21.3). Habitats and Microliahitats. — Foraging individ- uals of both species were encountered significant- ly more in forest than open land (Edible-nest Swiftlet: x' = 94.298, P < 0.001; Glossy Swiftlet: x' = 58.548, P < 0.001; Fig. 4). TABLE 2. Temporal (monthly) variation (%) in activities of Edible-nest and Glossy swiftlets at Chalis-ek, North Andaman Islands, India. Manchi and Sankaran • SWIFTLET FORAGING IN THE ANDAMAN ISLANDS 263 m OJ X fN o o o O o O o •— .— ( 0. o o o o o o o o O o Os fN in* nj g O o in O § o o < d V d d V d V d V d V d V d V d V d d d V d V d d V d V > o < > z < o r-- o 00 so q (N O z m m r- so q m Tf < ro q q q q q q O S q in fN q 00 o in (N d in d in in d C3 u. fN m m in 00 q m q q *■5 c od — — d sd fN d d A CN % JZ \'< U ■o flj (A c O S u > D a. d q d q d q d OO CN OS c5 c/5 o A ± 13.3 ± 10.9 Apr + 1 os + 1 q fN + 1 q +1 q CN II 5: + 1 q + 1 m o + 1 fN II ? c/5 3 o c at m 5 ± 2.7 ± 9.6 = 142 ± 16.1 ± I6.C ± 3.3 ± 3.2 = 299 at in r- d 00 d 00 d d >% A 1 r<5 q 00 s: X q X c: c A sd d X od sd fN d d O I— H 00 ’> O •c u a CN d 00 d m cn q od q od q q 00 m "O c T3 « — — (N — — — — in in fN ’> c c Mar + 1 +1 + 1 + 1 II + 1 + 1 + 1 + 1 II A (U O CO O sd q d oo q os fN sd q in q q (N c a> so in q CN q q d q q c: c t c cd C T + 1 + 1 in + 1 od + 1 II + 1 +1 d + I d + 1 ro II q q X fN q fN in r- q c: O) OS r- S; o CO od d q d od d q O On so (N 00 d — C13 00 in fN d sd fO SO c + 1 + 1 +1 + 1 II + 1 + 1 + 1 + 1 II g »r) ■q q fN s: X q o os k: o d d (N X d in d in q sd os q 00 q oo" (n q O c o oo nj — 00 — q — — d d fN Jan + 1 + 1 + 1 + 1 II +1 + 1 4-1 + 1 II c5 ’C cd (n q q (n q Os q ? > ri d d d d d r-- 00 ' — ' "u, D A c/5 'A c/5 O sz s. £ D, £ c. H. cd B B ’> p B £ o in sO II Lh O a. o in X fN £ .£ r-«. c o < -* (U OJ *o d u. U- U a. s: u- U. U m 5: tu CQ < H OJ B •A at on A. w u g. (A c cn c/5 c 00 JJ dJ c/5 A C/5 X A O o 'B O pa a 264 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 2. June 2010 TABLE 4. Spatial variation (%) in activities of Edible-nest and Glossy swiftlets at Chalis-ek, North Andaman Islands, India. Time .speni on an activity at different locations (.v ± SD) ANOVA Species Activity Hill base Hill top Open land-1 Open land-II F P Edible-nest Fly/glide 74.8 -h 15.1 72.9 -F 15.8 79.3 ± 20.8 73.9 -F 13.3 3.069 0.027 Swiftlet Feeding attempts 13.7 -h 10.9 12.4 -F 10.6 20.7 ± 20.8 26.1 4- 13.3 13.390 <0.001 Call 2.4 H- 5.9 2.3 -F 4.6 0 0 5.130 0.002 Follow 9.1 -F 11.4 12.4 -F 13.2 0 0 24.400 <0.001 n = 719 n = 333 n = 306 n = 67 n = 13 Glossy Swiftlet Fly/glide 90.4 4- 13.6 90.3 -F 14.9 74.00 ± 17.8 79.9 4- 18.6 57.552 <0.001 Feeding attempts 7.4 -h 12.2 6.7 -F 12.9 25.90 ± 17.7 19.8 4- 18.6 152.767 <0.001 Call 1.0 -F 4.6 1.2 -F 3.8 0.03 ± 0.6 0.05 4- 0.5 14.502 <0.001 Follow 1.2 -F 4.9 1.9 -F 5.4 0.10 ± 1.2 0.20 4- 1.5 18.562 <0.001 n = 1,265 n = 359 n = 41 1 n — 358 n = 137 Statistically, monthly variations in feeding at- tempts could not be shown for Edible-nest Swiftlets for Inside Stream Bank Canopy, Inside Forest Canopy, and 0-5 m Above Ground categories and for Glossy Swiftlets in Below Stream Bank Canopy. The Edible-nest Swiftlet had monthly variations in feeding attempts except in 5-30 m Above Ground. The Glossy Swiftlet had monthly variations in 0-10 m Above Forest Canopy, Inside Forest Canopy, 0-10 m Above Stream Bank Canopy, and Inside Stream Bank Canopy (Table 9). Statistically, diurnal variations in feeding attempts could not be shown for Edible-nest Swiftlets in Inside Forest Canopy, Inside Stream Bank Canopy, Below Stream Bank Canopy, and 0-5 m Above Ground, and for Glossy Swiftlets in Below Stream Bank Canopy. The Edible-nest Swiftlet had significant diurnal variations in feeding attempts in 0-10 m Above Stream Bank Canopy, >30 m Above Ground, and 5-30 m Above Ground microhabitats, whereas the Glossy Swiftlet had diurnal variations, except in >10 m Above Forest Canopy, 0-10 m Above Forest Canopy, and >30 m Above Ground microhabitats (Table 10). All microhabitats except Inside Forest Canopy and Inside Stream Bank Canopy were shared by the two species. Variations in microhabitat use could not be calculated for > 10 m Above Forest Canopy, Inside Forest Canopy, Inside Stream Bank Canopy, >30 m Above Ground, and 0-5 m Above Ground. Both species had variations in use of microhabitats 0-10 m Above Forest Canopy, 0-10 m Above Stream Bank Canopy, and 5-30 m Above Forest Ground (Table 11). Important microhabitats for Edible-nest Swiftlets in descending order were >10 m Above Forest Canopy, >30 m Above Ground, and Above Forest Canopy. Inside Forest Canopy and Inside Stream Bank Canopy were important microhabitats for Glossy Swiftlets. Foraging activities for both species were not correlated to mean temperature (Edible-nest Swiftlet: r = 0.096, P = 0.856; Glossy Swiftlet: r = 0.007, P = 0.989) and average rainfall (Edible-nest Swiftlet: r = —0.447, P = 0.375; Glossy Swiftlet: r = —0.216, P — 0.682). TABLE 5. Variation.s (.v ± SD, %) in feeding maneuvers by Edible-nest and Glossy swiftlets at Chalis-ek, North Andaman Island, India. Feeding maneuver Ediblc-nesl Swiftlel n = .S87 Glossy Swiftlet n = 67.1 Mann-Whitney test Z P Twi.st 44.12 ± 47.44 61.29 ± 31.26 -4.347 <0.001 Fly-pause 2.43 ± 8.82 0.17 ± 1.59 -9.149 <0.001 Roll 2.40 ± 9.59 0.15 ± 1.43 -8.053 <0.001 Flutter 25.57 ± 23.26 19.28 ± 23.85 -3.516 <0.001 Tail-wing-open 25.48 ± 22.92 19.11 ± 23.83 -3.619 <0.001 % Foraging attempt (Mean +- 2 SE) % Foraging attempt (Mean +- 2 SE) %Foraiging attempt (Mean +- 2 SE) Mamhi ciiul Saukaran • SWIFTLET FORAGING IN THE ANDAMAN ISLANDS 265 A I □ Edible-nest Swiftlet I □ Glossy Swiftlet B 1 □ Edible-nest Swiftlet I □ Glossy Swiftlet C I □ Edible-nest Swiftlet I □ Glossy Swiftlet FIG. 2. Spatio-temporal variation (A = monthly, B = diurnal, and C = location; %) in feeding attempts of Edible-nest and Glossy swiftlets at Chalis-ck, North Andaman Island, India. DISCUSSION Aerial and Foraging Behavior. — Edible-nest and Glossy swiftlets performed fly/glide, foraging attempts (twist, fly-pause, roll, flutter, and tail- wing-open), call, follow, preen, defecate, and carry nest material in decreasing order. When breeding seasonality of the Edible-nest Swiftlet based on Sankaran and Manchi (2008), and Manchi (2009) was considered, temporal varia- tions in activities of the two species near caves used for breeding appeared to be related to timing of breeding. Swiflets at Chalis-ek had chicks and eggs in the nest during April and May, and were more active near breeding sites. Edible-nest Swiftlets near caves used for breeding were inactive between 0700 and 1700 hrs, and 0800 and 1400 hrs, respectively, during early and late nest construction periods. Not requiring nest material might have led Edible-nest Swiftlets to spend most of the time exploring potential foraging areas away from breeding sites. Com- mencement of egg-laying during February fol- lowed by incubation, and brooding and feeding nestlings led to increased activity of Edible-nest Swiftlets near breeding sites. The active period of Edible-nest Swiftlets declined after May (Fig. 1), which was attributed to fledging occurring from most nests and avoidance of extreme weather conditions (Medway 1962). Medway (1962) reported wider foraging ranges of Edible-nest Swiftlets in the non-breeding season, leading to changes in activity near breeding sites. Swiftlets at Chalis-ek, traveled away from the caves even during the pre- incubation period, possibly due to non-availability of food near caves. Food availability for swiftlets depends on the density and diversity of insects which in turn depend on weather conditions (Johnson 1969). Glossy Swiftlets were more active in early and late morning hours, but did not have significant diurnal variations in activities (Fig. 2). Asynchronous breeding in the Andaman and Nicobar Islands by Glossy Swiftlets may have resulted in no variation in diurnal activities. Fly/glide was the most frequent behavior apparently used for scanning foraging areas while traveling to other foraging areas. Both species spent substantial time in fly/glide which depends on availability of in.sects. Insect density varies chronologically, spatially, and seasonally (John- ■son 1969), thus, affecting fly/glide in the two species. Feeding attempt was the second major TABLE 6. Temporal (monthly) variation (%) in foraging maneuvers of Edible-nest and Glossy swiftlets at Chalis-ek, North Andaman Islands, India. 266 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 2, June 2010 On ro o o o O o o in o o in CN 00 o Os 00 SO cn a o o o CN O o o o O cn 00 os OS CN SO in o r- O d d d d d d d d d d CL o CN o O O in o O < > V V V V V V V V < d d d d d d d d V d d o > z o < 00 o _ CN wn — r- in 00 2 so o — 00 so os 00 k. — ■ 00 OS p fO p c — sO o so cn o os CN CN vri CN d — CN d d d d CN cA ■o c u. so p p m p r-* 00 p 00 m os sO O CN in p O 00 CN CN d CN d d — d CN P m Tt Os p C3 CO ON d sd o sd sd d in (N d CN CN VO m cn CN Jun + 1 + 1 + 1 + 1 + 1 II + 1 o o + 1 + 1 II e 03 g 00 sd in d p p os d os cn p 0C3 cn p d p CN cn (N 00 00 p s: os r- p c: XJ 00 CN — 00 CN CN CN d CN CN sd sd CN m d so d m 00 c <: Q T + 1 + 1 + 1 + 1 + 1 II + 1 +1 O +1 + 1 11 s: 00 OS p p — — 5: cn r-- 00 ::: t: + 1 “ d d d in 00 CN o d in 00 00 p r-) d p d m r- 00 sd % Tt CN CN SO d CN S >> _ p 00 CN r-; Os p in sd in CO U £ u ci m + 1 in + 1 00 + 1 p + 1 + 1 p II c: + 1 os + 1 cn + 1 SO + 1 OS + 1 II u 8. u (N d (N CN m CN CN “ d d d d d sd d d d d Api + 1 + 1 + 1 + 1 +1 II + 1 o o + 1 + 1 II S(— ( ■Q 01) CN CN in CN CN c r-; 00 m p 00 s: 3 O u > c ja (N QO CN p p 5 3 lU c cn p in "t p r- P d d d d (N r- d p CN d m ~o CO E o d CN d d d CN d CN in d m 1 CN CN d (N so (N •M (N rj Os d •— < CN CN <3 CO OD c ~~ oh CO a + 1 + 1 +1 + 1 + 1 II + 1 + 1 +1 + 1 + 1 II T m + 1 + 1 + 1 + 1 +1 II + 1 + 1 + 1 + 1 + 1 11 (N C^l 00 o OJ 00 p p p p p •M 00 p p ” p p p C s o 00 CN d d d d o sd d a d d d d d d in d D CN CN in d CN V C'J (N so d CN c , e ON 00 p s: CN m p o c: > 3 1) 00 CN d CN CN CN CN CN CN CN SO sd d c-i d d d d sd 0) E T + 1 + 1 +1 + 1 + 1 li + 1 + 1 o + 1 + 1 II so — c O 03 in C c 2 > 3 c CL Im k. 3 05 D CL O (U CL O U o o 3 C i3 £ (U OJO 00 > c/5 c/5 c o CO -C UJ p E PJ CO O O Mimvhi and Sankamn • SWIFTLET FORAGING IN THE ANDAMAN ISLANDS 267 TABLE 8. Spatial variations in the proportionate activities of Edible-nest and Glossy swiftlets at Chalis-ek, North Andaman Island, India {P < 0.05 = significant difference). Time spent on different foraging maneuvers at different locations (,v ± SD) ANOVA Foraging Species maneuvers Hill ba.se Hill top Open land-1 Open land-11 F P Edible-nest Twist 43.0 -+■ 25.5 42.4 -E 27.6 56.6 -E 31.4 63.0 -E 35.0 5.977 <0.001 Swiftlet Fly-pau,se 2.4 8.8 2.4 -E 9.3 2.8 -E 6.9 2.4 -E 5.0 0.248 0.863 Roll 1. 5 7.8 3.3 “E 11.6 3.2 -E 7.6 0.6 -E 2.1 2.147 0.093 Flutter 26.5 20.7 25.2 H- 25.6 21.7 23.4 25.8 -E 26.4 1.141 0.332 Tail-wing-open 26.6 22.8 26.8 -E 23.3 15.7 + 19.6 8.5 -E 1 1.5 4.844 0.002 n = 587 n = 276 n = 256 n = 42 n = 13 Glossy Swiftlet Twist 65.7 -E 36.9 69.3 H- 33.6 56.5 -E 25.7 58.9 -E 32.4 6.407 <0.001 Fly-pause 0.3 -E 2.8 0 0.2 -E 1.4 0.07 -E 0.4 1.127 0.337 Roll O.l -E 1.2 0.2 -E 1.4 0.2 1.7 0 0.725 0.537 Flutter 21. 1 -E 31.9 14.4 -E 25.9 20.2 -E 18.6 20.3 -E 21.4 4.553 0.004 Tail-wing-open 12.7 -4- 25.3 16.1 -E 26.0 22.9 -E 20.6 20.7 -E 25.9 11.288 <0.001 n = 673 n z= 141 n = 134 n = 306 n 92 activity after fly/glide. All individual activities, except feeding attempts, had temporal variations in their proportions by month. No diurnal variation in feeding attempts and calls was recorded. Call and follow were presumed to be mating/pairing displays as well as a response to foraging competition. Call and follow activities by Edible-nest Swiftlets, and follow by Glossy Swiftlets were higher during nest construction (Jan-Feb). This suggests that follow must be more related to mating. Insect diversity was assumed to affect the occurrence of different foraging maneuvers with the presumption that different foraging maneuvers are used to prey on different type of insects. This dependence resulted in temporal (monthly and hourly) variations in the proportions of different foraging maneuvers. One factor considered was that Glossy Swiftlets have been shown to be more maneuverable than Edible-nest Swiftlets (Tarburton 1986), as it has both a higher wing and tail maneuverability index. This could help explain why Glossy Swiftlets were able to produce greater variability in feeding and flutter maneuvers. Roll alone by Glossy Swiftlets had diurnal variations; however, no diurnal variation was observed in foraging maneuvers by Edible-nest Swiftlets. Spatial var- iations were shown by twist and tail-wing-open in both species, and flutter alone by Glossy Swift- 0600 0800 1000 1200 1400 1600 I □ Edible-nest Swiftlet I □ Glossy Swiftlet Indian Standard Time n = 276 141 256 134 42 306 92 13 Edible-nest Swiftlet Glossy Swiftlet Hill base Hilltop Openland-l Openland-ll FIG. 3. Temporal variations in diurnal hock sizes of Edible-nest and Glossy swiftlets at Chalis-ek, North Andaman Island, India. FIG. 4. Variation in abundance of Edible-nest and Glossy swiftlets at four locations at Chalis-ek, North Andaman Island. India. 268 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 2, June 2010 C ■a c c ra p ~o c < o Z u ca U X) :a X u. o £ S L. '■B c O ■o c ra :/5 o £ c o. > O- O c a U c ca CQ o o fT UL« °l ^ < S E o 2 A o X < £ o 5 g: a § I U O == s ca ^ QQ 03 S I iy5 C/5 > OJ o o > ^ X) o < < 3 F E r- °° ^ F lo o If; rn A ' ' ^ I g P 3 .“F U O 4J U O [? tu U- (U ^ < < E E o 2 >> a. c c ca U Jii C ca ca CL o c ca U ? 2 m O rn 10 ^ — >A c X E E E o r^i u~j 00 o O TABLE 10. Temporal (diurnal) variations in activity of Edible-nest and Glossy swiftlets in different microhabitats at Chalis-ek, North Andaman Islands, India. Manchi and Saukuran • SWIFTLET FORACIING IN THE ANDAMAN ISLANDS 269 in 00 SO m in m in r- in rt in o r- so r- 00 •— (N 00 (N in p 00 (N o 00 *— o (N in CN 00 d> o d d m (N 00 sO 00 ON 00 OO o m o o m o o o n »— • «— m o o o O sO o o o 00 o O d d d d d d d d d d d d d V in iri + 1 +1 00 (N — ^ o6 m 00 O o o o o o o o o +1 +i OO 00 (N m + 1 On sd rJ — +10 0+10 OO 00 00 p m p p in (N p 00 m d p d p in ? m m os CN cn m d d i + 1 + 1 o o o o o o o II + l +1 + 1 + 1 + 1 o + 1 + 1 + 1 11 m CN 00 s: p p p so — > CL o c c3 u c 0^ o X o X (U ,§ § < s E ^ 1) 03 o u O li- eu < £ u. Jo $ < e < £ X < E 583 < E < £ t— 1 o IL 1) Jo < E < E X < £ 656 *0 o !E o o o o 72. o •;o p o o c 5 c 13 PP ro A m 1 in in + II s: A 7 o c + £ 13 CQ m A ro 1 in in II s: CO lU c UJ CO _o a 270 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 2. June 2010 TABLE 11. Association between Edible- North Andaman Island, India. nest (ENS) and Glossy swiftlets (GS) with microhabitat use at Chalis-ek, Microhabitat^ Species n Activity (.v ± SD) x2 p HaFC ENS 412 1 1.47 ± 10.09 _ GS 4 6.75 ± 2.06 AFC ENS 413 11.43 ± 9.38 109.901 <0.001 GS 277 8.32 ± 13.02 IFC ENS 0 0 GS 174 6.95 ± 7.53 ASC ENS 10 18.50 ± 13.19 228.748 <0.001 GS 209 24.82 ± 20.74 ISC ENS 0 0 GS 169 24.07 ± 19.54 BSC ENS 1 39 0.009 0.923 GS 1 12 HaG ENS 20 15.05 ± 12.53 GS 3 40.00 ± 26.00 LaG ENS 7 5.43 ± 5.29 48.672 <0.001 GS 65 18.95 ± 22.84 NG ENS 3 7.00 ± 5.29 GS 1 1 17.64 ± 8.58 HaFC: >10 m Above Forest Canopy; AFC: 0-10 m Above Forest Canopy: [FC: Inside Forest Canopy; ASC: 0-10 m Above Stream Bank Canopy: ISC: Inside Stream Bank Canopy; BSC: Below Stream Bank Canopy; HaG: >30 m Above Ground; LaG: 5-30 m Above Ground; NG: 0-5 m Above Ground; — = cannot be calculated. lets. Monthly variations in foraging maneuvers indicate significant seasonal variations in insect diversity and change in prey preference with breeding stages (Johnson 1969). Diurnal variation in insect diversity in the forest is uncertain, although differing slightly in open lands. This indicates that most foraging attempts in the microhabitats >10 m Above Forest Canopy and >30 m Above Ground may be because of more dispersion of insects at that height (Johnson 1969). Habitats and Microhahitats. — Edible-nest Swiftlets were more active in forested areas, whereas Glossy Swiftlets were more active in open paddy lands. The Edible-nest Swiftlet’s activity is in concordance with previous studies (Medway 1962, Harrison 1972, Hails and Amir- rudin 1981, Waugh and Hails 1983, Charles 1987, Lim and Cranbrook 2002, Nguyen et al. 2002). The Glossy Swiftlet is the only swiftlet species known to occur in all habitats. The diet of the Glossy Swiftlet in Malaysia (Laurie and Tomp- kins 2000) shows elasticity in food and foraging habitat selection unlike the Edible-nest Swiftlet. Eurther validity is given the concept of greater elasticity by Collins (2000) for Glossy Swiftlets on Palawan where it has resource partitioning caused by the presence of Pygmy Swiftlet (Collocalia trof’lodytes). In this situation, it was restricted in most of its feeding to <2 m above ground; a zone more likely to increase vulnera- bility to land-based predators. That this species flies so low when “forced” into ecological partitioning was reported from New Guinea by Schoddle and Hitchcock (1968), where it was sharing habitat with the Uniform Swiftlet {Aero- dramus vanikorensis). It has been shown on that same island, but at greater altitudes, to share space with the Mountain Swiftlet {A. hirundinaceus) in a similar way (Diamond 1972). Edible-nest Swiftlets were more active in early morning and late afternoon and were in larger foraging flocks near caves used for breeding. Frequency of feeding attempts by both species was correlated with foraging flock size. Edible-nest Swiftlets forage high above the canopy, while Glossy Swiftlets forage at the canopy level and near the ground. Johnson (1969) suggested that insects move upwards in air with an increase in ambient temperature. This may be the reason for the shift of Edible-nest Swiftlets to higher altitudes during foraging and Glossy Swiftlets, being smaller and more agile to maneuver in confined spaces, to move towards relatively cooler zones inside the canopy at the forest edges or stream banks close to open paddy lands for foraging. Variations in microhabitat use signify resource partitioning among Edible-nest and Glossy swift- Manchi and Sankanm • SWIFTLET FORAGING IN THE ANDAMAN ISLANDS 271 lets at Chalis-ek with Edible-nest Swil'tlets having affinity for categories, >10 m Above Forest Canopy, >30 m Above Ground, and Above Forest Canopy, and Glossy Swiftlets for Inside Forest Canopy and Inside Stream Bank Canopy in descending order. Variations in foraging maneu- vers indicate variations in food preferences by both species as demonstrated by gut content and bolus analysis (Medway 1962, Harrison 1972, Hails and Amiruddin 1981, Waugh and Hails 1983, Charles 1987, Collins 2000, Laurie and Tompkins 2000). The Edible-nest Swiftlet’s affinity towards >30 m Above Ground in open land for foraging was unusual as the species has been known to prefer forested areas. The plasticity in food selection by Edible-nest Swiftlets (Charles 1987, Laurie and Tompkins 2000) suggests two microhabitats, >10 m Above Forest Canopy and >30 m above ground, may have similar insect diversity. This demonstrates that forested habitat and microhabitats above forest canopy level are important for Edible-nest Swiftlets. Glossy Swift- lets also had affinity towards both available habitats. We infer, in concurrence with previous studies, that cave-dwelling Edible-nest Swiftlets are dependent on forests and will be adversely affected by the present rate of deforestation and habitat alteration. Changes in land use away from caves used for breeding sites will also lead to population declines of Edible-nest Swiftlets be- cause of its wider foraging ranges. However, habitat alteration around breeding caves may not have a marked effect on foraging activities of Glossy Swiftlets, which have more elasticity in habitat use. ACKNOWLEDGMENTS We acknowledge the Department of Environment and Forests, Andaman and Nicobar Islands for providing funds and active collaboration with Salim Ali Centre for Ornithology and Natural History (SACON), Coimbatore and for initiating this conservation program and study. We thank the Swiftlet Protection Team for assistance during the work. We also thank Dr. Krishnamoorthy Thiyagesan, AVC College, Mayiladuthurai, for valuable suggestions during statistical analysis, Archana Waran. Shobha Phaniraj, Rachna Chandra. Madhuri Ramesh, and Drs, P. A. Azeez, Padmanabhan Pramod, and Balakrishnan Peroth for com- ments on previous versions of the manuscript. LITERATURE CITED Altmann, J. 1974. Observational study of behaviour; sampling methods. Behaviour 49:227-265. Chantler, P. 1999. Family Apodidae (swifts and swift- lets). Pages .^87-418 in Handbook of the birds of the world. Volume 5. Barn Owls to hummingbirds (J. del Hoyo, A. Elliott, and J. Sargatal, Editors). Lynx Edicions, Barcelona, Spain. Chantler, P. and G. Drie,s.sens. 2()()0, Swifts - a guide to the swifts and treeswifts of the World. Pica Press, Sussex, United Kingdom. Charles, M. F. 1987. The management of bird’s nest caves in Sabah, Wildlife Section, Sabah Forest Department, Sabah, Malaysia. Collins, C. 2000. Foraging of Glossy and Pygmy swiftlets in Palawan, Philippines. Forktail 16:53-55. Diamond, J. M. 1972. Avifauna of the Eastern Highlands of New Guinea. Publications of the Nuttall Ornitholog- ical Club Number 12. Hails, C. J. and A. Amirrudin. 1981. Food samples and selectivity of White-bellied Swiftlets (Collocalia esculenta). Ibis 123:328-333. Harrison, T. 1972. The food of Collocalia Swiftlet (Aves, Apodidae) at Niah Great Cave in Borneo. Journal of Bombay Natural History Society 71:376-392. Johnson, C. G. 1969. Migration and dispersal of insects by flight. Methuen, London, United Kingdom. Langham, N. 1980. Breeding biology of the Edible-nest Swiftlet (Aerodramiis fuciphagus). Ibis 122:447^61. Lau, A. S. M. AND D. S. Melville. 1994. International trade in swiftlet nests. Traffic International and World Wide Fund for Nature, Cambridge, United Kingdom. Laurie, S. A. and D. M. Tompkins. 2000. The diets of Malaysian swiftlets. Ibis 142:596-602. Lim, C. K. and Earl of Cranbrook. 2002. Swiftlets of Borneo: builders of edible nests. Natural History Publication, Kota Kinabalu, Borneo. Manchi, S. 2009. Breeding ecology of the Edible-nest Swiftlet and the Glossy Swiftlet in the Andaman Islands. Thesis. Bharathiar University, Coimbatore. India. Manchi, S. and R. Sankaran. 2009. Impact of the 2004 earthquake on the limestone caves in North and Middle Andaman Islands, India. Cuixent Science 97: 1230-1234 Medway, L. 1962. The swiftlets (Collocalia) of Niah Cave. Sarwak. Part 2. Ecology and regulation of breeding biology. Ibis 104:228-245. Medway, L. 1963. The antiquity of trade in edible birds'- nest. Federal Museum Journal 8:36^7. Mitchell, K. 2001. Quantilative analysis by the Point- centered Quarter Method. Department of Mathematics and Computer Science, Hobart and William Smith Colleges, Geneva, New York. USA. Nguyen, Q. P. 1983. The food of Edible-nest Swiftlets Collocalia fneiphaga germani Oust, in Vietnam. Top Chi Sinh Vat Hoc 5 (3):5-8. Nguyen, Q. P., Y. V. Quang, and J. F. Volsin. 2002. The White-nest Swiftlet and the Black-nest Swiftlet; a Monograph. Societe Nouvelle Des Editions Boubee. Paris. France. Norberg, U. M. 1986. Evolutionary convergence in foraging niche and flight morphology in insectivorous 272 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 2. June 2010 aerial-hawking birds and bats. Omis Scandinavica 17:253-260. Sankaran, R. 1995. Impact assessment of nest collection on the Edible-nest Swiftlet (Collocalia fuciphaga) in the Nicobar Islands. Occasional Report. Salim Ali Centre for Ornithology and Natural History, Coimba- tore, India. Sankaran, R. 1998. The impact of nest collection on the Edible-nest Swiftlet (Collocalia fuciphaga) in the Andaman and Nicobar Islands. Report to lUCN. Salim Ali Centre for Ornithology and Natural History, Coimbatore, India. Sankaran, R. 2001. The status and conservation of the Edible-nest Swiftlet (Collocalia fuciphaga) in the Andaman and Nicobar Islands. Biological Conserva- tion 97:283-294. Sankaran, R. and S. Manchi. 2008. Conservation of the Edible-nest Swiftlets in the Andaman and Nicobar Islands. Occasional Report. Salim Ali Centre for Ornithology and Natural History, Coimbatore, India. SCHODDE, R. AND W. B. Hitchcock. 1968. Contribution to Papuasian ornithology 1. Report on the birds of the Lake Kutubu Area, Territory of Papua and New Guinea. Paper 13. Division of Wildlife Research and Technology, CSIRO, Australia. Spennemann, a. W. 1928. Die ebbaren Vogelnester von Grissee. Orn. Monatsb. Berlin 1 11-1 12. SPSS Institute Inc. 1998. SPSS for Windows. Version 10.0. SPSS Institute Inc., Chicago, Illinois, USA. Tarburton, M. K. 1986. A comparison of the flight behaviour of the White-rumped Swiftlet and the Welcome Swallow. Bird Behaviour 6: 72-84. Waugh. D. R. and C. J. Hails. 1983. Foraging ecology of a tropical aerial feeding guild. Ibis 125:200-217. The VMIsoii Journal of Ornithology 122(2):273-278, 2010 GOLDEN- AND BLUE-WINGED WARBLERS: DISTRIBUTION, NESTING SUCCESS, AND GENETIC DIEFERENCES IN TWO HABITATS JOHN L. CONFER,' - KEVIN W. BARNES,' AND ERIN C. ALVEY' ABSTRACT. — We analyzed phenotypic distribution, nesting success, and genetic purity of Golden-winged Warblers (Vermivoni clirysoptera) and Blue-winged Warblers {V. pinus) in two ecologically distinct nesting habitats: early- succession uplands, and swamp forests. The proportion of phenotypically pure Golden-winged Warblers in swamp forests (94%) differed significantly from uplands (53%) as did the proportion of Golden-winged Warbler pairs (93%) in swamp forests compared to uplands (48%). Only 1% of the phenotypically pure Golden-winged Warblers in swamp forests paired with either a hybrid or a Blue-winged Warbler, but 7% of the phenotypically pure Golden-winged Warblers in uplands formed a hybrid pair. The probability of nesting success for Golden-winged Warbler pairs in swamp forests (65%) was significantly higher than in uplands (37%). Mitochondrial DNA analyses indicate that all 10 phenotypically pure Golden- winged Warblers sampled from swamp forests had the ancestral Golden-winged Warbler haplotype, while 10 of 25 Golden- winged Warblers from uplands had the ancestral Blue-winged Warbler haplotype (P = 0.033). Swamp forests may provide a source habitat for Golden-winged Warblers with a high phenotypic and genotypic purity even in sympatry with Blue- winged Warblers. Received 27 August 2009. Accepted 23 December 2009. The Golden-winged Warbler (Vermivora chry- soptera) has declined 8.8% per year in U.S. Fish and Wildlife Region 5 from 1966 to 2007 (Sauer et al. 2008). This decline occurred in all areas that surround the site for this study, Sterling Forest State Park, New York, USA. The Golden-winged Warbler declined by 53 and 72% in New York (Confer 2008) and in Pennsylvania (Pennsylvania Breeding Bird Atlas 2009), respectively, during the last 20 years. This species has declined from —200 pairs to 15 pairs in New Jersey from 1998 to 2008, according to statewide surveys by the New Jersey Division of Fish and Wildlife (S. M. Petzinger, unpubl. data). The Golden-winged Warbler has been extirpated in Massachusetts and greatly declined in Connecticut following intrusion by the Blue-winged Warbler (V. pinus) despite 14,000 ha of suitable shrubland habitat on utility rights-of-ways (Confer and Pascoe 2003). Popula- tion declines of the Golden-winged Warbler usually follow arrival of the Blue-winged Warbler with subsequent competition and hybridization (Confer et al. 2003). However, Golden-winged Warblers have co-existed with Blue-winged War- blers for over a century (Eaton 1914) in southern New York and populations appear to be stable at — 125 breeding pairs in Sterling Forest State Park (Confer et al. 1998, this study). Initial surveys in 1999 and 2000 were conduct- ed in uplands, where both Golden-winged and ' Department of Biology, Ithaca College, 953 Danby Road, Ithaca. NY 14850, USA. ^Corresponding author; e-mail: Confer@ithaca.edu Blue-winged warblers occur frequently. It became apparent that both species also nested in two wetland communities described by the New York State Natural Heritage Program (Reschke 1990). The shrub swamp community supported about equal numbers of both species. The red maple (Acer ruhrum) hardwood forest community, or swamp forest, attracted primarily Golden-winged Warblers and our wetland surveys focused on this habitat. This study tests the hypothesis that favorable demographics for Golden-winged Warblers in swamp forest habitat account for its long-term stability and its co-existence with Blue-winged Warblers in Sterling Forest State Park, New York. Specifically, we hypothesize that Golden- winged Warblers will have higher nesting success, lower hybridization rates, and higher genetic purity in swamp forest habitats than in uplands. METHODS The principal author and a field crew of three to .seven people studied Golden-winged and Blue- winged warblers for 5 to 6 days a week for 10 weeks in Sterling Forest State Park. New York in 2001, 2()()3-2()06, and 2008. Territories of Golden-winged and Blue-winged warblers often were contiguous or overlapped and, at times, data could be acquired by sound from more distant teiTitories. Each territory was observed for 25-100 person-hrs based on 15-50 person-visits, includ- ing simultaneous observation of more than one 273 274 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2. June 2010 territory. Male Golden-winged Warblers were banded as early in the season as possible to assist delineation of temtorial boundaries and pairing status. At least 80% of the males were individ- ually recognizable within 3 weeks of the start of the season, assisted by a —50% return rate of banded males and distinctive phenotypes of occasional hybrids. The intervals between obser- vations were too long to be used for calculations of nesting success during 2002 and 2007. We included observations from 2002 to help estimate the number of birds and their phenotype in different habitats. We made observations each year in 3^ upland sites separated by 8-14 km, and 5-12 swamp forests separated by 4—12 km, depending on size of the field crew. We tried to obtain complete demographic values and blood samples for every pair in every patch, which provide an unbiased estimate of demographic values within each patch. Sites were selected because they had typical habitat and provided a large number of Vermivora territories in a contiguous area, facil- itating data collection. Upland sites consisted of a segment of rights- of-way (ROW) with a managed width of —20 m, another ROW segment with —80 m of managed width, a managed successional field, and an unmanaged area of secondary succession. These sites provide the total range of upland habitat used by Vermivora in the park. The two ROW sites were the only segments of ROW that provided continuous habitat with eight or more adjacent or over-lapping Vermivora territories. We estimate the upland sites we sampled probably supported more than half of all upland Vermivora territories in the park based on Geographic Information System (GlS)-based habitat classification of satellite imagery, surveys along all ROW in the park, and our visits to many small patches of disturbance habitat. Both ROWs have been managed in a stable vegetative condition for the last 40 years by a combination of herbicide application and cutting. Both ROW have about 60% herb cover. The two upland successional sites used for this study had —9-12 ha of successional habitat. One of the upland sties has been used for experimental habitat management for Golden-winged Warblers and this management created early secondary succession habitat. The second successional site was an abandoned gravel mine with .secondary succession field/shrub conditions. Most of the habitat used by Vermivora at this site was flooded by a large American beaver {Castor canadensis) pond built in summer 2006. Monitoring at this site has ended. All swamp forest patches were between —0.5 and 5 ha and supported 1-3 pairs of Vermivora. We monitored 12 patches of swamp forest, summed for all years, which provided a represen- tative sample of Vermivora demographics for swamp forests in general. Nests were observed at 2-5 day intervals. We visited nests more frequently when nestlings were expected to be 5-7 days of age and appropriate for blood sampling and near the time of fledging. Estimation of nesting success followed Mayfield (1961). We calculated nesting success starting with the first day of incubation, assuming incubation lasted 1 1 days and the nestling stage lasted 10 days. Many nests discovered during construction or egg-laying were abandoned. We attributed most of the high failure rate early in the nesting cycle to human-induced abandonment and omitted these nests from calculations of nest success. We found six second nest attempts of pairs whose first nest was predated after incuba- tion started. These re-nests are included in the sample size for nests. We used the criteria of Gill (1980) to dis- tinguish between phenotypically pure and hybrid birds. We used Chi-square, 2X2, contingency tests with Yeats correction for continuity with larger samples to test for differences in the distribution of pure and hybrid phenotypes in upland and swamp forests, and Fisher’s exact probability value for tests where some cells had expected values <5. We used a /-test to compare the probability of nesting success between upland and swamp forest habitats. Mitochondrial DNA analyses of Golden- winged and Blue-winged warblers shov/ little variation within two distinct haplotypes, presum- ably the ancestral mtDNA of each species. The two haplotypes differ from each other by 3-5% (Gill 1997, Shapiro et al. 2004, Dabrowski et al. 2005). Samples from Sterling Forest State Park collected in 2002 and 2003 showed extensive introgression with 12 of 28 phenotypically pure Golden-winged Warblers having the ancestral Blue-winged Warbler haplolype (Dabrowski et al. 2005). P. J. Hartman (pers. comm.), continuing mtDNA analyses of Sterling Forest State Park samples, found the same haplotype classification for four replicate samples analyzed by both Confur ei a/. • WARBLER HABITAT DIFFERENTIATION 275 TABLE 1. Abundance of resident warbler pairs by phenotypes in swamp forests and uplands in Sterling State Park Forest, New York, 2001-2{X)6 and 2008. Male and female phenotypes of pairs, males listed first Habitat GW" X GW BR X GW GW X BR GW X BW BW X BW BW X GW BR X BW Swamp forests 42 1 0 0 2 0 0 Uplands 46 2 3 1 40 1 2 ^ Abbreviations: GW = Golden-winged Warbler, BR = Brewster’s Warbler (hybrid), and BW ^ Blue-winged-Warbler. laboratories. Pooling results from these two laboratories, we compare the phenotype-haplo- type concordance of 35 phenotypically pure Golden-winged Warbler individuals in upland and swamp forest habitats in Sterling Forest State Park. RESULTS All upland sites included herbaceous patches of goldenrod {Solidago spp.) and grasses, and shrub patches of dogwood {Corniis spp.) and briers (Riihus spp.). Poison ivy {Toxocodendron radicans) and Asiatic bittersweet (Celastrus orhicidatus) were widely distributed and, at times, formed dense stands. Forests formed at least half of the boundary for all upland territories. Results from upland sites were pooled because vegetation varied as much within one site as among the sites, and because inspection of our data did not suggest any difference in population parameters among these sites. Red maple dominated the canopy in swamp forests of Sterling Forest State Park, and provided —60-80% canopy closure. Dead tree trunks suggest that elms {Ulmus spp.) were once domi- nant. Spice bush (Lindera benzoin) was the most widely distributed shrub followed in abundance by arrowwood (Viburnum recognitum) and wild raisin (V. cassinoides). Poison sumac (Toxocodendron vernix) and swamp azalea (Rhododendron visco- sum) occurred occasionally. Tussock sedge (Carex stricta) strongly dominated the herb layer and was the principal substrate for nests in wetlands. Dense patches of Phragmites occurred in several loca- tions, although Golden-winged Warblers seldom perched on this plant and only once used it as a nest substrate. Data from all swamp forests were pooled for the same rea.sons applied to the upland sites. Phenotypically pure Golden-winged Warblers during 2001-2006, and 2008 comprised 94% (85 of 90) of the resident individuals in swamp forests and 53% (99 of 186) in uplands (Table 1) (X" = 43.97; df = \, P < 0.001). The proportion of Golden-winged Warbler individuals in these two habitats was reflected in pairings among the pure and hybrid phenotypes. Phenotypically pure Golden-winged Warblers formed 93% (42 of 45) of all pairs in swamp forests but only 48% (46 of 95) in uplands (2G = 23.81; df = 1, F < 0.001). The rate of introgression due to hybrid pairings was marginally lower in swamp forests. Only 1% (1 of 85) of the phenotypically pure Golden- winged Warbler individuals formed a social pair with either a hybrid or a Blue-winged Warbler in swamp forests, but 7% (7 of 99) of the phenotypically pure Golden-winged Warblers formed a hybrid pair in uplands (Fisher’s exact one-tailed test; df = 1, F = 0.051). The probability of nesting success in uplands varied considerably among years and between Golden-winged and Blue-winged warblers (Ta- ble 2). This large variation may be partially due to differences in predator abundance, although the variation is certainly partially due to the small sample of nests per habitat per year. Eastern chipmunks (Tamias striatus) were at least com- mon on ROW, and became abundant following oak (Quercus spp.) mast years when as many as five to 10 chipmunks were often observed concun'ently in 2002 and 2004. Black rat snakes (Elaphe obsoleta) were consistently common on ROW, probably due to the warming effect of direct sunshine and the presence of shelter in exposed rock talus. In contrast, no chipmunks or black rat snakes were observed in swamp forests during this study. The abundance of these predators on most ROW in Sterling Forest State Park may account for the variable and compara- tively low nesting success of Golden-winged Warblers in uplands. The estimated probability of nesting success for 2001, 2003-2006, and 2008 (Table 2) for Golden- winged Warblers in swamp forests (65% of 23 nests) was greater than in uplands (37% of 32 nests) (one-tailed r-test; df = 1, F < 0.0485). Nesting success for Blue-winged Warblers in uplands was moderately high at 54% based on 28 276 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 2, June 2010 TABLE 2. Probability of nesting success for pairs of Golden-wi State Park Lorest, New York. Six re-nests are includecJ. nged Warblers and Blue-winged Warblers in Sterling Year Golden-winged X Golden-winged Warblers Blue-winged X Blue- winged Warblers Swamp forest Upland Swamp forest Upland 2001 1“ (373'^) 0.49 (4/4) 0 1 (2/2) 2003 0.37 (3/3) 0.7 (7/7) 1 (1/1) 0.34 (3/3) 2004 0.59 (3/3) 0.26 (5/5) 0 0.29 (5/5) 2005 0.57 (6/6) 0.29 (4/4) 0 0.77 (5/4) 2006 0.82 (6/6) 0.23 (9/7) 0.21 (2/1) 0.61 (10/9) 2008 0.44 (2/1) 0.32 (3/3) 0 0.42 (3/3) Totals 0.65 (23/22) 0.37 (32/30) 0.42 (3/2) 0.54 (28/26) Probability of nesting success (Mayfield 1961). Number of nests. '■ Number of pairs. nests, but not significantly better than for Golden- winged Warblers. The low value for nesting success of Blue-winged Warblers in swamps (Table 2) is based on only three nests and is not reliable. We collected >300 blood samples from adults and sets of nestlings to ascertain if swamp forests attracted a population of birds with greater genetic purity than uplands. Preliminary analyses of mitochondrial DNA indicate swamp forests sup- port a higher proportion of Golden-winged Warbler phenotype-haplotype matches than up- land habitat. All 10 phenotypically pure Golden- winged Warblers from swamp forests that have been analyzed had the ancestral Golden-winged Warbler haplotype, while 10 of 25 of the phenotypically pure Golden-winged Warblers from uplands had the ancestral Blue-winged Warbler haplotype (Table 3). These preliminary results strongly suggest swamp forests support a higher proportion of Golden-winged Warbler phenotype-haplotype matches than upland habitat. DISCUSSION Golden-winged Warblers expanded into the Northeast during the last 150 years, primarily into secondary succession shrublands on aban- doned farmland and most of the populations now nest in the North Central states, often on clear cut timberlands (Gill 1980, Huffman 1997). Studies throughout the Golden-winged Warbler range de.scribe use of anthropogenic, disturbance eco- systems on dry sites: e.g., Georgia (Klaus and Buehler 2001 ), northcentral New York (Confer et al. 2003), Pennsylvania (Kubel 2005; J. L. Larkin, unpubl. data), Virginia (Wikson et al. 2007), West Virginia (Canterbury and Stover 1999), and Wisconsin (Roth and Lutz 2004). However, other studies have found nesting by Golden-winged Warblers in wetlands; Michigan (Berger 1958), southern New York (this study). North Carolina (Rossell et al. 2003), and northern New York according to surveys by the New York State Department of Environmental Conservation (T. Q. Rush and T. J. Post, unpubl. data). Some speculate the historical abundance of beaver meadows and other wetlands provided the evolutionary crucible (Hunter et al. 2001 ). Both wet and dry Golden-winged Warbler territories are dominated by patches of herbs and shrubs and, usually, a forested boundary. Upland sites with these attributes have been described for central New York (Confer and Knapp 1981), northcentral New York (Confer et al. 2003), southern New York (this study), and clearcuts in Georgia (Klaus and Buehler 2001). GIS mapping of vegetation has shown Golden-winged Warbler territories include a diverse mosaic of succession- al habitat types, especially herbaceous openings, and are often surrounded by forest in a wetland in North Carolina (Rossell et al. 2003), and in wet and dry territories in Sterling Forest State Park (J. L. Confer and J. L. Larkin, unpubl. data). Wetlands with many patches of vegetation in both southern New York (this study) and northern New York, according to surveys by the New York State Department of Environmental Con.servation (T. Q. Rush and T. J. Post, unpubl. data), attracted a nearly pure Golden-winged Warbler population despite the presence of Blue-winged Warblers in nearby uplands. Our multi-year ob.servations in Sterling Forest State Park overcome extreme annual variation and show major differences in population demograph- Confer ei cil. • WARBLER HABITAT DIFFERENTIATION 277 TABLE 3. Distribution of ancestral Golden-winged or Blue-winged warbler mitochondrial DNA for birds with pure Golden-winged Warbler phenotype nesting in two habitats. Mitochondrial haplotype Habitat Golden-winged Warbler Blue-winged Warbler Swamp Forests 10 0“ Uplands 15 10 “ Fisher's exact probability test; P for two-tailed test = 0.033. ics between wet and dry habitats. The proportion of Golden-winged Warblers to Blue-winged Warblers and hybrid phenotypes is much higher in swamp forests than in uplands. The absence of ancestral Blue-winged Warbler mtDNA in the sample of phenotypically pure Golden-winged Warblers in swamp forests indicates a history of infrequent hybridization for Golden-winged War- blers in swamp forests. We observed a low frequency of hybrid pairings in swamp forests, which suggests the current rate of introgression is also lower in swamp forests than in uplands. Further, nesting success in swamp forests was —75% better than nesting success in uplands. Our results strongly indicate the unusual persistence of Golden-winged Warblers in the presence of Blue- winged Warblers in uplands of Sterling Forest State Park is the result of a source/sink dynamic between swamp forests and uplands. Continued stability of genetically pure Golden- winged Warblers in this region may require that swamp forests sustain a sufficiently large popu- lation to buffer stochastic events. We observed about 55 Golden-winged Warbler pairs during casual surveys in 2009 with 22 pairs in swamp forests. We extrapolate, using GIS-based habitat classification, to an estimate of —125 Golden- winged Warbler pairs in Sterling Forest State Park with —50 in swamp forests. The Sterling Forest State Park population alone may be too small to perpetuate the Golden-winged Warbler. Fortu- nately, Sterling Forest State Park is within the Hudson Highlands, a tri-state parkland of 28,000 ha. Red maple hardwood forests occur throughout this larger area. Management for swamp forests throughout the larger Hudson Highlands could help sustain the Golden-winged Warbler population in this region. The presence of Golden-winged Warblers in wetlands with a mosaic of vegetation in North Carolina and southern and northern New York with habitat segregation from Blue-winged Warblers in the New York studies may provide important exam- ples for management in a wide area of sympatry. Further analyses of genetic purity and population demographics comparing uplands and wetlands would be valuable. ACKNOWLEDGMENTS This work was conducted in collaboration with and financially supported by the New York State Office of Parks, Recreation and Historic Preservation, Non-game Unit of the New York State Department of Environmental Conservation, New York State Biodiversity Research Institute, New York State Wildlife Grant program, and The Sterling Forest Partnership. T. B. Lyons, R. W. Perry, J. R. Gell, P. G. Novak, and J. A. Hutchinson, all of the New York State Office of Parks, Recreation and Historic Preservation, and L. M. Stenzler, D. F. Westneat, M. A. and J. C. Yrizarri, and P. J. Hamill provided much appreciated assistance. P. J. Hartman provided a valuable review of the manuscript. The work of more than 50 field assistants, principally from Ithaca College, made this study possible. LITERATURE CITED Berger, A. J. 1958. The Golden-winged-Blue-winged warbler complex in Michigan and the Great Lakes area. Jack-pine Warbler 36:37-72. Canterbury, R. A. and S. M. Stover. 1999. The Golden- winged Warbler: an imperiled migrant songbird of the southern West Virginia coalfields. Green Lands 29: 44-51. Confer, J. L. 2008. Golden-winged Warbler (Vermivora chrysoptera). Pages 468^69 in The second atlas of the breeding birds of New York State (K. J. McGowan and K. Corbin, Editors). Cornell University Press, Ithaca, New York, USA. Confer, J. L. and K. Knapp. 1981. Golden-winged Warblers and Blue-winged Warblers: the relative success of a habitat specialist and a generalist. Auk 98:108-1 14. Confer, J. L. and S. M. Pascoe. 2003. The avian community on utility rights-of-ways and other man- aged shrublands in northeastern United States. Forest Ecology and Management 185:193-206 Confer, J. L., J. Gebhards, and J. Yrizarry. 1998. Golden-winged and Blue-winged warblers at Sterling Forest: a unique circumstance. Kingbird 39:50-55. Confer, J. L., P. E. Allen, and .1. L. Larkin. 2003. Effects of vegetation, interspecific competition, and brood parasitism on Golden-winged Warbler nesting success. Auk 121: 138-144. Dabrowski, a., R. Fraser, J. L. Confer, and J. L. Lovette. 2005. Geographic variability in mitochon- drial introgression among hybridizing populations of Golden-winged (Vermivora chrysoptera) and Blue- winged (V^. pinus) warblers. Conservation Genetics 6:843-853. 278 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2, June 2010 Eaton, E. H. 1914. Birds of New York. Part 2. University of the State of New York, Albany, USA. Gill, F. B. 1980. Historical aspects of hybridization between Blue-winged and Golden-winged warblers. Auk 97:1-18. Gill, F. B. 1997. Local cytonuclear extinction of the Golden-winged Warbler. Evolution 51:519-525. Huffman, R. D. 1997. Effects of residual overstory on bird use and aspen regeneration in aspen harvest sites in Tamarac National Wildlife Refuge, Minnesota. Dis- sertation. West Virginia University, Morgantown, USA. Hunter, W. C., D. A. Buehler, R. A. Canterbury, J. L. Confer, and P. B. Hamel. 2001. Conservation of disturbance-dependent birds in eastern North America. Wildlife Society Bulletin 29:440-455. Klaus, N. A. and D. A. Buehler. 2001. Golden-winged Warbler breeding habitat characteristics and nest success in clearcuts in the southern Appalachian Mountains. Wilson Bulletin 1 13:297-301. Kubel, J. 2005. Breeding ecology of Golden-winged Warblers in managed habitats of central Pennsylvania. Thesis. Pennsylvania State University, State College, USA. Mayfield, H. 1961. Nesting success calculated from exposure. Wilson Bulletin 73:255-261. Pennsylvania Breeding Bird Atlas. 2009. http://www. carnegiemnh.org/atlas/index.htm (accessed 14 July 2009). Reschke, C. 1990. Ecological communities of New York State. New York Natural Heritage Program. New York State Department of Environmental Conservation, Latham, New York, USA. http://www.dec.ny.gov/ animals/29389. html (accessed 17 July 2009). Rossell Jr., C. R., S. C. Patch, and S. P. Wilds. 2003. Attributes ot Golden-winged Warbler territories in a mountain wetland. Wildlife Society Bulletin 31:1099- 1104. Roth, A. M. and S. Lutz. 2004. Relationship between territorial male Golden-Winged Warblers in managed aspen stands in northern Wisconsin. Forest Science 50:153-161. Sauer, J. R., J. E. Hines, and J. Fallon. 2008. The North American Breeding Bird Survey, results and analysis 1966-2007. Version 5.15.2008. USGS, Pa- tuxent Wildlife Research Center, Laurel. Maryland, USA. Shapiro, L. H., R. A. Canterbury, D. M. Stover, and R. C. Fleischer. 2004. Reciprocal introgression between Golden-winged Warblers {Vermivora chrysoptera) and Blue-winged Warblers (Vc pinus) in eastern North America. Auk 121: 1019-1030. Wilson, M. D., B. D. Watts, M. G. Smith, J. P. Bredlau, AND L. W. Seal. 2007. Status assessment of Golden- winged Warblers and Bewick’s Wrens in Virginia. Technical Report Series, CCBTR-07-02. Center for Conservation Biology. College of William and Mary, Williamsburg, Virginia, USA. The Wilson Jounuil of Ornithology 1 22(2):279 -287, 2010 GENETIC AND MORPHOLOGICAL VARIATION OF THE SOOTY-CAPPED BUSH TANAGER {CHLOROSPINGUS P/LEATUS), A HIGHLAND ENDEMIC SPECIES FROM COSTA RICA AND WESTERN PANAMA TANIA CHAVARRIA-PIZARRO,' - GUSTAVO GUTIERREZ-ESPELETA,' ERIC J. EUCHS,' AND GILBERT BARRANTES' ABSTRACT. — We examined the effect of geographic isolation on morphology, genetic structure, and abundance of the Sooty-capped Bush Tanager {Chlorospingits pHeatus), an endemic species restricted to highlands ot Costa Rica and western Panama. Abundance and morphology were measured at five study sites and genetic variation was calculated from three microsatellite loci. We expected geographic discontinuities in this species' distribution to have an effect on its morphology and genetic structure. Genetic variation was higher within than between populations with no effect of geographic barriers on population genetic divergence in this species, indicating gene flow is high between populations. Unique alleles were detected in each population and G, values increased with geographic distance between populations. Some morphological traits differed between populations, which may be caused by adaptation to different selective pressures in each population. Molecular data did not differ between the two color morphs that coexist in two isolated populations, which were considered different species. Received 16 July 2009. Accepted 22 November 2009. Highland endemic avian species in the moun- tains of Costa Rica and western Panama cuiTently show naturally isolated and discontinuous distri- butions, reflecting their confinement to mountain peaks and their isolation from other highland regions in the Neotropics (Stiles 1983; Barrantes 2000, 2009). Morphological and molecular diver- gence documented for some of these endemic highland species in Costa Rica correspond to this natural isolation (Stiles 1983, Barrantes 2000, Barrantes and Sanchez 2000). For example, isolated populations of the Volcano Hummingbird iSelasphorits flammula) and Fiery-throated Hum- mingbird (Panterpe insignis) differ in coloration and morphological dimensions (Stiles 1983, 1985). Similarly, geographic bamers have affect- ed the genetic and morphological divergence among populations of the Black-and-yellow Phainoptila {Phainoptila melano.xantha) (Bar- rantes 2000, Barrantes and Sanchez 2000). The Sooty-capped Bush Tanager (Chlorospin- gus pileatiis) is endemic to the highlands of Costa Rica and western Panama. This species is naturally restricted to the summit of high mountains and primarily inhabits forest edges and open areas in the cloud forest and paramo. Sooty- capped Bush Tanagers forage in flocks of 5-20 individuals, often accompanied by other small ' Escuela de Biologia, Ciudad Universitaria Rodrigo Facio, Univer.sidad de Costa Rica, San Jose. Costa Rica. ^Corresponding author; e-mail: taniachavarria79@yahoo.com birds (Powell 1985, Stiles and Skutch 1989). A grayish-green morph of Sooty-capped Bush Tan- ager (lighter morph) coexists on two high volcanoes (Irazu and Tumalba) of central Costa Rica with the darker and more widely distributed morph. The grayish-green morph was first con- sidered a different species (Zeledon’s Chloro- spingus, Chlorospingits zeldoni) by Ridgway (1905), and Eisenmann (1955) and Slud (1964) continued C. zelecloni as a different species. However, Carriker (1910) suggested that C. zelecloni was a color morph of Sooty-capped Bush Tanager. Johnson and Brush (1972) analyzed the structure and pigment composition of feathers of both taxa and concluded C. zelecloni was a moiph of Sooty-capped Bush Tanager. However, no further investigations have been conducted to test this hypothesis (e.g., using molecular data). We used molecular data to analyze the effect of isolation (i.e., distance and barriers between populations) on genetic structure and moiphology of the Sooty-capped Bush Tanager. We .specifi- cally hypothesized that small populations isolated by effective geographical barriers have a notable reduction in genetic variation. In addition, we searched for genetic differences between the two color morphs on Irazii Volcano. To our knowl- edge, this is the first fine-.scale genetic analysis, using microsatellites of an avian species endemic to the isolated highland regions of Costa Rica and western Panama. Most species in this hotspot of diversity and endemism are restricted to a reduced portion of the highland forests and have low 279 280 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 2, June 2010 FIG. 1. Sampling localities in Costa Rican mountain ranges; Talamanca, Central Volcanic, and Tilaran. Locations of Monterverde, Irazu, and Cerro de la Muerte are indicated by black arrows. abundance (Jankowski and Rabenold 2007, Bar- rantes 2009). METHODS Field Data Collection. — We conducted field work in highland forests of three Costa Rican mountain ranges: Talamanca Mountain Range, Central Mountain Range, and Tilaran Mountain Range (Fig. 1) from July 2003 through December 2005. Censuses and tissue collections were conducted at five different sites: Monteverde (10° 19' N, 84° 47' W; 1,550 m asl) in the Tilaran Mountain Range; Poas Volcano (10° 1 1' N, 84° 13' W; 2,700 m asl), Irazu Vol- cano (09° 59' N, 83° 52' W; 3,200 m asl), and Barva Volcano (10° 08' N, 84° 06' W; 2,800 m asl) in the Central Mountain Range; and Macizo Cerro de la Muerte (09° 33' N, 83° 43' W; 3,100 m asl) in the Talamanca Mountain Range. Mountain ranges and field sites within mountain ranges are separated by geographic discontinuities of different magnitude such as watersheds and mountain passes (Fig. 1 ). We censused birds 4-7 times in each of the sites to account for possible fluctuations due to climatic conditions. Censuses started at 0545 hrs and all individuals heard and seen within 25 m of each side of the transect were recorded. Length of transects varied from 1.9 to 3.0 km among sites. Thus, abundance was expressed as individuals/km to allow comparison among sites. We used abundance categories based on mean values across censuses for each site. Categories were defined as very common (>20 individuals observed/census), common (10-19 individuals/ census), uncommon (5-9 individuals/census), and rare (<5 individuals/census). We calculated the area of the potential available habitat for Sooty- capped Bush Tanagers on each mountain range using Geographical Information Software (Arc- View GIS 3.3) on a digital elevation map of Costa Rica ( 1 :2()(),()()(); Lambert conformal con- Chaviinia-Pizcirro cf at. • SOOTY-CAPPED BUSH TANAGERS 281 ical projeclion) (Fig. I ). This area was estimated from the lower limit of the altitudinal distribution of Sooty-capped Bush Tanager to the summit of the mountains in each mountain range. We set 5-8 mist nets (12 m long) during each visit to a study site. The following measurements were taken for each individual captured: body mass, length of the beak from distal end of nares to the beak tip, width of the beak at front of nares, length of the culmen from the front of the skull to the beak tip, tarsus from the tibiotarsus joint to the distal end of the tarsometatarsus, flattened wing cord, and tail length. These morphological traits were chosen as they are expected to respond adaptively to ecological differences in habitat (Grant 1986, Price and Boag 1987) and were taken by the same person (GB). Birds were banded (plastic bands) to avoid re-measuring the same individuals. Moiphological data were first analyzed using a stepwise (forward) discriminant function analysis (DFA). This option excludes those variables that have no further effect in explaining the between-population variance. We used a MANOVA on those variables included in the DFA. Statistical analyses were performed in Statistica Version 6.0 (StatSoft Inc., Tulsa, OK, USA). Laboratory Procedures. — ^DNA was extracted from liver tissue of 56 individuals that had been obtained in previous years and maintained in the tissue collection at the Zoology Museum at the University of Costa Rica. We collected blood samples from five additional individuals for a total of 61 individuals (14 from Cerro de la Muerte, 9 from Irazu Volcano, 18 from Barva Volcano, 1 1 from Poas Volcano, and 9 from Monteverde). Tissue samples were preserved in 95% alcohol and stored at —20° C until DNA extraction. DNA was extracted using a commer- cial extraction kit for animal tissue and blood (Promega, Madison, WI, USA) following the specifications provided by the manufacturer. Genotypes of the 61 individuals were obtained using three microsatellite loci (Escpl, Escp4, and Escp6; Hanotte et al. 1994) isolated from the related species Emheriza schoeniclus (Yuri and Mindell 2002). The three microsatellites used were the most polymorphic for E. schoeniclus. Microsatellites were amplified using an Eppen- dorf Mastercycler thermocycler in a volume of 25 pL (12.5 pL of PGR Master Mix 2X PROMEGA, 2.5 pL [10 pm] of each primer, 5 pL of DNA, and 2.5 pL Fl20) following protocols for each locus as described in Hanotte et al. (1994). Genotypes were obtained using an automatic sequencer (ABl Prism 310, Applied Biosystems, Palo Alto, CA, USA). Alleles were sized using program GENOTYPER (3.7NT, Applied Biosystems, Palo Alto, CA, USA). We counted the number of alleles per locus and calculated their frequencies in each population. We tested deviations of these alleles from Hardy- Weinberg equilibrium with a Chi-square test using program GENALEX 6 (Peakall and Smouse 2006); sequential Bonferroni correction was applied due to performance of multiple simulta- neous tests (Rice 1989). The observed (Ho) and expected (He) heterozygosity per locus in each population were calculated using program ARLE- QUIN Version 3.1 (Schneider et al. 2000, Excoffier et al. 2005). An analysis of molecular variance (AMOVA) was used to assess population structure with 30,000 permutations to infer significance of E^i values (Weir and Cockerham 1984). AMOVA calculations were conducted using ARLEQUIN Version 3.1 (http://cmpg. unibe.ch/software/arlequin3). We also checked all loci for the presence of null alleles (Chakra- borty et al. 1992, Brookfield 1996) using Micro- Checker Version 2.2.3 (University of Hull, Hull, UK; http://www.microchecker.hull.ac.uk/; van Oosterhout et al. 2004). All loci were tested based on a dinucleotide repeat motif and 1,000 permu- tations. We further examined population structure and admixture using a model-based clustering method as implemented in the Bayesian clustering program STRUCTURE (Pritchard et al. 2000, Falush et al. 2003). We inferred population structure using a model which allowed for admixture and correlated allele frequencies. Each Monte Carlo Markov Chain used to infer population structure was based on 1,000,000 iterations and a burn-in of 100,000. Multiple runs were conducted changing the number of putative populations (K) between one and five. At least three independent runs were assessed to estimate the likelihood of the data and the posterior probabilities for each fixed number of populations (K). We calculated genetic distances between pop- ulations using two parameters: Nei's (Dm) genetic distance and Es, (Nei 1987, Nei and Kumar 2000). We correlated genetic distances (E^i values) with geographic distances (km) using a Mantel test and examined whether area of habitat available for 282 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 2, June 2010 each population con'elated with allele number and/or genetic distance {Dm) using ARLEQUIN. RESULTS Null Alleles. — ^We found no evidence of null alleles on the first two loci (Escpl and Escp4) with combined probability for all cases of L* > 0.05. The third locus (Escp6) showed a significant homozygote excess congruent with the presence of null alleles (P < 0.001). No evidence of scoring errors or allele drop-out was found for any loci. Observed allele frequencies for the third locus only deviated slightly (at 0.01 level) from frequencies estimated considering the presence of null alleles. AMOVA analyses were performed twice using only Escpl and Escp4 loci, and using all three loci. Results were similar with small deviations at the 0.001 level. Thus, we used all three loci for further analysis. Genetic Variation and Population Structure. — Microsatellite loci were highly polymorphic within each population relative to the number of genotyped individuals (Table 1 ). Allele number in any locus ranged from seven to 18 across populations. The Barva Volcano population had the highest average number (± SD) of alleles (9.33 ± 0.66), while Monteverde had the lowest (5.67 ± 0.66; Table 2). All populations had unique alleles, although number of these alleles varied across populations: four in Barva, three in Cerro de la Muerte and Monteverde, and two in Irazu and Poas populations (Table 1). Frequency of the allele Escp6 significantly deviated from Hardy-Weinberg expectations (lower observed values than expected) in Cerro de la Muerte {X^ = 64.36, df = 36, P = 0.003), Barva Volcano {X^ = 1 13.56, df = 45, P < 0.0001 ), and Monteverde {X^ = 48.0, df = 21, P = 0.001). All populations had high levels of genetic diversity ba.sed on observed heterozygosity, which did not deviate significantly from expected heterozygosity in all populations (Table 2). Ge- netic variation was higher within populations than between populations; 98% of the total variance was distributed within populations and only 2% among populations (tp^, = 0.0153; P > 0.17). The.se results are consistent with high levels of gene flow among populations. Genetic structure estimates among populations were similar if the third locus was removed due to the likelihood of null alleles (tps, = 0.0154; P > 0.18). Similarly, analyses using STRUCTURE showed no evidence of significant population structure. The highest likelihood was obtained when the number of populations was set to one (K = 1), which suggests complete admixture. Pairwise values ranged from 0.04 between Barva and Poas to 0.09 between Irazu and Monteverde (Table 3). values and geographic distances were not significantly correlated (Man- tel test, r = 0.49, P = 0.09); similar results were obtained using a simple linear correlation. We did not detect any genotypic difference between three individuals of the grayish-green morph and six individuals of the darker morph in the Irazu population. Both color morphs shared some genotypes. Morphological Variation. — We captured 72 individual Sooty-capped Bush Tanagers; 24 at Cerro de la Muerte (Estacion Biologica Cerro de la Muerte), 11 at Irazu, 18 at Barva, 10 at Poas, and nine at Monteverde. Culmen length, wing cord, tail length, beak width, and body mass differed among populations (DFA: f’20.168 = 2.63, P = 0.0003). The Monteverde population had the shortest culmen and wing length and the widest beak; the Irazu population had the longest tail (Wilks’ lambda = 0.64, df = 16/174, P = 0.04; Table 4). Population Size and Habitat Area. — Sooty- capped Bush Tanagers were very common (>20 individuals/census, corrected by distance) at Barva and Poas, and common at other sites (10- 19 individuals/census). The highland habitat available for Sooty-capped Bush Tanagers de- crea,sed from the Talamanca Mountain Range to the Tilaran Mountain Range. The Talamanca Mountain Range has 80.8% of the total natural highland habitat potentially available for this species (293,135 ha), while the Central Mountain Range has 17% and the Tilaran Mountain Range has 3%. Area of available habitat was not coiTelated with number of alleles (/• = 0.53, n — 5, P = 0.16) nor with genetic distance (Nei’s) (r = 0.62, /? = 5, P = 0. 1 1) in Sooty-capped Bush Tanager populations. DISCUSSION All populations of the Sooty-capped Bush Tanager had high genetic variation as indicated by genetic diversity indices and number of alleles per locus. The genetic diversity within popula- tions of Sooty-capped Bush Tanager is unexpect- edly high for an endemic species with a small and naturally fragmented geographical distribution (Tables I, 2). Within-population genetic variation TABLE 1. Allele frec|uencies for three microsatellite loci (Esc|.il, Esci.i4, Esc|.t6) in five populations of Sooty-capped Bush Tanager (/; — individuals with amplified microsatellites per loci in each population) in the highlands of Costa Rica. Chavanhi-Pizarro d al. • SOOTY-CAPPED BUSH TANAGERS 283 sO o o So o O c O c d d d d d •T u o o o LU o o o o o d d d d d Os oo — os 00 tn os o o O (N cn — — o d d d d d so >/~) o so o in o o — d d d d d o r- in os in o CN o (n d d d d d r-j o in CN o — o o n d d d d d so in o — O o o o d d d d 00 •t o sO o o o q o o d d d d - o o o o o o o o uj d d d o sO o o r- 4 o o o — d d d d d ri -t in r^! o o d d d d 00 r»“, ’1’ o o o o o o q q d d -i- sO r- in sO O q o o o d d d d ri r*-, r- o o o o o o o o d d d d os in o os (U V “O u. u V > > C3 > u > V > O CQ (U OJ > > > O c o s CQ u < H 1 U '3 N 03 d a > CC CQ c/5 'C3 O o, < C. Muerte 0.00 0.00 0.27 0.04 0.00 0.04 0.18 0.23 0.04 0.00 0.00 11 0.00 0.00 0.09 0.00 Irazu V. 0.00 0.06 0.07 0.43 0.07 0.00 0.14 0.14 0.07 0.00 0.00 7 0.00 0.00 0.00 0.00 BarvaV. 0.00 0.10 0.00 0.11 0.00 0.29 0.07 0.07 0.04 0.00 0.04 14 0.21 0.04 0.00 0.00 Poas V. 0.00 0.00 0.06 0.12 0.06 0.12 0.06 0.06 0.00 0.00 0.00 8 0.00 0.06 0.12 0.00 Monteverde 0.06 0.17 0.00 0.00 0.00 0.06 0.00 0.00 0.00 0.06 0.00 7 0.00 0.00 0.00 0.12 TABLE 2. Number of individuals in), allele number (Na), observed heterozygosity (Ho), and expected heterozygosity (He) calculated for each locus in Eve populations of Sooty-capped Bush Tanager in Costa Rica. 284 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122. No. 2. June 2010 r- to (N cn o o O O o d d d d d u X + I +1 +1 +1 +1 to nC 00 00 00 00 r- d d d d d (O r- to — — CN — (N d d d d d Ho +1 +1 +1 -f-l +1 a IT) to 00 o c/2 'O r- NO + 1 d d d d d C cd u B m 00 NO m NO V3 m 00 NO m NO •2 d d d d d 0 H id Z +1 + 1 +1 +1 +1 r- m m r-- r- 00 m NO NO 00 00 os to m to 00 to r- r-; p — to — (N — d +1 +1 + 1 +1 +1 p fO NO nO cq NO m NO 00 iri ON 00 OJ nO NO m 00 00 00 00 00 d d d d d to o (N o o X to (^1 NO to d d d d d nO o rd ON r- o 00 r- z 00 r- X to 00 ON 00 00 On d d d d On r-- o On O X r- 00 NO ON 00 d d d d d G Na ON - o r- o ON 00 ON X 00 r4 r- NO 00 00 r- OO d d d d d r- oo r- c 00 o X r- r-' 00 r- d d d d d a CJ u] Na 00 00 00 00 oo -t ON ty^ o On V 3 ■o u. o o > > > > > V s 'O N c Q cJ f3 Ui C3 QQ O CL is expected to be small compared to between- population variation in species with small ranges and fragmented distributions, mainly due to genetic drift or founder effects and small effective population size (Peterson et al. 1992, Avise 1994). The absence of genetic structure among popula- tions of Sooty-capped Bush Tanagers indicates gene flow may be relatively high among popula- tions, counteracting the effects of genetic drift, small effective population size, and local adapta- tion even in small populations (Avise 1994). However, time since population isolation may have also affected the near absence of genetic structure in this endemic species. Populations are expected to retain most of their original genetic composition (Templeton et al. 1995) if isolation is relatively recent, as it has been hypothesized for Costa Rican highland endemic species (Stiles 1983, Barrantes and Sanchez 2000, Weir 2006, Barrantes 2009), There were some genetic and morphological differences across populations of the Sooty- capped Bush Tanager. For example, the presence of unique alleles in each population indicates that processes of random drift, inbreeding, and perhaps local adaptation are causing the genetic differ- ences among populations. Genetic difference between populations (Fs, values) tend to increase with geographic distance (P = 0.09) suggesting some limitation in dispersal. However, gene flow among populations likely dilutes the effect of these processes. Schwartz et al. (1986) indicated that gene flow rates as low as one migrant per generation are generally effective to prevent loss of genetic variation from habitat fragmentation. Morphology differed among some Sooty- capped Bush Tanager populations. Moiphological differences may be caused by different abiotic and biotic pressures operating on different popula- tions. For example, the habitat of the Sooty- capped Bush Tanager in Monteverde is exposed to relatively intense trade winds year round (Clark et al. 2000). Birds in this population had shorter wings (Table 4), which are more efficient at maneuvering under strong wind conditions (Pen- nycuick 1968, Thomas 1996, Margjjerie et al. 2007). There is no clear explanation for dilTer- ences in other morphological traits (e.g., culmen, tail, and beak width), as fruit species and habitat structure appear similar across localities (Bar- rantes and Loiselle 2002). Morphological diver- gence between populations has been reported for other endemic birds in the highlands of Costa Rica Chavania-Pizano ci al. • SOOTY-CAIM^ED lUJSIl T'ANAOERS 285 TABLE 3. Geographic distance (km) (above diagonal) and values (below the diagonal) between five populations of Sooty-capped Bush Tanager in the highlands of Costa Rica. Site C. Muerte Irazii V. Barva V. Piiiis V. Monteverde C. Muerte — 47 74 91 142 Irazii V. 0.042 — 34 48 103 Barva V. 0.047 0.063 — 15 71 Poas V. 0.063 0.060 0.040 — 55 Monteverde 0.053 0.090 0.059 0.078 — (Stiles 1983, 1985; Barrantes and Sanchez 2000), indicating rapid morphological divergence among populations, if recent population fragmentation is assumed (Stiles 1983, 1985; Barrantes and Sanchez 2000; Barrantes 2009). The lack of concordance between morphology and genetics in this study is possibly caused, at least in part, by the intrinsic characteristics of the microsatellites. These genetic markers are neutral (or nearly so) whereas morphological traits are frequently under strong selective pressures (Zwaitjes 2003, San- tiago-Alarcon et al. 2006), and changes in traits presumed to be under selection often do not correlate with variation of microsatellites (Philli- more et al. 2008) or mitochondrial DNA markers (McCormack et al. 2008, Weir et al. 2008). We found no genetic differences between the two color morphs in the Irazii population. Genotypes vary among individuals of each morph and some genotypes are shared with individuals of the other morph. Our findings support the hypothesis of Carriker (1910) and Johnson and Brush (1972), who suggested that differences in coloration reflect two morphs of the same species. We found individuals of the grayish-green moiph vary greatly in coloration. All grayish-green birds also had yellow nasal tufts, and one had one yellow secondary feather in its right wing (GB, unpubl. data). Evidence that helps explain this polymorphism is elusive. Birds of both morphs use the same habitat, forage together, and apparently mate with each other (Johnson and Brush 1972; GB, unpubl. data). Johnson and Brush ( 1972) suggested that brighter coloration of grayish-green individuals is favored during peri- ods of high volcanic activity when ash in the atmosphere reduces visibility. However, informa- tion on abundance of each morph during high and low volcanic activity periods is needed to test this hypothesis. This polymorphism may have evolved recently by assortative mating with some repro- ductive barriers, but its detection only would be possible with detailed genetic and behavior studies (Yeh 2004, McGlothlin et al. 2005). The geographic distribution of the Sooty- capped Bush Tanager on only three mountain ranges, the number of naturally isolated popula- tions in = 5), as well as the low number of alleles analyzed (n = 3) imposed some restrictions on analyses and limited the statistical power of some results. For example, the low number of micro- satellite loci limited use of more detailed analyses to quantify gene flow. Similarly, lack of signif- icant correlation between geographic and pairwise genetic differences was likely caused by the small number of isolated populations. In addition, we did not survey Sooty-capped Bush Tanagers in western Panama where the Fortuna Mountain Pass may limit bird movement across this geographic discontinuity, affecting the scope of our findings. The geographic discontinuities that separate populations of Sooty-capped Bush Tanagers as well as the area of available habitat appear to have had little effect in shaping the genetic structure of this endemic species. The presence of divergent morphologies and color moiphs in some popula- tions and unique alleles in all populations of Sooty-capped Bush Tanager suggest some isola- tion and/or local adaptation of populations (Stiles 1983, 1985; Barrantes and Sanchez 2000). This TABLE 4. Morphological trails (mean ± SD) for five populations of Sooty-capped Bush Tanager in the highlands of Costa Rica. (Length = mm. body mass = g). Morphology trails Site Culmen length Wing cord Tail length Tarsus length Beak length Beak width Mass C. Muerte 15.35 ±1.11 71.6 ± 2.91 64.0 ± 4.65 23.85 ± 0.76 6.38 ± 0.32 6.38 ± 0.32 19.90 ± 1.44 Irazii V. 16.0 ± 0.84 70.3 1 ± 2.63 55.0 ± 2.20 24.21 ± 0.88 6.41 ± 0.28 6.42 ± 0.28 19.87 ± 0.87 Barva V. 15.22 ± 0.91 69.7 ± 2.70 61.66 ± 3.52 23.70 ± 0.76 6.40 ± 0.48 6.40 ± 0.48 20.28 ± 1.16 Poas V. 15.88 ± 1.15 69.5 ± 2.51 61.93 ± 2.09 23.68 ± 0.90 6.36 ± 0.38 6.36 ± 0.38 20.37 ± 2.57 Monteverde 13.58 ± 2.71 68.1 ± 4.35 55.57 ± 3.69 23.14 ± 1.15 6.43 ± 0.77 6.43 ± 0.77 17.65 ± 1.67 286 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2, June 2010 system of geographically isolated populations deserves special conservation attention, because it presents a unique opportunity to further study incipient causes of population divergence in an area of high endemism. ACKNOWLEDGMENTS We thank Johel Chaves, John McCormack, Jason Weir, and C. E. Braun tor helpful comments that largely improved the manuscript. We also thank Cesar Sanchez for helping with field work and Federico Valverde for logistic support. Financial support was provided by the Consejo Nacional para Investigaciones Cientificas y Tecnologicas (CON- ICIT), the Asociacion Ornitologica de Costa Rica, and the Vicerrecton'a de Investigacion, Universidad de Costa Rica. LITERATURE CITED Avise, J. C. 1994. Molecular markers, natural history and evolution. Chapman and Hall Press, New York, USA. Barrantes, G. 2000. Ecology and evolution of Phainoptila melano.xantha (Bombycilidae, Aves) in the highlands of Costa Rica and western Panama. Dissertation. University of Missouri-St. Louis, USA. Barrantes, G. 2009. The role of historical and local factors in determining species composition of the highland avifauna of Costa Rica and western Panama. Revista Biologi'a Tropical 57 (Supplement l):323-332. Barrantes, G. and J. Sanchez. 2000. A new subspecies of Black and Yellow Silky Flycatcher, Phainoptila melano.xantha (Bombycilidae, Aves), from Costa Rica. Bulletin of the British Ornithological Club 120:40-46. Barrantes, G. and B. A. Loiselle. 2002. Reproduction, habitat use and natural history of the Black-and- Yellow Silky Flycatcher (Phainoptila melano.xantha), an endemic bird of western Panama-Costa Rica highlands. Ornitologi'a Neotropical 13:121-136. Brookfield, J. F. Y. 1996. A simple new method for e.stimating null allele frequency from heterozygotc deficiency. Molecular Ecology 5:453—455. Carriker Jr., M. A. 1910. An annotated list of birds of Costa Rica, including Cocos Island. Annals of the Carnegie Museum 6:314-915. Chakraborty, R., M. De Andrade, S. P. Daiger, and B. Budowle. 1992. Apparent heterozygote deficiencies observed in DNA typing data and their implications in forensic applications. Annals of Human Genetics 56:45-47. Clark. K. L., R. O. Lawton, and P. R, Butler. 2000. The physical environment of Monteverde. Pages 10-25 in Monteverde: ecology and con.servation of a tropical cloud fore.st (N. M. Nadkarni and N. T. Wheelwright, Editors). University of Chicago Press, Chicago, Illinois, USA. Eisenmann, E. 1955. The species of Middle American birds. Transactions of the Linnean Society of New York 7:1-128. Excoffier, L., G. Laval, and S. Schneider. 2005. ARLEQUIN Version 3.0: an integrated software package for population genetics data analysis. Evolu- tionary Bioinformatics Online 1 :47-50. Falush, D., M. Stephens, and J. K. Pritchard. 2003. Inference of population structure using multilocus genotype data: linked loci and correlated allele frequencies. Genetics 164:1567-1587. Grant, P. 1986. Ecology and evolution of Darwin’s Finches. Princeton University Press, Princeton, New Jersey, USA. Hanotte, O., C. Zanon, A. Pugh, C. Greic, A. Dixon, AND T. Burke. 1994. Isolation and characterization of microsatellite loci in a passerine bird: the Reed Bunting Emheriza schoenielus. Molecular Ecology 3:539-530. Jankowski, J. E. and K. N. Rabenold. 2007. Endemism and local rarity in birds of neotropical montane rain forest. Biological Conservation 138:453—463. Johnson, N. K. and A. H. Brush. 1972. Analysis of polymorphism in the Sooty-capped Bush Tanager. Systematic Zoology 21:245-262. Marguerie, E., J. B. Mouret, S. Doncieux, and J. A. Meyer. 2007. Artificial evolution of morphology and kinetics in a flapping-wing mini-UAV. Bioinspiration and Biomimetics 2:65-82. McCormack, J. E., A. T. Peterson, E. Bonaccorso, and T. B. Smith. 2008. Speciation in the highland of Mexico: genetic and phenotypic divergence in the Mexican Jay (Aphelocoma iiltramarina). Molecular Ecology 12:423-434. McGlothlin, j. W., P. G. Parker, V. Nolan Jr., and E. D. Ketterson. 2005. Correlational selection leads to genetic integration of body size and an attractive plumage trait in Dark-eyed Juncos. Evolution 59: 658- 671. Nei, M. 1987. Molecular evolutionary genetics. Columbia University Press, New York, USA. Nei, M. and S. Kumar. 2000. Molecular evolution and phylogenetics. Oxford University Press, New York, USA. Peakall, R. and P. E. Smouse. 2006. GENALEX6: genetic analysis in Excel. Population genetic software for teaching and research. Molecular Ecology Notes 6:288-295. Pennycuick, C. j. 1968. A wind-tunnel study of gliding flight in the pigeon Cohunha livia. Journal of Experimental Biology 49:509-526. Peterson, A. T., P. Escalante, and S. A. Navarro. 1992. Genetic variation and differentiation in Mexican populations of Common Bush-Tanagers and Chestnut- capped Brush-Finches. Condor 94:244-253. Phillimore, a. B, 1. P. F. Owens, R. A. Black, J. CiiiTOCK, T. Burke, and S. M. Clegg. 2008. Complex patterns of genetic and phenotypic diver- gence in an island bird and the consequences for delimiting conservation units. Molecular Ecology 17:2839-2853. Powell, G. V. N. 1985. Sociobiology and adaptive significance of interspecific foraging flocks in the Ncotropics. Ornithological Monographs 36:713-731. Price, T. D. and P. T. Boag. 1987. Selection in natural populations of birds. Pages 257-287 in Avian genetics: Cluivarrki-Pizarro el al. • SOOTY-CAPPED BUSH TANAGERS 287 a population and ecology approach (F. Cooke and P. A. Buckley, Editors). Academic Press, London, United Kingdom. Pritchard, J. K., M. Stephens, and P. Donnelly. 2000. Inference of population structure using multilocus genotype data. Genetics 155:945-959. Rice, W. R. 1989. Analyzing tables of statistical tests. Evolution 43:223-225. Ridgway, R. 1905. New genera of Tyrannidae and Turdidae, and new forms of Tanagridae and Turdidae. Proceedings of the Biological Society of Washington 17:211-214. Santiago-Alarcon, D., S. M. Tanksley, and P. G. Parker. 2006. Morphological variation and genetic structure of Galapagos Dove (Zenaida galapagoensis) populations: issues in conservation for the Galapagos bird fauna. Wilson Journal of Ornithology 118:194- 207. Schneider, S., D. RossiNt, and L. Excoffier. 2000. ARLEQUIN Version 2000: a software for population genetics data analysis. Genetics and Biometry Labo- ratory, University of Geneva, Switzerland. Schwartz, O. A., V. C. Bleich, and S. A. Holl. 1986. Genetics and the conservation of mountain sheep Ovis canadensis nelsoni. Conservation Biology 37:179-190. Slud, P. 1964. The birds of Costa Rica. Distribution and ecology. Bulletin of the American Museum of Natural History 128:1^30. Stiles, F. G. 1983. Systematics of the southern forms of Selasphorus (Trochilidae). Auk 100:311-325. Stiles, F. G. 1985. Geographic variation in the Fiery- throated Hummingbird, Panterpe insignis. Ornitholog- ical Monographs 36:23-30. Stiles, F. G. and A. F. Skutch. 1989. A guide to the birds of Costa Rica. Cornell University Press, Ithaca, New York, USA. Templeton, A. R., E. Routman, and C. A. Phillips. 1995. Separating population structure from population histo- ry: a cladistic analysis of the geographical distribution of mitochondrial DNA haplotypes in tiger salamander, Amhystoma trigrinum. Genetics 140:767-782. Thomas, A. L. R. 1996. The Bight of birds that have wings and a tail: variable geometry expands the envelope of flight performance. Journal of Theoretical Biology 183:237-245. Van Oosterhout, C., W. F. Hutchinson, D. P. M. Wills, AND P. Shipley. 2004. Micro-Checker: software for identifying and correcting genotyping errors in micro- satellite data. Molecular Ecology Notes 4:535-538. Weir, B. S. and C. C. Cockerham. 1984. Estimating F- statistics for the analysis of population structure. Evolution 38:1358-1370. Weir, J. T. 2006. Divergent timing and patterns of species accumulation in lowland and highland neotropical birds. Evolution 60:845-855. Weir, J. T., E. Birmingham, M. J. Miller, J. Klicka, and M. A. Gonzalez. 2008. Phylogeography of a morphologically diverse neotropical montane species, the Common Bush-Tanager (Chlorospingus ophthal- micus). Molecular Phylogenetics and Evolution 47:650-664. Yeh, P. j. 2004. Rapid evolution of a sexually selected trait following population establishment in a novel habitat. Evolution 58: 166-174. Yuri, T. and D. P. Mindell. 2002. Molecular phylogenetic analysis of Fringillidae, New World nine-primaried oscines. (Aves: Passeriformes). Molecular Phyloge- netics and Evolution 23:229-243. ZwARTJES, P. W. 2003. Genetic variability in migratory and endemic island songbirds (genus Vireo): a comparative assessment using molecular and morphological traits. Conservation Genetics 4:749-758. The Wilson Journal of Orniihulogy 122(2):288-295, 2010 LONG-TERM CHANGES IN AVIAN COMMUNITY STRUCTURE IN A SUCCESSIONAL, EORESTED, AND MANAGED PLOT IN A REEORESTING LANDSCAPE ELIZABETH W. BROOKS' AND DAVID N. BONTER' --' ABSTRACT. — We examined a 35-year transition in the breeding bird community at a successional study site in a reforesting landscape in southwestern New York, USA. Changes in the successional plot were compared with those in two additional census plots, one in undisturbed forest and the other in a managed tree farm. The territories of 7,429 singing male songbirds were mapped on the census plots. The most dramatic changes in community structure were in the successional plot where total number of territories declined between the beginning of the study (.v = 95.8 territories, 1969-1973) and end of the study (x = 57.2 territories, 1999-2003); grassland/shrub nesting species were nearly extirpated, and the number of neotropical migrant territories increased from zero in 1969 to 30 in 2003. The average number of neotropical migrant territories in the undisturbed forest plot declined from the beginning of the study (.v = 54.0, 1975-1979) to the end (.v = 44.8, 2003-2007). The average number of territories increased in the managed tree farm from the beginning of the study (.v = 53.6, 1983-1987) to the end (.v = 104.0, 2003-2007) largely due to increases in abundance of temperate zone migrants and resident species. Counts of individual species in the census plots were not highly correlated with counts from regional Breeding Bird Survey routes. Received 28 March 2009. Accepted 20 October 2009. Long-term studies are required to better under- stand the effects of anthropogenic change on natural ecosystems (Collins 2001), and a need for detailed long-term studies in ornithology has been recognized (Wiens 1984). Relatively few studies follow populations for more than two decades; many long-term studies have not followed the same protocol every year (Hall 1984, Leek et al. 1988, McNulty et al. 2008), or have only sampled avian communities at end points of a long time period (Ambuel and Temple 1982, Kirk et al. 1997, Haney et al. 2008). Studies examining changes in community structure following anthro- pogenic or natural change have tended to focus on succession following timber harvests or forest fires (Hagan et al. 1997, Hobson and Schieck 1999, Venier and Pearce 2005). Recent attention has focused on the inOuence of forest fragmentation on bird communities (Ste- phens et al. 2004), but many areas are undergoing large-scale regeneration rather than forest frag- mentation. Little attention has been directed at avian community structure following agricultural abandonment and secondary succession (but see Johnston and Odum 1956), a widespread process in eastern North America. Many areas in New England support more forested land today than at any point in the last two centuries as marginal ' Braddock Bay Bird Ob.servatory, P. O. Box 12876, Rochester, NY 14612, USA. ^Cornell Laboratory of Ornithology, 159 Sapsucker Woods Road. Ithaca. NY 14850, USA. ’Corresponding author; e-mail: dnb23@cornell.edu agricultural lands are abandoned and allowed to undergo secondary succession, resulting in pro- found changes in bird communities (Hagan 1993). Grassland bird declines have been attributed, in part, to abandonment and reforestation of farmland (Norment 2002); the grassland area in New York and New England has declined by 60% since the 1930s (Vickery et al. 1994). Secondary succession and maturation of forests have been cited as mechanisms driving long-tenn changes in popula- tions of shaib specialists like the Eastern Towhee (Pipilo erythrophthalnms) (Hagan 1993) and spe- cies that prefer eaiiy-successional forest like the Least Flycatcher {Etnpichnax minimus) (Holmes and Sherry 1988). Additional studies of avian community changes during and following agricul- tural abandonment are warranted, and would benefit from long-term studies at focal sites. Intensive, long-term field studies can also be useful for assessing and validating the utility of larger-scale monitoring data. For instance, results from intensive studies have been compared with regional trends detected by the Breeding Bird Survey (BBS) (Holmes and Sherry 2001, Ballard et al. 2003, McNulty et al. 2008). These local, intensive studies may not be subject to many of the limitations often associated with BBS data such as road-bias or changes in observers (Kendall et al. 1996, Keller and Scallan 1999, Niemuth et al. 2007). We can have greater confidence in the results if trend estimates generated from different methodologies agree. We document a 35-year transition in the breeding bird community at a study site that has 288 Brooks cold Bonier • LONG-TERM COMMUNI'I'Y CHANC3HS 289 undergone seeondary succession in a region experiencing reforestation. In addition, we com- pare trends on the successional plot with addi- tional long-term censuses of a forested plot and on an actively managed tree farm. Our objectives were to; (1) quantify long-term changes in breeding bird communities on intensively studied plots that experienced different management regimes, and (2) compare changes in the avian communities on the census plots with regional trends measured by the Breeding Bird Survey. METHODS Intensive Census Plots. — Breeding bird temto- ries were mapped on three long-term census plots. Territories were defined by spot mapping (Sven- son et al. 1970), a reliable method of estimating populations of breeding birds (James and Warner 1982). At least eight visits were made to each plot at approximately weekly intervals during the breeding season from late May to late July annually. All observations were recorded by a single observer (EWB) throughout the course of the study. The census plots were searched during each visit on foot for ~ 1 .5 hrs, and temtories of singing males were mapped. All partial territories (territories extending beyond the boundaries of the study plots) were considered to be whole territories in the analysis. A 9.3-ha “successional” plot transitioned from an open agricultural field to secondary growth forest during the study (35 years, 1969-2003). The “forest” plot was a 16.6-ha area in mixed deciduous-coniferous forest that was relatively undisturbed during the course of the study (33 years, 1975-2007). The “plantation” plot included 10.7 ha in an actively managed Christ- mas tree farm; the habitat was maintained at a relatively early successional stage in the planta- tion plot, although vegetation height increased over the course of the study (25 years, 1983- 2007). All three plots were in Allegany County (42° 14' N, 77° 49' W) in southwestern New York State. BBS Trends. — ^The Breeding Bird Survey (BBS) is an annual survey of bird diversity and abundance conducted along hundreds ol roadside routes in North America during the breeding season each year. A route is composed of a series of 50 stops distributed across 39.4 km. A skilled volunteer observer records all individual birds that are heard or observed at each stop during a 3-min point count (Robbins et al. 1986). Data from the two BBS survey routes nearest to the census plots were analyzed for this study (routes 61045 and 61046). These BBS routes were surveyed by the same observer as the census plots (EWB) with the observer being the primary recorder on these routes beginning in 1991 (route 61045) and 1973 (route 61046). Data Analysis. — ^Population trends for each species on each census plot were calculated with Poisson regression using Proc GENMOD in SAS (SAS Institute 2003). Poisson models adequately fit the data in 137 of 141 runs (using the Deviance Chi-square method; models not fitting the data showed signs of overdispersion). Species were included in the analysis if they were detected in at least 5 years within a plot. Data from Tree Swallows (Tachycineta hicolor) and Eastern Bluebirds (Sialia sialis) were eliminated from analysis as the presence of these species was dependent upon the availability of artificial nesting boxes, and nest boxes were not consis- tently provided throughout the course of the study. Birds were grouped according to life history characteristics (Table 1) and species were classi- fied by migration strategy (neotropical migrant vs. temperate migrant/resident) and by preferred breeding habitat (woodland vs. grassland/shrub). Classifications were based on life history infor- mation available through the Patuxent Wildlife Research Center (Gough et al. 1998). Species richness was calculated within years as the total number of species detected on the plot. Counts for BBS routes 61045 and 61046 were downloaded from the BBS web site (Sauer et al. 2008). We tested for correlations between BBS counts and census counts for all years with census data except for 2006 when the BBS was not conducted on route 61046. CoiTclations coeffi- cients (/•) among counts over time for all census plot/BBS route combinations were calculated using Proc CORR in SAS. RESULTS Widespread changes in the avian community were detected on the successional plot where number of total territories declined (;(■ I = 138.8, P < O.OOl) and grassland-nesting species essen- tially disappeared. Numerous nesting species were lost from the plot within the first 10 years including Clay-colored Sparrow (Spizella pallida). Vesper Sparrow {Pooecetes gramineus). Savannah Sparrow {Passerculus sandwichensis), Henslow’s 290 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 2. June 2010 TABLE 1. Species, life-history classifications, number of territories recorded, and trend estimates for breeding birds recorded on long-term studies conducted in successional, forest, and tree farm (plantation) census plots. Species Breeding habitat" Migration status^ Successional n (trend)"" Forest n (trend) Plantation n (trend) Ruffed Grouse Bonasa umhellus w T 19(+) 6 Wild Turkey Meleagris gallopavo w T 11 Sharp-shinned Hawk Accipiter striatus w T 5 Cooper’s Hawk A. cooperii w T 5 Red-shouldered Hawk Biiteo lineatus w T 10 (+) Broad-winged Hawk B. platypterus w N 24 Mourning Dove Zenaida macroura s T 18 (+) 6 23 (+) Black-billed Cuckoo Coccyzus erythopthalmus w N 10 (-) Barred Owl Strix varia w T 12 Yellow-bellied Sapsucker Sphyrapicus varius w T 9 (+) Downy Woodpecker Picoides puhescens w T 16 Hairy Woodpecker P. villas us w T 17 Pileated Woodpecker Dryocopus pileatus w T 7 Alder Flycatcher Empidonax alnorum s N 41 Great Crested Flycatcher Myiarchus crinitus w N 10 (+) Eastern Kingbird Tyrannus tyrannus s N 9 13 Blue-headed Vireo Vireo solitarius w N 11 (+) 81 (+) Red-eyed Vireo V. olivaceus w N 8 Blue Jay Cyanocitta cristata s T 33 (+) 91 7 (+) American Crow Con’us hrachyrhynchos N/A T 19(+) Horned Lark Eremophila alpestris s T 8 Black-capped Chickadee Poecile atricapillus w T 32 (+) 112 Red-breasted Nuthatch Sitta canadensis w T 8 (+) 94 White-breasted Nuthatch S. carolinensis w T 11 (-) Brown Creeper Certhia americana w T 68 House Wren Troglodytes aedon s N 25 Winter Wren T. troglodytes w T 17 (+) Golden-crowned Kinglet Regulus satrapa w T 17 (+) 209 Veery Catharus fuscescens w N 22 (-) Hermit Thrush C. giittatus w T 18 (+) Wood Thrush Hylocichla niustelina w N 44(-) American Robin Turdiis migratorius s T 176 (-) 98 102 (-P) Gray Catbird Dumetella carolinensis s N 30 9 Brown Thrasher To.xostoma riifum s T I4(-) 17 Cedar Waxwing Bomhycilia cedrorum s T 102 (-r) 20 (+) 136 (-P) Blue-winged Warbler Vermivora pinus s N 8 (+) Nashville Warbler V. ruficapilla s N 26 Chestnut-sided Warbler Dendroica pensylvanica s N 17 (+) 26 (+) Magnolia Warbler D. magnolia w N 78 (+) 308 15 (+) Yellow-rumped Warbler D. corona ta w T 105 (-H 128 (-I-) 70 (+) Black-throated Green Warbler D. virens w N 12 337 (-) Blackburnian Warbler D. fusca w N 10 (+) 323 Prairie Warbler D. discolor s N 7 24 Ovenbird Seiurus aurocapilla w N 18 (+) 77 Mourning Warbler Oporornis Philadelphia s N 19 (+) Common Yellowthroat Geothlypis trichas s N 195 82 22 Canada Warbler Wilsonia canadensis w N 8 (-) Scarlet Tanager Piranga olivacea w N 18 (-) Eastern Towhee Pipilo erythrophthalmus s T 82 (-) 29 Chipping Sparrow Spizella passerina s T 375 (-) 69 430 (-1-) Field Sparrow S. pusilla s T 129 (-) 102 (-) Vesper Sparrow Pooecetes gramineus s T 15 Savannah Sparrow Passerculus sandwichensis s T 45 (-) Grasshopper Sparrow Ammodranius savannarum s N 26 (-) Song Sparrow Melo.spiza melodia s T 414(-) 39 378 {+) Brooks and Bonier • LONG-TERM COMMUNITY CHANGES 291 TABLE 1. Continued. Species Breeding liabilaf Migration status'’ Successional n (trend)’’ Forest n (trend) Plantation n (trend) White-throated Sparrow Zonotrichia alhicollis s T 18 (+) 17 Dark-eyed Junco Junco hyenialis w T 40 (+) 246 Northern Cardinal Cardinalis cardinalis s T 6 9 (-) Rose-breasted Grosbeak Pheucticiis ludovicianiis w N 9 (-) Indigo Bunting Passerine! cyanea s N 60 (+) 51 22 Bobolink Dolichonyx oryzivorus s N 31 Red-winged Blackbird Agelaius phoeniceus s T 30 Brown-headed Cowbird Molothriis ater s T 52 (-) 34 19 (+) Purple Finch Carpodacus purpureus w T 114 (-) 56 (+) 83 (-I-) House Finch C. mexicanus N/A T 7 American Goldfinch Spinus tristis s T 6 50 (+) “ Breeding habitat and migration classifications are from Gough et al. (1998), W = woodland nesting species, S = successional species including grassland and shrubland nesting birds. T = temperate, including permanent residents and short-distance migrant species. N = neotropical migrants. Trend sign indicates significant iP < 0.05) population increase (-I-) or decrease (-) over time, and period (.) = insufficient data. No sign indicates the lack of a significant directional trend. Sparrow (Ammodramiis henslowii). Bobolink (Dolichonyx oiyzivorus). Red-winged Blackbird {Agelaius phoeniceiis), and Common Crackle (Qiiiscaliis quiscula). The number of territories of neotropical migrants (/^i = 50.5, P < 0.001) and woodland-nesting species (x^i = 55.1, P < 0.001) increased despite the decline in total territories on the successional plot, (Fig. lA). Positive trends were detected for 17 species (49%, P < 0.05, Table 1) and declines were detected for nine species (26%). Total species richness varied considerably but increased over time (/^i = 23.2, P < 0.001, Fig. 2). The avian community on the forest plot remained relatively stable with significant de- clines detected for eight species (17%, Table 1) and increa.ses detected for 12 species (25%). No trends were detected in total territories {x^\ =0.1, P = 0.712) or in species richness (x^\ = 0.6, P = 0.430) over time. The number of territories of neotropical migrants declined (/^i = 9.2, P = 0.003) while territories of shorter-distance mi- grants and resident species increa.sed (x^i = 12.6, P < 0.001, Fig. IB). The total number of territories increased on the managed tree farm {x'^\ = 101.9, P < 0.001), where the number of breeding pairs significantly increased for 11 species (42%, Table 1) and declined for only three species (12%). Increases in number of territories were primarily among resident and temperate-zone migrants {x^\ — 52.1, P < 0.001) with grassland/shrub nesting species (x^i = 20.8, P < 0.001) and woodland nesting species (x^i = 55.8, P < 0.001) becoming more common over time (Fig 1C). Counts of species on local BBS routes were not highly correlated with counts of breeding territo- ries for each species on the census plots. Comparing counts on BBS route 61045 with the census plots, the average con-elation coefficients were greatest for the successional plot {x corre- lation, r = 0.25), followed by the managed tree farm {x = 0. 14) and forest plot {x — 0.07). Counts of individual species on BBS route 61046 were also best correlated with counts on the succes- sional plot (.V = 0.23), followed by the forest plot (T = 0.16) and plantation plot (.v = —0.06). On average, the counts of individual species on the two BBS routes were not highly correlated (.v = 0.17). DISCUSSION We documented profound changes in commu- nity structure and abundance of birds in an area undergoing secondary succession following agri- cultural abandonment, while a nearby, undis- turbed forest bird community remained relatively static. The importance of disturbance and succes- sion in structuring avian communities has long been appreciated despite the paucity of long-term studies documenting the dynamic changes in communities over time (reviewed by Brawn et al. 2001). Some form of disturbance is required to meet the habitat needs of many species, and the loss of grassland and shrub habitats in North America has contributed to great declines in birds 292 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 2, June 2010 B. Forest plot ■ Neotropical migrants ■ Woodland • Temperate/Resident • Shrub/Grassland FIG. I. Total territories (// = 7,429) mapped on intensive census plots over 35 years, where temperate migrants and resident species that nest in early successional habitats demonstrated the strongest trends, declining in the successional landscape (A), remaining relatively stable in an undisturbed forest plot (B), and becoming more common in the managed tree farm (plantation, C). that specialize in these environments (Askins 2000, Brawn et al. 2001), Clearing of forests in eastern North America created early-successional habitats that allowed for geographic spread and numerical increases of species adapted to these environments. Recent changes in land use patterns, however, have led to a dramatic shift in the structure of bird commu- nities. The area of cultivated cropland east of the Mississippi River declined by ~22% between 1950 and 2000 (Brown el al. 2005), largely due to the loss of small farms. These declines in small- Brooks and Bonier • LONG-TFRIVI COMMUNITY CHANGHS 293 * Successional • Forest o Plantation FIG. 2. Changes in species richness on census plots in successional. forested, and managed tree farm (plantation) habitats. scale agriculture contributed to similar declines in populations and distributions of early-succession- al birds, leading to local and regional extirpation of some species (Hunter et al. 2001). Johnston and Odum (1956) documented chang- es in the richness of species nesting in the Piedmont of Georgia on plots of various succes- sional stages with increasing richness recorded with increasing age of the vegetation. We also found an increase in species richness during the early stages of secondary succession. However, in contrast to the earlier study, density of nesting birds decreased significantly as our successional study site transitioned from grassland to forest over a 35-year period. The rapid decline in grassland- and shrub-nesting species was not matched by a similarly rapid increase in wood- land-nesting birds, many of which are neotropical migrants. This lack of replacement by woodland-nesting species may be a function of a general decline in the abundance of neotropical migrants in the region (Robbins et al. 1989). Holmes and Sherry (2001), in one of the few intensive, long-term studies of breeding bird communities, recorded a decline in the abundance of breeding birds in a 30- year study in relatively undisturbed, contiguous forest in New Hampshire. Declines were most pronounced for neotropical migrants. Other stud- ies have reported long-term declines in popula- tions of long-distance migrants at sites in West Virginia (Hall 1984), New Jersey (Leek et al. 1988), and Sweden (Enemar et al. 1994). In contrast, Askins and Philbrick (1987) detected a decline and subsequent rebound in abundance ol many neotropical migrants during a 32-year study at an isolated 23-ha woodlot in Connecticut. The amount of forest in the stirrounding landscape during the Connecticut study initially declined and then increased, suggesting a link between abundance of long-distance migrants and regional forest area. A general decline in abundance of long- distance migrants could explain the relatively shallow increase in neotropical migrant territories on our successional study plot as the habitat became more suitable for many of these wood- land-breeding birds. As in the earlier studies, the abundance of long-distance migrants declined over time on the undisturbed forest plot censused in our study, suggesting that local populations of neotropical migrants may be declining. Quantifying changes in the abundance and distribution of birds remains one of the major methodological challenges in avian ecology and conservation biology. Most population and trend estimates in North America are based on Breeding Bird Survey data, yet concerns over the adequacy of the BBS continue to be raised (Keller and Scallan 1999, Francis et al. 2005) and addressed through novel analytical techniques (e.g., Royle et al. 2005). Confidence in trend estimates generated from BBS data can be strengthened if these trends agree with data from other monitoring schemes. For instance, McNulty et al. (2008) compared long-term trends in an old-growth forest plot with regional BBS trends in a relatively static land- scape (the Adirondack Mountains of New York) and found agreement in the direction of the trend for 74% of the species monitored. Our compar- ison of bird counts on local BBS routes with counts on intensive census plots failed to show high agreement among the surveys. The greatest agreement, however, was between the BBS routes and the successional plot, a logical result in a region experiencing the abandonment of small- scale agriculture. Analyses comparing results of long-term, intensive research with broad-scale monitoring programs such as the BBS should provide a more complete picture of the status of and temporal changes in bird populations. Intensive studies may be the only way to monitor trends in abundance and distribution of .species that are hard to detect, or species that occur in low densities. For example. Clay-colored Sparrows were not detect- 294 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2, June 2010 ed on either BBS route in the current study, although they were recorded nesting within the successional and managed tree farm plots. Long-term, intensive studies can provide unique perspectives on ecological phenomena that are unattainable in most contemporary studies that occur over restricted temporal scales. More long-term research projects should be encouraged (Wiens 1984). Further, fluctuations in populations detected in this study and elsewhere (Holmes and Sherry 2001) highlight the value and necessity of conducting continuous, long-term studies when studying population trends. ACKNOWLEDGMENTS We especially thank the late Clarence Klingensmith for help in setting up the study plots, conducting the habitat surveys in 1969 and 1998, and for advice and encourage- ment through the years. We thank Cynthia Clements, Phillips Foster, Tom and Kathy Kent, and the late Alice Foster and Eddy Foster for permission to conduct the study on their land. We appreciate the assistance of Gaylord Rough, Dennis Smith, Rick Walker, and the late Robert Place. We thank Doris and the late Lou Burton, Peter Gradoni, and William Howe for a.ssistance in the field. Research was supported by a grant from the Alfred University Research Foundation during the 1970 field season. We thank C. E. Braun and two anonymous reviewers for comments that improved the manuscript. LITERATURE CITED Ambuel, B. and S. A. Temple. 1982. Songbird populations in southern Wisconsin forests- 1954 and 1979. Journal of Field Ornithology 53; 149-158. Askins, R. A. 2000. Restoring North America’s birds: lessons from landscape ecology. Yale University Press, New Haven, Connecticut, USA. ASKIN.S, R. A. AND M. .1. Philbrick. 1987. Effect of changes in regional forest abundance on the decline and recovery of a forest bird community. Wilson Bulletin 99: 7-21. Ballard, G., G. R. Geupel, N. NtJR, and T. Gardali. 2003. Long-term declines and decadal patterns in population trends of songbirds in western North America, 1979-1999. Condor 105: 737-755. Brawn, J. D., S. K. Robin.son, and F. R. Thomp.son III. 2001. The role of disturbance in the ecology and conservation of birds. Annual Review of Ecology, Evolution, and Systematics 32: 251-276. Brown, D. G., K. M. John.son, T. R. Loveland, and D. M. Tmeobald. 2005. Rural land-use trends in the conterminous United States. 1950-2000. Ecological Applications 15: 1851-1863. Collins, S. L. 2001. Long-term re.search and the dynamics of bird populations and communities. Auk I 18; 583- 588. Enemar, a., B. Cavai.lin, E. Nyholm, I. Rudebeck, and A. M. Thorner. 1994. Dynamics of a passerine bird community in a small deciduous wood, S. Sweden, during 40 years. Ornis Svecica 4; 65-104. Francis, C. M., J. Bart, E. H. Dunn, K. P. Burnham, and C. J. Ralph. 2005. Enhancing the value of the Breeding Bird Survey: reply to Sauer et al. (2005). Journal of Wildlife Management 69: 1327-1332. Gough, G. A., J. R. Sauer, and M. Iliff. 1998. Patuxent bird identification infocenter. Version 97.1. USGS, Patuxent Wildlife Research Center, Laurel, Maryland, USA. www.mbr-pwrc.usgs.gov/id/framlst/infocenter. html (accessed 1 March 2009). Hagan, J. M. 1993. Decline of the Rufous-sided Towhee in the eastern United States. Auk 1 10: 863-874. Hagan, J. M., P. S. McKinley, A. L. Meehan, and S. L. Grove. 1997. Diversity and abundance of landbirds in a northeastern industrial forest. Journal of Wildlife Management 61: 718-735. Hall, G. A. 1984. Population decline of neotropical migrants in an Appalachian forest. American Birds 38: 14-18. Haney, A., S. Apfelbaum, and J. M. Burris. 2008. Thirty years of post-fire succession in a southern boreal forest bird community. American Midland Naturalist 159: 421-433. Hobson, K. A. and J. Schieck. 1999. Changes in bird communities in boreal mixedwood forest; harvest and wildfire effects over 30 years. Ecological Applications 9: 849-863. Holmes, R. T. and T. W. Sherry. 1988. A.ssessing population trends of New Hampshire forest birds: local vs. regional patterns. Auk 105: 756-768. Holmes, R. T. and T. W. Sherry. 2001. Thirty-year bird population trends in an unfragmented temperate deciduous forest; importance of habitat change. Auk 118: 589-609. Hunter, W. C., D. A. Buehler, R. A. Canterbury, J. L. Confer, and P. B. Hamel. 2001. Conservation of disturbance-dependent birds in eastern North America. Wildlife Society Bulletin 29; 440—455. James, F. C. and N. O. Wamer. 1982. Relationships between temperate forest bird communities and vegetation structure. Ecology 63: 159-171. Johnston, D. W. and E. P. Odum. 1956. Breeding bird populations in relation to plant succession on the Piedmont of Georgia. Ecology 37: 50-62. Keller, C. M. E. and J. T. Scallan. 1999. Potential roadside biases due to habitat changes along Breeding Bird Survey routes. Condor 101: 50-57. Kendall, W. L., B. G. Peterjohn, and .1. R. Sauer. 1996. First-time observer effects in the North American Breeding Bird Survey. Auk 1 13; 823-829. Kirk, D. A., A. W. Diamond, A. R. Smith, G. E. Holland, AND P. Chytyk. 1997. Population changes in boreal forest birds in Saskatchewan and Manitoba. Wilson Bulletin 109: 1-27. Leck, C. F., B. G. Murray, and J. Swinebroad. 1988. Long-term changes in the breeding bird populations of a New-Jersey forest. Biological Con.servation 46: 145- 157. McNulty, S. A., S. Droege, and R. D. Masters. 2008. Long-term trends in breeding birds in an old-growth Brooks and Bonter • LONG-TERM COMMUNITY CHANGES 295 Adirondack forest and the surrounding region. Wilson Journal of Ornithology 120: 153-158. Niemuth, N. D., a. L. Dahl, M. E. Estey, and C. R. Loesch. 2007. Representation of landcover along Breeding Bird Survey routes in the northern plains. Journal of Wildlife Management 71: 2258-2265. Norment, C. 2002. On grassland bird conservation in the Northeast. Auk 119: 271-279. Robbins, C. S., D. Bystrak, and P. H. Geissler. 1986. The Breeding Bird Survey: its first fifteen years, 1965- 1979. Resource Publication 157. USDl, Fish and Wildlife Service, Washington, D.C., USA. Robbins, C. S., J. R. Sauer, R. S. Greenberg, and S. Droege. 1989. Population declines in North- American birds that migrate to the Neotropics. Proceedings of the National Academy of Sciences of the USA. 86: 7658-7662. Royle, J. A., J. D. Nichols, and M. Kery. 2005. Modelling occurrence and abundance of species when detection is imperfect. Oikos 1 10:353-359. SAS Institute. 2003. SAS Version 9.1. Cary, North Carolina, USA. Sauer, J. R., J. E. Hines, and J. Fallon. 2008. The North American Breeding Bird Survey, results and analysis I966-2(X)7, Version 5.15. USGS, Wildlife Research Center, Laurel, Maryland, USA. www.mbr-pwrc.usgs. gov/bbs/bbs.html (accessed 10 February 2009). Stephens, S. E., D. N. Koons, J. J. Rotella, and D. W. Willey. 2(K)4. Effects of habitat fragmentation on avian nesting success: a review of the evidence at multiple spatial scales. Biological Conservation 115: 101-110. SvENSSON, S., K. Williamson, C. Ferry, A. H. Joensen, D. Lea, H. Oelke, and C. S. Robbins. 1970. An international standard for a mapping method in bird census work recommended by the International Bird Census Committee. Audubon Field Notes 24: 722-726. Venier, L. A. and j. L. Pearce. 2005. Boreal bird community response to jack pine forest succession. Forest Ecology and Management 217: 19-36. Vickery, P. D., A. L. Hunter, and S. M. Melvin. 1994. Effects of habitat area on the distribution of grassland birds in Maine. Conservation Biology 8: 1087-1097. Wiens, J. A. 1984. The place of long-term studies in ornithology. Auk 101: 202-203. The Wilson Joiinuil of Ornithology 122(2):296-306, 2010 SCALE-DEPENDENT RESPONSE BY BREEDING SONGBIRDS TO RESIDENTIAL DEVELOPMENT ALONG LAKE SUPERIOR MICHELLE T. LORD' AND DAVID J. LLASPOHLER' ABSTRACT. — We examined the influence of shoreline residential development on breeding bird communities along forested portions ot Lake Superior and hypothesized that anthropogenic changes related to housing development would alter bird community structure compared to areas without human development. We used point counts to compare relative abundance of bird species in relation to residential development at coarse (along 1 km shoreline stretches with and without housing/cottage development) and fine (developed and undeveloped sides of a shoreline access road) spatial scales during the 2005 breeding season. More species had development related differences in abundance at the finer-scale analysis than at the coarse .scale. American Crows (Corvits hrachyrhynchos) and American Robins (Titrclus migratorius) were more abundant on the developed, shoreline side of shoreline access roads. Red-breasted Nuthatches (Sitta canadensis). Black- throated Green Warblers (Dendroica virens), and Red-eyed Vireos {Vireo olivaceus) were more abundant on the undeveloped, inland side of shoreline access roads. Several species were detected exclusively in developed or undeveloped forest areas. The pattern of development-related differences in relative abundance of bird species depended on the scale at which data were analyzed, suggesting that many species may respond to habitat differences within the 100 m scale quite distinct from how they respond to differences at the scale of thousands of meters. Received 27 August 2009. Accepted 23 December 2009. Rural population growth in many parts of the United States has increased in the last 15 years (Long and Nucci 1997, Gustafson et al. 2005). Growth in parts of the northern Great Lakes Region has been concentrated around inland lakes (Radeloff et al. 2001, Gonzalez- Abraham et al. 2006). Two-thirds of previously undeveloped inland lakes in northern Wisconsin (i.e., lakes with no residential housing) have become devel- oped with homes and cottages near the shoreline since 1965 (Lindsay et al. 2002, Elias and Meyer 2003). Housing development alters shoreline habitat through changes in habitat structure and plant species composition, which can create discontinuities in plant communities resulting in fragmentation (Brown 2003). These changes have the potential to influence many taxa including forest birds (Vale and Vale 1976, Mills et al. 1989, Theobald et al. 1997, Swenson and Franklin 2000). Many recent studies focusing on the ecological impacts of nearshore residential development have examined inland lakes with less attention on Great Lakes shorelines (Rottenborn 1999, Lindsay et al. 2002, Elias and Meyer 2003). Perimeters of both inland lakes and the Great Lakes provide an interface between aquatic and ' School of Forest Resources and Environmental Science, Michigan Technological University, Houghton, Ml 49931. USA. ^Current address: 73 Training Hill, Middletown. CT 06457. USA. 'Corresponding author; e-mail: michelle.ford@gza.com upland ecosystems but, because of their size and associated climatological influence, the abiotic and biotic environment along Great Lakes shore- lines often differs dramatically from inland lake riparian zones or adjoining upland habitats (Eichenlaub 1979, Albert et al. 1986). Thus, results from studies of inland lakes cannot be confidently extended to larger lake systems. North America’s Great Lakes (Huron, Ontario, Michigan, Erie, and Superior) contain 20% of the world’s and 95% of North America’s surface fresh water. The waters of the Great Lakes support the world’s largest freshwater commercial fishery, supply drinking water to millions of citizens in the United States and Canada, and support a multi- million dollar recreation industry (Great Lakes Information Network 2005). The amenity values associated with the 17,542 km of Great Lakes shoreline have likewise attracted residential development (Schnaiberg et al. 2002). Shoreline residential development on the Great Lakes has accelerated in recent years and now represents one of the fastest growing segments of rural housing expansion in the region (Orr 1997). The Keweenaw Peninsula is in the western Upper Peninsula of Michigan and is surrounded on three sides by Lake Superior (Fig. 1). Kewee- naw County is one of the least populated in Michigan and much of the shoreline remains undeveloped; however, residential development in this area has rapidly increa.sed over the past decade with population shifts to rural areas (Orr 1997). The Upper Peninsula has not yet felt the 296 Ford am! Flaspohler • BREEDING SONGIBRDS AND DEVELOPMENT 297 effects of rural in-migralion as greatly as other scenic places in the United States, but land use changes and population increases are evident (Orr 1997). We focused on differences between residential- ly developed and undeveloped shoreline areas of Lake Superior and measured vegetation habitat features associated with development known to influence bird presence and relative abundance. To our knowledge, this is the first study to examine the influence of Great Lakes shoreline residential development on forest breeding bird communities on any of the Great Lakes. Our objectives were to; ( 1 ) evaluate differences in forest vegetation between developed and unde- veloped areas, (2) examine if relative abundance of breeding bird species differs between residen- tially developed and undeveloped shoreline forest along Lake Superior, and (3) assess the spatial scale at which development influences bird presence and abundance. METHODS Study Sites. — The study was conducted in two residentially developed and three undeveloped shoreline areas along the eastern shore of Lake Superior in the Keweenaw Peninsula in the northern-most part of the Michigan’s Upper Peninsula (Fig. 1). Only east and southeast facing areas were used to minimize variability in soils, bedrock geology, site aspect, wind, and forest type. Developed areas included Hermit’s Cove (47° 15' N, 88° 07' W) and Rabbit Bay (47° 04' N, 88° 20' W) in Keweenaw and Houghton counties, respectively. These areas are similarly developed and are comprised predominately of seasonal cottages which are bounded on one side by Lake Superior and on the other by closed-canopy mixed hardwood and boreal transition forest with a gravel road along the forest side of the properties. Cottages at both sites are only on the lake side of the road and have a mixture of old (built prior to 1950) and newer homes, as well as seasonal and permanent residents. Most cottages are occupied primarily during the summer months (Manarolla 2005). Low-elevation and near-shore portions of the Keweenaw Peninsula are dominated by boreal transition forests composed primarily of paper birch (Betiila papyrifera), balsam fir (Abies halsamea), red maple (Acer riihrum), white spruce (Picea glaiica), northern white cedar (Thuja occidentalis), and yellow birch (Betula alleglia- nieusis). Average precipitation for Keweenaw County during the breeding sea, son from 1 May 2005 through 31 July 2005 was ~22.8 cm with a mean temperature of 14° C (Weather Under- ground Corporation 2005). The soils of the eastern shore of the Keweenaw Peninsula range from gravelly and rocky sands to sandy loams and are poorly drained in lower elevations near the shore (Albert et al. 1986). Undeveloped study areas. Smith Fisheries (47° 23' N, 87° 53' W), south Rabbit Bay (47° 02' N, 88° 21' W), and north Hermit’s Cove (47° 15' N, 88° 06' W) were .selected by visiting areas along the eastern shore in early spring and by using Geographic Information System (CIS) data and aerial photographs to identify areas with and without concentrated residential development near the shore. Undeveloped study areas were along sections of shoreline north of Hermit’s Cove and along sections of shoreline south of Rabbit Bay, and were placed ^300 m from the nearest residential stmcture. The Smith Fisheries study area was —48 km north of Hermit’s Cove, just north of the town of Lac La Belle, and was only accessible via logging roads and trails (Fig. 1). Thirty-six plots were established for vegetation and bird community sampling, 15 in undeveloped study areas and 21 in developed study areas. We collected data at two spatial scales: (1) broad-scale; developed and undevel- oped shoreline study areas, and (2) fine-scale: developed and undeveloped sides of the road accessing developed shoreline study areas. Measurement of Habitat Variables. — We mea- sured habitat characteristics for the broad scale comparison along transects spaced every 100 m in both developed and undeveloped shoreline study areas. Vegetation transect centers along undevel- oped shoreline study areas were 50 m inland. Transects along developed shoreline study areas were centered on the center of the unpaved road (50-75 m from the shore). We considered roads to be part of tbe vegetation disturbance associated with development. A second set of transects was oriented perpendicular to the shore, and extended 50 m in each direction toward and away from the shoreline for both developed and undeveloped shoreline study areas. This method sampled vegetation in both developed residential property and the undeveloped forest across the road from the development along the developed shoreline. Point-count centers for bird surveys were also placed in the center of the road so that equal 298 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2, June 2010 O' SmMti Fisheries Study Area FIG. 1. Study .sites in Houghton and Keweenaw countie.s, Michigan, USA. Ford and Flaspoldcr • BREEDING SONGBIRDS AND DEVELOPMENT 299 portions of the point-count area were in devel- oped/shoreline and undeveloped/inland sides of the road. Vegetation sampling methods were adapted from Noon (1981). Canopy cover and ground cover were estimated using an ocular tube approximately every 2 m along each transect. Coarse woody debris (CWD) was defined as any piece of downed wood >8 cm in diameter that crossed a transect. The total pieces of CWD that crossed a transect were recorded and divided by the length of the transect (100 m). Diameter at breast height (DBH) and species were recorded for all living trees >3 cm in diameter within arms length, —0.8 m, on either side of the transect. DBH data were converted to total stem count to compare the number of stems >3 cm in each transect. We also calculated deciduous and coniferous shrub/sapling density along each transect in developed and undeveloped areas. Each woody stem ^3 cm in diameter at breast height was recorded as either coniferous or deciduous. Average density for each shrub type was calculated as the total number of shrubs counted divided by the length of the transect. Total shrub density was calculated as the sum of both deciduous and coniferous shrubs divided by transect length. A gridded density board was used to measure understory density and calculate percent cover from 0 to 3 m above the forest floor (Noon 1981). Separate density readings were taken facing areas on each side of the road (i.e., facing the residential development along the shoreline and facing the undeveloped inland) by standing on the side of the road with the density board placed 1 1.3 m from the road edge. Readings were summed and converted to percent cover in four height catego- ries, 0-0.3, >0.3-1, >1-2, and >2 m, and total understory density 0-3 m. Bird Surveys. — We used 5 min 50-m fixed- radius point counts to estimate breeding bird relative abundance (Bibby et al. 2000, Sutherland et al. 2004) at developed and undeveloped sites. Counts started at sunrise, —0600 hrs, and ended by 1030 hrs between 6 June and 1 July 2005. Point-count centers were established every 200 m to minimize double counting individuals (Bibby et al. 2000). Point-count centers in developed shoreline study areas were in the center of the road at the center of vegetation transects, 50-75 m from the shoreline. Point-count centers within undeveloped shoreline study areas were 50 m from the shore to minimize wave noise interfer- ence. Twenty-one point-count stations were established in developed shoreline study areas and 15 in undeveloped shoreline study areas; all counts were done by M. T. Ford. The observer waited for 2 min prior to start of each point-count period (Reynolds et al. 1980). All birds detected by sight or sound within the 50- m radius within 5 min were recorded. Birds detected outside the 50-m radius or seen flying over were also noted, but were not included in analyses. All stations were visited two times in 2005, once early and once later in the breeding season with one of the counts conducted within 30 min of sunrise and the other later in the morning. Point counts were not conducted in rain or high winds (> 20 km/hr) (Robbins 1981). Bird species richness in developed and unde- veloped areas was estimated from the total number of species detected in 2005. Relative abundance for each species was defined as the maximum number of individuals detected be- tween the two visits at each station. Birds were also grouped into nesting/foraging guilds to investigate guild-level associations with develop- ment. Statistical Analysis. — ^We used t- and Wil- coxon-Mann-Whitney non-parametric tests to examine differences in bird abundance and vegetation attributes between developed and undeveloped areas. The more powerful /-test was used when substantial deviation from sym- metry was not detected in boxplots generated from the raw data (Quinn and Keough 2002). The Wilcoxon-Mann-Whitney test was used at the same significance levels to examine nesting/ foraging guild differences between developed and undeveloped areas. Understory density from 0 to 0.3 m and 1 to 2 m appeared to be asymmetrical and was analyzed using Wilcoxon- Mann-Whitney tests. The Shannon-Wiener spe- cies diversity index (//’) was used to compare species richness and evenness between developed and undeveloped areas (Magurran 2004). All data were tested at a significance level of a = 0.05 unless otherwi.se specified. RESULTS Broad Spacial Scale Comparison of Habitat Characteristics in Developed and Undeveloped Shoreline Study Areas. — ^Developed shoreline had higher deciduous and total shrub density than undeveloped areas. Canopy cover and CWD were 300 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122. No. 2. June 2010 TABLE 1. Vegetation characteristics Peninsula, Michigan. in developed and undeveloped shoreline study areas on the Keweenaw Characteristics Treatment Mean ± SD p Canopy cover' Developed 76.91 ± 22.33 <0.000 Undeveloped 87.82 ± 12.08 Ground cover Developed 64.68 ± 16.42 0.312 Undeveloped 61.56 ± 19.16 Coarse woody debris" Developed 8.38 ± 5.75 <0.000 Undeveloped 16.6 ± 10.16 Number of stems > 8 cm" Developed 357 ± 205.96 <0.000 Undeveloped 288 ± 166.13 Total shrub density" Developed 6.3 ± 10.62 0.016 Undeveloped 4.38 ± 9.24 Deciduous shrub density" Developed 9.38 ± 8.25 <0.000 Undeveloped 4.35 ±6.18 Coniferous shrub density Developed 3.23 ±4.16 0.194 Undeveloped 4.42 ± 6.05 Total understory density (UD) Developed 50.03 ± 28.25 0.87 Undeveloped 50.9 ± 33.21 UD 0-0.3 m Developed 78.19 ± 36.31 0.483 Undeveloped 74 ± 34.17 UD > 0.3-1. 0 m Developed 51.85 ± 37.2 0.473 Undeveloped 47.04 ± 40.85 UD > 1. 0-2.0 m Developed 27.83 ± 34.07 0.203 Undeveloped 35.63 ± 37.13 UD > 2.0-3.0 m Developed 42.24 ± 36.7 0.462 Undeveloped 46.93 ± 37.88 Total diameter at brea.st height (DBH) Developed 17.77 ± 15.87 0.498 Undeveloped 19.51 ± 59.66 DBH of deciduous trees Developed 16.65 ± 15.18 0.474 Undeveloped 21.41 ± 96.54 DBH of coniferous trees Developed 18.92 ± 16.47 0.635 Undeveloped 39.4 ± 13.52 ■' Significant (a = 0.05). greater in undeveloped than developed areas. Developed and undeveloped shoreline study areas did not differ in mean DBH for all trees together, deciduous trees, or coniferous trees. Stem counts were greater in developed than undeveloped shoreline study areas (Table 1). Small Spatial Scale Comparison of Developed Shoreline and Undeveloped Inland Sides of Shoreline Road. — ^The residentially developed shoreline side of the road had lower canopy cover, ground cover, CWD, stem counts, total understory density, and understory density from 0-0.3 to >0.3-1 m (Table 2). Mean DBH of coniferous trees, deciduous trees, and all trees pooled did not differ between developed and undeveloped sides of the road (Table 2). Birds. — We detected 291 individuals of 32 species for all point counts pooled within the 50-m radius point counts during the survey period (Table 3). Species richness was —34% greater in developed (29 species) shoreline study areas than undeveloped (18 species) shoreline study areas. Ninety percent (29 of 32 species) of all species detected were in developed shoreline study areas; only 56% (18 of 32 species) of all species detected were in undeveloped shoreline study areas. Fourteen species were found only in developed shoreline study areas and three species were only in undeveloped shoreline study areas (Table 3). The Shannon-Wiener diversity index (//') was 2.96 in developed and 2.45 in undevel- oped shoreline study areas. Species evenness was virtually identical in developed {E = 0.88) and undeveloped (E = 0.86) shoreline study areas. The Shannon-Wiener diversity index within developed shoreline study areas was 2.97 on the developed side of the road and 2.70 on the inland, undeveloped side of the road. Species evenness was similar between shoreline (E = 0.92) and inland (E = 0.87) sides of the road. FonI ciihI Flcispohkr • BREEDING SONGBIRDS AND DEVELOPMENT 301 TABLE 2. Vegetation characteristics in developed shoreline and undeveloped inland sides ol shoreline residentially developed areas on the Keweenaw Peninsula, Michigan. access roads in Characierislics Trealnienl Mean ± .SD r Canopy cover' Shoreline 69.49 ± 23.31 0.002 Inland 84.34 ± 18.79 Ground cover' Shoreline 68.16 ± 18.9 0.005 Inland 69.49 ± 12.81 Coarse woody debris" Shoreline 6.71 ± 6.16 0.008 Inland 10.05 ± 4.83 Number of stems > 8 cm" Shoreline 138 ± 79.53 <0.001 Inland 219.5 ± 126.5 Total shrub density Shoreline 5.51 ± 6.99 0.16 Inland 7.1 ± 6.66 Deciduous shrub density Shoreline 7.71 ± 8.07 0.066 Inland 1 1.05 ± 8.19 Coniferous shrub density Shoreline 3.32 ± 4.79 0.854 Inland 3.15 ± 3.48 Total understory density" Shoreline 43.86 ± 39.86 0.047 Inland 56.2 ± 40.33 UD 0-0.3 m" Shoreline 67.88 ± 40.81 0.01 Inland 88.5 ± 28.06 UD > 0.3- 1.0 m" Shoreline 41.11 ± 35.9 0.008 Inland 62.59 ± 35.73 UD > 1. 0-2.0 m Shoreline 25.32 ± 32.19 0.508 Inland 30.34 ± 36.08 UD > 2.0-3.0 m Shoreline 41.12 ± 35.07 0.784 Inland 43.37 ± 38.67 Total diameter at breast height (DBH) Shoreline 18.75 ± 16.21 0.194 Inland 17.15 ± 15.64 DBH of deciduous trees Shoreline 18.13 ± 14.13 0.21 Inland 15.97 ± 15.63 DBH of coniferous trees Shoreline 19.2 ± 17.56 0.77 Inland 18.68 ± 15.55 ^ Significant (y — 0.05). Broad Spatial Scale Comparison of Bird Species Relative Abundance. — Only Red-eyed Vireos (Vireo olivaceus), of 32 total species detected, were significantly (5 times) more abundant in developed than undeveloped shore- line study areas. Ground-nesting birds were equally abundant in developed and undeveloped shoreline study areas (P > 0.05). Shrub nesters were more abundant in undeveloped forest and canopy nesters were more abundant in developed shoreline study areas when species were pooled based on nest site location (P < 0.05) (Table 3). Fine-scale Comparison of Bird Species Relative Abundance. — ^Two species, American Crow (Cor- vus brachyrhynchos) and American Robin (Tur- dus migratorius) were statistically more abundant on the developed, shoreline side of the shoreline access roads. Three species, Black-throated Green Warbler (Dendroica virens), Red-eyed Vireo, and Red-breasted Nuthatch (Sitta canadensis) were more abundant on the undeveloped, inland side of shoreline access roads (Table 4). Differences in relative abundance of American Redstarts (Seto- phaga ruticilla) and Blackburnian Warblers {Den- droica fusca) approached significance {P = 0.10) with both detected more frequently on the undeveloped, inland side of point counts (Ta- ble 4). DISCUSSION Analyses of developed and undeveloped shore- line areas at the broad scale showed that only one species was significantly more abundant in developed shoreline areas and no species were significantly more abundant in undeveloped shoreline areas. Fine-scale plot-level analyses in developed areas suggested that two species were significantly more abundant on the developed. 302 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2. June 2UI0 TABLE 3. Bird species detected on point counts between May and July 2005 areas on the Keweenaw Peninsula, Michigan. in developed and undeveloped study Common name Species AOU Code Number of individual.s detected Developed Undeveloped Nesting guild Ground Shrub Canopy American Crow Corviis brachyrhynchos AMCR 5 0 X American Goldfinch Spinus tristis AMGO 1 0 X American Redstart Setophaga niticilla AMRE 14 10 X American Robin Tardus migratorius AMRO 9 0 X Black-and-white Warbler Mniotilta varia BAWW 7 3 X Black-capped Chickadee Poecile alricapillus BCCH 6 0 X Blackburnian Warbler Dendroica fusca BLBW 1 0 X Blue Jay Cyanocitta cristata BLJA 2 1 X Black-throated Green Warbler Dendroica virens BTNW 7 3 X Chipping Sparrow Spizella passerina CHSP 3 0 X Cape May Warbler Dendroica tigrina CMWA 0 1 X Common Raven Corvus corax CORA 0 2 X Chestnut-sided Warbler Dendroica pensylvanica CSWA 1 0 X Golden-crowned Kinglet Regains satrapa GCKl 6 1 X Hairy Woodpecker Picoides villosas HAWO 2 0 X Hermit Thrush Catharus guttafas HETH 3 10 X Mallard Anas platyrhynchos MALL 2 0 X Magnolia Warbler Dendroica magnolia MAWA 1 1 X Yellow-rumped Warbler D. coronata MYWA 18 4 X Nashville Warbler Vermivora ruficapilla NAWA 23 24 X Northern Parula Parula americana NOPA 24 12 X Ovenbird Seiurus aurocapilla OVEN 9 0 X Pine Warbler Dendroica pinus PIWA 3 0 X Red-breasted Nuthatch Sitta canadensis RBNU 5 3 X Ruby-crowned Kinglet Regulus calendula RCKI 0 1 X Red-eyed Vireo Vireo olivaceus REVI 14 3 X Ruby-throated Hummingbird Archilochus coluhris RTHU 2 0 X Song Sparrow Melospiza melodia SOSP 1 0 X Swainson’s Thrush Catharus ustidatus SWTH 3 8 X White-throated Sparrow Zonotrichia alhicoUis WTSP 7 0 X White-breasted Nuthatch Sitta carolinensis WBNU 1 0 X Winter Wren Troglodytes troglodytes WIWR 6 13 X shoreline portion rather than the undeveloped, inland portion of the shoreline access roads. Three species were more abundant on the inland, undeveloped side of the shoreline access roads. Our results suggest species may be responding to habitat alterations associated with residential development at the patch, rather than the larger stand/landscape scale. We found development- related differences in natural habitat features including canopy volume, shrub cover, forest lloor vegetative cover, and amount of CWD. Riparian areas in much of the U.S. represent concentrations of biodiversity and foci for resi- dential development in rural areas (Naiman et al. 1993, Schnaiberg et al. 2002, Bub et al. 2004, Gonzalez-Abraham et al. 2006). Shoreline vege- tation, compared to adjoining inland sites, is typically denser, more varied in species compo- sition, and offers greater vertical and horizontal structural diversity (Riffell et al. 2001). Rotten- born (1999) proposed that shoreline forests have unique vegetation characteristics compared to similar forests distant from the shoreline and these differences may shape patterns of bird community composition. Research has shown that riparian vegetation structure influences breeding bird communities (Willson 1974, Hostetler and Holling 2()()()) and altering these habitats through removal of native vegetation can change bird community structure. Vegetation structure fre- Font and Flaspohler • BREEDING SONGBIRDS AND DEVEEOBMENT 303 TABLE 4. Birds detected on undeveloped inland and developed shoreline sides of shoreline access roads in residentially developed areas on the Keweenaw Peninsula, Michigan. Total Species Treatment count Mean/ 10 ploCs ± .SD Zcalc“ AMCR" Inland 0 0 ± 0 2.03 Shoreline 5 23.8 ± 0.54 AMGO Inland 0 0 ± 0 0.97 Shoreline 1 4.8 ± 0.22 AMRE Inland 17 33.3 ± 0.66 1.91 Shoreline 7 28.6 ± 0.56 AMRO Inland 1 0 ± 0 1.97 Shoreline 8 38.1 ± 0.74 BCCH Inland 1 4.8 ± 0.22 1.37 Shoreline 4 19 ± 0.4 BLEW Inland 2 9.5 ± 0.3 1.74 Shoreline 0 0 ± 0 BLJA Inland 1 4.8 ± 0.22 1.01 Shoreline 2 9.5 ± 0.3 BTNW“ Inland 6 28.6 ± 0.56 2.04 Shoreline 2 9.5 ± 0.3 BAWW Inland 3 14.3 ± 0.36 0.04 Shoreline 5 23.8 ± 0.44 CHSP Inland 1 4.8 ± 0.22 0.57 Shoreline 2 9.5 ± 0.3 CSWA Inland 0 0 ± 0 0.97 Shoreline 1 4.8 ± 0.22 CORA Inland 0 0 ± 0 1.4 Shoreline 0 5.6 ± 0.23 GCKI Inland 4 19 ± 0.4 1.17 Shoreline 2 9.5 ± 0.3 HAWO Inland 1 4.8 ± 0.22 -0.02 Shoreline 1 4.8 ± 0.22 HETH Inland 0 0 ± 0 0.2 Shoreline 2 9.5 ± 0.3 MALL Inland 0 0 ± 0 0.97 Shoreline 2 9.5 ± 0.44 MAWA Inland 1 4.8 ± 0.22 1.4 Shoreline 0 0 ± 0 NAWA" Inland 14 70 ± 0.66 4.2 Shoreline 8 38.1 ± 0.59 NOPA^ Inland 21 100 ± 0.95 4.02 Shoreline 4 19 ± 0.4 OVEN Inland 6 28.6 ± 0.64 0.45 Shoreline 3 14.3 ± 0.36 PIWA Inland 1 4.8 ± 0.22 -0.02 Shoreline 1 4.8 ± 0.22 RBNU'* Inland 4 19 ± 0.4 2.22 Shoreline 1 4.8 ± 0.22 REVP Inland 10 47.6 ± 0,51 2.47 Shoreline 4 19 ± 0.40 RCKI Inland 1 2.8 ± 0.17 0.97 Shoreline 0 0 ± 0 RTHU Inland 2 9.5 ± 0.30 1.40 Shoreline 0 0 ± 0 SOSP Inland 0 0 ± 0 1.4 Shoreline 2 9.5 ± 0.3 TABLE 4, Continued. Total Species Treatment count Mean/IO plots ± S!) 7x-alc* SWTH Inland 2 9.5 -L 0.3 0.7 Shoreline 1 4.8 -+■ 0.22 WBNU Inland 1 4.8 0.22 0.97 Shoreline 0 0 0 WIWR^ Inland 5 24 0.44 3.58 Shoreline 1 5 -h 0.22 WTSP Inland 3 14 -t- 0.36 0.73 Shoreline 5 24 ~h 0.44 MYWA Inland 10 48 -h 0.68 0.71 Shoreline 9 43 -h 0.51 Mann Whitney Zerit = 1 .96 al a = 0.05 and Zerit — 1 .645 at ot — 0. 1 0. Ho is rejected when Zcalc > Zerit. quently influences bird species richness, relative abundance (Willson 1974, Hostetler and Holling 2000), and foraging behavior (Smith et al. 1998). We found development-related differences in natural habitat features including canopy volume, shrub cover, forest floor vegetative cover, and amount of CWD. Research has shown that natural habitat features such as canopy volume, shrub cover, and forest vegetative cover are influenced by residential development (Clark et al. 1984, Christensen et al. 1996, Woodford and Meyer 2003). Many studies (Willson 1974, Burke and Nol 1998, Sallabanks et al. 2000) have suggested that changes in the avian community in the presence of human development are the result of development-related changes in vegetation rather than direct human disturbance of birds. Elias and Meyer (2003) found that undeveloped shoreline in northern Wisconsin had higher percent canopy, subcanopy, understory cover, coarse woody debris, and percentage of shoreline overhung by trees and sbrubs compared to developed shoreline stretches. They also found that plant species richness and diversity were greater in developed than in undeveloped shore- line areas. Lindsay et al. (2002) compared bird community attributes in riparian areas along developed and undeveloped inland lakes in northern Wisconsin. They reported that granivo- rous and omnivorous bird species were more abundant along developed lakes than undeveloped lakes. In contrast, insectivorous species were more abundant along undeveloped than developed lakeshores (Lindsay et al. 2002). Rottenborn (1999) reported that shoreline development near riparian areas in California affected nearby riparian bird communities and concluded that 304 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 2, June 2010 species richness and diversity decreased as native vegetation was lost to residential development. Differences in vegetation characteristics be- tween developed and undeveloped shoreline areas may be responsible for most of the differences we observed in bird abundance. Deciduous shrub density and total shrub density were greater along developed shoreline, while mean canopy cover and mean number of CWD pieces were greater along undeveloped shoreline. The greater number of all but coniferous shrubs and saplings within developed shoreline areas is likely the result of greater light penetration from the more open forest canopy. Comparison of forest characteris- tics at the individual plot scale suggested that within-plot canopy cover, CWD, understory density from 0 to 0.3 m, >0.3 to 1 m, and overall understory density were greater on the undevel- oped side of shoreline access roads. The presence of development features such as lawns, houses, outbuildings, walkways and driveways may con- tribute to these differences. Christiansen et al. (1996) found that inland lake shoreline areas in northern Wisconsin with more cabins had fewer large pieces of dead and down wood in near shore waters, and speculated that many property owners remove snags and downed wood in or near the water’s edge. The same process may occur around cottages, decreasing the amount of CWD in these areas. All study areas have thin soils which, given the close proximity to winds off Lake Superior, may result in increased blow-downs in developed shoreline areas, height- ening the contrast in CWD abundance between developed and undeveloped shorelines. The higher abundance of CWD, canopy cover, understory density from 0 to 0.3 m and from >0.3 to 1.0 m available in the undeveloped inland side of shoreline access road census locations likely influenced the relative abundance of several species. Nashville Warblers (Vemiivora rufica- pilla) and Winter Wrens {Troglodytes troglodytes) were both more abundant in the undeveloped inland portion of shoreline access roads. Nashville Warblers nest at the ground level and Winter Wrens often place their nests in the roots of upturned trees or snags (Ehrlich et. al. 1988). The higher abundance of CWD and increased under- story density on the undeveloped, inland side of the road may be important habitat features for these two species. Northern Parula (Panda americana) have been shown to prefer riparian areas (Moldenhauer and Regelski 1996) and to be relatively tolerant of habitat disturbance short of clear cutting (Brooks 1940). Northern Parula, which build nests in the mid-canopy, have been shown to be positively correlated with at least 15% canopy cover and, when breeding, are heavily dependent on an abundance of epiphytes, particularly lichen (Us- nea spp.) in their northern range (Collins et al. 1982, Moldenhauer and Regelski 1996). Northern Parula were more abundant on the inland, undeveloped side of shoreline access roads than on the residentially developed side. This may be the result of lower abundance of epiphytic lichen related to canopy openings (Esseen and Rehhorn 1998) and/or preference by Northern Parula for increased canopy cover, which was greater on the inland side of point-count plots. Red-breasted Nuthatches depend on the avail- ability of cavities in dead standing trees or roots of upturned trees as well as sap from living conifers. The sap is used to coat the entrance of the cavity during the incubation period to protect against nest predators (Ghalambor and Martin 1999). The increased abundance of CWD in inland areas may be an indicator of dead, dying trees in the vicinity and perhaps of standing dead, cavity bearing trees making it an increasingly suitable habitat for Red- breasted Nuthatches. Given that our data were collected over a single year, we realize there are several limita- tions associated with our conclusions. Annual variations in weather, vegetation characteristics associated with development (i.e., possible col- lection of CWD by landowners for firewood or annual maintenance of vegetation within resi- dential properties), and local bird community all likely contribute to bird species response within our study areas. Another recognized limitation is the potential for birds to naturally nest >50 m from the shoreline, regardless of the presence of residential development. Species detection is another factor which may be identified as a limitation of our study; it is likely that differ- ences exist in detection of bird species between the more open, less vegetated residential shore- line areas and undeveloped areas with overall greater sub-canopy density. We believe the differences found in bird species abundance between residentially developed and undevel- oped shoreline areas are legitimate based upon the known species habitat preferences and the measured vegetation characteristics of each study area. Ford and Flaspo/ilcr • BREEDING SONGBIRDS AND DEVELOPMENT 305 It is essential to clarify the scale at which bird species respond to changes related to shoreline development to understand how development- related habitat alteration interacts with breeding bird habitat use. Our results suggest birds may respond more to development-related habitat changes at the scale of tens of meters, than at the scale ot the shoreline study areas we measured (i.e., hundreds of meters). Ornithologists and ecologists have long recog- nized the role that scale has in understanding habitat use in wildlife (Wiens 1989, Levin 1992). The differing patterns of species composition and relative abundance we found suggest development related changes did not elicit a strong response by most bird species at the scale of the 1 km shoreline. However, development related changes likely influence which areas (e.g., developed or undeveloped side of the road) within a shoreline area a bird chooses for breeding. Our study was conducted in a landscape matrix that is >95% forested, and any forest opening related to residential development represents a relatively small landscape disturbance. Landscape condition has also been shown to be a strong predictor of bird species presence (Saab 1999, Lee et al. 2002). Our results should be viewed in the context of the largely unfragmented suiTOunding forest. ACKNOWLEDGMENTS This research was supported by the USDA Mclntire- Stennis Program and the Copper Country Audubon Society. Robert Froese, Alec Lindsay, Hugh Gorman, and Tom Drummer were helpful in revising earlier drafts of this manuscript. Kurt Doran and Kevin Ford assisted with vegetation data collection. LITERATURE CITED Albert, D. A., S. R. Denton, and B. V. Barnes. 1986. Regional landscape ecosystems of Michigan. School of Natural Resources, University of Michigan, Ann Arbor, USA. Bibby, C. J., N. D. Burgess. D. A. Hill, and S. H. Mustoe. 2000. Bird census techniques. Academic Press, London, United Kingdom. Brooks, M. 1940. The breeding warblers of the central Allegheny Mountain Region. Wilson Bulletin 52:249- 266. Brown, D. G. 2003. Land use and forest cover on private parcels in the upper Midwest USA. 1970 to 1990. Landscape Ecology 18:777-790. Bub. B. R., D. J. Flaspohler. and C. J. Huckins. 2004. Riparian and upland breeding-bird assemblages along headwater streams in Michigan's Upper Peninsula. Journal of Wildlife Management 68:383-392. Burke, D. A. and E. Noi.. 1998. Intluencc of food abundance, nest-site habitat, and forest fragmentation on breeding Ovenbirds. Auk I 15:96-104. Chri,stensen, D. L., B. R. Herwio, D. E. Schindi.er, and S. R. Carpenter. 1996. Impacts of lakeshore residential development on coarse woody debris in north temperate lakes. Ecological Applications 6:1 143-1 149, Clark, K. L., D. L. Euler, and E. Arm.strong. 1984. Predicting avian community response to lakeshore cottage development. Journal of Wildlife Management 48:1239-1247. Collins, S. L., F. C. James, and P. G, Risser. 1982. Habitat relation.ships of wood warblers (Parulidae) in northern central Minnesota. Oikos 39:50-58. Ehrlich, P. R., D. S. Dobkin, and D. Wheye. 1988. The birders handbook: a field guide to the natural history of North American birds. Simon and Schuster Inc., New York, USA. Eichenlaub, V. 1979. Weather and climate of the Great Lakes Region. Notre Dame Press, South Bend, Indiana, USA. Elias, J. E. and M. W. Meyer. 2003. Comparisons of undeveloped and developed shorelands, northern Wisconsin, and recommendations for restoration. Wetlands 23:800-816. Esseen, P. and K. Renhorn. 1998. Edge effects on an epiphytic lichen in fragmented forests. Conservation Biology 12:1307-1317. Ghalambor, C. K. and T. E. Martin. 1999. Red-breasted Nuthatch (Sitta canadensis). The birds of North America. Number 459. Gonzalez- Abraham, C., V. C. Radeloff, R. B. Hammer. T. J. Hawbaker. S. 1. Stewart, and M. K. Clayton. 2006. Building patterns and landscape fragmentation in northern Wisconsin, USA. Landscape Ecology 22: 217-230. Great Lakes Information Network. 2005. Great Lakes Information Network, 12 October 2005. Ann Arbor, Michigan. USA. Gustafson, E. J., R. B. Hammer, V. C. Radeloff. and R. S. Potts. 2005. The relationship between environ- mental amenities and changing human settlement patterns between 1980 and 2000 in the midwestern USA. Landscape Ecology 20:773-789. Hostetler, M. and C. S. Holling. 2000. Detecting the scales at which birds respond to structure in urban landscapes. Urban Ecosystems 4:25-54. Lee, M., L. Fahrig, K. Freemark, and D. J. Currie. 2002. Importance of patch scale vs. landscape scale on selected forest birds. Oikos 96:1 10-1 18. Levin, S. A. 1992. The problem of pattern and scale in ecology: the Robert H. MacArthur Award Lecture. Ecology 73:194.3-1967. Lindsay, A. R.. S. S. Gillum, and M. W. Meyer. 2002. Intluence of lakeshore development on breeding bird communities in a mixed northern forest. Biological Conservation 107:1-11. Long, L. and A. Nucci. 1997. The 'clean break' revisited: is U.S. population again deconcentrating? Environ- mental Planning 29:1355-1366. 306 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 2. June 2010 Magurran, a. E. 2004. Measuring biological diversity. Blackwell Publishing, Oxford, United Kingdom. Manarolla, M. T. 2005. Breeding bird community structure in developed and undeveloped areas along the Lake Superior shoreline. Thesis. Michigan Tech- nological University, Houghton, USA. Mills, S. G., J. B. Dunning, and J. M. Bates. 1989. Effects of urbanization on breeding bird community structure in southwestern desert habitats. Condor 91: 416-428. Moldenhauer, R. R. and D. J. Regelski. 1996. Northern Parula (Panila americana). The birds of North America. Number 215. Naiman, R. j., H. Decamps, and M. Pollock. 1993. The role of riparian corridors in maintaining regional biodiversity. Ecological Applications 3:209-212. Noon, B. R. 1981. Techniques for sampling avian habitats. USDA, Forest Service, Technical Report RM-87. Rocky Mountain Forest and Range Experiment Station, Fort Collins, Colorado, USA. Orr, B. 1997. Land use change on Michigan’s Lake Superior shoreline: integrating land tenure and land cover type data. Journal for Great Lakes Research 23:328-338. Quinn, G. P. and M. J. Keough. 2002. Experimental design and data analysis for biologists. Cambridge University Press, Cambridge, United Kingdom. Radeloff, V. C., R. B. Hammer, P. R. Voss, A. E. Hagen, D. R. Field, and D. J. Mladenoff. 2001. Human demographic trends and landscape level forest man- agement in the northwest Wi.sconsin pine barrens. Forest Science 47:229-241. Reynolds, R. T., J. M. Scott, and R. A. Nussbaum. 1980. A variable circular-plot method for estimating bird numbers. Condor 82:309-313. Riffell, S. K., B. E. Keas, and T. M. Burton. 2001. Area and habitat relationships of birds in Great Lakes coastal wet meadows. Wetlands 21:492-507. Robbins, C. S. 1981. Bird activity levels related to weather. Studies in Avian Biology 6:301-310. Rottenborn, S. C. 1999. Predicting the impacts of urbanization on riparian bird communities. Biological Conservation 88:289-299. Saab, V. 1999. Importance of spatial scale to habitat use by breeding birds in riparian forests: a hierarchical analysis. Ecological Applications 9:135-151. Sallabanks, R., j. R. Walters, and J. A. Collazo. 2000. Breeding bird abundance in bottomland hardwood forests: habitat, edge, and patch size effects. Condor 102:748-758. SCHNAIBERG, J., J. RiERA, M. J. TURNER, AND P. R. VOSS. 2002. Explaining human settlement patterns in a recreational lake district: Vilas County, Wisconsin, USA. Environmental Management 30:24—34. Smith, R., M. J. Hamas, and M. E. Dallman. 1998. Spatial variation in foraging of the Black-throated Green Warbler along the shoreline of northern Lake Huron. Condor 100:474-484. Sutherland, W. J., I. Newton, and R. E. Green. 2004. Bird ecology and conservation. Oxford University Press Inc. New York, USA. Swenson, J. J. and J. Franklin. 2000. The effects of future urban development on habitat fragmentation in the Santa Monica Mountains. Landscape Ecology 15:713- 730. Theobald, D. M., J. R. Miller, and N. T. Hobbs. 1997. Estimating the cumulative effects of development on wildlife habitat. Landscape and Urban Planning 39:25-36. Vale, T. R. and G. R. Vale. 1976. Suburban bird populations in west-central California. Journal of Biogeography 3:157-165. Weather Underground Corporation. 2005. History for Houghton, Michigan. Weather Underground Corporation. I October 2005. Wiens, J. A. 1989. The ecology of bird communities. Cambridge University Press, Cambridge, United Kingdom. Willson, M. F. 1974. Avian community organization and habitat structure. Ecology 55:1017-1029. Woodford, J. E. and M. W. Meyer. 2003. Impact of lakeshore development on green frog abundance. Biological Conservation 110:277-284. The Wilson Joiinuil ofOrnilliolof'v 1 22(2):3()7 -3 1 7, 2010 SONGBIRD NEST SURVIVAL IS INVARIANT TO EARLY-SUCCESSIONAL RESTORATION TREATMENTS IN A LARGE RIVER ELOODPLAIN DIRK E. BURHANS,' " BRIAN G. ROOT,- TERRY L. SHAFFER,^ AND DANIEL C. DEY^ ABSTRACT. — We monitored .songbird nest survival in two reforesting, ~50-tia former cropland sites along the Missouri River in central Missouri from 2001 to 2003. Sites were partitioned into three experimental units, each receiving one of three tree planting treatments. Nest densities varied among restoration treatments for four of five species, but overall nest survival rates did not. Nest survival varied with day-of-year and with incidence of brood parasitism by Brown-headed Cowbirds (Mololhnis ater). Nest survival was higher early and late in the season, and parasitized nests experienced lower nest survival, despite few complete losses directly attributable to parasitism. Probability of parasitism was inversely related to distance to the nearest tree, and was much lower than in old field study sites in the same region. High cowbird parasitism frequencies are usually associated with landscapes low in forest cover, yet these sites in an agriculturally-dominated bottomland landscape experienced low (~0.8-24%) cowbird parasitism. The assumed negative relationship between landscape-level forest cover and cowbird parasitism needs further study in habitats other than forest. Received I September 2008. Accepted 26 November 2009. Studies of nest predation have noted variation in nesting success among habitat types within the same species or suites of species (McCoy et al. 1999, Lloyd and Martin 2005), or variation in nesting success among regions or fragment sizes (Donovan et al. 1995, Robinson et al. 1995). These differences could be attributed to predators, which are known to vary across regions (Thomp- son et al. 1999, Pietz and Granfors 2000). Predator identities may vary across habitat types within the same landscape or region (Chalfoun et al. 2002), although predator differences may not be reflected in differences in nesting success (Thompson and Burhans 2003). Nest predation is considered the largest cause of nest loss (Ricklefs 1969, Martin 1992), and Brown-headed Cowbird (Molothrus ater) brood parasitism, which also affects nest success, may also vary by region, landscape, or habitat type (Robinson et al. 1995, Burhans 1997, Burhans and Thompson 2006). Much of our present under- standing about the interactions of habitat and land.scape features affecting cowbird parasitism comes from studies in upland forested habitats ' Department of Fisheries and Wildlife Sciences, 202 Natural Resources Building, University of Missouri, Columbia, MO 6521 1, USA. ^U.S. Fish and Wildlife Service, 6010 Hidden Valley Road, Suite 101, Carlsbad, CA 9201 1, USA. ■’U.S. Geological Survey, Northern Prairie Wildlife Research Center, 8711 37th Street Southea.st, Jamestown, ND 58401, USA. ■•USDA Fore.st Service, Northern Research Station. 202 ABNR Building, Columbia. MO 65211. USA. ^Corresponding author; e-mail: burhansd@gmail.com (Brittingham and Temple 1983, Robinson et al. 1995, Donovan et al. 2000, Thompson et al. 2000). We monitored songbird nesting success at two sites having three contiguous 16.2-ha habitats of former cropland along the Missouri River. The habitats varied in densities of planted oaks {Quercus spp.) and in presence of a homogenous herbaceous cover crop. The treatments were sufficiently large to attract reasonable numbers of species of nesting birds across the different habitat types. Habitats were physically adjacent to one another, so that any differences in predation within the site should have been attributable to variation in habitat type only, and not to factors such as region, landscape context, or fragmenta- tion. Our specific goals were to examine: (1) whether nesting suceess varied among the planted habitat types, and (2) whether frequencies of cowbird parasitism varied among habitat types. METHODS Our central Missouri, USA research sites were managed by the Missouri Department of Conser- vation and were in predominately agricultural floodplain landscapes surrounded by agriculture or early-SLiccessional vegetation originating from Hoods of the mid-1990s. Plowboy Bend Conser- vation Area (38° 48' 05" N, 92° 24' 17" W) was a row-crop agriculture/tloodplain ecosystem west of the Missouri River’s main channel within a levee- protected floodplain. The suirounding landscape within a 5-km radius (Driscoll and Donovan 2004) centered on Plowboy Bend was ~24% cropland. 307 308 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2. June 2010 26% grassland, 34% forest, and 15% water (based on circa-2000 TM satellite data interpreted by the Missouri Resources Assessment partnership; www.cerc.usgs.gov/morap). Smoky Waters Con- servation Area (38° 35' 09" N, 91° 58' 03" W) was 72 km southeast of Plowboy Bend between the main channels of the Missouri and Osage rivers (a major tributary of the Missouri River). The surrounding landscape within a 5-km radius was —24% cropland, 23% grassland, 33% forest, and 18% water. The floodplain at Smoky Waters is subject to occasional flooding and has not been protected since a levee was breached in the 1993 and 1995 floods. Smoky Waters was flooded for 3 weeks in June 2001, and Plowboy Bend was flooded for 1 week during the same time period; Smoky Waters was also flooded for 1 week in early May 2002. The three 16.2-ha restoration treatment units at each site were cleared of herbaceous vegetation and disked in autumn 1999. Two of the three units were planted with swamp white oak (Qiierciis hicolor) and pin oak {Q. palustris) at a 9 X 9-m spacing (Dey et al. 2001, Shaw et al. 2003). Redtop grass {Agrostis gigantea) was planted in one of the two units, producing a low, dense ground cover that substantially reduced invasion of other herbaceous and woody vegetation. We refer to the combination of planted oaks and redtop grass as the “redtop” treatment. The “no redtop” treatment contained the same oak plant- ings, but without seeded ground cover, and had diverse mixtures of invasive herbaceous and woody growth compared to redtop habitats. The remaining unit at each site served as a “control” and was not planted to either oak or redtop (Shaw et al. 2003). Structurally, vegetation in the control treatments was most similar to that in the no redtop treatments except for the ab.sence of oaks. Planted oaks included conventional 1-year old, not tran.splanted bare root .seedlings, but also included “RPM®” (Root Production Method) oaks. These oaks were grown with a special root- pruning method and attained heights of >1.5 m within the first year of planting (Grossman et al. 2003, Dey et al. 2004). Both sites received the same three plantings in the same arrays but, at Plowboy Bend, the three units adjoined each other in a “pyramid” fashion, whereas at Smoky Waters they were adjoined linearly (redtop, no redtop, control). Nest Monitoring. — We searched for and moni- tored nests from late April to early August during 2001-2003. Nests were located by systematically searching potential nest sites and by observing behavior of adult birds (Martin and Geupel 1993). We devoted equal time to searching in each treatment unit, alternating searches among the three habitats according to visit and time of day within each site. We marked nest locations with plastic flagging placed >3 m from the nest. We focused on Dickcissel (Spiza americana). Field Sparrow {Spizella piisilla). Indigo Bunting (Passerina cyanea), Red-Winged Blackbird {Agelaius phoe- niceus), and Song Sparrow {Melospiza melodia) because they were the most common nesters at the sites. We monitored nests on average every 3- 4 days except during flooding in 2001 and 2002. We found nests of the five focal species in building (22%), laying (20%), incubation (46%), and nestling (12%) stages. We visited nests in early morning on the expected fledging date (based upon observed date of hatching or estimated age of nestlings) to look for evidence of fledging, such as fledgling begging calls, observation of fledglings, parents carrying food to fledglings, or parents chipping rapidly nearby. Nests found empty prior to this date were considered depredated unless we observed evidence of premature fledging. Nests were considered “successful” if they fledged at least one chick; fates of nests where we did not observe these activities were classified as “un- known” for the last interval between visits; this interval was censored to not bias the analysis (Stanley 2004). We approached nests and viewed their contents at the maximum distance possible (—2 to 4 m) to ascertain status, and were careful not to leave “dead end trails” leading to the nest (Martin and Geupel 1993). Interval fates other than “successful” and “depredated” included “abandoned” and “flooding.” Ne.st failures due to cowbird brood parasitism were from parental desertion (eggs remaining in abandoned parasit- ized nest) or complete brood loss after parasitism (starvation of host chicks or complete absence of host chicks or eggs in parasitized nests) and were classified as “cowbird.” “Disturbance” included cases where nests were tipped or removed from the substrate, presumably due to wind' or animal trampling, but the nest contents remained on the ground. Vegetation Measurements. — ^We used vertical density-board measurements taken between 20 June and 8 July in 2002 from 72 systematically Burhans el al. • SONGBIRD NEST SURVIVAL IN l-LOODPl.AINS 309 placed locations in each treatment. Vegetation data were collected in other years, but we chose 2002 data because samples were most complete in that year, and because 2002 represented a temporal midpoint in the study. Percent cover of vertical vegetation was estimated using a nine- increment density board (2.25 m tall X 0.25 m wide). Percent cover of living and dead vegetation was estimated at each 0.25-m increment from a distance of 15 m. We estimated percent cover in each increment for forb, grass, and woody vegetation, combining them to generate an estimate of mean total percent cover for each board measurement. Grand means were calculated for each sample over all of the 0.25-m increments for each vegetation type of interest (forb, grass, and total vertical vegetation). We focused on grass cover, forb cover, and grass height because visually these differences appeared to distinguish treatments from each other (despite the presence of planted oaks, woody cover was negligible within all treatments because the young seedlings were small). We added a category for “grass cover <0.25 m” using only the grass scores from the lowest increment to quantify low dense grass cover, which appeared to visually distinguish redtop treatments from other treatments in the field. We also added “Mean grass height,” which was the last-recorded increment having grass cover on the vertical density board. We randomly chose 68 of the 72 vegetation samples in each site X treatment combination due to inconsistent numbers of samples taken in each treatment. We measured the distance from ground to the bottom of the nest cup (±0.01 m; hereafter nest height) upon finding each nest because nests were, at times, toppled after predation. Nest height was measured in 0.0 1-m increments with marked sticks that were carried daily for nest-searching. We obtained the spatial coordinates (Universal Transverse Mercator) after termination of nesting using a Trimble Geoexplorer Global Positioning System (GPS) unit with an accuracy of 5-7 m. We measured the distance between each nest and the nearest tree (±:3 m in estimated height) using a Geographic Information System (GIS) to compare nest locations with locations ol previously mapped trees and forest patches. Data Analysis. — We used a general linear model (Proc MIXED, SAS Version 9.1, 2003) with an LSMEANS statement to calculate means and standard errors for each vegetation variable of interest. We included site as a random effect to account for differences between the two sites. We used Likelihood Ratio Tests to assess overall model significance against a null model that included only the intercept; we performed multi- ple comparisons tests for differences in least squares means among the three treatment types using a Bonferroni adjustment il the overall model was significant {P ^ 0.05). We investigated whether nest densities varied among restoration treatments by comparing ob- served numbers of nests in each habitat to expected numbers under the assumption ot no difference among restoration treatments. We used Chi-square goodness-of-fit tests to assess statisti- cal significance (P < 0.05). We used the logistic-exposure method (Shaffer 2004) to estimate and analyze nest survival. This approach uses a generalized linear model with binomial distribution (interval nest fate = 0 if failed and 1 if successful) and logit link function to model daily nest survival in terms of covariates potentially affecting nest survival (Dinsmore et al. 2002, Shaffer 2004). We treated each interval between nest checks as an observation, thereby allowing time-dependent covariates such as nest stage to change from one interval to the next during the nesting cycle (Shaffer 2004). We did not include observation days of building stage nests in the analysis. Intervals between nest observations were usually >1 day and varied in length, and we used the modified logit link function; g(0) = loge (0''V[1 - 0'^']), where 0 is the interval survival rate and t is the interval length in days (Shaffer 2004). The effective sample size, which follows from the model likelihood, is the sum of the total number of days that all nests were known to have survived and the number of intervals that ended in failure (Rotella et al. 2004). We used the GENMOD procedure in SAS (SAS Institute 2000) to estimate parameters of the logistic-exposure models. We developed a priori candidate models involv- ing (1) covariates related to our hypothesis about effects of the restoration treatment on nest survival, and (2) additional covariates that we believed might explain variation in ne.st survival (below). — Restoration treatment, which was defined as control, redtop, or no redtop. — Nest height, which has been positively related to nest survival for some of the same species in other studies in the region (Burhans et al. 2002, Burhans and Thompson 2006). 310 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 2. June 2010 — Nesting stage, for which we considered the stage (laying, incubation, nestling) confirmed at the terminal visit of the interval to be the stage tor that interval; e.g., if a nest transi- tioned from incubation to the nestling stage during an interval, a value of “nestling” was assigned to that interval. — Day-of-year, which may reflect seasonal variation in predator activity or abundance. We included both linear and quadratic terms (day-of-year, [day-of-year]“), the latter term to account for non-linear nest survival over the season in models with “day-of-year” (Grant et al. 2005). — Cowbird parasitism, which presumably affects nest predation because of increased begging calls and nest visits by hosts and/or cowbirds, which alerts predators to the presence of nests (Dearborn et al. 1998, Hauber 2000). — Year, site (Plowboy Bend or Smoky Waters), bird species, and site were included as categorical covariates to account for addi- tional unexplained variation. We considered models that corresponded to the covariates and reflected our hypotheses about relationships of habitat and other factors to nest survival: (1) day-of-year, (2) cowbird parasitism, (3) nest height, (4) nesting stage, and (5) habitat. Each model also included species, year, site, and year X site to reflect our study design and to reduce the chance of statistical confounding. We also considered a model that included only species, year, site, and year X site, and a global model that included all covariates. Nests were considered “successful” for the relevant interval if at least one young fledged. We considered complete nest losses due to nest predation, flooding, unknown weather events, or apparent trampling by animals, abandonment, or cowbird parasitism to be “failed” nesting at- tempts. Only one nest fledged a cowbird but not host young; it was considered “successful” for the interval for which it fledged the cowbird. We evaluated support by comparing multiple models, fit to the same data, using an information- theoretic approach (Burnham and Anderson 2002). We considered the model with the lowest value for Akaike’s Information Criterion (AIC) to be the best approximating model for the data, and considered models within two AIC units of the best model to repre.sent potential best models (Burnham and Anderson 2002). However, we report 95% confidence intervals (CIs) for each parameter estimate based on model-averaging to account for model-selection uncertainty (Burn- ham and Anderson 2002). We used model-based methods (Shaffer and Thompson 2007) to esti- mate nest survival (including 95% CIs) in relation to covariates. We computed species-specific period survival to show model-based relationships over the entire nest period in terms of the covariates by raising daily survival to a power equal to the average length of the nest cycle for each species (Field Sparrow = 23 days; Dickcissel = 25.5 days; Indigo Bunting, Red-winged Blackbird, Song Sparrow = 27 days; Ehrlich et al. 1988; D. E. Burhans, unpubl. data). We realize this average does not correspond to the exact nest cycle for each species, but it is close for most of the species, and our goal was to simplify graphic interpreta- tion of model-based results. We show model- based relationships between 26-day period sur- vival rates and covariates other than species by weighting the daily survival rate estimate for each species by the proportion of observed nests of that species. We computed model-averaged survival estimates for a given covariate by holding remaining covariates at their mean value. We calculated period mortality due to a specific cause (abandonment, predation, parasitism, disturbance) by multiplying the model-based average period mortality rate (1 — average period survival) by the proportion of nests in that category; for example, if the average period mortality over all nests was 0.75 and the proportion of abandoned nests was 0.10, period mortality due to abandonment would be estimated as 0.10 X 0.75 = 0.075. We used logistic regression to model the probability of cowbird parasitism, including only nests initiated by the second week of July, as cowbirds typically do not lay after this time in this region (Burhans 1997). We included bird species in each of the models similar to the nest survival models. Models included: (1) species only, (2) habitat (redtop, no redtop, control) and species, (3) site (Plowboy Bend or Smoky Waters) and species, (4) year and species, and (5) a global model with all of the above covariates. We also considered (6) a model incorporating distance to the nearest tree (>3 m in estimated height; range = 1-263 m.). Parasitism probability has been negatively associated with distance to the nearest tree in other studies (Clotfelter 1998), presumably because female cowbirds are known to survey for Biirlums cl al. • SONGBIRD NEST SURVIVAL. IN ELOODPL.AINS 31 1 TABLE I. Distribution of nests across restoration treatments for five songbird species in lower Missouri River floodplains, 2001-2003. Chi-square tests (df = 2, all tests) were used to assess whether nest number and densities varied with restoration treatment. Restoration treatment Specie.s Redlop No redtop Control r r Dickcissel 22 37 8 18.84 <0.001 Field Sparrow 23 5 22 12.28 0.002 Indigo Bunting 5 31 25 18.23 <0.001 Red-winged Blackbird 70 143 40 66.55 <0.001 Song Sparrow 9 15 8 2.69 0.26 host nests from trees (Hauber and Russo 2000, Saunders et al. 2003). RESULTS We analyzed data from 463 nests representing an effective sample size of 4,773. This represent- ed 67 Dickcissel nests (daily survival estimate assuming constant survival = 0.941; Cl = 0.913- 0.960), 50 Field Sparrow nests (0.958; 0.930- 0.975), 61 Indigo Bunting nests (0.957; 0.934- 0.972), 253 Red-winged Blackbird nests (0.951; 0.936-0.962), and 32 Song Sparrow nests (0.933; 0.888-0.961). Estimated daily survival rate of all species combined under the assumption of constant survival was 0.950 (Cl = 0.939-0.959) and average period survival over the entire nesting cycle was 0.26. Thus, average period mortality across the entire nest cycle from all sources was 0.74. Predation (215 losses) accounted for 63% of period mortality. Flooding (18 losses; 5%); abandonment, including that attributable to cow- birds (11 losses, 3%); and unknown (weather or animal trampling) events (9 losses, 3%) accounted for additional mortality. Complete nest losses due to cowbird parasitism occurred only through abandonment with desertions at one Field Spar- row nest and three Indigo Bunting nests (1%). Nests were not equally distributed among resto- ration treatments for all species except Song Sparrow (Table 1). The best approximating model of nest survival was that with parasitism (Table 2; model-estimat- ed period survival for parasitized nests: 0.10, Cl = 0.00-0.40; unparasitized nests: 0.27, 0.19- 0.35). Day-of-year and global models also re- ceived support (Table 2). We found only weak support for differences in nest survival attribut- able to restoration treatment (Table 2, Fig. IB). The site X year and species covariates were not related to specific hypotheses, but we include graphs of the model-averaged values to show the considerable variation (Fig. 1C, D). Cowbird Parasitism. — ^We analyzed 426 nests for probability of cowbird parasitism. Cowbird TABLE 2. Candidate models of dally nest survival for five songbird species in lower Missouri River floodplains, 2001- 2003. Models with a lower AAIC and a greater Akaike weight (w,) have greater support; K is the number of parameters in the model. Model Variable,s K AIC AAIC W, Parasitism Year X Site, Year, Site, Species, Parasitism" 11 1380.73 0 0.52 Day-of-year Year X Site, Year, Site, Species. Day-of-year" 12 1382.06 1.34 0.27 Global Year X Site, Year, Site, Species, Parasitism', Day-of-year", Nest height. Nesting stage. Habitat 18 1383.98 3.33 0.10 Species Year X Site, Year, Site, Species 10 1385.29 4.55 0.05 Nest height Year X Site, Year, Site, Species, Nest height 1 1 1386.74 6.01 0.03 Nesting stage Year X Site, Year, Site, Species, Ne.sting stage (laying, incubation, nestling) 12 1387.51 6.79 0.02 Restoration treatment Year X Site, Year, Site, Species, Restoration treatment (redtop, no redtop, control) 12 1388.28 7.56 0.01 Null 1 1420.55 39.76 0.00 “ Unparasitized = 0; parasitized = 1 . “ Includes day-of-year, (day-of-year)^. 312 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. 2, June 2010 FIG. 1. Model-averaged interval survival estimates (±95% confidence intervals) for covariates (A) day-of-year, (B) restoration treatment, (C) site X year, and (D) species from Missouri River floodplain reforestation sites, 2001-2003. parasitism frequency was 24.0% for Indigo Buntings {n = 55 nests); other species were rarely parasitized (Dickcissel 5.1%, n = 59 nests; Field Sparrow 4.6%, n = 44 nests; Red-winged Blackbird 0.8%, n = 239 nests; Song Sparrow 3.5%, n = 29 nests). The best-supported model included distance to nearest tree and species (Table 3). Probability of parasitism decreased with distance to the nearest tree (Fig. 2). Vegetation. — Mean forb cover varied overall among treatments (/" = 13.1, df = 2, P < 0.05; Table 4). Mean forb cover was lower in redtop versus no redtop treatments (t = 12.64, df = 2, adj. P = 0.019), but did not differ in pairwise multiple comparisons between other treatments. Grass cover <0.25 m varied overall among treatments (y~ = 8.1, dt = 2, P < 0.05; Table 4) but did not differ in any multiple-comparisons tests between pairs of treatments. DISCUSSION We found little support for our principal hypothesis about variability in nest survival due to differences in floodplain restoration treatments. Nest densities generally varied among habitat- treatment types and mean forb cover was lower in redtop treatments; the model relating nest survival to restoration treatment had the weakest support of any model except the null model. Twedt et al. (2002) working in Mississippi and Louisiana also evaluated songbird nest survival in relation to different floodplain restoration treat- ments. Their study considered plantings with a wider range of ages (2-10 years) and tree heights (~3- 15 m), and they also found no differences in nest survival among treatments. Daily nest survival of Indigo Bunting, Red-winged Blackbird, and Dick- cissel in their treatments with nest sample sizes >3 1 ranged from 0.919 to 0.948 (Twedt et al. 2002). Burhaus et ai • SONGBIRD NEST SURVIVAL IN FLOODF^LAINS 313 TABLE 3. Candidate models for iiieidence of cowbird brood parasitism in lower Missouri River tloodplains, 2001-2003. Models with a lower AAIC and a greater Akaike weight (« ,) have greater support; K is the number of parameters in the model. Model K AlC AAlC M’, Distance to nearest tree, species 6 130.89 0.00 0.95 Global, all variables 1 1 136.88 5.99 0.05 Species only 5 141.96 1 1.07 0.00 Site, species 6 143.91 13.02 0.00 Year, species 7 144.54 13.65 0.00 Restoration treatment, species 7 145.87 14.98 0.00 Null 1 169.36 38.48 0.00 Daily nest survival rates at our sites were comparable to other studies in central Missouri, but were both higher and lower than sites in northern Missouri. Average daily survival for Field Sparrows in a study at old-field sites <30 km away (Burhans et al. 2002) was 0.936 ± 0.004 (SE) compared with 0.958 from model-based estimates in the present study. Daily nest survival for Indigo Buntings at those same old fields was 0.955 ± 0.003 compared to our model-based estimate of 0.957. In contrast, daily survival of Indigo Buntings nesting in riparian forests and buffer strips in northeastern Missouri was 0.90 (Peak et al. 2004). Our model-based Field Sparrow (0.330) and Dickcissel (0.203) interval survival rates (Fig. ID) were lower than Mayfield success rates for Field Sparrows (0.472 ± 0.6 SE) and Dickcissels (0.297 ± 0.2) in Conservation Reserve Program sites in northern Missouri (McCoy et al. 1999). A negative relationship between cowbird par- asitism and nest survival was evident from the best model (Table 2), although only four complete nest losses, via nest abandonment, were directly attributable to cowbird parasitism. Nest abandon- ment attributed to cowbird parasitism has been noted frequently for Field Sparrows at more heavily-parasitized locations (Strausberger and Burhans 2001), but has rarely been documented for Indigo Buntings after clutch initiation (re- viewed in Burhans et al. 2000). We may have incorrectly attributed bunting abandonments that were due to death of parents or other disturbances to cowbird parasitism. That would not affect the outcome of the logistic-exposure survival analy- sis, however, because these nests would still be categorized as parasitized. Parasitized nests could FIG. 2. Model-based probability of cowbird parasitism in relation to distance from nearest tree for songbird species using Missouri River floodplain reforestation sites, 2001- 2003. have experienced lower survival if predators were attracted to them because of more frequent and louder begging calls by cowbird offspring (Dear- born 1999) or more frequent visits by host parents (Dearborn et al. 1998). However, we had few parasitized nests that survived sufficiently long to contain cowbird nestlings. A final explanation could be that both nest predators and cowbirds tended to find the same nests. Considerable variation in nest survival among years, sites, and dates was obvious from our most- supported models of daily survival. Nest survival in an earlier study of Field Sparrow and Indigo Bunting in Missouri old-field habitats (Burhans et al. 2002) was similarly higher in early May and August and lower in June and early July. Temporal patterns in nest survival likely reflect unmeasured variables such as within-season differences in predator activity or among-year differences in predator foraging patterns or abundance. Flooding may have contributed to important yearly mortality differences, particular- ly at one site. Sixteen nests were lost to flooding at Smoky Waters in 2001 and 2002 with 15 of those lost in 2001, whereas only two nests were lost to flooding at Plowboy Bend, both in 2001 (Fig. 1 C). We did not detect a strong effect of nest height on nesting success, although two other studies from the same region found nest survival increased with increasing nest height (Burhans et al. 2002, Burhans and Thompson 2006). The model-averaged parameter estimate for nest height (0.03, Cl = —0.10-0.15) suggested a weak positive effect, but the range of heights in the present study (0-2.10 m) was lower (0 to 5.5 and 314 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 2. June 2010 TABLE 4. Least-squares means (±95% Cl) for vegetation measurements from bottomland restoration sites in lower Missouri River floodplains, 2002. Vegetation variables (expressed as proportions), except for mean grass height, are mean cover values across four vertical board measurements taken for each sample. “ Vegetation variable Restoration treatment Redlop No redtop Control Mean grass cover 0.21 ± 0.16 0.16 ± 0.16 0.26 ±0.16 Mean forb cover 0.05 ±0.10 0.27 ± 0.10 0.16 ± 0.10 Mean total vegetation 0.27 ± 0.22 0.43 ± 0.22 0.42 ± 0.22 Mean grass height, m 0.80 ± 0.40 0.67 ± 0.40 1 .07 ± 0.40 Grass cover <0.25 m 0.81 ± 0.29 0.23 ± 0.29 0.43 ± 0.29 Sixiy-eighi sample.s w'ere taken for each treatment in each site. 0 to 3.6 m, respectively). The reduced range of nest heights in our study may not have been sufficient to reveal an effect, given the positive influence of extraordinarily tall nests on survival in Burhans et al. (2002). Probability of cowbird parasitism decreased with distance to the nearest tree (Fig. 2). This finding is in agreement with those of Clotfelter (1998), Hauber and Russo (2000), and Saunders et al. (2003). Johnson and Temple (1990) found that parasitism frequencies in open prairie habitats were higher near wooded edges, which presum- ably contained trees for perches, than far from wooded edges. Twedt et al. (2002) found higher parasitism frequencies in stands with tall cotton- wood (Populus spp.) trees compared to younger stands with shorter trees in Mississippi and Louisiana floodplain restoration sites. Our sites experienced low parasitism compared to more forested (—70%) sites in the region (Burhans et al. 2000, Burhans and Thompson 2006) despite having low (~33%) landscape-level forest cover. Field Sparrows experienced parasit- ism of 11.3% (n = 443 nests) in a study of cowbird parasitism in old fields within 30 km of our sites (Burhans and Thompson 2006) compared to 4.6% in this study; Indigo Buntings experi- enced parasitism frequencies of 48% (/? = 295 nests) in old fields compared to 23.6% in this study. Dickcissels were parasitized at 9.6% (/? = 242 nests; Winter 1999) in southwestern Missouri prairie fragments compared to 5.1% at our sites; Dickcissels at other sites in the Midwest have been parasitized at frequencies of 60-85% (Zimmerman 1983). Landscapes with low forest cover such as ours are usually associated with high cowbird abundance and parasitism (Robin- son et al. 1995, Donovan et al. 2000, Thompson et al. 2000). We observed Brown-headed Cowbirds only rarely during 5 years of breeding point counts at these sites (D. E. Burhans and B. G. Root, unpubl. data), whereas cowbird detections were an order of magnitude higher on point counts at old field sites 30 km distant during the same period (Burhans and Thompson 2006). Old fields and large river floodplains are different habitats, and may be prone to different levels of parasitism (Robinson and Herkert 1997), possibly because of differences in vegetation structure. Twedt et al. (2002) similarly worked at sites having low landscape-level forest cover (36%), but noted lower cowbird numbers in stands having short trees at their restoration sites. Cowbirds could be abundant in these large river landscapes, but possibly select habitats within them that have more or taller trees (Twedt et al. 2002). Studies have shown that regional and landscape effects such as forest cover constrain parasitism at more- local scales (Donovan et al. 2000, Thompson et al. 2000). However, most landscape-level studies have used forested habitats to document parasit- ism frequencies and cowbird numbers (e.g., Robinson et al. 1995). Several studies have looked at variation in cowbird numbers and parasitism among habitats (Hahn and Hatfield 1995, Straus- berger and Ashley 1997, Robinson et al. 1999), but the relationship between landscape-level cowbird abundance and allocation of cowbirds among habitats deserves further work. Much of our understanding about landscape determinants of cowbird parasitism comes from work in the 1980s-2000s in upland forested habitats (Britting- ham and Temple 1983, Robinson et al. 1995, Donovan et al. 2000, Thompson et al. 2000). Birds using forested bottomlands, abandoned bottomland fields, and grasslands within large Biirinms cl at. • SONGBIRD NEST SURVIVAL IN I LOODPLAINS 315 Hoodplain agricultural landscapes may differ from forests in habitat preference by cowbirds, and require different management considerations. We do not have data for livestock presence, but low parasitism in our study could be explained by the apparent scarcity of cattle operations in proximity to our sites. Nesting studies have shown decreases in parasitism with distance from grazed habitats (Goguen and Matthews 1999, 2000) despite cowbirds’ ability to commute between breeding and feeding areas (Thompson 1994). We found differences in habitat use by the species in the restoration treatments, but our data did not show that nest survival varied with planting treatment. Our findings and those of Twedt et al. (2002), who also studied agricultural floodplains low in regional forest cover, indicate that further study of interactions between land- scape- and habitat-level patterns in Brown-headed Cowbird abundance and parasitism frequencies are warranted. ACKNOWLEDGMENTS We thank F. R. Thompson for helpful suggestions about the data analysis. Many field assistants helped gather information for this study, including Marjanne Manker, Kristin Ellis, Charles Sharp. Heather Heitz, Faren McCord, Sarah Moody, and Tina Foglesong. Maria Furey mapped site-level geographic data used in the study, and Bill Dijak provided statistics from those data. Jennifer Grabner helped with the vegetation data and her crew gathered the vegetation samples. We thank Soda Popp for generously allowing use of his access road to Smoky Waters and Terry Bruns for help when vehicles got stuck. Wes Bailey and anonymous reviewers provided helpful comments on earlier drafts. Some equipment, vehicles, and computer support were courtesy of the USDA Forest Service, North Central Research Station. This work was funded through the University of Missouri’s Center for Agroforestry under cooperative agreements 58-6227-1-004 with the Agricul- tural Research Service and C R 826704-01-2 with the U.S. Environmental Protection Agency. The results presented are the sole responsibility of the principal investigators and/ or University of Missouri and may not represent the policies or positions of the EPA. Any opinions, findings, conclu- sions or recommendations expressed in this publication are those of the author(s) and do not necessarily reflect the view of the U.S. Department of Agriculture. LITERATURE CITED Brittingham, M. C. and S. A. Temple. 1983. Have cowbirds caused forest songbirds to decline? Biosci- ence 33:31-35. Burhans, D. E. 1997. Habitat and microhabitat features associated with cowbird parasitism in two forest edge cowbird hosts. Condor 99:866-872. Buriian.s, D, E. and F. R. Thomp.son III. 2006. Songbird abundance and parasitism differs between urban and rural shrublands. Ecological Applications 16:394- 405. Burhans, D. E., F. R. Thompson III, and .1. Faaborg. 2000. Costs of parasitism incurred by two songbird species and their quality as cowbird hosts. Condor 102:364-373. Burhans, D. E., D. Dearborn, F. R. Thomp.son HI. and J. Faaborg. 2002. Factors affecting predation at song- bird nests in old fields. Journal of Wildlife Manage- ment: 66:240-249. Burnham, K, P. and D. R, Anderson. 2002. Model selection and multimodel inference: a practical information-theoretic approach. Springer-Verlag, New York, USA. Chalfoun, a. D., M. J. Ratnaswamy, and F. R. Thompson III. 2002. Songbird nest predators in forest-pasture edge and forest interior in a fragmented landscape. Ecological Applications 12:858-867. Clotfelter, E. D. 1998. What cues do Brown-headed Cowbirds use to locate Red-winged Blackbird nests? Animal Behavior 55:1181-1189. Dearborn, D. C. 1999. Brown-headed Cowbird nestling vocalizations and risk of nest predation. Auk 116: 448^57. Dearborn, D. C., A. D. Anders, F. R. Thompson 111, and J. Faaborg. 1998. Effects of cowbird parasitism on parental provisioning and nestling food acquisition and growth. Condor 100:326-334. Dey, D., j. Kabrick, j. Grabner, and M. Gold. 2001. Restoring oaks in the Missouri River floodplain. Pages 8-20 in Proceedings of the 29"’ Annual Hardwood Symposium, 16-19 May 2001. French Lick, Indiana, USA. Dey, D. C., W. Lovelace, J. M. Kabrick, and M. A. Gold. 2004. Production and early field performance of RPM® seedlings in Missouri floodplains. Pages 59-65 in Proceedings of the 6* Walnut Council Research Symposium. USDA Forest Service, General Technical Report NC-243. North Central Research Station, St. Paul. Minnesota, USA. Dinsmore, S. j., G. C. White, and F. L. Knopf. 2002. Advanced techniques for modeling avian nest survival. Ecology 83:3476-3488. Donovan, T. M., F. R. Thompson 111. and J. R. Faaborg. 2000. Cowbird distribution at different scales of fragmentation: trade-offs between breed- ing and feeding opportunities. Pages 255-264 in Ecology and management of cowbirds (J. N. M. Smith. T. L. Cook, S. 1. Rothstein, S. K. Robinson, and S. G. Sealy, Editors). University of Texas Press. Austin, USA. Donovan, T. M., F, R. Thomp.son III. J. Faaborg, and J. R. Probst. 1995. Reproductive success of migratory birds in habitat sources and sinks. Conservation Biology 9:1380-1395. Drlscoll. j. L. and T. M. Donovan. 2004. Landscape context moderates edge effects: nesting success of Wood Thrushes in central New York. Conservation Biology 18:1330-1338. 316 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 2, June 2010 Ehrlich. P. R.. D. S. Dobkin, and D. Wheye. 1988. The birder’s handbook. Simon and Schuster Inc., New York, USA. Goguen. C. B. and N. E. Matthews. 1999. Review of the causes and implications of the association between cowbirds and livestock. Studies in Avian Biology 18:10-17. Goguen, C. B. and N. E. Matthews. 2000. Local gradients of cowbird abundance and parasitism relative to livestock grazing in a Western landscape. Conservation Biology 14:1862-1869. Grant, T. A., T. L. Shaffer, E. M. Madden, and P. J. PlETZ. 2005. Time-specific variation in passerine nest- survival: new insights into old questions. Auk 122: 661-672. Grossman, B. C., M. A. Gold, and D. C. Dey. 2003. Restoration of hard mast species for wildlife in Missouri using precocious flowering oak in the Missouri River floodplain, USA. Agroforestry Sys- tems 59:3-10. Hahn, D. C. and J. S. Hatfield. 1995. Parasitism at the landscape scale: cowbirds prefer forests. Conservation Biology 9:1415-1424. Hauber. M. E. 2000. Nest predation and cowbird parasitism in Song Sparrows. Journal of Field Ornithology 71:389-398. Hauber, M. E. and S. A. Ru.sso. 2000. Perch proximity correlates with higher rates of cowbird parasitism of ground nesting Song Sparrows. Wilson Bulletin 112: 150-153. Johnson, R. G. and S. A. Temple. 1990. Nest predation and brood parasitism of tallgrass prairie birds. Journal of Wildlife Management 54:106-1 1 1. Lloyd, J. D. and T. E. Martin. 2005. Reproductive success of Chestnut-collared Longspurs in native and exotic grassland. Condor 107:363-374. Martin, T. E. 1992. Breeding productivity considerations: what are the appropriate features for management? Pages 455-473 in Ecology and conservation of neotropical migrant landbirds (J. M. Hagan III and D. W. Johnston, Editors). Smithsonian Institution Press, Washington. D.C., USA. Martin, T. E. and G. R. Geupel. 1993. Protocols for nest monitoring plots: locating nests, monitoring success, and measuring vegetation. Journal of Field Ornithol- ogy 64:507-519. McCoy, T. D.. M. R. Ryan, E. W. Kurzeje.ski, and L. W. Burger. 1999. Con.servation Reserve Program: source or sink habitat for grassland birds in Missouri? Journal of Wildlife Management 63:530-538. Peak, R. G., F. R. Thompson 111, and T. L. Shaffer. 2004. Factors affecting songbird nest survival in riparian forests in a midwestern agricultural land.scape. Auk 121:726-737. PiETZ, P. J. and D. a. Granfors. 2000. Identifying predators and fates of grassland passerine nests using miniature video cameras. Journal of Wildlife Manage- ment 64:71-87. Ricklefs, R. E. 1969. An analysis of nesting mortality in birds. Smithsonian Contributions to Zoology 9:1- 48. Robinson, S. K. and J. L. Herkert. 1997. Cowbird parasitism in different habitats. Illinois Natural History Reports 348:2-3. Robinson, S. K., J. D. Brawn, S. F. Morse, and J. R. Herkert. 1999. Use of different habitats by breed- ing Brown-headed Cowbirds in fragmented mid- we,stern landscapes. Studies in Avian Biology 18:52- 61. Robinson, S. K., F. R. Thompson III, T. M. Donovan, D. R. Whitehead, and J. Faaborg. 1995. Regional forest fragmentation and the nesting success of migratory birds. Science 267:1987-1990. Rotella, j. J., S. J. Dinsmore, and T. L. Shaffer. 2004. Modeling nest-survival data: a comparison of recently developed methods that can be implemented in MARK and SAS. Animal Biodiversity and Conservation 27:187-205. SAS Institute Inc. 2000. Version 8. SAS Institute Inc., Cary, North Carolina, USA. Saunders, C. A., P. Arcese, and K. D. O’Connor. 2003. Nest site characteristics in the Song Sparrow and parasitism by the Brown-headed Cowbird. Wilson Bulletin 1 15: 24-28. Shaffer, T. L. 2004. A unified approach to analyzing nest success. Auk 121:526-540. Shaffer, T. L. and F. R. Thompson III. 2007. Making meaningful estimates of nest survival with model- based methods. Studies in Avian Biology 34:84- 95. Shaw, G. W., D. C. Dey, J. Kabrick, J. Grabner, and R. M. Muzika. 2003. Comparison of site preparation methods and stock types for artificial regeneration of oaks in bottomlands. Pages 186-199 in Proceedings of the 13* Central Hardwood Forest Conference, 1-3 April 2002, Urbana, Illinois (J. W. Van Sambeek, J. O. Dawson, F. Ponder Jr., E. F. Lowenstein, and J. S. Fralish, Editors). USDA Forest Service. General Technical Report NC-234. North Central Research Station, St. Paul, Minnesota. USA. Stanley, T. R. 2004. When should Mayfield data be discarded? Wilson Bulletin 116:267-269. Strausberger, B. M. and M. V. Ashley. 1997. Community-wide patterns of parasitism of a host “generalist” brood-parasitic cowbird. Oecologia I 12:254-262. Strausberger, B. M. and D. E. Burhans. 2001. Nest desertion by Field Sparrows and its possible inlJuence on the evolution of cowbird behavior. Auk 118:770- 776. Thompson III, F. R. 1994. Temporal and spatial patterns of breeding Brown-headed Cowbirds in the midwestern United States. Auk 1 1 1 :979-990. THOMP.SON 111, F. R. and D. E. Burhans. 2003. Predation of songbird nests differs by predator and between field and forest habitats. Journal of Wildlife Management 67:408-416. Thomp.son III. F. R.. W. Duak, and D. E. Burhans. 1999. Video identification of predators at .songbird nests in old fields. Auk 1 16:259-264. Thompson 111. F. R.. S. K. Robin.son, T, M. Donovan. J. R. Faaborg, D. R. Whitehead, and D. R. Larsen. Bwhcnis et al. • SONGBIRD NEST SURVIVAL IN I’LOODPLAINS 317 2000. Biogeographic. landscape, and local factors affecting cowbird abundance and host parasitism levels. Pages 271-270 in Ecology and management of cowbirds (J. N. M. Smith, T. L. Cook, S. I. Rothstein, S. K. Robinson, and S. G. Sealy, Editors). University of Texas Press, Austin, USA. Twedt, D. J., R. R. Wilson, J. L. Henne-Kerr, and D. A. Grosshuesch. 2002. Avian response to bottomland hardwood restoration: the first ten years. Restoration Ecology 1 0:645-6.^.‘>. Winter, M. 1999. Nesting biology of Dickcissels and Henslow’s Sparrows in southwestern Missouri prairie fragments. Wilson Bulletin I 1 1:315-526. Zimmerman, J. L. 1983. Cowbird parasitism of Dickcis.sels in different habitats at different nest densities. Wilson Bulletin 95:7-22. The Wilson Journal of Ornithology 122(2):3 18-325, 2010 CAROTENOID-BASED MALE PLUMAGE PREDICTS PARENTAL INVESTMENT IN THE AMERICAN REDSTART RYAN R. GERMAIN,'-^ MATTHEW W. REUDINK,'--' PETER P. MARRA,2 AND LAURENE M. RATCLIEEE' ABSTRACT. — We examined whether male plumage coloration signals parental quality in the American Redstart (Seiophaga ruticilla), a highly ornamented, migratory warbler. We measured the relationship between both adult male arrival date and phenotype (morphology, melanin- and carotenoid-based plumage), and parental care levels of both parents. Males with brighter flank feathers made more visits to the nest and spent more time at the nest, consistent with the 'good- parent hypothesis’. Female parental care (number of visits) was negatively correlated with intensity of red of her mate’s tail feathers and positively associated with her mate’s parental effort. These data indicate offspring of brighter males receive more care from both parents. Our results suggest carotenoid-based plumage traits of male American Redstarts may have an important role in intersexual signaling, and add to our understanding of the evolution of multiple ornaments. Received W July 2009. Accepted 20 January 2010. Three major models have been developed to explain patterns of association between male ornamentation and parental care. The ‘good- parent’ hypothesis predicts that when paternal care influences offspring viability directly, males should signal their ability to provide for offspring to prospective mates, and more attractive males should provide more care (Heywood 1989, Hoelzer 1989, Norris 1990, Hill 1999, Siefferman and Hill 2003). Some evidence suggests female parental care does not decrease in response to increased care by their mate, resulting in a net benefit to the offspring of ‘good-parent’ males (Sanz et al. 2000, Schwagmeyer and Mock 2003, but see Wright and Cuthill 1989, 1990; Markman et al. 1995). Alternatively, the ‘differential allocation’ hypothesis predicts females mated to more attractive males increase their own level of parental effort to avoid desertion by their mates and to enhance their fitness through production of high-quality offspring (Burley 1986, 1988; Studd and Robertson 1988; Badyaev and Hill 2002). In this model, attractive males restrict their parental investment to allocate more energy towards their own survival, resulting in a negative relationship between male attractiveness and care (Burley 1986, 1988; Mpller and Thornhill 1998). A third ' Department of Biology. Queen’s University, Kingston, ON K7L 3N6, Canada. -.Smithsonian Migratory Bird Center, National Zoolog- ical Park, 3001 Connecticut Avenue, Northwest, Washing- ton. D.C. 20008. USA. 'Current address: Trent University, Forensic Science Department. DNA Building, 2140 East Bank Drive, Peterborough. ON K9J 7B8, Canada. 'Corresponding author; e-mail: ryan.germain@qucensu.ca alternative, the ‘trade-off hypothesis’, similarly proposes that attractive males provide less care towards their offspring (to be able to seek extra mating opportunities), resulting in a negative relationship between male attractiveness and level of care (Williams 1966, Magrath and Komdeur 2003, Mitchell et al. 2007). Several studies have failed to find a relationship between male plumage and parental care (Rohde et al. 1999, Smiseth et al. 2001, Cooper and Ritchison 2005). Some of these studies, however, involve species with multiple ornaments whose signal function and context are not well understood. The American Redstart (Seiophaga ruticilla), a small (7-8 g) neotropical migratory warbler, is particularly suited for studies of male plumage coloration and parental care. American Redstarts are sexually dichromatic and yearling males exhibit delayed plumage maturation, suggesting strong sexual selection on plumage coloration. Adult males are black with bright orange plumage patches on their wings, tail, and sides of the breast (flanks), and a white or black breast, depending on bib size (Sherry and Holmes 1997). The plumage of adult male American Redstarts is highly variable, both in bib size (melanin-based plum- age) and orange (carotenoid) coloration (Lemon et al. 1992, Kappes et al. 2009, Reudink et al. 2()()9b). Recent evidence suggests tail feathers of American Redstarts have an important role in intrasexLial signaling during the non-breeding season, where males with brighter tails occupy higher quality winter territories (Reudink et al. 2009a). Additionally, both bib size (Perreault et al. 1997) and flank redness have been linked to a male’s ability to secure paternity at his nest during 318 Gcnnain ci uL • AMERICAN REDSTART I’LUMAGE AND PARENl AL CARE 319 the breeding season, and tail brightness to a male’s ehances of achieving polygyny (Reudink et al. 2009b). In particular, flank feathers may timction as a sexual signal in American Redstarts, as males will puff out their breast and flank leathers in a ‘‘flutt display” during courtship (Sherry and Holmes 1997). Our objectives were to quantify variation in plumage ornamentation, morphology, and arrival behavior of adult male American Redstarts, and levels of parental care provided by both parents to offspring to identify which features may provide a reliable signal to females of male parental effort in this highly ornamented songbird. We predicted male plumage coloration would correlate with levels of parental care, based on recent evidence suggesting that different plumage regions may be used to signal information in different contexts (Reudink et al. 2009a, b; Kappes et al. 2009); however, we made no a priori predictions as to which plumage region(s) might be important. METHODS Field Data and Feather Collection. — ^We con- ducted field work at the Queen’s University Biological Station in southeastern Ontario, Canada (44° 34' N, 76° 19' W), during two consecutive breeding seasons (May-Jul 2006- 2007). The 60-ha study site is a mixed deciduous forest, largely dominated by sugar maple {Acer saccluiriim) and eastern hop-hornbeam (Ostrya virginiana). We surveyed transects from 0600 to 1200 hrs EST each day from 1 May to 1 July across the study site to record the arrival date for each individual. We designated arrival date as the number of days after the arrival of the first male each season (i.e., first date of male an'ival = 0; male that anJved 5 days later = 5) to standardize arrival date across seasons. We captured adult male redstarts on their respective temtories using mist nests and song playback, usually within 7 days of arrival, and females during the nesting phase using mist nets near their nest location. All birds were banded with a Canadian Wildlife Service aluminum band (permit # I0766C) and a unique combination of three color bands (exclud- ing red and orange in males); we recorded unflattened wing chord length (±1.0 mm) as a measure of relative body size. We plucked a single tail feather (third rectrix; R3) and 12-15 orange flank feathers from each male (Quesada and Senar 2006) (CWS collection permit # CA 0154). We classified adult male bib size using a 1-5 scale ( I = small amount of black plumage, 5 = large amount of black plumage with interme- diates given half points) following Lemon et al. (1992). Parental Care Observations. — We monitored American Redstart pairs each season (~40 pairs/ .season) daily until their nest was found, and observed nests every 1-2 days following comple- tion of nest-building until hatching and recorded the number of young. We confirmed the number of nestlings again on day 5 after hatching and continued to monitor nests until fledging. Fifty nests were observed: 25 in 2006 and 25 in 2007. We erected ground-based video cameras (Canon ZR500) on tripods >5 m from the base of the nest tree on days 5 and 7 after hatching to record parental effort. Cameras were focused on the nest and its immediate (<20 cm) surroundings, and recording occuiTed between 0600 and 0900 hrs EST. Each recording lasted 120 min and the first 5 min were discarded to allow parents time to recover from human disturbance. The remaining 1 15 min were used to calculate provisioning rates of males and females, and the time each parent spent at the nest. We recorded and analyzed a total of 148 hrs of video from 38 nests (/? reduced due to high predation rates). A trained observer recorded the measures of parental care through binoculars at a distance >10 m from the nest using a stopwatch if there were more nests to be watched than cameras available (2006 — day 5: n = 8, day 7: n = 5; 2007 — day 5: n = 2, day 7: n = 2). There was no difference between video camera and trained observer nest watches in either the number of visits to the nest, or total time at the nest (2-sample t-test assuming unequal variance, all P values > 0.09). Nests depredated between days 5 and 7 were excluded from analysis (2006: n = 7; 2007: n = 5). We recorded both the total (i.e., sum of days 5 and 7) number of visits to the nest and the total nest attendance (in sec) for males and females as measures of parental care. Both variables are effective measures of parental effort in small insectivorous birds that make frequent trips to the nest (Saetre et al. 1995). Total time at the nest for males may include activities such as vigilance and removal of fecal sacs, but does not necessarily represent a direct benefit to rearing of offspring as it would for females (i.e., brooding); trips to the nest may serve as a more accurate representation of care, as males are continually bringing food. We divided both total number of visits per hour 320 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2. June 2010 and total nest attendance per hour by the number ot chicks present in the nest to standardize for brood size and recording time. We compared standardized parental care variables with raw data to ensure that parental effort/hr/chick was repre- sentative of total parental effort (linear regres- sions, all P < 0.0004). Six males in our study were polygynous; however, primary and second- ary nests were temporally separated, and we included only data from the primary nest (during the period when the male was caring exclusively for the first nest). All individuals were identifiable via color bands, and we were able to ensure that our data set contained no pair duplications across both years. Color Analysis. — ^We mounted feathers from each male on black paper (Colorline Ebony #142), and stacked flank feathers as they would naturally lie on the bird (Siefferman and Hill 2003, Quesada and Senar 2006). Mounting feathers on a black background is a standard technique (Siefferman and Hill 2003, Quesada and Senar 2006), and the paper rrsed had low (<5%) reflectance, indicating it would have a minimal impact on our analysis. We omitted samples from further analysis (n = 5) in instances where there were too few flank feathers or the orange area on the tail feather was too small to obtain reliable readings. We analyzed male plumage and parental care observations for both males and females from 16 nests in 2006 and 17 nests in 2007. We gathered reflectance spectra from flank and tail feathers of each male using a PX-2 pulsed xenon light source attached to an Ocean Optics USB20()0 spectrometer (Dunedin, FL, USA). We took 25 readings throughout the orange (caroten- oid-based) region for both flank and tail feathers following Reudink et al (20()9b). We gathered raw reflectance data into 10 nm bins from 320 to 700 nm using CLRI.0.3 (Montgomerie 2008), and averaged across the 25 measurements. We calculated brightness for tail and flank feathers by averaging percent reflectance from 320 to 700 nm (color variable I: ‘brightness’). The carotenoid- based plumage of American Redstarts consists of peaks in both the UV and red/orange regions of the spectrum. We used principal component analysis (PCA) to collapse the spectrum into a smaller set of independent variables that describe the measures of hue and chroma based on shape of the curve while controlling for variation in brightness (Cuthill et al. 1999, Montgomerie 2006, Stein and Uy 2006, Reudink et al. 2()09b). We verified this method by calculating hue and chroma using the equations: hue=arctan ([R415-510 ~ R.S20-4I.s)/R320-700]/|R573-70() ~ ^415-575)/ R320-7(X)J)5 LIV chl'oma = R320-4l.‘s/R320-7(X)> ‘ind red chroma = R575_7()(/R32o-7(X)- Flank PC I described 85.5% of the variation in shape of the reflectance curve, and loaded positively on short (UV) wavelengths and nega- tively on the longer (red/orange) wavelength region of the spectrum. Flank PC 1 associated negatively with both hue {r = 0.68, n = 32, P < 0.001) and red chroma (r = 0.99, n = 32, P < 0.001), and associated positively with UV chroma (r = 0.81, n - 32, P < 0.001). Tail feather PC I explained 48.4% of the variation, and loaded positively on both the shorter and longer wave- length regions of the spectrum, and negatively on intermediate wavelengths. Tail PC I associated positively with hue (/“ = 0. 16, /? = 33, P = 0.02) and red chroma (r = 0.61, n = 33, P < 0.001), but had no association with UV chroma (r < 0.001, n = 33, P = 0.94). The first principal component from both plumage regions was used to describe ‘redness’ (color variable 2 = ‘red- ness’), where birds with greater ‘redness’ have more negative flank PC I values, and more positive tail PC I values. Statistical Analysis. — All statistics were per- formed using JMP 7.0 (SAS Institute 2007) and R 2.6.1 for Windows (R Development Core Team 2007). We pooled data across both years of study and, in instances where data for the same individual were collected across 2 years (/; = 3), avoided pseudoreplication by randomly selecting 1 year to exclude. We used paired r-tests to compare the parental care variables across both observation days (for both males and females), as well as across gender, and linear regression to examine the relationship between a pair’s number of visits to the nest, and total nest attendance. All plumage variables were tested for co-linearity using Pearson’s pairwise analysis. We applied backwards stepwi.se multiple regressions with parental care variables (male visits/hr/chick, female visits/hr/chick, male time at the nest/hr/ chick, female time at the nest/hr/chick) as the response variable to calculate the factors that best predicted both male and female parental care. We used measures of male tail brightness, tail redness (PC I), flank brightness, flank redness (PC 1), bib score, body size, year, and arrival date as our main effects in each model. Germain et al. • AMERIC AN Rlil^S'l AR I' I’LUMACjl-; ANI) PARHN I'AL C'ARB 321 RESULTS Male American Redstarts did not show a difference in mean time (sec ± SD) spent al the nest between days 5 (108.21 ± 112.53) and 7 (77.98 ± 75.66) (paired Mest, ^32 = 1.65, P = 0.1 1); however, mean number of visits for males on day 7 ( 1 .92 ± 0.73) after hatching were greater than on day 5 (1.52 ± 0.81) (paired /-test, = — 2.89, P = 0.007). Females showed no difference in the mean number of visits to the nest between days 5 (1.56 ± 0.59) and 7 (1.72 ± 0.68) (paired Mest, 632 = —1.78, P = 0.09), but spent significantly more time at the nest on day 5 (778.04 ± 484.18 sec) than day 7 (616.68 ± 523.90 sec) (paired Mest, ^32 = 3.54, P = 0.001). Females did not visit the nest more or less frequently than males (paired Mest, — F27, P = 0.21 ), but did spend a greater amount of time at the nest (paired Mest, ^34 = —7.91, P < 0.0001 ). There was a significant positive relation- ship between each member of a breeding pair for number of visits (;~ = 0.21, F\ -i\ = 8.17, P = 0.008) and nest attendance (r = 0.30, F| 31 = 13.11, F = 0.001). Pairwise correlations between all male plumage variables revealed that tail brightness increased with decreasing tail redness (PC I) (/~ = 0.12, Fi.31 = 4.30, P = 0.05). No other plumage variables were significantly coirelated. Plumage Color and Provisioning. — Results of a backwards stepwise multiple regression revealed total number of male visits/hr/chick was signifi- cantly predicted by the brightness of flank feathers (Table 1 ). This data set contains one high value for flank brightness (Fig. 1 ). This point is not due to a technical eiTor in measurement, as readings taken before and after this point were within the normal range, and this point remained high after repeated measurements. This individual was also within the normal range for all other measures. The correlation between male flank brightness and number of visits/hr/chick remained significant when this point was excluded (Fig. I ). Flank brightness was also a significant predictor of total time spent by males at the nest (Table i ). Female visits were significantly predicted by year and mate’s tail redness (Table 1), where lemales in 2007 made more trips to the nest than those in 2006 (two-sample Mest assuming unequal vari- ance, t25 = —2.77, P = 0.01). There was a negative relationship between female visits to the nest and male tail redness, where females which made fewer trips to the nest were paired to males with redder tails (Fig. 2). None of the remaining variables related to arrival dale, morphology, or plumage predicted male or female levels ol parental care. DISCUSSION The brightness of flank feathers in adult male American Redstarts predicted their total number of visits to the nest, and total time spent at the nest. Our results are consistent with the ‘good- parent’ hypothesis and suggest that flank feather brightness may provide a reliable signal of parental effort in adult male American Redstarts. A significant negative correlation was found between number of female visits to the nest and redness of her mate’s tail feathers. We found a strong positive correlation between breeding partners for visits to the nest, and nest attendance. The ‘good-parent’ hypothesis predicts that at- tractive males should signal their ability to provide increased levels of parental care, compared to less attractive males (Heywood 1989, Hoelzer 1989). We show brightness of male carotenoid-based flank plumage is associated with levels of male parental care. Plumage brightness has been found to predict mating success in other species, such as Golden- collared Manakins (Manacus vitellinus) (Stein and Uy 2006). In addition, it has been linked to stronger immune response to novel antigens (Saks et al. 2003a), suggesting it has an important role in signaling male health status and, potentially, the ability to provide for offspring. The pairwise correlation analysis among all male plumage variables revealed only one signif- icant association; tail brightness increased with decreasing tail redness. This .suggests that in American Redstarts, plumage brightness does not necessarily coirespond to greater feather caroten- oid content (Saks et al. 2()()3b. Andersson and Prager 2006). Instead, plumage brightness may be influenced by the underlying structural matrix of feathers (Shawkey and Hill 2005); several studies have linked structural-based colors with condition/ individual quality (Doiiccl 2002. Doucet and Montgomerie 2003) as well as the level of male parental care (Siefferman and Hill 2003). However, the interactions between structural and pigment- based components of feather coloration remain poorly understood and further research is needed to elucidate patterns of plumage brightness. Recent work by Kappes et al. (2009) with a smaller sample of more southerly breeding 322 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122. No. 2. June 2010 TABLE 1. Multiple regression models examining male and female parental care variables in relation to standardized arrival date, wing length, bib score, flank feather brightness and redness, and tail feather brightness and redness. Initial models are shown for each model. Final models (i.e., models following removal of nonsignificant effects by stepwise backw'ard deletion) are also shown where predictor variables were significant. Sample size increased in the final models where removal ot nonsignificant predictor variables added additional individuals to the model that were missing information for these variables. Male visits/hr/chick (initial model, n = 31, = 0.46) Male visits/hr/chick (final model, /; = 32, r = 0.37) Male visits/hr/chick (initial model, « = 31, /- = 0.19) Male visits/hr/chick (final model, n = 32, r =0.1) Female visits/hr/chick (initial model, /; = 31, r = 0.42) Female visit.s/hr/chick (final model, n = 32, r = 0.33) Female visits/hr/chick (initial model, u = 31, r = 0.14) Female visits/hr/chick (final model, n = 33) American Redstarts, found males with brighter flanks provisioned less than duller males, and those with less red Hanks sired more total young. Flank brightness and tail brightness in our study, in agreement with Kappes et al. (2009), were not related suggesting the two plumage regions may be u.sed as signals in different contexts. Flank redness (PC I) is related to genetic paternity Predictor variables p F P Year 0.13 0.09 0.77 Standardized aixival 0.01 0.26 0.62 Wing 0.06 0.51 0.48 Bib score 0.06 0.15 0.70 Flank brightness 0.12 3.91 0.06 Flank redness (PC I) 0.03 0.92 0.35 Tail brightness -0.03 0.12 0.74 Tail redness (PC I) -0.04 1.37 0.25 Flank brightness 0.11 15.004 0.0006 Year 22.45 0.18 0.68 Standardized arrival 1.34 0.20 0.66 Wing 1.50 0.03 0.88 Bib score -6.74 0.14 0.71 Flank brightness 4.81 0.52 0.48 Flank redness (PC 1) 3.40 0.84 0.37 Tail brightness 1.31 0.02 0.89 Tail redness (PC 1) -2.19 0.24 0.63 Flank brightness 6.39 4.34 0.05 Year 0.69 4.15 0.05 Standardized arrival -0.02 0.69 0.42 Wing 0.02 0.10 0.76 Bib score 0.06 0.30 0.59 Flank brightness -0.02 0.17 0.69 Flank redness (PC I) 0.004 0.03 0.87 Tail brightness 0.03 0.19 0.67 Tail redness (PC 1) -0.05 3.24 0.09 Year 0.49 9.30 0.005 Tail redness (PC I) -0.05 8.16 0.008 Year 357.2 1.14 0.30 Standardized arrival -9.75 0.27 0.61 Wing 43.61 0.49 0.49 Bib score 64.12 0.3 1 0.58 Flank brightness -28.60 0.46 0.50 Flank redness (PC 1) 27.75 1.39 0.25 Tail brightness 61.82 1.11 0.30 Tail redness (PC I) -5.15 0.03 0.86 No significant predictors (indicating that it may act as a measure of genetic quality) (Reudink et a). 2()()9b), and our results suggest that Hank brightness may also act as an intersexual signal of parental care. Recent evidence suggests American Redstart tail feather coloration acts as an intrasexual signal. Specifically, tail brightness is related to both territory quality in wintering areas (Reudink et al. Germain et at. • AMERICAN REDSTART PLUMACE AND PARENTAL CARE 323 FIG. 1. Linear regression of male visits to the nest (standardized/hr/chick) on flank feather brightness. Gray regression line (r = 0.34, F\ t,q = 15.29, P = 0.0005) includes high value (open circle). Black regression line (r = 0.16, f'|,2y = 5.66, P = 0.02) excludes this individual. 2009a) and polyterritorial polygyny in breeding areas (Reudink et al. 2009b). We found no relationship between tail brightness and either male or female parental care; however, female visits to the nest were negatively correlated with the redness of male tail feathers. Plumage redness is often associated with greater carotenoid depo- sition into the feather structure; high quality birds would be assumed to have redder feathers (Saks et al. 2003b). However, the redness of American Redstart tail feathers has been found to signifi- cantly decrease across subsequent breeding sea- sons (Reudink et al. 2009b). One suggestion is that female American Redstarts alter their level of parental care based on the perceived age of their mates with females providing more care when paired with older males, although we do not currently have the data to address this possibility. Neither male bib size nor arrival date was correlated with any measure of parental care by males or females. Consi.stent with previous studies (Perreault et al. 1997, Reudink et al. 2()09b, but see Lemon et al. 1992), bib size does not appear to have a prominent signaling role; however, anival date has previously been shown to be an important predictor of reproductive success (Norris et al. 2004, Reudink et al. 2009c). Reudink et al. (2009c) demonstrated that early arriving males were more likely to achieve polygyny than those which arrived at a later date. However, when males occupy multiple territories and provide care Male tail feather redness (PC I) FIG. 2. Linear regression of female visits to the nest (standardized/hr/chick) on tail feather redness (PC I) (r = 0.15, Ti,3| = 5.60, P = 0.02). X-axis values are from low to high, where individuals with more negative score (e.g., — 10) have tails with lower redness than those with higher values. for multiple nests, there may be a limit to the amount of care one male can provide, as evidenced by male American Redstarts providing relatively less care at their second nest, than at their primary nest (Secunda and Sherry 1991; RRG, unpubl. data). The strong relationship between both female and male visits to the nest, and nest attendance suggests that offspring of more ornamented males are receiving the benefit of more care by both parents. Our results also suggest the potential for assortative mating, where females that provide more care pair with males which do the same. These data conform to models which suggest that equality of investment in species with biparental care is an evolutionarily stable strategy (offspring receive a benefit from both parents providing high amounts of care, reviewed in Wright and Cuthill 1989). Unfortunately, our data do not allow us to differentiate the effects of individual feeding effort from those of territory quality, as both male and female feeding rates could be affected by availability of food on their territory. It is also possible that male coloration is not indicative of parental care, but rather the association arises from brighter males obtaining higher quality teiritories and provisioning chicks more often with less effort. Male American Redstarts able to secure high quality territories during the over- wintering season arrive earlier in breeding areas 324 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 2, June 2010 and show higher levels of realized reproductive success (Reudink et al. 2009c). One possibility is that socially dominant males arriving earlier from high quality winter territories also obtain higher quality territories during the breeding season, in which case we should observe a relationship between aixival and provisioning. We did not observe a relationship between male arrival date and parental care, but direct measurements of territory quality are needed to discern how variations in individual and territory quality influence provisioning. The brightness of male American Redstart flank feathers may have an important role in intersexual signaling but further research is needed to clarify female mate choice in this system. Experimental approaches that measure female response to changes in male ornamentation would be most useful. Research on the ornamental qualities of female plumage to examine if females with brighter carotenoid-based plumage provide more for their offspring and pair with brighter males would also be informative. ACKNOWLEDGMENTS We thank the many field assistants who contributed to this study, and M. M. Osmond and T. D. Barran for assistance in video analysis. V. L. Friesen, J. J. Nocera, J. R. Foote. C. F. Richter, and K. E. Delmore provided helpful comments on earlier versions of this manu.script. We thank R. D. Montgomerie for use of color analysis equipment, and F. J. S. Phelan and F. L. Connor for logistical support at Queen’s University Biological Station. This research was supported by the Natural Sciences and Engineering Research Council of Canada. Queen’s University, National Science Foundation, Smithsonian Institution, Ontario Gov- ernment (MTCU), American Ornithologi.sts’ Union, Society of Canadian Ornithologists, Sigma Xi, and the American Museum of Natural History. LITERATURE CITED Ander,sson, S. and M. Prager. 2006. Quantifying colors. Pages 41-89 in Bird coloration. Volume I, Mechanics and measurements (G. E. Hill and K. J. McGraw, Editors). Harvard University Press, Cambridge. Mas- sachusetts, USA. Badyaev. a. V. AND G. E. Hill. 2002. Paternal care as a conditional strategy: distinct reproductive tactics as.sociated with elaboration of plumage ornamentation in the House Finch. Behavioral Ecology 13:591-.‘i97. Burley, N. 1986. Sexual selection for ae.sthetic traits in species with biparental care. American Naturalist 127: 41.3-445. BiiRi.HY. N. 1988. The differential allocation hypothesis: an experimental test. American Naturalist I.I2: 611-628. CooRER. S. W, AND G. Ri TCiiLSON. 2005. Nestling provisioning by male and female Yellow-breasted Chats: no relationships between morphology and parental care. Journal of Field Ornithology 76: 298- 302. CuTHiLL, I. C., A. T. D. Bennett, J. C. Partridge, and E. J. Maier. 1999. Plumage reflectance and the objective assessment of avian sexual dichromatism. American Naturalist 153: 183-200. Doucet, S. M. 2002. Structural plumage coloration, male body size, and condition in the Blue-black Gras.squit. Condor 104: 30-38. Doucet, S. M. and R. Montgomerie. 2003. Multiple sexual ornaments in Satin Bowerbirds: ultraviolet plumage and bowers signal different aspects of male quality. Behavioral Ecology 14: 503-509. Heywood, J. S. 1989. Sexual selection by the handicap mechanism. Evolution 43: 1387-1397. Hill, G. E. 1999. Mate choice, male quality, and carotenoid-based plumage coloration. Proceedings of the International Ornithological Congress 22: 1654- 1668. Hoelzer, G. a. 1989. The good parent process of sexual selection. Animal Behaviour 38: 1067-1078. Kappes, P. j., B. j. M. Stutchbury, and B. E. Woolfenden. 2009. The relationship between carot- enoid-based coloration and pairing, within- and extra- pair mating success in the American Redstart. Condor 111: 684-693. Lemon, R. E., D. M. Weary, and K. J. Norris. 1992. Male morphology and behavior correlate with reproductive success in the American Redstart {Setophaga nuicillci). Behavioral Ecology and Sociobiology 29:399^03. Magrath, M. J. L. AND J. Komdeur. 2003. Is male care compromised by additional mating opportunity? Trends in Ecology and Evolution 18: 424^30. Markman, S., Y. Yom-Tov, and J. Wright. 1995. Male parental care in the Orange-tufted Sunbird: behaviour- al adjustments in provisioning and nest guarding effort. Animal Behaviour 50: 655-669. Mitchell, D. P., P. O. Dunn, L. A. Whhtingham, and C. R. Freeman-Gallant. 2007. Attractive males provide less parental care in two populations of the Common Yellowthroat. Animal Behaviour 73:165-170. M0LLER, A. P. AND R. THORNHILL. 1998. Male parental care, differential parental investment by females and sexual .selection. Animal Behaviour 55: 1507-1515. Montgomerie, R. 2006. Analyzing colors. Pages 90-147 in Bird coloration. Volume 1. Mechanisms and measure- ments (G. E. Hill and K. J. McGraw, Editors). Harvard University Press, Cambridge, Massachusetts, USA. Montgomerie, R. 2008. CLR, Version 1.03. Queen’s University, Kingston, Ontario, Canada, (available at http://post.queensu.cii/~mont/color/). Norris, D. R., P. P. Marra, T. K, Kyser, T. W. Sherry, AND L. M. Ratcliffe. 2004. Tropical winter habitat limits reproductive success on the temperate breeding grounds in a migratory bird. Proceedings of the Royal Society of London, Series B 271: 59-64. Norris, K. .1. 1990. Female choice and the quality of parental care in the Great Tit Parns major. Behavioral Ecology and Sociobiology 27: 275-281. Germain el al. • AMERICAN REDSTART PEIJMACJE AND PARENTAL CARE 325 Pf.rreaui.t, S., R, E. Lemon, and U. Kuhni.e:in. 1997. Pattern.s and conelales of extrapair paternity in American Redstarts (Setophafia niiicilla). Behavioral Ecology 8; 612-621. Quesada. J. and J. C. Senar. 2006. Comparing plumage color measurements obtained directly from live birds and from collected feathers: the case of the Great Tit Pams major. Journal of Avian Biology 37: 609-616. R Development Core Team. 2007. R: a language and environment for statistical computing. R Foundation for Statistical Computing, Vienna, Austria, http:// www.R-project.org. Reudink. M. W., P. P. Marra, P. T. Boag, and L. M. Ratclife. 2009b. Plumage colouration predicts pater- nity and polygyny in the American Redstart. Animal Behaviour 77; 495-501. Reudink. M. W., C. E. Studds, T. K. Kyser, P. P. Marra, AND L. M. Ratcliffe. 2009a. Plumage brightness predicts non-breeding season temtory quality in a long-distance migratory songbird. Journal of Avian Biology 40:34-41. Reudink, M. W., P. P. Marra, T. K. Kyser, P. T. Boag, K. M. Langin, and L. M. Ratcliffe. 2009c. Non- breeding season events influence polygyny and extra- pair paternity in a long-distance migratory bird. Proceedings of the Royal Society of London, Series B 276:1619-1626. Rohde, P. A., A. Johnsen, and J. T. Lifjeld. 1999. Parental care and sexual selection in the Bluethroat, Luscina s. svecica: a field-experimental test of the differential allocation hypothesis. Ethology 105: 651- 663. Saetre, G. P., T. Fossnes, and T. Slagsvold. 1995. Food provisioning in the Pied Flycatcher; do females gain direct benefits from choosing bright-colored males? Journal of Animal Ecology 64: 21-30. Saks, L., O. Indrek, and P. Horak. 2003a. Carotenoid- based plumage coloration of male Greenfinches reflects health and immunocompetence. Oecologia 134: 301-307. Saks, L., K. McGraw, and P. FIorak. 2003b. How feather colour reflects its carotenoid content. Functional Ecology 17: 555-561. Sanz, j. j., S. Kranenharg, and J. M. Tinbf.rgen. 2000. Differential response by males and females to manipulation ol partner contribution in the Great Tit {Pams major). Journal of Animal Ecology 69: 74-84. SAS Institute. 2007. .IMP® Version 7.0.2. Duxbury, Pacific Grove, California, USA. SCHWAGMEYER, P. L. AND D. W. MOCK. 2003. HoW consistently are good parents good parents? Repeat- ability of parental care in the House Sparrow, Passer domesticHS. Ethology 109: 303-313. Secunda, R. C. and T. W. Sherry. 1991. Polyterritorial polygyny in the American Redstan. Wilson Bulletin 103:190-203. Shawkey, M. D. and G. E. Hill. 2005. Carotenoids need stnictural colours to shine. Biology Letters 1: 121-125. Sherry, T. W. and R. T. Holmes. 1997. American Redstart (Setophaga ruticilla). The birds of North America. Number 277 Siefferman, L. and G. E. Hill. 2003. Structural and melanin coloration indicate parental effort and repro- ductive success in male Eastern Bluebirds. Behavioral Ecology 14: 855-861. Smiseth, P. T., j. Ornborg, S. Andersson, and T. Amundsen. 2001. Is male plumage reflectance correlated with paternal care in Bluethroats? Behav- ioral Ecology 12: 164-170. Stein, A. C. and J. A. C. Uy. 2006. Plumage brightness predicts male mating success in the lekking Golden- collared Manakin (Manaciis vitellimis). Behavioral Ecology 17: 41-47. Studd, M. V. and R. j. Robertson. 1988. Differential allocation of reproductive effort to territorial estab- lishment and maintenance by male Yellow Warblers (Dendroica petechia). Behavioral Ecology and Socio- biology 23: 199-210. Williams, G. C. 1966. Adaptation and natural selection. Princeton University Press, Princeton, New Jersey. USA. Wright, J. and 1. Cuthill. 1989. Manipulation of .sex differences in parental care. Behavioral Ecology and Sociobiology 25: 171-181. Wright, J. and I. Cuthill. 1990. Manipulation of sex differences in parental care: the effect of brood size. Animal Behaviour 40: 462—47 1 . The iVilso?! Journal oj Ornithology 1 22(2):326-333, 2010 VARIATION IN PLUMAGE COLORATION OF NORTHERN CARDINALS IN URBANIZING LANDSCAPES TODD M. JONES,' AMANDA D. RODEWALD,'-" AND DANIEL P. SHUSTACK' - ABSTRACT. — Biologists know relatively little of how one of the most important avian phenotypic signals, feather coloration, may be affected by anthropogenic changes resulting from urbanization. We examined the relationship between urbanization and carotenoid-based plumage color of Northern Cardinals (Cardinalis cardinalis) in 13 riparian forests distributed across a rural-to-urban landscape gradient in central Ohio, USA. Feathers and morphometric measurements were collected from breeding territorial males (n = 129) and females (/? = 145) during March-August 2006-2008. Plumage brightness of males, but not temales, increased with body condition (i.e., size-adjusted mass) and declined with amount of urbanization surrounding forests in which cardinals bred. The extent plumage coloration reflected condition was partially mediated by landscape composition. Specifically, the relationship between brightness and body condition was most pronounced in the most rural land.scapes. The interdependency of male coloration and body condition may be more relaxed in urban than rural landscapes if carotenoid-rich foods from anthropogenic and/or invasive sources are more available, or are accessible to birds across a wide range of condition. Received 11 May 2009. Accepted 18 January 2010. Most studies of birds in human-dominated landscapes have described striking and relatively predictable patterns in response of species diver- sity and abundance to intensity of urban develop- ment (Blair 1996, Anderies et al. 2007, Devictor et al. 2007, Tratalos et al. 2007). Subtle consequences of urbanization, such as changes in behavior or coloration are less commonly examined. Biologists continue to have a poor understanding of how urbanization might influ- ence plumage coloration, despite a century of interest in causes and consequences of this important phenotypic signal in birds (Anders.son 1994, Bortolotti 2006). Most studies of avian coloration in urban systems have focused on detrimental effects of anthropogenic pollution upon levels of carotenoids in plant and animal matter (Sillanpaa et al. 2008) as well as plumage coloration, as in Great Tits (Pants major) (Eeva et al. 1998, Isaksson et al. 2005, Isaksson and Andersson 2007), The presence of certain exotic plants, some of which are positively associated with urban development (Borgmann and Rode- wald 2005), also can affect coloration by altering access to diet-derived pigments. For example, Mulvihill et al. (1992) showed that access to exotic honeysuckle fruits (Lonicera spp.) contain- ing rhodoxanthin, a carotenoid-group (xantho- ' The Ohio State University, School of Environment and Natural Resources, 2021 Coffey Road, Columbus, OH 43210, USA. ^Current address: Massachusetts College of the Liberal Arts. Biology Department, 375 Church Street. North Adam.s. MA 01247. USA. ’Corresponding author; e-mail; rodcwald.l@osu.edu phylls) pigment, changed retrix-tip colors of Cedar Wax wings (Bomhydlla cedrorum) from yellow to orange. However, the consequences of urbanization affecting avian coloration remain poorly understood despite these studies. One of the primary ways in which urban development might affect plumage coloration is by altering food resources. Cities generally contain a variety of novel food resources, including bird seed at feeders, fruit from exotic plants, and refuse that can change diets. These changes in food resources can directly affect coloration through increased availability of pig- ment-containing food sources. The link between diet and plumage coloration should be particularly strong for species that have carotenoid-based plumages, which are highly dependent on carot- enoid pigments obtained from food sources (Mpller et al. 2000, McGraw et al. 2001, Hill et al. 2002). However, access to pigments is not the sole determinant of coloration because individuals may differ in their ability to use ingested pigments (McGraw and Hill 2001). A variety of urban- associated factors, such as abundant human- commensal predators, oxidative stressors (e.g., pollution), and high avian densities, may constrain the ability of individuals to acquire or physiolog- ically sequester carotenoids for color production (Mpller et al. 2000, Isaksson and Andersson 2007). Consequently, plumage quality .is known to reflect general health or immunocompetence (Andersson 1994, Saks et al. 2003, Jawor and Breitwisch 2004, Dawson and Bortolotti 2006, Maney et al. 2008), individual quality (Hill 1991, Kristiansen et al. 2006), and quality of the environment in which birds live (Arriero and 326 Jones et al. • NORTHERN CARDINALS IN URBANIZING LANDSCAPES 327 TABLE 1. Land central Ohio, USA cover characteristics within a l-kni radius of 1 3 riparian-forest study sites in urbanizing landscapes in Study site Urban index I5uiidings‘ Fore.si width (m) Proportion of l-kni-radius area covered Agriculture Lawn Pavenienl Road Public Hunting -1.15 210 1 94 0.32 0.08 0.01 0.01 Prairie Oak.s -1.12 58 1 48 0.47 0.12 0.03 0.02 Three Creek.s -0.71 92 1 33 O.IO 0.10 0.04 0.02 South Galena -0.57 185 1 63 0.I4 0.30 0.02 0.01 Galena -0.48 360 277 0.I5 0.22 0.04 0.02 Elk Run -0.16 812 1 67 0.3 1 0.27 0.06 0.05 Woodside Green 0.32 1,227 1 04 O.l I 0.40 0.07 0.05 Rush Run 0.75 1,61 1 1 50 0.00 0.41 0.09 0.06 Cherrybottoni 0.76 997 1 65 0.02 0.36 0.16 0.07 Kenny 0.89 1,733 1 26 0.00 0.34 0.17 0.06 Casto 1.25 1,776 202 0.00 0.42 0.20 0.08 Lou Berliner 1.26 2,272 1 56 0.00 0.28 0.23 0.08 Tuttle 1.61 2,285 1 60 0.00 0.34 0.30 0.09 “ Number of buildings per landscape. Fargallo 2006). Thus, coloration might be a useful indicator of cryptic detrimental or beneficial effects of urbanization upon an ecosystem. We used the Northern Cardinal {Cardinalis cardinalis) as a model species to examine the extent to which coloration varied with individual body condition and the amount of urbanization surrounding forested parks in central Ohio, USA. The Northern Cardinal is a socially monogamous, dichromatic, year-round resident that exhibits a single annual molt immediately after the breeding season (Halkin and Linville 1999). Cardinals display a brilliant red carotenoid-based plumage that is highly dependent upon their diet (Linville and Breitwisch 1997, McGraw et al. 2001). The species is known to respond positively to urbanization, likely as a consequence of more abundant nesting substrates, fruit resources, bird feeders, and warmer winter temperatures than in comparable rural landscapes (Leston and Rode- wald 2006, Rodewald and Bakermans 2006). Previous studies indicate that forests within urban landscapes contain nearly three times the amount of fruit and nearby bird feeders than in rural areas (Atchison and Rodewald 2006, Leston and Rodewald 2006). Increased fruit re.sources are primarily from the presence of exotic and invasive species, such as Amur honeysuckle (Lonicera maackii) and multiflora ro.se (Rosa midtiflova). These increased food resources represent one of many ways that urbanization may inOuence the condition and coloration of cardinals. Specifically, we asked (1) does plumage coloration vary along a rural-to-urban landscape gradient, and (2) does plumage coloration reflect differences in condition among individuals? METHODS Study Sites. — Cardinals were studied at 13 riparian forest sites within the Scioto River Watershed of central Ohio, USA (Table 1, modified from Rodewald and Shustack 2008b). Study sites consisted of mature forest corridors >250 m long and >100 m wide with all study sites separated by >2 km. Riparian forests across the study area were generally similar in both shape and landscape configuration (i.e., they were narrow, linear, highly-connected forests in frag- mented landscapes). We described the amount of urbanization suiTounding each study site using an urban index developed for a larger concurrent study of avian communities across 35 sites (Rodewald and Shustack 2()08b). Landscape composition within a 1-km radius of each site was calculated ba.sed on digital orthophotographs (2002-2004) and exist- ing data from county auditors. An urban index was calculated for each site to represent the amount of urbanization surrounding that particu- lar forest. Positive values of the urban index indicate increasingly urban forests, which were generally surrounded by >800 buildings within the 1-km radius, and bordered by residential or commercial areas. Negative urban index values represent more rural forests, which were generally surrounded by <360 buildings and bordered by 328 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. 2, June 2010 agricultural fields or pastures (described in Rodewald and Shustack 2008b). Feather Sampling. — ^Feather samples were col- lected from 281 Northern Cardinals captured in mist nets during the breeding season from early March to mid-August 2006—2008. We measured right tarsus length (nearest 0. 1 mm), mass (g), and length of the right unflattened right wing chord (nearest 0.5 mm) for all individuals captured. Six to 10 feathers were collected from the breast of each male while 3-6 feathers were collected from the underwing coverts of both wings of females. These areas of feather collection were chosen because they are known to be important areas for cardinal coloration and have been used in previous studies (Linville and Breitwisch 1997; Wolfenbarger 1999a, c; Jawor et al. 2003, 2004). Collected feathers were labeled, stored in manila coin envelopes, and placed in a freezer until measured. Color Measurements. — Feathers from each individual were superimpo.sed in layers of three on 4.5 X 7.5 cm white note cards to imitate the plumage surface of a bird (Bennett et al. 1997, Quesada and Senar 2006). Superimposed feathers were photographed at a distance of 2 cm using a sLipra-macro feature of a Fujifilm FinePix S800()fd digital camera. Feathers were photo- graphed inside an 20 X 20 X 20 cm Digital Concepts Lighting Studio (with 2 20-watt lights directed at the feathers) to standardize lighting conditions. Images were imported into Adobe Photoshop CS 2™, a program which quantified hue, saturation, and brightness. Hue was defined as a certain point on the color spectrum and was measured on a circular scale in degrees with 0° (360°) being the purest red color (i.e., red with few blue or yellow tones). The program quantified saturation as a percentage of density with i)% being least den.se and 100% being most den.se (i.e., least amount of gray tones). Brightness was quantified on a percentage grayscale with 0% being black and 100% being white. The color- picker tool (3X3 eye-dropper sample) was used to randomly measure and quantify plumage coloration at 10 points within the largest overlap- ping area of the superimposed feathers. We examined the precision of our color measurements by repeating them 10 times and calculating the coefficient of variation for measurements for each individual. Within-individual measurements were preci.se for brightness (CV range = 0.5^. 4; mean ± SF = 1.93 ± 0.04) and saturation (CV range = 0.5-5. 7; mean ± SE = 1.82 ± 0.05). Hue measurements showed only small amounts of variation (CV range = 1.8-34.9; mean ± SE = 7.26 ± 0.25). Consequently, the mean of the 10 repeat measurements was used to represent plumage coloration for each individual. Our approach to color quantification is widely used (Kilner 1997, Kilner and Davies 1998, Dale 2000, Gerald et al. 2001, Eitze and Richner 2002, Bortolotti et al. 2003, Bezzerides et al. 2007, Surmaki and Nowakowski 2007), but it does not necessarily reflect how plumage is perceived by other cardinals. Also, this method is not sensitive to UV light as perceived by birds (Cuthill 2006), although previous studies involving carotenoid plumages reported positive correlations among yellow-red spectrum (500-700 nm) and UV reflectance peaks (Senar and Quesada 2006). Data Analysis. — ^We reduced the number of color variables by applying a principal compo- nents analysis to hue, saturation, and brightness measures separately for males and females. The first two factors for males explained 47% (eigenvalue = 1 .40) and 34% (eigenvalue = 1.01) of the variation in plumage color. The first factor (hereafter called “brightness”) was strong- ly correlated with brightness (—0.75) and satura- tion (0.84). whereas the second factor (hereafter referred to as “hue”) was strongly correlated with hue (0.89). Only the first factor was useful for females explaining 55% (eigenvalue = 1.67) of the variation in plumage color. This first factor (hereafter called “hue”) loaded positively for hue (0.86) and negatively for saturation (—0.85). We first developed a condition index to examine relationships between coloration and body condition using principal component analy- sis (PC A) on wing and tarsus length for males and females separately (SAS Institute 2002). The first factor explained 60% of the variation in body size in males (eigenvalue = 1.19) and was positively correlated with both wing and tarsus (0.77 for each). The first factor for females explained 62% of the variation in body size (eigenvalue = 1.25) and was positively correlated with both wing and tarsus (0.79 for each). We then regressed mass against this body-frame-size factor for. 129 males and 145 females. Mass was positively associated with body size for males (F| 127 = 1 1.92, P < 0.001), and residuals were used as a body condition index, which indicated if individuals were heavier (positive value) or lighter (negative value) than expected for a given body (frame) Jones et al. • NORTHliRN CARDINALS IN URBANIZING LANDSCARLS 329 size. Variation in female mass was not a function of body size (F1.143 = 20.72, F = 0.399), and we used mass alone to examine condition of females. We used mixed effects models to test how (I) condition was related to the urban index, year, and urban by year interaction (urban and year treated as fixed effects; site treated as random effect), and (2) how plumage coloration was influenced by condition, the urban index, and a condition by urban interaction (urban and condition treated as fixed effects; site and year treated as random effects). These models were used separately for males and females, and for each of the two principal components describing plumage color. We also considered how plumage variation among individuals might be related to urbaniza- tion for males and females, separately, by examining coefficient of variations for color metrics at each site, by male/female, and year. We used a mixed effect model with year as a random effect and the urban index as a fixed effect. RESULTS Our derived principal component (lower values indicate brighter plumage) indicated male plum- age was brighter as condition improved (condition (3 = -0.12 ± 0039 SE; Ei,,,, = 10.09, P = 0.002, Fig. I A), but declined with amount of urban development suiTounding forest patches (urban p = 0.17 ± 0.079 SE; = 4.92, P = 0.029, Fig. IB). We detected a significant condi- tion by urbanization interaction (P = 0.10 ± 0.036 SE; Fi.m = 7.77, P = 0.006, Fig. 2), indicating the relationship between brightness and condition was stronger in rural than urban landscapes. Condition varied among years (pa- rameter estimates were —1.99 ± 0.479 SE for 2006, -1.44 ± 0.479 SE for 2007 with 2008 as the reference year; F\,\\ \ = 9.54, P < O.OOl), but was not directly related to the urban index (Ci 1 1 1 = 1.76, P = 0.1869) nor via an urbanization by year interaction (C| m = 0.22, P = 0.8025). Hue was not significantly related to either condition (^i.iii = 9.49, P = 0.3484) or urbanization (F| III = 0.06, P = ().80()), nor was there a condition by urbanization interaction (E|,iii = 0.79, P = 0.376). We found no evidence that among-individual variation of male coloration at sites changed along the rural-urban gradient for either brightness (Fi.is = 0.01, P = 0.917) or hue 28 = 0.54, P = o'^^467). We found no evidence that plumage coloration of females was related to mass (/” 1.127 “ 0.50, P — 0.480), urbanization (A|,i27 = 0.28, P — 0.597), or a mass by urbanization interaction (/’I.I27 = 0.22, P = 0.638). Variation in coloration among females was not significantly related to urbanization (^1.29 = 1.19, /^ = 0.284). We found some evidence that mass varied with urbanization in some years (urban by year interaction estimates: —1.47 ± 0.75 SE for 2006; 0.60 ± 0.808 SE for 2007 with 2008 as the reference year; Ei 127 = 3.60, P = 0.0301), but mass was not directly related to either urbanization (E| 127 = 2.19, E = 0.1416) or year (E| 127 = 1.04, P — 0.3568). Female mass declined with urbanization in 2006, but showed no pattern in 2007 and 2008. DISCUSSION Brightness of male plumage increased with body condition, but declined as the amount of urbanization surrounding forests in which cardi- nals bred increased. The relationship between brightness and body condition was stronger in more rural than urban landscapes. The interde- pendency of coloration and body condition may be relaxed in urban landscapes if birds in comparatively poor condition have access to carotenoid-rich foods due to supplemental feeding or fruits of exotic shrubs. We know in our study system that urban development promotes invasion by exotic Amur honeysuckle (Borgmann and Rodewald 2005), which also is the largest source of fruit at our sites (Leston and Rodewald 2006). Mulvihill et al. (1992) reported that access to exotic honeysuckle fruits containing rhodo- xanthin, a carotenoid-group (xanthophylls) pig- ment was responsible for orange, rather than yellow, tipped retrices of Cedar Waxwings. The weaker interdependency of condition and colora- tion in urban landscapes may explain why we failed to find a direct relationship between male condition and urbanization, despite the reduced brightness of urban males. Others have identified links between plumage and individual quality (Linville et al. 1998. Wolfenbarger 1999b, Jawor and Breitwisch 2004), but we found the extent plumage coloration reflected condition was partially mediated by landscape composition. Our findings suggest plumage coloration may prove to be a less useful indicator of male quality in urban landscapes. The potential decoupling of coloration and quality 330 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 2. June 2010 FIG. 1 . Relationship of plumage brightness as described by the first principal component to (A) body condition and (B) urban index of 129 male Northern Cardinals in riparian forests of central Ohio, 2006-2008. may have important con.sequences given that coloration can serve as a usetiil cue in mate selection (e.g., sexual selection; Darwin 1871, Andersson 1994) and, in the case of cardinals, is known to promote assortative mating (Jawor et al. 2003). Our finding that among-individual variation in plumage was not related to urbanization provides insight into the assertion that synanthropic birds may overmatch resources in urban environments (Shochat 2004). Studies have shown that when individuals have more access or are less limited in .sequestering carotenoid resources, they converge on a similar plumage di.splay, resulting in less variation among individuals (Hill 1992). The comparatively dull plumage of urban birds suggests that densities may exceed resource levels. However, in the case of overmatching, one also would expect to find high variation among individuals in coloration, as only a few birds would have access to the best resources. Carotenoids are generally considered to be a limiting resource among birds displaying carot- enoid-based plumages (Mpller et al. 2000, Dale 2006), specifically for Northern Cardinals (Lin- ville and Breitwisch 1997). One would expect increased intraspecific competition would con- strain individual access to carotenoids in areas where carotenoids are especially limited due either to environmental gradients or density- Jones et al. • NORTHERN CARDINALS IN URBANIZING LANDSCAPES 331 Rural — — Urban PIG. 2. The relationship between body condition and plumage brightness (i.e., principal component I) was stronger for rural (n = 42) than urban (h = 86) male cardinals breeding in riparian forests of central Ohio, 2006-2008. dependence, thereby increasing variation in plum- age coloration among individuals. High variation in carotenoid-based coloration among individuals might be an indicator of the availability of reduced carotenoids in a particular environment. We found no evidence of greater variation in access to carotenoids than individuals in more rural forests despite greater densities of breeding pairs in urban landscapes (Leston and Rodewald 2006, Rodewald and Bakermans 2006). Supple- mental food resources associated with residential land, such as bird feeders and fruiting plants (Atchison and Rodewald 2006, Leston and Rodewald 2006) may permit high densities of cardinals to use similar or greater per capita food resources than birds living in more rural environ- ments. This scenario of apparent resource-match- ing is consistent with patterns reported by Rodewald and Shustack (2008a), which showed that survival, reproductive productivity, and energetic condition of cardinals were similar across the rural-to-urban gradient. The presence of a relationship between color- ation and urbanization suggests development surrounding forests inlluences the ability of cardinals to either acquire or sequester carotenoid pigments. However, there are a number of important caveats to our findings. First, given that coloration of cardinals may improve with age (Wolfenbarger 1996), we cannot exclude the possibility that different age structure in urban and rural populations contributed to the observed patterns. Second, there is a possibility that cardinals molted and grew feathers at different locations than where they were captured, which may reflect different food conditions. However, cardinals in the study system: (1) established breeding territories beginning in March, (2) generally molted in August while still on teiritory, and (3) had high site fidelity across years. Third, we focused only on plumage coloration, and other ornaments (e.g., mask, crown, and beak colora- tion) of cardinals also may be important indicators of quality (Jawor and Breitwisch 2004). Our study provides some evidence that urban disturbances can influence plumages of Northern Cardinals in our system, but continued study is needed to identify and understand the underlying mecha- nisms of plumage differences. ACKNOWLEDGMENTS This research was generously supported by a Schwab Associate Scholarship grant from The Ohio State Univer- sity. College of Biological Sciences, and a Research Experience for Undergraduates (REU) supplement from the National Science Foundation (DEB-0639429 to A. D. Rodewald). Field components of this research were partially supported by the National Science Foundation ( DEB-034()879 and DEB- 0639429 to A. D. Rodewald) and the Ohio Division of Wildlife. This work would not have been possible without dedicated efforts of L. J. Kearns, I. J. Ausprey, and J. R. Sinith-Castro. We appreciate helpful comments from W. M. Masters and P. G. Rodewald that improved this manuscript. We thank the following for their work in the field; B. T. Adams. Elizabeth Ames. Christine Austin, S. C. Buescher, L. M. Koerner, S. E. Lehnen, W. W. 332 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2, June 2010 Li, L. L. MacArthur, Angela Petersen, J. E. Price, E. C. Rogers, Kaitlin Uppstrom. Benjamin Van Allen, and R. M. Zajac. We are grateful to Franklin County Metro Parks, Columbus Parks and Recreation, Ohio Division of Wildlife, The Nature Conservancy, City of Bexley, Gahanna Parks and Recreation, and private landowners for their coopera- tion and access to sites. LITERATURE CITED Anderies, j. M., K. Madhusudan, and E. Shochat. 2007. Living in the city: resource availability, predation, and bird population dynamics in urban areas. Journal of Theoretical Biology 247:36-49. Andersson, M. B. 1994. Sexual selection. Princeton University Press, Princeton, New Jersey, USA. Arriero, E. and j. a. Fargallo. 2006. Flabitat structure is associated with the expression of carotenoid-based coloration in nestling Blue Tits (Panis caeruleus). Naturwissenschaften 93: 1 73-180. Atchison, K. A. and A. D. Rodewald. 2006. The value of urban forests to wintering birds. Natural Areas Journal 26: 280-288. Bennett, A. T. D., I. C. Cuthill, J. C. Partridge, and K. Lunau. 1997. Ultraviolet plumage colors predict mate preferences in starlings. Proceedings of the National Academy of Sciences of the USA 94: 8618-8621. Bezzerides, L. a., K. j. McGraw, R. S. Parker, and J. Husseini. 2007. Elytra color as a signal of chemical defense in the Asian Ladybird Beetle Harmonia a.xyridis. Behavioral Ecology and Sociobiology 61: 1401-1408. Blair, R. B. 1996. Land use and avian species diversity along an urban gradient. Ecological Applications 6:506-519. Borgmann, K. L. and a. D. Rodewald. 2005. Forest restoration in urbanizing landscapes: interactions between land uses and an exotic shrub. Restoration Ecology 13:334-340. Bortoloiti, G. R. 2006. Natural selection and coloration: protection, concealment, advertisement, or deception? Pages 3^0 in Bird coloration: evolution and function (G. E. Hill and K. J. McGraw, Editors). Volume 2. Harvard University Press, Cambridge. Massachusetts, USA. Bortolotti. G. R., K. j. Fernie, and J. E. Smits. 2003. Carotenoid concentration and coloration of American Kestrels {Falco sparveriiis) disrupted by experimental exposure to PCBs. Functional Ecology 17: 651- 657. Cuthill, 1. C. 2006. Color perception. Pages 3-35 in Bird coloration: mechanisms and measurements (G. E. Hill and K. J. McGraw, Editors). Volume 1. Harvard University Press, Cambridge. Massachu.setts, USA. Dale. J. 2000. Ornamental plumage does not signal male quality in Red-billed Queleas. Proceedings of the Royal Society of London, Series B 267: 2143-2149. Dali:. .1. 2006. Intraspecific variation in coloration. Pages 36-86 in Bird coloration: function and evolution (G. E. Hill and K. .1. McGraw, Editors). Volume 2. Harvard University Press, Cambridge, Mas.sachu.setts, USA. Darwin, C. [1871] 1981. The descent of man, and selection in relation to sex. Princeton University Press, Prince- ton, New Jersey, USA. Dawson, R. D. and G. R. Bortoloiti. 2006. Carotenoid- dependent coloration of American Kestrels predicts ability to reduce parasitic infections. Naturwis- senschaften 93: 597-602. Devictor, V., R. Julliard, D. Couvet, A. Lee, and F. Jiguet. 2007. Functional homogenization effect of urbanization on bird communities. Conservation Biology 21: 741-751. Eeva, T., E. Lehikoinen. and M. Ronka. 1998. Air pollution fades the plumage of the Great Tit. Functional Ecology 12: 607-612. Fitze, P. S. and H. Richner. 2002. Differential effects of a parasite on ornamental structures based on melanins and carotenoids. Behavioral Ecology 13:401-407. Gerald, M. S., J. Bernstein, R. Hinkson, and R. A. E. Fosbury. 2001. Formal method for objective assess- ment of primate color. Journal of Primatology 53: 79- 85. Halkin, S. L. and S. U. Linville. 1999. Northern Cardinal (Cardinalis cardinalis). The birds of North America. Number 440. Hill, G. E. 1991. Plumage coloration is a sexually selected indicator of male quality. Nature 350: 337-339. Hill, G. E. 1992. Proximate basis of variation in carotenoid pigmentation in male House Finches. Auk 109: 1-12. Hill, G. E., C. Y. Inouye, and R. Montgomerie. 2002. Dietary carotenoid predict plumage coloration in wild House Finches. Proceedings of the Royal Society of London, Series B 269: 1119-1124. ISAKSSON, C. AND S. Andersson. 2007. Carotenoid diet and nestling provisioning in urban and rural Great Tits Panes major. Journal of Avian Biology 38: 564-572. ISAKSSON, C., J. Ornborg, E. Stephensen, and S. Andersson. 2005. Plasma glutathione and carotenoid coloration as potential biomarkers of environmental stress in Great Tits. EcoHealth 2: 138-146. Jawor, j. M. and R. Breitwisch. 2004. Multiple ornaments in male Northern Cardinals, Cardinalis cardinalis, as indicators of condition. Ethology 110: 1 13-126. Jawor, J. M., S. U. Linville, S. M. Beall, and R. Breitwisch. 2003. Assortative mating by multiple ornaments in Northern Cardinals (Cardinalis cardina- lis). Behavioral Ecology 14: 515-520 Jawor, J. M., N. Gray, S. M. Beall, and R. Breitwisch. 2004. Multiple ornaments correlate with aspects ol condition and behavior in female Northern Cardinals, Cardinalis cardinalis. Animal Behaviour 67: 875-882. Kilner, R. 1997. Mouth colour is a reliable signal of need in begging Canary nestlings. Proceedings of the Royal Society of London, Series B 264: 963-968. Kilner, R. and N. B. Davies. 1998. Nestling mouth colour: ecological correlates of a begging signal. Animal Behaviour 56: 705-712. Kristiansen, K. O., J. O. Bustnes, I. Folstad, and M. Helberg. 2006. Carotenoid coloration in Great Black- backed Gull Lariis marinits reflects individual quality. Journal of Avian Biology 37: 6-12. Jones ct III. • NORTHERN CARDINALS IN URBANIZING LANDSCAPES 333 Lkston, L. F. V. AND A, D, Rodkwald. 2()()6. Are urban forests ecological traps for unticrstory birds? An examination using Northern Cardinals. Biological Conservation 131: 566-574. Linville, S. U. and R. Breitwisch. 1997. Carotenoid availability and plumage coloration in a wild popula- tion of Northern Cardinals. Auk 1 14: 796-800. Linville, S. U., R. Breitwisch, and A. J. Schilling. 1998. Plumage brightness as an indicator of parental care in Northern Cardinals. Animal Behaviour 55: 119-127. Maney, D. L., a. K. Davis, C. T. Goode, A. Reid, and C. Showalter. 2008. Carotenoid-based plumage color- ation predicts leukocyte parameters during the breed- ing season in Northern Cardinals (Cardiiuilis cardina- lis). Ethology 1 14: 369-380. McGraw, K. J. and G. E. Hill. 2001. Carotenoid access and intraspecific variation in plumage pigmentation in male American Goldfinches {Carduelis tristis) and Northern Cardinals (Ccirdinalis cardinalis). Functional Ecology 15:732-739. McGraw, K. J., G. E. Hill, R. Stradi, and R. S. Parker. 2001. The influence of carotenoid acquisition and utilization on the maintenance of species-typical plumage pigmentation in male American Goldfinches (Carduelis rristis) and Northern Cardinals (Cardinalis cardinalis). Physiological and Biochemical Zoology 74: 843-852. M0LLER, A. P., C. Biard, j. D. Blount, D. C. Houston, P. Ninni, N. Saino, and P. F. Sural 2000. Carotenoid dependent signals: indicators of foraging efficiency, immunocompetence or detoxification ability? Avian Poultry Science Review 1 1:137-159. Mulvihill, R. S., K. C. Parkes, R. C. Leberman, and D. S. Wood. 1992. Evidence supporting a dietary basis for orange-tipped rectrices in the Cedar Waxwing. Journal of Field Ornithology 63: 212-216. Quesada, j. and j. C. Senar. 2006. Comparing plumage colour measurements obtained directly from live birds and from collected feathers: the case of the Great Tit Pams major. Journal of Avian Biology 37: 609-616. Rodewald, a. D. and M. H. Bakermans. 2006. What is the appropriate paradigm for riparian forest conserva- tion? Biological Conservation 128: 193-200. Rodewald, A. D. and D. P. Shustack. 2008a. Consumer resource matching in urbanizing landscapes: are synanlhropic species over-matching? Ecology 89: 515-521. Rodewald, A. D. and D. P. Shu.stack. 2008b. Urban flight: understanding individual and population-level responses of Nearctic-neotropical migratory birds to urbanization. Journal of Animal Ecology 77: 83-91. Saks, L., 1. Ots, and P. Horak. 2003. Carotenoid-based plumage coloration of male Greenfinches reflects health and immunocompetence. Oecologica 134: 301-307. SAS Institute. 2002. SAS/STAT, Software Release 9.1. SAS Institute, Cary, North Carolina, USA. Senar, J. C. and J. Quesada. 2006. Absolute and relative signals: a comparison of melanin- and carotenoid- based patches. Behaviour 143: 589-595. Shochat, E. 2004. Credit or debit? Resource input changes population dynamics of city-slicker birds. Oikos 106:622-626. Sillanpaa, S., j. P. Salminen, E. Lehikoinen, E. Toivonen, and T. Eeva. 2008. Carotenoids in a food chain along a pollution gradient. Science of the Total Environment 406:247-255. SuRMACKl, A. AND J. K. NowAKOWSKi. 2007. Soil and preen waxes influence expression of carotenoid-based plumage coloration. Naturwissenschaften 94: 829- 835. Tratalos, j., R. a. Fuller, K. L. Evans, R. G. Davies, S. E. Newson, j. j. D. Greenwood, and K. J. Gaston. 