Skip to main content

Full text of "Western birds"

See other formats



Vol. 32, No. 1, 2001 



Volume 32, Number 1, 2001 

Historical Changes in the Abundance and Distribution of the American 
Avocet at the Northern Limit of Its Winter Range Mark A. Colwell, 


Tamar Danufsky, Ryan L. Mathis, and Stanley W. Harris 1 

Report of the California Bird Records Committee: 1998 Records 

Richard A. Erickson and Robert A. Hamilton 13 

Arizona Bird Committee Report: 1996-1999 Records 

Gary H. Rosenberg 50 

Recolonization of the Flicker and Other Notes from Isla Guadalupe, 
Mexico Paul R. Sweet, George F. Barrowclough, John T, Klicka, 
Liliana Montahez-Godoy, and Patricia Escalante- Pliego 71 


NOTES 

Orange Bishops Breeding in Phoenix, Arizona Thomas A. Gatz . . 81 

Breeding-Season Home Ranges of Spotted Owls in the San 
Bernardino Mountains, California Guthrie S. Zimmerman, 


William S. LaHaye, and R. J. Gutierrez 83 

First Report of the Gray Heron in the United States 

Kenneth M. Burton and Sean D. Smith 88 

Cloacal Inspection or Pecking in Allen's Hummingbird 

Janet L. Leonard 91 

Book Reviews Steve N. G. Howell, Robert A, Hamilton 93 

Featured Photo Steve N. G. Howell 97 


Cover photo by © Brian E. Small of Los Angeles, California: Nutting’s 
Flycatcher { Myiarchus nuttingi), Mason Regional Park, Irvine, California, 
January, 2001. If accepted, this will be the first record of Nutting’s 
Flycatcher for the State of California. 

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 include 
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 the Rocky Mountain and Pacific states 
and provinces, including Alaska and Hawaii, western Texas, northwestern Mexico, and the 
northeastern Pacific Ocean. 

Send manuscripts to Kathy Molina, Section of Ornithology, Natural 
History Museum of Los Angeles County, 900 Exposition Blvd., Los Angeles, CA 
90007. For matter of style consult the Suggestions to Contributors to Western Birds 
(8 pages available at no cost from the editor) and the Council of Biology Editors Style 
Manual (available for $24 from the Council of Biology Editors, Inc., 9650 Rockville 
Pike, Bethesda, MD 20814). 

Reprints can be ordered at author’s expense from the Editor when proof is returned or 
earlier. 

Good photographs of rare and unusual birds, unaccompanied by an article but 
with caption including species, date, locality and other pertinent 
information, are wanted for publication in Western Birds. Submit photos and 
captions to Photo Editor. Also needed are black and white pen and 
ink drawings of western birds. Please send these, with captions, to 
Graphics Manager. 


WESTERN BIRDS 


Volume 32, Number 1, 2001 



HISTORICAL CHANGES IN THE ABUNDANCE AND 
DISTRIBUTION OF THE AMERICAN AVOCET AT 
THE NORTHERN LIMIT OF ITS WINTER RANGE 

MARK A. COLWELL, TAMAR DANUFSKY, RYAN L. MATHIS, and STANLEY 
W. HARRIS, Department of Wildlife, Humboldt State University, Areata, California 
95521 

ABSTRACT: Humboldt Bay, California, is the northern limit of the winter 
distribution of the American Avocet (Recurvirostra americana ) on the Pacific coast. 
After the first record in 1935, avocets were uncommon (17 observations) until the 
early 1960s, when a wintering population of <100 birds became established in North 
(Areata) Bay. Numbers increased to approximately 1000 by the early 1990s but have 
since declined to approximately 500. From 1968 to 1985, avocets consistently used 
intertidal habitats and oxidation ponds in the northeast quarter of Areata Bay. 
Beginning in the mid-1990s they expanded their use of Areata Bay and into South 
Bay. During February and March 1998 and February 2000, up to 32 occasionally fed 
in flooded pastures adjacent to the bay; only two avocets had been observed there in 
the previous 40 years. At low tide, avocets aggregated in intertidal habitats of Areata 
Bay and South Bay. We hypothesize that they increased because of a rangewide 
population increase and that in Humboldt Bay they concentrate where small particle 
size of sediments makes for better feeding habitat. Altered habitat quality, especially 
during wet years (late 1990s), may have changed avocet distribution in Humboldt Bay. 

Understanding changes in the abundance and distribution of a species is 
a central theme of animal ecology, and research at the limits of an 
organism’s range can provide valuable insight into factors limiting a species’ 
distribution, especially when conducted over a long period. The American 
Avocet ( Recurvirostra americana) winters in intertidal and freshwater 
habitats along the Atlantic, Gulf, and Pacific coasts of North America 
(Robinson et al. 1997). Along the Pacific coast, Humboldt Bay, California 
represents the northern extreme of the species’ winter range (Evans and 
Harris 1994, Robinson et al. 1997); there are no winter records from the 
Pacific Northwest (Paulson 1993). Evans and Harris (1994) showed that 
avocets were a recent addition to the wintering avifauna of the northern 
California coast, first recorded in 1935. Beginning in the 1960s, a growing 
population of wintering avocets concentrated in the northeast quarter of 


Western Birds 32:1-12, 2001 


1 


CHANGES IN THE ABUNDANCE AND DISTRIBUTION OF THE AVOCET 


Humboldt Bay {Evans and Harris 1994). Since 1989, we have collected 
information on avocet abundance and distribution at Humboldt Bay and 
have noticed changes in the patterns reported by Evans and Harris (1994). 
Here, we summarize changes in avocet distribution and abundance on 
Humboldt Bay and discuss possible reasons for these changes. 

STUDY AREA 

Humboldt Bay (40° 46' N, 124° 14' W) is the largest bay between San 
Francisco Bay, California, 370 km south, and Coos Bay, Oregon, 335 km 
north (Barnhart et al. 1992). Costa (1982) characterized the bay as a tidal 
coastal lagoon with limited freshwater input. The bay consists of three 
sections, each of which is at the end of a different tributary. The largest 
section, Areata Bay, receives fresh water from Jacoby and Freshwater creeks. 
Broad intertidal habitats of Areata Bay consist of sediments varying from 
clayey silts of intertidal flats to sandy substrates lining main channels. Much of 
the lower intertidal reaches of the center of Areata Bay supports eelgrass 
[Zostera marina), but the extent of this habitat has been reduced by oyster 
culture (Barnhart et al. 1992). Oxidation ponds of the Areata marsh and 
sewage-treatment facility lie on the northeast shore of Areata Bay. A narrow 
shipping channel connects Areata Bay to Entrance Bay. The channel and 
Entrance Bay are dredged to provide access to ships, hence most intertidal 
habitats have steep banks and coarse substrates. The Elk River empties into 
the main channel and provides some intertidal flats, which are partially 
covered by eelgrass. South Bay consists of extensive intertidal habitats and 
substrates varying from sand to clayey silt. Large relatively pristine eelgrass 
beds cover much of the lower intertidal reaches of South Bay. Salmon Creek 
enters the southeast corner of South Bay through Humboldt Bay National 
Wildlife Refuge (HBNWR). Since 1850, people have converted much of the 
bay’s salt marsh and intertidal habitats to agricultural lands, especially pasture. 
Barnhart et al. (1992) described Humboldt Bay in detail. 

METHODS 

Our examination of avocet population size and distribution relies on 
several independent sources of data collected over the last 50 years. The 
quality of this information varies with the different objectives of researchers 
and unequal sampling effort. 

Field notes. Harris has kept regular field notes on bird distribution and 
abundance around Humboldt Bay since 1959. From these notes, we 
collated maximum annual estimates and locations of avocets. 

Christmas Bird Counts. Two CBC circles cover virtually all intertidal 
habitats of Humboldt Bay and abut at the harbor entrance. The Centerville 
CBC, conducted for 41 years, covers all of South Bay. Harris has coordi- 
nated this count and collated data since 1973. To the north, the Areata 
CBC, conducted since 1984, covers all of Areata Bay and the main channel 
leading to the bay entrance. Because we were able to identify most locations 
at which observers recorded avocets, CBC data offer an important historical 
perspective on the species’ use of the bay. 


2 


CHANGES IN THE ABUNDANCE AND DISTRIBUTION OF THE AVOCET 


Graduate theses. Two graduate theses provided data on avocet distribu- 
tions. In 1968 and 1969, Gerstenberg (1972) surveyed all shorebirds at nine 
locations on Humboldt Bay. Evans (1988) summarized avocet distribution 
baywide during three winters (1982-83 through 1984-85); he made 76 
surveys at 1 1 sites within Areata Bay and 42 surveys from 1 1 locations in 
South Bay. Since each used different survey methods and collated data 
differently, we summarized their information as follows. Using Gerstenberg’s 
(1972) maximum winter count of avocets at each of his survey locations, we 
summed all avocet observations at the nine locations and calculated the 
proportion at each site (yielding a proportional abundance). Evans (1988) 
summarized his results as the average monthly high- and low-tide count at 
each location. We used the maximum average winter (October-February) 
count for each location to represent use of a site. Next, we calculated the 
proportion of these summed maximum counts for each location. 

Baywide shorebird surveys. From 1989 to 1994 and again in 1998, 
Colwell coordinated multiple observers who surveyed Humboldt Bay simul- 
taneously to estimate seasonal use by shorebirds (Colwell 1994). Surveys 
were shore-based counts of intertidal habitats of the bay, adjacent pastures 
and freshwater sloughs, and ocean beaches fronting the bay. The survey 
protocol entailed four sequential half-hour scans (Altmann 1974) during 
which observers estimated the number of shorebirds using an area. The start 
of each half-hour scan was synchronized among observers to reduce the 
likelihood that birds moving between areas would be counted by more than 
one observer. We surveyed shorebirds during rising tides so that foraging 
birds were pushed by the advancing tide toward shore, improving observers’ 
estimates of abundance (Colwell and Cooper 1993). Observers conducted 
surveys from prominent observation points around the bay to maximize 
observation area and minimize overlap between adjacent observers (further 
details in Colwell and Cooper 1993, Colwell 1994). 

During the five years of surveys, sampling effort varied greatly with weather 
and the number of observers (Colwell 1994); many locations were not 
surveyed every year, although some sites received regular coverage, often by 
the same observers. Consequently, we selected the single maximum count of 
avocets at each site from 10 winter survey dates (November and February, 
1989-1994). We summed these maximum counts and calculated the propor- 
tional abundance of observations at each location. In this form, data were 
corrected for possible changes in avocet population size at Humboldt Bay. We 
summarized 1998 survey data separately by the same method. 

Winter 1 998-1 999 surveys. At 19 locations around the bay, we used a 
protocol similar to that of the shorebird surveys to esitmate winter (Novem- 
ber 1998-January 1999) shorebird numbers. Unlike baywide surveys (see 
above), which took place during flood tides, we did these surveys during ebb 
tides, when intertidal flats were exposed. Areas surveyed varied from 3.5 to 
52.1 ha; we based their boundaries on edges of channels, salt marsh, and 
other prominent habitat features. We summarized these data as the maxi- 
mum count of four surveys at each site and present data as the proportion at 
each location of total avocet observations (Danufsky 2000). 

Low-tide avocet mapping. At five tides <0.33 m (January, February, and 
October 1999; January and February 2000), we mapped avocets on high- 


3 


CHANGES IN THE ABUNDANCE AND DISTRIBUTION OF THE AVOCET 


resolution images of the bay, overlaid by a UTM-based grid system. Observ- 
ers mapped avocets from the same shoreline observation points used in the 
baywide surveys (see above) and from boats moving through the bay’s main 
channels. We marked the location of avocets on images by means of 
topographic features like channels, islands, and saltmarsh points and we 
recorded the number of avocets observed within each 500-m grid. We 
standardized data by calculating the number of avocets in a grid cell within 
five observation periods (January and February 1999 and 2000, October 
1999). For cells including intertidal habitats exposed at low tide, we 
calculated the average number of avocets per grid cell and variance. We 
compared observed avocet distributions to a random model by using the 
index of dispersion (/, variance-to-mean ratio; Krebs 1989). If avocets were 
randomly distributed among grid cells, this ratio equaled one; if they were 
aggregated it was greater than one; if they were evenly distributed it was less 
than one. We compared each survey’s / value to a random distribution by 
means of a two-tailed t test. 

RESULTS 

Population estimates. Over 40 years, avocet numbers have varied signifi- 
cantly (Kruskal-Wallis test, % 2 = 36.4, P = 0.00002) among the five-year 
intervals (Figure 1). Beginning in the 1950s, a small population established 
itself at Humboldt Bay. Numbers nearly doubled over the next two decades. 


1600 n 


1400 


1200 

Ifl 

15 

© 1000 

© 800 - 
© 

S 600 
© 
z ; 

400 


200 


0 1 — 
1955 


S 
* 

1960 1965 


1970 





1975 1980 1985 1990 1995 


Figure 1 . Estimates (maximum counts) of American Avocets wintering at Humboldt 
Bay, California, from 1955 to 1999, based onfield notes (1955-1967; 1970-1981; 
1995-1997), graduate theses (1968-1969; 1982-1985), shorebird surveys (1989- 
1994; 1998) and baywide mapping of avocet distributions (1999-2000). 


4 


CHANGES IN THE ABUNDANCE AND DISTRIBUTION OF THE AVOCET 


Tabic 1 Estimates of the Number and Spatial Distribution of Nonbreeding 
American Avocets at Humboldt Bay, California, Based on Bay-wide Map- 


ping During Diurnal Low Tides 



30-31 
Jan 1999 

27-28 
Feb 1999 

23 Oct 
1999 

29-30 
Jan 2000 

13 Feb 
2000 

Number of avocets 

543 

343 

182 

475 

479 

Mean (per grid celi) 

1.65 

1.04 

0.55 

1.44 

1.45 

Variance (per grid cell) 

62.33 

21.97 

17.43 

90.89 

70.97 

Dispersion index 

37.9 

21.1 

31.6 

63.1 

48.9 

t value 

758.1 

261.6 

217.2 

1118.1 

869.0 

P 

<0.01 

<0.01 

<0.01 

<0.01 

<0.01 

Observed low tide (m) 

-0.31,-0.51 

-0.02,-0.10 

0.10 

0.55, 0.49 

0.63 

Time of low tide 

1712, 1748 

1618, 1724 

1642 

1300, 1424 

1212 


By the early 1990s, estimates ranged from 850 to 1200 birds, after which 
numbers declined during the late 1990s. During January/February low-tide 
mapping in 1998-1999 and 1999-2000, we estimated an average of 460 
± 84 (range 343-543) avocets on Humboldt Bay (Table 1). 

Historical distributions in Humboldt Bay. Avocets occurred on 20% of 
Centerville CBCs over the count’s 41-year history; 98% of the 372 avocets 
occurred between 1994 and 1999 (Table 2). Except for records of no more 
than four individuals in 1969, 1975, and 1978, avocets first appeared on 
the Centerville CBC in 1994, when 39 were in the Eel River lowlands. In 
1995 none were reported; in 1996, when inclement weather compromised 
survey efforts, eight were at HBNWR adjacent to South Bay. From 1997 to 
1999, 98 to 99% of avocets occurred on HBNWR; no more than four were 
reported in the Eel River lowlands. 

Avocets occurred on all 16 Areata CBCs (Table 2). Total numbers varied 
greatly (mean 763 ± 379); high numbers reported from 1993 to 1995 almost 


Table 2 Averages and Ranges of American Avocet Numbers 
on the Centerville and Areata Christmas Bird Counts by Five- 
Year Periods 


5- Year Interval 

Centerville CBC 

Areata CBC 

1955-1959 

0° 



1960-1964 

0 

— 

1965-1969 

0.2 (0-1) 

— 

1970-1974 

0 

— 

1975-1979 

1 (0-4) 

— 

1980-1984 

0 

839° 

1985-1989 

0 

590 (500-800) 

1990-1994 

8 (0-39) 

825 (564-1806) 

1995-1999 

65 (0-165) 

665 (382-1176) 


“One estimate only; CBC began the last year of this interval. 


5 


CHANGES IN THE ABUNDANCE AND DISTRIBUTION OF THE AVOCET 



Pacific 

Ocean 


Mudflat 


Areata 


68/6982/8389/94 98 98/99 


10 12 Kilometers 


Figure 2. A 30-year history of changes in the proportional abundance of American 
Avocets around Areata Bay, California, based on field notes, graduate theses, 
shorebird surveys, and recent research. Data are from shore-based surveys and do not 
include the center of Areata Bay. 


6 


CHANGES IN THE ABUNDANCE AND DISTRIBUTION OF THE AVOCET 


certainly represent summed counts from different observers. A decline in 
avocet observations in Areata Bay paralleled the increase in South Bay. 

Over the past three decades, the spatial distribution of avocets in intertidal 
habitats of Humboldt Bay has changed (Figure 2). During the late 1960s 
(Gerstenberg 1972) and mid-1980s (Evans 1988), avocets used the north 
and east parts of Areata Bay nearly exclusively. During the early 1990s, 
baywide surveys revealed a similar pattern, but a few avocets (6.1%) 
occurred on the south and west shores of Areata Bay; only six occurred 
outside Areata Bay in the southwest corner of South Bay (1992 and 1993). 
During January and February 1998, avocet use of the north and east sectors 
of Areata Bay declined, whereas use of south and west areas increased 
sharply. From November 1998 to January 1999, avocets were more evenly 
distributed around Areata Bay than they were 10-30 years earlier. On 
numerous occasions during January and February 1998, S. Manion and 
Mathis observed 5-32 avocets feeding at low tide in flooded pastures north 
and west of Areata Bay. On 16 February 2000, Colwell observed 15 avocets 
feeding in flooded fields of HBNWR. 

Avocet densities during winter 1999 and 2000. From the total number 
of birds in each grid cell, we determined that avocets were concentrated 
(Table 1) along channels and on intertidal flats of Areata Bay and in the 
southeast corner of South Bay (Figure 3). 

DISCUSSION 

The number of nonbreeding avocets at Humboldt Bay clearly has varied 
greatly over the past 50 years. Although Grinnell and Miller (1944) listed no 
coastal records for the avocet north of Marin County (Novato), they omitted 
two specimens collected at Humboldt Bay 17 and 18 August 1935 
(Humboldt State University; Davis 1939). From 1945 through 1958, 17 
avocets occurred on Humboldt Bay on six occasions. Beginning in 1959, 
their numbers increased steadily from 75 (1959) to 800-1400 (1980s-90s). 
A similar increase occurred in the Point Reyes region, another important 
wintering area on California’s north coast (G. W. Page pers. comm., 
Shuford et al. 1989). More recently, the number of avocets wintering at 
Humboldt Bay appears to have declined by 50% to approximately 500. 

Virtually all early records of avocets were concentrated in the northeast 
quarter of Areata Bay. Nelson (1989) surveyed waterbirds in South Bay from 
July 1987 to June 1988 and listed avocets as seen irregularly, with numbers 
insufficient to warrant analysis. CBC data corroborate the early differences 
between Areata Bay and South Bay. Beginning in the 1990s, avocets 
expanded to the south and west shores of Areata Bay. CBC data show that 
in 1994 a small number of avocets began to feed on intertidal flats in the 
southeast corner of South Bay; at high tide they roosted in freshwater 
habitats of HBNWR. Most recently, during the winters of 1998 and 2000, 
avocets fed in flooded pastures adjacent to Areata Bay, which had rarely 
been observed previously (Harris unpubl. data). 

Establishment and growth of the Humboldt Bay population. At least 
two explanations, not mutually exclusive, exist for the establishment and 
increase in numbers of avocets at Humboldt Bay: (1) the avocet’s breeding 


7 


CHANGES IN THE ABUNDANCE AND DISTRIBUTION OF THE AVOCET 



Figure 3. Low-tide distribution of wintering American Avocets in intertidal habitats of 
Humboldt Bay during winter mapping surveys (see Table 1). Numbers represent the 
total number of avocets observed in each grid cell (500 x 500 m). 


8 





CHANGES IN THE ABUNDANCE AND DISTRIBUTION OF THE AVOCET 


population has increased in the region supplying winter visitors to Humboldt 
Bay; (2) nonbreeding habitat has changed, allowing avocets to expand their 
range into formerly unoccupied areas. On the basis of the Breeding Bird 
Survey, Robbins et al. (1986) reported that from 1965 to 1979 the 
population of the American Avocets increased significantly. This increase 
coincided with the major period of growth in the Humboldt Bay avocet 
population. Breeding Bird Survey data for 1966 to 1996 suggest that the 
population increased at >0.25% per year in Oregon, California, and 
Nevada, all breeding locales for color-marked birds observed at Humboldt 
Bay (Robinson and Oring 1996, Plissner et al. 1999). 

Evaluating the habitat hypothesis is difficult. Specimens show that 
nonbreeding avocets visited Humboldt Bay occasionally earlier in the 20th 
century (and probably before that). In the late 1950s local numbers began to 
increase. Evans and Harris (1994) speculated that construction of sewage- 
oxidation ponds in the northeast corner of Areata Bay in 1957 offered 
important habitat that was previously unavailable. Their habitat-limitation 
argument was based on spatial and temporal coincidence and seems 
plausible — numbers increased dramatically after the ponds’ construction, 
and avocets occurred only in their vicinity. From 1980s data, Evans and 
Harris (1994) described an ‘avocet home range’’ (868 ha) in the north-east 
corner of Areata Bay where avocets occurred predictably; they recorded 
only 30 observations of avocets outside this home range during 1142 hours 
observation. 

The habitat-limitation hypothesis posits that, prior to the construction of 
oxidation ponds, Humboldt Bay lacked one or more habitat requirements 
(e.g., food, water, sanctuary from predators) necessary to support wintering 
avocets (Evans and Harris 1994). Evans (1988) reported that avocets used 
these oxidation ponds to feed on cladocerans ( Daphnia magna), especially 
during October and November; most other times, however, avocets fed on 
nearby intertidal flats exposed during low tide. In addition to food, oxidation 
ponds may provide an important source of fresh water for avocets occupy- 
ing saltwater habitats (Evans and Harris 1994). Evans and Harris (1994) 
discounted the importance of fresh water because avocets breed in hypersa- 
line environments and have a well-developed salt gland (Mahoney and Jehl 
1985). In hypersaline habitats, however, avocets concentrate in areas of 
fresh water (L. W. Oring pers. comm.). This suggests that although avocets 
possess the physiological adaptation to tolerate these habitats, they pay an 
energetic cost in doing so, possibly influencing their distribution. 

Avocet distributions within Humboldt Bay and along the Pacific coast. 
An extension of the habitat-limitation hypothesis posits that the absence of 
avocets from some areas of Humboldt Bay stemmed from an absence of 
essential habitat requirements. A partial explanation for the patchy distribu- 
tion of avocets within the bay (Figure 3) may be related to the distribution of 
intertidal sediment types — avocets occurred in areas characterized by fine 
(silty-clay) sediments and were absent from areas with coarse (sandy) 
sediments (Thompson 1971). Danufsky (2000) found support for this 
hypothesis in an inverse relationship between avocet occurrence and sub- 
strate particle size. The aggregated baywide distribution of avocets probably 
stems from a relationship between substrate composition, prey availability, 


9 


CHANGES IN THE ABUNDANCE AND DISTRIBUTION OF THE AVOCET 


and the manner in which avocets feed. At Humboldt Bay, most avocets feed 
with a single scything maneuver (Evans and Harris 1994), a tactile foraging 
method of sweeping the bill through the upper film of sediment (Hamilton 
1975). Sediment type influences habitat use by avocets. For example, 
Quammen (1982) experimentally manipulated substrates at Newport Bay, 
California, and found that avocets moved more quickly through and scythed 
less (pecked more) where sand was added to a mud substrate. We hypoth- 
esize that avocet distributions are related to their scything for prey in fine 
substrates, which vary in their distribution in Humboldt Bay. 

If substrate particle size influences distributions of foraging avocets, then 
historical changes in avocet distribution in Humboldt Bay suggest that this 
habitat characteristic changed abruptly in the 1990s. In the mid-1990s, 
avocets began using areas of the bay that they rarely used during the period 
of population expansion (1960-1995). We speculate that wet years (1995- 
1999), coupled with logging of Humboldt Bay’s watershed, increased the 
amount of fine sediments in the bay, altering avocet foraging habitat. 
Interestingly, the February 1998 observations of avocets feeding in flooded 
pastures adjacent to Areata Bay coincided with the strongest (wettest) El 
Nino on record. 

Furthermore, we speculate that, in part, the northern limit of the avocet ’s 
winter range on the Pacific coast also may be related to the preference for 
feeding in fine sediments (Robinson et al. 1997). Specifically, the broad flats 
and fine sediments of Humboldt Bay are uncommon at bays along the 
Oregon and Washington coast. At least 17 major estuaries occur on the 
Oregon coast; Washington has five such estuaries. Fifteen of the estuaries in 
Oregon (e.g., Nestucca, Nehalem, and Rogue) are steep-banked with sandy 
substrates. Grays Harbor and Willapa Bay, Washington, and perhaps Coos 
Bay and Tillamook Bay, Oregon, offer broad intertidal flats and fine 
substrates of the type apparently favored by avocets. During migration, 
avocets occur occasionally on the Oregon-Washington coast. Paulson 
(1993) reported two avocets at Coos Bay on 12 December 1980, but they 
were not observed after that date. In summary, the combination of Humboldt 
Bay’s fine sediments, abundant prey (e.g., crustaceans, oligochaetes, poly- 
chaetes) linked to these substrates (Robinson et al. 1997), and moderate 
climate provide conditions sufficient to support avocets. These conditions 
are generally absent farther north on the Pacific coast. 

CONSERVATION IMPLICATIONS 

During January and February 1998 and February 2000, avocets fed in 
flooded pastures adjacent to Humboldt Bay. In this area, avocets had been 
observed in pastures rarely — Gerstenberg (1972) reported two in pastures 
during September 1969. Pastures commonly flood each winter following 
heavy rain. These observations strongly suggest that during winter 1998 
food availability in intertidal habitats was reduced because of habitat changes. 

In 1997-1998, winter use of flooded pastures coincided with two noteworthy 
environmental events. First, on 7 November 1997, an oil spill (-5000 gallons Bunker 
C fuel) occurred in Humboldt Bay, and oil contaminated intertidal flats throughout the 
west and southwest portions of Areata Bay, the main channel, and limited areas of 


10 


CHANGES IN THE ABUNDANCE AND DISTRIBUTION OF THE AVOCET 


South Bay. As yet, the effect of this oil on avocets remains undetermined, but oil may 
have killed some proportion of the avocets’ invertebrate prey, forcing the birds to feed 
in pastures. Second, the record rainfall generated by El Nino of winter 1998-1999, 
coupled with soil exposed by timber harvesting, may have increased the amount of 
sediments deposited on flats during winter storms. We speculate that during winter 
1998 these factors reduced invertebrate populations and habitat, forcing avocets to 
seek food in alternative foraging habitats. 

The hypothesis that avocets and their prey were influenced by heavy 
sedimentation in the bay is bolstered by similar observations for another bay- 
dependent species with a very different feeding ecology, the Black Brant 
(Branta bernicla nigricans). During winter, and more so during spring 
migration, thousands of Brant stage at Humboldt Bay where they feed 
nearly exclusively on eelgrass. During winter and spring 1998, Brant fed 
extensively in pastures adjacent to Areata Bay and South Bay. This pattern 
was observed only once before, during winter 1953, another El Nino year. 
During winter 1998-1999, K. Kovacs (pers. comm) and Mathis commented 
independently on the poor quality of eelgrass and its covering by sediment. 

Historical accounts of avocets and recent studies seeking to estimate their 
population and distribution within Humboldt Bay establish the foundation for 
long-term monitoring of avocets. Moreover, features of the American Avocet 
make it an ideal subject for monitoring changes in population size, distribu- 
tion, and habitat use at Humboldt Bay and elsewhere. Its conspicuous pied 
plumage, large size, and loose flocking tendencies make it easy to identify, 
observe, and count. Its use of fine sediment in intertidal habitats and 
dependence on prey taken from surface sediments render it a strong 
candidate as an indicator species for habitat change, which may influence 
other coastal waterbirds. We urge researchers at other locations to monitor 
avocets and expand our understanding of the relationships between avocet 
population size, distribution, and changes in the habitats in which they forage. 

ACKNOWLEDGMENTS 

For assistance, we thank the many individuals listed by Colwell (1994) and B. 
Acord, S. Bartz, K. Bentler, J. Bettaso, B. Brown, B. Condon, L. Connolly, N. Cull, 
K. Dhillon, E. Elias, C. Erickson, B. Frey, M. Gabriel, R. Green, P. Herrera, J. 
Liebezeit, L. Leeman, T. Leeman, M. Marriot, J. Mayer, S. McAllister, J. Moore, K. 
Nelson, M. Ortwerth, N. Pappani, J. Reilly, D. Ruthrauff, J. Saunders, L. Shannon, 
and J. Verschuyl. R. LeValley, N. War nock, and an anonymous reviewer improved the 
manuscript. We greatly appreciate the efforts of G. Eberly, D. Lee, and E. McCoy, 
who safely piloted us through channels of the bay. Partial support for our research 
came from a grant from the Oiled Wildlife Care Network, California Dept, of Fish and 
Game (Eureka office), and Humboldt State University. 

LITERATURE CITED 

Altmann, J. 1974. Observational study of behavior: Sampling methods. Behaviour 
49: 227-267. 

Barnhart, R. A., Boyd, M. J., and Pequegnat, J. E. 1992. The ecology of Humboldt 
Bay, California: An estuarine profile. U.S. Fish & Wildlife Serv. Biol. Rep. 1. 

Colwell, M. A. 1994. Shorebirds of Humboldt Bay, California: Abundance estimates 
and conservation implications. W. Birds 25:137-145. 


11 


CHANGES IN THE ABUNDANCE AND DISTRIBUTION OF THE AVOCET 


Colwell, M. A., and Cooper, R. J. 1993. Estimates of coastal shorebird abundance: 
The importance of multiple counts. J. Field Ornithol, 64:293-301. 

Danufsky, T. 2000. Winter shorebird communities of Humboldt Bay: Species diver- 
sity, distributions, and habitat characteristics. M.S. thesis, Humboldt State 
University, Areata, CA. 

Davis, J. M. 1939. More shorebirds from Humboldt Bay region. Condor 41:124. 

Evans, T. J. 1988. Habitat use and behavioral ecology of American Avocets 
(Recurui rostra americana ) wintering at Humboldt Bay, California. M.S. thesis, 
Humboldt State University, Areata, CA. 

Evans, T. J., and Harris, S. W. 1994. Status and habitat use by American Avocets 
wintering at Humboldt Bay, California. Condor 96:178-189. 

Gerstenberg, R H. 1972. A study of shorebirds in Humboldt Bay, California — 1968- 
1969. M.S. thesis, Humboldt State University, Areata, CA. 

Grinneli, J., and Miller, A. H. 1944. The distribution of the birds of California. Pac. 
Coast Avifauna 27. 

Hamilton, R. B. 1975. Comparative behavior of the American Avocet and the Black- 
necked Stilt (Recurvirostridae). Ornithol. Monogr. 17. 

Krebs, C. J. 1989. Ecological Methodology. Harper Collins, New York. 

Mahoney, S. A., and Jehl, J. R., Jr. 1985. Adaptations of migratory shorebirds to 
highly saline and alkaline lakes: Wilson’s Phalarope and American Avocet. 
Condor 87: 520-527. 

Nelson, E. T. 1989. The composition, distribution, and seasonal abundance of 
waterbirds using South Humboldt Bay, California, July 1987-June 1988. M.S. 
thesis, Humboldt State University, Areata, CA. 

Paulson, D. R. 1993. Shorebirds of the Pacific Northwest. Univ. of Wash. Press, 
Seattle. 

Plissner, J. H., Haig, S. M., and Oring, L. W. 1999. Within- and between-year 
dispersal of American Avocets among multiple western Great Basin wetlands. 
Wilson Bull. 111:314-320. 

Quammen, M. L. 1982. Influence of subtle substrate differences on feeding by 
shorebirds on intertidal mudflats. Mar. Biol. 71:339-343. 

Robbins, C. S., Bystrak, D., and Geissler, P. H. 1986. The Breeding Bird Survey: Its 
first fifteen years, 1965-1979. U. S. Fish & Wildlife Service Resource Publ. 157. 

Robinson, J. A., and Oring, L. W. 1996. Long-distance movements by American 
Avocets and Black-necked Stilts. J. Field Ornithol. 67:307-320. 

Robinson, J. A., Oring, L. W., Skorupa, J P, and Boettcher, R. 1997. American 
Avocet (Recurui rostra americana ), in The Birds of North America (A. Poole and 
F. Gill, eds.), no, 275. Acad. Nat. Sci., Philadelphia. 

Shuford, W. D., Page, G. W., Evens, J. G., and Stenzel, L. E. 1989. Seasonal 
abundance of waterbirds at Point Reyes: A coastal California perspective. W. 
Birds 20:137-265. 

Thompson, R. W. 1971. Recent sediments of Humboldt Bay, Eureka, California, final 
report. Petrol. Res. Fund PRF 789-G2. 


Accepted 2 September 2000 


12 


REPORT OF THE CALIFORNIA BIRD 
RECORDS COMMITTEE: 1998 RECORDS 


RICHARD A. ERICKSON, LSA Associates, One Park Plaza, Suite 500, Irvine, 
California 92614 

ROBERT A. HAMILTON, 34 Rivo Alto Canal, Long Beach, California 90803 


ABSTRACT: The California Bird Records Committee assessed 269 records of 92 
species in 1998, accepting 195 of them. New to California were the Bulwer’s Petrel 
( Bulweria buhverii), photographed on Monterey Bay, Monterey Co., the Bristle- 
thighed Curlew ( Numenius tahitiensis), photographed and videotaped at Crescent 
City, Del Norte Co., and Pt. Reyes, Marin Co., the American Woodcock ( Scolopax 
minor), photographed at Iron Mountain Pumping Plant, San Bernardino Co., the 
Iceland Gull ( Larus gloucoides), photographed at Bodega Harbor, Sonoma Co., and 
Otay, San Diego Co., the Bridled Tern ( Sterna anaethetus), sketched and described 
at Bolsa Chica, Orange Co., and the Olive-backed Pipit (Anihus hodgsoni ), photo- 
graphed at Southeast Farallon Island, San Francisco Co. The Harris’s Hawk 
(j Parabuteo unicinctus ) is no longer considered extirpated from California. With these 
additions, California’s bird list stands at 613 species, eight of which are nonnative. 

This 24th report of the California Bird Records Committee (hereafter the 
CBRC or the Committee) details the evaluation of 269 records of 92 
species. Although most records pertain to birds found in 1998, the period 
covered by this report spans the 25 years from 1974 to 1999. Accepted 
were 195 records involving 73 species. The acceptance rate of 72.5% was 
typical for the last decade but below the overall Committee average 
(Rottenborn and Morlan 2000). Sixty-three records were not accepted 
because of insufficient documentation or descriptions were inconsistent with 
known identification criteria. Six additional records were not accepted 
because of questions concerning the bird’s natural occurrence or establish- 
ment of introduced populations. Counties best represented by accepted 
records were Monterey (14 records), Kern (13), Orange (13), San Diego 
(13), Marin (11), San Francisco (11), and Riverside (10). Records were 
accepted from 20 other counties. 

Highlights of this report include the addition of six species new to the 
California list: Bulwer’s Petrel ( Bulweria bulwerii), Bristle-thighed Curlew 
( Numenius tahitiensis), American Woodcock ( Scolopax minor), Iceland 
Gull ( Larus glaucoides), Bridled Tern ( Sterna anaethetus), and Olive- 
backed Pipit (An thus hodgsoni). Potential first state records not accepted 
include the Lesser White-fronted Goose (Anser erythropus), Barnacle 
Goose ( Branta leucopsis), Falcated Duck ( Anas falcata), Ross’s Gull 
(. Rhodostethia rosea), and McKay’s Bunting ( Plectrophenax hyperboreus). 
Also reported here is the Committee’s decision to readmit the Harris’s Hawk 
( Parabuteo unicinctus) to the list of extant species in California. With these 
additions, California’s list stands at 613 species, of which eight are non- 
native and two have been extirpated in historical times. Potential first state 
records of the Shy Albatross (Thalassarche cauta), Glossy Ibis ( Plegadis 
falcinellus), Northern Bobwhite ( Colinus uirginianus), Slaty-backed Gull 
{ Larus schistisagus ), Eastern Bluebird ( Sialia sialis), and Gray Silky-fly- 


Western Birds 32:13-49, 2001 


13 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 

catcher ( Ptilogonys cinereus) are currently being considered. 

Other highlights include accepted records of the Wedge-tailed Shearwater 
{Puffinus pacificus), Neotropic Cormorant { Phalacrocorax brasilianus ), 
Whooper Swan ( Cygnus cygnus), Common Pochard (Aythya ferina), 
Gyrfalcon ( Falco rusticolis), Mongolian (Charadrius mongolicus) and 
Wilson’s (C, wilsonia ) plovers, Sooty Tern ( Sterna fuscata), Long-billed 
Murrelet ( Brachyramphus perdix), Ruby-throated Hummingbird 
( Archilochus colubris), Veery (Catharus fuscescens), and Field Sparrow 
{Spizella pusilla). Also accepted were California’s first spring Sulphur- 
bellied Flycatcher ( Myiodynastes luteiventris) and the first CBRC-reviewed 
Blue-headed Vireos { Vireo solitarius). The Committee’s decision not to 
accept a long-standing record of the Red-necked Stint ( Calidris ruficollis ; 
the state’s only purported specimen record) is reported. Species recorded in 
especially high numbers in 1998 included the Short-tailed Albatross 
( Phoebastria albatrus; 3 records), Lesser Black-backed Gull (Larus fuscus; 
10), Eastern Wood-Pewee ( Contopus virens; 5), Dusky-capped Flycatcher 
( Myiarchus tuberculifer; 8), Philadelphia Vireo ( Vireo philadelphicus; 8), 
Yellow-green Vireo ( Vireo flauouiridis; 8), Gray Catbird { Dumetella 
carolinensis; 16), Painted Bunting ( Passerina ciris; 7), and Common 
Grackle (Quisca/us quiscula ; 5). 

The list of species reviewed by the CBRC is posted at the Western Field 
Ornithologists’ web site (http://www,wfo-cbrc.org). This site also includes 
the entire California state list, the Committee’s bylaws, a reporting form for 
direct e-mail submission of records to the CBRC, the addresses of current 
Committee members, a photo gallery of recent submissions including 
several birds published in this report, a list of relevant publications by CBRC 
members, and other information about the CBRC, WFO, and Western 
Birds. 

All records reviewed by the CBRC (including copies of descriptions, 
photographs, videotapes, audio recordings and Committee comments) are 
archived at the Western Foundation of Vertebrate Zoology, 439 Calle San 
Pablo, Camarillo, California 93012, and are available for public review. The 
CBRC solicits and encourages observers to submit documentation for all 
species on the review list, as well as species unrecorded in California. 
Documentation should be sent to Guy McCaskie, CBRC Secretary, P O. 
Box 275, Imperial Beach, CA 91933-0275 (e-mail: guymcc@pacbell.net). 

Committee News. The Committee’s voting membership after the 6 
January 2001 annual meeting consisted of Jon L. Dunn, Richard A. 
Erickson (chairman), Kimball L. Garrett (vice chairman), Robert A. Hamilton, 
Bert McKee, Joseph Morlan, Michael A. Patten, Peter Pyle, Scott B. Terrill, 
and John C. Wilson. Guy McCaskie will serve as non-voting secretary. 
Recent Committee members who also voted on many of the records in this 
report include Matthew T. Heindel, Steve N. G. Howell, Alvaro Jaramillo, 
Guy McCaskie, Michael M. Rogers (outgoing secretary), Stephen C. 
Rottenborn, Mike San Miguel, and Daniel S. Singer. 

At its 15 January 2000 meeting the Committee decided to publish its 
Annotated List of the Birds of California — for the first time including 
nesting annotations created by the Committee — in California Fish and 


14 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


Game, to remove the Gray Catbird ( Dumetella carolirtensis ) and the 
species pair of the White/Black-backed Wagtails ( Motacilla albaAugens) 
from the CBRC review list, the latter to conform with all other species pairs/ 
groups where records must be attributed to a single species, to place the 
Harris’s Hawk ( Parabuteo unici rictus) on the review list but to review 
records in five-year increments (see species account below), and to review 
records of the Iceland Gull ( Larus glaucoides) in one-year increments (see 
species account below). Also, see Miscellaneous Decisions below. 

Format and Abbreviations. As in other recent CBRC reports, records are 
generally listed geographically, from north to south, and/or chronologically 
by first date of occurrence. Included with each record is the location, county 
abbreviation (see below), and date span. The date span usually follows that 
published in North American Birds (formerly American Birds or Field 
Notes ) but, if the CBRC accepts a date span that differs from a published 
source, the differing dates are italicized. Initials of the observer(s) responsible 
for finding and/or identifying the bird(s) — if known and if they have supplied 
supportive documentation — are followed by a semicolon, then the initials, in 
alphabetized order by surname, of additional observers submitting support- 
ive documentation, then the CBRC record number consisting of the year of 
observation and chronological number assigned by the secretary. All records 
are sight records unless otherwise indicated: a dagger (+) indicates the 
observer supplied a supportive photograph, (t) videotape, (§) a voice 
recording, and (#) a specimen record, followed by the acronym (see below) 
of the institution housing the specimen and its catalog number. 

An asterisk (*) prior to a species’ name indicates that the species is no 
longer on the CBRC review list. The first number in parentheses after the 
species’ name is the number of records accepted by the CBRC through this 
report; the second is the number of new records accepted in this report 
(because this number excludes records thought to pertain to returning 
individuals, it may be zero). Two asterisks (**) after the species’ total indicate 
that the number of accepted records refers only to a restricted review period 
or includes records accepted for statistical purposes only; see Roberson 
(1986) for more information. 

When individual birds return to a location after a lengthy or seasonal 
absence, each occurrence is reviewed under a separate record number, and 
Committee members indicate whether or not they believe the bird is the 
same as one accepted previously. Such decisions follow the opinion of the 
majority of members and, if a bird is considered a returning individual, the 
total number of records remains unchanged. 

Although the CBRC does not formally review the age, sex, or subspecies 
of each bird, information on these subjects is often provided during the 
review process (and in some cases a strong or unanimous consensus is 
achieved). We report much of this information. 

The CBRC uses standard abbreviations for California counties; those used 
in this report are ALA, Alameda; BUT, Butte; CC, Contra Costa; DN, Del 
Norte; F1UM, Humboldt; IMP, Imperial; INY, Inyo; KER, Kern; LAS, 
Lassen; LA, Los Angeles; MRN, Marin; MEN, Mendocino; MNO, Mono; 
MTV, Monterey; ORA, Orange; PLA, Placer; RIV, Riverside; SAC, Sacra- 


15 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


mento; SBE, San Bernardino; SD, San Diego; SF, San Francisco; SJ, San 
Joaquin; SLO, San Luis Obispo; SM, San Mateo; SBA, Santa Barbara; 
SCL, Santa Clara; SCZ, Santa Cruz; SIS, Siskiyou; SON, Sonoma; TRI, 
Trinity; VEN, Ventura; YUB, Yuba. A full list of county abbreviations is 
available on the WFO-CBRC web site. Other abbreviations used: I., island; 
L,, lake; Mt, mountain; n. miles, nautical miles; N.W.R., national wildlife 
refuge; Pt., point; R., river; W.M.A., wildlife management area. 

Museum collections housing specimens cited in this report, allowing 
access to Committee members for research, or otherwise cited are the 
California Academy of Sciences, San Francisco (CAS), Natural History 
Museum of Los Angeles County, Los Angeles (LACM), Pacific Grove 
Museum of Natural History (PGMNH), San Bernardino County Museum, 
Redlands (SBCM), San Diego Natural History Museum (SDNHM), Santa 
Barbara Museum of Natural History (SBMNH), and the Museum of Verte- 
brate Zoology, University of California, Berkeley (MVZ). 

RECORDS ACCEPTED 

YELLOW-BILLED LOON Gauia adamsii (65, 1). A first-winter bird was at Field’s 
Landing, HUM, 18 Jan 1998 (MHM; 1998-023). An alternate-plumaged adult off 
Otter Pt., Pacific Grove, MTY, 24 Nov 1998-19 Mar 1999 (DR; 1999-059) returned 
for its sixth winter, as summarized by Rottenborn and Morlan (2000). 

SHORT-TAILED ALBATROSS Phoehastria albatrus (7**, 3). The long-predicted 
increase in records of this endangered species in California waters may be underway. 
Immatures about 20 miles W of Bodega Head, SON, 28 Aug 1998 (GLf, DnWNf; 
BDP, DGS; 1998-148), 15 n. miles W of Pt. Reyes (38° 03' N, 123° 21' W), MRN, 
26 Oct 1998 (BMcKf; ADeM, SWe ; 1998-179), and 3 n. miles WNW of Pt. Pinos 
(36° 40'N, 122° 00'W), MTY, 21 Dec 1998 (SNGH, ADeM; RJA, SFB, RLBf, JEG, 
GGr, PKe, DLShf, MS, RLTf; 1998-231) were considered separate individuals by the 
Committee and provided the first accepted records since 1987, On the basis of 
photographs, Pyle considered the first two birds “somewhat worn whereas the third 
was immaculate (also per SNGH description),” leading him to conclude that the first 
two birds were >1 year old and the third was in its first year (ca. 6 months old). This 
also explains the pale eye crescents apparent on the first two birds but not the third, 
the first sign of adult plumage (H. Hasegawa pers. comm, to Pyle). A photograph of 
the third bird appeared in N. Am. Birds 53:203 and on the cover of Western Birds 
vol. 31, no. 1 (2000). 

BULWER’S PETREL Bulweria bulwerii (1, 1). One seen well by a boatload of 
observers on Monterey Bay, MTY, 26 July 1998 (JD, BMcKt; AB, GBf, JEG, JMof, 
DLSh, RLTf, DWf; 1998-119) represented the first well-documented record for 
North America — preceding a bird off North Carolina (LeGrand et al. 1999) by only 
days. Therefore the record was forwarded to the American Birding Association’s 
Checklist Committee for its review. The similar Jouanin’s Petrel (B. fallax) has been 
recorded as close as Hawaii (Clapp 1971) but is stockier (thicker winged, larger 
headed, and larger billed) and typically has proportionally shorter outer rectrices. The 
Committee’s unanimous acceptance of the record on a single circulation is a 
testament to the exceptional package of documentation compiled by veteran trip 
organizer Debra L. Shearwater. Photographs appeared in Field Notes 52:498 and 
519. 

WEDGE-TAILED SHEARWATER Puffinus pacificus (3, 1). A dark-morph indi- 
vidual was on Monterey Bay, MTY, 10-21 Oct 1998 (EWGf, TMcG; BBf, RLBff, 


16 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


DCf, ADeMt, PGf, JEG, AKf, GMcC, BMcKf, DR+, MMRt, DMS, DLSh, WDS, 
JSf, RLTf; 1998-162). Another exceptionally complete set of documentation (cf. 
Bulwer’s Petrel) made the decision of Committee members easy. The previous 
California records were from Monterey Bay, MTY, 31 Aug 1986 (Stallcup et al. 
1988, Bevier 1990) and the Salton Sea, RIV, 31 Jul 1988 (McCaskie and Webster 
1990, Pyle and McCaskie 1992). A photograph of the 1998 bird appeared on the 
cover of N. Am. Birds vol. 53, no. 1 (1999). 

MANX SHEARWATER Puffinuspuffinus (43, 6). A record of one seen from shore 
at Pigeon Pt., SM, 5 May 1996 (GD, NL; 1996-077A) received the requisite number 
of accept votes only on its fourth and final circulation. Five accepted records in 1998 
represented only half of the previous year’s record total (Rottenborn and Morlan 
2000): Cordell Bank. MRN, 11 Jul (MI; 1999-144), three on Monterey Bay, SCZ, 8 
Aug (MI, BMcK; 1998-021; BMcKf, MI; 1998-022; MI; 1999-134), and one at SE 
Farallon I , SF, 3 Oct (WR; 1999-032). One additional record from 1998 is still under 
Committee review. 

MASKED BOOBY Sula dactylatra (11, 1). The recent surge in records of this 
species continued with an adult at Aho Nuevo L, SM, 19 Jun-6 Aug 1998 (MF, MM, 
CAM, JM, JDP, RWR, MMR, SCR, DGS, FT; 1998-098). The bird was clearly a 
Masked Booby in the now limited sense that excludes the Nazca Booby (S. grand), 
newly recognized as a species (Pitman and Jehl 1998, Roberson 1998, AOU 2000). 
Field Notes 52:499 refers to a videotape that the CBRC has not seen. 

BLUE-FOOTED BOOBY Sula nebouxii (80**, 0). A subadult near Obsidian Butte, 
IMP, 19-28 Feb 1998 (JEP, DSP; RJN; 1998-111) was considered probably the 
same individual as one seen at Salton City, IMP, 29 Nov 1997 (Rottenborn and 
Morlan 2000). 

BROWN BOOBY Sula leucogaster (58, 7). Four records in 1998 were uncompli- 
cated, but records of three birds found earlier were not. The former were an immature 
at San Simeon State Beach, SLO, 19 Jan 1998 (TME, KH; 1998-054; what may 
have been the same bird was reported at Morro Bay, SLO, 20-22 Dec 1997 — N. 
Am. Birds 52:256), an adult at Morro Rock, SLO, 5 Jul 1998 (MG; 1998-122), an 
adult male at Pt. Reyes, MRN, 7 Jun 1998 (KB, GGa, KCK, LMLf, JM, RSf; 1998- 
088), and a second-year male at SE Farallon L, SF, 12-24 Oct 1998 (WR, PPf; 
1999-007). The Pt. Reyes bird was the northernmost ever recorded in California; one 
was in northwestern Washington 18 Oct-9 Nov 1997 (Field Notes 52:114-115). 

Most interesting of the other three records was an immature seen from a boat 12 
n. miles SSW of Santa Barbara, SBA, 31 Aug 1996 (BS; DG, KAH, AH, DKf, CAM, 
MMR; 1996-1 17A; Figure 1). The record received 80% acceptance in its first round 
of circulation as a Red-footed Booby (S. sula; 1996-117) before failing acceptance by 
the same margin. It was then resubmitted as a Brown Booby (1996-1 17 A) and gained 
a unanimous decision on its second round. Descriptions of the bird’s foot color varied, 
so observers were not of one mind concerning the bird’s identity. The Committee’s 
initial reliance on reported reddish feet (also suggested by photographs) eventually 
gave way to a full consideration of other characters (i.e., underwing pattern, dark 
brown coloration, bill color and bill/head shape, heavier structure, uniform color of 
central rectrices) and a realization that apparent foot color can be influenced greatly 
by lighting. 

An immature 15 miles SE of East Pt., Santa Rosa L, SBA, 30 Sep 1997 (BDf; 1998- 
069) was supported by a single color photograph and was accepted (unanimously) only 
after its third circulation (a second bird reported without documentation was not 
accepted). This may have been the same bird seen between Santa Rosa I. and Santa Cruz 
I., SBA, 19 Oct 1997 (Rottenborn and Morlan 2000). Another immature at Newport 
Beach, ORA, 7 Dec 1997-7 Feb 1998 (LRH; GMcC, JEP, DGS; 1997-211) was 


17 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


initially identified as a Red-footed Booby. Only after circulating the record three times did 
the Committee consider the entire date span to pertain to the same bird. 

NEOTROPIC CORMORANT Phalacrocorax brasilianus (10, 2). Adults were 
near Obsidian Butte, Salton Sea, IMP, 28 Apr-12 Jul 1998 (KZKf; GMcC, MAPf, 
SCR, DGS; 1998-079) and 3 miles S of Mecca, Salton Sea, RIV, 14 Jun-16 Jul 1998 
(MAPf; GMcC; 1998-097). 

TRICOLORED HERON Egretta tricolor (26**, 1). Returning adults were at the 
Tijuana R. estuary, SD, 25 Jul 1998-12 Feb 1999 (GMcC; 1998-112; same as 

1997- 184, Rottenborn and Morlan 2000) and Bolsa Chica, ORA, 31 Oct 1998-7 
May 1999 (RF, MTH, PKnf, VL, JM, MMRt, SS; 1998-200; same as 1997-167 and 

1998- 004, Rottenborn and Morlan 2000). Another adult was 4 miles S of Mecca, 
Salton Sea, RIV, 31 Jan-7 Mar 1998 (MAP; KLGf, CAM, GMcC; 1998-048). 

REDDISH EGRET Egretta rufescens (78, 3). An immature around Johnson Rd. 
and Colfax St. at the N end of the Salton Sea, RIV, 11 Jul -1 Aug 1998 (RC, GMcC, 
MJSM, DGS; 1998-101) was considered the same as an immature at Obsidian Butte, 
Salton Sea, IMP, 2-8 Aug 1998 (GMcC; MJSM; 1998-114). On the coast, an adult 
was at Seal Beach N.W.R., ORA, 17 Apr 1998 (JF; 1998-071), and an immature was 
at the Tijuana R. estuary, SD, 29 Aug- 4 Oct 1998 (GMcC; 1998-121). 

YELLOW-CROWNED NIGHT-HERON Nyctanassa uiolacea (18, 0). The adult 
that has been reported at La Jolla, SD, for nearly 20 years (Rottenborn and Morlan 
2000) was seen again 30 Apr-26 May 1998 (DGS; 1998-083). 

EMPEROR GOOSE Chen canagica (63, 1). One was on the Garcia R. Flats, MEN, 
28 Dec 1996-24 Feb 1997 (ERH; MGt, MMRt, PMS; 1997-091). Two birds were 
initially reported with marginal written documentation, but photographs received 
subsequently documented the presence of only one . 

TRUMPETER SWAN Cygnus buccinator (24, 2). Adults were 7.5 miles NNE of 
Marysville, YUB, 3 Feb 1994 (TM, BEWt; 1998-077) and on Rindge Tract, SJ, 17 
Dec 1995 (WH; 1998-215). A second adult accompanying the latter bird, also 
reported as a Trumpeter Swan, was not accepted. 

WHOOPER SWAN Cygnus cygnus (5, 1). An adult at Lower Klamath N.W.R., 
SIS, 28 Jan 1998 (TCB; 1998-026) was in the same area as ones seen in 1991-92 
and 1994 (McCaskie and San Miguel 1999) and may have been the same individual. 
Patten (2000) summarized recurring doubts concerning the provenance of Whooper 
Swans in North America. 

GARGANEY Anas querquedula (21, 1). One at the Salinas sewer ponds, MTY, 25 
Apr-4 May 1998 (CHo; DR, SCR, JSf; 1998-084) fit the pattern of males in spring 
established by most records of this species (Spear et al. 1988) and was the first to be 
found in Monterey County. 

COMMON POCHARD Aythya ferina (2, 1). Two males photographed at Bolsa 
Chica, ORA, 26 Dec 1994 (PK; 1995-082; Figure 2) made California’s second 
record, following a male at Silver Lake, SBE, during falls and winters between 
February 1989 and November 1992 (Patten 1993). The current record struggled 
through four rounds because the birds were identified from photographs well after the 
sighting, and the prints that accompanied the record through three rounds were 
inadequate for some members to rule out hybrids. Communication between Commit- 
tee members and the photographer and inspection of the original slides led to 
eventual acceptance by a 9-1 plurality. 

KING EIDER Somateria spectabilis (34, 1). Record 1998-014 involved a female 
seen at Laguna Beach, ORA, 23 Dec 1995 (JEP) and at Huntington Beach, ORA, 28 
Dec 1995 (DSP). 


18 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


MISSISSIPPI KITE Ictinia mississippiensis (27, 1). A first-summer bird at Furnace 
Creek Ranch, Death Valley, INY, 6-7 Jun 1998 (SCR, MJM, MMRf, MJSM, SBT; 
KSG, MSM; 1998-087) was at the site of more records than anywhere else in 
California but was the first in the state since 1994 (Howell and Pyle 1997). 

HARRIS’S HAWK Parabuteo unicinctus (4**, 4). Following the appearance of a 
number of Harris’s Hawks in southwestern California in 1994, the Committee 
packaged several records together under number 1996-080 so it could reconsider its 
designation of the California population as extirpated. After a single circulation, we 
decided to consider the individual reports separately to avoid the possibility of different 
members accepting different records and no single record achieving acceptance on its 
own. After three more circulations, four records had been accepted: up to eight at 
Borrego Valley, SD, 15 Apr 1994-22 Sep 1996 (RT; AME, RAE, PJ, PEL, CAM, 
GMcC, MAPf, DRf; 1996-080A; Figure 3), up to five (including two that constructed 
a nest) at Boulevard, SD, 1 Jun 1994-31 Oct 1995 (PEL, ADS; 1996-080B), one in 
the Coachella Valley, RIV, 31 Dec 1994-29 Jan 1995 (SJM; MAP; 1996-080D), and 
up to five at George AFB, SBE, 2-21 Jan 1995 (EACf, CAM, MAP, MSM; 1996- 
080E). Thus the Committee no longer considers the species extirpated from Califor- 
nia. Four more records from the same period are still under review. 

At its January 2000 meeting the Committee decided to review subsequent records 
only once every five years. Patten and Erickson (2000), using primarily CBRC data, 
summarized the reappearance of this species in California and northern Baja 
California, emphasizing the cyclic nature of this species’ occurrence in those areas 
over the past 150 years. Erickson et al. (in press) reported additional recent records for 
northern Baja California. 

*ZONE-TAILED HAWK Buteo albonotatus (54, 3). An adult was in Mission Viejo, 
ORA, 18 Oct 1975 (SJG; 1999-178), an adult was near Lake Henshaw, SD, 13 Jun 
1998 (PU; 1999-039), a returning adult (cf. Rottenborn and Morlan 2000) was at 
Goleta, SBA, 21 Oct 1998-24 Mar 1999 (FE; SS; 1998-206), and two adults were at 
San Diego Wild Animal Park, San Pasqual, SD, 19 Nov 1998-15 Mar 1999 (GMcC, 
JM; 1999-024), one judged to be a returning bird (cf. Rottenborn and Morlan 2000). 

GYRFALCON Falco rusticolus (8, 1). A gray-morph juvenile at Tule Lake N.W.R., 
SIS, 24 Nov 1998 (RE; 1999-058) was the fifth to be recorded in the California 
portion of the Klamath Basin. 

YELLOW RAIL Coturnicops noueboracensis (70, 2). Females were found in 
weakened condition in Manhattan Beach, LA, 20 Oct 1998 (#LACM 110747; 1999- 
097) and Santee, SD, 16 Dec 1998 (#SDNHM 50186; 1998-232). There were only 
four previous accepted records for southern California, three in the period 1896- 
1917. 

In comments, Patten noted the light weight of both specimens (perhaps merely a 
result of their poor condition) and questioned how the similar Swinhoe’s Rail (C. 
exqui situs) of east Asia was eliminated from consideration. Although the chance of 
Swinhoe’s Rail reaching California is remote and these specimens do match other 
Yellow Rail specimens, specimens for comparison with Swinhoe’s Rail are not 
available at either museum. This question seems relevant to other California speci- 
mens as well, so an eventual review of all records may be in order. 

MONGOLIAN PLOVER Charadrius mongolicus (8, 2). Juveniles were at Bodega 
Harbor, SON, 2-3 Sep 1996 (DnWNf; KB, LML, BDP; 1996-118) and the Eel River 
Wildlife Area, HUM, 2-3 Oct 1998 (SNGH, SWe; WB, AJ, JSM, GMcC, JM, MSM; 
1998-150). Previous California records, all coastal from Marin County through 
Ventura County, span the period 12 July to 25 September. 

WILSON’S PLOVER Charadrius wilsonia (8, 1). An adult male was at Coronado, 
SD, 27 Apr-1 May 1998 (NBB, BFf, PAG, RL, JOZ; 1998-072); reports of it as late 


19 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 



Figure 1. Immature Brown Booby, Sula leucogaster, 12 nautical miles south- 
southwest of Santa Barbara, Santa Barbara County, 31 Aug 1996 (1996-1 17A). This 
individual was originally identified as a Red-footed Booby, S. sula, and was nearly 
accepted as such. Note the seemingly red feet but also apparent whitish mottling on 
the underwing coverts and the bill’s shape (relatively thick with flat forehead and 
culmen) and color (bluish gray with no hint of pinkish). 

Photo by David Koeppel 



Figure 2. Male Common Pochard, Aythya ferina, one of two at Bolsa Chica, Orange 
County, 26 Dec 1994 (PK; 1995-082) that represent the second record for California. 


Photo by Peter Knapp 


20 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 



Figure 3. Adult Harris’s Hawk, Parabuteo uriicinctus, in Borrego Springs, San 
Diego County, 13 Jan 1995 (1996-080A), one of up to eight there beginning in April 
1994. On the basis of this record and three others, the species is no longer considered 
extirpated in California. 

Photo by Don Roberson 


as 2 May were not accepted. Surprisingly, of six California records in the latter half of 
the 20th century, only one other was south of Ventura County. 

AMERICAN OYSTERCATCHER Haematopus paUiatus (16, 1). An adult was at 
San Nicolas L, VEN, 28 Jun-3 Sep 1998 (RAH; 1998-117); nine of California’s 
accepted records are from the Channel Islands. 

BRISTLE-THIGHED CURLEW Numenius tahitiensis (2, 2). California’s first 
accepted records, of one at Battery Pt., Crescent City, DN, 14-16 May 1998 (ADBf; 
BED, GEf, GMcC, JM, DvWN, DRt; 1998-075) and one at Kehoe Beach, Pt. Reyes, 
MRN, 16-25 May 1998 (LMLf ; KB, LC, MF, BG, SCH, SNGHf, KCK, MJM, JSM, 
JLM, JM, BDP, MAP, RWR, JR, MMR, SCR, TR, MSMt, DGS, RSt; 1998-076), 
were part of a larger phenomenon involving 13-17 birds from Washington through 
central California and discussed in detail by Patterson (1998) and Mlodinow et al. 
(1999). Photographs of the accepted birds were published by Patterson (1998), in 
Field Notes 52:384, and on the cover of Western Birds vol. 30, no. 3 (1999). Two 
other records were not accepted (see below). 


21 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


f . J* 


aii 



Figure 4. Adult Iceland Gull, Larus glaucoides, at Bodega Harbor, Sonoma County, 
13 Jan 1985 (1985-007), the first accepted record for California. 

Photo by Albert Ghiorso 



BAR-TAILED GOD WIT Limosa lopponica (23, 2). An adult was at Ravenswood 
Open Space Preserve. SM, 12-29 Sep 1998 (RST; AME, MF, DGS; 1998-138), and 
a juvenile was at the confluence of Alviso and Coyote sloughs. SCL/ALA, 2 Oct 1998 
(SCR; 1998-204). 

AMERICAN WOODCOCK Scolopax minor (1. 1). One was at Iron Mt. Pumping 
Plant, SBE, 3-9 Nov 1998 (MAP, GMcC; RF. KSG. JM, DRt: 1998-177). Patten et 
al. (1999) provided details of this record, California's first. 

LITTLE GULL Larus minutus (72, 4). An adult was at Mystic L., near Lakeview, 
RIV, 15 Nov 1998-27 Mar 1999 (MAPt: DSC, RF, RL. CAM, GMcC, JM; 1998- 
181). An adult or bird in second alternate plumaqe at Alviso, SCL, 28 Apr-2 May 
1998 (SCR; MWE, MF, NL, MJM, RWR; 1998-070) was followed by one in first 
alternate plumage there 8-17 May 1998 (SCR; NL, JM; 1998-078). One in first 
alternate plumage at Lower Klamath N.W.R.. SIS, 14-16 Jun 1998 (ADB, RE; 1998- 
140) was at a less traditional time and place. 

BLACK-HEADED GULL Larus ridibundus (20, 0). An adult at Santa Barbara, 
SBA, 17 Nov-24 Dec 1997 (DSW; 1997-204} returned for its sixth winter at the 
same location. 

ICELAND GULL Larus glaucoides (2, 2). After nearly 15 years of intermittent 
review (cf. Heindel and Garrett 1995, Rottenborn and Morlan 2000) and the formal 
input of at least 12 outside consultants from throughout North America and Europe 
(E. A. T. Blom, Davis Finch, Roger Foxall, Daniel D. Gibson, Michel Gosselin, Peter 
J. Grant, Raymond Henson, Ted Hoogendoorn, Dennis Paulson, J. V. Remsen, Jr., 
Stuart Tingley, Thede Tobish), this species was finally added to the state list at the 
January 2000 meeting, on the basis of two records: an adult at Bodega Harbor, SON, 
30 Dec 1984-18 Jan 1985 (KFC, DDeS; JRAt, LCB, RAE, AGf, RLeBt, GMcC, 


22 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


Lo-rat -VivSma* 


I r 


2H-2.S 1W6 

Ob^i^lAk, tj._t-H.C_ S *£-0 o_ 


selr* 



\ 
no)t 

b rourw ■»poH 
«kcK^W -tUn^S 
<i*»A -^Aas 
<V» icL «A P<^2tLS+ 

4W-V p.nK 


Ao».. \ **-«o s-Ab^ 


• ‘b'vx c^yia. w i~t,OL» lAd-VVi 

Ci n AirctA Cn mp^rifc om ^ 

• tCU «- -Vf-WUrt A. Uerr^‘i. 

W 1+T» *5 t e_ ^ A^l 0 

— oJA t* V CLtV, Un 1 

<L<V W'V-C r' i&c*A-e 

bkt\ur +4 ^» 

• *%< r Cs.^> vA <Vv-C *^.«j-< s^o , V*^, 

C.O'fJW V, A. CX. 


9 MV\ 




V.k-^ 


Ca.^ < fiul\ CV?^ ^ 

4A«^ 4^4 t^cc i <5^, 


O-C kA\ 

• kcI Ua^vA 



b l ^ 
e-e-vA-«rs 



L-j'rt A-a 

CTH 

«yr4A*-A< Covty A ■ -. 

