WESTERN Vol. 44, No. 1, 2013 Western Specialty: Gray-crowned Rosy-Finch E h hoto by 45 Thomas A. Blackman of San Dicp, California: A JuLl Gray-crowned Rosy-Finch at the type locality of subspecies Lsacoaticte tephrocatis urn hr in a, Si. Paul Island, Pribilof Islands. Alaska., 7 August 2012. This subspecies of (he Gray -crowned Rosy-Finch is endemic to (he Pribilof Islands, SL. Mali hew Island, and Hall Island in the central Bering Sea. It is a large subspecies, —52 grams, almost as large as the subspecies of (he Aleutian Islands (L, f. griseottuefta) or an Evening Grosbeak (Coccethraustes vespertintts). It differs from subspecies griseottuch# in its blacker color, on both the back and underparts, Volume 44, Number 1, 2013 A Twenty- Year Investigation of the Effects of Fire on a Coastal Sage Scrub Bird Community David J . Moriarty 2 2011 Nevada Bird Records Committee Report Martin Meyers 17 Nesting Ecology and Nest Success of the Blue Grosbeak along Two Rivers in New Mexico Jean-Luc E. Cartron, Deborah M. Finch, David L. Hawksworth, and Scott H. Stoleson 33 Distribution and Movement Patterns of Individual Crested Caracaras in California Kristie N. Nelson and Peter Pyle 45 Conspecific Nest Aggression of the Pacific Wren on Vancouver Island, British Columbia Ann Nightingale and Ron Melcer, Jr. 56 NOTES First Record of a Surfbird in the Hawaiian Islands Eric A. VanderWerf 65 Common Nesting Habitats and Weights at Fledging of Wedge-tailed Shearwaters on Tern Island, Hawaii Phillip J. Howard, Sarah C. Harvey, Paula L. Hartzell, Pete Leary, and Ty J. Penally 69 An Apparent Long-Distance Flight by a Dusky Grouse in Montana Ronald J. Kienholz and Aleen M. Kienholz 76 Melanistic Adult Male Northern Harrier Wintering in Idaho Robert A. Miller, Neil Paprocki, and Elizabeth H. Urban 78 In Memoriam: Gale Monson Richard L. Glinski 80 Book Review Lauren Harter 82 Thank You to Our Supporters 83 President’s Message Edward R. Pandolfino 84 Featured Photo: Carotenism in the Hairy Woodpecker Ron LeValley and Jeff N. Davis 85 Front cover photo by © Sophie Webb of Felton, California: North- ern Gannet ( Morus bassanus ), Southeast Farallon Island, San Francisco County, California, 25 April 2012. It remained at least through February 2013 and is the first of this Atlantic species recorded in the North Pacific Ocean. Possibly the recent decrease of arctic ice allowed it to traverse the Arctic Ocean and enter the Pacific through the Bering Strait. Back cover: “Featured Photo” by © Beth Hamel of Morgan Hill, California: Hairy Woodpecker ( Picoides villosus ), Golden Gate Park, San Francisco County, California, 26 October 2012. The yellow rectrices represent an unusual case of carotenism, affect- ing both the type and distribution of carotenoid pigment. Yet such carotenistic Hairy Woodpeckers have been noted repeatedly in coastal northern California. Western Birds solicits papers that are both useful to and understandable by ama- teur field ornithologists and also contribute significantly to scientific literature. Send manuscripts to Daniel D. Gibson, P. O. Box 155, Ester, AK 99725; avesalaska@ gmail.com. For matters of style consult the Suggestions to Contributors to Western Birds (at www.westernfieldornithologists.org/docs/journal_guidelines.doc). WESTERN BIRDS Volume 44, Number 1, 2013 A TWENTY- YEAR INVESTIGATION OF THE EFFECTS OF FIRE ON A COASTAL SAGE SCRUB BIRD COMMUNITY DAVID J. MORIARTY, Biological Sciences Department, California State Polytechnic University, Pomona, California 91768; djmoriarty@csupomona.edu ABSTRACT: From 1983 to 2003 I examined the effects of fire on the bird com- munity of two 1.25-ha tracts of coastal sage scrub, Box Canyon and F Canyon, 0.4 km apart in a 31 -ha reserve in Los Angeles County, California. Wildfire burned Box Canyon in 1981 and both sites in 1989. 1 observed 90 species in F Canyon, 80 in Box Canyon, of which 73 were seen in both. The same species were common throughout the 20-year period. F Canyon had more species per count than Box Canyon in both summer and winter. Immediately after the 1989 wildfire, observations of some spe- cies of open habitat increased, and observations of some species confined to shrubs decreased. Effects of the 1989 fire on the sites were of short duration. Differences between the sites in number of species attributed to the 1981 fire in earlier studies are confounded by differences between the sites. Long-term studies provide insight into the organization and dynamics of avian communities (Holmes et al. 1986, Brawn and Robinson 1996, Herrera 1998). Understanding differences among sites in richness of bird species is important in fragmented habitats such as coastal sage scrub, which has been reduced by urbanization and whose biodiversity is the subject of concern (Westman 1981, McCaull 1994, Chase et al. 2000, Cox and Allen 2008). In a brief period, such as a single breeding season, differences in species richness between sites may represent random variation, be transitory, or may indicate a long-term difference (Chase et al. 2000). Long-term data address this ambiguity. The effect of fire on birds of southern California’s coastal sage scrub has been examined with varying results and conclusions. Wirtz (1982) and Mendelsohn et al. (2008) reported increased species richness in chaparral and coastal sage scrub, respectively, but Moriarty et al. (1985) and Stanton (1986) reported richness in burned coastal sage scrub decreased from that at an unburned site. Mendelsohn et al. (2008) suggested discrepancies between their study and that of Stanton (1986) may have been due to differences 2 Western Birds 44:2-16, 2013 THE EFFECTS OF FIRE ON A COASTAL SAGE SCRUB BIRD COMMUNITY Table 1 Ranges of Dates of Blocks of Counts and Number of Species Seen by Season at the Voorhis Ecological Reserve, Pomona, California, 1983-2003 F Canyon Box Canyon Dates n a Summer Winter Total Summer Winter Total Block total 20 Jan 1983- 16 May 1984 56 49 47 65 46 38 60 79 17 Oct 1986- 31 Dec 1988 41 38 29 44 31 18 37 54 30 Jul 1989- 11 Nov 1991 fa 39 46 37 57 38 27 49 66 6 Dec 1992- 31 Aug 1996 85 51 50 61 35 32 42 65 23 Jul 2000- 22 Jul 2001 18 26 24 31 20 21 28 37 19 Oct 2002- 30 Dec 2003 26 29 29 39 27 31 37 45 Total (all blocks) 265 78 71 90 66 60 80 97 “Number of paired counts in each block. fa The study area burned on 28 July 1989, two days before the start of the third count block. between the sites they studied in San Diego and Los Angeles counties, respectively. Differences unrelated to fire may have confounded the results. In this paper my objectives are (1) to describe species richness and com- munity composition of birds at two sites in coastal sage scrub over 20 years, (2) to compare the two sites with respect to richness and composition and examine how these differences affect the conclusions of previous studies of the effect of fire at these sites (Moriarty et al. 1985, Stanton 1986), and (3) to describe the effect on richness and composition of a wildfire that burned both sites in the sixth year of the study. METHODS Study Sites My study area was the 31 -ha Voorhis Ecological Reserve in the San Jose Hills on the campus of California State Polytechnic University, Pomona, Los Angeles County, California (34° 03'N, 117° 49' W). I established two 1.25- ha sites at approximately 300 m elevation and 0.4 km apart, connected by contiguous vegetation. The first site (Box Canyon) was burned in a wildfire on 21 August 1981 (Moriarty et al. 1985). The second site (F Canyon) was not burned and was used as a control by Moriarty et al. (1985) and Stanton (1986). The sites are both south-facing canyons and appear similar in their topography and vegetation structure (Moriarty et al. 1985, Stanton 1986). A second wildfire of apparent accidental origin burned the entire reserve 3 THE EFFECTS OF FIRE ON A COASTAL SAGE SCRUB BIRD COMMUNITY Table 2 Numbers of Times Each Species Was Observed at the Voorhis Ecological Reserve, Pomona, California, 1983-2003, by Site and Season Species 0 and residency status fa F Canyon Summer Winter Box Canyon Summer Winter 1. Anna’s Hummingbird, Calypte anna, P 98 117 115 110 2. California Towhee, Melozone crissalis, P 121 100 116 74 3. Western Scrub-Jay, Aphelocoma californica, P 100 87 111 83 4. Northern Mockingbird, Mimus polyglottos, P 106 79 94 48 5. House Finch, Haemorhous mexicanus, P 104 75 74 59 6. Spotted Towhee, Pipilo maculatus, P 86 61 59 25 7. Wrentit, Chamaea fasciata, P 74 64 44 30 8. Mourning Dove, Zenaida macroura , P 89 39 41 20 9. California Quail, Callipepla californica, P 53 68 46 24 10. Cactus Wren, Campylorhynchus brunneicapillus, P 81 53 4 1 11. Bushtit, Psaltriparus minimus, P 31 38 62 34 12. California Thrasher, Toxostoma redivivum, P 58 44 36 19 13. Lesser Goldfinch, Spinus psaltria, P 31 31 34 35 14. Bewick’s Wren, Thryomanes bewickii, P 40 38 33 16 15. Northern Flicker, Colaptes auratus, P 10 36 15 65 16. Red-tailed Hawk, Buteo jamaicensis, P 20 43 9 22 17. Rufous-crowned Sparrow, Aimophila ruficeps, P 29 37 24 12 18. Yellow-rumped Warbler, Setophaga coronata, W 3 27 5 54 19. Phainopepla, Phainopepla nitens, S 58 1 17 0 20. White-crowned Sparrow, Zonotrichia leucophrys, W 7 45 5 18 21. Nuttall’s Woodpecker, Picoides nuttallii, P 23 23 12 2 22. Black Phoebe, Sayornis nigricans, P 15 14 9 8 23. Song Sparrow, Melospiza melodia, P 16 14 7 6 24. House Wren, Troglodytes aedon, P 18 11 11 8 25. Ash-throated Flycatcher, Myiarchus cinerascens, S 24 1 17 2 26. Ruby-crowned Kinglet, Regulus calendula, W 3 29 0 4 27. Black-headed Grosbeak, Pheucticus melanocephalus, S 15 1 21 2 28. American Kestrel, Falco sparverius, P 7 6 6 12 29. American Robin, Turdus migratorius, W 2 17 0 14 30. Brown-headed Cowbird, Molothrus ater, S 16 1 15 1 31. Cooper’s Hawk, Accipiter cooperii, P 7 7 4 8 32. Blue-gray Gnatcatcher, Polioptila caerulea, W 4 15 2 1 33. Western Kingbird, Tyrannus verticals, S 8 1 11 1 (continued) 4 THE EFFECTS OF FIRE ON A COASTAL SAGE SCRUB BIRD COMMUNITY Table 2 ( continued ) Species 0 and residency status' 5 F Canyon Summer Winter Box Canyon Summer Winter 34. California Gnatcatcher, Polioptila calif ornica , P 11 5 5 2 35. Common Raven, Corvus corax, P 4 7 2 9 36. Cedar Waxwing, Bombycilla cedrorum , W 1 13 0 8 37. Black-chinned Hummingbird, Archilochus alexandri, S 10 1 10 0 38. Wilson’s Warbler, Cardellina pusilla, S 7 2 10 0 39. Cliff Swallow, Petrochelidon pyrrhonota, S 8 0 9 1 40. Bullock’s Oriole, Icterus bullockii, S 10 0 7 1 41. American Goldfinch, Spinus tristis, W 3 7 2 5 42. Loggerhead Shrike, Lanius ludouicianus, P 6 5 3 1 43. Hermit Thrush, Catharus guttatus, W 5 8 0 1 44. Western Bluebird, Sialia mexicana, W 0 8 0 10 45. Say’s Phoebe, Sayornis saya, W 0 1 3 10 46. Sharp-shinned Hawk, Accipiter striatus, W 0 8 0 5 47. Western Meadowlark, Sturnella neglecta, XP 4 3 3 6 48. Pacific-slope Flycatcher, Empidonax difficilis, S 8 1 4 1 49. Dark-eyed Junco, Junco hyemalis, W 1 4 0 7 50. Acorn Woodpecker, Melanerpes formiciuorus, P 3 6 2 1 51. Lazuli Bunting, Passerina amoena, XS 5 0 8 0 52. American Crow, Corvus brachyrhynchos, S 4 3 2 2 53. Northern Harrier, Circus cyaneus, W 0 7 0 3 54. Hooded Oriole, Icterus cucullatus, S 6 0 4 0 55. Turkey Vulture, Cathartes aura, XP 1 4 2 5 56. Golden-crowned Sparrow, Zonotrichia atricapilla, W 1 5 0 2 57. Cassin’s Kingbird, Tyrannus vociferans, XP 2 3 2 1 58. White-tailed Kite, Elanus leucurus, XP 4 1 0 1 59. Western Tanager, Pirartga ludoviciarta, XM 3 0 3 0 60. Rock Wren, Salpinctes obsoletus, W 0 5 0 0 61. Greater Roadrunner, Geococcyx calif ornianus, XP 4 0 1 0 62. Fox Sparrow, Passerella iliaca, W 1 3 0 1 63. Western Wood-Pewee, Contopus sordidulus, XS 2 1 2 0 64. Chipping Sparrow, Spizella passerina, XP 0 2 3 0 65. Costa’s Hummingbird, Calypte costae, XS 0 0 5 0 66. Yellow Warbler, Setophaga petechia, XS 3 1 1 0 67. Violet-green Swallow, Tachycineta thalassina, XP 2 1 1 0 68. Orange-crowned Warbler, Oreothlypis celata, XW 0 3 1 0 69. Townsend’s Warbler, Setophaga townsendi, XW 2 0 1 1 ( continued ) 5 THE EFFECTS OF FIRE ON A COASTAL SAGE SCRUB BIRD COMMUNITY Table 2 ( continued ) F Canyon Species 0 and residency status' 5 Summer Winter 70. Allen’s Hummingbird, Selasphorus sasin , XS 3 0 71. Swainson’s Thrush, Catharus ustulatus, XM 2 1 72. Band-tailed Pigeon, Patagioenas fasciata, XP 1 1 73. Vaux’s Swift, Chaetura vauxi, XM 1 0 74. Purple Finch, Haemorhous purpureus, XP 0 0 75. Lark Sparrow, Chondestes grammacus , XP 1 1 76. Lawrence’s Goldfinch, Spinus lawrencei, XP 1 1 77. Prairie Falcon, Falco mexicanus, XP 0 2 78. Great Horned Owl, Bubo virginianus, XP 2 0 79. Spotted Dove, Streptopelia chinensis, XP 1 1 80. European Starling, Sturnus vulgaris, XP 1 0 81. House Sparrow, Passer domesticus, XP 1 0 82. Gray Flycatcher, Empidonax wrightii, XS 1 0 83. Red-winged Blackbird, Agelaius phoeniceus, XP 0 0 84. White-throated Swift, Aeronautes saxatalis, XP 0 0 85. Golden Eagle, Aquila chrysaetos, XP 10 86. Lincoln’s Sparrow, Melospiza lincolnii, XW 0 1 87. Oak Titmouse, Baeolophus inornatus, XP 0 1 88. Warbling Vireo, Vireo gilvus, XM 0 1 89. Savannah Sparrow, Passerculus sandwichensis, XW 1 0 90. Mountain Chickadee, Poecile gambeli, XP 1 0 91. Cassin’s Vireo, Vireo cassinii, XS 10 92. Northern Rough-winged Swallow, Stelgidopteryx serripennis, XS 10 93. Sage Sparrow, Artemisiospiza belli, XP 0 1 94. Rufous Hummingbird, Selasphorus rufus, XM 1 0 95. Plumbeous Vireo, Vireo plumbeus, XS 0 0 96. MacGillivray’s Warbler, Geothlypis tolmiei, XM 0 0 97. Common Poorwill, Phalaertoptilus nuttallii, XP 0 0 Box Canyon Summer Winter 0 0 1 0 0 0 0 0 0 0 1 1 1 0 0 0 2 3 0 0 0 0 0 0 0 0 1 1 1 1 1 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 0 1 0 1 0 0 1 “Listed from most to least frequently seen; those in bold are the 35 most common species used in some statistical analyses. Numbers in bold represent species seen significantly more often in F Canyon than in Box Canyon within a season; numbers underlined are species seen significantly more often within season in Box Canyon (x 2 goodness of fit to 1:1 ratio with exact probabilities, all P < 0.05). b Residency status determined from data: P, permanent resident; S, summer (April-September) resident; W, winter (October-March) resident; M, migrant; X, categorized not from the infre- quent observations but from range maps (Sibley 2000) and personal observation. 6 THE EFFECTS OF FIRE ON A COASTAL SAGE SCRUB BIRD COMMUNITY including both sites on 28 July 1989. The fire was restricted mainly to the reserve. Adjacent undeveloped coastal sage scrub in the San Jose Hills, as well as developed areas, did not burn. The region encompassing the Voorhis Ecological Reserve has cool, wet winters, dry, hot summers, and is prone to fire (Westman 1981, Minnich 1983). The dominant native shrubs at both sites were the California Sagebrush ( Artemisia californica), California Buckwheat (Eriogonum fasciculatum), and Black Sage ( Salvia mellifera). Large shrubs and trees were present, including Coast Live Oak ( Quercus agrifolia ), Western Sycamore {Platanus racemosa ), Laurel Sumac ( Malosma laurina), California Walnut ( Juglans californica ), Elderberry ( Sambucus mexicana), and Toyon (. Heteromeles arbutifolia). Prevalent exotic herbs were Mediterranean Mustard ( Hirschfeldia incana), Black Mustard ( Brassica nigra), Tocalote [Centaurea melitensis), Red Brome ( Bromus rubens ), Ripgut Grass ( Bromus diandrus), and Horehound (Marrubium vulgare) (Moriarty et al. 1985, Stanton 1986, Clark 1990). Bird Counts Bird counts commenced on 20 January 1983, but I define 21 August 1981, the date of the first fire, as “day zero.” The 265 pairs of counts, of one count each in Box Canyon and F Canyon, were distributed in six blocks or periods of 1 to 3 years from 1983 to 2003 (Table 1). Within each block, paired counts were conducted approximately twice per week. Within a pair, each site was visited sequentially on one day, then the order of sites was switched on the next count day. Counts began shortly after sunrise, and entailed the observer standing quietly at a defined peripheral location that allowed an unobstructed view of the entire site. Birds flying over the sites were not recorded. Individual birds were not marked and density was not estimated, so analysis is restricted to presence/absence and numbers of species. I define counts from April through September as summer (n = 135), those from October through March as winter (n = 130). These broad seasonal cat- egories did not provide resolution to distinguish categories such as migrants or juveniles but did allow each commonly observed species to be categorized as a summer resident (observed in at least four of the six summer months, n = 12, Table 2) winter resident (observed in at least four of the six winter months; n = 16), or permanent resident (observed in at least four months of both seasons; n = 27). I categorized species not meeting any of these criteria (n = 42) (Table 2) on the basis of range maps (Sibley 2000) and personal observation. There were few observations of these 42 species on the 265 field days. Six were seen on 6 to 13 days, the remaining 36 on 5 or fewer days. Statistical Analysis Counts of the first block (1983-84, Table 1) were 50 min in duration; all later counts were 30 min in duration. For analyses involving the mean number of species per count, I adjusted the value for each count in the first block as an estimate of the number of species expected if the count had been for 30 min. As a basis for this adjustment, I made an additional 45 paired counts of 50 min duration between 7 November 2002 and 1 May 2003, recording the time a species was first seen. I used data from these 45 paired counts only for 7 THE EFFECTS OF FIRE ON A COASTAL SAGE SCRUB BIRD COMMUNITY the adjustment, not in any other analyses. In F Canyon, the mean number of species seen was 12.4 ± 0.4 (± SE) in 50 min, 1.8 ± 0.2 after the first 30 min. In Box Canyon, the mean number of species seen was 8.9 ± 0.4 in 50 min, 1.4 ± 0.2 after the first 30 min. I adjusted counts for F Canyon and Box Canyon separately because the number of species seen on these 45 paired counts in 50 min was higher in F Canyon (paired-sample t = 7.77, df = 44, P < 0.001). The number of species seen after 30 min was positively correlated with the total number of species seen in the 50 min count (Box Canyon, r = 0.44, df = 43, P = 0.001; F Canyon, r = 0.26, df = 43, P = 0.04). Therefore, I figured the adjustment for each count in the 1983-84 block by using ordinary least-squares linear regression to predict the number of species that would be seen after 30 min and subtracting that number from the total number seen in the count. For F Canyon, the adjustment function was adjusted number = total species seen - (-0.01 + 0.15 total species seen); for Box Canyon it was adjusted number = total species seen - (-0.41 + 0.20 total species seen). The phrase “adjusted number of species” indicates an analysis using the adjustment of the 1983-84 data. Data from the other five blocks were not adjusted because those counts were 30 min in duration. I analyzed differences in adjusted number of species among count blocks and sites with split-plot ANOVA. Each count (whole unit) was split into the two sites (F Canyon and Box Canyon) as subunits. I tested the whole-unit effect of count block by using unexplained variation among counts as an error term, the subunit effects of site and site-by-count-block interaction by using unexplained variation within counts as an error term. I used Tukey’s honestly significant difference (HSD) test for multiple comparisons among count blocks and tested for a difference between the sites within each count block by separate paired-sample t tests. Because analyses within a count are paired by date, I used the actual number of species per count rather than the adjusted number. I analyzed winter and summer data separately. Because the data were not normally distributed, I assessed correlation in the number of species observed between count blocks within canyons and seasons by the nonparametric Spearman rank correlation. The concomi- tant loss of power was minimal because 95% of the Spearman values were significant at P < 0.05. I used the Fisher exact test to evaluate differences within canyons by season in the proportional use of sites in the count block before the 1989 fire (1986-88) and the count block after the fire (1989-91). Means are reported ± one standard error. For statistical analyses I used SAS software version 9.2 (SAS Institute, Inc., Cary, NC). RESULTS Bird Species Richness and Community Composition Over the 20-year study I made 3753 observations representing 97 species and 12,619 individuals (Table 2). During the 16 months of the first count block (Table 1), 79 species were seen. The remaining 18 species were ac- cumulated over the next 19 years. Ninety species were seen in F Canyon, 80 in Box Canyon. Seventy-three species (75%) were seen in both canyons. No commonly seen species was restricted to a single canyon. Of the 24 species seen in only one canyon, 12 of these were seen once, seven were 8 THE EFFECTS OF FIRE ON A COASTAL SAGE SCRUB BIRD COMMUNITY seen twice, three were seen three times, and two were seen five times. The California Gnatcatcher, listed as threatened by the U.S. Fish and Wildlife Service, was seen 19 times beginning in 1994. A small number of species dominate the counts. The six most common species (Anna’s Hummingbird, California Towhee, Western Scrub-Jay, Northern Mockingbird, House Finch, Spotted Towhee) account for 36% of the 3753 observations. Addition of the next five species (Wrentit, Mourn- ing Dove, California Quail, Cactus Wren, Bushtit) encompasses 50% of the observations. The 35 most commonly seen species (Table 2, in bold) account for 90% of the observations. Fifty-one species were seen on fewer than 5% of the 265 count days. Numbers of species recorded in the six blocks of counts were significantly different in summer (F 5 12 g = 18.3, P < 0.0001) and in winter (F 5 12 4 = 21.8, P < 0.0001). The first count block (1983-84) and the last count block (2002-03) were more similar to each other and had more adjusted species per count than the four count blocks in between (Figure 1). However, results of the Tukey HSD multiple-comparisons tests (Figure 1) indicate some over- lap. For example, the 1992-96 block is always grouped with either the first or the last block in both canyons and in both seasons. Therefore, although the count blocks were a significant source of variation in the number of adjusted species observed, there was no block that was unique. 14 1 12 ■ 10 ■ c 13 o JJ In CD u CD Q_ w c: CD tt) n 2 Days since 21 August 1981 Figure 1. Mean adjusted species per count in F Canyon and Box Canyon, Voorhis Ecologi- cal Reserve, Pomona, California, in summer (top panel) and winter (lower panel) through six blocks of counts (Table 1), 1983-2003. Vertical bars at the means represent one standard error. Lower-case letters indicate groupings by the Tukey HSD test within a season, not be- tween seasons. Results of paired-sample t tests of the difference between F Canyon and Box Canyon for each count block in each season: NS, not significant ( P > 0.05); *, P < 0.05; **, P < 0.001. The vertical line indicates the date of the fire (28 July 1989) that burned both sites and the surrounding reserve. 9 THE EFFECTS OF FIRE ON A COASTAL SAGE SCRUB BIRD COMMUNITY Site Differences During the summer (Figure 1), the adjusted mean number of species per count was significantly higher at F Canyon (10.7 ±0.25) than at Box Canyon (8.3 ± 0.23; F 1 i 2 g = 64.6, P < 0.0001). There was no significant interaction between the site and count blocks (F 5 129 = 1 .9, P = 0. 1). When count blocks are examined individually (Figure 1), the canyons were significantly different in all except the 1989-91 and the 2000-01 block. During the winter (Figure 1), the adjusted mean number of species per count was also significantly higher at F Canyon (9.9 ± 0.27) than at Box Canyon (6.7 ± 0.24; F : 124 = 120.9, P < 0.0001). At this season, there was a significant (F 5 12 4 = 16.8, P < 0.0001) interaction between site and count block but no significant dif- ference in the 1989-91 and 2002-03 blocks (Figure 1). More species were seen in F Canyon than in Box Canyon on 214 (81%) of the 265 paired counts. The same number of species was seen on 21 counts (8%), and more species were seen in Box Canyon on the remaining 30 counts (11%). The ratio of 214:30 is significantly different from the 1:1 expected if the higher number of species seen was distributed randomly (sign test, P < 0.001). In comparisons within seasons, F Canyon had more species observed than Box Canyon in all count blocks except in winter of the 2002-03 block (Table 1). Of the 35 most common species, 16 were seen significantly more often in F Canyon than in Box Canyon (Table 2). Five of these (Spotted Towhee, Wrentit, Mourning Dove, Cactus Wren, California Thrasher) preferred F Canyon in both summer and winter. Nine species (Northern Mockingbird, California Quail, Bewick’s Wren, Red-tailed Hawk, Rufous-crowned Spar- row, White-crowned Sparrow, Nuttall’s Woodpecker, Ruby-crowned Kinglet, Blue-gray Gnatcatcher) preferred F Canyon only in the winter. The House Finch and Phainopepla preferred F Canyon only in the summer. Three species were seen significantly more often in Box Canyon (Table 2). The Bushtit preferred Box Canyon in the summer, whereas the Northern Flicker and Yellow-rumped Warbler preferred Box Canyon only in the winter. The remaining 16 species had no preference in either season (Table 2). Within canyons and seasons, there was significant positive correlation between count blocks in the ranks of the 35 most common species in 57 of the 60 comparisons (Table 3). In other words, the species observed most often in one block were the species observed most often in the other blocks. The three exceptions to this all involved the count block immediately after the 1989 fire (1989-91) in F Canyon in the winter (Table 3). In summary, the general results of these correlations are that the same species occurred at similar relative frequencies over the entire 20-year study. The species in Table 2 are listed by decreasing number of total observations, and this order is generally representative of both sites, both seasons, and the count blocks. Effects of the 1989 Wildfire The mean adjusted number of species per count immediately after the fire (1989-91) at both sites and in both seasons was similar to that both before the fire and much later after the fire (Figure 1). The 1989-91 block was the only block in which there was no significant difference between F Canyon 10 THE EFFECTS OF FIRE ON A COASTAL SAGE SCRUB BIRD COMMUNITY Table 3 Spearman Rank Correlations of the 35 Most Common Species 0 at the Voorhis Ecological Reserve, Pomona, California, 1983-2003 b 1983-84 1986-88 1989-91 1992-96 2000-01 2002-03 F Canyon 1983-84 0.77 0.26* 0.74 0.81 0.72 1986-88 0.90 0.23* 0.65 0.83 0.79 1989-91 0.45 0.34 0.34 0.13* 0.36 1992-96 0.83 0.79 0.56 0.78 0.83 2000-01 0.73 0.77 0.45 0.72 0.79 2002-03 0.76 0.77 0.37 0.77 0.80 Box Canyon 1983-84 0.46 0.62 0.53 0.35 0.40 1986-88 0.61 0.67 0.59 0.61 0.58 1989-91 0.78 0.65 0.36 0.47 0.50 1992-96 0.67 0.73 0.74 0.74 0.71 2000-01 0.46 0.64 0.60 0.67 0.79 2002-03 0.52 0.58 0.64 0.69 0.87 “See Table 2. b In the matrix for each canyon, the lower left triangle contains summer correlations, the upper right triangle winter correlations. All correlations except the three marked with an asterisk (*) are significant ( P < 0.05, n = 35). and Box Canyon in the number of adjusted species per count in both sum- mer and winter, although there was no significant difference in the summer of 2001-01 and the winter of 2002-03 as well (Figure 1). Examination of both seasons in Figure 1 suggests a tendency for a decrease in adjusted species per count in F Canyon and an increase in Box Canyon immediately after the fire, but statistical support for the trend is weak. In addition to the only three nonsignificant rank correlations involving the count block immediately after the fire, correlations involving the 1989-91 count block are significantly weaker than correlations involving other count blocks (Wilcoxon rank-sum test, exact P < 0.05). This indicates that of all the count blocks, the post-fire block was the most different. In F Canyon in the summer, the Wrentit and Cactus Wren (Table 4) were seen significantly more often before the fire (1986-88) than after the fire (1989-91), while the American Kestrel was seen significantly more often after the fire. In F Canyon in the winter, the Western Scrub-Jay, Spotted Towhee, Wrentit, California Quail, and Cactus Wren were seen significantly more often before the fire, while the House Finch and Lesser Goldfinch were seen significantly more often after the fire (Table 4). In Box Canyon in the summer, the only significant difference was the Rufous-crowned Sparrow, which was seen more often after the fire. In Box Canyon in the winter the Yellow-rumped Warbler and White-crowned Sparrow (both winter residents) were seen significantly more often after the fire (Table 4). Six species were only seen in the 1989-91 count block: the Rock Wren, White-throated Swift, Lark Sparrow, Golden Eagle, Cassin’s Vireo, and Prairie Falcon. 11 THE EFFECTS OF FIRE ON A COASTAL SAGE SCRUB BIRD COMMUNITY Table 4 Percentages of Counts in Which the 35 Most Common Species at the Voorhis Ecological Reserve Were Observed before (1986-88) and after (1989-91) the 1989 Fire by Site and Season 0 F Canyon Box Canyon Summer (n = 23) Winter (n = 18) Summer (n = 19) Winter (n = 20) 86-88 89-91 86-88 89-91 86-88 89-91 86-88 89-91 Anna’s Hummingbird 65 68 94 80 78 84 83 90 California Towhee 83 89 78 30 91 74 28 60 Western Scrub-Jay 91 42 89* 20* 70 68 67 55 N. Mockingbird 43 63 39 10 78 68 33 10 House Finch 61 74 17* 75* 48 63 39 35 Spotted Towhee 61 32 50* 5* 17 42 0 10 Wrentit 96* 5* 72* 0* 22 11 17 0 California Quail 39 5 33* 0* 0 11 0 0 Mourning Dove 35 74 6 20 30 16 11 15 Bushtit 9 11 6 5 22 26 22 5 California Thrasher 35 16 22 0 17 11 0 0 Cactus Wren 48* 0* 39* 0* 4 0 0 0 Lesser Goldfinch 13 42 0* 60* 13 21 0 10 Bewick’s Wren 22 16 39 20 9 37 6 35 Northern Flicker 9 11 22 20 13 11 56 75 Rufous-crn. Sparrow 4 16 0 5 0* 37* 0 0 Red-tailed Hawk 4 21 22 35 13 0 28 45 Yellow-rumped Warbler 0 11 22 45 0 5 6* 65* Phainopepla 35 16 0 0 22 26 0 0 White-crn. Sparrow 9 11 22 35 0 0 0* 45* Nuttall’s Woodpecker 13 32 11 0 9 5 0 5 House Wren 13 0 6 0 4 5 0 0 Black Phoebe 0 21 0 30 0 16 6 0 Ash-throated Flycatcher 13 11 0 0 17 5 0 0 Song Sparrow 13 0 11 0 4 0 0 5 Black-headed Grosbeak 17 0 0 0 13 0 0 0 Ruby-crn. Kinglet 0 11 11 10 0 0 0 5 American Robin 4 0 11 10 0 0 6 0 Brown-headed Cowbird 4 16 0 0 0 5 0 0 American Kestrel 0* 32* 0 15 4 11 6 15 Cooper’s Hawk 0 5 0 15 0 0 0 0 California Gnatcatcher 0 0 0 0 0 0 0 0 Blue-gray Gnatcatcher 4 0 39 5 0 0 0 0 Common Raven 0 5 0 25 0 5 6 40 Western Kingbird 0 16 0 0 0 11 0 0 “Asterisks (*) indicate a significant difference between 1986-88 and 1989-91 within the site and season (Fisher exact test, P < 0.05). 12 THE EFFECTS OF FIRE ON A COASTAL SAGE SCRUB BIRD COMMUNITY DISCUSSION Bird Species Richness and Community Composition The Voorhis Ecological Reserve contains some species typical of urban habitats and others dependent on the reserve’s shrub habitat, but its avifauna is dominated by several common species that are generally found in urban habitats. While they use fragments of sage scrub, they are not shrub-obligate species (Soule et al. 1988, Crooks et al. 2004). The six most common (Table 2) species (Anna’s Hummingbird, California Towhee, Western Scrub-Jay, Northern Mockingbird, House Finch, Spotted Towhee) fit this description. But some of the other common species, such as the Wrentit, California Quail, Cactus Wren, and California Thrasher, tend to be scrub specialists (Soule et al. 1988, Bolger et al. 1991, Crooks et al. 2001). Site Differences One interesting result of this study was the consistent difference in the number of adjusted species per count between the sites, F Canyon averag- ing two to three more species than Box Canyon. The amount of difference fluctuated over the count blocks and by season, but F Canyon had a higher adjusted mean number of species per count in all blocks, and the fact that F Canyon had more species in 81% of the counts is notable. I selected the sites after the 1981 fire because of their similar vegetation, physical similarity, close proximity, connection with continuous coastal sage scrub within the reserve, and because Box Canyon burned in the 1981 fire while F Canyon did not (Moriarty et al. 1985, Stanton 1986). Moriarty et al. (1985) and Stanton (1986) compared these two sites to assess the effect of the 1981 fire on avian species richness in coastal sage scrub. Both studies found significantly fewer species seen in Box Canyon than F Canyon and concluded that fire reduced avian species richness. The conclusions of Moriarty et al. (1985) and Stanton (1986) are sub- stantially mitigated by the long-term results I report here. The consistently higher number of bird species seen in F Canyon suggests that the sites are not equivalent and that F Canyon was not a “control” for the burned Box Canyon. The difference in the number of species per count attributed to fire effects was four by Moriarty et al. (1985), and 5.9 by Stanton (1986), but these differences may have been due, at least in part, to inherent differences between the sites. The long-term results (a difference of about three species) do not completely contradict the earlier (Moriarty et al. 1985, Stanton 1986) conclusions, but they certainly confound the conclusions to the point where an inference on the effect of fire on avian species richness in coastal sage scrub cannot be drawn from those studies. Furthermore, comparison of the count blocks immediately before and after the 1989 fire suggested the response to fire was different at the two sites and may also have varied by season. This finding demonstrates the difficulty of selecting appropriate comparison sites for study of the effect of uncontrolled fires (Diamond 1986). I attempted to match a burned and unburned area with respect to size, aspect, topography, proximity, and vegetation in order to control for any effect except the fire. Yet the long-term data strongly suggest the sites differ. If multiple burned and 13 THE EFFECTS OF FIRE ON A COASTAL SAGE SCRUB BIRD COMMUNITY unburned canyons were available after the 1981 fire, such replication may have detected the effect of site, although such replication would have been approximate and may not have demonstrated the effect (Hargrove and Picker- ing 1992, Underwood 2009). Adequate replication of sites in field situations requires prior knowledge of perturbations and detailed analysis (Silva-Lugo and Tanner 2010), and these conditions could not be met at my study sites. Given the two canyon’s similarity and proximity, what causes the effect? The data do not provide an answer to that question, but possibilities may be suggested. Perhaps the vegetation structure and/or composition differ in some way important to the birds that is not obvious to human observers. Perhaps predators such as Cooper’s Hawk spend more time in the Box Canyon area than F Canyon, resulting in smaller birds spending more time in F Canyon. The composition of the avian community at the two sites is similar. That is, the same species were observed in the two canyons at the same approximate relative frequency but more often in F Canyon than in Box Canyon. This suggests that the factors in which the sites differ may be subtle. It is possible that it was easier to detect species in F Canyon than in Box Canyon. There was nothing in the field experience that suggested such a bias, but it cannot be conclusively ruled out. The increase in the number of species seen at both sites and in both sea- sons in 2002-03 (Fig. 1) is consistent with these possibilities. However, such an interpretation is predicated on the 2002-03 data representing a biological effect and not random variation, and distinguishing between these possibili- ties would require more long-term data. As seen above, conclusions based on short-term (1-2 years) data without adequate control are subject to error. Effects of Wildfire Having discussed the importance of site effects and lack of replication in comparisons before and after the fires, I stress that this discussion of the effect of the 1989 fire is descriptive, anecdotal rather than inferential. In contrast to the conclusions of Moriarty et al. (1985) and Stanton (1986) after the 1981 fire, the number of species did not decline after the 1989 fire. There was an increase in species after the fire in Box Canyon, paralleling the findings of Mendelsohn et al. (2008) after San Diego County’s Cedar Fire of 2003, but only during the winter. The fire did appear to have some effect on the composition of the species. Not surprisingly, species associated with dense vegetation (e.g., Wrentit, California Quail, Spotted Towhee) were seen less often after the fire (Knick and Rotenberry 1995, Mendelsohn et al. 2008). The Cactus Wren was seen less often after the fire, presumably because of the destruc- tion of the Opuntia cactus with which the bird associates. Some species were seen more often after the fire, apparently taking advantage of the more open habitat. These included ground-feeding species such as the House Finch, Lesser Goldfinch, Rufous-crowned Sparrow, Yellow-rumped Warbler, and White-crowned Sparrow. The American Kestrel was also more frequently seen after the fire, perhaps because prey was easier to detect. A few species such as the Golden Eagle, Prairie Falcon, and Lark Sparrow, perhaps responding to open habitat, were seen only within the first two years following the fire. Some of the most common species (e.g., Anna’s 14 THE EFFECTS OF FIRE ON A COASTAL SAGE SCRUB BIRD COMMUNITY Hummingbird, California Towhee, Northern Mockingbird, Mourning Dove) did not show any significant change in their use of the sites after the fire. These four species are common in both shrub and urban habitat (Soule et al. 1988, Crooks et al. 2004), and this flexibility may extend to using burned and unburned areas. These apparent effects of the fire did not last long. By the start of the fourth count block, approximately three years post-fire, most of the effects described above were no longer apparent. The 1989 fire did not cause any long-lasting change to the bird community. The effects of fire on bird com- munities vary greatly, with the severity of the fire being critical to species’ responses (Smucker et al. 2005). In addition to severity, I suggest that the proximity of habitat suitable for common species also contributes to a fire’s effect on the community. In this study, adjacent undeveloped areas of coastal sage scrub habitat in the San Jose Hills did not burn. Many of the common species are also found in nearby urban areas (Mills et al.1989, Crooks et al. 2004). Therefore, there were numerous potential refugia as well as sources of birds to repopulate the reserve. The appearance of the California Gnatcatcher in 1994 (five years post-fire) suggests recovery of the scrub (Atwood and Bon- trager 2001). Results of this study largely parallel those of Mendelsohn et al. (2008) in low-elevation coastal sage scrub. But because of birds’ inconsistent response to fire (Smucker et al. 2005) and lack of replication in my study, it would not be appropriate to consider the effects I discuss here as general features of the response of coastal sage scrub birds to fire. Such replication may not be feasible, given the patchy distribution, dynamics, and disturbance of coastal sage scrub (Crooks et al. 2001, Cox and Allen 2008). Long-term studies contribute to our understanding of avian community dynamics (Holmes et al. 1986), and suggest questions for future work. Stud- ies of the breeding birds of the Voorhis Ecological Reserve could emphasize demographics (Brawn and Robinson 1996), particularly of the threatened California Gnatcatcher. Study of migrants or vagrants could address these species’ use of the reserve. Such research may help clarify bases of the dif- ferences between F Canyon and Box Canyon, as well as provide a baseline for comparison following any future fires. ACKNOWLEDGMENTS I am grateful to Patricia Farris, Cynthia Shannon, Arthur Davenport, Michelle Shaughnessy, Mary Haus, Kristen Schroeder, and Gail Drus for field support. Holly Vuong and anonymous reviewers improved the manuscript. LITERATURE CITED Atwood, J. L., and Bontrager, D. R. 2001. California Gnatcatcher (Polioptila cali- fornica), in The birds of North America (A. Poole and F. Gill, eds.), no. 574. Birds of N. Am., Inc., Philadelphia. Bolger, D. T., Alberts, A. C., and Soule, M. E. 1991. Occurrence patterns of bird species in habitat fragments: Sampling, extinction, and nested species subsets. Am. Nat. 137:155-156. Brawn, J. D., and Robinson, S. K. 1996. Source-sink population dynamics may complicate the interpretation of long-term census data. Ecology 77:3-12. Chase, M. K., Kristan, W. B. Ill, Lynam, A. J., Price, M. V., and Rotenberry, J. T. 15 THE EFFECTS OF FIRE ON A COASTAL SAGE SCRUB BIRD COMMUNITY 2000. Single species as indicators of species richness and composition in Cali- fornia coastal sage scrub birds and small mammals. Conserv. Biol. 14:474-487. Clark, C. 1990. Vascular plants of the undeveloped areas of California State Poly- technic University, Pomona. Crossosoma 4:1-7. Cox, R. D., and Allen, E. B. 2008. Stability of exotic annual grasses following restora- tion efforts in southern California coastal sage scrub. J. Appl. Ecol. 45:495-504. Crooks, K. R., Suarez, A. V., Bolger, D. T., and Soule, M. E.. 2001. Extinction and colonization of birds on habitat islands. Conserv. Biol. 15:159-172. Crooks, K., Suarez, A. V., and Bolger, D. T., 2004. Avian assemblages along a gradient of urbanization in a highly fragmented landscape. Biol. Conserv. 115:451-462. Diamond, J. 1986. Overview: Laboratory experiments, field experiments, and natural experiments, in Community Ecology (J. Diamond and T. J. Case, eds.), pp. 3-22. Harper & Row, New York. Hargrove, W. W., and J. Pickering. 1992. Pseudoreplication: a sine qua non for regional ecology. Landscape Ecology 6:251-258. Herrera, C. M. 1998. Long-term dynamics of Mediterranean frugivorous birds and fleshy fruits: A 12-year study. Ecology 68:511-538. Holmes, R. T., Sherry, T. W. and Sturges, F. W. 1986. Bird community dynamics in a temperate deciduous forest: Long-term trends at Hubbard Brook. Ecol. Monogr. 56:201-220. Knick, S. T., and Rotenberry, J. T. 1995. Landscape characteristics of fragmented shrubsteppe habitats and breeding passerine birds. Conserv. Biol. 9:1059-1071. McCaull, J. 1994. The Natural Community Conservation Planning Program and the coastal sage scrub ecosystem of southern California, in Environmental Policy and Biodiversity (R. E. Grumbine, ed.), pp. 281-292. Island Press, Washington, D.C. Mendelsohn, M. B., Brehme, C. S., Rochester, C. J., Stokes, D. C., Hathaway, S. A., and Fisher, R. N. 2008. Responses in bird communities to wildland fires in southern California. Fire Ecol. 4:63-82. Mills, G. S., Dunning, J. B., Jr., and Bates, J. M. 1989. Effects of urbanization on breeding bird community structure in southwestern desert habitats. Condor 91:416-428. Minnich, R. A. 1983. Fire mosaics in southern California and northern Baja Califor- nia. Science 219:1287-1294. Moriarty, D. J., Farris, R. E., Noda, D. K., and Stanton, P. A. 1985. Effects of fire on a coastal sage scrub bird community. Southwest. Nat. 30:452-453. Sibley, D. A. 2000. The Sibley Guide to Birds. Knopf, NY. Silva-Lugo, J. L., and G. W. Tanner. 2010. Testing control sites for fire ecology research. J. Torrey Bot. Soc. 137:263-276. Smucker, K. M., Hutto, R. L., and Steele, B. M. 2005. Changes in bird abun- dance after wildfire: Importance of fire severity and time since fire. Ecol. Appl. 15:1535-1549. Soule, M. E., Bolger, D. T., Alberts, A. C., Wright, J., Sorice, M., and Hill, S. 1988. Reconstructed dynamics of rapid extinctions of chaparral-requiring birds in urban habitat islands. Conserv. Biol. 2:75-92. Stanton, P. A. 1986. Comparison of avian community dynamics of burned and unburned coastal sage scrub. Condor 88:285-289. Underwood, A. J. 2009. Components of design in ecological field experiments. Ann. Zool. Fennici 46:93-111. Westman, W. E. 1981. Factors influencing the distribution of species of Californian coastal sage scrub. Ecology 62:439-455. Wirtz, W. O. 1982. Postfire community structure of birds and rodents in southern California chaparral. U.S. Dept. Agric. Gen. Tech. Rep. PSW-58. Accepted 22 Ma\ j 2012 16 2011 NEVADA BIRD RECORDS COMMITTEE REPORT MARTIN MEYERS, Secretary, Nevada Bird Records Committee, c/o Great Basin Bird Observatory, 1755 E. Plumb Lane #256, Reno, Nevada 89502; nbrc@gbbo.org ABSTRACT: This report covers the 115 records reviewed by the Nevada Bird Records Committee in 2011, of which 110 were endorsed. These 115 records cover sightings from 24 October 1971 through 26 November 2011, 37 in 2011, 20 antedat- ing the formation of the committee in 1994. Three species are added to the Nevada list (and to the committee’s review list): the Iceland Gull ( Larus glaucoides), Winter Wren (Troglodytes hiemalis ), and Sedge Wren (Cistothorus platensis ). The Northern Beardless-Tyrannulet (Camptostoma imberbe ) is removed from the Nevada list (and review list) because evidence to substantiate the record is insufficient. Ten species are re- moved from the review list because of the number of records or regularity of occurrence. The Nevada state list now stands at 488 species, of which 167 are on the review list. The Nevada Bird Records Committee (NBRC) began 2011 with 66 records pending review. During the year, an additional 114 reports were received and added to the database. The committee completed reviews of 115 records during the year, so at the end of 201 1 , the queue contained 65 pending records. Since the founding of the NBRC in 1994, a total of 789 records have been reviewed, of which 726 have been endorsed. At its founding in 1994, the committee decided not to review any sightings prior to that year but reversed that decision several years later. Fortunately, founding secretary James Cressman and his wife Marian continued to ac- cumulate documentation for “pre-committee” records. That accumulated documentation has been provided to the current committee, and one of our long-term goals is to organize and review as many of those records as possible. Of the 114 “new” records received and added to the database in 2011, 15 precede the committee, dating back to 24 October 1971. Of the 1 15 records reviewed by the committee in 201 1, 37 were of birds found during 2011; 20 preceded the committee’s founding. Reviewed re- cords thus extended from 24 October 1971 to 26 November 2011. Of the 110 endorsed records, 84 were supported by photographs. Five records, none supported by photographs, failed to gain endorsement. The endorse- ment rate for birds photographed and not photographed implies that records with photographs are more likely to be endorsed than are those with only written documentation, but records with only detailed written documentation can achieve committee endorsement as well. The NBRC has six voting members and a nonvoting secretary. In 2011, the committee welcomed new member Jeanne Tinsman, who joined con- tinuing members John Klicka, Tim Lenz, Carl Lundblad, Greg Scyphers, and Dennis Serdehely. The position of secretary continues to be held by Martin Meyers. At the close of 2011, Greg Scyphers rotated off the committee (term-limited) and was replaced by Will Richardson for 2012. The NBRC’s website at http:/gbbo.org/nbrc contains a statement of pur- pose, links to a downloadable submission form, the committee’s bylaws, the Nevada state checklist maintained by the NBRC, the review list, and answers to frequently asked questions. There is a link to a list of every submission to Western Birds 44:17-32, 2013 17 2011 NEVADA BIRD RECORDS COMMITTEE REPORT the NBRC, with the endorsement status and, if available, a photograph. New in 2011, all previous NBRC reports are available as PDF files. The reports up through 2007 (1994-96, 1997, 1998, 1999, 2000, 2004, 2005, and 2007) appeared in Great Basin Birds , published by the Great Basin Bird Observatory. REVISIONS TO THE NEVADA STATE LIST IN 2011 During the period covered by this report, three new species were added to the Nevada list, and one species was removed. Added were the Iceland Gull (. Larus glaucoides), Winter Wren (Troglodytes hiemalis, on the basis of a good description antedating the split of the Pacific [ T. pacificus] and Winter wrens; Chesser et al. 2010), and Sedge Wren ( Cistothorus platensis). Removed: the Northern Beardless-Tyrannulet ( Camptostoma imberbe) was on the Nevada list on the basis of a bird banded at Pahranagat National Wildlife Refuge (NWR), 28 April 1973. Other than a band number and a one-line comment containing no details, there is no documentation for this record. It was never submitted to the committee (it antedates the formation of the committee by two decades.) After the NBRC exerted considerable effort to find additional information without success, a motion to delete the spe- cies was discussed and received unanimous agreement in September 2011. These changes brought the count of species recorded in Nevada to 488 at the end of 2011. REVISIONS TO THE NEVADA REVIEW LIST IN 2011 The only species the NBRC added during 2011 to the list of species it reviews were the three species new to Nevada, discussed above. At the committee’s biennial meeting in September, it removed 10 species from the review list. Justification for removal from the review list included the total number of endorsed records, regularity of sightings, evidence of occurrence from similar nearby habitats in adjoining states, plus the personal experiences, knowledge, and judgment of the members. These species were removed: Eurasian Wigeon ( Anas penelope), Thayer’s Gull (Larus thayeri), Acorn Woodpecker (Melanerpes formiciuorus), Chestnut-collared Longspur (Calcarius ornatus), Magnolia Warbler ( Setophaga magnolia), Chestnut-sided Warbler (S. pensyluanica), Black-throated Blue Warbler (S. caerulescens), Palm Warbler (S. palmarum), Lark Bunting ( Calamospiza melanocorys), and Lawrence’s Goldfinch (Spinus lawrencei). There are currently 167 species on the Nevada review list, of which seven are exempt from review in some limited geographic area. Six of those seven exemptions are for very localized breeding populations. The seventh, the Broad-winged Hawk, is exempt in the Goshute Mountains, where multiple birds are ob- served annually in migration (Smith et.al. 2008). SPECIES ACCOUNTS For each species, the format is English name, scientific name, (total number of endorsed records of the species, number of records endorsed in this year’s report). Two asterisks after the total of records signify that the 18 2011 NEVADA BIRD RECORDS COMMITTEE REPORT number of records refers to a restricted review period, usually signifying that the species is no longer on the review list, was placed on the review list as a result of a perceived drop in population, or is exempt from review in some locations. Note that the total number of records for a species is not necessarily the total number of individual birds reported. After the heading for the species is each record of that species reviewed in 2011, in this format: NBRC record number, name of each submitter, date or range of documented dates, and location (county in parentheses.) If the record involved multiple individual birds, the number follows the county. “(P),” “(V),” or “(A)” following a submitter’s name indicates that he or she provided a photo, video, or audio recording, respectively. Discussion of a particular sighting follows that sighting’s data, whereas that related to the species in general is at the end of the species account. If there are multiple records of the species, the records are ordered by date of first sighting. Certain records in these accounts are noted as “establishing” records. The NBRC uses the term “establishing” record for the first NBRC-endorsed record for any species on the checklist of which there is not yet at least one endorsed record. Early in its history, the NBRC adopted an existing checklist (Titus 1996) based on numerous sources that constituted the most reliable information available at the time. All but 30 of the review species on the state list now have at least one endorsed record. The committee is pursuing documentation of these 30, and we have had some significant success in this quest. However, it will probably be a few years before we conclude that we have exhausted all possibilities. At that time, we will reevaluate the status of those species still without an endorsed record. EURASIAN WIGEON Anas penelope (15**, 2). 2011-053, Paul Lehman, 19 November 1990, Key Pittman Wildlife Management Area (WMA) (Lincoln). 2010-067, Martin Meyers, 5 November 2010, Lemmon Valley (Washoe). Species removed from the review list September 2011. GREEN-WINGED TEAL (EURASIAN) Anas crecca crecca (4, 1). 2011-028, Martin Meyers (P), 14 December 2006, Damonte Ranch Wetlands, Reno (Washoe). The NBRC’s bylaws permit, but do not require, reviews of rare subspecies of spe- cies otherwise common in Nevada, but only if the subspecies has previously been considered a full species by the AOU or is currently considered a full species by other similar international organizations. BLACK SCOTER Melanitta americana (6, 2). 2010-077, Dennis Ghiglieri (P), Rose Strickland, Dennis Serdehely, Greg Scyphers (P), 14-15 November 2010, west shore of Pyramid Lake (Washoe). 2010-078, Dennis Ghiglieri (P), Rose Strickland, 14 November 2010, south end of Pyramid Lake (Washoe). This is the first time the NBRC has received documentation of two Black Scoters nearby on the same day LONG-TAILED DUCK Clangula hyemalis (4**, 1). 2011-006, Alan de Queiroz (P), Fred Welden (P, Figure 1A), Nancy Devon (P), Martin Meyers (P, Figure IB), Fred Petersen (P), 3 February- 14 May 2011, Virginia Lake, Reno (Washoe). The plumage changed during the bird’s stay. The bird alternated between feeding on the Truckee River near downtown Reno and sleeping on a pond in a city park farther upstream, occasionally it visited a city park several miles from either of those locations. This species has been on the review list only since 2007, when it was added because of a dearth of recent records. 19 2011 NEVADA BIRD RECORDS COMMITTEE REPORT RED-THROATED LOON Gavia stellata (3, 1). 2011-003, Andrew Lee (P), Den- nis Serdehely, 1-17 January 2011, Las Vegas Bay, Lake Mead National Recreation Area (NRA) (Clark). LITTLE BLUE HERON Egretta caerulea (2, 1). 2007-100, NOT ENDORSED. 24 July 1993, Henderson Bird Viewing Preserve (Clark). The brief written description lacked some critical identification criteria. On its first round, the record received four votes to endorse and two not to endorse. On the second round, the record received no votes for endorsement. 2011-021, Tim Almond (P), Andrew Lee (P), 21 April 2011, Clark County Wet- lands Park (Clark). TRICOLORED HERON Egretta tricolor [1, 1). 2011-064, Vincent Mowbray, Mar- ian Cressman, Robert Rucker, 12-13 September 1992, Pahranagat NWR (Lincoln). This record established the Tricolored Heron on the Nevada state list. GLOSSY IBIS Plegadis falcinellus (3, 2). 2010-013, Carl Lundblad (P), 25 April 2010, Clark County Wetlands Park (Clark). 2010- 023, Greg Scyphers (P), Jon Dunn (P), Carl Lundblad (P), 21-22 May 2010, Dyer (Esmeralda). These two records followed on the heels of Nevada’s first state record, 24 May 2009 (Meyers 2010). WHITE-TAILED KITE Elanus leucurus (10**, 3). 2011-067, Vincent Mowbray, Mark Kasprzyk (P), 18 September 1982, Mormon Farm. Las Vegas (Clark). 2011- 020, Dennis Serdehely, 15 April 2011, private property at Warm Springs Natural Area (Clark). 2011-019, Scott Krammer (P), 20 April 2011, Las Vegas (Clark). A very small population of the White-tailed Kite breeds in the Pahranagat Valley, where the species is exempt from committee review. COMMON BLACK-HAWK Buteogallus anthracinus (8, 4). 2010-063, Richard Saval (P), 2 April 2004, Corn Creek (Clark). 2010- 030, Carl Lundblad (P), Martin Meyers (P), Greg Scyphers (P), Tim Almond (P, Figure 2), 29-30 May 2010, Meadow Valley Wash (Lincoln), 2 birds, observed (and photographed) engaging in talon-grappling and other displays. The species breeds in Utah less than 8 km from the Nevada border and approximately 48 km from Meadow Valley Wash (R. Fridell pers. comm.). It is probably only a matter of time before breeding of the Common Black-Hawk is confirmed in Nevada. 2011- 018, Aaron Ambos (P), 11 April 2011, Meadow Valley Wash (Lincoln). One member voted not to endorse this record, stating that he thought the bird was probably correctly identified but that some of the documentation was ambiguous and that the photograph was not completely convincing on its own. 2011-061, Greg Scyphers (P), 15 June 2011, Corn Creek (Clark). HARRIS'S HAWK Parabuteo unicinctus (5, 1). 2011-016, David Henderson, 6 April 2011, Pahranagat NWR (Lincoln). An additional four records of Harris’s Hawk, one of an apparent family group, are awaiting committee review. The question of origin is the main issue with occurrences of this species, used frequently for falconry. ZONE-TAILED HAWK Buteo albonotatus (13, 4). 2010-040, Eric Hough (P), Greg Scyphers (P), Diane Wong, 16 May-30 August 2010, Pahranagat NWR (Lin- coln), 2 birds, at least one of which remained through the summer. Hough described the birds defending territory near a nest. He observed the birds during multiple surveys between 16 May and 14 June 2010, but they did not nest successfully. 2011-031, Dennis Serdehely, 29 May 2011, Pahranagat NWR (Lincoln). 2011-065, David Henderson, 17 August 2011, Pahranagat NWR (Lincoln). One member voted not to endorse this record, stating that the identification was almost certainly correct but that other dark buteos were not adequately eliminated. 20 2011 NEVADA BIRD RECORDS COMMITTEE REPORT 2011-066, David Henderson, 17 August 2011, Pahranagat NWR (Lincoln). The NBRC endorsed four records of the Zone-tailed Hawk in 2011, representing a mini- mum of one adult and one immature bird, possibly as many as four different birds. As with the Common Black-Hawk, it is probably just a matter of time before breeding of the Zone-tailed Hawk is confirmed in Nevada. BLACK RAIL Laterallus jamaicensls (0, 0). 2008-081, NOT ENDORSED. 5 May 2008, Ash Meadows NWR (Nye). Carl Lundblad submitted documentation of a bird only heard in response to a tape during a survey. Several members commented that the documentation was convincing, but only one member found it sufficient for an establishing record. Lundblad is a member of the committee and was confident of the identification but recognized that the documentation would fall short, and even voted against endorsement. Nonetheless, he submitted the sighting to place the data into the permanent record, where it belongs, even without committee endorsement. UPLAND SANDPIPER Bartramia longicauda (2, 2). 2010-083, Vincent Mow- bray, 20 May 1985, Corn Creek (Clark). This record established the Upland Sand- piper’s occurrence in Nevada. The record received two votes against endorsement on the first round but only one on the second. A vote of 5-1 is sufficient for endorsement. 2010- 021, Carl Perretta, 13 May 2010, Hidden Valley (Clark). Interestingly, the voting on this record went the same way as that on 2010-083. STILT SANDPIPER Calidris himantopus (5**, 1). 2011-042, Andrew Lee (P), 8 August 2011, Clark County Wetlands Park (Clark), 2 birds. This species had been exempt from review in southern Nevada until 2009, when the exemption was re- moved because of a lack of recent reports. There remains just one endorsed record from northern Nevada. RED PHALAROPE Phalaropus fulicarius (7, 1). 2011-015, Greg Scyphers (P), 17 October 2010, Pahranagat NWR (Lincoln). HEERMANN’S GULL Larus heermanni (8, 2). 2010-054, Rob Lowry (P), 17 Octo- ber 2010, Washoe Lake (outside state park) (Washoe); bird in its second plumage cycle. 2011- 017, David Shen (P), 3 April 2011, Walker Lake (Mineral). This adult repre- sents the earliest spring record of Heermann’s Gull for Nevada; the other four spring records extend from 19 April through 3 June. The three fall records extend from 16 October through 28 October. Six of the eight records are from northern Nevada. THAYER’S GULL Lams thayeri (1 1**, 2). 2011-002, Andrew Lee (P), Rick Fridell (P), Martin Meyers (P), Robert Wilkin (P), David Syzdek (P), 1 January-5 February 2011, Hemenway Harbor, Lake Mead NRA (Clark), 2 birds. Thayer’s Gull had been exempt from review in northwest Nevada until 2007. The exemption was removed in 2007 after several years with very few reports. Since that time, there have been endorsed records in the northwest every year, with multiple endorsed records in all but one of those years. The committee voted to remove Thayer’s Gull from its review list in September 2011. This record of two individuals together is the only committee- endorsed record for southern Nevada. 2011-007, Alan de Queiroz (P), Martin Meyers (P), Dennis Serdehely (P), Rob Lowry (P), 8-9 February 2011, Virginia Lake, Reno (Washoe). ICELAND GULL Larus glaucoides (1,1). 2011-004, Fred Petersen (P), Martin Meyers (P, Figure 3), Greg Scyphers (P), Jon Dunn, Fred Welden (P), 24 January- 9 February 2011, Paradise Park, Reno (Washoe). This third-cycle “Kumlien’s” Gull was found by Fred Petersen on 24 January 2011 and observed and photographed by numerous birders from Nevada and California during its stay. Originally found at Paradise Park in Reno, it alternated between that location and Sparks Marina, often being seen at Paradise Park in the morning and Sparks Marina in the afternoon. The one vote against endorsing this record was more concerned with the taxonomic issues 21 2011 NEVADA BIRD RECORDS COMMITTEE REPORT Figure 1. This Long-tailed Duck (2011-006) was observed and photographed at three different locations around Reno from its initial discovery on 3 February 2011 to the last sighting on 14 May 2011. During that time, it went through substantial plumage changes, as seen in these photos, taken 7 February 2011 (A) and 14 May 2011 (B). Photos by Fred Welden (A) and Martin Meyers (B) than with the identification of this individual bird. As to the taxonomic issues regarding the Iceland (including “Kumlien’s”) and Thayer’s Gulls, the committee’s position is to follow the American Ornithologists’ Union’s current classification. LESSER BLACK-BACKED GULL Larus fuscus (10,3). 2010-061, Martin Meyers (P), 29 October 2010, Sparks Marina (Washoe). 2010- 065, Greg Scyphers (P), 1 November 2010, Carson City (Carson City). 2011- 022, Andrew Lee (P), 21 April 2011, Hemenway Harbor, Lake Mead NRA (Clark). Of the ten Nevada records, three have been judged to represent repeated occurrences of the same individual over three successive years. The first committee- 22 2011 NEVADA BIRD RECORDS COMMITTEE REPORT Figure 2. This Common Black-Hawk (2010-030) was with another of its species in Meadow Valley Wash, Lincoln Co., 29-30 May 2010 (photo taken 30 May). The two were observed engaging in talon-grappling and other displays. Photo by Tim Almond endorsed Lesser Black-backed Gull arrived in Nevada during the winter of 2007-2008 (Meyers 2010). Since that time, there have been at least two documented in the state every year except 2011, when only one was found. Through 2011, sightings in the northern portion of the state (6) slightly outnumbered those in the south (4). GLAUCOUS GULL Larus hyperboreus (7,1). 2011-005, Martin Meyers (P, Figure 4), Jon Dunn, 26-29 January 2011, Sparks Marina (Washoe). This beautiful large gull showed up to accompany Nevada’s first Iceland Gull (see 2011-004 above), and many of the observers who came to see the Iceland Gull were able to enjoy this bird as a bonus. While the Iceland Gull alternated between Paradise Park and Sparks Marina, the Glaucous Gull was observed at Sparks Marina only. BLACK-LEGGED KITTIWAKE Rissa tridactyla (3, 1). 2010-066, Greg Scyphers (P), Martin Meyers (P), 5 November 2010, The Willows, Pyramid Lake (Washoe). The two previous endorsed records for this species were from 1975 and 1995 (Meyers 2010). Several anecdotal reports precede the committee’s establishment, and we have some documentation for a few additional records from the late 1980s to the mid 1990s that we are reviewing. However, there is a gap of almost 15 years with no documented reports or even anecdotal reports preceding this 2010 record. We have documentation for two subsequent records for 2011 pending review. LEAST TERN Sternula antillarum (12, 1). 2011-024, Andrew Lee (P), 3 May 2011, Clark County Wetlands Park (Clark). All endorsed records fall between 29 April and 20 July, and all but two are for the southern part of the state. ARCTIC TERN Sterna paradisaea (5, 3). 2010-084, Paul Lehman, Shawneen Finnegan (P), 11 October 1990, Ash Meadows NWR (Nye). 2010-070, Greg Scyphers (P), 23 May 2010, Ash Springs (Lincoln). 23 2011 NEVADA BIRD RECORDS COMMITTEE REPORT Figure 3. This Kumlien’s Iceland Gull (2011-004) represents the first record of the Iceland Gull for Nevada. It was extensively documented (25 photos, three written descriptions, and additional opinions from experts were submitted.) The bird was found 24 January 2011 and remained until 9 February 2011, alternating between Paradise Park in Reno and Sparks Marina, Washoe Co. These photos were taken 25 January 2011. Photos by Martin Meyers 24 2011 NEVADA BIRD RECORDS COMMITTEE REPORT Figure 4. Many birders who visited the Reno/Sparks area in January 2011 to observe Nevada’s first Iceland Gull (2011-004) received a bonus with this Glaucous Gull (2011-005), as both were often seen sitting on the same set of floats at the Sparks Marina, Washoe Co. The Glaucous Gull was present 26-29 January; this photo was taken 28 January 2011. Photo by Martin Meyers 2010-042, Martin Meyers (P), 1 September 2010, The Willows, Pyramid Lake (Washoe). PARASITIC JAEGER Stercorarius parasiticus (5, 1). 2010-073, Greg Scyphers (P), 14 September 2010, Clark County Wetlands Park (Clark). GREAT GRAY OWL Strix nebulosa (0, 0). 2011-039, NOT ENDORSED. 6 July 2011, Mountain City (Elko). The documentation described a relatively brief sighting of a flying bird, viewed without binoculars at a distance reported as approximately 150 yards. Some details provided were consistent with the identification as this spe- cies, but the committee was unanimous in not endorsing what would have been an establishing record. The Great Gray Owl is on the Nevada list on the basis of a bird reported from Little Valley (Washoe County) in 1980 (Alcorn 1998). The committee has been unable to come up with any meaningful documentation for that record, but we have not yet given up hope of uncovering some. (See the note under “Revisions to the Nevada State List” above for discussion of related issues.) RED-HEADED WOODPECKER Melanerpes erythrocephalus (1, 0). 2010-037, NOT ENDORSED. 26 June 2010, Great Basin National Park (White Pine). This record was documented rather thoroughly, but the details were obtained piecemeal in response to repeated requests. During the first round, that bothered some members, two of whom voted not to endorse the record. In the second round, the dissenting opinions from the first round convinced all members that the documentation was not fully satisfactory for such a rare sighting, despite the likelihood that the identification was correct. ACORN WOODPECKER Melanerpes formicivorus (14**, 2). 2011-052, C. S. Lawson (P), 31 October 1971, Boulder City (Clark). 25 2011 NEVADA BIRD RECORDS COMMITTEE REPORT Figure 5. This Philadelphia Vireo (2010-059) at Floyd Lamb Park, Tule Springs, Clark Co., 19-23 October 2010 (photo taken 21 October), is the second to be satisfactorily documented in Nevada. This cooperative bird pleased birders who had been waiting for 15 years since the state’s first record (1995-22), at Indian Springs, Clark Co., 23 September 1995. Photo by Greg Scyphers Figure 6. The Purple Martin has been satisfactorily documented to the NBRC only eight times. This bird (2011-030) was at Dyer, Esmeralda Co., 28 May 2011. Photo by Greg Scyphers 26 2011 NEVADA BIRD RECORDS COMMITTEE REPORT Figure 7. This wonderful photo left no doubt in anyone’s mind about the identification of Nevada’s first recorded Sedge Wren (2011-101). The bird was found 15 October 2011 at a ranch in Dyer, Esmeralda, Co., and was gone by the next day. Photo by Greg Scyphers Figure 8. The fourth Yellow-throated Warbler (2011-001) to be endorsed by the NBRC was at Moapa National Wildlife Refuge, Clark Co., 2-16 January 2011 (photo taken 16 January). Photo by Greg Scyphers 27 2011 NEVADA BIRD RECORDS COMMITTEE REPORT 2011-032, Steve Ting (P), 4 June 2011, Verdi (Washoe). Species removed from the review list, September 2011. YELLOW-BELLIED SAPSUCKER Sphyrapicus varius (6, 3). 2010-074, Greg Scyphers (P), 26 October 2008, Miller’s Rest Stop (Esmeralda). 2010-075, Greg Scyphers (P), 7 November 2008, Corn Creek (Clark). 2010-089, Greg Scyphers (P), 6 December 2010, Corn Creek (Clark). LEAST FLYCATCHER Empidonax minimus (6, 1). 2011-037, Robert Williams (V), 18 June 2011, Ruby Valley (Elko). A beautiful close-up video of this bird, singing its ”che-bek” song repeatedly, supports this record. SULPHUR-BELLIED FLYCATCHER Myiodynastes luteiuentris (2,2). 2011-060, Vincent Mowbray, Kevin Wallace, 13 June 1987, Corn Creek (Clark). Another of the pre-committee records from the data provided by the Cressmans, this record established the Sulphur-bellied Flycatcher on the Nevada list. 2010- 041, Eric Hough, 27 July 2010, Pahranagat NWR (Lincoln). This record had trouble on its first two rounds, primarily because it would have represented an estab- lishing record. In both of those rounds, two members believed the identification was probably correct but that the documentation was insufficient for a first record. However, 2010-041 was brought up for an additional review after 2011-060 was endorsed as the first state record. On that single round of reconsideration, it was endorsed unanimously. SCISSOR-TAILED FLYCATCHER Tyrannus forficatus (6, 2). 2010-045, William Pratt, Greg Scyphers (P), James Hulsey (P), Jim Healy (P), Cathy Tobin, Michelle Tobin (P), 7 September-3 October 2010, Clark County Wetlands Park (Clark). Initially and again on 14 September 2010, this Scissor-tailed Flycatcher was reported from the Duck Creek section of the park. On 1 October 2010, what the committee judged to be the same individual was observed at the Henderson Bird Viewing Preserve. The distance between these two locations is less than one-half mile. 2011- 026, Rose Strickland, Dennis Ghiglieri (P), 13 May 2011, Miller’s Rest Stop (Esmeralda). PHILADELPHIA VIREO Vireo philadelphicus (2, 1). 2010-059, Joan Clarke, Rita Schlageter, Greg Scyphers (P, Figure 5), Martin Meyers (P), 19-23 October 2010, Floyd Lamb Park (Clark). The first Philadelphia Vireo the NBRC endorsed was found in 1995 (Cressman et al 1998). Nevada birders had been looking eagerly for another for 15 years before Clarke and Schlageter discovered the second. PURPLE MARTIN Progne subis (8, 4). 2010-072, Greg Scyphers (P), 7 September 2009, Key Pittman WMA (Lincoln). 2010- 026, Carl Lundblad (P), 23 May 2010, Ash Meadows NWR (Nye). 2011- 030, Dennis Serdehely, Greg Scyphers (P, Figure 6), 28 May 2011, Dyer (Esmeralda). 2011-033, Martin Meyers, Greg Scyphers, 4 June 2011, Dyer (Esmeralda). WINTER WREN Troglodytes hiemalis (1, 1). 2010-079, Michael J. San Miguel, Jr., 9 November 2000, Corn Creek (Clark), one of the classic desert migrant traps. San Miguel originally reported his sighting as of an “Eastern” Winter Wren, then, after the split of the Pacific Wren and Winter Wren was announced (Chesser et al. 2010), provided his complete notes on the sighting to the committee. The record was endorsed on a 5-1 vote, the one negative voice expressing concern that this first state record was based on a single observer providing only written evidence. That’s an issue the committee’s bylaws so far do not address. Another Winter Wren found at Pahranagat NWR in 2011 (amazingly, on the exact same date 11 years later) was seen and documented by many observers; the NBRC’s review of that record will be published in its report for 2012. SEDGE WREN Cistothorus platensis (1,1). 2011-101, Greg Scyphers (P, Figure 28 2011 NEVADA BIRD RECORDS COMMITTEE REPORT 7), Dennis Ghiglieri (P), Rose Strickland, 15 October 2011, Dyer (Esmeralda). Found by Scyphers at a private ranch at another of Nevada’s best migrant traps. BROWN THRASHER Toxostoma rufum (11, 2). 2010-086, Peter Shen (P), Bonnie Dunman, Jeanne Tinsman, 4 December 2010-29 March 2011, Corn Creek (Clark). 2011-038, Paul Lehman, Barbara Carlson, 4 July 2011, Miller’s Rest Stop (Es- meralda). When this sighting was first announced, some expressed surprise at the mid-summer date. However, Nevada has endorsed records of the Brown Thrasher for every season — including a previous one for July (see http:/gbbo.org/nbrc for specific dates.) CHESTNUT-COLLARED LONGSPUR Calcarius ornatus (8**, 1). 2011-044, C. S. Lawson, 28 January 1973, Floyd Lamb Park at Tule Springs (Clark). Lawson (1974) collected the specimen, which one member suggested should be reexamined. The eight endorsed records represent 40 individuals. Species removed from the review list, September, 2011. WORM-EATING WARBLER Helmitheros vermivorum (10, 1). 2011-010, Greg Scyphers (P), 26 September 2010, Pahranagat NWR (Lincoln). A submission for this species from the same general area on 8 October will be considered in the NBRC’s report for 2012, addressing the question of whether that sighting represents a con- tinuation of 2011-010. KENTUCKY WARBLER Geothlypis formosus (6, 1). 2010-071 , Greg Scyphers (P), Carl Lundblad (P), 26 September-1 October 2010, Pahranagat NWR (Lincoln). The records of this species are evenly divided between spring (all May) and fall (all September, with the 2010 record extending to 1 October.) MAGNOLIA WARBLER Setophaga magnolia (17**, 3). 2010-062, Richard Saval (P), 10 June 2006, Corn Creek (Clark). 2010-087, Greg Scyphers (P), 16 September 2010, Floyd Lamb Park (Clark). 2010- 085, Carl Lundblad (P), 1 October 2010, Crystal Springs (Lincoln). Species removed from the review list, September 2011. CHESTNUT-SIDED WARBLER Setophaga pensyluanica (15**, 2). 2011-055, Susanne Luther, Don Roberson, Joe Morlan, 27 May 1976, Lida (Esmeralda). 2011- 054, Jeri Langham, 1 Junel980, Dyer (Esmeralda). Species removed from the review list, September 2011. BLACK-THROATED BLUE WARBLER Setophaga caerulescens (17**, 8). 2011- 051, C. S. Lawson (P), 24 October 1971, Boulder City (Clark), 2 birds. 2011-043, Rebecca Benedetti, Ed Mark, Jane Thompson (P), 18 October 1992, Wadsworth (Washoe). 2011-048, John Brack (P), 22 October 1993, Miller’s Rest Stop (Esmeralda). 2011-056, Scott Terrill (P), 23 September 1995, Indian Springs (Clark). 2011-057, Marian Cressman (P), 11 December 1996, Corn Creek (Clark). 2011-027, NOT ENDORSED. 1 September 2002, Lahontan Diversion Dam (Churchill). This record was submitted 9 years after the sighting and contained brief field notes which had been taken at the time. The documentation strongly suggested a correct identification, but several members concluded that other species (particu- larly the Orange-crowned Warbler, Oreothlypis celata) had not been satisfactorily eliminated. An additional factor in the decision was the date, 1 September being earlier by two and a half weeks than any other endorsed record of the Black-throated Blue Warbler in Nevada, and earlier than any fall record for California (Hamilton et. al. 2007). The record received a 4-2 vote on the first round and a 2-4 vote on the second and final round. 2010-088, Greg Scyphers (P), 17 October 2010, Pahranagat NWR (Lincoln). 29 2011 NEVADA BIRD RECORDS COMMITTEE REPORT 2010-060, Rose Strickland, Dennis Ghiglieri (P), 25 October 2010, Miller’s Rest Stop (Esmeralda). 2010- 069, Andrew Lee (P), Peter Shen (P), 13 November-5 December 2010, Floyd Lamb Park (Clark). Note that five of the eight records endorsed in 20 1 1 are based on much earlier sightings. Species removed from the review list, September 2011. PALM WARBLER Setophaga palmarum, (15**, 4). 2011-046, John Brack (P), 2 October 1993, Miller’s Rest Stop (Esmeralda). 2011- 041, Scott Page, Gareth Pearson (P), 23 April 2006, Corn Creek (Clark). This record was documented with a single photograph (and almost no written descrip- tion.) One member did not find that sufficient for endorsement, but the remainder of the members voted to endorse. 2010- 080, Carl Lundblad (P), 1 October 2010, Pahranagat NWR (Lincoln). 2011- 011, Greg Scyphers (P), 17 October 2010, Pahranagat NWR (Lincoln). Species removed from the review list, September 2011. YELLOW-THROATED WARBLER Setophaga dominica (4, 3). 2010-081, Bruce Sorrie, Barbara Fearis, 22 April 1977, Kingston Campground (Lander). 2010- 082, Michael Perrone, 24 October 1979, Duckwater (Nye). 2011- 001, Bruce Lund (P), Richard Titus (P), Richard Saval (P), Martin Meyers (P), Greg Scyphers (P, Figure 8), 2-16 January 2011, Moapa NWR (Clark). This record is only the second of the Yellow-throated Warbler since the 1994 formation of the NBRC. This very cooperative bird remained for at least two weeks, permitting many Nevada birders to enjoy it. PRAIRIE WARBLER Setophaga discolor (5, 2). 2010-044, Greg Scyphers (P), Jon Dunn, Carl Lundblad (P), 5-6 September 2010, Dyer (Esmeralda). 2010-064, James Stuart (P), Richard Saval (P), Joyce Marie Stewart (P), 31 Octo- ber-3 November 2010, Red Rock Canyon National Conservation Area (Clark). The dearth of Nevada records of this species may seem surprising to California birders, but it should be noted that only about 5% of California’s records are from the interior (Hamilton et al. 2007). BLACK-THROATED GREEN WARBLER Setophaga uirens (2, 2). 2010-051, Andrew Lee (P), Richard Saval (P), Greg Scyphers (P), 30 September-2 October 2010, Floyd Lamb Park (Clark). This is the first record of the Black-throated Green Warbler to be submitted to the NBRC. Its unanimous endorsement establishes the species on the Nevada state list. 2010- 068, Andrew Lee (P), 6 November 2010, Sunset Park, Las Vegas (Clark). After Lee documented Nevada’s first Black-throated Green Warbler in September, 2010, he went out and found another one in November! CLAY-COLORED SPARROW Spizella pallida (8, 3). 2011-069, William Pratt, 31 August 2011, Henderson Bird Viewing Preserve (Clark). 2011- 078, Dennis Serdehely, 21 September 2011, Miller’s Rest Stop (Esmeralda). 2011-079, Andrew Lee (P), 6 October 2011, Sunset Park, Las Vegas (Clark), 2 birds. LARK BUNTING Calamospiza melanocorys (15**, 4). 2011-050, Marian Cress- man (P), 31 August 1996, Hiko (Lincoln). 2011-009, Greg Scyphers (P), 5 September 2010, Dyer (Esmeralda). 2010- 076, Greg Scyphers (P), Carl Lundblad (P), 19 September 2010, Highland Park, Tonopah (Nye). 2011- 077, Greg Scyphers (P), Martin Meyers (P), 17 September 2011, Dyer (Es- meralda), 2 birds. Species added to the review list in September 2007 and removed in September 2011. GRASSHOPPER SPARROW Ammodramus savannarum (5, 2). 2011-082, Rick Fridell (P), 29 September 2009, Miller’s Rest Stop (Esmeralda). A Grasshopper 30 2011 NEVADA BIRD RECORDS COMMITTEE REPORT Sparrow had been documented at this same well-known vagrant trap by Scyphers (2010-017) only one week before Fridell’s sighting (Meyers 2011). Excellent photos of both birds convinced all members of the committee that two different individuals were involved. 2010-057, Greg Scyphers (P), Martin Meyers (P), 24 October 2010, Key Pittman WMA (Lincoln). Some older field guides show the breeding range of the Grasshopper Sparrow as encompassing nearly the entire state of Nevada. All the field work for the Atlas of the Breeding Birds of Nevada (Floyd et.al. 2007), however, revealed only two birds, one “possible breeder” in Lander County and one “probable breeder” in White Pine County. In July 2001 , one year after field work for the atlas was completed, Walter Wehtje observed an adult carrying an object — perhaps a fecal sac — -in Lincoln County (Floyd et al. 2007). If the Grasshopper Sparrow does breed in Nevada, it is extraordinarily rare. And, as the number of endorsed records shows, it is a very rare visitor at any time. SWAMP SPARROW Melospiza georgiana (12, 5). 2011-013, Martin Meyers, 3 January 1998, Walker Lake (Mineral). 2010- 058, Martin Meyers, 17 October 2010, Corn Creek (Clark). 2011- 085, Martin Meyers, 16 October 2011, Dyer (Esmeralda). 2011-091, Paul Lehman, 30 October 2011, Pahranagat NWR (Lincoln). 2011-100, G. Scott Clemson, Rose Strickland, Dennis Ghiglieri (P), 24-26 No- vember 2011, Pahranagat NWR (Lincoln). SCARLET TANAGER Piranga olivacea (5, 1). 2010-056, Greg Scyphers (P), Martin Meyers (P), 24 October 2010, Pahranagat NWR (Lincoln). All but the single spring (23 May) record of the Scarlet Tanager for Nevada fall between 16 October and 26 November. COMMON GRACKLE Quiscalus quiscula (11, 1) . 2011-012, Greg Scyphers (P), 4 February 2010, Beatty (Nye). BALTIMORE ORIOLE Icterus galbula (5, 1). 2011-081, Rick Fridell (P), 16 September 2004, Corn Creek (Clark) LAWRENCE’S GOLDFINCH Spinus lawrencei (17”, 5). 2011-047, C. S. Lawson, (P), 1 October 1972, Corn Creek (Clark), 2 birds. 2011-045, Paul Lehman, Shawneen Finnegan (P), 14 October 1990, Indian Springs (Clark). 2011-049, Marian Cressman (P), 11 October 1992, Corn Creek (Clark). 2010- 048, Greg Scyphers (P), Dennis Serdehely, 5-9 September 2010, Dyer (Esmeralda). 2011- 059, Greg Scyphers (P), 22 May 2011, Dyer (Esmeralda), 2 birds, observed carrying nesting material, but the outcome of this attempt was not determined. This is the first known attempt of Lawrence’s Goldfinch to nest in Nevada. The 17 en- dorsed records represent 41 individual birds. Species removed from the review list, September 2011. ACKNOWLEDGMENTS The NBRC thanks everyone who contributed to the accounts contained in this report. All submissions, photos, advice, comments, and opinions are greatly appreci- ated. We apologize to anyone who may have been overlooked. Some of the contribu- tors on this list are no longer with us, but their contributions are no less appreciated. Tim Almond, Aaron Ambos, Rebecca Benedetti, Ned Bixler, John Brack, Bar- bara Carlson, Joan Clarke, G. Scott Clemson, James and Marian Cressman, Brian Day, Alan de Queiroz, Nancy Devon, Bonnie Dunman, Jon Dunn, Barbara Fearis, Shawneen Finnegan, Rick Fridell, Dennis Ghiglieri, Jim Healy, David Henderson, 31 2011 NEVADA BIRD RECORDS COMMITTEE REPORT Eric Hough, James Hulsey, Jeri Langham, Tony Leukering, Mark Kasprzyk, Scott Krammer, C.S. Lawson, Andrew Lee, Paul Lehman, Rob Lowry, Bruce Lund, Carl Lundblad, Susanne Luther, Bruce Mactavish, Ed Mark, Martin Meyers, Joe Morlan, Vincent Mowbray, Christina Nycek, Scott Page, Gareth Pearson, Carl Perretta, Michael Perrone, Fred Petersen, William Pratt, Peter Pyle, Don Roberson, Rya Ru- benthaler, Robert Rucker, Michael J. San Miguel, Jr., Richard Saval, Rita Schlageter, Greg Scyphers, Dennis Serdehely, David Shen, Peter Shen, Bruce Sorrie, Joyce Marie Stewart, Rose Strickland, James Stuart, Brian Sullivan, David Syzdek, Scott Terrill, Jane Thompson, Steve Ting, Jeanne Tinsman, Carolyn Titus, Richard Titus, Cathy Tobin, Michelle Tobin, Alan Wallace, Kevin Wallace, Fred Welden, Robert Wilkin, Robert Williams, Diane Wong. Special thanks to the entire staff of the Great Basin Bird Observatory for their gener- ous logistical support. Greg Scyphers and Dennis Serdehely reviewed early drafts of this report and contributed much-appreciated input. Kristie Nelson reviewed the first submitted draft thoroughly, and her comments were invaluable, as were those from Dan Gibson. Will Richardson performed the final review and made substantial im- provements. Thank you to editor Philip Unitt for his help with this report's final stages. LITERATURE CITED Alcorn, J. R., 1988. The Birds of Nevada, Fairview West Publishing, Fallon, NV. Chesser, R. T., Banks, R. C., Barker, F. K., Cicero, C., Dunn, J. L., Kratter, A. W., Lovette, I. J., Rasmussen, P. C., Remsen, J. V., Rising, J. D., Stotz, D. F., and Winker, K. 2010. Fifty-first supplement to the American Ornithologists’ Union Check-List of North American Birds. Auk 127:726-744. Cressman, J., Baepler, D., Biale, A., Crow, D., Hoskins, L., Neel, L., Ryser, F. Jr., and Eidel, J. 1998. Initial report of the Nevada Bird Records Committee: 1994-96 Records. Great Basin Birds 1:63-67 Floyd, T., Elphick, C. S., Chisholm, G., Mack, K., Elston, R. G., Ammon, E. M., and Boone, J. D. 2007. Atlas of the Breeding Birds of Nevada. Univ. of Nev. Press, Reno. Hamilton, R. A., Patten, M. A., and Erickson, R. A. 2007. Rare Birds of California. W. Field Ornithol., Camarillo, CA. Lawson, C. S., 1974. First Nevada record of Chestnut-collared Longspun Auk 91:432. Meyers, M. 2010. 2008/2009 Nevada Bird Records Committee Report, www.gbbo. org/nbrc/GBB_2 0 08_2 0 09_Report . pdf. Meyers, M. 2011. 2010 Nevada Bird Records Committee Report, www.gbbo.org/ nbrc/GBB_2010_Report.pdf. Smith, J. P., Farmer, C. J., Hoffman, S. W., Kaltenecker, G. S., Woodruff, K. Z., and Sherrington, P. F. 2008. Trends in autumn counts of migratory raptors in western North America, in State of North America’s Birds of Prey (K. L. Bildstein, J. P. Smith, E. Ruelas I., and R. R. Veit, eds.), p. 217-251. Nuttall Ornithol. Club and Am. Ornithol. Union Ser. Ornithol. 3. Titus, C. K. 1996. Field List of the Birds of Nevada. Red Rock Audubon Soc., Las Vegas, NV. Accepted 18 October 2012 32 NESTING ECOLOGY AND NEST SUCCESS OF THE BLUE GROSBEAK ALONG TWO RIVERS IN NEW MEXICO JEAN-LUC E. CARTRON, Department of Biology, MSC03 2020, 1 University of New Mexico, Albuquerque, New Mexico 87131-0001; jlec@unm.edu DEBORAH M. FINCH and DAVID L. HAWKSWORTH, USDA Forest Service, Rocky Mountain Research Station, 333 Broadway SE, Suite 115, Albuquerque, New Mexico 87102 SCOTT H. STOLESON, USDA Forest Service, Northern Research Station, 335 National Forge Rd., Irvine, Pennsylvania 16329-0267 ABSTRACT: From 1997 through 2008, we studied the nesting habits and nest success of the Blue Grosbeak ( Passerina caerulea) along the middle Gila River (1997-2001) and the middle Rio Grande (2000-2008) in New Mexico. A riparian forest of cottonwoods grows along both rivers, but the forest along the Rio Grande is a much more intensively managed ecosystem, with an understory dominated by saltcedar ( Tamarix spp.) and other non-native invasive plants, frequent wildfires, and large-scale attempts at remediation of the vegetation. Along the Gila River 100 (95%) of 105 nests were in native shrubs or trees, and the mean height of all nests was 3.4 m. Of 85 nests found along the Rio Grande, 54 (64%) were in saltcedar and 16 (19%) were in other non-native shrubs or trees. Mean nest height was 2.2 m, significantly lower than along the Gila River. Nests were typically found along edges along both rivers but were placed significantly farther from water along the Gila River. In spite of these differences in nest placement, the observed proportion of successful nests along the two rivers did not differ significantly: 28 (47%) of 60 nests along the Rio Grande, 36 (54%) of 67 nests along the Gila River. Overall, differences between the two sites in floristic composition and vegetation structure appeared to affect the placement of Blue Grosbeak nests more than they did nest success. During the last three decades, much has been published on the eco- logical role and threatened status of riparian areas in arid and semi-arid southwestern North America (e.g., Johnson et al. 1977, 1987, Hunter et al. 1988, Howe and Knopf 1991, Finch and Yong 2000, Smith et al. 2009a). In particular, southwestern riparian areas are important to birds at lower and middle elevations, where they concentrate key resources such as water, shade, insects and fruits for food, and suitable nest sites (Carothers et al. 1974, Stamp 1978, Ohmart and Anderson 1982, Rosenberg et al. 1982, Hunter et al. 1988, Cartron et al. 2000). In turn, alteration or loss of riparian areas through river regulation, groundwater pumping, woodcutting, overgrazing, and the spread of non-native, invasive plants has reduced ripar- ian bird species and communities (Unitt 1987, Knopf et al. 1988, Ohmart 1994, Cartron et al. 2000, Finch and Stoleson 2000). In the Southwest, the Blue Grosbeak ( Passerina caerulea ) is one of many birds strongly tied to riparian areas. Range-wide, the species occupies a diversity of habitats (Ingold 1993), but in southwestern North America, it breeds only in riparian areas (Johnson et al. 1987). Like other riparian birds, including the Southwestern Willow Flycatcher ( Empidonax trail lii extimus) (e.g., Owen et al. 2005, Sogge et al. 2008), the Blue Grosbeak uses not only native vegetation (e.g., Powell and Steidl 2000) but also Western Birds 44:33-44, 2013 33 NESTING ECOLOGY AND NEST SUCCESS OF THE BLUE GROSBEAK areas now dominated by the invasive saltcedar ( Tamarix spp.) (Hunter et al. 1988, Brown and Trosset 1989, Rosenberg et al. 1991, Ellis 1995). Beyond this finding, however, surprisingly little has been learned about the nesting habits of the Blue Grosbeak in southwestern riparian areas and how these habits may relate to anthropogenic changes in vegetation structure or floristic composition. Here, we analyze data on Blue Grosbeaks nesting along two main rivers of New Mexico, the Gila River and the Rio Grande. The middle reaches of these two rivers have been altered to different degrees and in different ways, which, in turn, have been addressed by different management strategies. The middle Gila River in the southwestern part of the state is an unregulated river that nonetheless has been seriously degraded in several places by uncon- trolled livestock grazing and by phreatophyte control (Boucher et al. 2003). To address local problems of the river deepening its channel, severe bank erosion, and loss of riparian vegetation, livestock grazing was discontinued within much of the riparian zone in the 1990s, followed by excavation and grading of the river banks and by planting of native vegetation. As a result, periodic flooding over the river’s bank was restored, and with it, sediment deposition and revegetation along the banks (Boucher et al. 2003). Compared to that along the middle Gila River, the riparian forest along the middle Rio Grande is a much more intensively managed ecosystem (Crawford et al. 1998, Cartron et al. 2008, Smith et al. 2009a, b). Since the completion of Cochiti Dam in 1973, the river no longer floods over its bank at most locations. Saltcedar and Russian olive (Elaeagnus angustifo- lia), two non-native plants introduced as ornamentals or for erosion control, have spread along the river, replacing most of the original, native understory vegetation. The lack of flooding and the spread of saltcedar and Russian olive have contributed to heavy fuel loads that enabled stand-replacing wildfires during the last decade (Cartron et al. 2008). To reduce the risk of fire, the understory vegetation now has to be removed mechanically at regular inter- vals, but this practice also promotes the spread of other non-native plants such as kochia ( Kochia scoparis). Here we focus primarily on a comparison of the Blue Grosbeak’s nesting habits and nest success along the Rio Grande and the Gila River. MATERIALS AND METHODS Study Areas Our study along the middle Gila River extended from 1997 to 2001, that along the middle Rio Grande from 2000 to 2008 (Figure 1). The Gila River study area is located in the Cliff— Gila Valley of Grant County, at elevations ranging from 1335 to 1420 m. It consisted of two disjunct sites (33° 1' N, 108° 35' W and 32° 46' N, 108° 34' W), of 50 ha and 25 ha both sup- porting a mosaic of brushy floodplain vegetation, narrow (width 10-200 m) patches of riparian forest, and fields along the Gila River and associated earthen irrigation ditches. Most patches of forest included in the study area consisted of stands of mature Fremont cottonwood ( Populus fremontu) with a canopy >25 m (reaching 40 m in places). The canopy or mid-story veg- 34 NESTING ECOLOGY AND NEST SUCCESS OF THE BLUE GROSBEAK Figure 1. Locations of study areas along the middle Gila River (1997-2001) and middle Rio Grande (2000-2008) in New Mexico. Black triangles represent sites of field work. 35 NESTING ECOLOGY AND NEST SUCCESS OF THE BLUE GROSBEAK etation also included Goodding’s willow ( Salix gooddingii), boxelder ( Acer negundo), velvet ash ( Fraxinus velutinus), Arizona walnut ( Juglans major), Arizona sycamore [Platanus wrightii), Arizona alder (Alnus oblongifolia), and Russian olive. In the understory were shrubs such as three-leaf sumac {Rhus trilobata), false indigo ( Amorpha fruticosa), and New Mexico olive (. Forestiera neomexicana), and forbs and grasses. Along the river, flood- ing over the banks promotes an early stage of open vegetation consisting of coyote willow ( Salix exigua), bluestem willow (S. irrorata), seepwillow (. Baccharis salicifolia), and cottonwood saplings. Most of the Gila River study area was on private land (the U Bar Ranch), where the riparian zone is flanked by dry and irrigated pastures used for livestock grazing and hay farming. Also included in the study area were lands managed by The Nature Conservancy and the Gila National Forest. The second study area was located along the middle Rio Grande from Cochiti Lake (35° 37' N, 106° 19' W, elevation 1608 m) in Sandoval County south approximately 235 km to the Bosque del Apache National Wildlife Refuge (33° 48' N, 106° 54' W, elevation 1372 m) in Socorro County (Figure 1). It consisted of 23 scattered sites representing a total of 454 ha along the river. Most of the 23 sites were narrow (width 50-250 m) patches of riparian vegetation, although the width of some of the sites in the south reached 400 m or more (maximum 600 m). The 23 study sites along the middle Rio Grande consisted of a mix of untreated cottonwood forest with a canopy of Rio Grande cottonwoods ( Populus deltoides ssp. wislizenii) and understory typically dominated by saltcedar and Russian olive, burned areas with little or no forest canopy, and patches of vegetation from which the understory was cleared sometime during the study. Within two years, burned sites were colonized or re-colonized by non-native plants including not only saltcedar and Russian olive but also kochia, Siberian elm ( Ulmus pumila), and tree of heaven {Ailanthus altissima) forming a dense shrub layer. Our study sites were variously under the administrative oversight of the Middle Rio Grande Conservancy District, the U.S. Bureau of Reclamation, the Bosque del Apache National Wildlife Refuge, and the city of Albuquerque’s Open Space Division. Land use adjacent to the riparian zone was mostly agricultural (pastures and crop fields). Field Methods Along both the Gila River and the Rio Grande, searches for Blue Grosbeak nests were part of a larger effort to study entire riparian bird communities, habitat associations, and responses to disturbance or management. We ac- tively searched for nests from the second week of May through the middle of August. Along the Rio Grande, stands of mature cottonwood and their edges were thoroughly searched, as were the more open river banks. In some areas the river-edge vegetation consisted of tall, nearly impenetrable thickets of willow and/or Russian olive. Although searches for nests were limited to the edges of those thickets, Blue Grosbeaks appeared to largely avoid the thickets’ interior, unlike other species such as the Spotted Towhee ( Pipilo maculatus). Along both rivers, we recorded nest height (m), nest substrate, nest-substrate height (m), distance between the nest and the trunk or main 36 NESTING ECOLOGY AND NEST SUCCESS OF THE BLUE GROSBEAK stem of the nest plant (m), diameter at breast height of the nest plant (cm), and distance to the nearest body of water (either the river or a riverside drain; m). Along the Gila River, we also measured the distance to the nearest edge. Nests were monitored every 3-5 days with use of binoculars, pole- mounted mirrors or video cameras, or 15x spotting scopes. Nests that were abandoned or destroyed were examined for evidence (e.g., cowbird eggs, mammal hairs) of the cause of nest failure. We considered a nest successful if (1) the parents were observed feeding one or more fledged young Blue Grosbeaks; (2) the parents behaved as if dependent young were nearby in the vicinity of a now-empty nest that had not been parasitized; or (3) Blue Grosbeak nestlings were in the nest within one or two days of the estimated date of fledging. We considered a nest to have failed if (1) the nest’s contents disappeared before fledging of young was possible (depredation), assuming the young require 9-10 days for fledging (Stabler 1959), (2) the nest contained no grosbeak young but contained cowbird eggs or chicks, (3) the nest was deserted after eggs had been laid (desertion), or (4) the nest was abandoned prior to egg laying (abandonment). Statistical Analyses For all analyses we used the statistical software Centurion XV. II (Statgraph- ics). We used t tests to identify significant differences between the Gila River and Rio Grande datasets in nest height, nest plant height, diameter at breast height of the nest plant, distance between nest and trunk or main stem, and distance between nest and nearest body of water. We assessed differences in proportions of successful and unsuccessful nests with chi-squared tests. RESULTS We found a total of 190 Blue Grosbeak nests during our study, 105 (55%) along the Gila River and 85 (45%) along the Rio Grande. In both areas, some pairs initiated nesting in late May, eggs were typically recorded from early June through mid-August, and young fledged through at least the end of August. The peak of the nesting season appeared to be in July, when 61 (72%) of the 85 Rio Grande nests and 73 (70%) of the 105 Gila River nests were active. One nest along the Gila River and one nest along the Rio Grande were found after mid-August and the end of our active nest searches. The late nest along the Rio Grande was discovered on 5 September 2003. It contained two Blue Grosbeak eggs and two cowbird eggs and was deserted. Nest Substrate and Nest Materials Along the Gila River, Blue Grosbeaks built their nests in plants of at least 20 species (Table 1). Most (70%) nests were found in four native species of shrub or tree, primarily boxelder, followed in order of decreasing frequency by seepwillow, Goodding’s willow, and netleaf hackberry ( Celtis reticulata). Only five nests were found in exotic vegetation, Russian olive or saltcedar. Atypical were nests built in plants such as stinging nettles ( Urtica dioica), common sunflower ( Helianthus annuus), and canyon grape ( Vitis arizo- nica). Along the Gila River we recorded the nest materials in only a few 37 NESTING ECOLOGY AND NEST SUCCESS OF THE BLUE GROSBEAK Table 1 Substrates and Heights of Blue Grosbeak Nests along the Middle Gila River (1997-2001) and Middle Rio Grande (2000- 2008) in New Mexico Plant species 0 No. (%) nests Mean Nest height (m) SD Range Middle Gila River Boxelder 29 (28) 6.73 3.88 0.3-15.0 Seepwillow 17 (16) 1.65 0.54 1. 1-3.1 Goodding’s willow 16 (15) 2.04 0.85 0.9-4. 3 Netleaf hackberry 12 (ID 2.46 1.23 1. 3-5.2 Fremont cottonwood 6 (6) 2.13 0.39 1.8-2. 8 New Mexico locust 4 (4) 1.47 0.67 0.7-2. 3 Russian olive 4 (4) 1.66 0.87 0.6-2. 7 Arizona sycamore 2 (2) 8.1 1.98 6. 7-9. 5 Arizona alder 2 (2) 3.15 1.06 2. 4-3. 9 Velvet mesquite 2 (2) 1.45 0.49 1. 1-1.8 Bluestem willow 2 (2) 2.65 0.64 2.2-3. 1 Saltcedar 1 (1) 2.4 — — Wingleaf soapberry 1 (1) 0.5 — — Velvet ash 1 (1) 0.8 — — Canyon grape 1 (1) 2.0 — — Stinging nettles 1 (1) 1.3 — — Catclaw mimosa 1 (1) 1.7 — — Common sunflower 1 (1) 0.9 — — Wild rose (Rosa sp.) 1 (1) 1.0 — — Acacia ( Acacia sp.) 1 (1) 0.9 — — Total 105(100) 3.36 3.12 0.3-15.0 Middle Rio Grande Saltcedar 54 (64) 2.12 1.22 0. 2-8.7 Russian olive 11 (13) 1.21 0.64 0.4-2. 5 Rio Grande cottonwood 7 (9) 6.64 5.27 1.4-15 Coyote willow 6 (7) 1.36 0.85 0.5-2. 5 Kochia 2 (2) 1.05 0.21 0.9-1. 2 Tree of heaven 1 (1) 1.03 — — White mulberry 1 (1) 0.50 — — Siberian elm 1 (1) 1.20 — — Desert olive 1 (1) 1.80 — — Goodding’s willow 1 (1) 0.96 — — Total 83(100) 2.24 2.24 0.2-15 a See text for the scientific names of nest plants; names of introduced species are in bold type. nests; these consisted of dried grasses, leaves, and, occasionally, also snake skin, used plastic flagging, and hair. Along the Middle Rio Grande, nests were found in 10 species of plants (Table 1). Most (82%) nests were built in non-native shrubs or trees, primarily saltcedar and Russian olive (e.g., Figure 2). Only 15 (18%) nests were found in native plants, mainly Rio Grande cottonwood and coyote willow. Nest materials included leaves and twigs of saltcedar, leaves, shredded bark, cot- ton, seed pods, and twigs from cottonwood trees, grass, forb stems, horse 38 NESTING ECOLOGY AND NEST SUCCESS OF THE BLUE GROSBEAK Figure 2. Nest of the Blue Grosbeak in a Russian olive along the middle Rio Grande, New Mexico. Photo by June Galloway, Rocky Mountain Research Station hair, snake skin, candy-bar wrappers, pieces of plastic bag, paper, and old flagging tape. Measurements of Nests’ Characteristics We pooled measurements of nests and nest plants for all years to maximize sample sizes and because there were no detectable annual differences in the five variables we measured (consensus t test, P > 0.5). Along the Gila River, observed nest height varied from 0.3 to 15 m, with a median of 2.1 m. Most (69%) nests were found at heights of <3 m, but nest height varied considerably by substrate (Table 1). Nests in boxelder were significantly higher than nests in the other three main substrates combined (t = 7.87, df = 1, P < 0.001). Nests also found at heights substantially higher than the mean included those in Arizona sycamore (Table 1). Along the middle Rio Grande, nest height varied from 0.16 m (one nest in a saltcedar) to 15 m (one nest in a cottonwood), with a median of 1.7 m. Most (86%) nests were placed at heights of <3 m. With the exception of those in cottonwoods, Blue Grosbeak nests were placed at mean heights of <2.5 m in every species of nest plant observed (Table 1). Nests were placed significantly lower along the Rio Grande than along the Gila River (Table 2). Additionally, nest plants were significantly shorter along the Rio Grande than along the Gila River, as they were in diameter at breast height. But the placement of the nest relative to the trunk or main stem did not differ. The mean distance of nests to water (either the river or a drain) was significantly smaller along the Rio Grande than along the Gila River. Of the 104 nests along the Gila River whose distance to the nearest 39 NESTING ECOLOGY AND NEST SUCCESS OF THE BLUE GROSBEAK Table 2 Characteristics of Blue Grosbeak Nests and Nest Plants along the Middle Gila River (1997-2001) and the Middle Rio Grande (2000-2008) in New Mexico Middle Gila River Middle Rio Grande Nest and nest-plant attributes Mean ± SD No. (%) nests Mean ± SD No. (%) nests Result of t test (P) a Nest height (m) Nest-plant 3.4 ±3.12 105 (100) 2.2 ± 2.2 83 (98) 0.006” height (m) Diameter of nest plant at breast 7.2 ±5.8 104 (99) 5.2 ± 3.3 77 (91) 0.006” height (cm) Distance from trunk or main 22.5 ±38.9 104 (99) 8.5 ± 13.5 77 (91) < 0.001” stem (m) Distance from 0.9 ± 1.5 113.9 ± 85 (81) 0.8 ± 2.2 67.1 ± 73 (86) 0.9 water (m) Distance from edge 177.8 93 (89) 57.4 80 (94) 0.025* (m) 4.7 ±8.3 104 (99) — — — “Levels of significance: *P < 0.05; **P < 0.01. edge was recorded, 83 (80%) were placed <5 m from the nearest edge; 48 (46%) were <1 m from the nearest edge (Table 2). Although distance to the nearest edge was not quantified during field work along the Rio Grande, nests placed along edges were also common, if not typical, in that study area. Nest Success We were able to determine the outcome of 60 nests along the Rio Grande and 67 nests along the Gila River (Table 3). The observed proportion of suc- cessful nests along the two rivers did not differ significantly (x 2 = 0.38, df = 1, P = 0.427). The cause of nest failure remained unknown in many cases, especially along the Rio Grande (Table 3). Along the Gila River, however, more than half of all failures were caused by predation. Less frequently, nest failure was caused by weather or the nest failed to produce any Blue Grosbeak fledglings but instead fledged cowbirds. Along the Rio Grande, cowbird parasitism, predation, and weather were all responsible for some of the failures. Along the Rio Grande, we were able to determine the outcome of 42 (78%) of the 54 nests in saltcedar, and of those 17 (45%) were successful, a proportion that was not significantly (x 2 = 0.15, df = 1, P = 0.694) different from that of nests in substrates other than saltcedar. DISCUSSION Several studies (Hunter et al. 1988, Brown and Trosset 1989, Rosenberg et al. 1991 , Ellis 1995) have shown the Blue Grosbeak to readily occupy, and nest in, saltcedar-dominated riparian habitats. Our study was not designed 40 NESTING ECOLOGY AND NEST SUCCESS OF THE BLUE GROSBEAK Table 3 Observed Total Number of Blue Grosbeak Nests, Successful Nests, and Nests that Failed by Cause along the Middle Gila River (1997- 2001) and the Middle Rio Grande (2000-2008), New Mexico Gila River Rio Grande Observed number of nests 105 (100) 85 (100) Number (%) of nests with known outcome 67 (64) 60 (71) Observed number (%) of nests that 36 (54) 28 (47) succeeded Observed number (%) of nests that failed 31 (46) 32 (53) Observed number (%) of nests that failed 17 (55) 3 (9) because of predation Observed number (%) of nests that failed 6 (19) 9 (28) because of cowbird parasitism Observed number (%) of nests that failed 2 (6) 0 (0) because of weather Observed number (%) of nests that failed 6 (19) 20 (62) from an undetermined cause to assess use of nest plants relative to their availability, but nesting Blue Grosbeaks did not seem to avoid saltcedar at any of the Rio Grande sites where it was dominant. Along the Rio Grande, we found the grosbeaks to nest not only in saltcedar-dominated vegetation but also to use saltcedar as the most frequent nest-supporting plant. In riparian areas, Blue Grosbeak nests are typically found close to the ground (Bent 1968, Rosenberg et al. 1991, Averill 1996). Powell and Steidl (2000), however, showed that nests can also be built high in trees where the understory vegetation is sparse. Along both the Rio Grande and the Gila River, most nests were constructed at heights of <3 m above ground, many of them at heights of only 1 m or less. We also detected nests in trees, however, and as Powell and Steidl (2000) reported, they tended to be found where little to no understory vegetation was present. Along the Rio Grande, we found seven nests in cottonwoods. Of those, only four were at heights of >2 m, and these were at sites where the shrub layer had been cleared. Tree nests were particularly common along the Gila River, in boxelders and Arizona sycamores. As the main nesting substrate, boxelder substantially raised the mean height of Blue Grosbeak nests along the Gila River. If nests in boxelder are excluded, the mean heights of nests along the Rio Grande (2.24 m) and along the Gila River (2.07 m) were similar. Boxelder tends to cast considerable shade, and along the Gila River we found it in stands with few shrubs growing underneath its foliage. Paralleling our finding that Blue Grosbeak nests were significantly higher along the Gila River than along the Rio Grande, Stoleson and Finch (2003) reported that along the Gila River Southwestern Willow Flycatchers nested primarily in boxelder and at heights greater than reported for other populations of that bird. Throughout much of its distribution, the Blue Grosbeak often nests along woodland edges adjacent to open areas (Ingold 1993), and an association between nests and wooded riparian edges has been noted in Arizona (e.g., 41 NESTING ECOLOGY AND NEST SUCCESS OF THE BLUE GROSBEAK LaRue 2005) and in California (Gaines 1974, Riparian Habitat Joint Venture 2004). That association was also evident in both of our study areas. Along the Gila River, where we measured distances between nests and the near- est edges, that association was particularly pronounced. In contrast, close proximity to water did not characterize many of the nests along either the Gila River or the Rio Grande. This finding is congruent with that of other studies (e.g., Pequegnat 1951) showing that Blue Grosbeaks can nest in the absence of nearby water. We found the mean distance to water to be significantly higher along the Gila River, where the riparian zone tended to extend farther from the river channel and where most Blue Grosbeak nests were along the edges between the riparian forest and fields rather than along the edges between the riparian forest and river bank (Stoleson pers. obs.). The differences between the Rio Grande and the Gila River we noted in the placement of Blue Grosbeak nests — in nest height, nest substrate, and distance from water — did not visibly translate into differences in nest success. Nest success was not significantly lower along the Rio Grande than along the Gila River despite the much greater representation of non-native plants in general and saltcedar in particular. Nesting in saltcedar was not associated with reproductive success lower than in other nests along the Rio Grande. Our study thus did not detect any direct harmful effect of saltcedar per se on the Blue Grosbeak’s nest success along the Rio Grande. This finding seems congruent with the fact that in terms of habitat quality, saltcedar cannot simply be dismissed as poor or unsuitable habitat. The quality of saltcedar as habitat for birds varies geographically; it is also better for some riparian species than for others (Sogge et al. 2008). ACKNOWLEDGMENTS We thank G. Bodner, K. Brodhead, P. Chan, J. Galloway, J. Garcia, B. Gibbons, J. Kelly, R. Hunt, M. Means, G. Sadoti, B. Trussel, H. Walker, H. Woodward, and numerous seasonal technicians for help recording nest and vegetation data. Access to study sites was provided by City of Albuquerque Open Space, the Middle Rio Grande Conservancy District, the U.S. Fish and Wildlife Service (Bosque del Apache National Wildlife Refuge), the Nature Conservancy, the Gila National Forest, Phelps Dodge Corporation, and the U Bar Ranch. Funding and additional support were provided by the Joint Fire Science Program, Ducks Unlimited, New Mexico State Parks, the U.S. Fish and Wildlife Service’s Bosque Improvement Initiative, the University of New Mexico, the U.S. Army Corps of Engineers, the Rocky Mountain Research Station’s Middle Rio Grande Ecosystem Management Group, the Gila National Forest, Phelps Dodge Corporation, the National Fish and Wildlife Foundation, and the Nature Conservancy. Special thanks also to Suzy Stephens, who drafted Figure 1, and to Steve Rothstein, Scott Durst, and two anonymous reviewers for their constructive and thoughtful comments on earlier drafts of the manuscript. LITERATURE CITED Averill, A. 1996. Brown-headed Cowbird parasitism of neotropical migratory song- birds in riparian areas along the lower Colorado River. M. S. thesis, Univ. of Ariz., Tucson. Bent, A. C. 1968. Life histories of North American cardinals, grosbeaks, buntings, towhees, finches, sparrows, and allies, part I. U.S. Natl. Mus. Bull. 42 NESTING ECOLOGY AND NEST SUCCESS OF THE BLUE GROSBEAK Boucher, P. F., Stoleson, S. H. Shook, R. S. Pope, R. D., and Monzingo, J. 2003. Riparian restoration on the Gila River, New Mexico, creates breeding habitat for Southwestern Willow Flycatchers. Studies Avian Biol. 26:135-142. Brown, B. T., and Trosset, M. W. 1989. Nesting-habitat relationships of riparian birds along the Colorado River in Grand Canyon, Arizona. Southw. Nat. 34:260-270. Carothers, S. W., Johnson, R. R., and Aitchison, S. W. 1974. Population structure and social organization of southwestern riparian birds. Am. Zool. 14:97-108. Cartron, J.-L. E., Lightfoot, D. C., Mygatt, J. E., Brantley, S. L., and Lowrey, T. K. 2008. A Field Guide to the Plants and Animals of the Middle Rio Grande Bosque. Univ. of New Mexico Press, Albuquerque. Cartron, J.-L. E., Stoleson, S. H., Stoleson, P. L. L., and Shaw, D. W. 2000. Ri- parian areas, in Livestock Management in the American Southwest: Ecology, Society and Economics (R. Jemison and C. Raish, eds), pp. 281-328. Elsevier Science, Amsterdam. Crawford, C. S., Ellis, L. M., and Molles, M. C. Jr. 1998. The middle Rio Grande bosque: An endangered ecosystem. New Mexico J. Sci. 36:277-299. Ellis, L. M. 1995. Bird use of saltcedar and cottonwood vegetation in the middle Rio Grande valley of New Mexico, U.S.A. J. Arid Environments 30:339-349. Finch, D. M., and Stoleson, S. H. (eds.). 2000. Status, ecology, and conservation of the Southwestern Willow Flycatcher. USD A Forest Service Gen. Tech. Rep. RMRS-60. Finch, D. M., and Yong, W. 2000. Landbird migration in riparian habitats of the middle Rio Grande: A case study. Studies Avian Biol. 20:88-98. Gaines, D. 1974. The nesting riparian avifauna of the Sacramento Valley, California, and the status of the Yellow-billed Cuckoo. M.S. thesis, Univ. of Calif., Davis. Howe, W. H., and Knopf, F. L. 1991. On the imminent decline of Rio Grande cot- tonwoods in central New Mexico. Southw. Nat. 36:218-224. Hunter, W. C., Ohmart, R. D., and Anderson, B. W. 1988. Use of exotic saltcedar (Tamarix chinensis ) by birds in arid riparian systems. Condor 90:113-123. Ingold, J. L. 1993. Blue Grosbeak (Guiraca caerulea ), in The Birds of North America (A. Poole and F. Gill, eds.), no. 79. Acad. Nat. Sci., Philadelphia. Johnson, R. R., Haight, L. T., and Simpson, J. M. 1977. Endangered species vs. endangered habitats: a concept, in Preservation and management of riparian habitat: A symposium (R. R. Johnson and D. A. Jones, tech, coords.), pp. 68-79. U.S. Forest Service Gen. Tech. Rep. RM-43. Rocky Mt. Forest and Range Exp. Sta., Fort Collins, CO. Johnson, R. R., Haight, L. T., and Simpson, J. M. 1987. Endangered habitats versus endangered species: A management challenge. W. Birds 18:89-96. Knopf, F. L., Johnson, R. R., Rich, T., Samson, F. B., and Szaro, R. C. 1988. Con- servation of riparian ecosystems in the United States. Wilson Bull. 100:272-284. LaRue, C. 2005. Blue Grosbeak ( Passerina caerulea ), in The Arizona Breeding Bird Atlas (T. E. Corman and C. Wise-Gervais, eds.), pp. 544-545. Univ. of New Mexico Press, Albuquerque. Ohmart, R. D. 1994. The effects of human-induced changes on the avifauna of western riparian habitats. Studies Avian Biol. 15:273-285. Ohmart, R. D., and Anderson, B. W. 1982. North American desert riparian ecosys- tems, in Reference Handbook on the Deserts of North America (G. L. Bender, ed.), pp. 433-479. Greenwood Press, Westport, CT. Owen, J. C., Sogge, M. K., and Kern, M. D. 2005. Habitat and sex differences in physiological condition of breeding Southwestern Willow Flycatchers ( Empi - donax traillii extimus). Auk 122:1261-1270. Pequegnat, W. E. 1951. The biota of the Santa Ana Mountains. J. Entomol. Zool. 42:1-84. 43 NESTING ECOLOGY AND NEST SUCCESS OF THE BLUE GROSBEAK Powell, B. F., and Steidl, R. J. 2000. Nesting habitat and reproductive success of southwestern riparian birds. Condor 102:823-831. Riparian Habitat Joint Venture. 