2007. Bird densities are associated with household densities. Global Change Biology 13: 1685-1695. Wolfenbarger, L. L. 1996. Fitness effects associated with red coloration of male Northern Cardinals (Cardinalis cardinalis). Dissertation. Cornell University, Ithaca, New York, USA. Wolfenbarger, L. L. 1999a. Female mate choice in Northern Cardinals: is there a preference for redder males? Wilson Bulletin 111: 76-83. Wolfenbarger, L. L. 1999b. Is red coloration of male Northern Cardinals beneficial during the non-breeding season?: a test of status signaling. Condor 111: 655- 663. Wolfenbarger, L. L. 1999c. Red coloration of male Northern Cardinals correlates with mate quality and territory quality. Behavioral Ecology 10: 80-90. The Wilson Journal oj Ornithology 122(2);334-339, 2010 ANNUAL PRECIPITATION AFFECTS REPRODUCTION OF THE SOUTHERN GREY SHRIKE (LANIUS MERIDIONALIS) ODED KEYNAN' AND REUVEN YOSEP--^ ABSTRACT. — We studied the breeding ecology of the Southern Grey Shrike (Lanins mericlionalis) at the Shezaf Nature Reserve, Arava Valley, Israel, an extremely arid desert with mean annual rainfall of 30 mm. We color-banded 128 shrikes during 2007-2009. The breeding season lasted from late February until late June. We found 34 nests and a correlation between years with amount of precipitation and number of breeding pairs. Second broods after successful broods were found only in 2007 and were all unsuccessful. Most nests were in Acac ia trees; pairs that nested in dry trees, lacking foliage, were more likely to fail. The average clutch size was 3.44; average number of nestlings was 1.95; and average number of nestlings that fledged was 1.24. These results are lower than other studies of the species, probably a result of the scarcity of food in the arid environment. Nest survival rate was higher during incubation than during rearing of nestlings (0.71 vs. 0.62), and total nest survival was 44.5%. The arid Arava Valley influences breeding success of Southern Grey Shrikes, and the present .severe drought could negatively influence persistence of the shrike population in the Arava Valley. Received 6 September 2009. Accepted 23 November 2009. Rainfall is known to have profound effects on the life history of many bird species and different populations within species (Lloyd 1999, Morrison and Bolger 2002, Mondajem and Bamford 2009). Water is a strong limiting factor in arid ecosys- tems, where rainfall is generally low, erratic, and unpredictable (Noy-Meir 1973). This leads to low primary productivity and low food availability when combined with high temperatures (Tieleman et al. 2004). Birds that reproduce in arid ecosystems respond to these limiting factors by reducing clutch size (Lloyd 1999, Tieleman et al. 2004) or breeding earlier in the season or by not breeding at all in regions where conditions are particularly unfavorable (Immelmann 1973). Arid ecosystems are characterized by high between-year variations in both amount and duration of productivity due to the unpredictabil- ity of rains (Illera and Diaz 2006). Thus, individual birds that live in these environments must have a high phenotypic flexibility to survive and reproduce (Ricklefs and Wikelski 2002). Lloyd (1999) showed that >50% of the species in South Africa’s arid zones had inter-seasonal and inter-annual variations in clutch size due to changes in food availability. Tieleman et al. (2004), in their re,search on life history variation along an aridity gradient in larks, showed that nest predation increa.sed while both growth rate and clutch size decreased with increasing aridity. ' Department of Zoology, George S. Wise Faculty of Life •Sciences, Tel Aviv University. Israel. ^ International Birding and Research Center in Eilat, P. O. Box 774, Eilat 88000. Israel. •’Corresponding author; e-mail: ryoseftieilatcity. co.il The Southern Grey Shrike {Lanins meridiona- lis) is a passerine (Laniidae). Birds in this family are known for their similarity to diurnal raptors in morphology, especially the strong and sharp bill, and their predatory behavior (Yosef 2008). Shrikes do not posses strong talons, in contrast to raptors, and impaling behavior evolved in the family as a way to dismember prey (Smith 1973). The Southern Grey Shrike was recently sepa- rated from the Great Grey Shrike (L. excnhitor) (Lefranc and Worfolk 1997). It is widely distributed in the Middle East, North Africa, and central Asia (Lefranc and Worfolk 1997); in Israel, it is resident along the Dead Sea Ritt Valley (Shirihai 1996). There are few studies of its reproductive biology in extremely arid ecosys- tems (e.g., Yosef 1992, Budden and Wright 2000); more research is needed to understand the species’ response to extreme aridity and a fluctuating, scarce food base. Our objectives were to: ( 1 ) characterize the breeding ecology of the Southern Grey Shrike in an extremely arid environment, (2) compare it to previous studies on populations of the same species in other and similar environments, and (3) characterize the changes in the reproductive biology within our population following changes in precipitation between years. We hypothesized that, similar to Tieleman et al. (2004), we would find an effect of aridity on reproductive effort and success of the Southern Grey Shrike. We predicted that, because of the reduced prey base in desert areas, the number of clutches of a pair of shrikes in the season would be low due to exposure and lack of appropriate breeding sites. 334 Keymin and Yosef - BRFEDING OJ- SOUTHERN C3REY SHRIKES 335 METHODS Study Area. — ^The study was conducted at the Shezaf Nature Reserve during April 2006 to July 2009 in the southern part of Israel, —30 km south ot the Dead Sea in the Arava Rift Valley between the settlements of Hazeva (30° 46' N, 35° 16' E) and Ein Yahav (30° 45' N, 35° 15' E). The Arava Valley is an extremely arid zone (UNESCO 1977) with a mean annual winter rainfall of 35 mm that occurs in a range of 6-9 days with large annual variations in total rainfall and temporal and spatial distribution (Anava et al. 2000). The mean temperature during summer is 38° C with daily temperatures that can rise to 49° C (Goldreich and Kami 2001). The flora of the Shezaf Nature Re.serve is dominated by Acacia trees {Acacia tortilis and A. raddiana) and scattered shrubs (ZiUa spiuosa, Lyciiim shawii, and Haloxylon persicum) that occur only in the dry river beds (wadis). Field Procedures. — We color-banded 128 Shrikes for individual recognition. Eorty-nine (38.3%) adults were captured in a treadle trap (Yosef and Lohrer 1992) and 79 (61.7%) nestlings were banded at the nest 10-11 days post-hatching. All shrikes were accustomed to presence of human observers. Territories were mapped by watching the shrikes’ movements in their home ranges and by watching aggressive behavior towards other shrikes; temitory size was calculat- ed as the minimum polygon bounded by defended points of the habitat (Yosef and Grubb 1992). Nests were located during the breeding season and on each visit were checked to ascertain stage of the reproductive cycle (egg laying, incubation, nestling, fledging). We gathered data for 34 nests (23 in 2007, 1 1 in 2008) and calculated their probability of survival from incubation to fledging using Mayfield’s method (Mayfield 1961, 1975). The average duration of incubation and nestling days were from Yosef (1992). We recorded the height of the nest above ground, the species of tree in which the nest was built, and its condition (dry with no foliage vs. green). Means are expressed (± SD), but groups were compared using Mann Whitney-f/ or y- tests. We chose P < 0.05 as the minimum acceptable level of significance. RESULTS Breeding Season and Nesting Attempts. — ^The breeding season lasted from late Eebruary until late June with peak breeding activity in March (Fig. 1.). Three pairs had second attempts after successful first broods in 2007. None of the.se attempts fledged any young. None of the pairs in 2008 had second nesting attempts even after having successfully fledged young from the first brood (Table I). Three nesting attempts were made following nest failure, and one female nested four times with three different males, but only the fourth attempt was successful (Table 1 ). We found 34 nests during the two breeding seasons. There was a significant decline in nesting attempts between 2007 and 2008 (23 vs. 1 1 ; = 4.24, df = \, P = 0.04; Table 1). We found a significant correlation between number of breed- ing pairs and amount of precipitation between 2007 and 2009 (Spearman correlation /y = 1.00, P < 0.0001, Fig. 2). Nest Construction. — ^Thirty-three (97%) nests were in Acacia trees {Acacia raddiana and A. tortilis) and one (2.9%) was in a Commicarpus plumhagineus bush. Twenty-six (76.5%) nests were in green trees and eight (23.5%) in dry trees {y~ = 9.529, df = 1, = 0.002, n = 34). Nests in dry trees may have been more likely to fail than those in trees with green foliage (Mann Whitney- U z = —1.445, P — 0.09, n — 34). Average nest height was 1.86 ± 0.64 m (range 0.47-3.37). Nest height did not affect breeding success (Spearman correlation P = 0.49, n = 34 for clutch size, 0.43 for hatching success, 0.47 for fledging success) Clutch Size and Incubation. — The average clutch size was 3.44 ± 0.89 eggs per nest (2-5) for the two breeding seasons and did not differ between years (Mann Whitney-U z = —0.59, P = 0.59, n = 28; Table 1 ). Only females brooded clutches while males foraged and brought food to the female at or near the nest. Males also guarded the area sim'ounding the nest tree. Hatching Success and Nestlings. — ^The average number of eggs that hatched per nest was 2.14 ± 1.28 (0-4) for all nests and did not differ significantly between years. Nestling success per nest, for pairs that fledged at least one young per nesting attempt, was 2.6 ± 1 .03 for the two breeding seasons and did not differ between years (Mann Whitney-U z = — 1.2, P = 0.23, n = 29) (Table 1). At least one nestling fledged from 20 (58.8%) nests during the entire study. The average number of fledglings per nest was 1.53 ± 1.48 (0-4) and did not differ between years (Mann Whitney-U z = —0.16, P = 0.89, /? = 17 nests). 336 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2, June 2010 FIG. 1. Average reproductive activity (nests/month) of Southern Grey Shrikes for 2007-2009 in the Shezat Nature Reserve, Arava Valley, Israel. Fledglings. — Young remained in the nest tree for 2-5 days post-fledging and both parents cared for them. The young remained with the parents for up to 6 weeks after fledging. Only five (6.3%) of 79 banded nestlings remained in the study area after leaving their parents, and only one (1.3%) fledgling from 2007 bred in the study area in 2008. Both parents cared for the fledglings in the first 3 days after they started to follow the parents in the territory but subsequently the parents separated. In three ca.ses (17.6%), females stopped caring for fledglings and started incubating a second clutch. In six other cases (35.4%), females left the breeding area and left the males to care tor the fledglings. In five cases (29.4%) both male and female cared for the fledglings for the entire post-fledging period and, in three cases (17.6%), all fledglings were lost after a few days and the parents separated. Nest Survival Rate. — Six of 34 nests watched during the incubation period failed while being exposed for 298.5 days for a daily mortality rate of 2% per nest day. Daily nest survival rate was 98%. The survival rate for the entire period was 71%. Seventy-three (74.5%) of 98 eggs laid in 28 nests that succeeded in hatching at least one egg, hatched. Eleven (39.2%) of 28 nests watched during the nestling period failed while being exposed for 408.5 days for a daily mortality rate of 2.3% per nest day. Daily nest survival rate was 97.3%. The survival rate for the entire period was 62.5%. Probability of survival of a nest from start of incubation until the young fledged was 0.71 X 0.745 X 0.625 = 0.329 (i.e., the probability that TABLE 1. Between season variances (.v ± SD) in the nesting ecology ot the Southern Grey Shrike in Shezaf Nature Reserve, Israel. Breeding parameter Breeding pairs Second brood due to failure Second brood after success Breeding attempts Average clutch size Average nestlings per nest Fledging success 2007 16 4 3 23 3.52 ± 0.89 2..50 ± 0.78 1.17 ± 1.30 2008 Statistic.s 9 2 0 1 1 3.27 ± 0.90 2.8 ± 1.03 1.36 ± 1.39 f „ = 34, f = 1,96. df = \. P = 0.16 Insufficient data Insufficient data ,j = 34, f = 4.24, df = \, P = 0.04 Mann Whitney-(7 n = 28, z = -0.59, P = 0.59 Mann Whitney-U /; = 29 z = — 1 .2, P = 0.23 Mann Whitney-U n = M,?.. = —0.16, P = 0.89 Kcynaii am/ YoscJ • BRCHDING ()!■ SOUTHERN GREY SHRIKES 337 FIG. 2. Number of breeding pairs of Southern Grey Shrikes in Shezaf Nature Reserve, Arava Valley, Israel as a function of annual precipitation (line) between 2007 and 2009, any nest would fledge at least one young was 32.9%). DISCUSSION Our study evaluated the hypothesis that we would find an effect of aridity on reproductive effort and success of the Southern Grey Shrike. Our results are consistent with our predictions that, because of the reduced prey base in deserts, the number of clutches laid by a pair in the season would be low because of exposure and lack of appropriate breeding sites. However, a longer and more detailed study is required to understand the ecological constraints that influence Southern Grey Shrike breeding capability in the desert and their ability to persist in this harsh environ- ment. The results of the 2007, 2008, and 2009 breeding seasons reflect the combined impact of arid habitats and multi-year drought on the breeding biology of the Southern Grey Shrike at the Shezaf Nature Reserve. Nest survival rate was higher in our study during the incubation period than during the nestling period. This supports our initial prediction that harsh conditions in the Arava affect the breeding success of shrikes. Differences between the incubation and brooding periods can be explained by the need for more food during rearing of the nestlings, a need that was harder to achieve the closer the season was to the summer when prey availability decreases greatly in the desert. We found no between-year variations in start or length of the breeding season. These findings suggest that, unlike many other avian species. Southern Grey Shrikes do not rely on rain as a stimulus for the onset or end of the breeding season. Differences among regional studies in length of the breeding period (Table 2) also cannot be explained by rainfall. This conclusion does not agree with Illera and Diaz (2006) and Mondajem and Bamford (2009), who suggested that rainfall was the cue that stimulated repro- ductive activities in birds living in unpredictable environments. Lloyd (1999), however, found rain as a stimulus for breeding activity for only some of the birds of the South African arid zones. We believe Southern Grey Shrikes do not use rainfall as a stimulus for timing and duration of the breeding season because of the wide range of habitats they occupy. Southern Grey Shrikes occur in many habitats that differ in their predictability. Thus, rainfall cannot be a reliable cue for the onset of the breeding season. The main between-year differences in our study were the number of breeding pairs at the study site and their breeding attempts. Some of the pairs in 2007 attempted second broods after success but 338 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122. No. 2, June 2010 TABLE 2. Life history variances (.v ± SD) in the current study in comparison to other reproductive studies of Southern Grey Shrike. Current study Budden and Wrislit 2000 Inbar 1995 Yosef 1992 Campos el al. 2000 Lepley et al. 2000 Location Hazeva. Israel Hazeva. Lsrael Ml. Hernion. Israel Negev Desert, Israel Northern Spain Caru-Seche. France Avg. rainfall (mm) 30 30 1,500 90 325 >500 Breeding season Feb-Jun Feb-Jun Unavailable Jan-Aug Mar-Jul Mar-Jun Max. # of broods 2 2 Unavailable 4 1 1 Avg. clutch size 3.44 ± 0.89 3.9 ± 0.57 4.6 ± 0.5 5.8 ± 0.6 5.7 ± 0.4 5.2 ± 1.1 Avg. hatchlings/nest 2.14 ± 1.28 1.09 Unavailable 4.7 ± 2.1 2.1 ± 2.4 3.1 Avg. fledglings/nest 1.53 ± 1.48 Unavailable Unavailable 3.6 ± 1.4 Unavailable 1.54 ± 1.9 all attempts failed. In contrast, none of the pairs in 2008 and 2009 had any signs of trying a second brood. This difference indicates that number of breeding attempts was most probably related to food abundance. Bird species that Hedge more than one brood decrease in breeding success as the breeding season advances (Verhulst and Nilsson 2008). This decrease in breeding success occurs mainly because of climatic factors that affect food availability (Takagi 2001, Verhulst and Nilsson 2008, Yosef and Zduniak 2008, Mondajem and Bamford 2009). It is possible that food was so scarce during 2008 that the pairs could invest in only one brood. Southern Grey Shrikes can fledge up to four broods in less arid environments, where food is more abundant and the duration of the breeding season is longer (Table 2). Far fewer pairs live and breed in the Arava Valley than in the Sede Boker area (Shirihai 1996). We suggest that conditions in the Arava are marginal for the Southern Grey Shrike and are possibly a limiting factor to its distribution. The severe droughts in the past few years have affected the numbers of breeding pairs in the Arava (Fig. 2). The number of pairs declined sharply between years (Table 1 ), reaching a low of three in 2009 (Fig. 2), and it appears shrikes a.ssess habitat quality before they initiate breeding (Yosef 2008). It is possible that some relocate to other breeding habitats if food availability is low. It is also possible that some individuals that nested during 2007 did not survive the harsh summer or were unable to continue to defend a good territory due to drought. George et al. (1992) reported that drought .severely affected total bird density and breeding success. They showed that some species recov- ered numbers in the year alter the drought but others did not. Bolger et al. (2005) found that an extreme drought dramatically reduced both the number of pairs breeding and breeding success in four species of passerine birds in southern California. The Southern Grey Shrike population in Shezaf Nature Reserve has declined drastically in the past 3 years, due to a severe drought. It will be of interest to ascertain whether the population can persist and will recover in the future if there are good rains. ACKNOWLEDGMENTS We thank Amotz Zahavi, Ido Izhaki. Moshe Inbar, Ronni Ostreicher, and the staff of Hazeva Field Study Center; and Harel Ben-Shachar of the Israel Nature Re.serves and Parks Authority. We thank Tom Cade and an anonymous reviewer for improving an earlier draft of the manuscript. LITERATURE CITED Anava, a.. M. Kam, a. Shkolnik, and A. A. Degen. 2000. Seasonal field metabolic rate and dietary intake in Arabian Babblers (Twdoides squamiceps) inhabiting extreme desert. Functional Ecology 14: 607-613. Bolger. D. T., M. A. Patten, and D. C. Bo.stock. 2005. Avian reproductive failure in response to an extreme climatic event. Oecologia 142: 398^06. Budden, a. and J. Wright. 2000. Nestling diet, chick growth and breeding success in the Southern Grey Shrike {Lanius nicridionalis). The Ring 22: 165-172. Campos, F., F. Guiterrez-Corchero, and M. A. Her- nandez. 2007. Fenologia reproductora y xito repro- ductor del alcaud: n real, Lunins meridiomdis, en zonas agricolas del Norte de Espania. Ecologia 21 : 167-174. George, T. L., A. C. Fowler. R. L. Knight, and L. C. McEwen. 1992. Impact of severe drought on grassland birds in western North Dakota. Ecological Applica- tions 2: 275-284 CiOi.DREiCH. Y. AND O. Karni. 2001. Climate' and precipitation regime in the Arava Valley, Israel. Israel Journal of Earth Sciences 50: 53-59. Ii.lera, J. C. and M. Diaz. 2006. Reproduction in an endemic bird of a semi arid island: a food mediated process. Journal of Avian Biology 37: 447—456. Keyiniii and Yosef - BREEDING OE SOUTHERN GREY SHRIKES 339 Immelmann, K, 1973. Role of the environment in reproduction as source ol "predictive” information. Pages 121-147 in Breeding biology of birds (D. S. Earner, Editor). National Academy of Sciences, Washington, D.C., USA. Inbar, R. 199.3. Shrike nesting in Mount Hermon, Israel. Proceedings Western Eoundation of Vertebrate Zool- ogy 6: 215-217. Lefranc, N. and T. Worfolk. 1997. Southern Grey Shrike. Pages 140-144 in A guide to the shrikes of the world. Yale University Press, New Haven, Connecti- cut, USA. LEPLEY, M., C. P. GUtLLAUME, A. NEWTON, AND M. TttEVENOT. 2000. Breeding biology of Southern Grey Shrike {Lanins meridionalis) in Crau-Seche (Bouches du Rhone, Prance). Alauda 68: 35—43. Lloyd, P. 1999. Rainfall as a breeding stimulus and clutch size determinant in South African arid-zone birds. Ibis 141: 637-643. MAYFtELD, H. 1961. Nesting success calculated from exposure. Wilson Bulletin 73: 255-261. Mayfield, H. 1975. Suggestions for calculating nest success. Wilson Bulletin 87: 456^66. Mondajem, a. and a. Bamford. 2009. Influence of rainfall on timing and success of reproduction in Marabu Storks Lentoplilos crumeniferus. Ibis 151: 344-351. Morrison, S. A. and D. T. Bolger. 2002. Variation in a sparrow’s reproductive success with rainfall: food and predator-mediated processes. Oecologia 133: 315-324. Noy-Meir. 1. 1973. Desert ecosystems: environment and producers. Annual Review of Ecology and Systematics 4: 25-5 1 . Ricklefs, R. E. and M. Wikelski. 2002. The physiology/ life history nexus. Trends in Ecology and Evolution 17: 462^68. Shirihai, H. 1996. The birds of ksrael. Academic Press, London, United Kingdom. Smi th, S. M. 1973. Eood manipulalion by young passerines and the possible evolutionary history of impaling by shrikes. Wilson Bulletin 85: 318-322. Takagi, M, 2001. Some effect of inclement weather conditions on the survival and condition of Bull- headed Shrike nestlings. Ecological Research 16: 55- 63. TIELEMAN, B. I., J. B. WtLLIAMS, AND G. H. Vl.SSER. 2004. Energy and water budgets of larks in a life history perspective: parental effort varies with aridity. Ecol- ogy 85: 1399-1410. UNESCO. 1977. Map of world distribution of arid lands. MAB Technician Note 7. UNESCO, Paris, Prance. Verhulst, S. AND J.-A. Nilsson. 2008. The timing of bird’s breeding seasons: a review of experiments that manipulated timing of breeding. Philosophical Trans- actions of the Royal Society London, Series B 363: 399-410. Yosef, R. 1992. Prom nest building to fledging of young in Great Grey Shrikes (Lanins excubitor) at Sede Boqer, Israel. Journal of Ornithology 133: 279-285. Yosef, R. 2008. Laniidae. Pages 732-796 in Handbook of the birds of the world. Volume 13 (J. del Hoyo, A. Elliott, and D. A. Christie, Editors). Lynx Edicions, Barcelona, Spain. Yosef, R. and T. C. Grubb Jr, 1992. Territory size influences nutritional condition in nonbreeding Log- gerhead Shrikes {Lanins Indovicianns) a ptilochronol- ogy approach. Conservation Biology 6: 447-449. Yosef, R. and P. E. Lohrer. 1992. A composite treadle/ Bal-Chatri trap for Loggerhead Shrikes. Wildlife Society Bulletin 20: 116-118. Yosef, R. and P. Zduniak. 2008. Variation in clutch size, egg size variability and reproductive output in the Desert Pinch (Rhodospiza obsolete). Journal of Arid Environments 72: 1631-1635. The Wilson Journal of Ornithology 122(2):340-345, 2010 HOME RANGE SIZES AND HABITAT USE OE NELSON’S AND SALTMARSH SPARROWS W. GREGORY SHRIVERT^' THOMAS P. HODGMAN,- JAMES P. GIBBS,' AND PETER D. VICKERY' ABSTRACT. — Nelson’s (Ammodramus nelsoni) and Saltmarsh (A. caudacutus) sparrows are sympatric breeders in tidal marshes of the southern Gulf of Maine. These sparrows hybridize, have different mating strategies, and males do not defend territories or provide parental care. We estimated and compared core area sizes, home range sizes, and habitat u.se between species and between males and females. We radio-marked 140 sparrows (63 Nelson’s and 77 Saltmarsh sparrows) during three breeding seasons (1999-2001) at Scarborough Marsh, Maine, USA. Home ranges of male A. nelsoni were 2.3 times larger (± SE) (1 19.68 ± 19.43 ha) than those of male A. caiidacutiis (52.85 ± 8.68 ha). Home range sizes of female Nelson’s and female Saltmarsh sparrows did not differ from each other (female Nelson’s home range = 43.58 ± 13.10 ha; female Saltmarsh home range = 27.8 1 ± 6.3 ha). More than 40% of male and 1 8% of female home ranges had two discrete core areas and, in most instances, each core area corresponded to a separate lunar cycle. We suggest that differences in mating strategies, densities, and adaptation to nesting in tidal marshes explain the larger home range estimates for male Nelson’s Sparrows. Female and male Nelson’s Sparrows' home ranges had more Spartina alterniflora cover and female Saltmarsh Sparrows’ home ranges had greater Juncus gerardii cover than random locations. Home ranges of female Saltmarsh Sparrows had less Spartina alterniflora cover and more Juncus gerardii cover than female Nelson’s Sparrows. We did not detect any differences in vegetation variables between male Saltmarsh and male Nelson's span'ow home ranges. Received 1 1 September 2009. Accepted 10 January 2010. Home range and resource use can occur at multiple spatial and temporal scales that may change during the annual cycle for most bird species. Movement patterns and space require- ments are important parameters to consider when attempting to understand populations and how they might respond to different management actions or habitat change scenarios (Barbour and Litvaitis 1993). Home range studies can provide valuable information about habitat selection, use of space, and in making predictions about how environmental change can alter a species’ avail- able habitat (White and Garrott 1990). Nelson’s {Antmoclrcimns nelsoni previously known as Nelson’s Sharp-tailed Sparrow [AOU 1998, Chesser et al. 2009] ) and Saltmarsh (A. caiulacutus previously known as Saltmarsh Sharp- tailed Sparrow [AOU 1998, Chesser et al. 2009]) sparrows are listed as conservation priorities due to our limited understanding of their basic ' .State University of New York. College of Environmen- tal Science and Forestry, I Forestry Drive, Syracuse, NY 13210. USA. ^Bird Group. Maine Department of Inland Fisheries and Wildlife, 650 State Street. Bangor, ME 04401, USA. 'Center for Ecological Research, P. O. Box 127, Richmond. ME 04357, USA. ■‘Current address: Department of Entomology and Wildlife Ecology. University of Delaware, Newark, DE 19717, USA. ^Corresponding author; e-mail: gshriver@udel.edu demographic parameters and habitat requirements (Rich et al. 2004). Previous re.search into the breeding ecology of these sparrows has described responses to landscape features (Benoit and Askins 1999, Shriver et al. 2004), dispersal patterns and survival rates (DiQuinzio et al. 2001), and second (Nocera et al. 2007) and third (Woolfenden 1956, Greenlaw and Rising 1994) order habitat selection (Johnson 1980). These studies have increased our understanding of the breeding ecology for these taxa from Long Island, New York, USA (Woolfenden 1956, Greenlaw and Rising 1994) to the Canadian Maritimes (Nocera et al. 2007), supporting the species split (AOU 1995). These aid in making predictions about how changes to available habitat may affect the persistence of populations of these .species. Our objectives were to; (1) use radiotelemetry to estimate species- and male/female-specific home range size, and (2) resource use to further our understanding of how these non-territorial and promi.scuous (Greenlaw and Rising 1994) species use their breeding habitat. METHODS We used mist nests to capture sparrows at Scarborough Marsh, Maine, USA (43° 34' N, 70° 22' W) during three breeding seasons (20 May-15 Aug) from 1999 to 2001. All sparrows were individually marked with a U.S. Geological Survey band and 3 color bands. We conducted 340 S/irivcr el a/. • NELSON'S AND SALTMARSIl SPARROW HOME RANGE SIZES 341 our study on a lOO-ha portion of this l,5()()-ha salt marsh. Spurt ina patens, S. alterniflora. Distich I is spicata, Jiinciis gerartlii, and salt pans (depres- sions with little or no vegetation and generally high salinity levels) dominated this area of Scarborough Marsh. We used radiotelemetry to relocate individuals in the marsh by homing during each of three breeding seasons. We attached 1-g radio trans- mitters {<5% body mass, Holohil Systems Ltd., Caip, ON, Canada) to adult males and females of both species using the thigh-harness method (Rappole and Tipton 1991). We assessed the fit of transmitters using a 30 X 25 X 25 cm, slotted, clear-plastic cage to view span'ows after trans- mitter attachment; if a harness and transmitter did not appear to fit properly, we removed it and released the sparrow. We attempted to locate radio-marked birds each day by homing using a receiver (Communication Specialists Inc., Orange, CA, USA) and 4-element, hand-held antenna. We recaptured as many birds as possible at the end of the season to remove transmitters. The surface complexity and zonation of the marsh vegetation in our study area facilitated accurate plotting of sparrow locations. We used a high resolution, color-infrared, aerial photograph to record locations of individual spaiTows each field day until batteries in transmitters failed or transmitters were removed. This geo-referenced photograph allowed us to plot locations of sparrows to within 3 m. We overlaid a transpar- ency on the photograph each field day, and marked date, sparrow locations, transmitter fre- quency, band number, species, and male or female on the overlay. We examined the dominant vegetation within a 1-m-radius circle of each sparrow location and categorized vegetation as: (1) Spartina alterniflora, (2) S. patens, (3) S. patens! S. alterniflora, (4) Jnncns gerarclii, (5) Distichlis spicatalS. patens, (6) un-vegetated salt pan, or (7) vegetated salt pan. We used the proportion of each vegetation type across all locations for a specific sparrow in the habitat use analysis. We also sampled 174 randomly located 1-m-radius circular plots to estimate the avail- ability of each vegetation type within the study area. We used the geo-referenced photograph as the base layer to digitize these data using ArcView 3.3 and used the animal movement extension (Hooge and Eichenlaub 1997) to estimate 50% (hereafter, core area) and 95% (hereafter, home range) fixed kernel home ranges. We used only those individuals with >20 locations for all analyses. We used univariate ANOVA to test for differences in core area and home range sizes between species and males/ females (Zar 1999). We calculated the number of sparrows that had two di.screte core areas (i.e., two 50% kernel home ranges) and measured the distance between the centroids for each core area. We used univariate ANOVA with Tukey HSD post-hoc tests to explore differences in vegetative cover between randomly located points, and locations of males and females for each species, and between the two species. We used a Kolmogorov-Smirnov goodness of fit test to examine if each variable fit a normal distribution (Zar 1999). All statistical tests were conducted in SPSS (Version 16.0) with alpha = 0.05. RESULTS We estimated core area and home range size for 63 radio-marked Nelson’s and 77 radio-marked Saltmarsh spaiTows at Scarborough Marsh, Maine in 1 999-200 1 . Core area and home range sizes of males were larger than for females, both for Nelson’s (core area: F|_62 = 8.5, P = 0.005; home range: Fi,e2 — ll-C P = 0.001) and Saltmarsh sparrows (core area: F\jf, = 4.2, P = 0.045; home range: F] 75 = 5.7, P = 0.019) (Table 1). Both core area and home range sizes of male Nelson’s Sparrows were larger than those of male Salt- marsh Sparrows (core area: F| 51 = 4.3, P = 0.043; home range: F1.51 = 10.6, P = 0.002) (Table 1). We found no difference in core area or home range sizes of female Nelson’s and female Saltmarsh sparrows (core area: F| 77 = 0.18, P = 0.674; home range: F\J■^ = 1.1, P — 0.289) (Table 1). Distances between core areas, for those individuals that maintained two core areas, were greater for male Nelson’s than for male Saltmarsh spaiTows (F| 54 = 4.5, P = 0.038), and were greater for female Nelson’s compared to female Saltmarsh sparrows (F| 34 = 8.0, P = 0.008) (Table 1). Female and male Nelson’s Sparrows' home ranges had more Spartina alterniflora cover than random locations (F| 237 = 16.2, P < 0.001) but did not differ between each other (F = 0.142). The amount of S. patens cover differed among Nelson’s Sparrows and random locations (F| 237 = 6.7, P = 0.002) with less S. patens in the home ranges of male Nelson’s Sparrows (F < 0.001, Table 2). Spartina patenslS. alterniflora cover differed among Nelson's Span'ows and random 342 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2, June 2010 TABLE 1 . Characteristics (.v ± SE) of radio-marked Nelson’s and Saltmarsh sparrows fixed kernel 50% core areas and 95% home ranges at Scarborough Marsh, Maine, USA, 1999 to 2001. Nelson’s Sparrow Saltmarsh Sparrow Variable Females Ui = 34) Males (« = 29) Females (n = 44) Males (H = 33) Core area, ha Home range, ha Home ranges with two cores Core area distance, m 6.15 ± 1.66 43.58 ± 13.10 9 (27%) 161.40 ± 38.85 18.42 ± 3.91 119.68 ± 19.4 13 (45%) 204.28 ± 25.61 5.25 ± 1.38 27.81 ± 6.30 8 (18%) 77.56 ±13.16 9.64 ± 1.67 52.85 ± 8.68 14 (42%) 158.99 ± 17.55 locations {F^ 2^l — 5.2, P = 0.006) with less Spartina patens! S. alterniflora in male Nelson’s Sparrow home ranges (P = 0.018, Table 2). Both male and female Nelson’s Sparrow home ranges had more vegetated pan cover than random locations (^1.237 = 38.1, P < 0.001, Table 2). Spartina alterniflora cover differed among Saltmarsh Sparrow home ranges and random locations {F\,2S2 — 8.4, P < 0.001, Table 2) with greater cover in male Saltmarsh Sparrow home ranges than in random locations {P < 0.001). Spartina patens cover differed among Saltmarsh Sparrow home ranges and random locations (F\.252 — 6.6, P = 0.002, Table 2) with less cover in home ranges of male Saltmarsh Sparrows than in random locations (P = 0.004). Spartina patensIS. alterniflora cover differed among Salt- marsh Sparrow home ranges and random locations (^1.252 = 5.5, P — 0.005, Table 2) with less cover in home ranges of male (P = 0.039) and female {P = 0.004) Saltmarsh Sparrows than in random locations (Table 2). Juncus gerardii cover dif- fered among Saltmarsh Sparrow home ranges and random locations (^1252 = 7.1, P = 0.001, Table 2) with more cover in female Saltmarsh Sparrow home ranges than in random locations [P = 0.001, Table 2). Both male and female Salt- marsh Sparrow home ranges had more vegetated pan cover than random locations (^1,237 = 48.2, P < 0.001, Table 2). Female Saltmarsh Sparrow home ranges had more Spartina alterniflora cover (7^1,79 = 4.5, P = 0.038) and more Juncus gerardii cover (F] 79 = 4.6, P = 0.035) than female Nelson’s Span'ow home ranges (Table 2). We did not detect any differences in vegetation variables between home ranges of male Saltmarsh and male Nelson’s span'ows (Table 2). DISCUSSION Home range sizes of both species were 2-10 times larger than those reported for other emberizid span'ows (Patterson and Petrinovich 1978, Reed 1985, Delany et al. 1995, Bechtoldt and Stouffer 2005, Jones and Bock 2005, Cox and Jones 2007). Male/female-specific home range size was approximately twice as large for Nelson’s than for Saltmarsh sparrows. Male Nelson’s Sparrow home range sizes were three times larger than those of females. All home ranges for both species and males and females overlapped within the 100-ha study area. Home range sizes for male Saltmarsh Spairows were twice those of females, but both were smaller than home ranges of male Nelson’s SpaiTOws. Wool- fenden (1956) estimated male Saltmarsh Sparrow home ranges to be 3^ times larger ( 1 .2-1.6 ha for TABLE 2. Proportion ± SE of .seven vegetation types within Nelson’s and Saltmarsh sparrow home ranges and at random points at Scarborough Marsh. Maine, 1999-2001. Nelson's Sparrow Saltmarsh Sparrow Vegetation type Females Males Females Males Random Spartina alterniflora S. patens S. patensIS. alterniflora .innens gerardii Disticlilis spicata Un-vegetated Pan Vegetated Pan 0.188 ± 0.021 0.276 ± 0.023 0.104 ± 0.02! 0. 1 86 ± 0.025 0.012 ± 0.007 0.005 ± 0.002 0.188 ± 0.021 0.288 ± 0.041 0.178 ± 0.021 0.044 ± 0.010 0.214 ± 0.030 0.006 ± 0.005 0.013 ± 0.005 0.203 ± 0.032 0.132 ± 0.017 0.288 ± 0.029 0.087 ± 0.013 0.283 ± 0.035 0.006 ± 0.003 0.004 ± 0.001 0.155 ± 0.021 0.216 ± 0.032 0.201 ± 0.025 0.081 ± 0.029 0.216 ± 0.042 0.001 ± 0.001 0.015 ± 0.006 0.235 ± 0.030 0.069 ± 0.017 0.435 ± 0.031 0.239 ± 0.030 0.1 15 ± 0.022 0.008 ± 0.006 0.041 ± 0.014 0.019 ± 0.009 S/irivcr et al. • NELSON'S AND SALTMARSII SPARROW HOME RANGE SIZES 343 males vs. 0.4 ha for females) than those of females. Greenlaw and Rising (1994) also report- ed a 4-fold difference in male (4.3 ha) versus female (1.1 ha) home range size from studies of Saltmarsh Spanows. It is difficult to compare our findings with other home range estimates given our use of radiotelemetry. The non-territorial mating system of these spaiTows (Woolfenden 1956, MuiTay 1969, Post and Greenlaw 1982, Greenlaw 1993, Shriver et al. 2007) generally results in more time spent by males searching for receptive females or females soliciting copulatory chases (Greenlaw and Rising 1994, Shriver et al. 2007). This, coupled with a lack of territorial boundaries, partially explains the large home ranges observed for both taxa in our study. Male Nelson’s SpaiTows had 33% larger home ranges than male Saltmarsh Sparrows which may be explained by differences in mating systems and breeding densities between the two .species. Male Nelson’s Sparrows spend a substantial amount of time guarding females (up to 43 hrs with the same female, Shriver et al. 2007) while male Saltmarsh Sparrows practice “scramble competition polyg- yny’’ (Greenlaw and Rising 1994) where small groups (3-6) of males chase receptive females, copulate, and then usually perch to search for another receptive female (W. G. Shriver et al., unpubl. data). Male Saltmarsh Sparrows have not been observed guarding females (Shriver et al. 2007). Twenty-seven percent of female Nelson’s Sparrows maintained two distinct core areas and the distance between these core areas was two times greater than for female Saltmarsh Sparrows. This may cause male Nelson’s Sparrows to travel farther to encounter females compared to male Saltmarsh Sparrows. We estimated there were three times more female Saltmarsh Sparrows than female Nelson’s Sparrows (1998 to 2001; W. G. Shriver et al, unpubl. data) in our study area. We suggest, based on our banding data from this site, greater dispersion of female Nelson’s Sparrows (especially in lower density areas), greater distances between female Nelson’s Sparrows' core areas, and the mate guarding behavior of male Nelson’s Sparrows, at least partially explain differences in home range sizes for these species. Other aspects of sharp-tailed sparrow breeding ecology appear to be synchronous with new moon tidal events (Shriver et al. 2007) and we reviewed the timing of the maintenance of each core area. More than half of the sparrows with two discrete core areas did not use them simultaneously, but instead core area use was timed with separate lunar cycles. Most nests typically flood during spring tides (Gjerdrum et al. 2005, Shriver et al. 2007), and it is apparent that re-nesting female sparrows moved within the study area to construct a second nest (during the subsequent lunar event). The large ranges (and multiple core areas) of males are likely a consequence of spatial shifts exhibited by females. Female Nelson’s Spairows do not re-nest immediately after a Hooding event (Shriver et al. 2007), and appear to move further to re-nest than female Saltmarsh Sparrows. Home ranges of male Nelson’s Sparrows, therefore, are larger and distance between cores further as they attempt to maintain close proximity to females during re-nesting attempts. Patterns in habitat use for these sparrows differed from studies of nest site selection (DiQuinzio et al. 2002, Gjerdrum et al. 2005, Shriver et al. 2007). We found spaiTows more frequently in Spartina alteniiflora and J uncus gerardii cover and less often in S. patens cover than was available in the study area. Vegetated pan cover was also more common within sparrow home ranges than was available in the study area. Saltmarsh Sparrow female home ranges had the greatest proportion of Juncus gerardii compared to random locations or other sparrow home ranges. Gjerdrum et al. (2008) also found Juncus gerardii was the key predictor variable in explaining variation in number of nests and fledgling production. Tidal marshes are limited globally in extent to 45,000 km" with 15,500 km" (34%) occumng along the Atlantic and Gulf coasts of North America (Bertness 1999, Mendelssohn and Mckee 2000, Greenberg et al. 2006). The coastlines of North America support the greatest number of endemic tidal marsh vertebrate species, including Saltmarsh Sparrows (Greenberg et al. 2006), increasing the global significance of these eco- systems. Resource professionals engaged in tidal marsh conservation have reason to be concerned about persistence of tidal marsh vertebrate .species given their geographically limited habitat and vulnerability to habitat loss and fragmentation resulting from sea level rise. Even moderate estimates (30 cm rise) of .sea level rise had negative effects on the estimated persistence of Seaside Sparrows (Ainniodranius maritimus) in Connecticut (Shriver and Gibbs 2004). Our study was conducted in one of the larger marshes in the Gulf of Maine where these species co-occur, near 344 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 2. June 2010 the northern limit of Saltmarsh Sparrows and the southern limit of Nelson’s Sparrows. These conditions certainly influence the spatial use patterns of these sparrows and our estimates of home range size for these taxa may not apply to smaller sites where these sparrows are relatively common and allopatric. Future home range studies at other sites using similar techniques would be valuable to examine the variation in these parameters given different marsh sizes and location within the range of each species. Space requirements and habitat use information similar to those pre.sented here and elsewhere (Gjerdrum et al. 2008) could then be used to model the effects of climate change on these species which would aid in conservation planning efforts. ACKNOWLEDGMENTS We thank J. J. Bacon, M. J. Baumflek, G. C. Hall, C. N. Jacques, B. J. MacCulloch, A. B. Sacerdote, and M. J. McElroy for assistance in the field. Jan Taylor (USFWS, Region 5) provided early and continued support for this project. This research was funded in part by The Maine Outdoor Heritage Fund, Maine Department of Inland Fisheries and Wildlife Federal Aid in Sport Fish and Wildlife Restoration - Project W82R. U.S. Fish and Wildlife Service, and a National Fstuarine Research Reserve Graduate Fellowship awarded to WGS. J. S, Greenlaw, C. S. Flphick, J. J. Nocera, and P. W. C. Paton provided valuable comments on early versions of this manu-script and are warmly acknowledged. LITERATURE CITED American Ornithologists’ Union (AOU). 1995. Fortieth supplement to the American Ornithologists' Union checklist of North American birds. Auk 1 12:819-8.50. American Ornithologi.sts’ Union (AOU). 1998. Check- list of North American birds. Seventh Edition. American Ornithologi.sts’ Union, Washington, D.C., USA. Barbour. M. S. and J. A. Litvaitis. 1993. Niche dimensions of New England cottontails in relation to habitat patch size. Oecologia 95: 321-327. Bechtoldt, C. L. and P. C. Stouffer. 2005. Home-range size, response to fire, and habitat preferences of wintering Henslow’s Sparrows. Wilson Bulletin 117: 21 1-225. Benoit, L. K. and R. A. Askins. 1999. Impact of the spread of Phra^mitcs on the distribution of birds in Connecticut tidal marshes. Wetlands 19: 194-208. BERTNE.SS. M. D. 1999. The ecology of Atlantic shorelines. Sinauer Associates Inc., Sunderland, Massachu.setts, USA. Chf;,s.ser. R. T., R. C. Bank.s, F. K. Barker, C. Cicero. J. L. Dunn. A. W. Kratter, I. J. Lovette. P. C. Ra.smu.s.sen. j. V. Rem.sen Jr., J. D. Rlsing, D. F. Stotz, and K. Winki;r. 2009. Fiftieth Supplement to the American Ornithologists’ Union checklist of North American birds. Auk 126:705-714. Cox, J. A. AND C. D. Jones. 2007. Home range and survival characteristics of male Bachman’s Sparrows in an old- growth forest managed with breeding season burns. Journal of Field Ornithology 78:263-269. Delany, M. F., C. T. Moore, and J. D. R. Progulske. 1995. Territory size and movement of Florida Grasshopper Sparrows. Journal of Field Ornithology 66: 305-309. DiQuinzio, D. A., P. W. C. Paton, and W. R. Eddleman. 2001. Site fidelity, philopatry, and survival of promiscuous Saltmarsh Sharp-tailed Sparrows in Rhode Island. Auk 1 18: 888-899. DiQuinzio, D. A., P. W. C. Paton, and W. R. Eddleman. 2002. Nesting ecology of Saltmarsh Sharp-tailed Sparrows in a tidally restricted .salt marsh. Wetlands 22: 179-185. Gjerdrum, C., C. S. Flphick, and M. Rubega. 2005. Nest site selection and nesting success in saltmarsh breeding sparrows: the importance of nest habitat, timing, and study site differences. Condor 107: 849- 862. Gjerdrum, C., C. S. Flphick, and M. A. Rubega. 2008. How well can we model numbers and productivity of Saltmarsh Sharp-tailed Sparrows (Ammodramus caii- dacutus) using habitat features? Auk 125: 608-617. Greenberg, R., J. E. Maldonado, S. Droege, and M. V. McDonald. 2006. Tidal marshes: a global perspective on the evolution and con.servation of their terrestrial vertebrates. BioScience 56: 675-685. Greenlaw, J. S. 1993. Behavioral and moiphological diversification in Sharp-tailed Sparrows (Ammodramus caiidaciilus) of the Atlantic Coast. Auk 1 10: 286-303. Greenlaw, J. S. and J. D. Rising. 1994. Sharp-tailed Sparrow (Ammodramus caiidaciitus). The birds of North America. Number 1 12. Hooge, P. N. and B. Eichenlaub. 1997. Animal movement extension to Areview. Version 2.0. USGS, Ala,ska Science Center-Biological Science Office, Anchorage, Alaska, USA. JOHN-SON, D. H. 1980. The comparison of usage and availability measurements for evaluating resource preference. Ecology 6 1 : 65-7 1 . Jones, Z. F. and C. E. Bock. 2005. The Botteri’s Sparrow and exotic Arizona grasslands: an ecological trap or habitat regained? Condor 107: 731-741. Mendelssohn, 1. A. and K. L. Mckee. 2000. Saltmarshes and mangroves. Pages 501-536 in North American terrestrial vegetation (M. Barbour and W. D. Billings, Editors). Cambridge University Press, . Cambridge, United Kingdom. Murray Jr., B. G. 1969. A comparative study of Le Conte’s and Sharp-tailed .sparrows. Auk 86: 199-231. Nocera, J. J., T. M. Fitzgerald, A. R. Hanson, and G. R. Milton. 2007. Differential habitat u.se by Acadian Nelson’s Sharp-tailed Sparrows: implications for regional con.servation. Journal of Field Ornithology 78: 50-55. PA iTitR.soN, T. L. AND L. Petrinovici I. 1978. Territory size in the White-crowned Sparrow (Zonotrichia leu- Shrivcr el al. • NELSON'S AND SALTMARSII SPARROW HOME RANGE SIZES 345 cophrys): measurement and stability. Condor 80: 97- 98. PcxsT, W. AND J. S. Greenlaw. 1982. Comparative costs of promiscuity and monogamy: a test of reproductive effort theory. Behavioral Ecology and Sociobiology 10: 101-107. Rappole, J. H. and A. R. Tipton. 1991. New harness design for attachment of radio transmitters to small passerines. Journal of Field Ornithology 62: 335-337. Reed, J. M. 1985. A comparison of the “flush” and spot- map methods for estimating the size of Vesper Spanow teiTitories. Journal of Field Ornithology 56: 131-137. Rich, T. D., C. J. Beardmore, FI. Berlanga, P. J. Blancher, M. S. W. Bradstreet, G. S. Butcher, D. W. Demarest, E. H. Dunn, W. C. FIunter, E. E. Inigo-Elias, j. A. Kennedy, A. M. Martell, A. O. Panjabi, D. N. Pashley, K. V. Rosenberg, C. M. Rustay, j. S. Wendt, and T. C. Will. 2004. Partners in Flight North American Landbird Conservation Plan. Cornell Laboratory of Ornithology, Ithaca, New York, USA. Shriver, W. G. and j. P. Gibbs. 2004. Projected effects of sea-level rise on the population viability of Seaside Sparrows (Ammodramus niciritimiis). Pages 397-410 in Species conservation and management: case studies (fj. R. Ak9akaya, M. Burgman, O. Kindvall, C. C. Wood, P. SJogren-Gulve, J. S. fJatfield, and M. A. McCarthy, Editors). Oxford University Press, Oxford, United Kingdom. Shriver, W. G., T. P. Hodgman, J. P. Gibbs, and P. D. Vickery. 2004. Landscape context influences salt marsh bird diversity and area requirements in New England. Biological Conservation 119: 545-553. Shriver, W. G., P. D. Vickery, T. P. FIodgman, and J. P. Gibbs. 2007. Flood tides affect breeding ecology of two sympatric sharp-tailed sparrows. Auk 124: 552- 560. White, G. C. and R. M. Garrott. 1990. Analysis of wildlife radio-tracking data. Academic Press, New York, USA. Woolfenden, G. E. 1956. Comparative breeding behavior of Ammospiza caudacuta and A. maritima. University of Kansas, Publications of the Museum of Natural History 10: 45-75. Zar, j. H. 1999. Biostatistical analysis. Fourth Edition. Prentice Hall, Upper Saddle River, New Jersey, USA. The Wiiso)! Journal of Ornithology 122(2):346-353, 2010 ASSESSING GENERALIZED EGG MIMICRY: A QUANTITATIVE COMPARISON OF EGGS OF BROWN-HEADED COWBIRDS AND GRASSLAND PASSERINES DWIGHT R. KLIPPENSTINE' - AND SPENCER G. SEALY'"' ABSTRACT. — Perceived similarity in appearance of eggs of the Brown-headed Cowbird (Molothrus ater) and several grassland passerine hosts has been suggested to represent a form of generalized mimicry evolved to impede egg discrimination. We tested this hypothesis by quantitatively comparing four parameters (ground color, maculation color, distribution of maculation, and maculation density) among cowbird eggs and those of six grassland passerines known to preferentially accept cowbird eggs over non-mimetic blue eggs. Cowbird eggs did not significantly differ in three of four parameters (mean ground color, and color and density of maculation) from the average grassland passerine egg measured for this community, in southern Saskatchewan, when all grassland passerine eggs were pooled. Cowbird eggs sufficiently overlapped grassland pas.serine eggs in these three parameters that 88% of cowbird eggs measured were statistically indistinguishable from among all grassland pas.serine eggs. The frequency at which cowbird eggs were misclassitied as eggs of each grassland passerine ranged from 8 to 48%, which suggests that a single cowbird egg is capable of mimicking the eggs of more than one species. These results support the hypothesis that cowbirds have evolved a generalized egg appearance that mimics the eggs of multiple grassland passerine hosts within a single community. Received 31 October 2008. Accepted 23 October 2009. Evolution of egg mimicry by obligate brood parasites is regarded as one of the most common responses to development of egg discrimination and ejection by intolerant hosts (Underwood and Sealy 2002, Kilner 2006). Egg mimicry evolves because ejection of unlike eggs by hosts provides a selective advantage to parasitic eggs that match the clutch in which they are laid (Davies and Brooke 1991, Soler et al. 2003). The extent of similarity gradually escalates through a co- evolutionary race as hosts refine their egg- discrimination abilities over subsequent genera- tions (Moksnes et al. 1990, Davies and Brooke 1991). The evolutionary equilibrium hypothesis posits that whenever the level of mimicry exceeds a species’ ability to discriminate between eggs, the host accepts the parasitic egg (Davies and Brooke 1988, Lotem and Nakamura 1998). Presumably, acceptance occurs becau.se of the greater cost of mistakenly ejecting the wrong egg (Davies and Brooke 1988, Lotem and Nakamura 1998). Egg mimicry has been documented or suspect- ed in several parasitic systems (Underwood and Sealy 2002). Most notably different groups of female Common Cuckoos (Ciiciiliis canonts) have evolved eggs that closely mimic those of a ' Department of Biological .Scienees, University of Manitoba, Winnipeg, MB R3T 2N2. Canada. ^Current address: Faculty of Medicine. University of Manitoba. Winnipeg, MB R3E 3P.3, Canada. ’ Corresponding author: e-mail: sgsealy@cc.umanitoba.ca preferred host (Davies and Brooke 1988, March- etti et al. 1998, Gibbs et al. 2000). The Brown- headed Cowbird {Molothrus ater, hereafter cow- bird) has been largely exempt from these speculations (Sealy et al. 2002, Mermoz and Ornelas 2004) because individual female cow- birds are known to parasitize multiple species (Fleischer 1985, Alderson et al. 1999). Rothstein (2001) suggested cowbird egg mimicry could not evolve due to the immense variability in egg appearance each female cowbird would encounter when parasitizing several species. He also argued that because cowbirds are known to parasitize hundreds of species it would be more beneficial to avoid hosts that discriminate among eggs than to enter a costly co-evolutionary race. A recent comparison of tolerant and intolerant hosts of the cowbird also found that egg appearance was not correlated with acceptance of cowbird eggs (Peer and Sealy 2004). The close resemblance in color and maculation pattern between cowbird eggs and those of many grassland passerines, despite skepticism, has been suggested to be a form of generalized egg mimicry (Peer et al. 2000, Davis et al. 2002, Frontispiece in Klippenstine and Sealy 2008). Generalized cowbird egg mimicry is the idea that cowbird eggs have been adapted to match the appearance of eggs laid by multiple hosts. The eggs of a parasite in traditional mimicry evolve to resemble those of a single, intolerant host species (Underwood and Sealy 2002). Generalized egg mimicry, however, incorporates the selective 346 K/ippcnsfilic and Scaly • COMPARISON OF COWBIRD AND GRASSLAND PASSERINE EGGS 347 torces provided by all hosts exhibiting egg discrimination. This causes the parasite’s eggs to assume the average color and maculation pattern of eggs from all intolerant hosts in proportion to the selective pressure provided by each host. A single parasite egg can be laid randomly in any host nest in true generalized egg mimici^ and still have a high probability of matching the host’s eggs. Indirect evidence of generalized cowbird egg mimicry has come from experiments that showed several grassland passerines accept cowbird eggs despite possessing an ability to eject less mimetic foreign eggs {Peer et al. 2000, Davis et al. 2002, Klippenstine and Sealy 2008). Therefore, the tolerance of parasitism observed in grassland passerines has, in part, been attributed to their inability to identify cowbird eggs among their clutch, i.e., evolutionary equilibrium hypothesis (Peer et al. 2000, Davis et al. 2002, Klippenstine and Sealy 2008). Generalized cowbird egg mim- icry is plausible because most grassland passerine eggs are similar in color and maculation pattern (Peer and Sealy 2004, Klippenstine and Sealy 2008). Grassland passerines also are historic hosts of cowbirds and, presumably, have been exposed to parasitism sufficiently long for the required co- evolutionary race to have occurred (Mayfield 1965, Rothstein 1994, Igl and Johnson 2007). We quantitatively compared color and macula- tion pattern of cowbird eggs to eggs of a community of grassland passerines known to accept cowbird eggs, but eject non-mimetic eggs, to examine the possibility of generalized cowbird egg mimicry (Klippenstine and Sealy 2008); Sprague’s Pipits (Anthus spragueii). Vesper Spar- rows (Pooecetes gramineus). Savannah Span'ows (Passerculiis sandwichensis), Baird’s Sparrows (Ammodramus hairdii). Chestnut-collared Long- spurs (Calcariiis ornatus), and Western Meadow- larks (Stwnella neglecta). We predicted if gener- alized cowbird egg mimicry has evolved that: (I) cowbird eggs would match the average color and maculation pattern present among the eggs of all six grassland passerines; (2) cowbirds would have incorporated enough of the community-wide variation that a proportion of their eggs would match those of each grassland passerine; and (3) an individual cowbird egg would be capable of mimicking the eggs of more than one species. METHODS Fieldwork was conducted in southcentral Sas- katchewan, Canada from 7 May to I July 2001 and 9 May to 15 July 2002. Sites were 60 km south of Regina within a lOO-knri area centered about the abandoned hamlet of Dummer (49° 50' N, 104° 49' W). Nests were found in cultivated and uncultivated land using predominantly a dragging rope as described by Klippen.stine and Sealy (2008). Photographs of Eggs. — ^Appearance of cowbird eggs and grassland passerine eggs was quantified using photographs taken with a 35-mm Pentax automatic focus camera and 100-speed color Kodak film for increased resolution (Fleischer and Smith 1992). Photographs of complete clutches were desired unless this was not possible because of partial predation, cowbird parasitism, or if eggs broke when handled. Cowbird eggs were photographed whenever a nest was naturally parasitized and were included in the photograph of the host’s eggs. All eggs were temporarily placed in an egg box, which measured —20 by 50 cm and had a foam mat on the bottom with six indentations that held the eggs in place during each photograph. Three color chips (red, green, and blue), each with three shades of their corresponding color (light, medi- um, and dark), were placed on the mat and included in the photograph (Fleischer and Smith 1992, Villafuerte and Negro 1998). The egg box was placed in a portable hood to exclude outside lighting and a camera stand within the hood was used to ensure consistent photography. Films were developed as 4 X 6 color prints with a matte finish to decrease glare. Photographs in 2001 were developed throughout the field season, but all photographs in 2002 were developed on the same day to reduce variation in color, contrast, and brightness caused by developing at different times. Prints were scanned with a flatbed scanner at 300 dpi for digital analysis. Measurements of Egg Parameters. — Four egg parameters were measured: ground color, macu- lation color, distribution of maculation, and total density of maculation. Eggs were isolated from their respective photographs using the magic wand tool of Adobe Photoshop 4.0 to enhance the accuracy of measurements. Maculation was measured using inverted black and white images of each egg. Markings were highlighted as bright white spots, leaving the ground color of the egg black. The proportion of maculation on the blunt and pointed ends of each egg was measured by centering a 20- X 30-pixel rectangular, selection box one-quarter the distance from the pointed end 348 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 2, June 2010 and one-quarter from the blunt end of each egg, and measuring the proportion of area covered in white compared to black using the Northern Eclipse Program (modified from Fleischer 1985). Density of maculation was calculated as the percent total area covered in maculation for both blunt and pointed ends of each egg on a scale of 0 to 100, i.e., percent X 100. The distribution of maculation was calculated as the percent of the total density of maculation occupying only the blunt end, i.e., blunt end maculation divided by total maculation multiplied by 100. Eggs with evenly distributed maculation had values close to 50 and values increased as the proportion of maculation on the blunt end of the eggs increased. Ground and maculation colors were measured from the color images of each egg using the RGB program (Villafuerte and Negro 1998). Three samples of ground color were captured using Adobe Photoshop 4.0, i.e., three square 10 X 10 pixels, along the length of the egg, and avoiding the area of brightness caused by glare of the camera’s flash. Three to six spots or scrawls were sampled in the case of maculation. All selected areas were transferred into the RGB program, which calculated the red, green, and blue value of each pixel on a scale of 0 to 256 (Villafuerte and Negro 1998). The average red, green, and blue values were assigned to the background color and color of maculation for each species’ egg. Preliminary analy.ses revealed that for eggs of all species the proportion of red, green, and blue was highly correlated to each other in both ground color and maculation color, respectively. The amounts of red, green, and blue for both maculation and ground color were summed because they were not independent; the new value approximated the darkness or lightness of the color on a scale of 0 to 768 with an egg’s color becoming darker as this number decreased. This was possible because ground color and macula- tion color of each species’ eggs examined tended to be brown, differing only in intensity of color, which would be seen as lightening or darkening. The correlation between red, green, and blue would have been lower or absent had different colors, i.e., blue, pink, etc., been abundant on the eggs. The extent to which the shade o( each color varied within a photograph due to minor differ- ences in illumination, development, and film was calculated by measuring the discrepancy between observed and theoretical values for the three shades of the color chips photographed with the eggs. A correction factor was computed that represented how much red, green, and blue was added or subtracted from the photograph during its exposure and development (Villafuerte and Negro 1998). This correction factor was calculat- ed for each photograph by averaging the differ- ences between theoretical and observed values for all three shades of each color. Statistical Analyses. — Eggs of all six grassland passerines were pooled to calculate the average parameters for this community. The total number of eggs for each species was reduced to 15 by randomly eliminating additional eggs to ensure that all species were represented equally when pooled. All reductions were repeated twice and the results were compared to ensure that no bias occurred during the initial reduction. Only 14 Sprague’s Pipit eggs were available for compar- ison, and the average of all 14 eggs was inserted to make up the missing value (Zar 1996). We opted to pool the eggs of each species in equal proportions rather than in proportion to their frequency of parasitism, relative abundance or frequency of cowbird egg ejection, i.e., as a measure of selective pressure, for two reasons: ( 1 ) these parameters do not vary significantly among species within this community (Davis 2003, Klippenstine and Sealy 2008); and (2) currently, levels of parasitism and relative abundance of each species varies between years and habitat structure (Davis 2003, Klippenstine and Sealy 2008). A single measure of these parameters may not accurately represent the selective pressure provided by each species currently or throughout their co-evolution with cowbirds. We used Discriminate Analysis (DA) to examine the extent of discrimination possible between cowbird eggs and the average grassland passerine egg using the categories of maculation (density and distribution of maculation) and color (ground and maculation color). An ANOVA was used to identify which parameter or parameters created the difference if discrimination was observed in any of these categories (Manly 1986). All four parameters were recombined, based on the latter analyses, to acquire the correct collection of parameters that minimized discrim- ination between cowbird eggs and the average grassland passerine egg. The predicted group membership function following a DA was used to assess overlap between cowbird eggs and grassland passerine eggs using minimum discrim- Klippciislinc and Sca/y • COMPARISON OF COWBIRD AND GRASSLAND PASSERINE EG(}S 349 TABLE 1. Mean parameter values for the average grassland passerine egg and Brown-headed Cowbird eggs. Species («) Density of maculation, % Disiribuiion of maculalion. % Ground color Maculation color Average grassland passerine (90) 23.5 ± 2.0 Brown-headed Cowbird (50) 28.7 ± 2.2 75.8 ± 2.2* 64.3 ± 1.9 472. 1 ± 9.4 453.6 ± 7.2 372.0 ± 7,1 362.1 ± 6.4 * Differs (P = 0.05) from eowbird eggs. ination parameters. This function works by attempting to re-assign each egg blindly to its correct species, listing the portion of eggs for each species that was misclassified as being the eggs of all other species (Manly 1986). The portion of misclassification between two species’ eggs represented the extent of overlap or “mimicry” between their eggs. The predicted group membership analysis was conducted on two scales to fully assess mimicry: ( 1 ) the frequency of misclassification that oc- curred when eggs of all six grassland species were compared simultaneously with eowbird eggs; and (2) the frequencies of misclassification that occurred when eowbird eggs were compared separately to eggs of each grassland species. We examined whether the sum of misclassification frequencies between eowbird eggs and eggs of each species was greater than the proportion of eowbird eggs misclassified when compared to all species simultaneously to learn whether individ- ual eowbird eggs mimic those of multiple species. If, however, each eowbird egg only matches the eggs of one species, the sum of misclassification frequencies involving each species should be equal to the proportion misclassified when all eggs were compared together. Differences in the frequency of misclassification between species were as,sessed using a Chi-square test. There was a tendency for greater discrimination because the same data used in the initial DA were used in the predicted group membership (Manly 1986). This analysis was conservative, as it was more likely to separate species than place them together. Statis- tical significance was set at alpha = 0.05 for all analyses, and all average values are given as mean ± SE. RESULTS Cowhird Eggs versus Average Grassland Passerine Egg. — Cowbird eggs were indistin- guishable in color (ground and maculation combined) from the average grassland passerine egg for this community (DA, yj2 = 1.8, E = 0.40; Table 1, Eig. 1), but were readily discernable in maculation (density and distribution of macula- tion combined; DA, x'2 = 13.5, P = 0.001). Dissimilarities in overall maculation resulted exclusively from cowbird eggs being far more evenly maculated than the average grassland passerine egg {P\ = 12.3, P = 0.001), as no difference in the total density of maculation was observed between the two groups (P\ = 2.5, P — 0.17, Table 1). Thus, minimum discrimination between cowbird eggs and the average grassland passerine egg was achieved by excluding distri- bution of maculation as a parameter (DA, xS = 2.8, P = 0.43 vs. xS = 13.5, P = 0.009 for all four parameters). Overlap BetM'een Cowbird Eggs and Grassland Passerine Eggs. — ^The predicted group member- ship analysis was significantly more likely to misclassify cowbird eggs on the basis of mini- mum discrimination parameters, i.e., color and total density of maculation, than eggs of all other grassland passerines (x‘i > 4.4, P < 0.05 for all species; Table 2). Overall, 88% of cowbird eggs were misclassified compared to 4 to 70% of grass- land passerine eggs (Table 2). Between 4 and 67% of grassland passerine eggs were misclassi- fied as being the eggs of another grassland passerine, an indication of the close resemblance between grassland passerine eggs (Table 2). Between 8 to 48% of cowbird eggs were misclassified, depending on the species, when cowbird eggs were compared to eggs of each species separately using minimum discrimination parameters (Table 3). The sum of all cowbird egg misclassifications for all six species greatly exceeded the proportion of cowbird eggs misclas- sified when all eggs were compared simulta- neously (186 vs. 88%; Tables 2, 3). The propor- tion of cowbird eggs misclassified did not differ (28 to 49%; X'4 = 6.0, P = 0.20) among eggs of Sprague’s Pipit, Vesper, Savannah, and Baird’s sparrows, and Chestnut-collared Longspurs (Ta- ble 3). A substantially lower proportion of cowbird eggs were misclassified as Western 350 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2, June 2010 Host species Host egg Cowbird egg Sprague’s Pipit Vesper Sparrow Savannah Sparrow Baird’s Sparrow Chestnut-collared Longspur Western Meadowlark FIG. 1. Example.s of remarkable similarities between gra.ssland pas.serine eggs and cowbird eggs in naturally parasitized nests. Meadowlark eggs (8%) than the other five species (X^i > ^ ^ species; Table 3). DISCUSSION Our results support the hypothesis that cowbird eggs have evolved to generally mimic the appear- ance of eggs from multiple grassland passerines. Cowbird eggs were indistinguishable from the ground and maculation colors, and total density of maculation as measured for the average grassland passerine eggs from this community in southern Saskatchewan (Table I ). All six grassland species examined have been experimentally demonstrated to eject non-mimetic eggs (Klippenstine and Sealy 2008). Thus, our results support the idea that ejection of discordant cowbird eggs has forced them to assume the average appearance of eggs from all intolerant hosts. Cowbird eggs so closely approximated the color and maculation density of grassland passer- KUppeusthw and Sealy • COMPARISON OF COWBIRD AND GRASSLAND PASSFZRINE EGGS 351 TABLE 2. Proportion of correct identification and misclassification observed for each species when compared simultaneously using minimum discrimination parameters (ground color, maculation color, and density of maculation). Eggs (%) mi.sclassiried £ IS . . . Species {n) Eggs correctly identified (%) Other grassland passerine eggs Brown-headed Cowbird eggs Other species overall Sprague’.s Pipit (14) 71 22 7 29 Vesper Sparrow (21) 43 57 0 57 Savannah Sparrow (34) 41 56 3 59 Baird's SpaiTow (37) 54 30 16 46 Chestnut-collared Longspur (40) 30 67 3 70 Western Meadowlark (24) 96 4 0 4 Brown-headed Cowbird (50) 12 88 * 00 00 * Differs (P = 0.05) from all other grassland passerine eggs. ine eggs that 88% ( 1 of every 1 . 1 eggs) of cowbird eggs were misclassified as belonging to another species (Table 2). Between 8 and 48% of cowbird eggs were misclassified depending on the grass- land hosts, when compared to eggs of each species separately (Table 3). Thus, any cowbird egg laid randomly in a Sprague’s Pipit, Vesper Sparrow, Savannah Sparrow, Baird’s Sparrow, or Chestnut- collared Longspur nest has between a “one-in- three” and a “one-in-two” chance of matching the hosts eggs; the chance of matching for Western Meadowlarks is significantly lower at approximately “one-in-ten” (Table 3). Effective mimicry requires most cowbird eggs to match the eggs of at least one host. This is especially true considering female cowbirds are thought to exhibit little to no host selection (Fleischer 1985, Alderson et al. 1999). “Generalized” mimicry necessitates that color and maculation density of cowbird eggs overlap to a significant extent with the eggs of all six species. Another important component of generalized mimicry is the idea that a single cowbird egg may mimic the eggs of more than one species. This contrasts with the elaborate mimetic system evolved by the Common Cuckoo. Different subgroups or “gentes” of female cuckoos in that TABLE 3. Proportion of nii.scla.ssificulion observed using minimum discrimination parameters (ground color, maculation color, and density of maculation) between cowbirds eggs and eggs of individual species. Cowbird Species (H) eggs misclassified as eggs of each grassland passerine i%) Sprague's Pipit ( 14) 28 Vesper Sparrow (21 ) 38 Savannah Sparrow (34) 36 Baird's Sparrow (37) 28 Chestnut-collared Longspur (40) 48 Western Meadowlark (24) 8* Total 186 * Differs (P = 0.05) from all other grassland eggs. system mimic the eggs of a preferred host (Davies and Brooke 1988, Marchetti et al. 1998, Gibbs et al. 2000). The benefit of gentes is that it allows cuckoos, as a species, to parasitize several species that exhibit egg discrimination while lowering competition for nests (Davies and Brooke 1988). Cowbird gentes are supported (Fig. 1), illustrating some of the remarkable similarities we discovered between cowbird eggs and the egg of each grassland passerine. However, the high frequency at which cowbird eggs were misclassified when compared to the eggs of each species (8 to 48%, sum total 186%; Table 3) is only possible if individual cowbird eggs are capable of matching the eggs of more than one species. We conclude that eggs of individual female cowbirds have not evolved to mimic those of a prefeixed species. Cowbird eggs were more evenly maculated than the average grassland passerine egg despite the strong resemblance in three of four parameters (Table 1 ). This discrepancy may ari.se because an unforeseen cost exists for cowbird eggs that are more unevenly maculated, e.g., increased rates' of depredation (Underwood and Sealy 2002, Kilner 2006). Cowbird eggs have similarly been shown to be rounder than host eggs to increase shell strength and reduce the potential for damage during laying or by intolerant hosts (Rohwer and Spaw 1988, Pieman 1989). Alternatively, diver- gences in maculation distribution might reflect the parameters used by grassland passerines to discriminate cowbird eggs. Rothstein (1982) found that Gray Catbirds (Dumctella carolinensis) were more responsive to changes in color than maculation. and American Robins (Turdus mi- gratoriiis) were more sensitive to color than size (also Underwood and Sealy 2006). Differences in 352 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2, June 2010 host discrimination alter the selective pressure placed on cowbird eggs by each host and modify which parameters are mimicked (Rothstein 1982, Underwood and Sealy 2006). Consequently, the distribution of maculation may not be imitated by cowbirds because grassland passerines discrimi- nate foreign eggs solely on the basis of color and density of maculation. Our results show that cowbird eggs were most readily distinguished from eggs of Western Meadowlarks on the basis of color and maculation density (Tables 2, 3). This corresponds with the level of intolerance observed for this species compared to the other five grassland passerines (Klippenstine and Sealy 2008). Western Mead- owlarks ejected 67% of cowbird eggs and 91% of the non-mimetic eggs added to their nests. The five remaining grassland passerines, in compari- son, accepted all or nearly all cowbird eggs and ejected about 20% of non-mimetic eggs (Klip- penstine and Sealy 2008). Western Meadowlarks apparently eject cowbird eggs more readily because of the lower extent of matching between the eggs of the two species. Peer and Sealy (2004) rated Western Meadowlarks eggs similar in color and maculation to cowbird eggs. They also suggested egg appearance was not correlated with this host’s intolerance of parasitism, but was related to their larger beaks. We suspect the difference arises because they compared the color and maculation of eggs using human observers. It is extremely difficult, however, for a human observer to assess the density of maculation independent of its distribution, which in this study had a profound affect. Consequently, discrepan- cies observed between the two studies emphasize the benefits of assessing egg appearance quanti- tatively. Generalized cowbird egg mimicry is not the only possible explanation of the close resem- blance between cowbird eggs and grassland passerine eggs. Concealment of eggs in the nesting habitat from depredation is a primary force directing the evolution of egg appearance in most species (Underwood and Sealy 2002, Kilner 2006). Therefore, the same forces promoting the evolution ot color and maculation density in grassland passerine eggs may have .selected the same pattern in cowbird eggs (Underwood and Sealy 2002, Kilner 2006). We propo.se two criteria to distinguish between resemblance due to nest concealment and cowbird egg mimicry. The first is that cowbird egg mimicry requires host species that exhibit egg discrimination and ejection. The presence of egg discrimination in all six species strongly supports the idea of cowbird egg mimicry (Klippenstine and Sealy 2008). A worthwhile test of this criterion would be to examine the extent of matching between cowbird eggs and eggs of grassland passerines that have not evolved egg recognition. The second criterion is that nest concealment would not necessarily make cowbird eggs better at matching grassland passerine eggs than eggs of any other grassland passerine. That cowbird eggs were significantly more likely to be misclassified than all other grassland passerine eggs (88% vs. 4-70%) supports generalized cowbird egg mimicry (Table 2). The benefits of mimicking grassland passerine eggs are obvious for cowbirds. Despite having evolved egg discrimination, all six grassland species fully or partially accept cowbird eggs (Davis et al. 2002, Klippenstine and Sealy 2008). This tolerance has translated into reasonable breeding success for cowbirds within this com- munity (Davis et al. 2002, Davis 2003, Klippen- stine and Sealy 2008).). However, we believe the value of generalized cowbird mimicry is mo.st likely limited to only a few ho.st assemblages. Communities where host eggs are highly diver- gent in appearance are probably not conducive to generalized mimicry (Rothstein 2001). General- ized cowbird egg mimicry, even for grassland passerines examined in this study, is feasible only because each host exhibits a relatively crude level of egg recognition (Klippenstine and Sealy 2008). Greater selective pressure will be placed on cowbirds to evolve more species-specific egg mimicry as continued parasitism forces grassland passerines to develop their ability to discriminate foreign eggs (Moksnes et al. 1990, Davies and Brooke 1991 ). Cowbird egg mimicry of grassland passerine eggs may eventually resemble the type of mimicry exhibited by the Common Cuckoo. Alternatively, as Rothstein (2001) suggested, cowbirds may in the future choose to avoid such troublesome hosts. ACKNOWLEDGMENTS We are indebted to the landowners who permitted us to conduct our research on their land. We thank M. A. Talbot and N. L. Haalboom for their hard work in the field; Cornelius Klippenstine for use of his farmhouse; C. S. Bjornsson and Erwin Huebner for technical assistance; N. C. Kenkel, M. A. Klippen.stine, D. M. Gillis, .1. F. Hare, and M. V. Abrahams for statistical advice; M. D. Kujath, L. C. Graham for commenting on earlier drafts of the manuseript; Klippeusliiw ami Scaly • COMPARISON OF COWBIRD AND GRASSLAND l>ASSERINE EGGS 353 and two anonymous reviewers for eonstruelivc comments that improved this manuscript. This study was funded by a Discovery Grant from the Natural Sciences and Engineer- ing Research Council of Canada to S. G. Scaly and through in-kind services provided by S. K. Davis and the Saskatchewan Watershed Authority (formerly Saskatche- wan Wetlands Conservation Corporation). This paper is dedicated to the memory of B. A. Klippenstine. LITERATURE CITED Alderson, G. W., H. L. Gibbs, and S. G. Sealy. 1999. Determining the reproductive behaviour of individual Brown-headed Cowbirds using microsatellite DNA markers. Animal Behaviour 58:895-905. Davies, N. B. and M. de L. Brooke. 1988. Cuckoos versus Reed Warblers: adaptations and counter adaptations. Animal Behaviour 36:262-284. Davies. N. B. and M. de L. Brooke. 1991. Coevolution of the cuckoo and its hosts. Scientific American 264:92-98. Davis, S. K. 2003. Nesting ecology of mixed-grass prairie songbirds in southern Saskatchewan. Wilson Bulletin 115:1 19-130. Davis, S. K., D. R. Klippenstine, and R. M. Brigham. 2002. Does egg recognition account for the low incidence of cowbird parasitism in Chestnut-collared Longspurs (Calcarius ornaliis)! Auk 1 19:556-560. Fleischer, R. C. 1985. A new technique to identify and assess the dispersion of eggs of individual brood parasites. Behavioral Ecology and Sociobiology 17:91-99. Flei.scher, R. C. and N. G. Smith. 1992. Giant Cowbird eggs in the nests of two icterid hosts: the use of morphology and electrophoretic variants to identify individuals and species. Condor 94:572-578. Gibbs, H. L., M. D. Sorenson, K. Marchetti, M. de L. Brooke. N. B. Davies, and H. Nakamura. 2000. Genetic evidence for female host-specific races of the Common Cuckoo. Nature 407:183-186. IGL, L. D. AND D. H. JOHN.SON. 2007. Brown-headed Cowbird, Molothrus ater, parasitism and abundance in the Northern Great Plains. Canadian Field-Naturalist 121:239-255. Kilner. R. M. 2006. The evolution of egg colour and patterning in birds. Biological Reviews 81:383^06. Klippenstine, D. R. and S. G. Sealy. 2008. Differential ejection of cowbird eggs and non-mimetic eggs by grassland passerines. Wilson Journal of Ornithology 120:667-673. Lotem, A. and H. Nakamura. 1998. Evolutionary equilibria in avian brood parasitism: an alternative to the “arms race-evolutionary lag” concept. Pages 22.3- 235 ill Parasitic birds and their hosts: studies in coevolution (S. I. Rothstein and S. K. Robinson, Editors). Oxford University Press, Oxford. United Kingdom. Manly, B. F. J. 1986. Multivariate statistical methods: a primer. Chapman and Hall. New York, USA. Marchetti, K., H. Nakamura, and H. L. Gibbs. 1998. Host-race formation in the Common Cuckoo. Science 282:471-472. Mayfield, H. 1965. The Brown-headed Cowbird, with old and new hosts. Living Bird 4:13-28. Mermoz, M. E. and J. F. Ornelas. 2004. Phylogenetic analysis of life-history adaptations in parasitic cow- birds. Behavioral Ecology 15:109-119. MOKSNES, A., E. R0SKAFT, A. T. Braa, L. Korsnes, H. M. Lampe, and H. C. Pedersen. 1990. Behavioral responses of potential hosts towards artificial cuckoo eggs and dummies. Behaviour 1 16:64-89. Peer, B. D. and S. G. Sealy. 2004. Correlates of egg rejection in hosts of the Brown-headed Cowbird. Condor 106:580-599. Peer, B. D., S. K. Robinson, and J. R. Herkert. 2000. Egg rejection by cowbird hosts in grasslands. Auk 1 17:892-901. PiCMAN, J. 1989. Mechanism of increased puncture resistance of eggs of Brown-headed Cowbirds. Auk 106:577-583. Rohwer, S. and C. D. Spaw. 1988. Evolutionary lag versus bill-size constraints: a comparative study of the acceptance of cowbird egg by old hosts. Evolutionary Ecology 2:27-36. Rothstein, S. 1. 1982. Mechanisms of avian egg recogni- tion: which egg parameters elicit responses by rejecter species? Behavioral Ecology and Sociobiology 11:229-239. Rothstein, S. I. 1994. The cowbird’s invasion of the far west: history, causes and consequences experienced by host species. Studies in Avian Biology 15:301-315. Rothstein, S. 1. 2001. Relic behaviours, coevolution and the retention versus loss of host defences after episodes of avian brood parasitism. Animal Behaviour 61:95-107. Sealy, S. G., D. G. McMaster, and B. D. Peer. 2002. Tactics of obligate brood parasites to secure suitable incubators. Pages 254-269 in Avian incubation: behaviour, environment, and evolution (D. C. Deem- ing, Editor). Oxford University Press, Oxford, United Kingdom. SoLER, J. J., J. M. Aviles, M. Soler, and A. P. Mgller. 2003. Evolution of host egg mimicry in a brood parasite, the Great Spotted Cuckoo. Biological Journal of the Linnean Society 79:551-563. Underwood, T. J. and S. G. Sealy. 2002. Adaptive significance of egg coloration. Pages 280-298 in Avian incubation: behaviour, environment, and evo- lution (D. C. Deeming. Editor). Oxford University Press, Oxford. United Kingdom. Underwood. T. J. and S. G. Sealy. 2006. Parameters of Brown-headed Cowbird Malotlinis ater egg discrim- ination in Warbling Vircos Vireo yilviis. Journal of Avian Biology 37:457^66. Vii.LAFUERTE. R. AND J. J. Negro. 1998. Digital imaging for colour measurement in ecological research. Ecology Letters 1:151-154. Zar. j. 1996. Biostalistical analysis. Third Edition. Prentice Hall, Upper Saddle River. New Jersey. USA. The Wilson Journal of Ornithology 122(2);354-360, 2010 FLIGHT FEATHER MOLT OF TURKEY VULTURES ROBERT M. CHANDLER, PETER PYLE,--' MAUREEN E. ELANNERY,' DOUGLAS J. LONG,''^ AND STEVE N. G. HOWELL'' ABSTRACT. — We document the molt sequence of flight feathers in Turkey Vultures {Cathartes aura) based on studies of captive and wild birds, and examination of museum specimens. We found an unusual pattern of primary replacement, which appears to be a modified form of Stajfelmauser, or stepwise wing molt. A Stajfehnauser-Wke strategy for replacement of the secondaries is also described. These patterns of feather replacement appear to be adaptations to maintain flying performance while replacing all primaries and most secondaries during each molt. To what extent molt patterns in Turkey Vultures reflect convergent adaptation for flight, rather than ancestral characters useful for phylogenetic studies, remains unknown. Received 15 June 2009. Accepted 12 January 2010. Maintaining optimal flight performance is essential for Turkey Vultures (Cathartes aura), a species adapted for long-term soaring (Rosser and George 1986, Tucker 1987). This species locates its primary diet of carrion by detecting odor plumes while aloft, and spends most daylight hours searching for food (Owre and Northington 1961, Stager 1964). It is a migratory species that can travel hundreds or thousands of kilometers seasonally between southern wintering and north- ern breeding areas (Stewart 1977, Kirk and Houston 1995, Bildstein and Zalles 2001). Little is known about the molt of Turkey Vultures despite its widespread distribution and abundance (Kirk and Mossman 1998). Identifying and understanding molt processes in birds can be accomplished in several ways. Miller (1941), Stresemann (1963), and Pyle (2005) used data from museum study skins of raptors to make their basic molt sequence hypotheses. Analyzing molt patterns from study skins can be difficult and may not represent the progression of a single individ- ual, and sample sizes of large birds are typically too low to undertake comprehensive analyses. Houston (1975) observed flight feather replace- ment in both captive and wild individuals of Old World vultures (Gyps), but molt in captive birds ' Department of Biological and Environmental Sciences, Georgia College and State University. Milledgeville. GA 31061. USA. ^Institute for Bird Populations, P. O. Box 1346, Point Reyes Station. CA 94956, USA. ' Department of Ornithology and Mammalogy, California Academy of Sciences, 55 Music Concourse Drive, San Francisco. CA 941 18. USA. ■* Department of Natural Sciences, Oakland Museum of California, 1000 Oak Street. Oakland. CA 94607, USA. 'PRBO Conservation .Science, 3820 Cypress Drive #1 1, Petaluma, CA 94954. USA. '’Corresponding author; e-mail: bob.chandler@gcsu.edu may not reflect that of wild birds (cf. Pyle 2005). Snyder et al. (1987) based their molt study of California Condors (Gymnogyps californiamis) on photographs of wild birds in flight. Bloom and Clark (2001) used banding records with some recaptures to study the molt of the Golden Eagle (Acitiila chrysaetos), but sample sizes of known individual birds are low in these studies. Ideally, a complete understanding of molt strategies would be based on the study of wild birds, captive individuals, and specimens; we present such an analysis of molt in Turkey Vultures. METHODS A. M. Rea began collecting molt sequence data on Turkey Vultures in 1972 as part of a larger study on the phylogenetic relationship of New World vultures (Rea 1983). Chandler joined the project in 1978, collecting feathers from three adult individuals held captive at the San Diego Natural History Museum (SDNHM) in San Diego, California, USA and studying molt year-round on several others. Flannery and Long tabulated molt data on five vultures (4 adults and 1 first year bird) between 2002 and 2004, which were held captive by the California Academy of Sciences (CAS) in San Francisco, California. The extended study period of captive individuals allows calculation ot feather replacement rates both within and between individuals to establish ranges of variation. Pyle and Howell visually examined hundreds of vultures in the field, and Pyle and Flannery examined molt in over 20 museum specimens. We classified ages of Turkey Vultures using head and beak coloration (Henckel 1981), and known history for captive individuals. Age coding of birds follows Pyle (1997); HY (hatching or first calendar year), SY (second calendar year), TY (third calendar year), and ASY (at least third 354 Chandler el al. • TURKEY VULTURE MOLT 355 calendar year). Chandler collected data from two captive (SDNHM) ASYs during 1972-1980 and a third ASY during 1979-1980. Flannery and Long collected data from four captive ASYs (3 males and 1 female known to be ages 7+ to 25+ years) and a female received as an FlY, housed at CAS during 2002-2004. The ASYs were from Okla- homa, Pennsylvania, and Texas; the HY, received with down still present on the head and the majority of the beak still dark, was from Tennessee and amved in December 2001. All eight individuals were kept in outdoor pens, ensuring they were acclimated to the natural local photoperiod and climate. Primaries, secondaries, and rectrices for the vultures studied by Chandler (SDNHM) were painted with a dot and slash code to identify the specific feather (a dot = 1 and a slash = 5), and with a year-specific color for each individual for each year of study. The aviary was checked several times daily for dropped feathers, which were labeled and dated. Wings were examined visually for the individuals at CAS, and dropped feathers were examined to tabulate wing molt data. Primaries and secondaries on each healthy wing of each individual were scored as old, missing, in pin, growing, or new. The percentage of feather length grown was recorded for each growing feather. Data were collected weekly during the first year of the study and biweekly or monthly in subsequent years. Drop dates for each primary on each bird were estimated by calcu- lating primary-specific growth rates and obtain- ing mean, back-calculated dates for each feather. Feathers not dropped that year were designated as being retained in both studies. Pyle and Howell recorded data from 1998 to 2006 on wing molt from Turkey Vultures in the field. Binoculars were used to score primaries as missing, growing, new, or old. Molt was assumed if the same primary was missing or growing on both wings simultaneously; individ- uals with a missing or growing primary on only one wing were not included in the data .set. Data were collected year-round and, although most data were obtained from central California, at least 10 birds each from Maryland, Indiana, Mexico, and El Salvador were scored. Vultures observed with dark heads and beaks, and no primary molt in June-December were classified as HYs and not included in the data set. Analysis was based on date of observation and progres- sion of molt, including either no molt, or a score representing the number of the primary that was growing on the observation date. Individuals with two primaries being replaced simultaneous- ly from different parts of the wing (e.g., an inner and an outer primary) were scored twice in the data set. Secondary molt patterns afso were scored when visible. We analyzed the data using a two-tailed Utest with JMP Version 5.1 statistical software (SAS Institute 2002) to examine whether or not there was a difference in mean dates of molt between captive and wild birds. More than 20 specimens at CAS, Museum of Vertebrate Zoology (MVZ), SDNHM, and the National Museum of Natural History (USNM) were examined. Primaries and secondaries were scored in the same manner as captive birds, and replacement sequences were inferred based on wear dines reflecting protracted molts (Pyle 2005). RESULTS Primaries. — ^All data indicate the 10 primaries of a Turkey Vulture molt sequentially from the innermost primary (PI) to the outermost primary (PIO). This process started for the SY at CAS in late January with PI and was completed in late July with PIO (Pig. 1). This SY then replaced PI- PS for a second time within the year in September-November (Pig. I ). The following year, as a TY, this individual continued the sequence where it had suspended the year before, replacing P4-P10 in order between early March and early August, and replacing PI -PS in August- September (Pig. 1). The captive ASYs at both locations replaced P5-P10 in March-September and replaced P1-P2 in July-August (Pig. 1). This pattern averaged later than the drop dates recorded for the younger bird as an SY and a TY. Replacement of PS and P4 in ASYs was more variable (Pig. 1). Based on the above results, data on P1-P4 from birds in the field were partitioned as either molting in August-November or in January- March (Pig. 2). The remaining primaries. P5- PIO had sequential mean dates of molt, beginning in April-May (P5) and finishing in September- October (PIO). This pattern was similar to that recorded for captive birds of all ages (Pig. 2). Mean dates of molt for P1-P4 and P7-P8 were not significantly different between captive and wild birds. However, mean drop dates were signifi- cantly earlier (two-tailed Me.st, P < 0.01) in 356 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122. No. 2, June 2010 Captive data 10 - A ♦ □ 9 - A ♦ □ 8 - El 7 - ♦ A o E5 CC 6 - A □ 5 □ ♦ A 4 - □ ♦ ► □ 3 - A t> A 2 - A «□ 1 - ib. □ ♦ A J F M A M J J A S 0 N D Mouth ♦ TY □ AHYavc FIG. I. Calculated feather drop date by month for Turkey Vultures. Data collected from Turkey Vultures held captive at the California Academy of Sciences. SY and TY data collected from same individual during 2002 and 2003. Open triangle represents assumed drop date for PI in a SY calculated using percentage feather grown; actual drop date occurred prior to acquisition of the bird. captive than in wild populations for P5, P6, P9, and PIO. This difference is most likely attribut- able to the higher energy requirements for breeding and food acquisition observed in wild populations that could suspend, slow down, or otherwise delay molt. Three vultures observed in the field were replacing PI during 20-29 January. All three Field obseivatious 3 T J 29 I » 26 I ■ < 97 I ■ 1 42 I ■ 1 19 I -■ 1 9 I •— I 10 t— ■ — I 1 1 « 1 1 F M A M J J 56 I ■- 26 I ■ 1 H 30 I ■ ( 28 I ■ 1 19 I ■ I T 1 1 1 A S O N Month D FIG. 2. Mean (± SE) feather drop date for Turkey Vultures by month based on field observations. Chaiullcr et al. • TURKEY VULTURE MOLT 357 were classified as SYs by head plumage and evenly worn primaries (P2-P10). Most of 19 vultures observed replacing P2-P3 in February and March (Fig. 2) were classified as SYs. In contrast, many of the 77 vultures observed replacing P1-P3 in August-November (Fig. 2) were ASYs. Vultures replacing P4 in February- April (/; = 19) included three SYs and seven ASYs. Two vultures that could not be partitioned into our defined categories were replacing P3 on 4 January and P4 on 7 December (Fig. 2). Both of these individuals were observed in central Cali- fornia. ASY vultures showing no primary molt (/? = 181) were observed between 6 September and 17 March. The mean (± SD) date for individuals showing no molt was 14 January ±49 days. Secondaries. — ^There were three consistent foci for initiating molt of secondaries based on data from the SDNHM captive birds, specimens, and field observations: the tertials and secondaries 1 and 5 (SI and S5, numbered proximally from outermost). Molt proceeded proximally from SI and S5 and distally from the tertials, and it may become irregular in the area from S8 to SI 1. Six secondaries were retained from 1973 to 1978 on captive ASYs (SDNHM), two for one bird and four for the other individual. Combined locations of retained secondaries for the two birds were: S4, S7 (twice), S8 (twice), and SIO. SI dropped regularly in April followed by S2 and S5 in May and June, respectively. S3 and S6 usually dropped in June, S4 and S7 in July, S8 in August or September (but at times as early as mid-July), and S9 and SIO dropped between June and late September; S9 always dropped before SIO. There was a tendency for molt to terminate at SI 1 and S12 with drop dates between May and June, and S12 following Sll. Smaller secondaries (S1-S4) dropped together in April and May and as late as August. The tertiary feathers, as with the smaller secondaries, had a tendency to drop together between May and August with no di.scernable sequence. There appeared to be .some interannual varia- tion in the number of secondaries retained. For example, 1979 was an atypical year of extraordi- narily high numbers of retained feathers. One vulture retained five secondaries and two rectri- ces, another had nine .secondaries and one rectrix retained that year. A possible rea.son is that the birds suffered from lead poisoning. The following year, after the birds had recovered from the apparent lead poisoning, no feathers were re- tained. Retained feathers dropped in May or June of the next year. Rectrices. — ^Tail feathers generally had a con- sistent pattern for the six feather pairs (SDNHM). On one side, numbering rectrices from the central feather R1 to the outermost feather R6, the sequence was R1-R6-R2-R4-R3-R5. R1 dropped between late April and May, R6 dropped regularly in May. The largest rectrix, R2, dropped in late May or June, followed by R4 in late June or July, R3 in July or August, and R5 in mid-August to early September. Tail feathers retained from the previous year dropped in May or early June of the next year. Three rectrices were retained during the first 6 years (1973-1978), a pair of R4s for one vulture and a single R5 for another. Feather Synchrony and Growth. — ^Time elapsed between dropping of paired (right and left) feathers was calculated for P5 in captive birds at SDNHM. The average elapsed time was 7 1/2 days after the first of the pair had dropped (range 1- 41 days). PI and P6 through PIO held to this average or had an even shorter elapsed time, at times less than 24 hrs. P2 through P4 had the greatest range with an average of less than 10 days for P2 to an average of 1 month for P3 and P4. The maximum time for an individual feather was 41 days between drop dates for P3. DISCUSSION Basic concepts of flight feather molt in diurnal soaring hawks, eagles, and Old World vultures have been presented by Miller ( 1941 ), Stresemann (1963, 1966), and Stresemann and Stresemann ( 1966). Later studies on wing molt by Brown and Amadon (1968), Houston (1975), Brown (1976), Edelstam (1984), Clark (2004), and Pyle (2005) have led to hypotheses about the role of flight feather replacement schemes that may be used to infer possible differences in flight strategies and aerial performance. Birds molt for specific reasons during certain times of the year and this molt is not random feather loss, but has a definite sequence that has an adaptive advantage in optimizing aerodynamics with respect to feather maintenance (Hedenstrom and Sunada 1999. Rohwer 1999, Pyle 2006). Large birds exhibit several remigial replace- ment .strategies detectable within a year and between years by wear patterns, contrasting feather appearance, and retained feathers (Pyle 2005, 2006, 2008). The most common strategy of molt among large birds that need to maintain 358 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 2, June 2010 flight, involves a Stajfelmauser pattern meaning “staggered molt" (Stresemann and Stresemann 1966) and also known as “stepwise molt”. Falconids typically replace all primaries and secondaries during each prebasic molt using multiple waves originating at P4 and S4, whereas larger accipitrids can retain primaries and sec- ondaries from a previous season as part of a Stajfelmauser sequence (Pyle 2005). The physio- logical demands of producing new feathers and the time required to complete a full molt are correlated to body mass and efficiency for flight requirements of larger birds (Shugart and Rohwer 1996, Pyle 2005). Sequential replacement of feathers in waves without having large gaps, increases the efficiency of wings and still allows larger birds to fly. A Stajfelmauser pattern occurs in a wide range of larger birds, e.g., pelicans, herons, and condors (Pyle 2006, 2008). As widespread and common as Turkey Vultures are throughout the Western Hemisphere, it is surprising that no comprehensive review has been published about their molt (but note comments and references by Kirk and Mossman 1998). For example, Stresemann and Stresemann (1966) mentioned that primaries are replaced irregularly, but Jackson (1988) correctly reported that, hairing some aberrant irregular patterns, the molt takes place annually and within a single year, being replaced in a serial inner-to-outer sequence. Rea (1983) included findings by Chandler on second- aries and rectrices. Our data produce a clearer picture of the flight feather molt of Turkey Vultures, which show similarities to other New World vultures, and the California Condor (Snyder et al. 1987). However, the molt of Turkey Vultures is more regular, more symmetrical between wings, and in a 1 -year rather than the condor’s 2-year cycle. There have been no comprehensive studies published on other cathartid vultures for additional comparisons or differences in molt in the Cathartidae. Further comparisons of flight feather molt with other species of birds can be made in two ways, either by comparing molt in avian tamilies related to the Cathartidae, or by comparing molt adapta- tions in other avian families that show morpho- logical convergence with soaring flight. Unfortu- nately, the proximate affinity of cathartid vultures to other avian groups (e.g., Falconitormes or Ciconii formes) remains unresolved (Ligon 1967, Rea 1983, Emsiie 1 988, Sibley and Ahlquist 1990, Avi.se et al. 1994, Helbig and Seibold 1995, Wink 1995, Livezey and Zusi 2007). The wing molt of Turkey Vultures, depending on phylogenetic relationships, may reflect an ancestral pattern. Molt patterns similar to those in Turkey Vultures are shared by some species of Ciconii formes (Bloesch et al. 1977, Shugart and Rohwer 1996), a proposed ancestral group to Cathartidae. Howev- er, similarities in molt patterns are also seen in several other taxonomic groups, and are consid- ered to be a widespread convergent molt adapta- tion in many taxa of larger soaring birds (Pyle 2006). Further data are needed to address this nature versus nurture issue. Molt Terminology. — ^Turkey Vultures have an unusual pattern of primary replacement (Figs. 1, 2) that could be interpreted in several ways. One inteipretation would be that the second molt cycle of SY birds involves replacement of P1-P3 (occasionally P1-P4) twice. This would necessi- tate labeling one of the replacements, either that of January-March or that of August-November, as a prealternate molt. Each successive prebasic molt typically would then begin at P4 (or P5) in the spring and finish with P1-P3 (or P1-P4) in the fall. However, prealternate molts are not known to occur among families thought to be close relatives of New World vultures. A second interpretation would be that replace- ment of P1-P3 in SYs in the fall represents an advanced start to the third prebasic molt, concur- rent with completion of the second prebasic molt. Each subsequent prebasic molt would then begin with PI in the fall, usually (but not always) suspend for the winter, and resume with P4 in the spring (Pyle 2008). This is essentially a modifi- cation of Staffelmauser, in that all primaries are replaced in 1 year via two waves. Incomplete-to- complete replacement of the secondaries also appears to follow a Staffelmauser-\\]s.Q strategy found in other large birds (Pyle 2006). However, molt of primaries differs from text-book Staffel- mauser in that replacement of PI in the second wave is delayed until after the preceding wave has replaced the middle primaries; the molt assumes an annual regularity not found in other species exhibiting Staffelmauser. The advancement of some or all primary molt in selected species of a family, to occur before another life-history event (migration, breeding, chick-feeding), has been documented among loons, puffins, gulls, and diurnal raptors (Howell 2001, Howell and Pyle 2005, Pyle 2005). However, in these groups the second prebasic wing molt typically occurs or Cluiiicller el al. • TURKEY VULTURE MOLT 359 starts lafcr in SYs than in subsequent ages, whieh is not the case in Turkey Vultures. A third interpretation would be to treat the first wing molt of Turkey Vultures as a preformative molt and the molt of P1-P3 of SYs in the fall as the start of the second prebasic molt. The idea of a preformative molt being limited to flight feathers has not previously been recognized, but Howell (in press) has argued that Staffelmaiiser can develop in two ways. One is the conventional view that the first wing molt pertains to the second prebasic molt (a normal schedule), the other is that the process of Staffelmaiiser is kick-started by insertion of a preformative wing molt (an accelerated schedule). Species exhibiting acceler- ated schedules include Osprey (Pandion haliae- tiis), and most if not all Pelecaniformes (Howell, in press). The first wing molt in these species starts earlier than subsequent wing molts, which is also true of Turkey Vultures. To what extent normal and accelerated schedules reflect ancestral traits or environmental responses remains un- known. ACKNOWLEDGMENTS R. M. Chandler especially acknowledges A. M. Rea for initially inviting him to participate in a molt sequence study. We thank W. S. Clark for his constructive review and L. D. Chandler for her help. Many museums and their curators and staff have allowed us to study specimens in their care. For this we thank; Philip Unitt at the San Diego Natural History Museum; S. L. Olson, H. F. James, and R. B. Clapp at the National Museum of Natural History, Washington. D.C.; J. P. Dumbacher at the California Academy of Sciences, San Francisco; and Carla Cicero and the late N. K. Johnson at the Museum of Vertebrate Zoology, University of California, Berkeley. We also thank the SDNHM which supported this research while A. M. Rea was Curator and R. M. Chandler was Collection Manager; as well as the California Academy of Sciences for the support of D. J. Long and M. E. Flannery. LITERATURE CITED Avise, j. C., W. S. Nelson, and C. G. Sibley. 1994. DNA sequence support for a close phylogenetic relationship between some storks and New World vultures. Proceedings of the National Academy of Science of the USA 91:5173-5177. Bildstein, K. L. and j. Zalles. 2001. Raptor migration along the Mesoamerican land corridor. Pages 119-141 in Hawkwatching in the Americas. (K. L. Bildstein and D. Klem Jr., Editors), Hawk Migration Association of North America, North Wales. Pennsylvania, USA. Bloesch. M., M. Dizerens, and E. Sutter. 1977. Die mauser der schwungfedern beim Weisslork Ciconia ciconia. Ornithological Beob. 74:161-188. Bloom. P. H. and W. S. Clark. 2001. Molt and sequence of plumages of Golden Eagles and a technique for in- hand ageing. North American Bird Bander 26:97-1 16. Brown, L. 1976. Classification and distribution. Birds of prey, their biology and ecology. Hamlyn Publishing Group Limited, London. Great Britain. Brown, L. and D. Amadon. 1968. Eagles, hawks, and falcons of the world. Country Life Books, Hamlyn Publishing Group Limited, London, Great Britain. Clark, W. S. 2004. Wave moult of the primaries in acciptrid raptors, and its use in ageing immatures. Pages 795-804 in Raptors worldwide. World Working Group on Birds of Prey (R. D. Chancellor and B. U. Meyersburg, Editors). Budapest, Hungary. Edelstam, C. 1984. Patterns of moult in large birds of prey. Acta Zoologica Fennica 21:271-276. Emslie, S. D. 1988. The fossil history and phylogenetic relationships of condors (Ciconiiformes: Vulturidae) in the New World. Journal of Vertebrate Paleontology 8:212-228. Hedenstrom, a. and S. Sunada. 1999. On the aerody- namics of moult gaps in birds. Journal of Experimental Biology 202:67-76 Helbig, a. and I. Seibold. 1995. Are storks and New World vultures paraphyletic? Molecular Phylogenetics and Evolution 4:315-319. Henckel, R. E. 1981. Ageing the Turkey Vulture HY to ASY. North American Bird Bander 6:106-107. Houston, D. C. 1975. The moult of the White-backed and Ruppell’s Griffon vultures. Gyps africaniis and G. ruppellii. Ibis 1 17:474^88. Howell, S. N. G. 2001. Molt of the Ivory Gull. Waterbirds 24:438-442. Howell, S. N. G. In press. Molt in North American birds. Peterson Reference Guides. Houghton Mifflin Co., Boston, Massachusetts, USA. Howell, S. N. G. and P. Pyle. 2005. Molt, age determination, and identification of puffins. Birding 37:412-418. Jackson, J. A. 1988. Turkey Vulture, Cathartes aura. Pages 25^2 in Handbook of North American birds. Volume 4. Diurnal Raptors (Part 1). (R. S. Palmer, Editor). Yale University Press, New Haven. Connecti- cut, USA. Kirk, D. A. and D. C. Houston. 1995. Social dominance in migrant and resident vultures at carcasses: evidence for a despotic distribution? Behavioral Ecology and Sociobiology 36:232-233. Kirk, D. A. and M. J. Mossman. 1998. Turkey Vulture (Cathartes aura). The birds of North America. Number 339. Ligon, j. D. 1967. Relationships of the catharlid vultures. University of Michigan, Museum of Zoology. Occa- sional Paper 651:1-26. Livezey, B. C. and R. L. Zusi. 2007. Higher-order phylogeny of modern birds (Theropoda. Aves: Neor- nithes) based on comparative anatomy. 11. Analysis and discussion. Zoological Journal of the Linnean Society 149:1-95. Miller, A. H. 1941. The significance of molt centers 360 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 2. June 2010 among the secondary remiges in Falconiformes. Condor 43:1 13-1 15. OwRE, O. T. AND P. O. Northington. 1961. Indication of the sense of smell in the Turkey Vulture, Cathartes aura (Linnaeus), from feeding tests. American Mid- land Naturalist 66:200-205. Pyle. P. 1997. Identification guide to North American birds. Part I. Slate Creek Press. Bolinas, California, USA. Pyle, P. 2005. Remigial molt patterns in North American Falconiformes as related to age, sex, breeding status, and life-history strategies. Condor 107:823-834. Pyle, P. 2006. Stajfelmauser and other adaptive strategies for wing molt in large birds. Western Birds 37:179- 185. Pyle, P. 2008. Identification guide to North American birds. Part II. Slate Creek Press, Point Reyes Station, California, USA. Rea. a. M. 1983. Cathartid affinities: a brief overview. Pages 26-54 in Vulture biology and management (S. R. Wilbur and J. A. Jackson, Editors). University of California Press, Berkeley, USA. Rohwer. S. 1999. Time constraints and moult-breeding tradeoffs in large birds. Proceedings of the Interna- tional Ornithological Congress 22:568-581. Rosser, B. W. C. and J. C. George. 1986. Slow muscle fibres in the pectoralis of the Turkey Vulture (Cathartes aura)', an adaptation for soaring flight. Zoologischer Anzeiger 3^:252-258. SAS 1N.ST1TUTE Inc. 2002. JMP Version 5.1. SAS Institute Inc., Cary. North Carolina, USA. Shugart, G. W. and S. Rohwer. 1996. Serial descendent primary molt or Staffelmauser in Black- crowned Night Herons. Condor 98:222-233. Sibley, C. G. and J. E. Ahlquist. 1990. Phylogeny and classification of birds: a study in molecular evolution. Yale University Press, New Haven. Connecticut, USA. Snyder, N. E. R., E. V. Johnson, and D. A. Clendenen. 1987. Primary molt of California Condors. Condor 89:468^85. Stager, K. 1964. The role of olfaction in food location by the Turkey Vulture (Cathartes aura). Los Angeles County Museum Contributions in Science 81:1-63. Stewart, P. A. 1977. Migratory movements and mortality rate of Turkey Vultures. Bird-Banding 48:122-124. Stresemann, E. 1963. Taxonomic significance of wing molt. Proceedings of the International Ornithological Congress 13: 171-175. Stresemann, E. 1966. Inheritance and adaptation in moult. Proceedings of the International Ornithological Con- gress 14:75-80. Stresemann. E. and V. Stresemann. 1966. Die Mauser der Vogel. Journal Eur Ornithologie 107:1-53. Tucker, V. A. 1987. Gliding birds: the effect of variable wing span. Journal of Experimental Biology 133:33- 58. Wink, M. 1995. Phylogeny of Old World and New World vultures (Aves: Accipitridae and Cathartidae) inferred from nucleotide sequences of the mitochondrial cytochrome b gene. Zeitschrift fur Naturforschung 50:868-882. Short Communications The Wilson Joiinuil of Ornitholoi’y l22(2):36l-365, 2010 The First Reported Case of Cooperative Polyandry in the Red-footed Booby: Trio Relationships and Benefits Lei Cao,' “ Guixia Zhao,' Shan Tang,' and Hanzhao Guo' ABSTRACT. — We report the first observations of cooperative polyandry in the Red-footed Booby (Siila siila), a trio that remained stable over at least 4 consecutive years. Both males were heterosexual; neither was dominant. One male was more sexually active, but mounting activity suggested shared paternity over the years. Trio males, compared to monogamous males, invested more in pair bond maintenance activity which possibly assisted in maintaining a stable breeding unit. Trio members, like monogamous adults, shared breeding duties evenly, reducing the total investment for each trio member relative to that of monogamous adults. The extra foraging effort of the trio enabled similar provisioning rates in good and bad years. We suggest trio formation improved female lifetime repro- ductive success due to enhanced nest, egg, and chick care; maintained stable food provisioning rates despite highly heterogeneous foraging conditions; and in- creased male survivorship due to sharing of breeding duties. Red-footed Booby trio formation is probably an “opportunistic deviation” from normal monogamy which is favored by differential costs and benefits to males and females. Received 15 July 2009. Accepted 10 January 2010. Most seabirds are monogamous (Lack 1967), i.e., cooperative mating systems are rarely ob- served and there is little information on parental investment or cost and benefits to the individuals involved (Stacey and Koenig 1990, Phillips 2002, Ludwigs 2004). Brown Skuas (Stercorariiis ant- arcticiis), forming polygynous and polyandrous trios, were no more reproductively successful than monogamous pairs (Phillips 2002). A Common Tern (Sterna hiriinclo) trio raised more chicks than comparable monogamous pairs through higher brooding and provisioning rates (Ludwigs 2004). The Red-footed Booby (Siila sula) is socially monogamous showing reversed sexual size di- morphism (Nelson 1978); both adults share parental duties equally (Weimerskirch et al. ' School of Life Sciences, University of Science and Technology of China, Hefei 230027, Anhui Province, China. "Corresponding author; e-mail: caolei@ustc.edu.cn 2006). We report the first description of cooper- ative polyandry in the Red-footed Booby, a trio that nested five times in the same tree during 2004-2007. We studied trio male bond mainte- nance activities and breeding behavior to test whether trio members shared parental duties as in monogamous pairs. METHODS Study Site and Field Work. — ^Dong Island (16° 40' 03" N, 1 12° 43' 52" E) is in the northern South China Sea and supports ~35,500 pairs of Red- footed Boobies (Cao 2005). A trio nested five times during 2004-2007: in March 2004, January 2005, November 2005, December 2006 (when all nests in the colony were destroyed by a typhoon), and March 2007. Trio male bond maintenance behavior during the nest-building period was observed for 115 hrs in 2006 and 124 hrs in 2007, standardized to 100 hrs for analysis. The trio and four adjacent monogamous nests (pair numbers: N9, N1 1, N12, and N13) were studied in 2007 from nest building to chick fledging to compare adult behavior. Nest construction started between 6 and 19 March and chicks fledged from 17 to 23 August. Observation periods/breeding durations of the five nests were: trio = 893 hrs/ 1 69 days; N9 = 873 hrs/ 1 64 days; N 1 1 = 874 hrs/ 167 days; N12 = 880 hrs/ 166 days; and N13 = 803 hrs/ 158 days; activities were standardized to 1,000 hrs for analysis (heterosexual mounting was standardized to 100 hrs). Age of birds was classified based on plumage (Nelson 1978) and males and females were classified using color of bare facial parts (James 2001) and voice (Lormee et al. 2003). Trio individuals were uniquely color marked. Measurements were taken of the trio and paired adults following Lowe (1989). Studies of breeding behavior focused on bond maintenance and division of labor. Bond mainte- nance was measured as frequency of sexual activities (sky-pointing and mounting) and mate preening per unit of time. Division of labor included nest attendance (% time spent by each 361 362 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 2, June 2010 TABLE 1. Trio male bond maintenance activities during nest building in 2006 and 2007. Values rounded to the nearest whole number for clarity of presentation; NS = not significant {P > 0.05). .Activity (n/100 hrs) 2006 2007 2006 vs. 2007 CTA o-B CTA VS. CfB ctA CTB OCA vs. o-B CfA o-B Sky-pointing 4 17 P < 0.01 2 31 P < 0.001 NS P < 0.05 Heterosexual mounting 23 43 P < 0.05 23 16 NS NS P < 0.001 Homosexual mounting 3 10 P < 0.05 2 5 NS NS NS Heterosexual preening 30 10 P < 0.001 32 13 P < 0.01 NS NS Homosexual preening 17 4 P < 0.01 24 9 P < 0.01 NS NS adult), territorial defense (frequency of flight- circuiting, fighting, wing-flailing, and forward head waving; Nelson 1978), nest construction (frequency of delivery of nesting material by each bird), incubation (hrs spent by each adult; data were standardized to 300 hrs for analysis), provisioning and foraging; provisioning was measured by the percentage of days that: ( 1 ) each adult fed the chick, and (2) foraging trips were undertaken by each adult during the observation period. Provisioning and foraging data collection occurred during the late chick-rearing period after brooding ceased. Chicks were weighed at 0700, 1600, and 2000 hrs daily from 60 to 88 days after hatching (peak body mass period before fledging; Cao 2005) to learn whether they had been fed and the amount of food delivered (calculated as the difference between successive weighings). Breeding success (fledglings produced divided by the number of eggs laid) was calculated for the trio and the general population by checking nests every 5 or 10 days. Data affected by the super typhoon in December 2006 were omitted from the calculation, as a typhoon of this strength, which destroyed all nests in the colony, is a rare occurrence (about once every 20 years). Statistical Analyses. — ^The binomial distribution was used to examine differences between activity levels of trio individuals and between those of trio and monogamous individuals (Tables 1, 2) using the equation: : = \p - 0.5l/y{ Var(y;)) based on the null hypothesis that a p-value of 0.5 is expected if there is no difference between activity levels (p = proportion of the combined activities attributable to one individual and Var(/7) — p(\ — p)/u \u = combined activities of the 2 individuals]). We calculated the number of standard deviation units (-) that the activity level of the trio member (.Y) was from the mean activity level of the monogamous sample (.v) using the equation z = (x - a:)/s, where s = standard deviation for the monogamous sample, to examine differences between division of labor activity levels of trio and monogamous pair individuals (Table 2). We used the r-score for sample sizes greater than 30, and the f-distribution for smaller sample sizes (Fowler and Cohen 1996); all tests were two-tailed. The r-test was used to examine differences between adult measurements (mass, wing length, etc) and between food deliveries to chicks. Data are presented as mean ± SD. RESULTS Measurements. — ^The trio consisted of adult males o*A and crB, and an adult female (9A). Body masses of o*A (948 g) and 9A (961 g) were similar, ~ 10% heavier than CB (856 g), but body masses of trio members were not different (P > 0.05) from those of paired adults (male = 852 ± 47.1 g. n = 17; female = 956 ± 77.9 g, n = 14); differences were not significant (P < 0.05) between culmen lengths (crA = 84.0 mm; crB = 85.5 mm; 9A = 85.0 mm; paired adults: male = 82.7 ± 2.64 mm, n = 17; female = 84.7 ± 2.94 mm, n - 14) or wing lengths {CA = 386 mm; o*B = 373 mm; 9A = 377 mm; paired adults: male = 377 ± 8.1 mm, n = 17; female = 382 ± 5.9 mm, n = 14). Breeding Behavior. — ^Trio males were hetero- sexual (Table 1); no agonistic behavior was observed, even when both males were present during copulation and sky-pointing. Male A mate- preened significantly more and o*B sky-pointed significantly more in both years, but ctB did significantly more mounting than cA in 1 year. Male A exhibited similar bond maintenance behavior in both years, but cB sky-pointed significantly more and did less heterosexual mounting in 2007. The trio indulged in far more bond maintenance activity than monogamous pairs, but frequency of heterosexual mounting was not significantly different (Table 2). SHORT COMMUNICATIONS 363 TABLE 2. Breeding behavior of trio and monogamous pair members during nest building to chick fledging in 2007. Values rounded to the nearest whole number for clarity of presentation; trio male duplicate data are for CfA and CB, respectively; NS = not significant {P > 0.05). Trio N9 Nil N12 NI3 p Sky-pointing (;;/ 1,000 hrs) Male(s) 5/48 0 10 8 0 NS/P < 0.01 Female 0 1 0 0 0 Total 53 1 10 8 0 P < 0.01 Heterosexual mounting (/t/100 hrs) Male(s) 23/16 23 34 30 14 NS/NS Total 39 23 34 30 14 NS Mate preening (between sexes) Male(s) 139/36 2 17 0 25 P < 0.01/NS (n/ 1.000 hrs) Female 3/10 12 13 0 44 NS/NS Total 188 14 30 0 68 P < 0.05 Nest attendance (hrs/ 1,000 hrs) Male(s) 384/486 397 311 372 408 NS/NS Female 411 404 488 371 467 NS Total 1,281 801 800 743 875 P < 0.01 cr/Q ratio 30:38:32 50:50 39:61 50:50 47:53 NS Territorial defense (/7/l,000 hrs) Male(s) 184/258 111 114 119 93 P < 0.01/P < 0.01 Female no 136 135 98 103 NS Total 551 247 249 217 197 P < 0.01 c/g ratio 33:47:20 45:55 46:54 55:45 47:53 NS Nest construction Nest material supply (/;/ 1,000 hrs) Male(s) 82/78 235 237 256 54 NS/NS Female 1 6 I 2 3 NS Total 161 241 238 258 56 NS Nest building («/l,000 hrs) Male(s) 24/27 13 5 8 5 P < 0.05/P < 0.05 Female 108 192 212 240 44 NS Total 159 205 217 248 49 NS Totals cf/g ratio 33:33:34 56:44 53:47 52:48 56:44 NS Incubation (hrs/300 hrs) Male(s) 69/89 125 118 151 137 P < 0.05/NS Female 142 175 182 149 163 NS ni sC d p d 00 p o ■3 c -3 + 1 T +1 d + 1 t 00 — ll ^ cn p p p ■3 in — sd ml d L. o 1 CJ C- r? -3 — d cn 3 U oo p cn u D. (N — ' rd — r-i r- — -3 + 1 d + 1 d + 1 sd J 3 in m, in m ' — ' E oc — p m, o > G in d d m in in m mi mi T3 u. r3 — m — oo (N i> “3 o m •— 00 3 "t in 00 p p ri p Olj •>2 oc rd o sd in p mi p d m . zz +1 + 1 + ! d + 1 d + 1 mi + 1 d +1 d c5 < NO 00 00 r' Uh 0^ r- O'! 00 nj — r- 4; ON si m 1 p r*" oc d — ! d d d 00 d sd 00 d 00 — sC m, NO mj r- m, sO m> sO in 'J'. ■73 15 u. w ■3 Q _ E 00 m s. •^C/5 5 o in ON — — 00 ^ +1 u _ - Wl 3 ni in p o r| p 'o d 00 ni OO ni p 00 00 d — ri — ■« « 3 3 + 1 d in + 1 d + 1 d +1 nO + 1 sd L. f-^ Cti U W c. nj + 1 T T n J. 1 oc 00 p r| 00 p ri 00 rr 00 3 C d d d sd d 00 d sd d sd iT. 5 -u -u O- y: E in — 00 — > V .£Z *3 Q. m o ri n p E p p p r-- 00 E ir w in d p 00 d ml O mi p d p 3 o +1 ml + 1 + 1 d + 1 3 + 1 + 1 oo On 00 00 — G 1 in o m. r- 1 O p — ON p ci LU 3 d ni d d sd d ri sd sd ri d oc r- sO 00 ON 00 sO r^ NO CQ < H o JC V5 u 1) L. u a 00 fU a. c/5 "3 0/ 3 < a. (i> cn o o O) u. O O cn r- o ri r' -i* TD 1 > CJ •0 Co CJj u. U 5 Q c "O (U 0 -a 0 -5 0 "2 0 ■y‘. >3 (J Sc c "Tv C3 CO *C3 0 r-“ ID u u § CJJ c c G C P 0 tS X) •o ID § CO U OC DISCUSSION Measured flight speeds for grebes are almost non-existent, and no data for either the Red- necked or Horned grebe are presented in recent Birds of North America compilations (Stout and Nuechterlein 1999, Stedman 2000). We found flight speeds in windless conditions averaged 61.2 km/hr for Red-necked Grebes and 55.5 km/hr for Horned Grebes (vs. 60 km/ hr for Eared Grebes [P. nigricollis]; Cullen et al. 1999). Air speeds for both species increased through the fall migration period, even when tail winds were considered. This increase may have relevance to the question of whether migrants fly at minimum power velocity (V^p), maxi- mum distance velocity (V,na), or somewhere in between (e.g., Alerstam 1990, 2003). Tail wind data, for example, suggest that Red-necked Grebes fly slower in August, just as they are arriving at the Great Lakes (where many stop and molt), and that speeds increase progressive- ly in Septeir,ber and October, as the grebes prepare to leave for wintering areas on the Atlantic Coast. The Horned Grebe had a similar pattern for September-October. This pattern might indicate that grebes in early fall empha- size energy conservation, whereas in late fall migration, when time to reach wintering areas is constrained, the ability to fly faster and cover long distances rapidly is more important. However, larger birds necessarily fly faster than smaller birds of the same species, because flight speed is proportional to mass (Rayner 1988). Data on weight changes in migrating Red- necked and Horned grebes are not available. Eared Grebes lay on large fat stores while at fall staging areas, so that body mass is 50% greater on departure than on arrival (Jehl 1988, 1997; Jehl et al. 2003; Jehl and Henry 2010). If Red- necked and Horned grebes show a similar pattern, the expected increase in air speed between arriving and departing birds would approximate 5%, which is much less than we found in early versus late fall samples. Addi- tional data on mass and body composition of these species at different seasons are needed to move the discussion forward. Regardless of whether the increased speed is due to behavioral adjustment, increased body mass, or both, seasonal differences in flight speed can be expected. Detailed interspecific or intra- 378 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 2, June 2010 specific comparisons must use data from the same time period or from birds in the same physiolog- ical condition. For this reason, the data in Table 1 should be used cautiously. ACKNOWLEDGMENTS We thank I. D. Wijayratne and R. S. Haataja for surveying; Jaw-chuan Lin for statistical analyses; J. V. Remsen. S. J. Stedman. and J. R. Jehl Jr. for reviews; Amy Slagle, Don Slagle, and M. J. Hester for logistical support; and the Copper Country Audubon Club (Houghton, Michigan, USA) for funding. Jehl and Stedman revised the final manuscript after the unexpected death of the senior author (LCB). Original data on flight speed and conditions are filed at the Museum of Natural History, Louisiana State University, Baton Rouge. LITERATURE CITED Alerstam, T. 1990. Bird migration. Cambridge University Press, Cambridge, United Kingdom. Alerstam, T. 2003. Bird migration speed. Pages 253-267 in Avian migration (P. Berthold, E. Gwinner, and E. Sonnenschein, Editors). Springer-Verlag, Berlin, Ger- many. Binford, L. C. 2006. Birds of the Keweenaw Peninsula, Michigan. Miscellaneous Publication Number 195. Museum of Zoology, University of Michigan, Ann Arbor, USA. Cullen, S. A., J. R. Jehl Jr., and G. L. Nuechterlein. 1999. Eared Grebe {Podiceps nigricollis). The birds of North America. Number 433 Jehl Jr., J. R. 1988. Biology of the Eared Grebe and Wilson’s Phalarope in the nonbreeding season: a study of adaptations to saline lakes. Studies in Avian Biology 12:1-74. Jehl Jr., J. R. 1997. Cyclical changes in body composition in the annual cycle and migration of the Eared Grebe {Podiceps nigricollis). Journal of Avian Biology 28:132-142. ^ Jehl Jr., J. R. and A. E. Henry. 2010. The postbreeding migration of Eared Grebes. Wilson Journal of Ornithology 122:217-227. Jehl Jr., J. R., A. E. Henry, and H. I. Ellis. 2003. Optimizing migration in a reluctant and inefficient flier: the Eared Grebe. Pages 199-209 in Avian migration (P. Berthold, E. Gwinner, and E. Son- nenschein, Editors). Springer-Verlag, Berlin, Ger- many. Rayner, j. V. M. 1988. Form and function in avian flight. Current Ornithology 5:1-66. SCHNELL, G. D. 1965. Recording flight-speed of birds by Doppler radar. Living Bird 4:79-87. Stedman, S. J. 2000. Horned Grebe (Podiceps cmritus). The birds of North America. Number 505. Stout, B. E. and G. L. Nuechterlein. 1999. Red-necked Grebe (Podiceps grisegena). The birds of North America. Number 465. The Wilson Journal of Ornithology 122(2):378-380, 2010 Thermoregulatory Behavior in Migratory European Bee-eaters {M crops apiaster) Reuven Yosef ABSTRACT. — Birds in hyper-arid environments have acute problems of energy and water balance, and thermoregulate both physiologically and behaviorally. 1 report on European Bee-eaters (Merops apiaster) engaged in a previously unreported thermoregulatory behavior of diving into the sea and in salt ponds with high levels of salinity. This behavior may also explain the previously reported, but unexplained, finding of bee- eaters inside a tiger shark (Galeocerdo ciivier) in the Red Sea. The.se observations should instigate future experiments on the subject of .selective use of salt water for evaporative cooling and thermoregulatory behavior by desert birds. Received 27 August 2009. Accepted 9 December 2009. ' International Birding and Research Centre in Eilat, P. O. Box 774, Eilat 88000. Israel; e-mail: ryo.sef@eilatcity.co.il Animals in deserts use a variety of mechanisms to maintain their body temperature within phys- iologically acceptable limits and minimize water loss (Louw and Seely 1982). Heat stressed birds engage in panting or gular fluttering to affect evaporative cooling (Tieleman 2002, pers. obs.). Other physiological adaptations in hot environ- ments include greater tolerance of tissues to high temperatures and dehydration, heterothermy (fluc- tuation of body temperature in response .to environmental stress), reducing insulation, and appropriate coloration (Louw and Seely 1982, Maclean 1996, Tieleman 2002). Behavioral thermoregulation is a suite of behavioral activities that minimize the exposure of the organism to heat stress and reduce water loss at high temperatures (Wunder 1979). It is SHORT COMMUNICATIONS 379 common to see birds exploit cooler terrestrial micro-climates and bathe in water to remove body heat by evaporative cooling (Dawson and Barthol- omew 1968, Cook 1997). The International Birding and Research Centre in Eilat, Israel has reclaimed a local garbage dump as a Bird Sanctuary to help migratory bird populations (Cherrington 2001, Tal 2002, Ro- senzweig 2003). The site has commercial salt ponds and a fresh water lake that are less than 50 m apart. Eilat is at the southernmost tip of Israel within the Sahara-Arabian desert belt characterized by extreme temperatures (mean = 9.5 to 39.4° C; Israel Meteorological Service, www.ims.gov.il), very low precipitation (mean annual rainfall = 17 mm), and high solar radiation (mean = 6.32 kWh m7day, Eaiman et al. 2003). I report observations of migrant European Bee- eaters (Merops apiaster) diving into the sea and salt ponds with high levels of salinity (>15%). OBSERVATIONS European Bee-eaters migrate through the Eilat bottleneck in both seasons in large numbers. They are among the last migrants in spring (May-Jun) and their passage at Eilat coincides with the peak heat waves of late spring-early summer when ambient temperatures can range from 1 8 to 48° C. Temperatures peak on many occasions during periods of khamsins, dry, hot, and dusty local wind that carries great quantities of sand and dust from the deserts with speeds to 160 km/hr, and a rise of temperatures as much as 20° C in 2 hrs. I observed 57 flocks of bee-eaters diving into the shallow salt ponds in the Bird Sanctuary or off-shore in the Red Sea in spring 2005-2009 for a total of 36 khamsin days with temperatures above 43° C. Similar observations were also made from the shores of the Red Sea. However, the outcome of this behavior far out at sea was too distant to describe properly. Bee-eaters were also present on days when temperatures were under 42° C, but were not ob.served diving into the salt ponds. All observations were conducted with Swarovski 15 X 56 binoculars. Bee-eaters perched in the shade of Prosopis alba and Acacia radciiana trees surrounding the salt pond at the southern extremities of the Bird Sanctuary. I considered them heat stressed because of their open beaks, tongues thrust out and gular fluttering. Every 15-20 min small groups of bee-eaters, 8-39 birds/flock, would take off and dive head first into the shallow salt pond. On most occasions the whole flock would take off in a circling Bight, calling, and diving into the salt pond. Rarely did single birds remain behind when the flock was engaged in diving. The bee-eaters flew into the shadiest parts of the trees and perched there for the next 30-45 min after flying from the water, or paddling to the shoreline. The bee-eaters spread their wings while wet and allowed the breeze and the low humidity of the area (15-18%) to dry the feathers during the period immediately after they perched. I believe the birds achieved evaporative cooling since bee-eaters did not engage in gular fluttering for the first 15-20 min after diving. The maximum depth of the salt pond was 200 mm. I observed 576 bee-eaters diving into the ponds with 389 flying from the water surface and 1 1 1 unable to do so and had to paddle ashore where they lay on the ground for several minutes with their wings stretched out before flying to a perch. The remaining 76 bee-eaters were dazed for many minutes, possibly from hitting the bottom of the pool. They surfaced from the dive but did not move, and had trouble reaching the shoreline. These birds were collected, rehabilitat- ed (allowed to rest in the shade and fed sugar water), and returned to the wild a day or two later. On days when I was not present at the Bird Sanctuary, 117 carcasses washed up on shore suggesting this is a common phenomenon. I saw no bee-eaters diving into the nearby fresh water lake which is much deeper (~ 4 m), and which has fish. DISCUSSION Ery (1984, 2001) reviewed bathing (a.k.a. splashing) by .several species of bee-eaters. The major reasoning forwarded by most published reports was that bee-eaters were feeding on insects or fish. Ery (2001:313) states that cooling, dislodging parasites, drinking or cleansing are alternatives to what seems “to be essentially some .sort of social comfort activity" in the gregarious bee-eaters. My observations of bee-eaters diving into salt ponds, devoid of fish or other possible food sources, during the hottest days of the season strengthens the argument that in the desert “splashing" has a thermoregulatory function. The described behavior may explain the finding of bee-eaters inside a tiger shark (Galeocerdo cuvier) in the Red Sea (Yosef et al. 2002). Perhaps the Bee-eaters involved had dived into the Red Sea for cooling and were unable to take off from 380 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 2. June 2010 the surface and were eaten by the shark, or had drowned and collected from the body of water. The latter appears to be the more realistic option because tiger sharks are considered to be noctur- nal foragers (Compagno and Niem 1998). The ability of some birds to discriminate fresh and saline waters is debated (Maclean 1996). However, most studies have investigated the tolerance of birds to saline water - drinking tolerance, body mass loss, serum and urine osmolarity, and fecal water loss (Maclean 1996). It has also been suggested that a possible added value could be the removal of ectoparasites in the salt water. However, this has yet to be proved. I was unable to find literature on how birds use water sources and modify their behavior to thermoregulate in extreme temperatures. King and Chastant (2008) report on cormorants (Pha- lacrocorax spp.) and darters (Auhinga spp.) that regurgitate water into their chicks open mouths and over their bodies on warm and windless days during the hottest hours to help them maintain their body temperatures. We still lack information on the ability of seawater to have a cooling effect on heat-stressed birds. My observations should instigate future experiments on the subject of evaporative water loss and thermoregulatory behavior in desert birds. ACKNOWLEDGMENTS I thank Berry Pin.show for di.scu.ssing the .subject. Susan Craig, Keith Bildstein, and an anonymous reviewer improved an earlier version of the paper. LITERATURE CITED Cherrington, M. 2001. To save a watering hole. Pages 8- 14 in The best American science and nature writing 2001 (E. O. Wilson, Editor). Houghton Mifflin Company, New York, USA. Compagno, L. J. V. and V. H. Niem. 1998. Carcharhinidae. Requiem sharks. In FAO identification guide for fishery purposes. The living marine resources of the western central Pacific. (K. E. Carpenter and V. H. Niem, Editors). FAO, Rome, Italy. Cook, W. E. 1997. Avian desert predators. Springer- Verlag, Heidelberg, Germany. Dawson, W. R. and G. A. Bartholomew. 1968. Temperature regulation and water economy of desert birds. Pages 357-394 in Desert biology. Volume 1 (G. W. Brown Jr., Editor). Academic Press, New York, USA. Faiman, D., D. Feuermann, P. Ibbetson, A. Zemel, A. Ianetz, V. Liubansky, and I. Setter. 2003. Data processing for the Negev radiation survey: Ninth year. TMY-V4. Israel Ministry of Energy and Infrastructure, Jerusalem. Fry, C. H. 1984. The bee-eaters. T & AD Poyser, Staffordshire, United Kingdom. Fry, C. H. 2001. Family Meropidae (Bee-eaters). Pages 286-341 in Handbook of the birds of the world. Volume 6, Mousebirds to Hornbills (J. del Hoyo, A. Elliott, and J. Sargatal, Editors). Lynx Edicions, Barcelona. Spain. King, D. T. and J. E. Chastant. 2008. Water regurgitation by adult Double-breasted Cormorants: a possible mechanism to assist in chick thermoregulation. Waterbirds 31:283-284. Louw, G. N. AND M. K. Seely. 1982. Ecology of desert organisms. Longman House, Essex, United Kingdom. Maclean, G. L. 1996. Ecophysiology of desert birds. Springer-Verlag, Heidelberg, Germany. Rosenzweig, M. L. 2003. Win-win ecology-how the earth’s species can survive in the midst of human enterprise. Oxford University Press, New York, USA. Tal. a. 2002. Pollution in a promised land-an environ- mental history of Israel. University of California Press, Los Angeles, USA. Tieleman, B. I. 2002. Avian adaptation along an aridity gradient-physiology, behavior and life history. Dis- sertation. Rijks Universiteit, Groningen, The Nether- lands. Wunder, B. a. 1979. Evaporative water loss from birds: effects of artificial radiation. Comparative Biochem- istry and Physiology Part A: Physiology 63:493-494. Yosef, R., D. Zakai, M. Rydberg-Heden, and R. Nikolajsen. 2002. An unusual record of a European Bee-eater Merops apiaster from Eilat-inside a tiger shark Galeocerdo ciivier. Sandgrouse 24:140-142. SHORT COMMUNICATIONS 381 The Wilson Joiinial of Ornithology 122(2):38l-384, 2010 Population Status of Chuck- will’ s-widow (Caprinmlgus carolinensis) in the Bahamas William K. Hayes, Elwood D. Bracey,“ Melissa R. Price,' Valerie Robinette,' Eric Gren,' and Caroline Stahala-^ ABSTRACT. — The Chuck-wiH’s-widow {Caprinuil- giis carolinensis) in the Bahama Islands has been regarded as a rare to uncommon winter visitor. We conducted breeding season surveys on the three largest northern islands (North Andros, Grand Bahama, and Great Abaco) to examine the status of this species. We encountered singing birds on most survey routes on all three islands, suggesting that sizeable breeding popula- tions are widespread in the northern Bahamas with an aggregate estimate of 500-1.000 pairs. Our density estimates were somewhat less than those from the primary range in the United States, suggesting either a lower carrying capacity in the Bahama Islands or recently established populations that have yet to reach carrying capacity. Received 24 August 2009. Accepted 3 January 2010. Much remains to be learned about Chuck- wilTs-widow (Caprimitlgus carolinensis) and other nightjars (Caprimulgiformes), largely be- cause of their nocturnal and secretive habits. The species breeds across much of the eastern United States and winters in southern Florida, Central America, the West Indies, and northern South America. Gaps remain in our knowledge of the species’ distribution, abundance, and habitat use despite its distinctive nocturnal song, which greatly facilitates detection (Straight and Cooper 2000). The species has gradually expanded its range in North America in the past century (Straight and Cooper 2000). The status of Chuck-will’s-widow in the Bahama Islands remains unclear. The species is considered a rare to uncommon winter visitor (Brudenell-Bruce 1975, Emien 1977, Buden 1987, White 1998, Straight and Cooper 2000). Northrop (1891) collected a male with enlarged testes ' Department of Earth and Biological Scicncc.s, Loma Linda University, Loma Linda, CA 92350, USA. -2305 Royal Palm Way, Treasure Cay, Abaco. Bahamas. -’Department of Biological Science, Florida State Uni- versity, Tallahassee. FL 32306, USA. “'Corresponding author, e-mail: whayes@llu.edu during the breeding season (15 May) on Andros, but was uncertain of nesting. Paterson (1972) observed a bird incubating a single egg on Andros, but Bond (1973:3, 1984:20) rejected the record as a misidentified Antillean Nighthawk {Chordeiles gundlachii). However, Buden (1992) and Buden and Sprunt (1993) heard calling birds and flushed a pair during visits to several Exumas cays in April through June, suggesting possible breeding. Calling birds subsequently heard from Eebruary to July on several islands, documented in North American Birds seasonal reports, led to increased speculation of nesting. Breeding was eventually confirmed for Grand Bahama (Norton 1999; Norton and White 2001, 2002) and Abaco (Norton 2000). However, several questions re- main unanswered. Were these breeding records extralimitai, or indicative of larger breeding populations? If the latter, how large might these populations be, and how widespread are they? The objective of our study was to clarify the population status of Chuck-will’s-widow in the Bahama Islands. Our results, when compared to similar surveys from the species’ continental range in the United States, provide inferences about the bird’s abundance and breeding status. METHODS We conducted breeding season surveys on the three largest northern islands: North Andros, Grand Bahama, and Great Abaco. We followed generally practiced protocols for surveys of nightjars (e.g.. Cooper 1981, Wilson 2008). Surveys were conducted along 8-20 km routes that included 9-12 brief stops (usually 10) of 2- 5 min duration at 1 km or greater intervals. Surveys, with a few exceptions, were completed during >50% moon illumination. Surveys were conducted during either the hour before sunrise or within 1-2 hrs following sunset, depending on when the moon was above the horizon. Weather varied from clear to overcast but with relatively low wind and absence of rain. We recorded the 382 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 2, June 2010 GPS (Global Positioning System) coordinates, time, and number of calling birds for each stop. We broadcast a recording of the song by speaker or whistled a cruder imitation of the song by mouth for most stops. We did not differentiate between spontaneous and solicited calls, as both were frequently detected. Calling birds could at times be heard at >500 m; we tried not to double- count individuals on consecutive stops, but this may have resulted in slight underestimation bias. We (MRP, VR, and EG) conducted seven surveys on North Andros during 2009. These included: 10 and 12 May (mornings) south of Nichol’s Town; 17 May (morning) between Mastic Point and San Andros Airport; 2 June (evening) west of Stafford Creek; and 3, 4, and 5 June (evenings) between Stafford Creek and Love Hill. These routes traversed primarily Caribbean pine (Piiuis carihaea) forest habitat with scattered openings, but coppice, marsh, and residential areas were present at some stops. WKH conducted three surveys during 2007 on Grand Bahama. These included: 1 July (morning) from Bassett Cove to the oil storage facility east of Riding Point; 1 July (evening, 12 stops, no moonlight) near the water towers at Lucayan Estates; and 2 July (morning, 9 stops) in Lucaya along West Beach and Midshipman roads. All routes were in primarily pine forest with scattered openings; however, the first included occasional coppice and residential areas, and the third transected a rural residential area. EDB conducted repeated surveys along two routes (9 stops each) between Treasure Cay and Leisure Lee on Great Abaco during 2009. These included four evening surveys on 5 May, 10 May (no moonlight), I June, and 5 June. Additional evening surveys repeated in July and August yielded no calling birds and were excluded from analysis. Routes included pine forest with .some coppice. CS conducted two surveys in 2009 further south on Great Abaco. These were 14 July (morning) along Great Abaco Highway between Crossing Rocks and Abaco National Park, and 15 July (morning) along Mid Road towards Hole-in-the-Wall. The routes passed through pine forest with variable understory, and several stops included coppice. We derived the following variables for each route: total number of calling birds; mean birds per stop along route; proportion of stops with bird detections; number of birds calling per occupied stop (with birds calling); and density (pains/knr). Numbers were adjusted for routes of 9 or 1 2 rather than 1 0 stops to reflect the standard 1 0-stop route (i.e.. birds/stop multiplied by 10). We assumed that all calling individuals were detected within a 500-m radius to estimate population density (calling males = pairs/kni"). Thus, the number of calling birds detected at each stop corresponded to a minimum density of pairs per 0.785 km- (area of circle with 500 m radius). This assumption is subject to .several factors influencing detection rate, including some birds being heard at greater distances and many individuals not calling during the brief stops. RESULTS We encountered 42 calling birds along the 14 survey routes (Table 1 ). The number of calling birds per 10-stop route varied from zero to 12.2. Calling birds were detected on four of the seven (57%) Andros routes and on 100% of the three Grand Bahama and four Abaco routes. Grand Bahama appeared to host the highest density of individuals. One nine-stop route through a rural residential area yielded 1 1 calling birds. The maximum number of calling birds recorded per stop was two, and this was consistent for all three islands. Three birds were heard at several stops on Grand Bahama but, in each case, one was judged to be a bird heard at the previous stop and was not counted. Repeated surveys along two routes on Great Abaco illustrated nightly and .seasonal variation. Calling birds after sun.set numbered four (5 May), five (10 May), six (1 Jun), and seven (5 Jun) for the 18 stops of both routes combined. No birds were detected on three additional surveys (22 and 30 Jul, 5 Aug), and these surveys were excluded due to lateness of season. Our mean estimate of calling birds for the first four surveys (5.5) was 21% below the maximum count (7), suggesting detection of 79% or less for individual surveys. DISCUSSION Standardized surveys provide the best under- standing of the relative abundance of Chuck- will's-widow, and only males are known to give the primary song (Straight and Cooper 2000). A number of factors can affect calling rates and influence detection rates and density estimates. Male calling peaks in late spring in continental populations and may decrease markedly in July (e.g.. Cooper 1981, Straight and Cooper 2000). Both calling rale and duration of calling through- TABLE I. Chuck-will's-widows on North Andros (7 survey routes). Grand Bahama (3 routes), and Great Abaco (4 routes), Bahama Islands. Values include .v ± SD and rangi within parentheses for birds per 10-stop route and associated variables. SI lORT COMMUNICATIONS 383 in NO 'e 00 f^i + 1 ? in CO ri 'E d d c. 00 o .3 — — c _ — * Q +1 +1 +1 00 00 — CO o d d c. •y; n O m 1 CO O r> 1 a — o CN 33 ;§ 3}- $ ro mi d. + 1 + 1 + 1 V‘. ON CO NO # oi GO CO n oT m, rn — d CL d 1 1 ri ! c o d y; C 00 00 CL in o d d d 'El +1 + 1 +1 -t o o — in d r\ r i 1 1 mj 1 £ 1 O 1 n 1 C, 00 CO y. -o — ini d 5 + 1 + 1 + 1 vO o mi d E o sC sO "E H ■X. CL c y o — OO r- CO O h- ir. Q d U. CJ £ a -C C 3 < < •o C3 1 . U O QD a P z a 6 u > •C 'U -3 O -a OJ rr. 3 -D a. V 3 c c f- Gill the night increase significantly with greater moonlight (Cooper 1981, Mills 1986, Wilson and Watts 2006, Woods and Brigham 2008). Our surveys suggest a substantia) breeding population of Chuck-will’s-widows exists on the large northern islands in the Bahamas. Abundance varied locally, but calling males were widespread on each of the large islands surveyed (North Andros, Grand Bahama, and Great Abaco). Birds calling in summer on New Providence (e.g., Norton and White 2002) and the Exumas (Buden 1992, Buden and Sprunt 1993) suggest breeding populations on these islands as well. Relative densities from our surveys (,v = 0.14- 0.56 birds/stop among the 3 islands) were generally lower than surveys within the primary range of continental populations. Cooper (1981), from 20 surveys in northern Georgia, reported 0- 1.75 calling individuals per stop {x = 1.37 for 6 nights with >0.5 lunar illumination). James and Neal (1986) listed counts of 1.77 and 2.32/stop in Searcy and Clay counties, Arkansas, respectively. Baumgartner and Baumgartner (1992) reported counts of 1.67 and 1.98/stop in northeast and southeast Oklahoma, respectively. More recently, Wilson (2008) reported on coordinated surveys from seven states having two or more surveys conducted in 2007. These included: Florida with 1.54 birds/stop, Alabama with 1.50, South Car- olina with 1.30, Georgia with 0.98, Mississippi with 0.90, Virginia with 0.44, and North Carolina with 0.42. Our surveys differed somewhat from tho.se of Wilson (2008) in that stops were spaced more clo.sely and listening time was briefer. We also used tape playback or vocal imitation to help stimulate calling. We do not know whether the current breeding population in the Bahama Islands represents a recent colonization or has been overlooked. The relatively low densities compared to continental populations suggest either a lower carrying capacity in the Bahama Islands or recently established populations that have yet to reach carrying capacity. Population densities at the expanding northern periphery of the continental range are generally low (Hunt 2008). Future surveys could improve our knowledge on these possibilities. Two birds collected from Cuba during the breeding .season with enlarged gonads suggest possible breeding there as well (Straight and Cooper 2000). Our density estimates provide for a rough approximation of the total population in the 384 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 2. June 2010 Bahamas. The estimates are limited by three major factors (see also Wilson 2008): (1) some birds were detected at distances >500 m, resulting in overestimation; (2) not all males present were calling, resulting in underestimation; and (3) survey routes may not have been representative of the habitats available on each island. We suspect our numbers are underestimates. We multiplied our estimates by the quantity of suitable habitat on each island based on detailed maps. As many as 214 pairs exist on Andros (0. 18 pairs/km“; 5,957 km" total area; 20% of total area assumed to be suitable habitat), 390 pairs on Grand Bahama (0.7 1 pairs/km"; 1 ,373 km" total area; 40% suitable habitat), and 192 pairs on Abaco (0.38 pairs/km"; 1,681 km" total area; 30% suitable habitat). New Providence (207 km") presumably hosts a breeding population, and smaller popula- tions may exi.st on other islands. We conclude the Bahama Islands supports an aggregate population in the vicinity of 500-1,000 pairs. Clearly, the recent scattered reports of singing and nesting birds reflect sizeable breeding popu- lations rather than extralimital records. Efforts should be undertaken to locate additional popula- tions on other islands and to examine whether populations are gradually increasing in size. ACKNOWLEDGMENTS We thank the following for their .support: Forfar Field Station and Insular Species Conservation Society (Andros research); Grand Bahama Power Company (Grand Ba- hama): and Bahamas National Trust and Friends of the Environment (Abaco). LITERATURE CITED Baumgartner, F. M. and A. M. Baumgartner. 1992. Oklahoma bird life. University of Oklahoma Press, Norman, USA. Bond, J. 1973. Supplement 18 to the Check-list ot birds of the West Indies (1956). Academy of Natural Sciences of Philadelphia. Pennsylvania. USA. Bond, J. 1984. Supplement 25 to the Check-list ot birds of the We.st Indies (1956). Academy of Natural Sciences of Philadelphia. Penn.sylvania. USA. Brudenele-Bruce. P. G. C. 1975. The birds of New Providence and the Bahama Islands. Collins, London, United Kingdom. Buden. D. W. 1987. The birds of the southern Bahamas. British Ornithologists’ Union, London, United King- dom. Buden, D. W. 1992. The birds of the Exumas, Bahama Islands. Wilson Bulletin 104:674-698. Buden, D. W. and A. Sprunt IV. 1993. Additional observations on the birds of the Exumas, Bahama Islands. Wilson Bulletin 105:514—518. Cooper, R. J. 1981. Relative abundance of Georgia caprimulgids based on call-counts. Wilson Bulletin 93:363-371. Emlen, j. T. 1977. Landbird communities of Grand Bahama Island: the structure and dynamics of an avifauna. Ornithological Monographs 24. Hunt, P. 2008. Northeast nightjar survey: 2008 results. Report to the New Hampshire Audubon Society, www. vtecostudies.org/PDF/Nightjar08.pdf (accessed 4 Au- gust 2009). James, D. A. and J. C. Neal. 1986. Arkansas birds: their distribution and abundance. University of Arkansas Press, Fayetteville, USA. Mills, A. M. 1986. The influence of moonlight on the behavior of goatsuckers (Caprimulgidae). Auk 103:370-378. Northrop, J. I. 1891. The birds of Andros Island, Bahamas. Auk 8:64-80. Norton, R. L. 1999. West Indies Region. North American Birds 53:436-437. Norton, R. L. 2000. West Indies. North American Birds 54:426-427. Norton, R. L. and A. White. 2001. West Indies. North American Birds 55:370-372. Norton, R. L. and A. White. 2002. West Indies. North American Birds 56:496-M97. Paterson. A. 1972. Nesting of Chuck-will’s-widow on Andros Island. Bahamas. Auk 89:676-677. Straight, C. A. and R. J. Cooper. 2000. Chuck-will’s- widow (Caprimiilgiis ccirolincn.'iis). The birds of North America. Number 499. White, A. W. 1998. A birder's guide to the Bahama Islands (including Turks and Caicos). American Binding Association, Colorado Springs, Colorado, USA. Wilson, M. D. 2008. The Nightjar Survey Network: program construction and 2007 Southeastern Nightjar Survey results. Center for Conservation Biology Technical Report Series, CCBTR-08-001 . College of William and Mary, Williamsburg, Virginia, USA. W1L.SON, M. D. AND B. D. Watts. 2006. The effect of moonlight on detection of Whip-poor-wills: implica- tions for long-term monitoring strategies. Journal of Field Ornithology 77:207-21 1. Woods, C. P. and R. M. Brigham. 2008. Common Poorwill activity and calling behavior in relation to moonlight and predation. Wilson Journal of Ornithol- ogy 120:505-512. SHORT COMMUNICATIONS 385 The Wilson Journal of Ornithology 122(2):385-387, 2010 Yellow Rails Wintering in Oklahoma Christopher J. Butler,' " Lisa H. Pham,' Jill N. Stinedurf,' Christopher L. Roy,' Eric L. Judd,' Nathanael J. Burgess,' and Gloria M. Caddell' ABSTRACT. — Yellow Rails (Cotnrnicops novebor- acensis) were recently discovered to migrate through southeastern Oklahoma in small numbers during fall with a few records into December and a single record in January. We made seven trips (3 in Nov 2008 and 1/ month from Dec 2008 through Mar 2009) to Red Slough Wildlife Management Area in McCurtain County (Oklahoma) to catch and band Yellow Rails. Twenty- five Yellow Rails were banded and birds were observed during each month. Rails were encountered in areas dominated by Sporohohts spp. averaging 44 cm in height in areas with 4 cm or less of standing water. Yellow Rails appear to overwinter in small numbers in McCurtain County, Oklahoma —300 km north of the Gulf Coast. Received 22 June 2009. Accepted 15 October 2009. Yellow Rails (Coturnicops noveboracensis) are secretive, nocturnal birds that breed in the northern United States and Canada with a disjunct population in southcentral Oregon (Bookhout 1995). The total population of Yellow Rails is estimated to be only 17,500 individuals (Butcher et al. 2007). The species is on the Audubon’s Society’s Red WatchList (Butcher et al. 2007) and is a species of special concern in Canada (Environment Canada 2006), and most of the states where breeding occurs (Grace et al. 2005). Yellow Rails generally leave breeding areas between mid-August and November (Stenzel 1982, Bookhout 1995, Popper and Stern 2000, Goldade et al. 2002) and winter along the coast, from Texas to North Carolina (Bookhout 1995). Several studies of Yellow Rails have been conducted in breeding areas but little is known about their migratory or wintering ecology (Bookhout 1995). The first Yellow Rail recorded in Oklahoma was collected on 7 March 1842 in what is now Delaware County (Tomer 1959). More than a century passed before Yellow Rails were again reported in Oklahoma. In 1954, one was reported ' Department of Biology, University of Central Okla- homa. Edmond, OK 73034, USA. ^Corresponding author; e-mail: cbutlerl l@uco.edu killed by a mowing machine (Heck and Arbour 2008) . These authors summarized published dates of occuiTence for the 20th century and reported one record from the 1 960s, three records from the 1970s, no records during the 1980s, and four records from the 1990s. W. D. Arbour began conducting regular rope drags to flush Yellow Rails beginning in 2001 at Red Slough Wildlife Management Area (WMA) in McCurtain County, Oklahoma. Yellow Rails were observed annually at this location from 2001 to 2008 (Heck and Arbour 2008). They are now considered to be regular autumn migrants from 15 October through 26 November in southeastern Oklahoma (OBRC 2004). Yellow Rails were also recorded during De- cember in 2 years (2004 and 2005) at Red Slough WMA and a single record exists from January (2004) at this location (Heck and Arbour 2008). We hypothesized that Yellow Rails may over- winter in small numbers in southeastern Okla- homa given the species has been recorded as late as mid-January in McCurtain County. METHODS Red Slough Wildlife Management Area (33° 44' 05" N, 94° 38 13" W; Fig. 1) is in McCurtain County, Oklahoma, and consists of —1,000 ha of moist soil management areas. 1.160 ha of (reforesting) bottomland hardwoods, and 165.6 ha of re.servoirs for a total of 2,325.6 ha (USDA 2009) . Red Slough WMA is cooperatively man- aged by the U. S. Forest Service, Natural Resources Conservation Service, Oklahoma De- partment of Wildlife Conservation, and Ducks Unlimited (USDA 2009). We traveled to Red Slough WMA seven times between November 2008 and March 2009 to capture and band Yellow Rails (3 times in Nov and once monthly from Dec through Mar). Vegetation was sampled at nine locations where birds were flushed during November 2008. A 1 X I ni" frame was placed at each location where we flushed Yellow Rails and vegetation height of 386 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2, June 2010 FIG. 1. Red Slough Wildlife Management Area (black star) in McCurtain County, Oklahoma. each species was measured as well as percent cover following Daubenmire (1959). Yellow Rails may have run before flushing (which could potentially bias the results), but the habitat appeared to be homogeneous and any bias was expected to be minimal. We began capturing Yellow Rails —30 min after sunset and continued our efforts for 3 hrs each night. We used a 12-m nylon rope weighted with bottles (filled with rocks) which two people dragged through damp, grassy areas where rails had previously been encountered. Rope-dragging caused rails to flush and we were able to shine a handheld spotlight on the area where the bird landed and catch them in a small handheld net. Each bird was banded with a USGS aluminum band, classified to age following Pyle (2008), and released. RESULTS Twenty-five Yellow Rails were banded from November 2008 through March 2009 (Table 1). One bird banded in November 2008 was recap- tured in December 2008 and one bird banded in TABLE 1. Trapping effort and capture.s of Yellow Rails, McCurtain County, Oklahoma, 2008-2009. # hours Month of sampling New birds Recaptures November 2008 18 13 2 (both banded earlier in Nov) December 2008 6 9 1 (banded in Nov) January 2009 6 1 1 (banded in Dec) February 2009 6 0 0 March 2009 6 2 0 December 2008 was recaptured in January 2009. A single Yellow Rail was flushed in February 2009 but we were unable to catch it. During November, most birds captured (9 of 13) were classified as adults (i.e., AHY) following Pyle (2008). In contrast, from December 2008 through March 2009, most birds captured (8 of 12) were classified as young (i.e., HY/SY). Fourteen genera in eight plant families were identified in the nine vegetation plots. Seven plots were dominated by grasses (60-100% cover), mainly Sporoholus spp. (dropseed), some identi- fied as Sporoholus compositus. Other grasses on these seven plots included Steinchisma hians (gaping grass) and Panicum spp. (panicgrass). One plot was dominated by flatsedge (Cyperus spp.) with 85% cover, and another by an unidentified Polygonaceae with 97.5% cover. Other plants encountered included Ambrosia spp. (ragweed), Ammannia spp. (toothcup), Andropo- gon virginiciis (broomsedge bluestem), Boltonia asteroides (white doll’s daisy), Eleocharis spp. (spike-rush), Hypericum drummondii (St. John’s wort), Juncus spp. (rush), O.xalis spp. (wood sorrel), and Tridens soictus (long-spike tridens). The dominant vegetation averaged (± SE) 44.1 ± 4.6 cm in height in areas where rails were captured. Ten Yellow Rails were captured in areas with no standing water and fifteen were captured in shallow water <4 cm deep. DISCUSSION Yellow Rails appear to winter in small numbers at Red Slough WMA in McCurtain County, Oklahoma. This is considerably farther north than previously reported wintering areas, because Red SHORT COMMUNICATIONS 387 Slough WMA is ~ 300 km north of the Gulf Coast. It is unknown whether this is a recent wintering range expansion or whether Yellow Rails have wintered here but not been previously detected. Most birds banded during winter (Dec- Mar) were HY/SYs, which suggests this species may undergo differential migration and young winter farther north than adults. Yellow Rails are reported to use the drier areas of Spartina spp. marshes along the coast (Book- hout 1995). Radiotelemetry in coastal Texas suggested that Yellow Rails prefer areas domi- nated by saltgrass {Distichlis spicata) and gulf cordgrass {Spartina spartinae) (Grace et al. 2005). Wayne (1905: 396-397) collected wintering Yellow Rails near Charleston, South Carolina away from the immediate coast in a “low, wet piece of open land with a dense growth of short, dead grass”. He collected 58 Yellow Rails between 1903 and 1918 at this freshwater locality (Post 2008). Heck and Arbour (2008) suggested Yellow Rails in Oklahoma occur primarily in fall panicum (Panicum dichotomiflorum) and long- leafed rushgrass {Sporobolus asper) during fall migration. We generally found Yellow Rails in areas dominated by Sporobolus spp. rather than Panicum spp., although the number of higher taxa (families and genera) at our sites was fairly high. It is possible that Yellow Rails choose wintering habitat based on structural components (e.g., areas dominated by grassy vegetation <50 cm in height with little or no standing water) rather than by dominant plant genera, as suggested for other rail species (Conway 1995, Melvin and Gibbs 1996). Yellow Rails are difficult to detect and may winter farther inland than previously reported. Given that they overwinter in small numbers in southeastern Oklahoma, observers elsewhere in the southeastern United States should be alert for the possibility of Yellow Rails overwintering in damp grassy areas. ACKNOWLEDGMENTS We thank W. D. Arbour, R. A. Bastarache, and the U. S. Forest Service for assistance in locating potential sites tor Yellow Rails. We thank E. J. Davis, M. E. Stone, B. M. Wilcox, B. D. Rising, J. A. Young, S. M. Still, D. R. Wood, and his ornithology class for assi.stance capturing Yellow Rails. We thank the Oklahoma State University Fore.st Resources Center for providing accommodations in McCurtain County. This research was supported by the Office of Research and Grants at the University of Central Oklahoma. T. A. Bookhout and an anonymous reviewer provided helpful comments on the manuscript. LITERATURE CITED Bookhout, T. A. 1995. Yellow Rail (Coiiiniicops nove- horacensis). The birds of North America. Number 139. Butcher, G. S., D. K. Niven, A. O. Panjabi, D. N. Pashley, and K. V. Rosenberg. 2007. WatchList: The 2007 WatchList for United States birds. American Birds 61:18-25. Conway, C. J. 1995. Virginia Rail (Rallus limicola). The birds of North America. Number 173. Daubenmire, R. 1959. A canopy-coverage method of vegetation analysis. Northwest Science 33:43-64. Environment Canada. 2006. Species of special concern: Yellow Rail. Quebec Region, Canadian Wildlife Service, Environment Canada, Quebec City, Quebec, Canada. http://www.qc.ec.gc.ca/faune/oiseaux_menaces/html/rale_ jaune_e.html (accessed 4 Februaiy 2009). Goldade, C. M., j. A. Dechant, D. H. Johnson, A. L. Zimmerman, B. E. Jamison, J. O. Church, and B. R. Euliss. 2002. Effects of management practices on wetland birds: Yellow Rail. USDI, Geological Survey, Northern Prairie Wildlife Research Center, James- town, North Dakota, USA. Grace, J. B., L. K. AuLAtN, H. Q. Baldwin, A. G. Billock, W. R. Eddleman, A. M. Given, C. W. Jeske, and R. Moss. 2005. Effects of prescribed fire in the Coastal Prairies of Texas. Open File Report 2005-1287. USDI, Geological Survey, Lafayette, Louisiana, USA. http://www.nwrc.usgs.gov/factshts/ 2005-1287-fire-in-coastal-texas-report.pdf (accessed 8 October 2009). Heck, B. A. and W. D. Arbour. 2008. The Yellow Rail in Oklahoma. Bulletin of the Oklahoma Ornithological Society 41:13-15. Melvin, S. M. and J. P. Gibbs. 1995. Sora (Porzana Carolina). The birds of North America. Number 250. Oklahoma Bird Records Committee (OBRC). 2004. Date guide to the occurrences of birds in Oklahoma. Fourth Edition. Oklahoma Ornithological Society, Tulsa, USA. Popper, K. J. and M. A. Stern. 2000. Nesting ecology of Yellow Rails in southcentral Oregon. Journal of Field Ornithology 7 1 :460^66. Post, W. 2008. Wintering ecology of Yellow Rails based on South Carolina specimens. Wilson Journal of Ornithology 120:606-610. Pyle. P. 2008. Identification guide to North American birds. Part 11. Anatidae to Alcidae. Slate Creek Press, Point Reyes, California, USA. Stenzel, j. R. 1982. Ecology of breeding Yellow Rails at Seney National Wildlife Refuge. Thesis. Ohio State University, Columbus, USA. Tomer, J. S. 1959. An Oklahoma record of the Yellow Rail. Auk 76:94-95. U.S. Department of Agriculture (USDA). 2009. Red Slough Wildlife Management Area. USDA, Forest Service, Hot Springs, Arkansas, USA. http://www.fs. fed.us/r8/ouachita/natural-resources/redslough/info. shtml (accessed 26 May 2009). Wayne, A. T. 1905. Notes on certain birds taken or seen near Charleston, South Carolina. Auk 22:395^00. 388 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 2, June 2010 The Wilson Journal of Ornithology 122(2):388-39L 2010 Breeding of the Giant Laughingthriish (Garrulax maximus) at Lianhuashan, Southern Gansu, China Jie Wang,* " Chen-Xi Jia,' Song-Hua Tang,* Yun Fang,* and Yue-Hua Sun*’^ ABSTRACT. — We describe the nest sites, nests, eggs, and incubation and provisioning behavior of the endemic Giant Laughingthrush (Garrula.x maximus) in a coniferous forest (2,850-2,950 m asl) at Lianhua- shan, southern Gansu, central China. We found seven shallow bowl-shaped nests in Picea-Ahies trees, 4.0 ± 1.5 m (x ± SD, n = 1) above the ground during May and June 2003, 2007, and 2008. Clutch size was 2.2 ± 0.4 unspotted blue eggs (2-3, n = 6) of which 1.4 ± 0.5 nestlings hatched (1-2, n = 7), and 1.0 ± 1.0 young fledged (0-2, n = 7). Three nests failed, possibly due to predation or abandonment during prolonged rainfall. Both males and females incubated clutches; nest attentiveness during the day decreased from 92.6 ± 0.9% before hatching to 59.4 ± 1.5% during days 3-7 of the nestling period. Both parents fed the nestlings ( 1 .0 ± 1.0 times/hr) and consumed the feces (0.3 ± 0.5 times/hr) during the 7-15 days after hatching Received 28 March 2009. Accepted 26 October 2009. The endemic Giant Laughingthrush {Garrulax maximus) is distributed from southern Gansu and southeastern Qinghai to southern and southeastern Tibet and northern Yunnan in southern and central China (Zheng 2005, Thompson 2007). Habitats occupied include open broadleaf and mixed broadleaf-coniferoLis forests, bamboo {Phyllosta- cliys spp.) scrub in broadleaf evergreen forest, oak (Quercus spp.) forest with scattered conifers, and relatively open broadleaf evergreen forest with shrub understory at 2,135^,115 m elevation (Thompson 2007). The Giant Laughingthrush is the largest laugh- ingthrush (Cheng et al. 1987, Rasmussen and Anderton 2005) and was previously treated as a subspecies of Spotted Laughingthrush (G. ocella- tus) (Ali and Ripley 1996). Little is known about its natural history and breeding except one ' Key Laboratory of Animal Ecology and Conservation Biology. Institute of Zoology. Chinese Academy of .Sciences. Beijing 1 00 1 01. China. ^Graduate School of Chinese Academy of Sciences, Beijing 100049, China. ^Corresponding author; e-mail: sunyh@ioz.ac.cn description based on a single nest in a bamboo clump, which contained two eggs (Ludlow and Kinnear 1944). We describe the nests, eggs, nestlings, and nesting behavior of this species in an alpine conifer-dominated forest in southern Gansu, China. METHODS Study Area. — Our study area was in the Lianhuashan Nature Reserve (34° 57' N, 103° 46' E) in southern Gansu Province, China. The forested area occurs on north, northeast, and northwest-facing slopes; only shrubs and grasses grow on south-facing slopes. Coniferous forest is the most prevalent cover type and is dominated by Dragon spruce {Picea asperata) and Earges fir {Abies fargesii). The other cover types are; (1) coniferous-deciduous forest, including spruce, fir, Himalayan birch {Benda utilis), and willow {Salix spp.); and (2) shrublands including willow, sea buckthorn {Hippophae rhamnoides), and barbeiry {Berheris spp.). The study area has been described by Sun et al. (2003). The mean annual tempera- ture is 5. 1-6.0° C, with a maximum of 34.0° C and minimum of —27.1° C. The climate is semiarid, and the annual precipitation is about 65 cm. Field Procedures. — We located seven nests (6 during incubation and 1 during the nestling stage) by following nesting-related activities of adults or by systematically checking individual trees in the conifer-dominated forest between 2,850 and 3,100 m elevation during three breeding seasons (Apr-Jul in 2003, 2007, and 2008). We inspected nests every 3-5 days to identify hatching and fledging dates and, if possible, cause of nest failure. Incubation or brood care was documented using data loggers (Tinytag Plus 2 Gemini Data Loggers, West Sussex, UK) at two nests, one of which was monitored for 194 hrs during 1 1 days. The probe was placed among the eggs to take a temperature reading every 15 sec. Length of on- and off-nest bouts, and behaviors of provisioning SHORT COMMUNICATIONS 389 TABLE 1. Giant Laughingthrush nests observed in the Lianhuashan Nature during May and June 2003, 2007, and 2008. DBH = diameter at breast height. Reserve, southern Gansu, central China Nest # Date found Stage Nest tree Nest Clutch size Nest fate Species DBH (cm) Height above ground (m) Distance to the trunk (m) Eggs Nestlings Fledglings 1 25 May 2003 Incubation Spruce 12 7.0 0.40 3 2 2 Successful 2 27 May 2003 Incubation Spruce 10 3.0 0.04 2 2 2 Successful 3 9 June 2003 Nestling Spruce 14 4.5 0.10 2 2 Successful 4 24 May 2007 Incubation Fir 32 3.8 2.00 2 1 1 Successful 5 9 June 2007 Incubation Spruce 13 4.0 0.05 2 1 0 Abandoned 6 26 May 2008 Incubation Fir 45 2.4 2.10 2 1 0 Depredated 7 3 June 2008 Incubation Spruce 26 3.5 2.20 2 1 0 Depredated young and removing feces were recorded at one nest by observation (33.7 hrs during days 6-10, 12-13, and 15 of the nestling period) and at another nest by video recorder (1.9 hrs on the 15th day of the nestling period). All observers and equipment were camouflaged using branches and leaves. Nest characteristics were measured after termi- nation of nesting and included height of the nest, distance of the nest to trunk, and height and diameter at breast height (DBH) of the nest tree. All data are expressed as A ± SD. RESULTS Nests were placed in Dragon spruce or Farges fir with three in branches of large conifers (DBH = 34 ± 10 cm) at a height of 2. 4-3. 8 m above ground and 2. 0-2.2 m from the tmnk. Four nests were in small spruce (DBH = 12 ± 2 cm) nearly touching the trunk at a height of 3. 0-7.0 m FIG. 1 . Nest with two eggs of the Giant Laughingthrush in a Dragon spruce (Picea asperata) at Lianhuashan, southem Gansu, central China. Photograph by Chenxi Jia. 390 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2, June 2010 On-nest bout Off-nest bout - )IG Nest attentiveness 0 O) 1-^ Q) 0 13 < 0 13 0 0 0 Day FIG. 2. Length of on- and off-nest bouts and nest attentiveness of the Giant Laughingthrush at one nest (found in 2007) during the incubation and nestling periods. Day 0 = batching day. The numbers on the asterisks are the time (hrs) monitored by data loggers whereas the numbers on/under the bars are the sample sizes of on/off-nest bouts. (Table 1). Nests were shallow and bowl-shaped (Fig. 1) with an inside diameter of 1 1.6 ± 0.7 cm ( 1 0.8- 1 2.2 cm, n = 4), outside diameter of 1 9.3 ± 0.9 cm (18.5-20.0 cm), inside depth of 5.1 ± 1.3 cm (4. 0-6. 5 cm), and outside height of 9.4 ± 1 . 1 cm (8.0-10.5 cm). Two nests were 78 and 80 g in wet weight, respectively. The inside bowl of the nest was lined with thin strands of bamboo, madder {Ruhia spp.), and spiraea (Spiraea spp.) stems, whereas the outer bowl and foundation were mainly of twigs, 25 cm (10-38 cm) in length, mostly of honeysuckle (Lonicera spp.) with a few from spruce, fir, and birch. Clutch size was 2.2 ± 0.4 eggs (2-3, n = 6) with 1.4 ± 0.5 nestlings hatched (// = 7) and 1.0 ± 1 .0 young fledged (n = 7). Ten unspotted blue eggs were 33.5 ± 1.5 mm (31.6-36.2 mm) in length, 22.9 ± 0.7 mm (21.6-24.4 mm) in diameter, and 8.8 ± 0.8 g (7.8-10.0 g) in weight. One nest was abandoned in the nestling stage, possibly due to prolonged rainfall (15-22 Jun 2007). One nest was possibly predated as one egg had a small hole in the eggshell during the incubation stage. The other nest was predated at 2344 hrs (Beijing Time, revealed by data logger) during the nestling stage with the nest overturned and feathers of the adult scattered on the ground. Mean length of on-nest bouts decreased from 59.4 ± 39.1 min (1.5-136.0 min) before hatching to 27.6 ± 18.0 min (3.0-72.5 min) during the nestling period, whereas mean off-nest bouts increased from 4.6 ± 2.8 min (0.5-10.5 min) to 15.6 ± 10.1 min (2.5-45.0 min) (Fig. 2). Nest attentiveness during the day decreased from 92.6 ± 0.9% before hatching to 59.4 ± 1.5% during days 3-7 of the nestling period. The female began diurnal activities at 0613 hrs (±28 min, n = 9) and began nocturnal brooding at 1931 hrs (±28 min, n = 9). Both parents provisioned nestlings and re- moved (or consumed) feces at a frequency of 1.2 ± 1.5 (0-6) and 0.5 ± 0.9 (0-3) times/hr, respectively, based on observation (32 hrs) and video recording (1.9 hrs) during days 7-15 of the nestling period. The nestling period appeared to be 17-18 days (2 nests). SHORT COMMUNICATIONS 391 DISCUSSION Nests in conifers were higher than the nest in the bamboo clump (1.2 m) reported by Ludlow and Kinnear (1994) and the nest of the Spotted Laughingthrush (<2 m) (Ali and Ripley 1996). Nest materials used by the Giant Laughingthrush were similar to those used by the closely related Snowy-cheeked Laughingthrush (G. sukatschewi) (Bi et al. 2003) and the Spotted Laughingthrush (Ali and Ripley 1996), but did not include moss, birch leaves, or strips of honeysuckle bark, which are frequently used by the Elliot’s Laughingthrush (G. elliotii) (Jiang et al. 2007). We did not find nests in bamboo clumps, which were quite short (<1.8 m) and scattered under the conifer trees or in small patches within the forest. The bamboo in this area had been nearly clear-cut by the local people 5 years earlier. The unspotted blue color of the eggs was similar to the Snowy-cheeked Laughingthrush (Bi et al. 2003) but possibly differed from the Spotted Laughingthrush, which has eggs that are spotless or with a few chocolate-brown specks near the broad end (Ali and Ripley 1996). Mean clutch size was similar to that of the Spotted Laugh- ingthrush (normal = 2 eggs, Ali and Ripley 1996), but less than the high-altitude Brown- cheeked Laughingthrush (G. henrici) (2.6 eggs, Lu et al. 2008), Snowy-cheeked Laughingthrush (3.5 eggs; Jie Wang, unpubl. data), and Elliot’s Laughingthrush (3.4 eggs, Jiang et al. 2007). Both parents incubated eggs and brooded nestlings, similar to most other laughingthrushes (Thompson 2007). Frequency of provisioning nestlings was lower than for the Snowy-cheeked Laughingthrush (1.2 vs. 7.9 times/hr; Bi et al. 2003). One reason may be that adult Giant Laughingthrushes seemed to be vigilant and prone to decrease their activities during the nestling stage, as the frequency of provisioning young estimated from video recordings (7 times in 1.9 hrs) was higher than from observation (31 times in 32 hrs). ACKNOWLEDGMENTS We are grateful to Dan Strickland for help with English and J,-L. Li, Y.-X. Jiang, Sh. -Q. Zhao, and the .staff of Lianhuashan Nature Reserve tor field assistance. We appreciate the constructive comments of K. E. Miller, C. E. Braun, and one anonymous reviewer on earlier drafts of the manuscript. The work was supported by National Natural Science Foundation of China (Grant 306201301 10). LITERATURE CITED Ali, S. and S. D. Ripley. 1996. Pages 29-31 in Handbook of the birds of India and Pakistan. Second Edition. Volume 7. Oxford University Press, New Delhi, India. Bi, Z.-L., Y. Gu, C.-X. Jia, Y.-X. Jiang, and Y.-H. Sun. 2003. Nests, eggs, and nestling behavior of the Snowy- cheeked Laughingthrush {Garrulax sukatschewi) at Lianhuashan Natural Reserve, Gansu, China. Wilson Bulletin 115:474^77. Cheng, T. H., Z. Y. Long, and B. L. Zheng. 1987. Pages 113-115 in Fauna Sinica (Aves. Volume 11. Passer- iformes, Muscicapidae II, Timaliinae). Science Press, Beijing, China. Jiang, Y.-X., Y.-Z. Zhu, and Y.-H. Sun. 2007. Notes on reproductive biology of Elliot’s Laughingthrush at Zhuoni, Gansu. Sichuan Journal of Zoology 26:555- 556. Lu, X., G.-H. Gong, and X.-H. Zeng. 2008. Reproductive ecology of Brown-cheeked Laughingthrush (Garrulax henrici) in Tibet. Journal of Field Ornithology 79:152- 158. Ludlow, F. and N.-B. Kinnear. 1944. The birds of southeastern Tibet. Ibis 86:43-86, 176-208, 348-389. Rasmussen, P. C. and J. C. Anderton. 2005. Page 413 in Birds of south Asia: the Ripley Guide. Volume 2. Smithsonian Institution, Washington, D.C., USA, and Lynx Edicions, Barcelona, Spain. Sun, Y.-H, J. E. Swenson, Y. Fang, S. Klaus, and W. SCHERZINGER. 2003. Population ecology of the Chi- nese Grouse, Bonasa sewerzowi, in a fragmented landscape. Biological Conservation 110:177-184. Thompson, H. S. S. 2007. Page 253 in Handbook of the birds of the world. Volume 12. Picathartes to tits and chickadees (J. del Hoyo, A. Elliott, and D. A. Christie. Editors). Lynx Edicions, Barcelona, Spain. Zheng, G.-M. 2005. Page 253 in A checklist on the classification and distribution of the birds of China. Science Press, Beijing, China. 392 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. 2, June 2010 The Wilson Journal of Ornithology 1 22(2):392-395, 2010 First Description of the Nest of Jocotoco Antpitta {Grallaria ridgelyi) Harold F. Greeney'-'''^ and Mery E. Juina J' - ABSTRACT. — The Jocotoco Antpitta (.Grallaria ridgelyi) is an endangered and poorly studied inhabitant ot montane bamboo (Chusc/uea spp.) thickets of extreme southeastern Ecuador. There is nothing known of the breeding biology of Jocotoco Antpitta apart from a single record of a dependent juvenile, and we describe a nest of this species for the first time. The nest was a bulky cup composed primarily of dead plant materials and tirmly supported by a large clump of epiphytes on the side of a dead trunk. The single nestling was provisioned at a rate of 1.96 feedings/hr during the final 5 days prior to Hedging. Two adults brought food to the nestling, often delivering prey from a nearby worm- feeding station created by the Jocotoco Foundation. The pair we studied may breed twice a year, but this may have been facilitated by their proximity to the artificial feeder. Received I Ma\ 2009. Accepted 17 September 2009. The Jocotoco Antpitta (Grallaria ridgelyi) is one of the rarest species of this poorly known genus that encompasses 31 species in the family Grallariidae (Remsen et al. 2009). Despite its large size and distinctive plumage, the Jocotoco Antpitta was only described 10 years ago (Krabbe et al. 1999), having been overlooked due to its propensity to forage in the thick undergrowth of Chu.sc/uea bamboo-covered slopes (Krabbe et al. 1999, Ridgely and Greenfield 2001, Heinz et al. 2005, Freile et al. 2010). The Jocotoco Antpitta is apparently confined to southeastern Ecuador and is considered Endangered by BirdLife Interna- tional (2009). The breeding biology of the Jocotoco Antpitta remains undocumented with only a single description of a juvenile accompa- nying two adults (Greeney and Gelis 2()()5b). We de.scribe the first nest of Jocotoco Antpitta from ' Yanayacu Biological Station & Center for Creative Studies. Cosanga, Napo Province, Ecuador, c/o 721 Loch y Amazonas, Quito. Ecuador. ^ Fimdacion de Conservacion Jocotoco, Avenida Los Shyris N37-146 y El Comcrcio. Quito, Ecuador. 'Current address; University of Nevada, Department of Biology 0314, 1664 North Virginia Street. Reno, NV 89.‘i.‘)7. USA. ■‘Corresponding author; e-mail: revmmo.ss@yahoo.com the Tapichalaca Biological Reserve in southeast- ern Ecuador. METHODS We made all observations at the Tapichalaca Biological Reserve (04° 30' S, 79° 10' W) near Valladolid in extreme southeastern Ecuador. The teiTain is typically steep for this elevation in the Andes (~ 2,500 m) with vegetation characterized by high epiphyte density, a canopy height of 15- 30 m, and a dense understory comprised largely of CIntsquea bamboo (Krabbe et al. 1999). We made nest measurements to the nearest 0.5 cm and behavioral observations at the nest using a tripod- mounted video camera placed ~ 5 m from the nest. We recorded 16.3 hrs of video from 1030 to 1800 hrs on 10 November 2008 (5 days prior to fledging) and from 0700 to 1715 hrs the day before fledging. RESULTS We observed two adult Jocotoco Antpittas on the morning of 5 November 2008 gathering food items at the worm feeder and black-light trap adjacent to the Tapichalaca Lodge (~ 2,500 m asl). Adults carried multiple invertebrates from this feeding area, leaving in the direction of the subsequently discovered nest. We again ob.served adults carrying prey in the same direction on 6 November. On 8 November 2008 at 1000 hrs MEJ found the nest, 30 m from the feeders, which contained a single large nestling. The nestling was well feathered at the time of discovery, cinnamon on the back, chestnut on the crown and upper breast, and finely barred with black on all visible areas (Fig. 1). Most of the ocular area was still bare with just a faint hint of the white malar pattern of adults. The upper mandible was dark with a pale orange tip while the lower mandible was mostly dull orange. The gape was orange but we did not examine the mouth lining or remove the nestling for closer inspection of plumage patterns. The nest was 3.6 m above the ground on the side of a vertical, rotting tree trunk that was 4.1 m tall SHORT COMMUNICATIONS 393 FIG. 1. Nestling Jocotoco Antpitta (G;'a//«n<7 in southeastern Ecuador, 3 days prior to fledging (Photo- graph by M. E. Juifia). (56 cm dbh) (Fig. 2). The nest was placed flush against the trunk and was supported primarily by a cluster of Tillansia bromeliads and secondarily by a 4-cm diameter, horizontal branch crossing under the nest. The nest was a large cup comprised predominantly of dead plant materials, mostly dicot and bromeliad leaves. The cup was lined internally with a thick layer of fine black rootlets and bare fern rachi.ses. Internal measurements of the nest cup were 15.5 X 14.0 cm in width (measured at perpen- dicular angles) and 12.5 (back) to 17.0 cm (front) deep. Externally, the nest was 23.5 by 26 cm in diameter and 18 cm tall. The area around the nest contained a dense understory and, along with the dense ~ 20 m canopy, provided nearly 90% shading of the nest. The nest was directly shaded, predominantly by Chusquea bamboo, as well as epiphytic ferns and orchids. The nest was on the northeast side of the trunk on a northeast facing slope where it received most of its sun in the mornings. FIG. 2. Nesting site of Jocotoco Antpitta {Grallaria ridgelyi) in southeastern Ecuador. The white arrow points to the nest itself (Photograph by M. E. Juina). Transcriptions of video at the nest revealed the nestling was brooded for only 7% of daylight hours. We were unable to ascertain if more than one adult brooded, but we confirmed that two adults brought food. Adult feeding visits to the nest were brief (mean ± SD = 2.2 ± 7.5 min), during which time they frequently probed sharply or rapidly into the nest lining. The nestling was provisioned at a rate of 1.96 feedings/hr with absences from the nest lasting 29.7 ± 24.4 min. The nestling often followed the adults to the rim of the nest as they left after delivering food during the day prior to leaving the nest. In the absence of adults, the nestling frequently stood and stretched its wings or shuffled about inside the nest, occasionally pecking at small invertebrates mov- ing around the nest. DISCUSSION The nest of Jocotoco Antpitta is a deep, bulky cup similar in form to the described nests of other Grallaria (Greeney et al. 2008). The nest was well supported by bromeliads, against the trunk and branch of a dead tree, most similar to the nest situations of Rufous Antpitta (G. rufula) (Greeney 394 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 2, June 2010 and Gelis 2005a), White-bellied Antpitta (G. hypoleiica) (Price 2003), Yellow-breasted An- tpitta (G. flavotincta) (Greeney et al. 2009), Variegated Antpitta (G. varia) (Protomastro 2000), Pale-billed Antpitta (G. carrikeri) (Wie- denfeld 1982), Great Antpitta (G. excelsa) {Koefed and Auer 2004), and some Scaled Antpitta (G. giiatimalensis) (Dobbs et al. 2001, 2003). The nest was composed predominantly of dead plant material, most similar to nests of Chestnut-naped Antpitta (G. nuchalis) (Juina et al. 2009), Chestnut-crowned Antpitta (G. ritficapilla) (Martin and Greeney 2006), Plain-backed Antpitta (G. haplonota) (Greeney et al. 2006), Watkins’s Antpitta (G. watkiusi) (Martin and Dobbs 2004), Variegated Antpitta (Quintella 1987, Protomastro 2000), Pale-billed Antpitta (Wiedenfeld 1982), and White-bellied Antpitta (Price 2003), as well as some nests of Moustached Antpitta (G. alleni) (Freile and Renjifo 2003, Londono et al. 2004, Greeney and Gelis 2006), and Scaled Antpitta (Miller 1963; Rowley 1966; Dobbs et al. 2001, 2003). Previous observations of an older, dependent fledgling Jocotoco Antpitta (Greeney and Gelis 2005b) were made on 30 November 2003 (HFG, Linpubl. data). We also observed young fledglings accompanying adults to the worm teeder on multiple days in July and November 2007 and in April 2008. Observations of juveniles, with the exception of the previously published report of Greeney and Gelis (2005b), were all near the worm feeder maintained by the Jocotoco Foun- dation which is presumably visited by only one pair of Jocotoco Antpittas (MEJ, pers. obs.). These records suggest at least two broods a year. The effects of these feeders are unstudied despite the many birding lodges in Ecuador that now use worm feeders for Gi'cillcivio antpittas. We believe it is possible the year-round surplus of food provided by the feeders in our study area may have encouraged this pair of Jocotoco Antpittas to reproduce throughout the year, a pattern which may not be common for pairs without dietary supplementation. The importance of this possibil- ity and the long-term repercussions on adult longevity and health deserve further study at this site and at others where worm feeders are used. Clutch size remains undocumented for Jocotoco Antpitta, but repeated sightings of single fledg- lings (Greeney and Gelis 2()05b, this study) combined with our observations of a single nestling, suggest this species may lay a smaller clutch than most other GraUaria (Greeney et al. 2008). Alternatively, single fledglings may reflect brood or clutch reduction from the typical two- egg clutches of the genus (Greeney et al. 2008). ACKNOWLEDGMENTS Field work was supported by the Jocotoco Foundation. We thank the staff and park guards of the Tapichalaca Biological Reserve, most especially Franco Mendoza and Timo Brockmeyer. Preparation of this manuscript was supported by National Geographic Grant W38-08. The PBNHS continues to aid our natural history work in Ecuador. This is publication Number 194 of the Yanayacu Natural History Research Group. LITERATURE CITED BirdLife International. 2009. Jocotoco Antpitta - Bird- Life species factsheet. http;//www.birdlife.org/datazone/ species/index.html (accessed 1 August 2009). Dobbs, R. C., P. R. Martin, and M. J. Kuehn. 2001. On the nest, eggs, nestlings, and parental care of the Scaled Antpitta (GraUaria giiatimalensis). Ornitologia Neotropical 12:225-233. Dobbs, R. C., P. R. Martin, C. Batista, H. Montag, and H. F. Greeney. 2003. Notes on egg laying, incubation and nestling care in Scaled Antpitta GraUaria guatimalensis. Cotinga 19:65-70. Freile, J. F. and L. M. Renjifo. 2003. First nesting records of the endangered Moustached Antpitta, GraUaria alleni. Wilson Bulletin 115:11-15. Freile, J. F., J. L. Parra, and C. H. Graham. 2010. Distribution and conservation of GraUaria and Gral- lariciUa antpittas (Grallariidae) in Ecuador. Bird Conservation International 21: In press. Greeney, H. E. and R. A. Gelis. 2005a. A nest of the Rufous Antpitta GraUaria ritfiUa depredated by a Turqoise Jay Cyanolyca turcosa. Cotinga 24:1 10-1 1 1. Greeney, H. E. and R. A. Gelis. 2005b. Juvenile plumage and vocalization of the Jocotoco Antpitta GraUaria ridgelyi. Cotinga 23:79-81. Greeney, H. E. and R. A. Gelis. 2006. Observations on parental care of the Moustached Antpitta (GraUaria alleni) in northwestern Ecuador. Ornitologia Neotrop- ical 17:313-316. Greeney. H. F., R. C. Dobbs, P. R. Martin, and R. A. Gelis. 2008. The breeding biology of GraUaria and Grallariciila antpittas. Journal of Field Ornithology 79:113-129. Greeney, H. F., M. E. Juina J., S. M. 1. Lliquin, and J. A. Lyons. 2009. First nest description of the Yellow- breasted Antpitta GraUaria flavotincta in northwest Ecuador. Bulletin of the British Ornithological Club 129:256-258. Greeney, H. E., R. A. Gelis, C. Dingle, F. J. Vaca B., N. Krabbe, and M. Tidwell. 2006. The ne.st and eggs of the Plain-backed Antpitta (GraUaria haplonota) in eastern Ecuador. Ornitologia Neotropical 17:601-604. Heinz, M., V. Schmidt, and M. Schaefer. 2005. New SHORT COMMUNICATIONS 395 distributional record for the Jocotoco Antpitta Gral- larici ridi’clyi in south Ecuador. Colinga 23:24-26. JuiN.A .1, M. E., J. B. C. Harris, and H. F. Greeney. 2(X)9. Description of the nest and parental care of the Chestnut- naped Antpitta (Grcillarici michalis) from southern Ecuador. Ornitologia Neotropical 20:305-310. Kofoed, E. M. and S. K. Auer. 2004. First description of the nest, eggs, young, and breeding behavior of the Great Antpitta (Gnillaria excelsa). Wilson Bulletin 1 16:105-108. Krabbe, N. and T. S. Schulenberg. 2003. Family Formicar- iidae (ground-antbirds). Pages 682-731 in Handbook of the birds of the world. Volume 8. Broadbills to tapaculos. (J. A. del Hoyo, A. Elliott, and D. A. Christie, Editors). Lynx Edicions, Barcelona, Spain. Krabbe, N., D. J. Agro, N. H. Rice, M. Jacome, L. Navarrete, and F. Sornoza. 1999. A new species of antpitta (Formicariidae: Gi allaria) from the southern Ecuadorian Andes. Auk 116:882-890. Londono, G. a., C. a. Saavedra-R., D. Osorio, and J. MartInez. 2004. Notas sobre la anidacion del Tororoi Bigotudo {Grallaria alleni) en la Cordillera Central de Colombia. Ornitologia Colombiana 2:19-24. Martin, P. R. and R. C. Dobbs. 2004. Description of the nest, egg and nestling of Watkin’s Antpitta Grallaria watkinsi. Cotinga 21:35-37. Martin, P. R. and H. F. Greeney. 2006. Description of the nest, eggs and nestling period of the Chestnut-crowned Antpitta Grallaria ruficapilla from the eastern Ecua- dorian Andes. Cotinga 25:47-49. Miller, A. H. 1963. Seasonal activity and ecology of the avifauna of an American equatorial cloud forest. University of California Publications in Zoology 66:1-78. Price, E. R. 2003. First description of the nest, eggs, hatchlings, and incubation behavior of the White- bellied Antpitta (Grallaria hypoleuca). Ornitologia Neotropical 14:535-539. Protomastro, j. j. 2000. Notes on nesting of Variegated Antpitta Grallaria varia. Cotinga 14:39^1. Quintela, C. E. 1987. First report of the nest and young of the Variegated Antpitta (Grallaria varia). Wilson Bulletin 99:499-500. Remsen Jr., j. V., C. D. Cadena, A. Jaramillo, M. Nores, J. F. Pacheco, M. B. Robbins, T. S. Schulenberg, F. G. Stiles, D. F. Stotz, and K. J. Zimmer. 2009. A classification of the bird species of South America. American Ornithologists’ Union. http://www. museum. lsu.edu/~Remsen/SACCBaseline.html (accessed 1 August 2009). Ridgely, R. S. and P. j. Greenfield. 2001. Birds of Ecuador. Cornell University Press, Ithaca, New York, USA. Rowley, J. S. 1966. Breeding records of birds of the Sierra Madre del Sur, Oaxaca, Mexico. Proceedings of the Western Foundation of Vertebrate Zoology 1:107- 204. Wiedenfeld, D. a. 1982. A nest of the Pale-billed Antpitta (Grallaria carrikeri) with comparative remarks on antpitta nests. Wilson Bulletin 94:580-582. The Wilson Journal of Ornithology 122(2):395-398, 2010 Nesting Ecology of the Grey-backed Shrike {Lanius tephronotus) in South Tibet Xin Lu,'-^ Chen Wang,' and Tonglei Yu' ABSTRACT. — We studied the breeding ecology of the Grey-backed Shrike (Lanius tephronotus) in alpine shrub habitats in south Tibet. Shrikes nested between late May and early July at elevations of 4,010-4,540 m with a delay in nesting time with increa.sed elevation. Nests were built mostly in bushes (83%), 0.7-2. 1 m above ground. Clutch size averaged (± SD) 4.12 ± 0.67, range = 3-5 eggs and was smaller for pairs that nested later at higher elevations. Incuhation by females began before clutch competition and lasted 15-18 days. Nestlings were cared for by both parents tor 14—15 days. Breeding success, considered as the number of nest attempts that fledged at least one young, was 45.8%. Shrikes nesting in alpine habitats, compared with their 'Department of Zoology, College of Life Sciences, Wuhan University, Wuhan, 430072, China. -Corresponding author; e-mail: luxinwh@gmail.com lowland congeners, experienced shorter breeding sea- sons, produced fewer broods, smaller clutches, and larger eggs. They followed a life history strategy that allowed them to compensate for reduced annual fecundity under harsh conditions. Received 29 June 2009. Accepted I November 2009. Lanins shrikes occur throughout the world except for South America and Australia (Lefranc and Worfolk 1997). Among the world’s Lanins species, the Grey-backed Shrike (L. tephronotus) breeds at the highest elevations in the Himalayas on the Tibetan plateau, and peripheral zones at elevations of 2,700 to 4,500 m. This species is a partial migrant and most shrikes winter at low elevations between 300 and 2,500 m within their 396 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122. No. 2. June 2010 breeding range while a few migrate to lowland Southeast Asia (Lefranc and Worfolk 1997). However, little information pertaining to its reproduction has been published (Tsering et al. 2002). We studied the breeding ecology of the Grey-backed Shrike in a shrub-covered alpine valley in south Tibet. Our objectives are to report on the adaptation of life history traits of shrikes to high-elevation environments, and provide base- line data for species conservation. METHODS Study Area. — Field data were gathered in the Xiongse Valley (29° 27' N, 91° 40' E) near Lhasa, Tibet during 1996 and 1999-2004. Shrub vegeta- tion characterizes the study area, including roses {Rosa sericea) and barberries (Berber is hem- leyana) at 3,980^,550 m, and Wilson Junipers (Sabina pingii) at 4,550-4,980 m on south-facing slopes; spiraeas (Spiraea japonica) at 4,200- 4,400 m, rhododendrons (Rhododendron nivale), and alpine willows (Salix sclerophylla) at 4,400- 5,100 m on north-facing slopes. Field Procedures. — Our study site was a 400-ha plot that covered the entire elevational range of the shrub vegetation. All Grey-backed Shrike nests located were described, and nest contents measured. Subsequent visits were made to examine nesting success. Several nests were checked infrequently because of their remoteness. We pooled the data of all years for analysis because of the small number of nests found during the study period. We chose P < 0.05 as the minimum acceptable level of statistical signifi- cance. Means are expressed ± SD. RESULTS Gray-backed Shrikes are the only Lanins species in Lhasa area. They arrived in the valley at 4,400 m in early May (2 to 12 May), singly (40% of 15 sightings) or in pairs (60%). Most departed by early October, but a few individuals were occasionally seen above 4,200 m until mid- December. Nesting elevations ranged between 4,010 and 4,540 m with 69% (25 of 36 nests) below 4,300 m. Shrikes tended to nest near streams (<5() m; 61% of 36 nests). A 1 82-ha plot on south-facing slopes was systemically searched and seven shrike nests were found, producing a breeding density of 3.8 pairs/ 100 ha. Shrikes nested in several plant species, of which most (61%) were in the two most common species, roses and barberries. Shrike nests con- structed in bushes were nearer the ground ( 1 .5 ± 0.4 m, range = 0.7-2.!, n = 30) than those in trees (2.8 ± 0.2 m, range = 2. 5-3.0, n = 6; t = 7.26, P < 0.001). Individual bushes selected for nesting had dense foliage. Mean vegetation cover around shrike nests was 34 ± 18% (range = 10- 85, n = 30), similar to the overall coverage (36%) in the study area. However, a few shrike nests were built in solitary tall bushes. Nests were cup- shaped and materials in the exterior layer of the nest cup consisted of thin twigs and roots of bushes. The interior layer consisted of feathers, sheep hair, and, at times, moss. Dimensions of seven nests were; exterior diameter 14.6 ± 0.9 cm (range = 13-16), interior diameter 9.0 ± 0.8 cm (range = 8-10), inside depth 6.8 ± 1.4 cm (range = 5-8.6), and outside height 15.3 ± 2.1 cm (range = 12-19). Females did not lay eggs immediately after they finished building the nest. The duration between completion of nest construction and laying of the first egg was 6 to 10 days (n = 5 nests). Egg- laying began in late May and ended by early July with a peak in early June (Fig. 1 ). Shrikes nesting at higher elevations initiated clutches later than those nesting at lower elevations (r = 0.41, = 32, P = 0.02). One egg was laid per day (n = 13 daily monitored eggs from 6 nests). Eggs were whitish with light-brown spots, especially on the large end. Fresh mass of 11 eggs was 5.0 ± 0.16 g (range = 4.7-5. 2). Measurements of 46 eggs from 12 nests averaged 25.7 ± 1.17 mm (range = 23.8-28.2) in length and 18.6 ± 0.48 mm (range = 17.8-20.0) in breadth. Clutch size varied from three to five eggs (4.12 ± 0.67, n = 25) with four being most common (56%). We detected a strong decline in clutch size over the season (/• = —0.43, n = 25, P = 0.03). However, it was not significant once we controlled for elevation (partial coiTelation coefficient, r = -0.28, n = 22, P - 0.18). Clutch size decreased significantly as elevations increased (r = —0.46, P — 0.02), but did not if we considered the egg-laying time (partial correlation coefficient, r — -0.34, P — 0.1 1 ). The decrease of clutch size was affected by both breeding date and nesting elevation. Incubation, considered as the period from the laying of the last egg to hatching of the first egg, started before clutch completion (n = 6 nests, 3 began with the first egg, 2 with the second egg, and 1 with the third egg). Only females incubated; SHORT COMMUNICATIONS 397 FIG. 1. Temporal distribution (all data combined) of date of laying of the first egg by Grey-backed Shrikes m the Xiongse Valley, south Tibet. Clutches initiated are arranged in 5-day periods. incubation lasted 15 days at 4,050 m (n = 1) and 17- 18 days at >4,350 m (/? = 3). Males defended the nest during incubation and, at times, fed incubating females. Nestlings were provisioned by both female and male parents. Fledging occurred at 14-15 days (14.3 ± 0.6, n = 3) of age and fledglings weighed 34.3 ± 0.6 g (range = 32.2-36.9, n = 6), which was 7 1 % of the adult female weight (48 . 1 g, average of 4 birds measured in the field). Fifty of 58 eggs (86.2%) from 14 clutches hatched successfully, resulting in a brood size of 3.71 ± 1.00 (/? = 14). Thirty of 33 hatchlings (90.0%) from nine clutches survived to fledging. Brood size at fledging was 3.36 ± 0.81 {n =11). Overall breeding success, measured as the per- centage of clutches that produced at least one fledged young, was 45.8% (11 of 24 known-fate attempts). The probability of a nest being successful calculated using Mayfield’s (1975) method was 0.455 (i.e., 45.5%). Two (15.4%) of 13 failed attempts were due to nest abandonment because of inclement weather (nestlings died in the nest after continuous rainfall), and 1 1 (84.6%) because of predation (damaged nest structure) likely by mountain weasels (Mustela ahaica), the most common carnivore in alpine areas. DISCUSSION Grey-backed Shrikes, compared to their Lhasa counterpart and other lowland congeners (Ta- ble 1), had a short breeding season (~ 50 days). They produced fewer and smaller clutches but larger eggs than their lowland relatives. Alpine shrikes had average clutch size similar to conspecifics in lowland Lhasa, but their maximum clutch size was one egg less. Shrike clutches within the alpine valley became smaller with increasing elevations. Invertebrates in the Tibetan plateau are scarce in terms of species diversity and population abundance decreases with increased elevations (Wang et al. 1992). Poor food avail- ability could prevent shrikes nesting at high elevation from producing larger clutches. Selec- tion should cause high-elevation shrikes to allocate more energy for each individual offspring (fewer but larger eggs) to improve their survival in harsh unpredictable alpine environments, which may compensate for reduced fecundity resulting from both fewer and smaller clutches. The shift in reproductive investment from offspring number toward offspring quality follows the general trend in avian life history shifts along elevational gradients (Badyaev 1997, Badyaev and Ghalam- bor 2001, Lu 2005, Bears et al. 2009). Badyaev and Ghalambor (2001) suggested that high- elevation birds provide greater parental care to their offspring in terms of longer incubation and nestling periods than their low-elevation counter- parts. Grey-backed Shrikes nesting in alpine habitats versus lowland Lhasa had longer incuba- tion and nestling periods (Table 1). Relative to other species of shrikes at the same latitude, Grey- backed Shrikes appeared to not be typically longer in these two parameters of parental care. This differentiation might be related to interspecific difference in body size or other traits. 398 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 2. June 2010 TABLE 1. Life history parameters of some Lanins shrikes nesting near 30° N. L. tephronotns L. tephronolus L. excubitor L. ludovicianus L. schach Locality Lhasa, China Lhasa, China Sede Boqer, Israel Florida, USA Yangzhou, China Latitude, N 29° 27' 29° 32' 30° 52' 27° 05' 32° 23' Elevation, m 4,300 3,650 475 11 10 Breeding season 50 65 120 160 100 length, days Broods per year 1 1 1^ 1-3 1-2 Clutch size (range) 4.1 (3-5) 4.0 (3-6) 5.8 (5-7) 4.3 (2-6) 5.9 (4-8) Egg volume, mm^ 8,916 10,153 8,689 8,484 Egg volume/female 196 161 183 149 body mass Incubation period. 16.7 (15-18) 15.3 16.8 (14-18) 17 14.0 (12-15) days (range) Nestling period. 14.3 (14-15) 12.9 17.2 (16-19) 15 12.5 days (range) Source This study Tsering et al. 2002 Yosef 1992 Yosef 2001 Yan and Ma 1991, Hu et al. 2007 Grey-backed Shrike nests were intermediate in probability to be successful (45.5%), compared with those of other species of shrikes (Red-backed Shrike [L. coUurio]: 35.3%, Horvath et al. 2000; Woodchat Shrike [L. senator]: 63.8%, Talagi 2001; Loggerhead Shrike [L. ludovicianus]: 60.0%, summarized by Yosef 2001). High- elevation shrikes, despite presumably higher juvenile survival, had lower annual productivity as a result of fewer and smaller clutches, and relatively low nesting success. This suggests that this species is more at risk if their habitats are altered by changes in land management activities or climate. ACKNOWLEDGMENTS We thank the Buddhists in the Xiongse Nunnery for logistic assistance and Guohong Gong. Liying Zhang, Xianhai Zeng, and Xiaoyan Ma for field assistance. We also thank Reuven Yosef and an anonymous reviewer for their comments on the manuscript. This work was conducted at the Field Research Station for Tibetan Wildlife, which is jointly administered by Wuhan University and Tibet University. LITERATURE CITED Badyaev, a. V. 1997. Avian life history variation along altitudinal gradients: an example with cardueline finches. Oecologia 1 1 1:365-374. Badyaev, A. V. andC. K. Ghalambor. 2001. Evolutional gradients: trade-off between parental care and fecun- dity. Ecology 82:2948-2960. Bear.s. H., K. Martin, and G. C. White. 2009. Breeding in high-elevation habitat results in shift to slower life- history strategy within a single species. Journal of Animal Ecology 78:365-375. Horvath R., R. Farkas, and R. Yosef. 2000. Nesting ecology of the Red-backed Shrike (Lanins coUnrio) in northeastern Hungary. Ring 22:127-132. Hu, J., T. P. Guan, C. Q. Zhou, and J. C. Hu. 2007. Growth of Rufous-backed Shrike (Lanins scliach). Sichuan Journal of Zoology 26:152-154. Lefranc, N. and T. Worfolk. 1997. Shrikes: a guide to the shrikes of the world. Pica Press, Sussex, United Kingdom. Lu, X. 2005. Reproductive ecology of Blackbirds (Tnrdns niernla ma.xiinns) in a high-altitude location, Tibet. Journal of Ornithology 146:72-78. Mayfield, H. F. 1975. Suggestions for calculating nest success. Wilson Bulletin 87:456^66. Talagi, M. 2001. Some effects of inclement weather conditions on the survival and condition of Bull- headed Shrike nestlings. Ecological Research 16:55- 63. TSERING, R.. L. E. JOHANNESSEN, and T. S0LHOY. 2002. Breeding biology of the Grey-backed Shrike (Lanins tephronotns). Proceedings of the International Orni- thological Congress 23:289. Wang B. H.. W. H. Yuan, C. M. Wang, F. S. Huang, S. H. Tang, and D. W. Lin. 1992. The Xizang insect fauna and its evolution. Henan Science and Technol- ogy Publishing House, Zhengzhou, China.- Yan, a. H. and j. S. Ma. 1991. Studies on ecology of the Brown-backed Shrikes (Lanins schacli). Chinese Journal of Zoology 26:30-32. Yo.SEF, R. 1992. From nest building to fledging of young in Great Grey Shrikes (Lanins twcnhilor) at Sede Boqer, Israel. Journal of Ornithology 133:279-285. Yosef, R. 2001. Nesting ecology of resident Loggerhead Shrikes in southcentral Florida. Wilson Bulletin 113: 279-284. SHORT COMMUNICATIONS 399 The Wilson Journal of Ornithology 122(2):399-402, 2010 The First Reported Hybridization of Abert’s and Canyon Towhees {Pipilo spp.) R. Roy Johnson'"^ and Steven L. Hopp^ ABSTRACT. — Two mixed pairs of towhees were found in irrigated desert yards near Tucson, Arizona. The first known towhee Fi hybrids, from a female PipUo aherti (Abert’s Towbee) and male P. fiiscns (Canyon Towhee) were studied from winter of 1998-1999 through summer 2000. This mixed pair raised at least eight young in three broods during the two breeding seasons. Young were so similar to P. fuscus that, if seen outside this context, they would probably not be identified as hybrids. A second mixed pair, also near Tucson, suggests that hybridization between P. aherti and P. fuscus may be more common than originally thought. Lack of previously reported hybridization between P. aherti and P. fuscus may be either due to internal or external isolating mechanisms, limited survival and longevity, or human failure to recognize hybrids. Received 2 September 2009. Accepted 10 January 2010. Hybridization among birds is relatively common, occurring in an estimated 10% of non-marine species (Grant and Grant 1992). The tendency to cross can be strong for some species pairs. For example, hybrid- ization is common in three of the four species of ‘ ‘red- eyed” towhees (Pipilo spp.) of the United States and Mexico (Sibley 1954; Sibley and West 1958, 1959; Sibley and Sibley 1964; Greenlaw 1996). Hybridiza- tion has not previously been reported in the complex of four species of towhees collectively known as “brown towhees.” The California Towhee (P. criss- alis) of the Pacific West does not occur sympatrically with the other three species of “brown towhees.” However, P. fuscus (Canyon Towhee) occurs throughout a portion of its range with both P. aherti in western Arizona and extreme northwestern Mexico, and with P. alhicollis (White-throated Towhee) of central Mexico (Davis 1951, Johnson and Haight 1996, AOU 1998). P. aherti and P . fuscus occur together throughout much of the range of P. aherti in southern and central Arizona, but P. aherti occurs in more mesic riparian bottoms (Tweit and Finch 1994) while P. fuscus generally inhabits 'Johnson & Haight Environmental Consultants, 3755 South Hunters Run, Tucson, AZ 85730, USA. ^Environmental Studies, Emory and Henry College, Emory, VA 24327, USA. •’'Corresponding author; e-mail: rroylois@aol.com adjacent riparianlands or drier uplands (Johnson and Haight 1996). OBSERVATIONS SLH discovered a mixed pair of “brown towhees” nesting on Tanque Verde Creek during February and March 1999, —24 km east of Tucson. The immediate area represents an eco- tone between the riparian mesquite (Prosopis velutina) bosque (woodland) along Tanque Verde Creek, typical P. aherti habitat, and upland Sonoran desertscrub, habitat often occupied by P. fuscus. This mixed pair of towhees raised at least three broods of F| hybrids during the next two breeding seasons (Appendix). The Abert’s Towhee was observed on 16 April 1999 carrying nesting material into a 3.4 m tall ornamental juniper (Juniperus spp.). The Abert’s Towhee was followed closely by the Canyon Towhee, suggesting the Abeit’s was the female because the male of a mated pair of Canyon Towhees has been reported to fly with the female as she carries nesting material (Johnson and Haight 1996; J. T. Marshall, pers. comm.). This was subsequently verified by extracting DNA from the feathers and ascertaining the young possessed the mtDNA of Abert’s Towhee (R. M. Zink, pers. comm.) The pair frequently engaged in sc/ueal-duets when greeting each other, a trait of both species (Marshall 1964). The nest was 1.9 m above ground and measured 7.6 cm tall with 12 cm outside, and 7.6 cm inside diameters. The nest contained three eggs on 27 April and on 8 May the nest contained three nestlings, e.stimated to be 4-5 days of age. The nestlings weighed 18.2, 25.5, and 26.5 g on 1 1 May; the smallest did not have its eyes open. We banded all 3 nestlings. We observed 22 provisioning trips to the nest on 12 May, 12 by the Abert’s and 10 by the Canyon Towhee. The Canyon Towhee flew south into desertscrub during forays from the nest, while the Abert’s Towhee flew north into the mesquite bosque for foraging. This difference in habitat preference by each adult was noted on numerous occasions. The young fledged on 18 or 19 May, and the adults immediately moved them into a wooded 400 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2. June 2010 region too thick for close observations. They led the fledglings back into open areas within ~1 week making them easier to observe. We observed parents feeding the young between 27 and 31 May. Particular young were generally favored by individ- ual parents, suggesting brood splitting (Marshall and Johnson 1968, Johnson and Haight 1996). However, on two occasions we observed a single fledgling being fed by both parents. One fledgling was observed bathing in a birdbath on 30 May. The female spent more time in the mesquite bosque during this period, and the male was more attentive to the young, leading us to suspect the pair was engaged in building a second nest. We observed the pair copulating on 29 May with the P. fiisciis in the superior position. The P. fusciis adult was observed alone in mid-June with one of the banded fledglings. The central breast spot of the young bird was evident at this stage but the throat patch was less conspicuous than that of the adult. The adult and young engaged in squeal -duet vocalization when meeting, a behavioral trait previously unre- ported between adult and young “brown towhees.” None of the banded young was observed following this initial post-fledgling period. We discovered both adult towhees feeding unbanded fledglings in late August 1999 and recorded their appearance (Appendix). We later saw only the adult mixed pair as they frequently visited the bird-feeding area together throughout the winter months. The parents were observed feeding three fledglings on 15 May 2000. A Cooper’s Hawk (Accipter cooperii) flew overhead as all five towhees fed on cracked corn in the open yard on 17 May. Both parents uttered a series of short, sharp, shrill peer or pink calls. These calls sounded identical in both parents and caused the fledglings to “freeze.” The young were observed sporadically (and described. Appendix) over the next few weeks and the adults remained on teiritory, but were not seen after November 2000. DISCUSSION Intrageneric Hybridization. — Several factors are important regarding this Tucson “brown towhee” hybridization. The.se include: (1) hybridization has not been reported earlier between P. aherti and P. fuscus in the Tucson region, although studies ot “brown towhees” have been conducted in this vicinity for more than 100 years (Bendire 1890; Marshall 1960, 1964; Marshall and John.son 1968; Johnson and Haight 1996); (2) the pair bond between congeneric species that are not sisters (Zink 1988, Zink and Dittmann 1991) was sufficiently strong to last 2 years and allow raising of at least three broods; (3) the female P. aberti selected a male P. fuscus at a locality where other P. aberti are not uncommon; and (4) the problem of detecting hybrids is exacerbated by the fact the Fj hybrids appeared so similar to P. fuscus. However, no interbreeding was found among several hundred color-banded birds in a multi-year study in an environment inhabited by both species, near Tucson, Arizona (Marshall 1960, pers. comm.). The residential yard near Tanque Verde Creek was in an ecotone which provided feeding sites. Both towhee species prefer feeding in open areas immediately adjacent to brush into which they may escape if disturbed (Marshall 1960, Dawson 1968, Marshall and Johnson 1968). Thus, the open yard between the brushy hillside and mesquite bosque provided an ideal situation for hybridiza- tion of P. fuscus and P. aberti. A second mixed pair of “brown towhees” occupied a territory during 2007-2009 in the foothills of the Tucson Mountains —12 km west of downtown Tucson (S. M. Russell, pers. comm.; RRJ, pers. obs.). The immediate surrounding landscape is a mixture of native upland Sonoran desertscrub with artificial plantings maintained by drip irrigation, forming habitat more mesic than native Sonoran desertscrub. This pair, similar to the Tanque Verde mixed pair, was first noted as they commonly appeared at and left a feeding station together. The pair would disappear into the adjacent desertscrub on leaving the feeding station. Male/female identity of this pair of P. fuscus and P. aberti remain unknown along with their breeding history. The pair built a nest in 2008 that failed to fledge young (S. M. Russell, pers. comm.). However, Russell noted an imma- ture “brown towhee” during the 2009 breeding season that he judged to have at least some P. aberti characteristics, e.g., black markings on the face. Russell noted only the single mixed towhee pair in the yard during this period with no pairs of either P. aberti or P. fuscus. During October and November Rus.sell noted two P. aberti (presum^ ably a pair) in this yard but no P. fuscus. Habitats are newly created through landscape disturbance that provide new niches. Processes are telescoped in time in human-disturbed habitats, compared to natural processes such as ecological succession. Both sets of mixed P. aberti and P. fuscus pairs in our study were in irrigated residential SHORT COMMUNICATIONS 401 yards. Disturbed habitat has had a profound influence on avian hybridization on the Great Plains in numerous passerine species. The entire Great Plains may now be considered a large hybrid zone between several related eastern and western species that nest in woody vegetation (Gill 1998). Shifting Habitat Selection by Abert’s Towhee. — A shift in habitat selection and local distribution by P. aberti has occuired recently in Tucson when arid desertscrub ecosystems were replaced by relatively mesic irrigated yards and parks. P . aberti was not listed as a resident of housing developments in Tucson’s suburban uplands in studies as recently as the 1980s (Tweit and Tweit 1986, Mills et al. 1989). P. aberti has now been recorded in suburban yards as far as 7 km from its original riparian habitat along the Santa Cruz River, increasingly encroaching on areas that had been occupied by P. fuscns (Tucson Audubon Society 2001-2009; RRJ, pers. obs.). These changes have increased the probability of P. aberti and P.fiiscHS occurring together and the possibility of hybridization. Whether P. aberti will displace P . fiiscus in much of suburban Tucson or the two species will coexist is unknown. A similar movement occurred by P. aberti in the Phoenix region during the 1900s. The species moved from the now dry Salt and Gila rivers into suburban yards and agricultural areas (Johnson et al. 1987, Rosenberg et al. 1987). However, in Phoenix, P. aberti does not come into contact with P. fuscus, which occurs only in more remote arid upland areas (Rea 1983, Johnson et al 1987). It is probable that disturbed native habitat in the Tucson area, as well as ecotonal factors, facilitat- ed the Tanque Verde towhee hybridization and Tucson Mountain mixed pairing. It is possible that hybridization is a recent phenomenon caused by newly created habitats stemming from human landscape alteration. It is also possible the difficulty in distinguishing hybrid offspring has led to an underestimation of the frequency of hybridization. Thus, a search should be undertak- en in avian collections for hybrid towhees. ACKNOWLEDGMENTS This paper benefited from discussions with J. T. Marshall, leading expert on “brown towhees,” and from discussions with L.T, Haight. Mitochondrial DNA analysis verifying the P. aberti was the female of the pair was conducted by R. M. Zink. S. M. and R. O. Russell assisted in banding young hybrids, and S. M. Russell provided information about the second mixed pair in the Tucson Mountains. We thank C. E. Braun, J. B. Dunning, and an unknown reviewer for criticjues of an early manuscript. LITERATURE CITED American Ornithologists’ Union (AOU). 1998. Check- list of North American birds. Seventh Edition. American Ornithologists’ Union, Washington, D.C., USA. Bendire, C. E. 1 890. Notes on Pipilo fuscus mesoleucus and Pipilo aberti, their habits, nests and eggs. Auk 7:22-29. Davis, J. 1951. Distribution and variation of the brown towhees. University of California Publications in Zoology 52: 1-120. Dawson, W. R. 1968. Pipilo aberti. Abert’s Towhee. Pages 632-638 in Life histories of North American cardinals, grosbeaks, buntings, towhees, finches, sparrows, and allies (O. L. Austin jR., Editor). U.S. National Museum Bulletin Number 237, Part 2. Gill, F. B. 1998. Hybridization in birds. Auk 1 15:281-283. Grant, P. R. and B. R. Grant. 1992. Hybridization of bird species. Science 256:193-197. Greenlaw, J. S. 1996. Spotted Towhee {Pipilo maculatus). The birds of North America. Number 263. Johnson, R. R. and L. T. Haight. 1996. Canyon Towhee {Pipilo fuscus). The birds of North America. Number 264. Johnson, R. R., L. T. Haight, and J. M. Simpson. 1987. Endangered habitats versus endangered species: a management challenge. Western Birds 18:89-96. Marshall Jr., J. T. 1960. Interrelations of Abert and Brown towhees. Condor 62:49-64. Marshall Jr., J. T. 1964. Voice in communication and relationships among brown towhees. Condor 66:345- 356. Marshall Jr., J. T. and R. R. Johnson. 1968. Pipilo fuscus mesoleucus. Canyon Brown Towhee. Pages 622-630 in Life histories of North American cardinals, grosbeaks, buntings, towhees, finches, sparrows, and allies (O. L. Austin Jr., Editor). U.S. National Museum Bulletin Number 237, Part 2. Mills, G. S., J. B. Dunning Jr., and J. M. Bates. 1989. Effects of urbanization on breeding bird community structure in southwestern desert habitats. Condor 91:416^28. Rea, a. M. 1983. Once a river: bird life and habitat changes on the middle Gila. University of Arizona Press. Tucson. Rosenberg, K. V., S. B. Terrill, and G. H. Rosenberg. 1987. Value of suburban habitats to desert riparian birds. Wilson Bulletin 99: 642-654. Sibley, C. G. 1954. Hybridization in the red-eyed towhees of Mexico. Evolution 8:252-290. Sibley, C. G. and F. C. Sibley. 1964. Hybridization in the red-eyed towhees of Mexico: the populations of the southeastern plateau region. Auk 81:479-504. Sibley, C. G. and D. A. West. 1958. Hybridization in the red-eyed towhees of Mexico: the eastern plateau populations. Condor 60:85-104. Sibley, C. G. and D. A. West. 1959. Hybridization in the Rufous-sided Towhees of the Great Plains. Auk 76:326-338. Tucson Audubon Society. 2001-2009. Tucson (Arizona) 402 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. 2. June 2010 bird count, www.tucsonbirds.org (accessed 19 August 2009). Tweit. R. C. .and D. M. Finch. 1994. Abeil’s Towhee (Pipilo aberti). The birds of Noith America. Number 111. Tweit. R. C. and J. C. Tweit. 1986. Urban development effects on the abundance of some common resident birds of the Tucson area of Arizona. American Birds 40:430-436. Zink, R. M. 1988. Evolution of brown towhees: allozymes, morphometries and species limits. Condor 90:72-82. Zink. R. M. and D. L. Dittmann. 1991. Evolution of brown towhees: mitochondrial DNA evidence. Condor 93:98-105. APPENDIX. — Characteristics of eight young Fi hybrids in three broods from Pipilo aberti X P. fusciis cross. Colors and characteristics were ascertained through field and photographic com- parisons with the parents, by examination of specimens in the University of Arizona bird collection, and from Birds of North America accounts (Tweit and Finch 1994, Johnson and Haight 1996). All characteristics are most like P. fuscus unless otherwise noted. General body color of all eight was intermediate between that of the two parents but most other characteristics were similar to adult P. fuscus. All showed throat patches varying from light to darker and striping around the patch varied from much to none. No young had the dark facial mask or uniform pinkish-brown of breast, flanks, and belly of female P. aberti. The only obvious P. aberti characteristic for any of the young was the dark rusty crissum (under-tail coverts) of five F, hybrids, in contrast to the lighter, buffy crissum of P. fiscus. Photographs available by using the second author’s complete name in a search engine or at http://zeeman.ehc.edu/envs/Hopp/index.html. Brood # 1 (May 1999; 3 Fi Hybrids Two with hint of rufous crown; third with uniform gray-brown crown. All with central breast spot. Two with more extensive, bright, rusty crissum, most like P. aberti. Brood # 2 (August 1999; 2 F) Hybrids) Both with slight but detectable rufous crown. Throat patch for both notably darker and less pronounced than in adult P. fuscus. One with slight breast spot. Tails of both longer than adult P. fuscus. Both with more extensive, bright rusty crissum, most like P. aberti. Brood # 3 (May 2000; 3 F, Hybrids) One with evident and two with slight rufous crown. One with central breast spot smaller and less conspicuous than adult P. fuscus, second with heavy streaking approaching a spot, and third without a spot. Tails of both similar in length to adult P. fuscus, slightly shorter than P. aberti. Bright rusty crissum in one, less evident in second, and third with no color difference between belly and crissum, appearing similar to front view of female Brown-headed Cowbird (Molothriis a ter). The Wilson Journal of Ornithology 1 22(2):402^05, 2010 Yellow-billed Cuckoo Hatched and Fed by a Red-winged Blackbird Ken Yasukawa' ABSTRACT. — I observed a Red-winged Blackbird (Agelaius phoeniceus) nest in southcentral Wisconsin, USA that received a Yellow-billed Cuckoo (Coccyziis ainericanus) egg. The cuckoo egg was laid after the blackbird clutch was complete and the female had begun incubation. The parasitic egg was considerably different in size and color, but was accepted by the female Red-winged Blackbird. The female, but not the male, afso cared for the nestling, which was strikingly different in plumage, gape markings, and begging behavior, and which grew despite its 3^ day hatching ' Beloit College. Department of Biology. Beloit, Wl 5351 I. USA: e-mail: yasukawa@bcloit.edu delay relative to its nest mates. Rapid development and exaggerated begging behavior of the Yellow-billed Cuckoo nestling may enable a parasitic cuckoo nestling to compete successfully with host nestlings for food. Received 24 August 2009. Accepted II December 2009. Yellow-billed (Coccyzus ainericanus) and Black- billed (C. erythropthahnus) cuckoos are the only altricial birds that lay in the nests of other birds as apparent facultative, as opposed to obligate, inter- specific brood parasites (Hughes 1999). Wyllie (1981) noted intraspecific parasitism also occurs. SHORT COMMUNICATIONS 403 FIG. 1. Nestling Yellow-billed Cuckoo (right) begging in a brood of Red-winged Blackbirds. and Fleischer et al. (1985) used egg protein analysis to verify that two female Yellow-billed Cuckoos contributed to a large clutch. The majority of records of interspecific parasitism are between Yellow- billed and Black-billed cuckoos (Nolan and Thomp- son 1975), but 1 1 other species are known to have been parasitized by Yellow-billed Cuckoos. Many of the unusual aspects of North American cuckoo reproductive ecology, including the tendency to vary clutch size in relation to food supply and rapid nestling development, may have facilitated evolu- tion of brood parasitism in cuckoos (Nolan and Thompson 1975). I report an unusual case of brood parasitism of a Red-winged Blackbird {Agelaiiis phoeniceiis) by a Yellow-billed Cuckoo and discuss an aspect of cuckoo nestling behavior that may have facilitated the evolution of parasitism. OBSERVATIONS On 21 May 2009 I found a Red-winged Blackbird nest containing four incubated eggs at Newark Road Prairie in southcentral Rock County, Wisconsin (42° 32' N, 89° 08' W). I checked this nest daily at ~ 0530 hrs CST. On 24 May I found an egg of a Yellow-billed Cuckoo in addition to the four blackbird eggs. The cuckoo egg was considerably larger than the blackbird eggs (Hughes [1999] gives a mean egg mass of 9.1 g for eastern populations of Yellow-billed Cuckoos, whereas Yasukawa and Searcy [1995] give mean masses of 4.02-4.07 g for Red-winged Blackbirds). The egg was pale bluish-green and unmarked, unlike the blackbird eggs, which were pale blue with dark brown blotches. All five eggs were warm to the touch, indicating the female was incubating them. On 28 May I observed one newly hatched blackbird nestling and by the next morning, there were three blackbird nestlings along with one blackbird egg and the cuckoo egg. When I checked the nest on the morning of 1 June, the Yellow-billed Cuckoo egg had hatched. The cuckoo nestling was thus 4 days younger than the first-hatched blackbird nestling and 3 days younger than the other two blackbird nestlings. The fourth blackbird egg did not hatch and showed no signs of development when 1 examined its contents on the morning of 3 June. The Yellow-billed Cuckoo nestling was strik- ingly distinct from the Red-winged Blackbird nestlings in appearance (Fig. I ) and behavior. Unlike its blackbird nest mates, when begging the cuckoo nestling rapidly vibrated its head, pro- duced a sustained hissing begging call while showing its distinctive gape with white spots. Happed its wings rapidly, and attempted to gain a position above the blackbird nestlings by standing 404 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 2, June 2010 on them. In contrast, the blackbird nestlings moved their heads only slightly as they stretched their necks and gaped. The blackbird gape lacked the white spots of the cuckoo, and the blackbird nestlings did not flap their wings or jostle for position. The blackbird begging call was a series of discrete chirps rather than a sustained hiss. I made a 1-hr video recording of the nestlings on 4 June to document the differences in nestling begging behavior between the two species. Unfortunately, a predator, probably a common raccoon (Procyon lotor), took the nestlings that night, and I could not continue my observations of this unusual case of brood parasitism to learn whether the fledgling would have received care. DISCUSSION Unusual reproductive behavior in Yellow-billed Cuckoos is associated with abnormal food abun- dance (Nolan and Thompson 1975) but, to the best of my knowledge, 2009 was not an outbreak year for eastern tent caterpillars (Malacosoma ameri- cana) or periodic cicadas {Magicicada spp.), and the numbers of annual cicadas (Tihicen spp.) did not seem unusual. The most recent outbreak of periodic cicadas in Rock County, Wisconsin was 2007 (Hamilton 2007). The egg and nestling of the Yellow-billed Cuckoo were quite different from those of the host Red-winged Blackbird, but the egg was accepted (incubated) and the female blackbird fed the nestling. I did not observe the male blackbird feeding nestlings. Nickell (1954) re- ported a similar case in Oakland County, Michigan. Furthermore, the 3^ day difference in hatching did not appear to disadvantage the cuckoo nestling, whereas a blackbird nestling experiencing the same hatching asynchrony would normally die of starvation (Holcomb and Twiest 1971, Strehl 1978). Red-winged Blackbirds accept Brown-headed Cowbird {Molothrii.s ater) eggs and care for cowbird nestlings and fledglings, and Brown- headed Cowbird parasitism is common on my study area (Clotfelter and Yasukawa 1999). Yellow-billed Cuckoos are observed on my study area in most breeding seasons as well, although they are not common. Given the number of studies of Red-winged Blackbird reproductive biology, extensive overlap in Red-winged Black- bird and Yellow-billed Cuckoo breeding ranges, and dearth of similar records in the literature. Yellow-billed Cuckoo parasitism of Red-winged Blackbirds is clearly rare. Nolan and Thompson (1975) suggested that North American cuckoos have traits that facilitate the evolution of brood parasitism (see also Hughes 1999). These traits include nomadic movements in response to food abundance, large egg size, variable clutch size in response to food availability, intraspecific brood parasitism, occa- sional use of abandoned nests of other species, short breeding cycle (17 days from egg laying to fledging is among the shortest of any species of bird), rapid nestling development (nestlings can raise their heads above the nest rim and are fully feathered within 2 hrs after hatching), extreme hatching asynchrony, and an unusual diet of food items that can be difficult to find. Yellow-billed Cuckoos occasionally lay eggs in nests of other species (called “incipient parasitism” by Nolan and Thompson 1975), and a necessary behavior is in place to allow evolution of brood parasitism if parasitism was to become more advantageous than parental care (Hamilton and Orians 1967). North American cuckoos have not evolved obligate brood parasitism, however, perhaps because species recognition, song learning, and ecological constraints act to maintain the advan- tage of parental behavior. Acceptance of a Yellow-billed Cuckoo egg and nestling by a Red-winged Blackbird may be another indication of the plastic nature of parental care in North American songbirds (Shy 1982). However, the exaggerated begging behavior of the nestling and its ability to overcome hatching 3^- days later may indicate they facilitated the Yellow-billed Cuckoo’s ability to secure food from the female Red-winged Blackbird despite the obvious differences in plumages, begging calls, and begging behavior of the two species. The ability to compete for food in novel and/or more exaggerated manners than host nestlings would also facilitate the evolution of obligate brood parasitism. ACKNOWLEDGMENTS 1 dedicate this .short communication to my mentor, Val Nolan Jr. I thank C. S. Anderson for field assistance and Beloit College for permission to work at Newark Road Prairie and financial support. S. G. Sealy and .1. M. Hughes provided background information on North American cuckoo parasitism, and E. D. Ketter.son, E. D. Clotfelter, and two anonymous reviewers made suggestions to improve earlier versions of the manuscript. SHORT COMMUNICATIONS 405 LITERATURE CITED Cl.OTFELTER, E. D. AND K. Yasukawa. 1999. Impact ot brood parasitism by Brown-headed Cowbirds on Red- winged Blackbird reproductive success. Condor 101:105-114. Fleischer, R. C., M. T. Murphy, and L. E. Hunt. 1985. Clutch size increase and intraspecific brood parasitism in the Yellow-billed Cuckoo. Wilson Bulletin 97: 125-127. Hamilton, K. 2007. Wisconsin Pest Bulletin 52 (10), 15 June 2007. Trade and Consumer Protection. Wisconsin Department of Agriculture, Madison, USA. pestbulle- tin.wi.gov/pests.jsp?categoryid=32&issueid = 84 (ac- cessed 1 December 2009). Hamilton 111, W. J. and G. H. Orians. 1967. Evolution of brood parasitism in altricial birds. Condor 67:361-382. Holcomb, L. C. and G. Twiest. 1971. Growth and calculation of age for Red-winged Blackbird nestlings. Bird-Banding 42:1-17. Hughes, J. M. 1999. Yellow-billed Cuckoo (Coccyzus onicf'icofius). The birds ol North Ameiica. Number 418. Nickell, W. P. 1954. Red-wings hatch and rai.se a Yellow- billed Cuckoo. Wilson Bulletin 66:137-138. Nolan Jr., V. and C. F. Thompson. 1975. The occurrence and significance of anomalous reproductive activities in two North American non-parasitic cuckoos Coccy- zus spp. Ibis 1 17:496-503. Shy, M. M. 1982. Interspecific feeding among birds: a review. Journal of Field Ornithology 53:370—393. Strehl, C. 1978. Asynchrony of hatching in Red-winged Blackbirds and survival of late and early hatched birds. Wilson Bulletin 90:653-655. Wyllie, I. 1981. The cuckoo. Universe Books, New York, USA. Yasukawa, K. and W. a. Searcy. 1995. Red-winged Blackbird {Agelaius phoeniceits). The birds of North America. Number 184. The H'ilsoii Journal of Ornithology 122(2):406-415. 2010 Ornithological Literature Robert B. Payne, Review Editor FIELD GUIDE TO THE SONGBIRDS OE SOUTH AMERICA; THE PASSERINES. By Robert S. Ridgely and Guy Tudor. University of Texas Press, Austin, Texas, USA. 2009: 750 pages and 122 color plates with facing-page maps. ISBN: 978-0-292-71979-8. $49.95 (paper).— There is much to celebrate with publication of this revamp and update of the first two volumes of The Birds of South America {BOSA: Oscine passerines, 1989; Suboscine passerines 1994) in a more compact format under a single cover. Undoubtedly the most anxiously awaited aspect of Songbirds is the tremendous contribution of Guy Tudor’s new art; illustrations of 406 species (some as only busts) have been added to the — 1,220 previously depicted ones, a whopping 33% increase. There remain not illustrated about 347 native species of South American passerines, roughly 18% of the authors’ total number of — 1,974 (excluding a few vagrants). Now repre- sented are all but one (I think) standing genus (Stymphalornis antbirds, endemic to southeast Brazil). The back cover calls attention to the need for “a more compact guide that birders can take into the field.” Given that they intend to show us how to identify 1,981 species of passerine birds, some of which are hard to see well, all while traipsing through the rainforests and de.serts on the Bird Continent, they may have set themselves up for disappointment. On some levels, they have succeeded, it’s just hard to say whether they have tackled the monster or the monster has tackled them. Weighing in at 1.28 kg, the book can be carried in the field by most birders, but not very comfortably by anyone. Perhaps most tellingly, almost no one has opted to tote it on recent South American birding walks where most folks would like to be carrying a field guide. I’ve not yet seen anyone with the plates cut out and bound .separately, all too often the fate of BOSA, but the specter has already come up. The ideal bulk may lie about halfway between Songbirds and the much smaller Birds of South America: Non- passerines (Erize el al. 2006, Princeton University Press). Tudor’s illustrations are wonderful; it’s hard to imagine ever seeing paintings of many of these birds that are more “spot-on.” In large measure because he is an avid and highly observant birder who understands the importance of accurate shapes and postures, Tudor is able to encapsulate what we need for field identification in the strict confines of afield guide. As is inevitably the case when a large number of species is treated, a few illustrations leave much to be desired (most of which are holdovers from BOSA). White-bearded Antshrike {Biatas nigropectus, Plate 20, #3) never appears uncrested (except in a museum tray) — the new, undersized male figure and a female bust, at appropriate scale, could have admirably replaced the poor BOSA ones; Plain-crested Elaenia {Elaenia cristata, Plate 44, #7), unrecognizable as is, shows a shaip, spiky crest virtually all the time; White-winged Cotinga (Xipholena atropur- purea, Plate 69, #7) is a much darker “blackish- purple” as described in the text; Black-collared Swallow (Atticora melanoleuca, Plate 76, #10) invariably looks black-and-white (not navy blue) in the field even at close range, and the posture is atypical; and all of the pipits (Anthus spp., Plate 84, perhaps with the exception of Correndera Pipit, A. correndera) remain unidentifiable in and of themselves. Other closely similar groups of birds, such as the dreaded “tyrannulets,” are well done al- though there are regrettable problems with a few species that occur in syntopy (together in the same habitat) through much of their ranges, particularly Rough-legged Tyrannulet (Phyllomyias burmeis- teri, Plate 41, #16); Greenish Tyrannulet (P. vire.scens, Plate 41, #17); and Mottle-cheeked Tyrannulet [Phylloscartes ventralis, Plate 46, #4). The most similar-looking small tyrannids, scat- tered over .several plates (mostly 41, 42, 46), could have been grouped in a much more useful manner for field identification, for starters by separating exclusively Andean from non-Andean species on different plates, then bunching the members of each broad group that actually overlap geographically and occur in the same habitats. Because the small South American tyrannids, too many of which are not illustrated, present the most complex field identification 406 ORNITHOLOGICAL LITERATURE 407 problems (female Sporophila seedeaters aside), I had hoped they would be given high priority in Songhii'cls. The choice was made to insert many ot the new illustrations at varying scales different from the rest of the (existing) figures on the plate. Some of the most problematic instances are Plate 52, #8, 11; Plate 53, #3, 4, 11, 13 relative to the themselves and the rest of that plate; Plate 61, #5, 7; Plate 7 1, #4; Plate 101, #2, 3; and Plate 1 14, #12 (Dickcissel [Spiza americaiia] should be significantly larger than Greenish Yellowfinch [Sicalis olivascens]'.). It is disheartening to see the stunning Araripe Manakin {Autilophia hoker- manui, Plate 64, #17) relegated to a corner- squeeze at about half-size, below its sister, Helmeted Manakin (A. galeata). Its illustration points to another unfortunate feature of all plates having new illustrations: something happened in the production stage so all of the new figures look faded relative to those in BOSA. Thus, the red and black on the Araripe Manakin are significantly paler or less saturated than these same colors on the Helmeted Manakin. A number of figures of females also have been newly illustrated, an important advancement, but a field guide really must treat all distinctive female plumages (of the species illustrated at all). Almost no juvenal plumages are shown or described. One of the few, Andean Laniisoma {Laniisoma huckleyi, Plate 66, #8), is mislabeled the adult female. I love the new illustrations of the mockingbirds (Plate 83) in flight; it would have been great to see much more of this kind of field-oriented illustra- tion. On the plates, each species illustration is numbered and some also have letters A, B, C, indicating the figure represents a distinctive subspecies. To interpret this information, one needs to go to the text. This cumbersome system could have been made much less so by simply putting the letters on the species’ map to inform users of where that taxon (or group of taxa) is expected. The South American distribution of each species is shown on a correspondingly numbered map on pages opposite or near its illustration, sometimes requiring a little hunting although I got used to it. I really like these maps. Several levels of geographic zoom are used to show most of the highly restricted ranges more clearly. The contrast between colors chosen for ranges, including migrants, and also rivers and political boundaries for most countries (to the state/department level in the tightest zooms), also works well. Without having looked at most of the maps critically, I think it’s fair to say the level of accuracy is admirably high. 1 happened to notice that the maps for Black Antbird {Cercomacra serva) and Blackish Antbird (C. iiigrescens) were switched (their names are bad enough!). Similarly, numbers labeling the maps and figures for Green Honey- creeper (Chlorophanes spiza) and Red-legged Honeycreeper (Cyaiierpcs caeruleus) are mixed up (numbers are correct on the maps and text, so switch the species’ numbers on the plate of your copy to match). Problems on the level of not showing Saffron-crested Tyrant-Manakin {Neo- pehna chrysocephahim) in northern Peru; of showing Long-winged Antwren {Mynnothenila longipennis) extensively between the Rio Napo in Peru and the Negro in Brazil where it is largely or completely absent; of missing the interface between Squamate Antbird (Myrmeciza squa- mosa, shown extensively north into Rio de Janeiro) and White-bibbed Antbird (M. loricata, shown far too extensively through interior Bahia, and ranging well into northern coastal Sao Paulo); of White-breasted Antbird (Rhegmatorhina hojf- mannsi) crossing (and extensively so) the Rio Juruena into northem Mato Grosso; of Seira do Mar Tyrannulet (Phylloscartes dijficilis) occumng all along the coast of southeast Brazil where known from only a few scattered points high in the mountains; and of showing the threatened Black-and-white Monjita {Xolmis dominicauus) over much too wide an area in Rio Grande do Sul (it was never known to be so amply distributed, even historically) — are somewhat more frequent. Maps for a few very poorly known birds are annoyingly problematic. For example, an undoc- umented, single-observer sight-record of the ultra- rare Kinglet Calyptura {Calyptura cristata. Plate 66, #6) in coastal northern Sao Paulo is mapped and cited in the text as if it were fact! This map should show only the single point of known occurrence (from late Oct 1996) just below Teresopolis, Rio de Janeiro with an aiTOW pointing to a red dot and at most a question mark anywhere else; the maximum zoom should have been provided here. In a field guide, the species accounts should foster transformation of the painting on the plate to a living bird by infusing vital information on habitat and behavior while enabling even the most difficult field identifications. My overall impres- 408 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 2, June 2010 sion is that over 90% of the species accounts in Songbirds are good to excellent; space limitations in the Journal precluded my elaboration of examples. Transcriptions of vocalizations of most species are offered, which is usually better than not having anything but not nearly as helpful as actually hearing the sounds. I highly recommend obtaining recordings (primary vocalizations of most species have been published or are available on the internet) ahead of travel to South America. The taxonomy used in Songbirds closely parallels that of the standing South American Checklist Committee (SACC) of the American Ornithologists’ Union. The SACC was obviously heavily referenced although it was not cited anywhere in a 16-page section toward the back of the book titled “Notes on Taxonomy and English Names.’’ However, the authors “express- ly note’’ (page 2) that they “adhere to the decisions on English names and taxonomy’’ that were recently published on behalf of the Interna- tional Ornithological Congress (IOC; Gill et al. 2006-2009, Birds of the World, Recommended English Names, Princeton University Press). The first bird in the book, Campo Miner (Geositfa poeciloptera), provides a poignant example of the taxonomic swirl that underlies Songbirds. It boils down to; (1) the IOC classification, which incorporated new information up to the end of 2004, maintaining the long-standing Geobates contrary to a 2003 publication suggesting it be merged into Geositta-, (2) the SACC invoking primarily genetic data published in 2005 to merge Geobates into Geositta; and finally (3) Songbirds incontrovertibly following the SACC, not the IOC. Ornithologists and birders frequently express their opinions that some taxa or populations of birds could or should be considered distinctive enough to merit species status. The unilateral splitting or lumping of species by authors of such works as Songbirds and other guide books in which no data or analyses of the case(s) are presented for peer-review or independent evalua- tion (now or a hundred years from now) reside in a bubble drifting around the perimeter of science, a “gray” subspecies of taxonomy. In Songbirds, Ridgely and Tudor are “gray’’ on —25 taxa traditionally considered subspecies. All are inter- esting cases worthy of focu.sed study, and the SACC and other taxonomic committees will soon consider not only all of them but also all of the equally or more interesting cases that, for one reason or another (we have no data to judge), were not mentioned in Songbirds. For one example, the Stigmatiira wagtail-tyrants go from two to four species absent any analysis at all but the well- supported (and accepted by the SACC) elevation to species status of Schistocichla lencostigma satiirata (Roraiman Antbird, in the Spot-winged Antbird complex) published in 2005 was not recognized. English names are not bound by scientific principles, and their u.sage for South American birds became particularly unstable after the mid- 1980s, when BOS A introduced changes to names of long standing. In Songbirds, the 2006 IOC classification is followed, even in cases where a novel name of Ridgely’ s from BOSA (e.g.. Noble Antthrush for Chamaeza nobilis) was rejected in favor of the traditional name (Striated Antthrush). But this has set up some conflicts. For example, a nomenclatural note on page 691 reads, “We rename Mitrephanes phaeocercus the Northern (rather than the Common) Tufted Flycatcher, prefeiTing to restrict the use of the name “Common’’ to species that truly are so.’’ This flycatcher was described from Mexico and most of its range lies in Middle America. Ridgely and Gwynne (1989, A Guide to the Birds of Panama, Princeton University Press) introduced the mod- ifier “Common,’’ and described its status in Panama as “Common.’’ Pointless self-reversal like this does not inspire confidence and promotes instability of names. True to form, Ridgely has not resisted unveiling a few English names that almost seem designed to shake things up. My vote for the standout is ... “Shrike-like Tanager’’ (in place of the entrenched White-banded Tana- ger, Neothraiipis fasciata), not to be confused with Lanio Shrike-Tanagers. In sum, there is a great deal of arbitrariness displayed throughout the taxonomy and English naming used in Songbirds. The new Ridgely and Tudor Field Guide to the Songbirds of South America should become a standard in the libraries of all birders, ornitholo- gists, and con.servationists interested in neotrop- ical birds, and it is sure to make birds more accessible to more people, which is at the grassroots level of effective bird conservation. If you ever dream of birding in South America, buy and thoroughly enjoy this book. Take it to South America with you and use it to ably identify the vast majority of passerine birds you will see. — BRET WHITNEY, Louisiana State Univer- sity Museum of Natural Science, 119 Foster ORNITHOLOGICAL LITERATURE 409 Hall, Baton Rouge, LA 70803, USA; e-mail; ictinia@earthlink.nel PALEOGENE BIRDS. By Gerald Mayr. Springer-Verlag, Berlin Heidelberg, Germany. 2009: 262 pages and 64 figures. ISBN: 978-3- 540-89627-2 ($189.00 USD, cloth; ~$25 per chapter in electronic form). — The Paleogene Period, from —65 to 23 million years ago, witnessed profound shifts in the earth system including the origin of permanent ice sheets in Antarctica. Global cooling from peak “greenhouse Earth” conditions is linked to a retreat of Eocene paratropical forests previously extending to the poles and the develop- ment of the “icehouse” conditions of the present. Across the same time period, shifts in the ecology, diversity, and distribution of nearly every part of the biota are inferred. New insights into every part of these Paleogene shifts have been rapidly accruing in recent years. The record of sea-surface temperatures and global climates has been significantly refined, and vertebrate paleontologists continue to further inform our knowledge of changing faunas. The rapid pace of new discoveries of fossil bird species in recent years belies the general idea of their impossibly fragmentary record. The early history of extant crown clade avian diversity, species of which are parts of the major bird lineages alive today, has been partially tracked in several short review articles eager to summarize the rush of new data recovered in recent years. Because relevant paleo-ornithological discoveries are moving forward at an arguably unprecedented pace, any summaries must be proximate, abbre- viate, and soon out of date. However, the last book-length review focusing on an overlapping time period and set of taxa is the 1999 Second Edition of A. Feduccia’s “The Origin and Evolution of Birds’’ (Yale University Press). With a few exceptions, including updating several chapters presenting the author’s views on the dinosaurian affinities of birds, most ol its content is from the 1996 First Edition. It is into this context that Gerald Mayr, Curator of Ornithology at the Senckenberg Institute, offers a significant contribution, a detailed new guide to fossil avian diversity from the first part of the Cenozoic. In —250 dense small-format pages, “Paleogene Birds" neatly presents insights into —40 million years of the early evolution of extant bird lineages. The book is largely organized taxonom- ically but also includes a brief overview of principal fossil localities, a review of debates concerning the relationships among major clades, and conclusions regarding the historical patterns of avian diversity in the Northern and Southern hemispheres. Mayr, who recognizes in his preface that any review in the present intellectual context will quickly need updating as new discoveries are made, quite thoroughly summarizes the earlier literature but focuses on key recent discoveries, a great many of which he has made. The commu- nity would profit from this book even if it were merely a synthesis of Mayr’ s own work. A prolific author, he has published over 150 papers since 1995, more than any other paleo-ornithologist over the same time period. Ornithologists of all kinds will be intrigued by the bizarre turns that avian evolution took in the first part of the Cenozoic. The book presents exceptional fossils that record novel ecologies and distributions for well known groups; e.g.. Old World hummingbird relatives, giant flight- less divers, and sea birds with bony pseudo- teeth. However, at -$189 USD for a new copy of this slim volume, many non-specialists may elect to visit their local libraries or hope for a lower-price soft cover edition. The intended primary readership of the book is the paleo- ornithological community. However, it will also serve as an important reference for other ornithologists and paleontologists seeking ac- cess to specific parts of the literature. The book seems organized for, and best approached by, skipping from section to section, rather than reading from cover to cover. Illustration print quality is excellent, although the specialist might wish for an oversized volume with color illustrations and larger photographs. One is well served by the detailed presentation of Mayr’s perspective as that perspective is often thought-provoking. Much of the data presented is fascinating, and comments made as small asides could be expanded into journal articles of considerable interest. Some of the more contro- versial ideas forwarded by Mayr and summarized in the book concern phylogenetic relationships among major lineages of birds; e.g., regarding the affinities of penguins with parts of former “Pelecaniformes.” While not all readers will agree with these hypotheses, all are supported by clearly presented evidence that is well positioned to inform new research questions. We still have much to learn about patterns and 410 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 2. June 2010 potential drivers of avian evolution. Mayr clearly stands with the rest of us in wondering what new bizarre forms we may discover tomorrow. It is striking how few avian taxa are generally shown in the skies of Paleogene ecosystem reconstructions. It is hard not to feel as if we have been missing significant patterns in vertebrate evolution across the Cenozoic greenhouse to icehouse transition. This book offers an important opportunity to renew our focus on the diversity and distributions of birds as significant parts of Paleogene environments. — JULIA A. CLARKE, Associate Professor, University of Texas, Department of Geological Sciences, 1 University Station Cl 100, Austin, TX 78712, USA; e-mail: Julia_Clarke@jsg. utexas.edu BIRDS OF BORNEO. By Susan Myers. Princeton University Press, Princeton, New Jer- sey, USA. 2009: 272 pages, 1,592 color illustra- tions, 630 maps. ISBN: 978-0-691-14350-7. $29.95 (paper). — At last, ornithologists and bird- watchers in Borneo have a complete and accurate guide to birds that they can fit in their pockets. For 30 years, folks have worked to this end, and it is extremely gratifying to have reached it. Starting in the late 1970s, Bertram Smythies’ magnificent handbook. The Birds of Borneo ( 1960, 1968), was out of print and the first modern field guide to the birds of southeast Asia (King et al. 1975) did not cover Borneo. Thus, there was a great need for a handy book on Bornean birds. The Sabah Society helped alleviate the problem by publishing an inexpensive version of Smythies (1981), but it was too big for field use. Charles Francis (1984) also helped by producing a pocket guide consist- ing of Smythies’ plates, but the illustrations without text were not adequate for identification of many species. For the last 1 5 years, Bornean bird enthusiasts have done pretty well with MacKinnon and Phillipps (1993), the first field guide to birds of the Indo-Malayan Archipelago. It has excellent illustrations by Karen Phillipps, but is now out of print. Also, because it includes all the birds of Sumatra, Java, and Borneo, it is too big and bulky for comfortable use in the field. Becairse it was published before knowledge of Bornean bird distribution and ecology had quite gelled, it also contains lots of small mistakes. Those problems are now history. The tremen- dous recent growth in .sophisticated birdwatching in Borneo, combined with comprehensive updates in the scientific literature (Smythies 1999, Shel- don et al. 2001), set the stage for a good field guide. The guide’s author, Susan Myers, has led Victor Emmanuel Nature Tours in eastern Asia for more than 15 years. Thus, she has been an integral part of the ecotourism phenomenon. She combines great experience in finding and identi- fying birds in Borneo with a grand perspective of Asian birds as a whole. The result of these congruent forces is a truly outstanding field guide. It is small, has extensive color illustrations of all the birds, a complete set of maps, and all the information needed on plumage, distribution, habitat, elevation, and behavior to sort out the species. It really is excellent. The quality of any field guide depends on insightful, pithy descriptions, and high quality illustrations. This book does well in both areas. A good test of its description quality is how well the author deals with difficult groups. Such a group in Borneo is the Cyornis flycatchers. These blue and brown birds of the forest understory look much the same, have nondescript songs, and often confuse newcomers (Smythies resorted to a key to deal with them). Myers, with simple, skillful descriptions of plumage, similar species, micro- habitats, and behavior makes it easy to sort them out. As for illustrations, each species is depicted individually next to its description and distribu- tion map, instead of in plates of similar taxa. Sexually dimorphic plumages, age variants, and occasionally alternative subspecies are provided. Most of the figures have been reproduced individually from the plates in Robson (2000), but 295 (of 180 species) were prepared specifi- cally for this book. This helps explain the large number of contributing artists (16) and the variation in illustration quality. Despite variation, only a couple of paintings are sub-par (e.g., Malaysian Rail-babbler [Enpetes macrocerus] and crows). All the illustrations are more than adequate for identification, and most are really good. 1 particularly liked the trogons, hornbills, bulbuls, babblers, and muscicapine flycatchers. My favorites are the wren-babblers. By virtue of its simplicity and thoroughness, this is an outstanding field guide. 1 really cannot wait to take it into the fore.st and give it a field test. No doubt it will hold up well. — FREDERICK H. SHELDON, Louisiana State University Muse- um of Natural Science and Department of Biological Sciences, Baton Rouge, LA, 70803, USA; e-mail: fsheld(®lsu.edu ORNITHOLOGICAL LITERATURE LITERATURE CITED Francis, C. M. 1984. Pocket guide to the birds of Borneo. Sabah Society and World Wildlife Fund Malaysia, Kuala Lumpur, Malaysia. King, B. F., E. C. Dickinson, and M. W. Woodcock. 1975. A field guide to the birds of south-east Asia. Collins, London, England. MacKinnon, J. and K. Phillipps. 1993. A field guide to the birds of Borneo, Sumatra, Java, and Bali. Oxford University Press, Oxford, United Kindom. Robson, C. 2000. A guide to the birds of southeast Asia. Princeton University Press, Princeton, New Jersey, USA. Sheldon, P. FI., R. G. Moyle, and J. Kennard. 2001. Ornithology of Sabah: history, gazetteer, annotated checklist, and bibliography. Ornithological Mono- graphs 52:1-285. Smythies, B. E. 1960. The birds of Borneo. Oliver and Boyd, London, England. Smythies, B. E. 1968. The birds of Borneo. Second Edition. Oliver and Boyd, London, England. Smythies, B. E. 1981. The birds of Borneo. Third Edition. Sabah Society and Malayan Nature Soci- ety, Kuala Lumpur, Malaysia. Smythies, B. E. 1999. The birds of Borneo. Eourth Edibon. Natural History Publications, Kota Kina- balu, Malaysia. BIRDS AND BATS OF PALAU. By H. Douglas Pratt and Mandy T. Etpison. Mutual Publishing L.L.C., Honolulu, Hawaii, USA. 2008; 290 pages. ISBN: 978-1-56647-871-7. $29.95 (paper). — Palau lies at the western edge of Micronesia. As a result of its relative proximity to the rich avifaunas of Asia, the Philippines, and New Guinea, Palau sustains a larger set of both resident and migrant birds than any other Micronesian island group. Most of Palau also is blessed with a rugged, raised limestone topogra- phy that has stifled deforestation and human habitation. It’s a great place for birdwatching. When I visited Palau in 1995 and 1997, I carried two field guides. The first was the rather small Field Guide to the Birds of Palau by John Engbring (1988, Conservation Office, Koror, Palau), which covered mainly resident species. The second was the more comprehensive A Field Guide to the Birds of Hawaii and the Tropical Pacific by H. Douglas Pratt, P. L. Bruner, and D. G. Berrett (1987, Princeton University Press, Princeton, NJ), which covered all island groups in Micronesia and Polynesia. H. Douglas Pratt and Mandy T. Etpison now have produced a nice, up- 41 1 to-date field guide for the birds (and bats) of Palau, whether resident or migrant. The text of this book has an informal style of writing, which may be good for some readers and not so good for others. There are no subheadings in the species accounts, and one must look through paragraphs of text to glean information on field marks, vocalizations, etc. The scattering of typos, un- italicized scientific names, grammatical errors, etc. is not enough to distract the average reader. The only maps are aerial photographs or satellite imagery; while attractive with their intense shades of blue and green, the lack of a clearly labeled, line-drawing map hinders one’s understanding of Palauan geography. The overall design of the book is lively and attractive. The species accounts feature magnifi- cent photographs, mainly by Etpison. Because nearly all of the photographs are unlabeled, however, you must make your own judgment about the age and gender of the bird depicted (if those attributes are determinable), not to mention what it was doing and where the photograph was taken. The single color plate of endemic birds, painted by Pratt, is excellent although it is placed, without warning and essentially unreferenced, in the introductory material. The 19 species of scolopacid sandpipers depicted on pages 199 to 207 are erroneously placed in the plover family (Charadriidae). Disagreement about common names of birds abounds in the tropical Pacific, so it is little surprise that some of these issues surface in Birds and Bats of Palau. Eor example, following the lead of F. Gill and M. Wright (2006, Birds of the World: Recommended English Names. Princeton University Press), Pratt and Etpison call Sterna hergii (page 224) the “Swift Tern’’ rather than the more conventional, non-apodid name Great Crested Tern. On pages 28 and 228-231, I congratulate the authors for not adopting Gill and Wright’s name “Angel Tern’’ for Gygis alha (= Gygis Candida), although they do use this simply heavenly name in the u.seful checklist of species on pages 278-282, which presents the distribution of each species on each major island. I’m not a big fan of proposing new common names in field guides, even though this practice has become widespread in these days of over- splitting to increase perceived endemism. The account of the kingfisher Todiramphus (often placed in the genus Halcyon) cinnamominus pelewensis helps explain my reservations. Usually 412 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 2. June 2010 known as the Micronesian Kingfisher with distinctive subspecies on Guam (T. c. cinnamo- miniis), Palau (7. c. pelewensis), and Pohnpei (7. c. reichenbachii), Pratt and Etpison now call it the Rusty-capped Kingfisher and regard it as endemic to Palau at the species level, even though they list it as 7. c. pelewensis in the species account, checklist, and index. Here I quote the first and last paragraphs of the species account (pages 88-90) as they lay out their thinking about this bird. “In most previous books this bird has been called Micronesian Kingfisher because it was regarded as one of 3 subspecies that make up that species. However, HOP has accumulated evidence (see below) that shows that those 3 supposed subspecies could not possibly interbreed success- fully and that there is really no such thing as the ‘Micronesian Kingfisher.’ Until the details are properly published, however, we continue to use the traditional taxonomy, but re-christen this endemic Palau soon-to-be species with a distinc- tive English name.’’ “Differences among the so-called ‘Micronesian Kingfishers’ include voice, adult and juvenile plumage coloration, and nesting habits. Those on Pohnpei are larger, and look very much like Rusty-caps as adults, but the Juveniles resemble the strikingly different Guam kingfisher. In that one, the white parts of the Rusty-cap’s plumage are a rich rusty chestnut, and. unlike the other 2, adult females are distinctive with a white belly, but Juveniles are like adults. These visual differences are what scientists call an ‘isolating mechanism.’ This is any behavioral, ecological, or physical characteristic that prevents two species from interbreeding, during the mating process or after pairs have formed. For example, Pohnpei kingfishers nest in cavities in termite nests, while the Rusty-capped uses tree cavities, since the termite nests in Palau are used by Collared Kingfishers. So even if, hypothetically, a Pohnpei bird and Rusty-cap would mate, they would not be able to decide where to nest! And we suspect that the 3 Micronesian forms might not recognize each other visually either.’’ Aside from the pitfalls of trying to mix popular writing with supposedly .scholarly thinking, and issues of biological versus phylogenetic species concepts, they mention nothing about whether there is molecular evidence for differentiation among these kingfishers. Changes in nomencla- ture are compromised until they are properly documented in a peer-reviewed publication. In the drawings that show the external parts of birds on page 29, the bill is divided into the “upper mandible’’ (= culmen or rostrum) and “lower mandible’’ (= mandible). For decades, ornithologists have puzzled other vertebrate biologists by using the word “mandible” (= lower Jaw) in describing the upper part of the bill. Also on page 29, the tarsal Joint (ankle) is called the knee, which of cour.se is actually one Joint higher, where the femur Joins the tibiotarsus and fibula. Similarly, they incoiTectly use “ankle” for the Joint were the toes articulate with the metatarsals. In spite of my criticisms, nobody interested in birds should visit Palau without this fine book. I wish that it had been available during my two trips there, and I hope that it becomes widely used by residents and visitors to this fascinating island group. — DAVID W. STEADMAN, Florida Museum of Natural History, University of Florida, Gaines- ville, FL 32611, USA; e-mail: dws@flmnh.ufl. edu THE ECOLOGY AND CONSERVATION OF ASIAN HORNBILLS: FARMERS OF THE FOREST. By Margaret F. Kinnaird and Timothy G. O’Brien. University of Chicago Press, Chi- cago, Illinois, USA, and London, UK. 2007: 315 pages, 8 pages of color photos, numerous figures throughout. ISBN: 978-0-226-43712-5. $45.00 (cloth). — Hornbills are easily the most conspicu- ous of Asian forest birds. Their huge size and massive, sculptured bills project a striking image that is advertised by the hornbill’s loud, demand- ing voice and whooshing wing beats. Dedicated fruit-eaters, hornbills have a dominant role in the seed-dispersal of trees and other plants. Sadly, both their large size and food habits predispose them to extirpation by hunting and forest loss, and hornbills are among the most threatened of Asian birds. This book offers an impressive compilation of field research and literature review on the evolution, ecology, and con.servation of these wonderful birds. Hornbills, despite their boister- ous nature, are for the most part hard to get to know. Imagine the challenges of studying a bird that spends its life out of reach high in the forest canopy, that roams over vast areas of forest tracking hundreds of species of fruits, and that nests deep within the tree-hollow of a forest giant. A few teams of researchers have met the.se ORNITHOLOGICAL LITERATURE 413 challenges over the past three decades. Kinnaird and O’Brien foremost among them and, as a result, the mysteries of the complex life ot the rainforest hornbills are revealed in this book. The authors begin with two chapters that introduce the hornbills, their phylogeny and biogeography. I was surprised to learn the radiation of 3 1 species and nine genera of Asian hombills, primarily rainforest dwellers, is be- lieved to have been derived from savannah and woodland inhabiting African hornbills. Speciation within the geographically complex southeastern Asian land mass and myriad islands has created a hot-spot of hornbill diversity centered on the Malaysian Peninsula and the islands ot Sumatra, Borneo, and Java, all once connected. Several unique features of hornbills are also discussed. Naturally, this includes a section on “Casque form and function’’ that weighs the evidence for the horny casque’s role in social signaling (species recognition, sex, and age), structural support for the long bill, and as a battering ram in one amazing species, the Helmeted Hornbill (Rhinoplax vigil), the males of which engage in aerial jousting with head-on collisions! The next two chapters characterize the forest habitat and diet of these hornbills. Habitats range from highly seasonal deciduous subtropical for- ests to evergreen tropical forests with only the most subtle seasonal variation. An important climatic influence is the occurrence once every few years of an El Nino Southern Oscillation drought that causes the synchronized fruiting of many species of trees which, at other times, withhold bearing fruit. This boom-or-bust food availability affects which years hornbills nest. Although capable and eager to hunt small vertebrates and insects with that deadly bill, forest hornbills achieve at best a dismal success rate and instead must rely on fruits to fill their dietary requirements. Dietary studies document 497 species of fruits consumed, with oily fruits favored. With some success, the authors tackle the seemingly impossible task of characterizing the temporal and spatial availability of fruit and how that availability affects hornbill abundance, community structure, and reproduction. Once they’ve established an ecological context for forest hornbills, the authors delve into a range of topics on hornbill reproduction and social systems, specifically monogamy, territoriality, and cooperative breeding. It’s worth highlighting one topic. A behavior for which hornbills are famous is that the female seals herself into the nesting cavity during incubation and nestling care. The brick-hard seal is made from a mortar of regurgitated fruit, feces, and other materials delivered by her mate and pasted to the rim of cavity entrance. Apart from fortifying the nest against predators, the .seal could serve other functions, such as creating a more suitable microclimate within the nest. However, the authors argue for a social explanation, that the female is forcing the male to be the sole provider for the family. Much of the field research on hornbills has focused on the evolutionary and ecological function of hornbills in seed dispersal-their role as farmers of the forest, as the book’s title suggests. The authors exhaustively review this research and then attempt to model populations of hornbills, not just for the sake of the hornbills (minimal viable populations) but for the forest they serve (ecological functional populations). This topic leads into the final chapters on the threats to hornbill survival and the outlook for the future. And threats there are aplenty. Asian hornbills and the forests that support them are in a disastrous situation through large-scale defores- tation and hunting driven by a dense and growing human population. The rate of forest loss in places like Sumatra is mind-boggling and depressing to any sensibility for biodiversity and wilderness. The authors map out a full range of realistic options for the conservation of hornbills and their forests, including parks and protected area, grass- roots activities, and species management. 1 was impressed by how much published research there is on hornbills, given the difficulty in studying canopy birds in tropical forests and the modest effort worldwide devoted to this field. 1 was also impressed by the command that Kinnaird and O’Brien have of their subject material. Throughout the book, the authors take care to bring together available data on each subject and to explore it in depth. The text is laden with numerical summaries and statistical tests that greatly bolster their explanations. This quantita- tive approach is enhanced by numerous tables and figures. The data are nevertheless limited, and the authors over-reach in their inteipretation in places, such as dismissing for lack of evidence the role of predation in the evolution of nest- sealing (see review by M. Whitmer, 2008, Auk 125: 996-997). Kinnaird and O’Brien are certain- ly entitled to their views by the wealth of 414 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 2. June 2010 irifoiTnation assembled here, but it may take decades more research by other scientists before some of these ideas are sorted out. There are eight pages of color photographs of hornbills by Tim Laman. The text is written in an easily read narrative style, although it is information-packed and therefore rather slow going. I didn’t notice any errors. This book will be a classic on hornbill biology, and on the ecology and evolution of a group of fruit-eating and seed-dispersing birds. — THANE K. PRATT, P. O. Box 420, Volcano, HI 96785, USA; e-mail: thane-linda@earthlink.net THE SECOND ATLAS OF BREEDING BIRDS IN NEW YORK STATE. Edited by Kevin J. McGowan and Kimberley Corwin. Cornell University Press, Ithaca, New York, USA, 2008: 712 pages, 25 two-page color paintings, 245 halftone drawings, 164 charts and graphs, and 512 color maps. ISBN: 978-0-8014- 4716-7. $59.95 (cloth). — This important work is the result of the first statewide re-survey of the distribution of breeding birds. The first survey, The Atlas of Breeding Birds of New York State, edited by Robert F. Andrie and Janet R. CaiTolI (1988), summarized the survey data collected between 1980 and 1985. For the Second Atlas, 1,207 volunteers surveyed essentially the same 5,279 blocks from 2000 to 2005, exactly 20 years later, compiling over half a million records representing 248 species and three hybrids. The result is an invaluable record, not only of the current distribution of breeding bird populations, but also the change in distribution over the past two decades. Change is occurring at a surprising rate; the Second Atlas documents changes, both increa.ses and decrea.ses, experienced by more than half of the state's species. The book opens with six introductory chapters. In the first, Kimberley Corwin introduces the project and the methodology. As in the original atlas, the spatial units were based on lO-km X 10- km ( 100 knV) squares, each divided into four 25- knr survey “blocks.” Thus, the entire state was delineated into a total ol 5,333 blocks. The operational definitions of possible breeding, probable breeding, and confirmed breeding are given, and there is a discussion of how the data were processed, including limitations and possible biases. The chapter concludes with 15 color maps that illustrate in great detail characteristics of the state, such as cities and major roads, river systems, and selected lakes (e.g., useful as landmarks); climate data such as precipitation and temperature; and habitat categories based on land use, forest types, and ecoregions. The latter is particularly useful for interpreting the distribution maps presented in the individual species accounts. Chapter 2, by Kevin McGowan and Benjamin Zuckerberg, is a thorough comparison of the first and second surveys. Survey effort is quantified not only project wide, but also by region and by ecozone in relation to species richness. The heart of this chapter is the amount of change for each species, based on the number of blocks in which they were detected. This information is presented in two tables, the first aixanged taxonomically for ease of finding a particular species, and the second in rank order of change. Additional tables are limited to those species in which the changes were statistically significant. An analysis of change with respect to breeding habitat revealed that grassland species by far had the greatest decline in distribution. In Chapter 3, Gregory Endinger and Timothy Howard introduce the reader to the different habitats of the state, categorized by ecoregions, ecozones, bird conservation regions, and natural and cultural ecological communities. This chapter is followed by an examination of land-use changes, presented by Charles Smith and Peter Marks. Both chapters should be of interest to New York residents and nonresidents alike, and they help place the distribution changes into perspec- tive. Chapter 5, by John Peterson, is a discussion of professional and amateur ornithology in New York and, in Chapter 6, Kenneth Rosenberg and Michael Burger identify conservation priorities regarding the state’s breeding birds. The main part of the book is the collection of species accounts, contributed by 41 authors. The text in each account consists of a brief introduc- tion to the natural history of the species, its status from the first atlas, the changes that have occurred during the past 20 years, and conservation implications. There are two maps: one showing the 2()()()-2()()5 distribution, and the other the 20- year change. There also is a table that summarizes the number of blocks for which breeding was possible, probable, and confirmed in 1980-85 and 2000-05, and a graph that shows the USGS Breeding Bird Survey trend in New York from 1965 through 2005. The layout is excellent. Each species account ORNITHOLOGICAL LITERATURE 415 covers two pages with the text and an attractive halftone portrait of the species on the left, and the maps, table, and graph on the right. Thus, the reader is presented each entire account without having to turn additional pages. This book is a remarkable accomplishment. It belongs on the shelf of anyone interested in the changing distribution of birds in the northeastern United States. However, because the volume is so beautifully done, especially the 25 color paintings of New York habitats, it will be equally at home on the coffee table. Finally, at a time when college textbooks command over $100 lor the paperback editions, the $59.95 list price for The Second Atlas represents a lot of value. — JOHN A. SMALL- WOOD, Montclair State University, Mont- clair, NJ 07043, USA; e-mail: smallwoodj@ montclair.edu Presidents of The Wilson Ornithological Society J. B. Richards 1888-1889 Lawrence H. Walkinshaw 1958-1960 Lynds Jones 1890-1893 Harold E. Mayfield 1960-1962 Willard N. Clute 1894 Phillips B. Street 1962-1964 Reuben M. Strong 1894-1901 Roger Tory Peterson 1964-1966 Lynds Jones 1902-1908 Aaron M. Bagg 1966-1968 Frank L. Burns 1909-1911 H. Lewis Batts Jr. 1968-1969 W. E. Saunders 1912-1913 William W. H. Gunn 1969-1971 T. C. Stephens 1914-1916 Pershing B. Hofslund 1971-1973 W. F. Henninger 1917 Kenneth C. Parkes 1973-1975 Myron H. Swenk 1918-1919 Andrew J. Berger 1975-1977 Reuben M. Strong 1920-1921 Douglas A. James 1977-1979 Thomas L. Hankinson 1922-1923 George A. Hall 1979-1981 Albert F. Ganier 1924-1926 Abbot S. Gaunt 1981-1983 Lynds Jones 1927-1929 Jerome A. Jackson 1983-1985 J. W. Stack 1930-1931 Clait E. Braun 1985-1987 J. M. Shaver 1932-1934 Mary H. Clench 1987-1989 Josselyn Van Tyne 1935-1937 Jon C. Barlow 1989-1991 Margaret Morse Nice 1938-1939 Richard C. Banks 1991-1993 Lawrence E. Hicks 1940-1941 Richard N. Conner 1993-1995 George Miksch Sutton 1942-1943 Keith L. Bildstein 1995-1997 S. Charles Kendeigh 1943-1945 Edward H. Burtt Jr. 1997-1999 George Miksch Sutton 1946-1947 John C. Kricher 1999-2001 Olin Sewall Pettingill 1948-1950 William E. Davis Jr. 2001-2003 Maurice Brooks 1950-1952 Charles R. Blem 2003-2005 Walter J. Breckenridge 1952-1954 Doris J. Watt 2005-2007 Burt L. Monroe Sr. 1954-1956 James D. Rising 2007-2009 John T. Emlen Jr. 1956-1958 E. Dale Kennedy 2009- THE WILSON JOURNAL OL ORNITHOLOGY Editor CLAIT E. BRAUN Editorial Board RICHARD C. BANKS 5572 North Ventana Vista Road Tucson, AZ 85750-7204 E-mail: TWILSONJO@comcast.net JACK CLINTON EITNIEAR SARA J. OYLER-McCANCE LESLIE A. ROBB Editorial NANCY J. K. BRAUN Assistant Review Editor ROBERT B. PAYNE 1306 Granger Avenue Ann Arbor, MI 48104, USA E-mail: rbpayne@umich.edu GUIDELINES FOR AUTHORS Please consult the detailed “Guidelines for Authors’’ found on the Wilson Ornithological Society web site (http://www.wiLsonsociety.org). All manuscript submissions and revisions should be sent to Clait E. Braun, Editor, The Wilson Journal of Ornithology, 5572 North Ventana Vista Road, Tucson, AZ 85750-7204, USA. The Wilson Journal of Ornithology office and fax telephone number is (520) 529-0365. The e-mail address is TWilsonJO@comcast.net NOTICE OF CHANGE OF ADDRESS Notify the Society immediately if your address changes. Send your complete new address to Ornithological Societies of North America, 5400 Bosque Boulevard, Suite 680, Waco, TX 76710. The permanent mailing address of the Wilson Ornithological Society is: %The Museum of Zoology, The University of Michigan, Ann Arbor, MI 48109, USA. Persons having business with any of the officers may address them at their various addresses given on the inside of the front cover, and all matters pertaining to the journal should be sent directly to the Editor. MEMBERSHIP INQUIRIES Membership inquiries should be sent to Timothy J. O Connell, Department of Natural Resources Ecology and Management, Oklahoma State University, 240 AG Hall, Stillwater, OK 74078; e-mail: oconnet@ okstate.edu 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 ornithological magazines from members and friends of the Society. Two members have generously established a fund for the purchase of new books; members and friends are invited to maintain the fund by regular contribution. The fund is administered hy the Library Committee. Jerome A. Jackson, Florida Gulf Coast Univeristy, is Chairman of the Committee. The Library currently receives over 200 periodicals as gifts and in exchange for The Wilson Journal of Ornithology. For information on the Library and our holdings, see the Society’s web page at http;// www.wilsonsociety.org. 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 ol Michigan, 1 109 Geddes Avenue, Ann Arboi, Ml 48109-1079, USA. Contributions to the New Book Fund should be sent to the Treasurer. This issue of The Wilson Journal of Ornithology was published on 18 May 2010. Continued from outside back cover 346 Assessing generalized egg mimicry: a quantitative comparison of eggs of Brown-headed Cowbirds and grassland passerines Dwight R. KUppenstirie and Spencer G. Sealy 354 Flight feather molt of Turkey Vultures Robert M. Chandler, Peter Pyle, Maureen E. Flannery, Douglas J. Long, and Steve N. G. Howell Short Communications 361 The first reported case of cooperative polyandry in the Red-footed Booby: trio relationships and benefits Lei Gao, Guixia Zhao, Shan Tang, and Hanzhao Guo 365 Further evidence of breeding by Shiny Cowbirds in North America Matthew J. Reetz, Jacob M. Musser, and Andrew W. Kratter 369 Circumstantial evidence for infanticide of chicks of the communal Smooth-billed Ani {Crotophaga ant) J. S. Quinn, A. Samuelsen, M. Barclay, G. Schmaltz, and H. Kahn 374 Flight speeds of migrating Red-necked and Fiorned grebes Laurence C. Binford and Joseph A. Youngman 378 Thermoregulatory behavior in migratory European Bee-eaters {Merops apiaster) Reuven Yosef 381 Population status of Chuck-will’s-widow {Caprimulgus carolinensis) in the Bahamas William K. Hayes, ElwoodD. Bracey, Melissa R. Price, Valerie Robinette, Eric Gren, and Caroline Stahala 385 Yellow Rails wintering in Oklahoma Christopher J. Butler, Lisa H. Pham, Jill N. Stinedurf, Christopher L. Roy, Eric L. Judd, Nathanael J. Burgess, and Gloria M. Caddell 388 Breeding of the Giant Laughingthrush {Garrulax maximus) at Lianhuashan, southern Gansu, Chint Jie Wang, Chen-Xi Jia, Song-Hua Tang, Yun Fang, and Yue-Hua Sun 392 First description of the nest of Jocotoco Antpitta {Grallaria ridgelyi) Harold E Greeney and Mery E. Juifia J 395 Nesting ecology of the Grey-backed Shrike {Lanius tephronotus) in South Tibet Xin Lu, Chen Wang, and Tonglei Yu 399 The first reported hybridization of Abert’s and Canyon towhees {Pipilo spp.) R. Roy Johnson and Steven L. Hopp 402 Yellow-billed Cuckoo hatched and fed by a Red-winged Blackbird Ken Yasukawa 406 Ornithological Literature Robert B. Payne, Book Review Editor The Wilson Journal of Ornithology (formerly The Wilson Bulletin) Volume 122, Number 2 CONTENTS June 2010 Major Articles 207 Not the Nice sparrow (The 2007 Margaret Morse Nice Lecture) ' Douglas W. Mock and P. L. Schwagmeyer 217 The postbreeding migration of Eared Grebes Joseph R. Jehl Jr. and Annette E. Henry 228 Postfledging and natal dispersal of Crested Ibis in the Qinling Mountains, China Yu Xiao-Ping, Xi Yong-Mei, Lu Bao-Zhong, LiXia, Gong Ming-Hao, Shi Liang, and Dong Rong 236 Morphological and genetic variation between migratory and non-migratory Tropical Kingbirds during spring migration in central South America Alex E. Jahn, Douglas J. Levey, Izeni Pires Parias, Ana Maria Mamani, Julian Quillen Vidoz, and Ben Ereeman 244 Red-cockaded Woodpecker male/female foraging differences in young forest stands Kathleen E. Eranzreb 255_ Foraging habits and habitat use by Edible-nest and Glossy swiftlets in the Andaman Islands, India Shirish S. Manchi and Ravi Sankaran 273 Golden- and Blue-winged warblers: distribution, nesting success, and genetic differences in two habitats John L. Confer, Kevin W. Barnes, and Erin C. Alvey 279 Genetic and morphological variation of the Sooty-capped Bush Tanager {Chlorospingus pileatus), a highland endemic species from Costa Rica and western Panama Tania Chavarria-Pizarro, Gustavo Gutierrez-Espeleta, Eric J. Euchs, and Gilbert Barrantes 288 Long-term changes in avian community structure in a successional, forested, and managed plot in a reforesting landscape Elizabeth W. Brooks and David N. Bonter 296 Scale-dependent response by breeding songbirds to residential development along Lake Superior Michelle T. Eord and David J. Elaspohler 307 Songbird nest survival is invariant to early-successional restoration treatments in a large river floodplain Dirk E. Burhans, Brian G. Root, Terry L. Shajfer, and Daniel C. Dey 318 Carotenoid-based male plumage predicts parental investment in the American Redstart Ryan R. Germain, Matthew W Reudink, Peter P Marra, and Laurene M. Ratclijfe 326 Variation in plumage coloration of Northern Cardinals in urbanizing landscapes Todd M. Jones, Amanda D. Rodewald, and Daniel P Shustack 334 Annual precipitation affects reproduction of the Southern Grey Shrike {Lanius meridionalis) Oded Keynan and Reuven Yosef 340 Home range sizes and habitat use of Nelson’s and Saltmarsh sparrows W. Gregory Shriver, Thomas P. Hodgman, James P Gibbs, and Peter D. Vickery Continued on inside back cover IU\U 0.10 for all comparisons; Fig. 2). The earliest Shiny Cowbird parasitism event was recorded on 29 September in a Chalk-browed Mockingbird nest and the latest, on 23 January in a Baywing nest. Even though Shiny Cowbirds started laying earlier than Baywings, they over- lapped considerably in breeding seasons (Fig. 3). Shiny Cowbird parasitism mostly (44%) occuiTed in December and January. Nearly 90% of all Baywing clutches were initiated during the same interval. Only 10% of Bay wing nesting attempts began too late to be parasitized by Shiny Cowbirds. Clutch Size, Incubation, and Nestling Periods. — Baywings laid, on average, 4.0 ± 0.1 eggs per nest (range = 2-5, mode = 4, « = 32 nests), and host clutch size did not vary with clutch initiation date (Spearman rank correlation: r = 0.22, Z = 1.23, P = 0.23, n = 32 nests). Initial clutch size in nests parasitized solely by Screaming Cowbirds was 5.3 ± 0.2 eggs, and final clutch size was 6.4 ± 0.2 eggs (n = 49 nests). Only two nests naturally parasitized by Shiny Cowbirds survived until hatching, and they were also parasitized by Screaming Cowbirds. Initial clutch size in these nests was 7.0 ± 1.0 eggs, and final clutch size was 8.0 ± 1.0 eggs (Table 1). Host incubation period was 13.0 ± 0.1 days ■^T-ooincNia)“2f^®^cNa) (MOOt-cNCNjO'^'^C^OO FIG. 1. Breeding sea.son of (A) Baywings and (B) Screaming Cowbirds in Reserve El Destino between 2002 and 2006. Bars represent the number of nesting attempts initiated (;? = 160) or the total number of Screaming Cowbird eggs laid (n = 686) during the breeding season divided into weekly intervals. with a range of 12-14 days and a mode of 13 (n = 42 clutches). Screaming and Shiny cowbirds had an incubation period of 12.0 ± 0.1 days (range = 1 1- 13 days, n = 40 and 14 eggs of Screaming and Shiny cowbirds, respectively). Average number of Baywing chicks at hatching was 3.2 ± 0.1 (range = 0—5, n = 60). Nests parasitized by Screaming Cowbirds only, had 1 .3 ± 0.2 parasite chicks at hatching (n = 37). Nests naturally parasitized by Shiny Cowbirds (n = 2) had one Shiny Cowbird and one Screaming Cowbird chick each. Combin- ing host and parasitic chicks, Baywings reared, on average, 4.2 ± 0.2 chicks per nest (range = 1-8, n = 60 nests; Table 1). The mean nestling period was 14 days (range = 12-16; n = 20 nests). 424 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122. No. 3. September 2010 2003 2004 (U > 2005 -5-3^ > > > 0 0 0 0 0 c c C C 0) 0 0 0 0 0 0) 0) ~3 “3 G> CD CO 0 CO 0 1^ 00 in CM 03 ID CM C33 CD CNJ 0 CN CM 0 CM 0 0 ■<— CM CM 0 CM FIG. 3. Seasonal distribution of Shiny Cowbird parasitism in Reserve El De.stino during the 2002-2006 breeding seasons. Bars represent the number of parasite eggs laid within each weekly interval in ne.sts of Baywings, Chalk-browed Mockingbirds, and House Wrens (n = 534). The dashed line indicates clutch initiation date of the earliest Baywing nesting attempt during this study (27 Nov). De Marsico cl al. • NESTING ECOLOGY OE' BAYWINCjS 425 TABLE 2. Length, width, and volume (.v ± SE) of Baywing, and Screaming and Shiny cowbird eggs. Values coiTesponding to host eggs represent average egg size and volume of all eggs in 98 clutches. Some Bay wing (Bw) nests in our sample were artificially parasitized with Shiny Cowbird eggs collected from Chalk-browed Mockingbird (CbM) and House Wren (HW) nests, and we present data on size and volume of Shiny Cowbird eggs in the three host species. Data from Chalk-browed Mockingbird and House Wren nests were provided by V. D. Fiorini and D. T. Tuero (unpubl. data). Specie.s Host n Length (cm) Width (cm) Volume (cm') Baywing 346 2.36 ± 0.01 1.78 ± 0.00 3.87 ± 0.02 Screaming Cowbird Bw 357 2.37 ± 0.01 1 .79 ± 0.00 3.94 ± 0.02 Shiny Cowbird Bw 1 1 2.30 ± 0.03 1.86 ± 0.00 4.11 ± 0.16 CbM 61 2.43 ± 0.02 1.89 ± 0.01 4.49 ± 0.05 HW 29 2.34 ± 0.00 1.86 ± 0.00 4.20 ±0.10 1.6, P = 0.20). Differences in egg volume were not attributable to larger Shiny Cowbird eggs coming from Chalk-browed Mockingbird nests because the results did not change when these eggs were removed from the data set (ANOVA; P2.491 = 8.9, P < 0.001). We found differences in growth parameters between host and parasite chicks, as well as between Screaming and Shiny cowbird chicks (Table 3, Fig. 4). Overall, both Screaming and Shiny cowbirds had higher growth constant (ANOVA: p2,45 = 6.5, P = 0.003), maximum instantaneous growth rate (ANOVA: P2.45 — 45.8, P < 0.001 ), and asymptotic weight (F2,45 — 37.3, P < 0.001) than Bay wing nestlings (Tukey post- hoc test, P < 0.05). Shiny Cowbirds had a higher weight at hatching (ANOVA: F2A5 — 6.7, P = 0.003) and lower age of maximum growth (^2,45 = 8.8, P < 0.001; Tukey post-hoc test, P < 0.001) than both Screaming Cowbird and host nestlings. Sex-specific growth curves for Screaming and Shiny cowbirds varied (Table 3, Fig. 4). Male Screaming Cowbird nestlings (77 = 4) did not differ from females (/? = 6) in weight at hatching (/-test: / = — 1 . 15, df = 8, P = 0.28), growth constant (/-test: / = —0.09, df = 8, P = 0.93), and age of maximum growth (/-test: / = —0.46, df = 8, P = 0.66; Table 3). However, male nestlings had a higher maximum instantaneous growth rate (/-test: / = —3.84, df = 8, P = 0.005) and reached a higher asymptotic weight than females (/-test: / = -5.53, df = 8, P < 0.001; Table 3). Sample sizes for Shiny Cowbirds were too low for statistical analyses, but growth curves suggest that differences in males and females similar to those of Screaming Cowbirds might occur in this species (Table 3, Fig. 4). Baywing Reproductive Success. — Only 35 (19%) of all nests fledged chicks and all but one produced host fledglings. The exception was a parasitized nest where all host eggs failed to hatch and only two Screaming Cowbird chicks fledged. The number of Baywing chicks fledged per nest was 3.0 ± 0.2 (77 = 34 nests, range = 1-5). Overall, host productivity in nests that survived the entire nesting cycle was 0.78 fledglings per egg laid (77 = 25 nests with known clutch size). Losses were mostly due to egg punctures by Screaming and Shiny cowbirds and hatching failures. On average, survival of host eggs was 0.92 (77 = 54 nests). Egg survival was higher in TABLE 3. Growth parameters (x ± .SE) of Bay wing, and Screaming and Shiny cowbird nestlings obtained from adjusted growth curves. Data for host nestlings were averaged over each brood. Data for parasite nestlings were based on individual nestlings of each gender. Sample sizes = 69 Baywings in 25 broods, 6 female and 4 male Screaming Cowbirds, and 5 female and 2 male Shiny Cowbirds. Screaming Shiny Parameter Baywing Male Female Male Female Hatch weight, g 3.6 ± 0.1 3.8 ± 0.4 3.3 ± 0.2 4.9 ± 0.3 4.4 ± 0.2 Growth constant/day 0.47 ± 0.01 0.50 ± 0.02 0.49 ± 0.02 0.5 1 ± 0.03 0.5 1 ± 0.03 Maximum growth rate, g/day 4.25 ± 0.07 6.37 ± 0.14 5.23 ± 0.22 6.10 ± 0.14 5.31 ± 0.32 Age of maximum growth, days 5.8 ± 0.1 6.2 ± 0.2 6.0 ± 0.2 5.3 ± 0.2 5.2 ± 0.1 Asymptotic weight, g 36.7 ± 0.4 51.6 ± 1.9 42.5 ± 0.6 48.3 ± 1.8 41.4 ± 1.0 426 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. 3, September 2010 Screaming Cowbird Shiny Cowbird O) sz D) 0 $ cn c C/) 0 Age (days) FIG. 4. Growth curves of Screaming and Shiny cowbird nestlings reared by Baywings. Points correspond to daily weights (x ± SE) as a function of age for male (black circles) and female (white circles) nestlings of each species, estimated from the corresponding adjusted logistic functions. The dotted line in each graph represents the growth curve of the host with adjusted daily weights (.v ± SE) averaged across all host nestlings within the same brood (/? = 25 broods with 1-5 host nestlings) Sample sizes were four and six female Screaming Cowbirds, and two and five female Shiny Cowbirds, respectively. nests that were not parasitized during the egg stage than in parasitized nests (Mann-Whitney (/-test: Z = -2.03, P = 0.042, n = 10 unparasitized and 44 parasitized nests). On average, hatching success was 0.88 {n = 54 nests). Hatching success did not differ between parasitized and unparasitized nests (Mann-Whit- ney (/-test: Z = 0.00, P > 0.99). There were no statistical differences in survival of Baywing chicks between parasitized {n = 20) and unpar- asitized nests (n = 8; Mann-Whitney (/-test: Z = — 1.52, P = 0.13). Nestling survival was, on average, 0.94 {n = 28 nests). Brood reduction occurred in six parasitized nests where one (n = 5) or two (/? = 1) host nestlings died due to starvation or nest crowding. Screaming Cowbird Reproductive Success. — About 89% (615/694) of the Screaming Cowbird eggs in Baywing nests failed as a result of nest predation or desertion (/? = 146 naturally parasitized nests found before or during host laying). Survival of Screaming Cowbird eggs in nests that survived until hatching was 0.93 {n = 48 nests). Hatching success was 0.62 (/? = 47 nests), and hatching failures were mostly due to parasitic females laying eggs too late for success- ful incubation. Hatching success of Screaming Cowbird eggs laid during host laying was 0.92 (// = 40 nests). Overall nestling survival was 0.77 (/? = 20 nests). Mortality of parasitic chicks occun'ed at six nests as a result of late hatching (/? = 4 nests) and partial nest predation (n = 2 nests). Screaming Cowbird productivity was 0.14 fledg- lings per egg laid (n = 20 nests). Shiny Cowbird Reproductive Success. — Nearly 91% of Shiny Cowbird eggs naturally laid (30/33, n = 26 nests) and 82% of those artificially placed in Baywing nests (40/49, n = 49 nests) failed due to nest predation or desertion. Combining natural and artificial parasitism. Shiny Cowbird egg survival was 0.95 (/? = 19 nests) and hatching success was 0.94 (/? = 18 nests). Nestling survival was 0.91 (n = 11 nests) and only one Shiny Cowbird chick died, presumably due to partial nest predation. Sample size was too small to estimate productivity. Only one of three Shiny Cowbird eggs naturally laid produced a fledgling {n — 3 nests). The others were laid before host laying and therefore rejected by hosts. DISCUSSION Nest Site Use. — Baywings used a wide variety of nesting sites, but the more commonly used were old, clo.sed nests built by several furnarid species. This ob.servation contrasts with a previ- ous study in a nearby area, which reported that De Mursico el al. • NESTING ECOLOGY Ol- FiAYWINGS 427 nest boxes and secondary cavities were preferred nesting sites (Fraga 1988, 1998). These differenc- es, however, could be explained by local variation in abundance of suitable nesting sites at the time Bay wings start to breed (Hoy and Ottow 1964). Some authors have observed Baywings fighting with owners to occupy their nests (Friedmann 1929, Jaramillo and Burke 1999), but we seldom observed nest piracy, suggesting that nesting sites were not limiting during our study (see also Hoy and Ottow 1964, Fraga 1988). It is not clear why in most cases Baywings use old nests of other species instead of building their own nests, but it is possible this behavior allows breeding pairs to save time and energy (Hauber 2002, Wiebe et al. 2007). Additionally, closed nests might offer better protection against predators than open cup nests (Martin and Li 1992, Auer et al. 2007). Breeding Seasons. — Our results indicate the breeding season of Baywings typically starts in early December and may extend to early March, considering nestling and postfledgling stages of late nesting attempts. Fraga (1998), in the Province of Buenos Aires, observed laying as early as October, but we did not observe any sign of nesting activity before late November. The breeding season of Screaming Cowbirds over- lapped extensively that of Baywings. However, female Screaming Cowbirds started parasitism in advance of host laying indicating that nest building and defense activities by the host may stimulate laying behavior of the parasite. It is also possible the parasite’s breeding season began earlier if Screaming Cowbirds in our study area also parasitize Brown-and-yellow Marshbirds, which breed from October to December (Mermoz and Reboreda 1996, Mermoz and Fernandez 2003). Shiny Cowbirds began to breed much earlier in our study area than Baywings, parasit- izing other hosts such as Chalk-browed Mocking- birds and House Wrens starting in late September (Fiorini and Reboreda 2006, Tuero et al. 2007). Friedmann (1929) suggested differences in time of breeding would be the cause of the low frequency of Shiny Cowbird parasitism in nests of Baywings. However, our data indicate only 10% of the nesting attempts of Baywings began after the Shiny Cowbird breeding season ended. Breeding seasons of Baywings and Shiny Cow- birds overlapped more extensively than previously estimated (90 vs. 65-80%; Fraga 1998), and female Shiny Cowbirds would have broad oppor- tunities for parasitizing this host. An alternative explanation for the low frequency of Shiny Cowbird parasitism of this host is that the parasite would have low reproductive success with Bay- wings (Fraga 1998), but this hypothesis remains untested. Incidence of Cowbird Parasitism. — Nearly all Baywing nests were parasitized by Screaming Cowbirds, and regularly more than once (Hoy and Ottow 1964, Mason 1980, Fraga 1998). These high levels of parasitism did not seem to be an artifact of nest boxes (Rattan 1997), as the frequency and intensity of parasitism by Scream- ing and Shiny cowbirds did not differ between nest boxes and natural sites. Our results did not appear to be influenced by the shortage of host nests in a particular year, as frequency and intensity of parasitism were consistent among consecutive breeding seasons. Females of other brood parasites parasitize nests within nearly exclusive breeding areas, which may allow them to reduce the costs associated with intraspecific competition (Langmore et al. 2007). Several observational and genetic studies indicate that cowbirds, however, lack territorial behavior (Rattan 1997, Hahn et al. 1999, Mermoz and Reboreda 1999, McLaren et al. 2003, Strausberger and Ashley 2005, Ellison et al. 2006). Thus, multiple parasitism can arise as a consequence of a limitation of suitable nests, high fecundity rates of female cowbirds, or high population density of the parasite relative to the number of host breeding pairs (Martinez et al. 1998, Strausberger 1998, McLaren et al. 2003, Ellison et al. 2006). Multiple parasitism by Screaming Cowbirds, which are constrained to locate and parasitize near a single host species, could also be the result of different females parasitizing the same nest and individual females parasitizing a nest more than once as a result of low nest availability. Baywing Reproductive Success. — Cowbird par- asitism reduces host fitness in several ways (Massoni and Reboreda 1998, Lorenzana and Sealy 1999, Zanette et al. 2005, Astie and Reboreda 2006). We found parasitized nests lost more host eggs than unparasitized nests, but hatching success and nestling survival did not differ between parasitized and unparasitized nests. Our results indicate survival of Baywing eggs and nestlings was higher than in previous reports (Eraga 1986, 1998), possibly because, during our study, most host egg and nestlings were lost due to nest desertion and predation rather than to egg punctures or competition with parasite nestlings. 428 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 3. September 2010 We rarely observed brood reduction in parasitized nests and nestling mortality due to parasitic mites, as reported previously (Fraga 1986). Regular cleaning of nest boxes may have contributed to decrease the incidence of ectoparasite infestations in our study population. However, given that most nests occurred in natural sites, we have confidence that box cleaning had little influence on our estimates of nestling survival. Costs of parasitism for Baywings are similar to those observed in other host species of the Shiny Cowbird that are larger than the parasite (Mermoz and Reboreda 2003, Astie and Reboreda 2006). Productivity is often lowered in hosts larger than the parasite by puncture or removal of host eggs by female cowbirds rather than by nestling mortality, as host young are seldom outcompeted by parasitic nestlings (Eckerle and Breitwisch 1997, Lichten- stein 1998, Lorenzana and Sealy 1999, Astie and Reboreda 2006), unless a marked hatching asynchrony exists (Soler et al. 1996, Dure Ruiz et al. 2008). Baywings are smaller than Screaming and Shiny cowbirds; thus, it is expected that host nestlings compete strongly for food with the parasite young. However, Baywings are cooper- ative breeders (Fraga 1991) and it is possible that helpers at the nest contribute to ameliorate the costs of parasitism by increasing nest provisioning rates (Fraga 1991). Nestling Growth. — We did not find differences in growth patterns between Screaming and Shiny cowbird nestlings in nests of Baywings. Overall, Screaming and Shiny cowbird nestlings did not differ in asymptotic weight, growth constant, and maximum instantaneous growth rate. This is consistent with similar comparisons between parasite nestlings of both species reared in Brown-and-yellow Marshbird nests (Mermoz and Fernandez 2003). However, parasite nestlings differed in weight at hatching and age of maximum growth, which could be the result of differences in egg size between both species (Nolan and Thompson 1978). A higher weight at hatching combined with a shorter incubation period could provide Shiny Cowbird nestlings a head start to overcome competition, which can be particularly important in nests of host species larger than the parasite (Astie and Reboreda 2006, Fiorini et al. 2009). However, evidence for a positive effect of egg size on nestling performance is equivocal (e.g., Schifferli 1973, Magrath 1992, Reed et al. 1999). Screaming Cowbird eggs and hatchlings were similar in mass to those of Baywings, although parasitic juveniles and adults are quite larger than the host (Mason 1987, Fraga 1998). Females in other parasite species have evolved eggs similar in size to those of the host, presumably in response to host discrimination against larger or smaller eggs (Mason and Rothstein 1986, Marchetti 2000). No experiments have been conducted on Baywings to test whether they are able to discriminate eggs by size. However, that they readily accepted larger Shiny Cowbird eggs artificially placed in their nests (this study) suggests Baywings do not discriminate foreign eggs on the basis of size. Screaming Cowbird eggs in Baywing nests usually have low chances of success as a consequence of poor synchronization of parasitism and high nest failure rates (Hoy and Ottow 1964, Fraga 1998, De Marsico and Reboreda 2008). Thus, a selective pressure may exist on female Screaming Cowbirds to increase the total number of eggs laid at the expense of laying smaller eggs (Fontaine and Martin 2006, Martin et al. 2006). Screaming Cowbird nestlings were sexually dimorphic in size and presented sex-specific growth curves. Males grew faster and were larger than females over the entire nestling period, consistent with a recent study of Brown-headed Cowbirds (Molothrus oter) (Tonra et al. 2008, but see Weatherhead 1989). Our sample sizes were too small to compare growth parameters between male and female Shiny Cowbirds, but our data indicate that sex-specific growth patterns similar to that observed in Screaming Cowbirds may occur in this species. The occun’ence of sex- specific growth curves among parasitic cowbirds implies that gender should be considered when analyzing the costs of competition and factors affecting nestling growth in host and parasite species. ACKNOWLEDGMENTS We thank Fundacion Elsa Shaw de Pearson for allowing us to conduct this study at Reserve El Destino. D. T. Tuero and V. D. Fiorini generously provided data for laying dates and egg size of Shiny Cowbirds. We are grateful to reviewers Rosendo Fraga and Peter Lowther for comments and suggestions that greatly helped to improve the manuscript. MCDM was supported by a fellowship from the Con.sejo Nacional de Investigaciones Cientificas y Tecnicas (CONICET). BM and JCR are re.search fellows of CONICET. This work was supported by research grants of Agencia Nacional de Promocion Cienti'fica y Tecnologica. Dc Marsico cl al. • NESTING ECOLOGY OE BAYWINGS 429 University of Buenos Aires, and the Student Grant Program of the Neotropieal Grassland Conservaney. LITERATURE CITED Astie, a. a. and J. C. Reboreda. 2006. Costs of egg punctures and parasitism by Shiny Cowbirds (Molo- thnis bonariensis) at Creamy-bellied Thrush (Tiirdits amaurochalinus) nests. Auk 123:23-32. Auer, S. K., R. D. Bassar, J. J. Fontaine, and T. E. Martin. 2007. Breeding biology of passerines in a subtropical montane forest in northwestern Argentina. Condor 109:321-333. Briskie, j. V. and S. G. Sealy. 1990. Evolution of short incubation periods in the parasitic cowbirds, Molo- rhnis spp. Auk 107:789-793. C.agnoni, M., a. M. Faggi, and A. Ribichich. 1996. La vegetacion de la Reserva El Destino (Partido de Magdalena, Provincia de Buenos Aires). Parodiana 9:25-44. Davies, N. B. 2000. Cuckoos, cowbirds and other cheats. T. & A. D. Poyser, London, United Kingdom. De MArsico, M. C. and J. C. Reboreda. 2008. Egg-laying behavior in Screaming Cowbirds: why does a specialist brood parasite waste so many eggs? Condor 110:143-153. Dure Ruiz, N. M., M. E. Mermoz, and G. J. Fernandez. 2008. Effect of cowbird parasitism on brood reduction in the Brown-and-yellow Marshbird. Condor 1 10:507- 513. Eckerle, K. P. and R. Breitwisch. 1997. Reproductive success of the Northern Cardinal, a large host of Brown-headed Cowbirds. Condor 99:169-178. Ellegren, H. 1996. First gene on the avian W chromosome (CHD) provides a tag for universal sexing of non-ratite birds. Proceedings of the Royal Society of London, Series B 263:1635-1641. Ellison, K., S. G. Sealy, and H. L. Gibbs. 2006. Genetic elucidation of host use by individual sympatric Bronzed Cowbirds (Molothrus aeneiis) and Brown- headed Cowbirds (M. ater). Canadian Journal of Zoology 84:1269-1280. Fiorinl V. D. and j. C. Reboreda. 2006. Cues used by Shiny Cowbirds {Molothrus bonariensis) to locate and parasitize Chalk-browed Mockingbird (Miinus satitr- niniis) nests. Behavioral Ecology and Sociobiology 60:379-385. Fiorini, V. D., D. T. Tuero, and J. C. Reboreda. 2009. Shiny Cowbirds synchronize parasitism with host laying and puncture host eggs according to host characteristics. Animal Behaviour 77:561-568. Fontaine, J. J. and T. E. Martin. 2006. Parent birds assess nest predation risk and adjust their reproductive strategies. Ecology Letters 9:428^34. Fraga, R. M. 1979. Differences between nestlings and fledglings of Screaming and Bay-winged cowbirds. Wilson Bulletin 91:151-154. Fraga, R. M. 1983. The eggs of the parasitic Screaming Cowbird {Molothrus rufoaxillaris) and its host, the Bay-winged Cowbird {M. badius): is there evidence for mimicry? Journal fur Ornitologie 124:187-193. Fraga, R. M, 1986. The Bay-winged Cowbird {Molothrus badius) and its brood parasites: interactions, coevolu- tion and comparative efficiency. Thesis. University of California, Santa Barbara, USA. Fraga, R. M. 1988. Nest sites and breeding success of Bay- winged Cowbirds {Molothrus badius). Journal fiir Ornithologie 129:175-183. Fraga, R. M. 1991. The social system of a communal breeder, the Bay-winged Cowbird Molothrus badius. Ethology 89:195-210. Fraga, R. M. 1996. Further evidence of parasitism of Chopi Blackbirds {Gnorinwpsar chopi) by the spe- cialized Screaming Cowbird {Molothrus rufoaxillaris). Condor 98:866-867. Fraga, R. M. 1998. Interactions of the parasitic Screaming and Shiny cowbirds {Molothrus rufoaxillaris and M. bonariensi.s) with a shared host, the Bay-winged Cowbird {Molothrus badius). Pages 173-193 in Parasitic cowbirds and their hosts. Studies in coevo- lution (S. I. Rothstein and S. K. Robinson. Editors). Oxford University Press, New York, USA. Friedmann, H. 1929. The cowbirds. A study in the biology of social parasitism. C. C. Thomas, Springfield, Illinois, USA. GenStat. 2007. GenStat Discovery Edition 3. Lawes Agricultural Trust, Rothamsted Experimental Station, Hertfordshire, United Kingdom. Griffiths, R., M. C. Double, K. Orr, and R. J. G. Dawson. 1998. A DNA test to sex most birds. Molecular Ecology 7:1071-1075. Hahn, D. C., J. A. Sedgwick, I. S. Painter, and N. J. Casna. 1999. A spatial and genetic analysis of cowbird host selection. Studies in Avian Biology 18:204-217. Hauber, M. E. 2002. Is reduced clutch size a cost of parental care in Eastern Phoebes {Sayornis phoebe)! Behavioral Ecology and Sociobiology 5 1 :503-509. Hoover, J. P. 2003. Multiple effects of brood parasitism reduce the reproductive success of Prothonotary Warblers. Protonotaria citrea. Animal Behaviour 65:923-934. Hoy, G. and J. Ottow. 1964. Biological and oological studies of the molothrine cowbirds (Icteridae) of Argentina. Auk 81:186-203. Hudson, W. H. 1874. Notes on the procreant instincts of the three species of Molothrus found in Buenos Aires. Proceedings of the Zoological Society of London 1 1:153-174. .lARAMlLLO, J. and P. Burke. 1999. New World blackbirds: the icterids. Princeton University Press. Princeton. New Jersey. USA. Kattan, G. H. 1997. Shiny Cowbirds follow the 'shotgun' strategy of brood parasitism. Animal Behaviour 53:647-654. Langmore. N. E., G. j. Adcock, and R. M. Kilner. 2007. The spatial organization and mating system of the Horsfield's Bronzed-cuckoos Chalcites basalis. Ani- mal Behaviour 74:403-412. Lanyon, S. M. 1992. Interspecific brood parasitism in blackbirds (Icterinae): a phylogenetic perspective. Science 255:77-79. 430 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. 3. September 2010 Lichtenstein. G. 1998. Parasitism by Shiny Cowbirds of Rufous-bellied Thrushes. Condor 100:680-687. Lorenzana, J. C. and S. G. Sealy. 1999. A meta-analysis of the impact of parasitism by Brown-headed Cow- birds on its hosts. Studies in Avian Biology 18:241- 253. Lowther. P. E. and W. Post. 1999. Shiny Cowbird (Molothriis bonariensis).'Yhe birds of North America. Number 399. Magrath, R. D. 1992. The effect of egg mass on the growth and survival of blackbirds: a field experiment. Journal of Zoology 227:639-654. Mahler, B., Y. Sarquis Adamson, A. G. Di Giacomo, V. A. Confalonieri, and J. C. Reboreda. 2009. Utilization of a new host in the host-specialist brood parasite Molothnis rufoaxillaris'. host switch or host acquisition? Behavioral Ecology and Sociobiology 63:1603-1608. Marchetti, K. 2000. Egg rejection in a passerine species: size does matter. Animal Behaviour 59:877- 883. Martin, T. E. and P. Li. 1992. Life-history traits of open- vs. cavity-ne.sting birds. Ecology 73:579-592. Martin, T. E., R. D. Bassar, S. K. Bassar, J. J. Fontaine, P. Lloyd, H. A. Mathewson, A. M. Niklison, and A. Chalfoun. 2006. Life-history and ecological correlates of geographic variation in egg and clutch mass among passerine species. Evolution 60:390- 398. Martinez, J. G., J. J. Soler, M. Soler, and T. Burke. 1998. Spatial patterns of egg laying and multiple parasitism in a brood parasite: a non-territorial system in the Great Spotted Cuckoo (Clamator glamdarius). Oecologia 117:286-294. Mason, P. 1980. Ecological and evolutionary aspects of host selection in cowbirds. Thesis. University of Texas. Austin, USA. Mason, P. 1986. Brood parasitism in a host generalist, the Shiny Cowbird. II. Host selection. Auk 103:61-69. Mason, P. 1987. Pair formation in cowbirds: evidence found for Screaming but not Shiny cowbirds. Condor 89:349-356. Ma.son, P. and S. I. Rothstein. 1986. Coevolution and avian brood parasitism: cowbird eggs show evolution- ary response to host discrimination. Evolution 40:1207-1214. Massonl V. AND J. C. Reboreda. 1998. Costs of brood parasitism and the lack of defenses on the Yellow- winged Blackbird-Shiny Cowbird system. Behavioral Ecology and Sociobiology 42:273-280. Mas.soni, V. AND J. C. Reboreda. 2002. A neglected cost of brood parasitism: egg punctures by Shiny Cowbirds during inspection of potential host nests. Condor 1()4;407-412. McLaren, C. M.. T. J. Underwood, and S. G. Sealy. 2003. Genetic and temporal patterns of multiple parasitism by Brown-headed Cowbirds (Molothm.s aler) on Song Sparrows (Melospiz.a melodia). Canadi- an Journal of Zoology 81:281-286. Mfrmoz, M. E. and G. j. Fernandez. 2003. Breeding success of a specialist brood parasite, the Screaming Cowbird, parasitizing an alternative host. Condor 105:63-72. Mermoz, M. E. and j. C. Reboreda. 1996. New host for a specialized brood parasite, the Screaming Cowbird. Condor 98:630-632. Mermoz, M. E. and J. C. Reboreda. 1999. Egg-laying behaviour by Shiny Cowbirds parasitizing Brown-and- yellow Marshbirds. Animal Behaviour 58:873-882. Mermoz, M. E. and J. C. Reboreda. 2003. Reproductive success of Shiny Cowbirds (Molothnis hoiuiriensis) parasitizing the larger Brown-and-yellow Marshbird (Pseudoleistes virescens) in Argentina. Auk 120:1 128- 1139. Miller, S. A., D. D. Dykes, and H. F. Polesky. 1988. A simple salting out procedure for extracting DNA from human nucleated cells. Nucleic Acids Research 16:1215. Nice, M. M. 1954. Problems of incubation periods in North American birds. Condor 56:173-197. Nolan, V. and C. F. Thompson. 1978. Egg volume as a predictor of hatchling weight in the Brown-headed Cowbird. Wilson Bulletin 90:353-358. Ortega, C. P. 1998. Cowbirds and other brood parasites. University of Arizona Press, Tucson, USA. Payne, R. B. and L. L. Payne. 1998. Brood parasitism by cowbirds: risks and effects on reproductive success and survival in Indigobirds. Behavioral Ecology 9: 64-73. Reboreda, J. C., M. E. Mermoz, V. Massonl A. A. Astie, and F. L. Rabuffetti. 2003. Impacto del parasitismo de cria del Tordo Renegrido (Molothnis hoiuiriensis) sobre el exito reproductive de sus hospedadores. Hornero 18:77-88. Reed, W. L., A. M. Turner, and P. R. Sotherland. 1999. Consequences of egg-size variation in the Red-winged Blackbird. Auk 116:549-552. Richner, H. 1991. The growth dynamics of sexually dimorphic birds and Fisher’s sex ratio theory: does sex-specific growth contribute to balanced sex ratios? Functional Ecology 5:19-28. Ricklefs, R. E. 1967. A graphical method of fitting equations to growth curves. Ecology 48:978-983. Schifferli, L. 1973. The effect of egg weight on the subsequent growth of nestling Great Tits Pams major. Ibis 115:549-558. Sick, H. 1985. Ornitologia brasileira: uma introducgao. University of Brasilia, Brazil. Siegel, S. and N. J. Castellan. 1988. Nonparametric statistics for the behavioral sciences. Second Edition. McGraw-Hill, New York, USA. Soler, M., J. G. Martinez, and J. J. Soler. 1996. Effects of brood parasitism by the Great Spotted, Cuckoo on the breeding success of the magpie host: an experi- mental study. Ardeola 43:87-96. Soren.son, M. D. and R. B. Payne. 2002. Molecular genetic perspectives on avian brood parasitism. Integrative and Comparative Biology 42:388^00. Strausberger. B. M. 1998. Temporal patterns of host availability. Brown-headed Cowbird brood parasitism, and parasite egg mass. Oecologia 1 16:267-274. Strau.sberger. B. M. and V. M. A.shley. 2005. Host use Dc Marsico et at. • NESTING ECOLOGY OF BAYWINGS 431 strategies of individual female Brown-headed Cow- birds Molothrus ater in a diverse avian community. Journal of Avian Biology 36:313-321. Tonra, C. M., M. E. Hauber, S. K. Heath, and M. D. Johnson. 2008. Ecological correlates and sex differ- ences in early development of a generalist brood parasite. Auk 125:205-213. Trine, C. L. 2000. Effects of multiple parasitism on cowbird and Wood Thrush nesting success. Pages 135-144 in Ecology and management of cowbirds and their hosts (J. N. M. Smith, T. L. Cook, S. I. Rothstein, S. K. Robinson, and S. G. Sealy, Editors). University of Texas Press, Austin, USA. Tuero, D. T., V. D. Fiorini, and J. C. Reboreda. 2007. Effects of Shiny Cowbird Molothrus honariensis parasitism on different components of House Wren T’/'og/or/vte.y wcr/w? reproductive success. Ibis 149:521- 529. Weatherhead, P. j. 1989. Sex-ratios, host-specific repro- ductive success, and impact of Brown-headed Cow- birds. Auk 106:358-366. WiEBE, K. L., W. D. Koenig, and K. Martin. 2007. Cost and benefits of nest reuse versus excavation in cavity-nesting birds. Annales Zoologici Fennici 44:209-217. Zanette, L., E. Macdougall-Shakleton, M. Clinchy, AND J. N. M. Smith. 2005. Brown-headed Cowbirds skew offspring sex ratios. Ecology 86:815-820. The Wilson Journal of Ornithology 122(3):432-438, 2010 BOBOLINK EGG MASS VARIABILITY AND NESTLING GROWTH PATTERNS BARBARA FREI,' ' DAVID M. BIRD,' AND RODGER D. TITMAN' ABSTRACT. — The Bobolink (Dolichonyx oryzivorus) is an obligate grassland species that is declining throughout its range in North America. There are few data available on Bobolink eggs and nestlings; this information is necessary for conservation planning efforts. Egg mass was recorded for 175 eggs from 37 nests in Quebec and eastern Ontario in 2006- 2007. Hatching asynchrony was evident with high between-clutch variation in egg mass. Egg mass did not differ with clutch size. Nest initiation date was positively correlated with smalle.st egg size and negatively correlated with within-clutch egg mass deviation. Nestling wing length, tarsus length, and mass were measured for 166 nestlings ranging from 2 to 10 days of age. Bobolink nestlings fledged below adult size and mass, achieving 87.7 ± 2.3%, 67.6 ± 1.5%, and 55.1 ± 0.4% of breeding adult tarsus length, mass, and wing length (± SE), respectively. Received 13 December 2008. Accepted 4 February 2010. The Bobolink (Dolichonyx oryzivorus) is an obligate grassland species that has recently received attention concerning population declines, mainly due to habitat loss and agricultural intensification (Bollinger et al. 1990, Herkert 1997). Studies have examined aspects of breeding ecology (e.g., Bollinger and Gavin 1992, Winter et al. 2004), but little information is available on eggs and nestlings, including linear nestling measurements. In addition, reported values are based on small sample sizes (e.g., mean egg mass, n — 5; Martin and Gavin 1995). Nestling growth rates and patterns in passerines are useful measures to examine the influence of environmental, parental, and social parameters to which the nestlings are subject. Knowledge of a time-frame of critical growth for body compo- nents may provide insight into vulnerable periods during the nesting cycle. Nestling growth may be affected by pre-hatching (e.g., egg mass and incubation) and post-hatching factors. Egg mass variability may result from physiological differ- ences among females or environmental variables such as geographic location, weather, and food availability (Christians 2002). As first hypothe- sized by Lack (1967), a trade-off exists between clutch size and egg size; females have a limited amount of energy for egg production that must be shared between the number of eggs and size of the eggs. In addition, some species exhibit seasonal change in egg size (Magrath 1992) or seasonal ' Avian .Science and Con.servation Centre, Department of Natural Resource Sciences, McGill University, Macdonald Campus, Ste-Anne-de-Bellevuc, QC H9X 3V9, Canada. ^Corresponding author; e-mail; barbara.frei@mail.mcgill.ca effects of within-clutch egg size variability (Takagi 2003). Nestling growth patterns may be described using several different parameters: ( 1 ) growth rate, (2) percent of adult weight attained by dependent young in the nest, and (3) shape of the growth curve (Ricklefs 1968a). Standard growth rate curves show mass increasing throughout the nestling period and reaching near-adult weights at fledging. There are two alternate curves where (a) nestlings attain a peak mass above adult weight and decrease to adult weight just before or after fledging, and (b) growth levels remain below adult values at fledging (Ricklefs 1968a). Growth rates may vary between individual broods of the same species, but growth patterns are species-specific (Ricklefs 1968a). For example, nestlings that fledge above adult weight are primarily oceanic species and aerial insectivores with long nestling periods, and are independent feeders post-fledging (Ricklefs 1968b). Conversely, ground-feeding species fledge below adult weight and rely on well-developed legs for escaping predators (Ricklefs 1968b). The goal of our study was to identify patterns and variability of egg mass and nestling growth in a population of Bobolinks by comparing: (1) egg mass within- and between clutches; (2) the relationship between (a) clutch size and egg mass and (b) nest initiation date and egg mass; (3) nestling growth between years of .5tudy, study sites and between broods; and (4) Bobolink nestling growth patterns by percentage of adult size and mass attained at fledging. METHODS Stiuly Area. — This study was conducted in 2006-2007 at three sites in southwestern Quebec 432 Frei Cl al. • BOBOLINK EGG MASS AND NESTLING GROWTH 433 and eastern Ontario with established Bobolink populations. The sites consisted of a private hay farm near Hemmingford, Quebec (45° 05' N, 73° 36' W), an agricultural park (Bois-de-la-Roche) in Senneville, Quebec (45° 26' N, 73° 56' W), and a wildlife conservation area (Atocas Bay) in Lefaivre, Ontario (45° 36' N, 74° 51' W). All sites consisted of multiple hayfields dominated by timothy (Phleiiiii pratense) and smooth brome (Bromiis ramosiis ssp. racemosiis) at Hemming- tord, and a broader mixture of forbs and grasses at Atocas and Bois-de-la-Roche. Measurements. — We measured and numbered eggs on day of discovery or as they were laid, and weighed them to the nearest 0. 1 g with a portable electronic balance. We took nestling measure- ments twice during consecutive visits 3^ days apart between 0700 and 1500 hrs EST. We marked individual nestlings on the initial visit by coloring their toenails using non-toxic perma- nent markers. Nestlings were separately classified by age during each visit, as asynchronous hatching resulted in different ages of nestlings within the same nest. Nestling age was classified using a combination of external physical traits; amount of down, feather tracts, feather eruption from sheaths, and opening of eyes. Photos were taken at each visit and compared to known-age nestlings to confirm classification. We measured length of the unflattened wing (wing length) to the nearest 0.5 mm using a ruler with a wing stop, and tarsometatarsus (tarsus) length to the nearest 0.1 mm with dial calipers. We weighed nestlings to the nearest 0. 1 g. The same person performed all measurements to reduce observer bias. We banded nestlings at > 7 days of age using aluminum bands. We monitored all nests until completion, and nests that fledged at least one young were recorded as successful. Statistical Analysis. — We checked egg and nestling data for normality using Shapiro-Wilks’ test (Zar 1999). All measurements were normally distributed, allowing use of untransformed data in subsequent analyses. We compared egg mass between years using Student’s r-test and between sites using one-way ANOVA (Zar 1999). We examined the variance of egg mass between and within clutches using one-way ANOVA. We examined if mean egg mass per clutch varied with changes in nest initiation date using linear regression. We also examined the variance of mean egg mass with clutch size using one-way ANOVA. We examined the effect of initiation date on mass of the smallest egg in each clutch using linear regression. We u.sed the standard deviation of mean egg mass per clutch as a measure of within-clutch egg mass variation. We regressed standard deviations with Julian initia- tion date to examine if within-clutch variation changed with initiation date of the nest. We measured a subset (/; = 102) of nestlings twice over the span of 10 days, obtaining a mixed longitudinal sample of nestling growth (Ricklefs 1983). We measured the remaining nestlings once [n = 64). We calculated mean age measurements for nestling wing length, tarsus length, and mass from 2 to 10 days of age, and examined variance across all ages using ANOVA. We used nested ANOVA in which year of study, study site, and brood were nested within nestling age. Missing values for day 8 at Bois-de-la-Roche, and day 10 for both Bois-de-la-Roche and Atocas were replaced with mean values for those dates. We calculated the percent of adult size and mass that was reached each day of nestling growth. Individual nestling measurements were divided by adult measurements, giving a percentage of adult growth. These were averaged to acquire mean (± SE) percentage of adult growth for nestlings of 2-10 days of age. Male and female Bobolinks are dimoiphic, but cannot be assigned to gender without genetic testing as nestlings. Thus, we averaged the reported adult male and female measurements from the breeding season to use as general adult measurements (Martin and Gavin 1995). We reported all mean values ± SE and used SPSS 16.0 (SPSS Inc., Chicago, IL, USA) and JMP 8.0.1 (SAS Institute Inc., Cary, NC, USA) for all analyses. RESULTS Hatching Phenology and Egg Measurements. — Three of 25 nests for which nestling measure- ments were obtained hatched synchronously (12%); 20 nests had one marginal nestling (80%), and two nests had two marginal nestlings (8%). Nestlings in three of four re-nests were each I day apart in age. Measurements were taken for 175 eggs from 37 nests. Sixteen of the 53 nests monitored were excluded from egg calculations as they were found either during the nestling stage or depre- dated prior to measurements being taken. Mean egg mass over both years and all sites (/; = 175) was 2.69 ± 0.05 g (1.6-3. 5 g). We found no 434 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 3, September 2010 FIG. 1. Linear regression with 95% Cl of smallest egg mass in Bobolink clutch compared to initiation date at Atocas Bay, Ontario, Bois-de-la-Roche, Quebec, and Hemmingford, Quebec during 2006 and 2007. differences between Bobolink egg mass between years of study (7175 = —1.87, P = 0.062) and among study sites {Fx■^^ = 1.45, P = 0.24). We found a between-clutch effect of egg mass across Bobolink females (F37 = 13.089, P < 0.001). The model sum of squares accounted for between-clutch variation only (12.773) and the total sum of squares accounted for both between- and within-clutch variation (16.824). Thus, be- tween-clutch variation accounted for 15.9% of the total variation. We found no differences in egg mass between clutch sizes (F37 = 0.830, P = 0.487) and mean egg ma.ss was not related with initiation date (r = 0.088; F37 = 1 .89, P = 0. 10). Egg mass of the smallest egg in the clutch increa,sed with initiation date (r = 0.13; /r3^ = 4.390, P = 0.045, [i = 0.021 ± 0.01 SE; Fig. 1 ). Within-clutch egg ma.ss variation was not related to clutch size (F37 = 0.246, P = 0.863), but decrea,sed with greater date of nest initiation (r = 0.277; F37 = 1 1 .5(X), P = 0.002; Fig. 2). Nestling Measurements and Growth. — We re- port mean ± SE nestling measurements for nestlings at ages 2-10 days (Table 1). Bobolink nestlings had ~ 5-10% larger tarsi in 2007 than in 2006 (F268 = ^-87 1, P < 0.001) and ~ 10-15% more mass in 2007 than in 2006 (F268 ~ 3.620, P < 0.001). Wing length did not differ between years (F268 = 1-382, P - 0.204). Year of study had an impact on two of the three variables tested; we tested among study sites within each year to prevent possible year effects. None of the variables for 2006 differed among study sites: nestling wing length (F268 = 1.221, F = 0.279), tarsus length (F268 = 0.518, F = 0.898), and mass (^268 0.530, F = 0.890). However, all variables differed among sites in 2007: nestling wing length (^268 “ 1.859, F = 0.033), tarsus length (F268 = 1.838 F = 0.036), and mass (F268 = 2.330, F = 0.005). Among site-differences were most evident between younger nestlings with Atocas > Bois- de-la-Roche > Hemmingford nestlings; there were no differences among sites at nestling age 8-10 days. All variables in all years differed between individual broods. Nestling growth as a percentage of adult size and mass increased predictably across nestling age (Fig. 3). Nestlings for all variables were below adult size or mass at day 10, just prior to or at Hedging. Nestling measurements as percentage of adult ± SE at day 10 were: wing length = 55.1 ± 0.4%, tarsus length = 87.7 ± 2.3%, and mass = 67.6 ± 1 .5%. DISCUSSION Asynchronous Hatching. — Hatching asynchro- ny was prevalent in the Bobolink population we studied. Most nests (80%) had a single late- Frei cl al. • BOBOLINK EGG MASS AND NESTLING GROWTH 435 FIG. 2. Linear regression showing within-ckitch deviation of Bobolink egg mass compared to initiation date at Atocas Bay, Ontario, Bois-de-la-Roche, Quebec, and Hemmingford, Quebec during 2006 and 2007. hatched nestling, resulting from incubation begin- ning with the penultimate egg, as previously described for Bobolinks (Martin 1974). Few nests had two late-hatched nestlings (8%), far fewer than in the closely related Red-winged Blackbird (Agelaius phoeniceus) where 8.1% of nests hatched synchronously, 41.9% had one late- hatched nestling, and 43.3% had two late-hatched nestlings (Forbes and Glassey 2000). Three of four re-nests in our study had nestlings staggered in age, a result of incubation beginning with the first egg laid. This, paired with smaller clutch size, suggests females shorten the re-nesting cycle as Bobolinks rarely re-nest (Martin and Gavin 1995). Egg Mass Variation Within- and Between Clutches. — Between-clutch variation of egg mass far exceeded within-clutch variation. This pattern is common for most bird species with an average of 70% of variation in egg size due to variation between- rather than from within-clutches (Chris- tians 2002). Egg size appears to reflect differences among individual females, and often increases with age and experience (Christians 2002). Overall, consistency of egg and clutch size is a consequence of selective pressures and energy expenditure limitations (Lack 1947). Egg Mass Variation with Clutch Size and Nest Initiation Date. — Our data did not support Lack’s (1967) trade-off hypothesis that greater energy TABLE 1. Mean (± SE) wing, tarsus, and mass measurements for Bobolink nestlings (/; Ontario, Bois-de-la-Roche, Quebec, and Hemmingford. Quebec during 2006 and 2007. = 268) at Atocas Bay, Nestling age (days) n Wing length (cm) Tarsus length (cm) Mas,s (g) 2 6 1 .02 ± 0.03 0.95 ± 0.03 6.5 ± 0.6 3 24 1 .28 ± 0.04 1.08 ± 0.04 7.4 ± 0.3 4 47 1.78 ± 0.04 1.31 ± 0.04 10.0 ± 0.3 5 39 2.38 ± 0.05 1.67 ± 0.04 14.2 ± 0.3 6 39 3.00 ± 0.05 1.86 ± 0.05 16.4 ± 0.3 7 33 3.34 ± 0.08 2.07 ± 0.09 17.3 ± 0.8 8 46 4.22 ± 0.06 2.22 ± 0.03 19.9 ± 0.9 9 27 4.72 ± 0.04 2.33 ± 0.04 21.5 ± 0.4 10 7 5.10 ± 0.04 2.25 ± 0.06 21.6 ± 0.5 436 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 3. September 2010 FIG. 3. Nestling growth represented as a percent of breeding adult Bobolink measurements ± SE for wing length, tarsus length, and mass for Bobolink nestlings pooled for sites and years. expenditure during laying results in larger clutch- es of smaller eggs. This hypothesis has been empirically supported in several studies (Black- burn 1991, Olsen et al. 1994), but others have failed to find a negative relationship between clutch size and egg size (Rohwer and Eisenhauer 1989). The opposite has been demonstrated to be true, in some cases, where smaller clutches contain smaller eggs (Galbraith 1988). Female Bobolinks appear to have laid larger eggs more consistently as the nesting season progressed, as nest initiation date was positively correlated with egg mass of the smallest egg and negatively conelated with within-clutch variation. In temperate regions, such as eastern Canada, ambient temperature increases throughout the breeding season resulting in decreased daily energy expenditure for homeotherms and in- creased food availability for insectivores (Avery and Krebs 1984, Eeva et al. 2000). In turn, this relaxes energy expenditure constraints related to egg-laying activities and may result in larger eggs later in the season. Mean egg mass of Common Blackbirds [Tunhis menila) was correlated with air temperature during the period when the species was undergoing rapid follicular growth (Magrath 1992), and Great Tits (Pams major) roosting in heated nest boxes laid larger eggs than those in cooled nest boxes (Nager and Zandt 1994). Few studies have investigated within-clutch .seasonal variation of egg size. Takagi (2003) in northern Japan during a study of Bull-headed Shrikes (Lanins bucephalus) found a seasonal increase in intra-clutch variation of egg volume due to a sea.sonal size decline in the smallest egg within the clutch. The inverse was true in the Bobolinks we studied; within-clutch variation decreased as the mass of the smallest egg increased over the season. Bobolinks in our study area began laying from mid- to late May. It is often cool and rainy during this period resulting in high energetic costs for laying females, leading to smaller smallest eggs and higher within-clutch variation earlier in the breeding season. Common Swifts (Apus apus) similarly lay lighter eggs during unfavorable weather (O’Conner 1979). Nestling Growth. — Bobolink young at fledging were well below adult size and mass. Thus, the Bobolink growth curve is a variant of the ‘standard’ growth curve in which growth levels off below adult weight (Ricklefs i968a). Bobolink young are often described as poor flyers for several days post-lledging (Martin and Gavin 1995), a result of leaving the nest prior to full feather emergence. Eastern Meadowlarks (Stnr- nella magna) (Lanyon 1995) and Savannah Sparrows (Passercnlns sandwichensis) (Wheel- wright and Rising 2008) are also ground-nesting grassland birds that fledge at below adult mass. This may be adaptive for ground-nesting birds, whose young are especially vulnerable to preda- tors and disturbance. Nestlings of species with Frei el al. • liOBOLlNK EGG MASS AND NESTLING GROWTH 437 well-protected nests generally remain in nests longer than young in exposed nests (Ricklefs 1968a). Growth ol wing, tarsus, and mass was greatest between nestling days 3 and 6. The inflection point occurred at about days 8-9, when growth leveled off below adult size and mass. Studies in New York and Wisconsin indicated the Bobolink nestling mass gain inflection point occuiTed 8 days after hatching (Martin and Gavin 1995). Tarsus length was the closest to adult size at fledging of the three variables measured. This accelerated growth of a specific nestling characteristic may be a reflection of the selective pressure to enable young to run from the nest by day 7 if disturbed (Martin and Gavin 1995). Eastern Meadowlarks and Savannah Sparrows also attain near-adult tarsus length at fledging and are known to escape from nests by running prior to fledging (Lanyon 1995, Wheelwright and Rising 2008). Incomplete growth prior to fledging is associated with provisioning by parents after the young leave the nest, when they complete their growth to adult size (Martin and Gavin 1995). This is the first report of linear nestling measurements, as well as the first comparison of egg size and within-clutch variation with clutch size and initiation date for Bobolinks. This study provides baseline data for future studies, as well as information on timing of critical growth for nestlings, which should prove useful for conser- vation and management planning for this declin- ing grassland species. ACKNOWLEDGMENTS Field assistance for this project was ably provided by Carine Lecoeur. Kate Robinson, Rachel Theoret-Gosselin, and Rachel Verkade. Funding for this project was provided by the Natural Sciences and Engineering Research Council (NSERC), Fondation de la faune du Quebec (FFQ), Canadian Wildlife Service (CWS). Bird Protection Quebec (BPQ), and McGill University. We thank Marie-Anne Hudson, Shawn Craik, and two anonymous reviewers who provided thoughtful criticisms to the manuscript. We thank Andrew Stairs, the City of Montreal, and Ducks Unlimited for use of their hayfields for the study. We are grateful to the McGill Bird Qbservatory (MBO) for training in banding and equipment use, and the Avian Science and Conserva- tion Centre (ASCC) for support. LITERATURE CITED Avery, M. 1. and J. R. Krebs. 1984. Temperature and foraging success of Great Tits Pams major hunting for spiders. Ibis 126:33-38. Blackburn, T. M. 1991. An interspecific relationship between egg size and clutch size in birds. Auk 108:973-977. Bollinger, E. K. and T. A. Gavin. 1992. Easiern Bobolink populations: ecology and conservation in an agricultural landscape. Pages 497-506 in Ecology and conservation of neotropical migrants (J. M. Hagan and D. W. Johnston, Editors). Smith.sonian Institute Press, Washington, D.C., USA. Bollinger, E. K., P. B. Bollinger, and T, A. Gavin. 1990. Effects of hay-cropping on eastern populations of the Bobolink. Wildlife Society Bulletin 18:142- 150. Christians, J. K. 2002. Avian egg size: variation within species and inflexibility within individuals. Biological Reviews 77:1-26. Eeva, T., S. Veistola, and E. Lehikoinen. 2000. Timing of breeding in subarctic passerines in relation to food availability. Canadian Journal of Zoology 78:67-78. Forbes, S. and B. Glassey. 2000. Asymmetric sibling rivalry and nesting growth in Red-winged Blackbirds {Agelaius phoeniceus). Behavioural Ecology and Sociobiology 48:413^17. Galbraith, H. 1988. Effects of egg size and composition on the size, quality and survival of Lapwing Vanellus vanellus chicks. Journal of Zoology 214:383-398. Herkert, j. R. 1997. Bobolink (Dolichonyx oryzivoms) population decline in agricultural landscapes in the midwestem USA. Biological Conservation 80:107- 112. Lack, D. 1947. The significance of clutch size. Ibis 89:302-352. Lack, D. 1967. The significance of clutch size in waterfowl. Wildfowl 9:67-69. Lanyon, W. E. 1995. Eastern Meadowlark (Sturnella magna). The birds of North America. Number 160. Magrath, R. D. 1992. Seasonal changes in egg-mass within and among clutches of birds: general explana- tions and a field study of the Blackbird Tiirdns menda. Ibis 134:171-179. Martin, S. G. 1974. Adaptations for polygynous breeding in the Bobolink {Dolichonyx oiyzivorons). American Zoologist 14:109-119. Martin, S. G. and T. A. Gavin. 1995. Bobolink (Dolichonyx oiyzivorons). The birds of North Amer- ica. Number 176. Nager. R. G. and H. S. Zandt. 1994. Variation in egg size in Great Tits. Ardea 82:315-328. Q’Conner. R. j. 1979. Egg weights and brood reduction in the European Swift (Apns apns). Condor 81:133-145. Olsen, P. D., R. B. Cunningham, and C. F. Donnelly. 1994. Is there a trade-off between egg size and clutch size in altricial and precocial nonpasserines? A test of a model of the relationship between egg and clutch size. Australian Journal of Zoology 42:323-328. Ricklefs, R. E. 1968a. Patterns of growth in birds. Ibis 1 10:419^51. Ricklefs, R. E. 1968b. Weight recession in nestling birds. Auk 84:30-35. Ricklefs, R. E. 1983. Avian postnatal development. Pages 1-83 in Avian biology (D. S. Earner and J. R. King. Editors). Academic Press. New York, USA. 438 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 3, September 2010 Rohwer, F. C. and D. I. Eisenhauer. 1989. Egg mass and clutch size relationships in geese, eiders, and swans. Ornis Scandinavica 20:43^8. Takagi, M. 2003. Seasonal change in egg-volume variation within clutch in the Bull-headed Shrike, Lanius bucephalus. Canadian Journal of Zoology 81:287- 293. Wheelwright, N. T. and J. D. Rising. 2008. Savannah Sparrow {Passerculus sandwichensis). The birds of North America. Number 45. Winter, M., D. H. Johnson, J. A. Shaffer, and W. D. SvEDARSKY. 2004. Nesting biology of three grassland passerines in the northern tallgrass prairie. Wilson Bulletin 116:211-223. Zar, j. H. 1999. Biostatistical analysis. Third Edition. Prentice Hall, Upper Saddle River, New Jersey, USA. The IVilson Journal of Orni/liology 1 22(3):439-446, 2010 BREEDING PHENOLOGY AND NESTING SUCCESS OF THE YUCATAN WREN IN THE YUCATAN PENINSULA, MEXICO JESUS VARGAS-SORIANO,'^ JAVIER SALGADO ORTIZ,- AND GRISELDA ESCALONA SEGURA^ ABSTRACT. — The Yucatan Wren (Campylorhynchus yucatanicus) is a highly restricted endemic species inhabiting the coastal scrub at the northern portion of the Yucatan Peninsula. We describe the breeding phenology and nesting success of this endangered species from April to September 2007 for a population at Ria Celestun Biosphere Reserve. We found 232 nests of which only 1 10 (47%) were active at either incubation or nestling stages. Yucatan wrens initiated nest building in late April, but clutch initiation occurred in early June and lasted until the end of July, resulting in a breeding season of 2 months. Nests were exclusively in coastal scrub and the transition between coastal scrub and black mangrove (Avicennia germinans) forest. Eleven species of trees were used as nesting substrate, but three included 75% of all nests found. Clutch size (x ± SD) was 3 ± 1.5 eggs with incubation and nestling periods averaging 16 ± 1.0 days, and 16.5 ± 1.9 days, respectively. Mayfield estimates of daily survival rate for incubation and nestling periods were 0.968 ± 0.005 and 0.975 ± 0.005, respectively with nesting success of 46%. The average number of fledglings per successful nest was 2.5 ± 1.3. Predation was the main cause of nest mortality accounting for 54% of the active nests. Parental care was provided by both parents, but participation of a third individual feeding nestlings was recorded at three nests, providing evidence for occasional cooperative breeding. Increasing human development in the coastal region of the Yucatan Peninsula may represent a serious threat to conservation of the Yucatan Wren due to habitat restriction and high dependency on three species of trees as nesting substrate. Received 3 December 2008. Accepted 4 Fehruar}’ 2010. Information on life history traits and social systems is scarce or lacking for many species of birds, especially those at tropical latitudes (Martin 1996, 2004). Estimation of productivity is one of the most important demographic parameters necessary to describe numerical fluctuations in space and time (Donovan et al. 1995, Jehle et al. 2004, Powell and Knutson 2006). Three approach- es have been suggested to measure bird produc- tivity: (1) percentage of nests constructed by females per year (Ricklefs and Bloom 1977, Anders and Marshall 2005), (2) nesting success (based on number of nestlings from individual broods that survive until the end of the reproduc- tive period as fledglings), and (3) estimation of reproductive success (based on number of indi- viduals that survive and reproduce in the follow- ing reproductive cycle). Estimation of reproduc- ' Laboratorio de Vida Silvestre y Colecciones Cientificas, Centro de Estudios de Desarrollo Sustentable y Aprove- chamiento de la Vida Silvestre, Universidad Autonoma de Campeche, Agustin Melgar s/n entre Juan de la Barrera y Calle 20, Colonia Buenavista, 24039, San Francisco de Campeche, Campeche, Mexico. ^Laboratorio de Ornitologi'a, Editlcio B4, Facultad de Biologia, Universidad Michoacana de San Nicolas de Hidalgo, Ciudad Universitaria, Morelia, Michoacan, Mex- ico. ^EI Colegio de la Frontera Sur Unidad Campeche, Calle 10 No. 264 Colonia Centro, 24000, San Francisco de Campeche, Campeche. Mexico. "Corresponding author; e-mail: abucefalo@hotmail.com live success provides information on population dynamics, and the probability of species extinc- tion or survival. Reproductive success and proportion of nests constructed by females have been difficult to estimate due to logistical limitations of following females during one or more reproductive seasons, or as a result of not knowing the final fate of individuals in time and space, particularly once they become independent of their parents and disperse to new areas (Ricklefs and Bloom 1977). However, nests are relatively easy to monitor and provide robust data on productivity. Thus, nesting success has been frequently used to evaluate population changes in bird communities (Ricklefs and Bloom 1977. Johnson and Shaffer 1990, Knadle et al. 2001, Jehle et al. 2004, Anders and Marshall 2005. Powell and Knutson 2006). Researchers can monitor nests to estimate nesting success and factors that influence repro- ductive success including vegetation structure, depredation, nest parasitism, and climatic condi- tions (Hazier 2004, Martin et al. 2006). This information is essential in understanding the ecology of species, which can suggest strategies for habitat conservation and management (Peter- son et al. 1993). The Yucatan Wren {Campylorhynchus yucata- nicus) occurs exclusively in the coastal portion of the Yucatan Peninsula in the states of Campeche and Yucatan, where it inhabits the thin line of 439 440 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 3. September 2010 FIG. I. Study Area showing transects at Ria Celestun Biosphere Reserve, Yucatan, Mexico. coastal scrub (Peterson and Chalif 1989, Howell and Webb 1995, Brewer 2001). It is endemic to the Yucatan Peninsula and considered by the Mexican Government as threatened and is includ- ed on the lUCN red list as near threatened (Diario Oficial de la Federacion 2002, BirdLife Interna- tional 2008). Information on the biology of this species has been limited to general descriptions of distribution and habitat use, and is largely anecdotal based on observations by Zimmerman (1957), who provid- ed a brief description of the nest and behavior of individuals from a population at Sisal, Yucatan. Our objectives were to obtain information on (1) breeding seasonality, (2) length of incubation and brooding periods, (3) clutch size, (4) parental care, and (5) nesting success. METHODS Study Arec/.— The study was conducted during April to September 2007 at Ria Celestun Bio- sphere Re.serve at the northwestern portion of the Yucatan Peninsula, Mexico (20° 46' and 21° 06' N, 90° I r and 90° 25' W) on the boundaries of the states of Campeche and Yucatan (Fig. I). The climate of the area is warm semi-dry with rain occurring mainly during summer. Annual precip- itation averages 750 mm with a dry season occurring from December to May and a wet season from June to November. The mean annual temperature is 26.5° C with May being the warmest month (average 29° C) (CONANP 2000). The most abundant types of vegetation at Ria Celestun Biosphere Reserve are coastal scrub, and mangrove and flooded lowland forests (CONANP 2000). Coastal scrub has two phases of develop- ment with the first adjacent to the beach and integrated by clumps of shrubs where the main species are bay cedar (Suriana inaritima) and sea rosemary {Toumefortia gnapluilodes). The second corresponds to scrub towards the interior of the dunes in a zone where the sand is less altered and canopy height ranges from 3 to 5 m. The most common plant species in this zone are white tree (Bnivaisia herlonderiana), wood luck seed (Tlie- vetia gaumeri), sea grape (Coccoloha uvifera), geiger tree (Cordia sehestena), black poison wood (Metopium brownie), peccary tree {Sideroxylon Wniras-Soriano cl al. • YUCATAN WREN NESTING SUCCESS 441 americiiniaii), joewood {Jcicquinia aiirantiaca), peccary wood (Caesalpinia vesicaria), blackhead (Pithecellobiiun keyeiise), wild sage {Liintcuui involiicrata), blacktorcli {Erithalis friiticosa), mexican cotton {Gossypiitm hirsiitiim), and cen- tury plant {A^cive angustifolia). Mangrove forest is abundant along the estuary and is represented by red (Rhizophoni mangle), black {Avicennia germinans), and white mangrove {Laguncitlaria racemosa) with tree heights rang- ing from 12 to 14 m. Other species such as pickleweed (Baris maritima) and sea purslane (Sesiivinni portiilacastnun) together with numer- ous species of grasses (Cyperaceae and Grami- neae) also occur along drainage channels forming patches, characterized by short trees ranging from 1.5 to 3 m in height, known as logwood patches. The flooded lowland forest merges between the coastal scrub and mangrove forest and is repre- sented by logwood tree (Haematoxyhim campe- chianum), milk tree (Cameraria latifolia), black poison wood, chicle tree (Manilkara zapota), gambo limbo (Biirsera simaruba), nance (Byrso- nima crassifolia), black olive tree (Bucida bu- ceras), ceiba tree (Ceiba aesculifolia), buttercup tree (Cochlospermum vitifolium), and button mangrove (Conocarpus erectus). Nest Searching and Monitoring. — We estab- lished line transects ranging from 4 to 6 km in length in each of the three main vegetation types within the study area. We visited each transect starting in April 2007 to locate wrens and nesting sites based on presence of old nests. Other wren species are known to maintain year-round nests, apparently used as dorms (Robinson et al. 2000, Yaber and Rabenold 2002). Yucatan Wrens are no exception, as old nests were found along each transect; old nests were used as evidence of wren presence and potential nesting sites. Each site was checked for nesting activity every third day with observations throughout morning (0630-1200 hrs EDT) and afternoon (1600-1930 hrs), following Martin and Geupel (1993). We found new nests, primarily by scanning trees along transects (Yucatan Wrens make nests of grasses and the yellow material contrasts with the dark back- ground of trees used as nesting substrate). We also used parental behavioral cues including observa- tion of birds carrying nesting material, alarm calls of wrens during both incubation and brooding, and parents delivering food to nestlings (Ralph et al. 1996). Few nests were found by accidentally flushing adults from the nest while walking. We recorded the number of nests found in each vegetation type and identified the species of trees used as nesting substrate. We checked nests every 2-4 days until failure or success to estimate clutch initiation, length of incubation and brooding periods, and final nest fate. We monitored nests daily when a transition, hatching or fledging, was anticipated. Yucatan Wren nests are dome shaped and bulky with a tubular entrance; we used a fiberscope (Provision model 960, Chicago Miniature Inc., Hackensack, NJ, USA.) to examine nest contents and avoid damage to nests and contents. We considered a nest failed if all eggs or nestlings disappeared and successful if one or more young fledged. We used three video cameras (SONY DCR- HC30/HC40) to record parental care and to document cooperative breeding as this behavior has been reported for other species of wrens (Rabenold 1984), Cameras were used at 45 of 1 10 active nests for a total of 200 hrs of video recording. Video cameras were used only at nests with nestlings. Each video was later reviewed with the aid of a TV monitor to calculate parental visitation rate, feeding, and presence of individ- uals other than the parental pair. Statistical Analyses. — We report means ± SD for all variables unless otherwise specified. Nesting success and daily nest survival probabil- ities were calculated following Mayfield (1961, 1975). RESULTS Timing of Breeding. — We found 232 nests throughout the 2007 breeding season of which 110 (47%) were active: laying (59), incubation (40), and nestling (11). The remaining 122 (53%) nests were abandoned during the building stage. Yucatan Wrens started nest building in early April, but most nests were built in June (50%) with fewer in May (35%), July (13.5%), and April (1.3%). Nests were build by both parents and the average time to build a nest was 6 ± 1.5 days (range 4-8, n = 149). Clutch initiation started in June despite early nest building. The earliest clutch recorded was on 1 1 June and the latest on 28 July; the breeding season extended for 2 months (Jun-Jul). There was a time lag of 19 ± 8.5 days (range 5-33, n = 24) between the end of nest building and laying of the first eggs. Old nests within temtories were not used for breeding. 442 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122. No. 3. September 2010 Cocos nucifera Maytenus phyllanthoides Coccoloba uvifera o Jacquinia aurantiaca o Avicennia gerrninans ^ Caesalpinia vesicaria Sunana mantima Sideroxylon celastrina Pithecellobium keyense Sideroxylon americanum Conocarpus erectus 0 10 20 30 40 50 60 70 80 Number of nests FIG. 2. Tree species used by Yucatan Wrens as nesting substrate at Ria Celestun Biosphere Reserve, Yucatan. Mexico. Clutch Size, Length of Incubation and Nestling Periods. — Clutch size averaged 3 ± 1.5 eggs (range 1-5; n = 99 nests), but half of the nests (49.5%) had four eggs. The proportion of nests with one egg was 10.1%, two eggs = 18.2%, three eggs = 21.2%, and five eggs = 1%. Yucatan Wrens typically laid one egg per day and females initiated incubation with laying of the last egg. The incubation period averaged 16 ± 1.0 days (range 15-17, n ~ 47). The earliest hatching date was 28 June and the latest was 1 1 August. Hatching was synchronous as all eggs within individual broods hatched within 24 hrs after the first egg hatched. Yucatan Wren nestlings fledged on average 16.5 ± 1.9 days after hatching (range 14-19; n = 5 1 ). The average number of young fledged per successful nest was 2.5 ± 1.3 (1^ young; n — 51 ). The earliest date of fledging was 14 July and the latest was 23 August. Parental Care. — Incubation feeding by Yuca- tan Wrens was not observed. Both parents fed nestlings at all nests observed (/? = 51) for parental care. The average number of provisioning visits/hr by both parents combined was 7 ± 1.6 (range 5-9 visits). We recorded the presence of a third adult feeding young at three nests, confirm- ing occasional cooperative breeding in this species. Fledglings remained clo.se to the nesting site for 2-3 days after fledging and both parents continued providing care to fledglings, indepen- dently of number. Nesting 5//cxr.v,v.— The Mayfield estimate of daily survival for the incubation period was 0.968 ± ().()()5 and 0.975 ± 0.005 for the nestling period. Estimated nesting success was 46%. Nest predation was the only cause of nesting failure, accounting for loss of 54% of all active nests. We found no evidence of brood parasitism by cowbirds of Yucatan Wren nests. Ectoparasites (Diptera: Muscidae) were observed on 20 juve- niles in seven nests, but no mortality was related to parasitism. Habitat Use and Nesting Substrate. — One hundred and fifty of the 232 nests (65%) were in coastal scrub, while the remaining 82 (35%) were in flooded lowland forest (between the coastal scrub and mangrove forest). Yucatan Wrens used 1 1 species of trees as nesting substrate (Eig. 2). However, only three species, button mangrove, peccary tree, and blackhead supported 75% of all nests found (Eig. 2). DISCUSSION The breeding season of Yucatan Wrens at Celestun was highly seasonal, occuiring over a period of ~ 3 months (May-Jul). Nest building started during the dry season in late April, but clutch initiation was not recorded until June suggesting initiation of breeding is closely related to seasonal occurrence of rainfall. Previous studies on avian breeding phenology have sug- gested that onset of breeding of tropical birds is closely connected with rainy seasons' (Ahumada 2001, Tarraoux and McNeil 2003). It is possible that rainfall may also be a primary stimulus for reproductive activity for Yucatan Wrens. We did not measure the rainfall regime for the study period, but Salgado-Ortiz et al. (2008) reported that 75% of the rain occurs from June to Vurgas-Soriano cl at. • YUCA I AN WREN NESTING SUCCESS 443 September in Celestun with a monthly maximum in September. Our observations of a time lag that averaged 20 days from the end of nest building to laying of the first egg in June supports the possibility that Yucatan Wrens may be using rain as an environmental cue to adjust timing of breeding. Clutch initiation of Yucatan Wrens extended over 2 months (Jun-Jul) with no evidence of second nesting attempts after either failure or success. The clutch initiation period is similar to that recorded for Cactus Wrens {Campylo- rhynchiis bninneicapilhis) in Arizona (Simons and Martin 1990), but shorter than that of Stripe-backed Wrens (C. nuchalis) in Venezuela (Piper 1994). The difference in duration (2 months) of the breeding season between Yucatan and Striped-backed wrens may be explained by seasonal effects occurring in the flooded savanna where the latter species occurs. The savanna remains relatively wet weeks after arrival of the first rains, presumably resulting in food being available for a longer period, which in turn ensures the possibility of more nesting attempts. The average clutch size of Yucatan Wrens was three eggs, similar to that of Rufous-naped Wrens (C. rufinucha) in Costa Rica (Skutch 1985), and slightly smaller than that of Cactus Wrens (4 eggs) in Arizona (Simons and Simons 1990). The smaller clutch size of the Yucatan Wren appears to fit the latitudinal trend in clutch size from temperate to tropical latitudes. However, the evidence is inconclusive because ~ 50% of Yucatan Wren nests had four eggs suggesting differences in clutch size may not be significant, as has been found for open cup nesting species (Martin 1996). The incubation period averaged 16 days, sim- ilar to that of the Cactus Wren; however, incubation started with the laying of the last egg by Yucatan Wrens in contrast to the first egg for Cactus Wrens (Farley and Stuart 1994). It is possible Yucatan Wrens begin incubation with the last egg to synchronize hatching. We observed that hatching of all eggs occurred within 24 hrs of hatching of the first egg, but were unable to obtain detailed description of hatching synchrony. The difference in initiation of incubation versus that of Cactus Wrens could be attributed to the large change of temperatures throughout the day occurring in the arid desert where Cactus Wrens occur (Mezquida 2004). Cactus Wrens starts breeding in March (Simons and Martin 1990), coinciding with frequent cold fronts from the north resulting in extreme weather conditions. Initiation of incubation with laying of the first egg could be a local adaptation to protect eggs from unstable weather conditions (Hansell 2000). Parental care, as confirmed by video cameras, was provided by both parents and an occasional adult helper at three nests. We were not able to capture and color-band individuals, but videotap- ing provided evidence of individuals in addition to the parents visiting active nests and providing food to nestlings. We suspect that cooperative breeding its more widespread in Yucatan Wrens, but confirmation requires study with color-banded individuals. Austad and Rabenold (1986) and Rabenold (1990) reported cooperative breeding has been documented in nine of the 13 species of Campy lorhyncluis, not including the Yucatan Wren. Our evidence is scarce, but the highly restricted coastal arid scrub where the Yucatan Wren occurs and dramatic seasonal changes have been suggested as environmental or causative factors for occurrence of cooperative breeding in other species (Rabenold 1990). Predation was the main cause of failure of Yucatan Wren nests. Stutchbury and Morton (2001) report nest depredation rates for tropical birds as a group are higher (65-80%) compared to those of temperate zone birds. The nest loss (54%) for Yucatan Wrens at Celestun is less than reported as typical for tropical birds. Nest depredation has been reported to generally be higher for open cup nests that for closed nests (Collias and Collias 1984, Martin 1995, Auer et al. 2007). Lower nest depredation of Yucatan Wrens could be a result of the dome-shaped nest. Clumping of nests has been reported to reduce the probability of depredation, increasing reproduc- tive success (Bowman and Hams 1980, Martin 1993). Yucatan Wrens built .several nests in the same tree; we did not study the direct effect of clumped nests on nesting success, but this behavior could explain the lower depredation rate of Yucatan Wren nests compared to Mangrove Warblers {Dendroica petechia Ivyanti) nesting in the same area (67-88%) (Salgado-Ortiz et al. 2008). The Mayfield estimate of daily survival rate of Yucatan Wren nests (0.978) was higher than reported for Mangrove Warblers (0.950) in the same area (Salgado-Ortiz et al. 2008), but was similar to that reported for Cactus Wrens (0.99) (Ricklefs 1968). Daily survival rates of Yucatan 444 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 3. September 2010 and Cactus wrens provide evidence to support the hypothesis that dome-shaped nests survive better than open cup nests (Martin 1996). Construction of several nests in clumps appears to be widespread behavior among the Troglody- tidae (Marquez-Valdelamar 1998). Nests built, but not intended for breeding are used as “dummy” nests, a strategy suggested to reduce depredation and increase chances of adult survival (Marquez- Valdelamar 1998, Robinson et al. 2000, Puga- Vazquez 2008). Robinson et al. (2000) demon- strated that only half of the nests constructed by Song Wrens (Cyphorhinus phaeocepholiis) in Panama were used for reproduction, while the others were constructed to confuse nest predators. Yucatan Wrens used 47% of the total new nests found over the breeding season. Some nests in our study were destroyed during the building stage by other individuals of the same species or uniden- tified predators, while others were finished, but remained unused for reproduction, supporting the “dummy” anti-predator strategy. Yucatan Wrens nested exclusively within coastal scrub and the transition to mangrove forest. This vegetation zone extends < 1 km inland. Yucatan Wrens mostly used three species of trees with button mangrove being used as the main nesting substrate. This indicates Yucatan Wrens are highly dependent on this tree species as nesting substrate. Previous studies have shown that other species of Troglodytidae construct nests mainly in trees with thorns, particularly of the genus Acacia. It has been hypothesized that use of thorny plants as nesting substrate is for protection from predators (Young et al. 1990, Marquez- Valdelamar 1998, Skutch 2001). This also has been suggested for Cactus Wren, which uses cactus (Opuntia spp.) as nesting substrate (Face- mire et al. 1990). Yucatan Wrens at Celesttin u.sed species lacking thorns, despite several thorny species occurring in coastal scrub. We believe the high use of button mangrove as nesting substrate reflects the greater abundance of this tree species within the preferred nesting habitat. Yucatan Wrens were not locally rare in our study area, but are highly habitat restricted and show marked preference in nesting substrate. Button mangrove is widely irsed as firewood by local people who are reducing the availability of this species. This use may have negative implications on future nesting success of Yucatan Wrens. The current human population increa.se and increasing land development resulting from the touri.st industry in coastal areas are becoming serious threats to future conservation of Yucatan Wrens and their habitat. ACKNOWLEDGMENTS We are grateful to Victor Nah Chin for help in the field and Rodolfo Noriega Trejo, R. E. Gongora Chi'n, and Rodolfo ColK Dfaz for help with plant identification. DUMAC, A.C., research field station, and personal from Celestiin Biosphere Reserve provided logistic support. The study was supported financially by El Colegio de la Frontera Sur (ECOSUR) and the National Council of Science (CONACyT Scholarship 225107 and SNl 21467). The study was conducted under research permit SGPA/ DGVS/03757/06. LITERATURE CITED Ahumada, J. a. 2001. Comparison of the reproductive biology of two neotropical wrens in an unpredictable environment in northeastern Colombia. Auk 1 18:191- 210. Anders, A. D. and D. M. R. Marshall. 2005. Increasing the accuracy of productivity and survival estimates in assessing landbird population status. Conservation Biology 19:66-74. Auer, S. K., R. D. Bassar, J. J. Fontaine, and T. E. Martin. 2007. Breeding biology of passerines in a subtropical montane forest in northwestern Argentina. Condor 109:321-333. Austad, S. N. and K. N. Rabenold. 1986. Demography and the evolution of cooperative breeding in the Bicolored Wren, Cainpylorhynchits griseu.s. Behaviour 97: 308-324. BirdLife International. 2008. Campy lorhynchus yiica- taniciis. In lUCN 2009. lUCN Red List of Threatened Species. Version 2009.2. . Bowman, G. B. and L. D. Harris. 1980. Effect of spatial heterogeneity on ground-nest predation. Journal of Wildlife Management 44:806-813. Brewer, D. 2001. Wrens, dippers, and thrashers: a guide to the wrens, dippers, and thrashers of the world. Yale University Press, New Haven, Connecticut, USA. CoLLlAS, N. E. and E. C. Collias. 1984. Nest building and bird behavior. Princeton University Press, Princeton, New Jersey, USA. CoMisibN Nacional de Areas Naturales Protegidas (CONANP). 2000. Programa de Manejo de la Reserva de la Biosfera Ria Celestiin. Mexico, D.F. Diario Oficial de la FederaciOn. 2002. Norma Oficial Mexicana NOM-059-ECOL-200I . Proteccion ambien- tal-Especies nativas de Mexico de flora y fauna silvestres-Categon'as de riesgo y espccificaciones para su inclusion, exclusion o cambio-Lista de especies en ric.sgo. Sccretan'a de Medio Ambiente y Recursos Naturales. Segunda Seccion del miercoles 6 de marzo dc 2002. Mexico, D.F. Donovan, T. M., F. R. Thompson 111. J. Faaborg, and J. R. Proust. 1995. Reproductive success of migratoiy Vargas-Soriano cl al. • YUCATAN WREN NES I ING SUCCESS 445 birds in habitat sources and sinks. Conservation Biology 9:1380-1395. Facemire, C. F., M. E. Fac'emire, and M. C. Facemire. 1990. Wind as a factor in the orientation of entrances of Cactus Wren nests. Condor 92:1073-1075. Farley, G. and J. Stuart. 1994. Atypical nesting sites of the Cactus Wren. Texas Journal of Science 46: 193- 195. FIansell, M. 2000. Bird nest and construction behaviour. Cambridge University Press, Cambridge, New York, USA. Hazler, K. R. 2004. Mayfield logistic regression: a practical approach for analysis of nest survival. Auk 121:707-716. Howell, S. and S. Webb. 1995. A field guide to the birds of Mexico and northern Central America. Oxford University Press, New York, USA. Jehle, G., a. a. Yackel Adams, J. A. Savidge, and S. K. Skagen. 2004. Nest survival estimation: a review of alternatives to the Mayfield estimator. Condor 106: 472^84. Johnson, D. H. and T. L. Shaffer. 1990. Estimating nest success: when Mayfield wins. Auk 107:595-600. Knadle, G. E., D. L. Brubaker, and K. S. Brubaker. 2001. Nest success is not an adequate comparative estimate of avian reproduction. Journal of Field Ornithology 72:527-536. Marquez-Valdelamar, L. 1998. Monografias de las especies mexicanas de la familia Troglodytidae (aves). Thesis. Facultad de Ciencias, Universidad Autonoma de Mexico. Mexico, D.F. Martin, T. E. 1993. Nest predation and nest sites: new perspectives on old patterns. Bioscience 43:523- 532. Martin, T. E. 1995. Avian life history evolution in relation to nest sites, nest predation, and food. Ecological Monographs 65:101-127. Martin, T. E. 1996. Life history evolution in tropical and south temperate birds: what do we really know? Journal of Avian Biology 27:263-272. Martin, T. E. 2004. Avian life-history evolution has an eminent past: does it have a bright future? Auk 121:289-301. Martin, T. E. and G. R. Geupel. 1993. Protocols for nest monitoring plots: locating nests, monitoring success, and measuring vegetation. Journal of Field Ornithol- ogy 64:507-5 19. Martin, T. E., R. D. Bassar, S. K. Bassar, J. J. Fontaine, P. Lloyd, H. A. Mathewson, A. M. Niklison, and A. Chalfoun. 2006. Life history and ecological correlates of geographic variation in egg and clutch mass among passerine species. Evolution 60:390- 398. Mayfield, H. F. 1961. Nesting success calculated from exposure. Wilson Bulletin 73:255-261. Mayfield, H. F. 1975. Suggestions for calculating nest success. Wilson Bulletin 87:456-466. Mezquida, E. T. 2004. Patrones de orientacion de los nidos de Passeriformes en una zona arida del Centro-Oeste de Argentina. Omitologfa Neotropical 15:145-153. Peterson, A. T., O. A. Flores-Villela, L. S. Leon- Paniagua, j. E. Llorhnte-Bousquets, M. a. Luis Martines, a. G. Navarro-Siguenza, M. G. Torres- ChAvez, and 1. Varga.s-Fernandez. 1993. Con.ser- vation priorities in Mexico: moving in the world. Biodiversity Letters 1:33-38. Peterson, R. T. and E. L. Chalif. 1989. Aves de Mexico, Guta de Campo. Diana, Mexico D.F. Piper. W. H. 1994. Courtship, population, nesting behavior, and brood parasitism in the Venezuelan Stripe-backed Wren. Condor 96:654-671. Powell, L. A. and M. G. Knutson. 2006. A productivity model for parasitized multibrooded song birds. Condor 108:292-300. Puga-Vazquez, I. L. 2008. Relacion de la anidacion de Uropsila leucogastra con factores ambientales en Calakmul, Campeche, Mexico. Thesis. Facultad de Ciencias Quhnico Biologicas, Universidad Autonoma de Campeche. San Francisco de Campeche, Cam- peche, Mexico. Rabenold, K. N. 1984. Cooperative enhancement of reproductive success in tropical wren societies. Ecology 65: 871-885. Rabenold, K. N. 1990. Helping and communal breeding in birds. Wilson Bulletin 102: 563-565. Ralph, C. J., G. R. Geupel, P. Pyle, T. E. Martin, D. F. Desante, and B. MilA. 1996. Manual de Metodos de Campo para el Monitoreo de Aves Terrestres. USDA, Forest Service, General Technical Report PSW-GTR- 159. Southwest Research Station, Albany, California. USA. Ricklefs, R. E. 1968. The survival rate of juvenile Cactus Wrens. Condor 70:388-389. Ricklefs, R. and G. Bloom. 1977. Components of avian breeding productivity. Auk 94:86-96. Robinson, T. R., W. D. Robinson, and E. C. Edwards. 2000. Breeding ecology and nest-site selection of Song Wrens in central Panama. Auk 1 17:345-354. Salgado-Ortiz, j., P. P. Marra, T. S. Sillett, and R. J. Robertson. 2008. Breeding ecology of the Mangrove Warbler {Dendroica petechia biyanti) and compara- tive life history of the Yellow Warbler subspecies complex. Auk 125:402-410. Simons, L. S. and T. E. Martin. 1990. Food limitation of avian reproduction: an experiment with the Cactus Wren. Ecology 71:869-876. Simons, L, S. and L. H. Simons. 1990. Experimental studies of nest-destroying behavior by Cactus Wrens. Condor 92:855-860. Skutch, a. F. 1985. Clutch size, nesting success, and predation on nests of neotropical birds, reviewed. Ornithological Monographs 36:575-594. Skutch, A. F. 2001. Life history of the Riverside Wren. Journal of Field Ornithology 72: 1-1 1 . Stutchbury, B. j. M. and E. S. Morton. 2001. Behavioral ecology of tropical birds. Academic Press. London, United Kingdom. Tarraoux, a. and R. McNeil. 2003. Influence of rain on the breeding and molting phenology of birds in a thorn woodland of northeastern Venezuela. Neotropical Ornithology 14:371-380. YAber, M. C. and K. N. Rabenold. 2002. Effects of 446 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 3. September 2010 sociality on short-distance, female-biased dispersal in tropical wrens. Journal of Animal Ecology 71:1042- 1055. Young, B. E., M. Kaspari, and T. E. Martin. 1990. Species-specific nest selection by birds in ant-acacia trees. Biotropica 22:310-315. Zimmerman, D. A. 1957. Some remarks on the behavior of the Yucatan Cactus Wren. Condor 59: 53-58. The Wilson Journal of Ornithology 1 22(3):447^54, 2010 BREEDING BIOLOGY AND NATURAL HISTORY OF THE SLATE-THROATED WHITESTART IN VENEZUELA ROMAN a. RUGGERA' -^ and THOMAS E. MARTIN' ABSTRACT. — We provide details on the breeding biology of the Slate-throated Whitestart (Myiobonts miniatus) from 126 nests found during seven breeding seasons, 2002-2008, at Yaeambii National Park, Venezuela. Nesting activity peaked in late April and May. Only the female built the nest and incubated the eggs. Males rarely visited the nest during these stages. Mean clutch size (2.1 ± 0.04 eggs, n = 93) was the smallest recorded for the Slate-throated Whitestart. Incubation and nestling period lengths were 15.3 ± 0.3 1 {n = 21 ) and 10.8 ± 0.24 (n = 7) days, respectively. Attentiveness (% of time on the nest) during incubation (59 ± 1.6%, n = 52) was similar to other tropical warblers and much lower than northern relatives. This caused a relatively low egg temperature (34.40 ± 0.33° C, n = 6 nests, 20 days) compared with north temperate birds. Both parents fed nestlings and increased their provisioning rates with nestling age. Growth rate based on nestling mass (k = 0.521 ± 0.015) was faster than for other tropical passerines but slower than northern relatives. Predation was the main cause of nesting failure and rate of predation increased with age of the nest. An estimated 15% of nests were successful based on an overall Mayfield daily predation rate of 0.053 ± 0.007. This study confirms a strong latitudinal variation in life history traits of warblers. Received 17 September 2009. Accepted 12 March 2010. New World wood warblers (Parulidae) are a small-sized and mostly insectivorous group of birds. Much work has been done on north temperate species, but resident tropical warblers are relatively poorly studied (but see Skutch 1945, 1954; Greeney et al. 2008; Cox and Martin 2009; Morales-Rozo et al. 2009). One of the most broadly distributed genera of this family is Myiobonts, which consists of 12 species (Sibley and Monroe 1990, Curson et al. 1994, Ridgely and Tudor 1994). The most widespread species of the genus is the Slate-throated Whitestart (M. minia- lus) which is a resident of mountain ranges between 700 and 2,500 m elevation in humid and wet forests (Ridgely and Tudor 1994). It occurs from Mexico to southern Bolivia, and unconfirmed data extend its distribution to northern Argentina (Di Giacomo 1995). It also is vagrant in southern Arizona, New Mexico, and west Texas (Dunn and Garrett 1997). However, the most northern breeding record is in the Sierra Madre Oriental, Mexico (McCormack et al. 2005). Slate-throated Whitestart populations are not threatened anywhere within their range and tolerate nesting in disturbed areas (Collins and Ryan 1994, Curson et al. 1994, Mumme 2010). ' uses Montana Cooperative Wildlife Re.search Unit. University of Montana, Missoula, MT 59812, USA. ^Current Address: CONICET-Instituto de Ecologi'a Regional (lER), Facultad de Cs. Naturales e Instituto M. Lillo. Universidad Nacional de Tucuman, C.C. 34, 4107 Yerba Buena, Tucuman, Argentina. ^Corresponding author; e-mail: raruggera@yahoo.com.ar For these reasons, the species is evaluated as one of Least Concern (lUCN 2009). The wide distribution of this species, which is almost equal to that of the entire genus, makes the Slate-throated Whitestart an interesting species for examining geographical variation. The Slate- throated Whitestart has great plumage and genetic variability along its range (Perez-Eman 2005). The 12 subspecies recognized show a dine in belly colors from red to orange and yellow from North to Central and South America (Curson et al. 1994). Yet, intraspecific and interspecific com- parisons in life-history traits of Myiohorus are scarce (Skutch 1945, 1954; Collins and Ryan 1994; Mumme 2010). M. m. ballux is the most common race in the Andes of Venezuela, Colombia, and part of Ecuador. Genetic information for our study site (Yacambu National Park, Venezuela) is lacking, but geographical proximity to Cubiro as well as the occurrence of some slight orange in the breast suggests M. m. ballux is also the race occurring in Yacambu (Phelps and Phelps 1950; J. L. Perez- Eman, pers. comm.). Our objective is to provide new information on the breeding biology of the Slate-throated White- start in a montane forest in northwest Venezuela. We compare our results with life-history traits recorded for other warblers, especially other Myiobonts species. METHODS Study Area. — This study was conducted during seven consecutive breeding seasons, from early March to early July in 2002 to 2008 at Yacambu 447 448 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 3. September 2010 National Park (09° 42' N, 69° 42' W) in Lara State, Venezuela. This is a montane forest consisting mainly of primary forest in the northern part of the Andes. The park ranges from 500 to 2,200 m, but our study plots were restricted to 1,350-2,000 m elevation (Fierro-Calderon and Martin 2007). The rainy season spans from mid- April to mid-July with a peak from May to June. Field Procedures. — Slate-throated Whitestart nests were found almost solely by observing parental behavior, especially during nest building when it is easiest to find them (Martin and Geupel 1993). Nests were checked every other day, except at stage-changing events (hatching or fledging), when they were checked every day or twice per day. Nest size was measured using a ruler with an accuracy of 0.01 cm for outer diameter (from edge to edge of nest) and height (exterior bottom-to-top of nest), and inner diam- eter (from edge to edge of cup) and height (bottom-to-top of cup). Eggs and nestlings were weighed with a portable electronic scale (ACCU- LAB, Elk Grove, IL, USA) with an accuracy of ±0.001 g. Mean egg mass was calculated only for eggs weighed between days 0 (last egg laid) and 2 of incubation. We measured the tarsus of nestlings using digital calipers (Mitutoyo, Kingsport, TN, USA) with an accuracy of 0.01 mm. We videotaped nests during incubation and nestling periods in 6-8 hr sessions, starting within 30 min of sunrise (Martin and Ghalambor 1999, Fierro-Calderon and Martin 2007). This allowed us to calculate parental attentiveness (percentage of time on nest for incubating or brooding), as well as on- and off-bout duration and parental provision rates. We sought to videotape incuba- tion in three periods to examine age-related changes in behaviors: early (2-3 days after the last egg was laid), middle (days 6-10) and late (2- 3 days before average hatching date) incubation. We recorded parental behavior in two stages during the nestling period: early (day 2-3 after hatch day) and pin break (day in which the eighth pin feather broke its sheath). The latter is possible because pin feather tips turn white the day before they break. Activity recorded in videotapes as well as observation of predation events during nest checks allowed us to identify some of the predators. We recorded egg temperatures by inserting thermisters on day 1 or 2 of the incubation period into one egg in a nest and the hole was sealed with glue (Weathers and Sullivan 1989). The wire was passed through the back of the nest and connected to a HOBO Stowaway XTI datalogger (Onset Corp., Bourne, MA, USA) that recorded temper- atures every 12-24 sec for 5-7 days per nest, unless the nest was depredated earlier (Martin et al. 2007). Statistical Analyses. — All means are reported with their standard errors (SE). Sample sizes are number of nests in all cases unless otherwise noted. Mean initiation date was calculated only for nests in which the exact egg laying date was known. Nesting season length was estimated as the number of days for the middle 90% of nests initiated (excluding the earliest and latest 5% of initiations) following Martin (2007). We calcu- lated the incubation period as the number of days between the last egg laid and the last egg hatched (Briskie and Sealy 1990, Martin 2002). Incubation periods were only calculated for nests in which the exact day the last egg was laid and exact hatch day were known. We quantified the nestling period as the number of days between the last egg hatching and the last nestling fledging, again based only on exact observations. We used analyses of variance (ANOVA) with significance at a = 0.05 to test changes in parental behavior during incubation and nestling stages, and used LSD post-hoc tests among periods. Nestling growth curves were calculated following Remes and Martin (2002). We included measurements of mass and tarsus length for nestlings in which the exact age was known for these analyses. Mortality and survival analyses of nests were calculated considering the elapsed time of the observations to allow inclusion of nests found in different stages following Mayfield (1975). Nests that failed because of experimental activities were excluded from estimation of daily survival or mortality. RESULTS Nest Description. — The highest nest located was at an elevation of 1,674 ± 5 m asl; we did not see or hear any Slate-throated Whitestarts above this elevation even though we worked to 2,000 m. Nests were in small nooks on banks or slopes and often built in open or disturbed areas- including road banks or human-made trails inside the forest. However, nests were also found on slopes within undisturbed forest. Nests were highly cryptic because they were usually hidden by surrounding shrubs, ferns, or grasses. Only the female was observed going to and from the nest (mainly in Riiggcni and Martin • BREEDING BIOL()C}Y Ol- A TROPICAL WARBLER 449 T3 B ro c B 00 CD c o c/3 CD JD E ID FIG. 1. Numbers of nests initiated (i.e., first egg laid) each week beginning 1 March for Slate-throated White- starts at Yacambii National Park, Venezuela during 2002- 2008, Only nests for which initiation date was observed are included (/; = 106). morning hrs) to build the nest. The male was nearby in the surrounding trees, usually singing and hunting insects. The first nests of the season were often built over 2 to 4 weeks in a rhythm that started with several nest-building visits per day that slowly increased over time to much more active building. Nests of later attempts were built over 3 to 4 days in much more active behavior. The nest was a dome-shaped construction entirely built with dry material including pieces of grass, several types of dry leaves, roots, and small twigs. The lining consisted of fine and soft dry fibers. The entrance was circular to elliptical (wider than higher). Nest orientation at 73 nests spanned from 1 to 334° with no orientation preferences. The internal dimensions of the cup averaged 49.1 ± 4.7 mm in diameter and 31.4 ± 9.5 mm in height (n = 66). The outer dimensions of the nest averaged 103.9 ±21.2 mm and 64.8 ± 25.0 mm for diameter and height, respectively. Nesting Chronology. — The nesting period of Slate-throated Whitestarts in Yacambu spanned from late March to late June based on 126 nests over seven breeding seasons. The earliest and latest nests across all years of study were initiated on 26 March and 23 June, respectively. Both nests were found before the first egg was laid, yielding precise documentation of these extremes, but even back-calculating nests that were already initiated when found did not exceed the.se extremes. The mean initiation date was 7 May (± 2.2 days, n = 106) and highest nesting activity occuired at the end of April and through May (Fig. 1 ). Re-nesting activities were observed within 6- 8 days after a nest failed and within a 15-m radius of the previous nest (// = 10). Banded individuals allowed confirmation that it was a re-nesting attempt by a particular pair of individuals. The nesting season of the Slate-throated Whitestart in Yacambii was 77 days, but this species appears to be single brooded; we did not observe any breeding activity of the same pair in the same season after a successful brood. Clutch Size and Eggs. — Mean clutch size for the Slate-throated Whitestart was 2.1 ± 0.04 eggs (/? = 93). We recorded one instance of a single egg clutch (1%), 15 nests had three eggs (16%), and the rest had two eggs (83%). Three-egg clutches did not exceed 25% of the clutches in any year, but three-egg clutches were observed in all years except 2004 (n = 14 in 2004). Eggs were laid on consecutive days and generally early in the morning (i.e., before 0800 hrs). The eggs were elliptical to sub-elliptical and the shell was dull white with iiTegularly distrib- uted dark red to brown spots. Mean egg mass was 1.61 ± 0.01 g (n = 152 eggs) and ranged from 1.20 to 1.94 g. The mean egg mass represented 19% of adult female body mass (10.1 ± 0. 15 g, /? = 20 females). Incubation Period. — Incubation began when the last egg was laid, but nest attentiveness was highly variable on that day. Eggs hatched synchronously in all cases. Only the female incubated, and the male rarely visited the nest to feed the female. The male did not visit during 6- 8 hrs of video observations at 31 of 52 nests sampled during incubation. Overall, the male averaged 0. 1 1 ± 0.02 trips/hr to the nest (n = 52) during incubation. The incubation period of the Slate-throated Whitestart averaged 15.3 ± 0.31 days (n = 21) varying from 14 to 18 days. Overall, nest attentiveness (percent of time spent on the nest incubating) during this stage averaged 59 ± 1.6% in = 52). Nest attentiveness increased (E2.49 = 15.8, P < 0.001) from early incubation to mid- incubation but did not continue to increase in late incubation (Eig. 2A). The increase in nest atten- tiveness was caused by reduction in duration of off-bouts with increasing age of the embryo (Fig. 2B; ^249 = 5.4, P = 0.008). Duration of on-bouts did not change (^2,49 = 1.49, P = 0.17) with age of the embryo (Fig. 2B). On- and off- bouts averaged 33.4 ± 1.63 min and 22.8 ± 2.65 min (// = 52 in both cases), respectively. Temperature of Slate-throated Whitestart eggs averaged 34.40 ± 0.33° C over 24-hr periods (/? = 450 THE WILSON JOURNAL OL ORNITHOLOGY • Voi 122, No. 3, September 2010 Incubation period FIG. 2. Mean ± SE for (A) nest attentiveness, and (B) on- and off-bout durations across three periods of incubation: early (days 1-5), middle (days 6-10), and late (days 11-17) for Slate-throated Whitestarts at Yacambti National Park, Venezuela. Sample sizes reflect numbers of nests. Different letters among periods indicate significant differences (a = 0.05) based on LSD tests of each bout type (on- vs. off-bouts) separately. 6 ne.sts, 20 days of sampling) based on an overall mean taken from the means of each nest. Nestling Period. — Hatchlings had sparse natal down with pink-orange skin, and were blind. The eighth primary feather broke its sheath on days 6 or 7. Brooding attentiveness was by females alone and averaged 63 ± 2.9% {n = 10) at the beginning of the nestling period (days 0-3 of age) and decreased dramatically (/^u.s == 92.6, P < 0.001 ) to 1 1 ± 4.9% (/? = 7) at pin break. Both females and males provisioned the nestlings at an average rate of 3.77 ± 0.44 trips/hr (// = 10) early in the nestling period (days 0-3), but provisioning rate markedly increased (F\,\5 — 25.7, P < 0.001 ) to 9.41 ± 1.18 trips/hr (// = 7) at pin break. Nestlings fledged in 10.8 ± 0.24 days in = 1) after hatching, ranging from 10 to 11.5 days. Growth rate ba.sed on nestling mass (Fig. 3A) was greater than when based on tarsus length (Fig. 3B). Nestling weights on the last 2 days in 12 ■ A 10 ■ O) C/) c/) cc 8 - 6 - 4 • 2 - 0 ■■ I I I /c= 0.521 +0.015 T 1 I 18 E E 15 cn 12 ■ c o c/) 9 3 CO ,C0 6 B .1 !i M' I I /c= 0.334 + 0.016 0 2 4 6 8 10 12 Nestling age (day) FIG. 3. (A) Mass and (B) tarsus length measurements of Slate-throated Whitestart nestlings at Yacambu National Park, Venezuela plotted against age. Growth rate constant (k) is indicated in each graph. the nest (days 10 and 1 1 ) averaged 10.78 ± 0.22 g (// = 5 fledglings) compared with mean adult mass (10.1 ± 0. 15 g for the female [n = 20] and 10.1 ± 0.1 1 g for the male [/? = 36]). Nest Survival. — Eighteen of 126 nests were excluded from nest survival analyses because of disruption by experimental activities or because they were abandoned before egg-laying. Overall daily mortality rate for the remaining nests was 0.065 ± 0.008 in = 108 nests; 1,084.5 exposure days) and only 15% of the nests were successful (ba,sed on a total nesting period of 30.2 days). Overall daily nest predation rate was 0.053 ± ().()()7. Daily nest predation rates increased with age of the nest from 0.037 ± 0.021 (/? = 81.5 exposure days) during egg-laying to 0.052 ± ().()()8 in = 736 exposure days) during incubation to 0.060 ± 0.015 in = 267.0 exposure days) during the nestling stage. Predators included mostly birds (e.g., small raptors, jays, and woodwrens) but also small mammals (e.g., squirrels and tayra [Eira harhara]). Rtiggcra amt Martin • BREEDING BIOIX3GY OE A TROPICAL. WARBLER 451 DISCUSSION The nesting season for Slate-throated White- start in Yacambil (late Mar to late Jiin) was almost identical to reports elsewhere in Venezuela (Collins and Ryan 1994), and Costa Rica (Skutch 1954, Mumme 2010). In general, the highly seasonal breeding activity began at the end of the dry season and continued through the rainy season. Behavior during nest building was consistent with previous observations reported by Skutch (1945, 1954) and Mumme (2010) in Central America, and for other Myioborus species (e.g., M. pictiis [Marshall and Baida 1974] and M. flaviverte.x [Morales-Rozo 2009]). Characteristics of Slate-throated Whitestart nests appear to be highly conservative along the broad range of the species. Slate-throated Whitestart nests in Ya- cambu were in the same types of places reported from Central America (Skutch 1954, McCormack et al. 2005, Mumme 2010) and Venezuela (Ewert 1975, Collins and Ryan 1994). Slate-throated Whitestart nests were quite similar to some other members of the genus, including M. melanoce- phalus (Greeney et al. 2008), M. torquatus (Skutch 1954), M. bnmniceps (Auer et al. 2007), and M. flavivertex (Morales-Rozo et al. 2009). However, some tropical Myioborus species, such as M. ornatiis, apparently build open cup nests (Curson et al. 1994; O. Cortes, pers. comm.). Painted Redstart (M. pictiis), the most northern species of the group, only occasionally builds an enclosed nest, and more often makes an open cup with the roof substituted by natural overhangs such as rocks, logs, and pine needles (Marshall and Baida 1974). Open-cup nests are common for warblers outside of the tropics. The placement of nests in small holes in the ground is also typical of some North American parulids, such as ground- nesting Vennivora, which do not build a roof (Martin 1993). Clutch size for Slate-throated Whitestart at our study site was relatively constant with 83% of nests containing two eggs. The mean clutch size in Yacambti (2.1 eggs) was relatively small compared with that of M. m. pallidiventris only 250 km from our .study site, which averaged 2.7 eggs based on six nests (Collins and Ryan 1994). This difference may reflect sample size issues, but it is notable that four of their six nests (67%) had three eggs, while only 15 of 93 nests (16%) had three eggs in our study. Moreover, we did not find more than 25% of nests with three eggs in any individual year. Other Myioborus species near the equator also had two-egg clutches. M. melanoce- phalus in Ecuador had two eggs in all of .seven nests (Greeney et al. 2008), and M. flavivertex in Colombia had two eggs in the only nest ob.served (Morales-Rozo et al. 2009). Basileuterus tristria- tus in Yacambu had a mean clutch size of 1.96 eggs and only one of 96 clutches had three eggs (Cox and Martin 2009). Thus, two eggs near the equator may be more typical of warblers, but the frequency of occurrence of three eggs needs further study in Myioborus. Variation in clutch size among latitudes is clear. M. m. aurantiacus in Costa Rica usually laid three eggs with mean clutch sizes of 2.84 eggs from 13 nests (Skutch 1954) and 2.89 from 82 nests (Mumme 2010). The only nest observed for M. m. hellmayri in Guatemala had three eggs (Skutch 1954), M. bnmniceps in Argentina had a mean clutch size of 2.6 eggs (Auer et al. 2007), while M. pictus in Arizona had a mean clutch size of 3.7 eggs (Marshall and Baida 1974). In contrast, warblers in the north temperate region typically have mean clutch sizes of 4-5 eggs (Martin 1988). Thus, tropical warblers have a strong reduction in clutch size compared to north temperate warblers, but variation within the tropics and subtropics both in Slate-throated Whitestart races and among congeners also appears evident. Incubation behavior and duration (15.3 days) of the Slate-throated Whitestart in Yacambu seem to be typical tor tropical and subtropical warblers (see Cox and Martin 2009) and differ substantially from northern relatives. The mean incubation period was about a day longer than for other .SLibspecies (Skutch 1954, Collins and Ryan 1994. Mumme 2010). However, other whitestart species and most tropical warblers have incubation periods ranging from 14 to 16 days (Martin et al. 2007, Cox and Martin 2009), and our results fit in the middle of this range. Longer incubation periods in the tropics and subtropics, compared to northern species, have been attributed to lower nest attentiveness which causes colder incubation temperatures (Martin 2002, Martin et al. 2007. Martin and Schwabl 2008, Londono 2009). Nest attentiveness of Slate-throated Whitestarts at our study site was relatively low hut similar to other tropical and southern warblers (Cox and Martin 2009). Skutch (1954) studied one nest of Slate- throated Whitestart in Guatemala during middle 452 THE WILSON JOURNAL OF ORNITHOLOGY • Voi 122, No. 3. September 2010 incubation, and both nest attentiveness (67.4%) and duration of bouts (on-bouts = 37.6 min; off- bouts = 18.2 min) were similar to our observa- tions for mid-incubation (Fig. 2). Length of the nestling period (10-1 1 days) had low variation within Yacambu and was similar to that observed for Slate-throated Whitestarts in other localities (Skutch 1954, Collins and Ryan 1994, Mumme 2010). Nestling periods in other species of Myioboriis are similar to Slate-throated Whitestarts (Skutch 1954, Marshall and Baida 1974, Auer et al. 2007, Greeney et al. 2008). Most warblers have a nestling period within a range of 9-13 days (Cox and Martin 2009) even when including north temperate and tropical species (Elliott 1969, Martin 1995, Remes and Martin 2002, Martinez et al. 2004, Di Giacomo 2005, Auer et al. 2007, Cox and Martin 2009). The duration of the nestling period appears to be a more conservative life-history trait within Paruli- dae than incubation period. Slate-throated Whitestarts in Costa Rica made 20.3 provisioning trips/hr, on average, during nestling ages 5-9 days (Mumme 2010), which is twice the rate that we observed (9.41 trips/hr) for this species in Yacambu on days 6-7. Despite the much higher provisioning rate, nestlings averaged the same weight on days 7-8 in both places: 9.5 g in Costa Rica (Mumme 2010) and 9.5 ± 0.17g {n = 24) in Venezuela (this study). The larger clutch size in Costa Rica (Mumme 2010) may account for the higher provisioning rate and similar nestling size compared with our site. The larger clutch size and similarly sized nestlings with higher parental provisioning rates in Costa Rica (Mumme 2010) compared with our site may suggest that food is more limiting at our site for this species. Growth rate estimates are scarce for tropical warblers. One estimate based on limited data for the Slate-throated Whitestart in Venezuela (Col- lins and Ryan 1994) found that nestlings grew at a rate (k = 0.522) based on mass almost identical to our estimate from Yacambu (k = 0.521, Fig. 3). Generally, tropical passerines grow more slowly than species in temperate regions (Ricklefs 1976, Cox and Martin 2009). Our estimated growth rate (Fig. 3) was slower than estimates for north temperate warblers (i.e., Remes and Martin 2002, Cox and Martin 2009), but faster than for many other tropical passerines (Ricklefs 1976). This relatively fast growth of nestlings given slow embryonic development (i.e., long incubation periods) demonstrates that developmental rates in embryonic and post-natal stages are not constrained by each other. Nest design is thought to be an important factor influencing predation rate and failure from weather (Collias and Collias 1984, Auer et al. 2007). The last may be especially important in tropical and subtropical montane forests where rains during the breeding season are quite intense. Enclosed nests in these environments may afford nest contents protection from rain (Collias and Collias 1984) as well as predation (Auer et al. 2007). The daily predation rate of Slate-throated Whitestarts in Yacambu was higher than the average for all non-cavity-nesting species studied at this site (0.041 ± 0.003, n = 38 spp.; TEM, unpubl. data). This average was remarkably similar to the average for 10 non-cavity-nesting species studied in Panama (Robinson et al. 2000). Thus, the domed ground nests of this warbler in Yacambu are readily found by predators. The Slate-throated Whitestart in Yacambu had a relatively high daily predation rate, small clutch size, long incubation period, and low nest attentiveness during incubation, typical of many tropical passerines (Skutch 1954; Martin et al. 2000, 2007; Cox and Martin 2009). Warblers appear to have strong divergences among relatives across geographic space, making them an impor- tant group for studying latitudinal changes in life- history strategies. Some traits, such as clutch size seem to have geographic variation even within the tropics within this species and genus. Other features of Slate-throated Whitestart nesting behavior, including nest construction and place- ment, nestling period length, and growth rate are relatively conservative. Focusing effort on par- ulids in tropical latitudes in particular could strengthen many aspects of life history theory. ACKNOWLEDGMENTS We thank Carlo.s Bosque for substantial logistical support in aiding this work. We also thank many assistants for valuable help in the field. We are grateful to Amy Stokes for help with data summary and T. E. Martin’s and Pedro Blendinger’s laboratories for helpful comments on the manuscript. We are also grateful to J. L. Perez-Eman, J. G. Blake, and two anonymous reviewers for valuable com- ments and improvements to the manuscript. This study was made possible in part by support under NSF grants DEB- 9981527, DEB-0543178, and DEB-0841764 to T. E. Martin. Permit numbers are DM/0000237 from EONACIT, PA-lNP-005-2004 from INPARQUES, and 01-03-03-1 147 from Ministerio del Ambiente. Any use of trade names is to Riiggeni aiul Mcirtiii • BREEDING BIOLOGY OF A TROPICAL WARBLER 453 aid method descriptions only and does not imply endorse- ment by the U.S. Government. LITERATURE CITED Auer, S. K., R. D. B.'Vssar, J. J. Fontaine, and T. E. Martin. 2007. Breeding biology of passerines in a subtropical montane forest in northwestern Argentina. Condor 109:321-333. Briskie, j. and S. G. Sealy. 1990. Evolution of short incubation periods in the parasitic cowbirds, Molo- thnis spp. Auk 107:789-793. CoLLiAS, N. E. and E. Collias. 1984. Nest building and bird behavior. Princeton University Press, Princeton, New Jersey, USA. Collins, C. T. and T. P. Ryan. 1994. Notes on breeding biology of the Slate-throated Redstart {Myioboms miniatiis) in Venezuela. Ornitologt'a Neotropical 5:125-128. Cox, W. A. and T. E. Martin. 2009. Breeding biology of the Three-Striped Warbler in Venezuela. Wilson Journal of Ornithology 121:667-678. CuRSON, J., D. Quinn, and D. Beadle. 1994. Warblers of the Americas: an identification guide. Houghton- Mifflin, New York, USA. Di Giacomo, A. G. 1995. Dos especies nuevas para la avifauna argentina. Hornero 14:77-78. Di Giacomo, A. G. 2005. Aves de la Reserva El Bagual. Pages 201^66 in Historia natural y paisaje de la Reserva El Bagual, Provincia de Formosa, Argentina. Inventario de la fauna de vertebrados y de la flora vascular de un area protegida del Chaco Humedo (A. G. Di Giacomo and S. F. Krapovickas, Editores). Temas de Naturaleza y Conservacion 4. Aves Argentinas/Asociacion Omitologica del Plata, Buenos Aires, Argentina. Dunn, J. L. and K. L. Garrett. 1997. A field guide to the warblers of America. Houghton-Mifflin, Boston, Massachusetts, USA. Elliott, B. G. 1969. Life history of the Red Warbler. Wilson Bulletin 81:184-195. Ewert, D. 1975. Notes on nests of four avian species from the Coastal Cordillera of Venezuela. Wilson Bulletin 87:106. Fierro-CalderOn, K. and T. E. Martin. 2007. Repro- ductive biology of the Violet-chested Hummingbird in Venezuela and comparisons with other tropical and temperate hummingbirds. Condor 109:680-685. Greeney, H. F., P. R. Martin, R. C. Dobbs, R. A. Gelis, A. BOcker, and H. Montag. 2008. Nesting ecology of the Spectacled Whitestart in Ecuador. Ornitologt'a Neotropical 19:335-344. lUCN. 2009. lUCN Red List of threatened species. Version 2009.1. www.iucnredlist.org. Londono, G. a. 2009. Eggs, nests, and incubation behavior of the Moustached Wren (Thryothorus geniharhis) in Manu National Park, Peru. Wilson Journal of Orni- thology 121: 623-627. Marshall, J. and R. P. Balda. 1974. The breeding ecology of the Painted Redstart. Condor 76:89-101. Martin, T. E. 1988. Nest placement: implications for selected life history traits, with special reference to clutch size. American Naturalist 132:900-910. Martin, T. E. 1993. Nest predation and nest sites: new perspectives on old patterns. BioScience 43:523-532. Martin, T. E. 1995. Avian life history evolution in relation to nest sites, nest predation and food. Ecological Monographs 65:101-127. Martin, T. E. 2002. A new view of avian life-history evolution tested on an incubation paradox. Proceed- ings of the Royal Society of London, Series B 269:309-316. Martin, T. E. 2007. Climate correlates of 20 years of trophic changes in a high-elevation riparian system. Ecology 88:367-380. Martin, T. E. and G. R. Geupel. 1993. Nest-monitoring plots-methods for locating nests and monitoring success. Journal of Field Ornithology 64:507-519. Martin, T. E. and C. K. Ghalambor. 1999. Males helping females during incubation. I. Required by microcli- mate or constrained by nest predation? American Naturalist 153:131-139. Martin, T. E. and H. Schwabl. 2008. Variation in mater- nal effects and embryonic development rates among passerine species. Philosophical Transactions of the Royal Society of London, Series B 363:1663-1674. Martin, T. E., P. R. Martin, C. R. Olson, B. J. Heidinger, and j. j. Fontaine. 2000. Parental care and clutch sizes in North and South American birds. Science 287:1482-1485. Martin, T. E., S. K. Auer, R. D. Bassar, A. M. Niklison, AND P. Lloyd. 2007. Geographic variation in avian incubation periods and parental influences on embry- onic temperature. Evolution 61:2558-2569. Martinez, W. E., V. D. Piaskowski, and M. Teul. 2004. Reproductive biology of the Gray-crowned Yellow- throat (Geothlypis poliocephala palpebralis) in central Belize. Ornitologia Neotropical 15:155-162. Mayfield, H. F. 1975. Suggestions for calculating nest success. Wilson Bulletin 87:456-466. McCormack, J. E., G. Castaneda Guayas.amin, B. Mila, and F. Heredia Pineda. 2005. Slate-throated Red- starts (Myioboms miniatus) breeding in Maderas del Carmen, Coahuila, Mexico. Southwestern Naturalist 50:501-503. Morales-Rozo, a., E. Rodriguez-Ortiz, B. Freeman, C. A. Olaciregui, and C. D. Cadena. 2009. Notas sobre el nido y los pichones del Abanico Colombiano (Myioboms flavivertex: Parulidae). Ornitologia Neo- tropical 20: 113-119. Mumme, R. L. 2010. Breeding biology and nesting success of the Slate-throated Whitestart (Myioboms miniatus) in Monteverde, Costa Rica. Wilson Journal of Ornithology 122: 29-38. P£rez-Eman, j. L. 2005. Molecular phylogenetics and biogeography of the neotropical redstarts (Myioboms-, Aves, Parulinae). Molecular Phylogenetics and Evo- lution 37:51 1-528. Phelps, W, H. and W. H. Phelps Jr. 1950. Lista de las aves de Venezuela con su distribucion. Parte 2. Passeriformes. Boletin de la Sociedad Venezolana de Ciencias Naturales 12:1^27. 454 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 3. September 2010 RemeS, V. AND T. E. Martin. 2002. Environmental influences on the evolution of growth and develop- mental rates in passerines. Evolution 56: 2505-2518. Ricklefs, R. E. 1976. Growth rates of birds in the humid New World tropics. Ibis 118:179-207. Ridgely, R. S. and G. Tudor. 1994. The birds of South America. Volume 2. University of Texas Press, Austin, USA. Robinson, W. D., T. R. Robinson, S. K. Robinson, and J. D. Brawn. 2000. Nesting success of understory forest birds in central Panama. Journal of Avian Biology 31: 151-164. Sibley, C. G. and B. L. Monroe Jr. 1990. Distribution and taxonomy of birds of the world. Yale University Press, New Haven, Connecticut, USA. Skutch, a. E. 1945. Studies of Central American redstarts. Wilson Bulletin 57:216-234. Skutch, A. F. 1954. Life histories of Central American birds: Families Fringillidae, Thraupidae, Icteridae, Parulidae and Coerebidae. Cooper Ornithological Society, Berkeley, California, USA. Weathers, W. W. and K. A. Sullivan. 1989. Nest attentiveness and egg temperature in the Yellow-eyed Junco. Condor 91: 628-633. The IVilsun Journal oj Ornilholof'v 1 22(3):455-464, 2010 REPRODUCTIVE BIOLOGY OF A GRASSLAND SONGBIRD COMMUNITY IN NORTHCENTRAL MONTANA STEPHANIE L. JONES, J. SCOTT DIENI,' AND PAULA J. GOUSE'" ABSTRACT. — Successful conservation of grassland bird populations requires basic information on their breeding biology; in particular, information from undisturbed native prairie over an extended period of time. We pre.sent data collected at Bowdoin National Wildlife Refuge in northcentral Montana on the reproductive biology of six grassland bird species that breed in mixed-grass prairie: Sprague’s Pipit (Anthiis spragneii). Savannah (Passercuhts sandwichensis). Grasshopper (Ammod ramus savannariim), and Baird’s (A. hairdii] sparrows, Chestnut-collared Longspur (Calcarius ornatus), and Western Meadowlark (Stiirnella neglecta). Basic measures of reproductive biology are presented, including estimates of nest density, phenology, fecundity, parasitism rates, and nest success, and how these and other characteristics varied across years. Nests {n = 1,494) of the six focal species accounted for 98% of all passerine nests found during 1997- 2007; Chestnut-collared Longspurs (51%) were the dominant breeding species. Total nest density across years ranged from 20 to 41 nests per 40 ha (CV = 26%) on unburned sites. Mean clutch initiation date and clutch size varied little across years; however, clutch size tended to decrease over the course of a season, regardless of bird species. Daily nest survival rates did not differ markedly among bird species, but did vary substantially among years, suggesting that year-dependent factors were affecting nest success among all species similarly. Received 20 December 2008. Accepted 28 February 2010. Declines in grassland bird populations have been consistently greater and geographically more widespread than in any other avifaunal group in North America (Knopf 1996, Askins et al. 2007, Sauer et al. 2008), which has been particularly evident for breeding endemics of the northern mixed-grass prairie (Knopf 1996). This has been attributed to many factors, including habitat conversion and fragmentation (Askins et al. 2007), woody vegetation encroachment (Houston and Schmutz 1999), removal of native grazers (Knopf 1994, 1996) and fire (Higgins 1984), and increased predation and parasitism (Basore et al. 1986, Davis and Sealy 2000). Conservation of grassland bird populations requires basic information on their breeding biology, especially data from undisturbed native prairie over an extended period of time. Large expansive areas of shortgrass and mixed-grass prairies still exist in the western portion of the Great Plains, including northcentral Montana, where these grasslands are relatively intact. In addition, northern prairie grasslands are funda- ' U.S. Fish and Wildlife Service, P. O. Box 25486 DFC, Denver, CO 80225, USA. “Redstart Consulting, 403 Deer Road, Evergreen, CO 80439, USA. ^U.S. Fish and Wildlife Service, Bowdoin National Wildlife Refuge, Malta, MT 59538, USA. ■‘Current address: U.S. Fish and Wildlife Service, Charles M. Russell National Wildlife Refuge, P. O. Box 1 10, Lewiston, MT 59457, USA. ’Corresponding author; e-mail: stephanie_jones(§>fws.gov mentally dynamic. Local temporal variability in grassland conditions are largely a function of weather, which can influence plant community structure and composition, both within- and between seasons (Igl and Johnson 1999, Winter et al. 2005, Askins et al. 2007). Fluctuations in habitat conditions result in notable shifts in local bird population densities (Igl and Johnson 1999, Winter et al. 2005) as many native grassland bird species adapt to shifting habitat conditions with nomadic behavior (Clark and Shutler 1999, Ahlering and Johnson 2006, Jones et al. 2007). Consequently, long-term studies focusing on the same population can be instrumental to under- standing how reproductive characteristics vary over time. We studied six species, two of which, Sprague’s Pipit {Anthus spragueii) and Baird’s Sparrow (Ammodramus hairdii) are breeding endemics to the northern mixed-grass prairie, and a third. Chestnut-collared Longspur (Calcarius ornatus), breeds only in the short to mixed-grass prairie of the western and northern Great Plains. These species have been the subject of recent studies on their reproductive biology (Davis and Sealy 1998, Winter 1999, Davis 2003), and on habitat and management effects (Madden et al. 1999. Davis 2005). However, few studies have occuired in Montana (Hill and Gould 1997, Robbins and Dale 1999, Green et al. 2002), or over an extended period of years. Savannah (Passercuhts .sandwi- chensis), and Grasshopper (Annnodramus .savan- na rum) sparrows, and Western Meadowlarks (Sturnella neglecta) have been well-studied 455 456 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 3, September 2010 throughout much of their range, but relatively few data for these species are available from the northern mixed-grass prairie (Wheelwright and Rising 1993, Lanyon 1994, Vickery 1996). We present 1 1 years of data on the reproductive biology of six grassland bird species that breed in mixed-grass prairie of northcentral Montana: Sprague’s Pipit; Savannah, Grasshopper, and Baird’s sparrows; Chestnut-collared Longspur; and Western Meadowlark. Our objectives were to characterize: ( 1 ) songbird community structure, (2) nesting phenology, (3) fecundity, (4) brood para- sitism rates and nest success, and (5) to describe how these characteristics varied across years. METHODS Study Area. — Our study was conducted during 1997-2007 at Bowdoin National Wildlife Refuge (NWR) in Phillips County, near the town of Malta in northcentral Montana (48° 24' N, 107° 39' W; —750 m asl). The study area consisted of five permanent sites (26-59 ha) totaling 218 ha of flat to gently rolling native mixed-grass prairie (Fig. 1). Data for this analysis were collected at four sites beginning in 1997. One of these sites partially (3 ha) burned in a wildfire in 1994, and was then prescribed burned during spring 2004; a second site was prescribed burned in spring 2007 and was not monitored during 2004. A fifth site, which had been prescribed burned during spring 2000, was added in 2001 . No other burning events have occurred on the study sites since refuge documentation began in 1936. Grazing by cattle was gradually phased out during 1973-1977. The climate is continental and semiarid, characterized by strong winds and high evapora- tion rates. Mean long-term annual and seasonal (May-Jul) precipitation totals are 34 and 18 cm, respectively (National Climatic Data Center 2007). Mild drought conditions prevailed during the study period; mean annual and seasonal precipitation totals were 32 and 14 cm, respec- tively (National Climatic Data Center 2007). The plant community was dominated by western wheatgrass {Pascopynim smithii), needle-and- thread (Stipa coinata), blue grama (Bouteloua gracilis), and clubmoss (Selaginella densa). Invasive plant species were isolated and sparse. Shrubs (Sarcobatus vermicidatus, Artemisia cana, Ceratoides lanata) were sparse, and trees absent, except sporadically outside the periphery of two study sites. Photographic records at Sprague’s Pipit nests (u = 5) documented Northern Harrier {Circus cycmeus), garter snake (Tliamnopliis spp.). West- ern Meadowlark, and deer mice (Peromyscus spp.) preying upon nestlings during this study (S. L. Jones, unpubl. data). Other potential nest predators have been observed on, or within, the immediate vicinity of our study sites, including: Jones el at. • NESTING OE GRASSLAND SONGBIRDS 457 American badger (Taxidea taxiis), long-tailed weasel (Mnstela fremita), red fox (Vidpes vulpes), coyote {Canis latnms), mice and voles (Zapus, Reithrodontomys, and Microtiis), Richardson’s ground squirrel (Spermophihis ricluirdsonii), bull- snake (Pituopliis melanoleucus), prairie rattle- snake (Crotcdus viridis), gulls {Lcinis spp.), Short- eared Owl (Asia flanimeiis). Loggerhead Shrike {Lanins ludovicianns), and Black-billed Magpie {Pica luidsonia). Nest Searching. — Sites were searched for nests 3-5 times per week from early May through late July in an attempt to locate all active nests each year. Search techniques (% successful) included behavioral observations (3%) (Martin and Geupel 1993), opportunistic foot flushes (36%), and rope dragging (61%) (Davis 2003, Dieni and Jones 2003). Nests were marked for relocation by placing a discrete strip of plastic flagging on the ground ~2.5 m on either side of the nest. Nests were monitored every 2-A days until the nesting outcome was ascertained. Nesting stage at initial discovery was assigned by candling eggs (Loke- moen and Koford 1996) or estimating age of nestlings (Jongsomjit et al. 2007). Nesting out- comes were: (1) successfully fledged at least one young of the parental species, (2) depredated, (3) abandoned (eggs or nestlings left permanently unattended), or (4) unknown. Observations of fledglings within 3 days of expected fledging, the presence of feces and feather scales in the nest, fledglings near the nest, and adults uttering alarm calls nearby or feeding new fledglings within 50 m of the nest were treated as evidence of reproduc- tive success. Predation was assumed when the nest, eggs or nestlings too young to fledge disappeared or were destroyed. Statistical Analysis. — Number of nests located for each focal bird species was summarized only for the three sites which had not been disturbed by prescribed fire to quantify annual variation of nesting density. Site effects were ignored given the relatively close spatial proximity of the study sites. The nest was the primary sample unit for the analysis of reproductive parameters and nest success. Clutch initiation dates (CID) were estimated by calculating the date of clutch initiation from egg or nestling age or from known dates of major events (e.g., hatching). Reproduc- tive parameters are reported using standard measures of central tendency. Data variation was summarized using the coefficient of variation (CV = SD/mean; Zar 1999); annual variation for many parameters was summarized only for the domi- nant species with adequate sample sizes. Trend lines (across years) were fit using the method of least squares; strength of trends was characterized using simple beta {h) and correlation (Pearson r) coefficients. We estimated nest daily survival rates (DSR) using the survival model in program MARK (White 2008) for the six focal bird species, and to examine how DSR varied temporally across years (1997-2007). Program MARK uses a generalized linear approach to modeling DSR and maximum likelihood estimation to derive model coefficients and sampling variances (Dinsmore et al. 2002, Rotella et al. 2004). Regression models were constructed using the logistic transformation (logit) as the link function and natural logs [ln(DSR/l-DSR)]. Bird species and year were treated as nominal-scale variables (6 and 11 levels, respectively) with each level introduced into the regression model as an artificial explan- atory variable and 0 or 1 coding; second-order interactions between species and year were added with the appropriate cross-product terms. One candidate model set was evaluated within an information-theoretic framework (Burnham and Anderson 2002). Program MARK was used to calculate Akaike’s Information Criterion coirect- ed for small sample sizes (AIC,.), ranking the fit of each model in ascending order of AIC^. values. Estimates for each species and year were calculated using the model-averaging approach (Burnham and Anderson 2002) weighted accord- ing to that model’s likelihood in the set (normal- ized Akaike weight). The intercept-only model (where the regression coefficient was set to zero) was included in a set with the covariate of interest (species or year), which served to reduce model- selection bias of the estimate (Burnham and Anderson 2002). A point estimate of nest success was calculated as DSR raised to the mean number of days of the nesting period (incubation and nestling stages combined) for each species (Rotella 2007). RESULTS Nine species of passerines were actively breeding on our study area. Nests (/; = 1,494) of the six focal species accounted for 98% of all passerine nests found during 1997-2007. Only a small number of nests of Horned Lark {Eremo- phila alpestris-, n = 5), Lark Bunting {Calamos- 458 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 3, September 2010 TABLE 1. Annual variation in nesting totals for six songbird species, Bowdoin National Wildlife Refuge, Montana. Numbers of nests discovered in the three undisturbed sites were summarized across years (1997-2006). Total nests were summarized across all songbird species encountered. Values are adjusted per 40 ha. Linear trends in nest totals across years are summarized using Pearson (/•) and beta (b) coefficients. Statistic Sprague's Pipit Savannah Sparrow Grasshopper Sparrow Baird's Sparrow Chestnut-collared Longspur Western Meadowlark Total nests Mean 2.7 4.6 2.7 1.5 15.4 2.6 30.4 CV 0.49 0.30 0.33 1.05 0.31 0.46 0.25 Min 1.3 2.8 1.5 0 10.7 0.8 20.6 Max 5.3 7.4 4.1 4.8 22.4 4.1 41.2 n 94 164 97 52 547 92 1,075 r 0.16 -0.64 0.14 -0.76 -0.53 -0.61 -0.56 h 0.1 -0.3 0.0 -0.4 -0.8 -0.2 -1.3 pizci melcmocorys\ n — 7), and Vesper Sparrow (Pooecetes gramineus\ n = 18) were found. Chestnut-collared Longspurs were the dominant breeding species accounting for over 51% of all passerine nests discovered over all years, followed by Savannah Sparrow (17%), Sprague’s Pipit (8.4%), Grasshopper Sparrow (8.1%), Western Meadowlark (7.8%), and Baird’s Sparrow (6%). Mean rank in nest abundance across all years followed a similar order: Chestnut-collared Long- spur (1.0, CV = 0.0), Savannah Sparrow (2.2, CV = 0.20), Western Meadowlark (3.9, CV = 0.27), Grasshopper Sparrow (4.1, CV = 0.19), Sprague’s Pipit (4.2, CV = 0.33), and Baird’s Sparrow (5.6, CV = 0.18). Annual variation in total nests was summarized collectively for three unburned study sites (Ta- ble 1). Annual variation (CV) in nest density ranged from 1.05 for Baird’s Sparrow to 0.30 for Savannah Span'ow among focal species. Variabil- ity in total songbird nest density was moderate (CV = 0.25) ranging from 41 nests/40 ha in 1997 to 21 nests/40 ha in 2001. A modest declining trend for total songbird nests/40 ha was evident {b = — 1.3, r = —0.56) across all years. This trend was pronounced for Baird’s SpaiTow (r = —0.76), but also evident for Savannah Sparrow (r = —0.64), WesteiTi Meadowlarks (r = —0.61), and Chestnut-collared Longspurs (r = —0.53). Nests were primarily discovered during incubation (67%) with frequency of nests discovered gener- ally decreasing thereafter. Relatively few nests were discovered during the nest-building phase (Fig. 2). Reproductive Parameters. — The CID frequency distribution for Chestnut-collared Longspurs was multimodal across all years, exhibiting three peaks. Bimodal distributions were evident for Build Laying- Incubation — Nestling 509 (Z) c 400 300 ^ 200 I 100 0 ■ 15 222 266 110 122 168 68 14 0 1-3 4-6 7-9 10-12 13-15 16-18 19-21 22-29 Days in nesting cycle FIG. 2. Frequency distribution of nest age-intervals at time of discovery, Bowdoin National Wildlife Refuge, Montana, 1997-2007 Oi = 1.494). Nests were primarily located during incubation with frequency of nests generally decreasing thereafter. Relatively few nests were located during the nest-building phase. Jones el al. • NESTING OF GRASSLANO SONGBIRDS 459 TABLE 2. Frequency distributions For clutch initiation dates (CID) for six songbird species at Bowdoin National Wildlife Refuge, Montana, 1997-2()()7. Values include peaks (modes), mean, and median nesting dates. Nesting period includes the range of activities observed across years for each species from the earliest C!D date to last nest fledged. Frequency distributions were multimodal for three species and generally positively skewed towards early season breeding for all species. Species Nest peaks Mean Median Earliest Latest Potential nesting period Sprague’s Pipit 23 May 5 Jun 25 May 7 May 31 Jul 07 May-25 Aug Savannah Sparrow 24 May, 27 Jun 6 Jun 2 Jun 11 May 19 Jul 1 1 May-09 Aug Grasshopper Sparrow 27 May 13 Jun 16 Jun 18 May 17 Jul 18 May-06 Aug Baird's Spanow 30 May 1 1 Jun 9 Jun 14 May 21 Jul 14 May-10 Aug Chestnut-collared Longspur 21 May, 6 Jun, 15 Jun 7 Jun 7 Jun 30 Apr 19 Jul 30 Apr- 10 Aug Western Meadowlark 24 May, 2 Jun, 7 Jun 4 Jun 4 Jun 30 Apr 14 Jul 30 Apr-07 Aug Savannah SpaiTOws and Western Meadowlarks; the remaining species were unimodal. All CID frequencies were positively skewed towards early- .season breeding. On average, Sprague’s Pipits were the earliest to nest (median date = 25 May) and Grasshopper Sparrows were the latest (medi- an date = 16 Jun) (Table 2). All species ended clutch initiation by the end of July. The earliest fledging date observed was 26 May (Chestnut- collared Longspur) and the latest was 25 August (Sprague’s Pipit). Variation in mean CID across years was low for Chestnut-collared Longspurs (SD = 2.9 days) and higher for Savannah Sparrows (SD = 5.6 days). The trend in CID across years was extremely weak for either species {\b\ < 0.20 d, IH < 0.12). Variability (CV < 0.18) in incubation and nestling phase duration was low, regardless of species (Table 3). Mean clutch size across all years ranged from 4.1 eggs (Chestnut-collared Longspurs) to 4.6 eggs (Sprague’s Pipits). Mean clutch size did not vary considerably across years for either Chestnut-collared Longspurs (CV = 0.04) or Savannah Sparrows (CV = 0.06). Clutch size tended to decrease over the course of a season, regardless of bird species, although Sprague’s Pipits and, to a lesser extent. Chest- nut-collared Longspurs had increased clutch sizes during the first month of the breeding season (Fig. 3). Hatching rate (% of total eggs laid that hatched) across all years ranged from 88% for Chestnut-collared Longspurs and Grasshopper Sparrows to 78% for Savannah Sparrows. Hatch- ing rates were similar for Sprague’s Pipits (85%), Baird’s Sparrows (84%), and Western Meadow- larks (82%). Mean number of nestlings across all years ranged between 3.5 for Savannah SpaiTows to 4.0 TABLE 3. Nest stage duration and reproductive parameters, by songbird species at Bowdoin National Wildlife Refuge. Montana, 1997-2007. Values represent mean ± CV (n; range). Fledged/nest represents the number of young fledged across all nests, regardless of nesting outcome. Reproductive parameters Nest stage (days) Species Incubation Nestling Clutch size No. of nestlings No. Hedged/ .successful nest Fledged/nest Sprague’s Pipit 12.2 ± 0.12 12.9 ± 0.18 4.6 ±0.17 4.0 ± 0.26 3.4 ± 0.35 1 .30 ± 1 .07 (85; 7-15) (19; 9-17) (129; 1-6) (97; 1-6) (49; 1-6) Savannah Sparrow 1 1.6 ± 0.1 1 lO.I ± 0.16 4.3 ± 0.23 3.5 ± 0.37 3.2 ± 0.41 1.23 ± 0.94 (184; 8-15) (73; 7-14) (260; 1-6) (191; 1-6) (101; 1-6) Grasshopper 10.9 ± 0.14 9.7 ± 0.17 4.3 ± 0.24 3.9 ± 0.28 3.6 ± 0.33 1.89 ± 1.56 Sparrow (86; 7-13) (26; 7-12) (123; 0^'-6) (95; 1-6) (64; 1-5) Baird’s Sparrow 1 1.0 ± 0.10 9.6 ± 0.1 1 4.3 ±0.19 3.7 ± 0.31 3.4 ± 0.36 1.52 ± 1.31 (62; 7-14) (27; 8-11) (90; 2-6) (70; 1-5) (40; 1-5) Chestnut-collared 10.9 ± 0.12 11.1 ±0.16 4. 1 ± 0.20 3.6 ± 0.27 3.4 ± 0.32 1 .50 ± 1 .37 Longspur (567; 7-15) (185; 7-15) (770; 1-7) (627; 0“-6) (342; 0^'-5) Western 12.9 ± 0.12 12.7 ±0.10 4.4 ± 0.30 3.8 ± 0.40 3.2 ± 0.47 1.24 ± 0.82 Meadowlark (78; 8-15) (31; 9-16) (119; 0“-7) (89; 1-6) (46; 1-5) ‘‘ Brown-headed Cowbird (Molothrus ater) only. 460 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 3, September 2010 (D N tfl SZ o o c cc CD 6.5 6.0 5.5 5.0 - 4.5 - 4.0 3.5 - 3.0 - 2.5 — ♦ Sprague's Pipit — ■ — Savannah Sparrow ii Grasshopper Sparrow — - X - - Baird's Sparrow — X — Chestnut-collared Longspur — • — Vttestern M eado \Mark 1-15 May 16-31 May 1-15 Jun 16-30 Jun 1-15 16-31 Jul Jul Date FIG. 3. Mean clutch size plotted against date by species across all years (1997-2007), Bowdoin National Wildlife Refuge, Montana. Mean clutch size generally decreased for each species as the nesting season progres.sed. for Sprague’s Pipits (Table 3). Grasshopper Spar- rows had the highest fledging rate ( 1 .9) across all nests in all years while Savannah Sparrows and Western Meadowlarks had the lowest (1.2) (Table 3). Variation in mean fledglings/nest across years was low for Chestnut-collared Longspurs (CV = 0.20) with a modest decreasing trend across years (b = —0.05, r = —0.59). This statistic was more variable for Savannah SpaiTows (CV = 0.45), but with no discernable trend (r = -0.09). Nest Success. — DSR across all species and years was estimated at 0.947 (SE = 0.002). There was little statistical support for species-specific variation in DSR (Table 4). Model-averaged estimates for each species (Table 5) indicated little variation in DSR among species. However, variation in DSR across years was evident, suggesting a decline during the study period (h = —0.002, r = —0.54) (Fig. 4). There was little TABLE 4. Model set exploring the relationship between bird species (6 levels) and year (1997-2007) at Bowdoin National Wildlife Refuge, Montana on nest daily survival rate (DSR) using Akaike’s Information Criterion, corrected for small sample sizes (AIC, ). Delta AIC,. (AlC,. model i - AIC, minimum). Akaike weight (w,), and number of parameters (K) are included for each model. Model A AIC, vv, K Year ().() 0.66 1 1 Specie.s -f year 1.-3 0.3 1 16 Intercept 6.8 0.02 1 .Specie.s 10.7 0.00 6 Species*ycar 10.7 0.00 63 statistical support for an interaction between species and year (Table 4). Nest depredation accounted for 82% of all known nest failures with most (71%) during the nestling stage. Desertion (abandonment) was judged to be the cause for 18% of all known nest failures. Cause of abandonment was discernable in only 27% of these cases; 25% {n = 38) were attributed to severe weather (e.g., heavy rain, hailstorms), 1% (n = 2) to observer activities, and 1% (« = 2) to Brown-headed Cowbird brood parasitism. Brown-headed Cowbird parasitism rates ranged from 2% for Chestnut-collared Longspurs to 26% for Western Meadowlarks (Table 6). DISCUSSION Songbird community structure based on nesting density was relatively stable across the 1 1-year period of this study. Chestnut-collared Longspurs were clearly the dominant nesting species every year. Savannah Span'ows were the second most abundant bird species during all but 2 years, while Baird's Sparrows were least abundant among focal species during all but 2 years. Few to no nests of shortgrass prairie obligates, such as Horned Lark or McCown’s Longspur {Calcarius mccownii) were located. Chestnut-collared Long- spurs generally breed in drier, shortgrass to short- mixed-grass prairie, while Sprague’s Pipits and Baird’s Sparrows range across the breadth of mixed-grass prairie grassland types (Knopf 1996). Our assemblage of breeding songbirds, in addition to plant community composition and structure (Dieni and Jones 2003), classify our study area as Jones el at. • NESTING OF GRASSLAND SONGBIRDS 461 TABLE 5. Estimated nest DSR for six bird species across all years at Bowdoin National Wildlife Refuge, Montana, 1997-2007. Standard errors (SE), lower (LCI) and upper (UCI) 95% confidence limits, and number of nests (/?) are included. Nest success equals DSR raised to the mean number of days of the nesting period (incubation and nestling stages combined) for each species. Species DSR SE LCI UCI n Nesting period Nest success Sprague's Pipit 0.947 0.003 0.942 0.952 128 25.1 0.27 Savannah Sparrow 0.946 0.004 0.938 0.953 255 21.7 0.25 Grasshopper Span'ow 0.948 0.004 0.940 0.955 121 20.6 0.39 Baird’s Sparrow 0.947 0.003 0.940 0.953 90 20.6 0.32 Chestnut-collared Longspur 0.947 0.002 0.943 0.951 759 22.0 0.31 Western Meadowlark 0.947 0.003 0.941 0.952 118 25.6 0.25 Totals 0.947 0.002 0.943 0.950 1,471 22.3 0.30 xeric short-mixed-grass prairie (Knopf 1996, Askins et al. 2007). Breeding Bird Survey data indicate long-term (1966-2007) declines in rangewide populations for all our focal bird species (Sauer et al. 2008). In particular, rangewide populations of two mixed- grass endemics, Sprague’s Pipit (-3.9, P = 0.00, n = 169) and Baird’s Sparrow ( — 3.4, P = 0.01, n = 141), as well as Chestnut-collared Longspur (—2.8, P = 0.00, n = 160) have exhibited steep declines (Sauer et al. 2008). We recorded general declines in nest density for most of our study species on undisturbed study sites, although no discemable trend was observed for Sprague’s Pipit or Grasshopper Sparrow. The decline in nesting density of Baird’s Sparrow may be a function of drier than normal climatic conditions that prevailed over most of the study period (National Climatic Data Center 2007). Our study area possessed none of the habitat conditions that aie commonly the result when the northern mixed-grass prairie is left idle (Madden et al. 1999). The prevalence of invasive plant species and woody-vegetation encroachment remained low, while forbs were abundant during the course of the study (S. L. Jones, unpubl. data). Moreover, xeric mixed-grass prairie historically had lengthy periods without grazing, and natural burn fre- quencies that often ranged between 8 and 26 years (Askins et al. 2007). We used nest density as an index to bird abundance to gauge community structure and trend, but nest density clearly does not equate directly to bird or even territorial density. Some species we studied have been documented to have multiple broods in a season. Our population was marked for only a few species and years (Jones et al. 2007); it was not possible in most cases to distinguish single broods by different individuals from multiple broods after a successful nest or re- 0.98 - „ 0.97 CE C/3 Q 0.96 B 2 0.95 CTJ 1 0.94 i_ 3 CO ^ 0.93 (0 2 0.92 CO OJ ^ 0.91 0.90 - T - — T r r- , ^ , , 1997 1999 2001 2003 2005 2007 Year FIG. 4. Annual variation in daily survival rates of songbird nests (DSR ± SE) at Bowdoin National Wildlife Refuge, Montana. There was considerable annual variation in DSR (annual mean = 0.946 ± 0.01 1 CV). A declining trend in DSR was evident (h = -0.002, r = -0.54). 462 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 3, September 2010 TABLE 6. Nesting outcome and parasitism rates for each bird species across all years, 1997-2007 at Bowdoin National Wildlife Refuge, Montana. Most monitored nesting attempts were unsuccessful with the majority of failures attributed to nest predation. Weather events (e.g., hail) were collectively the leading known cause of nest desertion. Parasitism by Brown-headed Cowbirds resulted in only a few nest failures. Brown-headed Cowbird parasitism Known causes of abandonment («) Species Successful Depredated Abandoned Unknown n Relative frequency Total cowbirds fledged Weather Observer disturbance Parasitism Sprague’s Pipit 0.37 0.57 0.06 0.00 128 0.02 2 1 0 0 Savannah Sparrow 0.40 0.43 0.17 0.01 260 0.13 17 15 1 1 Grasshopper Sparrow 0.52 0.35 0.12 0.01 123 0.04 2 3 0 0 Baird’s Sparrow 0.43 0.42 0.16 0.00 89 0.04 0 4 0 0 Chestnut-collared Longspur 0.44 0.48 0.07 0.01 770 0.02 3 11 0 1 Western Meadowlark 0.40 0.45 0.14 0.00 119 0.26 18 4 1 0 n 637 688 151 13 1,489 91 42 38 2 2 nesting after nest failure. However, Savannah Sparrows, Chestnut-collared Longspurs, and Western Meadowlarks all had multimodal CID distributions suggesting at least some community members may have had several clutches per year. None of the species we studied has been documented with more than two successful broods per year (Wheelwright and Rising 1993, Lanyon 1994, Vickery 1996, Hill and Gould 1997). Sprague’s Pipit, and Baird’s and Grasshopper sparrows had only single frequency peaks in our CID data, which suggests a tendency towards a single brood per season. This contrasted with Vickery (1996), who reported Grasshopper Spar- rows in general have a protracted breeding season, often producing >2 broods annually, even in the northern portion of their range. Sprague’s Pipit and Baird’s Sparrow have only rarely been documented to have multiple broods in a season (Robbins and Dale 1999, Green et al. 2002). Second broods after successfully fledging young have been documented twice for Sprague’s Pipit (Sutter 1996, Davis 2009). Re-nesting after nest failure has been rarely documented for this species (Sutter 1996, Davis 2009), as has polygyny (/? = 1) (Dohms and Davis 2009). Baird’s Sparrow in southwest Manitoba (Davis and Sealy 1998) and in our study had a .second, smaller nest initiation peak in mid to late June, suggesting .second broods or re-nesting attempts; .second broods have only been documented twice (Davis and Sealy 1998; S. L. Jones, unpubl. data). Mean CID for Savannah Sparrows and Chest- nut-collared Longspurs varied little across years. Mean clutch size also had little variation among seasons; however, a within-season declining pattern was evident for this parameter for all species. Similar patterns have been documented elsewhere for these bird species (Vickery 1996, Davis and Sealy 1998, Winter 1999, Davis 2003). Sprague’s Pipits were a notable exception in our study, where clutch size generally increased during the first month of the breeding season. A similar pattern was also reported for Sprague’s Pipit in Saskatchewan (Davis 2003). Clutch size in our study linearly decreased over the course of the breeding season for Western Meadowlarks, which contrasted with a study in Manitoba, where Western Meadowlarks had peak clutch size during mid-season (Lanyon 1994). Brown-headed Cowbird parasitism rates, ex- cept for Western Meadowlark, were relatively low in our study (2-13%). Much higher rates were reported for our focal species in southwest Manitoba: Sprague’s Pipit = 18%, Savannah Sparrow = 32%, Grasshopper Sparrow = 29%, Baird’s Sparrow = 37%, and Chestnut-collared Longspur = 14% (Davis and Sealy 2000). Our rate for Western Meadowlark (26%) was low compared to rates reported in Manitoba (43%) (Davis and Sealy 2000), and North Dakota (47%) (Koford et al. 2000), but were consistent with reported rates in Saskatchewan (25%) (Davis 2003). Chestnut-collared Longspurs and Spra- gue’s Pipits had the lowest parasitism rates in our study, which is consistent with findings in other studies (Hill and Gould 1997, Davis and Sealy 2000, Davis 2003). Several factors may be Jones el al. • NESTING OF GRASSLAND SONGBIRDS 463 responsible for the relatively low levels of parasitism found in our study area (i.e., large patches of continuous grassland, a landscape with few potential perches, relatively little shrub and tree cover, and no cattle grazing). All of these factors directly enhance Brown-headed Cowbird parasitism rates (Davis and Sealy 2000, Koford et al. 2000). Nest depredation is commonly the single most important factor affecting nesting success in cup- nesting passerines (Ricklefs 1969); depredation rates have been generally high throughout the mixed-grass prairie (Vickery et al. 1992; Winter 1999; Davis 2003, 2005; Winter et al. 2005). Nest predation accounted for the majority of observed nest failures in our study, mainly during the nestling stage. This is a common pattern among passerines; greater parental activity at the nest, as well as greater olfactory cues, noise, and move- ments by nestlings are thought to increase the likelihood of nest detection by predators (Halupka 1998). Overall nest success for Sprague’s Pipits was estimated to be much higher in southwest Manitoba (47%, Davis and Sealy 1998) than estimated in our study area (27%). Grasshopper Sparrow and Western Meadowlark nesting suc- cess rates were comparable to those reported by Lanyon (1994) and Vickery (1996). Our estimate of nest success (32%) for Baird’s Sparrow was higher than reported in Saskatchewan (26%, Davis 2003), but lower than in Manitoba (54%, Davis and Sealy 1998). Nest success for Chestnut- collared Longspurs was higher in Alberta (56%, Hill and Gould 1997) than estimated for our study area (44%; Table 6). DSR estimates did not differ markedly among bird species, suggesting that factors such as weather and the predator community were affect- ing all bird species equally. Therefore, tho.se elements affecting nest success that vary with year (i.e., predator community or weather) probably influenced the members of this grassland songbird community in a similar manner in a given year. ACKNOWLEDGMENTS This project was funded by the U.S. Fish and Wildlife Service, Nongame Migratory Bird Program, Region 6. Field support and resources were provided by J. E. Comely, C. R. Luna, D. M. Prellwitz. and the staff at Bowdoin National Wildlife Refuge. We thank the many dedicated field assistants who worked on this project over the years: Brice Adams, A. C. Araya, B. S. Atchley, R. R. Conover, Michael Friedrich, Jeanne Hammond, D. I. Hedegaard, Jessica Hinz, B. G. Larson, Rebecca Lind, L. B. Lingohr, S. N. Luttich, Kevin Payne, Heather Sauls, Caroline Stahala, Vicki Trabold, P. N. Wiederrick, and J. K. Wood. We are also grateful to M, T. Green, G. R. Guepel, and J. R. Ruth for insights that contributed to the design and conduct of this study. D. J. Ingold, D. H. Johnson, and an anonymous reviewer provided edits and comments that substantially improved this paper. We thank Mike Artmann for the map in Fig. 1. LITERATURE CITED Ahlering, M. A. AND D. H. Johnson. 2006. Conspecific attraction in a grassland bird, the Baird’s Sparrow. Journal of Field Ornithology 77:365-371. Askins, R. A., F. Chavez-Ramirez, B. C. Dale, C. A, Haas, J, R. Herkert, F. L. Knopf, and P. D. Vickery. 2007. Conservation of grassland birds in North America: understanding ecological processes in different regions. Ornithological Monographs 64: 1^6. Basore, N. S., L. B. Best, and J. B. Wooley Jr. 1986. Birds nesting in Iowa no-tillage and tilled cropland. Journal of Wildlife Management 50:9-28. Burnham, K. P. and D. R. Anderson. 2002. Model selection and multi-model inference: a practical information-theoretic approach. Second Edition. Springer-Verlag, New York, USA. Clark, R. G. and D. Shutler. 1999. Avian habitat selection: pattern from process in nest-site use by ducks. Ecology 80:272-287. Davis, S. K. 2003. Nesting ecology of mixed-grass prairie songbirds in southern Saskatchewan. Wilson Bulletin 115:119-130. Davis, S. K. 2005. Nest-site selection patterns and the influence of vegetation on nest survival of mixed-grass prairie passerines. Condor 107:605-615. Davis, S. K. 2009. Re-nesting intervals and length of incubation and nestling periods of Sprague’s Pipit. Journal of Field Ornithology 80:265-269. Davis, S. K. and S. G. Sealy. 1998. Nesting biology of the Baird’s Sparrow in southwestern Manitoba. Wilson Bulletin 110:262-270. Davis, S. K. and S. G. Sealy. 2000. Cowbird parasitism and nest predation in fragmented grasslands of southwestern Manitoba. Pages 220-228 in Ecology and management of cowbirds and their hosts: .studies in the conservation of North American passerine birds (J. N. M. Smith, T. L. Cook. S. 1. Rothstein, S. K. Robinson, and S. G. Sealy, Editors). University of Texas Press, Austin, USA. Dieni, j. S. and S. L. Jones. 2003. Grassland songbird nest site selection patterns in north-central Montana. Wilson Bulletin 115:388-396. Dinsmore. S. j., G. C. White, and F. L. Knopf. 2002. Advanced techniques for modeling avian nest survival. Ecology 83: 3476-3488. Dohms. K. M. and S. K. Davis. 2009. Polygyny and male parental care by Sprague’s Pipit. Wikson Journal of Ornithology 121:826-830. 464 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 3. September 2010 Green, M. T., P. E. Lowther, S. L. Jones, S. K. Davis, AND B. C. Dale. 2002. Baird’s Sparrow (Ammodra- miis bairdii). The birds of North America. Number 638. Halupka, K. 1998. Vocal begging by nestlings and vulnerability to nest predation in Meadow Pipits Anthiis pratensis: to what extent do predation costs of begging exist? Ibis 140; 144—149. Higgins, K. F. 1984. Lightning fires in North Dakota grasslands and in pine-savanna lands of South Dakota and Montana. Journal of Range Management 37:100- 103. Hill, D. P. and L. K. Gould. 1997. Chestnut-collared Longspur {Ccdcarius ornatiis). The birds of North America. Number 288. Houston, C. S. and J. K. Schmutz. 1999. Changes in bird populations on Canadian grasslands. Studies in Avian Biology 19:87-94. iGL, L. D. AND D. H. Johnson. 1999. Le Conte’s Sparrows breeding in Conservation Re.serve Program fields; precipitation and patterns of population change. Studies in Avian Biology 19:178-186. Jones, S. L., J. S. Dieni, M. T. Green, and P. J. Gouse. 2007. Annual return rates in breeding grassland songbirds. Wilson Journal of Ornithology 1 19:89-94. JoNGSOMJiT, D., S. L. Jones, T. Gardali, G. R. Geupel, AND P. J. Gouse. 2007. A guide to nestling development and aging in altricial passerines. USDl, Fish and Wildlife Service, Biological Technical Publication FWS/BTP-R6008-2007. Washington, D.C., USA. Knopf, F. L. 1994. Avian assemblages on altered grasslands. Studies in Avian Biology 15:247-257. Knopf, F. L. 1996. Prairie legacies-birds. Pages 135-148 in Prairie conservation: preserving North America’s most endangered ecosystem (F. B. Sam.son and F. L. Knopf, Editors). Island Press, Covelo, California, USA. Koford, R. R., B. S. Bowen, J. T. Lokemoen, and A. D. Kruse. 2000. Cowbird parasitism in grassland and cropland in North Dakota. Pages 229-235 in Ecology and management of cowbirds and their hosts: studies in the conservation of North American passerine birds (J. N. M. Smith, T. L. Cook, S. 1. Rothstein, S. K. Robinson, and S. G. Sealy. Editors). University of Texas Press, Austin, USA. Lanyon, W. E. 1994. We.stern Meadowlark (Sturnella neglecta). The birds of North America. Number 104. Lokemoen, J. T. and R. R. Koford. 1996. Using candlers to determine the incubation stage of passerine eggs. Journal of Field Ornithology 67:660-668. Madden, E. M., A. J. Han.sen, and R. K. Murphy. 1999. InlJuence of pre.scribed fire history on habitat and abundance of pas, serine birds in northern mixed-grass prairie. Canadian Field-Naturalist 1 13:627-640. Martin, T. E. and G. R. Geupel. 1993. Nest-monitoring plots: methods for locating nests and monitoring success. Journal of Field Ornithology 64:507-519. National Climatic Data Center. 2007. Monthly surface data, Malta, Montana. National Oceanic and Atmo- spheric Administration, Asheville, North Carolina, USA. http://lwfncdc.noaa.gov/oa/ncdc.html. Ricklefs, R. E. 1969. An analysis of nesting mortality in birds. Smithsonian Contributions in Zoology Number 9. Washington, D.C., USA. Robbins, M. B. and B. C. Dale. 1999. Sprague’s Pipit (Antbus spragiieii). The birds of North America. Number 439. Rotella, j. j. 2007. Nest survival models. Pages 17.1- 17.22 in Program MARK: “a gentle introduction” (E. G. Cooch and G. C. White, Editors). Sixth Edition. www.phidot.org/software/mark/docs/book/. Rotella, J. J., S. J. Dinsmore, and T. L. Shaffer. 2004. Modeling nest-survival data: a comparison of recently developed methods that can be implemented in MARK and SAS. Animal Biodiversity and Conservation 27: 187-205. Sauer, J. R., J. E. Hines, and J. Fallon. 2008. The North American breeding bird survey, results and analysis 1966-2007. Version 5.15. USGS, Patuxent Wildlife Research Center, Laurel, Maryland, USA. www.mbr- pwrc . usgs .go v/bbs/. Sutter, G. C. 1996. Renesting intervals in Sprague’s Pipit. Anthns spragueii. Canadian Field-Naturalist 1 10:694- 697. Vickery, P. D. 1996. Grasshopper Sparrow (Ammodramus savanna rnin). The birds of North America. Number 239. Vickery, P. D., M. L. Hunter, and J. V. Wells Jr. 1992. Evidence of incidental nest predation and its effects on nests of threatened gra.ssland birds. Oikos 63:281-288. White, G. C. 2008. Program MARK. Mark and recapture survival rate estimation (Version 5.1). www.wamercnr. colostate.edu/~gwhite/mark/. Winter, M. 1999. Relationship of fire history to territory size, breeding density, and habitat of Baird’s Sparrow in North Dakota. Studies in Avian Biology 19:171- 177. Winter, M., D. H. Johnson, and J. A. Shaffer. 2005. Variability in vegetation effects on density and nesting success of grassland birds. Journal of Wildlife Management 69:185-197. Wheelwright, N. T. and J. D. Rising. 1993. Savannah Sparrow (Pas.sercidiis sandwichensis). The birds of North America. Number 45. Zar, j. H. 1999. Biostatistical analysis. Fourth Edition. Prentice-Hall, Upper Saddle River, New Jersey, USA. The Wilson Journal ofOrnithology 1 22(3):465-474, 2010 MOUNTAIN BIKING TRAIL USE AFFECTS REPRODUCTIVE SUCCESS OF NESTING GOEDEN-CHEEKED WARBLERS CRAIG A. DAVIS,'"' DAVID M. LESLIE JR.,' W. DAVID WALTER,' AND ALLEN E. GRABER'* ABSTRACT. — We evaluated toraging and nesting behavior, territory size, and nest success of Golden-cheeked Warblers (Dendroica clirysoparia), a federally endangered songbird, relative to mountain biking trail use. We conducted our study at two mountain biking sites and two control sites at Fort Hood Military Base and in Austin, Texas, in spring 2002 and 2003. Teiritories ot male Golden-cheeked Warblers in biking sites (2.2 ha) were >1.5 times as large as those in non-biking sites (1.4 ha). Mayfield nest success in biking sites (n = 33) was 35% compared to 70% in non-biking sites {n = 22). Nest abandonment was three times greater in biking areas (15%) than non-biking areas (5%). Seven nests were depredated in biking sites, but only two nests were depredated in non-biking sites. Texas rat snakes (EUiphe ohsoleta) were the most frequent nest predator at biking sites, accounting for 71% of the predations. We conducted behavioral observations of male Golden-cheeked Warblers in biking (n = 139) and non-biking (/t = 204) sites. Males spent similar amounts of time in diurnal behaviors in biking and non-biking sites. We used video-camera systems to record female nesting behaviors at 17 nests in biking sites and 15 nests in non-biking sites. Nesting behaviors of females did not differ between biking and non- biking sites. The cumulative effect of disturbance from mountain biking trail use on Golden-cheeked Warbler foraging and nesting behavior appears to be minimal, but fragmentation and alteration of habitat by mountain biking trails may reduce quality of nesting habitat for Golden-cheeked Warblers. Received 23 November 2009. Accepted 27 March 2010. Mountain biking is one of the fastest-growing outdoor recreational activities in the United States, experiencing substantial growth in the last 20 years (Mosedale 2002, Pike Masteralexis et al. 2008). This increase has led to greater demands for quality outdoor experiences on trail networks in areas that often contain fragile environments (Flather and Cordell 1995). Land managers have raised concerns about the impacts of mountain biking activities on natural resources (Chavez et al. 1993, Symmonds et al. 2000). These concerns led to restrictions in mountain biking activities at all federal wilderness areas and many other public lands in the United States despite little scientific insight on their impacts to wildlife (Taylor and Knight 2003). The Golden-cheeked Warbler {Dendroica chry- soparia) is a federally endangered species that breeds exclusively in the mature juniper (Jimi- perus ashei)-o'dk (Quercus spp.) woodlands of ' Department of Natural Resource Ecology and Manage- ment. Oklahoma State University, Stillwater, OK 74078, USA. 'U.S. Geological Survey, Oklahoma Cooperative Fish and Wildlife Research Unit. Department of Natural Resource Ecology and Management, Oklahoma State University, Stillwater, OK 74078, USA. ■’Colorado Cooperative Fish and Wildlife Research Unit, 201 JVK Wagar Building, 1484 Campus Delivery, Colorado State University, Fort Collins, CO 80523, USA. ■"SWCA Environmental Consultants. 130 Rock Point Drive, Suite A. Durango, CO 81301, USA. ’Corresponding author; e-mail: craig.a.davis@okstate.edu central Texas (Ladd and Gass 1999). Some of its breeding range is used by mountain bikers (e.g.. Dinosaur Valley State Park in Somervell County, Lost Maples State Natural Area in Bandera County, Belton Lake Outdoor Recreation Area [BLORA] in Bell County, and Emma Long Metropolitan Park [ELMP] in Travis County). Increased levels of predation on passerine nests near mountain biking trails have been documented (Miller et al. 1998, Miller and Hobbs 2000). Mountain biking also may alter distribution and behaviors of Golden-cheeked Warblers. In partic- ular, disturbance from mountain biking may lessen the time and energy used for fitness- enhancing activities including feeding, nest atten- tiveness, mate attraction, and ten'itory defense (Knight and Gutzwiller 1995). Disturbance from mountain biking activities may pose a threat to successful breeding of Golden-cheeked Warblers; a Biological Opinion issued by the U.S. Fish and Wildlife Service (USDI 1993) and a regional “take” permit issued by the U.S. Fish and Wildlife Service under the umbrella oi the Balcones Canyonlands Con.serva- tion Plan recommended eliminating and restrict- ing mountain biking activities from Golden- cheeked Warbler habitat in areas of the Fort Hood Military Reservation and Travis County, respectively. Those recommendations were con- troversial, and further research of the impacts of mountain biking on Golden-cheeked Warblers was needed. Our objectives were to examine effects of mountain biking trail use on foraging 465 466 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 3. September 2010 and nesting behavior, territory size, and nest success of Golden-cheeked Warblers. METHODS Study Area. — We conducted this study at two mountain biking sites and two control sites from March to June, 2002 and 2003. The mountain biking sites were at BLORA in Fort Hood Military Reservation (31° 06' N, 97° 31' W) and ELMP (30° 19' N, 97° 50' W) in Balcones Canyonlands Preserve (BCP) in Austin, Texas. Control sites were at Training Area 13A (13A) on Fort Hood Military Base (30° 10' N, 97° 45' W) and Forest Ridge Preserve (FRP; 30° 24' N, 97° 55' W) in BCP. Recreational activities were restricted from both control areas, and military activities were restricted on both sites at Fort Hood. Mountain biking activities, as measured by TrailMaster photoelectric trail counters placed along trails, averaged 15 mountain bikers/day (5 bikers/day at BLORA and 17 bikers/day at ELMP) and ranged from zero to 30 bikers/day at biking sites (Davis and Leslie 2008). Mountain bikers accounted for >80% of the trail users at biking sites. Other users recorded at biking sites included motorized bikers (<2%), hikers and joggers (2%), and researchers (11%). Fort Hood and BCP are within the Crosstimbers and Southern Tallgrass Prairie and Edwards Plateau ecoregions, respectively. Fort Hood con- tains —21,850 ha of Golden-cheeked Warbler habitat and supports the largest known population of Golden-cheeked Warblers under one manage- ment authority (USDI 1993); the BCP contains — 10,600 ha of preserve properties with 5,300 ha of Golden-cheeked Warbler habitat (Reidy et al. 2008). Mountain biking sites ranged in size from 116 (ELMP) to 192 ha (BLORA), and control sites ranged in size from 190 (FRP) to 198 ha (13A). All sites were woodlands associated with steep limestone canyons and plateaus dominated by Ashe Juniper and a variety of oaks. Territory Size. — We used spot-mapping (Ralph et al. 1993) to delineate territories for color- banded and unbanded males at each site from 10 March to 24 June 2002 and 2003. Males were color banded from previous studies (at the BLORA site) and during our study. We used conspecific playback to capture males with 6-m mist nets. We banded males with unique color combinations of two or three plastic color bands and a numbered aluminum U.S. Geological Survey band. Individual locations of males were identified by standing directly under the location where a male was observed and using a handheld global positioning system unit (GPS; Garmin GPS III-I-, Garmin International, Olathe, KS, USA) to record Universal Transverse Mercator (UTM) coordinates. We recorded one location of each individual during a single visit, but attempted to obtain locations at different time periods through- out the day. We later plotted locations of individuals on 1:4,000 digital orthoquad maps with 1 .5-m resolution imagery and a 500 X 500-m grid overlay to distinguish territories throughout each year. We distinguished individual territories by identifying color bands. We relied on counter- singing by unbanded males that occupied adjacent territories as well as a general familiarity of an individual’s territory as the season progressed to facilitate delineation of territories. We attempted to obtain >30 locations per individual male (Laundre and Keller 1984, Bibby and Burgess 1992, Seaman et al. 1999). We used the Home Range extension in ArcView and 1 ,000 iterations of the bootstrap procedure to calculate the minimum number of male locations sufficient for a reliable minimum convex polygon (MCP) estimate (Rodgers and Carr 1998, ESRI 2000). We calculated that 33 locations were sufficient for estimating the MCP for males on the basis of the bootstrap procedure. We estimated MCP using the Home Range extension in ArcView for those males with a minimum of 33 locations. We characterized land cover types within each study site and each teiritory using the National Land Cover Database (NLCD) of 2001 to evaluate whether teiTitories differed by habitat composi- tion. Land-cover categories in the NLCD were reclassified into eight broad categories (water, bare ground, deciduous forest, evergreen forest, mixed forest, shrub/scrub, grassland, and wetland) to standardize our analysis across the four study sites. The proportions of landcover types from NLCD (i.e., forested, grassland) within each bird’s home range were extracted in ArcView. We walked all trails within each study area using a GPS unit (Trimble Geoexplorer 3, Trimble Navigation Ltd., Sunnyvale, CA, USA) to delin- eate trails. We downloaded trail locations and created a Geographic Information System (GTS) cover layer of trails for each study site using ArcView. Time Activity Budgets of Territorial Males. — We assessed the impact of mountain biking activity on foraging activities by conducting Davis et at. • MOUNTAIN BIKING AND GOLDEN-CHEEKED WARBLERS 467 behavioral observations of males on territories at each study site from March to June, 2002 and 2003. We conducted observations throughout the day. We minimized sampling bias due to diurnal variation in foraging behavior by collecting behavioral data equally during three time periods: early, sunrise to 1100 hrs; mid-day, 1101 to 1500 hrs; and late, 1501 hrs to sunset. Males were located by listening for a Type A- or B-song. We predominantly observed males using binoculars, but, in some cases, observed males directly. We attempted to observe each bird for 300 sec and discarded observation sessions <30 sec in length. We dictated behaviors into a handheld microcas- sette tape-recorder. We avoided bias toward conspicuous behaviors by not recording behaviors until 5 sec after locating a bird (Wunderle and Latta 1998). We used focal individual sampling (Alt- mann 1974, Martin and Bateson 1986) to collect behavioral data. We avoided autocorrelation for banded birds by not recording behavior from the same individual more than once in a single day. We did not collect behavioral observations during inclement weather (rain or high wind). We classified behaviors into nine categories (Robinson and Holmes 1982, Remsen and Rob- inson 1990): singing, preening, agonistic, mate guarding, perching (resting), locomotion (flying and hopping), foraging (searching, attacking, and food handling), feeding fledgling, and time-out (time when the bird was out-of-sight). The bird was considered in “time-out” when a focal male briefly disappeared behind vegetation during an observation. The tape recorder continued to run during the “time-out” period, and observations resumed when the bird reappeared. We ended the observation at the end of the last observation before the “time-out” if the bird did not reappear within the 300-sec period. We further examined foraging activities in 2003 by recording the amount of time a male was engaged in searching, attacking, and food- handling. Search behaviors were considered to be any movements leading up to locating and acquiring food, and included hops (leg powered movements <10 cm), short flights (<20 cm), medium flights (20 cm to 2 m), and long flights (>2 m). Attack maneuvers were any movements directed at a food item or the substrate that concealed the food item after it was located and included gleaning, reaching, hanging, leaping, sally-striking, sally-hovering, and flutter-chasing (Remson and Robinson 1990). Parental Behavior at Nests. — We assessed the impact of mountain biking trail use on nesting and parenting activities of Golden-cheeked Warblers by video-recording nests using miniature video camera systems (Fuhrman Diversified Inc., Sea- brook, TX, USA). We relied on video-recordings in 2002 provided by another researcher studying nest predation of Golden-cheeked Warblers at Fort Hood sites (M. M. Stake, University of Missouri, Columbia, USA); we collected nesting and parent- ing behavior in 2003 using eight miniature video camera systems at the Fort Hood and Austin sites. Protocols for video-recording were similar for 2002 and 2003. Each system consisted of a 32- by 32- by 60-mm camera with a small lens and six infra-red diodes (950 nm) that allowed filming at night, an articulating arm that supported the camera, and a clamp at the base of the arm for attachment to a substrate. A video recorder, allowing 24 hrs of footage (~7 frames/sec) on a 120-min VHS cassette tape, was connected to the camera with a 20-m cable and powered by a 12-V rechargeable battery. We used a hand-held monitor connected to the cable for positioning the lens at the nest during set-up and connected to the base set-up to check nest status each day without approaching the nest. We changed tapes daily and replaced batteries every other day. We generally found nests opportunistically during location of focal males. Most nests were found by following an adult to the nest during building and nestling stages. We set up cameras at nests, if accessible, and if eameras were available; nests found in the egg stage were prioritized for cameras but nests in the nestling stage were used as cameras became available. Cameras were typically installed in the afternoon to reduce risk of Brown-headed Cowbird (Molothrus ater) parasitism. We only installed cameras after the laying stage to reduce the risk of female abandonment (Stake 2000, Stake and Cimprich 2003, Stake et al. 2004, Reidy et al. 2008) and removed cameras if females did not resume activity at the nest within I hr. We were only able to monitor female activity at nests during incubation and nestling stages. Nests were monitored daily by video until nest fate was known (e.g., fledged young, depredated, aban- doned, failed due to weather). We watched tapes using a multispeed video player. Tapes for individual nests were viewed until the nest was no longer active. We recorded behaviors during the incubation stage, early 468 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122. No. 3, September 2010 nestling stage, and late nestling stage. Behaviors recorded included female incubating eggs, female brooding nestlings, female or male provisioning nestlings, female shading nestlings, female or male perching and scanning at the nest, and male feeding the female. Females brood recently hatched nestlings for long periods during the first 3 days after hatching and continue brooding for 1- 2 days thereafter (Ladd and Gass 1999); thus, we designated the early nestling stage as the first 5 days of the nestling stage. We recorded a variety of behaviors, but focused on incubation, brooding nestlings, and provisioning nestlings because changes in these behaviors may affect nest success (Gill 2007). Specifically, we focused on percent time spent incubating, number of incuba- tion boLits/day and length of incubation bouts during the incubation period, and percent time spent brooding and provisioning nestlings, num- ber of brooding and provisioning bouts/day, and length of brooding and feeding bouts during the early and late nestling stage. Nest Success. — We assessed effects of moun- tain bike trail use on Golden-cheeked Warbler nesting success by monitoring all nests located at the study sites. Most of the nests (90%) were monitored with video; we monitored nests without cameras every 2-5 days until near fledging at which time we visited daily to more accurately monitor nests and correctly identify nest outcome. We considered a nest to be successful if >1 host young fledged. We considered a nest depredated if a predator removed at least one egg or nestling while the nest was active, and a nest abandoned if the male and female were absent from the nest after several days and the eggs or young were still present. We were able to identify all predation events to species-level from watching tapes. Successful production of ^1 fledgling was confirmed by locating fledglings being fed by a parent (usually verified the day of or day after fledging), or by recording fledging by video surveillance. Statistical Analyses.— We calculated area of trails by multiplying the linear length of trails within each MCP by 2.5 m, considered the average width of trails (AEG. pers. obs.). We used one-way analysis of variance (ANOVA) to examine differences in territory size, amount of trail area within each MCP, and habitat compo- sition (% developed, % deciduous, % juniper, and % grassland) between biking and non-biking areas. Some banded males had multiple foraging observations within time periods, and we aver- aged behaviors for those individuals within each time period prior to conducting analyses. We analyzed time-activity data using multivariate analysis of variance (MANOVA) with a factorial an'angement (SYSTAT 2004). Treatment (biking and non-biking), time period (early, mid-day, and late), and year (2002 and 2003) were independent factors in the MANOVA. We used MANOVA because the dependent variables (i.e., individual behaviors) were not independent of each other (i.e., the amount of time engaged in one behavior influenced the amount of time engaged in other behaviors). Wilks’ lambda (X.) was the test criterion. We used univariate analysis of variance (AN- OVA) to examine differences in individual behaviors between years if the overall MANOVA was significant (Barker and Barker 1984). Behav- ioral data met assumptions of normality and homogeneity (Johnson and Wichern 1988, SY- STAT 2004). We used a similar MANOVA with factorial airangement to analyze foraging search behavior data. We used Kruskal-Wallis tests, because of our small sample size, to compare food capture techniques (i.e., proportion of attack modes used by each male during foraging) of males between biking and non-biking sites. There was considerable variability in amount of time each nest was videotaped (which could overemphasize the activities at some nests); thus, we calculated parental behavior parameters on a daily basis. We compared parental behavior parameters using Kruskal-Wallis tests. We used Mayfield’s (1961, 1975) method to calculate daily nest survival rates (DNS) for incubation and nestling stages, and nest success for biking (incubation: 1 1 1 nest-days, nestling: 146 nest-days) and non-biking (incubation: 69 nest-days, nestling: 108 nest-days) sites. We calculated standard errors for DNS following Johnson (1979). All statistical tests were consid- ered significant at P < 0.05. RESULTS Territory Size. — Territory size for male Golden- cheeked Warblers averaged 1.5 times larger, in biking than non-biking sites (Table 1). Territories ranged from 0.48 to 7.27 ha in biking sites and 0.24 to 4.33 ha in non-biking sites. Trails occupied twice the area within territories of Golden-cheeked Warbler males in biking than Davis el al. • MOUNTAIN BIKING AND GOLDEN-CHEEKED WARBLERS 469 TABLE 1 . Mean ± SE for minimum convex polygon (MCP) territory size (ha), trail area (ha) within MCPs, and habitat composition (%) within MCPs for Golden-cheeked Warblers at biking and non-biking sites near Austin and on Fort Hood, Texas, during spring 2()02-2()()3. Variable Biking III = ,20) Non-biking (« = 61) F- p Territory size (MCP) 2.16 ± 0.19 1.43 ± 0.16 6.01 0.016 Trail area within MCP 0.068 ± 0.009 0.029 ± 0.004 9.77 0.002 Habitat composition Developed 0.67 ± 0.42 0.00 ± 0.00 1.21 0.274 Deciduous 6.00 ± 1.50 3.48 ± 2.70 0.78 0.378 Juniper 92.00 ± 1.78 96.52 ± 2.70 2.05 0.155 Grassland 1.34 ± 0.83 0.00 ± 0.00 1.27 0.262 F and P values for analysis of variance comparisons between biking and non-biking sites: df = 1,89. non-biking sites (Table 1), Habitat composition of territories did not differ between sites (Table 1), Time Activity Budgets of Territorial Moles. — We made 400 total observations (1-7 observa- tions/male, 100-120 unique males) of male Golden-cheeked Warblers during the 2 years of study. Overall behavior did not differ between biking and non-biking sites (Wilks’ T = 0,97, P = 0,265; Table 2) and time periods (Wilks’ X = 0,95, P = 0.365), but differed between years (Wilks’ X = 0.87, P < 0.001). There were no significant two- or three-way interactions (Wilks’ X > 0.93, P > 0.116). Singing and provisioning nestlings were the only behaviors that differed between years. Male Golden-cheeked Warblers spent less time singing in 2002 than 2003 (8.3 vs. 16.0%) and more time provisioning nestlings in 2002 than 2003 (1.2 vs. 0.1%). Prey search behaviors of male Golden-cheeked Warblers did not differ between biking and non- biking sites (Wilks’ X = 0.97, P = 0.677; Table 3). The total number of search bouts during each observation period also did not differ between biking and non-biking sites (T|,93 = 0.03; P = 0.866; T ± SE = 26.2 ± 4.4 bouts vs. 27.2 ± 3.2 bouts). Male Golden-cheeked War- blers spent similar amounts of time engaged in attack techniques to capture prey at biking and non-biking areas (Table 4). Parental Behavior at Nests. — We video-record- ed 32 Golden-cheeked Warbler nests (17 in biking sites vs. 15 in non-biking sites) for 4,094.5 daytime hrs in 2002 and 2003 (incubation stage: 1.141.2 hrs for nests at biking sites and 828.2 hrs for nests at non-biking sites; nestling stage: 1,109.7 hrs for nests at biking sites vs. 1,015.4 hrs for nests at non-biking sites). Average recording time per nest was 110.6 ± 17.4 hrs (range = 15.2-432.8 hrs). Golden-cheeked War- blers spent similar amounts of time incubating eggs, brooding nestlings, and provisioning nest- lings in biking and non-biking sites (Table 5). The number of bouts and length of the bout for each of these behaviors also did not differ between biking and non-biking sites (Table 5). TABLE 2. Diurnal time activity budget (.v ± SE) for male Golden-cheeked Warblers at biking (n = 139) and non- biking (/; = 204) locations near Austin and on Fort Hood, Texas during spring 2002-2003. D.uy time (%) 2002 200.2 Behavior Biking Non-biking Biking Non-hiking Perching 37.5 -h 4.9 26.8 4- 3.5 29.9 4- 2.6 31.4 ± 2.2 Singing 7.8 0.9 8.7 -h 0.9 15.6 4- 1.3 16.3 ± 1.2 Preening 10.2 H- 3.3 7.6 4- 2.4 5.9 4- 1.3 6.7 ± 1.3 Agonistic 0.0 -f- 0.0 0.2 4- 0.1 0.6 4- 0.5 0.2 ± 0.1 Mate guarding 0.2 0.2 0.0 4- 0.0 0.2 4- 0.2 0.0 ± 0.0 Locomotion 38.8 5.1 49.7 4- 4.2 44.6 4- 3.1 42.0 ± 2.5 Foraging 4.0 -+■ 1.0 3.9 + 0.8 2.5 4- 0.5 3.0 ± 0.6 Food handling 0.9 -+■ 0.5 1.4 4- 0.7 0.6 4- 0.4 0.3 ± 0.2 Feeding fledgling 0.6 0.3 1.7 4- 0.6 0.1 4- 0.1 0.1 ± 0.1 470 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 3, September 2010 TABLE 3. Food searching behaviors (x ± SE) for male Golden-cheeked Warblers at biking {n = 41) and non- biking {n = 54) sites near Austin and on Fort Hood, Texas during spring 2003. Percent of total food searching time Behavior Biking Non-biking Hop 59.4 ± 5.0 59.8 ± 4.0 Short flight. <20 cm 10.7 ± 1.9 11.6 ± 2.2 Medium flight, 20 cm-2 m 5.4 ± 1.2 7.4 ± 1.3 Long flight, >2 m 24.5 ± 5.0 21.1 ± 4.3 Nest Success. — We monitored 33 nests in biking sites and 22 nests in non-biking sites. Twenty-one (64%) and 19 nests (86%) success- fully fledged at least one host young in biking and non-biking sites, respectively. Seven nests failed due to predation in biking sites, compared to two in non-biking sites. Texas rat snakes {Elaphe obsoletci) accounted for most of the depredation (71%) in biking sites, and American Crows {Corviis branchy tynchos) and fire ants {Solenop- sis invicta) each accounted for one depredation in biking sites. Texas rat snakes and Cooper’s Hawks (Accipiter cooperii) accounted for the two depredations in non-biking sites. Nest aban- donment was three times higher in biking sites (5 nests; 15%) than non-biking sites (1 nest, 5%). The abandonments occurred either late in the laying stage or early in the incubation stage. We did not document parasitism by Brown-headed Cowbirds. Daily nest survival for biking sites was 0.964 ± 0.018 for the incubation stage and 0.945 ± 0.019 for the nestling stage, whereas DNS for non-biking sites was 0.986 ± 0.014 for the incubation stage and 0.981 ± 0.013 for the nestling stage. The Mayfield nest success estimate in non-biking sites (70%) was twice as high as in biking sites (35%). DISCUSSION Our results suggest mountain biking trail use negatively impacted Golden-cheeked Warblers during the breeding season. Male territories were >1.5 times larger and nest success was substan- tially lower in biking than non-biking sites. Nest abandonment and predation were also higher in biking than non-biking sites. Several studies have suggested that habitat fragmentation negatively impacts Golden-cheeked Warbler nest survival (Fink 1996; Maas-Burleigh 1998; Peak 2007; Reidy et al. 2008, 2009). Peak (2007) reported nest survival of Golden-cheeked Warblers was lower in woodlands with high forest edge. Similarly, Reidy et al. (2009) reported that as edge factors including open edge density and amount of open edge increased, nest survival of Golden-cheeked Warblers substantially declined. Golden-cheeked Warbler temtories in biking sites in our study occurred in habitats more fragmented by trails than temtories in non-biking sites. It appears mountain biking indirectly affected Gold- en-cheeked Warblers by habitat alteration and fragmentation caused by occurrence of mountain biking trails. Territory size in many bird species is inversely related to food abundance (Mahler 1970, Gass et al. 1976, Salmonson and Baida 1977, Smith and Shugart 1987). The larger territories observed in biking sites in our study suggests that habitat quality was poorer in these sites than in non- biking sites. Fragmentation and alteration of habitats in biking sites may have reduced the quality of these habitats for Golden-cheeked Warblers. For example, arthropod abundance may have been low within forest edges and near biking trails because of altered microclimate and vegetation structure (Jokimaki et al. 1998, Kilgo 2005). However, Davis and Leslie (2008) reported TABLE 4. Foraging attack techniques {x ± SE) used by male Golden-cheeked Warblers at biking (n = 16) and non- biking (n = 21 ) sites near Austin and on Fort Hood, Texas during spring 2003. Percent Attack technique Biking Non-biking P‘ Gleaning Reaching Hanging Sally-.striking Sally-hovering Flutter-chasing 57.4 ± 10.1 47.4 ± 8.9 0.555 13.4 ± 7.2 29.3 ± 8.1 0.192 3.1 ± 3.1 4.3 ± 4.3 0.823 5.2 ± 3.6 1.1 ± 1.1 0.323 20.8 ± 8.6 16.1 ± 6.7 0.506 0.0 ± 0.0 1.7 ± 1.2 0.232 “ /'-value from Kru.skal-Wallis tc.st with I df. Davis et a!. • MOUNTAIN BIKING AND GOLDEN-CHEEKED WARBLERS 471 TABLE 3. Parental behavior (.v ± SE) for female Golden-cheeked Warblers nesting non-biking sites near Austin and on Fort Hood, Texas during daylight hours in spring 2002 and 200.3. in biking and Parameter Biking Non-biking /" Incubation stage*’ Incubation. % 67.40 ± 0.83 74.84 ± 0.93 0.105 No. of incubation bout.s/day 17.98 ± 2.38 14.05 ± 2.12 0.355 Length of incubation bout, min 32.39 ± 6.86 38.44 ± 3.22 0.298 Early nestling stage*" Brooding, % 36.59 ± 8.47 25.46 ± 7.42 0.257 No. of brooding bouts/day 16.82 ± 2.93 24.60 ± 8.06 0.345 Length of brooding bout, min 16.63 ± 5.72 12.60 ± 4.38 0.165 Feeding nestlings, % 7.46 ± 1.49 6.02 ± 0.93 0.852 No. feeding bouts/day 67.47 ± 17.07 79.65 ± 22.75 0.710 Length of feeding bout, min 1.38 ± 0.22 1.58 ± 0.48 0.850 Late nestling stage*" Feeding nestlings, % 4.93 ± 1.08 4.25 ± 0.71 0.446 No. of feeding bouts/day 62.22 ± 9.91 76.24 ± 18.04 0.608 Length of feeding bout, min 1.19 ± 0.19 0.87 ± 0.20 0.440 ““ P-value from Kruskal-Wallis test with 1 df. Sample sizes for incubation stage, early nesting stage, and late nestling stage in biking and non-biking sites were 8 and 7, 7 and 8, and 9 and 11 nests, respectively. that arthropod abundance from sweep-net samples of tree branches in biking and non-biking sites did not differ, and our study found no difference in foraging behaviors between biking and non-biking sites, which suggests that habitat quality was similar between biking and non-biking sites. Davis and Leslie (2008) did not distinguish between samples collected from forest edges or near biking trails and the impact of habitat fragmentation and alteration on arthropod abun- dance in biking sites is unknown. Male Golden-cheeked Warblers in our study spent similar amounts of time engaged in diurnal behaviors in biking and non-biking sites. Nesting behaviors of female Golden-cheeked Warblers also did not differ between biking and non-biking sites. These observations suggest disturbance from mountain biking does not affect daily activity budgets of Golden-cheeked Warblers, but our direct observations of Golden-cheeked Warbler encounters with mountain bikers did show the birds were disturbed by bikers. During three of the four observed encounters with bikers, the bird flushed >20 m in response to the passing biker. Moreover, three of five nests abandoned at biking sites were <2 m from the trail, suggesting disturbance from mountain biking has a role in nest abandonment. We cannot conclude that direct disturbance from mountain biking is benign to nesting Golden-cheeked Warblers. Several studies have shown that direct human disturbance can alter activity budgets of birds (Stalmaster 1983, Kaiser and Fritzell 1984, Burger 1994, Mikola et al. 1994, Knapton et al. 2000). Our limited number of observations of direct encounters of Golden-cheeked Warblers and mountain bikers likely does not indicate that such encounters are rare, but rather, they are difficult to observe. We expect variations in biker behaviors (e.g., speed, group sizes, and silent vs. talking bikers) will elicit variable responses in warblers. These factors should be considered when evaluating the impact of mountain biking on Golden-cheeked Warblers. Our estimate of nest success for Golden- cheeked Warblers in biking sites was lower than in non-biking sites (35 vs. 70%), but our nest success estimate for biking sites is comparable to other studies from Fort Flood (45% [Stake et al. 2004], 34% [Peak 2007], 40% [Reidy et al. 2009]) and Austin (40% [Reidy et al. 2009]). This lower nest success e.stimate in biking sites suggests those sites may provide lower quality nesting habitats for Golden-cheeked Warblers compared to non-biking sites. The nest success estimate for biking sites is still above the rate necessary to balance juvenile and adult mortality (25-30%; Donovan and Thompson 2001). Reidy et al. (2008) showed that Golden-cheeked Warblers can successfully reproduce in large protected urban sites, but they also suggested these habitats 472 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 3. September 2010 may not be able to sustain productive populations if further disturbance in the landscape or breeding patch continues. This also appears to be the case for biking sites in our study, which are currently able to maintain viable populations of Golden- cheeked Warblers. Additional use, fragmentation, and alteration of these habitats could impact productivity by affecting densities of breeding adults, pairing success, or predator abundance which may result in these habitats becoming inadequate for sustaining viable populations (Reidy et al. 2008). Reidy et al. (2009) suggested higher predator abundance and activity near edges contributed to low nest survival of Golden-cheeked Warblers near edges. The main predator of Golden-cheeked Warbler nests in our study was the Texas rat snake, which is a skilled tree climber (Pierce et al. 2008, Sperry et al. 2009) and major nest predator of Golden-cheeked Warblers (Stake et al. 2004, Reidy et al. 2008). Texas rat snakes prefer fragmented and edge habitats (Blouin-Demers and Weatherhead 2001, Sperry et al. 2009), and respond positively to increased edge and fragmentation (Sperry et al. 2009). Thus, it is likely that habitat fragmentation and alteration of biking sites by trails increased vulnerability of Golden-cheeked Warbler nests in those areas to predation by rat snakes and other edge-adapted predators. Sperry et al. (2009) suggested that Golden-cheeked Warbler nests that occurred in areas with many canopy openings, such as areas traversed by mountain bike trails, may be more at risk to rat snake predation. The direct impact of mountain biking on Golden-cheeked Warblers may be minimal, but the indirect impact from fragmentation and alteration of habitats from mountain bike trails may reduce the quality of nesting habitat for Golden-cheeked Warblers. In particular, Golden- cheeked Warblers nesting in habitats fragmented and altered by mountain biking trails may be more vulnerable to nest predation (Reidy et al. 2009) and possibly encounter lower prey abundance (Jokimaki et al. 1998, Kilgo 2005). Conservation efforts that curtail construction of new mountain biking trails in Golden-cheeked Warbler habitat and reduce the amount of forest open edge habitat created by existing mountain biking trails should promote recovery objectives. ACKNOWLEDGMENTS Funding was provided by the U.S. Army Fort Hood Military Base and U..S. Army Corps of Engineers Construction Engineering Research Laboratory. The U.S. Army Corps of Engineers Construction Engineering Research Laboratory, Jones Technologies, and the Okla- homa Cooperative Fish and Wildlife Research Unit (Oklahoma State University, Oklahoma Department of Wildlife Conservation, U.S. Geological Survey, U.S. Fish and Wildlife Service, and Wildlife Management Institute cooperating) administered the grant and provided logistical support. We also thank The Nature Conservancy of Fort Hood Field Office for their support. We appreciate the support of J. D. Cornelius, D. M. Herbert, T. J. Hayden, and C. E. Pekins during data collection at Eort Hood field sites, and D. M. Koehler and C. M. Abbruzzese for support during data collection at Austin field sites. We thank J. L. Sterling, M. M. Stake, K. T. Cutrera, V. D. Bump, and H. M. Oswald along with numerous volunteers for data collection; A. Reisch for recording parental behavior from video tapes; B. D. Hebert and T. J. Richardson for recording male behavior from microcassettes and entering behavior data; A. J. Eritchie for sorting and identifying arthropods; and A. D. Thomas for data entry. LITERATURE CITED Altmann, j. 1974. Observational study of behavior: samplings methods. Behaviour 49:227-265. Barker, H. R. and B. M. Barker. 1984. Multivariate analysis of variance (MANOVA): a practical guide to its use in scientific decision making. University of Alabama Press, Birmingham, USA. Bibby, C. j. and N. D. Burgess. 1992. Bird census techniques. Academic Press, Harcourt Brace and Company, London, United Kingdom. Blouin-Demers, G. and P. J. Weatherhead. 2001. Habitat use by black rat snakes (Elaphe obsoleta obsoleta) in fragmented forests. Ecology 82:2882- 2896. Burger, J. 1994. The effect of human disturbance on foraging behavior and habitat use in Piping Plover {Charadriiis melodus). Estuaries 17:695-701. Chavez, D. J., P. L. Winter, and J. M. Baas. 1993. Recreational mountain biking: a management perspec- tive. Journal of Park and Recreation Administration 11:29-36. Davis, C. A. and D. M. Leslie Jr. 2008. Effects of mountain biking activity on foraging and nesting behavior of Golden-cheeked Warblers. Final Report, Project Number MIPR #W81EWF63530473. U.S. Army Corps of Engineers Research and Development Center, Construction Engineering Research Laborato- ry, Champaign, Illinois, USA. Donovan, T. M. and E. R. Thomp.son 111. 2001. Modeling the ecological trap hypothesis: a habitat and demo- graphic analysis for migrant songbirds. Ecological Applications 11:871-882. Environmental Systems Research Institute (ESRI). 2000. ArcView GIS, Version 3.2a. Environmental Systems Re.search Institute, Redlands, California, USA. Eink, M. L. 1996. Factors contributing to nest predation within habitat of the Golden-cheeked Warbler, Travis Davi.s- Cl a/. • MOUNTAIN IMKING AND GOLDEN-Ci lEEKTD WARBEERS 473 County. Texas. Tliesis. Texas A&M University, College Station, USA. Fl.ather. C. H. .^no II. K. CORDELl,. 1995. Outdoor reereation: historieal and antieipated trends. Pages 3- 16 in Wildlife and reereationists: eoexistenee through management and researeh (R. U. Knight and K. ,1. Gutzwiller, Editors). Island Press, Washington, D.C., USA. Gass, C. L., G. Angeher, and J. Centra. 1976. Regulation of food supply by feeding territoriality in the Rufous Hummingbird. Canadian Journal of Zoology 54:2046- 2054. GtLL, F. B. 2007. Ornithology. Third Edition. W. H. Freeman and Company, New York, USA. Johnson, D, H. 1979. Estimating nest success: the Mayfield method and an alternative. Auk 96:651-661. Johnson, R. A. and D. W, Wichern. 1988. Applied multivariate statistical analysis. Second Edition. Pren- tice Hall, Englewood Cliffs, New Jersey, USA. JOKIMAKI, J., E. HUHTA, J. iTAMtES, AND P. RaHKO. 1998. Distribution of arthropods in relation to forest patch size, edge, and stand characteristics. Canadian Journal of Forest Re.search 28:1068-1072. Kaiser. M. S. and E. K. Fritzell. 1984. Effects of river reereationists on Green-backed Heron behavior. Jour- nal of Wildlife Management 48:561-567. Kilgo, j. C. 2005. Harvest-related edge effects on prey availability and foraging of Hooded Warblers in a bottomland hardwood forest. Condor 107:627-636. Knapton, R. W.. S. a. Petrie, and G. Herring. 2000. Human disturbance of diving ducks on Long Point Bay, Lake Erie. Wildlife Society Bulletin 28:923-930. Knight, R. L. and K. J. Gutzwiller (Editors). 1995. Wildlife and reereationists: coexistence through man- agement and re.search. Island Press, Washington, D.C., USA. Ladd, C. and L. Gass. 1999. Golden-cheeked Warbler (Dendroicci chrysoparia). The birds of North America. Number 420. Laundre, j. W. and B. L. Keller. 1984. Home-range size of coyotes: a critical review. Journal of Wildlife Management 48: 1 27- 1 39. Maas-Burleigh. D. S. 1998. Factors influencing demo- graphics of Golden-cheeked Warblers (Dendroicci chrysoparia) at Fort Hood Military Reservation, Texas. Thesis. University of Oklahoma, Norman, USA. Mahler, W. J. 1970. The Pomarine Jaeger as a brown lemming predator in northern Alaska. Wil.son Bulletin 82:130-157. Martin, P. and P. Bate.son. 1986. Measuring behavior: an introductory guide. Cambridge University Press, New York, USA. Mayfield, H. 1961. Nesting success calculated from exposure. Wil.son Bulletin 73:255-261. Mayfield, H. F. 1975. Suggestions for calculating nest success. Wilson Bulletin 87:456-466. Mikola, J., M. Miettinen, E, Lehikoinen, and K. Lehtila. 1994. The effects of disturbance caused by boating on survival and behavior of Velvet Scoter Melanilta fusca ducklings. Biological Conservation 67:1 19-124. Miller, J. R. and N. T. Hohils, 2000. Recreational trails, human activity, and nest predation in lowland riparian areas. Landscape and Urban Planning 50:227-236. Miller, S. G., R. L. Knight, and C. K. Miller. 1998. Innuence of recreational trails on breeding bird communities. Ecological Applications 8:162-169. Mosedale, j. 2002. Mountain biking in the Canadian Rocky Mountains: a situational analysis. Mountain Forum E-consultation for the UNEP/Bischkek Global Mountain Summit. Bishkek, Kyrgystan. Peak, R. G. 2007. Forest edges negatively affect Golden- cheeked Warbler nest survival. Condor 109:628-637. Pierce, J. B., R. R. Fleet, L. McBrayer, and D. C. Rudolph. 2008. Use of trees by Texas rat snake (Elaplie ohsoleta) in eastern Texas. Southeastern Naturalist 7:359-366. Pike Masteralexis, L., C. A. Barr, and M. A. Hums. 2008. Principles and practice of sport management. Third Edition. Jones and Bartlett Publishers, New York, USA. Ralph, C. J., G. R. Geupel, P. Pyle, T. E. Martin, and D. F. Desante. 1993. Handbook of field methods for monitoring landbirds. USDA, Forest Service, General Technical Report PSW-GTR-144. Pacific Southwest Research Station, Albany, California, USA. Reidy, j. L., M. M. Stake, and F. R. Thompson III. 2008. Golden-cheeked Warbler nest mortality and predators in urban and rural landscapes. Condor 110:458-466. Reidy, J. L., F. R. Thompson, and R. G. Peak. 2009. Factors affecting Golden-cheeked Warbler nest sur- vival in urban and rural landscapes. Journal of Wildlife Management 73:407^13. Remsen, j. V. AND .S. K. Robinson. 1990. A classification scheme for foraging behavior of birds in terrestrial habitats. Studies in Avian Biology 13: 144-160. Robinson, S. K. and R. T. Holmes. 1982. Foraging behav- ior of forest birds: the relationships among search tactics, diet, and habitat stmeture. Ecology 63: 1918-1931. Rodgers, A. R. and A. P. Carr. 1998. HRE: the home range extension for AreView™: user's manual. Beta te.st Version 0.9, July 1998. Ontario Mini.stry of Natural Resources, Centre for Northern Forest Eco- system Research, Thunder Bay, Ontario, Canada. Salmonson, M. G. and R. P. Balda. 1977. Winter territoriality of Townsend's Solitaires (Myadestes townsendi) in a pinyon-juniper-ponderosa ecosystem. Condor 49: 148-161. Seaman, D. E., J. J. Millspaugh, B. J. Kernohan, G. C. Brundige, K. j. Raedeke, and R. a. Gitzen. 1999. Effects of sample size on kernel home range estimates. Journal of Wildlife Management 6,3:739-747. Smith, T. M. and H. H. Shugart. 1987. Territory size variation in the Ovenbird: the role of habitat structure. Ecology 68:695-704. Sperry, J. H., D. A. Cimprich, R. G, Peak, and P. J. Weatherhead. 2009. Is nest predation on two endangered bird species higher in habitats preferred by snakes? Ecoscience 16:1 1 1-1 18. .Stake, M. M. 2000. Using video to identify Black-capped Vireo nest predators at Fort Hood, Texas. Endangered 474 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 3, September 2010 species monitoring and management at Fort Hood, Texas: 2000 Annual Report. The Nature Conservancy of Texas, Fort Hood Project, Austin, USA. Stake, M. M. and D. A. Cimprich. 2003. Using video to monitor predation at Black-capped Vireo nests. Condor 105:348-357. Stake, M. M., J. Faaborg, and F. R. Thompson III. 2004. Video identification of predators at Golden-cheeked Warbler nests. Journal of Field Ornithology 75:337- 344. Stalmaster, M. V. 1983. An energetics simulation model for managing wintering Bald Eagles. Journal of Wildlife Management 47:349-359. Symmonds, M. C., W. E. Hammitt, and V. L. Quisen- BERRY. 2000. Managing recreational trail environ- ments for mountain bike user preferences. Environ- mental Management 25:549-564. SYSTAT. 2004. SYSTAT 11. SYSTAT Software Incorpo- rated, San Jose, California, USA. Taylor, A. R. and R. L. Knight. 2003. Responses of wildlife to human activity: terminology and methods. Wildlife Society Bulletin 31:1263-1271. U. S. Department OF Interior (USDI). 1993. Biological opinion. USDI, Eish and Wildlife Service, Austin, Texas, USA. Wunderle, J. M. and S. C. Latta. 1998. Avian resource use in Dominican shade coffee plantations. Wilson Bulletin 110:271-281. The Wilson Jotinuil of Oniilliology 122(3);475-483, 2010 SURVIVAL, SITE FIDELITY, AND POPULATION TRENDS OF AMERICAN KESTRELS WINTERING IN SOUTHWESTERN FLORIDA DANIEL M. HINNEBUSCH,' ^ JEAN-FRANgOlS THERRIEN,' ^ MARC-ANDRE VALIQUETTE,' - BOB ROBERTSON, SUE ROBERTSON,’ AND KEITH L. BILDSTEIN' ABSTRACT. — The winter population ecology of American Kestrel (Falco span’erius), one of the most abundant and widely distributed raptors in North America, is poorly understood. We systematically searched a 225-km^ area in Cape Coral, southwestern Florida, for American Kestrels during 14 winters (1 Dec- 15 Mar, 1994-2008) to measure their annual apparent survival and to see il individuals returned to the same wintering area. We recaptured 101 of 2,958 banded kestrels during the study. We estimated annual apparent survival of 75% for males and 74% for females using a Cormack-Jolly- Seber model. These estimates are considerably higher than previous estimates for American Kestrels, but are similar to estimates reported for other species of Falco. Forty-six percent of the kestrels estimated to have survived were observed in the study area 1 year after recapture, based on year-specific color banding. All but six of 101 kestrels were recaptured within 1 km ot where they were banded. Four of five kestrels banded as nestlings and subsequently recaptured in the study area were banded in southeastern Pennsylvania, suggesting migratory connectivity. Eighty percent of the kestrels trapped were females, but the proportion of females decreased annually (-3 ± 1% per year). Overall, the population decreased by an average ot 7 ± 2% per year. Recent land-use change accompanied by increased human density and suburban expansion may be causing the observed trends. Received 29 October 2009. Accepted 9 March 2010. American Kestrels [Falco sparverius) occur in open habitats throughout most of North and South America (Smallwood and Bird 2002). The American Kestrel, historically abundant through- out most of its range, has declined significantly since the 1970s (Sullivan and Wood 2005, Farmer et al. 2008). Potential causes for this decline include West Nile virus, organophosphate poison- ing, increased predation by growing populations of Cooper’s Hawks [Accipiter cooperii), popula- tion declines of Northern Flickers [Colaptes auratus), a primary cavity nester; and land-use changes (Sullivan and Wood 2005, Farmer et al. 2006, Medica et al. 2007, Farmer et al. 2008). American Kestrels are partial migrants that undertake leapfrog migration (Roest 1957, Small- wood and Bird 2002). Individuals breeding at northern latitudes are more likely to migrate and move greater distances than individuals breeding farther south (Goodrich and Smith 2008). Banding returns indicate that most individuals that migrate to southwestern Florida breed in the northeastern United States or in southeastern Canada (Layne 1982). Individuals or, more rarely, pairs of ' Hawk Mountain Sanctuary, Acopian Center for Con- servation Learning, 410 Summer Valley Road, Orwigsburg, PA 17961, USA. -Universite Laval, Departement de biologie and Centre d'etudes nordiques, Quebec, QC GIK 7P4, Canada. ^ Deceased. ■'Corresponding author; e-mail: hinnebusch@hawkmtn.org and ospreyl984@yahoo.com kestrels maintain areas of exclusive use during winter (Cade 1955, Mills 1975). Males and females use different habitats in winter with males occupying areas with more woody vegeta- tion (Smallwood and Bird 2002). This may be the result of later-migrating males selecting subopti- mal habitat left vacant by females (Smallwood 1988), or because larger females dominate smaller males (Ardia and Bildstein 1997). Site fidelity by migratory birds can be advan- tageous because individuals benefit from being familiar with a previously occupied teiritory (Spaans 1977, Nichols et al. 1983). Familiarity with a territory likely allows for more efficient foraging and predator avoidance (Hinde 1956, Clarke et al. 1993, Yoder et al. 2004). Evidence supporting winter site fidelity in American Kestrels is limited to studies of small numbers of band recoveries and observation of uniquely marked individuals. Return rates of 19% (four of 21 individuals) and 17% (eight of 47) were reported at wintering sites in central Ohio (Mills 1975) and .southern Texas (Bolen and Derden 1980), respectively, for kestrels marked with patagial tags. Tabb (1977) reported a return rate of 2% (17 of 842) for kestrels banded at several sites in southern Florida, but that estimate is based entirely on recaptures because individuals could not be reliably identified without recapture. Several other species of raptors demonstrate winter site fidelity (Garrison and Bloom 1993, Harmata and Stahlecker 1993, Powers 1996). 475 476 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. 3. Seplemher 2010 Moreover, satellite tracking has provided evi- dence for winter site fidelity by North American falcons, including Peregrine Falcon (Falco pere- grimis) and Prairie Falcon (F. mexicamis), based on a small number of individuals (McGrady et al. 2002, Steenhof et al. 2005). Survival estimates are essential for effective species’ management and conservation (Crone 2001, Sandercock 2006). However, survival in the wild remains poorly known for the American Kestrel (Smallwood and Bird 2002). To date, two studies have estimated mortality rates using band recovery and territory occupancy (Roest 1957, Henny 1972), a technique that does not consider the probabilities of capture and detection biases and is further biased by territory vacancies that may not be the result of deaths of birds (Gould and Fuller 1995). We report the results of a 14-year study of American Kestrels wintering in southwestern Florida. Specifically, we: ( 1 ) estimate annual apparent survival, (2) discuss a decrease in size of the observed population characterized by a decrease in number of females, (3) provide evidence for strong winter site fidelity, and (4) discuss evidence for migratory connectivity be- tween winter territories in Florida and natal sites in the northeastern United States. METHODS Study Area. — We marked American Kestrels within —275 km^ containing 2,966 km of roads in the city of Cape Coral and adjacent Pine Island, west of Fort Myers, Lee County, Florida. The study site is on a peninsula along the Gulf of Mexico bordered to the east by the Caloosa- hatchee River and to the west by Matlacha Pass. Native cover includes slash pine (Pinii.s elliottii) woodlands and tidal marshes (Wesemann 1986, Millsap and Bear 2000). An extensive network of canals and roads was built in the area in the late I95()s (Bernard 1983, Zeiss 1983). The site now consists largely of single-family residences and unoccupied, but regularly mowed, lots (Millsap and Bear 2000). Parallel roads are -100 m apart. The human population in 2000 was 102,286 (376.1/knr). In 2006 it had grown 48% to 151,389 (556.7/km^) (USCB 2009). Data Collection.— BR and SR .searched Cape Coral for American Kestrels each winter ( 1 Dec- 15 Mar) for 14 years beginning in December 1994. Observers drove along each road in the study area at <25 km/hr while .searching for kestrels, which often hunted from prominent perches and were easily observed (Bildstein and Collopy 1987). We surveyed each road once each winter and recorded male/female, location, and band color of all kestrels observed on our initial visit to each road. We attempted to capture every kestrel observed by using a bal-chatri trap (25-cm-diameter, 7-cm- high) baited with two mice {Miis muscidus) (Berger and Mueller 1959). Unless previously banded, captured kestrels were marked with a USGS aluminum band and a year-specific plastic, coil-style, color band (red, blue, white, yellow, black, or green), except for 1994-1995 and 1995- 1996, when color bands were not used. Each color, except white, was used in 2 nonconsecutive years with bands being placed on different legs in different years. Wrap-around color bands were manufactured by Gey Band & Tag Company, Norristown, Pennsylvania, USA. A body-condi- tion index was calculated for each captured bird by dividing the cube root of body mass by wing chord, which was used as a wing loading index for Merlins {Falco cohunbarius) by Temple ( 1972). If we were unable to capture a bird, we returned to the same site later in the season, usually in early March, and again attempted to trap it. However, observations of kestrels observed on a road after our initial survey of that road were not included in analyses. We indicated sites at which we banded or observed kestrels as the nearest intersection of roads. We obtained information on banding and recovery locations outside the Cape Coral study area from the Bird Banding Laboratory, USGS, Patuxent Wildlife Research Center, Laurel, Mary- land, USA. We considered recovered banded kestrels to have been found during the breeding .season if they were found between 1 April and 3 1 August {sen. 0.10. Both apparent survival and recapture probability were allowed to vary with year and sex, as well as an interaction between year and sex, in the global model. We tested all possible models nested within the global model. The best model not included in model averaging [0(sex)/p(year + sex)| had AAIC<. = 2.25 and vt', = 0.13. Model AIC, AAIC.. Parameters Variance 0(.)/p(year) 1209.07 0.00 0.39 14 1 19.36 4)(.)/p(year + sex) 1209.64 0.57 0.30 15 1 17.92 0(sex)/p(year) 1210.72 1.65 0.17 15 1 18.99 recovered during the breeding season, although we were not able to ascertain if kestrels died in breeding areas or as migrants. Kestrels recovered during the breeding season included five in Quebec, one in New Brunswick, one in Maine, two in New York, two in Pennsylvania, one in Delaware, and one in South Carolina. Kestrels recovered during the non-breeding season includ- ed one each in New Brunswick, Ontario, New York, North Carolina, and South Carolina. No birds captured in Cape Coral were recaptured outside the study area and reported to the Bird Banding Laboratory. DISCUSSION The annual apparent survival estimates of 75% for males and 74% for females are considerably higher than those reported by Roest (1957) and Henny (1972) (43 and 55%, respectively). Both of these studies relied on band recoveries from dead birds or nest-site occupancy to evaluate annual mortality rates. Their estimates of annual apparent survival are based on a time when direct persecution by humans (Bildstein 2008) may have resulted in greater mortality. At least 24% of the recovered kestrel bands reported by Roest (1957) came from birds killed by people. Human-caused mortality of kestrels reported in Henny (1972) ranged from 16% for 1958-1965 to 48% for 1925-1945. Our estimates are similar to estimates reported for other species of Falco in North America, including 75-89% for Prairie Falcons (Enderson 1969, Steenhof et al. 2006), 62% for Merlins (Lieske et al. 2000), and 79-93% for Year FIG. 1. Number of observations of female (circles), male (.squares), and total (triangles) American Kestrels in Cape Coral, southwestern Florida, 1996-2008. “n" indicates number of kestrels and “Y” indicates the calendar year beginning in a winter season in the regression equations. 480 THE WILSON JOURNAL OL ORNITHOLOGY • VuL 122, No. 3, September 2010 FIG. 2. Proportion of color-marked American Kestrel cohorts observed in Cape Coral, southwestern Florida and the estimated proportion surviving each year. Points indicate the average proportion of marked cohorts observed in the study area after each year. Vertical lines are 95% confidence intervals. The curve indicates the estimated surviving population based on annual adult apparent survival estimates from the CIS model in this study. Peregrine Falcons (Court et al. 1989, Tordoff and Redig 1997). Survival estimates for Common Kestrels (Falco tinnunciiliis) in the United King- dom were 80% for males and 66% for females (Dobson 1987). To our knowledge, this study reports the first estimates of annual apparent survival for American Kestrels using a CJS model. We recognize use of year-specific color bands is not the most effective method to estimate annual apparent survival, recapture probability, and between-year inter-capture distances because it did not allow us to record individual encounter histories for kestrels that were not recaptured. Our estimate of apparent survival with the CJS model was based only on the 101 recaptured kestrels. We believe this estimate is reliable because it considers the low recapture probability. The decline in the numbers of kestrels observed in the study area does not seem to be attributable to a decrease in adult survival over the study period as no year effect was retained in the models. However, the decrease in numbers of kestrels observed started in 2002, 3 years after discovery of West Nile virus in New York (Lanciotti et al. 1999). American Kestrels are susceptible to infection by West Nile virus (Komar et al. 2003) and exhibit symptoms that may decrease survival in wild birds (Nemeth et al. 2006). Ninety-five percent of tested American Kestrels breeding in Pennsylvania had antibodies indicating they had been exposed to the virus 03 ■g > TD C a) n E Z3 Inter-capture distance (m) FIG. 3. Di.stances between capture and recapture sites between winter seasons for American Kestrels in Cape Coral, southwestern Florida. The greatest inter-capture distance was 7.9 km. Hiimehiisch cl a/. • WINTERINC} AMERICAN KESTRELS 481 (Medica et al. 2007). The population decrease in our study area was characterized by a decrease in females. We believe this change in sex ratio could be due to land-use changes at the study site. Female kestrels occupy more open habitat in winter than males and, over the course of our study, land use in our study area shifted from open areas to more woody cover due to suburban development. The increa.se in number of males observed in the study area was relatively small compared to the decrease in the number of females observed. The observed increase in human population in the study area, combined with the suburban expansion of Cape Coral, has reduced the occurrence of good wintering habitat for kestrels. Those two factors seem to negatively affect the overall Cape Coral kestrel population as well as skew the sex ratio toward males. American Kestrels wintering in Cape Coral, Florida exhibited strong winter site fidelity. Most kestrels that returned to the study area wintered a short distance (<1 km) from their previous wintering area, based on between-year inter- capture distances. The relatively large numbers of color-banded birds we saw, together with a small recapture rate, suggest the small number of recaptures was not due to emigration but rather to capture avoidance. Our observations of numerous color-banded kestrels in the area support strong winter site fidelity by American Kestrels. Our estimate of the proportion of birds observed in winters subsequent to capture is likely conservative due to loss of color bands (Nekson et al. 1980). Seven percent of the ke.strels observed in the last year of our study (2007-2008) had an aluminum band and no color band. It is unlikely these kestrels had been banded in years when we did not use color bands (1994-1995 and 1995-1996) because we estimate that >96% of the kestrels banded in those years had already died by 2007-2008. 482 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 3, September 2010 Four of five kestrels recaptured after being banded as nestlings outside the study area were banded within 42 km of each other in southeastern Pennsylvania, suggesting migratory connectivity in populations we studied. American Kestrels migrate singly or in small groups (Kerlinger et al. 1985), and juveniles migrate earlier than adults (Smallwood 1988, Mueller et al. 2000); thus. Juveniles do not have the opportunity to follow adults on migration, as has been observed for some flocking migrants (Maransky and Bildstein 2001). Strong connectivity between natal and wintering areas has important conservation impli- cations because a decrease in productivity in a local breeding population can drastically affect the population in wintering areas with which it is connected and vice versa. Recoveries of kestrels banded in the study area were distributed over a much wider range than were banding locations of nestlings recaptured in the study area, providing little evidence of connectivity between wintering and breeding areas for adults. We could not ascertain whether a recovered bird had died in its breeding area or as a migrant in most cases. We cannot clearly state there is a lack of migratory connectivity in adult kestrels. ACKNOWLEDGMENTS The unexpected death of Bob Robert.son provides us with the opportunity to recognize his endless devotion of time and energy to the study of American Kestrels. Danny Bystrak, at the Bird Banding Laboratory, provided information on banding and recovery locations for kestrels outside our study area. We especially appreciate the comments of C. E. Braun, C. J. Farmer, Brian Millsap, and Karen Steenhof on earlier versions of this manuscript. This is Hawk Mountain Sanctuary contribution to con.ser- vation science number 189. LITERATURE CITED Ardia, D. R. and K. L. Bildstein. 1997. Sex-related differences in habitat .selection in wintering American Kestrels, Falca span'erius. Animal Behaviour 53: 1305-1311. Berger, D. D. and H. C. Mueller. 1959. The bal-chatri: a trap for the birds of prey. Bird-Banding 30:18-26. Bernard, E. 1983. Lies that came true: tall tales and hard sales in Cape Coral. Fla. Anna Publishing Inc., Ocoee, Florida. USA. Bii.dSTEIN. K. L. 2008. A brief history of raptor conservation in North America. Pages 5-36 in State of North America's birds of prey ( K. L. Bildstein, J. P. Smith. E. Ruelas Inzunza. and R. R. Veit, Editors). Nultall Ornithological Club, Cambridge. Massachu- setts. USA. and American Ornithologists’ Union, Washington. D.C.. USA. Bildstein, K. L. and M. W. Collopy. 1987. Hunting behavior of Eurasian (Falco tiimiinciilus) and Amer- ican Kestrels (F. span’erius): a review. Pages 66-82 in The ancestral kestrel (D. M. Bird and R. Bowman, Editors). Raptor Research Report Number 6. Bolen, E. G. and D. S. Derden. 1980. Winter returns of American Kestrels. Journal of Field Ornithology 51:174-175. Cade, T. J. 1955. Experiments on winter territoriality of the American Kestrel, Falco sparverius. Wilson Bulletin 67:5-17. Clarke, M. F., K. Burke da Silva, H. Lair, R. Pocklington, D. L. Kramer, and R. L. McLaugh- lin. 1993. Site familiarity affects escape behaviour of the eastern chipmunk. Taniias striatus. Oikos 66:533- 537. Court, G. S., D. M. Bradley, C. C. Gates, and D. A. Boag. 1989. Turnover and recruitment in a tundra population of Peregrine Falcons. Ibis 131:487^96. Crone, E. 2001. Is survivorship a better fitness surrogate than fecundity? Evolution 55:261 1-2614. Dobson, A. P. 1987. A comparison of seasonal and annual mortality for both sexes of fifteen species of common British birds. Omis Scandinavica 18:122-128. Enderson, j. H. 1969. Peregrine and Prairie falcon life tables ba.sed on band-recovery data. Pages 505-509 in Peregrine Falcon populations: their biology and decline (J. J. Hickey, Editor). University of Wisconsin Press. Madison, USA. Environmental System Research Institute (ESRl). 2009. ArcGIS. Version 9.3. Environmental System Research Institute, Redlands, California, USA. Farmer, C. J., L. J. Goodrich, E. Ruelas Inzunza, and J. P. Smith. 2008. Conservation status of North Amer- ica’s birds of prey. Pages 303M19 in State of North America’s birds of prey (K. L. Bildstein, J. P. Smith. E. Ruelas Inzunza, and R. R. Veit, Editors). Nuttall Ornithological Club, Cambridge, Massachusetts, USA, and American Ornithologists’ Union, Washington, D. C., USA. Farmer, G. C.. K. McCarty, S. Robertson, B. Robert- son, AND K. L. Bildstein. 2006. Suspected predation by accipiters on American Kestrels {Falco span>erius) in eastern Pennsylvania, U.S.A. Journal of Raptor Research 40:294-297. Garrison, B. A. and P. H. Bloom. 1993. Natal origins and winter site fidelity of Rough-legged Hawks wintering in California. Journal of Raptor Research 27:1 16-118. Goodrich L. J. and J. P. Smith. 2008. Raptor migration in North America. Pages 37-149 in State of North America’s birds of prey (K. L. Bildstein, J. P. Smith, E. Ruelas Inzunza, and R. R. Veit, Editors). Nuttall Ornithological Club, Cambridge, Massachusetts, USA, and American Ornithologists’ Union, Washington, D.C., USA. Gould, W. R. and M. R. Fuller. 1995. Survival and population size estimation in raptor studies: a comparison of two methods. Journal of Raptor Research 29:256-264. Harmata, a. R. and D. W. Stahlhcker. 1993. Fidelity of migrant Bald Eagles to wintering grounds in southern Iliimchiisch cl al. • WINTERING AMERICAN KESTRELS 483 Colorado and northern New Mexieo. Journal of Field Ornithology 64:129-134. Henny, C. J. 1972. An analysis of the population dynaniies ot seleeted avian speeies: with speeial reterence to ehanges during the eurrejit modern pesticide era. Wildlife Re.search Report I. USDI, Fish and Wildlife Service, Washington. D.C., USA. Hinde. R. a. 1956. The biological significance of the territories of birds. Ibis 98:340-369. Kerlinger, P., V. P. Bingman. and K. P. Able. 1985. Comparative flight behaviour of migrating hawks studied with tracking radar during autumn in central New York. Canadian Journal of Zoology 63:755-761. Komar, N., S. Langevin, S. Hinten, N. Nemeth, E. Edwards. D. Hettler. B. Davis, R. Bowen, and M. Bunning. 2003. Experimental infection of North American birds with the New York 1999 strain of West Nile virus. Emerging Infectious Diseases 9:31 1-322. Lanciotti, R. S., J. T. Roehrig, V. Deubel, J. Smith, M. Parker, K. Steele, B. Crise, K. E. Volpe, M. B. Crabtree, J. H. Scherret, R. A. Hall, J. S. Mackenzie, C. B. Cropp, B. Panigrahy, E. Ostlund, B. Schmitt, M. Malkinson, C. Banet, J. Weissman, N. Komar, H. M. Savage, W. Stone, T. McNamara, AND D. J. Gubler. 1999. Origin of the West Nile virus responsible for an outbreak of encephalitis in the northeastern United States. Science 286:2333-2337. Layne, J. N. 1982. Analysis of Florida-related banding data for the American Kestrel. North American Bird Bander 7:94-99. Lieske, D. j., I. G. Warkentin, P. C. James, U. W. Oliphant, and R. H. M. Espie. 2000. Effects of popu- lation density on survival in Merlins. Auk 1 17:184-193. Maransky, B. P. and K. L. Bildstein. 2001. Follow your elders: age-related differences in the migration behav- ior of Broad-winged Hawks at Hawk Mountain Sanctuary. Pennsylvania. WiLson Bulletin 1 13:350-353. McGrady, M. j., T. L. Maechtle. J. J. Vargas, W. S. Seegar, and M. C. Porras Pena. 2002. Migration and ranging of Peregrine Falcons wintering on the Gulf of Mexico coast. Tamaulipas, Mexico. Condor 104:39-48. Medica, D. L., R. Clauser, and K. Bildstein. 2007. Prevalence of West Nile virus antibodies in a breeding population of American Kestrels (Fatco sparverius) in Penn.sylvania. Journal of Wildlife Diseases 43:538- 541. Mills, G. S. 1975. A winter population study of the American Ke.strel in central Ohio. Wifson Bulletin 87:241-247. Millsap, B. a. and C. Bear. 2000. Density and reproduction of Burrowing Owls along an urban development gradient. Journal of Wildlife Manage- ment 64:33^1. Mueller, H. C.. N. S. Mueller. D. D. Berger, G. Allez, W. Robichaud, and j, L. Kaspar. 2000. Age and .sex differences in the timing of fall migration of hawks and falcons. Wifson Bulletin 112:214-224. Nelson, L. J.. D. R. Ander,son, and K. P. Burnham. 1980. The effect of band loss on estimates of annual survival. Journal of Field Ornithology 51:30-38. Nemeih, N., D. Gould, R. Bowen, and N. Komar. 2006. Natural and experimental West Nile virus infection in five raptor species. Journal of Wildlife Diseases 42: 1-13. Nichols, J. D., K. J. Reinecke, and J. E. Hines. 1983. Eactors affecting the distribution of Mallards wintering in the Mississippi Alluvial Valley. Auk 100:932-946. Powers, L. R. 1996. Wintering Sharp-shinned Hawks (Accipiter striatus) in an urban area of southwestern Idaho. Northwestern Naturalist 77:9-13. Roest, a. I. 1957. Notes on the American Sparrow Hawk. Auk 74:1-19. Sandercock, B. K. 2006. Estimation of demographic parameters from live-encounter data: a summary review. Journal of Wildlife Management 70: 1504-1520. SAS Institute Inc. 2003. Statistical analysis system. Version 9.1.3. SAS Institute Inc., Cary, North Carolina, USA. Smallwood, J. A. 1988. A mechanism of sexual segrega- tion by habitat in American Kestrels (Falco sparverius) wintering in south-central Florida. Auk 105:36^6. Smallwood, J. A. and D. M. Bird. 2002. American Kestrel (Falco sparverius). The birds of North America. Number 602. Spaans, a. L. 1977. Are starlings faithful to their individual winter quarters? Ardea 65:83-87. Steenhof, K., M. R. Fuller, M. N. Kochert, and K. K. Bates. 2005. Long-range movements and breeding dispersal of Prairie Falcons from southwest Idaho. Condor 107:481-496. Steenhof, K., K. K. Bates, M. R. Fuller, M. N. Kochert, J. O. McKinley, and P. M. Lukacs. 2006. Effects of radiomarking on Prairie Falcons: attachment failures provide insights about survival. Wildlife Society Bulletin 34:1 16-126. Sullivan, B. L. and C. L. Wood. 2005. A plea for the common birds. North American Birds 59:18-30. Tabb, E. 1977. Winter returns of American Kestrels in southern Florida. North American Bird Bander 2:163. Temple, S. A. 1972. Systematics and evolution of the North American Merlins. Auk 89:325-338. Tordoff, H. B. and P. T. Redig. 1997. Midwest Peregrine Falcon demography, 1982-1995. Journal of Raptor Research 31:339-346. U. S. Census Bureau (USCB). 2009. State and county QuickFacts. Cape Coral, Florida. U.S. Census Bureau, Washington, D.C., USA, http://quickfacts.census.gov/ qfd/states/1 2/1 2 1 0275.html. Wesemann, T. 1986. Factors influencing the distribution and abundance of Burrowing Owls (Athene cunicu- laria) in Cape Coral. Florida. Thesis. Appalachian State University. Boone. North Carolina. USA. White, G. C. and K. P. Burnham. 1999. Program MARK: survival estimation from populations of marked animals. Bird Study 46:120-139. Yoder, J. M., E. A. Marsciiall. and D. A. Swanson. 2004. The cost of di.spersal: predation as a function of movement and site familiarity in Ruffed Grouse. Behavioral Ecology 15:469^76. Zeiss, B. 1983. The other side of the river: historical Cape Coral, B. Zeiss, Cape Coral. Florida. USA. The Wilson Journal of Ornithology 122(3):484-493, 2010 TEMPORAL AND SPATIAL SHIFTS IN HABITAT USE BY BLACK BRANT IMMEDIATELY FOLLOWING FLIGHTLESS MOLT TYLER L. LEWIS,' -' PAUL L. FLINT,' JOEL A. SCHMUTZ,' AND DIRK V. DERKSEN' ABSTRACT. — Each year thousands of Pacific Black Brant (Branta bernicla nigricans) undergo flightless wing molt in the Teshekpuk Lake Special Area (TLSA), Alaska, in two distinct habitats: inland, freshwater lakes and coastal, brackish wetlands. Brant lose body mass during wing molt and likely must add reserves upon regaining flight to help fuel their 2,500 km migration to autumn staging areas. We characterized movements and habitat use by Brant during post-molt (the period immediately following the recovery of flight) by (1) marking individual Brant with GPS (global positioning system) transmitters, and (2) conducting a series of replicate aerial surveys. Individuals molting in inland habitats promptly abandoned their molt wetland during the post-molt and moved into coastal habitats. Consequently, inland habitats were nearly deserted by early August when Brant had regained flight, a decrease of >5,000 individuals from the flightless period of early July. Conversely, coastal molting Brant largely remained in coastal habitats during the post-molt and many coastal wetlands were occupied by large flocks (>1,000 birds). Our results indicate that inland, freshwater wetlands were less suitable post-molt habitats for Brant, while coastal wetlands were preferred as they transitioned from flightless molt. The immediacy with which Brant vacated inland habitats upon regaining flight suggests that food may be limiting during molt and they are not selecting inland molt sites strictly for food resources, but rather a balance of factors including predator avoidance and acquisition of protein for feather growth. Our data clearly demonstrate that patterns of habitat use by Brant in the TLSA change over the course of the molt season, an important consideration for management of future resource development activities in this area. Received 27 July 2009. Accepted 17 February 2010. Failed- and non-breeding individuals of many Northern Hemisphere goose species migrate to distantly-located molting areas to undertake a flightless wing molt (Hohman et al. 1992, Kjellen 1994). The flightless molt period imposes a number of constraints, including restricted ability to escape predators or access unused habitats and food sources (Fox and Kahlert 2000, Kahlert 2006). Growth of new flight feathers requires intake of certain nutrients essential to feather development, especially proteins and sulfur-con- taining amino acids (Murphy and King 1984, Sedinger 1984, Heitmeyer 1988, Fox and Kahlert 1999). Fox and Kahlert (2005) suggested that geese alleviate these constraints by choosing wing molting habitats that balance low predation risk with food supplies high in protein and energy. The constraints of the flightless period are immediate- ly relaxed upon regaining flight, however, and habitat attributes necessary for molting, including open-water escape areas or availability of nutri- ents for feather growth, may no longer drive habitat selection. Northern Hemisphere geese migrate soon after completing their molt, often traversing consider- ' U..S. Geological Survey. Alaska Science Center, 4210 University Drive. Anchorage, AK 99508, USA. ^Current address: Institute of Arctic Biology, and Department of Biology and Wildlife, University of Alaska, 31 1 Irving Building, Fairbanks, AK 99775, USA. ’Corresponding author; e-mail: tlewis@usgs.gov able distances from arctic molting sites to more southerly staging and wintering areas (Hohman et al. 1992). However, because the flightless wing molt is an energetically costly event and geese often lose considerable body mass during the molt (Williams and Kendeigh 1982, Taylor 1993, Fox and Kahlert 2005, Portugal et al. 2007), geese likely need to accumulate new fat reserves to fuel their forthcoming migration. Thus, post-molt habitats may be chosen largely for food availabil- ity and, when a molt site offers sufficient food availability for birds to build reserves, individuals may opt to stay at that site beyond the time at which flight was recovered. Conversely, abrupt abandonment of a molt site following recovery of flight may suggest food is limiting and alternate sites offer superior food supplies, conferring a selective advantage to birds that change sites. Each year, thousands of Pacific Black Brant (Brantct bernicla nigricans\ hereafter Brant) migrate to the Teshekpuk Lake Special Area (TLSA), on the Arctic Coastal Plain of Alaska, where they undergo flightless wing molt (Derksen et al. 1982, Taylor 1995). TLSA and sun'ounding areas provide Brant with a choice beJween two distinct molting habitats in close proximity: coastal, brackish wetlands or inland, freshwater wetlands (Flint et al. 2008). Each habitat offers different plant species as potential forage for molting Brant, as well as different types of open- water escape habitat (Markon and Derksen 1994, 484 L('u/.v et al. • POST-MOLT ECOLOGY OF BLACK BRANT 485 70°50'0"N- 70"40'0"N' 153°30’0"W 153"0’0"W 152"30'0"W -70°50’0'’N -70“40’0"N 152WW FIG. 1. Molting goose region of the Teshekpuk Lake Special Area, Alaska, during 2006-2008. Lakes at which Brant were aerially surveyed are darkened and strata boundaries are depicted with bold lines. Flint et al. 2008). We examined the post-molt ecology of Brant in the TLSA to better understand the potential differences in food availability and quality between coastal versus inland habitats, as well as preference of Brant for each habitat type following the recovery of flight. We used a combination of population-level aerial surveys and individuals marked with GPS (global posi- tioning system) transmitters to examine changes in Brant distribution from molt to post-molt, probability of remaining on a molt site following recovery of flight, and types of habitats used post- molt. METHODS Classification of Coastal versus Inland Molt- er. — The TLSA is characterized by tundra vege- tation, continuous permafrost, and large perma- frost thaw lakes (Markon and Derksen 1994, Jorgensen et al. 2006). Flightless molting Brant occur primarily in the northeast section (Fig. 1) and are primarily subadults, non-breeders, and failed-breeders associated with breeding colonies in Alaska, Russia, and western Canada (Bollinger and Derksen 1996). Flint et al. (2008) divided the TLSA molting area into three strata based on distinct vegetative and landscape characteristics. The majority of our research was conducted in strata two (inland stratum) and three (coastal stratum; Lig. 1 ). The coastal stratum is character- ized by the occuirence of saltwater intrusion, elevations <1.5 m, and salt-tolerant plant com- munities, while the inland stratum has a lack of saltwater intrusion, slightly greater elevations (^4 m), and predominantly freshwater plant communities. Brant marked with transmitters were classified as either inland or coastal molters depending upon the stratum in which they molted. Additionally, some transmitter-marked Brant molted in areas near the TLSA, but outside the stratum boundaries. This latter group molted solely in coastal habitats and is included in the coastal molter classification. GPS Transmitters. — Our first method for char- acterizing movements and habitat use of Brant during the post-molt was through use of GPS 486 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. 3. September 2010 transmitters. We captured female Brant from nesting colonies on the Arctic Coastal Plain of Alaska, between Barrow (71° 18' N, 156° 44' W) and the Colville River Delta (70° 25' N, 150° 23' W), during two summers (2007-2008). We captured birds on nests using spring-loaded bow traps (Sayler 1962) and marked them with GPS transmitters. We removed eggs from nests of captured females changing their status to failed breeder and significantly increasing the likelihood they would migrate to the TLSA to undergo flightless wing molt. We marked 39 Brant from 11 to 17 June 2007 and 39 Brant from 12 to 20 June 2008. We experienced high rates of trans- mitter failure in both years of our study as verified via incidental recaptures of Brant with failed transmitters, recovery of defective transmitters from hunter-harvested individuals, and sudden loss of transmitter signals from flightless individ- uals (i.e., birds that could not move from our reception range). Thus, missing transmitters are largely attributable to transmitter failure rather than molt migration from our study area (Lewis and Flint 2008). GPS transmitters were externally affixed be- tween the shoulder blades of captured Brant via dual subcutaneous anchors, following Lewis and Flint (2008). We used two models of GPS transmitters in 2007 that differed only in power source: solar (/? = 19) and battery powered (n = 20). All 39 GPS transmitters used in 2008 were battery-powered. Estimated transmitter life was 45-60 days for battery-powered transmitters and 45-i- days for solar models. GPS transmitters were programmed to collect and store a location every 6 hrs and offload all stored locations via a daily VHF radio transmission. Offloaded data were collected weekly from fixed-wing aircraft equipped with wing-mounted two-element yagi antennas. Data were downloaded using an HR2600 receiver produced by HABIT Research (Victoria, BC, Canada). Aerial Surveys. — Aerial surveys, our second method for characterizing Brant movements and habitat use, encompassed most of the molting Brant population in the TLSA. We conducted replicate aerial surveys of 33 selected lakes within the TLSA during summers 2006-2008 (Fig. 1). The initial survey was during the first week of July and repeated on a weekly basis for 6 weeks, encompassing the core molting period of Brant (Derksen et al. 1982). The survey protocol followed that used by the U.S. Fish and Wildlife Service (USFWS) in their annual mid-July survey of molting geese in the TLSA as described by Flint et al. (2008). This protocol assumes that detection rate and observer error are constant on all surveys. We used the USFWS survey data from 1996 to 2005 to select the subsample of lakes for our repeated survey, choosing only lakes that were used by >200 Brant in >3 years. This resulted in a subsample of 33 lakes which typically contained >90% of the molting Brant counted in these years. All surveyed lakes were within the boundaries of either the inland or coastal stratum and were classified accordingly (Flint et al. 2008). Statistical Analyses. — We calculated two vari- ables using data from our birds marked with GPS transmitters to explore Brant post-molt ecology: probability of molt site abandonment and distance from coastline. Both variables are time-depen- dent. being measured in relation to time since onset of molt, and thus required that start date of molt be known. We used an algorithm, which assumed movement rates of Brant are significant- ly reduced during the flightless molt, to estimate each individual’s start date of molt. The algo- rithm, described by Lewis et al. (2010), identified the date at which movement rates distinctly changed, allowing us to objectively define a precise date of molt initiation for each of our transmitter-marked birds. We did not use this algorithm to estimate the end date of molt because many individuals did not increase their movement rates or move among wetlands despite presumably regaining flight. We used estimates from previous Brant research in the TLSA which found that partial flight is regained —23 days after shedding of flight feathers (Taylor 1995). We assumed for all analyses that Brant had regained partial flight by 23 days from the onset of flightlessness. We quantified timing of molt site abandonment for each of our transmitter-marked individuals. Most wetlands used during molt are separated by wide stretches of upland tundra which are essentially impassable to flightless Brant; thus, individuals are often confined to one wetland, or interconnected wetlands, until flight is regained. Individuals were assigned a primary molt wetland from the time molt started, as calculated by our molt algorithm, and their daily status classified as “present” until they moved to a different, unconnected wetland, at which point their status was classified as “abandoned”. This exercise produced a binary variable for each of our Lewis ei cil. • POST-MOLT ECOLOGY OF BLACK BRANT 487 transmitter-marked individuals, where each day since the onset of molt was classified as either present or abandoned. We used this binary variable in Kaplan-Meier survival models to estimate the daily probability of molt site abandonment for coastal and inland molters. The Kaplan-Meier analysis produced separate proba- bility curves for coastal and inland molters with each curve plotting the probability that Brant remained on their primary molt wetland versus time since onset of molt. We used a log-rank test of equality to examine the null hypothesis (a = 0.05) that there was no difference in probability curves of coastal versus inland molters (Pollock et al. 1989). We modeled the distance from coastline in relation to time since onset of molt to examine if coastal versus inland molters differed in their post-molt habitat use. We fit a series of general linear models (GLM) to describe the shape of time trends for the explanatory variable of ‘distance from coast’. Models examined included intercept, linear, quadratic, and cubic time function with the best-fitting model selected using Akaike’s Infor- mation Criterion (AIC; Burnham and Anderson 2002). The start of molt was defined as time zero with time (hrs) of all subsequent GPS locations added onto zero and all prior GPS locations subtracted from zero. The cubic function was the best-fitting time model (AAIC = 0) and all other models had a AAIC < 4.0. Thus, using ‘distance from coast’ as our explanatory variable, we next fit a series of general linear mixed models in which the cubic time function was included in all models. We used mixed models to account for repeated measures of individuals marked with transmitters by including subject as a random effect (Littell et al. 2000). The mixed models included the response variable ‘habitat’, defined as either coastal or inland depending upon the habitat strata in which each individual molted. The model set included a time only model, habitat as an additive variable with time, or all combi- nations of habitat by time interactions for a total of nine models. The best-fitting model was selected using AAIC and the null model (inter- cept-only) was compared against each model to assess model fit (Burnham and Anderson 2002). We examined the spatial distribution of Brant across strata by summarizing each aerial survey as the ratio of Brant counted inland to Brant counted coastally and used this ratio as our response variable to fit a candidate set of GLMs. We used FIG. 2. The probability individuals remained on their molt wetland by number of days since start of molt for Brant marked with GPS transmitters which molted at coastal versus inland wetlands, Teshekpuk Lake Special Area, Alaska, during 2007-2008. Brant typically regain flight 23-24 days after onset of molt. year (2006, 2007, 2008) and Julian date of each survey as our explanatory variables, and our candidate model set consisted of single variable models, the additive model, the interaction model, and the intercept-only null model for a total of five models. Akaike’s information Criterion (AICc), adjusted for small sample size, was calculated for each model in the candidate set, AAICc and AIC weights (vc,) were used to infer the relative support of each model, and the null model (intercept-only) was compared against each model to assess model fit. All analyses were performed using SAS Version 9.1 statistical software (SAS Institute 2003) and descriptive statistics are presented as mean ± SE. RESULTS GPS transmitters on 23 Brant functioned for ^23 days from onset of flightlessness allowing us to examine Brant behavior as they transitioned from their flightless molt. Sixteen of these 23 birds were classified as coastal molters and seven as inland molters. The probability that an individual remained on its molt wetland on a given day was significantly different for coastal versus inland molters (log-rank test, = 7.73, df = \, P = 0.005), as inland molters abandoned their molt site sooner than coastal molters (Fig. 2). Abandonment of molt sites by inland molters began 22 days after molt initiation and the probability of remaining at their molt site was 21% at 28 days. Abandonment of molt sites by coastal molters began 25 days after molt initiation 488 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 3, September 2010 TABLE 1. Candidate general linear mixed models evaluating variation in distance from coast (m) for Brant with GPS transmitters in the Teshekpuk Lake Special Area, Alaska, 2007-2008. The explanatory variables are time (t), included as a cubic function (t -h F + F) in all models, and habitat (h) defined as coastal or inland. Models are listed in order of AAIC. Model Parameters Log-likelihood AAIC Wj h + t + ht + t- -b ht- + t-"' 10 -3,5927.29 0 0.37 h + t + ht + F + t-’ + hF 10 -3,5927.34 0.09 0.36 h + t + ht + F + hF + F + ht^ 11 -3,5926.85 1.11 0.21 h + t + t- + t^ + ht^ 9 -3,5930.88 5.17 0.03 h + t + t- + ht- + 0 9 -3.5931.61 6.65 0.01 h + t + t- + hF + F + ht^ 10 -3,5930.83 7.09 0.01 h + t + ht + F + F 9 -3.5937.37 18.16 0.00 h + t + t- + t^ 8 -3,5938.63 18.67 0.00 t + F + t' 7 -3,5947.15 33.71 0.00 and the probability of remaining at their molt site was 86% at 28 days. The best-fitting model (AAIC = 0) to explain distance from the coast for Brant marked with GPS transmitters contained the explanatory vari- able of habitat, defined as coastal or inland molter, in the habitat*time {ht) and habitat* time*time (ht^) terms of the cubic function (Table 1). Two other models also received substantial support (AAIC 2): the model with habitat in the ht and ht’ terms and the model with habitat in all three terms of the cubic time function (Table 1). The null model, in which the habitat variable was not included, received no empirical support indicating that inclusion of habitat explains a significant amount of variation in distance to coast (Table 1). Plots of predicted values for each of the three well-supported models indicate they differ only at the extreme tails of the cubic functions. The biological inference for each of the three well-supported models is identical and we used model-averaged estimates, based on the three top models, for all further inference. Model- averaged parameter estimates (Table 2) indicate that by 23 days after initiation of molt, when many Brant have regained flight, inland molting Brant promptly moved into coastal habitats, while coastal molting Brant remained in coastal habitats (Fig. 3). Inland and coastal molters were at similar distances from the coast before the onset of molt and were ~8 and 2 km from the coast during the flightless molt, respectively (Fig. 3). The best-fitting model (AAIC^ = 0) to explain the proportion of Brant per habitat strata, as counted from aerial surveys, contained only Julian date as a response variable (Table 3). The additive model containing year and Julian date received moderate support, although its vv, was low relative to the top model (Table 3). All other models, including the null model, had AAIC^. values >20. These results indicate the distribution of Brant across habitat strata was strongly influenced by Julian date. The proportion of Brant in the coastal stratum increased by date beginning in mid to late TABLE 2. Weighted parameter estimates and unconditional standard error (SE) for general linear mixed models evaluating variation in distance from coast (m) for Brant with GPS transmitters in the Teshekpuk Lake Special Area, Alaska. 