C«M4.rb 0-*>b» — 
QlX-i-AV^ »v» s Hh 

-fcol .A t>v il* 'i 



o * t ’ 

aA* c_w^* ,!. A . T*A^-*v\ 


ZS 


Tix^ha>V| - 
2- F*c_fc> rv.ajTM 

Ct^e-wx, -C 


i*ii sV*nA v 


Al^»o 

Tu^«* c. 

M.c.CoL*»Wifi 
Cl<* e*^J 3* TTokH 
HwtOOj'A’ ^*»nc^ cA Cs.C , 
(^Sl "X a.vmrt.rv| ) 


Figure 5. First-winter Lesser Black-backed Gull, Lorus fuscus, at Obsidian Butte, 
Salton Sea, Imperial County, 25 Jan 1998 (1998-042), the first record of this 
plumage in California. 


Sketch by Michael A. Patten 


JM, DnWNt, JOlt, BDP, DGSt, RSt, SWif; 1985-007; Figure 4) and a juvenile at 
the Otay dump, SD, 17-25 Jan 1986 (REWt; LRBt, JLDf, JML, MJL, GMcC, DR, 
ASt; 1986-015). Perhaps not surprisingly, both of these birds were at the pale end of 
the spectrum (i.e., more like nominate glaucoides than the expected Kumlien’s Gull, 
L. g. kumlieni, of North America); birds closer in appearance to kumlieni (and thus 
pale Thayer’s Gulls, L. thayeri, as well) have not fared well in the CBRC review 
process (see Rottenborn and Morlan 2000 and Records Not Accepted below). 

Many Committee members expressed skepticism of the current taxonomic treat- 
ment of the Iceland/Thayer’s Gull complex in America (AOU 1973, 1983, 1998) — 
indeed, that skepticism was largely responsible for the temporary suspension of 
CBRC review of these records — but all eventually agreed that such taxonomic 
considerations are beyond the scope of the CBRC and that the treatment here 
matches better what has been done with extralimital records elsewhere in North 


23 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


Chic* trCfer# 

iinm 

obSs.rl L t,4£-&oa 


** StVe, *Tib'***y 

bdK *>@Sn 

“Via £ 




■ 

y*yVSh W*»> 


1^ Ste«f f fu*& 




|>r> 


2 O^C wWef 


Hlff esufc'k Hij 




foV 






d? 


i 2” 5 


Figure 6. Adult Bridled Tern, Sterna anaethetus, at Bolsa Chica, Orange County, 17 
Jul 1998 (1998-105), the first accepted record for California. 

Sketch by Marshall Iliff 


America. Because of the complex taxonomy and identification of this species, records 
will be handled differently than for other species; records will not begin circulation as 
received but will be held until the next annual meeting, when they may or may not 
undergo an expedited review. Rottenborn and Morlan (2000) addressed the 
Committee’s latest dealings with these birds. 

LESSER BLACK-BACKED GULL Larus fuscus (17, 8). California is not immune 
from the population expansion of this species in North America, as ten individuals 
were seen in 1998: a returning adult at Alviso, SCL, 11 Oct 1997-23 Jan 1998 
(MJM; MWE, MF, JM, RWR; 1997-150) and again in the Alviso area, SCL/ALA, 21 
Oct-8 Dec 1998 (SCR; MMRt; 1998-203), a second-winter bird at L. Cunningham, 
San Jose, SCL, 31 Oct 1997-31 Mar 1998 (SCRf, MMRt; MF, MTH, NL, MJM, 
GMcC, JM, RWR, DRf, RMSf, MSi; 1997-1 78), returning in its third winter 20 Dec 
1998-14 Mar 1999 (SCR; MJM, MMRf; 1999-001), an adult at L. Cunningham, 
San Jose, SCL, 8 Jan 1998 (BMcK; 1998-003), a returning adult at Doheny State 
Beach, ORA, 8 Oct 1997-4 Mar 1998 (JDWf; TC, MTHf, MF, GMcC; 1997-156) 
and again 4 Nov 1998-10 Mar 1999 (LMLf, JM, DSP; 1998-201) with a second 
adult there 30 Nov-7 Dec 1998 (DSP; MTH; 1998-207), an adult at the Whitewater 
R. delta, Salton Sea, RIV, 14 Feb-8 Mar 1998 (MAP; CAM, GMcC; 1998-049), an 
adult at Salton Sea State Recreation Area, RIV, 21 Jan-7 Mar 1998 (DSP; TC, KLGt, 
MTH, RL, CAM, GMcC, KM, MAP; 1998-025), an adult at Obsidian Butte, Salton 
Sea, IMP, 19 Jan 1998 (GMcC, PEL; 1998-039), a first-winter bird at Obsidian Butte, 
IMP, 24 Jan-8 Mar 1998 (MAP; KLG, MTHf, CAM, GMcC; 1998-042; Figure 5), 
and an adult at the Brawley landfill, IMP, 30 Dec 1998-1 Feb 1999 (SJT; 1999-048), 
considered probably different from one at the same locality 22-27 Jan 1996 
(McCaskie and San Miguel 1999). A photograph of the first San Jose bird was 
published in Field Notes 52:253. 


24 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 



Figure 7. Possible Blue-headed Vireo, Vireo solitarius, on San Nicolas Island, 
Ventura County, 20 Oct 1998 (1998-216), The extensively white-edged outer 
rectrices, blackish tertials with thick whitish edges, brilliant white throat, bluish head, 
and seemingly sharp demarcation between throat and auricular support identification 
as Blue-headed, but using known identification criteria most members could not rule 
out a bright Cassin’s Vireo. 

Photo by Sandra G. Harvill 


Several CBRC members expressed the concern that not all California birds may be 
of European/eastern North American origin (i,e., L. /. graellsii ) and that certain 
Asian taxa (e.g., heuglini, taymirensis) may be involved as well. In the end, the 
AOU’s (1998) regarding all of these taxa as subspecies of the Lesser Black-backed 
Gull rendered such concerns outside the scope of CBRC review. In the future, if the 
AOU defines a less inclusive Lesser Black-backed Gull, some or all of these records 
may be reconsidered. 

BRIDLED TERN Sterna anaethetus (1, 1). An adult at Bolsa Chica, ORA, 17 Jul 
1998 (MI; KSG, JEP; 1998-105; Figure 6) represents the first accepted record for the 
state; another record from Bolsa Chica 12 Jun 1993 (Erickson and Terrill 1996) has 
been resubmitted in light of the 1998 sighting. The Bridled Tern becomes one of five 
species on the state list supported by only sight records. 

After some initial confusion concerning potential differences between the west 
Mexican S. a. nelsoni and the Caribbean S. a. recognita/melanoptera (i.e., the bird 
had a pale collar and white on at least the two outer rectrices), the record was accepted 
unanimously on the second circulation. Ridgway (1919) said only that nelsoni is 
“similar to S. a. recognita but averaging larger, with relatively longer or more slender 
bill, and under parts of body tinged with pale gray. ” The tail pattern also eliminates the 
only other similar species, the Gray-backed Tern (S. lunata) of the central Pacific. 


25 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


SOOTY TERN Sterna fuscata (6, 0). A returning adult (cf. Rottenborn and Morlan 
2000) was at Bolsa Chica, ORA, 12-19 Apr 1998 (PEL; 1998-080). 

LONG-BILLED MURRELET Brachyramphus perdix (3, 1). Birds in “basic” 
plumage (first alternate?) at Pt. Saint George, DN, 21 Jul 1998 (ADB; 1998-1 18A), 
29 Jul 1998 (ADB; 1 998-1 18B), and 22-23 Aug 1998 (BED; 1998-118) were 
judged to be the same individual, contra N. Am. Birds 53:101. Rottenborn and 
Morlan (2000) reported the first accepted records of this species and summarized 
other records that remain unreviewed by the CBRC, including several specimens. 

RUDDY GROUND-DOVE Columbina talpacoti (69, 3). An immature male was 
found dead 22 km N of Blythe, RIV, 26 Nov 1992 (#LACM 107326; SC; KLG, MAP; 
1994-72), an immature male was at Furnace Creek Ranch, Death Valley, I NY, 7 Oct 
1998 (JHef; TH; 1998-212), and a male was at Independence, INY, 2 Nov 1998 
(RAHu; 1998-213). Patten took the Blythe specimen to Moore Laboratory of 
Zoology at Occidental College, there confirming its identity as C.t. eluta of western 
Mexico (the subspecies believed to account for all California records). Acceptance of 
the last bird was not unanimous, as Morlan felt that the description failed to rule out 
a Plain-breasted Ground-Dove (C. minuta), a southern species that seems very 
unlikely to reach California as a natural vagrant (its status in captivity is unknown). The 
larger size and black axillars of talpacoti are distinctive, as are purple markings on the 
wing coverts of minuta, 

GROOVE-BILLED ANI Crotophaga sulcirostris (11, 1). One at Desert Center, 
RIV, 4 Oct 1998 (MAP, BDS; 1998-155) provided the seventh fall record from the 
interior, conforming to the predominant pattern developing in California. 

BROAD-BILLED HUMMINGBIRD Cynanthus latirostris (54, 2). Lone females 
presumably wintering were at San Elijo Lagoon, SD, 5-8 Jan 1998 (MBS; MMR; 

1998- 027) and Goleta, SBA, 28 Nov 1998-28 Feb 1999 (RAH; GMcC, DR, SCR; 

1999- 062). 

RUBY-THROATED HUMMINGBIRD Archilochus colubris (4, 1). An immature 
female captured at SE Farallon L, SF, 25 Aug 1998 (PPf; SD, DH, MHf; 1999-008) 
fit the species’ incipient pattern of early fall vagrancy to California (the previous two 
fall records, also from this location, were on 21-22 Aug 1985 and 7 Sep 1994). 
Howell and Pyle (1997) discussed its separation from other small hummers, particu- 
larly the Black-chinned (A. alexandri ); Pyle (1997) authored the definitive treatment 
for this species in the hand (as all accepted California records have been). 

GREATER PEWEE Contopus pertinax (32, 1). An individual that wintered at 
Brock Research Center, IMP, 24 Dec 1998-27 Feb 1999 (MAPt; RF, KSG, KZKt, 
CAM, GMcC, JM, MMRt, MJSM, DMS; 1999-043) was identified as an adult 
because of its fresh, truncate rectrices and unworn tertials (the Greater Pewee is 
unique among North American Contopus in retaining the juvenal flight feathers 
during the first prebasic molt; Pyle 1997). A color photograph of this bird was 
published in N. Am. Birds 53:221. 

EASTERN WOOD-PEWEE Contopus virens (9, 5). A virtual invasion of this 
eastern flycatcher in spring/summer 1998 more than doubled the state’s accepted 
records, with individuals found at Bodega Head, SON, 3 Jun 1998 (ANW; PCo§, 
BDP; 1998-115). Pt. Reyes, MRN, 22 Jun 1998 (JM; TBf; 1998-004), Mono L. 
County Park, MNO, 4-8 Jul 1998 (SH|, RSt; MWE§, DL, DP§, JiP, DRt, AW; 1998- 
103), Galileo Hill Park, KER, 10 Jun 1998 (MTH; 1998-120), and South Fork 
Wildlife Area, KER, 27 Jun-9 Jul 1998 (SAL; MTH, MAP; 1998-102). With 
vocalizations critical to distinguishing this species from the Western Wood-Pewee (C. 
sordidulus), audio tape nicely supported the Bodega and Mono Lake records (photos 
of the latter bird supported the record but were not definitive). A photo of the Pt. 


26 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


Reyes bird showed a relatively pale breast and entirely yellow-orange lower mandible, 
marks highly suggestive of the species claimed, but probably more important to this 
record’s acceptance was an hour the finder spent listening to and describing the bird’s 
somewhat variable clearly whistled songs mixed with frequent call notes, none of 
which were like those of its western counterpart. The Galileo bird, a first for Kern 
County, gave only frequent “chip” call-notes with an occasional “Dusky-capped 
Flycatcher-like peeeeeah” (described as the typical Eastern Wood-Pewee song minus 
the final syllable). These vocalizations, combined with the bird’s appearance (the lower 
mandible was “all orange-yellow” and its “very crisp” plumage featured a “white 
throat, faint vest, and fairly light undertail coverts”) and the observer’s skill and 
experience levels, were considered adequate for CBRC endorsement. The bird at 
South Fork Wildlife Area was observed and heard singing by several observers. 

YELLOW-BELLIED FLYCATCHER Empidonax flaviventris (12, 1). A silent 
immature at California City, KER, 7 Sep 1998 (MTH; RC, MSM, MJSM, JCWt; 

1998- 124) was the fifth for Kern County, which now claims nearly half the state’s 
records. All California records but one (Carpinteria, SBA, 16-17 Oct 1987 ; Pyle and 
McCaskie 1992) are for September. Although sight records of this species are still 
somewhat contentious within the Committee because of the high potential for 
confusion with the Western Flycatcher complex ( E , difficilis/occidentalis) and the 
Acadian Flycatcher (E. uirescens ), this record received unanimous endorsement 
during the first round of voting. Photographs of this bird were not definitive, but they 
supported prolonged close observations by several experienced observers, and 
submission of the photos was considered important by at least one member. The 
spacing of the primaries on this bird was not inspected (cf. DeSante et al. 1985, 
Heindel and Pyle 1999), but independent review of museum specimens by Howell (in 
litt.) and Patten (in litt.) suggests that this feature’s utility may be limited because of 
overlap in the pattern of primary spacing of the Yellow-bellied and Western 
flycatchers. 

DUSKY-CAPPED FLYCATCHER Myiarchus tuherculifer (52, 7). Acceptance of 
one at Bolinas, MRN, 12 Jan-21 Mar 1998 (KH; PP; 1998-002) and one inland at 
Finney L„ IMP, 12 Jan-28 Mar 1998 (BDCW; KLGt, GMcC, MAP; 1998-051) 
yields a seasonal total of eight for 1997-98. In 1998-99, a returning bird was at 
Santa Cruz, SCZ, 19 Dec 1998 (SG; 1999-027, same as 1998-062), and new 
individuals were at Bodega Bay, SON, 22-27 Dec 1998 (BDP; JEPa, DnWN; 1999- 
005), nearby at Bodega Dunes Campground, SON, 23 Dec 1998-16 Jan 1999 
(DnWNf; BDP, JEPa; 1999-006), Los Osos, SLO, 12-19 Dec 1998 (TME; KH ; 

1999- 096), Ventura 28 Dec 1998-2 Jan 1999 (WW; DDesJt; 1999-029), and Long 
Beach, LA, 27 Dec 1998-28 Mar 1999 (KSG; RF, MJSM, MSMf, JM; 1999-016). 
Records of this southern flycatcher follow an upward trend and scatter fairly evenly 
northward along the coast to Sonoma County, with outliers farther north and in the 
interior. San Diego County claims only three records, suggesting that this corner of 
the state may lie southwest of the vector followed by most vagrants to California. All 
accepted records are thought to pertain to late fall migrants or, particularly, overwin- 
tering birds. 

GREAT CRESTED FLYCATCHER Myiarchus crinitus (42, 2). Individuals were at 
Twenty-nine Palms, SBE, 12 Sep 1998 (BiD, EACf; 1998-221) and in Manhattan 
Beach, LA, 18 Oct 1998 (KLf; KSG, TEW; 1998-208). The former record was the 
state’s third earliest and only the fifth inland. Despite being underexposed, photo- 
graphs helped this report gain unanimous acceptance since they showed important 
field marks, particularly pale edging to the inner tertials that appears thicker basally 
than expected on a Brown-crested Flycatcher (M. tyranrtulus ■ Heindel and Patten 
1996, Pyle 1997). 


27 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


SULPHUR-BELLIED FLYCATCHER Myiodynastes luteiventris (13, 1). 
California’s first spring record, at Gazos Creek, SM, 14 Jun 1998 (RST; BMcKt, 
AME; 1998-106), was the first of its kind for San Mateo County and the fourth for 
northern California. Photographs were inadequate to eliminate the similar Streaked 
Flycatcher (M. maculatus), but all three observers described the Sulphur-bellied’s 
distinctively thick blackish malar streak. Those with the best views also noted blackish 
streaks on a whitish background from the throat to the lower breast and flanks, 
unstreaked belly, undertail coverts washed with yellow, lack of yellow in the face, and 
only a very limited amount of pale on the lower mandible (cf . Howell and Webb 1995, 
Pyle 1997). A report of this bird continuing on 15 Jun 1998 received no CBRC 
support, as members were not convinced that the description ruled out a female 
Black-headed Grosbeak (Pheucticus melonocephalus). 

THICK-BILLED KINGBIRD Tyrannus crassirostris (14, 1). One at Pomona, LA, 
14 Oct 1998-1 Mar 1999 (MJSM; 1998-175) returned for its seventh winter at this 
location, while an immature at Half Moon Bay, SM, 19 Dec 1998-7 Mar 1999 
(AWK; GEC, MD, MWEt, RF, NL, LML f, MMa, JM, CL, DvWN, MMRf; 1998-233) 
was a first for San Mateo County and one of the northernmost records for this species. 
The age of the latter bird was inferred from the rufous edgings to the rectrices and 
wing coverts, fresh tertials, and bright yellow underparts (cf. Pyle 1997). 

WHITE -EYED VIREO Vireo griseus (38, 2). A singing male at Butano Creek, SM, 
18 Jun 1998 (DLSu; 1999-026) furnished a county first, while another at Pt. Saint 
George, DN, 27 Jun 1998 (ADB; 1998-141) was well north of the state’s previous 
northernmost records (in Marin County). Approximately 75% of accepted records are 
from spring (8 May to late June). 

YELLOW-THROATED VIREO Vireo flavifrons (69, 2), One at Pt. Reyes, MRN, 
13 Jun 1998 (AME; 1998-108) was in spring, season of approximately 70% of 
California’s records. The state’s third in winter was at Westminster, ORA, 29 Dec 
1998-15 Feb 1999 (KSG, SSo, RF, JM, GMcC, MTH, DR; 1999-018). 

BLUE-HEADED VIREO Vireo sohtarius (5**, 5). This recently recognized species 
was added to the review list in 1998 (records from 1997 and later), with the following 
records being the first to gain formal CBRC acceptance: likely immature female on SE 
Farallon I., SF, 1 Oct 1998 (PP; 1999-011), immature (likely male) banded on SE 
Farallon I., SF, 1 1 Oct 1998 (WR, PPt; 1999-012), immature male banded at the Big 
Sur R. mouth, MTY, 28 Sep 1998 (JBot; 1999-057), one at Arroyo Grande Creek, 
SLO, 14 Oct 1998 (BS; 1998-157), and a wintering bird in Orange, ORA, 31 Oct 
1998-16 Jan 1999 (DRW; 1999-095). At least some pre-1997 records will undergo 
a form of review, and anyone who has observed this species in the state in any year 
is urged to submit contemporaneous documentation to the Committee. 

Blue-headed Vireos are difficult to separate from bright Cassin’s Vireos (V cassinii) 
under field conditions (Heindel 1996, Pyle 1997), and evaluation of many records is 
troublesome, even with a fairly detailed description and/or photographs (see Fig- 
ure 7). Accepted birds were described as possessing or photographed exhibiting (1) an 
immaculate white throat sharply demarcated from the auriculars (no blending at the 
border), (2) white center of breast and belly, (3) blue or blue-gray head contrasting with 
a green back (some with greenish napes), and (4) lemon-washed flanks (often very 
bright). One bird in hand was reported to have outer rectrices with pale outer edges 
approximately 1 mm thick. Until identification criteria are better delineated, the 
Committee requests unusually detailed and precise descriptions of suspected Blue- 
headed Vireos. 

‘PHILADELPHIA VIREO Vireo philadelphicus (116, 8). Individuals found within 
the typical autumn window of occurrence (mid September to late October) were at 
Fairhaven, HUM, 2 Oct 1998 (BED, DRi; RF, JM, MMR, GMcC, MSM; 1998-152), 


28 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


Pt. Reyes, MRN, 30 Sep 1998 (PU; JFH, DGS; 1998-189), El Granada, SM, 4 Oct 
1998 (RST; 1999-046), Big Sur R. mouth, MTY, 27 Oct 1998 (JBo; 1999-056; 
banded immature), Los Angeles R. near Glendale, LA, 26 Sep-11 Oct 1998 (KLG; 
1999-037), and Huntington Beach, ORA, 12-22 Oct 1998 (KSG, MSM; 1998- 
199). One at Bodega Head, SON, 5-10 Sep 1998 (DnWN; KCK, BDP, DGS; 1998- 
134) arrived a day later than the state’s earliest fall record. A singing male at the 
Parker Creek Diversion Dam, MNO, 21 Jun 1998 (ES; DMaf, DP§, JiP; 1998-104) 
furnished California’s eleventh in spring, with Mono County accounting for six of 
these records (the previous five occurred at Oasis). Although photographs of this bird 
appear to be diagnostic, and the record was accepted unanimously, a recorded song 
included phrases that some members considered similar to those of a Warbling Vireo 
(V gilvus ) and thus seemingly atypical for a Philadelphia Vireo. 

YELLOW-GREEN VIREO Vireo flavouiridis (63, 8). This vireo’s impressive 
showing in fall 1998 matches the previous seasonal high set in fall 1988; as in that 
year, seven occurred along the coast. Individuals were at Fort Funston, SF, 30 Sep-4 
Oct 1998 (LC, MWE, MJM, JM; 1998-151), Lighthouse Field State Beach, SCZ, 15 
Oct 1998 (SG; 1998-220), Oso Flaco L., SLO, 28 Sep 1998 (BS; 1998-145), 
Oxnard Plain, VEN 13-16 Sep 1998 (GMcC, MSM; 1998-139), Oxnard Plain, VEN, 
26 Sep-4 Oct 1998 (SJT; JLD; 1998-229), Oxnard Plain, VEN, 29-30 Sep 1998 
(MSM; JLD; DSP; 1998-154), and Pt. Loma, SD, 17 Oct 1998 (GMcC; GLR; 1998- 
165). Few of these birds were aged specifically, but none was thought to be an adult. 
An immature at Galileo Hill Park, KER, 6 Oct 1998 (MTH; 1998-219) provided the 
second record for Kern County and only the state’s fourth inland. 

VEERY Catharus juscescens (10, 1). An immature banded at the Big Sur R. 
mouth, MTY, 21 Sep 1998 (JBof; 1998-174) made the state’s seventh fall record. 
This bird’s appearance was consistent with C. /. salicicola. 

GRAY-CHEEKED THRUSH Catharus minimus (20, 1). An immature at Galileo 
Hill Park, KER, 9 Oct 1998 (MTH; DSP, JCW; 1998-184; Figure 8) was the second 
identified inland in California, the first having been at the same park 14-18 Sep 1989 
(Patten and Erickson 1994). 

WOOD THRUSH Hylocichla mustelina (14, 1). An immature at California City, 
KER, 11-13 Oct 1998 (KSG, MTH, GMcC, MAP, MJSM, SBTt, JCW ; 1998-160) 
was a first for Kern County and the state’s fifth Wood Thrush inland. 

*GRAY CATBIRD Dumetella carolinensis (101, 16). After a single spring record 
in 1998 — of a one-year-old bird banded near the Carmel R. mouth, MTY, 20 Jun 
1998 (KN; 1998-109) — Gray Catbirds arrived in bulk in fall 1998; the final tally of 15 
records eclipsed the previous high seasonal total of seven from fall 1991 . Remarkably, 
seven of these were inland, including four in Kern County (just one previous fall 
occurrence there). Individuals were at Chico, BUT, 7-8 Oct 1998 (RW; HO, JOs, 
RER, MSk: 1998-167), Bodega Head, SON, 22 Sep-8 Oct 1998 (DnWN; DGS; 
1998-188), Pt. Reyes, MRN, 22-23 Sep 1998 (KB, JFH, DGS; 1998-156), SE 
Farallon I., SF, 31 Oct 1998 (WR; ISt; 1999-033; immature; Figure 9), Tunitas 
Creek, SM, 23 Oct 1998 (AW; 1998-168), near Ano Nuevo State Reserve, SM, found 
dead 30 Dec 1998, likely wintering (BMcKf; #LACM 111151; 1999-070), Galileo 
Hill Park, KER, 14 Oct 1998 (KSG; 1999-034), near Cantil, KER, 23 Oct 1998 
(MTH; 1998-21 8), California City, KER, 18 Oct-31 Dec 1998 (JHu; MTH, MAPf, 
DMS, GMcC, MSM, MJSM, JCWf; 1998-171; adult), California City, KER, 20 Oct- 
16 Nov 1998 (JHu; MTH, GMcC, MSM; 1998-180; immature), Iron Mt. Pumping 
Plant, SBE, 24 Oct 1998 (MAP; BiD; 1998-169; immature), Horsethief Springs, 
Kingston Range, SBE, 6 Oct 1998 (BiD; 1998-122), McGrath State Beach, VEN, 19 
Sep 1998 (NF; 1999-040), San Nicolas L, VEN, 1 1-17 Oct 1998 (RAH; WW; 1999- 
045; banded), and San Diego, SD, 27-28 Oct 1998 (GLR; GMcC; 1998-170). The 


29 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


»j-«k ) («w a WU~,» 



4 T- V 


&ft^-cAetlcf0{ s TV*i4 
? (Jcf r* 

H*tf t &r*t C+ y Ca. 

Figure 8. Gray-cheeked Thrush, Catharus minimus, at Galileo Hill Park, Kern 
County, 9 Oct 1998 (1998-184). Key features include dominance of cool gray tones 
to the upperparts, brownish-black spotting on white underparts, whitish border to the 
lower auriculars, and incomplete whitish eye-ring. 

Sketch by John C. Wilson 


number of accepted records indicates that the Gray Catbird is a rare, regular 
component of California’s avifauna, and records after 1999 will not be reviewed. 

WHITE/BLACK-BACKED WAGTAIL Motacilla ctlba/lugens (5, 3). In 1993, 
Howell and David Sibley were asked independently to review five vexing northern 
California records (1988-290, 1989-130, 1989-210, 1990-189, 1990-200); we 
thank them for their effort and advice, provided both in letters to the Committee and 
via publication of Sibley and Howell (1998), which included their updated opinions on 
all California records. The three records discussed here (among the five reviewed by 
Howell and Sibley) were originally considered by the CBRC in the early 1990s, then 
re-reviewed following publication of Sibley and Howell’s enlightening paper. Unfortu- 
nately, members were still unable to reach consensus on the age and sex of these 
birds, a step critical to identifying them to species (Pyle 1997). For two of these 
records, pitfalls inherent in the identification process were compounded by differing 
opinions regarding the likelihood that each pertained to a single bird that wintered in 
the same general area for four consecutive years. Finally, given that these wagtails are 
known to interbreed, at least occasionally (e.g., Kishchinski and Lobkov 1979, 
Badyaev et al. 1996), the potential for hybridization naturally affected members’ 
opinions (although Sibley and Howell suspected that “fewer than 5%” of the 216 
wagtail specimens they inspected might have been hybrids). Though records such as 
these frustrate the field ornithologist (and bird records committees), they are instruc- 
tive reminders that the natural world does not always conform neatly to our organizing 


30 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 



Figure 9. Gray Catbird, Dumetella carolinensis, on Southeast Farallon Island, San 
Francisco County, 31 Oct 1998 (1999-033), one of a record 16 accepted records in 
1998. The four inner greater coverts have been replaced and thus contrast with the 
duller outer greater coverts and primary coverts, a “molt limit” not shown by fall 
adults. Note also the lackluster remiges, typical of immatures. 

Photo by Ivan Samuels 


schemes and that some birds may simply not be identifiable to species, even by skilled 
observers studying them at close range. 

An early fall vagrant at the Eel River Wildlife Area, HUM, 1-3 Sep 1994 (NL, 
DEQt, DSa; 1994-133A) was submitted as a White Wagtail and received four 
“accept" votes, along with six votes for acceptance at the species-pair level. Upon its 
1998 re-submission the voting results after two rounds were identical to the original 
results (four votes for the White Wagtail, six votes for the species pair). Most members 
were nearly convinced that the bird was, in fact, an adult White Wagtail, but some 
were troubled that a well-regarded observer (one of many who observed this bird 
without submitting documentation) verbally reported having observed dark scapular 
lines not visible in two distant photographs. Sibley and Howell (1998) identified this 
individual only to the species-pair level. 

A widely seen and closely studied bird at Moss Landing, MTY, 21 Dec 1990-21 
Jan 1991 (DEG; JLD, SNGH, GMcC, PEL, RJO’Bt, MAP, DRt, DAS, SWe; 1990- 
200A; photograph in Am. Birds 45:317) was generally thought to be an adult female 
White or immature Black-backed Wagtail. The record was twice submitted as a 
“White/Black-backed Wagtail,” receiving unanimous support to the species-pair level 
each time. After the first review (as 1990-200), Morlan accepted it as a Black-backed, 
while Roberson and Pyle accepted it as a White that had wintered in the general area 
the preceding three years (Moss Landing, MTY, 23 Dec 1988-21 Jan 1989 [1988- 
290j; Sunset State Beach, SCZ, 3-11 Dec 1989 [1989-210]; Pajaro R. mouth/ 
Sunset State Beach SCZ/MTY 7 Nov-3 Dec 1990 [1989-210)). In the second review 
(as 1990-200A), Pyle’s opinion remained unchanged, while the remaining votes went 


31 



REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


to the species pair (Roberson was then off the Committee). Lars Svensson kindly 
reviewed a photograph of this bird but was unable to state anything conclusive 
regarding the appearance of birds of the subspecies ocularis of NE Siberia and NW 
Alaska (i.e., those known to reach California). Sibley and Howell (1998) identified this 
bird only to the species-pair level, opining that it was probably an adult female White 
Wagtail, although probably not the same individual involved in previous records from 
the general vicinity (because the appearance changed too rapidly from that ascribed to 
1990-189A, discussed in the following paragraph; S. Howell pers. comm.). 

One at the Pajaro R. mouth/Sunset State Beach, MTY/SCZ, 7 Nov-3 Dec 1990 
(BMe; JSM, JMS; 1990-189A) was seen by fewer observers and supported by much 
sparser details than were the preceding two records. It was first submitted as a White 
Wagtail and received three “accept” votes (all members accepting the species pair). In 
1993, commenting at the CBRC’s request, Sibley (in litt.) regarded the descriptions of 
this bird as “just too little to go on,” while noting that they suggested an immature 
White Wagtail; Howell (in litt.) simply considered the documentation insufficient to 
support a species-level identification. Sibley and Howell (1998) erroneously published 
this record as an immature White Wagtail (S. Howell in litt.). Upon its resubmission in 
1998, four members accepted it as a White Wagtail, with all members accepting the 
species pair. Among those accepting, Jaramillo, McCaskie, and Pyle reckoned it was 
likely a returning bird; Rottenborn, following Sibley and Howell (1998), considered it 
more likely an immature. 

OLIVE-BACKED PIPIT Anthus hodgsoni (1, 1). A first-fall bird observed and 
banded at SE Farallon I., SF, 26-29 Sep 1998 (RB, PC, WR; DMaf, PPf; 1999-010) 
marked the debut of this distinctive Old World species in California. This individual 
belonged the “expected” subspecies yunnanensis of northern Eurasia, which has 
accounted for all vagrant Olive-backed Pipits of known race in the New World. 
Capitolo et al. (2000) presented a full account of this record, including photographs 
and a summary of the North American records. 

SPRAGUE’S PIPIT Anthus spragueii (27, 1). Imperial County’s second record, 
comprising at least 12 wintering near Calipatria, IMP, 10 Jan-21 Feb 1998 (KLG, 
MTH, PEL, RL, CAM, GMcC, TRC, MAP; 1998-040), exceeded the largest flocks 
previously recorded in California (flocks of up to five near Needles, 2-27 Nov 1986, 
Langham 1991, and near Lancaster, LA, 22 Nov 1981-7 Mar 1982, Binford 1985). 
The birds were in a field of cut Bermuda Grass ( Cynodon dactylon) near the 
intersection of Sinclair Road and Highway 111. 

BLUE-WINGED WARBLER Vermivora pinus (26, 1). An immature male at 
Birchim Canyon, NW of Bishop, 1NY, 6 Sep 1998 (DP, JiP; JHe, TH; 1998-214) was 
the second recorded in Inyo County. This eastern warbler’s pattern of vagrancy is 
atypical in that inland records account for most of the state total (61%); perhaps even 
more unusual, 58% of the fall records are away from the coast. 

GOLDEN-WINGED WARBLER Vermivora chrysoptera (60, 1). A male at SE 
Farallon I., SF, 29-30 Sep 1998 (IS; RB, PP, WR; 1999-013) was the island’s third 
fall record and seventh overall. 

YELLOW-THROATED WARBLER Dendroica dominica (85, 3). One at the 
Carmel R mouth, MTY, 9 Sep 1998 (DR; 1998-142) was judged to be of the more 
regularly occurring subspecies albilora, while an immature male at San Nicolas I., 
VEN, 7-14 Sep 1998 (RAH; SGH; 1998-193) with some yellow in the lores was not 
ascribed a race (some albilora can show this). A bird presumed to be wintering was in 
Eureka, HUM, 14 Jan 1998 (MHM; 1998-022). 

GRACE’S WARBLER Dendroica graciae (34, 1). The state’s northernmost coastal 
record was at Jacks' Peak Regional Park, MTY, 12 Dec 1998-3 Feb 1999 (RFT; JA, 
DPH, CH, DR, MMR; 1999-004). 


32 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


PINE WARBLER Dendroica pinus (56, 1). A male that wintered at El Dorado Park, 
Long Beach, LA, 25 Nov 1998-10 Apr 1999 (KSG, CAM, MSM, MJSM, JM, SSo ; 
1998-202) began singing by the end of its stay. This bird was judged to be probably 
different from 1997-193, a likely immature male that wintered in a different part of 
this large park in 1997-98. Although this species winters in the state with some 
regularity, no individual is believed to have returned for a second year. One reported 
in Fountain Valley, ORA, 17-18 Nov 1998 (N. Am. Birds 53:106) was not submitted 
for review. 

WORM-EATING WARBLER Helmitheros uermiuorus (84, 1). One at San Pedro, 
LA, 23 Nov-14 Dec 1996 (DMH; KLG; 1997-025) required three circulations 
because a submitted photograph was not identifiable and the only submitted descrip- 
tion covered just the bird’s call notes and behaviors, plus a sketch depicting the 
undertail coverts. Additional documentation was eventually submitted and the record 
was accepted unanimously, but other widely seen birds routinely fare poorly in CBRC 
review for similar reasons (cf. the account for this species under Records Not 
Accepted). One reported at Moss Landing, MTY, 16 Nov 1998 ( N . Am. Birds 
53:102) was not submitted for review. 

MOURNING WARBLER Oporornis Philadelphia (106, 3). Accepted were an 
immature male banded at SE Farallon L, SF, 15-16 Sep 1998 (PP; WR|; 1999-014), 
an immature at Chorro Creek, Morro Bay, SLO, 20 Sep 1997 (JSR; WB- 1998-038), 
and an immature banded at San Nicolas I., VEN, 12-13 Sep 1998 (RAH; 1998-136). 
The first two birds were described as showing the distinctly yellow throats typical of 
this species, while the last had a throat described in the hand as being “pale gray with 
pale yellow tips; the yellow coloration was quite muted and would have been difficult 
to discern in the field.” While members were somewhat concerned with this descrip- 
tion, all distinctive measurements, appearances, and the call were consistent with a 
Mourning Warbler and eliminated congeners. In-hand photographs of this interesting 
bird were, unfortunately, destroyed by the developer. 

RED-FACED WARBLER Cardellina rubrifrorts (11, 1). One photographed at 
Bishop, INY, 20-21 May 1998 (BT; NBB, KSG, JHet, TH, DP, JiP; 1998-085) 
constituted the state’s eighth spring record, all but one at inland locales. This bird’s 
relatively dull plumage suggested perhaps a one-year-old female, but Pyle (1997) 
urged caution in assessing age and sex in this species. 

SCARLET TANAGER Piranga oliuacea. (99, 5). A female was photographed at 
Pt. Reyes, MRN, 1 Jun 1998 (RS; GGrt; 1998-096). A male observed there on 30 
Sep 1998 (RS; JFH, DGS, SCH, PU; 1998-158) was described by all viewers as 
having entirely black remiges and wing coverts, indicating an adult (it was termed a 
“young male” in N. Am. Birds 53:102). Additional fall males were at Galileo Hill 
Park, KER, 1 Oct 1998 (DSP; 1998-183), Chatsworth, LA, 27 Oct 1998 (JWSt; 

1998- 190; immature), and Huntington Beach, ORA, 31 Oct-1 Nov 1998 (KSG; 

1999- 113; immature). 

CASSIN’S SPARROW Aimophila cassinii. (40, 1). A male was voice-recorded as 
it sang and skylarked in the North Domenegoni Hills, RIV, 26 May 1995 (KFC§; 
1999-198). Whereas the species’ diagnostic primary song was captured on tape, 
several members noted that a second song (presumed to be from the same bird) 
sounded atypical for a Cassin’s Sparrow. Spring vagrants in May and June account for 
most California records. 

FIELD SPARROW Spizella pusilla (6, 1). Kern County’s first was photographed at 
Inyokern, KER, 1-4 Nov 1998 (SStf; MTH; 1998-205). Like previous California 
Field Sparrows’, this bird’s appearance was consistent with the pale western subspe- 
cies arenacea. 


33 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


SNOW BUNTING Plectrophenax nivalis (64, 2). One was at Areata Marsh, 
HUM, 28 Oct 1998 (PL; 1999-055), and a female was photographed on SE Farallon 
I., SF, 1 Nov 1998 (PPt; 1999-035). Each fit the predominant pattern of late fall 
records from the northern coast. 

PAINTED BUNTING Passerina ciris (65, 7). First-fall birds on the Shasta R. NE of 
Grenada, SIS, 19 Sep 1998 (RE; 1999-036) and at Pt. Reyes, MRN, 29 Sep 1998 
(GE; JMR; 1998-172) were county firsts. Other immatures were at Furnace Creek 
Ranch, Death Valley, INY, 24 Sep^lO Oct 1997 (RJN; JHe, TH, GMcC, MAP; 
1997-146, 1997-146A) and 21 Sep 1998 (GMcC; MSM; 1998-144), at Ridgecrest, 
KER, 24-27 Sep 1998 (LSu ; 1998-210), and near Cantil, KER, 6-7 Sep 1998 
(MTH; 1998-217). These records fit the predominant pattern of Painted Bunting 
records in California, of immatures occurring from late August through mid October. 

An adult male at Twentynine Palms, SBE, 8 Oct 1998 (BiD; 1998-223) was 
accepted 8-2 as a natural vagrant in the first round on the strength of its relatively 
remote desert location and appropriate date. While the Committee shares a general 
uneasiness that released/escaped birds may account for a considerable proportion of 
the state’s adult male Painted Bunting records (see this species’ account under 
Records Not Accepted, Natural Occurrence Questionable), this vote demonstrates 
that those concerns can be overcome. 

COMMON GRACKLE Quiscalus quiscula (44, 5). Records were of a weakened 
male taken as a specimen from Rohnert Park, SON, 15 Nov 1997 (SEt; 1999-101; 
^Sonoma State University 1985), a spring migrant photographed at Pt. Sur, MTY, 21 
Apr 1998 (JBo; DRf; 1998-089), a male photographed and voice recorded at 
Bishop, INY, 13 Sep 1998 (DPt§, JiP; 1998-211), a singing male videotaped in 
Wildomar, RIV, 21 Jan-15 Feb 1998 (CHaf; TRC, GMcC, CAM, MAP; 1998-041), 
and a singing male voice recorded at Twentynine Palms, SBE, 27 Mar-11 Apr 1998 
(MF, MAP, DGS, JOZ§; 1998-061). Like the previous California records, all birds 
identifiable to subspecies were of the expected Bronzed race versicolor. The Rohnert 
Park and Pt. Sur birds were county firsts. 

Through spring 1988, 15 of 22 accepted records were of apparent spring migrants 
occurring between 12 April and 12 June. From fall 1988 through fall 1998, only four 
of 22 accepted records fell within this spring window. During the latter period, 15 
records pertained to apparent fall migrants. Inyo County claims a remarkable 18 
records, more than triple the total for any other county. 

RECORDS NOT ACCEPTED, identification not established 

LEAST GREBE Tachybaptus dominicus. The description of a bird in Golden Gate 
Park, SF, 22 Jan 1998 (1998-043) was insufficient to establish its identity. 

STREAKED SHEARWATER Calonectris leucomelas. One reported on Monterey 
Bay, MTY, 6 Dec 1998 (1999-003) received no support. 

MANX SHEARWATER Puffin us puffinus. The documentation for a small black 
and white shearwater S of Santa Cruz I., SBA, 20 Sep 1998 (1998-143) was 
considered inadequate to support the record; the species remains unconfirmed in 
California waters south of Morro Bay or north of Bodega Bay. 

RED-TAILED TROPICBIRD Phaethon rubricauda. The report of one seen briefly 
from shore off south Vandenberg Air Force Base, SBA, 10 September 1998 (1998- 
038) received limited support through two circulations. 

BLUE-FOOTED BOOBY Sula nebouxii. A bird reported at Battery East, SF, 26 
Sep 1998 (1999-075) was not identified by all observers present. Several Committee 
members accepted that a booby was seen, but all but one agreed with those observers 
who considered the identification only probable. 


34 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


LESSER WHITE-FRONTED GOOSE Anser erpthropus. The report of one at Tule 
Lake N.W.R., SIS, 15 Mar 1998 (1998-057) received but a single vote to accept. 

TRUMPETER SWAN Cpgnus buccinator. The Committee’s traditional tough 
stand on this species continued, as evidenced by the following records not accepted: 
adult near Lower Klamath N.W.R., SIS, 14 Feb 1998 (1998-194), adult at Ash Creek 
W.M.A., LAS, 15 Feb 1998 (1998-195), adult at Cosumnes R. Preserve, SAC, 16 
Nov 1998 (1998-227), one heard only N of Marysville, YUB, 16 Dec 1998 (1998- 
228), and two adults and three imrnatures on Clair Engle L., TRI, 29 Dec 1998 
(1999-023). Only the last record received any accept votes. We appreciate the efforts 
of the Trumpeter Swan Society to compile reports of this species in California, but 
records lacking adequate documentation will ultimately fail to receive the Committee’s 
blessing. 

FALCATED DUCK Anas falcata. The report of a male seen briefly at Gray Lodge 
W.M.A., BUT, 15 Feb 1998 (1998-046) was interesting but received little support 
from Committee members. 

AMERICAN BLACK DUCK Anas rubripes. Votes on a record of a bird seen briefly 
at Cosumnes R. Preserve, SAC, 14 Mar 1997 (1997-069) were initially split almost 
evenly among “accepted," “not accepted — natural occurrence questionable,” and 
“not accepted — identification not established.” Votes for the last option dominated in 
the third and final circulation. 

KING EIDER Somateria spectabilis. One at Pt. Reyes, MRN, 29 Jan-16 Feb 
1997 (1997-061) received a split vote on its fourth and final circulation. Most 
Committee members agreed that this bird (seen by many, including at least one CBRC 
member) was probably identified correctly, but the documentation was considered 
inadequate. 

* ZONE-TAILED HAWK Buteo albonotatus. Individuals at Carlsbad, SD, 20 Nov 
1997 (1998-031) and 4.2 miles SE of Onyx Summit, SBE, 30 Aug 1998 (1998-124) 
were not documented to the Committee’s satisfaction. 

YELLOW RAIL Coturnicops noveboracensis. One seen briefly, and only in flight, 
at Fort Bragg, MEN, 5 Oct 1995 (1998-132) received majority acceptance initially 
but only a minority on the second and final round. 

GRAY-TAILED TATTLER Heteroscelus brevipes. A report of one at Bodega 
Head, SON, 30 May 1998 (1998-093) received no support, whereas one heard, and 
seen as a silhouette only, at Princeton Harbor, SM, 6 Jun 1998 (1998-164) received 
three and two “accept” votes on two rounds. Another report from Pt. Reyes during 
the same period is still under CBRC review. 

BRISTLE-THIGHED CURLEW Numenius tahitiensis. Reports of two birds at 
Kehoe Beach, Pt. Reyes, MRN, 6 May 1998 (1998-090) and one at Big Lagoon, 
HUM, 9 May 1998 (1998-081) corresponded with the 1998 landfall (see accepted 
records) but were not accepted by simple majorities on a single circulation. The Marin 
birds were identified about ten days after the sighting, and the Humboldt bird was seen 
only in flight. 

RED-NECKED STINT Calidris ruficoilis. The identity of a male stint in presumed 
first alternate plumage collected at the mouth of the Alamo R., Salton Sea, IMP, 17 
Aug 1974 (#SDNHM 38887; GMcC, JBut; 1984-085) remains unresolved. The 
record was published as a Red-necked Stint by McCaskie (1975), Roberson (1980), 
Garrett and Dunn (1981), AOU (1983), Veit (1988), and Small (1994), but the CBRC 
has never accepted it; since receiving a simple majority vote on its first circulation in 
1985, the record never reached even a split vote on four subsequent rounds. All have 
agreed that the bird was either a Red-necked or a Little Stint, but most have been 


35 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


unwilling to go beyond that. In 1993, the specimen was sent to the National Museum 
of Natural History where M. Ralph Browning and the late Claudia Wilds compared it 
with approximately 100 specimens of Red-necked and Little stints but were unable to 
identify it. Nevertheless, the record did receive three accept votes on its final circulation, 
and many members believe the specimen may yet be identified. The Committee would 
welcome the chance to review any significant reanalysis of the specimen. 

LITTLE STINT Calidris minuta. A reported juvenile at Eureka, HUM, 16-21 Sep 
1992 (1993-050, 1993-050A) received three to six “accept” votes during five 
circulations (the review process was restarted following receipt of better photographs), 
and most Committee members agreed that the bird may have been a Little Stint. In 
contrast, a bird reported from Abbott’s Lagoon, Pt. Reyes, MRN, 9 Aug 1998 (1998- 
125) received no support. 

WHITE-RUMPED SANDPIPER Calidris fuscicollis. One reported from the Santa 
Maria R. estuary, SBA, 6 Sep 1997 (1998-032) received but a single vote to accept 
on its second circulation. The observers failed to note a white rump, and the plumage 
described matched neither an adult or a (very early) juvenile. 

JACK SNIPE Lymnocryptes minimus. A report from Kern N.W.R., KER, 31 Dec 
1998 (1999-002) received no CBRC support. 

ICELAND GULL Larus glaucoides. Records not accepted included an adult at 
Areata, HUM, 6-23 Feb 1987 (1987-072), two juveniles at Alviso, SCL, 16-28 Jan 
1998 (1998-091), a juvenile at Moss Landing, MTY, 15-20 Feb 1998 (1998-059), 
a juvenile near Milpitas, SCL, 24 Feb 1998 (1998-107), and a second-winter bird at 
Ano Nuevo Pt., SM, 27 Feb 1998 (1998-056). The first bird resembled a typical 
Kumlien’s Gull (L. g. kumlieni ) but was not photographed and received only five or six 
votes to accept on each of four circulations. The other records scored more poorly, 
with only one to three votes to accept per record per circulation, except for six votes 
for the Alviso birds in the initial round. Rottenborn and Morlan (2000) provided more 
information on records of this species not accepted by the CBRC; Terrill et al. (1999) 
discussed the situation in the San Jose area in more detail. Howell (2000; in votes) 
emphasized another potential confounding factor in review of potential Iceland Gulls: 
hybrid Glaucous-winged (L. glaucescens) x Herring (L. argentatus) Gulls. 

ROSS’S GULL Rhodostethia rosea. An adult reported from Berkeley, ALA, 14 
Jan 1998 (1998-100) was described with too little detail for a first state record. 

RUDDY GROUND-DOVE Columbina talpacoti. Most members felt that a bird at 
Twentynine Palms, SBE, 24 Oct 1998 (1998-225) was likely this species, but by the 
third round of voting only four members considered the description adequate to 
support the record. Several comments referenced the Plain-breasted Ground-Dove 
(C. minuta), a situation discussed under Records Accepted. 

RUBY-THROATED HUMMINGBIRD Archilochus colubris. A female reported at 
Pt. Reyes, MRN, 26 Aug 1998 (1999-054) received no votes for acceptance after two 
voting rounds. In light of cautionary treatment by Pyle (1997), and considering other 
tantalizing encounters with Archilochus hummingbirds in the state that ultimately 
proved to involve likely or definite Black-chinneds, most members indicated unwilling- 
ness to accept sight records of this species without definitive observations of the 
shapes of the primaries (cf. Howell and Webb 1995, National Geographic Society 
1999, Sibley 2000). We urge observers claiming vagrant Ruby-throated Humming- 
birds to obtain definitive photographs (including the primaries) and/or in-hand data. 

GREATER PEWEE Contopus pertinax. Nine positive votes cast in the first round 
for a bird photographed at SE Farallon I., SF, 1 June 1998 (1998-116) would 
normally have constituted Committee endorsement of the state’s first spring record of 


36 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


a Greater Pewee, but Heindel requested recirculation to further consider a fairly bold 
breast pattern, relatively bold wing-bars, missing greater wing-coverts, and fresh 
tertials contrasting with wings that otherwise looked worn to most members {Figure 
10). Howell and Jaramillo, the members most familiar with the Greater Pewee and 
potentially confusing tropical species, considered this record to be fairly straightfor- 
ward; nonetheless, a majority voted not to accept in the third and fourth rounds, 
believing that the bird looked atypical (at least in photographs) and was inadequately 
supported by written details. Morlan and Patten opined that the bird may have been 
an Eastern Wood-Pewee (C. virens), partly because of an unprecedented incursion of 
that species to California in spring 1998. Given that the bird was seen at close range 
over the course of a day by at least two observers {only one of whom submitted a 
report), questions raised by the equivocal photographs might have been resolved by 
recording more detailed observations of the bird’s appearance and size. Comments on 
Figure 10 are welcome. 

EASTERN WOOD-PEWEE Contopus virens. A single observer briefly described a 
singing Eastern Wood-Pewee at Pt, Reyes, MRN, 1 Jun 1998 {1998-196). Though 
the distinctive “pee-a-wee” song was mentioned, the rest of the description was too 
superficial, with the problematic claim that the “head, upperparts, and tail were all 
brown.” The record therefore received no Committee support, although it could be 
revived with submission of corroborating information from others reported to have 
seen this bird. Also from the same observer were reports of a vocalizing bird at 
Greenhorn Summit, KER, 1 Aug 1998 (1998-197) and another bird at Pt, Reyes, 
MRN, 24 Sep 1998 (1998-198). Lacking convincing detail, these reports received no 
support. 

ALDER FLYCATCHER Empidonax alnorum. A widely studied immature 
Empidonax at Galileo Hill Park, KER, 7-11 October 1998 (1998-161; Figures 11 
and 12) may well have been an Alder Flycatcher, but whereas the calls (most 
commonly “a bunting-like pit” but with occasional “soft whits”) were atypical for 
western Willow Flycatchers (E. traillii brewsteri, E. t. adastus, E. t. extimus), they 
fell short of the more powerful “peep” or “beck” calls normally heard from Alder 
Flycatchers. After examining slides of 1998-161 with a large number of specimens of 
both the Willow and Alder at hand, Philip Unitt concluded that the western subspecies 
of Willow Rycatcher were eliminated on the basis of tertial pattern and that the bird 
was likely an Alder Rycatcher. Without specimens of first-fall eastern Willow Rycatch- 
ers (£. t. traillii) for comparison, however (juveniles certainly identified as E. t. traillii 
are almost lacking in collections), he was unwilling to state with certainty that the bird 
was an Alder Flycatcher. The record was defeated 1-9 after two rounds of voting, with 
most members opining that they would likely have voted differently if the bird had 
sounded like a typical Alder Flycatcher. Two singing birds in spring comprise 
California’s only accepted records. 

BLUE-HEADED VIREO Vireo solitarius. Individuals were not accepted from 
Mono L. County Park, MNO. 5 Sep 1995 (1999-128), Eureka, HUM, 19 Dec 1998 
(1999-116), Pt. Reyes, MRN, 19 Sep 1998 (1998-187), LosOsos, SLO, 15-22 Sep 
1998 (1999-071), San Nicolas I., VEN, 20 Oct 1998 (1998-216; Figure 7), and 
Pilarcitos Creek, SM, 1 Nov 1998 (1999-047). As might be expected, reports of this 
taxon have surged following the three-way split of the Solitary Vireo complex, but 
identification of this vireo is more problematic than many observers seem to 
appreciate, and the CBRC is unlikely to endorse records of incompletely or ambigu- 
ously described birds. The field marks discussed under Accepted Records are particu- 
larly important. Note also that most Blue-headed Vireos appear to pass through 
California from late September through October and that reports from outside this 
window may require exceptionally complete details for CBRC endorsement. 


37 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 



Figure 10. Possible Greater Pewee, Corttopus pertinax, on Southeast Farallon 
Island, San Francisco County, 1 June 1998 (1998-116). This bird’s crest and bill 
appear more or less typical for a Greater Pewee, but most members considered the 
whitish wing-bars and edging to the inner secondaries and tertials to be outside the 
species’ typical range of variation. 


Photo by Clyde Morris 


‘PHILADELPHIA VIREO Vireo philadelphicus. Descriptions of a bird at Galileo 
Hill Park, KER, 4 Oct 1998 (1998-153) failed to rule out the similar Warbling Vireo 
and received no votes for acceptance. 

YELLOW-GREEN VIREO Vireo flauouiridis. One reported at the Marin Head- 
lands, MRN, 5 Oct 1996 (1997-049) fell a vote shy of acceptance after four voting 
rounds, while one reported from Moss Beach, SM, 4 Oct 1998 (1998-226) received 
only four votes to accept in the first round. In each case, members expressed concern 
that a bright Red-eyed Vireo (V. oiivaceus) was not eliminated. Yellow-green Vireos 
are most reliably identified by the extension of yellow tones from the shoulder to 
behind the auriculars and by yellow edges to the remiges and rectrices (Terrill and 
Terrill 1981, Erickson and Terrill 1996), 

VEERY Cotharus fuscescens. Descriptions of a rusty Catharus thrush at the 
Carmel River mouth, MTY, 21-22 Sep 1998 (1999-052) pointed toward this 
species, but most members ultimately concurred that the observers’ views were too 
brief to permit the bird’s confident identification. 

BLUE-WINGED WARBLER Vermiuora pinus. One reported at Birchim Canyon, 
NW of Bishop, INY, 22 May 1998 (1998-082) was seen rather briefly by a lone 
observer and received only one “accept” vote during the first round. Apart from most 
members’ unwillingness to accept uncorroborated records from the observer, the 


38 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 



Figure 11. Alder/Willow Flycatcher, Empidonax alnorum/traillii at Galileo Hill 
Park, Kern County, 11 Oct 1998 (1998-161). Alder-like features evident in this 
photo include the whitish chin and throat, generally grayish underparts, prominent 
wing bars, fairly obvious eye-ring, and relatively short bill. 

Photo by Larry Sansone 


description of the wings and back was atypical for a Blue-winged Warbler, and the bird 
could not be refound by others later in the day. 

PINE WARBLER Dendroica pinus. Reports of brightly colored individuals at 
Arroyo de la Cruz, SLO, 25 Sep 1998 (1999-042) and Oceano, SLO, 1 Oct 1998 
(1999-041) each garnered four votes of acceptance during the first round. Members 
are generally skeptical of Pine Warblers reported in September and early October, and 
the CBRC is presently re-reviewing the three accepted records from before 13 
October. It appears that observers underestimate the variability of Blackpoll (D. 
striata) and Bay-breasted (D. castanea) warblers, which can appear nearly plain 
backed during their first fall. Observers should also be aware that a Pine Warbler’s tail 
is no longer than that of a Blackpoll or Bay-breasted (Pyle 1997); thus, descriptions of 
this feature should relate the undertail coverts to the tail tip. Early fall claims of the Pine 
Warbler should be backed by impeccable details, preferably by multiple observers, or 
by definitive photographs. 

WORM-EATING WARBLER Helmitheros uermivorus. A report of one in Ventura, 
VEN, 6-25 Jan 1997 (1997-085) ultimately foundered after four rounds, despite 
having been seen by multiple observers during an extended stay. The only description 


39 



REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


submitted for CBRC review pointed toward this rather distinctive warbler, but most 
members were concerned by mention of “a few small brown streaks at the top of the 
chest,” which this species would not show unless wet or otherwise mussed. This 
debacle again demonstrates how CBRC review can be undermined when observers 
assume that others will submit adequate supporting details. If additional details are 
received, the record will be re-reviewed. 

CONNECTICUT WARBLER Oporornis agilis. One reported at the Big Sur 
Ornithology Lab, MTY, 13 Sep 1998 (1998-192) was supported by fairly extensive 
notes that most members took as supporting the observer’s claim. Another person 
present on the day of the sighting, however, submitted a note to the Committee 
stating that the observer “considered the record tentative” on that day, and conveying 
an “impression [from the observer] that the bird was seen very, very briefly (1-2 
seconds) only.” In light of this testimony, the record garnered only four votes in the 
first round. 

MOURNING WARBLER Oporornis Philadelphia, The Committee withheld en- 
dorsement of individuals claimed at Santa Barbara I., SBA, 31 May 1974 (1980- 
021A; resubmitted after prior nonacceptance as a Connecticut Warbler), Ferndale, 
HUM, 31 Aug 1996 (1997-048), Hansen Dam Park, Los Angeles, LA, 24 Sep 1996 
(1997-026), Big Sur R. mouth, MTY, 19 Sep 1998 (1999-066), and Mt. Diablo State 
Park, CC, 16 Oct 1998 (1999-064). Most members opined that the bird at the Big 
Sur R. mouth was most likely a Common Yellowthroat ( Geothlypis trichas). The 
revived report from Santa Barbara 1., which four members voted to accept, seemingly 
suffered more from the baggage of prior submission (the observer having previously 
argued for another species) than from weakness of the documentation, which 
indicated an alternate-plumaged female Mourning Warbler. The remaining three 
reports involved birds described as showing little or no yellow in the throat, which in 
fall should indicate an adult female, which should account for a very small percentage 
of vagrants; an aberrant bird or hybrid (the latter unproven; Pitocchelli 1993, Heindel 
and Patten 1996); or the observer’s failure to detect pale yellow (cf. earlier discussion 
of 1998-136). Though most members considered these reports to be likely correct, 
observer inexperience and other factors ultimately turned the voting decisively against 
them. For further information on identification of this species and congeners see Pyle 
and Henderson (1990), Pyle (1997), and Dunn and Garrett (1997). 

SCARLET TANAGER Piranga oliuacea. Members generally agreed that a male 
was seen at Furnace Creek Ranch, Death Valley, INY, 10 Nov 1997 (1998-060), but 
after four rounds three members remained unconvinced that the record was ad- 
equately documented. 

SMITH’S LONGSPUR Calcarius pictus. A report of one at Galileo Hill Park, KER, 
10 Oct 1998 (1998-185) received no votes for acceptance, a few members suggest- 
ing that the details better fit a Vesper Sparrow (Pooecetes gramineus). A report from 
San Nicolas I., VEN, 18 Oct 1998 (1999-030) was thought to have greater potential 
of being correct, but after two rounds this report also received no votes to accept 
because the Vesper Sparrow and other longspurs were not convincingly ruled out. 

McKAY’S BUNTING Plectrophenax hyper bo reus. A nearly white bird associating 
with Dark-eyed Juncos ( Junco hyemalis) west of Orleans, HUM, 18 Sep 1998 
(1998-147) was considered by the Committee to have almost certainly been a 
partially albinistic Dark-eyed Junco. 

VARIED BUNTING Passerina versicolor. An adult male reported at Lancaster, 
LA, 21-23 Apr 1996 (1996-100) received no votes for acceptance after four rounds, 
the tallies being evenly split between “identification not established” and “natural 
occurrence questionable.” Although most members considered the identification to be 


40 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


likely correct, several were concerned that one description likened the bird’s size to 
that of a Lark Sparrow ( Chondestes grammicus ) and that the only other description 
appeared to have been written several weeks after the sighting. Members’ standards 
of acceptance are generally quite high for this species owing to its extreme rarity in 
California (two accepted records to date) and to its potential for escape from captivity. 
Hamilton (in press) has seen as many as 42 for sale in northwestern Baja California 
(Rosarito, 5 July 2000). For such species, observations should be detailed enough to 
detect signs of possible earlier captivity (e.g., atypical coloration in any feather tract, 
broken remex tips, frayed rectrices, abnormally long claws, bill deformities), and the 
description should indicate that such an inspection was conducted. 

PAINTED BUNTING Passerina ciris. The description of one reported at San Elijo 
Lagoon, SD, 20 Aug 1996 (1997-028) included relatively little detail, and the bill was 
described as being “very small and black.” Since the Painted Bunting’s bill is relatively 
large for a Passerina bunting and is typically dusky with a pinkish tinge, documenta- 
tion of this record was considered deficient by all but one member after the third round 
of voting. 

COMMON GRACKLE Quiscalus quiscula. Brown-headed individuals reported at 
the Honey Lake Wildlife Area, LAS, 11 May 1998 (1998-074) and at Panamint 
Springs, INY, 17 Oct 1998 (1999-063) received almost no Committee support 
because the descriptions did not match any post-juvenal plumage of the Common 
Grackle. Members opined that the latter description better fit a small female Great- 
tailed Grackle (Q. quiscula), with Patten suggesting the possibility of a small bird of the 
subspecies nelson i in a flock of the substantially larger monsoni; these races have 
colonized California rapidly in recent decades (Dinsmore and Dinsmore 1993, W. 
Wehtje pers. comm.) and appear to be forming a “hybrid swarm” north and west of 
the species’ historic range (Rea 1969, W. Wehtje pers. comm.). Another possibility 
mentioned was a Great-tailed Grackle x Brewer’s Blackbird (Euphagus cyano- 
cephalus), a hybrid combination believed to have occurred recently in Santa Maria, 
SBA (unpubl. CBRC record 1999-122). 

RECORDS NOT ACCEPTED, NATURAL OCCURRENCE 
QUESTIONABLE (IDENTIFICATION ACCEPTED) 

BROWN BOOBY Sula leucogaster. Another immature initially identified as a Red- 
footed Booby (see accepted records above) was at a fishing pier on Pt. Loma, SD, 17 
Nov 1997 (CTf ; AM|, SW; 1998-008). The bird was exceptionally tame (even for a 
booby), in abnormal plumage (white feathers on the mid breast), entangled in fishing 
line, and emaciated when taken into captivity that day. According to Meryl Faulkner, 
an experienced wildlife rehabilitator specializing in seabirds, the condition of the feet, 
bill, and feathers clearly indicated the bird had been in captivity for some time prior to 
17 Nov 1997 (McCaskie, in comments). The fourth and final vote on the record was 
five “accept” and five “not accept, natural occurrence questionable.” 

BARNACLE GOOSE Branta leucopsis. One was at Granite Bay, PLA, at least 1 
Jan-10 July 1998 (JAT; 1998-045). The record received no support as representing 
a natural vagrant, from the observer or the Committee, but the archiving of records of 
all long-distance migratory species is strongly encouraged. 

PYRRHULOXIA Cardinalis sinuatus. Although occurring at an “expected” time 
of year for natural strays of this species in California (May through July), a female at 
Pt. Loma, SD, 10 Jun 1998 (REW; 1999-044) “appeared ragged” and was close 
enough to the international border that six members considered it more likely an 
escapee. Recent investigations (Hamilton in press) have yielded multiple sightings of 


41 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


Pyrrhuloxias (and many other wild-caught birds) at pet stores in northwestern Baja 
California. 

PAINTED BUNTING Passerina ciris. Records not endorsed by the Committee 
include a spring adult male at Big Pine, INY, 9 Jun 1996 (JHe; 1996-105), a fall adult 
male at Victorville, SBE, 12 Oct 1997 (BiD; 1997-165), and an adult male with an 
adult female at Niland, IMP, 22-24 Dec 1998, with the female present to IS Jan 
1999 (MSM; GMcC, MAP; 1999-017). Interestingly, the observer of the Victorville 
bird also found the adult male at Twentynine Palms, SBE, 8 Oct 1998 (see accepted 
records above). The former record failed to gain acceptance because Victorville is far 
more urbanized than is Twentynine Palms (increasing, perhaps, the likelihood of 
escape/release) and the underparts and rump of the Victorville bird were described as 
being “salmon reddish” versus the Twentynine Palms bird’s “typical” shade of red. 
The pair at Niland included the first green Painted Bunting detected wintering in 
California (none of the seven prior reports of wintering adult males gained CBRC 
endorsement). Nine members considered natural occurrence unlikely in this case 
because the record involved a pair of adults and the likelihood that caged Painted 
Buntings are kept and sold in fair numbers in the Mexicali region. The Big Pine bird, 
which received seven votes for acceptance after four rounds, was well described by a 
seasoned observer as showing normal bright coloration, and both the date and 
location were perhaps ideal for a naturally occurring spring vagrant. It is, therefore, 
not surprising that this record was highly controversial. We provide the following 
condensation of Hamilton’s ongoing research into the Painted Bunting’s unusual 
status in California to help clarify the Committee’s reasoning in this case. 

Painted Buntings have, since 1962, established a pattern of fall vagrancy to 
California consistent with that exhibited by other eastern landbirds, with 69 accepted 
fall records between 17 August and 3 December. Green birds of unknown age 
excluded, immatures account for 30 of 41 fall records (73%). With the exclusion of 
eight fall records of adult males rejected by the Committee because of questionable 
natural occurrence, the proportion of immatures rises to 91%. A share of first-year 
birds in the range 73-91% matches expectations for naturally occurring vagrant 
songbirds (e.g., Robbins et al. 1959, Ralph 1971, Gathreaux 1982, DeSante 1983). 
During the remainder of the year, however, birds less than a year old account for just 
one of 19 records (5%), with colorful males accounting for 17 of the 18 records of 
adults (94%). Patterns of vagrancy such as these are, to our knowledge, unknown in 
nature. 