2004. The riparian bird conservation plan: A strategy for reversing the decline of riparian associated birds in California, version 2.0. Calif. Partners in Flight; www.prbo.org/calpif/pdfs/riparian_v-2.pdf. Rosenberg, K. V., Ohmart, R. D., and Anderson, B. W. 1982. Community organi- zation of riparian breeding birds: Response to an annual resources peak. Auk 99:260-274. Rosenberg, K. V., Ohmart, R. D., Hunter, W. C., and Anderson, B. W. 1991. The Birds of the Lower Colorado River Valley. Univ. of Ariz. Press, Tucson. Smith, D. M., Finch, D. M., Gunning, C., Jemison, R., and Kelly, J. F. 2009a. Post- wildfire recovery of riparian vegetation during a period of water scarcity in the southwestern USA. Fire Ecol. Special Issue 5: 38-55. Smith, D. M., Finch, D. M., and Hawksworth, D. L. 2009b. Black-chinned Hum- mingbird nest-site selection and nest survival in response to fuel reduction in a southwestern riparian forest. Condor 111:641-652. Sogge, M. K., Sferra, S. J., and Paxton, E. 2008. Tamarix as habitat for birds: Im- plications for riparian restoration in the southwestern United States. Restoration Ecol. 16:146-154. Stabler, R. M. 1959. Nesting of the Blue Grosbeak in Colorado. Condor 61:46-48. Stamp, N. 1978. Breeding birds of riparian woodland in south-central Arizona. Condor 80:64-71. Stoleson, S. H., and Finch, D. M. 2003. Microhabitat use by breeding Southwestern Willow Flycatchers on the Gila River, New Mexico. Studies Avian Biol. 26:91-95. Unfit, P. 1987. Empidonax traillii extimus: An endangered subspecies. W. Birds 18:137-162. Accepted 5 October 2012 44 DISTRIBUTION AND MOVEMENT PATTERNS OF INDIVIDUAL CRESTED CARACARAS IN CALIFORNIA KRISTIE N. NELSON, P. 0. Box 402, Lee Vining, California 93541; knelson@prbo . org PETER PYLE, The Institute for Bird Populations, P. O. Box 1346, Point Reyes Station, California 94956 ABSTRACT: There are now numerous records of the Crested Caracara (Caracara cheriway) from California and elsewhere well north of its breeding range, but whether or not they represent wild birds or escapees from zoos or falconers has been debated. Through 2011, the California Bird Records Committee (CBRC) had accepted 49 records that they considered to represent naturally occurring vagrants, but decisions concerning the number of individuals involved in these records were haphazard. Therefore, we assessed the date, location, age, molt status, and appearance of ca- racaras representing 60 observations specific to date and location in California and propose that these records involve only 11 individuals, recorded between 1 and 34 times throughout the state; a twelfth individual was recorded from December 2011 to April 2012. Our 11 -bird scenario was proposed and accepted by the CBRC in January 2012. This synthesis clarifies the species' pattern of occurrence in California: ten of the 11 individuals were first detected in fall or winter, eight individuals were first detected in their first or second years, four of these eight were later detected at appropriate ages elsewhere in California, and six individuals moved north within the state. These patterns are consistent with birds moving north as wild vagrants and so support the CBRC’s decision to accept the Crested Caracara as a naturally occurring species. We hope that our analysis will help other records committees evaluate the status of this species in other regions, perhaps revealing a similar pattern of natural vagrancy throughout North America. The Crested Caracara ( Caracara cheriway) has a long and perplex- ing history of vagrancy and occurrence in North America far north of its typical breeding range in Florida, Texas, Arizona, and Mexico. The earliest extralimital reports are from Monterey, California in 1837 (Prevost and des Murs 1855) and from Ontario, Canada, in 1892 (Brewster 1893), followed by scattered reports throughout the 20th century of individuals from across the continent (American Ornithologists’ Union [AOU] 1998, California Bird Records Committee [CBRC] 2007). Several factors resulted in opinions (e.g., AOU 1998) that these North American records could represent escapees from zoos or falconers. The species was not known to make significant migrations (Morrison 1996), the birds were not detected in greater numbers closer to their natural range as is typical of vagrants, and many extralimital records failed to match the seasonal or regional distribu- tion patterns expected for naturally occurring vagrants. Additionally, some Crested Caracara records have involved known or suspected escapees (Potter et al. 1980, AOU 1998). However, Grinnell and Miller (1944) believed that early California records likely pertained to wild vagrants. The CBRC (2007) was initially reluctant to add the species to its list of California birds because of concerns that the 12 reports prior to 2000 may have represented escapees, but eight more records from 2001 to 2003, coinciding with a spate of records elsewhere in North America (Brinkley Western Birds 44:45-55, 2013 45 DISTRIBUTION AND MOVEMENT PATTERNS OF CRESTED CARACARAS and Lehman 2003), resulted in the addition of the Crested Caracara to California’s main list of naturally occurring species in 2004 (San Miguel and McGrath 2005) and acceptance of 49 records between September 1987 and October 2011 (Table 1). Of these 49, the CBRC concluded that 18 involved the same bird as in other records, resulting in 31 accepted individuals, but these decisions were haphazard, and a thorough vetting of this issue was needed (Pyle et al. 2011). We therefore assessed the date, location, age, molt status, and appearance of Crested Caracaras documented in California and found that these 49 accepted records could be confidently assigned to just 11 individuals. Our analysis not only resulted in a more conservative assessment of vagrancy to California, it greatly clarified the species’ patterns of occurrence and movement within the state. METHODS We examined photographs and written documentation for 60 date/ location-specific observations of the Crested Caracara in California and southern Oregon between 1987 and 2011 (Table 1). We considered the date, location, age, molt pattern, worn/broken remiges, and cere shape to identify individuals whenever possible. Initially, we grouped the records by a conservative approach, favoring an assumption that they represent fewer individuals. For example, if two or more birds were of the correct age and molt status, did not differ in any feature we could evaluate, had no overlap- ping dates, and were present at locations between which one individual could have reasonably traveled, we considered them likely the same individual. This assumption was supported in some cases where uniquely marked individuals were confirmed on different dates at different locations, sometimes quite far apart (Pyle and Sullivan 2010; Figures 1, 2). Once all records were assessed in this manner some patterns emerged that added confidence to our deci- sions at the level of the individual. We assigned identification (ID) numbers to records to track individuals, those we assumed were of the same bird receiving the same ID number in Table 1 . Age determination was based on a comparison of descriptions and photo- graphs from each record to information presented in Pyle (2008). We categorize individuals as in their first cycle (fledging to beginning of second prebasic molt), second cycle (start of second to start of third prebasic molt), third cycle (start of third to start of fourth prebasic molt), or definitive cycle (individuals in definitive or “adult” plumage); we considered a bird’s age unknown when documenta- tion was insufficient to determine its plumage cycle. In the first plumage cycle, Crested Caracaras are washed brown or buff overall, have pale streaks or tips to the upperpart feathers, lack distinct blackish bars on the nape and breast, and have thinner whitish rectrices that are more heavily barred brownish basally (Figure 3). During the first cycle most individuals undergo a limited preformative molt of scattered body feathers in their first winter and spring (Pyle 2005a), but this does not alter their appearance substantially. In definitive plumage, caracaras are blackish and white, lack pale streaks on the upperparts, have bold black and white bars on the nape and breast, and have broader rectrices with darker bars that diminish basally (Figure 3). In their second and third cycles the birds are variably intermediate in appearance between those in the first and 46 DISTRIBUTION AND MOVEMENT PATTERNS OF CRESTED CARACARAS Table 1 Records of the Crested Caracara in California between September 1987 and April 2012 Representing 12 Individuals ID# a CBRC record Location Dates Age b 1 1987-267 Mono Lake, Mono Co. 13 Sep-16 Oct 1987 2C 1 1989-045 Shasta Valley, Siskiyou Co. 21 Oct 1988-26 Mar 1989 3C 1 1989-096 Smith River, Del Norte Co. 28-30 Apr 1989 3C or DC 1 NAB 44:490 Gold Beach, Oregon fall 1989-Apr 1990 DC 2 1993-196 Westmoreland, Imperial Co. 14 Dec 1993 1C 3 1995-021 Chula Vista, San Diego Co. 9 Feb 1995 DC 4 2005-026 Santa Barbara, Santa Barbara Co. 9 Oct 2001 U 4 2002-164 Long Beach, Los Angeles Co. 20 Oct 2001 U 4 Report near Gaviota, Santa Barbara Co. 24 Jan 2002 U 4 2002-147 Goleta, Santa Barbara Co. 30 Apr 2002 2C 4 Report Ventura, Ventura Co. 5 May 2002 U 4 2002-130 near Lakeview, Riverside Co. 4 Jul 2002 2C 4 2002-148 Vandenberg Air Force Base, Santa Barbara Co. 14-23 Jul 2002 2C 4 2002-192 Point Mugu, Ventura Co. 8 Aug 2002 2C 4 2002-154 near Marina, Monterey Co. 11-13 Aug 2002 2C 4 NAB 57:113 Santa Cruz, Santa Cruz Co. 16 Sep 2002 2C 4 2002-161 near Davenport, Santa Cruz Co. 21-27 Sep 2002 2C 4 NAB 57:113 Pescadero, San Mateo Co. 28 Sep 2002 U 4 NAB 57:113 Half Moon Bay, San Mateo Co. 21 Oct-6 Nov 2002 U 4 2002-209 Saticoy, Ventura Co. 9 Dec 2002-5 Jan 2003 2C 4 2004-118 near Petaluma, Sonoma Co. 16 Jul-1 Aug 2004 DC 4 2004-124 Manchester State Park, Mendocino Co. 20-24 Aug 2004 DC 4 2004-133 Jacoby Creek, Humboldt Co. 4-6 Sep 2004 DC 4 NAB 59:319 Eel River, Humboldt Co. Nov 2004-Jan 2005 DC 4 NAB 59:485 Curry County, Oregon Late Apr 2005 DC 4 2005-057 near Casper, Mendocino Co. 2 May 2005 DC 4 2005-070 Point Reyes National Seashore, Marin Co. 3-6 May 2005 DC 4 NAB 59:490 Rodeo Lagoon, Marin Co. 7 May 2005 DC 4 2005-071 Morro Bay, San Luis Obispo Co. 1-2 Jun 2005 DC 4 2005-086 Fort Dick, Del Norte Co. 13 Jun-12 Jul 2005 DC 4 2005-089 near Alton, Humboldt Co. 19 Jul 2005 DC 4 2005-097 Rodeo Lagoon, Marin Co. 2-3 Aug 2005 DC 4 2005-100 near Davenport, Santa Cruz Co. 14 Aug 2005-10 Apr 2006 DC 4 2006-129 Occidental, etc., Sonoma Co. 23 Apr-9 May 2006 DC 4 2006-078 Point Reyes National Seashore, Marin Co. 17-20 Jun 2006 DC 4 2006-084 S Humboldt Bay, Humboldt Co. 13-14 Jul 2006 DC 4 2007-076 near Ferndale, Humboldt Co. 8 Dec 2006-28 Mar 2007 DC (continued) 47 Table 1 ( Continued ) ID# a CBRC record Location Dates Age b 4 NAB 62:298 Ferndale, Humboldt Co. 30 Dec 2007 DC 4 2008-027 near Fort Dick, Del Norte Co. 31 Jan 2008-12 Sep 2009 DC 4 NAB 64:317 near Fort Dick, Del Norte Co. 29 Jan 2010-26 Feb 2012 DC 5 2004-074 Owen’s Lake, Inyo Co. 9 May 2004 U 6 2005-017 Finney Lake, Imperial Co. 1 Jan 2005 1C 7 2006-004 McGrath State Beach, Ventura Co. 5 Jan 2006 DC 7 2006-047 Point Sur, Monterey Co. 28-29 Mar 2006 DC 7 2006-051 Carmel Beach, Monterey Co. 10 Apr 2006-13 May 2007 DC 8 2006-042 Bixby Beach, Santa Barbara Co. 10 Jan 2006 1C 9 2006-127 Tijuana R Valley, San Diego Co. 9 Sep 2006-12 Feb 2007 2C 9 2007-144 Tijuana R Valley, San Diego Co. 22 Jun 2007-6 Jun 2008 3C 9 2008-093 Tijuana R Valley, San Diego Co. 15 Jul-21 Nov 2008 DC 9 2009-076 Tijuana R Valley, San Diego Co. 25 Apr-14 Nov 2009 DC 9 2011-139 Tijuana R Valley, San Diego Co. 25 Sep-1 Oct 2011 DC 10 2007-027 Hansen Dam, Los Angeles Co. 29 Jan-1 Feb 2007 1C 10 2007-083 near Goleta, Santa Barbara Co. 2-3 Feb 2007 1C 10 2007-101 Carmel, etc., Monterey Co. 25 Feb-1 Mar 2007 1C 10 2008-043 Ano Nuevo State Park, San Mateo Co. 14 Feb 2008 2C 10 2011-187 Santa Cruz, Santa Cruz Co. 27 Feb 2008 2C 10 2008-041 Morgan Hill, Santa Clara Co. 7 Mar 2008 2C 10 2008-080 Santa Barbara, Santa Barbara Co. 21 Mar 2008 2C 10 2009-082 Marina, Monterey Co. 4-23 Jul 2008 3C 11 2008-039 near Victorville, San Bernardino Co. 18-29 Feb 2008 2C 11 2008-047 near Weldon, Kern Co. 25-27 Mar 2008 2C 12 2011-239 Vandenberg AFB, Ventura Co. 19 Dec 2011-6 Jan 2012 DC 12 2012-005 Ballona Wetlands, Los Angeles Co. 13 Jan 2012 DC 12 2012-019 Piedras Blancas, San Luis Obispo Co. 10-21 Feb 2012 DC 12 2012-069 Pt. Sur, Monterey Co. 30 Mar 2012 DC 12 2012-052 near Davis, Yolo Co. 16-20 Apr 2012 DC “Identifying number of individual. b lC, first cycle; 2C, second cycle; 3C, third cycle; DC, definitive cycle; U, unknown. 48 DISTRIBUTION AND MOVEMENT PATTERNS OF CRESTED CARACARAS Figure 1. The notched P8, broken tip of P7, and the spiky P6 on the right wing confirmed the same individual in its second cycle (Table 1, ID #11) was photographed at Victorville, San Bernardino County, on 28 February 2008 (left) and then at the Kern River Preserve, Kern County, on 27 March 2008 (right). Photos by Stephen J. Myers (left) and Alison Sheehey (right) definitive cycles (Pyle 2008; Figure 3). Most in their second cycle but few in their third cycle, after completion of the third prebasic molt, are distinguishable from those in definitive plumage. The second, third, and definitive prebasic molts are complete; replacement of primaries and secondaries commences in the middle of each tract (at p4-p5 and s5, respectively), in western North America usually from February to April, and proceeds both distally and proximally from these molt centers until completed, usually from September to November (Pyle 2005b, 2008; Howell 2010). Within a cycle, the dark body feathering becomes increasingly bleached and brownish with wear in spring and summer, preceding replacement of body feathers from July through October. RESULTS We conclude that the most appropriate scenario accounting for the 60 caracara reports in California from September 1987 to October 2011 con- sists of 11 individuals, as outlined in Table 1 and accepted by the CBRC in January 2012 (Johnson et al. 2012). Clear patterns of vagrancy emerged when records were grouped according to this scenario. Eight of these 11 Crested Caracaras were first detected in their first or second cycle in the 49 DISTRIBUTION AND MOVEMENT PATTERNS OF CRESTED CARACARAS Figure 2. Adult Crested Caracara at the Ballona Wetlands, Los Angeles County, 13 January 2012 (left), and Piedras Blancas, San Luis Obispo County, 10 February 2012 (right). Note the large notch in the longest primary (P8), the fringed P7, and broken tip of P6 on the left wing in both images. These and other features indicated that this individual was first detected in Ventura County and subsequently observed and photographed in Monterey and Yolo counties (Table 1, ID #12), traveling north well over 800 km within 94 days. Photos by Christopher Taylor (left) and William Bouton (right) fall and winter, and four of these were later detected in later plumages at expected ages in subsequent years, adding confidence to our determinations of age and individual status. Two were first detected in definitive plumage in the winter, and one bird was of uncertain age when first detected in spring. Under this scenario, six individuals remained in the state for periods ranging from 5 weeks to multiple years, including 1 1 years and counting in the case of one individual (ID #4; Table 1). This individual accounted for over half of all California reports as it traveled from Santa Barbara County in Octo- ber 2001 to southern Oregon, back south to Marin and San Luis Obispo counties, then back north, finally settling in Del Norte County, California, from 2008 through 2012 (Figure 4). Most reports of adults, the age least expected to wander so far north naturally, pertained to birds (mainly ID #4) first detected in California in their first or second cycle, which then remained in the state to molt into definitive plumage (Table 1). Following the decision by the CBRC to accept 1 1 individuals, five additional records from December 2011 to April 2012 have been accepted which we conclude represent one additional individual (ID #12; Table 1, Figure 2). At least six individuals (ID #1,4, 9, 10, 11, and 12) moved north, usually 50 DISTRIBUTION AND MOVEMENT PATTERNS OF CRESTED CARACARAS Figure 3. Crested Caracaras in California showing plumage variation used to identify molt cycle. The bird on the left was photographed 10 September 2006 (CBRC 2006- 127) and has begun the second prebasic molt (so is technically in its second cycle) but still retains most of the plumage of the first cycle. The bird in the center, in its second cycle, was photographed 28 February 2008 (CBRC 2008-039). The bird on the right, photographed 17 September 2005, is in definitive plumage (CBRC 2005- 100). Note the progression from browner plumage with indistinct buff breast streaks to intermediate plumage to blacker plumage with white and black bars on the breast. Photos by Vic Murayama (left), Stephen J. Myers (middle), and Larry Selman (www.mostlybirds.com, right) during their first or second years, after which some moved south in the fall and north in the spring, at times returning to the same location in successive years. For example, in the winter of 2005-2006 ID #4 returned to the same field, using the exact same perches in Santa Cruz County as it had used in September 2002 when it was in its second cycle (D. Suddjian pers. comm.). It also returned to or near other coastal locations in Marin, Sonoma, Men- docino, Humboldt, and Del Norte counties on different dates as it traveled up and down the coast. ID #10 tended to spend summer and fall in the same area of San Diego, disappeared in winter, perhaps to Mexico, and returned to the same spot in San Diego the following spring or summer. Similarly, along the south coast of Oregon an adult Crested Caracara was seen near Floras Lake on approximately the same date for three consecutive springs (Mlodinow et al. 2007), suggesting one bird retracing a similar route annually. DISCUSSION Our scenario greatly alters our interpretation of the patterns of occurrence and movement of the Crested Caracara in California, not only by reduc- ing the total number of accepted individuals through April 2012 from 36 to 12, but by showing that most birds first arrive in California during their first or second fall or winter (as expected of wild vagrants) and subsequently may repeat their paths of migration. Three individuals were first detected in California in fall, eight during winter, and only one bird was first detected in remote Inyo County in spring; all birds previously thought to have shown 51 DISTRIBUTION AND MOVEMENT PATTERNS OF CRESTED CARACARAS up in summer can be linked to previous arrivals. This is consistent with the caracara’s pattern of wandering elsewhere outside its core range (M.J. Iliff in litt. ; see also Brinkley 2006, CBRC 2007). In Arizona, for example, its dispersal north and west of its breeding range during fall and winter has long been known (Phillips et al. 1964, Monson and Phillips 1981) and has increased notably in recent years (Stevenson and Rosenberg 2007, G. Rosenberg pers. comm.). In California, initial detections generally cluster during periods when vagrant Crested Caracaras were reported elsewhere in North America, providing evidence of periodic pulses of northward dispersal on a fairly large scale. One such event occurred in 2001 and 2002 when, in addition to the bird that was first detected in California (ID #4, Table 1), at least seven other vagrant caracaras were reported across the continent (Brinkley and Lehman 2003, San Miguel and McGrath 2005). A second, even larger pulse of Crested Caracaras dispersing north occurred during a roughly 4-year period beginning in 2005. From 2006 to 2008, when California experienced the largest influx of new arrivals (5), Oregon, Washington, British Columbia, New Mexico, Arizona, and Nevada also recorded vagrant Crested Caracaras in above-average numbers, as did the Great Plains and northeastern states (Brinkley 2006, CBRC 2007, Lehman and Brinkley 2009). Although the records may suggest increasing dispersal of Crested Caraca- ras into California, it is difficult to remove the effects of increased observer effort and documentation capability from this pattern, and we suspect that natural dispersal and vagrancy has long been occurring, an opinion shared by Grinnell and Miller (1944). Nevertheless, in the first report of the caracara in California, in 1837, Prevost and des Murs (1855) wrote, “M. le docteur Neboux l’a rencontre a Monterey (Haute-Californie), d’ou il en a rapporte plusieurs exemplaires ; en sorte que l’habitat de cet oiseau se trouve recule par le fait au nord, de pres de 10 degres” [Dr. Neboux encountered it at Monterey (Upper California), from where he reported several examples of it; by which fact the range of this bird is pushed back to the north, by nearly 10 degrees], “Several” caracaras at Monterey seem unlikely, and the ship on which Neboux was the physician and naturalist stopped at Magdalena Bay in Baja California and Mazatlan, San Bias, and Acapulco in western mainland Mexico after it left Monterey (Palmer 1918), a possible source of confusion. Alternatively, it is quite possible that expanding or retracting population sizes can explain the species’ irregularity of vagrancy throughout North America. The population of Florida may not be as prone to vagrancy as that farther west (Morrison and Dwyer 2012). Figure 4. Locations and dates of observation for one Crested Caracara remaining in California over 11 years (Table 1, ID #4). This individual was first detected in its first plumage cycle in Santa Barbara County in October 2001, moved between Riverside and San Mateo counties in 2002 and 2003 (A), traveled extensively between central California and southern Oregon from 2004 to 2006 (B), then moved north to settle in Humboldt and Del Norte counties from 2007 through 2012 (C). Gray dots represent reports from the earlier year(s) of each period, and different arrow types show major travel routes. Between migrations, this individual often returned to the same areas and, subsequent to 2003, was consistently documented in definitive plumage with the same cere/nostril shape, leading us to conclude that it was the same individual. 52 DISTRIBUTION AND MOVEMENT PATTERNS OF CRESTED CARACARAS A: 2001 - 2003 53 DISTRIBUTION AND MOVEMENT PATTERNS OF CRESTED CARACARAS Under our scenario there are over 12 instances of 6 individuals moving substantial distances to the north, often in short periods of time (Figures 1, 2, 4); Pyle and Sullivan (2010) documented that ID #10 dispersed north well over 450 km in 25 days, including over 125 km covered within 24 hours (Table 1). ID #4 traveled extensively in California over 8 years, mak- ing one northward flight of over 800 km within 11 days in 2005 (Figure 4, Table 1). ID #12 traveled north at least 800 km in early 2012 (Figure 2). These movements provide additional evidence that these birds are naturally occurring vagrants arriving from Mexico or Arizona, having dispersed north or northwest into California before continuing moving in those directions. Even the relatively stationary bird in San Diego County (ID #9) could have first reached this locale through a significant northward dispersal in its first cycle. In addition, records from the northern tier of the United States and southern Canada represent distances of 3000 km or more from the species’ typical range. As many states in the intervening region across the continent have also documented the Crested Caracara, it seems reasonable to infer that some or most of these records represent genuine vagrants. Although occurrence along the coast might seem illogical for a vagrant from mainland Mexico, it is clearly the pattern for the Crested Caracara in the West (e.g., Figures 2, 4). Most California records, both historic and modern, are for the immediate coast, as they are for Oregon, Washington, and British Columbia. A concentration along the coast might be expected if these birds tend to wander northwest. A similar trend in fall or winter coastal vagrancy is seen in other southern or Mexican vagrants to Califor- nia (CBRC 2007), including the Broad-billed Hummingbird (Cynanthus latirostris), Greater Pewee ( Contopus pertinax), Tropical Kingbird ( Tyran - nus melancholicus), Dusky-capped Flycatcher ( Myiarchus tuberculifer), Sulphur-bellied Flycatcher ( Myiodynastes luteiventris), and Streak-backed Oriole ( Icterus pustulatus). That observation effort along the coasts is greater than in interior regions is well known and likely has contributed to the coastal trend in Crested Caracara records as well. We hope that our analysis will help other records committees evaluate the status of the Crested Caracara in their regions, perhaps contributing to a similar overall pattern of natural vagrancy throughout North America. ACKNOWLEDGMENTS We are grateful to the photographers and observers who wrote detailed descrip- tions; without their contributions this analysis would not be possible. We also thank the Western Foundation of Vertebrate Zoology, especially Adam Searcy. Thanks to the CBRC for reviewing records and providing guidance. The manuscript greatly benefited from comments by Joan Morrison, Gary Rosenberg, Bob Gill, and Kimball Garrett. We thank Philip Unitt for help with the manuscript and for alerting us to the history and literature pertaining to the 1837 record. Finally we thank county bird records compilers and others who contributed information, including Alan Barron, David Suddjian, David Compton, John Sterling, and Oscar Johnson. LITERATURE CITED American Ornithologists’ Union (AOU). 1998. Check-list of North American Birds, 7th ed. Am. Ornithol. Union, Washington, D.C. 54 DISTRIBUTION AND MOVEMENT PATTERNS OF CRESTED CARACARAS Brewster, W. 1893. Capture of another Ardetta neoxena at Toronto, Ontario. Auk 10:363-364. Brinkley, E. S. 2006. The changing seasons: Far fetched. N. Am. Birds 60:332-339. Brinkley, E. S., and Lehman, P. E. 2003. Changing seasons: Unabashed bonanza. N. Am. Birds. 57:14-21. California Bird Records Committee (R. A. Hamilton, M. A. Pattern, and R. A. Erickson, eds.). 2007. Rare Birds of California Western Field Ornithologists, Camarillo, CA. Grinnell, J., and Miller, A. H. 1944. The distribution of the birds of California. Pacific Coast Avifauna 27. Howell, S. N. G. 2010. Molt in North American Birds. Houghton Mifflin Harcourt, New York. Johnson, O., Sullivan, B. L., and McCaskie, G. 2012. The 36th annul report of the California Bird Records Committee: 2010 records. W. Birds 43:164-188. Lehman, P. L., and Brinkley, E. S. 2009. The changing seasons: Cornucopia. N. Am. Birds. 63:16-32 Mlodinow, S., Irons, D., and Tweit, B. 2007. The spring season: Oregon/Washington region. N. Am. Birds. 61:499-504. Monson, G., and Phillips, A. R. 1981. Annotated checklist of the birds of Arizona. Univ. of Ariz. Press, Tucson. Morrison, J. L.1996. Crested Caracara ( Caracara cheriway ), in The Birds of North America (A. Poole and F. Gill, eds.), no. 249. Acad. Nat. Sci., Philadelphia. Morrison, J. L., and J. Dwyer. 2012. Crested Caracara ( Caracara cheriway), in The Birds of North America Online (A. Poole, ed.), no. 249. Cornell Lab Ornithol., Ithaca, NY; http://bna.birds.cornell.edu/bna/species/249. Palmer, T. S. 1918. Costa’s Hummingbird — its early history, type locality and name. Condor 20: 114-116. Phillips, A. R.,. Marshall, J., and Monson, G. 1964. The Birds of Arizona. Univ. of Ariz. Press, Tucson. Potter, E. F., Parnell, J. F., and Teulings, R. P. 1980. The Birds of the Carolinas. Univ. of North Carolina Press, Chapel Hill. Prevost, F., and des Murs, O. 1855. Oiseaux, in Voyage autour du Monde sur la Fre- gate la Venus, Commandee par Abel du Petit-Thouars. Zoologie: Mammiferes, Oiseaux, Reptiles, et Poissons, pp. 177-284. Gide et J. Baudry, Paris. Pyle, P. 2005a. Preformative molts in North American Falconiformes. Raptor Re- search 39:378-385. Pyle, P. 2005b. Remigial molt patterns in North American Falconiformes as related to age, sex, breeding status, and life-history strategies. Condor 107:823-834. Pyle, P. 2008. Identification Guide to North American Birds. Part 2. Slate Creek Press, Point Reyes Station, CA. Pyle, P., and Sullivan, B. 2010. Documenting repeated occurrences of individual birds with digital images. W. Birds 41:261-265 Pyle, P., Tietz, J., and McCaskie, G. 2011. Thirty-fifth report of the California Bird Records Committee: 2009 records. W. Birds: 42:134-163. Roberson, D. R. 1993. Fourteenth report of the California Bird Records Committee. W. Birds 24:113-166. San Miguel, M., and McGrath, T. 2005. Report of the California Bird Records Com- mittee: 2003 records. W. Birds 36:78-113. Stevenson, M., and Rosenberg, G. 2007. The fall season: Arizona. N. Am. Birds 62:121. Tweit, B., and Fix, D. 1990. The spring season: Oregon/Washington region. Am. Birds 44:490. Accepted 6 August 2012 55 CONSPECIFIC NEST AGGRESSION OF THE PACIFIC WREN ON VANCOUVER ISLAND, BRITISH COLUMBIA ANN NIGHTINGALE and RON MELCER, JR., Rocky Point Bird Observatory, Suite #170, 1581-H Hillside Avenue, Victoria, British Columbia, Canada V8T 2C1; corvid88@gmail.com ABSTRACT: Five of the ten wren species in North America are known to destroy nests of conspecifics. These include the Cactus Wren ( Campylorhynchus brunneica- pillus), Bewick’s Wren (' Thryomanes bewickii), Sedge Wren ( Cistothorus platensis), Marsh Wren ( Cistothorus palustris), and House Wren ( Troglodytes aedon). However, none of the Winter Wren complex, recently split as the Winter Wren ( Troglodytes hiemalis), Pacific Wren (T pacificus), and Eurasian Wren ( T. troglodytes), have been documented to do so in experiments or by observation of natural behavior. Here we present a detailed chronology of a nesting of the Pacific Wren — the first report of conspecific nest aggression in the Winter Wren complex. On 15 May 2011, in Victoria, British Columbia, Canada, a Pacific Wren approached another’s nest under video surveillance and removed two 9-day-old chicks. The nonparental adult returned to the nest, apparently attempting to kill and/or and remove the remaining two chicks, several times over 4.75 hours but was not successful. Although our findings are limited to a single event, they are consistent with those of other wrens. Birds are known to destroy the nests and eggs or remove the young from nests of other species, as well as conspecifics, to reduce competition for nests, food, perches, and, in polygynous species, for the male’s parental care (Fox 1975, Verner 1975, Pieman 1977a, Jones 1982). Within the order Passeriformes, members of the wren family (Troglodytidae) are par- ticularly well known for nest destruction and infanticide. This behavior has been observed in both experimental and natural settings in the House Wren {Troglodytes aedon) (Pribil and Pieman 1991), Marsh Wren ( Cistothorus palustris) (Pieman 1977a, b, Leonard and Pieman 1987, Kroodsma and Verner 1997), Bewick’s Wren (Thryomanes bewickii) (Pieman 1994), Cac- tus Wren (Campylorhynchus brunneicapillus) (Simons and Simons 1990), and Sedge Wren ( Cistothorus platensis ) (Pieman and Pieman 1980). Nest destruction has not been reported previously in the Winter Wren complex, recently split into three species, the Pacific Wren ( Troglodytes pacificus ), Winter Wren (T hiemalis), and Eurasian Wren {T. troglodytes). Other North American wren species that have not been documented destroying nests are the Canyon Wren {Catherpes mexicanus), Carolina Wren ( Thryothorus ludovicianus), and Rock Wren ( Salpinctes obsoletus). The Pacific Wren breeds in forests with dense undergrowth (Campbell et al. 1997), typically nesting in cavities built by other birds, crevices among the roots of upturned trees, and other spaces with relatively small entry holes. Occasionally, it builds its nest on ledges or in other more open areas (Campbell et al. 1997). Whatever the site, the construction is similar, resulting in an orb of grass and moss lined with plant down and fur. Hence the nature of the of Pacific Wren’s nest sites has provided few opportunities for observation. Here we describe in detail the first known instance of conspecific nest aggression in the Winter Wren complex. 56 Western Birds 44:56-64, 2013 CONSPECIFIC NEST AGGRESSION OF THE PACIFIC WREN STUDY AREA AND METHODS On 1 April 2011, we observed the beginning of nest construction on a ledge in a carport adjacent to a riparian area within the Georgia Depression on southeastern Vancouver Island, British Columbia (48° 35' 27" N, 123° 25' 38" W, 60 m above sea level). The ledge was 2.9 m from the ground, and one to three Pacific Wrens had used it as a winter roost for at least 5 years. The area is a mix of suburban and rural housing, with typical lot sizes ranging from about 0.05 ha on one side of Graham Creek to 0.4-4 ha on the other. The creek is within undeveloped private property and a protected linear natural park a minimum of about 75 m wide, extending about 750 m along the creek in both directions from the nest. The riparian habitat consisted of a structurally and compositionally diverse mixed forest on the steep banks of the creek; the dominant species of trees included Douglas-fir (Pseudotsuga menziesii), Grand Fir ( Abies grandis), Western Red Cedar ( Thuja plicata), Red Alder ( Alnus rubra), Big Leaf Maple ( Acer macrophyllum) , and Pacific Dogwood (Cornus sericea). A mixture of na- tive and nonnative plants, including Sword Fern ( Polystichum munitum), Indian Plum ( Oemleria cerasiformis), Oregon Grape ( Mahonia aquifolium), Vanilla Leaf ( Achlys spp.), English Ivy ( Hedera helix), and other herbs and forbs, contributed to a dense understory. The forest contained many downed trees with upturned roots and featured numerous natural cavities created by decay and primary cavity excavators such as woodpeckers, chickadees, and nuthatches. The Pacific Wren is a common year-round resident along this corridor. Video and Audio Recording We installed an Airlink 500W IP camera (AirlinklOl, Inc., Fremont, CA) 0.7 m from the nest, and the camera recorded video from a half hour before sunrise to approximately a half hour after sunset during nest building, court- ship, egg laying, and rearing of the young (1 April-23 May 2011). The device recorded audio sporadically through the cycle. We reduced video recording during incubation, then resumed the pre-dawn to post-dusk schedule when the chicks hatched. RESULTS Nest Construction The male (as determined by nest-building behavior) constructed the globular nest, apparently alone, for 7 days. During most of the construction, he was silent. As early as day 2, he produced vocalizations, including a chatter (http:// youtu.be/z3dWt-6z5NU) that varied in syntax and quality from his typical song. He sang both adjacent to and from within the nest. When a female arrived, the male left the nest but could be heard chattering from nearby while the female inspected the nest. The number of females that visited the nest during this stage is unknown. During nest construction, visiting birds approached by flying to the wires at the same level as the nest or landed on the wooden siding of the house and approached the nest from below. On day 57 CONSPECIFIC NEST AGGRESSION OF THE PACIFIC WREN 8 of construction, a female arrived at the partially completed nest carrying nest-lining material, indicating acceptance of the nest (Hejl et al. 2002). At this point, the nest was an open cup. When the female approached, the male left the nest but displayed nearby for about 1 minute. The female made two visits to the nest on this day with lining material, after which the male continued to work on the nest. Day 9 (9 April) was the last day that the male brought nesting material, but the female continued to line the nest and eventually to enclose the nest. The female first roosted in the nest on 14 April. During the night, she lost or plucked several feathers, some of which came out of the entry hole and rested on the outside of the nest. In the morning, the female retrieved the feathers and added them to the nest lining. Egg Laying Between 6:24 and 6:34 of 16, 17, and 18 April (4-16 minutes after sunrise), a presumed female arrived at the nest and stayed inside for ap- proximately 30 to 60 minutes. A wren returned to the nest on the evenings of 18-22 April and stayed throughout the night but did not remain at the nest through the entire day. At night, it often ejected feathers though the nest opening. Each morning, the presumed female recovered the feathers and returned them to the nest. The length of time she stayed at the nest increased each day, from 28 minutes on 16 April to 5 hours and 7 minutes on 22 April. We checked the nest on 21 April, observing three eggs, after this, we left the nest undisturbed for the remainder of the cycle. Pacific Wrens are believed to lay an egg a day, although in Britain the Eurasian Wren may skip a day (Hejl et al. 2002). If the eggs were laid every 1 or 2 days, the earliest date for the first egg was April 16 (the date that behavior suggested the first egg was laid) and the latest April 19. Incubation The female began incubating on 23 April, before the clutch was complete. She was in the nest for a total of between 7 and 8 hours during the day and stayed on the nest through the night. During daylight hours, she left the nest frequently, typically after 11 to 33 minutes of incubating. Between 18 April and 24 May, the female roosted in the nest each night except for the night of 22 May. The male visited the nest periodically, displaying and singing during incubation. While brooding the eggs, the female lost her tail and did not regrow it during the nestling period. The lack of a tail allowed the us to distinguish the parental female from the other wrens seen at the nest. Hatching and Rearing All nestlings hatched in this final clutch of four eggs. The female began removing fecal sacs on 10 May. In the Winter Wren complex, males generally feed nestlings less than the female does, although the male’s contribution increases as the nestlings age (Hejl et al. 2002). The male of this pair returned to visit the nest several times during incubation and nestling rearing, bringing food on only three occasions and successfully delivering it only twice, all on 11 May. The male 58 CONSPECIFIC NEST AGGRESSION OF THE PACIFIC WREN returned to the nest occasionally during the rearing period and could be observed looking into the nest. At no time did he attempt to enter the nest. Veiga (1990) found that in the House Sparrow ( Passer domesticus), bigamous males assist only the primary female in rearing of their young. The minimal paternal feeding of the young in the nest we video-recorded suggests that the parental male may have had another nest nearby. Conspecific Nest Aggression On 15 May, when the nestlings were 9 days old, they were attacked at length by a nonparental Pacific Wren. The three wrens visiting the nest could be distinguished by visible physical characteristics: the parental female had lost her tail and had no visible rectrices; the male had a normal tail; the intruding wren had a shortened fourth rectrix (R4) on the right side of the tail. Over 4.75 hours, the intruder attacked the nestlings seven times, the longest attack lasting 9 minutes. During the attacks, it removed two nestlings from the nest. There were several contacts between the intruder and the parental female during this period as well. The parental male was in the area and visited the nest on several occasions during and after the attacks. The chronology of the attacks was as follows: 6:19 A Pacific Wren other than the parents visited the nest. The parental female was at the nest feeding the young as the second bird approached. There was a physical altercation between the birds, after which both left the nest area. 6:24 The parental female returned with food and continued feeding the young. 6:28 The parental female returned with food but immediately left in the direction of the carport floor, possibly to chase the intruder. 6:31 The intruder returned to the nest. The young, sensing an adult at the nest, began begging for food. The intruder approached the nest opening slowly, then grabbed the nearest nestling with its foot, attempting to pull the chicks from the nest. The first nestling was able to retreat into the nest. The intruder then partially entered the nest, apparently pecking at the nestlings. Within 50 seconds, using its bill to grab the skin on the head of a nestling, the intruder was able to pull it through the opening (Figure 1). The nestling struggled and attempted to hang onto the nest with its feet as the adult pulled it away from the nest. Eventually, the nestling was supporting the full weight of the Figure 1. An intruding Pacific Wren successfully pulled a nestling from the nest by grasping the nestling’s skin in its bill. 59 CONSPECIFIC NEST AGGRESSION OF THE PACIFIC WREN adult, clinging to the nest while the adult pulled away. After a few seconds, both dropped out of view of the camera. Video of this event can be viewed at http://youtu.be/tuHcGaa4FZ4. 6:37 The parental female returned to the nest without food and immediately left. She then quickly returned with food and began feeding the nestlings again every 3 to 5 minutes. 6:47 A Pacific Wren with an intact tail, presumably the parental male, visited the nest and looked inside. The parental female arrived while it was there and did not interact with the visitor, supporting the idea that this wren was the parental male, which then flew off while the female fed the young. 6:56 The intruder returned and entered the nest. Motion in the nest suggested that the adult was pecking the young. The parental female returned within 1 minute and fought with the intruder for 30 seconds before both fell out of view of the camera (Figure 2). 6:59 The intruder returned and began removing material from the top of the nest. When the parental female returned, the intruder chased her away (Figure 3). 7:01 The intruder returned, entered the nest, and attacked the young again for about 8 minutes. Nest lining material was ejected during this attack. The intruder left the nest at 7:09. 7:10 The parental female resumed feeding the young and retrieved the ejected nest lining material on subsequent visits. 7:48 The intruder returned, entered the nest again, and attacked the young. 7:50 When the parental female returned, the intruder again chased her away and immediately returned to the nest to continue the attack. 7:54 The intruder returned and removed a second nestling from the nest (Figure 4), then departed. We found the two removed nestlings dead on the ground, one immediately below the nest and the other about 1 meter away. 7:58 The parental female returned but left without feeding the remaining young. 8:01 The parental female returned with food. While the female removed a fecal sac, the intruder was approaching the nest. 8:04 The intruder entered the nest and attacked the young. 8:07 The parental female returned and was chased off by the intruder, which also left the area. 8:10 The parental female returned with food but left immediately in the direction of the carport floor, possibly to interact with the intruder. Figure 2. The parental female (lower) pecked at the intruder’s wing when she found the intruder in the nest. The intruder and parent fought before leaving the view of the camera. 60 CONSPECIFIC NEST AGGRESSION OF THE PACIFIC WREN Figure 3. As well as attacking the young, the intruder (upper right) also began removing nest material. When the parental female returned, the intruder chased her from the nest. Figure 4. After a considerable struggle, the intruder removed a second nestling. 8:11 The intruder returned to the top of the nest. When the parental female ar- rived a few seconds later, the intruder once again chased her away and left the area. 8:14 The female returned, fed the young and settled into the nest for about 7 minutes, then resumed a schedule of feeding and brooding the young. She was absent from the nest for periods of up to 15 minutes. 8:41 A wren with a full tail, we presume the male, came to the nest, looked inside, and left. 9:12 The intruder returned to the nest and began a new attack, which continued uninterrupted for 9 minutes. Toward the end of this attack, the intruder began retrieving some of the nesting material that had been ejected and placed it in the nest. The intruder left at 9:21 9:23 The parental female returned, fed the young and stayed on the nestlings for 6 minutes. 9:29 Normal parental care (feeding forays and brooding) by the female resumed over the next hour and a half, the female staying in the nest with the young for up to 10 minutes at a time. During this period, the female left the nest 15 times, returning 1.5 to 16 minutes after leaving. 10:56 The intruder returned. The parental female was in the nest and hunkered down over the remaining nestlings. The intruder left. 10:57 The intruder returned, started climbing around the outside of the nest, then departed. 11:00 The female left the nest and within a minute the intruder returned and entered the nest for a final brief attack of about 1 minute, then left. 11:04 The parental female returned with food. 11:22 The parental male came to the nest and looked into the opening before leaving. 11:43 The parental female came to the nest with food, but left without delivering it. She returned 2 minutes later and entered the nest with the food. Feeding by the female continued throughout the day, interspersed with periods in which she remained in the nest, brooding the two remaining nestlings. 18:22 The male visited the nest, looking into the entry hole at the young. 61 CONSPECIFIC NEST AGGRESSION OF THE PACIFIC WREN Fledging of Remaining Nestlings The intruder made its final visit at 11:00 on 15 May. The nest was dam- aged during the attacks, with the entry hole significantly enlarged. There was no attempt by either parent to repair the nest. The parental female continued to attend to the remaining young, which successfully fledged on 23 May. The parental male visited the nest on several occasions between the attack and fledging. The parental female roosted in the nest on 23 May although there were no nestlings in it. The nestlings did not return. Once the nest was empty, the male returned and renovated the nest, adding more material to the exterior, including repairing and building up the edge of the opening and partially obscuring it with moss and grasses. Although he courted from the nest for several days, he did not attract a female for a second brood. We left the nest in place, and in early November two wrens began roost- ing on the top of it. They typically arrived within 1 minute of each other, sometimes virtually simultaneously. They used the site as a roost sporadically over the winter. DISCUSSION Using both video and audio recording devices, we have detailed conspecific aggression at a Pacific Wren nest on southern Vancouver Island, including the loss of two nestlings. Birds are known to destroy the eggs and nestlings of others, both heterospecific and conspecific. Infanticide by conspecific females is considered a form of resource competition (Hrdy 1979) or in- terference competition (Quinn and Holroyd 1989). The attackers do not typically receive nutrition from these attacks. Good nest sites are critical for successful breeding. In polygynous species, secondary females may attempt to displace the primary female, or vice versa, through infanticide (Hansson et al. 1997, Veiga 1990). The displacement of competitors through such behavior may increase the aggressor’s access to parental investment by the male, food resources, and nest sites by caus- ing the parental female to abandon the unsafe nest (Hansson et al. 1997). Polygynous males face a logical trade-off between a potential increase in re- productive output and exposure to nest failure through such nest destruction. Within its range, the Pacific Wren is most abundant in western Washington and southwestern British Columbia. During the breeding season, coastal habitats adjacent to the Georgia Depression support the densest populations on Vancouver Island (Wolf and Howe 1990, Campbell et al. 1997). At our study site (~1.5 hectares), Pacific Wrens were abundant, with generally two, sometimes three, singing males defending adjacent territories. Habitat quality and the density of the local Pacific Wren population could influence both the mating system (i.e. , polygyny) and level of competition for optimal nest sites and food resources. Although the Pacific Wren is usually monogamous, it may be polygynous in optimal habitats (Hejl et al. 2002). The scant attention the nestlings received from the parental male suggests that he probably had another nest nearby, possibly with the intruder. Both population density and competition for resources can influence the prevalence of infanticide (Hrdy 1979). Freed (1987) found that Rufous-and- 62 CONSPECIFIC NEST AGGRESSION OF THE PACIFIC WREN White Wrens ( Thryophilus rufalbus) remove nestlings from House Wren nests, and Alworth and Scheiber (1999) observed a female House Wren, in response to food shortage, usurp the territory of another House Wren dur- ing the nest-building stage. Food shortages can be caused by poor weather or by dense populations of birds competing for the same resources. Pacific Wrens forage on the forest floor, seeking food among leaf litter, decaying logs, and standing plants, and eat a wide variety of invertebrates including soft-bodied insects, insect larvae, and spiders (Hejl et al. 2002). At the nest we video-recorded, the female returned with food frequently, often several times within 5 minutes, suggesting that food was abundant in the area. We do not believe that the observed nest aggression was due to competition for food. As food was abundant, the need for additional support from the parental male also seems an unlikely reason for competition. The Pacific Wren’s flexibility in using a variety of natural and artificial nest sites should reduce competition for suitable locations. Under normal envi- ronmental conditions, there should be multiple nest sites appropriate for the Pacific Wren in the observed male’s territory. The spring of 2011, however, was uncharacteristically cool and wet in Victoria. There were several extended periods of heavy rain. On the day of the nest attack, the area was receiving a downpour. The observed nest was inside a carport, sheltered from rain and wind. If the intruder’s nest had failed because of the weather or depredation, she may have been seeking a more protected and secure nest site. If she re- moved the young from the sheltered nest, the parental female might abandon the site, making it available for the intruder to use. In this case, though, only two of four young were removed, and the parental female retained the nest. Although Pacific Wrens may rear more than one brood during a breeding season, neither the intruder nor the known parental female returned to use the subject nest after the two remaining young fledged. This could be because the parental female had lost young to an attacker, making it a higher-risk site for a second clutch. Interference competition may have successfully driven the parental female from the territory. Hejl et al. (2002) estimated renest- ing to begin within 4 to 6 days of a nest’s failing. Since a week had passed between the attack and the nest becoming available, the intruder may have already started another clutch, eliminating her interest in the site. On the basis of this single observation we can only speculate about the factors responsible for the attacks, but they are consistent with those of other species — in this case, competition for an optimal nest site. Video technology affords researchers the opportunity to observe behaviors that have previously been difficult to document. The use of non-invasive techniques such as video monitoring will likely continue to provide new in- sight into the behavior and life history of species that have previously been challenging to observe. The technology has its limitations, though, and it should be seen as an adjunct to, not a replacement for, more traditional observational methods. LITERATURE CITED Alworth, T., and Scheiber, I. B. R. 1999. An incident of female-female aggression in the House Wren. Wilson Bull. 111:130-132. Campbell, W. R., Dawe, N., McTaggart-Cowan, I., Cooper, J. M. , Kaiser, G. W., 63 CONSPECIFIC NEST AGGRESSION OF THE PACIFIC WREN McNall, M. C. E., and Smith, G. E. J. 1997. The Birds of British Columbia, vol. 3. Univ. Br. Columbia Press, Vancouver. Fox, L. R. 1975. Cannibalism in natural populations. Annu. Rev. Ecol. Evol. Syst. 6:87-106. Freed, L. A. 1987. Rufous-and-white Wrens kill House Wren nestlings during a food shortage. Condor 89:195-197 Hansson, B., Bensch, S., and Hasselquiest. D. 1997. Infanticide in Great Reed Warblers: secondary females destroy eggs of primary females. Animal Behavior 54:297-304. Hejl, S. J., Holmes, J. A., and Kroodsma, D. E. 2002. Winter Wren ( Troglodytes hiemalis), in The Birds of North America (A. Poole and F. Gill, eds.), no. 623. Birds N. Am., Inc., Philadelphia. Hrdy, S. B. 1979. Infanticide among animals: A review, classification, and examina- tion of the implications for the reproductive strategies of females. Ethol. Sociobiol. 1:13-40. Jones, J. S. 1982. Of cannibals and kin. Nature 299:202-203. Kroodsma, D. E., and Verner. J. 1997. Marsh Wren ( Cistothorus palustris), in The Birds of North America (A. Poole and F. Gill, eds.), no. 308. Acad. Nat. Sci., Philadelphia. Leonard, M. L., and Pieman, J. 1987. Nesting mortality and habitat selection by Marsh Wrens. Auk 104:491-495. Pieman, J. 1977a. Intraspecific nest destruction in the Long-billed Marsh Wren, Telmatodytes palustris palustris. Can. J. Zool. 55:1997-2003. Pieman, J. 1977b. Destruction of eggs by the Long-billed Marsh Wren (Telmatodytes palustris palustris ). Can. J. Zool. 55:1914-1920. Pieman, J. 1994. Evidence for interspecific egg destruction by Bewick’s Wrens. J. Field Ornithol. 65:479-481. Pieman, J., and Pieman, A. K. 1980. Destruction of nests by the Short-billed Marsh Wren. Condor 82:176-179. Pribil, S., and Pieman, J. 1991. Why House Wrens destroy clutches of other birds: A support for the nest site competition hypothesis. Condor 93:184-185. Quinn, M. S., and Holroyd, G. L. 1989. Nestling and egg destruction by House Wrens. Condor 91:206-207. Veiga, J. P. 1990. Infanticide by male and female House Sparrows. Anim. Behav. 39:496-502 Verner, J. 1975. Interspecific aggression between Yellow-headed Blackbirds and Long-billed Marsh Wrens. Condor 77:328-331. Simons, L. S., and Simons, L. H. 1990. Experimental studies of nest-destroying behavior by Cactus Wrens. Condor. 92:855-860. Wolf, A. T., and Howe, R. W. 1990. The Winter Wren in Wisconsin. Passenger Pigeon 52:103-112. Accepted 4 July 2012 64 NOTES FIRST RECORD OF A SURFBIRD IN THE HAWAIIAN ISLANDS ERIC A. VANDERWERF, Pacific Rim Conservation, 3038 Oahu Avenue, Honolulu, Hawaii 96822; ewerf@hawaii.rr.com On 9 April 2012, while I was monitoring Red-tailed Tropicbird (Phaethon rubri- cauda) nests on the cliffs near Halona Point along the southeastern coast of Oahu, Hawaii, a shorebird landed on a rocky ledge on the shoreline below me. The bird had a very distinctive tail pattern; the rectrices were bright white with broad black tips, which combined with other characters made it immediately obvious that it was a Surfbird ( Aphriza uirgata), a species not previously recorded in the Hawaiian Islands (Pyle and Pyle 2009). The bird was stocky, with short yellow legs and a short, thick (for a shorebird) bill that was orangish on the tomia and at the base of the mandible (Figure 1). The head, back, wings, and upper breast were mottled grayish brown, and the belly and lower breast were white with dark spots. It had a very prominent white wing stripe (Figure 2). I observed the bird for about an hour, during which time it moved slowly along the shoreline to the east, foraging intermittently for small invertebrates, primarily mollusks, on the rocks and in small tide pools. It was not wary and allowed me to photograph it within about 8 m. The photographs show feathers on the back, breast, and wing coverts of mixed ages. Most feathers were of the formative or basic plumages, be- ing plainer gray and having a narrow white fringe, but some of the scapulars were newer feathers of the alternate plumage, having a black center and a broader white fringe (Figure 1A). In addition, some of the wing coverts were much more worn and more brownish than others and probably were juvenal feathers (Figure IB), indicating the bird was in its first spring (O’Brien et al. 2006, Pyle 2008). The distribution of alternate and juvenal feathers was asymmetrical; the right side of the bird contained more alternate scapulars (Figure 1A), and the left side had more juvenal wing coverts (Figure IB). First-year shorebirds often undergo only a partial pre-alternate molt and do not make the northward migration, remaining on the wintering grounds during the breeding season (Johnson 1977, Johnson and Johnson 1983). The Surfbird was seen again by others and me on 10 different days between 7 April and 10 July, always within the same 300-m length of shoreline. It apparently moved over a much larger area, however, as it was not detected in this area on several occasions, and on three occasions I saw it flying to or from the west, once beyond Koko Head, a distance of 1.5 km. The bird sometimes was seen in the company of several Ruddy Turnstones (Arenaria interpres), one or two Pacific Golden-Plovers (Pluvialis fulva), and a Wandering Tattler ( Tringa incana). An obvious question regarding this record is where the bird came from. The Surfbird nests in Alaska and the Yukon and winters on mostly rocky shorelines from southern Alaska to southern Chile (Senner and McCaffery 1997, O’Brien et al. 2006). During the nonbreeding season it is very rare away from the Pacific coast of the Americas but recorded far outside its usual range, as evidenced by spring records from interior California, coastal Texas, Alberta, and Pennsylvania (Hayman et al. 1986, Senner and McCaffery 1997, Davis 2012). I had been visiting the area about once a week starting in late January to monitor tropicbird nests (VanderWerf and Young 2007), but I did not see the Surfbird before 9 April. Although the timing of the initial observation is consistent with the timing of the species’ northbound migration (Senner and McCaffery 1997), it seems unlikely that a bird of any age would leave a nonbreeding site on the mainland and fly a mini- mum of 3600 km to an island far from any known migration corridor. I think a more Western Birds 44:65-68, 2013 65 NOTES Figure 1. First-year Surfbird near Halona Point, Oahu, Hawaii. The feathers are of mixed ages, including many formative feathers (plain gray with narrow white fringe), some first-alternate scapulars (black center with broader white fringe), and a few very worn, more brownish juvenal wing coverts. The alternate scapulars were more prevalent on the bird’s right side (A, white arrows, 9 April 2012), but the left side contained more juvenal wing coverts (B, white arrows, 13 May 2012). Photos by Eric VanderWerj 66 NOTES Figure 2. Surf bird just prior to taking flight, Lanai Lookout, Oahu, 13 May 2012. Photo by Eric VanderWerf parsimonious explanation for the bird’s occurrence in Hawaii is that it came from the breeding grounds the previous summer or autumn and was simply overlooked until I saw it in April. A regular migratory pathway from Alaska to the Hawaiian Islands and beyond is facilitated in the autumn by favorable atmospheric patterns, and these tail winds are used by several shorebirds breeding in the Arctic, such as the Wandering Tattler, Bristle-thighed Curlew ( Numenius tahitiensis ), and Bar-tailed Godwit ( Limosa lapponica ; Gill et al. 2008), which occasionally are accompanied by vagrant species (Pyle and Pyle 2009). Furthermore, the bird was not observed for a 60-day period from 13 May to 10 July either, and, given the steep cliffs and high waves that make shoreline access difficult along much of the southeastern coast of Oahu (Figure 3), it easily could have been overlooked during the winter. I thank Peter Pyle for assistance in determining the bird’s plumage and age, Lance Tanino, Mike Ord, Peter Donaldson, Kurt Pohlman, David Kuhn, and Satoko Lincoln for providing information about their observations of the bird, Mike Lohr for bring- ing my camera on short notice, and Bob Gill, Peter Pyle, and Jack J. Withrow for comments on the manuscript. LITERATURE CITED Davis, A. 2012. Sightings, mid-March to mid-April 2012. Birding 44:20-22. Gill, R. E., Jr., Tibbitts, T. L., Douglas, D. C., Handel, C. M., Mulcahy, D. M., Gottschalck, J. C., Warnock, N., McCaffery, B. J., Battley, P. F., and Piersma, 67 NOTES Figure 3. Surfbird (in center foreground) near Halona Point, Oahu, Hawaii, on 10 April 2012. Though far outside the Surfbird’s usual winter range, the rocky volcanic shoreline of southeastern Oahu provides foraging habitat not unlike that found along parts of the Pacific coast of North and South America. Photo by Eric VanderWerf T. 2009. Extreme endurance flights by landbirds crossing the Pacific Ocean: Ecological corridor rather than barrier? Proc. Royal Soc. B 276:447-457. Hayman, P., Marchant, J., and Prater, T. 1986. Shorebirds: An Identification Guide to Waders of the World. Houghton Mifflin, Boston. Johnson, O. W. 1977. Plumage and molt in shorebirds summering at Enewetak Atoll. Auk 94:222-230. Johnson, O. W., and Johnson, P. M. 1983. Plumage-molt-age relationships in “over-summering” and migratory Lesser Golden-Plovers. Condor 85:406-419. O’Brien, M., Crossley, R., and Karlson, K. 2006. The Shorebird Guide. Houghton Mifflin, Boston. Pyle, P. 2008. Identification Guide to North American Birds, part 2. Slate Creek Press, Point Reyes Station, CA. Pyle, R. L., and Pyle, P. 2009. The Birds of the Hawaiian Islands: Occurrence, History, Distribution, and Status, version 1. B. P. Bishop Museum, Honolulu; http://hbs. bishopmuseum.org/birds/rlp-monograph (accessed 9 April 2012). Senner, S. E., and McCaffery, B. J. 1997. Surfbird ( Aphriza uirgata), in The Birds of North America (A. Poole and F. Gill, eds.), no. 266. Birds N. Am., Inc., Philadelphia. VanderWerf, E. A., and Young, L. C. 2007. The Red-billed Tropicbird in Hawaii, with notes on interspecific behavior of tropicbirds. Marine Ornithol. 35:81-84. Accepted 7 September 2012 68 NOTES COMMON NESTING HABITATS AND WEIGHTS AT FLEDGING OF WEDGE-TAILED SHEARWATERS ON TERN ISLAND, HAWAII PHILLIP J. HOWARD, URS Corporation, 4225 Executive Square, Suite 1600, La Jolla, California 92037; philliphoward85@gmail.com SARAH C. HARVEY, School of International Relations & Pacific Studies, University of California, San Diego, 9500 Gilman Drive, Mail Code 0519, La Jolla, California 92093-0519 PAULA L. HARTZELL, PETE LEARY, and TY J. BENALLY, U. S. Fish and Wildlife Service, 300 Ala Moana Blvd., Room 5-231, P. O. Box 50167, Honolulu, Hawaii 96850 The Wedge-tailed Shearwater {Puf firms pad ficus) is a common seabird in the Hawaiian Islands, breeding from Kure Atoll in the northwest to offshore islets near Maui in the south (Richardson 1957, Harrison 1990, Whittow 1997, Spear and Ainley 1999, Pyle and Pyle 2009). The highest concentration of breeding Wedge- tailed Shearwaters is in the northwestern Hawaiian Islands, where Pyle and Pyle (2009) estimated over 228,000 pairs. The largest colonies are on Laysan (150,000 pairs), Nihoa (35,000 pairs), and Lisianski (20,000 pairs) (Harrison 1990, USFWS 2005, Pyle and Pyle 2009). In the southeastern Hawaiian Islands, the population of the Wedge-tailed Shearwater has been estimated at roughly 67,000 pairs, with the largest colonies on O’ahu (30,550 pairs) and Ni’ihau (25,000 pairs) (VanderWerf et al. 2007, Pyle and Pyle 2009). The goals of our study were to estimate the number of Wedge-tailed Shearwaters fledging in 2010 on Tern Island, the main island in the atoll of French Frigate Shoals, in the northwestern Hawaiian Islands, to identify common nesting habitats in relation to vegetation cover and soil type, and to investigate dif- ferences between natural and artificial nests in fledglings’ weight and date of fledging. For one month (18 October-18 November 2010) we searched the entire island for chicks, checking every bush, burrow, building, pipe, and debris pile, and banding each chick with a uniquely numbered U.S. Geological Survey metal band. Additionally, we recorded information on nest type, soil type, and vegetation cover over the nest at time of capture. We categorized nests as natural burrows, natural ground nests, nest boxes/huts, other artificial sites, or unknown (Figures 1 and 2). Natural burrows consisted of burrows dug by Wedge-tailed Shearwaters in a natural setting without support or cover by man-made structures. Natural ground nests consisted of eggs laid on the bare ground, often under vegetation, without support or cover by man-made structures. Nest boxes/huts were man-made structures built for nesting of burrowing seabirds. Other artificial nests consisted of burrows or ground nests with support or cover by man-made structures. This category includes nests inside or under buildings, pipes, debris, or sidewalks. The category of “unknown” encompasses nests whose location could not be determined because the chick had become mobile and left its nest. Using stratified random sampling, we weighed 40 chicks (10 from each nest type) three times per week with a 1000-gram Pesola scale and monitored them for survival to assess differences in fledglings’ weight and departure date by nest type. We tested for statistical differences (a = 0.05) between nest types with a Mann-Whitney pairwise comparison in the program PAST, version 2.17b (Hammer et al. 2001). We banded 410 chicks in 2010 and found only two banded chicks dead before fledging, representing an estimated maximum number of 408 chicks fledged. This is 259 chicks fewer than in 2009 (667 chicks) and the lowest number of fledglings since 2005 (364 chicks) (USFWS unpublished data). Since 1990, however, the Western Birds 44:69-75, 2013 69 NOTES Figure 1. Common types of Wedge-tailed Shearwater nests on Tern Island. A, natural burrow; B, natural ground nest. 70 NOTES Figure 2. Common types of Wedge-tailed Shearwater nests on Tern Island. A, nest box/hut; B, artificial nest. 71 NOTES Wedge-tailed Shearwater population on Tern Island has increased slowly but steadily (USFWS unpublished data). As on most other islands where Wedge-tailed Shearwater nesting has been studied (Gallagher 1960, Shallenberger 1973, Byrd 1979, Byrd et al. 1983, Harrison 1990, Sievert 1996), the most common nest type on Tern Island was naturally dug burrows, which accounted for 62.0% (n = 254) of all nests. This type was followed by artificial nests at 23.2% (n = 95), natural ground nests at 7.1% (n = 29), unknown nest sites at 4.9% (n = 20), and artificial nest boxes at 2.9% (n = 12). Humus was the most common soil type for natural burrow and ground nests, accounting for 85.9% (n = 243) of all nests. Other soil types in which the shearwaters nested were fine sand and rock at 13.1 % (n = 37), all small rock at 0.7% (n = 2), and sand at 0.4% (n = 1). Although we took no standardized measurements, burrows dug in sand were apparently deeper than those in humus. No vegetation cover was the most common cover type for natural burrow and ground nests, accounting for 38.9% (n = 110) of all nests, followed by Tree Heliotrope (Heliotropium foertherianum ) at 35.3% (n = 100), Lepturus (Lepturus repens) at 21.6% (n = 61), native goosefoot (Chenopodium sp.) at 2.12% (n = 6), dead logs at 1.41% ( n = 4), and and Morning Glory ( Ipomoea pes-caprae) at 0.71% (n = 2). We found no significant differences by nest type in weight at initial banding, fledg- ing, or maximum weight achieved (Mann-Whitney pairwise comparison, P > 0.05; Figure 3). But natural burrows and artificial nests differed in the chicks’ date of fledging and minimum recorded weights. Chicks in artificial nests fledged significantly later and had lower minimum recorded weights than those in natural burrows (Figure 3). Fledgling success was 100% from all types of nests except natural ground nests, from which we recovered two dead chicks before fledging (Table 1). Proper conservation of burrow locations is essential for the long-term conservation of breeding colonies of the Wedge-tailed Shearwater, as most birds return to their natal island to breed and reuse burrows for nesting (Munro 1967, Shallenberger 1973). On Manana island, off O'ahu, out of 32 Wedge-tailed Shearwaters banded by Shallenberger (1973), 14 returned to the same burrow the following year to nest, 11 were in burrows less than 3 meters away, and 6 nested on the surface close to their burrow from the previous year. As we noted, ground nesting can be problematic for Wedge-tailed Shearwaters, as ground nests were the only nest type at which we found dead chicks. In another study on Tern Island, Sievert (1999) found ground-nesting Wedge-tailed Shearwaters to have a success rate of only 2% because they were more exposed to a hot microclimate than were chicks in shaded burrows. Exposure to the tropical sun on Tern Island increases incubating adults’ stress from heat and water loss, making the adult more likely to abandon the egg during the day (Howell and Bartholomew 1961, Sievert 1996). For conservation of the Wedge-tailed Shearwater on Tern Island, we recom- mend maintenance of existing nesting burrows and locations. Efforts should focus on conservation of areas with humus soil type and revegetation of bare areas with Tree Heliotrope and Lepturus. Conserving and restoring nesting habitat for the Wedge-tailed Shearwater may in turn help conservation efforts for the Tristram’s Storm-Petrel ( Oceanodroma tristrami; McClelland et al. 2008), a near-threatened (BirdLife International 2008) species of conservation concern (USFWS 2002), by reducing burrow-nesting seabirds’ competition for nest sites (McClelland et al. 2008). This study was made possible by funding from the U.S. Fish and Wildlife Service. We especially thank Keith Burnett, Patricia Jackson, Dan Rapp, Sarah Youngren, and all other Tern Island biologists who helped record data for this study. The find- ings and conclusions in this article are those of the authors and do not necessarily represent the views of URS, the University of California, San Diego, or the U.S. Fish and Wildlife Service. 72 NOTES 27-Nov 25-Nov 23-Nov 21 -Nov 19-Nov 17-Nov 15-Nov 1 3-Nov NB* Figure 3. Comparison by nest type of weights and dates of departure of Wedge-tailed Shearwater fledglings on Tern Island, French Frigate Shoals, 2010. NB, natural burrows; NG, natural ground nests; ANB, artificial nest boxes; AN, artificial nests. Asterisks indicate a significant difference between means of fledglings in artificial nests and natural burrows (Mann-Whitney pairwise comparison, P = 0.037) and in date of fledging (Mann-Whitney pairwise comparison, P = 0.037). LITERATURE CITED BirdLife International. 2012. Species factsheet: Oceanodroma tristrami; www. birdlife.org (21 October 2012). Byrd, G. V. 1979. Artificial nest structures used by Wedge-tailed Shearwaters on Kaua’i. ‘Elepaio 40:10-12. Byrd, G. V., Moriarty, D. I., and Brady, B. G. 1983. Breeding biology of Wedge-tailed Shearwaters at Kilauea Point, Hawaii. Condor 85:292-296. Gallagher, M. D. 1960. Bird notes from Christmas Island, Pacific Ocean. Ibis 102:489-502. Hammer, O., Harper, D.A.T., Ryan, and P. D. 2001. PAST: Paleontological statistics software package for education and data analysis. Palaeontologia Electronica 4(1); http://palaeo-electronica.org/2001_l/past/issuel_01.htm. Harrison, C. S. 1990. Seabirds of Hawaii: Natural History and Conservation. Cornell Univ. Press, Ithaca, NY. 73 NOTES Table 1 Weights and Dates of Departure of Wedge-tailed Shearwater Fledglings on Tern Island, French Frigate Shoals, 2010 Weight (g) Date of fledging Nest At banding Minimum Maximum At fledging Natural burrows NB1 450 395 575 395 22 Nov NB2 445 445 560 450 15 Nov NB3 515 435 650 435 15 Nov NB4 530 430 665 430 19 Nov NB5 565 452 600 452 26 Nov NB6 590 510 650 510 15 Nov NB7 635 475 635 475 17 Nov NB8 460 460 600 490 12 Nov NB9 540 435 590 435 17 Nov NB10 380 330 550 372 26 Nov Mean 511 437 608 444 18 Nov Natural ground nests NG1 510 450 530 450 17 Nov NG2 460 420 575 Found dead 19 Nov NG3 525 362 615 362 24 Nov NG4 520 500 615 500 15 Nov NG5 NG6 495 405 Found dead 18 Oct 600 405 19 Nov NG7 430 430 640 450 17 Nov NG8 440 390 620 390 22 Nov NG9 470 392 535 392 26 Nov NG10 555 380 580 380 19 Nov Mean 489 414 590 416 19 Nov Artificial nest boxes ANB1 590 375 590 375 22 Nov ANB2 420 381 585 402 11 Dec ANB3 530 520 610 565 5 Nov ANB4 470 400 590 400 17 Nov ANB5 465 372 620 372 4 Dec ANB6 500 451 620 451 6 Dec ANB7 655 465 660 465 8 Nov ANB8 560 405 570 405 19 Nov ANB9 590 450 690 450 22 Nov ANB10 575 361 590 361 30 Nov Mean 536 418 613 425 23 Nov Artificial nests AN1 450 312 585 312 2 Dec AN2 505 417 625 417 24 Nov AN3 455 412 575 412 26 Nov AN4 555 382 605 382 4 Dec AN5 405 302 465 302 2 Dec AN6 480 425 650 425 22 Nov AN 7 585 475 630 530 12 Nov AN8 480 340 620 340 17 Nov AN9 590 430 670 445 22 Nov AN10 550 432 650 432 26 Nov Mean 506 393 608 400 25 Nov 74 NOTES Howell, T. R., and Bartholomew, G. A. 1961. Temperature regulation in nesting Bonin Island Petrels, Wedge-tailed Shearwaters, and Christmas Island Shearwa- ters. Auk 78:343-354. McClelland, G. T. W., Jones, I. L., Lavers, J. L., and Sato, F. 2008. Breeding biology of Tristram’s Storm-Petrel Oceanodroma tristrami at French Frigate Shoals and Laysan Island, Northwest Hawaiian Islands. Marine Ornithol. 36:177-183. Munro, G. C. 1967. Birds of Hawaii. Charles E. Tuttle, Rutland, VT. Pyle, R. L., and Pyle, P. 2009. The Birds of the Hawaiian Islands: Occurrence, History, Distribution, and Status, version 1. B. P. Bishop Museum, Honolulu; http://hbs. bishopmuseum.org/birds/rlp-monograph (6 October 2011). Richardson R. 1957. The breeding cycles of Hawaiian seabirds. B. P. Bishop Mus. Bull. 218. SAS Institute, Inc. 2001. JMPIN, version 4.0.4. SAS Institute, Inc., Cary, NC. Sievert, P. R. 1996. Water and energy balance constraints on the nesting ecology of marine birds. Ph.D. dissertation, Univ. Pa., Philadelphia. Shallenberger, R. J. 1973. Breeding biology, homing behavior and communication patterns of the Wedge-tailed Shearwater Puffinus pacificus chlororhynchus . Ph.D. dissertation, Univ. Calif., Los Angeles. Spear, L. B., and Ainley, D.G. 1999. Seabirds in southeastern Hawaiian waters. W. Birds 30:1-32. U.S. Fish and Wildlife Service (USFWS). 2002. Birds of conservation concern 2002. U.S. Fish and Wildlife Service, Arlington, VA. U.S. Fish and Wildlife Service (USFWS). 2005. Seabird conservation plan, Pacific region. U.S. Fish and Wildlife Service, Migratory Birds and Habitat Programs, Pacific Region, Portland, OR. VanderWerf, E. A., Wood, K. R., Swenson, C., LeGrande, M., Eijzenga, H., and Walker, R. L. 2007. Avifauna of Lehua Islet, Hawaii. Conservation and value management needs. Pacific Science 61:39-52. Whittow, G. C. 1997. Wedge-tailed Shearwater ( Puffinus pacificus), in The Birds of North America (A. Poole and F. Gill, eds.), no. 305. Acad. Nat. Sci., Philadelphia. Whittow, G. C. 1979. The effects of heavy rain on nestling Wedge-tailed Shearwaters (Rabbit Island). ‘Elepaio 39:138-139. Accepted 9 November 2012 75 NOTES AN APPARENT LONG-DISTANCE FLIGHT BY A DUSKY GROUSE IN MONTANA RONALD J. KIENHOLZ and ALEEN M. KIENHOLZ, 280 North Low Bench Road, Gallatin Gateway, Montana 59730; plover4rjk@aol.com Two closely related species (previously regarded as several subspecies of the Blue Grouse, see Zwickel and Bendell 2005) constitute the genus Dendragapus, the Dusky Grouse (D. obscurus) of inland mountains and the Sooty Grouse (D. fuliginosus ) of coastal mountains. The ability of these birds to fly long distances is poorly known because observations of such behavior are at best serendipitous. In eastern Oregon, Anthony (1903) witnessed flights of both adult and immature Dusky Grouse from a mountain ridge to a nearby mountain slope and estimated the distance at “fully a mile and a half” (2.4 km). He noted that the flights were of “gradually descend- ing” trajectory and seldom sustained enough “to carry the birds to the top” of the mountain to which they were flying even though the latter was 400 feet lower than the ridge. In coastal British Columbia, Zwickel and Bendell (2004, 2005) concluded that level flight in excess of approximately 2 km was unlikely for the Sooty Grouse because “few islands more than approximately 2 km from a source population are inhabited.” Furthermore, these authors described an instance “in which a hen that flew out over a lake came down in the water after approximately 150 m.” To our Figure 1 . The probable path of a Dusky Grouse flying across the Gallatin River Canyon. Assuming relatively level flight, the bird likely launched somewhere near point A, elevation 2288 m. It then passed point B (where we were standing, elevation 2303 m) and continued onward for an uncertain distance. The upper portion of the figure shows topography between points A and B, with vertical lines at intervals of 0.5 km. 76 Western Birds 44:76-77, 2013 NOTES knowledge, there are no additional reports concerning the ability of either species to fly significant distances. Here, we describe the flight of a Dusky Grouse that appeared to be substantially longer than these previous estimates. Our observation was made on the morning of 18 September 2008 from the crest of an unforested ridge in Gallatin County, southwestern Montana (Figure 1). The ridge is situated above the canyon of the Gallatin River, and the elevation at the river’s level is approximately 1642 m. Weather conditions at the time provided unlimited visibility under full sun, and there was no wind. Initially, we noticed an unidentified bird flying toward us from the west in apparent sustained level flight. When first seen the bird was at a distance of about 2 km and on a plane essentially even with our elevation. We assumed it was a raptor, but as the bird got closer the flap-and-glide flight pattern of a grouse became apparent. When the grouse neared us, it flared, gained a few meters of altitude, passed over the ridge on which we were standing, and began a downslope glide out of sight to the northeast. Given that long ascending flight by this species is very unlikely (Caswell 1954, Zwickel and Bendell 2004, 2005, pers. obs.), the point from which the bird launched was presumably at a relatively high elevation on the west side of the canyon. Thus the shortest possible distance for a level flight across the canyon at our elevation was approximately 4.6 km (Figure 1). The bird’s path may have been along a gradual descent. If so, the grouse would have taken flight from a higher elevation, and the distance traversed would then exceed 5 km. Our observation suggests that Dusky Grouse are capable of longer sustained flights than previously thought. Notably, flights of >10 km are relatively common among the near relatives of this species, the Sharp-tailed Grouse ( Tympanuchus phasianellus; Connelly et al. 1998), Greater Prairie-Chicken (' Tympanuchus cupido; Johnson et al. 2011), and Greater Sage-Grouse ( Centrocercus urophasianus; Schroeder et al. 1999). It seems likely that flights such as the one we describe played a role in the Dusky Grouse’s colonization of isolated mountain ranges such as the Big Snowy and Judith Mountains in central Montana. We thank Oscar W. Johnson for his extensive review and assistance with text and graphics. Fred Zwickel and James Nygard offered helpful comments preparatory to the writing of this paper. We are grateful to reviewers Michael Schroeder and Daniel D. Gibson for constructive suggestions that improved the manuscript. LITERATURE CITED Anthony, A. W. 1903. Migration of Richardson’s Grouse. Auk 20: 24-27. Caswell, B. A. 1954. A preliminary study on the life history of the Blue Grouse in west central Idaho. M.S. thesis, Univ. Idaho, Moscow, ID. Connelly, J. W., Gratson, M. W., and Reese, K. P. 1998. Sharp-tailed Grouse (Tym- panuchus phasianellus ), in The Birds of North America (A. Poole and F. Gill, eds.), no. 354. Birds N. Am., Inc., Philadelphia. Johnson, J. A., Schroeder, M. A., and Robb, L. A. 2011. Greater Prairie-Chicken (' Tympanuchus cupido), in The Birds of North America Online (A. Poole, ed.), no. 36. Cornell Lab Ornithol., Ithaca, NY; http://bna.birds.cornell.edu/bna/ species/036. Schroeder, M. A., Young, J. R., and Braun, C. E. 1999. Greater Sage-Grouse ( Cen- trocercus urophasianus ), in The Birds of North America (A. Poole and F. Gill, eds.), no. 425. Birds N. Am., Inc., Philadelphia. Zwickel, F. C., and Bendell, J. F. 2004. Blue Grouse: Their Biology and Natural His- tory. NRC Research Press, Ottawa, Canada. Zwickel, F. C., and Bendell, J. F. 2005. Blue Grouse ( Dendragapus obscurus ), in The Birds of North America Online (A. Poole, ed.), no. 15. Cornell Lab Ornithol., Ithaca, NY; http://bna.birds.cornell.edu/bna/species/015. Accepted 15 October 2012 77 NOTES MELANISTIC ADULT MALE NORTHERN HARRIER WINTERING IN IDAHO ROBERT A. MILLER, NEIL PAPROCKI, and ELIZABETH H. URBAN, Idaho Bird Observatory and Raptor Research Center, Department of Biological Sciences, Boise State University, 1910 University Drive, MS1515, Boise, Idaho 83725; RobertMiller 7 @u . boisestate . edu Melanism (dark coloration) is a condition resulting from a greater than normal expres- sion of the eumelanin pigments in the plumage (Gill 1990). The dark coloration can be advantageous to raptors by increasing the feathers’ resistance to bacterial degradation (Goldstein et al. 2004). Conversely, abnormally dark pigmentation can reduce success in pairing by disguising key species-identification cues (Garcia 2003) and decrease lifetime reproductive success by increasing mortality (Kruger and Lindstrom 2001). Polymorphism in color, of which melanism is one example, occurs in at least 3.5% of avian species worldwide and in 22% of raptors of the family Accipitridae (harriers, hawks, eagles, kites, and Old World vultures; Galeotti et al. 2003). Galeotti et al. (2003) reported that polymorphism is more prevalent in species that are terrestrial, occupy diverse habitats, carnivorous, active day and night, or nomadic, suggesting that polymorphisms are under selective pressure. Mundy et al. (2004) traced melanism in the Parasitic Jaeger ( Stercorarius parasiti- cus) and Snow Goose (Chen caerulescens) to mutations within the melanocortin-1 receptor (MC1R) gene, finding the degree of melanism is correlated with the number of copies of the mutated allele. Manceau et al. (2010) illustrated how convergent phenotypes, such as those due to melanism, can arise via different mechanisms even in closely related species. It would therefore be unwise to assume melanism occurs the same way in all raptors, even though the selective pressures may be similar. We report a new observation of a melanistic Northern Harrier ( Circus cyaneus hudsonius ) in southwest Idaho. While surveying for raptors from a vehicle we observed a melanistic adult male Northern Harrier (Figure 1) on 9 January 2012 at 13:23 at ap- proximately 43.358° N, 116.314° W, along the boundary of the Morley Nelson Snake River Birds of Prey National Conservation Area and the U.S. National Guard’s Orchard Training Area. The bird was perched on the ground in a dense stand of Big Sagebrush ( Artemisia tridentata ; >25% ground cover) approximately 5 m west of the road. As our vehicle approached, the bird flushed from the ground and flew west approximately 100 m to land once again within dense sagebrush. The bird lacked the distinctive white rump patch expected on all age and sex classes of this species (Figure 1; Smith et al. 2011). For definitive identification and photographs we approached the bird again. We observed the low, buoyant flight style and strong dihedral expected for this species (Clark and Wheeler 2001). Characteristics identifying the bird as an adult male included black wing tips, dark trailing edge to the wing, and a black-tipped tail (Clark Figure 1 . Melanistic adult male N orther n Harrier in southwestern Idaho , 9 January 2012. Photos by Robert A. Miller 78 Western Birds 44:78-79, 2013 NOTES and Wheeler 2001). The bird had a yellow iris, characteristic of an adult (Hamerstrom 1968 via Smith et al. 2011). Simmons (2000) reported melanism in at least eight of the world’s 16 species of harriers, including the Northern Harrier and its Old World representative, the Hen Harrier (C. c. cyaneus). However, the distribution of melanistic harriers is generally localized and often near the edge of the species’ natural range (Simmons 2000). We are aware of three previous reports of melanistic Northern Harriers: an adult male in California (Howell et al. 1992), an adult female in Montana (Olson and Osborn 2000), and a juvenile male in Utah (Liguori 2009). Most likely, the bird we observed had developed a new mutation within the MC1R gene or related pathway, resulting in its unique plumage (Mundy et al. 2004). Alternatively, genes for melanism may be maintained within a small fraction of the western population of the Northern Harrier, consistent with the three previous observations. The bird might even have been a long-distance vagrant from a population where melanism is more prevalent, though no such population is known. This observation was made while we were working on a project supported by a Bureau of Land Management Challenge Cost Share Award, the Boise State Raptor Research Center, the Great Basin Landscape Conservation Cooperative, and a Graduate Science, Technology, Engineering, and Mathematics Fellows in K-12 Education Fellowship from the National Science Foundation to Paprocki (DGE-0742554). We thank J. Holderman and K. Warner for logistical support and J. Heath and B. Clark for comments on the manuscript. LITERATURE CITED Clark, W. S., and Wheeler, B. K. 2001. A Field Guide to Hawks of North America. Houghton Mifflin, Boston. Galeotti, P., Rubolini, D., Dunn, P. O., and Fasola, M. 2003. Colour polymorphism in birds: Causes and functions. J. Evol. Biol. 16:635-646. Garcia, J. T. 2003. Are simple plumage traits sufficient for species discrimination by harrier males? J. Avian Biol. 34:402-408. Gill, F. B. 1990. Ornithology. W. H. Freeman, New York. Goldstein, G., Flory, K. R., Browne, B. A., Majid, S., Ichida, J. M., Burtt, E. H. Jr., and Grubb, T. C. Jr. 2004. Bacterial degradation of black and white feathers. Auk 121:656-659. Hamerstrom, F. 1968. Ageing and sexing harriers. Inland Bird Banding News 40:43-46. Howell, S. N. G., Webb, S., Sibley, D. A., and Prairie, L. J. 1992. First record of a melanistic Northern Harrier in North America. W. Birds 23:79-80. Kruger, O., and Lindstrom, J. 2001. Lifetime reproductive success in Common Buzzard, Buteo buteo: From individual variation to population demography. Oikos 93:260-273. Liguori, J. 2009. Distant raptors. Birding 41:74-76. Manceau, M., Domingues, V. S., Linnen, C. R., Rosenblum, E. B., and Hoekstra, H. E. 2010. Convergence in pigmentation at multiple levels: Mutations, genes and function. Philos. Trans. Royal Soc. B 365:2439-2450. Mundy, N. I., Badcock, N. S., Hart, T., Scribner, K., Janssen, K., and Nadeau, N. J. 2004. Conserved genetic basis of a quantitative plumage trait involved in mate choice. Science 303:1870-1873. Olson, C. V., and Osborn, S. A. H. 2000. First North American record of a melanistic female Northern Harrier. J. Raptor Res. 34:58-59. Simmons, R. E. 2000. Harriers of the World: Their Behaviour and Ecology. Oxford Univ. Press, Oxford, England. Smith, K. G., Wittenberg, S. R. , Macwhirter, R. B., and Bildstein, K. L. 2011. Northern Har- rier ( Circus cyaneus), in The Birds of North America Online (A. Poole, ed.), no. 210. Cornell Lab Ornithol., Ithaca, NY; http://bna.birds.cornell.edu/bna/species/210. Accepted 18 September 2012 79 IN MEMORIAM GALE MONSON, 1912-2012 For many western ornithologists the most prominent event of 2012 was not the centennial celebrations of the state- hood of Arizona and New Mexico. It was the passing of Gale Monson, who would have celebrated his centennial year on 1 August 20 12. The patriarch of Arizona ornithology died peacefully on 19 February 2012 in Albuquerque with one of his daughters and his caregiver at his side. In addition to the 1964 classic The Birds of Arizona, which he co-authored with Allan Phillips and Joe Marshall, and the 1981 Annotated Checklist of the Birds of Arizona, co-authored with Al- lan, Gale’s written contributions number about 192 articles, mostly on birds, but a significant number also on another animal on which he was an authority, the desert bighorn sheep. But the writ- ten word was only part of his legacy, and perhaps not his greatest contribution. From his home state of North Da- kota with a B.S. in biology he came to Arizona in July of 1934 for his first professional job, surveying ranges for grazing on the Papago (Tohono O’odham) Indian Reservation for the U.S. Bureau of Indian Affairs. After that stint he worked for the U.S. Soil Conservation Service, surveying wildlife around Arizona until hiring on with the U.S. Fish and Wildlife Service in 1940. Between then and 1946 he spent half his time managing wildlife refuges in Arizona and New Mexico and the other half in the U.S. Army, receiving a Bronze Star in 1945 for service in Kunming, China. From 1946 to 1962 Gale managed four national wildlife refuges in southwestern Arizona: Havasu, Kofa, Cabeza Prieta, and Imperial. During his last seven years of federal service he was in Washington, D.C., working on issues affecting refuges. After his retirement from the U.S. Fish and Wildlife Service in October 1969 he and his wife, Sally, moved to Tucson. Settling in Tucson afforded Gale an opportunity to use his professional skills along with his gentle and giving nature for the good of the wildlife of Arizona, as well as for all the people fortunate enough to cross his path. Gale’s knowledge of Arizona’s wildlife was frequently tapped by local conservation leaders seeking his expertise for their causes and by local amateur naturalists who just liked getting out with him. Gale’s life spanned an important period in Arizona ornithology. Fortunately, his constitution was a wonderful match for the age. His science acumen was shaped to a large degree by his youthful days immersed in the nature of his family farm in North Dakota. This experience set in him a tone of reverence for nature that was evident when he stepped afield, and likely rivaled that of John Muir. In his early teens he was influenced by a traditional field ecologist, O. A. Stevens, and later by the developing science of wildlife management and Aldo Leopold. These inspirations along with his Gale Monson on Aztec Peak, Sierra Ancha, central Arizona, 1989. 80 Western Birds 44:80-81, 2013 IN MEMORIAM intellect and work ethic made him an outstanding naturalist who observed and recorded with accuracy and detail the natural world around him. During his later years Gale witnessed a dramatic increase in public interest in birds, from casual birdwatchers to budding scientists wanting to dig deeper into birds’ lives and habits. A deluge of numerous unknown individuals reporting species in new places had Gale scrambling to ensure that the integrity of Arizona’s avian records was not being compromised. In time he embraced this new guild of birder, but it was not without very close scrutiny and frequent expressions of doubt. But in the end it was his ability to share nature with others and to be open to new possibilities that bridged the generations of ornithologists. Gale’s retirement job as weekend supervisor at the Arizona-Sonora Desert Museum afforded many people a chance encounter with this friendly distinguished gentleman. I was the recipient of one of these happenstances in February 1973 after trapping and returning an Aplomado Falcon that had been stolen from the museum. I will always remember the unselfish interest he took in my studies of raptors, helping me secure funds from Tucson Audubon Society with a personal visit to the home of Edward Chalif, then president of the society. That first encounter with Gale was the beginning of three decades of field adventures, phone conversations, and letters... my relationship with the mentor of a lifetime! I learned so much about Gale from reading his journals — not, unfortunately, until last August, after his passing. They document 70 years of his life, from age 12 through 92. He wrote in books of various shapes and sizes, a literature that took up six feet of bookshelf and weighs 178 pounds (yes, I weighed the totality). Although not part of Gale’s 192 scientific publications, they offer a unique glimpse of the origins and maturity of a man. . .a carefree boy, a student and career man, a father and husband. . . the span of an incredible life. The entire collection of Gale’s journals is now housed at the Special Collections Library, University of Arizona. The centennial birthday party last August so many of us were anticipating was not meant to be, so on 18 March 2012 about 100 friends and family gathered in Tucson to share memories. Since there was not enough time for all attendees to share their tales about Gale, Bill Broyles and I decided to gather the stories into a book. 1 The 243-page volume is a collection of stories from 50 contributors about one of the most intriguing and influential naturalists of our time. This past fall I found myself wondering about Gale’s greatest contribution. After reviewing his personal journals, reading all the stories about Gale in the lives of others for the book, remembering the times we walked in nature together and talked about birds and life... after all this, I am left with the dear notion that his written scientific contributions are dwarfed by his wonderful ability to see a bigger picture, to embrace the human lives around him, to share his natural world, and in doing so spawn a legion of associates who carry on with the elements of life that fascinated him. He was the model mentor. Richard L. Glinski 1 Counting Birds with Gale Monson by Bill Broyles and Richard L. Glinski is available from the Arizona Field Ornithologists; please visit their website (www.AZFO.org) or contact Janet Witzeman (jlwitzeman@aol.com). 100% of the money raised by the sale of this book goes to the Gale Monson Research Grant fund established by Arizona Field Ornithologists to assist studies of Arizona’s avifauna. 81 BOOK REVIEW The Young Birder’s Guide to Birds of North America, by Bill Thompson III. 2012. Houghton Mifflin, New York. 368 pp., ~450 color photographs, 300 line drawings. Paperback, $15.95. ISBN 978-0-547-44021-7. The latest addition to the Peterson Field Guides series, The Young Birder’s Guide to Birds of North America is written especially for children 8-12 years old. In many ways it is a simplified field guide, aiming to ease kids into birding without overwhelm- ing them. The real success of this book, however, is in Thompson’s light-hearted writing style, which kids will find fun, engaging, and full of tidbits of information. The value of this book is not as much to help aspiring young birders become great at bird identification but to ignite a spark of interest in birds and bird watching. The original version of this book was published in 2008 and limited in scope to eastern North America. For this version, 100 species from western North America were simply added to give it a broader geographic range. I wonder why the publisher chose to create a guide to all of North America rather than creating a new regional guide for western North America. However, given that this is meant as a first field guide for kids, the important birds to cover are those most likely to be seen in back- yards, towns, and local parks. Considering this book covers only 300 species, this is achieved remarkably well. The Young Birder’s Guide is dense with information, with a full page for each species (with the exception of a handful of similar, closely related species that share a page) arranged in taxonomic order. This layout, combined with the lack of a “quick index” page pointing to the major groups of birds, will likely make it difficult for the user to find and identify an unknown bird. Still, each page is attractive and fun to read. Every species account includes one or two photographs, which are generally well chosen. An exception is the photograph for the female Costa’s Hummingbird (p. 178), which shows a female Anna’s Hummingbird. Line drawings by Julie Zickefoose and Michael DiGiorgio are beautiful, fun to look at, and often informative. The best parts of the species accounts are the “Wow!” inserts, containing some random tidbit of information about each species. Some of them are just for laughs; for example, the insert for the Cinnamon Teal (p. 61) informs the reader that the spice does not, in fact, come from ground-up ducks. Most, however, are informative and fascinating. The range maps used are the standard Peterson maps, which are fundamentally flawed in that they do not show range during migration. Most young readers, like most adult field-guide users, will skip right to the species accounts and begin flipping through the pages. One hopes, however, that the young birder whose interest has been sparked by the species accounts will eventually turn to the 38 pages of introductory material. These sections are filled with great tips on getting started, how to look at a bird, how to use binoculars, how to get involved, etc. The back content is useful as well, including a resources page, a very nice glos- sary, and an index. In order to field-test this guide with its intended audience, I lent it to a small sample of kids. The group of 5- to 7-year-olds, not surprisingly, flipped through looking at the pictures and happily pointed out the birds they knew. I also handed it to a group around 11-13 years old, none of whom had shown the least interest in nature, and they used it in a surprising way. They hardly seemed to notice the photograph but flipped through, reading the captions for the line drawings and the little “Wow!” blurbs. I could see a new world opened to those kids, and they were excited to learn about something completely different. One girl spent about 20 minutes going through the book page by page, enthralled, and remaining oblivious when her friends moved on to other topics of conversation. 82 Western Birds 44:82-83, 2013 BOOK REVIEW This, then, is the strength of this book. Thompson’s fun writing style, great pictures, and interesting facts are all perfectly tailored to get kids interested in birds. This is a book for kids who have never thought to look at birds and kids who are just getting started in birding. Once the beginner gets a little more advanced, and starts to wonder why he or she can’t identify all the birds in the neighborhood by using this book, then it’s time to move on to another field guide. By then, it will have accomplished exactly what it aimed to do: create a new birder. Lauren Harter THANK YOU TO OUR SUPPORTERS The board of Western Field Ornithologists and the editorial team of Western Birds thank the following generous contributors who gave to WFO’s publication, scholar- ship, and general funds in 2011. The generosity of our members in sustaining WFO is an inspiration to us all. Larry Allen, Rosemead, CA Pasadena Audubon Society, Altadena, CA Paul Baicich, Oxon Hill, MD Richard Banks, Alexandria, VA Carol Beardmore, Phoenix, AZ David Bell, Altadena, CA Gordon Cameron, Mendham, NJ Eugene Cardiff, Rialto, CA William Carter, Ada, OK Sally Cedarblade, San Francisco, CA Theodore Chase, Jr., Princeton, NJ Jon Dunn, Bishop, CA Ted Eliot, Sonoma, CA Sylvia Gallagher, Huntington Beach, CA Raymond Hadley, Ester, AK Ed Harper, Carmichael, CA John Harris, Oakdale, CA Jordan Harrison, La Grande, OR Gjon Hazard, Encinitas, CA Diana and William Herron, Flagstaff, AZ Richard Jeffers, Santa Clara, CA Tom Jervis, Santa Fe, NM Ron Kienholz, Gallatin Gateway, MT Sandy Koonce, Redlands, CA Kevin Kritz, Denver, CO Dave Krueper, Corrales, NM Bruce LaBar, Tacoma, WA Arthur Langton, Canoga Park, CA Mark Leggett, Palm Springs, CA Paul Lehman, San Diego, CA Carl Lundblad, Moscow, ID Guy McCaskie, Imperial Beach, CA Robert McNab, Laguna Niguel, CA Heather Medvitz, Calabasas, CA Martin Meyers, Truckee, CA Steven Mlodinow, Longmont, CO Joseph Morlan, Pacifica, CA David Nelson, San Rafael, CA Mike Newlon, Berkeley Heights, NJ Robert Polkinghorn, Orange, CA Dolores and William Pollock, Santa Barbara, CA C. E. Probst, Kula, HI Terry Reeser, Chino Hills, CA Don Reinberg, Mill Valley, CA Jeanne Ridgley, Groveland, CA Jim Royer, Los Osos, CA Ken Schneider, San Francisco, CA David Seay, La Jolla, CA Marius Seritan, Davis, CA Dave Shuford, Inverness, CA Marjorie Siegel, Mill Valley, CA Marvin and Carola Sohns, San Rafael, CA Jay Sooter, Eureka, CA Kevin Spencer, Klamath Falls, OR Brian Sullivan, Carmel Valley, CA Jim Tietz, Shaver Lake, CA Lew Ulrey, Boise, ID Philip Unitt, San Diego, CA Debbie Van Dooremolen, Henderson, NV Janine Watson, San Francisco, CA Bright Winn, San Francisco, CA Jay Withgott, Portland, OR Francisco Wong, Del Mar, CA Tom Wurster, Garden Grove, CA 83 WFO PRESIDENT S MESSAGE: STATE OF THE ORGANIZATION The beginning of my term seems an ideal time to report on the overall health of WFO and on our plans for the future. During the past year, WFO’s board of directors commissioned an independent review of our finances and also took a close look at how well we are positioned to meet our primary goal of promoting field ornithology throughout our region. As part of that process, we spent an entire day prior to the Petaluma conference reviewing the status of the organization and beginning work on long-term and short-term plans. The key conclusions from the financial review and our planning session: 1. WFO’s financial health is good. We have adequate cash on hand, and our finances are well managed. 2. While our annual membership dues cover the costs of publishing Western Birds, we depend on income from annual conferences, special field trips, and donations to cover the rest of the organization’s expenses. 3. Our membership level is stable, but we feel we are reaching only a small fraction of the people whose interests should make them ideal WFO members. 4. We should do more to promote Western Birds as a unique and welcoming outlet for publications on field ornithology. 5. To adapt to the digital era, we ought to offer electronic (in addition to physical) versions of our publications (e.g., Western Birds, Rare Birds of California). Over the course of the year I will share with you the specific actions WFO will take to address these issues. As one example of the changes we are making, we decided to reward our members by giving them first access to spots on our special field trips and early registration for our annual conferences The most popular workshops and field trips at our recent conferences filled in the first few weeks of registration, and our special field trips to Cuba (2010 and 2013) and to the south-central Sierra Nevada (2011) filled within days. From now on, WFO members will get early access to all such events. Life members and patrons will be able to register one week in advance of other WFO members, and other WFO members will be able to register one week in advance of the general public. Of course, this early access is dependent on our having current email addresses for members. If you are not sure we have your current email address, please send it to me (erpfromca@aol.com). We are also looking at a number of ways to reach out to people who would benefit from joining WFO but may not know about us. That’s where all of you can help. Send me your ideas about reaching these potential members. Best of all, take the time to tell your friends and acquaintances about us. When I meet people I think might be interested in WFO, I give them my 30-second “elevator speech,” the gist of which is that WFO is for scientists with an interest in birding or field ornithology, and for birders with an interest in the science of birds. I look forward to the coming year and hope to see many of you at our next annual conference (Olympia, Washington, 22-25 August). Ed Pandolfino 84 Western Birds 44:84, 2013 FEATURED PHOTO CAROTENISM IN THE HAIRY WOODPECKER RON LeVALLEY, P. O. Box 332, Little River, California 95456; ron@madriverbio . com JEFF N. DAVIS, Colibri Ecological Consulting, 11238 N. Via Trevisio Way, Fresno, California 93730; jdavis@colibri-ecology.com On 3 January 2009, at Russian Gulch State Park in Mendocino County, California, Le Valley encountered a female Hairy Woodpecker ( Picoides villosus) with yellow outer rectrices. He and others hypothesized that the bird had acquired the yellow adventitiously through staining. Yet on 1 November 2010, Le Valley was surprised to find what appeared to be a different female Hairy Woodpecker with yellow outer tail feathers at Little River, approximately 6.5 km south of the previous sighting. This bird remained in the area for much of the winter and was photographed in November 2010 and January 2011 (upper photo on this issue’s inside back cover). On 25 December 2011, a male Hairy Woodpecker with yellow outer tail feathers was photographed at the same location. When LeValley spread the word about these birds he received reports of a number of other sightings in Mendocino County. John Sterling had seen a bird matching this description at MacKerricher State Park on 18 March 2010. Becky Bowen observed a pair of Hairy Woodpeckers, both showing yellow outer tail feathers, along a trail between Highway 101 and Virgin Creek Beach on 2 January 2011. LeValley photo- graphed the male on 2 January 2011 (lower photo on this issue's back cover), and multiple observers reported seeing it and/or the female on various dates through 2012. Other reports came from the Navarro River estuary, where Tim Bray photo- graphed a male near a nest cavity on 19 April 2011 and Matt Coleman photographed a female on 28 April 2011. LeValley received reports from other coastal counties in California as well. Sean Brophy photographed a yellow-tailed male in Carlotta, Humboldt County, on 26 December 2010; Beth and Tom Hamel found another, a female that had been pres- ent at least since 12 March 2011 (J. Morlan pers. comm.), at Golden Gate Park, San Francisco, on 26 December 2012 (photo on this issue's outside back cover); Steve Shunk photographed one at Point Lobos State Park, Monterey County, on 8 January 2009. Preceding these birds by nearly a decade was a pair with yellow outer tail feathers at Mountain Lake, San Francisco, on 9 July 2000 (Andrew Rush; J. Morlan pers. comm.). LeValley also received photos of yellow-tailed Hairy Woodpeckers from outside of California. Sandy Shanks photographed one in Port Townsend, Washington, on 2 November 2012, and Shirley Powell photographed a juvenile at Kawortha Lakes in Ontario, Canada, on 27 July 2012. A yellow-tailed male had reportedly visited the feeder for years and was accompanied by the juvenile during the summer of 2012. Hairy Woodpeckers with yellow outer rectrices apparently have not been reported in the literature previously (Jackson et al. 2002, Pyle 1997). Birds with yellow instead of red crowns occur from time to time (Putnam 1998), and exposure to tannins, soot, or pitch can stain the white feathers of the Hairy Woodpecker brown (Jackson et al. 2002). The subjects of our paper, however, with the exception of one individual (back cover), exhibited yellow only on the outer tail feathers. The confinement of yellow to the outer tail feathers and the apparent persistence of this plumage pattern from year to year suggest the aberration results not from staining but from carotenism. Western Birds 44:85-87, 2013 85 FEATURED PHOTO Carotenism is an environmental or genetic disorder affecting the normal expression of carotenoid pigments (red, yellow, and orange) in the integument of birds and other animals. The environmentally induced form in birds results from the intake or lack thereof of certain carotenoid-containing foods just prior to or during molt. It affects only birds normally pigmented with carotenoids. Carotenism can result from (1) a change in the normal distribution or extent of carotenoid pigments, (2) an increase or decrease in carotenoid concentration, (3) a change in type of carotenoid pigment, or (4) the total absence of carotenoids (Davis 2007). Abnormal yellow coloration has also been called xanthochroism, xanthochromism, xanthism, and flavism. Although those terms indicate yellow coloration, they obscure the mechanism responsible for the abnormality. That is, abnormal yellow coloration often results not from a change in carotenoids but from the absence of dark melanin, which reveals the presence of carotenoid pigmentation (Harrison 1966, Hailman 1984). The absence of melanin, however, is more aptly termed amelanism (Davis 2007). A review of the literature and unpublished photos suggests that carotenism most commonly involves a change in the expression of carotenoid pigment type, whereby one color replaces another. Examples include yellow instead of red hind crowns in the Hairy Woodpecker (Putnam 1998), orange instead of yellow tail tips in the Cedar Waxwing ( Bombycilla cedrorum; Mulvihill et al. 1992), orange instead of yellow lores in the White-throated Sparrow (Zonotrichia albicollis ; Brooks 1994), red instead of orange plumage in the Baltimore Oriole (Icterus galbula; Flinn etal. 2007), and yellow instead of red plumage in the Northern Cardinal ( Cardinalis cardinalis; McGraw et al. 2003). In the Cedar Waxwing, White-throated Sparrow, and Baltimore Oriole, the anomalies have been shown or suggested to result from environmental factors alone, through consumption of rhodoxathin-containing honeysuckle fruit just prior to and during molt. The aberration in the Northern Cardinal, however, is thought to result from a genetic mutation that disrupts the normal metabolic pathway for carotenoid expression. Genetic mutation would seem a reasonable explanation for the yellow instead of red hind crown in the Hairy Woodpecker as well. Expression of carotenoid-based colors where none are normally present is less common. Examples include pink flushes in the plumage of the Ring-billed Gull ( Larus delawarensis; Hardy 2003), yellow or orange-buff underparts in the White-throated Sparrow (Brooks 1994), and red underparts and rump in the Rose-breasted Grosbeak ( Pheucticus ludouicianus ; Pittaway and Iron 2006). Hairy Woodpeckers with yellow outer tail feathers would fall into this category. This plumage anomaly involves both a change in the normal distribution of carotenoid pigments (from hind crown alone to hind crown and tail feathers) and a change in carotenoid pigment type (from red pigments alone to red and yellow pigments). That the yellow is restricted to the outer tail feathers, which don't normally show carotenoid-based color, suggests a genetic mutation is likely responsible for this pattern. Likewise, that the plumage pattern seems to persist in individuals for years also implies the aberration results from a mutation rather than from an effect of diet prior to and during molt. We know of no other case involving manifestation of two carotenistic effects presumably triggered by a genetic mutation. Furthermore, most cases involving a change in carotenoid pigment type result in a change in color in normally carotenoid-colored areas. This case, however, resulted in a new color in a normally uncolored area. The yellow-tailed Hairy Woodpeckers from coastal northern California could conceivably be attributable to a single mutant parent and its mutant progeny. How- ever, the birds from Washington and Ontario suggest this genetic anomaly can arise independently. More information is needed to evaluate the cause, heritability, and persistence of this fascinating coloration. Please send reports and photos of such birds to the authors. We thank Joseph Morlan and Steve Shunk for comments on a draft of this note. 86 FEATURED PHOTO LITERATURE CITED Brooks, E. W. 1994. Diet-induced color variation in the White-throated Sparrow. N. Am. Bird Bander 9:12-13. Davis, J. N. 2007. Color abnormalities in birds: A proposed nomenclature for bird- ers. Birding 39:36-46. Flinn, T., Hudon, J., and Derbyshire, D. 2007. Tricks exotic shrubs do: When Balti- more orioles stop being orange. Birding 39:62-68. Hailman, J. P. 1984. On describing color abnormalities in birds. Fla. Field Nat. 12:33-36. Harrison, C. J. O. 1966. Alleged xanthochroism in bird plumages. Bird-Banding 37:121. Hardy, L. 2003. The peculiar puzzle of the pink Ring-billed Gulls. Birding 35:498- 504. Jackson, J. A., Ouellet, H. R., and Jackson, B. J. 2002. Hairy Woodpecker ( Picoides villosus), in The Birds of North America (A. Poole and F. Gill, eds.), no. 702. Birds N. Am., Inc., Philadelphia. McGraw, K. J., Hill, G. E., and Parker, R. S. 2003. Carotenoid pigments in a mu- tant cardinal: Implications for the genetic and enzymatic control mechanisms of carotenoid metabolism in birds. Condor 105:587-592. Mulvihill, R. S., Parkes, K. C., Leberman, R. C., and Wood, D. S. 1992. Evidence supporting a dietary basis for orange-tipped rectrices in the Cedar Waxwing. J. Field Ornithol. 63:212-216. Pittaway, R., and Iron, J. 2006. Erythristic Rose-breasted Grosbeak. Ontario Birds 24:2-5. Putnam, C. 1998. Xanthochroistic Hairy Woodpecker in Gratiot Co. Michigan Birds Nat. Hist. 5:59. Pyle, P. 1997. Identification Guide to North American birds, part. 1: Columbidae to Ploceidae. Slate Creek Press, Bolinas, CA. Wing your way to... OLYMPIA, WASHINGTON 22-25 AUGUST 2013 Mark your calendar now for a joint conference of Western Field Ornithologists and the Washington Ornithological Society, hosted by Black Hills Audubon Society, in Olympia, Washington, 22-25 August 2013. This conference will feature workshops on fall warblers (Jon Dunn and Kimball Garrett), sapsuckers (Steve Shunk), gull identification (Mike Donahue), dragonfly identification (Dennis Paulson), wilderness first aid (Heath Wakelee), and eBird (Brian Sullivan). John Marzluff will give the keynote talk on corvids. Two afternoon science sessions will begin with a plenary talk by Dennis Paulson. And, of course, we will offer a full set of field trips (including pelagics) and both photo and bird-sounds sessions. Registration will open in early May; please watch www.westernfieldornithologists. org/conference.php for more details. Join us in Olympia! 87 World Wide Web site: WESTERN BIRDS www. westernfieldornithologists . org Quarterly Journal of Western Field Ornithologists President: Edward R. Pandolfino, 1328 49th St., Sacramento, CA 98519; ERPfromCA@aol . com Vice-President: David E. Quady, 39 The Crescent, Berkeley, CA 94708; davequady@att. net Past-President: W. David Shuford, P O. Box 69, Bolinas, CA 94924; dshuf ord@prbo . org Treasurer/Membership Secretary: Robbie Fischer, 1359 Solano Dr., Pacifica, CA 94044; robbie22@pacbell.net Recording Secretary: Liga Auzins, 12842 Safford East, Garden Grove, CA 92840; llauzins@yahoo . com Directors: Kenneth P. Able, Kimball L. Garrett, Daniel D. Gibson, Robert E. Gill, Ed Harper, Osvel Hinojosa-Huerta, Kurt Leuschner, Joyce Meyer, Frances Oliver, Dan Singer, Brian L. Sullivan, Debbie Van Dooremolen Editor: Philip Unitt, San Diego Natural History Museum, P. O. Box 121390, San Diego, CA 92112-1390; birds@sdnhm.org Assistant Editor: Daniel D. Gibson, 3705 Quartz Rd., Ester, AK 99725; avesalaska@gmail . com Associate Editors: Kenneth P. Able, Doug Faulkner, Thomas Gardali, Daniel D. Gibson, Robert E. Gill, Paul E. Lehman, Ronald R. LeValley, Dan Reinking Graphics Manager: Virginia P. Johnson, 4637 Del Mar Ave., San Diego, CA 92107; gingerj 5@juno . com Photo Editor: Peter LaTourrette, 1019 Loma Prieta Ct., Los Altos, CA 94024; petelat 1 @stanford . edu Featured Photo: John Sterling, 26 Palm Ave., Woodland, CA 95695; jsterling@ wavecable . com WFO Website: Joseph Morlan, 1359 Solano Dr., Pacifica, CA 94044; webmaster@ westernfieldornithologists . org Book Reviews: Lauren Harter, 2841 McCulloch Blvd N. #1, Lake Havasu City, AZ 86403; lbharter@gmail.com Membership dues, for individuals and institutions, including subscription to Western Birds: Patron, $1000.00; Life, $600 (payable in four equal annual installments); Supporting, $75 annually; Contributing, $50 annually; Family $40; Regular U.S. $35 for one year, $60 for two years, $85 for three years. Dues and contributions are tax-deductible to the extent allowed by law. Send membership dues, changes of address, correspondence regarding missing issues, and or- ders for back issues and special publications to the Treasurer. Make checks payable to Western Field Ornithologists. BACK ISSUES OF WESTERN BIRDS WITHIN U.S. $40 PER VOLUME, $10 FOR SINGLE ISSUES, INCLUDING SHIPPING AND HANDLING. OUTSIDE THE U.S. $55 PER VOLUME, $15 FOR SINGLE ISSUES, INCLUDING SHIPPING AND HANDLING. The California Bird Records Committee of Western Field Ornithologists revised its 10-column Field List of California Birds in July 2009. The list covers 641 species, plus 6 species on the supplemental list. Please send orders to WFO, c/o Robbie Fischer, Treasurer, 1359 Solano Drive, Pacifica, CA 94044. Price for 9 or fewer, $2.75 each, for 10 or more, $2.50 each, which includes tax and shipping. Order online at http://checklist.westernfieldornithologists.org. Published 28 February 2013 ISSN 0045-3897 Female yellow. (ailed ElaLry Woodpecker (Picoides viflosus) al Little River Headlands, Mendocino County, California. 3 November 2010, Photo by Ron LeVattey Male yellow-tailed Hairy Woodpecker (Pie&idts viit&sui) al MaoKerricher Slate Park, Mendocino County, California, 2 January 2011, Note the presence of Hie normally colored hind crown. Photo by Ron Li Val/ey Vol. 44, No. 2, 2013 Western Specialty: LeConle’s Th rasher Photo by © Sieve N. G. Howell of Hu l intis, California: Lc Conte's Thrasher {T&x&st&ma lec&nte i arenieofa) Guerrero Negro, Baja California Sue,. Mexico, 1 1 October 2012 The subspecies of LeContc’s Thrasher inhabiting the Vizcaino Desert of the wc&l -central Baja California peninsula, Tbxostome iecontei arenicoh* differs from that of tbe desert region* of the southwestern ITS. and northwestern Mexico, T. t. feemtiei, in its si ightly darker hack, distinctly darker breast, and less graduated tail, as well as by .f .$% in mitochondria] DNA. The vocalizations off. I. arenicoia remain unstudied. linglish names applied to this subspecies in the IDAUh include “Desert Thrasher ,"' 1 “Rosalia Thrasher , 11 and “Santa Rosalia Thrasher 1 "' (the type locality is Santa Rosalia Bay): today's birders ofteri call it the “Vizcaino Thrasher . 11 Volume 44, Number 2, 2013 Morphological and Molecular Evidence Confirm the First Definitive Eastern White-breasted Nuthatch ( Sitta c. carolinensis ) for New Mexico Matthew J. Baumann, Sabrina M. McNew, and Christopher C. Witt 90 Chronic Low Reproductive Success of the Colonial Tricolored Blackbird from 2006 to 2011 Robert J. Meese 98 Diet and Home-Range Size of California Spotted Owls in a Burned Forest Monica L. Bond, Derek E. Lee, Rodney B. Siegel, and Morgan W. Tingley 114 Human Food Subsidies and Common Raven Occurrence in Yosemite National Park, California Cara E. Brook, David P. Bernstein, and Elizabeth A. Hadly 127 Genetic Evidence for Mixed Maternity at a Lark Sparrow Nest Kevin Ellison, Jeremy D. Ross, and Juan L. Bouzat 135 NOTES Refutation of Wyoming Nesting Record of the Pacific Wren James M. Maley, Jacob R. Saucier, and Matthew D. Carling 141 Lack of Recovery of the Yellow-billed Magpie from the West Nile Virus in California’s Central Valley Edward R. Pandolfino 143 First Record of the Red-bellied Woodpecker in Nevada Kenneth M. Burton and Leslie Scopes Anderson 148 California Condor Foraging on a Live California Sea Lion Pup Mike Tyner, L. Joseph Burnett, and Mike M. Stake 151 In Memoriam: Richard W. Stallcup Jon Winter 155 Book Reviews Matthew J. Baumann and Cole Wolf 158 President’s Message Edward R. Pandolfino 161 Featured Photo: The Subspecies of the Song Sparrow on Southeast Farallon Island and in Central California Oscar Johnson, Peter Pyle, and Jim Tietz 162 Front cover photo by © Gary Lindquist of Visalia, California: Com- mon Cuckoo ( Cuculus canorus ), Watsonville, Santa Cruz Co., Cali- fornia, 1 October 2012, representing a first record for California. Elsewhere in western North America, this species is known only as a casual migrant in western Alaska. The unbarred rump, evident in other photos, best distinguishes the Common Cuckoo and Oriental Cuckoo (C. optatus ), also a vagrant to Alaska. Back cover: “Featured Photos” of Song Sparrows ( Melospiza melodia) on Southeast Farallon Island, San Francisco Co., Cali- fornia. Upper by © Peter Pyle, Point Reyes Station, California: subspecies melodia, 3 November 1993. Lower by © Oscar Johnson of Santa Barbara, California: subspecies morphna, 20 September 2010. Western Birds solicits papers that are both useful to and understandable by ama- teur field ornithologists and also contribute significantly to scientific literature. Send manuscripts to Daniel D. Gibson, P. O. Box 155, Ester, AK 99725; avesalaska@ gmail.com. For matters of style consult the Suggestions to Contributors to Western Birds (at www.westernfieldornithologists.org/docs/journal_guidelines.doc). WESTERN BIRDS Volume 44, Number 2, 2013 MORPHOLOGICAL AND MOLECULAR EVIDENCE CONFIRM THE FIRST DEFINITIVE EASTERN WHITE-BREASTED NUTHATCH ( SITTA C. CAROLINENSIS) FOR NEW MEXICO MATTHEW J. BAUMANN, SABRINA M. McNEW, and CHRISTOPHER C. WITT, Museum of Southwestern Biology and Department of Biology, University of New Mexico, Albuquerque, New Mexico 87131-0001; cwitt@unm.edu ABSTRACT: We report the first confirmed eastern White-breasted Nuthatch (Sitta carolinensis carolinensis ) from New Mexico. The bird was collected on 15 October 2011, at Boone’s Draw, Roosevelt County. Morphological and plumage characters matched those of the eastern White-breasted Nuthatch, the bill being markedly shorter and the back paler than those of the subspecies resident in New Mexico, S. c. nelsoni, and other western subspecies. The DNA sequence of the mitochondrial gene for NADH dehydrogenase subunit 2 (ND2) was an exact match to published sequences from Michigan and Pennsylvania and was phylogenetically nested within a monophyletic and deeply divergent eastern clade. The White-breasted Nuthatch ( Sitta carolinensis) is a common and geo- graphically variable resident of mature deciduous, mixed deciduous, and coniferous forests through most of North America. Grubb and Pravosudov (2008) recognized seven subspecies that constitute eastern (S. c. carolinen- sis), interior western (S. c. tenuissima, S. c. nelsoni, S. c. mexicana, and S. c. lagunae), and Pacific (S. c. aculeata and S. c. alexandrae) groups of populations. The seven subspecies are based on differences in bill size and shape and plumage coloration, but at least some of the variation in these characters is clinal. The eastern subspecies averages the shortest in bill length and palest in back color, while interior western and Pacific populations have longer bills and darker mantles (Phillips 1986, Pyle 1997, Sibley 2000, Wood 1992). There are three distinct vocal groups, based on typical call notes, that are concordant with boundaries between subspecies: (1) eastern United States and Canada (nasal yenk); (2) Great Basin, Rocky Mountains and Mexico (a rapid yijijijijiji or yiji-yiji-yiji-yiji), and (3) Pacific slope (high- pitched, drawn-out aaarn) (Gaines 1988, Sibley 2000, Dunn and Alderfer 2011). Recent multi-locus phylogenetic analyses of the White-breasted 90 Western Birds 44:90-97, 2013 EASTERN WHITE-BREASTED NUTHATCH IN NEW MEXICO Nuthatch uncovered four geographically distinct clades, with boundaries that are also consistent with vocal groups and subspecies: (1) eastern, (2) eastern Sierra Nevada and northern Rockies, (3) Rocky Mountain, Great Basin, and Mexico, and (4) Pacific (Spellman and Klicka 2007, Walstrom et al. 2012). Clades two and three are sister groups that together constitute vocal group 2. The deep phylogenetic structure and overall concordance between patterns of geographic variation in measurements, vocalizations, plumage, mitochondrial DNA, and nuclear DNA suggest that there may be more than one species of White-breasted Nuthatch. In New Mexico, the White-breasted Nuthatch is a common resident in practically all wooded areas of the state. It descends to lower elevations and withdraws from high peaks during the winter (Ligon 1961), but little else is known about its seasonal movements. The resident subspecies in New Mexico is the interior west S. c. nelsoni (AOU 1957, Phillips 1986), which has been placed in the Rocky Mountain, Great Basin, and Mexico phylo- genetic clade (Spellman and Klicka 2007). Only one previous specimen of the White-breasted Nuthatch from New Mexico (University of Arizona [UAZ] 5057) has been identified as any subspecies other than S. c. nelsoni. That specimen, collected in Socorro County by Gale Monson on 23 December 1941, was identified by Alden H. Miller as S. c. tenuissima (J. P. Hubbard pers. comm.), a subspecies constituting the eastern Sierra Nevada clade (Spellman and Klicka 2007) and for which the type locality is in the Panamint Mountains, Inyo County, California. Birders have reported White-breasted Nuthatches of the eastern subspecies in Curry and Sandoval counties on the basis of calls, but they were unable to obtain recordings or conclusive photos (W. H. Howe and C .R. Rustay pers. comm.). This phenotypically distinctive and genetically divergent eastern subspecies of the White-breasted Nuthatch has therefore not yet been fully documented in New Mexico (S. O. Williams pers. comm.). Here we describe a female eastern White-breasted Nuthatch collected on 15 October 2011 from the eastern plains of New Mexico, representing the first definitive state record of S. c. carolinensis. Morphological and molecular data independently support assignment of the specimen to the eastern clade. METHODS At 08:10 on 15 October 2011, Baumann, Nicholas D. Pederson, and Cole J. Wolf mist-netted a White-breasted Nuthatch at Boone’s Draw, Roosevelt County (34° 06' 41" N, 103° 31' 44" W), a first record of this species at this locality over three previous seasons of netting. We collected the bird (New Mexico Department of Game and Fish permit 3217; U.S. Fish and Wildlife Service permit MB094297-1) because measurements taken in the field and compared to Pyle (1997) suggested tentatively that it was of the eastern subspecies. We took the specimen to the Museum of South- western Biology (MSB) at the University of New Mexico, where Andrew B. Johnson prepared it as a study skin MSB 37575 and frozen tissue sample (NK1 74933). We compared the plumage patterns and bill measurements of MSB 37575 to those of specimens at the MSB, UAZ, and University of Kansas Museum of Natural History. 