2007-2008. Explanatory variable' Parameter estimate Unconditional SE Intercept 7,572.67 947.77 Habitat -5,350.29 1.074.16 Time 3.48 1.69 Habitat*Time -5.19 1.88 Time^ -1.48*10 5.87*10 ’ Habitat*Time^ 6.21*10 ’ 3.98*10 ■’ Time’ -3.47*10 6.81*10 Habitat*Time’ 5.47*10 3.67*10 " Habital is a categorical variable (coastal or inland) with inland as the reference value. Lewis el al. • POST-MOLT ECOLOGY Ol- BLACK BRANT 489 Pr^-moK Flightless moK Fllghtregained Days since start cff molt FIG. 3. Model averaged estimate of distance from coastline (in) by days since start of molt for Brant marked with GPS transmitters which molted at coastal versus inland wetlands, Teshekpuk Lake Special Area, Alaska, during 2007-2008. FIG. 4. The proportion of Brant in coastal versus inland strata based on aerial surveys, Teshekpuk Lake Special Area, Alaska, conducted during July-early August, 2006- 2008. Aerial surveys were conducted six times per year with each survey separated by ~1 week. Weekly surveys are summarized across years, producing weekly means ± SE. July as they began to regain flight (Fig. 4). This pattern of movement into the coastal stratum did not differ across years or by the interaction of year*date. Most Brant had regained flight by early August and >90% were in the coastal stratum, a 30% increase from the flightless molt period (Fig. 4). The increased proportion of Brant in the coastal stratum did not result from inland molters leaving the TLSA altogether. Rather, the number of birds in the coastal stratum increased or remained constant during late July, despite some coastal molters leaving the area. Thus, many inland molters moved into coastal stratum habitats upon regaining flight, similar to the pattern observed for our individuals marked with trans- mitters. Aerial surveys, however, demonstrated this pattern for much larger numbers of Brant TABLE 3. Candidate general linear models evaluating variation in ratio of Brant counted inland versus those counted coastally during aerial surveys conducted in the Teshekpuk Lake Special Area, Alaska, 2006 to 2008. Models are listed in order of AAIC. Model Parameters R' AAIC B7 Julian date 3 0.86 0 0.68 Julian date + Year Julian date + Year 5 0.90 1.53 0.32 + Julian date*Year 8 0.91 21.41 0 Null 2 0 28.00 0 Year 4 0.02 34.38 0 (>15,000 annually) and over a larger spatial area. Post-molt habitat use varied among wetlands within the coastal stratum with some being completely abandoned by Brant during the post- molt while others experienced increases of hundreds to thousands of Brant. Numbers of Brant either increased or remained stable at 40% of the 22 coastal wetlands for which sufficient aerial survey data were available during the post- molt period; conversely, post-molt Brant numbers decreased significantly at all nine inland stratum wetlands surveyed. The coastal wetland complex at the Smith River (Fig. 1 ) was particularly important, supporting >5,000 Brant annually during the post-molt, an increase of 2,000-3,000 birds above the wetland’s average annual molting population. DISCUSSION Brant regain flight when their primaries reach 70-75% of their full-grown length, which typi- cally occurs 23-24 days after onset of molt (Taylor 1995). Using this timeline, we found that individuals molting in inland, fre.shwater habitats abandoned their molt site almost immediately upon regaining flight, moving almost exclusively into coastal habitats. Inland wetlands were nearly abandoned by early August, a decrease of >5,000 individuals from the flightless period of early July, and the only remaining individuals were likely those which had yet to attain flight. 490 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 3. September 2010 Conversely, Brant molting in coastal habitats continued to reside on their molt wetlands for several days after regaining flight and did not switch habitats during the post-molt, remaining within coastal habitats. Many coastal wetlands were occupied by large flocks (>1,000 birds) of flying Brant during the post-molt. Our results indicate Brant almost universally prefer coastal habitats immediately following recovery of flight and inland, freshwater wetlands and lakes are less suitable post-molt habitats. Our results also corroborate earlier research from the 1980s, in which a small sample (/? = 18) of radio-marked Brant moved from inland to coastal sites during the post-molt period (Jensen 1990). This suggests the post-molt patterns observed in our study have persisted over a timescale spanning no less than multiple decades. Brant molting in inland habitats chose to leave their molt site sooner than those molting coastally, but the differential timing of molt site abandon- ment could alternatively be explained by differ- ential molt durations between the two habitat types. Habitat-specific factors, such as food quality, competition for resources, or presence of other stressors may influence the rate of feather growth and, by extension, the duration of molt (Hahn et al. 1992, Romero et al. 2005). Citril Finches {Carduelis citrinella) in higher quality habitats molted more rapidly than those in low quality habitats, despite the two habitats being only 5 km apart (Borras et al. 2004). Similarly, the longer residency of Brant at coastal molt sites may be due to hypothetically poorer habitat quality at these sites, rather than individuals choosing to remain past the recovery of flight. We cannot completely discount this alternative hypothesis, but we find it unlikely that quality differences between coastal versus inland habitats would shift molt durations to the extent that explains our observed differences in timing of molt site abandonment. Abandonment of coastal molt sites lagged that of inland sites by >5 days, a pronounced difference for a species that averages 23 days to regain flight, and >50% of coastal- molting individuals had not moved by day 30. The majority of Brant molted coastally and inland molting Brant promptly moved to coastal habitats following cessation ot molt, suggesting that coastal habitat quality is at least equivalent with inland habitats, if not superior. Overall, the patterns of post-molt movements and habitat use by Brant would not seem to support the hypothesis that variation in time spent at molting locations is negatively associated with habitat quality. Adult Brant collected in the TLSA from 1988 to 1989 lost >20% of their body mass and >80% of their lipid reserves during the flightless wing molt, completing molt with lipid reserves com- prising only 2-4% of body mass (Taylor 1993). More recent research (2005-2007) also docu- mented considerable body mass losses for molting Brant, although rates of mass loss were slower than during 1988-1989 (PLF, unpubl. data). Given significant declines in body mass and lipid reserves, molting Brant must regain body reserves during the post-molt period to fuel their forth- coming 2,500 km migration to their primary autumn staging area at Izembek Lagoon, Alaska (Vangilder et al. 1986, Dau 1992, Reed et al. 1998). Post-molt distributions of Brant are likely driven by availability and quality of food resources, and the unidirectional movement of Brant from inland habitats suggests that inland food resources are less able to fulfill post-molt requirements than coastal resources. The imme- diacy with which Brant vacated inland stratum habitats upon regaining flight also suggests food is limiting during molt, potentially due in part to Brant grazing pressure, and Brant are not selecting inland molt sites strictly for food resources. Brant are likely selecting inland molt sites for a balance of factors which includes protein supplies for feather growth and large open-water areas for predator avoidance (Fox and Kahlert 2005). The sedge Corex suhspathacea and grass Pucciiiellio phyrognodes grow in saline-influ- enced tundra habitats (Markon and Derksen 1994) and are preferred forage items of Brant across the arctic and subarctic (Sedinger et al. 2001, Person et al. 2003). These plant species in the coastal habitats of the TLSA often occur as ‘grazing lawns’, which are areas of high density plant growth created by frequent goose grazing (Person et al. 2003). Grazing lawns typically experience short-term reductions in food abun- dance due to the immediate effects of grazing, but their long-term maintenance requires frequent grazing to create positive feedbacks of increased plant growth rates and nutrient content (Ruess. et al. 1997, Person et al. 2003). The nitrogen recycled through goose feces contributes to the salinity tolerance of grazing lawns (Ruess et al. 1997). The pronounced post-molt shift of Brant into coastal habitats likely contributes to mainte- Lewis ct a!. • POST-MOLT ECOLOGY OF BLACK BRANT 491 nance and expansion of grazing lawns, increasing coastal tood availability and further prompting Brant to exploit coastal habitats. The coastal plants grazed by Brant also have lower carbon to nitrogen ratios (C:N), the main index of food quality for geese (Riddington et al. 1997, Hassall et al. 2001), than the common inland stratum forage plants of Carex aquatilis and Deschompsia caespitosa (JAS, unpubl. data). Brant, because of their small body size and con'espondingly small digestive tract, have high food passage rates and require high quality foods from which nutrients can be quickly extracted (Prop and Vulink 1992, Sedinger 1997). Coastal areas in the TLSA are heavily influenced by summer fogs, which locally reduce air temperature and solar exposure, potentially delaying plant phenology (Post and Stenseth 1999, Menzel et al. 2006). Plant quality decreases as the arctic summer progresses (i.e., C;N temporally increases), and delayed phenolo- gy in coastal areas would result in higher quality forage plants during the late season post-molt period (Crawley 1983, Van der Wal et al. 2000). The combination of higher quality plant species and delayed plant phenology in coastal habitats may also be important factors contributing to the post-molt shift of Brant into coastal habitats. The TLSA also supports large numbers of molting Greater White-fronted {Anser albifrons), Canada (Brcinta canadensis). Cackling (B. hutch- insii), and Lesser Snow (Chen caendescens caerulescens) geese along with Brant. The number of molting geese counted in the TLSA in recent years has ranged from 70,000 to 90,000, collectively making this one of the most signif- icant goose molting areas in the circumpolar arctic (Flint et al. 2008). The TLSA is currently designated for future leasing for oil and gas extraction (2008 Northeast NPR-A Record of Decision; available at http://www.blm.gov/ak/st/ en/prog/energy/oil_ga.s/npra.html). Planning to minimize the effects of development of molting areas on Brant populations will require a clear understanding of patterns of habitat use and how patterns may change seasonally. The only data currently available to assess patterns of habitat use of Brant in the TLSA comes from a single annual survey conducted during the peak flightless period in mid-July (Flint et al. 2008). These survey data are not suited to describing intra-annual patterns of habitat use, such as the post-molt habitat shifts we describe. Our data clearly demonstrate that patterns of habitat use by Brant in the TLSA change over the course of the molt sea.son, and that future management and development deci- sions must consider seasonal variation in distri- butions of Brant. ACKNOWLEDGMENTS The Bureau of Land Management and U.S. Geological Survey, Alaska Science Center, provided funding and assisted with logistics. D. A. Nigro provided logistic support and helicopter management during captures. D. C. Douglas developed our algorithm for estimating molt initiation. Blake Bartzen assisted with captures. R. B. Lanctot provided housing in Barrow. J. S. Hamilton of Arctic Air Alaska provided aerial support. This paper benefited from the reviews of D. A. Nigro and T. F. Fondell. Use of trade, product, or company names is solely for descriptive purposes and does not imply endorsement or criticism by the U.S. government. LITERATURE CITED BoLLtNGER. K. S. AND D. V. Derksen. 1996. Demographic characteristics of molting Black Brant near Teshekpuk Lake, Alaska. Journal of Field Ornithology 67:141- 158. Borras, a., T. Cabrera, J. Cabrera, and J. C. Senar. 2004. Interlocality variation in speed of moult in the Citrial Finch Serinus citronella. Ibis 146:14—17. Burnham, K. P. and D. R. Anderson. 2002. Model selection and multi-model inference: a practical information-theoretic approach. Second Edition. Springer- Verlag, New York, USA. Crawley, M. J. 1983. Herbivory: the dynamics of animal- plant interactions. Blackwell Scientific Publications. Oxford, United Kingdom. Dau, C. P. 1992. The fall migration of Pacific Fly way Brent Branta bernicla in relation to climatic conditions. Wildfowl 43:80-95. Derksen, D. V., W. D. Eldridge, and M. W. Weller. 1982. Habitat ecology of Pacific Black Brant and other geese moulting near Teshekpuk Lake. Alaska. Wild- fowl 33:39-57. Fi.int, P, L., E. j. Mallek, R. j. King. J. A. Schmutz. K. S. Bollinger, and D. V. Derksen. 2008. Changes in abundance and spatial distribution of geese molting near Teshekpuk Lake, Alaska: interspecific com- petition or ecological change? Polar Biology 31:549- 556. Fox, A. D. and .1. Kahlert. 1999. Adjustments to nitrogen metabolism during wing moult in Greylag Geese. Anser anser. Functional Ecology 13:661-669. Fox. A. D. AND J. Kahlert. 2000. Do moulting Greylag Geese An.ser anser forage in proximity to water in respon.se to food availability and/or quality? Bird Study 47:266-274. Fox, A. D. and .1. Kahlert. 2005. Changes in body mass and organ size during wing moult in non-breeding Graying Geese Anser an.ser. Journal of Avian Biology 36:538-548. Hahn, T. P., J. Swingle. J. C. Wingfield, and M. 492 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122. No. 3, September 2010 Ramenofsky. 1992. Adjustments of the prebasic molt schedule in birds. Ornis Scandinavica 23:314- 321. Hassall, M., R. Riddington, and A. Helden. 2001. Foraging behaviour of Brent Geese, Branta b. bernicla. on grasslands: effects of sward length and nitrogen content. Oecologia 127:97-104. Heitmeyer, M. E. 1988. Protein costs of the prebasic molt of female Mallards. Condor 90:263-266. Hohman, W. L., C. D. Ankney, and D. H. Gordon. 1992. Ecology and management of postbreeding waterfowl. Pages 128-189 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, Editors). University of Minnesota Press, Minneapolis, USA. Jensen, K. C. 1990. Responses of molting Pacific Black Brant to experimental aircraft disturbance in the Teshekpuk Lake Special Area, Alaska. Dissertation. Texas A&M University, College Station, USA. Jorgensen, M. T., Y. L. Shur, and E. R. Pullman. 2006. Abrupt increase in permafrost degradation in arctic Alaska. Geophysical Research Letters 33:L02503, doi: 10. 1029/2005GL024960. Kahlert, j. 2006. Factors affecting escape behaviour in moulting Greylag Geese An.ser anser. Journal of Ornithology 147:569-577. Kjellen, N. 1994. Moult in relation to migration in birds — a review. Ornis Svecica 4:1-24. Lewis, T. L. and P. L. Flint. 2008. Modified method for external attachment of transmitters to birds using two subcutaneous anchors. Journal of Field Ornithology 79:336-341. Lewis, T. L., P. L. Flint, J. A, Schmutz, and D. V. Derksen. 2010. Pre-moult patterns of habitat use and moult site selection by Brent Geese Branta bernicla nigricans’, individuals prospect for moult sites. Ibis 152: 556-568. Littell, R. C., G. a. Milliken, W. W. Stroup, and R. D. WoLFiNGER. 2000. SAS System for mixed models. Fourth Edition. SAS Institute, Cary, North Carolina, USA. Markon, C. J, and D. V. Derksen. 1994. Identification of tundra land cover near Teshekpuk Lake, Alaska using SPOT satellite data. Arctic 47:222-231. Menzel, a., T. H. Sparks, N. Estrella, E. Koch, A. Aasa, R. Ahas, K. Alm-Kubler, P. Bissolli, O. Braslavska, a. Briede, F. M. Chmielewski, Z. Crepinsek, Y. Curnel, a. Dahl, C. Defila, A. DrtNNELLY, Y. Filella, K. Jatczak, F. Mage, A. Mestre, 0. Nordli, j. Penuelas, P. Pirinen, V. RemiSova. H. Scheifinger, M. Stritz, a. Susnik, a. J. H. Van Vliet, F.-E. Wielgolaski, S. Zach, and A. ZUST. 2006. European phenological response to climate change matches the warming pattern. Global Change Biology 12:1969-1976. Murphy, M. E. and J. R. King. 1984. Sulfur amino acid nutrition during molt in the White-crowned Sparrow. Condor 86:314-332. Pi-R.soN, B. T.. M. P. Herzog, R. W. Ruess, .1. S. Sedinger, R. M. Anthony, and C. A. Babcock. 2003. Feedback dynamics of grazing lawns: coupling vegetation change with animal growth. Oecologia 135:583-592. Pollock, K. H., S. R. Winterstein, C. M. Bunck, and P. D. Curtis, 1989. Survival analysis in telemetry studies: the staggered entry design. Journal of Wildlife Management 53:7-15. Portugal, S. J., J. A. Green, and P. J. Butler. 2007. Annual changes in body mass and resting metabolism in captive Barnacle Geese {Branta leucopsis): the importance of wing moult. Journal of Experimental Biology 210:1391-1397. Post, E. and N. C. Stenseth. 1999. Climatic variability, plant phenology, and northern ungulates. Ecology 80:1322-1339. Prop, J. and T. Vulink. 1992. Digestion by Barnacle Geese in the annual cycle: the interplay between retention time and food quality. Functional Ecology 6:180-189. Reed, A., D. H. Ward, D. V. Derksen, and J. S. Sedinger. 1998. Brant (Branta bernicla). The birds of North America. Number 337. Riddington, R., M. Hassall, and S. J. Lane. 1997. The selection of grass swards by Brent Geese Branta b. bernicla’. interactions between food quality and quantity. Biological Conservation 81:153- 160. Romero, L. M., D. Strochlic, and J. C. Wingfield. 2005. Corticosterone inhibits feather growth: potential mechanism explaining seasonal down regulation of corticosterone during molt. Comparative Biochemistry and Physiology, Part A 142:65-73. Ruess, R. W., D. D. Uliassi, C. P. H. Mulder, and B. T. Person. 1997. Growth responses of Carex ramen.'ikii to defoliation, salinity, and nitrogen availability: implications for geese-ecosystem dynamics in western Alaska. Ecoscience 4:170-178. SAS Institute. 2003. SAS Version 9.1. SAS Institute, Cary, North Carolina, USA. Sayler, J. W. 1962. A bow-net trap for ducks. Journal of Wildlife Management 26:219-221. Sedinger, J. S. 1984. Protein and amino acid composition of tundra vegetation in relation to nutritional require- ments of geese. Journal of Wildlife Management 48:1128-1136. Sedinger, J. S. 1997. Adaptations to and con.sequences of an herbivorous diet in grouse and waterfowl. Condor 99:314-326. Sedinger, J. S., M. P. Herzog, B. T. Person, M. T. Kirk, T. Obritchkewitch, P. P. Martin, and A. A. Stickney. 2001. Large-scale variation in growth of Black Brant goslings related to food availability. Auk 1 18:1088-1095. Taylor, E, J. 1993. Molt and bioenergetics of Pacific Black Brant (Branta bernicla nigrican.s) on the Arctic Coastal Plain, Alaska. Dissertation. Texas A&M University, College Station, USA. Taylor, E. J. 1995. Molt of Black Brant (Branta bernicla nigricans) on the Arctic Coastal Plain, Alaska. Auk 1 12:904-919. Van der Wal, R., N. Madan, S. van Lieshout, C. Lewis el al. • POST-MOLT ECOLOGY OF BLACK BRANT 493 Dormann, R. Langvatn, and S. D. Albon. 2000. Trading forage quality for quantity? Plant phenology and patch choice by Svalbard reindeer. Oecologia 123:108-1 15. Vangilder, L. D., L. M. Smith, and R. K. Lawrence. 1986. Nutrient reserves of premigratory Brant during spring. Auk 103:237-241. Williams, J. E. and S. C. Kendeigh. 1982. Energetics of the Canada Goose. Journal of Wildlife Management 46:588-600. The Wilson Journal of Ornithology 122(3);494-502, 2010 ARTHROPOD FORAGING BY A SOUTHEASTERN ARIZONA HUMMINGBIRD GUILD DONALD R. POWERS,'*' JESSAMYN A. VAN HOOK,' ELIZABETH A. SANDLIN," AND TODD J. McWHORTER^-' ABSTRACT. — We tested the hypothesis that foraging for arthropods may be a viable source of energy when hummingbirds are competitively excluded from sources of nectar. We hypothesized that the Magnificent Hummingbird (Eugenes fulgens) relies more upon arthropods than the Blue-throated Hummingbird (Lampornis clemenciae) or Black- chinned Hummingbird (Archilochus alexandri) in southeastern Arizona. We were unable to quantify arthropod foraging by A. alexandri, but measured frequent arthropod foraging by both E. fulgens and L. clemenciae. E. fulgens engaged in more aerial llycatching than L. clemenciae, and their rate of flycatching attempts was higher than by L. clemenciae. Analysis of gut contents showed that E. fulgens consumes the greatest diversity of arthropods. Respiratory quotient measurements indicated E. fulgens catabolized a greater amount of fat/protein than the other species. Gut morphology of E. fulgens does not appear to differ from other hummingbirds suggesting hummingbirds in general may have the ability to use arthropods as an alternative energy source when access to floral energy is restricted. Our data are consistent with the hypothesis that the diet of E. fulgens includes more arthropods than other species with which they compete. Received 10 November 2009. Accepted 3 Eebruary 2010. Floral nectar has been assumed to be the primary energy resource of hummingbirds. This is reasonable from both ecological and physio- logical perspectives when one considers that many aspects of hummingbird and floral natural history are intertwined (Feinsinger 1983, Stiles 1995) , and that sugars in floral nectar are quickly and easily digested. It is also known that hummingbirds supplement their diets with small arthropods as floral nectar lacks many important nutrients (Baker 1977). However, arthropods are generally considered to be of little energetic importance for hummingbirds (Wolf and Hains- worth 1971). Hummingbirds often spend <10% of their day foraging for arthropods when nectar is in sufficient supply (Gass and Montgomerie 1981). There is evidence that arthropod consumption increases during specific periods of the natural cycle of hummingbirds such as during reproduc- tion, specifically by nesting females (Murphy 1996) , and during periods of low nectar availabil- ity (Chavez-Ramirez and Dowd 1992). Time spent on other activities, including nectar forag- ing, is likely reduced when time spent foraging for arthropods increases (Chavez-Ramirez and Dowd 1992). Stiles (1995) argued that hummingbirds ' Biology Department. George Fox University. 414 North Meridian .Street. Newberg. OR 97132. USA. ^Department of Ecology and Evolutionary Biology. University of Arizona. Tucson. AZ 85721, USA. ’School of Animal and Veterinary Sciences, University of Adelaide, Roseworthy Campus, SA 5371, Australia. 'Corresponding author; e-mail: dpowcrs@gcorgefox.cdu acquire a meaningful amount of energy from consumption of arthropods, which would be needed to meet energy demands when nectar intake is reduced. This is not difficult to imagine since arthropods have high energy content (Weathers and Sullivan 1989) and appear to be fully digested in a hummingbird’s digestive tract (Remsen et al. 1986). We know of no studies that examined the possibility that hummingbirds might use arthro- pods when excluded from a nectar source because of local social interactions. For example, an aggressive dominant species could restrict access of subordinate species to clumps of flowers contained within its territory forcing the subordi- nate species to seek alternative energy sources (Young 1971). This may reduce the energetic impact of competition if the subordinate could use arthropods as a primary energy source. This attribute would provide ecological separation between competitors so they can coexist (Rosen- zweig 1985, Sandlin 2000b) in areas that do not provide abundant nectar resources (Brown et al. 1978). Eugenes fulgens (Magnificent Hummingbird) is a member of a three-species hummingbird guild seasonally inhabiting the Chiricahua Mountains of southeastern Arizona. E. fulgens will readily feed on nectar in artificial feeders, but use declines when feeders are defended by dominant Blue- throated Hummingbirds (Lampornis clemenciae) (Pimm et al. 1985, Powers and Conley 1994, Sandlin 2()00a. Powers et al. 2003). This behavior is substantially different from that exhibited by 494 Powers el uL • AF^THROPOD FORAGING BY HUMMINGBIRDS 495 Black-chinned Hummingbirds {Archilochus alex- andri), another subordinate species, which con- sistently intrudes on L. clenienciae territories in attempts to gain a nectar reward (Powers and Conley 1994). There are few natural flowers available in many portions of the Chiricahua Mountains prior to onset of monsoon rains, and arthropods may be the only nearby alternative energy source for E. fulgens (Pimm et al. 1985). The idea that E. fulgens has the ability to subsist only on arthropods is not new. Marshall ( 1957:81 ) suggested that E. fulgens can inhabit pine-oak (Piniis spp.-Quercus spp.) woodlands because it “can dispense with moist habitat and flowers” and switch to arthropod consumption. We conducted this study to ascertain if arthropods could be used as an alternative energy source when nectar availability is restricted and if the use of arthropods for energy is a feasible strategy that allows E. fulgens to reduce compet- itive interactions with territorial L. clemenciae. Evidence from behavioral, morphological, and physiological experiments are presented to dem- onstrate the energetic importance of arthropod foraging to hummingbirds. METHODS Study Animals. — Males of three species were used in this study: E. fulgens (—7.5 g), L. clemenciae (-8.0 g), and A. alexandri (—3.0 g). All three species occur in southeastern Arizona during summer and use different foraging strate- gies (Pimm et al. 1985, Powers and Conley 1994, Sandlin 2000a). L. clemenciae is a dominant species, defending territories along borders of riparian canyons. A. alexandri is non-territorial at our study site and robs nectar from territories defended by L. clemenciae . E. fulgens is also non- territorial, but appears to forage as a trapliner (Powers 1996), avoiding many competitive inter- actions at dense flower aggregations. Study Area. — This study was conducted at the American Museum of Natural History’s South- western Research Station in the Chiricahua Mountains, Cochise County, Arizona (31° 50' N, 109° 15' W; 1,700 m asl). Habitat at the station is largely riparian, bordered by oak woodland and a mixed deciduous/coniferous forest. Pimm et al. (1985) provide a more detailed description of the habitat. This forest endures a pronounced dry season from November until monsoon rains begin in early July that can dramatically affect the amount of nectar available (Li and Brown 1999). Feeding Stations. — Twenty feeding stations in areas with vegetation that provided many poten- tial perching sites were established for behavioral observations within the boundaries of the field station. This increased the probability that arthro- pod foraging could be observed in conjunction with nectar foraging. Each individual feeding station consisted of either a Perky-Pet Glass Feeder (Model No. 203-CP, Woodstream Corp., Denver, CO, USA) with the perch and corolla removed, or a triplet feeder constructed from three smaller Perky-Pet plastic singlet feeders (Model No. 214, Woodstream Corp., Denver, CO, USA) connected with Velcro® strips. The feeders were suspended from either a tree branch or a wire hook attached to a vertical length of PVC pipe. The height of the feeders (distance from the ground to the bottom of the feeder) ranged from 0.74 to 2.10 m. A 0.52 M (18% weight/weight) sucrose solution, a concentration typical of nectars in many hummingbird flowers (Baker 1975), was used at all the feeding stations. Observations were made at four different time intervals for 97 hrs throughout the day during June and July 1998 and 1999. The timed intervals were: early morning (0500-0730 hrs MST), late morning (0800-1130 hrs), late afternoon (1500- 1800 hrs), and early evening (1815-1930 hrs). The time of day each feeding station was visited was randomized to ensure an unbiased examina- tion of both nectar and arthropod foraging over all times of day in all habitat types. Arthropod Foraging. — The amount of time birds spent foraging for arthropods near the feeding stations was recorded. Arthropod foraging was divided into two modes: perch foraging and aerial foraging. Foliage gleaning or other known modes of arthropod foraging (Stiles 1995) were not detected. Perch foraging occurs when birds forage for arthropods while sitting on a perch (no night is involved). This mode of arthropod foraging has been noted in a few hummingbird species (Pitelka 1942, Brice 1992). Typical behaviors include tongue extension, the bill opening, “gulping”, or the head darting back and forth. Arthropod foraging data were only taken when it was clear that birds were attempting to catch arthropods (often the vantage point made it possible to see arthropod capture). Time spent perch foraging was measured by starting a stopwatch when a bird opened its bill, extended the tongue, or when the bird gulped (indicating the initiation of an arthropod foraging bout). The 496 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 3, September 2010 number of attempts the bird made for arthropods was eounted with the stopwatch running and the measurement was terminated when the bird flew, vocalized, scratched, rubbed its bill, preened its feathers, or lost active interest. The approximate height of the perch where the event occurred was recorded after recording the time and the number of attempts. The second major mode of arthropod foraging, aerial flycatching, occurs during hovering or slow forward flight (Stiles 1995). We measured the duration of aerial flycatching bouts with a stopwatch from the time the bird left its perch until it returned to a perch or when the bout had clearly ended. The number of foraging attempts during a bout was estimated (we could not reliably count actual captures) by counting the number of head lunges (indicating a capture attempt). The height of the perch from which the bout was started was estimated visually. Crop and Gizzard Analysis. — Crop and gizzard data came from samples taken by dissection of 17 birds (// = 5 L. clemenciae , n = 6 E. fulgens, and n = 6 A. alexandri) during June and July 1996. Gut contents were preserved in 95% ethanol. Whole or nearly whole arthropod specimens were identified to the lowest possible taxon. Measure- ments of the length of whole arthropods were made to calculate an approximate size range for arthropod prey. The crop and gizzard contents were dried to a constant mass in a drying oven (65° C) so we could compare the relative mass of arthropod contents for each hummingbird species. Measurements were standardized to compare the amount of arthropod material in the gut between species by dividing the mass of arthropod material collected by bird metabolic body mass (gO.?.-); exponent approximately describes the relationship between body mass and metabolic rate; Calder and King 1974). Gut Allometry.—l\\Q gut mass and length of each species were compared with 10 and eight other hummingbird species, re.spectively (M.-V. Lopez-Calleja, C. Martmez del Rio, and J. E. Schbndube, unpubl. data). The intestines, from proventriculus to cloaca, were removed from birds killed by thoracic compression and immediately placed in 1.02% saline. The inte.stine was flushed with saline, blotted dry, and weighed to the nearest 0.01 g after any fat and me.sentery were removed. Total length was measured to the nearest 0.1 mm with a digital caliper after the intestine was gently straightened. The relationship Paragona gigas - Eugenes fulgens Utmpomis clemenciae Calypfe anna I Selasphorus rufus — Selasphorus plaryrercus I Archilochus alexandri — Archilochus colubris Cynanthus larirvsrris ■ Hylocharis leucoiis Amadlia ruiila I Chlorostilhon canivetti L Sephanoides sephanoides Colibri thalossinus 0.2 FIG. 1. Presumed phylogeny of 14 hummingbird species used in our phylogentic independent contrasts. A 0.2 branch-length scale is below the physlogenetic tree. The phylogeny is based on McGuire et al. (2007) with placement of C. latirostris based on Garcfa-Deras et al. (2008). between gut length, gut mass, and body mass was examined using standard linear least-squares regression as well as phylogenetically indepen- dent contrasts (PIC; Felsenstein 1985, Garland et al. 1992). We performed least-squares regressions of log gut mass versus log body mass and log gut length versus log body mass for standard analysis. We calculated PIC for log body mass, log gut mass, and log gut length (Garland et al. 1992). The phylogeny and relative branch lengths used in our PIC analysis were based on McGuire et al. (2007) and Garcia-Deras et al. (2008) (Fig. 1). The maximum branch length was arbitrarily set at 1.0 with shorter branch lengths scaled appropri- ately. PIC analysis is robust to branch length variation (Garland et al. 1999) and setting branch lengths in this manner are unlikely to influence PIC results. Absolute values of standardized contrasts were correlated with their branch lengths, and branch lengths were increased by a factor of 10 and log transformed. This made correlations non-significant and contrasts were Powers el al. • ARTHROPOD FORAGING BY HUMMINGBIRDS 497 weighted equally in subsequent analyses. Stan- dardized contrasts were positivized according to Garland et al. { 1992). We perlormed least-.squares regression through the origin on positivized contrasts ot log gut mass versus log body mass and log gut length versus log body mass. Respiratory Quotient. — Respiratory quotient (RQ) measures the proportion of carbohydrate fuel an animal catabolizes during the measure- ment process by comparing carbon dioxide production with oxygen consumption (RQ = VCO2W02). A measurement close to 1.0 represents pure carbohydrate use, while for birds an RQ close to 0.7 represents pure protein/fat use. Comparing RQs between species provides a view of the importance of nectar as a fuel at a given point in time. A bird foraging more on arthropods might be expected to have an RQ closer to 0.7. We predicted RQs for wild-caught E. fulgens would be lower than those for the other two species. Birds were captured in mist nets three times each day (morning, midday, and late afternoon) and randomly assigned to one of two groups: (1) those caught from the wild and not fed, and (2) those birds that were fed sucrose solution after being caught but prior to RQ measurements. Fed birds were allowed to drink their fill of 0.86 M sucrose solution, and served as a control because they were presumably catabolizing primarily carbohydrate. The RQ of unfed individuals was expected to reflect the “normal” metabolic substrate for a bird at a certain time of day. Measurements of O2 consumption (V02) and CO2 production (VCO2) were made using an open- circuit, positive-pressure respirometry system (Powers 1991). Body mass was measured to the nearest 0.1 g both before and after trials using a portable balance (Model No. LS 200, Ohaus Coip., Pine Brook, NJ, USA). The birds for each trial were kept for about 0.5 hr in a metabolism chamber within an environmental chamber at a constant temperature (27° C) prior to making measurements. Metabolism chambers consisted of a large Mason jar (effective volume: 800 ml) for E. fulgens and L. clemenciae or a small Mason Jar (effective volume: 380 ml) for A. ale.xandri. Metabolism chambers were placed in an environ- mental control chamber (I-35L, Percival Scientif- ic, Perry, lA, USA). Temperatures were recorded to the nearest 0.1 °C using a Physitemp Bat- 12 (Physitemp Instruments, Clifton, NJ, USA) and a Cu-Cn thermocouple. The flow rate of dry, CO2- free air through the metabolism chamber was 500 ml/min. The inlet air passed through soda lime and Drierite to remove ambient CO2 and water vapor, respectively, prior to passing through an O2 analyzer (Model No. S-3A, Applied Electrochemistry, Naperville, IL, USA) and a CO2 analyzer (Model No. LI-6262, LI-COR Biosciences, Lincoln, NE, USA). Data acquisition and analysis were with a Power Macintosh (Model No. 7200, Apple Computer Corp., Cupertino, CA, USA) using Warthog LabHelper and LabAnalyst software (Mark Chappell, University of Califor- nia, Riverside, CA, USA). Statistical Analysis. — We used nonparametric analysis of variance (Kuskal-Wallis test; Zar 1974) to evaluate differences among sample means. Post hoc testing was by multiple compar- isons using Mann-Whitney U-tests (Zar 1974) when more than two groups were compared with alpha values adjusted using the Bonferroni coiTection. Curves illustrating allometric relation- ships were plotted using simple linear regression and the extent of variation explained by the regression is reported as (Zar 1974). All statistical calculations used SPSS 16.0.1 (SPSS 2007). Values reported are means ± SD. RESULTS Arthropod Eoraging. — Individual A. ale.xandri were difficult to locate and observe foraging for arthropods, possibly because the presence of one or more aggressive L. clemenciae motivated these birds to remain inconspicuous. E. fulgens and L. clemenciae both exhibited a higher arthropod foraging rate aerially than while perched (E. fulgens: U\ ^\ = 21.5, P < 0.001; L. clemenciae: ^1.81 = 27.9, P < 0.001; Eig. 2). E. fulgens had a higher aerial foraging rate than L. clemenciae (^1.113 = 4.8, P = 0.03). L clemenciae perhaps compensated for the lower aerial foraging rate by spending more time in areal foraging than E. fulgens (f/|.25 = 6.93, P = 0.01; T = 0.09 ± 0. 1 min/hr for E. fulgens, x = 0.52 ± 0.6 min/hr for L. clemenciae). The majority of individual E. fulgens engaged in aerial arthropod foraging (87% of 93 observations) rather than perch foraging (13% of 93 observations). Individual L. clem- enciae, in contrast, divided arthropod foraging activity more evenly with 45% engaged in perch foraging and 55% foraging aerially (95 total observations). These birds would typically drink from a feeder and then perch in close proximity to the feeder. L. clemenciae would often engage in 498 THE WILSON JOURNAL OL ORNITHOLOGY • Voi 122, No. 3, September 2010 Species FIG. 2. Mean (± SD) number of insect capture attempts per second for E. fulgens, and L. clemenciae during a foraging bout. Numbers above the error bars are .sample sizes. The indicates a significant difference between the aerial and perch values for both E. fulgens and L. clemenciae. The “t” indicates a significant difference between aerial values for E. fulgens and L. clemenciae. this behavior for a short time and then fly to defend the nearest feeder or to feed from it. There was a marked difference in the average height from which the two species foraged for arthropods with E. fulgens perching at a greater height in both modes than L. clemenciae (aerial foraging height f/|,8i= 32.2, P < 0.001; x = 9.9 ± 4.7 m for E. fulgens, x = 3.6 ± 2.5 m for L. clemenciae', perch foraging height; f/1,37 = 5.23, P = 0.02; X = 5.9 ± 4.0 m for E. fulgens, x = 2.1 ± 1.2 m for L. clemenciae). Both species engaged in aerial foraging from a higher perch than when perch foraging {E. fulgens'. (7i,59 = 5.39, P = 0.02; L. clemenciae'. Ui_s9 — 5.84, P = 0.02). E. fulgens appeared to remain high in the tree canopy (especially in Juglans spp. and Platanus spp.), while L. clemenciae perched in lower tree limbs or understory shrubs where it was closer to feeders. Crop and Gizzard Contents. — Guts of E. fulgens contained the most arthropod taxa of our three study species (Table 1). The number of taxa represented (from 2 Classes and 4 Orders) is conservative because only arthropods that could be positively identified (to the lowest possible taxon) are reported. This meant identifying nearly whole specimens only. Absolute number of individuals in each of the taxa could not be calculated because most ingested arthropods were fragmented. All three identifiable members of the Order Hymenoptera were wasps. Several Homop- teran specimens were the same type of leafhopper (Family Cicadellidae). There were three insects and one arachnid specimen that could not be further identified. Numbers and size range of whole arthropods as well as the dry mass of all arthropod material collected from the three hummingbird species digestive tracts did not differ (dry mass: 6/2.16 = 2.096, P = 0.351; whole arthropods: //2.16 = 2.228, P = 0.328; Table 2). Gut Allometry. — The linear regressions of gut mass and body mass, and the phylogentically independent contrasts (PIC) of gut mass and body TABLE 1 Arthropod taxa found in crops and gizzards of hummingbird species in this study. Species A* B" C" D" E" F- G’ H" I' E. fulgens L. clemenciae A. alexandri X XXX XXX XX X X XX X “ A = Cla.ss In.secta, Order Hymenoptera (length .1.4 - .1.8 mm); B = Cla.ss Insecta, Order Hymenoptera (length 2..1 mm); C = Class Insecta. Order Hymenoptera (length 1.0 mm); D = Cla.ss Insecta. Order Homoptera, Family Cicadellidae; E = Cla.ss Insecta, Order Diptera; F = Class Insecta (unknown Order"); G = Class Insecta (unknown Order"); H = Cla.ss In.secta (unknown Order"); I = Class Arachnida. Order Araneae. Different from other listed categories. TABLE 2. Arthropods {x ± SD) in crops and gizzards of hummingbirds in this study. Species n Whole arthropods (#/individual) Arthropod length (range, mm) Dry ma.ss (mg/g"’") L. clemenciae E. fulgens A. alexandri 5 1.8 ± 3.0 2.47-2.55 0.9 ± 1.2 6 6.2 ± 7.3 1.02-5.75 1.6 ± 1.6 6 0.8 ±1.1 0.97-3.50 1 .6 ± 1 .0 Powers et cil. • ARTHROPOD FORAGING BY HUMMINGBIRDS 499 >' - 0.37UII - 1 .276 y = l)-UUO C D FIG. 3. The relationship between gut mass and length, and body mass in hummingbirds ranging in size from ~2 to 20 g. A and B are the standard linear regression and the linear regression of independent contrasts for gut mass as a function of body mass. C and D are the standard linear regression and the linear regression of independent contrasts for gut length as a function of body mass. Triangles in A and C are hummingbird species in this study. mass were both significant (Fig. 3A, B) indicating body mass is a key factor affecting gut mass. The linear regression and PIC of gut length and body mass was also significant (Fig. 3C, D) indicating gut length varied with body mass. Respiratory Quotient. — Respiratory quotient (RQ) within a species did not vary with time of day and the data for each species were pooled (Fig. 4). There was no difference between fed and unfed RQ for A. alexandri (Cue = 2.91, P = 0.09). There was a difference between fed and unfed RQ for both E. fulgens and L. clemenciae (E. fulgens: f/1,19 = 13.7, P < 0.001; L. clemenciae: (/| j9 = 12.39, P < 0.001). E. fulgens also had a lower unfed RQ than A. alexandri or L. clemenciae (7/2,39 — 6.88, P = 0.032). There was little variation in unfed RQ in E. fulgens. DISCUSSION Nectar carbohydrates have been assumed to be the primary energy source for hummingbirds, but arthropods may also provide significant amounts of energy in addition to protein and lipids. Most insects from sweep net samples (Weathers and E. fulgens L. clemenciae A. alexandri Species FIG. 4. Fed and unfed mean (± SD) respiratory quotient (RQ) for A. alexandri . E. fulgens. and L. clemenciae. The indicates a significant difference between fed and unfed birds of the same species. The “t” indicates significance between unfed RQ for E. fulgens and the other species. 500 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 3. September 2010 Sullivan 1989) conducted in the Chiricahua Mountains were from the Orders Coleoptera, Homoptera, and Diptera. Weathers and Sullivan (1989) showed these insects have an average energy value of about 25 kJ/g (dry mass). Karasov’s (1990) conservative estimate of the metabolizable energy of insets is 19.3 kJ/g (dry mass) compared to 16.4 kJ/g for nectar. Thus, energetically, arthropods have the potential to be more than a small supplement to a primarily carbohydrate diet. Hainsworth (1977) showed that if a hummingbird devotes equal time to nectar- ivory and flycatching, even low efficiency rates of ~4()% can provide more energy than nectar feeding. E. fulgens seemed to emphasize aerial foraging for arthropods. They spent the vast majority of their arthropod foraging time in this mode, had the highest frequency of capture attempts, and generally foraged from higher perches. L clem- encicie split their time more evenly between perch and aerial foraging, but actually spent more total time foraging for arthropods. It is possible that we actually observed a relatively small portion of their total arthropod-foraging activity because of the traplining behavior of E. fulgens. Sandlin (2000a) suggested that E. fulgens switches to arthropod foraging to avoid the con.sequences of competition. Successful territo- rial species such as L. cletnenciae maintain reliable nectar sources by defending them from invaders thereby gaining an energetic advantage (Powers et al. 2003). Thus, L. cletnenciae may prefer to perch close to feeders, facilitating quick responses to intruders. This would be an efficient way to supplement their nectar diet for L. cletnenciae foraging on arthropods while perched in a defensive position. Powers and Conley ( 1994) report that L. cletnenciae are perched about 70% of the time, allowing them to both guard their territory and to forage for passing insects when available. E. fulgens had the greatest diversity of ar- thropod species in its crop and gizzard (Table 2). The.se data corroborate the findings of Cottam and Knappen (1939) who examined gut contents of E. fulgens and L. cletnenciae from the nearby Huachuca Mountains of Arizona. They found 13 different taxonomic specimens in the guts of E. fulgens but only .seven in E. cletnenciae. The reason for this difference is unclear. E. fulgens forages for arthropods higher in the canopy than E. cletnenciae and the greater arthropod diversity in E. fulgens guts may reflect the diversity of arthropods available in the upper canopy. It is also possible the trapline-foraging behavior of E. fulgens diversifies the arthropod component of the diet (Colwell 1973, Feinsinger and Chaplin 1975). Our data show that E. fulgens and L. cletnencae forage differently for arthropods but there is no evidence for a difference in arthropods consumed. It is not possible to calculate the total contribution of arthropods to the diet of the hummingbirds in this study. It has been suggested the types of arthropods we found in the guts are digested quickly (Remsen et al. 1986, Stiles 1995). Thus, our measures of dry mass are probably a good index of short-term arthropod consumption and correspond reasonably well to the maximum rate arthropods were captured (range ~6 to 30 captures/hr assuming 100% capture efficiency and total arthropod digestion in 1 hr). The total digestible energy of arthropods consumed ranges from 0.2 to 0.4 kJ/day if arthropod dry mass is turned over in the gut hourly. Daily energy expenditure (DEE) is 82 kJ/ day for L. cletnenciae and 29 kJ/day for A. alexandri (Powers and Conley 1994). Our data suggest arthropods account for < 1 % of DEE assuming DEE for E. fulgens is similar to L. cletnencae. These data suggest there is no difference in arthropod consumption or energy derived from arthropods for our three humming- bird species. Our hummingbirds had gut masses and lengths that corresponded with their body size suggesting that none of the hummingbirds in this study has a unique gut design. Any ability to subsist on arthropods when nectar is scarce is likely a common trait in hummingbirds. All three hummingbird species had RQ values <0.85 indicating they were not strictly cataboliz- ing carbohydrate. The RQ of E. fulgens was statistically lower than the other two species but we are uncertain if the difference is truly enough to argue a difference in use of arthropods for energy. The low variability in E. fulgens RQ does at least hint at a difference in how the three species forage but the exact nature of this difference is unclear. The higher variability in the RQ of L. cletnenciae and A. ale.xandri could correlate with time since their last nectar meal. The lower variability in E. fulgens could support more consistent arthropod consumption but it could also be explained by the routine capture of Powers el al. • ARTHROPOD FORAGING BY HUMMINGBIRDS 501 traplining individuals that have not yet fed. RQ data alone are insufficient to draw conclusions regarding the importance of different metabolic substrates used by the hummingbirds in this study. Hummingbirds can rapidly switch between fatty acid and carbohydrate oxidation (Welch and Suarez 2007) making interpretation of a depres- sion in RQ difficult. E. fiilgens differs behaviorally in arthropod foraging from at least L. clemenciae but there is no strong evidence that arthropod consumption provides a disproportionate amount of energy compared to the other hummingbird species in this study. It is probable that all three species energetically benefit from arthropods but quanti- fying the total contribution of insects to their energy budget is difficult. We still do not have a complete understanding of the role arthropods have in hummingbird nutrition but are hopeful that future studies using techniques such as stable isotope analysis might provide addition insight. ACKNOWLEDGMENTS We thank the American Museum of Natural History and Southwestern Research Station for use of their research facilities. We also thank the Murdock Charitable Trust and Holman Endowment for the Sciences for providing funding for this project. We thank Amber Hamilton and N. A. Miller for assistance in collecting behavioral data in the 1999 field season. M. Victoria Lopez-Calleja, Carlos Martinez del Rio, and J. E. Schondube generously shared unpublished data on hummingbird gut morphology. We thank C. A. Engman and Ryan Lapour for editing assistance, and D. J. Kimberly and D. L. Swanson for technical advice. We thank W. W. Weathers and an anonymous reviewer for their constructive comments on this manuscript. LITERATURE CITED Baker, H. G. 1975. Sugar concentrations in nectars from hummingbird flowers. Biotropica 7:37-42. Baker, H. G. 1977. Non-sugar chemical constituents of nectar. Apidologie 8:349-356. Brice, A. T. 1992. The essentiality of nectar and arthropods in the diet of Anna’s Hummingbird (Ccilypte anna). Comparative Biochemistry and Physiology 101 A: 151- 155. Brown, J. H., W. A. Kodric-Brown, and A. Kodric- Brown. 1978. Correlates and consequences of body size in nectar-feeding birds. American Zoologist 18:678-700. Calder III, W. A. AND J. R. King. 1974. Thermal and caloric relations in birds. Pages 260-413 in Avian biology (D. F. Earner and J. R. King, Editors). Volume 4. Academic Press, New York, USA. Chavhz-Ramirez, F. and M. Dowd. 1992. Arthropod feeding by two Dominican hummingbird species. Wilson Bulletin 104:74.3-747. Cot.WELi., R. K. 1973. Competition and coexistence in a simple tropical community. American Naturalist 107:737-760. COTTAM, C. AND P. Knappen. 1939. Food of .some uncommon North American birds. Auk 56:138-169. Feinsinger, P. 1983. Coevolution and pollination. Pages 282-310 in Coevolution (D. J. Futuyma and M. Slatkin, Editors). Sinauer Associates, Sunderland, Massachusetts, USA. Feinsinger, P, and S. B. Chaplin. 1975. On the relationship between wing disc loading and foraging strategy in hummingbirds. American Naturalist 109:217-224. Felsenstein, j. 1985. Phylogenies and the comparative method. American Naturalist 125:1-15. Garci'a-Deras, G. M., N. Cortes, M. Money, A. G. Navarro, J. Garci'a-Moreno, and B, E. Hernan- dez-Banos. 2008. Phylogenetic relationships within the genus Cynanthus (Aves: Trochilidae), with em- phasis on C. doubledayi. Zootaxa 1742:61-68, Garland Jr., T., P. H. Harvey, and A. R. Ives. 1992. Procedures for the analysis of comparative data using phylogenetically independent contrasts. Systematic Biology 41:1 8-32. Garland Jr., T., P. E. Midford, and A. R. Ives. 1999. An introduction to phylogenetically based statistical methods, with a new method for confidence intervals on ancestral values. American Zoologist 39:374-388. Gass, C. L. and R. D. Montgomerie. 1981. Hummingbird foraging behavior: decision-making and energy regu- lation. Pages 159-194 in Foraging behavior: ecolog- ical, ethological and psychological approaches (A. C. Kamil and T. D. Sargent, Editors). Garland STOP Press, New York, USA. Hainsworth, F. R. 1977. Foraging efficiency and parental care in Colihri coruscans. Condor 79:69-75. Karasov, W. H. 1990. Digestion in birds: chemical and physiological determinants and ecological implica- tions. Studies in Avian Biology 13:391-415. Li, S.-H. and j. L. Brown. 1999. Influence of climate on reproductive success in Mexican Jays. Auk 1 16:924- 935. Marshall Jr., J. T. 1957. Birds of pine-oak woodland in southern Arizona and adjacent Mexico. Pacific Coast Avifauna 32:5-1 25. McGuire, J. A., C. C. Witt, D. L. Altshuler, and J. V. Remsen Jr. 2007. Phylogenetic systematics and biogeography of hummingbirds: Bayesian and maxi- mum likelihood analyses of partitioned data and selection of an appropriate partitioning strategy. Systematic Biology 56:837-856. Murphy, M. E. 1996. Nutrition and metabolism. Pages 31- 60 in Avian energetics and nutritional ecology (C. Cary, Editor). Chapman and Hall, New York, USA. Pimm, S. L., M. L. Rosenzweig, and W. Mitchell, 1985. Competition and food selection: field tests of a theory. Ecology 66:798-807. 502 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 3, September 2010 PiTELKA, F. A. 1942. Territoriality and related problems in North American hummingbirds. Condor 44:189-204. Powers, D. R. 1991. Diurnal variation in mass, metabolic rate, and respiratory quotient in Anna’s and Costa’s hummingbirds. Physiological Zoology 64:850-870. Powers, D. R. 1996. Magnificent Hummingbird {Eugenes fiilgens). The birds of North America. Number 221. Powers, D. R. and T. M. Conley. 1994. Field metabolic rate and food consumption of two sympatric hum- mingbird species in southeastern Arizona. Condor 96:141-150. Powers, D. R., A. R. Brown, and J. A. Van Hook. 2003. Influence of normal daytime fat deposition on laboratory measurements of torpor use in territorial versus nonterritorial hummingbirds. Physiological and Biochemical Zoology 76:389-397. Remsen Jr., J. V., F. G. Stiles, and P. E. Scott. 1986. Frequency of arthropods in stomachs of tropical hummingbirds. Auk 103:436-439. Rosenzweig, M. L. 1985. Some theoretical aspects of habitat selection. Pages 517-540 in Habitat selection in birds (M. L. Cody, Editor). Academic Press. New York, USA. Sandlin, E. A. 2000a. Cue use and resource subdivision among three coexisting hummingbird species. Behav- ioral Ecology 1 1 :550-559. Sandlin, E. A. 2000b. Foraging information affects the nature of competitive interactions. Oikos 91:18-28. SPSS. 2007. SPSS 16.0 for Macintosh. SPSS Inc., Chicago, Illinois, USA. Stiles, F. G. 1995. Behavioral, ecological and morpholog- ical correlates of foraging for arthropods, by the humming- birds of a tropical wet forest. Condor 97:853-878. Weathers, W. W. and K. A. Sullivan. 1989. Juvenile foraging proficiency, parental effort, and avian reproduc- tive success. Ecological Monographs 59:223-246. Wolf, L. L. and F. R. Hainsworth. 1971. Time and energy budgets of territorial hummingbirds. Ecology 52:979-988. Welch, K. C. and R. K. Suarez. 2007. Oxidation rate and turnover of ingested sugar in hovering Anna’s (Calypte anna) and Rufous {Selasphonts ntfus) humming- birds. Journal of Experimental Biology 210:2154— 2162. Young, A. M. 1971. Foraging for insects by a tropical hummingbird. Condor 73:36-45. Zar, j. H. 1974. Biostatistical analysis. Prentice-Hall Inc., Englewood Cliffs, New Jersey, USA. The IVilson Journal of Ornithology 1 22(3):5()3-5 1 2, 2010 ARMY ANT RAID ATTENDANCE AND BIVOUAC-CHECKING BEHAVIOR BY NEOTROPICAL MONTANE FOREST BIRDS SEAN O’DONNELL,''^ ANJALl KUMAR,' AND CORINA LOGAN- ABSTRACT. We quantified resident and migrant bird attendance at army ant swarm raids {n = 48) in a neotropical montane forest. All observations were during seasons when Nearctic migrant birds are present. Bird species differed in army ant raid-attending behavior. Resident bird species attended 2 to 54% of raids, while migrants attended at lower maximum frequencies (2 to 21% of raids affended per species). Some residenf and migrant bird species attended raids more frequently than expected based on capture rates in mist-net studies and point-count density surveys. Army ant raid attendance may be a regular element of foraging behavior for some resident species, and important in the wintering ecology of some Nearctic migrant species. The bird species that attended raids most frequently were predicted to show behavioral specializations for exploiting army ant swarms. Eight resident bird species (but no migrants) performed a specialized behavior, bivouac checking, by which birds sample army ant activity. Resident bird species’ frequencies of raid attendance were positively associated with frequency of checking bivouacs (/• = 0.68). We hypothesize the absence of obligate army ant-following birds in montane forests has favored performance of specialized behaviors for exploiting army ant raids by some resident birds. Received 3 October 2009. Accepted 9 March 2010. Neotropical army ants (Formicidae: Ecitoninae) are top predators, and a diverse array of animal species associate with army ant colonies (Franks 1982, Franks and Bossert 1983, Brady 2003, Koh et al. 2004). Birds attend army ant foraging-raids to feed on arthropods and small vertebrates that flee from the advancing ants. Birds primarily attend the swarm raids of Eciton burchellii and Labidus praedator (Willis and Oniki 1978, Wrege et al. 2005). Bird flocks at army ant raids often include multiple species, and their composition is largely distinct from sympatric mixed foraging flocks of insectivores (Willis 1972, Willis and Oniki 1978, Otis et al. 1986, Willson 2004, Peters et al. 2008). Some bird species are obligate army ant raid attendants that obtain most or all of their food at army ant swarms (Willis and Oniki 1978, Swartz 2001, Willson 2004, Brumfield et al. 2007). Other bird species attend raids opportunistically (Swartz 2001, Chaves-Campos 2003). Opportunistic army ant raid-attending bird species vary in their reliance on army ants (Willis 1972, Willis and Oniki 1978). Obligate army ant-following birds are agonistic toward other birds at raids in lowland forests. This interference competition reduces the value of ant raids as a food source ' Animal Behavior Program, Department of P.sychology, Box 351525, University of Washington, Seattle, WA 98195, USA. ^University of Cambridge, Department of Experimental Psychology, Cambridge CB2 3EB, United Kingdom. ^Corresponding author; e-mail: sodonnel@uw.edu to Other birds (Willis 1966, Willis and Oniki 1978, Brumfield et al. 2007). Obligate army ant-following birds are poorly represented or absent from montane forests (Willis and Oniki 1978, Brumfield et al. 2007). There are no obligate army ant-following birds at our study site near Monteverde, Costa Rica (Kumar and O’Donnell 2007). Birds from a diverse array of families attend army ant raids in the Monteverde area, including some resident and Nearctic migrant species (henceforth migrants; Vallely 2001, Kumar and O’Donnell 2007). We hypothesized that some montane bird species would exhibit behavioral specializations for exploiting army ant raids in the absence of local competition from obligate army ant-following birds. We asked whether some montane bird species attend raids more often than expected as a first test of this hypothesis. The frequency of raid attendance varies among Monteverde area birds, but attendance frequency alone does not account for possible effects of local abundance (Vallely 2001, Kumar and O’Donnell 2007, Peters et al. 2008). We extended our previous analyses estimating the effects of species’ relative abun- dance on army ant raid attendance in this study. Bird abundance estimates were derived from previously published mist-net captures and point-count den.sities (Young et al. 1998, Jan- kowski et al. 2009). We predicted the most frequent raid attendant birds would be more likely to exhibit specialized behaviors for exploiting army ant swarms (Willis 1972, Willis and Oniki 1978, Swartz 2001, Chaves-Campos 2003). We ascertained whether 503 504 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. 3. September 2010 montane forest birds perform bivouac-checking behavior to test this prediction. Bivouac-checking birds visit Eciton burchellii temporary nests (bivouacs) as the ant foraging raids start in the morning, and again in the evening. Bivouac- checking birds may assess army ant raid activity and direction in the morning, or whether ants are emigrating to a new bivouac site in the evening (Swartz 2001, Chaves-Campos 2003). We corre- lated bird species’ frequencies of bivouac check- ing with their frequencies of attendance at army ant raids to examine whether raid attendance frequency and bivouac-checking behavior were associated among bird species. METHODS Study Site. — Our sample sites spanned the continental divide in the Tilaran Mountain Range near Monteverde, Puntarenas Province, Costa Rica (10° 18' N, 84° 47' W). The Monteverde area includes forest reserves with adjacent pri- vately held continuous forest and associated forest fragments (Guindon 1997, Haber 2000, Harvey 2000, Jankowski et al. 2009). We collected data in fore.sts between 1,175 m above mean sea level (m asl) and 1,580 m asl elevation on the Pacific Slope, and between 985 and 1,680 m asl elevation on the Atlantic Slope. Our raid attendance samples included four life zones (Holdridge life zone system; Holdridge 1966, Guindon 1997, Young et al. 1998, Young and MacDonald 2000). Sampling Bird Flock Composition. — We col- lected data during seasons when migrants occur in the Monteverde area (Stiles and Skutch 1989, Garrigues and Dean 2007). The four field trips collectively spanned much of the period of migrant presence (mid-Sep to late May): 1 1 January-4 March 2005 (dry season), 5 October- I 1 December 2005 (late wet season/onset of dry season), 19 December 2007-1 January 2008 (early dry season), and 7-10 April 2008 (late dry season). We ob.served birds at 48 army ant raids during these trips. The army ant raids involved three swarm-raiding army ant species (Eciton hurchellii, n = 40; Lahidus praedator, n = 5; and L. spinninodis, n = 3). We did not separate army ant species in our analyses as the Lahidus species raids were distributed across all life zones, and army ant species had no measur- able effect on attending bird flock composition in the Monteverde area (Kumar and O’Donnell 2007). We observed a maximum of one swarm per day and alternated sampling dates among elevations to minimize order effects. We treated each army ant swarm as an independent data point because observations were separated in space and time. Observations were made during the diurnal active raid period of surface-foraging army ants. Start times ranged from 0900 to 1600 hrs local time (CST). We located flocks of army ant swarm-attending birds or foraging army ant columns by searching along established trails. We walked to swarm raid fronts where birds fed and positioned ourselves at the best location for unobstructed viewing of the swarm front; this was usually off to one side and facing in the direction the swarm was moving. If the raid or the bulk of bird activity shifted location during observations we walked to a new observa- tion spot while attempting to avoid disturbing the birds. We noted all bird species and the number of individuals that were present upon our amval (Coates-Estrada and Estrada 1989). The observa- tion sessions lasted for 1 hr except when raid activity ceased, heavy rainfall began, or ants traversed impassable terrain (.y ± SD observation time = 47 ± 18 min). We analyzed only those flocks for which all attending birds were identi- fied to species. We collected four types of data on bird flock composition and bird behavior at the army ant swarms. (1) Start and end time to the nearest min. (2) Latitude/longitude coordinates and elevation to the nearest 10 m asl were taken with a hand- held Geographic Positioning System (GPS) unit or an air pressure altimeter. Elevations were con- firmed from topographic maps. (3) We identified all birds seen in attendance to species based on plumage appearance, behavior, relative body size, geographic range, and vocalizations (Stiles and Skutch 1989, Garrigues and Dean 2007). A bird had to be observed collecting prey that was fleeing from ants to be counted as an attendant. Birds were categorized as residents or migrants. (4) We recorded the number of individuals of each bird species present. The birds were not banded and we could not distinguish individuals that left the raid and returned from newly arriving swarm attendants. Thus, we used the maximum number of individuals that were observed simultaneously as a conservative estimate of the number of attending birds from each species (Coates-Estrada and Estrada 1989). Bivouac-checking Observations. — We conduct- ed 15 systematic watches for bivouac-checking behavior at E. burchellii bivouacs following O’Donnell Cl al. • MONTANE BIRDS AT ARMY ANT SWARMS 505 Swartz (2001 ). We sal on (he ground or on folding chairs at sites 5 to 10 ni from the bivouac, facing the entrance of the bivouac shelter. We chose locations that were partially concealed by vege- tation but had an unobstructed view of the bivouac entrance. Our observation durations ranged from 60 to 120 min (.v = 87.4 min/session; 21.9 hrs of observation). Like Swartz (2001), we conducted 12 observation sessions in the morning (start times 0440 to 0800 hrs) and extended the protocol by conducting three bivouac watches in the afternoon (start times 1533 to 1630 hrs). We recorded bivouac-checking behavior and counted a bird as checking the bivouac if it first flew toward the bivouac site and perched within 5 m of the bivouac. The bird then had to either peer at the bivouac shelter entrance, or land among the ants when ants were active outside the bivouac to be counted. We noted bird arrival time to the nearest minute and, whenever possible, recorded the duration of each bird’s visit. We also recorded the species identity of birds that were observed performing bivouac-checking behavior while we were collecting other data (n = 5 observations; Swartz 2001). We used the same behavioral criteria for bivouac checking as during systematic observations. Dates of bivouac-checking observa- tions were: 15 November to 3 December 2005, 22 to 29 December 2007, 7 to 1 0 April 2008, and 5 to 25 July 2009. We made six systematic and two opportunistic bivouac-checking observations dur- ing seasons when Nearctic migrants are present at the field site. We could not identify birds as individuals as birds were not banded and many species were sexually monomorphic in plumage. We used only one visit/species/observation session to estimate frequencies of bivouac checking. All but one of our bivouac-checking observations were in the premontane wet forest life zone. We sampled bivouac checking at elevations from 1,100 to 1,430 m asl; one systematic ob.servation was conducted at a higher elevation (1,575 m asl) in the lower montane wet forest life zone, but no birds were recorded bivouac checking at this site. We also noted if birds vocalized while bivouac checking in 2008 and 2009. We used observations of bird attendance at a larger sample of army ant raids {n = 54) for comparisons of raid attendance with bivouac checking. These data were collected during several field trips between 25 January 2005 and 10 April 2008. These raids partially overlap with the migrant season raid sample, but all were observed in the same life zone and over a similar range of elevations as our bivouac-checking observations (1,100 to 1,475 m asl). Statistical Analyses. — We calculated two mea- sures of bird species attendance at swarm raids based on species richness and individual abun- dance at the 48 migrant season raids (Coates- Estrada and Estrada 1989): (1) the percent of raids at which each species was present, and (2) the percent of all raid-attending birds accounted for by each species. We used two independently collected data sets to estimate forest understory activity and relative abundance of bird species. We compared these with our abundance measure of army ant raid attendance (i.e., each species’ percent of all birds at raids) to identify species that were present at raids more often than expected by chance. The estimate of relative activity came from mist-net data from Monte- verde area forests (Young et al. 1998). We calculated the number of mist-net captures for each attending bird species from Young et al. (1998: table 2), summing total captures for each species over the four life zones that overlapped our sample area. Mist-net studies of bird abun- dance must be interpreted with caution (Remsen and Good 1996). For example, bird species may differ in probability of being caught in nets in the understory. We used an independently-derived estimate of relative abundance from standardized auditory and visual point-count data to comple- ment the mist-net data. The point-count data also came from sites largely overlapping with our observations (Jankowski et al. 2009: supplemen- tary table 1). We followed Peters et al. (2008) to calculate two indices of raid attendance for each bird species. We regressed each species’ proportion of the total birds observed at raids against its total captures in the mist-net studies (Young et al. 1998), and against its total/ha abundance estimat- ed from point counts (Jankowski et al. 2009). We calculated residuals from these two linear regres- sions and used the residuals as indices of raid attendance. We identified species as high-fre- quency army ant raid attendants when their index value (regression residual) was higher than the upper 95% confidence interval (Cl) value from the con-esponding linear regression model. We calculated the mist-net index separately for residents and migrants because resident birds were mist-netted year-round, including months when migrants are absent. We could not calculate 506 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 3. September 2010 a point-count index for migrants as they were not recorded in the point-count study, in part because migrants rarely vocalize outside their breeding ranges (Jankowski et al. 2009). We asked whether bird species differences in raid attendance were correlated with bivouac- checking frequency. We calculated the Pearson correlation between an index of raid attendance frequency (i.e., the point-count attendance index) and bivouac-checking frequency for the 31 resident bird species observed at army ant raids in the bivouac-checking sample elevation range. RESULTS Bird Flock Composition at Ant Raids. — We observed 54 species of birds attending raids during the period of migrant presence (Table 1). The number of bird species in flocks at army ant raids ranged from one to 17 (x ± SD = 4.73 ± 3.27), while the number of migrant species ranged from zero to six (0.79 ± 1.29). We recorded 11 species of migrants attending raids; two were thrushes (Turdidae), one was a vireo (Vireonidae), and eight were wood warblers (Parulidae) (Ta- ble 1). We recorded 13 bird species at raids that had not been noted as army ant raid attendants in the Monteverde area previously ( 1 1 new residents and 2 new migrants). Migrants participated in 21 (43.8%) of the flocks at army ant swarms. None of the swarm-attending flocks was comprised only of migrants. Resident Species Differences in Raid Atten- dance.— Resident bird species varied widely in frequency of raid attendance. Resident species ranged from 2.1 to 54.2% of army ant raids attended, and from 0.2 to 10.8% of the individual birds at army ant raids (Table 1 ). Species’ point- count densities and mist-net captures were positively but weakly correlated (r = 0.42, n = 54, P < 0.01). These measures provide different information on bird species’ baseline rates of occurrence in the habitat. Both measures of baseline density were non-significantly related to army ant raid attendance by the 43 resident bird species. Point-count estimates of bird species’ abundances were poor predictors of species occurrence at army ant raids {R^ = 0.004, df = 53^ P = 0.64). Similarly, capture frequency in mist-net studies was weakly associated with attendance at army ant raids {R^ = 0.06, df = 42, P = 0.10). Both indices of frequency of raid attendance indicated strong species differences in frequency of army ant raid attendance after accounting for density effects. There was high consistency between the two indices of which bird species attended army ant raids at higher than expected frequencies. Seven resident bird species’ frequencies at raids exceeded the 95% Cl value predicted by linear regression of raid attendance on mist-net capture frequency (Fig. 1). Six of these seven species (and one other) were signif- icantly more frequent at raids than expected based on point-count densities (Fig. 1). Migrant Species Differences in Raid Atten- dance.— Relative abundances of migrants at raids spanned an order of magnitude among species in terms of percent of raids attended (2.1 to 20.8%) and in percent of individual birds seen at raids (0.19 to 1.87%; Table 1). The Kentucky Warbler {Oporonis formosus) was the most frequent migrant at army ant raids by both measures. Mist-net capture frequency of migrant species was a poor predictor of their proportion of raid- attending birds {R~ = 0.07, n = 11, P = 0.43). Examination of residuals from this analysis suggests three migrant bird species attended raids more frequently than expected by mist-net captures (Fig. 2). Bird Species that Check Bivouacs. — We ob- served eight species of birds from eight families checking bivouacs in the Monteverde area (Ta- ble 1). Birds checked bivouacs during 10 of the 15 systematic watches (67%); birds were seen bivouac checking during 67% of both morning and afternoon watches. The number of species seen during an observation session ranged from zero to four {x — 1.47 species). We did not see more than one individual of a given species airive at the same time. Birds of three species were present at a bivouac simultaneously on one occasion, but were not seen interacting. Six species checked bivouacs both in the morning and afternoon (Table 1); the other two species checked in the morning but were .seen only once. Bird Behavior During Bivouac Checks. — We obtained duration data for 17 bivouac checks by six bird species (the bird was not identified on one timed visit). Bivouac-check durations pooled across species varied from a few seconds to several minutes (range = 5 to 360 sec; -.v ± SD = 103 ± 126 sec). Visit lengths did not differ significantly among species (Kruskal-Wallis test, X- = 7.6, df = 5, P = 0.18). Three species. Orange-billed Nightingale- Thrush {Catharus aurantiirostris), Rufous-and- white Wren (Thryothorus rufalhus), and Rufous- O'Donnell ct al. • MONTANE BIRDS AT ARMY ANT SWARMS 507 TABLE 1 . Frequencies of occurrence of birds as foragers at army ant swarm raids in montane forests near Monteverde, Costa Rica. Nearctic migrants are indicated by an asterisk *. Species not seen in previous studies of avian army ant raid attendance in the Monteverde area are indicated by a plus symbol +. Number of raids attended % of raids Number of birds % of Species (H = 48) attended {n = 316) birds Orange-billed Nightingale-Thrush Cathams aurantHrostris 26 54.2 58 10.8 White-eared Ground Sparrow Melozone leucotis 18 37.5 29 5.4 Blue-crowned Motmot Momolus momota 15 31.3 21 3.9 Brown Jay Cyanoconix morio 14 29.2 37 6.9 Slaty-backed Nightingale-Thrush Cathams fuscater 10 20.8 21 3.9 Rufous-and-white Wren Thryothoms mfalbus 10 20.8 13 2.4 Rufous-capped Warbler Basileiitenis rufifrons 10 20.8 11 2.1 Kentucky Warbler *Oporornis fonnosus 10 20.8 10 1.9 Immaculate Antbird Myrmeciza invnaculata 9 18.8 22 4.1 Wood Thrush *Hylocichla mustelina 7 14.6 7 1.3 Golden-crowned Warbler Basileutems culicivorus 6 12.5 8 1.5 Slate-throated Whitestart Myiobonis miniatus 6 12.5 7 1.3 Yellowish Flycatcher Empidonax flavescens 6 12.5 6 1.1 Three-striped Warbler Basileutems tristhatus 5 10.4 12 2.2 Ruddy Woodcreeper Dendrocincla homochroa 5 10.4 7 1.3 Wilson’s Warbler *Wilsonia piisilla 5 10.4 7 1.3 Swainson’s Thrush *Cathams ustulatus 5 10.4 5 0.9 Blue-throated Toucanet Aulacorhynchus prasinus 4 8.3 6 1.1 Chestnut-capped Brush Finch Arremon bmnneinucha 4 8.3 4 0.8 Ruddy-capped Nightingale-Thrush Cathams frantzii 4 8.3 4 0.8 White-throated Thrush Turdiis assimilis 3 6.3 8 1.5 Azure-hooded Jay Cyanolyca cucuUata 3 6.3 5 0.9 10 species including one migrant: Black-and-white Warbler *Mniotilta varia 2 4.2 2 or 3 0.4 or 0.6 22 species including 6 migrants*; Chestnut-sided Warbler. +Dendroica pensylvanica'. Black- throated Green Warbler, D. virens, Ovenbird, +Seiums aiirocapilla', Golden-winged Warbler, + Vennivora chtysoptera; Canada Warbler Wilsonia canadensis; Philadelphia Vireo, Vireo philadelphicus 1 2.1 1 or 2 0.2 or 0.4 508 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122. No. 3. September 2010 Total mist-net captures 0 100 200 300 400 FIG. I. Frequency of army ant raid attendance plotted against estimated local abundance for montane resident bird species (Monteverde, Costa Rica). Solid line: linear regression best fit. Curved dashed lines: 95% confidence intervals (Cl). Triangle symbols indicate species with raid attendance above the 95% Cl (0-b NT = Orange-billed Nightingale-Thrush; B J = Brown Jay; W-e GS = White-eared Ground Sparrow; lA = Immaculate Antbird; B-c M = Blue-crowned Motmot; S-b NT = Slaty-backed Nightingale-Thrush; R-c W = Rufous-crowned Warbler; R-w W = Rufous-and-white Wren). capped Warbler (Basileitteru.s rufifrons), either sang or called while bivouac checking. Ruddy Woodcreepers (Dendrocincla homochroa) and/or Orange-billed Nightingale-Thrushes called or sang in the vicinity of bivouac sites during eight observation sessions, even when birds were not ob.served checking the bivouac. Individuals of these species apparently circled the bivouac site at distanees of lO to 1 5 m, calling or singing from different perches. Raid Attendance and Bivouac Checking. — Biv- ouac checking was associated with high frequen- cies of army ant raid attendance. We observed 42 species of birds attending 54 army ant raids in the elevation range where we sampled bivouac checking. All bird species that checked bivouacs O’Donnell Cl al. • MONTANE BIRDS AT ARMY ANT SWARMS 509 Total mist-net captures - Migrants FIG. 2, Frequency of army ant raid attendance plotted against estimated local abundance for migrant bird species (Monteverde, Costa Rica). Solid line: linear regression best fit. Curved dashed lines: 95% confidence intervals (Cl). Triangle symbols indicate species with raid attendance above the 95% Cl (KW = Kentucky Warbler; WW = Wilson’s Warbler; WT = Wood Thrush). were also seen attending ant raids. The eight species observed checking bivouacs were among the most frequent raid attendants (Table 2). Bivouac-checking birds represented 19.0% of the bird species at these raids, but 57.7% of individual birds at the raids. Furthermore, the percent of bivouacs checked by a species correlated positively (r = 0.68, /? = 31 , P < 0.01) with the point-count raid attendance index lor the 31 resident bird species that attended raids in the same elevation range (Table 2). Six of our systematic watches and two opportunistic observations were in months when Nearctic migrant birds are wintering in the Monteverde area (Nov, Dec, and Apr.). No migrant birds were seen checking bivouacs (Table 2). DISCUSSION Resident Bird Species Differences in Redd Attendance. — Frequency of mist-net capture and point-count densities did not predict variation in raid attendance among bird species observed. Bird species density was not an important determinant of frequency of army ant raid attendance, and bird species must vary in their frequency of raid attendance for other reasons. The two indepen- dently-derived indices of raid attendance for residents identified similar lists of bird species that were at raids more frequently than expected (Peters et al. 2008). These species may depend more heavily on army ants for food, and they may exhibit specialized behaviors for exploiting army ant swarms. These behaviors could include mechanisms for locating and orienting to the ants or to other attending birds, or abilities to track army ant colony activity and movements such as bivouac checking (Willis and Oniki 1978, Swartz 2001, Chaves-Campos 2003). TABLE 2. Bird species observed attending army ant raids and checking Costa Rica. bivouacs in montane forests near Monteverde, Common name ScientiHc name (Family) Percent of bivouac check ob.servaiions in = 20) Afternoon bivouac- checking observed? Percent of raids attended (n = 54) Percent of birds al raids (n = -S27) Raid attendance index rank (n = .11 species') Orange-billed Nightingale- Thrush Cathanis aurantiirostris (Turdidae) 50 Yes 59.3 20.2 1 Ruddy Woodcreeper Dendrocinchi homochroa (Furnariidae) 30 Yes 5.6 0.9 14 Rufous-capped Warbler Basilenterus rufifrons (Parulidae) 15 Yes 35.2 7.6 5 Rufous-and-white Wren Thryothorus rufalhiis (Troglodytidae) 15 Yes 24.1 5.5 6 Blue-crowned Motmot Momotiis moinota (Momotidae) 10 Yes 38.9 8.9 4 Blue-throated Toucanet Aukicorhynchus prasums (Ramphastidae) 10 Yes 9.3 2.4 8 White-eared Ground Sparrow Melozone leiicoti.s (Emberizidae) 5 No 48.1 1 1.9 2 Chirqui Quail-Dove Geotrygon ch i riquen.'iis (Columbidae) 5 No 1.9 0.3 26 510 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 3. September 2010 Ecological Implications of Bivouac-checking Behavior. — Only obligate army ant-following birds checked bivouacs in lowland neotropical forests, even at sites with opportunistic species that frequently attended raids (Swartz 2001, Chaves-Campos 2003). Swartz (2001) suggested bivouac-checking behavior can be used to distin- guish obligate army ant-following bird species from opportunistic raid attendants. Performance of this specialized behavior by Monteverde resident birds indicates army ant swarms are an important food resource to some montane bird species. The relative importance of foraging at ant raids to bird diets has not been quantified, but species differences in raid attendance combined with bivouac-checking behavior indicate some specialization for foraging at army ant swarms. Two bird species were seen checking bivouacs only once, indicating that more montane birds would be seen performing this behavior with additional observations. Whether resident and migrant montane birds exhibit other specialized behaviors to enhance attendance at ant raids requires further study. Some resident attendants vocalized while checking bivouacs, including nightingale-thrushes (Catharus spp.). Calling and singing at bivouac sites by some residents may recruit other birds to attend raids (Chaves-Campos 2003). Bivouac-checking birds devote time and energy to sampling army ant behavior without an immediate food reward, and possibly expose themselves to predators. Bivouac-checking bouts were generally short in duration, consistent with this behavior incurring some cost (Swartz 2001). Some bird species check bivouacs in the morning (at the start of raiding) and in late afternoon (when most colony emigrations begin; O’Donnell et al. 2009), suggesting that birds track movements of ant colonies as well as their daily raiding activity. Individually marking and tracking birds will be necessary to identify how many bivouacs the montane resident birds check on a given day, and how far they travel among bivouacs (Willson 2004, Chaves 2008). We hypothesize that reduced (or absent) competition with obligate army ant-following birds in some montane forests removes biotic constraints on specialized raid-attending behavior for other species. Reasons tor obligate army ant- following bird ab.sence in the Monteverde area Pacific Slope are not known, but lack of local competition with obligate army ant-tollowing birds may enhance the value of army ant swarms to opportunistic raid-attending birds. It remains to be tested whether these behaviors represent local evolutionary adaptations or plastic expression of behavior based on learning or cultural transmis- sion. We predict that observations at other sites without obligate army ant-following birds will reveal specialized army ant raid-attending behav- iors in additional bird species. Ecological Implications of Migrant Raid Atten- dance.— Three species of migrants attended army ant raids more often than expected according to the mist-net capture-based index. The Kentucky Warbler and two thmshes (Wood Thrush [Hylo- cichla nmstelina] and Swainson’s Thrush, [Ca- tharus utitulatus]) have been recorded as army ant-raid attendants from Mexico to Colombia (Willis 1966, 1984; Greene et al. 1984; Coates- Estrada and Estrada 1989; Roberts et al. 2000; Meisel 2004), as have several other migrants we observed attending army ant swarms (Willis 1966, Hardy 1974, Greene et al. 1984, Roberts et al. 2000, Meisel 2004, Rios et al. 2008). Army ant raid attendance is apparently consistent across the wintering ranges of some migrant birds. Swarm- raiding army ants do not occur in the breeding ranges of the migrants we recorded as raid attendants (Watkins 1985). Attendance at army ant raids by wintering migrants is a striking example of plasticity in foraging behavior, and of differences in foraging ecology between breeding and wintering habitats. Changes in foraging ecology are an important component of niche- switching between breeding and wintering sites (Nakazawa et al. 2004). Studies at South Amer- ican sites are needed to learn if similar changes in foraging ecology, including army ant raid atten- dance, are exhibited by Austral migrant birds (Jahn et al. 2004). Migrants were not at raids without residents. Residents may initiate raid-attending flocks that are later joined by migrants, possibly in part because residents are locally more abundant (Young et al. 1998, Jankowski et al. 2009). Some resident birds may also be more effective at locating or tracking army ant raids, in part due to specialized behaviors like bivouac checking. Migrants did not check bivouacs but could use other behavioral mechanisms for exploiting ant raids. Data on young migrant bird interactions with army ants during their first migration into the geographic range of these ants could reveal important developmental patterns, including roles O Donnell et al. • MONTANE BIRDS AT ARMY ANT SWARMS 511 ot learning and social transmission in the performance of raid-attending behavior. All of the migrants we observed at raids are known to forage on other resources while wintering (Willis 1966, 1984; Di Giacomo and Di Giacomo 2006; SO’D and AK, pers. obs.). However, even occasional reliance on army ant swarm-raids as a food source may have implica- tions for migrant bird conservation. The extent of reliance of migrants on army ants as a food source may affect fitness of birds and winter survival by migrants based on changes in army ant density. Army ant-foraging behavior and population dy- namics are strongly affected by forest clearing and forest fragmentation (Franks 1982, Meisel 2006, Kumar and O’Donnell 2009). Habitat change effects on wintering migrants could be exacerbat- ed through changes in army ant communities (Koh et al. 2004). ACKNOWLEDGMENTS Thomas Scare and Sean Tully assisted with field observations. We thank Frank Joyce and Katy Van Dusen for continuing logistical support, and the many residents of the Monteverde area who informed us of army ant swarm activity and allowed access to their private lands for research. The Tropical Science Center, Ecolodge San Luis/ University of Georgia, Monteverde Conservation League, and the Monteverde Institute extended support and access to protected lands for data collection. Research was conducted under permits from the Ministry of the Environment and Energy, Republic of Costa Rica (Scien- tific Passports 0387 and 01667), and in accordance with the laws of the Republic of Costa Rica. We thank the Organization for Tropical Studies (OTS) for ongoing research support, including assistance in obtaining permits. This project was supported by a University of Washington ALCOR Fellowship and an OTS Post-Course Award to AK, and University of Washington Royalty Research Fund and NSF grants (IBN-0347315) to SO’D. National Geographic Television also supported the field research. Research was partially supported by the National Science Foundation while SO'D was working at the foundation. Any opinions, findings, and conclusions or recommendations are those of the authors and do not necessarily reflect the views of the National Science Foundation. Two anonymous reviewers made helpful comments that improved the paper. LITERATURE CITED Brady, S. G. 2003. Evolution of the army ant syndrome: the origin and long-term evolutionary stasis of a complex of behavioral and reproductive adaptations. Proceedings of the National Academy of Sciences of the USA 100:6575-6579. Brumreld. R. T., J. G. Tello, Z. A. Cheviron. M. D. Carling, N. Crochet, and K. V. Rosenberg. 2007. Phylogenetic conservatism and antiquity of a tropical specialization: army-anl-following in the typical ant- birds (Thamnophilidac). Molecular Phylogeny and Evolution 45:1-13. Chaves, J. 2008. Benefits of cooperative food search in the maintenance of group living in Ocellated Antbirds. Dissertation. Purdue University, West Lafayette, Indiana. USA. Chaves-Campos, j. 2003. Localization of army-ant swarms by ant-following birds on the Caribbean slope of Costa Rica: following the vocalization of ant birds to find the swarms. Ornitologia Neotropical 14:289-294. Coates-Estrada, R. and A. Estrada. 1989. Avian attendance and foraging at army-ant swarms in the tropical rainforest of Los Tuxtias, Veracruz, Mexico. Journal of Tropical Ecology 5:281-292. Di Giacomo, A. S. and A. G. Di Giacomo. 2006. Observations of Strange-tailed Tyrants (Alectrurus risora) and other grassland birds following army ants and armadillos. Journal of Eield Ornithology 77:266- 268. Eranks, N. R. 1982. Ecology and population regulation in the army ant Eciton burchelli. Pages 389-395 in The ecology of a tropical forest: seasonal rhythms and long term changes (E. G. Leigh Jr., A. S. Rand, and D. M. Windsor, Editors). Smithsonian Institution Press, Washington, D.C., USA. Franks, N. R. and W. H. Bossert. 1983. The influence of swarm raiding ants on the patchiness and diversity of a tropical leaf litter ant community. Pages 151-163 in Tropical rain forest: ecology and management (S. L. Sutton, T. C. Whitmore, and C. A. Chadwick, Editors). Blackwell Scientific Publications, Oxford, United Kingdom. Garrigues, R. and R. Dean. 2007. The birds of Costa Rica. Cornell University Press, Ithaca, New York. USA. Greene, E., D. Wilcove, and M. McFarland. 1984. Observations of birds at an army ant swarm in Guerrero, Mexico. Condor 86:192-193. Guindon, C. F. 1997. The importance of forest fragments to the maintenance of regional biodiversity surround- ing a tropical montane reserve, Costa Rica. Disserta- tion. Yale University, New Haven, Connecticut, USA. Haber, W. A. 2000. Plants and vegetation. Pages 39-94 in Monteverde: ecology and conservation of a tropical cloud forest (N. Nadkarni and N. Wheelwright, Editors). Oxford University Press. New York, USA. Hardy, J. W. 1974. Jays as army ant followers. Condor 76:102-1 19. Harvey, C. 2000. Colonization of agricultural windbreaks by forest trees: effects of connectivity and remnant trees. Ecological Applications 10:1762-1773. Holdridge, L. R. 1966. The life zone system. Adansonia 6:199-203. Jahn, A. E., D. J. Levey, and K. G. Smith. 2004. Reflections across hemispheres: a system-wide ap- proach to New World bird migration. Auk 121 1005- 1013. Jankowski, J. E., A. L. Ciecka, N. Y. Meyer, and K. N. Rabenold. 2009. Beta diversity along environmental gradients: implications of habitat specialization in 512 THE WILSON JOURNAL OL ORNITHOLOGY • Voi 122, No. 3, September 2010 tropical montane landscapes. Journal of Animal Ecology 78:315-327. Koh, L. P., R. R. Dunn, N. S. Sodhi, R. K. Colwell, H. C. Proctor, and V. S. Smith. 2004. Species coextinc- tions and the biodiversity crisis. Science 305:1632- 1634. Kumar, A. and S. O’Donnell. 2007. Fragmentation and elevation effects on bird-army ant interactions in neotropical montane forest of Costa Rica. Journal of Tropical Ecology 23:581-590. Kumar, A. and S. O’Donnell. 2009. Elevation and forest clearing effects on foraging differ between surface- and subterranean-foraging army ants (Formicidae: Ecitoninae). Journal of Animal Ecology 78:91-97. Meisel, J. E. 2004. The influence of microclimate and habitat area on the ecology of the army ant Eciton burchellii in tropical forest fragments. Dissertation. University of Wisconsin, Madison, USA. Meisel, J. E. 2006. Thermal ecology of the neotropical army ant Eciton burchellii. Ecological Applications 16:913-922. Nakazawa, Y., a. T. Peterson, E. Martinez-Meyer, AND A. G. Navarro-Siguenza. 2004. Seasonal niches of Nearctic-neotropical migratory birds: implications for the evolution of migration. Auk 121:610-618. O’Donnell, S., J. Lattke, S. Powell, and M. Kasparl 2009. Species and site differences in neotropical army ant emigration behavior. Ecological Entomology 34:476-482. Otis, G. W., S. C. Santana, D. L. Crawford, and M. L. Higgins. 1986. The effect of foraging army ants on leaf-litter arthropods. Biotropica 18:56-61. Peters, M. K., S. Likare, and M. Kraemer. 2008. Effects of habitat fragmentation and degradation on flocks of African ant-following birds. Ecological Applications 18:847-858. Remsen Jr., j. V. AND D. V. Good. 1996. Misuse of data from mist-net captures to assess relative abundance in bird populations. Auk 113:381-398. Rios, M., G. Londono, and L. Biancucci. 2008. Notes on birds that follow army ants in the Northern Andes. Ornitologia Neotropical 19:137-142. Roberts, D. L., R. J. Cooper, and L. J. Petit. 2000. Use of premontane moist forest and shade coffee agroeco- systems by army ants in western Panama. Conserva- tion Biology 14:192-199. Stiles, F. G. and A. F. Skutch. 1989. A guide to the birds of Costa Rica. Cornell University Press, Ithaca, New York, USA. Swartz, M. B. 2001. Bivouac checking, a novel behavior distinguishing obligate from opportunistic species of army-ant following birds. Condor 103:629-633. Vallely, a. C. 2001 . Foraging at army ant swarms by fifty bird species in the highlands of Costa Rica. Ornitho- logia Neotropical 12:271-275. Watkins II, J. F. 1985. The identification and distribution of the army ants of the United States of America (Hymenoptera, Formicidae, Ecitoninae). Journal of the Kansas Entomological Society 58:479-502. Willis, E. O. 1966. The role of migrant birds at swarms of army ants. Living Bird 5:187-231. Willis, E. O. 1972. The behavior of Plain-brown Woodcreepers, Dendrocincla fuliginosa. Wilson Bul- letin 84:377-420. Willis, E. O. 1984. Neotropical thrushes (Turdidae) as army ant followers. Ciencia y Cultura 36:1 197-1202. Willis, E. O. and Y. Oniki. 1978. Birds and army ants. Annual Review of Ecology and Systematics 9:243- 263. Willson, S. K. 2004. Obligate army-ant-following birds: a study of ecology, spatial movement patterns, and behavior in Amazonian Peru. Ornithological Mono- graphs 55:1-67. Wrege, P. H., M. Wikelski, J. T. Mandel, T. Rasswei- LER, and 1. D. CouziN. 2005. Ant birds parasitize foraging army ants. Ecology 86:555-559. Young, B. E. and D. B. MacDonald. 2000. Birds. Pages 179-222 in Monteverde: ecology and conservation of a tropical cloud forest (N. Nadkarni and N. T. Wheelwright, Editors). Oxford University Press, New York, USA. Young, B. E., D. Derosier, and G. V. N. Powell. 1998. Diversity and conservation of understory birds in the Tilaran Mountains, Costa Rica. Auk 1 15:998-1016. The Wilson Journal of Omilhology 1 22(3): 5 1 3~5 1 7, 2010 MAROON-FRONTED PARROT {RHYNCHOPSITTA TERRISI) BREEDING HOME RANGE AND HABITAT SELECTION IN THE NORTHERN SIERRA MADRE ORIENTAL, MEXICO SONIA GABRIELA ORTIZ-MACIEL,'- CONSUELO HORI-OCHOA,' AND ERNESTO ENKERLIN-HOEELICH' ABSTRACT. The Maroon-tronted Parrot (Rhynchopsitta lerrisi) is a threatened species endemic to pine-oak (Pinus ^'P^ -Quercus spp.) forests in “sky islands” of the Sierra Madre Oriental in Mexico. We measured parrot {n = 10) home ranges during the breeding season (1999—2001) using radiotelemetry. Home ranges varied among years: 1999 = 12,379.05 ha, 2000 — 8,633.82 ha, and 2001 = 4,736.75 ha. Differences in home range size may reflect variation in distribution, abundance, and patchiness of food. Daily movements in 1999 were estimated as 23.6 km versus 13 7 km in 2000 and 26.8 km in 2001. Habitat preferences within the P was for Pinus-Abies-Pseiidotsuga forest and in 2000 and 20C influenced by nesting behavior. Rapid landscape changes ma areas for long term conservation of Maroon-fronted Parrots. Maroon-fronted Parrots {Rhynchopsitta terrisi) occupy temperate coniferous and mixed conifer- ous-deciduous forests of the Sierra Madre Orien- tal, Mexico, ranging from Nuevo Leon and Coahuila in the north to Queretaro in the south. Breeding occurs in a small area in Nuevo Leon and Coahuila where parrots nest in cavities and crevices in limestone cliffs that are often several hundred meters high. Maroon-fronted Parrots nest in aggregations and known nesting colonies have a long history of use. Nesting occurs in summer and early fall. Once breeding and nesting season ends, usually in November, parrots migrate south in the Sierra Madre Oriental (Enkerlin-Hoeflich et al. 1999). Recent counts suggest a total population of —3,500 individuals (Valdes-Pena et al. 2008). These parrots are endemic and protected under Mexican Government regulations, as well as internationally (SEMARNAP 2000, Birdlife In- ternational 2007, CITES 2007). Maroon-fronted Parrots feed mainly on pine (Pinus spp.) seeds (especially those of Pinus strobiformis, P. montezumae, P. greggii, P. cembroides, and P. culminicola), and flowers and fruits of agave (Agave gentryi) during the breeding season. They may fly long distances searching for food, and annual movement patterns vary considerably because of temporally and spatially erratic nature of pine cone crops (Enkerlin-Hoeflich et al. 1999). These parrots ' Centro de Calidad Ambiental, CEDES 5o pi.so, ITESM, Avenida Eugenio Garza-Sada 2501 sur, C. P. 64849. Monterrey, Nuevo Leon, Mexico. ^Corresponding author; e-mail: sgom@itesm.mx 'iius forest varied among nesting seasons. Preference in 1999 1 it was for Pinus forest and chaparral. This variation may be / necessitate planting and conserving pine forest to regenerate Received 3 January 2008. Accepted 26 January' 2010. eat acorns of several oak species (SGOM, pers. obs.) during winter. Most pine forests in the region are relicts and regeneration is scant or absent once forests are destroyed by natural causes or cut for use by humans. Radiotelemetry has been used as a tool in estimating parrot home range size because monitoring movements of neotropical parrots is difficult as they usually fly long distances, which allows opportunities to select among different forest habitats (Gilardi and Munn 1998). Our objectives were to use radiotelemetry to: (1) measure home range size, and (2) identify habitats used by breeding Maroon-fronted Parrots during three breeding seasons. We also estimated areal requirements for different habitats during breed- ing to effectively conserve and manage Sierra Madre Oriental forests and habitats for this important species. METHODS Our study was conducted in the northern Sierra Madre Oriental from 100° 16' W, 25° 42' N to 99° 50' W, 24° 50' N. This mountain chain is 600 km north-south and 80 km east-west. Elevation varies from 2,000 to 2,500 m above sea level with several parallel ,series of precipitous limestone ridges along a northwest to southeast direction. Climate varies with elevation, ranging from extremely dry to semi-dry and hot to subalpine climates at the highest elevations. Annual precipitation averages 400 to 900 mm (INEGI 1986). We captured 10 adult Maroon-fronted Parrots (4 in 1999, 6 in 2000) during the breeding season 513 514 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 3. September 2010 TABLE 1. Vegetation types, area, and number of visits by Maroon-fronted Parrots during the nesting seasons 1999, 2000, and 2001 in the Sierra Madre Oriental, Mexico. Area used Vegetation type 1999 2000 2001 Pinus- Abies-Pseudotsuga 4,739 ha in = -- 118) 29,350 ha in = = 28) 11,376 ha in = = 32) Pinus 21,180 ha in = = 27) 3,504 ha in -- = 37) 1,533 ha in -- = 53) Pinus cenibroides 2,571 ha in = = 20) 2,849 ha in = = 3) 744 ha in = = 3) Pinus-Quercus 744 ha in = = 3) 9,986 ha in = = 0) 4,581 ha in = = 0) Chaparral 9,666 ha in = = 14) 1,957 ha in = = 213) 660 ha in = = 53) Totals 38,900 ha in -- = 182) 47,645 ha in = = 281) 18,894 ha in = = 141) with mist nets at the El Taray Sanctuary in Coahuila and fitted them with telemetry neck- collars (Holohil Systems Ltd., Carp, ON, Canada). Radios weighed either 20 g (Model AI-2C, n = 9; loop antennas) or 13 g (Model SI-2C, n = 1; whip antenna) and represented 4.3 and 2.8% of average bird weight, respectively. Both types of transmit- ters had 24 months of expected battery life. The average (± SD) distance for signal detection was 3.6 ± 1.3 km {n = 181; Ortiz-Maciel 2000). These types of radios had no apparent negative impact on parrots (Snyder et al. 1994, Meyers 1996, Salinas-Melgoza and Renton 2005). We noted no apparent detrimental effects of transmit- ter attachment during our study, and some birds with non-functioning transmitters (dead batteries) were observed at the nesting cliffs in 2007. We attempted to locate and follow radio- marked birds using 3-element yagi antennas and TRX-2000 receivers (Wildlife Materials Inc., Carbondale, IL, USA) to measure home ranges. We searched for parrots in high-elevation pine forests along valley roads during their breeding season, and also at nesting colonies and clay licks. We radio-located birds from dawn to sunset for 39 days from August to November 1999, 47 days from July to November 2000, and 43 days from May to October 2001. We established a bird’s position by recording the direction of the maximum signal, taking a bearing with a compass, and using a Global Positioning System (GPS) unit to ascertain the observer’s location. Only one observer took bearings because of equipment limitations. We moved along the road >3 km and took another bearing to more precisely identity the location of a bird. We did not observe radio-marked birds directly at times, but usually heard vocalizations of parrot Hocks, because daily tracking started at a roosting site. We also located parrots at nesting cliffs in their cavities. All bird locations were recorded on 1:50,000 scale maps of the Sierra Madre Oriental (McFarland 1991, Ortiz-Maciel 2000). Transmitter signals were not detected during all searches because pan'ots may have been out of the receiver’s range and behind ridges that blocked the transmitter’s signal. Site Characterization. — We plotted all loca- tions on a landscape image (from Landsat-TM 1993) and processed it with Arcview GIS 3.2 (ESRl 1999). Forest habitats, characterized fol- lowing Hori-Ochoa (1998), were: Pinits-Abies- Pseudotsuga, Piniis, Pinus cemhroides, Pinus- Quercus, Querciis-Pimis, and Quercus. Other habitats included submontane shrub, microphyl- lous desert shrub, desert shrub, chaparral, grass- land, and agricultural zones. We measured breeding home range size using the 50% kernel method for each field season (Worton 1989), and processed data using the Animal Movement extension from Arcview GIS 3.2 (ESRI 1999). We followed some parrots for an entire day and calculated total distances traveled by adding the distances between each location. Habitat selection was estimated for each breeding season with a Chi-square independence test by counting the number of visits to each vegetation type and comparing the areal propor- tions for the 95% minimum convex polygon (White and Gairott 1990) (Table 1). We calculated Bonferroni confidence intervals to identify use by vegetation types (Neu et al. 1974, Byers and Steinhorst 1984, White and Garrott 1990). Confidence intervals (95%) were calculated using only habitats where birds were observed including Piuus-Ahies-Pseiidotsuga, Pi- nu.s, Pinu.s cemhroide.s, and Piuus-Quercits forest and chaparral. Values are presented as means Ortiz- Made! et al. • M A[^OON-FRONTED PARROT HOME RANGES 515 TABLE 2. Pioportioii ol Maroon-tronted Parrol locations by habitat type, habitat available, and Bonferroni confidence intervals tor the 1999. 2()()(), and 2001 breeding seasons, Sierra Madre Oriental, Mexico Year and habitat type Pmportion of parrot locations (P,,/,,) Proportion of habitat available (P„,,) Bonferroni confidence iniervals 1999 Pinus- Ahie.'i-P.seudot.uigu forest' 0.65 0.12 0.56 < Pob. ^ 0.74 Pinus forest 0.15 0.54 0.08 < P,,!,, ^ 0.22 Pinus cenibroides forest 0.1 1 0.07 0.05 < Pnhs S 0.17 Pinus-Quercus forest 0.02 0.02 -0,01 < Poh'^ ^ 0.04 ChapaiTal 0.08 0.25 0.03 < Pnbs ^ 0.13 2000 Pinus-Ahies-Pseudotsuga forest 0.10 0.62 0.05 < Pohs ^ 0.15 Pinus forest" 0.13 0.07 0.08 < Poh. ^ 0,18 Pinus cenibroides forest 0.01 0.06 -0.01 < A.Av ^ 0.03 Pinus-Quercus forest 0.00 0.21 0.00 < Poh, ^ 0.00 Chaparral" 0.76 0.04 0.69 < 0.82 2001 Pinus-Ahies-Pseudotsuga forest 0.23 0.60 0.14 < Pob, ^ 0.32 Pinus forest" 0.38 0.08 0.27 < Pob, ^ 0.48 Pinus cenibroides forest 0.02 0.04 -0.01 < Pobs - 0.05 Pinus-Quercus forest 0.00 0.24 0.00 < Pobs ^ 0.00 Chaparral" 0.38 0.04 0.27 < Pobs ^ 0.48 ■“ Preferred habitat. (± SD) with P < 0.5. The null hypothesis for each bird was that habitat types were used in accordance with their availability. RESULTS Home Range Size. — Home range size obtained with 50% kernel methods varied among breeding seasons. Home range size in 1999 was 12,379.05 ha [n = 3 birds), 8,633.82 ha [n = 7 birds) in 2000, and 4,736.75 ha (/? = 4 birds) in 2001. Daily Movements. — We ob.served two parrots in 1999 move a mean (± SD) distance of 23.6 ± 5.7 km per day (« = 15 days). The mean distance moved per day for three birds was 13.66 ± 7.64 km per day (n = 6 days) in 2000; the mean distance was 26.82 ± 18.17 km (n = 5 days) for two parrots in 2001. We also located a frequently- used communal night roost in Sierra La Viga. Habitat Selection. — Maroon-fronted Parrots demonstrated preference for specific habitat types within the pine forest mosaic ba.sed on frequency of locations within 95% minimum convex poly- gon home ranges. There was a difference between habitat used and areal abundance (breeding seasons: 1999, P < 0.001, X-4 = 493.5; 2000, P < 0.001, = 3,721.9; 2001 P < 0.001, = 688.4). Bonferroni confidence intervals indicated some of the six forests habitats were used more than expected. Pinus-Ahies-Pseudotsuga forest was used most in 1999, whereas Pinus forest and chaparral were used most in 2000 and 2001 (Table 2). DISCUSSION Most Maroon-fronted Parrot locations were at the highest elevations of the mountains, where subfreezing temperatures persist, indicating they are similar to Thick-billed Parrots (Rhvnchopsitta pachyrhyncha) and are temperate in their habitat preferences (Snyder et al. 1994, 1999). Breeding activity of Maroon-fronted Pan'ots started nor- mally in 1999, but no young fledged and all nests failed by mid-September. Large home ranges that year reflected food scarcity similar to that reported for Thick-billed PaiTOts where breeding success and effort appeared tightly linked to food availability (Enkerlin-Hoeflich et al. 1999). Most used habitats were in high elevation Pinits-Ahies- Pseudot.suga forest generally considered to be less suitable for foraging because the small reward ratio of small seeds characteristic of tree species in these forests (Enkerlin-HoeHich et al. 1999). Average home ranges of breeding Maroon- fronted PaiTots were larger than those of other psittacines; juvenile Puerto Rican Amazon {Ama- 516 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 3, September 2010 zona vittata) have estimated home ranges of 1,372 ha (Lindsey et al. 1991), Ground Parrots {Pezoponis wallicus) of Australia and Tasmania have estimated home ranges of 9.2 ha (McFarland 1991), and juvenile Lilac-crowned Amazon (Amazona finschi) have home ranges of 4,674 ha (Salinas-Melgoza 2003). These home range dif- ferences may reflect that Maroon-fronted Parrots now inhabit disturbed areas and require larger areas to satisfy their basic needs. Continental parrot species have larger home ranges, compared with island species, because of more availability of habitats assuring refuge during periods of stress (Snyder et al. 1987). Maroon-fronted Parrots did not use all forest types equally; all of their activities were in five habitats (4 forest types and chaparral). Pinus- Ahies-Pseudotsiigci forest was used more than expected in 1999, whereas Pinus forest and chaparral were more used than expected in 2000 and 2001. The El Taray nesting cliff is in chaparral habitat which may influence apparent habitat preferences. The dependence of Maroon-fronted Parrots on coniferous forests increases their vulnerability with respect to habitat changes resulting from timber harvest and fires as noted >25 years ago (Lawson and Lanning 1981, Enkerlin-Hoeflich et al. 1999). These temperate forests support different vegeta- tion types and regeneration, management, or restoration varies depending on each plant com- munity. For example, after a fire, chaparral stands of 20-30 years in age have nearly com- plete shrub cover with few to no emergent trees (Brown et al. 2000). Thus, the overall productivity of habitats for parrots is decreasing given the slow or absent forest regeneration and replacement by oak chaparral with little food value for parrots. Spatial and temporal patchiness of food plants may be increasing which may lead to food- induced stress on the population. Maroon-fronted Parrots also have local pressure for one of their preferred food sources, pinion pine seeds, which are valuable at the local market and extensively harvested as an income supplement by rural families. Recent fires have devastated several important sites, including foraging and breeding areas; >90% of standing trees died in 1998 in an area of 5,000 ha (Brown et al. 2000). Another fire in 2006 destroyed 545 ha of standing trees (CONAFOR 2006a, b) including >70% of the El Taray Sanctuary (Manzano-Camarillo 2006). These habitat changes have directly affected Maroon-fronted Parrot nesting activities. Regional forest managers should conserve and re-establish Pinus forest communities. Manage- ment should include control of timber harvests, fires, reforestation, and restoration (Valdez-Ta- mez 1981). Eorest management should be imple- mented to manage and conserve critical habitats, increase size of forested areas, and provide alternative livelihood opportunities for local inhabitants via ecotourism or provision of pay- ment and incentives for ecosystem services and watershed protection. ACKNOWLEDGMENTS This research was funded by ITESM Campus Monterrey and Centro de Calidad Ambiental, E. Alexander Bergstrom Awards Program, American Bird Conservancy’s William Belton Grants Program, and the World Parrot Trust. We appreciate the suggestions of Noel Snyder on this manuscript and the knowledge he shared with us on Maroon-fronted Parrots. We also thank Museo de las Aves de Mexico for use of facilities at the El Taray Sanctuary; Fabian Lozano for support with the Geographic Information System Laboratory; Patricia Vela, Santiago Salazar, and Fabiola Yepez for helping with image processing; Roberto Mercado for help with statistics; and Javier Cruz, Jose Manzano, Julio Ronquillo, Romualdo Martinez, and Claudia Macias for field support. LITERATURE CITED Birdlife International. 2007. Maroon-fronted Parrot. BirdLife species factsheet. www.birdlife.org/datazone/ .species/index. html. Brown, S., M. Delaney, and D. Shoch. 2000. Assessing the potential for carbon sequestration at three forest fire restoration sites in Mexico. Environmental Policy and Institutional Strengthening Indefinite Quantity Contract Number PCE-I-00-96-00002-00. Prepared by Winrock International team for U.S. Agency for International Development Mission to Mexico, Little Rock, Arkansas, USA. Byers, C. R. and R. K. Steinhorst. 1984. Clarification on a technique for analysis of utilization availability data. Journal of Wildlife Management 48: 1050-1052. CITES. 2007. Convention on International Trade in Endangered Species of Wild Fauna and Flora. Appendices I, II, and III. www.cites.org/eng/app/ appendices. shtml. CONAFOR. 2006a. Comision Nacional Forestal. conafor. gob.mx/portal/docs/secciones/comunicaciQn/B-052006. pdf CONAFOR. 2006b. Comision Nacional Forestal. conafor. gob.mx/portal/docs/seccione.s/comunicacion/B- 1 22006. pdf Enkeri.in-Hoefijch, E. C., C. Macias-Caballero, T. Monterrubio-Rico, M. A. Cruz-Nieto, N. F. R. Snyder, D. Venegas-Holguin, J. Cruz-Nieto, G. Ortiz-Mciciel Cl al. • MAROON-FRONTED PARROT HOME RANGES 517 Ortiz-Maciel, J. Gonzalez-Elizondo, and E. Stone. 1999. Statu.s, di.stribucion, ecologia y con- servacion dc la.s cotorras serranas (Rhynchopsitta pachyrhyiiclui y R. terrisi) en el Norte de Mexico. Final report. In.stituto Tecnologico y de Estudios Siiperiores de Monterrey. Centro de Calidad Ambien- tal. Programa de Manejo Sostenible de Ecosistemas, Monterrey, Nuevo Leon, Mexico. Environmental Systems Research Institute (ESRI). 1999. Arcview GIS 3.2. Environmental Systems Research Institute Inc., Redlands, California, USA. Gilardi, J. D. and C. A. Munn. 1998. Patterns of activity, flocking and habitat use in parrots of the Peruvian Amazon. Condor 100:641-653. Hori-Ochoa. M. C. 1998. Variabilidad de las respuestas espectrales de acuerdo a los cambios fenologicos de la vegetacion del anticlinorio de Arteaga, Coahuila y Nuevo Leon, Mexico. Thesis. Institute Tecnologico y de Estudios Superiores de Monterrey, Nuevo Leon, Mexico. INSTITUTO NaCIONAL DE ESTADISTICA GeOGRAFIA E IN- formatica (INEGI). 1986. Sintesis Geografica del Estado de Nuevo Leon, Mexico. Monterrey, Nuevo Leon, Mexico. Lawson, P. W. and D. V. Panning. 1981. Nesting and status of the Maroon-fronted Parrot (Rhynchopsitta terrisi). Pages 385-392 in Conservation of New World parrots (R. F. Pasquier, Editor). Parrot Working Group Meeting, St. Lucia. Smithsonian Institution Press, Washington, D.C., USA. Lindsey, G. D., W. J. Arendt, J. Kalina, and G. W. Pendleton. 1991. Home range and movements of juvenile Puerto Rican Parrots. Journal of Wildlife Management 55:318-322. Manzano-Camarillo, M. 2006. Diagnostico de dafios por incendios forestales del predio El Taray, municipio de Arteaga, Coahuila. Technical report presented to Comision Nacional de Biodiversidad. Institute Tecno- logico y de Estudios Superiores de Monterrey. Centro de Calidad Ambiental. Programa de Recursos Fore- stales y Zonas Aridas, Monterrey, Nuevo Leon, Mexico. McFarland, D. C. 1991. The biology of the Ground Parrot, Pezoporus wallicus, in Queensland. I. Micro- habitat use, activity cycle and diet. Wildlife Research 18:169-184. Meyers, J. M. 1996. Evaluation of 3 radio transmitters and collar designs for Amazona. Wildlife Society Bulletin 24:15-20. Nfiu, C. W., C. R. Byers, and J. M. Peek. 1974. A technique for analysis of utilization availability data. Journal of Wildlife Management 38: 541-545. Ortiz-Maciel, S. G. 2000. Uso del paisaje por la cotorra serrana oriental (Rhynchopsitta terrisi). Thesis. In- stitLito Tecnologico y de Estudios Superiores de Monterrey, Nuevo Leon, Mexico. Salinas-Melgoza, A. 2003. Dinamica espacio-temporal de individuos Juveniles del loro corona lila (Amazona finschi) en el bosque seco de la costa de Jalisco. Thesis. Universidad Nacional Autonoma de Mexico, Mexico, D.F. Salinas-Melgoza, A. and K. Renton. 2005. Seasonal variation in activity patterns of juvenile Lilac-crowned Parrots in tropical dry forest. Wilson Bulletin 117:291-295. SEMARNAP. 2000. Secretaria de Medio Ambiente, Recursos Naturales y Pesca. Proyecto para la Con- servacion, Manejo y Aprovechamiento Sustentable de los Psitacidos en Mexico. SEMARNAP, Mexico, D.F. Snyder, N. F. R., J. W. Wiley, and C. B. Kepler. 1987. The parrots of Luquillo: natural history and conserva- tion of the Puerto Rican Parrot. Western Foundation of Vertebrate Zoology, Los Angeles, California, USA. Snyder, N. F. R., E. C. Enkerlin-Hoeflich, and M. A. Cruz-Nieto. 1999. Thick-billed Parrot (Rhynchopsitta pachyrhyncha). The birds of North America. Number 406. Snyder, N. F. R., S. E. Koening, J. Koschmann, H. A. Snyder, and T. B. Johnson. 1994. Thick-billed Parrot releases in Arizona. Condor 96:845-862. Valdes-Pena, R. a., S. G. Ortiz-Maciel, S. O. Valdez- Juarez, E. C. Enkerlin-Hoeflich, and N. F. R. Snyder. 2008. Use of clay licks by Maroon-fronted Parrots (Rhynchopsitta terrisi) in northern Mexico. Wilson Journal of Ornithology 120:176-180. Valdez-Tamez, V. 1981. Contribucion al conocimiento de los tipos de vegetacion, su cartografia y notas floristico-ecologicas del municipio de Santiago, Nuevo Leon, Mexico. Thesis. Universidad Autonoma de Nuevo Leon, Monterrey, Nuevo Leon, Mexico. White, G. C. and R. A. Garrott. 1990. Analysis of wildlife radio-tracking data. Academic Press, London, United Kingdom. WORTON, B. J. 1989. Kernel methods for estimating the utilization distribution in home-range studies. Ecology 70:164-168. The Wilson Journal of Ornithology 122(3):5 18-531 , 2010 SHORT-TERM EFFECTS OF FIRE ON BREEDING BIRDS IN SOUTHERN APPALACHIAN UPLAND FORESTS NATHAN A. KLAUS,' SCOTT A. RUSH,"-^ TIM S. KEYES,' JOHN PETRICK.^ AND ROBERT J. COOPER- ABSTRACT. — We investigated how variation in fire severity (control or no fire; low, medium, and high severity fires) and interval (1-2 years vs. 3-6 years after fires) affected habitat and avian abundance, species diversity, richness, and evenness in the southern Appalachian Mountains. Fire severity and interval had significant implications for both habitat and avian communities. Species richness within 2 years of fires was on average 26% higher in areas receiving medium and high severity treatments than in unburned control units. Species diversity and species richness were markedly greater 3-6 years after fires within high severity treatments (12 and 44%, respectively), compared to unbumed controls. Relative abundance and species evenness did not vary with fire severity or time since fire. The short-term effects of low severity fires, or high severity fires with short rotation periods (^2 years) may have limited positive effects on avian communities. Facilitation of disturbance regimes including mid to high severity fires, which foster uneven-aged forests, can be an effective conservation tool for restoring avian communities. Received 30 June 2009. Accepted 1 March 2010. Disturbance is a fundamental ecological pro- cess that has had a role in structuring and maintaining diversity within many ecological communities (Loucks 1970, Elliott et al. 1999, Brawn et al. 2001, Bond and Keeley 2005, Bond et al. 2005). Generally, disturbance increases diversity by creating a mosaic of habitats or SLiccessional stages within a landscape (Askins 2001, Brawn et al. 2001). Fire is one form of disturbance that has had a large part in structuring the ecological communities of the southern Appalachian Mountains (Lorimer 1980, Van Lear and Waldrop 1989, Waldrop et al. 1992). Suppression of fire on public lands during the past century has likely reduced diversity and structure of animal and plant communities (Abrams 1992, Lorimer 2001, Artman et al. 2005). Several federal land management agencies, in an effort to restore diversity and ecological function, now plan for increased use of prescribed fire in the southern Appalachian Mountains. However, effects of restoring fire on wildlife are largely unknown in these mostly hardwood forest systems. In particular, effects of varying fire .severity and interval on wildlife are largely unknown. Many bird species associated with early succession habitats and pine (Pinus spp.) savannas ' Nongame Con.servation Section, Georgia Department of Natural Resources, 116 Rum Creek Drive, Forsyth, GA 31029. USA. ^ D. B. Warnell School of Fore.st Resources, University of Georgia, Athens, GA 30602, USA. ’ Chattahoochee-Oconee National Forest, 1755 Cleveland Highway, Gainesville. GA 30501, USA. ■‘Corresponding author; e-mail; srush@uwindsor.ca would likely benefit from return of fire to the southern Appalachian Mountains. Species of concern include: Bachman’s Sparrow {Aimophila aestivalis). Northern Bobwhite {Colinus virginia- nus), Red-cockaded Woodpecker {Picoides bor- ealis), Prairie Warbler (Dendroica discolor), and Golden-winged Warbler {Vermivora chrysoptera) (Brewster 1886, Burleigh 1958, Stupka 1963, Hunter et al. 2001, Klaus 2004). Measured responses for other bird species may not be positive (Lang et al. 2002, Artman and Down- hower 2003, Tomcho et al. 2006), and the benefits of fire restoration may not integrate similarly across all members of avian communities, or in all habitats (Artman et al. 2001, Saab and Powell 2005, Tomcho et al. 2006, Greenberg et al. 2007). Thus, reintroduction of fire should be based on desired ecological condition. The historical suppression of fire and changes in agricultural practices within the eastern United States have been implicated in loss of early successional habitat (Askins 2001). Loss of this habitat type has had a significant impact on avian communities. Early succession song- birds in the southern Appalachians have the .strongest declines of any group of birds (Hunter et al. 1999, Sauer et al. 2005). Many of the species that require mature forest for nesting also use early succession habitat as fledglings and during molt (Anders et al. 1998, Vega Rivera et al. 1999, Marshall et al. 2003, Rush and Stutch- bury 2008). Restoration of early successional habitat may be a key feature of policies directed at conservation of many avian species (Dettmers 2003, Bulluck and Buehler 2006, Buehler et al. 2007). 518 Klaus et at. • FIRE AND BIRDS IN THE SOUTHERN APPALACHIAN MOUNTAINS 519 Prescribed fire is often cited as a likely tool for restoring and providing suitable habitat for many avian species (Partners in Flight Working Group 2002), but our current understanding of the appropriate application of fire in shaping forest ecosystems of the southern Appalachian Moun- tains remains limited. We currently know of only a few studies that have examined the effects of prescribed fire on songbirds in hardwood forests of the eastern United States (Aquilani et al. 2000; Artman et al. 2001, 2005; Greenberg et al. 2007); only one of these studies occurred in the southern Appalachians. Our objectives were to document how breeding season distributions of birds changed with variation in fire severity and time since fire in the southern Appalachians. METHODS Study Area. — This study was conducted within the Brasstown and Tallulah ranger districts of the Chattahoochee-Oconee National Forest (CNF), Murray County, Georgia, USA. U.S. Forest Service (USFS) personnel were consulted within the CNF and 22 potential study sites were identified at elevations from 600 to 1,500 m. All bum units were >200 ha and had burned in the last 8 years, either as a prescribed fire or as part of a wildfire. Vegetation sampling sites were placed into one of three fire severity groups on the basis of the extent of canopy mortality: low, medium, or high. Low severity burns were typical of most ‘fuel reduction’ prescribed fires in the southern Appalachians. Flame lengths were generally less than 1 m, resulting in <5% canopy mortality throughout the burn unit and low to moderate midstory mortality. Moderate to severe midstory and 5 to 20% canopy mortality characterized medium severity burns and flame lengths ranged from <0.5 m in mesic sites to 4 m along dry ridge tops, leaving a heterogeneous result. High severity bums were characterized by 20 to 50% canopy mortality, mostly along ridge tops, and severe to total midstory mortality except in ravines. We selected 12 of the 22 burn units, four of each treatment, based on similarity of forest type, elevation, and average aspect. All chosen burn units had been burned <6 years earlier. Half the 12 burn units selected (2 of each treatment) were in sites that were burned <2 years previously, and two were in sites that were burned 3-6 years previously. Forest types consisted primarily of mixed pine (Finns spp.)-hardwoods and drier oak- hickory (Qiiercus-Carya), although ridge tops occasionally included small stands of F. taeda, F. virginiana, F. rigida, and F. pungens. Four control units were also identified with the aid of USFS personnel. Control units were similar in forest type, elevation, and average aspect but had not burned in >20 years. Bird Surveys. — Surveys were conducted by USFS and Georgia Department of Natural Re- sources staff. Participants received several weeks of spring training on bird identification and point- count techniques prior to surveys. All participants had conducted bird surveys for several years prior to this study. Two transects were placed for each survey within each burn unit. Transects followed contours and were at least 200 m apart. All surveys were conducted between 0630 and 1100 hrs, 17 May-10 June 2004, following protocols established by Hamel et al. (1996), except that counts were conducted for 10 min and distance bands of 0-10, 11-25, 26-50, 51-100, and >100 m were used. Ten point counts were conducted on each survey transect, 20 per burn unit. Locations of individual point counts were: (1) separated by at least 200 m, (2) at least 100 m inside the burn unit, and (3) placed in upland mixed pine hardwood stands or oak stands. Ravines, road edges, and other cover types were avoided. Four replications of each treatment were surveyed for a total of 80 points per treatment. Habitat Measurements. — We conducted vari- able-radius vegetation surveys within each burn unit centered on each point count location. We measured the basal area using a 10-factor prism of live trees >25 cm diameter breast height (DBH); this metric is referred to as BASAL AREA. We visually estimated average canopy (CANOPY COVER) and herbaceous cover (HERBACEOUS COVER) to the nearest 10% for a 25-m radius around each plot center. We also measured shrubs (woody plants 1 to 7 m tall and <12 cm DBH) within a 3-m radius around the plot center by counting all stems (SHRUB DENSITY) and measuring average shrub height (SHRUB HEIGHT) to the nearest decimeter. Each burn unit was categorized based on known fire histories into one of three fire histories: (1) control [not burned within the last >20 years], (2) burned within 1-2 years of surveys, and (3) burned within 3-6 years of surveys. Data Analy.ds.-—V^e restricted our analysis to limit the repeated counting of the same individual at two adjacent survey sites and to include only those birds detected within 100 m of each survey 520 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 3, September 2010 point. The relative abundance of each species for each treatment was calculated as the average number of individuals detected/ha. We restricted each point-count survey to counts of individuals detected per distance band during each time interval for conservative estimates of relative abundance of each species. Detections of the same species in the same distance band but in different time intervals were treated as a possible recount of the same individual and were excluded from further analysis. All bird species observed at each survey point along each transect were used in calculations of bird species diversity, species richness, and species evenness. Species diversity was calculated as the Shannon-Weaver Diversity Index (//') (Shannon and Weaver 1963). Species richness was calculated as the number of individ- ual species observed per point during each point count, and species evenness was calculated as H' divided by the natural log of species richness (Magurran 1988). We used linear, mixed-effects models (Ime) in R (Version 2.7.1; R Development Core Team 2008) to compare habitat measurements, diversi- ty, species evenness, and species richness between treatments using the burn unit as the experimental unit. Mixed-effects models are appropriate for this data structure because they partition information into multi-level components. The data recom- mends level- 1 (survey point), level-2 (survey transect), and level-3 (burn unit) components in our case. Each component is estimated with the appropriate degrees of freedom, and the standard errors of other parameter estimates are appropri- ately adjusted (Raudenbush and Bryk 2002). Transect and burn unit effects were accounted for as random variables, and treatment as a fixed effect. We used q-q plots prior to analysis to examine the normality of each habitat metric. We log transformed BASAL AREA, SHRUB DEN- SITY, SHRUB HEIGHT, and HERBACEOUS COVER and arcsine square-root transformed CANOPY COVER to improve normality and homoscedasticity. We used conditional t-tests for each Ime to ascertain whether covariates were significantly different from zero (Pinheiro and Bates 2000). We also developed a series of orthogonal contrasts to compare means between the different levels of burn severity and times since fire. We used canonical correspondence analysis (CCA) of data collected at each of the survey points to examine the relationship between bird community structure and measured environmental variables (McCune and Grace 2002). This tech- nique was used to produce graphical presentations depicting relationships among abundance of individual species, treatments, and measured environmental gradients. Habitat variables ex- plaining a significant amount of variation (P ^ 0.10), as calculated by Monte Carlo permutation tests (1,000 random permutations of samples in the species data; package Vegan in R) (R Development Core Team 2008), were included in the CCA analyses. The means of these variables are represented by the origin in the resulting diagram. We constrained our analysis for clarity in presentation of bird assemblages to include only those species that were detected at >20% of all survey points. We conducted two separate analyses to facilitate comparisons, one comparing habitat and bird assemblage between control and treatments 1-2 years after fires, and a separate analysis with similar comparisons for control and treatments sampled 3-6 years after fires. We log transformed BASAL AREA, SHRUB DENSITY, SHRUB HEIGHT, and HERBACEOUS COVER and arcsine square-root transformed CANOPY COVER prior to using them in the CCA. RESULTS Fire and Habitat. — Fire severity affected hab- itat structure. Comparisons of treatments and unburned controls revealed CANOPY COVER and SHRUB DENSITY differed among treat- ments (^6.9 ~ 19.1, P < 0.001; F6 9 = 4.0, P = 0.03, respectively), (Fig. 1). High severity treat- ments had significantly less CANOPY COVER than control units both 1-2 years it = 6.8, df = 9, P < 0.001), and 3-6 years after fires (t = 7.9, df = 9, P < 0.001). CANOPY COVER did not differ relative to the number of years since fires it = 1.0, df = 9, P = 0.33) but, SHRUB DENSITY was significantly lower among high severity treatments 1-2 years after fires than 3- 6 years after fires {t = —4.4, df = 9, P = 0.002). Ejfect.s on Avian Diversity and Communities. — Sixty-three species were detected among the four treatments (Table 1 ). Species richness differed among several treatments (Pf,.9 = 4.9, P = 0.02) and was significantly higher 1-2 years after fires relative to the low severity burn units. It did not differ among medium and high severity burn units (t = -2.5, df = 9, P = 0.03, t = -2.3, df = 9, P = 0.05) (Fig. 2). Species richness 3-6 years Herbaceous cover (%) Slmib stems/m^ Canopy cover (%) Basal ai'ea (mMia) K/dns el al. • FIRE AND BIRDS IN THE SOUTHERN APf^ALACHIAN MOUNTAINS 521 o o o 0-1 00 C>' ■o O' O' O' O -I D Conti'ol I I Low Medium Higli I 1 to 2 year's I Low Medium Higli * 3 to 6 years Fire intensity/year's since fire FIG. I. Box and whisker plots of habitat characteristics measured at point-count locations within control and burn treatments. Whiskers represent maximum and minimum observations while boxes repre.sent the 25 and 75% quartiles. 522 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 3, September 201U TABLE 1. Relative abundance of species detected during surveys in the southern Appalachian Mountains. Estimates are individuals/ lOha (.v ± SE). Missing values indicate species not detected. Species Scientific name Ruffed Grouse Bonasa umbellus Wild Turkey Meleagris gallopavo Black Vulture Coragyps atratus Turkey Vulture Cathartes aura Sharp-shinned Hawk Accipiter striatus Broad-winged Hawk Buteo platyptenis Red-tailed Hawk B. jamaicensis Mourning Dove Zenaida macroura Yellow-billed Cuckoo Coccyzus americanus Barred Owl Strix varia Chimney Swift Chaetiira pelagica Ruby-throated Hummingbird Archilochus coliihris Red-bellied Woodpecker Melanerpes carolinus Downy Woodpecker Picoides puhescens Hairy Woodpecker P. villosiis Northern Flicker Colaptes aiiratus Pileated Woodpecker Dryocopus pileatus Eastern Wood-Pewee Contopus virens Eastern Phoebe Sayornis phoehe Great Crested Flycatcher Myiarchiis crinitus Treatment (years after fires) Low Medium High Control 1-2 3-6 1-2 3-6 1-2 3-6 0.1 ± 0.1 0.2 ±0.1 0.1 ± 0.1 0.2 ± 0.2 0.2 ± 0.2 0.3 ± 0.2 0.2 ± 0.2 0.3 ± 0.2 0.2 ± 0.2 0.1 ± 0.1 0.1 ±0.1 0.2 ± 0.1 0.1 ± 0.1 0.3 ± 0.2 0.3 ± 0.2 0.2 ± 0.1 0.6 ± 0.2 0.7 ± 0.2 1.0 ± 0.3 0.3 ± 0.2 0.2 ± 0.1 0.5 ± 0.2 0.3 ± 0.2 0.5 ± 0.2 0.1 ± 0.1 0.1 ± 0.1 0.3 ± 0.2 0.6 ± 0.2 0.2 ± 0.2 1.2 ± 0.4 0.6 ± 0.3 0.1 ± 0.1 0.1 ± 0.1 0.1 ± 0.1 0.1 ±0.1 0.1 ± 0.1 0.2 ± 0.1 0.2 ± 0.1 0.1 ± 0.1 0.1 ± 0.1 0.2 ± 0.1 O.I ±0.1 0.2 ± 0.1 0.6 ± 0.2 0.1 ± 0.1 1 .2 ± 0.3 0.6 ± 0.2 0.2 ± 0.1 0.1 ±0.1 0.1 ± 0.1 0.7 ± 0.3 0.4 ± 0.2 0.9 ± 0.3 1.6 ± 0.3 1.1 ± 0.2 1.5 ± 0.4 1.6 ± 0.3 1 .3 ± 0.3 1.5 ± 0.3 0.7 ± 0.2 0.2 ± 0.2 1 .4 ± 0.3 2.4 ± 0.4 0.1 ± 0.1 0. 1 ± 0. 1 0.2 ± 0.1 0.1 ± 0.1 0. 1 ±0.1 0.1 ± 0.1 0.2 ± 0.1 0.2 ± 0.1 0.5 ± 0.2 0.6 ± 0.2 Klaus et a!. • FIRE AND BIRDS IN THE SOUTHERN APPALACHIAN MOUNTAINS 523 TABLE 1. Continued. Treatment (years after fires) Species Low Medium High Scientific name Control 1-2 .3-6 1-2 3-6 1-2 3-6 Yellow-throated Vireo 0.2 ± 0.1 0.4 ± 0.2 Vireo flavifrons Blue-headed Vireo 0.6 ± 0.2 0. 1 ± 0. 1 1 .0 ± 0.3 0.7 ± 0.2 0.7 ± 0.2 1.3 ± 0.3 0.7 ± 0.2 V. solitariiis Red-eyed Vireo 5.6 ± 0.4 4.0 ± 0.5 4.4 ± 0.5 4.0 ± 0.5 5.8 ± 0.5 3.5 ± 0.6 5.4 ± 0.5 V. olivaceus Blue Jay 0.9 ± 0.3 0.6 ± 0.2 0.9 ± 0.3 0.9 ± 0.3 0.6 ± 0.2 0.6 ± 0.2 0.8 ± 0.3 Cyanocitta cristata American Crow 1.5 ± 0.3 4.6 ± 0.6 3.7 ± 0.4 1.3 ± 0.3 1.4 ± 0.3 1.3 ± 0.3 2.7 ± 0.4 Con’us hrachyrhynchos Carolina Chickadee 0.3 ± 0.2 0.5 ± 0.2 0.2 ± 0.1 0.7 ± 0.3 0.1 ± 0.1 0.4 ± 0.2 Poecile carolinensis Tufted Titmouse 1.0 ± 0.2 1.4 ± 0.2 2.5 ± 0.4 1.7 ± 0.3 1.6 ± 0.3 0.6 ± 0.3 2.1 ± 0.3 Baeolophus bicolor White-breasted Nuthatch 0.5 ± 0.2 0.2 ± 0.1 0.4 ± 0.2 0.4 ± 0.2 0.5 ± 0.3 0.6 ± 0.3 0.5 ± 0.2 SiUa carolinensis Carolina Wren 0.2 ± 0.1 0.3 ± 0.2 0.2 ± 0.1 0.6 ± 0.2 0.4 ± 0.2 1.3 ± 0.3 1.4 ± 0.3 Thryothorus kidovicianus Blue-gray Gnatcatcher 0.1 ± 0.1 0.2 ± 0.1 0.1 ± 0.1 0.6 ± 0.3 Polioptila caerulea Wood Thrush 0.1 ± 0.1 0.6 ± 0.3 0.2 ± 0.1 0.2 ± 0.1 0.6 ± 0.3 0.4 ± 0.2 0.3 ± 0.2 Hylocichla mustelina Brown Thrasher 0.1 ± 0.1 0.1 ± 0.1 0.1 ± 0.1 Toxostoma rufum Cedar Waxwing 0.1 ± 0.1 0.9 ± 0.5 0.2 ± 0.2 Bombycilla cedrorum Golden-winged Warbler 0.2 ± 0.2 Vennivora chrysoptera Northern Parula 4.0 ± 0.4 0.1 ± 0.1 Parula americana Chestnut-sided Warbler 0.4 ± 0.2 0.2 ± 0.2 1 .4 ± 0.4 0.8 ± 0.2 1 .5 ± 0.4 6.6 ± 0.5 Dendroica pensylvcmica Black-throated Blue Warbler 0.6 ± 0.3 1 .0 ± 0.3 D. caendescens Black-throated Green Warbler 1 . 1 ± 0.3 1.2 ± 0.3 1 .7 ± 0.4 1.4 ± 0.3 3.5 ± 0.5 0.8 ± 0.2 1 .0 ± 0.3 D. virens Blackburnian Warbler 0.4 ± 0.2 0.9 ± 0.3 0.3 ± 0.2 0.4 ± 0.2 D. fusca Yellow-throated Warbler 0.1 ±0.1 0.1 ± 0.1 0.9 ± 0.3 0.2 ± 0. 1 0.8 ± 0.3 D. dominica 524 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 3. September 2010 TABLE 1. Continued. Treatment (years after fires) Low Medium High Species Scientific name Control 1-2 3-6 1-2 3-6 1-2 3-6 Pine Warbler 0.2 ± 0.1 0.2 ± 0.1 0.3 ± 0.1 0.2 ± 0.1 D. piniis Prairie Warbler 0.1 ±0.1 1.0 ± 0.3 D. di.tcolor Cerulean Warbler 0.2 ± 0.2 D. cerulea Black-and-white Warbler 1.8 ± 0.3 0.6 ± 0.2 1.4 ± 0.3 1.5 ± 0.3 1.3 ± 0.3 1.0 ± 0.2 0.9 ± 0.3 Mniotilta varia American Redstart 0.1 ±0.1 0.1 ± 0.1 0.6 ± 0.2 Setophaga ruticilla Worm-eating Warbler 0.6 ± 0.2 0.3 ± 0.2 0.3 ± 0.2 0.1 ± 0.1 0.5 ± 0.2 0.2 ± 0.2 Helmitheros vermivorum Ovenbird 4.7 ± 0.5 1.1 ±0.3 2.3 ± 0.5 2.9 ± 0.4 1.2 ± 0.3 0.2 ± 0.2 Seiurus aurocapilla Kentucky Warbler 0.2 ± 0.1 0.1 ± 0.1 0.1 ± 0.1 0.1 ± 0.1 0.2 ± 0.1 0.1 ± 0.1 Opororni.^ formosus Common Yellowthroat 0.1 ± 0.1 Geothlypis trichas Hooded Warbler 0.9 ± 0.3 1.8 ± 0.3 2.5 ± 0.4 0.6 ± 0.2 1.3 ± 0.3 1.4 ± 0.4 Wilsonia citrina Canada Warbler 0.2 ± 0.1 W. canaden.sis Yellow-breasted Chat 0.2 ± 0.2 0.1 ± 0.1 2.9 ± 0.5 Icteria vireos Eastern Towhee 0.5 ± 0.2 0.4 ± 0.2 1.4 ± 0.4 0.8 ± 0.2 0.9 ± 0.3 4.2 ± 0.5 4.0 ± 0.5 Pipilo erythrophthcdmu.s Chipping Sparrow 1.2 ± 0.4 0.2 ± 0.1 Spizella pcisserina Field Sparrow 0.1 ± 0.1 S. pusilla Song Sparrow 0.1 ± 0.1 Melospiza melodici Dark-eyed Junco 0.1 ±0.1 0.2 ± 0.1 Jiinco hyertudis Scarlet Tanager 1 .2 ± 0.3 1 .4 ± 0.3 2.4 ± 0.5 1 .9 ± 0.3 2.4 ± 0.4 2.6 ± 0.4 2.3 ± 0.4 Piranga oHvacea Northern Cardinal 0.1 ±0.1 0.3 ± 0.2 0.8 ± 0.2 0.1 ± 0.1 0.1 ±0.1 0.6 ± 0.2 Ccirdbudi.'i cardimdis Rose-breasted Grosbeak 0.2 ± 0.1 0.4 ± 0.2 0.3 ± 0.2 0.6 ± 0.2 Pheucticus litdovicianus Klaus ei al. • FIRE AND BIRDS IN THE SOUTHERN APPALACHIAN MOUNTAINS 525 TABLE I. Continued. Treatnienl (years after fires) Species Low Medium High Scientific name Control 1-2 3-6 1-2 3-6 1-2 3-6 Blue Grosbeak Passerine! caeriilea 0.4 ± 0.2 Indigo Bunting P. cycinea 1.0 ± 0.3 0.7 ± 0.3 2.5 ± 0.5 2.6 ± 0.4 2.1 ± 0.3 4.0 ± 0.4 5.3 ± 0.6 American Goldfinch 0.1 ± 0.1 0.1 ± 0.1 1.0 ± 0.3 0.6 ± 0.3 Spilius tristis after fires, relative to all other treatments, was significantly higher among the high severity bum units (Fig. 2). Generally, species diversity (//') did not differ among treatments or relative to time since fire (F(,_g = 2.4, P = 0.12). However, species diversity was significantly greater in the high severity bum units 3-6 years after fires in contrast with the controls (/ = -3.0, df = 9, F = 0.1) (Table 2). Species evenness did not differ among treatments (F(, g = 0.43, P = 0.84) (Fig. 2). Canonical correspondence analysis for bird species assemblages and habitat measured 1- 2 years after fires indicated the overall relation- ship between species and environmental variables (all canonical axes) differed significantly from those derived randomly (Table 2). The primary axis (horizontal axis. Fig. 3) indicated strong positive relationships with BASAL AREA, SHRUB DENSITY, and CANOPY COVER, a moderate positive effect with SHRUB HEIGHT, and a moderate negative effect with HERBA- CEOUS COVER (Table 2). The secondary axis (vertical axis, Eig. 3), indicated a strong negative relationship with SHRUB HEIGHT and weaker relationships with SHRUB DENSITY, CANOPY COVER, and BASAL AREA. The habitat in high severity burn units 1-2 years after fires was best characterized by reduced CANOPY COVER and BASAL AREA (Eig. 3). The avian community associated with high severity burn units 1-2 years after fires was largely represented by early successional species including American Goldfinch (scientific names are in Table 1 ), Eastern Towhee, Indigo Bunting, and Eastern Wood-Pewee. These species were along the negative side of the primary axis. Ovenbirds, a species most closely associated with habitat characterized by higher CANOPY COV- ER, BASAL AREA, and HERBACEOUS COV- ER tended to be in a positive position relative to both the primary and secondary axes. Treatment ellipses overlapped among other species indicat- ing that habitat metrics and species pools were similar among treatments (Fig. 3). Relationships between species and habitat 3- 6 years after fires were significantly different for all variables except SHRUB HEIGHT (Table 2). This metric was omitted from further analysis. The primary axis of the CCA had a strong negative relationship with BASAL AREA and CANOPY COVER, and a strong positive rela- tionship with SHRUB DENSITY. The secondary axis had a strong negative relationship with HERBACEOUS COVER. High severity burn units were best characterized by habitat with less CANOPY COVER, lower BASAL AREA, and higher SHRUB DENSITY. Several species along the primary axis were depicted farthest from the centroid and included Yellow-breasted Chat, Chestnut-sided Warbler, Indigo Bunting, and Eastern Towhee (Fig. 4). Species such as Hooded, Black-and-White, and Black-throated Green war- blers were mapped highest on the vertical axis (Fig. 4). DISCUSSION The effects of fire severity can vary between species and across avian communities. Our results indicated both species richness and diversity increased relative to fire severity and time since fire. Relative abundance did not change consid- erably among treatments or with time since fire (Table 1, Fig. 2); this may have been a bias from our conservative analysis of repeat observations. Species evenness also remained stable among treatments. The species most negatively affected by fire was the Ovenbird, which is associated with closed 526 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 3, September 2010 VI VI O o VI o • o (D Dh CO o A o o (L> -H CLh W d o o 8 9 O o Vi (D Vi (D O 20% of surveys are shown. 528 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. 3, September 2010 CAI FIG. 4. Habitat associations relative to fire severity for species (BAWW = Black and White Warbler, etc.) detected on point-count surveys conducted 3-6 years after fires in northern Georgia. Community mapped using Canonical Correspondence Analysis (CCA) with habitat centroids based on linear combination scores derived from habitat measured at control units (thin solid line), low .severity (thick solid line), medium severity (thin dashed line), and high severity (thick dashed line). Boxed habitat metrics are mapped with arrows indicating relative locations within the community. The longer the di.stance between the metric and the community center (axis point 0, 0) the stronger the relationship with the community. Proximity of individual species to habitat metrics indicates .strength of association between relative abundance and habitat metric. Only those species detected on >20% of surveys are shown. species associated with disturbance and early successional habitat responded positively to fire severity. Species that responded positively to increa.sed fire severity included American Goldfinch, Chest- nut-sided Warbler, Eastern Towhee, Indigo Bun- ting, Prairie Warbler, and Yellow-breasted Chat. Most of these species had higher abundance 3- 6 years after fires, tracking decreased canopy cover and released growth of shrubs relating to increa.sed shrub density. Also included among these species were Golden-winged and Cerulean warblers, species that are both endangered in Georgia. Both species were encountered once and were found exclusively in either medium (Gold- en-winged Warbler) or high severity (Cerulean Warbler) treatments (Table 1 ). The physiognomic properties of habitats con- tinued 3-6 years after high severity fires. Continued canopy openness related positively with abundance of several woodpecker species; Hairy Woodpeckers and Northern Flickers had highest abundance in high severity treatments. Species characteristic of forests with more-open canopy, such as Yellow-billed Cuckoo, Eastern Wood-Pewee, and Blue-gray Gnatcatchers in- creased relative to fire severity, responses that may relate to suppressed understory, more-open conditions, higher insect abundance, and im- proved foraging habitat (Greenberg et al. 2007). Mid-story nesting species are thought to be relatively flexible in habitat .selection and may not respond to fire-driven habitat change (Artman et al. 2005). Our results indicated density' of several of these mid-story nesting species including Wood Thrush, Blue-headed Vireo, and Red-eyed Vireo did not differ relative to either fire severity or time since fire. However, American Redstart, a species that forages and nests in shrubs and vines, appeared to be absent from treatment burn units K/au.s et al. • FIRE AND BIRDS IN THE SOUTHERN ARPALACHIAN MOUNTAINS 529 during the first 2 years after fires. This association may reflect the temporary reduction of vertical structure during the first few years after a fire (Waldrop et al. 2007). Uneven-age forests may represent natural historical conditions within the southern Appala- chian Mountains (Lorimer 1980). Our results indicated that returning a fire regime to these forests may be an effective conservation tool for restoring avian diversity (Artman et al. 2005, Greenberg et al. 2007). However, effectiveness of fire restoration can vary (Artman et al. 2005). Continued fire suppression, or frequent applica- tion of low severity fires, may lead to forest maturation, have limited effects on species diversity, and can lead to declines in some species (Artman et al. 2005, Klaus et al. 2005). Applica- tion of high severity fires corresponding with short rotation periods may not allow habitat regeneration and can lead to decreased diversity. Application of higher severity fires over broader time intervals may be more effective in managing uneven-age forests. Changes in relative abundance of a particular species may not equate directly with habitat quality or persistence (Johnson 2007). We did not examine the effects of treatment on reproduc- tive success, survival, or other demographic measures and cannot provide predictions relating treatments to these population metrics. Our results reflect acute numerical relationships among individual species and across avian communities. Longer-term studies focused on addressing vari- ation in population demographics through several seasons after different fire severity treatments would greatly benefit our understanding of avian communities. CONSERVATION IMPLICATIONS Management directed at conservation of avian species should balance restoration of early successional habitat with conservation of mature forest (Dettmers 2003, Artman et al. 2005, Bulluck and Buehler 2006, Buehler et al. 2007). Our results, and those of others (Artman et al. 2005, Greenberg et al. 2007), indicate conserva- tion activities that foster avian diversity within the southern Appalachian Mountains may benefit from occasional high severity fires and a fire return interval of 3 < 10 years. We are not suggesting all burns should be of high severity, only that the full range of fire severity be used through time and space. Fire may be most beneficial to avian conserva- tion when it is used over relatively large spatial .scales (Artman et al. 2005, Greenberg et al. 2007). Managed fires will help develop a matrix of habitats and successional stages, and minimize the isolation of populations (Askins 2001, Brawn et al. 2001). A rationale against pre.scription of smaller-scale fires is that larger fires may be more easily controlled in habitat characterized by terrain similar to that in the Chattahoochee- Oconee National Forest. Larger burn units can make use of creeks and roads for fire breaks helping avoid undue disturbance such as plowing, which may be a vector for invasion of exotic species (Memam et al. 2006). Our study did not examine the cumulative impacts of repeated low intensity burns, or combinations of burn intensi- ties. However, our results suggest that medium and high intensity fires can have positive effects on bird communities and should be considered a valuable land management tool. ACKNOWLEDGMENTS We acknowledge support of the D. B. Warnell School of Forest Resources of the University of Georgia, the Nongame Conservation Section of the Georgia Department of Natural Resources, and the U.S. Forest Service. We thank J. S. Hender.son, and J. M. Wentworth for helping select study sites, D. G. Gregory and C. E. Hans for assisting with bird survey work, and M. N. Hayes for assistance with data entry. LITERATURE CITED Abrams, M. D. 1992. Fire and the development of oak forests in eastern North America; oak distribution reflects a variety of ecological paths and disturbance conditions. Bioscience 42:346-353. Anders, A. D., J. Faaborg, and F. R. Thompson 111. 1998. Postfledging dispersal, habitat use, and home-range size of juvenile Wood Thrushes. Auk 1 15:349-358. Aquilani, S. M.. D. C. Leblanc, and T. E. Morrell. 2000. Effects of prescribed surface fires on ground- and shrub-nesting neotropical migratory birds in a mature Indiana oak forest, USA. Natural Areas Journal 20:317-324. Artman, V. L. and J. F. Downhover. 2003. Wood Thrush (Hylodchla miistelina) nesting ecology in relation to prescribed burning of mixed-oak forest in Ohio. Auk 120:874-882. Artman, V. L., E. K. Sutherland, and J. F. Downhover. 2001. Prescribed burning to restore mixed-oak com- munities in .southern Ohio: effects on breeding bird populations. Conservation Biology 15:1423-1434. Artman, V. L., T. F. Hutchinson, and J. D. Brawn. 2005. Fire ecology and bird populations in eastern 530 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 3, September 2010 deciduous forests. Studies in Avian Biology 30:127- 138. Askins, R. a. 2001. Sustaining biological biodiversity in early successional communities: the challenge of managing unpopular habitats. Wildlife Society Bulle- tin 29:407-412. Bond, W. J. and J. F. Keeley. 2005. Fire as a global ‘herbivore’: the ecology and evolution of flammable ecosystems. Trends in Ecology and Evolution 20:387- 394. Bond, W. J., F. I. Woodward, and G. F. Midgley. 2005. The global distribution of ecosystems in a world without fire. New Phytologist 165:525-538. Brawn, J. D., S. K. Robinson, and F. R. Thompson III. 2001. The role of disturbance in the ecology and conservation of birds. Annual Review of Ecology and Systematics 32:251-276. Brewster, W. 1886. An ornithological reconnaissance in western North Carolina. Auk 3:94-112. Buehler, D. a., a. M. Roth, R. Vallender, T. C. Will, J. F. Confer, R. A. Canterbury, S. B. Swarthout, K. V. Rosenberg, and L. P. Bulluck. 2007. Status and conservation priorities of Golden-winged Warbler (Vennivora chrysoptera) in North America. Auk 124:1439-1445. Bulluck, L. P. and D. A. Buehler. 2006. Avian use of early successional habitats: are regenerating forests, utility right-of-ways and reclaimed surface mines the same? Forest Ecology and Management 236:76-84. Burleigh, T. D. 1958. Georgia birds. University of Oklahoma Press, Norman, USA. Dettmers, R. 2003. Status and conservation of shrubland birds in the northeastern U.S. Eorest Ecology and Management 185:81-93. Elliott, K. L., R. L. Hendrick, A. E. Major, J. M. Vose, AND W. T. Swank. 1999. Vegetation dynamics after a prescribed fire in the southern Appalachians. Forest Ecology and Management 114:199-213. Greenberg, C. H., A. L. Tomcho, J. D. Lanham, T. A. Waldrop, J. Tomcho, R. J. Phillips, and D. Simon. 2007. Short-term effects of fire and other fuel reduction treatments on breeding birds in a southern Appalachian upland hardwood forest. Journal of Wildlife Management 71:1906-1916. Hamel, P. B., W. P. Smith, D. J. Twedt, J. R. Woeher, E. Morris, R. B. Hamilton, and R. J. Cooper. 1996. A land manager’s guide to point counts in the Southea.st. USDA, Forest Service, General Technical Report SO- 120. Southern Re.search Station, New Orleans, Louisi- ana, USA. Hunter, W. C., R. Katz, D. Pashley, and B. Ford. 1999. Partners in flight bird conservation plan for the southern Blue Ridge (Physiographic Area 23). USDl, Fish and Wildlife Service, Atlanta, Georgia. USA. HtiNTER, W. C. D. A. Buehler, R. A. Canterbury, J. L. Confer, and P. B. Hamel. 2001. Conservation of disturbance-dependent birds in ea.stern North America. Wildlife Society Bulletin 29:440-455. .loilN.SON, M. D. 2007. Measuring habitat quality: a review. Condor 1 09:489-504. Klaus, N. A. 2004. Status of the Golden-winged Warbler in north Georgia, and a nesting record of the Lawrence’s Warbler. Oriole 69:1-7. Klaus, N. A., D. A. Buehler, and A. M. Saxton. 2005. Forest management alternatives and songbird breeding habitat on the Cherokee National Forest, Tennessee. Journal of Wildlife Management 69:222-234. Lang, J. D., L. A. Powell, D. G. Krementz, and M. J. Conroy. 2002. Wood Thrush movements, habitat use, and effects of forest management for Red-cockaded Woodpeckers. Auk 119:109-124. Lorimer, C. G. 1980. Age structure and disturbance history of a southern Appalachian virgin forest. Ecology 61:1169-1184. Lorimer, C. G. 2001. Historical and ecological roles of disturbance in eastern North American forests: 9,000 years of change. Wildlife Society Bulletin 29:425-439. Loucks, C. G. 1970. Evolution of diversity, efficiency, and community stability. American Zoologist 10:17-25. Magurran, a. E. 1988. Ecological diversity and its measurement. Princeton University Press, Princeton, New Jersey, USA. Marshall. M. R., J. A. DeCecco, A. B. Williams, G. A. Gale, and R. J. Cooper. 2003. Use of regenerating clearcuts by late-successional bird species and their young during the post-fledging period. Eorest Ecology and Management 183:127-135. McCune, B. and J. B, Grace. 2002. Analysis of ecological communities. MJM Software Design. Glenedon Beach, Oregon, USA. Merriam, K. E., j. E. Keeley, and J. L. Beyers. 2006. Euel breaks affect nonnative species abundance in California plant communities. Ecological Applications 16:515-527. Neimi, G. J. and j. M. Hanowski. 1984. Relationships of breeding birds to habitat characteristics in logged areas. Journal of Wildlife Management 48:438-443. Partners in Plight Working Group. 2002. Priority research needs for the conservation of neotropical migrant landbirds. Journal of Field Ornithology 73:329-339. PiNHEiRO, J. C. AND D. M. BATES. 2000. Mixed-effects models in S and S-Plus. Springer-Verlag. New York. USA. Raudenbush. S. W. and a. S. Bryk. 2002. Hierarchical linear models: applications and data analysis methods. Sage Publications, Thousand Oaks, California, USA. R Development Core Team. 2008. R: a language and environment for statistical computing. R Foundation for Statistical Computing, Vienna. Austria. Rush, S. A. and B. J. M. Stutchbury. 2008. Survival of fledgling Hooded Warblers (WHsonia citriua) in small and large forest fragments. Auk 125:183-191. Saab, V. A. and H. D. W. Powell. 2005. Fire and avjan ecology in North America: process influencing pattern. Studies in Avian Biology 30:1-13. Sauer. J. R., J. E. Hines, and J. Fallon. 2005. The North American Breeding Bird Survey, results and analysis 1966-2005. Version 6.2.2006. USGS, Patuxent Wild- life Re.search Center, Laurel, Maryland, USA. Klaus et al. • FIRE AND BIRDS IN THE SOUTHERN APPALACHIAN MOUNTAINS 53 I Shannon, C. E. and W. Weaver. 1963. Tlic muihemaiical theory of communication. University ol' Illinois Press, Urbana, USA. Smith, T. M. and H. H. Shugart. 1987. Territory size variation in the Ovenbird: the role of habitat structure. Ecology 68:695-704. Stupka, A. 1963. Notes on the birds of the Great Smoky Mountains during the breeding season. Migrant 17:60- 62. Tomcho, a. L., C. H. Greenberg, J. D. Lanham, T. A. Waldrop, J. Tomcho, and D. Simon. 2006. Effects of fuel reduction treatments on breeding birds in a southern Appalachian upland hardwood forest. USDA, Forest Service. General Technical Report PSW-GTR-203. Pa- cific Southwest Research Station, Portland, Oregon, USA. V.AN Lear, D. H. and T. A. Waldrop. 1989. History, use, and effects of fire in the southern Appalachians. USDA, Forest Service, General Technical Report SE- 54. Southeastern Forest Experiment Station, Asheville, North Carolina, USA. Vega Rivera, J. H., J. H. Rappole, W. J. McShea, and C. A. Haas. 1999. Postbreeding movements and habitat use of adult Wood Thrushes in northern Virginia. Auk 1 16:458-466. Waldrop, T. A., D. L. White, and S. M. Jones. 1992. Fire regimes for pine-grassland communities in the south- eastern United States. Forest Ecology and Manage- ment 47:195-210. Waldrop, T. A., D. A. Yaussey, R. J. Phillips, T. A. Hutchinson, L. Brudnak, and R. E. J. Boerner. 2007. Fuel reduction treatments affect stand structure of hardwood forests in western North Carolina and southern Ohio, USA. Forest Ecology and Management 255:31 17-3129. The Wilson Journal of Ornithology 122(3); 532-544, 2010 AVIAN COMMUNITIES OF THE ALTAMAHA RIVER ESTUARY IN GEORGIA, USA ROSS A. BRITTAIN, ' VICKY J. MERETSKY,' AND CHRIS B. CRAET' ABSTRACT. — We surveyed male breeding birds in five habitats (bottomland forest, maritime oak [Quercus spp.], pine [Finns spp.] forest, maritime shrub, saltmarsh) of coastal Georgia, USA using distance-sampling methods to estimate population densities. We examined species-habitat relationships using indicator species analysis (ISA). Acadian Flycatcher {Empidona.x virescens) in bottomland forest. Northern Parula (Paritla americana) in maritime oak. Brown-headed Nuthatch (Sitta piisilla) in pine forest. Clapper Rail (Rallus longirostris) in saltmarsh, and White-eyed Vireo (Vireo griseiis) in shrub habitat ranked highest for Partners in Flight (PIF) priority species by densities. The ISA indicated fewer PIF priority species were associated with saltmarsh, but more species (6) were unique to saltmarsh than any other habitat. Indicator species occurred more often in maritime oak than bottomland forest (8 vs. 6). but both habitats had similar numbers of PIF priority species (4). Shrub habitat covered the smallest area (~0.2%) and had three PIF priority species, including Painted Bunting (Passerina ciris), the only PIF species with extremely high priority in this study. Received 27 November 2009. Accepted 18 March 2010. Indicator species appeal to conservationists, land managers, and governmental agencies be- cause they provide an efficient method of monitoring the impacts of environmental distur- bances and management policies (Carignan and Villard 2002). Thus, indicator species analysis (ISA), based on combining frequency of occur- rence and mean abundance (Dufrene and Le- gendre 1997), is gaining use in conservation studies (Graham and Blake 2001, Kirk and Hobson 2001, Morissette et al. 2002, Mouillot et al. 2002, Grundel and Pavlovic 2007). The use of indicator species can be helpful if it includes monitoring many species that represent different taxa and life histories, the chosen species monitored have excellent quantitative regional data, and habitat-related trends can be distin- guished from natural variation (Carignan and Villard 2002). Once a suite of indicator species has been cho.sen to represent a community, monitoring activities need to efficiently and reliably estimate appropriate population parameters that signal potential impacts to conservation stakeholders (Thomas 1996). Two relatively simple, yet meaningful, avian abundance parameter estima- tions for modeling population dynamics are species density and size (Buckland et al. 2001). Population size is the density times the area of appropriate habitat, and research has focused on ' Indiana University. School of Public and Environmental Affairs, Room 410, Bloomington, IN 47405, USA. ^Current addre.ss: 3475 Winchester Drive, Greenwood, IN 46143, USA. ’Corre.sponding author: e-mail: rabritta@indiana.edu improving estimates of population density. Other important parameters, including population de- mographics and survival, are more difficult to assess because they require capturing individuals (Nisbet 2001). A common method used to estimate avian abundance has been count data (Hodges and Krementz 1996, Bajema et al. 2001 , Rosenstock et al. 2002). The point-count method is relatively simple to conduct but assumes bird detection remains constant across different habitats, observ- er abilities, weather conditions, and species characteristics (Rosenstock et al. 2002). Dis- tance-sampling methods yield more precise esti- mates of bird density than index methods, such as point counts, by adjusting for detection (Rosen- stock et al. 2002, Thompson 2002). A 7-year comparison of point-count methods versus dis- tance sampling found the latter method was more robust in large-scale, multispecies surveys (Nor- vell et al. 2003). Distance-sampling methods have been used in many studies to assess quantitative differences in habitat use by avian species to distinguish species- habitat relationships (Hodges and Krementz 1996, Estades and Temple 1999, Fletcher and Koford 2002), but the method does not work well with rare species (Buckland et al. 2001). Conversely, ISA is valuable for assessing .species-habitat correlations when species are rare, non-normality exists, where distance sampling may be logisti- cally difficult, and where there may be no detection of a species at many sample sites (Mouillot et al. 2002). Our objective was to examine avian communi- ties and densities of male breeding bird species in 532 Brinahi cl cil. • COASTAL GEORGIA AVIAN COMMUNI TIES 533 FIG. 1. Study area and sample sites on Sapelo Island, McIntosh County, Georgia and Clayhole Swamp, Glynn County, Georgia. five habitats of coastal Georgia, USA (tidal- freshwater broadleaf deciduous forest, saltmarsh, maritime scrub-shrub, maritime broadleaf ever- green forest, and maritime narrowleaf evergreen forest). METHODS Study Area. — We surveyed broadleaf deciduous bottomland forest affected by tides, saltmarsh, maritime scrub-shrub, maritime broadleaf ever- green forest, and maritime narrowleaf evergreen forest in coastal Georgia, USA (Fig. 1) for male breeding birds. We selected a broadleaf deciduous forest affected by tides (bottomland forest) at the Clayhole Swamp Wildlife Management Area adjacent to the Altamaha River, Glynn County, Georgia, managed by the Department of Natural Resources (GDNR) (Duberstein and Kitchens 2007). We selected four other habitats on 6,677- ha Sapelo Island, McIntosh County, Georgia, managed by GDNR. We selected saltmarsh habitat primarily on the southwestern portions of the island and maritime scrub-shrub habitat (shrub) based on linear landscape features, such as secondary dunes, on the edge between forest and saltmarsh (Johnson et al. 1974). Maritime broadleaf evergreen forests (maritime oak [Quer- cus spp.]) were dominated by live oak {Q. virginiana), primarily on the north end of the island (Johnson et al. 1974). Maritime narrowleaf evergreen forests (pine [Pinus spp.] forest) consisted of loblolly pine {P. taeda) with scattered slash pine (P. elliottii) and were distributed throughout the island (Johnson et al. 1974). GDNR manages Sapelo Island’s pine forests for sawtimber (old growth) using fire and timber harvest. We selected bird sample points at least 250 m apart. We used a grid in bottomland forest stalling from the southernmost accessible point via Honeygal Road. Saltmarsh sample points on Sapelo Island were randomly selected using a l-m^ grid overlain on a Geographic Information System (GIS) shapefile. Sample points in shrub habitat were in the ecotone between primary dunes and terrestrial forests on Sapelo Island. Oak and pine forest sample points were selected by driving one-lane sand roads and randomly locat- ing them between 10 and 200 m on a line directly into forest patches that were relatively consistent within a 100-m radius. 534 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 3, September 2010 Bird Sampling. — We conducted point counts following Ralph et al. (1993). We located 10 points in each habitat in 2006 (50 total) and an additional 20 points per habitat in 2007 (total of 150 points). Ten-minute counts were conducted twice at least 2 weeks apart at each point within 4 hrs of sunrise between 19 May and 9 June 2006, and between 17 May and 13 June 2007. The relative direction and distance of each detected bird from the sample point during each sampling event, excluding flyovers, were estimat- ed. Distances were verified in the field on the way to the next point by pacing the distance from the point just sampled to a detected bird. Detections of the same species only included birds detected simultaneously or with obvious differences in plumage. Observer bias issues were limited by having a single observer conduct all sampling events. Population Densities. — The density (# of birds/ km") of species sampled using point counts was estimated in program DISTANCE 5.0 (Thomas et al. 2006) and equation D = n/kv, where D — density of birds per unit area, n = number of birds detected, k = number of points sampled, and V = the detection function of the species with distance from the sample point following Buck- land et al. (2001). The detection function indicates the probability of detecting an individual as a function of distance from the sample points for a particular species (Buckland et al. 2001). The density of each species is calculated by fitting the frequency of detection as a function of distance into each of 12 theoretical models using the best fitting model (with the lowest Akaike’s Information Criterion [AIC] value) (Buckland et al. 2001). The frequency distribution of distance to detected (seen or heard) males for each species was pooled across all habitats to increase the sample size of detections. The cutoff for exclusion from DISTANCE analysis was <50 observations for a species across all habitats and <10 individuals of a species in a given habitat. We tested for differences in detection between habitat types using DISTANCE without pooling the frequency distribution across habitats, but only pooled results are presented due to lack of differences. Densities were considered to be different when there was a lack of overlap of the 95% confidence intervals (Gardner and Altman 1986, Hodges and Krementz 1996, Anderson et al. 2001, Fletcher and Koford 2002). DISTANCE requires detection functions to monotonically decrease with distance from the sample point so the detection function is either flat or decreases with distance (Thomas et al. 2006). DISTANCE constrains parameters of the detection function at fixed points to fit the function to the model, if the detection function increases with distance. These constraints distort the detection function which may be better adjusted by truncating the tail of the data or altering the intervals to smooth the detection function (Thomas et al. 2006). We also conducted the analyses by truncating detections at 80- and 90-m, but only the non-truncated results are reported due to lack of differences. Models were chosen based on the least number of constrained parameters to obtain monotonicity and the lowest AIC value. Statistics. — Species-habitat associations were assessed using indicator species analysis and 9999 Monte Carlo iterations (ISA; McCune and Grace 2002). ISA results include the relative abundance, relative frequency, highest indicator value, and P values of significant differences. Statistical results were considered significant at a < 0.05. Dufrene and Legendre (1997) recommend a cutoff threshold indicator value (IV) of 25 for a species to be considered truly indicative of a habitat. RESULTS Species Abundance. — We detected 56 species during the two breeding seasons (scientific names and Partners in Flight priority status are in Table 1). We encountered Northern Cardinals and Carolina Wrens more often than any other species (401 and 400 encounters for 5 habitats, respectively). Red-eyed Vireos were most abun- dant in bottomland forest (142), Northern Parula in maritime oak (180), Red-bellied Woodpecker in pine forest (116), Clapper Rail in saltmarsh (150), and Northern Cardinal in shrub (140) (Table 2). Bird Densities. — We estimated breeding male densities for 26 species overall (Table 2). None of the 14 species for which we estimated densities in bottomland forest was unique to this habitat, but Prothonotary Warbler and Barred Owl were only detected in bottomland forest. Tufted Tit- mouse densities were significantly higher in bottomland forest than all other species (172 Brittain et al. • COASTAL GEORGIA AVIAN COMMUNITIES 535 TABLE 1. Species detected during point counts in five habitats in coastal Georgia during 2006 and 2007. Species Scienufic name AOU code Great Blue Heron Ardeci heraciias* GBHE Great Egret A. alha^ GREG Snowy Egret Egretta thiiUA SNEG Yellow-crowned Night-Heron Nyctanassa violacea^ YCNH Red-shouldered Hawk Buteo lineatiis RSHA Red-tailed Hawk B. jamaicensis RTHA Clapper Rail Rallus longirostris^^^ CLRA Wilson’s Plover C ha rci drills wilsonia^^^ WIPE Willet Tringa semipahnata^^ WILL Whimbrel Numeniiis phaeopus WHIM Ruddy Turnstone Arenaria interpres RUTU Mourning Dove Zenaida macroura MODO Common Ground-Dove Columbina passerina^^ COGD Yellow-billed Cuckoo Coccyzus aniericanus^^ YBCU Barred Owl Strix varia BADO Ruby-throated Hummingbird Archilochus colubris RTHU Red-headed Woodpecker Melanerpes eni’throcephalus^^ RHWO Red-bellied Woodpecker M. carolinus RBWO Downy Woodpecker Picoides pubescens DOWO Pileated Woodpecker Dryocopus pileatus PIWO Eastern Wood-Pewee Coni opus V ire ns* EAWP Acadian Flycatcher Empidonax virescens* ACFL Great Crested Flycatcher Myiarchus crinitus GCFL Eastern Kingbird Tyrannus tyrannus EAKI White-eyed Vireo Vireo griseus** WEVI Yellow-throated Vireo V. flavifrons* YTVI Red-eyed Vireo V. olivaceus REVI Blue Jay Cyanocitta cristata BLJA American Crow Corvus brachyrhynchos AMCR Fish Crow C. ossifragus FICR Carolina Chickadee Poecile carolinensis** CACH Tufted Titmouse Baeolophus bicolor TUTI Brown-headed Nuthatch Sitta pusilla*** BHNU Carolina Wren Thryothorus ludovicianus CARW Marsh Wren Cistothorus palustris MAWR Blue-gray Gnatcatcher Polioptila caerulea BGGN Eastern Bluebird Sialia sialis EABL Northern Mockingbird Mimus polyglottos NOMO Brown Thrasher Toxostoma rufum BRTH Northern Parula Parula americana*** NOPA Yellow-throated Warbler Dendroica dominica*** YTWA Pine Warbler D. pinus** PIWA Prothonotary Warbler Protonotaria citred** PROW Common Yellowthroat Geothlypis trichas COYE Hooded Warbler Wilsonia citrina*** HOWA Yellow-breasted Chat Icteria virens* YBCH Eastern Towhee Pipilo erythrophthabnus** EATO Seaside Sparrow Ammodramus maritimus*** SESP Summer Tanager Piranga rubra* SUTA Northern Cardinal Cardinalis cardinalis NOCA Blue Grosbeak Passerina caerulea BLGR Painted Bunting P. ciris**** PABU Red-winged Blackbird Agelaius phoeniceus RWBL Boat-tailed Grackle Quiscalus major BTGR Brown-headed Cowbird Molothrus ater BHCO Orchard Oriole Icterus spurius** OROR ^ Local interest species (Hunter et al. 2001). Moderate priority species. High priority species. tttt Extremely high priority species. TABLE 2. Total detections (counts) and breeding male population densities (birds/km^) with 95% confidence interval (in parentheses) calculated using DISTANCE loi birds breeding in five habitats of coastal Georgia. THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. 3, September 2010 CN CN On IT) o o NO S' XI o O m (N 1 — r- ON 1 1 1 1 1 00 o 1—1 00 — 9—1 m m 00 in o '3* r4 CN CN r- NO CN "Tf CN ON m On — ' — -m’^'^’^NOONm ^ ON cn r- (N sD 00 lo 00 ON o in ON On ON o (N u .c •5-s. s .—K V S _ •rf s. ON in o NO ^ Tj- ON r- CN '3- in m r- 00 CN in 00 1 cn 00 — — 1 ON 1 1 — • 1 1 1 On 1 — 1 1 00 1 On o 1 On ON O CN 1 CN 00 ON r- CN tJ- — ^— s — —1 in r- CN r- ' — ' 00 — m m ON ON 00 ON oo ON CN m r- CN o o ON 00 — — CN — NO (N m On nO ^ m — NO m Tt — NO — fNlOOOfOOOOCNNOON o X 2 T3 ^ 1) t/5 1) > o 9 (U "O C (U c O > 3 E Q J H 3 S o o o 3 u •o OJ (/) w o O O c/5 >" U Oi ■a — OJ C 02 ^ p C 5 E ^ ^ > ^ ■R U D. O. -o-R ^ T3 O ^ U 2 ^ ^ -o fU ^ OJ o x: o -3 ^ ' H u- 3 _ O ^ >N 2 X § U * o o ^ w E .2 1) 2 ^ - 3 qj c/5 '— ^^o;Su>“oace:Q2o2Qa.cu> 'B < a OU 1^ « ■E "o 2 ^ ^-S c V j a o B * o (U O ^ -A o ^ 3 (U — 0^ X o u. U 3 O •c U 3 3 < o x> S— N 00 ■5-v OS o r-- 0 OS 5-v a in CN O O r- — o 1 Os — H 1 oo m 1 CN 00 m m 1 04 1 OS CN O 1 00 4 2: 1 SO 1 00 c-\ 00 m (N CN >r so r-- m in o m r-. 00 Os (N in m m C^l m O m cN in m 00 O'. o — so — so r- CN in — ' so Os 00 CN — CN in I m oi so r-* O'! I (N » no U o "O o (U o ^ • - - oij u 2^ c i a -s ? -5 = ^ o £ c CQ ttJ Z 03 Z c3 O O c o x: o -a E ^ ^ "O E o o o Z H ^ E o - LU ^ 25 cutoff; an additional 21 species were indicators accord- ing to the ISA but had IV scores < 25 (Fig. 2). Six species were indicative of bottomland forest and five species were indicative of saltmarsh. Shrub habitat had the fewest indi- cator species (4), but the species with the highest IV score in shrub habitat was the only extremely high priority species in the region, Painted Bunting. Eight species were indicators of mari- time oak and five were indicators of pine forest. DISCUSSION Implications for Conservation. — We estimated a proportionately high number of PIF priority species were indicative of bottomland forest (4), which occupies —10.0% of the coastal Georgia landscape (Brittain 2009). Prothonotary Warblers occurred in bottomland forest sites with high forest cover and basal area along tidal channels (Petit 1999). Similarly, Acadian Flycatchers and Yellow-billed Cuckoos were indicative of wetland forests typical of their southeastern habitats (Stevenson and Anderson 1994, White- head and Taylor 2002); Acadian Flycatchers were afso abundant in maritime oak. Yellow- throated Warblers may have been deterred from occupying bottomland forests, despite another study that observed them in that habitat type (Hall 1996), by the absence of large dead trees (Hamel 1992). Georgia saltmarshes had the lowest number of priority indicator species among the habitats studied (1), but the highest diversity of species unique to the habitat. Seaside SpaiTows occurred exclusively in tall smooth cordgrass {Spartina alterniflora) stands on the banks of larger estuaries with extensive mudflats (Post and Greenlaw 1994). Clapper Rails were detected in tall Spartina of the low marsh and the shrubby vegetation in the high marsh-upland ecotone (Eddleman and Conway 1998). Tall Spartina may be important as habitat for Seaside Span'ows and other saltmarsh birds, such as Marsh Wrens, by creating areas of dense vegetation for secure nests. Eastern Towhees and Yellow-breasted Chats in coastal Georgia, both PIF priority species, were more abundant in shrub habitat than in open forest (Greenlaw 1996, Eckerle and Thompson 2001). Common Ground-Doves were found only in coastal dunes despite known use of open pine, hummocks, and forest edges (Bowman 2002) common on Sapelo Island, likely due to the lack of bare soil in forest habitats (Hamel 1992). The PIF Conservation Plan defines quality shrub Bhitain et al. • COASTAL GEORGIA AVIAN COMMUNITIES 541 habitat as “largely forested areas with some edge and forest openings for buntings” (Hunter et al. 2001). However, shrub habitat, where Painted Buntings, Eastern Towhees, and Yellow-breasted Chats were far more common in this and another study (Springborn and Meyers 2005), was largely composed of coastal dunes. Maritime scrub-shrub habitat deserves conservation priority over other habitats as it occupies only —0.2% of the coastal Georgia landscape (Brittain 2009) and is impor- tant for the area’s only species of extremely high conservation priority. Northern Parula occupied mature hardwood forests with Spanish moss {Tillandsia usneoides) (maritime oak type) and floodplains with blue palm (Brohea annata) (bottomland forest type), as reported by Moldenhauer and Regelski (1996), but were far more abundant in maritime oak. Similarly, Eastern Wood-Pewee and Yellow- throated Vireo occuined in all three forest types, but were more common in maritime oak (McCarty 1996, Rodewald and James 1996). The ISA also revealed White-eyed Vireos were indicators of maritime oak rather than the shrub habitat for which they are listed by PIP (Hunter et al. 2001). A PIE Conservation Plan goal in pine forest includes emphasis on late successional stands with increasing disturbance regimes to enhance understory habitat quality (Hunter et al. 2001). Managed disturbances, such as prescribed fire and logging on Sapelo Island, maintain the open stands of mature pine habitat preferred by Brown-headed Nuthatches (Withgott and Smith 1998). However, this conservation plan may aid Brown-headed Nuthatches and Red-headed Woodpeckers but may harm Pine Warbler popu- lations dependent on a well-developed shrub layer (Rodewald et al. 1999, Smith et al. 2000). Pine forest occupies nearly half of the coastal Georgia landscape (Brittain 2009), and there may be opportunity for diverse strategies to protect associated species. Bird Densities. — Assumptions in distance sam- pling include complete detection of all individuals close to the sample point and that birds do not leave the sampling area prior to detection (Buck- land et al. 1993, Thompson 2002). The detection rates in our study generally increased within the first 5-10 m of the sample point before declining, indicating birds may have moved from the sample point as the observer approached, or the area of near-perfect detection extended ~5 m beyond the sample point itself. DISTANCE could not find acceptable detection functions without constraining parameters to obtain monotonicity for 10 species (38%). Constraining the parameters may have distorted the detection functions and skewed the results. However, alternative methods, such as truncation, did not resolve the constrained parameter issues and yielded results that were not significantly different than those calculated using the original data. The lack of good fit for 65% of the detection functions increased concerns about the model results, but truncation and grouping the data into intervals made the fit worse, as reflected by lower P-values in the Kolmogorov-Smirnov tests, and did not significantly change the results. Eorest structure may hamper detection of silent and sedentary individuals, whereas open habitats, such as saltmarsh, should allow for near complete detection (Pacifici et al. 2008). Thus, breeding male density estimates in forested habitats were likely underestimated. Detection variations among different habitats should be considered (Simons et al. 2007) by limiting the pooling of data to create the detection functions. However, we believed increased sample size outweighed detection differences between habitats. We tested each species without pooling the detection frequency data and the results showed no significant differences between pooled and non-pooled data for most species (not presented). The 95% confidence intervals for each of the species were significantly lower in the non-pooled data than pooled data for White-eyed Vireo (maritime oak and pine forest). Blue-gray Gnatcatcher (shrub), and Yel- low-throated Warbler (shrub). Yellow-throated Warblers also had a significantly higher density estimate for maritime oak using the non- pooled data. However, all models had to con- strain parameters to obtain monotonicity using non-pooled frequency data, and had lower Goodness of Eit P-values, indicating these models were less reliable than those using the pooled data. The lack of improved models and differences in density estimates between results using pooled versus non-pooled frequency data led us to conclude the pooled data were safer to use. Some breeding male densities appear high, particularly Blue-gray Gnatcatchers, Brown-head- ed Nuthatches in pine forest. Northern Parula in maritime oak, and White-eyed Vireos on Sapelo 542 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 3. September 2010 Island. Male Brown-headed Nuthatches, found almost exclusively in pine forests, were often observed traveling in family groups that may have skewed results by being more easily detected in family clusters than individually. Northern Par- ulas were definitively more abundant in maritime oak and the number of males captured during bird banding attempts was similar to the number of detected males (Brittain 2009). White-eyed Vireos and Blue-gray Gnatcatchers appeared curious about the observer’s presence at the sample point, which may have positively biased estimated densities. Our measurements of densities of six species (Acadian Flycatcher, Blue-gray Gnatcatcher, Northern Parula, Prothonotary Warbler, Red-eyed Vireo, White-eyed Vireo) were greater than those reported by Hodges and Krementz (1996). These differences may be the result of variations in study design or individual bird detection skills. Northern Parula densities in upstream riparian corridors were similar to densities in shrub habitat in this study (Hodges and Krementz 1996). Distance sampling indicated the PIF priority species with the highest densities were; Acadian Flycatcher in bottomland forest. Northern Parula in maritime oak. Brown-headed Nuthatch in pine forest. Clapper Rail in saltmarsh, and White-eyed Vireo in shrub habitat. Shrub habitat should have the highest conservation value since it had the highest proportion of indicator species that were also PIF priority species (75%), the only extremely high priority indicator species (Painted Bunting), and occupied a much smaller portion of the coastal land- scape (—0.2%) than other habitats. Maritime oak, bottomland forests, and saltmarsh cover moderate portions of coastal Georgia, but forest habitats had more PIF priority indicator .species than saltmarsh (5, 4, and 1, respectively). Protections given tidal wetlands under Section 404 of the Clean Water Act (e.g., bottomland forest and saltmarsh; PL 92-500, 33 USC 1251), combined with the high percentage of priority species, suggests avian conservation efforts should focus on conserving maritime oak habitat. Pine forest is relatively plentiful and was equal to shrub for the least number of PIF priority indicator species in upland habitat. ACKNOWLEDGMENTS We thank the National Oceanic and Atmospheric Administration (NOAA) and the National Estuarine Research Reserve System (NERRS) for funding this research through a Graduate Research Fellowship; the U.S. Environmental Protection Agency for providing funds for research in the Clayhole Swamp; the Georgia Department of Natural Resources, Wildlife Resource Division, that manages the Sapelo Island NERR; and Buddy Sullivan, Dorset Hurley, Aimee Gaddis, Fred Hay, and Jon Garbisch. LITERATURE CITED Anderson, D. R., W. A. Link, D. H. Johnson, and K. P. Burnham. 2001. Suggestions for presenting the results of data analyses. Journal of Wildlife Management 65:373-378. Bajema, R. a., T. L. DeVault, P. E. Scott, and S. L. Lima. 2001. Reclaimed coal mine grasslands and their significance for Henslow’s Sparrows in the American Midwest. Auk 118:422^31. Bowman, R. 2002. Common Ground-Dove (Columbina passerina). The birds of North America. Number 645. Brittain, R. A. 2009. Trophic status, habitat use and climate change impacts on avian species of coastal Georgia. Dissertation. Indiana University, Blooming- ton, USA. Buckland, S. T., D. R. Anderson, K. P. Burnham, and J. L. Laake. 1993. Distance sampling: estimating abundance of biological populations. Chapman and Hall, London, United Kingdom. Buckland, S. T., D. R. Anderson, K. P. Burnham, J. L. Laake. D. L. Borchers, and L. Thomas. 2001. Introduction to distance sampling: estimating abun- dance of biological populations. Oxford University Press, New York, USA. Carignan, V. AND M. A. ViLLARD. 2002. Selecting indicator species to monitor ecological integrity: a review. Environmental Monitoring and Assessment 78:45-61. Duberstein, J. AND W. KITCHENS. 2007. Community composition of select areas of tidal freshwater forest along the Savannah River. Pages 321-348 in Ecology of tidal freshwater forested wetlands of the southeast- ern United States (W. H. Conner, T. W. Doyle, and K. W. Krauss, Editors). Springer-Verlag, New York, USA. Dufrene, M. and P. Legendre. 1997. Species assemblages and indicator species: the need for a flexible asymmetrical approach. Ecological Monographs 67:345-366. Eckerle, K. P. and C. F. THOMP.SON. 2001. Yellow- breasted Chat (Icteria virens). The birds of North America. Number 575. Eddleman, W. R. AND C. J. Conway. 1998. Clapper Rail {Rcillus longirnstris). The birds of North America. Number 340. Britniiu et al. • COASTAL GEORGIA AVIAN COMMUNITIES 543 Estades, C. F. and S. a. Temple. 1999. Dcciduous-foresl bird conimunitie.s in a fragmented land.scape dominat- ed by exotic pine plantations. Ecological Applications 9:573-585. Fletcher, R. J. and R. R. Koford. 2002. Habitat and landscape associations of breeding birds in native and restored grasslands. Journal of Wildlife Management 66:101 1-1022. Gardner, M. J. a.nd D. G. Altman. 1986. Confidence intervals rather than P values: estimation rather than hypothesis testing. British Medical Journal 292:746- 750. Graham, C. H. and J. G. Blake. 2001. Influence of patch- and landscape-level factors on bird assemblages in a fragmented tropical landscape. Ecological Applica- tions 1 1: 1709-1721. Greenlaw, J. S. 1996. Eastern Towhee (Pipilo ery- throphthalmus). The birds of North America. Number 262. Grundel, R. and N. B. Pavlovic. 2007. Distinctiveness, use, and value of midwestern oak savannas and woodlands as avian habitats. Auk 124:969- 985. Hall, G. A. 1996. Yellow-throated Warbler (Dendroica dominica). The birds of North America. Number 223 Hamel, P. B. 1992. Land manager’s guide to birds of the south. The Nature Conservancy, Southeast Region, Chapel Hill, North Carolina, USA. Hodges, M. F. and D. G. Krementz. 1996. Neotropical migratory breeding bird communities in riparian forests of different widths along the Altamaha River, Georgia. Wilson Bulletin 108:496- 506. Hunter, W. C., L. H. Peoples, and J. A. Collazo. 2001. Partners in flight bird conservation plan for the south Atlantic Coastal Plain (physiographic area #03). American Bird Conservancy, The Plains, Virginia, USA. Johnson, A. S., H. O. Hillestad, S. F. Shanholtzer, and G. F. Shanholtzer. 1974. An ecological survey of the coastal region of Georgia. USDI, National Park Service Scientific Monograph Number 3. Kirk, D. A. and K. A. Hobson. 2001. Bird-habitat relationships in jack pine boreal forest. Forest Ecology and Management 147:217-243. McCarty, J. P. 1996. Eastern Wood-Pewee (Contopus Virens). The birds of North America. Number 245. McCune, B. and j. B. Grace. 2002. Analysis of ecological communities. MJM Software Design, Gleneden Beach, Oregon, USA. Moldenhauer, R. R. and D. j. Regelski. 1996. Northern Parula (Parula americana). The birds of North America. Number 215. Morissette, j. L., T. P. Cobb, R. M. Brigham, and P. C. James. 2002. The response of boreal forest songbird communities to fire and post-fire harvest- ing. Canadian Journal of Forest Re.search 32:2169- 2183. Mouillot, D., j. M, Culioli, and T. D. Chi. 2002. Indicator species analysis as a test ol non-random distribution of species in the context of marine protected areas. Environmental Conservation 29:385- 390. Nisbet, I. C. T. 2001. Detecting and measuring senescence in wild birds: experience with long-lived seabirds. Experimental Gerontology 36:833-843. Norvell, R. E., E. P. Howe, and J. R. Parrish. 2003. A seven-year comparison of relative-abundance and distance-sampling methods. Auk 120:1013- 1028. PaCIFICI, K., T. R. StMONS, AND K. H. POLLOCK. 2008. Effects of vegetation and background noise on the detection process in auditory avian point-count surveys. Auk 125:600-607. Petit, L. J. 1999. Prothonotary Warbler (Protonotaria citreci). The birds of North America. Number 408. Post, W. and J. S. Greenlaw. 1994. Seaside Sparrow (Ammodramus maritimus). The birds of North Amer- ica. Number 127. Ralph, C. J., G. R. Geupel, P. Pyle, T. E. Martin, and D. F. DeSante. 1993. Handbook of field methods for monitoring landbirds. USDA, Forest Service, General Technical Report PSW-GTR-144. Pacific Southwest Research Station, Albany, California, USA. Rodewald, P. G. and R. D. James. 1996. Yellow-throated Vireo {Vireo flavifrons). The birds of North America. Number 247. Rodewald, P. G., J. H. Withgott, and K. G. Smith. 1999. Pine Warbler {Dendroica pinus). The birds of North America. Number 438. Rosenstock, S. S., D. R. Anderson, K. M. Giesen, T. Leukering, and M. F. Carter. 2002. Landbird counting techniques: current practices and an alterna- tive. Auk 1 19:46-53. Simons, T. R., M. W. Alldredge, K. H. Pollock, and J. M. Wettroth. 2007. Experimental analysis of the auditory detection process on avian point counts. Auk 124:986-999. Smith, K. G., J. H. Withgott, and P. G. Rodewald. 2000. Red-headed Woodpecker (Melanerpes eiyrhro- cephalus). The birds of North America. Number 518. Springborn, E. G. and J. M. Meyers. 2005. Home range and survival of breeding Painted Buntings on Sapelo Island, Georgia. Wildlife Society Bulletin 33:1432- 1439. Stevenson, H. M. and B. H. Anderson. 1994. The birdlife of Florida. University Press of Florida. Gainesville, USA. Thomas, L. 1996. Monitoring long-term population change: why are there so many analysis methods? Ecology 77:49-58. Thomas, L., J. L. Laake, S. Strindberg, E. E. C. Marques, S. T. Bucki.and. D. L. Borchers, D. R. ANDER.SON, K. P. Burnham. S. L. Hedley, J. H. Pollard, J. R. B. Bishop, and T. A. Marques. 2006. DISTANCE 5.0 Release 2. 544 THE WILSON JOURNAL OL ORNITHOLOGY • Voi 122, No. 3, September 2010 Research Utrit for Wildlife Population Assess- ment, University of St. Andrews, United Kingdom. Thompson, W. L. 2002. Towards reliable bird surveys: accounting for individuals present but not detected. Auk 119:18-25. Whitehead, D. R. and T. Taylor. 2002. Acadian Flycatcher (Empidonax virescens). The birds of North America. Number 614. WiTHGOTT, J. H. AND K. G. SMITH. 1998. Brown-headed Nuthatch {Sitta pusilla). The birds of North America. Number 349. The Wilson Jounml oj Ornithology 122(3):545-555, 2010 INFLUENCE OF COVER AND EOOD RESOURCE VARIATION ON POST-BREEDING BIRD USE OF TIMBER HARVESTS WITH RESIDUAL CANOPY TREES MOLLY E. MCDERMOTT'- AND PETRA BOHALL WOOD' ABSTRACT. — We investigated avian use of clearcuts and two-age harvests during the post-breeding period in 2006 in the central Appalachians, West Virginia, USA with an intormation-theoretic approach to model selection. Cover variables appeared to be most important; e.g., vegetative vertical complexity had a strong positive relation with capture rates of mature forest birds and molting adults, as well as physical condition which supports a predator-avoidance hypothesis for habitat use. Basal area was a poor predictor of captures; residual trees near nets tended to depress capture rates. Food variables best explained capture rates for some species groups (e.g., early-successional insectivores and granivores, mature forest nesting adults, molting birds), but post-breeding habitat quality was based primarily on vegetative cover. Habitat use may depend on the bird s physical condition and molt status, and we found evidence for age-specific differences which may impact survival. Our study suggests important links between post-breeding habitat quality, molt status, physical condition, and bird age, and indicates a variety of response variables (relative abundance, survival, body condition) should be measured to assess avian habitat quality during the post-breeding period. Received 19 March 2009. Accepted 6 February 2010. Recent declines in populations of migratory birds that nest in both mature forest and early- successional habitats have stimulated interest in limiting factors and potential mortality sources throughout the year (Robbins et al. 1989, Hunter et al. 2001). Habitat loss and fragmentation in breeding and wintering areas are probable factors (Askins et al. 1990), but little is known about factors affecting survival during the post-breeding period (Baker 1993, Pagen et al. 2000), the interval between fledging of young and migration. Many bird species that nest in early-successional or mature forest habitats use early-successional habitat extensively during post-breeding. For example, many species of mature forest breeding .songbirds move into nearby shrublands after young have fledged (Pagen et al. 2000, Marshall et al. 2003, Vitz and Rodewald 2007). Juvenile Wood Thrush (Hylocichla mustelina) and Swain- son’s Thrush (Catharus ustulatus) are also known to move >1 km from their mature forest natal area into early-successional habitat before migra- tion (Anders et al. 1998, Vega Rivera et al. 1998, White et al. 2005, Dellinger 2007). One explanation for this behavior is predator avoidance; early-successional habitats may pro- vide increa,sed protection from predators (White et 'West Virginia Cooperative Fish and Wildlife Research Unit, P. O. Box 6125, 322 Percivai Hall, Division of Forestry and Natural Resources. West Virginia University, Morgantown, WV 26506, USA. ^Corresponding author; e-mail: mollyemcderrnott@gmail.com al. 2005). A dense understory provides cover for fledglings that, lacking the knowledge to detect and escape predators, often have high post- fledging mortality (Sullivan 1989, Anders et al. 1997). Adults may also seek protection and cover during the post-breeding period as they undergo pre-basic molt and are temporarily more vulner- able to predators (Pagen et al. 2000). More understory vegetation and increased vertical structure were associated with increased survival for fledgling Ovenbirds {Seiurus aurocapilla) (Vitz 2008), even where prey abundance was low (King et al. 2006). A second explanation is the re,source-selection, or optimal-foraging hypothesis (i.e., bird use of early-successional habitats is driven by greater food resource abundance) (White et al. 2005). Abundant food resources are needed by both adults and juveniles for accumulation of fat reserves in preparation for migration and comple- tion of molt (Murphy and King 1992, DeGraaf and Yamasaki 2003). Molt is nutritionally de- manding and requires energy and protein not only for synthesizing feathers, but for increasing metabolism to offset reduced insulation and flight efficiency (Jenni and Winkler 1994). Juvenile Swainson’s Thrush u.se of coastal scrub during post-fledging was best explained by fruit abun- dance variables, supporting the resource-selection hypothesis (White et al. 2005). Two-age harvesting or green-tree retention, where canopy trees are retained to more closely resemble natural disturbances, is becoming more common (Franklin et al. 1997, Rose and Muir 545 546 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 3, September 2010 1997, Vanha-Majamaa and Jalonen 2001, Vergara and Schlatter 2006), and is frequently used as an alternative to clearcut harvests in central Appala- chian hardwood forests (Miller et al. 2006). Mature residual trees and a regenerating age class provide vertical stratification for both shrub and canopy-associated birds (Duguay et al. 2000). The effects of two-age harvesting on breeding birds have been demonstrated (Baker and Lacki 1997, Boardman and Yahner 1999, Duguay et al. 2000, McDermott and Wood 2009), but no studies of which we are aware have examined avian use of stands with residual tree retention during the post- breeding period. The objectives of this study were to; (1) quantify how vegetative structure and food availability in timber harvests with a gradient of residual canopy trees affect post-breeding bird habitat use, and (2) examine the impacts of resource variation on avian body condition and molt status. Our study is unique in that we examine both mature forest and early-succession- al nesting bird species, investigate multiple types of food resources at the nest-level, and integrate molt and body condition variables. METHODS Study Areas. — Our study was conducted in West Virginia on the Monongahela National Forest (MNF) in Pocahontas and Tucker counties (38° 31' N, 79° 44' W), and the Wildlife and Ecosystem Research Forest (WERF) in Randolph County (38° 43' N, 80° 03' W). These managed forests in the central Appalachian Mountains are characterized by narrow, low valleys dissected by northeast-southwest ridges. We selected nine stands with a gradient of retained basal area ranging from 0 to 7 m^/ha for sampling. Stand age (4-7 years post-harvest) was a selection criterion, because there were few harvests in an early-successional stage. We chose stands on the MNF {n = 5) that were in close proximity to those on the WERF (/? = 4). Harvest area was 4.2-21 ha and elevation was 630- 1,090 m. The principal forest types were northern Allegheny hardwoods, cove hardwoods, mixed mesophytic, and oak-hickory (Quercus spp.- Carya spp.). Dominant overstory tree species included red oak (Q. rubra), red maple (Acer ruhruni), chestnut oak (Q. primis), and yellow poplar (Liriodendron tulipifera). The most abun- dant understory plants were blackberry (Ruhus spp.), black birch (Betula lenta), greenbrier {Smilax spp.), American beech (Fagus grand- ifolia), maples, and yellow poplar. Mist-Netting. — We used mist nets to sample bird habitat use because surveying birds by vision is difficult in dense vegetation, and many birds are secretive during the post-breeding period (Ralph et al. 2004). Our goal was to relate bird capture rates to microhabitat food and cover variables. We placed 10 mist nets (12 X 2.6 m with 30 mm mesh) in each stand, often along skid roads created during timber harvest and usually oriented perpendicular to the prevailing aspect. We arranged nets so that distance from edge was variable; a minimum of 20 m separated ends of nets, and each net was within 60 m of another net. We first sampled stands randomly and conducted repeat visits in the same order at least 3 weeks apart. A stand (n = 3) from each harvest type (clearcut, low-leave two-age, and high-leave two- age) was randomly chosen for the first round, and alternated for each sampling round. We sampled 2 consecutive days at each stand and visited each twice from 25 June to 15 August 2006. We opened nets at sunrise and closed after 4-5 hrs except in inclement weather, when mist nets were not operated. We checked nets every 30 min and recorded species, age, gender, wing chord, mass, fat, breeding condition (extent of brood patch or cloacal protuberance), molt status, and time of capture for each captured individual. Captured birds, except for Ruby-throated Hum- mingbirds {Archilochus colubris), were fitted with uses aluminum bands. We filled out molt cards per the British Trust for Ornithology for actively molting adults to obtain pre-basic molt scores. Food Resource Sampling. — We sampled fruits twice in each stand, once per round immediately after the first day of mist-netting. We obtained an index of fruit abundance by counting individual fruits along 12-m transects within 1 m of either side of each mist-net lane (Levey 1988). All fruits were counted, identified to species, and classified as unripe or ripe. We also collected arthropod samples, concur- rently with fruit sampling, to obtain biomass availability estimates (Hutto 1990, Smith and Rotenberry 1990). We used biomass as an index of food availability, not an absolute measure' of food consumed by birds (Hutto 1990, Duguay et al. 2000). We obtained four arthropod samples at each net location (// = 90), one from each of four saplings, each round. Saplings within 10 m of the McDevmon and W ood - POST-BREEDINCJ BIRD USE OF TIMBER HARVESTS 547 nets were beaten with a rubber mallet to dislodge arthropods (Schaulf 1997). This capture method was most suitable for sampling arthropods available to foliage-gleaning birds, the most common foraging guild captured in the pilot year (McDermott 2007). Samples were taken from two random tree saplings within mist-net height on each side of the net. We often used the same individual trees in the second round, as the time interval (~3 weeks) allowed sufficient time for repopulation (Linda Butler, Entomologist, West Virginia University, pers. comm.). We used a sturdy, light-colored sheet, 1 m“ in area to capture the arthropods. The sheet was held under the branches of one side of the sapling, and one person beat four limbs that were above the sheet from top to bottom (four beats per limb). Once specimens fell onto the sheet, small arthropods were collected with an aspirator into a glass vial of ethanol. We picked up larger insects with forceps and placed them into a vial. Specimens were stored in 70% ethanol. We identified all arthropods to Order, and classified each individual into one of three size categories: <3, 3-10, and >10 mm (Van Horne and Bader 1990). Specimens were subsequently placed on a tray and inserted into a drying oven at 60° C. We weighed each sample (by taxonomic Order, size, net, and round) to the nearest 0.1 mg to obtain an estimate of biomass after 48 hrs or when specimens achieved a constant mass. Arthropod richness at the Order level was calculated as a measure of arthropod diversity. Vegetation Sampling. — We recorded presence or absence of vegetative cover on four 12-m transects, 3 m and 6 m from the net paralleling each side of the net lane (Schemske and Brokaw 1981), at five equidistant intervals along each transect in height categories: 0-1 , >1-2 , >2-3 , >3-6 , >6-12 , >12-18 , >18-24 , and >24 m. A vertical complexity index (VCI) was calculated as a sum of all height cover class percentages divided by 100 and multiplied by a constant (Wood et al. 2005). We quantified the number of stems <7.6 cm dbh and percent ground cover categorized as leaf litter, woody, rock/bare ground, or herbaceous in two 2 X 2 m plots randomly placed on the tran.sect along each side of the net lane. We measured basal area with a 10- factor prism from the center of the net lane. Data Analyses. — We standardized captures for each species per 100 mist-net hrs, where 1 mist- net hr = 1 net open for 1 hr. Bird species that did not breed locally were excluded from analyses, and only first captures of individuals were used. We assigned species to two breeding habitat guilds: early-successional or mature forest ba.sed on Ehrlich et al. (1988) and observations from earlier studies in West Virginia (e.g., Duguay et al. 2000). Species were placed into foraging guilds (insectivorous or seasonally frugivorous) ba.sed on available life history data (e.g.. Birds of North America species accounts) and a cursory examination of fecal samples from captured birds. We performed separate analyses for hatch year (HY = born that spring) and adult (AHY >1 year of age) birds; juveniles are usually more suscep- tible to capture and may use habitats differently than adults. We occasionally caught family groups, but it was a small percent of overall captures (~5%; similar to Marshall et al. 2003), and these age groups were treated independently. We selected explanatory variables to model bird responses to resource availability that repre- sented the vegetative cover and food resources near nets, the main hypotheses for habitat use. We tested for pairwise correlation among all variables using Pearson’s product-moment correlation (Proc CORR; SAS Institute 2003). Percentage variables were arcsine transformed to approximate normal- ity. We reduced the number of explanatory variables within a category (fruit, arthropod, and vegetative cover) by dropping strongly coirelated variables {P > 0.7) to avoid multicollinearity. We selected the one that was less correlated with other model variables or most biologically meaningful when choosing between two coiTelated variables. Nine variables were retained for modeling (Ta- ble 1). We developed sets of a priori candidate models from these nine variables to explain avian capture rates and other metrics of interest based on a review of recent literature on post-breeding bird ecology. The 8-13 models developed for each species or group are described in McDermott (2007). Arthropod biomass was analysis-specific: prey size categories were chosen based on what is known of prey items consumed by each bird species. Total arthropod biomass included all prey sizes when multiple bird species were combined for a particular analysis. Response variables included capture rates of all adult and HY birds for species captured at >30% of nets or the top four species captured in each habitat guild, and capture rates of each age group classified by diet and breeding habitat (all 548 THE WILSON JOURNAL OL ORNITHOLOGY • Vd. 122, No. 3, September 2010 TABLE 1. Net-level vegetative and food resource variables used to model bird use of early-successional habitats during the post-breeding period in northcentral West Virginia, 2006. Resources were sampled concurrently with mist- netting from late-June through mid-August. Variable Code Value range -V ^ t SE Vegetative cover Understory stem density'' DENSE 34-282 121.42 ± 3.74 Residual basal area, m'/ha BA 0-11.48 3.02 ± 0.24 Vertical complexity index*’ VCI 42-117 74.92 ± 0.98 Fruit resources'’ Fruit abundance ERUIT 95-4,408 802.07 ± 41.57 Total ripe fruit RIPE 0-846 18.65 ± 5.29 Arthropod resources'* Arthropod richness, number of Orders RICH 5-12 8.33 ± 0.1 1 Total arthropod biomass, mg ARTH 15.3^07.9 100.04 ± 5.46 Lepidopteran larval biomass, mg LEP 0-373.3 16.19 ± 2.25 Seed resources Herbaceous ground cover, % SEED 3-100 47.6 ± 1.97 “ Number of .stems <7.6 cm dbh at two 2 X 2 m plots on either .side of each mist-net lane. ^ Vertical complexity index calculated as a .sum of all canopy cover class percentages divided by 100 and multiplied by 20. Fruit was measured along two 12 X I m transects on either side of each mist-net lane. Arthroptxis were sampled with a beating technique at 4 shrubs/net. seasonal frugivores and insectivores by habitat guild). We only analyzed age classes with >20 captures. A small percentage of adults met strict criteria for post-breeding (only one-third were not in full breeding condition and were actively molting); we included all adults captured during this period including late-breeders (30%) for both breeding habitat guilds. We analyzed molt score index and body condition in addition to the individual species and groups. Adult molt score (0-250; Cherry and Canned 1984) was reclassified to values of 0-6. We did not analyze molt .score as continuous because birds with a score of 0 or 250 are not molting flight feathers; these scores were reclassified as 0 because these birds would be similarly capable of flight. Adults in the heaviest stage of molt (.score = 100-140) are most likely to have compromised flight and higher energetic demands (Jenni and Winkler 1994); they received the highest recla.s- sified value (6). Each HY individual was given a molt score of 0-3 depending on the number of feather tracts in pin. Body condition was calculated by dividing mass (g) by wing length (mm) to account for body size (Winker 1995). Mass was regressed against wing length for each species, and the residuals were u.sed as the condition index. These values were reclassified to a score of 0^. We used Poisson regression (Proc NLMIXED; SAS Institute 2003) to model each response variable except body condition, which had a nomial distribution and was analyzed with linear regression. This procedure allows for a hierarchical (i.e., nested) approach to account for within-stand variability, and both Poisson and linear distribu- tions can be modeled. Analyses were conducted at the net scale because of variability in food and vegetative resources within stands and to examine microhabitat use patterns. We used an information-theoretic approach to select the best approximating models from each set of candidate models. The global model for each response variable was constructed using all variables (range = 7 to 9) relevant to each species or group. We used Akaike’s Information Criterion adjusted for small sample sizes (AlC,..) regardless of the sample sizeiparameter ratio, because second-order AIC,. converges to AIC as n becomes large (Burnham and Anderson 2002). We did not use QAIC to re-rank models, instead including a random effect in the regression models to account for any dispersion problems in the data (SAS Institute 2003). We re.scaled the resulting information criteria for each candidate model to obtain A, AIC^. (differences between the top model and every other model) and calculated Akaike weights (h’,) to better interpret the relative McDermott ami Wood • POST-BREEDING BIRD USE OF TIMBER HARVESTS 549 TABLE 2. Number ot iiuliviiluals captured for individual .species or groups of species analyzed (in bold) in regression models. Birds were captured by mist nets in northcentral West Virginia, June-August 2006. Species Foraging guild' Total captures Captures’’ AHY HY Early-successional species Chestnut-sided Warbler I 309 123 186 Common Yellowthroat I 42 26 16 Gray Catbird F 75 39 36 Indigo Bunting I 63 41 22 Early-successional insectivores I 520 245 275 Mature-forest species American Redstart I 58 26 32 Canada Warbler I 41 18 23 Hooded Warbler I 49 34 15 Red-eyed Vireo I 59 51 8 Veery F 68 39 29 Mature-forest insectivores I 440 227 213 Frugivores F 225 136 89 “ Foraging guilds analyzed were: I = primarily insectivores post-breeding and F = seasonal frugivores. HY = hatch year (juveniles), AHY = after hatching year (adults). likelihood of a model. Models with a A,- AIC < 2 were given equal consideration as competing models. RESULTS We captured 1,189 individuals: 579 HYs and 610 adults of 15 early-successional and 38 mature forest species. Insectivores predominated; season- al frugivores comprised 20% of species and individuals captured. Total captures per species (by age) or group of species analyzed ranged from 22 to 275 (Table 2). Early-Successioncil Bird Habitat Use. — The same cover and food variables were important for both age groups of early-successional insec- tivores (Table 3). Capture rates were negatively related to vertical complexity (Fig. 1 ), understory stem density, and basal area. Age groups had opposite relationships to food variables. Adults were positively related to arthropod biomass (Fig. 2) and negatively to Lepidopteran larval biomass, the opposite of HYs. We analyzed four early-successional bird species including three in.sectivores (Table 2). Chestnut-sided Warbler (Dendroica pensylvanica) captures were poorly predicted by all variables (Table 3). Five models with different variables had a similar likelihood of being selected for adults, while HY capture rates were best explained by the global model. Associations were the same as for all early-successional HY captures pooled because this species comprised 67% of these captures. A negative association with residual basal area best explained Common Yellowthroat adult {Geothlypis trichas) captures (Table 3, Fig. 3). Captures of adult Indigo Buntings (Passerina cyanea) were best explained by increases in vertical complexity. Several models competed to explain captures of HY Indigo Buntings. Of these, models with vertical complexity and basal area (negative relationships) had the highest collective weights. Models containing food variables were impor- tant for some species (Table 3). Adult Indigo Bunting captures were positively associated with herbaceous ground cover, which may represent availability of seeds. Lepidopteran larval biomass and total arthropod biomass were positive predic- tors of Chestnut-sided Warbler HY and adult captures, respectively. HY Indigo Bunting cap- tures were positively associated with Lepidopter- an larval biomass. We analyzed one seasonal frugivore in the early-successional habitat guild: Gray Catbird {Duinetella carolinensis). Ripe fruit was nega- tively related to Gray Catbird capture rates (Table 3), and basal area and stem density were positively related to adult captures. The global model best explained HY catbird captures; vertical complexity and basal area were positively related to captures whereas nearly all food variables were negatively related. 550 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 3. September 2010 TABLE 3. Regression analysis of bird capture rates by species, guild, and age group in northcentral West Virginia, 2006. Relationships with explanatory variables in competing models are presented. Response variable Explanatory variables Early-successional species Chestnut-sided Warbler, adults Chestnut-sided Warbler, HY Common Yellowthroat, adults Gray Catbird, adults Gray Catbird, HY Indigo Bunting adults Indigo Bunting HY LEP (-), ARTH (-b), BA {+). DENSE (+), VCl (-) GLOBAL"* BA (-) BA {+), VCI (+), RIPE (-), DENSE (+) GLOBAL VCl (-0, SEED (-b), RICH (-b), ARTH (-b) BA (-), VCl (-), DENSE (-), RICH (-), LEP (-b), SEED (-), ARTH (-) Early-successional insectivores Adults HY ARTH (-b), LEP (-), VCl (-), BA (-), DENSE (-) GLOBAL Mature forest species American Redstart, adults American Redstart, HY Canada Warbler, HY Hooded Warbler, adults Red-eyed Vireo, adults Veery, adults Veery, HY DENSE (-), ARTH (-b), VCl (-b) VCI (-b), ARTH (-), DENSE (-b), LEP (-b), BA (-b) BA (-), ARTH (-), DENSE (-), VCI (-) ARTH (-b), DENSE (-b), LEP (-b) BA (-), RICH (-b), DENSE (-), VCI (-b) ARTH (-b). FRUIT (-b) VCI (-), RIPE (-b) Mature forest insectivores Adults HY ARTH (-b), DENSE (-), VCl (+). LEP (-b), BA (-) GLOBAL All frugivores Adults HY BA (-b), RIPE (-), DENSE (-) FRUIT (-b), VCI (-) RIPE (-b), BA (+), DENSE (-), VCI (-b). FRUIT (-b) Molt index Adults HY RIPE (-b), ARTH (-), VCI (-b), BA (+), FRUIT (-b), LEP (-b), DENSE (-b) FRUIT (-b), ARTH (-b) Body condition index DENSE (-), LEP (-), VCI (-b), FRUIT (-) “ The global model, in which all variables were u.sed, wa.s the only competing model. Mature Forest Bird Habitat Use. — Adult insec- tivore capture rates were best explained by both cover and food variables (Table 3). Capture rates increased with increasing vertical complexity (Fig. 1), total arthropod biomass (Fig. 2), and Lepidopteran larval biomass, and decreased with increasing basal area (Fig. 3) and stem density. The global model best predicted capture rates of HYs. Cover variables had the same relationships as with adults, but arthropod biomass was inversely related to captures of HYs. Four of five species we analyzed were primary in.sectivores (Table 2). Cover variables dominated the top models for these species. Vertical complexity was one ot the top explanatory variables for both HY and adult American Redstart (Setophaga ruticilla) captures; (positive relationship). Red-eyed Vireo (Vireo olivaceus) adults were also positively associated with vertical complexity. The basal area model re- ceived the most support for captures of HY Canada Warblers (Wilsonia canadensis) and adult Red-eyed Vireos (Table 3) showing a negative relationship for these species (Fig. 3). The stem density model received a weight of 13% for Red- eyed Vireo adult captures and 20% for Hooded Warbler (Wilsonia citrina) adult captures, but showed opposite patterns: Red-eyed Vireo capture rates declined while Hooded Warbler rates increased with increasing understory stem density. Food variables were positively associated with captures of several species (Table 3): Lepidopter- McDermott am! Wood • POST-BREEDING BIRD USE OF TIMBER HARVESTS 551 (A) “liarly-successional HY - Early-successional adults • Mature forest HY -Mature forest adults Vertical complexity index -Body condition index ^ ^ — Adult molt index 40 60 80 100 120 Vertical complexity index FIG. 1. Relationship of vertical complexity of the vegetation in early-successional habitats of northcentral West Virginia in 2006 with (A) capture rate (captures/100 net hrs) of adult and hatch year (HY) birds of mature-forest and early-successional insectivorous species, and (B) body condition and molt index of captured adults. an larvae with adult Hooded Warbler and HY American Redstart captures; arthropod richness with Red-eyed Vireo adult captures; and arthro- pod biomass with American Redstart and Hooded Warbler adults. The Veery (Catharus fuscescens) was the only seasonal frugivore analyzed in the mature forest habitat guild. Adult Veery capture rate was FIG. 2. Captures/100 net hrs as a function of arthropod biomass for all mature-forest and early-successional adult insectivores, adult Veeries (seasonal frugivores), and early- successional hatch year (HY) birds in early-successional habitats of northcentral West Virginia in 2006. FIG. 3. Relationship between capture rate (captures/ 100 net hrs) and basal area for HY and adult frugivores, mature forest adults. Red-eyed Vireo adults, Canada Warbler hatch year (HY) birds, and Common Yellowthroats in early-successional habitats of northcentral West Virginia in 2006. positively associated with arthropod biomass (Fig. 2) and total fruit (Table 3). The model with vertical complexity (negative association) was ranked only 1.5 times better than that with ripe fruit (positive association) for HYs. However, these relationships were weak suggesting inade- quate model structure. Frugivore Habitat Use. — Nine species from different breeding habitat associations represented the seasonal frugivores. Total fruit abundance was positively associated with both age groups but was not a strong predictor in top models (Table 3). Age groups showed opposite relation- ships with ripe fruit: negative for adults and positive for HYs. The model with basal area (positive association) received 33% of the weight for adults, and basal area was the highest ranked cover parameter when considering all HY bird models (Fig. 3). Molt Score. — Molting HYs were positively associated with total fruit abundance (Table 3). This model was —2.5 times as likely to be selected as the model with arthropod biomass {w = 0.11). Abundance of ripe fruit best explained molt score of adults (vv = 0.69), but the global model also received support. All cover variables in the global model were positively related to adult molt, and vertical complexity was the highest ranked cover variable when considering all models. Complex vertical structure (a high index value) was associated with individuals in the most compromising stage of molt (Fig. 1 ). Body Condition Index. — Cover variables best explained variation in body condition index (Table 3). Low stem density (w = 0.15) and high 552 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 3, September 2010 vertical complexity {w — 0.15; Fig. 1) best predicted high body condition index values. Arthropod and fruit variables were negatively related to body condition. DISCUSSION Predator- Avoidance Hypothesis. — Cover vari- ables were important for post-fledging juveniles and adults using timber harvests. Birds may be selecting habitats with increased vegetative com- plexity that offers protection from predators. King et al. (2006) found that fledgling Ovenbirds had higher survival rates where understory vegetative structure was greatest. Thus, vertical complexity may be a good representation of protective cover. Thicker cover as measured by understory stem density had a negative influence on bird captures. Similarly, Vitz and Rodewald (2007) found a negative relationship between mature forest bird capture rates and density of shrub vegetation. Birds in our study were often traveling in foraging flocks that contained several species, and an open understory would facilitate movement of individ- uals. Adult Chestnut-sided Warblers, Gray Catbirds, and Hooded Warblers, species that nest in the harvested areas (McDermott 2007), had higher capture rates at nets with more understory stems. Thus, a higher understory stem density could serve as protective cover during nesting and post- breeding for these species. Early-successional bird habitat use outside of breeding tends to be less flexible than for those species nesting in mature forest (Rodewald and Brittingham 2004). Stand- level variables such as area and amount of edge (McDermott 2007, Schlossberg and King 2008) may explain more variation in habitat use by those species nesting in shrub-scrub. Residual trees in two-aged stands inhibit understory growth and limit stem density (Miller et al. 2006), but leave-trees provide raptor perches and residual trees may not offer predator protec- tion. This suggestion is supported by our data as captures were negatively associated with basal area; within stands, captures were greater at nets with fewer nearby trees. Resource-Selection Hypothesis. — This study provides evidence that post-breeding habitat use is influenced by food resource availability. For example, arthropod biomass positively predicted captures of adult insectivores and Veeries, which, like other seasonal frugivores, also consume arthropods. Lepidopteran larval biomass was a positive predictor of mature forest (e.g., Canada and Hooded warblers) and early-successional HY captures. Lepidopteran larvae are an important food source for nestling and adult songbirds (Van Horne and Bader 1990, Jones et al. 2003). Shrub- dwelling Lepidopteran larvae may have higher biomass in clearcuts than in mature forest because deciduous shrubs have high amounts of foliage (Van Horne and Bader 1990, Keller et al. 2003). Thus, some birds, especially mature forest indi- viduals, may be keying on this rich energetic source within early-successional habitats. Arthro- pods may be a fleeting resource because biomass is significantly higher in young clearcuts when they become fully stocked (by year 6 or 7) versus later serai stages (Keller et al. 2003). We found few strong relationships with total fruit abundance. Similarly, Vitz (2008) did not find evidence that mature forest birds use clear- cuts for fruit resources. White et al. (2005) found a positive relationship between thrush habitat use and fruit abundance, but we found only a weak positive relation for the Veery, the only thrush species analyzed individually in our study. We included herbaceous cover as a crude index of seed availability for seed-eating birds in the models. Indigo Bunting captures, particularly adults, were explained best by herbaceous cover. Thus, herbaceous plants may serve as a cue for seed availability for granivores. We found limited evidence for the resource selection hypothesis. We do not know accurately what constitutes available food for a bird without behavioral observations to quantify food resource selection (Hutto 1990). Foliage-associated arthro- pod biomass may have misrepresented available food, despite adjusting for potential prey size differences. Birds may not be able to key in on food resources at the net scale because of high within-stand variability in arthropod communities. We do not know how or if food is limiting during the post-breeding period. Food limitation likely varies from year to year and is more likely to occur during breeding because of food require- ments of rapidly growing nestlings (Nagy and Holmes 2005). Food and Cover Related to Condition, Molt, and Age. — The physical condition of a bird in theory is positively related to food resources and deposited body fat. However, food resource availability in our study was negatively associated with body condition of captured birds. Instead, a high index value was associated with high vertical McDermott and Wood - POST-BREEDING BIRD USE OF TIMBER HARVESTS 553 complexity. One explanation is that birds with a higher body condition index may use the most structurally complex habitat, which we infer offers the best protection and increases likelihood of survival. Birds in poor condition may select habitat with rich food resources to replenish fat reserves, while birds in better condition select microhabitat that provides protection from pred- ators (Moore and Aborn 2000). Habitat use may depend on physical condition, and both cover and food resources are important for post-breeding survival, which is influenced by body condition of fledglings in some species (Krementz et al. 1989, Naef-Daenzer et al. 2001 ). There was no evidence that body condition increased over the season; we sampled prior to pre-migratory fattening, and the mean condition index decreased over time due to a concomitant increase in HY captures. Age-specific differences were evident for most of the groups analyzed, particularly with food variables. Adults had a higher body condition index than juveniles for most species analyzed (MEM, unpubl. data), which may have contribut- ed to these differences. HY capture rates were negatively associated with arthropod biomass, while adult capture rates had a positive relation- ship. HYs are likely inexperienced in recognizing arthropod availability cues and may have a lower foraging efficiency than adults (Stevens 1985, Weathers and Sullivan 1989, Vanderhoff and Eason 2008) and thus, lower survival. These data bolster the argument that adults and HYs may be keying on different features within a stand, which may account for age-specific survival rates during the post-breeding period. Molting birds likely take advantage of both food and cover. Adults in heavy molt with diminished flight capabilities were associated with high vertical complexity, which likely provides protection from predators. Molting birds were positively associated with fruit variables. Changes in food choice have been ob.served in the field for molting passerines, and food availability may influence the timing and place of molt (Jenni and Winkler 1994). Increased energy require- ments during molt could be fulfilled by an abundance of fruits for the 13% of molting adults and 15% of molting Juveniles that were frugivo- rous or, for insectivores, an abundance of arthropods that are attracted to fruits (Sallabanks and Courtney 1992). Early-successional habitats, including those retaining canopy trees such as the two-age harvests we sampled, should be recognized for their conservation value. Avian survival may be compromised by ignoring bird habitat require- ments outside of breeding (King et al. 2006). Habitat use does not imply habitat quality; a variety of parameters (relative abundance, surviv- al, body condition) should be measured to evaluate post-breeding habitat quality for birds. ACKNOWLEDGMENTS This work was funded by the USDA Forest Service, Monongahela National Forest. The U.S. Geological Survey, West Virginia Cooperative Fish and Wildlife Research Unit provided field vehicles, equipment access, and logistical support. West Virginia University, USDA Forest Service, and P. D. Keyser ahso provided logistical support. We thank MeadWestvaco for allowing access to the Wildlife and Ecosystem Research Forest and USDA Forest Service for access to sites on the Monongahela National Forest. We are grateful to K. G, Fleyden, P. M. McElhone, J. S. Rehar, E. E. Samargo, M. B. Shumar, and J. C. Walker for help collecting data in the field and P. M. McElhone and M. B. Shumar for help sorting insects in the laboratory. We thank P. J, Herron for allowing us to work under his master banding permit. J. T. Anderson, C. M. Johnson, and Sue Olcott provided comments on an earlier version of this manuscript. LITERATURE CITED Anders, A. D., J. Faaborg, and F. R. Thompson III. 1998. Postfledging dispersal, habitat use. and home-range size of juvenile Wood Thrushes. Auk 1 15:349-358. Anders, A. D., D. C. Dearborn. J. Faaborg, and F. R. Thompson 111. 1997. Juvenile survival in a population of neotropical migrant birds. Conservation Biology 11:698-707. Askins, R. a., j. F. Lynch, and R. Greenberg. 1990. Population declines in migratory birds of eastern North America. Current Ornithology 7:1-57. Baker, M. D. and M. J. Lacki. 1997. Short-term changes in bird communities in response to silvicultural prescriptions. Forest Ecology and Management 96:27-36. Baker, R, R, 1993. The function of post-lJedging exploration: a pilot study of three species of pas.serines ringed in Britain. Ornis Scandinavica 24:71-79. Boardman, L. a. and R. H. Yahner. 1999. Wildlife communities associated with even-aged reproduction stands in two stale forests of Pennsylvania. Northern Journal of Applied Forestry 16:89-95. Burnham, K. P. and D. R. Ander.son. 2002. Model .selection and multimodcl inference: a practical- theoretic approach. Second Edition. Springer-Verlag. New York, USA. Cherry, J. D. and P. F. Cannell. 1984. Rate and timing of prebasic molt of adult Boreal Chickadees. Journal of Field Ornithology 55:487^89. DeGraaf, R. M. and M. Yama.saki. 2003. Options for managing early-successional forest and shrubland bird 554 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 3, September 2010 habitats in the northeastern United States. Forest Ecology and Management 185:179-191. Dellinger, T. A. 2007. Post-fledging ecology and survival of neotropical migratory songbirds on a managed Appalachian forest. Thesis. West Virginia University, Morgantown, USA. Duguay, J. P., P. B. Wood, and G. W. Miller. 2000. Effects of timber harvests on invertebrate biomass and avian nest success. Wildlife Society Bulletin 28:1123- 1 131. Ehrlich, P. R., D. S. Dobkin, and D. Wheye. 1988. The birder’s handbook. Simon and Schuster, New York, USA. Franklin, J. F., D. R. Berg, D. A. Thornburgh, and J. C. Tappeiner. 1997. Alternative silvicultural approaches to timber harvesting: variable retention harvest systems. Pages 111-139 in Creating a forestry for the 21st Century (K. A. Kohm and J. F. Franklin, Editors). Island Press, Washington, D.C., USA. Hunter, W. C., D. A. Buehler, R. A. Canterbury, J. L. Confer, and P. B. Hamel. 2001. Conservation of disturbance-dependent birds in eastern North America. Wildlife Society Bulletin 29:440-455. Hutto, R. L. 1990. Measuring the availability of food resources. Studies in Avian Biology 13:20-28. Jenni, L. and R. Winkler. 1994. Moult and ageing of European passerines. Academic Press Limited, Lon- don, United Kingdom. Jones, J., P. J. Doran, and R. T. Holmes. 2003. Climate and food synchronize regional forest bird abundances. Ecology 84:3024-3032. Keller, J. K., M. E. Richmond, and C. R. Smith. 2003. An explanation of patterns of breeding bird species richness and density following clearcutting in north- eastern USA forests. Forest Ecology and Management 174:541-564. King, D. 1, R. M. Degraaf, M. L. Smith, and J. P. Buonaccorsi. 2006. Habitat selection and habitat- specific survival of fledgling Ovenbirds (Seiurus auroeapilla). Journal of Zoology 269:414-421. Krementz, D. G., j. D. Nichols, and J. E. Hines. 1989. Post-fledging survival of European Starlings. Ecology 70:646-655. Levey, D. 1988. Spatial and temporal variation in Costa Rican fruit and fruit-eating bird abundance. Ecological Monographs 58:251-269. Marshall, M. R., J. A. Dececco, A. B. Williams, G. A. Gale, and R. J. Cooper. 2003. Use of regenerating clearcuts by late-successional bird species and their young during the post-fledging period. Forest Ecology and Management 183:127-135. McDermott, M. E. 2007. Breeding and post-breeding forest bird community dynamics in regenerating clearcuts and two-age harvests in the central Appala- chians. Thesis. West Virginia University, Morgan- town, USA. McDermoit, M. E. and P. B. Wood. 2009. Short- and long-term implications of clearcut and two-age silviculture for con.servation of breeding fore.st birds in the central Appalachians, USA. Biological Conser- vation 142:212-220. Miller, G. W., J. N. Kochenderfer, and D. B. Fekedulegn. 2006. Influence of individual reserve trees on nearby reproduction in two-aged Appalachian hardwood stands. Forest Ecology and Management 224:241-251. Moore, F. R. and D. A. Aborn. 2000. Mechanisms of en route habitat selection: how do migrants make habitat decisions during stopover? Studies in Avian Biology 20:34^2. Murphy, M. E. and J. R. King. 1992. Energy and nutrient use during moult by White-crowned Sparrows, Zonotrichia leiicophrys gcimbelii. Ornis Scandinavica 23:304-313. Naef-Daenzer, B., F. Widmer, and M. Nuber. 2001. Differential post-fledging survival of Great and Coal tits in relation to their condition and fledging date. Journal of Animal Ecology 70:730-738. Nagy, L. R. and R. T. Holmes. 2005. Food limits annual fecundity of a migratory songbird: an experimental study. Ecology 86:675-681. Pagen, R. W., F. R. Thompson III, and D. E. Burhans. 2000. Breeding and post-breeding habitat use by forest migrant songbirds in the Missouri Ozarks. Condor 102:738-747. Ralph, C. J., E. H. Dunn, W. J. Peach, and C. M. Handel. 2004. Recommendations for the use of mist nets for inventory and monitoring of bird populations. Studies in Avian Biology 29:187-196. Robbins, C. S., J. R. Sauer, R. S. Greenberg, and S. Droege. 1989. Population declines in North American birds that migrate to the Neotropics. Proceedings of the National Academy of Science of the USA 86:7658-7662. Rodewald, P. G. and M. C. Brittingham. 2004. Stopover habitats of landbirds during fall: use of edge- dominated and early-successional forests. Auk 121:1040-1055. Rose, C. R. and P. S. Muir. 1997. Green-tree retention: consequences for timber production in forests of the western Cascades. Ecological Applications 7:209-217. Sallabanks, R. and S. P. Courtney. 1992. Frugivory, seed predation, and insect-vertebrate interactions. Annual Review of Entomology 37:377^00. SAS Institute. 2003. Version 9.1. SAS Institute Inc., Cary, North Carolina, USA. SCHAUFF, M. E. 1997. Collecting and preserving insects and mites: techniques and tools. NHB-168. USDA, Sys- tematic Entomology Laboratory, National Museum of Natural History. Washington, D.C., USA. SCHEMSKE, D. W. AND N. Brokaw. 1981. Treefalls and the distribution of understory birds in a tropical forest. Ecology 62:938-945. SCHLOSSBERG, S. AND D. 1. KING. 2008. Are shrubland birds edge specialists? Ecological Applications 18:1325- 1330. Smith, K. G. and J. T. Rotenberry. 1990. Quantifying food resources in avian studies: present problems and future needs. Studies in Avian Biology 13:3-5. STEVEN.S, J. 1985. Foraging success of adult and juvenile starlings Sturnus vulgari.'i: a tentative explanation for McDermott cnid Wood - POST-BREEDING BIRD USE OF TIMBER HARVESTS 555 the preference of juveniles for cheiries. Ibis 127:341- 347. Sullivan, K. A. 1989. Predation and starvation: age- specific mortality in juvenile Juncos (Jitnco phaeno- tiis). Journal of Animal Ecology 58:275-286. V.ANDERHOFF, E. N. AND P. K. Eason. 2008. Comparisons between juvenile and adult American Robins foraging for mulberry fruit. Wilson Journal of Ornithology 120:209-213. Vanha-Majamaa, I. and J. Jalonen. 2001. Green tree retention in Fennoscandian forestry. Scandinavian Journal ot Forest Research Supplement 3:79- 90. Van Horne, B. and A. Bader. 1990. Diet of nestling Winter Wrens in relationship to food availability. Condor 92:413-420. Vega Rivera, J. H., J. H. Rappole, W. J. McShea, and C. A. Haas. 1998. Wood Thrush postfledging movements and habitat use in northern Virginia. Condor 100:69- 78. Vergara, P. M. and R. P. Schlatter. 2006. Aggregate retention in two Tierra del Fuego Nothofagus forests: short-term elfects on bird abundance. Forest Ecology and Management 225:213-224. ViTZ, A. C. 2008. Survivorship, habitat use, and movements for two species of mature forest birds. Dis.sertation. Ohio State University, Columbus, USA. ViTZ, A. C. AND A. D. Rodewald. 2007. Vegetative and fruit resources as determinants of habitat use by mature forest birds during the postbreeding period. Auk 124:494-507. Weathers, W. W. and K. A. Sullivan. 1989. Juvenile foraging proficiency, parental effort, and avian repro- ductive success. Ecological Monographs 59:223-246. White, J. D., T. Gardali, F. R. Thompson III, and J. Faaborg. 2005. Resource selection by juvenile Swainson’s Thrushes during the postfledging period. Condor 107:388-401. Winker, K. 1995. Autumn stopover on the Isthmus of Tehuantepec by woodland Nearctic-Neotropic mi- grants. Auk 112:690-700. Wood, P. B., J. P. Duguay, and J. V. Nichols. 2005. Cerulean Warbler use of regenerated clearcuts and two- age harvests. Wildlife Society Bulletin 33:851-858. The Wilson Journal of Onhihology 1 22(3):556-562, 2010 DISTRIBUTION, ABUNDANCE, AND STATUS OF CUBAN SANDHILL CRANES {GRUS CANADENSIS NESIOTES) XIOMARA GALVEZ AGUILERA'-' AND EELIPE CHAVEZ-RAMIREZ-^ ABSTRACT. — We conducted the first country-wide survey between 1994 and 2002 to examine the distribution, abundance, and conservation status of Sandhill Crane (Gnis canadensis nesiotes) populations throughout Cuba. Ground or air surveys or both were conducted at all identified potential areas and locations previously reported in the literature. We define the current distribution as 10 separate localities in six provinces and the estimated total number of cranes at 526 individuals for the country. Two populations reported in the literature were no longer present and two localities not previously reported were discovered. The actual number of cranes at two localities was not possible to evaluate due to their rarity. Only four areas (Isle of Youth, Matanzas, Ciego de Avila, and Sancti Spiritus) each support more than 70 cranes. The remaining locations each have less than 25 individuals. Sandhill Cranes appear to be declining and have almost disappeared in Pinar del Rio and Granma provinces, and in northern Matanzas Province. Identified threats to the remaining populations include habitat modification (woody plant encroachment, agricultural expansion, and fire suppression), predation due to wild hogs (Sus scrofa), dogs [Canis lupus familiaris), mongoose (Crossarchus spp.), and poaching. Received 3 November 2009. Accepted II February’ 2010. The Cuban subspecies of Sandhill Crane (Grus canadensis nesiotes) is one of many endemic bird subspecies and species present in Cuba. Little information regarding distribution or overall ecology (Meine and Archibald 1996) has been published recently or historically on this rare crane despite being one of the largest birds in the country, and the Caribbean. The Cuban Sandhill Crane has unique characteristics relative to the other subspecies. For example, most nests are constructed on dry land (Walkinshaw 1953, Galvez Aguilera et al. 2005), which differs from all other subspecies which nest primarily in association with wetlands (Tacha et al. 1992). The Cuban subspecies of Sandhill Crane is listed as critically endangered (lUCN 1994). Only four populations were known in the early 1990s and the population for Cuba was estimated at between 100 and 200 individuals by different authors. The first records of Sandhill Cranes in Cuba are from Poey (1851-1855) and Gundlach (1875, 1876) who described the distribution of this species in large savannahs in Cienega de Zapata and pine (Finns spp.) -dominated savannahs east of Guamutas and on the Isle of Pines (cuiTently ' Empresa Nacional Para la Protcccion de Flora y Fauna, Ministerio de Agricultura. Havana. Cuba. ^Platte River Whooping Crane Maintenance Trust Inc.. 661 1 West Whooping Crane Drive. Wood River, NE 68X8.3, USA. ’Current address: Conservacion de Flamencos y Aves, Ninos y Cn'as AC, C. 33 #503 x 6 y 72, Merida, Yucatan, C.P. 90070, Mexico. ■‘Corresponding author; e-mail: fchavez@whoopingcrane.org named Isle of Youth). Barbour (1943) reported cranes from south of Matanzas Province near Alacranes and Union de Reyes. Bangs and Zappey (1905), Read (1913), Walkinshaw and Baker (1946), and Garrido (1985) described the distri- bution on the Isle of Youth (Pines) as limited to open plains north of Cienega de Lanier. Walk- inshaw (1953) defined the distribution on the Isle of Youth as encompassing the region north of Cienega de Lanier from Siguanea westward to West Port and eastward to Pasadita. They are reported to extend to Sabana Grande on the Isle of Youth during winter. Garcia (1987) summarized the status of endemic subspecies in Cuba and considered cranes rare in all regions where they were previously considered abundant, such as in Guane, Mendoza, and Vinales. Garcia (1987) emphasized the need to protect cranes where they were still present at that time, including Pinar del Rio, Ciego de Avila, Camaguey, Cienega de Zapata, and the Isle of Youth. The presence of Sandhill Cranes has been documented over time in different regions of Cuba; however, few authors have made quantita- tive estimates of their abundance. The most documented population is that on the Isle of Youth. The status of Cuban Sandhill Cranes in Cuba at the beginning of our field work was not well known and many different estimates of crane numbers had been put forward from a low of 100 to a high of “about” 300 individuals (Meine and Archibald 1996). There have been reports over the past several years on the population size of cranes by different authors, although they are not in agreement. Cuban Sandhill Cranes, from the time 556 Galvez Aguilera and Chavez- Ramirez • CUBAN SANDHILL CRANE STATUS 557 Gundlach (1875) made his observations when cranes were believed to have been “plentiful”, are considered to have decreased (Walkinshaw 1953). Cranes were considered to be more common during the mid 1980s than formerly (GaiTido 1985), suggesting that numbers had increased between the 1950s and 1980s. Walk- inshaw (1953) suggested that during the early 1950s, Sandhill Cranes had almost disappeared from mainland Cuba. Numbers reported for Cuban Sandhill Cranes have fluctuated from 200 in 1953 (Walkinshaw 1953) to 120 in 1975 (Gan-ido and Garcia 1975) to more than 200 individuals in the early 1990s (XGA, pers obs.). Most reports citing numbers of cranes in Cuba are estimates based on only limited field research or tield surveys, which we believe underestimated actual crane numbers. Many of the 15 crane species, and several subspecies of more common cranes, are consid- ered to be threatened or endangered (Meine and Archibald 1996). The Cuban Sandhill Crane at the beginning of our work was considered to be critically endangered per the lUCN Red List (lUCN 1994), likely because no reliable informa- tion on numbers and distribution was available. The critically endangered designation was based on the belief the populations present were highly fragmented and each numbered less than 50 individuals. We undertook a study to evaluate the current status of the Cuban subspecies of Sandhill Crane in 1995 due to lack of reliable recent information and disagreement among different literature sources on distribution and abundance. Our overall objective was to define the distribution and estimate abundance of the Cuban Sandhill Crane throughout the country. Our specific objectives were to: (1) identify the presence and distribution of different popula- tion(s) or subpopulations of this subspecies in Cuba, (2) estimate the number of individuals present at each .separate location or population, and (3) evaluate the status and potential conser- vation problems for each of the distinct popula- tions. METHODS Country-wide Surveys. — We initially planned a nation-wide search; however, the entire country is not suitable for cranes and we limited our field surveys to areas deemed appropriate based on vegetation characteristics, topography, and per- sonal knowledge of different areas. A preliminary evaluation of Sandhill Crane distribution was defined by contacting field personnel of the Empresa Nacional para la Proteccion de Flora y Fauna (National Agency for the Protection of Flora and Fauna, hereafter Flora and Fauna) to identify areas where cranes had been observed in the previous 5 years. Areas reported in the literature as having cranes were included for preliminary evaluation, regardless of whether new or recent observations had been reported. Each site believed to have potential for cranes was visited and presence was confirmed through actual sighting of cranes from air or ground surveys. The number of cranes present at selected sites in each area was estimated from intensive and extensive ground counts. Aerial surveys were conducted throughout a large portion of the country in areas described on vegetation maps as having characteristics of potentially suitable habitat for Sandhill Cranes (open areas, savannahs, and sparsely forested areas). Open habitat types in Cienega de Zapata, and the provinces of Isle of Youth, Sancti Spiritus, and Camaguey were surveyed at low altitudes (40-60 m) during October 1995 using a fixed- wing airplane. Areas in the provinces of Pinar del Rio, Cienega de Birama, and the northern regions of Matanzas were surveyed from the air in January 1996. Areas in each province with appropriate vegetation characteristics, or actual sightings following the aerial surveys, were visited to conduct personal interviews with local inhabi- tants. Three hundred and forty interviews were conducted of between 10 and 20 individuals in each potential region who had lived in the area at least 20 years. Those interviewed included employees of Flora and Fauna, forest rangers, fishermen, farmers, cattle producers, and others who spent much of their time outdoors in areas likely to be used by cranes. Areas where interviews indicated cranes were likely to be present were surveyed extensively by driving and walking routes to estimate the number of cranes present. More intensive and standardized counts were conducted in areas where cranes appeared to be abundant, or widespread. Population Estimates. — Coordinated point counts were conducted in four areas that appeared to support the largest numbers of cranes to define their distribution and estimate population size. All counts were conducted during February between 1995 and 2002 with a different area sampled each 558 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 3, September 2010 TABLE 1. Cuban Sandhill Crane populations by location and province. Province Location Population' Estimated number Status Protected by'’ Pinar del Rio Guane 1 <10 Stable ENPFF/CGB Consolacion Sur 2t 0 Extirpated NA Matanzas Cienega de Majaguillar 3 <12 Unknown ENPFF Cienega de Zapata 4 120 Increasing EFI/(AP) Sancti Spiritus Cienega de las Guayaberas 5 71 Stable ENPFF (AP) Ciego de Avila Moron Norte 6 102 Stable ENPFF (AP) Jucaro 7 Present Stable? ENPFF Camaguey Sabana de Lesca 8 24 Stable CGB Cayo Romano 9 16 Unknown ENPFF (AP) Granma Cienega Birama lot 7 Not confirmed ENPFF (AP) Isle of Youth Los Indies (11) and Sabana Grande 11, 12 171 Increasing ENPFF (AP) (12) “ Figure 1. ^ ENPFF = Empresa Nacional para la Proteccion de Flora y Fauna (National Agency for the Protection of Flora and Fauna), EFI = Empresa Forestal Integral (Integrated Forestry Agency). CGB = Cuerpo de Guardabosques (Forest Ranger Corps). AP = Protected Area, t = not confirmed. year. Methods used for counts were those developed by the International Crane Foundation and used in Wisconsin (Meine and Archibald 1996) and Mississippi (Hereford et al. 2001) for other subspecies of Sandhill Cranes. Counts were conducted by systematically establishing obser- vation points 1 .5-2 km apart throughout the entire area to be sampled. Ground counts conducted in the four areas with the largest numbers of cranes consisted of between 16 and 40 points. Two to four observers at each point recorded the number of cranes observed, time of observation, direction from observation point, and direction of flight. Between 300 and 500 persons participated in the counts at each location surveyed. Two hours of observation were conducted before sunset during an evening followed by 2 hrs of observation the next morning beginning at sunrise. We recorded number of birds seen, time of observation, and direction of flight at each sample location to eliminate duplicate observations. The maximum number of birds, after eliminating duplicate counts, was considered the total for that area. RESULTS Distribution and Abundance. — We defined the distribution of Cuban Sandhill Cranes as consist- ing of 10 separate localities and presumed to consist of nine separate populations in six provinces (Table 1, Fig. 1). We estimated 526 cranes for the entire country. The total cranes counted in each of the areas with largest numbers of cranes varied from 71 in Guayaberas to 171 on the Isle of Youth (Table 2) for a total of 464 cranes in the four areas with the largest popula- tions. We estimated all other sites combined to have 62 cranes (Table 1) based on site visits. FIG. I. Distribution of Sandhill Crane populations in Cuba. Numbers represent the location of populations corresponding with data in Table 1. Galvez Aguilera am! Chavez- Ramirez • CUBAN SANDHILL CRANE STATUS 559 TABLE 2. Estimates of Sandhill Crane abundance in the four largest populations in Cuba. Categor>’ Isle of Youth Cienega de Zapata Guayaberas Moron Norie Survey period February 1998 February 1999 February 1998 February 2(K)2 # Count stations 52 26 16 32 Cranes obs. /point 5.3 6.3 9.1 6.5 Range 1-22 1-21 4-16 2-18 Estimated total cranes 171 120 71 102 ground-based surveys, direct observations, and knowledge ot local people. We confirmed crane presence for Jucaro (#7, Fig. 1) but have no estimate of abundance. We could not confirm presence of cranes during our study or based on local people’s knowledge, for two locations previously reported (Walkinshaw 1949). These included the region of Consolacion del sur in Pinar del Rio Province and Cienega de Birama in Granma Province (#’s 2 and 10, respectively; Fig. 1 ). Only a rough estimate of numbers is available for two other localities due to the rarity of crane observations in those areas. These localities are Cienega de Majaguillar in northern Matanzas Province and Cayo Romano key, off the northern coast of Camaguey Province (#’s 3 and 9, respectively; Fig. 1). Provincial Reports Pinar Del Rio Province. — Interviews with 30 local residents suggested crane presence in a small area over the entire time period they could remember. We confirmed crane presence in 1996 in the area of Guane (#1, Fig. 1) with no more than 10 individuals likely remaining. Locals suggested cranes were more common through the 1990s but had decreased significantly in recent years. Large areas were planted with Carribean pine (Pinus caribaea) in the 1970s, which apparently caused many areas to become wooded and less suitable for cranes. Our last visit to the area in 2002 confirmed the presence of one chick; however, even fewer cranes were apparent than in 1996. Threats to cranes in this province include habitat modification, presence of large agricultur- al operations including hog {Siis scrofa) farming, and extensive hunting activities, some of which are reported to be illegal. The area where cranes are present is under management and protection by Forestry Agency authorities, and hunting has been limited due to concerns expressed to authorities regarding the critical situation for cranes in the area. However, the low number of cranes present and the small area of habitat available will make it difficult for cranes to increase in this locality in the future. Matanzas Province. — Cienega de Zapata (#4, Fig. 1) in the southern portion of the province has long been known to support cranes. Crane presence was confirmed from aerial surveys in 1995 and, in 1996, five nests were found in different sections of the wetlands. We estimated 120 cranes during the ground point counts conducted in Cienega de Zapata in 1998. We recorded a new locality in the northern area of the province not previously reported as supporting cranes called Cienega de Majaguillar (#3, Fig. 1) with an estimate of <12 cranes. The situation of these cranes appears to be similar to that of the precarious population in Pinar del Rio. There are several threats to cranes in this area. Thei'e is loss of open savannah due to increases in woody shrubs and trees such as cayput (Galaleuca quinquenerva), believed to be a consequence of decreased use of prescribed fires since the 1 970s. Significant populations of wild hogs and dogs (Canis lupus familiaris) are present and could threaten Sandhill Cranes. Sancti Spiritus Province. — Sandhill Cranes are present in the Cienega de Guayaberas and, according to those interviewed, have been present for at least 40 years (#5, Fig. 1 ). A ground point count conducted in 1998 estimated 71 cranes. The area suitable for cranes is <12 knr and is decreasing due to invasion by exotic woody plants such as marabou (Dichrostachys glomer- ata), which eliminates open spaces. Large areas have also been modified for rice cultivation and this practice continues to increase. The presence ot water buffalo [Buhalus huhalis) has caused overgrazing in some areas further contributing to invasion of exotic woody plants. Ciego de Avila Province.— SandhiW Cranes have been reported in this province since the beginning of the 20th century. Crane presence was 560 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122. No. 3, September 2U 10 confirmed via aerial and ground surveys in 1995 and 1996 in the northern portion of the province (#6, Fig. 1), but not in the south (#7, Fig. 1). We did not obtain a count of cranes in the south but did confirm their presence. We estimated 102 and 101 cranes were present in the northern area of the province based on point counts conducted in 1997 and 2002, respectively. Chicks were observed in this area in 1995 and 1996. Northern Ciego de Avila supports one of the largest populations of cranes in Cuba and is a priority area for protection. The case for protection was made to government authorities. A 250,000-ha protected area was established in summer 2004 that will limit and control expansion of habitat alterations in the entire region. However, agriculture, cattle grazing, and tree plantations continue and could affect crane habitat in the future. Potential threats in this area include extensive livestock grazing, exotic plants and animals, and habitat fragmentation due to agriculture and tree plantations. Cattle grazing is expanding and there is an effort to establish non-native grasses for livestock forage. Feral water buffalo are present, which has led to overgrazing and erosion in some areas. Invasive plants include casuarina {Casiia- rina equisetifolia) and marabou, which are rapidly increasing and causing reduction of open savan- nahs. The hydrological regime of this area has been affected by deforestation of native plants, road construction, and sugar cane fields, which could flood crane nests and drown chicks. The presence of sugar cane fields provides cover for mongoose (Crossarchus spp.) and black rats (Rattiis rattus), which depredate crane eggs. Ccwuigiiey Province. — Historical Sandhill Crane localities (#8, Fig. 1) in this province included Sabanas de Lesca and Me.seta de San Felipe. Surveys of local people suggested cranes were no longer present but we observed six cranes flying in 1994. At least 12 pairs were observed in a 2001 ground survey in the Meseta de San Felipe. A previously unreported area was key Cayo Romano off the northern coast (#9, Fig. 1), which is believed to have had cranes for at least 30 years. No extensive count was possible during our work in this area but biologists who visited the site estimated that —16 cranes are present on this key. Principle threats in this area include possible future mining activities and military exercises (Cayo Romano) that occur in the area occupied by cranes. Wild dogs are known to be present in the area. Granma Province.— A single recent visual observation was reported in interviews for this province in the region of Birama in the delta of Rio Cauto (#10, Fig. 1). Interviews suggested cranes were common in the 1970s through the mid 1980s, but decreased in number through the early 1990s and are now believed to be completely absent. Our ground searches conducted yearly between 1999 and 2002 could not confirm the presence of cranes in this area. Habitat modifica- tions in areas of previous crane use include development of wetlands for rice production. We believe cranes have been extirpated from this province. Isle of Youth Province. — Isle of Youth is a well known area for Cuban Sandhill Cranes (#’s 11 and 12, Fig. 1). The type specimen for this subspecies was collected from the Isle of Youth (Bangs and Zappey 1905). Two ground counts were conduct- ed on the Isle of Youth using the same points in 1995 and 1998, yielding estimates of 1 15 and 171 cranes, respectively. The 1995 value is similar to that reported by Walkinshaw (1953) of 100 cranes for the Isle of Youth. We believe the 1998 count is more accurate as it was conducted over 2 days during one morning and one afternoon, compared to the 1995 count which was conducted only during the morning. Threats present include cattle grazing activity and herding during the nesting period. Illegal hunting and fishing activities cause disturbance to the cranes in this area. Conservation Status of Populations Eight of 10 areas supporting Sandhill Cranes are under some form of protection, either as protected areas or other schemes managed by Flora y Fauna. These include Birama, Cayo Romano, Norte de Moron, Cienega de Guayaber- as, Cienega de Zapata, Majaguillar, Jucaro, and Los Indios/Sabana Grande. Two more were under management by the Forestry Agency in Matanzas Province (Empresa Integral Forestal de Cienega de Zapata). Areas where Sandhill Cranes were present in Guane in Pinar del Rio Province and Lesca in Camaguey Province were also managed by the Forestry Agency. DISCUSSION We confirmed Cuban Sandhill Cranes to be more widespread than previously reported (Gar- rido 1985, Garcia 1987, Meine and Archibald 1996). However, presence of cranes in two areas Gcdvez Aguilera and Chavez- Ramirez • CUBAN SANDHILL CRANE STATUS 561 that previously supported them (Walkinshaw 1949) could not be confirmed. The abundance of cranes estimated during this study for Cuba is higher than all previously reported estimates, leading to reclassifying the Cuban Sandhill Crane from critically endangered in 1994 to vulnerable in 1997. Our estimate of 526 cranes is signifi- cantly greater than the estimated 103 individuals of the endangered Mississippi Sandhill Crane {Gnis canadensis pulla) (Hereford et al. 2001), but lower than the 4,000-6,000 estimated for Florida Sandhill Cranes (G. c. pratensis) (Tacha et al. 1994). The present distribution is broader than expected, but at least two previously documented sites appear to no longer support cranes. We have no explana- tion for causes of the disappearance of Sandhill Cranes from these areas, except to note significant habitat changes have occurred in those areas. A series of threats was identified for all of the present populations. At least two populations, those in Guane in Pinar del Rio and Birama in Granma, are so low that it is likely they have already disappeared or will do so soon. The estimates are unreliable for other localities and are possibly optimistic, and should be re-evaluated in the field (Cayo Romano and Sabana de Lesca in Camaguey, and Majaguillar in northern Matanzas Province). We could not identify causes of population disappearance or decreases, but important factors may include population isolation effects and degradation and decreases in available habitat. Decreases in size or fragmentation of habitat could contribute to population declines in some areas as alterations are visible throughout the entire country. The Mississippi Sandhill Crane’s low productivity and endangered status is attrib- uted in large part to habitat fragmentation (Valentine 1970; S. G. Hereford, pers. comm.). All existing populations are separated by distanc- es greater than the longest known dispersal movement by non-migratory Sandhill Cranes (up to 48.3 km in Florida; Nesbitt et al. 2002). We have limited data on potential movements and dispersal of Cuban Sandhill Cranes but, of 10 radio-marked birds on the Isle of Youth, a single crane moved >18 km during 1 year (Galvez Aguilera 2002). Even shorter movements (10.5 km by Florida Sandhill Cranes in Georgia; Bennett 1989) are reported for other non-migratory Sand- hill Cranes. Four significant populations appear to be stable and perhaps increasing; (1) Isle of Youth, (2) northern Ciego de Avila, (3) Cienega de Zapata in Matanzas, and (4) Cienega de Guayaberas in Sancti SpiritLis. Delineation of a new 250,()()()-ha protect- ed area in 2004 that encompasses all crane areas in northern Ciego de Avila Province resulted in the four largest populations being at least partially within the boundaries of a protected area. Howev- er, that does not guai'antee protection as several threats are present regardless of whether they are in or outside a protected area boundary. Woody plant encroachment continues in many areas within Reserves, due to elimination of natural fire regimes, and is a serious problem of concern for several federal agencies (FC-R, pers. obs.). Feral domestic (hogs and dogs) and introduced animals (mongoose) continue to be threats to nesting cranes in all four areas; however, the exact effect of each one of these predators has not been quantified. Predation is a problem that affects other non- migratoi'y populations and is the major cause of mortality for the Mississippi Sandhill Crane (S. G. Hereford, pers. comm.) We have been working directly in the Isle of Youth over the last few years and recently in Ciego de Avila to better evaluate the specific threats to cranes and develop appropriate or responsive management plans. Some Reserve personnel are trying to implement counts at regular intervals to document population trends. The trend and conser- vation status of Cuban Sandhill Cranes may not be entirely known until we can ascertain whether numbers are increasing or decreasing over time. Possible direct actions to enhance declining popu- lations could include exchange of individuals from one population to another or from potential captive breeding efforts. Captive breeding programs for cranes do not exist in Cuba, but are present and successful in other countries. Indirect actions to expand or improve nesting habitat could include management activities including instituting pre- scribed fire programs, and elimination and manage- ment of invasive woody plants, as has been done for the Mississippi Sandhill Crane (Valentine and Hereford 1997). Limiting agriculture and forestry plantation expansion in protected areas where cranes are present could also help expand or improve nesting habitats. Control of potential domestic and introduced predators is another action that could be taken in areas where predators are a problem. ACKNOWLEDGMENTS Many individual.s contributed to this effort. Some of those that provided support over several years include 562 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 3, September 2010 Miguel Magraner Fernandez, Jose Rivera Rosales, Fidel Quialia Gongora, Jose Osorio, Vladimir P. Fleites, Jose R. Rodriguez, Daniel Reit. Jorge Jerez, Amar Perez, Duniet Marrero, Pavel Martinez Arredondo, Leandro Torrella, Odey Martinez, and Frank Moreira. We thank the Empresa Nacional para la Proteccion de Flora y Fauna, territorial delegations of the Ministerio de Agricultura, the provincial delegations of CITMA in Sancti Spritus and Ciego de Avila, Cuerpo de Guardabosques Nacional, The Interna- tional Crane Foundation, and Jessica Dooley and Luis Ramirez for assistance with the manuscript. We thank the many individuals, organizations, and agencies that helped in a variety of ways. Funding for this project was provided in part by grants from the Tom D. and Catherine T. MacArthur Foundation, the National Geographic Society, and the World Watch Institute. We greatly appreciate the comments and suggestions by the editor and two anony- mous reviewers. LITERATURE CITED Bangs, O. and W. R. Zappey. 1905. Birds of the Isle of Pines. American Naturalist 39:179-215. Barbour, T. 1943. Cuban ornithology. Memoirs of the Nuttall Ornithological Club Number 9. Cambridge, Massachusetts, USA. Bennett, A. J. 1989. Movements and home range of Florida Sandhill Cranes. Journal of Wildlife Manage- ment 53:830-836. Galvez Aguilera, X. 2002. Distribucion y abundancia de Grus canadensis nesiotes en Cuba. Uso de habitat y reproduccion de una poblacion de esta specie en la Reserva Ecologica Los Indios, Isla de la Juventud. Dissertation. Universidad de la Habana, Cuba. Galvez Aguilera, X., V. Berovides, and E. Chavez- Ramirez. 2005. Nesting ecology and productivity of the Cuban Sandhill Crane on the Isle of Youth, Cuba. Proceedings of the North American Crane Workshop 9:225-236. Garcia, F. 1987. Las aves de Cuba, subespecies endemicas. Tomo II. Gente Nueva, La Habana, Cuba. Garrido, O. H. 1985. Cuban endangered birds. Pages 992- 999 in Neotropical ornithology (P. A. Buckley, M. S. Foster, E. S. Morton, R. S. Ridgely, and F. G. Buckley, Editors). Ornithological Monographs Number 36. Garrido, O. H. and F. GarcIa. 1975. Catalogo de las aves de Cuba. Academia de Ciencias, La Habana. Cuba. Gundlach, J. 1875. Neue beitrage zur ornithologie Cubas. Journal fur Ornithologie 23:293-340. Gundlach, J. 1876. Contribucion a la omitologi'a cubana. Impresora La Antilla, La Habana, Cuba. Hereford, S. G., T. E. Grazia, and J. N. Phillips. 2001. The 2000 Mississippi Sandhill Crane annual popula- tion census. USDI, Fish and Wildlife Service, Mississippi Sandhill Crane National Wildlife Refuge Report. Gautier, Mississippi, USA. lUCN. 1994. lUCN red list categories. lUCN, Gland, Switzerland. Meine, C. D. and G. W. Archibald. 