Painted Buntings are the single most abundant wild-caught species at bird markets 
that we have monitored in Baja California, with a high count of 40 colorful males and 
37 green birds at a Rosarito pet store 5 July 2000 (Hamilton in press). In addition, 
birders note apparent escapees (colorful males with abnormal appearance) south of 
the border with some regularity (Erickson et al. in press). Just as troubling, a colorful 
male collected at the Sagehen Field Station, NEV, 17-18 Apr 1985 (1988-229) was 
yellow below and scarred between the orbits, strong indications of prior captivity 
(Hawthorne 1972). This record’s remote northeastern location and spring timing 
might best be explained by a wild-caught escapee responding to migratory restless- 
ness and attempting northward migration (alternate scenarios have not been articu- 
lated). 

A fundamental CBRC objective is to distinguish between natural patterns of 
occurrence and those relating to captivity. Therefore, some members consider it best 
to withhold CBRC endorsement of adult male Painted Buntings outside of fall 
migration until it becomes clear that the preponderance of adult males in winter, 
spring, and/or summer results from some unique natural phenomenon, not simply 
wild-caught escapees adapting to their new surroundings in winter and yielding to 
Zugunrhue in spring. Whereas winter and late summer records generally receive little 


42 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 





Figure 12. Alder/Willow Flycatcher, Empidonax alnorum/traillii at Galileo Hill 
Park, Kern County, 11 Oct 1998 (1998-161). The bold tertial edging and eye-ring 
rule out western subspecies of the Willow Flycatcher, though not an eastern bird. The 
appearance of a gray nape contrasting with tire greener back results from exposure of 
the grayish bases of green-tipped nape feathers. 

Photo by Larry Sansone 


support, most members believe that Painted Buntings are likely to wander occasion- 
ally to the eastern deserts in late spring/early summer Furthermore, colorful male 
Painted Buntings may be more likely than females or nonbreeding green males to 
“overshoot” their breeding grounds. Seemingly supporting this hypothesis are photos 
in N, Am. Birds 53:343-344 of definitive alternate-plumaged males in Saskatchewan 
1-13 May 1999 {underparts not clearly visible) and near Aztec, New Mexico, 28 
April-2 May. Note, however, that an adult male photographed in northern Ontario 
15 May 1998 (N. Am. Birds 52:406) had faded red underparts, suggesting diet 
deficiencies and possible prior captivity. Because of such considerations, California’s 
five spring/early summer records of colorful males in the eastern deserts are 
particularly difficult to assess. Two accepted records from the 1980s (1986-262, 
1987-138) are presently being reassessed, while two later records (1996-092, 1996- 
105) are not accepted, and the fifth (2000-086) is circulating for the first time. All 
members recognize the potential for each of these records to pertain to a naturally 
occurring vagrant, but the lack of green birds causes a minority of members to 
withhold endorsing them at this time. 


43 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


With continued research and the accumulation of records, the Painted Bunting’s 
natural patterns of vagrancy should become better established, perhaps prompting 
reevaluation of certain records in light of more complete knowledge of the species’ 
natural history. In the meantime, we hope that increased understanding of the issues 
will help replace current hostility over the fate of individual records with the positive 
energy of a larger mystery to be solved. We strongly encourage the continued 
submission of documentation of Painted Buntings, as these records are needed to 
build a more complete picture of the species’ status and distribution in California. We 
also welcome responses to this discussion. 

RECORDS NOT ACCEPTED, Identification accepted but establishment 
of introduced population questionable 

TRUMPETER SWAN Cygnus buccinator. A neck-collared adult NW of Roseville, 
PLA, 22-23 Dec 1998 (BEW; REM; 1998-230) was captured in northern Idaho, 
released in southern Idaho in November 1996, and seen in various states and 
provinces in the interim. McCaskie and San Miguel (1999) explained the CBRC’s 
position on such records. 

MISCELLANEOUS DECISIONS 

The following decisions were made at the January 2000 meeting. 

WHITE IBIS Eudocimus a I bus (2,0). Two records at the Salton Sea in the 1970s 
are now considered as pertaining to the same individual. We infer an adult at the N end 
of the sea, RIV, 10-24 Jul 1976 and at the S end of the sea, IMP, 5 Aug 1976 (1 976- 
045) was the same as one at Unit 1, Salton Sea N.W.R., IMP, 25 Jun-14 Jul 1977 
(1978-049). These records were originally reported by Luther et al. (1979, 1983). 
The only other California record is of one collected in San Diego, SD, 15-20 Nov 
1935 (Huey 1936, Dunn 1988). 

GRAY SILKY-FLYCATCHER Ptilogonys cinereus. This species joined the Falcated 
Duck (Anas falcata ), Crested Caracara ( Caracara cheriway), and Oriental Greenfinch 
(i Carduelis sinica) on the Supplemental List of California birds (Patten and Erickson 
1994) on the basis of these three records “not accepted, natural occurrence question- 
able”: one photographed at Ventura, VEN, 9 Apr 1976 (1980-115, Binford 1985), 
one at Pt. Loma, SD, 24 May 1993 (1993-115, McCaskie and San Miguel 1999), 
and one photographed at Poway, SD, 10-12 Mar 1994 (1994-075, Howell and Pyle 
1997). A more recent record from Orange County is still under review. 

CORRIGENDA TO THE TWENTY-THIRD COMMITTEE REPORT 
(Rottenborn and Morlan 2000) 

The date of the accepted Wedge-rumped Storm-Petrel (1996-1 14) was not given. 
This bird was recorded 31 July 1996. 

CONTRIBUTORS 

Mark Ackerman, R. J. Adams, John R. Arnold, Jonathan Ausubel, John Ayres 
(JAy), Stephen F. Bailey, Alan Baldridge, Thomas C. Barber, Alan D. Barron, John 
Barrow (JBa), Tony Battiste, Louis R. Bevier, Laurence C. Binford, Shauna Bingham, 
Jim Booker (JBo), William Bouton, Bill Boyce, Ronald L. Branson, Greg Brinkley, N. 
Bruce Broadbooks, Eric Brooks, Ryan Burnett, Kenneth Burton, John Butler, Kurt F. 


44 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


Campbell, Phil Capitolo, Eugene A. Cardiff, Kris K. Carter, George E. Chaniot, Jr., 
Ryan Chornock, Janis Christian, Sue Clark, Therese R. Clawson, Luke Cole, Daniel 
S. Cooper, Don Cunningham, Jim Danzenbaker, Stephen J. Davies, Gary Deghi, A1 
DeMartini, Bill Deppe, David F. DeSante, Don DesJardin, Bruce E. Deuel, Bruce 
Dexter, Sandy Dierks, Vladimir Dinets, Matthew Dodder, Jon L. Dunn, Mark W. 
Eaton, Demian A. Ebert, Thomas M. Edell, Alan M. Eisner, Ray Ekstrom, F, Emerson, 
Richard A, Erickson, Sandy Etchell, Gil Ewing, Michael Feighner, George Finger, 
Shawneen E. Finnegan, Robbie Fischer, John Fitch, Brian Foster, Nick Freeman, 
Peter Gaede, Sylvia J. Gallagher, Gray Gallogly, Kimball L. Garrett, Douglas E. 
George, Bruce Gerow (BGe), Steve Gerow, Albert Ghiorso, Karen S. Gilbert, Martin 
Gilbert, Peter A. Ginsburg, Steven A. Glover, Dave Goodward, Ed Greaves, Eric W. 
Greisen, Jennifer E, Green, Bill Grenfell, George Griffeth, Mary Gustafson, Charity 
Hagan, Frank A. Hall, Robert A. Hamilton, Steve C. Hampton, Keith Hansen, Denise 
Hardesty, Sandra G. Harvill, David P Haupt, Karen A. Havlena, Loren R. Hays, Scott 
Hein, D. Mitchell Heindel, Jo Heindel, Matthew T, Heindel, Tom Heindel, Pablo 
Herrera, Michelle Hester, Craig Hohenberger, James F. Holmes, Waldo Holt, Andrew 
Howe, Steve N. G. Howell, Robert A. Hudson (RAHu), E. Rae Hudspeth, Joan 
Humphrey, John E. Hunter, Marshall Iliff, Richard Irvin, Alvaro Jaramillo, Curtis 
Johnson, H. Lee Jones, Paul Jorgensen, Robert J. Keiffer, Paul Keller, Alison Kent, 
Peter Knapp, David Koeppel, Andrew W. Kratter, Kenneth Z. Kurland, Keith C. 
Kwan, Debi Lamm, Jeri M. Langham, Kevin Larson, Stephen A. Laymon, Rick 
LeBaudour, Paul E. Lehman, Gary S. Lester, Lauren P. Lester, Nick Lethaby, Ron 
LeValley, Cindy Lieurance, Leslie M. Lieurance, Roger Linfield, Michael J. 
Lippsmeyer, Paul Lohse, Guy Luneau, Rolf E. Mall, Michael J. Mammoser, Tim 
Manolis (TM), Curtis A. Marantz, John S. Mariani, Dave Marquart (DMa), Bruce 
Marshall (BMa), Doug Martin (DM), John Martin (JMa), Jennifer L. Matkin, Robert E. 
Mauer, Jr. (REMa), Sean McAllister, Guy McCaskie, Todd McGrath, Bert McKee, 
Patrick McNulty, Anthony Mercieca, Bob Merrill, Peter J. Metropulos, Mark Miller, 
Kathy Molina, Joseph Morlan (JM), Clyde Morris, Michael H. Morris, Steve Morris, 
Jim Mountjoy (JMo), Dan Murley (DMu), Stephen J. Myers, Dan W. Nelson, David W. 
Nelson, Kristie Nelson, Richard J. Norton, Robert J. O’Brien, Jerry Oldenettel, Helen 
Ost, John Ost, Debby Parker, Jim Parker, John E. Parmeter (JEPa), Benjamin D. 
Parmeter, Michael A. Patten, Courtenay Peddle, Dharm S. Pellegrini, J.D. Phillips, 
James E. Pike, Jeff Poklen, Gary W. Potter, Peter Pyle, David E. Quady, Kurt 
Radamaker, Richard E. Redmond, Robert W. Reiling, David Rice (DRi), Will 
Richardson, Jean M. Richmond, Janet Robbins, Don Roberson, Geoffrey L. Rogers, 
Michael M. Rogers, Stephen C. Rottenborn, James S. Royer, Ruth A. Rudesill, 
Tamiko Ruhlen, Edd Russel, Ron M. Saldino, Ivan Samuels, Michael J. San Miguel, 
Mike San Miguel, Daan Sandee, Larry Sansone, Paul M. Saraceni, Brad Schram, 
David M. Shaw, Douglas G. Shaw, Debra L. Shearwater, W. David Shuford, David A. 
Sibley, Mary Simpson, Daniel S. Singer, Mike Skram, Arnold Small, Brenda D. Smith, 
Reed V. Smith, Ron Smith (RSm), Jim Snowden (JSn), John Sorensen, Steve 
Sosensky, Jean Marie Spoelman, Rich Stallcup, Susan Steele, Daniel M. Stoebel, 
Mary Beth Stowe, Emilie Strauss, Jeffrey W. Streb, David L Suddjian, Robert 
Sutherland (RSu), Mac Sutherlin, Lee Sutton, Ann D. Swart, Craig Taylor, Monte M. 
Taylor, Richard L. Ternullo, Scott B. Terrill, Robert Theriault, Ronald S. Thorn, 
Robert F. Tintle, Francis Toldi, Gerald L. Tolman, Bob Toth, John A. Trochet, Steve 
J. Tucker, Philip Unitt, Sven Wahlberg (SvW), Stan Walens (SWa), Bruce E. Webb, 
Sophie Webb, Dave Weber, Richard E. Webster, Walter Wehtje, Joel D. Weintraub, 
Alan N. Wight, David S. Wilcove, Brian D.C. Williams, Douglas R. Willick, Rob Willis, 
John C. Wilson, Steve Wilson, Adam Winer, Thomas E. Wurster, David G. Yee, James 
O. Zimmer. 


45 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


ACKNOWLEDGMENTS 

This report would not have been possible without the 27 1 observers who submitted 
reports to the Committee. Peter R. Bloom, Janet Linthicum, Eric Mellink, Kurt 
Radamaker, Brian J. Walton, and Thomas E. Wurster provided input critical to CBRC 
review of recent Harris’s Hawk records. M. Ralph Browning and the late Claudia 
Wilds examined a stint specimen. Chris Corben offered insight on Mongolian Plover 
identification. CBRC review of Iceland Gull records was assisted by Erik A. T. Blom, 
Davis Finch, Roger Foxall, Daniel D. Gibson, Michel Gosselin, the late Peter J. Grant, 
Raymond Henson, Ted Hoogendoorn, Dennis Paulson, J. V. Remsen, Jr., Stuart 
Tingley, and Thede Tbbish. David A. Sibley advised on wagtail identification. Drafts of 
this report were reviewed and improved by Kimball Garrett, Steve N. G. Howell, 
Alvaro Jaramilio, Gary S. Lester, Michael Patten, Peter Pyle, Don Roberson, Michael 
Rogers, and Mike San Miguel. Peg Stevens and Jon C. Fisher continued to archive the 
committee’s materials at WFVZ. We extend our thanks to all. 

LITERATURE CITED 

American Ornithologists’ Union. 1973. Thirty-second supplement to the American 
Ornithologists’ Union Check-list of North American Birds. Auk 90:411-419. 

American Ornithologists’ Union. 1983. Check-list of North American Birds, 6th ed. 
Am. Ornithol. Union, Washington, D. C. 

American Ornithologists’ Union. 1998. Check-list of North American Birds, 7th ed. 
Am. Ornithol. Union, Washington, D. C. 

American Ornithologists’ Union. 2000. Forty-second supplement to the American 
Ornithologists’ Union Check-list of North American Birds. Auk 1 17:847-858. 

Badyaev, A. V., Gibson, D. D., and Kessel, B. 1996. White Wagtail ( Motacilla alba) 
and Black-backed Wagtail ( Motacilla lugens), in The Birds of North America (A. 
Poole and F. Gill, eds.), nos. 236, 237. Acad. Nat. Sci., Philadelphia. 

Bevier, L. R. 1990. Eleventh report of the California Bird Records Committee. W. 
Birds 21:145-176. 

Binford, L. C. 1985. Seventh report of the California Bird Records Committee. W. 
Birds 16:29-48. 

Capitolo, P., Richardson, W., Burnett, R., and Pyle, P. 2000. First record of an Olive- 
backed Pipit in California. W. Birds 31:112-116. 

Clapp, R. B. 1971. A specimen of Jouanin’s Petrel from Lisianski Island, northwest- 
ern Hawaiian Islands. Condor 73:490. 

DeSante, D. F. 1983. Annual variability in the abundance of migrant landbirds on 
Southeast Farallon Island, California. Auk 100:826-852. 

DeSante, D. F., Johnson, N. K., Le Valley, R., and Henderson, R. P. 1985. Occur- 
rence and identification of the Yellow-bellied Flycatcher on Southeast Farallon 
Island, California. W. Birds 16:153-160. 

Dinsmore, J. J., and Dinsmore, S. J. 1993. Range expansion of the Great-tailed 
Graclde in the 1900s. J. Iowa Acad. Sci. 100:54—59. 

Dunn, J. L. 1988, Tenth report of the California Bird Records Committee. W. Birds 
19:129-163. 

Dunn, J. L., and Garrett, K. L. 1997. A Field Guide to Warblers of North America. 
Houghton Mifflin, Boston. 

Erickson, R. A., and Terrill, S. B. 1996. Nineteenth report of the California Bird 
Records Committee: 1993 records. W. Birds 27:93-126. 


46 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


Erickson, R. A., Hamilton, R. A., and Howell, S, N. G. In press. New information on 
migrant birds in northern and central portions of the Baja California Peninsula, 
including species new to Mexico, in Birds of the Baja California Peninsula: Status, 
Distribution, and Taxonomy (R. A. Erickson and S. N. G. Howell, eds.). Am. 
Birding Assoc. Monogr. Field Ornithol. 

Garrett, K., and Dunn, J. 1981. Birds of Southern California; Status and Distribution. 
Los Angeles Audubon Soc., Los Angeles. 

Gathreaux, S. A., Jr. 1982. Age-dependent orientation in migratory birds, in Avian 
Navigation (F. Papi and H. G. Wallraff, eds.), pp. 68-74. Springer-Verlag, Berlin. 

Hamilton, R. A. In press. Records of caged birds in Baja California. Appendix D in 
Birds of the Baja California Peninsula: Status, Distribution, and Taxonomy (R. A. 
Erickson and S. N. G. Howell, eds.). Am. Birding Assoc. Monogr. Field Ornithol. 

Hawthorne, V. M. 1972. Painted Bunting record for northeastern California. Calif. 
Birds 3:91-92. 

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. 

Heindel, M. T., and Patten, M. A. 1996. Eighteenth report of the California Bird 
Records Committee: 1992 records. W. Birds 27:1-29. 

Heindel, M., and Pyle, P. 1999. Identification of Yellow-bellied and “Western” 
flycatchers. Birders’ J. 8:78-87. 

Howell, S. N. G. 2000. Identification of Thayer’s-like gulls: The Herring x Glaucous- 
winged gull problem. Birders’ J. 9:25-33. 

Howell, S. N. G., and Pyle, P. 1997. Twentieth report of the California Bird Records 
Committee: 1994 records. W. Birds 28:117-141. 

Howell, S. N. G., and Webb, S. 1995. A Guide to the Birds of Mexico and Northern 
Central America. Oxford Univ. Press, Oxford, England. 

Huey, L. M. 1936. Noteworthy records from San Diego, California. Condor 38:121. 

Kishchinski, A. A., and Lobkov. E. G. 1979. Spatial relationships between some bird 
subspecies in the Beringiari forest-tundra (in Russian). Moskovskoe Obshchestvo 
i Spytatelei Prirody. Otdel Biol. Biull. Novaia Seriia 5:11-23. 

Langham, J. M. 1991. Twelfth report of the California Bird Records Committee. W. 
Birds 22:97-130. 

LeGrand, H. E., Jr., Guris, P., and Gustafson, M. 1999. Bulwer’s Petrel off the North 
Carolina coast. N. Am. Birds 53:113-115. 

Luther, J. S., McCaskie, G., and Dunn, J. 1979. Third report of the California Bird 
Records Committee. W. Birds 10:169-187. 

Luther, J. S., McCaskie, G., and Dunn, J. 1983. Fifth report of the California Bird 
Records Committee. W. Birds 14:1-16. 

McCaskie, G. 1975. A Rufous-necked Sandpiper in southern California. W. Birds 
6:111-113. 

McCaskie, G., and San Miguel, M. 1999. Report of the California Bird Records 
Committee: 1996 records, W. Birds 30:57-85. 

McCaskie, G., and Webster, R. E. 1990. A second Wedge-tailed Shearwater in 
California. W. Birds 21:139-140. 

Mlodinow, S. G., Feldstein, S., and Tweit, B. 1999. The Bristle-thighed Curlew landfall 


47 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


of 1998: Climate factors and notes on identification. W. Birds 30:133-155. 

National Geographic Society. 1999. Field Guide to the Birds of North America, 3rd 
ed. Natl. Geogr. Soc., Washington, D.C. 

Patten, M. A. 1993. First record of the Common Pochard in California. W. Birds 
24:235-240. 

Patten, M. A. 2000. Changing seasons: Winter season, December 1999 to February 
2000. N. Am. Birds 54:146-148. 

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 Research 
34:187-195. 

Patten, M. A,, McCaskie, G., and Morlan, J. 1999. First record of the American 
Woodcock for California, with a summary of its status in western North America. 
W. Birds 156-166. 

Patterson, M. 1998. The great curlew fallout of 1998. Field Notes 52:150-155. 

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. 

Pitocchelli, J. 1993. Plumage and size variation in the Mourning Warbler. Condor 
94:198-209. 

Pyle, P. 1997. Identification Guide to North American Birds, part I. Slate Creek Press, 
Bolinas, CA. 

Pyle, P., and Henderson, P, 1990. On separating female and immature Oporornis in 
fall. Birding 22:222-229. 

Pyle, P., and McCaskie, G. 1992. Thirteenth report of the California Bird Records 
Committee. W. Birds 23:97-132. 

Ralph, C. J. 1971. An age differential of migrants in coastal California. Condor 
73:243-246. 

Rea, A. M. 1969. The interbreeding of two subspecies of Boat-tailed Grackle 
Cassidix mexicanus nelsoni and Cossidix mexicanus monsoni in secondary 
contact in central Arizona. M. S. thesis, Univ. of Ariz., Tucson. 

Ridgway, R. 1919. The birds of North and Middle America, part VIII. U. S. Natl. Mus. 
Bull. 50. 

Robbins, C., Bridge, D., and Feeler, R. 1959. Relative abundance of adult male 
redstarts at an inland and a coastal locality during fall migration. Maryland Birdlife 
15:23-25. 

Roberson, D. 1980. Rare Birds of the West Coast. Woodcock Publ., Pacific Grove, 
CA. 

Roberson, D. 1986. Ninth report of the California Bird Records Committee. W. Birds 
17:49-77. 

Roberson, D. 1998. Sulids unmasked: Which large booby reaches California? Field 
Notes 52:276-297. 

Rottenborn, S. C., and Morlan, J. 2000. Report of the California Bird Records 
Committee: 1997 records. W. Birds 31:1-37. 

Sibley, D. A. 2000. The Sibley Guide to Birds. Knopf, New York. 

Sibley, D. A., and Howell, S. N. G. 1998. Identification of White and Black-backed 


48 


REPORT OF THE CALIFORNIA BIRD RECORDS COMMITTEE: 1998 RECORDS 


Wagtails in basic plumage. W. Birds 29:180-198. 

Small, A. 1994. California Birds: Their Status and Distribution. Ibis Publ., Vista, CA. 

Spear, L. B., Lewis, M. J., Myers, M. T., and Pyle, R. L. 1988. The recent occurrence 
of Garganey in North America and the Hawaiian Islands. Am. Birds 42:385- 


Stallcup, R., Morlan, J., and Roberson, D. 1988. First record of the Wedge-tailed 
Shearwater in California. W. Birds 19:61-68. 

Terrill, S. B., and Terrill, L. S. 1981. On the field identification of Yellow-green, Red- 
eyed, Philadelphia, and Warbling vireos. Continental Birdlife 2:144-149. 

Terrill, S. B., Rottenborn, S. C., Singer, D. S., and Roberson, D. 1999. Winter 
season: Middle Pacific coast region. N. Am. Birds 53:203-207. 

Veit, R. R. 1988. Identification of the Salton Sea Rufous-necked Sandpiper. W. Birds 
19:165-169. 


392. 


Accepted 1 1 November 2000 



49 


ARIZONA BIRD COMMITTEE REPORT: 

1996-1999 RECORDS 

GARY H. ROSENBERG, P. O. Box 91856, Tucson, Arizona 85752-1856 

ABSTRACT: The Arizona Bird Committee assessed 218 records, accepting 138. 
These included Arizona’s first records of the Leach’s ( Oceanodroma leucorhoa), 
Black (O. melania), and Least (O. microsoma ) Storm-Petrels, Short-tailed Hawk 
(Buteo brachyurus), Pacific Golden Plover ( Pluvialis fulua), Yellow-footed Gull 
( Larus livens), and Carolina Wren ( Thryothorus ludovicianus), bringing the number 
of bird species recorded in Arizona to 522. 

This is the fourth report of the Arizona Bird Committee (hereafter ABC) 
(see Speich and Parker 1973, Speich and Witzeman 1975, and Rosenberg 
and Witzeman 1998, 1999). This report covers records mainly from the 
period between 1996 and 1999 but also includes some from prior to 1996 
previously not reported on by the ABC. A total of 218 reports submitted to 
the ABC are addressed here, with 138 (63%) accepted. Seven species were 
added to the Arizona list, Leach’s Storm-Petrel ( Oceanodroma leucorhoa), 
Black Storm-Petrel ( Oceanodroma melania), Least Storm-Petrel 
( Oceanodroma microsoma) (first physical documentation), Short-tailed 
Hawk ( Buteo brachyurus), Pacific Golden-Plover ( Pluvialis fulua), Yellow- 
footed Gull ( Larus livens), and Carolina Wren ( Thryothorus ludovicianus). 
This brings the total number of species recorded in Arizona to 522, which 
includes four species that have been accepted by the ABC but have not been 
physically documented within the state: Leach’s Storm-Petrel, Swallow- 
tailed Kite (Elanoides forficatus), Black Swift ( Cypseloides niger), and Blue- 
headed Vireo ( Vireo so I i tar i us). 

Other highlights reported on here include acceptance of Arizona’s second 
Wandering Tattler ( Heteroscelus incanus), fifth Pomarine Jaeger 
( Stercorarius pomainus), second Black-billed Cuckoo ( Coccyzus 
erythropthalmus), third Eastern Wood-Pewee (Contopus virens), second 
Nutting’s Flycatcher ( Myiarchus nuttingi), first breeding Winter Wren ( Tro- 
glodytes troglodytes), third breeding Black-capped Gnatcatcher ( Polioptila 
nigiceps ), first photographed Blue-winged Warblers ( Vermiuora pin us), 
second and third photographed Prairie Warblers ( Dendroica discolor ), first 
sound-recorded Fan-tailed Warbler ( Euthlypis lachrymosa ), first winter 
Rufous-capped Warblers ( Basileuterus rufifrons), and third Field Sparrow 
( Spizella pusilla). 

The current Arizona Bird Committee (2000) consists of Chris D. Benesh, 
Troy Corman, Roy M. Jones, David Krueper, Narca Moore-Craig, Gary H. 
Rosenberg (secretary), Will Russell, and Mark M. Stevenson. Recent commit- 
tee members who also voted on records in this report include Doug Danforth, 
Kenn Kaufman, Chuck LaRue, G. Scott Mills, Dave Stejskal, and Carl Tomoff. 
Janet Witzeman serves in a non-voting capacity as assistant secretary. 

The list of species on the ABC’s review list can be found on its web page 
(http://personal.riverusers.com/~ghrosenberg/GaryRosenbergHomePage.html). 
Included on this web site as well are the ABC’s bylaws, a current list of 
committee members, a brief history of the ABC, the past two reports by the 


50 


Western Birds 32:50-70, 2001 


ARIZONA BIRD COMMITTEE REPORT: 1996-1999 RECORDS 


ABC (as published in Western Birds), a reporting form for electronic 
submission of reports to the ABC, a selection of photographs of rarities from 
Arizona, and a varying selection of photographs and discussions concerning 
field-identification topics (e.g., the Black-capped Gnatcatcher). This site is a 
work in progress. 

The ABC encourages observers to submit documentation for species on 
the review list, as well as new species for Arizona. All material should be sent 
to Gary H. Rosenberg, ABC secretary, P. 0. Box 91856, Tucson, AZ 
85752-1856 (e-mail ghrosenberg@theriver.com). The committee would 
like to emphasize the importance of submitting sightings directly to the ABC 
for review. The listing of reports, including those with written descriptions, 
on local list-servers on the Internet may not make it to the ABC. Only those 
reports submitted to the ABC, or to the regional editor for North American 
Birds (who turns the material on review-list species over to the secretary of 
the ABC), will be considered by the committee. The ABC thanks the many 
observers from Arizona and around North America who have submitted 
their documentation of sightings to the ABC. 

Each record listed below includes a locality, county (abbreviation: see 
below), date (span normally as published in Field Notes/North American 
Birds), and initial observer if known. Additional observers who submitted 
reports, photographs, and recordings are also listed. All records are sight 
records unless noted otherwise with a symbol for a photograph, sound 
recording, or specimen. It has not been customary for the ABC to review 
individuals returning for multiple years, but these dates are normally included 
within the accounts. The ABC emphasizes that the listing of reports under 
Reports Not Accepted does not necessarily mean that the members of the 
ABC “do not believe” the observer, simply that the documentation supplied 
to the committee for evaluation was not detailed enough to substantiate the 
sighting as a record. 

The ABC’s abbreviations for counties in Arizona are APA, Apache; COS, 
Cochise; COC, Coconino; GIL, Gila; GRA, Graham; GRE, Greenlee; LAP, 
La Paz; MAR, Maricopa; MOH, Mohave; NAV, Navajo; PIM, Pima; PIN, 
Pinal; SCR, Santa Cruz; YAV, Yavapai; YUM, Yuma. Other nonstandard 
abbreviations commonly used within this report include *, specimen; B.T.A., 
Boyce Thompson Arboretum; L.C.R.V., lower Colorado River valley; N.I.R., 
Navajo Indian Reservation; N.M., national monument; N.W.R., national 
wildlife refuge; ph., photograph; P.A.P., Pinal Air Park; P.R.D., Painted 
Rock Dam; S.P.R., San Pedro River; S.T.P., sewage treatment plant; s.r., 
sound recording; UA, University of Arizona; v.t, video tape. 

RECORDS ACCEPTED 

RED-THROATED LOON Gauia stellata. A single individual was at L. Havasu City, 
MOH, 25 Jan 1998 (TC). There have been about ten previous accepted records for 
Arizona. 

LEAST GREBE Tachyhaptus dominicus. A single individual was at a small pond in 
E. Turkey Creek, Chiricahua Mts., COS, 1-2 Aug 1997 (ph. DF; see NAS Field Notes 
52:101), and another was at Willcox, COS, 3 Nov 1998-10 Jan 1999 (DM; ph. 
MMS). These represent the ninth and tenth state records. 


51 


ARIZONA BIRD COMMITTEE REPORT: 1996-1999 RECORDS 


LEACH’S STORM-PETREL Oceanodroma leucorhoa. Although no physical 
documentation was obtained, excellent written details were submitted for at least one 
individual seen during Tropical Storm Nora at L. Havasu City, MOH, 26 Sep 1997 
(BH). This tropical storm brought unprecedented numbers of both Black and Least 
Storm-Petrels to L. Havasu, L.C.R.V. This sighting represents one of the few inland 
records of Leach's Storm-Petrel for the western U.S. (see Jones 1999). 

BLACK STORM-PETREL Oceanodroma melania. After Tropical Storm Nora as 
many as 40 were at L. Havasu City, MOH, 26 Sep 1997, with at least three 
remaining there until 30 Sep 1997 (ph. TC, BGr, BH; Figure 1; see VJ. Birds 
30:187). These represent the first records of this species from Arizona (see Jones 
1999). 

LEAST STORM-PETREL Oceanodroma microsoma . After Tropical Storm Nora 
aproximately 200 were on L. Havasu, MOH, 26 Sep 1997, with several remaining 
there until 30 Sep (ph. TC, BGr, BH, m.ob.; *UA; see W. Birds 30:188). The 
photographs and two specimens provided the first physical documentation for this 
species in Arizona; there were two previous accepted sight reports from the state 
(Rosenberg et al. 1991). Similar storms have brought this species to the Salton Sea 
previously (see Jones 1999, Kaufman 1977). 

REDDISH EGRET Egretta rufescens. Accepted records are of one at P.R.D., 
MAR, 24 Aug 1997 (CBa), one at Avra Valley S.T.P., PIM, 17 Jul 1998 (ph. MMS), 
and one at Gila Farms Pond, MAR, 10-11 Aug 1998 (ph. RMJ), bringing the total 
number of accepted records for the state to nine. 

WHITE IBIS Eudocimus albus. One was at Nogales S.T.P., SCR, 6 Jul-15 Sep 
1999 (JSa; ph. MMS, GHR); there were five previous accepted records for Arizona. 



Figure 1. These were some of the nearly 40 Black Storm-Petrels found on Lake 
Havasu after Tropical Storm Nora 27 September 1998. 


Photo by Bill Grossi 


52 


ARIZONA BIRD COMMITTEE REPORT: 1996-1999 RECORDS 



Figure 2. This Pacific Golden-Plover at the Western Sod Farm near Arizona City 8- 
10 August 1998 provided a first Arizona record. 


Photo by George Hentz 


ROSEATE SPOONBILL Ajaia ajaja. Single individuals were at Gillespie Dam, 
MAR, 19-22 Jul 1997 (SPe, MMS, DAI) and at Picacho Res., PIN, 25 Sep-4 Oct 
1997 {DC; ph. MMS); these represent the fourth and fifth records since 1992. 

ROSS’S GOOSE Chen rossii. A stunning blue-morph individual of this species at 
Nogales, SCR, 28 Dec 1998 (MP, ph. MMS) represents the first sighting of this form 
in Arizona. 

BRANT Branta bernicla. Three at Cornville, YAV, 6 Apr 1998 (RR; ph. RR, 
MMS), with one remaining until 19 Apr, represents the first report from Yavapai 
County; there are fewer than ten records ever for Arizona, all of the subspecies 
nigricans. 

WHITE-WINGED SCOTER Melanitta fusca. Accepted records are of one at 
Castle Rock Shores, L.C.R.V., LAP, 27 Dec 1985 (SGw), one in Tempe, MAR, 13- 
31 Mar 1996 (RMJ), one at Ganado L., N.I.R., APA, 28 Nov 1997 (MMS, BHo, 
GHR), and one at Green Valley, PIM, 3-6 Nov 1998 (RP, ph. MMS). This species is 
now seen in Arizona nearly annually and has been removed as a review species 
(Rosenberg and Witzeman 1998). 

BLACK SCOTER Melanitta nigra. A single individual at Parker Dam, LAP, 1 Feb- 
14 Apr 1996 (RMJ; ph. MMS) represents only a seventh record for Arizona. 

RED-SHOULDERED HAWK Buteo lineatus. Accepted records are of one in s.w. 
Phoenix, MAR, 17 Dec 1985 (RWi), one in Tucson, PIM, 20 Feb 1997 (RH), and one 
in Scottsdale, MAR, 14 Apr 1998 (JB1). Although this species is reported in Arizona 
nearly annually, very few are properly documented (see under Reports Not Accepted). 
The ABC encourages observers to attempt to document all sightings of this species in 
the state carefully, as many of the written descriptions of Red-shouldered Hawks 


53 


ARIZONA BIRD COMMITTEE REPORT: 1996-1999 RECORDS 



Figure 3. Arizona’s third Northern Jacana was at Arivaca Lake 19 October 1998. 

Photo by Mark Stevenson 



Figure 4. Arizona's first Yellow-footed Gull was at Wahweep Marina, Lake Powell, 23 
April 1999. 


Photo by Gary H. Rosenberg 


54 


ARIZONA BIRD COMMITTEE REPORT: 1996-1999 RECORDS 



Figure 5. Arizona’s second documented Nutting’s Flycatcher spent the winter of 
1997-1998 at Patagonia Lake State Park. 

Photo by Larry Sansorte 


submitted to the committee for review fail to eliminate confusing species such as the 
Broad-winged Hawk. 

BROAD-WINGED HAWK Buteo platypterus. Accepted records are of one at 
Kearns Canyon, N.I.R., NAV, 23 May 1992 (CL; ph. JBu), one at Cave Creek 
Canyon, COS, 13 May 1994 (ph. BZ), one at Granite Creek near Prescott, YAV, 4 
May 1997 (CT), and one at Comville, YAV, 9 Apr 1998 (TL, IT). As with Red- 
shouldered Hawk, we receive reports of this species nearly annually, but many of 
these fail to exclude similar species such as the Red-shouldered and Gray (B. nitidus) 
hawks. We encourage observers to use care in identifying this species in Arizona. 

SHORT-TAILED HAWK Buteo brachyurus. A dark-morph individual photo- 
graphed on Miller Peak in the Huachuca Mts., COS, 31 July-4 Sep 1999 (ph, RH; 
ph. GHR) established a first documented record for Arizona. A light-morph individual 
seen earlier at the same location 26 July-4 Sep 1999 (RH) has been accepted as well, 
given that the dark bird was photographed. There were three previous sight reports 
from the state (see Rosenberg and Witzeman 1998), and these will now be re- 
evaluated. 

AMERICAN GOLDEN-PLOVER Pluvialis dominica. Reports of one at Western 
Sod Farm near Arizona City, PIN, 27 Sep 1999 (RH, MMS; ph. MMS) and one in s.w. 
Phoenix, MAR, 16 Oct 1999 (TC) were the only ones accepted. This species will 


55 


ARIZONA BIRD COMMITTEE REPORT: 1996-1999 RECORDS 



Figure 6. This male Black-capped Gnatcatcher was found nesting at Brown Canyon 
22 April 1997. 

Photo by Gary H. Rosenberg 


remain a review species even though it is found annually because of the difficulty in 
separating it from the very similar Pacific Golden-Plover. 

PACIFIC GOLDEN-PLOVER Pluvial is fulva. A single individual at Western Sod 
Farm near Arizona City, PIN, 6-12 Aug 1998 (RH; ph. GHe, MMS: s.r. RMJ; Figure 
2) provided Arizona with its first record. This species is a regular fall visitor to the 
California coast, with adults and one-year-old birds arriving during the first week of 
August; therefore, one in Arizona at that season, while unprecedented, was not 
unexpected. This sighting does represent one of the few noncoastal records of this 
species. 

NORTHERN JACANA Jacana spinosa. A stunning adult at Arivaca L., PIM, 15- 
23 Oct 1998 (EB: ph. GHR, MMS; Figure 3) provided Arizona with its third accepted 
record; the previous two records are both for June. 

WANDERING TATTLER Heteroscelus incanus. A well-described individual at 
Tucson S.T.P., PIM, 12 Sep 1991 (BL) was only the second for Arizona; the state’s 
only previous tattler was found 18 Sep 1971 in Phoenix (see Witzeman 1997). 

UPLAND SANDPIPER Bartramia longicauda. A single bird at Western Sod Farm 
near Arizona City, PIN, 16 Aug 1999 (ph. RH) provided one of only a few accepted 
Arizona records; interestingly, this species is regular as a fall migrant east of the 
continental divide as close to Arizona as Las Cruces, New Mexico. 

POMARINE JAEGER Stercorarius pomarinus. An adult was at Paloma Ranch 
near Gila Bend, MAR, 8-12 Oct 1999 (ph. BGr, MMS; v.t. GHR). There were only 
four previous Arizona records. 

PARASITIC JAEGER Stercorarius parasiticus. One individual was on L. Havasu, 
MOH, during Tropical Storm Nora 26 Sep 1997 (RMJ; see Jones 1999). There are 


56 


ARIZONA BIRD COMMITTEE REPORT: 1996-1999 RECORDS 



Figure 7. This Blue-winged Warbler was mist-netted along the San Pedro River near 
Hereford 17 July 1999, 

Photo by Lisa Walraven 



Figure 8. Arizona’s second documented Prairie Warbler was at Wahweep, Lake 
Powell, 30 November 1997. 


Photo by Garj ; H. Rosenberg 


57 


ARIZONA BIRD COMMITTEE REPORT: 1996-1999 RECORDS 


fewer than ten previous sight reports from Arizona, many of which have not been 
reviewed by the ABC (see Rosenberg and Witzeman 1998). Also, a jaeger of 
undetermined species was at the Bill Williams arm of L. Havasu, LAP, 26 Nov 1999 
(CBa); the description, while detailed enough to determine the bird to be a jaeger, was 
not complete enough to establish the species. 

LAUGHING GULL Larus atricilla. An adult was at Patagonia L., SCR, 13-15 
Aug 1998 (GHR). There have been fewer than fifteen previous Arizona records. 

MEW GULL Larus canus. Single individuals were at Lake Havasu City, MOH, 22 
Feb 1998 (RH) and at Davis Dam, MOH, 29 Jan 1999 (JPi). Virtually all of the 
previous Arizona records are from the Colorado River during winter. 

YELLOW-FOOTED GLJLL Larus livens. A subadult bird at Wahweap, L Powell, 
COC, 21-23 Apr 1999 (CL; ph. GHR, MMS; v.t. GHR; Figure 4; see N. Am. Birds 
53:344) provided Arizona with its first record, although this species is regular at the 
Salton Sea and at the northern end of the Gulf of California. This bird also 
represented the first Utah record, and there is at least one record from Nevada (Lake 
Mead). 

GLAUCOUS GULL Larus hyperboreus. An immature at Fisher’s Landing along 
the L.C.R.V. n. of Yuma, YUM, 28 Nov-29 Dec 1992 (JT) is only the third accepted 
Glaucous Gull for Arizona. 



Figure 9. Arizona’s first winter record of Rufous-capped Warbler was along the San 
Pedro River near Hereford in December 1998, 


Photo by Jim Burns 


58 



ARIZONA BIRD COMMITTEE REPORT: 1996-1999 RECORDS 


BLACK-LEGGED KITTIWAKE Risso tridactyla. Accepted records are of one at 
Parker Dam, LAP, 11-19 Nov 1978 (BW) and of one adult at Willcox, COS, 15 Nov 

1998 (JHa; ph. GHe). This species was more regular during the 1970s and 1980s; 
there have been only five sightings in Arizona since 1990. 

ARCTIC TERN Sterna paradisaea. A well-described immature from Many Farms 
L,, APA, 19 Sep 1993 (CL) furnished only a fourth Arizona record. 

BLACK-BILLED CUCKOO Coccyzus erythopthalmus. An individual photo- 
graphed in Pinery Canyon, Chiricahua Mts., COS, 7 Oct 1984 (ph. RHi) represents 
the first physically documented record for Arizona. The only other accepted record for 
the state was from nearby Portal on 3 Oct 1984 (see Rosenberg and Witzeman 1998), 
suggesting the possibility that both records involved the same individual. 

GROOVE-BILLED ANI Crotophaga sulcirosrtis. Accepted records are of one at 
Buckeye, MAR, 30 Jul 1985 (TC) and one at Lake Montezuma, COC, 24 Sep 1998 
(DHo). This species is still casual in Arizona and remains on the review list. 

BERYLLiNE HUMMINGBIRD Amazilia betyllina. Accepted records are of one in 
Ramsey Canyon, COS, 29 May-13 Jun 1998 (CMa), another there 18-20 Jul 1999 
(BCa), and one in Miller Canyon, COS, 5 Jul-early Aug 1999 (MMS). In recent years 
this species appears to have become a regular summer visitor to feeders in the 
Huachuca Mts. and likely breeds in canyons such as Ramsey and Miller in small 
numbers. Observers need to be careful with the identification of the Berylline because 
of the recent discovery of a hybrid Berylline x Magnificent Hummingbird in Arizona 
(Heindel and Howell 2000). 

PLAIN-CAPPED STARTHROAT Heliomaster constantii. One was in Madera 
Canyon, SCR, 7 Jul-31 Aug 1997 (NC; ph. MMS). There have been about 20 
sightings of this Mexican species in Arizona, ten of which have been reviewed by the 
ABC. w 

LUCIFER HUMMINGBIRD Calothorax lucifer. Accepted records away from the 
Portal region are of one in French Joe Canyon, COS, 19 Jul 1997 (NCr) and a pair, 
with a female sitting on a nest, in Chino Canyon, SCR, 13-20 Apr 1997 (NC,DNe; 
ph. MMS, GHR), providing only a second nesting report from the state. The Lucifer 
Hummingbird remains a review species for all reports away from the Portal area. 

EARED TROGON Euptilotis neoxenus. Accepted records are of one along the 
Black River, APA, 13 Jun 1992 (DF), one in Haunted Canyon, near Globe, PIN, 1 
Jan-14 Mar 1996 (DPi, MMa), and one in Cave Creek Canyon, COS, 10-28 Nov 

1999 (JBo; ph. JOl). The birds in Haunted Canyon and along the Black River were in 
obscure canyons probably never birded before, making one wonder how many Eared 
Trogons wander around Arizona without being found. 

RED-HEADED WOODPECKER Melanerpes erythrocephalus. One at the P.A.P. 
pecan grove, PIN, 25 Oct 1997-Apr 1998 (PS; ph. MMS) provided only a seventh 
documented record from the state (Rosenberg and Witzeman 1998). 

EASTERN WOOD-PEWEE Contopus virens. One singing individual in Madera 
Canyon, SCR, 23 Jun-20 Aug 1998 (MMS, ph. MMS, s.r. GHR, RMJ) represents 
only a third documented record for Arizona (see Rosenberg and Witzeman 1999). 

LEAST FLYCATCHER Empidonax minimus. One well-described individual at 
Arivaea L., PIM, 28 Nov 1996 (NCr) represents one of the few accepted reports of 
this species for the state. Given the difficulty in distinguishing the Least from the Dusky 
Flycatcher, the ABC still encourages observers to document all sightings of this species 
physically. 

NUTTING’S FLYCATCHER Myiarchus nuttingi. A well-documented individual at 
Patagonia Lake State Park 14 Dec 1997-21 Mar 1998 (WR; ph. GHR, LSa, MMS, BZ; 


59 


ARIZONA BIRD COMMITTEE REPORT: 1996-1999 RECORDS 


s.r. GHR, RMJ; v.t. CDB; Figure 5; see cover N. Am. Birds 52, no. 2; 52:148) 
represents only the second confirmed report of this species from the United States. The 
only other documented occurrence was of an individual collected at Roosevelt Lake, 
PIN, 8 Jan 1952 (Monson and Phillips 1981). The ABC has not reviewed one mist- 
netted near Elgin, SCR, 1 5 Jul 1985, but photos published (Bowers and Dunning 1987) 
suggest that the bird was likely a worn Ash-throated Flycatcher (M. cinerascens). 
Observers are encouraged to attempt to document any report of this species in Arizona 
with tape recordings of voice, which are useful for acceptance by the ABC, and not rely 
solely on difficult-to-ascertain visual characters such as color of mouth lining. 

WHITE-EYED VIREO Vireo griseus. The only accepted record was of one in 
Portal, COS, 19 Aug 1998 (GHR); there were eleven previous records for Arizona. 

BELL’S VIREO Vireo bellii. A report from the confluence of the Salt and Verde 
rivers n.e. of Phoenix, MAR, 26 Dec 1993 (TC) represents one of the few accepted 
winter sightings in the state. 

PHILADELPHIA VIREO Vireo philadelphicus. Accepted records are of one in 
French Joe Canyon, COS, 14 Apr 1996 (CCa) and one in s.w. Phoenix, MAR, 16 
Oct 1999 (TC). The April report marks the earliest ever for spring in Arizona. 

RED-EYED VIREO Vireo olivaceus. One at Kingfisher Pond along the upper San 
Pedro R., COS, 16 May 1997 (DK) represents the only accepted record. This species 
has certainly declined over the past 20 years, and, because of its similarity to the 
Yellow-green Vireo (Terrill and Terrill 1981), it remains a review species. 

YELLOW-GREEN VIREO Vireo flavoviridis. Two July reports accepted, of one at 
Guadalupe Canyon, COS, 11 Jul 1988 (NMC; ph. MMS; s.r. CDB) and one at 
Patagonia, SCR, 21-24 Jul 1999 (RH, JLD, MMS). Four of the five accepted Yellow- 
green Vireos in Arizona were discovered in late June or July (Rosenberg and 
Witzeman 1999). 

BLUE JAY Cyanocitta cristata. One individual at River Mile 175 in the Grand 
Canyon, COC, 7 May 1998 (CL; ph. NBr) was the sixth recorded in Arizona. 

CAROLINA WREN Thryothorus ludovicianus. A singing individual at Cook’s 
Lake along the San Pedro R. near Dudleyville, PIN, 18 Jun 1999-fall 2000 (TK, TC; 
ph. TC, MMS; s.r. TC, GHR, CDB) provided Arizona with its first record. This species 
has been expanding its range in Texas and New Mexico and was not unexpected. 

WINTER WREN Troglodytes troglodytes. Although the Winter Wren is not a 
review species, the committee did review the first summer reports from Arizona. One 
individual was along the w. fork of Oak Creek Canyon, COC, 25 Jun 1994 (FBr) and 
has been reported from there in subsequent years. Another individual was seen 
carrying food (presumably to a nest) along Christopher Creek, COC, 21 Jul 1999 
(KN, TC). The known breeding location nearest Arizona is northern California or 
northern Idaho, but a singing male has been seen in the Jemez Mts. in northern New 
Mexico in recent years (B. Howe pers. comm.). 

BLACK-CAPPED GNATCATCHER Polioptila nigiceps. Accepted records are of 
two pairs in Brown Canyon, PIM, 22 Apr 1997 (NC, RT; ph. GHR, MMS; s.r. GHR; 
v.t. CDB; Figure 6), with at least one of the pairs attempting nesting and remaining 
there until Jan 1998, one male at Patagonia, SCR, 25 May 1998 (DS, RT; s.r. DS), 
one female in Chino Canyon, SCR, 23 Feb 1999 (RH), and a pair in California Gulch, 
SCR, 23-31 Jul 1999 (RH; s.r. CDB; ph., s.r. DS). This species has been an irregular 
breeder in Arizona, with the first record in the early 1970s (Phillips et al. 1973) and 
the most recent in the late 1980s (possibly early 1990s). Extreme caution should be 
exercised in identifying this species in Arizona because of apparent hybridization with 
the Black-tailed Gnatcatcher ( P. melanura). 


60 


ARIZONA BIRD COMMITTEE REPORT: 1996-1999 RECORDS 


WOOD THRUSH Hglocichla mustelina. One was at Agua Caliente Park, Tucson, 
PIM, 25-27 Oct 1997 (WLe, MMS). There are fewer than fifteen accepted records for 
Arizona. 

RUFOUS-BACKED ROBIN Turdus rufopal/iatus. One was in Guadalupe Can- 
yon, COS, on the late dates of 3-4 Jun 1980 (CBr), one was at Kino Springs, SCR, 
14 Dec 1986-1 Apr 1987 (RBa), one was well north in Prescott, YAV, 10 Dec 1988- 
6 Jan 1989 (LF), one was along Sonoita Cr. below Patagonia L., SCR, 19 Dec 1992 
(GH), one was netted along the San Pedro R. near Sierra Vista, COS, on the very late 
date of 5 Jun 1996 (ph. DK), one wintered along the San Pedro R. near Hereford, 
COS, 5 Dec 1998-2 Jan 1999 (JLe; v.r. CDB), two were at the B.T.A., PIN, 19-30 
Jan 1998 (CT, m.ob.), and one was at the Arizona-Sonora Desert Museum, PIM, 4- 
6 Nov 1999 (JKr; ph. MMS). This species is one of the more regular of the Mexican 
strays to occur in Arizona; the June record is the first in summer for the state. The 
Rufous-backed Robin is no longer a review species. 

BOHEMIAN WAXWING Bombycilla garrulus. A large flock of at least 150 
individuals in Flagstaff, COC, from early February into March {with one remaining to 
10 Apr 1984) was photographed 15 Mar 1984 (ph. JCo, TC). To our knowledge, this 
is the most recent sighting of this species in Arizona. 

BLUE-WINGED WARBLER Vermivora pinus. Accepted records are of one in 
Sycamore Canyon, SCR, 20 Dec 1998 (GSM), one along the Santa Cruz R. in 
Tucson, PIM, 13-28 Mar 1999 (RGr; ph. RMJ, MMS; v.t. GHR; see N. Am. Birds 
53:311), and one along the San Pedro R. near Hereford, COS, first seen 16 May 
1999 (HKo) and later (probably the same individual) netted 17 Jul (ph. MSM; Figure 
7). There were only five previous records for Arizona. 

GOLDEN-WINGED WARBLER Vermivora chrysoptera. An accepted report of 
one in Sabino Canyon, PIM, 15 Feb 1997 (CGo) provided one of the few winter 
records for the state. Other accepted records are of one along the San Pedro R. near 
Dudleyville, PIN, 18 Jun 1999 (TC) and one in Madera Canyon, SCR, 1-3 Nov 1999 
(ph. MMS). This species is still considered casual in Arizona with about 20 records. 

TENNESSEE WARBLER Vermivora peregrina. Records accepted by the commit- 
tee are of one along the S.P.R. s. of Charleston, COS, 24 Sep 1986 (DK), one along 
Granite Cr. near Prescott, YAV, 1 May 1997 (CT), and one in Cave Creek Canyon, 
COS, 24-26 May 1997 (CHo). The Tennessee Warbler has now been removed from 
the review list. 

MAGNOLIA WARBLER Dendroica magnolia. Accepted records are of one at the 
B.T.A., PIN, 29-30 Sep 1997 (RMJ) and one near Topock, MOH, 30 Nov 1997 
(DWi). There are fewer than 25 records for the state. 

CAPE MAY WARBLER Dendroica tigrina . Winter still appears to be the best 
season to find this warbler in Arizona; records accepted are of one in Patagonia, SCR, 
16 Dec 1984 (AMo), one at Fountain Hills, MAR, 22-24 Dec 1986 (RBr), and one 
at the Forty-Niners Country Club in n.e. Tucson, PIM, 28 Nov-3 Dec 1998 (RH, 
MMS). There are still fewer than 15 total records for Arizona. 

BLACK-THROATED GREEN WARBLER Dendroica virens. The only report 
submitted to the committee and accepted was of one at Portal, COS, 29 Oct 1993 
(ph. DJa). This species was once thought of as regular in Arizona, but this record 
represents only the second accepted record since 1990, with only a few reports not 
submitted to the committee. 

BLACKBURNIAN WARBLER Dendroica fusca. One accepted record of a male 
from along the S.P.R. e. of Sierra Vista, COS, 27 May 1999 (DEd). There are still fewer 
than 20 records for Arizona, about six of which have not been reviewed by the ABC. 


61 


ARIZONA BIRD COMMITTEE REPORT: 1996-1999 RECORDS 


YELLOW-THROATED WARBLER DencJroica dominica. One was at Reid Park in 
Tucson, PIM, 31 Oct-7 Dec 1997 (CGr; ph. GHR, MMS), representing about a 20th 
record for Arizona. 

PINE WARBLER Dendroica pinus. Winter is certainly the best season to find this 
warbler in the state; accepted reports were of one in Reid Park in Tucson, PIM, 26 
Dec 1997-15 Feb 1998 (ph. MMS), one along the Santa Cruz R., PIM, 5-25 Jan 

1998 (RBo; ph. RH), and one at Sweetwater Wetlands, Tucson, PIM, 23-26 Jan 

1999 (TC; ph. MMS). There were only four previously accepted records for Arizona 
(see Rosenberg and Witzeman 1999). 

PRAIRIE WARBLER Dendroica discolor. Amazingly, one was photographed at 
Wahweap, Lake Powell, COC, 30 Nov-1 Dec 1997 (ph. GHR, MMS; Figure 8), and 
another individual was found and photographed by the same observer the next day, 1 
Dec 1997, remaining until at least 11 Jan 1998, in Tucson, PIM (ph. MMS, GHR). 
There was only one previous documented record for the state. 

BAY-BREASTED WARBLER Dendroica castanea. The only accepted record is of 
one in Horseshoe Canyon, Chiricahua Mts., COS, 7 Nov 1998 (NMC). There were 
about 10 previously accepted records for Arizona. 

BLACKPOLL WARBLER Dendroica striata. One accepted record of one at 
Granite Reef Picnic Area along the Salt R. n.e. of Phoenix, MAR, 27 Oct 1991 (ph. 
SGa). A large percentage of the Arizona reports have not been submitted to the ABC. 

LOUISIANA WATERTHRUSH Seiurus motacilla. Accepted records are of one in 
Cave Creek Canyon, COS, 27 Dec 1992 (CSa), one in Sycamore Canyon, SCR, late 
Nov 1996-24 Jan 1997 (RH), one along the S.P.R. near Hereford, COS, 10 Sep 
1998 (DK), one in Sycamore Canyon, SCR, 1 Feb 1999 (ph, MMS), and one along 
Queen Creek near Superior, PIN, 7 Nov 1999 (MMS). It has been well demonstrated 
that this species is a rare but regular winter visitor to rocky streams in s.e. Arizona. It 
has been removed as a review species by the ABC, but sketch details are still requested 
for inclusion of reports in N. Am. Birds. 

KENTUCKY WARBLER Oporornis formosus. Accepted records are of one at 
Kino Springs, SCR, 28 Oct 1978 (GHR), one in Flagstaff, COC, 1 1 May 1997 (FBr), 
one in Scottsdale, MAR, 19 Nov 1997 (JB1), and one at the Roger Road Wastewater 
facility, Tucson, PIM, 8-15 Oct 1998 (ph. MMS). This species has been removed as 
a review species in Arizona, but sketch details are requested for inclusion of reports in 
N. Am. Birds. 

MOURNING WARBLER Oporornis Philadelphia. A well-described individual 
from Navajo, NAV, 1 Oct 1988 (DS) made the only accepted record. There are only 
five previously accepted records of the Mourning Warbler in Arizona, only one of 
which has been documented with a photograph or specimen. 

FAN-TAILED WARBLER Euthiypis lachrymosa. A singing male at Patagonia, 
SCR, 21-27 May 1997 (JBo, WR, BPr; s.r. DS) provided Arizona with its sixth 
record. 

RUFOUS-CAPPED WARBLER Basileuterus rufifrons. Two winter reports were 
accepted, of one along the S.P.R. near Hereford, COS, 4-30 Dec 1998 (MPr; ph 
JBu, GHR; Figure 9) and one in lower Sycamore Canyon, SCR, 23 Dec 1998 (ph. 
GHR, CDB). These represent the first winter records for Arizona and the ninth and 
tenth records overall. 

SCARLET TANAGER Piranga oliuacea. Reports of one breeding-plumaged male 
at the Hassayampa River Preserve near Wickenburg, MAR, 26 Apr 1998 (EHo) and 
one at Paloma, MAR, 25 Oct 1998 (RMJ) were the only ones accepted. There are 


62 


ARIZONA BIRD COMMITTEE REPORT: 1996-1999 RECORDS 


only 14 accepted records for the state, with an additional eiqht reports not submitted 
to the ABC. 

FLAME-COLORED TANAGER Piranga bidentata. Two accepted records of 
females were of one in Miller Canyon, COS, 29 May 1997 (TC) and one nesting, and 
apparently mated with a Western Tanager (P. ludoviciana), in Whitetail Canyon, 
COS, 16 May-mid-Jun 1998 (RT; ph. MMS). There have been only five previously 
accepted records for Arizona, three of which were of nesting pairs. The first recorded 
male, in April 1985 (Morse and Monson 1985), was also mated with a Western 
Tanager, and a hybrid male tanager at Bog Spring in Madera Canyon (1995 and 
1996) showed characteristics of both the Flame-colored and Western tanagers. 

FIELD SPARROW Spizella pusilla. One alona the S.P.R. near Hereford, COS, 12 
Dec 1998-21 Feb 1999 (CSm, CDB, GHR; ph. MMS, GHR) provided only a third 
Arizona record. 

LAPLAND LONGSPUR Calcarius lapponicus. Accepted records are of one near 
Hfrida, COS, 20 Feb-4 Mar 1984 (AMo) and one near Topock, L.C.R.V., MOH, 27 
Nov 1999 (RMJ). This species is still considered casual in Arizona with fewer than ten 
accepted records in all. 

RUSTY BLACKBIRD Euphagus carolinus. One accepted record of a bird seen on 
both sides of the Colorado R. near Ehrenberg, LAP, 24 Nov 1988 (ph. PL). 

COMMON GRACKLE Quiscalus quiscula. One on the Salt River Indian Res., 
MAR, 21 Dec 1987 (RBr) represents only the eighth accepted record for the state. A 
very low percentage of reports from Arizona have been submitted to the ABC, 

ORCHARD ORIOLE Icterus spurius. One female remained at Fountain Hills, 
MAR, 2 Jan-10 Feb 1992 (JSo, ph. SFi, PL). Additional accepted records are of one 
at the Phoenix Zoo, MAR, 20-24 Dec 1997 (RMJ; ph. RD) and one in the Avra 
Valley, PIM, 31 Jan-28 Feb 1998 (HMc; ph, MMS), As with the previous species, 
very few of the Arizona reports have been written up and submitted to the ABC. The 
Orchard Oriole is still casual in the state at best and remains a review species. 

STREAK-BACKED ORIOLE Icterus pustulotus. One well-described individual 
from the bottom of the Grand Canyon (RM 246), MOH, 22 Jan 1998 (CL) 
represented the first n. Arizona record. Critical in the identification (distinguishing the 
Streak-backed from the immature Bullock’s Oriole) was a combination of reddish 
orange in the face and a lack of orange-yellow in the tail (see Kaufman 1983 for a 
discussion of this complex). This record is only the fourth accepted of this species 
away from the lower S.P.R. , where it has bred in the recent past. 

BALTIMORE ORIOLE Icterus galbula. Accepted records are of one male in Chino 
Valley, YAV, 3 May 1991 (LMu), one male 6 mi. n. of Camp Verde, YAV, 29 Apr 
1998 (TLi), and another male in s.w. Phoenix 30 May-11 Jun 1998 (RWi, RMJ). 
There have been very few recent reports of this species from Arizona. Because of the 
difficulty of distinguishing females from Bullock’s Oriole this species remains on the 
review list. 

REPORTS NOT ACCEPTED 

RED-THROATED LOON Gauio stellate. The description of one seen at a great 
distance at P.R.D., MAR, 13 Jan 1993 did not rule out a Pacific Loon (G. pacified). 

LEAST GREBE Tachybaptus dominicus. Details of one reported in Scottsdale. 
MAR, 16 Apr 1998 did not rule out the Eared ( Podiceps auritus) or Pied-billed 
(Podilymbus podiceps ) grebes. 


63 


ARIZONA BIRD COMMITTEE REPORT: 1996-1999 RECORDS 


LEACH’S STORM-PETREL Oceanodroma leucorhoa. One reported after Tropi- 
cal Storm Nora from Lake Havasu, MOH, 27 Sep 1997 lacked details specific 
enough to rule out other storm-petrel species. A photograph submitted with the 
report was intriguing, but most committee members believed that a pale area 
appearing on the rump was a photographic artifact and that the bird in the photo was 
likely a Black Storm-Petrel. 

YELLOW-CROWNED NIGHT-HERON Nyctanassa violacea. An immature re- 
ported from Kino Springs, SCR, 10 Jul 1992 may have been correctly identified, but 
the description was not complete enough to rule out a Black-crowned Night-Heron 
( Nycticorax nycticorax). Furthermore, the observer was not confident of the identifi- 
cation, a refreshing admission! 

WHITE IBIS Eudocimus albus. A brief description of two at P.R.D., MAR, 23 Jul 
1995 was not detailed enough for acceptance, and the sighting coincided with the 
occurrence of a Sacred Ibis ( Threskiornis aethiopicus) there. A report of three adults 
at Mittry L. near Yuma, YUM, 26 Sep 1998 was thought by many of the committee 
(after two rounds) to be correct, but three members were still troubled by the lack of 
description of bill and leg characteristics. 

RED-SHOULDERED HAWK Buteo lineatus . Single reports from near Chino 
Valley, YAV, 9 Jan 1990, from Picacho Res., PIN, 5 Aug 1992, from near Mayer, 
YAV, 21 Feb 1999, and from near Liguarta, Gila Mts,, YUM, 15 Apr 1999, were all 
lacking details specific enough to eliminate similar species. 

BROAD-WINGED HAWK Buteo platypterus. An immature hawk reported from 
Patagonia, SCR, 2 Dec 1998 was described as lacking a “dark trailing edge border” 
to the wing, which committee members thought better fit a Gray Hawk. 

WHITE-TAILED HAWK Buteo albicaudatus. Two reports, one from near Kirkland, 
YAV, 30 Dec 1996, the other from the Empire Cienega Ranch, SCR, 6 Mar 1997 
were not detailed enough to accept. Because of the lack of substantiated reports this 
century (see Rosenberg and Witzeman 1998), the committee requires at least 
photographic evidence accompanying any report of this species. 

APLOMADO FALCON Falco femoralis. A description of a falcon from n.e. of 
Douglas, COS, 24 Sep 1999 lacked the detail to substantiate this sighting as the first 
credible record of the Aplomado Falcon for Arizona since 1940 (Monson and Philips 
1981). As the number of valid records from New Mexico has been on the increase in 
recent years, it is only a matter of time until an Aplomado Falcon is sighted again in 
Arizona. The ABC encourages that any report of this species include some sort of 
physical documentation. 

SAGE GROUSE Centrocercus urophasianus. One male reported near Globe, 
PIN, 15 Sep 1995 was thought perhaps to have been a raptor; the closest that the 
Sage Grouse occurs to Arizona is central Colorado. 

ELEGANT QUAIL Callipepla douglasii. An adult present for several months in 
Douglas, COS, first seen 17 Jan 1998, was considered by the committee to likely 
have been an escapee. As this species ranges north to within 100 miles of Arizona’s 
border with Sonora, it is possible that one could occur naturally in Arizona. The 
committee would like to see a pattern of vagrancy before accepting one as a first 
Arizona (and United States) record. 

AMERICAN GOLDEN-PLOVER Pluvialis dominica. The description of one 
reported from Willcox, COS, 16 Sep 1992 lacked sufficient detail to be accepted. 

YELLOW-FOOTED GULL Larus livens. An adult gull with a “black” mantle was 
described from Willow L., YAV, 8 May 1999, but the bird was seen at a distance of at 


64 


ARIZONA BIRD COMMITTEE REPORT: 1996-1999 RECORDS 


least 3000 feet, so critical features that would eliminate other dark-mantled species 
were not discerned. 

WESTERN GULL La rus occidentalis. A description of a juvenile or first-winter gull 
from Willow L., YAV, 16 Dec 1998 lacked enough critical detail to determine the 
species and did not eliminate a Herring Gull (L. argentatus smithsonianus). 

BLACK-BILLED CUCKOO Coccyzus erythropthalmus. An intriguing report of 
one near Carrizo, NAV, 26 May 1995 included a fairly good description, but the call 
was said to be a series of six to eight “coo” notes, incorrect for this species. 

GROOVE-BILLED ANI Crotophaga sulcirostris. The description of one reported 
from Camp Verde, YAV, 2 May 1998 lacked enough detail for acceptance. 

BUFF-COLLARED NIGHTJAR Caprimulgus ridgwayi. A very interesting report 
of a nightjar on the road at night (seen in the headlights) in Guadalupe Canyon, COS, 
28 Jan 1995 was perhaps of this species, but the details did not rule out other 
nightjars (particularly the Whip-poor-will, C. vociferus) which, at times, can show a 
tawny collar. As this record would be a first in winter for the U.S., the ABC was unable 
to accept it without more substantiation. 