91 EASTERN WHITE-BREASTED NUTHATCH IN NEW MEXICO Figure 1. Comparison of the pale bluish gray mantle of MSB 37575 (middle) and an eastern White-breasted Nuthatch (MSB 21144, left), with the dark bluish gray mantle of S. c. nelsoni (MSB 26967, right). Molecular Methods Following the manufacturer’s protocol, we extracted DNA from frozen muscle tissue of MSB 37575 with a Qiagen DNEasy kit. We amplified the gene for mi- tochondrial NADH dehydrogenase subunit 2 (ND2) by using the primers H6313 and L5219 (Sorensen et al. 1999) in a 15-pL reaction with 1 pL of DNA and the following reagents: 0.5 mM of each primer, 2 mM of each dNTP, 0.75 units AmpliTaq Gold (Life Technologies), 1.5 pL of MgCl, and 1.5 pL of Taq Gold buffer. Amplification reactions were done according to the following protocol: 95 °C for 8 min, (95 °C for 45 sec, 50 °C for 30 sec, 72 °C for 45 sec) x 35 cycles, 72 °C for 10 min. We visualized products of the polymerase chain reaction on a 1% agarose gel, then cleaned with Exo-Sap-It (USB Corporation). Using BigDye 3.1 chemistry (Life Technologies), we sequenced the products with external primers and read them with an ABI 3130 automated sequencer. We assembled the sequences and inspected them manually with Sequencher 4.7 (GeneCodes). We used the software packages Muscle (Edgar 2004) and Phyml (Guindon and Gascuel 2003), respectively, to align the sequence with previously published ND2 sequences and analyze it phylogenetically. We used the program Mega (Kumar et al. 2008) to calculate average uncorrected pairwise distances between MSB 37575 and the four major clades identified by Spellman and Klicka (2007). 92 EASTERN WHITE-BREASTED NUTHATCH IN NEW MEXICO Figure 2. Comparison of the pale salmon wash on the flanks and huffy wash on the belly and upper breast of MSB 37575 (middle) and an eastern White-breasted Nuthatch (MSB 21144, left) with the bluish gray flanks of S. c. nelsoni (MSB 26967 , right). RESULTS Plumage The forecrown of MSB 37575 is dark blue, while the hindcrown and nape are black, contrasting with the white face and sides of the neck. The black of the nape expands out toward the anterior edge of the mantle. The mantle is pale bluish gray, notably lighter than the dark bluish gray of S. c. nelsoni (Figure 1). The tertials of MSB 37575 are also pale bluish gray with black inner webs, while the tertials of S. c. nelsoni are darker blue and less contrasting than those of MSB 37575 (Figure 1). The greater coverts are pale blue with large black centers and white edging on the distal ends of the feathers. The primaries are brownish with paler tips. The rectrices, except the pale blue central pair, are black with large white corners. The throat, upper breast, and 93 EASTERN WHITE-BREASTED NUTHATCH IN NEW MEXICO lower belly are white with the belly and lower breast tinged buffy. The flanks of MSB 37575 are white with a pale salmon wash, while the flanks of S. c. nelsoni are bluish gray (Figure 2). The undertail coverts are white with limited rufous edging on some feathers. Measurements The exposed culmen of MSB 37575 measures 17.1 mm, which fits with our measurements for 13 female eastern White-breasted Nuthatches (mean 17.3 mm, SD 1.0; Table 1). It is below the range of our measurements of 15 female specimens of S. c. nelsoni from New Mexico and Arizona (range 17.5-19.8 mm; Table 1). Molecular Results We successfully amplified and sequenced a 968-base-pair fragment of the ND2 gene. Chromatograms of the sequence were clear and unambigu- ous, without double peaks or internal stop codons that would indicate the erroneous amplification of a nuclear pseudogene. The complete sequence is available at Genbank (accession no. JQ965152). Comparison with published sequences of the White-breasted Nuthatch confirmed that MSB 37575 is part of the eastern clade with 100% bootstrap support. Over the area sequenced, the ND2 gene from MSB 37575 is a 100% match with published haplotypes of S. c. carolinensis from Michigan and Pennsylvania (haplotypes Ell, E19, and E23 in Spellman and Klicka 2007). The average pairwise difference between the New Mexico specimen and sequences from other eastern specimens was 0.3% (Table 2). Average divergences between the New Mexico specimen and the other clades were substantially higher: 3.8% with the Pacific clade, 6.3% with the Rocky Mountain, Great Basin, and Mexico clade, and 6.9% with the eastern Sierra Nevada clade (Table 2). DISCUSSION MSB 37575 can safely be identified as an eastern White-breasted Nut- hatch on the basis of measurements, plumage color, and mitochondrial DNA sequence. The bill measurements are within the expected range of the eastern subspecies but below the range of S. c. nelsoni, resident in New Mexico, on the basis of measurements from a series of specimens. The most noticeable plumage features distinguishing eastern and western populations are the mantle color, flank color, and the degree of contrast on the tertials, and these differences are consistent between the sexes. In S. c. carolinensis and MSB 37575, the mantle and tertials are pale bluish gray, making the black inner webs of the tertials contrast sharply, while the mantle of western populations is dark bluish gray and the tertials are less contrasting (Figure 1). S. c. carolinensis and MSB 37575 show whitish flanks with a pale salmon wash, while S. c. nelsoni has bluish gray flanks (Figure 2). These plumage distinctions can be used in the field, with care, by birders or banders for subspecific identification and are described in several widely used identifica- tion guides (Pyle 1997, Sibley 2000, Dunn and Alderfer 2011). Molecular data corroborate the morphological analysis. The mitochondrial 94 EASTERN WHITE-BREASTED NUTHATCH IN NEW MEXICO Table 1 Measurements of Exposed Culmen (mm) of MSB 37575, Sitta c. carolinensis, and S. c. nelsoni, by Sex Subspecies or specimen Sex n Exposed culmen (mean, mm) SD Range MSB 37575 F — 17.1 — — S. c. carolinensis a F 13 17.3 1.0 15.0-18.5 S. c. nelsoni h F 15 18.3 0.6 17.5-19.8 S. c. carolinensis a M 16 17.6 0.9 15.9-18.7 S. c. nelsoni b M 15 19.5 0.7 18.1-20.5 a Specimens from Kansas in the University of Kansas Museum of Natural History b Specimens from New Mexico and Arizona in the Museum of Southwestern Biology. DNA sequence matches published sequences that are phylogenetically nested within the deeply divergent eastern clade, with 100% bootstrap support. Because the geographic sampling of published sequences is thorough, the only reasonable explanation for these data is that MSB 37575 originated within the range of S. c. carolinensis. The approximate western limit of S. c. carolinensis extends from central Texas north through central Oklahoma, central Kansas, northeastern Colorado, western Nebraska, and the Dakotas to southern Manitoba (AOU 1957, 1998, Grubb and Pravosudov 2008, Sullivan et al. 2009, Such and Such 2012; ). As northern populations of S. c. carolinensis move irruptively (Heintzelman and MacClay 1971), primarily in fall (Phillips et al. 1964, Phillips 1986), the occurrence of this subspecies in New Mexico should be expected, especially given the closeness of breed- ing populations and the subspecies’ apparent regularity in eastern Colorado during migration (Leukering et al. 2012). It is likely that many observed seasonal movements represent young of the year dispersing farther in higher numbers in late summer and fall (Grubb and Pravosudov 2008). MSB 37575 was in its first year on the basis of 30% skull ossification and the presence of a bursa of Fabricius. Previous reports suggest that other eastern White-breasted Nuthatches have occurred in New Mexico. C. R. Rustay (pers. comm.) heard a White- Table 2 Mean Levels of Percent Divergence (p) in the Mitochondrial Gene ND2 between the MSB 37575 and the Four Major Clades of Sitta carolinensis (1) (2) (3) (4) (1) MSB 37575 (2) Eastern clade 0.003 (3) Pacific clade 0.038 0.038 (4) Rocky Mountain, Great Basin, and Mexico clade 0.063 0.061 0.073 (5) Eastern Sierra Nevada clade 0.069 0.068 0.080 0.017 "As defined by Spellman and Klicka 2007. 95 EASTERN WHITE-BREASTED NUTHATCH IN NEW MEXICO breasted Nuthatch giving “yen/c” calls like those of S. c. carolinensis at Hillcrest Park on 18 December 2010 in Clovis. Similarly, White-breasted Nuthatches have been heard giving vocalizations reminiscent of the eastern subspecies in the Corrales bosque (-35° 14' N, 106° 35' W), Sandoval County, on several occasions between December and March (Williams and Howe 2011). As of December 31 2006, the Melrose migrant trap (-34° 26' N, 103° 48' W), Roosevelt County, had four records of the White-breasted Nuthatch on dates ranging from 1 August to 21 September (Par meter 2007), but none of these individuals was identified to subspecies. Additionally, Wyoming has a single report of a vocal eastern White-breasted Nuthatch from Rawhide Wildlife Habitat Management Unit, Goshen County, in April (Faulkner 2010). Four subspecies of White-breasted Nuthatch have been described from the eastern United States: S. c. carolinensis, S. c. cookei, S. c. atkinsi, and S. c. litorea (Oberholser 1917, AOU 1957, Phillips 1986, Dickinson 2003). An analysis of the eastern White-breasted Nuthatches by Wood (1992), however, found much of the variation to be clinal, with no appre- ciable difference in back coloration. For this reason, he proposed that all eastern White-breasted Nuthatches should be considered one subspecies, S. c. carolinensis. Spellman and Klicka (2007) subsequently found that all of the eastern populations are monophyletic by mtDNA, with little or no geographic structure. The eastern clade of the White-breasted Nuthatch may warrant species status on the basis of its deep level of molecular divergence, reciprocal monophyly, and morphological and vocal distinctness (Spellman and Klicka 2007, Walstrom et al. 2011). In conclusion, MSB 37575 represents the first confirmed eastern White- breasted Nuthatch for the state of New Mexico. Its occurrence there has been reported on several occasions, and it likely occurs with some regularity on the eastern edge of the state. White-breasted Nuthatches found in New Mexico should be scrutinized for potential vagrants of the eastern or other subspecies, particularly during fall migration and when found away from areas of known breeding. Identification should be based on a combination of characters, potentially including vocalizations, back and flank color, tertial pattern, bill length, and mitochondrial DNA. ACKNOWLEDGMENTS We thank Andrew B. Johnson for curation of MSB 37575. We thank Nicholas D. Pederson and Cole J. Wolf for their assistance in the field. We thank Natalie A. Wright and Mark B. Robbins for help obtaining measurements from the University of Kansas Museum of Natural History. The University of Arizona (George Bradley and Alex Badyaev) provided a loan of specimens. We thank John P. Hubbard for his assistance on previous records of the White-breasted Nuthatch in New Mexico. William H. Howe and Christopher R. Rustay provided valuable information about previous experience with probable eastern White-breasted Nuthatches in New Mexico. We thank Gregg Moore for allowing access to his property for our field work. We thank Doug Faulkner, Kimball Garrett, and Philip Unitt for perceptive comments on the manuscript. 96 EASTERN WHITE-BREASTED NUTHATCH IN NEW MEXICO LITERATURE CITED American Ornithologists’ Union. 1957. Check-list of North American Birds, 5th ed. Am. Ornithol. Union, Washington, DC. American Ornithologists’ Union (AOU). 1998. Check-list of North American Birds, 7 th ed. Am. Ornithol. Union, Washington, D.C. Dickinson, E. C., ed. 2003. The Howard and Moore Complete Checklist of the Birds of the World. Princeton Univ. Press, Princeton, NJ. Dunn, J. L., and Alderfer, J. K. 2011. National Geographic Field Guide to the Birds of North America, 6th ed. Natl. Geogr. Soc., Washington, DC. Edgar, R. C. 2004. Muscle: A multiple sequence alignment method with reduced time and space complexity. BMC Bioinformatics 5:113. Faulkner, D. W. 2010. Birds of Wyoming. Roberts and Co., Greenwood Village, CO. Gaines, D. 1988. Birds of Yosemite and the East Slope. Artemisia Press, Lee Vin- ing, CA. Grubb, T. C. Jr., and Pravosudov, V. V. 2008. White-breasted Nuthatch ( Sitta caro- linensis), in The Birds of North America Online (A. Poole, ed.), no 54. Cornell Lab Ornithol., Ithaca, NY; http://bna.birds.cornell.edu/bna/species/054. Guindon, S., and Gascuel, O. 2003 . A simple, fast, and accurate algorithm to estimate large phylogenies by maximum likelihood. Syst. Biol. 52:696-704. Heintzelman, D. S. and MacClay, R. 1971. An extraordinary autumn migration of White-breasted Nuthatches. Wilson Bull.. 83:129-131. Leukering, T., Schmoker, B., and Mlodinow, S.G. 2012. Fall migration: Colorado and Wyoming region. N. Am. Birds 66:127-132. Ligon, J. S. 1961. New Mexico Birds. Univ. of New Mexico Press, Albuquerque. Oberholser, H. C. 1917. Critical notes on the eastern subspecies of Sitta carolinensis Latham. Auk 34:181-187. Par meter, J. E. 2007. Annotated checklist of the birds of the Melrose migrant trap, Roosevelt County, New Mexico. New Mexico Ornithol. Soc. Bull. 35:2. Phillips, A., Marshall, J., and Monson, G. 1964. The Birds of Arizona. Univ. of Ariz. Press, Tucson. Phillips, A. R. 1986. The Known Birds of North and Middle America, part I. Hirun- dinidae to Mimidae; Certhiidae. A. R. Phillips, Denver. Pyle, P. 1997. Identification Guide to North American Birds, part I. Slate Creek Press, Bolinas, CA. Sibley, D. A. 2000. The Sibley Guide to Birds. Knopf, New York. Sorenson, M. D., Ast, J. C., Dimcheff, D. E., Yuri, T., and Mindell, D. P. 1999. Prim- ers for a PCR-based approach to complete mitochondrial genome sequencing in birds and other vertebrates. Mol. Phylogenet. Evol. 12:105-114. Spellman, G. M., and Klicka, J. 2007. Phylogeography of the White-breasted Nuthatch (Sitta carolinensis ): Diversification in North American pine and oak woodlands. Molecular Ecol. 16:1729-1740. Such, J., and Such, M. 2012. News from the field: spring 2012 (March-May). Colo. Birds 46:277-295. Sullivan, B. L., Wood, C. L., Iliff, M. J., Bonney, R. E., Fink, D., and Kelling, S. 2009. eBird: A citizen-based bird observation network in the biological sciences. Biol. Conserv. 142:2282-2292. Walstrom, W. V., Klicka, J., and Spellman, G. M. 2012. Speciation in the White- breasted Nuthatch ( Sitta carolinensis): a multilocus perspective. Molec. Ecol. 21:907-920. Williams, S. O., and Howe, W. H. 2011. New Mexico Ornithological Society Field Notes (Spring 2011). 50(2): 27. Wood, D. S. 1992. Color and size variation in eastern White-breasted Nuthatches. Wilson Bull. 104:599-611. Accepted 21 February 2013 97 CHRONIC LOW REPRODUCTIVE SUCCESS OF THE COLONIAL TRICOLORED BLACKBIRD FROM 2006 TO 2011 ROBERT J. MEESE, Department of Environmental Science & Policy, University of California, One Shields Avenue, Davis, California 95616; rjmeese@ucdavis.edu ABSTRACT: 1 studied the Tricolored Blackbird (Agelaius tricolor) in California’s Central Valley over six breeding seasons from 2006 through 2011 and documented fates of nesting attempts, reproductive success of colonies, and relative abundance of insect prey in foraging areas. I found widespread and chronic reproductive failures except in cases of relatively high insect abundance. My observations suggest that the productivity of Tricolored Blackbird colonies is food-limited and that the relatively high reproductive success at few colonies is primarily a function of unusually high abundance of insects in nearby foraging areas. The Tricolored Blackbird ( Agelaius tricolor) forms the largest breeding colonies of any North American songbird (Beedy and Hamilton 1999). Over 99% of the world’s population occurs in California, with small numbers also in Washington, Oregon, Nevada, and Baja California (Beedy and Hamilton 1999). In California, for decades the largest colonies and the vast majority of the population have occurred in the Central Valley (Neff 1937, Beedy and Hamilton 1999, Cook and Toft 2005). Over the 20 th century, the number of Tricolored Blackbirds plummeted from habitat loss through conversion to agriculture and urbanization, market hunting, poisoning, and the birds being shot as an agricultural pest (Neff 1937, Beedy and Hamilton 1999). This decline resulted in a petition by the Center for Biological Diversity to list this blackbird under the state and federal endangered species acts (CBD 2004). Although these petitions were declined, the Tricolored Blackbird is considered a federal species of con- servation concern (USFWS 2008) and a California Bird Species of Special Concern (Shuford and Gardali 2008). Recent statewide surveys and intensive monitoring of colonies in the Central Valley have shown that its abundance has continued to decline, fall- ing 35%, from about 400,000 to 250,000 individuals, from 2008 to 2011 (Kyle and Kelsey 2011, Meese 2011). Large numbers formerly nested in coastal marshes in southern California (Baird 1870, Neff 1937, Unitt 2004), but workers in this region have recently documented severe population declines in this population segment (Unitt 2004, Feenstra 2009). I report here on field work in the Central Valley over six breeding seasons, from 2006 through 2011, during which I sought to detect, monitor, and es- timate the productivity of each of the Tricolored Blackbird’s largest breeding colonies. Key goals were to examine the relationship between the blackbird’s reproductive success and insect abundance and to use these results to inform conservation decisions intended to increase the species’ numbers. 98 Western Birds 44:98-113, 2013 LOW REPRODUCTIVE SUCCESS OF THE COLONIAL TRICOLORED BLACKBIRD METHODS My field work from 2006 to 201 1 covered the Tricolored Blackbird’s breed- ing season from late March through July, addressing the following topics. Surveys of Previously Occupied Locations I began surveys in early spring in the San Joaquin Valley, where Tricol- ored Blackbirds breed for the first time each season (Hamilton 1998), then moved north to the Sacramento Valley as the season progressed (Figure 1). I surveyed by car on public roads the locations of all previously occupied colonies known by or reported to me. I was especially interested in document- ing those colonies where the nesting substrate is ephemeral, primarily grain fields near dairies, as these colonies were at risk from loss of nests by the harvest of triticale, a wheat ( Triticum ) x rye ( Secale ) hybrid grown for dairy cows (Beedy and Hamilton 1999, Cook and Toft 2005, pers. obs.; Figure 2). Surveys for New Locations I supplemented the surveys of existing colony locations with intensive searches for new ones in grain and weedy fields adjacent to dairies and in freshwater wetlands in state wildlife areas, national wildlife refuges, and accessible private duck clubs. In both the San Joaquin Valley and southern California, from the 1980s to the present, many of the largest colonies have been located adjacent to dairies (Beedy and Hamilton 1999). Therefore, to enhance colony detection, I used available GIS layers to map the locations of all dairies in the San Joaquin Valley, then transferred these locations to paper maps that I took into the field to guide survey efforts. After birds completed their breeding efforts in the San Joaquin Valley, I used similar methods to survey for colonies in the Sacramento Valley. I surveyed locations of previously documented colonies and sites that had been reported to me by numerous collaborators (state and federal agency personnel, observers posting messages to the central_valley_binds listserv, birders) or entered into the Tricolored Blackbird Portal (http://tricolor.ice. ucdavis.edu). Colony Monitoring I monitored colonies from within days of the arrival of adults and the initiation of breeding until breeding ceased and the adults had departed. In most cases, I monitored colonies twice a week to assess current condi- tions and the colony’s chronology to estimate the optimal times for making estimates of reproductive success and the size of the breeding population. Monitoring entailed making observations from the closest public road for colonies located on private property that I lacked permission to access, or from immediately adjacent roads for colonies located on public property or on private property I had permission to access. Estimating Area Occupied I estimated the dimensions of the available nesting substrate and the area occupied by breeding birds in one of two ways. On public or private property 99 LOW REPRODUCTIVE SUCCESS OF THE COLONIAL TRICOLORED BLACKBIRD Figure 1. Locations of Central Valley Tricolored Blackbird colonies studied from 2006 to 2011. Numbers correspond to records in Table 1. Some locations were occupied in more than 1 year (see Table 1). with access, I measured these dimensions directly with a hand-held GPS. On private property where I lacked access, I visually estimated these dimensions from a distance and sketched them in a field notebook. The area occupied by breeding Tricolored Blackbirds is reliably estimated from careful observation of the birds’ behavior, as birds leave and return to nests at intervals defined by the stage of the breeding cycle. The interval is longest when females are incubating eggs, shortest when females are building nests or when both adults are feeding young. I confirmed the perimeter of the occupied area as initially outlined by the birds’ behavior through subsequent monitoring. To 100 LOW REPRODUCTIVE SUCCESS OF THE COLONIAL TRICOLORED BLACKBIRD Figure 2. Tricolored Blackbird colony in field of triticale adjacent to a dairy in Kern County, California. measure the size of the occupied area, I plotted both the dimensions from visual estimates and the coordinates measured by GPS into Google Earth. After birds had finished breeding, I also searched accessible but apparently unoccupied areas to confirm the absence of nests. Estimating the Number of Breeding Birds I estimated the number of breeding birds in a colony either visually at the time of nesting and/or by sampling nests after the breeding season. I made visual estimates of the number of breeding birds each time I monitored a colony by carefully observing it for 5-30 minutes per visit. When possible, I observed colonies from multiple vantage points to enhance detection and to increase the precision of the estimate. At sites to which I was granted access, I re-entered colonies after the young had fledged and both young and adults had left the area and estimated nest densities by one of two methods depending on the nesting substrate. In relatively impenetrable substrates, such as some colonies established in milk thistle ( Silybum marianum) or in Himalayan blackberry ( Rubus armeniacus), I counted nests within samples of randomly placed 1- x 2-m sampling frames of 2-cm PVC pipe. In all other nest substrates, I counted nests within line transects 2 m wide and varying in length from 20 to 100 m. I marked the start and end points of transects by a handheld GPS unit and computed the transects’ lengths by GPS or by the length-measurement tool in Google Earth. Using the densities of nests in the sampled areas, I calculated the number of nests in the colony by multiplying the average number of nests per hectare by the number of hectares occupied by breeding birds. I assumed that on average each male breeds with two females (Beedy and Hamilton 1999) so multiplied the number of nests by 1.5 to estimate the number of breeding birds at a colony. This provided an independent estimate of the number of breeding birds for comparison with my visual estimate during monitoring. Estimating Relative Abundance of Insects Each time I monitored a colony I spent 15-60 minutes observing the flight lines of actively foraging birds, which typically move back and forth from the colony to foraging destinations along nearly single-file flight paths. These 101 LOW REPRODUCTIVE SUCCESS OF THE COLONIAL TRICOLORED BLACKBIRD flight lines are conspicuous and persistent (Hoffmann 1927, Beedy and Hamilton 1999) and typically consist of several hundred to several thousand birds (pers. obs.). The distances from the colonies to foraging destinations varied from tens of meters to 9 km (Table 1). After determining the foraging destinations, I walked or drove to the areas where the largest numbers of birds were actively foraging and, where permissible, I visually estimated the relative abundance of insects in terrestrial habitats by walking (at a pace of 1 step every 3-5 seconds) over the substrate where birds had been foraging moments before and counting the number of insects seen. I waded in rice paddies adjacent to large colonies in Yuba and Colusa counties and supplemented my visual observations by reaching into and sweeping the muddy substrate with my fingers extended in an attempt to detect aquatic insect larvae. I subsequently categorized the number of insects observed per minute as low (1-10), moderate (11-100), or high (>100; Table 1). I estimated the relative abundance of insects 3-6 times per site within a 4-week interval while the site was occupied by breeding birds. Estimating Reproductive Success I estimated reproductive success, defined as the average number of young fledged per nest, either by visual estimates or by sampling. Visual estimates were derived from the number of breeding birds estimated visually during monitoring and the number of fledglings observed at the end of the breed- ing season. As one male breeds, on average, with two females (Beedy and Hamilton 1999), each two nests have three birds associated with them, so the product of the number of breeding birds multiplied by 2/3 provides an estimate of the number of nests constructed. The visual estimate of the number of young fledged divided by the estimate of the number of nests constructed yields an estimate of the number of young fledged per nest. For accessible colonies, prior to the fledging of young, I estimated repro- ductive success by counting the contents of a random sample of nests when the average age of nestlings was 7 to 9 days and calculated reproductive suc- cess by dividing the number of young by the number of nests in the sample. Estimating the Number of Young Produced I estimated the number of young produced at a colony by repeated observations of young in groups (“creches”) following fledging and/or by multiplying the number nests constructed times the average reproductive success of sampled nests (see above). At most colonies, the number of fledged birds may be counted. After fledg- ing, the young spend a minimum of several days in groups, perching and begging conspicuously from the tops of vegetation at the colony’s margins (Beedy and Hamilton 1999, pers. obs.). Typically, young begin to leave the nesting substrate to perch high in nearby shrubs or trees approximately 4 days after fledging. However, creches remain within the colony’s boundar- ies for up to 2 weeks or more if there are no nearby taller shrubs or trees, as is often the case in the “silage belt” of the southern San Joaquin Valley (pers. obs.). 102 LOW REPRODUCTIVE SUCCESS OF THE COLONIAL TRICOLORED BLACKBIRD Statistical Analyses I used Spearman’s rank correlation to test for a relationship between insect abundance and reproductive success and Pearson’s product-moment cor- relation to test for a relationship between colony size and reproductive suc- cess. I used a Kruskal-Wallis test to evaluate potential relationships between reproductive success and breeding season, type of nesting substrate, and the location of colonies by county. I excluded Butte County from the analysis for counties as I estimated reproductive success at only one site there. Finally, I used a Mann-Whitney test to look for a difference in reproductive success between the San Joaquin and Sacramento valleys and between colonies near and distant from dairies. RESULTS During the 6 years of this study the Tricolored Blackbird’s reproductive suc- cess across all sites and years averaged 0.62 young fledged/nest (540,000 young fledged from 870,000 nests; Table 1). Only 11% of the colonies studied (5/47) fledged an average of 1 or more young per nest. The relative abundance of insects on foraging substrates varied greatly by site and year, from a low of zero insects per 10 minutes of searching in the case of larvae of the water scavenger beetle (Coleoptera: Hydrophilidae) or other insects in rice paddies in Colusa County in 2007 to a high of over 1000 individuals per 10 minutes of searching in the case of grasshoppers in open pasture near a colony in Merced County in 2010 (Table 1). Within a site and year, however, estimates of relative abundance of insects varied little: at any single colony, they did not differ spatially or temporally. Insect abundance was positively correlated with reproductive success (Spearman’s rank correlation coefficient p = 0.737, P « 0.05), and the colony with the highest reproductive success (1.44), situated in milk thistle in Merced County in 2010, was surrounded by open rangeland in which grasshoppers were extraordinarily abundant (Table 1). Colony size was positively correlated with reproductive success (r = 0.53, r 2 = 0.28). Reproductive success did not differ by type of nesting substrate (Kruskal-Wallis H = 6.049, P = 0.109), year (Kruskal-Wallis H = 8.8, P = 0.117), or county (Kruskal-Wallis H = 7.841, P = 0.165). The presence of a dairy did not affect reproductive success (Mann-Whitney U = 208, P = 0.298), and the reproductive success of colonies in the San Joaquin and Sacramento valleys was similar (Mann-Whitney U = 189.5, P = 0.41). DISCUSSION The Tricolored Blackbird’s reproductive success was chronically low throughout the Central Valley, in all six breeding seasons studied from 2006 to 2011, and in all types of nesting substrate. Higher reproductive success was associated with greater abundance of favored insect groups in foraging habitats surrounding colonies and was not associated with dairies. Both terrestrial and aquatic insects varied widely in abundance. Caterpillars and grasshoppers dominated the terrestrial prey and dragonflies and larval 103 Table 1 Distribution, Characteristics, and Reproductive Success of Tricolored Blackbird Colonies Studied in California’s Central Valley, 2006-2011 Distance to No. insect Nesting No. breeding primary foraging samples/relative Reproductive Year and site 0 County substrate birds Foraging substrates destinations Insect groups taken insect abundance success LOW REPRODUCTIVE SUCCESS OF THE COLONIAL TRICOLORED BLACKBIRD 3 f -c CO £ 00 X T — 1 T—i X T — 1 X X CM X co O O t-H ON ON 0 00 0 0 O O O O 0 0 0 0 0 & -2 0 ) 2 2 2 IS X, 0 ) nD 0 O X O X O X 0 ) TD O 0 ) TD P) 0 X 0 ) D X S \ X \ X \ X vX S s \ X vx S CO \ \ \ X X X CO ON o s I oo c » » II a c/f &J 8 | s 2 •A o c n CD "T a 42 o3 -4— 1 o S 0 ) ° _ £ £ in 0 ) o o c n CD CD in CD CD in 'S. a o m -3 1/5 !?n Q) cd 0 ) ■° o. 5 r. is ^ ■- g - g s 5 > -*d c/) O g 2 “ O 5- 3 cd cr Q) to CQ cd cr 0 ) 3 CQ in c/) c/> E*_c o - .y e - .y c .y c o - - e- - 2$^ 8 loJS S 42 3 js gi 8 rrt ^ 'T“I ^ m ^ ^ -i—i P - ! rrt P - ! ’5 t/> "5 in 'S in o o 5= o 5= CD jo q) <3 cd 42 £ — r r\ t r\ CD cd 3 2 3 33 3 33 3 33 3 cr ro^ 003000 , 3 2 s c e- P 42 3 = 42 35 3 o 3 , 3 - 7-5 3 -a -§ 33 _ _r 3 - o’ 3X33 33 07) S3 3 33 4S 33 42 to J 15 ^cx < < in 33 in , 3 — 3 IU 3 la J£ 2 S ^ 3 S C ^ m— C < ^ < jOO-J 3 , t/> •is 33 ffi C > 33 3 .is -x 42 in -3 33 X 3 3 X 3 45 c/5 S3 x — 42 2 j§ 3 X =5 <« < 0 p-N 0 0 0 0 0 0 0 0 0 V — J P*s 0 0 0 X 0 0 0 0 0 0 V ' X 0 0 3 f 0 X 0 0 0 co X T—i o' o’ o’ x’ x’ 06 0' x' 0 co o' O'- X X X X X X X X o o o LO - ffl 3 o _o 3 o p p C/) ■?s 1 -4— 1 fO U ^ 0 ) (J p J) to 2 -4— 1 to s -4— 1 t n c/> c^> (/) J) '($ -4— ' -4— 1 '(5 -I— 1 -4— 1 ‘S -4— 1 -4— 1 ’c0 -4— 1 -4—1 CO CJ (0 CO co CO s u u 0 u 3 x 3 CJ o X o o o o o X o o 3 o _o 3 in J 3 O CJ 3 in J 3 O CJ 3 x £ OJ OJ OJ ^d X o o CM CM [>. 00 o o CM X o CM X t". 00 CX3 00 o o CM ON 104 Distance to No. insect Nesting No. breeding primary foraging samples/relative Reproductive Year and site 0 County substrate birds Foraging substrates destinations Insect groups taken insect abundance success 11 Kern Cattails 2,000 Alkali shrublands, 0.5-6 km Beetles, weevils, 3/Low 0.04 alfalfa, Dairy caterpillars LOW REPRODUCTIVE SUCCESS OF THE COLONIAL TRICOLORED BLACKBIRD LO o T—i co LO LO T— 1 LO LO 00 o o o CM CM CM T—i CM o o o o o o O O O o CO o 00 o if o if o LO o 00 o \ 00 o oo o -J 00 3 CO Q> h-j (0 v- co '> in 12 ■> 12 ■> 12 ■> 12 ■> 12 1 CJ gd cd cT 12 ■> c n ■> co ■> co CD CD 5-h jd a CD G cr cd cd 2 jd CD a a o CD CD 5-h jd a 1 j-. -G OJ jd s p TD G cd T3 cd" J-H G cd" 5-h G S 3 S 3 5-h CD (U J-H CD 'G > ’■a CD 1 !-h J3 O) • (13 cd" M— . J-H _c 03 "ca jd 1 CO cd cd JD to cd to cd to cd to cd Id P "S > td CD cd" CJ 15 (13 ._ ■ cd cd M— . 15 < cd Oh 2 JZ co Oh Oh Oh CU C G < cd cd *-+— < *G 5-H ‘G < (13 < cd < (C *-t— I (13 O O O cm' _c/3 ’(3 -t— * -t— * (0 CJ o o o o o LO o o o o LO LO o o o o o o o o o^ o o o o o LO o' T—i LO T—i it 1 it t-H o' 00 lo' CO T—i CN1 lo' T—i ^0) 'H _Q3 'H 13 c 03 CD TD cd 13 cd (0 <13 G dG co CO CO cd .CJ ■-G co 12 "4-—* 1 1 CO G 2 03 fc H-H -4— i CO G 2 CD _C _Q3 13 £ 13 JD -4— i co 12 3* US (13 C 5-h CD EJ CJ co 12 -*—1 "2 -4— J -4— 1 cd "2 -4— 1 -4— 1 cd "2 -4— 1 -4—1 1 cd £ TD TD 13 G -4— 1 G cd CJ CJ CJ § G cd § G cd E cd I JD § CD ■ JO £ oo (13 JD £ CD X c r> y (fl £ 03 X O O o b ra CX (fl 03 Z Q) 03 5 u C o ^ ^ Q3 ^3 X s< i — 1 (/> X • 0) (ti Z 2“ ° £ 03 (3 <2 c/> C C 0) X 3 - 4—4 Cfl 3 O c n -t-j i o 0) eg 2 co c ■& s £ § •-2 la -H- 2^^ <2 to Z 3 cfl 3^ -4—4 3 3 O CJ 03 33 3 3 s-. 3 £ X °0 eo X 1 X CM o- X o- o CM T—i o X - 1 X X o X co T — 1 C 3 o O O o o t-H o o o o 03 ~3 03 33 O 00 03 -t— ' 3 s-. 03 33 O x> 03 3 r\ 3 ro 3 ro 3 ro 3 ro "03 IU S-. 0 ) 3 3 3 ro 3 \ 3 \ 3 \ CJ CJ o X- < CJ CJ CJ \ eo X X eo s \ X to Ic? ^ il: ir? — I- 03 03 3 Q3 Cl |S § s co X _2 03 co (0 3 <§ O 03 £ 3 CQ CO ti (« co 2 2 x .. 2 3 rH 03 CO CX = •- 03 13 CX fr.y^ & 2 2 ra 2 2 58 r. CO ^ CO CO ^ CO > 7^ CO 77=3 r ^H ^ ^3 ’"•* ^ *>» 3 3 § C >C gijT & _fS "3 JS - 3 JS "3 3 3 -§ X -§ X >§ X ^ ^ _c ra' x 3 x 3 2 2 cfl X cfl X Ifl X 3 3 3 3 , 3 tn X «3 "jg «3 “ “ JO .3 x < X lu X X ,u d, Ce X < 3 , 3 X 03 co 3 Ce X o o c — s o o o o o o o o 03 X o o c — J r~o o o o o o o o o 03 X i- .b LO o c — ^ X o o o o o o o o x x> T—i W X x' x' o' x' co x' o' o' 6 X X t". X CM T — 1 co CM H T — 1 CO § -fa "c £ - $ x - x <_ J a o o oo _Q3 3 _y X 15 2 JS £ CO '(5 y — 1 y — 1 CO u o CNJ CO '(3 -t — 1 -t — 1 CO CJ CO ‘ro -t-j -i— 1 CO CJ CO co JO o CJ CO .y i5 2 Z CM l>- o T—i O CM 3 X 3 CJ 2 Z CM CM co "(3 -*— > ~3 CJ 2 Z j) to X -4— < X c 3 >4 ^ X E C _ _03 ra to."* 13 -2 X 'H X ra £ u * .3 3 X X x ra x CJ x x 03 03 2 2 03 03 X 03 2 03 X 03 2 03 co 3 3 3 CQ X 03 2 03 CO X CM CM X CM X CM 106 LOW REPRODUCTIVE SUCCESS OF THE COLONIAL TRICOLORED BLACKBIRD !> 1 8 3 -a o 0 ) <_> o ^ 3 CX 0 ) Cd §> 8 •-3 C 3 js CO LO CM LO LO xo CM CT3 O CM T— H CM co LO CO O O O 0 O O O 0 O 0 0 O O ^ 0) r c/) id C O P) 3—4 03 D £ £ £ C) C \ 3 •— 03 x> ■ 03 3 C-J \ "O O \ \ C-J \ C.J O Z F u C C) ^t 1 LO LO CO 2, \ CO co LO xo CO 3 c n 3 3 -4-j i t/3 CX 3 O S— 4 CO -4—4 o 0 ) t/3 3 CO o g o) £ Q o 3 3 H — ' 5—. H — i fd (/> b.^ 3 3 c/3 0 ) -4—4 3 to -D 3 c/3 CO 3 'th 3 S-. O ux CO 3 X) X) 6 co .£! 3 3 -t— 1 0 ) X> 2 3 ^ c/> -t—i 1 3 3 O CJ a> 33 3 3 s- 3 £ 0 ) c n in xb co Ph > Ph > 0) Q) 0) q) ^ fa D- o) Ph a) Ft > a 5 a 5 « o „ o „ v — cx -3 c/f-3 c/f c/f -t-; *r; £ T) £ -S "S | s CJ - S 03 o 3 4-S 3 h -t— 1 c n C/3 33 JS 3 s -C 2 ^ ’fd -t— 1 -4— ' "3 -4—4 -4—4 C CJ fd 3 c fd T7b CJ CJ X JD s X3 T3 03 O S-4 03 £ O 0 O 0 03 S 03 £ £ £ O- o- 0 T—i T—i CM CM CN1 id fd 1 £ I in in cn HD HD HD C C C JS JS JS in in in in _ in ^ in (d Q) fd Q) fd j?' 2 J?' G CO 3 3 S3 3 3 3 c c c c c 3 < < < 03 03 O CX) I C/3 T3 3 JS 03 03 O) 3 3 Ph -4— 1 CO <£ , 3 3 3 3 3 < 03 '3 -(— ' -4—i 3 CJ t"- C30 CM 03- J2 J— I ^ 03 £ 03 03 3 03 g$ 03 JS 03 CX JS 03 3 — I 3» A CX ^ O CX O > C/3 m 03 -3 F\ -P 03 _2 -£.s> $ -£ 18 ,q> CO S p -g ’!> X 1 > 3 ' . 03 O 03 03 03 CQ cS « - 03 03 — S- > . . 03 Q3 3 CX 03 3 Cx 5 cx o X 5 r; -3 03 -£ % M S £ 03 cj _§ co 1 CX) I LO I c >. ^0) J) fd h 1J 3 XsC JJ fd cj P HD~ P-4 fd - 4 — < c/) in JZ -4— 1 2 -4— 1 P 3 \ . D • S 3 X 3 s § fd "3 T3 in CD C 3 D -4— 1 -(—J P-4 p O O CJ D DQ 0) 4 — 3 2 J-4 3 2 CT 3 LO o CM 03 _03 2 ~03 8.-8 o e X 3 03 ==! 3 'p. 03 1 fd o o co 1 co 1 LO c/f HD 3 & 3 & CO HD C/3 CO HD C fd £ 3 c fd HD 3 c fd HD C JS T3 3 c JS X) (d HD 3 3 in 3 3 co" CO CO fd JS -4—1 CO CO fd J-H HD HD P-4 P-4 p: JS 0) C CJ c CO 5 CO CD fd *-4— P-. D "CD JS co co P-4 P to JS CO CO Id P 3' CJ td p v; fd fd fd fd fd C C < Dh Ph CO CJ Ph CO C < pH 5 < LO c — s 0 O 0 O C — J 0 O 0 O c — J LO C — J O 0 O 0 O c — J LO of w 0' o' LO' O' cm' CM CM T—i 03 '3 -t— < -4—4 3 0 3 03 jc o CJ 03 CM 03 '3 -H -4—4 3 CJ 3 03 J 3 o CJ o CO 03 C o 3 (J O 3 CO E 03 03 CX) 107 LOW REPRODUCTIVE SUCCESS OF THE COLONIAL TRICOLORED BLACKBIRD water scavenger beetles dominated the aquatic insects foraging Tricolored Blackbirds chose. My attempts to estimate the relative abundance of aquatic insects were unsuccessful, as I observed no larval aquatic insects while sampling. The only aquatic insects observed in abundance were recently hatched dragonflies (giant green darners, Anax junius) at one site in Yuba County in 2007 and hatching caddisflies (Trichoptera) at one site in Kern County in 2009. The dragonflies hatched in mid-July, when only a single blackbird colony was known to be active, and although they were extremely abundant and easily captured, they were fed primarily to fledglings and emerged too late to result in high productivity at this colony. The caddisfly hatch was similarly dramatic but brief, lasting less than an hour, and did not appear to enhance the blackbirds’ productivity. It is possible that greater overlap between the Tricolored Blackbird’s breeding season and the peak of dragonfly hatching would have enhanced the blackbirds’ reproductive success, but aquatic insects, as a group, did not appear to contribute to high reproductive success at any colony I studied. Previous researchers have documented poor reproductive success of Tricolored Blackbird colonies, but mine is the first study to confirm low re- productive success at colonies throughout the Central Valley over a 6-year interval and the first to assess insect abundance at sites where breeding birds forage. The reproductive success of entire colonies can be reduced severely by both mammalian and avian predators (reviewed by Beedy and Hamilton 1999), but rates of predation are highly variable in space and time and, until recently, predators have not been known to cause sustained reproductive failures of multiple colonies across a wide geographic area. Since 2006, predation by Cattle Egrets ( Bubulcus ibis) has caused nearly complete repro- ductive failures of even very large colonies in Tulare County (Meese 2012), but predation by Cattle Egrets on Tricolored Blackbird eggs and nestlings is unknown outside of Tulare County. The lack of influence of substrate type on reproductive success also suggests that predation is not responsible for the widespread and sustained reproductive failures I documented, as some substrates with thorns or stinging hairs, such as Himalayan blackberries and stinging nettles ( Urtica dioica), appear to confer some defense against preda- tors, while others lacking such armaments, such as cattails ( Typha latifolia ), are believed to expose nesting blackbirds to higher rates of predation (Beedy and Hamilton 1999, Cook and Toft 2005). Beedy and Hayworth (1992) documented the reproductive failure of a colony in Merced County and found that lethal levels of selenium were likely responsible for mortality of nestlings and the failure of this colony. To my knowledge, theirs is the only study that has attributed a colony’s failure to environmental contaminants, but because the potential role of environmental contaminants in reducing the blackbird’s productivity has not been assessed, it deserves further study. Food limitation is known to reduce reproductive success in many groups of birds (reviewed by Martin 1987), but my study provides evidence of food- limited reproduction that is spatially widespread and temporally persistent. The Tricolored Blackbird’s reproductive success was low despite the diverse array of foods the species consumes, as documented in this and previous studies (Skorupa et al. 1980), and despite the suggestion that a diverse diet 108 LOW REPRODUCTIVE SUCCESS OF THE COLONIAL TRICOLORED BLACKBIRD might allow the blackbird to exploit alternate food sources when a preferred food is limited by weather or other conditions (Crase and DeHaven 1977). During my study, a generalized diet did not appear to provide insurance against food-limited reproductive failures, and relatively high reproductive success was associated with exceptional abundance of a narrow range of favored insect groups (Table 1). Reproductive success did not differ by major geographical region (San Joa- quin vs. Sacramento Valley) or by county of the Central Valley. These results may affect management intended to increase the numbers of tricolors as they suggest there is no temporal or geographical basis for making conservation investments. Thus, all else being equal, efforts intended to benefit a year’s first attempts at breeding in the San Joaquin Valley should be as effective as efforts intended to benefit subsequent attempts in the Sacramento Valley. I found no effect of nesting substrate on reproductive success. This result differs from that of Cook and Toft (2005), who found that the proportion of colonies suffering complete reproductive failure was greater in native wet- lands than in upland substrates. Those authors attributed the differences to rates of predation being higher in wetlands than in uplands. However, Cook and Toft (2005) did not measure insect abundance in nearby foraging areas and did not directly observe predators in blackbird colonies, so comparing my results to theirs is difficult. If rates of predation are substrate-dependent, as suggested by Cook and Toft (2005), the absence of substrate-related dif- ferences suggests that predators are not responsible for the chronically low reproductive success I documented. The Tricolored Blackbird’s low reproductive success from 2006 to 2011 may help to explain the decline in its abundance observed from 2008 to 2011, when methodologically similar statewide surveys of the species found a drop from 400,000 to 258,000 birds (Kelsey 2008, Kyle and Kelsey 2011). Although additional field work is needed to determine whether this recent decline is part of a longer-term trend, the chronically low reproduc- tive success through the breeding season of 2011, which did not figure into the results of the 2011 statewide survey, suggests that the decline in abundance will continue. Since the 1980s, numerous large dairies have been located in the southern San Joaquin Valley, and dairies appear to be attractive to Tricolored Black- birds, as for decades some of the largest colonies have been established in grain fields adjacent to dairies (Beedy and Hamilton 1997; Figure 2). The grains stored to feed the dairy cattle appear to provide an ad libitum food source, and the large fields of triticale provide nesting substrate, as the stems of triticale plants are taller and stronger than are the stems of other grains and capable of supporting blackbird nests. Although the stored grains may provide a superabundant food supply, however, they appear to be insufficient to sustain breeding because the reproductive success of colonies adjacent to dairies is as low as that of colonies distant from dairies. To form eggs, breed- ing females require relatively high levels of essential amino acids and essential fatty acids, and these essential compounds are found in higher proportion in insects than they are in grains (Carey 1996, Ramsay and Houston 1998). In addition, nestling blackbirds require animal foods and do not eat plant materials for the first 9 days of life (Crase and DeHaven 1977, Skorupa et 109 LOW REPRODUCTIVE SUCCESS OF THE COLONIAL TRICOLORED BLACKBIRD al. 1980), until they are almost ready to fledge. The dietary requirements of breeding female and nestling blackbirds may help to explain why the re- productive success of colonies adjacent to dairies is low despite the virtually unlimited availability of grains. Additional research is needed to assess the effects of a granivorous diet on clutch size, egg hatchability, rates of brood reduction (Beedy and Hamilton 1999), and nestling starvation. The apparent attraction to and use of stored and provided grains extends beyond the breeding season and dairies in the San Joaquin Valley. Aggre- gations of thousands of birds are also observed around dairies and feedlots in winter, especially at dairies at Point Reyes National Seashore in Marin County, around feedlots near Birds Landing in southern Solano County, and at several sites in Merced County (pers. obs.). Given the large number of birds at sources of grain available ad libitum, both the breeding and winter distributions may be influenced if not determined by the sources of these grains. Across much of its range, the Tricolored Blackbird may be largely dependent on grains provided for livestock as a replacement for natural foods that have been lost to agriculture and urbanization, although my results sug- gest that a granivorous diet is insufficient to support the species’ breeding. The colonies I studied represent most of the largest colonies documented during this 6-year interval and represent the entire geographic range of the largest colonies of the Tricolored Blackbird (Kelsey 2008, Kyle and Kelsey 2011). Although these colonies had the greatest proportional potential to contribute to the productivity of the species, the chronic poor reproduc- tive success of the largest colonies suggests that they may not serve as an effective core of a conservation strategy. Colonies adjacent to dairies, for decades many of the largest colonies, appear to serve as ecological traps (Dwernychuk and Boag 1972), fledging relatively few young in most years. Entire colonies are also lost when the triticale in which the birds nest is har- vested during normal agricultural operations (Beedy and Hamilton 1997, Cook and Toft 2005, pers. obs.), although I included none of these harvested colonies in this analysis. During the 6 years of my study, the Tricolored Blackbird has experienced chronic, widespread low reproductive success apparently because insects are insufficient. These results support the view expressed by DeHaven et al. (1975) that reductions in the Tricolored Blackbird’s abundance from the 1930s to the 1970s were due at least partially to limitation of its food supply through loss of foraging habitat. This loss of foraging habitat may result in a decline in productivity over a period of years that is difficult to detect, but that decline may ultimately lead to the situation where, despite the avail- ability of suitable nesting substrate, tricolors abandon colonies or decline to extinction in an area where they formerly were abundant. This mechanism is believed to be responsible for the decimation of the species in southern California (Unitt 2004, Feenstra 2009) and Baja California (Erickson et al. 2007, Erickson and de la Cueva 2008). The relationships I describe here suggest that the Tricolored Blackbird’s habit of colonial breeding and the requirements of egg-forming females and nestlings to consume insects place a great burden upon landscapes within the 9-km radius of a colony within which the birds forage (Hamilton and Meese 2006) and that food limitation, not predation, is responsible for widespread, 110 LOW REPRODUCTIVE SUCCESS OF THE COLONIAL TRICOLORED BLACKBIRD chronic low reproductive success. These results suggest that future invest- ments in conservation should stress nesting substrates that are free from possible destruction during harvest and are surrounded by secure, productive foraging habitats. The variety of foraging substrates Tricolored Blackbirds use suggests a degree of flexibility in choices of conservation strategies, as both agricultural (alfalfa and sunflowers) as well as native (non-irrigated pasture) foraging habitats supported colonies with relatively high reproductive output (this study and unpubl. data). Where the blackbird’s productivity is a priority in agricultural settings, crops that serve as foraging substrates should not be sprayed with insecticides so as to maximize insect abundance. In regions of extensive pasturelands or grasslands where secure nesting substrates may be absent, management should stress the provision and maintenance of secure nesting substrates. As with foraging substrates, there are several options for nesting substrates, as nesting tricolors use a wide variety of native and non-native vegetation (Table 1). The conservation of the Tricolored Blackbird will require strategic choices that take into account the species’ unique needs. California’s Central Valley, where for decades the majority of blackbirds have occurred (Neff 1937, Beedy and Hamilton 1997), has been transformed from its origin as a vast region of wetlands and vernal pools (Frayer et al. 1989), capable of supporting mil- lions of birds (extrapolated from Neff 1937), to an agricultural heartland that provides much of the nation’s fresh fruits, nuts, and dairy products (American Farmland Trust 1989) but which is unsuitable for foraging blackbirds. The future of the Tricolored Blackbird depends upon resources in California’s Central Valley that appear in most years to be in short supply, and this future is made uncertain by competition between the needs of a colonial species and land uses that limit the range of strategic conservation choices. ACKNOWLEDGMENTS I am grateful to the U.S. Fish and Wildlife Service, the California Department of Fish and Game (now California Department of Fish and Wildlife), the Jiji Foundation, Audubon California, and several Audubon chapters for supporting this work. I thank the many property owners, who wish to remain anonymous, for allowing me to survey and monitor blackbird colonies and assess insect populations on their properties. I thank Steve Simmons of Merced for helping to provide access to several colonies in Merced County and Ted Beedy and Dave Shuford for thoughtful and rigorous reviews of the manuscript. LITERATURE CITED American Farmland Trust. 1989. Risk, challenges, and opportunities: Agriculture, resources, and growth in a challenging Central Valley. Am. Farmland Trust, San Francisco, CA. Ammon, E. M., and Woods, J. 2008. Status of Tricolored Blackbirds in Nevada. Great Basin Birds 10:63-66. Baird, S. F. 1870. Geological Survey of California. Ornithology. Volume I. Land Birds. Univ. Press: Welch, Bigelow, & Co., Cambridge, UK. Beedy, E. C., and Hamilton, W. J. III. 1997. Tricolored Blackbird status update and management guidelines. U.S. Fish and Wildlife Service and Calif. Dept. Fish and Game, Sacramento; http://tricolor.ice.ucdavis.edu/reports. Ill LOW REPRODUCTIVE SUCCESS OF THE COLONIAL TRICOLORED BLACKBIRD Beedy, E. C., and Hamilton, W. J. III. 1999. Tricolored Blackbird ( Agelaius tricolor), in The Birds of North America (A. Poole and F. Gill, eds.), no. 423. Birds N. Am., Inc., Philadelphia. Beedy, E. C., and Hayworth, A. 1992. Tricolored Blackbird ( Agelaius tricolor) nesting failures in the Central Valley of California: General trends or isolated phenomena?, in Endangered and Sensitive Species of the San Joaquin Valley, California (D.F. Williams, S. Byrne, and T.A. Rado, eds.), pp. 33-46. Calif. Energy Comm., Sacramento. Carey, C. 1996. Female reproductive energetics, in Avian Energetics and Nutritional Ecology (C. Carey, ed.), pp. 325-374. Chapman Hall, London. Center for Biological Diversity. 2004. Petition to list Tricolored Blackbird under the state and federal endangered species acts and request for emergency action to protect the species; www.biologicaldiversity.org/species/tricolored_blackbird/ pdfs/petition.pdf. Cook, L. F., and Toft, C. A. 2005. Dynamics of extinction: Population decline in the colonially nesting Tricolored Blackbird Agelaius tricolor. Bird Conserv. Int. 15:73-88. Crase, F. T., and DeHaven, R. W. 1977. Food of nestling Tricolored Blackbirds. Condor 79:265-269. DeHaven, R. W., Crase, F. T. and Woronecki, P. P. 1975. Breeding status of the Tricolored Blackbird, 1969-1972. Calif. Fish and Game 61:166-180. Dwernychuk, L. W., and Boag, D. A. 1972. Ducks nesting in association with gulls — an ecological trap? Can. J. Zool. 50:559-563. Erickson, R. A., and de la Cueva, H. 2008. Nesting Tricolored Blackbird survey: Baja California 2008. Report to the U.S. Fish and Wildlife Service; http: //tricolor, ice . ucdavis . edu/ reports . Erickson, R. A., de la Cueva, H., and Billings, M. J. 2007. Nesting Tricolored Black- bird survey: Baja California 2007. Report to the U.S. Fish and Wildlife Service; http ://tricolor. ice . ucdavis . edu/ reports . Feenstra, J. 2009. Results of the southern California Tricolored Blackbird survey 2009; http :// tricolor, ice . ucdavis . edu/ reports . Frayer, W. E., Peters, D. D., and Pywell, H. R. 1989. Wetlands of the California Central Valley: Status and trends. U.S. Fish and Wildlife Service, Portland, OR. Hamilton, W.J. III. 1998. Tricolored Blackbird itinerant breeding in California. Condor 100:218-226. Hamilton, W. J. Ill, and Meese, R. J. 2006. Habitat and population characteristics of Tricolored Blackbird colonies in California. Report to Calif. Dept. Fish and Game, Habitat Conservation Planning Branch, 1416 Ninth St., Suite 1260, Sacramento, CA 95814; http://tricolor.ice.ucdavis.edu/reports. Hoffman, R. 1927 (reprinted 1955). Birds of the Pacific States. Houghton Mifflin, New York. Kelsey, R. 2008. Results of the Tricolored Blackbird 2008 census. Audubon Califor- nia, Winters, CA; http://tricolor.ice.ucdavis.edu/reports. Kyle, K., and Kelsey, R. 2011. Results of the 2011 Tricolored Blackbird statewide survey. Audubon California, Sacramento; http://tricolor.ice.ucdavis.edu/reports. Martin, T. E. 1987. Food as a limit on breeding birds: A life-history perspective. Annu. Rev. Ecol. Syst. 18:453-487. Meese, R. J. 2011. Reproductive success of Tricolored Blackbird colonies in 2011 in the Central Valley of California. Calif. Dept, of Fish and Game, Wildlife Branch, Nongame Wildlife Program Report 2011-08, Sacramento, CA; http://nrm.dfg. ca.gov/FileHandler.ashx?DocumentVersionID=62198. Meese, R. J. 2012. Cattle Egret predation causing reproductive failures of nesting Tricolored Blackbirds. Calif. Fish and Game 98:47-50. Neff, J. 1937. Nesting distribution of the Tricolored Red-wing. Condor 39:61-81. 112 LOW REPRODUCTIVE SUCCESS OF THE COLONIAL TRICOLORED BLACKBIRD Ramsay, S. L., and Houston, D. C. 1998. The effect of dietary amino acid composition on egg production in Blue Tits. Proc. Royal Soc. London B 265:1401-1405. Shuford, W. D., and Gardali, T., eds. 2008. California Bird Species of Special Con- cern: A ranked assessment of species, subspecies, and distinct populations of birds of immediate conservation concern in California. Studies of Western Birds 1. W. Field Ornithol., Camarillo, CA, and Calif. Dept. Fish and Game, Sacramento. Skorupa, J., Hothem, R. L., and DeHaven, R. W. 1980. Foods of breeding Tri- colored Blackbirds in agricultural areas of Merced County, California. Condor 82:465-467. Unitt, P. 2004. San Diego County bird atlas. Proc. San Diego Soc. Nat. Hist. 39. U.S. Fish and Wildlife Service. 2008. Birds of Conservation Concern 2008. U.S. Fish and Wildlife Service, Div. Migratory Bird Mgmt., Arlington, VA; http: //library. fws.gov/bird_publications/bcc2008.pdf. Accepted 27 October 2012 113 DIET AND HOME-RANGE SIZE OF CALIFORNIA SPOTTED OWLS IN A BURNED FOREST MONICA L. BOND, DEREK E. LEE, RODNEY B. SIEGEL, and MORGAN W. TINGLEY, The Institute for Bird Populations, P. O. Box 1346, Point Reyes Station, California 94956 (current address of Bond and Lee: Wild Nature Institute, P.O. Box 165, Hanover, New Hampshire 03755); rsiegel@birdpop.org ABSTRACT: Fire is pervasive in forests used by California Spotted Owls ( Strix occidentalis occidentalis) and their prey species. We assessed the diets and sizes of the breeding-season home ranges of seven Spotted Owls occupying burned forests in the southern Sierra Nevada 4 years after a fire and compared the results with data from previous studies in unburned forests within the range of the subspecies. Prey captured by owls in the burned area comprised 40.3% (by biomass) pocket gophers (' Thomomys spp.) and 25.9% northern flying squirrels ( Glaucomas sabrinus). In contrast, in unburned areas of the Sierra Nevada Spotted Owls fed primarily on fly- ing squirrels, or on both flying squirrels and woodrats ( Neotoma spp); in unburned southern California forests they fed overwhelmingly on woodrats. The owls’ mean home range in the burned forest covered 402 ha, an area similar to that recorded in unburned forests of the Sierra Nevada. Our results are consistent with hypotheses that the burned habitat in our study area was rich in gophers and that Spotted Owls foraging on gophers in burned forests do not require home ranges substantially larger than do owls in unburned forests. With currently available data we could not conclusively attribute variation in diet or home-range size to the influence of fire, so further testing is warranted. Use of rodenticides and herbicides in managing burned Spotted Owl habitat may reduce the owl’s key prey. The California Spotted Owl ( Strix occidentalis occidentalis) occurs in mixed-conifer and montane hardwood forests throughout the Sierra Nevada and the mountain ranges of southern California (Verner et al. 1992, Gutier- rez et al. 1995). Its vital rates are positively associated with large areas of older conifer-dominated forest and related conditions (Blakesley et al. 2005, Seamans 2005), and within its range the subspecies is viewed as an indicator for management of these forest types. Forest fire is one of the most important issues affecting the Spotted Owl’s habitat (Weatherspoon et al. 1992). California Spotted Owls typically occupy older forests with greater canopy cover (Bias and Gutierrez 1992, Call et al. 1992, Gutierrez et al. 1992, Moen and Gutierrez 1997, Bond et al. 2004), and this has led to the presumption that fires of moderate and high severity reduce the owl’s survival and pose a significant risk to its habitat (Weath- erspoon et al. 1992, U.S. Forest Service 2001, 2004, 2005). However, Spotted Owls can occupy territories and continue to reproduce in burned habitats, including those with large severely burned patches (Gaines et al. 1997, Bond et al. 2002, Jenness et al. 2004, Bond et al. 2009). Franklin et al. (2000) hypothesized that fire could enhance the abundance of and access to prey for the Northern Spotted Owl (S. o. caurina) in California by creating patchy openings in the forest canopy and increasing habitat edges. Indeed, four years after a large, patchy fire in the southern Sierra Nevada, California Spotted Owls selectively foraged in burned forest rather than unburned forest, with the strongest selection for severely burned patches (Bond et al. 2009). 114 Western Birds 44:114-126, 2013 CALIFORNIA SPOTTED OWLS IN A BURNED FOREST Prey composition is an important determinant of raptors’ use of space (Peery 2000). Zabel et al. (1995) found that prey species predicted the size of the Northern Spotted Owl’s home range in California better than the proportion of older forest in the home range, and Carey et al. (1992) re- ported that the prey base strongly influenced the size of the Northern Spotted Owl’s home range in Oregon. In both studies, owls with a greater percent biomass of larger prey in their diet had home ranges smaller than those of owls that consumed more of the smaller prey species. Ward et al. (1998) documented that in California Northern Spotted Owls selected foraging sites according to the distribution of larger prey (e.g., woodrats), which provided an energetic benefit to the owls. The California Spotted Owl’s major prey items by biomass are the dusky-footed and big-eared woodrats ( Neotoma fuscipes and N. macrotis), northern flying squirrel ( Glaucomys sabrinus), and pocket gophers (' Thomomys spp.; Thraikill and Bias 1989, Williams et al. 1992, Smith et al. 1999, Munton et al. 2002). Despite the importance of fire in shaping ecosystems in portions of the Spotted Owl’s range, little is known about the diets and home-range sizes of owls occupying recently burned forests. Understanding the California Spotted Owl’s prey base and space use in burned forests is important because fire is frequent in this subspecies’ habitat, and it is not unusual for owls to continue to occupy and breed in burned forests (Bond et al. 2002, 2009, Roberts et al. 2011). Furthermore, forest management after a fire can include removal of fire-killed trees and applications of rodenticide and herbicide in efforts to enhance the growth of tree seedlings (McGinnis et al. 2010, Swanson et al. 2011), possibly to the detriment of foraging Spotted Owls and their prey. Our objective was to assess the breeding-season diet and estimate the sizes of the home ranges of California Spotted Owls occupying territories in recently burned mixed-conifer forests and compare the results with those of previous studies of owls in unburned forests. We note that the breeding-season home range may differ greatly from the year-round home range, as we have reported previously on the expanded winter movements of the same owls addressed in this study (Bond et al. 2010). Our sample of owls in burned forests was relatively small (seven owls from four territories), our study included only a single breeding season, and the data from unburned forests we used for comparison were recorded in different years and in different areas. Therefore, rather than testing how fire affected diets and home-range sizes, we report for the first time basic information about the diet and home-range sizes of the California Spotted Owl in a burned forest and address differences from those in unburned forests qualitatively — a basis for developing hypotheses about fire effects to be tested in future studies. METHODS Study Areas We estimated the diets and home ranges of California Spotted Owls during the breeding season of 2006 in forests burned by the 2002 McNally Fire in the Greenhorn Mountains and Kern Plateau in the Sequoia National Forest, 115 CALIFORNIA SPOTTED OWLS IN A BURNED FOREST southern Sierra Nevada. Elevation ranged from 1500 m to 2500 m. In July and August 2002, the McNally Fire burned approximately 60,985 ha in the Sequoia and Inyo national forests, including 33,704 ha of conifer-dominated forests (Odion and Hanson 2006). Like most forest fires, the McNally Fire burned with variable severity, leaving a mosaic of severely burned, moder- ately burned, and unburned patches in the study area (Bond et al. 2009, 2010). Fires of high severity kill most existing vegetation and result in a forest structure dominated no longer by live trees but by herbs, shrubs, and dead trees, those of moderate severity cause patchwork mortality, and those of low severity rarely kill overstory trees and do not substantially alter the forest’s structure (Smith 2000). For comparison with our burned study area we obtained data from previ- ous studies of the diets and home-range sizes of California Spotted Owls in unburned forests elsewhere within the subspecies’ range (Thrailkill and Bias 1989, Call et al. 1992, Zabel et al. 1992, Smith et al. 1999, Zimmer- man et al. 2001, Munton et al. 2002). Of these studies the nearest area matching ours in elevation was in the southern Sierra Nevada in the Sierra National Forest at 1220-2925 m elevation, 150 km north of the McNally Fire area. There Munton et al. (2002) studied the owl’s diet and Zabel et al. (1992) estimated the sizes of home ranges of radio-marked owls during the breeding season of 1988. Farther north in the central Sierra Nevada, in the Eldorado (elevation 266-2257 m) and Tahoe national forests (670-1585 m), respectively, Thrailkill and Bias (1989) studied Spotted Owl diets during the 1986 and 1987 breeding seasons and Call et al. (1992) estimated the sizes of home ranges of radio-marked Spotted Owls during the 1987 breeding season. South of the McNally Fire area in the San Bernardino Mountains, Smith et al. (1999) assessed the owl’s diet (elevation 1000-2500 m) and Zimmermann et al. (2001) estimated home-range sizes by radio telemetry (elevation 2000-2500 m). In all the study areas summers are dry and hot, winters are wet and cold. At the Sierra Nevada study areas the vegetation is classified as Sierran Mixed Conifer (Allen 1988), dominated by ponderosa pine ( Pinus ponderosa), Jeffrey pine ( P jeffreyi), and white fir ( Abies concolor). Other common trees include sugar pine [P. lambertiana), incense-cedar ( Calocedrus de- currens), canyon live oak ( Quercus chrysolepis), and California black oak (Q. kelloggii). Above 2100 m, a transition zone is dominated by red fir (A. magnified). Study areas in the central Sierra Nevada also included Douglas- fir ( Pseudotsuga menziesii). In the San Bernardino Mountains study area the forest is dominated by bigeone Douglas-fir ( Pseudotsuga macrocarpa), canyon live oak, coast live oak ( Quercus agrifolia ), and black oak below 1500 m, by Jeffrey pine, ponderosa pine, white fir, black oak, canyon live oak, sugar pine, incense-cedar, and western juniper {Juniperus occidentalis) above 1500 m (Stephenson and Calcarone 1999). Historically, fire regimes varied but most fires were of mixed severity (Skinner and Chang 1992, Weatherspoon et al. 1992, Minnich et al. 1995). 116 CALIFORNIA SPOTTED OWLS IN A BURNED FOREST Field Methods Prior to our study, U.S. Forest Service (USFS) personnel surveyed for and located California Spotted Owls within and near the McNally Fire’s perimeter during three consecutive breeding seasons, 2003-2005 (W. Rannals and R. Galloway, unpubl. data). During the 2006 breeding season we surveyed for California Spotted Owls by standard methods described by Franklin et al. (1996). We selected four territories confirmed to be occupied by pairs of Spotted Owls. We selected these territories because they were inside or within 1 km of the fire perimeter, Spotted Owls had occupied them continuously since the fire, and there was sufficient road access for effective radio-tracking. We captured the adult owls with mist nets or noose poles and fitted them with color bands and backpack-style radio transmitters (batteries with a life expectancy of 1 year) that were designed to minimize contact with the owl’s back (AVM Instrument Company, Colfax, CA). Radio transmitters were affixed with Kevlar ribbon (0.63 cm wide; Bally Ribbon Mills, Bally, PA). With the harness, the transmitter units weighed <20 g, or <4% of each owl’s body mass. We considered each individual owl as a separate sample because Call et al. (1992) and Zimmerman et al. (2001) found that male and female California Spotted Owls forage independently. All four pairs attempted to nest in 2006, and we located each nest tree. Three attempts failed (Mill Creek, Speas Ridge, Burnt Ridge), and one pair successfully fledged one young (Poison Creek). From 22 May to 15 August 2006 we radio-tracked the seven owls (4 males and 3 females) by locating each owl nightly or every other night between 21 :30 and 04:00. We used a systematic design with a random start to record a representative sample of foraging locations. We randomly assigned each owl to a 1-hour time block on the first night we tracked it, then systemati- cally shifted each owl’s time block by 1 hour (or back to the first time block when the sequence was completed) on each subsequent night of sampling. To estimate the owls’ locations, we triangulated on signals from the radio transmitters from established stations and used at least three compass bear- ings of the strongest signals, taking all bearings within approximately 30 minutes (Guetterman et al. 1991). For comparison of the sizes of home ranges in our burned study area with those of owls occupying unburned areas, D. Call, T. Munton, and G. Zimmerman provided us with locations of foraging radio-marked Spotted Owls in the Tahoe (n = 7), Sierra (n = 8), and San Bernardino (n = 5) national forests, respectively, during the same seasonal window of mid-May to mid-August. These studies used field methods similar to ours to estimate locations of foraging (Call et al. 1992, Zabel et al. 1992, Zimmerman et al. 2001). In addition to estimating locations of nighttime foraging, every 7-10 days we recorded the location of each radio-marked owl’s daytime roost. To assess the owls’ diets, we collected regurgitated pellets at roost sites. We included multiple pellets found together in the same location as a single sample because remains of a single animal can be egested in more than one pellet (Forsman et al. 1984). 117 CALIFORNIA SPOTTED OWLS IN A BURNED FOREST Data Analyses We used digital maps and the GIS extension Fragstats (McGarigal et al. 2002) to calculate the area of suitable vegetation burned at high severity within a 2-km radius around the nest of each pair of owls. We chose this radius because >90% of each owl’s foraging locations fell within it. We de- fined suitable vegetation as conifer and hardwood forest, which are those the California Spotted Owl typically uses (Verner et al. 1992). We report area burned at moderate and high severity because at these severities a fire alters a forest’s structure substantially (Smith 2000) and because most managers consider these levels a threat to the California Spotted Owl (USFS 2004). Cover types and fire severity were derived from Landsat Thematic Mapper imagery developed by the USFS Pacific Southwest Region Remote Sensing Lab and available from the GIS Clearinghouse (http://www.fs.fed.us/r5/ rsl/ clearinghouse/) . We dissected all California Spotted Owl pellets collected from the McNally Fire study area and isolated all identifiable prey remains, including skulls, mandibles, other bones, bills, feathers, fur, and insect exoskeletons. We used the collection of the California Academy of Sciences, San Francisco, as a reference to identify remains of birds and mammals. We quantified the number of individuals of each species of prey in each sample as the high- est count of diagnostic skeletal parts of each species. For example, if there were three right and two left maxillae of Neotoma macrotis in a pellet we considered the number of individuals of this species as three. We estimated the number of insects from the largest number of heads or paired mandibles. We identified two prey taxa (Glaucomas sabrinus and N. macrotis) to spe- cies and four others (' Thomomys , Peromyscus, Microtus, and Scapanus) to genus only. The remaining prey items were identified to one of four broader taxonomic categories (diurnal squirrel, bird, bat, and insect). We calculated the percent frequency of each category in all pellet samples and estimated the percent biomass for each category from average weights (g) provided by Smith et al. (1999). For each of the four territories we tallied prey items by percent biomass but were unable to separate samples by sex because both members of a pair often roosted together in the same tree. The previously published studies of contents of California Spotted Owl pellets did not pres- ent the data by territory (Thrailkill and Bias 1989, Smith et al. 1999, Munton et al. 2002). Therefore, in addition to estimating the diet for each territory, we also pooled samples from all four territories for qualitative comparisons with those of other studies. To estimate the sizes of the owls’ home ranges, we used a kernel-density estimator, a widely used statistical method (Seaman and Powell 1996). Numerous methods for estimating home ranges with kernel estimators are available, but each method can generate a different estimate. In our analysis we used a fixed-bandwidth kernel estimator, which has been shown to have little bias in comparison to adaptive-bandwidth kernels (Seaman and Powell 1996). The choice of a smoothing parameter or bandwidth, h, to be used in an analysis affects the results critically, and there are several options for calculating the “optimal” bandwidth. The two most widely used methods for choosing h are the reference bandwidth, h re f (Worton 1995) and least- 118 CALIFORNIA SPOTTED OWLS IN A BURNED FOREST squares cross validation (LSCV; Silverman 1986, Wand and Jones 1995). The LSCV method produces more accurate estimates of simulated home ranges than does h re f (Seaman and Powell 1996), but LSCV does not always converge, preventing its use in some cases. We estimated the sizes of the owls’ home ranges by kernel density in a two-step process. First, we analyzed sample sizes to test whether the number of points where each owl was relocated was sufficient for unbiased estimation of its home range. We used a bootstrapping algorithm in which, for each owl, we drew at random an increasing number of relocation points (n) from which we built a kernel-density utilization distribution resulting in an area estimate. We began by sampling 5 points and continued until all the available points for each owl had been drawn. At each n, we repeated the process 200 times, to calculate an incremental mean and standard deviation of kernel-density area. We used the simple h ref method because small samples of points did not allow convergence with LSCV. We evaluated the simulation results for asymptotic convergence on home-range size as sample size increased, and if values for an individual did not converge, we excluded it from further analysis. We used owl locations from the unburned study areas that had been recorded during the same seasonal window of mid-May to mid-August, and ran the same power analyses by using bootstrapping simulations. Second, after the sample-size analysis, we calculated the final home-range size as the 95 th percentile of the fixed-kernel density estimated with LSCV For all these calculations we used the package AdehabitatHR (Calenge 2006) in R, version 2.14. RESULTS The average area of suitable (conifer and hardwood) forest burned at moderate severity within 2 km of the four Spotted Owl nests we studied in or near the McNally Fire area was 237.9 ha (SD = 119.1, range 110.9- 397.5); that which burned at high severity was 93.3 ha (SD = 35.5 ha, range 42. 1-1 18.8 ha). Thus an average of 23% (range 10-39%) of suitable vegetation within the 2-km radius burned at moderate severity and 9% (range 4-12%) burned at high severity (Table 1). We identified 199 individual prey items from 101 pellet samples from Spotted Owls in the McNally Fire area (range 7-35.5 pellet samples per owl territory). Samples from all four territories contained a substantial proportion of gophers and flying squirrels; big-eared woodrats were present in pellets at only two of the territories (Table 1). In the pooled sample of pellets from all four territories, the prey biomass comprised 40% gophers, 26% flying squirrels, and 11% woodrats. The dominant species of prey that Spotted Owls captured in the closest unburned study area (Sierra National Forest, 1988) was the northern flying squirrel, 46% by biomass; gophers represented 18% and woodrats 12% (Munton et al. 2002; Table 2). In the Eldorado National Forest Spotted Owls primarily captured woodrats (38% of prey by biomass) and flying squirrels (31%; Table 2). In the unburned San Bernardino National Forest study area, woodrats were the overwhelmingly dominant prey, constituting 74% of the owls’ diets by biomass (Smith et al. 1999; Table 2). 119 CALIFORNIA SPOTTED OWLS IN A BURNED FOREST Table 1 Percent Each Category of Prey Contributed to Total Dietary Bio- mass, Estimated Home Range, and Severity of Fire within 2 km of Four Cali- fornia Spotted Owl Nests Four Years after the 2002 McNally Fire, Sequoia National Forest, California Site Prey item Poison Creek Mill Creek Speas Ridge Burnt Ridge Thomomys spp. 18.9 48.6 40.4 53.3 Neotoma macrotis 35.4 0.0 0.0 8.3 Glaucomys sabrinus 20.7 35.9 35.3 11.6 Peromyscus spp. 5.2 8.6 2.4 1.6 sciurid 8.1 6.7 9.8 10.5 Microtus spp. 4.1 0.0 6.5 0.0 No. prey items 83 13 44 59 Male home-range size 718.0 (40) 252.8(51) 307.3 (48) 606.3 (53) (ha) (n) a Female home-range size 129.8 (33) 606.0 (45) 197.3 (34) (ha) (n) b Percent burned at 38.9 9.8 19.3 23.5 moderate severity Percent burned at high 11.6 3.7 9.0 11.5 severity Average elevation (m) 2167 1679 2052 1833 a n = Number of night-time relocation points per owl. b Female at Mill Creek was not radio-marked. The mean number of independent relocations we obtained for each for- aging Spotted Owl in the McNally Fire area was 43 (range 33-53; Table 1). Simulation results for all owls in the McNally Fire converged and so we included them in the analysis (Figure 1). We excluded data from three Spot- ted Owls in the sample from unburned forests because of lack of asymptotic convergence (TH5, TH6, and SB1; Figure 1). The size of the home range of a Spotted Owl in the McNally Fire area averaged 402.5 ha (SE = 88.7, range 129.8-718.0 ha; Tables 1 and 3). The individual with the largest home range was the sole male that reproduced successfully that year (Poison Creek), while his mate’s home range was the smallest (Table 1). Home-range sizes in unburned forests averaged 487.0 ha (SE = 63.9 ha) in the Tahoe National Forest, 529.0 ha (SE = 72.9 ha) in the Sierra National Forest, and 370.4 ha (SE = 58.7 ha) in the San Bernardino National Forest (Table 3). DISCUSSION Here we report the first data on the diet and the size of home ranges of the California Spotted Owl in a burned landscape. The size of an owl’s home range varies with many factors, including but not limited to forest structure, fire severity, key prey species and temporal fluctuations in their numbers, the owl’s foraging behavior, and whether it is nesting or feeding young (Carey et al. 1992, Zabel et al. 1995, Ward et al. 1998). We could not quantify home-range size rigorously as a function of diet, habitat, or other factors because our sample of owls and the length of our study were not sufficient. 120 CALIFORNIA SPOTTED OWLS IN A BURNED FOREST Table 2 Average Percent Each Category of Prey Contributed to Total Dietary Biomass in Pooled Samples of Pellets Regurgitated by Spotted Owls in an Area Burned by the McNally Fire, Sequoia National Forest, and in Unburned Areas in the Eldorado, Sierra, and San Bernardino National Forests, California 0 Eldorado NF b n = 139 Sierra NF C n = 1269 McNally Fire n = 199 San Bernardino NF d n = 8441 Thomomys spp. 7.8 18.4 40.3 10.4 Glaucomys sabrinus 30.7 45.6 25.9 3.0 Neotoma spp. 38.1 11.8 10.9 74.0 Diurnal squirrel 6.6 0.9 8.8 1.3 Bird 12.4 12.9 4.0 3.5 Peromyscus spp. 1.3 5.5 4.5 4.0 Microtus spp. 0.7 1.2 2.6 1.3 Scapanus spp. 2.1 1.5 1.5 0.3 Bat 0.0 0.1 0.8 0.1 Insect 0.3 0.6 0.6 1.2 Sylvilagus spp. 0.0 0.0 0.0 1.0 “Sample size is number of individual prey items. Bold numbers indicate greatest percent biomass by study area. b Data from Thrailkill and Bias (1989). “Data from Munton et al. (2002). d Data from Smith et al. (1999). Previous studies of the California Spotted Owl’s breeding-season diet (Thrailkill and Bias 1989, Smith et al. 1999, Munton et al. 2002) and home- range size (Call et al. 1992, Zabel et al. 1992, Zimmerman et al. 2001) were conducted in unburned areas. The dominant prey item we identified in pellets from the McNally Fire area was the pocket gopher; the northern flying squirrel and big-eared woodrat were also important. In comparison, the dominant prey California Spotted Owls took during the breeding season in an unburned area in the Sierra National Forest in the southern Sierra Nevada was the northern flying squirrel, proportions of pocket gophers and woodrats being smaller (Munton et al. 2002; Table 2). In the Eldorado National Forest in the central Sierra Nevada Spotted Owls consumed mainly woodrats and flying squirrels, while to the south in an unburned area in the San Bernardino National Forest, woodrats were the overwhelming prey (Smith et al. 1999). In the McNally Fire area, owl pellets contained woodrat remains in amounts comparable to those in the nearest unburned study area at simi- lar elevation (Sierra National Forest). But the average percent biomass of gophers in pellets from the McNally fire area was more than twice that in pellets from the unburned study area in the Sierra National Forest and more than 5 times higher than in pellets from the unburned Eldorado National Forest study area. Gophers are found in a variety of habitat types in the Sierra Nevada but are uncommon in mature and older forests with little or no herbaceous ground cover (Williams et al. 1992). They feed largely on 121 CALIFORNIA SPOTTED OWLS IN A BURNED FOREST A B Figure 1 . Mean (black line) and standard deviation (gray shading) of sizes of fixed-kernel home ranges for increasing random draws of n relocation points for representative radio- tracked California Spotted Owls from studies in multiple national forests in California. Plots were assessed for evidence of convergence in home range as n increased. Three home ranges did not converge, two in the Tahoe National Forest (A, B), and one in the San Bernardino National Forest (C). All other home ranges converged asymptotically (e.g., for a bird on San Bernardino National Forest, D). herbaceous grasses and forbs and during the winter frequently consume mountain whitethorn ( Ceanothus cordulatus) leaves and willow ( Salix spp.) stems (Williams et al. 1992). Four years after the McNally Fire, stands of severely burned forest had significantly greater cover of shrubs and herbs and more dead trees than unburned stands or stands burned at moderate or low severity (Bond et al. 2009). Elsewhere we reported that California Spotted Owls selectively foraged in burned forest over unburned, with the strongest selection for severely burned patches (Bond et al. 2009). Forest fire may have enhanced the habitat for gophers (sensu Williams et al. 1992), and the standing fire-killed trees offered perches for foraging Spotted Owls, resulting in greater biomass of gophers in owl pellets from the McNally Fire area than in unburned areas of the Sierra Nevada. In the McNally Fire area, Spotted Owl pellets contained less biomass of northern flying squirrels than did those from unburned areas of the Sierra 122 CALIFORNIA SPOTTED OWLS IN A BURNED FOREST Table 3 Characteristics of Home Ranges Estimated by the Fixed-Kernel Method of California Spotted Owls Four Years after the 2002 McNally Fire, Sequoia Na- tional Forest, and in Three Unburned Areas Study area Tahoe NF a Sierra NF b Sequoia NF McNally Fire San Bernardino NF C Latitude (° N) Elevation (m) Mean home range (ha) SE of home range Range of home range 40 670-1585 487.0 63.9 284.8-682.3 37 1220-2925 529.0 72.9 253.3-718.5 36 1500-2500 402.5 88.7 129.8-718.0 34 2000-2500 370.4 58.7 242.0-526.5 a From D. Call, unpublished data. fa From T. Munton, unpublished data. c From G. Zimmerman, unpublished data. Estimated from locations of night-time foraging recorded 22 May-15 August 2006. Nevada. Waters and Zabel (1995) found that the mean density of the flying squirrel in old-growth forest in northeastern California was greater than in young stands regenerating naturally after forest fire. On the basis of its habitat requirements (Waters and Zabel 1995, Meyer et al. 2007), the northern flying squirrel may take longer than the Spotted Owl’s other prey to recover after severe forest fire. Future research should test the assumptions about the effects of fire on the California Spotted Owl’s major prey species, with our results providing a basis for developing hypotheses. The mean size of a home range in the McNally Fire area was 402 ha, similar to that in the unburned study areas, as evidenced by overlapping standard errors. The mean home range in our burned study area, however, was 24% smaller than in the nearest unburned area of similar elevation (Sierra National Forest), and pellets contained a greater biomass of gophers. In comparison, owls in the Sierra National Forest study area had the largest mean home ranges and were foraging primarily on flying squirrels. Our results provide the basis for testable hypotheses on the effects of fire on the California Spotted Owl’s use of space. As the biomass of an individual gopher and northern flying squirrel is similar (Smith et al. 1999), future studies could test the hypothesis that owls foraging primarily on gophers and flying squirrels may not need to forage as widely in burned forests as in comparable unburned forests. Studies could examine whether fire enhanced habitat for gophers and degraded it for flying squirrels, while controlling for any spatial and temporal differences in densities of these two prey species. Our study of Spotted Owls occupying burned forests documented diets containing a majority of pocket gophers by biomass, with woodrats also a major prey item in one territory (Table 1). Shrubs and herbaceous vegetation provide important habitat for both of these key species of prey (Williams 1992), and these habitats can be abundant in a burned landscape (Bond et al. 2009). Therefore, post-fire management involving use of rodenticides to prevent damage to young trees or herbicides to suppress shrub growth within California Spotted Owl habitat may impair the owls’ foraging. 123 CALIFORNIA SPOTTED OWLS IN A BURNED FOREST ACKNOWLEDGMENTS We thank K. Bruno and L. Kehas for field assistance and W. Rannals and R. Gal- loway of the Sequoia National Forest for providing Spotted Owl locations prior to our field work and valuable support during our study. We are indebted to D. Call, T. Munton, and G. Zimmerman for providing us with their telemetry data and to J. P. Ward for advice on field methods and data analysis. The California Academy of Sci- ences graciously allowed us access to its collection for our pellet analysis. We thank R. Gill, D. Herter, and D. Wiens for helpful comments on the manuscript. This study was funded by grants from Environment Now and the Resources Legacy Fund Foundation to The Institute for Bird Populations and was conducted by the Institute's Sierra Ne- vada Bird Observatory. This is contribution 377 of The Institute for Bird Populations. LITERATURE CITED Allen, B. 1988. Sierran mixed conifer, in A Guide to Wildlife Habitat Relationships (K.E. Mayer and W.F. Laudenslayer, eds.), pp. 46-47. Calif. Dept. Fish and Game, Sacramento. Bias, M. A., and Gutierrez, R. J. 1992. Habitat associations of California Spotted Owls in the central Sierra Nevada. J. Wildlife Mgmt. 56:584-593. Blakesley, J. A., Noon, B. R., and Anderson, D. R. 2005. Site occupancy, apparent survival, and reproduction of California Spotted Owls in relation to forest stand characteristics. J. Wildlife Mgmt. 69:1554-1564. Bond, M. L., Gutierrez, R. J., Franklin, A. B., LaHaye, W. S., May, C. A., and Sea- mans, M. E. 2002. Short-term effects of wildfires on Spotted Owl survival, site fi- delity, mate fidelity, and reproductive success. Wildlife Soc. Bull. 30: 1022-1028. Bond, M. L., Gutierrez, R. J., and Seamans, M. E. 2004. Modeling nesting habitat selection of California Spotted Owls ( Strix occidentalis occidentalis ) in the central Sierra Nevada. Forest Sci. 50:773-780. Bond, M.L., Lee, D. L., Siegel, R. S., and Ward, J. P. 2009. Habitat use and se- lection by California Spotted Owls in a postfire landscape. J. Wildlife Mgmt. 73:1116-1124. Bond, M. L., Lee, D. E., and Siegel, R. B. 2010. Winter movements by California Spotted Owls in a burned landscape. W. Birds 41:174-180. Calenge, C. 2006. The package adehabitat for the R software: A tool for the analysis of space and habitat use by animals. Ecol. Modelling 197:516-519. Call, D. R., Gutierrez, R. J., and Verner, J. 1992. Foraging habitat and home- range characteristics of California Spotted Owls in the Sierra Nevada. Condor 94:880-888. Carey, A. B., Horton, S. P., and Biswell, B. L. 1992. Northern Spotted Owls: Influ- ence of prey base and landscape character. Ecol. Monogr. 62:223-250. Forsman, E. D., Meslow, E. C., and Wight, H. M. 1984. Distribution and biology of the Spotted Owl in Oregon. Wildlife Monogr. 87:1-64. Franklin, A. B., Anderson, D. R., Forsman, E. D., Burnham, K. P., and Wagner, F. W. 1996. Methods for collecting and analyzing demographic data on the Northern Spotted Owl. Studies in Avian Biology 17:12-20. Franklin, A. B., Anderson, D. R. Gutierrez, R. J., and Burnham, K. P. 2000. Climate, habitat quality, and fitness in Northern Spotted Owl populations in northwestern California. Ecol. Monogr. 70:539-590. Gaines, W. L., Strand, R. A., and Piper, S. D. 1997. Effects of the Hatchery com- plex fires on Northern Spotted Owls in the eastern Washington Cascades, in Proceedings of the Fire Effects on Rare and Endangered Species and Habitats Conference (J. N. Greenlee, ed.), pp. 123-129. Int. Assoc. Wildfire and Forestry, Coeur d’Alene, ID. 124 CALIFORNIA SPOTTED OWLS IN A BURNED FOREST Guetterman, J. H., Burns, J. A., Reid, J. A., Horn, R. B., and Foster, C. C. 1991. Radio telemetry methods for studying Spotted Owls in the Pacific Northwest. U. S. Forest Serv. Gen. Tech. Rep. PNW-GTR-272. Gutierrez, R.J ., Verner, J., McKelvey, K. S., Noon, B. R., Steger, G. N., Call, D. R., LaHaye, W. S., Gingham, B. B., and Senser, J. S. 1992. Habitat relations of the California Spotted Owl, in The California Spotted Owl: A Technical Assessment of its Current Status (J. Verner, K. S. McKelvey, B. R. Noon, R. J. Gutierrez, G. I. Gould, Jr., and T. W. Beck, tech, coords.), pp. 79-99. U. S. Forest Serv. Gen. Tech. Rep. PSW-GTR-133. Gutierrez, R. J, Franklin, A. B., and LaHaye, W. S. 1995. Spotted Owl (Strix oc- cidentalis), in The Birds of North America (A. Poole and F. Gill, eds.), no. 179. Acad. Nat. Sci., Philadelphia. Jenness, J. J., Beier, P, and Ganey, J. L. 2004. Associations between forest fire and Mexican Spotted Owls. Forest Sci. 50:765-772. McGinnis, T. W., Keeley, J. E., Stephens, S. L., and Roller, G. B. 2010. Fuel buildup and potential fire behavior after stand-replacing fires, logging fire-killed trees and herbicide shrub removal in Sierra Nevada forests. Forest Ecol. and Mgmt. 260:22-35. Meyer, M. D., Kelt, D. A., and North, M. P. 2007. Microhabitat associations of northern flying squirrels in burned and thinned forest stands of the Sierra Nevada. Am. Midland Nat. 157:202-211. Minnich, R. A., Barbour, M. G., Burk, J. H., and Fernau, R. F. 1995. Sixty years of change in Californian conifer forests of the San Bernardino Mountains. Conserv. Biol. 9:902-914. Moen, C. A., and Gutierrez, R. J. 1997. California Spotted Owl habitat selection in the central Sierra Nevada. J. Wildlife Mgmt. 61:1281-1287. Munton, T. E., Johnson, K. D., Steger, G. N., and Eberlein, G. P. 2002. Diets of California Spotted Owls in the Sierra National Forest, in Proceedings of a symposium on the Kings River Sustainable Forest Ecosystems Project: Progress and current status (J. Verner, K. S. McKelvey, B. R. Noon, R. J. Gutierrez, G. I. Gould, Jr., and T. W. Beck, tech, coords.), pp. 99-105 U.S. Forest Serv. Gen. Tech. Rep. PSW-183. Odion, D. C., and Hanson, C. T. 2006. Fire severity in conifer forests of the Sierra Nevada, California. Ecosystems 9:1177-1189. Peery, M. Z. 2000. Factors affecting interspecies variation in home-range size of raptors. Auk 117:511-517. Roberts, S. L., van Wagtendonk, J. W., Miles, A. K., and Kelt, D. A. 2011. Effects of fire on Spotted Owl site occupancy in a late-successional forest. Biol. Conserv. 144:610-619. Seaman, D. E. and Powell, R. A. 1996. An evaluation of the accuracy of kernel density estimators for home range analysis. Ecology 77:2075-2085. Seamans, M. E. 2005. Population biology of the California Spotted Owl in the central Sierra Nevada. Ph.D. dissertation, Univ. of Minn., St. Paul. Silverman, B. W. 1986. Density Estimation for Statistics and Data Analysis. Chap- man and Hall, London. Skinner, C. N., and Chang, C. 1996. Fire regimes, past and present. Sierra Nevada Ecosystem Project: Final report to Congress, vol II, pp. 1041-1069. Centers for Water and Wildland Resources, Univ. of Calif., Davis. Smith, J. K. 2000. Wildland fire in ecosystems: Effects of fire on fauna. U.S. Forest Serv. Gen. Tech. Rep. RMRS-GTR-42, vol. 1. Smith, R. B., Peery, M. Z., Gutierrez, R. J., and LaHaye, W. S. 1999. The relation- ship between Spotted Owl diet and reproductive success in the San Bernardino Mountains, California. Wilson Bull. 111:22-29. Stephenson, J. R., and Calcarone, G. M. 1999. Southern California mountains and 125 CALIFORNIA SPOTTED OWLS IN A BURNED FOREST foothills assessment: Habitat and species conservation issues. U.S. Forest Serv. Gen. Tech. Rep. GTR-PSW-172. Swanson, M. E., Franklin, J. F., Beschta, R. L., Crisafulli, C. M., DellaSala, D. A., Hutto, R. L., Lindenmayer, D. B., and Swanson, F. J. 2011. The forgotten stage of forest succession: Early-successional ecosystems on forest sites. Frontiers Ecol. Env. 9:117-125. Thrailkill, J. and Bias, M. A. 1989. Diets of breeding and nonbreeding California Spotted Owls. J. Raptor Res. 23:39-41. U.S. Forest Service. 2001. Sierra Nevada Forest Plan Amendment, Final Environ- mental Impact Statement. January 2001. U.S. Forest Service. 2004. Sierra Nevada Forest Plan Amendment, Final Supple- mental Environmental Impact Statement. January 2004. U.S. Forest Service. 2005. Final Environmental Impact Statement for Revised Land Management Plans: Angeles National Forest, Cleveland National Forest, Los Padres National Forest, San Bernardino National Forest. U.S. Forest Service, Pacific Southwest Region. September 2005. Verner, J., Gutierrez, R. J., and Gould, G. I., Jr. 1992. The California Spotted Owl: A technical assessment of its current status (J. Verner, K. S. McKelvey, B. R. Noon, R. J. Gutierrez, G. I. Gould, and T. W. Beck, tech, coords.), pp. 55-77. U. S. Forest Serv. Gen. Tech. Rep. PSW-GTR-133. Wand, M. P., and Joes, M. C. 1995. Kernel smoothing. Chapman & Hall/CRC, Boca Raton, FL. Ward, J. P., Jr., Gutierrez, R. J., and Noon, B. R. 1998. Habitat selection by North- ern Spotted Owls: The consequences of prey selection and distribution. Condor 100:79-92. Waters, J. R., and Zabel, C. J. 1995. Northern flying squirrel densities in fir forests of northeastern California. J. Wildlife Mgmt. 59:858-866. Weatherspoon, C. P, Husari, S. J., and van Wagtendonk, J. W. 1992. Fire and fuels management in relation to owl habitat in forests of the Sierra Nevada and southern California, in The California Spotted Owl: A technical assessment of its current status (J. Verner, K. S. McKelvey, B. R. Noon, R. J. Gutierrez, G. I. Gould, and T. W. Beck, tech, coords.), pp. 207-221. U. S. Forest Service General Technical Report PSW-GTR-133. Williams, D. F., Verner, J., Sakai, H. F., and Waters, J. R. 1992. General biology of major prey species of the California Spotted Owl, in The California Spotted Owl: A technical assessment of its current status (J. Verner, K. S. McKelvey, B. R. Noon, R. J. Gutierrez, G. I. Gould, and T.W. Beck, tech, coords.), pp. 207-221. U.S. Forest Serv. Gen. Tech. Rep. PSW-133. Worton, B. J. 1995. Using Monte Carlo simulation to evaluate kernel-based home range estimators. J. Wildlife Mgmt. 59:794-800. Zabel, C. J., Steger, G. N., McKelvey, K. S., Eberlein, G. P., Noon, B. R., and Verner, J. 1992. Home-range size and habitat-use patterns of California Spotted Owls in the Sierra Nevada in The California Spotted Owl: A technical assessment of its current status (J. Verner, K. S. McKelvey, B. R. Noon, R. J. Gutierrez, G. I. Gould, and T. W. Beck, tech, coords.), pp. 149-163. U. S. Forest Serv. Gen. Tech. Rep. PSW-GTR-133. Zabel, C. J., McKelvey, K. S., and Ward, J. P., Jr. 1995. Influence of primary prey on home-range size and habitat-use patterns of Northern Spotted Owls ( Strix occldentalis caurina). Can. J. Zool. 73:433-439. Zimmerman, G. S., LaHaye, W. S., and Gutierrez, R. J. 2001. Breeding-season home ranges of Spotted Owls in the San Bernardino Mountains, California. W. Birds 32:83-87. Accepted 26 November 2012 126 HUMAN FOOD SUBSIDIES AND COMMON RAVEN OCCURRENCE IN YOSEMITE NATIONAL PARK, CALIFORNIA CARA E. BROOK, Stanford University, Department of Biology, 371 Serra Mall, Stanford, California 94305-5020; cbrook04@gmail.com DAVID P. BERNSTEIN, SLAC National Accelerator Laboratory, 2575 Sand Hill Road, Menlo Park, California, 94025-7015 ELIZABETH A. HADLY, Stanford University, Department of Biology, 371 Serra Mall, Stanford, California 94305-5020 ABSTRACT : We examined the influence of the availability of human food on the distribution and foraging habits of the Common Raven ( Corvus corax), in Yosemite National Park, California. The raven arrived and established itself as a year-round resident in Yosemite in the 1960s, and its population has been increasing ever since. Surveys of nine sites with varying levels of human influence in Yosemite suggest that the Common Raven is most densely distributed in human-influenced regions, especially Yosemite Valley. It is largely absent from Yosemite’s Badger Pass Ski Area in the off season but uses the site as anthropogenic food becomes more available during the ski season. At Badger Pass and four recreational destinations ravens track the availability of human food temporally, preferentially foraging before or after human mealtimes. The Common Raven ( Corvus corax) ranges across North America, Eur- asia, and northern Africa, occupying natural landscapes as varied as deserts, forests, and grasslands, as well as human-dominated cityscapes (Boarman and Heinrich 1999). Noted for their opportunistic nature and adaptability to diverse habitats, ravens thrive in developed areas and respond favorably to human transformation of natural landscapes (Boarman and Heinrich 1999). Between 1966 and 2003, according to the Breeding Bird Survey, the raven population increased on the order of 1.5% per year and expanded its range into urbanized regions across the United States and Canada (Figure la; www.mbr-pwrc.usgs. gov/bbs/bbs.html). In the American West, over the past half-century, the raven’s expansion has been documented widely in conjunction with increasing human modification of the environment (Marzluff and Angell 2005). On the Olympic Peninsula, ravens cluster more densely and reproduce at higher rates near human settlements (Marzluff and Neather- lin 2006). In the Mojave Desert, the rate of juvenile ravens’ survival increases with their nest’s proximity to anthropogenic resources (Webb et al. 2004) such as landfills, trash bins, and runoff from irrigation (Boarman 1993), and Knight et al. (1993) reported corresponding population increases over the past 30 years — at rates greater than 1500% in some areas. Joseph Grinnell’s pioneering studies found that ravens were absent from Yosemite National Park in the early 20th century (Grinnell and Storer 1924). Though Grinnell noted that the raven was “common locally” in the Sierra Nevada, he acknowledged that, “conditions in the Yosemite region do not seem to be attractive to the Western Raven, as we saw only a single individual [in 6 years of field study] and that bird was doubtless a wanderer from farther south” (Grinnell and Storer 1924). Furthermore, Wallis (1950) highlighted a ranger’s observations of a “rare” Common Raven “which previ- ously had been unrecorded from Yosemite National Park.” Christmas Bird Western Birds 44:127-134, 2013 127 RAVENS SUBSIDIZED IN YOSEMITE (a) Percent Change, Ravens (1966-2003) L 1 Less than -1,5 □ -1.5 to -0.25 ■ >-0.25 to 0.25 ■ >0.25 to +1.5 ■ Greater than +1.5 (b) Humans and Ravens in Yosemite (1905-2009) O o o cf o c o c 03 E Z3 X "ra =3 C C < II 0 - 2.5 NPS Organic Act Mission 66 Human-Bear Management Plan 1 2005 Yosemite Valley Plan Resources Management Plan Concession Services Management Plan - 1.5 - 1 - 0.5 - 0 • II ZJ 0) < CD =3 (/) ”5 x j-? > c Q. c cr o O 00 O Figure 1 . Increases in the population of the Common Raven (a) from the Breeding Bird Survey, U.S. and Canada, 1966-2003, and (b) in Yosemite National Park from Christmas Bird Count data (solid circles), compared with rates of human visitation from National Park Service data (open circles) 1905-2009. Count records (http: / / netapp . audubon . org/ CBCObservation/Historical/ ResultsBySpecies.aspx?l) indicate that the raven established itself as a year- round resident in Yosemite in the late 1960s and has been increasing in population since the early 1980s (Figure lb). Coincidence of the timing of the raven’s population explosion and height- ened human visitation to Yosemite National Park (Figure lb) suggest that in Yosemite ravens may be following a pattern of expansion subsidized by 128 RAVENS SUBSIDIZED IN YOSEMITE human resources consistent with that elsewhere in the American West (Mar- zluff and Neatherlin 2006, Knight et al. 1993). We investigated the raven’s spatial and temporal reliance on human food in Yosemite. Specifically, we tested the hypotheses that ravens (1) are more abundant in human-influenced regions of Yosemite, especially Yosemite Valley, and (2) track human food subsidies temporally at a given site. METHODS To examine the spatial extent of human influence on the raven, we sur- veyed nine sites of varying human influence in Yosemite National Park from 27 June to 18 August 2009 (Figure 2a). At each site, we walked surrounding trails and noted ravens by sight or call between dawn (-05:30) and 09:30. Because of the large area investigated and difficulty of access to some sites, we intended these initial surveys as preliminary data for our subsequent more detailed study. We surveyed for a total of 1812 minutes across all sites. We also investigated the spatial and temporal interactions between avail- ability of human food and presence of foraging ravens from 28 November 2009 to 4 April 2010 at Badger Pass Ski Area in southern Yosemite National Park. Badger Pass is popular with skiers in winter but is largely abandoned during summer and fall. We surveyed for ravens at Badger Pass once in November, during the off season, as well as on days immediately preceding and following the ski area’s opening date of 18 December 2009. Because of difficulties of access to our field site, we visited Badger Pass only inter- mittently through the ski season and after the area’s closing in April. Each day of a survey, we arrived before dawn, commenced observations at first light (-06:30), and ended them after dark (-19:00). From the top deck of Badger Pass Day Lodge, we recorded the number of ravens foraging on the ground or deck of the lodge continuously through the day, so as to avoid over-representing the same raven in discrete counts. We also counted the number of people occupying the lodge’s picnic tables every 30 minutes. This count served as a proxy for availability of human food at Badger Pass, though we counted all people seated at the picnic tables regardless of whether or not they were eating. To further explore the relationship between the timing availability of hu- man food and the raven’s foraging habits, from 21 July to 22 August 2009 we visited four sites ravens frequented in Yosemite Valley: Upper Pines Campground, Swinging Bridge Picnic Area, Nevada Falls Overlook, and Valley View Turnout. We surveyed each site six times, with each survey tak- ing place on a randomly selected different day. We surveyed each site three times between dawn (-05:30) and 14:00 and three times between of 14:00 and dark (-20:30), so that all daylight hours were represented equally. For these surveys, we divided each recreation area into plots of 10 x 10 m and used a random-number generator to determine the plot in which to establish our observation point. One observer sat in a chair at the observation point, counted the ravens, and recorded their activity patterns, as well as counting people eating within a 100-m radius. We distinguished ravens at rest in the trees from those that were “foraging,” that is, on the ground or on a picnic table, fire pit, or cooler. As at Badger Pass, we noted ravens’ presence in the 129 RAVENS SUBSIDIZED IN YOSEMITE 2 . Upper Pines Campground 50 H 0 100 3 . Swinging 1 Bridge £ Picnic Area 2 3 X 3* Nevada Falls Overtook 5 . Valley View Turnout o 4 =8: 73 03 < CD 3 0) eg d ' (Q Figure 2. (a) Sites of surveys of raven distribution (gray circles) and raven foraging with respect to human food consumption (black numbers), (b) Numbers of people with food (gray) and numbers of foraging ravens (black) at sites depicted in (a) by time of day. All lines are averages from multiple visits to each site. Error bars show the standard error. Standard errors for human use at Badger Pass are larger because these data were recorded in half-hour counts instead of continuously through the day. Note also that the scales of the axes for the Valley View Turnout are different from those in the other graphs because of much lower numbers of ravens and people. The number of observer-hours for all sites combined was 288. site continuously through the day to avoid over-representing the same raven in discrete counts. We returned to the same observation point on subsequent visits. Any ravens flying over the study site were not included in our counts. For our spatial data, we used a 2-way analysis of variance (ANOVA) to test the independence of raven sightings on observational minutes (treated in the model as a continuous variable) and location (treated as a discrete variable). We divided the spatial data into two groups, sightings in the valley and sightings outside of the valley, to make a coarse delineation of variation in human influence on the scale of Yosemite as a whole, as the valley is more extensively developed than other regions of the park. We used linear regres- 130 RAVENS SUBSIDIZED IN YOSEMITE sion analysis to examine the relationship between sales of lift tickets at the Badger Pass Ski Area (a proxy for human visitation and availability of human food) and presence of foraging ravens the following day. We compared raven foraging with human visitation to Badger Pass the preceding day because the ravens scavenged primarily in the early morning on remnants left from the day before (ravens are not active after dark). For our time-series data, for both Badger Pass and Yosemite Valley, we used cross-correlation tests to quantify the lag time between availability of human food and the ravens’ foraging response. To create a time-series loop of 24 hours, we assumed people with food and foraging ravens were absent after nightfall at 20:30. Though people may have left food after dark, the raven is an exclusively diurnal animal. For cross-correlation tests we used MATLAB. We set statisti- cal significance for all tests at P < 0.05. The computer code with which we ran the ANOVA and cross-correlation analysis is available from the authors. RESULTS In summer, in 1812 observer-minutes of surveys across Yosemite National Park, we recorded 125 sightings of the raven. The number of sightings per observer-minute was significantly higher in Yosemite Valley than at all other visited sites (2-way ANOVA; F = 28.22, df = 1, P = 1.5 x 10 -5 ). During our winter surveys at Badger Pass, we observed few ravens dur- ing the November off season and the days preceding the ski area’s opening on 18 December 2009. Then raven foraging at the site increased after the opening and remained high through the ski season (108 days total). After Badger Pass closed at the end of the ski season on April 4, however, raven sightings decreased substantially (Figure 3). Thus, in a regression analysis, foraging ravens were correlated positively with the proxy for availability of human food, sales of lift tickets the preceding day (r = 0.780, F = 13.98, df= 1,P = 0.004). At the Badger Pass and Yosemite Valley sites surveyed daily, human food consumption adhered to predictable mealtime patterns. At Badger Pass, Swinging Bridge, and Nevada Falls, mean human food consumption peaked once a day, between 12:30 and 14:30. At Upper Pines Campground, mean human food consumption peaked twice in a given day, at 10:30 and 19:30. The pattern was more complex at Valley View Turnout, where visitors stopped for shorter periods at less defined mealtimes, though the availability of human food increased in the late afternoon and peaked at 17:30. Raven foraging was greatest at times of day that were slightly offset from peak human food consumption (Figure 2b): cross-correlation analysis revealed a lag time between the ravens’ foraging behavior and the peak of human food consumption. At those sites where human food consumption predictably peaked only once, at midday — Badger Pass, Swinging Bridge, and Nevada Falls — our cross-correlation function indicated lag times of 5, 6, and 5 hours, respectively. At Upper Pines Campground, where human food consumption peaked twice a day, the lag time was 3 hours. At our most complex site, Valley View, where human food was sporadic, the lag time was 2 hours. 131 RAVENS SUBSIDIZED IN YOSEMITE Figure 3. Rate of occurrence of the Common Raven at Badger Pass over the 2009-2010 ski season. Bars show the ski area’s dates of opening (18 December) and closing (4 April) The total number of observer-hours was 117.5. DISCUSSION As predicted, we found an association between raven and human presence in Yosemite National Park. Ravens occurred disproportionately in Yosemite Valley, the most developed region of the park, and ravens concentrated at Badger Pass Ski Area during the winter when people — and human food resources — were available. We interpret the patterns of raven foraging we observed as direct responses to human food left over from previous periods of peak consumption. At Badger Pass, ravens arrived abruptly on 19 De- cember, one day after the ski area opened and the first morning that leftover human food became available. As at Badger Pass, in our time-series studies in summer 2009, ravens frequently foraged in the early morning before people were eating; we infer that ravens were scavenging scraps remaining from the previous day and/ or night. At those sites with the longest lag time between the peaks of hu- man food consumption and raven foraging — Badger Pass, Swinging Bridge, and Nevada Falls — we speculate that ravens traveled to these sites to forage after human mealtimes, then departed for other locations. At Upper Pines Campground, with a 3-hour lag time, the ravens may have been forced to forage more quickly, depart, and then return following the second human mealtime. And at Valley View, a 2-hour lag time suggests a more immedi- ate and opportunistic foraging behavior. At Upper Pines Campground and Valley View Turnout, ravens resident to these sites may be able to forage more frequently and at shorter lag times. Our study is observational, not experimental, so we cannot prove a dif- fering pattern of availability of human food to be the cause of the observed lag times in raven foraging. However, while a variety of factors independent of human food availability might motivate peaks in the raven’s foraging ef- fort, one would expect such peaks to take place at similar times in similar 132 RAVENS SUBSIDIZED IN YOSEMITE habitats, all else being equal. The raven is a generalist that demonstrates considerable ingenuity in exploiting diverse habitats worldwide (Boarman and Heinrich 1999), so we assume the minor differences in habitat between our five observational sites — all located within or on the outskirts of Yosemite Valley — to be negligible from the perspective of the raven’s habitat prefer- ence. Therefore, we attribute differences in the pattern of raven foraging to differences in patterns of availability of human food. Further study of the diurnal movements and foraging patterns of individual ravens in Yosemite National Park — through radio-telemetry tracking — is needed to disentangle the human factors driving the raven’s activity patterns. The raven first arrived in Yosemite National Park in the 1960s, which, intriguingly, coincided with an increase in the park’s management of visitors’ effect on the surrounding ecosystem. The open-air landfills in Yosemite were closed in 1969 and 1970, and in 1975, all trash bins were bear-proofed (Figure lb; Greene 1987). While the park’s policy reduced human conflicts with bears (Harms 1980), raven populations continued to climb through the 1980s. Notably, the number of ravens recorded on the Christmas bird count in 1980 increased by over 200% during the same year in which overnight human visitation to Yosemite campgrounds increased 39% (Wendt 1981). Conversely, in 1998, those numbers declined 69% from the previous year (Wendt 1981), following the elimination of some 353 campsites by the severe flood of 1997. While the association between the raven’s expansion and history of wildlife management in Yosemite National Park will likely never be anything more than correlative, the link between contemporary ravens and people in Yo- semite is undeniable. Our study does not address the effects of an expanding raven population on the rest of the Yosemite ecosystem, but as ravens are known nest predators (Marzluff and Restani 1999), raptor mobbers (Dawson 1981), and predators (Kristan and Boarman 2003), we suspect that their effect on the dynamics of the surrounding ecosystem is significant. Previous studies have demonstrated increased rates of nest predation by corvids in more human-modified landscapes (Andren 1992), but patterns of predation by the raven specifically remain unclear (Marzluff et al. 2007). Further studies of the raven’s effects on declining songbird populations in Yosemite Valley (S. Stock pers. comm.), including experiments investigating predation on model nests and observations via motion-detecting cameras, are needed to elucidate the responsible factors. Although America’s national parks are commonly assumed to be pristine, we found that even in regions of minimal habitat alteration, large numbers of people may be affecting the dynamics of surrounding ecosystems in dramatic ways. While more obvious disturbances, such as the introduction of exotic species and land conversion, have received substantial attention in the ecological literature, the far-reaching effects of more subtle forms of human land use remain unknown. While the observable effects of availability of human food on large vertebrates like bears are widely recognized, our studies suggest that even micro-scale waste can have tangible effects on a natural habitat. As the first ecological investigation of the raven’s expansion into Yosemite to date, our study sheds light on the extent of a subtler human influence on a seemingly wild ecosystem. 133 RAVENS SUBSIDIZED IN YOSEMITE ACKNOWLEDGMENTS We thank Lily C. Li for aid in graphics and design. We acknowledge the Freeman Spogli Institute for International Studies and Stanford VPUE for funding and institu- tional support. We thank Donald Kennedy, Terry Root, members of the Hadly lab at Stanford, and colleagues in the National Park Service (Miriam Luchans, Linda Eade, Sarah Stock and Les Chow) for helpful comments on the study. Finally, we thank Jameson Chace and John Withey for valuable reviews of the manuscript. LITERATURE CITED Andren, H. 1992. Corvid density and nest predation in relation to forest fragmenta- tion: A landscape perspective. Ecology 73:794-804. Boarman, W. I. 1993. When a native predator becomes a pest: A case study, in Con- servation and Resource Management (S. K. Majumader, E. W. Miller, D. E. Baker, J. R. Pratt, and R. F. Schmalz, eds.), pp. 191-206. Pa. Acad. Sci., Easton, PA. Boarman, W. I., and Heinrich, B. 1999. Common Raven ( Corvus corax), in The Birds of North America (A. Poole and F. Gill, eds.), no. 476. Birds N. Am., Inc., Philadelphia. Dawson, J. W. 1981. Mating behavior in the Golden Eagle in non-fertilization con- texts. Raptor Research 16:136. Greene, L. W. 1987. Yosemite: The Park and Its Resources — A History of the Dis- covery, Management, and Physical Development of Yosemite National Park, CA. U.S. Dept. Interior Historic Resource Study. Government Printing Office, Denver. Grinnell, J., and Storer, T. I. 1924. Animal Life in the Yosemite. Univ. of Calif. Press, Berkeley. Harms, D. R. 1980. Black bear management in Yosemite National Park. Int. Conf. Bear Res. Mgmt. 4:205-212. Knight, R. L., Knight, H. A. L, and Camp, R. J. 1993. Raven populations and land- use patterns in the Mojave Desert, California. Wildlife Soc. Bull. 21:469-471. Kristan, W. B., and Boarman, W. I. 2003. Spatial pattern of risk of Common Raven predation on desert tortoises. Ecology 84:2432-2443. Marzluff, J. M., and Angell, T. 2005. In the Company of Crows and Ravens. Yale Univ. Press, New Haven, CT, Marzluff, J. M., and Neatherlin, E. 2006. Corvid response to human settlements and campgrounds: Causes, consequences, and challenges for conservation. Biol. Conserv. 