1996. The cranes: status survey and conservation action plan. lUCN, Gland, Switzerland. Nesbitt, S. A., S. T. Schwikert, and M. J. Folk. 2002. Natal dispersal in Florida Sandhill Cranes. Journal of Wildlife Management 66:349-352. POEY, F. 1851-1855. Apuntes sobre la fauna de la Isla de Pinos. En Memorias sobre la Historia Natural de la Isla de Cuba, acompanadas de sumarios Latinos y extractos en Erances, I. Imprenta de Barcina, La Habana: 424-431. Read, A. C. 1913. Birds observed in the Isle of Pines, Cuba. Oologist 30:130-131. Tacha, T. C., S. a. Nesbitt, and P. A. Vohs. 1992. Sandhill Crane {Grus canadensis). The birds of North America. Number 3 1 . Tacha, T. C., S. A. Nesbitt, and P. A. Vohs. 1994. Sandhill Crane. Pages 76-94 in Migratory shore and upland game bird management in North America (T. C. Tacha and C. E. Braun, Editors). International Association of Eish and Wildlife Agencies, Washing- ton, D.C., USA. Valentine, J. M. 1970. A colony of Sandhill Cranes in Mississippi. Journal of Wildlife Management 34:761- 768. Valentine, J. M. and S. G. Hereford. 1997. History of breeding pairs and nesting sites of the Mississippi Sandhill Crane. Proceedings of the North American Crane Workshop 7:1-9. Walkinshaw, L. H. 1949. The Sandhill Crane. Bulletin Number 29. Cranbrook Institute of Science, Bloom- field Hills, Michigan, USA. WALKIN.SHAW, L. H. 1953. Nesting and abundance of the Cuban Sandhill Crane on the Isle of Pines. Auk 70: 1-10. Walkinshaw, L. H. and B. W. Baker. 1946. Notes on the birds of the Isle of Pines, Cuba. Wilson Bulletin 58:133-142. The Wilson Jot15 sec and seldom dive deeper than ~2 m. Once out of the water, a quick shake removes any remaining water drops from their feathers (Attenborough 1998). The White-capped and the Rufous-throated dippers have contour feathers that show the same structural dimensions as contour feathers of typical terrestrial birds, whereas those of the White-throated Dipper appear to have developed a slight increase in water resistance. In contrast, the American Dipper and, in particular, the Brown Dipper with their larger (r + d)/r values have 568 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. 3, September 2010 attained a substantial increase in water repellency at some expense of their water resistance. The first two species, which peck their under-water prey from a rocky perch and do not dive, show no difference in feather structure from that of truly terrestrial birds, whereas the last three are true plungers with breast feathers that show adapta- tions to an aquatic lifestyle. The White-throated Dipper, among these three, has evolved an adaptation to resisting water penetration; the other two have developed greater water repellency, in particular the Brown Dipper. These diverging adaptations may be explained by differences in habitat; the White-throated Dipper is more frequently found along water courses at lower altitudes, where currents are relatively slow and the water is deeper. This would favor greater resistance to water penetration. American and Brown dippers, in contrast, typically inhabit higher mountain ranges, where currents are fast and shallow, which would favor water repellency over resistance to water penetration. The geo- graphical range of the White-throated Dipper, in its most eastern distribution, partly overlaps that of the Brown Dipper, which has probably evolved primarily in the isolation of the high Himalayan mountain ranges. Brown Dippers tend to breed at altitudes well over 5,000 m, whereas the White- throated Dipper prefers the 2,000 to 3,500 m range in Afghanistan and somewhat higher in the Himalaya. However, there is overlap in altitude, particularly at the western end of their overlap- ping ranges and during the non-breeding season (Ormerod and Tyler 2005). Adaptations of dipper feathers from their terrestrial origins to their under-water feeding habits in terms of both water repellency and resistance to water penetration are consistent with our hypothesis including the observation that no similar adaptations have been made by their non- diving congeners. It is likely that plunging is a recent behavior given the dippers’ passerine origins and that their feathers are primarily water repellent, and only partially adapted to either prevent water penetration or extend water repel- lency. Plunging is apparently, an evolutionary step that White-capped and Rufous-throated dippers have not completed. ACKNOWLEDGMENTS The authors express their gratitude to the Curators, Division of Birds, National Museum of Natural History, Smithsonian Institution, Washington, D.C. for permission to use dipper feathers for this study. The authors are also indebted to C. J. Dove for assistance with measurements. LITERATURE CITED Attenborough. D. 1998. The life of birds. Video® Episode 5. BBC Books, London, England. Cassie, a. B. D. and S. Baxter. 1944. Wettability of porous surfaces. Transactions of the Faraday Society 40:546-551. Elowson, a. M. 1984. Spread-wing postures and the water repellency of feathers; a test of Rijke’s hypothesis. Auk 101:371-383. Kennedy, R. J. 1970. Directional water-shedding properties of feathers. Nature 227:736-737. Kennedy, R. J. 1972. The probable function of flexules. Ibis 1 14:265-266. Moilliet, j. L. (Editor). 1963. Water proofing and water repellency. Elsevier, New York, USA. Ormerod, S. J. and S. J. Tyler. 2005. Family Cinclidae (Dippers). Pages 332—355 in Handbook of the birds of the world. Volume 10 (J. del Hoyo, A. Elliott, and D. A. Cristie, Editors). Lynx Edicions, Barcelona, Spain. Ruke, a. M. 1968. The water repellency and feather structure of cormorants, Phalacrocoracidae. Journal of Experimental Biology 48: 185-189. Ruke, A. M. 1970. Wettability and phylogenetic develop- ment of feather structure in water birds. Journal of Experimental Biology 52:469^79. Ruke, A. M. and E. H. Burger. 1985. Wettability of feathers and behavioural patterns in water birds. Proceedings of the Pan-African Ornithological Con- gress 6: 153-158. Ruke, A. M., W. A. Jesser, and S. A. Mahoney. 1989. Plumage wettability of the African Darter, (Anhinga melanogaster) compared with the Double-crested Cormorant, (Phalacroconix aiirifiis). O.strich 60:128- 132. Ruke, A. M., W. A. Jesser, S. W. Evans, and H. Bouwman. 2000. Water repellency and feather structure of the Blue Swallow (Hinmdo atrocaerulea). Ostrich 7 1 ; 143-145. Rut.SCHKE, E. 1960. Untersuchungen iiber Wasserfestigkeit und Struktur des Gefieders von Schwimmvogeln. Zoologische Jahrbiicher 87; 441-506. The Wilson Journal oj Ornithology l22(3):569-576, 2010 ANNUAL BIRD MORTALITY IN THE BITUMEN TAILINGS PONDS IN NORTHEASTERN ALBERTA, CANADA KEVIN P. TIMONEY'-' AND ROBERT A. RONCONP ABSTRACT. Open pit bitumen extraction i.s capable of cairsing mas.s mortality event.s of resident and migratory birds. We investigated annual avian mortality in the tailings ponds of the Athabasca tar sands region, in northeastern Alberta, Canada. We analyzed three types of data: government-industry reported mortalities; empirical studies of bird deaths at tailings ponds; and rates ot landing, oiling, and mortality to quantify annual bird mortality due to exposure to tailings ponds. Ad hoc selt-reported data from industry indicate an annual mortality due to tailings pond exposure in northeastern Alberta of 65 birds. The self-reported data were internally inconsistent and appeared to underestimate actual mortality. Scientific data indicate an annual mortality in the range ot 458 to 5,029 birds, which represents an unknown fraction of true mortality. Government-overseen monitoring within a statistically valid design, standardized across all facilities, is needed. Systematic monitoring and accurate, timely reporting would provide data useful to all concerned with bird conservation and management in the tar sands region. Received 17 November 2009. Accepted 5 May 2010. Global demand for unconventional energy sources such as coal bed methane, heavy oil, and bitumen has grown in recent years. Bitumen in northern Alberta, Canada, is extracted by two methods, in situ well-based approaches and truck and shovel open pit mining. The latter method produces “tails” during separation of bitumen from the sand. The tails, a mixture of process- affected water, residual hydrocarbons, brine, silts and clays, and metals are discharged into tailings ponds. The extent of tailings ponds in northeastern Alberta grew by 422% between 1992 and 2008 (Timoney and Lee 2009). The Athabasca tar sands development is one of the largest energy projects in the world. Production of bitumen is predicted to rise from the current 1.3 million barrels/day to three million barrels/day by 2018 (Alberta Energy 2009). Water bodies along migration routes attract many bird species as they afford foraging, roosting, nesting, and resting opportunities (Ron- coni 2006). A variety of deterrents have been used to discourage waterbirds from landing in tailings ponds such as floating and beach effigies, propane scare cannons, and sound-producing systems (Boag and Lewin 1980, Colder Associates Ltd. 2000, Ronconi and St. Clair 2006). Some birds that land at tailings ponds become oiled and a proportion of the oiled birds later die. Bird mortality rates from oiling have not been precisely 'Treeline Ecological Research. 21551 Township Road 520. Sherwood Park, AB T8E IE3, Canada. ^Department of Biology, Dalhousie University, 1355 Oxford Street. Halifax. NS B3H 4J1. Canada. ■’Corresponding author; e-mail: ktimoney@interbaun.com measured, but casualties appear to be high for gregarious species, particularly for diving birds (Clark 1984). Bird migration is affected by weather as birds are more likely to land when they encounter headwinds, low temperatures, and precipitation (Newton 2007). Storms may increase the likelihood of bird oiling at tailings ponds (Ronconi 2006), and inclement weather may increase the probability of mass mortality events. Oiled ducks may suffer from reduced insula- tion, increased metabolic rate, and hypothermia even from small amounts of oil (Hartung 1967, McEwan and Koelink 1973). Survival rates of rehabilitated birds may be as low as 1 to 20% for some species (Mead 1997). Birds from 43 species have died due to exposure to tailings ponds in the area, mostly waterbirds such as dabblers and divers: Mallard (Anas platyrhynchos). Common Goldeneye {Bucephala clangula). Northern Shov- eler {Anas clypeata). Lesser Scaup {Aythya affinis), American Coot {Fulica americana), grebes, mergansers, geese, and shorebirds includ- ing Semipalmated Sandpiper {Calidris piisilla). Pectoral Sandpiper (C. melanotos). Stilt Sandpiper (C. himanlopits), and Lesser and Greater Yellow- legs {Tringa flavipes, T. melanoleiica) (Shaip et al. 1975, Dyke et al. 1976, Gulley 1980, Ronconi 2006). Deaths of birds of prey, gulls, passerines, and other groups have also been documented. Mortality rates may be high even at small ponds: 27 dead birds were found at a 0.4-ha tailings pond lacking deterrents (Dyke et al. 1976). There may be continual “incidental take" of birds during the open water .season, especially at night when human observations are impractical. Oiled birds in tailings ponds have been observed to sink out of 569 570 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122. No. 3. September 2010 sight (Dyke et al. 1976), minimizing the chance of detection. Our objective was to provide estimates of annual bird mortality resulting from bitumen tailings pond exposure in northeastern Alberta, Canada through synthesis and analysis of avail- able data. These data included numbers reported by industry to government and scientific data on mortality and landing rates at tailings ponds. METHODS Study Area. — We studied avian mortality in the Athabasca bitumen (tar) sands region (geographic center at 57° 03' N, 111° 31' W, Fig. 1) in the lower Athabasca River watershed north of the city of Fort McMurray, Alberta, Canada. The 120.6 km^ of tailings ponds within the area of open pit mining, as of March 2008, covered 1.4 times the area of natural water bodies (84.9 kni') composed of the surface of the Athabasca River (50.4 km-) and lakes, ponds, and other river surfaces (34.5 km-) (Timoney and Lee 2009; KPT and RAR, unpubl. data). The area lies within a convergence zone of North American waterfowl llyways; millions of birds migrate through northeastern Alberta en route to and from local and distant breeding areas in northern Alberta, the Peace- Athabasca Delta, Mackenzie River Valley, and the arctic (Butterworth et al. 2002, Thomas 2002, USDI 2009b). Thirty-five species and species groups of waterbirds, and 29 other species have been observed on one lease (Syncrude # 17) at the natural water body Mildred Lake (Sharp et al. 1975). More than 16,000 birds were observed flying over one tailings pond during spring migration (Ronconi and St. Clair 2006) while more than 25,000 swans, geese, ducks. Sandhill Cranes (Grus canadensis), and gulls were ob- served in daylight during a fall migration at Syncrude Lease # 17 (McLaren and McLaren 1985). The total number of migratory birds passing through the lower Athabasca River Valley is unknown. Data Collection and Analyses.—Spot censuses and shoreline searches for dead birds of varying duration, extent, and frequency at tailings ponds of known areal extent during the open-water season (Gulley 1980, Van Meer and Arner 1985) were used to calculate bird mortalities per knr from which mortalities were adjusted to the 2008 areal extent ot tailings ponds. Mortality data were collated from three companies with tailings ponds (Suncor 1990-2008, Syncrude 2000-2007, and FIG. 1. Study area (modified from Timoney and Lee 2009). Areas (as of 19 March 2008) undergoing bitumen extraction are hachured; tailings ponds are black. Tailings pond names are MLSB = Mildred Lake Settling Basin; ANTP = Aurora North tailings pond; SATP = Shell Albian tailings pond; TIP = Tar Island Ponds 1 and 1 A; 8AEML = Suncor Millennium tailings ponds 8A and EML. Shell Albian 2000-2008). Data were obtained from reports produced by the companies (Syn- crude 2008), and from the Alberta government (Sustainable Resources Development) under a freedom of information request (K. P. Timoney, October 2008). These data, reported by company, year, and mortality type were analyzed to obtain mean annual mortality. We estimated the total number of birds landing and subjected to oiling during spring migration at the Albian Sands Muskeg River Mine: landings/hr (from Ronconi and St. Clair 2006) and oiled birds/day (from Ronconi 2006). RESULTS Mortality Rates Estimated by Systematic Sur- veys.— Systematic surveys for dead birds at tailings ponds (Table 1), used to calculate mor- tality per km^ (range 7.2 to 145.2 birds), were extrapolated to estimate total potential mortality based on 120.6 km^ of tailings ponds in 2008. An estimate based on the lowest observed mortality at Tinumcy ciiul Roiiconi • BIRD MORTALITY AT TAILINGS PONDS 571 TABLE 1. Estimated annual bird mortality/kmVyear in the Athabasca tar sands tailings ponds based on spot counts and systematic shoreline surveys. Site Area (km’)” Year Dead birds Dead birds/knr' Reference’’ Cumment.s MLSB 12.25 1984 94 7.68 L 3 scare cannons, human effigies MLSB 1 1.74 1985 84 7.15 1, 3 scare cannons, human effigies MLSB 1 1.58 1980-1983*’ 189.5 16.36 1, 3 scare cannons, human effigies Pond 1 1.86 1977 77 41.40 2 human effigies; fresh tailings received 95% of days (Apr-Ocl) Pond 1 1.86 1978 79 42.47 2 human effigies; fresh tailings received 100% of days (Apr-Oct) Pond 1 1.86 1979 270 145.16 2 human effigies with artificial lighting at night; fresh tailings received 92% of days (Apr-Oct) Pond lA 0.56 1977 43 76.79 2 deterrents?‘‘; fresh tailings received 34% of days (Apr-Oct) Pond lA 0.56 1978 31 55.36 2 deterrents?'*; fresh tailings received 25% of days (Apr-Oct) Pond lA 0.56 1979 33 58.93 2 deterrents?'*; fresh tailings received 0% of days (Apr-Oct) “ .Xreas of MLSB derived from planimetry of airphotographs ( 1 980, AS2 1 65- L (1980). 3; 1984, AS3051 -5; 1986, AS3356-280; Ponds 1 and 1 A areas derived from Gulley Dead birds/year 1980-1983 derived mathematically from reported values for 1984 and 1980-1984 (Van Meer and Arner 1985): 1980-1984 average mortality of 170.4 birds/year; total birds dying 1980-1984 = 170.4 X 5 = 852 birds; 1984 mortality of 94 birds; 1980-1983 average mortality = (852 — 94)/4 = 189.5 birds/ year, or 16.36 birds/km’; the average weighted mean mortality 1980-1985 = ((16.36 X 4) + 7.68 + 7.15)/6 = 13.38 birds/km’. The high estimate of annual mortality is the weighted mean mortality per km=; it is the sum of 500.38 birds/km’ for 12 years of data (1980-1983 comprises 4 years of data), 500.38/12 = 41.70 birds/km^ References: 1 = Van Meer and Amer (1985); 2 = Gulley (1980); 3 = Colder Associates Ltd. (2000). Queries sent to Suncor (17 Nov to 12 Dec 2008) regarding deterrents in use on Pond lA during 1977-1979; no reply received to date (4 May 2010). Syncrude’s Mildred Lake Settling Basin (MLSB) of 7.2 birds/km- in 1985 yielded an annual mortality of 863 birds. A medium estimate based on Syncrude’s MLSB (1980-1985) average mor- tality of 13.38 birds/km^ yielded an annual mortality of 1,614 birds. A high estimate based on the weighted mean mortality rate for all years at Syncrude’s MLSB and Suncor’ s Tar Island Ponds 1 and lA of 41.7 birds/km^ yielded an annual mortality of 5,029 birds. Industry-based Annual Mortality Reported to Government. — Annual mortality attributed to oiling over the period 2000 to 2007 ranged from 17 to 201 birds. The weighted mean (± SD) annual mortality was 65 ± 59 birds. (Table 2). Additional annual mortalities attributed to ’other’ and to ‘unknown’ causes (details in Table 2) averaged (± SD) 13 ± 9 (max 3 1 in 2007) and 16 ± 9 birds, respectively. Industry data had poor agreement with the government data released under the freedom of information request (Ta- ble 3); the mean difference was 19%. Annual Bird Mortality Estimated by Landing and Oiling Rates. — A spring landing rate of 121.44 birds/day in a 3.5 km- tailings pond was calculated, resulting in an estimated rate of 34.69 landings/km^/day during daylight hours only (low estimate. Table 4). We calculated 54.93 landings/ day (109.86 landings/km-/day) (high estimate. Table 4) from observations at a 0.50-km- area where deten'ent testing occuired (RAR, unpubl. data not previously reported in Ronconi and St. Clair 2006). Scaling for the total area of tailings ponds in 2008, about 125,513 to 397,408 birds may land during a 30-day spring migration period. Thirteen oiled waterbirds and shorebirds were found during the same period and at the same site (Ronconi 2006), most of which were covered in >50% oil, from which oilings/day and the proportion of landed birds becoming oiled were calculated (Table 4). We estimate that 286 to 905 birds may be oiled during spring migration at an overall oiling rate of 0.2278% for birds that landed on ponds. About 229 to 815 birds may die each spring due to oiling if an oiled bird is unable to land more than once, and 80 to 90% of oiled birds die (Discussion). DISCUSSION Uncertainties in Mortality Estimates. — Bird oilings may peak in August and September rather than in spring (Van Meer and Arner 1985). and it is reasonable to double the spring mortality to derive an annual mortality of ~458 to 1,630 birds. 572 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. 3, September 2010 TABLE 2. Bird mortalities attributed to oiling, ‘other’ and ‘unknown’ causes released by the Alberta Government for petroleum companies with tailings ponds in northeastern Alberta." Year Oiling" Other Unknown Suncor Syncrude Albian" Suncor Syncrude Albian Suncor Syncrude Albian 1990 103 0 0 1991 93 0 0 1992 194 0 2 1993 135 0 4 1994 87 0 1 1995 43 0 0 1996 72 0 4 1997 71 0 6 1998 80 0 3 1999 48 0 10 2000 193 8 0 2 1 12 7 0 2001 2 15 0 2 0 23 3 0 2002 17 20 1 6 3 1 2 1 2003 15 16 17 2 23 3 2 5 0 2004 10 33 2 0 5 2 2 9 1 2005 3 8 14 2 18 2 1 16 1 2006 3 57 3 4 8 7 3 8 4 2007 9 10 26 1 7 6 6 19 6 2008" 16 4 0 2 1 0 2 4 Recent Mean ± 31.5 ± 20.9 ± 12.4 ± 1.2 ± 8.9 ± 3.0 ± 6.2 ± 8.6 ± i.6 ± SD'’ 65.5 16.7 10.0 1.4 7.6 2.4 7.7 6.0 2.2 “ 'Other' includes electrocution, collisions, predation, fights with other birds, and natural causes; 'Unknown' includes incidents where company was not able to identify cause of death and incidents where cau.se of death was not reported. Tailings pond at Shell Albian began to fill in 2003; mortality due to oiling not expected prior to 2003. Calculations of recent tailings pond mortalities used years 2000-2007 for Suncor and Syncrude and years 2003-2007 for Shell Albian. Mortality means for 'other' and 'unknown' use the period 2000-2007. Mortality means are for each company and mortality type. Average mortalities by year, 2000-2007. oiling 64.8 ± 58.7. other 13.1 ± 9.3. and unknown 16.5 ±9.1. Values for 2008 were 'year to date' current to July 2008 with the exception of Syncrude for which the death of 1,606 ducks at the Aurora North tailings pond in April 2008 was not made public until 2009. This adjustment to the mortality estimate may be conservative as it does not include mortalities that occur before spring, between spring and fall migration, and after fall migration. Annual tailings pond mortality estimates derived from mortality surveys (863 to 5,029 birds) and landing-oiling rates (458 to 1,630 birds) are roughly of the same magnitude. Self-reported oiling mortality data from industry provide the lowest estimate (65 bird deaths/year) whereas Wells et al. (2008) provide a high estimate of 8,676 to 156,168 bird deaths/year. Wells et al. (2008) assumed that all birds that land at tailings ponds are oiled and that peak landing rates exist 24 hrs/day for 100 days. Our mortality estimates may be conservative given that 500,000 to one million birds die annually at oilfield wastewater ponds in the United States (USDI 2009a). Those wastewater ponds are similar to bitumen tailings ponds in their mixture of water, residual oil or bitumen, and salts. The presence of extensive tailings ponds TABLE 3. Annual bird mortality at Syncrude" as reported by the Alberta government and by Syncrude (2008). Source Year 2000 2001 2002 2fX)3 2004 2005 2006 2007 Alberta Government 17 20 28 44 47 42 73 36 Syncrude 20 21 31 44 69 55 46 35 DifferenceL % -18 -5 -11 0 -47 -31 37 3 “ Combined Mildred Lake and Aurora leases. Difference = (Govemmeni - Syncrude/Govemment) X 100; mean difference without regard to sign = 19%. Timoney and Ronconi • BIRD MORTALITY AT TAILINGS RONDS 573 TABLE 4. Rates of landing and the proportion of birds that subsequently become oiled at Shell Albian Sands tailings pond during April-May 2003.'' Number of birds landing Landings/hr Landings/day' Oiled birds Oiliiigs/day % Landed birds that became oiled Ducks 536 3.99 63.69 7 0.149 0.23 Shorebirds 444 3.30 52.76 4 0.085 0.16 Geese/Swans 10 0.07 1.19 1 0.021 1.79 Gulls 13 0.10 1.54 1 0.021 1.38 Other waterbirds 19 0.14 2.26 0 0.000 0.00 Overall L022 7.60 121.44 13 0.277 0.23 Data compiled from spring migration studies in the 3.5 knr main tailings pond at Shell Albian Sands, Muskeg River Mine {Ronconi 2006. Ronconi and St. Clair 2006) yielding a low estimate of 34.69 landings/km’/day; there were 54.93 landings/day within the 0.50 km* observation area yielding a high estimate of 109.86 landings/knr/day; 47 days of ob.servaiion for oiled birds. Loons, grebes, cranes, herons, cormorants, and coots. ^ Obser\ations for daylight hours only; average 15.97 hrs of daylight between 18 April and 29 May, 134.4 hrs of observation (source: www.almanac.com/rise for Fort McMurray, AB, Canada). containing bitumen, polycyclic aromatic hydro- carbons (PAHs), naphthenic acids, brine, heavy metals, and ammonia along an internationally significant migratory bird comdor poses long- term threats to migratory and resident birds (Schick and Ambrock 1974, Wells et al. 2008). Tailings ponds may pose the greatest threat in spring when warm effluent-fed tailings ponds provide open water at a time when natural water bodies remain frozen; however, a high risk of oiling may extend throughout the open water season (Van Meer and Amer 1985). There are four aspects of our estimates that influence their accuracy. First, no nocturnal observations of migrating or landing birds were made. Many birds migrate at night (Richardson 1971, Blokpoel 1973, Blokpoel and Burton 1973), but landing rates during darkness are unknown. Many birds migrate at night and mortality rates might be higher if data from night-time observa- tions were available. There are also no observa- tions for November through early April, when natural water bodies are frozen but large areas of tailings ponds remain unfrozen due to addition of warm tailings. The frequency of landings then is unknown, but is presumably greater than zero for resident birds. Higher rates of nocturnal landings and landings in non-migratory periods would increase our estimates. A second source of estimation error is that numbers of birds flying over, landing, becoming oiled, or being found dead are an unknown fraction of the true parameter values. Some of these parameters were estimated from the most recent and systematically collected data available (e.g., Ronconi 2006, Ronconi and St. Clair 2006); however, without data from other sites and years. there is no means to assess variation in rates of birds landing, becoming oiled, or dying. Third, we assumed a mortality rate of 80 to 90% for birds that came in contact with oil. Mortality rates of oiled birds are unknown (Clark 1984), but even very small amounts of oil may kill birds (Hartung 1967, McEwan and Koelink 1973) and survival rates of rehabilitated oiled birds may be as low as 1 to 20% (Mead 1997). Our estimates assume a small proportion (10-20%) of oiled birds may survive oiling. Finally, by scaling our estimates from individ- ual tailings ponds to the areal extent of tailings ponds in the region, we assumed that bird use and associated mortalities are similar across sites. This assumption remains untested and, for some species such as shorebirds, the extent of shoreline contaminated with bitumen may be a better predictor of mortality than extent of open water. Individual events may result in large variations in mortality. A migratory waterfowl mortality event at the Syncrude Aurora North tailings pond occuiTed in April 2008 at which 1,606 dead waterfowl were later found (CBC 2008, 2010). Provided that all dead waterfowl were found and no non-waterfowl died, the single event resulted in a mortality of 162 birds/km“ well in excess of our highest estimate (Table 1). The frequency of mass mortality events is unknown. Inconsistencies and Deficiencies in Reporting Bird Mortality. — We note three major shortcom- ings in the data provided by government and industry. First, mortality estimates based on mortality surveys and landing/oiling rates are far higher than those reported by government. Second, industry-reported mortalities often do not match mortalities reported by the government 574 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. 3, September 20 10 (Table 3), even though government and industry numbers should be identical. Third, the bird mortality data released by government lack detail. Only company name and total bird mortality for each year and general cause of death are reported; this results in loss of valuable data on location, date, and circumstances of specific incidents. Sampling design, including appropriate sample size, sampling effort, and accurate detection and identification of species is a critical aspect of an effective monitoring protocol (McComb et al. 2010). Numbers of bird mortalities are directly related to search effort and sampling design. Industry-reported data on bird deaths are prob- lematic as they are not systematic, repeatable, and statistically robust. Review of mortality data in Syncrude annual reports indicates that few of the ob.servations come from tailings ponds, which is likely where most oiled birds die. Syncrude (2006, 2008) reported underestimating mortality when explaining an increasing trend in bird mortality in recent years as partially attributable to improved monitoring and reporting practices. The Need for Improved Data. — Currently, neither the total number of birds migrating through the region nor the total annual bird mortality attributable to tailings ponds are known with sufficient scientific rigor. Data on mortalities during extreme weather events are lacking. The fate of lightly-oiled birds that continue migration, in particular to summer breeding areas, is unknown. Important questions remain about the variability in landing and oiling probability with season, weather conditions, time of day, and pond size and location. Questions also remain about mortality detection efficiency in relation to sampling effort, monitoring protocols, and tailings pond size. The ad hoc monitoring by industry, sanctioned by government, is inconsistent, cannot answer these questions, and undoubtedly under- estimates actual mortality. CONSERVATION IMPLICATIONS The pace and scale of development of the Athabasca tar sands is unprecedented in North American history. The industrial footprint and resultant habitat loss may double in 15 years and will certainly increase bird mortality rates. Open pit bitumen extraction may exert population-level impacts upon migratory and resident birds, and is capable of causing mass mortality events. Existing natural water bodies should be protected to help offset landings of birds in tailings ponds (Ronconi 2006) . Production of liquid tailings should be phased-out before this expansion of the industry occurs. Tailings ponds exceed the extent of natural water bodies in the area, continue to increase in extent, and lie along an internationally significant fly way; thus, they may pose a significant regional mortality risk. Harmful effects of tailings ponds are not limited to oiling of waterbirds. Ingestion of bitumen grit by waterfowl may be a significant route of exposure to contaminants (King and Bendell-Young 2000). Nesting Tree Swallows (Tachycineta bicolor) exposed to process-affected wetlands have higher mortality, hormonal stress, nestling parasitism, and reduced nesting success (Wayland and Smits 2004, Gentes 2006). The effects were attributed to PAH exposure, possibly through feeding on contaminated insects. A variety of strategies have been tested to reduce the attraction of birds to industrial developments (Brough and Bridgman 1980, Stevens and Clark 1997, Read 1999; and reviews by Bomford and O’Brien 1990, Donato et al. 2007) . The tar sands industry in northeastern Alberta has been using landing deterrents such as propane cannons and scarecrows for >30 years (Boag and Lewin 1980, Gulley 1980, Colder Associates Ltd. 2000). One of the long-standing problems of bird deterrents is habituation (Bom- ford and O’Brien 1990, Stickley et al. 1995, Conover 2001). A recent comparison of modem deten'ent techniques found the odds of landing at “protected” bitumen tailings ponds remained high relative to non-deterrent controls (Ronconi and St. Clair 2006). Overall, these authors observed no significant difference in the deter- rence value of industry standard versus radar- activated systems. The effectiveness of existing deterrents may be enhanced with development of compensation ponds (Read 1999, Donato et al. 2007), providing clean water and a positive stimulus for birds deterred from tailings ponds. Government should assume responsibility for development of systematic monitoring and re- search on tailings pond bird landing, oiling, and mortality rates. The work should be conducted by independent scientists using a statistically valid sampling design with emphasis on spring and fall migration. A rigorous and systematic monitoring plan standardized across all facilities is likely to yield a better understanding of the factors contributing to avian mortality at tailings ponds than ad hoc monitoring. These data would be Timoncy and Romani • BIRD MOR TALITY AT TAILINGS PONDS 575 valuable lo development and rerinenienl of effective mitigation strategies (e.g., deterrents and compensation ponds) as a component of an adaptive management approach towards reducing avian mortalities. Well-designed monitoring pro- grams help managers and policy makers to reach informed decisions based on facts (McComb et al. 2010). Systematic monitoring and accurate, time- ly reporting would provide data useful to all concerned with bird conservation and manage- ment in the tar sands region. ACKNOWLEDGMENTS We thank C. C. St. Clair, the editor, and two anonymous reviewers for helpful comments that improved this manuscript. RAR was supported by the Killam Trust, Dalhousie University, during the writing of this paper. LITERATURE CITED Alberta Energy. 2009. Oil sands. Alberta Energy, Edmonton, Alberta, Canada. Blokpoel, H. 1973. Bird migration forecasts for military air operations. Occasional Paper 16. Canadian Wildlife Service, Ottawa, Ontario, Canada. Blokpoel, H. and J. Burton. 1973. Weather and height of nocturnal migration in east-central Alberta; a radar study. Bird Banding 46:31 1-328. Boag. D. a. and V. Lewin. 1980. Effectiveness of three waterfowl deterrents on natural and polluted ponds. Journal of Wildlife Management 44:145-154. Bomford, M. and P. H. O’Brien. 1990. Sonic detenents in animal damage control: a review of device tests and effectiveness. Wildlife Society Bulletin 18:411^22. Brough T. and C. J. Bridgman. 1980. An evaluation of long grass as a bird deterrent on British airfields. Journal of Applied Ecology 17:243-253. Butterworth, E., a. Leach, M. Gendron, B. Pollard, and G. R. Stewart. 2002. Peace-Athabasca Delta Waterbird Inventory Program: 1998-2001. Ducks Unlimited Canada, Ottawa, Ontario, Canada. Canadian Broadcasting Corporation (CBC). 2008. Few survivors after 500 ducks take dip in Alberta oil sands waste. CBC News, Edmonton, Alberta, Canada. www.cbc.ca/canada/north/story/2008/04/30/ducks- follo.html?ref= rss Canadian Broadcasting Corporation (CBC). 2010. Syncrude ducks death trial. CBC News, Edmonton, Alberta, Canada, http://www.cbc.ca/canada/cdmonton/ story/201 0/03/24/f-edm()nton-i ndepth-syncrude-ducks- trial.html Clark, R. B. 1984. Impact of oil pollution on seabirds. Environmental Pollution (Series A) 33:1-22. Conover, M. 2001. Resolving human-wildlife conHicts: the science of wildlife damage management. CRC Press LLC, Boca Raton, Florida, USA. Donato, D. B., O. Nichols, H. Possingham, M. Moore, P. F. Ricci, and B, N. Noller. 2007. A critical review of the effects of gold cyanide-bearing tailings solutions on wildlife. Environment International 33:974-984. Dyke, G. R., D. A. Birdsall, and P. L. Sharp. 1976. Tesi of a bird deterrent device at a tailings pond, Athabasca Oil Sands, 1974. Professional Paper 1976-1. Syncrude Canada Ltd., Calgary, Alberta, Canada. Gentes, M.-L. 2006. Health assessment ol Tree Swallows (Tachycineta hicolor) nesting on Athabasca oil sands, Alberta. Thesis. University of Saskatchewan, Saska- toon, Canada. Colder Associates Ltd. 2000. Report on oil sands tailings pond bird deterrent system.s-a review of research and current practices. Suncor Energy Inc. (Oil Sands), Syncrude Canada Ltd., and Albian Sands Energy Inc., Calgary, Alberta, Canada. Gulley, J. 1980. Factors influencing the efficacy of human effigies in deterring waterfowl from polluted ponds. Thesis. University of Alberta, Edmonton, Canada. Hartung, R. 1967. Energy metabolism in oil-covered ducks. Journal of Wildlife Management 31:798-804. King, J. and L. I. Bendell-Young. 2000. The toxicolog- ical significance of grit ingestion to juvenile Mallard ducklings. Journal of Wildlife Management 192:181- 193. McComb, B., B. Zuckerberg, D. Vesely, and C. Jordan. 2010. Monitoring animal populations and their habi- tats: a practitioner’s guide. CRC Press, Boca Raton, Florida, USA. McEwan, E. H. and a. F. C. Koelink. 1973. The heat production of oiled Mallards and scaup. Canadian Journal of Zoology 5 1 :27-3 1 . McLaren, M. A. and P. L. McLaren, 1985. Bird migration watches on Crown Lease 17, Alberta, Fall 1984. Environmental Research Monograph 1985-2. Syncrude Canada Ltd., Calgary, Alberta, Canada. Mead, C. 1997. Poor prospects for oiled birds. Nature 390:449-450. Newton, I. 2007. The migration ecology of birds. Elsevier Ltd., New York, USA, Read, J. L. 1999. A strategy for minimizing waterfowl deaths on toxic ponds. Journal of Applied Ecology 36:345-350. Richardson, W. J. 1971. Spring migration and weather in eastern Canada: a radar study. American Birds 25:684-690. Ronconi, R. a. 2006. Predicting bird oiling events at oil sands tailings ponds and assessing the importance of alternate waterbodies for waterfowl: a preliminary assessment. Canadian Field-Naturali.st 120:1-9. Ronconi, R. A. and C. C, St. Clair. 2006. Efficacy of a radar-act ivatcd on-demand system for deterring wa- terfowl from oil sands tailings ponds. Journal of Applied Ecology 43:1 1 1-1 19. Schick, C. D. and K. R. Ambrock. 1974. Waterfowl investigations in the Athabasca Tar Sands Area. Canadian Wildlife Service, Ottawa, Ontario, Canada. Sharp, P. L., D. A. Bird.sai.l, and W. J. Richardson. 1975. Inventory studies of birds on and near Crown Lea.se Number 17. Athabasca Tar Sands, 1974. Environmental Re.search Monograph 1975-4. Syn- crude Canada Ltd., Calgary, Alberta, Canada. 576 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 3. September 2010 Stevens, R. G. and L. Clark. 1998. Bird repellents: development of avian-specific tear gases for resolution of human-wildlife conflicts. International Biodeterio- ration and Biodegradation 42:153-160. Stickley, a. R., D. F. Mott, and J. O. King. 1995. Short- term effects of an inflatable effigy on cormorants at catfish farms. Wildlife Society Bulletin 23:73-77. Syncrude. 2006. Annual report of oil sands development in 2005 and projected for 2006, Mildred Lake Oil Sands Mine. Alberta Environment, Edmonton, Al- berta, Canada. Syncrude. 2008. 2007 Annual reclamation progress tracking report, Mildred Lake and Aurora North oil sands mines. Alberta Environment, Edmonton, Canada. Thomas, R. 2002. An updated, provisional bird inventory for the Peace-Athabasca Delta, northeastern Alberta. BC Hydro, Burnaby, British Columbia, Canada. Timoney, K. P. and P. Lee. 2009. Does the Alberta tar sands industry pollute? The scientific evidence. The Open Conservation Biology Journal 3:65-81. U.S. Department of Interior (USDI). 2009a. Migratory bird mortality in oilfield wastewater disposal facilities. USDI, Pish and Wildlife Service, Cheyenne, Wyo- ming, USA. http://www.fws.gov/mountain-prairie/ contaminants/contaminants 1 b.html U.S. Department OF Interior (USDI). 2009b. Waterfowl breeding population and habitat survey strata esti- mates, Strata 13-18, 20, and 77. USDI, Pish and Wildlife Service, Patuxent, Maryland, USA. http:// mbdcapps.fws.gov/ Van Meer, T. and B. Arner. 1985. Bird surveillance and protection programme, summary of 1984 and 1985 activities. Syncrude Canada Ltd., Calgary, Alberta, Canada. Wayland, M. and J. Smits. 2004. The ecological viability of constructed wetlands at Suncor: population and health-related considerations in birds. Task 5. Pages 48-61 in Northern Rivers Ecosystem Initiative: collec- tive findings. Assessment of natural and anthropogenic impacts of oil sands contaminants within the Northern River Basins (F. M. CONLY, Compiler). Environment Canada, Saskatoon, Saskatchewan, Canada. Wells, J., S. Casey-Lefkowitz, G. Chavarria, and S. Dyer. 2008. Impact on birds of tar sands oil development in Canada’s boreal forest. Natural Resources Defense Council, New York, USA. The Wilson Journal of Ornilhology 122(3):577, 2010 ABUNDANCE AND DISTRIBUTION OF WATERBIRDS IN THE ELANOS OF VENEZUELA FRANCISCO J. VILELLA,'^ MARK S. GREGORY AND GUY A. BALDASSARRE- ABSTRACT. — The Llanos is a significant waterbird site in the Western Hemisphere, but abundance and distribution of waterbirds across this vast region are poorly known, which hampers conservation initiatives. We used point counts along road routes in the Llanos region of Venezuela to examine abundance and distribution of waterbirds during 2000-2(X)2 within five ecoregions across the Llanos. We detected 69 species of waterbirds and recorded 283,566 individuals, of which 10 species accounted for 80% of our observations. Wading birds (Ciconiiformes) represented the largest guild both in numbers of species (26) and individuals (55%), followed by waterfowl (26%), and shorebirds (11%). Five species comprised 62% ot all individuals: Cattle Egret (Bnhulciis ibis). White-faced Whistling Duck (Dendrocygna vidiiata). Black-bellied Whistling Duck (D. autiimnalis). Great Egret (Ardea alba), and Wattled Jacana (Jacana jacana). Wading birds were particularly ubiquitous with at least 21 of 26 species recorded in each of the ecoregions. Species richness (66), proportion of waterbirds detected (54%), and mean number of birds per route (1,459) were highest in the Banco-Bajio- Estero savanna ecoregion. Our study provides the most comprehensive data set available on waterbirds in the Llanos of Venezuela and highlights regions of special conservation concern. Received 20 April 2009. Accepted 7 October 2009. ' U.S. Geological Survey, Cooperative Research Unit, Department of Wildlife and Fisheries, Mississippi State University, Mississippi State, MS 39762, USA. ‘ SUNY College of Environmental Science and Forestry, Syracuse, NY 13210, USA. ^ Current address: Fundacion Jardln Botanico de Merida, Apartado 52, Merida 5101, Venezuela. ■* Corresponding author; e-mail: fvilella@cfr.msstate.edu Erratum The second author’s name was inadvertently left off of the article published in the March 2010 Issue (Volume 122: 102-115). We apologize for this error. The corrected author sequence is shown above. Short Communications The Wilson Journal of Ornithology 122(3):578-582, 2010 Piping Plover Foraging Distribution and Prey Abundance in the Pre-laying Period Jonathan B. Cohen'-^ and James D. Fraser* ABSTRACT. — Migratory birds arriving in breeding areas should select territories that maximize reproduc- tion and survival. Prey available prior to egg laying may be as important as prey availability for chicks later in the season. We sampled benthic and terrestrial prey items in Piping Plover (Charadrins melodiis) foraging habitat soon after they arrived in breeding areas in New York during 2001-2003. Benthic invertebrates in the sand flats were abundant and available to adults, whereas terrestrial arthropods typically used later in the season were sparse in all cover types. Foraging adults selected intertidal sand Bats over other micro- habitats. One benefit of nesting near sand flats apparently is abundant food upon arrival in breeding areas. Protecting habitat between arrival in breeding areas and territory establishment is uncommon but warranted for this threatened species. Received 4 September 2009. Accepted 17 March 2010. Choice of a breeding territory is a key determinant of avian fitness and is made based on proximate cues that, in good habitat, will be correlated with fitness benefits (Fretwell and Lucas 1970). These cues may be related to cover from predators (Benson et al. 2009), intraspecific social attraction (Harrison et al. 2009), nest site availability (Rhodes et al. 2009), and food availability (Earnst and Rothe 2004). Food abundance during egg-laying may be critical both to current and future avian reproductive success via effects on adult energy reserves (Martin 1987). Lack (1954) proposed that breeding should be timed so chicks should hatch at the time of greatest food abundance. However, habitat quality during territory establishment (i.e., between the start of the breeding season and egg-laying) and during brood-rearing may not be correlated (Nooker et al. 2005). Thus, adults cannot forecast availability of food for chicks by prey abundance leading up to and during egg-laying (Kirstan et al. 2007). ' Department of Fisheries and Wildlife Sciences, Virginia Tech, Blacksburg, VA 24061. USA. ^Corresponding author; e-mail: jocohenl@vt.edu Selecting habitat that does not provide adequate food for chicks is not a viable strategy, but food availability for the young may not necessarily be a proximate cue for decisions about territory location and size. Prey used by adults and chicks during the breeding season may be entirely different for some bird species, (Wilson et al. 2004) or partially overlap (Shealer 1998, Jiguet 2002). Diets of adult and chick Eurasian Dotterels (Charadrius morinellus) differed early in the breeding season and became more similar as soft-bodied prey became more prevalent (Gal- braith et al. 1993). Precocial birds may benefit more than altricial species from settlement decisions based on habitat conditions at time of teiritory establishment, rather than on a forecast of future conditions, because they can move their broods to high- quality foraging habitat soon after hatch (Koszto- lanyi et al. 2007). Our objective was to examine the types and distribution of prey available to Piping Plover (Charadrius melodus) adults upon arrival in breeding areas. Foraging adults tend to concentrate in intertidal flats upon arrival at East Coast breeding sites where those habitat features are available (Fraser et al. 2005), whereas after the breeding season chicks forage in a range of upland and intertidal habitats (Elias et al. 2000). We predicted that intertidal flats would provide abundant prey between arrival and egg-laying. METHODS Our study area was West Hampton Dunes, New York, USA (40 46.5' N, 72 43' W) from 2001 to 2003. This 2.8-km long barrier island village was 200-500 m wide, bounded by the Atlantic Ocean to the south and Moriches Bay to the north. The ocean shoreline of the island was characterized by a relatively linear high-wave energy intertidal zone with fresh wrack (washed-up organic debris) at the daily high tide line. The bay shoreline of the island was a low-wave energy intertidal zone, including portions of linear shoreline interrupted 578 SHORT COMMUNICATIONS 579 by 27 ha of irregularly shaped sand flats with fresh wrack at the daily high tide line. Piping Plover adults arriving to breed on the Atlantic Coast forage primarily on intertidal sand Hats and mudflats that may or may not be contiguous with sandy beaches used for nesting (Cairns 1982, Fraser et al. 2005). Nesting pair density is higher at sites that contain sand flats adjacent to nesting areas than at sites that do not (Patterson et al. 1991, Elias et al. 2000, Cohen et al. 2009), suggesting a fitness benefit of nesting near sand flats. Adults forage on benthic poly- chaetes and crustaceans (Cairns 1977) and terrestrial arthropods (Shaffer and Laporte 1994). Adults lead chicks to foraging sites after the eggs hatch where they prey primarily on adult arthropods (Cuthbert et al. 1999); chicks generally do not prey on benthic invertebrates until they are near fledging. Sand flats provide abundant arthropods in summer (Cohen et al. 2009), but nesting Piping Plovers arrive before insects typically emerge. Thus, prey for chicks cannot be a proximate influence on territory selection. We collected 10-cm diameter X 2-cm deep sediment cores from randomly-selected sites from the bay side intertidal flats using sections of PVC pipe. We collected all samples on 1 day, within a week of the first large annval of Piping Plovers. Piping Plovers forage on invertebrates at or just under the surface, and our cores represented the zone with the most readily available prey. Sediment samples were stored in 1-L Nalgene jars filled with 100% ethanol. We later sorted organisms from the sediment samples and counted them by general category (polychaete worms, crustaceans, insect larvae, and other organisms). We sampled adult arthropods in several cover types on the ocean and bay side of the island using paint stirrers coated, except for the handles, with Tanglefoot Insect Trap Coating (The Tanglefoot Company, Grand Rapids, MI, USA). These cover types included intertidal zone (the zone between the water and the day’s high tide line, which on the bay side contained the sand flats), fresh wrack (washed up organic debris from the most recent high tide), old wrack (clumps of sand-covered wrack deposited by earlier high tides), backshore (flat, dry sandy area between the mean high tide line and the dune, including bare and vegetated substrates), and dune (a low sandy ridge). We placed one pair of paint stirrers in each cover type on each of six transects, the first of which was placed at random within 100 m of the edge of the site and the rest were evenly-spaced at 42()-m intervals. We placed one stirrer vertically in the sand with the uncoated handle buried and the flat surface facing the water’s edge. The other stirrer was horizontal on the ground 10 cm south of the vertical stick with its long axis parallel to the water’s edge. The area exposed was 64.5 cm^ (21.5 X 3 cm) for the horizontal stick (coated on the upper side only) and 1 29 cm^ for the vertical stick (coated on both sides, Loegering and Fraser 1995). We collected sticky trap samples in 1 day within 1 week of collecting sediment samples. The sticks were exposed for 3 hrs between mid- falling and mid-rising tide, starting between 0700 and 1000 hrs, after which we recorded the number of arthropods in different taxa. We performed a census of Piping Plovers in the study area concuirent with arthropod sampling, and recorded the numbers that were foraging in 2002 and 2003. We calculated the proportion of birds in each cover type on the bay and ocean side of the island, including upland and intertidal areas. We also measured the area of the intertidal zones and upland zones (the aggregation of non- intertidal cover types) from aerial photographs taken before 5 May each year by tracing polygons on digitized versions of the photographs in ArcView 3.1 (ESRI, Redlands, CA, USA). We then calculated the percent area of those cover types. RESULTS We observed the first Piping Plovers at the site between 12 and 1 5 March with larger numbers (~ 20) arriving between 18 and 21 March. The earliest recorded nest at the site was on 17 April. We found low numbers of ten'estrial arthro- pods, and many sticky traps caught no arthropods (Table 1). By contrast, there were abundant benthic invertebrates in the bayside intertidal zone (Table I ). No .sediment cores were totally devoid of organisms. The benthos in the cores comprised mainly polychaete worms, crustaceans, and insect larvae with the proportions of each depending on year (Fig. I ). We observed 93 ± 1% (x ± SE) (/? = 46) of foraging Piping Plovers in the bay intertidal zone in 2002, which comprised 32% of the habitat area. The other 7 ± 1% were using the ocean intertidal zone which comprised 8% of the habitat area. The remaining 60% of the area was unused upland. We only observed seven foraging Piping Plovers 580 THE WILSON JOURNAL OL ORNITHOLOGY ♦ Vol. 122. No. 3, September 2010 Year ■ Polychaetes □ Insect larvae □ Crustaceans FIG. 1. Mean percent abundance of polychaete worms, insect larvae, and crustaceans in benthic cores from intertidal sand flats in a Piping Plover breeding area. West Hampton Dunes, New York, 2001-2003. in 2003, all of which were in the bay intertidal zone, which comprised 38% of the habitat area. DISCUSSION The prey base when Piping Plovers arrived in breeding areas largely consisted of intertidal zone benthos. We did not have the resources to sample and analyze the ocean intertidal zone benthos, but in casual observations we noted Piping Plovers foraging on worms there. The plovers were distributed between two habitats (sand flats and ocean intertidal zone) that provided benthic prey. The substrate in upland areas was dry sand and TABLE 1. Prey organisms in Piping Plover foraging .samples. West Hampton Dunes, New York, 2001-2003. habitat, caught in terrestrial sticky trap samples and benthic core Prey type Cover type Date n No. >0-' Organism counts Min Max Median Terrestrial Bay ITZ" 19 Mar 2001 5 0 (0%) 0 0 0 24 Mar 2002 5 2 (40%) 0 5 0 25 Mar 2003 5 0 (0%) 0 0 0 Ocean ITZ 19 Mar 2001 6 0 (0%) 0 0 0 24 Mar 2002 6 1 (17%) 0 1 0 25 Mar 2003 6 0 (0%) 0 0 0 UplantF 19 Mar 2001 40 1 (3%) 0 1 0 24 Mar 2002 43 20 (47%) 0 8 0 25 Mar 2003 48 4 (8%) 0 2 • 0 Benthic Bay ITZ 18 Mar 2001 6 6 (100%) 3 18 1 1 24 Mar 2002 6 6 (100%) 4 135 43 ' 25 Mar 2003 9 9 (100%) 1 1 349 34 " No. >0 = miniber of samples (sticky traps or sediment cores) with at least one prey organism. Uphmd = pooled samples from all stratii (cover types) at or above the tide line. These include fresh wrack, old wrack, backshorc (bare and vegetated strata), and dune. SHORT COMMUNICATIONS 581 would not be expected to host benthos. Terrestrial arthropods were virtually absent from all zones, even though large numbers are caught in sticky traps later in spring and summer (Cohen et al. 2009). Thus, early season prey abundance and distribution could affect Piping Plover territory location and size, and could help explain why nesting pair density of Piping Plovers is high where sand flats are available (Patterson et al. 1991, Elias et al. 2000, Cohen et al. 2009). There often is a tidal lag between the bay and ocean side of baiTier islands as water passes through tidal inlets, and the presence of bay side intertidal flats could increase the potential time each day that benthic organisms are available as prey. A connection between early season foraging opportunities and fitness may not require adults to forecast prey availability for chicks. Reproductive success in many avian studies decreases with time from onset of breeding (Martin 1987). Insect abundance during egg-laying and chick-rearing were uncon-elated for Tree Swallows (Tachyci- neta bicolor), but reproductive success was correlated with insect abundance during egg- laying by affecting timing of breeding and egg mass (Nooker et al. 2005). Thus, it likely benefits female Piping Plovers to nest adjacent to high- quality foraging habitat that allows them to quickly enter breeding condition. Maternal nutri- tion has been shown to influence quality of offspring of many avian species with offspring ranging from fully altricial to fully precocial (Blomqvist et al. 1997, Parker 2002, Reynolds et al. 2003, Verboven et al. 2003). Perhaps more importantly in our study area, intertidal flats usually host a higher abundance of arthropods in late spring than other cover types (Loegering and Fraser 1995, Elias et al. 2000, Cohen et al. 2009). Thus, the same cover type provides prey for adults upon amval and different but still abundant prey for chicks later in the season. However, chicks in .some breeding areas cannot access intertidal sand flats, and there is a difference between pre-laying and chick-rearing foraging areas. Piping Plover broods in Virginia without access to bay side flats starved, becau.se the ocean beaches were arthropod-poor and only chicks that accessed intertidal flats survived (Loegering and Fraser 1995). Some adults in that case settled within easy flight range of early- season foraging habitat but, when their eggs hatched, there was insufficient food for their chicks. Sand flats were used by foraging adults in a site adjacent to our study area in New York, but were not available to any broods (Cohen et al. 2009). Chick survival was unaffected in that ca.se (Cohen et al. 2009). Most strategies commonly used for Piping Plover conservation focus on enhancing survival of eggs and chicks to improve productivity (fledglings produced per breeding pair), and protections often are established after birds have settled on their territories (USDI 1996). However, true productivity is number of recruits into the breeding population per breeding pair (Martin 1987). The correlation between recruitment and fledglings produced depends in part on natal site philopatry which can be a function of density and habitat quality (Sedgwick 2004, Wrege et al. 2006), and which was low (~ 12%) at our site (Cohen et al. 2006). Site fidelity of adults and immigration of new breeders may depend on habitat conditions during pre-laying (Wrege et al. 2006). Thus, protecting habitat from disturbance or degradation during the pre-laying period is important. ACKNOWLEDGMENTS The U.S. Army Coips of Engineers, New York District, provided funding for this study, Amy Alsfeld, Elizabeth Eorbus, Sarah Gibson, Stephen Hartsfield, Jed Hayden, Emilie Masiello, Melissa Neely, Dawn Romaine, and Donald Wardwell assisted with data collection. Steven Deso assisted with laboratory work. LITERATURE CITED Benson, T. J., N. M. Anich, J. D. Brown, and J. C. Bednarz. 2009. Swainson's Warbler nest-site selec- tion in eastern Arkansas. Condor 1 1 1:694-705. Blomqvist, D., O. C. Johansson, and E. Gotmark. 1997. Parental quality and egg size affect chick survival in a precocial bird, the Lapwing Vanelliis vanelliis. Oeco- logia 110:18-24. Cairns, W. E. 1977. Breeding biology and behaviour of the Piping Plover (Charadrius melodits) in southern Nova Scotia. Thesis. Dalhousie University. Halifax, Nova Scotia, Canada. Cairns, W. E. 1982. Biology and behavior of breeding Piping Plovers. Wilson Bullelin 94:531-545. Cohen, J. B., J. D. Eraser, and D, H. Catlin. 2006, Survival and site fidelity of Piping Plovers on Long Island, New York. Journal of Field Ornitholosy 77:409^17. Cohen, J. B., L. M. Houghton, and J. D. Fraser. 2009. Nesting density and reproductive success of Piping Plovers in relation to storm- and human-created habitat changes. Wildlife Monographs 173. CUTHBERT, F. J.. B. SCHOLTENS. L. C. WeMMER, AND R. McLain. 1999. Gizzard contents of Piping Plover 582 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 3, September 2010 chicks in northern Michigan. Wilson Bulletin 111: 121-123. E.arnst, S. L. and T. C. Rothe. 2004. Habitat selection by Tundra Swans on northern Alaska breeding grounds. Waterbirds 27:224-233. Eli.as, S. P., J. D. Fraser, and P. A. Buckley. 2000. Piping Plover brood foraging ecology on New York barrier islands. Journal of Wildlife Management 64: 346-354. Fr.yser. J. D., S. E. Keane, and P. A. Buckley. 2005. Prenesting use of intertidal habitats by Piping Plovers on South Monomoy Island, Massachusetts. Journal of Wildlife Management 69: 1731-1736. Fretwell, S. D. and H. L. Lucas Jr. 1970. On territorial behavior and other factors intluencing habitat distri- bution in birds. Acta Biotheoretica 19:16-36. Galbraith. H., S. Murray, K. Duncan, R. Smith, D. P. Whitfield, and D. B. A. Thompson. 1993. Diet and habitat use of the Dotteral Charadrhis morinellus in Scotland. Ibis 135: 148-155. Harrison, M. L., D. J. Green, and P. G. Krannitz. 2009. Conspecifics influence the settlement decisions of male Brewer’s Sparrows at the northern edge of their range. Condor 1 1 1:722-729. JiGUET, F. 2002. Arthropods in the diet of Little Bustards Tetrax tetrax during the breeding season in western France. Bird Study 49:105-109. Kirstan 111. W. B., M. D. Johnson, and J. T. Rotenberry. 2007. Choices and consequences of habitat selection for birds. Condor 109:485-488. Kosztolanyi, a., T. Szekely, and I. C. Cuthill. 2007. The function of habitat change during brood-rearing in the precocial Kentish Plover Charadriits cdexandrinus. Acta Ethologica 10:73-79. Lack, D. 1954. The natural regulation of animal numbers. Oxford University Press, Oxford, United Kingdom. Loegering, J. P. AND J. D. Fraser. 1995. Factors affecting Piping Plover chick survival in different brood-rearing habitats. Journal of Wildlife Management 59: 646-655. Martin, T. E. 1987. Food as a limit on breeding birds: a life-history perspective. Annual Review of Ecology and Systematics 18:453-^87. Nooker, J. K., P. O. Dunn, and L. A. Whittingham. 2005. Effects of food abundance, weather, and female condition on reproduction in Tree Swallows {Tacliy- cineta bicolor). Auk 122:1225-1238. Parker, T. H. 2002. Maternal condition, reproductive investment, and offspring sex ratio in captive Red Junglefowl (Gallus gallus). Auk 119:840-845. Patterson, M. E., J. D. Fraser, and J. W. Roggenbuck. 1991. Factors affecting Piping Plover productivity on Assateague Island. Journal of Wildlife Management 55: 525-531. Reynolds, S. J., S. J. Schoech, and R. Bowman. 2003. Diet quality during pre-laying and nestling periods inOuences growth and survival of Florida Scrub-Jay (Aphelocoma coeridescens) chicks. Journal of Zoolo- gy, London 261:217—226. Rhodes, B., C. O’Donnell, and I. Jamieson. 2009. Microclimate of natural cavity nests and its implica- tions for a threatened secondary-cavity-nesting pas- serine of New Zealand, the South Island Saddleback. Condor 111: 462^69. Sedgwick, J. A. 2004. Site fidelity, territory fidelity, and natal philopatry in Willow Flycatchers (Empidonax trailli). Auk 121:1 103-1 121. Shaffer, F. and P. Laporte. 1994. Diet of Piping Plovers on the Magdelen Islands, Quebec. Wilson Bulletin 106: 531-536. Shealer, D. a. 1998. Differences in diet of chick provisioning between adult Roseate and Sandwich terns in Puerto Rico. Condor 100:131-140. U. S. Department of Interior (USDI). 1996. Piping Plover (Cluiradriiis melodu.s), Atlantic Coast Popula- tion. Revised Recovery Plan. USDI, Fish and Wildlife Service, Hadley, Massachusetts, USA. Verboven, N., P. Monaghan, D. M. Evans, H. Schwabl, N. Evans, C. Whitelaw, and R. G. Nager. 2003. Maternal condition, yolk androgens and offspring performance: a supplemental feeding experiment in the Lesser Black-backed Gull (Lcinis fuscus). Proceed- ings of the Royal Society of London. Series B 270: 2223-2232. Wrege, P. H., W. D. Shuford, D. W. Winkler, and R. Jellison. 2006. Annual variation in numbers of breeding California Gulls at Mono Lake, California: the importance of natal philopatry and local and regional conditions. Condor 108: 82-96. Wilson, L. J., F. Daunt, and S. Wanless. 2004. Self- feeding and chick provisioning diet differ in the Common Guillemot Uria aedge. Ardea 92: 197-208. SHORT COMMUNICATIONS 583 The Wilson Journal of Ornithology 1 22(3):.‘i83-5K7, 2010 Interspecific Song Imitation by a Cerulean Warbler Than J. Boves,'-* David A. Buehler,' and Phillip C. Massey^ ABSTRACT. — We report a male Cerulean Warbler (Dendroica cernlea) singing a highly accurate version of another species’ song (Hooded Warbler; Wilsonia citrina), as well as a typical Cerulean Warbler song. We performed playback experiments to examine the manner in which this bird used, and responded to, the different songs. He responded aggressively (wing flicks and garbled song), and with his Cerulean Warbler song, to recorded versions of two different types of Cerulean Warbler songs, including a type that he was not observed singing. He did not respond in any manner to playback of Hooded Warbler song. Received 5 October 2009. Accepted 2 February’ 2010. Few cases of high-quality interspecific imita- tion exist in the wild, outside of natural voice mimics. Differences in acquired or innate auditory templates are thought to limit interspecific song learning (Marler 1976, 1984). Interspecific imita- tion, when it does occur, typically is between closely-related species or species that compete for space and resources (Kroodsma 2004). Members of the Parulidae, of non-mimicking species, are among those best known to learn and copy songs of other species (Payne et al. 1984, Spector 1992, Martin et al. 1995). Males in many genera of Parulidae (including Dendroica) sing two distinct song types that may .serve different functions (Kroodsma 1981, Spec- tor 1992). Type I songs, in the case of Cerulean Warblers (Dendroica certdea), are primarily used in an intersexual context (e.g., mate attraction) and are sung more often early in the breeding season when the male is unmated (Woodward 1997). Type II songs are tho.se that typically have an intrasexLial function (e.g., territorial defense) and are sung more often later in the breeding season, after a mate is secured (Woodward 1997). We describe a case of a male Cerulean Warbler singing an indistinguishable version of another ' Department of Forestry. Wildlife, and Fisheries, University of Tennessee, Knoxville, TN 37996, USA. ■701 White Pine Drive. Hendersonville, NC 28739, USA. ^Corresponding author; e-mail: tboves@utk.edu species’ song (Hooded Warbler; Wilsonia citrina), as well as a typical Cerulean Warbler song. We also describe the results of playback experiments performed with this individual. OBSERVATIONS On 25-27 June 2009, we observed an after- second-year, unpaired male Cerulean Warbler in the North Cumberland Wildlife Management Area (36° 12' 41.1" N, 84° 16' 55.1" W), in the Cumberland Mountains of Tennessee, USA, sing- ing both a Cerulean Warbler song and an accurate copy of a Hooded Warbler song. We delineated this individual’s temtory and ascertained his pairing status based on extensive spot-mapping, nest searches, and banding efforts. Two other paired male Cerulean Warblers maintained territories adjacent to this individual but, to the best of our knowledge, this individual was unpaired and did not breed during the 2009 season. We did not observe him singing a Hooded Warbler song until 25 June, although he had been defending this territory since early May. The high density of Hooded Warblers, and the quality of this individ- ual’s imitation, served to mask this behavior from our observations until this occasion. We first ob.served this individual on 25 June 2009 at 1500 hrs singing a Cemlean Warbler song at a rate of six/min over a 5-min period (Fig. 1 ). At 1505 hrs, he began singing a Hooded Warbler song at a rate of eight/min over a 1 5-min period (Fig. 2). At 1525 hrs, he began to sing the Cerulean Warbler song again for a 10-min period. He continued to alternate between songs for the next 15 min. We then attempted to capture the bird using playback ot a similarly structured Cerulean Warbler song, a Cerulean Warbler decoy placed at 2 m, and a mist- net. We placed the speakers at ground level and adjusted the volume of the speakers to the approximate intensity level produced by the bird. The bird immediately responded with a typical Cerulean Warbler vocal response to conspecific playback (a garbled, softly sung. Cerulean Warbler song). He also performed the same Cerulean Warbler song that he was observed singing earlier. 584 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 3. September 2010 7 500- 7 000 6 500 6 000 5500- 5000 ^ 4500- I 5“ 4 000- c 0 3 1 3 500- u. 3 000- 2500- 2 000 1 500 1 000 0 500 —\ r 01 02 03 04 05 06 07 09 1 11 12 13 14 15 16 17 18 19 Time (s) FIG. 1. Spectroaram of Cerulean Warbler song, performed by the bird in question. This song was indistinguishable by ear from songs performed by other male Cerulean Warblers. A previously recorded version of this song type was played to the male Cerulean Warbler in question on 25 June 2009 and he responded aggressively. Physically, he responded typically as well, with nervous wing flicks and foraging imitations as he approached within 5 m of the speakers and decoy. The bird was caught shortly thereafter, banded, photographed, and released. He appeared to be a phenotypically normal, after-second-year male Cerulean Warbler (based on Pyle 2000), displaying traits that may be indicative of age and good condition (e.g., a strong breast band and large amount of white spotting on his tail; TJB, unpubl. data). The bird returned to singing a Hooded Warbler song 4 min after release. FIG. 2. .Spectrogram of Hooded Warbler song performed by the Cerulean Warbler in question. This song was indistinguishable by ear from actual Hooded Warbler songs. SHORT COMMLINICATIONS 585 8 500 6 000 7 500 7 000- 6 500 6 000 5 500- 5 000 f 4000-1 3 500- 3 000 2 500 2 000 1 500- 1 000- 0 500 09 1 11 Time (s) 12 13 14 15 16 17 18 19 FIG. 3. Spectrogram of alternative Cerulean Warbler song type used in playback experiment with the bird in question on 26 June 2009. He responded to this song type in the same aggressive manner as he responded to the Cerulean Warbler song type that he was heard singing. We returned to the location on 26 June and made sound recordings of both songs using a Dan Gibson Parabolic Microphone and a M-Audio Microtrack II digital recorder. We detected the bird at 0800 hrs, without use of playback, singing the same type of a Cerulean Warbler song. He sang this song for 15 min at a rate of six/min. After that, he sang his Hooded Warbler song for 9 min, at a rate of nine/min. Following our recording of the bird’s repertoire, we broadcast a recorded version of a similar Hooded Warbler song for 10 min and he did not respond, vocally or physically. We then broadcast an alternate version of a Cerulean Warbler song with no decoy (Fig. 3). The bird again responded within 2 min, approaching the speakers and di.splaying similar aggressive behavior (wing flicks, garbled song). He then retreated into the canopy and sang his Cerulean Warbler song (Fig. 1 ). We returned on 27 June to see if he would re.spond to Hooded Warbler playback and, once again, he did not. We did, however, observe him again singing the same type of Cerulean and Hooded warbler songs. We recorded individual male Cerulean War- blers earlier in the season for unrelated research and, after these observations, we recorded neigh- boring Hooded Warblers for comparison (Figs. 4, 5). We created spectrograms of all songs using Raven Pro acoustic analysis software (Cornell Laboratory of Ornithology, Ithaca, NY, USA). DISCUSSION This individual likely learned his song reper- toire from both Cerulean and Hooded warbler tutors. This would be possible at many locations in the Appalachian Mountains, as both Hooded and Cerulean warblers are relatively common in similar habitats. These two species, however, use different strata of the forest canopy (Cerulean Warblers use the overstory and Hooded Warblers use the understory), belong to separate, well- resolved genera (Lovette and Birmingham 2002), and have distinctly different songs. Our observa- tions differ from the pattern of interspecific learning suggested by Kroodsma (2004); imitators usually belong to closely related species or compete for space and resources. There are several species of wood-warblers within the same genus and occupying similar forest strata, which seem more likely to act as interspecific tutors for Cerulean Warblers (e.g.. Black-throated Green [Dendroica virens] or Blackburnian warblers [D. fusca\). Our playback experiment results suggest this individual used, and responded to. Cerulean Warbler song in an intrasexual (Type 11) manner. His response to Cerulean Warbler playback (wing nicks, garbled songs) was typical of teiTitorial behavior by Cerulean Warblers and, during playback bouts, he only performed his Cerulean 586 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 3, Septe))}her 2010 10 500- 10 000- 9 500- 9 000- 8 500- 8 000- 7 500- 7000- 6 500- 5 6 000- c 5 500- 4 D I 5 000- u. 4 500- 4 000- 3 500- 3 000- 2 500- 2 000- 1 500- 1 000- 0 500- 0 1 ^2 ^ oT ~04 05 06 o'? 08 ^ o'b 1 ^ 77 Time (s) — , . 1 1 . 1 . j— 12 13 14 15 16 1 9 FIG. 4. Spectrogram of Cerulean Warbler song performed by a neighboring Cerulean Warbler. Warbler song. This male was unpaired and singing at a late date (highly unusual for after-second-year birds in this region), suggesting that his unusual song repertoire was the cause of his lack of reproductive success. We were unable to ascertain how he used his Hooded Warbler song or how he perceived other Hooded Warbler songs. He did not respond (physically or vocally) to Hooded Warbler song, suggesting that he did not use or perceive Hooded Warbler song in a Type II manner. This individual may be at a disadvantage when defending a high-quality territory because his use N X (B D CT 7 000: 6 500- 6.000: 5.500- 5.000: 4.500 4 000: 3.500 3 000- 2.500- 2 000- 1 500- 1 000: 0 500 02 — I — 04 — I 06 — I — 0.8 1 Time (s) — I — 12 — I — 1 4 ■ '* I -■ 1 ' 1.6 1,8 FIG. 5. Spectrogram of Hooded Warbler song performed by a neighboring Hooded Warbler. SHORT COMMUNICATIONS 587 of an interspecific song may increase the chance ot conspecilic fights, therefore increasing the costs of defense (Martin et al. 1995). This was the case during the 2009 season; the territory he occupied was adjacent to only two conspecific teiTitories, which is low for appropriate habitat in the densely-populated Cumberland Mountains. His teiTitory also had lower than average DBH and canopy height, similar to areas often unoccu- pied by (or occupied by younger male) Cerulean Warblers in the area (TJB, unpubl. data). In the event this individual returns in 2010, we will attempt to observe his interactions with neighboring Hooded Warblers, record changes in song preference throughout the breeding season, and observe changes in responsiveness to play- back of different songs at different times during the season. ACKNOWLEDGMENTS This study was performed in conjunction with the Cerulean Warbler Forest Management Experiment, which is funded by the National Fish and Wildlife Foundation, U.S. Fish and Wildlife Service, the Nature Conservancy, and the National Council for Air and Stream Improvement. We thank D. K. Dawson and M. G. Efford for use of the microphone and recorder. We thank T. M. Freeberg, D. A. Spector, and an anonymous reviewer for helpful comments on this manuscript. LITERATURE CITED Kroodsma, D. E. 1981. Geographic variation and functions of song types in warblers (Parulidae). Auk 98:743-751 . Kroodsma, D. E. 2004. The diversity and plasticity of birdsong. Pages 108-131 in Nature’s music: the science of birdsong. Elsevier Academic Press, San Diego, California, USA. Lovette, 1. J. AND E. Birmingham. 2002. What is a wood- warbler? Molecular characterization of a monophyletic Parulidae. Auk 119:695-714. Marler, P. 1976. Sensory templates in species-specific behavior. Pages 314-329 in Simpler networks and behavior (J. Fentress, Editor). Sinauer, Sunderland, Massachusetts, USA. Marler, P. 1984. Song learning: innate species differences in the learning process. Pages 289-309 in The biology of learning (P. Marler and H. S. Terrace. Editors). Springer-Verlag, Berlin, Germany. Martin, P. R., J. R. Fotheringham, and R. J. Robertson. 1995. A Prairie Warbler with a conspecific and heterospecific song repertoire. Auk 112:770-774. Payne, R. B., L. L. Payne, and S. M. Dohlert. 1984. Interspecific song learning in a wild Chestnut-sided Warbler. Wilson Bulletin 96:292-294. Pyle, P., S. N. G. Howell, R. P. Yunick, and D. F. Desante. 2000. Identification guide to North American passerines. Slate Creek Press, Bolinas, California, USA. Spector, D. A. 1992. Wood-warbler song systems: a review of paruline singing behaviors. Current Orni- thology 9:199-238. Woodward, R. L. 1997. Characterization and significance of song variation in Cerulean Warblers. Thesis. Queen’s University, Kingston. Ontario, Canada. The Wilson Journal of Ornithology 122(3):587-591, 2010 Nests, Nest Placement, and Eggs of Three Philippine Endemic Birds Luis A. Sanchez-Gonzalez,'-^’^ Carl 01iveros,‘ “ Nevong Puna,-^ and Robert G. Moyle' ABSTRACT. — We describe the nest and eggs of two Philippine endemic passerines, the Rusty-faced Babbler {Robsoniits rabori) and the Blue-breasted Blue Fly- ' Natural History Museum and Biodiversity Research Center, Dyche Hall, University of Kansas, 1345 Jayhawk Boulevard, Lawrence, KS 66045, USA, ~9 Bougainvillea Street, Manuela Subdivision, Las Pinas City, Philippines 1740. ^46-C Brgy Maningning, Lacao Street, Puerto Princesa City, Palawan, Philippines 5300. ■‘Museo de Zoologia “Alfonso L. Herrera’’, Departa- mento de Biologia Evolutiva, Facultad de Ciencias, Universidad Nacional Autonoma de Mexico, Apartado Postal 70-399, Mexico D.F., 04510, Mexico. ^Corresponding author; e-mail: lasg@ku.edu catcher (Cyornis herioti). We also describe a novel type of nest and nest placement for the Short-crested Monarch (Hypothymis helenae), and provide the first description of the eggs of this species. We discuss similarities among eggs and nests of these species with their relatives, and the need for more information regarding natural history and breeding habits for the Philippine avifauna. Received 13 October 2009. Ac- cepted 10 February 2010. Ornithologists have long been aware of the distinctiveness of the Philippine avifauna, which is reflected in the high percentage of endemism in the archipelago (Dickinson et al. 1991). However, 588 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 3, September 2010 FIG. 1. Left. Nest of Robsonius rahori. Arrow shows the entrance location (Photograph by Nevong Puna). Right. Egg of Robsonius rabori (Photograph by Luis A. Sanchez-Gonzalez). Site: ~12 km WSW of Baler. Municipality of San Luis in Aurora Province (15° 39.22' N, 121° 30.44' E), June 2009. the Philippine avifauna is one of the most poorly known in the world (Kennedy et al. 2000). Breeding habits for some endemic species are well known, and the first descriptions of the breeding habits or nestlings of other .species are starting to appear in the literature (Strijk 2004). However, breeding habits for a large portion of the Philippine avifauna remain undocumented. Dickinson et al. (1991) estimated that —395 species may breed in the country and 157 of these apparently lack any breeding information. Not surprisingly, 77 of the.se poorly known species are endemic, which repre.sents about one half of the Philippine endemics. Knowledge of species natural history and nesting biology is fundamental for development of sound protection strategies. This information is also useful for taxonomic purposes. For example, studies on nest architecture can provide a new set of taxonomic characters used for phylogenetic analysis (Zyskowsky and Prum 1999), while data on clutch size, egg color, and egg patterns can be used to infer evolutionary relationships (e.g., Mayr 1931, Ford 1981, Zelenitsky and Modesto 2003). We report the first description of the nest, nest placement, and eggs for two species of Philippine endemic birds, Cyornis herioti and Robsonius rabori, and provide the first description of the eggs of Hypothymis helenae as well as a completely new type of nest placement for this species. METHODS Study Area. — All data for this paper were collected during a survey conducted from mid May to late June 2009 in the Sierra Madre Mountains of Aurora Province, in the northeastern part of Luzon Island. The main vegetation type in the area is dipterocarp forests, although montane forests are found above 1,200 m. The three sites were at elevations ranging from 50 m at Baler to 530 m at the two other sites visited. The average SHORT COMMUNICATIONS 589 FIG. 2. Left. Location of the nest cavity of Cyornis herioti. The white arrow shows the cavity in the rock (Photograph by Luis A. Sanchez-Gonzalez). Right. Nest and eggs of Cyornis herioti', Aurora National Park, Aurora Province (15° 67.80 N, 121° 33.61 E), May 2009. Photograph by Robert G. Moyle). rainfall is 300 mm per year and is evenly distributed throughout the year (Dickinson et al. 1991). Field Procedures. — Data were gathered during observation walks spanning the hours of avian activity. Observations were made with binoculars (7 X 35, 10 X 50) and we used a specialized field guide (Kennedy et al. 2000) for bird identifica- tions. NEST DESCRIPTIONS Robsoniiis rabori. — The Rusty-faced Babbler (Timaliidae) is an uncommon and local species known only from the northern half of Luzon Island (Collar and Robson 2007), where it occurs in rocky or part-rocky areas in lowland forest and second growth up to 1,000 m (Kennedy et al. 2000). Almost nothing is known about the breeding habits of this elusive species, and its breeding dates have only been extrapolated from observations of fledglings in February and March, and in May and August (Kennedy et al. 2000). We found a nest of this species on 1 8 June 2009 along an overgrown dirt logging road in second growth forest at 525 m, ~12 km WSW of Baler, Municipality of San Luis in Aurora Province (15° 39.22' N, 121° 30.44' E). The nest consisted of a large clump of dry sticks, branches, and leaves with a front entrance (8 cm in diameter). The dome was 19 cm wide and propped up by twigs 50 cm from the ground against a bank of rock and mud (Fig. 1 Left). The nest was <1 m from a 0.5 m high rock slide. A small stream was —30 m from the nest down a steep slope. The nest contained a single elongated egg (28.9 X 18.55 mm) with a whitish-cream base color and gray and light-brown blotches (Fig. 1 Right). The female occupant of the nest was collected (catalog number KUNHM 114678); another egg about to be laid was in its oviduct indicating a clutch size of two eggs. Both are now in the Natural History Museum of the University of Kansas (KUNHM). The nest architecture of this species is similar to that reported for a R. .sorsogoueusis (Collar and Robson 2007); they build large balls of clumped dried twigs and leaves. This architecture differs from babblers in the genera Trichastoma, Keiio- pio, and Turdinus, which currently stand close to Robsoniiis in linear classifications, but is similar to that of Gypsophilo and Napothera, which also build dome nests (Collar and Robson 2007). However, the eggs of R. rabori superficially 590 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 3. September 2010 FIG. 3. Nest and eggs of Hypothymi.s helenae inside a tree-fern stump. Baler, Aurora Province ( 14° 44' N, 121° 33' E), June 2009. Photograph by Luis A. Sanchez-Gonzalez. resemble those of the genus Tiirdinus, as illus- trated in the case of T. macrodactylus from Java (Hoogerwerf 1949). Cyornis herioti. — The Blue-breasted Blue Fly- catcher (Muscicapidae) is endemic to Luzon and Catanduanes islands where it inhabits primary and selectively-logged forests between 100 and 1,200 m (Taylor 2006). This sexually dimorphic species occurs in mixed species flocks, solitary or in pairs, and its breeding habitats are mostly unknown. The breeding season, which may occur in May, has been inferred from observations of individuals with enlarged gonads, nestlings, and recently Hedged juveniles. However, both the nest and eggs are unknown (Kennedy et al. 2000, Taylor 2006). A nest and eggs of this flycatcher were found on 23 May 2009 in Aurora National Park, Aurora Province (15 67.80 N, 121° 33.61 E). The nest was in a shallow cavity in a 4-m tall rock along the course of a river (Fig. 2 Left). The nest was cup- shaped and had thick walls (inside dimensions 60 X 50 mm; outer dimensions 100 X 120 mm). The inside was lined with fern fronds and the outside was covered with moss. Depth of the nest cup was 35 mm, whereas the total depth of the nest was —55 mm. The nest contained two broken eggs and one whole egg (21 X 16 mm), which were light brown with slightly darker specks (Fig. 2 Right). The female was observed incubating the eggs during the night. All attributes of the nest including placement in a shallow cavity and clutch size seem to be typical of nesting habits in the genus, as well as in other genera of Muscicapidae, such as Cyanoptila and Eiimyias (Taylor 2006). Hypothymi.s helenae. — The Short-crested Mon- arch is an endemic species of the Monarchidae that inhabits humid forests up to 1,000 m, where it usually occurs singly or in pairs in mixed species flocks. Breeding has been reported to occur around August, although some individuals taken in April and May also seemed to be in breeding condition as inferred from enlargement of the gonads (Kennedy et al. 2()()(), Coates et al. 2006). SHORT COMMUNICATIONS 591 Allen et al. (2006) described the nest of this species from Camiguin Norte Island but eggs were not observed. We found a nest of this species on 8 June 2009 in secondary growth forest at Baler, Aurora Province (14° 44' N, 121° 33' E). The nest was inside the broken trunk of a tree fern —70 cm trom the ground. The cavity was 55 mm deep and 65 mm in diameter. Nest material included dry leaves and small branches of tree fern (Fig. 3), all roughly shaped into a cup. The nest contained two eggs, one of which was laid the day after discovery of the nest. The eggs (21 X 16 mm) were bluish green with brown and dark brown blotches; the distal part of the egg was heavily blotched. The female, which has a different plumage from the male, was flushed from the nest at dusk of the second day. Nest architecture and placement of this species differs somewhat from its conspecific H. helenae personam from Camiguin Norte and its close relative H. azurea, for which nests are typically cup-shaped structures decorated outside with lichen and moss, usually in the fork of a tree branch (Hoogerwerf 1949, Allen et al. 2006, Coates et al. 2006), whereas the heretofore undescribed nest of H. helenae helenae from Luzon was placed in a hollow trunk. Nest architecture and nest placement within the genus Hypothymis, and specifically within H. helenae, needs further study. Natural history and phyloge- netic studies of the genus are also needed to examine the significance of these differences. ACKNOWLEDGMENTS LASG thanks Consejo Nacional de Ciencia y Tecnologi'a (CONACyT) for a postdoctoral fellowship in the program “Estancias postdoctorales y sabaticas al Extranjero para la consolidacion de grupos de investigacion” Segunda Etapa (Expediente 93730) at the University of Kansas. This study was conducted with financial support from the National Science Foundation (DEB 0743491 to RGM). Michael Andersen greatly improved the English. We afso thank C. E. Braun. Bob Kennedy, and an anonymous reviewer whose comments considerably improved this manuscript. LITERATURE CITED Allen, D., C. Espanola, G, Broad, C. Oijvhros, and J. C. T. Gonzalez. 2006. New bird records for the Babuyan Islands, Philippines, including two first records for the Philippines. Forktail 22:57-70. Coates, B. J., G. C. L. Dutson, C. E. Filardi, P. Clement, P. A. Gregory, and C. W. Moeliker. 2006. Family Monarchidae (monarch-flycatchers). Pages 244-329 in Handbook of the birds of the world. Volume 11. Old World flycatchers to Old World warblers (J. Del Hoyo, A. Elliott, and D. A. Christie, Editors). Lynx Edicions, Barcelona, Spain, Collar, N. J. and C. Robson. 2007. Family Timaliidae (babblers). Pages 70-291 in Handbook of the birds of the world. Volume 12. Picathartes to tits and chickadees (J. Del Hoyo, A. Elliott, and D. A. Christie, Editors). Lynx Edicions, Barcelona, Spain. Dickinson, E. C., R. S. Kennedy, and K. C. Parkes. 1991 . The birds of the Philippines. An annotated check-list. B.O.U. Check-list Number 12. British Ornithologists’ Union, Tring, United Kingdom. Ford, J. 1981. Evolution, distribution and stage of speciation of Rhipidura fuliginosa complex in Aus- tralia. Emu 81:128-144. Hoogerwerf, A. 1949. Een bijdrage tot de oologie van het eiland Java. Uitgave van de Kon. Plantentuin van Indonesie, Buitenzorg, Java, Indonesia. Kennedy, R. S., P. C. Gonzales, E. C. Dickinson, H. C. Miranda, and T. H. Fisher. 2000. A guide to the birds of the Philippines. Oxford University Press, Oxford. England. Mayr, E. 1931. Birds collected during the Withney South Sea Expedition XVI. Notes on the genus Rhipidura. American Museum Novitates 502:1-21, Striik, j. S. 2004. Description of the nest and nestling of Great Eared Nightjar Eurostopodus macrotis from Luzon. Philippines. Forktail 20:128-129. Taylor, B. 2006. Family Mu.scicapidae (Old World flycatchers). Pages 56-163 in Handbook of the birds of the world. Volume 1 1 . Old World flycatchers to Old World warblers (J. Del Hoyo, A. Elliott, and D. A. Christie, Editors). Lynx Edicions. Barcelona. Spain. Zelenitsky, D. K. and S. P. Modesto. 2003. New information on the eggshell of ratites (Aves) and its phylogenetic implications. Canadian Journal of Zool- ogy 81 :962-970. Zyskowsky, K. and R. Prum. 1999. Phylogenetic analysis of the nest architecture of neotropical ovenbirds (Furnariidae). Auk 118:891-911. 592 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 3. September 2010 The Wilson Journal of Ornithology 122(3):592-597, 2010 Nests, Eggs, and Young of the Azure-crowned Hummingbird {Amazilia cyanocephala) Juan Francisco Ornelas' ABSTRACT. — The first of several nests (previously undescribed) of the Azure-crowned Hummingbird {Amazilia cyanocephala) was found on 20 Lebruary 2009 along the Rio Pixquiac in Coatepec, Veracruz on the Gulf of Mexico side of the Sierra Madre Oriental of Mexico. The cup-shaped nest consisted of fibers and scales of tree ferns with the outside covered with liverworts and a few mosses and lichens. It was saddled on a thin, horizontal branch and had one “tail” of hanging liverworts, longer than the height of the nest cup, draped beneath the nest. It eventually contained two white, non-glossy eggs that were long-elliptical in shape, measuring roughly 13.5 X 7.5 mm. Other nests of the species discovered subsequently were similar in construction and placed on branches or substrates of different plant species. Observations of nest building, eggs, and incubation behavior at the first nest were generally consistent with descriptions for other Amazilia hummingbirds. Received 29 September 2009. Accepted 21 Januaty 2010. Little is known about the nests, eggs, incuba- tion, fledglings, and breeding behavior of mem- bers of the genus Amazilia (Skutch 1931, Harverschmidt 1952, Atwood et al. 1991), which comprises ~30 species in Central and South America (AOU 1998, Remsen et al. 2008). Published records of nests and eggs are available for less than half of the species (del Hoyo et al. 1999). Thus, knowledge of nesting of Amazilia is far from complete (Ornelas 1995, Weller 1998). The Azure-crowned Hummingbird (Amazilia cyanocephalct) is a common resident in Mexico from Tamaulipas south through the Isthmus of Tehuantepec to Chiapas, and through Central America to Honduras and northern Nicaragua (Howell and Webb 1995, Johnsgard 1997). This medium-sized hummingbird is characterized by the immaculate white throat and belly with golden green Hanks and indicated breast band, uniformly green to bronzy upperparts and olive-green ' Dcpartamento de Biologi'a Evoluliva, In.stitulo de Hcologi'a. AC. Carrctera antigua a Coatcpcc No. 351, El Haya. 91070 Xalapa. Veracruz., Mexico; e-mail: franciseo. ()rnela.s@inecol.edu.mx centers of under tail coverts, and a bright, glittering blue forehead and crown (more tur- quoise in female-type plumage). It is similar in size and general appearance to the Violet-crowned Hummingbird (A. violiceps), but A. cyanocephala has a blackish upper mandible and a reddish lower mandible with dark tip, whereas the bill of A. violiceps is fleshy red except for the dark tip (Weller 1998). Both taxa are apparently entirely allopatric (Johnsgard 1997); locally sympatric, the White-bellied Emerald (A. Candida) is much smaller and easily distinguished in the field by the dark subterminal band of its tail. The species is fairly common in oak (Quercus spp.) and pine (Finns spp.) woodlands, cloud forests, humid evergreen forest edges, pine savannas, gardens, and brushy and second growth habitats from 500 to about 2,400 m elevation (Davis 1972, Peterson and Chalif 1973, Howell and Webb 1995. Johnsgard 1997). Birds on the Atlantic Slope from southern Tamaulipas to Veracruz in Mexico, and in the interior from eastern Oaxaca and Chiapas to western Nicaragua are assigned to the cyanocephala subspecies. Resident birds of the Mosquitia coastal region in Honduras are assigned to the chlorostephana subspecies with crown glittering green to tur- quoise-green (Howell and Webb 1995). Range of the chlorostephana subspecies below 100 m is tied to the distribution of pine (P. caribae) savanna, stands of pine close to gallery forest, and edges of isolated evergreen rainforest (del Hoyo et al. 1999). Birds from Belize south to northern Nicaragua, formerly separated as guate- malensis, show strong intergradation with nomi- nate birds from Chiapas, Mexico (Howell and Webb 1995). Little information on nesting behavior of the Azure-crowned Hummingbird is available, but two species of the widespread genus Amazilia have been studied (Skutch 1931, Harverschmidt 1952), and major differences from these are unlikely (Johnsgard 1997). Azure-crowned Hum- mingbirds with enlarged gonads have been SHORT COMMUNICATIONS 593 FIG. 1. Ne.sl placement, nests, eggs, and nestlings of Azure-crowned Hummingbird (AmazUia cxanocephala) photographed along the Rio Pixquiac near Xalapa, Veracruz. (A) Completed nest first under construction on 20 February 2009 (nest #1, Table 1 ). (B) Two eggs laid in an interval of 2 days, 25 and 27 February (alternate laying). (C) Eggs hatched almost .synchronously after 15 days of incubation, the first egg hatched by noon on 15 March. (D) Eight-day old ne.stlincrs. (E) Incomplete nest (nest #2) built after failure of nest #1. (F) Ne.st #2 was not completed but the first egg was laid on 5 April. (G) and (H) present two views of nest #5. Photographs by the author. 594 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 3. September 2010 TABLE 1. Measurements (mm) of Amazilia cyanocephala nests and eggs (min and max exterior and interior diameter of cup-shaped nests). Nest measurements Egg measurements Nest # Total depth External diameter Inside diameter of cup Depth of cup Egg 1 Egg 2 Reference 1 45 45 X 45 28 X 32 20 13 X 7 14 X 8 This Study 2 30 44 X 45 27 X 28 17 This study 3 50 42 X 55 25 X 32 22 This study 4 40 51 X 57 30 X 34 22 This study 5 51 43 X 55 24 X 32 21 This study 6 45 44 X 51 25 X 32 20 This study 7 45 38 X 45 25 X 30 22 This study 8 50 45 27 22 14.2 X 9.0 14.2 X 9.2 Rowley (1984) reported from February to August (del Hoyo et al. 1999) and two nests were found in April in Oaxaca, Mexico (Rowley 1984). Nests of A. cyanocephala have been found before, but none has been described and observed in detail; detailed nesting information is unavailable for most species in this genus (del Hoyo et al. 1999). The objective of this paper is to describe the nest, eggs, and incubation of the Azure-crowned Hummingbird (A. c. cyanocephala) in Veracruz, Mexico. OBSERVATIONS The study area consisted of cloud forest and secondary riparian growth bordering the Rio Pixquiac (19° 30' 25" N, 96° 57' 39" W; 1,348 m above sea level) 6 km west of the city of Xalapa, Veracruz, Mexico. Rio Pixquiac is a small tributary of Rfo Consolapa, which traverses Coatepec on the eastern foothills of the Cofre de Perote volcano on the Atlantic side of the Sierra Madre Oriental. 1 observed a hummingbird at 0715 hrs on 20 February 2009 fly directly into a Phvllostachys bamboo bush at a distance of about 15 m. 1 trained binoculars at where the bird appeared to land and observed it sitting on a nest 1.5 m above the ground adding nesting material. 1 watched (he hummingbird for ~ 15-20 min, during which time it visited the nest twice. I was able to confirm the bird was building the nest at 0730 hrs the following morning. 1 watched the female building the nest for —30 min, during which time she made two visits. I returned later in the morning (at 1015 hrs) to obtain photographs of the nest. The female was not .seen during this 15- min visit. 1 checked the nest daily at —0800 hrs. The female made one trip to the nest every 30 min during the course of the early morning observation on 20 February, when the nest was first found, typically remaining for a minute or more. Her building rate early on the following morning was similar and the nest was nearing completion (Fig. lA). However, the female when watched from much closer range, showed signs of nervousness and flew from the nest at each visit. I stopped close-range observations on nest building to avoid cessation of activity and to reduce risk of nest abandonment. The female, at times, spent a short time during nest-building visits sitting motionless in the nest cup in an incubation posture on 25 February. I checked the nest in her absence by noon, and the nest contained one egg. The female amved a few minutes later, perching 5 m from the nest and caiTying material for the nest (lichens). I moved away from the nest after taking photographs and egg measurements, and the female entered the nest and placed nesting material on the outer wall. No signs of incubation occun'ed that day or the following day. A second egg (Fig. IB) was found (egg measurements in Table 1) on 27 February. Eggs were laid in the early morning and there was an interval of 2 days before the second egg was laid. Incubation began after the second egg was laid, after which the nest and eggs were checked twice daily, early in the morning and in the afternoon until the eggs hatched. Building (— 8-- 10 days) and maintenance of the nest, which consisted primarily of attaching lichens to the outside of the nest (there were few lichens at the start of observation), continued throughout the incubation period. The incubation period lasted 15 days and eggs hatched almost synchronously. The first egg SHORT COMMUNICATIONS 595 FIG. 2. Illustration of the construction of Aniazilia cyaiioceplictia nests collected along the Rio Pixquiac near Xalapa Veracruz. Open cup-shaped nest at the two ends of the supporting, horizontal Phyllosuichys bamboo branch kept the top of the nest cup level with respect to the ground (below). Nest cup on the top ol an old Helicoiiici infructescence (above) Drawing by Edmundo Saavedra. 596 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 3. September 2010 hatched (Fig. 1C) by noon on 15 March and the second hatched —1800 hrs that day. Eggshell remains were removed from the nest the following day. The female brooded the young after feeding and during the night on subsequent days. Nest- lings were photographed when they were 8 days of age (Fig. ID). 1 checked the nest early in the morning on 23 March, when the nestlings were 9 days of age, and found it empty. The nest was intact with no signs of disturbance or stealing of nesting material by other hummingbirds (Skutch 1931). 1 believe Social Flycatchers {Myiozetetes similis) attacked the nestlings, as 1 observed several perched near by and had witnessed their attacks on nestlings of Northern Rough-winged Swallow {Stelgidopteryx serripeuuis) in the same area. The nest was collected 2 days later to make a full description and illustration (Fig. 2). The female remained in the area for several days. The cup-shaped nest was built in a 2 m tall Phyllostachys bamboo, isolated by young second growth and Heliconia plants near the Rfo Pixquiac. It was saddled on a lower, slightly downward sloping branch 1.5 m above the ground. The horizontal supporting branch was imbedded within the nest wall and provided additional support on the lower side of the nest. One “tail” of nest material hung beneath the supporting branch (Fig. 2, nest measurements in Table 1). The nest cup was constructed in two layers like most hummingbird nests. The outer layer was comprised of strips of grasses, small dried pieces of leaves and flowers of Mimo.sa spp., scales from tree ferns {Alsophila firnia, Cyathea hicrenotci and Cyathea aff. fidva) (Cyatheaceae), whereas the somewhat thicker inner layer was comprised of Tillandsia deppeana seed down (Bromeliaceae) and many hairs from horses. The outside of the nest was heavily decorated with many pieces of greenish-white lichen, liverworts Calypogeja spp. (Calypogejaceae), and less so with mosses; all were adhered with spider web. 1 discovered the female building a second nest (nest #2; Table 1 ) on 2 April, only 30 m to the west of nest #1, fastened on the outmost, unsheltered branch of a Bougainvillea shrub, 1 .7 m above the ground. The nest was half completed and had no lichens on the outer wall (Fig. 1 E). I continued my observations of this nest by daily checking it at -0800 hrs. The female continued steadily at her growing nest during consecutive days. The first egg was laid on 5 April, but the nest was not completed (Fig. IF). A period of inclement weather was initiated the following day that presumably caused the female to interrupt building and nest attendance, resulting in failure to lay a second egg on 7 April. The single egg lay cold in the nest for an additional 3 days and on 10 April it was destroyed, leaving only eggshell fragments. The nest was collected 2 days later for measurements. The female was observed in the area for several days, but a third nest attempt was not observed. I am confident the female, which built the second nest, was the same individual whose nestlings were depredated. A female (apparently the same) nested in this area for 3 consecutive years; nest #3 was placed on a dried Heliconia infructescence (Fig. 2) and nest #4 was placed on a Bougainvillea shrub. I searched the Rio Pixquiac area on 10 April and located three more A. cyanocephala nests similar to those described. Nests were placed on a branch of a palm {Acrocomia spp.) tree (Fig. IG-H), a Heliconia infructescence, and on a branch of a Quercus tree, all no more than 3 m above the ground. The three nests were each active with two nestlings. Nests were collected and measured (Table 1) after young fledged. This study provides the first descriptions and measurements of the nests of AmazUia cyanoce- phala. Evidence is presented that supports the belief that predation and weather may be the main causes of nesting failure of tropical hummingbirds (Skutch 1931, FieiTO-Calderon and Martin 2007), as well as temperate hummingbirds (e.g., Baltos- ser 1986, 1996). ACKNOWLEDGMENTS I thank Mayra Ledezma for showing me the second nest. Eduardo Ruiz Sanchez for identifying the bamboo. Francisco Lorea Hernandez for identifying the plant material. Edmundo Saavedra for the nest illustration, and Fernando Gonzalez Garci'a, Elisa Peresbarbosa, C. E. Braun, and two anonymous reviewers for manuscript review and suggestions. 1 especially thank Julieta Ornelas, Ana Belen Ornelas, Elisa Peresbarbosa, and Jo.se Luis Cordova for assistance in the field. This study was partly funded by a grant (Ref. 61710) from the Consejo Nacional de Ciencia y Tecnologi'a (CONACyT) and research funds from the Dcpartamento de Biologi'a Evolutiva. Instituto de Ecologi'a, A.C. (Ref. 902-12-563). LITERATURE CITED American Ornithologist.s’ Union (AOU). 1998. Check- list of North American birds. .Seventh Edition. Washington, D.C.. USA. SHOr^T COMMUNICATIONS 597 Atwood. J. L., V. L. Frrz, and J. E. Bamesberger. 1991. Temporal patterns of singing activity at leks of the White-bellied Emerald. Wilson Bulletin 103:373-386. Baltosser, W. H. 1986. Nesting success and productivity of hummingbirds in southwestern New Mexico and southeastern Arizona. Wilson Bulletin 98:353-367. Baltosser, W. H. 1996. Nest attentiveness in humming- birds. Wilson Bulletin 108:228-245. D.avis. L. I. 1972. A field guide to the birds of Mexico and Central America. University of Texas Press, Austin, USA. DEL Hoyo, j., A. Elliott, and J. Sargatal (Editors). 1999. Handbook of the birds of the world. Volume 5. Barn-owls to hummingbirds. Lynx Editions, Barce- lona, Spain. Fierro-Calder6n, K. and T. E. Martin. 2007. Repro- ductive biology of the Violet-chested Hummingbird in Venezuela and comparisons with other tropical and temperate hummingbirds. Condor 109:680-685. Haverschmidt, E. 1952. Notes on the life history of AmaziUa fimbriata in Surinam. Wilson Bulletin 64:69- 79. Howell, S. N. G. and S. Webb. 1995. A guide to the birds of Mexico and northern Central America. Oxford University Press, New York, USA. Johnsgard, P. A. 1997. The hummingbirds of North America. Second Edition, Smithsonian Institution Press, Washington, D.C., USA. Ornelas, J. F. 1995. Radiation in the genus Aniazilicr. a comparative approach to understanding the diversifi- cation of hummingbirds. Dissertation. Univeristy of Arizona, Tucson, USA. Peterson, R. T. and E. L. Chalif. 1973. A field guide to Mexican birds. Houghton Mifflin Co., Boston, Mas- sachusetts, USA. Remsen Jr, j. V., C. D. Cadena, A. Jaramillo, M. Nores, J. F. Pacheco, M. B. Robbins, T. S. Schulenberg, F. G. Stiles, D. F. Stotz, and K. J. Zimmer. 2008. A classification of the bird species of South America. Part 4. Apodiformes. http://www.museum.lsu.edu/ ~Remsen/SACCBaseline04.html Rowley, J. S. 1984. Breeding records of some land birds in Oaxaca, Mexico. Proceedings of the Western Founda- tion of Vertebrate Zoology 2:74-224. Skutch, A. F. 1931. Life history of the Rieffer’s Hummingbird (AmaziUa tzacatl tzacatl) in Panama and Honduras. Auk 48:481-500. Weller, A. A. 1998. Biogeographie, geographische variation und taxonomie der gattung AmaziUa (Aves, Trochilidae). Dissertation. Rheinische Friedrich-Wil- helms-Universitat, Bonn, Germany. The Wilson Journal of Ornithology 122(3):597-599, 2010 Nest Description of the Garden Emerald {Chlorostilbon assimilis) from Costa Rica Luis Sandoval''^ and Ignacio Escalante’ ABSTRACT. — The Garden Emerald (Chlorostilbon assimilis) is endemic to southwestern Costa Rica and Panama, and knowledge about its’ reproductive habits is limited. We describe the nest and nestlings of the Garden Emerald based on a nest found in La Amistad International Park. The nest was built on an anthropo- genic substrate, and was similar to nests described for other emerald species. However, unlike other emeralds, the nest contained no lichens, rnos,ses, or ferns. The nestlings resembled adult female plumage, similar to that for other nestling emerald species. Received 24 August 2009. Accepted 3 February' 2010. The neotropical genus Chlorostilbon includes 17 species of small hummingbirds (AOU 1998, ' Escuela de Biologi'a, Universidad de Costa Rica. 2060 San Pedro de Montes de Oca, San Jose, Costa Rica. ^Corresponding author; e-mail: biosandoval(®hotmaiI. com Remsen et al. 2009), 1 1 of which have described nests (Wolf 1964, Hilty and Brown 1986, Stiles and Skutch 1989, Thomas 1994, Howell and Webb 1995, Stiles 1996, Oniki and Antunes 1998, Schuchmann 1999, Gamdo and Kirkcon- nell 2000). The Garden Emerald (Chlorostilbon a.ssimilis) is endemic to southwestern Costa Rica and Panama to western Darien, including pacific islands, and rarely from Bocas del Toro along the Caribbean coast of Panama (Ridgely and Gwynne 1989, Stiles and Skutch 1989). This hummingbird inhabits secondary forest and open areas with isolated shrubs at elevations up to 1,500 m (AOU 1998, Garrigues and Dean 2007). The nest of the Garden Emerald is undescribed (Stiles and Skutch 1989, Schuchmann 1999) and we provide the first nest description for this species. We also include a brief description of the nestlings. 598 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. 3. September 2010 FIG. 1 . Nest of Garden Emerald found on 28 March 2009 near the Altamira Ranger Station. La Amistad International Park. Costa Rica. (Collected on 18 April 2009; photograph by Luis Sandoval). OBSERVATIONS We found a nest of Garden Emerald on 28 March 2009 at the Altamira Ranger Station in La Amistad International Park, Buenos Aires, Pun- tarenas, Costa Rica (09 02.9' N, 83° 00.8' W; 1,375 m asl). The site is an open area with a park ranger station, surrounded by a garden, which included a live fence of Stachytarpheta spp. { Verbenaceae) bushes, Heliconia spp. patches, and young secondary forest. The nest was collected on 18 April 2009 after the chicks fledged, and was deposited in the Museo de Zoologia. Universidad de Costa Rica. The nest was a cup built on a hanging portion of steel mesh under the elevated floor of the ranger station, I m above ground. The egg cup was 28.0 X 23.5 mm in diameter, and 15.8 ± 2.9 mm (mean ± SD) deep. The outside of the nest measured 37.1 X 40.7 mm in diameter and 19.7 ± 3.5 mm tall. The nest was comprised of two layers. The outer layer included tree bark, small pieces of dry leaves, twigs, spider web, and downy seeds (mostly Asteraceae). The inner layer was comprised mostly of dry fine plant fibers and downy plant material (Pig. 1). The nest contained two completely feathered nestlings when found. The bills of the nestlings were black with reddish in the mandible base. The throat and breast of the nestlings were gray, and the backs were green with buffy-tipped feathers. DISCUSSION The nest structure, including the . two-layer construction, is similar to nests described for other species of emeralds (Schuchmann 1999). However, the nest of Garden Emerald, unlike other emerald nests, contained no lichens, mosses, or fern materials (Oniki and Antunes 1998, Schuchmann 1999, Garrido and Kirkconnell 2000). The feather color pattern of the Garden Emerald nestlings was similar to that described for others emerald species where the nestlings SHORT COMMUNICATIONS 599 resemble the adult female plumage (Wolf 1964, Stiles and Skutch 1989, Schuchmann 1999). The female did not attend the chicks during the night of 28 March, likely because the chicks filled the entire nest. We observed the female attending the nest and feeding both fledglings the following morning. The nest was built on an anthropogenic substrate, but we postulate that, in natural conditions. Garden Emeralds would build nests in the lower branches of bushes along trails, forest edge, or in forest gaps. Our observations reveal that Garden Emeralds can take advantage of human structures and successfully use them for nesting similar to wrens, swallows, or flycatchers (Stiles and Skutch 1989, Baicich and Harrison 2005, Sandoval and Barrantes 2009). ACKNOWLEDGMENTS We thank H. F. Greeney, J. B. Hams, C. E. Braun, and an anonymous referee for useful comments on an early version of the manuscript, and the Altamira Ranger Station administra- tive personnel in La Amistad International Park for support during our field work. We especially thank Sebastian Bonilla for bringing us the nest after the chicks fledged. LITERATURE CITED A-MERican Ornithologists’ Union (AOU). 1998. Check- list of North American birds. Seventh Edition. Amer- ican Ornithologists’ Union, Wa.shington, D.C., USA. Baicich, P. J. and C. O. Harrison. 2005. Nest, eggs, and nestlings of North American birds. Princeton Univer- sity Press, Princeton, New Jersey, USA. Garrido, O. and a. Kirkconnell. 2000. Field guide to the birds of Cuba. Cornell University Press, Ithaca, New York, USA. Garrigues, R. and R. Dean. 2007. The birds of Costa Rica; a field guide. Zona Tropical Publication, San Jose, Costa Rica. Hii.ty, S. and W. Brown. 1986. A guide to the birds of Colombia. Princeton University Press, Princeton, New Jersey, USA. Howell, S. and S. Webb. 1995. A guide to the birds of Mexico and northern Central America. Oxford Uni- versity Press, New York. USA. Oniki, Y. and a. Z. Antunes. 1998. On two nests of the Clittering-bellied Emerald Chlorostilbon aureoventris (Trochilidae). Ornitologfa Neotropical 9:71-76. Remsen Jr., j. V., C. D. Cadena, A. Jaramillo, M. Nores, j. E. Pacheco, M. B. Robbins, T. S. SCHULENBERG, F. C. StILES, D. F. STOTZ, AND K. J. Zimmer. 2009. A classification of the bird species of South America. American Ornithologists’ Union. www.museum.lsu.edu/~Remsen/SACCBaseline.html. Ridgely, R. S. and j. a. Gwynne Jr. 1989. A guide to the birds of Panama with Costa Rica, Nicaragua, and Honduras. Princeton University Press, Princeton, New Jersey, USA. Sandoval, L and G. Barrantes. 2009. Nest usurping occurrence of Piratic Flycatcher (Legatus leucophaius) in southwestern Costa Rica. Ornitologfa Neotropical 20:401-407. Schuchmann, K. L. 1999. Family Trochilidae (humming- birds). Pages 468-680 in Handbook of the birds of the world. Volume 5. Barn-owls to hummingbirds (J. del Hoyo, A. Elliott, and J. Sargatal, Editors). Lynx Edicions, Barcelona, Spain. Stiles, F. G. 1996. A new species of Emerald hummingbird (Trochilidae, Chlorostilbon) from the Sierra de Chiribiquete, southeastern Colombia, with a review of the C. mellisiigus complex. Wilson Bulletin 108:1- 27. Stiles, F. G. and A. F. Skutch. 1989. A guide to the birds of Costa Rica. Cornell University Press, Ithaca. New York, USA. Thomas, B. T. 1994. Blue-tailed Emerald hummingbird (Chlorostilbon mellisiigus) nesting and nestling devel- opment. Ornitologia Neotropical 5:57-60. Wolf, L. L. 1964. Nesting of the Fork-tailed Emerald in Oaxaca. Mexico. Condor 66:51-55. 600 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. 3. September 2010 The Wilson Journal of Ornithology 122(3):600-603, 2010 The Nest of the Cipo Canastero {Asthenes luizcie), an Endemic Furnariid from the Espinhago Range, Southeastern Brazil Henrique Belfort Gomes’’^ and Marcos Rodrigues" ABSTRACT. — The Cipo Canastero [Asthenes liiizae) is a recently described species from the Espinha^o Range, southeastern Brazil. We describe the nest, eggs, and nestlings of this species. Six nests were found in four different territories, two of which were active. All nests were in Vellozia niveo (Velloziaceae). Mean measurements were: height above ground = 21.8 cm, entrance diameter = 4.3 cm, nest length = 23.5 cm, nest height = 23.3 cm, and nest width = 20.3 cm. Both active nests were parasitized by Shiny Cowbirds (Molothriis bonariensis). Received 30 July 2008. Accepted 5 March 2010. The Cipo Canastero [Asthenes luizae Vielliard 1990; Furnariidae) is one of the four species of endemic birds of the Espinha90 Range (Vascon- celos 2008), an endemic bird area in South America (Stattersfield et al. 1998). Due to its restricted range and occurrence in a habitat subjected to significant anthropogenic (fire and cattle) and biological pressures (brood parasit- ism), this species has been classified globally as ‘vulnerable’ (Vielliard 1990, Collar et al. 1992, BirdLife International 2000). The recent discov- ery of the Cipo Canastero, an isolated Asthenes in southeastern Brazil, had great biogeographical significance, suggesting a historical relationship between the mountains of southeastern Brazil and the Andean-Chacoan-Pampean-Patagonian re- gions (Remsen 2003). Anedoctal data regarding six nests of the Cipo Canastero were briefly di.scussed by Studer and Teixeira (1993). They reported one active nest with two eggs and five empty nests. The species is considered high priority for study and conserva- tion due to its rarity, lack of biological data, and biogeographical importance (Stotz et al. 1996). Our objective is to provide detailed descriptions ' Rua Patagonia. 1155. Apartamento 101, 30320-080, Belo Horizonte, MG, Brazil. ’ Laboratorio de Ornitologia, Dcpartamento de Zoologia, ICB. Universidade Federal de Mina.s Gerai.s, Caixa Po.stal 486. 31270-901, Belo Horizonte, MG. Brazil. ’Corresponding author; e-mail: hbeltort@gniail.coni of the nest, eggs, and nestlings of the Cipo Canastero. METHODS Fieldwork was conducted at Alto da Boa Vista (19° 17' S, 43° 35' W), an area of rocky fields at an elevation of 1,230 to 1,355 m above sea level in the Serra do Cipo Range, District of Santana do Riacho, State of Minas Gerais, southeastern Brazil. The vegetation of the rocky fields is marked by sparse treelets and scrubs often with thick bark. Species of Velloziaceae, Eriocaula- ceae, Compositae, Rubiaceae, and Orchidaceae are typical of rocky substrates in the area (Giulietti et al. 1997). This vegetation type is characterized by high floristic diversity and an elevated number of endemic and endangered plant species (Giulietti et al. 1997). We randomly searched for nests in September- November 2005 within an area of ~ 100 ha where at least five pairs of adults had been daily observed vocalizing and foraging. The nests found were periodically monitored and their status recorded. We used calipers (0.05 mm precision) to measure the following parameters for each nest: length, height, and width of the nest, height above ground and entrance diameter (Fig. 1). Eggs were weighed with a 10 g spring balance (0.5 g scale), measured with calipers, and described regarding their format and color. We report mean ±SD for all varibles. RESULTS Six nests were found in four different teirito- ries, two of which were active and four were empty. Empty nests were found in territories with active nests. All nests were in ‘canela-de-ema! shrubs [Vellozio niveci) and had a spherical or cylindrical shape (Fig. 1). The external part of the nests was built of sticks without thorns (15- 55 cm in length), dry material of Barhacenio spp. (Velloziaceae) and Vellozia spp., and moss. The incubation chambers were covered by pilose material of a Cactaceae, feathers, dry fibers of SHORT COMMUNICATIONS 601 FIG. Measurement points for nests of Cipo Cana.stero: (I) height above ground, (2) entrance diameter. r w. V* / cilllunCC U Iclll IC' ICl , (3) PCSt length, (4) nest height, and (5) nest width. Two white eggs of the Cipo Canastero and one of Shiny Cowhird are depicted below the nest. Velloziaceae, and pilose material of a gall that develops in a shrub of Erythroxylum spp. Nests were 21.8 ± 8.2 cm (/? = 6) above ground and had an entrance diameter of 4.3 ± 0.5 cm (n = 4). Nest measurements were: length = 23.5 ± 4.3 by 20.3 ± 4.2 cm in width and a height of 23.3 ± 6.7 cm (// = 6). Both active nests were parasitized by Shiny Cowbirds (Molothms honariensis). One active nest had three eggs, one of Cipo Canastero and two of the cowbird, and the other had two Cipo Canastero eggs and one cowbird egg. The eggs of Cipo 602 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 3, September 2010 Canastero were oval-shaped and completely white, while those of the cowbird varied in color from pink with brown stains, to blue with brown spots (Fig. 1). Cipo Canastero’ s eggs measured 21.8 ± 0.2 X 17.8 ± 0.1 mm and had a mass of 3 ± 0.5 g (n = 3). One active nest was discovered on 21 Septem- ber and was under construction. This nest had three eggs (Cipo Canastero = 1, Shiny Cowbird = 2) on 5 October. The Shiny Cowbirds eggs hatched between 10 and 12 days later. A nestling cowbird was found alive and the other was dead on 18 October, and the Cipo Canastero egg continued in the nest. This egg was found broken on 28 October, without development. The nestling cowbird was no longer in the nest on 2 November. Both adults were marked with colored rings at this nest. Only one individual Cipo Canastero was observed on the nest during incubation, and the other was next to the nest. Both adults were observed delivering food for the nestling cowbird. The other active nest was found on 26 October and had three eggs (Cipo Canastero = 2, Shiny Cowbird = 1). The nest contained one cowbird nestling on 1 1 November and two nestlings of Cipo Canastero on 17 November. One nestling weighed 2.5 g and was 3.7 cm in length, and the other 2 g and 3.9 cm in length. The Cipo Canastero’s nestlings had orange skin and yel- lowish beak and palate, and were covered with gray down. Both Cipo Canastero’s nestlings had their eyes shut. The nestling Shiny Cowbird weighed 16 g on 17 November. Only the nestling Shiny Cowbird was in the nest on 20 November. The nestling Shiny Cowbird was not in the nest on 29 November and it was .seen next to two adults about 1 50 m from the nest on 1 3 December. DISCUSSION The nests and eggs of Cipo Canastero we de.scribe are similar to those briefly discussed by Studer and Teixeira (1993). These authors also found six nests of this species in Vellozici nivea. These findings suggest Cipo Canastero may specialize in using this plant species as a substrate for nesting. However, V. nivea does not occur in the entire range of the Cipo Canastero including the northern mountains in Minas Gerais, north of the Jequitinhonha River Valley (Vasconcelos 2008; R. Mello-Silva, pers. comm.). A possible explanation for occurrence of Cipo Canastero in areas where this plant is absent is that several other species of Vellozia are common and abundant in the rocky fields throughout the Espinhago Range (Giulietti et al. 1997). Several other shrubs of different families, but with architecture similar to V. nivea, also occur in the rocky fields and it is possible that Cipo Canasteros also build their nests in them. Zyskowski and Prum (1999) presented a phylogenetic analysis of nest architecture in Furnariidae. These authors had incomplete data about the nest of Cipo Canastero and they grouped this species in the same operational taxonomic unit as Canyon Canastero (A. pudibunda). Cactus Canastero (A. cactorum). Dusky-tailed Canastero (A. humicola), Rusty-vented Canastero (A. dor- bignyi), Berlepsch’s Canastero (A. berlepschi), Steinbach’s Canastero (A. steinbachi), Short- billed Canastero (A. baeri), and Patagonian Canastero (A. patagonica). Nests of this group are characterized by their domed vegetative structure, built of sticks, and the presence of an entrance tunnel. The detailed nest description presented here supports this arrangement. How- ever, it is possible that Asthenes is polyphyletic (Zyskowski and Prum 1999, Remsen 2003, Irestedt et al. 2006) and new analyses including the Cipo Canastero are needed to better under- stand the systematic position and the biogeo- graphical affinities of this species (Vasconcelos et al. 2008). Adult Shiny Cowbirds were observed around nests of Cipo Canastero at distances varying from 3 to 15 m. The Cipo Canastero seemed to ignore these species, but we observed a pair of Cipo Canastero chasing two individual Rufous- collared Sparrows {Zonotrichia capensis) that approached their nest. The lack of reaction of Cipo Canastero to presence of Shiny Cowbirds may suggest they do not recognize this species as a nest parasite. The number of nests we found is low, but the presence of Shiny Cowbird eggs in the nests demonstrates the Cipo Canastero is a host. Shiny Cowbird eggs had a shorter hatching period, from 10 to 12 days for the first nest and 15-16 days in the second nest. This .shorter hatching period allows Shiny Cowbird nestlings to develop sooner, providing them with higher chances of success. Early hatching in the first nest enabled the nestling of the parasite to make a hole in the egg of the host and, in the second, allowed the nestling cowbird (6.5 times larger in mass) to probably eject the Cipo Canastero nestlings from the nest or caused death by starvation. SHORT COMMUNICATIONS 603 Knowledge of the nesting substratum and reproductive period of species is fundamental for establishment of conservation programs. The Cipo Canastero remains common in Serra do Cipo, despite environmental impacts occurring during recent years within its area of occurrence (Bird- Life International 2000). Nest records, however, are rare and more detailed studies on reproductive seasonality and factors related to the Cipo Canastero’ s success are needed to optimize the conservation of this threatened and restricted- range species. ACKNOWLEDGMENTS We thank Geraldo Fernandes, Tadeu Guerra, Adilson and Maria Gomes. Marcelo Vasconcelos, and Denise Sarfar for support in the field. Virginia Braga drew the nest depicted (Fig. 1). Fabio Olmos, Flavio Rodrigues, Pedro Teixeira, and two anonymous reviewers helped improve the manu- script. Renato Mello-Silva provided important information on the distribution of Vellozici nivea. This work was supported by ‘Funda^ao O Boticario de Prote^ao a Natureza’ and Idea Wild. MR was supported by the Brazilian Research Council (CNPq) and HBG received a student fellowship from the same organizations. We thank the staff of the Brazilian Environmental Agency (IBAMA and CEMAVE), especially Katia Ribeiro and Joao Nascimento. LITERATURE CITED BirdLife International. 2000. Threatened birds of the world. Lynx Edicions, Barcelona, Spain. Collar, N. J., L. P. Gonzaga, N. Krabbe, A. Madrono 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 Edition. Smithsonian Institution Press, Washington, D.C., USA. Giulietti, A. M., J. R. PiRANi, AND R. M. Harley. 1997. Espinha9o Range region, eastern Brazil. Pages 397- 404 in Centres of plant diversity: a guide and strategy for their conservation. Volume 3. (S. D, Davis, V. H. Heywood, O. Herrera-MacBride, J. Villa-Lobos, and A. C. Hamilton, Editors). Information Press, Oxford, United Kingdom. Irestedt, M., j. EjeldsA, and P. G. P. Eric.son. 2006. Evolution of the ovenbird-woodcreeper assemblage (Aves: Furnariidae) - major shifts in ne.st architecture and adaptive radiation. Journal of Avian Biology 37:260-272. Remsen, j. V. 2003. Family Furnariidae (ovenbirds). Pages 162-357 in Handbook of the birds of the world. Volume 8. Broadbills to tapaculos (J. del Hoyo, A, Elliott, and D. A. Christie, Editors). Lynx Edicions, Barcelona, Spain. Stattersfield, A. J., M. J. Crosby, A. J. Long, and D. C. Wege. 1998. Endemic bird areas of the world: priorities for biodiversity conservation. BirdLife International, Cambridge, United Kingdom. Stotz, D. F., j. W. Fitzpatrick, T. A. Parker HI. and D. K. Moskovits. 1996. Neotropical birds: ecology and conservation. University of Chicago Press, Chicago. Illinois, USA. Studer, A. AND D. M, Teixeira. 1993. Notas sobre a biologia reprodutiva de Asthenes luizae Vielliard. 1990 (Aves, Furnariidae). Page 44 in Resumos do III Congresso Brasileiro de Omitologia (M. P. Cime, Editor). Editora da Universidade Catolica de Pelotas. Pelotas, Brazil. Vasconcelos, M. F. 2008. Mountaintop endemism in eastern Brazil: why some bird species from campos rupestres of the Espinha^o Range are not endemic to the Cerrado region? Revista Brasileira de Omitologia 16:348-362. Vasconcelos, M. F., S. D’Angelo Neto, and J. FjeldsA. 2008. Redescription of Cipo Canastero Asthenes luizae, with notes on its systematic relationships. Bulletin of the British Ornithologists’ Club 128:179- 186. Vielliard, J. 1990. Uma nova especie de Asthenes da serra do Cipo, Minas Gerais, Brasil. Ararajuba 1:121-122. Zyskowski. K. and R. Prum. 1999. Phylogenetic analysis of the nest architecture of neotropical ovenbirds (Furnariidae). Auk 116:891-911. 604 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 3, September 2010 The Wilson Journal of Ornithology 122(3):604-608, 2010 Nest Box Use by Great Tits in Semi-arid Rural Residential Gardens Motti Charter, ‘ Yossi Leshem,' Shay Halevi," and Ido Izhaki-^ ABSTRACT. — We studied use of nest boxes by Great Tits [Parus major) in rural village gardens in a semi-arid area. Great Tits occupied 46.6% of the nest boxes, and used nest boxes within higher tree densities and with more tree species in the vicinity. Breeding success was greater in nest boxes with higher plant density, more plant species, and greater height of trees in the vicinity of the nest. The presence of children or dogs (Canis lupus familiaris) near nest boxes did not affect breeding parameters. Syrian Woodpeckers (Den- drocopos syriacus) enlarged 38.0% of nest box entrances during the first year. House Sparrows (Passer domesticus) occupied 41.0% of the nest boxes with enlarged holes and none of those with normal holes. Great Tits occupied both types, but significantly fewer pairs breeding in nest boxes with enlarged holes succeeded in fledging at least one young, probably due to their eviction by the larger House Sparrows. Received 19 October 2009. Accepted 19 January 2010. Great Tits (Pants major) are one of the most studied passerines due to their wide distribution and use of nest boxes. Most studies have been in northern Europe (Sanz 1998) and, to a much lesser extent, in Mediterranean Europe (Beldal et al. 1998, Lambrechts et al. 2008), but rarely in the Middle East (Yavin 1987). More studies have been of Great Tits breeding in deciduous (van Balen 1973, Riddington and Gosler 1995) and coniferous forests (Orell and Ojanen 1983, Mand et al. 2005), than with those in suburban (Dhondt et al. 1984, Cowie and Hinsley 1987) and rural gardens (Perrins 1965, Riddington and Gosler 1995). Residential gardens around homes vary greatly in number of plant and tree species; studies in the.se locations facilitate a deeper understanding of the environmental factors birds use for nest site selection. Great Tits in Mediterranean regions are the smallest species of cavity breeders and may lose ' Zoological Department. Tel-Aviv University, Ramat- Aviv. Tel-Aviv 69978. Israel. ’ Moshav Ram-On 19205. Israel. department of Evolutionary and Environmental Biolo- gy. University of Haifa, Haifa 31905, Israel. ■•Corresponding author; e-mail: charterm@post.tau.ac.il nest sites to larger species, mainly House Sparrows (Passer domesticus) (Barba and Gil- Delgado 1990), unlike in Europe where Great Tits compete mainly with smaller species for nest sites and are the dominant species (Dhondt and Eykerman 1980, Minot 1981, Alatalo et al. 1986, Minot and Perrins 1986, Alatalo et al. 1987, Kempenaers and Dhondt 1991). We studied nest box use by Great Tits in rural village gardens in a Mediterranean climate and examined environmental factors, in particular, the presence of children, dogs (Canis lupus famil- iaris), plants, trees, grass, and size of entrance holes of nest boxes, that may be related to occupation and breeding success of Great Tits. METHODS The study site was in Moshav Ram-On in the semi-arid Jezreel Valley, Israel (32° 31' 55" N, 35° 15' 25" E), 80-90 m above sea level with an average annual precipitation of 453 mm. The 50- ha study site comprised 170 houses, each with a garden of 500 m^ and other small buildings used for storage. These were surrounded by agricultural areas, mainly fields of sweet corn, alfalfa, oats, wheat, grape vines, almond plantations, and olive groves. Eifty-eight wood nest boxes (15 cm wide X 15 cm long X 24 cm high) with an entrance of 30 mm diameter, at a height of 1 .5 to 2.0 m above the ground, were affixed during August 2006 to trees randomly with no more than one nest box per garden. Active nests (defined as a nest box in which Great Tits or Hou.se Sparrows built nests) were visited each week from 1 January 2007 to 1 June 2007. Laying date, clutch size, mean number of young per nest box after hatching, mean number of fledglings per laying (all pairs that laid at least 1 egg were included), and successful pairs (pairs that fledged at least 1 young) were recorded for each pair of Great Tits. The following data were recorded for each nest box within a 30-m radius; presence of dogs (yes/ no), presence of children (yes/no), plant density (rated I lowest to 5 highest, not including trees). SHORT COMMUNICATIONS 605 TABLE 1. Characteristics (x ± SE) of areas where Great Tits built nests in nest boxes in senii-arid rural gardens, 2007. Built ne.st Yes No Mann-Wtiilney U Plant density 3.04 ± 0.2 (/; = 23) 2.74 ± 0.2 (n = 23) U = 230.0, n = 46, P = 0.43 Tree density 3.26 ± 0.2 (n = 23) 2.43 ± 0.2 (/; = 23) U = 164.5, n = 46, P < 0.05 Number of grass species 0.94 ± 0. 1 (/( = 18) 1 .00 ± 0.0 in = 19) U = 161.5, n = 37, P = 0.30 Number of tree species 3.96 ± 0.5 (/; = 23) 2.26 ± 0.3 (;; = 23) U = 152.5, n = 46, P < 0.05 Number of plant species 6. 1 7 ± 1.4 (/; = 23) 4.55 ± 0.6 (/I = 22) U = 228.0, n = 45, P = 0.57 Height ot trees 3.65 ± 0.4 (n = 23) 2.96 ± 0.3 (n = 23) U = 184.5, 11 = 46, P = 0.07 Distance from homes 3.45 ± 0.3 (n = 22) 3.52 ± 0.3 (/! = 23) U = 284.5, 11 = 46, P = 0.91 Distance closest neighboring nest box 4.04 ± 0.3 (u = 23) 3.57 ± 0.3 (u = 23) U = 212.5, n = 46, P = 0.23 Distance to street 3.50 ± 0.4 (n = 22) 2.61 ± 0.4 (n = 23) U = 186.5, n = 45, P = 0.1 1 tree density (1 to 5), number of grass species, number of tree species, number of plant species, height of trees (1 to 5), distance from the house (1 to 5), distance to closest neighboring nest box (1 to 5), distance to street (1 to 5), and whether the entrance hole had been enlarged by Syrian Woodpeckers (Dendrocopos syriacus) (yes/no). Plant density, tree density, height of trees, distance from the house, distance to closest neighboring nest box, and distance to street were also rated. All statistical tests were two-tailed and all tests were non-parametric. Descriptive breeding data were analyzed using the Mann-Whitney t/-test. Spearman’s test was used to analyze rank coirela- tions among breeding parameters. Chi-square and Fisher’s exact tests were used for comparing rate of occupation and nest success, between nests of different entrance sizes, and nest boxes within the vicinity of dogs and/or children. Levels of significance were set at P < 0.05. Statistical analyses were performed using Statistica (Version 7.1) software and SPSS (Version 17). RESULTS Great Tits built nests in 46.6% of the nest boxes {n = 58), and laid eggs in 85.2% (n = 27) of these. They succeeded in Hedging at least one young in 74.1% of the nest boxes in which they laid eggs (n — 20). The mean laying date was 20 March 2007 (n = 23, SE 0.2), mean (± SE) clutch size was 7.13 ± 0.4 (/? = 15), mean number of Hedglings per laying pair was 4.9 ± 0.6 [n = 23), and mean number of fledglings per successful pair was 5.7 ± 0.4 (/? = 20). The breeding density was 5.4 pairs/ 10 ha. Great Tits built more nests in nest boxes that were in an area with higher tree density and a higher number of tree species, but no differences were found between nest box location and other habitat attributes (Table 1). Greater breeding success was observed (fledged at least 1 nestling) in nest boxes in the vicinity of higher plant density, more plant species, and greater height of trees (Table 2). Clutch size was not correlated with nest box location or any habitat attribute; number of Great Tit fledglings was negatively coiTelated with tree height in the vicinity of nests (Table 3). The presence of dogs did not affect clutch size (Mann- Whitney U = 4.0, n = S, P = 0.38), number of fledglings (Mann-Whitney U = 22.0, n = 15, P = 0.55), or rate of occupation of nest boxes ix- = 0.04, n = 30, P = 0.92). The presence of children also did not affect Great Tit clutch size (Mann-Whitney U = 16.0, n = 14, P = 0.57), number of fledglings (Mann-Whitney U = 28.0, /? = 1 8, P = 0.33) or rate of occupation of nest boxes {x~ = 0.46, n = 44, P = 0.46). Syrian Woodpeckers enlarged 38.0% of nest box (/? = 58) entrances during the first year. Great Tits equally occupied nest boxes with both enlarged and normal hole entrances (/’ = 0.16, n = 58, P = 0.69), but significantly fewer pairs breeding in nest boxes with enlarged holes succeeded in fledging at least one young (Fig. 1). House Sparrows occupied 41.0% of the nest boxes with enlarged entrance holes and none with normal holes (Fig. 1). DISCUSSION Differences in occupation and breeding success varied with local habitat attributes around the nest box even though 46.6% of the nest boxes overall were occupied by Great Tits. Tree number/ density/height and plant number/density had a 606 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122. No. 3. September 2010 TABLE 2. Characteristics (x ± SE) of areas that affected nest success of Great Tits fledging at least 1 young per nest in nest boxes in semi-arid rural gardens, 2007. Succeeded to fledge young Yes No Mann-Whiiney U Plant density 3.35 ± 0.3 (n = 17) 2.17 ± 0.3 (72 = 6) U = 23.0,' 72 = 12, P < 0.05 Tree density 3.47 ± 0.3 (72 = 17) 2.67 ± 0.4 (72 = 6) U = 32.5, 72 = 23, P = 0.18 Number of grass species 0.93 ± 0.1 (72 = 14) 1.00 ± 0.0 (72 = 6) U = 26.0, 72 = 18, P = 0.60 Number of tree species 4.47 ± 0.7 (72 = 17) 2.50 ± 0.4 (72 = 6) U = 26.5, 72 = 22, P = 0.08 Number of plant species 7.53 ± 1.7 (72 = 17) 2.33 ± 0.6 (72 = 6) U = 10.0, 72 = 22, P < 0.01 Height of trees 3.18 ± 0.2 (72 = 17) 5.00 ± 0.0 (72 = 6) U = 3.0, 72 = 23, P < 0.001 Distance from homes 3.25 ± 0.4 (72 = 16) 4.00 ± 0.6 (72 = 6) U = 36.0, 72 = 22, P = 0.34 Distance closest neighboring nest box 4.12 ± 0.3 (72 = 17) 3.83 ± 0.5 (72 = 6) U = 44.5, 72 = 23, P = 0.62 Distance to street 3.69 ± 0.4 (72 = 16) 3.00 ± 0.9 (72 = 6) U = 40.5, 72 = 22, P = 0.56 role in nest selection and success of breeding pairs. Trees are important for Great Tits, as their reproductive parameters increase in accordance with size of the woodland (Hinsley et al. 1999). Nestling diet in woodlands contained more caterpillars (the preferred prey of Great Tits) than in marginal habitats with fewer and younger trees (Riddington and Gosler 1995). Children, dogs, and proximity to streets did not appear to affect Great Tit breeding. In contrast. Park et al. (2005) found a difference in breeding near streets. Diet was not monitored in the present study but may be important, as both tree and plant densities may influence prey abundance. Gardens with a higher abundance of plants and trees are normally irrigated more, providing green leaves for cater- pillars. Breeding densities and success were within the range of other studies in gardens (Dhontdt et al. 1984, Cowie and Hinsley 1987), but lower than for pairs in deciduous forests (Cowie and Hinsley 1987). Clutch size was smaller than in most studies in Europe, and fits within the quadratic relation suggested by Sanz (1998). Syrian Woodpeckers were the primary cavity builders in the study site and, in some cases, enlarged natural cavities that were too small for them but which were used by Great Tits. Consequently, the enlarged entrances become available for larger species including House Sparrows. Great Tit and House SpaiTow diets do not overlap, but both use cavities and may compete for them. House SpaiTows in our study occupied 41.0% of the nest boxes with enlarged holes and none with normal holes. Great Tits occupied nest boxes with both enlarged and normal hole entrances similarly, but significantly fewer pairs breeding in nest boxes with enlarged holes succeeded in fledgling at least one young, most likely due to their eviction by larger House Sparrows. Barba and Gil-Delgado (1990) reported Great Tits occupied a diminishing number of nest boxes over a period of 3 years and suggested this was TABLE 3. The relation between habitat attributes, clutch size, and number of fledglings in semi-arid rural gardens, 2007. Clutch size" Number of fledglings" r = -0.12, n = 12, P = 0.71 r = -0.22, /; = 12, P = 0.30 r = -0.51, /; = 8, P = 0.19 r = -0.14, n = 12, P = 0.66 r = 0.29, // = 1 2, P = 0.37 r = 0.10, /I = 20, P = 0.67 Plant density Tree density Number of grass species Number of tree species Number of plant species Height of trees Distance from homes Distance close.st neighboring nc.st box Distance to street r = -0.31, n = 12, P = 0.32 r = 0.28, n = 12, P = 0.38 r = -0.02, n = 12, P = 0.96 r = -0.57, n = 12, P = 0.054 r = 0.07, n = 20, P = 0.76 r = -0.35, // = 15, P = 0.21 r = 0.10, n = 20. P = 0.69 r = 0.23, II = 20. P = 0.34 r = -0.51, n = 20. P < 0.05 r = 0.14, n = 19, P = 0.56 /• = 0.19, n = 20. P = 0.41 r = 0.01, n = 19, P = 0.97 ■* Spearman’s lest. SHORT COMMUNICATIONS 607 Normal entrances (r? = 36) Enlarged entrances (n = 22) Nest entrance size ■ Percent of Great Tit pairs « Nest boxes occupied by House Sparrovvs FIG. 1. Percent of Great Tit pairs that succeeded in fledging young = 7.92, n = 21, P < 0.01) and percent nest boxes occupied by House Sparrows {x~ = 13.91, n = 58, P < 0.001) from nest boxes with normal entrance hole (30 mm) and with entrance holes enlarged by Syrian Woodpeckers. due to an increase in House Sparrows and black rats (Rattus rattus) in the nest boxes. The normal entrance holes in our study prevented House Sparrows from entering and Great Tits were able to breed, whereas in the entrances that were enlarged, sparrows also were able to enter. Both this study and that of Barba and Gil-Delgado (1990) suggest that competition may be the reason for decreased occupation (Barba and Gil-Delgado 1990) and breeding success (this study), but both studies were purely observational. Experimental studies manipulating hole entrance size are needed to better understand whether competition between the two species affects their breeding success. ACKNOWLEDGMENTS We thank the residents of Moshav Ram-On for assistance and Ran Avrami for weather data. We especially thank Hava and Uri Ravid for technical assistance in the field, and Naomi Paz for editorial assistance. LITERATURE CITED Alatalo, R. V., L. Gustafsson, and A. Lundberg. 1986. Interspecific competition and niche shifts in tits Pants: evaluation of non-experimental data. American Natu- ralist 127:879-834. Alatalo, R. V., D. Erickson, L. Gustafs,son, and K. Larson. 1987. Exploitation competition intluences the use of foraging sites by tits: experimental evidence. Ecology 68:284-290. Barba, E. and J. A. Gil-Delgado. 1990. Competition for nest-boxes among four vertebrate species: an experi- mental study in orange groves. Ecography 13:183-186. Beldal. E. J., E. Barba, J. E. Gil-Delgado, D. J. Iglesias, G. M. Lopez, and J. S. Monros. 1998. Laying date and clutch size of Great Tits (Pants major) in the Mediterranean region: a comparison of four habitat types. Journal of Ornithology 139:269-276. CowiE, R. J. AND S. A. Hinsley. 1987. Breeding success of Blue Tits and Great Tits in suburban gardens. Ardea 75:81-90. Dhondt, A. A. AND R. Eyckerman. 1980. Competition and the regulation of numbers in Great and Blue tits. Ardea 68:121-132. Dhondt, A. A., R. Eyckerman, R. Moermans, and J. Huble. 1984. Habitat and laying date in Great and Blue tits. Ibis 126:388-397. Hinsley, S. A., P. Rothery, and P. E. Bellamy. 1999. Influence of woodland area on breeding success in Great Tits Pants major and Blue Tits Parus caeritleus. Journal of Avian Biology 30:271-281. Kempenaers, B. and A. A. Dhondt. 1991. Competition between Blue and Great tits for roosting sites in winter: an aviary experiment. Ornis Scandinavica 22:73-75. Lambrechts, M. M., A. Rielix. M. J. Galan, M. Cartan- SON, P. Perret, and j. Blondel. 2008. Double- brooded Great Tits (Pants major) in Mediterranean oak habitats: do first broods always perform better than second broods? Russian Journal of Ecology 39:516-522. Mand, R.. V. Tilgar, A. Lohmus. and A. Leivits. 2005. Providing nest boxes for hole-nesting birds-Does habitat matter? Biodiversity and Conservation 14:1823-1840. Minot, E. O. 1981. Effects of interspecific competition for food in breeding Blue and Great tits. Journal of Animal Ecology 50:375-385. Minot. E. O. and C. M. Perrins. 1986. Interspecific interference competition-nest sites for Blue and Great tits. Journal of Animal Ecology 55:331-350. Orell, M. and M. Ojanen. 1983. Effect of habitat, date of laying and density on clutch size of the Great Tit 608 THE WILSON JOURNAL OF ORNITHOLOGY • Vul. 122, No. 3. September 2010 Panis major in northern Finland. Holarctic Ecology 6:413-423. Park, Y., W. Lee, and S. Rhim. 2005. Influence of forest road on breeding of tits in artificial nest boxes. Journal of Forestry Research 16:301-302. Perrins, C. M. 1965. Population fluctuations and clutch- size in the Great Tit, Pams major L. Journal of Animal Ecology 34:601-647. Riddington, R. and a. G. Gosler. 1995. Differences in reproductive success and parental qualities between habitats in the Great Tit (Pams major). Ibis 137:371- 378. Sanz, J. J. 1998. Effects of geographic location and habitat on breeding parameters of Great Tits. Auk 1 15:1034- 1051. van Balen, j. H. 1973. A comparative study of the breeding ecology of the Great Tit Pams major in different habitats. Ardea 61:1-93. Yavin, S. 1987. Nest site selection of the Great Tit (Pams major). Thesis. Tel Aviv University, Israel. The Wilson Journal of Ornithology 122(3):608-61 1, 2010 Analysis of Nest Sites of the Resplendent Quetzal (Pharomachrus mocinno): Relationship between Nest and Snag Heights Dennis G. Siegfried,'"^ Daniel S. Linville,' and David Hille^ ABSTRACT. — The Resplendent Quetzal (Pharoma- chrus mocinno) is of particular conservation concern becau.se of its iconic status in Central American culture. This species is a secondary cavity nester and modifies abandoned woodpecker nest sites in dead tree trunks (i.e., snags). We used 1 1 historical nest sites, reported in 1969, from Atitlan, Guatemala and 10 recent nest sites from San Gerardo de Dota. Costa Rica to examine if a relationship exists between nest and snag height. There were significant differences between Costa Rica and Guatemala in both nest height (6.3 vs. 10 m, respec- tively; /-test|4 = -2.49, P = 0.042) and snag height (8.1 vs. 14.0 m, respectively; t-testn = —2.39, P = 0.033). There was no difference in nest heights relative to snag heights for Costa Rica (0.76) and Guatemala (0.77; /-test|7 = —0.20, P = 0.84). One aspect of conservation efforts for this species has been placement of nest boxes to provide nesting sites for additional pairs. Our results provide a better understanding of placement requirements for nest boxes to encourage their use anywhere within the range of the species. Received 3 December 2009. Accepted 5 April 2010. Cavity nesting-.species have challenges from finding the limited re.sottrce of a nest cavity (Cornelius et al. 2008) to predation events on eggs, chicks, or adults (Martin and Li 1992, Brightsmith 2005). Many cavity-nesting species in both temperate and tropical latitudes display ' Department of Biology, .Southern Nazarene University, 6729 Northwest 39"’ Expressway, Bethany, OK, 73008, USA. ^Quetzal Education Research Center, 1913-7050, Car- tago. Costa Rica. ’Corresponding author; e-mail: dsiegfri@snu.edu preferences for cavity height in relation to height of the tree or snag (i.e., relative nest height). Primary cavity-nesting species, including Red- headed Woodpecker (Melanerpes erythrocepha- lus). Northern Flicker (Colaptes ait rants), and Acorn Woodpecker (M. formicivonts), place nests at relative nest heights of 0.49-0.58 (Sedgwick and Knopf 1990, Hooge et al. 1999). Some secondary-cavity nesters in the temperate zone select relative nest heights with more variation: Black-capped Chickadees {Poecile atricapillits) and House Wren {Troglodytes aedon) at 0.37- 0.38, European Starlings {Slit nuts vulgaris) and American Kestrels (Falco sparveriits) at 0.5 (Sedgwick and Knopf 1990), and Eastern Screech-Owls {Megascops asio) at 0.31 (Belthoff and Ritchison 1990). Less information is available in tropical areas about selection of relative nest heights by .secondary-cavity nesters. The Resplendent Quet- zal {Pharomachrus mocinno) is an exclusive secondary-cavity nester (Johnsgard 2000) that, as an iconic species, has become the focus of con.servation efforts at many locations throughout Central America. More efficient planning and implementation has occurred with concomitant con.servation benefits for many other species in focusing on one vulnerable and highly mobile species (Wheelwright 1983, Powell and Bjork 1995). The Resplendent Quetzal throughout its range uses natural cavities created from stubs of fallen branches or abandoned nest cavities of other species, primarily woodpeckers, which SHORT COMMUNICATIONS 609 Resplendent Quetzals occupy and enlarge (Wheelwright 1983). The Acorn Woodpecker is the primary excavator in the Talamanca Mountain Range ot southern Costa Rica. This species excavates several locations for nesting and then abandons these sites as the level of decay increases (Koenig et al. 1995). We examined nest heights of currently active (in Costa Rica) and historical (in Guatemala; Bowes and Allen 1969) Resplendent Quetzal nests and compared them to heights of snags where they occur. We sought to ascertain if there were: ( 1 ) any differences in nest placement between countries, (2) consistent relative nest heights for Resplendent Quetzals, and (3) any implications for future conservation efforts. METHODS Study Site. — The San Gerardo de Dota Valley is on the Pacific Slope of the Talamanca Mountain Range, 85 km southeast of San Jose below the Cerro del Muerte on the Pan-American highway. The Quetzal Education Research Center (QERC; 9° 33' 08.67" N, 83° 48' 23.18" W; 2,200 m asl) is part of a 400-ha preserve consisting of home- steads pioneered in 1952 for dairy farming. Much of the pasture land was converted to fruit-tree orchards starting in 1984 and the last pasture areas were allowed to begin the process of succession to forest in 1996 (Neuenschwander 2001). The primary forest is dominated by oaks {Qiiercus spp.) with an understory dominated by Lauracea, Moracea, and Fabacea. Secondary forest areas have similar composition. The Savegre River bisects the valley and most human activity occurs in the associated riparian area. This area has been planted with eucalyptus (Eucalyptus spp.) trees for erosion control to supplement the remaining native species. Data Collection. — Our initial search for both active and past nests from January through April 2008 involved interviewing local guides and individuals in San Gerardo de Dota, Costa Rica who were familiar with nest locations. We searched along established trails to locate these and other nests. Eight of the 10 sites that we located were active and two were not being used although observations in previous years indicated these were active nest sites. We observed pair activity including exchange of occupancy or delivery of food items at a site to document it was active. We measured nest and snag height using a standard metric tape measure and clinometer (Suunto PM-5). We calculated relative nest height as the height of the nest divided by the height ot the snag. This calculation was also done lor the 1 1 reported sites from Guatemala (Bowes and Allen 1969). We used 21 nest sites for analysis, 10 from Costa Rica and 11 sites in Guatemala reported by Bowes and Allen (1969). All 21 sites were in trunks of dead trees (i.e., snags). Statistical Analysis. — We had unequal number ot sites in each country and a recognized difference in subspecies. Thus, we used two- sample r-tests assuming unequal variance to compare nest and snag heights, and the ratios between countries (Van Emden 2008). We used Pearson product-moment correlation to examine if the nest-to-snag relationship was linear across sites. We used Systat 1 1 (Systat 2004) for all of our calculations. RESULTS There was only one Resplendent Quetzal nest cavity per tree at each Costa Rica site. There were many woodpecker holes in a snag, but the majority, based on shape and depth, were relic foraging holes. There were two additional large holes in the case of tree # 4; one was 1 m above the active nest while the other was 650 cm below it facing a different direction. Tree # 2 had one hole 500 cm above the nest that was used by Spot-crowned Woodcreepers (Lepidoco- laptes ajfini.s). Nesting success was difficult to ascertain and it remains unclear how many nest sites were successful. However, we ob.served a nestling that left nest # 1 prematurely and died, and that nest # 5 was abandoned by the adults prior to fledging. Snag heights in Costa Rica averaged 8.1 m (5 = 2.78, range = 1.9—10.8 m, n = 10) and were lower than those used in Guatemala, where the average was 14.0 m (s = 7.57, range = 5.8- 29.0 m, n = 11; /-test,.-, = -2.39, P = 0.033). The average nest height of snags used in Costa Rica (6.3 m, .v = 2.53, range = 1.4-10.1 m, n = 10) was also lower than in Guatemala (10.8 m, 5 = 6.15, range = 4.0-24.4 m, n = 11; Mesti4 = — 2.49, P = 0.042). The product-moment corre- lation between snag height and nest height was high (/- = 0.976, P < 0.0001; Fig. 1). No difference was found in nest heights relative to snag heights for Guatemala (0.77) and Costa Rica (0.76; Mest|7 = -0.20, P = 0.84). 610 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 3, September 2010 D) 0) (/) (U 25 - 1 1 1 1 o 20 - r= 0.976 o - 15 - o - 10 - OO • o o - 5 - • • • Costa Rica • o Guatemala 0 - 1 1 1 1 ’ 1 ^ 0 5 10 15 20 25 30 Snag height (m) FIG. 1. Relationship of snag versus nest height of Respiendent Quetzais in two popuiations in Costa Rica and Guatemaia (/■ = 0.976, P < 0.000 i). DISCUSSION Successful nesting by Resplendent Quetzals is dependent on numerous factors; from males having sufficient covert lengths and display flights (Unger 1988) to defending and provision- ing chicks. One aspect that is crucial to successful nesting is finding suitable cavities. Resplendent Quetzals in the northern portion of their range excavate entire nests as there are no woodpeckers or their allies that produce a cavity of suitable size (Renner 2005). We observed Resplendent Quet- zals at San Gerardo de Dota making small holes in dead trees prior to mating, but nest cavities were developed from cavities initially excavated by Acorn Woodpeckers. The difference may be related to most nest sites at San Gerardo de Dota being in trunks of oaks, which have relatively hard wood. Finding abandoned Acorn Woodpecker nests appears to be important, but relative nest height may be of primary importance in .selecting a nest site. Average relative nest height was 0.65 (Renner 2005) when no previously excavated cavities were available, primarily due to a lack of sufficiently large woodpecker species. Average relative nest heights were 0.76 and 0.77, respec- tively in our study and the earlier study in Guatemala (Bowes and Allen 1969). Both our Costa Rica site and the Guatemala (Bowes and Allen 1969) site are within the range of a woodpecker species that is sufficiently large to excavate a cavity that may then be enlarged by the Resplendent Quetzal (Skutch 1944, Bowes and Allen 1969, Johnsgard 2000). The primary-cavity nester at San Gerardo de Dota is the Acorn Woodpecker. This species nests roughly half way up in a snag (Hooge et al. 1999). As the snag ages, material is lost from the top reducing the overall relative nest height. It is only after sufficient decay that the site is abandoned by woodpeckers and becomes available for Resplen- dent Quetzals. Nests are often used by Resplen- dent Quetzals for several years until the snag finally collapses. The continued presence of Resplendent Quet- zals is of primary interest to the inhabitants of San Gerardo de Dota as there are more than 12,000 visitors annually to the Quetzal Education Re- search Center. Conservation efforts at numerous locations throughout the range of the Resplendent Quetzal have included addition of artificial nest boxes (Bowes and Allen 1969, LaBastille 1974). Approximately four nest boxes per year, including replacements, have been in place since 1993 at San Gerardo de Dota. These boxes potentially could provide nesting sites for additional pairs, but none has been used. Thus, there may be sufficient numbers of natural sites for the number of nesting pairs. However, it is possible the SHORT COMMUNICATIONS existing boxes are inappropriately positioned. These boxes are mounted at the top of 2.5 m posts, which do not extend above the boxes like the snags above the natural nesting cavities. Further, the absolute height of the nest boxes is considerably below the average heights of natural nesting cavities recorded in our study. Conservation efforts for the Resplendent Quet- zal have implications for many other species in the tropical cloud forests of Central America following succession, which provides new areas tor nesting. Completion of the proposed Meso- American biological corridor in this area could be beneficial for Resplendent Quetzals and other species (Kaiser 2001). ACKNOWLEDGMENTS We thank the many individuals in San Gerardo de Dota who helped locate nest sites for this project. We also thank two anonymous reviewers who made many useful sugges- tions as well as the National Aeronautics and Space Administration (NASA) for providing funding for this research. LITERATURE CITED Belthoff, J. R. and G. Ritchison. 1990. Nest-site selection by Eastern Screech-Owls in central Ken- tucky. Condor 92:982-990. Bowes, A. and D. G. Allen. 1969. Biology and conservation of the quetzal. Biological Conservation 1:297-305. Brightsmith, D. J. 2005. Competition, predation and nest niche shift among tropical cavity nesters: ecological evidence. Journal of Avian Biology 36:74—83. Cornelius, C., K. Cockle, N. Politi, I. Berkunsky, L. Sandoval, V. Ojeda, L. Rivera, M. Hunter Jr., and K. Martin. 2008. Cavity-nesting birds in neotropical forests: cavities as potentially limiting resource. Omitologia Neotropical 19:253-268. Hooge, P. N., M. T. Stanback, and W. D. Koenig. 1999. 61 1 Nest-site selection in the Acorn Woodpecker. Auk 1 1 6:45-54. JOHNSGARD, P. A. 2000. Trogoiis and quetzals of the world. Smithsonian Institution Press, Washington, D.C., USA. Kaiser, J. 2001. Conservation biology: bold corridor project confronts political reality. Science 293:2196- 2199. Koenig, W. D., R, L. Mumme, M. T. Stanback, and F. A. PiTELKA. 1995. Patterns and consequences of egg destruction among joint-nesting Acorn Woodpeckers. Animal Behaviour 50:607-621. LaBastille, A. 1974. Use of artificial nest-boxes by quetzals in Guatemala. Biological Conservation 1 :64- 65. Martin, T. E. and P. Li. 1992. Life history of open- vs. cavity-nesting birds. Ecology 73:579-592. Neuenschwander, D. E. and L. R. Finkenbinder. 2001. The chainsaw and the white oak: from astrobiology to environmental sustainability. Radiations 7:5-10. Powell, G. and R. Bjork. 1995. Implications of intratropical migration on reserve design: a case study using Pharomachrus mocinno. Conservation Biology 9:354-362. Renner, S. C. 2005. The Resplendent Quetzal {Pharoma- chrus mocinno) in the Sierra Yalijux, Alta Verapaz, Guatemala. Journal of Ornithology 146:79-84. Sedgwick, J. A. and F. L. Knopf. 1990. Habitat relationships and nest site characteristics of cavity- nesting in cottonwood floodplains. Journal of Wildlife Management 54:1 12-124. Skutch, A. 1944. Life history of the quetzal. Condor 46:213-235. Systat. 2004. Systat Software, Version 11.00.01. Systat Software Inc., Chicago, Illinois. USA. Unger. D. 1988. Welche function hat der extreme sexuladimorphismus de Resplendent Quetzal {Phar- omachrus mocinno)? Universitiit Tubingen, Tubingen. Germany. Van Emden, FI. F. 2008. Statistics for terrified biologists. Blackwell Publishing Ltd., Malden, Massachusetts. USA. Wheelwright. N. T. 1983. Fruits and ecology of Resplendent Quetzals. Auk 100:286-301. 612 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. 3. September 2010 The Wilson Journal of Ornithology 122(3):612-614, 2010 Parasitism of a Blue-winged Teal Nest by a Northern Shoveler in South Dakota Thomas E. Lewis ' “ and Pamela R. Garrettson' ABSTRACT. — Nest parasitism in dabbling ducks is uncommon. Northern Shovelers (Anas clypeata) are considered infrequent parasites and Blue-winged Teal (A. discors) are rarely hosts to nest parasitism. We documented parasitism of a Blue-winged Teal nest by a Northern Shoveler. We reviewed 3,003 records from nests located in 1994 to 1997 in North Dakota and documented only one duck nest parasitized by a Northern Shoveler. Only nine of 1,494 Blue-winged Teal ne.sts were parasitized. We found no evidence of Northern Shovelers parasitizing Blue-winged Teal nests in our nest records and only one account in the literature. Received 12 January 2010. Accepted 24 March 2010. Northern Shovelers (Anas clypeata) and Blue- winged Teal (A. discors) are common upland nesting ducks in the northcentral United States and prairie Canada (Dubowy 1996, Rohwer et al. 2002). The estimated abundance of Northern Shoveler and Blue-winged Teal in the traditional survey area was well above the long-term average in 2009 (USDl 2009). Nest parasitism occurs when a female purpose- ly lays eggs in a nest to be raised by an individual that is not her mate (Sayler 1992). Nest parasitism is relatively uncommon among dabbling ducks, except for cavity nesting or ducks that crowd together on nesting islands (Sayler 1992). North- ern Shovelers are documented to parasitize nests of other ducks (Glover 1956, Weller 1959, Sayler 1992, Zimmerman et al. 2003). Blue-winged Teal nests may rarely be parasitized by other ducks, but documentation is lacking (Rohwer et al. 2002). We document the second reported case of a Northern Shoveler parasitizing a Blue-winged Teal nest. OBSERVATIONS We Hushed a Northern Shoveler hen from a nest on the morning of 1 2 May 2009. The nest was ' U.S. Fish and Wildlife Service, Division of Migratory Bird Management, I 1510 American Holly Drive, Laurel. MD 20708. USA. ^Corresponding author; thom_lewis@fws.gov in 0.3-m high grass, —0.4 m from a gravel railroad right-of-way and 20 m from a wetland near Aberdeen, South Dakota (45° 27' 28" N, 98° 24' 58" W). Examination of the nest revealed two small (—46^9 mm) and two medium (—52- 55 mm) eggs. The small eggs were tan in color and the medium eggs were olive-tan and lighter in color. During later visits the nest contained four small and four medium eggs on 14 May 2009, six small and four medium eggs on 16 May 2009, and seven small and four medium eggs on 17 May 2009. A hen Blue-winged Teal flushed from the nest on 17 May 2009 during our approach and down and breast feathers diagnostic of the species had been added to the nest. The ultimate fate of the nest could not be ascertained due to the temporary nature of our visit in South Dakota. DISCUSSION Northern Shoveler eggs are described as pale olive-buff to a pale greenish gray in color (Bent 1923b, Palmer et al. 1976b, Bellrose 1980). Average length for Northern Shoveler eggs ranges from 52.2 to 52.85 mm (Bent 1923b, Palmer et al. 1976b, Dubowy 1996). Blue-winged Teal eggs are reported as light buff-tan (Bennett 1938), creamy tan (Glover 1956, Palmer et al. 1976a), or cream (Rohwer et al. 2002) in color and average length ranges from 46.4 to 47.1 mm (Bent 1923a, Glover 1956, Palmer et al. 1976a). The color and length of the eggs we observed suggests the medium eggs belonged to Northern Shoveler and the small eggs to Blue-winged Teal. Broley (1950) de- scribes Blue-winged Teal down feathers with conspicuous white centers and dark tips and breast feathers with an elongated dark sub-terniinal area and light proximal portion. Feathers in our nest matched this description. Observation of the hen Northern Shoveler on 12 May 2009, the hen Blue- winged Teal and feathers in the nest on 17 May 2009, along with the egg differences strongly suggests the Northern Shoveler parasitized the nest of the Blue-winged Teal. Blue-winged Teal prefer nesting in upland sites (Bent 1923a, Bennett 1938, Glover 1956, Johns- SHORT COMMUNICATIONS 613 gard 1975, Palmer et al. 1976a, Bellrose 1980). Rohwer el al. (2002) reported Blue-winged Teal rarely nest in high densities on islands or on muskrat {Ondatra zihethicns) houses and nests are dispersed, not readily accessible, and nest para- sitism is rare, but documentation is lacking. Bent (1923a) described an account of a Blue-winged Teal nest containing four eggs of the teal and five eggs of another duck species. Bellrose (1980) included Blue-winged Teal as victims of Redhead (Aythya americana) parasitism. Lokemoen (1991) reported that Blue-winged Teal nests had parasit- ism in eight of 1 3 nests on islands, but zero of 54 nests on peninsulas. Zimmerman et al. (2003) found Blue-winged Teal nests parasitized at rates from 24 to 72% over a 7-year period on small islands in North Dakota. Glover (1956) reported only one of 186 Blue-winged Teal nests examined was parasitized by another duck which was a Northern Shoveler. We examined data for 3,003 nests collected as described by the Northern Prairie Wildlife Re- search Center ( 1993) that we visited in northcentral North Dakota from May to July 1994-1996 (Garrettson and Rohwer 2001). Of these nests, 1 ,494 were Blue-winged Teal and only nine (0.6%) were reported as parasitized. None was parasitized by a Northern Shoveler. We found only one case of Northern Shoveler parasitism, and egg morphology and color suggested conspecific parasitism (1 of 1 29 Northern Shoveler nests). Weller (1959) listed parasitic egg-laying by Northern Shoveler in nests of Mallard (Anas platyrhynchos), American Wigeon (A. america- na), Cinnamon Teal (A. cyanoptera), and Red- head. Fournier and Hines (2001 ) found two of 669 .scaup nests parasitized by Northern Shovelers. Zimmerman et al. (2003) reported Northern Shovelers parasitized three of 1,642 nests, but did not identify the parasitized species. Only Glover (1956) reported parasitic egg-laying by a shoveler in the nest of a Blue-winged Teal. Considering the abundance of these species and overlap in breeding range and nesting habitat, it is surprising that parasitism of Blue-winged Teal nests by Northern Shovelers is so rarely reported. ACKNOWLEDGMENTS Louisiana State University and Delta Waterfowl Foun- dation supported nest record collection and associated research. The manuscript benefited from the comments of W. P. Johnson, P. I. Padding, two anonymous referees, and the editor. The findings and conclusions in this article are those of the authors ajid do not necessarily repre.sent the views of the U.S. Fish and Wildlife Service. LITERATURE CITED Bellrose, F. C. 1980. Ducks, geese and swans of North America. Third Edition. Stackpole Books, Harrisburg, Pennsylvania, USA. Bennett, L. J. 1938. The Blue-winged Teal; its ecology and management. Collegiate Press Inc., Ames, Iowa, USA. Bent, A. C. 1923a. Blue-winged Teal. Pages 111-121 in Life histories of North American wild fowl. Part 1. U.S. National Museum Bulletin Number 126. Bent, A. C. 1923b. Northern Shoveler. Pages 135-143 in Life histories of North American wild fowl. Part I. U.S. National Museum Bulletin Number 126. Broley, J. 1950. Identifying nests of the Anatidae of the Canadian prairies. Journal of Wildlife Management 14:452M56. Dubowy, P. j. 1996. Northern Shoveler (Anas clypeata). The birds of North America. Number 217. Fournier, M. A. and J. E. Hines. 2001. Breeding ecology of sympatric Greater and Lesser scaup (Aythya inarila and Aythya affinis) in the subarctic Northwest Territories. Arctic 54:444M56. Garrettson, P. R. and F. C. Rohwer. 2001. Effects of mammalian predator removal on production of nesting ducks in North Dakota. Journal of Wildlife Manage- ment 65:398-405. Glover, F. A. 1956. Nesting and production of the Blue- Winged Teal (Anas discors Linnaeus) in northwest Iowa. Journal of Wildlife Management 20:28^6. JoHNSGARD, P. A. 1975. Waterfowl of North America. Indiana University Press, Bloomington. USA. Lokemoen, J. T. 1991. Brood parasitism among waterfowl nesting on islands and peninsulas in North Dakota. Condor 93:340-345. Northern Prairie Wildlife Research Center. 1993. Northern Prairie Wildlife Research Center: Center nest tile cooperator’s instniction manual. USDI, Northern Prairie Wildlife Re.search Center, Jamestown, Nonh Dakota. USA. Palmer, R. S., F. C. Bellrose, and D. S. Farner. 1976a. Blue-winged Teal (Anas discors). Pages 463^82 in Handbook of North American birds. Volume 2. Waterfowl (R. S. Palmer, Editor). Yale University Press. New Haven, Connecticut. USA. Palmer. R. S., F. C. Bellrose. L. K. Sowls. and A. W. SCHORGER. 1976b. Northern Shoveler {Anas clvpeaia). Pages 498-515 in Handbook of North American birds. Volume 2. Waterfowl (R. S. Palmer. Editor). Yale University Press, New Haven, Connecticut. USA. Rohwer, F. C., W. P, John.son. and E. R, Loos. 2002. Blue-winged Teal (Anas di.scor.'i). The birds of North America. Number 625. Sayler, R. D. 1992. Ecology and evolution of brood parasitism in waterfowl. Pages 290-322 in Ecology and management of breeding waterfowl (B. D. J. Batt. A. D. Afton. M. G. Anderson, C. D. Ankney, D. H. 614 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 3. September 2010 Johnson, J. A. Kadlec, and G. L. Krapu, Editors). University of Minnesota Press, Minneapolis, USA. U.S. Department of Interior (USDI). 2009. Waterfowl population status, 2009. USDI, Fish and Wildlife Service, Washington, D.C., USA. Weller, M. W. 1959. Parasitic egg laying in the Redhead (Aythya americana) and other North American Anati- dae. Ecological Monographs 29:333-365. Zimmerman, A. L., M. A. Sovada, T. K. Kessler, and R. K. Murphy. 2003. Nest parasitism on constructed islands in northwestern North Dakota. Prairie Natural- ist 35:65-80. The Wilson Journal of Ornithology 122(3):614-617, 2010 Nest Predators of Ground-nesting Birds in Montane Forest of the Santa Catalina Mountains, Arizona Chris Kirkpatrick’ and Courtney J. Conway--^ ABSTRACT. — We used time-lapse video cameras and track plates to identify nest predators of Red-faced Warblers (Cardellina riihrifrons) and Yellow-eyed Juncos (Junco phaeonotus) in high-elevation (> 2,300 m) forests of the Santa Catalina Mountains in southeastern Arizona. Mammals, especially gray fox (Urocyon cinereoargenteus) and cliff chipmunk {Ta- niias dorsalis), were the principal nest predators of Red- faced Warblers and Yellow-eyed Juncos within our study system, accounting for 89% of all nest depreda- tions. Our study is one of the first to use video cameras at real nests to document the prevalence of nest predators in montane forest ecosystems. Additional research is needed to learn if mammals are the dominant nest predators in other montane environments. Received 23 October 2009. Accepted 24 March 2010. Nest depredation is the most common cause of reproductive failure of small land birds (Ricklefs 1969, Martin 1992). However, we lack reliable data on the identity and relative importance of nest predator species for most birds (Thompson 2007). Information on nest predators is particu- larly sparse for birds that breed in high-elevation (>2,300 m) forests in sky island mountain ranges of the southwestern United States. For example, >89% of nest failures of ground-nesting birds in high-elevation forests of the Santa Catalina ' Arizona Cooperative Fi.sh and Wildlife Research Unit, .School of Natural Resources and the Environment, 325 Biological Sciences Ea.st, University of Arizona, Tucson, AZ 85721, USA. ^U.S. Geological Survey, Arizona Cooperative Fish and Wildlife Research Unit, School of Natural Resources and the Environment, 325 Biological Sciences East, University of Arizona, Tucson, AZ 85721, USA. ’Corresponding author; cconway@usgs.gov Mountains in southeastern Arizona is attributed to nest depredation (Kirkpatrick and Conway 2010). However, the identity of even the most common nest predators remains unknown. Iden- tification of nest predators is a critical first step in managing breeding populations of birds and in discerning how nest depredation shapes nest-site selection, life history strategies, and breeding behavior in birds (Weatherhead and Blouin- Demers 2004, Thompson 2007). We used time-lapse video cameras (henceforth “video cameras’’) to monitor nests of Red-faced Warblers (Cardellina rubrifrons) and Yellow- eyed Juncos (Junco phaeonotus) in montane forests of the Santa Catalina Mountains in southeastern Arizona. Use of video cameras to monitor bird nests provides a reliable method to identify common nest predators (McQuillen and Brewer 2000, Williams and Wood 2002), but the technology is expensive and labor intensive and often only a small sample of nests can be monitored (Thompson 2007). We augmented video camera monitoring with track plates (Wilson and Delahay 2001) to identify potential nest predator species within our study system. METHODS Study Area. — We conducted research in five 16-20 ha study plots (2,300-2,800 m elevation) within several forested drainages (Kirkpatrick and Conway 2010) and one forested ridge (Kirkpatrick and Conway 2005) in the Santa Catalina Moun- tains, Pima County, Arizona, USA. Temperatures within our study area varied from an average of 6.5° C in mid-April to an average of 17.6° C in mid-July (Decker and Conway 2009). SHORT COMMUNICATIONS 615 Nest Monitoring. — We searched lor and mon- itored nests ot Red-faced Warblers and Yellow- eyed Juncos in onr five study plots from late April to mid-July in 2005-2006. Red-faced warblers and Yellow-eyed Juncos are the two most common ground-nesting birds within our study area: they breed concurrently, share similar nest- site characteristics, and nest in close association (Kirkpatrick and Conway 2010). Thus, we were reasonably confident that nests of both species would be susceptible to the same community of nest predators. Video Cameras.— Wc monitored 16 Red-faced Warbler and 23 Yellow-eyed Junco nests with continuously-operated, time-lapse (5 frames/sec), video cameras equipped with infrared illumina- tion for night-time photography (Fieldcam TLV, Fuhrman Diversified Inc., Seabrook, TX, USA; McQuillen and Brewer 2000). The nesting stage at which we first set out video cameras at nests varied: building (19% of nests), egg laying (12%), incubation (54%), and nestling (15%). We attached a small (5X2X2 cm) camera via a small articulated arm to a small wooden stake buried in the ground ~ 1 m from the nest and positioned the camera so that it was 40-60 cm from the nest. We concealed the camera and its articulated arm with camouflage fabric and ran a cable (hidden under leaf litter) > 15 m from the video camera to a concealed location where we placed the video recorder and battery. We visited each video recorder every 24 hrs to change video tapes and batteries, and to check the status of the nest remotely with a hand-held video monitor. We occasionally approached nests along a defined trail to verify nest contents or to adjust cameras. We made no effort to mask our scent trail during these nest checks. Human scent trails leading to artificial ground nests in our study area did not increase the probability of nest depreda- tion (Kirkpatrick and Conway 2009). We left video cameras at nests until young fledged or nests failed; we then reviewed video tapes of all depredated nests to identify the species of nest predator and the date and time of each depreda- tion event. We assumed that multiple visits to a nest by the same predator species were the result of a single individual of that species. Track Plates. — We placed 38 track plates at 100-m intervals along the center of four of our five study plots during the middle of the bird breeding season (3 Jun 2007), alternating loca- tions of the track plates from one side of the drainage to the other. We placed the plates on relatively flat ground at a random distance <50 m from the drainage bottom (T ± SD: 26 ± 10 m). We chose a 50-m distance interval becau.se Red- faced Warblers and Yellow-eyed Juncos select nest sites that average 26 and 21 m, respectively, Irom drainage bottoms within our study area (Kirkpatrick and Conway 2010). We used a protocol similar to Cain (2001) to construct track plates from galvanized steel sheeting (60 X 91 cm) and taped a piece of contact paper (30 X 46 cm) to the middle of each plate. We sprayed a mixture containing two parts isopropyl alcohol and one part blue carpenter’s chalk (Drennan et al. 1998) on the track plate around the contact paper and placed a Japanese Quail {Coturnix japonica) egg covered in chicken {Callus gallus) egg yolk on the center of the contact paper. We calculated an index of predator activity based on the percentage of the 38 track plates that had > 1 identifiable track of a predator species following a 6-day exposure period (Dren- nan et al. 1998). RESULTS Video Cameras. — Nineteen of 39 nests with video cameras were successful, 17 were depre- dated (total of 18 depredation events; Table 1), two were abandoned due to inclement weather, and one was abandoned for an unknown reason. Both Red-faced Warbler and Yellow-eyed Junco nests were depredated by gray foxes {Urocyon cinereoargenteiis), cliff chipmunks {Tamias dor- salis), and Steller’s Jays {Cyanocitta stelleri). One Yellow-eyed Junco nest was depredated partially by a gray fox and a woodrat (likely Neotoma mexicana). Gray foxes and the woodrat depredat- ed nests at night (67% of all nest depredation events), whereas cliff chipmunks and Steller’s Jays depredated nests during the day (33% of all nest depredation events; Table 1). Gray foxes depredated video camera nests both early and later in the avian breeding season (27 Apr-19 Jul), whereas cliff chipmunks depredated video camera nests only later in the .season (1 Jun-1 Jul). Track Plates. — We identified tracks of six mammal and one bird species on track plates in early June 2007. Indices of predator activity were 50% for gray fox, 32% for mice (Peromyscus spp.), 32% for cliff chipmunk, 8% for Abert’s squirrel {Sciurus aberti), 5% for Steller’s Jay, 3% for bobcat {Lynx rufus), and 3% for domestic dog {Canis lupus familiaris). Pilot work using track 616 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 3, September 2010 TABLE 1. Depredation events at Red-faced Warbler and Yellow-eyed Junco video camera nests by four nest predator species in high-elevation (>2.300 m) forest within the Santa Catalina Mountains, Arizona (late Apr to mid Jul, 2005- 2006). Depredation events Nest predator Red-faced Warbler nests Yellow-eyed Junco nests Totals Time range (hrs) of depredations'* Gray fox 3 8 1 1 2004-0330 Cliff chipmunk 2 2 4 0645-1911 Steller’s Jay 1 1 2 0800-1551 Woodrat*’ 0 1 1 0200-0420^ Totals 6 12 18 Mountain Standard Time. Likely N. me.xicuna (Kirkpatrick and Conway 2006). '■ A woodrat depredated the brooding female Yellow-eyed Junco at 0200 hrs and then returned to the nest to depredate one of three nestlings at 0420 hrs. A gray fox depredated the two remaining ne.stlings (which were still alive) at 0218 hrs the next day (Kirkpatrick and Conway 2006). plates in early May 2007 also recordecJ rock squirrel (Spermophilus variegatiis) and Common Raven {Connis corax) within our study area (C. Kirkpatrick, unpubl. data). DISCUSSION Mammals, especially gray fox and cliff chip- munk, were the principal predators of Red-faced Warbler and Yellow-eyed Junco nests within our study system, accounting for 89% of all nest depredations documented on video. Gray fox and cliff chipmunk are common in montane forests throughout southeastern Arizona (Hoffmeister 1956, Lange 1960) and are known to depredate bird nests (Burt and Grossenheider 1976, Hart 1992). Other potential mammalian nest predators, including mice, Abert’s squirrels, bobcats, do- mestic dogs, and rock squirrels were present within our study area (documented by track plates and visual observations by field personnel) but were not recorded depredating nests by video cameras (perhaps because some of these species occur at low densities; C. Kirkpatrick, pers. obs.). Additional video camera monitoring may be necessary to ascertain whether some of these mammals (especially mice; Bradley and Marzluff 2003, King and DeGraaf 2006) are occasional nest predators within our study system. Some nest predators may avoid or be attracted to nests with video cameras and the track plates provide a means by which one can identify nest predators present in the study system that were not captured by video surveillance. Data from track plates do not provide an unbiased estimate of density or movement rates of nest predators because they are baited and some predators may avoid or be attracted to them, but they do provide information on the suite of potential nest predators and their relative levels of activity (relative to other locations where track plates are placed with similar methods). The prevalence of mammalian nest predators in our study system is surprising given that snakes are considered to be the principal nest predator of many New World passerine birds (Weatherhead and Blouin-Demers 2004), especially in the southern USA (Thompson 2007). We believe we did not record snakes as nest predators because our study area was: ( 1 ) >2,300 m above sea level in an environment thermally inhospitable to most snake species (M. J. Goode, University of Arizona, pers. comm.); and (2) beyond the northern range limits of several montane rattle- snake species (Stebbins 1985), one of which (twin-spotted rattlesnake [Crotahis pricei]) is known to occasionally depredate Yellow-eyed Junco nests at high elevation (2,560 m; Gumbart and Sullivan 1990). We observed only two snakes (Sonoran mountain kingsnake [Lampropeltis pyr- ometana] and Arizona black rattlesnake [Crotalus cerheriis]) near our study plots and none within the boundaries of our study plots despite thou- sands of person hours in the field from 2002 to 2009 (C. Kirkpatrick, unpubl. data). Our study is one of the first to use video cameras at real nests to identify nest predators and measure their relative importance in montane forest ecosystems (similar results were obtained in montane forests of northern Arizona where mammals caused 62% of diurnal nest depreda- tions; T. E. Martin, University of Montana, pers. comm.). Additional research is needed to ascer- tain if mammals are the principal nest predators in other high-elevation environments. Selection SHORT COMMUNICATIONS 617 should favor adaptations by birds to reduce exposure to numerous olfactory-oriented, noctur- nal nest predators (e.g., gray foxes, woodrats, mice) in montane environments, just as selection is thought to favor behaviors that reduce exposure to visually-oriented, diurnal predators in other ecosystems (Weatherhead and Blouin-Demers 2004). Identifying the commonalities in the suite of primary nest predators within different ecosys- tems may help explain differences in behavior and life history traits among species of birds. Montane forests in the Santa Catalina Mountains provide an ideal environment to examine these issues given that breeding populations of ground-nesting birds at higher elevations (>2,300 m) are contiguous with breeding populations at lower-elevations (>1,800 m) where snakes and other nest predators are likely to be more common. ACKNOWLEDGMENTS This research was funded in part by the U.S. Fish and Wildlife Service. We thank S. J. Sferra (U.S Bureau of Reclamation) for the loan of video cameras and M. H. Ali, M. M. Eastwood, D. D. LaRoche, N. M. Nardello, A. M. Purdy, E. T. Rose, T. A. Selvidge, D. D. Tracy, and Anna Westling-Douglass for assistance locating nests and oper- ating video cameras. R. L. Peterson and the University of Arizona Steward Observatory provided field housing and J. D. Taiz (U.S. Forest Service) provided logistical support. Any use of trade names is for descriptive purposes only and does not imply endorsement by the U.S. Government. LITERATURE CITED Bradley, E. and J. M. Marzluff. 2003. Rodents as nest predators: influences on predatory behavior and consequences to nesting birds. Auk 120:1 180-1 187. Burt, W. H. and R. P. Grossenheider. 1976. A field guide to mammals: field marks of all North American species found north of Mexico. Houghton Mifflin Company, Boston, Massachusetts, USA. Cain, J. W. 2001. Nest success of Yellow Warblers (Dendroica petechia) and Willow Flycatchers {Empi- donax iraillii) in relation to predator activity in montane meadows of the central Sierra Nevada, California. Thesis. California State University, Sacra- mento, USA. Decker, K. L. and C. J. Conway. 2009. Effects of an unseasonable snowstorm on Red-faced Warbler nest- ing .success. Condor 1 1 1:392-395. Drennan, j. E., P. B. Beier, and N. L. Dodd. 1998. Use of track stations to index abundance of sciurids. .lournal of Mammalogy 79:352-359. Gumbart, T. C. and K. a. .Sullivan. 1990. Predation on Yellow-eyed Junco nestlings by twin-spotted rattle- snakes. Southwestern Naturalist 35:367-368. Hart, B. E. 1992. Tamais dorsalis. Mammalian Species 399. Hoffmeister, D. F. 1956. Mammals of the Graham (Pinaleno) Mountains, Arizona. American Midland Naturalist 55:257-288. King, D. I. and R. M. DeGraaf. 2006. Predators at bird nests in a northern hardwood forest in New Hamp- shire. Journal of Field Ornithology 77:239-243. KiRKPATRtCK, C. AND C. J. CoNWAY. 2005. Reproductive success and habitat associations of riparian birds within the sky island mountains of southeastern Arizona. Wildlife Research Report #2005-07. USGS, Arizona Cooperative Fish and Wildlife Research Unit, Tucson, USA. KtRKPATRiCK, C. AND C. J. CoNWAY. 2006. Woodrat (Neotoma) depredation of a Yellow-eyed Junco (Junco phaeonotus) nest. Southwestern Naturalist 51:412- 414. Kirkpatrick, C. and C. J. Conway. 2009. Effects of researcher disturbance on nesting success of Red-faced Warblers: a species of conservation concern. Wildlife Research Report #2009-01. USGS, Arizona Coopera- tive Fish and Wildlife Research Unit, Tucson, USA. Kirkpatrick, C. and C. J. Conway. 2010. Importance of montane riparian forest and influence of wildfire on nest-site selection of ground-nesting birds. Journal of Wildlife Management 74:729-738. Lange, K. 1. 1960. Mammals of the Santa Catalina Mountains. American Midland Naturalist 64:436^58. Martin, T. E. 1992. Interaction of nest predation and food limitation in reproductive strategies. Current Ornithol- ogy 9:163-197. McQuillen, H. L. and L. W. Brewer. 2000. Methodo- logical considerations for monitoring wild bird nests using video technology. Journal of Field Ornithology 71:167-172. Ricklefs, R. E. 1969. An analysis of nesting mortality in birds. Smithsonian Contributions to Zoology 9:1-48. Stebbins, R. C. 1985. A field guide to western reptiles and amphibians. Houghton Mifflin Company, Boston, Massachusetts, USA. Thompson, F. R. 2007. Factors affecting nest predation on forest songbirds in North America. Ibis 149:98-109. Weatherhead, P. J. and G. Blouin-Demers. 2004. Understanding avian nest predation: why ornitholo- gists should study snakes. Journal of Avian Biology 35:185-190. Williams, G. E. and P. B. Wood. 2002. Are traditional methods of determining nest predators and nest fates reliable? An experiment with Wood Thrushes (Hvlo- cichhi inioitelina) using miniature video cameras. Auk 1 19:1 126-1 132. Wilson, G. J. and R. J. Delahay. 2001. A review of methods to estimate the abundance of terrestrial carnivores using field signs and observation. Wildlife Research 8:151-164. 618 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 3. September 2010 The Wilson Journal of Ornithology 122(3):6 18-620, 2010 Feeding a Foreign Chick: a Case of a Mixed Brood of Two Tit Species Toshitaka N. Suzuki^’^ and Yuko Tsuchiya^ ABSTRACT. — We report on a mixed brood of Great (Pants major) and Varied (Poecile varius) tits. The brood consisted of one Great Tit and five Varied Tits. Incubation and food provisioning to nestlings were by adult Varied Tits, but not by adult Great Tits. The foreign Great Tit nestling tended to receive more food than the average Varied Tit nestling. Provisioning of the foreign Great Tit by Varied Tits lasted even after fledging. We discuss the causes and outcomes of this rare phenomenon. Received 5 December 2009. Accepted 9 March 2010. Nestlings of different species are rarely present in the same nest (Shy 1982), apart from cuckoo- host interactions (Payne 1977, Rothstein 1990). Mixed broods have been reported among several species of passerines including titmice, nuthatch- es, flycatchers, and thrushes (e.g., Mackenzie 1950, 1954; Slagsvold 1998; Dolenec 2002; Robinson et al. 2005; Govoni et al. 2009). Two types of mixed brood rearing are known: only one species raises foreign young along with their own (Slagsvold 1998), and both pairs from different species raise the mixed brood together (i.e., nest sharing; Robinson et al. 2005, Govoni et al. 2009). The major cause of mixed broods in both cases has been explained as interspecific competition for a desirable nest site. The occurrence of mixed broods provides an opportunity to investigate the evolutionary origin of interspecific brood parasitism (Slagsvold 1998), and to examine the role of learning in species recognition (Hansen et al. 2008, 2009). However, detailed observations ot mixed broods have not been reported with the exception of studies in which researchers experimentally cross- foster eggs between different species (e.g., Slagsvold 1998). This .seems to be due to the ' Department of Life Seienee, Graduate Sehool ot .Science, Rikkyo University, 3-.34-1 Nishi-Ikebukuro, Toshima, Tokyo 171-8501. Japan. ^PREC Institute Inc., 1-7-15 Kyomachibori, Nishi, Osaka 5.5()-00().3, Japan. 'Corresponding author; e-mail: toshi.n.suzuki@gmail. com rarity of occurrence of mixed broods as well as insufficient opportunity for video recordings in past decades. We report a case of a naturally-occurring mixed brood of two tit species. The mixed brood consisted of one Great Tit {Parus major) and five Varied Tits {Poecile varius). The Varied Tit pair raised the foreign Great Tit with their own nestlings in the same nest box. Feeding events inside the nest box were video recorded. OBSERVATIONS We placed 25 nest boxes (17 X 17 X 30 cm, 3 cm entrance) in a mixed deciduous-coniferous forest in 2006 in Kamizawa, Nagano, Japan (36° 21' N, 138° 35-36' E) and monitored them during the breeding season. The forest consisted mainly of Japanese larch {Larix leptolepis), Mongolian oak {Qitercus mongolica), and giant dogwood {Swida controversa). Nine of the 25 boxes were used by Great Tits, one box by Varied Tits, and one contained a mixed brood of Great and Varied tits. We visited the mixed brood nest 1 1 times in the period of 1 month (5 times during the egg stage, 5 times during the nestling stage, and once on the day of fledging). Incubation by one of the Varied Tit adults (probably the female) was confirmed during the nest visitations, but not by either of the Great Tits. We identified six eggs in the nest and all eggs hatched synchronously on 16 May. The Varied Tit also sat on the naked chicks on the day of hatching. We recorded feeding events inside the nest box during the nestling period by placing a KPC-EX500 CCD camera (KT&C Co., Seoul, Korea) on the ceiling of the nest box connected to a GR-DF590 digital video camera (Victor Co., Kanagawa, Japan). Only the adult Varied Tits provisioned food (caterpillars, bees or spiders) for the nestlings for the entire 805 min recorded, and no adult Great Tits visited the nest box. We analyzed the provisioning behavior of the adult Varied Tits using the video recordings from inside the nest box. We could easily discriminate between the foreign Great Tit and Varied Tit chicks by morphology in the late nestling stage. SHORT COMMUNICATIONS 619 (/) D) c 'c g w ■> o L_ Q. 0) E D FIG. 1. Provisionings by adult Varied Tits to a foreign Great Tit nestling and Varied Tit nestlings in 480 min of video records. The bar representing the Varied Tit nestlings indicates the mean value for the five nestlings. The broken line indicates the expected number of provisionings if all nestlings are provided with food randomly. and used the video data recorded at this stage (14, 15, and 16 days of age) for analysis. We analyzed 154 provisioning events during 480 min of video recordings. The adult Varied Tits fed several different nestlings at each provisioning visit (x ± SD: 2.2 ± 0.7 nestlings; ranging from 1 to 4). The mean rate of provisioning visits was 8.8/hr. We tested whether food provisionings by adult Varied Tits were biased between the Great Tit nestling and the Varied Tit nestlings using a G- test (Sokal and Rohlf 1995). A significant difference was found (G-test with Williams’ correction: Gj = 6.36, P = 0.012; Fig. 1) when comparing the observed number of provisionings received by the Great Tit nestling and the five Varied Tit nestlings with the expected number of provisionings (i.e., the value if all nestlings obtain food randomly). This implies the adult Varied Tits provided more food to the Great Tit young than they did to young Varied Tits on average. All young fledged sequentially between sunrise and noon on 1 June when they were 17 days of age. The Great Tit and Varied Tit fledglings were confirmed to be fed by the adult Varied Tits after fledging. DISCUSSION This is the first report of an occurrence of a Great and Varied tit mixed brood, and the first video analysis of a naturally occurring mixed brood. The foreign Great Tit nestling appeared to be smaller than Varied Tit chicks, but it tended to obtain more food than the average Varied Tit nestling (Fig. I). This appears to be the result of active begging behavior by the Great Tit; it begged for food in 100% of the provisioning visits by the adults, whereas Varied Tit young begged only in 85% of the provisioning visits, on average. The Varied Tit adults fed the Great Tit even after fledging, suggesting they learned to recognize the foreign chick as if it was their own. This observed mixed brood is unlikely to be due to a new breeding strategy of Great Tits in this population because this is the only known case in our 4-year study (breeding seasons from 2006 to 2009), during which time we monitored 106 nest boxes of cavity-nesting birds. A previous report suggests that Great Tits have not evolved intraspecific brood parasitism (Kempenaers et al. 1995). It is more likely the mixed brood is a nonadaptive by-product of interspecific nest-site competition. Interspecific nest usurpations in our study site were often observed between Great Tits and Varied Tits during the beginning of the breeding season. It may be possible the observed mixed brood occurred either when Varied Tits took over the Great Tits’ nest in which one egg had already been laid, or when Great Tits failed to usurp the nest of Varied Tits. Feeding of a foreign chick is obviously costly for adult Varied Tits, but the fitness consequence for the foreign Great Tit remains unclear. The Great Tit nestling tended to obtain more food than the average Varied Tit nestling, but we cannot reject the possibility that quality and quantity of food may be insufficient for the foreign Great Tit. A previous report suggests the dietary needs of nestlings may be different between these two tit species (Mizutani and Hijii 2002). The Great Tit may also have other costs, such as problems with species recognition when subsequently trying to mate and breed. Species recognition is imprinted early in life, and may be irreversible once it has been establi.shed (Hansen et al. 2008, 2009). The occurrence of mixed broods may represent an initial step in evolution of interspecific brood parasitism (Slagsvold and Hansen 2001). The pre.sent observation provides an example that a foreign chick can be raised to fledging by a different species. Future field observations may reveal further examples, which may give us valuable insights into the evolution of interspe- cific brood parasitism. 620 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 3. September 2010 ACKNOWLEDGMENTS We are grateful to Masayoshi Kamioki, O. K. Mikami, Kihoko Tokue, Keisuke Ueda, and two anonymous reviewers for helpful comments on this manuscript. We also thank the Lorestry Agency and the Picchio Wildlife Research Center for permitting us to study at this site. LITERATURE CITED Dolenec, Z. 2002. A mixed brood of nuthatch (Sitta europaea) and Great Tit (Parus major) species. Natura Croatica 1 1 : 103-105. Govoni, P. W., K. S. Summervill, and M. D. Eaton. 2009. Nest sharing between an American Robin and a Northern Cardinal. Wilson Journal of Ornithology 121:424-426. HAN.SEN, B. T., L. E. JOHANNESSEN, AND T. SLAGSVOLD. 2008. Imprinted species recognition lasts for life in free-living Great Tits and Blue Tits. Animal Behaviour 75:92 1-927. Hansen, B. T., L. E. Johannessen, and T. Slagsvold. 2009. Interspecific cross-fostering affects mate guard- ing behaviour in Great Tits (Pams major). Behaviour 146:1349-1361. Kempenaers, B., R. Pinxten, and M. Eens. 1995. Intraspecific brood parasitism in two tit Pams species: occurrence and responses to experimental parasitism. Journal of Avian Biology 26:1 14-120. Mackenzie, J. M. D. 1950. Competition for nest sites among hole breeding species. British Birds 43:184-185. Mackenzie, J. M. D. 1954. Redstarts reared in tits’ nests. Scottish Naturalist 66:146-154. Mizutani, M. and N. Huh. 2002. The effects of arthropod abundance and size on the nestling diet of two Pams species. Ornithological Science 1:71-80. Payne, R. B. 1977. The ecology of brood parasitism in birds. Annual Review of Ecology and Systematics 8:1-28. Robinson, P. A., A. R. Norris, and K. Martin. 2005. Interspecific nest sharing by Red-brested Nuthatch and Mountain Chickadee. Wilson Bulletin 117:400- 402. Rothstein, S. 1. 1990. A model .system for coevolution: avian brood parasitism. Annual Review of Ecology and Systematics 21:481-508. Shy, M. M. 1982. Interspecific feeding among birds: a review. Journal of Field Ornithology 53:370-393. Slagsvold, T. 1998. On the origin and rarity of interspecific nest parasitism in birds. American Naturalist 152:264-272. Slagsvold, T. and B. T. Hansen. 2001. Sexual imprinting and the origin of obligate brood parasitism in birds. American Naturalist 158:354—367. SOKAL, R. R. AND F. J. Rohlf. 1995. Biometry. Third Edition. W. H. Freeman, New York, USA. The Wilson Journal of Ornithology 122(3):620-622, 2010 Infanticide by an Eastern Phoebe Gary Ritchison'-^ and Brandon T, Ritchison' ABSTRACT. — We videotaped an Eastern Phoebe (Sayornis phoebe) nest in central Kentucky on 22 June 2008 and documented a case of infanticide. An adult male phoebe, likely the resident male, was captured at the nest site on 21 June 2008. The resident female fed nestlings (3, ~8 days of age) 69 times during a 4.1 -hr taping .session (0815-1221 hrs EOT) the next day, but the resident male was not observed at the nest. Beginning at 0923 hrs, an intruding phoebe (likely a male) occasionally visited the nest, sometimes pecking at. but not feeding, the nestlings; this phoebe initiated a vigorous attack at 1151 hrs, pecking and pulling the head of one of the nestlings for 25 min before pulling it out of the nest. The intruding phoebe returned after a short absence, just before the end of the videotape, and began attacking another nestling. The nest was empty when checked the next day. Suitable territories and nest sites may be limiting factors for Eastern Phoebes, and ' Department of Biological Sciences, Eastern Kentucky University, Richmond, KY 40475, USA. ^Corresponding author; e-mail: gary.ritchison@eku.edu infanticide by non-breeding males may enhance their reproductive success if the resident female or a new female subsequently pairs with the male and initiates a new nest. Received II Januarv 2010. Accepted 24 March 2010. Infanticide has been reported in many species of birds, primarily in cooperative breeders, colonial breeders, and polygamous species (Freed 1986). Infanticide has been reported among socially monogamous species of birds, but primarily species where breeding opportunities may be limited by biased .sex ratios or availability of needed resources such as nest sites (Freed 1986; Robertson and Stutchbury 1988; Mpller 1988, 2004). Infanticide may be adaptive either by providing access to a needed resource, such as a nest site (Robertson and Stutchbury 1988), or by shortening the time until a new breeding attempt (Freed 1986). SHORT COMMUNICATIONS 621 Most reports ot infanticide among socially monogamous species have involved secondary cavity-nesting species where availability of nest sites may limit breeding opportunities (Freed 1986, Robertson and Stutchbury 1988, Kermott et al. 1991 ). We report a case of infanticide by an Eastern Phoebe {Sayornis phoebe), a non-cavity- nesting species with specific nest-site require- ments that may limit breeding opportunities of some individuals. OBSERVATIONS We studied Eastern Phoebes at the Blue Grass Army Depot (BGAD) from mid-March to mid- July 2008. The BGAD is in Madison County, Kentucky and encompasses 5,907 ha of pastures with trees and woodlands interspersed throughout. Concrete shelters (about 2.5 X 5 X 2.5 m; « = 54), each with two open doors (1 X 2 m) at opposite corners occur throughout the depot to provide depot personnel with protection in case of emergencies. The shelters are rarely visited by Army personnel, but phoebes readily enter the shelters to nest. We located a phoebe nest on 21 June 2008 with three nestlings ~8 days of age in a shelter that, for security reasons, we visited infrequently. An adult phoebe entering the shelter was captured in a mist net on the same day and banded with a USGS aluminum band and a unique combination of three colored-plastic bands. The phoebe was identified as a male based on the presence of a cloacal protuberance. Standard measurements were taken and the male, in apparently good condition, was released. The nest was videotaped from 08 1 5 to 1221 hrs EDT on 22 June 2008 with a camcorder placed 2 m from the nest. The camcorder was protected from the elements in the shelter so we did not retrieve the videotape until the next day (23 Jun 2008). We subsequently reviewed the tape and noted all visits to the nest by adult phoebes as well as any other recorded behaviors or vocalizations. An unbanded Eastern Phoebe, apparently the resident female, provisioned nestlings at a rate of 16.8 visits/hr during the 4.1 -hr videotape; the phoebe banded the previous day (assumed to be the resident male) was not observed at the nest. The female phoebe visited the nest and provi- sioned nestlings 35 times from 0815 to 1015 hrs, and always exhibited the same behavior, landing at the edge of the nest and quickly feeding a nestling. Another phoebe, also unbanded, but behaving differently, was first observed at 0923 hrs. Over the next 42 min, this phoebe hovered near the nest twice, briefly landed on the edge of the nest three times, and pecked the head of a nestling during two of those nest visits. During the third hour of taping (1015-1115 hrs), the resident female fed nestlings 16 times; the other phoebe entered the shelter and perched on the camcorder four times, but did not visit the nest. The resident female continued provisioning nestlings during the fourth hour of taping (1115- 1215 hrs). The ‘intruding’ phoebe entered the shelter and perched on the camcorder 21 times from 1115 to 1145 hrs, uttering ‘bipeaked vocalizations’ (Smith 1969) on multiple occasions and ‘initial peaked vocalizations’ (Smith 1969) twice. The female visited the nest three times while the other phoebe was on the camcorder; the presence of the intruding phoebe had no apparent effect on the female’s behavior. The intruding phoebe flew to the nest at 1151 hrs and began vigorously pecking and pulling one of the nestlings, focusing its attack on the nestling’s head. This phoebe pecked or pulled at the nestling about 1,250 times (~50 times/min) over the next 25 min. The resident female came to the nest ~5 sec after the attack was initiated, hovered briefly at the edge, then landed with beak agape and moved to within ~2 cm of the attacker for ~3 sec. The female then flew from the nest, returning 5 sec later and again briefly hovered near the nest. As the attack continued, the female left and re-entered the shelter and perched on the camcorder three times (but uttered no calls), and flew to the nest and provisioned nestlings five times, uttering no vocalizations and not interacting with the attacking phoebe. The intruding phoebe pulled the apparently dead nestling out of the nest at the conclusion of the 25-min attack, and returned to the nest —4.5 min later and began attacking a second nestling. The videotape ended 40 sec after this second attack began. The nest was empty when visited the next day with no sign ot adult or nestling phoebes in the shelter. We re-visited the shelter 12 days later; the nest was still empty and we observed no adult phoebes in or near the shelter. DISCUSSION Male Eastern Phoebes typically provision nestlings and defend nest sites and territories from conspecifics (Weeks 1994). Thus, the 622 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. 3. September 2010 absence of the male at our focal nest plus the presence of the intruding phoebe suggests the resident male was not present. The male we banded at the nest site, likely the resident male, appeared to be in good condition when released. The reason for the male’s absence is unknown, but could have been due either to abandonment of the territory (e.g., as a result of being captured in a mist net) or predation. Our observations indicate adult Eastern Phoe- bes at times engage in infanticide. Beheler et al. (2003:996) also noted instances of infanticide (‘probably by nonparental adults’) by Eastern Phoebes, and Weeks (1994) reported that infan- ticide at times occurred after a territorial male phoebe disappeared. Phoebes are sexually mono- morphic and both males and females utter ‘bipeaked’ and ‘initial peaked’ vocalizations (Smith 1969). Thus, we cannot be certain the intruding phoebe in our study was a male. However, the absence of the resident male suggests the infanticidal phoebe was a male. Infanticide by male phoebes may enhance repro- ductive success if the resident female or a new female subsequently pairs with the infanticidal male and initiates another nest. This behavior has been reported in other species of songbirds, including House Wrens {Troglodytes aedon; Freed 1986, Kermott et al. 1991), Tree Swallows (Tachycineta bicolor, Robertson and Stutchbury 1988), and Barn Swallows (Hirundo nistica', Mpller 2004). Suitable territories and nest sites may be limiting factors for Eastern Phoebes, and com- petition for these resources may be intense (Beheler et al. 2003). Phoebes occupy a variety of woodland and edge habitats (Graber et al. 1974), but the availability of suitable nest sites may be limited in many areas (Weeks 1994). Thus, some adult phoebes may be non-breeding floaters. Supporting this assumption is the rapid replacement of dead mates and rapid settlement of territories when new nest sites become available (Beheler et al. 2003). The disappear- ance of a territorial male repre.sents an oppor- tunity for non-breeding males to acquire the resources needed for reproduction. Phoebes are typically double-brooded (Weeks 1994) and, in our study area, second broods are generally initiated in late May or early June; new nests are rarely initiated after late June (GR, pers. obs.). A male phoebe acquiring a territory with 8-day- old nestlings on 22 or 23 June, as in our study, would likely have no opportunity to breed during the cun'ent breeding season if those nestlings developed normally and fledged (gen- erally when 16-18 days of age; Weeks 1994). Infanticide could potentially provide a male phoebe with a breeding opportunity. However, the nest in our study was empty when checked 12 days later and we observed no phoebes in or near the shelter, indicating the infanticidal male phoebe in our study did not subsequently breed in that shelter. ACKNOWLEDGMENTS We thank the U.S. Army for allowing us to conduct re.search at their facility, the EKU Research Committee for financial support, and C. E. Braun, H. P. Weeks Jr., and an anonymous reviewer for helpful comments. A short video clip of the infanticidal phoebe attacking and removing the nestling is posted at http://www.youtube.com/watch?- v=4mVQMpyqlrs. LITERATURE CITED Beheler, A. S., O. E. Rhodes Jr., and H. P. Weeks Jr. 2003. Breeding site and mate fidelity in Eastern Phoebes (Sayomis phoebe) in Indiana. Auk 120:990-999. Freed, L. A. 1986. Territory takeover and sexually selected infanticide in tropical House Wrens. Behavioral Ecology and Sociobiology 19:197-206. Graber, R. R., J. W. Graber, and E. L. Kirk. 1974. Illinois birds: Tyrannidae. Illinois Natural History Survey Biological Notes Number 86. 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. Condor 91:630-636. Moller, a. P. 1988. Infanticidal and anti-infanticidal strategies in the swallow Hirundo ru.sticci. Behavioral Ecology and Sociobiology 22:36.3-371. Moller, A. P. 2004. Rapid temporal change in frequency of infanticide in a passerine bird associated with change in population density and body condition. Behavioral Ecology 15:462^68. Robertson, R. J. and B. J, Stutchbury. 1988. Experimental evidence for sexually .selected infanti- cide in Tree Swallows. Animal Behaviour 36:749-753. Smith, W. J. 1969. Displays of Sayornis phoebe (Aves, Tyrannidae). Behaviour 33:283-322. Weeks, Jr., H. P. 1994. Ea.stern Phoebe (Sayornis phoebe). The birds of North America. Number 94. SnOr^T COMMUNICATIONS 623 The Wilson Journal of Ornithology 122(3):623-624. 2010 Marsh Wren Eats Small Fish Andrea J. Ayers'’^ and James W. Armacost Jr.' ABSTRACT. — We report an observation of a Marsh Wren [Cistothorus palustris) eating a small fish. On 29 November 2008, a Marsh Wren was observed catching and consuming a small fish, probably a mosquitofish (Ganibusia affinis), in a coastal marsh in Jefferson County, Texas. Marsh Wrens are known to eat invertebrates, primarily insects and spiders, and we believe this is the first record of a Marsh Wren capturing and consuming vertebrate prey. Received 24 August 2009. Accepted 22 January 2010. The Marsh Wren {Cistothorus palustris) is a small passerine that inhabits a variety of coastal and interior marshland habitats. Prior observations of the feeding habits of Marsh Wrens indicate they feed strictly on invertebrates, foraging in dense patches of shrubs or along the base of cattail (Typha spp.) stalks or other marsh vegetation, on or near the ground or water surface (Bent 1948, Kroodsma and Verner 1997). Marsh Wrens typically move constantly as they forage, gleaning prey from plant surfaces or from on or just below the water surface. Most prey items are too small for a person to identify from a few meters (Kroodsma and Verner 1997). While conducting fieldwork as part of a project to assess avian mortality and population-level effects associated with a transmission line in coastal marshes at the J.D. Murphree Wildlife Management Area (WMA) in Jefferson County, Texas, AJA made an opportunistic observation of a foraging Marsh Wren that captured and consumed a small fish. After reviewing Bent (1948) and Kroodsma and Verner (1997), we believe this is the first report of a Marsh Wren taking live vertebrate prey. OBSERVATIONS The J.D. Murphree WMA contains a diverse community of coastal wetlands, including fresh- water, intermediate, brackish, and saline wetlands (Texas Parks and Wildlife Department 2008). On ' Department of Biology, Lamar University, Beaumont, TX 77710, USA. ^Conesponding author; e-mail: andreajayers@hotmail. com 29 November 2008, at —1000 hrs CST, AJA observed a Marsh Wren perched above a roadside ditch that drained water from a former construc- tion site for the transmission line. A section of the ditch was partially overhung by matted vegetation that was residue from Hurricane Ike, which made landfall on the southeast Texas coast on 13 September 2008. The Marsh Wren was perched on a small dead limb —20 cm above and diagonally crossing the running water, looking downstream. The wren sat perched over the water observing small fish swimming below for a short time. While still perched, the wren bent down and plucked a fish from the water, never leaving its perch. The wren captured and handled the fish adeptly, positioning the fish in its beak to swallow it head-first. The wren proceeded to swallow the fish in small increments, alive and whole. We estimated the size of this fish to be approximately twice the length of the Marsh Wren’s beak. The exposed CLilmen of an adult Marsh Wren ranges in size from 12 to 14 mm (Ridgway 1904); thus, we estimate the fish to have been 24-28 mm in length. Afterwards the bird flew from the roadside ditch into denser marsh vegetation. This observa- tion was made from a distance of —4 m. using 8 X 40 binoculars, and the prey item, which had a silver, reflective body and a noticeable tail fin, was easily identifiable as a fish rather than an invertebrate. We speculate the fish was a mosquitofish {Ganibusia affinis), based on the type of marshland habitat. DISCUSSION This observation occurred 1 1 weeks after Hun-icane Ike made landfall on the southeast Texas coast. The J.D. Murphree WMA was covered by a storm surge that peaked at >4 m and persisted for several days (Berg 2009). This .salt water intrusion killed many plants in Texas coastal marshes, and abundance of aquatic inver- tebrates may have been reduced. A possible explanation for this Marsh Wren’s behavior may have been the lack of invertebrate prey, leading it 624 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 3. September 2010 to capture and consume a small vertebrate. Marsh Wrens typically forage in dense vegetation, making their foraging behavior difficult to observe (Kroodsma and Vemer 1997), and it is possible that vertebrates are a regular, but previously undocumented, component of their diet. We believe our observation of a Marsh Wren feeding on a small fish is the first record of this behavior and contributes new knowledge about the ecology of this species. Pieman et al. (1993) observed Marsh Wrens to destroy eggs and nestlings of competing birds in deep-marsh habi- tats, while Bump (1986) reported Marsh Wrens appeared to have sampled the contents in four of five cases when they punctured eggs. However, we consider consumption of egg contents incidental, and not primarily a foraging behavior. Larger species of wrens are known to take vertebrate prey on occasion. For example. Cactus Wrens {Campy- lorhynchus hrunneicapilliis) have been known to feed on amphibians and reptiles (Storer 1920, ProLidfoot et al. 2000). It isn’t surprising that Marsh Wrens, which feed mainly on aquatic invertebrates (Bent 1948, Kroodsma and Vemer 1997), would capture and consume small fish, given that other species of wrens are known to take vertebrate prey on occasion. ACKNOWLEDGMENTS We thank Golden Pass LNG Terminal LLC and Port Arthur LNG Terminal LLC for access and funding our research on bird interactions with transmission lines. We al.so thank the Texas Parks and Wildlife Department, J.D. Murphree Wildlife Management Area, for access to study sites. We thank Paul Nicoletto of Lamar University for assistance with fish identification and Steven Cardiff of the Louisiana State University Museum of Natural Science for assistance with beak measurements. We thank L. S. Johnson, Andrew Kasner, and an anonymous reviewer for comments that improved the manuscript. LITERATURE CITED Bent, A. C. 1948. Life histories of North American nuthatches, wrens, thrashers, and their allies. U.S. National Museum Bulletin Number 195. Berg, R. 2009. Tropical cyclone report: Hurricane Ike (AL092008) 1-14 September 2008. National Hurri- cane Center, Miami, Florida, USA. Bump, S. 1986. Yellow-headed Blackbird nest defense: aggressive responses to Marsh Wrens. Condor 88:328- 335. Kroodsma, D. E. and J. Verner. 1997. Marsh Wren (Cistothonis palustris). The birds of North America. Number 308. PiCMAN, J., M. L. Milks, and M. Leptich. 1993. Patterns of predation on passerine nests in marshes: effects of water depth and distance from edge. Auk 1 10:89-94. Proudfoot, G. a., D. a. Sherry, and S. Johnson. 2000. Cactus Wren (Campylorhynclius hrunneicapillus). The birds of North America. Number 558. Ridgway, R. 1904. The birds of North and Middle America: a descriptive catalogue of the higher groups, genera, species, and subspecies of birds known to occur in North America. Part 3. U.S. National Museum Bulletin Number 50. Storer, T. 1. 1920. Lizard eaten by Cactus Wren. Condor 22:159. Texas Parks and Wildlife Department. 2008. J.D. Murphree WMA. www.tpwd.state.tx.us/huntwild/ hunt/wma/find_a_wma/list/?id=40. SnOF^T COMMUNICATIONS 625 The Wilson Jounuil of Ornithology 1 22(3);625-626, 2010 Occasional Mimicry and Night-time Singing by the Western Curve-billed Thrasher (Toxostoma curvirostre pahneri) R. Roy Johnson' - and Lois T. Haight' ABSTRACT. — The first instance of vocal mimicry is reported for the western subspecies of Curve-billed Thrasher (Toxostoma curvirostre palmeri). A Curve- billed Thrasher engaged in countersinging with a migrating Black-headed Grosbeak (Pheucticus melano- cephahts) near Tucson, Arizona. Night-time singing by Curve-billed Thrasher is also documented for the first time. At least three responding Curve-billed Thrashers engaged in spontaneous song near Tucson, Arizona. Additional night-time singing was elicited by playback recordings Received 13 Fehritaiy 2010. Accepted 24 April 2010. The Curve-billed Thrasher {Toxostoma curvi- rostre) is placed in the Mimidae, a family that includes several species that mimic other species and/or sing at night (Bent 1948). Perhaps most widely known is the ability of Northern Mock- ingbird (Mimiis polyglottos) to mimic other species, and also sing both day and night during the breeding season (Demckson and Breitwisch 1992). Mimicry has been reported for the nominate subspecies group of the Curve-billed Thrasher in Texas (Rylander 1981, Tweit 1996), but neither vocal mimicry nor night-time song has been previously documented for the western subspecies (T. c. palmeri; Bent 1948, Tweit 1996). The only reference found for possible night-time singing by T. curvirostre was by Oberholser (1974; 654), “it seldom .sings at night,” also referenced by Tweit (1996) but with no specific documentation or reference to source by either publication. OBSERVATIONS The western Curved-billed Thrasher has a varied and melodious repertoire, but mimicry of another species was recorded only recently. A Curved-billed Thrasher mimicked a migrant Black-headed Grosbeak {Pheucticus melanoce- phalus) on a desert plot —23 km east of Tucson, ' John.son & Haight Environmental Con.sultants, 3755 South Hunters Run, Tucson, AZ 85730, USA. "Corresponding author; e-mail: rroyloi.s@aol.com Arizona. The Black-headed Grosbeak was singing near the top of a large ornamental eucalyptus {Eucalyptus spp.) tree on 23 April 2009 at —0830 hrs MST. Another apparent Black-headed Gros- beak was answering from a patch of Sonoran Desert scrub —35 m away. Singing during spring migration is not unusual for Black-headed Gros- beak but countersinging by the species in desert lowlands is unusual. The grosbeak singing in the eucalyptus tree behaved in typical fashion, commonly giving one or more chuck notes between songs, but the bird in desertscrub was not interspersing songs with chuck notes. On closer examination with binoculars, the second bird was observed to be a Curve-billed Thrasher near the top of a foothill palo verde {Parkinsonia microphylla) tree singing a song that was audibly indistinguishable from that of the grosbeak. T. c. palmeri has sung both spontaneously and elicited night-time song on this same desert plot. A Curve-billed Thrasher began spontaneous song on 21 February 2005 at 2212 hrs (sunset 1815 hrs). It sang two refrains of several notes, the refrains separated by -15 sec. The thrasher at site A was answered by a short song from a second thrasher -150 m distance (site B). Thrasher A sang another short song and was answered by a third thrasher (site C) -150 m in a different direction at 2213 hrs. Thrashers singing at sites A, B, and C roughly formed an equilateral triangle with spacing of -150 m between thrashers. This appioximates the spacing between T. curvirostre males engaged in diurnal temtorial song in this area during the breeding season. The moon was bright on the night of 21 February (full moon 24 Feb) but an -5 min shower had earlier produced a trace of rain at —2145 hrs. A thrasher at site A again sang spontaneously on the following night, 22 Febru- ary, and was answered by a thrasher at site B between 2127 and 2128 hrs. The sky was .similar to the previous night with .scattered clouds but the moon was bright and visible most of the time. 626 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 3. September 2010 DISCUSSION Diurnal song at this time of year is common for this early nesting species for which breeding readiness peaks in early February (Vleck 1993, Tweit 1996). The Curved-billed Thrasher was not among the 53 listed avian species known to respond to playback tapes by 1980 (Johnson et al. 1981). After not hearing spontaneous singing during the nights following 22 February, a Curved-billed Thrasher song tape was played (Colver et al. 1999) on 27 February. Following five playings of the tape, beginning at 2229 hrs with hesitation between each, a thrasher at site A responded. It sang a refrain of ~six notes at 2233 hrs. A similar song again came from site A at 2234 hrs. after playing the tape two more times. Despite further playing of the tape there was no additional response. A similar response from a thrasher at site A resulted when a tape was played on 28 February, beginning at 2325 hrs. Answers of a half-dozen notes were sung again at 2328 hrs and at 2329 hrs. No further attempts were made to elicit songs. Neither behavior discussed here has been recorded previously for this species. Because of its close relationship to other mimids that exhibit mimicry and/or nocturnal song, it would be of phylogenetic and general ornithological interest to devise a project, using playback tapes, to further investigate these phenomenon among the subspe- cies of T. curx’irostre. ACKNOWLEDGMENTS We thank R. C. Tweit, C. E. Braun, and two unknown reviewers for critiques of an early manuscript. LITERATURE CITED Bent, A. C. 1948. Life histories of North American nuthatches, wrens, thrashers, and their allies. U.S. National Museum Bulletin Number 195. Colver, K. J., D. Stokes, and L. Stokes. 1999. Stokes field guide to bird songs: western region. Time Warner Audio Books, New York, USA. Derrickson K. C. and R. Breitwisch. 1992. Northern Mockingbird (Mimus polyglottos). The birds of North America. Number 7. Johnson, R. R., B. T. Brown, L. T. Haight, and J. M. Simpson. 1981. Playback recordings as a special avian censusing technique. Studies in Avian Biology 6:68- 75. Oberholser, H. C. 1974. The bird life of Texas. University of Texas Press, Austin, USA. Rylander, M. K. 1981. Vocal mimcry in the Curve-billed Thrasher. Bulletin of the Texas Ornithological Society 14: 29-30. Tweit, R. C. 1996. Curve-billed Thrasher {Toxostoma cun’irostre). The birds of North America. Number 235. Vleck, C. M. 1993. Hormones, reproduction, and behavior in birds of the Sonoran Desert. Pages 73-86 in Avian endocrinology (P. J. Sharp, Editor). Journal of Endocrinology Ltd., Bristol, United Kingdom. The Wilson Journal of Oniilhology 122(3):627-633, 2010 Ornithological Literature Robert B. Payne, Review Editor HANDBOOK OF THE BIRDS OF THE WORLD. Volume 14. Bush shrikes to Old World span-ows. Edited by J. del Hoyo, A. Elliott, and D. Christie, 2009. Lynx Edicions, Barcelona, Spain. 2009: 893 pages, 51 color plates, many photo- graphs and maps. ISBN: 978-84-96553-50-7. 212 Euros (—$310.00 US) (cloth). — The volume continues the series on birds of the world with general accounts on all families and species of birds. This volume includes the Malaconotidae (bushshrikes), Prionopidae (helmetshrikes), Van- gidae (vangas), Dicruridae (drongos), Callaeidae (New Zealand wattlebirds), Notiomystidae (Stitchbird), Grallinidae (mudlarks), Struthideidae (formerly Corcoracidae, Australian mudnesters), Artamidae (woodswallows), Cracticidae (butcher- birds), Pityriaseidae (Bristlehead), Ptilonorhyn- chidae (bowerbirds), Paradisaeidae (birds-of-para- dise). Corvidae (crows), Buphagidae (oxpeckers), Stumidae (starlings), and Passeridae (Old World sparrows). The text includes a summary of recent molecular studies on systematic relationships, descriptions of each species and subspecies, their behavior and ecology, a list of references, and an index to the common and scientific names. The text was written by ornithologists who have conducted fieldwork on these birds and are familiar with them. All species are well illustrated in the color plates and many are shown in great supporting photographs. The family accounts describe and compare the highlights of the species accounts, which include taxonomy, subspecies and distribution including a map, habitat, food and feeding, breeding, movements, survival, status and conservation concern, and a bibliography. The length of the species accounts varies with what is known about the birds and with the style of the authors. The Madagascar vangas are amazingly diverse in morphology and most species are not well known. Biological information on these endan- gered birds might help in conservation plans; but the continuing loss of habitat in Madagascar, even in sites with nominal protection, is the main challenge to their survival. Cooperative breeding is widespread in butcher- birds, woodswallows, and Australian mudnesters. and the text describes the Rowleys’ previously unpublished study of Australian Magpie {Gym- norhina tibicen). Courtship displays of bowerbirds and birds-of-paradise are well described and illustrated in photographs. The extreme display structures and decorations of bowerbirds include the maze-like four-walled bower of Yellow- breasted Bowerbird (Chlaniydera lauterbachi) and the ornamental feathers of King-of-Saxony Bird-of-Paradise (Pteridophora alberti) in the bower of Archbold’s Bowerbird (Archboldia papuensis). Bowerbird young remain in the nest for 17-21 days, except Archbold’s Bowerbird which live in cold and wet high-altitude habitats and remain nestlings for 30 days. The 42 species of birds-of-paradise are fascinating in their sexual dimorphism and elaborate displays. None has become extinct since 1600, but three are globally threatened and others are declining in numbers in the wild, especially due to hunting. Conservation efforts have had success in protecting certain populations from hunting and habitat loss. Cap- tive-breeding programs now are feasible and are recommended while the wild populations are still healthy. The crows are the most species-rich group in Volume 14 with 24 genera, 123 species, and 359 taxa. The section on this family could stand alone as a readable book. Corvids have more than their share of cooperatively-breeding species; the best known is Elorida Scrub-Jay {Aphelocoma coer- idecens). Some corvids have a complex layer of social organization with the pair, the family group, and the flock (Red-billed Chough [Pvr- rhocorax pyrrhocorax]). Resident singles or pairs of Common Raven (Con’iis corox) chase single intruders from carcasses but, when the intruders recruit other vagrants to join them, the group can overcome the aggressiveness of the resident territorial birds. Ravens, American Crow (C. brachyrhynchos), and Rook (C. fnigUegus) each have more than 20 distinct calls of known function. The geographic variation in calls of Large-billed Crow (C. macrorhynchos) requires further study. Some corvids including the Com- mon Raven are good vocal mimics of other species in captivity and others are mimics in the 627 628 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122. No. 3, September 2010 wild (Eurasian Jay [Garnilus glandariiis])\ the function of mimicry in wild corvids remains a mystery. Common Ravens associate with gray wolf (Canis lupus) in Yellowstone National Park, USA and the scavenging corvids get early access to the carcass at wolf kills. Food-storing and retrieval is common in several corvids; one, Clark’s Nutcracker (Nucifraga columbiana), bur- ies pine (Pinus spp.) seeds in places that lose their snow cover at different times in spring allowing nutcrackers to have a steady supply of seeds through a season. Some crows are common and widespread, but the Hawaiian Crow (Cor\ms hawaiiensis) is now extinct in the wild; new birds from a captive breeding group may be re- introduced in the future, although birds reared and released in the 1990s soon died from predation and disease. Starlings have a wide range of breeding organizations, some are cooperative breeders, some are densely colonial, and some are polyg- ynous, all much like the unrelated New World blackbirds (Icteridae). Some Passeridae sparrows (here treated as 40 species), are highly colonial, as up to 20 nests of Spanish Sparrow {Passer hispaniolensis) have been built in a single nest of White Stork (Ciconia ciconia), and colonies of this sparrow have as many as 800,000 pairs. This volume includes a Foreword that discusses the influence of bird watchers on the interest in and concern for birds. “Birding past, present and future — a global view” compares the roles of great men and social change in times past, advances in field guides and the web as information resources, advances in optics, cam- eras, and recording devices, and the availability of local and overseas tours to see birds. The demographics and economic impact of birders have been surveyed mainly in the U.S. and U.K. In the U.S., 20 million people consider themselves birders, getting into the field at least 50 times a year. The economics of birding may amount to US$31.7 billion, counting vehicles, fuel, guides, food and lodging including ecotours, plus $1.9 billion on optics and $2.6 billion on nest boxes, feeders and bird food. Birds benefit from birders in projects for conservation, and a concern for responsible ecotourism includes training local guides and support of birding places. Birding is promoted as exciting and intellectually challeng- ing, and a passion for the natural world that is thought to benefit our physical and emotional health and wellbeing. The volumes of the Handbook of the Birds of the World are written clearly and well illustrated, and they are the highest standard source of information on birds. When Volume 16 is published in late 201 1, the series will be the first work to describe in words and illustrations all species of birds in the world. — ROBERT B. PAYNE, Professor Emeritus, University of Michigan, 1306 Granger Avenue, Ann Arbor, MI 48104, USA; e-mail: rbpayne@umich.edu GROUSE OF PLAINS AND MOUNTAINS: THE SOUTH DAKOTA STORY. By Lester D. Flake, John W. Connelly, Thomas R. Kirschen- mann and Andrew J. Lindbloom. South Dakota Department of Game, Fish and Parks, Pierre, South Dakota, USA. 2010: xv -i- 246 pages, and ~173 color photographs and 27 graphs. ISBN: 9780615350158. $15.00 (Soft cover). — This lav- ishly illustrated book with color photographs and graphs is intended for the general public but is prepared to also appeal to biologists, scientists, managers, and anyone interested in grouse in the prairies of North America. The authors are well known for their interest in grouse and several have impressive research and teaching credentials. This is a book about four species of grouse native to South Dakota: Sharp-tailed Grouse {Tympanuchus phasianellus). Greater Prairie-Chicken (T. cu- pido). Greater Sage-Grouse {Centrocercus uro- phasianus), and Ruffed Grouse {Bonasa umbel- lus). The latter is the least expected as one does not associate aspen {Populus spp.) forests with the grasslands and shrublands of the Great Plains. Greater Sage-Grouse have the smallest distribu- tion at present in South Dakota followed by Ruffed Grouse, Greater Prairie-Chicken, and Sharp-tailed Grouse with the most extensive range across the state. All four species are hunted with modest (sage-grouse) to expansive seasons (the other 3 species). This book is based on science and accumulated field experience. It has 13 chapters, six appendi- ces, and an extensive Literature Cited. The pre.sentation is pleasant and easily understood but the information is supported by citation of the scientific literature. The chapters are: 8outh Dakota’s Grouse Species, Physical Characteris- tics, Behavior, Mobility and Habitat, Nutritional Needs and Food Habits, Nesting Ecology, Broods, Survival and Mortality, Population Monitoring, Hunting Seasons, Grouse Hunting, Habitat Loss, OR N ITHOLOG ICA L LITER A I'U R E 629 and Further Thoughts. Each chapter is heavily illustrated (only color photographs and graphs are used), covers each of the species, and ends with a Review (= summary). Each chapter is interesting and well prepared and there is something for everyone. I was interested in all of the material but especially the material (pages 3-4) on Dusky Grouse (Demimgapiis obscurus), which may have been present historically, although poorly conducted transplants failed to establish a population after releases in 1969 and 1974. Greater Sage-Grouse are at the extreme eastern edge of their current (only 3 counties in South Dakota) and historical range in South Dakota. One wonders how this species survives in areas with low densities of big sagebrush {Artemisia tridentata subspp.). Silver sagebrush {A. cana) obviously will support sage- grouse; the presence of robust, taller grasses with more cover is likely important in supporting sage- grouse at the eastern and northern peripheries of their distribution. Ruffed Grouse are also an anomaly as they appear to have low dispersal capability but somehow found their way to the isolated Black Hills (most likely from the west). Management of public lands (USDA, Eorest Service) in the Black Hills for ponderosa pine { Finns ponderosa) does not benefit Ruffed Grouse, which has more affinity for species of aspen and other deciduous shrubs. Relatively little is known about Ruffed Grouse in South Dakota, which the authors acknowledge but try to fill the data gaps with published information from scientific studies in other states. The authors are candid about the role of hunting and are cautious about possible population effects (with literature citations) on grouse. There are two chapters (10 and II) that discuss data that can be obtained from hunter-harvested birds, as well as the joy of being afield with hunting dogs in the fall. They describe how data are collected (Chapter 9) and provide actual data on counts in spring as well as harvest estimates. They are quick to note the inadequacies of some surveys but do not discuss how state wildlife agencies place priorities on how funds are spent (obviously on species where hunter interest is high and where grouse populations are large). The photographs amply demonstrate that grouse also have value for bird watchers, especially in spring during breeding seasons. Scattered throughout the text and focused in the final chapter (13), the authors discuss some of the habitat issues involved with grouse populations. These include everything from West Nile virus (pond development for livestock and energy production which may enhance mo.squito distri- bution) to livestock grazing, federal land conser- vation programs, prairie and forest management, fences and roads to wind turbine farms. These issues are not restricted to South Dakota but occur throughout the Great Plains and the Intermountain West. I was hard pressed to find factual errors and any editorial glitches, but did find Dalke italicized on page 61, and a missing parenthesis on page 73. Use of Jargon such as ‘input’ is relatively minor throughout. With rare exceptions (i.e., page 208, hectares and acres), the authors do not use metric terms and we read about pounds and inches and miles. Several references in the Literature Cited (T. D. I. Beck and J. L. Beck, M. M. McDonald and J. E. McDonald) are not alphabetical. A nicer problem is that there may be too many similar photographs. One weakness is there is no Index. The stated goal of this book (Eoreword) is “to foster increased interest and appreciation for South Dakota’s four grouse species and the habitats that support them.’’ The book clearly meets this goal as it is very attractive, the material is accurate and current, it is well written and easy to read, and the color photographs keep inviting you to turn the next page. This book is a must for everyone interested in grouse and the northern Great Plains, and I highly recommend it. — CLAIT E. BRAUN, Grouse Inc., 5572 North Ventana Vista Road, Tucson, AZ 85750, USA; e-mail: sg-wtp(5)j uno.com FARMLAND BIRDS ACROSS THE WORLD. Edited by Wouter van der Weijden, Paul Terwan, and Adriaan Guldemond. Lynx Edicions, Barce- lona, Spain. 2010: 144 pages and 500 -t- color photographs, maps, and figures. ISBN: 13-978- 84-96553-63-7. $39.00 (hardcover). — Books that aim to highlight conservation issues can .some- times be downers that bring to mind Aldo Leopold’s famous quote: “Being an ecologist is living in a world of wounds.” Others can tend to act as generally uniriendly encyclopedias of inlormation, useful only to con.servation planners. Farmland Birds Across the World is a unique creation. It highlights the importance of agricul- tural habitats to global birdlife in a way that is accessible, informative, and engaging to the casual reader, ecologist, and conservationist alike. 630 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 3, September 2010 It does so using more than 500 stunning large format photographs of birds and the agricultural landscapes they inhabit that invite the reader on a journey around the globe. It is authored by seven experts in biology and links the beauty and natural history of birdlife to the challenging and often complex conservation issues facing birds in agricultural landscapes. One unique way they convey these stories is by peppering the chapters with separate boxed-off panels highlighting spe- cific conservation issues (e.g., declines in Asian vultures due to veterinary drugs), conservation- related natural history stories (e.g., the journey of the Black-tailed Godwit [Limosa lirnosa]) and interesting on-the-ground conservation efforts (e.g., novel planting schemes that benefit sky- larks). This makes the book easy to pick up, even for just a moment. The introduction is a short overview that highlights the complex, mutual relationships between farming and birds that can be both beneficial and negative. It highlights interesting paradoxes. For example, agriculture and its intensification represent perhaps the most impor- tant threat to bird populations globally, yet agricultural lands also represent the sole remain- ing habitat for many critically endangered species. Chapter 2 effectively begins the body of this book and represents an overview of issues related to farming such as factors affecting birds in crops, resources that crops can provide to birds, the ecosystem services provided by birds to crops, crop damage by birds, and how farming can adversely impact birds. The next six chapters focus on birdlife in specific agricultural environments: grasslands, cereal and other annual crops, rice fields, orchards and plantations, coffee and cacao, and farmyards. The first page of each chapter contains a global map of the distribution of each agricultural crop and some brief statistics impressively illustrating the global nature and extent of agriculture. For example, cereals are the world’s most important calorific crop, occupying approximately one-third of all arable land. The authors highlight how changes in the area and cultivation intensity of crops, the birdlife exploiting the.se crops, and the threats to bird populations vary by region. Notably, they draw on massive global bird and agricultural data bases to create figures illustrating the worldwide trends in human activity shaping bird populations. Each chapter ends with another unique feature, a bulleted .section outlining different conservation challenges and opportunities associated with issues like global climate change, agricultural policy, farming practices and, at times, a surpris- ing and conspicuous lack of information. For example, the authors (page 67) mention, “Al- though China is the world’s leading rice producer, little is known on birdlife in its rice fields than the occurrence of herons and egrets.’’ The final chapter discusses the future prospects for farmland birds, summarizing some of the grand conservation challenges and highlighting interesting and unique solutions to those prob- lems. Collectively, this volume manages to convey the authors’ excitement at the richness of farmland birdlife, their confusion due to the paradoxes associated with bird conservation and agricultural production, their fear at the often overwhelming threats facing farmland bird popu- lations, and finally relief there are potential solutions to improve the fortunes of farmland birds. My only significant complaint about this volume was the limited treatment given to the potential importance of biofuel production on avian biodiversity. The authors do give significant mention of the widespread impacts that oil-palm {Elaeis spp.) is having on tropical birds. However, second generation cellulosic biomass crops that may include cereals, trees and woody plants, and other crops have potential to radically reshape agricultural landscapes worldwide. To be fair, little is known about the ecological effects of many of these crops, but I felt the threat of biofuels to farmland birds should have made lists of conservation threats. It is estimated that over one-third of the earth’s surface has now been converted to agricultural uses. This makes the book a particularly timely resource. A notable asset of this volume is that it does a seamless job of integrating scientific ideas into the chapters in a way that is accessible to non-scientists. It is this feature that makes the book so broadly appealing, yet useful even to con.servation scientists. Biologists often think of themselves as storytellers and the authors of Farmland Birds Across the World have succeeded in telling fascinating and important stories that instill a .sense of global responsibility for the birds that are impacted by the growing of our food, fiber, and fuel. — BRUCE A. ROBERTSON, Research Associate, Great Lakes Bioenergy Research Center, Kellogg Biological Station, Michigan State University, 3700 East Gull ORNITHOLOGICAL LITERATURE 631 Lake Drive, Hickory Corners, Ml 49060, USA; e-mail: brucerobertson@hotmail.eom BIRDS OF EUROPE, SECOND EDITION. By Lars Svensson. Princeton Field Guides, Princeton University Press, Princeton, New Jersey, USA. 2009: 448 pages, 3,500+ color illustrations. Illustrations and captions by Killian Mullarney and Dan Zetterstrom. ISBN: 978-0-691-14392-7. $29.95 (paperback). — The text covers all 772 species ot birds occuiTing in Europe as well as 32 introduced species or variants and 118 vagrant species. In the United Kingdom, where it was published by Harper/Collins, it is known as the "Collins Bird Guide." The book is a substantial revision of the author’s first edition (1999). The book includes all bird species that breed or regularly occur in Europe, Africa north of 30° N, the Sinai Peninsula, the Middle East through Turkey, Israel, Jordan, Syria, Georgia, and Azerbaijan, and the Canary Islands and Madeira. Europe is bordered on the east by the Ural Mountains, Ural River, and Caspian Sea. The text and distribution maps are on one page and the illustrations are on the facing page. For each species the text has a description of shapes, male and female, young and old, breeding and nonbreeding, and geographic variation where appropriate. The text is clear, concise, and accessible, and generally more detailed and comprehensive than in North American field guides. For example, in one of the longer species accounts, the description of Rough-legged Hawk/ Rough-legged Buzzard (Biiteo lagopus) has 441 words, each well chosen. Key identification features are highlighted in italics, variations in plumage within a population are shown and described, and differences between similar species are noted; e.g., the wing pattern in flight is helpful in identification of larks. Habitat, range, and breeding status are described with notes on frequent hybridization. Many species have useful descriptions of feeding behavior and breeding biology. The Middle Spotted Woodpecker (Den- drocopos medium), unlike other related woodpeck- ers, calls but does not drum in territorial assertion. Voice is described in words, often with cadence- and pitch-perfect transliteration: the species- typical call of Rock Ptarmigan {Lagopus niuta) is “here comes the bride’’. Species status (breeding, resident/migratory, abundance) is noted in short form for Britain and Ireland. Each family has an informative account with comments such as the night vision of owls is not “surreal’’ but requires light from the moon or stars. Other useful sections are included on seawatching (and how to point out birds on a featureless sea.scape to other watchers), raptors in flight, and general advice for identification of waders/shorebirds and gulls. The quality of the illustrations is very good, the birds lifelike, the details precise, the selection comprehensive. Many birds are illustrated in more than one posture and in flight. The color plates include in small type many useful comments on identification. Because more than one standard taxonomic source is currently used in ornithology, the species recognized in the book reasonably follow “author’s preference’’ in recent revisions. For example, in the Heiring Gull (Larus argen- tatus) species complex, where 2 lA pages of color plates have 67 images, several species are recognized that previously were known as sub- species; these are American Herring Gull (L. smithsonianus). Yellow-legged Gull (L. luicha- hellis), Armenian Gull (L. armenicus), Audoin’s Gull (L. audouinii), and Caspian Gull (L. cachinnans). Some distinctive forms recognized as subspecies receive detailed accounts as well; e.g., “Steppe Buzzard’’ {Buteo buteo vulpinus). The owls are particularly well illustrated with new images for this edition, and the comments on these birds (“stately look’’, “deceptively gentle look’’, “grim look’’, “astonished look’’, “aus- tere look’’) capture the impression so the reader can identify the live owl in the field. Wallcreeper (Tichodroma murario) is described as “The Hoopoe \Upupa epops] of the rock face!” Hawfinch {Coccothraustes coccothraustes) has a big head; “the Hawfinch is a flying pair of nut- crackers!’’ A few plumages are shown that do not appear in North American field guides, e.g.. summer plumage of female Long-tailed Duck (Clangula hyemolis), and some compari.sons of North American species go beyond what is said in North American field guides, e.g., adult Buff- breasted Sandpiper (Tryngites suhruficolli.s) adults are similar to (and distinguishable from) iuvenile Ruff (Philomachus pugna.x). Most North Ameri- can Cotharus thrushes are illustrated with spotted Juvenile wing coverts, perhaps the appropriate plumage if most vagrants that appear in Europe are birds in their first autumn, but the text should indicate the age in these birds, as it does for first- winter New World warblers. Two pages on the 632 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 3. September 2010 pochards (Aythya spp.) show hybrids and the most similar non-hybrids with notes for identification. Vagrant species are illustrated along with similar resident or breeding species, or are in the back of the book, and accidentals are listed but not described. Introduced species are in the main section of the book (e.g., Canada Goose [Branta canadensis}) or in the back of the book (e.g., the estrildid finches with established populations). The book is an excellent field guide, small enough to use in the field, a clear and informative read on birds, and clearly the best book for identification of the birds of Europe. It also will be useful in North America especially in identi- fication of northern Palearctic vagrants in Alaska and elsewhere. It is highly recommended. — ROBERT B. PAYNE, Professor Emeritus, University of Michigan, 1306 Granger Avenue, Ann Arbor. MI 48104, USA; e-mail: rbpayne(s> umich.edu CONSERVATION BIOLOGY OE HAWAI- IAN EOREST BIRDS. Edited by Thane K. Pratt, Carter T. Atkinson, Paul C. Banko, James D. Jacobi, and Bethany L. Woodworth. Yale Uni- versity Press, New Haven, Connecticut, USA. 2009: 707 pages, 97 black and white plus 32 color illustrations. ISBN: 978-0-300-14108-5. $85.00 (cloth). — Hawaiian forest birds are famous among evolutionary biologists for their spectacular bill morphology and bright plumage. They are equally famous among conservation biologists for extinc- tion, declines, and numerous threats. This book touches on the former but focuses on the latter. It appears to largely summarize the USGS, Biolog- ical Research Discipline efforts from a recent grant. This may explain why they ignore much other research on Hawaiian birds. The book is organized into five parts. Part 1 is introductory and has three chapters that deal with origins, historic decline and extinction, and forest birds and Hawaiian culture. Part 2 on status, biology, and limiting factors has 1 I chapters on monitoring, trends, habitat loss and degradation, food exploitation, life history and demography, malaria and pox virus, genetics, mammalian predators. Hawaiian raptors, introduced birds, and epidemiology of malaria and pox. Part 3 is about applying research to management, including controlling invasive species, land.scape ecology, managing disease, controlling small mammals, and captive propagation. Part 4 deals with recovery programs for four species, one of which is likely extinct, a second is close behind, a third is rapidly declining even though it has been extensively studied for more than 20 years, and the fourth is a success story. Part 5 deals with the future, and covers the social and political obstacles and asks the question “Can Hawaiian forest birds be saved?” Space limits chapters in the book that can be reviewed. The origins and evolution chapter by Thane Pratt is masterful historical biogeography, both in understanding where Hawaiian birds came from as well as initial conditions they likely encoun- tered. We understand why these are the main vertebrates of Hawaiian forests, and a non-random set of potential colonists. They competed among themselves, which led to adaptive radiation, but did not compete with amphibians, reptiles, mammals, or social insects. There are two excellent chapters on disease. One deals with modeling the epidemiology of avian malaria and pox; the other concerns managing disease. The former is a particularly clear exposition of compartmentalized modeling for understanding transmission dynamics in a vector-transmitted disease such as malaria. The chapter on managing disease considers diverse techniques for controlling mosquitoes and pigs (Sus scrofa) to break transmission, along with immunizing the birds themselves and genetically modifying mosquitoes. Unfortunately, these chap- ters, as well as the one on genetics and conservation, ignored our prize-winning and widely cited 2005 paper (Condor 107: 753-764) documenting an increase in avian malaria in highest elevation forest where pigs were managed. This is a clear example that mosquitoes docu- mented in the forest may have originated from unmanaged lands, a point emphasized by the book. No specific adaptations were identified in the book, despite intensive studies that have been performed by several research groups. Banko and Banko, in their chapter on evolution and ecology of food exploitation, deal with comparative bill morphology as adaptation for particular diets, but they ignore work documenting that the liwi (Vestiaria coccinea) has undergone a niche shift with accompanying changes in bill morphology. Woodworth and Pratt, in their chapter on life history and demography, ignore nestling over- growth and seasonal variation in sex allocation in the Hawaii Akepa (Loxops coccineus coccineiis). ORNITHOLOGICAL LITERATURE 633 clear adaptations tor breeding during a seasonal decline in food. These have been dismantled by competition with introduced Japanese White-eyes (Zosterops japonicus), documented in Evolution- ary Ecology Research 10: 931-965 (2008) ( = EER 2008), referenced and dismissed in the book. The Banko and Banko chapter on food essentially ignored arthropod prey in Metrosi- deros polymorpha trees for insectivorous birds. This is the most important tree in wet to mesic Hawaiian forests, and EER 2008 shows that most birds use foraging substrates associated with the tree. They refer to an old dissertation, not the papers that document the diversity of prey with seasonal pattern of variation. A reader might think that Scotorythrci caterpillers are the most impor- tant food resource for all bird species. Like other islands, Hawaii has a preponderance of spiders. The chapters in the book on monitoring forest birds and identifying trends are thorough but misleading. The authors fail to recognize analysis of residuals as essential when modeling changes in populations over time. A regression model requires residual analysis to ensure the model fits the data. If it does not, the model needs to be corrected. The trends reported for Hakalau Eorest National Wildlife Refuge are incon'ect, as shown in our paper in the May 2010 issue of Condor (112: 213-221), because their residuals indicate significant lack of fit. The endangered Hawaii Akepa and the liwi, a species of concern, are significantly declining on the refuge. An example in the book that clearly indicates the lack of fit is Eig. 25.6 in the chapter on saving Hawaiian birds. The endangered Hawaii Creeeper (Oreoinystis mcmci) has a 21 -year trend that looks positive, but residuals are negative in the early and late years of the series, and positive in the middle, indicating lack of lit. The creeper is declining, not increasing, in models with an additional param- eter with appropriate residuals. The chapter by Jeff Eoster on introduced birds provides a marvelous historical and ecological perspective. Unfortunately, he dismisses EER 2008 because of the trend analysis. He argues that food limitation can only be established by measuring food. However, ornithologists have long recog- nized that food limitation can be inferred by stunted growth of nestlings, fault bars in wing and tail feathers that reflect nutritional stress during molt, lower fat, and changes in begging patterns. Woodworth and Pratt in the life history and demography chapter have a section on human- caused changes in demographic rates, but only report on nesting success. We (EER 2008), document significant change in juvenile survival and adult survival in the Hawaii Akepa associated with increase in numbers of introduced Japanese White-eyes. Chewing lice (Phthiraptera) are not even mentioned in the book although they contribute to parasite-mediated competition. Most chapters in this book have to be read critically, because they ignore important literature and accept misleading trend analysis. The answer to “Can Hawaiian forest birds be saved?” is “No” unless the immediate threat is recognized and managed. That threat is an introduced bird, which has stunted growth and lowered survival throughout the community (2009. Cun-ent Biolo- gy 19: 1736-1740). Malaria will kill remaining birds that are food limited because they will be unable to mount energetically expensive immune responses.— LEONARD A. FREED, Depart- ment of Zoology, University of Hawai’i, Honolulu, HI 96822, USA; e-mail: Ifreedt® hawaii.edu THE WILSON JOURNAL OF ORNITHOLOGY Editor CLAIT E. BRAUN 5572 North Ventana Vista Road Tucson, AZ 85750-7204 E-mail: TWILSONJO@comcast.net Editorial NANCY J. K. BRAUN Assistant Editorial Board RICHARD C. BANKS JACK CLINTON EITNIEAR SARA J. OYLER-McCANCE LESLIE A. ROBB Review Editor ROBERT B. PAYNE 1306 Granger Avenue Ann Arbor, MI 48104, USA E-mail: rbpayne@umich.edu GUIDELINES FOR AUTHORS Please consult the detailed "Guidelines for Authors” found on the Wilson Ornithological Society web site (http://www.wilsonsociety.org). All manuscript submissions and revisions should be sent to Clait E. Braun, Editor, The Wilson Journal of Ornithology, 5572 North Ventana Vista Road, Tucson, AZ 85750-7204, USA. The Wilson Journal of Ornithology office and fax telephone number is (520) 529-0365. The e-mail address is TWilsonJO@comcast.net NOTICE OF CHANGE OF ADDRESS Notify the Society immediately if your address changes. Send your complete new address to Ornithological Societies of North America, 5400 Bosque Boulevard, Suite 680, Waco, TX 76710. The permanent mailing address of the Wilson Ornithological Society is: %The Museum of Zoology, The University of Michigan, Ann Arbor, MI 48109, USA. Persons having business with any of the officers may address them at their various addresses given on the inside of the front cover, and all matters pertaining to the journal should be sent directly to the Editor. MEMBERSHIP INQUIRIES Membership inquiries should be sent to Timothy J. O'Connell, Department of Natural Resources Ecology and Management, Oklahoma State University, 240 AG Hall, Stillwater, OK 74078; e-mail: oconnet@ okstate.edu 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 ornithological magazines from members and friends of the Society. Two members have generously established a fund for the purchase of new books; members and friends are invited to maintain the fund by regular contribution. The fund is administered by the Library Committee. Jerome A. Jackson, Florida Gulf Coast Univeristy, is Chairman of the Committee. The Library currently receives over 200 periodicals as gifts and in exchange for The Wilson Journal of Ornithologv. For information on the Library and our holdings, see the Society’s web page at http:// www.wilsonsociety.org. 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, Mu.seum of Zoology, The University of Michigan, 1109 Geddes Avenue. Ann Arbor, Ml 48109-1079. USA. Contributions to the New Book Fund should be sent to the Treasurer. This issue of The Wilson Journal of Ornithology was published on I September 2010. Continued from outside back cover 569 Annual bird mortality in the bitumen tailings ponds in northeastern Alberta, Canada Kevin P. Timoney and Robert A. Ronconi 577 Abundance and distribution of waterbirds in the Llanos of Venezuela Francisco J. Vilella, Mark S. Gregory, and Guy A. Baldassarre Short Communications 578 Piping Plover foraging distribution and prey abundance in the pre-laying period Jonathan B. Cohen and James D. Fraser 583 Interspecific song imitation by a Cerulean Warbler Than J. Boves, David A. Buehler, and Phillip C. Massey 587 Nests, nest placement, and eggs of three Philippine endemic birds Luis A. Sdnchez-Gonzdlez, Carl Oliveros, Nevong Puna, and Robert G. Moyle 592 Nests, eggs, and young of the Azure-crowned Hummingbird {Amazilia cyanocephala) Juan Francisco Ornelas 597 Nest description of the Garden Emerald {Chlorostilbon assimilis) from Costa Rica Luis Sandoval and Ignacio Escalante 600 The nest of the Cipo Canastero {Asthenes luizae), an endemic furnariid from the Espinha90 Range, southeastern Brazil Henrique Belfort Gomes and Marcos Rodrigues 604 Nest box use by Great Tits in semi-arid rural residential gardens Motti Charter, Yossi Leshem, Shay Halevi, and Ido Izhaki 608 Analysis of nest sites of the Resplendent Quetzal {Pharomachrus mocinno)\ relationship between nest and snag heights Dennis G. Siegfried, Daniel S. Linville, and David Hide 612 Parasitism of a Blue- winged Teal nest by a Northern Shoveler in South Dakota. Thomas E. Lewis and Pamela R. Garrettson 614 Nest predators of ground-nesting birds in montane forest of the Santa Catalina Mountains, Arizona Chris Kirkpatrick and Courtney J. Conway 618 Feeding a foreign chick; a case of a mixed brood of two Tit species Toshitaka N. Suzuki and Yuko Tsuchiya 620 Infanticide by an Eastern Phoebe Gary Ritchison and Brandon T. Ritchison 623 Marsh Wren eats small fish Andrea J. Ayers and James W. Armacost Jr. 625 Occasional mimicry and night-time singing by the western Curve-billed Thrasher ( Toxostoma curvirostre palmer!) R. Roy Johnson and Lois T Haight 627 Ornithological Literature Robert B. Payne, Book Review Editor The Wilson Journal of Ornithology (formerly The Wilson Bulletin) Volume 122, Number 3 CONTENTS September 2010 Major Articles 417 Reproductive success and nestling growth of the Baywing parasitized by Screaming and Shiny cowbirds Maria C. De Mdrsico, Bettina Mahler, and Juan C. Reboreda 432 Bobolink egg mass variability and nestling growth patterns Barbara Frei, David M. Bird, and Rodger D. Titman 439 Breeding phenology and nesting success of the Yucatan Wren in the Yucatan Peninsula, Mexico Jesus Vargas-Soriano, Javier Salgado Ortiz, and Griselda Escalona Segura 447 Breeding biology and natural history of the Slate-throated Whitestart in Venezuela Roman A. Ruggera and Thomas E. Martin — ^ 455 Reproductive biology of a grassland songbird community in northcentral Montana Stephanie L. Jones, J. Scott Dieni, and Paula J. Gouse - j 465 Mountain biking trail use affects reproductive success of nesting Golden-cheeked Warblers 3 Craig A. Davis, David M. Leslie Jr., W. David Walter, and Allen E. Graber 1 475 Survival, site fidelity, and population trends of American Kestrels wintering in southwestern Florida J Daniel M. Hinnebusch, Jean-Frangois Therrien, Marc-Andre Valiquette, Bob Robertson, Sue Robertson, and 4 Keith L. Bildstein 484 Temporal and spatial shifts in habitat use by Black Brant immediately following flightless molt Tyler L. Lewis, Paul L. Flint, Joel A. Schmutz, and Dirk V Derksen 494 Arthropod foraging by a southeastern Arizona hummingbird guild Donald R. Powers, Jessamyn A. Van Hook, Elizabeth A. Sandlin, and Todd J. McWhorter 503 Army ant raid attendance and bivouac-checking behavior by neotropical montane forest birds Sean ODonnell, Anjali Kumar, and Gorina Logan 513 Maroon-fronted Parrot {Rhynchopsitta terrisi) breeding home range and habitat selection in the northern Sierra Madre Oriental, Mexico Sonia Gabriela Ortiz-Maciel, Consuelo Hori-Ochoa, and Ernesto Enkerlin-Hoejlich 518 Short-term effects of fire on breeding birds in southern Appalachian upland forests Nathan A. Klaus, Scott A. Rush, Tim S. Keyes, Johyi Petrick, and Robert J. Cooper 532 Avian communities of the Altamaha River estuary in Georgia, USA Ross A. Brittain, Vicky J. Meretsky, and Chris B. Craft 545 Influence of cover and food resource variation on post-breeding bird use of timber harvests with residual canopy trees Molly E. McDermott and Petra Bohall Wood 556 Distribution, abundance, and status of Cuban Sandhill Cranes {Grus canadensis nesiotes) Xiomara Galvez Aguilera and Felipe Chavez-Ramirez 563 The feather structure of Dippers; water repellency and resistance to water penetration Arie M. Rijke and William A. Jesser Continued on inside hack cover Wl/- Wilson Journal of Or ni tholog^ Volume 122y Number 4, December 2010 Published by the Wilson Ornithological Society THE WILSON ORNITHOLOGICAL SOCIETY FOUNDED 3 DECEMBER 1888 Named after ALEXANDER WILSON, the first American ornithologist. President — E. Dale Kennedy, Biology Department, Albion College, Albion, MI 49224, USA; e-mail: dkennedy@aIbion.edu First Vice-President— Robert C. Beason, P. O. Box 737, Sandusky, OH 44871, USA; e-mail; Robert.C.Beason@gmaiI.com Second Vice-President — Robert L. Curry, Department of Biology, Villanova University, 800 Lancaster Avenue, Villanova, PA 19085, USA; e-mail: robert.curry@viIIanova.edu Editor — Clait E. Braun, 5572 North Ventana Vista Road, Tucson, AZ 85750, USA; e-mail: TWILSONJO@ comcast.net Secretary — John A. Smallwood, Department of Biology and Molecular Biology, Montclair State University, Montclair, NJ 07043, USA; e-mail: smallwoodj@montclair.edu Treasurer — Melinda M. Clark, 52684 Highland Drive, South Bend, IN 46635, USA; e-mail: MClark@tcservices.biz Elected Council Members — Jameson F. Chace, Sara R. Morris, and Margaret A. Voss (terms expire 2011); Mary Bomberger Brown, Mary Garvin, and Mark S. Woodrey (terms expire 2012); Mark Deutschlander, Paul G. Rodewald, and Rebecca J. Safran (terms expire 2013). Living Past-Presidents — Pershing B. Hofslund, Douglas A. James, Jerome A. Jackson, Clait E. Braun, Mary H. Clench, Richard C. Banks, Richard N. Conner, Keith L. Bildstein, Edward H. Burtt Jr., John C. Kricher, William E. Davis Jr., Charles R. Blem, Doris J. Watt, and James D. Rising. Membership dues per calendar year are: Regular, $40.00; Student, $20.00; Family, $50.00; Sustaining, $100.00; Life memberships, $1,000.00 (payable in four installments). The Wilson Journal of Ornithology is sent to all members not in arrears for dues. THE WILSON JOURNAL OF ORNITHOLOGY (formerly The Wilson Bulletin) THE WILSON JOURNAL OF ORNITHOLOGY (ISSN 1559-4491) is published quarterly in March, June, September, and December by the Wilson Ornithological Society, 810 East 10th 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 OSNA, 5400 Bosque Boulevard, Suite 680, Waco, TX 76710. All articles and communications for publication should be addressed to the Editor. Exchanges should be addressed to The Josselyn Van Tyne Memorial Library, Museum of Zoology, Ann Arbor, MI 48109, USA. Subscriptions, changes of address, and claims for undelivered copies should be sent to OSNA, 5400 Bosque Boulevard, Suite 680, Waco, TX 76710, USA. Phone: (254) 399-9636; e-mail: business(gosnabirds.org. Back issues or single copies are available for $ 1 2.00 each. Most back issues of the journal are available and may be ordered from OSNA. Special prices will be quoted for quantity orders. All issues of the journal published before 2000 are accessible on a free web site at the University of New Mexico library (http://elibrary.unm.edu/sora/). The site is fully searchable, and full-text reproductions of all papers (including illustrations) are available as either PDF or DjVu files. © Copyright 2010 by the Wilson Ornithological Society Printed by Allen Press Inc., Lawrence, KS 66044, USA. COVER: Wilson’s ?\ovev {Characlrius wilsonia). Illustration by Don Radovich. @ This paper meets the requirements of ANSI/NISO Z39.48-1992 (Permanence of Paper). MCZ library DEC 0 9 2010 HARVARD UNIVERSITY FRONTISPIECE. Henslow's Sparrows (Ammodramus henslowii) arc tallgrass prairie specialists in North America. Analysis of yearly resettlement and abundance patterns from Breeding Bird Survey data indicated this species is less likely to return to previously used breeding areas than other grassland sparrows. Restricting analyses to single observer-collected data resulted in significant effects not detected in data aggregated from multiple observers during the same study period. Acrylic painting by Marco Antonio Pineda M of Puebla, Mexico. Wilson Journal of Ornithology Piihlished by the Wilson Ornithological Society VOL. 122, NO. 4 December 2010 PAGES 635-858 The Wilson Journal of Ornithology 122(4):635-645, 2010 BREEDING PATTERNS OE HENSLOW’S SPARROW AND SYMPATRIC GRASSLAND SPARROW SPECIES L. LYNNETTE DORNAK' ABSTRACT. Henslow’s Sparrows (Ammodramiis henslowii) are reported to have irregular patterns of return to breeding areas. I present data supporting these reports at rangewide extents, while testing potential biases inherent in the North American Breeding Bird Survey data. Two measures of population variability were used to show that Henslow's Sparrows are less likely to use breeding areas predictably and consistently, but have similar variance in numbers at occupied sites relative to other sympatric grassland sparrow species. 1 illustrate how restricting analyses to single-observer- collected Breeding Bird Survey data results in subtle but significant effects not detected in data aggregated from multiple observers through the study period. The most conservative analysis (single-observer, restricted distribution) showed that Henslow s Sparrows exhibited lower prevalence of occurrence than Grasshopper (Ammodramiis savannarum) (P < 0.001 ) and Savannah (Passercuhis sandwichensis) (P < 0.001) sparrows but no difference in variation of abundance (P > 0.05). These results suggest Henslow’s Sparrows are not returning to previously used breeding habitat from year-to-year. Grassland management should consider the behavior documented in this study and attempt to incorporate this facet of Henslow's Spairow biology into decisions that involve broad-scale land.scape de.sign. Received 27 Janiiaiy 2010. Accepted 8 May 2010. Artificial grazing regimes, drainage of wet- lands, large-scale agriculture, and alteration of natural fire regimes have changed North Ameri- can landscapes, and left behind only relict tracts of native prairie (Hyde 1939; Knopf 1988, 1994). The full extent of these impacts on native flora and fauna has only recently begun to be appreciated. Shifts are likely occumng continent-wide, but the most dramatically impacted habitat has been native grasslands. Since European settlement, 99.9% of native prairies have been lost (Samson and Knopf 1994) or altered (Vickery et al. 1994) in North America, and they are now termed a critically endangered habitat (Noss et al. 1995). Thirteen bird species within this biome have incuired ' Department of Geography, University of Kansas, Lawrence, KS 66045, USA; e-mail: picoide.s@ku.edu serious declines, far suipassing those of any other North American biome (Knopf 1994, Peterjohn and Sauer 1999, Robbins et al. 2002, Powell 2006), probably the result of continued conversion of prairie habitat to an artificial landscape. Henslow’s {Amwodramu.s henslowii). Grass- hopper [A. savanna rum), and Savannah (Pas.ter- ciiliis sandwichensi.s) sparrows are obligate grassland nesters (Vickery et al. 1994) with population declines across part or all of their breeding ranges (Peterjohn and Sauer 1999, Wells and Rosenberg 1999). None has national threatened or endangered species status, but Henslow’s and Grasshopper sparrows have been listed as Birds of Conservation Concern (USDl 2008) and Species ot Continental Importance (Rich et al. 2004). However, the species’ behavior and habitat use, as it relates to landscape configuration, must be 635 636 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 4, December 2010 understood more fully before successful manage- ment can be achieved. Hyde (1939:23) was the first to note the unpredictability of Henslow’s Sparrows appear- ance in breeding areas, writing “its presence in a given season cannot be certainly predicted,” a sentiment that has been repeated by other authors (Wiens 1969, Skipper 1998, Ingold et al. 2009). Most studies report low (Skipper 1998, Monroe and Ritchison 2005) or nonexistent site fidelity for Henslow’s Sparrows in breeding areas (J. L. Zimmerman cited in Pruitt 1996). This behavior is atypical compared to Grasshopper and Savannah sparrows which exhibit higher nest site fidelity (Bedard and LaPointe 1984, Wheelright and Rising 1993). These studies have used site fidelity as a measure of return rates to breeding areas in con- •secutive years, and have necessarily concentrated at only a few sites. Thus, a broader-scale assessment of annual resettlement patterns in breeding areas is necessary, but is lacking in the literature. Henslow’s Sparrows may be also erratic and opportunistic in selection of sites for breeding on a broad spatial scale. Previous research on this species has focused at local and regional scales with a few exceptions (e.g., Herkert 2007). The only studies that used landscape-level analysis were limited to sections of the species’ range (Bajema and Lima 2001, Cunningham and Johnson 2006, Thogmartin et al. 2006), and little information exists to provide a range-wide perspective for this species. 