BERYLLINE HUMMINGBIRD Amazilia beryllina. The following reports of this 
species lacked detail sufficient for acceptance: one in Guadalupe Canyon, COS, 27 
May 1985, one at Patagonia, SCR, 8 Jun 1985, one in Cave Creek Canyon, COS, 
31 Jul 1996, and one in Madera Canyon, SCR, 6 Aug 1998. It should be noted once 
again (see Rosenberg and Witzeman 1998) that virtually all Arizona records for this 
species are from a narrow elevation zone (5000-6000 feet) within oak woodland. 

EARED TROGON Euptilotis neoxenus. The description of a trogon from open 
pine-oak habitat at 7000 feet in the Huachuca Mts. (exact location not mentioned!), 
COS, 21 May 1989 certainly suggested this species, but at least three committee 
members believed the report was too marginal for acceptance. 

YELLOW-BELLIED FLYCATCHER Empidonax flavtventris. A report of one 
from Spruce Mt. near Prescott, YAV, 20 Aug 1991 was not detailed enough to 
determine the species within this difficult complex of flycatchers. 

LEAST FLYCATCHER Empidonax minimus. A winter report from Buenos Aires 
N.W.R., PIM, 19 Dec 1993 did not eliminate other Empidonax species. Another 
mid-winter report of a silent individual from Patagonia, SCR, 20 Feb 1999 had too 
many inconsistencies with the Least Flycatcher (general shape and eye-ring shape); 
the committee felt that Hammond’s and Dusky flycatchers were not eliminated. 

NUTTING’S FLYCATCHER Myiarchus nuttingi. On the heels of a well-docu- 
mented Nutting’s Flycatcher from Patagonia Lake during the winter of 1997-98 (see 
under Accepted Records), two additional reports of this species were received, one 
from Kino Springs, SCR, 29 Dec 1998, the other from Patagonia L., SCR, 11 Jan 
1999. In both cases, vocalizations were not described well enough (not heard in the 
Patagonia L. bird) to rule out the Ash-throated Flycatcher. Furthermore, other 
observers reported an Ash-throated Flycatcher from Kino Springs during the same 
period that showed a tail pattern more suggestive of a Nutting’s. The separation of 
Nutting’s and Ash-throated Flycatchers is subtle enough that only well-documented 
(tape-recorded or extensively photographed) individuals are acceptable to the ABC. 

GREAT CRESTED FLYCATCHER Myiarchus crinitus. A sketchy report of this 
species from along the Hassyampa R. near Wickenburg, MAR, 5 Sep 1998 was felt 
to have not dealt properly with the range of variation within this genus; most 
committee members felt that Brown-crested Flycatcher (M. tyrannulus ) was not 
eliminated. 


65 


ARIZONA BIRD COMMITTEE REPORT: 1996-1999 RECORDS 


GREAT KISKADEE Pitangus sufphuratus. An honest account of a strange 
flycatcher at the Grand Canyon, COC, 2 Aug 1995 was thought to apply to this 
species, but the description was not detailed enough to accept this report as a third 
Arizona record. 

FORK-TAILED FLYCATCHER Tyrannus savana. The description of one reported 
from along the S.P.R. near Dudleyville, PIN, 2 Aug 1995 lacked detail sufficient to 
substantiate a first Arizona record. 

WHITE-EYED VIREO Vireo griseus. A report from Safford, GRA, 25 JUN 1996 
did not include enough detail for acceptance. 

YELLOW-THROATED VIREO Vireo flavifrons. Details of one from s. of Willcox 
1 Apr 1993 were not sufficient to accept a sighting of this species nearly a month 
earlier than the next earliest record in s.e. Arizona. 

BLUE-HEADED VIREO Vireo solitarius. A description of an individual reported 
from Phoenix, MAR, 8 Jan 1990 did not adequately rule out Cassin’s Vireo. The 
committee urges extreme caution in the differentiation of these two species (see 
Heindef 1996). 

RED-EYED VIREO Vireo olivaceus. A report from Chiricahua N.M., COS, 4 Oct 
1985 did not rule out other vireos, such as the Yellow-green. 

YELLOW-GREEN VIREO Vireo flauoviridis. The details of one reported from 
Cave Creek Canyon, COS, 9 Mar 1982 were insufficient to support a sighting far 
outside the span of dates of the species’ regular occurrence in Arizona. 

BLACK-CAPPED GNATCATCHER Polioptila nigriceps. A single individual re- 
ported from Sabino Canyon, P1M, 18 Apr 1998 was well away from areas of previous 
occurrence, and the description did not rule out the common Black-tailed Gnat- 
catcher. Another report of one wintering in Sycamore Canyon, SCR, 26 Dec 1985 
was almost certainly correct (there are accepted reports from that locality from before 
and after this sighting), but the details submitted did not substantiate the record. One 
additional report from Sycamore Canyon 4 Jun 1994 also lacked the critical detail 
necessary to eliminate the Black-tailed or a hybrid between the Black-capped and 
Black-tailed. A male at Chino Canyon, SCR, from April 1995 through 1999 is 
suspected of being such a hybrid. Its head pattern was perfect for a Black-capped, but 
the amount of white in the tail, the bill length, and the vocalizations all appeared 
intermediate between the two species. It was originally found with a female Black- 
capped but was subsequently found mated with a female Black-tailed. 

VEERY Catharus fuscescens. Reports of individuals at Watson Wood near 
Prescott, YAV, 4 Apr 1997, at Madera Canyon, SCR, 6 Apr 1997, and at the 
Hassayampa River Preserve near Wickenburg, MAR, 17 May 1998 all lacked detail 
sufficient to eliminate other species in this complex genus. There is only one 
physically documented record of the Veery in Arizona away from historic nesting 
areas near Springerville (Rosenberg and Witzeman 1999). 

WOOD THRUSH Hylociehla musteline. One reported from Ramsey Canyon, 
COS, 18 May 1997 was not described adequately for acceptance. 

RUFOUS-BACKED ROBIN Turdus rufopalliatus . Reports of individuals at Here- 
ford, COS, 28 Mar 1981, at Tucson Mt. Park, PIM, 21 Dec 1988, and in Huachuca 
Canyon, COS, on the late date of 5 Jun 1999 were likely correct, but the details 
supplied did not substantiate the reports. 

BLUE MOCKINGBIRD Melanotis caerulescens. An individual seen in n.w. 
Tucson, PIM, 23 Sep 1992 was not accepted mainly because the observer felt it was 
likely an escapee and because the observation time was so short. The committee 


66 


ARIZONA BIRD COMMITTEE REPORT: 1996-1999 RECORDS 


believed it better to wait for additional early fall reports before perhaps reevaluating 
this report. 

RED-THROATED PIPIT Anthus ceruinus. A very intriguing report from two 
highly experienced observers of a calling individual heard only at the P.A.P. pecan 
grove 26 Oct 1991 was not accepted because several members of the committee (as 
well as the observers themselves) felt uncomfortable accepting such a rarity in the state 
(second Arizona record) on the basis of call alone. Whereas many on the committee 
had little doubt that the bird was correctly identified, they felt that an identification 
based solely on a call and descriptions written several years after the “sighting” did not 
constitute an acceptable record. It should be noted that the fall of 1991 was far and 
away the best ever for Red-throated Pipits in California, with several found at desert 
oases as in Kern County and Death Valley. 

GRAY SILKY-FLYCATCHER Ptilogonys cinereus. An interesting report of two 
adults and one juvenile along Ruby Road w. of Sycamore Canyon, SCR, 10 Aug 1993 
did not include enough information to support a first state record. Also, the brief 
descriptions of both the juvenile and female did not support the identification. 

BLUE-WINGED WARBLER Vermiuora pinus. A report of one from Tempe, 
MAR, 14 May 1995 was likely correct but not thorough enough for the record to be 
accepted. A description of one near Hereford, COS, 1 May 1998 also lacked enough 
critical detail for acceptance. 

TENNESSEE WARBLER Vermiuora peregrina. Reports not accepted by the ABC 
are of one from Yuma, YUM, 20 Dec 1986, one from the Empire Cienega Ranch, 
SCR, 6 May 1993, and one from along Granite Creek near Prescott, YAV, 14 Sep 
1994. 

BLACK-THROATED GREEN WARBLER Dendroica uirens. A report of one from 
Phoenix, MAR, 13 Aug 1981 did not adequately eliminate similar species such as 
Townsend’s (D. townsendi) and Hermit (D. occidentalis) warblers. 

BLACKBURNIAN WARBLER Dendroica fusca. A report of one in Sawmill 
Canyon, COS, 20 May 1999 was not detailed enough for acceptance. 

WORM-EATING WARBLER Helmitheros uermiuorus. A report from below Glen 
Canyon Dam, COC, 15 Jun 1992 was too brief for acceptance. This species has been 
removed as a review species. 

KENTUCKY WARBLER Oporornis formosus. Two reports were not accepted, 
one from Beaver Dam Creek, MOH, 24 May 1997 and one from near Kirkland, 
MOH, 22 Jun 1997. 

MOURNING WARBLER Oporornis Philadelphia. A report from Florida Wash, 
PIM, 31 Aug 1993 lacked certain critical features necessary to substantiate it. 

SLATE-THROATED REDSTART Myioborus miniatus. An odd report of one 
associating closely with a juvenile Painted Redstart (that apparently wanted to be fed) 
in Cave Creek Canyon, COS, 18 Jun 1995 lacked certain critical characters. Perhaps 
not coincidentally, a researcher at the Southwest Research Station in Cave Creek 
Canyon was apparently coloring the white parts of Painted Redstarts in the canyon, 
possibly accounting for the report. 

SCARLET TANAGER Piranga olivacea. A winter report from near the confluence 
of the Salt and Verde rivers, MAR, 10 Dec 1988 was likely correct, but the observer did 
not note characters to eliminate a greenish-yellow Summer Tanager definitively. There 
is only one winter record accepted for the state (see Rosenberg and Witzeman 1999). 

RED-HEADED TANAGER Piranga ergthrocephala. A male reported from Carr 
Canyon, COS, 14 May 1996 was, from the description of voice and plumage, 


67 


ARIZONA BIRD COMMITTEE REPORT: 1996-1999 RECORDS 


thought by most on the committee to likely have been a one-year-old Summer 
Tanager with a yellow-green body and red head. Some form of physical documenta- 
tion is needed for acceptance of a first Arizona (and United States) record. 

AMERICAN TREE SPARROW Spizelia arborea. Two reports, one from near 
Gadsden s. of Yuma, YUM, 20 Jan 1981, the other from Gila Farms Pond s. of 
Phoenix, MAR, 25 Dec 1994 were both too sketchy for acceptance. There still 
remains only one accepted southern Arizona record (see Rosenberg and Witzeman 
1999). 

FIELD SPARROW Spizelia pusilla. A report of a flock of five from Kitt Peak, PIM, 
24 Sep 1997 clearly pertained to some other sparrow, perhaps the Chipping. 

LAPLAND LONGSPUR Calcarius lapponicus. One reported from near Elfrida, 
COS, 4 Feb 1997 was likely correctly identified, but the details were too sketchy for 
acceptance. 

COMMON GRACKLE Quisca/us quiscula. A report of an individual seen only in 
flight at Springerville, APA, 26 Jun 1994 was almost certainly correct, but, given that 
the bird was never seen perched, the committee felt the report was best left not 
accepted. 

ORCHARD ORIOLE Icterus spiurus. Descriptions of one at Cave Creek Canyon, 
COS, 24 Apr 1994 and one at Madera Canyon, SCR, 21 May 1996 were both too 
brief to be accepted. 

BALTIMORE ORIOLE Icterus galbula. Two reports of females lacked enough 
detail to eliminate the similar Bullock’s Oriole: one at Springerville, APA, 24 Sep 
1985 and one at Prescott, YAV, 19 Aug 1987. 

PINE GROSBEAK Pinicola enucleator. A report of three from Sedona, YAV, 31 
Jan 1998 was intriguing but lacked enough detail for acceptance of an extralimital 
record of this species in Arizona. Similarly, a report from Glendale, MAR, 9 Mar 1999 
was not conclusive enough to support a first lowland record for the state. 

PURPLE FINCH Carpodacus purpureus. One reported from Sedona, YAV, 1 Apr 
1999 did not eliminate the similar Cassin’s Finch (C. cassinii). 

CORRIGENDA 

The following records were inadvertently published in the previous Ari- 
zona Bird Committee reports (Rosenberg and Witzeman 1998 and 1999) as 
“accepted” when in fact they were not accepted by the committee. The ABC 
apologizes for the errors: Eurasian Wigeon (Anas penelope) at the 
confluence of the Salt and Verde rivers, MAR, 10 Dec 1991; Red-shoul- 
dered Hawk 16 mi s. of Chino Valley, YAV, 9 Jan 1990; Red-shouldered 
Hawk at Picacho Res., PIN, 5 Aug 1992; Broad-winged Hawk at Barfoot 
Lookout, Chiricahua Mts., COS, 6 Aug 1995; Purple Gallinule ( Porphyrula 
martinica ) at Gila Farms Pond, MAR, 1 Aug 1991; American Golden- 
Plover at Willcox, COS, 18 Sep 1992; Tennessee Warbler at the Empire 
Cienega, SCR, 6 May 1993; Orchard Oriole in Cave Creek Canyon, COS, 
24 Apr 1994. 

CONTRIBUTORS 

D. Alexander, C. Babbitt, D. Bailey, J. Bartley (JB1), J. Bates, R. Bates, J. Bealer, 
C. D. Benesh, E. Bennett, T. Bishop, J. Bock, R. Bowers, R. Bradley, J. Braley, F. 


68 


ARIZONA BIRD COMMITTEE REPORT: 1996-1999 RECORDS 


Brandt, C. Brown, M. Brown, N. Brown, K. Brunson, J. Burns, W. Cady, S. 
Capawana, B. Carrell, C. Cathers, J. Chace, R. Chapman, D. Clark, M. Collie, J. 
Coons, T. Corman, N. Crook, D. Crumb, V. Dayhoff, B. Demaree, K. Diem, R. Ditch, 
J. L. Dunn, D. Edwards, F. Fekel, R. Ferguson, S. Finnegan, D. Fischer, L. Frederick, 

S. Ganley, B. Given, S. Goldwasser, S. Goodchild, C. Gordon, C. Green, R. 
Grohman, B. Grossi, P. Hammerton, J. Hand, T. & J. Heindel, G. Hentz, R. Hicks, 
J. Hirth, C. Hohenberger, D. Hook, E. Hough, B. Howe, R. Hoyer, C, Hunter, B. 
Jackson, D. Jasper, B. Johnson, R. Jones, C. Kangas, B. Kerr, C. Kolesar, H. Koons, 

T. Koronkiewicz, J, Krebs, J. Kreitzer, D. Krueper, C. LaRue, P. Lehman, W. Leitner, 
J. Levine, T. Linda, R. Long, B, Lyon, M. Mammoser, C. Marrantz, B. McAneny, L. 
McCloskey, A. McCready, H. McCrystal, V. Miller, G. S. Mills, S. Mlodinow, G. 
Monson, A. Moorhouse, N. Moore-Craig, D. Morrison, P, Moulton, L. Muelbach, G. 
Nation, D. Nelson, K. Newlon, P. Norton, J, Oldenettel, J. Owen, R. Palmer, J. Paris, 
D. Pearson, S. Peterson, D. Pierce, L. Piest, J. Pike, B. Pranter, M. Pretti, J. Price, B, 
Principe, R. Radd, G. H. Rosenberg, R. Ray, W. Russell, J. Saba, P. Salomon, C. 
Sandell, M. SanMiguel, L. Sansone, M. Sennett, C. Sherman, J. Simon, C. Smith, M. 
Sogge, J. Sommers, D. Stejskal, M. M. Stevenson, B. Sutton, D. Taylor, J. Taylor, R. 
Taylor, B. Thomen, I. Tolinson, C. Tomoff, M. Vandzura, J. Whetstone, B. Whitney, 
D. Wight, R. Witzeman, B. Wotten, B. Zimmer. 

ACKNOWLEDGMENTS 

I thank the many observers who submitted their sightings to the ABC for review; 
Arizona ornithology has benefited greatly from their efforts. Chris Benesh, Roy 
Jones, Narca Moore-Craig, Mark Stevenson, and Janet Witzeman all contributed 
greatly to the improvement of the manuscript. 

LITERATURE CITED 

American Ornithologists’ Union. 1998. Check-list of North American Birds, 7th ed. 
Am. Ornithol. Union, Washington, D.C. 

Bowers, R., and Dunning, J. B. 1987. Nutting’s Flycatcher in Arizona. Am. Birds 
4L5-10. 

Heindel, M. 1996. Field identification of the Solitary Vireo complex. Birding 28:458- 
471. 

Heindel, M., and Howell, S. N. G. 2000. A hybrid hummingbird in southeast Arizona. 
W. Birds 31:265-266. 

Jones, R. M. 1999. Seabirds carried inland by tropical storm Nora. W. Birds 30: 185- 
192. 

Kaufman, K. 1977. The changing seasons. Am. Birds 31:141-152. 

Kaufman, K. 1983. Identifying Streak-backed Orioles: A note of caution. Am. Birds 
37:140-141. 

Monson, G., and Phillips, A. R. 1981. Annotated Checklist of the Birds of Arizona, 
2nd ed. Univ. of Ariz. Press, Tucson. 

Morse, R., and Monson, G. 1985. Flame-colored Tanager in Arizona. Am. Birds 
39:843-844. 

Phillips, A. R., Speich, S., and Harrison, W. 1973. Black-capped Gnatcatcher, a new 
breeding bird for the United States; with a key to the North American species of 
PolioptUa. Auk 90:257-262. 

Rosenberg, G. H., and Stejskal, D. 1994. The Arizona Bird Committee’s Field 
Checklist of the Birds of Arizona. Ariz. Bird Committee, G. H. Rosenberg, P. O. 
Box 91856, Tucson, A Z 85752-1856. 


69 


ARIZONA BIRD COMMITTEE REPORT: 1996-1999 RECORDS 


Rosenberg, G. H., and Witzeman, J. L. 1998. Arizona Bird Committee Report, 
1974-1996: Part 1 (nonpasserines). W. Birds 29:199-224. 

Rosenberg, G. H,, and Witzeman, J, L. 1999, Arizona Bird Committee Report, 
1974-1996: Part 2 (passerines). W. Birds 30:94-120. 

Rosenberg, K. V., Ohmart, R. D., Hunter, W. C., and Anderson, B. W. 1991. Birds 
of the Lower Colorado River Valley. Univ. of Ariz. Press, Tucson. 

Speich, S,, and Parker, T. A. 1973. Arizona Bird Records, 1972. W. Birds 4-53-57. 

Speich, S. M., and Witzeman, J. L. 1975. Arizona Bird Records, 1973, with 
additional notes. W. Birds 6:145-155. 

Terrill, S. B., and Terrill, L. 1981. On the field identification of Yellow-green, Red- 
eyed, Philadelphia, and Warbling Vireos. Continental Birdlife 2:144-149. 

Witzeman, J., Demaree, S., and Radke, E. 1997. Birds of Phoenix and Maricopa 
County, Arizona. Maricopa Audubon Soc,, Phoenix. 

Accepted 29 September 2000 


70 


RECOLONIZATION OF THE FLICKER AND OTHER 
NOTES FROM ISLA GUADALUPE, MEXICO 

PAUL R. SWEET and GEORGE F. BARROWCLOUGH, Department of Ornithol- 
ogy, American Museum of Natural History, Central Park West at 79th Street, New 
York, New York 10024 

JOHN T. KLICKA, J. F. Bell Museum of Natural History, University of Minnesota, St. 
Paul, Minnesota 55108 (current address Barrick Museum, University of Nevada, Las 
Vegas, Nevada 89154) 

LILIANA MONTANEZ-GODOY and PATRICIA ESCALANTE-PLIEGO, Instituto de 
Biologia, Universidad Nacional Autonoma de Mexico, Apartado Postal 70-153, 
04510 Distrito Federal, Mexico 

ABSTRACT: During a visit to Isla Guadalupe from 31 May to 3 June 1996, we 
documented three species new to the island, the Barn Owl, Swainson’s Thrush, and 
Hooded Oriole, and established first breeding records for the European Starling and 
Western Meadowlark. Red-shafted Flickers are now breeding on the island, represent- 
ing a recent recolonization from the mainland following the extinction of the endemic 
population. We investigated the validity of Colaptes auratus rufipileus and con- 
cluded that it does not meet the standard for phylogenetic species but differs from C. 
a. collaris at the 75% level usually associated with subspecific rank. Damage to the 
cypress forest by goats continues, and all species dependent on these trees are 
threatened by loss of habitat. 

Isla Guadalupe is a volcanic oceanic island lying 260 km west of the coast 
of Baja California (29° 00' N, 118° 20' W). It is some 37 km long with an 
area of 250 km 2 and reaches an elevation of 1298 m. The history and status 
of the avifauna have been reviewed by Howell and Cade (1954) and more 
recently by Jehl and Everett (1985). Subsequent observations have been 
reported by Dunlap (1988), Oberbauer et al. (1989), Mellink and Palacios 
(1990), Howell and Webb (1992), and Gallo and Figueroa (1996). Here we 
report on birds observed and collected on a visit to Isla Guadalupe between 
31 May and 3 June 1996. Maps of the island, showing localities mentioned 
here, have been provided by Jehl and Everett (1985) and Moran (1996). 
Three of the species we recorded represent first records for the island; in 
addition, we confirmed the breeding of two recently colonizing species and 
noted the recolonization of a previously extirpated species whose taxonomic 
status we reexamined. 

ITINERARY 

We spent a few hours at the naval base (Melpomene Cove or Punta Sur) 
upon our arrival on the island and then traveled the dirt road from there to 
the airstrip (Campo Pista: 29° 01' N, 118° 17' W), where we spent one night 
and a morning observing and collecting. We then continued to the cypress 
grove (29° 07' N, 118° 20' W, 1165m), near the highest point on the island, 
where we spent two nights. One night was spent at the fishing village 
(Campo Tepeyac or Campo Oeste). We returned to the naval base by 
launch. 


Western Birds 32:71-80, 2001 


71 


NOTES FROM ISLA GUADALUPE 


CONSERVATION 

Unfortunately the decimation of the island’s vegetation by feral goats is 
continuing; we saw large numbers of goats throughout the island. They 
appeared to be most numerous in the central and northern portions of the 
island, nearer the major source of fresh water. Moran (1996) reviewed the 
history of goat damage to the flora of the island and estimated that the 
youngest of the endemic Guadalupe Cypress ( Cupressus guadalupensis ) 
were well over 100 years old. Apparently, in excess of 20,000 goats were 
removed from the island in 1971, reportedly leaving only 239 animals at 
that time (Agraz Garcia 1978). Moran (1996) estimated that the population 
had rebounded to at least 7000 in 1994. Low-intensity hunting by fisher- 
men and marines has had little impact on the population in recent years. In 
addition, attempts at protecting the remaining cypress forest fragment with 
a wire fence have failed. The fence is breached in many places, and a gate 
is missing; large herds of goats were observed trooping through the 
cypresses at all hours, leaving virtually no vegetation of consequence within 
the understory of the grove {Figures 1 and 2). In addition, the bark of many 
of the trees showed evidence of damage by goats. Clearly only the complete 
removal of goats from the island or a secure fence with a full-time warden will 
guarantee long-term survival of the cypresses and the birds dependent on 
them. We also saw feral dogs and cats on the island, but only in small 
numbers. 

ANNOTATED LIST 

Laysan Albatross ( Phoebastria immutabilis). First recorded breeding on 
Isla Guadalupe in 1986 (Dunlap 1988). Its status up to 1992 was reviewed 
by Gallo and Figueroa (1996). We briefly surveyed the rocky plateau above 
the naval base and observed six full-grown chicks and one adult bird. The 
young birds were close to fledging with down remaining only on the head 
and neck. 

Brandt’s Cormorant { Phalacrocorax pencillatus). A pair was seen flying 
past Islote Zapato at the south anchorage. 

American Kestrel ( Falco sparverius). Fairly common and observed 
throughout the island. 

Western Gull ( Larus occiderttalis). A few were seen along the coast 
between Campo Tepeyac and Melpomene Cove. 

Mourning Dove ( Zenaida macroura). Common throughout the island. 
Many were observed coming to drink at the fresh water spring below the 
north end of the cypress grove. 

Barn Owl ( Tyto alba). A single flank feather (Coleccion Nacional de Aves 
25550) found near the airstrip provides the first documentation of the 
occurrence of this species on the island, although we saw or heard none. 
There have been several anecdotal reports of large owls on the island, and 
previous authors assumed that these refer to the Great Horned Owl ( Bubo 
uirginianus). 

During Edward Palmer’s initial study of the avifauna of the island (Ridgway 
1876), two unidentified species of owls were observed but not collected. 


72 


NOTES FROM ISLA GUADALUPE 


Accounts of B. virginianus on the island appear to originate with Palmer’s 
assistant (Bryant 1887) who reported a “Horned Owl {Bubo)” and that “the 
Mexicans” reported a “large owl (“ Tecolote ”).” It is possible this represents 
a misunderstanding on behalf of the assistant: tecolote in Mexican Spanish 
could refer to any small owl, bubo (interpreted as Bubo?) to any large owl. 
Presumably the tecolote is the Burrowing Owl, and Bryant (1887) assumed 
that the large owl was a Great Horned. However, it now appears plausible 
that these early reports refer to the Barn Owl, a species that is known for its 
ability to colonize islands, including Isla Socorro some 500 km from the 
southern tip of Baja California. No Great Horned Owls were heard during 
our visit. 

Burrowing Owl ( Athene cunicularia). Fairly common; several pairs were 
seen between the south end and the cypress grove. 

Anna’s Hummingbird ( Calypte anna). Common in the cypress grove. 

Northern Flicker ( Colaptes auratus [cafer group]). Fairly common in the 
cypress grove; we found several active nests and collected two birds. A male 
(CNAV 24795) had enlarged testes and a female (CNAV 24794) had yolking 
eggs. The endemic form, C. a. rufipileus, has not been collected since 1906 
and has been considered extinct. We believe the birds we saw and collected 
were mainland Red-shafted Flickers and that Isla Guadalupe has been 
successfully recolonized by flickers following the extirpation of the endemic 
subspecies (see below). 

Rock Wren ( Salpinctes obsoletus). Abundant from the beaches up to the 
cypress forest. 

Swainson’s Thrush ( Catharus ustulatus ). We collected a male with 
moderately enlarged testes (CNAV 24746) in the cypress grove; no other 
individuals were seen or heard. This is the first record of this species on Isla 
Guadalupe. Presumably it represents a vagrant; Swainson’s Thrush occurs 
throughout Baja California as a migrant (Howell and Webb 1995) and has 
been recorded and presumed to be migrating in the mountains of Baja 
California as late as early June (Wilbur 1987). Escalante-Pliego compared 
the specimen to series at CNAV and tentatively identified it as nominate C. 
u. ustulatus, common in migration on the nearby mainland. 

Northern Mockingbird ( Mimus polyglottos). We saw two together in an 
open area near the cypress grove but could not determine if they were a pair. 
Bryant (1887) observed two birds in 1886, and subsequently individuals 
were seen near the airstrip in January 1988 (Howell and Webb 1992) and 
November 1989 (Mellink and Palacios 1990). 

European Starling ( Sturnus vulgaris). This species is now common on Isla 
Guadalupe. We saw a flock of over 100 birds near the spring below the 
cypress grove. Previous authors noted a few starlings on the island but 
suspected that they might represent only winter visitors (Howell and Webb 
1992). However, we confirmed nesting in tree cavities in the cypress grove 
and collected one juvenile (CNAV 24826). 

Guadalupe Junco ( Junco insularis). This distinctive, endemic taxon has 
been recognized as a species in recent treatments (Howell and Webb 1995). 
We found it common within the cypress grove but did not see it anywhere 
else on the island. At the time of our visit, fledglings in streaky juvenal 
plumage were abundant and some males were still singing. Although other 


73 


NOTES FROM ISLA GUADALUPE 


authors have found juncos down to sea level outside of the nesting season, 
we saw them only in the remnant stand of cypress. 

Western Meadowlark ( Sturnella neglecta). Jehl and Everett (1985) re- 
ported only a record from 1886, but Howell and Webb (1992) saw 35-40 
near the airstrip in January 1988 and suggested that the species might 
breed. Mellink and Palacios (1990) also observed it in November 1989. We 
found this species to be common throughout the island, particularly in the 
extensive grasslands of the central plateau, with flocks of up to 50 birds. 
Males were singing from prominent perches. We collected two birds, a male 
with enlarged testes (CNAV 24798) and a female with a regressing brood 
patch (CNAV 24836). Evidently this species has successfully colonized the 
island. 

Hooded Oriole ( Icterus cucullatus). We saw and collected a single male, 
which had slightly enlarged testes, in a solitary eucalyptus tree at Campo 
Pista (CNAV 24756). This is the first island record and presumably a 
vagrant; however, we were unable to reach the palm grove on the north 
slope of the island (Moran 1996), which perhaps could harbor a small 
breeding population. 

House Finch ( Carpodacus mexicanus). Common from sea level to the 
cypress grove. Recently fledged young were abundant during our visit and 
males were singing. 

SYSTEMATIC STATUS OF THE FLICKERS OF ISLA GUADALUPE 

Ridgway (1876) originally described the subspecies of the Red-shafted 
Flicker from Isla Guadalupe, emphasizing the “bright tawny forehead” — 
hence the name rufipileus — as the diagnostic character. He suggested rump 
color, bill length, wing length, tail length, and the amount of black on the 
underside of the tail as additional characters. Bryant (1886) treated the 
flickers of Isla Guadalupe as a full species. 

Here we reevaluate the taxonomic status and characteristics of the 
Guadalupe Flicker in order to determine the provenance of the flickers we 
observed on the island. A series of C. a. rufipileus is available in the 
ornithological collections of the American Museum of Natural History; these 
birds were collected in the months of May, July, and August. For compara- 
tive material of mainland flickers that might represent the source for a 
possible recolonizing population, we borrowed specimens from a number of 
museums (see Acknowledgments). To ensure that we were working with 
individuals from breeding populations, and not migrants, we used adult 
specimens of both sexes from late spring and summer from mainland 
breeding localities. Few specimens were available from any one locality for 
the breeding season; therefore, our mainland sample was composed of Red- 
shafted Flickers from the mountains of northern Baja California, the 
southern California ranges from San Diego north to the San Bernardinos, 
the coastal ranges north to Monterey, and a few birds from the Sierra 
Nevada. These samples represent C. a. collaris, a form with a range 
proximal to and a phenotype similar to the recent Guadalupe birds (a 
comparison of extremes of C. a. collaris and the paler C. a. canescens, 


74 


NOTES FROM ISLA GUADALUPE 


courtesy P. Unitt, indicates that the two recent island birds are not typical of 
the highly migratory canescens ). In all, our final sample consisted of one 
recent male and one recent female flicker from Isla Guadalupe, which we 
treated as of unknown origin, 1 1 males and 13 females from Isla Guadalupe 
from the late 19th and early 20th centuries, representing C. a. rufipileus, 
and 1 3 male and 1 5 female specimens from Baja California and southern 
California, representing C. a. col laris. 

Sweet and Barrowclough independently measured, for all specimens, 
culmen length (from base of skull), wing chord (bend of wing to longest 
primary), tail length (to tip of longest central rectrix), and width of the 
terminal black band on the ventral surface of one of the central reetrices; 
some measurements could not be taken because of molt or broken feathers 
or bill. In addition, we independently arranged the specimens into graded 
series of crown/forehead color, by sex, from most to least rufescent; we then 
assigned the value of one to the most reddish bird, the value two to the next 
bird in the series, etc. (R. W. Dickerman, pers. comm, suggested that foxing 
was not a significant problem in flickers). In this way, we obtained two 
estimates for each of four mensural characters and a color character for the 
Guadalupe, mainland, and two new flickers. We averaged the two values, 
reducing measurement sampling error. 

We used the computer package SAS (Windows version 6.12: procedure 
univariate; SAS Inst. 1988) to produce box plots for the crown color 
estimates. Because the color scores we assigned the birds are on an arbitrary 
scale, which is undoubtedly nonlinear, we used a nonparametric, descriptive 
technique for analyzing those data. Box plots simply encode frequency 
information and require no assumptions about the underlying distribution: 
the heavy vertical lines indicate the range, the solid lower and upper edges 
of the box correspond to the 25th and 75th percentiles of the distribution, 
the dashed line is the 50th percentile (median), and the cross is the mean of 
the distribution (Figure 3). Crown color, the principal character originally 
used to diagnose the Guadalupe Flicker, clearly varies substantially between 
the mainland and the island for both sexes. The two new specimens, a male 
and a female, both have scores unequivocally placing them with the 
mainland series. 

Ridgway (1876) also mentioned a pink wash on the rump as a character 
for the island race, but we were unable to find a consistent difference in rump 
color and did not investigate that trait further. In the four measurements, the 
island and mainland forms completely overlapped in culmen and tail length. 
In wing length and width of the black tail bar, the means of the two samples 
for males and females differed, though with substantial overlap in the 
distributions (box plots not shown here). However, neither measurement had 
as great a power for discrimination as the crown color. Because these four 
measurements showed varying patterns of correlation (Table 1), we used a 
principal-component (PC) analysis to determine whether there was a linear 
combination of measurements that, independent of crown color, yielded a 
segregation of the island birds. We adopted a PC rather than a discriminant- 
function analysis (DFA) because our sample sizes were small and PC analysis 
identifies the “natural” combinations of the measurements representing 


75 


NOTES FROM ISLA GUADALUPE 



Figure 1. View northwest from remaining cypress grove. Young trees do not exist and 
older trees (such as the one in foreground) have had pieces of bark stripped by goats. 

Photo by Paul R. Sweet 



Figure 2. View into remnants of cypress grove, showing lack of understory. 

Photo by Paul R. Sweet 


76 


NOTES FROM ISLA GUADALUPE 



Guadalupe Mainland Guadalupe Mainland 


Figure 3. Box plots for crown color scores for male and female specimens of Red- 
shafted Flickers from Isla Guadalupe (historical sample), the southwestern mainland, 
and two recent Isla Guadalupe specimens, (triangles). 


most of the variation in a set of data; in DFA, desired groupings are made on 
an a priori basis and high discrimination among groups that represents a 
very small fraction of the actual total variation is sometimes found. Using the 
matrix of character correlations, we found PC axes, separately for the two 
sexes, that summarized most of the variation in measurements with two 
vectors (SAS Inst. 1988: procedure princomp). For males, the first two PC 
axes explain 45% and 27% of the total variation, respectively; for females, 
the corresponding values are 41% and 38%. The character loadings on the 
first two PC axes are available from the authors. In Figure 4 we have plotted 


Table 1 Matrix of Correlation Coefficients between 
Mensural Measurements of Isla Guadalupe and Main- 
land Red-shafted Flickers 0 



Culmen 

Wing 

Tail 

Tail Tip 

Culmen 



-0.076 

-0.031 

0.292 

Wing 

-0.182 

— 

0.091 

-0.640 

Tail 

0.276 

0.515 

— 

0.372 

Tail Tip 

0.409 

-0.321 

0.198 

— 


a Males above and females below diagonal. 


77 


NOTES FROM ISLA GUADALUPE 



PCI 


Figure 4. Scores of Isla Guadalupe (solid circles) and mainland (open circles) Red- 
shafted Flickers on first two PC axes based on measurements only. The recent female 
from Isla Guadalupe is indicated by a triangle; the recent male specimen was 
incomplete and so cannot be included. 


the PC scores for the Isla Guadalupe, mainland, and recent flickers for which 
a full set of measurements were available. Unfortunately, some birds, 
including one of the recent specimens from Isla Guadalupe, had missing 
values and could not be included in the PC analysis. 

For both sexes, the first PC axis yielded a good, but not perfect discrimi- 
nation between the island and mainland specimens. In both cases the island 
birds have greater average PCI scores than do the mainland specimens. The 
pattern of character loadings indicates that for both sexes the PCI axis is 
largely a contrast of wing length vs. tail tip. For the males, sample location 
appears to be independent of the second PC axis; for females, island 
specimens have somewhat lower average scores on PC2 than do those from 
the mainland. The pattern of character loadings indicates that PC2 is at least 
in part a wing length plus tail length axis. The recent island female is clearly 
associated with the mainland specimens. 

Our analysis of crown color plus four measurements indicates that the two 
recent flickers collected on Isla Guadalupe have characteristics placing them 
with mainland flickers rather than with the historical population. Green way 
(1967) summarized the historical status of the Guadalupe Flicker; it was last 
collected in June 1906. None were recorded after that in spite of a number 
of expeditions and searches. However, Jehl and Everett (1985) reported that 
in late 1972 flickers were once again seen on the island; such observations 
have continued (e.g., Howell and Webb 1992), but the subspecific status of 
the recently observed birds was in doubt. The long historical gap in 
observations of flickers between 1906 and 1972, the continuing presence 
of flickers since then, and our character analysis, taken together, suggest that 
Isla Guadalupe was recolonized by mainland Red-shafted Hickers in the late 
1960s or early 1970s and that the endemic subspecies is extinct. 


78 


NOTES FROM ISLA GUADALUPE 


Like most birds, the Guadalupe Flicker was described before the invention 
of statistical tests and quantitative studies of population variability. In part our 
detailed analysis of these flickers was based on an interest in ascertaining 
whether the subspecies C. c. rufipileus actually designates a distinctive 
endemic population (e.g., Barrowclough 1982). One modern, statistical 
standard for the recognition of subspecies is the “75% rule” (e.g., Mayr 
1969). One interpretation of this rule is that subspecific recognition is 
warranted if 90% of one subspecies can be distinguished from 90% of the 
second (this is the 75% rule for symmetrical distributions). For crown color, 
the box plots indicate that females meet this standard, but males do not. For 
the PC analysis, based on measurements alone, the results are even less 
conclusive; no single PC axis can separate the island and mainland birds in 
accord with the 75% rule for either sex. For the females, a linear combina- 
tion of PCI and PC2 can do so, but the pattern of Figure 4 is such that it 
appears this would not hold up with an increased sample size. With the 
specimens currently available — that is, with few specimens from Isla 
Guadalupe in fresh (recently molted) plumage — the now extinct Guadalupe 
Flicker does not approach the 100% diagnosably different status required 
for recognition of a phylogenetic species (e.g., McKitrick and Zink 1988). 
However, females meet the 75% standard for subspecies. We therefore 
believe that subspecies status is (was?) warranted for this island form. It is 
clear that the flickers of Isla Guadalupe differed on average from those of the 
mainland; however, we did not find the degree of differentiation to be 
extensive. There may have been sufficient continuing gene flow to prevent 
complete isolation. An alternative possibility is that a series of fresh August 
or September specimens would have been largely diagnosable. Molecular 
methods may someday have the power to address such questions. 

ACKNOWLEDGMENTS 

Our trip to Isla Guadalupe would not have been possible without the assistance of 
the Armada de Mexico, which provided transport to and from Isla Guadalupe; we 
particularly thank Commandante Francisco Perez-Rico and the officers and crew of 
C-81 Zarco for their hospitality while we were at sea. The fishermen of Cooperativa 
de Langosteros y Abuloneros de Ensenada provided much needed logistic support on 
the island and taught us the noble art of “chivo” hunting. We thank the curatorial staffs 
of the U.S. National Museum of Natural History (Philip Angle), the Los Angeles 
County Museum of Natural History (Kimball Garrett), the California Academy of 
Sciences (Karen Cebra), and the Museum of Vertebrate Zoology (Carla Cicero) for 
providing loans of flickers under their care. We thank the Secretaria del Medio 
Ambiente, Recursos Naturales y Pesca, for providing permits. We are grateful to 
Jeffrey Groth for preparing two figures and to Philip Unitt, William Everett, and an 
anonymous reviewer for helpful comments on an earlier version of this paper. This 
research was supported in part by the Leonard J. Sanford Trust. 

LITERATURE CITED 

Agraz Garcia, A. A. 1978. La cabra cimarrona ( Capra hircus ) en la Isla de Guadalupe, 

B. C. Ganadero 3:35-49. 

Barrowclough, G. F. 1982. Geographic variation, predictiveness, and subspecies. 

Auk 99:601-603. 


79 


NOTES FROM ISLA GUADALUPE 


Bryant, W. E. 1887. Additions to the ornithology of Guadalupe Island. Bull. Calif. 
Acad. Sci. 2:269-318. 

Dunlap, E. 1988 Laysan Albatross nesting on Guadalupe Island, Mexico. Am. Birds 
42:180-181. 

Gallo-Reynoso, J.-P., and Figueroa-Carranza, A.-L. 1996. The breeding colony of 
Laysan Albatrosses on Isla de Guadalupe, Mexico. W. Birds 27:70-76. 

Greenway, J. C., Jr. 1967. Extinct and Vanishing Birds of the World, 2nd ed. Dover 
Publ., New York. 

Howell, S. N. G., and Webb, S. 1992. Observations of birds from Isla Guadalupe, 
Mexico. Euphonia 1:1-6. 

Howell, S. N. G., and Webb, S. 1995. A Guide to the Birds of Mexico and Northern 
Central America. Oxford Univ. Press, Oxford, England. 

Howell, T. R,, and Cade, T. J. 1954. The birds of Guadalupe Island in 1953. Condor 
56:283-294. 

Jehl, J. R., and Everett, W. T. 1985. History and status of the avifauna of Isla 
Guadalupe, Mexico. Trans. San Diego Soc. Nat. Hist. 20:313-336. 

Mayr, E. 1969. Principles of Systematic Zoology. McGraw-Hill, New York. 

McKitrick, M. C., and Zink, R. M. 1988. Species concepts in ornithology. Condor 
90:1-14. 

Mellink, E., and Palacios, E. 1990. Observations on Isla Guadalupe in November 
1989. W. Birds 21:177-180. 

Moran, R 1996. The Flora of Guadalupe Island, Mexico. Memoirs Calif. Acad. Sci. 19. 

Oberbauer, T. A., Cibit, C., and Lichtwardt, E. 1989. Notes from Isla Guadalupe. W. 
Birds 20:89-90. 

Ridgway, R. 1876. Ornithology of Guadalupe Island, based on notes and collections 
made by Dr. Edward Palmer. Bull. U.S. Geol. Geogr. Surv. Terr. 2:183-195. 

SAS Institute Inc., 1988. SAS Language Guide for Personal Computers, release 6.03 
ed. SAS Inst., Cary, NC. 

Wilbur, S. R. 1987. Birds of Baja California. Univ. Calif. Press, Berkeley. 

Accepted 30 October 2000 


80 


NOTES 

ORANGE BISHOPS BREEDING IN PHOENIX, ARIZONA 


THOMAS A. GATZ, U. S. Fish and Wildlife Service, 2321 W. Royal Palm Road, 
Suite 103, Phoenix, Arizona 85021 

In late summer of 1998 and 1999, 1 observed a small colony of approximately ten 
Orange Bishops (Euplectes franciscanus ) in north Phoenix, Arizona. The colony was 
in a grassy tree-lined partially channelized desert wash in a highly urbanized area south 
of the intersection of First Drive and Greenway Parkway. A small amount of standing 
water was present through most of the year. In June of 2000, the 1998-1999 site was 
still occupied and I located a second group of six birds approximately 1 . 1 km east of the 
first location in a small cattail marsh with surface water in the same wash southwest of 
the intersection of Seventh Street and Greenway Parkway. The cattail marsh formed 
as a result of water ponded below a street culvert. The bishops at both locations were 
feeding on seeds of Johnson Grass ( Sorghum hatepense), a nonnative species from 
the Mediterranean region. Males, and occasionally females, often perched in nearby 
native and nonnative trees. I visited these specific locations and the surrounding area 
frequently from 1994 to 1997 but observed no bishops prior to 1998- 

Evidence of breeding included alternate-plumaged males (Figure 1) displaying and 
chasing females from June through September, a male carrying a long blade of grass, 
apparently for nesting material, on 4 July, and a recently constructed but unoccupied 
nest in the cattail marsh on 6 August (Figure 2). The roundish nest, made of dried 



Figure 1. Adult male Orange Bishop in alternate plumage, Phoenix, Arizona, July 
2000. 


Photo by T. Gatz 


Western Birds 32 81-82, 2001 


81 


NOTES 



Figure 2. Nest of Orange Bishop, Phoenix, Arizona, August 2000. 

Photo by T. Gatz 


grass, was 10 cm tall by 7.5 cm wide with an opening near the top. It was woven 
around several stems of Johnson Grass, 1 m above the water surface. In September, 
I observed a female feeding a fledged young. 

This species, a member of the mainly African weaver family, the Ploceidae, is 
native to sub-Saharan Africa and is now well established in southern California with 
flocks of 50 to 100 birds routinely noted in some flood-control basins near Los 
Angeles [Garrett, K. L. 1998. Field separation of bishops { Euplectes ) from North 
American emberizids. W. Birds 29:231-232]. Both the California and Arizona birds 
are assumed to have originated from escaped cage birds. Small colonies of this species 
are now established in the West Indies and are also believed to have originated from 
captive birds (Raffaele, H., J. Wiley, O. Garrido, A. Keith, and J. Raffaele. 1998. A 
Guide to the Birds of the W 7 est Indies, p. 446. Princeton Univ. Press, Princeton, N.J.). 

Although no additional cattail marsh/Johnson Grass habitat occurs in the area 
immediately surrounding these sightings, a considerable amount of similar habitat, 
into which this species could spread, exists in riparian areas, around artificial 
impoundments for flood control, and in agricultural sections of central and southern 
Arizona. 

I thank Kimball L. Garrett and Scott Smithson for sharing useful information on 
Orange Bishops with me. Garrett and Kathy Molina reviewed an earlier draft of the 
manuscript. 

Accepted 14 October 2000 


82 


NOTES 


BREEDING-SEASON HOME RANGES 

OF SPOTTED OWLS IN THE 

SAN BERNARDINO MOUNTAINS, CALIFORNIA 

GUTHRIE S, ZIMMERMAN and WILLIAM S. LaHAYE, Department of Fisheries and 
Wildlife, University of Minnesota, 174 McNeal Hall, 1985 Buford Ave., St. Paul, 
Minnesota 55108 

R. J. GUTIERREZ, Department of Wildlife, Humboldt State University, Areata, 
California 95521 and Department of Fisheries and Wildlife, University of Minnesota, 
St. Paul, Minnesota 55108 


Home ranges of the Northern Spotted Owl ( Strix occidentalis caurina) in the 
Pacific Northwest (Forsman et al. 1984, Solis and Gutierrez 1990, Carey et al. 1992, 
Zabel et al. 1995), of the Mexican Spotted Owl [S. o. lucida) in the southwestern U.S. 
(Ganey and Baida 1989, Zwank et al. 1994, Ganey et al. 1999), and of the California 
Spotted Owl (S. o, occidentalis ) in the Sierra Nevada (Call et al. 1992, Zabel et al. 
1992) have all been quantified. No home-range estimates exist, however, for isolated 
populations of the California Spotted Owl in the southern portion of its range 
(Gutierrez and Pritchard 1992, Gutierrez et al. 1995). Therefore, we report breeding- 
season home-range size for two pairs of radio-marked Spotted Owls in the San 
Bernardino Mountains, which support the largest population of the subspecies in 
southern California (LaHaye et al. 1997). 

Our study area is approximately 140 km east of Los Angeles, California (34° 15' 
N, 117° 55' E). The San Bernardino Mountains are oriented east/west with elevations 
ranging from 800 to 3500 rn and are surrounded by desert and chaparral vegetation 
(Barbour and Major 1988). The climate is Mediterranean, with most precipitation 
falling during the winter in the form of snow above 2000 m and rain at lower 
elevations. Precipitation, influenced by elevation and slope aspect, ranges from 25 to 
100 cm (Minnich et al. 1995). Vegetation grades from Mojave desert scrub and 
coastal scrub at lower elevations to alpine at higher elevations (Barbour and Major 
1988). Within this continuum, local aspect and topography form a complex mosaic of 
forest, chaparral, desert, and wetland vegetation. In this range, Spotted Owls occur 
between 800 and 2600 m and occupy forests composed of Canyon Live Oak 
( Quercus chrysolepis), Black Oak (Q. kelloggii), Big-cone Douglas Fir ( Pseudotsuga 
macrocarpa ), White Fir (Abies concolor), Incense Cedar ( Calocedrus decurrens), 
Jeffrey Pine ( Pinus jeffregi), Ponderosa Pine (P. ponderosa), and Sugar Pine ( P. 
larnbertiana). 

Using radio telemetry, we monitored two pairs of Spotted Owls from July 1987 to 
August 1988. We captured the owls with noose poles or mist nets. Each owl was fitted 
with a radio transmitter (Telonics Inc., Mesa, Arizona) by means of a backpack harness 
(Guetterman et al. 1991). The total mass of the transmitter package was approxi- 
mately 18 g. We received transmitter signals through TR-2 receivers and four-element 
hand-held Yagi antennas (Telonics Inc.). Otis and White (1999) demonstrated that 
autocorrelation of telemetry locations does not bias estimates of home ranges when 
animals are considered the sampling unit. Nonetheless, we separated all owl locations 
in this study by at least 24 hours. We followed techniques outlined by Guetterman et 
al. (1991) and estimated nocturnal locations from triangulation of three to six 
compass bearings taken from fixed telemetry points. Triangulations resulting in 
polygons larger than 2 ha were removed from analyses. 

We defined a breeding-season home range as the area used by owls during their 
nightly activities (Burt 1943) between March and August. We used the program 
CALHOME (Kie et al. 1996) to estimate home ranges with a minimum convex 
polygon (MCP) and the program KERNELHR (Seaman et al. 1998) to produce fixed- 


Western Birds 32:83-87, 2001 


83 


NOTES 


and adaptive-kernel (Worton 1989) estimates. We used kernel estimators because 
they require no unrealistic assumptions about space use (Worton 1989) and perform 
better than other estimators in simulations (Worton 1995). The fixed-kernel estimator 
with least-squares cross validation outperformed the adaptive-kernel estimator in 
simulations (Seaman and Powell 1996, Seaman et al. 1999), so we focused our 
results and discussion on home ranges estimated with this procedure. The adaptive 
kernel (Ganey et al. 1999) and MCP have been used commonly as home-range 
estimators in studies of Spotted Owls (e.g., Forsman et al. 1984, Call et al. 1992, 
Zabel et al. 1995). Thus, given similar sample sizes, the adaptive, kernel and MCP 
provided estimates comparable to those of other studies. We calculated the 95% fixed 
kernel, 95% adaptive kernel, 95% MCP, and 100% MCP for individual owls. We 
combined locations from individuals of each pair to estimate the pairs’ home ranges 
(Table 1). We used the field techniques outlined by Franklin et al. (1996) to assess 
reproductive activity of radio-marked owls. 

The number of locations per owl varied from 51 to 65 (Table 1). Fixed-kernel 
estimates of individual owls’ home ranges varied from 223 to 654 ha. Considered 
individually or as a pair, the Pine Knot owls had a larger range than the Fawnskin owls. 
Fixed-kernel estimates for the pairs indicated that the Pine Knot pair’s home range 
was almost twice as large as the Fawnskin pair’s (Table 2). The Fawnskin pair nested 
during both years of the study, the Pine Knot pair during only the first year, but 60% 
of the latter’s telemetry locations were obtained the second year when they did not 
nest. Furthermore, locations for the Pine Knot pair during the first year were 
concentrated near the center of the home range and nest; the second year the female 
used a larger area. 

Kernel home ranges are estimates based on probability-density functions and are 
subject to sampling error (Seaman et al. 1999), which decreases as sample size 
increases. Seaman et al. (1999) reported that the bias and variance of kernel 
estimators reached an asymptote at >50 locations and recommended that home- 
range estimates be based on 50 or more locations. Alt home-range estimates in our 
analysis were based on >50 locations. However, our fixed- and adaptive-kernel 
estimates of home-range size for each pair may be slightly underestimated. By 
combining locations from members of a pair, we assumed that nocturnal locations of 
pairs are independent. If foraging locations for individuals of mated pairs are 
dependent, the sample size will be smaller than the sum of locations of members of the 
pair because the pair is acting as a single unit (Burnham and Anderson 1998:52). 


Table 1 Home-range Estimates (ha) Based on Nocturnal Radiote- 
lemetry Locations for Individual Spotted Owls during the Breeding 
Season in the San Bernardino Mountains, 1987-1988 





Estimation Method 0 


Territory 

Locations 

100% MCP 

95% MCP 

95% FK 

95% AK 

Fawnskin 

Female 

51 

153 

122 

223 

262 

Male 

65 

325 

229 

430 

568 

Pine Knot 

Female 

65 

660 

495 

654 

831 

Male 

63 

648 

377 

448 

619 


a MCP, minimum convex polygon; FK, fixed kernel; AK, adaptive kernel. 


84 


NOTES 


Table 2 Home-range Estimates (ha) Based on Nocturnal Radiote- 
lemetry Locations for Pairs of Spotted Owls during the Breeding 
Season in the San Bernardino Mountains, 1987-1988 


Estimation Method 0 


Territory 

Locations 

100% MCP 

95% MCP 

95% FK 

95% AK 

Fawnskin 

116 

325 

210 

333 

415 

Pine Knot 

128 

816 

632 

598 

810 


°MCP, minimum convex polygon; FK, fixed kernel; AK, adaptive kernel. 


However, Forsman et al. (1984) reported that paired Spotted Owls they studied 
foraged at the same locations only 4% to 10% of the time, indicating largely 
independent foraging behavior in this species. In contrast, if the pair’s foraging 
locations in our study were not independent, we believe that the bias in our estimates 
of home-range size is small because we have >50 independent locations for each pair. 

Pairs of Spotted Owls are more strongly associated with a central place (nest or 
roost area) during the breeding season than during the nonbreeding season (Forsman 
et al. 1984). Consequently, home ranges during the breeding season are smaller (e.g. , 
Forsman et al. 1984, Zabel et al 1992, Gutierrez et al. 1995). 

Our 100% MCP estimates of breeding-season home ranges for individual Califor- 
nia Spotted Owls are larger than estimates reported for the Mexican Spotted Owl 
(range 278-361 ha; Zwank et al. 1994. Willey and van Riper 1995, respectively) but 
smaller than most estimates for the Northern Spotted Owl (range 413-817 ha; Solis 
and Gutierrez 1990, Zabel et al. 1995, respectively). In general, our estimates are 
smaller than most previous estimates for California Spotted Owls. Reported home 
ranges for California Spotted Owls range from 289 ha in the southern Sierra Nevada 
(Zabel et al. 1992) to 2195 ha in the northern Sierra Nevada (Zabel et al. 1992). 
Zwank et al. (1994) reported home ranges about 30% smaller than ours for an isolated 
population of Mexican Spotted Owls in New Mexico. Our estimates of pairs’ home 
ranges differ from those of other studies in a pattern similar to that described for 
individuals. Differences in home-range size of Spotted Owls among study areas may 
be due to variations in habitat (Zabel et al. 1992), prey base (Zabel et al. 1995), 
foraging behavior, or weather. 

The variation in home-range size within our study is consistent with other studies of 
Spotted Owls (Forsman et al. 1984, Ganey and Baida 1989, Call et al. 1992, Zabel 
et al. 1995). Hypotheses for differences in size of home ranges include differences in 
habitat (Forsman et al. 1984, Carey et al. 1992), prey type and abundance ( Zabel et 
al. 1995), prey-renewal rates (Carey et a!.1992), and habitat fragmentation (Carey et 
al. 1992). Home-range size may also be related to nesting status, with nesting owls 
having smaller home ranges because they are strongly associated with the nest. 

Densities of the Dusky-footed Woodrat (Neotoma fuscipes), the primary prey of 
Spotted Owls in our study area (Smith et al. 1999), can vary greatly (Williams et al. 
1992). The habitat in both territories we studied is similar, and both pairs nested 
during the study. Therefore, we suspect that the differences in home-range size 
between the two pairs were due to differences in prey availability or random variation 
within the population. We could not investigate whether prey, habitat, nesting status, 
or the combination of these factors correlated with home-range size because our 
sample size is small. However, this study is the first to document home ranges of 
California Spotted Owls from isolated populations in southern California. 


85 


NOTES 


This study was funded by Snow Summit Sid Corporation and Bear Mountain Ltd. 
Additional support was provided by the University of Minnesota. We thank JoAnn 
Gronski, Mike Hollister, Richard Tanner, Richard Smith and John Stephensen for 
helping to monitor radio-marked Spotted Owls. J. L. Ganey, R. LeValley, and an 
anonymous reviewer gave helpful suggestions that improved the quality of the 
manuscript. 

LITERATURE CITED 

Barbour, M. G., and Major, J. 1988. Terrestrial Vegetation of California. Calif. Native 
Plant Soc., Sacramento. 

Burnham, K. P., and Anderson, D. R. 1998. Model Selection and Inference. A 
Practical Information-theoretic Approach. Springer-Verlag, New York. 

Burt, W. H. 1943. Territoriality and home range concepts as applied to mammals. J. 
Mammal. 24:346-352. 

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: 
Influence 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. 

Franklin, A. B,, Anderson, D. R., Forsman, E. D., Burnham, K. R, and Wagner, F. W. 
1996. Methods for collecting and analyzing demographic data on the Northern 
Spotted Owl. Studies Avian Biol. 17:12-20. 

Ganey, J. L., and Baida, R. R 1989. Home-range characteristics of Spotted Owls in 
northern Arizona. J. Wildlife Mgmt. 53:1159-1165. 

Ganey, J. L., Block, W. M., Jenness, J. S., and Wilson, R. A. 1999. Mexican Spotted 
Owl home range and habitat use in pine-oak forest: Implications for forest 
management. Forest Sci. 45:127-135. 

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 Service, Pacific Northwest Gen. Tech. Rep. 272. 

Gutierrez, R. J., and Pritchard, J. 1992. Distribution, density, and age structure of 
Spotted Owls on two southern California habitat islands. Condor 92:491-495. 

Gutierrez, R. J., Franklin, A. B., and LaHaye, W. S. 1995. Spotted Owl (Strix 
occidentalis), in The Birds of North America (A. Poole and F. Gill, eds.), no. 179. 
Acad. Nat. Sci., Philadelphia. 

Kie, J. G., Baldwin, J. A., and Evans, C. J. 1996. CALITOME: A program for 
estimating animal home ranges. Wildlife Soc. Bull. 24:342-344. 

LaHaye, W. S., Gutierrez, R. J., and Call, D. R. 1997. Nest-site selection and 
reproductive success of California Spotted Owls. Wilson Bull. 109:42-51. 

Minnich, R. A., Barbour, M. G., Burk, J. H., and Fernau, R. F. 1995. Sixty years of 
change in the California conifer forests of the San Bernardino Mountains. Cons. 
Biol. 9:902-914. 

Otis, D. L., and White, G. C. 1999. Autocorrelation of location estimates and the 
analysis of radiotracking data. J. Wildlife Mgmt. 63:1039-1044. 

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. 


86 


NOTES 


Seaman, D. E., Griffith, B., and Powell, R. A. 1998. KERNELHR: A program for 
estimating animal home ranges. Wildlife Soc. Bull. 26:95-100. 

Seaman, D. E , Millspaugh, J. J., Kernohan, B. J., Brundige, G. C., Raedeke, K. J., 
and Gitzen, R. A. 1999, Effects of sample size on kernel home range estimators. 
J. Wildlife Mgmt. 63:739-747. 

Smith, R, B., Peery, M. Z., Gutierrez, R. J., and LaHaye, W. S. 1999. The 
relationship between Spotted Owl diet and reproductive success in the San 
Bernardino Mountains, California. Wilson Bull. 111:22-29. 

Solis, D. M., Jr., and Gutierrez, R. J. 1990. Summer habitat ecology of Northern 
Spotted Owls in northwestern California. Condor 92:739-748. 

Willey, D. W., and van Riper, C., III. 1995. Home range characteristics of Mexican 
Spotted Owls in southern Utah. J. Raptor Res. 29:48-48. 

Williams, K. S., Verner, J., Sakai, H. F,, and Waters, J. R. 1992. General biology of 
the major prey species of the California Spotted Owl, in A Technical Assessment 
for the California Spotted Owl (J. Verner, K. McKelvey, B. R. Noon, R. J. 
Gutierrez, G. 1. Gould, Jr., and T, W. Beck, eds.). U. S. Forest Service, Pacific 
Southwest Res. Sta. PSW-GTR-133. 

Worton, B. J. 1989. Kernel methods for estimating the utilization distribution in 
home-range studies. Ecology 70:164-168. 

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., McKelvery, 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 A Technical Assessment for the California Spotted Owl. (J. 
Verner, K. McKelvey, B. R. Noon, R. J. Gutierrez, G. I. Gould. Jr., and T. W. 
Beck, eds.). U. S. Forest Service, Pacific Southwest Res. Sta. PSW-GTR-133. 

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

Zwank, P. J., Kroel, K. W., Levin, D. M., Southward, G. M., and Romme, R. C. 1994. 
Habitat characteristics of Mexican Spotted Owls in southern New Mexico. J. 
Field Ornithol. 65:324-334. 


Accepted 28 November 2000 


87 


NOTES 


FIRST REPORT OF THE GRAY HERON 
IN THE UNITED STATES 

KENNETH M. BURTON, P. O. Box 716, Inverness, California 94937 

SEAN D. SMITH, Saint Paul Island Tours, 1500 West 33rd Avenue Suite 220, 
Anchorage, Alaska 99503 


On 1 August 1999, at about 1400 ADT, Mike Greenfelder flushed a large Ardea 
heron from the shore of Weather Bureau Lake on St. Paul Island in the Pribilof 
archipelago, Alaska (57° 09' N, 170° 14' W). The bird disappeared into dense fog 
before it could be identified to species. Greenfelder and Smith searched fruitlessly the 
rest of the afternoon, and in the evening they were joined by Burton and several other 
birders. We finally saw the bird in flight near its original location. The heron alighted 
briefly at Rocky Lake, then continued flying west, calling repeatedly. Although the 
light was poor and our views were distant, we tentatively identified the bird as an adult 
Gray Heron, Ardea cinerea, in breeding plumage, on the basis of coloration and 
voice. We followed it by automobile to Southwest Point, where the road becomes 
essentially impassable and the heron disappeared behind bluffs, heading north as if to 
circumnavigate the island. 

The following morning, we found the bird standing in a hunched posture on a rock 
at Webster Lake (11 km from Weather Bureau Lake and 20 km from Southwest 
Point). We observed the now stationary bird for almost an hour through spotting 
scopes from a distance of about 300 m and took detailed notes, on which the 
following description is based. This period of observation confirmed the previous 
evening’s tentative identification. We were too far for photos and decided against 
closer approach until others had seen the bird and a video camera had been obtained. 
The heron was still in the same place when Burton saw it again several hours later, but 
it was wading actively when Smith returned with other observers. Before photos or 
videos could be obtained, the bird flushed inexplicably, and it was not seen again 
despite several days of intensive searching. 

This observation has been accepted on the Alaska unsubstantiated list and repre- 
sents the first report of a Gray Heron in the United States. 

In general appearance the bird resembled a Great Blue Heron ( Ardea herodias), 
a species with which all the observers are quite familiar, but had a less massive bill. The 
features that distinguished it from that species were its whitish, not rusty, thighs and 
complete lack of rufous on the shoulders. In flight, the upperparts showed stronger 
contrast between the black remiges and the gray back and wing coverts than is typical 
of a Great Blue Heron, and the underwings were a fairly uniform bluish gray with a 
pale leading edge. Its call, uttered frequently in flight, was a throaty “kraak,” higher 
pitched than the croak of a Great Blue Heron. 

The distal part of the bill was dull yellowish, becoming darker toward the base, 
especially on the culmen. The bird had a whitish crown, paler than the gray of the 
forehead and lores. Single black supraocular stripes extended posteriorly from the 
eyes, broadening and joining at the nape and forming nuchal plumes. White on the 
lower part of the face blended into gray on the sides of the neck; this was a paler gray 
than that of the mantle and secondary coverts. Several whitish plumes extended from 
the scapulars. At rest, the bird showed a black shoulder patch bordered on the sides 
with white, with no rufous anywhere on the wing. 

Short black streaks extended in two lines down the whitish foreneck. The breast, 
belly, thighs, and undertail coverts appeared creamy white overall; the sides of the 
belly were black. There was a dusky wash, from which a few creamy plumes 
protruded, across the breast and a dusky band across the lower belly. Grayish 


88 


Western Birds 32:88-90, 2001 


NOTES 


streaking extended along the flanks, and the undertail coverts showed sparse dingy 
streaking. The legs were dull yellowish green, paler than is typical of a Great Blue 
Heron. 

Our sighting was preceded by a week of predominantly southwest and west- 
southwest winds averaging 15-21 knots daily. We found no other Asiatic vagrants on 
St. Paul associated with this weather system, but similar conditions produced a 
Chinese Pond-Heron ( Ardeola hocchus) on 4 August 1996 (Hoyer and Smith 1997). 
The Pribilofs’ only other heron records are of single Black-crowned Night-Herons 
(Nycticorax nycticorax ) on 3 April 1979 and 6 August 1986. The early-August 
occurrence of three of the Pribilofs’ four heron records is notable. 

The Gray Heron breeds widely across Eurasia and Africa, withdrawing from most 
northern areas in winter (del Hoyo et al. 1992). In northeast Asia, the nominate race 
extends to Sakhalin Island and the Aldan River in the Russian Far East, roughly 3000 
km west of St. Paul, while A. c. jouyi reaches Korea and Japan (Dement’ev and 
Gladkov 1951, Howard and Moore 1991, del Hoyo et al. 1992). A. c. jouyi tends to 
be paler than A. c. cirierea on the neck and upperwing coverts, with no buff or mauve 
tinge on the neck (Cramp and Simmons 1977, Hancock and Kushlan 1984). 
However, these differences are subtle, relative, and obscured by intergradation and 
clinal variation (Cramp and Simmons 1977, Hancock and Kushlan 1984), making 
subspecific identification of the St. Paul bird impossible. 