130:301-314. Marzluff, J. M., and Restani, M. 1999. The effects of forest fragmentation on avian nest predation, in Forest Fragmentation: Wildlife Management and Implications (J.A. Rochelle, L.A. Lehmann, and J. Wisniewski, eds.), pp. 155-170. Konin- klijke Brill, Leiden, the Netherlands. Marzluff, J. M., Withey, J. C., Whittaker, K. A., Oleyar, M. D., Unfried, T. M., Rul- lman, S., and DeLap, J. 2007. Consequences of habitat utilization by nest preda- tors and breeding songbirds across multiple scales in an urbanizing landscape. Condor 109:516-534. Wallis, O. L. 1950. A new bird record — raven observed near Mariposa Grove. Yo- semite Nature Notes 29 (12): 117-1 18. Webb, W. C., Boarman, W. I., and Rotenberry, J. T. 2004. Common Raven juvenile survival in a human-augmented landscape. Condor 106:517-528. Wendt, C. W. 1981. Overnight use trends in Yosemite National Park. U.S. Govern- ment Memorandum. National Park Service, Yosemite National Park, CA. Accepted 4 November 2012 134 GENETIC EVIDENCE FOR MIXED MATERNITY AT A LARK SPARROW NEST KEVIN ELLISON, Department of Biological Sciences, University of Manitoba, Win- nipeg, Manitoba R3T 2N2 (current address: Wildlife Conservation Society, 301 North Willson Avenue, Bozeman, Montana 59715); kellison@wcs.org JEREMY D. ROSS, Department of Biological Sciences, Bowling Green State Uni- versity, Bowling Green, Ohio 43403; rossjd@bgsu.edu JUAN L. BOUZAT, Department of Biological Sciences, Bowling Green State Uni- versity, Bowling Green, Ohio 43403; jbouzat@bgsu.edu ABSTRACT: Among passerines not nesting in cavities, mixed maternity within a clutch is rare and has been recorded at <1% of the nests of four North American species. Using genetic methods, we report the first case of mixed maternity in the Lark Sparrow (Chondestes grammacus). Within an unusually large clutch in Texas, we found that the attending female shared the same mitochondrial haplotype with an unhatched egg and one nestling, but not with another nestling. Analysis with microsatellite DNA confirmed mixed maternity at the nest; the female and a nestling did not share alleles at five of nine loci analyzed. Various behaviors may lead to mixed maternity, including intraspecific reuse of nests, conspecific usurpation, and conspe- cific parasitism. Our study stresses the important roles that intensive nest monitoring and genetic tests can play in detecting cryptic reproductive strategies. Mixed maternity at nests is rarely reported among passerines (Yom-Tov 2001, Lyon and Eadie 2008). However, mixed maternity may be common but underdetected in certain species or populations; such a context could alter our interpretation of evolutionary adaptations such as egg recogni- tion, nest defense, and parenting behavior (Yom-Tov 2001). For instance, Latif et al. (2006) only recently recorded the first evidence of conspecific nest parasitism and egg ejection in the Song Sparrow ( Melospiza melodia) despite widespread study of that species’ nesting over the past century (see Smith et al. 2006). In addition to nest parasitism, multiple maternity at a nest can also arise via several mechanisms (see review in Lyon and Eadie 2008): nest reuse (i.e. , when a female lays among eggs left in an abandoned nest), errors in nest ownership — possibly paired with the need to lay (Harms et al. 1991), mate change during laying (Griffith et al. 2004), or usurpation (i.e., the forcible takeover of an active nest by a female other than the owner) (Sullivan et al. 1989; reviewed by Ellison 2008). The Lark Sparrow ( Chondestes grammacus) appears unusual in its propensity to reuse inactive nests constructed by other individuals (McNair 1984, 1985). Lark Sparrows also attempt to use active nests of other birds, sometimes resulting in successful usurpation of nests of the same or other species (McNair 1984, Ellison unpubl. data). Detecting such mixed maternity at a nest requires intensive monitoring, often encompassing the use of egg marking, video surveillance, or genetic testing (Latif et al. 2006). Clutches that exceed twice the species’ mean clutch size can alert researchers to the possibility of conspecific nest parasitism (Yom-Tov 1980). The mean number of eggs Martin and Parrish (2000) reported for the Lark Sparrow is 3.84 ± 0.70 (n = 209 nests). Among 200 nests of the Lark Sparrow (Ellison et al. Western Birds 44:135-140, 2013 135 GENETIC EVIDENCE FOR MIXED MATERNITY AT A LARK SPARROW NEST 2006), we encountered one with seven eggs and therefore suspected that multiple females had used it. To determine if there was evidence for mixed maternity in this nest, we examined mitochondrial DNA sequences and nuclear genotypic data of the adult female, eggs, and nestlings. We hypoth- esized usurpation, reuse, or intraspecific parasitism as possible explanations for any mixed maternity, and we considered the likelihood of each in light of the physical and genetic evidence. METHODS The Lark Sparrow nest was found on 28 April 2002 in Kinney Co. , Texas. It was on the ground, under an overhanging pad of prickly pear {Opuntia spp), and contained four eggs when discovered. Two days later, the nest held five Lark Sparrow eggs and we found another two eggs 25 and 38 cm outside the nest. We collected one egg outside the nest; the other was damaged and discarded in the field. On 11 May, we observed five hatch- lings were in the nest, and on 13 May we captured an adult female at the nest, banded her, and took a blood sample. This banded female was then observed carrying food to the nest and fledglings through 22 May. On 14 May we banded and sampled the blood of two nestlings (A and B); the third remaining nestling (C) was too small to be banded or bled. On 18 May A and B (the only young we detected) had fledged; we recaptured B and sampled its blood again. We therefore had samples from a nesting female, one egg, and two young from this nest. Blood samples were stored in Queen’s lysis buffer (Seutin et al. 1991) at 4°C. The egg was stored in a sealed bag at -20 °C. To prepare the egg for DNA extraction, we removed the shell while the egg was still frozen and sealed it in a 1.5-mL microcentrifuge tube. We thawed each egg’s contents separately and checked for any evidence of a developing embryo. DNA from blood and eggshell samples was extracted by a proteinase-K digestion followed by a standard phenol-chloroform extraction protocol (Sambrook et al. 1989, Strausberger and Ashley 2001). To check against possible con- tamination we concurrently extracted DNA from an unrelated Lark Sparrow from Nebraska, which we used as a positive control, and performed a blank extraction as a negative control. Using primer set L16743 (Tarr 1995) and H920 (5'-GTC CGG CAA CCA TTA CAC TA-3'; Ross 2011), we amplified 910 base pairs of the mitochon- drial control region (d-loop) from each sample. We purified the products of the polymerase chain reaction by ethanol precipitation and sequenced them with ABI BigDye Terminator version 3.1 (Applied Biosystems, Foster City, CA). For sequencing we used the H920 and two additional nested prim- ers, H598 (5'-TTC AAC CGA CCA CTT GTA TCT G-3' ; Ross 2011) and L437 (Tarr 1995). We sequenced the gene on an ABI377 Genetic Analyzer (Applied Biosystems) and checked the accuracy of the scoring visually, then exported the sequences to BioEdit (version 7.09, Hall 1999) to check those from the same sample for mismatched base pairs. We genotyped nine nuclear microsatellites from each sample: Dppl6 (Dawson et al. 1997), Map23 (Alderson et al. 1999), ESCpl (Hanotte et al., 1994), Asp09 and Aspl5 (Delany et al. 2000), Cup02 (Gibbs et al. 136 GENETIC EVIDENCE FOR MIXED MATERNITY AT A LARK SPARROW NEST 1999), Pdop 3 (Griffith et al. 1999), G/Olb (Rasner et al. 2004), and G/05 (Petren 1998). We had tested these microsatellites previously, finding them to be polymorphic in the Lark Sparrow. The polymerase chain reactions for amplification of the microsatellites followed conditions described by Ross (2011). We genotyped the microsatellites with an ABI377 Genetic Analyzer and scored them with GeneScan software (Applied Biosystems). RESULTS Sequences of the mitochondrial d-loop of the adult female banded at the nest, the eggshell, and nestling B were identical (Table 1). That of nestling A was different, with 9 base-pair substitutions in the DNA fragment of 910 base pairs. As expected, the mitochondrial d-loop of the control sample yielded a completely different haplotype (Table 1). The female shared at least one allele at each microsatellite locus with nestling B, which is consistent with a mother-offspring relationship (Table 1). Also, the female and nestling A did not share any alleles at five micro- satellite loci (Table 1). Although we were not able to genotype all the loci for the eggshell, genotypes at four loci (Map23, Dpp 16, Asp09, and G/05) were identical to those of the attending female. Since the embryo in the egg was undeveloped and we extracted the DNA from tissue remaining in the eggshell, we can assume that the genotypes of the eggshell correspond to those of the mother. Neither the blank extraction nor the negative controls of the polymerase chain reaction yielded any product. Thus we are confident that the products of DNA amplification from the eggshell were not from contamination. Table 1 Mitochondrial DNA Haplotypes and Microsatellite Genotypes of the Attending Female Lark Sparrow, Two Young Sampled in the Nest, and the Shell of an Egg Female Nestling A Nestling B Eggshell Nebraska sample mtDNA haplotype a HI H2 HI HI H3 Microsatellite qenotypes b Dpp 16 156/158 158/160 156/158 156/158 158/160 Map 23 143/145 141/143 143/145 143/145 143/143 ESCpl 169/179 129/155 151/179 — 127/127 Asp09 133/133 133/133 129/133 133/133 129/133 Aspl5 121/143 131/145 143/143 — 119/133 Cup02 112/112 112/112 112/112 — 110/112 Pdop3 91/91 83/147 91/91 — 119/119 G/Olb 216/224 220/222 216/216 — 216/220 G/05 185/185 187/187 185/185 185/185 185/189 “Haplotypes HI, H2, and H3 correspond to Genbank sequences FJ348339, FJ348340, and FJ348341. b Alleles of microsatellite genotypes are reported in number of base pairs; those matching the attending female are in bold. 137 GENETIC EVIDENCE FOR MIXED MATERNITY AT A LARK SPARROW NEST DISCUSSION As far as is known, mixed maternity in nests of North American passerines is rare. For cup-nesting species, conspecific parasitism or usurpation has been previously noted in 10, the Cave Swallow ( Petrochelidon fulua), Cliff Swal- low ( P pyrrhonota), Hooded Warbler (Setophaga citrina), Song Sparrow, Field Sparrow ( Spizella pusilla), Yellow-eyed Junco ( Junco phaeonotus), Red-winged Blackbird ( Agelaius phoeniceus), Yellow-headed Blackbird (. Xanthocephalus xanthocephalus ), Brewer’s Blackbird ( Euphagus cyano- cephalus), and Bullock’s Oriole ( Icterus bullockii), though at relatively few nests (Sullivan et al. 1989, Petter et al. 1990, Harms et al. 1991, Arnold and Owens 2002, Latif et al. 2006). Our genetic analyses revealed that an attending female Lark Sparrow was not related to at least one nestling (nestling A) at a nest where she was the biological mother of nestling B and an egg found outside the nest (Table 1). The mixed maternity of the clutch may reflect intraspecific reuse of the nest, as the nest could have been abandoned with one or more eggs in it when a second female took over the nest. Indeed, birds are more apt to abandon nests during laying when the parental investment in the nest is low and the benefit of choosing a more desirable nest site may outweigh the costs of abandonment (Hosoi and Rothstein 2000). However, this conclusion implies that the female that reused the nest had ejected at least one of her own eggs, since the genetic analyses revealed that the egg found outside the nest was laid by the nesting female. Mixed maternity of the clutch may also have arisen through usurpation; that is, the banded female may have usurped the nest, with one or more unrelated eggs within it (Lindell 1996). Attempted usurpation followed by the banded female regaining her nest and raising unrelated young alongside her own is equally likely. Under this scenario, a would-be usurper ejected part of the host’s clutch, laying at least one egg of her own (nestling A). We cannot say how many of the three new eggs in the nest were laid by either female. Neverthe- less, the banded female raised at least one related and one unrelated young. Mixed maternity at the nest could have also occurred as a result of con- specific parasitism. We are unable to distinguish conspecific usurpation and nest parasitism as each can result in the same outcome. We do note that, in several instances, Lark Sparrows have raised unrelated young, including those of other species. This occurred despite some nests containing eggs of the previous owner (McNair 1984, Ellison unpubl. data). Further study is needed to determine if Lark Sparrows can eject eggs. Peer et al. (2000) reported that they rejected 2 of 3 undersized non-mimetic eggs but appear unable to discriminate undersized mimetic model eggs, as they accepted 2. Ross (2011) closely monitored approximately 75 color-banded Lark Sparrows in Ohio. In each of three years, both females and males defended territories from conspecific invaders of both sexes. None of 14 nests known to be built by marked females was permanently usurped. On three occasions (in 51 territories studied in 2006 and 2007), however, when a fledgling barely capable of flight was flushed and produced an alarm call, it was defended by three adults, two being the juvenile’s social parents and the third an unbanded adult of unknown sex. These observations might signal 138 GENETIC EVIDENCE FOR MIXED MATERNITY AT A LARK SPARROW NEST a parasitic/usurping female offering parental support to what could be her offspring. As Lark Sparrows apparently saturated the suitable habitat at this site, females may have adopted strategies that allowed them to breed where no territories were available. In summary, we report here the first case of mixed maternity in a wild population of the Lark Sparrow. We cannot determine how the eggs of two females came to be in the same nest. Regardless, our observation reflects a behavior of interest and should serve as an added impetus for the consid- eration of alternative reproductive tactics that may be revealed by intensive monitoring of nests. ACKNOWLEDGMENTS We thank the land owners of Fort Clark Springs for the permission to study the Lark Sparrow on their property and the Ohio Department of Natural Resources for financial support. We also thank Mike D. Boyd for his assistance in the field. LITERATURE CITED Alderson, G., Gibbs, H. L., and Sealy, S. G. 1999. Parentage and kinship studies in an obligate brood parasitic bird, the Brown-headed Cowbird ( Molothrus ater), using microsatellite DNA markers. J. Heredity 90:182-190. Arnold, K. E., and Owens, I. P. F. 2002. Extra-pair paternity and egg dumping in birds: Life history, parental care and the risk of retaliation. Proc. Royal Soc. London B 269:1263-1269. Dawson, R. J. G., Gibbs, H. L., Hobson, K. A., and Yezerinac, S. M. 1997. Isola- tion of microsatellite DNA markers from a passerine bird, Dendroica petechia (the Yellow Warbler), and their use in population studies. Heredity 79:506-514. Delany, M. F., Geisel, J. T., and Brazeau, D. A. 2000. Genetic variability among populations of the Florida Grasshopper Sparrow. J. Wildlife Mgmt. 64:631-636. Ellison, K. 2008. Nest reuse by Vermilion Flycatchers in Texas. Wilson J. Ornithol. 120:239-244. Ellison, K., Sealy, S. G., and Gibbs, H. L. 2006. Genetic elucidation of host use by individual sympatric Bronzed Cowbirds (Molothrus aeneus) and Brown-headed Cowbirds (M. ater). Can. J. Zool. 84:1269-1280. Gibbs, H. L., Tabak, L. M., and Hobson, K. 1999. Characterization of microsatel- lite loci for a neotropical migrant songbird, the Swainson’s Thrush (Catharus ustulatus). Molec. Ecol. 8:1151 Griffith, S. C., Stewart, I. R. K., Dawson, D. A., Owens, I. P. F., and Burke, T. 1999. Extra-pair paternity in mainland and island populations of a socially monogamous bird, the House Sparrow ( Passer domesticus): Is there an “island effect”? Biol. J. Linnean Soc. 68:303-316. Griffith, S. C., Lyon, B. E., and Montgomerie, R. 2004. Quasi-parasitism in birds. Behav. Ecol. Sociobiol. 56:191-200. Hall, T. A. 1999. BioEdit: A user-friendly biological sequence alignment editor and analysis program for Windows 95/98/NT. Nucleic Acids. Symp. Ser. 41:95-98. Hanotte, O., Zanon, C., Pugh, A., Greig, C., Dixon, A. and Burke, T. 1994. Isolation and characterization of microsatellite loci in a passerine bird: The Reed Bunting Emberiza schoeniclus. Molec. Ecol. 3:529-530. Harms, K. E., Beletsky, L. D., and Orians, G. H. 1991. Conspecific nest parasitism in three species of New World blackbirds. Condor 93:967-974. Hosoi, S.A., and Rothstein, S. I. 2000. Nest desertion and cowbird parasitism: Evi- dence for evolved responses and evolutionary lag. Anim. Behav. 59:823-840. 139 GENETIC EVIDENCE FOR MIXED MATERNITY AT A LARK SPARROW NEST Latif, Q. S., Grenier, J. L., Heath, S. K., Ballard, G., and Hauber, M. E. 2006. First evidence of conspecific brood parasitism and egg ejection in Song Sparrows, with comments on methods sufficient to document these behaviors. Condor 108:452-458. Lindell, C. 1996. Patterns of nest usurpation: when should species converge on nest niches? Condor 98:464-473. Lyon, B. E., and Eadie, J. M. 2008. Conspecific brood parasitism in birds: A life- history perspective. Annu. Rev. Ecol. Evol. Syst. 39:343-363. Martin, J. W., and Parrish, J. R., 2000. Lark Sparrow ( Chondestes grammacus), in The Birds of North America (A. Poole and F. Gill, eds), no. 488. Birds N. Am., Inc., Philadelphia. McNair, D. B. 1984. Reuse of other species nests by Lark Sparrows. Southwest. Nat. 29:506-509. McNair, D. B. 1985. A Comparison of oology and nest record card data in evaluating the reproductive biology of Lark Sparrows, Chondestes grammacus. Southwest. Nat. 30:213-224. Peer, B. D., Robinson, S. K., and Herkert, J. R. 2000. Egg rejection by cowbird hosts in grasslands. Auk 117:892-901. Petren, K. 1998. Microsatellite primers for Geospiza fortis and cross-species ampli- fication in Darwin’s finches. Molec. Ecol. 7:1782-1784. Petter, S. C., Miles, D. B., and White, M. M. 1990. Genetic evidence of mixed re- productive strategy in a monogamous bird. Condor 92:702-708. Rasner, C. A., Yeh, P., Eggert, L. S., Hunt, K. E., Woodruff, D. S., and Price, T. D. 2004. Genetic and morphological evolution following a founder event in the Dark-eyed Junco, Junco hyemalis thurberi. Molec. Ecol. 13:671-681 Ross, J. D. 2011. The evolutionary history, demographic independence and conservation status of two North American prairie bird species: The Greater Prairie-chicken and the Lark Sparrow. Ph.D. dissertation, Bowling Green State University, Bowling Green, OH. Sambrook, J., Fritsch, E. F., and Maniatis, T. 1989. Molecular cloning: A laboratory manual. Cold Spring Harbor Laboratory Press, Cold Spring Harbor, NY. Seutin, G., White, B. N., and Boag, P. T. 1991. Preservation of avian blood and tissue samples for DNA analyses. Can. J. Zool. 69: 82-90. Smith, J. N. M., Keller, L. F., Marr, A. B., and Arcese, P. 2006. Conservation and biology of small populations: The Song Sparrows of Mandarte Island. Oxford University Press, New York. Strausberger, B. M. , and Ashley, M. V. 2001 . Eggs yield nuclear DNA from egg-laying cowbirds, their embryos and offspring. Cons. Genetics 2:385-390. Sullivan, K. A., Cole, J., and Villalobos, E. M. 1989. Intraspecific nest usurpation by a Yellow-eyed Junco. Wilson Bull. 101:654-655. Tarr, C. L. 1995. Amplification and determination of mitochondrial control-region sequences in oscine passerines. Molec. Ecol. 4:527-529. Yom-Tov, Y. 1980. Intraspecific nest parasitism in birds. Biol. Rev. 55:93-108. Yom-Tov, Y. 2001. An updated list and some comments on the occurrence of intra- specific nest parasitism in birds. Ibis 143:133-143. Accepted 9 March 2013 140 NOTES REFUTATION OF WYOMING NESTING RECORD OF THE PACIFIC WREN JAMES M. MALEY, JACOB R. SAUCIER, and MATTHEW D. CARLING, Museum of Vertebrates, Department of Zoology & Physiology, University of Wyoming, 1000 E. University Avenue, Laramie, Wyoming 82071 The nesting status of the Pacific Wren (Troglodytes pacificus) in Wyoming is not clear. According to several sources (Hellmayr 1934, AOU 1957, Peters 1960, AOU 1998, Dickinson 2003, Clements 2007), Wyoming is not part of the distribution, but this is contradicted by several other sources. On the basis of two immature specimens (USNM 228577 and 228578), collected by Alexander Wetmore in the Tetons on 27 August and 15 September 1910, Phillips (1986) suggested there might be an undescribed subspecies that nests in northwest Wyoming. Saucier examined these specimens and confirmed they are Pacific Wrens, already in adult plumage, although they appear lighter and grayer than specimens from elsewhere. Cary (1917) is the only additional paper we have found mentioning these specimens. We cannot exclude the possibility that they were migrants from elsewhere in the distribution (Toews and Irwin 2012), as information on the schedule of the Pacific Wren’s migration in the Rocky Mountains is still lacking. Wyoming has two winter records of the Pacific Wren but none for fall (D. W. Faulkner pers. comm.). According to Faulkner (2010), there is one record of nesting and five other summer reports of the Pacific Wren for Wyoming. These include two documented reports of singing birds in the Tetons during late June of 1981 and 1985; while nesting seems possible, there was no further evidence. The other three reports are undocumented (Faulkner 2010). The only physical confirmation of nesting in the state is reportedly a single specimen taken from a nest with two fledglings in the Freezeout Hills of central Wyoming in 1897 (Knight 1902, Ridgway 1904, McCreary 1939, Faulkner 2010). Unfortunately, at some point between 1897 and 1910, this specimen was removed from the University of Wyoming Museum of Vertebrates (UWYMV) and could not be located for examination (Faulkner 2010). Specimens housed at UWYMV were not properly curated for decades; many of them were removed and placed in a poorly organized teaching collection. While re- cently inventorying, organizing, and returning specimens from the teaching collection that belong in the vertebrate collection, Maley found the missing specimen (UWYMV 740). Charles W. Gilmore collected it on 15 July 1897 along Tepee Creek, Freezeout Hills, Carbon County. Gilmore was an undergraduate in the Department of Zoology hired to assist Wilbur C. Knight in collecting and preparing specimens (Faulkner 2010). Knight (1902) reported that Robert Ridgway identified the specimen. Notes written on other specimen tags indicate that Ridgway identified several specimens collected during the 1890s and now housed at UWYMV. Examining the specimen closely, we identified it as a juvenile House Wren ( T. aedon). It is a small, short-tailed wren that is slightly more brightly colored than an adult House Wren. The bird has gray scalloping on the breast, a whitish throat and breast, and buffy flanks and undertail coverts. It lacks the rich cinnamon coloration characteristic of adult (UWYMV 751, 752) and immature (USNM 228577, 228578) Pacific Wrens that we examined. The exposed culmen measures 7.8 mm, the wing chord 43.4 mm, the tail 17.5 mm, and the tarsus 15.6 mm. The bill and wing length are below the range of adults of either species (Pyle 1997). The tail is growing in and also shorter than the range of either species (Pyle 1997). It may not have been made Western Birds 44:141-142, 2013 141 NOTES clear to Ridgway that this was a juvenile taken from the nest, not one of the parents. We compared the plumage and measurements to a specimen of a juvenile House Wren (UWYMV 2191) that either was about to fledge or had recently fledged and found them to be very similar. Therefore we conclude that despite suspected nesting of Pacific Wren in Wyoming, this has yet to be confirmed. We thank M. Ralph Browning, Daniel D. Gibson, Douglas W. Faulkner, and Philip Unitt for their constructive reviews. LITERATURE CITED American Ornithologists’ Union. 1957. Check-list of North American Birds, 5 th ed. Am. Ornithol. Union, Baltimore, Maryland. American Ornithologists’ Union. 1998. Check-list of North American Birds, 7 th ed. Am. Ornithol. Union, Washington, D.C. Cary, M. 1917. Life zone investigations in Wyoming. Bull. U.S. Dept. Agric. Biol. Survey 42:1-95. Clements, J. F. 2007. The Clements Checklist of Birds of the World, 6 th ed. Cornell Univ. Press, Ithaca, NY. Dickinson, E. C. (ed.). 2003. The Howard and Moore Complete Checklist of the Birds of the World, 3 rd ed. Princeton Univ. Press, Princeton, NJ. Faulkner, D. W. 2010. Birds of Wyoming. Roberts and Co., Greenwood Village, CO. Hellmayr, C. E. 1934. Catalogue of birds of the Americas and the adjacent islands. Field Mus. Nat. Hist. Zool. Ser. 13, part 7. Knight, W. C. 1902. The birds of Wyoming. Univ. Wyo. Agric. Exp. Sta. Bull. 55:1-174. McCreary, O. 1939. Wyoming Bird Life, revised ed. Burgess, Minneapolis. Peters, J. L. 1960. Check-list of Birds of the World, vol. 9. Mus. Comp. Zool., Cambridge, MA. Phillips, A. R. 1986. The Known Birds of North and Middle America, part I. A. R. Phillips, Denver. Pyle, P. 1997. Identification Guide to North American Birds, part I. Slate Creek Press, Bolinas, CA. Ridgway, R. 1904. The birds of North and Middle America. Bull. U.S. Natl. Mus. 50, part 3. Toews, D. P., and Irwin, D. E. 2012. Pacific Wren (Troglodytes pacificus), in The Birds of North America Online (A. Poole, ed.), no. 720. Cornell Lab Ornithol., Ithaca, NY; http://bna.birds.cornell.edu/bna/species/720. Accepted 5 March 2013 142 NOTES LACK OF RECOVERY OF THE YELLOW-BILLED MAGPIE FROM THE WEST NILE VIRUS IN CALIFORNIA’S CENTRAL VALLEY EDWARD R. PANDOLFINO, 1328 49th Street, Sacramento, California 95819; erpfromca@aol.com The 2005 outbreak of West Nile virus (WNV) in California’s Central Valley was fol- lowed by declines in several species of birds, including the Loggerhead Shrike ( Lanius Iudovicianus), Western Scrub-Jay ( Aphelocoma californica ), Yellow-billed Magpie (Pica nuttalli), American Crow ( Corvus brachyrhynchos ), and Oak Titmouse ( Baeo - lophus inornatus) (Airola et al. 2007, Koenig et al. 2007, Pandolfino 2007, Crosbie et al. 2008, Pandolfino 2008a, Wheeler et al. 2009, Smallwood and Nakamoto 2009). Prior to the 2005 outbreak, the Loggerhead Shrike was acknowledged to be in decline across most of its range (Yosef 1996, U.S. Fish and Wildlife Service 2002), and the Yellow-billed Magpie had declined locally in some areas in the Coast Range and southern California (Roberson 1985, Lehman 1995, Koenig and Reynolds 2009). From 2008 to 2010 I monitored results from Central Valley Christmas Bird Counts (CBC) for signs of recovery among these species, finding that abundances of the Western Scrub-Jay, American Crow, and Oak Titmouse appear to have recovered to levels comparable to their pre-WNV level, while the Loggerhead Shrike and Yellow- billed Magpie showed no evidence of recovery (Pandolfino 2008b, 2009, 2010). For this analysis I used data from CBC circles (http://netapp.audubon.org/cbcob- servation/) and Breeding Bird Survey (BBS) routes (Sauer et al. 2011) in the Central Valley (Figure 1) from 1979 to 2012. To ensure that I was using data from within the normal range of the Yellow-billed Magpie, I included only CBC circles that had averaged at least one Yellow-billed Magpie per party hour prior to 2005 and BBS routes that had averaged at least 10 Yellow-billed Magpies per run prior to 2005. The numbers of Yellow-billed Magpies recorded on both CBCs (winter) and BBS routes (summer) every year since the 2005 WNV outbreak were lower than any recorded between 1979 and 2005 (Figure 2). The apparent effect of WNV can be seen on every CBC circle and BBS route (Table 1). Post-WNV numbers were lower on all 16 CBC circles and on all 10 BBS routes. The overall declines of 62% from CBC data and 56% from BBS data are comparable to prior published results for both the breeding season (declines of 22-83%; Koenig et al. 2007, Crosbie et al. 2009, Wheeler et al. 2009, Smallwood and Nakamoto 2009) and winter (42-64%; Airola et al. 2007, Pandolfino 2008b, Crosbie et al. 2009) and support the conclusion that recovery is not evident. Since 2006, anecdotal reports of local recovery of Yellow-billed Magpie numbers have appeared in various California list-serve discussions and web sites (e.g., http:// ebird.org/content/ca/news/yellow-billed-magpie-survey-report). At the suburban Cardinal Oaks Park east of Sacramento I noted that magpies, which were present in double-digit numbers prior to 2005, were completely absent during 2006 and 2007. Magpies began to reappear in this park in 2008 and were back to pre-2005 numbers by 2009. These anecdotal observations suggest that magpies may be reoccupying some prime locations where the local population was severely reduced or eliminated. However, data from CBCs and BBS routes showed no recovery on a broader scale. As noted above, even prior to WNV, the Yellow-billed Magpie had been extirpated from some localized areas in the Coast Range and southern California (Roberson 1985, Lehman 1995, Koenig and Reynolds 2009). Nevertheless, prior to 2005 the population was stable in the Central Valley and at the scale of the species’ entire range (Butcher and Niven 2007). The lack of recovery of the Yellow-billed Magpie in the Central Valley may be due to the continued presence of WNV. While there has been no outbreak on the scale Western Birds 44:143-147, 2013 143 1 ^^Redding AX 2 BX Susan ville *ATa G E cX Chico dX* E /X X* Yuba City __ Jf-3 ■ r..v X G RosevUle 8" Sacramento 10 9 Vacaville Elk Grove •Jfc A L I F O R N I A N apa ' V al I ej o * ’Fairfield H Concord * X . I- 13 # Antioch Stockton * San Leandro .Tracy ! CISCO • Livermore* 1 /' Modesto 14 15 • * San Mateo Fremont X • San Jose M erced Santa Cruz » Los Banos 4 <5 JOAQUIN Vt Figure 1. Approximate locations of the 16 CBC circles (filled circles) and 10 BBS routes (crosses) used for analyses of trends in numbers of the Yellow-billed Magpie. CBC circles: 1, Redding; 2, Red Bluff; 3, Chico; 4, Oroville; 5, Peace Valley; 6, Marysville; 7, Lincoln; 8, Folsom; 9, Sacramento; 10, Putah Creek; 11, Rio Cosumnes; 12, Wallace-Bellota; 13, Stockton; 14, Caswell-Westley; 15, La Grange-Water ford; 16, Los Banos. BBS routes: A, Red Bluff; B, Cohasset; C, Orland; D, Glenn; E, Oroville; F, Pennington; G, Zamora; H, Staten Island; I, Hughson; J, Westley. 144 NOTES Figure 2. Abundance of the Yellow-billed Magpie on Central Valley CBCs from count year 79 (winter 1978-79) to count year 112 (winter 2011-2012) and on BBS routes from 1979 to 2011. seen in 2005 since that year, WNV continues to be present in the region (Wheeler et al. 2009, Centers for Disease Control 2012). Infected magpies’ extremely high rate of mortality (Crosbie et al. 2008, Wheeler et al. 2009) and the apparent absence of seropositive birds (with antibodies to WNV) among magpies that survived the 2005 outbreak (Crosbie et al. 2008) suggest that there may be very few WNV-resistant individuals in the Central Valley. Another major WNV outbreak could cause a further loss of Yellow-billed Magpies in the region. Data on the numbers of infected birds and human beings suggest that 2012 saw the highest levels of WNV in the southern Central Valley since 2005 (Centers for Disease Control 2012). Therefore, monitoring of CBC and BBS data after 2012 may show further local declines in magpies there. Wheeler et al. (2009) noted that, with WNV now endemic to the Central Valley, localized species are less likely to recover from population losses. The entire range of the Yellow-billed Magpie lies within areas of persistent WNV. Therefore, recovery of the population to pre-WNV levels may be unlikely. In addition, the continuing conversion of oak savanna to urbanization and intensive agriculture such as vineyards (Zach et al. 2002) further reduces the habitat available for magpies. On a more positive note, the Yellow-billed Magpie population in the Central Valley has adapted well to human-altered landscapes such as rural residential and suburban areas. This provides some hope that, if a WNV-resistant 145 NOTES Table 1 Average Abundance of the Yellow-billed Magpie pre- and post-West Nile Virus Pre-WNV a Post-WNV b Change CBC circle Chico 5.7 ± 1.0 2.8 ± 0.5 -52% Caswell-Westley 22.9 ± 3.6 3.6 ± 0.7 -84% Folsom 1.7 ± 0.7 0.4 ± 0.04 -78% Lincoln 5.1 ± 0.6 1.1 ± 0.2 -78% Los Banos 6.4 ± 0.6 3.0 ± 1.4 -53% La Grange-Waterford 3.9 ± 0.5 2.2 ± 0.4 -43% Marysville 37.9 ± 20 1.6 ± 0.2 -96% Oroville 1.0 ± 0.1 0.2 ± 0.1 -75% Putah Creek 3.8 ± 0.2 2.4 ± 0.2 -38% Peace Valley 2.5 ± 0.2 0.5 ± 0.2 -79% Red Bluff 3.3 ± 0.3 1.6 ± 0.1 -52% Rio Cosumnes 3.7 ± 0.3 0.9 ± 0.1 -76% Redding 1.3 ± 0.1 0.6 ± 0.1 -49% Sacramento 6.1 ± 0.4 3.1 ± 0.3 -48% Stockton 1.4 ± 0.1 0.8 ± 0.2 -40% Wallace-Bellota 2.8 ± 0.2 1.7 ± 0.2 -39% All circles combined 4.2 ± 0.2 1.6 ± 0.1 -62% BBS route Cohasset 19 ± 2 13 ± 1 -30% Glenn 26 ± 3 3 ± 1 -87% Hughson 49 ± 5 33 ± 8 -34% Orland 35 ± 4 9 ± 2 -74% Oroville 40 ± 3 24 ± 3 -40% Pennington 29 ± 3 9 ± 2 -70% Red Bluff 11 ± 1 2 ± 1 -86% Staten Island 12 ± 3 7 ± 1 -44% Westley 60 ± 5 17 ± 7 -72% Zamora 54 ± 7 28 ± 6 -48% All routes combined 35 ± 2 15 ± 2 -56% a Birds per party-hour 1979-2005 for CBC data, birds per route 1979-2005 for BBS data. b Birds per party-hour 2006-2012 for CBC data, birds per route 2006-2011 for BBS data. subpopulation develops, it may find ample habitat in which to expand. In contrast to magpies in the Central Valley, those in the Coast Range and southern California do not readily use urbanized landscapes (Koenig and Reynolds 2009), and that population may be at higher risk of further declines and/or range contractions. I thank Dan Airola, Walter Koenig, Paul Lehman, and Philip Unitt for their helpful reviews and suggestions, which improved the focus and clarity of the paper. I am also indebted to the many volunteers who participate in Christmas Bird Counts and run Breeding Bird Surveys. These data sources are invaluable tools to help us understand avian population trends. LITERATURE CITED Airola, D. A., Hampton, S., and Manolis, T. 2007. Effects of West Nile virus on sensitive species in the lower Sacramento Valley, California: An evaluation using Christmas Bird Counts. Central Valley Bird Club Bull. 10:1-22. 146 NOTES Centers for Disease Control. 2012. West Nile virus: Statistics, surveillance, and control archive ; http : / /www. cdc .gov/ ncidod/ dvbid/ westnile/ surv&control . htm [accessed January 2013]. Crosbie, S. P., Koenig, W. D., Reisen, W., Kramer, V. L., Marcus, L., Carney, R., Pandolfino, E. R., Bolen, G. M., Crosbie, L. R., Bell, D. A., and Ernest, H. B.. 2008. A preliminary assessment: Impact of West Nile virus on the Yellow-billed Magpie, a California endemic. Auk 125:542-550. Koenig, W. D., Marcus, L., Scott, T. W., and Dickinson, J. L. 2007. West Nile virus and California breeding bird declines. Ecohealth 4:18-24. Koenig, W., and Reynolds, M. D. 2009. Yellow-billed Magpie ( Pica nuttalli), in The Birds of North America Online (A. Poole, ed.), no. 180. Cornell Lab of Ornithol., Ithaca, NY; http://bna.birds.cornell.edu/bna/species/180 [accessed February 2013]. Lehman, P. E. 1994. The Birds of Santa Barbara County, California. Vert. Mus., Univ. of Calif., Santa Barbara. Pandolfino, E. R. 2007. Review of the 107th Christmas Bird Count in the Central Valley of California: December 2006-January 2007. Central Valley Bird Club Bull. 10:53-59. Pandolfino, E. R. 2008a. Population trends of the Loggerhead Shrike in California: Possible impact of West Nile virus in the Central Valley. Central Valley Bird Club Bull. 11:37-44. Pandolfino, E. R. 2008b. Review of the 108th Christmas Bird Count in the Central Valley of California: December 2007-January 2008. Central Valley Bird Club Bull. 11:53-61. Pandolfino, E. R. 2009. Review of the 109th Christmas Bird Count in the Central Valley of California: December 2008-January 2009. Central Valley Bird Club Bull. 12:53-62. Pandolfino, E. R. 2010. Review of the 110th Christmas Bird Count in the Central Valley of California: December 2009-January 2010. Central Valley Bird Club Bull. 13:25-34. Roberson, D. 1985. Monterey Birds. Monterey Peninsula Audubon Soc., Carmel, CA. Sauer, J. R., Hines, J. E., Fallon J. E., Pardieck, K. L., Ziolkowski, D. J., Jr., and Link, W. A. 2011. The North American Breeding Bird Survey, results and analysis 1966-2010, version 12.07.2011. U.S. Geol. Survey Patuxent Wildlife Research Center, Laurel, MD. Smallwood, K. S., and Nakamoto, B. 2009. Impacts of the West Nile virus epizootic on the Yellow-billed Magpie, American Crow, and other birds in the Sacramento Valley, California. Condor 111:247-254. U.S. Fish and Wildlife Service. 2002. Birds of conservation concern 2002. Division of Migratory Bird Management, Arlington, VA. Wheeler, S. S., Barker, C. M., Fang, Y., Armijos, M. V., Carroll, B. D., Husted, S., Johnson, W. O., and Reisen, W. K. 2009. Differential impact of West Nile virus on California birds. Condor 111:1-20. Yosef, R. 1996. Loggerhead Shrike ( Lanius ludovicianus), in The Birds of North America (A. Poole and F. Gill, eds.), no. 231. Acad. Nat. Sci., Philadelphia. Zach, S., Ballard, G., Chase, M., Elliott, G., Gardali, T. Geupel, G. R., Heath, S., Humple, D., Lynes, M., Pitkin, M., Scoggin, S., and Stralberg, D. 2002. The oak woodland bird conservation plan: A strategy for protecting and managing oak woodland habitats and associated birds in California, version 2.0. PRBO Conservation Science, Stinson Beach, CA. Accepted 15 March 2013 147 NOTES FIRST RECORD OF THE RED-BELLIED WOODPECKER IN NEVADA KENNETH M. BURTON, P. O. Box 1038, Areata, California 95518; shrikethree@gmail . com LESLIE SCOPES ANDERSON, 64 E. 14 th St., Areata, California 95521 On the afternoon of 12 June 2012 Anderson found an unfamiliar woodpecker in cottonwoods (Populus sp.) near the Bressman Historic Cabin at Ruby Lake National Wildlife Refuge, Elko County, Nevada (40.2° N, 115.4° W). She observed the bird for approximately 20 minutes as it foraged, during which time she obtained several excellent photographs from which she was able to identify it as a female Red-bellied Woodpecker ( Melanerpes carolinus). Because Anderson did not realize the significance of her find, she did not report it. Two days later she shared a photo with Burton, who suspected it represented a first state record and notified the Nevada Bird Records Committee. The committee’s secretary, Martin Meyers, posted the sighting on the state birding e-list, but by then several days had elapsed and the bird was not refound. On the basis of Anderson’s Figure 1. Female Red-bellied Woodpecker ( Melanerpes carolinus ) at Ruby Lake National Wildlife Refuge, Nevada, 6 June 2012; ventral view. Photo by Leslie Scopes Anderson 148 Western Birds 44:148-150, 2013 NOTES Figure 2. Female Red-bellied Woodpecker (Melanerpes carolinus) at Ruby Lake National Wildlife Refuge, Nevada, 6 June 2012; dorsal view. Photo by Leslie Scopes Anderson definitive photographs, however, the committee subsequently voted unanimously in the first round to accept the record (2012-030), and it stands as the first for Nevada (M. Meyers pers. comm.). The photos clearly show the combination of red nape and nasal tufts; white-barred back and central rectrices; gray face, crown, and underparts; and nearly concealed red belly patch diagnostic of a female Red-bellied Woodpecker (Figures 1 and 2). The similar female Golden-fronted Woodpecker (M. aurifrons) has yellow head patches, 149 NOTES black central rectrices, and no belly patch. Some of the photos appear to show worn, brown primaries contrasting with fresher, blacker secondaries and secondary coverts, suggesting a bird in its first plumage cycle; during the preformative molt, a small pro- portion of individuals retain a variable number of outer primaries, while subsequent molts include all primaries (Pyle 1997). The Red-bellied Woodpecker’s normal range extends from the eastern United States up river valleys onto the Great Plains as far west as extreme northeast Colorado (South Platte River; Winternitz 1998), approximately 1100 km due east of Ruby Lake. Since the mid-1900s, the species has expanded its range north and west, likely as a result of climate change and supplemental feeding (Shackelford et al. 2000); Breed- ing Bird Survey data show strong population increases from 1966 to 2010 across most of its range (Sauer et al. 2011). Like most other woodpeckers, it is essentially nonmigratory (Shackelford et al. 2000), although northern populations are prone to some southward movement during cold winters (Winkler et al. 1995), and it is at that season that vagrants may be expected. In fact, there is a limited pattern of winter records west of the breeding range (e.g., at least two for Alberta and one for Idaho; http://birdscalgary.wordpress.com, http://geobirds.com, http://idahobirds.net), and it may be expected that continued population growth will result in an increase in extralimital records. A Red-bellied Woodpecker in Coeur d’Alene, Idaho, in the winter of 2002-2003 (Rogers 2003) is the only one recorded west of Ruby Lake, and there are no records in any of the other states bordering Nevada. The June date of the Ruby Lake bird makes the sighting all the more remarkable. We thank Kristie Nelson and Daniel D. Gibson for reviewing the manuscript. LITERATURE CITED Pyle, P. 1997. Identification Guide to North American Birds, part 1. Slate Creek Press, Bolinas, CA. Rogers, T. H. 1993. Idaho-western Montana region. Am. Birds 47:278-279. Sauer, J. R., Hines, J. E., Fallon, J. E., Pardieck, K. L., Ziolkowski, D. J., Jr., and Link, W. A. 2011. The North American Breeding Bird Survey, Results and Analysis 1966-2010, version 12.07.2011. U.S. Geol. Survey Patuxent Wildlife Research Center, Laurel, MD. Shackelford, C. E., Brown, R. E., and Conner, R. N. 2000. Red-bellied Woodpecker (Melanerpes carolinus ), in The Birds of North America (A. Poole and F. Gill, eds.), no 500. Birds N. Am., Inc., Philadelphia. Winkler, H., Christie, D. A., and Nurney, D. 1995. Woodpeckers: A Guide to the Woodpeckers of the World. Houghton Mifflin, New York. Winternitz, B. L. 1998. Red-bellied Woodpecker, in Colorado Breeding Bird Atlas (H. E. Kingery, ed.), pp. 252-253. Colo. Bird Atlas Partnership and Colo. Div. Wildlife, Denver. Accepted 24 February 2013 150 NOTES CALIFORNIA CONDOR FORAGING ON A LIVE CALIFORNIA SEA LION FUF MIKE TYNER, L. JOSEPH BURNETT, and MIKE M. STAKE, Ventana Wildlife Society, 19045 Portola Drive, Suite F-l, Salinas, California 93908; mikestake @ ventanaws . org California Condors (Gymnogyps californianus ) feed on a variety of wild and domestic carrion, but are not known to take live prey in the wild (Snyder and Sny- der 2000, Snyder and Schmitt 2002). There exists, however, an historical account (Townsend 1848) of a California Condor apparently attempting to take a live salmon along a river bank in the Pacific Northwest. However, the bird was collected before the fish was killed, and it is not clear if feeding was observed. A sister species, the Andean Condor ( Vultur gryphus), has, however, been reported to take live prey (Murphy 1925), and smaller species of New World vultures, the Black ( Coragyps atratus) and Turkey ( Cathartes aura) Vultures have also been reported to take live prey occasionally (Parmalee 1954, Kirk and Mossman 1998, Buckley 1999, Paves et al. 2008). While not providing clear evidence of taking live prey, Townsend's ac- count indicates the opportunistic foraging by California Condors is not confined to dead animals. Nevertheless, occasions for documenting California Condors with live prey have been limited, because, historically, few individuals have been available for observation. Even with intensive conservation efforts during the past 20 years that have resulted in increased numbers of birds in the wild, the practice of provisioning carcasses at designated feeding sites limits opportunities for observation of natural foraging, especially rare events such as attempts to take live prey. In 1997, Ventana Wildlife Society began releasing captive-reared California Condors along the central California coast. The restored population of this region, including condors released by the National Park Service at Pinnacles National Park, numbered 66 birds as of 2012. This population continues to be fed supplementally, like populations reintroduced in southern California, Arizona, and Baja California. Although in each region condors have been observed feeding on carcasses of land mammals other than those provided through the reintroduction program, the central California population is unique among reintroduced populations because it also feeds on carcasses of marine mammals naturally occurring along the ocean shore (Burnett et al. 2013). Because at least several hundred California Sea Lions ( Zalophus califor- nianus) congregate most days in a single rocky cove near Big Sur (36° 12' N, 121° 43' W), condors regularly find and consume sea lion carcasses at this site. We first observed condors feeding on sea lion carcasses at this location in 1999. Since then, the site has provided numerous opportunities for observation of condors foraging, not only of birds feeding on carcasses but also of interactions between condors and live sea lions. On 20 June 2010, Mike Tyner witnessed a California Condor attempting to for- age on a live sea lion at the Big Sur site. The observation began at 11:05 PDT with four condors perched on a large rock on the beach near a dozen or more sea lions. These birds, identified by numbered wing tags, had been feeding on carcasses at this location almost daily for more than a month. On the ground about 1 m below the condors was a live sea lion pup appearing weak and malnourished. One of the condors (#222, Figure 1), a 10-year-old breeding female, approached the sea lion and began pecking at the posterior of the pup. The condor grasped the pup with its beak as the sea lion struggled slowly in the direction of the water, leaving a trail of blood. The sea lion then turned toward a shady spot closely overhung by a second large rock about 1 m away. The pup did not strike back at the condor during the observation, and none of the adult sea lions in the area defended the pup. Eventually, both the sea lion and Western Birds 44:151-154, 2013 151 NOTES Figure 1 . Adult female California Condor that killed and consumed a debilitated pup of the California Sea Lion near Big Sur, California, 20 June 2010. Photo by Tim Huntington the condor were under the rock where the condor could be seen tugging at the sea lion pup. A few minutes later, the condor dragged the now lifeless pup from under the rock and began reaching with its head into the anal orifice. After feeding for 20 minutes, her crop appeared full, and she left the carcass. During the next hour, 11 other condors fed on the sea lion pup, reducing the carcass to mostly bones. Species of the vulture family are primarily scavengers, but many can kill live prey, as supported by our observation, those of Black Vultures feeding on live seal pups in Chile (Paves et al. 2008), and even observations of California Condors killing and eating live rats and mice in captivity at the Los Angeles Zoo (M. Clark pers. comm.). Nevertheless, special circumstances likely fostered the observation reported here. In June 2010, an unusually large number of sea lion pups was abandoned along the central California coast. Many of these were emaciated, a condition attributed by Melin et al. (2010) to anomalous oceanographic conditions in 2009, warmer sea-surface temperatures reducing the availability of sea lion prey within the normal foraging range of sea lions. This necessitated sea lions having to make longer foraging trips (Melin et al. 2010) under conditions that persisted through the period of pupping. The consequences were seen in females giving birth before reaching normal offshore pupping areas, females abandoning pups, and, ultimately, a nutritional deficit in many pups (Melin et al. 2010). The end result was an unusual abundance of malnourished pups in a predictable location that provided condors with consistent and easy op- portunities for foraging (Figure 2). Our observation has at least one important implication for condor management, namely, in further documenting the ability of condors in recovered populations to locate prey independent of human intervention and demonstrating their ability to take advantage of unusual circumstances opportunistically. For populations supported 152 NOTES Figure 2. Two California Condors feeding on the pup of a California Sea Lion near Big Sur, California. Photo by Tim Huntington by conservation management, such ability can be interpreted as a positive sign that condors will be able to forage successfully if supplemental feeding is scaled back in the future. However, as long as lead poisoning threatens the sustainability of condor populations (Finkelstein et al. 2012, Rideout et al. 2012), providing carcasses is important not only to ensure a source of uncontaminated food but also to facilitate other necessary management, such as trapping for health exams, providing treatments for lead exposure, and maintaining equipment for tracking. The risk of lead exposure for condors feeding on marine mammals along the central California coast is likely reduced, but other harmful contaminants might be present in marine carcasses (Bur- nett et al. 2013), and the threat of lead exposure persists for these birds when they forage on carcasses of land mammals (Sorenson and Burnett 2007). Our observation need not prompt changes in condor management but provides more information on the condor’s capabilities to take advantage of opportunities for foraging. On 30 November 2011, lead author Mike Tyner (age 35) was killed by a falling branch while monitoring a newly released California Condor at Big Sur. Mike was best known for his unassuming leadership, his love of outdoor adventure, and his dedica- tion to avian conservation. In five years as a field crew leader for Ventana Wildlife Society, Mike played an important role in the California Condor recovery program, by supervising releases, assisting with condor-rescue efforts during a major wildfire, training young field biologists, and contributing field observations that have advanced our knowledge of condor biology. LITERATURE CITED Buckley, N. J. 1999. Black Vulture ( Coragyps atratus), in The Birds of North America (A. Poole and F. Gill, eds.), no. 411. Birds N. Am, Inc., Philadephia. Burnett, L. J., Sorenson, K. J., Brandt, J., Sandhaus, E. A., Ciani, D., Clark, M., David, C., Theule, J., Kasielkie, S., and Risebrough, R. W. 2013. Eggshell 153 NOTES thinning and depressed hatching success of California Condors reintroduced to central California. Condor 115: in press. Finkelstein, M. E., Doak, D. F., George, D., Burnett, J., Brandt, J., Church, M., Grantham, J., and Smith, D. R. 2012. Lead poisoning and the deceptive recovery of the critically endangered California Condor. Proc. Natl. Acad. Sci. 109:11449-11454. Kirk, D. A., and Mossman, M. J. 1998. Turkey Vulture ( Cathartes aura), in The Birds of North America (A. Poole and F. Gill, eds.), no. 339. Birds N. Am., Inc., Philadelphia. Melin, S. R., Orr, A. J., Harris, J. D., Laake, J. L., DeLong, R. L., Guilland, F. M. D., and Stoudt, S. 2010. Unprecedented mortality of California sea lion pups associated with anomalous oceanographic conditions along the central California coast in 2009. Calif. Coop. Oceanic Fish. Inv. Rep. 51:182-194. Murphy, R. C. 1925. Bird Islands of Peru. Putnam, New York. Parmalee, P. W. 1954. The vultures: their movements, economic status, and control in Texas. Auk 71:443-453. Paves, H. J., Schlatter, R. P., and Espinoza, C. I. 2008. Scavenging and preda- tion by Black Vultures Coragyps atratus at a South American sea lion breeding colony. Vulture News 58:4-15. Rideout, B. A., Stalis, I., Papendick, R., Pessier, A., Puschner, B., Finkelstein, M. E., Smith, D. R., Johnson, M., Mace, M., Stroud, R., Brandt, J., Burnett, J., Parish, C., Petterson, J., Witte, C., Stringfield, C., Orr, K., Zuba, J., Wallace, M., and Grantham, J. 2012. Patterns of mortality in free-ranging California Condors (Gymnogyps califomianus). J. Wildlife Diseases 48:95-112. Snyder, N. F., and Schmitt, N. J. 2002. California Condor ( Gymnogyps california- nus), in The Birds of North America (A. Poole and F. Gill, eds.), no. 610. Birds N. Am., Inc., Philadelphia. Snyder, N. F., and Snyder, H. 2000. The California Condor: A Saga of Natural History and Conservation. Academic Press, San Diego. Sorenson, K. J., and Burnett, L. J. 2007. Lead concentrations in the blood of Big Sur California Condors, in California Condors in the 21 st Century (A. Mee and L. S. Hall, eds.), pp. 185-195. Nuttall Ornithol. Club, Cambridge, MA. Townsend, J. K. 1848. Popular monograph on the accipitrine birds of N.A. — No. 11. Literary Record and Journal of the Linnaean Association of Pennsylvania College 4:265-272. Accepted 14 March 2013 154 IN MEMORIAM RICHARD W. STALLCUP, On 15 December 2012, during the compilation of the Point Reyes Christmas Bird Count in Point Reyes Station, Rich Stallcup passed away in a San Rafael hospital from complications of leukemia. Rich was a co-founder of that count in 1970 and had partici- pated in it every year until 2012. The birding community in California lost a soul of incalculable value as a teacher, naturalist, tour leader, and indefatigable cheerleader for the natural world. His reputation extended far beyond the borders of California; he was well known in birding circles across North America. An ordinary memorial cannot express the depth of his influence on people and how many lives he touched dur- ing his time on this planet. In January 2013, more than 500 people, some from as far away as the east coast, attended a memorial service for Rich. Born Richard William Stallcup in Oakland, California, on 19 December 1944, Rich was educated in the Oakland public school system and attended California State University, Hayward. He had a single sibling, his older brother Kenneth. Married once, he had one daughter, Willow, and eventually two grandchildren. He spent his remaining years with his life partner, Heather Cameron. His interest in birds began at the early age of six, when his father Leland took him on a field trip to Drake’s Bay on the Point Reyes Peninsula. Throughout his life and international travels, Rich is remembered most for his love of Point Reyes and for living and spending most of his time in that region of California. Although he made over a thousand trips to Point Reyes he never lost his enthusiasm and fascination for the place. As he grew as a naturalist, his interest in things such as reptiles, amphibians, insects, and flora broadened his view of the natural world. He possessed a profound understanding of life on the Point Reyes Peninsula. Even the wandering to the outer point of a bird common only a few miles inland was of major interest to him. Beginning in the 1960s, Rich, along with confederates Guy McCaskie and C. J. Ralph, began to find rare vagrants regularly on outer Point Reyes, a phenomenon that has attracted birders to this area of California ever since. He wrote a field checklist (2000) of the Birds of the Point Reyes Seashore that totaled 470 species, of which well over half were vagrants, nearly all of which Rich had seen personally. Beginning in 1967 he also spent a great deal of time on the Farallon Islands, making frequent 2-week trips there between 1973 and 1978. In 1969, the Farallons became a national wildlife refuge. Rich’s views of life and nature took on a particular and perhaps inevitable orientation during the counterculture life of the late 1960s and early 1970s. Although drafted and trained as a U.S. marine during the Vietnam War, Rich abhorred the thought of killing another human being, and he was able get a discharge from further service. However, the experience had an indelible impact on his thinking. He realized that the web of life was entirely connected but that mankind had the single greatest influence on life, an influence seriously flawed by greed, ego, and a profound lack of understanding. Rich’s understanding of this principle guided him throughout his life as he sought to 1944-2012 Western Birds 44:155-157, 2013 155 IN MEMORIAM impart that connection to the many people he taught. As his classroom he preferred the field, where the immediacy of life was all around him. He fought many battles on behalf of wildlife that has no voice in the affairs of mankind. In 1976 Rich joined the formation of the early tour group Wings and led dozens of trips in western North America and Mexico, but being away from Point Reyes and his family for extended periods took its toll, and he began to limit his tour leading mainly to California and Arizona. In 1975 and 1976 I was fortunate enough to co-lead with Rich two trips sponsored by PRBO Natural History Excursions, both of which will be long remembered. His love of Arizona rivaled his love of California. Rich served as president of Western Field Ornithologists from 1975 to 1978 and was a founding member of the California Bird Records Committee, serving for 8 years. In 1967 he was the first resident bird bander at the fledgling Point Reyes Bird Obser- vatory, set up in a converted chicken coop on outer Point Reyes. From those early days PRBO grew into an international conservation and research organization with a large staff and large annual budget. Rich maintained a close association with PRBO through most of his adult life and joined the staff as a resident naturalist in 1997. In 1982 the American Birding Association bestowed on him the Ludlow Griscom Award for outstanding contributions to American ornithology. In that same year he began a regular column in PRBO’s newsletter (later called the Observer) that covered a vast range of topics relating to his personal observations and offering readers a unique personal perspective. He also served for 11 seasons as a co-editor of the quarterly report for the Middle Pacific Coast region in American Birds. The entire winter 2013 issue of the PRBO Observer (number 171) was dedicated to Rich and offers enriching glimpses of his life and legacy. It is a must read for everyone who knew him. In 1966, when the Guide to the Field Identification of Birds of North America by C. S. Robbins, B. Bruun, and H. S. Zim was published by Golden Press, it was the first such field guide beside the Peterson series to appear since the early 1940s. As did a number of his contemporaries, Rich found that much of the book did not represent birds he knew in life and took to correcting the plates with a felt-tip pen. The guide was ultimately revised in 1983, but problems persisted. Rich then set about detailing the flaws in the Golden Nature Guide, and the result was a slim volume titled Birds for Real, which Rich published himself in 1985. Not only did this book reveal how much Rich knew about the species he covered, it also set the stage for field guides to follow and what would be demanded to cover the subject of bird identification adequately. The market for field guides became very competitive, and today most publishers are keenly aware of what a successful book must deliver. Today we have a plethora of excellent field guides that cover North America, Mexico, and South America that met this challenge. The entire ornithological community owes a debt to Rich for taking the time to raise the standard with something that went far beyond a typical book review. Rich had a natural fascination with the sea and spent as much time as possible on boat trips off the shores of California. He was well known for his understanding of bird life at sea as well as of marine mammals. Yet for his first 25-30 boat trips he had a hard time with sea sickness. He eventually learned to control it without drugs, preferring to remain alert and vigilant, never knowing what might fly across the bow of the boat. Rich’s enthusiasm was infectious, and he was often as excited at seeing a bird he had seen many times as he was the first time he encountered it. No one ever left one of his boat trips without feeling enlightened about something. His experi- ence in the coastal waters of California resulted in Ocean Birds of the Nearshore Pacific (1990), in which he described the mechanics of seabirding and addressed the species’ identification with descriptions and photographs, including some species not formally recorded in California at the time. Ocean Birds of the Nearshore Pacific is also an excellent primer on coastal oceanography and includes marine mammals, sea turtles, and some fish. It remains useful to this day. He was a senior or junior author or a contributor on dozens of reports, papers, checklists, and books during his life. 156 IN MEMORIAM Rich’s connection to the natural world was nearly mystical, and he imparted deep insights to his many followers. That was his true legacy, not a list of publications. He followed in the steps of many 19 th century naturalists who were guided by their interest in natural history but who lacked a formal education in a given discipline, education that did not exist in the 19 th century. When natural history became a “science,” a great deal of what animates the natural world was lost at the expense of examining how the individual pieces worked and could be described. Rich never lost the importance of a wider connection, yet he understood and integrated the approach that “science” took. He had the soul of a poet, the mind of a scientist, and the spirit of a shaman — a combination of talents not often found in anyone, let alone a naturalist. Rich’s ability to find rare or uncommon birds was legendary. He often found birds where it seemed none were present, and for those who were fortunate to be in the field with him, this happened too often to be due merely to luck or chance. It was well beyond something that could be attributed to so mundane or trivial an explanation. There are too many such stories to be recounted here, but his ability was extraordi- nary. Some have said that he was the most gifted field ornithologist of his generation. Rich was a dear friend for 46 years, a colleague, co-author, and fellow birder. He will be greatly missed by all who were fortunate enough to know him. He was one of a kind, with a staggering talent that few of us can ever hope to imagine. Jon Winter 157 BOOK REVIEWS Raptors of New Mexico, edited by Jean-Luc Cartron. 2010. University of New Mexico Press. 728 pages, over 700 color photographs and 50 color distribution maps. Hardback, $50. ISBN 978-0-8263-4145-7. As the title suggests, this book reviews the birds of prey regularly occurring in New Mexico with additional information on the vagrant species that occasionally appear in the state. This is no small task given that New Mexico is the fifth largest state in the U.S. and boasts the fourth highest number of avian species recorded, so raptors are well represented in the Land of Enchantment. Forty-four documented raptor species, including breeding species as diverse as the Boreal Owl and Aplomado Falcon, make it unique among the 50 states. Each of the 37 regularly occurring species of raptors has its own lengthy section, ranging from 10 to 28 pages, that contains a detailed analysis of the species within New Mexico. Besides Cartron, the chapters are written by a number of authors who have studied a particular species in depth. Additionally, contributions of photographs came from over 100 individuals, so this book is truly a collaboration. This is a landmark publication for New Mexico ornithology as the first publication to take on only the raptors. For each species account and migration data the book pulls an enormous amount of information from the peer-reviewed literature, but it is worded in a fashion that amateur ornithologists and birders can easily follow. The editor breaks the book into the introduction and two main parts, the introductory chapters and the species accounts. Each section is extremely detailed, and the editor sets the stage for the species accounts well. The introduction and introductory chapters address the origin and definition of the term “raptor,” the plant communities of New Mexico, and long-term data on raptor migration from monitoring sites within the state. “Raptor” has a seemingly new meaning in the wake of new hypotheses about the birds’ relationships arising from recent molecular studies. It is essential then that Cartron introduces the definition of the term “raptor,” where it came from, and how it is applied in this book. He does a nice job of this by introducing the first ornitholo- gists to use “raptor” and the criteria they used to define one. The introduction also speculates about the potential phylogenetic relationships of the raptors and how these relationships may affect our understanding of their evolutionary history. The plant communities of New Mexico provide raptors a wide variety of habitats. The section addressing them provides a clear overview of their distribution, valuable for the reader interpreting habitat descriptions in the species accounts. For instance, it is important to know the difference between “montane conifer forest” and “Great Basin conifer woodland” because their avifaunas differ. The migration data presented in this section are impressive. Thirty-five pages are dedicated to long-term banding and observation of migrating raptors, primarily in the Sandia and Manzano mountains of central New Mexico. Migration timing, tem- poral shifts in abundance, mark-and-recapture studies, satellite-transmitter data, and measurements of birds captured in the field are just some of the topics covered in this section. A good example of the data in this chapter is Figure 2.5 (p. 43), which depicts the movements of 1 1 individual Golden Eagles that were fitted with satellite transmitters in the Manzano Mountains and subsequently found to move as far north as Alaska and south to Mexico. Such information makes this chapter valuable for anyone who wants to understand the movements of raptors through or within New Mexico. The 37 species accounts of raptors regularly occurring in New Mexico make up the bulk of this book, nearly 600 pages. They present detailed information on distribution, habitat associations, life history, status, and management. Also, each account has a large map of the species' distribution in New Mexico, with clear symbols representing breeding, migration, etc. Each species account is carefully constructed so that each presents roughly parallel information, but some accounts lack information that others 158 Western Birds 44:158-160, 2013 BOOK REVIEWS have. For instance, the account for Cooper’s Hawk has a section “Migration” under “Life History” while a similar section is lacking in the account for Swainson’s Hawk. As one might expect with many authors, the writing style differs noticeably from ac- count to account, but this in no way takes away from the information presented to the reader. Each account is well written, and it is obvious that the authors have a passion for their respective species. For example, in the account for Harris’s Hawk are several photographs of author James C. Bednarz holding fledglings or monitoring a nest. The species accounts end with ten pages dedicated to seven species of vagrants that have been confirmed in New Mexico. Each of these accounts contains information on the records from the state and speculation on population trends for the species as a whole. Some include photographs of the species from New Mexico, such as of the rufous morph of the Eastern Screech-Owl in Portales in 2003. It is hard to imagine a bookshelf of anyone interested in ornithology in New Mexico or the Southwest without Raptors of New Mexico. With detailed information on migration dynamics, 44 in-depth species accounts, and over 700 color photographs it is a major contribution to New Mexico ornithology and sets the bar for future publications dedicated to raptors. Matthew J. Baumann Twelve Hundred Miles by Horse and Burro: J. Stokley Ligon and New Mexico’s First Breeding Bird Survey, by Harley G. Shaw and Mara E. Weisen- berger. 2011. University of Arizona Press, Tucson, xv + 244 pp., 36 black-and-white photographs. Paperback, $26.95. ISBN 978-0-8165-2861-5. J. Stokley Ligon (1879-1961) was a New Mexico ornithologist and conservationist during one of the most dynamic periods in the state’s history. Dale Zimmerman noted that Ligon “Probably . . . covered New Mexico more thoroughly than any other natural- ist before or during his lifetime.” Born and raised on a ranch in Texas, Ligon was a self-trained ornithologist who spent his twenties drilling wells and fixing windmills in west Texas and southern New Mexico. He was well acquainted with trapping predators and fur-bearers from his time on the ranch and family hunting trips. Later in life he was paid to manage teams of trappers that extirpated the Mexican wolf and grizzly bear from New Mexico. He eventually softened his views on predator control and in 1927 successfully lobbied the state legislature to designate bears as game animals, rather than pests to be shot on sight. Ligon is probably best known to contemporary ornithologists and birders from his New Mexico Birds and Where to Find Them (1961), an ambitious but relatable work that describes the state’s bird life and birding locations, along with more general topics such as life zones and bird conservation. Twelve Hundred Miles by Horse and Burro is an account of a little-known period of Ligon’s life: his first job as a professional biologist working for the U.S. Biological Survey (USBS), to assess breeding waterbirds around New Mexico in the summer of 1913. The book begins by describing Ligon’s life prior to the start of his survey and his appointment to the USBS. The bulk of the book consists of day-by-day accounts of the survey, which contain many excerpts from Ligon’s field diary and an unpublished report on his work. As the book notes, Ligon’s diary entries are usually brief and fairly vague. Although Ligon likely kept a personal account of his travels, it seems to have been lost to history (Ligon sent his field diary and report to the USBS, and they are currently housed at the Smithsonian). In the diary he rarely mentions events that do not directly relate to the survey and does not provide complete species lists for each day. Most of his entries are focused on waterbirds, as per his instructions, so refer- ences to land birds are often scant. The authors compensate for the lack of detail by providing meticulously researched narration that retraces Ligon’s route as closely as possible while providing a historical context to the locations he visited. In addition to 159 BOOK REVIEWS Ligon’s personal writings, over a dozen of his pictures from the survey are included, and many are compared to current photos of the same location. The authors note that the book is meant to be not only an account of Ligon’s travels but also an explo- ration of the changes in the way of life and landscape of New Mexico over the past century. In order to accomplish this aim they revisited a number of sites along Ligon’s route and compared the habitats and wildlife Ligon reported to what they observed. One of the most appealing aspects of the book is its extensive use of primary sources. The authors faithfully transcribed seemingly every entry from Ligon’s diary and much of his formal report, including errors and edits made by Ligon and his su- pervisors. The authors infer that Ligon did not have access to any field guides before and during the survey because of frequent misspellings and the incorrect use of bird names in the diary. Besides the diary and report, details of Ligon’s correspondence with the USBS and other naturalists are also included. Descriptions of the letters or the letters themselves offer unique perspectives; in one instance, Ligon is chided for the poorly prepared and improperly labeled specimens he sent to the USBS. The authors also explore the occasionally amateur and naive nature of Ligon’s writings, due to his lack of formal scientific training. These shortcomings were apparently evident to his supervisors, who rejected some of his observations, including an April report of a Scarlet Ibis near the Gila River at close to 7000 feet elevation. The book generally does a good job correcting Ligon’s errors, but it does contain a number of its own. Although most are minor, some are puzzling. Lor example, under the species accounts it suggests that Scissor-tailed Llycatchers reached New Mexico by following forests along the Canadian River (the Scissor-tailed Llycatcher is just now expanding its range to the Canadian River basin), and refers to brood parasitism by cowbirds on Willow Llycatchers as “nest predation.” The authors specify that 20 species of shorebirds nest at Bosque del Apache National Wildlife Refuge, and that a similar number nest at Bitter Lake National Wildlife Refuge (the correct total is about five). The attempts to describe changes in habitat and birdlife in the time since Ligon’s journey are commendable but often fall flat. Ornithologists and serious birders will find many of the book’s reports uninformative (e.g., “I observed perhaps fifty coots... so they are still in the area”; p. 154) or too brief to be useful. This work provides a firsthand account of New Mexico’s first breeding bird survey and contains valuable insights to the development of one of the state’s first modern ornithologists. Those familiar with the region are not likely to gain new information on bird distributions, but the many photographs and writings from over a century ago make it a vivid and entertaining read for anyone interested in the history of ornithol- ogy in New Mexico and the southwestern United States. Cole J. Wolf 160 WFO PRESIDENT S MESSAGE: THE FUTURE OF FIELD ORNITHOLOGY When it comes to deciding which of the many worthy organizations we should support, the number of options seems overwhelming. None of us can afford to belong to all of them, so we try to choose those focused on the areas we care most about. While we hope WFO members appreciate direct benefits such as first option on special field trips and conference registration, we know that most of you belong to WFO because we work to promote field ornithology in a variety of ways. Among those, none is more important than fostering, encouraging, and teaching the next generation of field ornithologists. WFO has a substantial Scholarship Fund (thanks in large part to a generous gift from Pasadena Audubon Society). This fund provides young field ornithologists the opportunity to attend our conferences and to participate in field expeditions led by top ornithologists such as Ken Able, Jon Dunn, and Kimball Garrett. We think this is an ideal way to support their growing interest and to give them models to emulate. In recent years we have provided scholarships to seven very impressive young people: • John Garrett, now majoring in Ecology and Evolutionary Biology at the Uni- versity of California, Santa Cruz, has been making contributions to the field since age 9, when he wrote the “Kid’s Corner” feature for the Pasadena Audubon newsletter. He has served as an eBird editor for Los Angeles County. • Kimberlyann Harvey, an aspiring bird artist and one of our youngest recipients at 13, won a scholarship to attend last year’s conference in Petaluma. She is committed to becoming an ornithologist and recently read Proctor and Lynch’s Manual of Ornithology (a substantial tome) cover to cover. • Joyce Realegeno received a scholarship to attend our Sierra Vista conference. She has studied habitat preferences of corvids around Baldwin Hills, California. • Ioana Seritan also earned a scholarship to our Petaluma conference and was voted an ABA Young Birder of the Year in 2013. • Seventeen year-old Marcel Such will participate in the upcoming field expedition to northeastern California. Marcel is a co-editor of Colorado Birds and founded the Boulder City Teen Naturalists organization. • Ray VanBuskirk began banding birds at age 12 and created and runs a project studying rosy-finch ecology in New Mexico. • Cole Wolf joined the board of the Central New Mexico Audubon Society at 16 and has studied the nesting requirements of the Pinyon Jay. All of these recipients have in common participation in hands-on field research in ornithology, involvement in local ornithology/birding organizations, a commitment to teaching other young people about birds and nature, and an abiding interest in and dedication to conservation. For anyone concerned about the future of field ornithol- ogy, these young people should give cause for optimism. We would like to continue and even expand our scholarship offerings. If this is something you care about, you can help by encouraging young people you know to become WFO members and apply for a scholarship (www.westernfieldornithologists. org/scholarship.php). You can also donate to WFO, requesting that your donation go directly to our Scholarship Fund, or simply give a promising young person a gift membership to WFO. A student membership is only $10! Edward R. Pandolfino Western Birds 44:161, 2013 161 FEATURED PHOTO THE SUBSPECIES OF THE SONG SPARROW ON SOUTHEAST FARALLON ISLAND AND IN CENTRAL CALIFORNIA OSCAR JOHNSON, P. O. Box 21903, Santa Barbara, California 93121; henicorhina@yahoo . com PETER PYLE, The Institute for Bird Populations, P O. Box 1346, Point Reyes Sta- tion, California, 94956; ppyle@birdpop.org JIM TIETZ, PRBO Conservation Science, 3820 Cypress Drive #11, Petaluma, Cali- fornia 94954; jtietz@prbo.org The Song Sparrow ( Melospiza melodia) is one of the most morphologically variable birds of North America. As many as 52 subspecies have been named, 39 of which were recognized by the American Ornithologists’ Union (AOU 1957) and Paynter (1970) from Canada, the United States, Baja California, and central Mexico. In the latest taxonomic revision of the Song Sparrow, Patten and Pruett (2009) recognized 25 subspecies. The subspecies vary from small and pale in the desert Southwest (fallax ) to large and dark in the Aleutian Islands (maxima), with a wide range of intermediates and other variations. Although the Song Sparrow has little or no prealternate molt, the appearance of the basic plumage, especially in subspecies of more open and drier habitats, is affected by wear. The upperparts generally become paler (grayer or browner) and less distinctly streaked from fall to spring, while the underparts become whiter (less buff or brownish) and more distinctly streaked in spring. The subspecies are migratory to various degrees, with some being resident, while others are short- distance or medium-distance migrants (AOU 1957, Arcese et al. 2002). Grinnell and Miller (1944) detailed the distribution of 18 subspecies of the Song Sparrow in California, including 12 characterized as “permanent residents,” four found only in winter, and two that breed in California and show at least some evidence of seasonal movement. Subspecies santaecrucis was synonymized with gouldii in the fifth edition of the AOU checklist (1957) and by Pyle (1997), on the basis of informa- tion from Aldrich (1984) and other sources. Pyle (1997) categorized the remaining 17 subspecies into four subspecies groups, the northwest coastal Pacific ( rufina ) group, the California mainland (gouldii ) group, the California island ( clementae ) group, and the interior western (montana) group. Two other subspecies groups occurring north of Mexico, the Alaska island ( insignis ) and eastern ( melodia ) groups, contain subspecies without records in California at the time. Arcese et al. (2002) and Patten and Pruett (2009) synonymized five additional subspecies recognized in California by Grinnell and Miller (1944), realigned some of the subspecies according to subspecies groups, and recognized just five groups overall. Here we follow the subspecies taxonomy of Patten and Pruett (2009) with the exception of recognizing fisherella ( see below). We also maintain the interior western subspecies group of Pyle (1997), which was lumped with the eastern group by Patten and Pruett. Finding and identifying wintering or migrant subspecies of the Song Sparrow in Cali- fornia can be confounded by the presence and variability of local resident subspecies. Southeast Farallon Island, part of the Farallon National Wildlife Refuge, located 32 km off the coast of Marin County, lacks a resident population of the Song Sparrow and thus provides a unique opportunity for investigation of its migration and identification along the central California coast. PRBO Conservation Science (formerly known as the Point Reyes Bird Observatory) has censused birds and operated a banding station on the island since 1967 (DeSante and Ainley 1980, Richardson et al. 2003), allow- 162 Western Birds 44:162-170, 2013 FEATURED PHOTO ing banders to obtain detailed data on many of the migrant Song Sparrows reaching the island, including measurements and photographs. From 1967 to 2012, 97 Song Sparrows were recorded on the island, with 70 of those arriving from August to November, 25 from February to June, and one each in December and January. Four fall migrants remained through the winter, but none has oversummered. Of these 97 individuals, attempts were made to identify 36 to subspecies, either by biologists in the field or retrospectively by us from photographs and banding data. Twenty-five Song Sparrows were caught and banded, of which nine were photographed in hand, and two additional birds were preserved as specimens. We have used photographs and measurements of Song Sparrows from the island and compared them to series of specimens, published measurements, and morphological descriptions. We have also reviewed all specimens of the Song Sparrow collected in central California and housed at the California Academy of Sciences (CAS) and the Museum of Vertebrate Zoology (MVZ). Here we summarize the subspecies of the Song Sparrow on South- east Farallon Island and in central California on the basis of our best identifications. Subspecies of the northwest coastal Pacific group breed from coastal Alaska to Oregon, migrate to coastal northwestern and central California (Grinnell and Miller 1944, Patten and Pruett 2009), and were reported by DeSante and Ainley (1980) to be the most common subspecies group to reach the island. M. m. morphna breeds from coastal southwestern British Columbia to central western Oregon (Swarth 1923, Patten and Pruett 2009). Grinnell and Miller (1944) characterized morphna as a fairly common winter visitant to California, primarily to the northern half of the state west of the Sierra Nevada, with records as far south as San Bernardino and Los Angeles counties. Patten and Pruett (2009), however, indicated morphna to be less migratory, with only a few individuals moving into northwestern California. It appears that Patten and Pruett (2009) referred most of the birds identified as morphna by Grinnell and Miller (1944) to merrilli, which breeds primarily from mainland southeastern Alaska and interior central British Columbia south to eastern Washington and northern Idaho and occurs in winter as far south as southern California and Arizona (Phillips et al. 1964, Patten et al. 2003). M. m. morphna is characterized by its medium size, rufous upperparts with indis- tinct darker streaks on the back, and heavy diffuse rufous streaking on the breast; its appearance does not vary as much between fall and spring as does that of some other subspecies, perhaps because of its preference for shaded habitats. M. m. merrilli is a variable subspecies and may represent an intergrade swarm among the northwest coastal Pacific (where placed by Patten and Pruett 2009), interior western (where placed by Pyle 1997), and eastern subspecies groups, similar to subspecies Passerella iliaca altiuagans of the Fox Sparrow and Junco hyemalis cismontanus of the Dark- eyed Junco, which breed in this same region of British Columbia (Pyle 1997). Along the western boundary of its range, merrilli appears to grade into morphna (Patten and Pruett 2009), having the upperparts reddish but slightly grayer than in morphna (toward the interior western group; see below) and with the dark streaks more distinct (toward the eastern group; see below). The underpart streaking of merrilli tends to be finer than that of morphna, toward the interior western group. Seasonally, mer- rilli appears to wear from redder in fall to grayer in spring, so it could be more easily confused with morphna when in fresher fall and winter plumage. At Southeast Farallon Island, 31 Song Sparrows were identified at the time of ob- servation as of the northwest coastal Pacific group, and 19 of the 31 were identified as morphna. Our examination, however, indicates that only three records from the island have been satisfactorily documented as morphna, including a specimen collected 11 October 1969 (PRBO 330; Figure 1) and individuals photographed on 21 September 1983 and 20 September 2010 (see the lower photo on this issue’s back cover). Several Song Sparrow records from Southeast Farallon that have been referred to morphna appear instead to represent merrilli. These include a specimen collected 12 October 163 FEATURED PHOTO Figure 1. Specimens of two Song Sparrows collected on Southeast Farallon Island, 11 October 1969, Melospiza melodia morphna (PRBO 330; above) and 12 October 1969, M. m. merrilli or intergrade merrilli x morphna (PRBO 331; below). Note especially the grayer nape and face of PRBO 331. Figure 2. Song Sparrow of the subspecies merrilli banded on Southeast Farallon Island, 29 November 2008. In comparison to morphna, this merrilli has more gray on the face, finer streaks below, and more prominent streaks on the back. Photo by Jim Tietz Photo by Peter Pyle 1969 (PRBO 331; Figure 1) and individuals photographed 21-22 September 1989, 30 March 1990, and 22 October 2008-12 April 2009 (Figure 2). In a review of 107 specimens collected in California and labeled u morphna’ at CAS and MVZ, Pyle found that only 15 individuals appeared to fit pure morphna-, these were collected along the coast south to Santa Cruz County (MVZ 92895) and east to Yolo (CAS 51542) and Alameda (CAS 51569) counties. The remainder appeared Figure 3. Song Sparrow banded on Southeast Farallon Island, 26 September 1980. The dark coloration with reddish tones suggests rufina. Photo by Phil Henderson Figure 4. Song Sparrow at Buena Vista Park, San Francisco County, 10 February 2011. The long bill and dark plumage without reddish tones indicate caurina. Photo by Dominik Mosur 164 FEATURED PHOTO Figure 5. A large and reddish Song Sparrow suggestive of rufina at the mouth of the Little River, Mendocino County, 25 January 2013. Photo by Ron LeValley to be either pure merrilli, intergrades between morphna and merrilli, or in one case, an intergrade between morphna and rufina (see below). A specimen collected at Carmel, Monterey County (CAS 21554), was either morphna or a morphna x merrilli intergrade. Thus, on the basis of records from Southeast Farallon Island and this review of specimens, we agree with Patten and Pruett (2009) that merrilli rather than morphna is the more common migratory subspecies reaching most of California and that the winter range of morphna is restricted primarily to moister coastal habitats Figure 6. Song Sparrow banded on Southeast Farallon Island, 31 March 1991. The gray face, rufous wings and crown, and fine black streaks on a white underside lacking any buff appear closest to gouldii or possibly an intergrade with fisherella. See also upper photo on this issue’s inside back cover. Photo by Peter Pyle Figure 7 . Song Sparrow photographed on Southeast Farallon Island, 11 November 2012. The reddish coloration with distinct blackish and rufous streaking both above and below is intermediate between morphna and gouldii and thus the bird appears to be cleonensis of coastal northwestern California. Photo by Lukas Musher 165 FEATURED PHOTO of northwestern California, south to Santa Cruz and perhaps Monterey counties. We suspect that many of the Song Sparrows on Southeast Farallon identified as morphna may in fact have been merrilii. Pyle (1997) and Patten and Pruett (2009) placed three additional subspecies of the Song Sparrow in the northwest coastal Pacific group. These include kenaiensis, which breeds on the Kenai Peninsula of the south coast of Alaska, caurina, which breeds on the northern coast of the Gulf of Alaska, and rufina, which breeds in coastal British Columbia and southeastern Alaska. Some individuals of all three of these subspecies migrate south along the coast for the winter (Patten and Pruett 2009) and could show up in central California, though Grinnell and Miller (1944) recorded only caurina in California and primarily along the immediate coast in Del Norte and Humboldt counties. A specimen from Alameda County reported by Grinnell and Wythe (1927; Univ. Calif. Los Angeles 12148) as caurina is merrilii, as we confirmed by examin- ing photos of the specimen. The subspecies caurina is very dark, larger and longer-billed than morphna, with colder gray-brown rather than rufous upperparts and darker and browner underparts; it also has a shorter bill than the otherwise similar kenaiensis. In fall, the darker back streaking of caurina is even less distinct than that of morphna, and it practically disappears in spring, at which point the upperparts become duskier. Although similar to caurina, rufina is slightly smaller with an equally long bill and with more rufous tones to the plumage; rufina appears intermediate between caurina and morphna and intergrades with each. On Southeast Farallon Island, DeSante and Ainley (1980) identified two fall migrants as “cf. caurina,'’ one banded on 16 October 1970 and another observed 23-24 October 1972, but no photographs or descriptions were recorded to assess the identifications, and the wing chord of the former (64 mm) is small for caurina (Pyle 1997). We have identified two large and dark Song Sparrows from the island as either caurina or rufina. An individual banded and photographed on 4 October 1986 (see the lower photo on this issue’s inside back cover) was dark dusky brown above, suggestive of caurina, but had some reddish tones below and a short wing chord (63 mm), both suggesting rufina, and is thus probably best identified as caurina/ rufina. Another individual banded and photographed 26-27 September 1980 was similarly dark overall but showed significant reddish tones both above and below, and its small size (wing chord 63 mm) is more suggestive of rufina (Figure 3). A large individual banded on 13 September 1987 was likely caurina on the basis of a combination of the relatively long wing chord (73 mm; see Pyle 1997) and large mass (33.2 g), but no plumage characters were noted. Review of specimens from California at CAS and MVZ revealed only one individual closest to caurina, collected in Humboldt County (MVZ 87445), one individual closest to rufina, collected 14 December 1904 in Alameda County (CAS 60331), and one individual that appeared to be an intergrade rufina x morphna, collected 8 December 1901 in San Benito County (CAS 51580). In addition, we have examined photographs of individuals closest to caurina from San Francisco County (Figure 4), and closest to rufina from Mendocino County (Figure 5). Although Grinnell and Miller (1944) did not record rufina in California, records for coastal Oregon (Marshall et al. 2006) and the evidence we report here suggest it regularly, if rarely, reaches Southeast Farallon and coastal northern California in fall and winter, expected since both caurina and morphna do so. It should be noted, however, that the systematic relationships and morphological variation of the subspecies in this group are particularly poorly known (L. DeCicco pers. comm.), so any identification of these three subspecies out of range should be regarded as provisional. In central California, the subspecies gouldii of the California mainland group breeds in upland habitats along the California coast opposite Southeast Farallon from Lake and coastal southern Mendocino counties through Santa Cruz County. Grinnell and 166 FEATURED PHOTO Miller (1944) considered it a permanent resident, although Patten and Pruett (2009) reported a single vagrant gouldii south of its normal range, on Santa Cruz Island, Santa Barbara County, 31 October 1988 (San Diego Natural History Museum 45418); we agree with this identification after examining photos of the specimen. To the north of gouldii the subspecies cleonensis breeds from coastal southwest Oregon to central Mendocino County. Although Grinnell and Miller (1944) considered it essentially a resident, they noted an individual of cleonensis collected at Olema, Marin County, 17 September 1909 (Grinnell and Wythe 1927; MVZ 10570); we have examined this specimen and agree that it is cleonensis. Another individual at Bolinas, Marin County, November-December 2010 was identified in the field by Pyle as closest to cleonensis. The remaining five resident California subspecies in this group ( heermanni , pusil- lula, maxillaris, samuelis, and graminea) noted by Grinnell and Miller (1944) and recognized by Patten and Pruett (2009) occur as residents in specific bioregions such as San Francisco Bay marshes and in southern California, including on the Channel Islands. We consider these subspecies unlikely to reach Southeast Farallon Island and do not consider them further here. The subspecies gouldii has olive-brown upperparts with distinct dark brown back streaks in fall, becoming grayer on the face and redder on the wings with wear through spring. Its underpart streaking is darker brown and more distinct than in subspecies of the northwest coastal Pacific group. M. m. cleonensis is intermediate between morphna and gouldii in most respects, with more gray on the face and more distinct back streaks than in morphna but with more rufous in the plumage than in gouldii. It was considered part of the California mainland group by Pyle (1997) and part of the northwest coastal Pacific group by Patten and Pruett (2009), but given that it is intermediate between the two groups it could reasonably be placed with either. We have identified as closest to gouldii a Song Sparrow present on Southeast Faral- lon Island 31 March-1 April 1991 (see the upper photo on the inside of this issue’s back cover and Figure 6). Another individual on 17 October 1974 was described at the time of observation as representing a California coastal subspecies (i.e., gouldii), but was not photographed or measured extensively enough to confirm the subspe- cies. A Song Sparrow photographed on the island 11 November 2012 (Figure 7) was small, tinged brownish, and more heavily streaked on the back and underparts than merrilli. We have tentatively identified it as cleonensis, and it is possible that some of the other individuals on the island identified as morphna or gouldii were actually cleonensis. Thus, despite gouldii and cleonensis being considered resident or nearly so by Grinnell and Miller (1944), there does appear to be some propensity for these two subspecies to disperse or migrate. We suggest that gouldii is like some resident coastal species thought unlikely to reach the island by DeSante and Ainley (1980:77), including the American Crow ( Coruus brachyrhynchos ), Bewick’s Wren (Thryomanes bewickii), and Western Bluebird ( Sialia mexicana), that have been recorded since. The interior western group, including merrilli (see above), fisherella, and mon- tana, breeds in drier inland habitats from central British Columbia south to northern California and east through the Great Basin; it can be distinguished from other groups by relatively pale grayish upperparts and muted pale reddish to brownish streaking on the upperparts and underparts. Grinnell and Miller (1944) separated fisherella as occurring in inland northern California south to Tehama and Inyo counties, whereas Patten and Pruett (2009) synonymized it with montana of the Great Basin. On the basis of our examination of specimens we recognize fisherella but suspect its bound- ary with montana might be better defined biogeographically (see Pyle 1997:28) as along the crest of the Cascade Range and Sierra Nevada, rather than to the east of Oregon and California as considered by Grinnell and Miller (1944). M. m. fisherella is darker and browner above and has darker and browner (less reddish) and more distinct upperpart and underpart streaking than montana-, like merrilli, it appears to be heavily influenced by intergradation with surrounding subspecies montana, 167 FEATURED PHOTO merrilli , cleonensis , gouldii, and heermanni, showing characters of each of these subspecies where ranges approach one another (cf. Grinnell and Miller 1944). The pale reddish fallax, a resident of the desert Southwest (Grinnell and Miller 1944, Patten and Pruett 2009), could also be considered part of this group. Both fisherella and montana are migratory and have been recorded in winter as far south in California as Los Angeles and Imperial counties (Grinnell and Miller 1944, Patten et al. 2003). In the eastern San Francisco Bay region, specimens of fisherella have been taken in Sonoma, Napa, and eastern Alameda counties (Grin- nell and Wythe 1927, Grinnell and Miller 1944), and it appears to be regular in winter along the east flank of the coastal range south at least to Monterey County (CAS and MVZ specimens; D. Roberson pers. comm.). In coastal central California, a specimen closest to fisherella was collected at Berkeley, Alameda County (MVZ 124699), and a specimen that appears closest to montana was collected at Santa Cruz, Santa Cruz County (MVZ 95042), although Grinnell and Miller (1944:543) cautioned that such individuals may represent variants of fisherella or intergrades between fisherella and montana. We have identified from Southeast Farallon Island no Song Sparrows as fisherella or montana (but see Figure 6); however, these other records from the San Francisco Bay area and elsewhere suggest a potential for these two subspecies to occur there. The eastern group of the Song Sparrow contains two subspecies: the highly mi- gratory nominate melodia, including several subspecies synonymized by Patten and Pruett (2009), and atlantica, which is largely resident in salt marshes on the middle Atlantic coast and is not expected to reach the West. Nominate melodia breeds across much of eastern North America, west to northeastern British Columbia, and winters across much of the eastern United States east of the Rocky Mountains (AOU 1957, Arcese et al. 2002, Patten and Pruett 2009). Vagrants have occurred in Washington, where a specimen was collected in King County, 27 February 1978 (Paulson 1992), and twice in Arizona (Phillips et al. 1964). M. m. melodia is characterized by small to medium size, a short thick bill, medium brownish dorsal coloration largely lacking gray or reddish tones, heavy dusky dorsal streaking, a broad dark malar stripe, heavy and distinct dark brown ventral streaking, and a buff wash across the chest when fresh. Song Sparrows banded on Southeast Farallon Island on 3 November 1993 (see upper photo on this issue's back cover; wing 70 mm, weight 19.5 g) and 15-16 October 1995 (Figure 8; identified by Richardson et al. 2003 as kenaiensis; wing 70 mm, weight 19.7 g) match nominate melodia. Both individuals had distinct black streaking both dorsally and ventrally that contrasted strikingly with the ground color, much buff in the supercilium and chest, a contrasting dark malar stripe, and a short conical bill; both were on the large end for melodia and so may have come from west- ern populations (i.e., juddi, synonymized with melodia by Patten and Pruett 2009). Moreover, these individuals lacked the gray face of the somewhat similar California mainland group and were darker and more heavily streaked than montana of the interior western group (see above). A third individual on the island, photographed at a distance on 6 November 1992, also appears closest to nominate melodia , but the photograph is not detailed enough for us to be certain. We have also examined photographs of a Song Sparrow at Clatsop County, Oregon, 23 November 2010 (M. Patterson pers. comm., Figure 9) that represents nominate melodia. To our knowl- edge, these individuals represent the first records of the eastern subspecies group for California and Oregon. Given these records, others documented from Washington and Arizona, and the regularity with which many migratory eastern passerines occur on Southeast Farallon Island and the Pacific coast, we suggest that melodia may occur rarely but regularly in California and Oregon in fall and winter. Our analysis indicates that Song Sparrows can reach Southeast Farallon Island from a broad range of areas, representing at least seven subspecies: at least three records of morphna-, three of caurina/rufina, possibly including at least one each of 168 FEATURED PHOTO Figure 8. Although published by Richardson et al. (2003) as kenaiensis, this Song Sparrow banded on Southeast Farallon Island, 15 October 1995 represents nominate melodia, perhaps of the western populations or an intergrade with merrilli or montana. The very contrasting plumage, buff wash on the chest, and short conical bill are typical of melodia. The gray dorsum, however, is unusual for at least the eastern populations of melodia (D. A. Sibley pers. comm.). Photo by Peter Pyle caurina and rufina; at least four of merrilli ; one of cleonensis; one closest to goul- dii; and at least two of melodia. Our records of rufina and melodia from Southeast Farallon and elsewhere in coastal central California represent the first records of these subspecies from the state. Although no individuals of fisherella or montana have been documented from Southeast Farallon, fisherella, at least, appears to be uncommon in fall and winter in the eastern San Francisco Bay area and should be expected to occur on the island. The uncertainty evident in the identification of many Figure 9. Song Sparrow of subspecies M. m. melodia photographed at Warrenton, Clatsop County, Oregon, 23 November 2010. Note the contrasting plumage, thick bill, and broad dark malar stripe typical of melodia. Photo by Mike Patterson 169 FEATURED PHOTO of the birds we discuss underscores the fact that identification of subspecies can be difficult and it may not be realistic to identify all individuals to subspecies with certainty, particularly those outside of their normal range. The unique situation on Southeast Farallon Island has provided insight into the movements of the Song Sparrow, and we hope that this analysis will help others locate and document migratory subspecies in coastal central California. We thank Lucas DeCicco, David Sibley, Don Roberson, Steve N. G. Howell, Steven Mlodinow, Tony Leukering, and Keith Hansen for their comments and assistance in identifying subspecies of the Song Sparrow, and Phil Henderson, Odvin Lund, Lukas Musher, Mike Patterson, Ron LeValley, and Dominik Mosur for use of their photos. We thank the Museum of Vertebrate Zoology (Carla Cicero), the California Academy of Sciences (Maureen Flannery), the Donald R. Dickey Bird and Mammal Collection at the University of California, Los Angeles (Kathy Molina), and the Western Foundation of Vertebrate Zoology (Adam Searcy) for access to their collections. We thank the U.S. Fish and Wildlife Service for permission to work on the Farallon National Wildlife Refuge. Michael A. Patten, Daniel D. Gibson, and Philip Unitt provided very helpful reviews of the manuscript. This is PRBO contribution 1924 and IBP contribution 462. LITERATURE CITED Aldrich, J. W. 1984. Ecogeographical variation in size and proportions of Song Spar- rows ( Melospiza melodia). Ornithol. Monogr. 35. American Ornithologists’ Union. 1957. Checklist of North American Birds, 5th ed. Lord Baltimore Press, Baltimore. Arcese, P., Sogge, M. K., Marr, A. B., and Patten, M. A. 2002. Song Sparrow (Melospiza melodia), in The Birds of North America (A. Poole, F. Gill eds.), no. 704. Birds N. Am., Philadelphia. DeSante, D. F., and Ainley, D. G. 1980. The avifauna of the South Farallon Islands, California. Studies Avian Biol. 4. Grinnell, J., and Miller, A. H. 1944. The distribution of the birds of California. Pac. Coast Avifauna 27. Grinnell, J., and Wythe, M. W. 1927. Directory to the bird-life of the San Francisco Bay region. Pac. Coast Avifauna 18. Marshall, D. B., Hunter, M. G., and Contreras, A. L. 2006. Birds of Oregon: A General Reference. Ore. State Univ. Press, Corvallis. Patten, M. A., and Pruett, C. L. 2009. The Song Sparrow as a ring species: Patterns of geographic variation, a revision of subspecies, and implications for speciation. Syst. Biodiv. 7:33-62. Patten, M. A., McCaskie, G., and Unitt, P. 2003. Birds of the Salton Sea: Status, Biogeography, and Ecology. Univ. of Calif. Press, Berkeley. Paulson, D. R. 1992. Interior Song Sparrow in western Washington. Wash. Birds 2:42-43. Paynter, R. A., Jr. 1970. Subfamily Emberizinae, in Check-list of Birds of the World (R. A. Paynter, Jr., ed.), pp. 3-214. Mus. Comp. Zook, Cambridge, MA. Phillips, A., Marshall, J., and Monson, G. 1964. The Birds of Arizona. Univ. Ariz. Press, Tucson. Pyle, P. 1997. Identification Guide to North American Birds, part I: Columbidae to Ploceidae. Slate Creek Press, Bolinas, CA. Richardson, T. W., Pyle, P, Burnett, R., and Capitolo, P. 2003. The occurrence and seasonal distribution of migratory birds on Southeast Farallon Island, 1968-1999. W. Birds 34:58-96. Swarth, H. S. 1923. The systematic status of some northwestern Song Sparrows. Condor 25:214-222. 170 Wing your way to... OLYMPIA, WASHINGTON 22-25 AUGUST 2013 Mark your calendar now for a joint conference of Western Field Ornithologists and the Washington Ornithological Society, hosted by Black Hills Audubon Society, in Olympia, Washington, 22-25 August 2013. This conference will feature workshops on fall warblers (Jon Dunn and Kimball Garrett), sapsuckers (Steve Shunk), gull identification (Mike Donahue), dragonfly identification (Dennis Paulson), wilderness first aid (Heath Wakelee), and eBird (Brian Sullivan). John Marzluff will give the keynote talk on corvids. Two afternoon science sessions will begin with a plenary talk by Dennis Paulson. And, of course, we will offer a full set of field trips (including pelagics) and both photo and bird-sounds sessions. Conference registration is now live at www.westernfieldornithologists.org/ conference. php. Join us in Olympia! Sketch by Keith Hansen 171 World Wide Web site: WESTERN BIRDS www. westernfieldornithologists . org Quarterly Journal of Western Field Ornithologists President: Edward R. Pandolfino, 1328 49th St., Sacramento, CA 98519; ERPfromCA@aol . com Vice-President: David E. Quady, 39 The Crescent, Berkeley, CA 94708; davequady@att. net Past-President: W. David Shuford, P O. Box 69, Bolinas, CA 94924; dshuf ord@prbo . org Treasurer/Membership Secretary: Robbie Fischer, 1359 Solano Dr., Pacifica, CA 94044; robbie22@pacbell.net Recording Secretary: Liga Auzins, 12842 Safford East, Garden Grove, CA 92840; llauzins@yahoo . com Directors: Kenneth P. Able, Kimball L. Garrett, Daniel D. Gibson, Robert E. Gill, Ed Harper, Osvel Hinojosa-Huerta, Kurt Leuschner, Joyce Meyer, Frances Oliver, Dan Singer, Brian L. Sullivan, Debbie Van Dooremolen Editor: Philip Unitt, San Diego Natural History Museum, P. O. Box 121390, San Diego, CA 92112-1390; birds@sdnhm.org Assistant Editor: Daniel D. Gibson, 3705 Quartz Rd., Ester, AK 99725; avesalaska@gmail . com Associate Editors: Kenneth P. Able, Doug Faulkner, Thomas Gardali, Daniel D. Gibson, Robert E. Gill, Paul E. Lehman, Ronald R. LeValley, Dan Reinking Graphics Manager: Virginia P. Johnson, 4637 Del Mar Ave., San Diego, CA 92107; gingerj 5@juno . com Photo Editor: Peter LaTourrette, 1019 Loma Prieta Ct., Los Altos, CA 94024; petelat 1 @stanford . edu Featured Photo: John Sterling, 26 Palm Ave., Woodland, CA 95695; jsterling@ wavecable . com WFO Website: Joseph Morlan, 1359 Solano Dr., Pacifica, CA 94044; webmaster@ westernfieldornithologists . org Book Reviews: Lauren Harter, 2841 McCulloch Blvd N. #1, Lake Havasu City, AZ 86403; lbharter@gmail.com Membership dues, for individuals and institutions, including subscription to Western Birds: Patron, $1000.00; Life, $600 (payable in four equal annual installments); Supporting, $75 annually; Contributing, $50 annually; Family $40; Regular U.S. $35 for one year, $60 for two years, $85 for three years. Dues and contributions are tax-deductible to the extent allowed by law. Send membership dues, changes of address, correspondence regarding missing issues, and or- ders for back issues and special publications to the Treasurer. Make checks payable to Western Field Ornithologists. BACK ISSUES OF WESTERN BIRDS WITHIN U.S. $40 PER VOLUME, $10 FOR SINGLE ISSUES, INCLUDING SHIPPING AND HANDLING. OUTSIDE THE U.S. $55 PER VOLUME, $15 FOR SINGLE ISSUES, INCLUDING SHIPPING AND HANDLING. The California Bird Records Committee of Western Field Ornithologists revised its 10-column Field List of California Birds in July 2009. The list covers 641 species, plus 6 species on the supplemental list. Please send orders to WFO, c/o Robbie Fischer, Treasurer, 1359 Solano Drive, Pacifica, CA 94044. Price for 9 or fewer, $2.75 each, for 10 or more, $2.50 each, which includes tax and shipping. Order online at http://checklist.westernfieldornithologists.org. Published 15 June 2013 ISSN 0045-3897 Song Sparrow (Afotospisa rnefudkt) of (he coaslal California group, Southeast Farallon Island, San Francisco County, California, 31 March 1 99 ] . The small size, gray ground color of the head, and whilish underpays with black streaks i si spring (worn) basic plumage arc consistent with subspecies gouidti found along the coast of ccntrat California opposite- the Fa ra I Lon Islands. Photo hy Peter Pyte Song Sparrow (Afclvsftiza meiottiu) of llic Alaska group, Southeast Fata I Ion Island, San Francisco County,. California, 4 October 1 9K<3. The large size, long narrow bill, and very dark and sooty coloration arc con&istcnl with the sub- species Gavrina or rufina. breeding in southern and southeastern Alaska, The rusty tone in the streaking of underpins suggests rufina (not yet confirmed in California), but in lack of a specimen cauriria (regular in winter along the roast of northern California) cannot be c-k eluded. Photo by Odvin Land I Vol. 44, No. 3, 2013 Western Specialty: Common Poorwill Photo by ® Soon Logan of Sherman Oaks, California' Common Poorwill (Fhakicn&pfitus nuifaliiij Topanga State Park. Los Angeles County,. California, 23 June 2013 Nole (hat the incoming juvenile plumage of the young is considerably more rufous lhan the plumage of the adult, Volume 44, Number 3, 2013 Seabirds New to the Eastern Chukchi and Beaufort Seas, Alaska: Response to a Changing Climate? Robert H. Day, Adrian E. Gall, Tawna C. Morgan, John R. Rose, Jonathan H. Plissner, Peter M. Sanzenbacher, Jamie D. Fenneman, Katherine J. Kuletz, and Bridget H. Watts 174 Third Report of the Alaska Checklist Committee, 2008-2012 Daniel D. Gibson, Lucas H. DeCicco, Robert E. Gill, Jr., Steven C. Heinl, Aaron J. Lang, Theodore G. Tobish, Jr., and Jack J. Withrow 183 The Importance of Agriculture to Long-billed Curlews in California’s Central Valley in fall W. David Shuford, Gary W. Page, Gary M. Langham, and Catherine M. Hickey 196 The 37 th Annual Report of the California Bird Records Committee: 2011 Records Kristie N. Nelson, Stephen C. Rottenborn, and Scott B. Terrill 206 NOTES Rare and Unusual Birds Observed on Tern Island, French Frigate Shoals, Northwestern Hawaiian Islands Phillip J. Howard, Sarah C. Harvey, and Steven M. Ritt 237 A Winter Record of a Little Bunting from Interior Oregon Alan Contreras and Craig Turner 243 Book Reviews Hector Gomez de Silva and Dave Krueper 246 Featured Photo: Dark-faced Common Murres off Central California in Fall and Winter Peter Pyle 250 Front cover photo by © Steven C. Heinl of Ketchikan, Alaska: Ash- throated Flycatcher (Myiarchus cinerascens), Ketchikan, Alaska, 7 November 2012, representing the first record for Alaska. Back cover: “Featured Photos” by © Alan Wight of Petaluma, California: Common Murres (Uria aalge ) over Cordell Bank off Bodega Bay, California, 7 October 2012, showing variation in the pattern of the head in basic plumage. Note also the white trailing edge to the secondaries on the whiter-headed bird, lacking on the darker-headed bird. Western Birds solicits papers that are both useful to and understandable by amateur field ornithologists and also contribute significantly to scientific literature. The journal welcomes contributions from both professionals and amateurs. Appropriate topics in- clude distribution, migration, status, identification, geographic variation, conservation, behavior, ecology, population dynamics, habitat requirements, the effects of pollution, and techniques for censusing, sound recording, and photographing birds in the field. Papers of general interest will be considered regardless of their geographic origin, but particularly desired are reports of studies done in or bearing on North America west of the 100th meridian, including Alaska and Hawaii, northwestern Mexico, and the northeastern Pacific Ocean. Send manuscripts to Daniel D. Gibson, P. O. Box 155, Ester, AK 99725; avesalaska@ gmail.com. For matters of style consult the Suggestions to Contributors to Western Birds (at www. western fieldornithologists.org/docs/journal_guidelines.doc). WESTERN BIRDS Volume 44, Number 3, 2013 SEABIRDS NEW TO THE EASTERN CHUKCHI AND BEAUFORT SEAS, ALASKA: RESPONSE TO A CHANGING CLIMATE? ROBERT H. DAY, ADRIAN E. GALL, TAWNA C. MORGAN, JOHN R. ROSE, ABR, Inc. — Environmental Research & Services, P. O. Box 80410, Fairbanks, Alaska 99708-0410; bday@abrinc.com JONATHAN H. PLISSNER, PETER M. SANZENBACHER, ABR, Inc.