1 examined abundance data from the entire breeding range for Henslow’s Sparrows, and compared it to similar data for Grasshopper and Savannah sparrows. The focus was not to compare habitats used by each species; instead, 1 analyzed patterns of resettlement in habitat the birds already had assessed as suitable for nesting. Grasshopper and Savannah sparrows were chosen for comparison, not because of similarity or dissimilarity of habitat choices, but because they are obligate grassland species that nest within the breeding range of Henslow’s Sparrow. 1 u.sed North American Breeding Bird Survey (BBS) data (Sauer et al. 2007) to address two objectives. First, I examined how Henslow’s Sparrow populations vary in time and space. A secondary objective was to assess the impact of using different subsets of BBS data because these data have been criticized for inherent biases. This information is critical to understanding this species’ natural history, as well as conservation implications that may be derived. 1,700 FIG. I . Henslow’s (A), Grasshopper (B), and Savannah (C) sparrow distributions from Breeding Bird Survey route data collected during 2000-2007. Henslow’s, Grasshopper, and Savannah sparrows were observed on 1 32; 1, 387; and 1,765 routes, respectively. METHODS Study Area. — The study area included the entire breeding ranges of Henslow’s, Grasshopper, and Savannah sparrows (Fig. 1). All bird abundance data were derived from the BBS, a monitoring Donutk • POPULATION DYNAMICS OF HENSLOW’S SPARROWS 637 system created in the 1960s with the goal of understanding long-term trends in North Ameri- can breeding bird populations (Sauer et al. 2007). BBS data are collected annually, on fair-weather June mornings, on 4,100 standardized roadside census routes across the United States, Canada and, most recently, Mexico. Sampling points are spaced evenly along the survey route (Robbins et al. 1986), which is 39.4 km in length; 50 sampling points are located every 0.8 km along the route. Observers record all birds seen or heard during 3 min at each stop. Data are available for all years between 1966 and the present, but I used data from 2000-2007 to assure maximal route density and consistency. Count Data. — I used two measures to describe the magnitude of yearly fluctuations of resettle- ment across ranges of these species: prevalence of occurrence (proportion of years present) and variation in abundance. The former describes how consistently a species returns to a given BBS route year after year, and is calculated as the ratio of the number of years in which a species was detected on a route to the number of times during the study period the route was surveyed. The latter measures year-to-year variation in number of individuals of a species at the site and is estimated as the coefficient of variation: the standard deviation data divided by the mean abundance for each route across all years in which the route was surveyed. Routes sampled in only 1 year of the study period were excluded from analyses because the coefficient of variation was undefined. These two variables describe the consistency of occur- rence and abundance, but saturation effects of variation in abundance may influence my results. The BBS data were divided and analyzed in four groups to consider potentially inherent biases (geographic and observer). (1) BBS routes were considered across the entire breeding range of each species. (2) Analysis was constrained to BBS routes within the recorded breeding range of Henslow’s Sparrows. 1 also examined BBS route data collected by single observers across all years in the study period, removing possible biases originating from differences in observer consis- tency. These analyses were conducted (3) across the entire range of each species, and (4) only within the distributional area of Henslow’s Sparrow. The smallest data set had 56 routes and, to have equal samples size for each analysis, I randomly selected 50 routes per species for each data set. Statistical Analysis. — Shapiro-Wilk normality tests and Levene’s test for homogeneity of variances were performed on all data sets. At least one sample within each data set had non- normally distributed data and unequal variances. Both negative and positive skews were common within data sets, and transformation of the data was not possible. Thus, I used Kruskal-Wallis nonparametric analysis of variance to test for differences among means, which provided con- servative and consistent testing for differences among all groups. Kruskal-Wallis test statistics were evaluated with ot = 0.05. Mann-Whitney (7-tests were performed to differentiate between significant groups whenever significant x~ values were obtained. I used the non-parametric Ken- dall’s tau statistic to analyze the relationship between prevalence of occurrence and variation in abundance. A Bonferroni correction was applied to Mann-Whitney (7-tests and Kendall’s tau con'elations, and all effects were reported at a = 0.0167. Statistical analyses were performed using SPSS, Version 16.0 (2007). All maps were created in ArcGIS, Version 9.2 (ESRI 2009). RESULTS Prevalence of Occurrence. — Significant differ- ences were found among species in all data sets tested for prevalence of occuirence (P < 0.001). There were significant differences between spe- cies for the multiple-observer data set, both at the full extent of the species’ ranges and within the distribution of Henslow’s SpaiTOws (Table 1). Mann-Whitney (7-tests for the full extent of the species’ ranges indicated the median prevalence of occuiTence for Henslow’s SpaiTows (0.50) was lower than for Grasshopper (0.78, P = 0.006), and Savannah ( 1.00, P < 0.001) spaiTows; Grasshop- per Span'ows, did not differ significantly in prevalence from Savannah Sparrows (P = 0.022, Table 2). Results constrained to within the distributional area of Henslow’s Sparrow were similar, and significant for all comparisons (Tables 1, 2; Fig. 2). There were significant among-group differenc- es in the single-observer data sets for both the full- range extent and within the distributional area of Henslow’s Sparrow (P < 0.001, Table 1). Median Henslow’s Sparrow prevalence (0.50) for full- range extents was significantly lower than for Grasshopper (0.88, P < 0.001), and Savannah (1.00, P < 0.001) sparrows (Table 2). Grasshop- per Sparrows had lower prevalence than Savannah 638 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4. December 2010 TABLE 1. Comparisons (Kruskal-Wallis tests) of Henslow's. Grasshopper, and Savannah sparrows between multiple- and single-observer groups, and between data collected across the full extent of each species’ breeding range versus restricted to Henslow's Sparrow’s distributional area. Breeding Bird Survey data were collected from 2000 to 2007 on 50 randomly selected routes. (*) indicates P < 0.05. Full extent of species’ ranges Henslow’s Sparrow distributional area •Multiple observers Single obsen'er Multiple observers Single observer H if P « df P HUP H di P Prevalence of occurrence 26.9 2 <0.001* 39.9 2 <0.001* 45.4 2 <0.001* 45.1 2 <0.001* Variation in abundance 9.0 2 0.01 1* 12.6 2 0.002* 2.7 2 0.26 8.1 2 0.018* Sparrows {P — 0.005, Table 2). The same results occurred when these data were restricted to the distributional area of Henslow’s Sparrow (Tables 1, 2; Fig. 3). Variation in Abundance. — Results were less consistent between data sets than for prevalence comparisons. There were differences between species (P = 0.011, Table 1) for the multiple- observer data set across the full extent of the species’ ranges. Grasshopper Sparrows had higher variability in abundance (median = 0.54) than Henslow’s Sparrows (0.42, P = 0.004), but no sig- nificant differences existed between Henslow’s and Savannah (0.50, P — 0.070) sparrows, or between Savannah and Grasshopper spaiTows (P = 0.18, Table 2). No differences were apparent between species (Table 1, Fig. 2) when the data were limited to the distributional area of Henslow’s Span'ow. The single-observer data sets revealed a similar pattern, but with significant differences between species in analysis of the full extent (P — 0.002) and within the distribution of Henslow’s Span'ow (P = 0.018, Table 1). There was no difference between Henslow’s SpaiTows and Grasshopper (0.44, P = 0.090) or Savannah {P = 0.21, Table 2) sparrows. Grasshopper Sparrows, however, did (median = 0.47) have higher variability than Savannah Sparrows (0.39, P < 0.001). The same results were obtained for analyses limited to single-observer routes within the distribution of Henslow’s SpaiTow (Tables 1, 2; Fig. 3). Correlation of Prevalence and Variation. — Prevalence of occuiTence was significantly and positively correlated to variation in abundance for Henslow’s Span'ows in the multiple-observer (i = 0.37, P < 0.001) and single-observer (x = 0.32, P = 0.003) data sets. This relationship did not hold for either Grasshopper or Savannah spaiTOws in any data set (all P > 0.05). TABLE 2. Comparisons (Mann- Whitney U-tests) of Henslow’s (HESP), Grasshopper (GRSP), and Savannah (SASP) sparrows between multiple- and single-observer groups, and between data collected across the full extent of each species’ breeding range and restricted to Henslow’s Spaixow’s distributional area. Breeding Bird Survey data were collected from 2000 to 2007 on 50 randomly selected routes. One comparison for variation in abundance is lacking because of nonsignificant results (Kruskal-Wallis analysis). (*) indicates P < 0.0167. r-'ull e.xtent of .species' ranges Henslow’s .Sparrow distributional area Multiple observers Single ob.server Multiple observers Single observer U z p u p u : P u p Prevalence of occurrence HE.SP X GRSP 856.5 -2.7 0.006* 649.5 -4.2 <0.001* 582.0 -4.7 <0.001* 625.5 -4.4 <0.001* HESP X SASP 505.0 -5.2 <().()() 1* 422.5 -5.9 <0.001* 364.5 -6.3 <0.001* 347.0 -6.4 <0.001* GRSP X SASP 938.0 -2.3 0.022 879.0 -2.8 0.005* 998.5 - 1 .9 0.045 933.5 -2.5 0.014* Variation in abundance HESP X GRSP 8.30.5 -2.9 ().()()4* 1,003.5 -1.7 0.09 1,152.5 -0.7 0.50 HESP X SASP 987.5 - 1 .8 ().()7() 1,068.0 -1.3 0.21 1 .()09.() -1.7 0.10 GRSP X SASP 1 .056.5 - 1 .3 0.18 697.0 -3.8 <0.001* 815.0 -3.0 0.003* Dornuk • POPULATION DYNAMICS OF IIENSLOW’S SPARROWS 639 Coefficient of variation of abundance 0.000-0.125 0.125-0.321 0.321 -0.502 0.502-0.707 0.707- 1.404 Prevalence of occurrence o 0.250-0.333 O 0.333-0.500 O 0.500-0.714 Q 0.714-0.875 Q 0.875- 1.000 N FIG. 2. Henslow's (A), Gra.sshopper (B), and Savannah (C) sparrow distributions showing prevalence (circle size) and abundance (shading of circles). Data are restricted to within the Henslow's Sparrow's breeding distribution and were collected by multiple observers from 2000 to 2007 on 50 randomly selected Breeding Bird Survey routes. 640 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4. December 2010 Coefficient of variation of abundance 0.000-0.125 0.125-0.322 0.322-0.514 0.514-0.774 0.774- 1.404 Prevalence of occurrence O 0.250-0.333 O 0.333-0.500 O 0.500-0.714 O 0.714-0.875 O 0.875- 1.000 N FIG. 3. Hen.slow’s (A), Grasshopper (B), and Savannah (C) sparrow distributions showing prevalence (circle size) and abundance (shading of circles). Data are restricted to within the Henslow’s Sparrow’s breeding distribution but were collected by single observers only for given routes from 2()()() to 2007 on 50 randomly selected Breeding Bird Survey routes. Doniak • POPULATION DYNAMICS OF HENSLOW’S SPARROWS 641 DISCUSSION Resettlement. — Site fidelity describes the like- lihood ot an individual’s return to a particular site from year to year. Often, the decision for adult birds to return to a particular site is based on the individual’s experiences, or the experiences of conspecitics, at the site in previous years (Hilden 1965). This measurement is ideal for finer-scale analyses, but broad-scale studies require consid- eration of resettlement of an area by groups of individuals. More importantly, global patterns exhibited by these groups across space and time can provide insight into differences across their geographic ranges. Previous studies have evaluated nest site fidelity in Henslow’s, Grasshopper, and Savannah sparrows (Bedard and La Point 1984, Skipper 1998, Jones et al. 2007), although, to my knowledge, this study is the first to examine variation in resettlement patterns across their breeding ranges. Henslow’s Sparrows, in all analyses, had the lowest prevalence of occun'ence among the three species. They were, as a group, less likely to return to a BBS route from I year to the next. Henslow’s Sparrows when detected, with the exception of one test, had variability in abundance not distinguishable from those of Grasshopper and Savannah sparrows. Both Grass- hopper and Savannah sparrows had greater numbers of high-prevalence routes than Hen- slow’s Sparrows. These across-species differences in variation in occurrence might be best explained by habitat choices and social behavior of each .species. Henslow’s Sparrows require large, open grass- lands (Hyde 1939, Smith 1968, Thogmartin et al. 2006). Vegetation structure, including aspects of vegetation height (Wiens 1969, Herkert 1994a) and density (Wiens 1969, Zimmerman 1988), is an important factor for nest site selection by Hen- slow’s Sparrows (Graber 1968, Bajema and Lima 2001, Powell 2006). Fields must remain relatively undisturbed for several consecutive years to achieve vegetation structure typical of Henslow’s Sparrow’s breeding habitat. Thus, this sparrow avoids fields frequently disturbed by haying (Graber 1968, Cully and Michaels 2000), burning, or grazing on frequent rotations (Bollinger 1995). Similarly, Grasshopper and Savannah spaiTows prefer natural habitat to managed landscapes (Owens and Myers 1973, Dale et al. 1997). These species are less strict in their habitat preferences. unlike Henslow’s Sparrows, and take advantage of a greater variety of habitats, including grazed, cultivated (Owens and Myers 1973), and hay fields (Graber 1968, Dale et al. 1997). These differences in habitat preferences among the species may explain differences in prevalence of occurrence. The narrowness of breeding habitat preferences of Henslow’s Sparrow within the dynamic grassland biome possibly motivates groups of individuals to seek unused habitat when previously settled locations are no longer suitable (Reinking et al. 2000). For example, an area that is optimal habitat in 1 year may experience a disturbance, such as a late summer burn, resulting in unsuitable habitat for subsequent breeding seasons, thereby encouraging Henslow’s Spar- rows to seek new nesting locations. Settling suitable habitat when encountered upon arrival in the breeding area, rather than homing to a previously used site and then relocating, would also be advantageous for species that use highly variable resources (Johnson and Grier 1988). Grasshopper and Savannah span'ows use a wider variety of habitats (Smith 1968, Wiens 1969, Owens and Meyers 1973, Dale et al. 1997), and may resettle an area in successive breeding seasons, even though the habitat has been altered from previous years. The ability to be opportu- nistic and flexible when choosing a nesting location may be advantageous for species whose nesting habitat is unpredictable and patchy (Wiens 1973, Cody 1985, Johnson and Grier 1988). Henslow’s Span'ows, in the relatively few areas where they occurred consistently, had higher variation in abundance. They had less variation in abundance in areas where they were not consistently pre,sent creating a positive association between prevalence and variability. This pattern might be explained by the effect of clustering, combined with opportunistic settling. The earliest literature on Henslow’s Sparrows noted the tendency of the species to nest in loose colonies, especially when occupying large patches of habitat (Hyde 1939, Graber 1968, Wiens 1969). Individuals of nomadic species, such as Hen- slow’s Sparrows, arriving to an area may use the presence of conspecifics which have already .settled there as a means to evaluate habitat quality, especially when knowledge of nest success from previous years within that nesting location is lacking (Bollinger and Gavin 1989, Ahlering et al. 2006). Thus, clustering when paired with opportunistic location selection may 642 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4. December 2010 produce a pattern of iiTegularity of settlement with regularity of abundance. Unlike Henslow’s Spar- rows, no significant relationship exists between variability, prevalence, and abundance for either Grasshopper or Savannah sparrows. Biases in BBS Data. — Use of BBS data is a timely and cost-effective way of answering relatively short-term, broad-scale questions (Sauer et al. 2005, Winter et al. 2006). The BBS is the only source for range-wide, standardized data in North America; it is commonly used to estimate abundances and year-to-year fluctuations across species’ breeding ranges (Bibby et al. 2000, Diefenbach et al. 2003, Herkert 2007). Although widely used, the BBS is not without potential bias. One criticism of the BBS is that of change of observers across time. Link and Sauer (1998) suggested that differences in ability among observers may influence trends detected along routes over time. I found results differed when data were limited to single-observer routes. Specifically, Savannah Sparrows had significantly higher prevalence rates than Grasshopper Sparrows in the single-observer data set, a difference not detected in analyses of the multiple-observer data set. The difference was significant in both the full extent of the species’ ranges and in the area; thus, reduced to the distributional area of Henslow’s Sparrow, the effect was not a result of geographic extent. Removing routes surveyed by multiple observers also affected variation in abundance. There were contrasting differences between species across the full extents of the species’ ranges. Grasshopper Sparrows had the highest variability, but the ranks of the other two species shifted. The multiple-observer data set in ana- ly.ses constrained to within the distribution of Henslow’s Sparrow showed no species differences, but the single-observer data set revealed signifi- cant differences. Thus, results of several analyses changed when route data collected by multiple observers were removed. This variation is likely a result of change in observers and their relative abilities (Link and Sauer 1998). Sauer et al. (1994) suggested abun- dance patterns are best represented by analyses of single-observer data only. However, even single observers may, through time, improve in identifi- cation and detection skills, lor example by learning a song (Link and Sauer 1998) or the opposite (e.g., with declining hearing abilities); these elfects cannot be measured, but may be reduced by comparisons with multiple-observer data. Detection. — Other concerns not addressed with- in the analyses of this paper should be considered. It is possible that Henslow’s SpaiTows are not detected as easily as Grasshopper or Savannah sparrows. The species may be too rare in some areas, resulting in detection difficulties (Wells and Rosenberg 1999, DeVault et al. 2002). In other locations where individuals are more abundant, lack of activity (Diefenbach et al. 2007, Confer et al. 2008), or failure to detect individuals may cause discrepancies, thereby affecting count data. Henslow’s SpaiTow’s cryptic appearance and secretive behavior (Hyde 1939), in addition to its insect-like song (Leftwich and Ritchison 2000), may lead to detection problems during surveys. Detection may be further restricted when singing declines after mate pairing (Leftwich and Ritch- ison 2000) or as a result of Henslow’s Sparrows’ proclivity for nocturnal singing (Walk et al. 2000). Henslow’s SpaiTows may not nest near roadsides as a response to woody vegetation (Patten et al. 2006) or traffic volume (Forman et al. 2002), yet BBS data are collected entirely from roadside routes. It is possible that populations of birds are not surveyed accurately because of species’ behavior and surveying techniques. Bajema et al. (2001) suggest applying a correction factor to estimate Henslow’s Span'ow abundance to counter these detection problems. CONSERVATION IMPLICATIONS Henslow’s SpaiTow was once thought to be prevalent throughout the western extent of its breeding range, although it now occupies less than 1% of this original area (Robbins et al. 2002). Its need for regular, although infrequent, habitat disturbance to maintain a particular serai stage may facilitate this decline in present human- dominated landscapes (Pruitt 1996). This habitat in the pre-European landscape was maintained by fire (natural and artificial), and grazing by large herbivores, creating a mosaic of habitats on the landscape (Knopf 1994, Umbanhowar 1996). Burning, grazing, and mowing have all been recommended as suitable management practices (Pruitt 1996), but the timing, extent, and frequency of these disturbances are critical factors that affect suitability of the habitat for breeding Henslow’s Sparrows (Herkert 1994a, Powell 2006). The patterns documented in this study are strongly suggestive of a species that has adapted to be able to track optimal sites in a shifting habitat mosaic. Henslow’s Sparrows are also considered area- Dornak • POPULATION DYNAMICS OF IIENSLOW'S SPARROWS 643 sensitive and affected by patch size of breeding habitat (Herkert 1994b, Walk and Warner 1999, Oleary and Nyberg 2000, Thogmartin et al. 2006). The CLiiTent approach of conserving large isolated patches of high-quality habitat may be too narrowly focused on this concept. Renfrew and Ribic (2008), in a study of Bobolinks (Dolichony.x on’zivonis), and Grasshopper and Savannah spairows, conclud- ed that patch-size might only be particularly restrictive when the entire landscape is heavily fragmented. Thus, conserving smaller areas in addition to large, contiguous habitat patches within the landscape matrix would create more total grassland habitat and better suit Henslow’s Spar- rows, as it enhances the overall quality of the landscape and fosters its nomadic movements and colonial behavior (Hora and Koford 2006, Ribic et al. 2009). Acquiring smaller, but functional, plots may be easier and more economical, thus producing more immediate conservation impacts. Ultimately, losses of suitable grassland nesting habitat on regional scales may result in extinction of obligate grassland species (Vickery et al. 1994), and particularly Henslow’s Sparrows, which appear to move much more broadly across regional land- scapes than the other two species analyzed. ACKNOWLEDGMENTS I could not have accomplished this project without the aid of A. T. Peterson. 1 express my gratitude for his support with project design and manuscript refinements. I also thank S. L. Egbert for professional and academic guidance, and Y. J. Nakazawa, L. L. Rausch, and Xingong Li for valuable technical support. Data were made available by the Breeding Bird Survey. Financial support was provided by the Department of Geography and by the Experimental Program to Stimulate Competitive Research (EPSCoR). LITERATURE CITED Ahlering, M. A., D, H. Johnson, and J. Faaborg. 2006. Conspecific attraction in a grassland bird, the Baird’s Sparrow. Journal of Field Ornithology 77:36.6-371, Bajema, R. a. and S. L. Lima. 2001. Landscape-level analysis of Henslow’s Sparrow (Ammodramus heiislo- wii) abundance in reclaimed coal mine grasslands. American Midland Naturalist 145:288-298. Bajema. R. A., T. L. DeVault, P. E. Scott, and S, L. Lima. 2001. Reclaimed coal mine grasslands and their significance for Henslow’s Sparrows in the American Midwest. Auk 118:422-431. Bedard, J. and G. LaPointe. 1984. Banding returns, arrival times, and site fidelity in the Savannah Sparrow. Wilson Bulletin 96:196-205. Bibby, C. j., N. D. Burgess, D. A. Hill, and S. H. Hustoe. 2000. Bird census techniques. Academic Press, London, United Kingdom. Bollinger, E. K. 1995. Successional changes and habitat selection in hayfield bird communities. Auk 1 12:720-730. Bollinger, E. K. and T, A. Gavin. 1989. Succe.ssional changes and habitat selection in hayfield bird com- munities. Auk 106:584-594. Cody, M. L. 1985. Habitat selection in open-country birds. Pages 191-226 in Habitat selection in birds (M. L. Cody, Editor). Academic Press, Orlando, Florida. USA. Confer, J. L., R. E. Serrell, M. Hager, and E. Lahr. 2008. Field tests of the Rosenberg-Blancher method for converting point counts to abundance estimates. Auk 125:932-938. Cully Jr., J. F. and H. L. Michaels. 2000. Henslow’s Sparrow habitat associations on Kansas tallgrass prairie. Wilson Bulletin 112:115-123. Cunningham, M. A. and D. H. Johnson. 2006. Proximate and landscape factors influence grassland bird distri- butions. Ecological Applications 16:1062-1075. Dale, B. C., P. A. Martin, and P. S. Taylor. 1997. Effects of hay management on grassland songbirds in Saskatchewan. Wildlife Society Bulletin 25:616-626. DeVault, T. L., P. E. Scott, R. A, Bajema, and S. L. Lima. 2002. Breeding bird communities of reclaimed coal-mine grasslands in the American Midwest. Journal of Field Ornithology 73:268-275. Diefenbach, D. R., D. Brauning, and j. a. Mattice. 2003. Variability in grassland bird counts related to observer differences and species detection rates. Auk 120:1168-1179. Diefenbach, D. R., M. R. Marshall, J. A. Mattice, and D. Brauning. 2007. Incorporating availability for detection in estimates of bird abundance. Auk 124:96- 106. Environmental Research Systems (ESRI). 2009. Arc- Map, Version 9.2. Environmental Research Systems Inc., Redlands, California, USA. Forman, R. T. T., B. Reineking, and A. M. Hersperger. 2002. Road traffic and nearby grassland gird patterns in a suburbanizing landscape. Environmental Manaae- ment 29:782-800. Graber, j. W. 1968. Passerherhidus heuslowii hendslowii (Audubon): Western Henslow’s Sparrow. Pages 779- 788 in Life histories of North American cardinals, grosbeaks, buntings, towhees, finches, sparrows, and allies (A. C. Bent, Editor). Smithsonian Institute, U.S. National Museum Bulletin 237. Herkert, J. R. 1994a. Status and habitat selection of the Henslow’s Sparrow in Illinois. Wilson Bulletin I06:35-J5. Herkert, J. R. 1994b. The effects of habitat fragmentation on midwestern grassland bird communities. Ecological Applications 4:461^71. Herkert, J. R. 2007. Conservation Reserve Program benefits on Henslow’s Sparrows within the United States. Journal of Wildlife Management 7 1 -2750- 2751. Hidden, O. 1965. Habitat selection in birds. Annales Zoologici Fennici 52:53-75. Horn, D. J. and R. R. Koford. 2006. Could the area- sensitivity of some grassland birds be affected by 644 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4, December 2010 landscape composition? Proceedings of the North American Prairie Conference 19:109-116. Hyde, A. S. 1939. The life history of Henslow’s Sparrow, Passerherbiilus henslowii (Audubon). University of Michigan, Ann Arbor, Museum of Zoology Miscella- neous Publications 41:1-79. iNGOLD, D. J., J. L. Dooley, and N. Lavender. 2009. Return rates of breeding Henslow’s Sparrows on mowed versus unmowed areas on a reclaimed surface mine. Wilson Journal of Ornithology 121:194-197. Johnson, D. H. and J. W. Grier. 1988. Determinants of breeding distributions of ducks. Wildlife Monographs 100. Jones, S. L., J. S. Dieni, M. T. Green, and P. J. Gouse. 2007. Annual return rates of breeding grassland songbirds. Wilson Journal of Ornithology 119:89-94. Knopf, F. L. 1988. Conservation of steppe birds in North America. Ecology and Conservation of Grassland Birds 7:27-H. Knopf, F. L. 1994. Avian assemblages on altered grasslands. Studies in Avian Biology 15:247-257. Leftwich, C. and G. Ritchison. 2000. Singing behavior of male Henslow’s Sparrows (Ammodmmus henslowii). Bird Behavior 18:1-7. Link, W. A. and J. R. Sauer. 1998. Estimating population change from count data: application to the North American Breeding Bird Survey. Ecological Applica- tions 8:258-268. Monroe. M. S. and G. Ritchison. 2005. Breeding biology of Henslow’s Sparrows on reclaimed coal mine grasslands in Kentucky. Journal of Field Ornithology 76:143-149. Noss, R. F., E. T. Large, and J. M. Scott. 1995. Endangered ecosystems of the United States: a preliminary assessment of loss and degradation. USDl, National Biological Service 06II-R-0I (MR). Wash- ington D.C., USA. O’Leary, C. H. and D. W. Nyberg. 2000. Treelines between fields reduce the density of grassland birds. Natural Areas Journal 20:243-249. OwEN.S, R. A. AND M. T. Myers. 1973. Effects of agriculture upon populations of native passerine birds of an Alberta fescue grassland. Canadian Journal of Zoology 51:697-713. Paiten, M. a., E. Shochat, D. L. Reinking, D. H. Wolfe, AND S. K. Sherrod. 2006. Habitat edge, land management, and rates of brood parasitism in tallgrass prairie. Ecological Applications 16:687-695. Peterjohn, B. G. and j. R. Sauer. 1999. Population status of North American grassland birds from the North American Breeding Bird Survey. 1966-1996. Studies in Avian Biology 19:27^4. Powell, A. F. L. A. 2006. Effects of prescribed burns and bison (Bo.s bison) grazing on breeding bird abundances in tallgrass prairie. Auk 123:183-197. Pruht. L. 1996. Henslow's Sparrow: status assessment. USDl. Fish and Wildlife Service. Bloomington, Indiana. USA. Reinking. D. L.. D. A. Wiedenfeld, D. H. Wolfe, and R. W. Rohrbaugh Jr. 2()()(). Distribution, habitat use. and nesting success of Henslow’s Sparrow in Okla- homa. Prairie Naturalist 32:219-232. Renfrew, R. B. and C. A. Ribic. 2008. Multi-scale models of grassland passerine abundance in a fragmented system in Wisconsin. Landscape Ecology 23:181-193. RtBic, C. A., R. R. Koford, j. R. Herkert, D. H. Johnson, N. D. Niemuth, D. E. Naugle, K. K. Barker, D. W. Sample, and R. B. Renfrew. 2009. Area sensitivity in North American grassland birds: patterns and process- es. Auk 126:233-244. Rich, T. D., C. J. Beardmore, H. Berlanga, P. J. Blancher, M. S. W. Bradstreet, G. S. Butcher, D. W. Demarest, E. H. Dunn, W. C. Hunter, E. E. Inigo-Elias, J. a. Kennedy, A. M. Martell, A. O. Panjabi, D. N. Pashley, K. V. Rosenberg, C. M. Rustay, J. S. Wendt, and T. C. Will. 2004. Partners in Flight North American landbird conservation plan. Laboratory of Ornithology, Cornell University, Ithaca, New York, USA. Robbins, C. S., D. Bystrak, P. Henslow, and P. H. Geissler. 1986. The Breeding Bird Survey: its first fifteen years, 1965-1979. USDl, Fish and Wildlife Service, Washington, D.C., USA. Robbins, M. B., A. T. Peterson, and M. A. Ortega- Huerta. 2002. Major negative impacts of early intensive cattle stocking on tallgrass prairies: the case of the Greater Prairie-Chicken {Tyinpaniichiis ciipido). North American Birds 56:239-244. Samson, F. and F. L. Knopf. 1994. Prairie conservation in North America. BioScience 44:418^21. Sauer, J. R., B. G. Peterjohn, and W. A. Link. 1994. Observer differences in the North American Breeding Bird Survey. Auk 1 1 1:50-62. Sauer, J. R., J. E. Hines, and J. Fallon. 2007. The North American Breeding Bird Survey, results and analysis 1966-2006. Version 10.13.2007. USGS, Patuxent Wildlife Research Center, Laurel, Maryland, USA. Sauer, J. R., J. Casey, H. Laskowski, J. D. Taylor, and F. Jane. 2005. Use of survey data to define regional and local priorities for management on national wildlife refuges. USDA, Forest Service. General Technical Report PSW-GTR-191. Pacific Southwest Re.search Station, Albany, California, USA, Skipper, C. S. 1998. Henslow’s Sparrows return to previous nest site in western Maryland. North American Bird Bander 23:36^1 . Smith, W. P. 1968. Passerherbiilus henslowii susiirrans (Brewster): Eastern Henslow’s Sparrow. Pages 776- 778 in Life histories of North American cardinals, grosbeaks, buntings, towhees, finches, sparrows and allies (A. C. Bent. Editor). Smith.sonian Institute, U.S. National Museum Bulletin 237. SPSS Institute Inc. 2007. SPSS for Windows, Version 16.0. SPSS Institute Inc., Chicago, Illinois. USA. Thogmartin. W. E., M. G. Knut.son, and j. R. Sauer. 2006. Predicting regional abundance of rare grassland birds with a hierarchical spatial count model. Condor 108:25^6. Umbanhowar Jr., C. E. 1996. Recent fire history of the northern Great Plains. American Midland Naturalist 135:1 15-121. Donuik • POPULATION DYNAMICS OF HENSLOW’S SPARROWS 645 U.S. Department of Interior (USDD. 2008. Birds of conservation concern 2008. USDI. Fish and Wildlife Service, Division of Migratory Bird Management, Arlington, Virginia, USA. Vickery, P. D., M. L. Hunter, and S. M. Melvin. 1994. Effects of habitat area on the distribution of grass- land birds in Maine. Conservation Biology 8:1087- 1097. W.ALK, J. W. AND R. E. Warner. 1999. Effects of habitat area on the occurrence of grassland birds in Illinois. American Midland Naturalist 141:339-344. Walk, J. W., E. L. Kershner, and R. E. Warner. 2000. Nocturnal singing in grassland birds. Wilson Bulletin 1 12:289-292. Wells, J. V. and K. V. Rosenberg. 1999. Grassland bird conservation in northeastern North America. Studies in Avian Biology 19:72-80. Wheelwright, N. T. and J. D. Rising. 1993. Savannah Sparrow (Passerculiis sundwichensis). The birds of North America. Number 45. Wiens, J. A. 1969. An approach to the study of ecological relationships among grassland birds. Ornithological Monographs 8:1-93. Wiens, J. A. 1973. Intertenatorial habitat variation in Grasshopper and Savannah sparrows. Ecology 54:877- 884. Winter, M., D. H. Johnson, J. A. Shaffer, T. M. Donovan, and W. D. Svedarsky. 2006. Patch size and landscape effects on density and nesting success of grassland birds. Journal of Wildlife Management 70:158-172. Zimmerman, J. L. 1988. Breeding season habitat selection by the Henslow’s Sparrow (Ammod ramus henslowii) in Kansas. Wilson Bulletin 100:17-24. The Wilson Journal of Ornithology 122(4):646-654, 2010 INFLUENCE OF WOODY VEGETATION ON GRASSLAND BIRDS WITHIN RECLAIMED SURFACE MINES BRET M. GRAVES,'-' AMANDA D. RODEWALD,' AND SCOTT D. HULL- ABSTRACT. — We examined the influence of woody vegetation on reclaimed surface mines on relative abundance of Grasshopper Sparrows {Aminocl ramus savannarum), Henslow’s Spanows (A. henslowii). Eastern Meadowlarks (Stiirnella magna). Savannah Sparrows (Passerculus sandwichensis). Bobolinks (Dolichonyx oiyzivoriis), and Dickcissels (Spiza americana) as well as nest-site selection and nesting success of Grasshopper and Henslow’s sparrows and Eastern Meadowlarks. Grasshopper and Henslow's sparrows were the most abundant grassland species on reclaimed mines. Numbers of Grasshopper. Henslow’s, and Savannah spanows. and Bobolinks were negatively a.ssociated with percent cover of woody vegetation within 100 m of survey locations. Only Grasshopper Sparrows responded to woody vegetation at nest- patch scales, as random locations had >2.5 times as much woody cover as nest locations. Daily nest survival (DNS) was negatively associated with amount of woody vegetation within 100 m of Grasshopper (DNS 0.76 ± 0.001 SE) and Henslow’s sparrow nests (DNS 0.94 ± 0.020 SE). but only marginally negatively related to daily nest survival of Eastern Meadowlark nests (DNS 0.87 ± 0.006 SE). Avoidance of woody vegetation by grassland birds and the comparatively lower daily nest survival of Grasshopper and Henslow’s sparrow nests near woody vegetation suggests managers of reclaimed surface mines who manage to conserve grassland birds should direct efforts towards reducing woody encroachment. Received 22 June 2009. Accepted 14 July 2010. Grassland birds have declined more rapidly than any other group of birds in the Midwest over the past 30 years (Peterjohn and Sauer 1999, Cully and Michaels 2000, Vickery and Herkert 2001, Sauer et al. 2005). Among the most precipitously declining are area-sensitive species including Grasshopper Sparrow (Ammodramiis savan- nanini), Henslow’s Sparrow (A. henslowii), and Eastern Meadowlark (Stiirnella magna) (Herkert 1994). These declines have been largely attributed to changes in agricultural practices (Johnson and Igl 2001), habitat loss (Askins 1993), and habitat degradation and fragmentation (Herkert 1994, Cully and Michaels 2000, Vickery and Herkert 2001). Large gains in native and managed grassland habitats in the Midwest are unlikely to occur and successful conservation of grassland birds will require consideration of new opportu- nities to provide quality grassland habitat for area- sensitive grassland-breeding species. Reclaimed surface mines represent one opportunity to manage for grassland bird species (Whitmore and Hail 1978, Whitmore 1980, Wray et al. 1982, Bajema et al. 2001, DeVault et al. 2002, Monroe and Ritchison 2005. Galligan et al. 2006). Reclaimed surface mines represent a conservation ' .School of Environment and Natural Resource.s, The Ghio State University. 2021 Coffey Road. 210 Kottman Hall, Columbus. OH 4.^210, USA. Hfureau of Wildlife Management, Wisconsin Depart- ment of Natural Resources, 101 South Webster Street, P. O. Box 7921. Madison. W1 53707, USA. 'Corresponding author; e-mail: graves. 1 2@o.su.edu paradox in that grassland birds occupy them even though they are highly disturbed, dominated by exotic grasses, and vulnerable to invasion by exotic woody vegetation such as autumn olive (Elaeagmis iimbellata). Woody encroachment is a key management issue with the potential to impact the conservation value of reclaimed surface mines for grassland- breeding birds. Reforestation of mined areas to the original deciduous forest landscape is often not po.ssible due to poor soil conditions (Wray et al. 1982, Brothers 1990, DeVault et al. 2002, Scott et al. 2002), but invasion by woody species such as autumn olive and black locust (Robinia pseiidoa- cacia) commonly occurs if left unmanaged (Rum- mel and Brenner 2003). The decision to maintain open grasslands by removing woody vegetation remains controversial despite known negative effects of woody vegetation on grassland birds (Bajema et al. 2001, Ribic and Sample 2001, Bakker et al. 2002, Fletcher and Koford 2002, Grant et al. 2004). Removal of woody vegetation within reclaimed surface mines may alleviate nest predation by redistributing movements of common mammalian nest predators including common raccoon (Procyon lotor), striped skunk (Mephitis mephitis), and Virginia opossum (Didelphis vir- giniana) (Winter et al. 2000). However, some grassland species such as Vesper (Pooecetes gramineiis) and Clay-colored (Spizella pallida) sparrows, and Common Yellowthroats (Geothlypis trichas) may be tolerant of woodland encroach- ment at the landscape .scale (Grant et al. 2004). 646 Graves el a/. • GRASSLAND BIRD CONSERVATION ON RECLAIMED MINES 647 Effective encouragement and conservation of grassland birds on reclaimed surface mines requires an understanding of the influences that unmanaged woody plant encroachment can have on distribution, nest-site selection, and nesting success ot obligate grassland birds. We focused on six obligate grassland bird species receiving national and regional conservation attention (Walk and Warner 2000, Bajema et al. 2001, Vickery and Herkert 2001, DeVault et al. 2002); Henslow's SpaiTow, Grasshopper Sparrow, East- ern Meadowlark, Savannah Span'ow {Passerciiliis sandwichensis). Bobolink (Dolichonyx oryzi- voriis), and Dickcissel (Spiza americana). We examined: ( 1 ) how woody vegetation features were related to habitat use and distribution of Grasshopper, Savannah, and Henslow’s spaiTows, Eastern Meadowlarks, Bobolinks, and Dickcissels breeding within reclaimed surface mines; (2) the extent woody habitat features were related to nest- site selection of Grasshopper and Henslow’s sparrows, and Eastern Meadowlarks; and (3) if these woody habitat features affected daily nest survival of Grasshopper and Henslow’s spaiTows, and Eastern Meadowlarks. METHODS Study Area. — This study was conducted on reclaimed surface mines within wildlife manage- ment areas in eastern Ohio, USA, managed by the Ohio Department of Natural Resources (ODNR), Division of Wildlife. Collectively, these wildlife management areas represent the range of re- claimed surface mine habitats available to grass- land birds in eastern Ohio. Tri-Valley Wildlife Management Area (WMA) in northern Muskin- gum County is —6,100 ha comprised of —50% open land (grasslands, wetlands, and food plots), 40% woodland, and 10% brushland. Approxi- mately 2,600 ha of reclaimed grasslands are within Tri-Valley WMA. Woodbury WMA in southern Coshocton County is —7,600 ha of —35% openland (grasslands, wetlands, and food plots), 8% brushland, and 57% woodland. Ap- proximately 1,200 ha of reclaimed grasslands are within Woodbury WMA (ODNR 2007). Dominant vegetation types in the study area were non-native cool .season grasses including fescue (Festiica spp.), redtop (Agrostis gigantea), timothy (Phleum pratense), and orchard grass (Dacty'lis glomerata) as well as native warm season grasses including switchgrass (Pauiciim virgatum), Indian grass (Sorghastnim nutans), big bluestem (Andropogon gerardii), and little blue- stem {Schizachyriuni scopariurn). Native and non- native forbs {Melilotus spp., Trifoliuin spp., Solidago spp., Lotus corniculatus), and woody vegetation including autumn olive, black locust, blackberry (Ruhus allegheniensis), multiflora rose (Rosa inultiflora), lespedeza (Lespedeza spp.), (lowering dogwood (Cornus florida), and pines (Pinus spp.) were also well represented. Surveys. — Grassland birds were surveyed from May through July 2005 and 2006 at 101 different point-count locations in Tri-Valley (24 in 2005 and 29 in 2006) and Woodbury (24 in 2005 and 24 in 2006). We used a systematic random sampling design for point-count locations. Starting loca- tions (6 per study site) of point-count survey transects were randomly selected using a Geo- graphic Information System (GIS); subsequent point-count locations were systematically placed 250 m apart until it was not possible to place the next point 250 m from the previous location (i.e., bander such as a woodland area or large wetland complex). All point-count locations (n = 101) were recorded with a handheld Global Positioning System (GPS) unit. We performed grassland bird surveys at each of the 101 point-count locations three times each from 15 May to 15 July in both 2005 and 2006 between 0630 and 1100 hrs EST on days without fog, precipitation or winds >25 kph. Two trained observers first allowed a 2-min period for birds to adjust to surveyor presence at each point-count location followed by a 6-min survey period when all singing male grassland birds were recorded within a 100 m radius. The time of detection of each bird was recorded during each point-count survey along with species code and straight-line distance (m) to each bird detected with the use of a laser rangefinder (Ransom and Pinchak 2003). Nest Survival. — We randomly selected eight of 16 5-ha study plots similar in topography within the grassland units containing varying amounts of woody encroachment for nest searches (4 within Woodbury and 4 within Tri-Valley). Field teams used a .spot-mapping protocol (Bibby et al. 2000) to identify breeding territories to facilitate locat- ing nests. Grids were visited eight times during morning hours as part of the spot-mapping efforts from 1 May to 16 June 2005 and 2006, and an observer walked the entire 5-ha plot on parallel transects (.separated by 75 m). These spot-maps helped field teams extensively search plots for nests of Grasshopper and Henslow’s sparrows. 648 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4, December 2010 and Eastern Meadowlarks from 1 May to 1 August 2005 and 2006. We initially conducted nest searches using rope-dragging techniques where a 30-m rope was dragged behind two individuals while a third walked several meters behind the rope looking for birds that flushed from nests. We also relied on behavioral cues to find nests and carefully followed adults that were carrying food and nesting material until a nest was located. Observers systematically searched known breeding territories obtained from spot maps to locate nests. Some nests were located by chance during bird surveys and while taking vegetation measurements. We marked nest locations with a handheld GPS unit as well as with one colored flag placed 5 to 10 m from the nest depending on how concealed the nest was to avoid flushing the attending adult. We checked nests every 3^ days during the incubation period and every 2-3 days when young neared the fledging stage. Nests were visited from different angles during each nest check to avoid creating paths that may attract predators. The number of eggs and/or nestlings, date the nest was checked, and the species name were recorded. A nest was considered active if at least one egg was present. A typical nest period for the three focal species was 26 days with 8- 10 days post-hatching to fledge young (Ingold 2002). We considered a nest successful if at least one young fledged or if it was active for >8 days post-hatching (Vickery 1996) and there were no obvious signs of predation. A nest was considered lost to predation if the eggs were removed, cold or broken or if the nest was disturbed and the nestlings were no longer in the nest during the first week post hatching (Ingold 2002). GIS Analysis.— 'We estimated the amount of woody vegetation within 100 m of nests and survey points using National Agricultural Imagery Program (NAIP) aerial photographs from 2005 and 2006 of Coshocton and Muskingum counties, which were projected using ArcGIS 9.1 (ESRl 2005). The percentage of woody vegetation and number of woody patches were estimated for circular plots with a lOO-m radius centered on each nest (n = 81) and on all GPS .survey locations in = 101). We visited plots and directly compared vegetation shown in images to actual vegetation on plots observed during field visits to verify the size of woody plants detectable in aerial photographs. Ground-truthing revealed woody vegetation patches <4 m^ in area were not reliably detected from aerial photographs; conse- quently patches of woody vegetation >4 m" in size were defined as woody vegetation patches. Each woody patch of vegetation >4 m“ within 100 m of each nest and survey point was digitized and the area of each polygon was used to calculate the total area of woody vegetation within a 100-m radius suiTOunding each nest and survey location. The total percent of woody vegetation was calculated by dividing the total area (m‘) of woody vegetation by the total area of the circle (31,415 m") multiplied by 100. Aerial photo- graphs, projected in ArcGIS 9.1, were used to measure the distance (m) from each nest and each survey location to the nearest woodland edge. Statistical Analyses. — We compared an a priori set of woody vegetation variables (percent cover of woody vegetation, number of woody patches, and distance to woodland edge) on bird survey locations to identify which woody vegetation features were used as cues by grassland birds in selecting suitable habitat. We u.sed 101 survey locations as replicates to identify woody habitat features used by Grasshopper, Henslow’s, and Savannah spairows. Eastern Meadowlarks, Bobo- links, and Dickcissels as cues in selecting suitable habitat for breeding locations. The relationship between relative abundance and woody vegetation was tested using regression analysis in Proc GENMOD (SAS Version 9.1; SAS Institute 2002) using generalized linear models. Bird point-count data were transformed when appro- priate to minimize non-normality and variance heterogeneity. A log transformation was applied to counts of Grasshopper and Henslow’s spar- rows, and Eastern Meadowlarks and a normal distribution was used. A Poisson distribution was u.sed for counts of Savannah Sparrows, Bobolinks, and Dickcissels. Percent woody vegetation, number of woody vegetation patches >4 m^, and the distance from plot center to the nearest woodland edge were moderately correlated (r < 10.51) and were incorporated into a principal components analysis (Proc FACTOR in SAS Version 9.1; SAS Institute 2002) to characterize woody vegetation surround- ing survey locations using a single parameter. The first factor of the three principle components explained >56% of the variation in woody vegetation among survey points (Eigenvalue = 1 .69). This first factor loaded positively for percent woody cover (0.88) and number of woody patches (0.70), and loaded negatively for distance to woodland edge (—0.65). (iraves cl at. • GRASSLANr:) HU^D CONSERVATION ON RECLAIMED MINES 649 We examined an a priori set of woody vegetation variables (percent cover of woody vegetation, number of woody patches, distance to nearest woody patch, and distance to woodland edge) at individual nests and random locations to learn which woody vegetation variables influenced grassland birds during nest-site selection. We used 81 nests as replicates to identify woody habitat features used by individual birds (Grasshopper and Henslow's spaiTows, and Eastern Meadowlarks) to select nest-site locations for each species indepen- dently during individual nest attempts. We trans- formed the variables when appropriate to minimize non-normality and variance heterogeneity. Percent woody vegetation, number of woody vegetation patches, and distance from plot center to the nearest woody patch and woodland edge were used in a discriminant function analysis (Quinn and Keough 2002) (DFA; Proc CANDISC in SAS Version 9.1; SAS Institute 2002) to examine which variables best discriminated between nest and random plots. Variables that best discriminated between nest and random plots were interpreted as potential habitat cues used by grassland birds in selecting locations for nest placement (Quinn and Keough 2002). We examined how woody habitat features in nest plots and at nests influenced nest success by incorporating the first principal component de- scribing woody vegetation suiTounding nest sites (PCwood) into the logistic-exposure method (Shaffer 2004) to model daily nest survival rates of Henslow’s and Grasshopper sparrows, and Eastern Meadowlarks. This approach models the success or failure of nests during each interval between nest checks and evaluates the probability of nest success over a range of values for influential categorical and continuous explanatory variables. Nest losses to all sources were classified as failures. Ne.sts abandoned prior to laying or found after depredation had already occurred were excluded. We fit the model (PCwood interacting with species) with Proc GENMOD (SAS Version 9.1; SAS Institute 2002) using a binomial response distribution (interval nest fate = 1 if success, and 0 if fail) and provided the user-defined logit link function (g(6) = loge (0'Vl 1 — 0'^'])) where t = the length of the interval (Shaffer 2004). RESULTS Grasshopper and Henslow’s sparrows were the most abundant species on Tri-Valley and Wood- bury WMAs during 2005 and 2006. We detected 1,491 birds on point counts during the study. including 732 Henslow’s Sparrows, 465 Grass- hopper Sparrows, 209 Eastern Meadowlarks, 39 Bobolinks, 31 Dickcissels, and 15 Savannah Sparrows. Mean relative abundance ranged from 2.33 birds per survey point for Henslow’s Sparrows at Tri-Valley to 0 for Savannah Sparrows and Bobolinks at Woodbury. The amount of woody vegetation was negatively associated with abundance of Grasshopper Spar- rows (Beta estimate = -0.14 ± 0.041 SE; x~ = 11.34, P < 0.001), Henslow’s Sparrows (Beta estimate = —0.15 ± 0.051 SE; = 7.73, P = 0.005), Savannah Sparrows (Beta estimate = -2.47 ± 0.975 SE; x' = 7.38, P = 0.007), and Bobolinks (Beta estimate = —2.35 ± 0.745 SE; X“ = 1 1.48, P < 0.001 ), but not significantly for Eastern Meadowlarks (Beta estimate = —0.11 ± 0.241 SE; x^ = 0.21, P = 0.65) or Dickcissels (Beta estimate = -0.57 ± 0.704 SE; x“ = 0.84, P = 0.36) (Fig. 1). Forty-five Grasshopper Span'ow, 18 Henslow’s Sparrow, and 1 8 Eastern Meadowlark nests were monitored during the nesting seasons. Habitat variables discriminated between random locations and nest sites of Grasshopper Span'ows (Canon- ical con-elation = 0.25, Likelihood ratio = 0.936; Wilks’ Lambda F3 ,42 = 3.24, P = 0.024) (Table 1). Random locations had >2.5 times as much woody vegetation as ne.st locations of Grasshopper Sparrows (F| 144 = 5.41, P = 0.021). Nest placement by Grasshopper Sparrows was not associated with number of woody patches (F| 144 = 1.75, P = 0.19) or distance to woodland edge (F| 44 = 0.96, P = 0.33). Habitat variables did not discriminate between random plot loca- tions and nests of Henslow’s SpaiTows (Wilks' Lambda F3 j 14 = 1.12, P = 0.35) or Eastern Meadowlarks (Wilks’ Lambda F3 , ,4 = 1.35, P = 0.26) (Table 1). Twenty of 45 (44%) Grasshopper Spairow nests, sixteen of 18 (89%) Henslow's Sparrow nests, and nine of 18 (50%) Eastern Meadowlark nests were successful (i.e., fledged at least 1 young or young remained in the nest >8 days post-hatching). The amount of woody vegetation surrounding nest locations was significantly associated with daily nest survival of Grasshopper and Henslow’s spaiTows, and Eastern Meadow- larks (Beta estimate = -1.19 ± 0.457 SE; x" = 6.57, P = 0.010), despite all three species differing in daily nest survival rates (Beta estimates: Eastern Meadowlark = -1.55 ± 1.17 SE, Grasshopper Spari'ow = —2.39 ± 1.13 SE, 650 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 4, December 2010 FIG. 1. Relative abundance of grassland bird species (A-F) within 100-m radius of 101 point-count locations on reclaimed surface mines in relation to amount of woody vegetation (PC wood) in eastern Ohio, 2005-2006. Amount of woody vegetation increases from —3 to 4 along the x axis. Hen.slow’.s Sparrow = 0.00; “ 11.71, P — 0.003). There was no evidence of significant interactions and all three species responded similarly to woody vegetation surrounding their nests = 4.32, P = 0.12). Mean ± SE daily nest survival rate was 0.76 ± 0.001 for Grasshopper Sparrows, 0.94 ± 0.020 for Henslow’s Sparrows, and 0.87 ± 0.006 for Eastern Meadowlarks. Daily nest survival of Grasshopper and Henslow’s sparrows was nega- Graves el al. • CJRASSLAND liIRD CONSERVATION ON RECLAIMED MINES 651 TABLE I. Mean percent woody vegetation, number of woody patches, and distance to woodland edge (± SH) for nest- site locations of Grasshopper and Henslow’s sparrows, and Eastern Meadowlarks and random vegetation plots. Grasshopper Sparrow Herislow's Sparrow Eastern Meadowlark Random Variable v ± SE F P .v ± SE F P J ± SE F P ^ ± SE % Woody vegetation 1 .5 H- 0.29 5.4 0.021 2.3 -F 1.38 1.0 0.325 1.5 -F 0.64 2.2 0. 1 37 4.0 -F 0.70 # Woody patches 3. 1 Distance to H- 0.49 1.8 0.188 2.7 1.04 1.4 0.233 1.9 -F 0.42 3.6 0.061 4.2 -F 0.51 woodland 165.3 -F 1 1.70 1.0 0.331 161.1 16.09 0.7 0.396 201.8 -+■ 19.39 0.7 0.391 181. 1 H- 9.49 lively associated with amount of woody vegeta- tion (PCwood) surrounding nest locations (Fig. 2). Eastern Meadowlark daily nest survival was marginally associated with amount of woody vegetation in proximity to nest locations (Fig. 2). DISCUSSION Our findings suggest woody vegetation nega- tively influenced abundance, nest placement, and nesting success of some grassland bird species on reclaimed surface mines. Our results are consis- tent with previous studies of the effects of woody vegetation on grassland bird occurrence in the midwestern U.S.; however, our study is unique as we specifically examined reclaimed surface mines (but see Bajema et al. 2001 and Galligan et al. 2006). The abundance of obligate grassland birds has been shown to be negatively related to woody vegetation in Minnesota and North Dakota (Grant et al. 2004, Winter et al. 2006, respectively), southwestern Missouri (Winter and Faaborg 1999), Illinois (O’Leary and Nyberg 2000), eastern South Dakota (Bakker et al. 2002), and West Virginia (Wray et al. 1982). Our failure to detect an association between woody vegetation and abundance of Eastern Meadowlarks and Dickcissels was possibly an artifact of the relatively large territory size of these species and use of fixed radius sampling (detections >100 m were not included in the analysis). Daily nest survival by species PCwood FIG. 2. Daily survival rates of Grasshopper and Henslow’s sparrow, and Eastern Meadowlark nests in relation to amount of woody vegetation (PCwood) in proximity to nest-site locations at Tri-Valley and Woodbury Wildlife Management areas in eastern Ohio, 2005-2006. Amount of woody vegetation increa.ses from - 1 to 3 along the x axis. 652 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122. No. 4. December 2010 Winter and Faaborg (1999) also found no significant association between woody vegetation and density of these two species. The small number of observations on surveys may have limited our ability to identify patterns in cases where we did not detect significant associations (e.g.. Savannah Sparrow and Dickcissel). Numbers of Grasshopper, Henslow’s, and Savannah sparrows, and Bobolinks were nega- tively related to percent cover of woody vegeta- tion within 100 m of survey locations, but only Grasshopper Sparrows showed strong evidence of selecting nest patches with lower amounts of woody vegetation than random locations. Our analysis failed to discriminate between Eastern Meadowlark and Henslow’s Sparrow nests and random locations, perhaps because of small sample size {n = 18 nests) or the conservative nature of our comparison of used plots to random plots versus used plots to non-used plots. Nest locations of Eastern Meadowlarks had more than two times less woody cover and fewer woody patches within 100 m than random locations (Aldredge and Griswold 2006, Buskirk and Mill- spaugh 2006, Johnson et al. 2006, Thomas and Taylor 2006). The literature on effects of woody vegetation on nest-site selection of grassland birds has yet to reveal consistent relationships, and investigations show active avoidance of woody vegetation during nest-site selection of Eastern Meadowlarks, Henslow’s and Grasshopper spar- rows, and Dickcissels (Hull 2000, Winter et al. 2000, Hubbard et al. 2006), positive associations with trees for Grasshopper Spairows (Sutter and Ritchison 2005), or no association between nest- site .selection and presence of woody vegetation for Vesper and Clay-colored sparrows (Grant et al. 2006). Grassland nesting passerines may use the ab.sence of woody vegetation as a cue during nest-site selection given that certain predators (i.e., mice) and brood parasites (i.e., Brown- headed Cowbird [Molothru.s ater]) frequently use woody vegetation both as cover and visual perches (With 1994). Daily nest survival was related to the amount of woody vegetation surrounding nests of Grasshop- per and Henslow’s sparrows and marginally related to Eastern Meadowlark nests, which is consistent with other studies conducted elsewhere in the Midwest (Winter et al. 2000, Scheiman et al. 2003, Bollinger and Gavin 2004, Davis 2005; but see Grant et al. 2006). Woody vegetation and woodland edges may support a diverse predator community not otherwise present in grassland systems (Renfrew and Ribic 2003). Woody elements are generally thought to attract predators because they provide perches for avian predators and promote use by other predators due to cover and other resources (Thompson and Burhans 2003, Renfrew et al. 2005). We suggest efforts to remove woody vegetation will enhance the value of reclaimed surface mines for grassland birds given the apparent avoidance of woody vegetation and the overall low daily nest survival of Grasshopper Span'ows and Eastern Meadowlarks, coupled with results from other studies relating low nesting success to presence of woody vegetation. Removal of encroaching woody plants will increase grassland area and reduce edges and fragmentation. It may also decrease numbers of woodland predators by eliminating cover or perches and reducing move- ment corridors. This should make grassland units more productive and attractive to a diverse community of grassland-nesting birds. Our study emphasizes the role of woody vegetation in grassland bird management on reclaimed surface mines, but managers should also remain attentive to landscape and patch-scale issues given their known influence on grassland birds (Ribic and Sample 2001, Bakker et al. 2002, Winter et al. 2005). ACKNOWLEDGMENTS R. .1. Gates and Daniel Shiistack provided statistical advice, and Jacob Straub provided GIS expertise. We thank Kacy Ray, Michelle Schroeder, Beth Geboy, and Jennifer Haney for assistance in the field. Financial support for this project was provided by the Ohio Department of Natural Resources, Division of Wildlife and by the Ohio Biological Survey through their small grants program. LITERATURE CITED Aldredge, .1. R. and J. Griswold. 200b. Design and analysis of resource selection studies for categorical resource variables. Journal of Wildlife Management 70:337-346. Askins, R. a. 1993. Population trends in grassland, shrubland, and forest birds in eastern North America. Current Ornithology I 1:1-34. Bajema, R. a., T. L. DeVault, P. E. Scott, and S. L. Lima. 2001. Reclaimed coal mine grasslands and their significance for Henslow's Sparrows in the American Midwest. Auk 1 18:422-431. Barker, K. K.. D. E. Nauct-e, and K. F. Higgins. 2002. Incorporating landscape attributes into models for migratory grassland bird con.servation. Conservation Biology 16:1638-1646. Graves et al. • GRASSLAND BIRD CONSERVATION ON RECLAIMED MINES 653 Bibby, C. J., N. D. Burgess, D. A. Hill, and S, H. Mustoe. 2000. Bird census techniques. Academic Press, San Diego, California, USA. Bollinger, E. K. and T. A. Gavin. 2004. Responses ol nesting Bobolinks (DoUchonyx oryzivonts) to habitat edges. Auk 121:767-776. Brothers, T. S. 1990. Surface-mine grasslands. Geograph- ical Review 80:209-225. Buskirk, S. W. and J. J. Millspaugh. 2006. Metrics for studies of resource selection. Journal of Wildlife Management 70:358-366. Cully, J. F. and H. L. Michaels. 2000. Henslow’s Sparrow habitat associations on Kansas tallgrass prairie. Wilson Bulletin 112:115-123. D.avis, S. K. 2005. Nest-site selection patterns and the influence of vegetation on nest survival of mixed-grass prairie passerines. Condor 107:605-616. DeVault, T. L., P. E. Scott, R. A. Bajema, and S. L. Lim.a. 2002. Breeding bird communities of reclaimed coal-mine grasslands in the American Midwest. Journal of Field Ornithology 73:228-275. ESRl. 2005. ArcGIS Version 9.1. Environmental Systems Research Institute, Redlands, California, USA. Fletcher, R. J. and R. R. Koford. 2002. Habitat and landscape associations of breeding birds in native and restored grasslands. Journal of Wildlife Management 66:101 1-1022. Galligan, E. W., T. L. DeVault, and S. L. Lima. 2006. Nesting success of grassland and savanna birds on reclaimed coal mines of the midwestern United States. Wilson Journal of Ornithology 118:537-546. Grant. T. A., E. Madden, and G. B. Berkey. 2004. Tree and shrub invasion in northern mixed-grass prairie: implications for breeding grassland birds. Wildlife Society Bulletin 32:807-818. Grant, T. A., E. Madden, T. L. Shaffer, P. J. Pietz, G. B. Berkey, and N. J. Kadrmas. 2006. Nest survival of Clay-colored and Vesper sparrows in relation to woodland edge in mixed-grass prairies. Journal of Wildlife Management 70:691-701. Herkert, j. R. 1994. The effects of habitat fragmentation on midwestern grassland bird communities. Ecological Applications 4:461^7 1 . Hubbard, R. D,, D. P. Althoff, K. A. Blecha, B. A. Bruvold. and R. D. Japuntich. 2006. Nest site characteristics of Eastern Meadowlarks and Grasshop- per Sparrows in tallgra.ss prairie at the Fort Riley Military Installation, Kansas. Transactions of the Kansas Academy of Science 109:168-174. Hull, S. D. 2000. Effects of management practices on grassland birds: Eastern Meadowlark. USDI, Geolog- ical Survey, Northern Prairie Wildlife Research Center, Jamestown. North Dakota. USA. Ingold, D. J. 2002. Use of reclaimed stripmine by grassland nesting birds in east-central Ohio. Ohio Journal of Science 102:56-62. Johnson, C. J., S. E. Nielsen, E. H. Merrill, T. L. McDonald, and M. S. Boyce. 2006. Re.source selection functions based on use-availability data: theoretical motivation and evaluation methods. Journal of Wildlife Management 70:347-357. Johnson, D. H. and L. D. Icjl. 2001. Area requirements of grassland birds: a regional perspective. Auk 118:24- 34. Monroe, M. S. and G. Rlichison. 2005. Breeding biology of Henslow’s Sparrow on reclaimed coal mine grasslands in Kentucky. Journal of Field Ornithology 76:143-149. Ohio Department of Natural Resources (ODNR). 2007. Woodbury Wildlife Management Area. ODNR, Division of Wildlife, Columbus, Ohio, USA. O’Leary, C. H. and D. W. Nyberg. 2000. Treelines between fields reduce the density of grassland birds. Natural Areas Journal 20:243-249. Peterjohn, B. G. and j. R. Sauer. 1999. Population status of North American grassland birds from the North American Breeding Bird Survey, 1966-1996. Studies in Avian Biology 19:27^4. Quinn, G. P. and M. J. Keough. 2002. Multivariate analysis of variance and discriminant analysis. Pages 425-442 in Experimental designs and data analysis for biologists (G. P. Quinn and M. J. Keough, Editors). University Press, Cambridge, United Kingdom. Ransom, D. and W. E. Pinchak. 2003. Assessing accuracy of a laser rangefinder in estimating grassland bird density. Wildlife Society Bulletin 31:460-463. Renfrew, R. B. and C. A. Ribic. 2003. Grassland passerine nest predators near pasture edges identified on videotape. Auk 120:371-383. Renfrew, R. B., C. A. Ribic, and J. L. Nack. 2005. Edge avoidance by nesting grassland birds: a futile strategy in a fragmented landscape. Auk 122:618-636. Ribic, C. A. and D. W. Sample. 2001. Associations of grassland birds with landscape factors in southern Wisconsin. American Midland Naturalist 146:105- 121. Rummel, S. M. and F. j. Brenner. 2003. Relationships between vegetation and bird populations in reclaimed mine grasslands and native grasslands. Journal of The Pennsylvania Academy of Science 76:56-61. SAS Institute. 2002. SAS/STAT Software. Release 9.1. SAS Institute Inc., Cary, North Carolina, USA. Sauer, J. R., J. E. Hines, and J. Fallon. 2005. The North American Breeding Bird Survey, results and analysis 1966-2004. Version 2005.2. USGS, Patuxent Wildlife Research Center, Laurel. Maryland, USA. SCHEIMAN, D. M., E. K. Bollinger, and D. H. Johnson. 2003. Effects of leafy spurge infestation on grassland birds. Journal of Wildlife Management 67:1 15-121. ScoiT, P. E., T. L. DeVault. R, A. Bajema. and S. L. Lima. 2002. Grassland vegetation and bird abundances on reclaimed midwestern coal mines. Wildlife Society Bulletin 30:1006-1014. Shaffer, T. L. 2004. A unified approach to analyzing nest success. Auk 121:526-540. Sutter, B. and G. Ritchison. 2005. Effects of grazing on vegetation structure, prey availability, and reproduc- tive success of Grasshopper Sparrows. Journal of Field Ornithology 76:345-351. Thomas, D. L. and E. J. Taylor. 2006. Study designs and tests for comparing resource use and availability. II. Journal of Wildlife Management 70:324-336. 654 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4, December 2010 Thompson, F. R. and D. E. Burhans. 2003. Predation of songbird nests differs by predator and between field and forest habitats. Journal of Wildlife Management 67:408^16. Vickery, P. D. 1996. Grasshopper Sparrow {Ammodramus savannarum). The birds of North America. Number 239. Vickery, P. D. and J. R. Herkert. 2001. Recent advances in grassland bird research: where do we go from here? Auk 1 18:1 1-15. Walk, J. W. and R. E. Warner. 2000. Grassland management for the conservation of songbirds in the midwestem USA. Biological Conservation 94:165- 172. Whitmore, R. C. 1980. Reclaimed surface mines as avian habitat islands in the eastern forest. American Birds 34:13-14 Whitmore, R. C. and G. A. Hall. 1978. The response of passerine species to a new resource: reclaimed surface mines in West Virginia. American Birds 32:6-9. Winter, M. and J. Faaborg. 1999. Patterns of area sensitivity in grassland-nesting birds. Conservation Biology 13:1424-1436. Winter, M., D. H. Johnson, and J. Faaborg. 2000. Evidence for edge effects on multiple levels in tallgrass prairie. Condor 102:256-266. Winter, M., D. H. Johnson, and J. A. Shaffer. 2005. Variability in vegetation effects on density and nesting success of grassland birds. Journal of Wildlife Management 69:185-197. Winter, M., D. H. Johnson, J. A. Shaffer, T. M. Donovan, and W. D. Svedarsky. 2006. Patch size and landscape effects on density and nesting success of grassland birds. Journal of Wildlife Management 70:158-172. With, K. A. 1994. The hazards of nesting near shrubs for a grassland bird, the McCowns’ Longspur. Condor 96:1009-1019. Wray, T., 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. The Wilson Journal of Ornithologv 122(4):655 665, 2010 EFFECTS OF BREEDING STAGE AND BEHAVIORAL CONTEXT ON SINGING BEHAVIOR OF MALE INDIGO BUNTINGS MATTHEW D. BECKETT' AND GARY RITCHISON'- ABSTRACT. — We studied the effects of breeding stage and behavioral context on the singing behavior of male Indigo Buntings (Passerina cyanea\ n = 15) during the 2004 breeding season in Madison County, Kentucky, USA, to better understand how males with a single-song repertoire vary the characteristics of their song to convey different information. Playback experiments were conducted in 2005 in territories of focal males (n = 14) to further examine the possible effect of male-male interactions on singing behavior. We analyzed 10,919 songs of 15 male Indigo Buntings with songs consisting of a series ot figures that were usually paired (i.e., phrases). Mean song duration was 2.30 ±0.13 (SE) sec (range = 1.44- 3.40 sec) with males varying song duration by varying the number of figures and phrases in each song. Singing rates varied significantly (P < 0.0001) among breeding stages and were highest prior to pairing, suggesting singing has a role in mate attraction. Singing rates also differed (P = 0.013) during playback experiments with rates higher during playback and post- playback periods (Jf = 4.4 songs/min) than during the pre-playback period (x = 2.9 songs/min). These results suggest that singing also has a role in territory defense. Songs of male buntings tended to be shorter prior to pairing and were generally longer in duration after pairing. Playback experiments revealed that bunting songs were longer {P = 0.03) during and after playback (x = 2.6 sec) than during the pre-playback period (x = 2.0 sec). These results suggest male Indigo Buntings vary singing rates and song duration to convey different information, appearing to use shorter songs uttered at high rates to attract mates and longer songs to convey aggression during male-male interactions. Received 21 December 2009. Accepted J June 2010. Males of most species of passerines have multi- song repertories (MacDougal-Shackleton 1997). These repertoires may be important in male-male interactions (song matching; Stoddard et al. 1992), preventing habituation (Krebs et al. 1978, Yasukawa 1981), providing information about motivational state (Falls 1969), and repelling intruders (Krebs et al. 1978, Yasukawa 1981). Song repertoires may also have intersexual functions. For example, female Atlantic Canaries {Serinus canaria) exhibit more nest building behavior when subjected to playback of larger repertoires (Kroodsma 1976). Males of some species, such as Ovenbirds (Seiurus aurocapillcv. Weary and Lemon 1988) and Henslow’s Sparrows {Ammodramus henslowii, Left- wich and Ritchison 2008), have a single song type. Males with single-song repertoires could potentially use singing to convey different information, but may do so by altering the characteristics of their single song (Morton and Young 1986, Leftwich and Ritchison 2008). However, few investigators have examined how male songbirds with single-song repertoires vary their singing behavior or the characteristics of their songs during the breeding season and in different behavioral contexts. Male Indigo Buntings (Passerina cyanea) have a repertoire of one song type or theme (Payne ' Department of Biological Sciences, Eastern Kentucky University, Richmond, KY 40475, USA. ^Corresponding author; e-mail: gary.ritchison@eku.edu 2006), and little is known about how their singing behavior varies during the breeding season and in different social contexts (Thompson 1972, Shio- vitz 1975). We examined the effects of breeding stage and behavioral context on their singing rates and singing behavior to better understand ( 1 ) the possible functions of singing by male Indigo Buntings, and (2) how males with a single-song repertoire might vary their singing behavior to convey different information. METHODS We studied Indigo Buntings from 26 April to 20 August 2004 and 1 9 May to 5 June 2005 at the Central Kentucky Wildlife Management Area, 17 km southeast of Richmond, Madison County, Kentucky, USA. Male buntings were observed and recorded in 2004, and playback experiments were conducted in 2005. The study area consisted primarily of wooded edges, scattered trees, and lields in early stages of succession. Male buntings were captured using playback of conspecific songs to lure them into mist nets. Captured buntings were banded with a U.S. Geological Survey aluminum band plus a unique combination of three colored plastic bands to aid in individual recognition. Male age (second-year [SY] or after- ,second-year [ASY]) was a.ssigned using plumage differences (Pyle 1997, Payne 2006). Teiritory boundaries were delineated by mon- itoring movements of focal males and noting 655 656 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4, December 2010 locations of interactions with conspecifics. Focal individuals were observed and recorded 2-3 times per week during observation periods of 20-30 min. All observations were conducted during the period from just after sunrise to 1100 hrs (EDT) when singing rates of Indigo Buntings are relatively constant (Thompson 1972). Inclement weather may influence singing behavior so buntings were not observed when it was raining. The date, time, and focal bird’s identity were noted on tape prior to each session. Recordings were made using a Sony (TCM-50DV) cassette recorder with a Sennheiser shotgun microphone (Model ME 88), typically at a distance of 10-30 m from focal males. All songs of focal males were recorded during each observation period and the breeding stage was noted. Songs were categorized during each observation, as: ( 1 ) spontaneous advertising (no conspecific males heard or observed), (2) long- range interaction, where a conspecific male was singing in an adjacent or distant territory >50 m away, and (3) short-range interaction, where a conspecific male was singing <50 m away. Eocal male songs were classified as long- and short- range up to 1 min after a conspecific was last heard singing. We divided the breeding season into six stages: (1) pre-pairing (no female present on male’s territory), (2) nest building (period of nest construction by the female), (3) egg-laying (from the day the first egg was laid to the day the penultimate egg of a clutch was laid), (4) incubation (from the day the last egg was laid to the day before first egg hatched), (5) nestling (from the day the first egg hatched to the day before young fledged), and (6) post-fledging (from the day the young leave the nest to the time when fledglings leave male’s territory). Female buntings frequently produce two broods during the breeding season and often began new nests several days after young from a previous nest Hedged. Thus, the post-Hedging period ended when a female began building a new nest. An attempt was made to locate and monitor all nests in the territories of focal males to identify breeding stages. Nests were located by observing movements of females during nest building and nestling provisioning, and by searching likely nest sites in focal territories. Nests were found at several stages and we, at times, backdated to ascertain the dates of breeding stages. We used the following durations for backdating: 8 (lirst nest of the season) or 2 (later nests) days for nest building, 1 day for each egg in a clutch for egg laying, 12 days for incubation, and 9 days for the nestling period (Bradley 1948, Payne 2006). Playback experiments were conducted in 2005 in territories of focal males (/? = 14) to further examine the possible effect of male-male interac- tion on the singing behavior of male Indigo Buntings. All songs used for the playback experiments were taken from commercially-avail- able recordings (Common Birds and Their Songs, Houghton Mifflin Company; More Binding by Ear — Eastern/Central, Houghton Mifflin Company, Boston, MA, USA). Three tapes, each with songs of a different male, were used in the experiments (Tape 1 means: song length = 1.8 sec, number of figures = 4.3, number of phrases = 5.3; Tape 2 means: song length = 2.9 sec, number of figures = 12.5, number of phrases = 7.5; Tape 3 means: song length = 2.8 sec, number of figures = 14.7, number of phrases = 9.3; definitions of figure and phrase are below). Tapes I, 2, and 3 had three, four, and three different versions of each male’s song, respectively. Male buntings on all three tapes sang at a rate of four songs/min. Recordings of males singing at higher rates would have increased the probability of the songs of focal males overlapping those on the recordings, making subsequent song analysis more difficult. A speaker was placed near the center of the focal male’s territory to simulate singing by an intruding male during each experiment. Each experiment consisted of a 3-min pre-playback period and a 6-min playback/post-playback period (with songs played back for 3 min and observa- tions continuing for an additional 3 min after playback ended). All songs recorded during each observation period were analyzed. Characteristics of songs measured included song duration, number of figures (“a sound which produces a single, complete, and distinct impression uninterrupted by silence greater than two centiseconds”; Shiovitz 1975:133), and the number of phrases (“a subdivision of a song based upon recogniz- able groups of sequential figures”; Shiovitz 1975:133; Fig. 1). In addition, we noted if songs had “squeaky” notes added or were uttered at higher than normal volume during the playback experiments. High-frequency “squeaky” notes are occasionally inserted between phrases in .songs of male Indigo Buntings during territorial disputes (Emlen 1972, Thompson 1972, Margo- Beckett and Ritehisott • SINGING BEHAVIOR Ol' INDIGO BUN TINGS 657 ( ^ r Phrases or n I 1 nrn A/ \ N X o c; (U cr 0) Time (sec) FIG. 1. Spectrograph of a typical song of a male Indigo Bunting showing the terminology used. Hash et al. 1994). Song volume was subjectively categorized as either normal or high volume. We attempted to maintain a constant distance from focal males during playback experiments to permit better estimation of song volume. Record- ings of the songs of focal males were subsequent- ly analyzed using sound-analysis software (Raven Version 1.2.1, Cornell Laboratory of Ornithology, Ithaca, NY, USA). Possible variation in characteristics of songs of male Indigo Buntings with male age, breeding stage, and behavioral context were examined using repeated measures analysis of variance with Stu- dent-Newman-Keuls (SNK) post-hoc tests to exam- ine possible differences among means. All analyses were conducted using the Statistical Analysis System (SAS Institute 1999). All values are presented as mean ± standard emor. Mean values were calculated using the means for each focal male to avoid bias due to different sample sizes. RESULTS We found no differences in characteristics of songs of SY and ASY male Indigo Buntings (song duration, number of figures, and number of phrases; all P > 0.17). Thus, songs of all males (SY and ASY) were combined for subsequent analyses. All male Indigo Buntings (/? = 15) in our study had a repertoire of a single song type. Mean song duration was 2.30 ± 0.13 sec and the mean number of figures per song was 9.95 ± 0.5 1 . The mean number of phrases per song was 6.15 ± 0.46. Songs of male buntings exhibited both intra- (Fig. 2) and inter-individual (Fig. 3) variation. Each male had a typical song that included a series of figures (phrases), usually paired, and usually uttered in the same order. The number of figures in each phrase occasionally varied, and figures (phrases) typically used later in songs were often omitted (Fig. 2). Songs of some males (5 of 15, 33%) were unique and shared few or no song figures with other males. Males in adjacent temtories typically shared some and, at times, all of the figures used in their songs and formed song neighborhoods where males used either the same or similar song figures (10 of 15, 67%; Fig. 4). One SY male in our study altered his song shortly after arriving in the study area (Fig. 5). Effects of Breeding Stage and Natural Interac- tions on Stngitig Behavior. — Singing rates (songs/ min) of male Indigo Buntings varied significantly among breeding stages (F5.45 = 15.1. P < 0.001) with rates highest during the pre-pairing stage (SNK test, P < 0.05; Fig. 6). Singing rates during the other breeding stages (nest-building, egg- laying, incubation, nestling, and post-fledging) did not differ (SNK te.st, P > 0.05; Fig. 6). 658 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 4. December 2010 Time (sec) FIG. 2. Spectrographs illu.strating the typical variation that exists in the single song of a male Indigo Bunting. 8 6 4 2 8 6 4 2 8 6 4 2 8 6 4 2 8 ■ 6 ■ 4 ■ 2 - ahs tchison • SINGING BEHAVIOR OE INDIGO BUNTINGS 659 0.5 1 0 1 5 2.0 2.5 3.0 3.5 4 0 Time (sec) istrating the typical variation among songs of different male Indigo Buntings. 8 6 4 2 8 6 4 2 8 6 4 2 8 6 4 2 r'.» M JOURNAL OF ORNITHOLOGY • VoL 122, No. 4. December 2010 J I I I L Male 1 Male 16 0.5 1.0 1.5 2.0 2.5 3.0 Time (sec) 3.5 4.0 the close matching of songs of .several neighboring male Indigo Buntings in a Becken and RUchison • SINGING BEHAVIOR OF INDIGO BUNTINGS 66 1 N >^ O c 0 13 O" 0 Time (sec) FIG. 5. Spectrographs of two songs uttered by a second-year (SY) Indigo Bunting male illustrating song switching. The first song was used immediately after arrival on his territory. Several days later, the second song incorporated figures and phrases from those of a neighboring male. Differences in the mean duration of songs uttered by male Indigo Buntings during different breeding stages approached significance (Fs^i = 2.2, P = 0.075) with longer songs during the nest- building period (Fig. 7). Songs were shorter in duration during post-fledging and pre-pairing periods than during other breeding stages (Fig. 7). Differences among breeding stages in the mean number of figures (F^2,\ = 2.1, F = 0.099) and phrases (F531 = 2.1, P = 0.094) per song also approached significance with more figures and phrases per song during nest-building and egg- laying stages and fewer during post-fledging and FIG. 6. Mean singing rates (± SE) of male Indigo Buntings during different breeding stages. pre-pairing stages (Fig. 8). None of the charac- teristics of songs of male buntings, including duration and number of figures and phrases per song differed among behavioral contexts (short- range interaction, long-range interaction, or spon- taneous singing; P > 0.24). Effects of Song Playback on Singing Behav- ior.— We compared singing behavior of male Indigo Buntings prior to presentation of the stimulus (pre-playback period) to that during and after presentation of the stimulus (playback and post-playback periods combined). Singing rates (songs/min) differed (F,jo = 8.3, P = 0.013) between periods with higher rates during the 4 Pre-pairing Nest- Egg- Incubation Nestling Post- building laying fledging Breeding stage FIG. 7. The mean duration of .songs (± SE) uttered by male Indigo Buntings during different breeding stages. 662 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, Nu. 4. December 2010 FIG. 8. Mean number of figures and phrases (± SE) in songs of male Indigo Buntings during different breeding stages. playback and post-playback periods (x = 4.4 ± 0.4 songs/min) than during the pre-playback period (.v = 2.9 ± 0.6 songs/min). The character- istics of songs also differed between periods with longer songs (/^uo = 6.4, P = 0.03) that included more figures iF\,\o = 6.6, P = 0.028) and phrases (^1.10 = 5.0, P = 0.049) uttered during and after playback (x duration = 2.6 ± 0.1 sec; x number of figures = 1 1.4 ± 0.1; x number of phrases = 6.8 ± 0.2) than prior to playback (x duration = 2.0 ± 0.1 sec; X number of figures = 8.8 ± 0.4; x number of phrases = 5.5 ± 0.2). Male buntings also uttered a significantly higher percentage (6’i.io = 6.3, P — 0.026) of high-volume songs during the post-presentation period (x = 11.4 ± 3.7%) than the pre-presentation period (none uttered). Comparison of pre- and post-presenta- tion periods revealed no difference in use of songs including high-frequency ‘squeaky’ notes (Fijo = 0.02, P = 0.9). DISCUSSION Song Matching. — All male Indigo Buntings in our study had a repertoire of one song type. Songs of some male buntings were unique, as also noted by previous investigators (Thompson 1970, Payne et al. 1988). Most males with adjacent territories in our study shared some, if not all, of the figures in their songs with conspecifics, and formed song neighborhoods where males in adjacent territories used the same or similar song figures. Song neighborhoods have been reported in other populations of Indigo Buntings (Thompson 1970, Payne and Westneat 1988). Song neighbor- hoods may develop because incorporating figures Li.sed by resident males into their .songs may be beneficial for young male buntings with SY males that song match more likely to successfully acquire and defend a territory, acquire a mate, and fledge young (Payne 1982, 1983). Song Function. — Our results indicate singing by male Indigo Buntings functions in mate attraction with singing rates declining significant- ly after pairing. Thompson (1972) also reported male Indigo Buntings sang more songs per bout and more bouts/hr when unmated. Previous studies suggest male singing rates represent a performance-related trait under positive sexual selection (Gil and Gahr 2002) that can influence mate choice and timing of mate acquisition (Alatalo et al. 1990, Hoi-Leitner et al. 1995). Otter et al. (1997) reported that singing rates represented accurate signals of male quality in Black-capped Chickadees {Poecile atricapillus). Singing rates of male buntings and males of other species of songbirds (e.g., Albrecht and Oring 1995, Ritchison 1995, Balsby 2000) decline after pairing. Thus, high singing rates prior to pairing may provide females with information concerning male status (paired or not) and their quality. Male Indigo Buntings in our study continued singing after pairing, but at lower rates. Singing by male buntings, both before and after pairing, as for males of many species of songbirds (e.g., Krebs et al. 1978, Nowicki et al. 1998), is likely important in territorial defense. Male Indigo Buntings typically airive in breeding areas ~1- 2 weeks before females (Payne 2006; GR, pers. obs.) and establish breeding territories. During this period, and throughout our study, males were at times observed counter-singing with neighbor- ing males and chasing conspecifics near temtory boundaries while singing; males responded to playback by increasing their singing rates. These observations indicate that singing by male Indigo Buntings has an important role in establishing and defending territories. Singing rates of male Indigo Buntings in our study were lowest during the fertile periods of their mates (nest-building and egg-laying periods) and remained relatively low during the incuba- tion, nestling, and post-fledging periods. Some investigators have suggested singing may be important for stimulating female reproductive cycles (Hinde and Steel 1976) or guarding mates (Mpller 1988). Singing by male Indigo Buntings does not appear to serve the.se functions because singing rates in our study were lowest during female fertile periods. Males of most songbird species do not sing or sing al low rates during the fertile periods of their mates (Gil et al. 1999). Beckett and Ritehison • SINGING BEHAVIOR OF INDIGO BUN I INGS 663 Song Characteristics. — Songs of male Indigo Buntings in our study tended to be shorter prior to pairing and were generally longer in duration after pairing. In addition, bunting songs were signifi- cantly longer during and after playback than during the pre-playback period. These results suggest male Indigo Buntings may use shorter songs to attract females, and longer songs to convey aggression during male-male interactions. Although females in some species appear to prefer males that sing longer songs (e.g., White-throated Sparrows, Zonotrichia albicollis', Wasserman and Cigliano 1991), females in other species prefer males that sing at faster rates (e.g., European Pied Flycatcher, Ficedula hypoleuca\ Gottlander 1987, Alatalo et al. 1990). If mate choice decisions by female Indigo Buntings are influenced by singing rates, then one possible explanation for the tendency of males to sing shorter songs prior to pairing is that such songs allow males to sing at higher rates (i.e., more songs per unit time). In addition, however, given that longer songs appear to convey an increased tendency for aggression, male buntings may also sing shorter songs prior to pairing to signal their non-aggressive intentions to potential mates. Similarly, Staicer (1996) suggest- ed that male Adelaide’s Warblers (Dendroica adelaidae) may use songs with characteristics that convey less aggression (‘appeasing songs’) during interactions with females. Male buntings in our study often added additional phrases to their songs during encoun- ters with other males, both in response to playback and when chasing conspecifics. Emlen (1972) noted that buntings used ‘lengthened’ songs during agonistic encounters, and Saunders (1929:48) witnessed “male Indigo Buntings chasing each other about from tree to tree, and both of them singing prolonged songs in flight.’’ In addition, Shiovitz (1975) found that playback of long variations of songs elicited stronger responses by male buntings than shorter varia- tions. Shiovitz (1975) also suggested the first few figures of the songs of male buntings likely functioned as a “sign on’’ in gaining the attention of conspecifics, whereas the remainder of the song seemed to convey aggression. Thus, male Indigo Buntings appear to increase song duration to convey aggression, and perhaps the likelihood of interacting, during intrasexual encounters. Males of other species also appear to use longer songs to convey an increased likelihood of aggression during intrasexual encounters (e.g., McGregor and Horn 1992, Balsby and Dabelsteen 2001, Leitao et al. 2006, Lattin and Ritehison 2009). We found male Indigo Buntings at times added ‘squeaky notes’ to songs and sang with increa.sed volume during male-male interactions. Thompson (1972) also indicated male Indigo Buntings inserted high-frequency (about 9 kHz), ‘squeaky’ notes into songs when an intruding male was present in a territory. Emlen (1972) noted that male buntings at times added high-pitched ‘.squeak’ notes when responding aggressively to playback of conspecific songs. These results suggest male buntings alter the characteristics of their single song to convey an increased likeli- hood of aggression, i.e., singing longer, louder songs that at times include ‘squeaky’ notes. In further support of this hypothesis, male buntings in our study at times inserted high-frequency ‘squeaky notes’ into songs during territorial chases with conspecifics and uttered songs at higher than normal volume when countersinging at territory boundaries (MDB and GR, pers. obs.). Changes in characteristics of songs during in- trasexual encounters have also been reported in other species of songbirds. For example, male Barn Swallows (Hintndo rustica) emphasize ‘rattles’ in their songs during aggressive encoun- ters, indicating that ‘rattles’ are an important component of competitive interactions between males (Galeotti et al. 1997). Similarly, American Pipits {Anthus rubescens) incorporate ‘snarr’ notes, a rasping element with a broad frequency range, into their songs during territorial disputes (Rehsteiner et al. 1998). Our results suggest the single song type of male Indigo Buntings serves several functions, includ- ing mate attraction and temtory defense. In addition, male buntings vary song rates and song duration during the breeding season and in different behavioral contexts, appearing to use shorter songs uttered at high rates to attract mates and longer songs to convey aggression during male-male interactions. ACKNOWLEDGMENTS We (hank E. R. A. Cramer, C. E. Braun, and an anonymous reviewer for comments that improved our paper, and the Eastern Kentucky University Research Committee for providing funds to support our research. LITERATURE CITED At.ATALO, R. V., C. Glynn, and A. Lundberg. 1990. Singing rate and female attraction in the Pied 664 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 4. December 2010 Flycatcher: an experiment. Animal Behaviour 39: 601-603. Albrecht. D. J. and L. W. Oring. 1995. Song in Chipping Sparrows, Spizella passerimr. structure and function. Animal Behaviour 50:1233-1241. B.alsby, T. J. S. 2000. The function of song in Whitethroats Sylvia communis. Bioacoustics 11:17-30. B.alsby. T. J. S. and T. Dabelsteen. 2001. The meaning of song repertoire size and song length to male Whitethroats Sylvia communis. Behavioural Processes 56:75-84. Bradley, H. L. 1948. A life history study of the Indigo Bunting. Jack-Pine Warbler 26:103-1 13. Emlen, S. T. 1972. An experimental analysis of the parameters of bird song eliciting species recognition. Behaviour 41:130-171. Falls, J. B. 1969. Functions of territorial song in the White-throated Sparrow. Pages 207-232 in Bird vocalizations (R. A. Hinde, Editor). Cambridge University Press, Cambridge, United Kingdom. Galeotti, P., N. Saino, R. Sacchi, and A. P. Mdller. 1997. Song correlates with social context, testosterone and body condition in male Barn Swallows. Animal Behaviour 53:687-700. Gil, D. and M. Gahr. 2002. The honesty of bird song: multiple constraints for multiple traits. Trends in Ecology and Evolution 17:133-141. Gil, D., j. a. Graves, and P. J. B. Slater. 1999. Seasonal patterns of singing in the Willow Warbler: evidence against the fertility announcement hypothesis. Animal Behaviour 58:995-1000. Gori lander, K. 1987. Variation in the song rate of the male Pied Elycatcher Ficedula hypoleuca: causes and consequences. Animal Behaviour 35:1037-1043. Hinde, R. A. and E. Steel. 1976. The effect of male song on an oestrogen dependent behavior in the female Canary {Serinus canarius). Hormones and Behavior 7:293-304. Hoi-Leitner, M., H. Nechtelberger, and H. Hoi. 1995. Song rate as a signal for nest site quality in Blackcaps (Sylvia atricapilla). Behavioral Ecology and Sociobi- ology 37:399-405. Krebs, J. R., R. Ashcroft, and M. Webber. 1978. Song repertoires and territory defen,se in the Great Tit. Nature 271:539-542. Kroodsma, D. E. 1976. Reproductive development in a female songbird: differential stimulation by quality of male song. Science 192:574-575. La ttin, C. and G. Ritchi.son. 2009. Intra- and intersexual functions of singing by male Blue Grosbeaks: the role of within-.song variation. Wilson Journal of Ornithol- ogy 121:714-721. Leftwich. C. and G. Ritchi.son. 2008. Singing behavior of male Henslow's Sparrows (Ammoclramus lienslowii): description and causation. Bird Behavior 18:1-7. Leitao. a.. C. T. Cate, and K. Riebei,. 2006. Within-.song complexity in a .songbird is meaningful to both male and female receivers. Animal Behaviour 71:1289- 1 296. MacDou(;ai.-Shackleton, S. A. 1997. Sexual selection and the evolution of song repertoires. Current Ornithology 14:81-124. Margoliash, D., C. Staicer, and S. A. Inoue. 1994. The process of syllable acquisition in adult Indigo Buntings (Passerine! cyanea). Behaviour 131:39-64. McGregor, P. K. and A. G. Horn. 1992. Strophe length and response to playback in Great Tits. Animal Behaviour 43:667-676. Moller, a. P. 1988. Spatial and temporal distribution of song in the Yellowhammer Emheriza citronella. Ethology 78:321-331. Morton, E. S. and K. Young. 1986. A previously undescribed method of song matching in a species with a single song "type”, the Kentucky Warbler (Oporornis formosus). Ethology 73:334-342. Nowicki, S., W. a. Searcy, and M. Hughes. 1998. The teiTitory defense function of song in Song Sparrows: a test with the speaker occupation design. Behaviour 135:615-628. Otter, K. A., B. Chruszcz, and L. Ratcliffe. 1997. Honest advertisement and song output during the dawn chorus of Black-capped Chickadees. Behavioral Ecol- ogy 8:167-173. Payne, R. B. 1982. Ecological consequences of song matching: breeding success and intraspecific song mimicry in Indigo Buntings. Ecology 63:401-41 1. Payne, R. B. 1983. The social context of song mimicry: song-matching dialects in Indigo Buntings (Passerina cyanea). Animal Behaviour 31:788-805. Payne, R. B. 2006. Indigo Bunting (Pas,serina cyanea). The birds of North America. Number 4. Payne, R. B. and D. F. Westneat. 1988. A genetic and behavioral analysis of mate choice and song neigh- borhoods in Indigo Buntings. Evolution 42:935-947. Payne, R. B., L. L. Payne, and S. M. Doehlert. 1988. Biological and cultural success of song memes in Indigo Buntings. Ecology 69:104-1 17. Pyle, P. 1997. Identification guide to North American birds. Part I. Slate Creek Press, Bolinas, California, USA. Reh.steiner, U., H. Geisser, and H. U. Reyer. 1998. Singing and mating success in Water Pipits: one specific song element makes all the difference. Animal Behaviour 55:1471-1481. Ritchison, G. 1995. Characteri.stics, use, and possible functions of the perch songs and chatter calls of male Common Yellowthroats. Condor 97:27-38. SAS iN.STtTUTE. 1999. SAS user's guide: statistics. SAS Institute Inc., Cary, North Carolina, USA. Saundf.r.s, a. a. 1929. Bird song. The University of the State of New York, Albany. USA. Shiovitz, K. a. 1975. The process of species-specific song recognition by the Indigo Bunting, Passerina cyairea, and its relationship to the organization of avian acoustical behavior. Behaviour 55: 128-179. Staicer, C. A. 1996. Acoustical features of .song categories of the Adelaide's Warbler (Denilroica aclelaiclae). Auk 1 13:771-783. .Stoddard, P. K., M. D. Beecher, S. E. Campbell, andC. L. Horning. 1992. Song-type matching in the Song Sparrow. Canadian Journal of Zoology 70:1440-1444. Beckett and Ritehison • SINGING BEHAVIOR OF INDIGO BUNTINGS 665 Thompson, W. L. 1970. Song variation in a population of Indigo Buntings. Auk 87:58-71. Thompson, W. L. 1972. Singing behavior of the Indigo Bunting, Passerina cyanea. Zeitschrift fur Tierpsy- chologie 31:39-59. W.’XSSERMAN, F. E. AND J. A. CiGLIANO. 1991. Soilg output and stimulation of the female in White-throated Sparrows. Behavioral Ecology and Sociobiology 29:55-59. Weary, D. M. and R. E. Lemon. 1988. Evidence against the continuity-versatility relationship in bird song. Animal Behaviour 36:1379-1383. Yasukawa, K. 1981 . Song and territory defense in the Red- winged Blackbird. Auk 98:185-187. The Wilson Journal of Ornithology 122(4):666-673, 2010 DEVELOPMENT OE INCUBATION TEMPERATURE AND BEHAVIOR IN THRUSHES NESTING AT HIGH ALTITUDE MARTIN L. MORTON' - AND MARIA E. PEREYRA' ABSTRACT. — Onset of incubation was studied in three Hermit Thrushes (Catharus guttatus) and one American Robin (Turclus migratorius), all with four-egg clutches, at a high altitude site in the Sierra Nevada, California, USA. Behavior of laying females at the nest was measured from continuous recordings of internal egg temperatures of first-laid eggs. Full nocturnal nest attentiveness began immediately with the first egg. Daytime attentiveness increased steadily during laying as toraging time away from the nest decreased. On-off bouts by tending females in daytime increased in frequency and decreased in duration until the last egg was laid. Time on the nest could not be directly equated to occurrence of incubation because eggs were not uniformly warmed to exceed the temperature threshold required for embryonic development (physiological zero). Incubation began, both day and night, after laying of the second egg. It increased steadily thereafter with percentage of daytime devoted to incubation lagging well behind that of nighttime. Nest attentiveness and incubation temperatures reached maxima about the time of clutch completion and were continued during following days. Received 17 December 2009. Accepted 2 April 2010. Development of nest-attentiveness and incuba- tion behaviors in female birds is a physiological- ly-driven process that begins during pre-laying and laying periods as females spend progressively greater amounts of time on the nest, and less time in foraging or other outside activities. The actual onset of the incubation phase, in which eggs reach temperatures sufficiently high to sustain embry- onic development, corresponds with development of a functional brood patch and sustained application of heat to the eggs. The rate at which these behaviors develop is influenced by underlying physiological mecha- nisms associated with gonadal maturation, follic- ular development and ovulation, and subsequent morphological changes with the appearance of a functional incubation patch. Subsequent behaviors surrounding the start of a clutch, although in large part driven by gonadal development initially stimulated by lengthening days (Wingfield 1984, Sharp et al. 1998, Dawson 2008), can be modulated by environmental conditions including availability and suitability of nest sites (Morton 1978, 2002), food supply (Hahn et al. 1995) and, in montane settings, storms (Morton 2002, Hahn et al. 2004). Once egg laying begins, the transition from laying to incubation progresses in an orderly fashion under the influence of a changing hormonal environment that eventually becomes dominated by prolactin (Vleck 1998). The development of incubation behavior marks ' Department of Biological Sciences, University of Tulsa, TuLsa, OK 74104, USA. ^Corresponding author; e-mail: marty-morton@utulsa.edu the transition between pre-nesting and maternal phases of reproduction, and is usually considered to occur gradually (Haftom 1978, 1981, 1988; Zerba and Morton 1983; Oppenheimer and Morton 2000; Wang and Weathers 2009). Be- cause embryonic development does not occur until eggs are warmed above a threshold of 25- 27° C (White and Kinney 1974), termed the physiological zero temperature (PZT), the amount of time the female applies heat during laying, and prior to completion of a clutch, becomes a critical determinant of embryonic growth, the length of the incubation period, and subsequent amounts of hatching asynchrony within a clutch (Mead and Morton 1985). Much importance has been at- tached to the age-size hierarchy of hatchlings and many hypotheses have been advanced to explain its assumed adaptive value (Clark and Wilson 1985; Stoleson and Beissinger 1995, 1999). The widespread occurrence of hatching asynchrony and potential for maternal shifts in timing of incubation onset to exaggerate or suppress age- size hierarchies in hatchlings makes accurate measurements of events that accompany shifts in nest attentiveness between laying and incubation particularly important. We demonstrate using egg temperature (Tegg) data from two species of thrush (Turdidae) that, although daytime attentiveness is frequently used in ecological studies to mark the start of the incubation period (and hence embryonic develop- ment), the.se are not equivalent conditions. In particular, we show the behavioral (daytime and night-time attentiveness) and physiological com- ponents of incubation (heat application) are decoupled early in laying, and only gradually 666 Morton and Pcrevra • ONSET OF INCUBATION IN THRUSHES 667 converge into actual incubation, as defined by the maintenance of eggs at temperatures above the threshold for development. METHODS Nests from three Hermit Thrushes {Cathariis guttatiis) and one American Robin {Turdus migratohits) were used in a study conducted at Tioga Pass (37.8° N, 1 19.2° W; 3,000 m asl), in the Sierra Nevada, California, USA. All nests were in lodgepole pine (Piniis contorta). We inserted the tip of a 36-ga copper-constantan thermocouple into the center of the first egg within a few hours after it was laid, then glued the insulated leads to the shell. Lead wires were threaded through the nest bottom and connected to a hidden strip-chart recorder set to move at 2 cm/ hr and calibrated to record between —5 and 45° C over a chart width of 10 cm. We moved well away from the nest during thermocouple installation (about 30 min), but left behind a dummy egg of the size and coloration of the thrush egg in case the laying female returned. Air temperature (Ta) was measured with a hygrothermograph placed in a vented box, in the shade, near the base of the nest tree. Temperature data for both Tegg and Ta were collected during the first 10 days of June in 1982 or 1985. Each nest was visited daily for quick inspections of nest contents and of the recording equipment. RESULTS All four thrushes laid three additional eggs on successive days after the first egg was implanted. Tegg data indicated that none of the birds returned to their nests on the first day (Day 1 ) until after sunset (Fig. 1). They remained on the nest throughout the night (~8 hrs) then left, presum- ably to forage, before returning to lay their second egg. One individual (Fig. lA), for example, first returned to her nest on Day 1 at 2001 hrs PST, remained there for the entire night and left at 0356 hrs to begin Day 2. She was away from the nest for 268 min then began an attentive period at 0824 hrs. She laid the second egg during the next 168 min and made three on-off bouts before leaving the nest until 1945 hrs when night-time attentiveness was resumed. Behavior at the nest was much the same on Day 3 except the daytime attentive period began earlier (0754 hrs) and lasted longer (366 min). Continuous daytime attentiveness by the female on Day 4 began at 0936 hrs whereupon she laid her fourth egg, but left the nest thereafter only for relatively brief periods of foraging; full time attentiveness, both day and night had begun. Our method did not permit us to record time spent on the nest by females laying their first egg, but it can be deduced (Fig. 1) that daytime attentiveness in- creased with laying order. The same pattern of long periods away from the nest on laying days (Fig. lA), also occurred in the other two Hermit Thrushes (Fig. IB, C) except that one female (B) omitted the lengthy afternoon period of foraging a day earlier. The American Robin exhibited an attentiveness pattern during laying that resembled Hermit Thrushes. She was on the nest full time at night after the first egg was laid, but did not continue with an extensive late afternoon absence after Day 2 or in the early morning after Day 3 (Fig. ID). Females of both species did not sit on the nest continuously throughout their daytime periods of nest attentiveness (hatched bars), but left period- ically to feed, as indicated by a sawtooth pattern to Tegg tracings. Frequency of these feeding or on-off bouts increased as laying proceeded and stabilized once the birds were in full-time incubation (Table 1). Night-time attentiveness ended before sunrise as indicated in the four mornings shown for each bird (Fig. lA, B, C, D). It ended 48.3 ± 1 1 .5 min {x ± SD) before sunrise for the three Hermit Thrushes, and 27.2 ± 5.2 min for the American Robin. Night-time attentiveness resumed after sunset for the five nights shown. The mean was 36.2 ± 8.5 min after sunset for Hermit Thrushes and 44.2 ± 5.2 min for the American Robin. A Hermit Thrush, on the day of her first egg, settled for the night at 1824 hrs, nearly an hour before sunset (Fig. 1C, Day 1) when a storm with hail and cold rain occurred. Females were on their nests during the first night after laying had begun, but chart recordings of Tegg indicated they were applying only enough heat to keep the egg between Ta and the threshold for development (Fig. 2). There was no daytime nest attentiveness following egg laying on the first day. Time spent on the nest increased as laying proceeded (shown for 2 days later. Fig. 3A) and females began to exhibit rhythmic on-off bouts during which time eggs were at times wanned above PZT. Neglect of eggs still occurred for considerable periods in daytime. For example. Day 3 (Fig. 3A) began with the female’s depar- ture at 0348 hrs (arrow 1 ) and during the 668 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 4. December 2010 Hermit Thrush 0 3 6 9 12 15 18 21 24 Hour of day American Robin 0 3 6 9 12 15 18 21 24 Hour of day FIG. I. Representation of nest attentiveness at Tioga Pass, California in four thrushes. All four laid an egg in the morning hours of Days I through 4. One day following clutch completion (Day 5) is also shown. Black bars indicate when females were tending their nests at night, white bars when they were absent during daytime, hatched bars when they were tending their nests in daytime which includes on-off foraging bouts. Time of thermocouple installation occurred on Day 1 at time X. SR marks time of sunrise, SS time of sunset (PST). subsequent 170 min, during the coldest part of the day, Tegg dropped to ambient and followed it until she returned to the nest (arrow 2) and began re-warming the eggs. This period of daytime nest attentiveness continued for 366 min whereupon the female departed (arrow 3) for 418 min before coming back on for the night (arrow 4). Thus, there were two daytime periods off the nest, one TABLE I . Number and duration of on-off bouts during periods of daytime nest attentiveness by three Hermit Thrushes and one American Robin. Eggs were laid by both species on Days 1 through 4. Number of on-off bout.s/day Duration of on ■off bouts/day (min) Day Hermit Thrush Mean ± SD American Robin Total Hermit Thrush Mean ± SD American Robin Mean ± SD 1 2 F' 3.7 ± 2.9 p 7 23.2 ± 13.7 65.8 ± 52.3 3 7.3 ± 4.0 12 24.6 ± 22.2 37.7 ± 49.6 4 15.3 ± 1.5 17 10.2 ± 5.4 14.1 ± 5.5 5 21.0 ± 1.7 17 10.5 ± 4.2 13.8 ± 4.6 6 19.7 ± 0.6 16 1 1.2 ± 8.3 13.5 ± 5.6 7 21.0 ± 1.7 16 1 1.0 ± 5.3 13.2 ± 2.8 ■' Assiimcil from observations of American Robins by Schant/. (iy.f9). Morton and Percvra • ONSET OF INCUBATION IN THRUSHES 669 Hour of day Hour of day FIG. 2. Egg temperatures (Tegg) in two nests (A = Hermit Thrush; B = American Robin) on the first night following onset of laying (Day 1). Arrow 1 shows when night-time attentiveness began, and arrow 2 when it ended. Concurrent record of air temperatures (Ta) also shown. Shaded horizontal bars indicate physiological zero temperature (PZT). in the morning and one in the afternoon (white bars, Fig. I ) when the female was, presumably, mainly engaged in foraging. An example of this same female’s attentiveness pattern is shown 4 days later (Fig. 3B) when she was in full incubation mode. She left the nest in daytime (beginning at arrow 1 ) only for foraging bouts and remained on the nest at mid-day for extended periods when there was exposure to sun. Eventually, well after sun.set, she began the period of night-time attentiveness (arrow 2). Tegg seldom dropped below PZT at this time; embryos were usually sufficiently warm for growth and development to occur. The percentage of total daytime and nighttime that eggs were heated above PZT increased gradually, but daytime incubation lagged behind that of nighttime (Fig 4). Night-time nest occu- pancy time was 100% from the first egg onward FIG. 3. Tegg and Ta obtained at a Hermit Thrush nest on the day the third egg was laid (A) and 4 days later when regular, full-time incubation was occurring (B). Shaded horizontal bars indicate PZT. See text for explanation of arrows. and daytime occupancy was not (Fig. 1 ). Incuba- tion onset also occurred more rapidly at night. For example, during the night that Hermit Thrush A was on her first egg it was not warmed sufficiently for development to occur (Fig. 2A). Tegg during the next night, after the second egg had been laid, exceeded PZT 70.8% of the time and on the following night it was 98.7% (Fig. 4A). In contrast, in the daytime hours during which the second egg was laid, PZT was exceeded for only 21 ol the 168 min (12.5%) the female was on the nest, or 2.2% of the total daytime period (Fig. 4A). PZT was exceeded in the daytime hours, during which the third egg was laid, for 126 ot the 366 min (34.4%) that a female was on the ne.st (Fig. 3A), or 13.1% of the total daytime period (Fig. 4A). Time that eggs were held above PZT was at or near 100% by the third night for Hermit Thrushes, but daytime heat tran.sfer to eggs was not maximized at 92 to 98% until the day following clutch completion (Fig. 4A, B, C). The American Robin followed a night-time pattern similar to that of Hermit Thrushes, but reached the daytime maximum above PZT (98 to 100%) a day ahead of them (Fig. 4D). THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4. December 2010 670 FIG. 4. Percent of time that Tegg was above PZT during nighttime and during daytime for three Hermit Thrushes (A, B, C) and an American Robin (D). Egg laying occurred on Days 1 through 4. Hermit Thrush females were attentive to their nests about 38% of the time on Day I (1 hr of daytime for laying and 8 hrs of nighttime in the first 24 hrs). This increased steadily on each laying day thereafter as more time became devoted to the nest during daylight hours (Fig. 5). In contrast, no incubation occurred on Day 1, but it began to increase thereafter with eggs being heated in the beginning mostly at night (Fig. 4) then more so in daylight, summing eventually to 80% on Day 4 before reaching a sustainable maximum of ~95% on Day 5 (Fig. 5). The American Robin also increa.sed its incubation time with laying, reaching 98% on Day 4. Frequency of on-off bouts increa.sed during laying until Day 5 for Hermit Thrushes and Day 4 for the American Robin (Table I). The slightly higher percentage of time eggs were held above PZT by the American Robin occurred because, unlike Hermit Thrushes (Fig. 3B), its eggs seldom cooled to below 28° C when it was off feeding during the cold hours of early morning. DISCUSSION Nest-attentive behavior in thrushes differed greatly with time of day. Nocturnal occupancy of nests began with the first egg and was fully FIG. .“i. Percent of time that Hermit Thrush females [n = 3) were tending their nests (filled circles) and warming their eggs above PZT (open circles). Egg laying occurred on Days 1 through 4. Morton and Pereyra • ONSET OF INCUBATION IN THRUSHES 671 sustained thereafter (Fig. 1). Diurnal attentive- ness, in eontrast, developed only gradually and was not complete until after the final egg was laid (Fig. 1). Daytime episodes of attentiveness were intemipted by short periods off the nest for foraging (on-off bouts). These increased in frequency and decreased in duration as laying proceeded and became stable at time of clutch completion, a behavioral marker of entry into full- time incubation (Table 1 ). Aside from those brief forays off the nest there were other, much longer portions of the day, that were presumably devoted mainly to foraging. These lengthy periods of nest neglect occupied more than 90% of daytime hours on the first-egg day (Day 1), decreasing to about 15% on the last-egg day (Day 4) for Hermit Thrushes and to 0% for the American Robin. We hypothesize that females required these long feeding periods to meet the nutritional demands of ovogenesis and that these demands were not trivial. For instance, data on egg weights and body weights of thrushes indicate that clutch mass may exceed one-third of a non-laying female’s body mass (Jones and Donovan 1996, Sallabanks and James 1999). Mountain White-crowned Sparrows (Zonotrichia leucophrys oriantha) on this same study area, had a four-egg clutch mass about equal to 45% of a female’s body mass, and resources for ovogenesis were accumulated before and during laying. Body mass increased by about 5% per day in the 3 days prior to first ovulation and decreased only slightly each day thereafter with each new egg. This is about twice the rate of mass gained in females (and males) engaged in autumnal pre- migratory fattening in this same population, indicating that laying females are intensely hyperphagic (Morton 2002). A trade-off between daytime nest neglect and extended periods for foraging by laying thrushes seems necessary, but what could be the value of their early nocturnal attentiveness? This question was examined for Red-winged Blackbirds (Age- laius phoeniceus) by Clotfelter and Yasukawa (1999). They concluded early nest attentiveness did not provide increased egg viability or protection from predators, but may have reduced parasitism by Brown-headed Cowbirds (Molo- thrus ater). Cowbirds were rarely present at Tioga Pass and none of their eggs was detected in 92 Hermit Thrush and 30 American Robin clutches (Morton and Pereyra 2004). Our visits to thrush nests were too infrequent to allow comment on timing of predation (nocturnal or diurnal), but exposure to freezing conditions at night or early morning, as well as .severe storms, was known to cause mortality in eggs and nestlings of other species (Morton 2002). Nocturnal attentiveness might preserve egg viability, at least under high altitude conditions. We found on our study area, however, that interspecific variation occurred. Nocturnal attentiveness by four species (Rock Wrens [Salpinctes ohsoletus]. Song Sparrows [Melospiza melodia]. Hermit Thrushes, and American Robins) began early in the laying sequence (Oppenheimer and Morton 2000, Morton 2002, this study), whereas in two others (Dusky Flycatchers [Empidonax oberholseri]. White-crowned Sparrows) it was delayed until after the penultimate egg was laid (Zerba and Morton 1983, Morton and Pereyra 1985). Early onset of nocturnal attentiveness occurs in a wide range of passerines, including many cavity nesters (Wang and Weathers 2009), but its adaptive significance seems unresolved. A key finding of our study was that the pattern of heat transfer to eggs sufficient for embryonic development differed from that of nest attentive- ness. For example, all four thrushes began night- time sitting after the first egg (Fig. 1), but none supplied sufficient warmth for incubation to begin until after the second egg (Figs. 2, 4). Hermit Thrushes were attentive for more than half of the daylight hours by Day 3 (Fig. 4A, B, C), but PZT was usually exceeded only briefly during periods that the female was on the nest (Fig. 3A). Mean fraction of daytime where Tegg was sufficiently high for embryonic development on Day 3 was only 23.7 ± 9.5% for Hermit Thrushes and 38. 1 % for the American Robin (Fig. 4). Females were developing the behavioral rhythm of incubation but full contact of eggs with the incubation patch in daytime was withheld because, during the night-time portion of Day 3, PZT was exceeded by all four thrushes at close to 100% of the time (Fig. 4). This indicates the Day 3 patch was functionally developed but incubation behavior was not with daytime incubation lagging behind nighttime. Total nest attentiveness behavior and of incu- bation (daytime plus nighttime) for the 4 days of laying and for 3 days thereafter increased gradually during laying, but with different trajec- tories (Fig. 5). Hermit Thrushes foraged for at least 2 hrs in early morning before returning to lay their final egg (Fig. 1), and then, immediately thereafter, began full-time incubation. Apparently, 672 THE WILSON JOURNAL OF ORNITHOLOGY • V'oL 122, No. 4, December 2010 the American Robin dealt more easily with nutritional demands because her prolonged early morning feeding periods ended on Day 3 (Fig. ID), and incubation behavior was fully developed by Day 4 (Fig. 4D). Incubation onset in European Starlings {Stunnis vulgaris) appeared to be pre-programmed (Meijer 1990), a conclusion supported by our study. For example, day-night timing of heat transfer to eggs during the laying period was much the same for all three Hermit Thrushes and their close relative, the American Robin (Fig. 4). Incubation devel- oped gradually during the laying period in all four birds with full behavioral and physiological capacities converging at the time of clutch completion (Fig. 5). These results agree with those obtained for other species of passerines inhabiting different environments (Wang and Weathers 2009), and with early field observations of Hermit Thrushes (Stanwood 1910) and Amer- ican Robins (Howell 1942). We suggest this consistency is due to strongly expressed neuroen- docrine and endocrine influences (Goldsmith 1982, Sharp 1997, Shaip et al. 1998, Sharp and Blache 2003, Vleck 1998) that have been tuned by natural selection to entrain incubation beha- viors that optimize embryonic development and survival. Results quite different from ours have been published. For example, incubation onset by European Pied Flycatchers (Ficechda hypoleuca) occurred at any time within a 4-day span, from the day of clutch completion to 3 days prior (Potti 1998). Even more plasticity was reported in a recent, large study of American Robins wherein females were presumed to begin incubating at any time within a 7-day span, from 4 days after clutch completion to 2 days prior (Rowe and Weather- head 2009). Accurate information on incubation onset in both of these studies was crucial to the hypotheses presented, yet this key event was defined as the date when eggs first felt warm to the touch. Obviously, according to this criterion, birds could be classified as “incubating” when eggs were under partial incubation, or full-time incubation, or even no incubation if Tegg was below PZT. We hope that more complete information, as provided by egg temperatures, will be obtained before data from studies are incorporated into current paradigms of timing and control of incubation onset. Studies of incubation have frequently suffered from inconsistencies in terminology and effec- tiveness of methods, and there has been a failure of investigators to use knowledge about underly- ing mechanisms. This lack of physiological insight is unfortunate because it can be an impediment to understanding ecological effects (Denny and Helmuth 2009) and even the proxi- mate causation of behavioral phenomena (Drick- amer 1998); avian reproduction has model-system potential for collaborations between evolutionary ecologists and endocrinologists (Wingfield et al. 2008). We believe the adaptive value of early incubation onset probably does not lie in hatching asynchrony (Mead and Morton 1985, Morton 2002). It has been shown, for example, that partial incubation before clutch completion can help maintain egg viability (Veiga 1992, Stoleson and Beissinger 1999), and reduce the magnitude of microbial trans-shell infections (Cook et al. 2003). ACKNOWLEDGMENTS This research was done with the permission of the LSDA, Forest Service and the California Department of Fish and Game. LITERATURE CITED Clark, A. B. and D. S. Wilson. 1985. The onset of incubation in birds. American Naturalist 125:603-61 1. Clotfelter, E. D. and K. Yasukawa. 1999. The function of early onset of nocturnal incubation in Red-winged Blackbirds. Auk 116:417^26. Cook, M. I.. S. R. Beissinger, G. A. Toranzos, R. A. Rodriguez, and W. J. Arendt. 2003. Trans-shell infection by pathogenic micro-organisms reduces the shelf life of non-incubated bird's eggs: a constraint on the onset of incubation? Proceedings of the Royal Society of London, Series B 270:2233-2240. Dawson, A. 2008. Control of the annual cycle in birds: endocrine constraints and plasticity in response to ecological variability. Philosophical Transactions of the Royal Society of London, Series B 363:1621- 1633. Denny, M. and B. FIelmuth. 2009. Confronting the physiological bottleneck: a challenge from ecomecha- nics. Integrative and Comparative Biology 49:197- 201. Drickamhr, L. C. 1998. Vertebrate behavior: integration of proximate and ultimate causation. American Zoologist 38:.39-42. Goldsmith, A. R. 1982. Plasma concentrations of prolactin during incubation and parental feeding throughout repeated breeding cycles in Canaries (Seriniis canar- ii(.s). journal of Endocrinology 94:51-59. Haflorn, S. 1978. Egg-laying and regulation of egg temperature during incubation in the Goldcrest Regiilus regains. Ornis Scandinavica 9:2-21. Hafi'orn, S. 1981. Incubation during the egg-laying period in relation to clutch-size and other aspects of Morton and Pcrcvra • ONSET OF INCUBATION IN THRUSHES 673 reproduclion in the Great Tit Parns major. Ornis Scandanavica 12:1 69- 1 85. Hafforn, S. 1988. Incubating female passerines do not let the egg temperature fall below the ‘physiological zero temperature' during their absences from the nest. Ornis Scandanavica 19:97-110. Hahn. T. P., J. C. Wingfield, R. Mullen, and P. J. Deviche. 1995. Endocrine bases of spatial and temporal opportunism in Arctic-breeding birds. Amer- ican Zoologist 35:259-273. Hahn, T. P.. K, W. Sockm.an, C. W. Breuner, and M. L. Morton. 2004. Facultative altitudinal movements by Mountain White-crowned Sparrows {Zonotrichia leii- coplnys orianthu) in the Sierra Nevada. Auk 121: 1269-1281. Howell, J. C. 1942. Notes on the nesting habits of the American Robin (Tiirdiis migratioriiis L.). American Midland Naturalist 28:529-603. Jones, P. W. and T. M. Donovan. 1996. Hermit Thrush {Catharu.s guttatus). The birds of North America. Number 261. Mead, P. S. and M. L. Morton. 1985. Hatching asynchrony in the Mountain White-crowned Sparrow (Zonotrichia leucophrys oriantha): a selected or incidental trait? Auk 102:781-792. Meijer, T. 1990. Incubation development and clutch size in the starling. Ornis Scandanavica 21:163-168. Morton, M. L. 1978. Snow conditions and the onset of breeding in the Mountain White-crowned Sparrow. Condor 80:285-289. Morton, M. L. 2002. The Mountain White-crowned Sparrow: migration and reproduction at high altitude. Studies in Avian Biology 24:1-236. Morton, M. L. and M. E. Pereyra. 1985. The regulation of egg temperatures and attentiveness patterns in the Dusky Flycatcher (Empidona.x oherhol.xeri). Auk 102: 25-37. Morton, M. L. and M. E. Pereyra. 2004. Clutch sizes and nesting habits of birds at Tioga Pass. Western Birds 35:62-70. Oppenheimer. S. D. and M. L. Morton. 2000. Nesting habits and incubation behavior of the Rock Wren. Journal of Field Ornithology 71:650-657. Pom, J. 1998. Variation in the onset of incubation in the Pied Flycatcher (Ficedida hypolenca): fitness conse- quences and constraints. Journal of Zoology, London 245:335-344. Rowe, K. M. C. and P. J. Weatherhead. 2009. A third incubation tactic: delayed incubation by American Robins (Tnrdns migratoriii.s). Auk 126:141-146. Sallabanks. R. and F. C. James. 1999. American Robin (Tnrdns migratorins). The birds of North America. Number 462. SCHANTZ, W. E. 1939. A detailed study of a family of robins. Wilson Bulletin 51:157-169. Sharp, P. J. 1997. Neurobiology of the on.set of incubation behaviour in birds. Pages 193-202 in Frontiers in environmental and metabolic endocrinology (S. K. Maitra, Editor). University of Burdwan. Burdwan, India. Sharp, P. J. and D. Blache. 2003. A neuroendocrine model for prolactin as the key mediator of seasonal breeding in birds under long- and short-day photope- riods. Canadian Journal of Physiology and Pharma- cology 81:350-358. Sharp, P. J., A. Dawson, and R. W. Lea. 1998. Control of luteinizing hormone and prolactin secretion in birds. Comparative Biochemistry and Physiology Part C 1 19: 275-282. Stanwood, C. j. 1910. The Hermit Thrush; the voice of the northern woods. Bird-Lore 12:100-103. Stoleson, S. H. and S. R. Beissinger. 1995. Hatching asynchrony and the onset of incubation in birds, revisited. CuiTent Ornithology 12:191-270. Stoleson, S. H. and S. R. Beissinger. 1999. Egg viability as a constraint on hatching synchrony at high ambient temperature. Journal of Animal Ecology 68:951-962. Veiga, j. P. 1992. Hatching asynchrony in the House Sparrow: a test of the egg-viability hypothesis. American Naturalist 139:669-675. Vleck, C. M. 1998. Hormonal control of incubation/ brooding behavior: lessons from wild birds. Proceed- ings of the WPSA European Poultry Conference. Israel 10:163-169. Wang, J. M. and W. W. Weathers. 2009. Egg laying, egg temperature, attentiveness, and incubation in the Western Bluebird. Wilson Journal of Ornithology 121:512-520. White, F. N. and J. L. Kinney. 1974. Avian incubation. Science 186:107-1 15. Wingfield, J. C. 1984. Environmental and endocrine control of reproduction in the Song Sparrow. Melos- piza melodia. General and Comparative Endocrinolo- gy 56:406^16. Wingfield, J. C., M. E. Visser, and T. D. Williams. 2008. Introduction. Integration of ecology and endo- crinology in avian reproduction: a new synthesis. Philosophical Transactions of the Royal Society of London, Series B 363:1581-1588. Zerba, E. and M. L. Morton. 1983. The rhythm of incubation from egg laying to hatching in Mountain White-crowned Sparrows. Ornis Scandanavica 14: 188-197. The Wilson Journal oj Ornithology 122(4):674-680, 2010 FLEXIBILITY IN NEST-SITE CHOICE AND NESTING SUCCESS OF TURDUS RUFIVENTRIS (TURDIDAE) IN A MONTANE FOREST IN NORTHWESTERN ARGENTINA SILVIA B. LOMASCOLO,' " A. CAROLINA MONMANY,--^ AGUSTINA MALIZIA,^ AND THOMAS E. MARTIN^ ABSTRACT. — We studied the consequences of nest-site choice on nesting success under differing disturbance levels for the Rufous-bellied Thrush (Turdus rufiventris). We compared nest-site choice and nest success between a disturbed site and an undisturbed site in a montane subtropical forest in northwestern Argentina. We found no overall difference in daily predation rate (DPR) between the disturbed and undisturbed sites. However, DPR of nests on bromeliads was significantly lower at the microhabitat level than on other types of subtrates at the disturbed site. T. rufiventris used bromeliads for nesting more often than expected by chance at the disturbed site. DPR did not differ between substrates at the undisturbed site and T. rufiventris used all substrates according to their availability. Nests had higher predation at the disturbed site when DPR on non-bromeliad substrates was compared between disturbed and undisturbed sites. Nest fate was independent of nest height. Our results suggest T. rufiventris' flexibility in nest-site choice, as reflected by increased use of the safest sites, i.e., bromeliads, in the disturbed site compared to the undisturbed site, may allow this species to survive in an otherwise much riskier habitat. Our results illustrate how microhabitat-scale effects can mediate landscape scale effects. Received 24 October 2009. Accepted 7 April 2010. Understanding habitat influences on nesting success of birds may be key to their successful conservation, given the sensitivity of this life stage to habitat disturbance (Martin 1992, Easton and Martin 2002). Nest success is influenced by nest-site choice in large part because nest-site characteristics can influence nest predation rates (Martin and Roper 1988, Martin 1993, Holt and Martin 1997, Martin 1998, De Santo et al. 2002, Easton and Martin 2002, Mezquida and Marone 2002, Kellett et al. 2003, Fontaine et al. 2007). Nest predation may be one of the main agents of natural selection influencing evolution of life history traits and nest-site choice (reviewed by Lima 2009, Martin and Briskie 2009). Flexibility in choosing a nest site may allow birds to optimize fitness by exploiting safer substrate types in disturbed conditions, but studies of nesting flexibility and their consequences for nest success are rare. Nest-site choice, including nesting substrate and nesting height, may be evolutionarily conser- ' lADIZA-CCT MENDOZA. CONICET, Avenida Ruiz Leal .s/n, CC 507, 5500 Mendoza, Argentina. ^ Univer.sidad de Puerto Rico, Departamento de Biologi'a, CN 235, P. O. Box 70377, San Juan, PR 00936, USA. ’LIEY-IER, Univer.sidad Nacional de Tucuman, Argen- tina, CC 34, Yerba Buena, 4107 Tucuman, Argentina. ‘‘uses, Montana Cooperative Wildlife Research Unit, Avian Studies Program, 205 Natural Science, University of Montana. Missoula, MT 59812, USA. ^Corresponding author; e-mail: .slomascolo(§>gmail.com vative in many species (Martin 1988, Martin and Roper 1988), which may constrain plasticity of choices. Some bird species use the same nest sites in disturbed and undisturbed forest patches, even though this increases predation in disturbed forest patches (Holt and Martin 1997, Easton and Martin 2002). Plastic changes in nest-site selection in response to predation risk have been observed in some species (e.g., Marzluff 1988, Eggers et al. 2006, Peluc et al. 2008). Variation in predation risk is often associated with disturbance of habitats, which also generally reflects differing habitat structure; both differing predation risk and habitat structure could influence nest-site choice (Martin 1992). Few studies have compared nest site choices under differing predation risk or differing habitat structure. Plasticity of nest-site choice is particularly interesting for tropical and subtropical birds, because they are often thought to have more specialized niches (MacArthur 1972). Conse- quently, their flexibility of nest-site choice to varying predation risk or habitat structure may be constrained. No study has examined plasticity of nest-site choice by a tropical or subtropical bird. Thus, examination of flexibility of nest-site choice and consequences of that choice on nesting success under differing disturbance levels for tropical or subtropical birds is needed. Our objective was to test whether nest-site choice and nesting success of the Rufous-bellied Thrush {Turdus rufiventris', Turdidae) differed 674 Lomciscolo cl al. • RUFOUS-BELLIED THRUSH IN MONTANE l-ORESI'S 675 between a disturbed site and an undisturbed site in a montane subtropical forest in northwestern Argentina. Tiirdiis rufiventris is a common bird that consumes fruits of many species in this environment (Malizia 2001 ), potentially having an important role in regeneration of native plants. It is widely distributed in central and northern Argentina (Ridgely and Tudor 1989, de la Pena and Rumboll 1998), and is one of the few native forest species that inhabits semi-urban locations (de la Pena and Rumboll 1998, Ferretti et al. 2005). The ability of members of the genus Tiirdus, including T. rufiventris (e.g., de la Pena and Rumboll 1998, Ferretti et al. 2005), to live in human-disturbed areas makes them good model species to study how birds adapt to habitat disturbance. Understanding the adaptability of species to disturbance is becoming increasingly important as anthropogenic habitat alterations become increasingly common and intact habitats increasingly rare (Pimm et al. 2001). Examining the effect of habitat characteristics in nest-site choice flexibility and nesting success is important in a threatened but understudied environment such as the subtropics (de la Pena 1979, Mezquida and Marone 2002). We examined nest-site preferences and consequences for nesting success of T. rufiventris between undisturbed and disturbed forest sites in northwestern Argentina. METHODS The study was conducted between October 1997 and January 1998 at El Rey National Park (hereafter El Rey; 24°42'S, 64°38'W) in Salta Province, and at Sierra de San Javier Biological Park (hereafter San Javier; 24°47'S, 65°22'W) in Tucuman Province. Both sites are in north- western Argentina and are part of the Yungas ecosystem represented by a subtropical montane forest (Brown 1995, Brown et al. 2001). An old cattle farm, El Rey was declared a national park in 1948 and has been pre.served intact with no human activity since then. It encompasses 44,162 ha, is 200 km from the nearest city, and no significant logging has been conducted in the area; it represents the undisturbed site. The canopy is dominated by Cinnamonuun porphyrium (Laur- aceae), Blepharocalix salicifolius (Myrtaceae), Cedrella lilloi, and C. angustifoUa (Meliaceae). A second stratum is comprised of species typically <20 m high, including Allophyllus edulis (Sapindaceae), Zanthoxylum coco (Ruta- ceae), and Primus tucunianensis (Rosaceae) (Blake and Rouges 1997). San Javier was created by the National University of Tucuman in 1973 and covers 14,100 ha; it is 15 km west of the city of San Miguel de Tucuman and 2 km northwe.st of the city of Yerba Buena. The area has been subject to different land u.ses including .selective logging focused on the most valuable timber species (Cedrela lilloi) affecting a large propor- tion of the sieiTa, modern agriculture (sugar cane, citrus, horticulture, floriculture), and expanding urbanization (Grau et al. 2008). Trekking and biking trails cross the San Javier site. The secondary forest where we sampled occurs near agricultural and urban sectors (Grau et al. 1997, Aragon and Morales 2003), and represents our disturbed site. Common canopy species at San Javier are Parapiptadenia exelsa (Fabaceae), Cinnamomurn porphyrium (Lauracae), Juglans australis (Juglandaceae), and Myrsine laetevirens (Myrsinaceae) while the subcanopy is dominated by Piper tucumanum (Piperaceae), Allophylus edulis (Sapindaceae), and Psychotria cartagenen- sis (Rubiaceae) (Grau et al. in press). The native pioneers Heliocarpus popayanensis (Malvaceae), Tecoma stans (Bignoniaceae), Solanum ripariwn (Solanaceae), and exotic colonizers including Morus spp. (Moraceae), Ligustrum luciduni (Oleaceae), and Citrus spp. (Rutaceae) (Grau and Aragon 2000) are also common at San Javier. Bromeliads are one of the main groups represent- ed among the abundant epiphytes both at San Javier and at El Rey (Blake and Rouges 1997). We located nests following Martin and Geupel (1993). This method involves detecting and following birds canying food or nesting material to the nest, or following female calls. Nest searching was done daily from 0630 to 1245 hrs from October to December. Nests were checked every 2-4 days to record nest stage (e.g., building, incubating, nestlings), number of eggs or nestlings when nests were accessible, nesting substrate, and nest height (Martin and Geupel 1993). The nest was considered failed or fledged if, after three checks, no bird activity was recorded at the nest (the last check had to be at least 30 min), depending on the stage recorded at the last active visit. We were unable to assign the cause of failure of many nests due to inaccessibility, but fledging was confirmed whenever possible by looking for parents with food or fledglings near the nest shortly after the assumed fledge date. We calculated nest success following Mayfield (1961 ) using the approach outlined by Hensler and 676 THE WILSON JOURNAL OL ORNITHOLOGY • Voi 122. No. 4. December 2010 Nichols (1981). We compared daily predation rates between sites and between substrates using a Chi-square test based on program CONTRAST (Hines and Sauer 1989). The two most common substrates used at San Javier, bromeliads and Psychotria ccirtagenensis, seemed to differ substantially in mean height, and the differences in daily predation rate found between the substrates could be attributed to substrate height and not to type of substrate per se. Thus, we tested whether the difference in substrate height was statistically significant using a Mann-Whitney f/-test, and whether predation was independent of substrate height in San Javier using a Chi-square test. We measured substrate availability based on the observed substrates used by T. rufiventris to examine if nests were randomly placed. We followed the quarter method (Matteuci and Colma 1982), modified to include bromeliad availability. We established two 200 m-long transects at random in the study area. The starting point of the transect was at the end of a random number of steps chosen by a person not familiar with the study site. A second random number indicated the orientation of the transect as measured by a compass. We established four quadrants every 10 m along each of these transects. We projected an imaginary line 17-m high (the maximum nest height recorded during this study) from the center point, and recorded the plant species closest to the imaginary line in each quadrant. We recorded when a bromeliad was the closest substrate, but not the substrate supporting the bromeliad. Only plants with >1.5 cm diameter at breast height (dbh) were recorded because no nest was found on plants with a smaller dbh. At least 160 plants were counted along each tran.sect. We analyzed these data using a Z-test for comparing two proportions (Zar 1999) to examine if nest site was chosen according to plant availability. RESULTS We found 44 nests of Rufous-bellied Thrushes on 12 different substrates (Fig. I A) at San Javier of which only 27 had at least one egg laid. Only tho.se 27 nests were u.sed to calculate nest success. All 44 nests at San Javier were included in the substrate use analyses. Sixty-five nests with at least one egg laid were found at El Rey on 19 different substrates (Table I). The.se nests were u.sed for estimating nest success and substrate u.se. Nests at San Javier were most common on Psychotria carthagenensis and bromeliads. How- ever, bromeliads were used in higher proportion than expected according to their availability at San Javier while the other substrates (including P. carthagenensis) were used in lower proportion than expected (Z = 10.8, df = 2, P < 0.0001; Table 2). Substrates at El Rey were used as expected by their availability (Z = 0.66, df = 2, P = 0.26; Table 2). Overall daily predation rate (DPR ± SE) at San Javier (0.0684 ± 0.0095, n = 27, exposure days = 248.5) did not differ from El Rey (0.0592 ± 0.0089, n = 65, exposure days = 709; X- = 0.669, df = \, P = 0.414). Nest success on P. carthagenensis, the most common understory shrub at San Javier, was not different from success on all substrates other than bromeliads (X“ = 0.5, df = P = 0.48) and substrates were combined for further analysis. Nests on bromeliads at San Javier were more successful (DPR = 0.0308 ± 0.0175 nests/day, n = 7, exposure days = 97.5) than nests on other substrates (DPR = 0.1126 ± 0.0257 nests/day, n — 20, exposure days = 151) (X^ — 6.92, df = 1, P = 0.0085). Success of nests on bromeliads at El Rey (DPR = 0.0522 ± 0.0207 nests/day, n — 9, exposure days = 115) was not significantly different from success on other substrates (DPR = 0.0606 ± 0.0098, n = 56, exposure days = 594) (X" = 0. 17, df = P = 0.68). We believe that lack of difference between substrates in DPR at El Rey was not affected by low sample size given the significant difference found for San Javier, where sample size was even smaller. Bromeliads were used more often at San Javier than at El Rey, and the high nest success on bromeliads may increase overall nest success at San Javier. Thus, we compared nest success between San Javier and El Rey without considering nests on bromeliads. Non-bromeliad nests experi- enced higher daily predation rates at San Javier than at El Rey (X" = 4.076; df = 1 ; P = 0.044). Nests built on P. carthagenensis at San Javier were significantly lower (.T = 2.1 m) than those on bomeliads (x = 8.9 m) (Mann-Whitney U = 10, P = 0.0004). However, nest success was independent of nest height (X" = 0.50, df = 5,' P = 0.48). Nests occurred between 1 and 17 m above ground with a bimodal distribution (Fig. I A). Most nests were between 1 and 2 m above the ground with a second peak of nests at 10 m. The two most common substrates mostly explained the distribution of nests at different heights (Fig. IB). LomascoU) cl al. • RUFOUS-BELLIED THRUSH IN MONTANE I-ORESTS 677 Nest height (m) 14 n Lisi Pitu Psca Aled Mosp Lilu Vine BIsa Citrus Myla Brom Cipo (1,5) (20) (2.1) (2.6) (2.8) (4.3) (5.0) (6.5) (8.0) (8.0) (8.9) (12.5) Substrate (average nest height in m) FIG. 1. (A) Distribution of Tardus rufivenlris nest heights. (B) Number of Tardus rafiventris nests found per substrate species (n = 42). Numbers in parentheses show mean height at which nests were built on each substrate. Lisi: Ligastram sinense; Pitu: Piper tucumanuiv, Psca: Psychotria carthagenensis-, Aled: Allophylus edalis\ Mosp: Moras spp., Lilu: Ligastram lacidanv. Vine: unidentified vine; Blsa: Blepharoccdyx salicifolias\ Citrus: Citrus spp.; Myla: Myrsine laetevireas', Brom: bromeliads; Cipo: Cinnamonuiai porphyriam. 678 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4, December 2010 TABLE 1. Substrates used by Tiirdus rufiventris at El Rey National Park, Salta, Argentina. Substrate species # Nests Family Acacia aroma 3 Fabaceae A. visco 1 Fabaceae Allophyhis edulis 21 Sapindaceae Celtis spinosa 1 Celtidaceae Condalia buxifolia Enterolobium 3 Rhamnaceae contortisiliquum 1 Fabaceae Eugenia uniflora 1 Myrtaceae Gledit.sia amorphoides 2 Fabaceae Nectandra pichurim 1 Lauraceae Pogonopu.^ tuhulosus 1 Rubiaceae Samhuciis peruviana 2 Adoxaceae Scutia buxifolia 5 Rhamnaceae Side roxy Ion ohtusifolium 1 Sapotaceae Urera caracasana 2 Urticaceae Vassohia hreviflora 1 Solanaceae Xylo.sma puhescens 1 Salicaceae Unidentified bromeliad 9 Bromeliaceae Unidentified fern 1 Unidentified moss 2 Unidentified Myrtaceae 2 Myrtaceae Unidentified vine 4 Total 65 DISCUSSION The high variability of substrate type and nest height u,sed by Rufous-bellied Thrushes is con- trary to that expected if nesting was conserved within a species (Knight and Fitzner 1985; Dhindsa et al. 1988; Martin 1988, 1993; but .see Forstmeier and Weiss 2004, Eggers et al. 2006, Peluc et al. 2008). This species nested on at least 12 substrate types, including understory shrubs, bromeliad epiphytes, canopy trees, and exotic as well as native species. Turdiis rufiventris al.so placed nests at a wide range of heights from 1 .5 to 17 m. Rufous-bellied Thrushes did not seem to specialize on one substrate type, but were selective of nesting substrate, particularly at the site where microhabitat selectivity was associated with potential fitness benefits. Turdus rufiventris favored bromeliads, the substrate associated with highest nest success at San Javier, in accordance with the expectation that birds should maximize nest success (Martin 1998). In contrast, this species was not selective with respect to substrate at El Rey, where predation rates did not differ between nest sites. Our results on nest-site choice and nest success in relation to disturbance differed depending on the scale at which we analyzed our data. Our results at a large scale did not support the hypothesis that anthropogenic disturbance in- creases the threat of nest predation (i.e., overall DPR was similar for San Javier vs. El Rey) (Holt and Martin 1997, De Santo et al. 2002, Easton and Martin 2002, Kellett et al. 2003). However, when examined at the microhabitat scale, we found that predation rate was much higher at the disturbed than at the undisturbed site. The higher predation rate at the disturbed site on non-bromeliad nest sites suggests predation rates were higher in the disturbed habitat but birds compensated through nest-site choice. These results confirm the impor- tance of examining ecological patterns at multiple .scales (Holling 1992, Levin 1992); if we had ignored microhabitat scale mechanisms, large scale disturbance effects would have been misin- terpreted. What makes bromeliads a safer nesting site at San Javier, and why is this not the case at El Rey? We showed that substrate height is not the determining factor; thus, bromeliads must be influencing other aspects of nest-site structure that affect predation risk. We believe bromeliad TABLE 2. Nest success and number of nests of Turdus rufiventris (Turdidae) on bromeliads and other substrates, and substrate availability at El Rey National Park, Salta, and Sierra de San Javier Biological Park, Tucuinan, Argentina. El Rey National Park .Sierra dc San Javier Biological Ptirk Overall DPR 0.0592 ± 0.0089 0.0684 ± 0.0095 DPR on bromeliads 0.0522 ± 0.0207 0.0308 ± 0.0175 DPR on other substrates 0.0606 ± 0.0098 0. 1 1 26 ± 0.0257 Number of nests on bromeliads (% of total nests) 9 (14%) 7 (26%) Number of nests on other substrates (% of total nests) 56 (86%) 20 (74%) Substrate availability Bromeliads (%) 7 3 Others (%) 93 97 Lomdscolo Cl al. • RUFOUS-BELLIED THRUSH IN MONTANE E’ORESTS 679 structure may make nests less visible and less vulnerable to predation by white-eared opossums (Didelpliis al biventer) and black rats {Rattiis rottus), which are common around houses in disturbed areas, but not to predation by House Wrens (Troglodytes aedon). Plush-crested Jays (Cyanocorax chrysops), and brown capuchin monkeys (Cehiis apella), common nest predators in the undisturbed site (Auer et al. 2007). In addition, the relevant structural feature may be one that only influences predation risk in disturbed areas. Predator communities most likely differ between disturbed and undisturbed sites (Auer et al. 2007; J. P. Jayat, pers. comm.); thus, the relevant structural feature must be one that conceals nests from predators that occur in disturbed sites, either exclusively or at least in higher numbers. Our results suggest a potential mechanism by which birds could adapt to the negative effects of human disturbance on nest predation risk. Turdus rufiventris apparently adjusts its nest-site choices to compensate for elevated predation risk in disturbed habitats. Additional studies of other species that compare both microhabitat choices and microhabitat-predation relationships in dis- turbed versus undisturbed areas are necessary to establish the generality of this mechanism. Studies of a range of species that vary in adaptability to human disturbance would be particularly helpful. Our ability to make general conclusions from this study is somewhat limited given the inclusion of only one disturbed site and one undisturbed site. Additional studies incorporating more replication would be valuable. ACKNOWLEDGMENTS SL was supported by an undergraduate student fellow- ship from Secretan'a de Ciencia y Tecnica from Universidad Nacional de Tucuman. Mitchell Aide made valuable comments on the manuscript. J. J. Lomascolo, T. A. and V. A. Monmany, and A. M. Garzia helped with transpor- tation and general support. I. M. Paz Posse helped with fieldwork. LITERATURE CITED Aragon, R. and J. M. Morales. 2003. Species compo- sition and invasion in NW Argentina secondary forests: effects of land use history, environment and landscape. Journal of Vegetation Science 14:195-204. Auer, S. K., R. D. Bassar, J. J. Fontaine, and T. E. Martin. 2007. Breeding biology of passerines in a subtropical montane forest in northwestern Argentina. Condor 109:321-333. Blake, J. G. and M. Rouges. 1997. Variation in capture rates of understory birds in El Rey National Park, northwestern Argentina. Ornitologia Neotropical 8:185-193. Brown, A. D. 1995. Las selvas de montana del noroeste de Argentina: problemas ambientales e importancia de su conservacion. Pages 9-18 in Investigacion, conserva- cion y desarrollo en selvas subtropicales de montana (A. D. Brown and H. R. Grau, Editors). Universidad Nacional de Tucuman, Argentina. Brown, A. D., H. R. Grau, L. R. Malizia, and A. Grau. 2001. Argentina. Pages 623-659 in Bosques nublados del Neotropico (M. Kappelle and A. D. Brown, Editors). INBio, San Jose, Costa Rica. de la Pena, M. R. 1979, Enciclopedia de las aves argentinas. Volume 7. Libren'a y Editorial Colmegna S. A., Santa Fe, Argentina. de la Pena, M. R. and M. Rumboll. 1998. Birds of southern South America and Antarctica. Princeton University Press, Princeton, New Jersey, USA. De Santo, T. L., M. F. Willson, K. E. Sieving, and J. J. Armesto. 2002. Nesting biology of tapaculos (Rhi- nocryptidae) in fragmented south-temperate rainforests of Chile. Condor 104:482-^95. Dhindsa, M. S., P. E. Komers, and D. A. Boag. 1988. Nest height of Black-billed Magpies: is it determined by human disturbance or habitat type? Canadian Journal of Zoology 67:228-232. Easton, W. E. and K. Martin. 2002. Effects of thinning and herbicide treatments on nest-site selection by songbirds in young, managed forests. Auk 119:685- 694. Eggers, S., M. Griesser, M. Nystrand, and j. Ekman. 2006. Predation risk induces changes in nest-site selection and clutch size in the Siberian Jay. Proceedings of the Royal Society of London, Series B 273:701-706. Ferretti, V., P. E. Llambias, and T. E. Martin. 2005. Life history differences between populations of a neotropical thrush challenge food limitation theory. Proceedings of the Royal Society of London. Series B 272:769-773. Fontaine, J. J., M. Martel, H. M. Markland, A. Niklison, K. Decker, and T. E. Martin. 2007. Testing ecological and behavioral correlates of nest predation. Oikos 116:1887-1894. Forstmeier, W. and I. Weiss. 2004. Adaptive plasticity in nest-site selection in respon.se to changing predation risk. Oikos 104:487-499. Grau, H. R. and R. AragUn. 2000. Arboles invasores de la Siena de San Javier, Tucuman, Argentina. Pages 5-20 in Ecologi'a de arboles exoticos en las Yungas (H. R. Grau and R. Aragon. Editors). LIEY-Universidad Nacional de Tucuman, Argentina. Grau, H. R., M, F. Arturi, A. D. Brown, and P. G. Acenolaza. 1997. Floristic and structural patterns along a chronosequence of secondary forest succession in Argentinean subtropical montane forests. Forest Ecology and Management 95:161-171. Grau, H. R., L. Paolini, A. Malizia, and J. Carilla. In press. Distribucion, estructura y dinamica de los 680 THE WILSON JOURNAL OF ORNITHOLOGY • Voi 122, No. 4, December 2010 bosques de la Sierra de San Javier. Pages 000-000 in E.stado Ambiental de la Sien'a de San Javier (M. E. Hernandez, Editor). Universidad Nacional de Tucu- man, Argentina. Grau, H. R., M. E. Hernandez., J. Gutierrez, N. I. Gasparri, C. Casaveccia, E. E. Flore, s-1valdi, and L. Paolini. 2008. A peri-urban neotropical forest transition and its consequences for environmental services. Ecology and Society 13:35. Hensler, G. L. and J. D. Nichols. 1981. The Mayfield method of estimating nesting success — a model, estimators and simulation results. Wilson Bulletin 93:42-53. Hines, J. E. and J. R. Sauer. 1989. Program CONTRAST: a general program for the analysis of several survival or recovery rate estimates. Fish and Wildlife Technical Report 24. USDI, Fish and Wildlife Service, Wash- ington, D.C., USA. Holling, C. S. 1992. Cross-scale morphology, geometry and dynamics of ecosystems. Ecological Monographs 62:447-502. Holt, R. F. and K. M.artin. 1997. Landscape modification and patch selection: the demography of two secondary cavity nesters colonizing clearcuts. Auk 1 14:443^55. Kellett, D. K., R. T. Alisauskas, and K. R. Mehl. 2003. Nest-site selection, interspecific associations, and nest success of King Eiders. Condor 105:373-378. Knight, R. L. and R. E. Fitzner. 1985. Human disturbance and nest site placement in Black-billed Magpies. Journal of Field Ornithology 56:153-157. Levin, S. A. 1992. The problem of pattern and scale in ecology. Ecology 73:1943-1967. Lima, ,S. L. 2009. Predators and the breeding bird: behavioral and reproductive flexibility under the risk of predation. Biological Reviews 84:845-513. MacArthur, R. H. 1972. Geographical ecology: patterns in the distribution of species. Harper and Row, New York, USA. Malizia, L. R. 2001. Seasonal tJuctuations of birds, fruits, and flowers in a subtropical forest of Argentina. Condor 103:45-61. Martin, T. E. 1988. Processes organizing open-nesting bird assemblages: competition or nest predation? Evolutionary Ecology 2:37-50. Martin, T. E. 1992. Breeding productivity considerations: what are the appropriate habitat features for manage- ment? Pages 455^73 in Ecology and conservation of neotropical migrants (J. M. Hagan and D. W. Johnston, Editors). Smithsonian Institution Press, Washington, D.C., USA. Martin, T. E. 1993. Nest predation and nest sites. New perspectives on old patterns. BioScience 43:523-532. Martin, T. E. 1998. Are microhabitat preferences of coexisting species under selection and adaptive? Ecology 79:656-670. Martin, T. E. and J. V. Briskie. 2009. Predation on dependent offspring: a review of the consequences for mean expression and phenotypic plasticity in avian life history traits. The year in evolutionary biology 2009. Annals of the New York Academy of Sciences 1168:201-217. Martin, T. E. and G. R. Geupel. 1993. Nest monitoring plots: methods for locating nests and monitoring success. Journal of Field Ornithology 64:506-519. 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. Marzluff, j. M. 1988. Do Pinyon Jays alter nest placement ba.sed on prior experience? Animal Behaviour 36:1- 10. Matteuci, S. D. and a. Colma. 1982. Metodologi'a para el estudio de la vegetacion. Monograph Number 22. General Secretary of the OEA, Washington, D.C., USA. Mayfield, H. 1961. Nesting success calculated from exposure. Wilson Bulletin 73:255-261. Mezquida, E. T. and L. Marone. 2002. Microhabitat structure and avian nest predation risk in an open Argentinean woodland: an experimental study. Acta Oecologica 23:313-320. Peluc, S., S. T. Sillett, j. T. Rotenberry, and C. K. Ghalambor. 2008. Adaptive phenotypic plasticity in an island songbird exposed to novel predation risk. Behavioral Ecology 19:830-835. Pimm, S. L., M. Ayres, A. Balmford, G. Branch, K. Brandon, T. Brooks, R. Bustamante, and R. Costanza. 2001. Can we defy nature’s end? Science 293: 2207-2208. Ridgely, R. S. and G. Tudor. 1989. The birds of South America. Volume 1. University of Texas Press, Austin, USA. Zar, j. H. 1999. Biostatistical analysis. Fourth Edition. Prentice-Hall, Upper Saddle River, New Jersey, USA. The Wilson Jounuil of Ornithology 1 22(4);68 1-688, 2010 BREEDING BIOLOGY OE THE TAIWAN BARBET {MEGALAIMA NUCHALIS) IN TAIPEI BOTANICAL GARDEN SHANG-YAO LIN,'^ FANG-CIAN LU,' FU-HSIEN SHAN,' SHU-PING LIAO,' JING-LING WENG,' WEI-JEN CHENG,' AND CHAO-NIEN KOH'-' ABSTRACT.— We studied the breeding biology of the Taiwan Barbet (Megalainia intchalis) in the Taipei Botanical Garden (TBG) during the 2008 and 2009 breeding seasons. Breeding pairs produced a mean of 1.8 broods per season and had a mean clutch size of 3.0 eggs. The mean incubation period was 13.8 days, and both parents shared incubation. The mean nestling period was —27.5 days, considerably shorter than that of other Asian barbets. The shorter nestling period may be related to the low fledgling success rate, multiple broods per season, and constant animal matter food in TBG throughout the entire nesting period. Ambient temperature influenced the amount of time spent by adults inside the nest cavity incubating eggs and brooding nestlings. Adult males provided more food to nestlings than adult females, while adult females cleaned the nest more often than males. The overall egg-to-fledgling success rate was low at 32.8%, because of infertile eggs and failure of eggs with embryos to hatch, abandonment, anthropogenic disturbances, predation, weather disturbances, and other unknown factors. Received 7 January 2010. ■Accepted 15 April 2010. The Asian barbets (Family Megalaimidae) are one of the least studied groups compared to barbets in Africa and Central and South America (Yahya 1988). The Asian barbets belong predom- inantly to the genus Megalaima, which includes 28 species and is centered in Southeast Asia (Short and Horne 2001). Barbets are mostly frugivores and are important pollinators and seed dispersers. Some of the more insectivorous barbets, such as White-cheeked Barbet (M. viridis), have a vital role in controlling insect pests in agricultural regions (Yahya 2000). Barbets are primary tree-cavity nesters that excavate their own nesting and roosting holes; thus, losses of dead trunks or branches have detrimental effects on this group (Short and Horne 2001). Gradual shrinkage of forest cover may have a far-reaching effect on barbets as is the case of the Red-crowned Barbet (M. rafflesii), which has nearly been eradicated from Thailand because of forest clear-cutting (Wells 1999). Detailed information on the breeding biology of most Megcdahna spp. is almost non-existent. The Taiwan Barbet (M. tniclutli.s) is an endemic species restricted to Taiwan, and was a subspecies of the Black-browed Barbet (M. oorli niicluilis) until recent genetic comparisons (Feinstein et al. 2008), and morphological and behavioral differ- ' Forest Protection Division. Taiwan Forestry Research Institute. Number 53 Nanhai Road. Zhongzheng District. Taipei City. Taiwan (R.O.C.). ^8011 Ryan Road, Unit 105. Richmond. BC V7A 2E4, Canada. ^Corresponding author; e-mail: nien@tfri.gov.tw ences among the subspecies (Collar 2006) gave it full species status. The Taiwan Barbet is com- monly distributed throughout Taiwan at low- to mid-elevations. They commonly occur in broad- leaf forests in both natural and man-made habitats, including parks, botanical gardens, school yards, and trees along sidewalks. This sexually monomorphic species begins breeding in April and continues into August (Fang 2008); detailed information on its breeding biology and parental roles is extremely limited. Koh and Lu (2009) investigated characteristics of nest trees and nest cavities of Taiwan Barbets in Taipei Botanical Garden (TBG). They discovered the species nested mostly in Ciunamomum camphora trees and in dead standing trees or live trees with dead branches. The only other research on this species was of its general biology in Yangming- shan National Park, Taiwan (Ho 1990). Primary cavity nesters such as the Taiwan Barbet can be affected by land management practices, especially in urbanized environments (James and Kannan 2009) where urban trees may be trimmed or completely removed during beautification actions or when the trees po.se a safety hazard for pedestrians or vehicles in traffic. The Taiwan Barbet is commonly sold in pet stores and origins of these captures are questionable, and most likely illegal. Our objective is to present information on the breeding biology and behavior of the relative- ly poorly known Taiwan Barbet. More specifical- ly, we ascertained the length of the incubation and brooding periods, parental roles, and hatching and fledging success rates in Taipei Botanical Garden. 681 682 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4, December 2010 METHODS Study Area. — This study occun'ed in Taipei Botanical Garden (TBG) (25° 01 ' 53" N, 121° 30' 39" E) from March to September 2008 and 2009. The botanical garden encompasses 8 ha of urban green space in central Taipei City, Taiwan and has >1,500 living plant species. The monthly tem- perature ranged from 14.2 to 30.1° C, averaging 22.7° C in 2008 and 2009; monthly precipitation ranged from zero to 1668.5 mm, averaging 209.3 mm (TBG, unpubl. data). This area of high plant diversity and abundance in the heavily urbanized city center provides excellent habitat for urban wildlife, especially avian species. Eorty- five avian species have been reported in TBG with 30 resident species and the other escapees, migrants, and transients (Koh 1998). Koh (1998) noted the year-round presence of Taiwan Barbets and categorized their status as resident. The data base from the Wild Bird Society of Ilan in Taiwan (http://wildbird.e-land.gov.tw) indicates Taiwan Barbets were observed almost every month from December 2003 to October 2005, suggesting the presence of a stable population in TBG. Nest Searching, Observing, and Monitoring. — Searches for paired barbets and their nests were conducted opportunistically on foot within TBG at least once each day from early March to mid- September in 2008 and 2009. Searches were based on visual and auditory cues, and nests were located by following adults. The gender of the adults was ascertained by observing their posi- tions during copulation (male mounting the female from rear) and was identifiable by learning the key features between individuals in the pair. Adults and fledglings were captured in mist nets, measured, and color-banded during the nestling- provisioning stage and after fledging, respective- ly. The condition and status of nest cavities were inspected using an endo.scope. The endoscope is custom made with a pinpoint camera at the end of an extendable pole of up to 6 m in length with a small monitor attached to the camera via exten- sion cables. This equipment allowed us to look inside the nest, and videotape and photograph the nest’s current conditions, whether it was a nesting cavity, abandoned, or incomplete. The endoscope also allowed us to identify the dates eggs were laid, clutch size, and .stages of nestling develop- ment. Breeding behaviors, including cavity exca- vation, copulation, nestling provisioning, nest cleaning, and time of entrance and exit ot parents were observed and recorded using spotting scopes and binoculars from distances that would not disrupt the birds’ natural behaviors and actions. Time spent inside the nest cavity during the incubation and nestling periods by adults was defined as incubation time and brooding time, respectively. The brooding time did not include time spent by adults inside the nest provisioning nestlings or nest sanitation. The incubation period was measured from the day the last egg was laid to the day the first egg hatched and was subdivided into two stages: early (1-6 days after incubation) and late (7-15 days after incubation). The nestling period was from the day the first egg hatched to the day the first nestling fledged. The nestling period was subdi- vided into two stages: early (1-13 days after hatching) and late (14-27 days after hatching). The nestling period for nest sanitation and nestling provisioning data was subdivided into three stages: early (1-10 days after hatching), middle (11-20 days after hatching), and final (21 days and onward after hatching). Nest sanitation was recorded when adults removed wood chips mixed with fecal matter from the nest cavity, and nestling provisioning was recorded when adults brought food into the nest cavity. Identification of food type was performed to the best ability of the observers. Eood types fed to nestlings were categorized as “plant matter’’, “animal matter’’, and “unknown’’. All variables were grouped into temporal periods of 4 hrs each, from 0600 to 1000 (morning), 1000 to 1400 (midday), and 1400 to 1800 hrs (afternoon), to evaluate mean differences of the variables during different periods of the day and between the parents in each time period (Yahya 1988). Success rates were calculated by dividing the number of nestlings or fledglings by the number of eggs or nestlings in the previous life stage, respectively. We removed the nest cavity and retrieved three eggs of ~1 week of age to conduct physical measurements of the eggs from one abandoned and Hooded nest. Hourly temperatures were obtained from the weather stations at TBG. The associated temperatures for each incubation and brooding period (hrs and days) were averaged to obtain the mean hourly and daily temperature for each incubation and brooding period. Statistical Analysis. — Statistical analyses were performed using SPSS for Windows Version 12.0 (SPSS Institute Inc. 2000) and Microsoft Office Excel 2003. Examination of data normality was Lin et al. • TAIWAN BARBET BREEDING BIOLOGY 683 TABLE 1 . Incubation and brooding period.s, and success Garden. Taiwan. rates of Taiwan Barbets during 2008-2009 in Taipei Botanical Mean ± SE («) Range Number of eggs laid 3.0 ± 0.1 (1 1) 2.()-4.0 Number of eggs hatched 2.0 ± 0.3 (13) O.O^.O Egg hatching success rate, % 44.0 ± 12.5 (9) 0.0-100.0 Number of nestlings successfully fledged LI ± 0.2 (15) 0.0-3.0 Fledging success rate — relative to number of eggs, % 32.8 ± 9.9 (11) 0.0-100.0 Fledging success rate — relative to number of hatchlings, % 45.4 ± 13.7 (9) 0.0-100.0 Length of incubation period, days 13.8 ± 0.3 (6) 13.0-15.0 Length of nestling period, days 27.5 ± 1.0 (6) 23.0-29.0 conducted using the one-sample Kolmogorov- Smirov test. Means were compared using paired r-tests or one-way ANOVA for data with normal distributions. Either Mann-Whitney f/-tests or Kruskal- Wallis tests were used for data with non-normal distributions. Tukey’s HSD tests were used, when appropriate, to examine which means of groups significantly differed from one another. Means ± SE are presented. RESULTS The breeding season of Taiwan Barbets began in April when the first observed pair at the study site laid their first egg and ended in August with fledging of the last nestling. We located eight and 20 broods in 2008 and 2009, respectively. Identified pairs in 2008 and 2009 produced a mean of 1.8 broods/pair (range = 1. 0-3.0, n — 1 3 ) and had a brood success rate of 7 1 .4% (n = 28). The last egg of the .season in 2008 was laid on 6 August. However, due to natural weather disturbances, this nest was abandoned and the season terminated on 17 August. The first egg in 2009 was estimated to have been laid on 20 April, and the last nestling fledged on 6 August. Courtship Behavior and Copulation. — Court- ship feeding was observed four times in 2009. Males fed unidentified berries on three separate occasions to females and an unidentified insect on another occasion. Copulation may or may not occur in conjunction with courtship feeding. The male mounted the female in one occasion while still holding three berries in his beak. The female selected one berry after copulation from the male’s beak and the male ate the rest. The average copulation attempt was 10.8 ± 4.0 sec in = 5). An unique behavior ob.served during the breeding season was the beak-tapping action performed by an adult male at the cavity entrance in = 4). Cavity E.xcavation. — The adult male signifi- cantly excavated the nest cavity longer than its partner. Males {n = 4) excavated for 17.6 ± 1.7 min/hr compared to 6.9 ± 0.8 min/hr for females (paired /-test, / = 5.447, df = 3, E < 0.05). The time required to complete a nesting cavity for one particular pair was 22 days. Egg Laying and Clutch Size. — Eemales laid an egg per day (n = 6), and the mean clutch size was 3.0 ± 0.1 (range = 2.0-4.0 eggs, n = 11; Table 1). We opened the nest cavity and mea- sured three eggs from a flooded and abandoned nest cavity. The eggs were white in color and subelliptical-shaped. Mean egg length was 26.9 ± 0.5 mm (range = 26.1-27.7 mm, n = 3) and the mean width was 19.8 ± 0.2 mm (range = 19.4- 20.0 mm, n = 3); mean egg weight was 5.4 ± 0.4 g (range = 5.0-5. 8 g, n = 2). One other egg was addled when retrieved and weighed 3.0 g. Incubation Period. — The incubation period began when the last egg of the clutch was laid and lasted 13.8 ± 0.3 days (range = 13.0- 15.0 days, n = 6; Table 1). Time spent by adults incubating eggs throughout the incubation period was 34.1 ± 2.3 min/hr (range = 21.6-43.3 min/ hr, n = 9). Males and females incubated equally in early and late stages. Incubation occurred most frequently in the early morning and decreased as the day progressed (range = 29.5^7.4 min/hr; Fig. 1). Eggs were most frequently incubated by adults during 0600 to 1000 hrs (ANOVA, E = 3.837, df = 25, P < 0.05) when compared to 1000 to 1400 and 1400 to 1800-hr periods (Tukey test, P < 0.05). Mean ambient temperature was lowest at 23.4° C during early mornings and increased to 29.5° C from 1300 to 1400 hrs, and then decreased to 27.2° C from 1700 to 1 800 hrs (TBG, unpubl. data; Fig. 1 ). Males incubated longer than females during the 0600 to 1000 and 1400 to 1800-hr periods but the differences were not 684 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4, December 2010 U o <0 Uh 3 u 22.5 ■ ■o 20.0 S 17.5 15.0 12.5 10.0 7.5 5.0 2.5 0.0 Time periods FIG. 2. Hourly brooding time by Taiwan Barbets in different hours of the nestling period and the mean daytime temperature for each day during 2008—2009 in Taipei Botanical Garden, Taiwan. The mean temperature was calculated by averaging the daytime temperatures of each brooding period. Error bars are ± SE. late afternoon (Fig. 2). Temperatures were coolest in the early morning and late afternoon and warmest in mid-day (Fig. 2). Females continued to be the primary parent staying overnight inside the nesting cavity during the nestling period (/; = 4) while a male was ob,served once with the nestlings overnight. Only 45.4 ± 13.7% of the nestlings fledged successfully (n = 9; Table 1). Nestling Provision, Food Types, and Nest Sanitation. — Frequency of nestling provision and nest cleaning increased as brooding time de- creased. The overall food provisioning rate was 1.3 ± 0.2 times/hr and increased during the nestling period (range = ().3-3.0 times/hr. n = 8). On average, males delivered food more frequently but not more significantly than females during the nestling period. The greatest amount of animal matter (27.1%) was fed to nestlings during the middle stage of the nestling period and decreased to 22.3% in the final stage. There were no 686 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4. December 2010 differences in food provisioning rates between animal and plant matter during the nestling period. Plant matter consisted entirely of berries from many different plant species, while animal matter consisted of mostly insects, except for four special small vertebrates, one gecko and three lizards. The insects provided to nestlings were: Coleoptera ( 1 1.5%), Hemiptera (7.9%), Hymenop- tera (1.8%), Mantodea (1.3%), Odonata (<1.0%), Lepidoptera (<1.0%), Orthoptera (<1.0%), and Blattaria (<1.0%); the remainder were insects of unidentified Orders (74.8%) (/? = 8). Nest sanitation was a frequent and important task performed throughout the nestling period, averaging 0.9 ± 0.1 times/hr (range = 0.5-1. 2 times/hr, n = 8) by both parents. Females maintained a higher nest sanitation rate than males throughout the nestling period in all three time periods (paired /-test, / = —4.206, df = 7, P < 0.01), especially in the early stage (paired /-test, / = —4.376, df = 4, P < 0.05). Adults cleaned the nest cavity more frequently and significantly during the 0600 to 1000-hr period than in the 1400 to 1800-hr period (Tukey test, P < 0.05). DISCUSSION The breeding season of Taiwan Barbets began in April, similar to other Megalaima spp. that live at similar latitudes. The breeding seasons of some Megalaima spp. may start as early as January and end as late as October, but commonly occur from March to July (Short and Horne 2001 ). Ho ( 1990) reported the breeding season of Taiwan Barbets in Yangmingshan National Park, Taiwan concluded in late August. The breeding .season at TBG also terminated in August. The average breeding season length of Taiwan Barbet in TBG was >3 months in length. We hypothesize that length of the breeding sea, son is affected by weather disturbances, number of broods per pair, length of incubation and nestling periods, and the repro- ductive condition of adult females after each brood. Egg Laying and Clutch Size. — The dimensions of Taiwan Barbet eggs are smaller than eggs of M. oorti nnchalis documented by Short and Horne (2001). More egg data are needed to provide a more preci.se range and mean of egg dimensions. The mean clutch of 3.1 eggs was greater than for M. viridis, an Asian barbet of similar body size (Yahya 1988). However, M. viridis has larger eggs and higher egg hatching success rate than Taiwan Barbets. We hypothesize the Taiwan Barbet’ s large clutch size helps to compensate for low egg survival. Incubation Period. — Eggs of the same clutch are laid on consecutive days and hatch on the same day, as observed for M. viridis and M. rubricapilla (Yahya 1988). Time required for Taiwan Barbet eggs to hatch is —14 days, within the range of 12 to 15 days for other Asian barbets (Short and Horne 2001). Adults were frequently inside the nest incubating eggs in the early morning and, by late afternoon, their time inside the cavity diminished. This temporal difference is most likely influenced by ambient temperatures. Temperatures were cooler in early morning and late afternoon and warm during the mid-day, which correspond with more and less time by adults incubating the eggs, respectively. M. viridis and M. rubricapilla parents are also more attentive during cool mornings and rainy days than during the afternoon (Yahya 1988). The effect of ambient temperature fluctuations on behaviors of breeding adults supports the hypothesis that optimal temperature has an important role in affecting the success of breeding pairs (White and Kinney 1974). Males incubated more in morning and afternoon periods than females; we hypothesize males relieve females after their overnight stay inside the cavity and become the primary incubator in the early morning. Males again take on the role of the primary incubator in late afternoon as the female forages and prepares for her nocturnal incubation duty in the evening. Nestling Period. — The nestling period of —27.5 days for Taiwan Barbets is short compared to 35-38 days for five small- to large-sized Megalaima spp. (Short and Horne 2001). We hypothesize the shorter nestling period may be related to the low fledgling success rate, multiple broods per .season, and steady animal matter food provided by the parents throughout the entire nesting period. Martin and Li (1992) commented that length of the nestling periods was inversely related to risks of nest failure and number of broods the parents produced in a breeding season. The lledgling failure rate in our study was relatively high compared to other Megcdaima spp. (Yahya 1988), and parents did produce multiple broods per breeding season. Taiwan Barbets supplied nestlings with animal matter food during all stages of the nesting period, especially during the middle stage. Adult M. rubricapilla and Liu et al. • TAIWAN BARBET BREEDING BIOLOGY 687 M. viriclis ted their nestlings more plant matter food than animal matter food for the majority of the nesting period (Yahya 1988). We suggest that Taiwan Barbet nestlings are provided with a constant source of animal protein and can develop faster than other Megalaima spp. Provisioning and Nest Sanitation.— Males at our study site have a slightly larger role in provisioning nestlings than females, while fe- males have a larger role in nest sanitation than males throughout the day. Overall, the parents more or less cooperated with each other in nestling provisioning and nest sanitation. A female disappeared during the nestling period on one occasion, leaving the male to do all nest sanitation and nestling provisioning. The three nestlings died in the male’s care. This result suggests that fledging success may be greatly influenced by the performance of either parent due to their cooperative roles in nestling provi- sioning and nest sanitation. Success and Mortality Rates. — Yahya (1988) reported the egg hatching success for M. viridis and M. riibricapilla to be 71.1 and 80.0%, respectively, which are higher than that of Taiwan Barbets in our study. Causes for failures of Taiwan Barbets were infertile eggs and failure of eggs with embryos to hatch (40.0%), human influences (25.0%), weather disturbances (15.0%), and unknown egg disappearances (20.0%), possibly a result of nest predation. The fledgling success rate was also lower than for M. viridis and M. riibricapilla at about 75.7 and 75.0%, respectively (Yahya 1988). The specific causes of nestling mortality are unknown, but 27.2% of the nestlings died inside the nest and were removed from the nest cavity by the parents, and 36.4% of the nestlings starved to death because parents had not returned to the nests. Mortalities of the remaining unfledged nestlings were of unknown disappearances. Potential fac- tors that reduce fledgling success of cavity nesters include nest parasites, nest desertion, starvation of young, predation, and hyperthermia (Ricklefs 1969, Nilsson 1986, LaBranche and Walters 1994). The specific causes that result in low egg hatching and fledgling success rates remain unclear and are a priority for future researchers. The workload differences between parents and a slightly more important role of females than males during incubation to nestling periods provide a new perspective into the relatively unknown breeding biology of Taiwan Barbets. Low breeding success rates merit attention and protection, especially for threats the species has from city beautification processes and the pet trade industry. ACKNOWLEDGMENTS We thank Hsiao-Wei Yuan and Tzung-Su Ding for comments on earlier drafts. We also thank C. E. Braun for proofreading the English of the manuscript, and two anonymous reviewers for commenting on the manuscript. We are grateful to C.-W. Chen, Y. S. Ho, W.-C. Yeh, and Y.-M. Chen for assistance with field work. We thank M.-S. Huang, H.-W. Chou, S.-Y. Peng, and L.-S. Lung for support. We are also grateful to K.-K. Kuo for supplying us with a professional endoscope to investigate the conditions inside nest cavities. We sincerely thank S.-Y. Chen and the many volunteers who contributed to the successful completion of this study. Our study was supported by grants from the Council of Agriculture and National Science Council, Taiwan. LITERATURE CITED Collar, N. J. 2006. A taxonomic reappraisal of the Black- browed Barbet Megalaima oorti. Forktail 22:170-173. Fang, W.-H. 2008. A guide to all birds of Taiwan. Owl Publishing House, Taipei, Taiwan. Feinstein, J., X. Yang, and S.-H. Li. 2008. Molecular systematics and historical biogrography of the Black- browed Barbet species complex (Megalaima oorti). Ibis 150:40-49. Ho, Y.-C. 1990. The biological study of Muller's Barbet Megalaima oorti imchalis in Yangmingshan National Park. Thesis. National Taiwan University, Taiwan. James, D. A. and R. Kannan. 2009. Nesting habitat of the Great Hornbill (Biiceros hicornis) in the Anaimalai Hills of .southern India. Wilson Journal of Ornithology 121:485^92. Koh, C.-N. 1998. Report on the bird fauna in the Taipei Botanical Garden. Taiwan Journal of Forest Science 13:69-78. Koh, C.-N. and F.-C. Lu, 2009. Preliminary investigation on nest-tree and nest-cavity characteristics of the Taiwan Barbet (Megalaima niichalis) in Taipei Botanical Garden. Taiwan Journal of Forest Science 24:213-219. LaBranche, M. S. and J. R. Walters, 1994. Patterns of mortality in nests of Red-cockaded Woodpeckers in the sandhills of .southcentral North Carolina. Wilson Bulletin 106:258-271. Martin, T. E. and P. Li. 1992. Life history traits of open- vs. cavity-nesting birds. Ecology 73:579-592. NilS-SON, S. G. 1986. Evolution of hole-nesting in birds: on balancing selection pressures. Auk 103:432-435. Ricklefs, R. E. 1969. An analysis of nesting mortality in birds. Smithsonian Contributions to Zoology 9:1^8. Short, L. L. and J. F. M. Horne. 2001. Toucans, barbets, and honeyguides. Oxford University Press, Oxford, United Kingdom. 688 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4, December 2010 SPSS Institute Inc. 2000. SPSS for Windows. Version 12.0. SPSS Institute Inc., Chicago, Illinois, USA. Wells, D. R. 1999. The birds of the Thai-Malay Peninsula. Volume I. Non-passerines. Christopher Helm, London, United Kingdom. White, F. N. and J. L. Kinney. 1974. Avian incubation. Science 1 86: 107-1 15. Yahya. H. S. a. 1988. Breeding biology of barbets, Megalaima spp. (Capitonidae: Piciformes) at Periyar Tiger Reserve, Kerala. Journal of the Bombay Natural History Society 85:493-511. Yahya, H. S. A. 2000. Food and feeding habits of Indian barbets, Megalaima spp. Journal of the Bombay Natural History Society 97:103-1 16. The ll 'ilson Journal of Ornithology 1 22(4):689-698. 2010 BREEDING BIOLOGY OF THE GOLDEN-FACED TYRANNULET {ZIMMERIUS CHRYSOPS) IN VENEZUELA WILLIAM GOULDING' AND THOMAS E. MARTIN' ABSTRACT.— We present the first detailed information on the breeding biology of the Golden-faced Tyrannulet (Ziiumernis chiysops). Information was gathered from 96 nests in Yacambn National Park, Venezuela during the 2002 to _008 breeding seasons. The enclosed nest was similar to descriptions of nests of other species in the genus. Eggs were laid on alternate days with mean (± SE) clutch size of 1.98 ± 0.02 (n = 45) and fresh weight of I.6I6 ± 0.020 g (/; = 48). Only the female incubated and the incubation period averaged 16.9 ± 0.3 days {n = 10). Nest attentiveness {% time on the nest) averaged 66.0 ± 1.6% {n = 40) and increased from early to mid- and late-incubation. Incubation behavior yielded an average 24-hi egg tempeiature ot 34.88 ± 0.45° C (n = 7 nests, 43 days). The nestling growth rate constant for body mass (k — 0.285 ± 0.01 1) was slow even for tropical tyrannids. The nestling period for nests where exact hatch and fledging days were observed ranged from 17 to 19 days with an average of 18.0 ± 0. 2 days (n = 9). Both females and males fed nestlings at a rate that increased over the nestling period with a mean of 4.41 ± 0.65 trips/hr {n = 10) during days 1 and 2 after hatching, and 14.93 ± 2.36 trips/hr {n = 6) at pin-break (days 10-11). Daily predation rates were similar in egg-laying (0.052 ± 0.025; n = 76.5 exposure days) and incubation periods (0.068 ± 0.010; n = 575.5 exposure days), but were lower during the nestling period (0.039 ± 0.010; n = 377.0 exposure days). The total daily predation rate (0.057 ± 0.007; n = 989.0 exposure days) indicated only 12% of nests were successful. These breeding biology parameters forZ chiysops differ substantially from other tyrant-flycatchers and temperate species, further highlighting the diversity within the Tyrannidae. Received 12 December 2009. Accepted 24 May 2010. The Golden-faced Tyrannulet (Zimmerius chrysops) is a member of the complex and large New World Family Tyrannidae. Tyrannids are diverse in behavior and occupy a broad range of habitats and niches, reaching their greatest diversity in the Neotropics (Ridgely and Tudor 1994, Hilty 2003, Fitzpatrick 2004), where they comprise >20% of passerine species (McGowan 2004). General trends in avian life histories cannot be clarified without studies of tropical species to compensate for the historical bias toward temper- ate species (Martin 1996, 2004; Stutchbury and Morton 2001). The paucity of information on tropical bird species is particularly evident for tyrannids. Information on the breeding biology of most tropical tyrannids is sparse despite their prominence in the neotropical avifauna, and diversity of form and function. Zimmerius chry'sops is an example of the sparse information on tropical tyrannids. Published breeding biology information on Z. chrysops is little more than a description of the enclosed, dome-shaped nest with a side entrance, and observed breeding dates (Hilty and Brown 1986, Best et al. 1996). Yet, it is broadly distributed with a range that extends from Peru, north through ' uses. Montana Cooperative Wildlife Research Unit. University of Montana. Missoula, MT 59812, USA. -Current address: Indooroopilly, Brisbane 4068 Queens- land, Australia. ^ Corresponding author; e-mail: willgoulding@yahoo.com.au the Andes to Colombia and Venezuela, and occurs from —300 to 2,400 m asl (Hilty and Brown 1986, Ridgely and Tudor 1994, Hilty 2003). The species is associated with the middle to upper forest layers, secondary growth, forest edges, as well as with plantation and garden habitats (Fjeldsa and Krabbe 1990). Our objectives are to: (1) present detailed reproductive biology information for Z. chrysops, and (2) compare this information with available temperate and tropical avian life history information. METHODS We studied Golden-faced Tyrannulets during the March-June breeding seasons in 2002-2008 in Yacambu National Park, Lara State, Venezuela (09° 42' N, 69° 42' W). Yacambu contains premontane and montane cloud forest of the northern Andes and encompasses a gradient from 500 to >2,300 m asl (Fierro-Calderon and Martin 2007). Peak precipitation in Yacambu occurs from late April through July (Fierro-Calderon and Martin 2007) with clouds and rain usually enveloping habitats after the morning period, producing cool wet conditions. We worked with equal field effort in the same areas across years in a 1 ,350 to 2,000 m asl re.search area. Nests were located primarily by observing parental behavior, but also by systematic search- ing. Dimensions of newly constmeted nests were measured before use and weather affected their 689 690 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4. December 2010 shape. Eggs were described following Preston (1953) as clarified by Palmer (1962). Eggs were weighed for fresh egg mass between days 0 and 3 of incubation, and opportunistically later. The incubation period was defined as the period between the last egg laid (day 0) and the last chick hatched (Nice 1954, Martin 2002). Nestling period was defined as the number of days between the last egg hatching and the last chick leaving the nest. Video cameras were placed at a discreet distance (>5 m) from the nest and camouflaged with vegetation to minimize any influence on parental behavior. Nests were filmed from 30 min after sunrise for 6 to 8 hrs (Martin and Ghalambor 1999, Martin 2002). Filming occuiTed on days 2-3 in early incubation, during the middle of the incuba- tion period, and ~2-3 days before expected hatch day (early, middle, and late incubation). Nests found after laying were filmed opportunistically on other days (Martin 2002). Nest attentiveness at individual nests was measured as the percent time females were on the nest based on the number of minutes spent incubating divided by the total number of minutes a nest was filmed (Martin 2002). Incubating egg temperatures (°C) were mea- sured by inserting a thermistor into the egg center through a small opening in the shell that was sealed w'ith adhesive (Weathers and Sullivan 1989, Martin et al. 2007). HOBO Stowaway XTI dataloggers (Onset Corp., Bourne, MA, USA) were connected to the thermistor by discreetly inserting a fine wire through the nest- wall. Thermistors were inserted within the first 2 days of incubation and dataloggers recorded temperatures in 12-24 sec intervals for 5-7 days per nest following Martin et al. (2007). Nests were videotaped during the early nestling period (days 1-3), during pin-break when flight feathers broke their sheaths (days 10-11), and the late nestling period preceding expected fledging (day 15 and above). Nests were opportunistically filmed on other days. Videotapes were analyzed for parental brooding attentiveness (%) based on the number of minutes brooding divided by the total number of minutes a nest was filmed, as well as the number of provisioning trips/hr (Martin and Ghalambor 1999, Martin et al. 200()b). Analysis of variance (ANOVA) was used to test for changes in incubation and nestling attentiveness between stages. Changes in brooding and provi- sioning rates with ne.stling age were analyzed using correlation coefficients (/■). Chicks were measured every second day (from day located) at the same time of day and at pin- break. Standard measurements taken were mass, tarsal length, wing chord, and primary feather pin measurements. These data were used to calculate the growth rate constant (k) following Ricklefs (1967) and Remes and Martin (2002). Overall nesting success and predation rates for egg-laying, incubation, and nestling periods were calculated using the Mayfield method (Mayfield 1961, 1975; Hensler and Nichols 1981; Johnson 2007). Chi- square tests of independence (X^) were used to test for differences in daily predation rates among nest stages (egg laying, incubation, and nestling periods). Egg and nestling masses were measured using an ACCULAB (Elk Grove, IL, USA) portable electronic scale with a precision of 0.001 g; all measurements were taken using Mitutoyo Digital Calipers (Kingsport, TN, USA) with a precision of 0.01 mm. SPSS Version 15.0 (2006) was used for statistical tests and means are given ± one standard error (SE). RESULTS Breeding Habitat. — Golden-faced Tyrannulets nested most commonly near 1,400 m asl and were not encountered above 1,500 m asl. This altitudi- nal zone contained primary and secondary growth premontane and montane cloud forest with patches of regenerating coffee {Cojfea arabica) plantations. Active nesting pairs were encountered near forest edges including roads, forest trails, water bodies, and areas of historical disturbance inside the forest boundary. Areas where nests were found contained suitable resources for nest construction (moss and lichen) on tree trunks. Nest Construction and Sites. — Pairs were ob- served exploring potentially suitable nest loca- tions together prior to the nest-building process. Only one of the pair, presumably the female, constructed the nest (// = 88 nest-building observations), but the mate generally accompa- nied the nest-building individual. Nests were typically constructed in a ball of moss on the side of a tree, under a branch, or in an epiphyte (Fig. 1). Two di.stinct building approaches were observed for nest construction. The first and more common approach involved building into an existing lump of moss or lichen. The second involved building the globular nest onto a bare trunk surface. In the first approach, when a suitable portion of moss or lichen was identified. Goiilc/ing and Martin • GOLDEN-FACED TYRANNULET BREEDING BIOLOGY 691 FIG. 1. Typical nest (with eggs and nestling) of the Golden-faced Tyrannulet with incubating female (Photo- graphs by W. Goulding). the adult would open a cavity first with the beak and head and, when at the mid-body section, would use its wings to thrust apart the moss and widen a cavity. The adult would collect materials to create the inside of the nest once a cavity was established. The adult used more of the other nesting materials for construction inside with the moss in situ. However, in the second approach where the nest was created on a relatively bare surface, initial construction involved slinging the nest material with spider webs from a point on the bark or from a small piece of moss or lichen. Females were observed repeatedly descending to the ground to scour exposed banks and road cuttings for rootlets and spider webs during construction (e.g., 6 visits in 10 min to the same location on a bank). They also loudly called from beside the nest after building visits with the nearby male usually responding. Females were observed re-using nesting material from previous unsuccessful nests, particularly the lining, but new material was often used. The small size of the nest often resulted in destruction of the nest during predation events. The mean height of nests above ground was 5.43 ± 0.42 m (n = 55). The external measure- ments were 1 16.14 ± 4.27 mm (/; = 10) in height FIG. 2. Seasonal distribution of nest initiation (date the first egg is laid in a nest) for the Golden-faced Tyrannulet at weekly (7 day) intervals. by 70.83 ± 2.50 mm (« = 6) in width. Nest cups had an external depth of 48.25 ± 2.87 mm (/? = 10) with internal depth and internal diameter being 32.85 ± 2.55 mm {n = 10) and 40.57 ± 2.71 mm (n = 7), respectively. Analysis of 11 nests showed an outer supporting structure of moss into which spider webs, black rootlets, lichen, and spider egg-casings of varying propor- tions were inserted with internal layers of finer dark rootlets. The final internal layer was a golden lining of soft, comose seed-arils fitting descrip- tions of liana seeds in the genus Odontadenia (Apocynaceae) (Orlando Vargas Ramirez, pers. comm.). The golden lining was characteristic of all Z. chn’sops nests that we observed. The earliest nest was initiated (i.e., first egg laid) on 13 March and the latest on 15 May with a mean initiation date of 11 April ± 2.2 days (n = 61) across years (Fig. 2). Nest-building activity began in February. Two nests that fledged chicks in early May 2006 were re-used in February 2007 with new material added. Eggs and Clutch Size. — Eggs were elliptical in shape and cream-white with liver-brown spots and markings mostly concentrated in a ring toward the obtuse end (Fig. 1). Eggs were 17.17 ± 0.25 mm in length by 13.42 ± 0.1 1 mm in width (/? =11). Eggs weighed between 1.286 and 1.861 g with a mean mass of 1.616 ± 0.020 g (/? = 48). Fresh egg weight represented 18% of mean adult body mass of 9.0 ± 0.2 g {n = 25). Clutch size was usually two with only one of 45 nests found prior to incubation having a single egg, yielding a mean clutch size of 1.98 ± 0.02 eggs (n = 45). The mean number of days observed between nest completion and laying of 692 THE WILSON JOURNAL OL ORNITHOLOGY • Vul. 122, No. 4, December 2010 Early Middle Late Incubation period ■> CD C 'c o w > o 25 20 15 10 5 0 B • • Si** • : 5 10 15 Nestling age (day) 20 FIG. 3. Average (A) nest attentiveness, and (B) on- and off-bout durations across three periods of incubation: early (days 1-5), middle (days 6-11), and late (days 12-18). Sample sizes reflect numbers of nests. FIG. 4. The change with nestling age in (A) female brooding behavior (% time spent brooding) and (B) rates that parents visit the nest to provision nestlings. the first egg was 4.71 ± 0.47 (n = 1). Eggs were laid in the morning on alternate days. Incubation. — Females incubated the eggs with- out help from males. Males were rarely observed visiting nests during incubation, and did so at only two of 40 nests with video observations of 6-8 hrs each, yielding an average of 0.01 ± 0.01 visit.s/hr (n = 40). Females were not observed incubating prior to laying the last egg and eggs hatched synchro- nously. Nest attentiveness averaged 66.0 ± 1.6% {n = 40) and was lower (ANOVA, F2.37 = 6.8, P = 0.003) in early incubation compared with middle and late incubation periods (Fig. 3A). Incubation on-bouts averaged 33.35 ± 1.94 min (n = 40) and did not change over the incubation period (ANOVA, F2.37 = 1.1, F = 0.3; Fig. 3B). Length of off-bouts averaged 16.19 ± 1.59 min (n = 40) and decreased from early to later incubation stages (ANOVA. F2.37 = P = 0.055; Fig. 3B). Incubation behavior yielded an average 24-hr egg temperature of 34.88 ± 0.45° C (/? = 7 ne.sts, 43 days of .sampling). The incubation period was 16.9 ± 0.3 days (/? = 10) for nests found prior to beginning of incubation and where exact hatch was observed. Nestling Period. — Both adults were observed provisioning chicks and the male was also observed passing food to the female when she was brooding. Adults regularly fed the nestlings fruit that appeared to be mistletoe berries (possibly Antidaphne viscoidea or Phoradendron spp.). which were regularly found stuck near the beak of chicks. Females regularly regurgitated seeds when incubating and brooding, taking care to eject them outside the nest. The percentage of time females brooded decreased through the nestling period (r == —0.86, P < 0.001), and stopped after the eighth primary pin feather broke its sheath on days 10-11 (Fig. 4A). Provisioning rates increased over the nestling period (r = 0.80, P < O.OOl, Fig. 4B). Parents visited the nest an average of 4.41 ± 0.65 trips/hr (n = 10) during days I and 2 after hatching, and 14.93 ± 2.36 trips/hr (/; = 6) at pin-break. Chicks had gray down distributed lightly over the head and body, and orange skin and beak at hatching (Fig. I). The nestling period for nests where exact hatch and fledging days were Goiilcling and Marlin • GOLDEN-FACED TYRANNULET BREEDING BIOLOGY 693 10 1 8 - S 6 - c/) CO I 4 2 ^ 0 A .1* ..I'* .i • i I • • /c= 0.285 + 0.011 E E O) c CO CO CT3 18 15 - 12 - 9 - 6 - O) c 0) ■O i— o o O) c 3 55 50 45 40 35 30 25 20 15 10 5 B !• :ii M :i /f= 0.200 + 0.012 .1 • • • • • • • • • I • A-= 0.201 + 0.010 0 5 10 15 Nestling age (day) FIG. 5. Relation.ship.s of (A) mas.s, (B) tarsus length, and (C) wing chord length plotted against age for Golden- faced Tyrannulets and their estimated growth rate constants (A). The dashed lines represent mean adult sizes. observed ranged from 1 7 to 19 days with an average of 18.0 ± 0.2 days (n = 9). Nestling mass on days 10 and 1 1 (i.e., pin-break) was 6.966 ± 0.152 g in = 9). The growth rate constant (A.) for body mass was low (Fig. 5 A) and nestlings were close to mean adult weight of 9.0 ± 0.2 g (/; = 25) at fledging (Fig. 5A). Growth rate based on tarsus length was slower at k = 0.200 ± 0.012 and tarsus size was essentially the same length as adults at 16.26 ± 0.17 mm (/; = 22) at fledging (Fig. 5B). Growth rate based on wing chord was k = 0.201 ± 0.010, but wing chord at fledging was less than adult size (Fig. 5C) of 52.03 ± 0.62 mm in = 20). Nesting Success and Predation. — Predation accounted for 91.6% of failures with remaining failures attributed to nests failing, abandonment, and poor nest condition that allowed chicks to fall out (re-used nest). Birds seemed to be a main predator of nests with punctured egg remains at times found near the nest. Nest predation often coincided with presence of army ants iLabidiis praedator) and large mixed-bird flocks feeding near the nest. One nest was also filmed being predated by a capuchin monkey iCehus oliva- ceus). Snakes were not recorded in video observations and rarely observed at this elevation in Yacambti across all years (TEM, pers. obs.). The total daily predation rate was 0.057 ± 0.007 in = 989 exposure days) with a total daily mortality rate of 0.064 ± 0.008. An estimated 12% of nests were successful based on a total nesting period of 37 days. The daily nest predation rate during egg-laying (0.052 ± 0.025; n = 76.5 exposure days) did not differ iX- = 0.4, P = 0.6) from the incubation period (0.068 ± 0.010; n = 575.5 exposure days). However, the daily predation rate during the nestling period (0.039 ± 0.010; n = 377 exposure days) was lower iX~ = 4.2, P = 0.029) than during incubation. DISCUSSION Golden-faced Tyrannulets in Yacambti Nation- al Park exhibited breeding biology parameters different from most typical tyrannids (Fitzpatrick 2004). The nest appears globular, as also de- scribed for congeners (e.g., Z. acer [Beebe et al. 1917], Z. iinprohiis [Hilty 2003 j. and Z. vili ssinnis jSkutch 1960]), but varied in appearance depend- ing on location of attachment and amount of exposed nest. Nests of Z. clirysops, compared with congeners, were slightly larger than those of Z. vilissinnis (Skutch 1960), but had a shallower nest-CLip than ob.served for Z. acer (Beebe et al. 1917; this was a single observation). Nest construction and materials were virtually the same as those observed for Z. vilissinnis, including the soft seed-aril lining (Skutch I960), and similar to Z. acer (Beebe et al. 1917). The use of epiphytic 694 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4. December 2010 plant resources for nesting by neotropical birds is well recognised (Nadkarni and Matelson 1989, Fierro-Calderon and Martin 2007). Golden-faced Tyrannulets show flexibility in nesting height with 71% of nests in Yacambil below the 8-12 m height range observed in Colombia (Hilty and Brown 1986), and a nest recorded ~45 m above ground in an emergent Ceiba tree in Ecuador (Cisneros-Heredia 2006). Our observations do not seem to reflect biases towards finding low nests because we found the majority of nests through parental behavior. We found nests to the tops of canopies by watching parental behavior. The lower height of nests at our site might reflect the lower stature of the disturbed forest areas and edges in which the species occurred at our site. Differences between nesting height in disturbed or regrowth areas and primary forest have been observed in other tyrannids (e.g., Z. vilissinuis [Skutch 1960] and Todiroslnim chrysocrotaphum [Hilty and Brown 1986]), indi- cating structural differences in habitat influence nest height variation. Nest architecture seems to be an evolutionarily conservative character that has been used to indicate phylogenetic relationships (Mobley and Prum 1995, Sheldon and Winkler 1999, Zys- kowski and Prum 1999). Enclosed nests are thought to be a derived character from the more common open nests reported for the majority of tyrannids (Collias and Collias 1984, Fitzpatrick 2004). It is also thought to be a character that evolved independently in different tyrannid sub- families (Fitzpatrick 2004). The sparse available information on nests of species of Zimmerius indicates only subtle differences between conge- ners. However, nests of the majority of species in the genus have yet to be investigated. Nest similarity may support a close relationship be- tween Camptostoma (Haverschmidt 1954) and Zimmeriii.s (unless evolved independently; Lanyon 1988). These two genera, together with the closely related Phyllomyias virescens and others in that genus that may be discovered to have enclosed nests (Fitzpatrick 2004, Ohlson et al. 2008, Rheindt et al. 2008), repre.sent the only enclosed nest-builders in the elaeniine clade proposed by Ohl.son et al. (2008). Despite differences in external materials, they have similarly attached and shaped immobile nests lined with ‘plant wool’ (not feathers) with a simple side-opening near the top (Beebe et al. 1917, Haverschmidt 1954, Skutch 1960, this study). The observed peak of breeding activity of Z. chtysops is just prior to and overlapping the peak precipitation period from late April through July in Yacambu (Fieiro-Calderon and Martin 2007). This is a pattern in avian breeding biology observed at other neotropical sites (Skutch 1950, 1960; Marchant 1959; Best et al. 1996; Medeiros and Marini 2007), including some species in lowland habitats in Venezuela (Cruz and Andrews 1989). This peak fits within the general breeding period of January to June described for tyrant- flycatchers north of Peru (Fitzpatrick 2004). The season is relatively short compared with other passerine species in this area (e.g., Biancucci and Martin 2008, Niklison et al. 2008, Cox and Martin 2009). Skutch (1950) found from his observations in Central America that, in general, avian populations at higher altitudes exhibited a more marked and narrow breeding season than in lowlands. Z. chrysops is thought to breed in Colombia from April until November at the lower elevation of 1,000 m asl (Hilty and Brown 1986). A record in February from Eoja Province in Ecuador also reports nesting activity by the species at a similar altitude to Yacambu sites (Best et al. 1996). The short season and absence of any observations of birds re-nesting in the same season after a successful breeding attempt indi- cates this species is single-brooded, typical of tyrannids (Eitzpatrick 2004). Eggs of Z. chtysops conform to the typical egg coloration observed in most tyrannids (Von Ihering 1904, Skutch 1960, Wetmore 1972). They are slightly larger than an egg of Z. acer measured by Beebe et al. (1917) and slightly shorter and wider than eggs of Z. vilissimus observed by Skutch (1960). The clutch size of two is on the lower edge of the range of 2-6 observed in the family and the 2^ of most tropical tyrannids (Fitzpatrick 2004). The 17-day incubation period is long relative to the range of 12 to 16 days known for most temperate and tropical tyrannids, and approaches a few unusual tropical species that have suspend- ed pendulant nests and long incubation periods (Fitzpatrick 2004). However, the incubation period was near the average for 33 other non- tyrannid species studied in Yacambii National Park (Martin and SchwabI 2008). The 18-day nestling period akso is in the upper range for tyrannids, which is from 12 to 24 days with most being 14 to 17 days (Fitzpatrick 2004). Our observations support that mistletoe fruit in the Gouh/ing and Martin • GOLDEN-FACED TYRANNULET BREEDING BIOLOGY 695 area (e.g.. Antidapline viscoidew, Restrepo et al. 2002, Kelly et al. 2004) comprise an important portion of the diet of nestlings, as with Z. vilissimus (Skutch 1960), and not only for adults tor which this has already been observed with others in the genus (Fjeldsa and Krabbe 1990, Alonso and Whitney 2001, Fitzpatrick 2004). Slow growth has been found in other tropical species that feed fruit to their offspring (Morton 1973). A comparison of the growth rate constant (A.) for mass with that published and summarized for other tropical and temperate species (Ricklefs 1976, Willis et al. 1978, Oniki and Ricklefs 1981, Starck and Ricklefs 1998, Remes and Martin 2002) reveals it is slow even for the slowest growing of tropical tyrannids {k < 0.3). Tropical tyrannids from Central America and Brazil had growth rates (k) extending from a low of 0.248 in another species {Mionectes macconnelli) that builds enclosed nests (Willis et al. 1978) to rates >0.4, similar to those observed for some faster growing tyrannids breeding in North America such as Tyrannus tyranmis {k = 0.438) and Sayornis phoebe (k = 0.425) (Murphy 1981), Tyrannus verticalis {k = 0.416) and T. forficatiis (k = 0.394) (Murphy 1988), Empidonax minimus (k — 0.499-0.505) (Briskie and Sealy 1989), and E. oberholseri (k = 0.425) (Pereyra and Morton 2001). Other non-tyrannid passerines investigated in Yacambii National Park also had much higher growth rates (Niklison et al. 2008, Biancucci and Martin 2008, Cox and Martin 2009). Incubation temperature is a reflection of parental attentiveness with cooler temperatures (low attentiveness) associated with longer incu- bation period lengths (Lyon and Montgomerie 1985, Martin 2002, Hepp et al. 2006, Martin et al. 2007). The observed 24-hr incubation temperature for Z. chrysops is below the optimal temperature range for embryonic growth (White and Kinney 1974, Webb 1987). However, egg temperature was close to the expected value for a tropical species with this length of incubation period and nest attentiveness (Martin et al. 2007: figure 2). Skutch (I960) considered Z. vilissimus in Costa Rica to have high nest attentiveness for a .small flycatcher. He observed mean on and off-bouts during incubation of 32.5 and 12.5 min, respectively with overall estimates of attentiveness of 67 and 72%. This indicated higher attentiveness (and shorter off-bouts) for this congener than for Z. dvysops in Yacambti National Park. Attentiveness for Z. chrysops was in the middle of the range of 39 other species studied in Yacambu (Martin and SchwabI 2008), but was lower than for most temperate passerines (Conway and Martin 2000: appendix 1 ; Martin 2002; Chalfoun and Martin 2007). Z. vilissimus had higher provisioning rates than those observed for Z. chrysops (Skutch 1960). However, both had mean feeding rates at least two-fold higher than other species studied in Yacambu National Park (Fierro-Calderon and Martin 2007, Biancucci and Martin 2008, Nikli- son et al. 2008, Cox and Martin 2009). This could reflect compensation for lower protein levels in fruit fed to nestlings (Morton 1973), although flycatchers seem to feed at higher rates than other species (Martin et al. 2000b). Consequently, predation levels could be influenced through greater nest detection due to increased parental visits (Skutch 1949; Martin et al. 2000a, b). However, predation rates were lower during the nestling period than during earlier periods when parents were less active. Nesting success for Z. chrysops was among the lowest observed for a member of the Tyrannidae when compared with other southern hemisphere and neotropical species. Skutch (1985) summa- rized data from different species in the humid Neotropics showing that 39.7% of nests found were successful with tyrannids exhibiting a range of 12.1 to 50%. Mionectes oleagineus in Trinidad had the lowest tyrannid nesting success (12.1%), a trend observed in other small species in Trinidad (Snow and Snow 1979). Low reproductive estimates were observed in Suiriri islerorum (>13.6%, mean = 20.8%) and S. ajfmis (19%) in tropical savannah of Brazil (Lopes and Marini 2005, Franca and Marini 2009). In Costa Rica. 36% of nests of Zimmerius vilissimus were successful and fledged chicks (Skutch 1985). Other non-tyrannid neotropical species in Panama were observed to have low nesting success similar to Z. chtysops (e.g., some formicariids and a thamnophilid species; Robinson el al. 2000). However, nesting success was generally higher (23-74%) and daily predation rates lower for other tyrannid species in the Neotropics and southern hemisphere; e.g., Argentina (Mezquida and Marone 2001, Auer et al. 2007). Puerto Rico (Toixes Baez and Collazo 1992), Brazil (Willis et al. 1978, Aguilar et al. 2000, Medeiros and Marini 2007), and Panama (Robinson et al. 2000, Dyrcz 2002). The extremely poor nesting success of Z. chrysops reflects high predation pressure, and may reflect the tendency for this species to nest 696 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4, Deceniber 2010 along forest edges where predation levels are high (Sdderstrom 1999). Multiple re-nest attempts would be expected to occur to compensate for high predation (Martin 1995, Fitzpatrick 2004), but the short breeding season (i.e.. Fig. 2) and our observations suggest that only 2-3 nesting attempts are made. The small clutch size of Z. chrysops with a relatively long incubation period and slow nest- ling growth suggest this species is at the ‘slow’ end of the slow-fast life history gradient (Martin 1996, 2004). The solely neotropical distribution, lack of morphological specialization (Traylor 1977), and indication that only conservative differences in nest architecture occur within the genus, place importance on investigating repro- ductive parameters of the diversity of Zimmeriiis species. Phenotypic similarity or dissimilarities would be valuable for understanding the finer scale processes influencing reproduction and broader life history variation within tyrannids; e.g., localized predation rates imposed upon otherwise phylogenetically close and similar species (Skutch 1949, 1985; Bosque and Bosque 1995; Martin 1995; Martin et al. 2000a). ACKNOWLEDGMENTS We thank L. A. Biancucci, Karolina Fieno-Calderon, A. M. Niklison. M. J. Foguet. A. A. Majewska, and R. A. Ruggera for discussions and assistance in the field. We also thank John Kanowski, Elena Arriero. Marline Maron. and C. A. McAlpine for help with the manu.script. This study was made possible in part by support under NSF grants DEB-()543178 and DEB-0841764 to T. E. Martin. Permit numbers were DM/0000237 from FONACIT. PA-INP-005- 2004 from INPARQUES, and 01-03-03-1 147 from Mini.s- terio del Ambiente. Specific equipment identities are provided to aid specific methods and do not represent an endorsement of these companies by USGS. LITERATURE CITED Aguilar. T. M., M. Maldonado-Coelho, and M. A. Marini. 2000. Nesting biology of the Gray-hooded Flycatcher {Mionectes niflveiitri.s). Ornitologia Neo- tropical 1 1 :223-230. Alonso. J. A. and B. M. Whitney. 2001. A new Zimmerin.s tyrannulet (Aves: Tyrannidae) from white sand forests of northern Amazonian Peru. Wilson Bulletin I 13:1-9. Auer. S. K.. R. D. Bassar. J. J. Fontaine, and T. E. Martin. 2007. Breeding biology of passerines in a subtropical montane forest in northwestern Argentina. Condor 109:321-333. Beebe. W.. G. I. Hartley, and P. G. Howes. 1917. Tropical wildlife in British Guiana. Volume 1. New York Zoological Society, New York, USA. Best, B. J., M. Checker. R. M. Thewlis, A. L. Best, and W. Duckworth. 1996. New bird breeding data from southwestern Ecuador. Omithologia Neotropical 7:69- 73. Biancucci, L. A. and T. E. Martin. 2008. First description of the breeding biology and natural history of the Ochre- breasted Brush Finch (Atlapeles seminifus) in Vene- zuela. Wilson Journal of Ornithology 120:856-862. Bosque, C. and M. T. Bosque. 1995. Nest predation as a selective factor in the evolution of development rates in altricial birds. American Naturalist 145:234-260. Briskie. j. V. and S. G. Sealy. 1989. Determination of clutch size in the Least Elycatcher. Auk 106:269-278. Chalfoun, a. and T. E. Martin. 2007. Latitudinal variation in avian incubation attentiveness and a test of the food limitation hypothesis. Animal Behaviour 73:579-585. Cisneros-Heredia, D. a. 2006. Notes on breeding, behaviour and distribution of some birds in Ecuador. Bulletin of the British Ornithologists Club 126:153- 164. COLLIAS, N. E. AND E. C. COLLIAS. 1984. Nest building and bird behavior. Princeton University Press, Princeton, New Jersey, USA. Conway, C. J. and T. E. Martin. 2000. Evolution of passerine incubation behavior: inlluence of food, temperature, and nest predation. Evolution 54:670- 685. Cox. W. A. AND T. E. Martin. 2009. Breeding biology of the Three-Striped Warbler in Venezuela: a contrast between tropical and temperate parulids. Wilson Journal of Ornithology 121:667-678. Cruz, A. and R. W. Andrews. 1989. Observations on the breeding biology of passerines in a seasonally flooded savanna in Venezuela. Wilson Bulletin 101:62-76. Dyrcz, a. 2002. Breeding ecology of the Social (Myioze- tetes .'iiniili.s) and Rusty-margined (M. cayanensis) flycatchers at Barro Colorado Island, Republic of Panama. Ornitologia Neotropical 13:143-151. Fierro-CalderOn, K. and T. E. Martin. 2007. Repro- ductive biology of the Violet-chested Hummingbird in Venezuela and comparisons with other tropical and temperate hummingbirds. Condor 109:680-685. Fitzpatrick, J. W. 2004. Family Tyrannidae (Tyrant Flycatchers). Pages 170-257 in Handbook of the birds of the world. Volume 9. Cotingas to pipits and wagtails (J. del Hoyo. A. Elliott, and D. A. Christie, Editors). Lynx Edicions, Barcelona, Spain. FjeldsA, j. and N. Krabbe. 1990. Birds of the High Andes. Zoological Museum. University of Copenhagen. Den- mark. Franca, L. F. and M. A. Marini. 2009. Low and variable reproductive success of a neotropical tyrant- flycatcher. Chapada Flycatcher (Siiiriri islerorinn). Emu 109:265-269. Haver.schmidt. F. 1954. The nesting of the Ridgway Tyrannulet in Surinam. Condor 56:139-141. Hensler, G. L. and j. D. Nichol.s. 1981. The Mayfield method of estimating nesting success: a model, cstimalors and simulation results. Wilson Bulletin 93:42-53. Gouhliiii' and Marlin • GOLDEN-FACED TYRANNULET BREEDING BIOLOCJY 697 Hepp, G. R.. R. a. Kennamer. and M. H. Johnson. 2006, Maternal effects in Wood Ducks: incubation temper- ature influences incubation period and neonate phe- notype. Functional Ecology 20:307-314. Hilty, S. J. 2003. Birds of Venezuela. Second Edition, Princeton University Press, Princeton, New Jersey, USA. Hilty, S, J, and W. L. Brown. 1986. A guide to the birds of Colombia. Princeton University Press, Princeton, New Jersey, USA. Johnson, D. H. 2007. Estimating nest success: a guide to the methods. Studies in Avian Biology 34:65-72. Kelly, D, L., G. O'Donovan, J. Feehan, S. Murphy, S. O. Drangeid, and L. Marcano-Berti. 2004. The epiphyte communities of a montane rain forest in the Andes of Venezuela: patterns in the distribution of the flora. Journal of Tropical Ecology 20:643-666. Lanyon, W. E. 1988. A phylogeny of the thirty-two genera in the Elaenia assemblage of tyrant flycatchers. American Museum Novitates 2914:1-57. Lopes, L. E. and M. A. Marini. 2005. Low reproduc- tive success of Campo Suiriri {Siiiriri ajfmis) and Chapada Flycatcher (S. isleronim) in the central Brazilian Cerrado. Bird Conservation International 15:337-346. Lyon, B. E. and R. D. Montgomerie. 1985. Incubation feeding in Snow Buntings: female manipulation or indirect male parental care? Behavioral Ecology and Sociobiology 17:279-284. Marchant, S. 1959. The breeding season in S.W. Ecuador. Ibis 101:137-152. Martin, T. E. 1995. Avian life history evolution in relation to nest sites, nest predation and food. Ecological Monographs 65:101-127. Martin, T. E. 1996. Life history evolution in tropical and south temperate birds: what do we really know? Journal of Avian Biology 27:263-272. Martin, T. E. 2002. A new view for avian life history evolution tested on an incubation paradox. Proceed- ings of the Royal Society of London, Series B 269:309-316. M.artin. T. E. 2004. Avian life-history evolution has an eminent past: does it have a bright future? Auk 121:289-301. Martin, T. E. and C. K. Ghalambor. 1999. Males feedini^ females during incubation. 1. Required by microcli- mate or constrained by nest predation? American Naturalist 153:131-139. Martin, T. E. and H. Schwabl. 2008. Variation in maternal effects and embryonic development rates among passerine species. Philo.sophical Transactions of the Royal Society of London, Series B 363:1663-1674. Martin, T. E., J. Scott, and C. Menge. 2000a. Nest predation increases with parental activity: .separating nest site and parental activity effects. Proceedings of the Royal Society of London. Series B 267:2287- 2293. Martin, T. E., P. R. Martin, C. R. Olson, B. J. Heidinger, and j. j. Fontaine. 2000b. Parental care and clutch sizes in North and South American birds. Science 287:1482-1485. Martin, T. E., S. K. Auer, R. D. Bassar, A. M. Niklison, AND P. Lloyd. 2007. Geographic variation in avian incubation periods and parental influences on embry- onic temperature. Evolution 61:2558-2569. Mayfield, H. 1961. Nesting success calculated from exposure. Wilson Bulletin 73:255-261. Mayfield, H. F. 1975. Suggestions for calculating nest success. Wilson Bulletin 87:456^66. McGowan, K. J. 2004. Introduction: the world of birds. Pages 1.1-1.114 in Handbook of bird biology (S. Podulka. R. Rohrbaugh Jr., and R. Bonney. Editors). Cornell Laboratory of Ornithology, Ithaca, New York. USA. Medeiros, R. C. S. and M. A. Marini. 2007. Biologia reprodutiva de Elaenia chiriqnensis (Lawrence) ( Aves, Tyrannidae) em Cerrado do Brasil Central. Revista Brasileira de Zoologia 24:12-20. Mezquida, E. T. and L. Marone. 2001. Factors affecting nesting success of a bird assembly in the central Monte Desert, Argentina. Journal of Avian Biology 32:287-296. Mobley, J. A. and R. O. Prum. 1995. Phylogenetic relationships of the Cinnamon Tyrant, Neopipo cinnamomea, to the tyrant flycatchers. Condor 97:650-662. Morton, E. S. 1973. On the evolutionary advantages and disadvantages of fruit eating in tropical birds. American Naturalist 107:8-22. Murphy, M. T. 1981. Growth and aging of nestling Eastern Kingbirds and Eastern Phoebes. Journal of Field Ornithology 52:309-316. Murphy, M. T. 1988. Comparative reproductive biology of kingbirds (Tyrannus spp.) in eastern Kansas. Wilson Bulletin 100:357-376. Nadkarni, N. M. and T. j. M.atelson. 1989. Bird use of epiphyte resources in neotropical trees. Condor 91:891-907. Nice, M. M. 1954. Problems of incubation periods in North American birds. Condor 56:173-197. Niklison, A. M., J. I. Areta, R. A. Ruggera. K. L. Decker. C. Bosque, and T. E. Martin. 2008. Natural history and breeding biology of the Rusty-breasted Antpitta [Grallariciila fernigineipectiis). Wilson Jour- nal of Ornithology 120:345-352. Ohlson, j., j. FjeldsA, .and P. G. P. Ericson. 2008. Tyrant flycatchers coming out in the open: phylogeny and ecological radiation of Tyrannidae (Aves, Passeri- formes). Zoologica Scripta 37:315-335. Oniki, Y. and R. E. Ricklefs. 1981. More growth rates of birds in the humid New World tropics. Ibis P3-349- 353. Palmer, R. S. (Editor). 1962. Handbook of North American birds. Yale University Press. New Haven. Connecticut. USA. Pereyra, M. E. and M. L. Morton. 2001. Nestling growth and thermoregulatory development in subalpine Dusky Flycatchers. Auk 118:116-136. Preston, F. W. 1953. The shapes of bird eggs Auk 70:160-182. RemeS, V. and T. E. Martin. 2002. Environmental influences on the evolution of growth and develop- ment rates in pas.serines. Evolution 56:2505-2518. 698 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 4, December 2010 Restrepo, C., S. Sargent, D. J. Levey, and D. M. Watson. 2002. The role of vertebrates in the diversification of New World mistletoes. Pages 83- 98 in Seed dispersal and frugivory: ecology, evolution, and conservation (D. J. Levey, W. R. Silva, and M. Galetti, Editors). CAB International, Wallingford, United Kingdom. Rheindt, F. E., j. a. Norman, and L. Christidis. 2008. Phylogenetic relationships of tyrant-flycatchers ( Aves: Tyrannidae), with an emphasis on the elaeniine assemblage. Molecular Phylogenetics and Evolution 46:88-101. Ricklefs, R. E. 1967. A graphical method of fitting equations to growth curves. Ecology 48:978-983. Ricklefs, R. E. 1976. Growth rates of birds in the humid New World tropics. Ibis 118:179-207. Ridgely, R. S. and G. Tudor. 1994. The birds of South America, Volume 2. The suboscine passerines. University of Texas Press, Austin, USA. Robinson, W. D., T. R. Robinson, S. K. Robinson, and J. D. Brawn. 2000. Nesting success of understory' forest birds in central Panama. Journal of Avian Biology 31:151-164. Sheldon. F. H. and D. W. Winkler. 1999. Nest architecture and avian systeinatics. Auk 1 16:875-877. Skutch, a. F. 1949. Do tropical birds rear as many young as they can nourish? Ibis 91:430^55. Skutch, A. F. 1950. The nesting seasons of Central American birds in relation to climate and food supply. Ibis 92:185-222. Skutch, A. F. 1960. Life histories of Central American birds. Volume 2. Pacific Coast Avifauna Number 34. Skutch, A. F. 1985. Clutch size, nesting success, and predation on nests of neotropical birds, reviewed. Ornithological Monographs 36:575-594. Snow, B. K. and D. W. Snow. 1979. The Ochre-bellied Flycatcher and the evolution of lek behavior. Condor 81:286-292. SoderstrOm, B. 1999. Artificial nest predation rates in tropical and temperate forests: a review of the effects of edge and nest site. Ecography 22:455-463. SPSS. 2006. Version 15.0.0. SPSS Inc., Chicago, Illinois, USA. Starck, j. M. and R. E. Ricklefs. 1998. Data set of avian growth parameters. Pages 381-423 in Avian growth and development (J. M. Starck and R. E. Ricklefs, Editors). Oxford University Press, New York, USA. Stutchbury, B. j. M. and E. S. Morton. 2001. Behavioral ecology of tropical birds. Academic Press, San Diego, California, USA. Torres Baez, P. and J. A. Collazo. 1992. Biologia reproductiva y datos preliminares sobre patrones de crecimiento del Jui de Puerico {Myicircinis antillarum) en Cabo Rojo, Puerto Rico. Caribbean Journal of Science 28:30-37. Traylor Jr., M. A. 1977. A classification of the tyrant Ilycatchers (Tyrannidae). Bulletin of the Museum of Comparative Zoology 148:129-184. Von Ihering, H. 1904. The biology of the Tyrannidae with respect to their systematic arrangement. Auk 21:31 3-322. Weathers, W. W. and K. A. Sullivan. 1989. Nest attentiveness and egg temperature in the Yellow-eyed Junco. Condor 91:628-633. Webb, D. R. 1987. Thermal tolerance of avian embryos: a review. Condor 89:874-898. Wetmore, a. 1972. Birds of the Republic of Panama. Part 3. Smithsonian Institution Press, Washington, D.C., USA. White, F. N. and J. L. Kinney. 1974. Avian incubation. Science 186:107-1 15. Willis, E. O.. D. Wechsler, and Y. Oniki. 1978. On behavior and nesting of McConneH’s Flycatcher (Pipromorpha macconnelli): does female rejection lead to male promiscuity? Auk 95:1-8. Zyskowski, K. and R. O. Prum. 1999. Phylogenetic analysis of the nest architecture of neotropical Ovenbirds (Furnariidae). Auk 1 16:891-91 1. The Wilson Journal of Oniithology 122(4):699-7()5. 2010 HABITAT TYPE IS RELATED TO NEST MASS AND FLEDGING SUCCESS OF ARCTIC WARBLERS JULIE C. HAGELIN,'’^ MARGARET L. PERRY,' BEN S. EWEN-CAMPEN,'-^ DEREK S. SIKES,- AND SUSAN SHARBAUGH^ ABSTRACT. Relatively little is known about Arctic Warblers (Phylloscopus borealis) that breed in central Alaska. We monitored Arctic Warbler populations in two adjacent but distinct habitat types in central Alaska (high elevation, 'open shrub and lower elevation, ‘dense shrub’). We collected 95 nests over three breeding seasons to learn more about nest- building behavior, nest mass, composition, fledging success, and nest parasites. Females were the sole builders of ground nests, which were primarily comprised ot moss, grass, and a lining of moose (Aices alces) hair. Dry weight of nests was 20 g. but differed up to ~3-fold within each habitat type each season. Nests from open shrub habitats were more massive and contained less moose hair lining than nests in dense shrub. Open shrub nests fledged more young during the most productive breeding season. We report the first record of the parasitic blowfly Protocalliphora tundrae in Arctic Warbler nests and for Alaska. Blowfly parasitism (55% of nests with hatchlings) was similar in both habitat types and did not correlate with fledging success, or nest mass. Nests with greater amounts of moss tended to have lower levels of blowfly infestation. Received 26 September 2009. Accepted 10 June 2010. The Arctic Warbler (Phylloscopus borealis) is an understudied old-world warbler that winters in southeast Asia and has an expansive subarctic breeding range that extends from Norway through Siberia into central Alaska (Tucker 1949, Price and Beck 1989, Lowther and Sharbaugh 2008). Little is known about its cryptic ground-nesting habits throughout its range. Alaska is the only location in North America where Arctic Warblers (Fig. 1 A) breed and these populations are consid- ered a separate subspecies (P. b. kennicotti) (Lowther and Sharbaugh 2008, Reeves et al. 2008). Documentation of Arctic Warbler nests in Alaska prior to our observations included a report on 1 1 nests near Nome (Price and Beck 1989) and two nests in Denali National Park by Murie (1956). Our objectives are to present detailed information about: ( 1 ) nest construction, (2) orientation, (3) composition, (4) fledging success, and (5) nest parasites in two adjacent, but distinct habitat types (high elevation, ‘open shrub’ and low elevation, ‘dense shrub’) in a central Alaskan population of P. b. kennicotti. ' Swarthmore College Department of Biology, Swarth- more, PA 19081, USA. "University of Alaska, Museum of the North, Fairbanks, AK 99775, USA. ■’Alaska Bird Observatory, Fairbanks, AK 99701, USA. ■’Department of Organismic and Evolutionary Biology, Harvard University, Cambridge, MA 02138, USA. ’Corresponding author; e-mail: jhageli 1 @swarthmore.edu METHODS Study Site. — We studied a breeding population of Arctic Warblers between 2004 and 2006 in the Tangle Lakes region of the Denali Highway in interior Alaska (63° N, 146° W). All data were collected from four 10-ha plots, each 150 m from the Denali Highway at mileposts 23, 26, 30, and 33, respectively. Pairs of plots represented two distinct habitat types of the region. ‘Open shrub’ plots were at higher elevation (1,121-1,154 m) and included north and east-facing slopes. This habitat type contained meadow-like clearings suiTounded by 1-2 m tall willow shrubs (Sali.x spp.) and dominated by gramminoids (Deschamp- sia caespitosa and Carex spp.), lupines (Lupinus arcticus), and burnet (Sanguisorba spp.). Nests were collected from open shrub plots in all 3 years. Dense shrub plots were at lower elevations (910-1,012 m) with flat ground or a gradual, east- facing slope. This habitat type was dominated by dwarf birch (Betula nano) and willow. Nests were collected in dense shrub habitats only in 2004 and 2005, as birds typically nested in these locations at much lower densities. Nest Monitoring, Collection, and Composi- tion.—VJc located most nests by systematically searching the ground in areas where we observed singing males. Each nest was monitored at least once every 4 days through laying, hatching, and rearing of young. Daily nest checks began 2 days prior to the expected fledging date to ascertain the number of young successfully fledged. Observations of nest building were made each season and we collected each nest once fledging 699 700 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4. December 2010 KIG. I. (A) Color-banded adult Arctic Warbler (Phylloscopiis borealis keniiecolli). Males and females are monomorphic. (Photograph by Ron Teel). (B) Opening of Arctic Warbler nest in the side of a tussock in open shrub habitat. (Photograph by Sue Guers). (C) Ne.st with fledglings associated with willows in dcn.se shrub habitat. (Photograph courtesy of Alaska Bird Ob.servatory). Ilagcliii Cl al. • ARCTIC WARBLER NEST MASS 701 was complete (between 20 Jul and I Aug). We first recorded nest orientation (in 2005 and 2006 only) by obtaining a compass bearing from a stick inserted straight into the nest opening prior to removing the nest from the ground. Nests are a domed mass of material that can be removed by lifting up the sides and the bottom, once the edge of the nest is located within the suiTounding tundra. We carefully removed non-nest material that clung to the outside of the nest during extraction and retrieved any bits of nest material left in the depression. Each nest was placed in a separate plastic collection container for transport to the laboratory, then transferred to a paper sack and dried overnight in an oven at 135° C. Dry nest mass was measured using an analyt- ical balance. Two key components of nests were measured each year: nest mass and mass of moose {Alces dices) hair, which comprised the lining of many nests. Our relatively large sample of nests from open shrub habitat (n = 80) allowed us to examine differences across years. Twenty-six nests from open shrub habitat were dissected further in 2006 to quantify the mass of moss and grass, two key components of nest structure. We calculated the percent composition of hair, moss, or grass as a proportion of nest mass. A snowstorm on 21 June 2006 caused birds to abandon the eight nests (and clutches) that we had located at that point in the season. However, Arctic Warblers began re-nesting after snow began to melt (on 23 Jun 2006) and we subsequently located 22 new nests. We collected all 30 nests, including the eight that had been abandoned. This sample enabled us to examine any differences between nests built before and after the storm (e.g., With and Webb 1993). Nest Parasites, Incidental Arthropods, and Fledging Success. — We removed, identified, and counted all puparia of an abundant nest parasite (a blowfly) after nests were dried but before they were weighed. We also identified incidental arthropods. Puparia are the rigid outer shells of developing blowfly larvae. The larval stage obligately feed on the blood of nestlings (Sab- rosky et al. 1989). Counting puparia is a conservative measure of parasitism, because it avoids double-counting flies that have successful- ly pupated into adults. All nests that failed prior to hatching were excluded from analyses (n = 8 that failed during the Jun 2006 storm, n — 3 depredated or abandoned during other years), as puparia only occur after blowfly larvae have fed on nestlings. All arthropod specimens were deposited in the University of Alaska Museum Insect Collection. Statistical Analyses. — The mass of nest com- ponents and parasite counts exhibited strongly non-normal distributions, which made parametric statistics inappropriate. All data are reported as medians and 25-75% interquartile ranges [IQR]. We used Wilcoxon rank sum tests to examine differences in nest construction and fledging success between habitat types. We used Rayleigh tests to ascertain whether orientation of nest openings deviated from a random distribution (Wilkie 1983). We examined differences between years within each habitat type using Kruskal- Wallis ANOVA. We used Spearman's rho (p) tests to assess non-parametric correlation(s), including relationship(s) between nest mass, different nest components, fledging success, fledge date, and number of puparia per nest. RESULTS Nesting Behavior. — Our observations are based on 95 nests (/? = 80 from open shrub habitat, n = 1 5 from dense shrub habitat) collected over 3 years (n = 24 in 2004, n = 41 in 2005, n = 30 in 2006). All nests were built on the ground with domed roofs and circular side entrances (Fig. IB), and at times incorporated willow stems (Fig. 1C). Each nest consisted of an outer layer of moss and coarse grass with an inner layer of finer materials including grasses and moose guard hairs, which formed a tightly-compressed, disc-shaped mat upon which eggs were laid. We witnessed construction of three nests in early June 2005. We observed two color-marked pairs at two nests for ~3 hrs. Males did not caiTy nesting material or enter the nest during construc- tion. However, males flew to and from nests in close proximity to the female that was gathering nest material. The male perched above the nest while the female entered and built the structure itself. Females typically gathered materials within 100 m of the nest; we did not observe travel between study plots. Chasing and physical interactions between the resident male and other males occurred frequently in the air directly above the nest during construction. The storm on 21 June 2006 blanketed the study area with 30 cm of snow which persisted for 5 days; eight nests identified at that stage of the season failed. Only one individual at each of six nests was color-marked. Two marked birds 702 THE WILSON JOURNAL OF ORNITHOLOGY • Voi 122, No. 4. December 2010 ( 1 male and 1 unknown, each with an unmarked mate) re-nested in the study area after snowmelt. One nest was depredated, but the other fledged young. We only included the initial nests from the two marked birds in our data set, as re-nests lacked statistical independence. Nest Construction. — Nests differed noticeably between habitats. Those in open shrub plots were —30% more massive (/? = 68, median [IQR] = 20.7 g [17.3-24.0]) than those in dense shrub {n = 13, median [IQR] = 15.2 g [13.4-20.0]; df = 1, X' = 6.96, P = 0.0083). Nests in dense shrub contained nearly four times the amount of moose hair (n = 13, median [IQR] = 1.1 g [0.4-1. 3]) than nests in open shrub (n = 68, median [IQR] = 0.3 g [0.0-0.8]; df = 1, x' = 6.18, P = 0.013). We measured differences in nest mass up to —3-fold every season (range: 2. 0-3. 5 fold) within each habitat. There was no coirelation between nest mass and hair mass in either habitat type (open shrub: n = 68, Spearman’s p = —0.15, P = 0.22; dense shrub: n = 13, Spearman’s p = — 0.02, P = 0.95). Orientation of nest entrances did not differ from a random distribution (open shrub plots: n — 62, mean vector (p) = 48°, vector length (r) = 0.13, Rayleigh test Z = 0.82, P = 0.82; dense shrub plots: n = 7, p = 228°, r = 0.40, Z = 2.02, P = 0.13). Entrances of nests built after the 2006 snowstorm were also randomly oriented (n = 20, p = 42°, r = 0.26, Z = 1.0, P = 0.37). Mass of nests and moose hair did not differ between pairs of plots within each habitat type (df = 1, 0.055 < X" ^ 1-85, 0.17 < P < 0.82). Dissection of 26 nests from open shrub habitat in 2006 indicated moss and grass accounted for most of the nest mass (83.4%), and that each was present in similar proportions (median [IQR] = 45.5% [31.0-54.4] and 37.9% [35.9-55.1], re- spectively). Moose hair, which was primarily incorporated into the nest cup, comprised 2.0% [0.7^. 2] of nest mass. Any remaining mass (—15%, or —3.0 g) resulted from small pieces of leaves, twigs, and feather dander. Nest mass also correlated positively with moss and grass contained in the nest structure (moss: n = 24, Spearman’s p = 0.62, P = 0.0013; grass: n = 24, Spearman’s p = 0.38, P = 0.065). Nests from open shrub habitat in 2006 were significantly lighter (by 3.2 g or 14.5%) than tho.se from the previous two seasons (median (IQR I 2006: n = 28, 18.9 g [13.9-21.9], 2005: n = 33, 22.1 g (17.4-28.21, 2004: n = 7, 22.1 g [20.8-25.7]; df = 2, x^ = 8.13, P = 0.017). The analysis pooled pre-storm (n = 8) and post-storm {n = 20) nests from 2006, as all comparisons (of nest mass, mass of moose hair, moss, grass) failed to reach statistical significance (df = 1, 0. 16 < X‘ < 0.44, 0.51 < P < 0.68). Moose hair in nests from open shmb did not differ over 3 years (df = 2, x“ = 2.23, P = 0.33). Nest mass and moose hair in dense shrub did not differ over 2 years (2004, 2005; both analyses: df = 1, x" = 0.18, P = 0.67). Fledging Success, Fledge Date, and Nest Construction. — The median number of young fledged in open shrub habitat differed each year (df = 2, x" = 24.33, P < 0.0001). Birds were most successful in 2005 [n = 29, median [IQR] = 6 young [5-6]), followed by 2004 {n = 15, median [IQR] = 4 young [4-5], and the 2006 storm (n = 20, median [IQR] = 4 young [3-5]; pairwise comparisons: df = 1, 4.39 < — 22.74, 0.0001 < P < 0.036). Open shrub habitat had significantly greater fledging success in 2005 than dense shrub habitat (/? = 7, median [IQR] = 3 young [0-6], df = 1, X' = 5.46, P = 0.019). The two habitats did not differ in 2004 (dense shrub: n = 8, median [IQR] = 4 young [3-5], df = 1, X“ ~ 0.018, P = 0.89). We detected no yearly differences in fledging success in dense shrub habitat (df = 1, X“ = 0.22, P = 0.64). The two plots within each habitat type also did not differ (df = 1, open shrub: X“ = 0.23, P = 0.62; dense shrub: X' = 0.032, P = 0.86). Median fledge date in open habitat differed significantly between years {n = 14, 2004: 20 July; n = 32, 2005: 22 July; n = 20, 2006: 27 July; df = 2, x^ = 34.49, P < 0.001). The storm disrupted breeding in 2006 and resulted in a median fledge date that was 5-7 days later than in previous years (df = 1, X“ ~ 13.48, P < 0.001). Median fledge dates in dense shrub habitat were identical to those reported for 2004 and 2005. Nest mass in open shrub habitat coirelated positively with three measures of reproductive success when all years were combined (n = 60, young fledged per nest: Spearman’s p = 0.28, P = 0.039; young hatched: Spearman’s- p = 0.33, P = 0.012; clutch size: Spearman’s p = 0.30, P = 0.024). Nest mass al.so correlated negatively with Hedge date (n = 60, Spearman’s p = —0.26, P = 0.05). The relationships were not detectable, however, when the disrupted 2006 .season was excluded (/? = 44, 0.03 < Spearman’s p < 0.09, 0.60 < P < 0.86), or when each year was analyzed separately (7 < /? < 33, —0.48 < I/agclin et al. • ARCTIC WARBLER NEST MASS 703 Spearman’s p < 0.15, 0.27 < P < 0.74). Hair mass in open shrub nests did not correlate with any measure of reproductive success (/? = 60, — 0.11 < Spearman’s p < 0.12, 0.35 < f < 0.49). Nest mass, hair mass, reproductive success or fledge date of nests from dense shrub had no detectable relationships (7 < // < 14, 0.078 < Spearman’s p < 0.20, 0.52 < P < 0.80). Nest Parasites, Incidental Arthropods, and Fledging Success. — We extracted 1,241 parasitic blowfly puparia and 187 adults from Arctic Warbler nests. The blowfly was identified as Protocalliphora tundrae. Incidental arthropods included three families of beetles (Order: Cole- optera) (Leiodidae: Catops alpinns, Carabidae: Patrobus foveocollis, Staphylinidae: Quediiis brunnipennis). Forty-six of 84 nests that successfully hatched young (55%) contained blowfly puparia, and the proportion of parasitized nests was similar in both habitats (open shrub habitat: 39 of 70 [56%], dense shrub: 7 of 14 [50%]). Most nests contained few puparia (median [IQR] = 1 [0-22]). The greatest number of puparia recorded in one nest was 157 (from open shrub habitat in 2006). Blowfly parasitism did not differ by habitat type (open shrub plots: n = 69, median [IQR] = 1 [0- 23]; dense shrub plots: n = 14, median [IQR] = 1 [0-13]; df = l,f = 0.29, P = 0.59) or by study plot (df = 3, = 3.76, P = 0.29). There was no correlation between puparia count per nest and fledging success (n = 81, Spearman’s p = — 0.056, P = 0.62) or nest mass (n = 73, Spearman’s p = 0. 12, F = 0.29). Parasite incidence per nest in open shrub habitat differed significantly between years with nests in 2004 containing few puparia (n = 15, median [IQR] = 0.0 [0. 0-0.0]) compared to those in 2005 (n = 33, median [IQR] = 4 [0-28] and 2006 (/? = 21, median [IQR] = 12 [0-28]; df = 2, X" = 15.74, P = 0.0004). The mass of moss contained in open shrub nests tended to be negatively associated with puparia count (/? = 17, Spearman’s p = —0.45, P = 0.07). Nests from dense shrub plots had no correlations between nest characteristics (mass, hair), puparia count, or young fledged (12 < /; < 13, —0.07 < Spearman’s p < 0.26, 0.40 < P < 0.83). DISCUSSION Nesting Behavior and Nest Composition. — The 95 nests we examined add new details to our understanding of Arctic Warblers that breed in central Alaska. First, consistent with other studies of Phyllo.scopus warblers, including other popu- lations of P. borealis (Barlein 2006, Clement 2006), females are the primary, if not sole, nest builder while males sing nearby and defend the site. Second, nests weigh ~20 g, and over 83% of nest mass is attributable to moss and grass. Our observations of nest location, shape, and lining of animal hair are consistent with other reports from Alaska (Murie 1956, Price and Beck 1989), and other Phylloscopus species (Bi 2004). Third, our data indicate nest openings were randomly oriented. Exposure can influence a nest’s micro- climate (With and Webb 1993), and several ground-nesting species exhibit clear patterns with regard to orientation, including birds that build domed nests or breed at high latitudes (e.g.. Burton 2007, Long et al. 2009). The position of openings of Arctic Warbler nests may be based on other parameters that we did not measure, including shelter near the nest site. It is also possible orientation is opportunistic given the uneven topography of tundra hummocks. Nest Construction. — Nests in open shrub habitat were more massive and contained less moose hair lining than those in dense shrub. Two non- exclusive explanations may account for this result. We hypothesize that local availability of nest materials differs between habitat types. For example, we often saw moose hair snagged on branches in dense shrub habitat, but moose were less common and had less opportunity (fewer shrubs of adequate height) to snag hair in open shrub habitat. The extensive breeding range and variety of nesting habitats of Phylloscopus borealis (Lowther and Sharbaugh 2008) are also consistent with an opportunistic focus upon local resources during nest building (Barlein 2006). An inverse relationship between nest mass and moose hair is suggestive of a thennal trade-off between the two habitat types. Pilot studies conducted on nests in situ, but after fledging, failed to detect any positive correlation between nest mass or the mass of moose hair lining and the insulation quality of nests (/? = 41 nests, unpubl. data). We believe this warrants additional research, prefeira- bly during incubation or brooding, as thennal properties likely result from a combination of factors, including amount and kind(s) of nest material, local topography of the breeding site, and parental behavior (Skowron and Kern 1980). Fledging Success, Fledge Date, and Nest Construction.— Htih'xm type was associated with 704 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 4, December 20IU a measurable difference in fledging success. Birds in open shrub plots during 2005 fledged twice the number of young as those in dense shrub. Open shrub plots at our study site also had a greater nest density (Sharbaugh et al. 2007), which supports the conclusion that open shrub habitat is generally more favorable to Arctic Warblers. The snow storm on 21 June 2006 interrupted breeding, as the median fledge date in 2006 occurred 5-7 days later than previous years. Nests were significantly lighter only in 2006, suggesting birds built somewhat smaller nests due to the shortened season caused by the storm. We detected no difference, however, between nests built before and after the 2006 storm, possibly due to our small sample of pre-storm nests {n = 8). It is unclear why Arctic Warblers exhibit such striking (~3-fold) differences in nest mass within each habitat type. Nest mass coixelated positively with fledging success and negatively with fledge date in open shrub habitat when all years were combined. These patterns were no longer evident, however, when the disrupted 2006 season was removed, or in smaller (yearly) samples. Nest mass in other avian species can indicate quality of the female nest builder (e.g., Soler et al. 1998, Mainwaring et al. 2008), or enhanced parenting effort (e.g., Szentirmai et al. 2005, Broggi and Senar 2009). Nest Parasites, Incidental Arthropods, and Fledging Success. — The puparia collected repre- sent the first Alaska record of the avian blowfly Protocalliphora tundrae. Our observations in- crease the number of known avian hosts of P. tundrae from two (Snow Bunting [Plectrophena.x nivali.x] and Savannah Sparrow [Passerculus sandwichensi.s]) to three, all of which are ground-nesting birds (Sabrosky et al. 1989). P. tundrae is considered a rare tundra inhabitant of northern Canada and the Yukon (Sabrosky et al. 1989). Incidental arthropods included Catops alpinus, a dung and carrion scavenging beetle known to occur in alcid nests (Perreau 1998, Peck and Cook 2002), and two other beetles that are generalist predators (Patrobus foveocollis and Quedius hrunnipennis). Q. hrunnipennis in Alas- kan tundra is unusual, as it has been reported to be a mature forest specialist from Alberta (Pohl et al. 2007). Blowfly parasitism of young varied annually but did not correlate with fledging success. This result is consistent with both observational and experimental investigations of other Protocalli- phora species and passerine hosts (Roby et al. 1992, Wittmann and Beason 1992). One recent investigation, however, suggests blowflies con- tribute to mortality and reduced movement soon after fledging (Strebyl et al. 2009). There are at least two different, non-exclusive explanations for the negative relationship between blowfly puparia and moss mass that warrant future study. First, moss may create a cool microhabitat. Cool environments can cause Pro- tocalliphora to delay or fail during larval development (Bennett and Whitworth 1991). Second, moss may cause nests to retain moisture, and Protocalliphora larvae, unlike other blow- flies, survive poorly when humidity is high (Bennett and Whitworth 1991). ACKNOWLEDGMENTS The Alaska Bird Observatory provided all field support for student interns (MLP, BSE-C). Pete Eisner, Sue Guers. Anne Ruggles, and David Shaw provided valuable support and suggestions during this project. We sincerely thank the following entomologists for identifications they provided: Terry Whitworth (Protocalliphora), Jan Ruzicka (Catops). Robert Davidson (Patrobus foveocollis), and Ales Smetana (Quedius). We thank Xin Lu for comparative details on nest building in Phylloscopus and Richard Ring for assistance during nest collection. We acknowledge Melissa and Nina Sikes for counting puparia. Nicholas Buttino and Sarah Hunter-Smith assisted with nest dissections. Undergraduate summer research stipends were provided by the Robert K. Enders Eield Biology Award (BSE-C) and the Lande Research Award (MLP) at Swarthmore College. LITERATURE CITED Barlein, F. 2006. Family Sylviidae (Old World Warblers). Pages 492-578 in Handbook of birds of the world. Volume 1 1 (J. del Hoyo, A. Elliott, and D. A. Christie, Editors). Lynx Edicions, Barcelona, Spain. Bennett, G. F. and T. L. Whitworth. 1991. Studies on the life history of some species of Protocalliphora (Diptera: Calliphoridae). Canadian Journal of Zoology 69:2048-2058. Bi, Z. L. 2004. Study of the vocalizations, breeding ecology and nest evolution of the genus Phylloscopus. Dissertation. Institute of Zoology, Chinese Academy of Sciences. Beijing, China. Broggi, J. and J. C. Senar. 2009. Brighter Great Tit parents build bigger nests. Ibis 151:588-591. Biirton. N. H. K. 2007. Intraspecific latitudinal variation in nest orientation among ground-nesting pas.serines; a study using published data. Condor 109:441^46. Clement. P. 2006. Arctic Warbler. Pages 661-662 in Handbook of birds of the world. Volume 1 1 (J. del Hoyo, A. Elliott, and D. A. Christie. Editors). Lynx Edicions. Barcelona. Spain. Long. A. M., W. E. Jen.sen. and K. A. With. 2009. Ila^eliii d al. • ARCTIC WARBLER NEST MASS 705 Orienlation ol Grasshopper Sparrow and Eastern Meadowlark nests in relation to wind direction. Condor 1 1 1 : 395-399. Lowther, P. E. and S. Sharbaugh. 2008. Arctic Warbler {Phylloscopiis borealis). The birds of North America. Number 590. MAtNWARiNG, M. C.. C. M. H. Benskin, and I. R. Hartley. 2008. The weight of female-built nests correlates with female but not male quality in Blue Tit Cyanistes caeruleiis. Acta Ornithologica 43:43—48. Murie, A. 1956. Nesting records of the Arctic Warbler in Mount McKinley National Park. Alaska. Condor 58:292-293. Peck, S. B. and J. Cook. 2002. Systematics, distributions, and bionomics of the small carrion beetles (Coleop- tera: Leiodidae: Cholevinae: Cholevini) of North America. Canadian Entomologist 134:723-787. Perreau. M. 1998. Coleopteres commensaux des oiseaux Alcidae au nord-est de la Siberie (Coleoptera, Silphidae, Agyrtidae et Leiodidae Cholevinae). Bulle- tin de la Societe Entomologique de France 103:83-89. PoHL. G. R.. D. W. Langor, and J. R. Spence. 2007. Rove beetles and ground beetles (Coleoptera: Staphylinidae, Carabidae) as indicators of harvest and regeneration practices in western Canadian foothills forests. Bio- logical Conservation 137:294—307. PRtCE, T. and D. Beck. 1989. Observations on the breeding of the Arctic Warbler in Alaska. Condor 91:219-221. Reeves. A. B., S. V. DROVETSKt, and I. V. Fadeev. 2008. Mitochondrial DNA data imply a stepping-stone colonization of Berengia by Arctic Warbler Phyllo- scopiis borealis. Journal of Avian Biology 39:567-575. Roby, D. D., K. L. Brink, and K. Wlitmann. 1992. Effects of bird blowfly parasitism on Eastern Bluebird and Tree Swallow nestlings. Wilson Bulletin 104:630- 643. Sabrosky, C. W., G. F. Bennett, and T. L. Whitworth. 1989. Bird blow flies (Proiocalliphora) in North America (Diptera: Calliphoridac) with notes on the Palearctic species. Smithsonian Institution Press, Washington, D.C., USA. Sharbaugh, S., R. Ring, S. Guers, and D. Shaw. 2007. Breeding ecology and habitat associations of the Arctic Warbler (Phyllo.scopus borealis) in central Alaska. Final Report 2004-2006. Alaska Bird Obser- vatory, Fairbanks, USA. http://www.ala.skabird.org/ ?page_id = 249 Skowron, C. and M. Kern. 1980. The insulation of nests of selected North American songbirds. Auk 97:816- 824. SoLER, J. J., J. J. Cuervo, and A. P. Moller. 1998. Nest building is a sexually selected behaviour in the Barn Swallow. Animal Behaviour 56:1435-1442. Strebyl. H. M., S. M. Peterson, and P. M. Kapfer. 2009. Fledging success is a poor indicator of the effects of bird blow flies on Ovenbird survival. Condor 1 1 1:193- 197. SZENTIRMAl, 1., J. KOMDEUR, AND T. SZEKLEY. 2005. What makes a nest-building male successful? Male behavior and female care in Penduline Tits. Behavioral Ecology 16:994-1000. Tucker, B. W. 1949. Kennicott’s Willow Warbler Acanthopneuste borealis kennicotti (Baird) Pages 330-339 in Life histories of North American thrushes, kinglets, and their allies (A. C, Bent, Editor). U.S. National Museum Bulletin 196. Wilkie, D. 1983. Rayleigh test for randomness of circular data. Applied Statistics 32:31 1-312. With, K. A. and D. R. Webb. 1993. Microclimate of ground nests: the relative importance of radiative cover and windbreaks for three grassland species. Condor 95:401—413. WiTTMANN, K, AND R. C. Beason. 1992. The effect of blowfly parasitism on nestling Eastern Bluebird development. Journal of Field Ornithology 63:286- 293. The li 'ilsoii Journal of Oniilhology 122{4):706-715. 2010 HOME RANGE, HABITAT USE, AND NEST SITE CHARACTERISTICS OF MISSISSIPPI KITES IN THE WHITE RIVER NATIONAL WILDLIFE REFUGE, ARKANSAS TROY J. BADER’ AND JAMES C. BEDNARZ' ABSTRACT. — We located 39 Mississippi Kite {Ictinia mississippiensis) nests during the 2004 and 2005 breeding seasons in the White River National Wildlife Refuge in the Mississippi Alluvial Valley of Arkansas, USA. Radio transmitters were placed on seven adult and eight juvenile kites; 649 locations based on ground telemetry and 393 locations from aircraft were recorded for six adults {n = 5 males, 1 female) and five juveniles (/; = 2 males, 3 females). The mean 90% kernel home range was 3,098 ha for adult kites {n = 6) and 439 ha (n = 5) for juveniles. Radio-marked Mississippi Kites significantly used mature forest (65.6%), second growth (15.5%), and water (10.0%) relative to availability (59.6, I 1.5, and 7.1%, respectively). Agriculture fields (6.5%) and wetlands (1.5%) were used significantly less within the home ranges of kites relative to availability ( 1 7.6 and 3.9%, respectively). Tree height, diameter at breast height (dbh), and height of nest tree emergence above the surrounding trees were significantly greater for nest sites than randomly-selected overstory trees. Nest trees were significantly closer to the edge than randomly-selected trees. Most (57%) Mississippi Kite nests were in crotches of secondary branches in the nest tree. Conservation of super-emergent trees and mature forests is needed for nesting and foraging areas. Second-growth forest should be allowed to mature (>75 years) and uniform canopies should be avoided. Received 7 August 2009. Accepted 28 April 2010. The Mississippi Kite (Ictinia mississippien.sis) breeds in bottomland and riparian areas through- out the Mississippi River Valley and southeastern North America (e.g., Kalla 1979, Evans 1981, Whitmar 1987, Barber 1995, St. PieiTe 2006). This raptor also inhabits riparian areas, shelter- belts, oak (Querciis spp.)-shrub prairies, and savannas in the western United States (Parker 1988). This kite has adapted to nesting in urban golf courses, parks, and neighborhoods primarily in the western portion of its range (Parker 1988). The ability of the western population to adapt to encroachment of humans has allowed it to increase in number and expand into new areas (Parker and Ogden 1979). Kite populations in the eastern United States were affected by severe habitat alteration in the late 18()()s and early 190()s. Gosselink and Lee (1989) estimated that, at time of European settlement, there were ~8() million ha of forested freshwater wetlands in the United States with the majority in the Mississippi Alluvial Valley. The area of bottomland hardwood forests in the Mississippi Alluvial Valley decreased to 4.8 million ha by 1937, and to only 2.1 million ha by 1970 (Gosselink et al. 1990). ' Department of Biological Sciences, Arkansas State University. P. O. Box 599, Jonesboro. ,AR 72467, USA. ’Current address: USDA-ARS, Stuttgart National Aqua- culture Research Center, P. O. Box 1050, Stuttgart, AR 72160. USA. 'Corresponding author; e-mail: troybader@hotmail.eom Several researchers have described habitat use and nest site characteristics of Mississippi Kites, often based on small sample sizes, in Tennessee (Kalla 1979), Illinois (Evans 1981), Missouri (Whitmar 1987, Barber 1995), and Arkansas (St. Pieire 2006) within the Mississippi Alluvial Valley; almost all of this work remains unpublished. Allan and Sime (1943) observed daily activities of 169 Mississippi Kites in the panhandle of Texas and subjectively estimated home-range size to be ~250 ha. Barber (1995) using telemetry estimated mean home-range size of kites along the Mis- sissippi River in Missouri as 1,160 ha (/? = 10), much larger than estimates from the Texas panhandle. Barber (1995) compared the number of telemetry locations within specific vegetation types to estimates of the proportions of these types available in the surrounding landscape. Forested habitats fragmented with agricultural fields and developments are not ecologically equivalent to the relatively contiguous-forested habitats once used by kites in the 180()s and early 19()()s. The vast changes in the landscape probably affected home-range size and habitat use, and may have influenced reproductive success of this species (Bader and Bednarz 2009). A better understanding of how Mississippi Kites use landscapes that are present today will be useful to conserve and enhance habitats for (his species. Our objectives were to examine: (1) use of habitats and landscapes, and (2) selection of nest sites by Mississippi Kites in (he Mississippi Alluvial Valley. 706 Bader and Bednarz • llOMli RANGE: AND llABi rA'I’ OE' MISSISSIPPI KI TES 707 Wildlife Refuge and vicinity in Arkansas, 2004-2005. METHODS Study Area. — White River National Wildlife Refuge (WRNWR; Fig. I) is in Arkansas, Desha. Monroe, and Phillips counties in southeastern Arkansas. It is —64,750 ha in size and is divided into two units by Arkansas Highway I. The South Unit, where the majority of our work was conducted, is the larger of the units (41,440 ha). This unit extends from St. Charles, Arkansas south to the confluence of the White River and the Arkansas River Canal (Fig. I). The WRNWR consists mostly of bottomland hardwoods open to the public for recreational use, hunting, and fishing and is managed for both game and non- game wildlife species (Bader 2007). Nest Searching. — We began searching for Mississippi Kite nests on 9 and 23 April 2004 and 2005, respectively. Once a Mississippi Kite was located, we watched it for possible breeding activities such as carrying sticks or copulation, which indicated kites were nesting nearby. We recorded the location of each nest with a Global Positioning System (GPS) receiver and left the site immediately to minimize disturbance. We periodically returned to areas where Mississippi Kites were seen and continued searching until a nest was lound or kites abandoned the area, indicating they were no longer attempting to breed. Capturing and Radio-marking Kites. — We used a mist-net system with a live Great Horned Owl {Bubo virginianus) or Red-shouldered Hawk (Buteo lineal us) near nests with nestlings (7- 30 days of age) to capture adult Mississippi Kites 708 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122. No. 4. December 2010 ( Hamerstrom 1963, Barber et al. 1998). The mist- net system consisted of two 2.6 X 6 m mist nets (72 mm mesh) connected to a pulley system on telescoping metal poles. This allowed the top of the net to be elevated ~7 m above ground with the bottom of the lower net —1.5 m off the ground. The total dimensions of the system using two mist nets were 5.2 m in height X 6 m wide (Bader 2007). We placed a U.S. Geological Survey (USGS) aluminum band and two or three plastic color bands on legs of each kite for unique identifica- tion. Selected kites were also fitted with a radio transmitter using a modified figure-eight leg harness (Radley 2002). Mississippi Kites were fitted with a 6.0-g transmitter (Holohil Systems Ltd., Carp, ON, Canada). We recorded several linear measurements, mass, and obtained a blood sample from each kite, before it was released. Gender of each kite was assigned based on plumage characteristics (Parker 1999), and veri- fied by DNA analysis in the laboratory (Bader 2007). We left the nest site immediately once a kite was released to minimize any potential stress on adult and nestling kites. We climbed to each nest and banded the nestling(s) when they were 3-4 weeks of age. The mass of each nestling was measured, and a USGS aluminum band and two or three plastic color bands were placed on their legs for identification of individuals. Each nestling was also fitted with a radio transmitter using the procedure described for adults and returned to the nest. Gender Identification. — A 0.05-cc sample of blood was drawn from the brachial vein of each kite using a hypodermic needle and a small capillary tube (Faaborg et al. 1995) and placed in Longmire Solution (Longmire et al. 1988). The protocol outlined by Donohue and Dufty (2006) was followed to extract DNA from the blood and i.solate the CHD-Z and CHD-W bands (Griffiths et al. 1998) to identify the gender of captured kites. Radiotelenietry of Adult and Juvenile Kites. — Triangulation was used to obtain locations from the ground and a fixed-wing aircraft was used to identify point locations from the air. Receiver sites for triangulations were established —500 m apart along roads or all-terrain vehicle trails near nest sites and compass azimuths were recorded in the direction of the strongest signal from the radio-marked kite. Compass azimuths from at least three receiver sites were taken within a 5- min period to estimate the error ellipse around each location. Eight triangulations with a >10- min interval between each triangulation were completed 2-3 days each week. The 10-min period allowed time for the kite to move between triangulation attempts, minimizing autocorrelation between locations (Swihart and Slade 1985). The triangulation azimuth was entered into the OTA Triangulation Program (Ripper et al. 2007), which provided coordinates for the estimated location. We could calculate an eiTor ellipse using this program that theoretically has a 95% probability of including the true location of the kite. We eliminated all estimated locations that had an error ellipse greater than 10 ha to improve accuracy. This relatively large error ellipse criteria was chosen to accommodate the high mobility of kites (typical home range size >3,000 ha). We obtained locations from the air (DeVault et al. 2003) for each radio-marked kite 2-3 times per flight depending on the amount of flight time available. Locations of individual kites were recorded >20 min apart. Flights were conducted 2-3 days each week using a Cessna 172 or 182 equipped with a pair of “side-looking” 4-element yagi antennae (Bader 2007). Habitat Use. — Point locations were entered into Geographic Information System (GIS) ArcView software and home-range polygons were devel- oped for each kite. Two types of home-range polygons were generated using point locations; minimum convex polygon (MCP) and 90% kernel (White and Garrott 1990). We used the 90% probability of use contours as estimates of home- range size. The home-range polygon of each kite was overlaid onto USGS Digital Orthophotograph Quarter Quadrangles (DOQQs) taken in fall 2004. ArcMap 8.3 at a scale of 1 :3,000 was used to classify habitats into seven cover types within each home range. The mature forest cover type consisted of tracts of large trees that had relatively even canopy with few openings and had not been recently logged (>30 year). Second-growth cover type consisted of young forest with large openings in the canopy, abandoned fields, and recently- logged forest. The water cover type included all open waterways (e.g., lakes, bayous, and rivers). The wetland cover type included all forested and non-forested swamps and low-lying vernal areas. The agriculture cover type included all types of agriculture fields (e.g., crop and hay fields). The urban cover type consisted of noticeably devel- oped areas (e.g., rural towns). The road cover type Bcicler ami Bcdmirz • HOME RANGE AND IIARITAT OE' MISSISSIRRI KITES 709 included highways, county roads, and obvious private roads. Polygons were drawn around each cover type and the sum of all polygons within a cover type was used to calculate the number of hectares available of that cover type within a given home range. Bailey’s confidence intervals (Bailey 1980, Cherry 1996) were used to examine whether kites used cover types within their home range in proportion with that available. We used the Chi- square distribution and a Bonfen'oni correction based on the number of available cover types to construct a 95% confidence interval (Cl) around the proportion of telemetry locations within each cover type. The confidence interval was compared to the proportion of habitat that was available within the kite’s home range to delineate the level of kite use. Nest-site Characteristics. — Vegetation data were collected at every nest site after young fledged or the nest failed and at an equal number of sites selected using a random-number table. The distance (0-250 m) was paced while using a compass to follow the random azimuth to locate the reference sample site and the closest overstory tree was chosen as the plot center. The same vegetation data were collected within a 0.04-ha circular plot around nest and random trees. We followed the BBIRD Protocol (Martin et al. 1997) and included species of the tree, tree height, tree diameter at breast height (dbh), distance to forest edge, dominant and co-dominant trees within the surrounding area, and species and height of each tree in the overstory and within the circular plot. Data collected on nest-site characteristics includ- ed nest height, orientation of nest in tree, and position of nest in tree (primary or secondary fork). We collected data on perch trees, defined as the tree(s) in which kites perched in the morning and frequently during the day in the immediate vicinity (<200 m) of their nests. Measurements collected at perch trees were the same as tho.se taken for nest trees, but also included distance from nest and azimuth to nest. We recorded whether the tree was dead or alive and classified it as a primary or secondary perch tree. The primary perch tree was the tree that one or both kites used the majority (>50%) of the time. The secondary perch tree, if present, was the tree u.sed by one or both kites less frequently (<50%). Statistical Analysis. — We used a paired r-test (SAS Institute 1999) to compare Mississippi Kite nest sites with sites selected at random (Sokal and Rohlf 1995), and a /-test to compare characteris- tics of primary and secondary perch trees. RESULTS We located 21 Mississippi Kite nests in 2004 and 18 in 2005. We captured seven adult kites, each from separate nests (4 in 2004, 3 in 2005), in 22 trapping attempts. We made 1 1 attempts to capture chicks in 2004 and 2005 and captured nine (at ages of ~3.5^ weeks) from nine nests. Analysis of DNA to identify gender of captured adult and nestling kites was conducted on 14 blood samples (6 adults and 8 nestlings). Five male and one female adult kites, and four male and four female nestling kites were identified. Five of six adults (83%) were identified correctly to gender based on plumage coloration. Radio-marked Mississippi Kites were located using ground receiver sites from 9 July to 28 August 2004 and from 14 July to 20 August 2005. Locations of each radio-marked kite were record- ed from a fixed-wing aircraft from 28 July to 19 August 2004 and from 20 July to 26 August 2005. We recorded 649 locations from the ground (2004 = 449 locations, 2005 = 200 locations) and 393 locations from the air (2004 = 134 locations, 2005 = 259 locations). MCP and 90% kernel home ranges were generated for each radio-marked kite (Table 1, Fig. 2). The mean (± SE) MCP home range size for adult male radio-marked kites was 3,289.1 ± 1,419.3 ha; n = 5). Home-range data were collected for only one female and the MCP home range was 1,899.8 ha. The mean (± SE) MCP home range for male juvenile kites was 536.2 ± 10.8 ha; n = 2) and 508.4 ± 316.4 ha; n = 3) for female juvenile kites. The mean (± SE) home range using the 90% kernel method for adult male radio-marked kites was 3,567.3 ± 1,393.5 ha; n = 5). The home range size for the one radio-marked adult female kite using the 90% kernel method was 752.2 ha. The mean 90% kernel home range for male juvenile kites was 299.0 ± 23.0 ha (/; = 2) and 533.0 ± 266.7 ha (/? = 3) for female juvenile kites. We estimated habitat use within home ranges of 12 Missi.ssippi Kites (6 adults. 