Gray Herons disperse widely after the breeding season (Hancock and Kushlan 
1984), which typically extends into the third week of July in northern Europe (Cramp 
and Simmons 1977) and early July in Japan (Brazil 1991). Migratory movement of 
most populations is to the southwest (Moreau 1972, Hancock and Kushlan 1984), 
though some individuals may move in other directions (Ali and Ripley 1968). There 
are many records of vagrants, and in North America the Gray Heron has occurred in 
the lesser Antilles and Bermuda (Hancock and Kushlan 1984, American Ornitholo- 
gists’ Union 1998), although the open sea is a major barrier to dispersal (Cramp and 
Simmons 1977). The Gray Heron has experienced exponential population growth in 
east Asia in recent decades (del Hoyo et al. 1992), increasing the likelihood of strays. 

The closely related and very similar Great Blue Heron is resident in southeastern 
and south-coastal Alaska west to Prince William Sound, approximately 1400 km east- 
northeast of St. Paul (AOU 1998). It wanders north and west after the breeding 
season but so far has not been recorded west of Kodiak Island. 

We thank our employer, Tanadgusix Corporation, for enabling us to spend the 
summer on St. Paul, Mike Greenf elder for his sharp eyes, and Daniel D. Gibson, 
Steven C. Heinl, Peter Pyle, and Theodore G. Tobish for reviewing the manuscript. 

LITERATURE CITED 

Ali, S,, and Ripley, S. D. 1968. Handbook of the Birds of India and Pakistan, vol. 1. 
Oxford Univ, Press, Bombay. 

American Ornithologists’ Union. 1998. Check-list of North American Birds, 7th ed. 
Am. Ornithol. Union, Washington, D.C. 

Brazil, M. A. 1991. The Birds of Japan. Smithsonian Inst. Press, Washington, D.C. 

Cramp, S., and Simmons, K. E L. 1977. Handbook of the Birds of Europe, the 
Middle East and North Africa, vol. 1. Oxford Univ. Press, Oxford, England. 

Del Hoyo, J., Elliott, A., and Sargatal, J. 1992. Handbook of the Birds of the World, 
vol. I. Lynx Edicions, Barcelona. 

Dement’ev, G. P., and Gladkov, N. A., eds. 1951. Birds of the Soviet Union, vol. II. 
Gosudarstvennoe Izdatel’stvo, Moscow. 


89 


NOTES 


Hancock, J., and Kushlan, J. 1984. The Herons Handbook. Harper & Row, New 
York. 

Howard, R., and Moore, A. 1991. A Complete Checklist of the Birds of the World, 
2nd ed. Academic Press, San Diego. 

Hoyer, R. C., and Smith, S. D. 1997. Chinese Pond-Heron in Alaska. Field Notes 
51:953-956. 

Moreau, R. E. 1972. The Palearctic-African Bird Migration Systems. Academic 
Press, London. 


Accepted 5 October 2000 


NOTES 


CLOACAL INSPECTION OR PECKING 
IN ALLEN S HUMMINGBIRD 

JANET L. LEONARD, Joseph M. Long Marine Laboratory, 100 Shaffer Road, 
University of California, Santa Cruz, California 95060 

Cloacal pecking, in which a male pecks at the cloaca of a female, causing her to 
void sperm, was first described in the Dunnock ( Prunella modularis), a species with 
a very complex group-breeding system (Davies 1983). It has been interpreted as a 
mechanism whereby males ensure paternity of a female’s offspring by eliminating the 
sperm of previous mates. Comparable behavior has been observed occasionally in 
other species of birds (N. Davies pers. comm.), but its taxonomic range and ecological 
context are not yet clear. On 14 April 2000 at approximately midday, I observed a 
somewhat similar behavior between two Allen’s Hummingbirds ( Selasphorus sasin ) 
in a suburban backyard in Half Moon Bay, California. It was a bright sunny day, and 
the birds were observed from a window of the house from about 3 meters without 
binoculars. When I first noticed them, an adult male Allen’s Hummingbird and what I 
took to be a female Allen’s were interacting face to face and less than 30 cm apart 
about 3 to 4 meters above the ground. While I watched, the male flew to a position 
below the female and, facing her, and put the tip of his bill in the general area of her 
cloaca. The pair remained in this position while sinking slowly, in hovering flight, 
almost exactly vertically, until the male’s tail was approximately 30 cm above the 
ground, at which point the male and female separated and flew into cover where I lost 
sight of them. The episode was prolonged, taking about 30 seconds, and appeared to 
require coordination between the two birds to allow the slow, sinking, vertical 
hovering flight. Allen’ s Hummingbird is a summer resident in central California, and 
the first adult male of the season that could be definitely identified as Allen’s rather 
than the transient Rufous Hummingbird (S. rufus) was observed in this location on 19 
March in 2000 and on 16 March in 1999. Both Allen’s and Anna’s (Calypte anna) 
Hummingbirds appear to establish breeding territories in or near the yard, although I 
have observed no nests. 

There are several potential explanations for this behavior. The first would be a 
simple cloacal inspection by the male, which might serve to assess the reproductive 
condition and/or recent mating history of the female, serving a function analogous to 
the anogenital sniffing common in mammalian courtship and social behavior. The 
second would be to stimulate the female to void sperm from a previous mating, as in 
the Dunnock. No droplet of semen was observed but this might have been missed 
given the rapid movement of the male and female at the bottom of the descent. A third 
possibility, given the long tongue and characteristic feeding behavior of humming- 
birds, is that the male might actually remove sperm from the reproductive tract of the 
female (Jeffrey R. Baylis pers. comm.). Physical removal of sperm by males was first 
demonstrated in damselflies (Waage 1979) and is known from other insects (Simmons 
and Siva-Jothy 1998). All of these explanations focus on the phenomenon of sperm 
competition, in which males continue to compete to fertilize a female’s eggs even after 
mating (for birds see Birkhead 1998). The observations reported here suggest that 
novel mechanisms of sperm competition remain to be described in birds. 

LITERATURE CITED 

Birkhead, T. R. 1998. Sperm competition in birds: Mechanisms and functions, in 
Sperm Competition and Sexual Selection (T. R. Birkhead and A. P. Moller, eds.), 
pp. 579-622. Academic Press, San Diego. 


Western Birds 32:91-92, 2001 


91 


NOTES 


Davies, N. B. 1983. Polyandry, cloaca-pecking and sperm competition in dunnocks. 
Nature 302:334-336. 

Simmons, L. W., and Siva-Jothy, M. T. 1998. Sperm competition in insects: 
Mechanisms and the potential for selection, in Sperm Competition and Sexual 
Selection (T. R. Birkhead and A. P. Moller, eds.), pp. 341-434. Academic Press, 
San Diego. 

Waage, J. K. 1979. Dual function of the damselfly penis: Sperm removal and transfer. 
Science 203:916-918. 


Accepted 14 October 2000 



Burrowing Owl Sketch © by Marni Fylling 


92 


BOOK REVIEWS 


Birds of North America, by Kenn Kaufman. 2000. Houghton Mifflin, New York. 
384 pages. Paperback, $20. ISBN 0-395-96464-4. 

The year 2000 saw the publication of two completely new field guides to North 
American birds — the Sibley guide and the Kaufman guide. This simultaneity compels 
us to draw parallels — but are we comparing apples with apples? Before the details of 
the Kaufman guide are addressed, some background information should be appreci- 
ated. Kaufman is a well-known figure in North American birding circles. His accom- 
plishments include editorship of the sadly short-lived but pioneering journal Continen- 
tal Birdlife (1979-1981) and authorship of numerous publications including the 
Peterson series Field Guide to Advanced Birding (1990). Such associations with the 
cutting edge of bird identification may have led to certain expectations from his new 
field guide. 

This guide, however, is aimed at entry-level and beginning birdwatchers, that is, 
anybody. This huge demographic is largely excluded (albeit inadvertently) by most 
other popular field guides. Excluded? This depends on how one defines a birdwatcher. 
Kaufman’s intent is to enhance appreciation of birds in the general populace, many of 
whom we, as more advanced practitioners, might not even consider as birdwatchers. 
Conservation depends ultimately on political decisions, and the greater the public’s 
awareness of birds, the greater the chance for their conservation. Ornithologists 
constitute a tiny minority that can help bring about conservation of endangered 
habitats or species. But if everybody appreciates birds, at some level, then the grass- 
roots support for conservation could be monumental. A dream, perhaps, but a worthy 
one. 

Departing from the seminal Peterson field guides, Kaufman chose to use photo- 
graphs for illustrations. Making the leap from a bird one sees to a field-guide painting 
is one experienced birders take for granted, but 1 suspect most people relate more 
easily to photographs. Really good photographs are not always easy to find, so 
Kaufman edited photos using a computer and his extensive field experience. 

The book is compact (it fits into a back pocket) and arranged in conventional field- 
guide format. Short introductory sections in user-friendly prose cover birding basics, 
bird topography, field marks (ten photos of House Finches convey variation more fully 
than in any other guide with which I am familiar), taxonomy, and geographic 
distribution. Then come the species accounts, with photos arranged opposite succinct 
text. Comparable poses were chosen for similar species, when possible, but photos 
cannot work as well as good paintings to facilitate direct comparisons. Most species 
are represented by at least two images that help convey differences in posture (useful 
for beginners) and plumage. Taxonomy and nomenclature, but not sequence, follow 
the American Ornithologists’ Union (through the 2000 supplement to the 1998 
checklist). First come ducks, then other swimming birds, aerial waterbirds, birds of 
prey, chickenlike birds, wading birds, shorebirds, medium-sized landbirds, and finally 
other landbirds (arranged in seven groupings). One can question why flycatchers or 
sparrows are not “typical songbirds,” but the divisions may be helpful to a beginner. 
For example, “everybody” knows what a duck is, so that’s a good place to start, and 
then comes the coot — “non-birders” at a local park 1 visit in my attempts to identify 
hybrid gulls often ask “what’s this black duck?” Give them the National Geographic 
Society (NGS) guide and see them struggle to even find, let alone identify, an 
American Coot. The Kaufman guide should get them to this identification on page 2, 
in the pictorial table of contents. 

The very first page has color-coded tabs that help locate a particular group. An 
important innovation is the pictorial table of contents at the front of the book, as one 
turns the first page. Photos of a selection of species from each group are presented 
right away, so that someone wanting to identify a bird can get (one hopes) to the right 


Western Birds 32:93-96, 2001 


93 


BOOK REVIEWS 


group quickly, without wading through introductory text that might frustrate a “non- 
birder.” The ability to put a name to a bird one sees is a powerful hook, and the 
pictorial table of contents is a good way to facilitate this step in incipient birders. Each 
of these groups (ducks, aerial waterbirds, etc.) is introduced in the text by a pictorial 
section that further narrows the choices and provides background information. 
Additional information is provided for problem groups: the summary of gulls on pages 
68-69 is as helpful and concise as one could wish for a beginner, while the detailed 
range map (p. 273) for the Black-capped and Carolina chickadees is useful for these 
commonly seen garden birds. 

On the basis of a few random comparisons, the color range maps seem on a par 
with the NGS for overall accuracy and further discriminate between common and 
scarce occurrence by different shades. The text for each species opens with a synopsis 
of abundance, habitats, and habits, followed by field marks, then voice. It seems user- 
friendly for the intended audience and could also benefit more experienced birders: 
the Short-tailed Shearwater is “much like Sooty Shearwater, not always identifiable;” 
or Cassin's Vireo is “Like a duller version of Blue-headed Vireo; best separated by 
range. Where they overlap some may not be identifiable.” 

Which species are covered? Although not stated in the introduction, it appears that 
all regularly breeding North American species are included, plus all regular migrants, 
commoner “vagrants,” and even some very rare vagrants that could be found in 
gardens or at feeders (e.g., the Green Violet-ear and Fieldfare). Oddballs from Alaska 
islands and the like are not included. 

What of the illustrations, the edited photos? In general these are well chosen, and 
pointers highlight field marks, but many minor problems pervade the photos, and in 
some cases their captions. This aspect of the book could have benefited from more 
thorough review. As examples, juvenile White-tailed Kites don’t have ruby-red eyes, 
the wing-tips of the perched adult Glaucous-winged Gull look too pale, the wing-tips 
of the standing adult Western Gull are too short (primaries 8-10 appear to be 
lacking!), the right-hand illustration of the “Ashy Storm-Petrel” on page 99 is a dead 
ringer in shape for a Black Storm-Petrel, the wing-tips of the left perched Calliope 
Hummingbirds fall well short of a strongly graduated tail (rather than falling slightly 
beyond a more squared tail), the right wing of the center Ruby-crowned Kinglet on p. 
287 appears deformed, the “immature” Black-throated Sparrow is actually a juvenile, 
and so on. While these facts could annoy or irritate more advanced birders, they may 
not detract from the overall goal of the book. The photos also are a potentially useful 
resource for more experienced birders, but the fact they’ve been “edited” cautions 
against their being used too literally. 

Most birders you and I know may not need this new guide, but what about the kid 
next door, or the curious onlookers you attract at the local park? This book has 
potential for a huge audience, and I congratulate Kaufman for his pioneering spirit 
and broad-based goals. The back cover proclaims that this guide “cuts through the 
clutter to focus on the essentials.” I agree, but the success of this guide will be 
measured by how it works for the intended audience. While I prefer the Sibley Guide 
and the NGS, I believe the Kaufman guide is more useful for beginning birders or “non 
birders,” and I suspect we all can learn from this book and its approach to birding. 

Steve N. G. Howell 


94 


BOOK REVIEWS 


The Sibley Guide to Birds, by David Allen Sibley. 2000. Alfred A. Knopf, New 
York. 544 pp. Paperback. $35.00. ISBN 0-679-45122-6. 

“There is order in the universe, and birds are no exception.” Welcome to the Sibley 
guide (p. 10) and the sixth-order tantric chakra, related to the act of seeing, both 
physically and intuitively. This long-awaited work provides an American answer to the 
increasingly detailed volumes covering other continents, particularly Europe. What- 
ever quibbling follows, let me emphasize that no student of birds is too green or jaded 
to benefit greatly from the art and observations in this guide. 

Sibley treats approximately 744 native and vagrant species plus 66 introduced and 
domestic species. The third edition of the National Geographic Society’s Field Guide 
to the Birds of North America (NG3) treats about 60 additional vagrants and species 
of far offshore waters, so most serious birders will want both guides. I would have 
preferred more complete coverage, particularly of species that pose identification 
pitfalls (e.g., the Red-tailed Tropicbird, which likely occurs regularly in North Ameri- 
can waters), but the Sibley guide’s omission of ultra-rare species can be viewed as a 
progressive step toward a less ego-driven world — one where birders more clearly 
recognize that our most profound contributions to field ornithology and conservation 
are made through such communitarian undertakings as breeding-bird atlases, Christ- 
mas Bird Counts, and purposeful bird-banding. Such projects require sound ground- 
ing in the local avifauna rather than esoteric knowledge of Siberian specialties. 

Acknowledging an increasingly chaotic state of affairs around the species level, and 
that most subspecies cannot be identified reliably in the field, Sibley opted to portray 
intraspecific variability by a system of “natural ecological regions” in place of 
subspecific designations. Thus, Gambel’s White-crowned Sparrow ( Zonotrichia 
leucophrys gambelii) becomes the “west taiga” form, while the partially migratory 
Puget Sound [Z. 1. pugetensis) and sedentary NuttalPs (Z. /. nuttalli) races constitute 
the “Pacific” group. Motivated readers are directed to admirably detailed subspecies 
accounts posted at www.sibleyart.com. This approach may be as valid as any, given 
the deep split between practitioners of the phylogenetic and biological species 
concepts, but I perceive no compelling reason why currently recognized subspecies 
names were not correlated to Sibley’s alternative system in the guide — what is gained 
by banishing this basic information to a web page that many readers will never visit? 

Fourteen pages of introductory material warrant a close read, as they outline 
Sibley’s approach to the subject material and provide unusually insightful information 
on bird topography, hybrids, aberrant plumages, molt, and other topics central to 
birding and ornithology. Species accounts stretch vertically, and each is headed by at 
least two in-flight illustrations, with flight silhouettes for selected species. Each family 
or major group begins with a summary page or spread describing the group’s 
characteristics and showing small images of each species, arranged by genera. Not 
only will newer birders find their match quicker, but I expect that those who start by 
learning the genera, rather than each species independently, will cultivate a deeper 
knowledge of birds and their identification. This is one subtle, but important, way in 
which the Sibley guide has the potential to advance American birding. 

The benefits of the book's distinctive layout (e.g., visual appeal and ease of 
comparison between plates) are not without costs. Most importantly, the rigid 
columnar format limits the possibility of image sizes being modified to show greater 
detail where warranted, or to otherwise use space judiciously. Readers will find 
scattered inconsistencies of scale, such as those between the Least and American 
bitterns and the Bonaparte’s and Black-headed gulls, and note that sexual dimorphism 
is lacking in the genus Accipiter. Groups that could have benefited from larger image 
sizes include the tropiebirds, boobies, several hawks and gulls, and the Oporornis and 
Wilsonia warblers. 

Birders have long expected field guides to feature in-flight depictions of seabirds, 
ducks, hawks, and gulls, but I was somewhat skeptical about the value of extending this 


95 


BOOK REVIEWS 


convention to every species. As it turns out, these flight shots significantly improve 
Sibley’s presentation by showing important wing and tail details and helping to convey 
each species’ unique character. Considering the author’s reputation for identifying 
birds by sounds as seemingly anonymous as a flight note, I was pleased to find the 
voice descriptions unusually detailed and helpful. The occasional eastern bias creeps 
through, however, such as the comparison of the Orange-crowned Warbler’s call note 
to that of a Field Sparrow — western readers can substitute here the Black-chinned 
Sparrow (or a simple “tsit”). Special sections focused on topics such as the identifica- 
tion of swans, scaup, peeps, Spizella spa nows, and meadowlarks are stimulating and 
informative, often discussing plumages, structural features, and even behaviors that 
open both the eyes and the mind. Reminders that some birds may not be identifiable 
appear with welcome regularity. 

Above all, Sibley is a formidable and attentive field ornithologist, and his distinctive 
artwork melds the birder’s eye for proportion, posture, and feather patterns with the 
artist’s gift for breathing depth and life into two dimensions. He employs annotated 
pointers to draw attention to key identifying features more quickly and effectively than 
text ever could. Particularly impressive plates include the shorebirds, hummingbirds 
(despite some tail-pattern miscues that could have benefited from a check of museum 
specimens), Empidonax flycatchers, chickadees through creepers, pipits and wag- 
tails, tanagers, sparrows, and icterids. Larophiles will appreciate 26 nicely executed 
Herring Gull images, but Sibley’s semi-impressionistic style fails to capture fine details 
characteristic of certain other medium-sized and large gulls, such as Thayer’s, and 
stark the white backgrounds detract from all of the gull and tern plates. The warblers 
are generally excellent — exceptionally dull Pine Warblers were a piece of cake for me 
on a recent Maryland trip — although western Orange-crowned Warblers appear too 
coarsely streaked and lack their distinctive pale marginal coverts. Apparent printing 
irregularities in my copy include overly intense rufous tones (on the Brown Thrasher, 
for example) and washed-out greens on the vireos. When will publishers care enough 
to reproduce colors faithfully? All things considered, however, Sibley’s artwork 
outshines that available elsewhere, and no other guide approaches the range of 
plumage variation depicted here. 

Range maps constitute the book’s only true disappointment. Among numerous 
errors evident in southern California and adjacent areas, the Yellow-rumped Warbler, 
Summer Tanager, and Green-tailed Towhee are shown breeding along the southern 
California coast, the coastal range of the Chestnut-backed Chickadee stops around 
San Francisco Bay (this species occurs south to Santa Barbara County), the west coast 
is mistakenly shown as a “main migration route” for the American Golden-Plover, 
Baja California Sur lacks its wintering Clay-colored Sparrows, and many seabirds 
strictly pelagic in this region are shown ranging along our immediate coast. Addition- 
ally, the use of green dots to illustrate broad patterns of rare occurrence is problem- 
atic, at least in California. For example, the Painted and American redstarts are 
depicted similarly in southern California although the latter species is about a hundred 
times more common than the former, and the exceptionally scarce Cape May Warbler 
is granted considerably more California dots than most regularly occurring eastern 
vagrants. For now, the maps in NG3 are far more readable and reliable — another 
good reason not to relegate “old faithful” to the shelf just yet. 

An accomplished guide can facilitate the student’s journey toward the rarified 
echelon of wisdom and consciousness represented by the seventh, or crown, chakra. 
Most birders alive today were assisted along their paths by Roger Tory Peterson’s 
remarkable combination of observational, writing, and artistic talents, and it is 
gratifying to see David Sibley emerge from Peterson’s long shadow to lead a new 
generation of seekers toward birding’s next level. 

Robert A. Hamilton 


96 


FEATURED PHOTO 


FIELD IDENTIFICATION OF FEMALE ALLEN S 
AND RUFOUS HUMMINGBIRDS 

STEVE N. G. HOWELL, Point Reyes Bird Observatory, 4990 Shoreline Highway, 
Stinson Beach, California 94970 


Probably no other pair of North American bird species poses greater field 
identification problems than the Rufous (Selasphorus rufus) and Allen’s {S. sasin ) 
hummingbirds. Observations of anything other than rufous-backed adult males (at 
least outside of known breeding ranges and breeding seasons) are usually lumped as 
“Rufous/Allen’s,” since even a small percentage of adult male Rufous are fully green 
backed, matching the pattern of adult male Allen’s (McKenzie and Robbins 1999). 
Females and immatures are generally considered unidentifiable to species in the field 
although, with excellent views and comparative experience, some immature males 
may be distinguishable by details of rectrix ’width and shape (Stiles 1972, Pyle 1997). 
While researching identification criteria for a photographic guide to North American 
hummingbirds, I found two features helpful in the separation of adult female Allen’s 
and Rufous hummingbirds. 

The most striking feature pertains to the race sedentarius of Allen’s Hummingbird, 
which is endemic to southern California. This taxon is distinguished from nominate 
sasin by measurements, especially bill length, which averages longer in sedentarius, 
but no consistent plumage differences have been reported (Grinnell 1929, Stiles 
1972, Mitchell 2000). Examination of adult female museum specimens revealed that 
sedentarius (n = 24; upper photo) has relatively dull and poorly contrasting cinna- 
mon-rufous sides and flanks with moderate to extensive mottling of iridescent green 
spots. By contrast, nominate sasin (n = 60; lower photo) has brighter cinnamon- 
rufous sides and flanks that contrast more sharply with a white forecollar and white 
median stripe on the underparts. Only about 30% of nominate sasin (n - 60), as well 
as 30% of adult female Rufous Hummingbirds (n = 50), have one to a few iridescent 
green spots on the chest sides, rarely if ever approaching the extent of spotting on 
sedentarius; most have solidly cinnamon-rufous sides and flanks that lack iridescent 
green spotting, a pattern not shown by sedentarius. 

Sedentarius is resident on most of the California Channel Islands, whence it spread 
in the 1960s and 1970s to the adjacent mainland coast (Wells and Baptista 1979). In 
recent years, its range has expanded inland through much of the coastal slope of Los 
Angeles County and parts of Orange County (Hamilton and Willick 1996, Gallagher 
1997), and resident populations of presumed sedentarius also occur along the coast 
nearly to the edge of the breeding range of nominate sasin in Ventura County (K. L. 
Garrett pers. comm., D. E. Mitchell, pers. comm.). There is also some post-breeding 
movement of sedentarius upslope into nearby mountains (Stiles 1972), and the taxon 
has occurred casually south to San Diego County (Unitt 1984). Identification criteria 
proposed here may help document the nesting ranges of these two subspecies, and 
also the detection of wintering Selasphorus hummingbirds other than sedentarius in 
southern California. I encourage observers and banders in a position to examine 
known sedentarius (e.g., on the Channel Islands) to test the usefulness of flank color 
and pattern and to examine whether this feature is helpful for immatures (of which I 
did not see adequate samples); immatures of nominate sasin do not show extensive 
green mottling on their sides and resemble adult females in flank color. 

A second feature worthy of critical examination in the separation of adult female 
Allen’s and Rufous hummingbirds is uppertail-covert pattern and color. About 30% of 
specimens of both sedentarius and nominate sasin show relatively extensive rufous 


Western Birds 32:97-98, 2001 


97 


FEATURED PHOTO 


edgings on their uppertail coverts (forming a distinctive rufous “rump patch”), in 
contrast to narrow or virtually absent rufous edgings on adult female Rufous (n = 50). 
Thus, on adult females, extensive rufous on the uppertail coverts suggests Allen’s 
Hummingbird, a difference illustrated, although not mentioned specifically, by Sibley 
(2000). A smaller sample suggests the same trend in immature females, but because 
10% of immature female Rufous (n = 20 specimens) have fairly distinct rufous fringes 
on their uppertail coverts, the presence of a rufous “rump patch” is not diagnostic of 
immature female Allen’s. Also note that immature males of both Allen’s and Rufous 
hummingbirds typically have extensively rufous uppertail coverts, so correctly deter- 
mining the age and sex of any bird is a prerequisite to specific identification (see Pyle 
1997). 

I thank personnel at the California Academy of Sciences (the late Luis Baptista, 
Douglas J. Long), the Museum of Vertebrate Zoology, University of California, 
Berkeley (Carla Cicero, Ned Johnson), the American Museum of Natural History, 
New York (R. Terry Chesser, Jacqueline Weicker), the National Museum of Natural 
History (Smithsonian Institution), Washington D. C. (James Dean, Gary R. Graves), 
and the Western Foundation of Vertebrate Zoology (Jon Fisher) for access to 
specimens in their care. Kimball L. Garrett (Natural History Museum of Los Angeles 
County) and Peter Pyle helped with specimen examination. I am indebted to Larry 
Sansone for use of his photo of a female s edentarius, taken in Los Angeles County, 
California, March 1992, and to Ian C. Tait for his photo of a female sasin taken in 
Marin County, California, June 1971. Garrett and Don E. Mitchell contributed 
information on the range of sedentarius, and the manuscript benefited from review 
by Robert A. Hamilton, Mitchell, and Pyle. This is contribution number 945 of the 
Point Reyes Bird Observatory. 

LITERATURE CITED 

Gallagher. S. R. 1997. Atlas of Breeding Birds, Orange County, California. Sea and 
Sage Audubon Press, Irvine, CA. 

Grinnell, J. 1929. An island race of the Allen Hummingbird. Condor 31:226-227. 

Hamilton, R. A., and Willick, D. R. 1996. The Birds of Orange County, California: 
Status and Distribution. Sea and Sage Press, Irvine, CA. 

McKenzie, P. M., and Robbins, M. B. 1999. Identification of adult male Rufous and 
Allen’s hummingbirds, with specific comments on dorsal coloration. W. Birds 
30:86-93. 

Mitchell, D. E. 2000. Allen’s Hummingbird ( Selasphorus sasin), in The Birds of 
North America (A. Poole and F. Gill, eds.), no. 501. Birds N. Am., Philadelphia. 

Pyle, P. 1997. Identification Guide to North American Birds, part 1. Slate Creek 
Press, Bolinas, CA. 

Sibley, D. A. 2000. The Sibley Guide to Birds. Knopf, New York. 

Stiles, F. G. 1972. Age and sex determination in Rufous and Allen’s hummingbirds. 
Condor 74:25-32. 

Unitt, P. 1984. The birds of San Diego County. San Diego Soc. Nat. Hist. Mem. 13. 

Wells, S. A., and Baptista, L, F. 1979. Breeding of Allen’s Hummingbird ( Selasphorus 
sasin sedentarius ) on the southern California mainland. W. Birds 10:83-85. 


98 


WESTERN FIELD ORNITHOLOGISTS 
26TH ANNUAL MEETING 

27-30 September 2001 • Reno, Nevada 

Registration over the Internet will be available on the WFO website, 
www.wfo-cbrc.org, by 1 June 2001. For conference information, contact 
Lucie Clark at luclark@sierra.net; 335 Ski Way #300, Incline Village, NV 
89451; 775-831-2909) 

Call for Papers and Poster Presentations 

Guidelines: 

1 . Oral and poster presentations should reflect original research, or summa- 
rize existing unpublished information, and be presented in a manner that will 
be of interest to serious amateur field ornithologists. Talks and posters 
relating to the following general themes are especially solicited for the 
current meeting, but other topics are also welcome. 

• Systematics, biogeography, and geographic variation of birds of the 
Pacific Coast region, the North American interior, and the interface 
between the two 

• New information on field identification problems relevant to the birds 
of western North America and the eastern Pacific Ocean 

• Techniques for field study of birds, including censusing, monitoring, 
and other studies; results of studies resulting from the application of 
such techniques 

• Ecology, population biology, and conservation of birds in the state of 
Nevada or any of the bioregions or habitats it represents (Great Basin, 
Mojave Desert, Sierra Nevada, Columbia Plateau) 

2. We expect to allot 20 minutes per oral presentation, which should 
include 5 minutes for questions and discussion; longer time slots (30 
minutes) are negotiable. 

3. Posters should fit within a width of 6 feet. 

4. An abstract of your presentation or poster should be submitted electroni- 
cally to Ted Floyd (tedfloyd57@hotmail.com) or as hard copy (Ted Floyd, 
Great Basin Bird Observatory, 1 East First Street, Suite 500, Reno, NV 
89501), no later than 30 June 2001. All abstracts should be submitted in 
the following format: 

• Your Last Name, Your First Name. Your affiliation (if any), complete 
mailing address, e-mail address (optional). Title of Your Talk. Brief (300 
word maximum) summary of the goals, results, and conclusions of your 
study. 

We look forward to seeing you in Reno! 


99 


Silent Auction: 

Top-of-the-Line Kowa Spotting Scope 

WFO is conducting a silent auction of a Kowa TSN-824 spotting scope (straight, 82 
mm, fluorite lens) with a TSE-27, 20-60x zoom eyepiece. All proceeds from the 
auction will go to the publication fund for Western Birds. The suggested retail price 
of the scope and eyepiece is $1395; some discount retailers sell them for as low as 
$1200. Bidding for the scope and eyepiece will start at $1100. The successful bidder 
will be announced at the Saturday night dinner at WFO's meeting in Reno (29 
September 2001). Bidders will have until the day before the dinner to submit bids; 
sealed bids will be accepted at the meeting. Mail-in bids should be postmarked no later 
than 15 September 2001 and sent to: 

Lucie Clark, Recording Secretary 
Western Field Ornithologists 
335 Ski Way #300 
Incline Village, NV 89431 

Good luck! 

Mike San Miguel 

President, Western Field Ornithologists 


100 


World Wide Web site: 

WESTERN BIRDS www.wfo-cbrc.org 

Quarterly Journal of Western Field Ornithologists 

President: Mike San Miguel, 2132 Highland Oaks Dr., Arcadia, CA 91006; 
sanmigbird@aol.com 

Vice-President: Daniel D. Gibson, University of Alaska Museum, 907 Yukon Dr., 
Fairbanks, AK 99775-6960 

Treasurer/Membership Secretary: Dori Myers, 6011 Saddletree Lane, Yorba 
Linda, CA 92886 

Recording Secretary: Lucie Clark, 9889 Tahoe Blvd., #56, Incline Village, NV 
89451 

Directors: Kimball Garrett, Daniel D. Gibson, Bob Gill, Gjon Hazard, Dave Krueper, 
Mike San Miguel, W. David Shuford, Mark K. Sogge, David Yee 

Editor: Philip Unitt, San Diego Natural History Museum, P.O. Box 121390, San Diego, 
CA 92112-1390; birds@sdnhm.org 

Associate Editors . Daniel D. Gibson, Robert A. Hamilton, Ronald R. LeValley, 
Tim Manolis, Kathy Molina, Mark K. Sogge 

Graphics Manager: Virginia P. Johnson, 4637 Del Mar Ave., San Diego, CA 92107 

Photo Editor: Peter La Tourrette, 1019 Loma Prieta Ct., Los Altos, CA 94024 

Featured Photo: Robert A. Hamilton, 34 Rivo Alto Canal, Long Beach, CA 90803 

Book Reviews: Steve N.G, Howell, Point Reyes Bird Observatory, 4990 Shoreline 
Highway, Stinson Beach, CA 94970 

Secretary, California Bird Records Committee: Guy McCaskie, P.O. Box 275, 
Imperial Beach, CA 91933-0275; guymcc@pacbell.net 

Chairman, California Bird Records Committee: Richard A. Erickson, LSA Associ- 
ates, 1 Park Plaza, Suite 500, Irvine, CA 92614; richard.erickson@lsa-assoc.com 


Membership dues, for individuals and institutions, including subscription to Western 
Birds- Patron, $1000.00; Life, $400.00 (payable in four equal annual installments); 
Supporting, $60 annually; Contributing, $34 annually; Family, $26; Regular U S. 
$22 for one year, $41 for two years, $60 for three years, outside U.S. $27 for one 
year, $51 for two years, $73 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 orders for back issues and special publications to the Treasurer. Make 
checks payable to Western Field Ornithologists. 

Back issues of Western Birds within U.S. $24 per volume, $6.00 for single issues, 
plus $1.00 for postage. Outside the U.S. $30 per volume, $7.50 for single issues. 

The California Bird Records Committee of Western Field Ornithologists recently 
revised its 10-column Field List of California Birds (January 2000). The last list covered 
606 accepted species; the new list covers 613 species. Please send orders to WFO, 
c/o Dori Myers, Treasurer, 6011 Saddletree Lane, Yorba Linda, CA 92886. 
California addresses please add 7.75% sales tax. 

Quantity: 1-9, $1.50 each, includes shipping and handling. 10-39, $1.30 each, add $2.00 
for shipping and handling. 40 or more, $1.15 each, add $4.00 for shipping and handling. 


Published June 15, 2001 


ISSN 0045-3897 






Vol. 32, No. 2, 2001 




Volume 32, Number 2, 2001 

Recent Bird Records From the Grand Canyon Region, 1974-2000 


Charles T. LaRue, Lara L. Dickson, Nikolle L. Brown, 

John R. Spence, and Lawrence E. Stevens 101 

Occurrence Patterns of Peregrine Falcons on Southeast Farallon 
Island, California, by Subspecies, Age, and Sex 
Sasha Earnheart-Gold and Peter Pyle 119 

NOTES 

Low-Elevation Nesting by Calliope Hummingbirds in the Western 

Sierra Nevada Foothills Brian D. C. Williams 127 

Rapid Second Nesting by Anna’s Hummingbird Near Its Northern 

Breeding Limits Ann Scarf e and J. Cam Finlay 131 

Nesting of Brandt’s Cormorants in the Northern Gulf of California 

Juan Ceroantes-Sanchez and Eric Mellink 134 

A Potential Threat to Bald Eagles in Baja California Sur, Mexico 

Gustavo Arnaud, Edgar Amador, and Marcos Acevedo 136 

Book Review Howard L. Cogswell 137 

Featured Photo Walter Wehtje 141 


Cover photo by © Martin Meyers of Truckee, California: Le Conte’s 
Sparrow {Ammodramus teconteii). Miller’s Rest Stop, Nevada, October, 
2000. If accepted, this will be the second record of Le Conte’s Sparrow 
for the state of Nevada. 


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 include 
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 the Rocky Mountain and Pacific states 
and provinces, including Alaska and Hawaii, western Texas, northwestern Mexico, and the 
northeastern Pacific Ocean. 

Send manuscripts to Kathy Molina, Section of Ornithology, Natural 
History Museum of Los Angeles County, 900 Exposition Blvd., Los Angeles, CA 
90007. For matter of style consult the Suggestions to Contributors to Western Birds 
(8 pages available at no cost from the editor) and the Council of Biology Editors Style 
Manual (available for $24 from the Council of Biology Editors, Inc'", 9650 Rockville 
Pike, Bethesda, MD 20814). 

Reprints can be ordered at author’s expense from the Editor when proof is returned or 
earlier. 

Good photographs of rare and unusual birds, unaccompanied by an article but 
with caption including species, date, locality and other pertinent 
information, are wanted for publication in Western Birds. Submit photos and 
captions to Photo Editor. Also needed are black and white pen and 
ink drawings of western birds. Please send these, with captions, to 
Graphics Manager. 


WESTERN BIRDS 

Volume 32, Number 2, 2001 



RECENT BIRD RECORDS FROM THE GRAND 
CANYON REGION, 1974-2000 

CHARLES T. LaRUE, 3525 W. Lois Lane, Flagstaff, Arizona 86001 

LARA L. DICKSON, Grand Canyon National Park, 823 N. San Francisco St., Suite 
B, Flagstaff, Arizona 86001 

NIKOLLE L. BROWN, 7779 N. Leonard, Clovis, California 93611 

JOHN R. SPENCE, Glen Canyon National Recreation Area, P. O. Box 1507, Page, 
Arizona 86040 

LAWRENCE E. STEVENS, P. O. Box 1315, Flagstaff, Arizona 86002 

ABSTRACT: We report information on 100 species from the Grand Canyon 
region from 1974 to 2000; of these, 98 are from the Colorado River and 18 species 
are new to the Grand Canyon region. We compiled new seasonal information on 33 
species and breeding information on 11 species. Changes are due, among other 
factors, to a change in habitats resulting from the construction and operation of Glen 
Canyon Dam. Flow regulation from the dam increased water clarity, and numbers of 
many waterbirds (primarily Anseriformes) subsequently increased. For example, the 
Common Goldeneye did not occur in Grand Canyon before the late 1980s, but since 
then it has become the most abundant wintering waterfowl. Stabilization of the river’s 
flow has also allowed the growth and expansion of riparian vegetation. Birds we 
believe are responding to this increase in vegetation include Bell’s Vireo, Sharp- 
shinned Hawk, and wintering Empidonax flycatchers and sparrows. Also, there has 
been an increase in coverage, e.g., bird-monitoring river trips in winter. Some species 
are increasing in number and/or expanding their ranges in Arizona, such as the 
Double-crested Cormorant, Cattle Egret, Ross’s Goose, White-winged Dove, Anna’s 
Hummingbird, Black Phoebe, Brown-crested Flycatcher, Crissal Thrasher, Summer 
Tanager, and Common Crackle. Last, improved field identification has helped in 
discerning closely related species. 

This report discusses bird observations from the Grand Canyon region from 
1974 through 2000. We report information on 100 species, of which 18 are 
new to the region. The Grand Canyon, located in northern Arizona, covers 
approximately 1000 km 2 and several life zones from lower Sonoran along the 
Colorado River (elevation 350 m) to boreal forests (2800 m; Figure 1). The 
physiographic Grand Canyon falls within the complex boundaries of six land- 


Western Birds 32:101-118, 2001 


101 



RECENT BIRD RECORDS FROM THE GRAND CANYON REGION, 1974-2000 



Figure 1 . The Grand Canyon region, showing the Colorado River corridor and major 
tributaries. 


management jurisdictions, including the National Park Service (Grand Canyon 
National Park, Lake Mead National Recreation Area, Glen Canyon National 
Recreation Area), the Bureau of Land Management* the U. S. Forest Service 
(Kaibab National Forest), the Navajo Nation, the Hualapai Nation, and the 
Havasupai Nation (Billingsley and Hampton 1999). 

Bird diversity and distribution in the Grand Canyon region have been 
summarized by Bailey (1939) and Brown et al. (1984, 1987, 1993); Stevens 
et al. (1997) and Sogge et al. (1998) made additional studies of riparian and 
waterbirds along the Colorado River. The completion of Glen Canyon Dam 
in 1963 eliminated the nearly annual floods that scoured the banks of the 
Colorado River through the Grand Canyon and allowed the growth and 
expansion of native and nonnative riparian thickets (primarily tamarisk, 
Tamarix ramosissima ) the length of the corridor (Turner and Karpiscak 
1980, Johnson 1991). Flow regulation has increased the populations of 
waterbirds (primarily Anseriformes) along the Colorado River in lower Glen 
Canyon and the upper Grand Canyon, where density is a function of water 
clarity (Stevens et al. 1997, Glen Canyon N.R.A. files). 

The majority of observations we report were recorded during investiga- 
tions of ecological impacts of Glen Canyon Dam on the river and adjacent 
riparian zones. Most of our data were obtained during projects funded by the 
Bureau of Reclamation and administered by Glen Canyon Environmental 


102 


RECENT BIRD RECORDS FROM THE GRAND CANYON REGION, 1974-2000 


Studies and Grand Canyon Research and Monitoring Center. Several 
projects were administered by Glen Canyon National Recreation Area, 
including 12 bird-monitoring trips in the Grand Canyon from 1997 to 1999 
during January, February, April, May, and June and a bird-monitoring 
project in Glen Canyon N.R.A. from 1992 to present. Sources of our other 
records include >200 research and commercial river trips from 1975 
through 1999 by Stevens, projects on both the river and canyon rims by 
Grand Canyon National Park staff, and personal time and personal commu- 
nications. 

Trips on the Colorado River were conducted via motor and rowboats 
during all seasons. River trips were from river mile (RM) 0 at Lee’s Ferry to 
RM 2S0 at Pearce Ferry (elevation 950 m to 350 m). We covered the Glen 
Canyon reach (RM 0 to RM -16 at Glen Canyon Dam) via motorboat. 
Locations on the Colorado River are given in river miles from Lee’s Ferry 
(Stevens 1983) followed by L (left bank looking downstream) or R (right 
bank), where applicable. Records are given in a seasonal order; for example, 
observations of a winter resident start in the fall and go through the spring. 
Observers are noted in the text with their initials for each species, and their 
full names are given alphabetically in the acknowledgments. 

Our reasons for reporting these observations include new species for the 
Grand Canyon region, new seasonal records, new breeding records, changes 
in status, including distribution and abundance, rarely encountered species, 
and range expansion. The published sources of status and distribution 
information from which we determined the significance of our records are 
Woodbury and Russell (1945), Phillips et al. (1964), Monson and Phillips 
(1981), Brown et al. (1984, 1987, 1993), Brown (1985), Stevens et 
al.(1997), and Sogge et al. (1998). For a broader regional perspective, we 
referred to the account by Rosenberg et al. (1991) of the lower Colorado 
River Valley (LCRV), from Davis Dam to Mexico (primarily the Arizona- 
California border) and to Monson and Phillips (1981) for the state of 
Arizona. For current information on the status of unusual species and 
records in Arizona, we referred to Rosenberg and Witzeman (1998, 1999). 

We report 18 species new to the Grand Canyon region: the Least Bittern, 
Greater White-fronted Goose, Ross’s Goose, Trumpeter Swan, Sanderling, 
Caspian Tern, White-winged Dove, Yellow-bellied Sapsucker, Eastern 
Phoebe, Scissor-tailed Flycatcher, Hutton’s Vireo, Blue Jay, Northern Parula, 
Prairie Warbler, Prothonotary Warbler, Bobolink, Common Grackle, and 
Streak-backed Oriole. Our data bring the total number of waterbird species 
in the Grand Canyon region to 91, adding six species (and one corrected 
identification) to the list by Stevens et al. (1997). 

SPECIES ACCOUNTS 

COMMON LOON Gavia immer. Sogge et al. (1998) reported one previous record 
from the river corridor. Six of our seven records indicate that the Common Loon is a 
rare transient primarily in fall. We observed individuals at Lee’s Ferry on 8 October 
1992 (ACP), RM 51.5 on 19 October 1992 (LES), RM -3.6 on 2 November 1998 
(JRS, CTL), RM -1.8L on 7 November 1996 (BKR, TMH), RM -12.0 on 12 
November 1993 (ACP, BKR), and RM 1.4 on 12 February 1999 (PGF, RKR, LLD). 


103 


RECENT BIRD RECORDS FROM THE GRAND CANYON REGION, 1974-2000 


A very unusual summer occurrence was of one at RM 214 on 24 July 1998 (NLB, 
photo). The Common Loon is an uncommon fall and winter migrant at lakes 
throughout Arizona and is most frequent in the LCRV. 

PACIFIC LOON Gauia pacifico. One was at Lee’s Ferry 9-26 February 2000 
(KSB, BRB, CTL). There is one other record from the Grand Canyon region, also 
from Lee’s Ferry in winter. This species is rare but regular in winter on lakes in the 
LCRV and elsewhere in central and southern Arizona. 

HORNED GREBE Podiceps auritus. We report the first records from the river 
corridor, seven from Glen Canyon: one at Lee’s Ferry 27-31 October 1996 (CTL), 
one each at RM -13.5L and -2.5L on 7 November 1996 (BKR, TMH), one at Lee’s 
Ferry on21 December 1997 (JRS), two at Lee’s Ferry on 1 January 1998 (JRS, BKR, 
CEG), one at RM -3.5 on 8 January 1997 (JRS, BKR, CTL), and two at RM -9.4 on 
30 March 1995 (ACP). This species is a rare to uncommon migrant and winter visitor 
throughout Arizona. 

AMERICAN WHITE PELICAN Pelecartus erythrorhynchos. One was at the base 
of Glen Canyon Dam on 30 January 1995 (ACP); this is the first winter record of the 
species from northern Arizona. Five were at RM 209.0 on 12 April 1999 (BHD, 
LLD, JAH). A flock of approximately 40 birds was at RM 8 on 8 May 1999 (reported 
to LES by TV). Nine were at RM 133 on 28 May 1999 (reported to LES by AP). A 
flock of 50 was at RM 120 in late August 1990 or 1991 (reported to LES by SH). Two 
river guides reported (to LES) a flock of 200 birds at RM 150 in mid-summer 1998. 
There are five previous records from the Grand Canyon region. The White Pelican is 
a rare but irregular transient throughout the year in Arizona. 

BROWN PELICAN Pelecartus occidentalis. One was observed in Glen Canyon at 
RM -14.5 on 9 June 1992 (ACP), and another was at RM 170R on 4 August 2000 
(NLB, photo). There is one other report from the Grand Canyon region. This species 
is a rare but irregular late summer and fall wanderer to lakes in the LCRV and 
elsewhere in Arizona. 

DOUBLE-CRESTED CORMORANT Phalacrocorax auritus. Brown et al. (1984) 
reported this species as a common permanent resident in the upper portion of Lake 
Mead and an irregular transient along the river upstream from there. Beginning as 
early as 1985 this species began appearing in winter at the base of Glen Canyon 
Dam, and since 1992 the numbers have steadily increased to a high of 22 individuals 
in 1999 (JRS). 

LEAST BITTERN Ixobrychus exilis. We report the first record of this species from 
the Grand Canyon region: one at Quartermaster Canyon, RM 260. 1L, upper Lake 
Mead, on 22 January 1999 (CTL, JRS, photo NLB, BHD). This species is rare in 
winter in the upper LCRV. There are two other northern Arizona records. 

GREAT BLUE HERON Ardea herodias. Incubating birds were seen on 10 nests at 
RM 275.0 on 7 April 2000 (CTL, LLD). Four pairs were on nests in tamarisks at the 
mouth of Burnt Springs Canyon (RM 259R) on 15 May 1990 (LES). The species nested 
in tamarisks at RM 0L (Lee’s Ferry) in 1998 and 1999, successfully raising young both 
years (LES, JRS, NLB). Another nesting attempt (on a ledge 100 m above the river) at 
RM -13. 3R in 1998 was unsuccessful (BKR, TMH). These are the first nesting attempts 
in the Grand Canyon region of this common wanderer along the Colorado River. 

CATTLE EGRET Bubulcus ibis. Four records: one at Marble Canyon Lodge on 24 
March 1999 (LLD, JAH), one at Glen Canyon Dam on 21 April 1994 (JDG), one at 
Lee’s Feny on 25 April 1995 (JDG), and three between RM -8.8 and -10.8 on 26 
April 1995 (ACP). There were two previous records from the Grand Canyon region. 
This species has recently become more abundant as a migrant in northern Arizona 
(CTL unpubl. data). 


104 


RECENT BIRD RECORDS FROM THE GRAND CANYON REGION, 1974-2000 

GREATER WHITE-FRONTED GOOSE Anser albifrons. The first record of this 
species from the Grand Canyon region is of a single bird observed at RM 158 on 11 
September 1997 (reported to LES by SH). The White-fronted Goose is a rare fall 
migrant in northern Arizona; there are seven records from the adjacent Navajo Nation 
(CTL unpubl. data). 

ROSS’S GOOSE Chen rossii. One was observed at RM 2 on 8 March 1993 
(reported to LES by KJB, CG), and another (or the same) was seen at RM -12.0 on 
19 March 1993 (ACP, photo JRS). These are the first records of this species from the 
Grand Canyon region. Ross's Goose is a rare winter resident in the LCRV and has 
recently increased throughout Arizona as a migrant and winter resident. 

TRUMPETER SWAN Cygnus buccinator. Two first-year birds that had been 
painted bright pink on one side were observed between RM 7 and 9.5 on 14 and 15 
February 1997 (LES), observed and photographed 25-27 February 1997 between 
RM 4 and 20.5 by VM (reported to LES), and at RM 5 on 10 March 1997 (DH), 
Presumably the same two birds were seen near RM 6 on 12 January 1998 (JRS, NLB, 
KME, CTL), at RM 10.6 on 4 January 1999 (LES) and 8 January 1999 (DGS, NLB, 
CTL), and at RM 12 on 22 January 1999 (reported to LES by RGB). These marked 
individuals originated in Wyoming, through a cooperative effort to augment the 
population there (S. Patla pers. comm.). These are the first records of this species 
from the Grand Canyon region. 

WOOD DUCK Aix sponsa. We accumulated nine records from the river corridor, 
seven from the Glen Canyon reach and Marble Canyon, two from RM 180-194. Five 
sightings were in spring (from 22 April to 8 June), four in fall or early winter (24 
September to 30 December). To our knowledge, these are the first records of this 
species from the river corridor. This species is sparse in fall and winter throughout 
Arizona, a status reflected in these Grand Canyon region records. 

EURASIAN WIGEON Anas penelope. An adult male was at RM -1.8L on 13 
March 1998 (CEG, BKR). There was one previous report from the river corridor. 
There are more than 50 records of this rare migrant and winter visitor in Arizona. 

MALLARD Anas platyrhynchos. The first breeding record of this species in the 
river corridor was in 1983 (Brown et al. 1987), and the population increased greatly 
after 1986 (Stevens et al. 1997). At present, Mallards breed commonly from Glen 
Canyon Dam to the Little Colorado River (RM 61) but apparently not farther 
downstream. Nests with up to 10 eggs have been found in dense stands of horsetail 
(. Equisetum spp.) slightly above the typical high water mark from 15 April through 15 
June (LES). The relatively long incubation period and the nests’ proximity to the 
water’s edge make this species highly susceptible to inundation as dam operations 
change. 

A probable female Cinnamon Teal { Anas cyanoptera) with three young was 
observed at RM 47. 8L on 8 June 1987 (LES, WB). Although this and the Blue-winged 
Teal (A. discors) are common migrants, neither has been recorded previously 
breeding on the Colorado River in the Grand Canyon region. 

GREATER SCAUP Aythya marila. Since 1995 up to 30 individuals have been 
found wintering from Lee’s Ferry upstream to RM -7.0 (Glen Canyon N.R.A. files). 
These are the first records from the river corridor in the Grand Canyon region. 
Monson and Phillips (1981) reported four records for Arizona based on specimens but 
noted that the species may be more common in the state. It has since been recognized 
as a rare but regular winter resident in the LCRV 7 and elsewhere in Arizona (Brown 
1985), a status reflected in the Grand Canyon region. 

SURF SCOTER Melanitta perspicillata. There is one prior record from the Grand 
Canyon region. We report an additional four: a male at Lee’s Ferry from 19 


105 


RECENT BIRD RECORDS FROM THE GRAND CANYON REGION, 1974-2000 


September 1991 to 26 April 1994 (ACP, JDG), a female at Glen Canyon Dam on 22 
October 1993 (ACP), two birds at Lee’s Ferry on 4 November 1995 (ACP), and one 
at Lee’s Ferry on 26 April 1995 (ACP). This species is a rare but regular migrant and 
winter visitor throughout Arizona. 

WHITE- WINGED SCOTER Melanitta fusca. We report five records from the Glen 
Canyon reach: one at Lee’s Ferry from 21 November 1996 to 8 February 1997 
(CTL, JRS), four at Lee’s Ferry on 10 December 1990, with two there on 31 
December 1990 (LES), one at Lee’s Ferry on 25 January 1998 (CTL et al.), and one 
at RM -14.5 on 27 February 1995 (ACP). These observations represent the first 
records from the river corridor in the Grand Canyon region. This species is a rare 
migrant and winter visitor in Arizona, 

LONG-TAILED DUCK Clartgula hpemalis. We report eight records: one at Glen 
Canyon Dam on 30 November 1993 (ACP), one at Lee’s Ferry on 10 December 
1990 (NCK), an immature at Glen Canyon Dam from 2 January to 2 March 1999 
(CTL, RKR, CEG, LLD, JAH), a male near Lee’s Ferry from 16 January to 4 March 
1992 (ACP, NCK), one on the Glen Canyon reach from 27 January to 9 April 1996 
(JEH, JDG, JRS, ACP, BKR), one at RM -9.9 from 27 February to 19 March 1995 
(ACP, MKS), one at Lee’s Ferry on 19 March 1993 (ACP), and one at Glen Canyon 
Dam 27-28 March 2000 (CTL, BRB). These represent the first records from the river 
corridor and indicate that this species is a sparse winter resident in the Glen Canyon 
reach. This species is rare and irregular in winter in the upper LCRV. 

BUFFLEHEAD Bucephala albeola. One was seen at Lee’s Ferry on 23 June 1995 
(JDG), a female was below Glen Canyon Dam on 24 June 1998, and a male was at 
RM -9.5 on 25 and 26 June 1998 (CTL, NLB). These are the first summer stragglers 
reported from the Grand Canyon region. A few have also been noted in summer in 
the LCRV. 

COMMON GOLDENEYE Bucephala clangula. Brown et al. (1987) did not list this 
species as occurring in Grand Canyon. Since the late 1980s, however, it has become 
the most abundant wintering waterfowl along the Colorado River in the Grand 
Canyon, being observed commonly from Glen Canyon Dam to RM 61 from 
November through March (Stevens et al. 1997, Spence et al. 1998). The largest 
single-day count from the Glen Canyon reach was 2380 on 8 January 1998 (JRS, 
CTL, TMH). Records of nonbreeding birds remaining into summer include nine below 
Glen Canyon Dam on 3 June 1999 (CTL), single birds at RM -4.6 and RM -9.8 on 
23 June 1995 (JDG), one male and two females in the Glen Canyon reach from 19 
May to 19 August 1998 (CTL, NLB), and two below Glen Canyon Dam on 16 August 
2000 (CTL). This species has increased in the LCRV since 1960, 

BARROW’S GOLDENEYE Bucephala islandica. Since 1992, this species has 
been found annually in winter below Glen Canyon Dam (Spence et al. 1998) with up 
to 80 present in some years. The birds are generally present from late October to early 
March; however, two lingered until 15 June in 1999 (CTL, LLD). A female observed 
at RM 66.0 on 11 January 1999 (CTL, BHD) is the only record of this species below 
Lee’s Ferry. Barrow’s Goldeneye is a rare but irregular winter visitor below dams in the 
LCRV. 

HOODED MERGANSER Lophodptes cucullatus. One to four males were seen at 
Lee’s Ferry from 20 November 1990 to 11 February 1991, and two males were 
observed there on 7 January 1992 (LES, NCK). From 1992 through 1999, 24 
records of 61 individuals were recorded from mid-November through April during 
waterfowl surveys in the Glen Canyon reach (Glen Canyon N.R.A. files). Brown et al. 
(1987) listed this species as casual in the region. To our knowledge these are the first 
records from the river corridor and indicate that this species is an uncommon migrant 
and winter resident in the Glen Canyon reach, as it is in the LCRV. 


106 


RECENT BIRD RECORDS FROM THE GRAND CANYON REGION, 1974-2000 

RED-BREASTED MERGANSER Mergus senator. Brown et al. (1984, 1987) 
listed four records from the Grand Canyon region and described the species as casual 
in the region; however, our observations represent the first records from the Colorado 
River corridor. Four were seen at RM -6.2 on 13 February 1998 (CTL, CEG). Single 
males were seen at RM 64.0 on 5 April 1999 and RM 221 on 13 April 1999 (CTL, 
LLD, JAH, BHD). A mortally wounded female, probably struck by a Peregrine Falcon, 
was examined at RM 211L on 30 April 1994 (LES). The Red-breasted Merganser is 
a sparse migrant throughout Arizona and is occasionally a common migrant in LCRV. 

OSPREY Pandiort haliaetus. We report the first winter records from both the 
Grand Canyon region and northern Arizona. Individuals were observed at Lee’s 
Ferry on 10 December 1990 (NCK) and in January 1991 (KJB), in Glen Canyon at 
RM -14.0 on 9 January 1998 (JDG) and on 18 January 1997 (JRS). 

SHARP-SHINNED HAWK Accipiter striatus. We accumulated seventeen winter 
records from the river corridor in January and February of 1998 and 1999 (CTL, 
NLB, JRS, LLD, RKR, DGS). These and four other winter records (Brown et al. 
1984, Sogge et al. 1998) indicate that this species is an uncommon winter resident of 
the river corridor, probably exploiting small birds wintering in the dense riparian 
vegetation that has proliferated since the building of the dam. 

COMMON BLACK-HAWK Buteogallus anthracinus. We report three records: 
individuals were seen at RM 60.5 on 10 April 1998 (LES), RM 205.0 on 19 April 
2000 (SEO, DS, DT), and Pumpkin Spring (RM 212) on 28 June 1997 (reported to 
LES by JC). These three stragglers and a few previous records from the Grand 
Canyon (B. T. Brown pers. comm.) represent the first records in northern Arizona 
aside from the extreme northwest corner (cf. Monson and Phillips 1981). Addition- 
ally, the Black Hawk is a casual transient and recent summer visitor in the LCRV and 
a summer resident of permanent streams, primarily in central Arizona. 

RED-SHOULDERED HAWK Buteo lineatus. An immature, likely of the distinc- 
tive California race B. I. elegans , was seen at Quartermaster Canyon, RM 260. 1L, 
upper Lake Mead, 26-27 February 1998 (CTL, RKR, NLB, CAL). This represents 
the second record from the Grand Canyon region. This species is a rare transient in 
Arizona that appears to be nearly annual in its occurrence. 

ZONE-TAILED HAWK Buteo albonotatus. Listed as casual by Brown et al. 
(1984), this hawk is now confirmed as a breeding species. Six successful breeding 
attempts were recorded from 1995 to 1999 on the south rim (Grand Canyon N.P. 
files, reported to LLD by EFL). 

VIRGINIA RAIL Rallus iimicola. One was seen at RM -8.7L on 2 January 1999 
(RKR, CTL), one. was seen at RM 52R on 10 January 2000 (JRS), and one was heard 
at RM 246. 0L on 21 January 1998 (CTL, JRS, BHD, NLB, KME). These are the first 
winter records from the Grand Canyon region, though the species is a common to 
abundant winter resident in the LCRV. 

SORA Porzana Carolina. One and two were heard calling at RM 246. 0L and 
RM 265. 0L, respectively, on 22 January 1998 (CTL, JRS, NLB). Nine and one were 
seen at RM 260. 0L and RM 265. 0L, respectively, on 27 February 1998 (CTL, RKR, 
CAL, NLB). One was observed at RM 259.5R on 4 April 2000 (CTL, LLD). There were 
two previous records from the Grand Canyon region. The Sora is a common winter 
resident and transient in the LCRV and a common transient throughout Arizona. 

COMMON MOORHEN Gallinula chloropus. Three were at RM 260. 0L on 22 
January 1998 and 27 February 1998 (JRS, NLB, BHD, CTL), two were there on 13 May 
1998 (NLB, CTL), one was at RM 255.5R on 3 April 2000 (CTL, LLD), and one was at 
Lee’s Ferry on 22 July 1994 (JDG). There was one previous record from the Grand 
Canyon region. This species is a fairly common but local permanent resident in LCRV. 


107 


RECENT BIRD RECORDS FROM THE GRAND CANYON REGION, 1974-2000 


SOLITARY SANDPIPER Tringa solitaria. One was observed at RM -6.5 on 10 
September 1992 (CTL). Although this is only the third record from the river corridor, 
the species is a fairly common migrant throughout Arizona. 

SPOTTED SANDPIPER Actitis macularia. Brown et at (1984) and Sogge et al. 
(1998) each reported one winter record from the river. We accumulated seven 
additional winter records from RM 1 to 73, indicating that this species winters in small 
numbers along the Colorado River. Additionally, three adults with one young were 
observed at RM -8.0 on 1 July 1998 (CTL, NLB), and one adult and one young were 
observed at RM 55.5 R in early July 1987 (LJES). These represent the third and fourth 
breeding records from the river corridor. 

SANDERLING Calidris alba. An immature bird was photographed by John 
Blaustein at RM 11 OR on 28 August 1974. This bird was initially identified as a 
Semipalmated Sandpiper ( Calidris pusilla) and is the basis for that species’ being 
ascribed to the Grand Canyon (B. T. Brown pers. comm., Brown et al. 1984, 1937, 
Stevens et al. 1997). A photograph of this individual in The Hidden Canyon 
(Blaustein 1999) is labeled as a Semipalmated Sandpiper, although it is identifiable as 
an immature Sanderling. Figure 2 is a more diagnostic photograph of the same bird. 
Sanderling is a sparse migrant through the region, and this photograph documents 
the only record from the Grand Canyon region. 

CASPIAN TERN Sterna caspia. We report the first records from the Grand 
Canyon region. One was seen at RM 171.3 on 9 April 1999 (LLD, JAH, CTL, BHD), 
another at Pearce Ferry (RM 280) on 13 May 1998 (NLB, CTL). Although a rare to 
fairly common transient in the LCRV, the Caspian Tern is sparse in northern Arizona. 

WHITE-WINGED DOVE Zenaida asiatica. We report the first records from the 
Grand Canyon region. Single birds were seen at Lee’s Ferry on 27 April 1982 (GMS), 



Figure 2. Sanderling at river mile 110 on 28 August 1974. 

Photo by John Blaustein 


108 


RECENT BIRD RECORDS FROM THE GRAND CANYON REGION, 1974-2000 


8 May 1999 (CEG), and 20 May 1992 (JRS). This species is a rare transient in 
northern Arizona, where records have recently increased (CTL unpubl, data). 

INCA DOVE Columbina inca. One was at RM 98. OR on 15 October 1993 (JB, 
CG), two were at Phantom Ranch (RM 87. 8R) on 17 October 1998 (NLB), one was 
at Lonely Dell, near Lee’s Ferry, on 24 October 1982 (ARL, CTL), and three were at 
Lee’s Ferry on 2 December 1995 (JDG). There are two previous records from the 
Grand Canyon region. This species is an occasional transient in northern Arizona. 

COMMON GROUND-DOVE Columbina passerina. One was observed at Lee’s 
Ferry on 14 October 1991 (SRG). It represents the second record of this species from 
the Grand Canyon region and northern Arizona. 

YELLOW-BILLED CUCKOO Coccpzus americanus. One was at RM -14.2 on 6 
June 1995, and another (or the same) was at Lee’s Ferry on 21 June 1995 (JDG). 
Brown et al. (1984) reported only one record from the Grand Canyon region since 
1950. This species has declined sharply in the western U.S. (Hughes 1999). 

GROOVE-BILLED ANI Crotophaga sulcirostris. One was at Cardenas Marsh, 
RM 71.0L, from 10 to 13 October 1993 (photo CG, Figure 3; JB). Another was 
photographed at the mouth of Salt Trail Canyon in the Little Colorado River gorge 6 
miles above the confluence on 21 October 1992 (reported to CTL by PFR), A third 
was at RM 80.5R on 23 October 1992 (LES). There was one prior record from the 
Grand Canyon region. This is a casual transient primarily in summer and fall in central 
and southern Arizona and elsewhere in the Southwest (Mlodinow and Karlson 1999). 

FLAMMULATED OWL Otus flammeolus. This species is a common summer 
resident of pine forest on the canyon rims, occurring as late as 25 October (on the South 
Bass Trail at the top of the Redwall Limestone in 1999, LES). A migrant was mist-netted 


Figure 3. Groove-billed Ani at Cardenas Marsh, river mile 71, on 10 October 1993. 




\ 


Photo by Chris Geanious 


109 


RECENT BIRD RECORDS FROM THE GRAND CANYON REGION, 1974-2000 


at RM 72R in dense riparian vegetation of Goodding willow (Salix gooddingii) and 
tamarisk on 5 May 1975 (LES). This species is occasionally found at lowland sites 
throughout Arizona, including two prior records from the bottom of Grand Canyon. 

LONG-EARED OWL Asio otus. An adult with three fledglings and the nest site (on 
the ground in dense tamarisk) were found at Cardenas Marsh (RM 71.0L) on 6 June 
1998 (NLB, TJT, CTL). Although this species was reported calling throughout the 
night at Cardenas Marsh from 1 to 3 May 1975 (SWC) and sporadically in subsequent 
years (LES), the 1998 record represents one of the few breeding records from Grand 
Canyon and the first verified from the Colorado River corridor. The Long-eared Owl 
breeds sparsely throughout Arizona. 

NORTHERN SAW- WHET OWL Aegolius acadicus. One was observed at Kwagunt 
Marsh at RM 55. 5R on 20 February 1996 (MJK, JB). Although irregularly present 
(occasionally in large numbers) on the canyon rims, and casually at lowland sites in the 
LCRV, this is the first report from the river corridor. 

COMMON POORWILL Phaiaenoptilus nuttallii. One was seen RM 246. OR, 
upper Lake Mead, on 22 February 1999 (LLD, RKR, CTL). This is the first winter 
record from the Grand Canyon region; however, whether this bird overwintered 
(perhaps hibernating) or was an early migrant is unknown. The presence of an 
overwintering population downstream in the LCRV has yet to be demonstrated. 

WEIIP-POOR-WILL Caprimulgus uociferus. Two birds were heard at Swamp 
Point on the north rim on 27 June 1999 (reported to LLD by JPD, JVJ, NEP). This 
is the third record from the Grand Canyon region; the two previous records were from 
the same location (Brown et al. 1984). This is the northernmost locality for this 
species in Arizona. 

ANNA’S HUMMINGBIRD Calypte anna. Single males were seen at RM 196. OR 
and 198. OR on 16 January 1999, RM 200.5Ron 17 January 1999, and RM 207.8L 
on 12 April 1999 (CTL). These are the first records from the Colorado River corridor 
and are likely associated with recent population increases and expansion in the LCRV, 
elsewhere in Arizona (Monson and Phillips 1981, Rosenberg et al. 1991), and in 
several other western states (DeSante and George 1994). 