— Environ- mental Research & Services, P. O. Box 249, Forest Grove, Oregon 97116-0249 JAMIE D. FENNEMAN, University of British Columbia, 2329 West Mall, Vancouver, British Columbia, Canada V6T 1Z4 KATHERINE J. KULETZ, U.S. Fish and Wildlife Service, 1011 East Tudor Road, Anchorage, Alaska 99503-6199 BRIDGET H. WATTS, Fairweather Science, 9525 King Street, Anchorage, Alaska 99515 (current address: 247 NW 19th St., Newport, Oregon 97365) ABSTRACT: Seabirds at high latitudes may respond to climate change in a vari- ety of ways, including range contractions or expansions and/or seasonal or annual shifts in distribution. Since 2006, three species of seabirds have been reported in the eastern Chukchi Sea for the first time: the Short-tailed Albatross ( Phoebastria albatrus), Northern Gannet ( Morus bassanus ), and Rhinoceros Auklet ( Cerorhinca monocerata). Sometime prior to 2006, the Ancient Murrelet ( Synthliboramphus antiquus ) expanded its usual maritime range north into the eastern Chukchi and now has reached the Beaufort Sea. The gannet appears to have entered the Pacific via the Northwest Passage, whereas the other three species have moved north from the Pacific. Whales, other seabirds, and diatoms have been recorded moving between the Atlantic and Pacific via the Northwest Passage in the past 15 years as sea ice has retreated and the passage has opened. Because of broad-scale changes to the Chukchi ecosystem and because of increased sampling of the region, we anticipate that additional seabirds will be recorded in the Chukchi and Beaufort seas from the North Pacific and possibly the North Atlantic. The Pacific arctic marine region is undergoing rapid physical and biological change as a result of climate change (Grebmeier et al. 2006a, b, Overland 174 Western Birds 44:174-182, 2013 SEABIRDS NEW TO THE EASTERN CHUKCHI AND BEAUFORT SEAS, ALASKA and Wang 2011, Sigler et al. 2011, Grebmeier 2012). Climate-change models suggest that the increases in air and water temperatures are great- est at high latitudes (IPCC 2007), influencing biological change at multiple trophic levels. For example, the part of the northern Bering Sea south of St. Lawrence Island has changed from an arctic to a subarctic environment in recent years: the cover of sea ice has decreased, blooms of phytoplankton within the ice are diminished, and the abundance and biomass of commu- nities of benthic organisms have been reduced, leading to reduced use by benthic-feeding marine mammals and an increase in populations of pelagic fishes in that region (Grebmeier et al. 2006b). Because of warming and the northward advection of water from the Bering Sea, boreal and subarctic benthic invertebrates also have expanded their distributions northward into the Chukchi Sea, as have the gray whales ( Eschrichtius robustus) that feed on them (Grebmeier 2012). Climate change also is expected to alter distri- butions of species such as ice-pupping seals and marine fishes in the Pacific Arctic (Sigler et al. 2011), and there is evidence that, in the Okhotsk Sea (eastern Russia), the productivity of both planktivorous and piscivorous sea- birds is changing with climate-associated changes in the marine ecosystem (Kitaysky and Golubova 2000). Seabirds, which are highly mobile, may be expected to be prominent bellwethers of change in arctic marine ecosystems. One facet of changes in these ecosystems is range contractions or expansions and/or seasonal or annual shifts in seabirds’ distributions. In this paper, we detail recent records of species new to the marine avifauna of the Chukchi and Beaufort seas, species that may be responding to climate change. Here, we report and interpret these observations with respect to changing environmental conditions in arctic Alaska. RECORDS Short-tailed Albatross A Short-tailed Albatross was observed by Watts and photographed by S. Nelson (Denver, CO) in the northeastern Chukchi Sea at 7 1.3°N, 163.22°W on 6 August 2012. What presumably was the same bird was seen and photo- graphed by C. Pham (U.S. Fish and Wildlife Service [USFWS]) nearly 200 km to the southwest at 70.03° N, 166.99° W on 17 August 2012. It was identifiable as a juvenile by its plumage that was evenly dark gray over the entire body and wings; the large bubblegum-pink beak also is clearly visible in the photographs. This is the first record of this species in the eastern, and probably the entire, Chukchi Sea. Photos documenting both sightings have been deposited with the Alaska Checklist Committee at the University of Alaska Museum. The Short-tailed Albatross is a subarctic-boreal Pacific species known to occur in the Bering Sea since the 19th century. For example, Nelson (1887) “found them common about Bering Straits in summer,” seeing “a number” near Big and Little Diomede islands in 1881. However, he did not see any albatrosses during his subsequent cruise through nearly the entire Chukchi Sea, including visits to Kotzebue Sound and Barrow in Alaska, North Cape 175 SEABIRDS NEW TO THE EASTERN CHUKCHI AND BEAUFORT SEAS, ALASKA (now Cape Schmidt) in Siberia, and Wrangel Island from July to September 1881. Although this species was still abundant at that time, it was not seen north of Bering Strait, suggesting that it somehow was limited in its north- ward distribution in those days., During an extensive cruise throughout the Chukchi Sea in the late 1920s, Jacques (1930) did not record the Short- tailed Albatross; although its numbers were declining at that time, they still were in the thousands (Hasegawa and DeGange 1982). Likewise, surveys in the eastern Chukchi Sea in the 1960s (Swartz 1967) and extensive surveys there in the 1970s and 1980s (Watson and Divoky 1972, Divoky 1987) found none. Finally, in a compilation of 2463 incidental sightings of the Short-tailed Albatross from 1988 to 2004, Piatt et al. (2006) recorded only two even as far north as the Chirikov Basin (between St. Lawrence Island and Bering Strait) but suggested that the Anadyr Current, flowing north, could attract shelf-edge specialists such as this albatross to the region. On the other hand, Portenko (1972) assumed that eight albatrosses observed by M. M. Sleptsov off Cape Serdtse-Kamen (northern Chukchi Peninsula) on 4 September 1939 were Short-tailed, although the species by then was on the verge of extinction (Murie 1959, Hasegawa and DeGange 1982). Therefore, we believe it more probable that these birds instead were Laysan Albatrosses {P. immutabilis), a species that Murie (1959) realized in the 1930s he was mistaking for the Short-tailed in the Aleutian Islands. Northern Gannet Rose and M. Akpik (Wainwright, AK) identified a Northern Gannet at 71.85° N, 161.80° W, 190 km west-northwest of Barrow, on 16 August 2010, but the observation was brief and no photos were taken. The mostly white plumage with a golden wash on the head and black distal parts of the wings indicated that the bird was an adult. This is the first report of this species in the Pacific sector of the Arctic Ocean. On the basis of this obser- vation and one of presumably the same bird the next day off Barrow ( North American Birds 65: 143), the Alaska Checklist Committee added the species to the Alaska unsubstantiated list in 2010 (Gibson et al. 2013). Because the Northwest Passage was not open in the intervening period (see below), we assume that an adult Northern Gannet found 21 months later at Southeast Farallon Island on 25 April 2012 ( North American Birds 66:551, Western Birds 44[lj: cover) was either the same bird or another that had arrived the same season and subsequently reached California waters via Bering Strait, Bering Sea, and the eastern North Pacific. Ancient Murrelet The Ancient Murrelet first was recorded in the Chukchi Sea on 14 Sep- tember 1976, with one bird seen just north of Bering Strait (North Pacific Pelagic Seabird Database, U.S. Geological Survey, Anchorage, AK; see Kessel 1989). Kessel (1989) also mentioned Day’s numerous records in and near Bering Strait in late September 1985 and reported two birds near Cape Lisburne on 26 August 1987. The first record of substantial numbers is by Fenneman, who saw a total of -50-100 birds scattered in small groups of two to six throughout the eastern Chukchi Sea from the last week of 176 SEABIRDS NEW TO THE EASTERN CHUKCHI AND BEAUFORT SEAS, ALASKA Lisburne ^ Point Hope \ 3/1 cd%- Kotzebue 100 Barrow Wamwrignt Point Lay Cape Serdtse-Kamen ■*o CD ", CHUKOTSK PENINSULA Big & Little ■ ■ Diomede ^ Islands %■ Group Size ,e T ofc® Wales SEWARD Chirikov PENINSULA Basin . Nome Cape / (Jlyakhpen Bering Sea St. Lawrence Island Figure 1 . Locations of Ancient Murrelets in the Chukchi and Beaufort seas of Alaska, 1976-2012. Size of circle indicates number recorded during 10-minute transects. September to mid-October 2006. Subsequently, in 4 of the last 6 years, ornithologists with ABR, Inc., and the USFWS recorded hundreds of Ancient Murrelets in the Chukchi (Figure 1, Table 1). Photos documenting one of the sightings have been deposited at the University of Alaska Museum with the Alaska Checklist Committee. In addition to these records from the Chukchi Sea, Plissner and Sanzen- bacher saw two at 71.15° N, 152.59° W, and a group of three at 71.12° N, 152.28° W, ~80 km northwest of Cape Halkett, in the western Beaufort Sea, on 21 September 2010. Also in the Beaufort, A. Bankert (USFWS) recorded a group of four at 71.71° N, 154.66° W, ~80 km northeast of Bar- Table 1 Summary of Records of the Ancient Murrelet in the Chukchi Sea, 2006-2012 Year Total number of birds Range of dates 2006 -50-100 late September to mid-October 2007 68 5 September-6 October 2008 0 2009 0 2010 289 31 August-8 October 2011 110 25 August-24 October 2012 198 10 August-22 October 177 SEABIRDS NEW TO THE EASTERN CHUKCHI AND BEAUFORT SEAS, ALASKA row, on 28 August 2012. In addition, J. C. George (Department of Wildlife Management, North Slope Borough, Barrow, AK, in litt.) saw “a few pairs” of Ancient Murrelets -25 km north of Cooper Island at 71.42° N, 155.60° W on 10 September 2012; neither he nor the Inupiat whalers he was with had ever seen them before. Photos documenting one of the Beaufort sight- ings have also been deposited at the University of Alaska Museum. These observations represent the first records of this species for the Beaufort Sea. The Ancient Murrelet is a boreal Pacific species that has been known from the northern Bering Sea only since 1976; earlier workers in the region did not record it, in spite of many extensive surveys (Nelson 1887, Jacques 1930, Swartz 1967, Portenko 1973, Watson and Divoky 1972, Divoky 1987). Its nearest breeding colonies are located in the Aleutian Islands, -1300 km south of Bering Strait (Gibson and Byrd 2007), except for one record of probable breeding on the Pribilofs (Lehman 2005). In September 1985, substantial numbers occurred near St. Lawrence Island and as far north as Bering Strait, where birds were seen flying north into the southern Chukchi (Kessel 1989, Day 1992). The species was not recorded in the Chukchi Sea again until 2006, however, suggesting that whatever envi- ronmental conditions may have been limiting its northward distribution in the intervening period had disappeared or been ameliorated. The Chukchi Sea was surveyed little during the intervening period, but seabird biologists working in the northern Bering during this period did not record the Ancient Murrelet either. However, land-based observers in the northern Bering Sea, at Gambell, St. Lawrence Island, recorded small numbers each fall from 2001 to 2007, then a substantial increase to over 100 annually from 2008 through 2012 (Lehman 2005, P. E. Lehman in litt.). The Ancient Murrelet has been noted in the Chukchi in 5 of the 7 years from 2006 to 2012, and sometimes in substantial numbers, suggesting the exploitation of new postbreeding habitat and/or a true range expansion. In addition, the recent records from the Beaufort Sea imply a further northward/eastward expan- sion of the range. Currently, the Ancient Murrelet is rare to uncommon in the eastern Chukchi Sea between early August and late October and casual to very rare in the Beaufort Sea from late August to late September. Rhinoceros Auklet Fenneman identified a single Rhinoceros Auklet -30 km west-northwest of Kotzebue in Kotzebue Sound, southeastern Chukchi Sea, in August 2006 (exact date unavailable). He saw the bird at close range for several minutes but was unable to photograph it. On the basis of the dull plumage and the dull bill with no horn evident, it was either a fresh juvenile or a non- or post-breeder. It differed from a Tufted Puffin in that it had a long, sloping forehead and lacked any sign of the puffin’s facial pattern, which would have been evident in a bird at least 1 year old at this time. This is the first report of this species in the Chukchi Sea. The Rhinoceros Auklet is an amphipacific species, breeding on both the eastern and western coasts of the North Pacific Ocean with only a few small colonies along the intervening north coast (AOU 1998, Gibson and Byrd 2007). The records nearest Kotzebue Sound are of one bird recorded at Cape 178 SEABIRDS NEW TO THE EASTERN CHUKCHI AND BEAUFORT SEAS, ALASKA Ulyakhpen (just west of Sireniki, Chukotka, Russia), in the northwestern Ber- ing Sea, on 11 July 1988 (Konyukhov et al. 1998) and one seen at western St. Lawrence Island on 28 August 2008 ( North American Birds 63:140). RESPONSE TO A CHANGING ENVIRONMENT The Northern Gannet presumably was able to move west into the Chukchi Sea as climate change opened the Northwest Passage, greatly reducing the extent of summer sea ice in the Arctic Ocean in recent years. The passage was open enough for unaided travel by ships in 1998, 2007, 2010, and 2012 (Corbyn 2007; http://en.wikipedia.org/wiki/Northwest_Passage, accessed 17 April 2013; http://earthobservatory.nasa.gov/IOTD/view. php?id=78797, accessed 17 April 2013), suggesting that the gannet maybe the vanguard of new species arriving from the Atlantic. During the past 15 years, movements presumed to be through the Northwest Passage also have been recorded for Bowhead Whales ( Balaena mysticetus ) from Alaska and Greenland traveling both directions in 2010 (Heide-Jorgensen et al. 2012); Gray Whales seen in the Mediterranean Sea in 2011 (http://en. wikipedia. org/wiki/Northwest_Passage; accessed 15 March 2013) and in Namibia in 2013 (www.guardian.co.uk/environment/blog/2013/may/14/grey- whale-walvis-bay-namibia#; accessed 10 June 2013); and a North Pacific diatom ( Neodenticula seminae) recorded in the northwestern North Atlantic in 1998 — the first record in the North Atlantic in 800,000 years (Corbyn 2007, Reid et al. 2007). In addition, a juvenile Great Black-backed Gull ( Larus marinus ) photographed at Barrow 8-10 October 2010 presumably came through the Northwest Passage ( North American Birds 65:143), and the proportionately large bill of an adult Atlantic Puffin ( Fratercula arctica) on 20 June 2008 in the Okhotsk Sea, eastern Russia, pointed to subspecies F. a. naumanni , which occurs from Greenland to Svalbard, so Kharitonov (2009) inferred the bird to have arrived in the Russian Far East via the North- west Passage. Hence the gannet we report might not have been the first seabird to transit the Northwest Passage in recent years. Kharitonov (2009) cited records of other Atlantic alcids in the Russian Far East but suggested that they instead reached the Pacific via the Northeast Passage (i.e. , across the ocean north of Asia), which also has been open recently. There are several reasons why we believe that these birds arrived via the Northwest Passage rather than the Northeast Passage (also called the Northeast Sea Route) across northern Russia. First, the straight-line distance along 73° N from central Davis Strait (between Canada and Greenland) to the central Chukchi Sea is only -3800 km, whereas the distance along 73° N from the northern tip of Norway to the central Chukchi Sea is -5900 km — over 50% longer. Second, in Europe, wandering Northern Gannets wander west, not east, from western Russia (Cramp and Simmons 1980); likewise, those in eastern North America wander west and have been recorded across the continent to central Canada and even Victoria Island (AOU 1998), which is not far from the Canadian sector of the Beaufort Sea and abuts the Northwest Passage. Third, if correct, the identification of the Atlantic Puffin to the subspecies from the northernmost Atlantic implies that it would have a much shorter journey to the Chukchi Sea and North Pacific via the Northwest Passage than it would via the 179 SEABIRDS NEW TO THE EASTERN CHUKCHI AND BEAUFORT SEAS, ALASKA Northeast Passage. Finally, in Europe, Great Black-backed Gulls also wander west and south from western Russia (Cramp and Simmons 1983), whereas in eastern North America, nonbreeding Great Black-backed Gulls occasionally summer as far north and west as Baffin Bay, Hudson Bay, and the Canadian Prairie provinces (AOU 1998), all of which are much closer to the Chukchi Sea than is their nearest breeding site in the northwestern corner of Russia. The expansion of the Ancient Murrelet’s range into the Chukchi Sea, and even the Beaufort Sea, appears to represent post-breeding dispersal. In the Aleutian Islands, which are the species’ nearest breeding grounds, the peak of fledging of chicks is the second week of July, with all fledging completed by the end of July (Byrd and Day 1986). The range of records from 10 August to 24 October, most of which are from late August to mid-October, is ~l-3 months after chicks have fledged. Hence it is not yet clear whether these records are of nonbreeding adults, postbreeding adults, and/or juve- niles; many of the sightings at Gambell in late August and early September, at least, have been of adults in alternate plumage (P. E. Lehman in litt.). The Short-tailed Albatross, Ancient Murrelet, and Rhinoceros Auklet all are North Pacific species that reached the Chukchi Sea presumably in response to lessened ice and warmer temperatures. The composition of the seabirds of the Chukchi Sea has changed dramatically in the past 37 years, reflecting the effects of climate change on the community as a whole: formerly composed primarily of piscivorous seabirds and benthic- feeding seaducks, the community has changed to one composed primarily of planktivorous seabirds (Gall et al. , unpubl. data). (Surprisingly in the face of this ecological change, of these three new species from farther south, only the Ancient Murrelet is a planktivore.) These changes to the structure of the seabird community have been accompanied by dramatic changes in the timing of sea ice and the reduction of its extent, and the abundance and/or biomass of zooplankton have increased because of the warmer temperatures (Lane et al. 2008, Matsuno et al. 2011). Because of this broad-scale change to the Chukchi ecosystem and because of increased study at sea, we may expect that additional species of seabirds will be recorded in the Chukchi and Beaufort seas from the North Pacific and possibly the North Atlantic. ACKNOWLEDGMENTS The research by ABR, Inc., was funded by ConocoPhillips Company, Shell Explora- tion and Production Company, and Statoil USA E&P, Inc., as part of the Chukchi Sea Environmental Studies Program (CSESP). Research by the USFWS was funded in part by the North Pacific Research Board (projects 637 and B64) and the Bureau of Ocean Energy Management (Interagency Agreement M10PG00050). The conclusions do not necessarily represent the views of the funding companies or agencies. We particularly thank scientists Caryn Rea (ConocoPhillips), Michael Macrander (Shell), Steinar Eldoy (Statoil), and Sheyna Wisdom (Fairweather Science) for support and feedback during all phases of the CSESP. We also thank personnel of ABR, Inc., for collecting data from 2008 to 2012, Elizabeth Labunski (USFWS) for collecting and processing data, and the many contributors to the North Pacific Pelagic Seabird Database over the past four decades. We also thank J. Craighead George of the North Slope Borough for providing the unpublished photos of Ancient Murrelets off of Barrow. The manuscript was improved greatly by the reviews of D. D. Gibson, S. C. Heinl, and M. L. Kissling. 180 SEABIRDS NEW TO THE EASTERN CHUKCHI AND BEAUFORT SEAS, ALASKA LITERATURE CITED AOU (American Ornithologists’ Union). 1998. Check-list of North American Birds, 7th ed. Am. Ornithol. Union, Washington, DC. Byrd, G. V., and Day, R. H. 1986. The avifauna of Buldir Island, Aleutian Islands, Alaska. Arctic 39:109-118. Corbyn, Z. 2007. Atlantic invaders. Nature Rep.: Climate Change 6:82-84. Cramp, S., and Simmons, K. E. L. (eds.). 1980. The Birds the Western Palearctic, vol. 2: Hawks to bustards. Oxford Univ. Press, Oxford, England. Cramp, S., and Simmons, K. E. L. (eds.). 1983. The Birds of the Western Palearctic, vol. 3: Waders to gulls. Oxford Univ. Press, Oxford, England. Day, R. H. 1992. Seabirds at sea in relation to oceanography. Ph.D. thesis, Univ. of Alaska, Fairbanks. Divoky, G. J. 1987. The distribution and abundance of birds in the eastern Chukchi Sea in late summer and early fall. Report for National Oceanic and Atmospheric Administration, Outer Continental Shelf Environmental Assessment Program (NOAA/OCSEAP), Arctic Environmental Information and Data Center, Anchor- age, AK; available from www.arlis.org. Gibson, D. D., and Byrd, G. V. 2007. Birds of the Aleutian Islands, Alaska. Nuttall Ornithol. Club and Am. Ornithol. Union Series in Ornithol. 1. Gibson, D. D., DeCicco, L. H., Gill, R. E., Jr., Heinl, S. C., Lang, A. J., Tobish, T. G., and Withrow, J. J. 2013. Third report of the Alaska Checklist Committee, 2008-2012. W, Birds 44:183-195. Grebmeier, J. M. 2012. Shifting patterns of life in the Pacific arctic and sub-arctic seas. Annu. Rev. Mar. Sci. 4:63-78. Grebmeier, J. M., Cooper, L. W., Feder, H. M., and Sirenko, B. I. 2006a. Ecosystem dynamics of the Pacific-influenced northern Bering and Chukchi seas in the Amerasian Arctic. Progr. Oceanogr. 71:331-361. Grebmeier, J. M., Overland, J. E., Moore, S. E., Farley, E. V., Carmack, E. C., Cooper, L. W., Frey, K. E., Helle, J. H., McLaughlin, F. A., and McNutt, S. L. 2006b. A major ecosystem shift in the northern Bering Sea. Science 311:1461-1464. Hasegawa, H., and DeGange, A. R. 1982. The Short-tailed Albatross, Diomedea albatrus, its status, distribution, and natural history. Am. Birds 36:806-814. Heide- Jorgensen, M. P., Laidre, K. L., Quakenbush, L. T., and Citta, J. J. 2012. The Northwest Passage opens for Bowhead Whales. Biol. Lett. 8:270-273. IPCC (Intergovernmental Panel on Climate Change). 2007. Climate change 2007: Synthesis report (R. K. Pachauri and A. Reisinger, eds.). IPCC, Geneva, Swit- zerland ; http : //www. ipcc . ch/publications_and_data/ar4/syr/ en/contents . html . Jacques, F. L. 1930. Water birds observed in the Arctic Ocean and the Bering Sea in 1928. Auk 47:353-366. Kessel, B. 1989. Birds of the Seward Peninsula, Alaska. Univ. Alaska Press, Fairbanks. Kharitonov, S. P. 2009. [The first record of the Atlantic Puffin (Fratercula arctica) in the Pacific and routes of alcid vagrancy between the Atlantic and Pacific oceans]. Vestnik Severo-Vostochnogo Nauchnogo Tsentra DVO RAN [Notes from the Northeastern Science Center, Far Eastern Div., Russian Acad. Sci.] 3:105-107. [In Russian with English summary.] Kitaysky, A. S., and Golubova, E. G. 2000. Climate change causes contrasting trends in reproductive performance of planktivorous and piscivorous alcids. J. Anim. Ecol. 69:248-262. Konyukhov, N. B., Bogoslovskaya, L. S., Zvonov, B. M., and van Pelt, T. I. 1998. Seabirds of the Chukotka Peninsula, Russia. Arctic 51:315-329. Lane, P., Llinas, P., Smith, S., and Pilz, D. 2008. Zooplankton distribution in the western Arctic during summer 2002: Hydrographic habitats and implications for food chain dynamics. J. Mar. Syst. 70:97-133. 181 SEABIRDS NEW TO THE EASTERN CHUKCHI AND BEAUFORT SEAS, ALASKA Matsuno, K., Yamaguchi, A., Hirawake, T., and Imai, I. 2011. Year-to-year changes of the mesozooplankton community in the Chukchi Sea during summers of 1991, 1992 and 2007, 2008. Polar Biol. 34:1349-1360. Murie, O. J. 1959. Fauna of the Aleutian Islands and Alaska Peninsula. N. Am. Fauna 61. Nelson, E. W. 1887. Report upon natural history collections made in Alaska between the years 1877 and 1881. U.S. Army Signal Corps, Arctic Ser. 3. Overland, J. E., and Wang, M. 2011. Large-scale atmospheric circulation changes are associated with the recent loss of sea ice. Tellus 62A:l-9. Piatt, J. F., Wetzel, J., Bell, K., DeGange, A. R., Balogh, G. R., Drew, G. S., Geer- naert, T., Ladd, C., and Byrd, G. V. 2006. Predictable hotspots and foraging habitat of the endangered Short-tailed Albatross (Phoebastria albatrus) in the North Pacific: Implications for conservation. Deep-Sea Res. (Part II) 53:387-398. Portenko, L. A. 1972. [Birds of the Chukchi Peninsula and Wrangel Island], parti. Izadtel’stvo Nauka, Leningrad [English translation 1989, Amerind, New Delhi], Portenko, L. A. 1973. [Birds of the Chukchi Peninsula and Wrangel Island], part II. Izadtel’stvo Nauka, Leningrad [English translation 1989, Amerind, New Delhi], Reid, P. C., Johns, D. G., Edwards, M., Starr, M., Poulin, M., and Snoeijs, P. 2007. A biological consequence of reducing arctic ice cover: Arrival of the Pacific diatom Neodenticula semina in the North Atlantic for the first time in 800,000 years. Global Change Biol. 13:1910-1921. Sigler, M. F., Renner, M., Danielson, S. L., Eisner, L. B., Lauth, R. R., Kuletz, K. J., Logerwell, E. A., and Hunt, G. L., Jr. 2011. Fluxes, fins, and feathers: Relation- ships among the Bering, Chukchi, and Beaufort seas in a time of climate change. Oceanography 23:250-265. Swartz, L. G. 1967. Distribution and movements of birds in the Bering and Chukchi seas. Pac. Sci. 21:332-347. Watson, G. E., and Divoky, G. J. 1972. Pelagic bird and mammal observations in the eastern Chukchi Sea, early fall 1970, in WEBSEC-70: An ecological sur- vey in the eastern Chukchi Sea, September-October 1970. U.S. Coast Guard Oceanogr. Rep. 50:111-172. Accepted 18 June 2013 182 THIRD REPORT OF THE ALASKA CHECKLIST COMMITTEE, 2008-2012 DANIEL D. GIBSON, P.O. Box 155, Ester 99725-0155; avesalaska@gmail.com LUCAS H. DeCICCO, 1010 Latouche St., Anchorage 99501; lhdecicco@gmail.com ROBERT E. GILL JR., 3014 Knik Ave., Anchorage 99517; rgill@usgs.gov STEVEN C. HEINL, 2603 4th Ave., Ketchikan 99901; stevencheinl@gmail.com AARON J. LANG, 40208 Alpenglow Circle, Homer 99603; birdingak@gmail.com THEODORE G. TOBISH JR., 2510 Foraker Drive, Anchorage 99517; tgtljo@gmail.com JACK J. WITHROW, University of Alaska Museum, Fairbanks 99775; jj withrow@alaska . edu ABSTRACT: The third report of the Alaska Checklist Committee outlines 15 spe- cies and three subspecies added to, and one species and one subspecies deleted from, the Checklist of Alaska Birds, resulting in a net total of 499 species and 117 additional subspecies currently recognized as occurring or having occurred naturally in Alaska. The Checklist of Alaska Birds is founded on the “Inventory of the species and subspecies of Alaska birds” (Gibson and Kessel 1997), plus additions, deletions, status changes, etc., reflected in the previous Alaska Checklist Committee (AKCLC) reports (Gibson et al. 2003, 2008) and on the informa- tion included in this third committee report. During the period 2008-2012, 14 species and three subspecies were added to the Checklist of Alaska Birds, and one species and one subspecies were deleted. First- and second-round votes on a 15th species (Common Chiffchaff, Phylloscopus collybita), under way at the close of 2012, were completed in March 2013, resulting in a net total of 499 species and 117 additional subspecies currently recognized by the AKCLC as occurring or having occurred naturally in Alaska. Organized in 2000 by founding members Gibson, Heinl, and Tobish, the AKCLC grew during the years 2008-2012: Lang joined in 2009, Withrow in 2010, Gill in 2011, and DeCicco in 2012. As presently constituted, the committee comprises six voting members and a nonvoting secretary. We post a new edition of the Checklist of Alaska Birds early in each new year at the University of Alaska Museum website, where the 19th edition (January 2013) can be found at www.universityofalaskamuseumbirds.org. Because preserved examples of avian species and subspecies (archived voucher specimens) make available manifold data about birds that can only be conjectured from representations (photos, videos, etc.), we include here details of first Alaska specimens of birds already on the Alaska list — speci- mens brought to our attention, re-evaluated, or obtained during this period of coverage. Subspecies listed in parentheses are inferred from photos on the basis of characteristics of plumage, phenology, and/or geographic range. Author, year of publication and type locality are not included for taxa (spe- cies or subspecies) discussed in a previous AKCLC report. Authorities for outlines of nesting range include Vaurie (1959, 1965), American Ornitholo- gists’ Union [AOU] (1998), Dickinson (2003), and Dickinson and Remsen (2013). Maintained separately, the Alaska “unsubstantiated” list comprises Western Birds 44:183-195, 2013 183 THIRD REPORT OF THE ALASKA CHECKLIST COMMITTEE, 2008-2012 species (currently 24) for which we have on file compelling sight reports but no Alaska specimen or readily identifiable photo. ADDITIONS TO THE ALASKA LIST AND FIRST ALASKA SPECIMEN RECORDS Anser albifrons. Greater White-fronted Goose. ADDITIONAL SUBSPECIES: An- ser albifrons sponsa Banks, 2011 {Hooper Bay, Alaska}. Breeds in western and southwestern Alaska. Holotype is U.S. National Museum of Natural History (USNM) 380325, ad. cf , 11 May 1942, Hooper Bay, Alaska, C. E. Gillham. Banks (2011) described this Greater White-fronted Goose as similar to A. a. gambelli of interior and northern Alaska but averaging smaller, and similar to A. a. elgasi of the Cook Inlet area but smaller in all dimensions; subspecies sponsa nests in western Alaska, in the Yukon-Kuskokwim delta area and in the Bristol Bay lowlands; it winters in California and western Mexico. Pterodroma solandri Gould, 1844 {Bass Strait}. Providence Petrel. Breeds on islands in Tasman Sea. Monotypic. First Alaska RECORD (photos): Ten+ birds, 15 September 2011, vicinity of 53° 16'N, 171° 05' E, off Attu Island, Aleutian Islands, B. E. Cooper andG. B. Mackiernan (Cooperand Mackier nan 2012, including photos). NOTES: In its nonbreeding season the Providence Petrel occurs widely in the central Pacific Ocean (Onley and Scofield 2007), north at least as far as main islands of Japan (Ornithological Society of Japan [OSJ] 2012), and it has been reported recently, but not confirmed, off California, Oregon, and Washington (see Cooper and Mackiernan 2012) and off British Columbia (Dunn et al. 2012). Egretta tricolor (Statius Muller, 1776) {Cayenne}. Tricolored Heron. Breeds from southern U.S. to northern South America. Egretta tricolor ( ruficollis ) Gosse, 1847 {Jamaica}, including occidentalis Huey, 1927 {Scammon Lagoon {= Laguna Ojo de Liebre], Baja California Sur} (see Hellmayr and Conover 1948, Erickson and Howell 2001, Unitt 2004). Range of the species except South America. First Alaska RECORD (photo): One adult, 22-24 May 1985, Chickamin River, mainland southeast Alaska, P. Bethel (in litt . , 2008; photo AKCLC). NOTES: Nearest nesting Tricolored Herons are in Baja California (see Erickson and Howell 2001). The species was for- merly (until about 2000) a regular winter visitor along the coast of southern California, but there are only four records for northern California (P. Unitt, in litt., 2013). There have been several Oregon records, one in spring (Marshall et al. 2006). Ardeola bacchus. Chinese Pond-Heron. Breeds from India to China and Mongolia and from southeast Asia to Sulawesi. Monotypic. FIRST ALASKA SPECIMEN: University of Alaska Museum (UAM) 26000, ad. cf in nuptial plumage, 20 May 2009, Attu Island, Aleutians, J. J. Withrow and D. W. Sonneborn (Withrow and Sonneborn 2011, including photo). HISTORY IN ALASKA: One bird in nuptial plumage, 4-9 August 1996, at St. Paul Island, Pribilof Islands (Hoyer and Smith 1997, including photos), provided the first and only previous Alaska record; one present 14-15 July 2011 at Gambell, St. Lawrence Island, provided the only other Alaska record (see N. Am. Birds [NAB] 65:673 — where photos [p. 713] erroneously dated 14-15 June). Nycticorax nycticorax. Black-crowned Night-Heron. Nycticorax nycticorax nyc- ticorax. Breeds in the Old World (Eurasia from Europe to China, Japan, Taiwan, the Sundas, and Philippines; north Africa). FIRST ALASKA SPECIMEN: UAM 15000, ad. cf , 11 May 2002, Attu Island, Aleutians, K. Winker and D. W. Sonneborn (Gibson and Byrd 2007). HISTORY IN ALASKA: Prior Alaska records, all since the 1970s, were outlined by Gibson and Kessel (1992, 1997), who inferred occurrences of nominate nycticorax and of subspecies hoactli (Gmelin, 1789) {Valley of Mexico} (Breeds in the New World). Four additional Aleutian specimens of nominate nycticorax were 184 THIRD REPORT OF THE ALASKA CHECKLIST COMMITTEE, 2008-2012 Figure 1. Eurasian Oystercatcher (Haematopus ostralegus [oscu/ans]), 26 May-13 June 2012, Buldir Island, Aleutians. Photo by R. A. Dugan collected in 2006 (Gibson and Byrd 2007). There is no Alaska specimen of hoactli. Finally, one Black-crowned Night-Heron was reported 1 1-29 July 201 1 at St. George Island, Pribilofs (see NAB 64:673). Circus cyaneus. Northern Harrier. Circus cyaneus cyaneus. Breeds in Europe, central and northern Asia. ADDITIONAL SUBSPECIES (specimen): UAM 9062, juv. <$ , remains salvaged in June 1999, Attu Island, Aleutians, D. W. Sonneborn. Identified as this species by C. J. Dove, at USNM, the specimen comprises a distal right wing; its length (chord 318 mm) points to this subspecies (Gibson and Byrd 2007) rather than to the North American C. c. hudsonius. Gallinula chloropus (Linnaeus, 1758) {England}. Common Moorhen. Breeds from Europe to Japan, in sub-Saharan Africa, in Madagascar, and on many isolated oceanic islands and archipelagos, from, e.g., Mauritius and the Seychelles to the Philippines, Palau, and northern Marianas. Gallinula chloropus chloropus, including indica Blyth, 1842 {Calcutta) (see Vaurie 1965). Europe to Japan. First Alaska RECORD (specimen): UAM 27369, juv. cf , 12-14 October 2010, Shemya Island, Aleutians, M. T. Schwitters (Withrow and Schwitters 2012, including photos). NOTES: In the wake of the AOU’s decision (Chesser et al. 2011) to split the New World Common Gallinule (Gallinula galeata) from the Old World Common Moorhen (G. chloropus), the Shemya specimen provides the first record of the latter species within the political limits of North America. The Common Moorhen has been recorded twice as far east as Kamchatka (October 1974 and October 1981, Artyukhin et al. 2000) and once in the Commander Islands (April 1956, Marakov 1962), which lie 320 km northwest of Shemya. Haematopus ostralegus Linnaeus, 1758 {Oland Island, Baltic Sea}. Eurasian Oystercatcher. Eurasia. Haematopus ostralegus (osculans) Swinhoe, 1871 {Liao- tung Peninsula, Manchuria}. Breeds in northeastern China, Korea, Sea of Okhotsk, and Kamchatka. First Alaska RECORD (photos; Figure 1): One bird, 26 May-13 185 THIRD REPORT OF THE ALASKA CHECKLIST COMMITTEE, 2008-2012 June 2012, Buldir Island, Aleutians, I. L. Jones, M. Goh, R. A. Dugan (photos NAB 66:570). NOTES: Two Eurasian Oystercatchers seen together on 19 June 1883 at Bering Island (Stejneger 1885) seem to provide the only record in the Commander Islands (see Johansen 1961, Artyukhin et al. 2000). Himantopus himantopus. Black-winged Stilt. Himantopus himantopus himan- topus. Breeds in central and southern Eurasia (east to Honshu); also Africa and Madagascar. First Alaska specimen: UAM 18462, ad. 9, 16 May 2003, St. George Island, Pribilofs, K. Sundseth and J. J. Weicker (see NAB 57 :389-390, photo p. 431). HISTORY IN Alaska: There have been two other Alaska records — of one photographed during the period 24 May-3 June 1983 at Nizki Island, Aleutians, C. F. Zeillemaker and others (Zeillemaker et al. 1985), and inferred to be this subspecies (Gibson and Kessel 1997); and two birds together 1 June 2003, one still present on 9 June 2003, at Shemya Island, Aleutians, G. V. Byrd (Gibson and Byrd 2007). Tringa semipalmata (Gmelin, 1789) {New York}. Willet. Breeds in south-central Canada and north-central U.S. and along Atlantic coast of southeastern Canada and U.S. Tringa semipalmata ( inornata ) (Brewster, 1887) {Larimer Co., Colorado}. South-central Canada and north-central U.S. FIRST ALASKA RECORD (photos; Figure 2): One bird, 22-30 June 2012, mouth of Kenai River, Kenai Peninsula, T. A. and L. K. Burke and others (photos AKCLC). The bird’s long and rather slender dark bill are consistent with subspecies inornata (J. L. Dunn, in litt. , 2012) HISTORY IN ALASKA: Transfer from the unsubstantiated list. Hartlaub (1883) inferred from a description of its feet that a “ Totanus ” reported to have been taken at Portage Bay [= Portage Cove, Haines] in 1882 was this species, but he never saw the specimen; Gabrielson and Lincoln (1959:828) included reference to Hartlaub ’s report in their hypothetical list. There were two 20th-century reports: up to two birds 8-9 August 1961 at Minto Lakes, west of Fairbanks, by W. T. Van Velzen (1963); and one with Greater Yellow- legs (T melanoleuca), Dunlins ( Calidris alpina), golden-plovers ( Pluvialis ), and Black Turnstones ( Arenaria melanocephala) on 29 April 1998 at Seward, by W. C. Shuster (in litt., 1998). NOTES: In the Pacific Northwest the Willet is generally rare or uncom- mon in coastal estuaries in Washington (Wahl et al. 2005) and very rare in spring, summer, and fall on the southern coast of British Columbia (Campbell et al. 1990). Gallinago solitaria Hodgson, 1831 {Nepal}. Solitary Snipe. Breeds in southern Siberia and from northeastern Mongolia to northeastern China and Kamchatka. Gallinago solitaria japonica Bonaparte, 1856 {Japan}. Northeastern Mongolia to northeastern China and Kamchatka. FIRST ALASKA RECORD (specimen): UAM 27000, ad. cf, 24 May 2010, Attu Island, Aleutians, J. J. Withrow and D. W. Sonneborn (Withrow and Sonneborn 2011). HISTORY IN ALASKA: Transfer from the unsubstan- tiated list. A sighting on 10 September 2008 at St. Paul Island, Pribilofs, provided the only previous Alaska report (Bieber and Schuette 2009, including photos of marginal quality). NOTES: East of Kamchatka the Solitary Snipe is a rare straggler in the Commander Islands (Johansen 1961), and it has been recorded at Karaginskiy Island (Artyukhin et al. 2000), off the northeast coast of Kamchatka. Leucophaeus atricilla (Linnaeus, 1758) {Bahama Islands}. Laughing Gull. Breeds on Atlantic coast from southeastern Canada to Venezuela, also at Salton Sea and on Pacific coast of Mexico. Monotypic. First Alaska RECORD (photos): One bird (first cycle), 5-7 January 2010, city of Kodiak, Kodiak Island, R. A. Macintosh and oth- ers (photos NAB 64:309). HISTORY IN ALASKA: Transfer from the unsubstantiated list. Early in the 20th century, Reichenow (1909) reported one collected 6 October 1906 on the Kenai River — a specimen that “reached the museum [Berlin] in such bad condition that it was thrown away” (A. Reichenow, in litt., 1915, fide Gabrielson and Lincoln 1959:829). Recent reports included one bird, probably in its second summer, 29 June 2002, Yakutat, P. M. Suchanek and others; one adult in summer plumage, 186 THIRD REPORT OF THE ALASKA CHECKLIST COMMITTEE, 2008-2012 31 August 2004, Icy Strait, S. T. Zimmerman and others; and one adult in summer plumage, 30 May 2009, Auke Bay, G. B. van Vliet. Heinl and Piston (2009) judged the details of a 1976 report from Ketchikan (on file AKCLC; report noted by AOU 1998) insufficient to eliminate Franklin’s Gull (L. pipixcan). NOTES: In the Pacific Northwest the Laughing Gull has been recorded three times, spring to fall, in coastal Washington (Wahl et al. 2005), and there have been three summer reports from the southern coast of British Columbia (see Field Notes 52:493, 1998; J. D. Fenneman, in litt. , 2013). Streptopelia decaocto (Frivaldszky, 1838) {Turkey). Eurasian Collared-Dove. Breeds from Europe to western China; Burma to eastern China. Streptopelia deca- octo decaocto. Europe to western China. FIRST ALASKA RECORDS (specimens; photos AKCLC): Added to the Alaska checklist in 2010 following first Alaska reports in 2006, 2007, and 2008 (Gibson et al. 2008, Heinl and Piston 2009) and invasion in 2009 (NAB 63:641; see also photos at NAB 63:555). Currently a rare or uncommon resi- dent in virtually every town from Metlakatla and Ketchikan north and west to Yakutat and Cordova. Specimens: UAM 27777, cf , 25 September 2010, Ketchikan, J. and J. Raymond; and UAM 30669, cf , 5 November 2012, Ketchikan, L. H. DeCicco and S. C. Heinl. NOTES: Following release in the Bahamas in the 1970s, this species spread quickly to Florida (Smith 1987, Romagosa 2012) and beyond (Cecile 2004, Cole and McCaskie 2004, Leukering and Gibbons 2005, Mlodinow et al. 2006, Hamilton et al. 2007). As noted by Romagosa and McEneaney (1999), collared-doves are kept in captivity and individual doves found in the wild might be the result of local release (see also Glover et al. 2001, Cole and McCaskie 2004). In at least one case in Alaska (three at Petersburg 20 July 2006), the birds were thought to have been local escapes (Gibson et al. 2008). The incursion in 2009, however, clearly represented another step in the continued spread of this species across North America. Myiarchus cinerascens (Lawrence, 1851) {Western Texas}. Ash-throated Fly- catcher. Breeds in western U.S., Baja California, and western Mexico. Monotypic (see Patten et al. 2003). First Alaska RECORD (photos; this issue’s cover): One bird, 7-17 November 2012, Ketchikan, S. C. Heinl (photos AKCLC), A. W. Piston, W. Young. HISTORY in Alaska: Transfer from the unsubstantiated list. One prior Alaska report, of one bird (photos, of marginal quality, AKCLC), 20 July 1999, Auke Bay, G. B. van Vliet and others (see Gibson et al. 2003). NOTES: The Ash-throated Flycatcher nests as far north as south-central Washington but is only a casual visitant in coastal Washington, mid-May to late November (Wahl et al. 2005). It is as well a very rare visitant in coastal southern British Columbia, early summer to late autumn, where there had been 36 records by the late 1990s (Campbell et al. 1997). Vireo solitarius (Wilson, 1810) {Bartram’s Woods, Philadelphia, Pennsylvania}. Blue-headed Vireo. Breeds from northwestern to eastern Canada and in north-central and northeastern U.S., south to Georgia. Vireo solitarius ( solitarius ). Northwestern to eastern Canada and north-central and northeastern U.S. FIRST ALASKA RECORD (six photos; the best of which is reproduced in Figure 3): One bird, 30 September and 2 October 2012, Middleton Island, Gulf of Alaska, S. C. Heinl, L. H. DeCicco (photos AKCLC), G. H. Rosenberg (photos AKCLC), and T. G. Tobish Jr. Peter Pyle wrote (in litt., 2012), “I have no problem endorsing it as an HY {hatch-year] male Blue-headed Vireo. I’m very confident of HY by primary coverts, and even if it was an AHY [after hatch year] I think it is too bright for AHY male Cassin’s [Vireo V. cassini/]. For an HY it is confirmed as Blue-headed.” HISTORY IN ALASKA: One prior Alaska report, a bird identified by song but not seen, 17 June 1999, Yukon River floodplain near Weshrinarin Creek, T. Rinaldi (Handel et al. 2009:546 and C. M. Handel, in litt., 2011). NOTES: The Blue-headed Vireo nests as close to Alaska as extreme southeastern Yukon Territory (Alexander et al. 2003) and northeastern British Columbia (Campbell et al. 1997). 187 THIRD REPORT OF THE ALASKA CHECKLIST COMMITTEE, 2008-2012 VI Figure 2. Willet ( Tringa semipalmata [inornate]), 22-30 June 2012, Kenai River mouth. Photo by L. K. Burke Cistothorus palustris (Wilson, 1810) {Philadelphia, Pennsylvania}. Marsh Wren. Breeds from northeastern, central, and southwestern British Columbia and western Washington to central Canada and from north-central to southeastern U.S. Subspe- cies in Alaska not known (see NOTES). FIRST ALASKA RECORD (photos; Figure 4): One bird, 25 October and 28 October 2009, in extensive 1.5-m reeds on Palmer Hay Flats State Game Refuge, at 61° 30' N, 149° 20' W, D. L. Crowson (photos AKCLC and NAB 64:134), R. A. Winckler, and B. L. Friest. NOTES: Subspecies C. p. plesius Oberholser, 1897 {Fort Wingate, New Mexico) nests as far north as the southern interior of British Columbia (Campbell et al. 1997); C. p. iliacus (Ridgway, 1903) {Wheatland, Knox Co., Indiana} nests as far northwest as Alberta (AOU 1957) and is likely the form now found nesting in northeastern British Columbia (Campbell et al. 1997) and found recently, and possibly nesting, in southeastern Yukon Territory (Alexander et al. 2003). Subspecies browningi Rea, 1986 {Pitt Meadows, British Columbia}, which nests in southwestern British Columbia and west-central Washington (Rea in Phillips 1986), is probably sedentary (P. Unitt, in litt. , 2013). Phylloscopus collybita (Vieillot, 1817) {France}. Common Chiffchaff. Breeds mainly in western and central Eurasia. Phylloscopus collybita ( tristis ) Blyth, 1843 {Calcutta}. Siberia from the Yenisei to Lake Baikal, and east in the Russian Far East as far as the Kolyma River. FIRST ALASKA RECORD (photos): One bird, 6-7 June 2012, Gambell, St. Lawrence Island, P. E. Lehman, K. J. Zimmer, and others (Lehman and Zimmer 2013; K. J. Zimmer, C. Irrigoo, and P. Schoenberger photos AKCLC). The AKCLC split its first vote on this identification, with two members voting to add and the other four voting to relegate the species to the unsubstantiated list (28 January 2013). After receiving an array of influential comments, from L. Svensson, P. Alstrom, P. Kennerley, P. Pyle, and others, we voted unanimously in a second round to add this species to the main list (6 March 2013). We did not accept the identification of an autumn bird reported as possibly this species (Lehman and Zimmer 2013:435) at St. Lawrence Island on 30 September 2011. NOTES: Subspecies P. c. tristis winters 188 THIRD REPORT OF THE ALASKA CHECKLIST COMMITTEE, 2008-2012 Figure 3. Blue-headed Vireo (Vireo solitarius [ solitarius ]), 30 September and 2 October 2012, Middleton Island, Gulf of Alaska. Photo by G. H. Rosenberg chiefly from the Middle East to India (Cramp and Brooks 1992). In the Russian Far East it is reported to occur as far east as western Chukotka (Brazil 2009). It is a casual migrant in Japan (since 1996 — OSJ 2012). Luscinia sibilans (Swinhoe, 1863) {Macao}. Rufous-tailed Robin. Breeds from south-central and southeastern Siberia to northeastern China and in the Russian Far East to Sakhalin, Ussuriland, and Kamchatka. Monotypic. First Alaska RECORDS (specimen and photos): UAM 24600, second-year 9, 4 June 2008, Attu Island, Figure 4. Marsh Wren ( Cistothorus palustris [ssp.]), 25 and 28 October 2009, Palmer Hay Flats. Photo by D. L. Crowson 189 THIRD REPORT OF THE ALASKA CHECKLIST COMMITTEE, 2008-2012 Aleutians, D. W. Sonneborn; and one bird (photos), 8 June 2008, St. Paul Island, Pribilofs, L. H. DeCicco and others (all DeCicco et al. 2009, including photos of second bird). HISTORY IN ALASKA: Transfer from the unsubstantiated list. A sighting on 4 June 2000 at Attu Island provided the first Alaska report of this species (NAB 54:317, including photos of marginal quality; see Gibson et al. 2003, DeCicco et al. 2009), and a sighting of one on 6 and 7 September 2012 at St. Paul Island, Pribilofs, D. Gochfeld, S. Schuette, and others, provided the first fall report (NAB 67[l]:in press; photos AKCLC). NOTES: The Rufous-tailed Robin nests widely in Kamchatka as far north as the Palana and Karaga rivers, near the isthmus (Artyukhin et al. 2000), and occurs as a migrant throughout Japan (OSJ 2012), Turdus iliacus Linnaeus, 1766 {Sweden}. Redwing. Breeds in Iceland, the Faeroe Islands, northern and eastern Europe, and in northern Russia as far east as Lake Baikal, the Lena River, and the lower Kolyma River. Turdus iliacus {iliacus). Range of the species except Iceland and the Faeroes. FIRST ALASKA RECORD (photos; Figure 5): One bird, 15-27 November 2011, Seward, Kenai Peninsula, J. Herbert, C. A. Griswold, and others (Griswold 2012; photos NAB 66:194). NOTES: The Redwing is unknown in the Russian Far East beyond the Kolyma (see Portenko 1973, Kistchinski 1980, Artyukhin et al. 2000) but is a straggler in Japan, where there have been more than 12 records (OSJ 2012). A Redwing in Washington (December 2004-February 2005, NAB 59:316 and 370; Wahl et al. 2005) provides the only prior record in western North America. Oreothlypis ruficapilla. Nashville Warbler. Oreothlypis ruficapilla ridgwayi. Breeds in southwestern Canada and western U.S. First Alaska SPECIMEN: UAM 26264, first-fall 9, 21 September 2009, captured aboard ship southeast of St. Matthew Island, Bering Sea, at 59° 20' N, 169° 00' W, L. H. DeCicco. Subspecies identification by C. Cicero (Museum of Vertebrate Zoology, in litt. , 2010). HISTORY IN ALASKA: Gibson et al. (2008) referred to the four previous Alaska records, as well as two historical reports, all in autumn, 5 September (2004, St. Lawrence Island, Lehman 2005) to 11 November (2005, Ketchikan, see Heinl and Piston 2009). One observed on 19 October 2009 in Juneau by L. H. DeCicco (NAB 64:132, photos AKCLC) — one month after he collected Alaska’s first specimen — provides the only additional record. Zonotrichia leucophrys. White-crowned Sparrow. Zonotrichia leucophrys pugetensis Grinnell, 1928 {Parksville, Vancouver Island}. Breeds from southwestern British Columbia south, west of the Cascade Range, in Washington and Oregon to northwestern California. ADDITIONAL SUBSPECIES (specimen): UAM 30685, ad. 9, 17 November 2012, Ketchikan, S. C. Heinl. HISTORY IN ALASKA: Specimen preceded by first Alaska reports in 2000 (immature at a Juneau feeder 9 December 2000-April 2001, G. B. van Vliet) and 2011 (adult photographed at Ketchikan during the period 5-20 November 2011, J. L. Dunn, L. H. DeCicco, and others; NAB 66:151). Leucosticte arctoa (Pallas, 1811) {Russian Altai}. Asian Rosy-Finch. Breeds in the Altai, Sayan, and Tannu Ola mountains, Mongolia, and in the Russian Far East from the Stanovoi Mountains east to Sakhalin, the Kuriles, and Kamchatka. Leucosticte arctoa (brunneonucha) Brandt, 1842 {Kuril Islands}. Stanovoi Mountains east to Sakhalin, the Kuril Islands, and Kamchatka. Winters to Manchuria, Ussuriland, Korea, Sakhalin, Kuriles, and Japan south to Kyushu (OSJ 2012). First Alaska RECORD (photos): One bird, 30 December 2011, Adak Island, Aleutians, I. J. Helmericks (photo NAB 66:370 and Birding 44[6]:32, 2012; additional photos AKCLC). NOTES: The Asian Rosy-Finch is unknown in the Commander Islands (Artyukhin et al. 2000), where the westernmost Gray-crowned Rosy-Finches occur ( Leucosticte tephrocotis griseonucha, including maxima Brooks, 1915 {Copper Island, Com- mander Islands} — Vaurie 1959). 190 THIRD REPORT OF THE ALASKA CHECKLIST COMMITTEE, 2008-2012 ADDITIONS TO THE LIST OF SPECIES UNSUBSTANTIATED IN ALASKA Oceanodroma monorhis (Swinhoe, 1867) {near Amoy, China}. Swinhoe’s Storm- Petrel. One bird was identified among 250+ Fork-tailed Storm-Petrels (O. furcata), 15 Leach’s Storm-Petrels (O. leucorhoa), and ~10 Northern Fulmars ( Fulmarus glacialis ); 5 August 2003; Gulf of Alaska south of Kodiak Island, at 56° 03' N, 153° 42' W; S. Terrill, L. Terrill, R. Terrill, R. D. Spight, and S. T. Zimmerman. We judged brief video (on file AKCLC) insufficient to substantiate this identification (see Chesser et al. 2010). Morus bassanus (Linnaeus, 1758) {Scotland}. Northern Gannet. An adult was identified off Point Barrow — on two successive days by observers unknown to each other and on separate ships — on 16 August 2010 at 71° 50' N, 161° 48' W by J. R. Rose (see Day et al. 2013) and on 17 August 2010 at 71° 18' N, 156° 47' W by C. Leedy. No photos. NOTES: North of its breeding range this species has occurred regularly in western Greenland (Boertmann 1994) and widely elsewhere in the North Atlantic (Mowbray 2002), but we found no prior report of occurrence in the Pacific sector of the Arctic Ocean. Botaurus stellaris (Linnaeus, 1758) {Sweden}. Eurasian Bittern. One bird was observed briefly in flight on 13 June and 25 June 2012 at Buldir Island, Aleutians, by R. A. Dugan, but the only photos (on file AKCLC), which point to this species, are distant flight shots in fog. Acrocephalus dumetorum Blyth, 1849 {India}. Blyth’s Reed Warbler. One bird was identified on 9 September 2010 at Gambell, St. Lawrence Island, by R. L. Ake, P. E. Lehman, A. J. Lang, and others (Lehman and Ake 2011, including photos). Of the many photos taken, most are of poor quality or flight shots (on file AKCLC). None show clearly the wing formula, and similar species in this difficult genus, such as the Large-billed Reed Warbler (A. orinus ) and Eurasian Reed Warbler (A. scirpaceus), can- not be entirely eliminated. Peter Kennerley (in litt . , 2011; see Kennerley and Pearson 2010) wrote, “To achieve acceptance, in my opinion, you would need crisp, sharp photographs and ideally the bird would need to be trapped, measurements taken, and feathers or blood sampled for comparative DNA evidence. As discussed above, and again in my opinion, I would not accept the record for a first for the USA and North America. It would need to be iron-clad in measurements and DNA, and this would require trapping and examination in the hand.” DELETIONS FROM THE CHECKLIST OF ALASKA BIRDS Bucephala clangula. Common Goldeneye. DELETE ADDITIONAL SUBSPECIES B. c. clangula. A specimen (USNM 237500) collected 27 November 1914 at St. Paul Island, Pribilofs (Hanna 1916), and long accepted as the only Alaska record of this Old World subspecies (see AOU 1957), has been reidentified as an example of New World B. c. americana (R. C. Banks, USNM, on specimen label). Charadrius alexandrinus. Snowy Plover. DELETE SPECIES. In view of the AOU’s decision (Chesser et al. 2011) to split into two species the Old World Kentish Plover (C. alexandrinus ) and New World Snowy Plover (C. nivosus), we re-examined the supporting evidence (photos of one bird, 23-24 May 1991, Nome River mouth, Seward Peninsula; on file AKCLC) for the sole Alaska report (Gibson and Kessel 1992) and unanimously agreed that, although of a “Snowy Plover” sensu lato, the photos do not point conclusively to either species. 191 THIRD REPORT OF THE ALASKA CHECKLIST COMMITTEE, 2008-2012 Figure 5. Redwing ( Turdus iliacus [iliacus]), 15-27 November 2011, Seward. Photo by C. A. Griswold ACKNOWLEDGMENTS We thank Jon L. Dunn, Peter Kennerley, Peter Pyle, and Pavel S. Tomkovich for valuable contributions to our assessments of various records and reports included herein, and we are pleased to acknowledge the review comments of Paul E. Lehman, Jon L. Dunn, and editor Philip Unitt. LITERATURE CITED Alexander, S. A., Doyle, F. I., Eckert, C. D., Griinberg, H., Hughes, N. L., Jensen, M., Johnson, I., Mossop, D. H., Nixon, W. A., and Sinclair, P. H. 2003. Birds of the Yukon Territory. Univ. British Columbia Press, Vancouver. American Ornithologists’ Union. 1957. Check-list of North American Birds, 5th ed. Am. Ornithol Union, Baltimore. American Ornithologists’ Union. 1998. Check-list of North American Birds, 7th ed. Am. Ornithol. Union, Washington, DC. Artyukhin, Yu. B., Gerasimov, Yu. N., and Lobkov, E. G. 2000. [Class Aves — birds, in Catalog of vertebrates of Kamchatka and adjacent waters], (R. S. Moiseev and A. M. Tokranov, eds.), pp. 73-99. Kamchatka Printing, Petropavlovsk- Kamchatskiy, Russia. [In Russian] Banks, R. C. 2011. Taxonomy of Greater White-fronted Geese (Aves: Anatidae). Proc. Biol. Soc. Washington 124:226-233. Bieber, G., and Schuette, S. 2009. First record of Solitary Snipe ( Gallinago solitaria ) for North America on Saint Paul Island, Alaska. N. Am. Birds 63:178-181. Boertmann, D. 1994. Meddelelser om Gronland. An annotated checklist to the birds of Greenland. Bioscience 38. Brazil, M. 2009. Birds of East Asia. Princeton Univ. Press, Princeton, NJ. Campbell, R. W., Dawe, N. K., McTaggart-Cowan, I., Cooper, J. M., Kaiser, G. W., and McNall, M. C. E. 1990. The Birds of British Columbia, vol. 2: Diurnal Birds of Prey through Woodpeckers. Royal British Columbia Mus., Victoria. Campbell, R. W., Dawe, N. K., McTaggart-Cowan, I., Cooper, J. M., Kaiser, G. W., McNall, M. C. E., and Smith, G. E. J. 1997. The Birds of British Columbia, vol. 3: Flycatchers through Vireos. Univ. British Columbia Press, Vancouver. 192 THIRD REPORT OF THE ALASKA CHECKLIST COMMITTEE, 2008-2012 Cecile, R. I. 2004. Range expansion of the Eurasian Collared-Dove in Iowa. Iowa Birdlife 74:70-73. Chesser, R. T., Banks, R. C., Barker, F. K., Cicero, C., Dunn, J. L., Kratter, A. W., Lovette, I. J., Rasmussen, P. C., Remsen, J. V. Jr., Rising, J. D., Stotz, D. F., and Winker, K. 2010. Fifty-first supplement to the American Ornithologists’ Union Check-list of North American Birds. Auk 127:726-744. Chesser, R. T., Banks, R. C., Barker, F. K., Cicero, C., Dunn, J. L., Kratter, A. W., Lovette, I. J., Rasmussen, P. C., Remsen, J. V. Jr., Rising, J. D., Stotz, D. F., and Winker, K. 2011. Fifty-second supplement to the American Ornithologists’ Union Check-list of North American Birds. Auk 128:600-613. Cole, L. W., and McCaskie, G. 2004. Report of the California Bird Records Com- mittee: 2002 records. W. Birds 35:2-31. Cooper, B. E., and Mackiernan, G. B. 2012. First record of Solander’s Petrel (Ptero- droma solandri) for Alaska. N. Am. Birds 65:704-708. Cramp, S., and Brooks, D. J. (eds.). 1992. The Birds of the Western Palearctic, vol. 6: Warblers. Oxford Univ. Press, Oxford, England. Day, R. H., Gall, A. E., Morgan, T. C., Rose, J. R., Plissner, J. H., Sanzenbacher, P. M., Fenneman, J. D., Kuletz, K. J., and Watts, B. H. 2013. Seabirds new to the eastern Chukchi and Beaufort seas, Alaska: Response to a changing climate? W. Birds 44:174-182. DeCicco, L. H., Heinl, S. C., and Sonneborn, D. W. 2009. First North American records of the Rufous-tailed Robin (Luscinia sibilans). W. Birds 40:237-241. Dickinson, E. C. (ed.). 2003. The Howard and Moore Complete Checklist of the Birds of the World, 3rd ed. Princeton Univ. Press, Princeton, NJ. Dickinson, E. C., and Remsen, J. V. Jr. (eds.). 2013. The Howard and Moore Com- plete Checklist of the Birds of the World, 4th ed., vol. 1: Non-passerines. Aves Press, Eastbourne, England. Dunn, J. L., Gibson, D. D., Garrett, K. L., Iliff, M. J., Lockwood, M., Pittaway, R., Sibley, D., and Zimmer, K. J. 2012. Twenty-third report of the ABA Checklist Committee. Birding 44(6): 28-29. Erickson, R. A., and Howell, S. N. G. (eds.). 2001. Birds of the Baja California Peninsula: Status, distribution, and taxonomy. Am. Birding Assoc. Monogr. Field Ornithol. 3. Gabrielson, I. N., and Lincoln, F. C. 1959. Birds of Alaska. Stackpole, Harrisburg, PA. Gibson, D. D., and Byrd, G. V. 2007. Birds of the Aleutian Islands, Alaska. Nuttall Ornithol. Club and Am. Ornithol. Union Series in Ornithol. 1. Gibson, D. D., Heinl, S. C., and Tobish, T. G. Jr. 2003. Report of the Alaska Checklist Committee, 1997-2002. W. Birds 34:122-132. Gibson, D. D., Heinl, S. C., and Tobish, T. G. Jr. 2008. Report of the Alaska Checklist Committee, 2003-2007. W. Birds 39:189-201. Gibson, D. D., and Kessel, B. 1992. Seventy-four new avian taxa documented in Alaska 1976-1991. Condor 94:454-467. Gibson, D. D., and Kessel, B. 1997. Inventory of the species and subspecies of Alaska birds. W. Birds 28:45-95. Glover, S. A., Terrill, S. B., Singer, D. S., and Roberson, D. 2001. The winter season. Middle Pacific Coast region. N. Am. Birds 55:225. Griswold, C. 2012. Alaska’s first Redwing! Winging It 24(2):6-7. Hamilton, R. A., Patten, M. A., and Erickson, R. A. (eds.). 2007. Rare Birds of California. W. Field Ornithol., Camarillo, CA. Handel, C. M., Swanson, S. A., Nigro, D. A., and Matsuoka, S. M. 2009. Estima- tion of avian population sizes and species richness across a boreal landscape in Alaska. Wilson J. Ornithol. 121:528-547. Hanna, G. D. 1916. Record of birds new to the Pribilof Islands including two new to North America. Auk 33:400-403. 193 THIRD REPORT OF THE ALASKA CHECKLIST COMMITTEE, 2008-2012 Hartlaub, G. 1883. Beitrag zur Omithologie von Alaska. J. fur Ornithol. 31:257-286. Heinl, S. C., and Piston, A. W. 2009. Birds of the Ketchikan area, southeast Alaska. W. Birds 40:54-144. Hellmayr, C. E., and Conover, B. 1948. Catalogue of birds of the Americas, part I, no. 2. Zool. Series Field Mus. Nat. Hist. 13 (publ. 615). Hoyer, R. C., and Smith, S. D. 1997. Chinese Pond-Heron in Alaska. N. Am. Birds 51:593-596. Johansen, H. 1961. Revised list of the birds of the Commander Islands. Auk 78:44-56. Kennerley, P., and Pearson, D. 2010. Reed and Bush Warblers. Christopher Helm, London. Kistchinski, A. A. 1980. [Birds of the Koryak Highlands]. Izdatel’stvo Nauka, Moscow. [In Russian] Lehman, P. [E.] 2005. Fall bird migration at Gambell, St. Lawrence Island, Alaska. W. Birds 36:2-55. Lehman, P. E., and Ake, R. L. 2011. Blyth’s Reed Warbler [Acrocephalus dumeto- rum) at Gambell, Alaska: First record for North America. N. Am. Birds 65:4-12. Lehman, P. E., and Zimmer, K. J. 2013. A Siberian Chiffchaff ( Pinylloscopus collybita. tristis) at Gambell, Alaska. N. Am. Birds 66:428-435. Leukering, T., and Gibbons, B. 2005. The changing seasons. N. Am. Birds 59:386-394. Marakov, S. V. 1962. [Rare and new birds in the Commander Islands]. Ornitologiya 5:166-167. [In Russian] Marshall, D. B., Hunter, M. G., and Contreras, A. L. (eds.). 2006. Birds of Oregon: A General Reference. Oregon State Univ. Press, Corvallis. Mlodinow, S., Irons, D., and Tweit, B. 2006. Spring migration. Oregon and Wash- ington region. N. Am. Birds 60:429. Mowbray, T. B. 2002. Northern Gannet ( Morus bassanus), in The Birds of North America (A. Poole and F. Gill, eds.), no. 693. Birds N. Am., Philadelphia. Onley, D., and Scofield, P. 2007. Albatrosses, Petrels & Shearwaters of the World. Princeton Univ. Press, Princeton, NJ. Ornithological Society of Japan. 2012. Check-list of Japanese Birds, 7th ed. Ornithol. Soc. Japan, Tokyo Patten, M. A., McCaskie, G., and Unitt, P. 2003. Birds of the Salton Sea: Status, Biogeography, and Ecology. Univ. Calif. Press, Berkeley. Phillips, A. R. 1986. The Known Birds of North and Middle America, part I: Hirun- dinidae to Mimidae; Certhiidae. A. R. Phillips, Denver. Portenko, L. A. 1973. [Birds of the Chukotski Peninsula and Wrangel Island]. Part II. Izdatel’stvo Nauka, Leningrad. [In Russian] Reichenow, A. 1909. Appendix B. — Birds. In Cruises in the Bering Sea (P. Niedieck; translated by R. A. Ploetz), pp. 240-245. Rowland Ward, London. Romagosa, C. M. 2012. Eurasian Collared-Dove [Streptopelia decaocto), in The Birds of North America Online (A. Poole, ed.), no. 630 (revised). Cornell Lab Ornithol., Ithaca, NY; http://bna. birds. Cornell. edu/bna/species/630. Romagosa, C. M., and McEneaney, T. 1999. Eurasian Collared-Dove in North America and the Caribbean. N. Am. Birds 43:348-353. Smith, P. W. 1987. The Eurasian Collared-Dove arrives in the Americas. Am. Birds 41:1370-1379. Stejneger, L. 1885. Results of ornithological explorations in the Commander Islands and in Kamtschatka. U.S. Natl. Mus. Bull. 29. Unitt, P. 2004. San Diego County bird atlas. San Diego Soc. Nat. Hist. Proc. 39. Van Velzen, W. T. 1963. Sight records of Willets in Alaska. Condor 65:167. Vaurie, C. 1959. The Birds of the Palearctic Fauna. Passeriformes. H. F. & G. Witherby, London. 194 THIRD REPORT OF THE ALASKA CHECKLIST COMMITTEE, 2008-2012 Vaurie, C. 1965. The Birds of the Palearctic Fauna. Non-Passeriformes. H. F. & G. Witherby, London. Wahl, T. R., Tweit, B., and Mlodinow, S. G. (eds.). 2005. Birds of Washington: Status and Distribution. Oregon State Univ. Press, Corvallis. Withrow, J. J., and Schwitters, M. T. 2012. First North American record of the Common Moorhen ( Gallinula chloropus) confirmed by molecular analysis. W. Birds 43:259-265. Withrow, J. J., and Sonneborn, D. W. 2011. Important recent bird records from Attu Island, Alaska. W. Birds 42:115-119. Zeillemaker, C. F., Eltzroth, M. S., and Hamernick, J. F. 1985. First North American record of the Black-winged Stilt. N. Am. Birds 39:241. Accepted 7 July 2013 Sketch by Maksim Dementyev 195 THE IMPORTANCE OF AGRICULTURE TO LONG-BILLED CURLEWS IN CALIFORNIA’S CENTRAL VALLEY IN FALL W. DAVID SHUFORD, GARY W. PAGE, Point Blue Conservation Science, 3820 Cypress Drive #11, Petaluma, California 94954; dshuford@pointblue.org GARY M. LANGHAM, National Audubon Society, 1200 18 th Street, Suite 500, Washington, DC 20036 CATHERINE M. HICKEY, Point Blue Conservation Science, 3820 Cypress Drive #11, Petaluma, California 94954 ABSTRACT: The Long-billed Curlew ( Numenius americanus ) — a large shorebird of continental conservation concern — is a migrant and winter resident in California’s Central Valley. The size of the curlew’s North American breeding population has been estimated recently, but little is known about its abundance and habitat needs at migratory stopovers and wintering areas. Following two broad-scale surveys of the curlew in the central and southern portions of the Central Valley in fall in 2007 and 2008, we coordinated a survey of it throughout the valley in August 2009, recording 20,469 curlews in 195 flocks. On all three surveys, during this otherwise arid season, curlews were found primarily in irrigated alfalfa and irrigated pasture. There was a strong, positive relationship between curlew abundance by subregion of the Central Valley and the subregion’s proportion of the entire valley’s acreage of both alfalfa and irrigated pasture. Identifying the habitat features important to curlews at both fine and landscape scales, documenting the birds’ movements (within and between seasons) in the Central Valley, and monitoring their populations is needed to aid in the conservation of this shorebird at risk. The Long-billed Curlew ( Numenius americanus) is a migrant and winter resident in California’s Central Valley, where it concentrates primarily in agricultural lands. The U.S. Shorebird Conservation Plan categorized the Long-billed Curlew as “highly imperiled” because of population declines, low population size, and threats on the nonbreeding and breeding grounds (Brown et al. 2001). Initial rough estimates of the curlew’s rangewide population size of 20,000 to 55,000 individuals, based on expert opinion (Morrison et al. 2001, Fellows and Jones 2009), have been superseded by newer estimates of about 110,000 to 165,000 individuals breeding in the United States (Stanley and Skagen 2007) and 139,000 to 183,000 in North America as a whole (Jones et al. 2008), extrapolated from a sam- pling of the population on the breeding grounds. Yet there is a paucity of information on the species’ abundance, concentration sites, and habitat use at migratory stopovers and in its winter range, where it spends about nine months of the year. Anecdotal evidence suggests that the vast expanses of pastures, alfalfa fields, and fields of harvested rice of the interior valleys of California are important habitats for the Long-billed Curlew during migration or winter. The first coordinated attempt to quantify its abundance across key valleys in September 2007 estimated at least 30,000 individuals: 65% in the Central Valley, 35% in the Imperial Valley, and <1% each in the Antelope Valley, San Jacinto Valley, and Carrizo Plain (Page et al. 2007). The roughly 19,000 curlews tallied in the Central Valley on that survey and another in September 196 Western Birds 44:196-205, 2013 THE IMPORTANCE OF AGRICULTURE TO LONG-BILLED CURLEWS 2008 were mainly in its central and southern portions; in both cases, cover- age was minimal in the extensive rice country of the Sacramento Valley to the north (Shuford et al. 2009). To better document the status of the Long-billed Curlew in the Central Valley, we surveyed the entire valley on a broad scale in August 2009. Here we report on the species’ abundance, distribution, and habitat-use patterns in 2009 and compare them with the patterns found on less extensive fall surveys in 2007 and 2008. We also recommend future research to inform better management practices and conservation of the Long-billed Curlew in the Central Valley. STUDY AREA AND METHODS Our study area was California’s Central Valley (~640 km long by 64 km wide), dominated by agriculture. We divided this vast region into 121 agri- cultural survey areas of varying size across four subregions: the Sacramento Valley (27 areas), Delta (29), San Joaquin Basin (22), and Tulare Basin (43) (Figure 1). These subregions differ slightly from those used for previous surveys (Shuford et al. 2009; see below). From 7 to 10 August 2009, over 100 volunteers and biologists searched 116 of the 121 areas for curlews. In addition, biologists with government agencies searched the following wetland complexes embedded within the agricultural areas: Sacramento National Wildlife Refuge (NWR), Delevan NWR, Sutter NWR, Colusa NWR, North Central Valley Wildlife Manage- ment Area (WMA; Llano Seco Unit), Upper Butte Basin Wildlife Area (WA; Howard Slough, Little Dry Creek, and Llano Seco units), Oroville WA, and Gray Lodge WA in the Sacramento Valley; Yolo Bypass WA in the Delta re- gion; San Joaquin River NWR, San Luis NWR (Freitas, Kesterson, San Luis, East Bear Creek, and West Bear Creek units), Merced NWR (Merced, Arena Plains, and Snobird units), North Grasslands WA (China Island and Salt Slough units), Los Banos WA, Volta WA, and Grasslands WMA in the San Joaquin Basin; and Mendota WA, Kern NWR, and Pixley NWR in the Tulare Basin. Observers recorded the size and location of each flock and classified the birds’ habitat by crop type or other habitat, average plant height (<10 cm, 10-20 cm, >20 cm), moisture (flood irrigated, damp, dry), and field condition (tilled, growing, dormant, fallow). They also recorded the behavior (feeding, roosting, flying) of all curlews seen. We used the same protocol and survey methods as in 2007 and 2008 (Shuford et al. 2009), enabling us to make comparisons of crop use by curlews among the three years. To evaluate the influence of the availability of key crops on curlew abundance, we obtained county crop data from the National Agricultural Statistics Service (2010). We grouped data on crop acreage and curlew abundance by four subregions of the valley (and the counties they comprise): Sacramento Valley (Glenn, Butte, Colusa, Sutter, Yuba, Placer), Delta (Yolo, Sacramento, Solano, Alameda, Contra Costa, San Joaquin), San Joaquin Basin (Stanislaus, Merced, Madera), and Tulare Basin (Fresno, Kings, Tulare, Kern). For the three of these four subregions used in Shuford et al. (2009), we modified the subregional boundaries of previous survey areas slightly to more closely match county boundaries and crop data available by county. In 197 THE IMPORTANCE OF AGRICULTURE TO LONG-BILLED CURLEWS Figure 1. Patterns of abundance and distribution of the Long-billed Curlew over 121 survey areas in 4 subregions of the Central Valley in August 2009. 198 THE IMPORTANCE OF AGRICULTURE TO LONG-BILLED CURLEWS a few instances where agricultural survey areas included substantial parts of two counties that were in different subregions, we apportioned the number of curlews in each county to the correct subregion. To compare curlew abundance to the availability of key crop types, we calculated the proportion of total curlews recorded in each of the four subre- gions in August 2009 (see Figure 1) and the proportion of the valley’s acreage of both alfalfa and irrigated pasture in these subregions for the crop year of 2009. We then used Spearman rank correlations to assess the significance of these relationships at the subregion and county levels and to evaluate the effect of the size of survey areas and subregions on curlew abundance. RESULTS Abundance and Distribution On the August 2009 survey, we recorded 20,469 curlews in 195 flocks in agricultural areas, and none in embedded wetland areas, of the Central Valley. Of the total, 7% were in the Sacramento Valley, 24% in the Delta, 17% in the San Joaquin Basin, and 53% in the Tulare Basin. We found curlews in 51 of 1 16 areas surveyed, but the birds concentrated in relatively few of these. The 23 areas that each held >250 birds and the 12 of these that held >500 birds collectively accounted for 88% and 68%, respectively, of all curlews recorded. Patterns of curlew concentration varied by subregion. With concentration gauged by the number of areas that held >250 curlews, 68% of the curlews in the Sacramento Valley were in 2 of 27 surveys areas, 94% in the Delta were in 6 of 29 areas, 86% in the San Joaquin Basin were in 3 of 22 areas, and 87% in the Tulare Basin were in 12 of 43 areas (see Figure 1). In addition, the curlews occurred mainly in the western and central portions of the Central Valley (Figure 1), extensively irrigated by flooding. There was no significant relationship between curlew abundance and the size of either individual survey areas (Spearman’s p = 0.09, P = 0.32) or the larger subregions (Spearman’s p = 0.80, P = 0.20). Habitat Use The 20,469 curlews recorded in August 2009 concentrated in just a few habitats: 91.4% were in various agricultural crops, 5.1% were roosting in wastewater, water storage, or agricultural evaporation ponds, and 3.5% were in flight. Of those in agricultural fields, 91.1% were in alfalfa, 4.7% in other hay crops, 3.4% in irrigated pasture, 0.6% in rice fields, and 0.2% in miscellaneous crops (Figure 2). The birds found roosting in various types of ponds likely had been foraging earlier in agricultural fields. Curlews used alfalfa fields with a broad range of crop heights. Alfalfa fields with crop heights of <10 cm held 55% of all curlews, those of 10-20 cm held 38%, and those >20 cm held 7%. Use of the latter height class may have been underestimated, as curlews in tall alfalfa at times can be difficult to see. Curlew abundance by subregion was strongly associated with the propor- tion of the entire valley’s coverage of both alfalfa (Spearman’s p = 1.00, P < 0.001) and irrigated pasture (Spearman’s p = 1.00, P < 0.001) in the subregion (Figure 3). At the county level, there was also a strong relationship 199 THE IMPORTANCE OF AGRICULTURE TO LONG-BILLED CURLEWS Sep ’07 Sep '08 Aug '09 Census Period Figure 2. Percentage of curlews in various agricultural crops on three surveys of the Central Valley in fall, 2007-2009. Data for 2007 and 2008 from Shuford et al. (2009). See Methods for differences in survey coverage by year. Subregion Figure 3. Proportion of the total harvested acres of alfalfa and irrigated pasture in the Central Valley in the 2009 crop year within four subregions of the valley and the proportion of Long-billed Curlews counted in those subregions in August 2009 (see Methods). SV, Sacramento Valley; DE, Delta; SJB, San Joaquin Basin; TB, Tulare Basin. Crop data from NASS (2010). 200 THE IMPORTANCE OF AGRICULTURE TO LONG-BILLED CURLEWS between curlew abundance and the extent of alfalfa (Spearman’s p = 0.74, P < 0.001) in the county; the county-level relationship was not significant for irrigated pasture (Spearman’s p = 0.28, P = 0.25) but it was for the combined coverage of alfalfa and irrigated pasture (Spearman’s p = 0.62, P < 0.01). DISCUSSION Survey Challenges The curlew’s behavior and patterns of habitat use may affect the accuracy of counts. As some curlews equipped with radio transmitters move >30 km within the Central Valley every day (K. Sesser unpubl. data), there is a strong likelihood that we may have either under- or overcounted curlews in particular survey areas as birds moved within or among them. The issue of movement was exacerbated by the overall survey lasting several days. Also, our ability to locate birds in early fall may have been enhanced by the heavy reliance of foraging curlews on two crops — irrigated alfalfa and pastures — during this driest time of the year. Similarly, the tendency of these large birds to gather in big flocks, their conspicuousness in flight, and their loud flight calls all make them easy to detect. On balance, we judge that the count of about 20,500 curlews underestimated the number in the Central Valley in August 2009. Although we asked observers to drive all roads through potential curlew habitat in their survey areas, this was not always possible, and not all available habitat was visible from the roads that were driven. Patterns of Abundance Variation in patterns of curlew abundance at various geographic and time scales appears to reflect primarily the extent of suitable habitat and, in some cases, the extent of its coverage on surveys. The 20,469 curlews recorded in August 2009, when we surveyed 116 of 121 agricultural areas in the entire Central Valley, was only modestly higher than the 19,063 in September 2007 and 18,775 in September 2008, when we surveyed 55 and 41, respectively, of 88 agricultural areas in the central and southern Central Valley (Shuford et al. 2009). The small differences among surveys likely reflect the curlew’s limited use of the Sacramento Valley in August and September. Despite the vastly greater coverage of the Sacramento Valley in August 2009 than in the prior two Septembers, we found only 1398 individuals there in 2009, all in agricultural areas, versus 3204 and 1658, respectively, in 2007 and 2008, when all were in embedded wetlands. It also may be that we covered some of the best areas for curlews in the central and southern Central Valley in 2007 and 2008 such that increased coverage in those regions in 2009 did not change the totals substantially. Possibly, curlew numbers had not reached their seasonal peak in the valley by the time of our August survey, but multi-year coastal surveys show curlews begin returning to central California by late June and the numbers reaching their winter plateau by early August (Shuford et al. 1989). Other notable patterns of abundance are detailed below by subregion. Sacramento Valle];. The limited availability of habitats suitable for the cur- lew in the Sacramento Valley at the time of fall surveys likely explains much 201 THE IMPORTANCE OF AGRICULTURE TO LONG-BILLED CURLEWS of the patterns of abundance observed. The extent of alfalfa and irrigated pasture in the Sacramento Valley is much less than in other subregions of the Central Valley (Figure 3), and the roughly 200,000 ha of cultivated rice fields, a habitat curlews also use in the Sacramento Valley (Elphick and Oring 1998), are not very suitable for them in late summer and early fall. At this time, the rice plants are dense and tall, and fields are drained just prior to harvest, which typically occurs from late August to mid-October. The limited use of rice in early fall is illustrated by our August 2009 survey, when we covered much of the rice-growing region of the Sacramento Valley but found only 116 curlews in rice fields, all in one survey area. Flooded wetlands in the Sacramento Valley also contribute to the extent of habitat for curlews, particularly for roosting. Our August 2009 survey, however, was at the seasonal nadir of available water on managed wetlands there, and we did not record any curlews on the many managed wetlands that were surveyed across this region. By contrast, many wetlands begin to be flooded in September for use by early migrating ducks. Greater availability of flooded wetlands may explain why we found use by roosting curlews in September 2007 and 2008 but not in August 2009. Delta. Large numbers of curlews concentrated in the Delta in southern Yolo and northern Solano counties on all surveys, especially in September. On the basis of survey areas each with >250 curlews as a gauge of concen- tration, 8874 curlews were in 5 areas, 9080 were in 5 areas, and 4538 were in 6 areas in September 2007, September 2008, and August 2009, respectively. Four to five adjacent survey areas in Yolo and Solano counties held 86-100% of these curlews on the three surveys. San Joaquin Basin. Moderate to large numbers of curlews were concen- trated in this region on all surveys. On the basis of survey areas with >250 curlews, 5024 curlews were in 6 areas, 2111 were in 4 areas, and 2978 were in 3 areas in September 2007, September 2008, and August 2009, respectively. The specific areas that held large numbers of curlews were not as consistent from survey to survey as they were in the Delta (Shuford et al. 2009; Figure 1). Tulare Basin. Because coverage of the three broad-scale surveys was inconsistent, it is difficult to evaluate the patterns of curlew concentration in this subregion. When coverage was nearly complete in August 2009, the 12 areas that each held >250 curlews, collectively accounting for 9385 individuals, were widely distributed across the basin. Over all surveys there was some indication of concentration in Kings and Tulare counties (Shuford et al. 2009; Figure 1), but more surveys with adequate coverage are needed to confirm this pattern. Patterns of Habitat Use The curlew’s strong affinity for foraging in irrigated alfalfa and pasture is demonstrated by the strong relationship between its abundance by subregion of the Central Valley and the proportion of the entire valley’s coverage of both alfalfa and irrigated pasture in each subregion (Figure 3). At a finer level, there was also a strong relationship between curlew abundance by county and the extent of alfalfa but not the extent of irrigated pasture. This 202 THE IMPORTANCE OF AGRICULTURE TO LONG-BILLED CURLEWS may be because over 90% of curlews in agricultural fields in August 2009 were in alfalfa. The curlew’s relative use of crops likely varies from year to year. Alfalfa, for example, accounted for 91% of the curlews in agricultural fields in August 2009 versus 48% in September 2007 and 69% in September 2008 (Figure 2). Another crop important to curlews in fall was irrigated pasture, which accounted for 42% of curlews in September 2007 and 21% in September 2008. Similarly, K. Sessser (unpubl. data) found that 10 curlews tagged with radio transmitters made significant use of alfalfa and grasslands (irrigated and native pasture combined) in the Central Valley from June to October. As they do patterns of abundance, seasonal changes in habitat suitability also may explain differences in the curlew’s habitat use across years and seasons. The greater use of alfalfa in August than in September (Figure 2) may reflect a decrease in irrigation of that crop toward the end of its growing season. In the San Joaquin Valley, irrigation of alfalfa decreases substantially by late September because of cessation or reduction of water deliveries to farmers, shifting of irrigation to higher-value crops when water is scarce, and a lessened need to irrigate alfalfa as cooler day and night temperatures reduce plant growth and evapotranspiration rates (H. Calvillo pers. comm.). In some cases farmers cease irrigating alfalfa fields to avoid crop losses. The risk of early rains in October spoiling the harvested alfalfa left in the field to cure before it is baled may not justify irrigating another crop in September (anonymous farmer pers. comm.). Use of irrigated pasture was greater in September than in August, but it is unclear if irrigation of pastures extends on average later in the season than for alfalfa. Future Research There is still much to be learned from additional broad-scale surveys of curlews. It would be valuable to survey the entire Central Valley in winter to see if substantial numbers of curlews winter in Sacramento Valley rice fields and in grasslands of the foothills or smaller interior valleys adjacent to the Central Valley. Such seasonal shifts are suggested by numbers much lower in winter than in fall in the Solano County portion of the Delta (Shuford et al. 2009), fall movements of some radio-tagged curlews from the Delta to other areas of the Central Valley (K. Sesser unpubl. data), the shifting of some curlews from various habitats on the valley floor in fall to native grasslands in foothills on either side of the Central Valley in winter (K. Sesser unpubl. data), and anecdotal evidence of curlew numbers increasing substantially from fall to winter in the Carrizo Plain (Shuford et al. 2009), a large valley in the dry inner Coast Ranges west of the southern San Joaquin Valley. Given the substantial annual variation in winter rainfall, standardized surveys at some sites from fall through winter in both wet and dry years should reveal seasonal changes in curlew use and whether patterns of use vary much with climatic conditions. Because of the difficulty of surveying large regions comprehensively and repeatedly, ensuring the conservation of the Long-billed Curlew will require more focused population monitoring. It will likely need to rely on volunteer- based ground surveys to sample potential habitat, as aerial surveys are 203 THE IMPORTANCE OF AGRICULTURE TO LONG-BILLED CURLEWS unlikely to prove fruitful — curlews often fly well before the approach of a plane (D. Shuford and G. Page pers. obs.). Monitoring of curlews via counts at nighttime roosts, where they often congregate in large numbers (Shuford et al. 2009), may also be worth exploring. Although curlews clearly use flood-irrigated alfalfa fields and pastures, na- tive pastures after the onset of seasonal rains, and rice fields flooded in winter, it would be valuable to understand the finer features of these habitats that favor use by curlews and their preferred prey (e.g., optimal plant densities and heights, depth and period of flooding, soil moisture and penetrability). If conditions favorable for curlews are also ones that enhance crop yields, it might be possible to devise a program of economic incentives for farmers to maintain crops and management practices that benefit them and curlews. Further, it would be instructive to assess the landscape features that affect curlew abundance, such as whether within a circumscribed area the acreage of a favored crop must exceed a threshold before curlews occupy it, or if the proximity of crop types or wetlands enhances their value to curlews out of proportion to the combined acreages of these habitats. ACKNOWLEDGMENTS We greatly appreciate the help of the following people who surveyed curlews in the Central Valley, without which this work would not have been possible: Chad Aakre, JoEllen Arnold, Pat Bacchetti, Bob Barnes, Eisso Mansvelt Beck, Melanie Bernal, Ken Biehle, Mel Bolinder, Robin Boyle, Laurie Brignolo, Kathy Brown, Nancy Bruce, Steve Brueggemann, Ed Burns, Eric Caine, Devin Calhoun, Mimi Calter, Walt Carnahan, Mike Carpenter, Giacomo Catalina, Rich Cimino, Bill Clark, Jan Clark, Karen Colbelli, Terry Colborn, Barbara Coley, Roger Coley, Chris Conard, Dan Cooper, Karen Cor- belli, Steve Cordes, Mike Curry, Jeff Davis, Sharon DeCray, Al DeMartini, Joe Devine, Ryan DiGaudio, Rob Doster, Jim Dunn, Phil Eager, Madeline Elsea, Sid England, Binny Fischer, Chuck Fischer, Tim Fitzer, Joe Frank, Scott Frazer, David Froba, John Fulton, Dorothy Furseth, Nancy Gamble, David Garza, Frank Gibson, Dan Gillman, Steve Glover, Sarah Goins, Matt Gould, Jihadda Govan, Linda Greene, Rich Greene, Ken Griggs, Jessica Groves, Kevin Guse, Brian Hansen, Rob Hansen, Fred Hanson, John Harris, Carmen Hashagen, Ken Hashagen, Phil Henderson, Brandon Hill, Paul Hofmann, Rob Holbrook, Shannon Holbrook, Tom Horner, Steve Huckabone, Susan Hult, Jennifer Isola, Jaleh Janatpour, Andrea Jones, Joanne Katanic, Tim Keldsen, Amy Kelsey, Rodd Kelsey, Evan King, Rhiannon Klingonsmith, Dan Kopp, Karl Kraft, Tony Kurz, Dean Kwasny, Jeri Langham, Sami La Rocca, Jim Laughlin, Meg Laws, Steve Laymon, Roderick Lee, Kelly Lesher, Leonard Liu, Carol Lombardi, Steve Lombardi, Stephen Long, John Luther, Bill Lydecker, Jay McEntee, Walt Mclnnis, Len McKenzie, Todd McNicholas, Jane Manning, Chad Martin, Rick Melnicoe, Rick Meredith, Susanne Methvin, Joe Mikelmis, Shawn Milar, Jeff Miller, Bill Moffat, Tucker Moffat, Ted Murphy, Bob Nay, Melissa Odell, Joseph O’Neil, Elizabeth Palmer, Ed Pandolfino, Juan Parra, Chuck Peck, Ed Penny, Dave Quady, Kris Randal, Megan Renshaw, Harold Reeve, Sharon Reeve, Orien Richmond, Jeanne Ridgley, Terres Ronneberg, Jim Ross, Sean Rowe, Tim Ruckle, Vance Russell, Jennifer Rycenga, Nancy Sage, Sal Salerno, Mike Sanders, Rusty Scalf, Matt Schaap, Mary Scheidt, John Schick, Jeff Seay, Deb See, Kristin Sesser, Alison Sheehey, Allison Shultz, Joe Silveira, Zachary Smith, Jim Snowden, Lara Sparks, Daniel Strait, Steven Summers, Simon Thornhill, John Thornhill, Andy Tomaselli, Don Turkal, Yvonne Turkal, Lisa Twiford, Laura Valoppi, Dave VanBaren, Karen Velas, Gail Wake lee, Heath Wakelee, Bobby Walsh, Nils Warnock, Bruce Webb, Jordan Wellwood, Liz West, Ed Whisler, 204 THE IMPORTANCE OF AGRICULTURE TO LONG-BILLED CURLEWS Denise Wight, Kerry Wilcox, Pam Williams, John Wilson, Mike Wolder, Gary Woods, Dennis Woolington, David Yee, Alicia Young, and Lowell Young. We thank Karen Velas for help in coordinating volunteers. We are very grateful to Kristin Sesser for sharing data on patterns of distribution and habitat use of transmitter-tagged curlews in the Central Valley and to Hugo Calvillo, Michelle Gilbert, Elizabeth Palmer, and Dan Putnam for providing insight on factors influencing the timing of cessation of irrigation of alfalfa in the Central Valley in fall. Many thanks to Julie Howar for producing the map of curlew abundance and to Jennifer Roth for conducting the statistical tests. The manuscript was improved by thoughtful reviews by two anonymous reviewers and helpful comments and edits by Dan Reinking and Philip Unitt. Funding was provided by the S. D. Bechtel, Jr. Foundation and the David and Lucile Packard Foundation. This is contribution 1885 of Point Blue Conservation Science. LITERATURE CITED Brown, S., Hickey, C., Harrington, B., and Gill, R., eds. 2001. The United States shorebird conservation plan, 2nd ed. Manomet Center for Conservation Science, P.O. Box 1770, Manomet, MA 02345; www.shorebirdplan.org/ plan-and-council . Elphick, C. S., and Oring, L. W. 1998. Winter management of Californian rice fields for waterbirds. J. Appl. Ecol. 35:95-108. Fellows, S. D., and Jones, S. L. 2009. Status assessment and conservation action plan for the Long-billed Curlew ( Numenius americanus). Biol. Tech. Publ. FWS/ BTP-R60 12-2009. U.S. Fish and Wildl. Serv., Washington, DC. Jones, S. L., Nations, C. S., Fellows S. D., and McDonald, L. L. 2008. Breeding abundance and distribution of Long-billed Curlews ( Numenius americanus) in North America. Waterbirds 31:1-14. Morrison, R. I. G., Gill, R. E., Jr., Harrington, B. A., Skagen, S. K., Page, G. W., Gratto-Trevor, C. L., and Haig, S. M. 2001. Estimates of shorebird popula- tions in North America. Canadian Wildl. Serv. Occas. Paper 104. Environment Canada, Ottawa. National Agricultural Statistics Service (NASS). 2010. County agricultural commis- sioners’ data 2009. NASS California Field Office, P.O. Box 1258, Sacramento, C A 95812; www. nass . usda . gov/Statistics_by_State/ California/Publications/ AgComm/Detail/index . asp . Page, G. W., Shuford, W. D., Langham, G. M., and Molina, K. C. 2007. Results of the 13-16 September 2007 survey for Long-billed Curlews in interior valleys of California. A report of PRBO Conservation Science, Audubon California, and the Natural History Museum of Los Angeles County. Available from PRBO Conservation Science, 3820 Cypress Drive #11, Petaluma, CA 94954. Shuford, W. D., Langham, G. M., Page, G. W., and Hickey, C. 2009. Distribution, abundance, and habitat use of Long-billed Curlews in California’s Central Val- ley from broad-scale surveys in 2007 and 2008. Central Valley Bird Club Bull. 12:29-44. Shuford, W. D., Page, G. W., Evens, J. G., and Stenzel, L. E. 1989. Seasonal abun- dance of waterbirds at Point Reyes: A coastal California perspective. W. Birds 20:137-265. Stanley, T. R., and Skagen, S. K. 2007. Estimating the breeding population of Long- billed Curlew in the United States. J. Wildl. Mgmt. 71:2556-2564. Accepted 23 April 2013 205 THE 37 th ANNUAL REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 2011 RECORDS KRISTIE N. NELSON, P.O. Box 402, Lee Vining, California 93541; knelson@pointblue . org STEPHEN C. ROTTENBORN, H. T. Harvey and Associates, 983 University Ave., Bldg. D, Los Gatos, California 95032; srottenborn@harveyecology.com SCOTT B. TERRILL, H. T. Harvey and Associates, 983 University Ave., Bldg. D, Los Gatos, California 95032; sterrill@harveyecology.com ABSTRACT: The California Bird Records Committee reached decisions on 254 records representing 309 individuals of 75 species and three species pairs documented since the 36 th report (Johnson et al. 2012), endorsing 210 records of 264 individuals. The recent split of the Xantus’s Murrelet into the Scripps’s Murrelet ( Synthliboram - phus scrippsi ) and Guadalupe Murrelet (S. hypoleucus), combined with first accepted state records of the Taiga/Tundra Bean-Goose (Artser fabalis/serrirostris), Common Crane {Grus grus), Common Ringed Plover (Charadrius hiaticula), and Common Snipe ( Gallinago gallinago ) outlined in this report, brings California’s total list of ac- cepted species to 649 as of the end of 2 0 1 1 , 1 0 of which are established introductions . Other notable records detailed in this report include those of the western Atlantic subspecies of the Common Eider (Somateria mollissima dresseri ), White-chinned Petrel ( Procellaria aequinoctialis), and Red-flanked Bluetail ( Tarsiger cyanurus). This 37 th report of the California Bird Records Committee (CBRC or the committee), a formal standing committee of Western Field Ornitholo- gists, summarizes decisions on 254 records of 75 species and three species pairs involving 309 individuals. The committee accepted 210 of the 254 records, involving 264 individuals of 72 species and three species pairs, for an acceptance rate of 83%. We consider 20 records to represent returning or continuing birds that were accepted previously. Forty-two reports of 28 species were not accepted because the identification was not adequately documented, and two reports of two species were not accepted because natural occurrence was questionable. Reports of multiple individuals together are given the same record number for purposes of review; we report the total number of accepted individuals, which may be greater than the number of accepted records. Although the majority of the records we summarize pertain to birds documented in 2011, the period covered spans the years 1971 through 2012. As of the end of 2011, the California list stood at 649 species. In 2012, the committee accepted first California records of the Northern Gannet (. Morus bassanus), Common Cuckoo (Cucu/us canorus), and Gray Hawk ( Buteo plagiatus); the details of these records will be published in the next report. These records, combined with the recent split of the Xantus’s Mur- relet into the Scripps’s Murrelet ( Synthliboramphus scrippsi ) and Guadalupe Murrelet (S. hypoleucus) (Chesser et al. 2012), bring the total number of accepted species on California’s state list as of August 2013 to 652. At its 2013 annual meeting, the committee removed the Yellow-crowned Night- Heron (Nyctanassa uiolacea), Harris’s Hawk ( Parabuteo unicinctus ), Lesser Black-backed Gull ( Larus fuscus), and Parakeet Auklet ( Aethia psittacula) from its review list. 206 Western Birds 44:206-236, 2013 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS Species accounts are organized with English and scientific names first, followed in parentheses by the total number of individuals accepted for California and the number of new individuals accepted in this report. Fol- lowing the heading are accounts for records accepted (as applicable), fol- lowed by records not accepted (as applicable) because identification was not established or because natural occurrence was questionable. An asterisk (*) preceding the species’ name indicates that the CBRC discontinued review- ing records of the species after 2011. A double asterisk (**) following the number of accepted records indicates that the species has been reviewed for a restricted interval, so the number of accepted records does not represent the total number of records for the state. Date ranges for each record are those accepted by the CBRC, and instances where these differ from those published in North American Birds are indicated with italics. A dagger (t) following an observer’s initials indicates submission of a photograph, (S) indicates submission of a sketch, (§) indicates submission of audio record- ings, (t) indicates submission of a video, and (#) precedes a specimen num- ber. The absence of a symbol following the observer’s initials indicates the submission of a sight record based on written description only. Additional details regarding minutiae of formatting and abbreviations may be found in previous CBRC reports, at http://www.californiabirds.org/, and in CBRC (2007). Also available at the website is the California bird list, the review list, committee news, recent photos of rare birds in California, the CBRC’s bylaws, a form for querying the CBRC database, and all annual reports from 1996 through 2009. Age terminology follows that used by CBRC (2007). Observers are encouraged to submit documentation for all species on the CBRC’s review list, sending it to Guy McCaskie, CBRC secretary, R O. Box 275, Imperial Beach, CA 91933-0275 (e-mail: secretary@califor- niabirds.org). In recent years, the proportion of records supported only by photographs, without any written documentation, has increased consider- ably. Even minimal written details on a bird’s appearance, accompanying photographs, can assist significantly in documenting records of rare birds. Documentation of all CBRC records is archived at the Western Foundation of Vertebrate Zoology, 439 Calle San Pablo, Camarillo, CA 93012, and is available for public review. SPECIES ACCOUNTS TAIGA/TUNDRA BEAN-GOOSE Anser fabalis/serrirostris (1, 1). California’s first bean-goose was at Unit 1 of the Salton Sea National Wildlife Refuge (N.W.R.), IMP, 9 Nov 2010-12 Jan 2011 (AKaf; BoMt, DLt, CAMt, GMcC, VMt, LSt, CTf, EGKt, JSmt, JLD, KZKf, TABlf, JMt, TMEt, BKSt, MSat; 2010-141; Figure 1). The bird associated closely with one adult and two immature Greater White-fronted Geese (A. albifrons) among much larger numbers of Snow (Chen caerulescens ) and Ross’s (C. rossii) geese foraging in agricultural fields. Three color photos were pub- lished in N. Am. Birds (one on the cover of 65[1] and two on 65:199). Few records that the committee has accepted have engendered so much controversy with respect to identification, owing to the intricacies and uncertainties of intraspecific variability and taxonomy. In 2007, the American Ornithologists’ Union North American Clas- sification Committee split the Bean Goose into two species, the Taiga Bean-Goose and 207 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS Tundra Bean-Goose, on the basis of differences in color, size, proportions, behavior, and evidence that the two taxa were largely allopatric in their breeding and winter ranges (Banks et al. 2007). Bean-geese breed at high latitudes in eastern Europe and Asia, and as implied by their English names, the Tundra Bean-Goose generally breeds on tundra farther north, the Taiga Bean-Goose in taiga farther south. So defined, both species occur as rare visitors to western Alaska. In North America outside Alaska, the Taiga Bean-Goose has been recorded on four occasions — along the Iowa/Nebraska border in 1984-1985, in Quebec in 1987, in Nebraska in 1998, and in Washington in 2002; the Tundra Bean-Goose has been reported outside Alaska only twice, in Quebec in 1983 and the Yukon in 2000 (Banks et al. 2007). Authors addressing the identification of the Taiga and Tundra bean-geese include Kurechi et al. (1983), Oates (1997), Mlodinow (2004), and Brown (2010). The main features distinguishing them are the head and bill shape (slightly concave culmen with shallowly sloping forehead on the Taiga vs. steeper forehead and convex or straight culmen on the Tundra), the shape of the “grinning patch” along the tomium (longer, thinner, and less conspicuous on the Taiga vs. shorter, deeper, and more pronounced on the Tundra), the ratio of culmen to head length (>1 on Taiga vs. <1 on Tundra), the ratio of culmen length to bill depth at the base (>2 on Taiga vs. <2 on Tundra), and call notes (deeper and shorter with a nasal “ang” quality in the Taiga vs. longer, higher, and less nasal in the Tundra). Although the average Taiga Bean-Goose has more yellow-orange on the bill than does the Tundra, there is overlap in this character. Identification criteria are clouded by intraspecific variability. Within each species, body, neck, and bill size seem to vary clinally, decreasing from east to west, and the difference in size between some western females and eastern males is considerable. Variation in the bean-goose complex has been categorized by as many as seven sub- species, but Sangster and Oreel (1996) classified the birds as two monotypic species encompassing this clinal variation. On the basis of molecular analyses, Ruokonen et al. (2008) suggested a realigned split between A. middendorffii of the eastern taiga and A. fabalis of the tundra and western taiga. They cautioned, “in the absence of diagnostic morphological traits, morphological measurements or genetical analyses are needed for reliable identification of the taxa especially in wintering areas.” Initially, CBRC members agreed that the Salton Sea bird was likely a Taiga Bean- Goose but were divided over whether the Tundra Bean-Goose could be definitively eliminated. Despite excellent photos, the features of the bird in question did not lend themselves to an easy identification. Various observers reported that the “grinning patch” on the tomium seemed to be a better fit for a Tundra and that in some postures the bill and head shape suggested the Tundra, with a more acute angle between the forehead and bill and a short-looking bill. The bird’s foraging behavior was also more typical of a Tundra than a Taiga bean-goose. Its close association with other geese (especially the Greater White-fronted) likely influenced its choice of foraging areas more than any innate habitat preferences. Vocalizations, which would have assisted considerably with the identification, were apparently not heard or recorded. The committee solicited opinions from authorities in Europe and Asia, but none conveyed strong opinions that the bird was a Tundra Bean-Goose. Rather, some considered it likely, others definitively, a Taiga Bean-Goose: Mariko Parslow (in litt.) and Masayuki Kurechi (in litt.) concluded, for example, that multiple measurements (photos) ruled out the Tundra Bean-Goose. Kurechi considered the birds most similar to the Taiga Bean-Geese breeding around the lower Kolyma River and wintering (at least in part) in Shimane Prefecture, Japan, a connection established via satellite telemetry. These birds breed farther west than and are smaller than those that have been classified as subspecies middendorffii. Kurechi considered the Salton Sea bird to be about the same size as or slightly smaller than representatives of the lower Kolyma population and opined that it therefore likely originated even farther west (on the basis of the dine of size decreasing from east to west). Such populations have not been well 208 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS Figure 1 . The size and proportions of the head and bill of California’s first bean-goose, shown here with an adult Greater White-fronted Goose {Anser albifrons) at the Salton Sea National Wildlife Refuge, Imperial Co., 16 Nov 2010 (2010-141), suggested that it was likely a Taiga Bean-Goose (A. fabalis). However, its characters did not neatly fit any known population of either that species or the Tundra Bean-Goose (A. serrirostris), and the committee ultimately accepted it as a Taiga/Tundra Bean-Goose. Photo by Kenneth Z. Kurland Figure 2. The bright rufous coloration and features of the bill processes of this female Common Eider ( Somateria mollissima) at Crescent City, Del Norte Co., 20-29 Nov 2011 (2011-182), indicate the subspecies dresseri, which is typically restricted to northeastern North America and thus was unexpected in California. Photo by Larry Sansone 209 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS studied, either morphologically or genetically. Parslow noted that some large male Tundra Bean-Geese (rossicus) and specimens of Scandinavian and eastern European origin resemble the Salton Sea bird, being small overall but resembling middendorf- fii in bill and head shape. Nial Moores (in litt.) commented that middendorffii is a shorter-distance migrant than Taiga Bean-Geese breeding farther west so western birds might be better candidates for vagrancy to California. Although European experts (e.g., Dan Brown, in litt.) largely agreed that the bird was a better match for the Taiga than for the Tundra, some commented that the short-necked appearance of the Salton Sea bird was quite unlike the long-necked, sometimes swanlike look typical of the Taiga Bean-Goose. Ultimately, six CBRC members voted to accept the record as pertaining to a Taiga/ Tundra Bean-Goose only, leaving the specific identity unconfirmed. Brown (2010) and Luokonen et al. (2008) concluded that some bean-geese may be impossible to identify without measurements or genetic analysis. All members agreed that the bird's natural occurrence was not an issue. The plum- age was not abnormally worn and neither hallux was missing. There are apparently very few bean-geese in captivity, possibly none in North America (Mlodinow 2004). “TRUMPETER SWAN Cygnus buccinator (147, 57). One juvenile was with Tun- dra Swans (C. columbianus ) at Lake of Pines, NEV, 20-26 Dec 2008 (JML; DRf; 2009-069). Five adults were at Modoc N.W.R., MOD, 26 Nov-1 Dec 2009 (SCRf; 2011-029; two of these lingered until 23 Jan 2010 (SCRf; 2011-030). Four first- winter birds were near Nelson, BUT, 11-12 Jan 2011 (BEWf; 2010-157); as many as 11 had been reported at this location 24 Nov 2010-11 Jan 2011 (N. Am. Birds 65:156 & 335), but the CBRC received documentation for these four. Two adults were on Waltz Road near Sheridan, PLA, 19-24 Jan 2011 (DRf; 2011-173). Two adults were northeast of Tipton, TUL, 8-10 Dec 2011 (DFat; SSumf, MESf; 2011- 271). Fourteen (10 adults and 4 first-winter birds) at Modoc N.W.R., MOD, 28-30 Dec 2010 (SCRf; 2011-031) and up to 24 (14 adults and 10 first-winter birds) there 28-30 Dec 2011 (SCRf; 2011-240) likely included returning birds accepted from the area of Alturas, MOD, in previous years. Clearly, the Trumpeter Swan now winters regularly in eastern Modoc County. The group of 24 on 30 Dec represented the highest concentration ever recorded in California and included a leucistic first-winter bird that was completely white like an adult but had extensive pink in the bill and bright yellow feet. Such leucistic individuals occur with some regularity, particularly in Wyoming and Montana (McEneaney 2005). A family group of two adults and three first-winter birds was near Durham, BUT, 31 Dec 2011-9 Jan 2012 (JLf; 2011- 231). The committee has discontinued reviewing records of this species after 2011. IDENTIFICATION NOT ESTABLISHED: Most members thought the identification of an adult and a first-winter bird reported from American Valley, PLU, 18 Dec 2010 (2011-178) may have been correct, but the description was not thorough enough to eliminate the possibility that the birds were Tundra Swans (C. columbianus). NATU- RAL OCCURRENCE QUESTIONABLE: The CBRC considered one at Auburn, PLA, 5 Mar 2011 (RPt; 2011-172) the same as one there 6-9 Apr 2009 (2009-122) and one of two there 17 Feb-25 Mar 2010 (2010-014; Johnson et al. 2012). Although both birds may have escaped or been released by a nearby waterfowl breeder (Johnson et al. 2012), they have been recorded only seasonally at this location. FALCATED DUCK Anas falcata (3, 1). An adult male at Colusa N.W.R., COL, 8 Dec 2011-10 Feb 2012 (GFt, DT; DBrf, DWNf, RLBf, JMf, CAMf, GMcC, VMt, LSt, TABef, KRf^, SLS, DWAf, J D; 2011-205) thrilled observers from all over North America, providing California with its first record since 2003. Photos showed possible tears or holes in the webbing, and a bent toe, on the left foot. Although such injuries could have been suffered in the wild, they prompted some to speculate that the bird may have been held in captivity. The vast majority of North American 210 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS records of Old World ducks such as the Falcated Duck, Baikal Teal (A. formosa), Smew (. Mergellus albellus), and Common Pochard {Aythya ferina ) away from Alaska are of adult males, presumably the most desirable to waterfowl fanciers (some can be readily purchased over the Internet). But adult males are also the most readily identifiable, and in the absence of bands, clipped halluces, or excessive cage wear, in recent years the CBRC has typically accepted records of flight-capable individuals such as this one. In addition, the date span of this record fits a naturally occurring wintering bird, and this record was accepted unanimously on the first round. After an expected absence through the breeding season, the bird returned to Colusa N.W.R. on 2 Dec 2012. KING EIDER Somateria spectabilis (40, 2). Single females were near Marshall on Tomales Bay, MRN, 22-29 Jan 2011 (LHf, DWt; 2011-019) and off the Pt. Reyes lighthouse, MRN, 5-7 Jun 2011 (DSSt; 2011-082), making 2011 the first year since 1993 with multiple records. The vast majority of California records are from fall to early spring; the bird off Pt. Reyes was only the sixth King Eider recorded in June. COMMON EIDER Somateria mollissima (2, 1). An adult female was at Crescent City, DN, 20-29 Nov 2011 (ADBt; RLBt, BPt, LSI, KPA; 2011-182), the location of California’s only prior record, of an adult male 5-18 Jul 2004 (2004-101; Cole et al. 2006). Whereas the adult male was of the race that occurs closest to California (u-nigrum of Alaska), the female had bright rufous plumage and rounded, not pointed, tips to the long bill processes (Figure 2), indicating some other subspecies. Information compiled by Ken Able led to the conclusion that the bird was most likely of the race dresseri, representing a first record for the Pacific Ocean. The breeding and winter ranges of dresseri are in northeastern North America, so the identification led to evaluation of the bird’s provenance during a second round of voting. Abie’s research revealed that several hundred examples of dresseri are held in captivity, including at least two private collections in Washington state. The Crescent City female showed no abnormal wear or other signs of captivity, and there are single records of dresseri from Illinois and Wisconsin and two from Colorado (AOU 1957). The committee considered natural vagrancy likely, and the record was accepted unanimously on the second round. SMEW Mergellus albellus (3, 0). IDENTIFICATION NOT ESTABLISHED: A report of a male on the Smith River, DN, 5 km upstream from its mouth 18 Nov 2011 (2011-186) garnered little support, as the description did not eliminate other, more likely species such as the Bufflehead ( Bucephala albeola). YELLOW-BILLED LOON Gavia adamsii (91 , 3). A second-year bird was upstream from Parker, Arizona, on the Colorado River, SBE, 15 Jan-4 Aug 2011 (LHat, DVPf; AEK, SRf, JPSf, JWet; 2011-008). Alternate-plumaged adults were at Bat- tery Godfrey, Presidio National Park, San Francisco, SF, 14 Oct 2011 (HC, MEa, PS; 2011-171) and on salt pond A4 in Sunnyvale, SCL, 25 Oct-2 Nov 2011, found dead (specimen not preserved) on 4 Nov 2011 (MJMamf, LSI, DEQ; 2011-155), the latter being the first recorded for Santa Clara County. IDENTIFICATION NOT ESTABLISHED: An alternate-plumaged adult reported from Pt. Pinos, MTY, 21 Jun 2008 (2009-080) may well have been correctly identified, but the documentation lacked detail sufficient for some members to accept an unseasonal summer record. A northbound migrant reported from Pt. Piedras Blancas, SLO, 10 Apr 2011 (2011- 087) and two migrants from that location 1 May 2011 (2011-088) were likewise not documented adequately. SHORT-TAILED ALBATROSS Phoebastria albatrus (37**, 5). A second- or third-year bird was seen 65 km west-southwest of Pigeon Pt., SCZ, 28 Mar 2011 (GSMf; 2011-265) during a National Oceanic and Atmospheric Administration re- search cruise. Single all-dark or mostly dark birds, likely in their first year, were seen 84 km southwest of Ano Nuevo Pt., SM, 21 Apr 2011 (RMe; 2011-061), 10 km 211 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS Figure 3. California’s third Great-winged Petrel ( Pterodroma macroptera ), off Santa Cruz Co., 26 Aug 2011 (2011-122), exhibited the extensive pale feathering on the face and forehead, stout bill, and broad, dark underwings that collectively distinguish this species from other dark gadfly petrels. Photo by Greg Lavaty Figure 4. This White-chinned Petrel (Procellaria aequinoctialis ) at Cordell Bank, Marin Co., 16 Oct 2011 (2011-149), shown here behind a Northern Fulmar (Fulmarus glacialis), was the third recorded in California. Photo by Thomas A. Blackman 212 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS Figure 5. California’s first Common Crane (Grus grus ) was at Lake Earl, Del Norte Co., 5-8 May 2011 (2011-065). Although the species is held in (and has occasionally escaped from) captivity in North America, seven committee members accepted that this bird had occurred naturally. Photo by Larry Sansone off Fort Bragg, MEN, 15 May 2011 (JWh; RFof, KAH, RHut, RJKf; 2011-073), 3 km west of Bodega Canyon, SON, 19 Nov 2011 (DSSf; 2011-183), and from Southeast Farallon I., SF, 6 Nov 2011 (OJ S; JRT; 2011-174). As the population of the Short- tailed Albatross continues to increase, this species is now regular in Alaska waters and rare, but increasingly frequent, off British Columbia, Washington, Oregon, and California (Howell 2012). GREAT-WINGED PETREL Pterodroma macroptera (4, 1). One 16 km southwest of Table Rock, SCZ, 26 Aug 2011 (ABof, DLS; RFf, JKf, GLt; 2011-122; Figure 3) showed more pale feathering on the crown than one seen off Santa Cruz, SCZ, 18 Sep 2010 (2010-104; Johnson et al. 2012), suggesting that these were different individuals. All four of California’s records through 2011 are from the period 21 Jul-18 Sep, and all have been of the pale-faced subspecies gouldii. HAWAIIAN PETREL Pterodroma sandwichertsis (14, 1). One was 11 km south- southwest of Southeast Farallon I., SF, 13 Aug 2011 (MD S, DW; DKuf, JMyt; 213 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS 2011-117). While the sketch depicted a bird with a dark “cowl” more typical of the Galapagos Petrel (P phaeopygia ), photos showed the dark hood was limited, with a pale notch at the lower rear portion of the auriculars, as characteristic of the Hawaiian (Force et al. 2007, Pyle et al. 2011b). Given the challenges of assessing such features on the ocean, photos of any Hawaiian/Galapagos Petrel in California are desirable. GALAPAGOS/HAWAIIAN PETREL Pte rod rom a phaeopygia/sandwichensis (26, 2). Single birds 73 km southwest of Pt. Piedras Blancas, SLO, 10 May 2011 (MH; 2011-081) and 145 km south of San Clemente I., LA, 16 Aug 2011 (JSF, PEL; 2011-121) were accepted as Galapagos/Hawaiian Petrels. The views (and thus the descriptions) of these birds were not adequate to allow identification to species. It is evident that the Hawaiian Petrel is a regular component of California’s offshore avifauna, whereas the Galapagos Petrel ranges north only to 20° N (Howell 2012). Nevertheless, the CBRC accepts as a Galapagos/Hawaiian Petrel any record of a bird of this pair that cannot be definitively identified as one species or the other. WHITE-CHINNED PETREL Procellaria aequinoctialis (3, 2). Single individuals were 11 km north of San Miguel I., SBA, 6 Sep 2011 (TMcGt; WTH, DKef, ASet, TWt; 2011-126; photo published in N. Am. Birds 66:200) and at Cordell Bank, MRN, 16 Oct 2011 (TMcGt; TABlf, GTf; 2011-149; Figure 4). Although both birds were in similar stages of wing molt, careful study of photos revealed that the molt of the October bird was slightly behind that of the September bird, confirming that they were different individuals. California had only one previous record, off Half Moon Bay, SM, 18 Oct 2009 (2009-194; Pyle et al. 2011a). With three records within three years, we wonder whether California can expect a continued increase in the White-chinned Petrel or if the records represent a pulse, like that of the Shy Albatross ( Thalassarche cauta). GREAT SHEARWATER Puffinus gravis (9, 1). One in a large flock of Sooty Shearwaters {P. griseus) off Half Moon Bay, SM, 30 Jul 2011 (JSFf; TMcGt, SBT; 2011-110; photo published in N. Am. Birds 65:573 and 682) provided a first record for San Mateo County. MAGNIFICENT FRIG ATEBIRD Fregata magnificens (12**, 0). IDENTIFICATION NOT ESTABLISHED: One at Newport Beach, ORA, 10 Sep 2011 (CW; DMcH, JM, SW; 2011-130) was described as having an all-black head and white underparts, consistent with a female Magnificent Frigatebird. However, the majority of the com- mittee believed that the documentation was not thorough enough, especially given the brevity of the sighting, to eliminate the possibility of a female Great Frigatebird ( F. minor). An immature at Pt. Loma, SD, 30 Oct 2011 (PWf; 2011-161) was also not accepted beyond “frigatebird sp. ” As both the Great Frigatebird and Lesser Frigatebird (. F. ariel ) have occurred in California, the committeee requires that any record of the formerly frequent Magnificent clearly eliminate the two other Pacific frigatebirds; it is considering whether or not to formally accept such birds as “frigatebird sp.” MASKED BOOBY Sula dactylatra (16, 1). A subadult was at Dana Pt., ORA, 17-24 Dec 2011 (RBMcN; CTf, VMf, BJSf, CAM, OJt; 2011-217). MASKED/NAZCA BOOBY Sula dactylatra/granti (11, 1). One in its first spring at Pt. La Jolla, La Jolla, SD, 23 Apr 2011 (SW; RR S; 2011-059) was described as having a pale gray bill, suggesting the Masked, but it lacked a complete pale collar, suggesting the Nazca. The committee agreed with the observers that this individual was best not identified to species. Identification criteria for juveniles and first-year birds were proposed by Pitman and Jehl (1998), but many of these birds may not be identifiable to species in the field (CBRC 2007). *BROWN BOOBY Sula leucogaster (125, 3). In a detailed analysis for the corn- 214 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS mittee of the ages, sexes, and timing of occurrence of the Brown Booby at Southeast Farallon I., SF, in 2006 and 2007 (PCt; LCt, AKt, DEQt, JRT, MMRf, MBr ; 2007- 158), Peter Pyle and Jim Tietz concluded that up to seven individuals had reached the island: an adult male 25 Jan-22 May 2006, an adult female 28 Apr 2006-31 Dec 2007, an adult female 26 Sep 2006-30 May 2007, a second-year female 20 Oct 2006-15 Feb 2007, an adult male 30 May 2007, a second-year female 5-18 Aug 2007, and an adult female 5 Sep 2007. The four birds observed in both 2006 and 2007 were the same as those in record 2007-033 (Singer and Terrill 2009). The two adult males were of the pale-headed subspecies brewsteri. The committee reviews records of Brown Booby only through 2007. NEOTROPIC CORMORANT Phalacrocorax brasilianus (33, 8). At least two adults were at Parker Dam, SBE, between late 2010 and early 2011. An adult was there on 26 Dec 2010 (LHat, DVPt; 2010-185). Record 2011-007 included an adult on 16 Jan 2011 (with a second just over the state line in Arizona), an adult on 17 Jan 2011, and an adult on 4 Feb 2011 that showed two generations of feathers, distinguishing it from 2010-185 (DVPt; SRf; 2011-007). The committee voted that at least two individuals were represented by these records, including single individuals at Parker Dam 3 Dec 2011-21 Jan 2012 (TABe; DKaf, CAMf, DVP; 2011-201) and 2 km away at Quail Hollow on the Colorado River, SBE, 17 Apr 2011 (TABet; 2011-054). For several years three or four have been on the Arizona side of L. Havasu several kilometers from Parker Dam, and some individuals occasionally stray over the state line into California, making it difficult to know how many different individuals are involved in these records. An adult was southwest of Lack and Lindsey roads, south end of the Salton Sea, IMP, 27 Mar-9 Apr 2011 (BoMf; 2011-043). At Ramer L., IMP, an adult and a second-year bird were attending a partially constructed nest 8-9 Apr 2011; the nest was completed and both birds were seen at it through 14 Jun. On 23 Jun, the nest was gone and only the second-year bird, which lingered in the area to 3 Sep 2011, was seen (GMcC; BoMf, JMLf, BJSt; 2011-048). California has one previous nesting “attempt” — by a single bird that built a nest near the mouth of the New R., Salton Sea, IMP, but did not attract a mate (CBRC 2007). Given the increase in numbers in Arizona, where the species is now resident as far north as Phoenix (Radamaker and Corman 2008), we can expect further attempts. A second- year bird was at Headgate Dam, SBE, near Parker, Arizona, 20-23 Jul 2011 (DVPt; 2011-106). An adult was at Fig Lagoon near Seeley, IMP, 2 Sep-11 Dec 2011 (GMcC; 2011-124), and another was nearby at Sunbeam L., IMP, 3-11 Dec 2011 (GMcC; 2011-206). IDENTIFICATION NOT ESTABLISHED: The report of a juvenile on Southeast Farallon I., SF, 6 Sep 2010 (2011-036) exemplified the difficulty in identifying recently fledged cormorants. The bird was well described, photographed, and reported as the first Neotropic for northern California and the entire California coast in N. Am. Birds (65:158). The dark lores and supraloral area, acute angle of the posterior edge of the dull yellow gular pouch, mostly dark maxilla, dark breast, and long tail (barely shorter than the neck) pointed toward the Neotropic, though the bird’s structure appeared intermediate between that of a Neotropic and a Double- crested Cormorant ( P. auritus). A year later, the observer inspected a dead juvenile cormorant on Southeast Farallon I. that looked very similar to the claimed Neotropic but whose measurements were consistent with the Double-crested, which breeds on the Farallones. He therefore retracted the identification as a Neotropic, and the committee unanimously agreed on the record’s second round. Documentation of one reported at Ramer L., IMP, 20 May 2011 (2011-075) was inadequate for acceptance. One reported at the north end of L. Havasu, SBE, 28 Oct 2011 (2011-165) was likely identified correctly, but given the distance (more than 1 km) and brevity of the observation, the record was not accepted. TRICOLORED HERON Egretta tricolor (61**, 3). Adults were at San Elijo La- 215 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS Figure 6. Features distinguishing California’s first Common Snipe (Gallinago gallinago, the lower bird) from Wilson’s Snipe (G. delicata, the upper bird) include the much broader white tips to the secondaries, the more extensive white on the underwing coverts, and the overall much paler underwing. These birds were shot by a hunter near Lakeview in the San Jacinto Valley, Riverside Co., 11 Dec 2011 (2011-215). Photo by Kimball L. Garrett goon, SD, 9-15 May 2011 (EGKf, JKt, SBrt, GMcC; 2011-070), Desert Center, RIV, 22 May 2011 (CMcG t; CAMf; 2011-077), and L. Havasu, SBE, 10 Aug-17 Sep 2011 (LHat; DVP, JWet; 2011-115), the last being the first recorded in San Bernardino County. YELLOW-CROWNED NIGHT-HERON Nyctanassa violacea (59, 7). Two adults at Famosa Slough/San Diego River mouth, SD, 28 Apr-7 Sep 2011 (JPet; PEL, BLCt ; 2011-062) were presumably the same birds present in the vicinity intermittently since 27 Apr 2006 (2006-056, 2007-166, 2008-092, 2009-085, and 2010-041; Johnson et al. 2012). A third adult at this location 5 Jul 2011 (JPet; 2011-100) was considered the same bird present here 15 May-8 Aug 2010 (2010-042; Johnson et al. 2012) and 9 Jul-20 Aug 2009 (2009-112; Pyle et al. 2011a). A second-year bird was in Del Mar, SD, 6 May-10 Jul 2011 (PEL; GMcC, EWf, SBrt, JPet; 2011-066). An adult was at San Elijo Lagoon, SD, 22 Jun 2011 (JK, SSut; 2011-097). Two chicks were in a nest at Imperial Beach, SD, 27 Jun-20 Jul 2011 (JPet; 2011-096); one was found long dead on 28 Jul (SDNHM #52127), while the second was seen as a fledgling on 1 Aug. A juvenile at L. Balboa, Encino, LA, 26 Aug 2010 (PNt; 2010-092) was accepted on the third round after it was decided the blurry photos proved the bird was not the hybrid Yellow-crowned x Black-crowned Night-Heron ( Nycticorax nyctlcorax) seen 32 km away at Malibu Lagoon around the same time. Two adults at the Imperial Beach Sports Park, Imperial Beach, SD, 17 Dec 2011-29 Jul 2012 (PEL; GMcC, JPet; 2011-254) were presumably the same as some of those previously accepted from this location, where the Yellow-crowned Night-Heron has nested annually since 2006. IDENTIFICATION NOT ESTABLISHED: A juvenile reported at Pacific Beach, 216 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS Figure 7. This bird’s relatively bright coloration in relation to a Western Wood-Pewee (Contopus sordidulus), including a thin eye ring, and crisp white edges to the wing coverts and tertials, caught the attention of Debby and Jim Parker, who found this Eastern Wood-Pewee (C. uirens) in Birchim Canyon near Bishop, Inyo Co., 24-26 Oct 2011 (2011-159). Photo by Jon L. Dunn Figure 8. This Great Crested Flycatcher (. Myiarchus crinitus ) in Zzyzx, San Bernardino Co., 21 Sep 2011 (2011-136) was only the sixth to be found in California’s interior. Photo by Brad Singer 217 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS SD, 10-22 Nov 2011 (SMR; 2011-189) may have been of this species, but it was seen only in poor light at dusk, so several key features could not be