6 juveniles). Mississippi Kites significantly used mature forest (95% Cl = 0.611-0.698; Table 2) more than its availability (0.596). Water was also significantly used (95% Cl = 0.074-0.130; Table 2) more relative to its availability (0.071). Mississippi 710 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4, December 201 U TABLE 1. Minimum convex polygon (MCP) and 90% kernel home range data for Mississippi Kites radiotracked in the White River National Wildlife Refuge, Arkansas, 2004-2005. Band # Gender Age- # of triangulations # of aerial locations Total locations MCP (ha) 90% Kernel (ha) Year 745-57320 M AHY 32 20 52 1,398.64 1,594.23 2004 745-57323 L HY 53 17 70 78.37 55.75 2004 745-57318 L AHY 103 21 124 1,899.81 752.19 2004 745-57322 M HY 47 19 66 547.00 321.96 2004 745-57319 M AHY 1 12 18 130 626.10 431.14 2004 745-5732 1 F HY 48 19 67 1,125.44 975.75 2004 745-57324 M AHY 54 20 74 6,225.47 7,770.00 2004 745-57326 M AHY 95 60 155 7,235.58 5,896.23 2005 745-57327 M AHY 22 13 35 959.68 2,145.07 2005 745-57329 F AHY 5 18 23'-' 497.44 1,192.65 2005 745-57328 M HY 0 1 1 1 U 0 0 2005 745-5733 1 M HY 35 72 107 525.42 275.94 2005 745-57332 _b HY 10 26 36'-- 548.27 545.45 2005 745-57330 F HY 33 61 94 321.37 567.61 2005 745-57333 F HY 0 0 o-^ 0 0 2005 AHY = alter hatch year or breeding adult; HY = Hatch year or nestling. No blood was collected. ■■ Few locations because kite was depredated (not included in calculation of MCP or kernel means). Locations while chick was still in the nest (not included in calculation of MCP or kernel means). No data collected before nestling was depredated (not included in calculation of MCP or kernel means). Kites also significantly used .second growth (95% Cl = 0.123-0.190; Table 2) more relative to its availability (0.115). Agriculture fields and wet- lands were significantly used less relative to availability (Table 2). Mississippi Kites used both roads and urban areas in proportion to their availability (Table 2). The mean nest tree height (33.26 m) was significantly taller (/ = 5.37, df = 33, P < 0.0001; Table 3) than paired randomly-selected trees (26.77 m). Mean dbh of nest trees (75.58 cm) was significantly greater (t = 8.44, df = 33, P < O.OOOl; Table 3) than paired randomly-.selected trees (48.34 cm). Mean height of nest-tree emergence (5.29 m) above the surrounding canopy was significantly greater {[ = 3.44, df = 33, P = 0.0016; Table 3) than the mean height of randomly-selected tree emergence (0.70 m). The mean (± SE) height of Mississippi Kite nests was 26.89 ± 0.82 m, while the mean surrounding canopy height was 27.68 ± 0.74 m, indicating kites tended to build nests almost level with the height of the surrounding canopy. The mean (± SE) distance (61.06 ± 15.65 m) to the forest edge from Mississippi Kite nests was significantly less (t = 4.16, df = 33, P = 0.()()()2; Table 3) than the distance (126.19 ± 16.18 m) to the forest edge from randomly-.selected sites. Other measured variables did not differ between nest and random sites (Table 3). Twenty of 35 nest sites (57%) were in crotches of .secondary branches. The remaining 15 nests were on primary branches of nest trees. Nests were randomly placed in any direction throughout the nest trees. Mississippi Kite nests were found in six species of trees with Overcup oak {Quercus lyrata\ n = II) used most commonly. Sweetgum {Liquidamber styracifliuv, n = 7), willow oak {Q. phellos) ill = 7), Nuttall oak (Q. nuttallir, n = 3), American sycamore (Platcmus occidentals-, n = 2), and white oak {Q. allxv, /? = I ) were used less frequently as nesting sites. Mississippi Kites used numerous species of trees for perching. Black willow {Salix nigra-, n = 9) and water hickory (Carya aquatica-, n = 8) were the two most common species used for primary perches. Overcup oak (n = 5) and Nuttall oak (// = 4) were also commonly used for primary perches. Sweetgum (/; = 3), slippery elm (Ulinus rubra-, n = 3), willow oak (// = 2), and water oak {Q. nigra-, n = 2) were also used as primary perch trees. Nineteen (52.8%) of 36 primary perch trees (Table 4) were live with a dead limb used as the perch, six were live trees with sparsely-leaved limbs, six were live trees with live limbs used as perches, and five were dead trees. Secondary perch trees included water hickory (/; = 3), overcup oak (// = 2), water oak (/; = 1), and sLigarberry (Celiis laevigata-, n = 1). One dead tree, three live trees with sparsely-leaved Iknkr and Hcdnarz • IlOMli RANGE AND HABITAT Ol' MISSISSlIMM Krn;S 71 i f, •7’ Mississippi Kite nest 0 250 500 llMetei s 1,000 Kite locations I I MCP 90% Kernel home range FIG. 2, Minimum Convex F^olygon (MCP) and 90% kernel home range of a radio-tracked Mississippi Kile (Band # 745-57331) in the White River National Wildlife Refuge, Arkansas, 2005. 712 THE WILSON JOURNAL OF ORNITHOLOGY • Voi 122. No. 4. December 2010 TABLE 2. The proportion of cover types within home ranges available and used by radio-tracked adult (/? = 6) and juvenile (n = 6) Mississippi Kites in the White River National Wildlife Refuge, Arkansas, 2004-2005. Cover type ha Locations in cover type Proportion of cover type available Proportion of observed use 95% Confidence Interval of use Agriculture 3,683.83 57 0.176 0.065* 0.044-0.090 Mature forest 12.497.36 577 0.596 0.656* 0.61 1-0.698 Second growth 2.417.41 136 0.1 15 0.155* 0.123-0.190 Road 64.83 2 0.003 0.002 0.000-0.01 1 Wetland 821.65 13 0.039 0.015* 0.006-0.029 Urban 38.59 6 0.002 0.007 0.001-0.018 Water 1,488.13 88 0.071 0.100* 0.074-0.130 * Significantly different from amount available. limbs, two live trees with a dead limb, and one live tree with live limbs were all used as .secondary perch trees. There were no significant differences in dbh, tree height, distance to edge, or distance to nest between primary and secondary perch trees (Table 4). Primary and secondary perch trees were significantly closer {t = 5.6, df = 31, P = <0.001 ) to an edge and had a significantly larger (t = 7.5, df = 34, P < 0.0001) dbh than trees selected at random (Tables 3, 4). There was no significant difference in height of primary and secondary trees and trees selected at random. DISCUSSION The mean 90% kernel home range size of male Mississippi Kites (3,567 ha, n = 5) was larger than the one female monitored (752 ha). We suggest this difference may be real because males are more likely to venture long distances from the nest during their daily foraging activities (Dunstan et al. 1978, Barber 1995). Males range widely, probably to gain access to habitat types or food resources that are not available near the nest site. Access to a larger foraging area and possibly to areas with more abundant prey makes the male’s role as a food provider to the nestlings vital to success of the nest. Female kites may have smaller home ranges because of their role in the primary defense of the nest; they afso tend to the nestlings more than males, similar to other female raptors (Newton 1978, Collopy 1984, Bader 2007). Barber (1995) estimated the mean male MCP Mississippi Kite home range size in Missouri as 1,357 ha (/7 = 6), notably smaller than the mean male MCP home ranges that we estimated (3,289 ha). The mean female kite home range in Missouri was 865 ha (/; = 4; Barber 1995), also considerably smaller than the MCP home range of the one female that we radiotracked ( 1 ,900 ha). We suggest these difference may be due to either differences in land.scapes of the respective study areas (the Missouri study was conducted in a more fragmented, forest-agriculture landscape) and the techniques used (no aerial telemetry was used in the Missouri study), or both factors. TABLE 3. Habitat characteri.stics al Mi.ssi.ssippi Kite ne.st .sites {n = 39) and sites selected at random (/; = 39) in the White River National Wildlife Refuge. Arkansas, 2()04-2()()5. Variable Nest .site Random site p’ Mean ± .SE Range Mean ± SE Range Tree height (m) 33.26 ± 0.97 24.09^3.07 26.77 ± 0.92 14.30^1.42 ■ <0.0001* DBH (cm) 75.58 ± 3.32 52.70-1 13.80 48.34 ± 2.25 20.30-79.00 <0.0001* No. of trees in 0.04-ha plot 3.56 ± 0.30 1 .00-7.00 3.53 ± 0.31 1 .00-8.00 0.9378 Mean tree heighP (m) 27.68 ± 0.74 20.77-37.57 26.85 ± 0.72 20.25-37.32 0.3535 Tree emergence"' (m) 5.29 ± 0.93 -8.53-17.85 0.70 ± 0.76 1 1.20-9.39 0.0016* Distance to forest edge (m) 61.06 ± 6.13 0-142.50 126.19 ± 16.18 0-3 1 1 .00 0.0002* * Significanlly tlifferenl from silcs selected at random. !’ £ 0.05. " /’ ^ probability that means are dilTerenl based on a paired t-tesl. Mean iree height - mean height of trees in a ().l)4-ha plot surrounding the nest or random tree. ^ Tree emergence - the difference between mean height of nest trees above surrounding trees in 0.04 ha plot or mean height of random trees above surrounding trees in 0.04-h:i plot. liuilcr and Bednarz • HOME RANGE AND HABITAT OE MISSISSIPPI KITES 713 TABLE 4. Characteristics ot primary (n — 36) and secondary (n = 7) perch trees ol' Mississippi Kites in the White River National Wildlife Refuge. Arkansas, 2004-2005. Primary Secondary Variable Mean ± SE Range Mean ± SE Range Tree height (m) 30.41 ± 1.68 14.3-66.5 27.23 ± 3.31 18.6-41.46 DBH (cm) 74,45 ± 2.77 36.1-1 1 1.8 84.14 ± 7.43 55.5-103.8 Distance to nest (m) 82.68 ± 8.66 1 1.3-231.0 78.53 ± 10.02 45.0-120.0 Distance to edge (m) 18.74 ±3.17 0-77.3 29.89 ± 16.49 0-124.5 Our data documented that Mississippi Kites significantly used open waterways more than available to kites within their home ranges (Table 2). Open water likely provided habitat for prey items such as dragonflies (Odonata) and frogs (Anura). Barber (1995) did not report this cover type to be more abundant within kite home ranges compared to reference areas, but it was a large part of most home ranges that he sampled. Agricultural fields were used less within the kite home ranges in Arkansas than available (Table 2), but were significantly more abundant in kite home ranges in Missouri (Barber 1995). Soybeans, milo, com, cotton, and fallow fields were included in the agriculture cover type in the Missouri study. Rice, soybeans, and cotton were the primary crops farmed in the areas surrounding the WRNWR. Wetland areas were significantly underused (Table 2), but we suggest wetlands indirectly may be an important area for foraging kites. Wetlands likely produce prey, but this prey is not restricted to these areas and probably moves into other cover types where kites may forage on them. Wet areas are used as foraging habitats by other raptors such as the Red-shouldered Hawk (Bed- narz and Dinsmore 1981, Bloom et al. 1993, Howell and Chapman 1997). Mature forests were significantly u.sed more within the home ranges of Mississippi Kites in Missouri (Barber 1995) and in Arkansas (Table 2). In Tennessee, Kalla (1979) reported wooded areas in the floodplain provided preferred nesting sites for kites and key habitat for prey species. Mature forests provide habitat for dragonflies and cicadas (Homoptera), the two most common prey items delivered to nests in the WRNWR (Bader 2007). Red-shouldered Hawks were demonstrated to use forested habitats in California (Bloom et al. 1993) and bottomland hardwood forests in Georgia (Howell and Chapman 1997) more than available. Lai'ge contiguous tracts of mature forests seem to be an important cover type for kites providing both nest sites and prey. Second growth was used significantly more within home ranges of kites in Arkansas (Table 2), but not in Missouri (Barber 1995). This cover type is common in numerous landscapes, and can provide foraging habitat and marginal nesting habitat for kites. Timber stands classified as second growth can provide nest sites if super-emergent trees are not removed during timber harvest operations or are created during timber haiwest. Fallow fields, which consist of early successional growth, can provide habitat for aerial insects and other prey items. The second growth cover type was not common in kite home range in Missouri and only accounted for 3.8% of home ranges and 1 .2% of the sampled suiTOunding areas. Kites used urban areas and roads in proportion to availability within their home range. Both urban areas and roads were near other cover types u.sed by kites. Kites may have used thermals produced in these areas and probably did not actively select these sites for foraging. Mississippi Kites selected trees taller than the mean overstory canopy height for their nest tree (Table 3). U.se of super-emergent trees for nesting allows this long-winged raptor efficient access to the ne.st (Barber et al. 1998). Kites in WRNWR tended to build nests almost level with the top of the suiTOLinding canopy. Barber (1995) and Whit- mar ( 1 987) reported nests were higher than the tops of the suiTOLinding trees. Placing nests at or slightly above the tops of the surrounding trees may provide a view of the surrounding area and potential predators (Whitmar 1987, Barber 1995). Placing nests slightly below the surrounding canopy height would give the nest more protection from the wind and make the nest less conspicuous to predators. Mississippi Kite nest trees have a larger dbh than randomly-selected trees (Table 3). Barber (1995) also reported the dbh of nest trees to be significantly larger than random trees. The large 714 THE WILSON JOURNAL OF ORNITHOLOGY • Voi 122. No. 4, December 2010 dbh of nests trees may be correlated with age and size ot the super-emergent trees typically used for nesting sites. Mississippi Kites selected nest sites closer to the forest edge than random sites and all forest edges comprised of lakes, bayous, or the White River. This pattern was also documented during research previously conducted on the refuge (St. Pierre 2006). These bodies of water provide open areas for kites to forage for aerial insects. Barber (1995) reported kite nests were located closer to field edges more frequently than water edges. Field edges were uncommon within the interior of WRNWR, where waterways comprised the ma- jority of the available edges. Whitmar (1987) reported kite nests in the Mississippi River Valley in Missouri had a mean distance to an edge of 49.2 m and a mean distance to water of 96.3 m. Edge habitats, particularly natural edges, seem to be an important attribute of Mississippi Kite nesting habitat. Kite nests were more commonly built on secondary than on primary branches. Barber ( 1995) akso documented this pattern for kite nests in Missouri and suggested this may be an anti- predator defense. Our data and those reported by Whitmar (1987) indicated kite nests were seem- ingly randomly placed throughout the nest tree. CONSERVATION IMPLICATIONS Mature tracts of forests should be conserved because this habitat type provides nesting sites and a foraging habitat for Mississippi Kites. Large, super-emergent trees should be con.served for kite nesting habitat. Management that pro- motes development of additional large, super- emergent trees will likely be beneficial to kites. Timber harvests consistent with kite conservation should be conducted .so uneven-height canopy forests are created. We recommend that second- growth forest should be allowed to develop longer between harvests (>75 years) to provide foraging and nesting habitat for kites. Managers should avoid even-aged regeneration of dear-cuts and fallow fields. Natural waterways, including wet- lands, should be con.served for production of prey and as foraging areas. ACKNOWLEDGMENTS This project was t'uncicci primarily by the U.S. Fish anti Wildlife Service (USFWS) and Arkansas Game and Fish Commission (AGFC) through a State Wildlife Grant. Substantial additional funds were provided by Arkansas State University. We thank K. L. Rowe (AGFC) for support and assistance on this project. We thank the USFWS, especially R. E. Hines and staff at the White River National Wildlife Refuge, for help on the logistics for this project. We are especially grateful to our research technician. W. D. Edwards for hard work and dedication to this project. We greatly appreciate the assistance in the field and in the laboratory provided by J. D. Brown. T. J. Benson. D. J. Baxter. T. R. Edwards. E. J. Bader, and A. M. St. Pierre. LITERATURE CITED Allan. P. F. and P. R. Sime. 1943. Distribution and abundance of the Mississippi Kite in the Texas panhandle. Condor 45:1 10-1 12. Bader, T. J. 2007. Reproductive success, causes of nesting failures, and habitat use of Swallow-tailed and Mississippi kites in the White River National Wildlife Refuge, Arkansas. Thesis. Arkansas State University, Jonesboro, USA. Bader, T. J. and J. C. Bednarz. 2009. Reproductive success and causes of nesting failures for Mississippi Kites: a sink population in eastern Arkansas? Wetlands 29:598-606. Bailey. B. J. R. 1980. Large sample simultaneous confidence intervals for the multinomial probabilities based on transformation of the cell frequencies. Technometrics 22:583-589. Barber, J. D. 1995. Mississippi Kite habitat use and reproductive biology in the Mississippi flood plains of southeast Missouri. Thesis. University of Missouri, Columbia, USA. Barber. J. D., E. P. Wiggers, and R. B. Renken. 1998. Nest-site characterization and reproductive success of Mississippi Kites in the Mississippi River floodplains. Journal of Wildlife Management 62:1373-1378. Bednarz, J. C. and J. J. Dinsmore. 1981. Status, habitat use, and management of Red-shouldered Hawks in Iowa. Journal of Wildlife Management 45:236-241. Bi.oom, P. H., M. D. McCrary, and M. J. Gibson. 1993. Red-shouldered Hawk home-range and habitat use in southern California. Journal of Wildlife Management 57:258-265. Cherry. S. 1996. A comparison of confidence interval methods for habitat u.se-availability .studies. Journal of Wildlife Management 60:653-658. COLLOFY, M. W. 1984. Parental care and feeding ecology of Golden Eagle nestlings. Auk 101:753-760. DeVault, T. L.. W. L. Stephens, B. D. Reinhart, O. E. Rhodes Jr., and I. L. Brisbin Jr. 2003. Aerial telemetry accuracy in a forested landscape. Journal of Raptor Research 37:147-151. Donohue. K. C. and A. M. Dui-ty Jr. 2006. Sex determination of Red-tailed Hawks (Biiteo jamaicensis c(iliinis) using DNA analysis and morphometries. Journal of Field Ornithology 77:74-79 Dunstan, T. C., j. H. Harper, and K. B. Phipps. 1978. Habitat use and hunting strategics of Prairie Falcons, Rcd-tailcd Hawks, and Golden Eagles. Final Report submitted to USDI, Bureau of Land Management, Denver, Colorado, USA. Bader and Bednarz • HOME RANGE AND IIARITA T Ol' MISSISSIPPI KITES 715 Evans, S. A. 1981. Ecology and behavior ol the Missi.ssippi Kite (Icrinia niississippiensis) in .southern Illinois. Thesis. Southern Illinois University, Carbondale, USA. Faaborg, J,, P. G. Parker, L. Delay, T. De Vries, J. C. Bednarz, S. M. Paz, J. Naranjo, and T. A. Waite. 1995. Conlirmation of cooperative polyandry in the Galapagpos Hawk (Biiieo galapagoensis). Behavioral Ecology and Sociobiology 36:83-90. Gosselink, j. G. and L. C. Lee. 1989. Cumulative impact assessment in bottomland hardwood forests. Wetlands 9:89-174. Gosselink, J. G., G. P. Shaffer, L. C. Lee, D. M. Burdick, D. L. Childers, N. C. Leibowitz, S. C. Hamilton, R. Boumans, D. Cushman, S. Fields, M. Koch, and J. M. Visser. 1990. Landscape conserva- tion in a forested wetland watershed. Can we manage cumulative impacts? Bio.science 40:588-600. Griffiths, R., M. C. Double, K. Orr, and R. J. G. Dawson. 1998. A DNA test to sex most birds. Molecular Ecology 7:1071-1075. Hamerstrom, F. 1963. The use of Great Horned Owls in catching Marsh Hawks. Proceedings of the Interna- tional Ornithological Congress 13:866-869. Howell, D. L. and B. R. Chapman. 1997. Home range and habitat use of Red-shouldered Hawks in Georgia. Wilson Bulletin 109:131-144. Kalla, P. I. 1979. The distribution, habitat preference, and status of the Mississippi Kite (Ictinia mississippiensis) in Tennessee, with reference to other populations. Thesis. East Tennessee State University, Johnson City, USA. Longmire, j. L., a. K. Lewis, N. C. Brown, J. M. Buckingham, L. M. Clark, M. D. Jones, L. J. Meincke, j. Meyne, R. L. Ratliff, F. A. Ray, R. P. Wagner, and R. K. Moyzis. 1988. Isolation and molecular characterization of a highly polymorphic centromeric tandem repeat in the family Falconidae. Genomics 2:14-24. Martin, T. E., C. R. Paine, C. J. Conway, W. M. Hochachka, P. Allen, and W. Jenkins. 1997. BBIRD field protocol. USGS, Montana Cooperative Wildlife Research Unit, University of Montana, Mis.soula, USA. Newton, I. 1978. Feeding and development of Sparrow- hawk Accipiler nisus nestlings. Journal of Zoology 184:465-487. Parker, J. W. 1988. Mississippi Kite Ictinia mississip- piensis. Pages 166-186 in Handbook of North American birds (R. S. Palmer, Editor). Volume 4. Yale University Press, New Haven, Connecticut, USA. Parker, J. W. 1999. Mississippi Kite (Ictinia mississip- piensis). The birds of North America. Number 402. Parker, J. W. and J. C. Ogden. 1979. The recent history and status of the Mississippi Kite. American Birds 33:1 19-129. Radley, P. H. 2002. Artificial nest structure use, post- fledging habitat u,se, and dispersal of Barn Owls (Tyto alba pratincola) in the Delta Region of Arkansas. Thesis. Arkansas State University, Jonesboro, USA. Ripper, D., J. C. Bednarz, and D. E. Varland. 2007. Landscape use by Hairy Woodpeckers in managed forests of northwestern Washington. Journal of Wildlife Management 71:2612-2623. SAS Institute. 1999. SAS/STAT user's guide. Version 8.2. SAS Institute Inc., Cary, North Carolina, USA. SoKAL, R. R. AND F. J. Rohlf. 1995. Biometry: the principles and practices of statistics in biological research. W. H. Freeman and Company, New York, USA. St. Pierre, A. M. 2006. Reproductive ecology of Swallow- tailed and Mississippi kites in southeastern Arkansas. Thesis. Arkansas State University, Jonesboro, USA. Swihart, R. K. and N. a. Slade. 1985. Testing for independence of observations in animal movements. Ecology 66:1 176-1 184. White, G. C. and R. A. Garrott. 1990. Analysis of wildlife radio-tracking data. Academic Press, New York. USA. Whitmar II, W. 1987. Quantification of the nesting site components in the Mississippi Kite (Ictinia mississip- piensis). Thesis. Southeast Missouri State University, Cape Girardeau, USA. The li ilson Journal of Ornithology 1 22(4);7 16-724, 2010 BIOLOGICAL, GEOGRAPHICAL, AND CULTURAL ORIGINS OF THE LOON HUNTING TRADITION IN CARTERET COUNTY, NORTH CAROLINA STORRS L. OLSON, HORACE LOFTIN,- AND STEVE GOODWIN' ABSTRACT. — A tradition of shooting Common Loons (Gavici immer) for food and for bone fishing lures was established on Shackleford Banks, North Carolina, by the mid- 19th century. This strongly ingrained tradition continued to be maintained, primarily by residents ot nearby Markers Island, when inhabitants of the banks moved inland about 1899. The practice probably arose because, on the east/west-tending Shackleford Banks, loons migrating northward in spring flew sufficiently low over land to be within shotgun range. Spring loon shooting, although illegal since 1918, grew to the point that dozens of hunters might be present on the banks on a given day. A strict law enforcement crackdown on this activity began in 1950, and the banks were effectively shut down for loon shooting. Loons continued to be shot opportunistically nearby, but a growing cultural intolerance of this practice brought the loon hunting tradition to an end. We document the existing memories and the few, scattered written sources concerning this unique local interaction between humans and birds. Received 30 March 2010. Accepted 23 June 2010. On 6 May 1950, eleven federal and state wildlife enforcement officers staged a long planned, coordinated raid on Shackleford Banks, Carteret County, North Carolina, apprehending nearly 100 hunters, of whom 78 were formally charged with illegally shooting loons (Anony- mous 1950) in violation of the Migratory Bird Treaty Act of 1918. This locally notorious event marked both the apogee and the rapid demise of a demographically narrow but strongly ingrained regional tradition, dating to the previous century, in which loons were hunted both for food and for use of their bones in making fishing lures. Loons do not usually factor so directly in human economy, nor does a similar loon hunting tradition .seem to have evolved eksewhere, so we investi- gated whether some unique combination of biological and physiographical conditions may have given rise to the loon hunting tradition in Carteret County. We also attempted to document through interviews and published and unpublished local records, how the values and attitudes of one small community were molded by the perception of loons as an exploitable natural resource. The Carteret County loon hunting tradition has es.sentially been eradicated because of changes in the economic, legal, and ethical environment. Knowledge of this aspect of regional wildlife ' Division of Birds, National Mu.seiim of Natural History, .Smithsonian Institution. P. O. Box 37012, Washington, D.C. 20013. USA. ’Carteret County Historical Society, 1008 Arendell Street, Morehead City, NC 28557, USA. 'P. O. Box 2027, Beaufort, NC 28516, USA. ■* Corresponding author; e-mail: olsons@si.edu heritage is in danger of vanishing, hence the importance of its historical documentation. METHODS We interviewed residents of Carteret County, North Carolina, who had participated in loon hunting, or who had firsthand knowledge of some aspect of the circumstances under which loons were obtained or were directly employed for human use. We gathered numerous other sources that, although published, are of a very restricted or evanescent nature and mostly unavailable in major research libraries. These publications are held in the libraries of the Carteret County Historical Society, Morehead City, NC, and the Core Sound Waterfowl Museum and Heritage Center, Harkers Island, NC; photocopies of pertinent portions, as well as all of our interview notes, have been filed in the Smithsonian Institution Archives, Washington, D.C. GEOGRAPHICAL CONSIDERATIONS The key component of this story is a physio- graphical feature, Shackleford Banks, a long, 13.8 km (8.6 mi), narrow, barrier island with its northern shore separated from Harkers Island by Back Sound and whose southern shore is washed by the open Atlantic (Fig. 1). Unlike most Atlantic barrier islands that tend to run north to south, Shackleford Banks is oriented nearly east to west. The eastern end of the island is sheltered by Cape Lookout and lies somewhat farther south than the western end. This east/west orientation is of great significance when considering the spring migratory pathway of loons. 716 Olson et al. • LOON HUN LING IN NOR'LI I C'AROIJNA 717 FIG. 1. Outline map ot Carteret County, North Carolina, USA, showing localities mentioned in the text. Black emphasizes islands of particular importance in the history of loon hunting. Inset: Carteret County (in black) in relation to other counties of North Carolina. Shackleford Banks was settled by the early 1700s and later came to have a considerable human population concentrated in two communi- ties, Diamond City and Wade’s Shore (or Wade’s Hammock), whose inhabitants were mainly shore whalers who supplemented their livelihoods taking mullet (striped mullet Miigil cephalus) and porpoises (bottle-nosed dolphin Tursiops tnincatiis). A growing tendency for people to move off the banks to areas closer to commerce and other amenities was greatly accelerated in the late 19th century when the settlements were devastated, and Diamond City’s high protective sand dune eroded away, by the great San Ciriaco hurricane that hit the outer banks on 17 August 1899. With the realization that the banks no longer provided a secure place to live, the Shacklefordians migrated inland, settling as far west as Salter Path on Bogue Banks, in the “Promise’ Land’’ section of Morehead City, and to the northeast as far as Cedar Island. The majority, however, retreated to Markers Island. Shackleford Banks was essentially uninhabited by 1902 and was used mainly as pasturage for hooved stock for most of the 20th century. Much of the preceding was summarized from Stick (1958). Shackleford Banks became part of the Cape Lookout National Seashore in 1966 (Public Law 89-366). LOON BIOLOGY AND BEHAVIOR AS PERTAINS TO HUNTING The Common Loon (Gavia immer) is a laige, fish-eating waterbird that breeds on fresh-water lakes of boreal forests across North America and winters at sea along both coasts as far south as Mexico. It is a heavy bird with males reaching a maximum of 7.6 kg (16.8 lbs) and averaging almost 6 kg (13 lbs), with females averaging 4.7 kg ( 10.3 lbs) (Evers et al. 2010). There is considerable variation in weight (Evers et al. 2010), but spring migrants would probably seldom weigh much less than 4.5 kg (10 lbs). Loons probably were taken opportunistically for food in any subsistence culture, but they were seldom included among game birds desired by North American commercial waterfowlers or sportsmen. Not only are loons difficult to hunt, but because they are fish-eaters, there is a perception that they would taste “fishy’’ and not be particularly desirable for the table, much as fish-eating mergansers (Mergus) are generally disdained. Loons are foot-propelled diving birds that spend their entire lives at the surface of or 718 THE WILSON JOURNAL OL ORNITHOLOGY • VuL 122, No. 4, December 2010 underneath the water, except when flying or incubating eggs. The Common Loon has a very high wing-loading and requires a long run across water to get airborne (Savile 1957, Evers et al. 2010). Common Loons do not experience an energetically costly wing molt on the lakes of their northern breeding areas, which freeze over soon after the young have fledged. Thus, molt is delayed until they arrive in coastal wintering areas where all flight feathers are molted simultane- ously and the birds are flightless until the new feathers grow in (Woolfenden 1967). Loons are not only reluctant to fly at .sea in winter because it is energetically costly, but for considerable periods of time they are incapable of flight. Loons are solitary or thinly distributed in winter and do not respond to decoys as do waterfowl. They also ride low on the surface of the water, making them a difficult target for hunters. Pursuit by boat, even rapid motorized craft, may be ineffectual because once a loon dives it is difficult to predict when and where it may briefly surface. The shooting journal of George Henry Mackay ( 1929) exemplifies the difficulty of shooting loons under normal circumstances and their greater vulnerability during their northward flight in spring. Mackay was an avid hunter of waterfowl and shorebirds as well as an amateur ornithologist and conservationist who kept a detailed Journal, including numbers of all species shot, of his hunting trips in New England from 1865 to 1922. Of the thousands of birds recorded in his diary, he shot only 22 loons, of which 19 were taken at West Island, Rhode Island, on dates spanning 13 to 25 April, when birds were evidently passing over land shortly after departing on their spring flight north. In spring migration, loons lly northward either along the coast or overland (Evers et al. 2010), but once having left the water they gain altitude and fly far above shotgun range. Birds along most of the north/south-trending eastern Atlantic Coast reach such high altitudes over water, perhaps offshore out of sight of land, and are thus not su.sceptible to shooting. Along .Shackleford Banks, however, loons accu- mulate off the east/west running ocean shoreline in spring and take off directly over land where they can be intercepted by gunners. The original Shacklefordians took advantage of this Haw in the loons’ usual immunity to put meat on their table. Markers Island later became the focal point for the continuation of loon hunting because of its proximity to the banks and the greater number of former Shackleford residents there (Eig. 1). Many of the same conditions apply to Bogue Banks. Residents of the settlement of Salter Path in particular were known for shooting and relishing loons (Stephens 1984, Dudley 1993). Eikewise, banksmen who settled on the mainland at Broad Creek crossed the sound to Bogue Banks for morning loon shoots near what is now known as Emerald Isle (Stephens 1984). We were told, however, that more loons cross at Shackleford Banks than farther west and this may come from a preference of loons to continue their northward flight over the waters of Core Sound, rather than over land. LOON HUNTING IN CARTERET COUNTY, NORTH CAROLINA We constructed the following history from interviews with local residents conducted in 2008 through 2010, augmented by a few scattered, generally recondite published sourees. Most respondents’ personal reeol lections date to the 1940s and early 1950s. Exactly when the loon hunting tradition arose is unrecorded, but one respondent’s father told him that it was estab- lished by at least the mid- 1800s. That seems reasonable considering that the practiee must have been in place well before the Shaekleford diaspora following the hurricane of 1899, as also suggested by Guthrie (1950). On Markers Island, according to Paul and Paul (1996:13), the “sport... has been a part of their lives since the earliest days on the Island.’’ Guthrie (1993:31) considered it “a favorite sport... since the conception of ‘Diamond City’ over a hundred years ago.’’ Loon hunting oecurred in spring, beginning in March but mostly in April through mid-May. The usual procedure was for hunters to depart Markers Island before daylight and leave their boats on the protected inner shore of Shackleford Banks. They dispersed on foot across the dunes to points along the Atlantic beach (Fig. 2) to await the flight of loons coming off the ocean that usually began at first light. Hunts occurred spontaneously as there was no deliberate organization. The recollection of one resident of Markers Island was that shooting began so regularly at daybreak that “it was as good an alarm cloek as you could ask for’’ and that at one point during World War II the shooting sounded like an invasion. The loons flew at their lowest altitude early in the morning and attained greater heights as the day progressed. Olson Cl al. • LOON IIUN TING IN NOR 11 1 CAROLINA 719 FIG. 2. Markers Island resident Otis C. Willis (b. November 1918, d. June 1996) with the results of a loon hunt in 1943. Barely visible in the background in the photograph on the left is the Cape Lookout lighthouse and two boats proceeding northward through “The Drain’’ or Barden Inlet, which was created during the hurricane of September 1933. Thus, the photograph was taken at the easternmost end of Shackleford Banks. This photograph was incorporated as a patch entitled “The Loon Hunt” (illustrated in Amspacher 2009:94) in a community quilt on display in the Core Sound Waterfowl Museum and Heritage Center, Markers Island, North Carolina. Wind direction had an effect on flights of loons. A northeast wind was favored for shooting because a head wind made it easier for loons to take off from the water and fly directly over the banks. A southwest tailwind would keep birds flying for a greater part of the day and was more advanta- geous for birds once airborne, but those birds would likely have had to take off farther out to sea and turn before crossing the banks. We were told that certain hunters once sank barrels in the sand for concealment (or shelter?), but no blind was usually deemed necessary beyond perhaps crouching behind any convenient large object that might have washed ashore. Anti- submarine buoys that had broken loo.se were mentioned as one such object used during and immediately after WW 11. At the height of the loon shooting, hunters spaced themselves so regularly along the beach that there was hardly a place a loon could cross the banks without being in range of .someone's shotgun. At times hunters were forced to stand behind one another. Loons are large and tough, and heavy shot, usually #2 or #4, was used. Even experienced gunners could be foiled at times, as one former shooter remembered expending a box of 25 shotgun shells without bringing down a single loon, although he then bagged five with his next five shots. The object was to shoot the birds so they would fall on the beach or in the dunes behind. Wounded birds that fell in the water might escape either by diving and swimming away, or, when dead, might float off if winds or cunents did not bring them back to the beach. Birds were not field dressed and were taken back intact, which meant an arduous trek back over the island with a heavy 720 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 4. December 2010 load. One respondent recalled a day. when he was too small to carry a gun, that he had accompanied his father and uncle to the banks on a morning when they shot and retrieved 18 loons. That was probably near the practical upper limit for two people to carry because 18 loons together would have weighed well over 90 kg (200 lbs). That day’s take was transported to the boat dangling from a pole carried over the shoulders of the hunters, a common practice, save for one loon that the youngster dragged behind him. Later, on Markers Island, the boy sold the birds for 50 cents apiece or whatever anyone would give him. There was no fall loon shoot because birds would be flying high and sometimes at night when arriving from the north and crossing land. But once the local appetite for loons had been established, birds might be taken opportunistical- ly, other than in spring, on the banks. One respondent told of becoming adept at shooting loons on the water with a pistol. Stephens (1984:124) reported that Bogue Islanders “used to go to the [Bogue] sound and firelight them in November,” which would also involve shooting birds on the surface of the water, but in that case at night with aid of a lantern. USES OF LOONS Food. — The primary impetus for hunting loons was for food. For 19th century residents of Shackleford Banks who had survived a winter on salt fish, root crops, and such grits and flour as they may have received in trade for mullet and whale oil in the previous months, a loon represented a sizeable chunk of fresh meat as well as a welcome change in diet. Loons do not migrate north until well after most ducks and geese have left for breeding areas and the “hardy group of fisher-folk” on Markers Island “would rather have a ‘loon-in-the-pot’ than the more popular goose, canvasback, or red-head duck” (Paul and Paul 1996:13-14). We made a point of asking whether cormorants (Plialcicrocorca) were ever shot for food. It was admitted that a cormorant would probably be just as good to eat as a loon but they were not taken. One of Stephens’ (1984:124) sources said that “1 love the skin the best in the world” and Dudley (1993:137), whose .source may have been the same, mentions that the birds were sometimes “picked.” The prospect of plucking a loon is daunting and would require a great amount of work, so that a person would really have to desire the skin to go to the effort. Fat is what gives any kind of meat most of its flavor and most of a loon’s fat is either attached to the skin or sequestered among the viscera. A loon with the skin on would likely be much more fishy-tasting than one that had the skin and feathers removed, although Dudley (1993:137) condescendingly avers that “fishy taste [is] barely discernable to the natives [of Markers Island and Salter Path].” By far the greater number of loons prepared for cooking were skinned and usually only the breast and thighs were consumed. We have been assured that the end result did not taste fishy. We inquired about loons of an acquaintance who came originally from Cedar Island and who had returned to Carteret County in retirement. Me related that he had never shot a loon but would know exactly what to do if someone gave him one, and he proceeded to describe the entire process from carcass to stew bowl. Jocularly known as “Markers Island turkey” (Amspacher 1987:217), additional sources for loon recipes (e.g., Stephens 1984, Amspacher 1987) are basically similar. The meat is soaked in water with salt or baking soda, browned in fried-out salt pork, and stewed to a thick gravy with some combination of onions, potatoes, rutabagas, and cornmeal dumplings. The process, which was at times done outside in large iron pots, would usually take most of a day. Fishing Lures. — A secondary, but once impor- tant use of loons, was for fishing lures made from the bones, which were used for trolling with handlines, mainly for bluefish (Pomatomus salt- atrix) and Spanish mackerel {Scomberomorus nuiculatus). M. Loftin remembers that in the 193()s loon bone lures were almost the only ones u.sed for trolling in the Beaufort area. In addition to bones from birds shot for food, during WWII when tankers along the east coast were being sunk by German U-boats, bones for lures also came from loons that died from spilled oil and were cast up on the beach. Almost all of our sources, both oral and published (e.g., Dudley 1993:137, Guthrie 1993, Paul and Paul 1996) insist that leg bones were used for the fishing lures and it seems from the descriptions that the tibiotarsus (drumstick) was meant. The femur of a loon is extremely short and curved, and the tarsometatarsus (foot) is flat and solid, so that neither of those bones would have been useful. That wing bones were also used is evident from the report that: “Watermen also Olson cl al. • LOON HUNTING IN NORT H CAROLINA 721 FIG. 3. Bones (A, B) of Common Loon (Gavia inimer) and trolling lures (C-E) made from loon bones. A: complete left tibiotarsus and fibula; B: complete right humerus; C, D: from Dudley (1993:136); E: from collection of Roy H. Lewis. Stacy. North Carolina (photograph by J. R. Humphrey). The notch in D is clearly from hook wear. want the wing and leg bones of the loons. They bleach them in the sun and cut them into 2 'A -inch [63 mm] lengths for fishing lures. The hollow bones are sometimes the only lures that will catch bluefish or Spanish mackerel” (Dudley 2001: 160). Furthermore, Dudley (2001:165), relates that ‘‘in olden days the hunters sold the leg and wing bones to Cheek’s Hardware in Morehead City for 10 cents a piece.” The only loon bone lures that we have been able to trace (Fig. 3 C-E) were clearly not made from any leg bone. They are all simple, completely terete cylinders that can only have been made from the humerus, the upper wing bone (Fig. 3B). The bone was beveled at one or both ends and the only one we examined personally (Fig. 3E) appeared to have been pared down to a smaller diameter. These bone cylinders were wired to a hook and the wire was said to have been twisted at the leading end of the bone in such a way as to give it a movement under water that was attractive to fish. An elderly resident of Markers Island who agreed to demonstrate how to make a lure from a loon bone protested that the leg bones presented to him were too small for the purpose, lending further support to the available evidence that it was the humerus that was fashioned into lures. It is possible that the tibiotarsus was also used, as its flattened proximal surfaces (Fig. 3A) would probably have imparted an erratic motion under water that might have been attractive to fish. Such a lure would have had thinner walls and been more subject to splitting than one made from a humerus, which may account for why none seems to have survived if they ever existed. The walls of the humerus of a loon are much heavier and denser than in any bone of a duck or goose, which, had they been used for lures, would probably not have lasted long before being broken. Following the demise of loon hunting, the supply of loon bones was greatly diminished and following WW II, loon bone lures were replaced by metal spoons and other trolling devices. Loon bone lures have now nearly completely disap- peared from tackle boxes and tool sheds of Markers Islanders. LORE AND LEGENDS OE LOONS IN CARTERET COUNTY It is natural that an activity that was such an integral part of local traditions as loon hunting would be incoiporated into local storytelling and literature. More or less apocryphal stories about loon hunting on Shackleford Banks are still in circulation and may be recognized in several variants. One involves a newcomer to the sport who was taken to the banks one morning and witnessed a loon flying parallel to the beach and having each hunter along the line shoot feathers out of it until it kept on flying off into the distance nearly nude, provoking the hope that it would freeze to death. A version of this is told by Paul and Paul (1996). As loons were normally shot Hying over the island, not parallel to it, and loons are not prone to losing many feathers when shot, this is likely a carryover from some dove hunter's tale. A similar anecdote that may be related involves a champion skeet shooter being invited to try his prowess on the banks who utterly failed to bring down a loon. In contrast with those failures, another story concerns a hunter returning from the banks to Markers Island with a large haul. He repaired to the local movie theater, where he knew a large portion of the population would be, and announced that he had a skiff full of loons if anybody wanted one. Whereupon, the 722 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 4, December 2010 theater emptied as people made haste for the skiff to get their loon. Shooting and eating loons has also been incorporated into local humor, such as the joke about the man who was arrested and confessed to having shot a “seagull.” Before sentencing, the judge asked the defendant why he wanted to shoot a gull in the first place, and the answer was that if it was cooked right it was “nigh bout goods a loon” (Williamson 2009:44). Even poetic inspiration has been invoked in loon lore. We were told by several residents of Markers Island that they recalled a high school boy in the 1950s who had written what we were first told was a story but were later assured was a poem about loon hunting, but from the perspective of the loon. We have not been able to trace a copy of said poem but it must have been an evocative one for its existence to be recalled more than half a century later. A more concrete literary contri- bution to the subject is a mystery novel. Shooting at Loons (Maron 1994), set in Carteret County, in which the prediliction of certain Markers Island fishermen to poach loons illegally and make a stew of them is mentioned more than once. TME END OF TME LOON MUNTING TRADITION It became illegal to take or possess loons in 1918 with the adoption of the Federal Migratory Bird Treaty Act. Knowledge and enforcement of the law would doubtless have been scant and sporadic at first and the hardship years of the Great Depression, followed by World War II, would have made a limited local penchant for loon meat seem like a paltry Federal offense. But afterwards the Shackleford loon hunt became a victim of its own success. Too many people were participating in the hunt and were being joined to .some extent by “sportsmen” from “off,” whose main interest in loons would have been in extending their season to be afield with a gun. This created an additional problem of birds left dead on the beach by shooters who did not share the locals’ appetite for stewed loon. Such conspicuous illegal activity could no longer be ignored. In the late 1940s an undercover wildlife agent settled on Markers Island with his family, where he came to be trusted by local residents, who conversed with him about fishing and hunting, including loon shooting. It was later perceived that his main purpose on Markers Island, however. had been to set up the inevitable raid on the Shackleford Banks loon shoot (e.g., Guthrie 1993), a perception that was reinforced by the tact that he left the island before or immediately after the raid. The raid occurred on Saturday, 6 May 1950, and was led by Robert Malstead, a Federal agent of the U.S. Fish and Wildlife Service, who was joined by “Andrew Jones, chief state game protector” and nine other state agents. Some 100 loon hunters were “rounded up,” but only 78 were actually charged, of whom “approximately 70” were from Markers Island. Fifty loons were seized as evidence and Jones “estimated that 200 more were killed.” We were told that some 25 or 30 other hunters escaped, which does not seem unlikely given the low ratio of agents to hunters. One of our sources, a teenager at the time, was among the escapees. The agents estimated that hunters were scattered “150 feet [45.7 m] apart for a distance of almost two miles [3.2 km]” and fired from “2000 to 2500 shots” in the first 15 min after the first shot at 0430 hrs. The math does not add up but it is clear there was a lot of shooting by a lot of people on that day. Decades later, Malstead provided his somewhat dramatized recollection of that day, complete with a photograph of 10 of the agents (the 1 1 th was presumably the photographer) with some of the confiscated loons (Dudley 2001:159- 163) but we have taken all portions in quotes above from a more immediate source (Anony- mous 1950). Those hunters who were convicted in federal court were fined $25, which is equivalent to about $225 in 2009 dollars, an amount that would have made an impression on those who had to earn that $25 by their labors on the waters. One may imagine that $1,750 of the entire 1950 economy of Markers Island had a much more severe impact than would be conveyed simply by converting that figure into its modern spending power. In retrospect, the May 1950 raid was a great deterrent success for law enforcement. There may have been no particular stigma (in fact perhaps the opposite) for most of those apprehended as loon shooters in Carteret County, but the public could no longer be unaware that such activity was egregiously illegal. The enforcers kept up their vigilance on Shackleford Banks for the next 2- 3 years, after which any .semblance of the loon shoots of old was abandoned. Loon shooting persisted and probably still goes on to a limited extent. Munting on Shackleford Olsun et al. • LOON HIINTINC] IN NOR I II CAROLINA 723 Bunks required leaving one’s boat on the sound side to hunt on the ocean side, and, as a federally protected reserve, shooting there became too risky. But the occasional loon would still be taken along the “island shore,’’ meaning the south shore of Harkers Island. One such incident resulted in a falling loon hitting a power line on the way down and knocking out electricity for part of the island, which “did not go over’’ well among some of those in positions of authority at the time. Even after the law had come down hard in 1950, residents of Salter Path and elsewhere on Bogue Banks continued to “sneak to the beach hills and elsewhere in an attempt to get a loon.... Some of them didn’t kill many the whole spring. But if they killed two or three, they divided them with their family connections. It was a big thing. They called in their sister or father or mother or something like that’’ and still do so “if they can get a hold of them’’ (Stephens 1984:124). Dudley (1993:137) reported that at “Salter Path, the hunters sat on high knolls or dunes anticipating loon flights over low-lying areas or swashes. With the pressure of development and high-rise condo- miniums, this practice has nearly ceased. Today, loon hunters position themselves around their homes, or wherever they can obtain visibility of loon flights, and the shooting continues. Gun shots are still heard, and game wardens still abound.’’ Given the pace of development on Bogue Banks since that passage was written, the incidence of surreptitious loon shooting there has by now probably all but vanished. EPILOGUE An impassioned, even lyrical, paean evoking the emotions and near ritual importance of the loon hunt was printed in the Carteret County News- Times shortly after the great raid of May 1950 (Guthrie 1950). Guthrie contended that the laws that overregulated hunting had come into effect because of “fancy” “upstate” sportsmen, where- as the Carteret County loon shooters participated in the “only exciting sport left untouched by the law. Their father had done it, their grandfather before, even before their forefather had moved from the outer banks. You weren’t a man until you had shot a loon.” Hours before daylight, fathers would rouse sons too small to carry a gun from their beds to cross the sound to the banks and participate in the Harkers Island rites of spring. In the later days of shooting on the banks, “the excitement of the forbidden hunt [would add] zest to the sport” (Paul and Paul 1996:13). One of our respondents, now a pillar of the Harkers Island community, stirring up recollections over 60 years old, admitted that although he did not really know what it was like to be addicted to anything, he thought that he had been a loon hunting addict. The motivation was not the thought of a meal, it was the uniqueness of “the sport.” Old traditions die hard, but shooting and eating loons as part of the economy and culture of coastal Carteret County is now a nearly forgotten part of the past. The Common Loon has become the object of such romantic and mythical sentimentality in most of North America (Evers et al. 2010) that there could now be no tolerance for even a small unsanctioned harvest of this species. The day is not far away when no one on Harkers Island will recall the taste of stewed loon, just as there is almost no one alive who remembers how to fashion a trolling lure from a loon bone. This interesting and unique local tradition between humans and birds now exists almost only as a historical phenomenon that deserves to go on record. ACKNOWLEDGMENTS We are deeply indebted to Ira Lewis, James Lewis, James Lewis Jr., Richard Lewis, Frank Moore, and James Rose for sharing their recollections with us. We are most grateful to Barbara Willis Yeomans for supplying the photographs of her father, Otis Willis. Roy H. Willis allowed us to study and photograph a loon bone lure in his collection and Jack Dudley allowed us to reproduce figures of lures from his book. We thank Jack Spencer Goodwin, Johanna R. Humphrey, and Rodney Kemp for assistance with the manuscript. We are indebted to the staff of the Carteret County Historical Society. Morehead City, NC. and the Core Sound Waterfowl Museum and Heritage Center, Harkers Island, NC, for numerous considerations. We thank Brian Schmidt and Christina Gebhard for their creative work with the figures. We also thank G. R. Graves for his comments on an earlier manuscript and H. M. Reeves for his most insightful remarks and for providing an important reference. LITERATURE CITED Amspacher, K. W. (Editor). 1987. Island born and bred. Tenth Printing. Harkers Island United Methodist Women, Harkers Island, North Carolina, USA. AM.SPACHER, K. W. (Editor). 2009. Our Down East memories. Down East community quilt. Quilt Number 1. Core Sound Waterfowl & Heritage Mu,seum, Harkers Island. North Carolina, USA. Anonymous. 1950. Game protectors .seize 78 hunters at Cape Lookout. Carteret County News-Times, Tues- day, 9 May 1950: page 1. 724 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4, December 2010 Dudley, J. 1993. Carteret waterfowl heritage. Decoy Magazine. Burtonsville. Maryland, USA. Dudley, J. 2001. Wings; North Carolina waterfowling traditions. Coastal Heritage Series, Morehead City, North Carolina, USA. Evers, D. C., J. D. Paruk. J. W. McIntyre, and J. L. Barr. 2010. Common Loon Gavia iminer. The birds of North America. Number 313. Guthrie, J. L. 1993. The last loon hunt. Mailboat 3^:31- 33. [A short-lived newsletter published on Markers Island.) Guthrie, S. 1950. Is it a crime to kill a loon for food? Carteret County News-Times, Friday, 12 May 1950. Mackay, G. H. 1929. Shooting journal of George Henry Mackay. John C. Phillips, Cambridge, Massachusetts, USA (Undated reprint by Kessinger Publishing [no city of publication] with 3 intercalated, unnumbered pages of "Introduction” by H. M. Reeves dated 1 July 2008). Maron, M. 1994. Shooting at loons. Mysterious Press, New York, USA. Paul, M. and G. Paul. 1996. Carteret County, NC: folklore, facts, and fiction. Beaufort Historical Asso- ciation. Beaufort, North Carolina, USA. Savile, D. B. O. 1957. Adaptive evolution in the avian wing. Evolution 11:212-224. Stephens, K. H. R. 1984. Judgment land: the story of Salter Path. Published by the author, Havelock, North Carolina, USA. Stick, D. 1958. The Outer Banks of North Carolina, 1584- 1959. University of North Carolina Press. Chapel Hill, USA. Williamson, Sonny [C. L). 2009. Salt spots for breakfast. Published by the author, Marshallberg, North Carolina, USA. Woolfenden, G. E. 1967. Selection for a delayed simultaneous wing molt in loons (Gaviidae). Wilson Bulletin 79:416-420. The Wilson Journal of Oniitlioloi'y 122(4):725-737, 2010 MIGRANT SONGBIRD SPECIES DISTRIBUTION AND HABITAT USE DURING STOPOVER ON TWO ISLANDS IN THE GULE OE MAINE REBECCA W. SUOMALA,' SARA R. MORRIS, ^ KIMBERLY J. BABBITT,' AND THOMAS D. LEE' ABSTRACT. — We compared the distribution of migrant bird species between two islands in the Gulf of Maine to e.xamine it ditterences in habitat resulted in differences in avian species composition and relative abundance during stopover. Ninety-one species were captured on both islands and those species captured on only one island were either breeding species or rare visitors to the islands. Ditterences in bird species distribution between islands were species-specific and consistent among sampling periods for nearly all species. Twelve species were captured more frequently on Star Island and 1 1 species more frequently on Appledore Island. Stopover species distribution appeared to be related to habitat structure, vegetation, diet, and habitat area. Scrub-shrub/open habitat breeding species and forest breeding species were not evenly distributed between islands. Island use was most clo.sely associated with breeding habitat. All but two of the eight species that breed in scrub-shrub or open habitat were captured more frequently on Star Island. Ten of the species more common on Appledore Island breed in forested habitat. Nine of the 1 1 species more common on Appledore Island are area- sensitive in breeding areas, suggesting potential area sensitivity during migration. Differential habitat use indicates a large number of stopover sites in a wide variety of habitats are necessary to meet migration needs of passerine species. Received 17 November 2006. Accepted 22 April 2010. Suitable stopover habitat is critical to a successful migration (Moore et al. 1995) and loss of stopover habitat may be a factor in long-term population declines (Sherry and Holmes 1995, Finch and Wang 2000, Hutto 2000). Migrants have increased risk of mortality during migration (Bairlein 1992; Moore et al. 1992, 1995; Hutto 1998; Sillett and Holmes 2002.) without habitat that can satisfy their needs to replenish critical energy supplies, reorient (Baird and Nisbet 1960), escape adverse winds and weather, recover from muscle fatigue or injury, and avoid the dehydra- tion of daytime flight. More information is needed on habitat requirements of passerines during migration (Moore et al. 1992, 1995; Sherry and Holmes 1995; Hutto 2000; Petit 2000) to understand what constitutes valuable stopover habitat and influences habitat selection (Moore et al. 1995, Dunn 2000). Migrants show differential habitat use including species-specific di.stribution among stopover hab- itats (Hutto 1985, Moore et al. 1990, Winker 1995), but the basis for habitat selection is not clear. Finch and Wang (2000) reported bird densities during migration were affected by ' Department of Natural Resources and the Environment, 206 Nesmith Hall, University of New Hampshire, Durham, NH 03824, USA. ’Canisius College, Department of Biology, 2001 Main Street, Buffalo, NY 14208, USA. Current address: New Hampshire Audubon Society, 84 Silk Farm Road, Concord, NH 03301, USA. ■* Corresponding author; e-mail: bsuomala@nhaudubon.org vegetation structure and habitat type, but few studies have examined fine-scale habitat relation- ships during migration (Rodewald and Britting- ham 2004). There is some evidence that songbirds may select stopover habitat with vegetative characteristics similar to their breeding habitat (Parnell 1969, Petit 2000). Other factors affecting stopover habitat use may include suitability to a species’ foraging method (Bairlein 1992), food availability (Hutto 1985, Moore et al. 1995), dietary shifts (Parrish 2000), predation and competition (Moore et al. 1992), and site characteristics including area (Moore et al. 1995), temperature, and humidity (Winker 1995). Star and Appledore islands are part of the Isles of Shoals on the New England coast, USA. They provide a unique opportunity to examine the relationship between habitat and migrant stopover patterns because large numbers of songbirds stop at the archipelago during spring and fall migra- tions (Morris et al. 1994, 1996), and the islands are <1 km apart. Weather and body mass can influence selection of stopover sites (Moore and Simons 1992) and the potential confounding effect that different weather conditions have on migration behavior can be eliminated by compar- ing two adjacent sites. Both islands are catego- rized as maritime shrub thicket (Sperduto and Nichols 2004), which encompasses a range from low-growing shrubs to tall thickets; comparison of these islands presents an opportunity to examine the relationship of birds and stopover habitat at a fine scale. Species richness and abundance in 725 726 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 4. December 2010 migration may be correlated with patch size {Martin 1980, Blake 1986, Somershoe and Chandler 2004), as has been shown during the breeding season when number of species and individuals increases with area (Freemark and Merriam 1986, Blake and Karr 1987, Faaborg et al. 1995). Star Island is smaller than Appledore Island allowing consideration of area effects on species composition during stopovers. Our objectives were to examine: (1) whether there were differences in species richness, com- position, and distribution between the islands; and (2) whether a species’ breeding habitat or food habits were related to stopover habitat use. We predicted more species would be captured on Appledore Island and that habitat differences between islands would result in differences in species composition and frequency of capture. METHODS Study Site. — The Isles of Shoals is a group of nine small islands and several ledges in the Gulf of Maine, 14.5 km southeast of Portsmouth, New Hampshire (42° 58' N, 70° 36' W), and 9.7 km from the nearest point of the mainland. Appledore Island, Maine is the largest of the islands at 33.6 ha and Star Island, New Hampshire is 13.4 ha. The minimum shore-to-shore distance between Star and Appledore islands is ~0.7 km and the study sites were 1 .6 km apart. Bird Surveys. — Bird surveys were conducted simultaneously on both islands in 1999 and 2000 using mist nets during spring and fall migration following the protocol established for the Apple- dore Island Migration Banding Station (Morris et al. 1994, 1996). There were four sampling periods: 11 May-8 June 1999, 10 May-8 June 2000, and 16 August-30 September 1999 and 2000. Weather permitting, mist nets (6 or 12 X 2.6 m, 4 shelves, 30 mm mesh) were operated during daylight hours with the nets opened just before sunrise, closed at sunset, and checked every 30 min throughout the day. Up to five nets were operated on Star Island and up to 10 nets were operated on Appledore Island. Captured birds were brought to a central location on each island and banded with USGS aluminum bands. Habitat Surveys. — We collected habitat data al net sites rather than at random locations distrib- uted over each entire island. There is extreme landscape variability on both islands with habitat ranging from rocky splash zones or mowed lawns to thick scrub-shrub relatively close to the nets. We characterized the natural vegetation that surrounded the mist nets to ensure data were representative of the habitat used by the species captured. These data should not be considered representative of the habitat on the entire island. Vegetative structure and composition at each site were measured using a modified version of James and Shugart (1970) and Winker et al. (1992). We established eight circular survey plots, each centered on a pair of mist nets (or a single net, if there was only one in a given area). There were three plots on Star Island and five on Appledore Island covering all net locations. Each plot was 0.04 ha (radius = 1 1 .35 m) and was divided into two semicircles, separated by the net lane (1-2 m) to remove the non-vegetated lane from analysis. We established six 1.5 m-diameter circular subplots in each plot using a stratified random design so each subplot was entirely within the larger plot. Subplots were located by extending a random distance into alternating halves of the plot from six equidistant points (3.94 m apart) along the diameter of the plot, excluding the outer 1 .5 m. We established a 1 m-wide transect in each half of the plot to measure shrub density. Transects were parallel to the net lane and 0.3 m from the edge of the net lane to reduce the influence of trimming along net lanes. We defined trees as woody stems with a diameter at breast height (dbh) >3.0 cm and shrubs as any woody vegetation <3.0 cm dbh and over 0.75 m in height. We recorded live tree species, height, and dbh (0.1 -cm diameter tape) in each plot. We recorded only dbh for dead trees. We recorded species and height for each live shrub stem in the transect, and counted stems separately if they branched within 0. 1-0.2 m above the ground. We counted the number of stems for dead shrubs. Heights for live trees and shrubs were measured to the nearest 0.1 m. Separate measurements were taken for each species in a cluster. We measured herbaceous cover, canopy cover, and vegetation density in each subplot. We visually estimated herbaceous cover in each subplot to the nearest 5%. We visually estimated canopy cover from the center of each subplot and recorded cover to the nearest 5% at six l-m height classes beginning with 0-1.0 m (to account for low woody plant cover that functions as canopy for ground-feeding birds). We used a density board (Nudds 1977) placed at the center of a Sinmuila cl ciL • MIGRATION STOPOVER DISTRIBUTION AND HABITAT USE 727 subplot to measure vegetation density and esti- mated percent cover from 2 m distant, rather than the recommended \5 m (Nudds 1977), because of the extreme density of the shrub vegetation. We chose a random direction using eight possible points of the compass and rejected points that were outside the full plot after walking 2 m from the board. We recorded vegetation cover to the nearest 5% in each of the four sections of the density board (0-0.5, 0.5- 1.0, 1.0-1. 5, 1.5-2.0m). All habitat data were collected between 6 May and 10 June 2000 by RWS to eliminate variability among observers. Subplot measurements were obtained between 5 and 10 June to ensure that vegetation density and canopy cover measure- ments were at the same time after leaf out. Data Analysis. — We used Systat Version 10 (SPSS Inc. 2000) for statistical analyses. The analyses, except bird species richness, include only those species that do not breed on the Isles of Shoals (i.e., only stopover migrants). We used raw counts of birds captured for all Chi-square tests. We used a heterogeneity Chi-square analysis (Zar 1999) to test for differences in species distribution between islands and analyzed each sampling period separately. We used Freeman- Tukey deviates in Systat to examine which species were responsible for significant Chi- square tests. We removed those species from the analysis until there was no difference {P > 0.05) between islands. Expected values for individual species were calculated using the total number of migrant birds captured on each island. The heterogeneity test was the prime analysis used in examining differences between islands, even though expected values for some .species were below the desired criteria for Chi-square tests. We u.sed a Chi-square goodness-of-fit test for each species as a .secondary analysis to compare number of individuals captured between islands without violating the expected value criteria. Expected values were >5 in all tests except Purple Pinch [Carpodacus purpureu.s) and Savan- nah Sparrow (Passerciilus sandwichensi.s) in spring 2000. We followed Jones et al. (2002) and the recommendations of Gotelli and Ellison (2004), and did not use Bonferroni corrections in the analyses to reduce the likelihood that real differences or relevant patterns would be missed. Bonferroni coiTections are not as practical for diverse communities requiring multiple statistical tests (Moran 2003) such as the number of migrant birds in this study. It is unlikely the results from all tests will be spurious (Moran 2003). Species were included in the distribution analysis if there were > 1 0 individuals banded on at least one island in any two sampling periods, and were analyzed in only those sampling periods when their numbers met this minimum criterion. The only exception was Purple Pinch, which is an irruptive migrant present in large numbers during fall 1999 only. We included it in the spring 2000 analysis, although only eight individuals were banded, to examine if the return migration in spring had the same distributional pattern. Recaptures were not included in the distribution analysis. We used a Mann-Whitney f/-test to compare mean number of tree and shrub stems per plot between islands for each species with more than 20 stems on one of the islands. We computed the mean per plot for each island and compared means between islands using a Mann-Whitney f/-test for the habitat data; live tree and shrub density (all species combined), dead tree and shrub density, tree and shrub height, and tree dbh. A mean for each individual plot was calculated first for the latter three measures, which were used to calculate the mean per plot for the entire island. We divided tree heights into five 1-m height classes beginning with 2.0 m and shrub heights into five 0.5-m height classes beginning with 0.5 m. We compared the distribution of trees and shrubs by height class between islands with a heterogeneity Chi-square analysis, and used a Mann-Whitney (/-test to examine differences among mean number of stems per plot in each height class. We divided tree dbh into seven size classes: 3-3.9, 4-4.9, 5-5.9, 6-6.9, 7-10.9, II- 15.9, and >16.0 cm. The larger size classes with few trees were combined to create suitably sized groupings. We compared the distribution of trees by dbh class between islands with a heterogeneity Chi- square analysis. We used ANOVA to compare percent herbaceous cover between islands and M ANOVA with associated univariate tests, when the M ANOVA was significant, to compare mean vegetation density at four height classes and canopy cover at six height cla.s.ses between islands. We assigned habitat types to each bird .species for breeding habitat to compare species differenc- es between islands with habitat associations in breeding areas. Habitat classifications were: early successional, forest, forest edge, low forests at high elevations and in bogs, scrub-shrub, and open areas (including grasslands and emergent wet- 728 THE WILSON JOL'RNAL OF ORNITHOLOGY • Vol. 122, No. 4. December 2010 lands). We also examined vegetation type and area sensitivity in breeding areas. Area sensitivity information was obtained from Robbins et al. (1989) and Birds of North America species accounts (Poole and Gill 1993-1997, 1998- 2002); breeding habitat information was primarily from DeGraaf and Yamasaki (2001) supplement- ed by Ehrlich et al. (1988). We classified fall trugivory into four categories using the mean percent of fruit in fecal samples during fall migration on Block Island, Rhode Island (Pairish 1997); low 0-29, medium 30-59, medium-high 60-89, and high 90-100%. RESULTS Bird Species Richness. — We banded 5,443 birds (3,833 migrants) on Star Island and 8,372 (6,604 migrants) on Appledore Island during the four sampling periods. Mean (± SD) captures for Star Island were 1,234 ± 28.3 in spring and 1,487 ± 255.3 in fall, and for Appledore Island were 2,140 ± 302.6 in spring and 2,046 ± 233.3 in fall. Ninety-nine species were captured on Star Island and 102 on Appledore Island; 91 of the species were captured on both islands. Mean (± SD) species richness for all sampling periods com- bined was 68 ± 4.65 for Star Island and 74 ± 3. 1 6 for Appledore Island. Mean (± SD) species richness during spring was 64 ± 0.71 on Star Island and 71 ± 0.71 on Appledore Island; during fall it was 72 ± 0.71 on Star Island and 76 ± 2.12 on Appledore Island. Eight species were captured only on Star Island and I I species only on Appledore Island. Some species were captured infrequently and were rare visitors to the Isles of Shoals or unusual captures in songbird mist nets. More of these “accidental” species would be expected on Appledore Island due to probability (more net hours). Species richness on the two islands was identical at 70 species with 100% overlap in species composition between the islands when we removed breeding and accidental species from the analysis. Bird Species Disirihiilion. — The distribution of migrant species between islands differed in each sampling period (spring 1999: = 242.27, df = 28, P < 0.001 ; .spring 2000: f = 206. 19, df - 33, P < O.OOl; fall 1999: yj = 325.92, df = 31. R < O.OOl; fall 2000: f = 400.69, df = 34, P < O.OOl). The difference between islands was attributed to a fairly consistent group of species (Table I). Twelve species were captured more frequently than expected on Star Island (scientific names in Table 1): Traill’s Flycatcher, Ruby- crowned Kinglet, Cedar Waxwing, Magnolia Warbler, Yellow-rumped Warbler, Wilson’s War- bler, Yellow-breasted Chat (fall, not tested in spring). Savannah Sparrow (spring, not tested in fall), Lincoln’s Sparrow (fall but not spring). Swamp Sparrow, White-throated Sparrow (fall but not spring), and Purple Finch. Eleven species were captured more frequently than expected on Appledore Island (Table I ): Red-eyed Vireo, Brown Creeper (fall, not tested in spring), Veery (fall but not spring), Swainson’s Thrush (spring but not fall). Black- throated Blue Warbler, Black- and-white Warbler, American Redstart (fall but not spring), Ovenbird, Northern Waterthrush, Mourning Warbler (spring but not fall), and Canada Warbler (spring but not fall). Habitat. — Thirteen species of shrubs and eight species of trees were recorded on Star Island and 15 species of shrubs and nine species of trees on Appledore Island (Table 2). The most common shrub species in the netting area on Star Island were chokeberry, raspberry, and rose; the most common trees were winterberry, sumac, and bayberry (scientific names in Table 2). The most common shrubs in the netting area on Appledore Island were chokecheri'y, winterberry, and aiTow- wood; the most common trees were chokecheiry, serviceberry, and arrowwood. Only raspberry and winterberry were in the top five species on both islands; the other dominant species were typically scarce or absent on the other island. Baybeiry and huckleberry were recorded only on Star Island, and apple, arrowwood, black cherry, and pin cherry were recorded only on Appledore Island. The mean number of trees on Appledore Island was higher than on Star Island, but this difference was not significant for all heights combined (Table 3). One Star Island plot had 228 trees, but the mean of the two remaining plots was only 65 trees. No plot on Appledore Island had <142 trees. The distribution of trees over five height classes differed (y~ — 375.5, df = 4, R < 0.001) between islands (Fig. 1). Appledore Island had more trees per plot in the highest three classes and Star Island had more in the lowest class. The difference between islands was not significant for the 3-3.9 m height class. There were also fewer dead trees per plot on Star Island than Appledore Island (Table 3). Trees on Star Island had a smaller mean dbh per plot (range = 3.0-10.7 cm) than Appledore Island (range = 3.0-33.2 cm; Table 3). The distribution Snoniala cl at. • MIGRAI ION STOPOVER DISTRIBUTION AND HABITAT USE 729 ot trees over seven dbh classes dit't'ered (x“ = 50.6, dt = 6, P < O.OOl) between islands with fewer trees on Star Island in all dbh classes, and fewer than 10 trees in the four classes greater than 6.0 cm. Star Island had shorter mean tree height per plot than Appledore Island (Table 3). There were only six trees taller than 4.0 m on Star Island (3 plots), whereas Appledore Island had 529 that were 4.0 m or taller (5 plots). Shrub density was higher on Star Island than Appledore Island, primarily because of a differ- ence in the number of small-diameter shrubs (Table 3). Star Island had more dead shrubs per plot than Appledore Island but the difference was not significant (Table 3). Overall mean shrub height did not differ between islands (Table 3). However, distribution of shrubs over five height classes (Fig. 1) differed (yj = 729.5, df = 4, P < 0.001) between islands. Star Island had signifi- cantly more shrubs per plot in the lowest three classes and Appledore had significantly more in the highest height class. There was no difference (P = 0.881) between islands in the 2-2.4 m class. There was a slight difference between islands in vegetation density for all four heights combined (P4.43 = 2.48, P = 0.058). Star Island had significantly higher mean percent cover than Appledore Island in the lowest two height classes (0-0.5 and 0. 5-1.0 m) but there was no significant difference in the two upper heights. Percent canopy cover differed (p6.4i = 4.25, P = 0.002) between islands. Star Island had higher percent cover in the lowest two height classes from 0.0 to 2.0 m, and Appledore Island had higher percent cover in the highest three classes (3.1 m and higher); one height class (2.1- 3.0 m) did not differ between islands. Star Island had mean (± SD) herbaceous cover of 52.5 ± 34.2% and Appledore Island had a mean cover of 35.8 ± 35.0%, but the difference was not significant (Pi,46 = 2.69, P = 0.1 12). Comparison of Bird Distribution with Habi- tat.— There were fewer trees on Star Island and they were smaller in height and dbh than Appledore Island which had only six trees taller than 4.0 m (3 plots), whereas Appledore Island had 529 that were 4.0 m or taller (5 plots). Star Island had more shrubs in all height classes except the tallest (>2.5 m). and had the highest vegetation density in all height classes except the highest (1. 5-2.0 m). Appledore Island had more canopy cover except in the 0-2.0 m range. Thus, Star Island more closely resembled a typical scrub-shrub habitat in structure while taller thickets with higher canopy cover characterized Appledore Island. This difference allowed us to compare differences in bird distribution between islands in relation to habitat (Table 4). There were significant differences in distribu- tion between islands iy' = 4.5, df = 1, P = 0.033) when bird species were grouped based on breeding habitat as ( 1 ) scrub-shrub or open, or (2) forest. Six of the eight species that breed in scrub-shrub or open habitat were captured significantly more frequently on Star Island (Table 4); one of the other two showed no difference between islands (Chestnut-sided War- bler) and the other was more numerous on Appledore Island (Mourning Warbler). Six of 32 forest-breeding species were more common on Star Island, 10 were more common on Appledore Island, and 16 showed no difference between islands (Table 4). Nine of 12 forest- breeding species that are area-sensitive in breeding areas were more common on Appledore Island, and the other three species showed no difference between islands (Table 4). DISCUSSION Star Island more closely resembled a typical scrub-shrub habitat in structure with denser shrub vegetation, more trees, and greater canopy cover <3.0 m. Appledore Island had greater tree density, canopy cover, and shrub density >3.0 m. Bird species richness and composition varied minimally between islands; however, relative abundance of some species varied greatly proba- bly because of vegetation differences. These results are consistent with Rodewald and Britting- ham (2004) who reported relative abundance during migration varied among habitats despite broad habitat use by all species. Differences in species distribution between Star and Appledore islands were species-specific and consistent between sampling periods. Bairlein (1983) and Moore et al. (1990) also reported high consistency of species-specific habitat distribu- tions and that migrants show selective use of certain habitats. There was no consistent relation- ship to winter habitat but there was to breeding habitat. Most migrant species that breed in scrub-shrub and open habitat were more numerous on Star Island, except for Chestnut-sided and Mourning warblers. The shorter vegetation on Star Island TABLE 1. Bird species distribution between Star Island (SI) and Appledore Island (AP) based on banding data from 1999 and 2000. H/g = Chi-square heterogeneity and goodness-ot-tit analyses; H = species responsible for the difference between islands. A — indicates no difference between the islands and blank = capture numbers insufficient for analysis in that sampling period. Island = location where relatively more of that species were captured. 730 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 4, Decenib^f 2010 00 c/5 O CQ W) I I I * * I W) I X T o (N — rs CM ^ 1 0. c/5 1 < * * * 1 * OX) 1 OX) X T O CN cu < Dh c/0 <; * * I W) OX) I c/5 * * Ot) I I ^ T ^ — ro IT) — m — — — I + + * I I I OX) OX) OX) I I I T ^ T T T T * * * * OX) * I OX) OX) 1 ^ T ^ T I I I I I I sO O ^ >/5 O CN — (N (N CN mONcncNicN-^'— — (N OS‘/5’OOOsD(N<7N'OOsCNU^i'*:^'aNfN'0^00 — — — ^ — 1 Oh j 1 Oh cu Oh 0- 1 1 , 1 c/5 < c/5 < < < < c/5 + * * + * * * * * * * * * * * * -S? OX) OX) OX) OX) OX) OX) OX) 1 OX) 1 T T K X ^ ^ ^ ^ T T X cn o r-- m r-- m 00 iTj oo 00 r- — X lo (N — (N ON o r^. (N rn fO (N (N 00 r-- __ t/5 (N r^, On C^i >/5 r- O ro Y* rs» 00 o to * * * * * * 1 * * * OX) OX) OX) OX) OX) 'S? X 1 T 1 T T (0^ Y- m r- oo tT) X o O 0 ^ 0- -H C/5 < C/5 * * OX) I I * + * OX) OX) 1^ ^ 7 T * 00 I I 4- * Ot) 00 I I I 7 T T D- < ■x- * 00 , 0. Oh a. >-H CL, <<<£/)< * * ** 6XJ OXJ M OJj au ^ ^ X I ^ T T o o o 00 X O o in o O 00 ro O' in •Y - H o o V (L) -C o -C OX) aj c TTr ■s ^ s (D E p > > "O C3 3 Z CL ' “O (U “O * c 5 = S ^ o 5 JZ H -i- ^ o ^ ra 2 p -C "2 p (j cj OJ ^ Si - C G ^ prOTD-O o-TT-C^ L) P Ucd-— j^ggl-Odg^ _iujQacLCc:DiQaOoi>oo 2^ s:i c3 3 03 ^ G C u. E 3 ^ S ^ JZ "O c/5 ^ G p ca o .x: u z z u _G JZ u. 3 G ^ QQ _G -D u. C3 C J=; 3 (U ^ 3 _ O C ^ OX) G C3 ^ S 5 _G X) ^ p -P ^ o "p -o CL G E ^ 3 V X _ ^ Y X o ^ — G G G _E J2 > CO CQ _G X 3 ^ c 3 2 o E 3 G X B c3 X — "O a 3 G G 3 3 t: 3 c/i *o > G j- G u C X ^ c3 "£ 5> ^ > > G 3 OX) c CO E 5 -a 3 t/) E O ■" § CQCD cs on > ^ O s t X) C3 O- P O > ■ ^ o ^ § 00 g_ t/> c/5 !s Q- o e O . ■- & : J OO ; O I "S o o -g (U E £i S ti. •g E u “ B- O C3 3 Qi PO Q- with its greater density at lower heights and searcity of taller trees may be more suitable for scrub-shrub species than Appledore Island. Bair- lein (1983) also reported differences in species along a gradient of vegetation density, and Rodewald and Brittingham (2004) reported shrub-sapling breeding species were most abun- dant in early successional and edge habitat during fall stopover. Ten of the 1 1 migrant species captured significantly more frequently on Appledore Island breed in forested habitat with the exception being Mourning Warbler. However, not all forest- breeding species were more numerous on Apple- dore Island. Nine of the 10 forest-breeding species that were more numerous on Appledore Island are area-sensitive in breeding areas. No area-sensitive species was more numerous on Star Island. Other studies have reported evidence of area-sensitivity during migration (Martin 1980, Cox 1988, Somershoe and Chandler 2004). The influence of patch size on stopover habitat use is also influenced by microhabitat and landscape (Petit 2000). The influence of habitat area in our study was not reliably separated from habitat structure. For example, area-sensitive species may be less flexible in migration stopover habitat than other forest breeding species and, thus, more common on Appledore Island because of its more forest- like habitat. Microhabitat factors likely contributed to the distribution patterns of forest-breeding species. American Redstart breed in habitat with a dense sapling (2. 5-9. 9 cm dbh) understory (DeGraaf and Yamasaki 2001). Appledore Island had far more trees in this size class than Star Island. Appledore Island also had more and larger trees than Star Island. Black-and-white Warblers were recorded only on large limbs of larger trees during spring migration in Kentucky (Mason 1979) and were more common on Appledore Island. Brown Creeper was the only forest-breeding species more common on Appledore Island that is not area-sensitive. It forages primarily on large- diameter trees with deeply furrowed bark (Hejl et al. 2002). This feeding substrate is almost totally lacking on Star Island but is present on Appledore Island. Fruit availability also seemed to affect the distribution of species. Three of the six species more numerous on Star Island that were not scrub- shrub or open habitat breeders appeared to be influenced by presence of fruit-bearing shrubs. 732 THE WILSON JOURNAL OF ORNITHOLOGY • Vul. 122, No. 4, December 2010 TABLE 2. Tree and shrub species composition on Star and Appledore islands sampling plots (0.04 ha) in 2000. Differences in means based on Mann-Whitney CZ-tests for between islands. A — indicates insufficient numbers for analysis (20 or fewer stems). Species with <10 stems (total trees and shrubs combined) on both islands were not analyzed*^. Common name Scientific name Star Island Shrubs Trees Appledore Island Shrubs Trees Difference shrubs trees Mean per plot Mean per plot Mean per plot Mean per plot Apple Pyrus malus 0.6 4.0 Arrowwood Viburnum recognitum — — 56.0 23.6 + + Northern bayberry Myrica pensyl van ica 35.3 12.0 — — + + Blackberry Rubus allegheniensis 12.3 — 9.4 — — Black cherry Primus serotina — — — 10.0 * Chokeberry” Pyrus spp.” 430.7 — 0.8 — Chokecherry Primus virginiana 3.3 9.0 87.8 55.0 * * Huckleberry Gaylussacia baccata 11.3 — — — — Pin cherry Pyrus pensylvanica — — 2.8 11.8 — Poison ivy Toxicodend ran radicans 19.0 1.0 — 0.4 -h Raspberry Rubus ideaus 296.0 — 30.2 — — Ro.se'’ Rosa spp.'’ 158.0 — 13.4 — * Serviceberry Amelanchier spp. 1.0 6.7 2.0 46.6 * Staghorn sumac Rhus typhina 1.0 19.3 5.4 9.6 — - Winterberry Ilex verticillata 76.3 68.3 57.2 36.2 - - + = 0.05 < P < 0.10. * P < 0.05, ** P < 0.01, and — = differences not significant. “ Red and purple chokeberry (P. arbutifolia. P. florihimda) and possible hybrids. Virginia rose {R. virginiana) on both islands, pa.sture rose [R. Carolina) only on Appledore Island. *■ Species with fewer than 10 stems: carrion-flower (Smilax herbacea). common elder iSambiiciis canadcn.sis). bittersweet nightshade iSalaniim dulcamara), red cedar Uuniperus virginiana). red maple (Acer rubrum), Virginia creeper (Parthenoci.'isus quinquefolia). The Yellow-rumped Warbler feeds on bayberry during fall migration (Parrish 1997); this shiub was moderately common on Star Island but absent from the Appledore Island plots. The other two species, Cedar Waxwing and Purple Finch, were two of the four highly frugivorous species present in the fall (Table 4). Additionally, the Yellow- breasted Chat was both highly frugivorous and a scrub-shrub breeder, and was also more numerous on Star Island. Both vegetation structure and species compo- sition can affect habitat suitability and use (Rice et al. 1984, Moore et al. 1995, Finch and Wang 2000). There were differences in plant species and TABLE 3. Differences in vegetation physiognomy between Star Island and Appledore Island based on a Mann- Whitney U-test of means in each category; df = I for all tests. There were three plots (0.04 ha) on Star Island and five plots on Appledore Island. Data collected in 2000. Category Star Lsland Appledore Island u p # stems Mean/plot ± SD # stem.s Mean/plot ± SD Live trees Density 358 1 19.3 ± 95.30 986 197.2 ± 36.69 lO.OO 0.456 Height' (m) 3. 1 ± 0.48 4.3 ± 0.44 15.00 0.025 Dbh” (cm) 4.4 ± 0.44 5.4 ± 0.71 14.00 0.053 Dead trees 13 4.3 ± 2.08 120 24.0 ± 14.05 15.00 0.025 Live shrubs Density 3,139 1,046.3 ± 80.71 1,341 268.2 ± 132.14 O.OO 0.025 Height' (in) 1.3 ± 0.27 1.5 ± 0.23 1 1 .00 0.297 Dead shrubs 919 306.3 ± 231.65 437 87.4 ± 65.89 2.00 0.101 “ Mean per plot for each island was calculated from means of each individual plot on that island. Suonia/a e/ al. • MIGRATION STOPOVER DISTRIBUTION AND HABITAT USE 733 (A) o Q. 6.0 (B) 900.0 800.0 o 700.0 - Q. (1) 600.0 - CL CO n 500.0 - 3 CO 400.0 - c (ti (D 300,0 - 200.0 - 100.0 - 0.0 - 0.5-0.9 1.0-1. 4 15-1,9 2.0-2. 4 Height class (m) I 1 1 T r LI dii_ ■ Star □ Appledore >2.5 FIG. I , Distribution of (A) trees and (B) shrubs by height class on Star Island. New Hampshire, and Appledore Island, Maine, in 2000. Bars are means ± SD; * = P < 0.05, and + = 0.05 < P < 0. 10. vegetation .structure between islands; the relative importance of these two factors is difficult to assess (Rice et al. 1984). Such was the case with Yellow- rumped Warbler, which feeds on bayberry during fall migration (Parrish 1997) and was more common on Star Island, the only island where bayben-y occuiTed. However, the effect of habitat structure cannot be distinguished from the presence of bayberry (Kwit et al. 2004). The primary factor is likely structural differences for most species because migrants must adapt to a wide variety of plant species along their migration route, and species differences between islands were primarily based on breeding habitat staicture. Seasonal shifts in habitat have been ob.served at other sites (Winker et al. 1992, Weisbrod et al. 1993, Wang et al. 1998), but only Blue-headed Vireo in our study had a seasonal switch between islands. Thus, most species in this study were captured more frequently on the same island 734 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122. No. 4. December 2010 TABLE 4. Ecological classifications for each avian species. SI = Star Island, AP = Appledore Island, - = no ditterence between islands, and blank = species not present in sufficient numbers to analyze. Area sensitivity without parentheses is from Robbins et al. (1989); in parentheses is from individual species accounts in The Birds of North America. Species More captures spr/fall Area sensitivity (breeding) General category* Breeding habitat Veg Special type'’ characteristics Frugivory (fall)* Northern Flicker /- No F-FE M Eastern Wood-Pewee No F-FE D-M Yellow-bellied Flycatcher -/- (No) F-SF C Forests and bogs Medium Traill’s Flycatcher SI/SI (No) SS-FE D Med-hi Least Flycatcher -/- (Yes?) ES-F-FE D-M Medium Eastern Phoebe /- (No) FE D-M Med-hi Blue-headed Vireo AP/Sl No F C-M Medium Philadelphia Vireo /- (No) F-FE-ES D-C-M Med-hi Red-eyed Vireo AP/AP Yes F-FE D-M Med-hi Red-breasted Nuthatch -/- (No) F C-M Mature forests Brown Creeper ZAP No F C-D-M Mature forests Low Golden-crowned Kinglet /- (No) F C Low Ruby-crowned Kinglet Sl/Sl (No) F C-M Low Veery -ZAP Yes F D Dense understory Med-hi Swain.son’s Thrush APZ- (Yes?) F-FE-ES C-M Med-hi Cedar Waxwing SI/Sl No FE D FruitZberry producers High Nashville Warbler -Z- (No) ES-F-FE C-D-M Medium Northern Parula -Z- Yes F-FE C-D-M Mature forests Medium Chestnut-sided Warbler -z- No SS-ES-FE D Dense low growth Magnolia Warbler SIZSI No F-ES C-M Dense understory Med-hi Black-throated Blue Warbler APZAP Yes F D-M Dense saplingZshrub Medium understory Yellow-rumped Warbler SIZSI (No) F C-M Med-hi Black-throated Green Warbler -Z- No F C-M Mature forests Medium Blackburnian Warbler -Z (No) F C-M Mature forests Blackpoll Warbler -z- (No) F-SF C Low northern spruce Medium Black-and-white Warbler APZAP Yes F D-M Forests with saplings Low American Redstart -ZAP No (Yes?) ES-F-FE D Thick sapling understory Low Ovenbird APZAP Yes E D-M Mature forests Medium Northern Waterthrush APZAP Yes F-FE D WetlandZpond edges Med-hi Mourning Warbler APZ- (No) SS-ES D-M Dense saplingsZshrubs Med-hi Wilson's Warbler SIZSI (No) SS-SF D Low Canada Warbler APZ- Yes F D-M Den.se understory Low Yellow-brea.sted Chat ZSI No SS-FE D No overstory High Savannah Sparrow SIZ (No) 0 G Lincoln’s Sparrow -ZSI (No) SS-EE-SF D-M Med-hi Swamp Sparrow SIZSI (No) SS D-G Wetlands Medium White-throated Sparrow -ZSI (No) FE-ES-SF D-C-M Brushy understory Med-hi Ro.se-breasted Grosbeak -Z Yes FE-F D High Baltimore Oriole -Z- No FE D Purple Finch SIZSI (No) F-FE C-M High “ Habitat categories: ES = early succe.ssional. F = forest. FE = forest edge, SF = short forests at high elevation and in bogs. SS = scrub/shrub, and O = open areas, grasslands, emergent wetlands. Vegetation type (veg type), C = coniferous. D = deciduous, G = grassland/sedge/reed, and M = mixed coniferous-deciduous. ' Frugivory ba.sed on mean percentage of fruit in fecal samples (Parrish 1997): low 0-29. medium .70-.S9. medium-high (med-hi) 60-89. high 90-100%. and blank = not in analysis. during both .spring and fall. The consi.stency of habitat .selection by a wide range of migrant species between spring and fall supports the hypothesis that habitat .selection in passerines is endogenous (Bairlein 1983, Morton 1990, Winker 1995). Migrant distribution might be expected to change between spring and fall, after the large proportion of young migrants have had more exposure to a variety of habitats on both their southward and return migrations, if habitat Siiomala cl al. • MIGRATION STOPOVER DISTRIBUTION AND HABITAT USE 735 selection was learned. Additionally, the correla- tion ot stopover habitat with breeding habitat indicates innate preferences as suggested by Moore et al. { 1995). The apparent association with breeding habitat by migrants during stopover is consistent with Petit’s (2000) analysis of habitat selection in five studies. Smaller physiognomic and plant species differences between islands were associated with differences in species distribution between is- lands. Our study suggests stopover habitat use was related to habitat structure, plant species, diet (extent of frugivory), and habitat area, but the predominant factor(s) may differ among species. Habitat differences between these two islands were smaller in structure and species type than described in other stopover habitat studies (e.g., Bairlein 1983, Moore et al. 1990, Rodewald and Brittingham 2004); but differences between is- lands were associated with differences in bird species distribution. Nocturnal migrants typically land when it is still dark and are unlikely to make stopover decisions based on these small habitat differences. Migrants may conduct habitat explo- ration flights during the day (Moore and Simons 1992, Wiedner et al. 1992), but few birds banded on one island were recaptured on the other island (Suomala 2005) indicating little movement after initial landing. We suggest .several mechanisms to explain the detection of small habitat differences: (1) migrants may be landing at the Isles of Shoals later in the morning than migrants at other locations and are able to see detailed habitat features; (2) migrants may be able to see certain features in the dark; and/or (3) there is a non-visual component of stopover habitat selection, similar to the vocal social cuing that Betts et al. (2008) reported was more important than vegetation structure in habitat choice for breeding site .selection. CONSERVATION IMPLICATIONS Priority stopover habitat for the bird communi- ty may be difficult to define due to differing habitat preferences of individual species (Hutto 2000, this study). Protection of stopover sites with a wide variety of habitats, both forest and scrub- shrub, is important to meet the needs of migrant passerines (Weisbrod et al. 1993, Rodewald and Brittingham 2004, this study). The specie.s-specific differences in habitat use in this study suggest migrants are not using the Isles of Shoals as simply emergency stops, landing at random out of desperation. Scrub-shrub habitat is naturally occurring on the coastal plain and a good source of berry-producing shrubs, which may be one of the best areas for migrants to deposit fat (Parrish 1997). Other studies have demonstrated the importance of edge and shrub- land habitats with fruit to fall migrants (Parrish 1997; Suthers et al. 2000; Rodewald and Britting- ham 2002, 2004), but the management of scrub- shrub habitat for migrants may be overlooked as an important priority. Land use changes along the coast are occurring rapidly with resulting deterioration of stopover habitat (Mabey and Watts 2000). Coastal devel- opment creates a landscape that is often inhospi- table to migrating birds. Conservation efforts for migrant songbirds should encompass the remain- ing small patches of natural habitat that may be critical to migrant survival (Blake 1986). Educa- tion and local land use policies may be effective tools for maintaining small habitat patches on private lands (Mabey and Watts 2000). ACKNOWLEDGMENTS This project was made possible by support from the Star Island Corporation and New Hampshire Audubon, grants from the Eastern Bird Banding Association and the University of New Hampshire Graduate School, and the generous donations of many individuals. This effort would not have been possible without help of volunteers at New Hampshire Audubon, the Star Island banding station, and the Appledore Island Migration Banding Station, supported in part by the Shoals Marine Laboratory. This paper is contribution # 14 of the Appledore Island Migration Banding Station and contribution # 153 of the Shoals Marine Laboratory. LITERATURE CITED Baird, J. and I. C. T. Nisbet. 1960. Northward fall migration on the Atlantic Coast and its relation to offshore drift. Auk 77:1 19-149. Bairlein, F. 1983. Habitat .selection and a.ssociation of species in European pas.serine birds during southward, post- breeding migrations. Omis Scandinavica 14:239-245. Bairlein, F. 1992. Morphology-habitat relationships in migrating songbirds. Pages 356-369 in Ecology and con.servation of neotropical migrant landbirds (J. M. Hagan and D. W. Johnston. Editors). Smithsonian Institution Press, Washington, D.C., USA. BETT.S, M. G., A. S. Hadley, N. Rodenhouse, and J. J. Nocera. 2008. Social information trumps vegetation structure in breeding-site selection by a migrant songbird. Proceedings of the Royal Society of London. Series B 275:2257-2263. Blake, J. G. 1986. Species-area relationship of migrants in i.solated woodlots in east-central Illinois. Wilson Bulletin 98:291-296. 736 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4. December 2010 Blake, J. G. and J. R. Karr. 1987. Breeding birds of isolated woodlots: area and habitat relationships. Ecology 68:1724-1734. Cox, J. 1988. The influence of forest size on transient and resident bird species occupying maritime hammocks of northeastern Florida. Florida Field Naturalist 16:25-34. DeGraaf, R. M. and M. Yamasaki. 2001. New England wildlife: habitat, natural history, and distribution. University Press of New England, Hanover, New Hampshire, USA. Dunn, E. H. 2000. Temporal and spatial patterns in daily mass gain of Magnolia Warblers during migratory stopovers. Auk 117:12-21. Ehrlich, P. R., D. S. Dobkin, and D. Wheye. 1988. The birders handbook: a field guide to the natural history of North American birds. Simon and Schuster Inc., New York, USA. Faaborg, j., M. Brittingham, T. Donovan, and J. Blake. 1995. Habitat fragmentation in the temperate zone. Pages 357-380 in Ecology and management of neotropical migratory birds: a synthesis and review of critical issues (T. E. Martin and D. M. Finch. Editors). Oxford University Press, New York, USA. Finch, D. M. and Y. Wang. 2000. Landbird migration in riparian habitats of the middle Rio Grande: a case study. Studies in Avian Biology 20:88-98. Freemark, K. E. and H. G. Merriam. 1986. Importance of area and habitat heterogeneity to bird assemblages in temperate forest fragments. Biological Conservation 36:1 15-141. Gotelli, N. j. and a. M. Ellison. 2004. A primer of ecological statistics. Sinauer Associates Inc., Sunder- land, Massachusetts, USA. Hejl, S. j., K. R. Newlon. M. E. McFadzen, J. S. Young, AND C. K. Ghalambor. 2002. Brown Creeper (Certhia ainericana). The birds of North America. Number 669. Hutto, R. L. 1985. Seasonal changes in the habitat distribution of transient insectivorous birds in south- eastern Arizona: competition meditated? Auk 102: 120-132. Hutto, R. L. 1998. On the importance of stopover sites to migrating birds. Auk 115:823-825. Hutto, R. L. 2000. On the importance of en route periods to the conservation of migratory landbirds. Studies in Avian Biology 20:109-1 14. James, F. C. and H, H. Shugart Jr. 1970. A quantitative method of habitat description. Audubon Field Notes 24:727-736. Jones, J., C. M. Francis, M. Drew, S. Fuller, and M. W. S. Ng. 2002. Age-related differences in body ma.ss and rales of mass gain of passerines during autumn migratory stopover. Condor 104:49-58. Kwn, C., D. J. Levey, C. H. Greenberg, S. F. Pearson, J. P. McCarty, S. Sargent, and R. L. Mumme. 2004. Fruit abundance and local distribution of winter Hermit Thrushes (Cutharus I’ullatu.s) and Yellow- rurnpcd Warblers (Dendroiea eoronata) in South Carolina. Auk 121:46-57. Mabey, S. E. and B. D. Watts. 2000. Conservation of landbird migrants: addressing local policy. Studies in Avian Biology 20:99-108. Martin, T. E. 1980. Diversity and abundance of spring migratory birds using habitat islands on the Great Plains. Condor 82:430-439. Mason, W. 1979. Habitat selection by the Parulidae during spring migration along the South Fork Creek in Glasgow, KY. The Kentucky Warbler 55:39^2. Moore, F. R. and T. R. Simons. 1992. Habitat suitability and stopover ecology of neotropical landbird migrants. Pages 345-355 in Ecology and conservation of neotropical migrant landbirds (J. M. Hagan and D. W. Johnston, Editors). Smithsonian Institution Press, Washington, D.C., USA. Moore, F. R., P. Kerlinger, and T. R, Simons. 1990. Stopover on a gulf coast barrier island by spring trans- gulf migrants. Wilson Bulletin 102:487-500. Moore, F. R., S. A. Gauthreaux Jr., P. Kerlinger, and T. R. Simons. 1992. Stopover habitat: management implications and guidelines. Pages 58-69 in Status and management of neotropical migratory birds (D. M. Finch and P. W. Stangel, Editors). USDA, Forest Service, General Technical Report RM-229. Rocky Mountain Station, Fort Collins, Colorado, USA. Moore, F. R., S. A. Gauthreaux Jr., P. Kerlinger, and T. R. Simons. 1995. Habitat requirements during migration: important link in conservation. Pages 121- 144 in Ecology and management of neotropical migratoi7 birds: a synthesis and review of critical issues (T. E. Martin and D. M. Finch. Editors). Oxford University Press, New York, USA. Moran, M. D. 2003. Arguments for rejecting the sequential Bonferi'oni in ecological studies. Oikos 100:2. Morris, S. R., M. E. Richmond, and D. W. Holmes. 1994. Patterns of stopover by warblers during spring and fall migration on Appledore Island, Maine. Wilson Bulle- tin 106:713-718. Morris, S. R., D. W. Holmes, and M. E. Richmond. 1996. A ten-year study of the stopover patterns of migratory passerines during fall migration on Appledore Island. Maine. Condor 98:395^09. Morton, E. S. 1990. Habitat segregation by sex in the Hooded Warbler: experiments on proximate causation and discussion of its evolution. American Naturalist 135:319-333. Nudds, T. D. 1977. Quantifying the vegetative structure of wildlife cover. Wildlife Society Bulletin 5:1 13-117. Parneix, j. F. 1969. Habitat relations of the Parulidae during spring migration. Auk 86:505-521. Parrish, J. D. 1997. Patterns of frugivory and energetic condition in Nearctic landbirds during autumn migra- tion. Condor 99:681-697. Parrish, J. D. 2000. Behavioral, energetic, and conserva- tion implications of foraging plasticity during migra- tion. Studies in Avian Biology 20:53-70. Pinrr, D. R. 2000. Habitat use by landbirds along Nearctic- neotropical migration routes: implications for con.ser- vation of stopover habitat. Studies in Avian Biology 20:1.5-33. Poole, A. and F. Gii.i, (Editors). 1993-1997. The birds of North America. The Academy of Natural Sciences, Philadelphia, Pennsylvania and The American Orni- thologists' Union, Washington, D.C., USA. Siioniala el al. • MIGRATION STOPOVER DISTRIBU TION AND HABITAT USE 737 Poole, A. and F. Gill (Editors). 1998-2002. The birds of North America. The Birds of North America Inc., Philadelphia, Pennsylvania, USA. Rice, J., B. W. Anderson, and R. D. Oiimart. 1984. Comparison of the importance of different habitat attributes to avian community organization. Journal of Wildlife Management 48:895-91 1. Robbins, C. S., D, K. D.awson, and B. A. Dowell. 1989. Habitat area requirements of breeding forest birds of the Middle Atlantic States. Wildlife Monographs 103. Rodewald, P. G. and M. C. Brittingham. 2002. Habitat use and behavior of mixed species landbird flocks during fall migration. Wilson Bulletin 114:87-98. Rodewald, P. G. and M. C. Brittingham. 2004. Stopover habitats of landbirds during fall: use of edge- dominated and early-successional forests. Auk 121: 1040-1055. Sherry, T. W. and R. T. Holmes. 1995. Summer versus winter limitation of populations: what are the issues and what is the evidence. Pages 85-120 in Ecology and management of neotropical migratoi'y birds: a synthesis and review of critical issues (T. E. Martin and D. M. Finch. Editors). Oxford University Press, New York, USA. SiLLETT, T. S. and R. T. Holmes. 2002. Variation in survivorship of a migratory songbird throughout its annual cycle. Journal of Animal Ecology 71:296-308. SOMERSHOE. S. C. AND C. R. CHANDLER. 2004. Use of oak hammocks by neotropical migrant songbirds: the role of area and habitat. Wilson Bulletin 1 16:56-53. Sperduto, D. D. AND W. F. Nichols. 2004. Natural communities of New Hampshire. New Hampshire Natural Heritage Bureau and The Nature Con.servancy, Concord, New Hampshire, USA. SPSS Institute Inc. 2000. Sysiat Version 10.0. SPSS Institute Inc.. Chicago, Illinois, USA. SuOMALA, R. W. 2005. Migrant songbird species distribu- tion, behavior, and habitat use during stopover: a comparison of two islands in the Gulf of Maine. Thesis. University of New Hampshire, Durham, USA. SUTHERS, H. B., J. M. Bickal, AND P. G. Rodewald. 2000. Use of successional habitat and fruit resources by songbirds during autumn migration in central New Jersey. Wilson Bulletin 1 12:249-260. Wang, Y., D, M. Finch, F. R, Moore, and J. F. Kelly. 1998. Stopover ecology and habitat use of migratory Wilson's Warblers. Auk 1 15:829-842. Weisbrod, A. R., C. J. Burnett, J. G. Turner, and D. W. Warner. 1993. Migratory birds at a stopover site in the Saint Croix River Valley. Wilson Bulletin 105:265-284. Wiedner, D. S., P. Kerlinger, D. A. Sibley, P. Holt, J. Hough, and R. Crossley. 1992. Visible morning flight of neotropical landbird migrants at Cape May. New Jersey. Auk 109:500-510. Winker, K. 1995. Habitat selection in woodland Nearctic- neotropic migrants on the Isthmus of Tehuantepec. I. Autumn migration. Wilson Bulletin 107:26-39. Winker, K., D. W. Warner, and A. R. Weisbrod. 1992. Migration of woodland birds at a fragmented inland stopover site. Wilson Bulletin 104:580-598. Zar, j. H. 1999. Biostatistical analysis. Fourth Edition. Prentice-Hall Inc., Upper Saddle River, New Jersey, USA. The Wilson Journal of Ornithology 122(4):738-743, 2010 CONCENTRATED MIGRATORY MORNING ELIGHT AT LAKE PONTCHARTRAIN, LOUISIANA, USA PETER H. YAUKEY' ABSTRACT. — I describe large volume morning flights of migrant songbirds in autumn along the south shore of Lake Pontchartrain, Louisiana, USA. These flights have at times exceeded 10,000 birds/hr since their discovery in 2002, ranking the site among the continent’s highest volume regular migratory songbird concentration locations. Migrants move in "reverse” direction northeastward up a peninsula in Lake Pontchartrain, and fly across a 9-km water gap to reach the north shore, usually struggling into a headwind. Flights are dominated by Yellow and Yellow-rumped warblers (Dendroica petechia and D. coronata). Eastern Kingbird {Tyrannus tyranniis). Indigo Bunting {Passerina cyanea), Cedar Waxwing (Bomhycilla cedronim), and American Robin (Tiirdus migratorius) during their respective migratory peaks. These species have exceeded 800 birds/hr, and 70 other small landbird species have been recorded. Flights of >100 birds/hr were more frequent after nights with more hours of northerly winds. Received 24 August 2009. Accepted 30 April 2010. The Nearctic-neotropical migratory system involves the movement of billions of birds southward within the Western Hemisphere, some staying within temperate latitudes and others traveling to the Neotropics. The major barrier to the large contingent of migrants headed from eastern North America to tropical America is the Gulf of Mexico, which many species cross in a single extended flight. The dynamics of this trans- Gulf movement have been most extensively studied in spring. Northbound migrants normally arrive on the central Gulf Coast during daylight, often after having flown >6 hrs longer than in a normal overland nocturnal flight. Arriving trans- Gulf migrants normally over-fly coastal regions and disperse into habitats >100 km inland, causing near-coastal areas to be labeled the “coastal hiatus” (Lowery 1945). Their reasons for passing far inland before alighting are unknown, but the paucity of forest cover in the wetland landscape adjacent to the coast is a likely factor. Fall movements are even more poorly understood. Able (1972) noted that weather conducive to forward migration is more limited in fall than spring when prevailing southeasterly winds impede crossing the Gulf except when broken temporarily by the passage of cold fronts with their accompanying northerly tailwinds. The objectives of this paper are to: (1) report the existence of an unusual migratory movement phenomenon at Lake Pontchartrain, Louisiana, and (2) discuss the insights it provides into the dynamics of autumnal migration on the Gulf Coast. The volume of bird movement discovered to occur here ranks this newly discovered site as ' Department of Geography, University of New Orleans, New Orleans, LA, 70148, USA; e-mail: pyaukey@uno.edii one of the major migratory songbird concentration locations in the Western Hemisphere. METHODS Study Site. — Lake Pontchartrain is ~100 km north of the Gulf of Mexico, from which it is separated by urban New Orleans and expansive wetlands (Fig. 1). The city extends 10-20 km southward from the central portion of the lake, while wetlands border the lake’s southwestern and southeastern flanks, the latter on the Bayou Sauvage National Wildlife Refuge. The landscape outside the southern huiricane-protection levee of the city changes abruptly to bottomland hardwood and swamp forests, which transition southward through shrub-dominated wetlands into open fresh and brackish marshes, and finally to salt marsh that is flanked in some places by bamer islands and headlands. Several large semi-tidal lakes and inlets penetrate nearly as far northward as the city. The corridor of the Mississippi River winds through the city before heading southeast to the Gulf, its <3 km fringe of flanking field and woodlots forming the largest contiguous area of dry land south of the city. Lake Pontchartrain is roughly oval in shape, 39 km north-south and 65 km east-west at its widest, but is pinched by the extension of a peninsula northward near its eastern end that reduces the ovewater distance to 9 km. This peninsula ends at two adjacent points of land. South Point and Pointe aux Herbes (Fig. 1C, D), from which railroad and highway bridges depail for the north shore. Migratory songbird movements are greatest on this peninsula, especially progressing northward up its western Hank and departing from either of these two points to cross the lake. 738 Ycmkey • MORNING FLIGHT OF MIGRANTS 739 FIG. 1. Study area showing locations where major morning movements have been recorded: (A) University of New Orleans; (B) Bayou Sauvage National Wildlife Refuge bicycle path; (C) South Point; (D) Pointe aux Herbes. Map by Steve Stevens and John Adams. Northward curvature of the Gulf shoreline wraps open wetland and coastal lakes north to the east of Lake Pontchartrain, and essentially no forest cover lies east or southeast of its eastern end. Pine {Pinus spp.) forest is absent south of Lake Pontchartrain, but is common on its north shore. Field Procedures. — Preliminary scouting of morning flight was conducted by visiting several coastlines and land-marsh interfaces near Lake Pontchartrain in fall 2002. Systematic collection of data began in fall 2003, when volunteer observers were stationed at Pointe aux Herbes (Fig. ID) to document movement of birds across the water gap and find correlations with weather patterns recorded at Lakefront Airport (Fig. 1 between sites A and B). Additional visits were made to sites C and D through 2009, mostly on days with northerly winds with the frequency of visitation declining after Hurricane Katrina in 2005. Observers counted birds passing their observation point in 15-min intervals starting near dawn, and typically lasting 1-3 hrs. I used surface wind data beeause of their hourly resolution, despite their differing at times from winds aloft at the altitude of nocturnal migration. The number of nocturnal hours of northerly (between WNW and ENE) winds was tabulated preceding each flight count because, in the case of the cold front passages that usually preceded the diurnal movements noted, the number of hours of post-frontal conditions before daybreak may affect the opportunity for nocturnal migrants to be wind-drifted south of the lake. Bird counts conducted over >30 min were used to calculate hourly passage rates. RESULTS The presence of significant diurnal song bird movements along the Lake Pontchartrain penin- sula was not discovered until 22 November 2002, although local observers had infrequently visited this peninsula in previous years and occasionally noted migratory birds moving in anomalous directions. A continuous heavy movement of 740 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4. December 2010 TABLE 1. Peak rates of hourly and daily passage for selected species at Lake Pontchartrain, Louisiana, USA. Species Peak hourly passage (date) Peak daily passage'' (date) Eastern Kingbird American Robin Cedar Waxwing Yellow Warbler Yellow-rumped Warbler Indigo Bunting 25.000 (5 Sep 2003) 12.000 (28 Nov 2004) 3.500 (22 Nov 2002) 800*' (31 Aug 2004) 4.500 (22 Nov 2002) 880 (15 Oct 2002) 35.000 (5 Sep 2003) 20.000 (28 Nov 2004) 7.000 (22 Nov 2002) 1.200 (23 Aug 2009) 12,500 (23 Nov 2002) 1.000 (15 Oct 2002) ■' All counts were obtained during <3 hrs of counting. E.xtrapolated from 41 min observation. songbirds northeastward along the western flank of the peninsula (Fig. IB) was witnessed on the morning of 22 November 2002 following a cold front passage. Frequent timed counts over 2 hrs produced an estimated total of 25,000 birds dominated by American Robins (Tiirdus migra- torins). Cedar Waxwings (Bombycilla cedrorum), and Yellow-rumped Warblers (Dendroica coro- nata). These birds were flying low into the NNE wind, using a narrow tree row that separated the lake levee from the landward marshes as a windbreak. The next morning another 18,500 birds were estimated to pass up the shore in 3 hrs at this site, dominated by the same species. Over the next few weeks, as the fall migration dwindled, similar but smaller bird movements were tracked up the peninsula and found to be crossing the lake at South Point and Pointe aux Herbes (Fig. 1C, D). The lakeshore along this western flank of the peninsula is not readily accessible, being behind gated roads within the Bayou Sauvage National Wildlife Refuge. A few morning songbird flights detected earlier in the same fall at a site 0.5 km south of the lakefront and near the western base of the peninsula .seemed in retrospect to have represented similar patterns of movement, but for neotropical migrant species (Fig. lA). On 5 September 2002, —1,400 Ea.stern Kingbirds (Tyninnus tyrannus) were count- ed in 1 hr flying ENE parallel to the lakefront, and on 15 October, —1,350 small songbirds, predom- inantly Indigo Buntings (Passerina cyanea), had been seen passing northward toward the nearby lakeshore in 68 min; their behavior upon reaching the shore was not identified. Flights exceeding 1 00 birds/hr were observed on 37 days each fall during 2()02-2()()4, and occurred as early as 8 August and as late as 28 January. Sub.sequently, flights have been detected as early as 13 July 2008 (proceeding up the Missi.ssippi River south of the lake). The largest flights have occurred during peak passages of Yellow Warblers (Dendroica petechia) and Eastern Kingbirds in August-September, Indigo Buntings in October, and American Robins, Cedar Waxwings, and Yellow-rumped Warblers in November. Counts during northeast winds have repeatedly exceeded 1,000 birds/hr, and six individual species have peaks of >500/hr with robin and kingbird counts suipassing 10,000/hr (Table 1). Species identifica- tion of many of the less frequent species was difficult until 2004, when more visits were made to site C, which has the best viewing conditions. Seventy-two species of small migratory land birds, mostly nocturnal migrants, were identified as they crossed or attempted to cross the lake at sites C and D from 2003 to 2009. All significant movements recorded at the lakefront have been northeastward, except for seven oriented at a bearing west of north in departing across the water from the western point (South Point, site C). These directions result in a longer overwater crossing and have only occuiTed when morning winds were from west of north. Wind directions were analyzed for flights from 2002 through 2004, the most consistent period of field observation (Eig. 2). Elights peaking at >100 birds/hr occuired with differing frequency when different numbers of hours of northerly winds had occurred the night before (yy = 23.0, n = 108, P = 0.0017). They occurred on a majority of mornings following >9 hrs of northerly winds, but less than a tenth of those with <2 hrs. Northerly winds are unu.sual on the Gulf Coast except after frontal passages, which were infrequent in August but frequent each November during the study. Thus, the northerly winds conducive to mass movements at the lake are more common during late fall periods dominated by temperate zone migrants. Another source of Yaukcv • MORNING FLIGHT OF MIGRANTS 741 70 0-2 3-5 6-8 9-11 Nocturnal hours of north wind FIG. 2. Percentage of days with morning movements >100 birds/hr that followed nights with different hourly frequencies of northerly winds. northeast winds in early fall has been the occasional tropical cyclone that has entered the eastern Gulf of Mexico; Tropical Storm Henry produced northeast winds during the massive kingbird movement on 5 September 2003 (Table 1), while 550 km to the ESE near the Florida coast. DISCUSSION Why are there regular northward movements of migrant birds across Lake Pontchartrain on mornings after fall cold fronts? There are three potential explanations. (1) The movements repre- sent forward migration of birds attempting to head eastward, but detouring northward to avoid the open Gulf and marshes ahead to the east. (2) The movements represent migrants not attempting forward progress, but searching locally for suitable stopover habitat in which to rest or feed. (3) The birds are making a corrective movement to compensate for wind drift or navigation eiTors. The first hypothesis of detoured forward migration is appealingly simple, but contradicts the basic widespread recognition that circum-Gulf migration in Louisiana is westbound (e.g., Lowery 1974). This has been established by radar (Able 1972), visual studies of morning flight (Duncan and Weber 1985), and cage orientation studies (Sandberg and Moore 1996). Other weaknesses include: (1) birds on some dates move primarily NW, (2) it seems unlikely that large numbers of birds would migrate forward into headwinds (yet most heavy Lights are during NE winds), and (3) birds seldom approach the overwater crossing up the east side of the peninsula as would be expected if birds were detouring from the south as proposed. The second hypothesis, that birds are .seeking foraging and resting habitat is consistent with the patchiness of forest and scrub habitats south of the lake. The overwhelming dominance of north- bound birds at the peninsula could be a conse- quence of shoreline configuration — only north- bound birds are able to approach the peninsula overland, and the lakeshore concentrates birds approaching from that direction. At least one species. Tree Swallow (Tachycineta bicolor), appears to commute across the lake in its daily search for food. This winter resident regularly forms nocturnal roosts in late fall in sugar cane {Saccharum ojficinaruin) fields 80 km to the southwest that may include over a million birds (Seymour 2009). I attributed large Tree Swallow flights (e.g., 18,000 in 1.5 hrs on 6 November 2004) to roost dispersal and omitted them from this analysis, but it is difficult to imagine most other songbirds crossing the lake as commuters. The capacity of swallows for strong flight as they depart across the lake contrasts with the greater struggle of many other songbirds. Some species have so much difficulty that they appear to make repeated failed departures in short order; for instance 100 departures by Golden-crowned and Ruby-crowned kinglets (Regulus satrapa and R. calendula) in 7.5 min on 2 November 2005 appeared to involve many repeat attempts to cross the lake. It is difficult to imagine that searching behavior would have such a strong northeastward impulse (or into-wind impulse) that birds would be willing to cross 9 km of water into a headwind, expending energy and potentially exposing them- selves to accipiters and falcons that are common on the peninsula in autumn. This would also fail to explain why other resident songbirds do not engage in such flights (only Rock Pigeons \Coiumba livia\ do so regularly, and usually southbound), and why Lights of Yellow-rumped Warblers and American Robins dwindle precipi- tously in mid-winter despite remaining numerous in the region. Searching for food may contribute to cross-lake movements by some species, but does not appear to be a sufficient explanation for most. The third hypothesis, corrective movement, is favored in part because it has been cited as the explanation for coastal retreat movements alon^ the Atlantic shoreline of northeastern North 742 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 4, December 2010 America including at Cape May, New Jersey (Baird and Nisbet 1960) and in Sweden at Falsterbo, a site where diurnal reverse-directional movements are also massive in fall (Alerstam and Ulfstrand 1975, Alerstam 1978, Akesson et al. 1996). Swainson’s Thrushes {Cathanis ustulatus) on the northern Gulf Coast have also been observed to orient inland in flight cages in the early morning when they are in lean body condition in fall, as did lean Red-eyed Vireos (Vireo olivaceus) released at night (Sandberg and Moore 1996, Sandberg et al. 2002). That migrants would exhibit a directional response to remove themselves from the coastal 100 ± km of Louisiana is also consistent with the knowledge they appear to find these areas unsuitable during spring, normally flying over them when arriving from a spring Gulf crossing (Lowery 1945, Gauthreaux 1971). Fall migrants intending to winter in the region could also backtrack if they find the coastal wetlands unsuitable wintering habitat. The northward direction of both neotrop- ical and temperate migrants could be either a reflection of attempting to backtrack to an earlier position on the migratory route before habitat became inhospitable, or of attempting to compen- sate for wind drift by retreating into the wind, which is generally northerly after nights of heavy migration in fall. Other authors have reported changes in diurnal flight orientation to compen- sate for nocturnal wind drift (Baird and Nisbet 1960, Able 1977, Gauthreaux 1978, Moore 1990), and apparent reversal of migratory orientation over inhospitable areas (DeSante 1973, Thorup and Rabpl 2007). Migrants might retreat because wind drift has carried them farther south than intended on that date in their migration in addition to avoiding inhospitable habitat. Some large flights at South Point have involved species that did not arrive in numbers elsewhere in coastal Louisiana until weeks later (e.g., 7,000 Cedar Waxwings on 22 November 2002). The occurrence of large movements of Amer- ican Robins and Cedar Waxwings along the Gulf Coast is interesting because the Gulf Coast forms the southern edge of their wintering ranges at this longitude. The large (lights of Yellow-rumped Warblers, occurring mainly in late November, may also represent temperate winterers rather than trans-Gulf migrants, as the bulk of other trans- Gulf migrant species have departed the northern Gulf Coast by I November (Lowery 1974). The other major concentration sites of migrants in North America usually involve birds well north of their southern range limits, at latitudes traversed by a large portion of the population during their migration. Large flights near these species’ southern range limits are surprising, because relatively small proportions of their populations migrate this far south. The evidence suggests the movements observed along Lake Pontchartrain are predominantly birds retreating from coastal wetlands. If correct, this adds an important detail to the dynamics of migration near the Gulf Coast. It indicates the “coastal hiatus” phenomenon for neotropical migrants long documented for spring migration has a counterpart in fall. In the case of temperate zone migration, it indicates that overshoot of wintering areas by a short distance (—100 km) along the Gulf Coast is a common phenomenon. In either case, the species involved must incur a stressful upwind flight to terra firma that may take several hours and deplete energy reserves. It also implies, in the case of neotropical migrants, that habitats lying Just north of the coastal wetlands might be especially important as stop- over sites forming the last stops before the demanding trans-Gulf flight. The numbers of birds involved in the Lake Pontchartrain flights make it one of only a handful of places in North America where diurnal movements of migrant song birds (excluding swallows) are documented to regularly exceed 1,000 birds/hr. These include sites along the Atlantic shoreline from Long Island Sound to Delaware Bay (e.g.. Lighthouse Point, Connecti- cut and Cape May, New Jersey), on the shorelines of Lake Erie and Lake Ontario (e.g.. Holiday Beach, Ontario, and Braddock Bay, New York), at Tadoussac on the north shore of the St. Lawrence River in Quebec, and along the Gulf of Mexico shoreline near Veracruz, Mexico. It is unusual in its inclusion of a 9 km elective overwater flight (as opposed to return to land from the ocean), although southbound Blue-gray Gnatcatchers {Polioptila caerulea) regularly make an overwater crossing at Smith Point on the Texas coast (John Arvin, pers. comm.). The predominant species are those known to be at least partially diurnally migratory, as is the case elsewhere. Most have large regular diurnal movements at one or more of these other localities, except apparently the Indigo Bunting. Large flights are characteristically ac- companied by other species, including a wide variety of nocturnal migrants. However, many Yaukcv MORNING FLIGHT OF MIGRANTS 743 understory species apparently do not cross the lake as no attempts have been observed by catbirds, thrashers, Catharus and Hylocichla thrushes, and wrens; the only sparrows have been those ot the most open habitats (e.g.. Savannah Sparrow [Passerciilus scmdwichensis]). If flights represent northward coiTective move- ments, they might be less common along the eastern half of the northern Gulf Coast, which lacks the wide expanse of coastal wetlands that lie south ot Lake Pontchartrain. Broad wetlands border the coast for 500+ km west from the mouth of the Mississippi River throughout Louisiana and the upper coast of Texas, and these would seem good candidates for reverse flights. However, without a funneling geographic feature, these flights might not be concentrated as they are at Lake Pontchartrain. The 40-km shoreline of Lake Pontchartrain apparently turns an otherwise diffuse northward movement into a spectacle. If one assumes that birds arrive at South Point or Pointe aux Herbes after accumulating along the lakeshore as they arrive from the south and detour eastward, then arrival of only five northbound birds/hr at each 0.1 km of lakefront (estimating the detection range of an urban observer on the ground) would accumulate into a flight of 2,000 birds/hr at the peninsula. These concentrating factors are probably lacking elsewhere in the coastal wetlands, but study of movement patterns of migrants throughout the area would be worthwhile, and may provide insights regarding which areas are most critical to their preparations to cross the Gulf of Mexico. ACKNOWLEDGMENTS I thank Laura Alexander, Jessica Campbell, Oscar Camp- bell, Laura Dancer, Pete Dunne, David Mizrahi, Walter Ellison, Duncan Evered, David Juran, David Little, Craig Matthews, David Muth, B. Mac Myers, R. D. Purrington, Melissa Ryce, Denise Spindell, Steve Stevens, Paul Sykes, and Phillip Wallace for helpful discussions and/or assistance in data collection. David Muth. J. Van Rem.sen Jr., Ken Able, and an anonymous referee provided helpful reviews of the manuscript. The Bayou Sauvage National Wildlife Refuge was helpful in allowing access to the study site. LITERATURE CITED Able, K. P. 1972. Fall migration in coastal Louisiana and the evolution of migration patterns in the Gulf region. Wilson Bulletin 34:231-241. Able, K. P. 1977. The orientation of passerine nocturnal migrants following offshore drift. Auk 94:320-330. Akesson, S., L. Kaki.s.son, G. Walinuer, and T. Alerstam. 1996. Bimodal orientation and the occurrence of temporary reverse bird migration during autumn in southern Scandinavia. Behavioral Ecology and Sociobiology 38:293-302. Alerstam, T. 1978. Reoriented bird migration in coastal areas: dispersal to suitable resting grounds? Oikos 30:405-408. Alerstam, T. and S. Ulfstrand. 1975. Diurnal migration of passerine birds over south Sweden in relation to wind direction and topography. Ornis Scandinavica 6:135-149. Baird, J. and I. C. T. Nisbet. 1960. Northward fall migration on the Atlantic Coast and its relation to off- shore drift. Auk 77:1 19-149. DeSante, D. F. 1973. Analysis of the fall occurrences and nocturnal orientation of vagrant wood warblers (Parulidae) in California. Dissertation. Stanford Uni- versity, Palo Alto, California, USA. Duncan, R. A. and W. C. Weber. 1985. The Yellow Warbler: a diurnal circum-Gulf fall migrant. Florida Field Naturalist 13:20-22. Gauthreaux Jr., S. A. 1971. A radar and direct visual study of passerine .spring migration in southern Louisiana. Auk 88:343-365. Gauthreaux Jr., S. A. 1978. Importance of daytime flights of nocturnal migrants: redetermined migration follow- ing displacement. Pages 219-227 in Animal migration, navigation, and homing (K. Schmidt-Koenig and W, T. Keeton, Editors). Symposium held at the University of Tubingen, 17-20 August 1977. Proceedings in Life Sciences. Springer-Verlag, Berlin, Germany. Lowery Jr., G. H. 1945. Trans-Gulf spring migration of birds and the coastal hiatus. Wil.son Bulletin 57:92-121. Lowery Jr., G. H. 1974. Louisiana birds. Louisiana State University Press, Baton Rouge. USA. Moore, F. R. 1990. Evidence for redetermination of migratory direction following wind displacement. Auk 107:425-^28. Sandberg, R. and F. R. Moore. 1996. Migratory orientation of Red-Eyed Vireos, Vireo olivacens. in relation to energetic condition and ecological context. Behavioral Ecology and Sociobiology 39:1-10. Sandberg, R., F. R. Moore, J. Bi.ackman. and M. Lohmu.s. 2002. Orientation of nocturnally migrat- ing Swainson’s Thrush at dawn and dusk: importance of energetic condition and geomagnetic cues. Auk 1 19:201-209. Seymour, M. 2(K)9. Chasing the storm. Louisiana Department of Wildlife and Fisherie.s, Baton Rouge. USA. http:// www.wlf.louisiana.gov/onlinestore/laconscrvationist/ pcekinside Thorup, K. and j. Rab0l. 2007. Compensatory behaviour after displacement in migratory birds: a meta-analysis of cage experiments. Behavioral Ecology and Socio- biology 61 :825-841. The Wilson Journal of Ornithology 122(4):744-754, 2010 NIGHT MIGRANT FATALITIES AND OBSTRUCTION LIGHTING AT WIND TURBINES IN NORTH AMERICA PAUL KERLINGER,' " JOELLE L. GEHRING," WALLACE P. ERICKSON,' RICHARD CURRY,^ AAETAB JAIN,' AND JOHN GUARNACCIA^ ABSTRACT. — Avian collision latality data from studies conducted at 30 wind farms across North America were examined to estimate how many night migrants collide with turbines and towers, and how aviation obstruction lighting relates to collision fatalities. Fatality rates, adjusted for scavenging and searcher efficiency, of night migrants at turbines 54 to 125 m in height ranged from <1 bird/turbine/year to ~7 birds/turbine/year with higher rates recorded in eastern North America and lowest rates in the west. Multi-bird fatality events (defined as >3 birds killed in 1 night at 1 turbine) were rare, recorded at <0.02% (n = 4) of ~25.000 turbine searches. Lighting and weather conditions may have been causative factors in the tour documented multi-bird fatality events, but flashing red lights (L-864, recommended by the Federal Aviation Administration [FAA]) were not involved, which is the most common obstruction lighting used at wind farms. A Wilcoxon signed-rank analysis of unadjusted fatality rates revealed no significant differences between fatality rates at turbines with FAA lights as opposed to turbines without lighting at the same wind farm. Received 30 May 2006. Accepted 29 June 2010. Songbirds often collide with communication towers, lighthouses, skyscrapers, and other struc- tures during nocturnal migration with fatalities, at times, numbering in the hundreds or even thousands of birds in a single night (Banks 1979, Avery et al. 1980, Trapp 1998, Kerlinger 2000). Research suggests lighting has a primary role in attracting or disorienting night-migrating songbirds at those structures, especially during overcast, foggy, or rainy conditions (Cochran and Graber 1958, Caldwell and Wallace 1966, Avery et al. 1976). Many studies, attempting to assess how lights influence bird behavior, have focused on communication towers. Larkin and Erase (1988) used tracking radar to show that, during fog and low cloud ceiling, night migrants circled a >3()5-m tall communication tower with Federal Aviation Administration (FAA)-approved ob- struction lighting, but departed the tower when lights were extinguished. Gauthreaux and Belser (2006) used marine surveillance radar and infra- red scopes to compare night migrant activity at two tall (>305 m), lighted communication towers with guy wires and at a .separate control plot. One ' Curry & Kerlinger L.L.C., P. O. Box 453. Cape May Point, NJ 08212. USA. ^ Michigan Natural Features Inventory. Stevens T. Mason Building. P. O. Box 30444, Lansing, Ml 48909, USA. 'Western EcoSystems Technologies Inc.. 2003 Central Avenue. Cheyenne. WY 82001. USA. ^ Curry & Kerlinger L.L.C., 1734 Susquehannock Drive, McLean. V A 22101, USA. ^302 Bryn Mawr Drive SE. Albuquerque. NM 87106, USA. 1407 Finniown Road. Waldoboro, ME 04572. USA. ''Corresponding author; e-mail: pkerlinger@comca.st.net of the towers had flashing and steady-burning red lights, the other had only flashing white strobe lights. They found that birds flew in straight flight paths over the control plot but, at the lit towers, flight paths were curvilinear and birds concen- trated near the towers. More birds concentrated at the tower with flashing and steady-burning red lights than at the tower with flashing white strobe lights. This led Gauthreaux and Belser (2006) to suggest that research was needed to improve understanding of the attraction of different obstruction lighting systems to night-migrating .songbirds. A recent study in Michigan by Gehring et al. (2009) of 24 communication towers with different heights, support systems, and lighting found that towers lit at night with only tJa.shing red or white lights had significantly fewer avian fatalities than towers lit with a combination of steady-burning and fla.shing lights. These results suggest avian fatalities can be reduced, perhaps by 50-71%, at communi- cation towers supported by guy wires by replacing steady-burning lights with flashing lights. Avian collision fatalities are well documented at wind turbines (turbine rotor or tower), but the estimated total number of birds killed annually is small relative to communication towers and other structures (Erickson et al. 2005, National Re- search Council 2007). Wind turbines >60 m in height are often equipped with FAA-approved obstruction lighting, but not all turbines need to be lit, as long as gaps between lighted structures do not exceed 0.8 km (FAA 2()()()). Thus —25-33% of turbines at most wind farms should have obstruction lighting on the nacelle (where the 744 Ker/iiigcr et al. • NIGHT MIGRANT FATALITIES AT WIND TURBINES 745 I FIG. 1 . Wind turbine showing location of Federal Aviation Administration obstruction lights on the nacelle. blades are attached to the rotor hub; Fig. 1) adjacent to where the rotors are attached. The FAA (2000) recommends flashing red lights be placed on the turbine nacelle but flashing white and steady-burning red lights have also been used on a limited number of turbines. The flashing red lights are model L-864, a red strobe, LED (light- emitting diode), or pulsating incandescent light that flashes 20-40 times/min with an intensity of 2,000 candela. Steady-burning red lights are classified as model L-810, an incandescent red lighting that has a minimum intensity of 32.5 candela. L-810 lights are red and occasionally are modified to flash. Flashing white lights are classified as model L-865 and are a white strobe typically set at 40-60 flashes/min with an intensity of up to 2,000 candela. These same types of obstruction lights are used on communi- cation towers (FAA 2000, Gehring et al. 2009). Concerns about avian mortality have prompted fatality studies at wind farms across the United States and Canada. Most studies are available as reports to meet permit requirements and are often reviewed by state and federal wildlife agencies. This paper reviews existing information on night migrant mortality at wind farms derived from 31 studies at 30 wind farms. Our objectives were to: (1) examine the incidence of multi-bird fatality events at individual wind turbines and their relationship with lighting, and (2) examine whether disproportionately greater numbers of fatalities occur at turbines equipped with FAA- approved lights as opposed to turbines without approved lights. METHODS Data were extracted from post-construction, avian fatality studies conducted at 30 wind farms across the United States and Canada (Table 1). Studies prior to 1995 were not included because turbines lacked obstruction lighting and were <50 m in height (to maximum blade tip height). Excluded studies in this category were conducted at Altamont Pass Wind Resource Area, San Gorgonio Pass, and Tehachapi Mountains, all in California. The results of studies at several small wind farms conducted after 1995 were also not included because lighting or search methodolo- gies could not be verified. Other wind farms in the United States and Canada were not included because they have not been studied or, if they have been studied, reports were not available. Data extracted included geographic region; turbine nameplate production in megawatts (MW); turbine height (to the maximum blade tip height); total number of turbines in the wind farm and the subset of turbines studied; lighting type, including the total number of turbines with lighting and the number of lit turbines studied; study duration and search interval during migra- tion seasons; unadjusted number of nocturnal migrant carcasses found; and estimated fatality rate (birds/turbine/year). Carcasses were assigned as night-migrating songbirds or similar species (i.e., cuckoos, etc.) based on migration tendency (nocturnal vs. diurnal) and date recorded. Methodologies for studying collision mortality at wind farms varied, but basically followed accepted practices (e.g., Anderson et al. 1999). Carcass searches were usually conducted every week to 1 month in the western United States, whereas they were usually conducted every 1- 2 days to 1 month in the eastern United States and Canada. Some studies used different search intervals at different subsets of turbines. Searchers TABLE I. Estimated avian migrant mortality per year (spring and fall) from 31 studies at 30 wind farms (Buffalo Mountain studies were done at each of the 2 types of turbines on site). 746 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. !22, No. 4. December 2010 V- - C c ” I >>-a >» ^ O (N OJ W) -O 0- 3 O = > O c 00 O tU O O r , o CN U (N ac H 1£ IS 'mm 3 3 V V c o ^ I c/3 -§ ^ — r- tS rs 1) u >> sC — ri o — O'! C^l ■a "o u u OX) OXJ c c 13 13 CA c/3 -3 u- u- o^ ^ :> ON ON :> r-- 0 c 15 u. 3 1 c^i I so C O £ o P "O 1) OXJ c — u. 4 4 O o ON OC \0 — : o o o o (N o (N 2003 D X k. o r-- o o 04 *3 *3 3 04 o "3 D D D ”3 m D in o Ui C C (— D o o D O O 5 OX) o C4 OX) c/3 c/: c/3 04 c d D D D 3 r- 3 X o o ‘C O 3 D m m > CD D C 0) c -D ^ u. u. u. 3 3 3 ^ >v £ £ 15 15 U. Urn 3 3 V V V V V ^ m o 3 X JO V3 3 X C 00 c r— D 00 o (N 3 D 04 £ 4 in 1 — d 04 S >> m s > lO — — (N 5 — ^ — •o "O “O ODD “ " " o ojj bij 00 ^ c c c — 13 13 13 ? c/3 c/3 ;/3 "O D Oil c 3 c3 u. u. E r- r- -C D c/3 C 3 O E Z ^ 00 vO ro ^ 00 so r^i > ^ ^ I s s so vO lO sO sC — d d 4 (N ^ ON r- so o^ sO r- ro r^i \C — d <1, u o Cj S 0^: / C/5 D CA X c c O' d. 3 o o D -d O C _D D D ■£ X - u -C 3 c c/3 D 0/j 3 X C '' P - c/3 D X w a. 3 5 > c J .i . QC 5 ^ U ^ D C O z £ 'd c E u 3 V vn (N D C V o o o (N 3 D C O c/3 C s: o D C E 3 I r- o o CN "3 D >v r- o o "3 D r- o o r4 >. D C D o .ii Q c/5 H U- u D ^ ^ .£ "d Ij JD C C 3 Ip Ip CN — — I V V (N c o (N "3 D D O X D C E Urn 3 V "O >> •o 3 D C o WJ •o D *- D II 3 ^ Ic — ^ c/3 D ^ SO «- fNj >v -a ^ ^ “O C 00 ^ 3 3 w - D D D (N 3 D >» OJ) c c o D D C I 3 "O c T3 d D X -3 3 3 *3 D N *5 3 E u 3 D .£ "o _0 D D 3 — CC3 3 .§ .§ “X ^ )’ c/3 c/3 D ( D D ' I I (N cN O ^ Z V o o z z 00 On 04 ON r- >>>>>% 3 3 3 ^ c/3 X >> >. c X c 3 -2 3 X rp fs) c/3 c/3 c/3 w c 1=; -C 3 c § c/3 c/3 (*S U !- C c c C_/ ^ E •2 o 3 D D ? >v C o £ o £ P in t-' — 04 04 r-- 00 so sd 2 ^ :::2, ^ r- — in 2 so d d w'lncNcn — "S "O X X X ^ D D D D Um u, u u i- OX) OX) 1 OX) OX) OX) OX) OX) OX) c p X) ^ c c c c 3 _c E E 3 04 E E E E E E c/3 c/3 •w' CA CA CA CA CA CA E 3 O 3 E E E 3 3 UL. X (X X X X X X X ON SO On On On o (n 00 ro ^ Cl in 04 04 ro S E ^ m 00 ^ E 00 so 04 cn m oo (N CN — . ^ CN in so so in ^ ^ ^ ^ J s s s s s in in in in m W P E 3 ca CO D D C 3 3 D cn £ E D to 3 X O U X .2 ^ s ^ 53 ^ S < oa >- Z 61) Lh 3 C D U UJ TABLE 1. Continued. Keiiingcr ct al. • NIGHT MIGRANT FATALITIES AT WIND TURBINES 747 ^ 3 00 ^ • 00 C ^ o. >.TD -O ^ (N 2^ o O o ^ o (N a> wj c ^ •;;; (U (U E E 5 15 u u, 5 E % cd ■o O S § oi) .E o fe .E ^ Z - u- ^ >-» 0) - (U Z ^ C/2 CJ o 00 r-i os OO 1 in 00 00 — * s — > > 2 00 00 s — ‘ 00 s nO m sO in d d d I r-- 04 00 q (^i < ^ CL S C E c o 5 o o CN % CS c/2 tZ) CS QJ T3 O c/2 >» C3 >, C3 o ^ 3 -Q 0) I (N — — r- oo m (n X 3 cd OS ' — OS (N CN CN ' — r- 00 00 o 1) *0 C4 X) X 7 X 15 fU u X "O X ~o 0) u. o t «u L_ f 04) c 0) L. (D u- X o u V Urn 04) 04) o 04) 04) 04) 04) OJD o u 04) 04) g 3 so 3 3 3 “S 3 E >> 3 3 -3 IE 1 15 15 15 cd CJ 15 15 X 15 -5 C/l (Zl c/2 c/2 c/2 3 c/2 c/2 cd c/2 c/2 -E cd 3 C3 3 o _cd Cd E Cd cd ll U. PL PL PL z PL PL PL PL 00 H > > z dj u uT C3 O (TD T3 -O O O CL CL O O 2 ^ y: y. B w o o ir c/3 ■o -o qj CD u O O CL CL U (U 3 3 O CJ 3 2 ^ U c A cx. A ■a o u o CL (D c C 3 0 ’E 3 3 3 ■ w'' 00 y y y y y y :/) , 0 B B B C/l B a B ■ (U 0 0 0 3 0 0) 0 L. k. u. u 3 L. L. u. - 3 3 3 3 . _< 3 3 3 3 3 3 3 00 3 3 3 3“ P" cr O’ 3 O’ O' O' ■- ^ u u o o O' On W) c c U -C -C U U U CL CL CL CL -2 -2 U. L. (U > o ro O o CN o c C “ 00 ii (N (N 04 ^ - 00 ?; O OJ (U ^ > U U c/:i O O o o (N o o o c (N (N - >- sO r- '3- ro so os 0- O' ro so 0- 0 ro '3- 0 0 Os *3- 0 00 0 0 04 0 0 — 0 m 04 04 00 ro VO 0 04 0 r-- m d d d d d d d d d d d d d ro d d d d d d 7 =" = u u •— oc 00 sO ro m 3" ro 0 0- 0- ro 04 0 0 0 c^t r-- 0- — ro ro ro r- so 3* ro 0 so VO 00 os m 0 0 ^ 3 0“i 10 o| ro — 04 ro 04 so 3; ro q — oc r- q 10 04 >0 Hi d d d d d 3" d d oi d carcasses found at lit turbine ■3’ ro 3" ro CO '3* fO 04 00 04 3- 00 00 VO 04 VO 04 ro ro 04 04 -i- vY r-* r^i so 00 or^ — VO o — r-- o GO O 00 O ^ «oOriO04 — o d d o o o 0^ ^ ro — sO On a^ sO VO n fo — 00 00 m r- o vO o — 04 o o o o o r- SO d d o o (N O ro — SO r- o o 103 O — 06 C<', Tt m (N so 00 (N vo r- — m 103 o Os Tj- o o V03 00(NiO3Or-- m 00 103 00 d d d — d ^ Os so ^ o r-j (^1 3 *0 o o o 0) O) 0^1 ro 00 t—4 y 0 >v sz ” CJ ’3 Ui 3 “O CJ y so 00 0 0 0 0 0 0 0 3 04 y 04 00 o o >- >- > > u S ME, OJ y 01 >-' y 04 CJ 00 y CJ 04) Un 3 CJ y CJ OX) CJ OX) uT CJ Cj' CJ 3 3 00 0 s X z d CJ 0 u z d CJ JT CJ L. z 72 5 CJ -C CJ u ;o c5 2 CJ ;o 5 72 5 CJ • E 3 0 0 3 a u- (/; u y L. 0 3 3 CJ y 0 3 3 CJ y y y CL 3 CJ LO CJ CL CJ CJ CL jj CL 3 3 (y5 CJ CJ ca 3 3 3 3 3 0 .4_) to 3 s S U U s S S S S s S 4; O Ui 3 3 months to 5 years. There was variation in search interval among studies. We estimate the total number of individ- ual turbine searches for all studies combined to be ~25,000 during spring and fall migration seasons. Incidence of Multi-bird Fatality Events. — The number of cai'casses found in the studies ranged from zero to 36/year, but when the number of turbines studied was considered, the average at turbines ranged from zero to ~5/turbine/year. Only four studies reported multi-bird fatality events (defined as >3 birds killed in one night at one stiTicture), but they were at turbines with lighting other than flashing red or at turbines with ancillary, non-EAA-type lighting. Fourteen freshly-killed birds ( 1 1 warblers, 2 flycatchers, and 1 vireo) were recovered on 19 May 1999 at two adjacent turbines in a section of the wind fai'm at Buffalo Ridge in southwestern Minnesota (Johnson et al. 2000) where eveiy other turbine was lit with steady- buming red lights. This event may have been related to a severe thunderstorm that occuired the night before the carcasses were discovered. This was the only multi-bird fatality event reported at the Minnesota site during 4 years of study. Two multi-bird fatality events (3 fatalities at turbine # 2 on 10 Oct 2000, 7 fatalities at turbine # 3 on 31 Oct 2002) were recorded at Buffalo Mountain I in Tennessee (Nicholson et al. 2005) where each of three turbines had a pair of flashing white lights. The 10 October 2000 mortality event occuiTed during clear, cold, windy conditions following passage of a strong cold front, and the 31 October 2002 mortality event occurred during mild, rainy weather. About 27 night-migrating passerines were found dead on the morning of 23 May 2003 at Mountaineer in West Virginia (Kerns and Ker- linger 2004) in the vicinity of three turbines and an electrical substation that was brightly lit at night by at least four sodium-vapor lamps (steady- burning white flood lights). Heavy fog occuired the night before this fatality event was di.scovered by maintenance workers, who alerted searchers. Seventeen of the fatalities were discovered at turbine # 23, which was —50 m from the .substation, five at the substation (collisions with 750 THE WILSON JOURNAL OF ORNITHOLOGY • Voi 122, No. 4, December 2010 fencing), and three and two respectively at the flanking turbines, # 22 and 24, which were ~250 and 125 m, respectively from the substation. No lurther multi-bird fatality events occurred at the substation and adjacent turbines, including foggy nights, after the sodium-vapor lamps were extin- guished. Few fatalities were found at searched turbines throughout the study, including those that had flashing red lights. Comparison of Lit and Unlit Turbines at Same Site. — The FAA does not recommend that all turbines at a wind farm be lit, but we were able to test whether disproportionate numbers of fatalities occurred at turbines with obstruction lighting. There were two sites at which none of the turbines was lit and seven at which all of turbines were lit (Table 1). The two completely unlit sites had 0.55-0.60 MW turbines in the 59- 61 m height range, and estimated fatality rates of <1 night migrant/turbine/year. Sites where all turbines were lit had 0.66-1.65 MW turbines in the 79-1 17 m height range, and estimated fatality rates of <1 to ~7 night migrants/turbine/year. Sites in eastern North America reported slightly greater estimated fatality rates (Table 1). This appeared to be the case for turbines of roughly equal height. Wind farms (n = 22) with both lit and unlit turbines had 0.36-2.00 MW turbines in the 54-125 m height range. Estimated fatalities ranged from <1 to ~4 birds/turbine/year with greater rates in eastern North America. Absolute numbers of fatalities were sufficiently large at 13 of 22 sites to make quantitative comparisons of fatality rates at lit and unlit turbines (Table 2). The rates were roughly the same at sites with flashing red lights, but may have been slightly greater at turbines lit with steady-burning red lights. We did not find that a significantly greater proportion of wind farm studies (Table 2) report- ed greater numbers of fatalities at lit turbines versus unlit turbines than expected by chance (Wilcoxon sign-rank test, n — 20, Z = 0.18, P — 0.43). This lack of evidence of a consistent trend with respect to fatality rates at lit versus unlit turbines across North America indicates it is highly unlikely flashing red lights are associated with greater fatalities. All but one of the studies reviewed that tested for differences in fatality rates at lit and unlit turbines reported no significant differences. The one exception was at Maple Ridge in New York, where a marginally significant difference was reported (0.10 > P > 0.05) in one of two statistical te.sts. No significant differences in collision mortality were reported at wind farms with lit and unlit turbines where some or all lit turbines had other than flashing red lights. These included Crescent Ridge, Illinois (Kerlinger et al. 2007), Buffalo Ridge, Minnesota (Johnson et al. 2000, 2002), and Buffalo Mountain, Tennessee (Fiedler et al. 2007). Unadjusted mortality at Erie Shores, Ontario (James 2008), where all lit turbines had steady-burning red lights, was four times greater at lit than at unlit turbines (0.4/turbine vs. 0.1/ turbine). The unadjusted rates, when turbines along the shore of Lake Ontario were removed from the calculations, were similar (0.35/turbine at lit vs. 0.28/turbine at unlit). DISCUSSION Collision fatality data from 30 wind farms in North America demonstrate that fatality rates of night migrating birds at these structures are relatively low, ranging between ~1 to 7/turbine/ year. A comparison of estimated fatalities by geographic region showed a gradient with fatal- ities increasing from western to eastern North America. This mirrors continent-wide studies of bird migration patterns (Lowrey and Newman 1966, Gauthreaux et al. 2003) that show the density of migration in central and eastern states is greater than recorded in western states. What is striking about the data from wind farms is the relative absence of large-scale fatality events, similar to those recorded at tall communication towers supported by guy wires, where collisions of hundreds of birds at times occur in a single night (Avery et al. 1980, Kerlinger 2000). Only four incidents (<0.02% of searches) were reported of multi-bird fatality events at wind turbines (/? = >3 carcasses at a single turbine on a single night) during ~25,000 turbine searches in all studies combined. That so many studies and so many searches have been conducted at wind turbines without recording large-scale fatality events strong- ly suggests the probability of large-scale fatality events occurring is extremely low. Ear more systematic research has been conducted at wind turbines than at communication towers or other structures, which supports the finding that large- scale fatality events rarely occur at wind turbines. We believe there are three reasons why large- scale fatality events apparently do not occur at wind turbines and why fatality rates at wind turbines are much lower than has been reported for communication towers. First, communication Kerlingcr el al. • NIGI 11' MIGRANT FATALITIES AT WIND TURBINES 751 towers at which lai'ge-scale fatality events and lai'ge numbers ot fatalities have been reported are taller than wind turbines. Wind turbines have rarely exceeded 125 m in height (this study), while communication towers for which large-scale fatal- ities have been found largely exceed 150 m and often exceed 305 m (Avery et al. 1980, Shire et al. 2000). Thus, communication towers extend into altitudes where more night migrants fly (Kerlinger and Moore 1989), than do wind turbines. Second, all communication towers for which lai'ge numbers of night migrant fatalities and large-scale events have been noted have guy wires (Shire et al. 2000), whereas wind turbines do not. The third reason for higher fatality rates and large-scale fatality events at communication towers opposed to wind turbines may be related to the types of lights that are placed on communication towers and wind turbines. A majority of commu- nication towers equipped with aviation obstruction lights have both steady-burning red (L-810) lights and flashing red (L-864) lights (FAA 2000). Wind turbines are most often equipped with only flashing red (L-864) lights (FAA 2000). The four multi-bird fatalities at single (or adjacent) turbines during a single night occurred at turbines with flashing white lights, steady-burning red lights, or ancillai7 facility lighting (sodium vapor lamps). Flashing red lights were not implicated in these multi-bird fatality events at turbines. Gehring et al. (2009) recently demonstrated for communication towers that steady-burning red lights attract night migrants, but flashing red lights do not. Gehring et al. (2009) also demon- strated towers at heights ranging from 116 to 146 m, supported by guy wires and equipped with only flashing red lights, experienced 50-70% fewer fatalities than towers with both flashing and steady burning red lights. The average fatality rate was 17.5 carcasses found per 40 days of peak migration (20 days in spring and 20 days in fall) per communication tower with both flashing and steady-burning red lights. The estimated fatality rate at these towers, when adjustments for searcher efficiency and carcass removal by scavengers is included (equal to about a 2-fold increase; Gehring et al. 2009), is likely to be as high as 70 night migrants/tower/year when adjustments were made to include the entire spring and fall migration seasons. Gehring et al. (2009) reported 74 carcasses/40 days of searching/ year (spring and fall seasons) without carcass removal or searcher efficiency adjustments at towers >305 m in height with guy wires. Adjustments for carcass removal and searcher efficiency, as well as adjustments that include the entire spring and fall migration seasons, increased these estimates to —300 birds/tower/year. Thus, communication towers equipped with guy wires and a combination of flashing and steady-burning red lights have fatality rates that are one to two orders of magnitude greater than wind turbines. We strongly suggest that reported fatalities of night-migrating birds are minimal at wind tur- bines based on results reported for communication towers (Gehring et al. 2009), especially when compared to tall, communication towers with guy wires. Note that Smallwood et al. (2010) estimat- ed mortality at wind turbines was greater when novel scavenger removal methods were used. We did not find evidence that large-scale fatality events occur at wind turbines or that the flashing red lights normally used on wind turbines cause large numbers of fatalities of night migrants. Our results, combined with those reported by Gehring et al. (2009), strongly suggest that wind turbines be equipped only with flashing red lights (strobe or LED) and that steady burning red lights not be used on turbines. It is prudent that post-construc- tion fatality studies continue at wind turbines, especially those taller than those we report, and at turbines erected in geographic areas where studies have not been conducted and greater numbers of birds may migrate. ACKNOWLEDGMENTS We thank the diligent researchers, who spent thousands of hours in the field collecting data, and the wind-power industry, which funded the studies analyzed. LITERATURE CITED Anderson, R. L., M. Morrison, K. Sinclair, and M. D. Strickland. 1999. Studying wind energy/bird inter- actions: a guidance document. Metrics and methods for determining or monitoring potential impacts on birds at existing and proposed wind energy sites. National Wind Coordinating Committee. Washington, D.C., USA. Avery, M., P. Springer, and .1. Cassel. 1976. The effects of a tall tower on nocturnal bird migration — a portable ceilometer study. Auk 93:281-291. Avery, M. L., P. F. Springer, and N. S. Dailey. 1980. Avian mortality at man-made structures: an annotated bibliography. USDI, Fish and Wildlife Service. FWS/ OBS-80/54. Washington. D.C., USA. Banks, R. C. 1979. Fluman related mortality of birds in the United States. USDI, Fish and Wildlife Service, Scientific Report Wildlife 215. Washington, D.C., USA. 752 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 4. December 2010 Caldwell, L. and G. Wallace. 1966. Collections of migrating birds at Michigan television towers. Jack- Pine Warbler 44:117-123. Cochran, W. and R. Gr.aber. 1958. Attraction of nocturnal migrants by lights on a television tower. Wilson Bulletin 70:378-380. Derby, C., A. Dahl, W. Erickson, K. Bay, and J. Hoban. 2007. Post-construction monitoring report for avian and bat mortality at the NPPD Ainsworth Wind Farm. Report to the Nebraska Public Power District, Columbus, Nebraska, USA. Western EcoSystems Technology Inc., Cheyenne, Wyoming, USA. Erick-SON, W., K. Kronner, and B. Gritski. 2003. Nine Canyon Wind Power Project avian and bat monitoring report: September 2002-August 2003. Report to the Nine Canyon Technical Advisory Committee and Energy Northwest, Benton County, Washington, USA. Western EcoSystems Technology Inc., Chey- enne, Wyoming, USA. Erickson, W. P., G. D. Johnson, and D. P. Young. 2005. A summary and comparison of bird mortality from anthropogenic causes with an emphasis on collisions. Pages 1029-1042 in USDA, Forest Service, General Technical Report PSW-GTR-191. Albany, California, USA. Erickson, W. P., G. D. Johnson, M. D. Strickland, and K. Kronner. 2000. Avian and bat mortality associated with the Vansycle Wind Project. Umatilla County, Oregon: 1999 study year. Report to Umatilla County Department of Resource Services and Development, Pendleton, Oregon, USA. Western EcoSystems Tech- nology Inc., Cheyenne, Wyoming, USA. Erickson, W. P., J. Jeffrey, K. Kronner, and K. Bay. 2004. Stateline Wind Project Wildlife Monitoring Final Report: July 2001-December 2003. Technical report peer-reviewed by and submitted to FPL Energy, the Oregon Energy Facility Siting Council, and the Stateline Technical Advisory Committee, Umatilla County. Oregon and Walla Walla County. Washing- ton. USA. Western EcoSystems Technology Inc., Cheyenne, Wyoming, USA. Federal Aviation Adminlstration (FAA). 2000. Ob- struction marking and lighting. AC 70/7460- IK. U.S. Department of Transportation. Washington. D.C., USA. Fiedler, J. K.. T. H. Henry. R. D. Tankersley, and C. P. Nicholson. 2007. Results ol' bat and bird mortality monitoring at the expanded Buffalo Mountain Wind- farm. 2005. Tennessee Valley Authority, Knoxville. USA. Gauthreaux Jr., S. and C. Belser. 2006. Effects of artificial night lighting on migrating birds. Pages 67- 93 in Ecological consequences of artificial night lighting (C. Rich and T. Longcore. Editors). Island Press. Washington. D.C., USA. Gauthreaux Jr.. S. A., C. G. Belser. and D. A. van Bi.aricom. 2003. Using a network of WSR-88D weather surveillance radars to define patterns o( bird migration at large spatial scales. Pages 335-346 in Avian migration (P. Bcrthold. E. Gwinner. and E. Sonnenschein. Editors). Springer-Verlag. Berlin. Germany. Gehring, j. L., P. Kerlinger, and a. M. Manville II. 2009. Communication towers, lights, and birds: successful methods of reducing the frequency of avian collisions. Ecological Applications 19: 505-514. Howe, R. W., W. Evans, and A. T. Wolf. 2002. Effects of wind turbines on birds and bats in northeastern Wisconsin. Report to Wisconsin Public Service Corporation and Madison Gas and Electric Company, Madison, Wisconsin, USA. University of Wisconsin- Green Bay, USA. Jain, A. A. 2005. Bird and bat behavior and mortality at a northern Iowa windfarm. Thesis. Iowa State Univer- sity, Ames, USA. Jain, A., P. Kerlinger, R. Curry, and L. Slobodnik. 2007. Annual report for the Maple Ridge Wind Power Project: post-construction bird and bat fatality study- 2006. Report to PPM Energy and Horizon Energy and Technical Advisory Committee for the Maple Ridge Project Study, McLean, Virginia, USA. Curry & Kerlinger LLC, McLean, Virginia, USA. Jain, A., P. Kerlinger, R. Curry, and L. Slobodnik. 2009a. Annual report for the Maple Ridge Wind Power Project: post-construction bird and bat fatality study - 2007. Report to PPM Energy and Horizon Energy and Technical Advisory Committee for the Maple Ridge Project Study. Houston, Texas, USA. Curry & Kerlinger LLC. McLean, Virginia, USA. Jain, A., P. Kerlinger, R. Curry, and L. Slobodnik. 2009b. Annual report for the Maple Ridge Wind Power Project: post-construction bird and bat fatality study-2008. Report to PPM Energy and Horizon Energy and Technical Advisory Committee for the Maple Ridge Project Study, Houston, Texas, USA. Cuny & Kerlinger LLC. McLean, Virginia. USA. Jain, A., P. Kerlinger, R. Curry, L. Slobodnik, J. Histed, and J. Meacham. 2009c. Annual report for the Noble Clinton Windpark LLC: postconstruction bird and bat fatality study-2008. Report to Noble Environmental Power LLC, Bliss, New York. USA. Curry & Kerlinger LLC, McLean, Virginia, USA. Jain, A., P. Kerlinger, R. Curry, L. Slobodnik, J. Quant, and D. Pursell. 2009d. Annual report for the Noble Bliss Windpark LLC: postconstruction bird and bat fatality study-2008. Report to Noble Environmen- tal Power LLC, Bliss, New York, USA. Curry & Kerlinger LLC, McLean, Virginia, USA. Jain. A., P. Kerlinger. R. Curry, L. Slobodnik, A. FUER.ST. and C. Hansen. 2009e. Annual report for the Noble Ellenburg Windpark LLC: postconstruction bird and bat fatality study-2008. Report to Noble Environ- mental Power LLC. Bliss, New York, USA. Curry & Kerlinger LLC, McLean, Virginia, USA. James, R. D. 2004. Bird observations at the Pickering wind turbine. Ontario Birds 21:84-97. James. R. D. 2008. Erie Shores Wind Farm. Port Burwell. Ontario: fieldwork report for 2006 and 2007, during the first two years of operation. Report to Environment Canada. Ontario Ministry of Natural Resources, Erie Shores Wind Farm LP-McQuarrie North American, and AIM PowerGen Corporation, Ontario, Canada. Kerlini^cr cl al. • NIGHT MIGRANT I’ATALITIES AT WIND TURBINES 753 James, R. D. and G. Coady. 2003. Bird monitoring at Toronto’.s Exhibition Place wind turbine. Ontario Birds 22:79-88. Johnson, G., W. Erickson, J. White, and R. McKinney. 2003. Avian and bat mortality during the first year of operation at the Klondike Phase I Wind Project, Sherman County, Oregon. Draft report to Northwest- ern Wind Power. Goldendale, Washington, USA. Western EcoSystems Technology Inc., Cheyenne, Wyoming, USA. Johnson, G. D., W. P. Erickson, M. D. Strickland, M. F. Shepherd, and D. A. Shepherd. 2000. Avian monitoring studies at the Buffalo Ridge. Minnesota Wind Resource Area: results of a 4-year study. Report to Northern States Power Company. Minneapolis, Minnesota. USA. Western EcoSystems Technology Inc., Cheyenne, Wyoming, USA. Johnson, G. D., W. P. Erickson, M. D. Strickland, M. F. Shepherd, D. A. Shepherd, and S. A. Sarappo. 2002. Collision mortality of local and migrant birds at the large-scale wind power development on Buffalo Ridge, Minnesota. Wildlife Society Bulletin 30:879- 887. Kerlinger, P. 2000. Avian mortality at communications towers: a review of recent literature, research, and methodology. Report to the USDI, Fish and Wildlife Service, Washington, D.C., USA. Curry & Kerlinger LLC, McLean, Virginia, USA. Kerlinger. P. 2001. Avian mortality study at the Green Mountain Windpower Project, Garrett, Somerset County, Pennsylvania: 2000-2001. Report to National Wind Power, Oakland, California, USA. Curry & Kerlinger LLC, McLean, Virginia, USA. Kerlinger, P. 2002a. An assessment of the impacts of Green Mountain Power Corporation's wind power facility on breeding and migrating birds in Searsburg, Vermont. Report NREL/SR-500-28591 to National Renewable Energy Laboratory, U.S. Department of Energy, Golden, Colorado. USA. Cape May, New Jersey, USA. Kerlinger, P. 2002b. Avian fatality study at the Madison Wind Power Project, Madison, New York. Report to PG&E Generating, Athens, New York, USA. Curry & Kerlinger LLC, McLean, Virginia, USA. Kerlinger, P. and R. Curry. 2000. Impacts of a small wind power facility in Weld County, Colorado, on breeding, migrating, and wintering birds: preliminary results and conclusions. Pages 64-69 in Proceedings of the National Avian-Wind Power Planning Meeting III. National Wind Coordinating Committee/RESOLVE, Washington, D.C., USA. Kerlinger. P. and F. R. Moore. 1989. Atmospheric structure and avian migration. Current Ornithology 6:109-142. Kerlinger. P., R. Curry, A. FIasch, and J. Guarnaccia. 2007. Migratory bird and bat monitoring study at the Crescent Ridge Wind Power Project. Bureau County, Illinois: September 2005-August 2006. Report to Orrick, Herrington, and Sutcliffe LLP, Washington, D.C, USA. Curry & Kerlinger LLC, McLean, Virginia, USA. Kerlinger, P., R. Curry, L. Culr, A. Jain, C. Wilk- erson, B. Fischer, and A. Ha.sch. 2006. Post- construction avian and bat I'alality monitoring study for the High Winds wind power project, Solano County, California: two year report. Report to High Winds LLC and FPL Energy, Livermore. California, USA. Curry & Kerlinger LLC, McLean, Virginia, USA. Kerns, J. and P. Kerlinger. 2004. A study of bird and bat collision fatalities at the Mountaineer Wind Energy Center, Tucker County, West Virginia: annual report for 2003. Report to FPL Energy and the MWEC Technical Review Committee, Vero Beach, Florida, USA. Curry & Kerlinger LLC, McLean, Virginia, USA. Koford, R., a. Jain, G. Zenner, and A. Hancock. 2005. Avian mortality associated with the Top of Iowa Wind Farm. Report to the Iowa Department of Natural Resources. Ames, USA. Iowa State University, Ames. USA. Larkin, R. and B. Frase. 1988. Circular paths of birds flying near a broadcasting tower in cloud. Journal of Comparative Psychology 102:90-93. Lowrey Jr, G. H. and R. J. Newman. 1966. A continent- wide view of bird migration on four nights in October. Auk 83:547-586. National Research Council. 2007. Environmental impacts of wind-energy projects. NRC, Committee on Environmental Impacts of Wind Energy Projects. The National Academies Press, Washington, D.C., USA. New Jersey Audubon Society. 2008. Post-construction wildlife monitoring at the Atlantic County Utilities Authority-Jersey Atlantic Wind Power Facility: peri- odic report covering work conducted between 1 January and 30 September 2008. Report to the New Jersey Board of Public Utilities and New Jersey Clean Energy Program, Newark, New Jersey, USA. New Jersey Audubon Society, Cape May Court House. New Jersey, USA. Nicholson, C. P., R. G. Tankersley Jr., J. K, Fiedler, and N. S. Nicholas. 2005. Assessment and prediction of bird and bat mortality at wind energy facilities in the southeastern United States: final report. 2005. Tennessee Valley Authority, Knoxville, USA. ScHNELL. G. D.. E. A. Mosteller, and j. Grzybowski. 2007. Post-construction avian/bat risk assessment and fatality study for the Blue Canyon II Wind Power Project. Oklahoma: summary of first-year findings. Report to Horizon Wind Energy. Houston. Texas. USA. Sam Noble Oklahoma Museum of Natural History, University of Oklahoma. Norman, USA Shire, G. G., K. Brown, and G. Winegrad. 2000. Communication towers: a deadly hazard to birds. American Bird Con.servancy, Washington. D.C.. USA. Smallwood, K. S., D. A, Bell, S. A, Snyder, and J. E. DiDonato. 2010. Novel scavenger removal trials increase wind turbine-caused avian fatality estimates. Journal of Wildlife Management 74:1089-1097. Stantec CONSUI.TING. 2008. 2007 spring, summer, and fall post-construction bird and bat mortality study at the 754 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 4. December 2010 Mars Hill wind farm, Maine. Report to UPC Wind Management LLC, Cumberland, Maine, USA. Stantec Consulting, Topsham, Maine, USA. Stantec Consulting. 2009. Post-construction monitoring at the Mars Hill wind farm, Maine-year 2, 2008. Report to First Wind Management LLC, Portland, Maine, USA. Stantec Consulting, Topsham, Maine, USA. Tierney, R. 2007. Buffalo Gap I Wind Farni avian mortality study, February 2006-January 2007, final survey report. Report to AES West Inc., Albuquerque, New Mexico, USA. TRC, Albuquerque, New Mexico, USA. Trapp, J. L. 1998. Bird kills at towers and other man-made structures: an annotated partial bibliography (1960- 1998). USDI, Fish and Wildlife Service, Washington, D.C., USA. TRC Environmental Corporation. 2008. Post-construc- tion avian and bat fatality monitoring and grassland bird displacement surveys at the Judith Gap Wind Energy Project, Wheatland County, Montana. Report to Judith Gap Energy LLC, Chicago, Illinois, USA. TRC Environmental Corporation, Laramie, Wyoming, USA. WiLCOXON, F. 1945. Individual comparisons by ranking methods. Biometrics 1:80-83. Young Jr., D. P., W. P. Erickson, R. E. Good, M. D. Strickland, and G. D. Johnson. 2003. Avian and bat mortality associated with the initial phase of the Eoote Creek Rim Windpower Project, Carbon County, Wyoming: November 1998-June 2002. Report to Pacificorp Inc., Portland, Oregon, SeaWest Wind- power Inc., San Diego, California, and Bureau of Land Management, Rawlins District Office, Rawlins, Wyoming, USA. Western EcoSystems Technology Inc., Cheyenne, Wyoming, USA. Young Jr., D. P., W. P. Erickson, K. Bay, S. Nomani, and W. Tidhar. 2009. Mount Storm Wind Energy Facility, phase 1 post-construction avian and bat monitoring, July-October 2008. Report to NedPower Mount Storm LLC, Houston, Texas, USA. Western EcoSystems Technology Inc., Cheyenne, Wyoming, USA. The H'ilsoii Jouriuil oJOniilhology 122(4):755-761, 20 10 DEVELOPMENTAL CHANGES IN SERUM ANDROGEN LEVELS OE EASTERN SCREECH-OWLS {MEGASCOPS ASIO) CORINNE P. KOZLOWSKI' -^ AND D. CALDWELL HAHN^ ABSTRACT. — We studied androgen production during development in nestling Eastern Screech-Owls (Megascops asio) and hypothesized that gender and hatch order might inOuence serum levels of testosterone and androstenedione. Testosterone levels were highest immediately after hatching and declined significantly in the 4 weeks leading to fledging. The average level of testosterone for 1-7 day-old owls was 3.99 ± 0.68 ng/ml. At 22-28 days of age, the average testosterone level for nestling owls was 0.83 ± 0. 1 8 ng/ml. Testosterone levels did not differ between males or females. The average testosterone level for male nestlings was 2.23 ± 0.29 ng/ml and 2.39 ± 0.56 ng/ml for female nestlings. The average level of androstenedione tor nestling owls was 1.92 ±0.11 ng/ml and levels remained constant throughout development. Levels were significantly higher in males than females. The average androstenedione level was 1.77 ± 0.16 ng/ml tor male nestlings and 1.05 ± 0.24 ng/ml tor female nestlings. Hatching order did not affect levels of either androgen. Our results provide a foundation for future studies of androgen production by nestling owls. Received 19 January 2010. Accepted 30 June 2010. Androgens influence many aspects of adult bird physiology and behavior, including metabolic rate (Hannsler and Prinzinger 1979), lipid storage (Wingfield 1984), timing of molt (Runfeldt and Wingfield 1985), song production (Hunt et al. 1997), and aggression (Hau et al. 2000). Andro- gens also influence the behavior of nestling birds, as testosterone levels in earlier-hatching White Stork nestlings {Ciconia ciconia) and Tawny Owl nestlings {Strix aluco) are correlated with elevated aggression towards siblings (Sasvari et al. 1999, 2006). Testosterone enhances the probability of aggressive behavior in young Domestic Chickens (Callus gallus domesticus) (Young and Rogers 1978) and facilitates aggression in Black-headed Gull chicks (Chroicoceplialus ridihundus) (Groothuis and Meeuwissen 1992, Ros et al. 2002). Testosterone also facilitates begging be- havior in nestling Atlantic Canaries (Serinus Canaria) (Buchanan et al. 2007) and European Pied Flycatchers (Ficedida hypoleuca) (Goodship and Buchanan 2006). Sex steroids are present early in avian devel- opment and are important in sexual differentia- tion. Testicular hormone synthesis in Domestic Chicken embryos has first been detected on day 6 of incubation, and peaks in plasma levels of testosterone occur on day 14 (Woods and Weeks ' St. Louis Zoo, Research Department, 1 Government Drive, St. Louis, MO 63110, USA. ^University of Missouri-St. Louis, Biology Department, 8001 Natural Bridge Road, St. Louis, MO 63121, USA. ^USGS, Patuxent Wildlife Researeh Center, 12100 Beeeh Forest Road, Laurel, MD 20708, USA. “Corresponding author; e-mail: corinnekozIowski@umsl.edu 1969). Similarly, androgen production begins on day 8 of incubation in embryos of male Japanese Quail (Coturnix japonica), and androgen concen- trations in plasma reach a plateau before hatching (Ottinger and Bakst 1981, Schumacher et al. 1988). Less is known concerning the maturation of the hypothalamic-pituitary-gonadal (HPG) axis in altricial birds. Sex steroids are present in the plasma of altricial hatchlings of several species (Hutchison et al. 1984, Adkins-Regan et al. 1990, Schlinger and Arnold 1992, Silverin and Sharp 1996). However, the typical levels of androgens produced by nestling birds, whether androgens are produced primarily in the gonads or the adrenal glands, or how these levels change through development have not been studied in detail. Many field studies sample individuals only once (e.g., Fargallo et al. 2007, Gil et al. 2008). However, a number of studies suggest that circulating androgen levels can vary during development. Testosterone levels of ne.stling Great Tits (Panis major) decline significantly during development (Silverin and Sharp 1996), and circulating androgen levels initially increase, then decrea.se post-hatching in nestling Zebra Finches (Taeniopygia guttata) (Adkins-Regan et al. 1990). Evidence concerning differences between male and female nestlings varies among species. Male and female Common Kestrel (Falco tinnunculus) (Fargallo et al. 2007), European Pied Flycatchers (Goodship and Buchanan 2006), Zebra Finches (Naguib et al. 2004), and nestling Lesser Black- backed Gulls (Lams fuscus) (Verboven et al. 2003) show no difference in circulating androgen levels. However, nestling male Black Coucals 755 756 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 4. December 2010 {Centropus grillii) have higher levels of circulat- ing androgens than females (Goymann et al. 2005) , and nestling female Atlantic Canaries (Weichel et al. 1986) and Japanese Quail (Ottinger et al. 2001) have higher levels of circulating androgens than males. Hormonal differences have also been reported among nestlings in regards to hatching order and may be due to social interactions (Wingfield et al. 1990). Second-hatching chicks were shown to have higher levels of circulating testosterone 1 day after hatching, as well as during aggressive encounters with older siblings in obligately siblicidal Nazca Boobies (Siila grcinti) (Ferree et al. 2004). First-hatched White Stork nestlings are more aggressive, receive more food, and have higher plasma levels of testosterone than their siblings (Sasvari et al. 1999). Similarly, plasma testosterone levels decreased with hatching order in 3-day-old Tawny Owls raised by females in poor condition. Owlets raised by females in good condition, however, had similar levels of testos- terone, regardless of hatching order (Sasvari et al. 2006) . Hatching order in other species is not related to plasma androgen levels (Naguib et al. 2004). We describe developmental patterns of andro- gen production by Eastern Screech-Owls (Mega- scops asio) during the nestling phase. Androgens, including testosterone and androstenedione, are produced primarily by the gonads. The adrenal glands are also a source of androgens, specifically androstenedione and dehydroepiandrosterone (DHEA), but production is minimal compared to the gonads. Most androgenic effects are mediated by testosterone and dihydrotestosterone. Andro- stenedione also circulates in the bloodstream at relatively high levels, but it has a significantly lower affinity for the androgen receptor compared to testosterone and dihydrotestosterone, and a relatively low biological activity in vivo. Instead, androstenedione serves primarily as a substrate for the production of testosterone and estrogens (Hadley and Levine 2007). Several studies have characterized levels of testosterone in nestling birds (Fargallo et al. 2007. Gil et al. 2008), but little is known about typical serum levels or developmental patterns of androstenedione. We measured .serum testosterone and andro- stenedione levels weekly in captive nestling Eastern Screech-Owls from hatching until Hedg- ing. Our objectives were to: ( I ) characterize levels of testosterone and androstenedione produced by nestling owls, (2) compare how these levels changed during the developmental period between hatching and fledging, (3) ascertain whether levels differ between male and female nestlings, and (4) whether levels differ among nestlings in regard to hatching order. METHODS Species and Study Area. — The Eastern Screech- Owl is a small owl (males are typically 160 g, females 200 g) with a distribution that extends east of the Rocky Mountains and south from the Canadian boreal forest to the Tropic of Cancer in Mexico (Gehlbach 1994). Clutches normally consist of 4-6 eggs; the first 2-3 eggs are laid 1 day apart with increasing intervals afterwards (Gehlbach 1994). Onset of incubation varies among females and depends on environmental conditions, but may begin with the first, second, or third egg. Incubation typically lasts 28 days; eggs hatch asynchronously over 2-3 days, and sibling feeding hierarchies are established. The young fledge between 28 and 30 days of age (Gehlbach 1994). The owls used in this study are part of a captive colony of breeding birds maintained at Patuxent Wildlife Research Center (PWRC) in Laurel, Maryland, USA. Owls are fed two mice per bird daily and Nebraska Bird of Prey Diet fortified with Vionate. They are housed in 12 X 3 m outdoor flight cages that contain nest-boxes, and pairs are kept together year-round. Egg-laying begins in March. Females solely incubate, and males provide food to incubating females and nestlings after hatching. Serum Collection and Androgen Assays. — Fif- ty-two blood samples were collected weekly during spring 2007 from 26 nestlings in nine clutches. Samples were collected only from nestlings hatching from first-laid clutches, and nestlings hatched between 28 April and 13 May 2007. Individual owls ranged from 2 to 27 days of age, and clutches contained 2-5 nestlings. Seven owl nestlings were sampled once, 12 were sampled twice, and 7 were sampled three times. No birds were sampled more that three times during the 4 weeks of blood collection. Serum samples were obtained by venipuncture from the brachial vein using a 27-gauge needle until the last week before Hedging, when they were taken from the jugular vein using a 25-gauge needle. A separate blood sample was also taken for identification of males and females Just before Hedging. Seium samples Kozlowski imd Hahn • SERUM ANDROCJEN LEVELS IN NESTEINCJ OWLS 757 were frozen until shipped to the St. Louis Zoo for hormone analysis. Serum samples were transferred to 2-ml eppendorf tubes and diluted 1:2 with phospho- buftered saline (PBS). Samples were mixed with a vortex and an equal volume of 100% ethanol was added to precipitate proteins and lipids following Kozlowski and Ricklefs (2010). Samples were immediately homogenized upon adding ethanol and allowed to incubate at room temperate for 10 min. Samples were then spun at 12,282 g in a centrifuge for 10 min. The supernatant was poured into a sterile cryule tube and frozen at —70° C until an assay was performed. All samples were analyzed using radioimmu- noassay (RIA) in the Endocrinology Laboratory at the St. Louis Zoo. Ethanol extracts were thawed and spun in a centrifuge at 12,282 g for 10 min to remove any remaining lipids in preparation for assay. Testosterone and androstenedione concen- trations were measured using commercially avail- able coated-tube radioimmunoassay kits (Coat-A- Count © Total Testosterone 1251 Kit, and Coat-A- Count © Direct Androstenedione 1251 Kit, Diag- nostic Products Corporation, Los Angeles, CA, USA). The lowest detection limit was 0.05 ng/ml and upper limit was 40 ng/ml in the testosterone assay. The lowest detection limit of the andro- stenedione assay was 0.04 ng/ml and the upper detection limit was 8.7 ng/ml. Both kits used in this study have highly specific antibodies. The testosterone antibody cross-reacts as follows: 5(3- androstan-3a, 17(3-diol: 0.4%; androstenedione: 0.5%; 5(3-androstan-3p, 17p-diol: 0.2%; 5a-dihy- drotestosterone: 3.3%; 5( 10)-estren-17a-ethinyl- 17P-ol-3-one: 0.2%; 4-estren-17a-methyl-17P-ol- 3-one: 1.1%; 4-estren- 1 7-ol-3-one: 20%; 19- nortestosterone: 20%; ethisterone: 0.7%; 19- hydroxyandrostenedione: 2.0%; 1 1-ketotestoster- one: 16%; methyltestosterone: 1.7%; norethin- drone: 0.1%; 1 1 P-hydroxyyestosterone 0.8%; and triamincinolone: 0.2%. The androstenedione an- tibody cross-reacts as follows: androsterone: 0.14%; DHEA: 0.16%; progesterone: 0.16%; spironolactone: 0.11%; 5a-dihydrotestosterone: 0.21%; and testosterone: 1.49%. Cross-reactivi- ties for all other compounds are below 0.1%. Assays were conducted following kit directions with the exception that kit standards, which are supplied in human serum, were replaced by standards diluted in 10% steroid-free calf serum. Standard diluent was added to extracted serum samples, and steroid-free owl serum extract was added to standards and quality controls to equalize the matrices of standards and samples. Owl .serum extract and calf serum were stripped of steroids using dextran-coated charcoal (DCC# 6241, Sigma Chemical, St. Louis, MO, USA) prior to use. All samples were analyzed in the same assay in duplicate for both testosterone and androstene- dione. Mean ± SE intra-assay variation of duplicate samples was 8.96 ± 1.21 for testoster- one and 9.66 ± 2.42 for androstenedione. Assay Validation. — We added a known amount of radioactively labeled hormone to a serum sample before extraction to measure extraction efficiency of both testosterone and androstenedi- one, and measured the amount of radioactivity after the extraction process. Serum was pooled from the experimental samples and divided into 10 samples, each containing 15 pL of serum. In five samples each for testosterone and andro- stenedione, 15 pL of 1-125 labeled hormone and 15 pL of PBS were added to the 15 pL of serum. Control samples consisted of 15 pL of labeled hormone and 30 pL of PBS. Two samples of 15 pL of 1-125 hormone were set aside to measure the total radioactivity present in the sample. Serum and control samples were extract- ed as described, and the supernatant was transferred from each sample to an individual 12 X 75 plastic test tube. The total amount of radioactivity in each sample was measured in both serum and control samples, and compared to the total count tubes to calculate the recovery percentage in each sample. Percent recovery was 91-97% for both hormones and did not differ between control and serum samples (testosterone: t = 1 .22, df = 8, P = 0.26; androstenedione: t = 0.73, df = 8, P = 0.49). Pour samples that contained high levels of hormone were diluted by 1/2, 1/4, and 1/8 with steroid-free owl serum extract for both testoster- one and androstenedione. Serial dilutions of extracted .serum samples were parallel to the standard curve (test of equal slopes, P > 0.34 for testosterone; P > 0.29 for androstenedione) (Zar 1999), supporting that no additional substances in the extract were cross reacting with the antibody. We assessed the accuracy of both assays by adding a known amount of either testosterone or androstenedione to four serum extracts containing low values of hormone. Addition of known amounts of the androgens at two dosage levels resulted in recovery of 101 ± 3.6% of added testosterone and 96 ± 2.9% for androstenedione. 758 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4, December 2010 O) c c o c 0) o c o o c 0) O) o •D c 05 £ 0) CO 3.5 3.0 2.5 2.0 1.5 1.0 0.5 0.0 • Male O Female \ \ Testosterone Androstenedione FIG. I. Concentration {x ± SE) of testosterone and androstenedione in the serum of male and female Eastern Screech-Owl nestlings (n = 19 males, 7 females). Identification of Males and Females. — We followed Griffiths et al. (1998) to identify each individual nestling Screech-Owl as male or female by amplifying introns of the genes CHD- W and CHD-Z through PCR. Statistical Procedures. — All statistical analysis was conducted using JMP 7.0.2 © (SAS Institute 2008). Blood samples were collected weekly, and nestlings were not sampled on the same day during each week; thus, nestlings were divided into 7-day age classes. These classes ranged from 1-7 to 22-28 days of age. General Linear Models were used to assess the relationship between testosterone and androstenedione in relation to age-class, hatch-order, brood size, and male/ female. A separate analysis was conducted for each androgen. Nest of origin and individual owl were treated as random effects. Hatch-order, male/female, brood size, and all two-way inter- actions were included as fixed effects. Non- significant fixed effects were dropped from the model. A Tukey HSD test was used to separate means when results were significant. RESULTS Both testosterone and androstenedione were detected in the serum of nestling Screech-Owls. The average concentration of testosterone in nestling owls was 2.30 ± 0.26 ng/ml and levels ranged from 0.17 to 7.59 ng/ml. The average concentration of androstenedione was 1 .92 ± 0.1 1 ng/ml and ranged from 0.98 to 3.51 ng/ml. Testosterone and androstenedione levels were unaffected by both brood size and hatch-order, and these effects were dropped from both statistical models. Seven of the nestlings in our study were 5.0 O) c c o 4.0 03 L_ O 3.0 o c o o c (D O) O k_ C 03 E D w 2.0 1.0 0.0 • Testosterone O Androstenedione 14 21 28 Age (days) FIG. 2. Weekly concentration (± SE) of testosterone and androstenedione in the serum of Eastern Screech-Owl nestlings: n = \2 nestlings in age-class 1; n = 18 nestlings in age-class 2; /t = 17 nestlings in age-class 3; /t = 5 nestlings in age-class 4. females, and 19 were males. No differences in serum testosterone were observed between male and female nestlings (7^1,46 = 0.68, P = 0.41). Average testosterone levels were 2.23 ± 0.29 ng/ml for male nestlings and 2.39 ± 0.56 ng/ml for female nestlings. In contrast, serum androstenedione levels differed significantly between male and female nestlings. Male owls had higher levels of andro- stenedione than females (F| 4^ = 5.173, P = 0.030). Average androstenedione levels were 1.77 ± 0.16 ng/ml for male nestlings and 1.05 ± 0.24 ng/ ml for female nestlings (Fig. 1). Testosterone levels did not differ between male and female nestlings, and the data were pooled for analysis. Nestling testosterone concentrations declined with age (P3.46 = 11.489, P < 0.001) (Fig. 2). In contrast, androstenedione levels did not vary by age for either male (^3,46 = 1.58, P = 0.21 ) or female (^3,46 = 0.37, P = 0.78) nestlings. Testosterone levels were higher than androstene- dione levels during the 2 weeks following hatching. However, by weeks 3 and 4, testoster- one and androstenedione were present at similar levels. Average testosterone levels in chicks 1- 7 days of age measured 3.99 ± 0.68 ng/ml. By week 4 (days 22-28), levels declined to less that one-quarter of their initial level, to 0.83 ±0.18 ng/ ml. Androstenedione concentrations averaged 1.72 ± 0.38 ng/ml for 1-7 day old male nestlings and 1.01 ± 0.65 ng/ml for 1-7 day old female nestlings. Average androstenedione levels by 22- 28 days of age were 1.39 ± 0.15 ng/ml for males and 1 .25 ± 0.58 for females. Kozlonski and Halm • SERUM ANDROGEN LEVELS IN NESTLING OWLS 759 DISCUSSION Testosterone and androstenedione are both measurable in nestling Screech-Owl serum from week 1 post-hatching until fledging. Our results demonstrate both a significant change in the level of testosterone during development and a difference in the pattern of testosterone and androstenedione during development. Testosterone is highest during the first week following hatching, and then declines. Androstenedione levels remain relatively constant from hatching through fledging. The HPG axis has been studied in detail in precocial species, but less is known about honnone production by altricial nestlings. The ovary and testes actively produce steroid hormones, including testosterone, during early embryonic development through post-embryonic development in precocial Domestic Chickens (Tanabe et al. 1979) and Mallards {Anas platyrhynchos) (Tanabe et al. 1983). The adrenal glands also appear to be important for androgen synthesis in precocial species. Testosterone is found at high concentra- tions in the adrenal gland shortly before hatching in both Domestic Chickens (Tanabe et al. 1979) and Mallards (Tanabe et al. 1983). Developmental studies of the HPG axis in altricial species are restricted to passerines. Limited work suggests the HPG axis of these species matures later in development. Silverin and Sharp (1996) demonstrated the hypothalamus and pitui- tary gland in Great Tits become functional 9 days after hatching, approximately halfway through their nestling phase. Injections of GnRH at this age elicit increases in male testosterone or female estradiol. The adrenal gland may also be an important source of androgens for altricial nestlings. Castration of nestling Zebra Finches did not significantly reduce plasma androgen levels, suggesting these androgens may arise primarily from the adrenal gland (Adkins-Regan et al. 1990). Our study suggests nestling Screech-Owls are capable of producing measureable levels of both testosterone and andro- stenedione. However, whether these androgens are gonadal or adrenal is unknown. Further study is needed to ascertain if development of the HPG axis in other altricial species, including owls, is similar to passerine species. We found a decrease in nestling testosterone levels during post-hatching development similar to other studies (Adkins-Regan et al. 1990, Silverin and Sharp 1996). Testosterone levels in Great Tit nestlings are highest in newly hatched birds, and reach basal levels by 3 days of age (Silverin and Sharp 1996). Male Zebra Finches have higher levels of plasma testosterone during days 2-14 of development than in older nestlings (Adkins-Regan et al. 1990). Similarly, Schlinger and Arnold (1992) found male Zebra Finches had higher levels of plasma androgens at 7-9 days of age compared to 4-6-day-old birds or 10- 13-day- old birds. Elevated testosterone levels are also characteristic of the period surrounding hatching in several precocial species, including Japanese Quail (Ottinger and Bakst 1981, Schumacher et al. 1988) and Domestic Chickens (Tanabe et al. 1979). This suggests testosterone may have a role in hatching. In contrast, androstenedione levels of nestling owls did not vary during post-hatching development. Androstenedione levels of precocial species change during development, but little is known about developmental patterns in altricial nestlings. Androstenedione levels in Domestic Chickens are elevated in pre-pubescent males (Culbert et al. 1977), and then decrease between 9 and 16 weeks of age when testicular maturation occurs (Sharp et al. 1977). Whether similar changes occur during nestling owl development is unknown. We found no differences in testosterone levels between male and female nestling Screech-Owls at any stage of development. This is similar to the findings for several other species (Groothuis and Meeuwissen 1992, Verboven et al. 2003) including Common Kestrels (Fargallo et al. 2007). In contrast, we detected a difference in androstenedi- one level, as male Screech-Owls had higher levels of androstenedione than females from hatching through fledging. There is limited evidence that male nestlings of other species may also maintain higher levels of androstenedione than females. Male European Starling {Sturnus vulgaris) nest- lings have been shown to produce higher levels of DHEA (a precursor of androstenedione) than female nestlings (Chin et al. 2008). Elevated levels of testosterone in nestling birds have been corre- lated with behavioral changes, including enhanced aggression (Sasvari et al. 1999, 2006), and begging intensity (Goodship and Buchanan 2006). Howev- er, it is unknown whether higher levels of androstenedione in male nestlings would produce similar effects. Androstenedione has a much lower affinity for the androgen receptor compared to testosterone and a relatively low biological activity (Hadley and Levine 2007). Whether higher levels of androstenedione in male nestlings have any physiological effects is unknown. 760 THE WILSON JOURNAL OF ORNITHOLOGY • Vul. 122, No. 4. December 2010 Hatching order was not associated with differ- ences in androgen levels in nestling Screech- Owls. The correlation between hatching order and serum androgens can be associated with the presence of sibling feeding hierarchies. First- hatched nestlings in several species have higher plasma levels of testosterone, are more aggres- sive, and receive more food than younger siblings (Sasvari et al. 1999, 2006). Sibling feeding hierarchies are often established in wild Screech- Owl broods, and nestlings vigorously compete for food resources (Gehlbach 1994). However, the owls in this study were from a captive colony where they have an abundant food supply. High androgen levels are known to be immunosuppres- sive and reduce growth for nestlings, in addition to increasing their aggressive behavior (Ros 1999, Fargallo et al. 2007 ). We have observed that captive owl nestlings show little sibling competition, and lack of a correlation between androgens and hatching order may be associated with reduced competition. Future work will characterize the androgen levels of Screech-Owl nestlings when food resources are restricted. This will allow us to ascertain if androgen levels are affected by hatching order when there is strong sibling competition among siblings for food. ACKNOWLEDGMENTS We thank USGS Patuxent Wildlife Research Center and G. W. Smith for use of the owl colony and research support to DCH, and the Maryland Ornithological Society, American Mu.seum of Natural History Frank W. Chapman Fund, and Sigma Xi for research support to CPK. We thank J. E. Bauman and the St. Louis Zoo Research Department for access to endocrinology laboratory facilities and R. E. Ricklefs for access to genetic laboratory facilities. We thank W. C. Bauer and M. M. Maxie for care of the owls, and Kelly Amy and Nathan Rolls for assistance in collecting blood samples. Helpful reviews of the manuscript were provided by B. A. Rattner. D. J. Hoffman, J. .1. Atwood, and two anonymous reviewers. Use of trade, product, or firm names does not imply endorsement by the U.S. Government. LITERATURE CITED Adkin.s-Rf.gan, E.. M. Abdelnabi. M. Mobarak, and M. A. Ottinger. 1990. Sex .steroid levels in developing and adult male and female Zebra Finches (Poephila fpittatci). General and Comparative Endocrinology 78:93-109. Buchanan, K. L., A. R. Goi.dsmith. C. A. Hinde, S. C. Grifmth, and R. M. Kii.nf.r. 2007. Does testosterone mediate the trade-off between nestling begging and growth in the Canary (Seriinis canarin)! Hormones and Behavior .32:664-671. CuLBERT, J., P. J. Sharp, and J. W. Wells. 1977. Concentrations of androstenedione. testosterone and LH in the blood before and after the onset of spermatogenesis in the cockerel. Reproduction and Fertility 51:153-154. Chin, E. H., A. H. Shah, K. L. Schmidt, L. D. Sheldon, O. P. Love, and K. K. Soma. 2008. Sex differences in DHEA and estradiol during development in a wild songbird: jugular versus brachial plasma. Hormones and Behavior 54:194-202. Fargallo, J. A., J. Marti'nez-Padilla, A. Toledano- Diaz. j. Santiago-Moreno, and j. a. DAvila. 2007. Sex and testosterone effects on growth, immunity, and melanin coloration in nestling Eurasian Kestrels. Journal of Animal Ecology 76:201-209. Ferree, E. D., M, C. Wikelski, and D. J. Anderson. 2004. Hormonal correlates of siblicide in Nazca Boobies: support for the challenge hypothesis. Animal Behavior 46:655-662. Gehlbach, F. R. 1994. The Eastern Screech Owl: life history, ecology and behavior in the suburbs and countryside. Texas A & M University Press, College Station, USA. Gil, D., E. Bulmer, P. Celis, and M. Puerta. 2008. Increased sibling competition does not increase testosterone or corticosterone levels in nestlings of the Spotless Starling (Stiirnus iinicolor). Hormones and Behavior 54:238-243. Goodship, N. M. and K. L. Buchanan. 2006. Nestling testosterone is associated with begging behaviour and fledging success in the Pied Flycatcher, Ficedula hypoleitca. Proceedings of the Royal Society of London, Series B 273:71-76. Goymann, W., B. Kempenaers, and j. C. Wingfield. 2005. Breeding biology, sexually dimoiphic develop- ment and nestling testosterone concentrations of the classically polyandrous African Black Coucal, Cen- tropus grillii. Journal of Ornithology 146:314-324. Griffiths, R., M. C. Double, K. Orr, and R. J. Dawson. 1998. A DNA test to sex most birds. Molecular Ecology 7:1071-1075. Groothuis, T. G. G. and G. Meeuwissen. 1992. The influence of testosterone on the development and fixation of form of displays in two age classes of young Black-Headed Gulls. Animal Behaviour 43:189-208. Hadley, M. E. and J. E. Levine. 2007. Endocrinology. Sixth Edition. Pearson Prentice-Hall, Upper Saddle River, New Jersey, USA. Hannsler, I. AND R. Prinzinger. 1979. The influence of the .sex hormone testosterone on body temperature and metabolism of Japanese Quail. Experientia 35:509-5 10, Hau, M., M. Wikelski, K. K. Soma, and J. C. Wingfield. 2000. Testosterone and year-round territorial aggres- sion in a tropical bird. General and Comparative Endocrinology 117:20-33. Hunt, K. E., T. P. Hahn, and J. C. Wingfield. 1997. Testosterone implants increase song but not aggression in male Lapland Longspurs. Animal Behaviour .54:1 177-1 192. Hutchison, J. B., J. C. Wingfiei.d, and R. E. Hutchison. 1984. Sex steroid differences in plasma concentrations Koz/ow.ski ami Hahn • SERUM ANDROCJEN LEVEES IN NESTLINCi OWLS 761 ot steroids during the sensitive period tor brain ditterentiation in the Zebra Finch. Journal of Endocri- nology 103:363-369. Kozlowski. C. P. and R. E. Ricklefs. 2010. Egg size and yolk steroids vary across the laying order in cockatiel clutches: a strategy for reinforcing brood size hierarchies? General and Comparative Endocrinology 168:460-465. Naguib. M., K. Riebel, A. Marzal, and D. Gil. 2004. Nestling immunocompetence and testosterone covary with brood size in a song bird. Proceedings of the Royal Society of London, Series B 271:833-838. Ottinger, M. a. and M. R. Bakst. 1981. Peripheral androgen concentrations and testicular morphology in embryonic and young male Japanese Quail. General and Comparative Endocrinology 43:170-177. OiTiNGER, M. A., S. Pitts, and M. A. Abdelnabi. 2001. Steroid hormones during embryonic development in Japanese Quail: plasma, gonadal, and adrenal levels. Poultry Science 80:795-799. Ros, a. F. H. 1999. Effects of testosterone on growth, plumage pigmentation, and mortality in Black-headed Gull chicks. Ibis 141:451-459. Ros, A. F. H., S. J. Dieleman, and T. G. G. Groothuis. 2002. Social stimuli, testosterone, and aggression in gull chicks: support for the challenge hypothesis. Hormones and Behavior 41:334—342. Runfeldt, S. and J. C. WtNGFiELD. 1985. Experimentally prolonged sexual activity in female sparrows delays termination of reproductive activity in their untreated mates. Animal Behavior 33:403^1 1. SAS Institute. 2008. JMP 7.0.2. SAS Institute Inc., Cary, North Carolina, USA. SasvAri, L., Z. Hegyi, and P. Peczely. 1999. Brood reduction in White Storks mediated through asymme- tries in plasma testosterone concentrations in chicks. Ethology 105:569-582. SasvAri, L., Z. Hegyi, and P. Peczely. 2006. Maternal condition affects the offspring testosterone concentra- tion in early nestling period: an experimental test on Tawny Owls. Ethology 1 12:355-361. SCHLINGER, B. A. and A. P. ARNOLD. 1992. Plasma sex steroids and tissue aromatization in hatchling Zebra Finch: implications for sexual differentiation of singing behavior. Endocrinology 130:289-299. Schumacher, M., J. Sulon, and J. Balthazart. 1988. Changes in serum concentrations of steroids during embryonic and post-hatching development of male and female Japanese Quail, Cotuniix coturnix japonica. Journal of Endocrinology 1 18:127-134. Sharp, P. J., J. Culbert, ANt) J. W. Wells. 1977. Variation in stored and plasma concentrations of androgens during sexual development in the cockerel. Journal of Endocrinology 74:467-476. Silverin, B. AND P. J. Sharp. 1996. The development of the hypothalamic-pituitary-gonadal axis in juvenile Great Tits. General and Comparative Endocrinology 103:150-166. Tanabe, Y., T. Yano, and T. Nakamura. 1983. Steroid hormone synthesis by the testes, ovary, and adrenals of embryonic and post-embryonic ducks. General and Comparative Endocrinology 49:144-153. Tanabe, Y., T. Nakamura, K. Fujioka, and O. Doi. 1979. Production and secretion of sex steroid hormones by the testes, the ovary, and the adrenal glands of embryonic and young chickens (Callus domesticus). General and Comparative Endocrinology 39:26-33. Verboven, N., P. Monaghan, D. M. Evans, H. Schwabl, N. Evans, and C. Whitelaw. 2003. Maternal condition, yolk androgens and offspring performance: a supplemental feeding experiment in the Lesser Black- Backed Gull (Larus fiiscus). Proceedings of the Royal Society of London, Series B 270:2223-2232. Weichel, K., G. Schwager, P. Heid, H. R. GOttinger, AND A. Pesch. 1986. Sex differences in plasma steroid concentrations and singing behaviour during ontogeny in canaries (Serinus canaria). Ethology 73:281-294. Wingfield, J. C. 1984. Androgens and mating systems: testosterone-induced polygyny in normally monoga- mous birds. Auk 101:665-671. Wingfield, J. C., R. E. Hegner, A. M. Dufty, and G. F. Ball. 1990. The challenge hypothesis: theoretical implications for patterns of testosterone secretion, mating systems, and breeding strategies. American Naturalist 136:829-846. Woods, J. E. and R. C. Weeks. 1969. Ontogenesis of the pituitary-gonadal axis in the chick embryo. General and Comparative Endocrinology 13:242-254. Young, C. E. and L. J. Rogers. 1978. Effects of steroidal hormones on sexual, attack, and search behavior in the i,solated male chick. Hormones and Behavior 10:107- 1 17. Zar, j. H. 1999. Biostatistical analysis. Fourth Edition. Prentice-Hall. Upper Saddle River, New Jersey. USA. The Wilson Journal of Ornithology 122(4):762-766, 2010 STRESS RESPONSIVENESS DECREASES WITH AGE IN PRECOCIAL, JUVENILE CHUKAR MOLLY J. DICKENS’ - AND L. MICHAEL ROMERO' ABSTRACT. — Development of the hypothalamic-pituitary-adrenal (HPA) axis and subsequent corticosterone (CORT) release in newly hatched birds is a balance between limiting exposure to the detrimental effects of CORT on growth and development, and the necessity of mounting an acute stress response. We measured the stress responsiveness of juvenile Chukar (Alectoris chukar) 20 to 60 days post-hatch prior to molting into full adult plumage. The integrated CORT response during 60 min of restraint in these individuals decreased with age. Comparisons to mean adult integrated CORT values imply the youngest juveniles have greater responses than adults while the oldest juveniles have reduced CORT responses. This pattern is currently an anomaly within the altricial-precocial range and suggests specific tradeoffs, such as molting into adult plumage, may affect timing of HPA suppression during development. Received 23 November 2009. Accepted 20 April 2010. The acute stress response in adult animals is an adaptive physiological reaction to noxious stimuli and is hallmarked by release of glucocorticoids following a hormone cascade along the hypotha- lamic-pituitary-adrenal (HPA) axis (Sapolsky et al. 2000). Corticosterone (CORT) is the predom- inant glucocorticoid secreted in avian species and, after CORT is released into the system, negative feedback mechanisms act to quickly shut down CORT secretion (Dallman and Bhatnagar 2001). Excessive CORT release, either due to a disrupted HPA axis or altered negative feedback sensitivity, can lead to physiological and behavioral pathol- ogies (McEwen 2005). Glucocorticoids, in terms of development, have severe effects on growth as they promote muscle breakdown; elevated con- centrations of glucocorticoids lead to muscle wasting and decreased growth rates (Wingfield 1994, Morici et al. 1997, Sapolsky et al. 2000). In addition, glucocorticoid concentrations can affect key aspects of neuronal development in the central nervous system (Sapolsky and Meaney 1986). Elevated CORT can also affect feather develop- ment (Romero et al. 2005, DesRochers et al. 2009). Development of the HPA axis and the stress response system in young birds post-hatch is considered to be a delicate balance between benefiting from the ability to mount an acute stress response and limiting the damaging effects of glucocorticoids. This balance requires tight regulation due to the impact of glucocorticoids on growth and feather development (Walker et al. 2005, Wada et al. 2007). Altricial laboratory rodents display a clear and easily defined “stress hyporesponsive period” or SHRP, beginning ' Department of Biology, Tufts University, Medford, MA 021.55, U.SA. ^Corresponding author; e-mail: molly.dickens@tufts.edu 2 days postnatal (reviewed by Sapolsky and Meaney 1986). A hyporesponsive period has also been noted in many altricial and semi-altricial avian species including Magellanic Penguins {Spheniscus magellanicus) (Walker et al. 2005), White-crowned Sparrows {Zonotrichia leucophrys nuttalli) (Wada et al. 2007), Northern Mocking- birds {Mimus polyglottos) (Sims and Holberton 2000), American Kestrels (Falco sparverius) (Love et al. 2003), Slender-billed Prions {Pachyp- tila belcheri) (Quillfeldt et al. 2009), and Atlantic Canaries (Serinus canaria) (Schwabl 1999). Hatchlings were shown to be fully capable of mounting an adult-like response within days of hatching by precocial species including Mallards {Anas platyrhynchos) (Holmes et al. 1990), chickens {Gallus gallus domesticus) (Freeman 1982), and Domestic Turkeys (Meleagris gallo- pavo) (Wentworth and Hussein 1985), and in semi-precocial Great-winged Petrels (Pterodroma macroptera) (Adams et al. 2008) and the semi- altricial Snowy Owl {Bubo scandiacus) (Romero et al. 2006), The HPA response, in most studies of altricial and semi-altricial avian species, slowly increases with age and fledglings about to leave the nest are close to adult-like responses (Schwabl 1999, Love et al. 2003, Walker et al. 2005, Bias et al. 2006, Wada et al. 2007). Chickens, a precocial species, show a clear but short hyporesponsive period immediately following hatching (Freeman 1982). However, there are limited data on development of the stress response in precocial species that match studies tracking development of the stress- respon.se system in altricial species. Our objective was to measure the stress responsiveness of juvenile Chukar {Alectoris chukar), within an age range in which they were 762 Dickens and Romero • STRESS RES[T)NSI VENESS OF WIED CIIUKAR 763 post-hatch but had yet to develop full adult plumage, to track the pattern of their HPA development. METHODS Study Species. — Chukar are non-migratory de- sert-adapted ground birds. They were introduced into North America from the steep, arid mountains of southwestern Eurasia as game birds, and can now be found extensively throughout the arid mountains of the American west (Christensen 1996). We captured wild juvenile and adult Chukar from a remote region of the Mojave Desert within the U.S. Navy China Lake Naval Air Weapons Station (CLNAWS) near Ridge- crest, California (117°37'W, 36° 3' N). Limited human access into CLNAWS assured minimal, non-experimental human interaction with our population. Juvenile Chukar in desert environ- ments can be relatively easy to capture due to limited dispersal during summer months and faithfulness to a specific water source (Larsen et al. 2007). Chukar are precocial and leave the nest with their mother immediately after hatching, and are fully capable of flight within 2 weeks of hatching (Christensen 1996). We trapped Chukar at human- made rainwater catchments known as “guzzlers” and one spring-fed pool from mid-July to August in 2006 and 2008. The guzzler system and trapping technique at CLNAWS have been described (Delehanty et al. 2004). Field Procedures. — We weighed Chukar im- mediately upon capture and u,sed the absence of a full black collar and eye band coloration as an indication of their juvenile status (Christensen 1996). We captured 31 juvenile Chukar with body mass ranging from 115 to 390 g. Juveniles representing the range of masses were captured evenly across the sampling period. These masses, according to the growth chart in Pis (2003), indicate juveniles were between 20 and 60 days of age. Adults in this region are, on average, 500 g (based on raw data collected for Dickens et al. 2009a, b; Pig. I). Juveniles typically reach adult size at ~12 weeks of age (Christensen 1996). Our study was limited by method of capture and time period in which it was conducted. We did not obtain samples from chicks immediately post- hatch as we were not following nests, and trapping began within a period in which most chicks were at least a few weeks of age. Capture-Stress Protocol. — Birds captured were provided with shade, ad libitum water, and were removed from the trap within 3 hrs of entering. Time spent in the trap was not ideal and we previously justified this experimental methodology (Dickens et al. 2009a). All individuals for which we were able to obtain baseline samples (samples taken within 3 min of approaching the trap, n = 27) demonstrated baseline samples well below their respective peak restraint-induced CORT concentra- tions, and all baseline concentrations were within the range expected for this species. We sampled CORT responsiveness of all individuals by placing them into an opaque bag and taking blood samples at 15, 30, and 60 min to measure the capacity for an endocrine stress response. All animal use complied with AALAC guidelines and was approved by the Tufts Institutional Animal Care and Use Committee. CORT was measured using radioimmunoassay (RIA) (Wingfield et al. 1992) as previously used for this species (Dickens et al. 2009a, b). Statistical Analysis. — We calculated the relative total CORT release over the 60-min restraint period as an integrated CORT value using measurements at 15, 30, and 60 min (Romero 2004). This measurement allowed us to compare the overall capacity of the HPA axis to release CORT over the 60-min restraint period. We performed a regression comparing each individu- al’s integrated CORT versus their mass (Graph- Pad, Prizm 4b). We also analyzed a group of adults sampled upon first capture from the wild to compare adult mass and the adult stress respon- siveness. We present these results as means ± SE (subset of data reanalyzed from that in Dickens et al. 2009a, b). RESULTS We found a correlation between body mass and integrated CORT concentrations that was signifi- cantly different from zero (Eig. I; r = 0.32, P < 0.001) for the 31 juveniles captured and sampled. Adult Chukar (/? = 19), captured during the same years as the juveniles, had a mean integrated CORT response of 2,188.0 ± 142.5 ng/mL/min and a mean body mass of 453.3 ± 12.2 g. There was no con-elation between adult mass and integrated CORT concentrations (r = 0.002, P = 0.904). DISCUSSION CORT response decreased with age in the precocial Chukar for the age range investigated 764 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 4, December 2010 FIG. 1. Individual points (n = 31) represent a single juvenile’s integrated CORT measurement versus body mass. The regression line is significantly different from zero (r~ = 0.32, P < 0.001). The open circle represents the adult mean values ± SE for each parameter {n = 19). and the heaviest, oldest, birds had the lowest response. The average adult response was in the middle of the juvenile response range. The youngest birds had relatively higher stress re- sponses than adults while the oldest juveniles, closest to full maturity, had a relatively lower response than that of adults. This pattern contrasts to that observed in altricial and semi-altricial species for which the HPA response becomes more similar to adult-like responses as the nestlings reach Hedging (e.g.. Love et al. 2003, Walker et al. 2005, Wada et al. 2007). The slow development of the HPA respon.se in these altricial species fits well with the developmental hypothesis summarized by Bias et al. (2006). Altricial species are either completely dependent or substantially dependent on parents for all aspects of their lives (Ricklefs 1979). They are also completely constrained to the nest and their capacity to e.scape a stressor, such as a predator, is extremely limited. Having an HPA response to an ine.scapable stimulus would lead to elevated CORT concentrations and do more harm to their growth and development than good for their recovery from the stressor (Bias et al. 2006). Precocial young, in contrast, are out of the nest and exposed to the same stressors as adults. Precocial juveniles of a variety of species are as fully capable of a quick escape from a predator as are adults (Herrel and Gibb 2004) and, for Chukar, all major requirements for quick flight e.scape occur within 8 days post-hatch (Jackson et al. 2009). Juveniles can physically escape as well as adults, and it also seems likely they initiate an adult-like CORT respon.se. Obtaining an adult- like response early in life is the pattern ob.served in the limited studies that investigate precocial young. There is a hyporesponsive period -in chickens immediately following hatch but chicks are able to mount a full CORT respon.se within the first day (Freeman 1982). Full responsiveness by the semi-precocial Great-winged Petrel was mea- sured after hatch and chicks remained responsive throughout the juvenile period (Adams et al. 2008). However, Mallard duckling respon.ses. Dickens am! Romero • STRESS RESPONSIVENESS OE WILD CIIUKAR 765 immedialely alter hatch, were more robust and actually decreased to match the adult response by day 21 (Holmes et al. 1990). We also observed a more robust response by younger than older Juvenile Chukar. We were not able to sample individuals earlier than 20 days after hatch and do not know whether Chukar hatchlings show some hyporesponsive period immediately after hatch similar to altricial species. We also cannot discuss the potential HPA response profile of the birds during this earlier developmental stage. However, the striking difference that our data confirm is that precocial young can have an adult or an even greater CORT response within a few weeks of hatching. The capacity of the juvenile Chukar response appears to decrease beyond the adult response as they become older. These data imply a hidden tradeoff in elevating CORT concentrations during the late part of the Juvenile stage. Preventing excessive CORT exposure outweighs the benefits of having the capacity for an adult CORT response. The tradeoff may reflect continued growth during this time, although the rate of growth is less than in younger Juveniles (Pis 2003). Alternatively, successful molting of Juve- nile feathers to grow adult feathers during this time may be sufficiently important to trump full HPA responsiveness. Chukar begin prebasic molt at 5 weeks of age, and the molt is nearly complete by 16 weeks (Christensen 1996). CORT concen- trations, both baseline and stress-induced concen- trations, are low during seasonal molt in adults (Romero and Remage-Healey 2000). Elevated CORT concentrations can lead to decreased protein deposition in feathers (Romero et al. 2005) and, perhaps as a result, reduced feather quality (DesRochers et al. 2009). Juveniles are ~250 g in mass at 5 weeks of age (the time of Prebasic molt according to Christensen 1996), if we extrapolate age from mass in the graph provided by Pis (2003), this age relates to the point at which the line (Fig. I) decreases below the adult average. Older Juveniles are growing adult flight feathers and replacing other body feathers to form the black band and collar typical of this species. Elevated CORT concentrations in Barn Owl {Tyto alba) chicks can lead to decreased deposition of pigment in growing feathers (Roulin et al. 2008). Maintaining low CORT during this period of feather replacement may be important for not only building strong adult flight feathers but also for acquiring appropriate adult coloration. More research is necessary to ascertain whether this pattern of HPA responsiveness is unique to Chukar, present in precocial species in general, or present only in precocial species that have distinct adult coloration patterns. This pattern is currently an anomaly in the growing literature investigating how the HPA develops within altricial-precocial species and may help understand the specific tradeoffs that affect when birds should suppress HPA responsiveness or express full adult-like responses. ACKNOWLEDGMENTS We thank the China Lake Naval Air Weapons Station for the opportunity to conduct these studies, T. G. Campbell for his support on the base, and D. J. Delehanty, K. A. Earle, Christine Lattin, J. L. Brubaker, and Joshua Leisenring for help in the field. Funding was provided by NSF Grant # IOB-0542099 to LMR and Sigma Xi Grants- in-Aid of Research to MJD. LITERATURE CITED Adams, N. J., J. F. Cockrem, E. J. Candy, and G. A. Taylor. 2008. Non-precocial Grey-faced Petrel chicks (Pterodroma macroptera gouldi) show no age-related variation in corticosterone responses to capture and handling. General and Comparative Endocrinology 157:86-90. Blas, j., R. Baos, G. R. Bortolotti, T. A. Marchant, AND F. Hiraldo. 2006. Age-related variation in the adrenocortical response to stress in nestling White Storks {Ciconia eiconia) supports the developmental hypothesis. General and Comparative Endocrinology 148:172-180. Christensen, G. C. 1996. Chukar (Alectoris chukar). The birds of North America. Number 258. Dallman, M. F. AND S. Bhatnagar. 2001. Chronic stress and energy balance: role of the hypothalamo-pitui- tary-adrenal axis. Pages 179-210 in Handbook of physiology; Section 7: the endocrine system; Volume IV: coping with the environment: neural and endocrine mechanisms (B. S. McEwen and H. M. Goodman, Editors). Oxford University Press. New York, USA. Delehanty, D. J., S. S. Eaton, and T. G. Campbell. 2004. From the field: Mountain Quail fidelity to guzzlers in the Mojave Desert. Wildlife Society Bulletin 32:588-593. Desrochers. D. W., j. M. Reed. J. Awerman. J. A. Kluge, J. Wilkinson, L. 1. van Griethuijsen. J. Aman, and L. M. Romero. 2009. Exogenous and endogenous corticosterone alter feather quality. Com- parative Biochemistry and Physiology — Part A: Mo- lecular & Integrative Physiology 152:46-52. Dickens, M. J., D. .1. Delehanty, and L. M. Romero. 2009a. Stress and translocation: alterations in the stress physiology of translocated birds. Proceedings of the Royal Society. Series B 276:2051-2056. 766 THE WILSON JOURNAL OF ORNITHOLOGY • VoL 122, No. 4, December 2010 Dickens, M. J., K. A. Earle, and L. M. Romero. 2009b. Initial transference of wild birds to captivity alters stress physiology. General and Comparative Endocri- nology 160:76-83. Freeman, B. M. 1982. Stress non-responsiveness in the newly-hatched fowl. Comparative Biochemistry and Physiology — Part A: Physiology 72:251-253. Herrel, a. and a. C. Gibb. 2004. Ontogeny of perfomiance in vertebrates. Pages 1-6 in Symposium on the ontogeny of performance in vertebrates held at the 7th Interna- tional Congress of Vertebrate Morphology. University of Chicago Press, Boca Raton, Florida, USA. Holmes, W. N., J. Cronshaw, and K. E. Rohde. 1990. Stress-induced adrenal steroidogenic responsiveness of the Mallard duck {Anas platyrhynchos) during early post-natal development. Comparative Biochemistry and Physiology — Part A: Physiology 92:403^08. Jackson, B. E., P. Segre, and K. P. Dial. 2009. Precocial development of locomotor performance in a ground- dwelling bird (Alectoris chukar): negotiating a three- dimensional terrestrial environment. Proceedings of the Royal Society, Series B 276:3457-3466. Larsen, R. T., J. T. Flinders, D. L. Mitchell, E. R. Perkins, and D. G. Whiting. 2007. Chukar watering patterns and water site selection. Rangeland Ecology and Management 60:559-565. Love, O. P., D. M. Bird, and L. J. Shutt. 2003. Plasma corticosterone in American Kestrel siblings: Effects of age, hatching order, and hatching asynchrony. Hor- mones and Behavior 43:480-488. McEwen, B. S. 2005. Stressed or stressed out: what is the difference? Journal of Psychiatry & Neuroscience 30:315-318. Morici, L. a., R. M. Elsey, and V. A. Lance. 1997. Effects of long-term corticosterone implants on growth and immune function in juvenile alligators. Alligator mississippiensis. Journal of Experimental Zoology 279:156-162. Pis, T. 2003. Energy metabolism and thermoregulation in hand-reared Chukars (Alectoris chukar). Comparative Biochemistry and Physiology — Part A: Molecular & Integrative Physiology 136:757-770. Quillfeldt, P., M. Poisbleau, O. Chastel, and j. E. Masello. 2009. Acute stress hyporesponsive period in nestling Thin-billed Prions Pachyptila helcheri. Journal of Comparative Physiology A: Neuroethol- ogy, Sensory, Neural, and Behavioral Physiology 195:91-98. Rk’KLEFS, R. E. 1979. Adaptation, constraint, and compro- mi.se in avian postnatal development. Biological Reviews of the Cambridge Philosophical Society 54:269-290. Romero, L. M. 2004. Physiological stress in ecology: lessons from biomedical re.search. Trends in Ecology and Evolution 19:249-255. Romero. L. M.. D. W. Holt, M. Maple.s, and J. C. Wingfield. 2006. Corticosterone is not correlated with nest departure in Snowy Owl chicks (Nyctea scandiaca). General and Comparative Endocrinology 149:119-123. Romero, L. M. and L. Remage-Healey. 2000. Daily and seasonal variation in response to stress in captive starlings (Sturnus vulgaris): corticosterone. General and Comparative Endocrinology 1 19:52-59. Romero, L. M., D. Strochlic, and J. C. Wingfield. 2005. Corticosterone inhibits feather growth: potential mechanism explaining seasonal down regulation of corticosterone during molt. Comparative Biochemistry and Physiology — Part A: Molecular & Integrative Physiology 142:65-73. Roulin, a., B. Almasi, a. Rossi-Pedruzzi, a. L. Ducrest, K. Wakamatsu, I. Miksik, j. D. Blount, S. Jenni-Eiermann, and L. Jenni. 2008. Corticoste- rone mediates the condition-dependent component of melanin-based coloration. Animal Behaviour 75: 1351-1358. Sapolsky, R. M. and M. j. Meaney. 1986. Maturation of the adrenocortical stress response: neuroendocrine control mechanisms and the stress hyporesponsive period. Brain Research Reviews 1 1:65-76. Sapolsky, R. M., L. M. Romero, and A. U. Munck. 2000. How do glucocorticoids influence stress responses? Integrating permissive, suppressive, stimulatory, and preparative actions. Endocrine Reviews 21:55-89. Schwabl, H. 1999. Developmental changes and among- sibling variation of corticosterone levels in an altricial avian species. General and Comparative Endocrinol- ogy 116:403^08. Sims, C. G. and R. L. Holberton. 2000. Development of the corticosterone stress response in young Northern Mockingbirds (Mimus polyglottos). General and Com- parative Endocrinology 1 19:193-201. Wada, H., T. P. Hahn, and C. W. Breuner. 2007. Development of stress reactivity in White-crowned Sparrow nestlings: total corticosterone response in- creases with age, while free corticosterone response remains low. General and Comparative Endocrinology 150:405-413. Walker, B. G., J. C. Wingfield, and P. D. Boersma. 2005. Age and food deprivation affects expression of the glucocorticosteroid stress response in Magellanic Penguin (Spheniscus magellanicus) chicks. Physiolog- ical and Biochemical Zoology 78:78-89. Wentworth, B. C. and M. O. Hussein. 1985. Serum corticosterone levels in embryos, newly hatched, and young turkey poults. Poultry Science 64:2195-2201. Wingfield, J. C. 1994. Modulation of the adrenocortical response to stress in birds. Pages 520-528 in Perspec- tives in comparative endocrinology (K. G. Davey, R. E. Peter, and S. S. Tobe, Editors). National Research Council of Canada, Ottawa, Ontario, Canada. Wingfield, J. C., C. M. Vleck, and M. C. Moore. 1992. Seasonal changes of the adrenocortical response to stress in birds of the Sonoran Desert. Journal of Experimental Biology 264:419-^28. Short Communications The Wilson Journal of Ornithology 122(4):767-77l, 2010 Composition Dynamics of an Avian Assemblage Fritz L. Knopf ABSTRACT. — I summarize the 10-year composition dynamics of a riparian avifauna on the Arapaho National Wildlife Refuge, Colorado, USA. Only 15 of the 89 breeding bird species (19,393 individuals) were present each of the 10 years whereas 25 species were only recorded in a single year. Number of species in any given year ranged from 30 to 57 (34 to 64% of the 10- year cumulative species total). The number of species detected >1 and <10 years represented over half of all species with no pattern of either increasing or decreasing across years of study. Short-term definitions of species richness preclude the ability to define the major dynamic within an avian assemblage — species that use the site an intermediate number of years. Understanding the ecological prerequisites for these species should provide greater insight into current physical habitat and climate changes occurring on-site. Received 16 February 2010. Accepted 24 June 2010. Ornithologists have historically inventoried a local avifauna to define the biological significance of an area: the more species present, the presumed greater ecological diversity and conservation significance of the locale. The number of species present is often referred to as species richness, or alpha diversity when viewed from broader spatial scales (Knopf and Samson 1994). Ecological studies have used species richness to interpret avian responses to changes in landscapes resulting from either natural processes or management practices (Conroy and Noon 1996, Brand et al. 2008, Johnson et al. 2009). Annual variation in species richness and relative species composition define conservation significance of an area (Fleishman et al. 2006). However, among-year variation in species com- position is generally ignored and the number of species recorded across the 2-4 years of most studies is often presented as a single number to define conservation significance. Species immi- gration and emigration are assumed to be equal in ‘ U.S. Fish and Wildlife Service, National Ecology Research Center, 1300 Blue Spruce Drive, Fort Collins, CO 80521, USA. ^Current address: 713 Boulder Circle, Fort Collins, CO 80524, USA; e-mail: FLKnopf@Yahoo.com such studies, although data addressing this assumption are notably lacking. I noted that species composition within an avian assemblage varied annually while species richness remained comparatively stable during a 10-year study of the breeding-bird assemblage in a shrub willow {Salix spp.) association along the Illinois River in the Arapaho National Wildlife Refuge, 1980-1989. The patterning of composi- tion change was much more dynamic than the standard assumption that individuals of some species colonize an area as individuals represent- ing other species leave. I briefly summarize the 10-year composition dynamics of avifaunal change with the intent to discourage use of species richness derived from short-term studies in assessments of conservation significance or efficacy of management actions at local sites. METHODS My coworkers and I studied the breeding bird assemblage along the Illinois River floodplain within the Arapaho National Wildlife Refuge (ANWR), Jackson County, Colorado, USA, from 1980 through 1989. The refuge is at an elevation of 2,485-2,516 m, within an intermountain glacial basin of upland vegetation dominated by sage- brush (Artemisia spp.). The riparian shrub assem- blage comprises species of willows (Salix bebbi- ana, S. caudata, S. exigua, S. geyeriana, S. monticola, S. planifolia, S. pseudocordata, and 5. wolfii) that grow to heights of 6 m (Cannon and Knopf 1984). Photographs of the study area are in Knopf et al. (1988: 281) and Sedgwick and Knopf (1992:724). We inventoried the breeding bird assemblage annually in the latter half of June, well after spring migrants had left ANWR. Each year, two observers each surveyed bird species and individ- uals at 200 (185 in 1980) points spaced at 100-m intervals (to reduce visual overlap between survey points) along each side and at random distances perpendicular to the river bank. The survey was conducted from the stream bank where the willow stand was less than 20 m wide. The radius of the 767 768 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4. December 2010 TABLE 1. Number of bird species recorded on 400, 10-min survey plots along the Illinois River floodplain on the Arapaho National Wildlife Refuge, Colorado, 1980- 1989. Year No. species No. individuals 1980“ 30 1,079 1981 35 1,231 1982 50 1,867 1983 45 1,841 1984 57 2,309 1985 45 1,866 1986 45 1,955 1987 43 2,154 1988 42 2,225 1989 57 2,866 1980-1989 89 19,393 ‘‘ Sampling effort was 370 plots in 1980. survey circle was not restricted to allow detection of larger, wary species. Avian surveys were conducted from one-half hour before sunrise until 1000 hrs MDT each morning except in periods of inclement weather. An observer moved to each survey point, waited 1 min and then recorded the species of each bird detected within the following 10-min interval. Each observer participated in multiple years of the surveys. RESULTS We detected 1,079-2,866 individuals represent- ing 30-57 avian species during surveys of the Illinois River floodplain in any given year, 1980 through 1989 (Table 1). Cumulative totals of 19,393 individuals representing 89 species were recorded across the 10-year period. The number of species recorded in any specific year ranged from 34 to 64% of the collective number of species recorded across the 10-year period. Detection of a given species in the floodplain also varied among years (Table 2). Only 15 (17%) of the cumulative 89 species were seen in all 10 years. Many more, 25 (28%), species were seen in only one of the 10 years. The number of detections within a species, ranged from >4,000 Yellow Warblers {Dendroica petechia) (Buckland et al. 1993) to a single record for many species that were only recorded in 1 year. Over half (49 of 89, 55%) of the species recorded on ANWR were present > 1 but <10 years. The number of species detected in an intermediate number of years showed neither a clear pattern of increasing nor decreasing with increasing years of study. DISCUSSION Ten of the 89 species present within the floodplain were waterfowl reflecting the manage- ment objective of ANWR to enhance breeding habitats for waterfowl by promoting taller herba- ceous vegetation to encourage nesting. Six raptor species foraged within the floodplain although most nested off the study area. Most cavity- nesting species did not have on-site nest sub- strates, but did forage as a part of the breeding TABLE 2. Number of years a species was recorded on 400. 10-min survey plots along the Illinois River floodplain on the Arapaho National Wildlife Refuge, Colorado. 1980-1989. No. years- No. species Specie.s'’ 10 15 AMRO, BBMA, BCCH, BHCO, BRBL, GWTE, LISP, RWBL, SASP, SOSP, WeSP, WEME, WIFE, WISN, YEWA 9 7 AMGO, GADW, HOWR, KILL, MALL, YHBL, WIPH 8 3 AMWI, BCNH, MODO 7 6 CITE, COGR, CONI, NOEL, SPSA, SWHA 6 8 BWTE, AMCR, COYE, FOSP, LESC, NOSH. TRSW, VEER 5 4 BASW, BRSP, NOPI, SORA 4 6 AMCO, GBHE, GTTO, NOHA, SEOW. WIWA 3 9 AMKE, BEKI, DUEL, GHOW, PISI, REDH, SWTH, VESP, WAVI 2 6 CAGO, CLSW, GRCA, MAWR. RCKI, GRSG 1 25 AMBI, BTHU, BUOR, CANV, CHSP, COFL, CORA, DOWO, EAKI, EUST, FOTE, GRHE, GRYE, HOLA, INBU, LABU, LEOW, OCWA, RNPH, RTHA, SATH, SSHA, VIRA, WEKl, YBCU Sampling etTurl was 370 plots in 1980. Species codes in the Appendix. SHORT COMMUNICATIONS 769 bird assemblage, as did Green-tailed Towhee {Pipilo chloruriis) and Sage Thrasher (Oreos- coptes momanits) that nested in the sagebrush uplands. Greater Sage-Grouse [Centrocerciis uro- phasianiis) were not known to nest within the floodplain, but occasionally moved broods into the riparian vegetation. Collectively, these species contributed a meaningful component to the species richness and conservation significance of the site, although the floodplain only provided one component of the habitat required for breeding. Earlier studies on ANWR included both basic studies of single species, i.e., Yellow Warbler (Knopf and Sedgwick 1992, Howe and Knopf 2000) and Willow Flycatcher (Enipidonax traillii) (Sedgwick and Knopf 1988, 1989, 1992), and applied studies of grazing-response guilds (Knopf et al. 1988, Stanley and Knopf 2002). Those studies focused on species that occurred at densities adequate to insure sample sizes to support statistically rigorous analyses — species that were present all 10 years. Their certain presence revealed these species were more tolerant of ecologic or climatic change occumng locally within the study area across the decade. Extending avian surveys on ANWR another 10 years would result in no increase in the number of species with >95% probability of appearance each year. Twenty-five species were present only one of 10 years. These species were likely non-breeders as spring migrants had left the refuge before the late June surveys. Some individuals may have been post-breeding dispersers from other locales, whereas others were stragglers as they were well outside documented breeding ranges. Extending avian surveys on the ANWR another 10 years would increase the cumulative list of avian species for the site to 100 species or more across a 20-year period. This latter component of species richness is not ‘rich’ at all, but peripheral to any meaningful definition of the breeding bird assem- blage or management actions on the ground. Interpreting or monitoring the significance of an area for avian conservation using a list of species recorded during a short-term study is problematic in that one component of the assemblage (species present every year) is eco- logically tolerant of subtle environmental changes and another component is only present inciden- tally. Declines or increases in numbers of both common and rare/accidental species would only be expected following relatively catastrophic changes in the landscape. The larger problem, however, is that short-term studies are unable to distinguish “incidental” species from those that periodically use the site during the breeding season. Research investigations focusing on the ecological prerequisites for .species that use an area intermittently should provide greater insight into more subtle physical and climate changes that drive species habitats within a local avian assemblage. The ability to identify these species already exists where standardized protocols to monitor breeding birds have been in place for multiple years. ACKNOWLEDGMENTS 1 am especially grateful to R. W. Cannon, J. F. Ellis, W. H. Howe, T. E. Olson, and J. A. Sedgwick for conducting avian surveys. E. C. Patten. Project Leader/Manager, graciously provided open access to Arapaho National Wildlife Refuge. This work was completed when I was a research wildlife biologist with the U. S. Fish and Wildlife Service, and I am grateful for the intellectual support of my administrators and academic peers at the National Ecology Research Center during the 1980s. LITERATURE CITED Brand, L. A., G. C. White, and B. R. Noon. 2008. Eactors influencing species richness and community composi- tion of breeding birds in a desert riparian corridor. Condor 1 10:199-210. Buckland, S. T., D. R. Anderson, K. P. Burnham, and J. L. Laake. 1993. Distance sampling. Chapman and Hall, London, United Kingdom. Cannon, R. W. and F. L. Knopf. 1984. Species composition of a willow community relative to seasonal grazing histories in Colorado. Southwestern Naturalist 29:234-237. Conroy, M. J. and B, R. Noon. 1996. Mapping of species richness for conservation of biological diversity: conceptual and methodological issues. Ecological Applications 6:763-773. Fleishman, E., R. F. Noss, and B. R. Noon. 2006. Utility and limitations of species richness metrics for conser- vation planning. Ecological Indicators 6:54.3-553. Howe, W. H. and F. L. Knopf. 2000. The role of vegetation in cowbird parasitism of Yellow Warblers. Pages 200-203 in Ecology and management of cowbirds and their hosts (J. N. M, Smith. T. L. Cook. S. 1, Rothstein. S. K. Robinson, and S. G. Sealy, Editors). University of Texas Press. Austin. USA. Johnson, T. N., R. D. Applegate. D. E. Hoover. P. S. Gipson, and B, K. Sandercock. 2009. Evaluating avian community dynamics in restored riparian habitats with mark-recapture models. Wilson Journal of Ornithology 121:22-40. Knopf, F. L. and F. B. Samson. 1994. Conservation of 770 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 4, December 2010 avian diversity in riparian corridors. Conservation Biology 8:669-676. Knopf, F. L. and J. A. Sedgwick. 1992. An experimental study of nest site selection by Yellow Warblers. Condor 94:734-742. Knopf, F. L., J. A. Sedgwick, and R. W. Cannon. 1988. Guild structure of a riparian avifauna relative to seasonal cattle grazing. Journal of Wildlife Manage- ment 52:280-290. Sedgwick, J. A. and F. L. Knopf. 1988. A high incidence of Brown-headed Cowbird parasitism of Willow Flycatchers in Colorado. Condor 90:253-256. Sedgwick, J. A. and F. L. Knopf. 1989. Region-wide polygyny in Willow Flycatchers. Condor 91:473^75. Sedgwick, J. A. and F. L. Knopf. 1992. Describing Willow Flycatcher habitats: scale and gender perspec- tives. Condor 94:720-733. Stanley, T. R. and F. L. Knopf. 2002. Vegetation and avian responses to grazing in a shrub-willow flood- plain. Conservation Biology 16:225-231. APPENDIX. Alphabetical list of species codes (Table 2). Code Species AMBI American Bittern (Botaurus lentiginosus) AMCO American Coot (Fulica americana) AMCR American Crow {Corvus brachyrhynchos) MAKE American Kestrel {Falco sparverius) AMGO American Goldfinch {Spinus tristis) AMRO American Robin (Turdus migratorius) AMWl American Wigeon (Anas americana) BASW Bam Swallow (Hirundo rustica) BBMA Black-billed Magpie (Pica hudsonia) BCCH Black-capped Chickadee (Poecile atricapillus) BCNH Black-crowned Night-Heron (Nycticorax nycticorax) BEKl Belted Kingfisher (Megaceryle alcyon) BHCO Brown-headed Cowbird (Molothrus ater) BRBL Brewer’s Blackbird (Euphagus cyanocephalus) BRSP Brewer’s Sparrow (Spizella breweri) BTHU Broad-tailed Hummingbird (Selasphorus platycercus) BUOR Bullock’s Oriole (Icterus bullockii) BWTE Blue-winged Teal (Anas discors) CAGO Canada Goose (Branta canadensis) CANV Canvasback (Aythya valisineria) CHSP Chipping Sparrow (Spizella passerina) CITE Cinnamon Teal (Anas cyanoptera) CLSW Cliff Swallow (Petrochelidon pyrrhonota) COFL Cordilleran Flycatcher (Empidonax occidentalis) COGR Common Crackle (Quiscalus quiscula) CONI Common Nighthawk (Chordeiles minor) CORA Common Raven (Corvus corax) COYE Common Yellowthroat (Geothlypis trichas) DOWO Downy Woodpecker (Picoides pubescens) DUFL Dusky Flycatcher (Empidonax oberholseri) EAKl Eastern Kingbird (Tyrannus tyrannus) EUST European Starling (Sturnus vulgaris) FOSP Fox Sparrow (Passerella iliaca) FOTE Forster’s Tern (Sterna forsteri) GADW Gadwall (Anas strepera) GBHE Great Blue Heron (Ardea herodias) GHOW Great Homed Owl (Bubo virginianus) GRCA Gray Catbird (Dumetella carolinensis) GRHE Green Heron (Butorides virescens) GRYE Greater Yellowlegs (Tringa melanoleuca) GTTO Green-tailed Towhee (Pipilo chlorurus) GWTE Green-winged Teal (Anas crecca) HOLA Horned Lark (Eremophila alpestris) HOWR House Wren (Troglodytes aedon) SHORT COMMUNICATIONS 771 APPENDIX. Continued. Code Species INBU Indigo Bunting {Passerimi cyanea) KILL Killdeer (Charadrius vocifents) LABU Lazuli Bunting (Passerimi amoena) LEOW Long-eared Owl (Asia otus) LESC Lesser Scaup (Aythya ajfinis) LISP Lincoln’s Sparrow (Melospiza lincolnii) MALL Mallard (Anas platyrhynchos) MAWA MacGillivray’s Warbler (Oporornis tolmiei) MODO Mourning Dove (Zenaida macroura) NOEL Northern Flicker (Colaptes auratus) NOHA Northern Harrier (Circus cyaneus) NOPI Northern Pintail (Anas acuta) NOSH Northern Shoveler (A. clypeata) OCWA Orange-crowned Warbler (Vermivora celata) PISI Pine Siskin (Spinus pinus) RCKI Ruby-crowned Kinglet (Regulus calendula) REDH Redhead (Aythya americana) RNPH Red-necked Phalarope (Phalaropus lobatus) RTHA Red-tailed Hawk (Buteo jamaicensis) RWBL Red-winged Blackbird (Agelaius phoeniceus) GRSG Greater Sage-Grouse (Centrocercus urophasianus) SASP Savannah Sparrow (Passerculus sandwichensis) SATH Sage Thrasher (Oreoscoptes montanus) SEOW Short-eared Owl (Asia flammeus) SORA Sora (Porzana Carolina) SOSP Song Sparrow (Melospiza melodia) SPSA Spotted Sandpiper (Actitis macularius) SSHA Sharp-shinned Hawk (Accipiter striatus) SWHA Swainson’s Hawk (Buteo swainsoni) SWTH Swainson’s Thrush (Catharus ustulatus) TRSW Tree Swallow (Tachycineta bicolor) VEER Veery (Catharus fuscescens) VESP Vesper Sparrow (Pooecetes gramineus) VIRA Virginia Rail (Rallus limicola) WAVI Warbling Vireo (Vireo gilvus) WCSP White-crowned Sparrow (Zonotrichia leucophiys) WEKI Western Kingbird (Tyrannus verticalis) V/EME Western Meadowlark (Sturnella neglecta) WIFE Willow Flycatcher (Empidonax traillii) WIPH Wilson’s Phalarope (Phalaropus tricolor) WISN Wilson’s Snipe (Gallinago delicata) WIWA Wilson’s Warbler (Wilsonia pusilla) YBCU Yellow-billed Cuckoo (Coccyz.us americanus) YEWA Yellow Warbler (Dendroica petechia) YHBL Yellow-headed Blackbird (Xanocephalus xanocephalus) Ill THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 4. December 2010 The Wilson Journal of Ornithology 122(4):772-776, 2010 A Digital Spot-mapping Method for Avian Field Studies Kevin E. Jablonski,' "^ Stacy A. McNulty,^ and Matthew D. Schlesinger^ ABSTRACT. — Avian mapping, also known as spot mapping or territory mapping, is a breeding season bird- survey technique that traditionally uses paper maps on which locations of birds are recorded. This method is often considered the most accurate in yielding a density, but has been criticized as being inefficient, time consuming, and inexact. We describe a novel digital- mapping method, incorporating a hand-held computer and high-accuracy global positioning system receiver (GPS), used in an ongoing boreal birds study. Digital mapping surpassed our expectations as to efficiency, flexibility, and work flow. We expect this method will become increasingly useful in many types of field studies, especially as costs decrease (currently ~$2,100 for the field receiver used). Received 4 January 2010. Accepted 23 June 2010. Avian mapping, also known as spot mapping or temtory mapping, is a widely used bird-survey technique in which an observer attempts to record the exact locations of individual birds within a defined study plot. The method emerged in the early 20th century as part of a naturalistic, observational approach to field biology, and was first described by Williams (1936:376) who wrote that “by the use of weekly maps ... it is possible to build up a series . . . covering the nesting season, which will show concentrations of records about certain localities.” Subsequent re.searchers ex- panded on Williams’s idea and offered methods of statistical Interpretation of mapping data (Ken- deigh 1944, Pough 1947, Hall 1964). However, field methods for mapping in ornithology have remained essentially static since Robbins (1970) offered an international standard (Verner 1985, Manuwal and Carey 1991, Bibby et al. 2000). The mapping method most commonly used involves establishment of a physical grid, 8-20 ha in size, using flagging or stakes to mark gridline ' SUNY ESF-244 Illick Hall. I Fore.stry Drive, Syracu.se, NY 13210. USA. ^Adirondack Ecological Center, SUNY ESF, 6312 State Route 28N, Newcomb, NY 128.32, USA. ’New York Natural Heritage Program. 62.3 Broadway, 3th Floor, Albany. NY 12233, USA. ’Corresponding author; e-mail: kevinJablonski@gmail.com intersections. An observer walks slowly along the gridlines noting the location of observed birds on a paper map using a set of symbols to indicate species, M/F, and behavior. The observer com- piles the maps from each visit after 5-10 visits, creating one map per species. This enables delineation of individual birds’ teiritories and estimation of breeding-bird density and diversity, as well as habitat structure and composition at bird locations or territories (Robbins 1970, Manuwal and Carey 1991, Ralph et al. 1993, Bibby et al. 2000). The main limitations of the paper mapping field method as currently applied concern two areas: (1) the large time investment required for collecting and compiling data, and (2) the chances of making errors during data recording and interpretation. It can be time-consuming to create a physical grid using flagging and stakes over 8- 20 ha, especially in thick vegetation or rough terrain. Compiling several maps for each bird species requires hours of data preparation and invites human eiTor in interpreting and transfer- ring data from one map to another. Recording detailed bird locations, behavior, and movements on maps scaled for practical field use can be inaccurate on paper maps (Verner and Milne 1990). We studied habitat associations of lowland boreal birds in Adirondack Park, New York, USA with 10-min variable circular plot point counts in 2008. However, we decided that a more location- explicit approach was wairanted for subsequent data collection after the initial field season. Our study was focused on a group of rare, cryptic birds in wet and often densely vegetated terrain. Thus, we sought a method that would provide more llexibility in detecting rare birds and allow us to spend more time at high-diversity sites than allowed by fixed-point counts. These require- ments led us to use mapping as a technique for the 2009 field sea.son. We suspected that paper-based mapping was not sufficiently efficient, accurate, or llexible for our needs and would be laborious to incorporate into our existing Geographic Infor- mation System (GIS). We (1) de.scribe a novel SHORT COMMUNICATIONS 773 TABLE 1. Target species of lowland boreal bird study. Spring Pond Bog, Adirondack Park, New York. Species Common name Picoides circticiis Black-backed Woodpecker Contopus cooperi Olive-sided Flycatcher Empidomix flaviventris Yellow-bellied Flycatcher Pehsoreus canadensis Gray Jay Poecile luidsoniciis Boreal Chickadee Dendroica tigrina Cape May Warbler D. palmanim Palm Warbler D. castanea Bay-breasted Warbler Melospiza lincolnii Lincoln’s Sparrow Eitphagits carol inLis Rusty Blackbird digital-mapping method that incoiporates a high- accuracy global positioning system receiver (GPS), a hand-held computer, and a GIS-derived grid, and (2) discuss the advantages over tradi- tional paper mapping as well as potential drawbacks. METHODS Our study area was the Spring Pond Bog Preserve, owned by The Nature Conservancy, in northern Adirondack Park, New York (44° 22' 23" N, 74° 30' 21" W). We applied the digital- mapping method to our study of habitat associa- tions of lowland boreal birds, a suite of 10 species that occur in the scattered lowland boreal habitats of the Adirondacks (Table 1 ). We established five 9-ha plots and visited each plot five times from 28 May to 8 July 2009, starting 15 min after sunrise. We created a GIS prior to entering the field using ArcGis Desktop 9.2 (ESRI 2007) that contained orthographic aerial photographs, wet- lands, roads, trails, and other geospatial data layers. We created five 300- X 300-m digital plots centered on point-count locations where we had recorded high target bird diversity in 2008. We divided each of these plots into a grid of 900 10- X 10-m .squares (Fig. 1 ), and created a point layer for recording bird observations. We created drop- down menus for attributes such as date, species, seen/unseen, M/F, and behavior to facilitate data entry in the field (Fig. 2). We loaded the.se data layers into ArcPad 7.1 on a Trimble Nomad handheld computer. The Nomad was equipped with a Wide-Area Augmentation System (WAAS)-enabled Garmin lOX Bluetooth GPS receiver. We conducted plot visits in much the same way as paper mapping. However, we used the Nomad along with a compass for navigation and data recording. We selected a random starting point and path through each plot by spinning a compass, and then proceeded at a calm, quiet pace through the plot using the Nomad’s “Tracklog” feature to en.sure that we followed grid lines so that we came within 20 m of every grid point. A full 9-ha plot visit lasted 90-120 min. No flagging or .stakes were necessary to mark the plot or grid. When we detected an individual of one of the target species, we ascertained its exact location and noted behavioral data related to territoriality or nesting activity. We then walked to the bird’s location and, after ensuring the chosen GPS accuracy measure. Positional Dilution of Preci- sion (PDOP), was below the pre-established threshold of 3.5, created a new data point using the Nomad touch screen. We decided on a PDOP threshold of 3.5 based on recommendations in the literature (i.e., Kennedy 2002) and expected accuracy under variable field conditions (from open, boggy peatlands with low relief to dense conifer Vv'etlands). We used the drop-down menus in creating this point (Fig. 2) to record attributes of the detection. We changed the map view according to navigational or data-recording needs throughout the visit. For example, when locating an individual bird, we zoomed to a scale of ~ 1:400 (in which a grid square is 2.54 cm wide), whereas while navigating we used a scale of 1:1,000-2,000. For comparison, a typical paper map with a 9-ha grid would be scaled to 1:1,600 (in which a grid square is 0.64 cm wide). Each week we backed up the data set to our desktop GIS. RESUFTS We recorded 153 individual detections of nine of the 10 target lowland boreal bird species in 25 plot visits. The number of target individuals detected per visit ranged from 3.8 to 10.2 (.v = 6.12, a = 2.44) while total target species recorded per grid ranged from 2 to 7 (.v = 5.0. a = 1.87). We compiled species detection maps for all sites after uploading the Nomad data to a desktop GIS (Fig. I). DISCUSSION The digital spot-mapping method was easy to use and work flow efficiency exceeded our expectations for spatial and attribute accuracy, and data acquisition and storage speed. Drop- down menus made field data entry fast and 774 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 4. December 2010 f±llSL^\ Legend A RUBL YPWA FIG. I. Example of 300- X 300-m (9-ha) digital grid with accumulated detections of Rusty Blackbird (RUBL) and Palm Warbler (YPWA) at site SPB08, Spring Pond Bog, Adirondacks, New York. unambiguous, unlike the complex system of paper map symbols an observer must correctly interpret to record breeding bird attributes. Dense or diverse bird populations can render paper map- ping time-consuming and confusing, if not unusable. Digital mapping has no functional upper limit, as existing data points do not preclude recording additional observations at the same location. The map can be scaled in the field to show more or less detail, and dense bird populations can be recorded without clutter. Errors of translation can be substantially reduced, as there is no need to interpret and compile multiple paper maps, especially il that compila- tion occurs after the details of field observations fade from memory. The Bluetooth GPS quickly and consistently obtained accurate locations, keeping us on the path of travel with no need to mark the plot in advance. The Nomad straps safely to one hand, enabling the observer to enter data, u.se binocu- lars, and navigate over obstacles and through wetlands while maintaining the travel path. Battery power lasted for —20 hrs of field use. We kept the electronic devices charged with an AC adapter for our vehicle. We did not directly compare the two methods but, in our experience, digital-mapping data can mirror data acquired by typical paper mapping studies. A review of recent studies that used the traditional mapping method (Fink et al. 2006, Toms et al. 2006, Villard et al. 2007, Romdal and Rahbek 2008, Gale et al. 2009) indicates that each could have incorporated digital mapping with little or no change to other aspects of the study while achieving a probable gain in efficiency in data acquisition and processing. Digital mapping can be also adapted to record unlimited additional data point attributes, whereas paper mapping would require additional data sheets for any additional data. We encountered some problems with the digital-mapping method related to the specific combination of device-software interface. ArcPad SHORT COMMUNICATIONS 775 ^ ArcPad Terr detect ♦•I Date 7/2/09 ▼ Site SPB04 Grid_no 549 Species YBFL ▼ Seen q N Y ArcPad Terr detect Sex Terri Terr2 Terr3 Nestl Nest2 M ▼ Singing ▼ 'W J Carrying Food Carrying Matter Copulation Flush Nest Found Page 1 Page 2 ^ Page 3 ^ Page 1 Page 2 ^ Page 3 1 i ► FIG. 2. Screenshot of Trimble Nomad running ArcPad 7.1 with drop-down menus used to enter data in lowland boreal bird mapping study. Spring Pond Bog, Adirondacks, New York. 7.1 is cumbersome to use in the field without access to an ArcGIS-equipped computer because changes to menus, data layers, or data sets must be made using ArcGIS. The operating system of the Nomad (Windows Mobile 6) hindered data processing, especially in customization and add- ing functionality. The interface between ArcGIS and ArcPad was limited, requiring us to download GIS shapefiles using Windows Sync rather than ArcGIS. A second limitation of digital mapping is cost of implementation. The Trimble Nomad-Garmin GPS combination costs ~$2,100 U.S., which does not include the cost of the ESRI software license or desktop computer. However, as recent devel- opments with GPS-enabled personal devices such as the Apple iPhone indicate, a combination of improved and cheaper technology, as well as free, customizable, open-source GIS software, is likely to substantially lower this cost in the near future. The digital spot-mapping method is useful for scientific studies requiring accurate bird locations and may also be suitable for citizen science applications (Silvertown 2009). Many birding enthusiasts, just as they have long been equipped with binoculars and paper bird guides, now carry smart mobile phones with electronic bird guides from groups such as National Audubon Soci- ety. For example, the Apple iPhone application BirdsEye (http://ebird.org/content/ebird/news/ BirdsEye) provides real-time access to the Cornell Laboratory of Ornithology’s eBird data base, which contains millions of citizen-reported rec- ords showing where and when birds were observed in North America. The capacity to record information about birds with a digital spot-mapping application combined with a GPS, camera, and recorder to capture bird photographs and songs for later identification may enhance information gathering and minimize interpretation errors. The potential of digital spot-mapping should improve as hand-held computing also improves. The mapping device of the near future will likely be a small, affordable, hand-held, touch-screen computer, equipped with a GPS unit capable of sub-meter accuracy, simple open-source GIS software loaded with (or accessing wirelessly) maps and high-resolution aerial photographs, and capability for syncing with a central storage server/computer for data transfer. The potential applications of such a device are myriad, and 776 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4, December 2010 digital mapping can be expected to revolutionize aspects of field ornithology. ACKNOWLEDGMENTS We thank The Nature Conservancy for property access, and the Adirondack Park Agency and NYS GIS Clearing- house for data layers. The manuscript was improved with comments from C. J. Ralph and two anonymous reviewers. Funding was provided by SUNY ESF. the Edna Bailey Sussman Foundation, the Northern New York Audubon Society, and the Wildlife Conservation Society. LITERATURE CITED Bibby, C. J., N. D. Burgess, D. A. Hill, and S. Mustoe. 2000. Bird census techniques. Academic Press, London, United Kingdom. Environmental Systems Research Institute (ESRI). 2007. ArcGIS Desktop, Version 9.2. Environmental Systems Research Institute, Redlands, California, USA. Fink, A. D., F. R. Thompson III, and A. A. Tudor. 2006. Songbird use of regenerating forest, glade, and edge habitat types. Journal of Wildlife Management 70:180-188. Gale, G. A., P. D, Round, A. J. Pierce, S. Nimnuan, A. Pattanavibool, and W. Y. Brockelman. 2009. A field test of distance sampling methods for a tropical forest bird community. Auk 126:439-448. Hall, G. A. 1964. Breeding bird census — why and how. Audubon Field Notes 18:413^16. Kendeigh, S. C. 1944. Measurement of bird populations. Ecological Monographs 14:67-106. Kennedy, M. 2002. The Global Positioning System and GIS. Taylor and Erancis Inc., New York, USA. Manuwal, D. a. and a. B. Carey. 1991. Methods for measuring populations of small, diurnal forest birds. USDA, Eorest Service, General Technical Report PNW-GTR-278. Portland, Oregon, USA. PouGH, R. H. 1947. How to take a breeding bird census. Audubon Magazine 49:290-297. Ralph, C. J., G. R. Geupel, P. Pyle, T. E. Martin, and D. F. Desante. 1993. Handbook of field methods for monitoring landbirds. USDA, Forest Service, General Technical Report PSW-GTR-144. Albany, California, USA. Robbins, C. S. 1970. Recommendations for an international standard for a mapping method on bird census work. Audubon Field Notes 24:723-726. Romdal, T. S. and C. Rahbek. 2008. Elevational zonation of afrotropical forest bird communities along a homogeneous forest gradient. Journal of Biogeography 36:327-336. SiLVERTOWN, J. 2009. A new dawn for citizen science. Trends in Ecology and Evolution 24:467^71. Toms, J. D., E. K. A. Schmiegelow, S. J. Hannon, and M. ViLLARD. 2006. Are point counts of boreal songbirds reliable proxies for more intensive abundance estima- tors? Auk 123:438-154. Verner, j. 1985. Assessment of counting techniques. Current Ornithology 2:247-302. Verner, J. and K. A. Milne. 1990. Analyst and observer variability in density estimates from spot mapping. Condor 92:313-325. Villard, M., E. K. a. Schmiegelow, and M. K. Trzcinski. 2007. Short-term response of forest birds to experimental clearcut edges. Auk 124:828-840. Williams, A. B. 1936. The composition and dynamics of a beech-maple climax community. Ecological Mono- graphs 6:317—108. SNORT COMMUNICATIONS 111 The Wilson Joiinuil of Ornithology 122(4):777-780, 2010 Variation of Type B Song in the Endangered Golden-Cheeked Warbler ( Denclroica chrysoparia ) Wendy J. Leonard,' Jayne Neal,' and Rama Ratnarn^'^ ABSTRACT. — The Golden-cheeked Warbler (Den- droica chrysoparia), a member of the Parulidae, uses a two-category song system referred to as Type A and Type B. However, songs recorded from adult males in Bexar County, Texas, USA, did not resemble the typical A or B songs reported for this species. We report an analysis of these atypical songs from six adult males. Quantitative features of the measured songs and syllable structure suggest this atypical song is more comparable to a B song than an A song with some modifications. The atypical song has dropped a terminal syllable of the known B song, and selectively modified one of the syllables to include frequency modulations (up and down slurs). Further, the known B song was not heard in the investigated sites. We conclude the atypical song is a variant of the B song. Received 10 December 2009. Accepted 30 June 2010. The Golden-cheeked Warbler (Dendroica chry- sopctria) (GCWA), a member of the Parulidae, has a two-category song system, referred to as “first category” and “second category” (Spector 1992). “First category” songs, used by unmated males during daylight, are used in disputes with other males and use declines after mating (Spector 1992). These songs, which can be simple and used near females, are usually sung at low rates (Bolsinger 2000). The “second category” song is used during the dawn chorus or before sunrise, and is more common later in the breeding season (Spector 1992). This song is more complex and faster than the first category song and is preceded with short call notes (chips) (Pulich 1976, Spector 1992, Bolsinger 2000). It can also be more variable than the first category song (Bolsinger 2000). Bolsinger (1997, 2000) uses the more common terminology ‘A song’ (for first category songs) and ‘B song’ (for second category songs). One of the authors (JN) repeatedly heard a GCWA song in spring and summer of 2008 and 2009 that was not a typical A or B song. These ' San Antonio Parks and Recreation Natural Areas, 21395 Milsa Road, San Antonio, TX 78256, USA. "Depailment of Biology, University of Texas at San Antonio, One UTSA Circle, San Antonio, TX 78249, USA. ^Corresponding author; e-mail: rama.ratnam@utsa.edu birds were visually identified and confirmed to be GCWA. Males sang these songs throughout the season, although they were heard more frequently later in the breeding season. It was not evident whether the song was Type A or B. We recorded song sequences in 2009 from six individuals at three sites in northern Bexar County, Texas, USA. Our goal was to quantita- tively compare the new song with reported songs for this species to correctly classify this song as an A or B song. METHODS Songs were recorded from adult male GCWAs on City of San Antonio, Texas properties. These sites included Grey Forest (29°37'39"N, 98°42' 4"W), a City Park (29° 38' 9" N, 98° 38' 24" W), and a site near Camp Bullis (29° 37' 12"N, 98° 35'24"W). We recorded over a distance of — 10 km in areas of deciduous woodlands and live oak/juniper {Quercus/Juniperus). Recordings were made using omnidirectional microphones (MKE-2, Sennheiser Electronic Corporation, Old Lyme, CT, USA) connected to a high resolution digital audio recorder (Model 722T, Sound Devices EEC, Reedsburg, WI, USA) that ampli- fied and digitized signals at 44.1 kSamples/sec, and stored them as WAV files in the built-in hard drive. Songs were analyzed using Audition (Adobe Systems Inc., San Jose, CA. USA) and MATLAB (The MathWorks Inc., Natick, MA, USA). A band pass filter (sixth order. Type II Chebychev, 2.5 to 10.5 kHz) was used to eliminate energy outside the frequency band of GCWA songs. Spectrograms were generated for each segment using a Fast Fourier Transform (lOO-msec segments, Hann windowed, 95% overlap). Individual songs were extracted from the sequences by visually scanning the spectro- grams. Songs were divided into syllables, and each .syllable was classified and given a lower-case letter designation. Quantitative analysis of songs followed Bolsinger (1997, 2000). Histograms 778 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4, December 2010 TABLE 1. Components of known GCWA and atypical GCWA songs. Control data (A song, B song) are from recordings deposited at Macaulay Library. Atypical songs are from 6 males in northern Bexar County, Texas recorded in May and June 2009. Values in parentheses are ranges. Song complexity Song rate Pause duration Number of chips Frequency of Duration (sec) (elements/song) (songs/min) (sec) preceding song modal intensity (Hz) A song' L9U1.6-L9) 3" 5.T (3.9-6.0) 10.2“ (1-3-13.9) 0.7“ (0-1) 5,354“ (4,580-5,980) B song' 2.0“(L8-2.3) 5'’ (4-6) 6.8“ (3. 