COSTA’S HUMMINGBIRD Calypte costae. Brown et al. (1984) and Sogge et al. 
(1998) reported this species along the river corridor primarily from RM 140 
downstream. A remarkable influx occurred in April and May of 1999. We recorded 60 
males throughout the corridor, including 16 from RM 0 to 65 (LLD, JAH, CTL, NLB). 
Single males were also seen as far upstream as RM -8.4R on 26 April 1999 and RM 
-10.0L on 27 April 1999 (LLD, CTL). This influx may be related to the severe 
drought that prevailed in the desert Southwest during the winter. An influx of Costa’s 
Hummingbirds occurred in 1984 (Brown et al.1987), which was also an exception- 
ally dry winter in the Southwest. One was observed at RM -6.0 on 10 April 1984 
(BTB, KDG). 

YELLOW-BELLIED SAPSUCKER Sphyrapicus varius. A first-winter individual 
was at Lee's Ferry from 16 to 30 January 1994 (CTL), and an adult male was at RM 
1.0R on 12 February 1999 (RKR, CTL, LLD). These are the first reports of this 
species from the Grand Canyon region. The Yellow-bellied Sapsucker is a rare winter 
visitor in the southwestern half of Arizona, including the LCRV. 

WILLOW FLYCATCHER Empidonax trail l ii. The decline and virtual disappear- 
ance of the breeding population of this species along the river corridor has been 
documented by Sogge et al. (1997) and Brown (1988). A territorial male was 
observed at RM -7.0L from 27 May to 15 June 1999 (LLD, CTL, JAH); it is the first 
to be noted from the Glen Canyon reach (M. Sogge pers. comm.). 


110 


RECENT BIRD RECORDS FROM THE GRAND CANYON REGION, 1974-2000 

HAMMOND S FLYCATCHER Empidonax hammondii. One seen at RM 193. 8R 
on 23 February 1998 (CTL, RKR) represents the first winter record from the Grand 
Canyon region and northern Arizona. The species is a rare and irregular winter 
resident in the LCRV. 

GRAY FLYCATCHER Empidonax wrightii. The first winter record from the 
Grand Canyon region is of one at RM 207. 5L on 21 February 1999 (CTL, RKR). The 
Gray Flycatcher is a locally uncommon winter resident in the LCRV. 

DUSKY FLYCATCHER Empidonax oberholseri. One was seen and heard at 
RM 49. 1R on 18 February 1998 (RKR, CTL, NLB, CAL). This represents the first 
winter record of the Dusky Flycatcher from the Grand Canyon region. The species is 
a rare and irregular winter resident in the LCRV. 

BLACK PHOEBE Sayornis nigricans. We report the first winter and breeding 
records from the Glen Canyon reach. At RM -12.0L, one bird was seen on 2 January 
1999 (RKR, CTL) and two were seen on 3 and 30 January 2000 (RKR, LLD, JDG). 
Another bird was at RM -11. 5R on 2 February 1999 (CTL, LLD). A bird was seen at 
a newly constructed nest at RM -12 . OR on 20 April 1999; this nest contained two eggs 
on 16 June 1999, and a second nest found the same day at RM -13.4L contained four 
eggs (CTL, JAH). The nest at RM -12. OR contained four eggs on 27 April 2000 
(CTL). A family group was seen at the base of Glen Canyon Dam on 7 July 1997 (CTL, 
DGS, BKR). This species has expanded its range in Arizona, particularly in the LCRV, 
and our records suggest this expansion includes the Grand Canyon region. 

EASTERN PHOEBE Sayornis phoebe. We report the first records of this species 
from the Grand Canyon region. Individuals were observed at RM 1.0R and 1.8R on 
1 April 1999 (CTL, JAH, KGM, BHD, MC, LLD) and at RM 155.2 on 8 April 1999 
(CTL, JAH, LLD, BHD). The Eastern Phoebe is an uncommon migrant/winter 
resident in much of Arizona, but its status in northern Arizona is unclear. 

VERMILION FLYCATCHER Pyrocephalus rubinus. A male was at Lonely Dell, 
Lee’s Ferry, from 25 May to 4 June 1997 (LSB, CEG). This is the sixth record from 
the river corridor. Although a common summer resident in southern Arizona, this 
species is a sparse transient in northern Arizona. 

BROWN-CRESTED FLYCATCHER Mgiarchus tyrannulus. Brown et al. (1984, 
1987) speculated on this species’ occurrence in Grand Canyon. Sogge et al. (1998) 
confirmed its restricted range in the immediate area of RM 198, and LES observed 
one to two pairs in the cottonwoods at Supai Campground in Havasu Canyon 
periodically during the 1980s and 1990s. From 1997 to 1999 we recorded 12 
observations (a maximum of 18 individuals) from RM 194 to 204 (CTL, LLD, NLB, 
JAH, TJT). Remarkably, one was observed at RM 5.2R on 30 May 1999 (reported to 
LLD by TJT, LA, KB). Our records suggest this species is continuing its expansion into 
the Grand Canyon region, as noted in the LCRV (Rosenberg et al. 1991) and 
elsewhere in the Southwest (Johnson 1994). 

EASTERN KINGBIRD Tyrannus tyrannus. One was observed at Lee’s Ferry on 
14 July 1994 (JDG). This is the fourth record from the river corridor of this rare 
annual migrant in Arizona. 

SCISSOR-TAILED FLYCATCHER Tyrannus forficatus. We report the first record 
of this species from the Grand Canyon region: one at RM 74R on 14 June 1996 
(reported to LES by SH, JC). A different (or the same) bird was reported at Thunder 
River ten days later (reported to LES by TF). This species is a recurring summer visitor 
(over 45 records) in Arizona. 

NORTHERN SHRIKE Lanius excubitor. There was one previous record from the 
river corridor. We report four additional records: an adult at RM 5.8R on 17 
December 1999 (CTL, CAB), and at Lee’s Ferry an adult on 1 January 1998 (JRS), 


111 


RECENT BIRD RECORDS FROM THE GRAND CANYON REGION, 1974-2000 

an immature on 15 January 1997 (CTL), and an adult found dead on 1 February 
2000 (CAB; specimen to Univ. of Ariz.). This species is a regular and uncommon 
winter resident in northern Arizona, 

BELL’S VIREO Vireo bellii. Both Brown et al. (1984) and Sogge et al. (1998) 
reported this species as a breeding resident as far upstream as RM 50, with occasional 
vagrants as far as Lee’s Ferry. We report three records above Lee’s Ferry: two at 
RM -8.8L on 2 June 1993 (JRS), two at RM -8.3 on 6 June 1995 (JDG), and one at 
RM -2.3L on 7 July 1997 (CTL, NLB). The presence of two birds each in June of 
1993 and 1995 in Glen Canyon may indicate breeding. Our records suggest that the 
range expansion in the Grand Canyon region documented by Brown et al. (1983) is 
continuing. 

CASSIN’S VIREO Vireo cassirtii. One was seen at RM 193.8R on 10 April 1999 
(CTL, LLD, JAH) and at RM 198. OR on 11 April 1999 (CTL). These records are 
nearly a month earlier than the known spring passage in northern Arizona (Phillips et 
al. 1964) and may represent individuals that wintered locally. This species is known to 
winter in the LCRV and is a common migrant throughout Arizona. 

HUTTON’S VIREO Vireo huttoni. We report the first record of this species from 
the Grand Canyon region and the northernmost record from Arizona. One was heard 
singing and later observed at RM 204. 1R on 21 February 1999 (CTL, LLD, RKR). 
Hutton’s Vireo is a common summer and uncommon winter resident in central and 
southeastern Arizona and a casual transient and winter visitor to the LCRV. 

RED-EYED VIREO Vireo olivaceus. One was obseived at Lonely Dell, Lee’s Ferry, 
on 23 September 1998 (CTL, NLB). There were six previous records from the Grand 
Canyon region. This species is a sparse transient throughout Arizona, but reports 
have declined in recent years. 

BLUE JAY Cyanocitta cristata. One was observed at RM 175.8 on 7 May 1998 
(CTL, PW, photo NLB). This represents the first record of this species from the Grand 
Canyon region and the sixth report from Arizona (Rosenberg and Witzeman 1999). 

TREE SWALLOW Tachycineta bicolor. One was seen at RM 204.0 on 24 
February 1998, and two were seen at RM 223.0 on 25 February 1998 (CTL, RKR, 
CAL). These represent the earliest spring records from the Grand Canyon region and 
were probably early spring arrivals. 

NORTHERN ROUGH-WINGED SWALLOW Stelgidopteryx serripenrtis. One 
was at RM 208.7 on 21 February 1999 (AMD, CTL, LLD), and six were between 
RM 260.2 and Pearce Ferry on 26 and 27 February 1998 (CAL, RKR, NLB). These 
represent the earliest records from the Grand Canyon region and were probably early 
spring migrants. Additionally, one was seen at Lee’s Ferry on 16 January 2000 (CAB) 
during a period of unusually mild weather. Breeding by this species on upper Lake 
Mead was recorded by Brown et al. (1993) and R. McKernan (pers, comm,). We 
report nests farther upstream than has previously been noted. At RM 196.0, a nest 
was observed 10 May 1999 (NLB, BHD), and adults with fledglings were observed on 
10 June 1998 (NLB, PW). 

CACTUS WREN Campylorhynchus brurtneicopillus. This species is a rare to 
uncommon permanent resident in the upper Lake Mead area (Brown et al. 1984) and 
the middle reaches of Peach Springs Wash (LES). A nest containing three eggs was 
found at RM 207. 5L on 12 April 1999 (CTL). To our knowledge, this is the farthest 
upstream Cactus Wren breeding has been reported. 

HOUSE WREN Troglodytes aedon. We report the first seven known winter 
records from the Grand Canyon region: one each at RM 203. 0L, 204. 3R, and 
209. 0L on 17 January 1999 (CTL, NLB, RKR), Spencer Canyon, RM 246. 0L, on 


112 


RECENT BIRD RECORDS FROM THE GRAND CANYON REGION, 1974-2000 

18 January 1999 and 26 February 1998 (CTL, NLB, RKR), and RM 209. 0L on 20 
January 1998 (JRS, CTL) and 21 February 1999 (LLD, RKR). This species is a fairly 
common winter resident in the LCRV but is sparse in winter in northern Arizona. 

WINTER WREN Troglodytes troglodytes. Both Brown et al (1984) and Sogge et 
al. (1998) considered this species to be a rare transient and winter visitor in riparian 
vegetation of the Colorado River and its lower tributaries. Our series of records 
indicates that it may be more common. We recorded 21 observations of 26 individuals 
from the length of the river corridor in 1998 and 1999. These records extend the 
known period of occurrence in the Grand Canyon region from 16 October 1999 (at 
Indian Gardens; CTL) to 1 April 1999 (at RM 5.8R; JAH). The Winter Wren is a rare 
winter resident throughout riparian areas in central and southern Arizona. 

BLUE-GRAY GNATCATCHER Polioptila eaerulea. There was one previous 
winter record from the river corridor, and we add two more. Three were at Lee’s Ferry 
on 5 December 1995 (JDG), three at RM -11. OR on 30 January 2000 (JDG). This 
species is common in southern Arizona in winter but sparse in the northern portions. 

HERMIT THRUSH Catharus guttatus. There were three previous winter records 
(Brown et al. 1984, Sogge et al. 1998). From 1998 to 2000, we accumulated 12 
winter records extending from RM -7.0 to 265.0, with the majority below RM 200. 
These records establish that this species is an uncommon winter resident in riparian 
vegetation along the river corridor, primarily in the lower western portions of the 
Grand Canyon. The Hermit Thrush is a common winter resident in riparian areas in 
central and southern Arizona and the LCRV. 

NORTHERN MOCKINGBIRD Mimus polyglottos. At “Dox Seep" at RM 70L on 
1 June 1995, an adult Blue-gray Gnatcatcher was observed landing on the head of a 
fledgling Northern Mockingbird and unsuccessfully attempting to place a geometrid 
moth larva in the panting mockingbird’s bright yellow mouth (LES). This was the first 
evidence of nesting mockingbirds known to us in the river corridor in the Grand 
Canyon, although singing males are heard here sporadically. 

SAGE THRASHER Oreoscoptes montanus. One was observed at RM 196. OR on 
20 February 1999 (CTL, LLD). This is the third record from the Colorado River in the 
Grand Canyon region and likely represents an early spring migrant. This species is a 
migrant and sparse summer resident on adjacent arid plateaus. 

CRISSAL THRASHER Toxostoma crissale. The status and distribution of this 
species in northern Arizona are poorly known (Phillips et al. 1964, Brown et al. 
1984), but recent field work is clarifying its distribution in the region Brown et al. 
(1984) speculated that it is a locally rare resident in the western Grand Canyon and 
noted only two records from the Colorado River. Sogge et al. (1998) reported four 
additional records from the Colorado River from RM 173.8 to 204. 5R during winter 
and spring surveys. We add five records: one each at RM 246. OR on 22 January 
(CTL, JRS) and 26 February 1998 (RKR, CAL), seven at RM 265.0L on 27 February 
1998 (CTL, RKR, CAL), and one at RM 45. 5R on 2 April 1998 (NLB, CTL). Last, 
one remained at Lonely Dell, Lee’s Ferry, from 30 November 1999 to 9 February 
2000 (CAB, CTL). The Arizona breeding-bird atlas project documented breeding 
across north-central and northwestern Arizona from the Echo Cliffs west (T. Corman 
pers. comm.), and there are several winter records from the western Navajo Nation 
(CTL unpubl. data). It may be concluded that this species is a sparse permanent 
resident throughout much of northwestern Arizona. In the Grand Canyon region, it is 
found primarily on plateaus in open sage desert and juniper woodlands, and along the 
Colorado River from RM 175 downstream to Grand Wash Cliffs (RM 280). 

EUROPEAN STARLING Sturnus vulgaris. Although nonnative starlings are 
widespread throughout the United States, their distribution and ecological role in 


113 


RECENT BIRD RECORDS FROM THE GRAND CANYON REGION, 1974-2000 


Grand Canyon are poorly known. Although single birds are occasionally seen during 
summer, a flock of 50 birds was seen for a full day at Granite Park (RM 209L) on 13 
November 1993 (LES). This record corresponds to a marked period of migratory 
passage in northeastern Arizona (CTL unpubl. data). 

BOHEMIAN WAXWING Bombycilla garrulus. One was observed at Lonely Dell, 
Lee’s Ferry, on 20 May 1992 (JRS). This represents the third record from the Grand 
Canyon region and is perhaps the latest in spring that this species has been reported 
from Arizona (cf. Phillips et al. 1964). The Bohemian Waxwing is a sparse and erratic 
winter visitor to northern Arizona. 

ORANGE-CROWNED WARBLER Vermivora celata. There were no winter 
records prior to our observations. From 1998 to 2000, we accumulated 23 winter 
records from RM -10 to 265. Although these records extend the length of the river 
conidor, the majority are from the lower western end. These records establish that this 
species is an uncommon winter resident in riparian vegetation along the Colorado 
River in the Grand Canyon region, perhaps becoming so only since completion of 
Glen Canyon Dam. Furthermore, it is a common winter resident along the LCRV. 

NORTHERN PARULA Parula amerieana. We report the first record of this 
species from the Grand Canyon region. An immature male was at Lee’s Ferry from 19 
November to 29 December 1999 (CTL, CEG). The Parula is a rare but regular 
migrant and winter resident in the LCRV and throughout Arizona, with reports 
increasing in recent years. 

PRAIRIE WARBLER Dendroica discolor. We report the first record of this 
species from the Grand Canyon region. An adult male was seen at Spencer Canyon, 
RM 246. 0L, on 22 January and 26 February 1998 (JRS, CTL, KME, CAL, RKR). 
This species is casual in Arizona. 

AMERICAN REDSTART Setophaga ruticilla. We report three records: a male at 
RM 131. 6L on 13 May 2000 (LLD, BTM, LKM), one at Lonely Dell, Lee’s Ferry, on 
3 June 1986 (CTL), and one at RM 11.3R on 13 October 2000 (JDG). There were 
four previous records, two from the river corridor, one in spring, one in fall, and two 
fall records from the rim. This is a sparse but regular transient throughout Arizona. 

PROTHONOTARY WARBLER Protonotaria citrea. We report the first record of 
this species from the Grand Canyon region. One was observed at RM -14. 3R on 26 
May 1994 (JDG). This species is a sparse transient throughout Arizona. 

COMMON YELLOWTHROAT Geothlypis trichas. We report the first winter 
record of this species from the Grand Canyon region. An adult male was at RM 50. 0L 
on 10 January and 14 February 1999 (CTL, RKR, PGF). The species is a fairly 
common winter resident in the LCRV. 

PAINTED REDSTART Myioborus pictus. Brown et al. (1984) reported three 
previous records from the Grand Canyon region. We report one at RM 31.5R on 4 
April 1999 (reported to CTL by CBN, BJB, TD). These are the only reports of this 
species from northern Arizona. Our record corresponds to the unusual occurrence of 
this species at nearly 20 lowland sites in Arizona in the spring of 1999 (Rosenberg and 
Benesh 1999). 

SUMMER TANAGER Piranga rubra. This species has recently colonized and 
become a rare summer resident in lowland riparian areas in the Grand Canyon 
(Brown et al. 1984). Sogge et al. (1998) reported singing males as far upstream as 
RM46.7 but no evidence of nesting. We accumulated four sightings that seem to 
indicate that this expansion is continuing upstream into the Glen Canyon reach: a 
subadult male at RM -13. 4L on 22 May 1998 (CTL), a subadult male at RM -3. OR on 
25 May 1999 (CTL, LLD), one at RM -6.1R on 21 June 1995 (JDG), and an adult 
male at RM -8.6R on 21 July 1993 (JRS). Although this species has recently 


114 


RECENT BIRD RECORDS FROM THE GRAND CANYON REGION, 1974-2000 


increased in the Grand Canyon region, it has decreased dramatically in the LCRV. 

CHIPPING SPARROW Spizella passerina. We report four winter records: two at 
RM 209. 0L on 21 January 1998 (JRS, CTL), two on 25 February 1998 (NLB, RKR), 
and one there on 18 January 1999 (CTL, NLB), and two at RM 196. OR on 23 
February 1998 (RKR, CTL}. There were two previous winter records from the river 
corridor. The species is rare in northern Arizona in winter. 

CLAY-COLORED SPARROW Spizella pallida. We report the first record of this 
species from the Grand Canyon region. One was observed at RM 1.0R (Paria Beach) 
on 2 May 1995 (JDG). 

BREWER’S SPARROW Spizella breweri. We report the first winter records of this 
species from the Grand Canyon region. One was at RM 5.2R on 12 January 1998 
(CTL, NLB), 25 were at RM -3. OR on 13 February 1998 (CTL, CEG), and two were 
at RM 56. 1R on 19 February 1998 (RKR, CAL, NLB). This species is generally 
sparse in winter in northern Arizona, and our records may be a result of favorable 
weather and/or food conditions. 

BLACK-THROATED SPARROW Amphispiza bilineata . This species is known as 
a common summer resident in desert habitats throughout northern Arizona with 
rather abrupt arrival and departure from the region in mid-March and early September 
(Phillips et al. 1964). Although it is not known as a winter resident in this area (Brown 
et al. 1984, Sogge et al. 1998), we accumulated 11 records of more than 20 
individuals from desert scrub adjacent to the Colorado River from October through 
February. From these records, it is apparent that this species winters in small numbers 
in the lowlands along the Colorado River in the Grand Canyon region. 

GRASSHOPPER SPARROW Am mod ram us sauannarum. One was seen at 
RM 1.0R on 27 and 31 October 1996 (CTL). This observation is the first report 
from the river corridor and the third from the Grand Canyon region. This species is 
a sparse migrant in northern Arizona (CTL unpubl. data) and more common 
elsewhere in the state. 

FOX SPARROW Passerella iliaca. “Slate-colored” individuals were observed at 
Lee’s Ferry on 27 October 1996 (CTL), RM -7.0L on 3 February 2000 (CTL, JRS), 
RM -8.8L on 18 March 1996 (JRS), and Spring Canyon (RM 204. 4R) on 11 April 
1999 (CTL, JAH). There were two previous records from the river corridor. This 
species is a rare transient and winter resident throughout Arizona. 

LINCOLN’S SPARROW Melospiza lincolrtii. Both Brown et al. (1987) and Sogge 
et al. (1998) speculated that this species may winter in the Grand Canyon region. 
During winter survey trips along the Colorado River in January and February, we 
recorded 60 in 1998 and 16 in 1999 (CTL, JRS, NLB, RKR, LLD). These records 
indicate that this species winters in the riparian vegetation along the Colorado River 
in the Grand Canyon region with high year-to-year variability. It is common in winter 
in southern Arizona. 

SWAMP SPARROW Melospiza georgiana. The two previous records were of one 
in October (Brown et al. 1984) and one in April (Sogge et al. 1998). We report six 
records, the first winter records from the Grand Canyon region. Individuals were 
observed at RM -8.7L on 4 February 2000 (CTL), RM 1.6R on 17 February 1998 
(CTL. RKR) and 30 November 1999-23 February 2000 (CTL), Phantom Ranch 
(RM 87.8R) on 20 February 1998 (RKR), RM 204.3R on 24 February 1998 (RKR), 
and RM 209. OR on 25 February 1998 (RKR, NLB). This species is a rare winter 
resident in northern Arizona. 

WHITE-THROATED SPARROW Zonotrichia albicollis. We report seven records. 
Individuals were observed at RM -14.3R on 29 October 1998 (CTL), at Lee’s Ferry 


115 


RECENT BIRD RECORDS FROM THE GRAND CANYON REGION, 1974-2000 


from 5 December 1994 to 7 March 1995 (JDG) and on 5 December 1995 (JDG, 
CTL), at RM 87. 8R on 15 January 1998 (JRS), and at RM -0.5L on 27 March 2000 
(CTL). There were four previous records from the Grand Canyon region, two from the 
Colorado River corridor. The White-throated Sparrow is an uncommon winter 
resident throughout Arizona but is considered rare in the LCRV. 

GOLDEN-CROWNED SPARROW Zonotrichia atricapilla. Individuals were ob- 
served at RM -8. 8R on 8 January 1998 (CTL, JRS), upper Deer Creek on 13 January 
1999 (CTL), RM 136.3R on 2 April 1997 (CTL, JRS, AF), and RM 53.0R on 5 April 
1999 (CTL, JAH). There were four previous records from the Grand Canyon region. 
This species is an uncommon winter visitor throughout Arizona, although less 
numerous in the northern portions. 

BOBOLINK Dolichonyx oryzivorus. The first record of this species from the 
Grand Canyon region is of one at Lonely Del!, Lee’s Ferry, on 15 September 1994 
(JDG). The Bobolink is a sparse, primarily fall, migrant throughout Arizona. 

COMMON GRACKLE Quiscalus quiscula. We report the first records of this 
species from the Grand Canyon region. One was observed with a female Great-tailed 
Grackle (Q. mexicanus) at RM 198. OR on 9 and 10 June 1997 (CTL, NLB, MDY), 
another was in the Lee’s Ferry area from 28 May to 31 July 1997 (CTL, CEG, DGS), 
with two present there on 29 and 31 July 1997 (CTL, CEG, DGS), and another was 
at RM 217.8 on 16 November 1993 (LES). This species is a recent sparse transient 
to Arizona (LaRue and Ellis 1992). 

STREAK-BACKED ORIOLE Icterus pustulatus. One was seen and heard at 
Spencer Canyon, RM 246. 0L, upper Lake Mead, on 22 January 1998 (JRS, CTL, 
NLB, KME). This represents the first record of this Mexican oriole from the Grand 
Canyon region. This species is known for its rare breeding (Corman and Monson 
1995) and rare but regular winter northward dispersal from Mexico into southern 
Arizona (Monson and Phillips 1981, Rosenberg and Witzeman 1999). This is the 
northernmost occurrence in Arizona. 

RED CROSSBILL Loxia curuirostra. One was seen at RM 51. 5L on 13 January 
1998 (CTL, JRS). There were no previous records from the river corridor. This 
species is a rare irregular transient at lowland sites throughout Arizona. 

ACKNOWLEDGMENTS 

The following is the list of observers with initials cited in text: LA, Lawrence Abbott; 
KB, Karen Barnett; JB, Jeffery Bennett; CAB, Catherine A. Bland; BRB, Ben R. 
Bobowski;-KSB, Katie S. Bobowski; WB, W 7 illiam Boecklen; BJB, Byron J. Boyle; 
RGB, Robert G. Bramblett; CB, Christopher Brod; BTB, Bryan T. Brown; NLB, 
Nikolle L. Brown; KJB, Kelly J. Burke; LSB, Lori S. Bush; SWC, Steven W. 
Carothers; MC, Matt Clifford; JC, Jerry Cox; TD, Tim Dale; JPD, John P. DeLong; 
BHD, Brian H. Dierker; LLD, Lara L. Dickson; AMD, A. Murphy Doty; KME, Kristin 
M. Enos; TF, Thomas Fergason; AF, Aaron Flesch; PGF, Peter G. Friederici; SRG, 
Steve R. Ganley; CG, Chris Geanious; CEG, Christine E. Goetze; JDG, John D. 
Grahame; KDG, Kathy D Groschupf; KG and DG, Kenton and Diane Grua; TMH, 
Tom M. Haberle; DH, Dan Hall; SH, Sharon Hester; JEH, John E. Hildebandt; JAH, 
Jennifer A. Holmes; SMJ, Sarah M. James; MJK, Michael J. Kearsley; JVJ, Jessica 
V. Jewell; NCK, Natasha C. Kline; ARL, Anne R. LaRue; CTL, Charles T. LaRue; 
EFL, Elaine F. Leslie; CAL, Casey A. Lott; BTM, Brian T. Meiering; LKM, Lisa K. 
Miller; VM, Velma McMeekin; KGM, Ken G. McMullen; CBN, Clay B. Nelson; SEO, 
Sue E. Ordway; NEP, Nicolas E. Pappani; AP, Alan Peterson; ACP, A. Clive Pinnock; 
RKR, Roger K. Radd; BKR, Brenda K. Russell; PFR, Patrick F. Ryan; DS, Drifter 
Smith; DGS, Dwight G. Smith; JRS, John R. Spence; MKS, Mark K. Sogge; LES, 


116 


RECENT BIRD RECORDS FROM THE GRAND CANYON REGION, 1974-2000 


Lawrence E. Stevens; GMS, Gary M. Stolz; DT, DrewThate; TJT, Timothy J. Tibbitts; 
TV, Tom Vail; PW, Peter Weiss; MDY, Mike D. Yard. 

We acknowledge Clive Pinnock and John Grahame for their years of field work at 
Glen Canyon National Recreation Area. We are grateful to the following individuals 
for their assistance in the field: Tony Anderson, Ken Baker, Jeff Behan, Jeff Bennett, 
Steve Bledsoe, Lewis Boobar, Kelly Burke, Matt Clifford, Sarah Davidson, Brian 
Dierker, Murphy Doty, Kristin Enos, Peter Friederici, Chris Geanious, Jimmy Hall, 
Tom Haberle, Jennifer Holmes, Casey Lott, Ken McMullan, Lars Niemi, Nels Niemi, 
Roger Radd, Brenda Russell, Dwight Smith. Tom Swartz, Tim Tibbitts, Peter Weiss, 
and Mike Yard. We thank Elaine Leslie, Pat Ryan, and Tim Tibbitts for providing 
records. Special thanks to John Blaustein for providing the Sanderling photograph 
and to Chris Geanious for providing the Groove-billed Ani photograph. Portions of 
these projects were funded by Glen Canyon Environmental Studies, Grand Canyon 
Research and Monitoring Center, and National Park Service grants to Spence and 
Stevens. We thank Dan Gibson, Gale Monson, Gary Rosenberg, and Janet Witzeman 
for their thoughtful and thorough reviews. 

LITERATURE CITED 

Bailey, F. M. 1939. Among the Birds in the Grand Canyon Country. U.S. Natl. Park 
Service, Grand Canyon, AZ. 

Behle, W, H., and Higgins, H. G. 1959. Birds of Glen Canyon, in Ecological Studies 
of Flora and Fauna in Glen Canyon (A. M. Woodbury, ed.), pp. 107-133. 
Anthropol. Papers 40, Glen Canyon Ser. 7, Univ. Utah. 

Benesh, C. D., and Rosenberg, G. H. 1998. The winter season. Arizona region. Field 
Notes 52:234-237. 

Billingsley, G. H,, and Hampton, H. M. 1999. Physiographic rim of the Grand 
Canyon, Arizona. U.S. Geol. Surv. Open-File Rep. 99-30, scale 1:250,000. 

Blaustein, J. 1999. The Hidden Canyon: A River Journey. Chronicle Books, San 
Francisco. 

Brown, B.T. 1988. Breeding ecology of a Willow Flycatcher population in the Grand 
Canyon, Arizona. W. Birds 19:25-33. 

Brown, B. X, Carothers, S. W., and Johnson, R. R. 1983. Breeding range expansion 
of Bell’s Vireo in Grand Canyon, Arizona. Condor 85:499-500. 

Brown, B. T., Carothers, S. W., Haight, L. X, Johnson, R. R., and Riffey, M. M. 
1984. Birds of the Grand Canyon region: An annotated checklist, 2nd ed. Grand 
Canyon Nat. Hist. Assoc. Monogr. 1. 

Brown, B. X, Carothers, S. W., and Johnson, R. R. 1987. Grand Canyon Birds. 
Univ. of Ariz. Press, Tucson. 

Brown, B. T., Carothers, S. W, Johnson, R. R., Riffey, M. M., and Stevens, L. E. 
1993. Checklist of the Birds of the Grand Canyon Region. Grand Canyon Nat. 
Hist. Assoc., Grand Canyon, AZ. 

Brown, D. E. 1985. Arizona Wetlands and Waterfowl. Univ. of Ariz. Press, Tucson. 

Corman, T., and Monson, G. 1995. First United States nesting records of the Streak- 
backed Oriole. W. Birds 26:49-53. 

DeSante, D. F., and George, T. L. 1994. Population trends in the landbirds of western 
North America, in A century of avifaunal change in western North America (J. R. 
Jehl, Jr., and N. K. Johnson, eds.), pp. 173-190. Studies Avian Biol. 15. 

Hughes, J.M., 1999. Yeliow-biiled Cuckoo (Coccyzus americanus), in The Birds of 
North America (A. Poole and F, Gill, eds.), no. 418. Birds N. Am., Philadelphia. 


117 


RECENT BIRD RECORDS FROM THE GRAND CANYON REGION, 1974-2000 

Johnson, N. K. 1994. Pioneering and natural expansion of breeding distributions in 
western North American birds, in A century of avifaunal change in western 
North America (J. R. Jehl, Jr., and N. K. Johnson, eds.), pp. 27-44. Studies 
Avian Biol. 15. 

Johnson, R. R. 1991. Historic changes in vegetation along the Colorado River in the 
Grand Canyon, in Colorado River Ecology and Dam Management (G. R. 
Marzolf, ed.), pp. 178-206. Natl. Academy Press, Washington, D.C. 

LaRue, C. T., and Ellis, D. H. 1992. The Common Grackle in Arizona: First specimen 
record and notes on occurrence. W. Birds 23:84-86. 

Mlodinow, S. G., and Karlson, K.T. 1999. Anis in the United States and Canada. N. 
Am. Birds 53:237-245. 

Monson, G., and Phillips, A. R. 1981. Annotated Checklist of the Birds of Arizona, 
2nd ed. Univ. of Ariz. Press, Tucson. 

Phillips, A., Marshall, J., and Monson, G. 1964. The Birds of Arizona. Univ. of Ariz. 
Press, Tucson. 

Rosenberg, G. H., and Witzeman, J. L. 1998. Arizona Bird Committee report, 
1974-1996: Part 1 (nonpasserines). W. Birds 29:199-224. 

Rosenberg, G. H., and Witzeman, J. L. 1999. Arizona Bird Committee report, 
1974-1996: Part 2 (passerines). W. Birds 30:94-120. 

Rosenberg, G. H., and Benesh, C. D. 1999. Spring migration. Arizona region. N. 
Am. Birds 53:309-311. 

Rosenberg, K. V., Ohmart, R. D., Hunter, W. C., and Anderson, B. W. 1991. Birds 
of the Lower Colorado River Valley. Univ. of Ariz. Press, Tucson. 

Sogge, M. K., Tibbitts, T. J., and Petterson, J. R. 1997, Status and breeding ecology of 
the Southwestern Willow Flycatcher in the Grand Canyon. W. Birds 28:142-157. 

Sogge, M. K., Felley, D., and Wotawa, M. 1998. Annotated species list and summary 
in riparian bird community ecology in the Grand Canyon — Final report. U. S. 
Geol. Surv., Colo. Plateau Field Station, N. Ariz. Univ., Flagstaff, AZ 86011. 

Spence, J. R., LaRue, C. T., Muller, J. R., and Brown, N. L. 1998. 1997 avian 
community monitoring along the Colorado River from Lee’s Ferry to Lake Mead. 
Final report submitted to Grand Canyon Monitoring and Research Center, 
Flagstaff; order from Glen Canyon National Recreation Area, P. O. Box 1507, 
Page, AZ 86040. 

Stevens, L. 1983. The Colorado River in Grand Canyon: A Comprehensive Guide. 
Red Lake Books, Flagstaff, AZ. 

Stevens, L. E., Buck, K. A., Brown, B. T., and Kline, N, 1997. Dam and geomorphic 
influences on Colorado River waterbird distribution, Grand Canyon, Arizona. 
Regulated Rivers: Research & Management 13:151-169. 

Turner, R. M., and Karpiscak, M. M. 1980. Recent vegetational changes along the 
Colorado River between Glen Canyon Dam and Lake Mead, Arizona. U.S. Geol. 
Surv. Prof. Paper 1132. 

Woodbury, A. M., and Russell Jr., H. N. 1945. Birds of the Navajo Country. Bull. 
Univ. LItah 35 (41), Biol. Ser. 9 (1). 


Accepted 12 December 2000 


118 


OCCURRENCE PATTERNS OF PEREGRINE 
FALCONS ON SOUTHEAST FARALLON ISLAND, 
CALIFORNIA, BY SUBSPECIES, AGE, AND SEX 

SASHA EARNHEART-GOLD and PETER PYLE, Point Reyes Bird Observatory, 
4990 Shoreline Hwy., Stinson Beach, California 94970 


ABSTRACT: We summarize observations of 201 Peregrine Falcons ( Falco 
peregrinus ) at Southeast Farallon Island during the fall and winter from 1990 to 1999 
by age, sex, and subspecies. The northwestern subspecies F. p. pealei and the 
continental subspecies F. p. anatum occurred with roughly equal frequency. We 
recorded 10 individuals of the arctic subspecies F. p. tundrius. During fall, adults 
occurred significantly earlier than immatures. Males tended to occur earlier than 
females; anatum tended to occur earlier than pealei. Four to six birds per year (of both 
anatum and pealei) remained through the winter. Under the assumption that 
wintering individuals returned each year to the maximum extent possible allowed by 
the observed variation by age, sex, and subspecies, their survival rate was 0.78. 

Although the Peregrine Falcon (Falco peregrinus) was recently removed 
from the federal endangered-species list (Federal Register 64 FR 46541 
46558), there is still much interest in this species (e.g., Ratcliffe 1980, Cade 
et al. 1988). On Southeast Farallon Island, 42 km off San Francisco, 
California, numbers of Peregrines increased significantly during fall migration 
from 1974 to 1993 (Pyle and DeSante 1994); the negative (but nonsignifi- 
cant) curvilinear relationship found in that analysis suggested that the increase 
was leveling off by the early 1990s. This species has also become a regular 
winter resident at Southeast Farallon (Pyle and Henderson 1991). 

Currently, there are about 19 recognized subspecies of the Peregrine 
Falcon in the world (White and Boyce 1988). The AOU (1957) listed two 
subspecies in California, P. f. anatum, which breeds and winters throughout 
the state, and P. f. pealei, which is reported to winter “rarely to California.” 
Grinnell and Miller (1944) listed three specimens of pealei from California, 
as far south as San Diego County (Swarth 1933), but Beebe (1960) seemed 
to question most reports of this subspecies south of Washington, and Hunt 
et al. (1975) believed few if any Peregrines migrate down the Pacific coast. 
Anderson et al. (1988) listed two records of pealei banded as nestlings in 
British Columbia and recovered in San Diego and Santa Cruz; they sug- 
gested on the basis of sight reports that it is a regular visitor to Southeast 
Farallon and elsewhere in California. 

A third subspecies, P f. tundrius, breeding in arctic North America, was 
described by White (1968). He suggested that it is a rare or uncommon 
migrant along the west coast of North America and listed one specimen (of 
213 tundrius examined) from California (March, Del Ray, Fresno County) 
and four specimens from Baja California (March and April). Anderson et al. 
(1988) listed a second California specimen of a banded tundrius recovered 
on San Miguel Island, and Hamilton and Willick (1996) listed two records of 
banded birds from Orange County. Otherwise, little has been published 
about the occurrence of pealei and tundrius in California. 

During the past ten years Point Reyes Bird Observatory biologists have 
made an effort to identify Peregrines on Southeast Farallon Island by 


Western Birds 32:119-126, 2001 


119 


OCCURRENCE PATTERNS OF PEREGRINE FALCONS ON FARALLON ISLAND 


subspecies, age, and sex. Here we summarize our findings from 1990 to 
1999 for both fall and winter, when a majority of migrant and resident 
individuals were recorded. Arrivals from spring and summer are excluded 
because sample sizes are much smaller (Pyle and Henderson 1991) and 
critical subspecies, age, and sex determinations were not recorded. 

METHODS 

The standardized daily census for migrant and wintering birds on South- 
east Farallon, uninterrupted since 1968, has been described by DeSante and 
Ainley (1980) and Pyle and Henderson (1991). Birds censused each day 
during the fall/winter period (15 July-1 March) were identified as arrivals if 
banding data and/or plumage observation, coupled with data from previous 
days, suggested that they had arrived at the island that day. Between 1990 
and 1999 the size, shape, and plumage of most arriving and winter resident 
Peregrine Falcons were examined critically for subspecies, age, and sex. We 
defined as winter residents those remaining for at least 21 days between 15 
December and 1 March. 

All subspecific, age, and sex determinations of Peregrines were made 
either by Pyle (78.6% of 126 identifications) or long-term interns (including 
Earnheart-Gold) that had been trained for this by Pyle. Plumage criteria used 
in these determinations included those presented by Brooks (1926), Beebe 
(1960), Brown and Amadon (1968), White (1968), and White and Boyce 
(1988), supplemented by examination of specimens at the California 
Academy of Sciences (CAS), San Francisco, and Museum of Vertebrate 
Zoology (MVZ), Berkeley. Determinations were usually made in the order 
age, subspecies, sex. During most (91.2%) determinations there were one to 
five other Peregrines at Southeast Farallon on the same day, often interact- 
ing with each other, allowing direct comparisons of size and plumage. 
Peregrines for which determinations were not made were not observed 
adequately or were difficult to categorize by known criteria. There is much 
variation within each subspecies, leading to overlap in size and/or plumage, 
especially between anatum and each of the other two subspecies. Because 
of this intrasubspecific variation we identified only typical examples and left 
many birds unidentified to subspecies. The following criteria were used to 
identify subspecies: 

F. p. anatum : Medium sized; stocky build. Adults with medium-dark gray upper- 
parts, contrastingly black hood extending to the bill, and buffy- or rosy-tinged 
underparts with medium-heavy spotting. Immatures medium brownish, sometimes 
tinged rufous, with little or no buffy in the crown, a medium-thick malar stripe, and a 
thin buffy terminal band on the tail. 

F. p. pealei: Large; medium-stocky build. Adults with medium-dark grayish upper- 
parts blending to a slightly darker crown and hood, often interrupted by a white band 
at the base of the bill, and whitish underparts with heavy barring throughout the breast 
and belly. Immatures dark gray to blackish without buffy in the crown, a thick malar 
stripe, and a thin or no buffy terminal band on the tail. 

F. p. tundrius: Medium-small; slender and long-winged in build. Adults with 
medium-pale grayish to bluish upperparts contrasting somewhat with a blackish 
crown and hood, interrupted by a white band at the base of the bill, and white 


120 


OCCURRENCE PATTERNS OF PEREGRINE FALCONS ON FARALLON ISLAND 


underparts with light barring on the belly but little or no barring on the upper breast. 
Immatures brownish with extensive buffy in the crown and nape, a reduced malar 
stripe, and an extensive buffy terminal band on the tail. 

We used simple and multiple linear regression and analysis of variance 
(ANOVA) to help describe the occurrence patterns and look for significant 
variation in these patterns among subspecies, ages, and sexes. 

RESULTS 

We counted 201 Peregrine Falcons during fall and winter, 1990-1999 
(Table 1); 126 birds identified to subspecies included 56 of anatum, 57 of 
pealei , and 10 of tundrius. Of 87 birds that were sexed, 47 were females 
and 40 were males, and of 121 birds that were aged, 50 were adults and 71 
were immatures. There was no significant linear trend in number of arrivals 
during the 10-year period taken as a whole (Figure 1; linear regression; t = 
0.363, P = 0.719). Evidently, however, fewer Peregrines arrived immedi- 
ately after El Nino years (1992 and 1997-1998), followed by gradual 
increases in arrivals until the next El Nino (Figure 1). For example, the 
positive linear trend in arrivals between 1993 and 1998 was significant ( t = 
7.86, P = 0.001). The 10-year trend for identified anatum was slightly and 
nonsignificantly positive ( t = 0.147, P = 0.887), and that for identified 
pealei was positive and nearly significant ( t = 2.193, P = 0.060). It should 
be noted that biases related to the 75 Peregrines (37.3%) not identified to 
subspecies could affect these analyses. 

All ten individuals of tundrius noted during the study period were 
observed and identified by Pyle. They were recorded in only five of the ten 


Table 1 Arrival Dates of Peregrine Falcons on Southeast 
Farallon Island during Fall and Winter (15 July-1 March), 
1990-1999 


Subspecies/Age/Sex n 

Mean 

Minimum 

Maximum 

Total 

201 

14 Oct 

22 Jul 

31 Dec 

F. p. anatum 

56 

14 Oct 

6 Aug 

1 Dec 

Male 

18 

30 Sep 

6 Aug 

23 Nov 

Female 

18 

22 Oct 

13 Aug 

27 Nov 

Adult 

26 

3 Oct 

6 Aug 

1 Dec 

Immature 

28 

24 Oct 

13 Aug 

27 Nov 

F. p. pealei 

57 

20 Oct 

26 Aug 

3 Dec 

Male 

18 

24 Oct 

7 Sep 

24 Nov 

Female 

26 

20 Oct 

26 Aug 

3 Dec 

Adult 

22 

13 Oct 

13 Sep 

13 Nov 

Immature 

34 

26 Oct 

7 Sep 

3 Dec 

F. p. tundrius 

10 

19 Oct 

2 Oct 

17 Nov 

Male 

3 

11 Oct 

7 Oct 

14 Oct 

Female 

3 

14 Oct 

7 Oct 

22 Oct 

Adult 

2 

11 Oct 

7 Oct 

14 Oct 

Immature 

8 

21 Oct 

2 Oct 

17 Nov 


121 


OCCURRENCE PATTERNS OF PEREGRINE FALCONS ON FARALLON ISLAND 


^ 30 n 



Year 


Figure 1. Numbers of Peregrine Falcon arriving (see text) on Southeast Farallon 
Island, by year, 1990-1999. 


years, 1992 (two birds), 1993 (one), 1994 (one), 1995 (two), and 1997 
(four), despite a similar amount of observation effort each year. Both adults 
(see Table 1) were males, recorded on single days in 1997. Seven of the ten 
birds were recorded between 7 and 22 October. There are also at least six 
records of this subspecies (all immatures) observed and described or photo- 
graphed (e.g., Figure 2) by Pyle or other biologists between 26 September 
and 6 November, 1981-1988. All birds identified as tundrius fit the 
description above, and most were seen in direct comparison with other 
Peregrines. 

Examination of mean arrival dates (Table 1) reveals several patterns 
according to subspecies, age, and sex. In all three subspecies identified males 
apparently tended to arrive before females, but this pattern was not 
signifcant for the species as a whole (F<i,99) = 1 .50, P = 0.299; F(3,83) = 1.62, 
P = 0.207, adjusting for subspecies) or for pealei (P(i,42) = 0.38, P = 0.543). 
It was almost significant for anatum (F( i,34) = 3.98, P = 0.054). In all three 
subspecies identified adults arrived before immatures, and this pattern was 
significant for the species as a whole (Fu,i54) = 11.32, P = 0.001; F(3,ii8) = 
15.14, P < 0.001, adjusting for subspecies), for anatum [F( 1 , 52 ) = 6.87, P = 
0.012), and for pealei (Fa, 54) = 7.32, P = 0.009). The mean arrival date of 
anatum was not significantly earlier than that of pealei (Fa.nu = 1 .29, P = 
0.259) when all age/sex groups were combined. When the ages and sexes 
were separated, this comparison was significant for males (Fa, 34) = 6.53, P 
= 0.015) but not for females, immatures, or adults (F< 1.72, P> 0.197). As 
in analyses of subspecies, it should be noted that biases among birds not 
identified to age (n = 80; 39.8% of the sample) or sex (n = 114; 56.7% of 


122 


OCCURRENCE PATTERNS OF PEREGRINE FALCONS ON FARALLON ISLAND 



Figure 2. Immature Peregrine Falcon of the tundra subspecies F. p. tundrius, Southeast 
Farallon Island, 6 November 1987. 

Photo by Scot Anderson 


the sample) might affect these results. The sample of tundrius was too small 
for these analyses to be meaningful. 

The winter resident population consisted of four to six individuals, up to 
three of anatum and three of pealei of various ages and sexes (Table 2). 
Under the (perhaps tenuous) assumption that winter residents return in 
consecutive years, the table implies a minimum of 13 wintering individuals 


Table 2 Winter Resident Peregrine Falcons on Southeast Farallon Island, 
1990-1999 


Year Summary by subspecies/age/sex 


1990- 91 

1991- 92 

1992- 93 

1993- 94 

1994- 95 

1995- 96 

1996- 97 

1997- 98 

1998- 99 

1999- 00 


Four adults; subspecies unknown 

anatum: 2 adults, 1 immature; pealei: 2 adults 

anatum: 2 adults; pealei: 1 adult Cf, 1 adult 9 

anatum ; 1 adult <f , 1 adult 9; pealei: 1 adult <f, 1 adult 9 

anatum: 1 adult C f, 1 immature C f; pealei: 1 adult (f, 2 immature 9 

anatum: 1 adult cf, 1 adult 9, 1 immature 9; pealei: 1 adult cf, 

1 adult 9, 1 immature 9 

anatum: 1 adult C f, 1 adult 9; pealei: 1 adult 9, 1 immature 

anatum: 1 adult C f, 1 adult 9; pealei: 1 adult cf, 1 adult 9 

anatum: 1 adult cf, 1 immature cf, 1 immature 9; pealei: 1 adult cf, 
1 immature 

anatum: 1 adult <$ , 1 adult 9; pealei: 1 adult 9, 1 immature cf, 

1 immature 9 


123 


OCCURRENCE PATTERNS OF PEREGRINE FALCONS ON FARALLON ISLAND 

over the 10-year period, five of ana turn (two males and three females) and 
eight of pealei (three males and five females). If returning winter resident 
adults of the same sex represented the same individuals, six adults (two of 
anatum and four of pealei ) disappeared and 21 adults returned for an 
annual adult survival rate of 0.78 (0.86 for anatum and 0.69 for pealei) over 
the 10-year period. 

DISCUSSION 

Our finding no linear trend in Peregrine numbers between 1990 and 
1999 suggests that populations have stabilized since recovery from their 
pesticide-related reductions from the 1950s to the 1970s and that our 
proportions may represent historical distributions in population size and 
range. It is also possible, however, that distributions have shifted since the 
population bottlenecks of the 1950s-1970s, especially if seabird popula- 
tions or other prey resources have changed. Our data further suggest that 
mortality of Peregrines is higher and/or fall and winter populations shift 
during El Nino periods (e.g., 1992 and 1997-1998), perhaps in response to 
reduced abundance of seabird prey off central California during these events 
(Ainley and Boekelheide 1990). 

The regularity of pealei in California (similar in abundance to anatum at 
Southeast Farallon) has not been previously appreciated in the literature 
(e.g., AOU 1957, Beebe 1960, Hunt et al. 1975), although suspected by 
Anderson et al. (1988) and confirmed by museum specimen examination 
(C. White pers. comm.) and unpublished data collected by the Santa Cruz 
Predatory Bird Research Group (B. Walton pers. comm.). It is possible that 
the abundant alcid populations at Southeast Farallon (Ainley and Boekelheide 
1990) provide a food resource for pealei, an alcid specialist where it breeds 
(Beebe 1960). Because the colony at Southeast Farallon represents the 
southernmost point of such alcid abundance (Carter et al. 1992), the island 
may represent the southern limit of the regular winter range of pealei. 
Individuals of pealei, though, winter or wander as far south as San Diego and 
even Baja California but are greatly outnumbered by anatum south of 
Southeast Farallon (B. Walton pers. comm.). 

Our data suggest that tundrius is an uncommon but regular transient 
down the California coast from late September through mid November, as 
proposed by White (1968) and Anderson et al. (1988) and confirmed with 
unpublished data from the Santa Cruz Predatory Bird Research Group (B. 
Walton pers. comm.). It is possible that these birds winter in Baja California 
(White 1968), although residents of anatum of that region, particularly 
immatures, may resemble tundrius in size and plumage more closely than 
those of California-breeding populations (C. White, M. A. Patten, pers. 
comm.; Pyle, specimen examination at CAS and MVZ). We suggest that the 
identification of the four specimens of tundrius from Baja California listed 
by White (1968) be reconfirmed. 

Our data on arrival patterns of anatum and pealei by age and sex are 
consistent with patterns widely known in birds: smaller males depart nesting 
grounds for winter areas slightly before larger females, and adults migrate 
and arrive on winter grounds well before immatures. The more local and 


124 


OCCURRENCE PATTERNS OF PEREGRINE FALCONS ON FARALLON ISLAND 


southern anatum arriving before the more northern pealei, especially 
among males, is also to be expected (Anderson et al. 1988). 

ACKNOWLEDGMENTS 

We thank the U.S. Fish and Wildlife Service, managers of the Farallon National 
Wildlife Refuge, for supporting our work at Southeast Farallon Island, Douglas J. Long 
at CAS and Ned K. Johnson at MVZ for permitting our examination of Peregrine 
specimens, and Michael Patten for information on Peregrines in California and Baja 
California. We especially thank Clayton White and Brian Walton for expert advice 
during reviews of the manuscript and Walton further for sharing unpublished data 
collected by the Santa Cruz Predatory Bird Research Group. This is Point Reyes Bird 
Observatory contribution 930. 

LITERATURE CITED 

Ainley, D. G., and Boekelheide, R. J. 1990. Seabirds of the Farallon Islands: Ecology, 
Dynamics, and Structure of an Upwelling-system Community. Stanford Univ. 
Press, Stanford, CA. 

American Ornithologists' Union. 1957. Check-list of North American Birds, 5th ed. 
Am. Ornithol. Union, Baltimore. 

Anderson, C. M., Roseneau, D. G., Walton, B. J., and Bente, P. J. 1988. New 
evidence of a Peregrine migration on the west coast of North America, in 
Peregrine Falcon Populations: Their Management and Recovery (T. J. Cade, J. 
H. Enderson, C. G. Thelander, and C. M. White, eds.), pp. 507-516. Braun- 
Brumfield, San Francisco. 

Beebe, F. L. 1960, The marine Peregrines of the northwest Pacific coast. Condor 
62:145-189. 

Brooks, A. 1926. Notes on the status of the Peale Falcon. Condor 28:77-79. 

Brown, L., and D. Amadon. 1968. Eagles, Hawks and Falcons of the World. 
McGraw-Hill, New York. 

Cade, T. J., Enderson, J. H., Thelander, C. G., and White, C. M. (eds.). 1988. 
Peregrine Falcon Populations: Their Management and Recovery. Braun- 
Brumfield, San Francisco. 

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 on the northern and central California coasts in 1989-1991. U.S. 
Fish & Wildlife Serv., 6924 Tremont Rd., Dixon, CA 95620. 

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. 

Hamilton, R. A., and Willick, D. R. 1996. The Birds of Orange County, California: 
Status and Distribution. Sage & Sea Press, Irvine, CA. 

Hunt, W. G., Rogers, R. R., and Slowe, D. G. 1975. Migratory and foraging behavior 
of Peregrine Falcons on the Texas coast. Can. Field-Nat. 89:111-123. 

Pyle, P., and Henderson, R. P. 1991. The birds of Southeast Farallon Island: 
Occurrence and seasonal distribution of migratory species. W. Birds 22:41-84. 

Pyle, P., and DeSante, D. F. 1994. Trends in waterbirds and raptors at Southeast 
Farallon Island, California, 1974—1993. Bird Populations 2:33-43. 


125 


OCCURRENCE PATTERNS OF PEREGRINE FALCONS ON FARALLON ISLAND 


Ratcliffe, D. 1980. The Peregrine Falcon. Poyser, Stafford, England. 

Swarth, H. S. 1933. Peale Falcon in California. Condor 35:233-234. 

White. C. M. 1968. Diagnosis and relationships of the North American tundra- 
inhabiting Peregrine Falcons. Auk 85:179-191. 

White, C. M., and Boyce, D. A., Jr. 1988. An overview of Peregrine Falcon 
subspecies, in Peregrine Falcon Populations: Their Management and Recovery 
(T. J. Cade, J. H. Enderson, C. G. Thelander, and C. M. White, eds.), pp. 789- 
810. Braun-Brumfield, San Francisco. 


Accepted 14 May 2001 


NOTES 

LOW-ELEVATION NESTING BY CALLIOPE 
HUMMINGBIRDS IN THE WESTERN 
SIERRA NEVADA FOOTHILLS 

BRIAN D. C. WILLIAMS, 8200 Turner Dr., Granite Bay, California 95746 


The Calliope Hummingbird breeds uncommonly to fairly commonly in the Sierra 
Nevada and other high mountain ranges in California. In the Sierra it typically nests 
above 4000 feet elevation (Grinnell and Miller 1944, Gaines 1992, pers. obs.), 
usually near moist meadows or other relatively level and wet sites with a mixture of 
deciduous and coniferous trees, shrubs, and flowering plants (Grinnell and Miller 
1944, Verner et al. 1980, Gaines 1992, pers. obs.). Although it can be locally fairly 
common as a migrant in the more arid lowlands during spring migration (pers. obs.), 
it generally avoids dense forests, dry ridges, or other relatively exposed and hot sites 
for breeding. Consequently, 1 was quite surprised to find a Calliope Hummingbird 
raising young in relatively arid Blue Oak (Quercus douglasii) woodland well below 
1000 feet elevation. 

On 6 June 1993, I found a female Calliope Hummingbird feeding a dependent 
fledgling in a California Buckeye ( Aesculus californica ) near Granite Bay in Folsom 
Lake State Park, Placer County, elevation 420 feet, at the western edge of the Sierra 
Nevada foothills. Both birds were clearly smaller than Anna’s Hummingbird ( Ca\ypte 
anna), a common resident in the area, and they were also smaller but chunkier than 
the slender-necked Black-chinned Hummingbird ( Archilochus alexandri), also fairly 
common there. Both Calliopes had buffy flanks with faint buffy breast bands (see 
Kaufmann 1990). They both had a pattern of symmetrical columns of dark throat 
spots, not concentrated in the center of the throat and lacking apparent iridescence. 
The female Calliope’s primaries extended just slightly beyond her short tail, and the 
four central retrices lacked white tips. I distinguished the juvenile mostly by behavior 
as it made only short uncoordinated flights of no more than a few feet and often had 
difficulty perching, though its plumage also seemed buffier plumage than the adult 
female’s (Baltosser 1994). The juvenile usually remained stationary in the buckeye 
until the female returned to feed it, inserting her bill into its gape. To make sure of the 
identification and to document this rare event, I called Bill Grenfell, a local wildlife 
photographer, who took a few photographs on the afternoon of 6 June (Figure 1). 
The birds were quite approachable, and we were able to stand closer than 12 feet 
without flushing either of them. They also appeared to ignore other birds in the area, 
except for a Black-chinned Hummingbird (unknown sex) that was chased from the 
area by the female Calliope on my single return visit about 10:00 on 7 June. On that 
visit I returned to the same tree and immediately found the juvenile still perched there; 
presumably the nest was nearby. The female was still making repeated foraging trips 
and returned regularly to feed the juvenile. 

The Calliope Hummingbird is a regular spring migrant at Folsom Lake State Park, 
and two to five adult males can usually be found daily in the Granite Bay-Beek’s Bight 
area from mid-April to early May. They concentrate in an area of mixed oak woodland 
and patchy Chamise (Adenostoma fasciculatum ) chaparral where sticky monkey- 
flowers ( Mimulus aurantiacus ) and buckeyes are numerous and Indian paintbrush 
{Castilleja sp.), yerba santa ( Eriodictyon californicum), and other flowering plants 
are scattered about in the chaparral. This female and her fledgling, however, were 
about 0.5 mile southwest of that area within a woodland dominated by the Blue Oak 
and situated on a gentle north-facing slope with a sparse understory of Poison Oak 


Western Birds 32:127-130, 2001 


127 


NOTES 



Figure 1. Female Calliope Hummingbird feeding fledgling in California Buckeye, 
Folsom Lake State Park, 6 June 1993. 


Photo by Bill Grenfell 


(' Toxicodendron diuersilobum) and flowering California buckeye. The woodland s 
canopy closure is approximately 80% (visually estimated), though the Blue Oak 
canopy is generally high, not very dense, and admits plenty of filtered light for a 
ground cover of mostly nonnative grasses and forbs. Otherwise there are no nearby 
concentrations of typical hummingbird-pollinated plants within at least 100 m. The 
site is also at the edge of an almost level drainage that stays moist relatively late into 
the spring and probably still contained some open water at the time of nest initiation. 
The north aspect, high canopy cover, relatively mesic conditions, and position at the 
bottom of a local basin (into which cooler air collects) likely give this spot one of the 
coolest microclimates within roughly a half-mile radius. The Calliope Hummingbird 
generally prefers cool microhabitats (Calder and Calder 1994). Gaines (1992), 
however, reported the species using arid ridges in Yosemite, and in 1978 Ted Beedy 
(pers. comm.) found two nests in junipers on a dry slope about 2 miles south of 
Highway 20 and >100 m from Yosemite Creek (>6000 feet elevation). 

Surprisingly. I later learned of a previous breeding attempt by a Calliope Humming- 
bird about 0.5 mile from the site, described above. On 20 April 1985 Tim Fitzer, Jack 
Wilburn, and Dan Brown saw a female Calliope fly from a nest. Excited by their 
discovery, Wilburn and Brown returned two days later to photograph the nesting 
Calliope (Figure 2). On a later visit Wilburn did not see the female and assumed the 
nest was abandoned. It is not known whether the bird laid any eggs or fledged young. 
The date of probable nest initiation in 1985 was very near that in 1993. If the juvenile 
fledged on 6 June 1993 (the last possible date) and incubation (15-16 days) and 
fledging (18-21 days) took 33-37 days (Calder and Calder 1994), the last possible 
dates for egg laying were 1-5 May. 

The only other probable or possible extralimital nesting records I was able to locate 


128 



NOTES 



Figure 2. Calliope Hummingbird on nest at Beek’s Bight, Folsom Lake State Park, 22 
April 1985. 

Photo by Dan Brown 


were by San Miguel (1985) and in the editors’ files for the Middle Pacific Coast region 
of American Birds and its various successors (1955-1991; 1997-1998). I subse- 
quently confirmed all of these records with the observers. Bob Yutzy found one at 
Lake Shasta, Shasta Co., on 21 June 1980. Richard A. Erickson found a pair 
copulating at 2000 feet elevation 1 mile north of Hyampom, Trinity Co., 16 June 
1983 (Le Valley and Evens 1983, Harris 1991). Jeri M. Langham found a male along 
French Hill Road about 1000 feet elevation near Shingle Springs, El Dorado Co., 8 
June 1985. David G. Yee found a female apparently on territory at 2000 feet elevaton 
near Mokelumne Hill, Calaveras Co., 1 June 1985. Yutzy reported a pair visiting a 
feeder at 1300 feet elevation west of Redding in mid-July 1985, and a female at a 
feeder in Redding 11 June 1987; he also saw a female on a nest at about 2000 feet 
elevation along Gilrnan Road northeast of Lake Shasta on 31 May 1997 and has seen 
the species a few other times in late May sightings at his feeders near Shasta (> 1 000 
feet) west of Redding (pers. comm.). Even though the various observers considered 
these records either out of range or at the margins of the species’ range, all of these 
localities are in or near coniferous forest (Ponderosa Pine, Pinus ponderosa, or 
Douglas Fir, Pseudotsuga menziesii; the Gilman Road nest was in a Douglas Fir), 
nesting habitat more typical Blue Oak woodland. 

Records accumulated over the years suggest the American River may serve as a 
migratory corridor, hummingbirds roughly following the progression of blooming 
flowers to more typical nest sites at higher elevations (mostly above 5000 feet in this 
part of the Sierra). This phenomenon was hypothesized by San Miguel (1985) for the 
Kaweah River of the southern Sierra and may be a common strategy of migrating 
hummingbirds. Possibly the nestings at Folsom Lake were responses to locally 
favorable food supplies (e.g., Sealy 1979) and/or suitable microclimate and nesting 
sites (e.g., Walley 1977, Roberson 1993), but the role of premature reproductive 


129 


NOTES 


development (in this and other extralimital nesting species) is unknown. Either way, I 
urge other observers to watch for signs of nesting Calliope Hummingbirds at other 
favorable sites outside of the recognized breeding range, especially in the Sierra 
Nevada foothills where other montane birds may nest locally well outside of their 
known range. Such nesting also raises the possibility of multiple broods (see also the 
hint by Calder and Calder 1994), as it seems possible for a Calliope Hummingbird 
nesting at a low elevation early to nest at a higher elevation later in the year, in June 
and July as is typical for the species (Orr and Moffitt 1971). 

Thanks to Bill Grenfell for quickly responding to my plea for photos, to David Yee, 
Don Roberson, and Steve Glover for providing copies of the editors’ files for 
American Birds and its successors, and to the many observers who, by recording and 
submitting their sightings, make distributional studies possible. Thanks also to Tim 
Manolis, Steve Speich, Tim Fitzer, and Jack Wilburn for information on previous local 
records, and to Dan Brown for contributing a photo of the nest. Tim Manolis and Ted 
Beedy provided very helpful suggestions on an earlier draft. 

LITERATURE CITED 

Baltosser, W. H. 1994. Age and sex determination in the Calliope Hummingbird. W. 
Birds 25:104-109. 

Calder, W. A., and Calder, L. L. 1994. Calliope Hummingbird (Stellula calliope ), in 
The Birds of North America (A. Poole and F. Gill, edsj, no. 135. Acad. Nat. Sci., 
Philadelphia. 

Gaines, David. 1992. Birds of Yosemite and the East Slope. Artemisia Press, Lee 
Vining, CA. 

Grinnell, J., and Miller, A. H. 1944. Distribution of the birds of California. Pac. Coast 
Avifauna 27. 

Harris, S. W. 1991. Northwestern California Birds. Humboldt State Univ. Press, 
Areata, CA. 

Kaufmann, K. 1990. Advanced Birding. Houghton Mifflin, Boston. 

LeValley, R., and Evens, J. 1983. Middle Pacific Coast region. Am. Birds 37:1022- 
1026. 

Orr, R. T.. and Moffitt, J. 1971. Birds of the Lake Tahoe Region. Calif. Acad. Sci., 
San Francisco. 

Roberson, D. 1993. Northern Parula, in Atlas of the Breeding Birds of Monterey 
County (D. Roberson and C. Tenney, eds.), p. 402. Monterey Peninsula 
Audubon Soc., Carmel, CA. 

San Miguel, G. L. 1985. Lower elevation breeding in the Sierra foothills. W. Tanager 
52 (3): 1-4. 

Sealy, S. G. 1979. Extralimital nesting of Bay-breasted Warblers: Response to forest 
tent caterpillars? Auk 96:600-603. 

Verner, J., Beedy, E. C., Granholm, S. L., Ritter, L. V., and Toth, E. F. 1980. Birds, 
in California wildlife and their habitats: Western Sierra Nevada (J. Verner and A. 
S. Boss, tech, coords.), pp. 75-319 . Gen. Tech. Rep. PSW-37. USDA Forest 
Service, Pacific Southwest Forest and Range Experiment Station, Berkeley, CA. 

Walley, W. J. 1977. An extra-limital nesting of the Wood Thrush in Manitoba. Blue 
Jay 35:82-86. 


Accepted 8 February 2001 


130 


NOTES 


RAPID SECOND NESTING BY 
ANNA S HUMMINGBIRD NEAR ITS 
NORTHERN BREEDING LIMITS 

ANN SCARFE, 4090 Gordon Head Road, Victoria, British Columbia V8N 3Y1 
J. CAM FINLAY, 270 Trevlac PL, Victoria, British Columbia V9E 2C4 


Anna’s Hummingbirds ( Calypte anna) were first reported in Victoria, British 
Columbia, in December 1944 and again each winter until 13 January 1947. 
Sightings continued through the 1950s (Taylor and Harper 1987), increased in 
numbers each year, and by 1970 the first individual was noted on the Victoria 
Christmas bird count. Numbers increased on subsequent counts, with 177 noted in 
1998 (D. Pearce pers. comm.). With these data and our banding records over the past 
two years, we estimate a minimum of 300 Anna’s Hummingbirds spend the winter 
and spring on the southeastern tip of Vancouver Island (greater Victoria). Given these 
numbers, we suspect this species might exhibit nesting overlap here as it does in 
Arizona and California (Russell 1996, Maender et al. 1996). 