assessed. The committee removed this species from the review list at its 2013 annual meeting. GLOSSY IBIS Plegadis falcinellus (26, 2). An adult at Owens L., INY, 11-12 Jul 2009 (CHot; DHot, BStf; 2009-125) went through four rounds of voting before being accepted unanimously. The appearance of red in the eyes in photos raised concerns of a hybrid with the White-faced Ibis (P. chihi), but the apparent red was likely a photographic artifact, as the bird’s features otherwise fit a Glossy Ibis perfectly. An adult was at the San Jacinto Wildlife Area near Lakeview, RIV, 18-28 Jun 2011 (CMcG t; CAMf, GMcC, LSf, DAB; 2011-085). IDENTIFICATION NOT ESTAB- LISHED: The unseasonal report of one at Unit 1 of the Salton Sea N.W.R., IMP, 8 Mar 2010 (2010-024) was supported by only blurry photos and a scant description. The committee had accepted only one March record, from the Prado Basin, RIV, 17 Mar 2008 (2008-044; Pike and Compton 2010). The description as “purplish” of the facial skin of a basic-plumaged adult on Rd. 102 southeast of Woodland, YOL, 27 Aug 2010 (2010-093) suggested a hybrid. BLACK VULTURE Coragyps atratus (7, 3). One at L. Casitas, VEN, 15-30 Oct 2011 (ABef, BEDf, MPt; 2011-158) was evidently the same as the one that wandered through Goleta, SBA, 10-13 Sep 2009 (2009-156), Santa Paula, VEN, 29 Nov-4 Dec 2009 (2009-221), L. Casitas/Ojai, VEN, 4 Jan-22 Feb 2010 (2010-016), and then back to Goleta, SBA, 18 Jul-6 Nov 2010 (2010-073), as revealed by the molt sequence and pattern of facial skin evident in photos (Johnson et al. 2012). The same characters indicated that one at Buellton, SBA, 14Jan2011 (BLSt; 2011-004) was a different individual. The flight feathers of a Black Vulture at the San Diego Wild Animal Park, San Pasqual, SD, 6 Mar 2011 (EGKf, TRSt; 2011-020) were extremely worn, leading four members to question its natural occurrence in the first round of voting, but alternative explanations (feather mites, electrocution) for this wear are possible, and the committee unanimously accepted the record in round two. This same bird (identified by patterns of wear) appeared in Goleta, SBA, 16-22 Apr 2011 (RHirf; 2011-089); such long-distance dispersal to the northwest further suggested that the bird had occurred naturally. Another was at Lone Pine, INY, 5 Aug 2011 (JLDf; 2011-112). IDENTIFICATION NOT ESTABLISHED: One reported near the Miramar landfill, San Diego, SD, 7 Jun 2011 (2011-086) was perched and seen only from a moving car by naked eye. HARRIS’S HAWK Parabuteo unicinctus (64, 12). A sizeable incursion of Harris’s Hawks, the largest in California since 1994, occurred in 201 1 . Single adults were near Jamul, SD, 22 Mar-26 May 2011 (JKt, MMt; 2011-042), near Boulevard, SD, 30 Mar 2011 (EAE; 2011-044), at Afton Canyon, SBE, 27 Mar 2011 (MC; 2011-046), and at Borrego Springs, SD, 11 Apr-7 Dec 2011 (WTH, SESf; 2011-050), the latter being considered the same bird as 2009-172 and 2010-124. One in its first spring was at Blythe, RIV, 14 Apr 2011 (JMMct; 2011-051). A pair of adults at Jacumba, SD, 18 Apr 2011-31 Dec 2012 (TCt; PEL, PKot, DDiT, GMcC, MBr, SSo, JKt, EGKf, DWAf; 2011-058) nested in an acacia tree; the female was nest-building in early May, incubating by 11 May, and feeding a single chick by 21 Jun. The chick fledged around 1 Aug and was seen through 31 Dec 2012 (PKot; JKt, PEL, EGKt; 2011-138). This pair was observed nest-building again in Feb 2012. Two additional Harris’s Hawks were also at Jacumba, SD, 12 Oct 2011-6 Jan 2012 (PEL; CAMt, LSt, RoR; 2011-154). An adult was at the north end of Poe Road, south end of the Salton Sea, IMP, 28 Apr 2011 (DJt; 2011-063). Two adults were suspected of nesting (though no evidence was observed) in McCain Valley near Boulevard, SD, 5 May-9 Aug 2011 (SMRt; EMt, RoR; 2011-067). Despite the obvious influx of Harris’s Hawks in 2011, there was still some resistance to acceptance of the records for Jamul and Afton Canyon on the grounds that these records were outside of the 218 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS species’ historical distribution and outside the distribution of records during the 1994 incursion (Patten and Erickson 2000). One member also expressed concern that the bird at Blythe may have been descended from birds transplanted along the Colorado River. NATURAL OCCURRENCE QUESTIONABLE: An adult at San Joaquin Wild- life Sanctuary, Irvine, ORA, 25-29 Nov 2011 (BCt; BEDf, JEP, CAMt; 2011-191) was unwary and approachable, and was near the coast where vagrancy of Harris’s Hawks is less likely than in inland desert areas. Given the incursion of 2011, most members thought that this bird might have occurred naturally. The committee removed this species from its review list at its 2013 annual meeting. COMMON CRANE Grus grus (1, 1). One at Lake Earl, DN, 5-8 May 2011 (ADBt; KRf, AJt, CCott, CAMt, GMcC, DWNf, MMRf, LSI; 2011-065; photo published in N. Am. Birds 65:569) provided California’s first record and the first for the west coast south of Alaska (Figure 5). Retained juvenal secondaries, along with pre-definitive features of the head and bill, suggested this bird was in its second or possibly third spring (P. Pyle pers. comm., 2013). This record engendered considerable discussion of the bird’s likely provenance. Most North American records considered to be of naturally occurring birds are from the middle of the continent (e.g., Nebraska, Kansas, and New Mexico) in March and early April with Lesser Sandhill Cranes (G. canadensis canadensis ) that breed in Alaska and northeastern Russia. Presumably, Common Cranes occasionally take up with these Sandhills and remain with them through the winter and migration. There are also several records of Common Cranes stopping during migration near Fairbanks, Alaska, in May (also with Lesser Sandhill Cranes). In early May, when the Lake Earl bird appeared, any naturally occurring Common Crane that had wintered in North America should already be in Alaska or Russia. Thus the timing of this bird’s occurrence and the fact that it was alone (rather than with Sandhills) were cause for concern. Moreover, Common Cranes are held in captivity in North America, and a known escapee from New York, which eventually bred with a Sandhill in New Jersey, set a precedent for wandering by escaped Common Cranes. However, the Lake Earl bird showed no abnormal wear, was unbanded, and had no clipped toes or other signs that it had ever been held in captivity. Furthermore, because some Common Cranes do not breed until they are 4-6 years old (Makatsch 1970, Glutz von Blotzheim et al. 1973), an immature bird might stray from a “normal” pattern of migration, both temporally and geographically. On the basis of plumage pattern, a Common Crane that appeared in the Queen Charlotte Islands, British Columbia, from 3 July into August 2011 may well have been the same as that at Lake Earl, suggesting that it was a lost wanderer. Though the committee agreed the bird’s origin could not be known with certainty, after two rounds of voting only two members questioned its natural occurrence. ’’"AMERICAN GOLDEN-PLOVER Pluvialis dominica (53, 1). A juvenile was at the Madera water-treatment plant, Madera, MAD, 21 Sep 2007 (GWf; 2011-127). The committee reviews records of the American Golden-Plover only from 2004 to 2009. LESSER SAND-PLOVER Charadrius mongolus (11, 1). An alternate-plumaged adult at Bolsa Chica, ORA, 25 Jun-2 Jul 2011 (BEDf; CAMt, EGKt, GMcC, TRSt, MMet; 2011-092) provided Orange County with its first record and was the first of this species recorded in California since 2005. This record also established an early date; the earliest previous occurrence was 10-12 Jul 2005 near McKinley ville, HUM (2005-083; Iliff et al. 2007). All California records have been of fall migrants. WILSON’S PLOVER Charadrius wilsonia (20, 4). Single males at the Tijuana River mouth near Imperial Beach, SD, 24 Mar 2011 (MSat; JLD, EGKt, GMcC; 2011-041) and at Bolsa Chica, ORA, 29 Mar 2011 (PKnt; 2011-105) were the first recorded in California in March; the previous early date was 2 Apr 2010 at the Santa Margarita River mouth, SD (2010-030; Johnson et al. 2012). Adult females were 219 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS Figure 9. This vireo photographed at Southeast Farallon Island, San Francisco Co., 9 Oct 2009 (2009-200) was in its first fall by the brown juvenal primary coverts contrasting with the replaced greater coverts of the formative plumage (see text). Note how different images of the same bird, in different posture and lighting, can make the malar contrast appear distinct or indistinct; because of the ambiguity of this feature, the committee concluded that it could have been either a first-fall male Cassin’s Vireo ( Vireo cassirtii) or a first-fall female Blue-headed Vireo (V solitarius), not accepting the record as the latter. Photos by Matt Brady (A) and Jim Tietz (B) Figure 10. Discovered on Christmas Day 2011 (photo taken 27 Dec), this adult Blue Jay ( Cyanocitta cristata) remained in Chico, Butte Co., through spring 2012 (2011-228). Photo by John T. Lewis 220 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS Figure 11. Two Northern Wheatears (Oenanthe oenanthe) were found on California’s coast in fall 2011. (A) The one photographed in Malibu, Los Angeles Co., 23 Sep 2011 (2011-137), appears to be an adult (and thus a female) because of its black wings and tail, broad and fresh primaries, and gray wash to the back and lesser coverts. (B) The other, at Anchor Bay, Mendocino Co., 2-5 Oct 2011 (2011-145) shows browner and more pointed wings and duller brown plumage, indicating an immature in its first fall. Photos by Daniel Tinoco (A) and Robert J. Keiffer (B) Figure 12. This Wood Thrush (. Hylocichla mustelina) brightened up the undergrowth at Shoshone, Inyo Co., 25 Oct-7 Nov 2011 (2011-157). Photo by Nancy Overholtz 221 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS at Border Field State Park, SD, 23 Jun 2011 (MSat; 2011-090) and Carpinteria Salt Marsh, SBA, 23 Jul-11 Sep 2011 (PAGaf; WTF t, CAMt, DWNf, JLDf, OJt, SBTf; 2011-109); although both birds were heavily worn, different patterns of wear confirmed that they were different individuals. Eleven of the state’s 20 accepted Wil- son’s Plovers have occurred since 2005, suggesting a recent upswing in occurrence or perhaps increased coverage of beaches by researchers monitoring the Snowy Plover (C. alexandrinus) and Least Tern ( Sternula antillarum). COMMON RINGED PLOVER Charadrius hiaticula (1, 1). An alternate-plumaged one-year-old male at the Davis Wetlands, YOL, 19-26 Aug 2011 (ToEf; SCHf, LBt, DWNt, CAMt, LSt, GMcC, BYf, JLDf, LHf, JMf, JCSt; 2011-118; Sterling and Easterla 2012), provided the first accepted record for California and the second for western North America outside western Alaska, the first being of a vocal immature at Port Susan Bay, Washington, 23 Sep 2006 (Aanerud 2011). The one at Davis also called frequently, facilitating the identification. See Sterling and Easterla (2012) for details of this record. California’s one previous report, from at Pt. Reyes, MRN, 9 Sep 1996, was of a bird well seen and heard by an experienced observer (1997-071; Rottenborn and Morlan 2000), but the committee was reluctant to accept a first state record without photos or documentation from more observers. UPLAND SANDPIPER Bartramia longicauda (30, 1). An Upland Sandpiper call- ing and displaying near Bellota, SJ, 25-28 Jun 2011 (LPf; 2011-093) was the first recorded in the Central Valley. Previous records were all of spring (15 May-13 Jun) or fall (8 Aug-28 Oct) migrants, but that for Bellota fits neither of those categories. This species has been declining over much of its range, and it has been recently extir- pated as a breeding species in Washingon and Oregon (P. E. Lehman pers. comm.). BAR-TAILED GODWIT Limosa lapponica (39, 1). A bird acquiring alternate plum- age was at the Santa Barbara harbor, SBA, 21 May 2011 (WTFf; 2011-076; photo published in N. Am. Birds 65:570). As summarized by Hamilton et al. (2007), the vast majority of California’s Bar-tailed Godwits have been fall migrants from Monterey County northward. The one in Santa Barbara was the first recorded in May and only the second apparent spring migrant, the first being near Crescent City, DN, 3-5 Jun 1984 (1984-186; Dunn 1988). This was also a first for Santa Barbara County and only the ninth south of Monterey County. LITTLE STINT Calidris minuta (19, 6). One in its first fall with extensive forma- tive feathering at Abbott’s Lagoon, Pt. Reyes, MRN, 16-17 Dec 2010 (AKf; KHat; 2011-114) is the latest recorded in California; the previous latest record was of one in formative plumage collected at Harper Dry L., SBE, on 21 Nov 1988 (1990-210; Heindel and Garrett 1995). The combination of dark legs, lack of webbing between the toes, long primary projection, and bill length/shape eliminated the Western (C. mauri ) and Least (C. minutilla) Sandpipers. Distinction from the Red-necked Stint (C. ruficollis), which is very similar in formative and basic plumages, rested primarily on the rusty-brown edging on the retained juvenile tertials and upperwing coverts, secondarily on body shape and a split supercilium. Alternate-plumaged adults were at the Alexandre Dairy near Fort Dick, DN, 6-8 Jul 2011 (ADBf; KMBf, KRt, JCSt; 2011-101), Davis Wetlands, YOL, 15-17 Jul 2011 (ToEf; SCHf, LH, SBTf, MBr, MMef; 2011-104), and Piute Ponds on Edwards Air Force Base, LA, 23-25 Jul 2011 (KLGt; NFf, MFrf, KH-Lf, CAMt, LSt; 2011-108; photo published in N. Am. Birds 65:715). While the timing and southward progression of these records led to speculation that at least two of them represented the same bird, comparison of photos suggests that these were all different individuals. Other alternate-plumaged adults at Crab Park, HUM, 15-16 Aug 2011 (DCol; 2011-221) and Mad River County Park, HUM, 21 Aug 2011 (DCol, RFo; 2011-119) were different individuals on the basis of plumage differences noted by Daryl Coldren, who saw both birds. Five individuals 222 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS in 2011 bested the previous single-year high of three in 2010. WHITE-RUMPED SANDPIPER Calidris fuscicollis (25, 1). One acquiring alternate plumage along Wheeler Road at the north end of the Salton Sea, RIV, 8 May 2011 (ADf; 2011-069) provides the earliest accepted record for California; the previous early date was 17 May 1986 at Stockton, SJ (1986-341; Langham 1991). The com- mittee has taken an especially conservative approach to all reports of this species prior to mid-May, but the photos of 2011-069 were adequate for acceptance by all but one member, who thought the bird might be a Western Sandpiper. CURLEW SANDPIPER Calidris ferruginea (41, 2). A bird acquiring alternate plum- age was at the J Street Marina, Chula Vista, and the nearby saltworks at the south end of San Diego Bay, SD, 24-27 Apr 2011 (JMLt; BLC, GMcC, MSat; 2011-057), while an alternate-plumaged bird was in adjacent Imperial Beach, SD, 24-30 Jul 2011 (GMcC; ABL, CMcG f, AnMf, VMf, DWNf, LSt, TABlf, BEDf, CAMf, JNf; 2011-107; photo published in N. Am. Birds 65:715). IDENTIFICATION NOT ESTABLISHED: After two rounds of review, most members found the photos of a somewhat distant sandpiper at Owens Lake, INY, 16 Sep 2007 (2007-208) insufficient for identification and thought that the bird might have been a Stilt Sandpiper (C. himantopus). COMMON SNIPE Gallinago gallinago (1, 1). A hatch-year bird shot by a hunter in the San Jacinto Valley, about 1.5 km northwest of Lakeview, RIV, 11 Dec 2011 (KLGf; 2011-215) is the first recorded in California and the first for western North America away from the Aleutian Islands/Bering Sea region. Although the hunter retained the specimen, Kimball Garrett took numerous measurements (including the width of the outermost rectrix, much broader in the Common than in Wilson’s), pho- tographs, and a small sample of tissue for genetic analysis. The features distinguishing the Common Snipe and Wilson’s Snipe (G. delicata) have been reviewed by Carey (1992), Carey and Olsson (1995), Bland (1998, 1999), Dunn and Alderfer (2007), and Pyle (2008). Most are difficult to impossible to assess in the field, but whiter underwings (with less dark barring) and broader white trailing edges to the wings (see Figure 6) may be noticeable. LITTLE GULL Hydrocoloeus minutus (107, 4). The CBRC considered an adult at L. Perris and the San Jacinto Wildlife Area, RIV, 6-27 Mar 2011 (CAM; DFuf, JLD; 2011-023) the same as one at this location 1-14 Mar 2010 (2010-019; Johnson et al. 2012) and 1-10 Mar 2009 (2009-056; Pyle et al. 2011a). Individuals in full juvenal plumage were at China L., KER, 29 Aug 2011 (SLSf; JLDf; 2011-123) and L. Havasu, SBE, 10 Sep 2011 (LHa S, DVP; 2011-131). Juvenal-plumaged birds had been previously recorded in California on four occasions, all during the period 15 Aug-5 Sep. First-winter birds were at Prado Regional Park, SBE, 14-18 Dec 2011 (HBK; TABef, BrSt, CAM, DWAf, BJSt; 2011-212) and L. Los Carneros in Goleta, SBA, 29 Dec 2011 (LWf; 2011-230). BLACK-TAILED GULL Larus crassirostris (3, 0). An adult was at Malibu, LA, 12 Feb 2011 (JFt; 2011-015). A majority of the CBRC considered this bird probably the same as the one 60 km to the southeast in Long Beach, LA, 8-21 Nov 2010 (2010-140; Johnson et al. 2012). IDENTIFICATION NOT ESTABLISHED: An adult reported near Vernalis, STA, 28 Mar 2011 (2011-052) may have been an adult or near-adult Lesser Black-backed Gull (L. fuscus). Reports from the north end of the Salton Sea, east of Avenue 86, RIV, 9 Apr 2011 (2011-049) and Salt Creek, east shore of the Salton Sea, RIV, 30 May 2011 (2011-080) were inconclusive. ICELAND GULL Larus glaucoides (11, 4). Second-winter birds apparently of the subspecies kumlieni were at the Ogier Ponds north of Morgan Hill, SCL, 4 Dec 2010 (SCRt; 2011-026) and 5-9 Mar 2011 (SCRt; RWRf; 2011-027, same as 2011-026), at Pillar Pt. Harbor, Half Moon Bay, SM, 4 Feb 2011 (AJt; 2011-017), 223 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS Figure 13. The bold head and throat pattern of a male Golden-winged Warbler (' Vermivora chrysoptera) often gives an initial impression of a chickadee rather than a warbler. This individual was photographed in Birchim Canyon near Bishop, Inyo Co., 2 Jun 2011 (2011-083). Photo by Debby Parker and at the mouth of Pilarcitos Creek in Half Moon Bay, SM, 24 Feb 2011 (AJt; 2011-260). An even paler second-winter bird at the Davis water-treatment plant, YOL, 19 Feb-14 Mar 2011 (ToEf; 2011-242) could have been a paler kumlieni or possibly nominate glaucoides. These four individuals were larger than most Iceland Gulls previously accepted in California, being similar in size to many Thayer’s Gulls (L. thayeri). While there may have been a bias in the past toward considering only smaller birds (i.e., females and smaller males) to be acceptable as Iceland Gulls, the committee recognizes that only a small percentage of California's apparent Iceland Gulls are very small, dainty, round-headed, and fine-billed. All four individuals were also in their second year, the age at which the melanism in the wingtips reaches its maximum, and thus less likely than first-cycle birds to have pale wingtips as a result of fading and wear. First-year birds that are as pale as some Kumlien’s Gulls on the east coast are not very rare in California in late winter, but most such birds have solid darker tail bands and secondary bars suggesting faded Thayer’s. Previously accepted records extended from 30 Dec to 7 Mar; these new records expanded this interval to 4 Dec-14 Mar. IDENTIFICATION NOT ESTABLISHED: An adult at the mouth of Salmon Creek, SON, 9 Jan 2008 (2011-120) may have been a Kumlien’s Gull, as the photos showed a smallish gull with dark gray, not black, in the primaries. However, the images did not show the bird from multiple angles, making it difficult for members to assess the bird’s size and shape; the bird may have been a small Glaucous-winged Gull (L. glaucescens). One photographed at the Recology Composting Facility, STA/ SJ, 27 Mar 2011 (2011-264) could not be identified conclusively. 224 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS LESSER BLACK-BACKED GULL Larus fuscus (95, 16). In 2011, 18 individu- als were recorded, two of them being considered continuing birds, as follows: at the Nimbus Fish Hatchery, SAC, an adult 9 Jan-5 Feb 2011 (BGf; CCoaf; DBrt; LPf; 2011-002); at Pillar Pt. Harbor, SM, a first-winter bird 18 Jan 2011 (PAt; 2011-150); at Mecca Beach, RIV, adults 1-4 Feb 2011 (CMcGt; 2011-012, considered the same as one there 10-28 Feb 2010 [2010-015; Johnson et al. 2012] and 18 Jan-15 Feb 2009 [2009-013; Pyle et al. 2011a]), and 1 Fet^l3 Mar 2011 (CMcG; TABef, BrSt; 2011-013); at Lower Otay L., SD, a first-winter immature 5-8 Feb 2011 (PEL; JKt, GMcC, LSI, BJSt, VMf, CAMf; 2011-014); at the Davis water-treatment plant, YOL, a first-winter immature 21 Feb - 14 Mar 201 1 (ToEf; 201 1-246; photo published in N. Am. Birds 65:513); at the American Ave. landfill, FRE, a first-winter immature 3 Mar 2011 (GWf; 2011-028, a first for Fresno County); at Salt Creek on the east shore of the Salton Sea, RIV, a first-winter immature 12-13 Mar 2011 (CMcG; TABef, BrSt; 2011-024); at Clear L., LAK, one likely in itsfourth spring 20 Mar 2011 (FEHt; DEWt; 2011-047), presumably the same as an adult at Redbud Park, Clear L., LAK, 21 Nov 2011-20 Jan 2012 (FEHt; 2011-185); at Obsidian Butte, south end of the Salton Sea, IMP, a one-year-old 30 May-30 Jun 2011 (DDiTf; GMcC, JPawt; 2011-079) and three-year-olds 18 Sep-9 Nov 2011 (GMcC; 2011-133) and 20 Dec 2011 (GMcC; 2011-227); at the South Wilbur Flood Area, KIN, a two-year-old 7 Sep 2011-18 Feb 2012 (MESf; SLSf, JFLf, SSumf, ESHf; 2011-125); at the Norco Egg Ranch, near Lakeview, RIV, an adult 19 Nov 2011-13 Jan 2012 (HBK; TABef, CAMf, BJSt; 2011-193, considered the same as the one there 28 Dec 2010-26 Mar 2011, 2010- 188; Johnson et al. 2010) and a second-winter bird 20 Nov 2011-10 Jan 2012 (HBK; CAMf, BrSt, AEK; 2011-194); at Clear L. south of Lakeport, LAK, one in its third or fourth fall 21 Nov 2011 (FEHt; 2011-262;); at the Yolo County Landfill, YOL, an adult 10 Dec 2011 (SCHf; 2011-214); and at Alviso, SCL, an adult 18 Dec 2011 (DMdt; 2011-234). IDENTIFICATION NOT ESTABLISHED: At the Alexandre Dairy near the Smith River, DN, a first-winter bird 3 Dec 2010 and 30 Jan 2011 (2010-189). Records of the Lesser Black-backed Gull have increased dramatically in recent years, and the committee removed this species from its review list at its 2013 annual meeting. SLATY-BACKED GULL Larus schistisagus (43, 5). An adult was at the Potrero Hills landfill, SOL, 25 Feb-8 Mar 2011 (RMu S; 2011-245); although the committee rarely accepts records of this species lacking photographs, the thorough details and annotated sketch were convincing. Other adults were at the Pilarcitos Creek mouth, Half Moon Bay, SM, 11 Mar 2011 (AJf; 2011-257) and Pacific Commons Linear Park in Fremont, ALA, 17-29 Nov 2011 (NAf; MP, JLDf, DEQt, JTin, AD; 2011- 180), the latter preceding California’s previous earliest fall date of 2 Dec. An advanced third-spring bird was at Pomponio State Beach, SM, 4 Mar 2011 (OJt; 2011-021). One in its second spring was at Half Moon Bay, SM, 5 Mar 2011 (DSSf; 2011-102). IDENTIFICATION NOT ESTABLISHED: Features apparent in photos of an adult gull at Ferry Pt., Miller/Knox Regional Shoreline, CC, 11 Feb 2011 (2011-243) were consistent with a Slaty-backed Gull, but the photos did not show the primary pat- tern well, and no written details accompanied the report. Accompanying the recent proliferation of digital photography, which has been a boon to the documentation of bird records in general, has been a decrease in written details, and many records submitted to the committee lack any description of what the observers noted in the field. Even minimal details on a bird’s appearance, accompanying photographs, can assist significantly in documenting rare bird records. SOOTY TERN Onychoprion fuscatus (13, 1). An adult at Bolsa Chica, ORA, 29 Apr-6 Aug 2011 (BEDf; 2011-095) was presumably the same as one at this location 4 Apr-18 Jul 2010 (2010-062; Johnson et al. 2012) and 28 Jul-30 Aug 2009 (2009-154; Pyle et al. 2011a). An adult was at the Tijuana River mouth, SD, 23 Jun 2011 (MSa; 2011-091). 225 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS THICK-BILLED MURRE Uria lomvia (50, 0). IDENTIFICATION NOT ESTAB- LISHED: One reported at Pt. Pinos, MTY, 29 Jan 2011 (2011-018) was seen rather briefly in flight. The description of a dark head with contrasting white chin, white flanks, and upperparts more blackish than the brown-tinged upperparts of a Common Murre {U. aalge) all fit the Thick-billed nicely. However, the bill was not scrutinized or described in detail. Moreover, while white flanks should allow a Thick-billed Murre to be distinguished from Common Murres of the race calif ornica, some individuals of the race inornata, which breeds in Alaska and British Columbia, have white flanks, and inornata could conceivably reach California in winter. The committee is reluctant to accept fly-by records of the Thick-billed Murre. LONG-BILLED MURRELET Brachyramphus perdix (29, 1). One in formative or basic plumage was 200 m off the north jetty of the harbor at Crescent City, DN, 3 Feb 2011 (JJf; 2011-011). SNOWY OWL Bubo scandiacus (61, 1). A record of one at Delevan N.W.R., COL, 9 Jan 1971 was accepted on the basis of old field notes taken shortly after the observation but only recently submitted (FH; 1971-501). ELF OWL Micrathene whitneyi (3**, 0). One of a pair (2010-050; Johnson et al. 2012) that nested at Corn Spring in eastern Riverside County in 2010 returned to the nest site 17 Apr-28 May 2011 (TABe; DEQ, RT§; 2011-056). See Johnson et al. (2012) for more information on the Elf Owl’s recent status in California. RUBY-THROATED HUMMINGBIRD Archi loch us colubris (14, 1). A first-fall male visited a feeder in Bolinas, MRN, 30 Sep-1 Oct 2009 (KHat; 2009-247), the site of two other records of the Ruby-throated Hummingbird (Singer and Terrill 2009, Johnson et al. 2012). BROAD-BILLED HUMMINGBIRD Cynanthus latirostris (81, 3). A first-fall female at Palomarin near Bolinas, MRN, 6 Oct 2011 (RDf; 2011-259) was unusual by being relatively early, a female, and from northern California. One near Mission Valley in San Diego, SD, 6 Nov- 2 Dec 2011 was more typical, being a male in early winter from the southern coast (MP; PELt; 2011-170). Another adult male was at Cotton- wood Springs, Joshua Tree National Park, RIV, 10 Mar 2011 (CHaf; 2011-022). RED-BELLIED WOODPECKER Melanerpes carolinus (0, 0). IDENTIFICATION NOT ESTABLISHED: First state records require thoroughly convincing documenta- tion to gain acceptance. A reported Red-bellied Woodpecker at Caswell State Park, STA, 17 Sep 2011 (2011-134) was seen and described only briefly by a single ob- server. Most members concluded other species of “ladder-backed” woodpeckers such as the Nuttall’s Woodpecker ( Picoides nuttallii) or a female Williamson’s Sapsucker (Sphyrapicus thyroideus) were not adequately eliminated. Although generally prone to only limited vagrancy, this species is expanding its range north and west, and records from near Calgary, Alberta, Coeur d’Alene, Idaho, and Ruby Lake National Wildlife Refuge, Nevada (Burton and Anderson 2013), demonstrate its capability of wandering west. CRESTED CARACARA Caracara cheriuiay (11, 0). An adult in the Tijuana River valley, SD, 25 Sep-1 Oct 2011 (GN; MSn|; 2011-139) was presumably the same as the one seen there sporadically 9 Sep 2006-14 Nov 2009 (2006-127, 2007- 144, 2008-093, and 2009-076; Pike and Compton 2010, Nelson and Pyle 2013). GYRFALCON Falco rusticolus (11, 0). IDENTIFICATION NOT ESTABLISHED: Most members considered the submitted video of a large falcon at Smith River, DN, 12-14 Nov 2007 (2007-259) inadequate to support the identification as this species. EASTERN WOOD-PEWEE Contopus virens (12, 1). A pewee in Birchim Canyon, INY, 22-26 Oct 2011 caught the attention of two observers by the late date and the 226 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS bird’s relatively bright plumage, which included a smooth pale gray nape and rear auriculars contrasting with a darker cap, a whitish throat, thin bright eye ring, and a pale mandible. Suspecting an Eastern Wood-Pewee, the observers kept tabs on the bird, which finally was heard vocalizing on the fourth day of observation, confirming its identity (DPaf, JPar=j= ; JLDt, JoHf, TH, CHof, RoH, NOt; 2011-159; Figure 7). DUSKY-CAPPED FLYCATCHER Myiarchus tuberculifer (88, 6). One first-winter bird at Shoshone, INY, 5-7 Nov 2011 was relatively early (LSWt; 2011-166). Other immatures wintered at Aviara Golf Course in Carlsbad, SD, 11 Dec 2011-2 May 2012 (SBrt; DWAf, TABf, CGt, CAMf, GMcC, BJSf; 2011-211), San Lorenzo River in Santa Cruz, SCZ, 17 Dec 2011-16 Jan 2012 (PBf; SGt, JLRt, PSt; 2011- 220), and Golden Gate Park, SF, 27 Dec 2011-28 Jan 2012 (ASHf; MEat, OJt§, JMrf, KSf, LSf ; 2011-225). Another of uncertain age was at Demuth Park in Palm Springs, RIV, 20-27 Mar 2011 (BWf; TABef, JLD, GMcC, SJMf; 2011-034). One at Veterans Memorial Park in Bell Gardens, LA, 2 Feb-21 Mar 2011 (RBt; BEDt, KLGt; 2011-010) returned the following fall, being noted 5 Nov 2011-6 May 2012 (RBt; CAMf, ASet; 2011-232). Finally, one returned for its fifth consecutive winter to La Mirada Creek Park, LA, 1 Dec 2011-25 Feb 2012 (JRwtt; JSFff; 2011-233; 2008-040, 2008-187, 2009-222, and 2011-037; Johnson et al. 2012). IDENTI- FICATION NOT ESTABLISHED: One reported on Clark Mt., SBE, 6 July 2011 (2011-103) would have represented a first summer record for California. While many members thought the identification may have been correct, the observer’s incomplete and brief view and inconclusive photographs were not sufficient to confirm such an unprecedented date and location. GREAT CRESTED FLYCATCHER Myiarchus crinitus (55, 1). One at Zzyzx, SBE, 21-22 Sep 2011 was only the sixth recorded for California’s interior (BrSt; TABet, JLDt, CAMf, JEPt; 2011-136; Figure 8). IDENTIFICATION NOT ESTABLISHED: A Myiarchus seen by two observers on 26 and 27 May 2010 in Olivenhain, En- cinitas, SD, may have been correctly identified but no vocalizations were heard, and some members were concerned the bird may have been a vagrant Brown-crested Flycatcher (M. tyrannulus). There is only one accepted spring record of the Great Crested Flycatcher for California. THICK-BILLED KINGBIRD Tyrannus crassirostris (20, 1). One at the mouth of Poggi Canyon in Otay Valley, Chula Vista, SD, 25 Oct 2011-25 Mar 2012 (DWAt, TJf, GMcC, RBMcNf, JPt; 2011-184) returned for its second documented winter (it was an adult when first found; 2010-176, Johnson et al. 2012). A first-year bird at Middle Ranch on Santa Catalina I., LA, 19 Nov 2011-3 Apr 2012 (LFf; JKif; 2011-192) represented California’s first record from an offshore island. BLUE-HEADED VIREO Vireo solitarius (68, 2). One was on Santa Barbara I., SBA, 16-17 Sep 2011 (NAL; PAGat; 2011-147); another was at Shay Park, HUM, 19 Sep 2011 (DCol; 2011-209). One at Pt. Loma, SD, 20 Nov 2011-28 Jan 2012 (PEL; 2011-177) had returned for its second winter (2011-001, Johnson et al. 2012). IDENTIFICATION NOT ESTABLISHED: Reports from near Alton, HUM, 8 Dec 2011 (2011-204), Arroyo Grande Cr., SLO, 30 Sep 1997 (2008-193), Lake Merced, SF, 3 Oct 2009 (2009-199), and Southeast Farallon I., SF, 9 Oct 2009 (2009-200) were not accepted on their fourth rounds of voting. These and other reports of birds seemingly intermediate between the Blue-headed and Cassin’s ( V. cassinii) vireos prompted several committee members to re-examine identification criteria of these two species through study of specimens and consultation of outside experts. Findings indicated some characters previously thought to distinguish these two species can be shared, but the majority of birds of known age should be identifiable (P. Pyle in lift., 2011). The extent of white edging in the outer rectrices is less reliable for identification than previously thought, although within each age/sex 227 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS class it averages broader in the Blue-headed than in Cassin’s (P. Pyle in litt . , 2011). Head color, often the primary feature used to distinguish these two species, is also variable; the brightest 2-3% of adult male Cassin’s Vireos overlap with the dullest 2-3% of immature female Blue-headed Vireos, illustrating the importance of age and sex in accurate identification. Malar contrast, sharper in the Blue-headed than in Cassin’s, is helpful, though apparent contrast can vary with a bird’s posture. Some individuals, even those well photographed, remain unidentifiable because of overlap between bright males of Cassin’s and dull females of the Blue-headed (Figure 9). The best features to age these vireos are the color of the primary coverts (browner and contrasting with the brighter replaced greater coverts in first-year birds, duskier, with greenish edging, and not contrasting with the wing coverts in adults), and the narrower and more worn and pointed rectrices of first-year birds (Pyle 1997). YELLOW-GREEN VIREO Vireo flavoviridis (103, 3). First-fall individuals were at Neary Lagoon in Santa Cruz, SCZ, 16-19 Oct 2011 (SGf; JoGt, CKf, SLf, WNt, SBTf; 2011-148) and Southeast Farallon I., SF, 18 Oct 2011 (MD; MBrf, MElf; 2011-251). A third of uncertain age was in Goleta, SBA, 23 Oct 2011 (DMC; RFC; 2011-152). All three represented dates and coastal locations typical for this vagrant in California. IDENTIFICATION NOT ESTABLISHED: A brief report of one in Chula Vista, SD, 10 Sep 2010 (2011-263) lacked diagnostic details, mentioning the bird had pale yellow flanks rather than the extensively yellow underparts extending up to the sides of the neck as expected on a Yellow-green Vireo, and the date was unusually early. BLUE JAY Cyanocitta cristata (15, 1). An adult found on 25 Dec 2011 in Chico, BUT, remained through 8 May 2012, the latest date the species has been documented in California (MFif; TABef, SBTf; 2011-228; Figure 10). WINTER WREN Troglodytes hiemalis (8, 3). One in Malibu, LA, 18 Dec 2005 (KLG; 2010-164) was heard vocalizing and seen briefly on a Christmas Bird Count. Two found by the same observer in Sacramento County were supported with video and sound recordings, one along the American River Parkway 28 Sep-5 Dec 2009 (CCotf§; 2011-219), the other at the Consumnes River Preserve 27 Nov 2011-25 Feb 2012 (CCof^ 2011-241). Vocalizations most readily distinguish the Winter Wren from the Pacific Wren (T pacificus), until recently considered conspecific (Toews and Irwin 2008, Chesser et al. 2010). Although there is some variation in the common call notes of the Winter Wren, no overlap has yet been reported in the quality of the voices of the Winter Wren and the Pacific Wren. IDENTIFICATION NOT ESTAB- LISHED: A wren photographed at Stovepipe Wells in Death Valley, INY, 2 Nov 2002 (2010-163) was not heard vocalizing, and most members declined to accept a record based on plumage characteristics alone. In an analysis of 26 adult Winter Wren and about 200 adult Pacific Wren specimens, Peter Pyle (in litt., 2013) found considerable overlap in plumage characteristics of both species, reinforcing the limitations that ex- ist in identifying either species on plumage characteristics alone. A report from near Blythe, RIV, 16 Apr 2011 (2011-060) was not accepted because the description of the call was incomplete and how the Pacific Wren was eliminated was not explained. With only eight records (none for spring) so far accepted, the committee still takes a cautious approach with this species until its identification, schedule, and distribution are better understood. Observers of this species should record any calls, if possible, or otherwise describe the calls and plumage in detail. RED-FLANKED BLUETAIL Tarsiger cyanurus (2, 1). Biologists working on San Clemente I., LA, found this remarkable vagrant on 6 Dec 2011 (JRuf; JTSt, SDNHM #53312; 2011-202); it represents only the second record for North America outside of western Alaska. The first for California (and the lower 48 states) was also from an off-shore island, Southeast Farallon I., 1 Nov 1989 (Patten and Erickson 1994). 228 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS After a few days of searching for and not refinding the bird on San Clemente I., the observers finally discovered it stabbed on a cactus spine — the victim of a Loggerhead Shrike (Lanius ludovicianus). Philip Unitt nonetheless prepared the headless remains of this first-fall female as a specimen. Remarkably, this is the third Old World species of the Muscicapidae recorded on San Clemente I., following records of the Stonechat ( Saxicola torquatus) in Oct 1995 (2005-005; Cole et al. 2006) and the Bluethroat (. Luscinia svecica) in Sep 2008 (2008-116; Pike and Compton 2010). Red-flanked Bluetails of the nominate race breed primarily in northern taiga from the Kamchatka Peninsula and Japan west to Finland and winter in southeast Asia. In North America, besides the two records listed above and one near Vancouver, British Columbia, 13 Jan-Mar 2013 (http : //blog . aba . org/ 2013/0 1/abarare-red-flanked-bluetail-british- columbia.html), it is a very rare spring and fall vagrant to western Alaska islands. NORTHERN WHEATEAR Oenanthe oenanthe (13, 2). Two were photographed (Figure 11): an adult female at Malibu, LA, 23 Sep 2011, found near dusk and gone the next day (DTif; 2011-137), and a first-fall female near Anchor Bay, MEN, 2-5 Oct 2011 (PK ; KAH, RJKt, DTo, RTr, JWh; 2011-145). Although two accepted records seem remarkable for a single season, there is a precedent: the falls of 1988, 1992, 1995, and now 2011 have all had two accepted sightings of this Old World rarity, which had not been recorded in California since 2001. WOOD THRUSH Hylocichla mustelina (28, 2). Two fall vagrants were found in California’s interior, where half of California’s Wood Thrushes have occurred. One at Galileo Hill, KER, 1 Oct 2011 (SLS; AH, KHL; 2011-140) is California’s earliest in fall by eight days. Another was in Shoshone, INY, 25 Oct-7 Nov 2011 (LWf; PELf, CAMt, NOt; 2011-157; Figure 12). RUFOUS-BACKED ROBIN Turdus rufopalliatus (17, 1). One was at Picacho State Recreation Area, IMP, 13 Nov 2011 (GjH; 2011-198). CURVE-BILLED THRASHER Toxostoma curuirostre (28, 1). One of the subspe- cies palmeri singing in El Monte, LA, 10 May-6 Jul 2011 was the second for Los Angeles County, the fourth for the coastal slope of California, and the first for Cali- fornia in summer (CAMf§; BEDf, JLDt, JSFt, KLG; 2011-072). Another returned for its second winter (23 Nov-3 Dec 2011) to Black Meadow Landing, Colorado R., SBE (TABet; 2011-199; previously accepted as 2011-003; Johnson et al. 2012). IDENTIFICATION NOT ESTABLISHED: A thrasher with a deformed, abnormally long bill was 16 km southeast of Holtville, IMP, 21 Jan-3 Feb 2011 (2011-009). It was either a Bendire’s (T. bendirei) or a Curve-billed Thrasher and unfortunately did not vocalize. SNOW BUNTING Plectrophenax nivalis (124, 2). One photographed near Orick, HUM, 14-15 Nov 2010 (TKf; LETf; 2010-148) had considerable white in the rump, tail, and secondaries, apparent dark bases to the outer greater coverts, and was pale overall, raising the question of its being a McKay’s Bunting ( P. hyperboreus ) or a hybrid McKay’s x Snow Bunting. Outside expert J. Rogers (in litt., 2010), however, advised that this bird was within the range of a first-fall male Snow Bunting. See Rogers (2005) for more information on distinguishing these species. Another apparent first-winter male Snow Bunting was near Kneeland, HUM, 15 Nov 2011 (MJMazf; 2011-210). WORM-EATING WARBLER Helmitheros vermivorum (120, 1). One was at Ramer Lake, IMP, 26-27 Nov 2011 (CrSf; HBK, BKrf, KZKt, CAMf, GMcC; 2011-195). GOLDEN-WINGED WARBLER Vermivora chrysoptera (75, 2). A male photo- graphed in Birchim Canyon, near Bishop, INY, 2 Jun 2011 was the 25th recorded in eastern California (JParf, DPa; NOf, CHot, RoH; 2011-084; Figure 13). A first- fall female on Santa Barbara I., SBA, 20-21 Oct 2010 (WTFf; 2010-181) was the 229 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS second for the Channel Islands. CONNECTICUT WARBLER Oporornis agilis (115, 3). One was at the fish docks on Pt. Reyes, MRN, 12-13 Sep 2011 (SCH; 2011-128). One in its first fall at Oxnard, VEN, 2-5 Oct 2011 (SCrt; JCa, DCt, JLD, HBK, CAM; 2011-144) was initially found and photographed by a 12-year-old observer. A third on Southeast Farallon I., SF, 20 Oct 2011(MBr, MD, DMxf; 2011-252), also in its first fall, is one of latest recorded in California. MOURNING WARBLER Geothlypis Philadelphia (141, 1). One was in Eureka, HUM, 14-16 Sep 2011 (SEMf, JSa, SBT; 2011-168). IDENTIFICATION NOT ESTABLISHED: A singing Geothlypis lacking white eye arcs, observed near Brooks, YOL, 19 May 2011 (2011-074), may have been a Mourning Warbler, but several committee members had concerns about the identification. The relative length of the undertail coverts was not noted, and description of the hood pattern, in which the black was concentrated on the lores yet absent from lower part of the breast, may have been better for the MacGillivray’s Warbler (G. tolmiei ) or a hybrid. CAPE MAY WARBLER Setophaga tigrina (20**, 12). Twelve records accepted in 2011 were far above the average reported in California in recent years. A spring male was photographed in Birchim Canyon near Bishop, INY, 14 May 2011 (DPat; BJK, NOf; 2011-071). Ten, apparently all in their first fall, were found between 11 Sep and 23 Oct: at Pt. Reyes, MRN, 11-25 Sep (SBT; 2011-162), Lake Merced, SF, 14-15 Sep (DMo; RFi, AIM, JMr, LSt; 2011-129), Laguna Niguel, ORA, 17 Sep (RBMcNf; 2011-132), Fort Mason, SF, 26-27 Sep (DMOf; MEat; 2011-143), at Southeast Farallon I., SF, 28 Sep (MBrf, DMxf, JRT; 2011-249) and 1 Oct (MBrf, DMxf, JRT; 2011-250; different individual), Lighthouse Field, SCZ, 30 Sep (JMyt; 2011-141), Twenty-nine Palms, SBE, 30 Sep (TABef§; CAMf; 2011-146), Santa Barbara I., SBA, 22-25 Oct (PAGaf; WTFf, MVif; 2011-156), and the campus of the University of California, Santa Cruz, SCZ, 23 Oct (JoGS; 2011-151). Finally, a first-winter male was found on the Pt. Reyes Christmas Bird Count in Inverness, MRN, 17 Dec 2011 (SCyt; 2011-226). The committee added this species to the review list in 2011 on the basis of a significant downward trend in California reports. From 1980 to 1995, California averaged 5.5 reports per year, but from 2000 to 2009, that number dropped to 3.3 (J. Dunn, in litt.). The Cape May Warbler’s population cycles with outbreaks of the Spruce Budworm ( Choristoneura fumerifana) in the boreal forest, so 2011 may have been a good summer for budworms. IDENTIFICATION NOT ESTABLISHED: The description of a bird observed briefly in Los Angeles, LA, 17 May 2011 (2011-078) was incomplete and unconvincing. PINE WARBLER Setophaga pinus (103, 2). A male was at Laguna Niguel Regional Park, ORA, 12-13 Nov 2011 (RBMcNf; 2011-175), and a female was in Santa Maria, SBA, 2-3 Feb 2011 (JMC; 2011-039). Two males returned for a second winter: one to Palos Verdes Estates, LA, 31 Oct 2011-14 Feb 2012 (MBy; RJN, JEP; 2011-236), another to Hansen Dam, LA, 11 Nov 2011-16 Mar 2012 (KLGf; 2011-237; previously accepted as 2011-025 and 2010-159, respectively; Johnson et al. 2012). !|! YELLOW-THROATED WARBLER Setophaga dominica (139, 7). Singing males were at the Big Sur River mouth, MTY, 5 Jun 2010 (MiSf; 2011-045) and Vanden- berg Air Force Base, SBA, 20 Apr 2011 (JMC; 2011-055). In fall, one was at Pt. Reyes, MRN, 10 Oct 2009 (NSf; 2009-246), at Lake Merced, SF, 5 Sep 2011 (HCf; 2011-135), and at Santa Barbara I., SBA, 19 Oct 2011 (NAL; WTFf; 2011-176). One first found on the relatively early fall date of 13 Sep 2011 at Ferry Park in San Francisco, SF, remained through at least 30 Dec 2011 (JHyf; PBaf, TABef, DMof, DEQf, DSSf, LSt, JMr, MVef; 2011-142). It had a yellow spot on the lores and a relatively long bill, suggesting S. d. dominica, but a white chin, typical of the more 230 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS frequent subspecies S. d. albilora. A small amount of yellow on the lores is not unusual for S. d. albilora (Pyle 1997, Jaramillo 1993), making subspecific identification more challenging. McKay (2008) suggested synonymizing S. d. albilora, arguing that the morphological and genetic differences between albilora and dominica are weak. Also in winter was one at Park La Brea, LA, 30 Dec 2011 (BKf; 2011-238). An older record of one at Pt. Reyes, MRN, 10 Oct 2009 was also accepted (NSf; 2009-246). The committee reviews records of this species through 2011. GRACE’S WARBLER Setophaga graciae (61, 2). A fall migrant was at the Ocean Meadows Golf Course in Goleta, SBA, 5 Nov 2011 (DMC; 2011-224), and two winter records were accepted, of a first-year bird at Greenwood Cemetery in San Diego, SD, 16 Dec 2011-18 Feb 2012 (DBzf; PEL, GMcC, RBMcNf, BJSt, MSnt; 2011- 216) and an adult at Bella Vista Open Space in Goleta, SBA, 26 Oct 2011-26 Feb 2012 (DMCt; KRf; 2011-223), the latter returning for its fourth consecutive winter (2009-046, 2009-217, and 2010-179; Johnson et al. 2012). LE CONTE’S SPARROW Ammodramus leconteii (34, 0). IDENTIFICATION NOT ESTABLISHED: The description of a sparrow at the Palo Alto Baylands, SCL, 10 Nov 1976 (1976-501) suggested Le Conte’s Sparrow but was written several days after the observation and after field guides were consulted. The location is now known to host a small wintering population of Nelson’s Sparrow (A. nelsoni ), which was unknown at the time of the observation. There is only a single accepted record of Le Conte’s Sparrow for the mainland San Francisco Bay area (CBRC 2007, www. califor niabirds . org/ cbrc_book/ update . pdf) . PYRRHULOXIA Cardinalis sinuatus {21 , 2). A female was in Inyokern, KER, 7-8 May 2011 (SLS; KHL, CAMf, BStf; 2011-068), and an adult male was in a backyard in Niland, IMP, 5-12 Aug 2011 (TABef, MJBf, HBK, CAMf, DMxf, GMcC; 2011-111). TAINTED BUNTING Passerina ciris (107, 0). IDENTIFICATION NOT ESTAB- LISHED: A single-sentence description suggested an adult male was in Ventura, VEN, 17 Oct 1980 (2011-261) but was too brief to convince the committee, which discontinued reviewing records of this species after 2004. RUSTY BLACKBIRD Euphagus carolinus (24**, 6). An adult male at Bodega Bay, SON, 18 Nov 2010 (DBrf; 2010-166) caused pause for a few members because of its overall shiny black appearance — a characteristic usually associated with the Brewer’s Blackbird (E. cyanocephalus) . However, it had thin pale rusty edges to the tertials, back and upper breast and a faint black mask across the face with black lores, as well as structural differences including a slender bill — traits in which the adult male Rusty differs from Brewer’s. A female at Clear Lake Riviera, LAK, 23-28 Oct 2011 was a first for Lake County (VMit; JWh; 2011-164). Males were at Ridgecrest, KER, 5-14 Nov 2011 (KHLf, CAM, AShf, SLSt; 2011-167) and Galileo Hill, KER, 18 Nov 2011 (SLSf; 2011-181). A female was at the San Joaquin Wildlife Sanctuary, ORA, 26Nov2011 (BA; TABet, BEDf, VPf, SSof; 2011-190), and another spent much of the winter at Thousand Palms Oasis, RIV, 7 Dec 2011-26 Feb 2012 (GSf; TABet, CAMt, GMcC, CMcGt, SJMt; 2011-203). IDENTIFICATION NOT ESTABLISHED: The description of a supposed Rusty Blackbird at Ridgecrest, KER, 1 Nov 2011 (2011- 169) lacked confirming details. A report from the Woodlands Wastewater Treatment Plant, YOL, 28 Oct 2010 (2010-190) barely described what the bird looked like. COMMON GRACKLE Quiscalus quiscula (87, 3). Spring vagrants were in southeastern California at the River Lodge Resort along the Colorado River 6 km southwest of Parker Dam, SBE, 16 Apr 2011 (TABet; 2011-053) and Big Pine, INY, 4 Jun 2011 (JoHt; THt; 2011-083). A male in Cypress, ORA, 9 Nov 2011 (BEDt; 2011-213) represented a date typical for a fall vagrant. IDENTIFICATION 231 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS NOT ESTABLISHED: A male icterid photographed in Sierra County, 3 Jul 2011 (2011-196) did not show the typical coloration of subspecies versicolor of the Com- mon Grackle, lacking any apparent bronzy tones to the body or bluish tones to the head, and most committee members suspected the bird could be a Brewer’s Blackbird (. Euphagus cyanocephalus) x Great-tailed Grackle (Q. mexiccmus), a hybrid that has occurred in California before (Rogers and Jaramillo 2002). BLACK ROSY-FINCH Leucosticte atrata (16, 1). An adult male visiting a feeder in Aspendell, INY, 21 Feb 2011 was the sixth recorded at this location (BStf; JLD, CAM, JParf, DPa, SLS; 2011-016). All of California’s records are from Inyo and Mono counties. COMMON REDPOLL Acanthis flammea (81, 1). One that hit a window in June Lake, MNO, on 1 Dec 2011 (BaMf, KNNt; 2011-197, LACM #115935) was— at the time — unprecedented for California so early and so far south. This bird was also the first of an invasion of the Common Redpoll during the winter of 2011-2012 that was remarkable not only in the number of individuals but also in how far south many of them strayed. Most of these records will be covered in the 2012 annual report. CORRIGENDA It was recently brought to our attention that CBRC record 2004-124, of a Crested Caracara observed 20-24 Aug 2004 in Mendocino County, had been observed at two different locations over 60 km apart. Unfortunately, all publications referencing this record (Cole et al. 2006, CBRC 2007, and Nelson and Pyle 2013) mention only the initial location. The bird was first observed flying north near Manchester S. P. 20 Aug 2004, then seen and photographed feeding on a deer carcass in Westport 22-24 Aug 2004. If the bird followed the coastline, it flew over 75 km. The list of people who contributed observations listed in the committee’s 36 th report (Johnson et al., 2012) inadvertently omitted Diana L. Humple, Barbara E. Kus, Emilie Strauss, and Glen Tepke. DATA SOLICITATION Sightings for 2011 published in N. Am. Birds for which the CBRC has received no documentation are of up to 1 1 Trumpeter Swans near Durham, BUT, Jan 2011 (documentation of only four was submitted) (N. Am. Birds 65:335); a Yellow-billed Loon at Table Bluff, HUM, 3 May (N. Am. Birds 65:512); an Iceland Gull in Half Moon Bay, SM, 14 Mar ( N . Am. Birds 65:514); single first-cycle and second-cycle Lesser Black-backed Gulls on 9 Mar, and an adult 8-14 Mar, at Folsom Lake, PLA, and one along the American River Parkway, SAC, 12 Mar (N. Am. Birds 65:514); and a Long- billed Murrelet off Duncan’s Landing, SON, 4 Dec ( N . Am. Birds 66:341). We welcome submission of documentation for these reports. ACKNOWLEDGMENTS The committee thanks the following persons for advice on records summarized in this report: Mariko Parslow-Otsu, Masayuki Kurechi, Nial Moores, Martin Garner, Dan Brown, Thede Tobish, Tristan Reid, and Richard Millington for comments (and Leo Ohtsuki for a translation of Kurechi’s comments) on the bean-goose, Tiago 232 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS Rodrigues for comments on the Common Snipe, Ken Able for his thorough analysis of the Common Eider, Peter Pyle and Jim Tietz for analysis of Brown Booby records on Southeast Farallon L, Peter Pyle, Dave Compton, and Oscar Johnson for their analysis of Black Vulture records, Jason Rogers and Paul Lehman for their comments on the genus Plectrophenax, and Matt Heindel for his comments on Blue-headed Vireo identification. Peter Pyle compiled comments and photographs of specimens for identification of the Blue-headed Vireo and Winter Wren. Collections staff at the California Academy of Sciences (CAS), Western Foundation of Vertebrate Zoology (WFVZ), and Museum of Vertebrate Zoology (MVZ) helped the committee in various ways and we thank them for their aid. We extend special thanks to James R. Tietz for updating the table of records published in Rare Birds of California and to Joseph Morlan for maintaining the corrigenda to Rare Birds of California and for developing and updating the data query, all of which are available on the CBRC s website, www. http://californiabirds.org/. The following past and present CBRC members provided valuable comments on drafts of the manuscript: Jon L. Dunn, Kimball L. Garrett, Oscar Johnson, Paul E. Lehman, Guy McCaskie, Peter Pyle, Daniel S. Singer, and James R. Tietz. This report benefited considerably from reviews and comments by Alan Contreras, Daniel D. Gibson, Kurt Radamaker, and Philip Unitt. Finally, the CBRC would not exist without the cooperation of birders and ornithologists throughout California. We especially thank the following 265 people who contributed observations for records included in this report: Kenneth P. Able, Peter Adriaens, Douglas W. Aguillard, Bruce Aird, Richard Aracil, Noah Arthur, Patricia Bacchetti (PBa), Richard Barth, Alan D. Barron, David Batzler (DBz), David A. Bell, Thomas A. Benson (TABe), A. Bertke (ABe), Mark J. Billings, Thomas A. Blackman (TAB1), Len Blumin, Steven Brad (SBr), Abraham Borker (ABo), Matt Brady (MBr), Dan Brown (DBr), Philip Brown, Kenneth M. Burton, Martin Byhower (MBy), Dick Cabe, Phil Capitolo, Scott Carey (SCy), Barbara L. Carlson, Jeff Cartier (JCa), Scott Cartier (SCr), Jamie M. Chavez, Daryl Coldren (DCol), Luke Cole, David M. Compton, Chris Conard (CCo), Travis Cooper, Hugh Cotter, Rebecca F. Coulter, Bill Crowe, Mike Curry, Brian E. Daniels, David Davis, A1 DeMartini, Mark Dettling, Jochen Dierschke, Ryan DiGaudio, Dean DiTommaso (DDiT), Matthew Dodder, Jon L. Dunn, Todd Easterla (ToE), Mark Eaton (MEa), Tom M. Edell, Elias A. Elias, Megan Elrod (ME1), Dane Fagundes (DFa), Linda Farley, Jon S. Feenstra, Robbie Fischer (RFi), Jon Fisher, Mike Fisher (MFi), Rick Fournier, Rob Fowler (RFo), Mary Freeman (MFr), Nick Freeman, Gary Fregien, Wes T. Fritz, Dave Furseth (DFu), Peter A. Gaede (PAGa), John Garrett (JoG), Kimball L. Garrett, Steve Gerow, Greg Gillson, Brian Gilmore, Peter A. Ginsburg (PAGi), Cory Gregory, Charity Hagen (CHa), Frank Hall, Steve C. Hampton, Keith Hansen (KHa), Alvaro Jaramillo, Michael Harrison, Lauren Harter (LHa), Raymond Hasey, Karen A. Havlena, Jack Hayden (JHy), Floyd E. Hayes, Gjon Hazard (GjH), Jo Heindel (JoH), Tom Heindel, Kelli Heindel-Levinson (KHL), Doug Henderson (DHe), Jason Henderson (JHe), Ron Hirst (RHir), Alan S. Hopkins, Debbie House (DHo), Chris Howard (CHo), Rosie Howard (RoH), Andrew Howe, Lisa Hug, W. Terry Hunefeld, Eugene S. Hunn, Richard Hubacek (RHu), Jeff Jacobsen, Oscar Johnson, Tom Johnson, David Juliano, Alan Justice, A1 Kalin (AKa), Eric G. Kallen, Doug Karalun (DKa), Joel Karvonen, Robert J. Keiffer, Jay Keller, Barbara J. Kelly, Clay Kempf, David Kennedy (DKe), Howard B. King, Julie King (JKi), Pat Kirkpatrick, Andy Kleinhesselink, Peter Knapp (PKn), Paul Konrad (PKo), Alexander E. Koonce, Barbara Krause (BKr), Bhaskar Krishnamachari, Kenneth Z. Kurland, Tony Kurz, Dave Kutilek (DKu), Sarah Lane, Jerry M. Langham (JML), James Laughlin, Andrew B. Lazere, Rick LeBaudour (RLB), Greg Lavaty, Paul E. Lehman, Nick A. Lethaby, John F. Lockhart, Dan Lockshaw, Michael J. Mammoser (MJMam), Emily Mastrelli, Curtis A. Marantz, Marcie Mason, Dan Maxwell (DMx), M. J. Mazurek (MJMaz), Sean E. McAllister, Guy McCaskie (GMcC), Chet McGaugh (CMcG), Todd McGrath (TMcG), Dorothy McHaney (DMcH), Jim McHaney, David McIntyre (DMcI), 233 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS Jimmy M. McMorran (JMMc), Robert B. McNab (RBMcN), Anthony Mercieca (AnM), Ryan Merrill (RMe), Alex Merritt (AIM), Martin Meyers (MMe), Bartshe Miller (BaM), Bob Miller (BoM), Vivki Miller (VMi), G. Scott Mills, Joe Morlan (JMr), Dominik Mosur (DMo), Vic Murayama, Chris Murray (CMu), Roger Muskat (RMu), Jean Myers (JMy), Stephen J. Myers, Wendy Naruo, David W. Nelson, Kristie N. Nelson, Pornpat Nikamanon, Janice Nordenberg, Richard J. Norton, Gary Nunn, Nancy Overholtz, Matthew Page, Daniel D. Palmer, Michael Park, Debby Parker (DPa), Jim Parker (JPar), Jim Pawlicki (JPaw), Jim Pea (JPe), David Pereksta, Roger Perkins, Vincent Phan, James E. Pike, Linda Pittman, Bob Power, David E. Quady, Roger Radd (RoR), J. Leighton Reid, Robert W. Reiling, Sandy Remley, Richard Rideout, Steven M. Ritt, Deren Ross, Kerry Ross, Stephen C. Rottenborn, Richard Rowlett (RiR), Jonathan Rowley (JRw), Jethro Runco (JRu), Matt Sadowski (MSa), Mike San Miguel (MSanM), Larry Sansone, Paul Saraceni, Jesse Sargent (JSa), Adam Searcy (ASe), Ken Schneider, Brad K. Schram, Debra L. Shearwater, Alison Sheehey (ASh), Ginny Short, Brad Singer (BrS), Daniel S. Singer, James P. Smith, Jerry Smith (JSm), Susan E. Smith, Pete Sole, Craig Solgaard (CrS), Steve Sosensky (SSo), Lea Squires, B. J. Stacey, Mark E. Stacy, Justyn T. Stahl, Mike Stake (MiS), Trent R. Stanley, Bob Steele (BSt), Susan L. Steele, John C. Sterling, Mark Stratton (MSn), Noah Strycker, Steven Summers (SSum), Sherman Suter (SSut), Linda Swanson, Dan Tankersley, Glen Tepke, Christopher Taylor, Scott B. Terrill, Richard Thomas, James R. Tietz (JRT), Jerry Ting (JTin), Daniel Tinoco (DTi), Dorothy Tobkin (DTo), Richard Trissel (RTr), Leslie E. Tucci, David Vander Pluym (DVP), David E. Quady, Martijn Verdoes (MVe), Matt Victoria (MVi), Brad Waggoner, Len S. Warren, Lynn Watson, Bruce E. Webb, Elsa Weber, Douglas E. Weidemann, John West (JWe), Stan Walens, Jerry White (JWh), Phillip White, David Wimpfheimer, Gary Woods, Tim Worfolk, Connie Wulkowicy, Stan Wulkowicy, David G. Yee, and Bob Yutzy. LITERATURE CITED Aanerud, K. 2011. Eighth report of the Washington Bird Records Committee. Wash. Birds 11:35-55. American Ornithologists’ Union [AOU]. 1957. Check-list of North American Birds, 5 th ed. Lord Baltimore Press, Baltimore. Banks, R. C., Chesser, R. T., Cicero, C., Dunn, J. L., Kratter, A. W., Lovette, I. J., Rasmussen, P. C., Remsen, J. V., Jr., Rising, J. D., and Stotz, D. F. 2007. Forty-eighth supplement to the American Ornithologists’ Union Check-List of North American Birds. Auk 124:1109-1115. Bland, B. 1998. The Wilson’s Snipe on the Isles of Scilly. Birding World 11:382-385. Bland, B. 1999. The Wilson’s Snipe on Scilly revisited. Birding World 12:56-61. Brown, D. 2010. Identification and taxonomy of the bean geese. Birding World 23:110-121. Burton, K. M., and Anderson, L. S. 2013. First record of the Red-bellied Woodpecker in Nevada. W. Birds 44:148-150. California Bird Records Committee [CBRC] (R. A. Hamilton, M. A. Patten, and R. A. Erickson, eds.). 2007. Rare Birds of California. W. Field Ornithol., Camarillo, CA. Carey, G. J. 1992. The status and field identification of snipe in Hong Kong. Hong Kong Bird Report 1992, pp. 139-152. Carey G., and Olsson, U. 1995. Field identification of Common, Wilson’s, Pintail and Swinhoe’s Snipes. Birding World 8:179-190. Chesser, R. T., Banks, R. C., Barker, F. K., Cicero, C., Dunn, J. L., Kratter, A. W., Lovette, I. J., Rasmussen, P. C., Remsen, J. V., Jr., Rising, J. D., Stotz, D. F., and Winker, K. 2010. Fifty-first supplement to the American Ornithologists’ Union Check-list of North American Birds. Auk 127:726-744. Chesser, R. T., Banks, R. C., Barker, F. K., Cicero, C., Dunn, J. L., Kratter, A. W., 234 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS Lovette, I. J. , Rasmussen, P. C., Remsen, J. V., Jr., Rising, J. D., Stotz, D. F., and Winker, K. 2012. Fifty-third supplement to the American Ornithologists’ Union Check-list of North American Birds. Auk 129:573-588. Cole, L., Nelson, K. N., and Sterling, J. C. 2006. The 30 th report of the California Bird Records Committee: 2004 records. W. Birds 37:65-105. Delacour, J. 1951. Taxonomic notes on the Bean Geese, Anser fabalis Lath. Ardea 39:135-142 Dunn, J. L. 1988. Tenth report of the California Bird Records Committee. W. Birds 19:129-163. Dunn, J. L., and Alderfer, J. 2007. Featured photo: Wilson’s and Common Snipes. W. Birds 38:153-156. Force, M. P., Webb, S. W., and Howell, S. N. G. 2007. Identification at sea of Hawai- ian and Galapagos petrels. W. Birds 38:242-248. Glutz von Blotzheim, U. N., Bauer, K. M., and Bezzel, E. 1973. Handbuch der Vogel Mitteleuropas, vol. 5. Akademische Verlagsgesellschaft, Frankfurt, Germany. Heindel, M. T. 1996. Field identification of the Solitary Vireo complex. Birding 28:458-471. Heindel, M. T., and Garrett, K. L. 1995. Sixteenth annual report of the California Bird Records Committee. W. Birds 26:1-33. Howell, S. N. G. 2012. Petrels, Albatrosses and Storm-Petrels of North America. Princeton Univ. Press, Princeton, NJ. Iliff, M. J., McCaskie, G., and Heindel, M. T. 2007. The 31 st report of the California Bird Records Committee: 2005 records. W. Birds 38:161-205. Jaramillo, A. 1993. Subspecific identification of Yellow-throated Warblers. Birders Journal 2:160 Johnson, O., Sullivan, B. L., and McCaskie, G. 2012. The 36 th report of the California Bird Records Committee: 2010 records. W. Birds 43:164-188. Kurechi, M., Yokota, Y., and Otsu, M. 1983. Notes on the field identification of An- ser fabalis serrirostris and A. /. middendorffii. Tori 32:95-106 (in Japanese). Langham, J. M. 1991. Twelfth report of the California Bird Records Committee. W. Birds 22:97-130. Makatsch, W. 1970. Der Kranich (Grus grus u. a.), 2nd ed. Neue Brehm-Bucherei 229. Zeimsen Verlag, Lutherstadt, Germany. McEneaney, T. 2005. Rare color variants of the Trumpeter Swan. Birding 37: 148-154. McKay, B. 2008. Phenotypic variation is clinal in the Yellow-throated Warbler. Condor 110:569-574. Mlodinow, S. G. 2004. Bean Goose (Anser fabalis) at Hoquiam, Washington: A first state record. N. Am. Birds 58:298-300. Nelson, K., and Pyle, P. 2013. Distribution and movement patterns of individual Crested Caracaras in California. W. Birds 44:45-55. Oates, J. 1997. Identification of Taiga Bean Goose and Tundra Bean Goose. Birding World 10:421-426. Patten, M. A., and Erickson, R. A. 1994. Fifteenth report of the California Bird Records Committee. W. Birds 25:1-34. Patten, M. A., and Erickson, R. A. 2000. Population fluctuations of the Harris’ Hawk ( Parabuteo unicinctus) and its reappearance in California. J. Raptor Res. 34:187-95. Pike , J . E . , and Compton , D . M . 2 0 1 0 . The 34th report of the California Bird Records Committee: 2008 records. W. Birds 41:130-159. Pitman, R. L., and Jehl, J. R. Jr. 1998. Geographic variation and reassessment of species limits in the “Masked” Boobies of the eastern Pacific Ocean. Wilson Bull. 110:155-170. Pyle, P. 2008. Identification Guide to North American Birds, part II. Slate Creek Press, Point Reyes Station, CA. 235 THE 37 th ANNUAL REPORT OF THE CBRC: 2011 RECORDS Pyle, P., Tietz, J., and McCaskie, G. 2011a. The 35 th report of the California Bird Records Committee: 2009 records. W. Birds 42:134-163. Pyle, P., Webster, D. L., and Baird, R. W. 2011b. Notes on petrels of the Dark-rumped Petrel complex ( Pterodroma phaeopygia/ sand wichert sis) in Hawaiian waters. N. Am. Birds 65:364-367. Radamaker, K., and Corman, T. 2008. Status of Neotropic Cormorant in Arizona with notes on identification and ageing. Arizona Birds Online 3:6-11; http:// azfo . org/journal/volumes/ AZFO_Vol3_6- 1 0 .pdf . Rogers, J. 2005. Birding in Alaska: Identifying McKay’s Bunting. Birding 37:618- 626. Rogers, M. M., and Jaramillo, A. 2002. Report of the California Bird Records Com- mittee: 1999 records. W. Birds 33:1-33. Rottenborn, S. C., and Morlan, J. 2000. Report of the California Bird Records Com- mittee: 1997 records. W. Birds 31:1-37. Ruokonen, M., and Aarvak, T. 2011. Typology revisited: Historical taxa of the Bean Goose-Pink-footed Goose complex. Ardea 99:103-112. Ruokonen, M., Litvin, J., and Aarvak, T. 2008. Taxonomy of the Bean Goose-Pink- footed Goose. Molec. Phylog. Evol. 48:554-562. Sangster, G., and Oreel, G. J. 1996. Progress in taxonomy of Taiga and Tundra Bean Geese. Dutch Birding 18:310-316. Singer, D. S., and Terrill, S. B. 2009. The 33 rd report of the California Bird Records Committee: 2007 records. W. Birds 40:158-190. Sterling, J. C., and Easterla, T. B. 2012. First documented record of a Common Ringed Plover ( Charadrius hiaticulo ) for California. W. Birds 43:274-275. Towes, D. P. L., and Irwin, D. E. 2008. Cryptic speciation in a holarctic passerine revealed by genetic and bioacoustic analyses. Molec. Ecol. 17:2691-2705. Accepted 4 September 2013 236 NOTES RARE AND UNUSUAL BIRDS OBSERVED ON TERN ISLAND, FRENCH FRIGATE SHOALS, NORTHWESTERN HAWAIIAN ISLANDS PHILLIP J. HOWARD and SARAH C. HARVEY, Garcia and Associates, 2601 Mis- sion Street, Suite 600, San Francisco 94110; philliphoward85@gmail.com STEVEN M. RITT, 505 Upper Paw Paw Road, Marshall, North Carolina 28753 The Northwestern Hawaiian Islands are a series of 10 small islands and atolls that extend northwest for 2000 km from the main Hawaiian Islands (Pyle and Pyle 2009). The islands are part of the Hawaiian Islands National Wildlife Refuge, Papahanaumokuakea Marine National Monument, and provide a sanctuary for millions of nesting seabirds and thousands of wintering shorebirds (Pyle and Pyle 2009). Tern Island is located at the northwestern tip of the French Frigate Shoals and is the largest (~12 hectares) and only human-made island in the shoals. The Northwestern Hawaiian Islands’ location in the Pacific Ocean offers exceptional opportunity for assessment of patterns of avian migration and vagrancy. The is- lands lie within the trans-Pacific migration route of several shorebirds from Alaska that migrate long distances (Gill et al. 2005) and is within the influence of several seasonal macro-meteorolgical processes (e.g., Aleutian Low, Hawaiian High, and associated synoptic disturbances) that affect bird movement and migration along the North Pacific coasts (e.g., Christoforou and Hameed 1997). From June through December 2010, Howard and Harvey were stationed on Tern Island, monitoring seabirds as part of the U.S. Fish and Wildlife Service’s Northwestern Hawaiian Islands Nesting Seabird Monitoring Study. While on the island, we noted several rare and ususual species, including the first Fork-tailed Swift ( Apus pacificus) recorded in the French Frigate Shoals and first Japanese White-eye (Zosterops ja- ponicus) recorded in the Northwestern Hawaiian Islands. We first observed a Short-eared Owl ( Asio flammeus) roosting on the island on 14 June and saw it sporadically until 16 July (Figure 1). Short-eared Owls, including some collected and confirmed by size as the migratory subspecies A. f. flammeus (Pyle and Pyle 2009), have been reported regularly from the Northwestern Hawaiian Islands. Having obtained only photo documentation of the bird, we cannot exclude its having been A. f. sandwichensis, a slightly smaller and darker subspecies resident in the main Hawaiian Islands but thought by Pyle and Pyle (2009) to be absent from the Northwestern Hawaiian Islands. Our observation is unusual in that it is only the second in summer of nine previous records of this species from French Frigate Shoals, though there are scattered June records of this owl from Midway (Pyle and Pyle 2009). We observed two Least Terns ( Sternula antillarum) sporadically from 31 July through 12 August, when we found one dead. Unfortunately, we lacked facilities to preserve this specimen and the carcass had to be discarded. The remaining bird was last seen on 31 August. No breeding or nesting was observed. The Least Tern is an occasional nonbreeding visitor and rare breeding visitor to both the Northwestern and main Hawaiian Islands (Pyle and Pyle 2009). The similar Little Tern (S. albifrons) has also been documented on the Northwestern Hawaiian Islands, where it is also an occasional nonbreeding visitor and rare breeder (Pyle and Pyle 2009). The birds we observed showed no contrast between the gray back and uppertail coverts, as expected for the Least Tern (Figure 2), whereas the Little Tern shows a distinct contrast between the gray back and white uppertail coverts (Melgar 2001). Although currently undocumented on the Hawaiian Islands, Saunders’s Tern (Sternula saundersi ) differs from the similar Least Tern by its square white forehead patch and olive or reddish- Western Birds 44:237-242, 2013 237 NOTES Figure 1. Short-eared Owl (Asio flammeus) on Tern Island, French Frigate Shoals, Northwestern Hawaiian Islands, 14 June 2010, presumably the same individual seen sporadically until 16 July 2010. Photo by Sarah Harvey brown legs with any yellow being restricted to the feet (Melgar 2001). The birds we observed had completely yellow legs and the white forehead patch extended back into the black forehead, as in the Least Tern (Figure 2). Howard and Harvey observed a single adult White-tailed Tropicbird ( Phaethon lepturus) flying over the island on 12 August. The bird did not stop to forage or roost, and we took no photos. The combination of black outer primaries contrasting with white primary coverts, a relatively small bill, and fast snappy wingbeats distinguished this from other species of tropicbird (Harrison 1983). The White-tailed Tropicbird breeds in the main Hawaiian Islands but has only recently established itself as a breeding species in the Northwestern Hawaiian Islands, as a result of human settlement on Midway. Nonbreeding individuals disperse widely through the Pacific (Pyle and Pyle 2009). Amerson (1971) hypothesized that this species could nest on La Perouse Pinnacle in the French Frigate Shoals, but we saw none there during three one-day trips to the pinnacle that coincided with the species’ breeding season. 238 NOTES Figure 2. Least Tern (Sternula antillarum) on Tern Island, French Frigate Shoals, Northwestern Hawaiian Islands, 31 July 2010. Differs from the similar Little Tern (S. albifrons) by lack of contrast between the gray back and uppertail coverts and from Saunders’s Tern (S. saundersi) by the completely yellow legs and white forehead patch extending back into the black forehead. Photo by Sarah Harvey We saw a swift on 30 October as it flew over the island at sunset. Later that evening we found what was likely the same individual perched on the outside mesh covering of a window and from photographs we were able to identify it as a Fork-tailed or Pacific Swift ( Apus pacificus; Figure 3). The bird remained perched there overnight and was last seen on the morning of 31 October. This species is considered a very rare vagrant in the Hawaiian Islands, with no known observation on the main southeastern islands and only two records in the Northwestern Hawaiian Islands, both at Midway, in 1995 and 1999 (Pyle and Pyle 2009). The Fork-tailed Swift breeds from central Siberia east to Kamchatka, southern China, and the Malay Peninsula and winters south to Australia (Higgins 1999, Brazil 2009, DSEWPaC 2013). Though its main migration route is along the coast of Asia, there are casual records of this species as far north as the Bering Sea and Gulf of Alaska, in the Yukon, Canada, and south to the Marshall and Mariana islands (Gibson and Byrd 2007, Pyle and Pyle 2009). Almost all of these records, like ours, are from autumn after most birds have arrived on their nonbreeding grounds in Australia (Hig- gins 1999, Brazil 2009, DSEWPaC 2013). We first noted a Great Blue Heron ( Ardea herodias) on 4 October, observing it foraging until finding it dead on 6 November (Figure 4). Unfortunately, we lacked facilities to preserve this specimen and the carcass had to be discarded. The Great Blue Heron is considered a vagrant throughout the Hawaiian Islands, with very few observations on the main Hawaiian Islands and only one confirmed record (September 1988) in the Northwestern Hawaiian Islands, this also on Tern Island (Pyle and Pyle 239 NOTES Figure 3. Fork-tailed Swift ( Apus pad ficus) on Tern Island, French Frigate Shoals, Northwestern Hawaiian Islands, 30 October 2010. Photo by Phillip Howard 2009). The similar Gray Heron (A. cinerea ) has been recorded in the westernmost Aleutian Islands and the Pribilof Islands in Alaska, and in the Northwestern Hawaiian Islands subsequent to our observation (Maynard 2010). The Gray Heron, however, lacks the rufous thighs and leading edge of the wings characteristic of the Great Blue Heron (Figure 4). Howard and Harvey observed a single Japanese White-eye ( Zosterops japonicus) for about a minute on 15 November as it foraged among bushes. It flew off, not to be seen again, and no photos were taken during the observation period. The Japanese White-eye was first introduced to O’ahu in 1929 and has rapidly expanded its popu- lation and range since. It is now considered the most abundant landbird in the main Hawaiian Islands, but as of 2010 had not spread into the Northwestern Hawaiian Islands (Pyle and Pyle 2009). Our observation apparently establishes the first record for this species in the Northwestern Hawaiian Islands. Although in its native range northern populations of the Japanese White-eye are migratory, the specimen records indicate, and most literature assumes, that only the nonmigratory, nominate subspecies, Z. j. japonicus, occurs on the Hawaiian Islands (Pyle and Pyle 2009). However, records from far at sea (185-490 km south-southwest of Kaua'i) and on Johnston Atoll (1200 km from Kaua’i) indicate a tendency for individuals of the Hawaiian populations to disperse (Pyle and Pyle 2009). This study was funded by the U.S. Fish and Wildlife Service. The findings and 240 NOTES Figure 4. A carcass of a Great Blue Heron ( Ardea herodias), presumably the same individual found alive and photographed on Tern Island, French Frigate Shoals, Northwestern Hawaiian Islands, 4 October 2010. The rufous thighs and leading edge of the wings (right panel) distinguish this from the similar Gray Heron (A. cinerea ), which has also been recorded from the archipelago (see text). Photo by Phillip Howard, 6 November 2010 conclusions in this article are those of the authors and do not necessarily represent the views of the U.S. Fish and Wildlife Service. We especially thank Paul Lehman and Robert Gill for guidance and review of the manuscript. LITERATURE CITED Amerson, A.B.,Jr. 1971. The natural history of French Frigate Shoals, Northwestern Hawaiian Islands. Atoll Res. Bull. 150:1-383. Brazil, M. 2009. The Birds of East Asia. Christopher Helm, London. Breese, P. L. 1959. Information on Cattle Egret; a bird new to Hawaii. Elepaio 20:33-34. Christoforou, P, and Hameed, S. 1997. Solar cycles and the Pacific “centers of ac- tion.” Geophys. Res. Lett. 24:293-296. Department of Sustainability, Environment, Water, Population and Communities (DSEWPaC). 2013. Apus pad ficus, in Species Profile and Threats Database. Department of Sustainability, Environment, Water, Population and Communities, Canberra, Australia (www.environment.gov.au/sprat). Gibson, D. D., and Byrd, G. V. 2007. Bird of the Aleutian Islands, Alaska. Nuttall Ornithol. Club and Am. Ornithol. Union Series in Ornithol. 1. Gill, R. E., Jr., Piersma, T., Hufford G., Servranckx R., and Riegen A. 2005. Cross- ing the ultimate ecological barrier: evidence for an 11 000-km-long nonstop 241 NOTES flight from Alaska to New Zealand and eastern Australia by Bar-tailed Godwits. Condor 107:1-20. Harrison, P. 1983. Seabirds: An Identification Guide. Houghton Mifflin, Boston. Higgins, P. J. (ed) 1999. Handbook of Australian, New Zealand and Antarctic Birds, vol. 4: Parrots to Dollarbird. Oxford University Press, Melbourne, Australia. Maynard, B. 2010. Gray Heron — Alaska. American Birding Association Peeps Online (http : / /birding . typepad .com/peeps/2 010/0 5/gray-heron-alaska . html) . Melgar, C. 2001. Identification and occurrence of Least and Little Terns in Hawai’i. Birding Hawaii (www.birdinghawaii.co.uk/leasttern2.htm). Pyle, R. L., and Pyle, P. 2009. The Birds of the Hawaiian Islands: occurrence, history, distribution, and status, version 1. B. P. Bishop Museum, Honolulu (http://hbs. bishopmuseum . org/birds/ rip-monograph) . Swick, N. 2011. Rare bird alert: May 27, 2011. American Birding Association blog (http : / /blog . aba . org/ 2011/05/ rare-bird-alert-may-2 7-2011. html) . Accepted 3 July 2013 White-tailed Tropicbird Sketch by George C. West 242 NOTES A WINTER RECORD OF A LITTLE BUNTING FROM INTERIOR OREGON ALAN CONTRERAS, 1150 Darlene Lane, No. 171, Eugene, Oregon 97401; acontrer56@gmail.com CRAIG TUMER, 3439 NE Sandy Boulevard, PMB 231, Portland, Oregon 97223; craig@greatskua . com On 28 January 2013 we were birding in central Wallowa County, extreme north- eastern Oregon. Late in the afternoon we entered the town of Joseph in order to check the town’s feeders a second time that day. There was 3-4 inches of snow on the ground and birds were concentrated. As we checked one feeder from the car, Contreras noticed a sparrow-like bird in a tree and was astonished to see a Little Bunting ( Emberiza pusilla), a species he had seen only once before, at Gambell, Alaska. Turner agreed with the identification and got out of the car and obtained several diagnostic photos, the best of which is Figure 1 . Kyle Bratcher relocated the bunting later that day in an area of dense, leafless shrubbery near an active feeder, but it quickly disappeared. Trent Bray subsequently found the bird about a block away and obtained additional photos (see Oregon Birds 39[1]:3, 2013), and eventually Kendrick Moholt joined the four of us for reasonable views of the bird in dense shrubbery, where we obtained a few additional photos. The bunting was reported by a few observers, and missed by many others, during the following two days, but it was not found again thereafter until 23 March, when relocated by Adrian and Christopher Hinkle. It was still skittish, but a few people saw it through 25 March 2013, when it was last reported. This is the first record of the Little Bunting for Oregon and, to our knowledge, the first winter record and the first noncoastal record in North America. We made no major effort to age or sex the bird, as this is difficult to do in the field, and our photos show insufficient detail. The Little Bunting was first documented in North America on 6 September 1970 when one was found on a U.S. Coast Guard icebreaker in the Chukchi Sea, approxi- mately 150 miles off Icy Cape, Alaska (Watson et al. 1974). It now occurs casually but almost annually in fall in northwestern Alaska, with most reports at Gambell, St. Lawrence Island, between late August and early October (Howell et al. in press). A high count of 10 individuals was reported there in autumn 2007 (N. Am. Birds 62:134, 2008) . This species remains exceptionally rare elsewhere in western Alaska, where known in the Pribilof and Aleutian islands. There is only one spring record, of a single bird at Gambell 2-4 June 2008 (Radamaker and Powell 2010). The first North American record south of Alaska was of one photographed at Point Loma, San Diego County, California, 21-24 October 1991 (McCaskie 1993). Since then, single individuals have been found at Southeast Farallon Island, California, 27-28 September 2002 (Cole and McCaskie 2004) and 14 November 2012 (www. californiabirds.org), and at Rancho San Jose de Castro, Baja California Sur, Mexico, 8 October 2008 (Radamaker and Powell 2010). The date of 14 November represents the latest seasonal record for North America. The Little Bunting breeds widely across northern Eurasia, from northeastern Fen- noscandia (Byers et al. 1995, Svensson et al. 2009) east across northern Russia to Anadyrland in the Russian Far East (Flint et al. 1984, Gibson and Byrd 2007, Brazil 2009) . Migration is primarily through the eastern half of Asia, but a few vagrants reach western Europe and Iceland annually (Lewington et al. 1991, Howell et al. in press). The Little Bunting is a rare migrant offshore in Japan (Shimba 2007, Brazil 2009) and a vagrant to Wrangel Island (Portenko 1973), which lies at the western edge of the Chukchi Sea, north of the species’ breeding range in the Russian Far East. Western Birds 44:243-245, 2013 243 NOTES Figure 1. Little Bunting, Joseph, Wallowa Co., Oregon, 28 January 2013. Photo by Craig Turner Wintering is mainly from Nepal and northeast India across central Asia and south- ern Siberia to central coastal China and southern Taiwan (Byers et al. 1995, Brazil 2009); small numbers winter in Japan (Shimba 2007). Flint et al. (1984) reported the Little Bunting to prefer montane and nearby areas in winter. The Oregon record is therefore slightly farther north than the species normally winters, although perhaps in typical semi-montane habitat, as Joseph lies at an elevation of 4190 ft at the base of the 9900-ft Wallowa Mountains. It is worth noting that the western breeding population of this species makes a long eastward movement in fall, moving from Fennoscandia and northwestern Russia to eastern Siberia before turning southeast (Byers et al. 1995), a pattern roughly analo- gous to that of the westernmost Gray-cheeked Thrushes (Catharus minimus), which breed on the Chukchi Peninsula (Portenko 1973) and migrate primarily east across northwestern North America before turning south or southeast east of the Rockies (American Ornithologists’ Union 1998). The populations of the Little Bunting breed- ing in eastern Russia apparently do not follow such a pattern in fall, although they are much closer to North America and the probable source of records at Gambell as early in fall as late August, an established pattern of occurrence (Lehman 2005). The immediate source of a particular vagrant is, of course, speculative. 244 NOTES Thanks to M. Ralph Browning and Paul E. Lehman for helpful comments on a draft of this note. LITERATURE CITED American Ornithologists’ Union. 1998. Check-list of North American Birds, 7th ed. Am. Ornithol. Union, Washington, DC. Brazil, M. 2009. Birds of East Asia. Princeton Univ. Press, Princeton, NJ. Byers, C., Curson, J., and Olsson, U. 1995. Sparrows and Buntings: A Guide to the Sparrows and Buntings of North America and the World. Houghton Mifflin, Boston. Cole, L. W., and McCaskie, G. 2004. Report of the California Bird Records Com- mittee: 2002 records. W. Birds 35:2-31. Flint, V. E., Boehme, R. L., Kostin, Y. V., and Kuznetsov, A. A. 1984. A Field Guide to Birds of the USSR. Princeton Univ. Press, Princeton, NJ. Gibson, D. D., and Byrd, G. V. 2007. Birds of the Aleutian Islands, Alaska. Nuttall Ornithol. Club and Am. Ornithol. Union Series in Ornithol. 1 Howell, S. N. G., Lewington, I., and Russell, W. In press. Rare Birds of North America. Princeton Univ. Press, Princeton, NJ. Lehman, P. E. 2005. Fall bird migration at Gambell, St. Lawrence Island, Alaska. W. Birds 36:2-55. Lewington, I., Alstrom, P., and Colston, P. 1991. A Field Guide to the Rare Birds of Britain and Europe. HarperCollins (Domino), Jersey. McCaskie, G. 1993. A Little Bunting reaches California. W. Birds 24:95-97. Portenko, L. A. 1973. Birds of the Chukchi Peninsula and Wrangel Island, part II. Izadtel’stvo Nauka, Leningrad [English translation 1989, Amerind, New Delhi], Radamaker, K. A., and Powell, D. J. 2010. A Little Bunting reaches Baja California Sur. W. Birds 41:55-58. Shimba, T. 2007. A Photographic Guide to the Birds of Japan and Northeast Asia. Yale Univ. Press, New Haven, CT. Svensson, L., Mullarney, K., and Zetterstrom, D. 2009. Birds of Europe, 2nd ed. Princeton Univ. Press, Princeton, NJ. Watson, G. E., Angle, J. P, and Browning, M. R. 1974. First North American record of Little Bunting in eastern Chukchi Sea. Auk 91:417. Accepted 10 April 2013 245 BOOK REVIEWS Moments of Discovery: Natural History Narratives from Mexico and Central America, edited by Kevin Winker. 2010. University Press of Florida. 401 + xiii pp. Clothbound, $75.00. ISBN 978-0-8130-3417-1. ' Moments of Discover]; is a compilation of 20 autobiographic chapters by biologists who traveled in Middle America, mainly for ornithological field work, from the 1930s to the 1990s. The authors include big names in ornithology of this region: Miguel Alvarez del Toro (the only Mexican author), Robert F. Andrle, John M. Bates, Lula C. Coffey, Walter W. Dalquest, Robert W. Dickerman, Stephen W. Eaton, Ernest P. Edwards, John T. Emlen, Jr., Paul D. Haemig, Joyce Heck, Joe T. Marshall, Jr., Paul S. Martin, Don Owen-Lewis, A. Townsend Peterson, John H. Rappole, Charles G. Sibley, Walter A. Thurber, Dwain W. Warner, and Kevin Winker. This collection is intended to preserve a record of events that have gone mostly unrecorded in print. The majority of the travels described took place in Tamaulipas and Veracruz, with most others in central Mexico, Chiapas, and Central America; relatively few pages deal with field work in Baja California and Sonora. The black-and-white photographs included in some chapters are interesting. The map of Middle American states and countries inserted between the preface and chapter 1 should have been larger and should have indicated the location of places mentioned in each of the chapters; as it stands I do not think it is at all useful. I enjoyed reading this book for several reasons: because of the historical importance of many of these events in Middle American ornithology, because sometimes they document ecosystems now long gone, because sometimes the biological observations brought back pleasant memories or nicely described idiosyncrasies of Mexican culture, and because the descriptive styles of some chapters are works of art in themselves and the anecdotes are humorous and fun to read. I have a greater respect for some of these authors and a greater knowledge of their contributions to the study of Middle American natural history after reading these autobiographies. However, almost every chapter, the author’s own included, contains errors. In the preface Kevin Winker admits “Each chapter is given mostly as the author wrote it. My editorial hand has generally been light." I think this decision was unfortunate because it compromises the accuracy of the information. (I can provide a list of these errors in case Dr. Winker is interested in publishing a second, revised edition.) Alvarez del Toro’s beautifully written chapter recounting his first trip to El Triunfo is an English translation of an account he published in his 1990 autobiography /Asf Era Chiapas!: 42 Ahos de Andanzas por Montanas, Selvas y Caminos en el Estado, published by the John D. and Catherine T. Mac Arthur Foundation. The translation is very good, except that both mentions of goldfinches at El Triunfo refer to the Brown-backed Solitaire (. Myadestes occidentalis) — in Mexico jilguero refers to the Brown-backed Solitaire but in Spain to the European Goldfinch (Carduelis carduelis). There are a couple of typos, and the date of the trip is lacking. Dickerman’s chapter has a surprising typo, with the name of a subspecies of the Red-winged Blackbird ( Agelaius phoeniceus arthuralleni) that he named himself misspelled (p. 91)! In most of the other chapters, by far the majority of errors are misspellings of Spanish names or words, extremely disappointing for a book on Latin America. This is one more of many books with total disregard for getting Spanish words and names right. It is beyond me why the authors and editor would not have the courtesy or try harder to be culturally and editorially accurate. Another source of error is scientific inaccuracy. There was an attempt to provide the current (in 2010) English and scientific names of species the first time they are mentioned in each chapter. This worthy goal was usually, but not always, achieved. 246 Western Birds 44:246-249, 2013 BOOK REVIEWS Examples of this problem plagued Eaton’s chapter: both the Boat-tailed Grackle (Quis- calus major) and Great-tailed Grackle (Q. mexicanus) are mentioned in Guatemala; there is a potentially confusing section mentioning both Passerini’s ( Ramphocelus passerinii) and Scarlet-rumped Tanagers when, to be precise, the author was writing about Cherrie’s Tanager ( R . costaricensis); “a large, spotted rodent” is named “an agouti (. Dasyprocta ),” when if it had a spotted coat it must have been a paca (Agouti); an “ Elaenia ” mentioned in the highland forest of “Sierra Colorado,” (actually Cerro Colorado, Hidalgo) is clearly an error (the Greenish Elaenia, Myiopagis viridicata, is possible there, though); and both the Cactus Wren (Campylorhynchus brunneicapil- Ius) and Chestnut-capped Brush-Finch (Arremon brunneinucha ) are mentioned near Hutizilac, Morelos, when it is actually related species that occur there. Edwards’s chapter well describes the Tawny-collared Nightjar (Antrostomus [in 2010 Caprimulgus } salvini) but does so under the names “Buff-collared Nightjar (■ Caprimulgus ridgwayi).” An endnote to Martin’s chapter mentions collecting an Eastern Towhee in Tamaulipas, but this refers to specimen Denver Museum of Natural History 48875, a Spotted Towhee. Sibley’s chapter mentions collecting a pair of Aplomado Falcons ( Falco femoralis) near Rancho Guirocoba, Sonora; this would be shocking if correct. I looked it up through www.ornisnet.org: the falcon that Sibley collected near Rancho Guirocoba is a Bat Falcon ( F. rufigularis) (which makes more biogeographic sense). Sibley’s chapter also seems to treat maculatus and erythrophthalmus as interchangeable names for the Spotted Towhee, which they are not (but nevertheless the summary of Sibley's findings on Pipilo systematics in Mexico is absolutely fascinating). In sum, this book provides fun and absorbing reading for those interested in the historical, cultural, and biological aspects of Middle American travels by earlier natural- ists, though it would have benefited greatly from a heavier editorial hand. Hector Gomez de Silva The Travails of Two Woodpeckers: Ivory-Bills & Imperials, by Noel F. R. Snyder, David E. Brown, and Kevin B. Clark. 2009. University of New Mexico Press, xiv +170 pp., 27 color photographs, 21 halftones, 2 maps, 2 tables, 1 chart. Clothbound, $34.95. ISBN 978-0-8263-4664-3. The Travails of Two Woodpeckers examines the historical perspectives and the apparent ultimate extermination of two of the most iconic bird species in North America. The authors thoughtfully weave an intriguing story for each of the species’ demise, replete with historical accounts, natural-history accounts, along with conserva- tion theory and endangered-species management. The book has been broken up into three main chapters, the first focusing on the Ivory-billed Woodpecker (Campephilus principalis ), the second on the Imperial Woodpecker (C. imperialis), and the third addressing conservation issues, past and present. In the account of the demise of the Ivory-billed Woodpecker, we are taken through the history of the species’ discovery and description, the description of its preferred habitat, and some of its population dynamics and life history. By the time the alarm bell for the Ivory-billed Woodpecker sounded, because of a significant population decline in the early the 20 th century, only limited information existed on crucial life-history requirements. One of the most respected naturalists of the era was James Tanner, who believed that habitat loss was the most compelling reason for the Ivory-bill’s decline in the U.S. Tanner believed that the Ivory-bill was a specialist in foraging, requiring huge tracts of virgin bottomland forest to find sufficient food. The authors believe that this assertion by Tanner started the perceived misbelief that sustained logging 247 BOOK REVIEWS of primary forest was the primary cause for the decline of the species. The authors of The Travails of Two Woodpeckers make a compelling alternative argument, that the true driver of the of the species’ ultimate demise was instead unrelenting depre- dation in the form of collecting for museums, subsistence hunting, curiosity hunting, and shooting through ignorance. They believe that the most crucial negative effect of timbering on the woodpecker may not have been to deplete its food supply but to facilitate an increase in shooting: lumber roads improved access to forests and led to substantially increased numbers of people in the forests. The Ivory-billed Woodpecker in Cuba (possibly a distinct species?) faced a similar threat, although that population managed to persist into the late 20 th century. The Imperial Woodpecker of the Sierra Madre Occidental of Mexico was the larg- est woodpecker in the world. It was recorded less than 40 miles south of the U.S. border in Chihuahua near the Mesa de Guacamayas, south and east of the Chiricahua Mountains, Arizona. The first specimens were collected by a contract hunter in the Sierra Bolanos in the state of Jalisco, and were shown to the ornithological world at the Zoological Society of London in 1832. They were the size of ravens, nearly two feet in length! Unlike the Ivory-bill, it was noted early on that the Imperial was hunted regularly by indigenous peoples as well as by the Mexicans at the time. Carl Lumholtz wrote (p. 71), “The giant woodpecker is seen in the more remote parts, but it is on the point of being exterminated, because the Tarahumares consider his one or two young such a delicacy that they do not hesitate to cut down even large trees to get at their nests. The Mexicans shoot them because their plumage is thought to be beneficial to health. It is held close to the ears and the head in order to impart its supposed magnetism and keep out the maleficent effects of the wind. In the pairing season these birds keep up a chattering noise, which to my ears was far from dis- agreeable, but very irritating to a Mexican whom I employed. He used to shoot the birds because they annoyed him.” Indeed, a local nickname of the bird was “un gran pedazo de came” (a great piece of meat). Although habitat destruction in the form of intensive logging certainly degraded the Imperial Woodpecker’s preferred habitat, Snyder, Brown, and Clark present a convincing argument that human depredation was actually the driving force behind the species’ decline. While visiting Tutuaca in southern Chihuahua on a WFO-organized field trip in 2003 (a site also visited by the authors in preparation for this book), I was able to listen to an elder Mexican in the remote ejido telling of his childhood and the Imperial Woodpecker. From my field notes on that trip, “Don Bernardino stated that at the age of 15 (he was born in 1947), it was not uncommon for him and his father to see between 1 and 3 Imperial Woodpeckers ( Campephilus imperialis ) in one day while on their way out to the main highway which links Hermosillo and Chihuahua City (Hwy. 16) while getting provisions for their house. That places the dates at around 1962 when Bernardino saw the birds. He mentioned that the early populations of the Thick-billed Parrots were larger, and that the individual flocks were much more extensive than they are now, with groups of 50-60 seen commonly during the day. He (Bernardino) has huge concerns that their numbers may continue to decrease in much the same way that the population of Imperial Woodpeckers did.... He now only sees small flocks of parrots daily, and he does not want this bird to disappear as the ‘Carpintero imperial’ did. It is sobering and humbling to speak with a man who has seen a species of bird go extinct in his backyard while ‘on his watch.’” The last section of the book presents an overview of the demise of the two wood- peckers, with valuable insights on conservation biology and lessons learned during their research for the book. The most intriguing idea the authors present is the warn- ing that sometimes explanations for declines of endangered species become widely accepted before alternative explanations are examined and the original hypothesis is fully tested. Snyder certainly has a great deal of experience with this tenet, and he and 248 BOOK REVIEWS his co-authors made it clear that the consequences of misidentification of causes of declines and lack of identification of limiting factors can mislead conservation efforts and waste initiative and money. Overall I was impressed with this book, with both the historical accounts of the species and the analyses of their declines, as well as with the conservation biology presented throughout. I recommend this book for those interested in learning more about these two charismatic species, as well as those who have an interest in biological investigation and conservation biology. It’s a good read, and one which will give WFO members close to northern Mexico a reason to dream about what could have been. Dave Krueper Wing your way to... SAN DIEGO, CALIFORNIA 9-12 OCTOBER 2014 The 39th annual conference of the Western Field Ornithologists will be held in San Diego, California, 9-12 October 2014, at the Marriott Courtyard hotel at Liberty Station, less than 1 mile from the San Diego airport and Fisherman’s Landing, from which our pelagic trips will depart. We are still in the early stages of planning for this meeting, but expect a wide diversity of presentations, workshops, social events, and full-day and half-day field trips, both terrestrial and pelagic. The conference will include a symposium on avifaunal change in western North America to be published subsequently as a volume of WFO’s monograph series Studies of Western Birds. We welcome and solicit contributions to this symposium and peer-reviewed publication. To inquire about participating, please contact Dave Shuford at dshuford@pointblue.org or Bob Gill at rgill@usgs.gov. Please watch www.westernfieldornithologists.org/conference.php for details as they become available. But plan now to join us next year in San Diego! 249 FEATURED PHOTO DARK-FACED COMMON MURRES OFF CENTRAL CALIFORNIA IN FALL AND WINTER PETER PYLE, The Institute for Bird Populations, P. O. Box 1346, Point Reyes Sta- tion, California 94956; ppyle@birdpop.org ABSTRACT : While most basic-plumaged Common Murres (Uria aalge) show white facial plumage and tips to the secondaries, some individuals off central California from September through November have dark faces and secondaries. To investigate the plumage state and age of such birds, I examined specimens and photographs of Common Murres taken off central California and conclude that outgoing and in- coming alternate feathers as well as early-replaced and late-replaced dark basic and formative feathers all may contribute to this variation. Hyper melanism may account for a small proportion of cases. The early onset of prebasic molt — before hormone signaling switches from dark feathers typical of the alternate plumage to white feath- ers typical of the basic plumage — may be responsible for most of the darkest-faced birds in basic plumage. This type of asynchrony of molt and the signal for change in plumage appears to occur more often in second-cycle birds, which initiate the prebasic molt earlier than do older adults, and it may also explain plumages in chicks and juveniles molting during May and June. Common Murres in basic plumage from colonies in central California appear to acquire dark facial feathers more often than birds from more northerly colonies, which could relate to earlier breeding and molt in central California. The definitive prealternate molt of Common Murres ( Uria aalge) breed- ing in central California can commence as early as November, rather than extending from January through April as in most other alcids and waterbirds (Ainley et al. 2002, Pyle 2008, 2009). The definitive prebasic molt in these murres occurs just after breeding, from July through September, at which time they lose the fully dark-headed alternate plumage and typically gain white basic feathering in the face, including the sides of the nape, posterior portion of the auriculars, subauricular area, chin, and throat. Because of the early start of the prealternate molt, full basic plumage may therefore be worn for as little as one to two months from September to November, as is the case in the more southern populations of the Common Murre breeding in the Atlantic (Harris and Wanless 1990). Populations of the Common and Thick-billed ( U . lomvia) murres that breed farther north undergo prealternate molt later and can retain full basic plumage into March (Ainley et al. 2002, Gaston and Jones 1998, Pyle 2008). The earlier timing of the prealternate molt in central California populations may reflect the earlier breeding cycle of these birds, as adults begin occupying colonies as early as late October (Ainley and Boekelheide 1990), much earlier than at northern colonies (Gaston and Jones 1998, Ainley et al. 2002). Although most basic-plumaged Common Murres show white faces, indi- viduals with partially or completely dark faces occur off central California from September through November. The upper image on this issue’s outside back cover shows a typical white-faced Common Murre in basic plumage; the lower shows a partially dark-faced individual. Both were photographed 250 Western Birds 44:250-261, 2013 FEATURED PHOTO on 7 October 2012 near Cordell Bank off Bodega Bay, California. A dark face has been used as a field mark to distinguish the Thick-billed from the Common Murre in basic plumage (Roberson 1980, Stallcup 1990, Gaston and Jones 1998), so the occurrence of dark-faced basic-plumaged Common Murres off central California needs to be further understood, and other field marks are needed to distinguish the two murre species (CBRC 2007). Dark-faced adult Common Murres in fall could result from any or all of three molt/plumage scenarios: (1) a late or incomplete definitive prebasic molt, resulting in worn retained dark alternate feathers in the face; (2) an early definitive prealternate molt, resulting in fresh or incoming dark alter- nate feathers in the face; and/or (3) definitive basic feathers in the face that are dark or partially dark. Similarly, first-cycle murres might show dark or partially dark rather than white formative feathers in the face. To investigate these possibilities, I examined 221 specimens of post-juvenile Common Murres collected in California and housed at the California Academy of Sciences (CAS; 168 specimens) and Museum of Vertebrate Zoology (MVZ; 53 specimens). I also observed and photographed Common Murres during nine pelagic field trips off central California between Fort Bragg and Half Moon Bay from August through November in 2011 and 2012. METHODS I used the presence of emerging or pin feathers to indicate active molt in the faces of murre specimens. For each specimen in full basic or formative plumage (in which evidence of molt was lacking), and for murres observed in the field, I scored the face as (1) entirely or almost entirely white, (2) primar- ily white with sparse dark flecking, (3) primarily white with moderate dark flecking, or (4) primarily dark with some white feathering (Figure 1). The amount of white on the tips of juvenal and basic secondaries also varied; on specimens I scored this trait as (1) wide, (2) moderate, (3) thin, or (4) absent (Figure 2). I categorized specimens as first-cycle or adult on the basis of dif- ferences in color of the primary coverts, bill size and color, and bill width at the gape (Pyle 2008, 2009; Figure 1); in the field, I aged murres by bill size and color. Second-cycle murres could not be aged by plumage after comple- tion of the second prebasic molt. However, I inferred that specimens with bill widths at the gape in the bottom quarter of the range for birds after their first cycle (19.6-20.7 mm of the full range 19.6-23.8 mm; Pyle 2009) were in their second cycle. This approach is consistent with results of bill-width data from other alcids that can be confirmed as second-cycle by plumage or other features. Sex was assumed from information on specimen labels. RESULTS Among adult specimens of the Common Murre from central California, 36 were collected while undergoing definitive prebasic molt. These birds had worn, dark alternate head feathers mixed with incoming white feathers. They ranged in date from 8 July (CAS 10020 with a few incoming white basic feathers) to 2 October (CAS 15796 with a few remaining alternate feathers). In addition, a one-year-old individual (MVZ 145340) collected from 251 FEATURED PHOTO Figure 1 . Four categories of variation in the facial plumage of the Common Murre in central California: (1) entirely or almost entirely white (CAS 43160, 12 February 1901), (2) primarily white but with sparse dark flecking (CAS 15784, 4 October 1909), (3) primarily white but with moderate dark flecking (CAS 16082, 13 December 1909), and (4) primarily dark but with some white (CAS 16078, 22 November 1909). Left, side view, and right, ventral view of the same specimens. All specimens were collected in Monterey Bay. CAS 15784 is in formative plumage; the other three specimens are in basic plumage. Carmel Bay on 16 May was molting in white facial feathers with black tips (Figure 3A and B), as was another one-year-old, also collected 16 May, from Admiralty Island, Alaska (MVZ 69). Concurrent replacement of back, rump, and underpart feathers, as well those of the head, indicated that they were undergoing the second prebasic rather than the first prealternate body molt. Thirty specimens of adults had been collected while undergoing defini- tive prealternate molt of the head, with the incoming feathers dark. The range of dates for these was 8 November (CAS 16077 with a few incoming alternate feathers) to 17 January (MVZ 75130 with a few basic feathers remaining and being replaced). The range of dates for 64 adults collected in fully dark (alternate) head plumage was 9 December (CAS 16070) to 27 August (CAS 15478 and 15482). I also observed a Common Murre with fully dark alternate head plumage off Half Moon Bay on the earlier date of 22 Nov 2011, the only such individual of more than 100 murres I scrutinized on this date. Two murres collected in June and July (CAS 68 and 75127) showed fully dark heads except for white tipping to some of the alternate face feathers (Figure 3C and D). Sixty-one specimens of basic-plumaged adults, without outgoing or 252 FEATURED PHOTO Figure 2. Four categories of variation in the width of white tips on the secondaries of the Common Murre in central California: (1) wide (CAS 15784, 4 October 1909), (2) moderate (CAS 16072, 26 November 1909), (3) thin (CAS 15793, 15 October 1909), and (4) absent (CAS 16078, 22 November 1909). All specimens were collected in Monterey Bay. incoming alternate feathers, had been collected in California between 16 September (CAS 15786) and 15 March (CAS 10024, after second cycle). Other than the latter specimen and those collected 12 February (CAS 43160, after second cycle; Figure 1) and 24 February (MVZ 17756, second cycle), all basic-plumaged murres were collected between mid-September and early January. Twenty-nine first-cycle birds in formative plumage (lacking juvenal or, apparently, first alternate body feathers) were collected between 18 August (CAS 48176) and 9 April (MVZ 145352). Birds in complete basic and formative plumage showed varying amounts of dark in the face (Figures 1 and 4; images on inside and outside back cover). In all cases, the dark coloration resulted from white-based basic feathers with dark tips of varying width (Figure 1). By contrast, alternate feathers were entirely dark. Murres with facial-plumage scores of 2 or 3 typically showed white on the sides of the nape and in the auricular and malar regions but dark-tipped throat feathers. By contrast, those with a plumage score of 4 typically displayed dark on the sides of the nape and in the auricular and malar regions but whiter throats (Figures 1 and 4). The mean facial score was greater in basic-plumaged (2.26, n = 61) than in formative-plumaged (1.76, n = 29) Common Murre specimens (Figure 5; ANOVA, P = 0.006). Facial-plumage scores of specimens were similar to 253 FEATURED PHOTO Figure 3. Variation in the molts and plumages of the Common Murre in central California. A (side view) and B (ventral view): MVZ 145340 (Carmel Bay, 16 May 1925), collected during presumably the second prebasic molt, with incoming feathers dark-tipped. C (side view: CAS 75127 (25 July 1903, San Francisco Bay) and D (ventral view): CAS 43166 (Monterey Bay, 19 June 1897), showing white tipping to alternate face feathers. E and F: CAS 18072 (Southeast Farallon Island, 30 May 1911), showing completely dark plumage. those for both basic-plumaged (n = 209) and formative-plumaged (n = 100) murres scored on four pelagic trips off Half Moon Bay and Bodega Bay between 30 September and 6 November, when few or no alternate feathers should be expected on the basis of specimen data (Figure 5; P = 0.30 and 0.78, respectively). By contrast, both basic-plumaged (n = 55) and formative-plumaged (n = 31) murres scored off Fort Bragg in northern California on 6 November 2011 showed significantly less dark in the face than did specimens and live birds scored off central California (Figure 5 ;P< 0.007 for each of four comparisons); none of the murres off Fort Bragg had facial scores of 3 or 4. Among specimens in basic plumage, faces of females averaged slightly darker than those of males (mean score: female 2.35, n = 26; male 2.15, n = 34), but this difference was not significant (P = 0.432). The amount of white on the tips of the secondaries averaged less in basic-plumaged (mean score 1.66, n = 61) than in formative-plumaged (mean score 1.41, n = 29) specimens, but this difference was not significant (ANOVA; P = 0.084). As with facial score, the sexes had a similar amount of white in the secondaries (P = 0.278). Among basic-plumaged birds, the mean score for white in the secondaries was 1 .07 for facial score 1 (n = 15), 254 FEATURED PHOTO Figure 4. Dark-faced Common Murres (facial score 4) collected on Monterey Bay. From top to bottom, CAS 15788 (formative plumage, 25 October 1909), CAS 15792 (basic plumage, 15 October 1909), CAS 16071 (basic plumage, 26 November 1909), and CAS 16078 (basic plumage, 22 November 1909). Left, side view, and right, ventral view of the same specimens. Bill-width measurements indicated that 15792 and 16071 were in second basic plumage, whereas 16078 was an older bird. The extent of white in the secondaries (see Figure 2) in these four specimens was absent, thin, thin, and absent, respectively. 1.27 for facial score 2 (n = 23), 1.50 for facial score 3 (n = 14), and 2.78 for facial score 4 (n = 9), indicating that murres with darker faces also showed less white in the secondaries (linear regression, P < 0.001). The mean secondary score for birds with facial score 4 was also significantly greater than those of each of the other three groups (ANOVA, P < 0.001), while secondary scores of the other three groups did not differ from each other (P > 0.098). Most basic-plumaged murres with facial scores of 1 through 3 showed wide or moderate white secondary tips (only one specimen in each class showed thin tips), whereas of 8 murres with facial score 4, none showed wide white tips, four showed moderate tips, two showed thin tips, and two showed no white tips (CAS 15978 and 16078; Figure 4; see also lower photo on outside and upper photo on inside of this issue’s back cover). Similarly, formative-plumaged murres with facial scores 1 through 3 all had wide or moderate white tips to the secondaries, whereas the one specimen with a facial score of 4 (CAS 15788) showed no white in the secondaries (Figure 4). Specimens of murres with a facial score of 4 also appeared to average darker in the flanks and underwing coverts than did those of scores 1 to 3 (Figures 1 and 4; images on inside and outside back cover), but I did not quantify these characters. Eight basic-plumaged specimens and one formative-plumaged specimen 255 FEATURED PHOTO Basic Plumage E O r o Q. o £ O r o Q. O 100 80 60 40 20 0 100 80 60 40 20 0 I I I 1 2 3 Formative Plumage I CM JSL Facial Plumage Score □ Specimens □ Central CA □ Fort Bragg Figure 5. Distribution of facial scores (see Figure 1) of Common Murres in basic and formative plumage on the basis of specimens and birds studied in the field. The latter were scored between 30 September and 6 November, when birds should have few or no alternate feathers (see text). Birds observed on four trips from Bodega Bay to Half Moon Bay (30 September-6 November, 2011 and 2012) and on one trip off Fort Bragg (6 November 2011) analyzed separately. See text for statistical comparisons. Sample sizes for basic plumage: 61 specimens, 209 live birds off central California, and 55 live birds off Fort Bragg; for formative plumage: 29 specimens, 100 live birds off central California, and 31 live birds off Fort Bragg. received a facial score of 4 (Figure 4), representing 13% of basic-plumage specimens and 3% of formative-plumaged specimens. On the four field trips off central California between 30 September and 6 November, 19 of the 209 (9%) basic-plumaged birds but none of the 100 formative-plumaged birds scored had dark faces. By bill width, live of the eight (63%) dark-faced specimens in basic plumage were in their second cycle (cf . Figure 4), whereas only 19 of 58 (33%) birds with facial scores 1 to 3 were in their second cycle; overall, faces were significantly darker in second-cycle birds (mean 2.63) than in older birds (mean 2.03; ANOVA, P = 0.019). A dark-faced bird in basic plumage photographed in Monterey Bay 15 September 2012 (inside back cover, top image) appears to have a medium-small bill, also sug- 256 FEATURED PHOTO gesting a second-cycle bird. On the other hand, another dark-faced adult, photographed 2 October 2012 off Half Moon Bay (inside back cover, bottom photograph) had a large bill and was attended by a young bird in formative plumage, suggesting it was a male at least four years of age, on the basis of minimum age at first breeding (Gaston and Jones 1998, Ainley et al. 2002); breeding males rather than females attend their young after fledging. Finally, of 33 specimens of juvenile murres at CAS and MVZ, three (9%) showed dark faces, while all five murres collected as chicks showed dark natal down in the face (Figure 6), corresponding to descriptions of the natal down of Common Murre (Gaston and Jones 1998, Ainley et al. 2002). DISCUSSION On the basis of this study, dark-faced and partially dark-faced Common Murres collected and observed off central California (Bodega to Monterey bays) during the fall may be completing their prebasic molt (September), initiating their prealternate molt (November), or in full basic or formative plumage (September through November). Of murres in full basic or forma- tive plumage, most have white or mostly white faces. A small proportion, however, perhaps 10-15% of those in basic plumage and <5% of those in formative plumage, have mostly dark faces. Many of these also have little or Figure 6. Variation in the head plumage of chicks and juveniles of the Common Murre. The nestling (top, CAS 88095) was collected on Southeast Farallon Island 7 June 1964; the juveniles (middle, CAS 15545; bottom, 15546) were collected on Monterey Bay 19 August 1909. CAS 88095 shows the pattern typical of natal down; CAS 15545 was one of only three of the sample of 33 juveniles showing dark auriculars. 257 FEATURED PHOTO Figure 7. Completely dark Common Murre with two in typical alternate plumage, photographed 24 July 2013, 14 km southeast of the Farallon Islands, flying south. That the dark bird was carrying a fish indicates it was breeding, perhaps on the Farallon Islands. Photo by Dru Deulin NOAA/ONMS/ACCESS no white on the trailing edge of the secondaries and appear to have darker underparts. Field observers should beware that such dark-faced birds can be confused with Thick-billed Murres. The distribution of dark and white in the face and secondaries suggests that the mechanisms resulting in the darkest-faced birds (facial score 4) may be different than those that result in the whiter-faced categories (facial scores 1 to 3). Dark formative or basic feathering in the faces of Common Murres could result from asynchrony of the cycles of molt and deposition of melanin in the growing feathers. When a molt is early or late relative to the hormone cycle that signals pigment deposition, the resulting feathers may resemble those associated with earlier or later plumages (Pyle 1997, 2008, Howell 2010). Common Murres undergoing prebasic molt early, for example, might begin replacing feathers before hormone signaling switches the color from dark to white. Such birds might be expected to show more dark in the nape and auriculars and more white in the chin and throat, since within a molt dorsal head feathers are often replaced prior to ventral feathers (Pyle 2008). The darkest-faced murres (facial score 4) showed such a pattern, having completely or largely dark napes and auricular regions and whiter chins and throats (Figure 4; photos on inside and lower outside of back cover), suggest- ing that they initiated their prebasic molt early, before hormone signaling had switched from dark to white. A similar mechanism may result in reduced or no white tips to the secondaries and darker underparts and underwing coverts in these birds. Being unconstrained by breeding, one-year-old alcids typically undergo the second prebasic molt earlier than older birds undergo subsequent prebasic molts (Pyle 2009), and this difference might account for the high proportion of darker-faced second-cycle birds I observed. For 258 FEATURED PHOTO example, the one-year-old murres collected 16 May (MVZ 145340 and MVZ 69) appeared to be undergoing early second prebasic molts that would have resulted in facial scores of 4 (Figure 3 A and B). Older birds with darker faces (Figure 4) may represent 2- to 4-year-olds that had not bred yet or breed- ers that failed early the previous year, which could result in earlier prebasic molt. Possibly, this pattern of pigmentation extends to nestlings and chicks, resulting in darker-faced downy chicks and juveniles that hatched early (Fig- ure 6). Some breeding birds also appear to show dark faces (lower inside back cover); another possible explanation for dark faces in these and other birds might be a later-than-normal signal for deposition of melanin imposed upon typical timing of prebasic molt. Although Ainley et al. (2002) reported females to molt earlier than males, I found only a slight and nonsignificant trend for females to show faces darker than those of males. Murres undergoing a late prebasic molt, on the other hand, might start acquiring partially dark faces as hormonal signaling switches back to produc- ing feathers typical of alternate plumage. Because molt of ventral tracts is later (see above), these birds might show white napes and auricular regions but partially dark chins and throats. Most murres of intermediate categories (facial scores 2 or 3) showed this pattern (Figure 1), indicating they may have undergone a late prebasic or preformative molt the previous fall, and/ or that hormone signaling for dark feathers may have been early. Likewise, an early prealternate molt, before hormone signaling had fully switched to dark, might produce white in the faces, as shown by the two specimens with white-tipped alternate feathers (Figure 3C and D). That early-replaced basic feathers appeared to be white with dark tips, and early-replaced alternate feathers dark with white tips, suggests that hormone signaling can switch quickly, as feathers are growing, as also suggested by such patterns in some ducks and ptarmigan (Pyle 2005, 2007), although some murres with facial scores 2 and 3 also appeared to have dark-tipped white throat feathers (Figure 1), opposite of what might be expected from a late prebasic or preformative molt. If asynchrony of the hormones stimu- lating molt and melanin deposition does explain variation in the patterns of the Common Murre’s face, these results suggest that the signaling cycle produces dark-tipped white facial feathers in May and June, white facial feathers from July through September, white-tipped dark or dark-tipped white feathers in October, and dark feathers from November through April, although individual and age-related variation in the timing of this signaling is likely. If these interactions extend to the natal period, it could explain the dark faces developed by downy chicks in May and June and the dark faces of some early-hatched juveniles that may undergo the prejuvenal molt in early June. The precise timing and mechanisms of birds' pigment-deposition cycles are largely unknown and in need of further study. Alternatively, the darkest-faced murres in formative and basic plumage could represent a polymorphism rather than asynchrony of pigmentation and molt cycles (cf. Sibley 2000:243). I examined or observed few murres intermediate between the whiter-faced categories (facial scores 1 through 3) and darkest-faced category (facial score 4). In both basic and formative plumage only the darkest-faced birds lacked white on the tips of the secondar- ies, and these birds also appeared to show more dark in the underparts and 259 FEATURED PHOTO underwing coverts. The single dark-faced specimen in formative plumage (Figure 4) is difficult to explain by the hypothesis of asynchronous hormone signaling and molt cycles. A completely dark individual was collected near the Farallon Islands on 30 May 1911 (CAS 18072; Figure 3E and F), and another was photographed 24 July 2013 near the Farallon Islands (Figure 7). Presumably these two birds represent hyper melanism. However, the incidence of dark faces in central California in fall, 9-15% of basic-plumaged murres, is greater than typically attributable to a plum- age abnormality such as complete or partial hypermelanism, and murres breeding farther north appear to show less dark in the face in general (see below). These observations suggest that asynchrony of the hormone signal- ing and molt cycles rather than polymorphism likely accounts for most or all dark-faced Common Murres in basic plumage off central California. Com- pletely dark plumage, as seen in Figures 3E and F and Figure 7, also could result from extreme asynchrony of hormone cycles producing completely dark underparts during the previous prebasic molt, coupled with dark head feathering from the prealternate molt. In this regard, it is interesting that the darkest-faced birds also had darker basic feathers elsewhere, suggesting that signaling for dark feathers in birds undergoing early prebasic molt affects not just the head feathers replaced during the prealternate molt. In summary, outgoing and incoming alternate feathers, and early-replaced and late-replaced basic feathers, all may contribute to the occurrence of dark-faced Common Murres off central California from September through November, as may hypermelanism in a small proportion of individuals. The lack of dark-faced murres observed off Fort Bragg on one pelagic field trip (6 November 2011) suggests the murres in that area may have come from other breeding populations, such as those in Mendocino, Humboldt, and Del Norte counties, northern California (Carter et al. 1992, Manuwal et al. 2001). At the largest of these colonies, on Castle Rock, Del Norte County, breeding appears to occur three weeks to a month later than at Southeast Farallon Island (R. Golightly and C. Strong pers. comm., Manuwal et al. 2001). The later breeding could result in later second and definitive prebasic molts, bringing molt into closer synchrony with the hormone cycle signaling for a white face. Further study of the complex interaction between molts and feather coloration of the Common and Thick-billed murres throughout their breeding ranges is needed for an understanding of geographic variation in facial plumages, and whether the cycles’ interactions are influenced more by genetic or environmental factors. ACKNOWLEDGMENTS I thank Maureen Flannery (California Academy of Sciences) and Carla Cicero (Mu- seum of Vertebrate Zoology) for assistance and access to specimens, Debi Shearwater for invitations to lead pelagic field trips for Shearwater Journeys, Rick Golightly and Craig Strong for information on the Castle Rock murre colony, and Alan Wight, Al Jaramillo, and Dru Devlin for use of their photographs. Jeff N. Davis, Ian Jones, Al Jaramillo, and Steve N. G. Howell reviewed the manuscript or provided feedback on murre plumages. I am forever indebted to Rollo Beck, who collected and prepared the great majority of specimens examined for this study. This is publication 463 of the Institute for Bird Populations. 