1-9.1) 8.1“ (4.5-17.1) 5.1“ (0-14) 4,863“ (4,280-5,420) Atypical song (2009) 2.1 (L3-2.3) 5.8 (4-7) 5.8 (3.4-11.2) 9.0 (2.5-15.7) 3.5 (0-12) 4,675 (4,200-5,080) Data from recordings deposited at Macaulay Library, Cornell Laboratory of Ornithology. Audio record numbers: J. S. Bolsinger (A song birds = 4, B song birds = 3; circa 1994) 109301, 109338, 109340, 109734, 109736, 109749-109752, 109754, 109760, 109761; W. H. Gunn (B song bird = 1; c. 1977) 68956, G, A. Keller (A song bird = 1; circa 1986) 40560, 40561, T. A. Parker III (A song bird = 1; circa 1986) 38337, 38344, 38347, R. C. Stein {B song bird = 1; circa 1961) 9190- 9195. *’ Data from Bolsinger (1997:38, table 1.1). were constructed for the following parameters: (1) song complexity (defined as the number of syllables per song), (2) number of chips preceding a song, (3) song duration, (4) pause duration (between songs in a bout), (5) song rate, and (6) frequency of modal intensity (defined as the frequency component with the greatest energy). Pause durations exceeding the 90th percentile were excluded from histograms as they indicated gaps between bouts. A and B song recordings made between 1961 and 1994 were obtained from the Macaulay Library (Table 1). The recordings served as control data and were analyzed similarly for number of chips, song duration, pause duration, song rate, and frequency of modal intensity. The frequency of modal intensity was measured from the spectrum of a syllable rather than the entire song (as adopted by Bolsinger 1997). The syllable was selected so it had comparable duration and frequency bandwidth across the set of atypical and control songs. Statistical comparisons of parameters were performed among three groups: (1) the atypical 2009 recordings, (2) control A songs, and (3) control B songs. Song parameters were averaged over all songs for a given bird, and then averaged over the number of birds. Thus, sample sizes reflect the number of individual birds recorded, and not the total number of recorded songs. The goal was to ascertain if the atypical 2009 recordings were similar to the A song or the B song by comparing for differences in means of the measured parameters. All comparisons were performed using the Kruskal-Wallis test (Sokal and Rohlf 1995) to look for significant differences between means of the groups. A pair-wise multiple comparison test was used to examine which pairs of groups differed in their means. The multiple comparison procedures (with a Bonfer- roni correction) reported confidence intervals at a-level of 0.01. RESULTS Two hundred and fifty-nine songs were record- ed from six adult GCWA males from 17 May to 13 June 2009. Numbers of songs produced by each bird (in decreasing order) were: 102, 63, 37, 26, 17, and 14. A total of 161 control A songs was obtained from recordings of six birds (numbers of songs were 67, 24, 21, 19, 16, 14), and 1 15 control B songs were obtained from recordings of five birds (numbers of songs were 35, 33, 19, 18, 10). Representative control A and B songs (Fig. lA, B, respectively) were compared with atypical songs (2 examples shown in Fig. 1C, D). A visual comparison of the spectrograms suggests the atypical song more closely resembles the known B song than the A song. However, there are differences in the number of syllables and the structure of the song. The known B song has five syllables labeled a through e (Fig. IB) whereas the atypical song has only four syllables and lacks the terminal e syllable. Syllable a is markedly less intense in the atypical song and lacks the buzzy nature in the known B song. In some instances it was so faint as to be barely audible above the background noise. Syllable b is well-preserved and has been used to align the songs (Fig. IB, C, D), and syllable d has been modified in duration but is otherwise unchanged. The greatest modifi- cation occurs in syllable c which in the known B song is a complex of two tonal elements centered between 6 and 4 kHz (Fig. IB). This syllable (marked as c' in Fig. 1C, D) in the atypical song has been lengthened and includes complex frequency modulations (up and down slurs); it is more variable across individuals than other syllables (Fig. 1C, D). SHORT COMMUNICATIONS 779 A 7,6 6 2,6 B 7S 6 - 26 Song A (Bobinger 1097) d Song B (Botsinger 1997) * t) j e c 7.5 6 26 5 26 0.5 s«c FIG. 1. Song spectrograms of (A) known A song, and (B) known B song. These songs were recorded between 1961 and 1994 and served as control data. C and D are atypical songs reported in this study. The known B and atypical songs are aligned on the offset of syllable b. The similarities and differences between the known A, B, and atypical songs were quantified by analyzing song parameters (Table 1). Song complexity was similar for the atypical song and the known B song with the A song demonstrating less complexity. The A song is not normally preceded by chips and a comparison of the mean number of chips was possible only between the B song and the atypical song. Mean number of chips was significantly higher in the B song than the atypical song, but the number of chips produced had a similar range. Frequency of modal intensity was estimated from the d syllable for all songs. This syllable in the control A song was identified as the pre-terminal syllable (Fig. lA) as it most closely resembles the d syllable in the atypical and control B songs. The mean frequency of modal intensity for the known B song (4,803 Hz) and atypical song (4,675 Hz) were closer to one another than to the A song (5,354 Hz), but they all significantly differed from one another, as did mean song durations for all pairs. There were no significant differences between song rates of the atypical and A or B songs, and the pause duration between the atypical and B songs. All other paired comparisons were significantly dii'ferent. DISCUSSION The atypical GCWA songs recorded in 2009 in northern Bexar County are different from the B song reported for this species in a number of respects, but the spectrogram analysis, syllable similarity, and structure suggest this atypical song is likely to be a variation of the known B song (Fig. 1 ). Further, A songs are not usually preceded by chips or calls notes whereas the known B song and the atypical song were frequently preceded by chips. These findings suggest the atypical song is a B or “second category” song (Spector 1992) with a markedly modified c syllable, and lacking the terminal e syllable found in the control B song. The typical A song was heard infrequently during the field investigations, but was not recorded. The atypical B song was heard more frequently during the later part of the GCWA breeding season. This agrees with the description of A and B song use by Spector (1992). A or “first category” songs dominate early in the breeding season and then decline after pairing (Spector 1992). B or “second category” songs are more common later in the breeding season (Spector 1992). We note the typical B song was not heard from these six birds during field investigations. We propose the atypical song is the B song which has undergone variation over time. This has important implications for conservation. Accurate population estimates are essential for assessing the status of the federally listed endangered GCWA; these assessments often rely on auditory detections. Our study sites are in a region of Bexar County that is experiencing rapid urban develop- ment with clearing of oak and Ashe juniper iJuniperus ashei) woodlands necessary for breed- ing. Further investigations over several seasons are needed to identify the ecological significance and permanence of the song change. Are the changes in song structure temporary resulting from a loss of GCWA breeding habitat, a change in vegetation, or increased habitat fragmentation (Laiolo and Telia 2005)? Or, are the changes permanent resulting from changes in morphology (Derryberry 2009)7 Our work is preliminary and was conducted late in the season. A more detailed study will be conducted next season to fully characterize this B song variant, but there is 780 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4. December 2010 sufticient evidence to warrant larger and detailed surveys of the species over its breeding range, particularly in isolated and fragmented habitats. ACKNOWLEDGMENTS The authors thank J. S. Bolsinger for helpful discussions, Macaulay Library for audio recordings of Dendroica songs, L. B. Hawkins for making an audiovisual recording of a GCWA singing the new song, and M. D. Valero for comments on the manuscript. The research was funded by the Norman Hackerman Advanced Research Program (NHARP) (Grant Number 010115-0101-2007) from" the Texas Higher Education Coordinating Board (to RR) and UTSA STARS (to RR). LITERATURE CITED Bolsinger, J. S. 1997. Patterns of use and variations in the songs of the Golden-cheeked Warbler (Dendroica chrysoparia). Thesis. University of Massachusetts, Amherst, USA. Bolsinger, J. S. 2000. Use of two song categories by Golden-cheeked Warblers. Condor 102:539-552. Derryberry, E. P. 2009. Ecology shapes birdsong evolution; variation in morphology and habitat ex- plains variation in White-crowned Sparrow song. American Naturalist 174:24-33. Laiolo, P. and j. L. Tella. 2005. Habitat fragmentation affects culture transmission: patterns of song matching in Dupont’s Lark. Journal of Applied Ecology 42:1183-1 193. PULICH, W. M. 1976. The Golden-Cheeked Warbler, a bioecological study. Texas Parks and Wildlife Depart- ment, Austin, USA. SoKAL, R. R. AND F. J. Rohlf. 1995. Biometry. W. H. Freeman and Co., New York, USA. Spector, D. a. 1992. Wood warbler song systems: a review of paruline singing behaviors. Current Orni- thology 9:199-238. The Wilson Journal of Ornithology 122(4):780-783, 2010 Nocturnal Social Cues Attract Migrating Yellow-breasted Chats Mark G. Alessi,'-^ Thomas J. Benson,' and Michael P. Ward' ABSTRACT. — We tested the hypothesis that migrat- ing birds use nocturnal conspecific song when selecting stopover habitat using data from the Yellow-breasted Chat (Icteria virens). We broadcast nocturnal song in unsuitable habitat (i.e., a manicured orchard) and alternated broadcast nights with nights where no song was broadcast. We caught significantly more individu- als (8 males, 7 females) on mornings following treatments relative to control nights when no songs were broadcast (2.5 vs. 0 bird.s/morning, respectively). Eleven of 15 (73%) chats were removed from the nets after sunrise (mean = 32 min after sunrise, range = 5- 60 min), and birds were captured on overcast, cloudy, and clear mornings. Only one bird was recaptured at the site, and only one male was detected singing at the site, suggesting individuals quickly left the area. Conspecific nocturnal songs for chats appear to be an important cue for selecting stopover habitat. Received H March 2010. Accepted 9 July 2010. Reliable cues of high-quality habitat are important for migratory birds becau.se energetic and fitness costs of not locating suitable habitat ' Illinois Natural History .Survey, Institute of Natural Resource Sustainability, University of Illinois, Champaign, IF 61820, USA. ^Corresponding author; e-mail: mgalessKG'gmail.com are likely high. Migratory birds use a suite of navigational cues to migrate between wintering and breeding areas (Able and Bingman 1987); however, the specific cues used to select stopover habitat remain largely unknown. Studies suggest that both visual (Moore et al. 1995) and acoustic cues (Ward and Schlossberg 2004, Mukhin et al. 2008) may be involved. However, many noctur- nally migrating passerines select stopover loca- tions prior to sunrise (Cochran et al. 1967, Moore and Aborn 2000) when visual cues may be absent or difficult to evaluate. Cues that facilitate locating suitable habitat are likely highly valuable for migrating birds, and one may be nocturnal song. Nocturnal song by diurnally active birds appears to be uncommon; however, recent research suggests nocturnal song may be an important habitat location cue (Herremans 1990, .Schaub et al. 1999, Mukhin et al. 2008). Several diurnally active species that sing at night have specific nocturnal songs. Eor example, research on Common Nightingales {Luscinia megarhyu- chos) suggests males sing a whistle song at night to attract females (Glutz von Blotzheim 1988, Amrhein et al. 2002, Roth et al. 2009). The SHORT COMMUNICAl IONS 781 nocturnal song of Yellow-breasted Chats {Icteria vireiis) is also given at a frequency and amplitude that may allow detection farther than their diurnal song (Canterbury 2007). Female and male chats may use this song as an indicator of suitable habitat and mates. However, the relative impor- tance of conspecific vocalization versus habitat cues is unknown. We broadcast the nocturnal song of male Yellow-breasted Chats, a song that is consistently sung at night during spring migration (Canterbury 2007), in unsuitable habitat during spring migra- tion. The shrubland habitat in which chats stopover is patchy, and the probability of a migrating individual quickly locating suitable habitat at dawn is unlikely; thus, birds would benefit by using pre-dawn cues to locate appro- priate areas. We hypothesized that both male and female chats use nocturnal song as an indicator of suitable habitat, and that both males and females would select areas with conspecific nocturnal song regardless of habitat suitability. We predict- ed that chats attracted to unsuitable habitat would quickly leave the site once they found it was unsuitable. METHODS This experiment was conducted on Univer- sity of Illinois property in southeast Urbana, Champaign County, Illinois, USA during spring migration between 29 April and 15 May 2009. The site was a 210 X 30 m linear strip of six rows of apple (Mains spp.) trees spaced 6.5-m apart with mowed strips of grass between each row of trees. The habitat north and east of the site was residential and commercial property, while the habitat to the we.st and south was a mixture of agricultural land with additional patches of apple trees and mature forest. The area was selected because it offers woody vegetation, but lacked the size and dense shrub layer that chats require for both breeding (Thompson and Nolan 1973) and stopover habitat (Parnell 1969). Yellow-breasted Chats are not known to nest within the Champaign city limits (Kleen et al. 2004), and the nearest shrubland with breeding chats is ~16 km away. The site is routinely monitored during statewide bird surveys and other research projects, and chats have yet to be recorded at this location. We alternated between control nights, when no songs were broadcast, and treatment nights with song playbacks when the weather allowed (i.e.. not raining). We used a mixture of commercially available chat songs (Elliott et al. 2000) and chat songs we recorded from individuals in east-central Illinois. The recording included a male whistle song, the song only used at night (Canterbury 2007). The entire recording was 100 sec in length and was repeated continuously using a Eox Pro FX3 playback system. The whistle song was imbedded in two locations during the 100-sec playback: 15 sec and 80 sec, and was 0.5 sec in duration. We placed the system in the center of the site in a weatherproof plastic container and directed the device at a 45° angle southward toward the sky. Songs were broadcast on treat- ment nights from sunset to 1.5 hrs after sunrise. We placed mist nets approximately every 30 m on the site to increase the probability of capturing all chats moving through the vegetation; nets were numbered 1 (South net) to 6 (North net) and were opened for 1.5 hrs before and after sunrise each morning for both treatment and control nights. Nets were checked every 20 min for birds. We banded all captured chats with a USGS band, recorded gender, and the net number in which each individual was captured prior to releasing them at the site. We conducted 5-min, unlimited- radius point counts each morning in two locations at the site after closing the nets to quantify the number of singing chats. One point was at the southernmost end of the strip while the other was at the northernmost end. We conducted a Chi- square test to examine if chats were captured more often than expected on mornings following treatments. RESULTS We captured 15 chats (8 males, 7 females) during four of six mornings that we broadcast chat song at night and captured no new individuals during any of the five control mornings (x' = 12.5, P < 0.001). Eight of 15 chats (53%) were captured in nets closest to the speaker (chat song), while seven (47%) were captured in the south- ernmost nets. Eleven of 15 (73%) chats were removed from the nets after sunrise (mean = 32 min, range = 5-60 min), and birds were captured on overcast, cloudy, and clear mornings. Only one male chat was heard singing during point counts and it was during the morning he was captured. No male chats were recorded singing on control days. Only one individual was recaptured; a female 3 days after initial capture. 782 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 4. December 2010 DISCUSSION Migrating chats were captured only on morn- ings following nights when conspecific vocaliza- tions were broadcast; no birds were captured during mornings following nights with no song. Both migrating male and female chats responded to experimental addition of nocturnal male song, suggesting male chat song functions as a cue for selection of stopover habitat. Male chats did not establish territories and vacated the area. The lack of appropriate habitat did not preclude chats from briefly stopping, as long as conspecific songs were present. Nocturnal song of diurnally active birds has been documented, primarily in species that breed in patchy or ephemeral habitat (Barclay et al. 1985, Walk et al. 2000). For example, male Common Nightingales breed in ephemeral habitat (Hewson et al. 2005) and also sing nocturnally for mate attraction (Amrhein et al. 2002). Male chats also appear to primarily use long-range whistle songs only at night (Canterbury 2007). Male chats will sing continuously at night for up to 6 days after arrival in breeding areas, potentially provid- ing abundant social cues early in the season when most females are returning from wintering locations (Canterbury 2007). We broadcast nocturnal song between sunset and 1 .5 hrs after sunrise, and chats were attracted either during the night or shortly after sunrise. Selecting stopover habitats at sunrise may be adaptive if increased visibility allows for better habitat as.sessment (Bolshakov and Bulyuk 1999); these results suggest that, at least for chats, individuals are selecting habitat based on social cues. Visibility can also have a role in affecting the importance of social cues (Herremans 1990); birds in our study were captured during cloudy, overcast, and clear mornings, but whether visibil- ity had a role in why birds were captured is unknown. Conspecific attraction has received increasing attention in recent years, and is important for both habitat selection and planning conservation and management actions (Ahlering and Faaborg 2006). Several studies have demonstrated that birds use conspecific attraction to assess and select habitat (Mukhin 2004, Serrano et al. 2004, Ward and Schlossberg 2004, Ahlering 2005, Betts et al. 2008, Ktitorov et al. 2010). Most conspecific attraction studies have not attempted to attract birds to unsuitable habitat, nor have studies focused exclusively on nocturnal song (but see Mukhin et al. 2008, Ktitorov et al. 2010). Attracting birds to unsuitable rather than suitable habitat is a more reliable method of assessing the importance of acoustic social cues (Chemetsov 2006, Betts et al. 2008). Heterospecifics may also use these cues while selecting habitat (Mbnkkonen et al. 1999). We did capture other shrubland bird species on mornings following treatments, including Orchard Orioles {Icterus spurius). Brown Thrashers {Toxostoma rufum), and Field Sparrows (Spizella pusilla), but either did not capture or caught fewer individuals following control nights. These species do not commonly breed in developed areas. A separate playback experiment is necessary to examine if these species use chat song in addition to vegetation structure to select suitable habitat. Many of the bird species that breed in patchily distributed habitats in North America are known to sing at night (Barclay et al. 1985, Walk et al. 2000) and may be using nocturnal heterospecific song as a cue during migration; future research is needed to address conspecific and heterospecific attraction in shrubland birds. ACKNOWLEDGMENTS We thank D. G. and M. N. Barron, M. R. Heimbuch, and D. C. Elbert for assistance in the field. We are grateful for comments from J. D. Brawn, D. A. Enstrom, and P. J. Weatherhead. We also thank C. E. Braun and three anonymous reviewers for comments on the manuscript. This research was conducted under Institutional Animal Care and Use Committee protocol #06248 and banded under Federal permit #06507. LITERATURE CITED Able, K. P. and V. P. Bingman. 1987. The development of orientation and navigation behavior in birds. Quarterly Review of Biology 62:1-29. Ahlering, M. A. 2005. Settlement cues and resource use by Grasshopper and Baird’s sparrows in the upper Great Plains. Dissertation. University of Missouri, Columbia, USA. Ahlering, M. A. and J. Faaborg. 2006. Avian habitat management meets conspecific attraction: if you build it, will they come? Auk 123:301-312. Amrhein, V., P. Korner, and M. Naguib. 2002. Nocturnal and diurnal singing activity in the nightingale: correlations with mating status and breeding cycle. Animal Behaviour 64:939-944. Barclay, R. M. R., M. L. Leonard, and G. Friesen. 1985. Nocturnal singing by Marsh Wrens. Condor 87:418-422. Bett.s, M. G., a. S. Hadley, N. Rodenhouse, and J. J. Nocera. 2008. Social information trumps vegetation SHORT COMMUNICATIONS 783 structure in breeding-site selection by a migrant songbird. Proceedings of the Royal Society of London, Series B 275:2257-2263. Bolshakov, C. V. and V. N. Bulyuk. 1999. Time of nocturnal flight initiation (take-off activity) in the European Robin Erithacus rubecula during spring migration: direct observations between sunset and sunrise. Avian Ecology and Behaviour 2:51-74. Canterbury, J. L. 2007. Songs of the wild: temporal and geographical distinctions in the acoustic properties of the songs of the Yellow-breasted Chat. Dissertation. University of Nebraska, Lincoln, USA. Chernetsov, N. 2006. Habitat selection by nocturnal passerine migrants en route', mechanisms and results. Journal of Ornithology 147:185-191. Cochran, W. W,, G. G. Montgomery, and R. R. Graber. 1967. Migratory flights of Hylocichla thrushes in spring: a radiotelemetry study. Living Bird 6:213-225. Elliott, L., L. Stokes, and D. Stokes. 2000. Stokes field guide to bird songs: Eastern Region. Time Warner Audio Books, New York, USA. Glutz von Blotzheim, U. N. 1988. Handbuch der vogel mitteleuropas. Band 11/1. AULA — Verlag, Wiesba- den, Germany. Herremans, M. 1990. Body-moult and migration overlap in Reed Warblers (Acrocephalus scirpaceus) trapped during nocturnal migration. Gerfaut 80:149-158. Hewson, C. M., R. J. Fuller, and C. Day. 2005. An investigation of habitat occupancy by the nightingale Luscinia megarhynchos with respect to population change at the edge of its range in England. Journal of Ornithology 146:244-248. KLEEN, V. M., L. CORDLE, AND R. A. MONTGOMERY. 2004. The Illinois breeding bird atlas. Special Publication Number 26. Illinois Natural History Survey, Urbana, USA. Ktitorov, P., A. Tsvey, and A. Mukhin. 2010. The good and the bad stopover: behaviours of migrant Reed Warblers at two contrasting sites. Behavioral Ecology and Sociobiology 64:1135-1143. Monkkonen, M., R. Hardling, j. T. Forsman, and j. Tuomi. 1999. Evolution of heterospecifics attraction: using other species as cues in habitat selection. Evolutionary Ecology 13:91-104. Moore, F. R. and D. A. Aborn. 2000. Mechanisms of en route habitat selection: how do migrants make habitat decisions during stopover? Studies in Avian Biology 20:34-42. Moore F. R., S. A. Gauthreaux, P. Kerlinger, and T. R. Simons. 1995. Habitat requirements during migration: important link in conservation of neotropical landbird migration. Pages 121-144 in Ecology and manage- ment of neotropical migratory birds (T. E. Martin and D. M. Finch, Editors). Oxford University Press, New York, USA. Mukhin, A. L. 2004. Night movements of young Reed Warblers (Acrocephalus scirpaceus) in sum- mer: is it postfledging dispersal? Auk 121:203- 209. Mukhin, A., N. Chernetsov, and D. Kishkinev. 2008. Acoustic information as a distance cue for habitat recognition by noctumally migrating passerines during landfall. Behavioral Ecology 19:716-723. Parnell, J. F. 1969. Habitat relations of the Parulidae during spring migration. Auk 86:505-521. Roth, T., S. Sprau, R. Schmidt, M. Naguib, and V. Amrhein. 2009. Sex-specific timing of mate searching and territory prospecting in the nightingale: nocturnal life of females. Proceedings of the Royal Society of London, Series B 276:2045-2050. SCHAUB, M., R. Schwilch, and L. Jenni. 1999. Does tape- luring of migrating Eurasian Reed Warblers increase number of recruits or capture probability? Auk 116:1047-1053. Serrano, D., M. G. Forero, J. A. DonAzar, and J. L. Tella. 2004. Dispersal and social attraction affect colony selection and dynamics of Lesser Kestrels. Ecology 85:3438-3447. Thompson, C. F. and V. Nolan Jr. 1973. Population biology of the Yellow-breasted Chat (Icteria virens L.) in southern Indiana. Ecological Monographs 43:145-171. Walk, J. W., E. L. Kershner, and R. E. Warner. 2000. Nocturnal singing in grassland birds. Wilson Bulletin 112:289-292. Ward, M. P. and S. Schlossberg. 2004. Conspecific attraction and the conservation of territorial songbirds. Conservation Biology 18:519-52. 784 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4. December 2010 The Wilson Journal of Ornithology 122(4):784-788, 2010 Range Expansion of the White-breasted Waterhen {Amaiirornis phoenicurus) into Micronesia Donald W. Buden’-'’ and Stanley Retograh ABSTRACT. — We report the first occurrence of the White-breasted Waterhen {Amaiirornis phoenicurus) in Oceania based on discovery of a breeding population on Woleai Atoll, Yap, Micronesia. Islanders believe it first arrived on Woleai in the early 1970s, at a time when the species was beginning a period of range expansion that may be ongoing. Yap Outer Islanders report the waterhen is present also on nearby Eauripik, Ifaluk, and Laraulep atolls, where it arrived after it colonized Woleai. We summarize extralimital, nonbreeding dis- persal of A. phoenicurus and report a recent first record for Palau. Received 15 January 2010. Accepted 28 April 2010. The White-breasted Waterhen, Amaiirornis phoenicurus is known to breed from Pakistan, India, Sri Lanka, and the Maidive Islands east through southeast Asia to Japan and the Philip- pines, and south to Christmas and Cocos (Keel- ing) islands (Australian Indian Ocean Territory), and the Lesser Sundas (Taylor 1996). Steadman (2006) reported that it is not indigenous to tropical Oceania now or prehistorically, and makes no mention of any vagrant sightings of it. We report the first known occurrence of this species in Oceania based upon a breeding population on Woleai Atoll, Yap, Micronesia, hitherto known only to island residents. LOCALITY Woleai Atoll (n7°23'N, 143° 55' E) is within Yap State, one of four states (along with Chuuk, Pohnpei, and Kosrae) comprising the Federated States of Micronesia (FSM) (Fig. 1). The FSM and the Republic of Palau, immediately to the west, are in the Caroline Islands, which span —3,000 km of the west-central Pacific Ocean between the Philippines and Indonesia to the west and the Marshall Islands to the east. Woleai Atoll ' Division of Natural Sciences and Mathematics. College of Micronesia-FSM. P. O. Box 159, Kolonia. Pohnpei. Federated States of Micronesia 96941. "Yap State Department of Education. P. O. Box 220, Colonia. Yap. Federated States of Micronesia 96943. ’Corresponding author; e-mail: don_buden@conifsm.lm comprises 22 low (4 m maximum elevation asl) coralline islands distributed along a barrier reef around a 23.4 km' lagoon. The total land area is —4.5 km'; Falalop is the largest island ( 1 .53 kni'). The nearest land area is Ifaluk Atoll, 50 km to the east, Eauripik Atoll is 115 km to the southwest, and Faraulep (= Feshaiulep) Atoll is 150 km to the northeast. Sorol Atoll and the high islands of Yap proper are —390 and 660 km to the northwest, respectively. The vegetation on Fala- lop, Woleai, consists largely of coconut {Cocos micifera) agroforest, secondary woodland (shrubs and small trees), and marsh vegetation in swampy areas where people often cultivate taro (Cyrto- spernia chaniissonis). Archaeological data for Woleai are lacking, but the earliest known dates of occupation for Fais Island and Ngulu, Ulithi, and Lamotrek atolls among Yap Outer islands range from 1,200-1,900 BP (Intoh 1997, Rainbird 2004). The national census in 2000 recorded 975 residents on Woleai (Yap Branch Statistics Office 2002). OBSERVATIONS On 18 June 2009, while collecting insects on Falalop Island, Woleai Atoll, DWB heard several A. phoenicurus calling at different locations in low, wet areas along the airstrip and in the vicinity of the settlement. He saw none of the birds during —6 hrs on the island, and identified the calls in retrospect. He queried several local residents and learned only that the bird was called Kuwakuwa, an onomatopoeic rendering of its call, and that it probably arrived on the island about 30 years ago. The residents’ descriptions of the bird were surprisingly vague, considering its distinctive appearance. In response to our request to teachers on Woleai for a photograph or a specimen, Ellisa Siyaneral and Julian Yangerely captured a bird, which died shortly thereafter, in mid-July 2009. The .speci- men was frozen and subsequently delivered to SR on 19 August via the government field trip ship. Eddie Haleyalig delivered photographs of the SHORT COMMUNICATIONS 785 FIG. 1 . The Federated States of Micronesia and islands where Amaurornis phoeniciirus has been reported. specimen, taken by his daughter Jennifer, to DWB, who identified it as A. phoeniciirus. DWB prepared the bird as a study skin and depo- sited it in the collection of the Natural History Museum, Tring, UK (BMNH # 2010.13.1). We distributed a questionnaire among Woleai residents, and 22 of 30 respondents reported seeing nests of the Kuwakuwa, 20 of 20 indicated nests were in or near taro patches, 23 of 31 reported seeing eggs of this species, and 21 of 26 observed young, all on Falalop Island. Angelburt Igemera, a student at College of Micronesia, reported having seen a nest with eggs on Seliap Island, and student Dickson Tiwelfil reported hearing and observing the Kuwakuwa in taro patches on Wottegai Island. These and other responses indicate the White-breasted Waterhen is present on all of the inhabited islands of Woleai Atoll. We are confident these reports all pertain to A. phoeniciirus as there are no other rallids on the atoll and no other species with which it is likely to be confused. The extent of this species’ distribution else- where among the Yap Outer Islands is uncertain. Interviews with islanders indicate this species is present at least on Eauripik, Ifaluk, and Faraulep atolls, and that it probably arrived on these islands from Woleai. Julius Sapelalut (pers. comm.). Chief Public Defender, FSM Public Defenders Office, and former resident of Ifaluk, reported that A. phoeniciirus was not seen on Ifaluk until the early-to-mid- 1980s, by which time it was already well established on Woleai. Breeding has not been confirmed on Ifaluk, but is probable because the birds are heard calling year-round (J. Sapelalut, pers. comm.). John Haglelgam (pers. comm.), a professor at COM, reported his sister recalled flushing a Kuwakuwa from a nest containing one egg in a taro patch on Eauripik Atoll in 2008, and Mickael Rapiy (pers. comm.) reported encounter- ing this species on Faraulep Atoll a number of years ago, after first having seen it on Woleai ~2 years earlier. The distribution and abundance of A. phoeniciirus in central Yap State atolls requires further study to supplement these anec- dotal accounts. Additionally, Mark Vereen, a resident of Ngermetengel, on the west side of Babeldaob Island, Palau, observed a White-breasted Water- hen in his yard daily from 30 October to 3 November 2009. The bird was initially identified by Doug Pratt from photographs taken by M. Vereen, documenting the first record of A. phoeniciirus for the Republic of Palau. It was 786 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 4. December 2010 usually seen at the edge of dense vegetation, and not tar from an extensive wetland area. It often mingled with domestic chickens and was occa- sionally observed feeding on earthworms in a perpetually wet area of the yard. The disappear- ance of the bird was coincidental with the acquisition of two young dogs that roamed the yard freely. RANGE EXPANSION AND DISPERSAL The appearance of the White-breasted Water- hen in Micronesia apparently is coincident with a broad range expansion by this species. Amaur- ornis phoeniciinis is resident year-round through- out most of its breeding range, but northern birds are migratory and winter as far south as Sumatra, and as far west as the Arabian Peninsula (Taylor 1996). Quantitative data on migration and dis- persal are largely lacking, but Desmond Allen and P. D. Round (correspondence, April 2010) recorded 106 trapped during banding surveys at Dalton Pass, Luzon Island, Philippines, during November and December 2009. It was the most numerous of 12 species of rallids encountered during the survey. To what extent these were birds originating from outside the Philippines, is unknown. The waterhen was rarely seen in Japan until the 1970s, but was then observed with increasing frequency, and first recorded breeding in Kyushu in 1982 (Nakamura 1987). It is now a year-round resident in Japan’s Ryukyu Islands and Haha-jima in the Ogasawara (= Bonin) Islands and continues to advance northward into the Japanese main islands (Brazil 1991, Taylor 1996). The waterhen is now so common on Iriomote Island that it may be displacing the Slaty-legged Crake (Ralliiia ei(riz.(moides) there (Taylor 1996). Sakaguchi and Ishihira (1998) attributed an increase in the number encountered on Iriomote Island during March-May to an influx of breeding adults, and Nakamura (1987) indicated —73% of the birds recorded in Japan were from March to August with the highest frequency in May. The White-breasted Waterhen was first record- ed in Korea in 1970 and a failed breeding attempt was recorded there in 2001 (Park 2002, Moores 2007) . The first recorded successful nesting in Korea was in 2008 and two young were observed at Namyang Bay on 30 August (Edelsten et al. 2008) . The waterhen was first recorded on Seram in the Moluccas in 1987 (Bowler and Taylor 1989), and may now be resident there (Coates and Bishop 1997). Colonization of Australian Indian Ocean Territory islands may have been at approximately the same time. Two young pro- duced by a breeding pair in 1993, and again in 1994, and incidental sightings of other adults were the first records of A. phoenicurus on Christmas Island (Carter 1994), and the waterhen probably became established as a resident breeder on Cocos (Keeling) Islands sometime between the late 1990s and 2004 (Johnstone and Darnell 2004, Dooley 2005, Reid and Hill 2005, Hadden 2006). The White-breasted Waterhen was first recorded on the Arabian Peninsula in Oman in 1977 (Gallagher and Rogers 1980), and Pedersen (2008) considered it a rare to uncommon winter visitor in Dubai (United Arab Emirates). One photographed on 30 October 2009 is the first record for Saudi Arabia and the northernmost record for the Arabian Peninsula (Lobley and Roberts 2010), but the birds have not yet bred anywhere on the peninsula (M. C. Jennings, pers. comm.). Examples of long distance dispersal or vagran- cy in this species include: one found alive in the Seychelles on 24 December 1991 (Skeirett 1994); one found dead in central Mongolia and another in southern Mongolia in May 2004 and June 2004, respectively (Gombobaatar et al. 2005); several records from the Russian Far East, including one found dead on Bolshoy Pelis Island on 20 June and another in the Olginskiy District on 15 July 1987 (Nazarov and Kazykhanova 1988); one observed in the Sikhote-Alin Nature Reserve on the Pacific Coast of Sikhote Mountains in 1996 (Elsukov 1999) and another on Sakahlin Island on 20 September 1994 (Kozin 1995); one was captured, banded, and released on the Avacha River near Elizovo, northwest from Petropavlosk- Kamchatskiy, Kamchatka Peninsula on 1 Novem- ber 1996 (Gerasimov 1996). DISCUSSION We do not know exactly when White-breasted Waterhens first arrived on and colonized Woleai Atoll. The con.sensus among islanders we inter- viewed suggests the species became established- in the early 1970s, which would roughly coincide with the early stages of this species’ ongoing range expansion. None of the elders we inter- viewed recalled seeing it any earlier, and none was recorded anywhere in Micronesia by Baker (1951), who included records of other species encountered on Woleai by Japanese ornithologists SHORT COMIVIUNICAl IONS 787 during the early 19()()s. That a resident population has gone unreported from Woleai for over 30 years is readily explained by the absence of knowl- edgeable observers. Commercial flights to the atoll have been, and continue to be, infrequent and iiTegular. The FSM and Yap State government ships also visit Woleai irregularly, but usually spend only a brief part of a day before continuing on. Comparisons of photographs of the bird from Woleai with specimens at the American Museum of Natural History by Mary LeCroy, and at the Smithsonian Institution by StoiTS Okson, indicate the Woleai specimen is within the range of variation of the nominate subspecies. The other subspecies tend to have either less (A. p. leiicoinelanus, Sulawesi, Moluccas, and Lesser Sundas) or more (A. p. insularis and A. p. midnicohariciis, Andaman and Nicobar Islands) white in the forehead (Taylor (1996). The bird photographed at Palau also appears to be an example of the nominate subspecies (photographs examined by DWB and Doug Pratt). The water- hen population on Woleai could have originated from the nearest sedentary population of A. p. phoenicurus in the Philippines, but more likely from migratory northeast Asian birds that strayed east from their usual migratory path. The White- breasted Waterhen appears to have bypassed Palau, Yap, and the Marianas, high islands with seemingly suitable habitat much nearer to poten- tial source populations, in establishing a foothold in Micronesia. The waterhen’s furtive and skulk- ing habits may limit sightings, but its loud and frequent vocalizations make it hard to overlook. Possibly the recent Palau sighting is a harbinger of wider colonization of Micronesia. ACKNOWLEDGMENTS We thank Carlo Cu.stodio. Guy Kirwan, Ali.son Pirie, Elena Ilyashenko. Robert Meese, Doug Pratt, Robert Pry.s- Jone,s, Bruce Robert, and Philip Round for a.ssistancc with the literature search. Information on recent colonizations and extralimital records was provided by Walter Boles (for Australian Indian Ocean Territory islands), Adrian Skerrett and Robert Prys-Jones (Seychelles Islands). Mike Jennings (the Arabian Peninsula), Sundev Gombobaatar (Mongolia), Kiyoaki Ozaki (Japan), and Vladimir Loskot, Yuri Gerasimov, and Valentin Ilyashenko (English translations of records from Ru.ssia). We thank Mary LeCroy and Storrs Olson for comparing photos of the Woleai bird with examples of different subspecies of A. phoenicurus at the American Museum of Natural History and the National Museum of Natural History, respective- ly; Ellisa Siyaneral and Julian Yangerelyaro for obtaining the first specimen of A. phoenicurus from Woleai, John Talugmai for delivering it to Yap, and Eddie Haleyalig and daughter Jennifer for the photographs, which made the initial identification possible. We thank Julius Sapelalut for information on the status of this species on Ilaluk Atoll and are grateful also to the many other residents and former residents of Yap Outer Islands who shared their local knowledge of the Kuwakuwa. Addi- tionally, we thank Doug Pratt for calling our attention to the recent sighting of A. phoenicurus from Palau and Mark Vereen for contributing this record, and Desmond Allen and P. D. Round for contributing bird banding data from the Philippines. We also appreciate the helpful comments by D. Pratt, R. Prys-Jones, and an anonymous reviewer on an earlier draft of the manuscript. LITERATURE CITED Baker, R. H. 1951. The avifauna of Micronesia, its origin, evolution, and distribution. University of Kansas Publications, Museum of Natural History 3:1-359. Bowler, J. and J. Taylor. 1989. An annotated checklist of the birds of Manusela National Park, Seram: birds recorded on the Operation Raleigh Expedition. Kukila 4:3-29. Brazil, M. 1991. The birds of Japan. Smithsonian Institution Press, Washington, D.C., USA. Carter, M. 1994. Birds of Australia’s Christmas Island. Wingspan 13:18-21. Coates, B. J. and K. D. Bishop. 1997. A guide to the birds of Wallacea: Sulawesi, the Moluccas and Lesser Sunda Islands, Indonesia. Dove Publications. Alderly, Queensland, Australia. Dooley, S. 2005. Twitchers' corner. Wingspan 15:41. Edelsten, T, S. Kyu-Sik, and R. Newlin. 2008. Birds Korea bird news (August), http://www.birdskorea.org/ Birds/Birdnews/BK-BN-birdnews-2008-08.shtml Elsukov, S. V. 1999. Birds. Terrestrial animals' cadastre of the Sikhote-Alin Nature Reserve and northern Primorye. Vladivostok:29-75. Gallagher, M. D. and T. D. Rogers. 1980. On some birds of Dhofar and other parts of Oman. Journal of Oman Studies, Special Report 2:347-385. Gerasimov, Y. N. 1996. Accidental flight of Awaurornis phoenicurus to Kamchatka. The Russian Journal of Ornithology, Express-issue 1996 (5):6. Gombobaatar, S., D, Sumiya. C. W. Leahy, and E. Potapov. 2005. New records of bird species in Mongolia. Amphibians, reptiles and birds. Scientific Journal for Ornithology 2:121-128. Hadden, D. 2006. Cocos (Keeling) Island birds. Wimzspan 16:34-37. Intoh, M. 1997. Human dispersals into Micronesia. Anthropological Science 105:15-28. .IOHN.STONE, R. E. and J. C. Darnell. 2004. Appendix B. Annotated checklist of birds from Cocos-Keelins Islands. Pages 477-496 in Handbook of Western Australia birds. Volume 2. Passerines (Blue-winged Pitta to Goldfinch) (R. E. Johnstone and G. M. Storr, Editors). Western Australian Museum. Canberra. Kozin, A. N. 1995. Record of the new rallen in Sakhalin 788 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4. December 2010 White-breasted Crake. Bulletin of the Sakahlin Museum 2. Yuzno-Sakhalinsk:289. Lobley, G. R. and P. Roberts. 2010. First record of White-breasted Waterhen Amaurornis phoenicurus in Saudi Arabia. Sandgrouse 32:55-56. Moores, N. 2007. Selected records from Socheong Island, South Korea. Forktail 23:102-124. Nakamura, K. 1987. Range expansion of White-breasted Waterhen, Amaurornis phoenicurus, into Japan and its colonization. Bulletin of the Kanagawa Prefectural Museum (Natural Science) 17:1-11. Nazarov, Y. N. and M. G. Kazykhanova. 1988. Another finding in the USSR, the White-breasted Crake Amaurornis phoenicurus Pennant. Page 141 in Rare birds of the Far East and their protection (N. M. Litvenenko, Editor). Vladivostok, Russia. Park, Jin-YOUNG. 2002. Current status and distribution of birds in Korea. Dissertation. Kyung Hee University, Seoul, South Korea. Pedersen, T. 2008. Birds of the United Arab Emirates — an annotated checklist, http://www.tommypedersen.com/ UAEchecklist-annotated.htm Rainbird, P. 2004. The archaeology of Micronesia. Cambridge University Press, Cambridge, United Kingdom. Reid, J. R. W. and B. M. Hill. 2005. Recent surveys of the Cocos Buff-banded Rail (Gallirallus philippensis andrewsi). Report to the Australian Government Department of the Environment and Heritage. Center for Resource and Environmental Studies, Australian National University, Canberra. Sakaguchi, N. and ISHiHiRA. 1998. Distribution and population dynamics of the White-breasted Waterhen Amaurornis phoenicurus in Iriomote Island, Okinawa, southern Japan. Island Studies in Okinawa 6:25^0. Skerrett, a. 1994. Seychelles Bird Records Committee. Birdwatch (Seychelles) 9:15-19. Steadman, D. W. 2006. Extinction and biogeography of tropical Pacific birds. University of Chicago Press, Chicago, Illinois, USA. Taylor, P. B. 1996. Eamily Rallidae (Rails, gallinules, and coots). Pages 108-209 in Handbook of birds of the world. Volume 3 (Hoatzin to Auks) (J. del Hoyo, A. Elliott, and J. Sargatal, Editors). Lynx Edicions, Barcelona, Spain. Yap Branch Statistics Office. 2002. Yap state census report, 2000 FSM census of population and housing. Division of Statistics, Department of Economic Affairs, National Government, Colonia, Yap, Federat- ed States of Micronesia. The Wilson Journal of Ornithology 122(4):788-791, 2010 First Reproductive Record of Wilson’s Plover in Baia de Todos os Santos, Northeastern Brazil Vitor O. Lunardi'"^ '^ and Regina H. Macedo“ ABSTRACT. — The Wilson’s Plover (Charadrius wilsonia) is widely distributed along the coast of the Americas. We present the first reproductive record in Bai'a de Todos os Santos, Brazil, broadening the southernmost limit of its breeding area along the Atlantic Coast to 12 44' S, 38° 45' W. We recorded a family with a subadult in 2007, and a family with chicks and a nest in 2008. The female invested more time in incubation than the male in 96 hrs of diurnal nest observations. There were 102 interruptions during incubation due to the approximation of domestic animals (cattle and horses). The nest was abandoned ' Programa de Pos-Graduay'ao em Ecologia, IB, Campus Universitario Darcy Ribeiro. Universidade de Brasilia, Brasilia. DF. Brazil. 70910-900. ^ Departamento de Zoologia, IB. Campus Universitario Darcy Ribeiro. Universidade de Brasilia, Brasilia, DF, Brazil. 709 1 0-900. ’Current address: Departamento de Ciencias Animals, Universidade Federal Rural do .Semi-Arido, Mossoro, RN, Brazil. 59625-900. ■'Corresponding author; e-mail: lunardi.vitor@gmail.com 9 days after egg laying. An experiment with artificial nests suggests that 30% of Wilson’s Plover nests may be destroyed by free-ranging domestic animals in this Bala. Received II Januaiy 2010. Accepted 5 June 2010. The Wilson’s Plover {Charadriu.s wilsonia) is a widely distributed Charadriidae associated with coastal habitats in the Americas with resident populations in numerous sites. It breeds on the Atlantic Coast, between the eastern USA and northeastern Brazil, in the Caribbean Islands, and on the Pacific Coast, between Mexico and northwestern Peru (Corbat and Bergstrom 2000). The species exhibits plumage differences through- out its range, which attracted the attention of systematists, leading to subdivision of the species into five subspecies (Wiersma 1996, Corbat and Bergstrom 2000, Grantsau and Lima 2008). Wilson’s Plover is seemingly monogamous (Cor- bat and Bergstrom 2000), and both males and SIIOR r COMMUNIC A HONS 789 females are eapable of performing ineubalion (Bergstrom 1981). Historically, the breeding range of Wilson’s Plover has been contracting and reproductive populations in the current northernmost limit are threatened, mainly due to destruction of breeding areas and interference during reproduction and while resting. However, little is known about threats outside this region (Corbat and Bergstrom 2000). Records of reproduction and information about populations of Wilson's Plover in the Southern Hemisphere are scarce and restricted to a few sites (Tovar 1968, Rodrigues et al. 1996, Wiersma 1996, Corbat and Bergstrom 2000, Grantsau and Lima 2008). We describe: (1) the first breeding record of Wilson’s Plover in the southernmost limit of its breeding range on the Atlantic Coast (Bata de Todos os Santos, Brazil), and (2) features of the species’ reproductive behavior and poten- tial threats to the study population. METHODS Monitoring of Nests and Families. — We recorded breeding of Wilson’s Plover in the coastal region of Saubara, Bahia, western coast of Baia de Todos os Santos, Brazil (12° 44' S, 38° 45' W). This region is characterized by a complex of salt marshes (restingas), mangroves (Rhizophora, Laguncularia, Avicennia), suprati- dal salt flats (apicum), and intertidal areas. The climate is humid tropical with distinct dry (Sep- Feb) and rainy (Mar-Aug) seasons. We com- pared photos of individuals from the study site for species identification with specimens depos- ited in the Museu de Zoologia da Universidade de Sao Paulo-MZUSP (specimens: 41395, 41396, 42745, 8(3034, and 80035). We based our comparisons on the Wilson’s Plover sub- species identification key to the Atlantic Coast developed by R. Grantsau (Grantsau and Lima 2008). OBSERVATIONS The first evidence of breeding of Wilson’s Plover in the region was the record of a family, comprising a pair with a subadult, foraging in the intertidal zone on 17 December 2007. The following year, we found a Wilson’s Plover nest with two eggs on 13 August 2008 in a restinga area. The next morning, taking advantage of the absence of the pair from the nest, we took nest and egg measurements and recorded a third egg FIG. 1. Ne.sl and chick of Wilson's Plover at Bai'a de Todos os Santos. Brazil. Photographs by Vitor O. Lunardi. (Fig. I). We performed 96 hrs of focal observa- tion of the nest to examine parental investment in incubation, between 0600 and 1800 hrs. during the first 8 days of incubation ( 14 to 21 Aug 2008). Observations were conducted from a fixed point inside of Ihe mangrove, 26 m from the nest, to avoid investigator disturbance. The nest with the eggs was abandoned on 22 August 2008, probably due to the intense disturbance by domestic animals (cattle and horses). The nest with eggs remained intact until 30 August 2008, when they were found destroyed. On 27 August 2008 we found another family of Wilson’s Plover with three chicks in an apicum area. Each chick (Fig. 1 ) was measured, weighed, and marked with a metal band (CEMAVE/Brazil). This family remained for the next 7 days at the 790 THE WILSON JOURNAL OL ORNITHOLOGY • Vot. 122, No. 4, December 2010 Days FIG. 2. Number of artificial nests destroyed, predated, and intact in a restinga area at Bai'a de Todos os Santos, Brazil, during 25 days. same site with the aggregated chicks always accompanied by the adults. Experiment to Verify Predation Intensity. — We observed Wilson’s Plover pairs defending territo- ries in a restinga area between May and October 2007 and 2008, also used as pasture by free- ranging cattle and horses from the local commu- nity. We conducted an experiment with artificial nests to estimate nest loss considering these animals may destroy nests. We built 30 artificial nests (cavities in the soil similar in shape and size to Wikson’s Plover nests) in January 2009, placed in a series, 30 m apart, along a strip of restinga. We added three Common Quail (Coturnix cotur- nix) eggs to each artificial nest. We made three daily visits to each nest {0600, 1200, and 1800 hrs) over 25 days (approximate incubation time for the species) to record nests that were: ( 1 ) intact, (2) destroyed (with at least 1 destroyed egg and tracks of domestic animals), and (3) predated (absence or destruction of at least 1 egg; no signs of domestic animals). Nine of 30 artificial nests (Fig. 2) were destroyed with 77.7% of occurrences during the day (()6()()-180() hrs). Fifteen nests were predated with 66.7% of the.se events at night (18()()-()6()() hrs). Monitoring of Nests and Families. — The only nest found consisted of a 25-mm cavity in sandy soil with a diameter of 102 mm and fragments of dry vegetation around the eggs. It was next to horse feces and a shrubby false buttonweed (Spermaeoce verticillala L., Rubiaceae; height 17.5 cm) surrounded by grasses (Fig. 1). The three eggs measured: length = 34.9, 34.8, and 34.2 mm; width = 25.2, 25.0, and 24.4 mm; and mass = 14.0, 14.0, and 13.5 g. The eggshells were colored light cream with dark brown stains over the entire surface, but with a higher concentration on the blunt end. The female invested 53.3 hrs (93.3%) in incubation during the observation period while the male invested 3.8 hrs (6.7%). The female was observed incubating between 0600 and 1733 hrs, and the male after 1714 hrs. The female interrupted incubation 125 times (T ± SD = 1.29 ± 1.41 /hr, n = 96) of which 102 (81.6%) were caused by approach of domestic animals, mainly cattle. After leaving the nest, the female often exhibited alarm and distraction behaviors: squatting, head-up (followed by vocalization), mock-brooding and/or broken-wing. In 54.3% of these interruptions the male approached the nest, exhibiting head-up and crouch-run display (be- havioral descriptions in Bergstrom 1988b). The female was reluctant to return to the nest after disturbances. The measures obtained for the three chicks were: right tarsus = 23.5, 23.3, and 22.1 mm; CLilmen = 17.3, 16.8, and 16.1 mm; and mass = 17.0, 16.5, and 15.5 g. The beaks were black; legs and feet grayish; plumage upperparts were brownish and black in a .scalloped pattern, and underparts and neck were white. The age and gender of the chicks could not be assigned. DISCUSSION We present the first breeding record of Wilson’s Plover in Baia de Todos os Santos, Brazil. This is the southernmo.st record on the Atlantic Coast for the species, representing an extension of their breeding range (Corbat and Bergstrom 2000, Grantsau and Lima 2008). Grantsau and Lima (2008) recorded breeding for Wilson’s Plover (C. wilsonia brasiliensis) in Mangue Seco, Brazil, ~23() km north of Bala de Todos os Santos. The plumage of the males and females in our study area was similar to that of C. vv. brasiliensis deposited at MZUSP, and suggests our records correspond to this subspecies. Bergstrom (1988a) reported Wilson’s Plover nests mainly near vegetation and occasionally near objects (e.g., cow manure, stones), emphasizing that both may function as a protection barrier against the wind. Objects and vegetation provided protection for the nest that we found againsl SHORT COMMUNICATIONS 791 easterly and southeasterly winds, which are predominant in August, when we found the nest. The number, dimensions, and color of the eggs and size of the chicks we report are similar to those previously published for the species (Ro- drigues et al. 1996, Corbat and Bergstrom 2000, Grantsau and Lima 2008). The female invested more time in diurnal incubation than the male. However, the male may have invested in noctur- nal incubation, since they initiated this activity at the end of the day. Previous observations for the species indicate males incubate mainly at night (Thibault and McNeil 1995). The experiment with artificial nests revealed that, in addition to natural threats (predators), use of coastal areas for grazing is another hazard to the reproductive success of Wilson’s Plover in Baia de Todos os Santos. Trampling by roaming animals was previously considered a potential threat for Wilson’s Plover eggs and recent hatched nestlings at Ilha do Curupu, Maranhao, Brazil (Rodrigues et al. 1996). Plovers leave their nests when disturbed and are extremely reluctant to return when intruders are present near nest sites. The constant interruption of incubation by inter- ference may lead to inadequate egg temperature regulation, and exposing eggs to predation and overheating (Corbat and Bergstrom 2000). The female often left its nest when disturbed by free- ranging cattle and horses. Both females and males invested actively in predator avoidance behaviors. We assume the intense disturbance was the primary cause of nest and egg desertion. Our observations confirm Baia de Todos os Santos as a breeding area of Wilson’s Plover, extending the limit of its breeding range. Use of coastal areas for grazing is critical to consider in any project directed at conserving this population. ACKNOWLEDGMENTS This research was supported by fellowships from Conselho Nacional de Desenvolvimeiilo Cienti'fico e Tecnologico and Coordena^'ao de Aperfeiij'oamento de Pessoal de Ni'vel Superior. We are grateful to CEMAVE/ ICMBio for authorization to capture and band the birds. We thank L. F. Silveira for authorization and access to the collection of MZUSP, J. G. Jardirn for taxonomic identification of S. verticillata. and D. G. Lunardi for revision of an earlier version of the manuscript. We thank two anonymous referees for helpful and constructive comments on the manuscript. LITERATURE CITED Bergstrom, P. W. 1981. Male incubation in Wilson's Plover {Cha rad fills wilsonia). Auk 98:835-838. Bergstrom, P. W. 1988a. Breeding biology of Wilson's Plovers. Wilson Bulletin 100:25-35. Bergstrom, P. W. 1988b. Breeding displays and vocaliza- tions of Wilson’s Plovers. Wilson Bulletin 100:36-49. Corbat, C. A. and P. W. Bergstrom. 2000. Wilson’s Plover (Charadriits wilsonia). The birds of North America. Number 516. Grantsau, R. and P. C. Lima. 2008. Uma nova sub- especie de Charadriits wilsonia (Aves, Charadrii- formes) para o Brasil. Atualidades Ornitoloaicas 142:4-5. Rodrigues, A. A. F., D. C. Oren, and A. T. L. Lopes. 1996. New data on breeding Wilson's Plovers Charadrius wilsonia in Brazil. Wader Study Group Bulletin 81:80-81. Thibault, M. and R. McNeii.. 1995. Day- and night-time parental investment by incubating Wilson's Plovers in a tropical environment. Canadian Journal of Zoology 73:879-886. Tovar. H. 1968. Areas de reproduccion y distribucion de las aves marinas en el litoral peruano. Boleti'n del Instituto del Mar del Peru 1 :523-526. WiERSMA, P. 1996. Species accounts. Family Charadriidae (plovers). Pages 410^42 in Handbook of the birds of the world. Volume 3 (Hoatzins to Auks) (.1. del Hoyo, A. Elliott, and J. Sargatal. Editors). Lynx Edicions. Barcelona, Spain. 792 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4, December 20 10 The Wilson Journal of Ornithology 1 22(4):792-795, 2010 Hatching Synchrony, Green Branch Collecting, and Prey Use by Nesting Harpy Eagles {Hcirpia harpyja) Adrian S. Seymour, Graham Hatherley," Francisco Javier Contreras,-^ James Aldred,"* and Fergus Beeley”^ ABSTRACT. — We observed an occupied Harpy Eagle (Harpia harpyja) nest over three separate periods in eastern Venezuela. Both eggs in the clutch hatched on the same day, and two nestlings competed in the nest for 14 days before one succumbed. The female adult collected green branches 45 times over 60 days of observation. Green branch deliveries were positively associated with prey deliveries and our observations best support the ‘nest sanitation by covering prey’ hypothesis for the adaptive significance of green branch collecting. Prey delivery rate to the nest averaged one delivery every 2.4, 2.1, and 3.7 days in the three observation periods. Three-toed sloths (Bradypus tridactylus) and wedge-capped capuchins (Cehus olivaceiis) were the most common prey items brought to the nest. Received 14 April 2010. Accepted 23 July 2010. The Harpy Eagle (Harpia harpyja) is one of the largest and most emblematic eagles in the world, and is a species of considerable conservation concern throughout its range in Central and South America (Vargas et al. 2006). However, remark- ably little is known of its ecology and behavior. The main reasons for this lack of knowledge are that Harpy Eagles are hard to locate and observe in the rainforest habitats in which they live. Much of what is known about their behavior comes from observations during nesting and chick rearing when their movements become centered on a nest (summarized by Schulenberg 2009). However, becau.se of the difficulty in ob.serving nests (they are usually positioned 20 m or more up in the canopy), only a few studies have been made and behavior at the nest is still poorly understood. For ' 13a Northumberland Road. Bristol. B.S6 7AU. UK. ’Emirates Natural History Unit. 59 Cotham Hill, BS6 6JR. UK. 'Motel Taguapire. Via Rio Grande. El Palmar, Estado Bolivar, Venezuela. ■‘Canopy Acce.ss Ltd.. 81a Hill Road, Clevedon, BS2I 7F’L. UK. ’’98 Coombe Dale. Bristol, B.S9 2JE, UK. ‘’Corresponding author; e-mail: aseymour_uk@yahoo.co.uk example, Harpy Eagles have been observed to collect green branches (i.e., branches with green foliage) during the nesting period (Rettig 1978), a behavior seen in many species of raptors. The adaptive signifance of this behavior remains unknown. Several hypotheses have been put forward including: (1) nest sanitation by repelling iivsects with the volatile components of leaf chemistry, (2) nest sanitation by covering prey remains, (3) advertisement of nest occupancy, (4) maintenance of humidity in the nest, and (5) provision of shade for the nestlings (Wimberger 1984). Little is known about factors affecting prey use by Harpy Eagles during nesting, and only limited data are available (summarized by Schulenberg 2009). Harpy Eagles are distributed over a large geographic area, and the composition of prey species varies considerably in different parts of its range (Emmons and Feer 1997). It is probable that regional variations in relative abundance and ease of catching prey also occur. Our objectives were to: ( I ) use quantitative and qualitative ob.servations of Harpy Eagle nesting behavior to explore the adaptive significance of green branch collecting, (2) report prey use by a pair of nesting Haipy Eagles in a fragmented forest habitat in Venezuela, and (3) report an unusual case of hatching synchrony and extended sibling competition. OBSERVATIONS We observed an active Harpy Eagle nest for 60 days over three separate periods (2 May-2 Jun 2009, 5 Aug-3 Sep 2009, 8 Jan- 10 Feb 2010) in the Sierra Imataca of Bolivar State, ea.stern Venezuela (07° 56.926' N, 61 43.244' W). We monitored the nest from a blind on an arboreal platform 25 m from the ground ~6() m from the nest tree, a tall (>35 m) silk-cotton or ceiba tree (Ceiha pentaii- dra). The nest was also monitored using a small remote camera fixed to a branch ~3 m above the nest platform. SI lORT COMMLINICA'I IONS 793 Green Branch Collecting. — Harpy Eagles col- lecled branches with green leaves on 46 separate occasions in the 60 days the nest was observed. The lemale collected the green branches in all but one occasion. Most often the female would anange the green branches over prey remains. She did not cover the chick(s) with green branches and hypothesis # 5 is not supported by our observations. The ambient humidity of the rain- forest habitat is high, suggesting that hypothesis # 4 is unlikely to explain the main function of green branches in the nest. One or more green branch deliveries appeared more likely to occur on days when prey was also delivered, but this association was not statistically significant (2X2 contingen- cy table: y- = 2.39, df = 1, P = 0.12). However, there was a period of 3 consecutive days during an exceptionally long fasting period when the female’s green branch collecting behavior was atypical in that ( 1 ) it was much more frequent than usual (branch collecting observed 13 times in 3 days compared to the expected frequency of 5.7 based on the average green branch delivery rate on ‘normal’ delivery days), and (2) the female jettisoned 38% of these green branches before reaching the nest, which was not normally done. If these 3 days of unusual behavior are omitted from the statistical analysis, the association between green branch collecting and prey delivery be- comes significant (2X2 contingency table: x~ — 3.94, df = 1, P < 0.047). These observations support hypothesis # 2. We observed flies and other winged insects frequently flying around the nest. The female and the remaining chick responded to flying insects by vigorously shaking their heads or snapping with their beaks. The female or chick would react in this way to flying insects at least once on 9 of the 21 observation days (i.e., 43% of days) in the first observation period (chick 0-21 days of age). This contrasts with the second observation period (chick 95- 112 days of age) when they reacted to flying insects only once in 15 observation days (i.e., 7% of days). We saw the female peck at crawling insects in the nest at least once in 10 of the 21 observation days (i.e., 48% of days) during the first observation period, while in the second period, we observed this behavior on only 3 of 15 ob.servation days (20% of days). Responses to arthropods and green branch collecting were more frequent in the first period when the chick was more vulnerable (average 1.3 branch deliverie.s/ day in the first period vs. 0.9 deliveries/day in the second observation period). The.se observations support both (#’s 1 and 2) nest sanitation hypotheses for green branch collection by Harpy Eagles. Most of the green branches were from the C. pentandra tree which supported the nest, but some were also taken from a nearby fig (Ficus spp.) tree. Branches were taken from another tree species on only one occasion. Prey Delivery. — We recorded 23 prey deliver- ies in the 60 days the nest was monitored. The male Harpy Eagle delivered 12 prey items to the nest in 29 days (average of 1 delivery every 2.4 days) in the first observation period. He delivered five pale-throated three-toed sloths (Bradypus tridcicty’lus), four wedge-capped capu- chins (Cehiis oiivaceus), and three unidentifiable prey items. Seven prey deliveries were recorded in 15 days (average 1 delivery every 2.1 days) in the second observation period, four by the male and three by the female. Three of the seven prey delivered were capuchin monkeys, two were sloths, and two could not be identified. Eour prey deliveries were observed in 15 days in the third observation period (average 1 delivery every 3.7 days). Two deliveries were made by the male, and the other two were made by the female. It was not possible to identify these prey items as the only video recorded of these deliveries was from positions that did not allow us to see into the nest. Synchronous Hatching. — Both eggs in the Harpy Eagle nest hatched on the same day (12 May 2009). Two chicks were seen in the nest up to 14 days after hatching, at times seen struggling to climb on top of each other. Only one of the chicks was ever observed being fed by the female at a time, although we could not tell if it was always the same chick. DISCUSSION Our observations of Harpy Eagles at the nest suggest that green branch collection is important for nest sanitation. Harpy Eagle behavior in response to arthropods (head shaking and peck- ing) was much more frequent in the first period of observation than subsequent observation periods suggesting that arthropods were more problematic when the chick(s) were newly hatched. This observation can be explained by a change in arthropod abundance/activity around the nest between different observation periods or, alterna- tively, the female Harpy Eagle may have been more diligent in removing potentially dangerous arthropods when the chick was very small and 794 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 4, December 2010 presumably more vulnerable to parasitism or being bitten and stung. Little is known about the effect of parasites and diseases on Harpy Eagle nestlings. Arthropod ectoparasites such as botflies in the genus Philonus can have a highly detrimental effect on survival ot nestlings of many bird species in the Neotropics (Dudaniec and Kleindorfer 2006) including small Falconiformes (Delannoy and Cruz 1991). Stinging and biting arthropods such as ants and wasps may also pose a threat to a harpy chick when it is small. Rettig (1978) observed a nesting female Harpy Eagle to snap at and kill insects, particularly bees and wasps that got close to the nest. The female Harpy Eagle in Rettig’s (1978) study preferred collecting branch- es of Mora trees rather than those of the C. pentandra tree in which it was nesting. These observations suggest high selectivity for particular kinds of green branches, but we could not tell whether this selectivity was based on arthropod- repelling quality or some other factor such as the ea.se at which the branches could be broken. The ‘nest sanitation by repelling insects with volatile components of leaf chemistry’ hypothesis cannot be ruled out. However, there was no consistency in branch choice between studies and in.sects were observed on the nest even when green branches were present (i.e., insect repelling capability of the green branches did not seem to be strong). The most common prey delivered to the nest in Rettig's (1978) study were sloths and capuchin monkeys. Sloths have been shown to be important components of Harpy Eagle diets in widely different parts of its range including Panama (Touchton et al. 2002) and the Brazilian Amazon (Galetti and de Carvalho 2000). Red-howler monkeys (Alnaatta seniculus) were common in our study area with .several groups living within a 5()0-m radius of the nest. Red-howler monkeys can be taken by Harpy Eagles (Rettig 1978, Peres 1990, Sherman 1991), and the similar-sized mantled howler (A. palliata) is the preferred primate prey of Harpy Eagles on Barro Colorado Island in Panama (Touchton et al. 2002). Howev- er, we did not observe any howler remains in the nest. Rettig (1978) ob.served a prey delivery rate of one prey every 3.5 days in the first part of the nesting period when the male was doing all the hunting and one prey every 2.5 days later in the nesting period when both adults were hunting. We observed the male Harpy Eagle deliver prey on average once every 2.4 days in the first period, and both parents delivered at an average rate of once every 2.1 days in our second observation period. There are many possible explanations for this apparently higher delivery rate at our study site, including differences in prey abundance or even ease of catching prey. One reason for increased prey delivery could be that the forest fragmentation characterizing our study area re- sulted in large amounts of forest edge habitats which may be favored by Harpy Eagles hunting monkeys (Eason 1989). Many eagle chicks show siblicidal behavior, but hatching is usually asynchronous (Simmons 1986) making competition between chicks much more asymmetric. Our observations of synchro- nous hatching differ from those of Rettig (1978) who only observed one of two Harpy Eagle eggs in the clutch hatch. Our obseiwation is unexpected as more equally-sized competing chicks have a similar potential to damage each other, which could result in the reduced survival of both competitors. More data are needed to quantify Harpy Eagle behavior and evaluate hypotheses for the adaptive significances of these behaviors. However, unless methods for finding and monitoring Harpy Eagle nests are greatly improved so that large sample sizes can be obtained for reasonable effort, our understanding of Harpy Eagle behavior is likely to remain obscure. ACKNOWLEDGMENTS We are grateful for perniis.sion to film the nest event granted by Centro Nacional Autonomo de la Cinemato- grafia (CNAC) of Venezuela's Ministry of Culture (No. CNAC/ PCIA No. 095-2009). We thank Senor Rafael Garcia and his son Larry for permission to conduct the study on their land, and for their great hospitality and kindness. We also thank Alejandro Lamus and Jose Hurtado for their highly competent, professional assistance and good companionship, and Emilio Perez for arranging permits. Finally we thank C. E. Braun and two anonymous referees for helpful suggestions on early versions of the manuscript. LITERATURE CITED Df.i.annoy, C. a. and a. Cruz. 1991. Philomis parasitisjii and nestling survival of the Puerto Rican Sharp- shinned Hawk. Pages 93- 1 (H in Bird-parasite inter- actions: ecology, evolution, and behavior (.1. E. Love and M. Zuk, Editors). Oxford University Press, Oxford, United Kingdom, OtiDANiEC, R. Y. AND S. Kleindorkbr. 2006. Effects of the parasitic Dies of the genus Philorni.s (DipteraiMusci- dac) on birds. Emu 106:13-20. SHORT COMMUNICATIONS 795 Eason, P. OH*-). Harpy Eagle allenipts predation oti adult howler monkey. Condor 9 1 :469-47(). ENtMONS. L. H. AND F. Feer. 1997. Neotropieal rainforest mammals: a field guide. Second Edition. University of Chicago Press, Chicago, Illinois, USA. G.ALEtTi, M. AND 0. DE Carvai.ho, 2000. Slotlis in the diet ot a Harpy Eagle nestling in eastern Amazon. Wilson Bulletin 112:535-536. Peres, C. A. 1990. A Harpy Eagle successfully captures an adult male red howler monkey. Wilson Bulletin 102:560-561. Rettig, N. L. 1978. Breeding behavior of the Harpy Eagle (HarpUi harpyja). Auk 95:629-643. Schulenberg, T. S. 2009. Harpy Eagle (Harpia harpyja). Neotropical birds online (T. S. Schulenberg, Editor). Cornell Laboratory of Ornithology. Ithaca, New York, USA. http://neotropical.birds.cornell.edu/portal/specie,s/ over\'iew?p_p_spp=206 1 3 StiERMAN, P. T. 1991. Harpy predation on a red-howler monkey. Folia Primatologica 56:53-56. Simmons, R. 1986. Offspring quality and the evolution of cainism. Ibis 130:339-357. Touciiton, J. M., Y. HstJ, and A. PAtj.ERONi. 2002. Foraging ecology of reintroduced captive-bred sub- adult Harpy Eagles {Harpia harpyja) on Barro Colorado Island. Panama. Ornitologia Neotropical 13:365-379. Vargas, J. D., D. Whitacre, R. Mosquera, J. Albu- querque, R. Piana, j. M. Thiollay, C. Marquez, J. E. Sancmez, M. Lezama-Lopez, S. Midence, S. Matola, S. Aguilar, N. Rettig, and T. Sanaioiti. 2006. Status and current distribution of the Harpy Eagle {Harpia harpyja) in Central and South America. Orintologia Neotropical 17:39-55. Wimberger, P. H. 1984. The use of green plant materials in bird nests to avoid ectoparasites. Auk 101:615-618. The Wilson Journal of Ornithology 122(4):795-799, 2010 Screaming Cowbird Parasitism of Nests of Solitary Caciques and Cattle Tyrants Alejandro G. Di Giacomo,' Bettina Mahler,^ and Juan C. Reboreda^-^ ABSTRACT. — The Screaming Cowbird (Molothrus rufoaxillaris) is one of the most specialized brood parasites with only three known hosts: Baywing (Agelaioicles hadiiis), the main host throughout most of its range, and two alternative hosts in some areas of its distribution, Chopi Blackbird (Gnorimopsar chopi) and Brown-and-yellow Marshbird (Pseiuloleisies vir- escens). We studied Screaming Cowbird parasitism in northeast Argentina where this parasite uses Baywings and Chopi Blackbirds as hosts. We monitored 69 nests of Baywings, 251 of Chopi Blackbirds, 31 of Solitary Caciques (Cacicus soli lari us), and 30 of Cattle Tyrants {Macheiornis rixosa). The frequency of Screaming Cowbird parasitism on Baywing nests was 80% and was 46% for Chopi Blackbirds. We recorded one event of Screaming Cowbird parasitism on one nest of Solitary Caciques and three events of Screaming Cowbird parasitism on one nest of Cattle Tyrants. The identities of parasitic eggs in both hosts were confirmed by sequencing the mtDNA control region. We propose ' Departamento de Conservacion. Aves Argentinas-Aso- ciacion Ornitologica del Plata. Matheu 1246, C1249AAB Buenos Aires, Argentina. "Departamento de Ecologi'a, Genetica y Evolucion, Facultad de Ciencias Exactas y Naturales. Universidad de Buenos Aires, Pabelldn II Ciudad Universitaria, CI428EGA Buenos Aires, Argentina, •’Corresponding author; e-mail: reboreda@ege.fcen.uba.ar these events of parasitism resulted from recognition errors by Screaming Cowbird females that regularly parasitize Baywings and Chopi Blackbirds. The nest of Solitary Caciques had been frequently visited by a pair of Baywings before Screaming Cowbird parasitism occuiTed, and the nest of Cattle Tyrants was near an active Chopi Blackbird nest that had been previously parasitized by Screaming Cowbirds. Received 5 Janu- ary 2010. Accepted 7 April 2010. The Screaming Cowbird (Molothru.s ntfoo.xil- lari.s) i.s a specialized brood parasite, which exclusively u.ses the Baywing (Agelaioides ha- diiis) as a host throughout most of its range (Ma.son 1980, Fraga 1998). It also parasitizes the Brown-and-yellow Marshbird (P.seudoletstes vir- e.scens) in .some areas (Mermoz and Fernandez 2003) and the Chopi Blackbird (Gttornuojisctr chopi) (Sick 1985, Di Giacomo 2005). Other species have been reported as hosts of the Screaming Cowbird, but there is general agree- ment that most of these reports were based on mis-identified Shiny Cowbird (M. honarien.sis) eggs (Friedmann 1963, Mason 1980, Fraga 1986). One hypothesis posited to explain host .speci- ficity in brood parasites is that females imprint on 796 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122. No. 4. December 2010 their foster parents and, once mature, search for nests of the same species in which to lay their eggs (Nicolai 1964, Slagsvold and Hansen 2001). This hypothesis has been directly supported by e.xperiments with captive brood-parasitic Village Indigobirds {Vidua chalyheata) bred in captivity and foster-reared by their normal host or by an experimental foster species and tested as adults for host choice (Payne et al. 2000). Indirect evidence in support of this hypothesis include: ( 1 ) the host specificity of Common Cuckoo (Cucidus canorus) females, but not males (Marchetti et al. 1998), and (2) the association between host species and parasite’s mitochondrial, but not nuclear DNA (Gibbs et al. 2000). Host imprinting and occurrence of recognition errors when parasitic females search for host nests (i.e., to lay eggs in nests of hosts other than the foster parents) may result in use of new hosts. This may lead to the formation of host specific races, if only females imprint on their hosts as by Common Cuckoos (Gibbs et al. 2000) or in speciation, if both males and females imprint on their hosts (i.e., indigobirds \ Vidua spp.]; Payne et al. 2002, Sorenson et al. 2003). Friedmann and Kiff (1985) suggested host imprinting may also occur in cowbirds and individual females may be host specialists. The evidence, however, is controversial for Brown- headed Cowbirds (M. ater). One study showed no genetic differentiation in mtDNA between cow- bird chicks raised by two different hosts (Gibbs et al. 1997), whereas results from another study at the same site, showed that half of the females laid in nests of a single host (Alderson et al. 1999). MtDNA haplotype frequency distributions in Shiny Cowbirds differed among parasitic chicks from nests of four hosts (Mahler et al. 2007), which indicate the occurrence of a non-random laying behavior by females of this parasite. A recent study (Mahler et al. 2009) of the host specialist Screaming Cowbird, showed that fre- quency distributions of mtDNA haplotypes dif- fered between cowbird chicks from nests of Bay wings and Chopi Blackbirds. This result indicates host choice by females is not random and female Screaming Cowbirds may preferen- tially parasitize nests of the foster parents. The number of mutations between directly related haplotypes in the latter study was small, which indicates that haplotype divergence in the popu- lation is relatively recent in evolutionary terms. Mahler et al. (2009) suggested a few Screaming Cowbirds may have switched recently from Baywings to Chopi Blackbirds, giving rise to a new race that only uses the latter host. Successful host switches require females to lay eggs in nests of a host other than the foster parent (recognition eiTors), that the new host successfully rears parasitic females, and these females subse- quently use the new foster parent as a host. We present direct evidence through genotyped eggs of Screaming Cowbird parasitism of two new hosts: Solitary Cacique {Cacicus solitarius) and Cattle Tyrant (Machetornis rixoso). Our study was conducted in an area where Screaming Cowbirds parasitize Baywings and Chopi Blackbirds. We propose parasitism of the new hosts resulted from recognition errors by Screaming Cowbird females that regularly parasitize Baywings and Chopi Blackbirds. METHODS Study Area. — The study was conducted at ‘Reserva Ecologica El BaguaF in the Province of Eormosa, Argentina (26° 18' 17.5" S; 58° 49' 51.1" W). This reserve is an open savanna of 3,300 ha in the eastern or humid Chaco region. Average annual rainfall at the study site is 1,350 mm and mean monthly temperatures vary from 16.9° C in July to 26.7° C in January. Data Collection and Analy.si.s. — We monitored nests of potential hosts of Shiny and Screaming cowbirds during the 1997-1998 to 2008-2009 breeding seasons. We found 69 nests of Bay- wings, 251 of Chopi Blackbirds, 31 of Solitary Caciques, and 30 of Cattle Tyrants. Most nests were found during construction, laying, and incubation, and were visited every 2-3 days until chicks fledged or the nest failed. Individual eggs were marked with waterproof ink and assigned to the host or to Shiny or Screaming cowbirds on the basis of background color, spotting pattern, and shape (Fraga 1983). We also genetically identified 27 and 31 Screaming Cowbird eggs in nests of Baywings and Chopi Blackbirds, respectively, by sequencing the control region of the mtDNA (Mahler et al. 2009). Screaming Cowbird eggs from nests of Solitary Caciques and Cattle Tyrants were also genetically identified. The initial assignment of Screaming Cowbird eggs on the basis of color and spotting pattern was confirmed by genetic analysis in all cases. Frequency of parasitism was calculated as number of nests with parasite eggs or chicks divided by total number of SHORT COMMUNICAl IONS 797 nests toiind, and intensity of parasitism as number ot parasitic eggs per parasitized nest. Generic Identificcition of Screoining Cowhird — We sequenced a 600 base pair fragment of the mtDNA control region using primers MBO-LI and MBO-H2 (Mahler et al. 2007). We collected Screaming Cowbird eggs and artificially incubat- ed them tor 48 hrs to obtain some embryonic development. We kept eggs at -20° C until they were processed. Embryonic tissue was obtained from the eggs and stored in DMSO buffer for posterior DNA extraction following a standard ethanol protocol (Miller et al. 1988). PCR reactions were performed as described by Mahler et al. (2009). We sequenced the amplified products on an Applied Biosystems Model 3100 Genetic Analyzer using ABI Big Dye™ Termi- nator Chemistry. We compiled the sequences in Bioedit Version 7.0.5. 3 software (Hall 1999) and compared them with those deposited in the EMBL, GenBank, under accession numbers EU199785-EUI99795 (Mahler et al. 2009). RESULTS Screaming Cowbirds regularly parasitized Baywings and Chopi Blackbirds at our study site. The frequency of parasitism of Baywing nests was 80% (n = 69 nests) with an intensity of parasitism of 2.4 ± 0.2 eggs per parasitized nest (x ± SE, n = 51 nests). Erequency and intensity of parasitism of Chopi Blackbird nests were 46% (/? = 251 nests) and 3.0 ± 0.2 eggs (/? = 116 nests), respectively. Shiny Cowbirds parasitized only one of 31 nests of the Solitary Cacique (3%). We recorded one event of Screaming Cowbird parasitism in one nest of Solitary Caciques in December 2007. Baywings had visited this nest several times during the previous days and had tried, unsuc- cessfully, to usurp the nest. Screaming Cowbird parasitism occurred before host laying and, after parasitism, the caciques deserted the nest. The egg was collected and incubated, DNA extracted, and the control region sequenced. The .sequence of the mtDNA control region corresponded to Scream- ing Cowbirds’ haplotype H5 (Mahler et al. 2009). We also monitored 30 nests of Cattle Tyrants but Shiny Cowbirds parasitized none of them. We recorded three events of Screaming Cowbird parasitism of one Cattle Tyrant nest during laying in November 2008. This nest was <1 m from an active Chopi Blackbird nest that was also multiple parasitized by Screaming Cowbirds. Cattle Ty- rants incubated the Screaming Cowbird’s eggs of which one hatched but the chick disappeared from the nest 48-72 hrs after hatching. We collected the other two eggs, one of which was rotten and we could not extract DNA from it. The .sequence of the mtDNA control region of the other egg corresponded to Screaming Cowbirds’ haplotype HI (Mahler et al. 2009). DISCUSSION We confirmed Screaming Cowbird parasitism of two new hosts. Solitary Cacique and Cattle Tyrant, in an area where this parasite regularly uses Baywings and Chopi Blackbirds. Solitary Caciques and Cattle Tyrants were previously reported as parasitized by Shiny Cowbirds, but until now they had not been mentioned as parasitized by Screaming Cowbirds (Ortega 1998). We propose these parasitism events resulted from recognition errors by female Screaming Cowbirds that regularly parasitize Baywings and Chopi Blackbirds. A pair of Baywings had frequently visited the nest of Solitary Caciques before Screaming Cowbird parasitism occuiTed, and the nest of the Cattle Tyrants was near an active nest of Chopi Blackbirds that had been previously parasitized by Screaming Cowbirds. Baywings seldom build nests but rather exploit a wide variety of covered nesting sites, including old nests built by other species and holes in trees (Eraga 1988, De Marsico et al. 2010). Baywings at our study site usually use old nests of Greater Thornbirds {Pliacellodomns ruber) and Little Thornbirds (P. sihilatrix) but, at times they usurp and use nests of Solitary Caciques and becards {Pachyromphus spp.) (Di Giacomo 2005). This may promote recognition eiTors by Screaming Cowbird females searching for nests occupied by its main host. Baywings also u.se holes in trees, which may have favored colonization of hole- nesting Chopi Blackbirds. The first step in the process of colonization of a new host would be occurrence of recognition errors when parasitic females search for host nests (i.e., to lay eggs in nests of a host species other than their foster parents). We show this type of error may occur at low frequencies in Screaming Cowbirds. However, the low frequency of para- sitism of Solitary Caciques and Cattle Tyrants also suggests the other steps necessary for parasitism of a new host (i.e., rearing parasitic females that subsequently use the same species as 798 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4. December 2010 a host; Payne et al. 2000, 2002; Sorenson et al. 2003) have not yet occurred in these hosts. The three known hosts of Screaming Cowbirds {Baywings, Chopi Blackbirds, and Brown-and- yellow Marshbirds) are cooperative breeders (Orians et al. 1977, Fraga 1991, Di Giacomo 2005), which suggests competition for food with nest mates is critical for Screaming Cowbirds. Thus, this parasite has only been successful in hosts’ nests where competition is less intense, as breeding pairs have one or more helpers at the nest that contribute to chick provisioning (De Marsico and Reboreda 2008). This hypothesis requires further investigation. ACKNOWLEDGMENTS We thank Alparamis SA and Aves Argentinas/ Asociacion Omitologica del Plata for allowing us to conduct this study at Reserva El Bagual. We thank Peter Lowther and Spencer Sealy for helpful comments on a previous version of this manuscript. BM and JCR are Research Fellows of CONICET. This study was supported by Agencia Nacional de Promocion Cientifica y Tecnolo- gica (Grant 06-00215) and the University of Buenos Aires (Grant XI 84). LITERATURE CITED Alderson, G. W., H. L. Gibbs, and S. G. Sealy. 1999. Determining the reproductive behaviour of individual Brown-headed Cowbirds using microsatellite DNA markers. Animal Behaviour 58:895-905. De Marsico, M. C. and J. C. Reboreda. 2008. Dif- ferential reproductive success favours strong host preferences in a highly specialized brood parasite. Proceedings of the Royal Society, Series B 275:2499- 2506. De Marsico, M. C., B. Mahler, and J. C. Reboreda. 2010. Reproductive success and nestling growth of Baywings parasitized by Screaming and Shiny cow- birds. Wifson Journal of Ornithology 122:417-431. Di Giacomo, A. G. 2005. Aves de la Reserva El Bagual. Pages 201-465 in Historia natural y paisaje de la Reserva El Bagual, Provincia de Formosa, Argentina (A. G. Di Giacomo and S. F. Krapovickas, Editors). Temas de Naturaleza y Conservacion 4. Aves Argentina.s/Asociacion Omitologica del Plata, Buenos Aires. Fraoa, R. M. 1983. The eggs of the parasitic Screaming Cowbird (Molothrus nifoa.xilloris) and its host, the Bay-winged Cowbird (M. hculiii.s): is there evidence for mimicry? Journal fiir Ornitologie 124:187-193. Fraga. R. M. 1986. The Bay-winged Cowbird (Molothru.s hcuHns) and its brood parasites: interactions, coevolu- tion and comparative efficiency. Thesis. University of California at Santa Barbara. California, USA. Fraga, R. M. 1988. Nest sites and breeding success of Bay- winged Cowbirds (Molothrus hadius). Journal liir Ornithologie 129:175-183. Fraga, R. M. 1991. The social system of a communal breeder, the Bay-winged Cowbird .Molothrus hadius. Ethology 89:195-210. Fraga, R. M, 1998. Interactions of the parasitic Screaming and Shiny cowbirds (Molothrus rufoa.xillaris and M. bonariensis) with a shared host, the Bay-winged Cowbird (Molothrus hadius). Pages 173-193 in Parasitic cowbirds and their hosts. Studies in coevo- lution (S. I. Rothstein and S. K. Robinson, Editors). Oxford University Press, New York, USA. Friedmann, H. 1963. Host relations of the parasitic cowbirds. U.S. National Museum Bulletin 233. Friedmann, H. and L. F. Kief. 1985. The parasitic cowbirds and their hosts. Proceedings of the Western Foundation of Vertebrate Zoology 2:225-304. Gibbs, H. L, P. Miller, G. Alderson, and S. G. Sealy. 1997. Genetic analysis of Brown-headed Cowbirds Molothrus ater raised by different hosts: data from mtDNA and microsatellite DNA markers. Molecular Ecology 6:189-193. Gibbs, H. L., K. Marchetti, M. D. Sorenson, M. de L. Brooke. N. B. Davies, and H. Nakamura. 2000. Genetic evidence for female host-specific races of the Common Cuckoo. Nature 407:183-186. Hall, T. A. 1999. BioEdit: a user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids Symposium Series 41:95-98. Mahler, B., V. A. Confalonieri, I. J. Lovette, and J. C. Reboreda. 2007. Partial host fidelity in nest selection by the Shiny Cowbird (Molothrus bonariensis), a highly generalist avian brood parasite. Journal of Evolutionary Biology 20:1918-1923. Mahler. B., Y. Sarquis Adamson, A. G. Di Giacomo, V. A. Confalonieri, and J. C. Reboreda. 2009. Utilization of a new host in the host-specialist brood parasite Molothrus rufoaxillaris: host switch or host acquisition? Behavioral Ecology and Sociobiology 63:1603-1608. Marchetti, K., H. L. Gibbs, and H. Nakamura. 1998. Host-race formation in the Common Cuckoo. Science 282:471^72. Mason, P. 1980. Ecological and evolutionary aspects of host selection in cowbirds. Thesis. University of Texas, Austin, USA. Mermoz, M. E. and G. j. Eernandez. 2003. Breeding success of a specialist brood parasite, the Screaming Cowbird, parasitizing an alternative host. Condor 105:63-72. Miller, S. A., D. D. Dykes, and H. F. Polesky. 1988. A simple salting out procedure for extracting DNA from human nucleated cells. Nucleic Acids Research 16:1215. Nicolai, J. 1964. Der brutparasitismus dcr Viduinae als ethologisches problem. Pragungsphanomene als fak- toren der rassen- und artbildung. Zeitschrift fiir Ticrpsychologie 21:1 29-204. Orians, G. H., C. E. Orians, and K. J. Orians. 1977. Helpers al the nest in some Argentine blackbirds. Pages 137-151 in Evolutionary ecology (B. Stone- hou.se and C. Perrins, Editors). Macmillan Press, London, United Kingdom. SHORT COMMUNICATIONS 799 Ortega, C. P. 1998. Cowbirds and other brood parasites. University of Arizona Press, Tucson, USA. Payne, R. B., L. L. Payne, J. L. Woods, and M. D. Sorenson. 2000. Imprinting and the origin of para- site-host species associations in brood-parasitic indigobirds. Vidua chalxbeata. Animal Behaviour 59:69-81. Payne, R., B. K. Hustler, R. Stjernstedt, K. M. Sefc, AND M. D. Sorenson. 2002. Behavioural and genetic evidence of a recent population switch to a novel host species in brood-parasitic indigobirds Vidua chaly- heaia. Ibis 144:373-38.5. Sick, H. 1985. Ornitologia brasileira: uma introduc9ao. University of Brasilia, Brazil. Slagsvold, T. and B. T. Hansen. 2001. Sexual imprinting and the origin of obligate brood parasitism in birds. American Naturalist 158:354-367. Sorenson, M. D., K. M. Sefc, and R. B. Payne. 2003. Speciation by host switch in brood parasitic indigo- birds. Nature 424:928-931. The Wilson Journal of Ornithology 1 22(4):799-803, 2010 Snakes are Important Nest Predators of Dickcissels in an Agricultural Landscape Page E. Klug,''“"^ L. LaReesa Wolfenbarger,‘ and John P. McCarty' ABSTRACT. — We used video cameras to monitor 33 Dickcissel (Spiza americana) nests during 2003-2004 in the highly-fragmented, agricultural ecosystem of eastern Nebraska and western Iowa. Nine nests fledged young, 20 were completely depredated, three were partially depredated, and one was abandoned due to ants. Nine snakes, six small mammals, six common raccoons (Procyon lotor), two Brown-headed Cowbirds (Molothrus ater), and one American mink (Neovison vison) were documented as nest predators. These results suggest a diversity of predators is responsible for depredation of Dickcissel nests with snake predation being an important cause of nest failure. Received 2] December 2009. Accepted 24 May 2010. Nest predation is the leading cause of repro- ductive failure for many bird species across numerous regions and habitats (Martin 1993), and may be an important factor in avian population declines (Heske et al. 2001 ). Grassland birds are of special concern becau.se they have undergone declines greater than any group of birds in North America (Knopf 1994) and have shown decreasing reproductive rates with in- creased habitat fragmentation (Herkert et al. 2003). Nearly 80% of grasslands in the Great Plains of North America have been converted to agricultural u.se, and <4% of the tallgrass prairie ' Department of Biology, University of Nebraska at Omaha, Omaha, NE 68182, USA. "Current address: Department of Biology. University of Notre Dame, Notre Dame, IN 46556, USA. ■’ Corresponding author; e-mail: pklug@nd.edu remains (Samson and Knopf 1994). Grassland birds nesting in the remaining grasslands often experience low reproductive success due to landscape features in agroecosystems that attract a diversity of nest predators (Bergin et al. 2000). Previous studies have identified predators of grassland bird nests, but more studies are necessary to understand how predator communi- ties differ among regions and what factors may contribute to these differences (Weatherhead and Blouin-Demers 2004). At least two video studies of nesting birds that have occun'ed in North American grasslands were in areas that could be considered tallgrass prairie, but were conducted at the edge of the historical tallgrass prairie range. Thompson and Burhans (2003) documented predators of shrub-nesting birds in old fields in Missouri, and Renfrew and Ribic (2003) docu- mented predators of five grassland species in grazed pastures in the Driftless Area of Wiscon- sin; both studies were conducted along the ecotone with eastern deciduous forests. Pietz and Granfors (2000) studied nest predation on 10 species of grassland pas,serines and Grant et al. (2006) studied two passerine species in mixed- grass prairie in North Dakota. To date, no studies have been conducted in the highly fragmented, agricultural ecosystem in the core of the former tallgrass prairie biome. Our objective was to identify predators of Dickcis.sel (Spiz.a americano) nests within small grassland patches interspersed within intensive row crop agriculture. Dickcissels are of con.servation concern and have shown 800 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 4. December 2010 TABLE 1. Nest predators recorded on video with corresponding and nestling stages. number of Dickcissel nests depredated at incubation Predator Incubation Nestling # Events Snakes Fox snake (Elaphe viilpinci) 3 2 5 Plains garter snake (Thanmophis rcidi.x) 0 2 2 Eastern yellowbelly racer {Coluber constrictor flavivenrris) 0 1 1 Bullsnake (Pituophis catenifer) 1 0 1 Small mammals Thirteen-lined ground squirrel (Spermophilus tridecemlineatiis) 2 1 3 Franklin’s ground squirrel {S. frtmklinii) 0 1 1 Cricetid rodent (Peromyscus spp.) 2 0 2 Mid-sized mammals Common raccoon {Procyon lotor) 4 2 6 Other Brown-headed cowbird (Molothrus ater) 1 1 2 American mink (Neorison vison) 0 1 1 Totals 13 1 1 24 population declines similar to other grassland passerines, but their predator community has not been documented on video. Dickcissels are an ideal model for understanding how nest predator communities vary among regions, landscapes, or habitat types because they inhabit marginal grasslands in extremely fragmented landscapes as well as relatively unfragmented landscapes. METHODS We monitored Dickcissel nests at DeSoto National Wildlife Refuge, Boyer Chute National Wildlife Refuge, Allwine Prairie Preserve, Cum- ing City Cemetery and Nature Preserve, and a privately owned Conservation Reserve Program (CRP) parcel in eastern Nebraska and western Iowa (King et al. 2009). The grasslands were managed with prescribed burning but no livestock grazing. Vegetation in the grasslands was either dominated by native warm-sea.son grasses or by exotic cool-season grasses with variation in forb density (Klug 2005). We video-documented predators of Dickcissel nests in 15 grassland fragments ranging from 4 to 45 ha (.v ± SE = 20.3 ± 3.1). The landscape within 1,600 m of grassland fragments included six major land cover types based on the evaluation of orthophotographs in ArcMap CIS (scale of 1:1500) by Klug et al. (2009). We classified the landscape as an agroecosystem becau.se 20-64% (.v ± SE = 52.5 ± 3.9) of the surrounding land was in crop production (e.g., corn or soybeans). Grasslands covered 2-43% (.v ± SE = 17.8 ± 2.5) and trees 6^2% (.V ± SE = 15.7 ± 2.5) of the landscape. Human development covered 0.5-22% (.v ± SE = 4.0 ± 1 .7), and roads <2% (T ± SE = 1 .2 ± 0. 1 ), whereas wetlands and water bodies covered 1- 14% (T ± SE = 5.4 ± 1.0) of the landscape. We used 24-hr time-lapse video systems to monitor nests and identify predators removing or consuming eggs and nestlings. We constructed six battery-powered video systems, which included a camera (V-1214-IR Weatherproof Bullet Camera with infrared illumination, Marshall Electronics, El Segundo, CA, USA) attached to a time-lapse video recorder (SSC-960 Time Lapse VCR, Samsung Electronics America Inc., Ridgefield Park, NJ, USA) by 8.5 m of cable (buried under vegetation and litter) and powered by a sealed lead acid battery connected by a power converter (375w PowerVerter, Tripp-Lite World Headquar- ters, Chicago, IE, USA). The camera was attached to a wood dowel and placed within 10 cm of the nest so the camera was completely obscured by vegetation, but vegetation did not completely obscure the field of view. We minimized disturbance at the nest by setting up the system in <20 min between lOOO and 1600 CST, in dry, moderate temperatures. We set up systems during both incubation and nestling stages, but only after females had incubated for 3-5 days to minimize abandonment. Nest contents were checked re- SHORT COMMUNICA I IONS 801 TABLE 2. Nest predators ot grassland passerines documented on videt) and Wisconsin. in Nebraska/Iowa, North Dakota, Missouri, Nebraska Predator North Dakota^ North Dakota' Missouri*^ Wisconsin' Snake (Elaphe rulpina, E. ohsoleia, Thaninophis spp., Coluber constrictor. Lanipropeltis getnla, L. calligaster, Pitnophis catenifer) 9 2 33 3 Common raccoon (Procyon lotor) 6 1 4 8 Squirrel (Sperniophiliis tridecemUneatus. S. franklinii, Scinrns niger) 4 13 1 1 1 4 Brown-headed Cowbird (Molotlirns ater) 2 2 5 1 1 Mouse or vole (Peroniyscns spp., Zapns spp., Microtns spp.) 2 2 3 2 Long-tailed weasel/American mink (Mnstela frenata, Neovison vison) 1 1 1 American badger (Taxidea taxiis) 2 3 7 Bird ot prey (Biiteo spp., Circus cyaneiis, Tyto alba) 2 1 3 2 Red fox/coyote/domestic dog (Vnlpes vitlpes, Canis latrans. C. lupus fatniiiaris) 2 1 Domestic cat (Pells catiis) 1 Striped skunk (Mephitis mephitis) 1 Virginia opossum (Didelphis virginiana) ] White-tailed deer (Odocoileus virginianus) 2 7 Unidentified 1 7 1 Totals 24 27 30 46 24 Pietz and Grantors (2000). Grant et al. (2006). Thompson and Burhans (200.t). Renfrew and Ribic (2003). motely by attaching a portable monitor to the system every 24 hrs given that the tape ran out after 24 hrs. We archived the tape to identify the predator using a high-resolution monitor and slow motion function if the contents differed from the previous day. We recorded time, nest stage, nest contents, and contents removed for each depre- dation event. We estimated daily survival rates (DSR) for nests using Program MARK with the design matrix tools and the logit-link function (Dinsmore et al. 2002), and calculated variance using the Delta method (Powell 2007). We also estimated DSR tor nests without video systems to better understand the influence of the equipment on nest survival. Means ± SE are presented. RESULTS We video-monitored 33 nests over 2 years (2003 = 15 nests, 2004 = 18 nests) for 287 observation days during incubation (// = 148 days) and nestling (// = 139 days) stages. Nests were monitored an average of 12.9 ± 1.0 days. We placed cameras at 19 nests that recorded both incubation and nestling stages, 1 1 nests during the incubation stage only, and three nests during the nestling stage only. The DSR of nests monitored with video was 0.946 ± 0.01 for an overall survival of 0.333 ± 0.07 if we extrapolated to a 20-day nesting cycle (i.e., the average length of a Dickcissel nesting cycle). The probability of predation for nests with cameras was similar to nests not monitored with video systems (DSR = 0.934 ± 0.005, overall survival rate = 0.255 ± 0.03). Twenty of the nests monitored completely failed due to depredation, three nests were partially depredated before eventual Hedging of young, nine nests fledged young with no depre- dation, and one nest was abandoned after an ant infestation. Nine snakes, six small mammals, six common raccoons {Procyon Infor), two Brown-headed Cowbirds (Molotlirus ater), and one American mink (Neovison vison) were documented as nest predators (Table I ). We documented one instance of a multiple predation at one nest (i.e., a Brown- headed Cowbird removed three Dickcissel eggs prior to a thirteen-lined ground squirrel \Spernw- philiis tridecewUneatus] consuming the remaining 2 Brown-headed Cowbird and 1 Dickcissel egg)". We documented partial predation by a fox snake (Elaphe vulpiiia) late in the nesting cycle at two nests, which caused the surviving nestlings to fledge prematurely. A fox snake at another nest was documented consuming one egg and leaving three, which later hatched and Hedged. We 802 THE WILSON JOURNAL OL ORNITHOLOGY • Voi 122, No. 4. December 2010 observed one female Brown-headed Cowbird removing Dickcissel nestlings without laying eggs. We recorded 13 and 1 1 depredations at the incubation and nestling stages, respectively (Ta- ble 1 ). Depredation by cricetid rodents and raccoons were nocturnal, whereas depredation by ground squirrels, mink, and cowbirds were diurnal. All depredations by snakes were diurnal except two by fox snakes, which were nocturnal. DISCUSSION The potential predator community in agricul- tural systems is diverse; thus, the composition of the community, and identity of the dominant predator, will ultimately dictate nest predation risk (Klug et al. 2009). We found at least 10 species responsible for predation of Dickcissel nests, including both obligate grassland species (e.g., ground squirrels and snakes) and more wide- ranging habitat generalists (e.g., common raccoon and Brown-headed Cowbird; Table 1). Our sam- ple size of video-monitored nests (/; = 33) is consistent with other studies in the grasslands of North America (Table 2), but it is likely that additional predator species would be detected with a larger sample size. For example, striped skunk (Mephitis mephitis), Virginia opossum (Didelphis virginiana), American Crow (Corvus brachyrhynchos), and coyote (Canis latrans) have been documented as nest predators (Renfrew and Ribic 2003, Peterson et al. 2004, Staller et al. 2005), but were not documented at our study sites regardless of their presence on predator surveys (Klug et al. 2009). However, given the consisten- cy of snake events across years, it is unlikely a larger sample would change our conclusion that snakes are among the most important nest predators for grassland birds in this region. Information about the predator community is important for understanding variation in nest predation and to increase productivity of nesting birds (Lahti 2009). However, recognizing the importance of the predator community also brings challenges. Studies that document predation of grassland bird nests consistently discover wide- ranging generalist predators including Brown- headed Cowbird and common raccoon. Thus, efforts directed at increasing reproductive success of birds have often focused on decreasing the population size of species that have increased due to agricultural land use (Heske et al. 1999, Kosciuch and Sandercock 2008). Our study, and other nest predation studies in grassland habitats. has revealed considerable nest predation by snakes and ground squirrels, which are essential members of prairie communities (Table 2). Rec- ommendations directed at decreasing the influ- ence of resident predators in grasslands where birds are nesting are not straightforward given that many of these species are also targets of conservation. For example, fox snakes and Franklin’s ground squirrels (Spermophihis frank- linii) are native members of the tallgrass prairie (Martin et al. 2003) and efforts to control these predators to increase bird populations must proceed with caution. ACKNOWLEDGMENTS We thank the staff of Boyer Chute and DeSoto National Wildlife refuges for their support. We also thank the American Museum of Natural History Frank M. Chapman Memorial Fund, Sigma Xi Grants-In-Aid of Research. Center for Great Plains Studies Grant-In-Aid for Graduate Students, University of Nebraska at Omaha, USDA Biotechnology Risk Assessment Research Grants Program, and the U.S. Fish and Wildlife Service for financial support. LITERATURE CITED Bergin, T. M., L. B. Best, K. E. Freemark, and K. J. Koehler. 2000. Effects of landscape structure on nest predation in roadsides of a midwestern agroecosystem: a multiscale analysis. Landscape Ecology 15:131-143. Dinsmore, S. J., G. C. White, and F. L. Knopf. 2002. Advanced techniques for modeling avian nest survival. Ecology 83:3476-3488. Grant, T. A., E. M. Madden, T. L. Shaffer, P. J. Pietz, G. B, Berkey, and N. j. Kadrmas. 2006. Nest survival of Clay-Colored and Vesper sparrows in relation to woodland edge in mixed-grass prairies. Journal of Wildlife Management 70:691-701. Herkert, j. R., D. L. Reinking, D. A. Wiedenfeld, M. Winter, J. L. Zimmerman, W. E. Jensen, E. J. Finck, R. R. Koford. D. H. Wolfe. S. K. Sherrod. M. A. Jenkins, J. Faaborg, and S. K. Robinson. 2003. Effects of prairie fragmentation on the nest success of breeding birds in the midcontinental United States. Conservation Biology 17:587-594. Heske, E. J., S. K. Robinson, and J. D. Brawn. 1999. Predator activity and predation on songbird nests on forest-field edges in east-central Illinoi.s. Landscape Ecology 14:345-354. Heske, E. J.. S. K. Robinson, and J. D. Brawn. 2(X)1. Nest predation and neotropical migrant songbirds: piecing together the fragments. Wildlife Society Bulletin 29:52-6 1 . Klug, P. E. 2005. The effects of local grassland habitat and surrounding landscape composition on the predators of grassland bird nests. Thesis. University of Nebraska, Omaha, USA. Klug, P., L. L, Wolfenbarger, and J. P. McCarty. 2009. The nest predator community of grassland birds SHORT COMMUNICATIONS 803 responds to agroecosysteni habitat at multiple scales. Ecography 32:973-982. Knopf. F. L. 1994. Avian assemblages on altered grasslands. Studies in Avian Biology 15:247-257. Kosciuch, K. L. and B. K. Sandercock. 2008. Cowbird removals unexpectedly increase productivity of a brood parasite and the songbird host. Ecological Applications 18:537-548. Lahti. D. 2009. Why we have been unable to generalize about bird nest predation. Animal Conservation 12:279-281. Martin. J. M.. E. J. Heske, and J. E. Hofmann. 2003. Franklin s ground squirrel (Spennophiliis franklinii) in Illinois: a declining prairie mammal? American Midland Naturalist 150:130-138. Martin. T. E. 1993. Nest predation and nest sites. Bioscience 43:523-533. PlETZ. P. J. and D. a. Granfors. 2000. Identifying predators and fates of grassland passerine nests using miniature video cameras. Journal of Wildlife Manage- ment 64:71-87. Peterson. B. L., B. E. Kus. and D. H. Deutschman. 2004. Determining nest predators of the Least Bell's Vireo through point counts, tracking stations, and video- photography. Journal of Field Ornithology 75:89- 95. Powell, L. 2007. Approximating variance of demographic parameters using the della method: a reference for avian biologists. Condor 109:949-954. Renfrew. R. B. and C. A. Ribic. 2003. Grassland passerine nest predators near pasture edges identified on video- tape. Auk 120:371-383. Samson, F. and F. Knopf. 1994. Prairie conservation in North America. Bio.science 44:418^23. Staller, E., W. Palmer, J. Carroll, R. Thornton, and D. Sisson. 2005. Identifying predators at Northern Bobwhite nests. Journal of Wildlife Management 69:124-132. Thompson III, F. R. and D. E. Burhans. 2003. Predation of songbird nests differs by predator and between field and forest habitats. Journal of Wildlife Management 67:408-416. Weatherhead, P. j. and G. Blouin-Demers. 2004. Understanding avian nest predation: why ornitholo- gists should study snakes. Journal of Avian Biology 35:185-190. The Wilson Journal of Ornithology 122(4):803-806, 2010 Interspecific Cavity-sharing Between a Helmeted Woodpecker {Dryocopus galeatiis) and Two White-eyed Parakeets {Aratinga leucophthalma) Kristina L. Cockle’-^-'* ABSTRACT. — Cavity-nesting birds may frequently compete for a limited supply of nest and roost cavities in trees, but interspecific sharing of these cavities has rarely been reported. The globally vulnerable Helmeted Woodpecker (Dryocopus galeatus), a little-known Atlantic Forest endemic, is believed to be threatened by nest-site competition; however, little is known about its ecology or natural history. 1 report an ob.servation of a female Helmeted Woodpecker roosting with two White-eyed Parakeets (Aratinga leucophthalma) in their non-excavated (natural) nest cavity at Cruce Caballero Provincial Park. Argentina, and discuss possible impli- cations for ecology and con.servation of this rare ' Center for Applied Conservation Research. Department of Forest Science, University of British Columbia, 2424 Main Mall, Vancouver, BC. V6T IZ4, Canada; e-mail: kri.stinacockle@gmail.com -Proyecto Selva de Pino Parana, Velez Sarsfield y San Jurjo S/N, San Pedro, Misiones, (3352), Argentina. -’Fundacion de Historia Natural Felix de Azara. Departa- mento de Ciencias Naturales y Antropologi'a. Universidad Maimonides. Valentm Virasoro 732 (CI405 BDB), Buenos Aires, Argentina. woodpecker. Received 22 Januarv 2010. Accepted 20 April 2010. Tree cavities are u.sed by many bird species for nesting and roosting, and supply of cavities may be a key re.source that limits populations (Newton 1994) and structures communities (Martin et al. 2004, Aitken and Martin 2008). Competition within species for a limited supply of cavities may lead to physical conflict resulting in injuries to adults and death of their young (Snyder et al. 1987, Heinsohn and Legge 2003). Cavity limita- tion may also create conditions that favor co- operative breeding (Heinsohn and Legge 2003). Cavity usurpation is common among competing .species (Snyder et al. 1987, Renton and Brighu smith 2009. Strubbe and Matthysen 2009), but cavities could also be shared between species (Robinson et al. 2006). Interspecific cavity- shaiing would allow individuals to avoid costs as.sociated with aggressive usurpation of cavities. 804 THE WILSON JOURNAL OL ORNITHOLOGY • Vul. 122. No. 4. December 2010 Sharing could also provide benefits including predator vigilance. The Helmeted Woodpecker (Dtyocopits galeatiis) is rare, little-known, globally vulnerable, and endemic to the Atlantic Forest of Paraguay, Argentina, and Brazil (BirdLife International 2008). It is considered a bird of tall forests (Winkler et al. 1995) and may be most common in mature forest in the Province of Misiones, Argentina (Bodrati and Cockle 2006). Little else is known about its natural history or ecology, but competition for nest sites with the larger and more abundant Lineated Woodpecker (D. lineatiis) has been pro- posed as a key thieat (BirdLife International 2008). OBSERVATIONS 1 observed a female Helmeted Woodpecker roosting overnight with a pair of White-eyed Parakeets (Aratinga leiicophthalma), while the latter were incubating in a naturally occurring, non-excavated cavity in a large live branch of grapia (Apuleia leiocarpa). The site was in Cruce Caballero Provincial Park (26° 3 L S, 53° 58' W), Province of Misiones, Argentina, in primary mixed Atlantic Forest with laurel (Nectandra and Ocotea spp.), guatambti {Balfourodendron riedelianum), and Parana pine {Araucaria an- giistifolia) (Cabrera 1976). The Helmeted Wood- pecker is scarce in the park, and the White-eyed Parakeet is abundant (Bodrati et al. 2010). A pair of White-eyed Parakeets laid at least two eggs in the cavity in December 2008 (Cockle et al. 2010), but their nest failed. The cavity was empty when checked for contents on 21 October and 2 November 2009. The cavity was shared for at least two consecutive nights on 3 and 4 December 2009 by a Helmeted Woodpecker and a pair of White- eyed Parakeets. The cavity contained one white egg on 3 December 2009. The following day, 1 arrived at the cavity at 0520 hrs, 20 min before dawn. The sky was dark, cloudy, and threatening to rain. A female Helmeted Woodpecker appeared from inside the cavity at 0525 hrs. She remained at the cavity entrance until 0537 hrs when she tlew out. clung briefly to the roots of an epiphytic giiembe {Philodendron spp.) about 2 m distant, then returned to the cavity. She remained in or near the cavity entrance until 0752 hrs and 1 could clearly distinguish her brown face from the chin to above the eye. She left the cavity three times to peck at the giiembe roots or move up the branch within I m of the cavity entrance. Three times she seemed to glean insects just inside the cavity entrance. She did not descend to the floor of the cavity (egg chamber) and remained visible Just inside the entrance. It rained from 0616 to 0625 hrs but the sun shone directly on the cavity from 0712 hrs onwai'ds. At 0752 hrs the woodpecker became agitated, jumping forward and backward inside the cavity. She flew out of the cavity and disappeared into the forest. Two White-eyed Parakeets immediately emerged from the bottom of the cavity and flew away in silence. Neither the woodpecker nor the parakeets vocalized that morning, and there was no sign of aggression. 1 returned to the cavity at 0510 hrs the following day accompanied by A. A. Bodrati. The sky was clear and the nest tree was illuminated by bright moonlight. We could immediately see the Helmeted Woodpecker in the cavity, apparently awake. She flew from the nest cavity 10 min before sunrise at 0530 hrs. The two White-eyed Parakeets emerged immediately from the bottom of the nest. They stayed 6 min at the entrance and then flew away in silence. The cavity was 12.4 m above ground in a live healthy grapia that measured 30 m in total height and 87 cm in diameter at breast height. It had one lateral entrance 6 cm wide X 38 cm tall. The cavity was 90 cm deep vertically and 14 cm deep horizontally. An epiphytic giiembe higher on the branch was rooted inside the cavity. 1 believe the parakeets roosted in the egg chamber and incubated the egg, and the woodpecker roosted above them near the cavity entrance. J. M. Klavins showed me another roost cavity at Cruce Caballero Provincial Park used by a Helmeted Woodpecker in May 2005. That cavity seemed to have been formed where a branch broke off the tree, and either a woodpecker excavated the knot-hole or the cavity formed through decay processes. It was about 12 m high in the main trunk of a live healthy tree. 59 cm in diameter, 24 m tall, at the edge of the park. It was later occupied by a nest of bees in 2006-2007, a roosting male Lineated Woodpecker in 2008 and 2009, and a successful nest of Scaly-headed Parrots {Pionu.s niaxiniiliani) in 2009 (KLC and J. M. Lammertink, unpubl. data). DISCUSSION Reports of interspecific cavity-sharing are uncommon. Skutch (1969) reports a cavity in Central America to which a pair of Streak-headed Woodcreepers {Lepidocolaptes souleyetii) and a SHORT COMMUNICATIONS 805 Tawny- winged Woodcreeper (Dein/rocincla ana- hotiiia) brought nest material alternately; the Tawny-winged Woodcreeper incubated the two eggs and raised the single nestling, a Streak- headed Woodcreeper. Robinson et al. (2006) report a mixed brood of Red-breasted Nuthatches (Sitta canadensis) and Mountain Chickadees (Poecile gambeli) raised in the same cavity by parents of both species in North America. I observed several incidences of probable interspe- cific nest-usurpation during tour breeding seasons studying cavity-nesting birds in the Atlantic Forest, but this is the first time I observed simultaneous interspecific cavity-sharing among birds, and it appears to be the first such report for South America. Much remains to study about the role of tree cavities in the ecology and conservation of the Helmeted Woodpecker (Winkler et al. 1995). Apparent nests were reported in a cavity 7.8 cm in diameter, 2.3 m high in an unidentified tree at Iguazu National Park in Misiones in spring 1985 (adults seen entering and leaving cavity without food) and by the side of a road in adjacent Igua9u National Park in Brazil in November 1988, but no further descriptions are available (Collar et al. 1992; Chebez 1994; Winkler et al. 1995; Hernan Casahas, pers. comm.). A photograph of the first cavity by Martin Adamovsky (published in Chebez 1994) shows it to have a circular entrance, apparently excavated by a woodpecker. Juan Mazar Barnett (pers. comm.) reported a male quietly excavating a cavity in a dead branch of a live laurel (Nectandra spp.) tree at Puerto Peninsula, Misiones, on 31 October 1994. A female was nearby at times, always passive, and did not excavate. Madrono Nieto and Esquivel (1995) reported two pairs nesting in spring 1994 at Reserva Natural del Bosque Mbaracayu in Paraguay without further details. The roost hole shared with the White-eyed Parakeets was not excavated by a woodpecker. The Helmeted Woodpecker has a narrow-based, weak-tipped bill and forages mainly by scaling and pecking (Brooks et al. 1993). Further studies may find that Helmeted Woodpeckers frequently reuse old cavities, entering into competition with secondary cavity-nesting birds such as parrots. Cavities may be in short supply in the Atlantic Forest: high-grade logging has already removed the largest trees with the greatest potential for cavity formation from nearly all remaining forest (Cockle et al. 2010). ACKNOWLEDGMENTS I thank A. A. Bodrati for finding the White-eyed Parakeet nest in 2008 and sharing in my observations in 2009; Kathy Martin for di.scussions and supervision; D. W. Cockle for building the cameras used to inspect cavities; J. M. Klavins, J. M. Segovia, and E. A. Jordan for sharing their observations or helping in the field; and Kathy Martin, A. R. Norris, Hernan Casanas, Juan Mazar Barnett and J. M. Lammertink for unpublished data or comments on the manuscript. This work was financed by a Re.search Grant from the British Ornithologists’ Union, a Conservar La Argentina grant from Aves Argentinas and BirdLife International, a Namkoong Family Fellowship, a Donald S. McPhee Fellowship, Rufford Foundation Booster Grant, Columbus Zoo and Aquarium Conservation Fund, National Science and Engineering Research Council of Canada (NSERC) Canada Graduate Scholarship, Killam Pre- doctoral Fellowship, Neotropical Bird Club Conservation Award to A. A. Bodrati, and research support from NSERC and Environment Canada to Kathy Martin. Equipment was loaned or donated by Environment Canada, Idea Wild, and the Area de Manejo Integral de la Reserva de Biosfera Yaboty. Research was authorized by the Misiones Minis- terio de Ecologia, RNR y Turismo. LITERATURE CITED Aitken, K. E. H. and K. Martin. 2008. Resource selection plasticity and community responses to experimental reduction of a critical resource. Ecology 89:971-980. BirdLife International. 2008. Threatened birds of the world 2008. CD-ROM. BirdLife International, Cam- bridge, United Kingdom. Bodrati, A. and K. Cockle. 2006. Habitat, distribution, and conservation of Atlantic Forest birds in Argentina; notes on nine rare or threatened species. Omitologia Neotropical 17:243-258. Bodrati, A., K. Cockle, J. M. Segovia, I. Roesler, J. I. Areta, and E. Jordan. 2010. La avifauna del Parque Provincial Cruce Caballero, Provincia de Misiones, Argentina. Cotinga 32:41-64. Brooks, T. M., R. Barnes, L, Bartrina, S. H. M. Butchart, R. P. Clay, E. Z, Esquivel, N. I. Etcheverry, J. C. Lowen, and j. Vincent. 1993. Bird surveys and conservation in the Paraguayan Atlantic Forest: Project CANOPY '92 final report. Study Report Number 57. BirdLife International. Cambridge. United Kingdom. Cabrera, A. L. 1976. Enciclopedia Argentina de agricul- tura y jardinen'a. Second Edition. Tomo II. Fa.sciculo I. Regiones fitogeograficas Argentinas. Editorial Acme S. A. C. 1., Buenos Aires, Argentina. Chebez, J. C. 1994. Los que se van: especies Argentinas en peligro. Editorial Albatros, Buenos Aires, Argentina. Cockle, K., K. Martin, and K. Wiebe. 2010. Selection of nest trees by cavity-nesting birds in the neotropical Atlantic Forest. Biotropica: On-line. Collar, N. J., L, P. Gonzaga, N. Krabbe, A. Madrono 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. Third Edition. 806 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122. No. 4. December 2010 International Council for Bird Preservation, Cam- bridge. United Kingdom. Heinsohn, R. and S. Legge. 2003. Breeding biology of the reverse-dichromatic, co-operative parrot Eclectus ror- atiis. Journal of Zoology, London 259:197-208. Madrono Nieto, A. and E. Z. Esquivel. 1995. Reserva Natural del Bosque Mbaracayu: su importancia en la conservacion de aves amenazadas, cuasi amenazadas y endemicas del Bosque Atlantico del Interior. Cotinga 4:52-57. Martin, K., K. E. H. Aitken, and K. L. Wiebe. 2004. Nest sites and nest webs for cavity-nesting communities in interior British Columbia, Canada: nest characteristics and niche partitioning. Condor 106:5-19. Newton, 1. 1994. The role of nest sites in limiting the numbers of hole-nesting birds: a review. Biological Conservation 70:265-276. Renton, K. and D. Brightsmith. 2009. Cavity use and reproductive success of nesting macaws in lowland forest of southeast Peru. Journal of Field Ornithology 80:1-8. Robinson, P. A., A. R. Norris, and K. Martin. 2006. Interspecific nest-sharing by Red-breasted Nuthatch and Mountain Chickadee. Wilson Bulletin 1 17:400^02. Skutch, a. F. 1969. Life histories of Central American birds. III. Families Cotingidae, Pipridae, Formicarii- dae, Furnariidae, Dendrocolaptidae, and Picidae. Pacific Coast Avifauna 35. Snyder. N. F. R. J. W. Wiley, and C. B. Kepler. 1987. The parrots of Luquillo: natural history and conservation of the Puerto Rican Parrot. Western Foundation of Vertebrate Zoology, Los Angeles, California, USA. Strubbe. D. and E. Matthysen. 2009. Experimental evidence for nest-site competition between invasive Ring-necked Parakeets {Psittacula krameri) and native nuthatches {Sitta europaed). Biological Conservation 142:1588-1594. Winkler, H., D. A. Christie, and D. Nurney. 1995. Woodpeckers: a guide to the woodpeckers of the world. Houghton Mifflin Company, Boston, Massa- chusetts, USA. The Wilson Journal of Ornithology 122(4):806-809. 2010 Response to Nestling Throat Ligatures by Three Songbirds Gabrielle L. Robinson,' Courtney J. Conway,' ""'' Chris Kirkpatrick,' and Dominic D. LaRoche' ABSTRACT. — We attempted to collect diet samples using throat ligatures from nestlings of three songbird species in a riparian woodland in .southeastern Arizona from May to August 2009. We had success with Song Sparrows (Melospiza melodiu). observed adult Yellow- breasted Chats {Icteria virens) reclaim food from nestlings, and discontinued the use of throat ligatures when we observed an adult Abert’s Towhee iPipilo aherfi) remove two, 3^-day-old ligatured nestlings from its nest. Previous studies have reported problems (e.g., aggression toward nestlings by adults) with throat ligatures, but we are the first to document removal (and subsequent nestling mortality) in response to this technique. We urge investigators to exercise caution when using throat ligatures on species for which evidence of the safety and efficacy of this method are lacking, especially when nestlings are small in size relative to adults. Received 18 February 2010. Accepted 29 April 2010. ' School of Natural Re.sources and the Environment. University of Arizona, Tuc.son, AZ 85721, USA. ’ U.S. Geological Survey, Arizona Cooperative Fish and Wildlife Research Unit. University of Arizona, Tucson. AZ 85721. USA. 'Corresponding author; e-mail: cconway(®usgs.gov Documenting the diet of birds is an important component of many studies designed to test hypotheses related to ecology and evolution of birds. Documenting the diet of nestling birds in an unbiased manner is particularly challenging, especially when investigators want to identify prey items to taxonomic levels beyond Order or want precise estimates of biomass. Investigators have used throat ligatures to collect diet samples from nestling birds since the 1930s (Kluijver 1933, Owen 1956, Mellott and Woods 1993). Fecal analysis, the use of artificial nestlings to collect food, analysis of stomach contents, and visual observation (Evans 1964) are often less effective or more invasive than throat ligatures. However, researchers have experienced several problems with throat ligatures related to both quality of the sample obtained and safety of the nestlings sampled. For example, food may slip past the ligature if the ligature is too loose (Owen 1956, Orians 1966), or food is disgorged if the sampling period is too long or the rate of provisioning too high (Orians 1966, Johnson et al. 1980). Moreover, parents will, at times. SHORT COMMUNICATIONS 807 remove and eat the food from the mouth of the ligatured nestling (Robertson 1973). Ligatures have also been tound to affect normal gaping (Johnson et al. 1980) and begging behavior (Orians and Horn 1969), and to occasionally cause death by strangulation (Moore 1986, Mellott and Woods 1993). Several authors have described accounts of aggressive behavior by adults toward nestlings with throat ligatures. Robertson (1973) mentions that Red-winged Blackbirds {AgeUiiiis plweni- ceiis) in a few cases attempted to remove ligatures from nestlings. Little et al. (2009) reported that adult Bobolinks (Dolichony.x oryzivorus) pecked and pulled at nestlings’ throat ligatures at most nests sampled. Furthermore, aggressive behavior by adults such as grasping the ligature and forcibly pulling the nestlings’ heads upwards or sideways occuired at half of the nests. Despite this aggression. Little et al. (2009) did not observe any nestling mortality resulting from the behavior of the adults. We provide the first documentation of nestling mortality resulting from removal by adults of nestlings outfitted with throat ligatures. METHODS We monitored nests of Abert’s Towhees (Pipilo aberti), Yellow-breasted Chats {Icteria virens), and Song Sparrows (Melospiza melodia) from May to August 2009 in a riparian woodland along the Santa Cruz River at Tumacacori National Historic Park (31° 34' 03" N, 111° 03' 03" W), Santa Cruz County, Arizona, USA. The elevation is ~ 1,000 m and the plant communities are cottonwood-willow iPopuhis-SaUx) riparian for- est along the river and mesquite-hackberry {Prosopis-Celtis) woodland in the adjacent up- land. We attempted to collect diet samples from nestlings of the focal species by attaching throat ligatures via the collar method (Kluijver 1933, Johnson et al. 1980). We constructed ligatures from 20-22 gauge, solid-core, black or blue plastic-coated copper wire (RadioShack Corp., Forth Worth, TX, USA) and bent the wire into a U-shape which we fitted around the nestlings’ necks sufficiently tight to prevent swallowing but sufficiently loose to allow normal breathing and gaping behavior. We weighed (g) each nestling before applying the ligature and returned nestlings to their nest within 10 min. We used binoculars to observe each sampled nest from a distance of — 15 m until we saw adults return with food. We approached the nest after each feeding visit to remove food from the nestlings’ mouths with tweezers. We tried to collect samples for 2-4 consecutive feedings at each nest before removing the ligatures to maximize the size of the food sample, and to prevent handling the nestlings and disturbing the nest on more than one occasion. We compensated for the missed feedings during sampling by feeding each nestling a comparable portion of wax worms (pyralid larvae) after we removed the ligatures. RESULTS All three nests with ligatured Song Sparrow nestlings (3-8 days of age) yielded diet samples with no apparent problems. We observed adults at four of five Yellow-breasted Chat nests (nestlings of 2-5 days of age) remove food from ligatured nestlings’ mouths before leaving the nest when the food was not swallowed (i.e., diet samples were biased; sensii Robertson 1973). We sampled (with black ligatures) our first Abert’s Towhee nest containing two 3-4-day-old nestlings on 29 July 2009 at 1052 hrs (MST). Nestling “A” weighed 7.0 g and nestling “B” weighed 16.7 g. Both adults amved at the nest with food —25 min after we applied the ligatures, and attempted to feed the nestlings for 2-3 min. We observed at least one of the adults consume food that was originally placed in a nestling’s mouth, similar to that observed for Yellow-breasted Chats. We approached the nest to collect diet samples after the adults left and found that only nestling "B” remained in the nest. We searched under the nest but could not find nestling “A.” We took a diet sample from nestling “B” and retreated to observe the nest from a distance. An adult returned to the nest -10 min later with food and spent — I min trying to feed the nestling. We then watched the adult remove the nestling from the nest and depart from the nest before we lost sight of the adult. We immediately approached the nest, began searching for nestling “B,” and found nestling “B” 2 m from the nest on the ground. The ligature was missing, and we concluded the adult had carried the nestling by grasping the ligature thereby forcing the ligature to open and slide oft the nestling s neck. We examined nestling "B,” found no sign of injury, returned it to the nest, and moved away for observation. The adult returned about 5 min later to feed the nestling and then shaded the nestling for about 5 min. The adult appeared to resume typical behavior as soon as the ligature was removed 808 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4. December 2010 from the remaining nestling, suggesting the ligature elicited the adult’s response and not the nestling itself. We did not find nestling “A” and assumed that it died after being removed from the nest. DISCUSSION Nest sanitation is the process by which parent birds rid their nests of foreign debris (Welty 1982). Several authors have described parent birds attempting to sanitize (remove) banded nestlings when parents perceived leg bands as foreign debris in the nest (Lovell 1945, Berger 1953, Brackbill 1954). No published account exists of a throat ligature eliciting nestling removal by adults despite the superficial resem- blance of throat ligatures to leg bands on nestling birds. This behavior has been observed with Florida Scrub-Jays (Aphelocoma coenilescens) (Reed Bowman, Archbold Biological Station, pers. comm.). Ours is the first study to report an instance of nestling removal by adults (and subsequent nestling mortality) in response to use of throat ligatures. We believe other researchers that have reported aggressive behavior of adults toward ligatured nestlings did not observe removal of young because adults in these studies were physically incapable of removing nestlings due to the relative mass of the nestlings. For example. Little et al. (2009) described aggressive behavior (but not removal) of ligatured, 6-day-old Bobolink nest- lings that would have weighed 53% of adult body mass (Martin and Gavin 1995). Towhee nestlings in our study were 16 and 37% of adult body mass based on average adult body mass of 44.7 g (Tweit and Finch 1994). Nest substrate may also affect an adult’s ability to successfully remove nestlings from a nest and may explain interspe- cific variation in this behavior. We encourage future investigators to observe parental behavior at several nests before they use throat ligatures widely on a species for which evidence of the safety and efficacy of this method is lacking, e.specially on younger ne.stlings that weigh <50% of adult body mass. Some species accept throat ligatures without problems (e.g.. Song Sparrows in our study) and other species experience problems that may bias diet samples (e.g.. Yellow-breasted Chats in our study). The color or type of material used to make ligatures may affect the probability of removal and deserves further study. We used both black and blue ligatures, but birds may respond differently to other colors similar to the way that some birds respond differently to different band colors (Burley 1988). We are concerned that removal of young by adults in response to use of throat ligatures on nestlings may be relatively wide- spread in birds based on recent observations of this behavior by Florida Scrub-Jays (Reed Bow- man, pers. comm.), the observations described here of Abert’s Towhees, and the recent report of aggression toward ligatured nestlings by Bobo- links (Little et al. 2009). ACKNOWLEDGMENTS We thank J. M. Moss and Lisa Carrico (U.S. Park Service, Tumacacori National Historical Park) for logistical support. Funding was provided by the Arizona Game and Fish Department from a Wildlife Conservation Fund grant and U.S. Geological Survey from a Park Oriented Biological Support grant. Any use of trade names is for descriptive purposes only and does not imply endorsement by the U.S. Government. LITERATURE CITED Berger, A. J. 1953. Reaction of female Horned Larks to banded young. Bird-Banding 24:19-20. Brackbill. H. 1954. Red-eyed Vireo throws banded young out of nest. Bird-Banding 25:61. Burley, N. 1988. Wild Zebra Finches have band-color preferences. Animal Behaviour 36:1235-1237. Evans, F. C. 1964. The food of Vesper, Field, and Chipping sparrows nesting in an abandoned field in southeastern Michigan. American Midland Naturalist 72:57-75. Johnson, E. J., L. B. Best, and P. A. Heagy. 1980. Food sampling biases associated with the “ligature meth- od.” Condor 82:186-192. Kluijver, H. N. 1933. Bijdrage tot de biologic en ecologie van den Spreeuw (Sturniis vulgaris vulgaris L.) gedurende zijn voortplantingstijd. Wageningen, The Netherlands. Little, L. P., A. M. Stong, and N. G. Perlut. 2009. Aggressive response of adult Bobolinks to neck ligatures on nestlings. Wilson Journal of Ornithology 121:441^44. Lovell, H. B. 1945. Banded Song Sparrow nestlings removed by parent. Bird-Banding 16:144-145. Martin, S. G. and T. A. Gavin. 1995. Bobolink (Dolichonyx oryzivorus). The birds of North America. Number 176. Mellott, R. S. and P. E. Woods. 1993. An improved ligature technique for dietary sampling in nestling birds. Journal of Field Ornithology 64:205-210. Moore, J. 1986. Dietary variation among nestling Starlings. Condor 88:181-189. Orians, G. H. 1966. Food of nestling Yellow-headed Blackbirds, Cariboo Parklands, British Columbia. Condor 68:321-337. SHORT COMMUNICATIONS 809 Orians, G. H. and H. S. Horn. 1969. Overlap in food and foraging of tour species of blackbirds in the potholes of central Washington. Ecology 50:930-938. Owen, D. F. 1956. The food of nestling jays and magpies. Bird Study 3:257-265. Robertson, R. J. 1973. Optimal niche space of the Red- winged Blackbird. 111. Growth rate and food of nestlings in marsh and upland habitat. Wilson Bulletin 85:209-222. Tweit, R. C. and D. M. Finch. 1994. Abert’s Towhee (Pipila aberti). The birds of North America. Number 1 11. Welty, j. C. 1982. The life of birds. Saunders College Publishing, Philadelphia, Pennsylvania, USA. EDITORS of The Wilson Bulletin (1888-2005) and The Wilson Journal of Ornithology (2006-2010) Lynds Jones 1888-1900 Frank L. Burns 1901 Lynds Jones 1902-1924 T. C. Stephens 1925-1938 Josselyn Van Tyne 1939-1948 David E. Davis 1949-1950 George M. Sutton 1950-1951 Harrison B. Tordoff 1952-1954 Keith L. Dixon 1955-1958 H. Lewis Batts Jr. 1959-1963 George A. Hall 1964-1973 .John P. Hubbard 1974 Jerome A. Jackson 1975-1978 Jon C. Barlow 1979-1984 Keith L. Bildstein 1985-1987 Charles R. Blem 1988-1997 Robert C. Reason 1998-2000 John A. Smallwood 2001-2003 James A. Sedgwick 2004-2006 Clait E. Braun 2007- The Wilson Journal oj Ornithology 122(4):8 10-81 1, 2010 William and Nancy Klamm Service Award for 2010: Jerome A. Jackson In 1971, a young ornithologist presented his first paper at a Wilson Ornithological Society meeting on Dauphin Island, Alabama. This first paper compared the breeding biology of Red- bellied (Melanerpes caroliniis) and Red-headed (M. erythrocephalus) woodpeckers; thus his passion for woodpeckers and his enthusiasm for the Wilson Ornithological Society have been intertwined from early in his ornithological career. Jerry Jackson’s work for the Society has been as extensive as it was varied. In 1973 Jerry was elected Treasurer, succeeding Bill Klamm in the position. Bill mentored Jerry during that year and Jerry's continuing friendship with the Klamms was largely responsible for their bequest to the Society’s endowment. He served 1 year as Treasurer before he was elected Editor of the Wilson Bulletin, a position he held for 4 years. In 1979 he was elected Second Vice-President, and he rose through the ranks to serve as President from 1983 to 1985. He has hosted not one, but two Wilson Society meetings: one at Mississippi State University in 1977 and another at Florida Gulf Coast University in 2002. Another of JeiTy’s passions is hi.story. In 1988, Harold Mayfield, George Hall, and Jen'y wrote a history of the Wilson Ornithological Society to mark its first century. Jeny’s account of the last third of the first century not only summarized the 810 William and Nancy Klamm Service Award 81 1 growth ot the Wilson Bulletin and the Society’s grants and prizes, but also included anecdotes that provide a llavor of the Society and the people who shaped it during that time. In that paper, Jerry wrote about his tenure as Treasurer, saying "Quite honestly, I was out of my element!" As well as writing about the history of the Society, Jerry himself has shaped it through his long tenure on the nominating committee, his more than three decades of active participation on the Wilson Council, and his work on the conservation, history and archives, and Klamm committees. With sincere gratitude and great affection, we are proud to present the 2010 William and Nancy Klamm Service Award to Jerome A. Jackson for his four decades of service to the Wilson Ornithological Society. — Sara R. Morris, Richard C. Banks, Robert C. Bea.son, Robert L. Curry, and E. Dale Kennedy (Klamm Service Award Com- mittee). The Wilson Journal of Ornithology 122(4):812-819, 2010 Ornithological Literature Robert B. Payne, Book Review Editor MOMENTS OF DISCOVERY. NATURAL HISTORY NARRATIVES FROM MEXICO AND CENTRAL AMERICA. Edited by Kevin Winker. University Press of Florida, Gainesville, Florida, USA. 2010: 384 pages, 50 black and white illustrations. ISBN: 978-0-8130-3417-1. $ 75.00 (hardcover). — This book, edited by Kevin Winker and authored by some of the most outstanding and prolific ornithologists and mam- malogists of the 20'*’ century in the northern Neotropics, covers a set of memories, experienc- es, and points of views of field work in Central America. It also contains what might be the last writings of some charismatic explorers including Miguel Alvarez del Toro, the Coffeys (Ben and Lula), Charles Sibley, John Emlen, and Dwain Warner. Twenty contributions that narrate field adventures that roughly cover from the late 1930s to the 1990s are presented almost unedited. Most of the chapters refer to field collecting experienc- es in Mexico, understandable in terms of close- ness and access for USA-based researchers during the 20'*’ century. However, a handful depicts experiences in Belize, Guatemala, Costa Rica, Panama, and El Salvador. Field work constitutes possibly the most charming part of an ornithologist’s life and, for the vast majority of researchers, the first touch of a foreign fauna becomes an enchanting and breath-taking experience. Therefore, after the confrontation to a new culture, new food, new friends and colleagues, and the addictive combi- nation of fear, excitement, and especially an awesome tropical fauna, everything together and surrounded by an unknown language, most get their hearts trapped in the tropics; sometimes literally because some of the authors even married Latin American women! The book might seem a kind of anthology, as I am sure it was intended. However, it is much more than the transcription of the stories that otherwi.se you would only hear around the fireplace in a camp, and Joined by a series of tequila shots. After a careful reading it is clear the collection of papers repre.sents very important moments in natural sciences, becau.se among the lines one discovers the intricate system of interrelationships among ornithological research groups, and individuals that started in the post- second world war period and flourished well beyond the 1970s. For a person passionate about the history of ornithological exploration, like me, the book helps us understand diverse aspects of the sometimes cryptic development of ornithology, that usually has to be reconstructed and tracked through dates and places depicted on the labels of specimens in ornithological collections. For instance, the ~35 years of collecting expertise of Bob Dickerman in Mexico and Central America have yielded more than 8,000 specimens, deposited in at least 16 museums in Canada, USA, and Mexico. Several have represented new taxa and, as a whole, they are one of the most complete avifaunal surveys developed by any bird collector in the region. Another interesting aspect is nostalgia. Anyone, anywhere in the world that is confronted with a series of memories of field workers, arrives at the point of environmental deterioration. Through the pages we get to places that are now long gone in a preserved state, and also learn about the evolution of road systems in Mexico and Central America, and the problematic security for developing field work in many regions at present. One also realizes that something has not changed at all: the hospitality of locals even in the most remote localities. It is, however, not a simple book to read. After the first few chapters, and understanding that the Los Tuxtlas and Tamaulipas regions are most favored by researchers (given the close proximity to Mexico and the fact that some of the authors participated in the same expeditions), the nan-a- tions get to common places, and listings of taxa collected or observed become boring. Neverthe- less, 1 really enjoyed the different ways the authors perceived the culture and the environ- ment, and clear differences in perception arise between those that participated in one or a few collecting trips, and those that spent long periods in or moved to the region. The major bias of this book is that it largely represents the “gringo visitor’’ point of view 812 ORNITHOLOGICAL LITERATURE 813 (only one ot the authors is Mexican while the others are North American or European), and therefore presents a situation in which U.S. researchers come to the tropics as leaders, teachers, or trainers, and local roles are those of field assistants and guides, leaving little room for true scientific collaboration among locals. This issue was true for several decades until the late 1970s and the book is really pitiful for not having writings by the late Bernardo Villa, Mario Ramos, and Allan Phillips (as noted by the Editor’s Afterword). It should have included more representatives of recent generations of zoologists, especially locals, for which true scientific collaboration is the way to proceed in field work across a broad spectrum of activities that include collecting, measuring, recording, and censusing populations. However, this is not Kevin Winker’s fault, given that several authors failed to submit collaborations. Also, common allusions to Dwain Warner’s influence along the chapters is understandable, given that Winker belonged to that research group and authors associated with it were more prone to respond than others. Some details might have improved the book. For example, photographs are scarce or absent in most of the chapters. Scientific names (appar- ently added by Winker) are at times misspelled and, at other times, erroneously assigned (e.g., Stiirnus for Meadowlarks [Sturnella], or Quisca- lus major assigned to Boat-tailed Crackles from Chiapas [indeed Q. mexicamis]). In spite of these minor issues, the book is worthwhile to read, and a must in natural history libraries, because it also represents a nice addition to the growing literature on the history of ornithology and biological exploration. Many readers will read it with nostalgia, others trying to recall unknown aspects of the Central American natural environ- ment. The main audience must be the younger generation of field naturalists in the Neotropics, because among its pages, they will find experi- ence, scholarship, and passion through the very words of those field workers that contributed (and some still do) to the development of Latin American science. I thank C. E. Braun for comments and help in improving the English. — ADOLFO G. NAVARRO-SIGUENZA, Mu- seo de Zoologia, Facultad de Ciencias, Uni- versidad Nacional Autonoma de Mexico, Apartado Postal 70-399, Mexico D. F. 04510, Mexico; e-mail: fcvgOKgiunam.mx WYTHAM WOODS: OXFORD’S ECOLOGI- CAL LABORATORY. Edited by Peter S. Savill, Chistopher M. Peirins, Keith J. Kirby, and Nigel Fisher. Oxford University Press, Oxford, UK. 2010: xvii + 263 pages, 80 illustrations, 8 pages of color plates. ISBN: 978-0-19-954320-5. $99.00 (cloth). — There isn’t, or at least there shouldn’t be, an ornithologist anywhere who would fail to recognize Wytham Woods as the site of possibly the single most productive study of avian populations ever conducted, that of Great {Parus major) and Eurasian Blue (Cyanistes caeruleus) tits, started in 1946 by John Gibb shortly after David Lack, the new Director of the nearby Edward Grey Institute (EGI), decided to give up studying European Robins (Erithicus rubecula) in favor of nest-box studies. However, this is but one of the many distinguished studies conducted in what is, as described by Chris Pemns in the Introduction to this volume, a “very ordinary... - piece of ash-maple-hazel woodland typical of central England’’ that’s not even particularly large, being only about 1,250 ha (2,750 acres) in size compared to, say, Hubbard Brook in New Hampshire — perhaps the closest analogue to Wytham Woods in the U.S.— which is 3,160 ha (6,952 acres). Yet, as Montagu Bertie, the fifth Earl of Abingdon, might have commented after the Enclosure Act of 1814 allowed him to buy out the various small freeholders and consolidate most of what is now the Wytham Woods estate, “mighty oaks from little acorns grow’’ — and scientifically, at least, the area ranks among the mightiest plots of land on earth. This fascinating book provides a history of Wytham Woods accessible to professionals and general readers alike, from its glaciated begin- nings to recent controversies over land-use practices that eventually wrestled control of the estate from the Oxford University Forestry Department, whose management focused entirely on improving the quantity of timber production, to a committee representing the interests of the broader university, thus opening up the estate to a wide range of ecological investigations. And an impressive set of studies they are, many of which are succinctly summarized here in chapters written by key figures focusing on the trees^ flowers, invertebrates, birds, and mammals as well as overall land use and conservation management. (I’m not sure why reptiles and amphibians were left out unless it’s a reflection of a relative paucity of herpetologists in the U.K.). 814 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122, No. 4. December 2010 Among the many eminent ecologists and evolu- tionary biologists who have done important work in Wytham Woods, besides Lack and Gibb, are Dennis Chitty, Nick Davies, Charles Elton, Paul Feeny, Mike Hassell, Alex Kacelnik, Bernard Kettlewell, John Krebs, Hans Kruuk, David Macdonald, Ian Newton, Chris Perrins, Mick Southern, Richard Southwood, and George Var- ley, an incomplete list that offers incontrovertible proof that one needn’t necessarily conduct a study in some exotic locale to be internationally successful. One of the reoccurring themes of this book is the immense value of long-term studies as a means of generating data that can be used to address new questions such as climate change as well as provide a fertile venue for applying new technologies like radio-tracking, passive integrat- ed transponders, and molecular genetics, to name a few. In most cases, such studies are not conducted by a single individual, but rather by a series of students and researchers. Nowhere has this process been conducted more successfully than in Wytham Woods, where some 40 students have completed theses involving some aspect of the tit study alone, each contributing his or her piece to a study that has successfully defined the cutting edge of avian ecology for decades. Of course, facilitating this is the U.K. system of higher education, which gives a Ph.D. student 3 years, and only 3 years, to conduct original research and write a thesis — half the length of time it takes to get an ecology Ph.D. in the U.S. Yet the system clearly works, and even if the seemingly frenetic time schedule followed by Ph.D.s in the U.K. is unrealistic for most American students, the potential advantages of plugging into an established study, as demonstrat- ed by the dozens of students who have success- fully worked on the Wytham Wood tits, deserve to be more widely recognized by U.S. academics. And still there remain much to be done. Interestingly, although the original idea of putting out nest boxes for birds came from foresters who were hoping to decrea.se los.ses to defoliating caterpillars by increasing lit populations, it’s still unclear whether increasing tit densities reduces caterpillars or not. This and several other examples di.scussed in the book illustrate that the precise consequences of many, it not most, ecological interactions are at best uncertain. Apparently neither the American nor the British models of higher education are likely to put ecologists out of business any time soon! However, if we are ever to understand complex ecological interactions, surely key insights will come from places like Wytham Woods where there has been intense long-term study of multiple components of the ecosystem. All ecologists will find much of interest in this book, whether they have spent time at Oxford and the EGI or not. For those of us who are farther along in our careers, the historical perspective provided by the book is fascinating. For those starting out, the book offers numerous examples of how workers at Wytham Woods have launched their own successful careers while taking advan- tage of the wealth of prior information provided by earlier generations of students. Particularly standing to benefit is anyone in a position to initiate or continue long-term studies, whether they be of trees or birds. Doing so can be difficult, both because of the problems of obtaining continuous funding and the difficulties of envi- sioning the value of work that may not be fully realized for 25, 50, or even 100 years — long after one has (hopefully) obtained whatever degrees one is going to get. But then, this is also long after most, if not all, of our published works will have outlived their usefulness, and what better legacy could any of us leave than data that can be followed up on by future generations of students, allowing them the means to answer questions which we have yet to even formulate? — WAL- TER D. KOENIG, Senior Scientist, Cornell Laboratory of Ornithology, 159 Sapsucker Woods Road, Ithaca, NY 14850, USA; e- mail; wdk4@cornell.edu ALL ABOUT BIRDS. A SHORT ILLUS- TRATED HISTORY OF ORNITHOLOGY. By Valerie Chansigaud. Princeton University Press, Princeton, New Jersey, USA. 2010: 239 pages, many color and black-and-white illustra- tions. ISBN: 978-0-691-14519-8. $29.95 (hard- cover).— This book was previously published in 2009 as '"The Hi.stoiy of Ornithology" in the United Kingdom by New Holland Publishers, when it was translated from French into English from its 2007 "Histoire de Tornithologie" , published in France by Dclachaux & Niestle. Chapters are arranged by century with a focus on ornithology in Europe since the 1500s, and the text consists mainly of biographical sketches ORNITHOLOGICAL LITERATURE 815 ot historically important ornithologists. The index lists 769 people, mainly ornithologists, but also illustrators and administrators. The bibliography cites nothing published by the ornithologists who are sketched in the book; instead, it cites only secondary sources, 14 for the history of ornithology, 16 for the history of illustration, and 1 1 for the history of natural history. A colored timeline at the end of the book summarizes the historical points of refer- ences, voyages and discoveries, the major events in ornithology by country, and the institutional- ization ot ornithology in museums and societies with an emphasis on France. In comparison, a recent book by Michael Walters (2003) has an extensive bibliography, mainly of primary sources with 334 entries. The two books are about the same length (255 pages and more text per page in Walters). Both books draw on the comprehensive book by Stresemann (1951, 1975), which developed the history of ornithology from classical times through the first half of the past century. Both the Chansigaud and Walters books are arranged first by century and within century by country of the ornithol- ogists. Chansigaud includes photographs of early ornithologists, whereas the corresponding photo- graphs in Walters are larger, sharper, and with more contrast; apparently both books used the same image sources. Chansigaud includes imag- es of period paintings and photographs of bird specimens as they were painted for the early monographs. The bibliographic sketches in Walters are more detailed and interesting to read, the deposition of their specimens in museums is reported, and the historical and social links between ornithologists are explored in Walters (2003) but generally not in Chansi- gaud (2010). For example, Walters describes the positive intellectual and institutional influence of Buffon on the early evolutionary ornithologist Lamarck and the brilliant and “most odious of men”, Cuvier, who publicly humiliated the older Lamarck in Lamarck’s own lecture room. Lamarck subsequently lost his students, who were his last source of financial support; his daughters then had to beg for food. When he died he “was buried in a pauper’s grave”, and Cuvier’s obituary of Lamarck was so vicious that the Academie Fran9aise in publishing a memorial volume to Lamarck refused to publish Cuvier’s contribution. Chansigaud does not develop the theme of influences and the give- and-take between the early ornithologists. In another example, Chansigaud’s account on Hume and his work in India has more on Hume’s substantial political and educational accomplishments than on his ornithological contributions. In contrast, Walters notes that Hume accumulated a collection of 63,000 bird skins and 19,000 eggs (now in the British Museum [Natural History]) and the largest ornithological library in Asia; Hume’s manu- script book on the ornithology of India was never published because a servant sold the manuscript in the market as waste paper. Walters (2003) includes appendices of the systematic aiTangements of birds from Charleton (1668) through Linnaeus, Brunnich, Latham, Pallas, Gloger, and others including Vieillot, Temminck, Bonaparte, Reichenow, and Fiirbin- ger. These lists make fascinating material on the history of systematic ornithology, but this topic is not part of Chansigaud’s book. The book ends with a mention of the important early molecular systematics work of Charles Sibley up to 1990. Chansigaud includes many photographs of birds, both early black-and-white illustrations and more recent photographs of color plates and mounted specimens, each in its historical context; these illustrations are an important feature of the book. Another positive feature is the index of historically important ornithologists, and the biographic details are generally accurate. Both Chansigaud’s and Walters’s books include sections on the beginnings of American orni- thology (Bartram, Wilson, Audubon, Bonaparte. Cones, Baird, and others in North America; Marcgraf, Don Felix d’Azara, von Spix, von Ihering, and Schomburgk through Hudson in South America); Chansigaud gives more atten- tion to the development of birding. The history of ornithology in America also has been developed recently by Barrow (1998) and Weiden.saul (2007) with their accounts limited to the ABA area of North America. For a look at early ornithology and ornithologists, the facts in the book appear generally accurate, but as is evident in the titles of the United Kingdom and French editions, AU About Birds is a set of historical biographies and, as in the subtitle of this edition, is not all about birds. — ROBERT B. PAYNE. Professor Emeritus. University of Michigan. 1306 Granger Avenue, Ann Aibor, MI 48104, USA; e-mail: rbpayne(® umich.edu 816 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4. December 2010 LITERATURE CITED Barrow, M. V. 1998. A passion for birds. American ornithology after Audubon. Princeton University Press, Princeton, New Jersey, USA. Stresemann, E. 1951. Der entwicklung der ornitholo- gie. F. W. Peters, Berlin, Germany. Stresemann, E. 1975. Ornithology from Aristotle to the present. Translation of Stresemann (1951) by Hans J. and Kathleen Epstein, with a Foreword and an Epilogue on American Ornithology by Ernst Mayr. Harvard University Press, Cambridge, Massachusetts, USA. Walters, M. 2003. A concise history of ornithology. Yale University Press, New Haven, Connecticut, USA. Weidensaul, S. 2007. Of a feather: a brief history of American birding. Harcourt, Orlando, Florida, USA. INVISIBLE CONNECTIONS: WHY MI- GRATING SHOREBIRDS NEED THE YEL- LOW SEA. By Peter Battley, Brian McCaffery, Danny Rogers, Jae-Sang Hong, Nial Moores, Ju Yung-Ki, Jan Lewis, and Theunis Piersma. Photographs by Jan van de Kam. CSIRO Pub- lishing, Collingwood, Victoria, Australia. 2010: 160 pages, 240 photographs. ISBN: 978-0-643- 09659-2. $49.95 AU (paper). — No other group of migratory birds is as perilously wedded to a relative handful of critical stopover sites as the shorebirds. Crisscrossing the globe, undertaking some of the longest nonstop migrations ever documented (including, in the case of the Bar- tailed Godwit [Liniosa Icipponica], the longest), shorebirds depend on abundant, concentrated food resources to fuel their hemispheric odysseys. Few such stopover sites are as important as the Yellow Sea, through which most of the East Asian-Australasian flyway funnels every year. Millions of shorebirds commuting between Alaska, Siberia, central and .south Asia, Indonesia, New Zealand, and Australia congregate on the sea’s vast and fecund tidal flats. A third of all the Byway’s shorebirds gather along the Yellow Sea each spring; for the godwits. among others, the entire Byway population concentrates there during migration. The importance of the Yellow Sea for such imperiled species as Nordmann’s Green- shank {Tringa guttifer), or the Spoon-billed Sand- piper (Eurynorhynchus pygnieiis)-\he world’s rarest shorebird-cannot be overstated. The Yellow Sea is also home to 600 million people, many of whom, like the shorebirds. depend upon healthy shellfish populations for their livelihoods. But as this book lays out in words and images, both fishermen and birds face an uncertain future. South Korea in particular has embarked on a campaign of tidal “reclamation,” resulting in the loss of hundreds of square kilometers of formerly rich foraging habitat. The infamous Saemangeum Project alone has impact- ed 400 km" of formerly prime shorebird habitat. Invisible Connections is a beautifully illustrated plea on behalf of birds whose lives are linked to the health of the Yellow Sea ecosystem, and whose migrations encompass more than half the globe. Written by seven noted shorebird biologists from the U.S., South Korea, Australia, New Zealand, and the Netherlands, the meaty text is almost secondary to the stunning images of Dutch photographer Jan van de Kam, whose obsession with shorebirds pays glorious dividends here. The combination is an eloquent argument for preserv- ing this migratory phenomenon. Van de Kam’s photographs are arrestingly beautiful. Following the migrants from their tundra breeding areas in Siberia and Alaska, to such starkly lovely sites as Roebuck Bay in northwest Australia, where he captures green- shanks and kangaroos in the same frame, van de Kam shows the full sweep of the East Asian- Australasian Byway. His images are rich in shorebird behavior, but most striking are those- especially photographs taken on the Yellow Sea- that show the profound impact of humans on shorebird habitat, and the linked fortunes of people, especially shell fishermen, and birds. (Among the more fascinating are a series of photos of former shorebird hunters near Shanghai who now use traditional tools like decoys and bamboo whistles to trap migrant shorebirds in clap nets for scientists, instead of for the pot.). This is an English-only version of an originally trilingual (English, Chine.se, and Korean) edition first published in 2008 by Wetlands International to mark the Ramsar Conference in South Korea. Like its predecessor, this book is meant to capture the eye and sympathies of the general public, a task for which it is well-suited. The photographs are the obvious hook for non-specialists, but the text is engagingly written and does an excellent job of conveying both the immense drama of the trans/circLim-Pacific migration, and the critical nature of the Yellow Sea to its migrants. The opening chapter, by Alaskan biologist Brian McCaffery, sums up the volume’s theme: ORNITHOLOGICAL LITERATURE 817 “Time is Running Out.” Subsequent chapters include “Shorebird Lifestyles,” “Flyways,” “The Tundra,” “A Southern Holiday?,” “Tidal Flat Specialists,” and “International Partner- ships.” Nial Moores, founder of Birds Korea, and South Korean conservationist Ju Yung-Ki, end the book with “The Heart of the Fly way,” directly discussing the threats and opportunities in the Yellow Sea region. I have just one criticism. For a book at whose heart lies geography, the general absence of maps (except for a few small ones that show overall flyway patterns), is puzzling and unfortunate. Most strangely, there is no map of the Yellow Sea itself, leaving those unfamiliar with the region at a loss to easily locate the many sites mentioned in the text and captions. This single failing aside. Invisible Connections is a gem-and conservationists every- where can only hope it has the desired effect. — SCOTT WEIDENSAUL, 778 Schwartz Valley Road, Schuylkill Haven, PA 17972, USA; e- mail: scottweidensauK® verizon.net A PHOTOGRAPHIC GUIDE TO THE BIRDS OF JAMAICA. By Ann Haynes-Sutton, Audrey Downer, Robert Sutton, and Yves-Jacques Rey- Millet. Princeton University Press, Princeton, New Jersey, USA. 2010: 304 pages, 6 figures, 650 color photographs, and 220 maps. ISBN: 978- 0-691-14391-0. $29.95 (paper). — Peter Vogel and Catherine Levy asserted that “Conservation of habitat is the primary measure that will save the birds of Jamaica... Most non-government organi- zations are lacking the financial and human resources to carry out the work that is necessary, such as advocacy addressed to the government and decision makers” (Raffaele et al. 1998:30, A Guide to the Birds of the West Indies, Princeton University Press, Princeton, NJ, USA). These words seem particularly prescient today, as the Jamaican government is poised to sell the Font Hill Nature Preserve, an extraordinary stand of old-growth mangroves (especially black man- grove, Avicennia genninans), ponds, and beaches, for an 800-unit resort hotel (full disclosure: I’ve worked there since 1986). Field guides can’t by themselves protect threatened habitats, wide- spread in Jamaica, but do comprise an important tool to educate people able to influence decision makers. The new Photographic Guide to the Birds of Jamaica is the first book to occupy the niche of comprehensive bird guide for Jamaica. It’s beautifully produced, compact (127 X 190 mm [5" X 7.5"|), and durably bound, using semi- gloss paper in an easy-to-use format with typically a page per species, or two pages in the case of most endemics. Rey-Millet treats us to gorgeous photographs — up to five per species, some filling the page — typically single-flash, and perfectly exposed. The multiple photographs accompanying many accounts highlight variation in posture, vantages on plumage, and age/male or female. A comprehensive, stand-alone Jamaican field guide will benefit Jamaicans in particular, and is long overdue considering the island’s species richness and endemism: 307 bird species — 127 breeding and 180 regular migrants. Forty-eight species (> 15% of the avifauna) are either endemic species or subspecies, or Caribbean endemics. Jamaica has more endemic bird spe- cies— vividly described in the first chapter of Roland H. Wauer’s A Birder’s West Indies: an Island-by-Island Tour (1996, University of Texas Press, Austin, USA) — than any other Caribbean island including much larger Cuba and, for reasons not entirely understood, but probably linked to the Caribbean region’s complex geolog- ical history. Jamaica has its own tody (Todus todus), and four endemic genera: Trochilus (two streamertail hummingbird species), Loxipasser anoxanthus (Yellow-shouldered Grassquit), Eu- neornis campestris (Orangequit), and Nesopsar nigerrimus (Jamaican Blackbird). Several taxa challenge birdwatchers, including 23 sandpipers, 25 wood warblers, 12 pigeons/doves, and eight tyrannid flycatchers. This new photographic field guide is organized into an Introduction (19 pages), and .sections on “Birding in Jamaica” (8 pages), “How to use this book” (3 pages), and species accounts (236 pages). The introduction covers Jamaica’s loca- tion, climate, geology and geomorphology, to- pography, and habitats (12 pages); composition and origins of the avifauna; bird movements including winter, austral, and elevational migra- tions; conservation issues and threats, reinforcing the importance of habitat and invasive species; and a history of Jamaican ornithology including bibliography. The birding section briefly de- scribes and maps 16 locations with distinctive habitats (including photographs) and species. This birding section also provides useful information for birders, such as trip tips, issues of security. 818 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122. No. 4. December 2010 hazards, and — usefully — the lead author’s e-mail address. The species accounts are standardized for quick reference to all species names, taxonomy (e.g., monotypic, polytypic), body length, plumage descriptions, similar species, voice, and habitat and behavior — foraging, diet, and nesting. Each account also contains a boxed description of geographic range and status in Jamaica (e.g., endemic, resident, winter visitor, transient; and abundance category); standard country map, color-coded and contoured for up to two levels of abundance/reliability of occurrence; and rele- vant lUCN redlist category (critically endangered, endangered, vulnerable, near threatened). Seven helpful appendices include “Probably extinct species” (up to 6 in total, 3 illustrated, but not the 2 macaws and a parrot), “Vagrants” (77 of the total 307 species listed with plumage descriptions), “Species endemic to Jamaica” (30 species), “Subspecies endemic to Jamaica” (19 species), “Caribbean endemic species and sub- species recorded in Jamaica” (18 species), and a list of species with commercially available vocalizations. The index includes just Latin binomial and English common species names, and a quick index inside the back cover includes 65 of the primary bird groups, mostly families. I found Just a handful of errors, mostly typographic, but Colombia was misspelled. Haynes-Sutton and Rey-Millet intended and accomplished, in collaboration with Downer and Sutton (neither of whom, sadly, lived to enjoy the finished product), a major improvement on Downer’s and Sutton’s 1990 edition (Birds of Jamaica — A Photographic Field Guide, Cam- bridge University Press, Cambridge, UK). The new guide most importantly includes photos of all Jamaican species except the vagrants: 627 bird photos of 219 species, compared to just 64 bird photos of 47 species, emphasizing endemics, in the Downer and Sutton edition. The new book also includes or expands “habitats and behavior” sections for all photographed species. Several Caribbean field guides benefit from regional comprehensiveness, but at the expen.se of local detail. Most notable for decades was James Bond's Birds of the West Indies (on my bookshelf the 5th Edition, 1993, Houghton Mifflin Company, Boston, MA, USA). A guide of the same title was published by Helm in 1998, and by Harper Collins in a new 2010 edition, with standard-format illustrations of 632 specie.s — including visitors and vagrants, and all residents except for Trinidad and Tobago — but lacking illustrations of racial variation among islands, and with short species accounts. Two linked guides are also valuable: Raffaele et al. (1998) presents excellent accounts of all species including migrants, at the costs of bulkiness for a field guide and separation of species plates from accounts. Raffaele et al. (2003, Princeton Field Guides: Birds of the West Indies, Princeton University Press, Princeton, NJ, USA) subsequently produced an efficient field guide by shortening species accounts, located on opposite pages to illustrations, in a smaller, durable package. The latter two guides facilitate comparisons by grouping similar species on plates (basically the same plates, except for some larger format, single-species illustrations in the 1998 guide). The new Jamaican guide best captures the birds’ vivid colors, and lavishes more attention on Jamaica’s birds by focusing geographically, including mapped species distributions within the island and more detailed species accounts in some cases. Eor example, more information is available to distinguish Sad from Stolid flycatch- ers (Myiarchus barhirostris, and M. stolidus, respectively), and 1 learned that the Orangequit ranges to sea level only on Jamaica’s wetter northeast coast. No electronic app (e.g., for iPhone) is yet available for Jamaican birds as for Birds of Dominican Republic and Haiti (http:// press.princeton.edu/blog/20 1 0/06/02/the-birds-of- the-dominican-republic-and-haiti-iphone-app- available-all-proceeds-go-to-disaster-relief/). This new photographic guide will facilitate identifying and appreciating all regularly encoun- tered Jamaican birds. The more we, and especially Jamaicans, come to know the island’s avifauna the better, since environmental threats are here to stay.— THOMAS W. SHERRY, Professor, Department of Ecology and Evolutionary Biology, 400 Boggs Hall, Tulane University, New Orleans, LA 70118, USA; e-mail: tsherry@lulane.edu DO HUMMINGBIRDS HUM? EASCINAT- ING ANSWERS TO QUESTIONS ABOUT HUMMINGBIRDS. By George C. West and Carol A. Butler. Rutgers University Press, Piscat- away. New Jersey, USA. 2010: 185 pages, 32 figures, and 8 color plates. ISBN: 978-0-8135- 4738-1. $21.95 (paper). — This book contains 87 questions about hummingbirds organized into ORNITHOLOGICAL LITERATURE 819 nine chapters, plus four appendices and an extensive reference section. Hummingbirds are the favorites of many birders, casual observers, gardeners, homeowners, and ornithologists piqu- ing curiosity with their extreme lifestyles and biology, tiny size, bright colors, and engaging behaviors. George West and Carol Butler have teamed up to produce this fourth in the Rutgers Animal Q&A Series. Butler authored the three previous titles in this series. West is a Professor Emeritus of Zoophysiology at the University of Alaska, Fair- banks, and a researcher on the physiology and ecology of birds. He bas publisbed more than one hundred scientific papers, provided hundreds of illustrations for others, and recently published the revised and updated A Birder's Guide to Alaska. The book is organized into nine main chapters: Hummingbird Basics, Systems and Senses, Feath- ers and Bones, Reproduction, Flight and Migra- tion, Dangers and Defenses, Attracting and Feeding, Identifying and Photographing, and Research and Conservation. There are four appendices covering hummingbird garden plants, places to see hummingbirds, hummingbird orga- nizations, and recommended reading and web sites. A useful feature is tbe References section, which follows the organization of the book and provides the literature support for many of the answers. This particular section might have been made easier to use by boldfacing the chapter titles, as the questions in each of the nine chapters restart numbering again from 1. Thus, there are as many as nine questions with the same number in this section. As a reviewer, I read the book sequentially, but the format allows, and is intended for, readers to simply look up a topic for which they seek an answer. The answers are thorough and well researched, but they are also self-contained for the most part even when a topic is complex, using repetition of material to facilitate as clear an understanding as possible by reading only the answer to the relevant question. There are internal references to other answers for more complex topics. As a hummingbird researcher who gives frequent public programs, 1 get asked a lot of questions, and it appears that virtually every possible question about hummingbirds has been included in this book. Obvious questions like "How many species of hummingbirds are there in the world?" and “Where is the best place to bang my hummingbird feeder?" to less typical ones like "When do hummingbirds molt?" and, of cour.se, the question that titles the book, are covered with equal depth and comprehensiveness. As an Eastern birder, I do notice a western flavor to many of the answers, but given that the majority of U.S. species are western this makes perfect sense. In a few instances, exceptions that the eastern Ruby-throated Hummingbird {Archi- lochus colubris) presents are overlooked, includ- ing for example the role that insects may have in their diet for fat deposition prior to migration. One minor quibble is that my personal favorite bird name. Rainbow Starfrontlet {Coeligena iris), did not make the list of "Descriptive Hummingbird Names". Few actual eiTors are evident. The most imme- diately apparent is the top photograph on color plate F, which is captioned as a Chestnut-breasted Coronet {Boissonneaua matthewsii) but is clearly not that species, but rather is a Shining Sunbeam {Aglaeactis cupripemus). Another eiTor appears in Figure 23, where breeding and winter ranges, and migration routes, are shown for both Rufous Hummingbird (Selasphoriis rufus) and Ruby- throated Hummingbird. There is a migration aiTow for Ruby-throated from the Yucatan of Mexico to the Florida Peninsula. Robinson et al. (1996, The Birds of North America. Number 204.) do not depict this migration route, and the species is extremely scarce anywhere in the Caribbean. These few minor issues and errors could easily be remedied in a second printing but, regardless. I highly recom- mend this book for anyone who has an interest in hummingbirds, or interacts with the public about birds in any way. — ALLEN T. CHARTIER, 1442 West River Park Drive. Inkster. Ml 48141, USA; e-mail: amazilial@comcast.net The Wilson Journal of Ornithology 122(4):820-832, 2010 PROCEEDINGS OE THE NINETY-FIRST ANNUAL MEETING JOHN A. SMALLWOOD, SECRETARY The Ninety-first Annual Meeting of the Wilson Ornithological Society was held from Thursday, 20 May, through Sunday, 23 May 2010, on the campus of Hobart and William Smith Colleges in Geneva, New York. Mark E. Deutschlander, Chair of the Biology Department at Hobart and William Smith (HWS) Colleges, chaired the Local Com- mittee, which also included Kristi Hannam of SUNY Geneseo, Sara Morris of Canisius College, Margaret Voss of Behrend College at Penn State Erie, John Waud of the Rochester Institute of Technology, Erica Cooney-Connor and Nancy Wilde of HWS Conference and Events, and Pat Heieck of HWS Dining Services. The Council met from 0900 to 1600 hrs on Thursday, 20 May, at the Finger Lakes Institute on the HWS campus. That evening there was an ice-breaker social for conferees and guests at the HWS Scandling Center. The opening session on Friday convened in the Albright Auditorium at 0841 hrs with welcoming remarks from Mark E. Deutschlander, WOS President E. Dale Kennedy, and WOS Second Vice-President and Chair of the Scientific Pro- gram Committee, Robert L. Curry. After present- ing several items of information for the conferees, Vice-President Curry concluded the opening ceremony by introducing Edward H. Burtt Jr., who in turn introduced the Margaret Morse Nice Lecture and the plenary speakers, Robert B. Payne and Laura Payne, who delivered an informative presentation, “Brood parasitism in cuckoos, cowbirds, and African finches.” The scientific program included 44 papers organized into 10 sessions, 30 posters, and two symposia consisting of six papers on migratory physiology and energetics, and six papers on the effects of energy development on birds. At the conclusion of oral presentations on Friday after- noon, the conferees were treated to a screening of a remarkable documentary film on the search for Ivory-billed Woodpeckers {Canipephilus princi- palis) near the town of Brinkley, Arkansas. The evening concluded with a reception in the Vandervort Room of the Scandling Center, which coincided with the poster session. The Local Committee hosted birding forays in the vicinity of the conference site each day from Thursday through Sunday. In addition, longer trips were scheduled for Thursday to the Cornell Laboratory of Ornithology; for Friday and Satur- day, on a tour of the Finger Lakes Region known as the Ontario Pathways; and on Sunday, to the Montezuma Wetlands Complex. The conferees enjoyed a social hour prior to the annual banquet, which also was held in the Vandervort Room. The evening events included an enjoyable dinner, and afterwards WOS Presi- dent E. Dale Kennedy joined those assembled in thanks to the many persons whose hard work had resulted in a successful conference. President Kennedy also thanked the three elected members of Council who had completed their terms of office, Robert S. Mulvihill, Timothy J. O’Connell, and Mia R. Revels, and welcomed the three newly elected members of Council, Mark E. Deutsch- lander, Paul G. Rodewald, and Rebecca J. Safran. The following WOS awards and commendations also were presented: MARGARET MORSE NICE MEDAL (for the WOS plenary lecture) Robert B. Payne and Laura Payne, “Brood parasitism in cuckoos, cowbirds, and African finches.” EDWARD’S PRIZE (for the best major article in volume 121 of The Wilson Journal of Ornithology) W. Andrew Cox and Thomas E. Martin, “Breeding biology of the Three-striped Warbler in Venezuela: a contrast between tropical and temperate parulids.” STORRS L. OLSON PRIZE (for the best book review in volume 121 of The Wilson Journal of Ornithology.) Richard C. Banks, “The white-cheeked geese: Branta canadensis, B. maxima, B. "fawrensis", B. hutchinsii, B. leucopareia, and B. minima. Ecophysiographic relationships, biogeography, and evolutionary considerations. Volume 2. West- ern taxa, biogeography, and evolutionary consid- erations.” 820 ANNUAL REPORT 821 WILLIAM AND NANCY KLAMM SERVICE AWARD (for clislingLiished service to the Wilson Ornitho- logical Society) Jerome A. Jackson. LOUIS AGASSIZ EUERTES AWARDS Jessica Oswald, University of Florida, “The biogeography of birds in Peruvian tropical dry forests.” Christopher Tonra, University of Maine, ”De- temiining the relative influence of environmental and endogenous factors on the transition from non- breeding to breeding in migratory songbirds.” GEORGE A. HALL/HAROLD E. MAYFIELD AWARD Nanette Mickle, of Woodbridge, Virginia, “Tracking long-distant songbird migration (Pur- ple Martins) using geo-locators.” PAUL A. STEWART AWARDS Allison Cox, University of Missouri, “Natal dispersal of Red-bellied Woodpeckers in a fragmented landscape.” Kira Delmore, University of British Columbia, “Divergent migratory behaviors as post-zygotic barriers to interbreeding in hybrid zones.” Luke Powell, Louisiana State University, “Us- ing home range estimates and 30 years of mist-net captures to determine the effect of land use history on dispersal of Amazonian birds.” Barry Robinson, University of Alberta, “Arctic Peregrine Falcons as indicators of trophic dynam- ics in coastal ecosystems.” Taza Schaming, Cornell University, “The impact of whitebark pine mortality on Clark’s Nutcracker demography and habitat use.” Daphna Shaw, University of Florida, “Male parental investment, relatedness, population densi- ty, and territorial proximity as factors promoting extra-pair paternity in the Northern Mockingbird.” Deborah Visco, Tulane University, “Ontoge- netic mortality rates in a neotropical understory pas.serine.” Kara-Anne Ward, University of Windsor, “Female mate search tactics in Long-tailed Manakins: a novel tracking system for monitoring social interactions.” ALEXANDER WILSON PRIZE (for best student paper) Stephanie G. Wright, Villanova University, “Hybrid chickadee vocalizations change as the hybrid zone moves northward in southeastern Pennsylvania.” LYNDS JONES PRIZE (for best student poster presentation) Jason M. Townsend, SUNY-College of Envi- ronmental Science and Forestry, “Mercury bio- accumulation in Calhanis thrushes along an elevational gradient.” NANCY KLAMM BEST UNDERGRADUATE STUDENT ORAL PAPER AWARD Christina Masco, Cornell University, “Individ- uality and recognition in the Great Black-backed Gull, Lams marinus." NANCY KLAMM BEST UNDERGRADUATE STUDENT POSTER AWARD (Their posters judged a tie, two students share this year’s award.) Anne Lugg, Kutztown University, “Aging House Wren nestlings based on feather tract development, wing chord, and head length.” Jordan S. Kalish, Ohio Wesleyan University, “Bacteria and fungi in the plumage of birds of prey.” WILSON ORNITHOLOGICAL SOCIETY TRAVEL AWARDS Aubrey Alamshah, Ohio Wesleyan University, “Seasonal changes in maintenance behavior of the House Sparrow {Passer doniesticiis)." Katie Glower, University of Minnesota, “The effects of rotational grazing on relative abundance of grassland songbirds.” James Garabedian, Frostburg State University, “Habitat associations of over-wintering birds in restored hayfields in the Manassas National Battlefield Park, Virginia.” Jessie Hogue-Morgenstern, Denison University, “The relationship of migrant avian frugivory and honeysuckle (Lonicera) management in the fall.” Gustavo Londoho, University of Florida, “Breeding biology of five species of high eleva- tion Andean tanagers.” Lawrence Long, The Ohio State University. Suivey ol Lake Erie island passerine nest predation.” Anne Lugg. Kutztown University, “Aging House Wren nestlings ba.sed on feather ti'ac't development, wing chord, and head length.” Adrian Monroe, Oklahoma State University, 822 THE WILSON JOURNAL OL ORNITHOLOGY • Vul. 122, No. 4. December 2010 "Winter bird response to heterogeneity-based management in a tallgrass prairie." Erica Mueller, Montclair State University, “A spectrographic analysis of American Kestrel (Fcilco sparve fills) vocalizations. Do broodmates sound more like each other than non-related broods?" Benjamin Padilla, Gordon College, "The effects of natural and human induced edges on foraging flocks of wintering songbirds in southern New England." Monika Parsons, University of Maine, "Quan- tifying incubation behavior of Common Eiders nesting on Jordan’s Delight Island, Maine." Ashley Rathman, Kutztown University, "The effect of Wood Thrush hosts on the survival of Brown-headed Cowbird eggs and nestlings.” Sonya Richmond, University of Toronto, "In- Huence of adult age on Rose-breasted Grosbeak nestling provisioning rates in stands harvested by single-tree selection." Marla Steele, Oklahoma State University, "Effects of cold front passage on migrant raptors at Shirakaba Pass, Nagano Prefecture, Japan, autumn 2000-2009." Brynne Stumpe, Canisius College, "The effects of a residential wind turbine on bird behavior.” Emily Thomas, Penn State University, “Effects of oil and gas development on songbirds." Judit Ungvari-Martin, University of Florida, "Variation in understory forest bird communities of Amazonian forests on different soils." Gretchen Wagner, CUNY Hunter College, "Patterns of Brown-headed Cowbird parasitism on Eastern Phoebes: a ten-year comparison." Whitney Wiest, University of Delaware, “De- velopment of a bird community integrity index to monitor salt marsh condition at national wildlife refuges.” Marie Wilson, University of the South, "The effects of exurbanization on the food and habitat of Pileated Woodpeckers (Dryocopiis pileatiis)." Stephanie Wright, Villanova University, "Hy- brid chickadee vocalizations change as the hybrid zone moves northward in southeastern Pennsyl- vania." Selection committee for the Nice Medal: James D. Rising (Chair), Robert C. Beason, Robert L. Curry, Jerome A. Jackson, E. Dale Kennedy, Sara R. Morris; for the Edwards Prize: Clait E. Braun (C’hair), James A. Cox, and Kevin Winker; for the Olson ITizc: Clait E. Braun (Chair), and Slorrs L. Olson; for the Klamm Service Award: Sara R. Morris (Chair), Richard C. Banks, Robert C. Beason, Robert L. Cumy, and E. Dale Kennedy; for the Fuertes, Hall/Mayfield, and Stewart Awards: Carla J. Dove (Chair), John Bates, Terry Chesser, Greg Farley, Rob Faucett, Russ Green- berg, James Hare, Rebecca Holberton, Tom Jensen, E. Dale Kennedy, Sarah Kingston, Kevin McCracken, Chris Milensky, Patricia Escalante Pliego, Shane Pruett, Nate Rice, Scott Robinson, B. A. Schreiber, Sarah Sonsthagen, John Swaddle, Keith Tarvin, Pepper Trail, Jeff Walters, James Whatton, Chris Witt, Margaret A. Voss, and Douglas W. White; for the Alexander Wilson Prize, the Lynds Jones Prize, and the Nancy Klamm undergraduate presentation awards: Ro- bert C. Beason (Chair), Edward H. Burtt Jr., John C. Kricher, Sara R. Morris, Mia R. Revels, John A. Smallwood, Margaret A. Voss, Doris J. Watt, and Douglas W. White; and for the WOS Travel Awards: Timothy J. O’Connell (Chair), Greg Keller, and Mia R. Revels. COMMENDATION WHEREAS THE WILSON ORNITHOLOGI- CAL SOCIETY held its annual meeting in Geneva, New York, on the campus of Hobart and William Smith Colleges, overlooking the stunningly beautiful Seneca Lake, and RECOGNIZING that the Committee on the Scientific Program, under the adroit direction of Robert L. Cuiry, assisted by Robert C. Beason, Susan B. Smith, and Scott H. Stoleson, arranged and executed an exemplary exposition of oral and poster presentations, which included symposia exploring the physiology and energetics of migratory birds, and the effects of energy development on birds, and RECOGNIZING that the Committee on Local Arrangements, through the efforts of Kristi Hannam, Sara Morris, Margaret Voss, John Waud, Erica Cooney-Conner, Nancy Wilde, and Pat Heieck, and particularly Chairperson Mark Deutschlander, who answered every request with a "yes” and whose many able assistants made tho.se events come to pass, provided ah excellent conference venue in a comfortable and welcoming campus environment in which dining experiences exceeded expectations, and WHEREAS the conferees found this meeting both informative and enjoyable, THEREFORE BE IT RESOLVED that the Wilson Ornithological Society commends the Committee on the Scientific Program, the Com- ANNUM. REPOR I 823 mittee on Local Arrangements, and Hobart and William Smith Colleges tor a most suceessfiil and rewarding meeting in Geneva. BUSINESS MEETING President E. Dale Kennedy called the annual business meeting to order at 1 136 hrs, Friday, 21 May 2010 in the Geneva Room of the Warren Hunting Smith Library on the campus of Hobart and William Smith Colleges, Geneva, New York. Noting that a quorum was present, she introduced Secretary John A. Smallwood, who presented a synopsis of the Council meeting, which had taken place the previous day. As of 30 April 2010, the society’s membership stood at 1,579 individuals, including 209 students, 216 international mem- bers, and 128 new members of all membership categories. While the renewal cycle was still underway, at that time the retention rate was 89%. In addition, 250 libraries and institutions sub- scribed to The Wilson Journal of Ornithology. Secretary Smallwood then asked those assembled to stand in recognition of the following members who were recently deceased: William A. Burnham and Carl D. Marti, both of Boise, ID. Secretary Smallwood reported the society is in very good financial state with a healthy endowment of —$2,500,000, and noted the value of these investments was increasing as the economy im- proves. On behalf of Council, the Secretary offered sincere gratitude to the WOS Finance, Audit, and Investment Committee, chaired by Allan Keith, for its astute management of societal assets. Secretary Smallwood further reported that the Membership Committee, chaired by Timothy J. O’Connell, had revised and updated the WOS membership recruitment poster. Noting that the poster was currently on display, the Secretary invited those assembled to “check it out,’’ to join the society and, if already a member, to encourage nonmember colleagues to Join. It was the Secretary’s pleasure to report that Council had unanimously re-elected Clait E. Braun as Editor of The Wilson Journal of Ornithology, and he congratulated Editor Braun for doing a superb Job, noting the Journal had grown to nearly a thousand pages, the quality is first rate, and his record for collecting page charges from authors has been outstanding. Finally Secretary Smallwood announced that Council had proposed two changes to the Bylaws. First, under Article III, Section and Section 6, Council proposed to delete the sentences that require the President and the Treasurer to hold specific positions relative to OSNA. Second, Council proposed that the references to the society’s Journal be changed from the former name. The Wilson Bulletin, to the current name. The Wilson Journal of Ornithology. These pro- posed changes to the Bylaws will be voted upon by the membership at the next annual meeting, in Kearney, Nebraska, March 2011. Secretary Smallwood then gave the Boor back to President Kennedy, who summarized the Treasurer’s report in the absence of the Treasurer (Melinda M. Clark), and introduced Editor Braun, who summarized the Editor’s report. The written reports of the Treasurer and the Editor are included below. Sara R. Moms, Chair, presented the report of the Nominating Committee, which also included Robert C. Beason, Mary Bomberger Brown, and Jameson Chase: President, E. Dale Kennedy; First Vice-President, Robert C. Beason; Second Vice- President, Robert L. Curry; Secretary, John A. Smallwood; Treasurer, Melinda M. Clark; Mem- bers of Council for 2010-2013 (three nominees for three positions), Mark E. Deutschlander, Paul G. Rodewald, and Rebecca J. Safran. Having asked for additional nominations from the floor and hearing none. President Kennedy closed the nominations as a result of a motion by Robert L. Curry, seconded by Jerome A. Jackson, which was unanimously passed by voice vote. Further, Timothy J. O’Con- nell moved and Jerome A. Jackson seconded that the Secretary cast a single unanimous vote for the slate of nominees. By acclamation it became so. Second Vice-President Curry updated tho.se assembled on the 201 1 meeting, which will be a Joint meeting with the Association of Field Ornithologists and the Cooper Ornithok:»gical Society, and will occur 9 to 13 March, in Kearney. Nebraska. Mary Bomberger Brown is serving as the local host. Vice-President Curry also an- nounced that preparations were underway for the 2012 meeting: the WOS will be participating in the North American Ornithological Conference in Vancouver. British Columbia, 14-18 August. This meeting will take place on the campus of the University of British Columbia, and Kathy Martin will serve as the local chair. Having completed the business at hand. Pres- ident Kennedy inquired if any one present had additional items ol business. Because no one did, Jed Burtt moved and Ted Davis seconded that we adjourn. This came to pass at 1204 hrs. 824 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4, December 2010 REPORT OF THE TREASURER Statements of Revenues and Expenses for the Year Ending 31 December 2009 2009 2009 2010 12 Months Actual Annual Budget Annual Budget Revenues Contributions $ Student Travel Research Fund Van Tyne Library Book Fund Sales - back issues Sales - books - Van Tyne Library Subscriptions Page charges Royalties BioOne Electronic Licensing Mailing list rental income Memberships Other income Total revenues from operations $ Expenses Journal publication costs: Editorial honorarium $ Editor travel Editor supplies/Mailings Editorial assistance Copyright expense Printing - Journal Artwork Printing color plates S Operating expenses: Postage & mailing - back issues S Storage back issues Van Tyne Library - expenses OSNA management services Credit card fees Travel expenses-OSNA rep Travel expenses - general and Treasurer Travel expenses-Ornith Council Meeting expenses Membership expenses Accounting fees Insurance - D&O Oftlce supplies Postage - general Other expenses Filing fees Discretionary expenses $ Awards Student banquet $ Membership awards Hall/Mayfield Stewart Fuertcs Klamm Service Award Wilson. Lynds Jones. Klamm Student travel grants 1.240 $ 1.800 $ 1,800 720 300 500 212 100 110 282 240 200 593 600 750 13.065 15,100 12,000 40,045 18,400 34,000 7,837 4,230 1,000 23,557 17,000 22,500 85 222 270 56,277 37,000 26,000 3,894 3,500 147,807 $ 98,492 $ 99,130 4,000 $ 4.000 $ 6,000 982 1,700 1,350 2,768 1.700 2,620 21,123 24,000 18,500 - 100 - 88,637 100.000 90,000 750 1.000 750 3,428 3,600 3,500 121,688 $ 136,100 $ 122,720 0 $ 440 $ 0 - 3.50 240 726 1 .000 1.200 14,193 24.300 16,000 3,730 2.200 2.100 - 1 .500 500 370 800 1.000 - 300 250 100 3,750 3,000 - 2.000 2,000 4.165 4,500 4,000 1 .500 1 ,650 1 .475 - 500 305 - 200 200 - 160 100 10 5 10 - 3.000 . 3,000 24,794 $ 46.655 $ 35,380 2.415 $ 1 .200 $ 1.200 - 600 600 1 .000 1 ,000 1 ,000 2.000 3.000 4.000 2.500 2,500 5,000 2.000 1 .000 700 700 700 4,700 9,000 5,000 ANNUAL REPOR I 825 Assets Cash Investments: Merrill Lynch - cash Coamerica - Van Tyne checking Van Tyne Univ. Michigan account Sutton Fund - cash equivalents Howland Mgmt - cash equivalent Total cash and cash equivalents Other Investments: Merrill Lynch - common stocks Merrill Lynch - corp bonds Merrill Lynch - mutual funds Sutton Fund - equities Sutton Fund - corp bonds Howland Mgmt - equities Howland Mgmt - fixed income Total Other Investments Total Assets FUND BALANCES Fund Balances Restricted funds - Sutton Fund Unrestricted funds Net Income Fund balance - Klamm Total Fund Balances STATEMENT OF FINANCIAL POSITION 3 1 December 2009 $ 150,452 452 1.050 28,164 193,342 $ 373,460 ii 595.227 33.043 37.155 I 12,950 9.676 1 . 1 80.604 1 14.858 2.083.513 2.456.973 $ $ i 150.790 491,069 326.310 1.488.804 2,456.973 Award expenses 1 .739 3,()()() 2,553 $ 15,054 $ 2 1 ,800 $ 19,853 Contrihutions Support - Ornith Council $ 9,300 $ 9,300 $ 15,000 Support - Ornith Council, restricted to revision _ _ costs Am Bird Conservancy Dues AAZN dues 250 250 $ 9,300 $ 9,550 $ 15,250 Total Expenses $ 170,836 $ 214,105 $ 193,203 Expenses in excess of revenues before $ (23,029) $ (115,613) $ (94,073) investment income Investment activity Revenues Investment earnings - budgeted $ 0 $ 0 $ 0 Realized gain/loss - ML (14,008) 35,000 27,081 Realized gains/losses-Howland (47,404) 40,000 (20,000) Realized gains/losses - Sutton (3,033) 5,300 (2,000) Unrealized gain/loss - ML 216,075 21,000 (150.000) Unrealized gain/loss - Howland 143,664 5,000 (200,000) Unrealized gain/loss - Sutton 15,416 9,800 (40,000) Investment earnings - ML 23,411 27,000 41,760 Investment earnings - Howland 30,241 63,000 32,695 Investment earnings - Sutton 4,217 3,900 3,673 Total revenues from investment activity $ 368,579 $ 210,000 $ (306,791) Investment fees $ 19,240 $ 28,500 $ 14,366 Investment revenues in excess of expenses $ 349,339 $ 181,500 $ (321,157) Total revenues in excess of expenses $ 326,310 $ 65,887 $ (415,230) Melinda M. Clark, Treasurer 826 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4. December 2010 EDITOR’S REPORT We received 206 new manuscripts in 2009 and published 124 dating from 2004 (1), 2005 (3), 2007 (21), 2008 (85), and 2009 (14) with a total published page count of 914. This included 28 Book Reviews, 1 Memorial, 1 Service Award, Proceedings of the Ninetieth Annual Meeting, List of Reviewers, Index, and the Table of Contents for Volume 121. Most papers published were from North America north of Mexico but the published material included 33 studies from 16 counties from outside of this area. Fifteen were from South America (Argentina = 4, Bolivia = 1, Brazil = 3, Chile = 1, Colombia = 1, Peru = 4, Venezuela = 1), 12 were from Central America (Costa Rica = 4, Ecuador = 4, Mexico = 3, Panama = 1 ), and the remainder were from other countries (Canary Islands and Spain = 2, China = 1, India = 2, Israel = 1). Over half of all manuscripts received in a specific year are rejected, withdrawn, or encouraged elsewhere in the initial year of receipt. Others may linger and eventually be withdrawn by the authors or the Editor for failure of authors to respond. Electronic submission of all manuscripts is now encouraged effective as of 1 March 2010. We earlier encouraged electronic submission from authors outside of North America. All reviewers are contacted electronically and a few still request and return paper copy. Presently, we return paper copy to authors of first reviews and marked manuscripts, as it is still easier to protect reviewer confidentiality and to provide personal edits of manuscripts by paper copy. All subsequent correspondence and edits are done electronically. This process has some lingering issues (extremely slow delivery of postal mail to some countries in Central and South America, limited use of English as a first language) resulting in .some delays in receipt of materials and understanding what is needed. We identified submission of .several LPU’s (Least Publishable Units) in 2009 and tried to discourage this practice. We al.so strongly dis- couraged manu.scripts for which the ‘important' findings were published el.sewhere. The.se issues are not easily di.scerned but reviewers can be helpful in identifying questionable manu.scripts. We receive manuscripts (as do most journals) that were originally submitted elsewhere. We treat the.se manu.scripts fairly but try to learn why they were discouraged by other Journals. We afso were threatened with legal action concerning author sequence on a paper after it was published. Signed copyright forms (showing author sequence) were available for all authors and the action was terminated. Thus, there are valid reasons why we insist on having signed copyright forms on file before any manuscript is published. The editing and publication process has become more time consuming each year, as there are many details to consider. I understand why other Journals have changed proces.se.s and procedures complete with large numbers of Associate or Assistant Editors and a variety of editorial staff. Everything is more complex and one must be alert to what is best for a particular situation. Allen Press continues to be very helpful and to do outstanding work preparing and printing The Wilson Journal of Ornithology. Their work is deeply appreciated, as it makes us look compe- tent. The support of WOS Council and the Publications Committee is also essential to making our life easier. Clait E. Braun, Editor The reports of the standing committees are as follows: REPORT OE THE JOSSELYN VAN TYNE MEMORIAL LIBRARY COMMITTEE Janet Hinshaw, WOS Librarian, has provided the following statistics regarding use of the Van Tyne Library and the status of the library. Loan.s Number of loans: three; number of members to whom loans were .sent: three; number of photo- copies or scans provided: six. Acquisitions E.xchanges: A total of 120 publications was received by exchange from 198 organizations or individuals. We have lost several exchanges in the past few years and have had to subscribe to keep those Journals current. Gifts: We received 24 publications from 20 organizations. Subscriptions: We al.so received 33 publica- tions from 23 sub.scriptions. We spent a total of .$660.84 on sub.scriptions in 2009. Donations: A total of 13 items was received from members and friends. The.se included 12 Journal issues and one translation. Donors includ- ed P. Ames and J. Hinshaw. ANNUAL REPORT 827 Purchases: one journal issue and three books, tor $144.95 in credit from Buteo Books. Dispersal Back Issues. Due to lack of demand in recent years and having to clear out of the store room, we have disposed of the bulk of our stock of back issues of The Wilson Bulletin. Gifts Sent Out: one book was donated to Dan Hanley of the University of Windsor. Duplicates: 26 books were sold for $2,600.00 credit with Buteo Books. We remind members that we can lend books or provide paper or electronic copies of articles from items in the library. The Google scanning project has not yet begun for the bird library, but is tentatively scheduled for fall/winter 2010. I thank our student worker, Theresa Gorman, for day-to- day operations, and our secretary, Bev Dole, for assistance with photocopying. Jerome A. Jackson, Chair REPORT OF THE PUBLICATIONS COMMITTEE The Publications Committee of the Wilson Ornithological Society consists of Mary Bomber- ger Brown (chair), Richard Banks, Michael Hamas, Fred Lohrer, and Clait Braun (as editor of The Wilson Journal of Ornithology). During the past year, we focused our attention on three areas, all with the goal of continuing to improve the quality and stature of The Wilson Journal of Ornithology. We acquired free access to The Wilson Journal of Ornithology on BioOne (www.bioone.org) for Wilson Ornithological Society members. We worked with Christina Berger at Allen Press and Christopher Schneider at the Schneider Group/ OSNA to transfer WOS member information to BioOne; there is now a link to BioOne on the Wilson Ornithological Society web page. This free society member access to BioOne offers us several benefits: ( 1 ) it offers a tangible benefit to membership; (2) it brings more attention to the society, potentially increasing membership; and (3) it increases journal usage. There is no cost to the Wilson Ornithological Society for this service. The Table of Contents of The Wilson Journal of Ornithology is now distributed via e-mail to all OSNA members before the journal is published. Early release of the Table of Contents offers us several benefits: (I) it advertises The Wilson Journal of Ornithology and The Wifson Ornitho- logical Society to OSNA members; (2) it helps members stay abreast of the journal contents; (3) it reminds members, including lapsed members, to renew their society membership; and (4) it encourages authors to consider submitting their manuscripts to the Journal for publication. There is no cost to The Wilson Ornithological Society for this .service. The guidelines for authors submitting manu- scripts to The Wil.son Journal of Ornithology and links to the guidelines on the web page were updated. The guidelines now clarify the use of the American Ornithologists’ Union checklist for names of birds that occur in Mexico and the rest of North America, and the International Ornitho- logical Congress checklist for names of birds that occur outside of North America. The prefeired font for manuscript submissions has been clarified as have details of figure/illustration preparation. We continue to work with the Schneider Group/ OSNA to address the issue of the size of the print run for The Wilson Journal of Ornithology. We have not pursued adding a Twitter feed to The Wilson Ornithological Society web page, al- though the service is available through OSNA at no cost to the society. Mary Bomberger Brown, Chair REPORT OF THE SCIENTIFIC PROGAM COMMITTEE The Committee on the Scientific Program consisted of WOS Second Vice-President Robert L. Curry (Chair), Robert C. Reason, Susan B. Smith, and Scott H. Stoleson. Paper sessions were moderated by Greg Farley, Elise Ferree, Mary Garvin, Jerome A. Jackson, Rachel Muheim, Timothy J. O’Connell, Daniel Shustack, Charles Smith, Lindsay A. Walters, and Douglas W. White. The symposium on new perspectives in migratory physiology and energetics was moder- ated by Susan B. Smith and Mark E. Deutsch- lander, and the symposium on the effects of energy development on birds was moderated by Scott H. Stoleson. PAPER SESSIONS Symposium: New Perspectives in Migratory Physiology and Energetics Ulf Bauchinger and Scott McWilliams, Uni- versity of Rhode Island, “Tissue turnover rate 828 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4, December 2010 determines extent of phenotypic flexibility in organ size of migrating birds.” Lauren S. MacDade and Paul G. Rodewald, The Ohio State University, and Kent A. Hatch, Long Island University, “Contributions of emer- gent aquatic insects to refueling by spring migrant songbirds.” Scott R. McWilliams, University of Rhode Island, and Barbara Pierce, Sacred Heart Univer- sity, “The fat of the matter: how dietary fatty acid composition affects energy expenditure of birds during migration.” Jen Owen, Michigan State University, “Links between energetic condition, immune function, and disease susceptibility in migrant birds.” Edwin Price, University of Western Ontario, “Towards a theoretical framework for avian fat storage.” Robert J. Smith, University of Scranton, “Arrival and transition into the breeding period: the fitness consequences of timing and condition for a landbird migrant.” Symposium: The Effects of Energy Development on Birds Gregory A. George, James Sheehan, and Petra Bohall Wood, West Virginia Cooperative Fish and Wildlife Research Unit, and Harry Edenborn, National Energy Technology Laboratory, “Avian responses to gas well development in West Virginia.” Nels Johnson, The Nature Conservancy, “Can natural habitats survive the energy revolution?” Todd Katzner, National Aviary, Trish Miller and Michael Lanzone, Powdermill Nature Re- serve, David Brandes, Lafayette College, and Robert Brooks, Pennsylvania State University, “Threats to migrating Golden Eagles from development of wind energy.” Jeff Larkin, Joseph Grata, and Joe Duchamp, Indiana University of Pennsylvania, and Laura Patton, Kentucky Department of Fish and Wildlife Resources. “Developing guidelines for the crea- tion of Golden-winged Warbler breeding habitat on reclaimed surface mines in southeastern Kentucky.” Emily H. Thomas, Pennsylvania State Univer- sity, “Effects of oil and gas development on songbirds.” Kim Van Fleet, Audubon Pennsylvania, “En- ergy development and fragmentation effects.” General Sessions William H. Barnard, Norwich University, “Blood parasites in overwintering Rusty Black- birds.” Sean R. Beckett, Glenn A. Proudfoot, and Mary Ann Cunningham, Vassar College, “Large-scale movement and migration of Northern Saw-whet Owls.” J. Alan Clark, Fordham University, “Tracking birds migrating at night through an urban-rural coiridor and quantifying the effects of light and noise pollution.” Katie Clower, University Minnesota, “The effects of rotational grazing on relative abundance of grassland songbirds.” Robert L. Curry, Villanova University, “Var- iation in chickadee morphology through time in a moving hybrid zone.” Christine Eldredge, Brian Monaco, and Mark E. Deutschlander, Hobart and William Smith Col- leges, “Seasonal energetic differences between migratory, non-migratory, and irruptive song- birds.” Harold Eyster, Chelsea, Michigan, “Tolmie’s MacGillivray’s Warbler: the story of a name.” Jeanne M. Fair, Los Alamos National Labora- tory, and Ellen Paul, The Ornithological Council, “Guidelines to the use of wild birds in research.” Greg H. Farley, Andree Brisson, Alex Galt, and Stephanie Kane, Fort Hays State University, “Patterns in migratory songbird capture between 1966-1980 and 1995-2009 at an autumn banding station in the High Plains.” George L. Farnsworth, Dorothy B. Engle, Brett E. Schrand, and Christopher C. Stobart, Xavier University, and Rebecca B. Desjardins, North Carolina State Museum of Natural Sciences, “Seasonal and geographic variation in nestling sex ratio in the Northern Mockingbird.” Elise Ferree and Janis Dickinson, Cornell University, “Lifetime survival and reproductive success of extra-pair and within-pair Western Bluebirds {Sicilia mexicana)." James E. Garabedian and Frank Ammer, Frostburg State University, and J. Edward Gates, University of Maryland, “Habitat associations .of over-wintering birds in restored hayfields in the Manassas National Battlefield Park, Virginia.” Douglas A. Gross, Pennsylvania Game Com- mission, “Pennsylvania boreal conifer forests and their bird communities: past, present, and future potential.” Jessie Hogue-Morgenstern and James Marshall, ANNUAL REPORT 829 Denison University, "The relationship of migrant avian Irugivory and honeysuckle (Lonicera) management in the fall." Jerome A. Jackson and Bette J. S. Jackson, Florida Gulf Coast University, "Winter associa- tion of birds with the invasive exotic Brazilian pepper (Schinus terebinthifolius)." Gustavo Adolfo Londoho, University of Flor- ida, "Breeding biology of five species of high elevation Andean tanagers." Lawrence C. Long and James Marshall, The Ohio State University, "Survey of Lake Erie island passerine nest predation.” Christina Masco and Thomas D. Seeley, Cornell University, "Individuality and recogni- tion in the Great Black-backed Gull, Larus marinus." Adrian P. Monroe and Timothy J. O’Connell, Oklahoma State University, “Winter bird re- sponse to heterogeneity-based management in a tallgrass prairie." Erica G. Mueller and John A. Smallwood, Montclair State University, “A spectrographic analysis of American Kestrel (Falco sparvehus) vocalizations. Do broodmates sound more like each other than non-related broods?" Rachel Muheim, Lund University, John B. Phillips, Virginia Polytechnic Institute and State University, Deiring Hall, Blacksburg, Virginia, and Mark E. Deutschlander, Hobart and William Smith Colleges, “Sunrise and sunset polarized light cues calibrate the magnetic compass in migratory songbirds.” Christopher Norment, SUNY College at Brock- port, “Eifteen years of research on grassland birds in New York State: habitat relations and manage- ment implications.” Timothy J. O’Connell and Andrew D. George, Oklahoma State University, “Influence of seeded, exotic grasslands on wintering birds in mixed- grass prairie.” Sarah E. Pabian and Margaret C. Brittingham, Penn,sylvania State University, “Complex rela- tionships between forest songbirds and soil conditions.” Monika Parsons and Frederick A. Servello, University of Maine, and Cynthia S. Loftin, USGS Maine Cooperative Fish and Wildlife Research Unit, “Quantifying incubation behavior of Common Eiders nesting on Jordan’s Delight Island, Maine.” Joel Ralston, University of Albany, “Integrat- ing species distribution models and genetic data to lorecasi the effects of climate change on genetic diversity in boreal birds.” Sonya Richmond and J. Malcolm, University of Toronto, E. Nol, Trent University, and D. Burke, Ontario Ministry of Natural Resources, “InlJu- ence of adult age on Rose-breasted Grosbeak nestling provisioning rates in stands harvested by single-tree selection.” Sarah T. Saalfeld and Warren C. Conway, Stephen E. Austin State University, David A. Haukos, Texas Tech University, and William P. Johnson, Texas Parks and Wildlife Department, “Nest suceess of Snowy Plovers in the Southern High Plains of Texas.” Matthew S. Savoca, David N. Bonter, Benja- min Zuckerberg, and Janis L. Dickinson, Cornell University, and Julie C. Ellis, Tufts Cummings School of Veterinary Medicine, “Nesting density an important factor affecting chick growth and survival in the Herring Gull {Larus argentatus)." Richard R. Schaefer and D. Craig Rudolph, USES Southern Research Station, “Southeastern American Kestrel nesting habitat and nestling provisioning in the West Gulf Coastal Plain.” Chad L. Seewagen and Christopher G. Gu- glielmo. University of Western Ontario, “Oven- bird spatial behavior at an urban stopover site: movement patterns, stopover durations, and the influence of airival condition.” Daniel P. Shustack, Massachusetts College of Liberal Arts, and Amanda D. Rodewald, The Ohio State University, “Nest predation reduces benefits to early clutch initiation in an urbanizing landscape.” Dennis G. Siegfried, Daniel S. Linville, and David Hide, Southern Nazarene University, “Analysis of nest sites of the Resplendent Quetzal (Paromachrus mociiuw): a relationship between nest and snag heights.” Charles R. Smith, Cornell University, "The elements of science-based, adaptive bird conser- vation and the public trust doctrine.” Marla L. Steele and Timothy J. O’Connell, Oklahoma State University, “Effects of cold front passage on migrant raptors at Shirakaba Pass, Nagano Prefecture, Japan, autumn 2000-2009.” Scott H. Stoleson, USFS Northern Research Station, “Does habitat choice in the post-breeding .sea.son affect physiological condition of forest- interior songbirds?” Judit Ungvari-Martin, Ari Martinez, and Scott K. Robinson, University of Florida, “Variation in 830 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 4, December 2010 understory forest bird communities of Amazonian forests on different soils.” Gretchen F. Wagner and Mark E. Hauber, CUNY Hunter College, “Patterns of Brown- headed Cowbird parasitism on Eastern Phoebes: a ten-year comparison.” Lindsey A. Walters, Canisius College, and Natalie Dubois and Thomas Getty, Michigan State University, “Influences of mate quality and population growth on House Wren clutch size.” Rebecca J. Whelan and Tera C. Levin, Oberlin College, Jennifer C. Owen, Michigan State University, and Mary C. Garvin, Oberlin College, “Short-chain carboxylic acids from Gray Catbirds (Dumetellci carolinensis) uropygial secretions vary with testosterone levels and photoperiod.” Douglas White and E. Dale Kennedy, Albion College, “Gaps in nocturnal incubation in House Wrens {Troglodytes aedon)." Whitney A. Wiest and W. Gregory Shriver, University of Delaware, and Hal Laskowski, U.S. Eish and Wildlife Service, “Development of a bird community integrity index to monitor salt marsh condition at national wildlife refuges.” Marie E. Wilson, Jordan M. Casey, David G. Haskell, and Nicholas Hollingshead, University of the South, “The effects of exurbanization on the food and habitat of Pileated Woodpeckers (Dryo- copus pdeatus)." Stephanie G. Wright and Robert L. Curry, Villanova University, “Hybrid chickadee vocali- zations change as the hybrid zone moves north- ward in southeastern Pennsylvania.” POSTERS Aubrey Alamshah and Edward H. Burtt Jr., Ohio Wesleyan University, “Seasonal changes in maintenance behavior of the House Sparrow ( Passer domesticus)." Andrew C. Alba and Gregory B. Cunningham, St. John Eisher College, “The fox in the henhouse; responses of chickens {Callus domes- ticus) to the scent of a predator.” Catherine C. AIsford and Dorothy 1. Fatunmbi, Canisius College, Brenda S. Keith and Richard S. Keith, Kalamazoo Nature Center, and Sara R. Morris, Canisius College, “Does age matter? A study of whether age affects stopover ecology of migrant warblers.” Robert C. Reason, Accipiter Radar Corpora- tion, “Remote radar monitoring of bird activity via the network.” Martyna Boruta and Mark E. Deutschlander, Hobart and William Smith Colleges, “Ultraviolet reflectance and age/sex variation in Yellow Warblers {Dendroica petechia)." W. P. Brown and M. E. Zuefle, Kutztown University, “Are bird images returned from search engine results accurately identified?” Jameson E. Chace, Salve Regina University, and Steven D. Eaccio, Vermont Center for Ecostudies, “Evaluation of breeding habitat quality for a population in decline: the Canada Warbler in northeastern Vermont.” Kailey Chidester, Bethany Bashaw, and Mark E. Deutschlander, Hobart and William Smith Colleges, “Determining land quality for migra- tory birds near the Lake Ontario shoreline: a study of daily mass gain during fall and spring migration.” Kelly Eox, Alyssa Caffarelli, and Lindsey Walters, Canisius College, and Brian S. Maitner, Rice University, “Nest defense and parental investment in the House Wren.” Casey Eranklin, Jordan Youngman, Mianna Molinari, and Mark E. Deutschlander, Hobart and William Smith Colleges, “Energetic condi- tion of spring migrating North American warblers at a northern stopover site: a study of Breeding and Insurance hypotheses.” Samuel Georgian, Paul Overdorf, and Mark E. Deutschlander, Hobart and William Smith Col- leges, “Determining the role of the Insurance Hypothesis in the stopover ecology of two kinglets and two flycatchers.” Erin Greenlee and Ian Hamilton, The Ohio State University, “Distributional shifts of breed- ing birds in Ontario, Canada.” Nathan Grosse, Christopher Norment, and Mark Norris, SUNY Brockport, and Heidi Ken- nedy, NY Department of Environmental Conser- vation, “Area effects removed: vegetation char- acteristics and grassland bird abundance in a western New York field.” Margret I. Hatch, Penn Slate Worthington Scranton, and Robert J. Smith, University of Scranton, “Annual variation in arrival of long- distant migrants and a comparison of first capture dates and condition between verified breeders and presumed migrants.” Jordan S. Kalish and Edward H. Burtt Jr., Ohio Wesleyan University, “Bacteria and fungi in the plumage of birds of prey.” Rachel Kushner, Christy Conley, and James S. Marshall, Denison University, “Patterns of inter- ANNUAL REPOR I 831 terence competition in overwintering central Ohio birds.” Anne Liigg, W. Brown, D. Alexander, M. Zuefle, and T. Underwood, Kutztown University, “Aging House Wren nestlings based on feather tract development, wing chord, and head length.” Mianna Molinari and Mark E Deutschlander, Hobart and William Smith Colleges, “The influence of temperature on energetic condition in spring migrating warblers.” Brad Mudrzynski and Christopher Norment, SUNY College at Brockport, “Early successional habitat characteristics and fruit abundance influ- ence stopover use of fall migrating songbirds.” Laura M. Niczyporowicz and Todd J. Under- wood, Kutztown University, and Roland R. Roth, University of Delaware, “The effectiveness of constant effort mist-netting in estimating abun- dance and reproductive success of a Wood Thrush population.” Nathan Olszewski, Kevin Sobol, and Lindsey Walters, Canisius College, “Male response to egg color in the House Wren.” Paul Overdorf and Mark E. Deutschlander, Hobart and William Smith Colleges, “Using wing morphology and mass to model the potential flight distances of passerines at a stopover site near Lake Ontario.” Benjamin J. Padilla, Ashley Ballou, and Gregory S. Keller, Gordon College, “The effects of natural and human induced edges on foraging flocks of wintering songbirds in southern New England.” Ashley N. Rathman and William P. Brown, Kutztown University, and Roland R. Roth, University of Delaware, “The effect of Wood Thrush hosts on the survival of Brown-headed Cowbird eggs and nestlings.” Susan B. Smith, Villanova University, “An assessment of refueling rates and diet of songbirds during migratory stopover at the Braddock Bay Bird Observatory.” Kevin Sobol, Nathan Olszewski, and Lindsey Walters, Canisius College, “Egg color and female quality in the House Wren.” Brynne A. Stumpe and Sara R. Morris, Canisius College, “The effects of a residential wind turbine on bird behavior.” Jason M. Townsend, SUNY College of Envi- ronmental Science and Eorestry, and Charles T. Driscoll, Syracuse University, “Mercury bioac- cumulation in Catharus thrushes along an eleva- tional gradient.” Mitch Walters, Cornell University, Marcela Liljesthrom, Centro Austral de Investigaciones Cientificas, CADICONICET, Argentina, and Ca- ren Cooper, Cornell University, “Comparing female characteristics to egg and yolk size in the Chilean Swallow (Tachycineta meyeni)." ATTENDANCE Alabama: Tuscaloosa, Marie E. Wilson. Arizona: Tucson, Clait E. Braun, Nancy J. K. Braun. California: Concord, Naho Kermeen. District of Columbia: Carla Dove. Delaware: Newark, Whitney Wiest. Elorida: Gainesville, Gustavo Londono, Judith Ungvari-Martin; Jacksonville, Bynne Stumpe; Naples, Bette Jackson, Jerome Jackson. Indiana: Bristol, Doris Watt. Kansas: Hays, Andree Brisson, Greg Earley, Cate Healy. Maine: Orono, Monika Parsons. Maryland: Bethesda, Ellen Paul; Brandywine, Rachael Kushner; Columbia, James Smith; Frost- burg, James Garabedian. Massachusetts: Chilmark, Allan Keith; East Falmouth, Ted Davis; Haverhill, Benjamin Pa- dilla; North Adams, Daniel Shustack; Pocasset, John C. Kricher. Michigan: Albion, Dale Kennedy, Doug White; Ann Arbor, Bob Payne, Laura Payne; Chelsea, Harold Eyster; East Lansing, Jen Owen. Minnesota: Minneapolis, Katie Glower. Missouri: Columbia, Andrew Cox; St. Louis, William Condit. New Hampshire: Bow, Christie Eldredge; Manchester, Jay Pitocchelli. New Jersey: Lincoln Park, Erica Mueller; Lyndhurst, Michael Newhouse; Pennington, Ste- phanie Wright; Randolph, John Smallwood, Mary Anne Smallwood. New Mexico: Albuquerque, Erin Greenlee. New York: Albion, Brad Mudrzynski; Armonk, J. Alan Clark; Brockport, Nathan Grosse, Chris- topher Norment; Bron.x, Oriana Chan, Kyle Wright; Brooklyn, Jesse Ross; Buffalo, Catherine Alsford, Lindsey Walters; Geneseo, Kristina Hannam; Geneva, Bethany Bashaw, Mark Deutschlander, Casey Franklin, Samuel Georgian, Becky Jones, Mianna Molinari, Brian Monaco, Paul Overdorf, Bob Taylor, Jordan Youngmann; Grand Island, Sara Morris, Kevin Sobol; Ham- burg, Kelly Fox; Hurley, Kailey Chidester; 832 THE WILSON JOURNAL OL ORNITHOLOGY • VoL 122. No. 4. December 2010 Ithcica, Caren Cooper, Elise Ferree, Jane Graves, Heidi Lovette, Christina Masco, Colleen McLinn, Matthew Savoca, Charlie Smith, Scott Sutcliffe, Jason Townsend, Mitch Walters; Menands, Joel Ralston; New York, Chad Seewagen; Newburgh, Gretchen Wagner; Niagara Falls, Dorothy Fa- tunmbi; Poughkeepsie, Glenn Proudfoot; Roches- ter, Andrew Alba; Rye, Christy Conley; Utica, Judith McIntyre; West Seneca, Alyssa Caffarelli, Nathan Olszewski. Ohio: Cincinnati, Dottie Engle, George Farns- worth; Columbus, Lawrence Long, Paul Rode- wald; Delaware, Aubrey Alamshah, Edward Burtt Jr., Jordan Kalish; Granville, Jessie Hogue- Morgenstern; Medina, Amanda Fawcett; Oberlin, Mary Garvin; Sandusky, Robert Season; Wester- ville, James Marshall. Oklahoma: Edmond, Dennis G. Siegfried; Norman, Tamaki Yuri; Stillwater, Adrian Mon- roe, Tim O’Connell, Marla Steele; Tahlecjuah, Mia Revels. Pennsylvania: Barto, George Gregory; Car- isle, Kim Vanfleet; Dunmore, Margaret Hatch; Erie, Margaret Voss; Harrisburg, Nels Johnson; Indiana, Jeff Larkin; Kutztown, Bill Brown, Todd Underwood; Levine, Scott Stoleson; Markleys- biirg, Frank Ammer; Mertztown, Ann Lugg; Orangeville, Douglas Gross; Philadelphia, Laura Niczyporoxicz; Pittsburgh, Todd Katzner, Fred McCullough, Michael Sobkowiak; Reading, Ash- ley Rathman; Scranton, Rob Smith; Sugar Grove, Pam Stoleson; University Park, Sarah Pabian; Villanova, Bob Curry, Susan Smith; Warren, Emily Thomas; Williamsport, Nathan Frank; Rhode Island: Kingston, Ulf Bauchinger, Scott McWilliams; Newport, Jameson Chace. Texas: Nacogdoches, Sarah Saalfeld, Rick Schaefer. Virginia: Alexandria, Dick Banks; Shipman, Allen Hale Vermont: Northfield, William Barnard; Will- iston, Sean Beckett; Canada, Ontario: Ajax, Sonya Richmond; London, Eddy Price; Mississauga, Silu Wang; Newmarket, Mike Van den Tillaart; Simcoe, Erica Dunn, David Hussell; Toronto, James Rising, Trudy Rising. Sweden: Lund, Rachel Muheim. The Wilson Jounuil oj Oniilhology 1 22(4):833-834, 2010 REVIEWERS FOR VOLUME 122 Referees are the life blood ol a journal as editors depend on them to help identify manuscripts with merit and offer suggestions to improve the data analysis, overall science, and writing. These individuals receive little recognition but are extremely important in the process of improving the science and quality of what is published. We thank all of those listed below who served as referees for manuscripts processed (accepted and published, withdrawn, rejected) after I July 2009 through completion ot the December 2010 issue of Volume 122. (Those shown in boldface reviewed more than one manuscript m this period.) The Wilson Ornithological Society and the editorial staff are indebted to and thank each person who served as a reviewer. — Clait E. Braun, Editor K. P. Able, N. J. Adams, J. Aguilar-Amat, J. A. Ahumada, A. D. Anders, R. G. Anthony, D. P. Armstrong, E. Arriero, L. Aubry, S. K. Auer, J. K. Augustine, D. Badzinski, V. Baglione, M. C. Baker, G. Ballard, W. H. Baltosser, G. Barrantes, J. Bart, J. Bates, R. C. Beason, G. Beauchamp, M. Bekoff, J. R. Belthof, M. Benowitz-Freder- icks, K. S. Berg, T. M. Bergin, M. Bidwell, R. O. Bierregaard, K. Bildstein, T. R. Birkhead, C. A. Bishop, J. G. Blake, J. A. Blakesley, L. A. Blanc, C. R. Blem, P. H. Bloom, C. W. Boal, C. E. Bock, E. K. Bollinger, J. S. Bolsinger, D. N. Bonter, T. A. Bookhout, C. A. Botero, W. S. Boyd, J. E. Bradley, L. A. Brennan, D. Brightsmith, D. W. Buden, D. E. Burhans, D. B. Burt, E. H. Burtt Jr., P. H. Butchko, C. J. Butler, B. S. Cade, T. Cade, C. Caffrey, J. W. Cain, S. Calme, B. R. Chapman, J. Chaves-Campos, N. Chernetsov, R. T. Chesser, R. Cintra, R. B. Clapp, W. Clark, C. E. Clarkson, N. J. Collar, C. T. Collins, J. C. Connelly, L. M. Conner, R. N. Conner, C. J. Conway, M. I. Cook, D. W. Coulton, R. Covas, J. Cox, S. Craig, E. R. A. Cramer, H. Q. P. Crick, A. Cmz, J. F. Cully, F. J. Cuthbert, L. D’Alba, W. E. Davis, D. K. Dawson, D. C. Dearborn, K. L. Decker, C. A. Delgado, M. M. Delgado, R. DeLotelle, E. P. Derryberry, A. Desrochers, T. L. DeVault, A. A. Dhondt, J. Diamond, A. S. Dolby, L. Dos Angos, C. J. Dove, W. Duckworth, K. M. Dugger, P. O. Dunn, J. B. Dunning Jr., W. R. Eddleman, B. A. Eichhorst, J. C. Eitniear, K. S. Elli.son, L. Ellison, C. S. Elphick, A. Engilis Jr., R. M. Erwin, R. H. M. Espie, D. L. Euler, J. Faaborg, A. Farnsworth, T. W. Fawcett, E. Fernandez-Juricic, K. Fierro-Calderon, J. Ejeldsa, C. R. Foster, M. S. Foster, R. M. Fraga, C. Francis, M. R. Francisco, P. C. Frederick, M. Frederiksen, R. W. Furness, G. A. Gale, V. Garcia, S. A. Gauthreaux, J. L. Gehring, D. D. Gibson, A. Giese, J. D. Gilardi, F. B. Gill, M. D. Giovanni, C. B. Goguen, H. B. Gomes, O. Gordo, J. B. Grand, D. J. Green, M. C. Green, R. Greenberg, P. D. Greger, M. Gregg, R. J. Gutierrez, M. J. Guzy, J. H. Haffer, B. A. Hahn, P. B. Hamel, B. T. Hansen, A. Harmata, R. E. Harness, J. B. C. Harris, B. J. Hatchwell, W. M. Healy, S. G. Herford, S. K, Herzog, E. J. Heske, B. Heinrich, K. Higgins, J. E. Hines, K. A. Hobson, P. A. R. Hockey, R. W. Hoffman, C. S. Holling, R. T. Holmes, S. B. Holmes, G. Holroyd, J. P. Hoover, S. N. G. Howell, R. L. Hutto, L. D. Igl, D. Ingold, N. Itonaga, F. C. James, J. M. Jawor, J. R. Jehl, B. Jobin, D. H. Johnson, L. S. Johnson, O. W. Johnson, C. D. Jones, S. L. Jones, R. Kannan, P. Kappes, A. C. Kasner, J. L. Kellermann, C. Kellner, R. S. Kennedy, D. T. King, C. K. Kirkpatrick, G. Kirwan, N. A. Klaus, D. Klem Jr., S. T. Knick, G. Kobriger, W. D. Koenig, E. M. Kofoed, R. R. Koford, N. K. Krabbe, D. E. Kroodsma, J. A. Kushlan, H. Kylin, Q. Latif, S. C. Latta, J. J. Lawler, A. T. Lee, J. D. Ligon, J. Liguori, C. K. Lim, B. C. Livezey, P. E. Llambias, H. Lloyd, J. P. Loegering, J. T. Lokemoen, G. A. Londono, S. C. Lougheed, 1. J. Lovette, J. Lowen. P. E. Lowther, J. E. Loye, R. S. Lutz, B. E. Lyon, E. A. MacDougal-Shackleton, R. H. F. Macedo, J. N. Mager III, D. L. Maney, M. A. Marini, W. R. Marion, P. P. Marra, L. B. Martin, P. R. Martin, T. E. Martin, T. L. Master, P. Matyjasiak, J. E. McCormack, D. McCracken, S. McGehee. K. J. McGowan, A. Mee, D. J. Mennill, J. M. Meyers. B. Mila, K. E. Miller, B. A. Millsap, C. M. Miskelly, S. Mitra, A. P. Mpller, C. E. Moorman, M. L. Morrison, R. I. G. Morrison, D. H. Morse, E. S. Morton. R. Mo.s.s, R. L. Mumme, T. G. Murphy, L. D. Murray. L. N. Naka, Z. Nemeth. S. A. Nesbitt, P. Newell, D. Newstead, N. Nielsen- PincLis, G. Niemi, I. C. T. Nisbet, R. Noske, K. P. Oh, G. H. Orians, L. W. Oring. S. J. Ormerod, J- M. Paillisson, J. D. Paruk, P. W. C. Paton. R. B. Payne, A. T. Pearse, N. G. Perlut, B. P. Peterjohn, L. J. Petit, W. Post, R. Poulin, B. Pranty, H. D. Pratt, T. Price, H. Prior, M. S. Pruett. B. Pugesek, C. M. Raley, C. J. Ralph, J. H. Rappole, J. T. 833 834 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 4, December 2010 Ratti, L. J. Redmond, K. P. Reese, H. M. Reeves, J. M. Reid, J. L. Reidy, R. Renfrew, K. Renton, M. Restani, R. E. Ricklefs, C. D. Rittenhouse, S. K. Robinson, E. Roche, A. D. Rodewald, R. D. Rodgers, J. J. Roper, R. N. Rosenfield, J. Rotenberry, R. R. Roth, D. C. Rudolph, T. J. Sachtleben, L. Sandoval, V. Santharam, J. C. Scarl, G. D. Schnell, E. A. Schreiber, K-L. Schuchmann, T. S. Schulenberg, P. Scott, S. G, Sealy, T. W. Seamans, P. Seddon, J. Sedinger, J. A. Shaffer, S. R. Sheffield, L. M. Siefferman, L. E. Silveira, J. A. Smallwood, C. R. Smith, G. W. Smith, T. Smulders, D. A. Spector, D. Sperry, M. T. Stanback, D. F, Stauffer, D. W. Steadman, S. J. Stedman, K. Steenhof, L. Stein, J. L. Stephens, I. Stewart, R. Stiehl, D. F. Stotz, P. C. Stouffer, B. M. Strausberger, B. N. Strobel, J. D. Styrsky, K. S. Summerville, D. L. Swanson, D. A. Swanson, P. W. Sykes, M. K. Tarburton, C. F. Thompson, M. B. Tingley, J. A. Tobias, P. W. Trail, E. N. Vanderhoff, D. E. Varland, C. S. Vaughan, C. M. Vleck, S. von Oettingen, H. Wada, J. S. Wakeley, J. W. Walk, B. Walker, H. S. Walter, J. R. Walters, D. Watt, P. J. Weatherhead, W. W. Weathers, H. P. Weeks, W. Wehtje, J. T. Weir, R. Weiss, C. White, L. A. Whittingham, J. W. Wiley, S. K. Willson, M. D. Wilson, R. E. Wilson, J. C. Wingfield, J. E. Woodford, K. Yasukawa, P. D. Yaukey, S. Yezerinac, R. Yosef, K. J. Zimmer, R. M. Zink, K. Zyskowski Editor’s Comments The rate of submissions has increased in 2010 and it appears we will again receive >200 new manuscripts in the calendar year. This increases our need for reviewers. All manuscripts (very few are initially discouraged as inappropriate prior to the review process) are sent to a minimum of two subject area referees. We select potential reviewers from perusal of the Literature Cited, our file of subject area reviewers, and those offered by the authors of specific manuscripts. In a perfect world, we would expect to receive positive responses from about two of every three referees contacted to review a specific manuscript. This only rarely happens and we may have to contact up to 12 (or more) potential referees before we find someone to review a particular manuscript. This problem is not unique to our Journal as other editors have similar problems. The most recent Issue of Science (2010, 329: 1466) has a commentary (Bcittling the Paper Glut) on publishing and the responsibility of authors of published papers to be willing to review manuscripts. Their proposed solution would be to have authors of submitted manuscripts provide three reviews of other manu.scripts. Our goal is to provide reviews and editorial recommendations on suitability of manuscripts in a timely manner. Searching for someone to agree to review manuscripts slows the process. We ask those contacted about reviewing a manuscript to consider the need for competent and timely reviews. At the least, reviewers contacted are asked to respond either yes or no instead of not responding at all. The areas of delay in the process of publishing for The Wilson Journal of Ornithol- ogy are: reviewer availability, timeliness of reviews, and timeliness in author revision. Some manuscripts that are encouraged fail to be returned in a timely manner. Some never are returned and become de facto rejections or withdrawals. We presently do not have a backlog of accepted manuscripts waiting for publication. Those being accepted in September and October into early November this year will be published in the March Issue of the next year with few exceptions. Clait E. Braun Editor, 2010 The ITilson Journal of Ornilholoi'y 1 22(4):835-858, 2010 Index to Volume 122, 2010 Compiled by Leslie A. Robb This index includes leterences to genera, species, authors, and key words or terms. References are made to scientific names ot all vertebrates mentioned within the volume and other taxa mentioned prominently in the text, in addition to avian species. Nomenclature follows the AOU Check-list of North American Birds (Seventh Edition) and F. Gill and M. Wright and supplements thiough 2010 (Birds of the World, Recommended English Names, Princeton University Press, Princeton, New Jersey, USA and Oxford, United Kingdom). Reference is made to books reviewed and announcements as they appear in the volume. A Abundance abundance and distribution of waterbirds in the Llanos of Venezuela, (577), 102-115 of Griis canadensis nesiotes, 556—562 Accipiter cooperi, 168-173, 290, 470, 475 striatus, 290, 518-531, 768, 771 Actitis macidarius [macidaria, I 15], 768, 771 Aegithalos caudatus, 2 1 2 Aerodramiis brevirostris, 259 fuciphagus inexpectatus, 259-272 hirundinaceus, 270 maximus, 259 salanganiis, 259 spodiopydius assimilis, 158 spp., 153 unicolor, 259 vanikorensis, 270 Agelaioides badius, 417^31 (Frontispiece), 795, 797 Agelaius phoeniceus, 187, 291, 307-317, 402^05, 435, 532-544, 671, 768, 771, 807 Aguilera, Xiomara Galvez and Felipe Chavez-Ramirez, Distribution, abundance, and status of Cuban Sandhill Cranes (Grus canadensis nesiotes), 556- 562 Aimophila aestivalis. 518 Aix spon.ta. 1 87 AJaia ajaja, 1 14 Alces alces, 699, 701-703 Aldred, James, .see Seymour, Adrian S., Graham Hatherley, Francisco Javier Contreras, , and Fergus Beeley Alectoris chukar, 762-766 Alessi, Mark G., Thomas J. Benson, and Michael P. Ward, Nocturnal social cues attract migrating Yellow- breasted Chats, 780-783 Alopex lagopus, 2 Alophoxiiis pallidiis, 1 76 Alouatta palliata. 794 seniculus, 794 Alvey, Erin C., see Confer, John L., Kevin W. Barnes, and Amaitrornis phoenicurus. 784-788 phoenicunis insidaris, 787 phoenicurus leiicomelanus, 787 phoenicurus midnicobaricus, 787 Amazilia Candida, 592 cyanocephala, 592-597 cyanocephala chlorostephcma, 592 cyanocephala cyanocephala. 592-597 cyanocephala guatemalensis, 592 rutila, 496 violiceps, 592 Amazon, Lilac-crowned, see Amazona finschi Orange-winged, see Amazona amazonica Puerto Rican, see Amazona vittata Amazona amazonica, 88-94 finschi, 516 vittata, 515 Amazonetta brasiliensis, 108 Ammodramus bairdii, 144, 346-353, 455^64 caudacutus, 340-345 henslowii, 290, 635-655, (Frontispiece) lecontii, 139-145 maritimus, 343, 532-544 nelsoni, 340-345 savannarum, 144, 290, 455^64, 635-654 Amos, Anthony F., see Foster, Charles R., , and Lee A. Fuiman Anas acuta, 768, 771 americana, 376, 613, 768, 770 clypeata, 569, 612-614, 768, 771 crecca, 376, 768, 770 cyanoptera. 613, 768, 770 discors, 108, 111,612-614, 768, 770 laysanensis, 9 platyrhynchos, 19, 302, 565, 569, 613, 759, 762, 768 771 strepera. 768, 770 Anhima cornuta, 105, 114 Anhinga, see Anhinga anhinga Anhinga anhinga, 105, 108, 111, 114 spp., 380 Ani, Smooth-billed, see Crotophaga ani Anich, Nicholas M. and Bryan M. Reiley, Effects of a flood on foraging ecology and population dynamics of Swain.son’s Warblers, 165-168 Anser albifrons, 49 1 ant, army, see Eciton biirchellii. Labidus praedator. Labidus spinninodis fire Red lire], see Solenopsis invicta Antbird, Immaculate, see Myrmeciza immaculata Anthus rubescens. 663 spragueii, 346-353, 455-464 Antpitta, Chestnut-crowned, see Grallaria ruficapilla Chestnut-naped, see Grallaria nuchalis 835 836 THE WILSON JOURNAL OL ORNITHOLOGY • Vol. 122, No. 4, December 2010 Great, see Grallaria excelsa Jocotoco. see Grallaria ridgehi Moustached, see Grallaria alleni Pale-billed, see Grallaria carrikeri Plain-backed, see Grallaria haplonota Rufous, see Grallaria riifula Scaled, see Grallaria giiatinialensis Variegated, see Grallaria varia Watkins’s, see Grallaria watkinsi White-bellied, see Grallaria hypoleiica Yellow-breasted, see Grallaria jlavotincta Anumhius anniimbi, 42 1 , 422 Aphelocoma coerulescens, 808 Apiis apus, 436 spp., 153 Acjuila adalberti, 364 chrysaetos, 354 Aranudes cajanea. 105, 115 Aramiis giiarauna. 105, 1 15 A rat Inga leucophthahna, 803-806 Archilochus alexandri, 494-502 colubris, 302, 496, 5 1 8-544 Ardea alba. 96, 102, 106, 1 14, 532-544 cocoi, 108, 114 herodias, 96, 232, 532-544, 768, 770 Arenaria interpres, 532-544 Arremon brunneinucha. 507 Arundinicola leucocephala, 105, 115 Asia jlammeiis, 457, 768, 77 1 otus, 768, 771 Asthenes baeri, 602 berlepschi. 602 cactoriim, 602 dorbignyi, 602 humicola, 602 liiizae, 600-603 patagonica, 602 pudibitnda. 602 steinbachi, 602 Athene ciinicuaria, 51-59 cimicuaria hypugaea, 51-59 Atterbcrry-Jones, Megan R. and Brian D. Peer, Cooperative breeding by Red-headed Woodpeckers, 160-162 Aidacorhynchus prasinits. 507, 509 Ayers, Andrea J. and James W. Armacost Jr., Marsh Wren eats small fish, 623-624 Aythya affinis. 569, 768, 771 ainericana, 613, 768, 771 inarila. 376 valisineria. 768, 770 B Babhitt, Kimberly J., .see Suomala. Rebecca W., Sara R. Morris. . and Thomas D. Lee Babbler. Rusty-faced, see Robsoniiis rahori Bader, Troy J. and James C. Bednarz. Home range, habitat use, and nest site characteristics of Mississippi Kites in the White River National Wildlife Refuge. Arkansas. 706-715 badger, American, sec Taxidea taxus liaeolophus bicolor. 1 26- 134, 171.51 8-544 Baldassarre. Guy A., see Vilella, Francisco, J., Mark S. Gregory, and Baldwin. Heather Q., Clinton W. Jeske, Melissa A. Powell, Paul C. Chadwick, and Wylie C. Barrow Jr., Home- range size and site tenacity of overwintering Le Conte’s Sparrow in a fire managed prairie, 139-145 Bao-Zhong, Lu, see Xiao-Ping, Yu, Xi Yong-Mei, , Li Xia. Gong Ming-Hao, Shi Liang, and Dong Rong Barbet, Black-browed, see Megalaima oorti nuchalis Red-crowned, see Megalaima rafflesii Taiwan, see Megalaima nuchalis White-cheeked, see Megalaima viridis Barclay, M., see Quinn, J. S., A. Samuelsen, , G. Schmaltz, and H. Kahn Barnes. Kevin W., see Confer, John L., , and Erin C. Alvey Barrantes, Gilbert, see Chavarn'a-Pizarro, Tania, Gustavo Gutierrez-Espeleta, Eric J. Euchs, and Barrow Jr., Wylie C., see Baldwin, Heather Q., Clinton W. Jeske, Melissa A. Powell. Paul C. Chadwick, and Bartramia longicauda, 1 1 5 Basileuterus culicivorus, 507 rufifrons. 507, 509 tristriatus, 45 1 , 507 Baywing, see Agelaioides hadiiis beaver, American, see Castor canadensis Beckett, Matthew D. and Gary Ritchison. Effects of breeding stage and behavioral context on singing behavior of male Indigo Buntings, 655-665 Bednarz. James C., see Bader, Troy J. and Bednarz. James C.. see Pappas, Sara, Thomas J. Benson, and Bee-eater. European, see Merops apiaster White-fronted, see Merops bullockoides Beeley, Fergus, see Seymour. Adrian S., Graham Hatherley, Francisco Javier Contreras, James Aldred. and behavior army ant raid attendance and bivouac-checking behavior by neotropical montane forest birds, 503-512 arthropod foraging by a southeastern Arizona humming- bird guild. 494-502 cavity-sharing between a Dryocopus galeaius and two Aratinga leucophthalma, 803-806 Coccyzus americanus hatched and fed by an Agelaius phoeniceus, 402-405 cooperative breeding of Todiramphiis veneratus, 46-50 cooperative polyandry in Sula siila, 361-365 costs and benefits of foraging alone or in mixed-species aggregations for Sterna forsteri. 95-101 foraging habits by Aerodramus fuciphagus inexpectatus and Collocalia esculenta affinis in the Andaman Islands, India, 259-272 hanging behavior of Corvus cornix. 183-185 hatching synchrony, green branch collecting, and prey use by nesting Harpia harpyja. 792-795 incubation temperature and behavior in thrushes nesting at high altitude, 666-673 infanticide by Crotophaga ani. 369-374 INDEX TO VOLUME 122 837 infanticide by Sayoniis phoehe, 620-622 nociin nal social cues attract niigratin^ icti^rici viri^ns^ 780-783 ohseivations of a possible foraging tool by Corviis corciw 181-182 Picoides borealis inale/temale foraging differences in young forest stands, 244-258 provisioning rates of Myiohorus miniutus in Monteverde, Costa Rica, 29-38 response to nestling throat ligatures by three songbirds, 806-809 site fidelity of Fatco span’eriits wintering in southwest- ern Florida, 475-483 thermoregulatory behavior in migratory Merops tipi- aster, 378-380 Benson, Thomas J., see Alessi, Mark G., , and Michael P. Ward Benson, Thomas J., see Pappas, Sara, , and James C, Bednarz Berardelli, Daniele, Martha J. Desmond, and Leigh Murray, Reproductive success of Bunowing Owls in urban and grassland habitats in southern New Mexico, 51-59 Biamonte, Esteban, A new bird species for Costa Rica: Sapphire-throated Hummingbird (Lepidopyga coer- ideogiilaris), 194-195 Bildstein, Keith L., see Hinnebusch, Daniel M., Jean- Fran9ois Therrien, Marc-Andre Valiquette, Bob Robertson, Sue Robertson, and Binford, Laurence C. and Joseph A. Youngman, Flight speeds of migrating Red-necked and Horned grebes, 374-378 Bird, David M., see Frei, Barbara, , and Rodger D. Titman Bittern, American, see Botaurus lentigiiiosiis Least, see Ixobiychiis exilis Pinnated, see Botaurus pinnatus Blackbird, Brewer's, see Euphagus cyanocephalus Chopi, see Gnorimopsar chopi Common, see Turdus merula Red- winged, see Agelaius pboeniceus Rusty, see Euphagus carolinus Yellow-headed, see Xanocephalus xanocephalus Blackcap, European, see Sylvia atricapilla blowfly, see Protocalliphora tundrae Bluebird, Eastern, see Sialia sialis Western, see Sialia mexicana bluefish, see Pomatomus saltatrix Boal, Clint W, see Strobe], Brad N. and bobcat, see Lynx rufus Bobolink, see Dolichonyx orxzivorus Bobwhite, Northern, see Colinus virginianus Boiga spp., 175 Bombycilla cedrorum. 290, 326, 329, 518-531, 725-743 Bonasa umbelliis. 2, 9, 290, 518-531 Bonter. David N., see Brooks, Elizabeth W. and Booby, Nazea, see Sula granti Red-footed, see Sula sula Botaurus lentiginosus, 768, 770 pinnatus. 108, I 14 botfly, see Philornis spp. Boves, Than J., David A. Buehicr, and Phillip C. Massey, Interspecific song imitation by a Cerulean Warbler, 583-587 Bracey, Elwood D., see Hayes, William K., . Melissa R. Price, Valerie Robinette, Eric Gren, and Caroline Stahala Bradypus tridactylus, 792, 793 Brant, Pacific Black, see Brania bernicia nigricans Branta bernicia nigricans, 484^93 canadensis. 491, 768, 770 hutcliinsii, 491 Initchinsii leucopareia, 2, 10 Braun, Clait E., review. Grouse of plains and mountains: the South Dakota story, by Lester D. Flake, John W. Connelly, Thomas R. Kirschenmann, and Andrew J. Lindbloom, 628-629 Braun, Clait E., see Kaler, Robb S. A., Steve E. Ebberl, , and Brett K. Sandercock breeding biology age of first breeding of Nipponia nippon in the Qinling Mountains, China, 228-235 annual precipitation affects reproduction of Lanins meridionalis, 334—339 breeding phenology and nesting success of Campvlo- rhynebus yucataniciis in the Yucatan Peninsula, Mexico, 439-446 cooperative breeding by Melanerpes erthrocephalus. 160-162 cooperative breeding by Todiramphus veneratus. 46-50 first reported case of cooperative polyandry in Sula sula. 361-365 first reproductive record of Charadrius wilsonia in Bata de Todos os Santos, northeastern Brazil, 788- 791 nest box use by Pants major in semi-arid rural residential gardens, 604-608 of Athene cunicuaria in urban and grassland habitats in southern New Mexico, 51-59 of Garrulax maximus at Lianhuashan, southern Gansu. China, 388-391 of Larus dominicanus on the Brazilian coast, 39—45 of Megalaiina nuchalis in Taipei Botanical Garden, 681- 688 of Melanopareia torquata. 1 62- 1 65 of Myiohorus miniatus in Monteverde, Costa Rica, 29- 38 of Myiohorus miniatus in Venezuela, 447-454 o{ Zimmerius cinysops in Venezuela, 689-698 parental care, food provisioning, and nest defense by Carpococcyx renauldi in northeastern Thailand 17.3-177 Passer domesticus associated with reduced Petrocheli- don pyrrhonota nesting success. 135-138 Brittain. Ross A., Vicky J. Merclsky, and Chris B. Craft. Avian communities of the Altamaha River estuary in Georgia, USA. 532-544 Brooks. Elizabeth W. and David N. Bonter. Long-term changes in avian community structure in a successional. forested, and managed plot in a reforesting landscape. 288-295 Bubalus huhalis. 559. 560 838 THE WILSON JOURNAL OF ORNITHOLOGY • Vol. 122, No. 4, December 2010 Bubo scandiacus [Nyctea scandiaca, 3, 10], 762 virginianus, 707, 768, 770 Bubulcus ibis, 102, 106, 108, 114, 212 Bucephala clangula, 377, 569 Buceros hydrocorax, 48 Buden, Donald W. and Stanley Retogral, Range expansion of the White-breasted Waterhen {Amaurornis phoenicurus) into Micronesia, 784-788 Buehler, David A., see Boves, Than J., , and Phillip C. Massey buffalo, water, see Bubalus bubalis Bulbul, Puff-throated, see Alophoxius pallidus bullsnake, see Pituophis catenifer [melanoleucus] Bunting, Conrmon Reed, see Emberiza schoeniclus Indigo, see Passerina cyanea Lark, see Calamospiza melanocorys Lazuli, see Passerina amoena Painted, see Passerina ciris Snow, see Plectrophenax nivalis Burgess, Nathanael J., see Butler, Christopher J., Lisa H. Pham, Jill N. Stinedurf, Christopher L. Roy, Eric L. Judd, , and Gloria M. Caddell Burhans, Dirk E., Brian G. Root, Terry L. Shaffer, and Daniel C. Dey, Songbird nest survival is invariant to early-successional restoration treatments in a large river floodplain, 307-317 Burhinus bistriatus, 105, 106, 108, 115 Bushtit, Long-tailed, see Aegithalos caudatus Buteo galapagoensis, 46, 364 jamaecensis, 68, 370, 518-544, 768, 771 lineatus, 68-74, 290, 532-544, 707, 713 lineatus alleni, 70 lineatus lineatus, 70 lineatus texanus, 70 platypterus, 171, 290, 518-531, spp., 801 swainsoni, 768, 771 Butler, Christopher J., Lisa H. Pham, Jill N. Stinedurf, Christopher L. Roy, Eric L. Judd, Nathanael J. Burgess, and Gloria M. Caddell, Yellow Rails wintering in Oklahoma, 385-387 Butorides striatus, 114 virescens, 108, 114, 768, 770 C Cacicus solitarius, 795-799 Caciques, Solitary, see Cacicus solitarius Caddell, Gloria M., see Butler, Christopher J., Lisa H. Pham, Jill N. Stinedurf, Christopher L. Roy, Eric L. Judd, Nathanael J. Burgess, and Cairina moschata, 1 1 4 Calamospiza melanocorys, 457 Calcarius mccownii, 460 ornatus, 346-353. 455-464 Calidris himantopus, 569 mauri, 15-22 melanotos, 569 pusilla [semipalmatus, 19], 19, 569 spp., 109, 114, 115 Callosciurus finlaysonii, 175 Calypte anna, 496 Campylorhynchus brunneicapillus, 443, 444, 624 nuchalis, 443 rufinucha, 443 yucatanicus, 439-446 Canary, Atlantic, see Serinus canaria Canastero, Berlepsch’s, see Asthenes berlepschi Cactus, see Asthenes cactorum Canyon, see Asthenes pudibunda Cip6, see Asthenes luizae Dusky-tailed, see Asthenes humicola Patagonian, see Asthenes patagonica Rusty-vented, see Asthenes dorbignyi Short-billed, see Asthenes baeri Steinbach’s, see Asthenes steinbachi Canis latrans, 457, 801, 802 lupus, 181, 182 lupus familiaris, 556, 559-561, 604-608, 615, 616, 801 Canvasback, see Aythya valisineria Cao, Lei, Guixia Zhao, Shan Tang, and Hanzhao Guo, The first reported case of cooperative polyandry in the Red-footed Booby: trio relationships and benefits, 361-365 Caprimulgus carolinensis, 381-384 capuchin, wedge-capped, see Cebus olivaceus Caracara plancus, 44 Caracara, Southern Crested, see Caracara plancus Cardellina rubrifrons, 35, 614-617 spp., 29, 36 Cardinal, Northern, see Cardinalis cardinalis Red-capped, see Paroaria gularis Cardinalis cardinalis, 168-173, 291, 326-333, 518-544 Carduelis citrinella, 490 Carpococcyx radiceus, 173, 175, 176 renauldi, 173-117 viridis, 173, 175 Carpodacus mexicanus, 291 purpureus, 291, 725-737 Castor canadensis, 274, 276 cat, domestic, see Felis catus Catbird, Gray, see Dumetella carolinensis caterpillar, eastern tent, see Malacosoma americana Cathartes aura, 354-360, 518-531 Catharus aurantiirostris, 503-512 frantzii, 507 fuscater, 507, 509 fuscescens, 290, 549-552, 725-737, 768, 771 guttatus, 82-87, 290, 302, 666-673 spp., 743 ustulatus, 82-87, 302, 507, 510, 545, 725-737, 742, 768, 771 Catops alpinus, 703, 704 Cebus apella, 679 olivaceus, 693, 792, 793 Centrocercus urophasianus, 2, 768, 769, 771 Centropus grillii, 755 spp., 173 Cercibis oxycerca, 108, 114 Certhia americana, 290, 725-737 Ceryle rudis, 48 Chadwick, Paul C., see Baldwin, Heather Q., Clinton W. INDEX TO VOLUME 122 839 Jeske, Melissa A. Powell, , and Wylie C. BaiTow Jr. Chaetura chapmani, 157 pelagica, 157, 518-531 spp., 153 Chandler, Robert M., Peter Pyle, Maureen E. Flannery, Douglas J. Long, and Steve N. G. Howell, Flight feather molt of Turkey Vultures, 354-360 Charadrius alexandrinus, 23, 25 alexandrinus alexandrinus, 25 alexandrinus nivosus, 25 collaris, 105, 115 melodus, 23, 25, 578-582 morinellus, 578 semipalmatus, 23-28 vociferus, 62, 768, 771 wilsonia, 532-544, 788-791 wilsonia brasiliensis, 790 Charter, Motti, Yossi Leshem, Shay Halevi, and Ido Izhaki, Nest box use by Great Tits in semi-arid rural residential gardens, 604-608 Chartier, Allen T., review. Do hummingbirds hum? Fascinating answers to questions about humming- birds, by George C. West and Carol A. Butler, 818- 819 Chat, Yellow-breasted, see Icteria virens Chavama-Pizarro, Tania, Gustavo Gutierrez-Espeleta, Eric J. Fuchs, and Gilbert Barrantes, Genetic and morphological variation of the Sooty-capped Bush Tanager (Chlorospingus pileatus), a highland endemic species from Costa Rica and western Panama, 279-287 Chavez-Ramirez, Felipe, see Aguilera, Xiomara Galvez and Chen [Anser, 19] caerulescens caerulescens, 491 Chen, Yang, see lie, Wang, , Lu Nan, Fang Yun, and Sun Yue-Hua Cheng, Wei-Jen, see Lin, Shang-Yao, Fang-Cian Lu, Fu- Hsien Shan, Shu-Ping Liao, Jing-Ling Weng, , and Chao-Nien Koh Chickadee. Black-capped, see Poecile atricapillus Boreal, see Poecile hudsonicus Carolina, see Poecile carolinensis Mountain, see Poecile gambeli Chicken, Domestic, see Callus gallus domesticus chimpanzee, see Pan spp. chipmunk, cliff, see Tamias dorsalis eastern, see Tamias striatus Chloroceryle aenea, 106, 109, 115 amazona, 115 americana, 109, 115 inda, 106, 109, 115 Chlorospingus pileatus, 279-287 zeldoni, 279 Chlorospingus, Zeledon’s, see Chlorospingus zeldoni Chlorostilbon assimilis, 597-599 canivetti, 496 Chordeiles gundlachii, 381 minor, 768, 770 Chroicocephalus ridibundus, 755 Chuck- will’s- widow, see Caprimulgus carolinensis Chukar, see Alectoris chukar cicada, annual, see Tibicen spp. periodic, see Magicicada spp. Ciconia ciconia, 233, 755, 756 Cinclus cinclus, 563-568 leucocephalus, 563-568 mexicanus, 563-568 pallassi, 563-568 schulzi, 563-568 Circus cyaneus, 456, 768, 771, 801 Cistothorus palustris, 62, 532-544, 623-624 Clangula hyemalis, 377 Clarke, Julia A., review, Paleogene birds, by Gerald Mayr, 409^10 climate change phenology of migration in relation to climate change, 116-125 Coccyzus americanus, 402M05, 518-544, 768, 771 erythropthalmus, 290, 402 Cockle, Knstina L., Interspecific cavity-sharing between a Helmeted Woodpecker {Dryocopus galeatus) and two White-eyed Parakeets {Aratinga leucophthalma), 803-806 Cohen, Jonathan B. and James D. Fraser, Piping Plover foraging distribution and prey abundance in the pre- laying period, 578-582 Colaptes auratus, 128, 131, 475, 518-531, 608, 725-737, 768, 771 Colibri thalassinus, 496 Colinus virginianus, 518 Collocalia esculenta qffinis, 259-272 Coluber constrictor, 801 constrictor flaviventris, 800 Columba livia, 62, 741 Columbina passerina, 532-544 Condor, California, see Gymnogyps califomianus Confer, John L., Kevin W. Barnes, and Erin C. Alvey, Golden- and Blue-winged warblers: distiibution, nesting success, and genetic differences in two habitats, 273-278 conservation status of Crus canadensis nesiotes, 556-562 Contopus cooperi, 773 virens, 518-544, 725-737 Contreras, Francisco Javier, see Seymour, Adrian S., Graham Hatherley, , James Aldred, and Fergus Beeley Conway, Courtney J., see Kirkpatrick, Chris and Conway, Courtney J., see Robinson, Gabrielle L., , Chris Kirkpatrick, and Dominic D. LaRoche Cooper, Robert J., see Klaus, Nathan A., Scott A. Rush, Tim S. Keyes, John Petrick, and Coot, American, see Fulica americana Coragyps atratus, 41, 44, 518-531 Cormorant, Double-crested, see Phalacrocorax auritus Neotropic, see Phalacrocorax brasilianus Reed, see Phalacrocorax africanus Corvus brachyrhynchos, 290, 296, 301, 302, 470, 518-544 768, 770, 802 corax, 3, 181-185, 302, 616, 768, 770 840 THE WILSON JOURNAL OL ORNITHOLOGY • IW. 122. No. 4. December 2010 coroiie. 184 o.^sifragiis. 532-544 Coiurnicops noveboracensis, 385-387 Cotiirni.x cotiirnix, 790 japonica. 615, 755, 756, 759 Coucal, Black, see Centropus grillii Cowbird, Brown-headed, see Mololhni.id Courtney J. Conway Feeding a foreign chiek: a case ol a mixed brood of two Tit species Toshitaka N. Suzuki and Yuko Tsuchiya Infanticide by an Eastern Phoebe Gary Ritchison and Brandon T. Ritchison Marsh Wren eats small fish Andrea J. Ayers and James W. Arrnacost Jr. Occasional mimicry and night-time singing by the western Curve-billed Thrasher {Toxostoma curvirostre palmeri) R. Roy Johnson and Lois T. Haight Ornithological Literature Robert B. Payne, Book Review Editor Major Articles NUMBER 4 Breeding patterns of Henslow’s Sparrow and sympatric grassland sparrow species L. Lynnette Dornak Influence of woody vegetation on grassland birds within reclaimed surface mines Bret M. Graves, Amanda D. Rodewald, and Scott D. Hull Effects of breeding stage and behavioral context on singing behavior of male Indigo Buntings Matthew D. Beckett and Gary Ritchison Development of incubation temperature and behavior in thrushes nesting at high altitude Martin L. Morton and Maria E. Pereyra Flexibility in nest-site choice and nesting success of Turdus rujiventris (Turdidae) in a montane forest in northwestern Argentina Silvia B. Lomdscolo, A. Carolina Monmany Agustina Malizia, and Thomas E. Martin Breeding biology of the Taiwan Barbet {Megalaima nuchalis) in Taipei Botanical Garden Shang-Yao Lin, Fang-Cian Lu, Fu-Hsien Shan, Shu-Ping Liao, Jing-LingWeng, Wei-Jen Cheng, and Chao-Nien Koh Breeding biology of the Golden-faced Tyrannulet [Zimmerius chrysops) in Venezuela William Goulding and Thomas E. Martin Habitat type is related to nest mass and fledging success of Arctic Warblers Julie C. Hagelin, Margaret L. Perry, Ben S. Ewen-Campen, Derek S. Sikes, and Susan Sharbaugh Home range, habitat use, and nest site characteristics of Mississippi Kites in the White River National Wildlife Refuge, Arkansas Troy J. Bader and James C. Bednarz Biological, geographical, and cultural origins ol the loon hunting tradition in Carteret County, North Carolina Storrs L. Olson, Horace Loftin, and Steve Goodwin Migrant songbird species distribution and habitat use during stopover on two islands in the Gulf of Maine Rebecca W. Suomala, Sara R. Morris, Kimberly J. Babbitt, and Thomas D. Lee Concentrated migratory morning Hight at Lake Pontchartrain, Louisiana, USA Peter H. Yaukey Night migrant fatalities and obstruction lighting at wind turbines in North America Paul Kerlinger, Joelle L. Gehring, Wallace P. Erickson, Richard Curry, Aaftab Jain, and John Guarnaccia Developmental changes in serum androgen levels of Eastern Screech-Owls {Megascops asio) Corinne P. Kozlowski and D. Caldwell Hahn Stress responsiveness decreases with age in precocial, juvenile Chukar Molly J. IDickens and L. Michael Romero 767 111 111 780 784 788 792 795 799 803 806 809 810 812 820 833 834 835 Short Communications Composition dynamics of an avian assemblage Fritz L. Knopf A digital spot-mapping method for avian field studies Kevin E. JabLonski, Stacy A. McNulty, and Matthew D. Schlesinger Variation of Type B song in the endangered Golden-cheeked Warbler {Dendroica chrysoparia) Wendy J. Leonard, Jayne Neal, and Rama Ratnam Nocturnal social cues attract migrating Yellow-breasted Chats Mark G. Alessi, Thomas J. Benson, and Michael P. Ward Range expansion of the White-breasted Waterhen {Amaurornis phoenicurus) into Micronesia Donald W Buden and Stanley Retogral First reproductive record of Wilson’s Plover in Baia de Todos os Santos, northeastern Brazil Vitor O. Lunardi and Regina FI. Macedo Hatching synchrony, green branch collecting, and prey use by nesting Harpy Eagles {Harpia harpyjd) Adrian S. Seymour, Graham Hatherley, Francisco Javier Contreras, James Aldred, and Fergus Beeley Screaming Cowbird parasitism of nests of Solitary Caciques and Cattle Tyrants Alejandro G. Di Giacomo, Bettina Mahler, and Juan C. Reboreda Snakes are important nest predators of Dickcissels in an agricultural landscape Page E. Klug, L. LaReesa Wolfenbarger, and John P. McCarty Interspecific cavity-sharing between a Helmeted Woodpecker {Dryocopus galeatus) and two White-eyed Parakeets {Aratinga leucophthalma) Kristina L. Cockle Response to nesding throat ligatures by three songbirds Gabrielle L. Robinson, Courtney J. Conway, Chris Kirkpatrick, and Dominic D. LaRoche Editors of The Wilson Bulletin {iSS8-xoo^) and The Wilson Journal of Ornithology (2006-Z010) William and Nancy Klamm Service Award for zoio Ornithological Literature Robert B. Payne, Book Review Editor Proceedings of the Ninety-First Annual Meeting Reviewers for Volume izz Editor’s Comments Index to Volume izz Contents of Volume izz -A * \ . 1 fci'V» *flf *■— - Iff— >' £: 9 THE WILSON JOURNAL OL ORNITHOLOGY Editor CLAIT E. BRAUN 5572 North Ventana Vista Road Tucson, AZ 85750-7204 E-mail: TWILSONJO(S)comcast.net Editorial NANCY J. K. BRAUN Assistant Editorial Board RICHARD C. BANKS JACK CLINTON EITNIEAR SARA J. OYLER-McCANCE LESLIE A. ROBB Review Editor ROBERT B. PAYNE 1306 Granger Avenue Ann Arbor. MI 48104, USA E-mail: rbpayne(S)umich.edu GUIDELINES LOR AUTHORS Please consult the detailed “Guidelines for Authors” found on the Wilson Ornithological Society web site (http://www.wilsonsociety.org). All manuscript submissions and revisions should be sent to Clait E. Braun, Editor, The Wilson Journal of Ornithology, 5572 North Ventana Vista Road, Tucson, AZ 85750-7204, USA. The Wilson Journal of Ornithology office and fax telephone number is (520) 529-0365. The e-mail address is TWilsonJO@comcast.net NOTICE OF CHANGE OF ADDRESS Notify the Society immediately if your address changes. Send your complete new address to Ornithological Societies of North America, 5400 Bosque Boulevard, Suite 680, Waco, TX 76710. The permanent mailing address of the Wilson Ornithological Society is: %The Museum of Zoology, The University of Michigan. Ann Arbor, MI 48109, USA. Persons having business with any of the officers may address them at their various addres.ses given on the inside of the front cover, and all matters pertaining to the journal should be sent directly to the Editor. MEMBERSHIP INQUIRIES Membership inquiries should be sent to Timothy J. O'Connell. Department of Natural Resources Ecology and Management, Oklahoma State University, 240 AG Hall, Stillwater, OK 74078; e-mail: oconnet@ okstate.edu THE JOSSELYN VAN TYNE MEMORIAL LIBRARY The Josselyn Van Tyne Memorial Library of the Wilson Ornithological Society, housed in the University of Michigan Mu.seum 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 ornithological magazines from members and friends of the Society. Two members have generously established a fund for the purchase of new books; members and friends are invited to maintain the fund by regular contribution. The fund is administered by the Library Committee. Jerome A. Jackson, Florida Gulf Coast Univeristy, is Chairman of the Committee. The Library currently receives over 200 periodicals as gifts and in exchange for The Wilson Journal of Ornithology. For infonnation on the Library and our holdings, see the Society’s weh page at http://www.wilsonsociety.org. With the usual exception of rare books, any item in the Library may be boiTowed 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. I 109 Geddes Avenue, Ann Arhor, MI 48109-1079. USA. Contributions to the New Book Fund should be sent to the Treasurer. This issue of The Wilson Journal of Ornithology was published on I December 2010. 767 111 111 780 784 788 792 795 799 803 806 809 810 812 820 833 834 835 Continued from outside back cover Short Communications Composition dynamics of an avian assemblage Fritz L. Knopf A digital spot-mapping method for avian field studies Kevin E. Jablonski, Stacy A. McNulty, and Matthew D. Schlesinger Variation of Type B song in the endangered Golden-cheeked Warbler {Dendroica chrysoparia) Wendy J. Leonard, Jayne Neal, and Rama Ratnam Nocturnal social cues attract migrating Yellow-breasted Chats Mark G. Alessi, Thomas J. Benson, and Michael P. Ward Range expansion of the White-breasted Waterhen [Amaurornis phoenicurus) into Micronesia Donald W Buden and Stanley Retogral First reproductive record of Wilson’s Plover in Bala deTodos os Santos, northeastern Brazil Vitor O. Lunardi and Regina FI. Macedo Hatching synchrony, green branch collecting, and prey use by nesting Harpy Eagles {Harpia harpyja) Adrian S. Seymour, Graham Hatherley, Francisco Javier Contreras, James Aldred, and Fergus Beeley Screaming Cowbird parasitism of nests of Solitary Caciques and Cattle Tyrants Alejandro G. Di Giacomo, Bettina Mahler, and Juan C. Reboreda Snakes are important nest predators of Dickcissels in an agricultural landscape Page E. Klug, L. LaReesa Wolfenbarger, and John P. McCarty Interspecific cavity-sharing between a Helmeted Woodpecker {Dryocopus galeatus) and two White- eyed Parakeets {Aratinga leucophthalma) Kristina L. Cockle Response to nestling throat ligatures by three songbirds Gabrielle L. Robinson, Courtney J. Conway, Chris Kirkpatrick, and Dominic D. LaRoche Editors of The Wilson Bulletin (iSSS-zoo^) and The Wilson Journal of Ornithology (2006-2010) William and Nancy Klamm Service Award for 2010 Ornithological Literature Robert B. Payne, Book Review Editor Proceedings of the Ninety-First Annual Meeting Reviewers for Volume 122 Editor’s Comments Index to Volume 122 Contents of Volume 122 The Wilson Journal of Ornithology (formerly The Wilson Bulletin) Volume 122, Number 4 CONTENTS December 20 1C Major Articles 635 Breeding patterns ol Henslows Sparrow and sympatric grassland sparrow species L. Lynnette 'Dornak 646 Influence of woody vegetation on grassland birds within reclaimed surface mines Bret M. Graves, Amanda D. Rodewald, and Scott D. Bull 655 Effects of breeding stage and behavioral context on singing behavior of male Indigo Buntings Matthew D. Beckett and Gary Ritchison 666 Development of incubation temperature and behavior in thrushes nesting at high altitude Martin L. Morton and Maria E. Pereyra 674 Flexibility in nest-site choice and nesting success of Turdus rufiventris (Turdidae) in a montane forest in northwestern Argentina Silvia B. Lomdscolo, A. Garolina Monmany, Agustina Malizia, and Thomas E. Martin 68 1 Breeding biology of the Taiwan Barbet {Megalaima nuchalis) in Taipei Botanical Garden Shang-Yao Lin, Eang-Cian Lu, Eu-Hsien Shan, Shu-Ping Liao, Jing-Ling Weng, Wei-Jen Cheng, and Chao- Nien Koh 689 Breeding biology of the Golden-faced Tyrannulet {Zimmerius chrysops) in Venezuela William Goulding and Thomas E. Martin 699 Habitat type is related to nest mass and fledging success of Arctic Warblers Julie C. Hagelin, Margaret L. Perry, Ben S. Ewen-Campen, Derek S. Sikes, and Susan Sharbaugh 706 Home range, habitat use, and nest site characteristics of Mississippi Kites in the White River National Wildlife Refuge, Arkansas Troy J. Bader and James C. Bednarz 716 Biological, geographical, and cultural origins of the loon hunting tradition in Garteret County, North Carolina Storrs L. Olson, Horace Lojiin, and Steve Goodwin 725 Migrant songbird species distribution and habitat use during stopover on two islands in the Gulf of Maine Rebecca W. Suomala, Sara R. Morris, Kimberly J. Babbitt, and Thomas D. Lee 738 Concentrated migratory morning flight at Lake Pontchartrain, Louisiana, USA Peter H. Yaukey 744 Night migrant fatalities and obstruction lighting at wind turbines in North America Paul Kerlinger, Joelle L. Gehring, Wallace P Erickson, Richard Curry, Aajiab Jain, and John Guarnaccia 755 Developmental changes in serum androgen levels of Eastern Screech-Owls {Megascops asio) Gorinne P. Kozlowski and D. Caldwell Hahn 762 Stress responsiveness decreases with age in precocial, juvenile Chtikar MollJ J. Dickens and L. Michael Romero Continued on inside hack cover • [. I * f ■■ t * .. « * " I % I ^-4*- 4 P ■ - I MCZ ERNST MAYR LIBRARY lllllll III lllllll III III 3 2: )44 118 689 4 05