The first documented nest of Anna’s Hummingbird in Canada was found near 
Duncan, 50 km north of Victoria, in July 1958. In Victoria a nest was first seen 30 
years later on 29 February 1988 (Campbell et al. 1990). Since then, nests have been 
found in increasing numbers each spring. 

Our study was conducted in Scarfe’s private garden approximately 1.5 km 
northeast of the University of Victoria, near the south end of Vancouver Island, from 
February to mid April in both 1998 and 1999. The female discussed here that laid two 
clutches was caught by means of a feeder within a wire cage trap. We banded her and 
colored the top of her head orange with Liquitex Value Series Basics acrylic color, a 
water-based paint that lasts about five months. Nest observations were made with the 
aid of a mirror on a long handle, with 7,8, and 9x binoculars and a 27x spotting 
scope. We obtained temperature data from the weather station on the campus of the 
University of Victoria. 

By the end of February 1998, Scarfe had located four active Anna’s Hummingbird 
nests in the greater Victoria region. When first spotted on 19 February, one of these 
(nest 1) was being built by a female 3.3 m above ground in Scarfe’s yard. The nest sat 
on a bare Douglas fir ( Pseudotsuga menziesii ) branch, 4 cm in diameter, with dense 
needle coverage above it and on its north and south sides. The north-facing wall of 
Scarfe’s residence was 2 m away, providing shelter from winds from the southeast. 
This nest contained two eggs when was first checked for eggs on 21 February. By 3 
March, nestlings were being fed, the first of which fledged on 20 March, the second 
on 23 March. The adult female was never trapped or color marked. 

On 22 March, only 33 m to the southwest of nest 1, Scarfe discovered a second 
female building a nest on the topmost, unsheltered branch of an apple tree ( Malus 
sp.), 4.75 m above ground. This female was trapped, banded, and color marked on 
23 March. That day we watched the unmarked female feeding one newly fledged 
juvenile. Also on 23 March, the marked female was seen feeding two newly fledged 
young (estimated 1-3 days out of the nest) in dense shrubbery about 8 m from the new 
nest tree. On 25 March we noted this female sitting on her nest at 10:00 when she left 
to feed the nearby fledglings about 7 m away. At 10:30 that day, we noted two eggs 
in the new nest. 

On 30 March, the marked female was incubating eggs and feeding her two 
fledglings, which were now more than 25 m away from nest 2 but still calling for food. 
By 3 April only one fledgling, being fed by the female, could be located. The next day 
it too disappeared. 


Western Birds 32:131-133, 2001 


131 


NOTES 


The marked female’s eggs hatched on, or just before, 14 April. We noted the 
female feeding nestlings and a small bill poking above the nest’s rim on that day. The 
two nestlings were present on 23 April, but the nest was empty on 27 April. The nest 
remained intact. The marked female remained around for several days and then 
disappeared. Over the summer both nests were gone. Feeders were available all 
summer but were seldom used. 

All four nests of Anna’s Hummingbirds that Scarfe found in 1998 had incubating 
females by the end of February. Another nest 2.5 km southeast of Scarfe’s residence 
fledged two young on 28 and 29 March 1998. 

The marked female was trapped again on 18 September 1998 and remarked. On 
12 February 1999 this marked female was seen building a nest in the exact spot where 
the unmarked female built the previous year. An egg was present on 19 February, 
almost a year from when the unmarked female began laying in 1998. A second egg 
appeared the next day. Both had hatched by the morning of 7 March, but the nestlings 
died that day. 

The southeast side of Vancouver Island has a Mediterranean-type climate with cold, 
wet winters. January temperatures sometimes fall below 5° C at night, but by mid 
February, overnight low temperatures seldom go below freezing and, if so, to only 
-1° or -2° C. 

In both 1998 and 1999 eggs were laid on or just before 19 February. Since the 
mean temperature for February was 5.5° C in 1998 but only 3.6°Cin 1999, it would 
seem that similarities in time of nest construction and egg laying in the two years were 
not entirely temperature dependent. We suggest that the early breeding has been 
aided by the increased numbers of hummingbird feeders being left filled all year, a 
result of public knowledge of the presence of this species. 

Because Anna’s Hummingbirds do raise two broods in California and Arizona 
(Russell 1996) and have been confirmed breeders on southern Vancouver Island for at 
least 40 years it was not surprising to find two broods being raised in Victoria. 

In 1999, the marked female selected the exact site used in 1998 by the unmarked 
female the previous year. Because this was a more sheltered site than the marked 
female had used the previous year, perhaps it was selected to avoid the effects of cold 
stress. Calder (1974) found that an Anna’s Hummingbird selected a sheltered nest site 
located so that the temperature within the nest was slightly warmer than the 
surrounding ambient air. 

Overlap of two broods (a female feeding late nestlings and incubating a second 
clutch simultaneously) was observed in the Broad-tailed Hummingbird ( Selasphorus 
platycercus ) by Bailey (1974). 

The use of the same spot two years in a row by an Anna’s Hummingbird is relatively 
uncommon but does occur. One of the earliest citations is by Bendire (1895). W. M. 
Tyler (in Bent 1940) noted that several other species of hummingbirds, including the 
Ruby-throated ( Archilochus colubris), Costa’s (C. costae), Rufous (S. ru/us), and 
Allen’s (S. sasin), have renested on the same site in two or more consecutive years. 
Tim Manolis (pers. comm.) has observed Black-chinned Hummingbirds (A. alexandri) 
reusing the same site in California. 

We express our appreciation to Environment Canada, Vancouver Island Special 
Production Office, for supplying temperature data. Thanks to Stephen M. Russell, 
William A. Calder, Tim Manolis, and an anonymous reviewer for providing construc- 
tive comments on the manuscript. We thank the James L. Baillie Memorial Fund of 
Bird Studies Canada for funding assistance. 

LITERATURE CITED 

Bailey, A. H. 1974. Second nesting of Broad-tailed Hummingbirds. Condor 76:350. 

Bendire, C. E. 1895. Life histories of North American birds from the parrots to the 
grackles. U. S. Natl. Mus. Spec. Bull. 3. 


132 


NOTES 


Bent, A. C. 1940. Life histories of North American cuckoos, goatsuckers, humming- 
birds and their allies. U. S. Natl. Mus. Bull. 176. 

Calder, W. A. Jr. 1974. The thermal radiant environment of a winter hummingbird 
nest. Condor 76:268-273. 

Campbell, R. W., Da we, N. K., McTaggart-Cowan, I., Cooper, J. M., Kaiser, G. W., 
and McNall, M. C. E. 1990. The Birds of British Columbia, vol. 2. Royal Br. 
Columbia Mus., Victoria. 

Maender, G. J., Hiett, K. L., and Bailey, S. J. 1996. Nesting Anna’s Hummingbirds 
in urban Tucson, Arizona. W. Birds 27:78-80. 

Russell, S. M. 1996. Anna’s Hummingbird ( Calypte anna), in The Birds of North 
America (A. Poole and F. Gill, eds.), no. 226. Acad. Nat. Sci., Philadelphia. 

Taylor, K., and Harper, C. 1987. Anna’s Hummingbird (Calypte anna) on Vancouver 
Island. Victoria Nat. 43 (6):9— 1 1 . 


Accepted 8 January 2001 


NOTES 


NESTING OF BRANDT S CORMORANTS 
IN THE NORTHERN GULF OF CALIFORNIA 

JUAN CERVANTES-SANCHEZ, C/ Nueva Zelanda 69 6A, 28035 Madrid, Spain 

ERIC MELLINK, Centro de Investigacion Cientifica y Educacion Superior de 
Ensenada, Apdo. Postal 2732, 22860 Ensenada, Baja California, Mexico (U.S. 
mailing address: P.O. Box 434844, San Diego, California 92143-4844) 


In Mexico, Brandt’s Cormorant ( Phalacrocorax penicillatus) breeds mostly on 
islands and offshore rocks along the Pacific coast of the peninsula of Baja California 
(Everett and Anderson 1991, Grinnell 1928, Wilbur 1987). The southernmost 
known colony on the Pacific coast is at the north end of Isla Margarita (approximately 
24° 30' N, 112° W; Guzman in Everett and Anderson 1991). Colonies in the Pacific 
can contain thousands of individuals. 

In the Gulf of California documented Brandt’s Cormorant breeding has been 
restricted to the Midriff Islands. In this region it nests or has nested on Roca Vela, on 
islands in Bahia de los Angeles, including Calavera, Smith, and Flecha, on Roca 
Partida (near Isla Partida Norte), Isla Partida (or Cardonosa), Isla Salsipuedes, Isla San 
Esteban, and Isla San Pedro Martir (Ainley et al. 1981, Banks 1963, Everett and 
Anderson 1991, Osorio-Tafall and del Toro-Aviles 1945, Tershy and Breese 1997, 
van Rossem and Hachisuka 1937, E. Palacios pers. comm.). In addition, van Rossem 
(1945) suspected some offshore rocks near Guaymas as a breeding locality for this 
species. Brandt’s Cormorant breeding colonies in the Gulf of California are small, 
rarely exceeding 250 pairs, and the total population in this area has been estimated 
at 500 to 1000 pairs (Everett and Anderson 1991). It is possible, however, that some 
colonies have been overlooked, through confusion of this species with the Double- 
crested Cormorant ( P. auritus ) (Russell and Monson 1998, Wilbur 1987). 

Here we report a breeding colony of Brandt’s Cormorant on Isla San Jorge in the 
northern Gulf of California, Mailliard (1923) and Mellink and Palacios (1993) 
previously reported on the breeding birds of this island, and neither recorded nesting 
Brandt’s Cormorants. Isla San Jorge (31° 01' N, 113° 15' W) comprises a series of 
small bare rocky islets located in the northeastern Gulf of California. 

On 9 November 1 999 we found a small breeding colony of Brandt’s Cormorant on 
the northeastern side of the southernmost islet (Figure 1). The colony, located on a 
30° slope, about 10 meters above mean sea level, was surrounded by a small 
congregation of California Sea Lions ( Zalophus californianus ) and some nesting 
Brown Boobies ( Sula leucogaster). Double-crested Cormorants were nesting on the 
upper ridges of the islet. We counted about 40 Brandt’s Cormorants, including 
breeding adults and first-year birds, and 24 nests, 19 of which contained one to four 
eggs. On 31 January 2000 we recorded 31 adults and 13 juveniles. Five adults were 
still attending nests. A small chick, yet to fledge, was visible in one of these. On 9 
January 2001 there were 39 adults and 26 young in the colony. Eight of the adults 
were attending nests. 

Our field work was supported by a grant from Conacyt (Mexico) to Mellink. We 
thank the Sociedad Cooperativa Ejidal Bahia de San Jorge for logistical support, 
Mario Machado and Carmelo Gil for transportation to the island, and Eduardo 
Palacios and Kimball Garrett for editorial assistance. 

LITERATURE CITED 

Ainley, D. G., Anderson, D. W., and Kelly, P. R. 1981. Feeding ecology of marine 
cormorants in southwestern North America. Condor 83:120-131. 


134 


Western Birds 32:134-135, 2001 


NOTES 



Figure 1 . Brandt’s Cormorant colony at Isla San Jorge, showing chicks three quarters 
grown and one incubating adult. Also visible are Blue-footed and Brown Boobies and 
California Sea Lions. 


Photo by Eric Mel link 


Banks, R. C. 1963. Birds of the Belvedere Expedition to the Gulf of California. Trans. 
San Diego Soc. Nat. Hist. 13:49-60. 

Everett, W. T., and Anderson, D. W. 1991. Status and conservation of the breeding 
seasons on offshore Pacific islands of Baja California and the Gulf of California. 
Int. Council Bird Protection Tech. Publ. 11:116-139. 

Grinnell, J. 1928. A distributional summation of the ornithology of Lower California. 
Univ. Calif. Publ. Zool. 32:1-300. 

Mailliard, J. 1923. Expedition of the California Academy of Sciences to the Gulf of 
California in 1921. Proc. Calif. Acad. Sci. 12:443-456. 

Mellink, E., and Palacios, E. 1993. Notes on breeding coastal waterbirds in northwest- 
ern Sonora. W. Birds 24:39-37. 

Osorio-Tafall, B. F., and del Toro-Aviles, M. 1945. Notas sobre la distribution de 
Phalacrocorax penicillatus (Brandt) en el Golfo de Cortes y la costa occidental 
de Baja California. Rev. Soc. Mex. Hist. Nat. 6:85-93. 

Russell, S. M., and Monson, G. 1998. The Birds of Sonora. Univ. of Ariz. Press, 
Tucson. 

Tershy, B. R., and Breese, D. 1997. The birds of San Pedro Martir Island, Gulf of 
California, Mexico. W. Birds 28:96-107. 

Van Rossem, A. J. 1945. A distributional survey of the birds of Sonora. Occ. Pap. 
Mus, Zool. La. State Univ. 21. 

Van Rossem, A. J., and Hachisuka, M. 1937. A further report on birds form Sonora, 
Mexico, with descriptions of two new races. Trans. San Diego Soc. Nat. Hist. 
8:321-336. 

Wilbur, S. R. 1987. The Birds of Baja California. Univ. of Calif. Press, Berkeley. 

Accepted 12 February 2001 


135 


NOTES 


A POTENTIAL THREAT TO BALD EAGLES 
IN BAJA CALIFORNIA SUR, MEXICO 

GUSTAVO ARNAUD, EDGAR AMADOR, AND MARCOS ACEVEDO, Centro de 
Investigaciones Biologicas del Noroeste, S.C., Apdo. Postal 128, La Paz, Baja 
California Sur, Mexico 23000 

The Southern Bald Eagle ( Haliaeetus leucocephalus leucocephalus ) is a resident 
species in the state of Baja California Sur (Grinnell 1928), occurring in mainly coastal 
habitats. Nesting is now known only from Bahia Magdalena- Alrnejas on the west 
coast (Henny et al. 1993, Amador-Silva and Guzman-Poo 1994, Rodriguez-Estrella 
et al. 1995) where no more than three pairs are found annually. The physiography of 
Baja California Sur is composed of extensive plains and hills, the principal mountain 
ranges are the sierras la Giganta and la Laguna, and predominant vegetation is desert 
scrub, which covers 92% of the state area (INEGI 1995). Freshwater habitats are 
scarce and largely temporary across the state. 

During bimonthly visits to the Sierra de la Giganta that commenced in June 1995 and 
continue to date, we have observed Bald Eagles on two occasions: On 13 April 1996 
we observed two adults, and on 27 March 1997 a single adult, soaring over a canyon 
located at 24° 50' N and 1 1 1° 00' W (80 km west of Bahia Magdalena), at 250 meters 
above sea level. This canyon includes a series of permanent fresh-water ponds (locally 
known as “pozas”) ranging in size from 35 to 100 m long and from 1 to 3 m deep. 
Unlike most other oases in Baja California Sur, these ponds contain abundant fish of at 
least two introduced species ( Oreochromis urolepis hornorum and Tilapia spp.). 

We have observed Ospreys ( Pandion haliaetus ) fishing in these ponds on several 
occasions and it seems likely that Bald Eagles may also utilize this resource, at least 
occasionally. Local informants report that fisherman occasionally shoot Ospreys, 
arguing that these birds take the fish they use for food. The hunting of Osprey is illegal, 
suggesting that fish-eating Bald Eagles could also be targeted, a potentially serious 
threat to the relatively small and vulnerable population nesting at Bahia Magdalena. 
Education of the fishermen in this area regarding the rarity of Bald Eagles in Baja 
California could be effective in helping to ensure the future of the peninsula’s last 
remaining nesting population of this endangered raptor. 

We thank Eduardo Palacios and Leopoldo Moreno for their critical review. Charles 
Henny and Peter Bloom also reviewed and improved this note. 

LITERATURE CITED 

Amador-Silva, E., and Guzman-Poo, J. 1994. El Aguila Calva (Haliaeetus 
leucocephalus ) en Isla Santa Margarita, Baja California Sur, Mexico. Revista de 
Investigaciones Cientificas, Serie Ciencias del Mar UABCS 5(l):33-35. 

Garcia, E. 1981. Modificaciones al sistema de clasificacion climatica de Koppen. 

Universidad Autonoma de Mexico, Mexico City. 

Grinnell, J. 1928. A distributional summation of the ornithology of Lower California. 
Univ. Calif. Publ. Zool. 32:1-300. 

Henny, C. J., Conant, B., and Anderson, D. W. 1993. Recent distribution and status 
of Bald Eagles in Baja California, Mexico. J. Raptor Res. 27:203-209. 

INEGI (Institute) Nacional de Estadistica Geografia Informatica). 1995. Sintesis 
Geografica de Baja California Sur. Publicaciones INEGI, Mexico City. 

Poole, A.F. 1989. Ospreys. Cambridge Univ. Press, Cambridge, England. 
Rodriguez-Estrella, R., Donazar, J, A., and Hiraldo, F. 1995. Fishermen and their 
gear may threaten Bald Eagles at Magdalena Bay, B.C.S. , Mexico. J. Raptor Res. 
29:144. 

Accepted 15 November 2000 


136 


Western Birds 32:136, 2001 


BOOK REVIEW 


The California Condor: A Saga of Natural History and Conservation, by 

Noel Snyder and Helen Snyder. 2000, Academic Press, San Diego. 410 pp., 118 
color and 21 black-and-white photos, 22 tables, 8 graphs, 4 maps. Hardback, 
$29.95. ISBN 0-12-654005-5. 

This nearly folio-sized book of 4 Vi pounds is aptly subtitled, given the history of the 
major study and conservation programs directed toward this species since the 1940s. 
A prologue tells of the authors’ first experience with California Condors and expresses 
the book’s goal, “to give the reader an appreciation of both the basic biology of the 
condor and the dynamics of condor conservation from a viewpoint mainly inside the 
conservation and research program.” The book is then organized into six “parts,” the 
first of which is “historical and background matters.” chapter 1, on “perspectives,” 
summarizes the species’ natural history and compares it with that of the Andean 
Condor and several Old World vultures. Chapter 2 provides an excellent account of 
the ceremonial and other uses made of California and Andean Condors by native 
peoples within their ranges. 

Part 1 concludes with the 39-page chapter 3, “Condor Research and Conservation 
in the Early-Mid 20th Century.” Much of this is derived directly from the same 
authors’ 1989 account “Biology and Conservation of the California Condor (Current 
Biology 6:175-267). Some of the text is identical, but other parts are rewritten and 
updated where, appropriate. As in their earlier account, the Snyders divided studies 
preceding theirs into segments that they now call “eras,” each named for one or two 
of the chief investigators or proponents: Finley (early 1900s), Robinson-Easton 
(1930s), Koford (the first detailed field study, 1939-46), Miller-McMillan (late 1950s 
into 1970s), [Fred] Sibley (1966-1969), and Wilbur (1970s). Throughout this 
chapter, the Snyders trace the developing knowledge of the condor’s behavior and 
ecology, along with changes in ideas as to its apparent decline. Each era is also 
credited with a special advance in knowledge or a particular emphasis that had a 
subsequent effect on the effort to conserve the species. Interaction of the sometimes 
conflicting ideas is discussed freely whenever it seems appropriate, and the authors 
give their own evaluation of the major advances and diversions. Thus the stage is set 
for Part II. 

This is titled “Struggles to Launch a New Program,” and its three chapters are 
“Battles in the Political Arena,” “Africa and Peru,” and “Development and Testing of 
Research Techniques.” Noel Snyder, of the U.S. Fish and Wildlife Service, and John 
Ogden, of the National Audubon Society, were named as joint managers of an 
expanded condor-conservation program beginning in early 1980. Noel’s wife Helen 
was also involved as an Audubon employee in much of the work of the research 
center, established in Ventura. Although they left those positions in 1986, and the 
recovery [advisory] team was subsequently discontinued by the Fish and Wildlife 
Service, it seems to me the Snyders are the best choice to tell the story of the often 
controversial efforts to conserve this endangered species, and this book is it! 

Chapter 4 details the campaign to establish and expand refuges and wilderness 
areas for the condor’s nesting. The Snyders believed from the outset that the very 
mobile condors should be studied first by radio-telemetry to discern the foraging range 
of individuals, and to aid in learning just what factors were involved in their population 
decline. They describe the controversies they encountered, including a tense meeting 
at the home of one of the chief proponents of the “hands-off” approach, and a “gag 
order” from their superiors in the Fish and Wildlife Service that prohibited them from 
contacting high-level California Department of Fish and Game officials directly. When 
freer communications were allowed in 1982, a more reasoned plan emerged. A 
permit issued in 1980 by the California Department of Fish and Game allowed the 
center to capture and radio-tag 10 birds, with the stipulation that one female be 


Western Birds 32:137-140, 2001 


137 


BOOK REVIEW 


retained as a mate for the then lone California Condor in captivity — the first admission 
by “hands-off” proponents that any efforts should be made toward building a captive 
population. But after the second chick handled died from stress, the California 
Department of Fish and Game permit was revoked, and only after two years was a 
new one issued. 

Without a permit to handle any California Condors, Snyder and Ogden went to 
Peru to observe, and sometimes participate in, research on the Andean Condor, thus 
learning much about capturing, radio-tagging, and behavior of condors. They also 
visited European research teams in South Africa, Zimbabwe, and Namibia working 
with the Cape Griffon and the Lappet-faced Vulture, the latter species showing 
breeding characteristics similar to the two condors’. Chapter 5, describing these trips, 
includes some interesting accounts of encounters with local peoples and even a 
narrow escape from being targets of a victorious but not yet disbanded army unit! 

Chapter 6 details the development of techniques for capturing condors (they 
selected cannon-netting), handling, blood-sampling, and sexing the birds, and radios 
for tracking of the birds so marked. The excitement attendant on establishing methods 
that would provide the data needed on the movement and lives of the marked birds 
comes through very nicely in this chapter. 

Part 111 is titled “Research Results of the New Program.” Its six chapters take up 
more than 100 pages, organized under censusing, movements and food, nest sites, 
breeding behavior, breeding effort and success, and mortality. Only a few highlights 
can be mentioned from this meaty basis for the rescue program that was finally 
adopted. The use of frequent photos of flying condors proved by far the most reliable 
tool for counting the number of birds still alive. California Condors obtained little food 
near their nest areas but traveled great distances to ranch lands around the southern 
San Joaquin Valley. When nesting, the adults used parts of that foraging range closer 
to their nest sites, but some young birds roved over the entire range. Characteristics 
of nest sites of different pairs varied, but most were difficult of access by humans or 
ground-based predators. Ravens, however, were watchful of many sites and quick to 
enter for a meal if the egg was left unguarded. Most pairs that lost an egg or young 
chick readily initiated another breeding effort the same year, but almost always at a 
different site and often miles away. This was not possible to detect without telemetry. 
Courtship postures are described, and one photo from the wild and one of a captive 
pair are included. 

Tables provide data pertinent to courtship, copulation, incubation shifts, percent- 
age of daylight hours adults were at or near nest sites, number of feedings of young by 
male and female parents, pairs breeding in the 1980s and the success of each, and 
dates when individual photo-documented birds were last seen. These are reported pair 
by pair or by individual, by means of identification codes of which only two are 
explained in the book (pp. 213-214). The other codes also probably have a 
geographic indication and were no doubt left unpublished to minimize potential 
disturbance when birds were still free-flying and nesting in those areas. Those familiar 
with California geography can guess at several others, so the continued secrecy may 
be of questionable value. More disconcerting, however, is the lack of any list of tables 
in the front of the book. Diligent readers should prepare such a list for themselves as 
not all of the text references to them are within two or three pages of the table. 

Chapter 12 discusses the known and potential causes of death of condors — from 
shooting, collisions with wires and wind turbines, possible effects of eating poisoned 
ground squirrels or coyotes, to cyanide guns set for coyote control. Strong evidence is 
presented that the predominant factor killing wild condors, at least in recent years, has 
been lead poisoning from bullet fragments ingested with food. The case histories of 
three radio-marked birds and one other found dead by a ranch foreman build the case 
like an intriguing detective story. This culminates in a step-by-step description of a 
January 1984 lab demonstration to skeptical state and federal officials that the last 


138 


BOOK REVIEW 


bird had, indeed, died from the cyanide-tracerite powder from a coyote-poisoning 
device. Only then was any restriction placed on the widespread use of such devices in 
the condors’ foraging range. 

In part IV, Conservation in the 1980s, chapter 13 surveys several efforts made to 
improve suboptimal nest sites by altering the slope of the cavity, building an external 
“porch” where the nearly fledged young could exercise their wings without falling off, 
by reducing the number of ravens in the area, etc. Of three ways to reduce the 
likelihood of condor deaths from lead poisoning, neither the conversion of the 
foraging range into a no-hunting zone nor prohibiting use of lead ammunition by 
hunters was ever tried, because of bureaucratic hurdles and lack of time. The third 
method, providing clean lead- and poison-free carcasses for the wild condors to eat, 
was tried on the Hudson Ranch in southwestern Kern County. But the condors shifted 
locations when deer-hunting season began in other parts of their range and so 
continued to eat meat with lead embedded. There is also a list of 30 steps taken from 
1937 to 1992, mostly by the federal government, that set aside or improved 
protection of areas to benefit the California Condor. 

Chapter 14, “Formation of a Captive Flock,” documents the controversies and 
slow stages in 1982-84 of this effort, the success in replacement clutching when 
removal of first-laid eggs of wild pairs was finally allowed and then hatched in 
incubators, and the questions as to the fraction of the population to which such efforts 
could be applied. Only after the winter of 1984—85 saw the disappearance of one or 
both members of all but one of the breeding pairs was the emphasis shifted to rapid 
development of a captive breeding flock, by taking the nine remaining birds into 
captivity. Another interplay of opposing views is described without rancor for those 
who didn’t agree with that goal and insisted that at least a few birds be left in the wild. 
Foremost among these proponents was the National Audubon Society, which sued to 
prevent the last captures. Most of chapter 15 is devoted to this topic, after precursor 
events described in chapter 14. This part ends with a summary of the “alternative 
plan” dependent on captive breeding to preserve the species “near term” but with 
eventual releases into the wild after safe numbers and genetic variety were in the 
population and after release techniques had been tested on Andean Condors 
temporarily released in the area. 

In part V, “Restoration,” chapter 16 describes the facilities built to house captives 
and the methods developed to promote pairing and at the same time preserve or even 
heighten the genetic diversity of this small population. Table 16 (p. 320) shows for 
each facility the number of mated pairs, eggs laid, and young fledged each year from 
1988 through 1998. The overall total, obtained only by adding yearly totals, was 141 
fledglings — very impressive from a start of only 9 adults and 18 immatures in 1987! 
Elsewhere in the chapter results are tabulated by each parent bird, and various 
comparisons are made of reproductive output with the few pairs followed in the wild 
(mostly in the early 1980s). Incubation periods averaged 57.2 days, slightly shorter 
than that of Andean Condors. This chapter also contains a host of other natural 
history information that would have been extremely difficult to obtain from wild birds 
even though critical to the planning for their survival. 

Techniques and results (through 1998) of “Releases to the Wild” (chapter 17) are 
described, with several appropriate photos. Results for each of 61 birds released in 
California and 28 in northern Arizona are summarized in table 22. The gist of the 
findings is that young raised by captive parents are much better at surviving when free 
to fly than those that are fed early in life only by “puppets” (gloves of very realistic 
condor head shapes) operated by hidden keepers. Those that undergo “aversive 
conditioning” to rectangular human-built structures fare somewhat better than those 
that don’t. Various release locations already tried, and several others with potential, 
are described. So also is the question of food subsidy and the conditions that might 
make it work by reducing the likelihood of the birds foraging widely and eating 


139 


BOOK REVIEW 


carcasses containing lead fragments. The authors give their own recommendations 
for releases and general captive propagation in this chapter, and those interested in 
the ultimate survival of the species should read them carefully. 

Part VI, “A General Evaluation,” includes in chapter 18 the authors’ philosophy of 
how efforts to conserve endangered species in general should be organized, with 
interplay of private and government units, intensive research and captive breeding if 
needed, but overall allowing for diversity of input and implementation guided by well- 
constructed recovery teams. These are distinguished from those that rely just on 
“recovery plans,” many of which have been written with inadequate supporting 
knowledge and are too detailed to allow for the constant alteration necessary as new 
facts emerge. The Snyders warn of the danger of goal substitution — of the preserva- 
tion of recovery programs being substituted for the recovery of the species itself. The 
chapter ends with nine recommendations specific to the recovery of the California 
Condor. The authors say, “[we] remain optimistic that unmanaged and viable wild 
condor populations can be reestablished, but we are deeply concerned about the 
recent rate of progress toward that goal.” 

Maybe those who want to know just about the California Condor’s natural history, 
without wading through this history of the research and conservation efforts directed 
toward its survival, will tire of all the accounts of controversies. However, for those 
interested in complex species-conservation problems and the associated negotiations 
and politics, this is a really valuable story. Many additions to the natural history saga are 
to be found only embedded in the conservation saga that forms the bulk of the book. 
Most technical aspects of the book’s production are also of high quality: heavy glossy 
paper, attractive layout, superb photos (drawings in two cases where more pertinent), 
a generous set of acknowledgments (and a special list of credits in front for all the 
photos), a bibliography of 382 cited references, and a thorough index. I noted only 
about three typographical errors. In short, anyone interested in the California Condor 
or in the conservation of endangered species in general should study this book. 

Howard L, Cogswell 


140 


FEATURED PHOTO 


RANGE EXPANSION OF THE GREAT-TAILED 
GRACKLE IN WESTERN NORTH AMERICA 

WALTER WEHTJE, Geography Program, Department of Earth Sciences, University 
of California, Riverside 92521 


The spread of the Great-tailed Grackle (Quiscalus mexicanus) is one of the most 
impressive range expansions to occur in North America during the 20th century. The 
U,S. breeding range was limited to southernmost Texas in 1900 (Ridgway 1902), but 
a century later Great-tailed Grackles breed in 19 states, from Arkansas in the East 
north to Minnesota and west to California and Oregon (Dinsmore and Dinsmore 
1993, Scheuering and Ivey 1995, Price 1997, Granlund 1999), with sightings in 
Washington, British Columbia, Montana, North Dakota, Wisconsin, and Ontario 
(Dinsmore and Dinsmore 1993, Granlund 1999). 

The northward spread of Great-tailed Grackles in North America has occurred on 
three fronts.- Birds from south Texas, of the subspecies Q. m. prosopidicola, moved 
north and east and are currently found east of a north-south line through western 
Texas and eastern Colorado. A separate population from Chihuahua invaded New 
Mexico in 1913, with breeding confirmed in eastern Arizona in 1937 and Colorado 
in 1973 (Bailey 1928, Phillips et al. 1964, Stepney 1975). Phillips (1950) described 
this population as a new subspecies, Q. m. monsoni, with a range that encompassed 
north-central Mexico, western Texas, southern and central New Mexico, and eastern 
Arizona. By 2000, monsoni was present in New Mexico, southeastern Nevada, 
Arizona, and California. The third subspecies, Q. m. nelson i , originated in coastal 
Sonora, Mexico, and adjacent areas (Friedmann et al. 1957), arriving in southern 
Arizona by the late 1930s (Phillips et al. 1964). After moving into Arizona, nelsoni 
expanded west and was first noted in California in 1964 (McCaskie et al. 1966); it is 
now found in Arizona, California, and southwestern Nevada. The subspecific affinities 
of Great-tailed Grackles breeding in Oregon, Idaho, Utah, and central Nevada have 
not been established (Dinsmore and Dinsmore 1993, Scheuering and Ivey 1995). 

Great-tailed Grackle subspecies differ in both size and color. Adult males of Q. m. 
monsoni are nearly 20% heavier than adult males of Q. m. nelsoni, with average 
(flattened) wing lengths of 191 mm vs, 167 mm (14% longer) and tail lengths of 210 
mm vs. 165 mm (27% longer; W. Wehtje unpublished data; monsoni n = 7, nelsoni 
n = 8). Adult females of monsoni are at least 10% heavier than females of nelsoni, 
with average (flattened) wing lengths of 150 mm vs. 136 mm (10% longer) and tail 
lengths of 150 mm vs. 126 mm (19% longer; W. Wehtje unpublished data; monsoni 
n = 23, nelsoni n = 6). In addition to the size differences, females of the two 
subspecies differ in plumage color: nelsoni shows pale grayish buffy underparts, 
monsoni darker brownish gray underparts (Phillips 1950, Rea 1969). 

Clockwise from upper left, birds shown in the featured photo on the back cover are 
an immature female nelsoni (26 January 2000, Oxnard, California), adult male 
nelsoni (10 December 1999, Tucson, Arizona), immature female monsoni (12 
December 1999, Arlington, Arizona), adult male monsoni (2 June 2000, Bill 
Williams National Wildlife Refuge [NWR], Arizona), adult female monsoni (10 
December 1999, Tucson), adult female monsoni (5 June 2000, Pena Blanca Lake, 
Arizona), adult female monsoni (2 June 2000, Bill Williams NWR), and immature 
female nelsoni (15 May 1994, Salton Sea NWR, California). It bears noting that 
recent specimens of adult female nelsoni in fresh plumage are, apparently, lacking 
from collections. 

In an ideal world, identifying Great-tailed Grackles to subspecies in the western US 


Western Birds 32:141-143, 2001 


141 


FEATURED PHOTO 


would be simple: large males and dark females would be of the subspecies monsoni, 
and small males and pale females would be of the subspecies nelsoni. In reality, this 
approach is complicated by plumage differences between adult and immature birds, 
and the degree to which female Great-tailed Grackles fade during the spring. The 
presence of intergrades complicates the picture further. 

In common with several other icterids, Great-tailed Grackles have distinct first-year 
and adult plumages (Pyle 1997). Immature birds have duller plumages, shorter wings, 
and shorter tails than adults. Juvenile males molt into their first basic plumage, a dull 
black, by late summer. Not until their second fall do they molt into the long-tailed and 
glossy plumage of the adult males; therefore, meaningful comparisons between males 
can be made only with birds of the same age. The featured photo illustrates the size 
difference between adult males ( monsoni flat wing 184 mm, tail 201 mm; nelsoni flat 
wing 174 mm, tail 174 mm). 

Determining the subspecies of female Great-tailed Grackles is more difficult, While 
size differences are useful, plumage characteristics are of greater importance. Deter- 
mining a bird’s age is essential to assigning it to either nelsoni or monsoni. Adult 
females have cream-colored irises, while immatures have khaki-colored irises with 
dark flecking until at least June of their second year. Immature females are smaller and 
duller than adults, and their feathers tend to wear and fade more rapidly and 
extensively than those of adults. The difference between adults and immatures 
becomes most pronounced in late spring, when immatures can have pale gray 
underparts with no hint of brown in them. By late summer both adults and immatures 
have faded underparts and should not be identified to subspecies. 

The two females on the upper row of the photograph are newly molted immatures. 
The left-hand bird is an unusually gray nelsoni (flat wing 134 mm, tail 122 mm), and 
the right-hand bird is a more or less typical monsoni (flat wing 139 mm, tail 127 mm). 
The size difference between adult females would normally be greater than this, but 
even a more typical nelsoni would be obviously paler than the rich brown character- 
istic of immature female monsoni. 

The bird in the lower right-hand corner of the featured photo is a fresh adult female 
monsoni showing underparts that are a deeper brown than ever shown by female 
nelsoni. The two middle birds demonstrate the range of fading shown by adult female 
monsoni by early June (these birds were collected within three days of each other). 
While each of these birds has faded significantly, neither is as worn as the immature 
nelsoni on the far left, which was collected at the Salton Sea in mid-May. Birds in this 
extreme environment tend to have completely pale underparts by late spring. 

In most of central California and Nevada, the plumage of grackles is often visibly 
faded by March and April, when these migratory populations arrive from southern 
Arizona and southern California. Since these grackles head back south in late summer 
before undergoing prebasic molt, they are difficult to observe in fresh plumage on the 
breeding grounds and therefore difficult to identify with certainty. Judicious collecting 
early in the nesting season should shed light on the interaction between the two 
subspecies as they continue to spread across western North America. 

Another confounding factor in distinguishing between the two western subspecies 
is the presence of intergrades between them. Interbreeding between monsoni and 
nelsoni was observed in Arizona in the mid 1960s (Rea 1969) and has continued to 
the present. In California, male Great-tailed Grackles tend to be intermediate in size 
between the two subspecies, while females are as dark as monsoni but closer in size 
to nelsoni. The degree of interbreeding in other states is not known. 

I would like to thank the curators and collection managers of the following 
institutions for the opportunity to examine their Great-tailed Grackle skins: Carnegie 
Museum, Los Angeles County Museum of Natural History, Museum of Southwest 
Biology, Museum of Vertebrate Zoology, San Bernardino County Museum, San 
Diego Natural History Museum, Santa Barbara Museum of Natural History, University 


142 


FEATURED PHOTO 


of Arizona Vertebrate Museum, University of California Santa Barbara Vertebrate 
Museum, and the Western Foundation of Vertebrate Zoology. I also thank the 
agencies and individuals who gave me permission to collect grackles. A Frank M. 
Chapman Memorial Grant from the American Museum of Natural History funded 
portions of this research. Finally, I thank Robb Hamilton for suggesting I write this 
article and Amadeo Rea for reviewing and improving the manuscript. 

LITERATURE CITED 

Bailey, F. M. 1928. Birds of New Mexico. N. M. Dept. Game and Fish, Santa Fe. 

Dinsmore J. J., and Dinsmore, S. J. 1993. Range expansion of the Great-tailed 
Crackle in the 1900s. J. Iowa Acad. Sci. 100:54-59. 

Friedmann, H., Griscom, L., and Moore, R. T. 1957. Distributional checklist of the 
birds of Mexico. Pac, Coast Avifauna 29. 

Granlund, J. 1999. Western Great Lakes region. N. Am. Birds 53:281-283. 

McCaskie, G., Stallcup, R., and DeBenedictis, P. 1966. Notes on the distribution of 
certain icterids and tanagers in California. Condor 68:595-597. 

Phillips, A. R. 1950. The Great-tailed Grackles of the Southwest. Condor 52:78-81. 

Phillips, A., Marshall, J., and Monson, G. 1964. The Birds of Arizona. Univ. of Ariz. 
Press, Tucson. - 

Price, J. 1997. Changing seasons. Natl. Audubon Soc. Field Notes 51:832-835. 

Pyle, P. 1997. Identification Guide to North American Birds, part I, Columbidae to 
Ploceidae. Slate Creek Press, Bolinas, CA. 

Rea, A. M. 1969. The interbreeding of two subspecies of Boat-tailed Grackle 
Cassidix mexicanus nelsoni and Cassidix mexicanus monsoni in secondary 
contact in central Arizona. M.S. thesis, Univ. Ariz., Tempe. 

Ridgway, R. 1902. Birds of Middle and North America, part II. Bull. U. S. Natl. Mus. 
50:1-834. 

Scheuering, E. J., and Ivey, G. L. 1995. First nesting of the Great-tailed Grackle in 
Oregon. Wilson Bull. 107:562-563. 

Stepney, P. H. R., and Power, D. M. 1975. First recorded breeding of the Great-tailed 
Grackle in Colorado. Condor 77:208-210. 


HELP SUPPORT THE WFO PUBLICATION FUND 

Western Field Ornithologists recently established a Publication Fund, which will be 
used specifically to increase the size of issues of Western Birds, include more color 
photographs, and initiate a new monograph series. We have already received 
financial support from the WFO Board, and we are pursuing sponsorship agreements 
with major optical companies. Via this announcement, we also are soliciting the 
financial support of our members and readers. Please send donations, payable to 
Western Field Ornithologists and earmarked “Publication Fund,” to Dori Myers, WFO 
Treasurer, 6011 Saddletree Lane, Yorba Linda, CA 92886. Thanks very much for 
your support! 


143 


Wing Your Way to . . . 

Reno, Nevada 

Western Field Ornithologists’ 
26 th Annual Meeting 

September 27-30, 2001 



Come join us in Reno for an exciting weekend of birding, 
panels, paper sessions, and vendor displays. 


Contact Lucie Clark at 775-831-2909 
or via e-mail: luclark@sierra.net 
VISIT OUR WEBSITE 
www.wfo-cbrc.org 


World Wide Web site: 

WESTERN BIRDS www.wfo-cbrc.org 

Quarterly Journal of Western Field Ornithologists 

President: Mike San Miguel, 2132 Highland Oaks Dr., Arcadia, CA 91006; 
sanmigbird@aol.com 

Vice-President: Daniel D. Gibson, University of Alaska Museum, 907 Yukon 
Dr., Fairbanks, AK 99775-6960 

Treasurer/Membership Secretary: Dori Myers, 6011 Saddletree Lane, Yorba 
Linda, CA 92886 

Recording Secretary: Lucie Clark, 9889 Tahoe Blvd., #56, Incline Village, NV 
89451 

Directors: Kimball Garrett, Daniel D. Gibson, Bob Gill, Gjon Hazard, Dave 
Krueper, Mike San Miguel, W. David Shuford, Mark K. Sogge, David Yee 

Editor: Philip Unitt, San Diego Natural History' Museum, P.0. Box 121390, San 
Diego, CA 92112-1390; birds@sdnhm.org 

Associate Editors: Daniel D. Gibson, Robert A. Hamilton, Ronald R. LeValley, 
Tim Manolis, Kathy Molina, Mark K. Sogge 

Graphics Manager: Virginia P. Johnson, 4637 Del Mar Ave., San Diego, CA 92107 

Photo Editor: Peter La Tourrette, 1019 Lorna Prieta Ct., Los Altos, CA 94024 

Featured Photo: Robert A. Hamilton, 34 Rivo Alto Canal, Long Beach, CA 90803 

Book Reviews: Steve N.G. Howell, Point Reyes Bird Observatory, 4990 Shoreline 
Highway, Stinson Beach, CA 94970 

Secretary, California Bird Records Committee: Guy McCaskie, P.O. Box 275, 
Imperial Beach, CA 91933-0275; guymcc@pacbell.net 

Chairman, California Bird Records Committee: Richard A. Erickson, USA Associates, 
1 Park Plaza, Suite 500, Irvine, CA 92614; richard.erickson@lsa-assoc.com 


Membership dues, for individuals and institutions, including subscription to Western 
Birds: Patron, $1000.00; Life, $400,00 (payable in four equal annual installments); 
Supporting, $60 annually; Contributing, $34 annually; Family, $26; Regular U.S. 
$22 for one year, $41 for two years, $60 for three years, outside U.S. $27 for one 
year, $51 for two years, $73 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 orders for back issues and special publications to the Treasurer. Make 
checks payable to Western Field Ornithologists. 

Back issues of Western Birds within U.S. $24 per volume, $6.00 for single issues, 
plus $1.00 for postage. Outside the U.S. $30 per volume, $7.50 for single issues. 

The California Bird Records Committee of Western Field Ornithologists recently 
revised its 10-column Field List of California Birds (January 2000). The last list covered 
606 accepted species; the new list covers 613 species. Please send orders to WFO, 
c/o Dori Myers, Treasurer, 6011 Saddletree Lane, Yorba Linda, CA 92886. 
California addresses please add 7.75% sales tax. 

Quantity: 1-9, $1.50 each, includes shipping and handling. 10-39, $1.30 each, add $2.00 
for shipping and handling. 40 or more, $1.15 each, add $4.00 for shipping and handling. 


Published September 15, 2001 


ISSN 0045-3897 








Vol. 32, No. 3, 2001 



Volume 32, Number 3, 2001 

Distribution and Abundance of Winter Shorebirds on Tomales Bay, 


California: Implications for Conservation John P. Kelly 145 

A Targeted Mist Net Capture Technique for the Willow Flycatcher 
Mark K, Sogge, Jennifer C. Owen, Eben H. Paxton, 

Suzanne M. Langridge, and Thomas J. Koronkiewicz 167 

NOTES 

Further Evidence for a Population Decline in the Western 

Warbling Vireo Thomas Gardali and Alvaro Jaramillo 173 

Brandt’s Cormorant Sinks At Sea Terence R Wahl 177 

First Record of the European Golden- Plover from the Pacific 

Andrew W. Piston and Steven C. Hein l 179 

Book Reviews Steve N. G. Howell, Gregory H. Golet 182 

Featured Photo Jon L. Dunn and Kimball L. Garrett 186 


Cover photo by © Jim Burns/NATURAL IMPACTS of Scottsdale, 
Arizona: Rufous-capped Warbler ( Basileuterus rufifrons), French 
Joe Canyon, Arizona, August, 2000. 


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 include 
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 the Rocky Mountain and Pacific 
states and provinces, including Alaska and Hawaii, western Texas, northwestern Mexico, 
and the northeastern Pacific Ocean. 

Send manuscripts to Kathy Molina, Section of Ornithology, Natural 
History Museum of Los Angeles County, 900 Exposition Blvd., Los Angeles, CA 90007. 
For matter of style consult the Suggestions to Contributors to Western Birds (8 pages 
available at no cost from the editor) and the Council of Biology Editors Style Manual 
(available for $24 from the Council of Biology Editors, Inc., 9650 Rockville Pike, 
Bethesda, MD 20814). 

Reprints can be ordered at author’s expense from the Editor when proof is returned or 
earlier. 

Good photographs of rare and unusual birds, unaccompanied by an article but 
with caption including species, date, locality and other pertinent 
information, are wanted for publication in Western Birds. Submit photos and 
captions to Photo Editor. Also needed are black and white pen and 
ink drawings of western birds. Please send these, with captions, to 
Graphics Manager. 


WESTERN BIRDS 


Volume 32, Number 3, 2001 



DISTRIBUTION AND ABUNDANCE OF WINTER 
SHOREBIRDS ON TOMALES BAY, CALIFORNIA: 
IMPLICATIONS FOR CONSERVATION 

JOHN P. KELLY, Cypress Grove Research Center, Audubon Canyon Ranch, Marshall, 
California 94940 


ABSTRACT: 1 analyzed the distribution and abundance of wintering shorebirds 
(Scolopacidae, Charadriidae, and Recurvirostridae) in Tomales Bay, California, on the 
basis of 57 baywide counts conducted over 10 years, from 1989-90 to 1998-99. 
Tomales Bay supports up to 20,689 shorebirds in early winter, thus qualifying as a 
wetland of “regional” importance in the Western Hemisphere Shorebird Reserve 
Network. Minimum overall shorebird abundance fell as low as 1291 in late winter. 
Tomales Bay supported approximately a third of the wintering shorebirds in the Point 
Reyes/Bodega area in early winter. Observations of tidally structured flock move- 
ments of several species suggested that the northern and southern ends of Tomales 
Bay are occupied by different wintering groups. In association with cumulative 
seasonal rainfall, most species declined in abundance significantly in midwinter. The 
Sanderling and Marbled Godwit increased with cumulative rainfall in the north and 
south bay, respectively, suggesting weather- related influxes from outer coastal beaches. 
After accounting for the effects of cumulative seasonal rainfall and a 10-year trend in 
annual rainfall, I detected no trends in species’ abundances. Foraging and roosting 
shorebirds at the northern end of the bay were vulnerable to direct disturbance from 
concentrated recreational use. Long water-residence times in southern Tomales Bay 
suggest that shorebirds there may be particularly vulnerable to toxic spills or anthro- 
pogenic eutrophication. The closeness of San Francisco Bay implies a high potential 
for invasion of nonnative organisms established there, which could alter the availabil- 
ity of benthic prey to shorebirds in Tomales Bay. Shorebird feeding habitat at the 
deltas of Walker and Lagunitas creeks may be adversely affected by heavy rainfall 
leading to the deposition of sediment. Daily influxes of roosting gulls from a local 
landfill were associated with reduced shorebird use of tide flats. Shorebirds’ use of 
open tide flats developed for mariculture is reduced, although floating oyster bags 
provide roosting areas during high tides. Breaching levees that isolate historic 
wetlands may increase shorebird use in some areas. The likelihood of regular or 
episodic intraseasonal movements among Point Reyes/Bodega area wetlands sug- 
gests Tomales Bay and other nearby wetlands are worthy of broad protection. 


Western Birds 32:145-166, 2001 


145 




DISTRIBUTION AND ABUNDANCE OF WINTER SHOREBIRDS ON TOMALES BAY 


Tomales Bay is one of California’s largest and least disturbed estuaries. It 
has been recognized as a wetland of “regional importance” on the basis of 
its supporting up to 20,000 shorebirds, a criterion for inclusion in the 
Western Hemisphere Shorebird Reserve Network {WHSRN; Harrington 
and Perry 1995, Page and Shuford 2000). An analysis of shorebird 
abundance and distribution on the Pacific coast of the United States 
identified Tomales Bay as a key wetland for shorebird conservation because 
it supports in at least one season at least 1% of the total population of 8 of 
13 shorebird species that concentrate in estuaries and brackish wetlands 
(Page et al. 1999). In spite of this recognition, little has been published on 
the spatial or temporal variations in shorebird abundance on Tomales Bay, 
or the birds’ vulnerability to environmental threats. Kelly and Tappen (1998) 
reported previously on the value of Tomales Bay to other winter waterbirds. 

Shuford et al. (1989) analyzed the results of up to 10 consecutive years of 
shorebird counts in the adjacent Point Reyes area between 1965 and 1982 
but provided little information on shorebird use of Tomales Bay. As part of 
a statewide survey, Jurek (1974) reported on five years of monthly shorebird 
counts at the Walker Creek delta, in the northern part of the Tomales Bay, 
but did not survey abundances baywide. Over recent decades, Tomales Bay 
has been suggested for protection as habitat for shorebirds (Smail 1972, 
National Oceanic and Atmospheric Administration 1987, Neubacher et al. 
1995, Page and Shuford 2000). However, with one exception dealing with 
the effects of aquaculture on shorebirds (Kelly et al. 1996), evaluations of 
conservation issues in Tomales Bay have not directly addressed patterns of 
shorebird use. 

In this paper, 1 (1) present results from 10 years of baywide winter 
shorebird censuses on Tomales Bay; (2) compare these results with other 
studies to provide a historical and geographical perspective on the impor- 
tance of Tomales Bay to shorebirds; (3) evaluate species’ distributions within 
the bay with regard to the importance of particular areas used by shorebirds; 
(4) examine processes that influenced abundance trends over the 10 years of 
study, and (5) identify needs for conservation of wintering shorebirds on 
Tomales Bay. I address these objectives with regard to all Scolopacidae, 
Charadriidae, and Recurvirostridae associated with Tomales Bay and imme- 
diately adjacent seasonal wetlands (Figure 1). 

STUDY AREA 

Tomales Bay floods the lower 20 km of the fault-rift Olema Valley on the 
central California coast, about 45 km northwest of San Francisco (Figure 1 ; 
Galloway 1977). Approximately 18% of the bay’s 28.5-km 2 area is inter- 
tidal, providing primarily sand or mud flats suitable for foraging shorebirds 
and cobblestone beaches along the east shore. In general, sediments grade 
from primarily fine to coarse sand in the northern reaches of the bay to 
muddier substrates in southern portions of the bay (Daetwyler 1966). 

Two primary points of freshwater inflow, Lagunitas Creek at the south 
end of the bay and Walker Creek near the north end of the bay, are 
associated with large tidal deltas suitable for foraging shorebirds (Figure 1). 
Numerous smaller delta marshes and tidal flats occur where small perennial 


146 


DISTRIBUTION AND ABUNDANCE OF WINTER SHOREBIRDS ON TOMALES BAY 


Dillon 



Figure 1. Shorebird count areas on Tomales Bay, California. 1, White Gulch; 
2, Lawson’s Meadow; 3, Sand Point; 4, Tom’s Point; 5, Walker Creek delta; 6, North 
Marshall; 7, South Marshall; 8, Lagunitas Creek delta; 9, Giacomini Pasture; 10, 
Inverness shoreline. Arrows indicate length of shoreline in each count area; labels 
indicate extent (ha) of exposed tidal flat at MLLW. 


and ephemeral streams enter the bay. Adjacent seasonal wetlands suitable 
for shorebirds are normally limited to about 15 ha of a 200-ha diked pasture 
at the south end of the bay and approximately 20 ha of wet meadow 
surrounded by sand dunes at the north end of the bay. Additional seasonal 
wetlands, used by shorebirds during periods of heavy flooding, occur in 
agricultural areas northeast of the bay. 

Most (95%) of the annual rain falls when wintering shorebirds are present, 
from October through April, with 55% falling from December through 
February (Audubon Canyon Ranch, unpublished data). Constraints on tidal 
exchange with the ocean, imposed by the linear shape of the bay, interact 
with winter runoff to create contrasting habitats at the two ends of the bay. 
In the southern half of the bay, variably high levels of freshwater inflow 
during winter, low flows in summer, and long water-residence times result in 
highly variable salinities (Hollibaugh et al. 1988). Salinities range from nearly 


147 



DISTRIBUTION AND ABUNDANCE OF WINTER SHOREBIRDS ON TOMALES BAY 

fresh after heavy winter runoff to slightly hypersaline in late fall (Smith and 
Hollibaugh 1997). In northern Tomales Bay, regular tidal mixing maintains 
salinities that consistently reflect those of nearshore waters along the outer 
coast. The difference between mean high and mean low tides is about 1.1m, 
with an average annual maximum tide swing of about 2.5 m (U. S. National 
Oceanic and Atmospheric Administration harmonics and correction tables; 
Tides and Currents, Nautical Software, Inc.). 

METHODS 

I divided Tomales Bay into 10 subareas within which all shorebirds could 
be counted in 60 to 90 minutes (Figure 1). These areas include almost all of 
the intertidal flats in the bay, with the exception of a few small areas at creek 
mouths along the west shore. Teams of qualified observers counted all 
shorebirds by species in all subareas simultaneously. 

Observers counted during rising tides, at tide levels between 0.76 and 
1.22 m above mean lower low water (MLLW) at Blake’s Landing (Figure 1). 
Count-area protocols were coordinated so that adjacent areas were surveyed 
simultaneously. Abundances generally represented counts of foraging indi- 
viduals, with estimates of large mobile flocks made only rarely. During three 
workshops held to examine and reduce observer bias in estimating flock size, 
biases were inconsistent and without a clear central tendency. Therefore 
observer bias was assumed not to affect overall estimates of abundance. The 
time and direction of all flock movements, departures, and arrivals during 
count periods were recorded and examined later to minimize chances of 
birds being double counted. Counts were conducted only on days with 
weather suitable for using telescopes to identify shorebirds. 

Each year from 1989-90 to 1998-99, we completed approximately 
three counts in each of two intraseasonal periods (n = 57): early winter 
(mean 2.8, range 1-4; 1 November-19 December) and late winter (mean 
2.9, range 2-3; 15 January-4 March). I summarized results within 
intraseasonal groups to compare early- and late- winter population levels. 
Most individual shorebirds were identified to species. Occasionally, individu- 
als were not identified to species and were recorded in pooled species groups 
such Least/Western sandpiper. Individuals recorded in pooled species groups 
were allocated to single species groups in direct proportion to the number of 
identified birds of each species in each count area on each day (Page et al. 
1999). Such adjustments were made, however, only if the number of 
identified individuals exceeded the number of undifferentiated individuals and 
was >50 for groups of two species or >100 for groups of three species. 

Observations of shorebird flocks revealed restricted movements in the 
middle section of the bay. Of 19 flocks of several shorebird species observed 
departing as the tide inundated the feeding area at Walker Creek delta, 
100% were oriented to the northwest and only one flock flew by from farther 
south. In southern Tomales Bay, 56 flocks of several species were observed 
departing north from feeding areas during rising tides, but only one flock 
(2%) was seen flying north past Cypress Point in Marshall (Figure 1). These 
observations suggest that several species winter in discrete groups. In 
addition, a reciprocal translocation of color-banded Dunlins, the most 


148 


DISTRIBUTION AND ABUNDANCE OF WINTER SHOREBIRDS ON TOMALES BAY 


abundant species in the bay, revealed that wintering populations at the two 
ends of the bay were separate (Kelly 2000). Therefore, I analyzed the results 
separately for subregions of Tomales Bay north (areas 1-6) and south (areas 
7-10) of Cypress Point, Marshall, in addition to analyzing them for the bay 
as a whole (Figure 1). 

I transformed abundance data into densities, based on the extent of 
exposed mudflat in each count area, estimated at MLLW with a planimeter 
on a bathymetric chart (Figure 1; U. S. National Oceanic and Atmospheric 
Administration Chart 18643, 16th ed., Dec 1995). Estimates of shorebird 
preferences based on densities among different habitat types may be 
sensitive to different measures of availability (Warnock and Takekawa 
1995), particularly if birds concentrate along the edge of the rising tide 
(Recher 1966, Burger et al. 1977) and the linear extent of the tide line is not 
closely correlated with areal extent of the habitat. However, suitable shore- 
bird habitat on Tomales Bay is limited primarily to tidal flats, and each count 
area represented a similar proportion of the area available for foraging along 
the tide line at any particular tide level. Consequently, densities based on 
different widths of tidal exposure did not significantly alter estimates of 
relative shorebird use by count area. 

To examine the effects of count-area location, intraseasonal period, and 
year on patterns of shorebird use, I used mixed-model analysis of covariance 
of shorebird densities, controlling for days since the beginning of each 
intraseasonal (early or late winter) period and treating year as a random 
effect. Shorebird densities were log-transformed to improve normality and 
stabilize the variance of residuals. Significant count-area effects were fol- 
lowed by multiple pairwise comparisons within the northern and southern 
subregions of the bay. I inspected bivariate scatterplots of shorebird abun- 
dance on cumulative seasonal rainfall, using Cleveland’s robust locally 
weighted regression algorithm (LOWESS; Cleveland 1979, Chambers et al. 
1983), to look for nonlinear thresholds and trends. I selected this method 
because it uses locally weighted least squares and a robust fitting procedure 
to define smoothed points that are relatively insensitive to outlying values, 
and it allows a flexible degree of smoothing by adjusting the proportion of 
data scanned (/) for each fitted value. To improve the normality of abun- 
dances used in the analysis, I based trends for the Semipalmated Plover, 
Black Turnstone, and Least Sandpiper on the natural logarithms (In) of the 
counts; distributions of other species’ abundances were approximately 
normal without this transformation. Annual and cumulative seasonal rainfall 
were derived from daily rainfall measured at Cypress Point (Figure 1; 
Audubon Canyon Ranch, unpubl. data). 1 estimated annual trends in species 
abundance within subregions as partial regression coefficients on year, 
independent of the linear effect of cumulative seasonal rainfall. I compared 
these results with correlated trends in annual rainfall. 

RESULTS 

Tomales Bay supported as many as 20,689 shorebirds in early winter, 
with minimum numbers falling as low as 1291 in late winter. The most 
abundant winter shorebird in Tomales Bay was the Dunlin, accounting for 


149 


Table 1 Mean, Minimum, and Maximum Abundances of Wintering Shorebirds on Tomales Bay, 1989-90 to 1998-99° 

North bay South bay Baywide 

urintpr I ate winter Earlu winter Late winter Early winter Late winter 


CD 

cn 

c 

ro 

PC 



c 


CD 

£ 


CD 

cn 

c 

(C 

PC 


U4 

to 

c 

(C 

CD 

s 


UJ 

CO 

c 

CO 

CD 


# 

00 

LD 


t-H 

* 

* 

eg 

« 

* 

ON 

» 

# 

00 


4t 

* 

CM 

to 

# 

CO 

co 

eg 

o 

rH 


oo 

CO 

t — ) 

CM 

i 

eg 

oo 


i 

rH 

o 


i 

o 

o 

d> 

1 

O 

1 

O 

o 

i 

i 

o 








^ ^ 




„ 



g 

o 

Co 

g 

to 


CO 

eg 

on 

CO 

CM 

t-H 

t-H 

o 

o 

T-H 

CO 

T-H 

o 

i — 1 


o 

O 

g. 

t-H 






g 


CM 




o 

i — ) 

o 

CO 


CO 

rH 

Ot 

+ 

to 

o 

t-H 

CM 

t-H 

+ 














eg 




CO 

t-H 



to 


tn 


co 

i 


CM 

CO 

eg 

t-H 


oo 



co 

eg 

t-H 

UO 

t-H 

t-H 

o 

co 

J. 

00 

t-H 

Qt 

CO 

i 

o 


1 

o 

1 

o 

1 

eg 

o 

i 

CO 

at 

1 

o 

a> 

o 

00 

t-H 


g 

oo 

3" 

o 

00 

3 

3 

3 

t-H 

CO 

t-H 

g. 

_g 

eg. 

lO 


g 

T 1 

3 

g 

g 

g 

g. 


+ 


co 

t-H 

CO 

o 

VO 

o 

co 

t-H 

+ 

CTt 



t-H 

co 

eg 


t-H 

t-H 




t-H 





w 


» 

* 





t-H 

o 

o 

o 

g 

3 

o 

t-H 

O 

eg 

o 


to 

g 

g 

g 

CO 

g 

g 

t-H 

00 

t-H 

g. 

g 

g 

ot 

o 

o 

o 

o 

eg 

o 

lO 

ID 

t-H 

o 

+ 

CO 




t-H 




e- 

t—H 





UJ 

to 


c 

> CO 
CD 

£ 


o 

o 

o 

o 

LO 

t-H 

o 

vq 

00 

eg 

co 

t-H 

to 

g 

g 

g 

si 

eg 

g 

t-H 

o 

g 

g 

g 

g 

55 

o 

o 

o 

i>* 

t-H 

00 

o 

t-H 

eg 

+ 

+ 

+ 



C 


CD 

2 


# 

# 

* 


I> 


00 

VO 




# 

* 

# 

» 


o 

to 

g 

g 

co 

o 

g 

g 

t-H 

g. 

l-H. 

g 

o 

co 


co 


o 


* 

^ 

_ _ _ 





CM 

LO 

CO 

CM 

t-H 

O 

00 

g 

O 

o 


t-H 




T—l 

235 

eg 

t-H 

+ 


) -c 

a UJ 
s Lfi 

§ 
- CD 
3 

J ^ 


00 


Ot 

00 


t-H 


g 

o 

ON 

t-H 




o 

g 

eg 

00 

CM 

g 

g 

3 

g 








g 


+ 


co 

t-H 

rH 

t-H 

+ 

eg 

283 

eg 


CO 

o 


cn 

<D 

‘o 

CD 

a 

to 


CD 
> 

O 

CL rrs 


CD 

-2 S 
3 -2 
C CL 
a i 
fs c 

O CD 
3 -O 
C"n 

” C 0 
CD cn tx 

3 C 

CD .2 CO 
jo 5 o 

-2 -2 
jo cl g 

cq < 


rj i— 

.8 

c _o 
E c b 

o c 
~n cd 

.2 O 
§ .2 
cl y 

c0 

CL 


cn 

3 

.5 

~a 

c 

a 

x 

CD 


a 

3 

25 s-. “> 

H--, CD 3 
j2 -a 

0 ft, o 

1 rj 

c/) 


3 


o 

CX 

E 

V3 

CO 

3 


-a 

0? 

f 0 v- 

£ "o 

-= a 

<t3 L- 
CL 2 


co 
3 

4 

O 

o 

3 
CO 
3 

T! 
"3 

_ E 

Ci) a 


EoSU 

CO £ 


a 
c 
a 

C 

E 

~ o 
8 S 
§ 13 

< S 

c P 
co i- 

O 3 

■g 8 

E ^ 

< 


CO 

O g 

^ o 

«n j-a 

f gs 

5 -2 cl 
o cd o 


£ E 


5 cj 

»rr CO 

0) c 

p— 1 .r* -t— * 


a) ~ 

S3 ft -2 

3 a 2 

S’? o 

S § £ 

s.f2 

£ a 
8 

co 


o 


CO 

3 

CL 

O 

a 

-c ■ 

cl 

co 

3 


£1 § 

■g £ 

Mg 

£ 


co 

3 

C 

a 

_u 

EC 

£ 

a 


CJ CO 

-si 

=3 CD 

-9 | 

cn 3 


Marbled Godwit 

Umosafedoa 799 (78.0) 916 (90.2) 85 (14.1) 152 (19.0)*** 885 (80.4) 100-1564 1067 (96 8) 73-2233 

Ruddy Turnstone 

Arenaria interpres 3 (0.8) 3 (1.2) 2 (1.7) + (0.1) 4 (2.0) 0-54 3 (1 2) 0-26 

Black Turnstone 





* 

# 

ov 

CO 

# 

* 








o. 

LO 

# 

« 

«- 





* 




LO 

CM 

T — t 

# 

# 




MD 




CO 

CM 

LO 





CM 


oo 

1 

1 

00 

OS 

LO 




T-H 

T-H 

1-H 

VO 

CM 

i-H 

to 

CM 

LO 

o 

CO 

1 

1 

1 

i-H 

CM 

| 

1 

1 

1 

1 

1 

o 

o 

o 

i-H 

CM 

O 

LO 

o 

o 

o 

o 

o 

t-H 

os 

c\T 

os' 

00 

CO 

crT 

o 

o 

i-H 

to 

o 

o 

c\i 

O 

CO 


o 

c\i 

o 

o 




VO 

r^. 

o 

CO 

i-H 

' ' 







t— H^ 

i-H 

CO 

' " 




LO 

+ 

CM 

i-H 

to 


LO 

CM 

LO 

o 

+ 

CO 




t-H 

r^ 

CO 

CO 







vO 

CM 


c*- 









t-H 


t-H 












CM 











to 









CM 

i-H 

Ch 








OS 

to 

1-H 

cm' 








CM 

OS 

00 

t-H 








OO 

CO 

CO 

1 

LO 




CM 


Os 

1 

1 

1 

CO 

1-H 

00 



t-H 

CO 

t-H 

o 

00 

t-H 

t-H 

CO 

CO 


o 

1 

1 

1 

[>• 

Or 

00 

o 

| 

1 

1 

1 

i-H 

o 

o 

t-H 

i-H 

t-H 

T— H 

o 

o 

o 

o 

Co 

t-H 

o 

CO 

i-H 

CO 


i — H 

i-H 

i-H 

o 

VD 

o 

t-H 

t-H 

CM 



CO 

CM 

o 

o 




CO 

Os 

to 

00 

i-H 

■ — 






' ' 

i-H 

1-H 

to 

*■ — " 








' ' 

~ — ' 






o* 

+ 

CM 

CM 

CO 

o 

o 


CO 

-1- 

o 

lO 



to 

00 

Or 

o 

00 








LO 

CM 

CM 









i-H 

T— H 

to 







CO 

o 

o 

LO 

LO 

* 

* 

to 

* 

«• 

* 

* 

to 

to 

o 

o 

CO 

o 

o 

00 

i-H 

oo 

|H 

to 

o 

o 

o 




1 — 1 

CO 

LO 

lO 









1-H 


t-H 





CM 

+ 

o 

CM 


os 

CO 

t-H. 

i-H 

o 

o 

i-H 



i-H 

o 

os 

CO 

i — i 







i-H 


i-H 

to 






os 

o 

o 

to 

t-H 

t-H 

CM 

O 

t-H 

o 

o 

si 

o 

o 

CM 

t-H 

CO 

qs 

CO 

qs 

t-H 

t-H 

CM 

i-H 

o 

o 

o 

00 

i-H 

o 

o 

69 

564 

564 

O 

CO 

OS 

CM 

55 

+ 

o 

o 


i-H 

LO 

o 

OS 

* 

CO 

LO 

If 

o 

if 

* 

Ot 

to 

o 

o 

t-H 

lO 

o 

o 

td 

LO 

os 

Os 

t-'i 

cti 

o 

o 




lO 

o 

LO 

Os 

’ " 

— 1 ' 






" — " 

i-H 


i-H 









*■ — 


' 





CO 

+ 

CM 

Os 

i-H 

Cs 

CM 

LO 


o 

+ 

CM 



LO 

i-H 

[>• 

O 

i-H 







LO 

00 

CM 

i-H 






LO 

r-H 

o 

t-H 


to 

os 

CM 

CM 

i-H 

o 

LO 

o 

i-H 

o 

CO 

Os 

LO 

i-H 

00 

CM 

CO 

tO 

CM 

o 

o. 