260 FEATURED PHOTO LITERATURE CITED Ainley, D. G., and Boekelheide, R. J., eds. 1990. Seabirds of the Farallon Islands: Ecology, dynamics, and structure of an upwelling-system community. Stanford Univ. Press, Stanford, CA. Ainley, D. G., Nettleship, D. N., Carter, H. R., and Storey, A. E. 2002. Common Murre (Uria aalge), in The Birds of North America (A. Poole and F. Gill, eds.), no. 666. Birds N. Am., Inc., Philadelphia. California Bird Records Committee (CBRC; R. A. Hamilton, M. A. Patten, and R. A. Erickson, eds.). 2007. Rare Birds of California. W. Field Ornithol., Camarillo, CA. Carter, H. R., McChesney, G. J., Jaques, D. L., Strong, C. S., Parker, M. W., Takekawa, J. E., Jory, D. L., and Whitworth, D.L. 1992. Breeding populations of seabirds in California, 1989-1991. U.S. Fish and Wildlife Service, Northern Prairie Wildlife Research Center, Dixon, CA. Gaston, A. J., and Jones, I. L.. 1998. The Auks: Alcidae. Oxford Univ. Press, Ox- ford, England. Harris, M. P., and Wanless, S. 1990. Moult and autumn colony attendance of auks. Br. Birds 83:55-66. Howell, S. N. G. 2010. Molt in North American Birds. Houghton Mifflin Harcourt, New York. Manuwal, D. A., Carter, H. R., Zimmerman, T. S., and Orthmeyer, D. L., eds. 2001. Biology and conservation of the Common Murre in California, Oregon, Wash- ington, and British Columbia. Volume 1: Natural history and population trends. U.S. Geol. Survey, Biol. Res. Div., Information and Technology Report SGS/ BRD/ITR-2000-0012, Washington, D.C. Pyle, P. 1997. Identification Guide to North American Birds, part 1. Slate Creek Press, Bolinas, CA. Pyle, P. 2005. Molts and plumages of ducks. Colonial Waterbirds 28:207-218. Pyle, P. 2007. Revision of molt and plumage terminology in ptarmigan (Phasianidae: Lagopus spp.) based on evolutionary considerations. Auk 124:508-514. Pyle, P. 2008. Identification Guide to North American Birds, part 2. Slate Creek Press, Point Reyes Station, CA. Pyle, P. 2009. Age determination and molt strategies in North American alcids. Marine Ornithol. 37:219-225. Roberson, D. 1980. Rare Birds of the West Coast. Woodcock Publ., Pacific Grove, CA. Stallcup, R. 1990. Ocean Birds of the Nearshore Pacific. Point Reyes Bird Observa- tory, Stinson Beach, CA. 261 World Wide Web site: WESTERN BIRDS www. westernfieldornithologists . org Quarterly Journal of Western Field Ornithologists President: Edward R. Pandolfino, 1328 49th St., Sacramento, CA 98519; ERPfromCA@aol . com Vice-President: David E. Quady, 39 The Crescent, Berkeley, CA 94708; davequady@att. net Past-President: W. David Shuford, P O. Box 69, Bolinas, CA 94924; dshuf ord@prbo . org Treasurer/Membership Secretary: Robbie Fischer, 1359 Solano Dr., Pacifica, CA 94044; robbie22@pacbell.net Recording Secretary: Liga Auzins, 12842 Safford East, Garden Grove, CA 92840; llauzins@yahoo . com Directors: Kenneth P. Able, Thomas A. Blackman, Jon L. Dunn, Kimball L. Garrett, Daniel D. Gibson, Robert E. Gill, Ed Harper, Kurt Leuschner, Joyce Meyer, Frances Oliver, Dan Singer, Catherine P. Waters Editor: Philip Unitt, San Diego Natural History Museum, P. O. Box 121390, San Diego, CA 92112-1390; birds@sdnhm.org Assistant Editor: Daniel D. Gibson, 3705 Quartz Rd., Ester, AK 99725; avesalaska@gmail . com Associate Editors: Kenneth P. Able, Doug Faulkner, Thomas Gardali, Daniel D. Gibson, Robert E. Gill, Paul E. Lehman, Ronald R. LeValley, Dan Reinking Graphics Manager: Virginia P. Johnson, 4637 Del Mar Ave., San Diego, CA 92107; gingerj 5@juno . com Photo Editor: Peter LaTourrette, 1019 Loma Prieta Ct., Los Altos, CA 94024; petelat 1 @stanford . edu Featured Photo: John Sterling, 26 Palm Ave., Woodland, CA 95695; jsterling@ wavecable . com WFO Website: Joseph Morlan, 1359 Solano Dr., Pacifica, CA 94044; webmaster@ westernfieldornithologists . org Book Reviews: Lauren Harter, 2841 McCulloch Blvd N. #1, Lake Havasu City, AZ 86403; lbharter@gmail.com Membership dues, for individuals and institutions, including subscription to Western Birds: Patron, $1000.00; Life, $600 (payable in four equal annual installments); Supporting, $75 annually; Contributing, $50 annually; Family $40; Regular U.S. $35 for one year, $60 for two years, $85 for three years. Dues and contributions are tax-deductible to the extent allowed by law. Send membership dues, changes of address, correspondence regarding missing issues, and or- ders for back issues and special publications to the Treasurer. Make checks payable to Western Field Ornithologists. BACK ISSUES OF WESTERN BIRDS WITHIN U.S. $40 PER VOLUME, $10 FOR SINGLE ISSUES, INCLUDING SHIPPING AND HANDLING. OUTSIDE THE U.S. $55 PER VOLUME, $15 FOR SINGLE ISSUES, INCLUDING SHIPPING AND HANDLING. The California Bird Records Committee of Western Field Ornithologists revised its 10-column Field List of California Birds in July 2009. The list covers 641 species, plus 6 species on the supplemental list. Please send orders to WFO, c/o Robbie Fischer, Treasurer, 1359 Solano Drive, Pacifica, CA 94044. Price for 9 or fewer, $2.75 each, for 10 or more, $2.50 each, which includes tax and shipping. Order online at http://checklist.westernfieldornithologists.org. Published 30 September 2013 ISSN 0045-3897 Common Murrc (Uria aalge), MonlBrcy Bay, California, I 5 September 2012. This individual shows a particularly dark head and lacks while on Line trailing edge of the secondaries, The medium-small bill suggests the bird may be in its second cycle. Photo by Ah'arn J arum ilia Common Murrc s {Uria aalge}, off Half Moon Bay,. California, 2 October 2011. The juvenile appeared to be following this dark -headed adult, suggesting ihaL the latter was a male at least four years of age. Photo by Peter Pyle Vol. 44, No. 4, 2013 Western Specialty: Lawrence’s Goldfinch Photos by © Thomas A Blackman of San Diego, California Lawrence’s Goldfinch {Spims fawrencei) Poway t San Diego County,. California, 12 February 201 3 '['bis remarkable pair of photographs, with males on the left and Females on the right, remind one of the meticulously painted plates of a field guide. Lawrence's Goldfinch is one of the primary species taking advantage of the wild, flowers that proliferate the year after a fire- and thus has capitalized on the wildfires that have swept so much of southern California since 2-0 02. Volume 44, Number 4, 2013 Movements of the Mangrove Warbler in Baja California Sur Cheryl L. Schweizer and Robert C. Whitmore 262 History of the Red-necked Stint Breeding in North America Lucas H. DeCicco, Nils Warnock, and James A. Johnson 273 Estimating the Number of Territorial Males in Low-Density Populations of the Sooty Grouse James D. Bland 279 Apparent Extirpation of the Sooty Grouse from the Sky Islands of South-Central California James D. Bland 294 NOTES The Distribution of Bubo uirginianus pinorum North and West to Washington Robert W. Dickerman 309 Records of the Black Merlin in New Mexico, with Comments on Its Identification Robert W. Dickerman 312 Behaviors of Nestling and Juvenile Black Vultures in Northwestern Mexico Marcelino Montiel-Herrera and Juan Pablo Gallo-Reynoso 316 Book Reviews Jeffrey L. Lincer 319 Featured Photo: Hybridization of the Black-footed and Laysan Albatrosses Cameron Rutt 322 Erratum 334 Index Daniel D. Gibson 336 Front cover photo by © Luke Eberhart-Phillips, of Bielefeld, Germa- ny: presumed male Red-necked Stint ( Calidris ruficollis) brooding two of four chicks, Kigluaik Mountains, Seward Peninsula, Alaska, 30 June 2013, representing the species’ first confirmed nesting in North America in 37 years. Back cover: “Featured Photos” by © Cameron Rutt of Blooming Glen, Pennsylvania: hybrid Laysan x Black-footed Albatrosses (Phoebastria immutabilis x P. nigripes) on Laysan, northwestern Hawaiian Islands. Top, presumed F x hybrid; bottom, presumed F 2 backer oss to the Black-footed Albatross, paired with a likely female Black-footed Albatross. Western Birds solicits papers that are both useful to and understandable by amateur field ornithologists and also contribute significantly to scientific literature. The journal welcomes contributions from both professionals and amateurs. Appropriate topics in- clude distribution, migration, status, identification, geographic variation, conservation, behavior, ecology, population dynamics, habitat requirements, the effects of pollution, and techniques for censusing, sound recording, and photographing birds in the field. Papers of general interest will be considered regardless of their geographic origin, but particularly desired are reports of studies done in or bearing on North America west of the 100th meridian, including Alaska and Hawaii, northwestern Mexico, and the northeastern Pacific Ocean. Send manuscripts to Daniel D. Gibson, P. O. Box 155, Ester, AK 99725; avesalaska@ gmail.com. For matters of style consult the Suggestions to Contributors to Western Birds (at www. western fieldornithologists.org/docs/journal_guidelines.doc). WESTERN BIRDS Volume 44, Number 4, 2013 MOVEMENTS OF THE MANGROVE WARBLER IN BAJA CALIFORNIA SUR CHERYL L. SCHWEIZER and ROBERT C. WHITMORE, P. O. Box 6125, West Virginia University, Morgantown, West Virginia 26506 (current address P. O. Box 257, Ethridge, Montana 59435); Cheryl_Schweizer@yahoo.com ABSTRACT: Mangrove forests are one of the most productive ecosystems in the world. Despite this, over half of the world’s mangroves have been lost through human activities. As suitable habitat declines, mangrove birds are forced into small isolated patches, exposing them to the dynamics of small populations. Our primary objective was to quantify local movement of Mangrove Warblers of the apparently sedentary subspecies Setophaga petechia castaneiceps, endemic to mangroves of Baja California Sur. In 2010, we captured and color-banded 108 breeding adult Mangrove Warblers at 16 sites, then surveyed all surrounding mangroves during the following winter and breeding seasons. We found no movement from one stand of mangroves to another, but we did find territory switching within a stand from winter to the breeding season. The rate of replacement of birds in a territory was high, sug- gesting that the proportion of floaters is high. We found no significant changes in population density by season or sex. The underlying framework for the conservation of fragmented populations is founded on the principles of island biogeography (MacArthur and Wilson 1967), that the probability of a species’ occurrence in a patch of habitat varies as a function of the patch’s size and isolation (Prugh et al. 2008). Simberloff (1974) stated, “any patch of habitat isolated from similar habitat by different, relatively inhospitable terrain traversed only with difficulty by organisms of the habitat patch may be considered an island.” The wildlife of habitat islands risks extirpation through random variation in demographic rates and environmental conditions, loss in genetic heterozygosity, edge ef- fects, cultural erosion, and human disturbance (Burkey 1995, Ludwig 1996). Small populations inhabiting habitat islands are especially vulnerable to these pressures (Smith et al. 2006). Rates of habitat fragmentation are so high that virtually all natural (continuous) terrestrial habitats and protected areas (such as nature preserves and parks) are certain to become habitat islands (Groombridge and Jenkins 2002). Worldwide, over 50% of mangroves have been destroyed (Holguin et al. 2006). The mangrove forest of Baja California Sur, Mexico, is a mosaic of small islands as a result of habitat fragmentation due to human activity. Within 20 years (1972-1992), 65% of mangrove forests in Mexico were destroyed (Herrera-Silveira and Ceballos-Cambranis 2000), and a further 70,000 ha were lost from 1993 to 2000 (SEMARNAT 2003). In northwestern Mexico, 262 Western Birds 44:262-272, 2013 MOVEMENTS OF THE MANGROVE WARBLER IN BAJA CALIFORNIA SUR 2300 ha were lost from the early 1970s to 2005 (Ruiz-Luna et al. 2010), Furthermore, this region is affected by seasonal hurricanes, which damage mangroves by reducing their stem density and basal area and by uprooting trees (Kovac et al. 2001). Not only are tropical depressions common, 78% of Mexico’s national total of hurricanes occur here (Flores-Verdugo et al. 1992). Some stands of mangroves in this area are weakened, being homogeneous in tree structure and diversity, broken by gaps, or in poor health from lack of benthic microbes, heightening the need for these unique habitats to be studied and their biodiversity preserved. In 2003, mangrove management in Mexico was regulated by Norma Oficial Mexicana NOM-022-SEMARNAT-2003, which “established the specifications for preservation, conservation, sustainable use and restoration of the coastal wetlands in mangrove areas” (SEMARNAT 2003). This law allowed for the enforcement of mangrove protection, until the agreement in 2004 to add Article 4.43, which states, “the prohibition of work and activities set out in paragraphs 4.4 and 4.22 and the limits set out in paragraphs 4.14 and 4.16 may be excluded if there is a preventative report of environmental impact statement, or if the case establishes compensation arrangements for the benefit of wetlands and it obtains corresponding authorization for change of land use” (Diario Oficial de la Federacion, 3 May 2004). This addition drew immediate attention from many environmental organizations calling for revision of the article. In 2007 another article was added (Article 60 TER of the Wildlife General Law), in which “it is forbidden to remove, fill, transplant, prune, or conduct any work or activity which directly or indirectly affects mangroves” (Diario Oficial de la Federacion, 1 February 2007). Under this addition, Mexico prohibited nearly all urban development harmful to mangrove ecosystems. Few birds specialize in mangrove forests, and detailed research on the ecology of these specialists is sparse (Hogarth 1999, Luther and Greenberg 2009) . One of the many subspecies of the Yellow Warbler (Browning 1994), Setophaga petechia castaneiceps is endemic to the mangroves of Baja California Sur. It is an apparently sedentary (although no published data have confirmed this) habitat specialist confined to mangroves year round (Curson et al. 1994). Although Baja California Sur has dry and rainy seasons, Cox (1968) considered the lack of extreme seasonality to be a major factor in tropical birds’ ability to maintain territories and often monogamous pair bonds year round. As the Mangrove Warbler specializes on a unique yet declining habitat, more research on its patterns of movement is needed. Rates of immigra- tion and emigration into and out of mangrove stands in part determine the subspecies’ susceptibility to the risks run by isolated, small populations. Presumably the number of birds living in one patch is limited by space and competition for resources such as nesting sites and food (e. g., Holmes 2010) . Low numbers, coupled with isolation and a sedentary lifestyle (pos- sibly no immigration and emigration), may be depressing genetic variation (Harrison and Hastings 1996, Callens et al. 2011). Inbreeding may be making the populations more susceptible to environmental changes. Five named hurricanes crossed Baja California Sur between 2001 and 2009 (NOAA 2011), displacing birds, destroying habitat, and possibly adding to 263 MOVEMENTS OF THE MANGROVE WARBLER IN BAJA CALIFORNIA SUR genetic complications. Also, increasing fragmentation of mangroves may be increasing edge effects. Because these factors may threaten the viability of S. p. castaneiceps, we attempted to quantify its local movement. STUDY AREA The study area included the known range of S. p. castaneiceps, which inhabits patches of mangroves on both coasts of Baja California Sur north to San Ignacio Lagoon and Estero el Coyote (Pond Lagoon) on the Pacific coast and to Ejido San Lucas (27.23° N) on the Gulf of California (Browning 1994, Dunn and Garrett 1997) (Figure 1). On the basis of locations reported by Whitmore et al. (2000), we established within this range 16 study sites in five general areas: La Paz, Mulege, and Ejido San Lucas, comprising 13 sites along the Sea of Cortez, and Magdalena Bay and Puerto Adolfo Lopez Ma- teos, comprising 3 sites along the Pacific coast (Figure 1). Each study site was a discrete mangrove stand separated by desert from the next closest stand. The stands of mangrove at all study sites on the Sea of Cortez were less than l. 8 km (straight line distance) in maximum length or width. Their area ranged from 2.4 to 6 ha, except for site LP6, which covered 25. 1 ha. On the Pacific coast, Magdalena Bay supports a large continuous mangrove stand stretching (straight-line distance) approximately 115 km along the coast. Our largest site (67.8 ha) was at Magdalena Bay. The entire study area spanned a total straight- line distance of 390 km and total area of 127.5 ha. The distance between neighboring study sites averaged 6.6 km but ranged from 0.9 to 31.8 km. Arid Baja California Sur has an annual precipitation <300 mm and mean temperatures of 20-22°C (Ruiz-Luna et al. 2010). Its mangroves are of three species, Auicennia germinans (black mangrove), Rhizophora mangle (red mangrove), and Laguncularia racemosa (white mangrove) (Hogarth 1999, Flores-Verdugo et al. 1992, Ruiz-Luna et al. 2010). Unlike those elsewhere, in Baja California Sur mangrove trees seldom exceed 5 m in height (Felger et al. 2001). Our study areas were dominated by red mangrove (pers. obs.). METHODS Data Recording At each study site, we captured and banded Mangrove Warblers during the breeding season (May 2010) by using recordings of their songs to lure them into mist nets. When broadcasting the songs, we assessed the birds’ territoriality by their behavior, considering them to be defending a territory if they fluttered their wings, flew at the speaker or observer, chirped, chased other birds, or countersang to the recording. We banded each bird with a U.S. Geological Survey aluminum band and unique combination of color bands for identification of individuals. We re- corded locations of capture via a GPS unit with a minimum accuracy of 10 m. After being banded, birds were released at the point of capture. All birds captured were adult by plumage and skull pneumatization. We observed no juveniles or fledglings; apparently nestlings had not yet fledged by the time of our banding, and the broadcasting of song targeted territorial adults. When possible, we sexed the birds by coloration and presence of brood patch or cloacal protuberance (Howell and Webb 1995, Pyle 1997). 264 MOVEMENTS OF THE MANGROVE WARBLER IN BAJA CALIFORNIA SUR Figure 1. Five major study areas (San Lucas, Mulege, Puerto Adolfo Lopez Mateos (PALM), Magdalena Bay, and La Paz) and specific sites of study of movements of the Mangrove Warbler in Baja California Sur, Mexico. We recorded a pair if both a female and male were captured at the same location and defended the territory together. As reported by Salgado-Ortiz et al. (2008), at Celestun, Yucatan, subspe- cies bryanti defends territories and maintains permanent pair bonds and a stable population density year round. Since we were interested in movements year round, we searched for banded birds in November 2010 as well as in the following breeding season (April 2011). We revisited each point of original capture (net location) at least once in November and April. At each location, we broadcasted Mangrove Warbler vocalizations for 15 minutes or until we detected a bird (banded or unbanded). We recorded all individuals detected during the 15 minutes within 10 m of the 265 MOVEMENTS OF THE MANGROVE WARBLER IN BAJA CALIFORNIA SUR point (Hutto et al. 1986) and noted their behavior. We chose a small radius because of the density of the vegetation and to ensure 100% detection within that radius. We spent more than 15 minutes at the location if needed to confirm if a bird was banded, but we recorded no individuals after the 15-minute survey. To standardize our effort at resighting, we recorded net hours and ob- servation hours to keep effort as equal as possible (Winker et al. 1997) per location of original capture. We attempted to resight the birds at the same times of the day as the original capture and under desirable field conditions, with slight differences due to seasonal changes in the times of sunrise and sunset. By standardizing effort at resighting by location of original capture (regardless of how many birds were captured there) we attempted to keep the distance the broadcasts could be heard equal, reducing the chance the birds would be drawn in to the area of broadcast. Data Analyses We excluded from our analyses the six individuals we could not sex. In- dividuals that were detected multiple times during a resighting survey were only counted once. If a Mangrove Warbler followed the observers from one survey point to a neighboring point (which was obvious), we counted it only once at the point where it first actively defended a territory. We defined resighting success as the percent of banded birds resighted during each survey. Because we visited some study sites more often than oth- ers during the follow-up surveys, we needed to express the percent of birds resighted by a measure of effort. To this end, we divided the total number of individuals resighted by number of individuals we attempted to resight. We calculated population density by dividing the number of individuals detected during the 15-minute surveys by the area of the site. Using a repeated-measures analysis of variance (ANOVA) in SAS (version 9.2), we tested for significant differences in densities by sex category (male, female, and pair) and survey season We used ArcMap 9.3.1 to measure the distances banded birds moved from their location of original capture. We confirmed birds’ switching territories and replacing others in a territory by comparing their locations to their previous locations by GPS. RESULTS Capture, Territoriality, and Resighting Success During the breeding season of 2010, we captured and banded 74 adult males and 34 adult females (28 pairs, Figure 2) at 57 locations within the 16 study sites. Every individual captured was actively defending its territory. Chipping was the main territorial cue Mangrove Warblers used both in winter and the breeding season; song was secondary. Even in the morning during the breeding season, song was infrequent and the birds communicated more consistently with their chip notes. Playing song and chip notes during both follow-up surveys led to aggression by both the male and female territory holders in which they chased off intruders (including migrant Yellow War- blers). Adult males’ responses involved more flying and chasing, whereas females’ aggression was harsh chipping accompanied by hopping through the mangrove trees looking for the unrecognized individual. 266 MOVEMENTS OF THE MANGROVE WARBLER IN BAJA CALIFORNIA SUR 30 25 - c 20 t/S C3 3 15 T3 C 10 5 - Banded Male Re-sighted Male ■ La Paz [5 Mulege □ Magdalena Bay PALM □ San Lucas Banded Re-sighted Banded Pair Re-sighted Female Female Pair Figure 2. Number of Mangrove Warblers banded (May 2010) and resighted (November 2010, April 2011) in Baja California Sur by site. PALM, Puerto Adolfo Lopez Mateos. We resighted 43 marked males and 13 marked females, representing 52% of all birds banded. Nine marked males and two marked females were resighted in both November and April, as were four pairs that remained together through the year of this study. The highest number of resighted banded birds (25) was in the La Paz study area; the lowest (6) was at Mag- dalena Bay (Figure 2). The percentage of resighted banded birds, both males (75%) and females (56%), was greatest at Mulege (Table 1). The percentage of banded males resighted was 58%, that of females 38% (Table 1). The percentages of males (37%, n = 22) and females 20%, n = 5) resighted during November were greater than those during the following April (30% and 14%, respectively, Table 1). Population Density The density of adult males, adult females, and pairs was greatest in the San Lucas study area in the breeding season of 2010 (2.2/ha, 1.7/ha, and 1.7/ha, respectively) and in November 2010 (1.5/ha, 1.0/ha, and 1.0/ha, respectively) (Table 2). In April 2011, these densities were highest at Mulege (1.1/ha, Table 2). All sites and seasons combined, the average density of adult males was 1.3/ha, that of adult females 0.9/ha, and that of pairs 0.9/ ha. The density of males was significantly higher than that of females and pairs during all three seasons surveyed (P = 0.02). There was no significant change in overall density from season to season. Territory Switching We detected no movements from one patch of mangroves to another. From November 2010 to April 2011, however, eight banded males moved their territories within a patch, by an average distance of 0.59 km. We ob- served movement of only one female, to the closest neighboring territory with a male. 267 MOVEMENTS OF THE MANGROVE WARBLER IN BAJA CALIFORNIA SUR Table 1 Percentages of Banded Mangrove Warblers Resighted by Study Period, Area, and Sex November 2010 April 2011 Total % Resighted % Resighted of % Resighted % Resighted of % Resighted of Study area by effort number banded by effort number banded numberbanded and site Male Female Male Female Male Female Male Female Male Female La Paz 58.30 33.3 48.3 25.0 48.3 12.5 48.3 12.5 66.7 37.5 LP1 0 0 0 0 0 0 0 0 0 0 LP3 100 100 50 50 100 LP4 100 50 100 50 100 0 100 0 100 50 LP6 63.6 100 58.3 33.3 50 0 50 0 83.3 33.3 LP7 16.7 0 10 0 30 50 30 50 40 50 Maqdalena Bay MB1 11.8 0 11.1 0 27.8 28.6 27.8 28.6 27.8 28.6 15.4 0 15.4 0 23.1 0 23.1 0 23.1 0 MB3 0 0 0 0 40 66.7 40 66.7 40 66.7 Puerto Adolfo Lopez Mateos 42.9 25 42.9 25 42.9 25 Mulege 40 12.5 33.3 11.1 50 55.6 50 55.6 75 55.6 MU2 60 0 60 0 40 100 40 100 80 100 MU3 0 50 0 50 100 50 100 50 100 50 MU4 100 0 33.3 0 33.3 50 33.3 50 66.7 50 MU5 0 0 0 0 50 0 50 0 50 0 MU6 0 0 0 0 100 0 100 0 100 0 San Lucas 11.1 16.7 11.1 14.3 33.3 28.6 33.3 28.6 44.4 28.6 SL1 0 0 0 0 0 0 0 0 0 0 SL2 33.3 50 33.3 50 33.3 100 33.3 100 66.7 100 SL3 0 0 0 0 40 0 40 0 40 0 Total 37.3 20 29.7 14.7 43.2 32.4 43.2 32.4 58.1 38.2 Territory Replacement Forty-seven males and 21 females vacated their territory during the study. Vacancies of males were filled by another male in 32 cases (68%). An open- ing for a female was filled 14 times (67%). Unbanded males and females were more likely to fill vacant territories (n = 25) than were banded males (n = 7) and females (n = 1). Of the 28 pairs of which we banded both the male and the female, we resighted 4. Of the remaining 24 pairs, 11 were not observed again, while of the 13 of which one of the pair was resighted, 1 1 individuals were resighted with a new mate and 2 males remained in their territories but we did not see a mate with them. The majority of the territory replacements took place between November 2010 and April 2011; only four took place between May and November 2010. DISCUSSION Our results confirm that S. p. castaneiceps is a year-round resident. Ter- ritorial birds, both males and females, moved minimally and defended the territory year round, much like other tropical passerines (Greenberg and Gradwohl 1986, 1997, Lefebvre et al. 1992, Morton et al. 2000, Salgado- Ortiz et al. 2008). Although not all banded individuals were resighted, and we observed new unbanded adult Mangrove Warblers during both follow-up 268 MOVEMENTS OF THE MANGROVE WARBLER IN BAJA CALIFORNIA SUR Table 2 Population Density 0 of the Mangrove Warbler in Baja California Sur March-July 2010 November 2010 April 2011 Study area and site Male Female Pairs Male Female Pairs Male Female Pairs La Paz 0.72 0.22 0.22 0.72 0.37 0.35 0.47 0.32 0.42 LP1 0.42 0.42 0.42 1.27 0.84 0.84 0.42 0 0 LP3 0.54 0 0 0.54 0.27 0.27 0.27 0.54 0.27 LP4 1.11 0.55 0.55 2.77 1.39 1.39 1.94 0.83 0.83 LP6 0.48 0.16 0.16 0.36 0.24 0.20 0.40 0.32 0.32 LP7 1.70 0.34 0.34 0.85 0.17 0.17 1.02 0.85 0.85 Magdalena 0.27 0.10 0.10 0.15 0.10 0.10 0.13 0.13 0.13 Bay MB1 0.50 0.15 0.15 0.35 0.23 0.23 0.23 0.19 0.19 MB3 0.12 0.07 0.07 0.02 0.02 0.02 0.07 0.10 0.10 Puerto Adolfo Lopez Mateos 1.08 0.62 0.62 — — — 0.77 0.77 0.77 Mulege 1.37 1.03 0.91 0.91 0.23 0.23 1.14 1.14 1.14 MU2 3.55 2.13 2.13 2.13 0 0 2.13 2.84 2.84 MU3 2.86 5.71 2.86 2.86 2.86 2.86 2.86 2.86 2.86 MU4 0.93 0.62 0.62 0.62 0 0 0.93 0.93 0.93 MU5 0.68 0.34 0.34 0.34 0 0 0.68 0.34 0.34 MU6 1.14 1.14 1.14 1.14 1.14 1.14 1.14 1.14 1.14 San Lucas 2.23 1.73 1.73 1.49 0.99 0.99 0.99 0.99 0.99 SL1 5.88 5.88 5.88 5.88 0 0 0 0 0 SL2 1.72 1.15 1.15 1.72 1.72 1.72 1.15 1.15 1.15 SL3 2.36 1.89 1.89 0.94 0.47 0.47 0.94 0.94 0.94 “Number of individuals per hectare. surveys, we infer that adults of S. p. castaneiceps are sedentary, only juve- niles dispersing. Failure to resight could be due to mortality, loss of bands, or missed detections. On the east coast of Baja California Sur, failure to resight banded birds cannot be due to their emigration from study sites because we surveyed all mangrove stands in that region. During the initial surveys in the breeding season of 2010, the unbanded adults may have been present as floaters, missed because of their nonterritorial behavior. The density of mangrove stands makes Mangrove Warblers difficult to detect. Despite the high rate of replacement of individuals, territory occupancy and density of S. p. castaneiceps remained fairly stable through the year, much like those of S. p. bryanti (Salgado-Ortiz et al. 2008). This pattern is consistent with other sedentary tropical birds (Cox 1985, Greenberg and Gradwohl 1986, 1997, Gorrell et al. 2005). We observed only one addition of a territorial pair to a mangrove patch. The density of males appearing greater than that of females and pairs may be due males’ more assertive defense of their territories, leading to a higher detection rate. Of the 43 banded males we resighted, all adults, 8 (19%) moved to another territory. In all of these cases, the male was seen with a new female. One male switched his territory to a neighboring territory (abandoning his female), paired with the female already in that territory, and recruited an additional female from a neighboring territory from which the male had disappeared. Territory switching by tropical birds that are territorial year round is com- mon; for example, Greenberg and Gradwohl (1997) reported that 37.5% 269 MOVEMENTS OF THE MANGROVE WARBLER IN BAJA CALIFORNIA SUR of Checker-throated Antwrens (. Epinecrophylla fulviventris) of known age switched territories. Vacant territories are filled within a day (Levin 1996, Morton et al. 2000). Mangrove Warblers maintain type A territories within which courtship, mating, nesting, and foraging all occur. Territory switching may be due to vacancy of a territory with more resources or an available mate (Morton et al. 2000, Fedy and Stutchbury 2004). Whether the original territorial males were displaced or died and subsequently replaced by another individual is unknown. Experiments with the sedentary White-bellied Antbird ( Myrmeciza lon- gipes) found that when adult males were temporarily removed from their territories, some neighboring territorial males switched to the new open territories (Fedy and Stuchbury 2004). Switching to territories where more food (Gorrell et al. 2005) or foraging substrate (Morton et al. 2000) is avail- able has also been recorded in other species of resident territorial birds, presumably helping the birds increase their survivorship or fitness. We observed a high rate of replacement of individuals (53% of males, 62% of females). In habitats where birds maintain a stable population density and are territorial year round, it is common that vacant territories are filled within a matter of hours (Greenberg and Gradwohl 1997, Morton et al. 2000, Fedy and Stutchbury 2004). In November, we observed many Yellow Warblers throughout the mangrove sites but we could not confirm whether these were migrant Yellow Warblers from farther north, immature Mangrove Warblers, or floaters. We do not know where or when these individuals moved to the vacant territories, but it is plausible they were floaters. A high incidence of floaters could be due to lack of dispersal of young, the limited extent of habitat suitable for territory establishment, or possibly a response to an increase in the density of males (Smith et al. 2006). Other studies, however, report that floaters are uncommon among sedentary territorial birds (Levin 1996, Morton et al. 2000, Fedy and Stuchbury 2004). It is possible that the individuals that gained access to the territories were the young from the previous year (as a function of density dependence; Woolfenden and Fitzpatrick 1984). We do not know whether the young disperse out of their natal mangrove stand, as we captured no juveniles. Salgado Ortiz et al. (2008) found that fledglings of S. p. bryanti remained in natal territories for an average of 27.4 days but did not investigate their movements or dispersal after this time. In a resident island population of the Song Sparrow ( Melospiza melodia), Smith et al. (2006) found that 35% of territory replacements were by first- year floaters. They also reported that about 25% of the displaced territory holders became floaters in their previously held territory. These territory replacements differed by the male’s age, males 2 and 3 years old being more likely to take over territories than those 1 or >4 years old. Given the rate of destruction and fragmentation of mangroves worldwide, continued research on mangrove specialists is crucial. Information on dis- persal of young Mangrove Warblers is needed for a better understanding of site fidelity, floaters, and possibly founding of new populations as a result of tropical storms. Furthermore, estimates of the population size and a better understanding of basic life history would aid any efforts toward the conserva- tion of both the Mangrove Warbler and mangrove ecosystems. 270 MOVEMENTS OF THE MANGROVE WARBLER IN BAJA CALIFORNIA SUR ACKNOWLEDGMENTS We thank T. Tuncer and S. Jaquier for assistance with collection of data, P. B. Wood, P. Turk, J. Burkhart, and E. Palacios for all their advice and help, and Z. Kaiser, X. Schweizer, and D. Whitmore for patience and support. Our study was done under permits 20580 from the U.S. Geological Survey and SGPA/DGVS/01392/11 from the Secretaria de Medio Ambiente y Recursos Naturales. LITERATURE CITED Browning, M. R. 1994. A taxonomic review of Dendroica petechia (Yellow Warbler) (Aves: Parulidae). Proc. Biol. Soc. Washington 107:27-51. Burkey, T. V. 1995. Extinction rates in archipelagoes: Implications for populations in fragmented habitats. Cons. Biol. 9:527-541. Callens, T., Galbusera, P., Matthysen, E., Durand, E. Y., Githiru, M., Huyghe, J. R., and Lens, L. 2011. Genetic signature of a population fragmentation varies with mobility in seven bird species of a fragmented Kenyan cloud forest. Molec. Ecol. 20:1829-1844. Cox, G. W. 1968. The role of competition in the evolution of migration. Evolution 22:180-192. Cox, G. W. 1985. The evolution of avian migration systems between temperate and tropical regions of the New World. Am. Nat. 126:451-474. Curson, J., D. Quinn, and D. Beadle. 1994. New World Warblers. A. & C. Black, London. Dunn, J. L., and Garrett, K. L. 1997. A Field Guide to Warblers of North America. Houghton Mifflin, Boston. Fedy, B. C., and Stutchbury, B. J. M. 2004. Territory switching and floating in White- bellied Antbird ( Myrmeciza lonqipes), a resident tropical passerine in Panama. Auk 121:486-496. Felger, R. S., Johnson, M. B., and Wilson, M. F. 2001. The Trees of Sonora, Mexico. Oxford Univ. Press, New York. Flores- Verdugo, F., Gonzalez-Farias, F., Zamorano, D. S., and Ramirez-Garcia, P. 1992. Mangrove ecosystems of the Pacific coast of Mexico: Distribution, struc- ture, litterfall, and detritus dynamics, in Coastal Plant Communities of Latin America (U. Seeliger, ed.), pp. 269-288. Academic Press, San Diego. Gorrell, J. V., Ritchison, G., and Morton, E. S. 2005. Territory size and stability in a sedentary neotropical passerine: Is resource partitioning a necessary condition? J. Field Ornithol. 76:395-401. Greenberg, R., and Gradwohl, J. 1986. Constant density and stable territoriality in some tropical insectivorous birds. Oecologia 69:618-625 Greenberg, R., and Gradwohl, J. 1997. Territoriality, adult survival, and dispersal in the Checker-throated Antwren in Panama. J. Avian Biol. 28:103-110. Groombridge, B., and Jenkins, M. 2002. World Atlas of Biodiversity: Earth’s Living Resources in the 21 st Century. Univ. of Calif. Press, Berkeley. Harrison, S., and Hastings, A. 1996. Genetic and evolutionary consequesnces of metapopulation structure. Trends Ecol. Evol. 11:180-183. Herrera-Silveira, J., and Ceballos-Cambranis, E. 2000. Manglares: Ecosistemas valiosos. Biodiversitas 19:1-10. Hogarth, P. J. 1999. The Biology of Mangroves. Oxford Univ. Press, New York. Holguin, G., Gonzalez-Zamorano, P., de-Bashan, L. E., Mendoza, R., Amador, E., and Bashan, Y. 2006. Mangrove health in an arid environment encroached by urban development — a case study. Sci. Total Env. 63:260-274. Holmes, R. T. 2010. Avian population and community processes in forest ecosys- tems: Long-term research in the Hubbard Brook experimental forest. Forest Ecol. Mgmt. 262:20-32. 271 MOVEMENTS OF THE MANGROVE WARBLER IN BAJA CALIFORNIA SUR Howell, S. N. G., and Webb, S. 1995. A Guide to the Birds of Mexico and Northern Central America. Oxford Univ. Press, New York. Hutto, R. L., Pletschet, S. M., and Hendricks, P. 1986. A fixed-radius point count method for nonbreeding and breeding season use. Auk 103:593-602. Kovac, J. M., Blanco-Correa, M., and Flores-Verdugo, F. 2001. A logistical regres- sion model of hurricane impacts in a mangrove forest of the Mexican Pacific. J. Coastal Res. 17:30-37. Lefebvre, G., Poulin, B., and McNeil, R. 1992. Settlement period and function of long-term territory in tropical mangrove passerines. Condor 94:83-92. Levin, R. N. 1996. Song behavior and reproductive strategies in a duetting wren, Thryo- thorus nigricapillus: I. Removal experiments. Anim. Behav. 52:1093-1106. Ludwig, D. 1996. The distribution of population survival times. Am. Nat. 147:506-526. Luther, D. A., and Greenberg, R. 2009. Mangroves: A global perspective on the evo- lution and conservation of their terrestrial vertebrates. BioScience 59:602-612. MacArthur, R. H., and Wilson, E. O. 1967. The Theory of Island Biogeography. Princeton Univ. Press, Princeton, NJ. Morton, E. S., Derrickson, K. C., and Stutchbury, B. J. M. 2000. Territory switch- ing behavior in a sedentary tropical passerine, the Dusky Antbird ( Cercomarcra tyrannina). Behav. Ecol. 11:648-653. National Oceanic And Atomopheric Administration (NOAA). 2011. Historical hurricane tracks (www.csc.noaa. gov/hurricanes/#app=3d30&3e3d-selected lndex=0). Prugh, L., Hodges, K. E., Sinclair, A. R. E., and Brashares, J. S. 2008. Effect of habitat area and isolation on fragmented animal populations. Proc. Natl. Acad. Sci. 105:20770-20775. Pyle, P. 1997. Identification Guide to North American Birds, part I. Slate Creek Press, Bolinas, CA. Ruiz-Luna, A., Cervantes Escobar, A., and Berlanga-Robles, C. 2010. Assessing dis- tribution patterns, extent, and current condition of northwest Mexico mangroves. Wetlands 30:717-723. Salgado-Ortiz, J., Marra, P. P., Sillett, T. S., and Robertson, R. J. 2008. Breeding ecology of the Mangrove Warbler {Dendroica petechia bryanti ) and compara- tive life history of the Yellow Warbler subspecies complex. Auk 125:402-410. Secretaria de Medio Ambiente y Recursos Naturales. 2003. Norma Oficial Mexi- cana NOM-022-SEMARNAT-2003, que establece las especificaciones para la preservacion, conservacion, aprovechamiento sustentable y restauracion de los humedales costeros en zonas de manglar. Diario Official de la Federation, 10 April 2003. Simberloff, D. S. 1974. Equilibrium theory of island biogeography and ecology. Annu. Rev. Ecol. Syst. 5:161-188. Smith, J. N. M., Keller, L. F., Marr, A. B., and Arcese, P. 2006. Conservation and biology of small populations: The Song Sparrow of Mandarte Island. Oxford Univ. Press, New York. Whitmore, R. C., Whitmore, R. C., and Whitmore, M. W. 2000. Distributional notes on the Mangrove Warbler (Dendroica petechia castaneiceps ) near the northern edge of its range in eastern Baja California Sur, Mexico. W. N. Am. Nat. 60:228-229. Winker, K., Escalante, P., Rappole, J. H., Ramos, M. A., Oehlenschlager, R. J., and Warner, D. W. 1997. Periodic migration and lowland forest refugia in a “sed- entary” neotropical bird, Wetmore’s Bush-Tanager. Cons. Biol. 11:692-697. Woolfenden, G. E., and Fitzpatrick, J. W. 1984. The Florida Scrub Jay: Demography of a Cooperative-Breeding Bird. Princeton Univ. Press, Princeton, NJ. Accepted 12 June 2013 272 HISTORY OF THE RED-NECKED STINT BREEDING IN NORTH AMERICA LUCAS H. DeCICCO, U.S. Fish and Wildlife Service, Migratory Bird Management, 1011 East Tudor Rd., Anchorage, Alaska 99503; lhdecicco@gmail.com NILS WARNOCK, Audubon Alaska, 441 West Fifth Ave., Suite 300, Anchorage, Alaska 99501 JAMES A. JOHNSON, U.S. Fish and Wildlife Service, Migratory Bird Management, 1011 East Tudor Rd., Anchorage, Alaska 99503 ABSTRACT : The largely palearctic Red-necked Stint has been documented breed- ing in the Nearctic Region only in Alaska, from which 12 records were published from 1909 to 1975. In summer 2012 we found a family of Red-necked Stints in the Kigluaik Mountains of the Seward Peninsula, in tundra of the dwarf shrub mat type with >50% cover of bare rock. The photographs obtained are the first published of the Red-necked Stint breeding in Alaska. The Red-necked Stint ( Calidris ruficollis) breeds in three disjunct regions, all in northern Russia: (1) the central and eastern portions of the Taimyr Peninsula, (2) the area south and upland of the Lena River delta, and (3) from the western side of the Kolyma lowlands east to the Chukchi Peninsula (the presumed core of the breeding range) and south to northern Kamchatka (Lappo et al. 2012, Figure 1). After breeding, Red-necked Stints from central Siberia are thought to migrate south by interior flyways, while those breed- ing in northeastern Russia and Alaska are thought to migrate south along the East Asian-Australasian flyway (Piersma et al. 1996). The winter range extends from India east to China and south to Australia and New Zealand (Morozov and Tomkovich 1984, Piersma et al. 1996). In the Nearctic Region, the Red-necked Stint occurs regularly only in Alaska, where it is a rare spring and fall migrant in southwestern Alaska, a rare spring migrant, rare breeder, and very rare fall migrant in western Alaska, and a very rare spring migrant and summer visitant in northern Alaska (Kessel and Gibson 1978, Kessel 1989). The last published record of breeding was in 1975 (Kessel and Gibson 1978). Like most calidridine sandpipers, the Red-necked Stint generally lays a 4-egg clutch (Cramp and Simmons 1983, Piersma et al. 1996) in a small cup on the ground, lined with grass and willow leaves. Depending on loca- tion, the species begins laying eggs from late May to mid-June; eggs hatch from mid-June to early July (Portenko 1972, Kessel 1989). In this paper, we report on the habitat and behavior of Red-necked Stints that nested on the Seward Peninsula in 2012 and review the status of the species’ breeding in North America. 2012 BREEDING RECORD On 30 June 2012, while searching for nests of the Red Knot (C. canu- tus) in the Kigluaik Mountains, about 45 km northwest of Nome, Seward Peninsula (64.50° N, 165.41° W; Figure 1), Warnock observed a small drab shorebird flush and give an alarm call. Given the Western Sandpipers Western Birds 44:273-278, 2013 273 HISTORY OF THE RED-NECKED STINT BREEDING IN NORTH AMERICA (C. mauri) in the area, he first assumed the bird was that species. Soon a second, more brightly colored bird (Figure 2), readily identified as a Red- necked Stint and presumably the male (Pavel Tomkovich pers. comm. 2012), appeared, calling. Subsequently, both adults performed a distraction display characterized by a low, hunched-over posture with slightly spread wings (Figure 3). We then observed four very young chicks (1 or 2 days old). We photographed the birds (still and video) and recorded their vocalizations. The audio recording is archived in and accessible through Cornell University’s Macaulay Library (catalog number ML1 76409). On 4 July we observed a drab adult and a single chick, 1.5 km from the original location. Given that young Western Sandpiper chicks of a similar size can move over 0.5 km in a day (Ruthrauff and McCaffery 2005), we assumed these stints were from the brood we saw initially. This family occurred on an alpine granitic ridge 16 km inland from Norton Sound and 300 m above sea level. The surrounding habitat was mainly dwarf shrub mat (Kessel 1979), comprising dwarf shrubs ( Empetrum nigrum , Salix spp., Dryas spp.), lichens, and sedges ( Carex spp.); >50% of the area was bare ground or rock-covered (Figure 4). This location lies within the transition zone between wet graminoid and shrub-dominated lowland tundra and granitic Dry as-dominated alpine tundra (Kessel 1979). Other breeding Figure 1. Locations of recorded breeding of the Red-necked Stint in Alaska; numbers correspond to records listed in Table 1 . Hatched areas represent the eastern segment of the Red-necked Stint’s breeding distribution in Russia. 274 HISTORY OF THE RED-NECKED STINT BREEDING IN NORTH AMERICA birds observed in this area included the American and Pacific golden-plovers (. Pluuialis dominica and fulua), Red Knot, Rock Sandpiper ( Calidris pti- locnemis), Western Sandpiper, Horned Lark ( Eremophila alpestris), and Lapland Longspur ( Calcarius Iapponicus). REVIEW OF NORTH AMERICAN BREEDING RECORDS The Red-necked Stint was first recorded breeding in Alaska at Nome on 10 July 1908, when A. H. Dunham collected a pair of adults with two young (Thayer 1909). Thirteen attempts of the Red-necked Stint to nest in Alaska have now been recorded, including the one we describe here (Table 1). These observations are centered geographically on the Seward Peninsula, particularly near Cape Prince of Wales. The only record elsewhere is from Barrow (Table 1, Figure 1). Almost half of them, six of 13, were found in the 1930s and 1940s, a time when museum-funded exploration in the region was widespread. The largest temporal gap, 37 years, is between the ultimate and penultimate breeding records (Table 1). It is unclear if so long a lapse is due to population fluctuation or the dearth of visits by ornithologists to the Wales area during the intervening decades. The breeding habitat of the Red-necked Stint in Alaska appears to be variable but consistently within dwarf shrub mat and dwarf shrub meadow (Kessel 1979). For example, Bailey (1926:32) reported birds “on the high tundra at the base of Wales Mountain,” including a pair building a nest along a stream. Kessel (1989), in the same area three decades later, observed a Red-necked Stint nest at 150 m elevation along a creek draw dotted with large rocks. Other nests near Wales were found at a “mountain dry place” (Hanna 1940:123) and on flat tundra “plains” (Ford 1934:232). In the Palearctic, the highest densities of breeding Red-necked Stints tend to be close to the coast and by rivers (Lappo et al. 2012). Portenko (1972) and Piersma et al. (1996) described the species’ typical breeding habitat as upland tundra at low elevations, often in drier, rocky areas. With principal nesting areas in the eastern Palearctic and satellite nesting areas in Alaska, the distribution of the Red-necked Stint is not anomalous biogeographically. Other species that share this pattern include three other shorebirds — the Eurasian Dotterel ( Charadrius morinellus), Lesser Sand- Plover (C. mongolus), and Common Ringed Plover (C. hiaticula ); one passerine — the Red-throated Pipit ( Anthus ceruinus); and one loon — the Arctic ( Gavia arctica). This pattern is bracketed by species of palearctic provenance that occur less regularly in western Alaska and have not been found breeding, such as the Terek Sandpiper ( Xenus cinereus ) and Great Knot ( Calidris tenuirostris), and by those that have well-established breed- ing populations in Alaska, such as the Bluethroat ( Luscinia svecica) and Eastern Yellow Wagtail ( Motacilla tschutschensis) (Kessel and Gibson 1978, Kessel 1989). These distributional patterns may be a result of the current geographic proximity of North America and Asia, separated by only 80 km at the Bering Strait, or a remnant of a time when the continents formed one continuous landmass (Hopkins 1959). 275 HISTORY OF THE RED-NECKED STINT BREEDING IN NORTH AMERICA Figure 2. Presumed male Red-necked Stint attending chicks (1 or 2 days old), Kigluaik Mountains, Seward Peninsula, Alaska, 30 June 2012. Photo by Lucas H. DeCicco Figure 3. Presumed female Red-necked Stint in distraction display posture, Kigluaik Mountains, Seward Peninsula, Alaska, 4 July 2012. Photo by Lucas H. DeCicco 276 HISTORY OF THE RED-NECKED STINT BREEDING IN NORTH AMERICA Figure 4. Granitic Dryas-dominated alpine tundra, habitat of a family of Red-necked Stints, Kigluaik Mountains, Seward Peninsula, Alaska, June 2012. Photo by Lucas H. DeCicco ACKNOWLEDGMENTS We thank those who assisted in the 2012 field work: Fernando Diaz, Luke Eberhart- Phillips, Nicholas R. Hajdukovich, Andrea Minoletti, Daniel R. Ruthrauff, and Noah Warnock. We thank Pavel S. Tomkovich for providing information on the Russian breeding grounds. The editorial comments provided by Daniel D. Gibson, Daniel R. Ruthrauff, and Robert E. Gill improved this manuscript greatly, and they are much appreciated. The findings and conclusions in this article are those of the authors and do not necessarily represent the views of the U.S. Fish and Wildlife Service. Table 1 Documented Breeding of the Red-necked Stint in Alaska Record 0 Date Locality Evidence Reference 1 10 Jul 1908 Nome Pair and 2 young Thayer 1909 2 14 Jun 1922 Cape Prince of Wales Pair building nest Bailey 1926 3 23 Jun 1933 Cape Prince of Wales Nest with 4 eggs Ford 1934 4 15 Jun 1939 Cape Prince of Wales Set of eggs Hanna 1940 5 16 Jun 1941 Cape Prince of Wales Set of eggs Bailey 1948 6 14 Jun 1945 Cape Mountain, Wales Nest with 4 eggs Bailey 1948 7 20 Jun 1945 Cape Prince of Wales Set of eggs Bailey 1948 8 17 Jun 1946 Cape Prince of Wales Set of eggs Bailey 1948 9 21 Jun 1968 24 km NW of Nome Pair with 3 young Kessel 1989 10 23 Jun 1968 Safety Sound, Nome Adult with brood Kessel 1989 11 30 Jun 1974 Cape Mountain, Wales Nest with 4 eggs Kessel and Gibson 1978 12 28 Jun 1975 Barrow Nest Kessel and Gibson 1978 13 30 Jun 2012 Kigluaik Mountains Pair with 4 young this publication “Numbers correspond with locations mapped in Figure 1 . 277 HISTORY OF THE RED-NECKED STINT BREEDING IN NORTH AMERICA LITERATURE CITED Bailey, A. M. 1926. A report on the birds of northwestern Alaska and regions adjacent to Bering Strait, part VII. Condor 28:31-36. Bailey, A. M. 1948. Birds of arctic Alaska. Colo. Mus. Nat. Hist. Pop. Ser. 8:1-317. Cramp, S., and Simmons, K. E. L. (eds.). 1983. The Birds of the Western Palearctic, vol. 3. Oxford Univ. Press, Oxford, England. Ford, E. R. 1934. Rufous-necked Sandpiper nesting in Alaska. Auk 51:232. Hanna, W. C. 1940. Rufous-necked Sandpiper nesting on Seward Peninsula, Alaska. Condor 42:122-123. Hopkins, D. M. 1959. Cenozoic history of the Bering land bridge. Science 129:1519-1528. Kessel, B. 1979. Avian habitat classification for Alaska. Murrelet 60:86-94. Kessel, B. 1989. Birds of the Seward Peninsula, Alaska: Their Biogeography, Sea- sonality, and Natural History. Univ. Alaska Press, Fairbanks. Kessel, B., and Gibson, D. D. 1978. Status and distribution of Alaska birds. Studies Avian Biol. 1. Lappo, E. G., Tomkovich, P. S., and Syroechkovskiy, E. E., Jr. 2012. [Atlas of Breed- ing Waders in the Russian Arctic]. Institute of Geography, Russian Academy of Sciences, Moscow (in Russian with English summaries). Morozov, V. V., and Tomkovich, P. S. 1984. [Distribution and breeding habitats of the Red-necked Stint Calidris ruficollis (Pall.)]. Biol. Sciences 4:42-48 (in Russian). Piersma, T., van Gils, J., and Wiersma, P. 1996. Family Scolopacidae (sandpipers, snipes, and phalaropes), in Handbook of the Birds of the World (J. del Hoyo, A. Elliott, and J. Sargatal, eds.), vol. 3, pp. 444-487. Lynx Edicions, Barcelona. Ruthrauff, D. R., and McCaffery, B. J. 2005. Survival of Western Sandpiper broods on the Yukon-Kuskokwim Delta, Alaska. Condor 107:597-604. Portenko, L. A. 1972. Birds of the Chukchi Peninsula and Wrangel Island, vol. 1. Translation 1981 by Amerind Publishing Co. Pvt. Ltd., New Delhi. Thayer, J. E. 1909. Limonites ruficollis in Alaska. Condor 11:173. Accepted 1 7 August 2013 278 ESTIMATING THE NUMBER OF TERRITORIAL MALES IN LOW-DENSITY POPULATIONS OF THE SOOTY GROUSE JAMES D. BLAND, California Department of Fish and Wildlife, 1812 Ninth Street, Sacramento, California 95811; BlandJim@yahoo.com ABSTRACT: Sierra Sooty Grouse ( Dendragapus fuliginosus sierrae) are chal- lenging to census because they occur at low densities, are cryptically colored, and live quietly in the forest canopy most of the year. I developed a census method that ac- counts for several aspects of Sierra Sooty Grouse breeding biology that hinder accurate estimates, including seasonality of singing, anomalous singing by yearling males, low population density, and clumped dispersion of breeding males. Within 167 km 2 near Pinecrest, Tuolumne County, California, I conducted landscape-scale censuses along a network of line transects from 2006 to 2009 and detected 22 clusters of breeding males (hooting groups). I then used spot-mapping methods to estimate the number of individual males within hooting groups. Territorial display by transient (yearling) males lasted only a few days and became uncommon after 1 May; persistently territorial males became increasingly reluctant to display after mid-May. Thus limiting the census period to 1 May- 15 June maximizes detections of persistently territorial males, and a minimum interval of 5 days between repeated censuses minimizes misidentifica- tion of transient males as territorial. In the 13 hooting groups that I spot-mapped, the number of persistently territorial males averaged 4.9, and the distance from the center of a territory to the center of the nearest neighboring territory averaged 209 m. The probability of a persistently territorial male being detected on a single census visit averaged 0.71. Three repetitions of the group-scale census within a hooting season were sufficient to detect 98% of persistently territorial males. The density of territorial males was much lower (~0.6 male/km 2 ), and the distribution of males' territories was much more clumped, than reported in other regions. The number of persistently territorial males was static from 2009 to 2011. There is no standardized widely used method for censusing the Sooty Grouse ( Dendragapus fuliginosus), although many censuses have been undertaken on an ad hoc basis. Agencies mandated to monitor Sierra Sooty Grouse (D. /. sierrae) require efficient, unbiased methods of census. This subspecies is difficult to census because the birds are cryptic and remain quietly in the forest canopy most of the year, the population density is low, the distribution of breeding males is clumped, the frequency of song varies through the breeding season, and occasionally yearling males display terri- torially. I developed auditory census methods that account for these aspects of the grouse’s breeding biology and used them in repeated censuses in Tuolumne County in 2009 and 2011. The Sierra Sooty Grouse occupies relatively open and arid forests, unlike most of the other subspecies of the Sooty Grouse, which typically occupy humid coastal forests (Bendell and Zwickel 1984). Population densities of D. /. sierrae are also lower (Bendell and Zwickel 1984, Zwickel and Bendell 2004), and the subspecies is more arboreal than the others (most males’ breeding display occurs in the forest canopy). Bland and Gardner (2013) found that the Sierra Sooty Grouse is closely associated with large trees and mature forest, with an open canopy. Where forest cover is open and popula- tion densities are low, male Sooty Grouse congregate in spring at traditional Western Birds 44:279-293, 2013 279 CENSUS TECHNIQUES FOR THE SOOTY GROUSE breeding sites (Lewis 1985a), forming what Bendell and Elliott (1967) called hooting groups. Available evidence indicates this is the case throughout the range of the Sierra Sooty Grouse (Bland 1993, unpubl. data). Zwickel and Bendell (2004:157) reported that hooting groups usually consist of 2 or 3 males but can include up to 7 or 8. Hooting groups are not considered classical leks, primarily because they lack a communal display arena (Lewis 1985a). The locations of individuals’ territories, and by extension hooting areas, vary little from year to year (Bendell 1955, McNicholl 1978) or from generation to generation (males can live >14 years, Zwickel et al. 1989), until the group dies out or vegetation succession makes the site unsuitable (Zwickel and Bendell 1985). The clumped distribution of breeding males constrains the methods and efficiency of a census. Large areas of apparently suitable breeding habitat may be unoccupied and so unproductive to census. Conversely, detection of a single hooting male readily leads to detection of the other territorial males in a group because all engage in countersinging (Zwickel and Bendell 2004:157). In the Sierra Nevada, peak hooting occurs between mid-April and early June (Zwickel and Bendell 2004:156, Bland unpubl. data). During this period, territorial males usually hoot throughout the day, with peaks in hooting around sunrise and sunset (Bendell 1955, Stewart 1967, McNicholl 1978, Zwickel and Bendell 2004). Hooting is often audible at distances up to 500 m (Hjorth 1970) and up to 1 km when the sound is not attenuated by topography or air movement (Doerr at al. 1984, Bland unpubl. data). Quiet or resting territorial males hoot or flutter their wings in response to broadcast of recorded calls of a female (Stirling and Bendell 1966, McNicholl 1981, Niederleitner 1987) and to a human observer’s intrusion into their territory (Zwickel and Bendell 2004: 160), unless startled or threatened by the observer. Early in the hooting season, a few nonterritorial males (yearlings and nonterritorial adults, Zwickel and Bendell 2004: 154) also hoot for a few days, potentially inflating early-season counts of territorial males (McNicholl 1981). I refer to these as transient males, as opposed to persistently territorial males. Late in the hooting season, quiet or resting territorial males become reluctant to respond to recorded calls of a female (McNicholl 1981) or intru- sions by observers, potentially deflating late-season counts based on calls. There have been three previous censuses of the Sierra Sooty Grouse. Hoffmann (1956) conducted a repeated census at Sagehen Creek, but his results are doubtful because of his informal method and small sample size (<6 males). Bendell and Zwickel (1984) ranked Sooty Grouse densities on a subjective 0-5 scale at two locations. Bland (1993) counted the number of territorial males within hooting groups at six locations. Outside California, most published Sooty Grouse censuses have been based on labor-intensive banding or area searches, with the objective of achieving a total count (Bendell 1955, Bendell and Elliott 1967, Boag 1966, Donaldson and Bergerud 1974, Redfield 1975, Zwickel and Bendell 2004). The effort required to complete such censuses is difficult to determine in advance because it is predicated on gradual depletion of undetected individu- als. The roadside “call counts” conducted by some state game agencies are indices of populations of adult males at regional scales and are generally capable of detecting only large changes or differences (Zwickel 1982, Fox 280 CENSUS TECHNIQUES FOR THE SOOTY GROUSE et al. 2009). Zwickel (1982) provided the basic outline for an audio census of male Sooty Grouse but offered few procedural details. There remains a need for an efficient, standardized method for censusing the Sooty Grouse. The California Department of Fish and Wildlife (CDFW) requires popula- tion data for establishing bag limits for the Sooty Grouse and monitoring populations thought to be in decline, and the U.S. Forest Service (USFS) requires such data to support its bioregional Management Indicator Species program (USFS 2008). The objectives of this paper are to present my findings on aspects of the Sooty Grouse’s breeding biology that affect auditory censuses, describe my census procedures, and report the results of a repeated census and estimate of density. My census protocol could be adopted by agencies for local-scale monitoring, and the supporting research could serve as a methodological foundation for developing probabilistic census methods or monitoring pro- grams at a regional scale. STUDY AREA My censuses took place in an area of 167 km 2 centered near Pinecrest Peak, Stanislaus National Forest, Tuolumne County, California (38° 14' N, 119° 56' W), on the western slope of the Sierra Nevada (Figure 1). Peaks and ridges within the area reach 2600-2800 m elevation. I set the lower boundary of the study area at 1775 m elevation, -360 m below the pre- sumed lower limit of grouse in the region (Grinnell et al. 1918). In this area, the predominant forest types include Sierran mixed-conifer forest (<-2200 m), red fir ( Abies magnified) forest (-2200-2700 m), and lodgepole pine (. Pinus contorta )/ subalpine conifer forest (>-2700 m). The eastern third of the study area lies within Emigrant Wilderness Area and is relatively pristine. Elsewhere, logging, including clear-cutting, has been widespread but generally at a small scale (1-10 ha). The entire area is usually covered with 1-4 m of snow from late November through early April, and in 2011 it received unusually high snowfall (153% of normal in April, Cooperative Snow Survey, California Department of Water Resources). METHODS My census methods incorporate two complementary procedures, one for counting groups of hooting grouse at the landscape scale (landscape- scale transects), another for counting individual grouse within those groups (group-scale transects). I controlled for observers’ variability in skill and acu- ity of hearing by limiting the number of observers (four for landscape-scale censuses, three for group-scale censuses) and mentoring each observer on his or her first two censuses. Landscape-Scale Census Landscape-scale censuses extended from 15 April to 7 June, 2006-2009, along a network of line transects that intersected all forested portions of the study area (Figure 1). I laid out landscape-scale transects with GIS software 281 CENSUS TECHNIQUES FOR THE SOOTY GROUSE Figure 1. Pinecrest study area. Shaded area indicates >10 % forest cover; solid lines, landscape-scale census transects; circles enclosing stars, hooting groups censused in both 2009 and 2011; circles enclosing dots, hooting groups censused in 2009 only; solid dots, hooting groups assessed by area searches or incomplete censuses, 2006-2009; circle enclosing an x, a hypothetical hooting male, surrounded by gridded areas representing maximum distances of audibility of 500 m (Hjorth 1970) and 1000 m (Doerr et al. 1984). (Manifold version 8, Manifold Software Limited, Hong Kong) by drawing them over a topographic map of the study area. Each transect was drawn to minimize loss or gain of elevation, skirt dangerous cliffs and streams, and be traversable by foot within 4-6 hours (often by snowshoe). Adjacent transects, and parallel portions of loop transects, were spaced -400-800 m apart (measured as horizontal distance on a map) to ensure that more than one transect would fall within the range (-500-1000 m) of audibility of hooting grouse located anywhere in forested portions of the study area (Figure 1). Transect routes were downloaded to hand-held Global Positioning System (GPS) units (Model GPSMAP 60CSx, Gar min International, Inc., Olathe, KS), which were used in the field so that the predetermined routes could be navigated as closely as possible. A single observer walked each transect once, stopping every 300-500 m at prominent spurs or vantage points to listen for hooting grouse. If no hooting was audible, the observer broadcast a recorded 22-syllable cackle of a female (provided by J. Bendell, Univ. Toronto) in each cardinal direction by using an MP3 audio player (Model SA4111, Philips Electronics, Andover, MA) and amplifying speakers (Model 40-1441, Radio Shack, Fort Worth, TX). Where hooting was detected, the observer did an area search to locate as many displaying grouse in the area as possible, using recorded calls of a female when necessary to re-initiate display. I defined a hooting group as all males a person could hear displaying from at least one previously detected grouse’s territory (this definition works well where grouse densities are low 282 CENSUS TECHNIQUES FOR THE SOOTY GROUSE and predominant ridges are parallel, but it does not perform as well where densities are higher or ridges converge in a tight “U”). I considered both hooting and wing-fluttering to be territorial display behavior (Zwickel and Bendell 2004). The observer recorded coordinates of displaying grouse with a hand-held GPS unit while standing beneath the grouse’s songpost tree. If a displaying grouse could not be observed directly because of obstructing foliage, the location of its apparent songpost tree was recorded. Later I used location coordinates to lay out group-scale transects in GIS, as described be- low. After recording locations of all displaying grouse at the site, the observer continued along the landscape-scale transect. Landscape-scale censuses were conducted during all daylight hours, except during storms or when air movement was strong enough to attenuate or obscure distant display sounds. Group-Scale Census I also laid out group-scale transects in advance with GIS, using locations of displaying grouse recorded during landscape-scale censuses. Each group- scale transect was positioned on a georeferenced orthophotograph (digital orthophoto quarter quadrangle, Stanislaus National Forest). I then centered a transparent 100-m grid, with total dimensions of 1500 x 1500 m (an area slightly larger than the area occupied by a typical hooting group), over the locations of the grouse in the hooting group, then trimmed the photograph to the 1500 x 1500 m grid to create a map of the hooting group. I labeled the grid lines with the applicable Universal Transverse Mercator values and drew a group-scale transect line on the map so that it (1) wove through suc- cessive territories, passing by known songposts at a distance of -50-100 m, and (2) maintained a curtain of tree foliage between the observer and all known songposts. In 20 1 1 , the year of the second census, I also ensured that transects extended 100 m beyond the furthest known territory at each end of a transect, thereby positioning observers where any additional territorial males would likely be detected — territories that constitute a hooting group are typically distributed more or less linearly along an elevation contour (Bland and Gardner 2013). Transect routes were downloaded to hand-held GPS units, which observers used in the field to navigate the predetermined routes. An observer slowly walked the full length of a group-scale transect (larger groups required longer transects), then retraced it in the opposite direc- tion. The initial pass served to rouse as many individuals as possible into heightened territoriality and display. Where no hooting was audible within 200-300 m, the observer broadcast one or two cackle calls at ~150-m intervals to induce any nearby silent males to display (excessive use of re- corded calls can cause males to move, complicating spot-map analysis). If a displaying male was not visible from the transect (obscured by foliage), the observer estimated its location from the locations of display sounds from a series of points along the transect. Estimated locations are acceptable for such counts because males’ territories are exclusive and typically encompass 0.6-2. 1 ha (Zwickel and Bendell 2004:200). If necessary, unresponsive males were approached unobtrusively to no closer than -75 m and stimu- lated with a recorded cackle call of a female. Visual contact was avoided to ensure continued hooting, which increases display (hence, detection) of additional males (McNicholl 1978). Locations of displaying grouse were 283 CENSUS TECHNIQUES FOR THE SOOTY GROUSE marked by hand onto the hooting-group map. When an observer was un- certain whether a truncated hoot sequence or quick wing-flutter might have been detected, he or she noted the location on the hooting-group map as a possible detection. Songpost trees were usually distinguishable on aerial photographs by their large size. Group-scale censuses took place between 08:00 and 14:00, and occasionally in late afternoon, but never during storms or when a continuous strong wind would have attenuated or obscured display sounds. Prior to a repetition of a group-scale census, I realigned transect lines where necessary so that they passed within 50-100 m of newly detected territories and maintained a curtain of tree foliage between the observer and all known songposts. In 2009, 1 conducted group-scale censuses of 15 hooting groups on five occasions between 19 April and 9 June, 2009. The average interval between these censuses was 8.3 days and the minimum was 3 days, except for 1 day on one occasion and 2 on another. Five additional groups were censused only 1-4 times because of limited manpower, and 2 were assessed only by initial area searches during landscape-scale censuses (these 7 groups were omit- ted from group-scale analyses). In 2011, I conducted group-scale censuses from 4 May to 16 June at 11 of the 15 hooting groups censused in 2009. Census Timing A primary objective of the 2009 group-scale censuses was to document patterns of hooting activity through the peak hooting period, in order to determine the optimal period and time of day for future censuses. To identify when anomalous hooting by transient males was most frequent, I compared the frequency of one-time detections (transient males cease displaying after a few days) across a series of 10-day intervals. To determine the approximate date when persistently territorial males became too reluctant to display to be censused reliably, I analyzed the proportion of males detected only on return runs of censuses (were responsive only after neighboring males had been roused into display). Data Analysis I estimated the number of persistently territorial males constituting hooting groups with standard spot-mapping methods (Kendeigh 1944): by plotting detection locations on repeated censuses, identifying clusters of detections (within areas equal to a typical territory, ~1.5 ha), and using records of si- multaneous singing to distinguish between two or more adjacent territories. I considered males that displayed on any two repetitions of a group-scale census to be persistently territorial. I estimated territory centers from mean locations of spot-map detections, and hooting-group centers from mean lo- cations of territory centers. I used Clark and Evans’ (1954) nearest-neighbor method to assess the distribution of males’ territories because it has been used in studies of the Sooty Grouse previously published. Clark and Evans’ R provides a measure of the degree to which observed distribution patterns deviate from a randomly distributed population of equal density, where values range from zero to 2.1491, with zero indicating maximum clumping, one random spacing, and 2.1491 even spacing. 284 CENSUS TECHNIQUES FOR THE SOOTY GROUSE It was not feasible to estimate territory size because five repetitions of group-scale censuses resulted in too few detection locations. Instead, I mea- sured the maximum spread of detection points attributed to each territorial male, which served as a rough index of territory size. This was helpful for distinguishing between (tallying) territories on spot maps because detections that were too widely dispersed were unlikely to be from a single persistently territorial male during peak hooting season, although activity areas do ex- pand after peak hooting season (Lewis 1985b, Bland unpubl. data). When analyzing the probability of detection, I considered each census visit to a persistently territorial male to be an experimental trial. I estimated cumu- lative detection probability as P* = 1 - (n K 1 - p } ), where K = the number of census repetitions, and p = the probability of detection on census repetition j (MacKenzie et al. 2005). I used a chi-squared test to assess whether time of day (before 11:00 versus after 11:00) influenced detection rates. In order to determine the minimum number of census visits and minimum time interval between visits, I analyzed the 2009 group-scale data by using only three of the five repetitions of a census (four when it was necessary to determine if a male that had displayed on only the second or third census had persisted). I omitted census repetitions conducted earlier than 1 May, later than 15 June, and <5 days after a previous census (2 of the 15 hooting groups had to be omitted from this analysis because the number or timing of repetitions of the census could not be reconciled with the revised criteria). Because this approach proved to be effective and efficient (see Results), I used it for censuses in 2011 and to analyze census data from both years. RESULTS Landscape-Scale Census I detected 22 hooting groups in the study area, along 500 km of landscape- scale transects. The average distance between the center of a group to that of the nearest neighboring group was 1916 m (range 1239-3676 m, SD = 680). Group-Scale Detection Rates and Timing The probability of an individual territorial male being detected on a single census visit averaged 0.71 (2009 census, 3 repetitions, 61 individuals [2 omitted, see footnote e in Table 1], range 0.33-1.0, SD = 0.24). The cu- mulative probability of detection was 0.92 after the second repetition, 0.98 after the third repetition, and 0.99 after the fourth repetition. Persistently territorial males were equally detectable before and after 1 1:00 (x 2 = 0.273, P = 0.60; 195 trials between 07:00 and 10:59, 161 trials between 11:00 and 18:00). Because detection rates were high, three or four repetitions were sufficient to detect all, or nearly all, persistently territorial males (Figure 2). In 2009, I started group-scale censuses early in the hooting season (19 April), while display by transient males was still relatively common (McNicholl 1981). One-time detections of displaying males (presumed transients) were common in late April (23% of detections), but they declined rapidly by 1 May and remained relatively rare thereafter (5-9 % of detections, Figure 3). A start 285 CENSUS TECHNIQUES FOR THE SOOTY GROUSE date of 1 May therefore avoided most detections of transient males. When I specified a minimum of 3 days between censuses, four males that were de- tected twice within 3 days were misclassified as persistently territorial. When I specified a minimum interval between censuses of 5 days, these individuals were detected only once and so classified as transient. Persistently territorial males became increasingly reluctant to display after mid-May (Figure 3), but broadcast of calls of a female continued to stimulate them to display reliably Table 1 Numbers of Persistently Territorial Male Sooty Grouse Estimated in 2009 and 2011 Estimated number of persistently territorial males 0 Hooting group 2009 initiaP 2009 final 0 2011 d Punch Bowl 4 4 2 Pike’s Peak 9 10 e ’ f 7 Pinecrest Peak Trail 5 4 /. s 6 Gooseberry 7 6 f 7 Waterhouse Trail 8 h h Pinecrest Peak 10 10 f 10 f Bull Run Rock 3 h h Herring Cr. Reservoir 2 2 3 Aspen Meadow 7 5 / 5 Flying Horse Boneyard 6 6 4 Lift 8 2 2 i Mill Creek Ridge 3 3 3 Cow Creek 3 2 2 Mini-Gargoyles 5 5f i Gargoyles 4 4 3 15-group total (mean) 78 (5.2) — — 13-group total (mean) 67 (5.5) 63 (4.9) — 11 -group total (mean) — 56 (5.1) 52 (4.7) “Detected on at least two repetitions of a census. b Data recorded 19 April-9 June 2009, in five repetitions of the census and a minimum of 3 days between repetitions. “Data limited to 1 May-9 June (as early as 28 April in two cases) 2009, in three repetitions of the census (plus a fourth when necessary to confirm persistence of individuals detected on only the second or third census) and a minimum of 5 days between repetitions (to exclude transients). d Data recorded 4 May-16 June 2011, in three repetitions of the census and a minimum of 5 days between repetitions. “Includes two individuals that were first detected at the periphery of the hooting group on the fifth census. Subsequent observations outside formal censuses confirmed these individuals per- sisted for at least two more weeks. This shortcoming in placement of transects was remedied by amending the protocol to ensure that group-scale transects extended 100 m beyond the furthest known territories at each end of a transect. ^One-time detection on the second or third census necessitated addition of a fourth census. sTime did not permit an additional census, so an available census preceding peak hooting was used. h Omitted from analysis because the 5 repetitions of the original census could not be reconciled with the criteria for number or timing of repetitions in the final protocol. 'Too remote to be recensused with the manpower available in 2011. 286 CENSUS TECHNIQUES FOR THE SOOTY GROUSE until ~15 June, so the dates for the end of the censuses I used in 2009 (9 June) and 2011 (16 June) were appropriate (also see Stirling and Bendell 1966). Anecdotally, I observed that the final decline in hooting activity ap- peared to coincide with late-morning temperatures rising above ~20 °C. Unusually deep accumulation of snow did not delay peak hooting season in 2011; the likelihood of persistently territorial males being detected during the first 15 days of May (the earliest period for which data were available in both years) was actually higher in 2011 than in 2009 (x 2 = 4.77 , P = 0.029). Census Estimates Without correction for census timing, the 2009 group-scale censuses re- sulted in an estimate of 78 persistently territorial males at 15 hooting groups (Table 1) and 17 one-time detections of transient males. Corrected for timing, the estimate for 13 of these groups (two could not be reconciled with the corrected timing) was 63 persistently territorial males (Table 1), with an aver- age group size of 4.9 individuals (range 2-10, SD = 2.7) and 13 one-time detections of transient males. A fourth repetition of the group-scale census was necessary at 6 (46%) of the 13 groups. The average distance between centers of nearest neighboring territories within a group was 209 m (range 75-638 m, SD = 166 m), and the spread of points for an individual male averaged 85.9 m (SD = 44.0 m, 2-5 detections/individual). The 9 groups Figure 2. Proportion of persistently territorial males detected and confirmed over 4 successive censuses. Solid line represents initial detection; dashed line, second (confirmatory) detection (2009 census data, 13 hooting groups, 61 individuals [2 omitted], minimum of 5 days between each repetition of the census). 287 CENSUS TECHNIQUES FOR THE SOOTY GROUSE 1-10 May 21-30 May 11-20 Jun Figure 3. Seasonal patterns of propensity of transient and persistently territorial males to display (data recorded earlier in 2009 than in 2011). Solid line represents the proportion of males detected displaying on only one occasion in 2009 (presumed yearlings or nonterritorial adults); dashed line, the proportion of persistently territorial males displaying only on the return run of censuses in 2011 (after the entire group had been stimulated with broadcasts of the call of a female). for which censuses were incomplete contained an estimated 38 territorial males, with an average group size of 4.2 individuals (range 1-11, mode = 3, SD = 3.3). The density of breeding males across the entire 167-km 2 study area was roughly 0.6 Am 2 , based on the combined results of group-scale censuses at 13 groups in 2009 and area searches or incomplete censuses at the 9 remaining groups from 2006 to 2009. The distribution of males’ territo- ries (those detected by area searches, complete censuses, and incomplete censuses pooled), was highly clumped at the landscape scale ( R = 0.42, c = 11.12, P = <0.001), although the significance of this test could be somewhat less than calculated because population size was estimated (Clark and Evans 1954). In 201 1, 1 recensused 11 of the 13 hooting groups censused in 2009, and estimated there were 52 persistently territorial males, a result not significantly different from the 2009 estimate for those 11 hooting groups (x 2 = 0.308, P= 0.579, Table 1 ) . There were just 8 one-time detections of transient males in 2011, and a fourth census was necessary at only 1 (9%) of the groups. The similarity in 2009 and 2011 estimates suggests the Pinecrest population of the Sooty Grouse remained static over those 3 years. 288 CENSUS TECHNIQUES FOR THE SOOTY GROUSE DISCUSSION Variation in Song Frequency, Detectability, and Environmental Interference Even among persistently territorial males the frequency of display is variable (Stirling and Bendell 1966, McNicholl 1978, Niederleitner 1987). However, most published studies of diurnal variation in hooting have been conducted under natural circumstances (Bendell 1955, Stewart 1967, McNicholl 1978), so the patterns reported do not apply to audio-simulated grouse. Stirling and Bendell (1966), for example, found that between the hours of 10:00 and 14:00, when the frequency of hooting among unstimulated males was relatively low, the broadcasting of recorded calls of a female increased the count of hooting males by 110%. Under normal circumstances (without audio stimulation), territorial males cease displaying when they are resting, foraging, or engaged in maintenance activities. When a predator or intruder appears (conspecific or otherwise, including human) a male’s response can range from silence to heightened display, depending on the type and mag- nitude of the threat and past habituation (McNicholl 1983). Some studies suggest males with a lower social status or lower-quality territory display less vigorously (without audio stimulation, McNicholl 1978, Lewis 1986). My findings show that when recorded cackle calls of a female are broadcast properly at the scale of a group census (1) the average probability of detect- ing a persistently territorial male increases to 0.71 (range 0.33-1.0), (2) all territorial males can be detected in three or four repetitions of the census, and (3) audio stimulation is equally effective in the morning and afternoon. In the only similar study, Stirling and Bendell (1966) found broadcast of a female’s calls sufficiently effective for all territorial males within audible range to be detected in <2 man-days of effort (three repetitions of the census per day). Failure to detect displaying males is of greatest concern with landscape- scale censuses because they are conducted only once and depend on au- dibility beyond 300 m. However, because the objective of landscape-scale censuses is to detect groups of countersinging males rather than individuals, success is achieved by detection of even a single member of a group. If the probability of detecting a single male at the scale of the group is 0.71, the probability of detecting at least one male in a group of audio-stimulated males must also be very high, even though detection distances are greater and sound propagation is less consistent. Furthermore, because landscape-scale transects are spaced -400-800 m apart, the 1000-m-diameter “audible zone’’ surrounding each group is traversed by an observer three or four times (Figure 1, x for hypothetical hooting male). To control for attenuating topography, I aligned landscape-scale transects with dominant ridges and spaced them <800 m apart. Two lines of anecdotal evidence from concurrent field work on the Sooty Grouse indicate my landscape-scale censuses were effective. Radiotelemetry technicians who traversed much of the study area during the hooting seasons of 2007-2009 reported no additional hooting sites. Similarly, habitat technicians who searched for fecal droppings in 138 plots of 0.1 ha distributed evenly throughout the study area identified just 4 or 5 sites where the number, diameter, and texture of droppings (Zwickel and Bendell 2004:136) suggested persistent springtime roosting. I visited 289 CENSUS TECHNIQUES FOR THE SOOTY GROUSE each of these areas during the subsequent hooting season and detected no hooting grouse. I believe if any hooting groups went undetected during landscape-scale censuses they were small, comprising only one or two in- dividuals. Males that have few other hooting grouse nearby tend to display less frequently (pers. obs.), presumably because they are less stimulated. Omission of one or two groups of one or two individuals would have had relatively little effect on the total number estimated. Number and Distribution of Males The overall density of the Pinecrest Sooty Grouse population, roughly 0.6 male/km 2 , is much lower than the 10-30 males/km 2 typically reported elsewhere (primarily British Columbia, Zwickel and Bendell 2004:212). The only published report of a lower density, 0.16 male Am 2 , is from a clearcut in coastal Alaska (Doerr et al. 1984). Most areas in which grouse breed at Pinecrest are in unharvested or selectively harvested forest (Bland and Gardner 2013). The only previous California estimate is 2 males Am 2 , from the coastal northwest (D. /. fuliginosus, Zwickel and Bendell 2004). Bendell and Zwickel (1984) attributed the relatively low densities of southern populations to lower interspersion of forbs, shrubs, grasses, and coniferous and broad-leaved trees. At Pinecrest these components of vegetation are well interspersed, but at a smaller patch scale (Franklin et al. 2002) than in clearcut forests or mosaics of habitats farther north, where most studies of the Sooty Grouse have taken place. The distribution of males’ territories at Pinecrest is more clumped (aggre- gated) than reported elsewhere. An R value of 0.42 indicates the average distance between nearest neighbors is 42% of what would be expected if territories were randomly distributed across the study area. In two previous studies (both on Vancouver Island), territory spacing was found to be uni- form ( R = 1.32-1.98; Bendell and Elliott 1967, Lewis and Zwickel 1981) or random {R = 0.99-1.28; Bendell and Elliott 1967). However, the areas covered in these studies were only 0. 1-2.9% the size of my Pinecrest study area. When I assessed the distribution of territories at Pinecrest at the scale of the smaller plots used in these studies, it also appeared to be random (or indeterminable because a plot encompassed only a single territory). Lewis (1985a) argued that clumped distribution results from a patchy distribution of the components of high-quality habitat, rather than breeding behavior rep- resenting an exploded lek, as suggested by McNicholl (1978). At Pinecrest, males’ territories are typically associated with openings in a more or less continuous forest canopy (Bland and Gardner 2013), and a spatially explicit habitat model indicates suitable habitat for breeding males is widespread, yet much is unoccupied (Bland and Gardner 2013). Uncensused Population Components My census methods estimate the number of persistently territorial males, not the total grouse population. Females, nondisplaying adult males, and yearling males cannot be censused reliably by visual or auditory means (Zwickel 1982). On the basis of banding studies or hunters’ harvest, females typically constitute -47-53% of the population under normal circumstances 290 CENSUS TECHNIQUES FOR THE SOOTY GROUSE (Zwickel and Bendell 2004). In coastal British Columbia, Bendell and Elliott (1967) and Redfield (1975) reported that -11-36% of the males are year- lings, with the proportion highest in areas where the population is sparse or increasing. Under normal circumstances, about 4% of yearling males display (Zwickel and Bendell 2004). A few nonterritorial, nondisplaying adult males have also been identified in intensively studied populations, but their relative abundance remains poorly known (Bendell and Elliott 1967, Lewis 1984). Management Implications The census method I have developed is an effective and relatively efficient means of monitoring Sooty Grouse populations, applicable to assessing the effects of development and forest management at a local scale. Monitoring at a regional scale, however, requires less labor-intensive methods. The method most widely advocated for regional population monitoring is sampling by point counts (Ralph et al. 1995), which has been widely adopted for multispecies monitoring (Manley et al. 2004). Point counts, however, were designed pri- marily for passerines, and are not well suited for the Sooty Grouse. Ideally, the findings of this study would be used to develop a regional monitoring program specifically for the Sooty Grouse. Alternatively, they could be used to design supplemental procedures for existing multispecies point-count programs, thereby improving their effectiveness for tracking Sooty Grouse populations. ACKNOWLEDGMENTS Financial support was provided by the U.S. Fish and Wildlife Service’s State Wildlife Grants Program, CDFW’s Game Bird Heritage Program, and the USFS’s Region 5 Management Indicator Species and Get Wild programs. Administrative support was provided by Scott Gardner and Jesus Garcia of CDFW and Adam Rich of Stanislaus National Forest. Kristy Blackburn, Chance Hildreth, Adrian Taylor, Alex Taylor, and Seth Taylor assisted with census work, and Tom Gardali and Brent Campos provided helpful editorial suggestions. LITERATURE CITED Bendell, J. F. 1955. Age, breeding behavior and migration of Sooty Grouse. Trans. N. Am. Wildlife Conf. 20:367-381. Bendell, J. F., and Elliott, P. W. 1967. Behavior and the regulation of numbers in Blue Grouse. Can. Wildlife Serv. Rep, Ser. 4. Canadian Wildlife Service, Environment Canada, Ottawa, Ontario. Bendell, J., and Zwickel F. C. 1984. A survey of the biology, ecology, abundance, and distribution of the Blue Grouse (genus Dendragapus), in Third International Grouse Symposium (P. J. Hudson and T. W. I. Lovel, eds.), pp. 163-192. World Pheasant Assoc., Reading, England. Bland, J. D. 1993. Forest grouse and Mountain Quail investigations: A final report for work completed during the summer of 1992. Report to Wildlife Mgmt. Div., Calif. Dept. Fish and Wildlife, 1812 Ninth St, Sacramento, CA 95811. Bland, J. D., and Gardner, S. 2013. Habitat assessment and monitoring protocol for Sooty Grouse ( Dendragapus fuliginosus) in the Sierra Nevada. State Wildlife Grant T-16, Segment 1 (F07AF00074). Report to Wildlife Branch, Calif. Dept. Fish and Wildlife, 1812 Ninth St., Sacramento, CA 95811. Boag, D. A. 1966. Population attributes of Blue Grouse in southwestern Alberta. Can. J. Zool. 44:799-814. 291 CENSUS TECHNIQUES FOR THE SOOTY GROUSE Clark, P. J., and Evans, F. C. 1954. Distance to nearest neighbor as a measure of spatial relationships in populations. Ecology 35:445-453. Doerr, J. G., Brighenti, J. M., Barescu, C. L., and Morin, M. 1984. Use of clearcut and old-growth forests by male Blue Grouse in central southeast Alaska, in Fish and Wildlife Relationships in Old-Growth Forests: Proceedings of a Symposium Held in Juneau, Alaska, 12-15 April, 1982 (W. R. Meehan, T. R. Merrell, and T. A. Hanley, eds.), pp. 309-313. Am. Inst. Fishery Res Biol., Juneau, AK. Donaldson, J. L., and Bergerud, L. T. 1974. Behavior and habitat selection of an insular population of Blue Grouse. Syesis 7:115-127. Fox, T. T., B. D. Dugger, and K. M. Dugger. 2009. Pilot study: Survey of Sooty Grouse (Dendragapus fuliginosus fuliginosus) abundance in western Oregon. Report to Dept. Fisheries and Wildlife, Nash Hall 104, Ore. State Univ., Corval- lis, OR 97331. Franklin, J. F., T. A. Spies, R. Van Pelt, A. B. Carey, D. A. Thornburgh, D. R. Berg, D. B. Lindenmayer, M. E. Harmon, W. S. Keeton, D. C. Shaw, K. Bible, and J. Chen. 2002. Disturbance and structural development of natural forest ecosystems with silvicultural implications, using Douglas-fir forest as an example. Forest Ecol. Mgmt. 155:399-423. Grinnell, J., Bryant, H. C., and Storer, T. I. 1918. The Game Birds of California. Univ. Calif. Press, Berkeley. Hjorth, I. 1970. Reproductive behaviour in Tetraonidae with special reference to males. Viltrevy 7:183-596. Hoffmann, R. S. 1956. Observations on a Sooty Grouse population at Sage Hen Creek, California. Condor 5:321-337. Kendeigh, S. C. 1944. Measurement of bird populations. Ecol. Monogr. 14:69-105. Lewis, R. A. 1984. Non-territorial adult males and breeding densities of Blue Grouse. Wilson Bull. 96:723-725. Lewis, R. A. 1985a. Do Blue Grouse form leks? Auk 102:180-184. Lewis, R. A. 1985b. Use of space by territorial male Blue Grouse. Wilson Bull. 97:97-101. Lewis, R. A. 1986. Aggressiveness, incidence of singing, and territory quality of male Blue Grouse. Can. J. Zool. 64:1426-1429. Lewis, R. A., and Zwickel, F. C. 1981. Differential use of territorial sites by male Blue Grouse. Condor 83:171-176. MacKenzie, D. I., Nichols, J. D., Royle, A., Pollock, K. H., Bailey L. L., and Hines L. E. 2005. Occupancy Estimation and Modeling: Inferring Patterns and Dynamics of Species Occurrence. Elsevier Academic Press, Oxford, England. Manley, P. N., Zielinski, W. J., Schlesinger, M. D., and Mori, S. R. 2004. Evaluation of a multiple-species approach to monitoring species and ecosystem conditions at the eco regional scale. Ecol. Appl. 14:296-310. McNicholl, M. K. 1978. Behaviour and social organization in a population of Blue Grouse on Vancouver Island. Ph.D. dissertation, Univ. Alberta, Edmonton. McNicholl, M. K. 1981. Caution needed in use of playbacks to census bird popula- tions. Am. Birds 35:235-236. McNicholl, M. K. 1983. Reactions of male Blue Grouse to intrusions by an observer. J. Field Ornithol. 54:77-83. Niederleitner, J. F. 1987. Use of early successional, midsuccessional, and old-growth forests by breeding Blue Grouse ( Dendragapus obscurus fuliginosus) on Hard- wicke Island, British Columbia. Can. J. Zool. 65:151-155. Ralph, C. J., Sauer, J. R., and Droege, S. 1995. Monitoring bird populations by point counts. Gen. Tech. Rep. PSW-GTR-149, U.S. Forest Service, Albany, CA. Redfield, J. A. 1975. Comparative demography of increasing and stable populations of Blue Grouse. Can. J. Zool. 53:1-11. 292 CENSUS TECHNIQUES FOR THE SOOTY GROUSE Stewart, P. A. 1967. Hooting of Sitka Blue Grouse in relation to weather, season, and time of day. J. Wildlife Mgmt. 31:28-34. Stirling, I., and Bendell, J. F. 1966. Census of Blue Grouse with recorded calls of a female. J. Wildlife Mgmt. 30:184-187. U. S. Forest Service. 2008. Sierra Nevada management indicator species monitoring implementation package. U. S. Forest Service, 1323 Club Dr., Vallejo, CA 94952. Zwickel, F. C. 1982. Blue Grouse, in CRC Handbook of Census Methods for Ter- restrial Vertebrates (D. E. Davis, ed.), pp. 63-65. CRC Press, Boca Raton, FL. Zwickel, F. C., and Bendell, J. F. 1985. Blue Grouse — effects on, and influences of, a changing forest. Forestry Chronicle 61:185-188. Zwickel, F. C., and Bendell, J. F. 2004. Blue Grouse: Their Biology and Natural His- tory. NRC Research Press, Ottawa, Canada. Zwickel, F. C., Boag, D. A., and Bendell, J. F. 1989. Longevity in Blue Grouse. N. Am. Bird Bander 14:1-4. Accepted 19 September 2013 Female Mount Pinos Sooty Grouse ( Dendragapus fuliginosus howardi), Giant Sequoia National Monument, California, 31 May 1992. Photo by James D. Bland 293 APPARENT EXTIRPATION OF THE SOOTY GROUSE FROM THE SKY ISLANDS OF SOUTH-CENTRAL CALIFORNIA JAMES D. BLAND, Department of Life Sciences, Santa Monica College, 1900 Pico Boulevard, Santa Monica, California 90405 and California Department of Pish and Wildlife, 1812 Ninth Street, Sacramento, California 95811); Bland_jim@yahoo.com ABSTRACT: The Mount Pinos Sooty Grouse (Dendragapus fuliginosus howardi) is endemic to south-central California and a species of special concern to the Cali- fornia Department of Pish and Wildlife. Historically, it ranged from Kings Canyon in the southern Sierra Nevada of Presno County (36° 45' N) south and west to the Mt. Pinos region of Kern and Ventura counties (34° 46.5' N). On the sky islands of Kern and Ventura counties it has not been sighted since 1993. Between 2002 and 2009, I surveyed in all known or potential historic habitat on these sky islands in spring, when males sing. I found no evidence of grouse in that region but did confirm the southernmost breeding sites in the main southern Sierra Nevada. Extirpation from the sky islands appears to have coincided with a proliferation of livestock grazing, timber harvesting, rural development, and fire suppression. Perhaps these activities altered the spatial pattern of seasonal habitats, increasing the grouse’s exposure to predation, or perhaps the removal of large trees, which the grouse uses as territorial songposts, rendered the sky islands unsuitable. The Mount Pinos Sooty Grouse ( Dendragapus fuliginosus howardi), endemic to California, historically ranged from the vicinity of Kings Canyon in the southern Sierra Nevada of Fresno County (36° 45' N) south and west to the Mt. Pinos region of Kern and Ventura counties (34° 46.5' N) (Grin- nell and Miller 1944). Reports from the San Jacinto Mountains, Riverside County (27 May 1971) and Big Pine Mountain, Santa Barbara County (8 June 1938) are anomalous and too poorly supported to be reliable (Garrett and Dunn 1981, Lentz 1993). North of Kings Canyon, subspecies howardi is replaced by the Sierra Sooty Grouse (D. /. sierrae). Dickey and van Ros- sem (1923) described howardi, distinguishing it from sierrae primarily on the basis of tail measurements (tail longer and more graduated) and plum- age color and pattern (paler above, with coarser vermiculation and barring, and darker and browner below). In California, the Sooty Grouse is closely associated with forests where firs {Abies, Pseudotsuga) are an important component of the canopy (Grinnell and Miller 1944). Fir needles and buds are the species’ primary foods, especially in winter (Zwickel and Bendell 2005). In the southern Sierra Nevada, north of Kern Gap (South Fork of the Kern River), where its habitat is generally contiguous, it occurs in small numbers at suitable locations (Bendell and Zwickel 1984, Bland 2008). South of Kern Gap, where its habitat is limited to isolated mountaintops in the extreme southern Sierra Nevada and northern Transverse Ranges — sky islands — it is considered extremely rare or possibly extinct (AOU 1957, Bland 2008). In 2005, the U. S. Forest Service (2005) assessed the Sooty Grouse as “historic/ potential” in the Mt. Pinos area but “apparently secure” elsewhere. In 2008, the California Department of Fish and Wildlife classified D. /. howardi as a species of special concern throughout its range (Bland 2008). In this paper I report the results my surveys for the Mount Pinos 294 Western Birds 44:294-308, 2013 EXTIRPATION OF THE SOOTY GROUSE SOUTH OF THE SIERRA NEVADA Sooty Grouse, review the scant literature on its historic status and habitat associations, and infer factors contributing to its decline in the sky islands. STUDY AREA The principal survey area was centered ~125 km northwest of Los An- geles, California, and spanned an arc of sky islands ~135 km long from the Piute Mountains in north-central Kern County to Pine Mountain in west- central Ventura County (Figure 1). I also searched immediately north of this area (north of Kern Gap, Figure 1) to confirm the southernmost breeding sites in the main Sierra Nevada (southern Tulare County). I delineated 10 survey sites in the sky islands by plotting a contour line at 1830 m (6000 ft) elevation, the approximate lower limit of white fir (A. concolor-, Mayer and Laudenslayer 1988) around all mountains in the grouse’s historic and potential range (from north to south, the Piute Mountains, Breckenridge Mountain, Bear Mountain, Tehachapi Mountains, Liebre Twins, Tecuya Mountain, Mt. Pinos and nearby peaks, Frazier Mountain, Alamo Mountain, and Pine Mountain). The Mt. Pinos survey site encompassed Mt. Pinos and Sawmill Mountain, Grouse Mountain, and Cerro Noroeste (Mt. Abel). All four of these peaks are located within an area approximately equal to the average annual home range of the Sierra Sooty Grouse (~12 km 2 , Bland and Gardner 2013). In the Tehachapi Range, I delineated two survey sites, an eastern site encompassing Tehachapi Peak and Cummings Mountain and a western site at Liebre Twins (Blue Ridge) on Tejon Ranch. The types of conifer forest represented within the survey area included fir, mixed conifer- fir, mixed conifer-pine, ponderosa pine ( Pinus ponderosa)-fiY, sugar pine (. P. lambertiana), Jeffrey pine {P jeffreyi), and ponderosa pine (vegetation “alliances” on the U. S. Forest Service’s vegetation maps). I consider the first five of these to be potential breeding habitat. Forests dominated by Jeffrey pine and ponderosa pine are generally too arid to be good breeding habitat. METHODS I surveyed between 15 April and 31 May, 2002-2005 and 2009, dur- ing peak “hooting” season for the Sooty Grouse (~mid-April-mid-June in central California, Bland 2013), when territorial males hoot throughout the day (Stewart 1967, Zwickel and Bendell 2004:156). Museum collectors of the 20 th century recognized peak hooting as a good time to search for the Mount Pinos Sooty Grouse: 81% of the sky island specimens were collected from 15 May to 15 June (Table 1). The full hooting season extends from about March to August, with some regional variation (Zwickel and Bendell 2004:156). When actively singing, males emit a sequence of 5-7 hoots every 1-10 min; the hoots are audible for at least 300 m and as far as 1 km under ideal conditions (Doerr at al. 1984, Zwickel and Bendell 2004:151, Bland unpubl. data). I used a survey technique first reported by Niederleit- ner (1987), which I have adapted, standardized, and used successfully with low-density populations in the Sierra Nevada (Bland 2013), including those of D. /. howardi in the southern Sierra Nevada (Bland 1993). I walked (or occasionally drove) an elevation contour through suitable breeding habitat, 295 EXTIRPATION OF THE SOOTY GROUSE SOUTH OF THE SIERRA NEVADA Piute Mountains Breckenridge Mountain Kern Co. Bear Mountain Tehachapi , Tehachapi Mt Pinos and \ & Mountains adjoining peaks Tecuya -3^ Noroeste 1 Sawmill Mtn Mountain A > W'~ Liebre Twins ... Frazier Mt Pinos A . Mountain Pine Mountain Ventura Co. Alamo Mountain Los Angeles Co. \ Figure 1. Survey area for the Mount Pinos Sooty Grouse. Names in bold are survey sites; thick solid lines, survey routes; thin solid lines, 1830 m elevation contour; thin dashed lines, mountains too arid to qualify for survey; circles enclosing dots, confirmed hooting sites. stopping every 300to500mat prominent spurs or vantage points to lis- ten for hooting grouse and then to broadcast a recorded call of a female in each cardinal direction (22-syllable cackle call provided by J. Bendell, Univ. Toronto). Recorded calls of a female stimulate quiet territorial males to sing (Stirling and Bendell 1966, Bland 2013). As I walked, I scanned the ground for fecal droppings and feathers, especially under large trees. I initiated surveys before 08:00, took a 1- to 2-hr break mid-day, and continued until a given transect was completed. Where patches of good habitat extended more than -1000 m up or down 296 Table 1 Specimens of the Mount Pinos Sooty Grouse Collected from the Sky Islands and Cataloged in the Ornithological Information System 0 Collection date Sex Collection location Specimen no. b 10 July 1903 M Tehachapi Peak USNM 186760 2 July 1921 M Mt. Pinos UCLA 5388 2 July 1921 F Mt. Pinos UCLA 5389 2 July 1921 M Mt. Pinos UCLA 5390 28 May 1922 M Mt. Pinos CAS 66681 28 May 1922 M Mt. Pinos UCLA 6608 28 May 1922 F Mt. Pinos FMNH 157167 28 May 1922 M Mt. Pinos UCLA 6609 30 May 1922 M Mt. Pinos UCLA 6610 30 May 1922 M Mt. Pinos UMMZ 122240 15 May 1923 M Mt. Pinos UCLA 12691 17 May 1923 M Mt. Pinos MVZ 100092 18 May 1923 M Mt. Pinos AMNH 750626 18 May 1923 M Mt. Pinos UCLA 12709 5 June 1927 (chick) Mt. Pinos UCLA 21838 21 May 1928 (egg set) Mt. Pinos, north slope WFVZ 3098 29 May 1928 F Mt. Pinos UCLA 22588 29 May 1928 F Mt. Pinos, near summit MVZ 100091 29 May 1928 M (chick) Mt. Pinos UMMZ 122350 29 May 1928 F Mt. Pinos MCZ 253515 29 May 1928 F Mt. Pinos UCLA 22587 29 May 1928 F Mt. Pinos UCLA 22589 30 May 1928 M Mt. Pinos MCZ 253516 30 May 1928 M Mt. Pinos, north slope UCLA 22590 30 May 1928 M Mt. Pinos, north slope UCLA 22591 30 May 1928 F Mt. Pinos, north slope UCLA 22592 31 May 1928 M Mt. Pinos, north slope UCLA 22593 31 May 1928 M Mt. Pinos, north slope UCLA 22596 31 May 1928 M Mt. Pinos MCZ 253517 28 May 1929 M Mt. Pinos, 1.5 mi ESE of summit MVZ 54125 11 June 1929 M Mt. Pinos, 2 mi NE of summit MVZ 54126 11 June 1929 F Mt. Pinos, 2 mi NE of summit MVZ 54127 12 June 1929 M Mt. Pinos, 2 mi NE of summit MVZ 54128 13 June 1929 F Mt. Pinos, 1 mi NE of summit MVZ 54129 15 June 1929 F Mt. Pinos UCLA 27464 15 June 1929 M Mt. Pinos UCLA 27469 15 June 1929 F Mt. Pinos UCLA 27470 29 June 1929 F Mt. Pinos UCLA 27471 26 May 1930 (skeleton) Mt. Pinos LACM 21093 11 July 1931 F Head of Cuddy Valley, ~2 mi. ENE MVZ 54131 of Mt. Pinos 2 Dec. 1933 F Mt. Pinos, summit FMNH 157168 23 Sep. 1976 F (feather) Mt. Abel (Cerro Noroeste), north SBNHM AVI 0506 slope, 3-3.5 mi. WNW of Mt. Pinos "Accessed through www.ornisnet.org, 30 June 2013. Richardson (1904) reported “one was taken” in the Piute Mountains “during the summer of 1903.” However, no corresponding specimen exists in any major public museum, including the Museum of Vertebrate Zoology (C. Cicero pers. comm.), where Richardson was later employed. This bird might not have been prepared as a study skin. b AMNH, American Museum of Natural History, New York; CAS, California Academy of Sci- ences, San Francisco; FMNH, Field Museum, Chicago; LACM, Natural History Museum of Los Angeles County, Los Angeles; MCZ, Museum of Comparative Zoology, Harvard Univer- sity, Cambridge, MA; MVZ, Museum of Vertebrate Zoology, University of California, Berkeley; SBNHM, Santa Barbara Museum of Natural History, Santa Barbara; Dickey Collection, Uni- versity of California, Los Angeles; UMMZ, University of Michigan Museum of Zoology, Ann Arbor; USNM, National Museum of Natural History, Smithsonian Institution, Washington, DC. 297 EXTIRPATION OF THE SOOTY GROUSE SOUTH OF THE SIERRA NEVADA slope, I established an additional transect parallel to the first ~400 m further into the patch. On the steep north face of the ridge comprising Mt. Pinos, Sawmill Mountain, Grouse Mountain, and Cerro Noroeste, where many of the most reliable historic observations originated, I surveyed three parallel transects over two breeding seasons (2002-2003): one along the upper rim (-777-807 m elevation), a second -200 m downslope, and a third where patches of appropriate habitat were present -500-1200 m below the sum- mits of Mt. Pinos and Sawmill Mountain. North of the Kern Gap, I used line-transect methods as described above in the Greenhorn Mountains from Portuguese Pass south to Woodward Peak (23-24 April 2002) and located survey stations where habitat appeared suitable along major Forest Service roads in the main Sierra Nevada from Sherman Pass south to Bartolas Creek (29-31 May 2004). For the historical review, I consulted published literature on the grouse’s status and habitat associations and www.ornisnet.org for records of speci- mens held by major North American museums. RESULTS Survey I completed 190 km of survey transects throughout the sky islands but detected no Sooty Grouse vocalizations, feathers, or fecal droppings (maps and GPS tracks of survey routes are available by request). North of the Kern Gap, I confirmed singing males at Sunday Peak in the southern Greenhorn Mountains (24 April 2002) and at Cherry Hill in the main Sierra Nevada (29 May 2004, Figure 1). These sites were the grouse’s southernmost sup- posed breeding sites at the time of the surveys (B. Barnes and T. Benson pers. comm.). Historic Status Little information is available regarding the historic status of the Mount Pinos Sooty Grouse, particularly south of Kern Gap. North of Kern Gap, the Death Valley Expedition of 1891 (Fisher 1893) reported that Sooty Grouse were “nowhere common” (the expedition observed or collected the species at Monache Meadow, Independence Creek, Big Cottonwood Meadow, Olancha Peak, Halsted Meadow, Horse Corral Meadow, Mineral King, and elsewhere). Grinnell and Miller (1944) considered it “locally common in suitable parts of [the] main southern Sierra Nevada.” In the late 1970s, Bendell and Zwickel (1984) ranked the density of Sooty Grouse at Crescent Meadow, northern Tulare County, as 3 on a scale of 0-5. In 1992, 1 counted only 0.5 hooting male/km along 24.7 km of line transects through known breeding habitat in northern Tulare County (Bland 1993). Most information regarding the historic status of D. /. howardi south of the Kern Gap comes from the vicinity of Mt. Pinos. Grinnell and Miller (1944) surmised it was “always, within history, sparsely represented on the more southwesterly, outlying islands of occurrence." Presumably the “more southwesterly, outlying islands” included Mt. Pinos and closely adjoining Sawmill and Grouse mountains, Cerro Noroeste, Frazier Mountain, and 298 EXTIRPATION OF THE SOOTY GROUSE SOUTH OF THE SIERRA NEVADA possibly the Tehachapi Mountains. Perhaps Grinnell and Miller reserved comment on status on the “northeasterly” islands (e.g., Piute Mountains, Breckenridge Mountain) because reliable information was not available. The only historic report from the Piute Mountains is by Richardson (1904), who observed Sooty Grouse “several times” in the summer of 1903. The only historic report from the Tehachapi Mountains is by L. Goldman, who observed and collected just a single male (Table 1) during three days’ survey work “on the summit and lower slopes” (Smithsonian Institution archives, collection 7176, box 37, folders 7 and 8). At Mt. Pinos, E. Nelson was the first to report “a few about the summit” in October 1891 (Fisher 1893). In autumn, Sooty Grouse are generally difficult to detect, being cryptic and silent, so observing “a few” in October could be interpreted as relatively abundant. In the summer of 1904, Grinnell (1905) noted “there must have been quite a number of them around though we actually saw but two” (he also observed many feathers, “dust-wallows,” and other “grouse signs”). In 1928, Pemberton (1928) reported hearing grouse hooting “quite commonly on the flatter upper part of the mountain.” He also noted, “many hooters” could be heard “well down on the cliff-like north slope.” He conjectured, “there are not many [on Mt. Pinos] and I believe that the number is less than one hundred.” On 21 May 1928, he photographed a nest on the north slope of Mt. Pinos and collected the eggs. Pemberton (1928) described the nest, eggs, and local environment. In the 13 years from 1921 to 1933, museum collectors took at least 40 specimens from Mt. Pinos (Table 1). Holt (1936) wrote that the grouse “was plentiful in bygone times on Mount Pinos and neighboring peaks. For the past six years, however, it has been feared that the bird was extinct here. As recently as last October, a U. S. Biological Survey party searched for two weeks without finding any indication that a single one remained. Then, on October 14, three hunters were caught with four Mount Pinos Sooty Grouse that they had shot on the northwest slope of Mount Pinos ... Were these the last four grouse in the Mount Pinos region? Probably not; it is a large, rugged area, and a few could easily escape the notice of searching parties. The hunters who took the birds state that a part of the flock was permitted to escape. If this observation is correct, and if grouse exist in numbers sufficient to enable them to withstand the depredations of natural enemies, they may repopulate the area in time. At best, however, their situation is critical.” The “two young hens and a pair of adults” mentioned by Holt were appar- ently prepared as study skins for the Santa Barbara Natural History Museum, but were lost along with all associated records in a fire on 12 April 1962 (P. Collins pers. comm.). No report of the survey Holt mentioned exists in the national archives (E. Alers pers. comm.). Grinnell and Miller (1944) reported that D. /. howardi was “said to have become very scarce of late years on Mount Pinos.” The AOU (1957) listed the subspecies as “extremely rare (possibly extinct) in the Tehachapi Mountains, Mount Pinos, and Frazier Mountain areas.” From 24 to 29 May 1964, Abbott (1965) reported hear- ing two grouse hooting on the northwest slope of Mt. Pinos, the only ones he had detected after searching Mt. Pinos, Cerro Noroeste, and Frazier Mountain “often” over the preceding 30 years. Abbott’s report is the last evidence of breeding behavior on the sky islands. A feather collected by D. 299 EXTIRPATION OF THE SOOTY GROUSE SOUTH OF THE SIERRA NEVADA Smith and J. Hamber on 23 September 1976 (Table 1) is the last physical evidence of the Sooty Grouse anywhere in the region. Since 1965, there have been only three credible reports of sightings documented in detail sufficient to eliminate misidentification of the Band- tailed Pigeon ( Patagioenas fasciata) or Mountain Quail {Oreortyx pictus) and placed on file at an appropriate institution: 23 September 1976 by J. Hamber and D. Smith at Cerro Noroeste (~5.5 km northwest of Mt. Pinos), 2 August 1992 by L. Allen at Sawmill Mountain (~1.8 km west of Mt. Pinos), and 1 May 1993 by R. Chichester also at Sawmill Mountain (Appendix). Other sightings reported since the 1970s (see Weiss 1979, Lentz 1993, and Bland 2008 for examples) lack documentation sufficient for a judgment of whether a Band-tailed Pigeon or Mountain Quail was misidentified, as frequently happens. Focused searches in the vicinity of Mt. Pinos in 1978 (Bendell and Zwickel 1984), 1979 (Weiss 1979), and 1988-1992 (Lentz 1993) failed to detect grouse. Habitat Associations Grinnell and Miller (1944) described the habitat of the Mount Pinos Sooty Grouse as “chiefly slopes clothed sparsely with white firs; but also other coniferous trees ... as well as open or brushy ground interspersed with, or adjacent to, forest.” First-hand records of the grouse’s habitat use in the sky islands are limited to the vicinity of Mt. Pinos (2697 m), where Grinnell (1905) found “grouse signs only among the firs on the north side” and Pem- berton (1928) found grouse “only on the higher portions of the mountain ... where the silver fir [white fir] reaches its best development.” Abbott (1965) “heard two grouse hooting in the heavy white fir on the northwest slope at about 8700 feet elevation” (2652 m). Pemberton encountered hooting males “on the flatter upper part of the mountain” and also “well down the cliff-like north slope.” Grinnell (1905) described the upper slopes of Mt. Pinos as “gentle,” “not heavily timbered,” with “groups of Jeffrey pines ... being interspersed with openings, either bare or brushy. In places above 8000 feet there are extensive low thickets of [snowberry, Symphoricarpos rotundifolius ] and, especially in moist spots, masses of gooseberry [Ribes cereum and R. montigenum], the latter more particularly on the shaded north side near springs.” “Acres of a low composite shrub [ Chrysothamnus nauseosus] cover the otherwise bare rolling area about the summit.” Con- versely, “the steep north slope, from the summit down about 2000 feet, is quite heavily timbered with California white fir,” “with a few fox-tail pine [limber pine, P. flexillis] at the summit of the main ridge ... and down the north slope among the firs.” Apparently, Pemberton was under the impres- sion hooting males moved down slope as the hooting season progressed, but other researchers have shown that locations of adult males’ breeding territories are static (Bendell 1955, McNicholl 1978, Bland 2013). Perhaps the males Pemberton heard early in the season were yearlings, which hoot for only a brief period early in the hooting season (McNicholl 1981, Bland 2013), often at low-quality sites (Jamieson and Zwickel 1983). Where I have observed the Mount Pinos Sooty Grouse’s breeding habitat north of Kern Gap (Sunday Peak, Poison Meadow, Sherman Peak, Big Meadows, Stony Creek, Big Baldy Ridge, Mineral King), it is similar in 300 EXTIRPATION OF THE SOOTY GROUSE SOUTH OF THE SIERRA NEVADA most respects to that of the Sierra Sooty Grouse on the western slope of the central Sierra Nevada, where I have studied Sooty Grouse habitat in detail. At Pinecrest (Tuolumne County), males establish breeding territories on or near steep slopes in mature fir or mixed conifer forest with an open canopy and where some trees exceeding 1 m in diameter (at breast height) are available for use as territorial songposts (Bland and Gardner 2013). In the southern Sierra Nevada, such habitats occur on north-tending slopes down to -1550 m elevation (U. S. Forest Service vegetation maps, fir and mixed-conifer forest types), in areas where large trees have not been elimi- nated by timber harvesting or catastrophic fire. Habitat associations south of Kern Gap were probably similar, although on the sky islands forests are drier, and the canopy is generally lower and more open. The Mount Pinos Sooty Grouse is probably much less migratory than northern populations of the Sooty Grouse, many of which migrate altitu- dinally between separate breeding and winter ranges (Bendell and Zwickel 1984, Zwickel and Bendell 2004). Within the range of D. /. howardi, suit- able habitat occurs in relatively narrow belts along the two main ranges of the southern Sierra Nevada (Great Western Divide and main Sierra Nevada) and around the peaks of the isolated sky islands, so substantial seasonal altitudinal migration is unlikely. Seasonal movements are probably similar to those of Sierra Sooty Grouse, which on the western slope winter and breed in one general area (Bland and Gardner 2013). The persistence of wintertime droppings (Zwickel and Bendell 2004:136) in males’ breeding territories at Sunday Peak and Mineral King (pers. obs.) suggests there is little seasonal migration in the southern Sierra Nevada. At Mt. Pinos, the only museum specimen collected in winter was collected at the summit, where other specimens were collected during the breeding season (Table 1). Grinnell (1905) mentioned reports from Mt. Pinos of grouse “in the pines down even as low as the sawmill ... in winter,” but the site of the sawmill (today’s McGill Campground, Kane 2008) is just 4.3 km east of the summit and well within potential breeding habitat (-2250 m, mixed conifer forest; U. S. Forest Service vegetation maps). DISCUSSION Given the results of my survey, and the lack of documented sightings since 1993, it can reasonably be concluded that the Mount Pinos Sooty Grouse has been extirpated from the sky islands. I believe I would have detected any breeding males present, since my methods have been effective for detecting territorial males in sparse populations throughout the Sierra Nevada, including those of D. /. howardi in northern Tulare County (Bland 1993). At Sunday Peak and Cherry Hill, males responded to my recorded cackle calls, and distances of initial detection ranged from -375 to 525 m. At 20:15 on 23 April 2002, 1 heard hooting on Sunday Peak from the shoulder of Rancheria Road (USFS 24S15), from a distance of -1.4 km. If a viable breeding population had been present in the sky islands, singing males should have been detected. Where population densities are low, male Sooty Grouse occur in clusters (Lewis 1985, Bland 2013) and counter- sing throughout the day (Stewart 1967, Zwickel and Bendell 2004:156), 301 EXTIRPATION OF THE SOOTY GROUSE SOUTH OF THE SIERRA NEVADA making detection relatively straightforward. It is possible that one or more silent grouse could have gone undetected, but the lack of hooting during peak hooting season implies that any such individuals did not constitute a viable breeding population. Although I was unable to survey some patches of potential habitat at the Breckenridge Mountain, Tehachapi Mountains, Alamo Mountain, and Pine Mountain survey sites because of limited access or time, habitat conditions or the lack of prior records suggest these areas had low potential for occupancy. In 2012, eight years after I surveyed the Greenhorn Mountains, N. J. Schmitt discovered a single male hooting ~1.5 km north of Shirley Peak in the southern Greenhorn Mountains, -7.5 km south of the birds I recorded at Sunday Peak. The bird was observed by several other people, including myself, and was photographed by J. Dunn and A. Sheehey, 29 April-27 May 2012. It was heard hooting again by several people in 2013, between late April and early May (N. J. Schmitt pers. comm.). On 13 May 2012, I searched this area and found no evidence (accumulations of droppings) of ad- ditional territorial males, and only a limited area of potential breeding habitat (only a few trees in the area, also known as Sawmill Ridge, exceeded 1 m in diameter at breast height). I believe this is a lone individual that dispersed to Shirley Peak, possibly from nearby Sunday Peak, and does not represent a group breeding at Shirley Peak. Nonetheless, the area should be monitored to assess whether grouse do successfully recolonize the Greenhorn Moun- tains south of Greenhorn Summit. The Mount Pinos Sooty Grouse could have been widespread -1500- 4000 yrs ago, when the region’s climate was consistently cool and moist, and fir, the primary food of the Sooty Grouse, was more widespread and abundant (Laudenslayer and Skinner 1995). Recurrent droughts beginning -1500 years ago (Graumlich 1993) might have contracted its range to the extent observed by early naturalists. As grouse habitat receded up the slopes of the sky islands, the population in smaller patches may have been reduced below the level of viability, being maintained only by emigration from larger patches. Perhaps no single island was large enough to sustain a population without genetic exchange with other islands. The three largest islands — Mt. Pinos and adjoining peaks, the Piute Mountains, and the Tehachapi Moun- tains — were known to be occupied by grouse at the beginning of the 20 th century, although available records suggest few existed in the Piute and Tehachapi mountains. Second-hand reports suggest the fifth-largest island, Frazier Mountain, might also have been occupied by a few grouse (Willett 1933). The lack of early 20 th -century records from other islands might be attributable to a lack of survey effort. Breckenridge Mountain supported an area of moist montane forest comparable to that of the largest islands, so it might also have supported grouse prior to heavy timber harvesting beginning in the late 1800s. Indigenous people or settlers could have extirpated grouse from the smaller islands by hunting before scientists became aware of them, or perhaps smaller islands were occupied only intermittently. Moist montane forest on Bear Mountain and the Liebre Twins might have been too little to support a persistent breeding population, but the locations of these islands suggest they could have served as stepping stones for dispersal between larger islands. Tecuya Mountain, Alamo Mountain, and Pine Mountain were 302 EXTIRPATION OF THE SOOTY GROUSE SOUTH OF THE SIERRA NEVADA probably all too small and too arid to support breeding grouse in modern times. Reports submitted orally to the Forest Service from Alamo and Pine mountains in the 1970s (Weiss 1979) were probably of misidentified Band- tailed Pigeons or Mountain Quail. Populations that persisted into the mid- 1900s were small and isolated and could have been extirpated by unseasonal snowfall, catastrophic fire, or disease. Increased human activity could also have expedited their disap- pearance, including the effects of livestock grazing (Mussehl 1963, Zwickel 1972, Zwickel and Bendell 2005), timber harvesting (Bland and Gardner 2013), rural development (Zwickel and Bendell 2005), illegal hunting, and fire suppression. Livestock grazing was well established in the sky islands by the mid-1800s (Morgan 1914, Stephenson and Calcarone 1999, Gos- sard 2005). In the early 1900s, flocks of sheep numbering in the thousands grazed many areas (Gossard 2005:72, Morgan 1914:34, Kane 2008:174, 211). In 1891, Fisher (1893) observed “the destruction of vegetation by sheep [in the southern Sierra Nevada] is a potent cause of the scarcity of ground-inhabiting birds.” In October, 1934, Grinnell (1935) noted, “there is . . . potentially much moisture-loving vegetation [~ 10 km north of Greenhorn Summit]; but heavy grazing in an extra dry season had produced conditions by autumn of this year, distressingly barren on and about the little meadows. ” Bergerud (1988) demonstrated the importance of low herbaceous growth for the insect foods and escape cover needed by grouse chicks. Timber harvesting began in the sky islands in the mid- 1800s to serve the needs of mines, ranches, settlements, and the army outpost at Fort Tejon (Morgan 1914, Laudenslayer and Darr 1990, Kane 2008). Between 1875 and 1925, sawmills were operating on at least four of the sky islands (Breckenridge Mountain, Piute Mountain, Tehachapi Mountains, and Mt. Pinos and adjoining peaks; Gossard 2005, Morgan 1914, Kane 2008). Large-scale commercial logging commenced in the 1950s and 1960s, but it was very limited in Los Padres National Forest (Mt. Pinos and adjoining peaks, Tecuya, Frazier, Alamo, and Pine mountains; Baker and Stewart 1996, Stephenson and Calcarone 1999). Bergerud (1988) described how intensive timber harvesting and other alterations of habitat structure at a similar scale, including wildfire, can disrupt the “predator-cover complex” of grouse, heightening their risk of predation as they navigate through an altered mosaic of habitats. Timber harvesting is negatively associated with habitat occupancy by breeding male Sierra Sooty Grouse, in part because the trees they select for territorial songposts, averaging ~1 m in diameter at breast height, are valuable timber (Bland 2006). Perhaps a century or more of logging in the sky islands eliminated elements of habitat structure that were essential for grouse or fragmented the habitat to the extent that the grouse were exposed to unsustainable rates of predation. Fire suppression since the early 1900s has also altered grouse habitats. Where fires are suppressed, undergrowth proliferates and large trees suffer increased mortality due to competition (Parsons and DeBenedetti 1979). Secondarily, where fires have been suppressed, successive wildfires are more frequent and more intense, often killing most or all trees. Historically, low- intensity wildfires at intervals of 15 to 30 years thinned forest understories and created patchy overstories (Taylor and Halpern 1991). This maintained 303 EXTIRPATION OF THE SOOTY GROUSE SOUTH OF THE SIERRA NEVADA the fine-scale heterogeneity of the canopy of old-growth mixed-conifer for- ests of the Sierra Nevada (Fites-Kaufmann et al. 2007), to which southern populations of the Sooty Grouse have apparently adapted. With fire sup- pression, increasingly large areas of grouse habitat became either overgrown (e.g., Mt. Pinos) or ruined by catastrophic fires (e.g., Piute Mountains). There are now substantial obstacles to natural recolonization from the southern Sierra Nevada. Habitat suitability in the sky islands continues to decline as a result of timber harvesting, wildlfire, and residential development. Drought and associated insect outbreaks have increased the mortality of co- nifers (Van Mantgem and Stephenson 2007, Bentz et al. 2010), creating an impetus to “sanitation log” afflicted trees (Graham et al. 1999). Under some scenarios of climate change, moisture- and temperature-sensitive elements of montane forest could be further eroded by a warming climate (Lenihan et al. 2003). The 2008 Piute fire caused substantial long-term damage to the first stepping stone to potential recolonization. Many decades, even centuries, will be necessary for the return of features of old forest such as trees >1 m diameter and large downed logs. The next major stepping stone, the Tehachapi Range, is rapidly being transformed by low-density residential development, even at higher elevations. Reintroduction by translocation from the Sierra Nevada might also be problematic, even if habitat conditions improve. Miller and Benson (1930) considered the Mount Pinos Sooty Grouse phenotypically uniform across the Kern Gap. However, those from Mt. Pinos may differ to some extent in mitochondrial DNA haplotype from those from the southern Sierra Nevada (G. Barrowclough pers. comm.). ACKNOWLEDGMENTS The California Department of Fish and Wildlife supported my field work with a series of Game Bird Heritage Grants, and Santa Monica College approved a teach- ing schedule that allowed for spring field work. Keith Axelson, Bob Barnes, Teresa Benson, Mike Foster, Rick Howard, John Kelly, Larry Layne, Clark Moore, Meenakshi Nagendran, Joe Sovinsky, and Michael White helped with access to study areas or other logistics. Ellen Alers, Carla Cicero, Paul Collins, James Dean, and Craig Ludwig helped search for old museum records, and Janet Hamber provided additional archived material. George Barrowclough, Jim Bendell, Clait Braun, Jon Dunn, Kimball Gar- rett, Greg Gerstenberg, Kathy Molina, Adam Rich, and Philip Unitt provided helpful comments on earlier drafts. LITERATURE CITED Abbott, W. G. 1965. Blue Grouse persists on Mount Pinos in southern California. Condor 67:85-86. American Ornithologists’ Union. 1957. Check-list of North American Birds, 5th ed. Lord Baltimore Press, Baltimore, MD. Baker, M., and Stewart, W. 1996. Ecosystems under four different public institutions: A comparative analysis, in Sierra Nevada Ecosystem Project: Final Report to Congress (D. C. Erman, ed.), vol. II, pp. 1347-1367. Univ. of Calif., Davis. Bendell, J. F. 1955. Age, breeding behavior and migration of Sooty Grouse. Trans. N. Am. Wildlife Conf. 20:367-381. Bendell, J. F., and Zwickel, F. C. 1984. A survey of the biology, ecology, abundance, and distribution of the Blue Grouse (genus Dertdragapus), in Third International 304 EXTIRPATION OF THE SOOTY GROUSE SOUTH OF THE SIERRA NEVADA Grouse Symposium (P. J. Hudson and T. W. I. Lovel, eds.), pp. 163-192. World Pheasant Assoc., Reading, England. Bentz, B. J., Regniere, J., Fettig, C. J., Hansen, E. M., Hayes, J. L., Hicke, J. A., Kelsey, R. G., Negron, J. F., and Seybold, S. J. 2010. Climate change and bark beetles of the western United States and Canada: Direct and indirect effects. BioScience 60:602-613. Bergerud, A. T. 1988. Increasing the numbers of grouse, in Adaptive Strategies and Population Ecology of Northern Grouse (A. T. Bergerud, and M. W. Gratson, eds.), pp. 686-731. Univ. of Minn. Press, Minneapolis. Bland, J. D. 1993. Forest grouse and Mountain Quail investigations: A final report for work completed during the summer of 1992. Report to Wildlife Mgmt. Div., Calif. Dept. Fish and Wildlife, 1812 Ninth St., Sacramento, CA 95811. Bland, J. D. 2006. Features of the forest canopy at Sierra Sooty Grouse courtship sites. Report to Wildlife Mgmt. Div., Calif. Dept. Fish and Wildlife, 1812 Ninth St., Sacramento, CA 95811. Bland, J. D. 2008. Mount Pinos Sooty Grouse, in California Bird Species of Special Concern: A Ranked Assessment of Species, Subspecies, and Distinct Populations of Birds of Immediate Conservation Concern in California (W. D. Shuford and T. Gardali, eds.), pp. 102-106. Studies of Western Birds 1. W. Field Ornithol., Camarillo, CA. Bland, J. D. 2013. Estimating the number of territorial males in low-density popula- tions of the Sooty Grouse. W. Birds 44:262-276. Bland, J. D., and Gardner, S. 2013. Habitat assessment and monitoring protocol for Sooty Grouse (Dendragapus fuliginosus) in the Sierra Nevada. State Wildlife Grant T-16, Segment 1 (F07AF00074). Report to Wildlife Branch, Calif. Dept. Fish and Wildlife, 1812 Ninth St., Sacramento, CA 95811. Dickey, D. R., and van Rossem, A. J. 1923. Description of a new grouse from southern California. Condor 25:168-169. Doerr, J. G., Brighenti, J. M., Barescu, C. L., and Morin, M. 1984. Use of clearcut and old-growth forests by male Blue Grouse in central southeast Alaska, in Fish and Wildlife Relationships in Old-Growth Forests: Proceedings of a Symposium Held in Juneau, Alaska, 12-15 April, 1982 (W. R. Meehan, T. R. Merrell, and T. A. Hanley, eds.), pp. 309-313. Am. Inst. Fishery Res. Biol., Juneau, AK. Fisher, A. K. 1893. Report on the ornithology of the Death Valley Expedition of 1891, comprising notes on the birds observed in southern California, southern Nevada, and parts of Arizona and Utah. N. Am. Fauna 7:7-158. Fites-Kaufmann, J. A., Rundel, P., Stephenson, N., and Weixelman, D. A. 2007. Montane and subalpine vegetation of the Sierra Nevada and Cascade Ranges, in Terrestrial Vegetation of California (M. G. Barbour, T. Keeler-Wolf, and A. A. Schoenherr, eds.), pp. 456-501. Univ. of Calif. Press, Berkeley. Garrett, K., and Dunn, J. 1981. Birds of Southern California: Status and Distribution. Los Angeles Audubon Soc., Los Angeles. Gossard, G. 2005. Three Valleys of the Tehachapi: Bear, Bright, and Cummings. Arcadia Publ., Mount Pleasant, SC. Graham, R. T., Harvey, A. E., Jain, T. B., and Tonn, J. R. 1999. The effects of thinning and similar stand treatments on fire behavior in western forests. Gen. Tech. Rep. PNW-GTR-463. U.S. Forest Service, Portland, OR. Graumlich, L. 1993. A 1000-year record of temperature and precipitation in the Sierra Nevada. Quaternary Res. 39:249-255. Grinnell, J. 1905. Summer birds of Mount Pinos, California. Auk 22:378-383. Grinnell, J. 1935. Winter Wren and Pileated Woodpecker on the Greenhorn Moun- tains, California. Condor 37:44-45. Grinnell J., and Miller, A. H. 1944. The distribution of the Birds of California. Pac. Coast Avifauna 27. 305 EXTIRPATION OF THE SOOTY GROUSE SOUTH OF THE SIERRA NEVADA Holt, W. B. 1936. Grouse and condors. Santa Barbara Mus. Nat. Hist. Leaflet 11:80-84. Jamieson, I. G., and Zwickel, F. C. 1983. Spatial patterns of yearling male Blue Grouse and their relation to recruitment into the breeding population. Auk 100:653-657. Kane, B. K. 2008. A View from the Ridge Route, vol. 4: The Resources. Jostens Printing and Publishing, Visalia, CA. Laudenslayer, W. F., and Darr, H. H. 1990. Historical effects of logging on the forests of the Cascade and Sierra Nevada ranges of California. Trans. W. Sec. Wildlife Soc. 26:12-23. Laudenslayer, W. F., and Skinner, C. N. 1995. Past climates, forests, and disturbances of the Sierra Nevada, California: Understanding the past to manage for the future. Trans. W. Sec. Wildlife Soc. 31:19-26. Lenihan, J. M., Drapek, R., Bachelet, D., and Neilson, R. P. 2003. Climate change effects on vegetation distribution, carbon, and fire in California. Ecol. Appl. 13:1667-1681. Lentz, J. E. 1993. Breeding birds of four isolated mountains in southern California. W. Birds 24: 201-234. Lewis, R. R. 1985. Do Blue Grouse form leks? Auk 102:180-184. Mayer, K. E., and Laudenslayer, W. F. 1988. A Guide to Wildlife Habitats in California. Calif. Dept, of Forestry and Fire Protection, Sacramento. McNicholl, M. K. 1978. Behaviour and social organization in a population of Blue Grouse on Vancouver Island. Ph.D. dissertation, Univ. Alberta, Edmonton. McNicholl, M. K. 1981. Caution needed in use of playbacks to census bird popula- tions. Am. Birds 35:235-236. Miller, A. H., and Benson, S. B. 1930. The summer resident birds of the boreal and transition life-zones of Mount Pinos, California. Condor 32:101-104. Morgan, W. M. 1914. History of Kern County, California. Historic Record Co., Los Angeles. Mussehl, T. W. 1963. Blue Grouse brood cover selection and land-use implications. J. Wildlife Mgmt. 27:547-555. Niederleitner, J. F. 1987. Use of early successional, midsuccessional, and old-growth forests by breeding Blue Grouse (Dendragapus obscurus fuliginosus ) on Hard- wicke Island, British Columbia. Can. J. Zool. 65:151-155. Parsons, D. J., and DeBenedetti, S. H. 1979. Impact of fire suppression on a mixed- conifer forest. Forest Ecol. Mgmt. 2:21-33. Pemberton, J. R. 1928. The nesting of Howard’s Blue Grouse. Condor 30:347-348. Richardson, C. H. 1904. A list of summer birds of the Piute Mountains, California. Condor 6:134-137. Stephenson, J. R., and Calcarone, G. M. 1999. Southern California mountains and foothills assessment: Habitat and species conservation issues. Gen. Tech. Rep. PSW-172, U.S. Forest Service, Albany, CA. Stewart, P. A. 1967. Hooting of Sitka Blue Grouse in relation to weather, season, and time of day. J. Wildlife Mgmt. 31:28-34. Stirling. I., and Bendell, J. F. 1966. Census of Blue Grouse with recorded calls of a female. J. Wildlife Mgmt. 30:184-187. Taylor, A. H., and Halpern, C. B. 1991. The structure and dynamics of Abies mag- nified forests in the southern Cascade Range, USA. J. Veg. Sci. 2:180-200. U.S. Forest Service. 2005. Final environmental impact statement, vol. 2 (appendices), land management plans, Angeles National Forest, Cleveland National Forest, Los Padres National Forest, San Bernardino National Forest. U.S. Forest Service, Vallejo, CA. Van Mantgem, P. J., and Stephenson, N. L. 2007. Apparent climatically induced increase of tree mortality rates in a temperate forest. Ecol. Lett. 10:909-916. 306 EXTIRPATION OF THE SOOTY GROUSE SOUTH OF THE SIERRA NEVADA Weiss, S. 1979. Blue Grouse study report. Report to Los Padres National Forest, 34580 Lockwood Valley Rd., Frazier Park, CA 93225. Willett, G. 1933. A revised list of the birds of southwestern California. Pac. Coast Avifauna 21. Zwickel, F. C. 1972. Some effects of grazing on Blue Grouse during summer. J. Wildlife Mgmt. 36:631-634. Zwickel, F. C., and Bendell, J. F. 2004. Blue Grouse: Their Biology and Natural His- tory. NRC Research Press, Ottawa, Canada. Zwickel, F. C., and Bendell, J. F. 2005. Blue Grouse (Dendragapus obscurus), in The Birds of North America Online (A. Poole, ed.), no. 15. Cornell Lab of Ornithol., Ithaca, NY; http://bna.birds.cornell.edu/bna/species/015. Accepted 11 September 2013 APPENDIX Documented observations of the Mount Pinos Sooty Grouse in the sky islands region since 1964. Written reports on file at an appropriate institution and includ- ing detail sufficient to exclude a misidentified Band-tailed Pigeon or Mountain Quail. Photocopies of these reports are available by request. 23 September 1976, J. Hamber, D. Smith, and M. Hasey, upper north slope of Mt. Abel (Cerro Noroeste), below the ski tow and lodge, with Camp Condor in view below, -2135 m elevation (34° 51' N 199° 12' W). D. Smith: “We heard the sound of a good-sized bird moving in a tree over our heads. I saw the bird fly to a nearby Jeffrey pine just down the side of the slope. The others heard it — sounding like the rustling of a shopping sack — as it landed, giving all of us a good profile of its full body. Bird was approx, size of a Rhode Island red chicken- — large body, small head, good size, square end tail with a light band at end — body appeared to be speckled and dark. Gray with no warm color apparent to us. Saw for about 3 or 4 seconds — then bird flew, flapping 3 or 4 times then coasting, dropping rapidly down canyon, disappearing into trees below. Wings were curved steeply down at ends as it coasted. I heard no call as it moved out. Picked up a few body feathers from the slope [specimen AV10506, Table 1] — may or may not be associated. Appearance of the bird was definitely that of a galliform.” J. Hamber: “The bird was first seen by Dick Smith as it flew from one tree to the Jeffrey pine.” “I moved upslope slightly and found the bird perched at an oblique angle on a bare area of a large branch of the tree about 40 feet from the ground and 6 feet from the trunk of a mature Jeffrey pine. The canopy of the tree shaded the bird, but no part was obscured by branches. In the binoculars [from -15 m], I observed a large, 15- to 18-inch bird with a small head; small, stout bill; short neck, plump body and fairly long, wide tail — squared off on the end. Definitely a gallinaceous bird. The bird was dark, brownish gray all over with some lighter mottling on the back. The tail was dark with a light l/2''-3/4" band on the end of the tail. The bird was in view for about 10 seconds. It then left the branch, sailed down the ravine on decurved wings and was lost to sight a few seconds later. ” 2 August 1992, L. Allen, -1 km SW of the peak of Sawmill Mountain, -2600 m elevation (34° 49' N, 119° 9' W). “The bird flushed [at a distance of 7-8 m with very loud wingbeats ... circled behind me and disappeared from view.... The head, dorsal surface of the body, and wings appeared mottled in no particular pattern but gave the overall impression of a medium-brown plumage. The tail was fanned as the bird banked; the dorsal color extended about halfway down the tail, beyond which the tail was a uniform dark brown, with a narrow, white (or pale) terminal band. 1 May 1993, R. Chichester, Upper northeast slope of Sawmill Mountain, -2600 m elevation (34° 48.9' N, 119° 9.6' W). “The birds were 10-15' away when I jumped 307 EXTIRPATION OF THE SOOTY GROUSE SOUTH OF THE SIERRA NEVADA them. My observations were side view and mainly rear view as they flew over the crown of the hill. The color pattern of the tail and body of the birds was definitely that of grouse. I noted the banded color pattern of the tail as it was fanned out in flight.” Afterward, a second person, M. Chichester, heard “several diagnostic booms” ... “fol- lowed by a few low clucks.” R. Chichester conveyed orally to M. Chichester (in litt.) that the flushed birds produced an “explosive, frightening sound,” were “double the size of quail,” “brown/gray” overall, with a “fanned gray tail” that had a “band on the tip.” Male Mount Pinos Sooty Grouse (Dendragapus fuliginosus howardi), Giant Sequoia National Monument, California, 31 May 1992. Photo by James D. Bland 308 NOTES THE DISTRIBUTION OF BUBO VIRGINIANUS PINORUM NORTH AND WEST TO WASHINGTON ROBERT W. DICKERMAN, Museum of Southwestern Biology, University of New Mexico, Albuquerque, New Mexico 87131; bobdickm@unm.edu Dickerman and Johnson (2008) described subspecies Bubo uirginianus pinorum of the Great Horned Owl, designating a type specimen from the Sandia Mountains of north-central New Mexico and outlining the range as “plains of the Snake River of Idaho, south ... at increasing elevations, to Arizona and New Mexico.” Bubo u. pinorum is darker dorsally and more heavily barred ventrally than the pale B. v. palles- cens of the desert Southwest or B. v. subarcticus of the Canadian prairie parklands but paler and more lightly barred than in B. v. lagophonus or B. u. saturatus to the northwest or in B. u. pacificus of coastal California. Here I trace the distribution of B. v. pinorum further, extending it north and west to central and northern Washington. Oberholser (1904) resurrected the name B. u. occidentals Stone (a synonym of B. u. subarcticus Hoy; Dickerman 1991) for the Great Horned Owls of south-central Canada and the U.S. from Montana, northeastern California, and Nevada east to eastern Minnesota and south to central Kansas (AOU 1957) because they just did not fit his other subspecies. His action was in large part a result of not recognizing migration in this species in the West (Oberholser 1904, Dickerman 2002, Dickerman and Johnson 2008). The name soon became entrenched in the literature (e.g., AOU 1910, 1931, 1957, Linsdale 1936, Grinnell and Miller 1944), and museums used the name occidentalis in arranging specimens. Thus in my search for specimens of pinorum I first looked for specimens identified as occidentalis, In 1991, when I began my studies of New Mexico’s Great Horned Owls, I visited the Utah Museum of Natural History, before my concepts were established. 1 scored specimens there as “pj” (pinon/juniper = pinorum) or as “pall.” (B. u. pallescens). In later visits to other collections, I distinguished the five subspecies that occur in the Rocky Mountain region, B. u. pinorum, pallescens, lagophonus, subarcticus, and saturatus. I identified over 46 specimens of B. u. pinorum from Washington, Oregon, Cali- fornia, Nevada, and Utah; those from Washington, Oregon, California, and Nevada are plotted in Figure 1. Some dots in Washington and Oregon represent birds that are not on their nesting grounds, so Figure 1 does not reflect exact nesting range, but it does reflect the overall range of the subspecies. Bubo u. pinorum is largely seden- tary; specimens from nonbreeding sites presumably represent only local movement. The type specimen of B. v. lagophonus Oberholser is one of four Great Horned Owls collected by Charles Bendire at Walla Walla, Washington, in the winter of 1881-1882. These four specimens, at the U.S. National Museum of Natural His- tory, include two of B. v. lagophonus and two of B. v. pinorum; lagophonus is a migrant at its type locality. I have seen 11 Great Horned Owls taken by Bendire at Walla Walla, and six of them represent pinorum. There are, scattered through several museums, at least 14 specimens of pinorum, including almost fledged juveniles, from Walla Walla County, Washington, and adjacent Umatilla County, Oregon. The dots from Nevada and California all represent birds collected in the nesting season, so Figure 1 represents the nesting range of pinorum in those states. Bubo u. pinorum probably occurs in southeastern British Columbia as well, although to date I have seen no specimens from there. A similar study is needed to outline the northeastern portion of the range of pinorum in Idaho, Montana, and Wyoming, and possibly western North and South Western Birds 44:309-311, 2013 309 NOTES Figure 1. Distribution of specimens of Bubo virginianus pinorum in Washington, Oregon, California, and Nevada. Dakota, but there are far fewer specimens from those areas, and they are far more scattered. Salvage of specimens by state game departments is to be encouraged. I am grateful to the curators and collection managers at the museums where I examined specimens: the American Museum of Natural History, New York; Burke Museum of Natural History and Culture, University of Washington, Seattle; Museum of Comparative Zoology, Harvard University, Cambridge, Massachusetts; Museum of Southwestern Biology, University of New Mexico, Albuquerque; Museum of 310 NOTES Vertebrate Zoology, University of California, Berkeley; San Diego Natural History Museum; Utah Museum of Natural History, University of Utah, Salt Lake City; and the United States National Museum of Natural History, Washington, D.C. Thanks also to reviewers M. Ralph Browning and Daniel D. Gibson. LITERATURE CITED American Ornithologists’ Union. 1910. Check-list of North American Birds, 3 rd ed. Am. Ornithol. Union, New York. American Ornithologists’ Union. 1931. Check-list of North American Birds, 4 th ed. Am. Ornithol. Union, Lancaster, PA. American Ornithologists’ Union. 1957. Check-list of North American Birds, 5 th ed. Am. Ornithol. Union, Baltimore. Dickerman, R. W. 1991. On the validity of Bubo virginianus occidentalis Stone. Auk 108:964-965. Dickerman, R. W. 2002. The taxonomy of the Subarctic Great Horned Owl, Bubo virginianus subarcticus, nesting in the United States. Am. Midland Nat. 148:198-199. Dickerman, R. W., and Johnson, A. B. 2008. Notes on Great Horned Owls nesting in the Rocky Mountains, with description of a new subspecies. J. Raptor Res. 42:20-28. Grinnell, J., and Miller, A. H. 1944. The distribution of the birds of California. Pac. Coast Avifauna 37. Linsdale, J. M. 1936. The birds of Nevada. Pac. Coast Avifauna 23. Oberholser, H. C. 1904. A revision of the American Great Horned Owls. Proc. U. S. Natl. Mus. 27:177-192. Accepted 11 July 2013 311 NOTES RECORDS OF THE BLACK MERLIN IN NEW MEXICO, WITH COMMENTS ON ITS IDENTIFICATION ROBERT W. DICKERMAN, Museum of Southwestern Biology, Department of Biology, University of New Mexico, Albuquerque, New Mexico 87131; bobdickm@unm.edu The nesting range of the Black Merlin ( Falco columbarius suckleyi), adapted to the temperate rain forest, has been reported as southeast Alaska (Gibson and Kessel 1997), coastal British Columbia (AOU 1957), and adjacent Washington (Wahl et al. 2005). Not all of this distribution has been documented with specimens collected in the nesting season, although some nesting individuals have been trapped and examined in Washington state (C. M. Anderson, fide C. M. White, in litt. , 2013). Migrants and wintering birds have been recorded in Oregon (Marshall et al. 2003) and California (AOU 1957). Elsewhere, specimens have been reported from Arizona (Monson and Phillips 1981), New Mexico (Jewett 1944, Friedmann 1950), Colorado (Bailey 1942), Nevada (Alcorn 1943), Idaho (Burleigh 1972, Haak and Sawby 2012), Utah (Behle 1985, Haney and White 1999), Wisconsin (Friedmann 1950), and New Jersey (Capainolo and Pitocchelli 1990). From New Mexico, eight specimens have been previously identified as F. c. suck- leyi, by Jewett (1944), Friedmann (1950), John P. Hubbard, or me (Table 1). Four of the eight New Mexico specimens are housed in the Museum of Southwestern Biology (MSB). To reassess the subspecific identifications of the New Mexico specimens, I reviewed the taxonomic treatments of the darker-plumaged Merlins of North America by Ridgway (in Baird and Ridgway 1874), Swarth (1935), Friedmann (1950), Temple (1972a, b), Palmer (1988), Hamilton and Schmitt (2000), Wheeler (2003), and Warkentin et al. (2005). Ridgway’s very brief description of Falco columbarius variety Suckleyi [in Baird and Ridgway 1874) was poor, but he did designate two syntypes, a male (U.S. National Museum [USNM] 4477) from Shoalwater Bay, Washington, and a female (USNM 5882) from Fort Steilacoom, Washington. Friedmann (1950), presumably using these type specimens among others, men- tioned that F. c. suckleyi has only three light-colored bars in the outer rectrices, in contrast to the three or four in F. c. columbarius, without specifying the frequency of three versus four bars in columbarius. In his count of the number of bars Friedmann apparently disregarded the most basal, which is largely hidden by the undertail coverts. In a major revision of the species, Temple (1972a, b) analyzed the plumage of the Merlin on the basis of five regional samples, eastern, central, and western taiga [F. c. columbarius, including F. c. bendirei ), “prairie parkland’" [F. c. richardsoni ), and “coastal forest” [F. c. suckleyi). No single character distinguished the coastal forest sample from all the others, but a combination of characters did. In over 90% of 104 coastal forest specimens the light bars did not extend to the inner edge of the outermost primary, thus being spots rather than bars, whereas in >90% of the specimens of the other four regional samples they did so. About 90% of the coastal forest specimens had the crown solid dark or with dark streaking far more extensive than the lighter background coloration, a pattern characterizing <20% of the specimens in the other four samples, in which light coloring was more prominent. Less diagnostic was the number of light bars in the outer rectrix; 100% of specimens of F. c. suckleyi had four or fewer bars, while only 9% of F. c. columbarius did so. Temple (1972a) took colorimetric readings of the five small regional samples and of all sex/age classes. Regardless of sex/age class, F. c. suckleyi was far darker dorsally (ventral data not presented) in “brightness” (15.7 ± 0.3 versus 17.0-17.9 ± 0.4-0. 6, 312 Western Birds 44:312-315, 2013 NOTES means ± standard error) than specimens of the other four samples. There was a small degree of overlap in dominant wavelength (23.4 ± 0.3 versus 24.1-25.8 ± 0.4-0. 7), especially with the “eastern taiga” sample of nominate columbarius. Four of the eight specimens from New Mexico are in the Museum of Southwestern Biology; I borrowed the others from the museums holding them, the U.S. National Museum of Natural History (USNM), Slater Museum, University of Puget Sound (PSM), and the Department of Biology, New Mexico State University (NMSUB), except for one specimen in the Museum of Zoology, University of Michigan (UMMZ 175043). To gain a better appreciation of the plumage characters of F. c. suckleyi, I examined seven additional specimens collected on Vancouver Island, British Co- lumbia, and preserved in the Museum of Vertebrate Zoology (MVZ), University of California, Berkeley: MVZ 15627, <$, Errington, 2 September 1910; MVZ 15629, §, Errington, 10 September 1912; MVZ 15630, S, Errington, 21 September 1910; MVZ 15631, Errington, 22 September 1910; MVZ 89861, imm. Comox, 9 March 1941; MVZ 99935, Quatsino, 31 August 1934; MVZ 99942, <$, Seal Islands (Comox), 3 August 1927. I consider 15629, 15630, 15631, 99935, and 99941 to be “ultra-typical standards” of F. c. suckleyi. I pooled females and immature males in all comparisons of plumage characteristics. I evaluated all specimens on the basis of two characters defining F. c. suckleyi, both of which are exemplified by the five “ultra- typical” specimens from Vancouver Island: (1) four light-colored bars in the outer rectrix, and (2) light markings on the inner web of the outermost primary appear as spots, rather than as bars as in F. c. columbarius. These two criteria accompanied the dark color and pattern of the crown and the darkness of the dorsum and venter. My reassessments of the eight New Mexico specimens are as follows. 1. USNM 194613, immature S, Lake La Jara [Rio Arriba Co.], 18 September 1904, collected by J. H. Gaut. This specimen had been identified by W. W. Cooke (in Bailey 1928) and J. W. Aldrich (label annotation, 1947) as F. c. columbarius, then as F. c. suckleyi by Friedmann (1950) and J. P. Hubbard (in notes, 1991), and finally as columbarius x suckleyi by Hubbard in 2012. On the basis of the barring on the inner web of the outer primary and the only moderately dark upperparts and ventral streaking it is F. c. columbarius. 2. PSM 10055, 5, [Ghost Ranch, near] Abiquiu [Rio Arriba Co.], 26 December 1946, shot by an unknown employee of Ghost Ranch. This specimen became 2234 in the collection of S. G. Jewett (1944), who identified it as F. c. suckleyi. J. P. Hubbard identified it as F. c. columbarius intermediate toward suckleyi in 2012. I consider it F. c. columbarius x suckleyi because of its only moderately heavy ventral streaking, mainly barred inner web of the outer primary, and having four tail bars. 3. UMMZ 175043, $, Reserve [Catron Co.], 12 October 1927, collected by H. H. Kimball. Originally identified on its label as F. c. columbarius. In 1962 and 2012 J. P. Hubbard identified it as F. c. columbarius x F. c. suckleyi on the basis of its dark coloration, but its outer primaries are barred. I have not examined it. 4. MSB 4613, $, Albuquerque, Bernalillo Co., 16 September 1979, collected by J. Johnson, prepared by D. L. Pennington. I identified this specimen as F. c. suckleyi in 1995, as did J. P. Hubbard in 1998 and 2012. On the basis of the spotted inner webs of the outer primaries, darkness of the venter, and four tail bars it does appear to be F. c. suckleyi. 5. MSB 6738, §, Albuquerque, Bernalillo Co., 4 January 1990, received from Wildlife Rescue, Inc., of Albuquerque. In plumage color, this specimen looks like F. c. columbarius dorsally and F. c. suckleyi ventrally, as I noted in 1995 and did J. P. Hubbard in 1998. On the basis of its more barlike maks on the inner web of the outer primaries, less deep ventral coloration, and five tail bars I now regard it as F. c. columbarius x F. c. suckleyi. 6. MSB 20581, on label (but immature S by size), Manzano Mountains, Tor- 313 NOTES ranee Co., 20 January 1998, received from Shirley Kendall. Though J. P. Hubbard identified this specimen as F. c. suckleyi in 1998, I consider it F. c. columbarius x F. c. suckleyi on the basis of the barred inner webs of the outer primaries and darkness of the underparts. 7. NMSUB 3858, Albuquerque area, Bernalillo Co., 20 March 1999, received from M. Hamburg. As did J. P. Hubbard in 2012, I identify this specimen as F. c. columbarius intermediate toward F. c. suckleyi on the basis of the inner web of the outer primary being both spotted and barred and the outer rectrix having five pale bars. 8. MSB 23665, immature $, Villanueva, San Miguel Co., 3 October 2000, found by G. Sena, sent to K. Ramsey, from whom the MSB received it. Again, this speci- men may be regarded as F. c. columbarius x F. c. suckleyi, not qualifying as clearly suckleyi because of its insufficiently dark venter, mixture of spots and bars on the inner webs of the outer primaries, and five tail bars. My results support the suggestion of Wheeler (2003) that away from the subspecies’ usual range, any prospective F. c. suckleyi needs to be compared rigorously to the standards. It demonstrates that F. c. suckleyi cannot be identified definitively in the field or casually in the museum (as evidenced by the misidentifications I correct here, including some of my own!). For example, three of the four specimens from southern California illustrated by Hamilton and Schmitt (2000) are identified in error as F. c. suckleyi: they have too many tail bars, and the markings visible on primaries are bars rather than spots. Likewise, my results strongly suggest that the numerous records of suckleyi from the interior of North America be re-examined on the basis of these more stringent criteria. Many of the records based on observations or photographs will prove to be inconclusive. Nevertheless, F. c. suckleyi does occur casually at least as far south and east as New Mexico, as attested by MSB 4613, collected at Albuquerque on 16 September 1979. Special thanks are due to John P. Hubbard, who spearheaded much of this project and who kindly loaned me his notes and gave advice. I thank Carla Cicero of the Museum of Vertebrate Zoology, Peter Houde of the Department of Biology, New Mexico State University, Las Cruces, Gary Shugart of the Slater Museum of Natural History, University of Puget Sound, Tacoma, and James Dean and Craig Ludwig of the U. S. National Museum of Natural History, Washington, DC, for the loan of speci- mens. Thanks also to reviewers Clayton M. White, M. Ralph Browning, and Daniel D. Gibson. The Museum of Southwestern Biology acknowledges the rehabilitators of New Mexico who provided four of the specimens used in this study. LITERATURE CITED Alcorn, J. R. 1943. Additions to the birds of Nevada. Condor 45:40. American Ornithologists’ Union. 1957. Check-list of North American Birds, 5 th ed. Am. Ornithol. Union, Baltimore. Bailey, A. M. 1942. The Black Pigeon Hawk in Colorado. Condor 44:37. Bailey, F. M. (with W. W. Cooke). 1928. Birds of New Mexico. New Mexico Dept, of Game and Fish, Santa Fe. Baird, S. F., and Ridgway, R. 1874. On some new forms of North American birds. Bull. Essex Inst. 5:197-204. Behle, W. H. 1985. Utah birds: Geographic distribution and systematics. Utah Mus. Nat. Hist. Occas. Paper 5:1-147. Burleigh, T. D. 1972. Birds of Idaho. Caxton, Caldwell, ID. Capainolo, P., and Pitocchelli, J. 1990. Eastern records of Falco columbarius suck- leyi. Kingbird 40:221-225. Friedmann, H. 1950. The birds of North and Middle America. U. S. Nat. Mus. Bull. 50, part 11. 314 NOTES Gibson, D. D., and Kessel, B. 1997. Inventory of the species and subspecies of Alaska birds. W. Birds 28:45-95. Haak, B. A., and Sawby, S. 2012. Recent nesting and identity of the Merlin in Idaho. W. Birds 43:105-108. Hamilton, R. A., and Schmitt, N. J. 2000. Identification of Taiga and Black Merlins. W. Birds 31:65-67. Haney, D. L., and White, C. M. 1999. Habitat use and subspecific status of Merlins, Falco columbarius , wintering in central Utah. Great Basin Nat. 59:266-276. Jewett, S. G. 1944. The Black Pigeon Hawk in New Mexico. Condor 46:206. Marshall, D. B., Hunter, M. G., and Contreras, A. L. 2003. Birds of Oregon. Ore. State Univ. Press, Corvallis. Monson, G., and Phillips, A. R. 1981. Annotated Check-list of the Birds of Arizona. Univ. Ariz. Press, Tucson. Palmer, R. S. 1988. Handbook of North American Birds, vol. 5. Yale Univ. Press, New Haven, CT. Swarth, H. S. 1935. Systematic status of some northwestern birds. Condor 37:199-204. Temple, S. A. 1972a. Systematics and evolution of the North American Merlins. Auk 89:325-338. Temple, S. A. 1972b. Sex and age characteristics of North American Merlins. Bird- Banding 43:191-196. Wahl, T. R., Tweit, B., and Mlodinow, S. G. 2005. Birds of Washington. Ore. State Univ. Press, Corvallis. Warkentin, I. G., Sodhi, N. S., Espie, R. H. M., Poole, A. R., Oliphant, L. W., and James, P. C. 2005. Merlin ( Falco columbarius), in The Birds of North America Online (A. Poole, ed.), no. 44. Cornell Lab Ornithol., Ithaca, NY; http://bna. birds .Cornell . edu/bna/ species/ 044 . Wheeler, B. K. 2003. Raptors of Eastern North America. Princeton Univ. Press, Princeton, NJ. Accepted 11 July 2013 Wing your way to... SAN DIEGO, CALIFORNIA 9-12 OCTOBER 2014 The 39th annual conference of the Western Field Ornithologists will be held in San Diego, California, 9-12 October 2014, at the Marriott Courtyard hotel at Liberty Station, less than 1 mile from the San Diego airport and Fisherman’s Landing, from which our pelagic trips will depart. We are still in the early stages of planning for this meeting, but expect a wide diversity of presentations, workshops, social events, and full-day and half-day field trips, both terrestrial and pelagic. The conference will include a symposium on avifaunal change in western North America to be published subsequently as a volume of WFO’s monograph series Studies of Western Birds. We welcome and solicit contributions to this symposium and peer-reviewed publication. To inquire about participating, please contact Dave Shuford at dshuford@pointblue.org or Bob Gill at rgill@usgs.gov. Please watch www.westernfieldornithologists.org/conference.php for details as they become available. But plan now to join us next year in San Diego! 315 NOTES BEHAVIORS OF NESTLING AND JUVENILE BLACK VULTURES IN NORTHWESTERN MEXICO MARCELINO MONTIEL-HERRERA and JUAN PABLO GALLO-REYNOSO, Labo- ratorio de Ecofisiologia, Centro de Investigation en Alimentacion y Desarrollo A. C., Carretera al Varadero Nacional km. 6.6, Colonia Las Playitas, Guaymas, Sonora, Mexico CP 85480; marcelino.montiel@estudiantes.ciad.mx The Black Vulture ( Coragyps atratus) is a common carrion-eating bird in much of the Americas (Ogada et al. 2012). Studies in North, Central, and South America have described general behaviors such as nesting, bathing, and drinking (McHargue 1981, Stolen 2000, Sazima 2011). But there are few reports of the vocal sounds made by vultures (Blumstein 1990) and no published sonograms. Here we describe some characteristics of the nest of a Black Vulture in Mexico and publish the first sonograms of the vocalizations of two nestlings and a juvenile. Torote Mountain (27° 53' 50" N, 110° 52' 30" W) is located near “Las Playitas” beach at Guaymas Bay, Sonora, Mexico. This mountain is covered by desert vegeta- tion and contains many small caves. On 7 April 2012, in a difficult climb to reach one cave ~60 m above sea level, we found a nest of the Black Vulture with a brood of two chicks. We measured the cave’s dimensions and placed a motion-detecting camera (Wildview Xtreme Series, STC-TGL4M) at the nest three times for three days each time from March to May 2012 to photograph the vultures’ behavior. The camera was set to take three photographs at 1 -minute intervals each time its motion sensor activated it. The camera remained posted on the ground ~5 m away from the entrance of the cave, tied to a branch of a Palo Blanco ( Acacia willardiana). In May and June, we photographed the brood and recorded its vocalizations with a Fujifilm Finepix S2000 HD digital camera. We analyzed the audio recordings with the software Raven Lite 1.0 (Charif et al. 2006). The sonograms shown here are samples of recordings of calls of two young chicks vocalizing simultaneously and one call of a juvenile. The nest cave was in solid rock with a floor of soil. The outer part or entrance to the cave was well illuminated and wide (Figure 1A), while the inner part provided nar- row access to a small darker chamber of depth ~2 m (Figure IB). During the first two months after hatching, the nestlings moved around the entrance of the cave, where their parents fed them by regurgitation several times during daylight (about every 2-3 hours) (Figure 1C, D). The nestlings were active both day and night; in daylight the birds took dirt baths and sometimes lay on their sides on the ground with their legs fully extended (Figure IE). They grew their black juvenile plumage at an age of ~3 months (Figure IF) yet did not fledge until the age of ~5 months. Notice that the young vulture’s beak is fully black in contrast to that of the adult, whose beak tip is white (Figure 1C, D, F). We recorded what were presumably defense or warning calls, described for the adult Black Vulture as grunts or hisses (www.allaboutbirds.org/guide/Black_Vulture/ sounds). Sample recordings obtained from a distance of 3 m from these two young birds are shown in Figure 2 A. When threatened, the nestlings bent forward with their heads down and spread their down-covered wings (Figure IB). Bouts of hissing lasted ~10 sec, followed 2 sec later by a short (~600 msec) hiss (Figure 2A). The birds then remained silent for several minutes before starting another round of hissing. The frequency range of their vocalizations was up to 2 kHz (Figure 2A). The sonograms show the chicks’ calls to be similar in pattern to the warning calls of juveniles and adults. Only one offspring survived. On 30 June 2012 we observed it ~50 m from the nest. When the bird saw one of us, it jumped up on the rocks and bushes since it was not mature enough to fly. It reached a flat rock, then regurgitated, and made hisses of ~500 msec in a range of frequency up to 9 kHz (Figure 2B). Its hisses were shorter than its calls as a nestling but otherwise similar, as indicated by its sonogram. 316 Western Birds 44:316-318, 2013 NOTES Figure 1. Nest of the Black Vulture found inside a cave on Torote Mountain, Sonora, (a) Entrance to cave. Note the birds near the center of the photo, (b) Two young Black Vultures in what we called their defensive position; these birds exhibited this behavior and hissed when threatened, (c, d) Feeding by regurgitation, (e) Common manner of resting, (f) Young Black Vulture ~3 months after hatching. Note that the sonogram shown in Figure 2B includes sounds made by other local birds and ambient noises. Our report may help clarify some aspects of development of young Black Vultures. We are grateful to Veronica Araiza Sanchez for technical advice and to Aurea Orozco-Rivas and Dorothy Pless for editing the manuscript. Montiel-Herrera held a scholarship from the Centro de Investigacion en Alimentacion y Desarrollo, A. C. LITERATURE CITED Blumstein, D. T. 1990. An observation of social play in Bearded Vultures. Condor 92:779-781. Charif, R. A., Ponirakis, D. W., and Krein, T. P. 2006. Raven Lite 1.0 User’s Guide. Cornell Laboratory of Ornithology, Ithaca, NY. McHargue, L. A. 1981. Black Vulture nesting, behavior, and growth. Auk 98: 182-1 85. 317 NOTES A ' — l — 1 — 1—1 — 1 — l — 1 — 1 — 1—1 — l — 1 — 1 — 1—1 — l — 1 — 1 — 1—1 — I — 1—1 — 1—1 — I — 1—1 — 1 — 1 — I — 1 0.5 1 1.5 2 2.5 3 3.5 t(s) Figure 2. Sonograms of calls of the nestling and juvenile Black Vultures, (a) Relative amplitude of the vocalizations by the two birds inside the cave and frequencies and shapes of their warning calls, (b) Relative amplitude and frequencies of the vocalizations produced by the vulture shown in Figure If. This is the pattern of a characteristic hissing produced by a Black Vulture (black lines at 0.7 and 3.1 sec). Note that the amplitude spikes at ~2.2 sec intervals were made from another local bird, and they were intentionally not subtracted to provide a visual reference of another shape of the sound produced by other bird. Ogada, D. L., Keesing, F., and Virani, M. Z. 2012. Dropping dead: Causes and consequences of vulture population declines worldwide. Ann. N.Y. Acad. Sci. 1249:57-71. Sazima, I. 2011. Black Vulture (Coragyps atratus ): Bath and drink. Rev. Bras. Ornitol. 19:81-84. Stolen, E. D. 2000. Foraging behavior of vultures in central Florida. Fla. Field Nat. 28:173-181. Accepted 18 September 2013 318 BOOK REVIEWS Hawks in Flight (2nd edition), by Pete Dunne, David Sibley, and Clay Sutton. 2012. Houghton Mifflin Harcourt, Boston. 335 pp., numerous color photographs and pen-and-ink drawings. Clothbound, $26.00. ISBN 978-0395709597. This is the second edition of a previous work that I’ve always found useful under many conditions. This edition includes 11 new species and contains useful sketches, emphasizing important topographic traits, and exquisite photos, giving the reader an opportunity to see the raptor under a variety of realistic conditions. It is logically ar- ranged, first by taxonomic group, covering all the raptors that breed in North America, and then calling out those with limited ranges in Florida, the Southwest, and Texas. It is a reference that can be useful in one’s library or backpack. It has the kind of detail in which one can enjoyably get lost from the armchair, but its layout, graphics, and text also make it useful in the field. It’s a book for both the experienced raptor biologist and the serious beginner. Although I’ve worked with raptors for almost 4 decades, I still found many interesting facts, revealing perspectives, and identification hints. The stated aim of Hawks in Flight is to be a book that “integrates an array of carefully selected photographs,... superb illustrations, and a clear, informative text and takes raptor identification to a higher level” and “places in the users’ hands an identification skill set that would otherwise take years to master. ” I liked this edition right from the beginning, as I read the first chapter, titled “The Flight Identification of Raptors: From the Shotgun to the Sublime.” The book goes beyond simply telling the reader what to look for to distinguish one species from another and tells the reader how to look for those features and how to process that information. The basic format is (1) a description of each species, its range, migration, and behavior, (2) a detailed account of its identifying field marks, and (3) an interesting and useful summary section that compares/contrasts similar-looking species. Chapter 2, “Buteos that Migrate,” is long but, like other taxonomically clustered chapters, includes a section on “Migration,” which is very informative. In looking to future publications/editions, the phenomenon of migration might enjoy a comfortable location if it were honored with its own chapter (although I liked the specificity of migration text as it relates to each species of raptor). I welcomed the guidance that the authors give the reader about not jumping to conclusions: “Buteo identification, like the identification of all birds in flight, relies on a number of hints and clues before a judgment is made.” To this end, chapter 2 contains a particularly useful section on the subspecies and morphs of the Red-tailed Hawk. Chapter 3, “Accipiters,” is especially useful to beginners and experienced raptor- philes alike. I knew I was going to enjoy this chapter when I read the chapter subtitle “The Artful Dodgers.” The description of how each accipiter responds to the buffeting of the wind brought to mind what I’ve observed but not really seen (thanks!). Much of the same can be said for chapters 4 (Falcons: Birds that Measure Distance by the Horizon), 5 (Pointed-winged Kites: The Wind-Given Form), 6 (Northern Harrier: The Great Fooler), 7 (Eagles and Vultures: Big Black Birds), 8 (Crested Caracara: A Chimera in Big Black Bird Clothing), and 9 (Osprey: The Fish Hawk). Following that sequence of taxonomy-based chapters is a treat for those hawk watchers who find themselves in the Southwest (chapter 10), Florida (chapter 11), or the Rio Grande and environs (chapter 12). The last chapter (13) is a brief introduction to “Other Birds that Soar.” The bibliography is a welcome inclusion, and although there are very few citations more recent than 2000, most readers will find it valuable. I liked the clever turns of phases in descriptions of raptors and their habitat, whether it be accipiters called “short-range interceptors,” the preferred habitat of the Golden Eagle described as “terrain that is at odds with the horizon,” or a kite likened to a “sonnet... wrapped in feathers.” In speaking about falcons, the authors rightly point out that “to discerning eyes, the way a falcon signs its name is how it flies.” When Western Birds 44:319-321, 2013 319 BOOK REVIEWS comparing the Merlin with the American Kestrel, that signature “is not a matter of degree, it is quantum.” The reader will appreciate the authors sharing facts that are obvious only if the observer has studied many raptor species over a wide spectrum of geography and conditions. For instance, in the chapter on buteos, they relate that “Red-tailed are capable of hover hunting in place, but over much of North America only Red-tailed and Ferruginous hawks are adept at kiting” (p. 24), and then they go on to clearly distinguish between kiting and hovering ; terms which are often misused by even the experienced raptor biologist. This book’s success is a function of four things: (1) the authors’ long history of observing and analyzing raptors in the field, (2) the recent advances in digital photography, (3) the clever and revealing sketches, and (4) thoughtful and careful organization of a meaningful text, punctuated by memorable metaphors. The outstanding level of accomplishment and educational value of this book would not have been possible had there not been the synergism of the recognizable expertise of Pete Dunne, David Sibley, and Clay Sutton. The many strengths of the book include a comprehensive list of North American raptor taxa, descriptions of flight styles and behavior, a detailed look at plumage variation within species (including effects of wear, fading, and molt), and a text that is extremely well written and user-friendly. No important weaknesses come to mind. This is a well-edited, practical, logical contribution to the raptor literature that takes the time to provide meaty, and often entertaining, text. It is a must-read reference book and field assistant. In my opinion, the book meets and exceeds its goals. Jeffrey L. Lincer RICA, La Mesa, California The Crossley Guide: Raptors by Richard Crossley, Jerry Liguori, and Brian Sullivan. 2013. Princeton University Press, Princeton, NJ. 286 pp., numerous com- puter-assisted color photographs in typical habitats, some in the form of a quiz, and species accounts with range maps. Paperbound, $29.95. ISBN 978-0691157405. One of a new series of “ID Guides,” The Crossley Guide: Raptors addresses 34 species. The book is organized into three parts: the first provides computer-assisted images of the raptors, from several angles, in expected habitat/settings, with side- by-side comparisons of the sexes, ages, morphs, and similar species. The photos of each species cover two pages for scarce and localized species or four pages for more common or widespread species. These photos, generally, give the reader an opportunity to see the raptor under a variety of realistic conditions. Some of these plates are in the form of a quiz. The second part consists of species accounts, broken down by overview, flight style, size and structure, plumage, geographic variation, molt, similar species, status and distribution, migrations, and vocalizations. The third part provides the answers to the photo quizzes in the first part. This is a reference and teaching tool that will be most usable from the armchair or in a unique classroom situation. It’s a book for the serious beginner, although I did find a few new “facts.” I put “facts” in quotes because there’s an awkward lack of citations and references, as if the authors didn’t rely on the considerable contribu- tions of other raptor biologists and their previous publications. This omission often struck me as a bit egocentric, nonscientific, and shallow. For instance, the authors indicated that the Falco peregrinus pealei is “partly migratory, with some individu- als wintering as far south as Mexico (p. 253),” contradicting the previous paragraph stating that “Peale’s is a resident [italics added] across the Pacific Northwest coast to the outer Aleutian Islands.” The authors write that the book “is intended as a halfway house between reality and old school traditional guides. It takes you ‘out of the field’ where you can enjoy the beauty of the outdoors and bird from the armchair.” More on that later. 320 BOOK REVIEWS I found the preface a bit off-putting in that it smacked of egocentricity in not giving the giants who came before the authors proper recognition and/or appreciation. The introduction is useful, along with the detailed “raptor topography,” although some of the lines connecting the topographic features and their names are hard to see. The first part of the book, “Raptor Images,” is an interesting approach and may be useful to the beginner, but I felt like it was needlessly touted/oversold in more than one instance. In general, the book needed better editing. In many instances, the text editing was incomplete, if not haphazard. In some cases, it was difficult to match up the colors as described in the text with the colors in the photos. Other examples include typos, repeated sentences, sentences so shortened they are not understand- able, facts repeated in the same section, half words, poor punctuation, etc. The use of the current 4-letter acronym is helpful, and I enjoyed the clever presentation of a variety of lighting conditions (black and white, looking into the sun, and sunrise lighting). Those should be helpful to the user. I liked the way that the species accounts are organized. The unique first paragraph is commonly told from either the perspective of an uncomfortable raptor biologist/ birder in the field or from the perspective of a hungry raptor — an approach that’s innovative and entertaining and should tend to keep the reader’s attention. The distinctions between the juvenile and adult plumages are, usually, made quite well. Discussions of what species might be confused with another were helpful and, in some cases, revealing; the instances of hybridization were interesting (but not confirmable without references). The species accounts, like the first part of the book, could have benefited from better editing (e.g., poor use of punctuation, apparent inconsistency between the text and the range map, apparently contradictory statements). The inclusion of a glossary is a nice, and useful, touch. An index is usually a wel- come addition to any book. But this book’s index essentially duplicated the table of contents in that it was limited to the species of raptors. If there is a future reprinting, the authors might consider either leaving the index out (and saving a page) or extend- ing its coverage to include major topics. As with most collaborations, I believe that the combined talents of each of the authors contributed positively to a better product than would have been possible without that collaboration. Does the book meet its goals? It is an interesting, if not well-edited, book that will get the reader “out of the field." It can accomplish that goal, but why would we want to do that? Given the need to get kids back into their environment, this seems like a goal that completely dismisses the “nature deficit disorder” that plagues the children of today, who will never be the environmental stewards of tomorrow if they continue to spend all their time in the armchair and not in the woods. Although I would not recommend this book for the experienced raptor biologist/ enthusiast, it can certainly be appreciated from one’s armchair, especially by those who are just beginning to get interested in raptors. It also has potential utility as a text for a class or workshop on raptors. One concern I have is the way the book is bound. The cover is glued to the pages adjacent to it in a way that makes me wonder if it will hold together. It may, but I have not seen this kind of construction in the past and, certainly, wouldn’t test its strength by taking it into the field, even if it stayed in my backpack. Jeffrey L. Lincer RICA, La Mesa, California 321 FEATURED PHOTO HYBRIDIZATION OF THE BLACK-FOOTED AND LAYSAN ALBATROSSES CAMERON RUTT, P. 0. Box 223, Blooming Glen, Pennsylvania 18911; cameronrutt@gmail . com ABSTRACT: Although the Laysan ( Phoebastria immutabilis) and Black-footed Albatrosses ( P. nigripes) have been known to hybridize for more than a century, little has been published regarding plumage variation of the hybrid progeny. During six months of field work on Laysan, Hawaii, I noted 13 possible hybrids (five presumed ¥i hybrids, three possible F 2 backcrosses with the Black-footed Albatross, and at least four possible F 2 backcrosses with the Laysan Albatross). Apparent F 2 backcrosses with the Black-footed Albatross differ from it most noticeably in their black-and-white underwings and much more extensive white circling the face. Apparent F 2 backcrosses with the Laysan Albatross differ from that species most noticeably in their extensive gray smudging throughout the body and darker underwing coverts. Apparent F 2 backcrosses interbreed with the Black-footed Albatross, the first evidence of any hybrid pairing with that parental species. The Laysan ( Phoebastria immutabilis) and Black-footed Albatrosses ( P. nigripes ) have been known to hybridize for more than a century (Rothschild 1900, Fisher 1948). The breeding ranges of these two species overlap widely in the northwestern Hawaiian Islands, particularly on Midway Atoll and Lay- san Island (Awkerman et al. 2008, 2009, Pyle and Pyle 2009). Although hybrids receive mention in a few contemporary field guides (Sibley 2000, Dunn and Alderfer 2006), because they are not illustrated or often observed at sea, hybrid phenotypes are not well known to many field ornithologists, though a few photographs were published by Awkerman et al. (2009), Pyle and Pyle (2009), and Howell (2012). Only McKee and Pyle (2002) have discussed the subject of hybridization between the Laysan and Black-footed Albatrosses with regard to plumage variation in each species. In North American waters, presumed hybrids have been found very infre- quently (McKee and Pyle 2002, Howell 2012). Although a few have been reported off California (Debra Shearwater, pers. comm.), none have been confirmed there (Hamilton et al. 2007), and several bleached Black-footed Albatrosses have been misidentified as hybrids (Figure 1; McKee and Pyle 2002). Hybrids have also caused confusion with the Short-tailed Albatross ( P. albatrus): Roberson (1986) suggested that one of the first Short-tailed Alba- trosses reported in California in the 20th century was one of these hybrids (Helm 1980, McKee and Pyle 2002). Despite extensive research, I am aware of just one individual documented at sea, a presumed first-generation (FJ hybrid photographed by Robert L. Pitman off the Aleutian Islands, Alaska, on 2 June 2002 (Howell 2012: figure A2a.2, Sophie Webb pers. comm.). In this paper I aim to update and expand upon McKee and Pyle (2002), presenting further photographs of presumed ¥i hybrids, possible second- generation (F 2 ) backcrosses with both parental species, another bird of inde- terminate hybrid origin, and a particularly unusual aberrant plumage aspect of the Black-footed Albatross. These variations may be sources of confusion 322 Western Birds 44:322-333, 2013 FEATURED PHOTO Figure 1. Apparently a very bleached Black-footed Albatross. Aside from the conspicuously pale head, which could suggest hybrid origin, the bill structure and plumage, with brown tones throughout, are consistent with the Black-footed Albatross. West of Kruzof Island, near Sitka, Alaska, May 2012. Photo by Linda Behnken with other rare or vagrant albatrosses in the North Pacific (Roberson 1986, McKee and Pyle 2002, Howell 2012). MECHANICS OF HYBRIDIZATION Hybridization in tubenoses is very rare (Howell 2012), and — despite being North America’s only well-documented example — successful hybridization between the Black-footed and Laysan Albatrosses is also rare (Fisher 1972, Howell 2012). During six months (September 2011-March 2012) of field work on Faysan, Hawaii, I noted only 13 possible hybrids (five presumed Fi hybrids, three possible F 2 backcrosses with the Black-footed Albatross, at least four possible F 2 backcrosses with the Faysan Albatross, and one bird of unknown hybrid origin) among approximately 150,000 and 24,000 breed- ing pairs of the Faysan and Black-footed Albatrosses, respectively (Pyle and Pyle 2009). McKee and Pyle (2002) suspected that hybrids arise through forced extra-pair copulation of male Black-footed with female Laysan Alba- trosses, rather than through bonded mixed pairs. Alternatively, it is possible that hybrid progeny originate from females that are willing to accept these extra-pair copulations (Tristan McKee pers. comm.), although there is no direct evidence for either scenario. Brief interspecific courtship dances have 323 FEATURED PHOTO been reported (Fisher 1972, McKee and Pyle 2002), yet I never witnessed any such dances on Laysan, nor did I detect any evidence of bonded mixed pairs, supporting the hypotheses of either forced or voluntary extra-pair copulations. HYBRIDS In addition to being intermediate in plumage aspects, hybrids differ from the parental species in vocalizations, bare-part coloration, and behavior (Fisher 1972; Table 1). Behaviorally, the two species have numerous dif- ferences in their courtship dances (Fisher 1972). The Black-footed and Laysan Albatrosses also walk with noticeably different postures; Black-footed Albatrosses walk with the neck partially extended in front of the body and the head held low, while Laysan Albatrosses maintain a nearly vertical neck posture and the head held high (Fisher 1972). Furthermore, preferences for breeding habitat differ: the Laysan Albatross prefers the vegetated, interior portions of islands, the Black-footed Albatross more exposed outer sandy beaches (Fisher 1972). All of these reported behavioral differences were consistent with my observations on Laysan. When possible in the field, I attempted to assess apparent hybrids’ plum- age aspect, bare-part coloration, walking posture, colony composition, and breeding behavior. Unfortunately, my departure in March precluded the opportunity to determine whether chicks successfully fledged and pre- vented documentation of plumages of hybrid progeny. Without marking of individuals or genetic confirmation, these observations of phenotypes are all necessarily tentative. Categorizations of presumed hybrids, especially of possible backcrosses, are inferences. “CLASSIC” PRESUMED F : HYBRIDS In all respects — physically, behaviorally, and reproductively — the five pre- sumed Fj hybrids appeared more closely aligned to the Laysan than to the Black-footed Albatross. If the forced-copulation hypothesis is correct — a female Laysan Albatross being fertilized in an extra-pair copulation with a male Black- footed Albatross — it could explain the behavioral and reproductive associations with the Laysan Albatross, as an Fj chick would have been raised by a pair of Laysan Albatrosses (McKee and Pyle 2002). The main physical difference between F : hybrids and the Laysan Albatross was the former having pearly to smoky gray head and neck feathering (Figure 2), which ended rather abruptly on the lower breast, contrasting markedly with the white belly. The pale pat- terning on the head, with a white nose band and lower white eye-arc, was more reminiscent of the Black-footed Albatross, but the white across the forehead was more extensive and extended into the eyebrow. Structurally, the bill shape matched that of the Laysan Albatross, appearing longer and narrower than that of the Black-footed Albatross (Pyle 2008) and having a more obvious saddle to the culminicorn. Bare parts were generally darker than those of a Laysan Albatross; the bills of Fi hybrids had a dusky tip (darker than the blue-gray of an average Laysan Albatross) and a dark base to the culminicorn, lacking the yellowish-orange base of the bill of an adult Laysan Albatross. The legs of Fi 324 (/) Q) O Q) a co 13 -j— ' c Q) (C Oh 0 ) o , i-H co 0 ) CO CO O - 4—1 (C -O "O 0 ) -t— * O o -x: o JV CQ X c cC co S' J "O 'C -O X "O 0 ) £ s 0 ) 0 - Oh CO 0 ) CL 3^ o c 0 ) _c Oh C '?3 0 ) £ H 0 ) _c -t— j CO c X co 3 CO c co 5 cC •c 0 ) 'C cj I CO a) CO CO O co co O S-h 16 x> d < CJ (-H s § ^ co CM H>> U-H G JD t9 ’co CO O Oh -J 0 ) -G CO 5 8 co •+- ; 0 ) td £< 03 "O -X 0) o -t; ro O -O O M— t CM ' r i . dd ^ CJ JD JO t 2 CQ CO -V co 2: O dC Oh ^ co J3 'G -G u3 30 CD £ 3 0 ) S-h Oh CO CO o S-H -(— 1 (O £ < 30 0 ) -t — 1 o 0 M— i 1 dd CJ JO PQ co co O S-h Id X) C CO co (O 0 ) -t — 1 CJ (0 S-h CO x: CJ 30 0 ) CJ g 30 0 ) 5-h x: 1 g X) 0 ) .> ‘co c 0 ) ~x 0 ) CD G (D > (D G 6 P 30 33 co £>- cn CO S-H (13 e -p e (0 > o 3 £ -S £ -a p S’ p £ 3 (is '— J 03 p 3=: o ^x: in J ^ 03 ro e ^ § O v +- 1 3 TD (13 03 "O s- - (13 JS 03 a o o.-g (a 3 "O e o,d & £ gi 3 co .£ _o 5 C33 •S3 03 -C 3x^ +-. +-■ e a cn o -g c -j, ro ro -g I.S I _J CO cn O-B-Sss ~ £ c o) ,— ] cn d3 -O "O Si £ (13 (0 D 1 C -C 5 ±3 O 5 O id CD -O i£ -X 03 "D £ co cn -d e 3 O' J-H (13 cn -C !> o 03 CO c -a e (13 3 .£ d ^ <0 QJ cn T3 >> C 33 (13 — 1 Q3 03 rtJ tS -G J 3 CO 76 £ p _c cn 0 G 03 O 3 JS o JD 03 03 J2 3 T3 +3 G 3 (13 •° 03 c J (13 -5 S" 5 ro "G J 8 03 3 -33 ~0 •G P C 3 o 5— X) x: 5— CO dO (0 CD cj cO J2 £ G 3 .3 sz c p 03 -C X 03 .S 03 O f 2 £ £ 0-. -i..! +± c £ ° X3 03 T3 03 P -C "G -*— 1 O "S o 3 M— < _ G3 > J O 03 CQ ^ ^ 'i 03 03 >4 +3 G 2 -g J cn 3 33 03 P -k C co^ S' ■£ 3 i-e -O 03 Q P j£ C x o O 3 £ . -G CD ~t — 1 CD -P id -§ x: > .11 c -o' 8 p P s X o ^ 03 O P Q) 7 £ s- 33 O O) CO C jO C J cQ d 3 ro| S S 03 G 0 ) ^ £ 8 c ~ -O £ -a 33 _P O (ri (0 G- CL 33 -G U .13 ■G ^ ^ G CQ ^ S 03 C £ g’C. C 3 Q) ■SP C/) CO CD dD CD G — 1 o o G i u- O CJ o jg ;§ cq £ 92 ■ i P — , CJ ~o CD 1 O 0 M— < 1 x: o JO CQ CD o ^ CO tO CD _>3 S^l Cl, (5 c?5 C 03 o 00 '£ £ 03 cc -a c (13 -a c (13 -O 03 cn O C 03 S' £ S O cn (13 c cn 03 C O 0.-^1 Oh T3 03 c, c ca 5 ro p 3 S' .-g 4- 2 _G o o 3 03 p -£ 3^ 03 ,33 > £ O jy rope 03 ^ -£ 3X o X 03 -S 03 => P 3 33 O S oh ro x -a 03 ., O' 3 03 £ ~o c c (13 <13 cn T3 O' ro ro p _] -G d p SJ s-H _33 O i or P 03 S cn _!S S CL ro 8 co c ‘I J CO d5 x: o CD co fO X) CD CO G tO J- c o Is ", co —G .SQ tO 5 ■— 1 O CD 1=3 dd CD ^ CD C D td S-H CD CD X) S-H td CD > S-H -3 CD O CD G 5 td CD td CO >> (D O S-H X) £ dd 2 -4— < dd "G CD S-H td td > £ Q td O CO CD co t< £ g 03 G3 03 £ -S <0 s JD PL P 03 _ cn P ro o jp c CD CD G -S 3 5 o e >T= M — -G ~ 3 S 33 O ^ b'CTl 5 3 c Q ro ro 03 5— CO >* CD O CD x: CJ S X4 > ^ (O S £ -p 3 1 'c75 P 03 O 03 CO p 03 P ££ J P s G3 s o -g u ro ST (13 03 £ -G 5 o cn P -Q co (0 r7 CD .i=5 d -i—i Xh CO tO i— < CJ CO X CD co cO co G X) CD tO U * G r >< £ ro U 3 > CO _ P G • d 5— ^ CD 5— -*— 1 1 CD fO -O CD G D dd ^ ^■ss X ro > CO CD CD Z£ to _ro 2 3 p 'P O 3 03 ro 3 "ro £ £ 3 ro 3i <13 3^-2 3 oo 3 03 _ro ,3 ^ cn CO^ .£ C P ‘S' £ X ^ co dd G Q 5 d -g in .2 3 & d £3 £ 33 j" G fO fO x: cd di ^ *!> fO > X) dd CD .52 cn n a —3 (O O 5- -id CD CO -pp ^ o CO ■-S co ■ £ £ ro co- cl ro £ o > CJ JS -a 3 O) 3 -a 3 S-H (13 -P (13 _P CO P _c '& 33 -a 3 3 -C 4— < 3 3 -O -- e P p 0 P G G d ro p -a td td td cn CO P CO CO CO O' c ■£ (13 .b -C td td td -j cpd P 3 ® J — ] J CD CD (D X § 3 dd dd dd — 1 8 .2 13 13 13 cn 3>s ro <£> ro co § p 'Si > (D co G CD x Jr, CD d=l G to co to CD dd tO Ph _C cn ^ dd > p- ^ to dd G X) co G CD > G co 65 ^ x j g CD -S' dd .b . to CD ^ - 7 ^ Gd .5=2 dd cd to dd co (D ^ cn _2 <0 3 G 1 ° 2 p -=6 -X c ro g G3 §- _S33 x 33 - P ro P m s- 1 3 h JO -G > -a 3 g 3 ro 03 _ p _c dd s-h * G O q to £ £ g 1 (13 V ^ -G -g si ® dd G C to CD > to (D dd G D CD 16 CD (D cj fO UU Oh O G O Xi fO S-H O O p ro _o p o cc .3 03 MM 3^ cn 3 cn P CL O ro 4-4 _c 03 3 p -a CL-G 3 -G -G .£P p" -a (13 03 -a S ' -3 G (13 ^ £ d 8 -Oh cn 3=1 CO 3 -2 -a 3 03 JC p ro cn 3>s (13 3 -X _o 0 03 1 -C cn P O CO p M d 'd o 2 £ s-h ro CO 13 V* \ ^ S 2 JP to Q CD dd CD -t — 1 1 o o M— I dd CJ jd PQ CD dd dd G td co cn S 2 dd G G O J-H td cn G '3 CO ^ dd 3 G £ td co *-•— ! CD dd G O G ^ s-h cd CD s-h dd cn S-H -2 •£ - 3 3 p 33 g"-; e" oiC§ cn P ”0. cn — ■ e s> 3 ro Oh J 03 03 sC -X 03 -X O >= JX £j £ -a £3 p ro 04 3 — ro _o P . — — -X 33 3D (0 3 -X 03 -2 -O -P CO ’Ss td S-H JdO dd S-H td Q cd td dd CD g x: CD -(-3 dd X id CD > cn G o dd co 13 G E PQ CD > td -i— T O Gh O co G CJ G cj o) JD CD > J •So d -G 2 - 5 - 3 G- o o -X -y o P S 2 d ^ ro o 03 -C > CO -dd ro 2 £ "p P -C 8 G Z -3 CD S-H G -t— < co O CD cn G 13 td 325 FEATURED PHOTO hybrids were darker than those of the Laysan Albatross and more blue-gray, with gray or dark purplish feet, generally lacking pink tones. These hybrids also showed fairly extensive dark gray smudging around the legs and flanks, areas that are cleanly white on a typical Laysan Albatross. Finally, the underwings of Fi hybrids that I was able to observe showed reduced white and more extensive black, particularly in the humerals (upper photo on this issue’s back cover). Behaviorally and reproductively, these F : hybrids were also more similar to the Laysan Albatross. Fj hybrids walked with the head held high, like a Laysan Albatross. Minimal observations of these hybrids’ courtship dances also suggested they were more like those of the Laysan Albatross (Fisher 1972). Similarly, their territories were in locations dominated by Laysan Albatrosses (average 76% of the nearest neighboring birds); one hybrid was in a predominantly Black-footed Albatross colony (nearest neighbors 28% Laysan Albatrosses), whereas three were located in the more vegetated inte- rior of the island, in areas occupied almost exclusively by Laysan Albatrosses. I detected only one of the five Fj hybrids engaged in successful reproduc- tion (Figure 3). It was paired with a Laysan Albatross and was incubating and/or brooding a chick from 24 December 2011 through at least 11 February 2012. POSSIBLE BLACK-FOOTED ALBATROSS BACKCROSSES In order for this phenotype to be generated, I speculate that Fi hybrids bred with a Black-footed Albatross. And as with the origin of Fj hybrids, I hypoth- esize that this backcrossing resulted from extra-pair copulations of a male Black-footed Albatross with a female hybrid, as all the aforementioned evidence suggests an otherwise strong affinity of Fi hybrids for the Laysan Albatross. Unlike F x hybrids, possible F 2 backcrosses with the Black-footed Albatross were physically, behaviorally, and reproductively more similar to that spe- cies. These F 2 hybrids appeared paler than the whitest extreme of the oldest male Black-footed Albatrosses (lower photo on this issue’s back cover). The head and neck of these hybrids were gray-brown, and they had more gray in the upperparts as well. A white face encircled the eye and bill (not limited to a white nose band); this bold facial pattern may seem novel for this spe- cies pair, yet it mirrors the paler cheeks of the Black-footed Albatross. The belly of F 2 hybrids was white, with a cutoff to the darker gray neck as in Fj hybrids (Figure 4). Compared to that of Fi hybrids and Laysan Albatrosses, the bill appeared comparatively stout and lacked the distinct saddle to the culminicorn, matching the Black-footed Albatross. The bill was significantly paler than that of the average Black-footed Albatross, being drab, pale pink with a dark base to the culminicorn, again as in Fi hybrids. Their dark gray or flesh-colored legs and feet were also paler than those of a Black-footed Albatross, with some mottling. Although I obtained only a single look at one of these hybrids’ underwings, it had strikingly white coverts with black primary coverts and humerals, much as in a pure Laysan Albatross, and very unlike a Black-footed Albatross. These three apparent F 2 hybrids all closely matched an “unidentified albatross” photographed on Laysan in 1995, McKee and Pyle’s (2002) figure 19, which those authors cautiously 326 FEATURED PHOTO left unresolved without the aid of any behavioral information. Although not previously described well, these hybrids are not just of recent occurrence; Peter Pyle (pers. comm.) photographed a similar bird on Laysan as early as 1984. In addition to their plumage characteristics, these hybrids acted like and associated with Black-footed Albatrosses. I observed all three of these pos- sible F 2 hybrids walk with a crouching posture, like a Black-footed Albatross. Additionally, one of these hybrids engaged in an extended bout of courtship dancing with a Black-footed Albatross (upper photo on this issue’s inside back cover), matching the style, complexity, and tempo of the Black-footed Albatross’s display and unlike that of a Laysan Albatross (Fisher 1972). These possible F 2 hybrids were strongly associated with Black-footed Albatrosses, which were 93% of their nearest neighbors. If these birds were, in fact, the result of hybridization, these observations belie the previously reported preference of hybrids to socialize only with Laysan Albatrosses (Fisher 1972, McKee and Pyle 2002), an observation based solely upon Fj hybrids. All of these F 2 hybrids were found on exposed outer sandy beaches typical of the Black-footed Albatross’s habitat. Two of these three F 2 hybrids were paired with Black-footed Albatrosses and shared parental duties (Figure 5), successfully hatching chicks (26 De- cember 2011-17 February 2012; 20 November 2011-17 February 2012). No evidence of a hybrid successfully breeding with a Black-footed Albatross has been reported previously (Awkerman et al. 2008). POSSIBLE LAYSAN ALBATROSS BACKCROSSES Identifying possible F 2 backcrosses with the Laysan Albatross was difficult, as it was impossible to delineate the extreme in darkness of a genetically pure Laysan Albatross (McKee and Pyle 2002). Because of this lack of clear distinction, multiple F 2 backcrosses with the Laysan Albatross may have passed unnoticed — unlike the other hybrid phenotypes, which were very conspicuous. Individuals into which Black-footed Albatross DNA may have introgressed from an interspecific hybrid of a past generation had more ex- tensive gray smudging overall, including the cheek, crown, lower hindneck, bend of the wing, uppertail coverts, and particularly around the lower belly, vent, and undertail coverts (lower photo on this issue’s inside back cover). Additionally, multiple birds showed reduced white in the underwing coverts, and one individual appeared to have largely dark underwings. Otherwise, these birds suggested a dark Laysan Albatross. UNKNOWN HYBRID COMBINATION On 8 March 2012, I observed briefly in flight a fourth type of possible hybrid. Its head and neck were like those of a Laysan Albatross, but it had a sooty body and undertail coverts, dark underwings, and grayish uppertail coverts, much like an apparent hybrid photographed on Midway Atoll 28 December 2006 (Figure 6). 327 FEATURED PHOTO Figure 2. The most commonly detected hybrid phenotype, a presumed F x hybrid. These hybrids looked and acted much like the Laysan Albatross, but note the largely pearl gray head and neck, with white facial patterning like that of a Black-footed Albatross. Laysan Island, 18 December 2011. Photo by Cameron Rutt Figure 3. Another presumed Fx hybrid, this was the only individual engaged in successful reproductive behavior (here seen incubating), although actual parentage of the chick is unknown. This hybrid shared duties with a Laysan Albatross and successfully hatched a chick. Laysan Island, 25 December 2011. Photo by Cameron Rutt 328 FEATURED PHOTO Figure 4. This possible F 2 backcross with a Black-footed Albatross looked noticeably grayer than that species in flight, and it had an obvious white face. The underwing pattern was strikingly unlike that of a Black-footed Albatross, with largely white underwing coverts and black primary coverts and humerals like those of a Laysan Albatross. Laysan Island, 12 November 2011. Photo by Cameron Rutt Figure 5. Another possible F 2 backcross with a Black-footed Albatross, this was one of two that paired with a Black-footed Albatross and successfully hatched chicks. If these birds are correctly identified as hybrids, this is the first documentation of a hybrid breeding with a Black-footed Albatross. Laysan Island, 11 February 2012. Photo by Cameron Rutt 329 FEATURED PHOTO Figure 6. Although this presumed hybrid could at first suggest a different species, the bill structure and color, as well as the head pattern, are all consistent with a Laysan Albatross. It is unclear how this phenotype might arise, although it appears most similar to an extremely dark version of a possible F 2 backcross with a Laysan Albatross. The pattern of the underwings would help categorize it. Midway Atoll, 28 December 2006. Photo by Ryan DiGaudio ABERRANT BLACK-FOOTED ALBATROSSES Two albatrosses on Laysan had plumage aspects that could suggest hybrid- ization; however, they seem more likely to represent aberrant Black-footed Albatrosses (Figures 7 and 8). Nearly identical, these birds were like pale, scaly, and frosty versions of a Black-footed Albatross, looking largely beige and silver. Their white faces recalled the pattern of a Black-footed Albatross and were much like those of the possible F 2 backcrosses with the Black- footed Albatross, but their bills were bright pink. Unlike these F 2 hybrids and the Laysan Albatross, however, their upperwings had dark patches and their underwings appeared much like a uniform, paler version of a Black- footed Albatross’s. The underwing pattern may be critical in distinguishing these birds from possible F 2 backcrosses with the Black-footed Albatross, which could appear similar, particularly at sea, where other details would be less apparent. These individuals were remarkably similar to a leucistic Black-footed Albatross photographed off Santa Barbara, California (Howell 2012: figure 25); one may have been the same individual. McKee and Pyle (2002) addressed aberrant plumage aspects of the Black-footed and Laysan Albatrosses more exhaustively. 330 FEATURED PHOTO Figure 7. Initially thought to be a hybrid, this aberrant Black-footed Albatross had the white face of possible F 2 backcrosses with the Black-footed Albatross (Figure 8) but lacked any other characteristics of the Laysan Albatross. It associated with Black- footed Albatrosses and walked with that species’ crouching posture, as did the bird pictured in Figure 8. Laysan Island, 27 November 2011. Photo by Cameron Rutt Figure 8. An individual different from that in Figure 7, though nearly identical, distinguishable by different apparent molt limits (or, perhaps, leucism) in the primaries. This aberrant Black-footed Albatross could cause confusion at sea, but it lacked the white and black underwing pattern of all presumed Fx and possible F 2 hybrids. Laysan Island, 15 December 2011. Photo by Cameron Rutt 331 FEATURED PHOTO CONCLUSION Although rare on their breeding grounds and scarcely observed at sea, presumed hybrid Black-footed x Laysan Albatrosses may be categorized by at least three different but consistent phenotypes. Table 1 outlines criteria for distinguishing these three phenotypes from the parental species. In par- ticular, underwing pattern may be critical in revealing hybrids or classifying problematic individuals. Most previous documentation of hybrids stems from Midway, so my observations on Laysan Island complement those descrip- tions. Further observation of hybrids may reveal that at sea their distribution tracks that of the Laysan rather than the Black-footed Albatross, concen- trated in the Gulf of Alaska and around the Aleutians. In any case, birders throughout the North Pacific should watch for these birds and document them when possible, as hybrid albatrosses are truly rare at sea. ACKNOWLEDGMENTS I thank the American Bird Conservancy, U.S. Fish and Wildlife Service, and the Papahanaumokuakea Marine National Monument for funding and facilitating my time on Laysan. Andrea Kristof, Robby Kohley, and David Tafoya all located hybrids, and Jack Toriello assisted me in monitoring their breeding behavior. I thank Andrea Kristof, especially, as well as Pete Leary and Robby Kohley, for entertaining conversations with me about hybrid albatrosses. Peter Pyle provided thoughtful input, first-hand knowledge, and motivation that were critical in crystallizing my field impressions. I’m grateful to Oscar Johnson, Tom Johnson, Tristan McKee, Peter Pyle, Sophie Webb, and Stephanie Wheeler for reviewing previous drafts of the manuscript, adding many helpful suggestions and edits, as well as Linda Behnken and Ryan DiGaudio for gra- ciously allowing use of their photos. Finally, thanks to Linda Behnken, Michael Force, Kimball Garrett, Steve N. G. Howell, Paul Lehman, Ryan Merrill, Debi Shearwater, Scott Terrill, Thede Tobish, Bill Tweit, David Vander Pluym, and Sophie Webb for responding to questions about occurrences of hybrids off North America. LITERATURE CITED Awkerman, J. A., Anderson, D. J., and Whittow, G. C. 2008. Black-footed Albatross (. Phoebastria nigripes), in The Birds of North America Online (A. Poole, ed.), no. 65. Cornell Lab of Ornithology, Ithaca, NY; http://bna.birds.cornell.edu/ bna/ species/065doi : 1 0 . 2 1 7 3/bna .65. Awkerman, J., Anderson, D., and Whittow, G. C. 2009. Laysan Albatross ( Phoe- bastria immutabilis), in The Birds of North America Online (A. Poole, ed.), no. 66. Cornell Lab of Ornithology, Ithaca, NY; http://bna.birds.cornell.edu/bna/ species/066doi: 10.217 3/bna. 66. Dunn, J. L., and Alderfer, J., eds. 2006. Field Guide to the Birds of North America. Natl. Geogr. Soc., Washington D.C. Fisher, H. I. 1948. Interbreeding of Laysan and Black-footed albatrosses. Pacific Sci. 2:132. Fisher, H. 1. 1972. Sympatry of Laysan and Black-footed Albatross. Auk 89:381-402. Hamilton, R. A., Patten, M. A., and Erickson, R. A. 2007. Rare Birds of California. W. Field Ornithol., Camarillo, CA. Helm, R. C. 1980. A Short-tailed Albatross off California. W. Birds 11:47-48. Howell, S. N.G. 2012. Petrels, Albatross & Storm-Petrels of North America. Prince- ton Univ. Press, Princeton, NJ. 332 FEATURED PHOTO McKee, T., and Pyle, P. 2002. Plumage variation and hybridization in Black-footed and Laysan Albatrosses. N. Am. Birds 56:131-138. Pyle, P. 2008. Identification Guide to North American Birds, Part II: Anatidae to Alcidae. Slate Creek Press, Point Reyes Station, CA. Pyle, R. L., and Pyle, P. 2009. The birds of the Hawaiian Islands: Occurrence, history, distribution, and status, version 1. B. P. Bishop Museum, Honolulu; http://hbs. bishopmuseum . org/birds/ rip-monograph . Roberson, D. 1986. Ninth report of the California Bird Records Committee. W. Birds 17:49-77. Rothschild, W. 1900. The Avifauna of Laysan and the Neighboring Islands, with a Complete History to Date of the Birds of the Hawaiian Possessions. Porter Ltd., London. Sibley, D. A. 2000. The Sibley Guide to Birds. Knopf, New York. THANKS TO WESTERN BIRDS’ REVIEWERS AND ASSOCIATE EDITORS Peer review is a critical step in the publication of a scientific journal. I thank the fol- lowing people for their generosity in taking the time to provide this essential service sustaining the scientific quality of Western Birds for volume 44: Kenneth P. Able, Dan Airola, Edward C. Beedy, James Bendell, M. Ralph Browning*, Brent Campos, Jameson F. Chace, Amy Cilimburg, Mark A. Colwell, Alan Contreras, Jeff N. Davis, Jon L. Dunn*, Scott Durst, John Eadie, Douglas W. Faulkner*, Alissa Fogg, Kimball L. Garrett*, Jennifer Gervais, Daniel D. Gibson*, Robert E. Gill Jr.*, Murrelet Halterman, Steven C. Heinl*, Dale Herter, Jennifer A. Holmes, Grainger Hunt, Oscar Johnson, Ian Jones, Stephanie L. Jones, Michelle L. Kissling, Walter Koenig, Paul E. Lehman*, Bruce Lyon, Michael McGrann, Steve Mlodinow, Kathy C. Molina, Joseph Morlan, Joan Morrison, Kristie N. Nelson, Javier Salgado Ortiz, Gary W. Page, Michael A. Patten, Elizabeth Porzig, Peter Pyle*, Kurt Radamaker, Gary H. Rosenberg, Daniel R. Ruthrauff, Leo Salas, Michael A. Schroeder, Bryan Sharp, W. David Shuford, Steve Shunk, John Tautin, Francis X. Villablanca, Sophie Webb, Clayton M. White, David Wien, John C. Withey, and Jack J. Withrow. Asterisks designate reviewers who reviewed more than one paper. As always, I must thank our hard-working associate editors, Kenneth P. Able, Doug Faulkner, Thomas Gardali, Daniel D. Gibson, Robert E. Gill, Paul Lehman, Ron LeValley, and Dan Reinking, plus featured-photo editor John Sterling, all of whom serve also as reviewers. Western Birds is not possible without their dedication. Philip Unitt 333 ERRATUM In Western Birds 44(3), Figures 5, 6, and 7 in the article Dark-faced Common Murres of Central California in Fall and Winter ( Western Birds 44:250-259, 2013) were inadvertently omitted; they are reproduced here. I apologize for this unfortunate lapse to author Peter Pyle and the readers of Western Birds. Philip Unitt Basic Plumage 100 n o £ 60 O 12 3 4 Formative Plumage 100 c 80 o t 60 O Facial Plumage Score □ Specimens □ Central CA □ Fort Bragg Figure 5. Distribution of facial scores (see Figure 1) of Common Murres in basic and formative plumage on the basis of specimens and birds studied in the field. The latter were scored between 30 September and 6 November, when birds should have few or no alternate feathers (see text). Birds observed on four trips from Bodega Bay to Half Moon Bay (30 September-6 November, 2011 and 2012) and on one trip off Fort Bragg (6 November 2011) analyzed separately. See text for statistical comparisons. Sample sizes for basic plumage: 61 specimens, 209 live birds off central California, and 55 live birds off Fort Bragg; for formative plumage: 29 specimens, 100 live birds off central California, and 31 live birds off Fort Bragg. 334 Western Birds 44:334-335, 2013 ERRATUM Figure 6. Variation in the head plumage of chicks and juveniles of the Common Murre. The nestling (top, CAS 88095) was collected on Southeast Farallon Island 7 June 1964; the juveniles (middle, CAS 15545; bottom, 15546) were collected on Monterey Bay 19 August 1909. CAS 88095 shows the pattern typical of natal down; CAS 15545 was one of only three of the sample of 33 juveniles showing dark auriculars. Figure 7. Completely dark Common Murre with two in typical alternate plumage, photographed 24 July 2013, 14 km southeast of the Farallon Islands, flying south. That the dark bird was carrying a fish indicates it was breeding, perhaps on the Farallon Islands. Photo by Dru Devlin NOAA/ONMS/ACCESS 335 WESTERN BIRDS, INDEX, VOLUME 44, 2013 Compiled by Daniel D. Gibson Acanthis flammea, 232 Accipiter striatus, 2-16 Acrocephalus dumetorum, 191 Aeronautes saxatalis, 2-16 Agelaius phoeniceus, 2-16 tricolor, 98-113 Aimophila ruficeps , 2-16 Albatross, Black-footed, 322-333 Laysan, 322-333 Short-tailed, 175-176, 211 americana, Bucephala clangula, 191 Ammodramus lecontell, 231 savannarum, 30-31 Amphispiza belli , see Artemisiospiza belli Anas crecca, 19 falcata, 210 penelope, 19 Anderson, Leslie Scopes, see Burton, K. M. Anser albifrons, 184 fabalis/serrirostris, 207-210 Aphelocoma californica, 2-16 Aphriza uirgata, 65-68 Apus pacificus, 239 Aquila chrysaetos, 2-16 Archilochus alexandri, 2-16 colubris, 226 Ardea herodias, 239 Ardeola bacchus, 184 Artemisiospiza belli, 2-16 Asio flammeus, 237 Auklet, Rhinoceros, 178 Baeolophus inornatus, 2-16 Bartramia longicauda, 21, 222 Baumann, Matthew J., Book review: Raptors of New Mexico, 158-159 Baumann, Matthew J., McNew, Sabrina M., and Witt, Christopher C., Morphological and molecular evidence confirm the first definitive Eastern White-breasted Nuthatch {Sitta c. carolinensis) for New Mexico, 90-97 Bean-Goose, Taiga/Tundra, 207-210 Benally, Ty J., see Howard, P. J. Bernstein, David P., see Brook, C. E. Bittern, Eurasian, 191 Blackbird, Red-winged, 2-16 Rusty, 231 336 Tricolored, 98-113 Black-Hawk, Common, 20 Bland, James D., Estimating the num- ber of territorial males in low-densi- ty populations of the Sooty Grouse, 279-293; Apparent extirpation of the Sooty Grouse from the sky islands of south-central California, 294-308 Bluebird, Western, 2-16 Bluetail, Red-flanked, 228-229 Bombycilla cedrorum, 2-16 Bond, Monica L., Lee, Derek E., Siegel, Rodney B., and Tingley, Morgan W. , Diet and home-range size of California Spotted Owls in a burned forest, 114-126 Booby, Brown, 214 Masked, 214 Masked/Nazca, 214 Botaurus stellaris, 191 Bouzat, Juan L., see Ellison, K. Brachyramphus perdix, 226 Brook, Cara E., Bernstein, David P, and Hadly, Elizabeth A., Human food subsidies and Common Raven occurrence in Yosemite National Park, California, 127-134 brunneonucha, Leucosticte arctoa, 190 Bubo scandiacus, 226 virginianus, 2-16, 309-311 Bucephala clangula, 191 Bunting, Lark, 30 Lazuli, 2-16 Little, 243-245 Painted, 231 Snow, 229 Burnett, L. Joseph, see Tyner, M. Burton, Kenneth M., and Anderson, Leslie Scopes, First record of the Red-bellied Woodpecker in Nevada, 148-150 Bushtit, 2-16 Buteo albonotatus, 20-21 jamaicensis, 2-16 Buteogallus anthracinus, 20 Calamospiza melanocorys, 30 Calcarius ornatus, 29 Calidris ferruginea, 223 Western Birds 44:336-343, 2013 INDEX fuscicollis, 223 himantopus, 21 minuta, 222 ruficollis, 273-278 Callipepla californica, 2-16 Calypte anna, 2-16 costae, 2-16 Campylorhynchus brunneicapillus, 2-16 Caracara, Crested, 45-55, 226 Caracara cheriway, 45-55, 226 Cardellina pusilla, 2-16 Cardinalis sinuatus, 231 Carling, Matthew D., see Maley, J. M. carolinensis, Sitta carolinensis, 90-97 Carpodacus mexicanus, see Haemorhous mexicanus purpureus, see Haemorhous pur- pureus Cartron, Jean-Luc E., Finch, Deborah M., Hawksworth, David L., and Stoleson, Scott H., Nesting ecol- ogy and nest success of the Blue Grosbeak along two rivers in New Mexico, 33-44 castaneiceps, Setophaga petechia, 262-272 Cathartes aura, 2-16 Catharus guttatus, 2-16 ustulatus, 2-16 caurina, Melospiza melodia, 162-170 Cerorhinca monocerata, 178 Chaetura uauxi, 2-16 Chamaea fasciata, 2-16 Charadrius alexandrinus, 191 hiaticula, 222 mongolus, 219 niuosus, 191 wilsonia, 219 Chickadee, Mountain, 2-16 Chiffchaff, Common, 188 chloropus, Gallinula chloropus, 185 Chondestes qrammacus, 2-16, 135-140 Circus cyaneus, 2-16, 78-79, 185 Cistothorus palustris, 188 platensis, 28 clangula, Bucephala clangula, 191 Clangula hyemalis, 19 cleonensis , Melospiza melodia, 162-170 Colaptes auratus, 2-16 Collared-Dove, Eurasian, 187 columbarius, Falco columbarius, 309-311 Condor, California, 151-154 Contopus sordidulus, 2-16 virens, 226 Contreras, Alan, and Turner, Craig, A winter record of a Little Bunting from interior Oregon, 243-245 Coragyps atratus, 218, 316-318 Cormorant, Neotropic, 215 Corvus brachyrhynchos, 2-16 corax, 2-16, 127-134 Cowbird, Brown-headed, 2-16 Crane, Common, 218 crecca, Anas crecca, 19 Crow, American, 2-16 Curlew, Long-billed, 196-205 cyaneus, Circus cyaneus, 185 Cyanocitta cristata, 228 Cygnus buccinator, 210 Cynanthus latirostris, 226 Davis, Jeff N., see Le Valley, R. Day, Robert H., Gall, Adrian E., Morgan, Tawna C., Rose, John R., Plissner, Jonathan H., Sanzen- bacher, Peter M., Fenneman, Jamie D., Kuletz, Katherine J., and Watts, Bridget H., Seabirds new to the eastern Chukchi and Beaufort seas, Alaska: Response to a changing climate?, 174-182 decaocto, Streptopelia decaocto, 187 DeCicco, Lucas H., Warnock, Nils, and Johnson, James A., History of the Red-necked Stint breeding in North America, 273-278 DeCicco, Lucas H., see Gibson, D. D. Dendragapus fuliginosus, 279-308 obscurus, 76-77 Dicker man, Robert W., The distribu- tion of Bubo virginianus pinorum north and west to Washington, 309-311; Records of the Black Merlin in New Mexico, with comments on its identification, 312-315 Dove, Eurasian Collared-, 187 Mourning, 2-16 Spotted, 2-16 dresseri, Somateria mollissima, 211 Duck, Falcated, 210 Long-tailed, 19 Eagle, Golden, 2-16 Egretta caerulea, 20 tricolor, 20, 184, 215 337 INDEX Eider, Common, 211 King, 211 Elanus leucurus, 2-16, 20 Ellison, Kevin, Ross, Jeremy D., and Bouzat, Juan L. , Genetic evidence for mixed maternity at a Lark Spar- row nest, 135-140 Emberiza pusilla, 243-245 Empidonax difficilis, 2-16 minimus, 28 wrightii, 2-16 Erratum, 334-335 Euphagus carolinus, 231 Falco columbarius, 312-315 mexicanus, 2-16 rusticolus, 226 sparuerius, 2-16 Falcon, Prairie, 2-16 Fenneman, Jamie D., see Day, R. H. Finch, Asian Rosy-, 190 Black Rosy-, 232 House, 2-16 Purple, 2-16 Finch, Deborah M., see Cartron, J.-L. E. Flicker, Northern, 2-16 Flycatcher, Ash-throated, 2-16, 187 Dusky-capped, 227 Gray, 2-16 Great Crested, 227 Least, 28 Pacific-slope, 2-16 Scissor-tailed, 28 Sulphur-bellied, 28 Fregata magnificens, 214 Frigatebird, Magnificent, 214 Gall, Adrian E., see Day, R. H. Gallinago gallinago, 223 solitaria, 186 Gallinula chloropus, 185 Gallo-Reynoso, Juan Pablo, see Montiel-Herrera, M. Gannet, Northern, 176, 191 Gavia adamsii, 211 stellata, 20 Geococcyx californianus, 2-16 Geothlypis formosus, 29 Philadelphia, 230 tolmiei, 2-16 Gibson, Daniel D., DeCicco, Lucas H., Gill, Robert E. Jr., Heinl, Steven C., Lang, Aaron J., Tobish, Theo- dore G. Jr., and Withrow, Jack J., Third report of the Alaska Checklist Committee, 2008-2012, 183-195 Gill, Robert E. Jr., see Gibson, D. D. Glinski, Richard L., In Memoriam: Gale Monson, 80-81 Gnatcatcher, Blue-gray, 2-16 California, 2-16 Godwit, Bar-tailed, 222 Goldeneye, Common, 191 Golden-Plover, American, 219 Goldfinch, American, 2-16 Lawrence’s, 2-16, 31 Lesser, 2-16 Gomez de Silva, Hector, Book review: Moments of Discovery: Natural History Narratives from Mexico and Central America, 246-247 Goose, Greater White-fronted, 184 Taiga/Tundra Bean-, 207-210 gouldii, Melospiza melodia, 162-170 Grackle, Common, 31, 231 Grosbeak, Black-headed, 2-16 Blue, 33-44 Grouse, Dusky, 76-77 Sooty, 279-308 Grus grus, 219 Gull, Black-tailed, 223 Glaucous, 23 Heermann’s, 21 Iceland, 21, 223-224 Laughing, 186-187 Lesser Black-backed, 22, 225 Little, 223 Slaty-backed, 225 Thayer’s, 21 Gymnogyps californianus, 151-154 Gyr falcon, 226 Hadly, Elizabeth A., see Brook, C. E. Haematopus ostralegus, 185 Haemorhous mexicanus, 2-16 purpureus, 2-16 Harrier, Northern, 2-16, 78-79, 185 Harter, Lauren, Book review: The Young Birder’s Guide to Birds of North America, 82-83 Hartzell, Paula L., see Howard, P. J. Harvey, Sarah C., see Howard, P. J. Hawk, Common Black-, 20 Cooper’s, 2-16 Harris’s, 20, 218 Red-tailed, 2-16 Sharp-shinned, 2-16 Zone-tailed, 20 Hawksworth, David L., see Cartron, 338 INDEX J.-L. E. Heinl, Steven C., see Gibson, D. D. Helmitheros vermivorum, 29, 229 Heron, Black-crowned Night-, 184 Chinese Pond-, 184 Great Blue, 239 Little Blue, 20 Tricolored, 20, 184, 215 Yellow-crowned Night-, 216 Hickey, Catherine M., see Shuford, W. D. himantopus, Himantopus himanto- pus, 186 Himantopus himantopus, 186 Howard, Phillip J., Harvey, Sarah C., and Ritt, Steven M., Rare and unusual species observed on Tern Island, French Frigate Shoals, northwestern Hawaiian Islands, 237-242 Howard, Phillip J., Harvey, Sarah C., Hartzell, Paula L., Leary, Pete, and Benally, Ty J., Common nesting habitats and weights at fledging of Wedge-tailed Shearwaters on Tern Island, Hawaii, 69-75 howardi, Dendragapus fuliginosus, 294-308 Hummingbird, Allen’s, 2-16 Anna’s, 2-16 Black-chinned, 2-16 Broad-billed, 226 Costa’s, 2-16 Ruby-throated, 226 Rufous, 2-16 Hydrocoloeus minutus, 223 Hylocichla mustelina, 229 Ibis, Glossy, 20, 218 Icterus bullockii, 2-16 cucullatus, 2-16 galbula , 31 iliacus, Turdus iliacus, 190 inornata, Tringa semipalmata, 186 Jaeger, Parasitic, 25 japonica, Gallinago solitaria, 186 Jay, Blue, 228 Western Scrub-, 2-16 Johnson, James A., see DeCicco, L. H. Johnson, Oscar, Pyle, Peter, and Tietz, Jim, The subspecies of the Song Sparrow on Southeast Farallon Island and in central California, 162-170 Junco, Dark-eyed, 2-16 Junco hyemalis, 2-16 kenaiensis, Melospiza melodia, 162-170 Kestrel, American, 2-16 Kienholz, Aleen M., see Kienholz, R. J. Kienholz, Ronald J., and Kienholz, Aleen M., An apparent long- distance flight by a Dusky Grouse in Montana, 76-77 Kingbird, Cassin’s, 2-16 Thick-billed, 227 Western, 2-16 Kinglet, Ruby-crowned, 2-16 Kite, White-tailed, 2-16, 20 Kittiwake, Black-legged, 23 Krueper, Dave, Book review: The Travails of Two Woodpeckers: Ivory-Bills & Imperials, 247-249 Kuletz, Katherine J., see Day, R. H. Lang, Aaron J., see Gibson, D. D. Langham, Gary M., see Shuford, W. D. Lanius ludovicianus, 2-16 Larus crassirostris, 223 fuscus, 22, 225 glaucoides, 21, 223-224 heermanni, 21 hyperboreus, 23 schistisagus, 225 thayeri, 21 Laterallus jamaicensis, 21 Leary, Pete, see Howard, P. J. Lee, Derek E., see Bond, M. L. Leucophaeus atricilla, 186-187 Leucosticte arctoa, 190 atrata, 232 LeValley, Ron, and Davis, Jeff N., Carotenism in the Hairy Wood- pecker, 85-87 Limosa lapponica, 222 Lincer, Jeff L., Book reviews: Hawks in Flight and The Crossley Guide: Raptors, 319-321 Longspur, Chestnut-collared, 29 Loon, Red-throated, 20 Yellow-billed, 211 Luscinia sibilans, 189-190 Magpie, Yellow-billed, 143-147 Maley, James M., Saucier, Jacob R., and Carling, Matthew D., Refuta- tion of Wyoming nesting record of the Pacific Wren, 141-142 339 INDEX Martin, Purple, 28 McNew, Sabrina M., see Baumann, M. J. Meadowlark, Western, 2-16 Meese, Robert J., Chronic low repro- ductive success of the colonial Tricolored Blackbird from 2006 to 2011, 98-113 Melanerpes carolinus, 148-150, 226 erythrocephalus, 25 formiciuorus , 2-16, 25, 28 Melanitta americana, 19 Melcer, Ron, Jr., see Nightingale, A. melodia , Melospiza melodia, 162-170 Melospiza georgiana, 31 lincolnii , 2-16 melodia, 2-16, 162-170 Melozone crissalis , 2-16 Mergellus albellus , 211 Merlin, 312-315 merrilli, Melospiza melodia, 162-170 Meyers, Martin, 2011 Nevada Bird Records Committee Report, 17-32 Micrathene whitneyi, 226 Miller, Robert A., Paprocki, Neil, and Urban, Elizabeth H., Melanistic adult male Northern Harrier winter- ing in Idaho, 78-79 Mimus polyglottos, 2-16 Mockingbird, Northern, 2-16 Molothrus ater, 2-16 Montiel-Herrera, Marcelino, and Gallo- Reynoso, Juan Pablo, Behaviors of nestling and juvenile Black Vultures in northwestern Mexico, 316-318 Moorhen, Common, 185 Morgan, Tawna C., see Day, R. H. Moriarty, David J., A twenty-year inves- tigation of the effects of fire on a coastal sage scrub bird community, 2-16 morphna, Melospiza melodia, 162-170 Morus bassanus, 176, 191 Murre, Common, 250-259, 334-335 Thick-billed, 226 Murrelet, Ancient, 176-178 Long-billed, 226 Myiarchus cinerascens, 2-16, 187 crinitus, 227 tuberculifer, 227 Myiodynastes luteiventris, 28 Nelson, Kristie N., and Pyle, Peter, Dis- tribution and movement patterns of individual Crested Caracaras in California, 45-55 Nelson, Kristie N., Rottenborn, Ste- phen C., and Terrill, Scott B., The 37 th annual report of the California Bird Records Committee: 2011 records, 206-236 Night-Heron, Black-crowned, 184 Yellow-crowned, 216 Nightingale, Ann, and Melcer, Ron, Jr., Conspecific nest aggression of the Pacific Wren on Vancouver Island, British Columbia, 56-64 Numenius americanus, 196-205 Nuthatch, White-breasted, 90-97 Nyctanassa uiolacea , 216 nycticorax, Nycticorax nycticorax, 184 Nycticorax nycticorax, 184 occidentalis, Strix occidentalis, 114-126 Oceanodroma monorhis, 191 Oenanthe oenanthe, 229 Onychoprion fuscatus, 225 Oporornis agilis, 230 Oreothlypis celata, 2-16 ruficapilla, 190 Oriole, Baltimore, 31 Bullock’s, 2-16 Hooded, 2-16 osculans, Haematopus ostralegus, 185 Owl, Elf, 226 Great Gray, 25 Great Horned, 2-16, 309-311 Short-eared, 237 Snowy, 226 Spotted, 114-126 Oystercatcher, Eurasian, 185 Page, Gary W., see Shuford, W. D. Pandolfino, Edward R., WFO Presi- dent’s Message: State of the Organization, 84; Lack of recovery of the Yellow-billed Magpie from the West Nile Virus in California’s Central Valley, 143-147; WFO President’s Message: The Future of Field Ornithology, 161 Paprocki, Neil, see Miller, R. A. Parabuteo unicinctus, 20, 218 Passer domesticus, 2-16 Passerculus sandwichensis, 2-16 Passerella iliaca, 2-16 Passerina amoena, 2-16 caerulea, 33-44 340 INDEX ciris, 231 Patagioenas fasciata, 2-16 Petrel, Galapagos/Hawaiian, 214 Great-winged, 213 Hawaiian, 213 Providence, 184 Solander’s, see Providence Swinhoe’s Storm-, 191 White-chinned, 214 Petrochelidon pyrrhonota , 2-16 Pewee, Eastern Wood-, 226 Western Wood-, 2-16 Phaethon lepturus, 238 Phainopepla, 2-16 Phainopepla nitens, 2-16 Phalacrocorax brasilianus, 215 Phalaenoptilus nuttallii, 2-16 Phalarope, Red, 21 Phalaropus fulicarius, 21 Pheucticus melanocephalus, 2-16 Phoebastria albatrus, 175-176, 211 immutabilis, 322-333 nigripes, 322-333 Phoebe, Black, 2-16 Say’s, 2-16 Phylloscopus collybita, 188 Pica nuttalli, 143-147 Picoides nuttallii, 2-16 villosus, 85-87 Pigeon, Band-tailed, 2-16 pinorum, Bubo uirginianus, 309-311 Pipilo maculatus, 2-16 Piranga ludouiciana, 2-16 olivacea, 31 Plectrophenax nivalis, 229 Plegadis falcinellus, 20, 218 Plissner, Jonathan H., see Day, R. H. Plover, American Golden-, 219 Common Ringed, 222 Kentish, 191 Lesser Sand-, 219 Snowy, 191 Wilson’s, 219 Pluvialis dominica , 219 Poecile gambeli, 2-16 Polioptila caerulea, 2-16 californica, 2-16 Pond-Heron, Chinese, 184 Poorwill, Common, 2-16 Procellaria aequinoctialis, 214 Progne subis, 28 Psaltriparus minimus, 2-16 Pterodroma macroptera, 213 phaeopygia/sandwichensis, 214 sandwichensis, 213 solandri, 184 Puffmus gravis, 214 pad ficus, 69-lb pugetensis, Zonotrichia leucophrys, 190 Pyle, Peter, Dark-faced Common Murres of central California in fall and winter, 250-259; see Johnson, O.; see Nelson, K. N. Pyrrhuloxia, 231 Quail, California, 2-16 Quiscalus quiscula, 31, 231 Rail, Black, 21 Raven, Common, 2-16, 127-134 Redpoll, Common, 232 Redwing, 190 Reed Warbler, Blyth’s, 191 Regulus calendula, 2-16 ridgwayi, Oreothlypis ruficapilla, 190 Rissa tridactyla, 23 Ritt, Steven M., see Howard, P. J. Roadrunner, Greater, 2-16 Robin, American, 2-16 Rufous-backed, 229 Rufous-tailed, 189-190 Rose, John R., see Day, R. H. Ross, Jeremy D., see Ellison, K. Rosy-Finch, Asian, 190 Black, 232 Rottenborn, Stephen C., see Nelson, K. N. ruficollis, Egretta tricolor, 184 rufina, Melospiza melodia, 162-170 Rutt, Cameron, Hybridization in Black- footed and Laysan albatrosses, 322-333 Salpinctes obsoletus, 2-16 Sandpiper, Curlew, 223 Stilt, 21 Upland, 21, 222 White-rumped, 223 Sand-Plover, Lesser, 218 Sanzenbacher, Peter M., see Day, R. H. Sapsucker, Yellow-bellied, 28 Saucier, Jacob, R., see Maley, J. M. Sayornis nigricans, 2-16 saya, 2-16 Schweizer, Cheryl L., and Whitmore, Robert C., Movements of the Man- grove Warbler in Baja California Sur, 262-272 Scoter, Black, 19 Scrub-Jay, Western, 2-16 341 INDEX Selasphorus rufus, 2-16 sasin, 2-16 Setophaga caerulescens, 29-30 coronata, 2-16 discolor, 30 dominica, 30, 230 graciae, 231 magnolia, 29 palmarum, 30 pensylvanica, 29 petechia, 2-16, 262-272 pinus, 230 tigrina, 230 townsendi, 2-16 uirens, 30 Shearwater, Great, 214 Wedge-tailed, 69-75 Shrike, Loggerhead, 2-16 Shuford, W. David, Page, Gary W., Langham, Gary M., and Hickey, Catherine M., The importance of agriculture to Long-billed Curlews in California’s Central Valley in fall, 196-205 Sialia mexicana, 2-16 Siegel, Rodney B., see Bond, M. L. sierrae, Dendragapus fuliginosus, 279-293 Sitta carolinensis, 90-97 Smew, 211 Snipe, Common, 223 Solitary, 186 solitarius, Vireo solitarius, 187 Somateria mollissima, 211 spectabilis, 211 Sparrow, Chipping, 2-16 Clay-colored, 30 Fox, 2-16 Golden-crowned, 2-16 Grasshopper, 30 House, 2-16 Lark, 2-16, 135-140 Le Conte’s, 231 Lincoln’s, 2-16 Rufous-crowned, 2-16 Sage, 2-16 Savannah, 2-16 Song, 2-16, 162-170 Swamp, 31 White-crowned, 2-16, 190 Sphyrapicus varius, 28 Spinus lawrencei, 2-16, 31 psaltria, 2-16 tristis, 2-16 Spizella pallida, 30 passerina, 2-16 sponsa, Anser albifrons, 184 Stake, Mike M., see Tyner, M. Starling, European, 2-16 Stelgidopteryx serripennis, 2-16 Stercorarius parasiticus, 25 Sterna paradisaea, 23, 25 Sternula antillarum, 23, 237 Stilt, Black-winged, 186 Stint, Little, 222 Red-necked, 273-278 Stoleson, Scott H., see Cartron, J.-L. E. Storm-Petrel, Swinhoe’s, 191 Streptopelia chinensis, 2-16 decaocto, 187 Strix nebulosa, 25 occidentalis, 114-126 Sturnella neglecta, 2-16 Sturnus vulgaris , 2-16 suckleyi, Falco columbarius, 309-311 Sula dactylatra, 214 dactylatra/granti, 214 leucogaster, 214 Surfbird, 65-68 Swallow, Cliff, 2-16 Northern Rough-winged, 2-16 Violet-green, 2-16 Swan, Trumpeter, 210 Swift, Fork-tailed, 239 Vaux’s, 2-16 White-throated, 2-16 Synthliboramphus antiquus, 176-178 Tachycineta thalassina, 2-16 Tanager, Scarlet, 31 Western, 2-16 Tarsiger cyanurus, 228-229 Teal, Green-winged, 19 Tern, Arctic, 23, 25 Least, 23, 227 Sooty, 225 Terrill, Scott B., see Nelson, K. N. Thrasher, Brown, 29 California, 2-16 Curve-billed, 229 Thrush, Hermit, 2-16 Swainson’s, 2-16 Wood, 229 Thryomanes bewickii, 2-16 Tietz, Jim, see Johnson, O. Tingley, Morgan W., see Bond, M. L. Titmouse, Oak, 2-16 Tobish, Theodore G. Jr., see Gibson, D. D. Towhee, California, 2-16 342 INDEX Spotted, 2-16 Toxostoma curvirostre, 229 redivivum, 2-16 rufum, 29 Tringa semipalmata, 186 tristis, Phylloscopus collybita, 188 Troglodytes aedon, 2-16 hiemalis, 28, 228 pacificus, 56-64, 141-142 Tropicbird, White-tailed, 238 Turner, Craig, see Contreras, A. Turdus iliacus, 190 migratorius, 2-16 rufopalliatus , 229 Tyner, Mike, Burnett, L. Joseph, and Stake, Mike M., California Condor foraging on a live California Sea Lion pup, 151-154 Tyrannus crassirostris, 227 forficatus, 28 uerticalis, 2-16 vociferans, 2-16 Urban, Elizabeth H., see Miller, R. A. Uria aalge, 250-259, 334-335 lomuia, 226 VanderWerf, Eric, A., First record of a Surfbird in the Hawaiian Islands, 65-68 Vermivora chrysoptera, 229 Vireo, Blue-headed, 187, 227-228 Cassin’s, 2-16 Philadelphia, 28 Plumbeous, 2-16 Warbling, 2-16 Yellow-green, 228 Vireo cassinii, 2-16 flavoviridis, 228 giluus, 2-16 philadelphicus, 28 plumbeus , 2-16 solitarius, 187, 227-228 Vulture, Black, 218, 316-318 Turkey, 2-16 Warbler, Black-throated Blue, 29 Black-throated Green, 30 Blyth’s Reed, 191 Cape May, 230 Chestnut-sided, 29 Connecticut, 230 Golden-winged, 229 Grace’s, 231 Kentucky, 29 MacGillivray’s, 2-16 Magnolia, 29 Mangrove, 262-272 Mourning, 230 Nashville, 190 Orange-crowned, 2-16 Palm, 30 Pine, 230 Prairie, 30 Townsend’s, 2-16 Wilson’s, 2-16 Worm-eating, 29, 229 Yellow, 2-16, 262-272 Yellow-rumped, 2-16 Yellow-throated, 30, 230 Warnock, Nils, see DeCicco, L. H. Watts, Bridget H., see Day, R. H. Waxwing, Cedar, 2-16 Wheatear, Northern, 229 White-eye, Japanese, 240 Whitmore, Robert C., see Schweizer, C. L. Wigeon, Eurasian, 19 Willet, 186 Winter, Jon, In Memoriam: Richard W. Stallcup, 155-157 Withrow, Jack J., see Gibson, D. D. Witt, Christopher C., see Baumann, M. J. Wolf, Cole J., Book review: Twelve Hundred Miles by Horse and Burro: J. Stokely Ligon and New Mexico’s First Breeding Bird Survey, 159-160 Woodpecker, Acorn, 2-16, 25, 28 Hairy, 85-87 Nuttall’s, 2-16 Red-bellied, 148-150, 226 Red-headed, 25 Wood-Pewee, Eastern, 226 Western, 2-16 Wren, Bewick’s, 2-16 Cactus, 2-16 House, 2-16 Marsh, 188 Pacific, 56-64, 141-142 Rock, 2-16 Sedge, 28 Winter, 28, 228 Wrentit, 2-16 Zenaida macroura, 2-16 Zonotrichia atricapilla, 2-16 leucophrys, 2-16, 190 Zosterops japonicus, 240 343 World Wide Web site: WESTERN BIRDS www. westernfieldornithologists . org Quarterly Journal of Western Field Ornithologists President: Edward R. Pandolfino, 1328 49th St., Sacramento, CA 98519; ERPfromCA@aol . com Vice-President: David E. Quady, 39 The Crescent, Berkeley, CA 94708; davequady@att. net Past-President: W. David Shuford, P O. Box 69, Bolinas, CA 94924; dshuf ord@prbo . org Treasurer/Membership Secretary: Robbie Fischer, 1359 Solano Dr., Pacifica, CA 94044; robbie22@pacbell.net Recording Secretary: Liga Auzins, 12842 Safford East, Garden Grove, CA 92840; llauzins@yahoo . com Directors: Kenneth P. Able, Thomas A. Blackman, Jon L. Dunn, Kimball L. Garrett, Daniel D. Gibson, Robert E. Gill, Ed Harper, Kurt Leuschner, Joyce Meyer, Frances Oliver, Dan Singer, Catherine P. Waters Editor: Philip Unitt, San Diego Natural History Museum, P. O. Box 121390, San Diego, CA 92112-1390; birds@sdnhm.org Assistant Editor: Daniel D. Gibson, 3705 Quartz Rd., Ester, AK 99725; avesalaska@gmail . com Associate Editors: Kenneth P. Able, Doug Faulkner, Thomas Gardali, Daniel D. Gibson, Robert E. Gill, Paul E. Lehman, Ronald R. LeValley, Dan Reinking Graphics Manager: Virginia P. Johnson, 4637 Del Mar Ave., San Diego, CA 92107; gingerj 5@juno . com Photo Editor: Peter LaTourrette, 1019 Loma Prieta Ct., Los Altos, CA 94024; petelat 1 @stanford . edu Featured Photo: John Sterling, 26 Palm Ave., Woodland, CA 95695; jsterling@ wavecable . com WFO Website: Joseph Morlan, 1359 Solano Dr., Pacifica, CA 94044; webmaster@ westernfieldornithologists . org Book Reviews: Lauren Harter, 2841 McCulloch Blvd N. #1, Lake Havasu City, AZ 86403; lbharter@gmail.com Membership dues, for individuals and institutions, including subscription to Western Birds: Patron, $1000.00; Life, $600 (payable in four equal annual installments); Supporting, $75 annually; Contributing, $50 annually; Family $40; Regular U.S. $35 for one year, $60 for two years, $85 for three years. Dues and contributions are tax-deductible to the extent allowed by law. Send membership dues, changes of address, correspondence regarding missing issues, and or- ders for back issues and special publications to the Treasurer. Make checks payable to Western Field Ornithologists. BACK ISSUES OF WESTERN BIRDS WITHIN U.S. $40 PER VOLUME, $10 FOR SINGLE ISSUES, INCLUDING SHIPPING AND HANDLING. OUTSIDE THE U.S. $55 PER VOLUME, $15 FOR SINGLE ISSUES, INCLUDING SHIPPING AND HANDLING. The California Bird Records Committee of Western Field Ornithologists revised its 10-column Field List of California Birds in July 2009. The list covers 641 species, plus 6 species on the supplemental list. Please send orders to WFO, c/o Robbie Fischer, Treasurer, 1359 Solano Drive, Pacifica, CA 94044. Price for 9 or fewer, $2.75 each, for 10 or more, $2.50 each, which includes tax and shipping. Order online at http://checklist.westernfieldornithologists.org. Published 31 December 2013 ISSN 0045-3897 The possible F 2 hybrid albatross on the left danced with a (likely female) Black ■fooled Albatross, dancing with Ihe vigor, speed, and dual whig lift of a Blaek’I'ootcd and unlike the Laysan. This hybrid, photographed on Laysan island 15 E>ecemher 2 (J 1 l, five years after being handed, was unsuccessful in attracting a male in 20 IE -20 1 J. Photo by Cameron Ritlf A possible F 2 Laysan Albatross baekeross, photographed on Laysan Island IH December 20 n. Such birds may be impossible to distinguish from i I l.j dark extreme of the Laysan Albatross, although the latter may represent H'2 backe losses or birds with small proportions of B I auk- footed genes. This bird showed extra gray in areas predominantly white in the Laysan Albatross, including extensively dark bum crisis. Photo tov Cameron Ri tit