39 

1-H 

CM 

393 

Os 

i-H 

o 

t-H 

642 

o 

CO 

CO 

29 

00 

+ 

o 


CJ 

-C 

CL 

<D 

<L> 

O 

c 

CD 


c T2 
to 


So 


o 


O u, 'W c— 

£ -o o cx *5 


CD 

a 

'a 
-o 


3 Q) 


a _ </> cn 
-53 o 'P .£ 
^ 9^2 "5 

CL^ C3 

s 

DC tO 


W X-L. ■ — 
§ E 


TO ■ ; ; 

.2 to W 
*o F 


C -2 


o 

c 


c 

| U S U 3 

> _] Q 


.U5 

'IC 

"o 

CJ 


o 

a 


a 

G 

</5 

t/5 

3 <u 

E .£• 
P e 
to 
c 
o 

e 
£ 
o 
CJ 


~a 

o 

c 


O CD 

£> a 

o o 

c 

jg 

.£ "o 

' — CD 

CJ -a 

CD 

DC 


a 

~Q 

O 


<u 


G 
G 2 
p <3 

O K3 
-C D- 

‘L'S 

DC 


a 

‘C 

a 

o 


t/D 

3 

CL 

O 


O 

£ 


Total 6709 (450.8) 4062 (343.2)*** 5047 (570.4) 1877 (332,0)*** 11,756 (91 7.0) 4827-20,689 5939 (604.2) 1291-13,787’** 

Q Ba s ed on 29 counts in early winter and 28 counts in late winter. Asterisks specify significant differences between early and late winter; ‘P < 0.05; **P < 0.01- ***P < 0 001 + Mean 
abundance <0.5. See Figure 1 for count areas included in north and south bay subregions. ’ ’ 

fa SE, standard error. 


DISTRIBUTION AND ABUNDANCE OF WINTER SHOREBIRDS ON TOMALES BAY 


over half (53%) of all shorebirds in early winter and nearly a third (29%) in 
late winter. The second most abundant species was the Western Sandpiper, 
representing 13% in early winter and 20% in late winter (Table 1). The total 
number of shorebirds declined significantly from early to late winter in 
northern ( t = 4.67, df = 56, P < 0.001) and southern ( t = 4.80, df = 54, P 
< 0.001) Tomales Bay, as did several individual species (Table 1). In contrast 
to this pattern, the Marbled Godwit increased significantly during winter in 
southern Tomales Bay, and the Sanderling increased significantly in the 
northern bay and baywide (Table 1). 

Maximum annual abundance varied by a factor of 8.9 in the Dunlin, 8.0 
in the Marbled Godwit, 6.7 in the Killdeer, 6.2 in the Greater Yellowlegs, 4.4 
in the Sanderling, 4.2 in the Western Sandpiper, 3.5 in the Spotted 
Sandpiper, 3.4 in the Willet, 3.2 in the Black Turnstone, 2.3 in the Black- 
bellied Plover, and 4. 1 in all species combined. Maximum annual abundance 
varied substantially more in some species, by a factor of 17.8 in the 
dowitchers, 16.7 in the Semipalmated Plover, 15.5 in the Least Sandpiper, 
and 12.0 in the Pacific Golden-Plover, reflecting proportional changes in 
smaller populations and/or difficulties in detecting inconspicuous birds. 
Densities in most species varied significantly from year to year (Table 2). 

Winter shorebird densities differed significantly from count area to count 
area within both subregions of Tomales Bay (Table 2). Most species showed 
significant annual and intraseasonal shifts in distribution of densities among 
count areas, indicated by significant interaction terms in analyses of covari- 
ance (Table 2). Although average midwinter declines in count-area density, 
weighted equally for large and small areas, were significant only for the 
Dunlin and dowitchers (Table 2), several species decreased significantly in 
overall abundance on the basis of north bay, south bay, and baywide totals 
(Table 1). 

Northern Tomales Bay supported more shorebirds, overall, than southern 
Tomales Bay in early winter (t= 2.92, df=55, P< 0.05) and late winter (t = 
4.57, df = 55, P < 0.001). The Walker Creek (area 5) and Lagunitas Creek 
(area 8) deltas not only supported the greatest concentrations of shorebirds 
but also the highest densities for more shorebird species than other areas in 
Tomales Bay (Table 2, Figure 2). The northeast shoreline from Sand Point 
(area 3) to Vincent’s Landing south of Tom’s Point (area 4) also supported 
relatively high abundances of several species (Table 2, Figure 2). Shorebird 
use of seasonal wetlands at Lawson’s Meadow (area 2) and Giacomini 
Pasture (area 9) was highly variable; consequently, mean shorebird densities 
in these areas could not be distinguished clearly from those in other areas 
(Table 2). Greater Yellowlegs concentrated primarily in the extreme south 
end of Tomales Bay, while Sanderlings and Marbled Godwits concentrated 
primarily at the north end of the bay (Figure 2). The Pacific Golden-Plover, 
Snowy Plover, and Red Knot occurred exclusively in northern Tomales Bay. 

Cumulative seasonal rainfall was associated with significant decreases in 
the use of Tomales Bay by most shorebird species but significant increases by 
the Marbled Godwit and Sanderling (Figure 3). Bivariate plots suggested that 
25-30 cm of rain must fall in a season to trigger the declines of the Western 
Sandpiper, Least Sandpiper, and dowitchers (Figure 3). Alternatively, rain- 
fall correlations might be the result of intraseasonal changes in abundance 


152 


(0 

CQ 

<n 

Jl) 

<0 

E 

£ 

c 

(A 

"O 

s-. 

0) 

S-. 

o 

_c 

CO 

CO 

G 
» <— < 
s- 
05 

-4—1 

c 


C/5 

0) 


C/5 

c 

0) 

Q 

c 

o 


u 

Q) 

y= 

UJ 

E 

o 

TD 

C 

<5 

cc 

> 

S-. 

<0 

£ 

T3 

c 

(0 

S 

TD 

O 

■ 

5-i 

05 

CL 

"53 

c 

o 

1/5 

(0 

05 

m 

fO 

S-i 


tO 

0) 

S-i 

< 

C 

D 

O 

CJ 

H — 1 

o 

in 

-4-1 

CJ 

05 

M — 1 
M— i 

UJ 

CM 

JM 

3 

(2 


ON 


OO 


CO 

rt3 

e 

03 

■ 4—1 

c 

3 

o 

o 

S' 

ro 

-O 

-C 

t3 

o 

CO 


CN] 

O 


< 4 . 

* 

* ^ >" 



* 

> 

co 

CN 

II 

* 

We 

„ 

co 



< 

> 

C 

* CD 

. 'S ” 

>; 


co : 

-Q 

< 

«- 

> >2 
< < 

■» ~ 


, - co 

> 

: >; 

» •'co 

# 

CO 


< . - 

^ » 

O 

CJ 

> < 

5 ~< 

< 

* 

* 

> > 

z 

*■ * ~ 

* * 

* " * 

« 

* 

: co 

* * 

* * 

< 

< > 

< > 

> 

< < 

< < 


CO 


LO 


</5 
CO 
C55 
w 
fO 
- 4 — < 

c 

3 

O 

o 

S' 

CO 

-Q 

_c 

-c 

o 


CO 


Ccj 


0 

CO 

CO 

II 

g 

-o 

1 

CJ 


co 

CD 

O 

CD 

D< 

CO 


CQ CQ < §<§ CQ CJ CQ CQ CJ CQ 


§ gcgCCQCJ gCQ§gg | 

< CQ<<<<“3 <<§<< < 

< < < CQ < CQ < CQ < < CQ CQ CQ 


CJ 


CJ 


UJ 


"UJ 


W _l CETZr 


> 


>" 
c o 


> 

CO 


% 

* 

CO 

< 


> 

* 

* 

< 


^ > 
<£ co 
< < 

: ’ * " 

> > 
co < 

s':' 

% > 

— * 

* * 

> < 


Q 

CJ DQ < CQ CQ CQ CJ 
CQ 

CQCJCJCQCQCQCJCJCQCJ 

gcQ<ca“)<g 

^ c 5 CJ 

<^g<<CQ<<<< 

^ CQ < CQ < < < 

CJ ^ 

CQ<^<<CQCQCQD3^ 

< < < CQ < < < 

<^<CQCQCQCQ<^^ 

CJ 

cj cq < < 00 < aj 

CQOU§§§fflggS 

QCQ<CQ§<^ 

CQCJ§CQ<<CJ^^CQ 

tirg; @3 

©3 ©3©3 


> 

CO 

< 

* 

t • - 

* "co 

kf- 
^ * 

> > 

* # 

< < 


1—4 ■> 

s J 

22 CL 
CL "TJ 

"S -i 

-1 
r> 03 

1C CL 


co 

< 

* 

!L > 
Jr co 
^ < 

# *-< 

# * 
>- co 
< < 


* * 

# # 

< < 


o 

23 135 
CQ CO 


05 

E djj ~ 
2 :§ 


> 

co 


£ 


> 

■» 

* 

< 


3 

-o 

o 

CJ 

"3 

<15 


£ 


> 5 - 


>- CO 

< < 


< < 


05 

c 

o 

"S 

£ o’ 

pH = 

t— 1— 

<15 

iS "a 
-S § 

CQ CO 


> 

CO 

* 

. ^ 

> . - 
co : 

< 82 

- <c 

* 

* m. ** 

> : 
<c > 


•» * 

< < 


05 

.& n 
D. 05 
"O Q. 

= '5- 

03 -t-j 

co ^ 
c 

J— 

CD _ 

« IS 




CO 

< 


> 

co" 


ro 

co 


<55 


05 


C 

3 

Q 


CO 

< * 

* 

» 

s co 

* - CO 

^ < 

< s 

* 

# # 
> > 


< < 


cr> 

CD 

‘G 

0 ^ ty> 

CL 

CO 

i- -O 
CD CD 
C u, 

Q < 


CD 

CD 

E 

03 


O 

c 


rB 

■£ ^ 
S UO 
> QJ 
co CO 
03 CO 
CD 3 
L; to 
£0 _ 
•- wT 

■f- CD 
S CJ 
ro <—■ 

t J Er 


'5 CD 
CO^CD 

■a K 

CO 

C S 

CJ 

S Q 
c A 
JU 

< A 
C DQ 

CD 

ii 

J .. 

'T' ^ 

CD 

jt: 

c c 
'§ £ 

CD <D 
4=5 J 3 
ra 03 

TJ "2 
c C- 

ro ra 
GJ cn 

<0 A 

CD A 

i- 

s* 2 
-s 

03 

J- CD 

ra v> 

CL S 
<D £ tf5 

w c c 

■SSS 

t’d s 
- 

X 

0 ) CD 


E 

O 

u 


03 

c 

13 

o 

o 


CL O ^ 

e 

0 - I 

0 c 

CD cd ^ 

U5 X £ 

5 ^ cp 
*5 ^9 E 

p, -> CD 

CD 

-SJ -g 

CL B . 

v U (D 
"== iC CJ 

£ c J 2 

c C02 

-S' 2 .8 

1 g,E 


5 

o 


CD g) 

(3 ra 
^ » » 
o) S c 
Is 2 ra 

™ 5® 

S 2 e 

ra 3 g 

^_i i/> 

§ ^>05 

o 

Li _c e 

"“I 


o 

CD 


3 

O 


03 

~o 


1- c 

° 03 

Cn CD 

$.g:i 
- -c ■£• 

C O C 

ra c to 

4 sL 
g,^S ' s 

.25 ^ 05 
1 C 5 » 


c 

(0 

u 

»EC= 

c 

CO 


03 

j=: 

CD 

•a 

o 

u 

X- 

03 

ti 

_Q3 

_D 

03 


U 

03 


U 

fO 

CD 

-a 

_o 

X -4 

03 

CL 

& 

fa 

>-4 

X) 

,03 

LU 


§ 

S 

3 

1 — 

c 

§ 

0 ) 

03 


CD 

X 

E 

CD 

CJ 

03 

? 

03 

X 

E 

CD 

> 

o 

s 

X-. 

03 
"6 
§ 
>5 
V— ( 

ca 

CD 


cn 

a 

c 

c 

‘Ei 

03 

x 

03 

CJ 

.C 

"to 

CO 

fd 

T 3 


.2 ^4 

ra 2 
> o 

8 o 


8 Cl - 
.1 J 
ca i -1 

cj O 
o V 
co Cl, 

CO * 

K3 X 

S 0 
H o 

-o V 

TJ .2 

<D -t2 

.S E 
Six 


153 


DISTRIBUTION AND ABUNDANCE OF WINTER SHOREBIRDS ON TOMALES BAY 



3 4 5 6 7 8 

Count Location 


3456789 10 

Count Location 


Figure 2. Mean winter abundance of shorebirds, by count area, in northern (solid 
bars) and southern (striped bars) Tomales Bay, California, 1989-90 to 1998-99. 
Error bars indicate standard errors; numbers above bars indicate maximum abun- 
dance observed during the study period. 


unrelated to rainfall. When count date and year were included as covariates, 
residual (uncorrelated) effects of cumulative seasonal rainfall in the north bay 
were significant for the Black-bellied Plover (b = -1 ,31, P < 0.05), Willet (b 
= 2.56, P < 0.05), Dunlin (b = -27.30, P < 0.05), and (In) Least Sandpiper 
(b = -0.06, P < 0.001); residual effects in the south bay were significant for 
the (In) Least Sandpiper (b = -0.06, P < 0.05). Conversely, intraseasonal 
trends independent of cumulative seasonal rainfall and year were significant 
for the Willet (north bay: b = -1.74, P < 0.01; south bay: b = 1.01, P < 
0.05), Dunlin (north bay: b = -17.96, P < 0.01; south bay: b = -22.25; P 
< 0.01), and (In) Black Turnstone (south bay: b = -0.03; P = 0.01). 


154 


w 


DISTRIBUTION AND ABUNDANCE OF WINTER SHOREBIRDS ON TOMALES BAY 


® 140 
c 120 
-o 100 

E 80 

< 60 

| 40 

S 20 

0 

123456789 10 123456789 10 



400 
g 350 



a) 

u 

c 

ro 

■o 

c 

D 

■Q 

< 

C 

ro 


5 
4 
3 
2 
1 
0 

123456789 10 


Red Knot 


19 


0 o ■ 0 0 0 0 0 0 



123456789 10 



123456789 10 


Count Location 


Count Location 


Figure 2 ( continued ). 


After the linear effects of cumulative seasonal rainfall and date are 
accounted for, residual annual trends were not significant (P > 0.05), except 
in five cases. In the south bay, significant linear trends independent of 
cumulative seasonal rainfall and date suggested 10-year declines in the 
Black-bellied Plover ( b = -5.62; P < 0.001), (In) Semipalmated Plover (b = 
-0.32; P<0.05), (In) Least Sandpiper (b = -0.32; P<0.05), and Dunlin (b = 
-223.21; P < 0.001), and increases in the Willet (b = 14.43, P < 0.01) and 
Marbled Godwit (b = 10.68, P < 0.05). In northern Tomales Bay, significant 
10-year declines independent of cumulative rainfall and date were evident in 
the Willet (b = -13.10, P < 0.05) and Dunlin (b = -330.26, P < 0.01). 
Although these changes in shorebird use were independent of cumulative 
seasonal rainfall, they might have reflected escalating effects of successive 
increases in annual rainfall (b = 5.3 cm/year, P < 0.05; Figure 4). The 10- 
year duration of this study was too short to distinguish between the influence 
of annual rainfall trend and possible underlying (correlated) population 


155 


DISTRIBUTION AND ABUNDANCE OF WINTER SHOREBIRDS ON TOMALES BAY 



123456789 10 123456789 10 



123456789 10 123456789 10 


d) 4000 

0 

<§ 3000 ^ 

1 800 

§ 600 

c 
<0 
CD 


All 

Shorebirds^ 

3041 



1 2 3 4 5 

Count Location 

Figure 2 ( continued ). 


Count Location 


trends. Annual trends in maximum (fall) abundance were nonsignificant for 
all species (n = 10, P > 0.05), after differences associated with cumulative 
rainfall through November and in the previous year were accounted for. 

DISCUSSION 

Winter Abundance and Distribution 

My results confirm Tomales Bay as a wetland of “regional” importance on 
the basis of its supporting more than 20,000 shorebirds, a criterion for 
inclusion in the WHSRN (Harrington and Perry 1995, Page and Shuford 
2000). Relative abundances of wintering shorebird species were similar to 
those reported for other Point Reyes area wetlands (Page et al. 1983, Page 
et al. 1992). Concurrent counts conducted in 1989-90 in other nearby 
wetlands indicated that Tomales Bay supported approximately a third of the 


156 


DISTRIBUTION AND ABUNDANCE OF WINTER SHOREBIRDS ON TOMALES BAY 





Cumulative Cumulative Cumulative 

seasonal rainfall (cm) seasonal rainfall (cm) seasonal rainfall (cm) 


Figure 3, Bivariate plots of baywide shorebird abundance (In) on cumulative seasonal 
rainfall in Tomales Bay, Lines represent LOWESS trends, with smoothing parameter 
/ = 0,6. Slope of linear regression ( b ) is indicated in each plot: *P < 0.05, **P < 0.01, 
***P < 0.001. Between northern and southern Tomales Bay linear slopes did not 
differ significantly (P < 0,05). Abundances of the Marbled Godwit and Black 
Turnstone are not ln-transformed. 


157 


DISTRIBUTION AND ABUNDANCE OF WINTER SHOREBIRDS ON TOMALES BAY 



Figure 4. Means and standard errors of cumulative seasonal rainfall measured on 
shorebird count days on Tomales Bay, from 1989-90 to 1998-99. Dashed line 
indicates total annual rainfall (July-June) across years. Horizontal dashed line indi- 
cates long-term average annual rainfall. 


wintering shorebirds in the greater Point Reyes area, which includes Bolinas 
Lagoon, Drake’s and Limantour esteros, Estero Americano, and Bodega 
Harbor (this study and unpublished data from P. Connors and G. Page). 

The only other baywide counts of shorebirds in Tomales Bay are aerial 
surveys conducted from August 1968 to April 1970 (Smail 1972). Un- 
known differences in detectability, however, preclude comparisons with 
these aerial counts, which detected far fewer shorebirds than my study. The 
aerial counts did indicate that Tomales Bay supported more shorebirds than 
other Point Reyes area wetlands, including Drake’s and Limantour esteros 
and Bolinas Lagoon (Smail 1972). Jurek (1974) conducted counts at the 
Walker Creek delta from 1969-70 to 1973-74 (area 5, Figure 1) and 
reported abundances that do not differ significantly from those I report; 
however, population trends could be obscured by substantial variability in 
shorebird use from year to year and count area to count area (Table 2). 

Among several species, shorebirds wintering at the northern and southern 
ends of Tomales Bay appear to represent different groups (this study, Kelly 
2000). The apparent separation of shorebirds into local wintering groups in 
the northern and southern portions of Tomales Bay is consistent with the 
spatial scale of winter site fidelity demonstrated by other studies along the 
California coast (Kelly and Cogswell 1979, Warnock and Takekawa 1996). 

Although the composition of shorebird species in the northern and 
southern portions of the bay was similar, habitat conditions differed substan- 


158 


DISTRIBUTION AND ABUNDANCE OF WINTER SHOREBIRDS ON TOMALES BAY 


tially. Habitat differences resulted from spatial and temporal hydrographic 
gradients in the estuary that affect salinity (lower and more variable in the 
south), sediment size (finer in the south), turbidity (greater and more variable 
in the south), and flood disturbance (more frequent and intense in the south; 
Daetwyler 1966, Hollibaugh et al. 1988). Such differences are likely to 
influence the composition, abundance, or availability of invertebrate prey 
(Johnson 1971, Wolff 1969, 1983) and, consequently, may be associated 
with spatial or temporal patterns of shorebird use. For example, Marbled 
Godwits concentrated in northern Tomales Bay, where substrates are 
composed predominantly of well-sorted medium and fine sand, but not in 
the southern portion of the bay, where substrates are composed of finer 
particles (Daetwyler 1966). Within the Point Reyes/Bodega area, godwits 
also concentrate in nearby Bodega Harbor (Page et al. 1983), where 
sediments are similar to substrates in northern but not southern Tomales Bay 
(Standing et al. 1975). 

Midwinter influxes of the Sanderling and Marbled Godwit were signifi- 
cantly associated with cumulative seasonal rainfall. These species occur 
commonly on nearby beaches along the outer coast and are known to switch 
habitats with the tides (Shuford et al. 1989, Connors et al. 1981, Colwell 
and Sundeen 2000). My study suggested an intraseasonal influx of these 
species into estuarine tide flats, correlated with cumulative rainfall and 
independent of tides. Connors et al (1981) linked tide-associated move- 
ments between habitats with enhanced foraging profitability and found that 
variation in the timing of these movements varies seasonally to enhance 
profitability. Thus midwinter influxes of Sanderlings and Marbled Godwits 
into Tomales Bay might reflect an intraseasonal increase in the profitability 
of foraging in tide flats over foraging on beaches. Alternatively, movements 
into Tomales Bay may result from a thermal advantage of refuge from 
harsher weather along the outer coast. 

The significant midwinter declines in several species using Tomales Bay 
suggest local or regional movements to alternative wintering areas. At 
Humboldt Bay Colwell (1994) also reported midwinter declines in several 
shorebirds. Nonmigratory movements of wintering shorebirds, ranging from 
use of local alternative habitats to interestuarine movements to large-scale 
flights from the coast inland, have been documented at several California 
estuaries, but in different areas intraseasonal patterns are not necessarily 
similar (Page et al. 1979, Ruiz et al 1989, Shuford et al. 1989, Warnock et 
al. 1995). 

Four of six principal wetlands in the Point Reyes/Bodega area are 
considered, independently, to have potential regional importance by the 
WHSRN (Harrington and Perry 1995, Page et al. 1999). These are Bolinas 
Lagoon (Page et al. 1979), Drake’s and Limantour esteros (Page et al. 
1983, Shuford et al. 1989), Tomales Bay (this study), and Bodega Harbor 
(Page et al. 1983, Ruiz 1987). The closeness of these areas, along with 
other locally important wetlands, within approximately 60 km of coast line 
indicates substantial importance of this area to wintering and migrating 
shorebirds. Movements of wintering shorebirds among wetlands in the Point 
Reyes/Bodega area have been suggested by my results and by those of 
others (Page et al. 1979, Myers et al. 1980, Shuford et al. 1989, Ruiz et al. 


159 


DISTRIBUTION AND ABUNDANCE OF WINTER SHOREBIRDS ON TOMALES BAY 


1989). Although strong site fidelity often distinguishes wintering groups of 
coastal shorebirds (War nock and Takekawa 1996, Kelly 2000), the ability to 
respond with intraseasonal or episodic movements among wetlands may be 
crucial to maximizing winter survival in the face of unpredictable changes in 
foraging conditions (Connors et al. 1981, Myers 1980, Ruiz et al. 1989, 
Colwell 1993, Warnock et al. 1995). Thus, regional complexes of wetlands 
such as those in the Point Reyes/Bodega area may need broad protection 
and cooperation by the numerous agencies responsible for their manage- 
ment (Laubhan and Fredrickson 1993). 

Page et al. (1992) reported on the use of Tomales Bay by migrating 
shorebirds in the context of other sites in the Pacific Flyway. Further 
information on shorebird use during spring and fall is needed for the 
importance of Tomales Bay to migrating shorebirds to be understood fully. 
The spatial scale and flexibility of stopover-site selection by migrating 
shorebirds within systems of coastal wetlands is poorly understood but might 
involve short-flight hopping (Iverson et al. 1996) or local movements related 
to changing habitat conditions, as found at inland wetlands (Skagen and 
Knopf 1993, 1994). If so, Point Reyes/Bodega area wetlands may provide 
extended stopover support for migrating shorebirds. 

Conservation Implications 

Shorebirds tend to concentrate where the density (Goss-Custard et al. 
1977, Bryant 1979) or availability (Recher 1966, Goss-Custard 1984) of 
invertebrate prey is greatest. Midwinter declines in several shorebird species 
observed in this study may be associated with the adverse effects of heavy 
freshwater runoff and associated sediment deposition on prey (Wolff 1983, 
Holland 1985, Anima et al. 1988, Nordby and Zedler 1991). Because of 
water-residence times as long as four months, prey in southern Tomales Bay 
may be particularly vulnerable to the effects of toxic spills or anthropogenic 
eutrophication (Smith and Hollibaugh 1989, 1997). Midwinter changes in 
shorebird use of key feeding areas at Lagunitas and Walker creeks are 
consistent with a hypothesis that deposition of sandy sediment during heavy 
runoff reduces the suitability of some areas to foraging shorebirds. Core 
samples collected in these areas 4-7 weeks after a major flood in 1982 
revealed a surface layer of sandy material that was several centimeters thick 
and virtually devoid of invertebrates (Anima et al. 1988). Gerstenberg 
(1979) reported reduced shorebird use in Humboldt Bay after siltation of tide 
flats during heavy rains. Quammen (1982) found reduced feeding success 
and abandonment of preferred feeding areas by Dunlins, Western Sandpip- 
ers, and dowitchers when tidal feeding areas were treated with a thin layer of 
sand. Periods of increased sedimentation in Tomales Bay have been associ- 
ated with logging and intensive farming in the watershed (Rooney and Smith 
1999) and could also result from development, overgrazing, or other 
disturbance of upland soils (Storm 1972). 

Daily, tidal, or intraseasonal movements of shorebirds to alternative sites 
in seasonal wetlands and upland fields have been reported in several areas 
along the central and northern coast of California (Gerstenberg 1979, Ruiz 
etal. 1989, Rameretal. 1991, Colwell 1993). Near Tomales Bay, however, 
alternative foraging areas are relatively limited, and use of alternative 


160 


DISTRIBUTION AND ABUNDANCE OF WINTER SHOREBIRDS ON TOMALES BAY 


habitats was generally limited to seasonal wetlands at Lawson’s Meadow and 
Giacomini Pasture (Figure 1). Breaching levees that isolate the Giacomini 
Pasture would restore approximately 150 ha of tidal marsh and at least 36 
ha of tidal flats (elevation less than 1.2 m above MLLW) for shorebirds at the 
southern end of the bay (Philip Williams and Associates et al. 1993). The 
protection of routinely used seasonal wetlands at Lawson’s Meadow adja- 
cent to the north end of the bay may be especially important. During heavy 
rainfall and flooding, some shorebirds flew to seasonal wetlands 5-15 km 
northeast of the bay (pers. obs.). 

Direct disturbance by people or dogs can have detrimental effects on the 
continuing use of foraging (Burger 1986, Burger and Gochfield 1991) and 
roosting (Burger 1981, Kirby et al. 1993) areas by wintering shorebirds. In 
Tomales Bay, aggregations of roosting and feeding shorebirds along the 
shoreline below sand dunes at Brazil Beach, and on offshore tide flats north 
of Tom’s Point (areas 3 and 4, Figure 1), were frequently disturbed by people 
from an adjacent 1000-site campground (Shannon and Associates 1998), 
who come for clamming or other activities (pers. obs.; Page and Shuford 
2000). Dune beaches near the mouth of Tomales Bay may provide critical 
winter habitat for the threatened Western Snowy Plover ( Charadrius 
alexandrinus niuosus; U.S. Fish and Wildlife Service 1995). Direct distur- 
bance of roosting and feeding Snowy Plovers by recreational users in this 
area was evident during winter; during spring and summer such disturbance 
may prevent nesting by this species in otherwise suitable habitat (Page and 
Shuford 2000). 

Gulls may interfere with shorebird foraging (Thompson and Lendrem 
1985, Warnock 1989, Amat and Aguilera 1990). During winter low tides at 
the Walker Creek delta (area 5, Figure 1), large concentrations of gulls were 
inversely correlated with shorebird use of exposed tidal flats, suggesting that 
gulls displaced shorebirds from otherwise suitable foraging habitat (Kelly et 
al. 1996). The gulls arrived daily from .a garbage dump approximately 20 km 
to the northeast. Possible indirect adverse effects of local landfills on 
shorebird use of estuaries should be investigated further, and the use of such 
landfills by wintering gulls could be reduced. 

Currently, leases for commercial oyster culture occupy 208 ha of Tomales 
Bay, but the limit placed on mariculture leases by the Marin County local 
coastal plan is 365 ha. (Tom Moore, Calif. Dept. Fish and Game, pers. 
comm.). Shorebirds’ use of open tide flats decreases when these are 
developed for mariculture (Kelly et al. 1996). The two most abundant 
shorebirds in Tomales Bay, the Dunlin and Western Sandpiper, avoided 
areas devoted to mariculture significantly, although Willets were attracted to 
oyster plots. Expansion of mariculture, or its redistribution to be concen- 
trated more heavily on particular tidal strata, could further reduce foraging 
opportunities and possibly abundances of wintering shorebirds. The occa- 
sional use by roosting Dunlin and Western Sandpipers of floating oyster 
culture bags in subtidal areas during high tides (pers. obs.) suggests possible 
benefits to shorebirds. 

The invasive nonindigenous European green crab (Ca rein us maenas ) has 
become established in Tomales Bay and can alter the abundances of benthic 
invertebrate prey dramatically, although the consequences for shorebirds 


161 


DISTRIBUTION AND ABUNDANCE OF WINTER SHOREBIRDS ON TOMALES BAY 


are unknown (Grosholz and Ruiz 1995, Grosholz et al. 2000). The closeness 
of San Francisco Bay implies a high potential for invasion by nonindigenous 
cordgrass (Spartina alterni flora), which is established there and known to 
affect shorebirds adversely (Goss-Custard and Moser 1988, Daehler and 
Strong 1996). These concerns, along with those related to watershed 
effects, limited availability of alternative habitats, and human disturbance 
associated with recreation, mariculture, and garbage dumps, indicate broad 
management needs for protecting shorebird foraging areas in Tomales Bay. 

ACKNOWLEDGMENTS 

The following observers provided expertise in the field: Russ Agnew, Sarah Allen, 
Bob Baez, Norah & Hugh Bain, Gordon Bennett, Gay Bishop, Ellen Blustein, John 
W. Boyd, Brian Bullick, Ken Burton, Yvonne Chan, Lisa Cohen, Rigdon Currie, Ann 
& Eric Davis, Mark Dean, Linda Devere, Jim Devore, John Dillon, Caroline Dutton, 
Lew & Marilyn Edmondson, Steve Engel, Gayanne Enquist, Katie Etienne, Jules 
Evens, Gary A. Falxa, Katie Fehring, Grant & Ginny Fletcher, Carol Fraker, Dan 
Froehlich, Nicole Gallagher, Robert Gleason, Quinton Goodrich, Gayle Greeley, Sue 
Gueres, Catherine M. Hickey, Terry Horrigan, Lisa Hug, Maggie Hynes, Jeri 
Jacobson, Lynnette Kahn, Mary Ellen King, Richard Kirschman, Carol Kuelper, 
Darlene Lam, Sarah Lynch, Flora Maclise, Aspen Mayers, John Me Donagh, Dan 
Murphy, Terry Nordbye, David Press, Jim Rappold, Ellen Sabine, Dave Schurr, Craig 
Scott, Anne Spencer, Rich Stallcup, Judy Temko, Stephen Thai, Janet Thiessen, 
Wayne Thompson, Gil Thomson, Forest Tomlinson, Tanis Walters, Diane Williams, 
David Wimpfheimer, and Chris Wood. Susan Kelly, Sarah Tappen, and Katie Etienne 
provided valuable assistance in the coordination of shorebird counts and management 
of data. Peter Connors, Gary Page, David Shuford, and Rich Stallcup provided 
valuable advice in designing the monitoring program. Thomas R. Famula, Robert E. 
Gill, William J. Hamilton III, Lewis W. Oring, and Wesley W. Weathers provided 
valuable review of the manuscript. This paper is a contribution of Audubon Canyon 
Ranch. 

LITERATURE CITED 

Amat, J. A., and Aguilera, E. 1990. Tactics of Black-headed Gulls robbing egrets and 
waders. Anim. Behav. 39:70-77. 

Anima, R. J., Bick, J. L., and Clifton, H. E. 1988. Sedimentologic consequences of 
the storm in Tomales Bay, in Landslides, Floods, and Marine Effects of the Storm 
of January 3-5, 1982, in the San Francisco Bay Region, California (S. D. Ellen 
and G. F. Waczorek, eds.), U.S. Geol. Surv., Menlo Park, CA. 

Bryant, D. M. 1979. Effects of prey density and site character on estuary usage by 
overwintering waders (Charadrii). Estuarine and Coastal Mar. Sci. 9:369-384. 

Burger, J. 1981. The effect of human activity on birds at a coastal bay. Biol. Conserv. 
21:231-241. 

Burger, J. 1986. The effect of human activity on shorebirds in two coastal bays in 
northeastern United States. Env. Conserv. 13:123-130. 

Burger, J., and Gochfeld, M. 1991, Human activity influence and diurnal and 
nocturnal foraging of Sanderlings ( Calidris alba). Condor 93:259-265. 

Burger, J., Howe, M., Hahn, A., Caldwell, D., and Chase, J. 1977. Effects of tide 
cycles on habitat selection and habitat partitioning by migrating shorebirds. Auk 
94:743-758. 


162 


DISTRIBUTION AND ABUNDANCE OF WINTER SHOREB1RDS ON TOMALES BAY 


Chambers, J. M., Cleveland, W. S., Kleiner, B., and Tukey, P. A. 1983. Graphical 
Methods for Data Analysis. Wadsworth & Brooks/Cole, Pacific Grove, CA. 

Cleveland, W. S. 1979. Robust locally weighted regressions and smoothing 
scatterplots. J. Am. Stat. Assoc. 74:829-836. 

Colwell, M. A. 1993. Shorebird community patterns in a seasonally dynamic estuary. 
Condor 95:104-114. 

Colwell, M. A. 1994. Shorebirds of Humboldt Bay, California: Abundance estimates 
and conservation implications. W. Birds 25:137-145. 

Colwell, M. A., and Sundeen, K. D. 2000. Shorebird distribution on ocean beaches 
of northern California. J. Field Ornithol. 71:1-15. 

Connors, P. G., and Maron, J. L. 1989. Estero Americano bird population study. 
Long-term Detailed Wastewater Reclamation Studies, Santa Rosa Subregional 
Water Reclamation System. Tech, memorandum to the city of Santa Rosa, 100 
Santa Rosa Ave., Santa Rosa, CA 95403. 

Connors, P. G., Myers, J. P., Connors, C. S. W., and Pitelka, F. A. 1981. Interhabitat 
movements by Sanderlings in relation to foraging profitability and the tidal cycle. 
Auk 98:49-64. 

Daehler, C. C., and Strong, D. R. 1996. Status, prediction and prevention of 
introduced cordgrass Spartina spp. invasions in Pacific estuaries, USA. Biol. 
Cons. 78:51-58. 

Daetwyler, C. C. 1966. Marine geology of Tomales Bay, central California. Pac. Mar. 
Sta. Res. Rep. 6:1-169. 

Galloway, A. J. 1977. Geology of the Point Reyes Peninsula, Marin County. Calif. 
Div. Mines Geol. Bull. 202. 

Gerstenberg, R. H. 1979. Habitat utilization by wintering and migrating shorebirds on 
Humboldt Bay, California. Studies Avian Biol. 2:33-40. 

Goss-Custard, J, D. 1977. Optimal foraging and size selection by worms in Red- 
shank, Tringa totanus , in the field. Anim. Behav. 25:10-29. 

Goss-Custard, J. D. 1984 Intake rates and food supply in migrating and wintering 
shorebirds, in Behavior of Marine Animals, vol, 6, Shorebirds: Migration and 
Foraging Behavior {J. Burger and B. L. Olla, eds.), pp. 233-270. Plenum, New 
York' 

Goss-Custard, J. D., and Moser, M. E. 1988. Rates of change in the numbers of 
Dunlin, Calidris alpina, wintering in British estuaries in relation to the spread of 
Spartina anglica. J. Appl. Ecol. 25:95-109. 

Grosholz, E. D., and Ruiz, G. M. 1995. Spread and potential impact of the recently 
introduced European green crab, Carcinus maenas , in central California. Mar. 
Bioi. (Berlin) 122:239-247. 

Grosholz, E. D., Ruiz, G. M., Dean, C. A., Shirley, K. A., Maron, J. L., and Connors, 
P. G. 2000. The impacts of a nonindigenous marine predator in a California bay. 
Ecology 81:1206-1224. 

Harrington, B., and Perry, E. 1995. Important shorebird staging sites meeting 
Western Shorebird Reserve Network criteria in the United States. Report of the 
Western Hemisphere Shorebird Reserve Network, Manomet Observatory, P. O. 
Box 1770, Manomet, MA 02345. 

Holland, A. F. 1985. Long-term variation of macrobenthos in a mesohaline region of 
Chesapeake Bay Estuaries 8:93-113. 


163 


DISTRIBUTION AND ABUNDANCE OF WINTER SHOREBIRDS ON TOMALES BAY 


Hollibaugh, J. T., Cole, B. E., Dollar, S. J., Hager, S. W., Vink, S. M., Kimmerer, W. 
J., Obrebski, S., Smith, S. V., Valentino, M., and Walsh, W. W. 1988. Tomales 
Bay, California: A macrocosm for examining biogeochemical coupling at the 
land-sea interface. Eos 69:843-845. 

Holway, D. A. 1990. Patterns of winter shorebird occurrence in a San Francisco Bay 
marsh. W. Birds 21:51-64. 

Iverson, G. C., Warnock, S. E., Butler, R. W., Bishop, M. A., and Warnock, N. 1996. 
Spring migration of Western Sandpipers along the Pacific coast of North 
America: A telemetry study. Condor 98:10-21. 

Johnson, R. G. 1971. Animal-sediment relations in shallow-water benthic communi- 
ties. Mar. Geol. 11:93-104. 

Jurek, R. M. 1974, California shorebird survey 1969-1974. Spec. Wildlife Invest. 
Rep. Pro. W-54-R, Job III-l. Calif. Dept. Fish and Game, 1416 Ninth St., 
Sacramento, CA 95814, 

Kelly, J. P. 2000. Foraging distribution and energy balance in wintering Dunlin. Ph. 
D. dissertation, Univ. of Calif., Davis. 

Kelly, J. P., Evens, J. G., Stallcup, R. W., and Wimpfheimer, D. 1996. Effects of 
aquaculture on habitat use by wintering shorebirds in Tomales Bay, California. 
Calif. Fish & Game 82:160-174. 

Kelly, J. P., and Tappen, S. L. 1998. Distribution, abundance, and implications for 
conservation of winter waterbirds on Tomales Bay, California. W. Birds 29:103- 
120 . 

Kelly, P. R., and Cogswell, H. L. 1979. Movements and habitat use by wintering 
populations of Willets and Marbled Godwits. Studies Avian Biol. 2:69-82. 

Kirby, J. S., Clee, C., and Seager, V. 1993. Impact and extent of recreational 
disturbance to wader roosts on the Dee estuary: Some preliminary results. Wader 
Study Group Bull. 68:53-58. 

Laubhan, M. K., and Fredrickson, L. H. 1993. Integrated wetland management: 
Concepts and opportunities, in Trans. N. Am. Wildlife and Natl. Resources Conf. 
58 (R. E. McCabe and K. A. Glidden, eds.), pp. 323-334. Wildlife Mgmt. Inst., 
Washington, D.C. 20005. 

Myers, J. P. 1980. Sanderlings Calidris alba at Bodega Bay: Facts, inferences and 
shameless speculations. Wader Study Group Bull. 30:26-32. 

National Oceanic and Atmospheric Administration. 1987. Gulf of the Farallones 
National Marine Sanctuary management plan (J. Dobbin Associates, Inc., eds.). 
U. S. Dept. Commerce, NOAA, Marine and Estuarine Mgmt. Div., Washington, 
D. C . 20235, 

Neubacher, D., Dell’Osso, J., Livingston, D., and Murray, R. 1995. Tomales Bay/ 
Bodega Bay watershed boundary study. U.S. Govt. Printing Office 1995-685- 
430, Point Reyes National Seashore, Point Reyes, CA 94956. 

Nordby, C. S., and Zedler, J. B. 1991. Responses of fish and macrobenthic 
assemblages to hydrologic disturbances in Tijuana estuary and Los Penasquitos 
Lagoon, California. Estuaries 14:80-93. 

Page, G. W., and Shuford, D. 2000. Southern Pacific Coast Regional Shorebird Plan, 
version 1.0. U. S. Shorebird Planning Council, U. S. Shorebird Conservation Plan. 
Point Reyes Bird Observatory, 4990 Shoreline Hwy., Stinson Beach, CA 94970. 

Page, G. W., Stenzel, L. E., and Kjelmyr, J. E. 1999. Overview of shorebird 
abundance and distribution in wetlands of the Pacific coast of the contiguous 
United States. Condor 101:461-471. 


164 


DISTRIBUTION AND ABUNDANCE OF WINTER SHOREBIRDS ON TOMALES BAY 


Page, G. W., Shuford, W. D., Evens, J. G., and Stenzel, L. E. 1983. The distribution 
and abundance of aquatic birds in wetlands of the Point Reyes to Bodega area. 
Report to the Point Reyes-Farallones National Marine Sanctuary. Point Reyes 
Bird Observatory, 4990 Shoreline Hwy., Stinson Beach, CA 94970 

Page, G. W., Shuford, W. D., Kjelmyr, J. E., and Stenzel, L. E. 1992. Shorebird 
numbers in wetlands of the Pacific Flyway: A summary of counts from April 1988 
to January 1992. Point Reyes Bird Observatory, 4990 Shoreline Hwy., Stinson 
Beach, CA 94970. 

Page, G. W., Stenzel, L. E., and Wolfe, C. M. 1979. Aspects of the occurrence of 
shorebirds on a central California estuary. Studies Avian Biol. 2:15-32. 

Philip Williams & Associates, Ltd., Wetlands Research Associates, Strong Associates, 
and Butler, L. J. 1993. An evaluation of the feasibility of wetland restoration on 
the Giacomini Ranch, Marin County. Philip Williams & Associates, Pier 35, The 
Embarcadero, San Francisco, CA 94113. 

Quammen, M. L. 1982. Influence of subtle substrate differences on feeding by 
shorebirds on intertidal mudflats. Mar. Biol. (Berlin) 71:339-343. 

Ramer, B. A., Page, G. W., and Yoklavitch, M. M. 1991 . Seasonal abundance, habitat 
use, and diet of shorebirds in Elkhorn Slough, California. W. Birds 22:157-174. 

Recher, H. F. 1966. Some aspects of the ecology of migrant shorebirds. Ecology 
47:393-407. 

Rooney, J. J., and Smith, S. V. 1999. Watershed land use and bay sedimentation. J. 
Coastal Res. 15:478-485. 

Ruiz, G. M. 1987. Interactions among shorebird, crab, and their invertebrate prey 
populations. Ph. D. dissertation, Univ. of Calif., Berkeley. 

Ruiz, G. M,, Connors, P. G., Griffin, S. E., and Pitelka, F. A. 1989. Structure of a 
wintering Dunlin population. Condor 91:562-570. 

Shannon and Associates. 1998. Lawson’s Landing Master Plan. Shannon and 
Associates, P. O. Box 322, Bodega Bay, CA 94923. 

Shuford, W. D., Page, G. W., Evens, J. G., and Stenzel, L. E. 1989. Seasonal 
abundance of waterbirds at Point Reyes: A coastal California perspective. W. 
Birds 20:137-265. 

Skagen, S. K., and Knopf, F. L. 1993. Toward conservation of midcontinental 
shorebird migrations. Cons. Biol. 7:533-541. 

Skagen, S. K., and Knopf, F. L. 1994. Migrating shorebirds and habitat dynamics at 
a prairie wetland complex. Wilson Bull. 106:91-105. 

Smail, J. 1972. The birds of Tomales Bay, in Tomales Bay Environmental Study: 
Compendium of Reports (R. Corwin, ed.), pp. 124-134. Conservation Founda- 
tion, Washington, D. C. 

Smith, S. V., and Hollibaugh, J. T. 1989. Carbon-controlled nitrogen cycling in a 
marine “macrocosm”: and ecosystem-scale model for managing cultural eutrophi- 
cation. Mar. Ecol. Prog. Ser. 52:103-109. 

Smith, S. V., and Hollibaugh, J. T. 1997. Annual cycle and interannual variability of 
ecosystem metabolism in a temperate climate embayment. Ecol. Monogr. 
67:509-533. 

Standing, J., Browning, B., and Speth, J. W. 1975. The natural resources of Bodega 
Harbor. Calif. Dept. Fish and Game Coastal Wetland Ser. 1 1 . 

Storm, D. R. 1972. Hydrology and water quality aspects of the Tomales Bay Study, 
in Tomales Bay Environmental Study: Compendium of Reports (R. Corwin, ed.), 
pp. 84-101. Conservation Foundation, Washington, D.C. 


165 


DISTRIBUTION AND ABUNDANCE OF WINTER SHOREBIRDS ON TOMALES BAY 


Thompson, D. B. A., and Lendrem, D. W. 1985. Gulls and plovers: Host vigilance. 
Behaviour 33:1318-1324. 

U. S. Fish and Wildlife Service. 1995. Proposed designation of critical habitat for the 
Pacific coast population of the Western Snowy Plover. Fed. Reg. 60:11767- 
11809. 

Warnock, N. 1989. Piracy by Ring-billed Gulls on Dunlin. Wilson Bull. 101:96-97. 

Warnock, N., Page, G. W., and Stenzel, L. E. 1995. Non-migratory movements of 
Dunlins on their California wintering grounds. Wilson Bull. 107:131-139. 

Warnock, S. E., and Takekawa, J. Y. 1995. Habitat preferences of wintering 
shorebirds in a temporally changing environment: Western Sandpipers in the 
San Francisco Bay estuary. Auk 112:920-930. 

Warnock, S. E., and Takekawa, J. Y. 1996. Wintering site fidelity and movement 
patterns of Western Sandpipers Calidris mauri in the San Francisco Bay 
estuary. Ibis 138:160-167. 

Wolff, W. J. 1969. Distribution of non-breeding waders in an estuarine area in relation 
to the distribution of their food organisms. Ardea 57:1-27. 

Wolff, W. J. 1983. Estuarine benthos, in Ecosystems of the World 26: Estuaries and 
Enclosed Seas (B. H. Ketchum, ed.), pp. 151-182. Elsevier Scientific, 
Amsterdam. 


Accepted 2 June 2001 



Western Sandpipers 


Sketch by George C. West 


A TARGETED MIST NET CAPTURE TECHNIQUE 
FOR THE WILLOW FLYCATCHER 


MARK K. SOGGE, JENNIFER C. OWEN, EBEN H. PAXTON, SUZANNE M. 
LANGRIDGE, and THOMAS J, KORONKIEWICZ, U.S. Geological Sun/ey, Forest 
and Rangeland Ecosystem Science Center, Colorado Plateau Field Station, P, O. Box 
5614, Flagstaff, Arizona 86011 (current address of Owen Department of Biological 
Sciences, University of Southern Mississippi, Hattiesburg, Mississippi 39406-5018; 
current address of Langridge Department of Environmental Studies, University of 
California, Santa Cruz, California 95064) 

ABSTRACT: We developed a targeted mist-netting technique designed to capture 
Willow Hycatchers (Empidonax traillii ) at their breeding sites. The technique uses a 
variety of conspecific vocalizations to lure territorial flycatchers into mist nets. Songs 
and calls are broadcast from a portable CD player with speakers placed on both sides 
of the net. By playing vocalizations commonly heard during territorial interactions, 
and switching the sound output from one speaker to the other, flycatchers are readily 
drawn into the nets. This capture technique is highly effective, captures birds of both 
sexes, and worked at sites throughout the Willow Flycatcher’s breeding range and on 
its Central American wintering grounds. 

Mist netting is an important tool in many avian research projects. Passive 
netting by deploying and periodically checking nets works well for many 
studies, especially those set where vegetation is relatively short (approxi- 
mately net height), the target species are active at the heights sampled by the 
net, and/or when multiple bird species are sought. Because these conditions 
do not apply to all areas or studies, targeted trapping methods have been 
developed for many species (e.g., see McClure 1984). By exploiting behav- 
ioral characteristics of a particular species, targeted techniques can be 
especially effective, even under difficult netting conditions. 

Breeding Willow Flycatchers ( Empidonax traillii ) can be captured readily 
via passive mist netting, especially in relatively short, linear habitats or where 
nets can be placed near active nests (Sedgwick and Klus 1997). In 1996, we 
initiated a large-scale Willow Flycatcher research program that required the 
capture of hundreds of adult flycatchers over much of the species’ range. 
Furthermore, we needed to capture both male and female flycatchers to 
study sex-based survivorship and movements. This was especially challeng- 
ing in habitats where thick vegetation in the lower strata and tightly 
interwoven upper-canopy vegetation often limit the number and location of 
net lanes. The Southwestern Willow Flycatcher (£. t. extimus), our primary 
target bird, is a federally listed endangered species, so the technique also had 
to minimize damage to habitat and pose little risk of capture-related injury. 

The Willow Flycatcher is an aggressively territorial species (Sedgwick 
2000, Sogge 2000), singing and approaching in response to playback of 
conspecific songs ( fitz-bews ) and calls. This behavior underlies current 
survey protocols (Sogge et al. 1997) and suggests it can be taken advantage 
of during banding efforts (McClure 1984: 222). However, our preliminary 
efforts (1994 and 1995) to capture flycatchers by broadcasting fitz-bews 
from a tape player placed under a mist net met with only limited success. 
Here, we describe a more effective capture technique that includes a variety 


Western Birds 32:167-172, 2001 


167 


A TARGETED MIST NET CAPTURE TECHNIQUE FOR THE WILLOW FLYCATCHER 


of flycatcher songs and calls broadcast from a CD player through speakers 
set on both sides of a mist net. 

METHODS 
Study Area 

We concentrated our banding effort in central Arizona at two main study 
areas, the Salt River and Tonto Creek inflows to Roosevelt Lake (Gila 
County) and the lower San Pedro River and its confluence with the Gila River 
(Pinal County). At both areas, vegetation ranges from 8 to 15 m tall and is 
composed of dense stands of saltcedar ( Tamarix ramosissima ) and broad- 
leaf riparian woodland, primarily of Goodding willow { Salix gooddingii) and 
Fremont cottonwood ( Populus fremontii ). 

Mist Nets 

Because of dense vegetation, we typically used short (length 2.6 and 6 m, 
height 2.6 m, mesh 38 mm) mist nets suspended between aluminum poles 
(diameter 1.6 cm, height 122 cm). The height at which we placed the nets 
(3-4 m at top of net) depended on the height and density of the understory 
and mid-canopy vegetation. In general, we stayed at least 10 m from active 
nests. To provide flight lanes for the flycatchers, we tried to place nets so 
vegetation did not encroach within 2 to 3 m above and on the sides of the 
net; nets closely surrounded by vegetation worked less well. 

Vocalizations 

Flycatchers use a variety of vocalizations during their natural interactions 
and in response to broadcast recorded vocalizations (vocalizations below 
follow terminology in Sedgwick 2000). Highly agitated and aggressive 
flycatchers sing at increased rates and use emphatic fitz-bews, creets, whits, 
writ-tus ( wee-oos of Stein 1963), and trills { churr of Stein 1963). In some 
intense confrontations (e.g., physical chase and aggression), flycatchers 
combine these into patterns sounding roughly like a high-pitched squeaker 
toy. To obtain vocalizations for our broadcasts, we tape-recorded an array of 
Willow Flycatcher vocalizations and from these selected a series that in- 
cluded fitz-bews, creets, and whits of varying rate and pitch, writ-tus, and 
trills. One recording included an array of calls made during an aggressive 
encounter between a flycatcher and a Brown-headed Cowbird (Mo lot hr us 
ater). We transferred the recorded vocalizations from cassette tape to audio 
CD by using a PC-based CD-ROM writer. The final CD contained 12 tracks, 
each track with a 10- to 60-second series of one or more vocalizations. 

Broadcast Equipment 

To broadcast vocalizations, we used a portable CD player and two 
speakers. One speaker was placed on each side of the mist net, 2 to 5 m 
from the net perpendicularly. Where possible, speakers were set in dense 
vegetation or on branches 1 to 2 m above ground. We used commercially 
available monaural amplified speakers and adjusted playback volume with 


168 


A TARGETED MIST NET CAPTURE TECHNIQUE FOR THE WILLOW FLYCATCHER 


the CD player’s controls. The speakers and CD player were connected via 
monaural audio cables 8-15 m (25-50 ft) long with Vs-inch male mono or 
stereo jacks on both ends. We switched sound output between the two 
speakers during a capture attempt by using a handmade toggle switch or 
simply plugging and unplugging the appropriate speaker wire into the CD’s 
headphone jack. 

The Capture Process 

The net/speaker combination was placed in a suitable open area near a 
song perch or nest. Banders sat quietly 5 to 15 m away from the net (in 
dense vegetation whenever possible), where they could clearly see the entire 
net and surrounding area. Once the vocalizations were broadcast, flycatch- 
ers generally paid little attention to the bander, even when he or she was 
clearly visible. The capture process typically began with the bander broad- 
casting fitz-bews and/or whits to bring a flycatcher closer to the net, then 
trying different tracks to find which vocalizations elicited the strongest 
response from that particular flycatcher. This usually stimulated the fly- 
catcher to become more aggressive and move toward the speakers, flying 
close to whichever speaker was broadcasting. By then switching immedi- 
ately to the speaker on the opposite side of the net, the flycatcher could be 
drawn into the net as it flew across to pursue the “moving” vocalization. The 
bander, able to watch the entire area around the net, could readily manipu- 
late the speaker output as the flycatcher moved near the net and could 
approach and remove the bird immediately after it hit the net. 

RESULTS 

The targeted mist-netting technique was very effective in capturing 
territorial adult Willow Flycatchers. From 1996 through 2000, we captured 
492 adult flycatchers at our primary study sites in Arizona (Luff et al. 2000), 
99 elsewhere in Arizona, and 364 in other states ranging from the Pacific 
Northwest to the northeastern and southeastern United States. There were 
no capture-related injuries or mortalities among these 955 flycatchers. At 
our primary study sites, we captured and banded an average of 60 to 75% 
of the adult population at each site. Males generally responded more 
strongly than females and were therefore captured more often. Outside of 
Arizona, where the sex of the captured flycatcher was not important to our 
objectives, we captured many more males (262; 72%) than females (102; 
28%). At our primary study sites in Arizona, where we focused on catching 
both sexes, the proportion of captured males (280; 57%) to captured 
females (212; 43%) was much closer. 

Targeted netting also worked during the nonbreeding season. Lynn and 
Whitfield (2000) reported target-netting 59 wintering flycatchers in El 
Salvador and Panama. In Costa Rica, we captured and banded 82 wintering 
flycatchers — over 90% of the local wintering population at two study sites 
(Koronkiewicz and Sogge 2000); the ratio of males to females was approxi- 
mately equal (USGS unpubl. data). As on the breeding grounds, there were 
no capture-related injuries or mortality. 


169 


A TARGETED MIST NET CAPTURE TECHNIQUE FOR THE WILLOW FLYCATCHER 
DISCUSSION 

This targeted capture technique works well because it takes advantage of 
the Willow Flycatcher’s strongly territorial nature and aggressive response to 
conspecific vocalizations. Three factors were important in developing our 
particular capture protocol: incorporating a variety of flycatcher vocaliza- 
tions, using a CD player, and using multiple speakers to manipulate the 
flycatcher into the net. A variety of recorded vocalizations was useful 
because individual flycatchers responded best to different vocalizations. One 
series of vocalizations, recorded when a flycatcher chased a cowbird from its 
territory, was particularly effective in eliciting aggressive response and 
chasing behavior, especially once a flycatcher was near the net. Other types 
of calls recorded during territorial squabbles between flycatchers also pro- 
duced this effect, though not as strongly. A few flycatchers did not respond 
strongly to any vocalizations and simply sang or called from above the nets 
or elsewhere within their territory. Although we initially used portable tape 
players when netting, we soon switched to CD players because a CD could 
hold multiple tracks and allowed rapid, silent, and relatively motion-free 
switching from one vocalization to another. Furthermore, CD broadcast had 
better clarity and durability than did that of tape players. Use of two speakers 
dramatically improved capture success over use of one. By alternately 
switching the sound output from one side of the net to the other, we more 
naturally mimicked a conspecific territorial intrusion and were better able to 
manipulate the responding flycatcher into the net. Flycatchers focusing 
intently on and pursuing the source of the vocalizations flew close to and/or 
perched near the broadcasting speaker. Placing the speaker high (over 2 m) 
often led to flycatchers flying back and forth over the top of the net, so the 
positioning of speakers relative to the net was very important. Furthermore, 
long speaker wires (>10 m) are important because they allow the bander to 
move away from the net, limiting disturbance. 

Frequently, both birds of a pair reacted to the broadcast vocalizations, and 
we sometimes caught the second bird after its mate was captured. When 
females were targeted, nets placed 10-20 m from the nest were more 
effective than those placed farther away within the territory. Females often 
responded aggressively and sometimes sang during the capture process, so 
aggression and song do not always indicate the sex of the responding bird. 
The combination of CD vocalizations and aggressive response by territory 
holders sometimes stimulated neighboring males to intrude into the territory 
and be caught. 

A particularly important aspect of this technique is its low impact on the 
flycatchers. Based on voluntary reporting, the overall injury and mortality 
rates of passerines during mist-netting/banding projects are 0,2% and 
0.4%, respectively (Canadian Bird Banding Office, unpubl. data). Even such 
relatively low rates could be problematic in work with an endangered 
species. Our targeted mist-netting technique, coupled with highly experi- 
enced staff, has allowed us to avoid any capture-related injuries or mortality 
among the >1000 flycatchers captured on the breeding and wintering 
grounds. Because a bird’s degree of entanglement in a net (and thus difficulty 
of removal) is directly proportional to its time in the net (McClure 1984), the 


170 


A TARGETED MIST NET CAPTURE TECHNIQUE FOR THE WILLOW FLYCATCHER 


bander’s being present when the bird is captured and removing it immedi- 
ately helps minimize stress and avoid injuries. The technique also minimizes 
our time within a flycatcher’s territory; birds not previously targeted for 
capture were usually caught within 10 minutes (broadcast time). 

We believe this technique can be readily adapted to other species as well. 
Using only songs and calls from commercially available tapes and CDs, we 
have used this technique to capture Alder (£. alnorum), Cordilleran 
(£. occidentalis), and Dusky (£. oberholseri ) Flycatchers. Johnson et al. 
(1981) identified at least 51 bird species that are responsive to playback 
recordings and might therefore respond well to targeted capture. Persons 
attempting to use this technique on other species can start with basic songs 
and calls from their target species, then augment with additional vocaliza- 
tions where possible. 

ACKNOWLEDGMENTS 

This work has been funded in part with federal financial assistance from the U.S. 
Bureau of Reclamation (Boulder City, Phoenix, and Salt Lake City offices), the 
National Park Service, and the U.S. Geological Survey, and with funding from the 
Arizona Game and Fish Department Heritage Program. Development of this tech- 
nique was possible due to the hard work and creativity of the 1996-2000 USGS 
banding crews; we extend our sincere thanks to Michelle Davis, Robert Emerson, 
Heather English, Kerry Kenwood, Therese Littlefeather, Jennifer Luff, Andrew 
McIntyre, Michael Moore, Renee Netter, and John Semones. Darrell Ahlers, Terry 
Doyle, Charles Drost, Linda Sogge, and an anonymous reviewer provided helpful 
comments on drafts of the manuscript. 

LITERATURE CITED 

Johnson, R. R., Brown, B. T„, Haight, L. T., and Simpson, J. M. 1981. Playback 
recordings as a special avian census technique. Studies Avian Biol. 6:68-75. 

Koronkiewicz, T. J., and Sogge, M. K. 2000. Willow Rycatcher (Empidonax traillii) 
winter ecology in Costa Rica: 1999/2000. Unpublished report to the U.S. 
Bureau of Reclamation (order from U.S. Geological Survey, P.O. Box 5614, 
Ragstaff, AZ 86011). 

Luff, J. A., Paxton, E. H., Kenwood, K. E., and Sogge, M. K. 2000. Survivorship and 
movements of Southwestern Willow Flycatchers in Arizona — 2000. Unpub- 
lished report to the U.S. Bureau of Reclamation (order from U.S. Geological 
Survey, P.O. Box 5614, Flagstaff, AZ 86011). 

Lynn, J. C., and Whitfield M. 2000. Winter distribution of the Willow Rycatcher 
( Empidonax traillii ) in Panama and El Salvador. Unpublished report to U.S. 
Bureau of Reclamation (order from Southern Sierra Research Station, P.O. Box 
1316, Weldon, CA 93283). 

McClure, E. 1984. Bird Banding. Boxwood Press, Pacific Grove, CA. 

Sedgwick, J. A. 2000. Willow Flycatcher ( Empidonax traillii), in The Birds of North 
America (A. Poole and F. Gill, eds.), no. 533. Birds N. Am., Inc.., Philadelphia. 

Sedgwick, J. A., and Klus, R. J. 1997. Injury due to leg bands in Willow Rycatchers. 
J. Field Ornithol. 68:622-629. 

Sogge, M. K. 2000. Breeding season ecology, in Status, Ecology, and Conservation 
of the Southwestern Willow Rycatcher (D. M. Finch and S. H. Stoleson, eds.), 
pp. 57-70. USDA Forest Serv. Gen. Tech. Rep. RMRS-GTR-60. 


171 


A TARGETED MIST NET CAPTURE TECHNIQUE FOR THE WILLOW FLYCATCHER 


Sogge, M. K., Marshall, R. M., Tibbitts, T. J., and Sferra, S. J. 1997. A Southwestern 
Willow Flycatcher natural history summary and survey protocol. Nat. Park Serv. 
Tech. Rep. NPS/NAUCPRS/NRTR-97/12 (order from U.S. Geological Survey, 
P.O. Box 5614, Flagstaff, AZ 86011). 

Stein, R. C. 1963. Isolating mechanisms between populations of Traill’s Flycatchers. 
Proc. Am. Philos. Soc. 107:21-50. 


Accepted 10 May 2001 



Willow Flycatcher 


Sketch by John Schmitt 


172 



NOTES 


FURTHER EVIDENCE FOR A POPULATION 
DECLINE IN THE WESTERN WARBLING VIREO 

THOMAS GARDALI, Point Reyes Bird Observatory, 4990 Shoreline Highway, 
Stinson Beach, California 94970 

ALVARO JARAMILLO, San Francisco Bay Bird Observatory, P. O. Box 247, Alviso, 
California 95002 

Recently, Gardali et al. (2000) reported that mist-net capture rates of breeding and 
migrating western Warbling Vireos ( Vireo gilvus swainsonii) had declined in Marin 
County, California, from 1979 to 1997. Evidence of a population decline from a 
single site, though, may be misleading and simply represent site-specific changes in 
abundance. Here we report a trend in Warbling Vireo abundance from 12 years 
(1987-1998) of mist-net data at another site in California, the Coyote Creek Field 
Station of the San Francisco Bay Bird Observatory. 

The Coyote Creek Field Station is situated at the extreme south end of San 
Francisco Bay, north of San Jose, Santa Clara County (37° 20' N, 122° 10' W). The 
site is 80 krn southeast of the Marin County site reported on by Gardali et al. (2000). 
The habitat at the Coyote Creek Field Station is a mix of valley riparian forest and an 
open weedy field. The latter is within a non-concrete flood-control channel that flows 
periodically during wet winters. This overflow channel is mowed every several years to 
prevent woody vegetation from colonizing. The forest habitat includes an old stand 
immediately adjacent to Coyote Creek as well as two riparian restoration sites, one 
planted in 1986, the other in 1994. The forested habitats are dominated by Fremont 
cottonwood ( Populus fremontii), willows (Salix spp.), boxelder (Acer rtegundo), 
western sycamore ( Platanus racemosa), and coyote brush ( Baccharis pilularis ). For 
a more detailed description of the study area see Otahal (1995). 

Warbling Vireos were sampled through standardized-effort mist-netting. A total of 
48 nets at fixed locations (36 prior to 1995) were operated once weekly throughout 
the year. Nets were opened 45 minutes before sunrise and kept open for 5 hours 
during each day of operation. A variable number of additional nets were also operated 
as part of other research projects over the years, particularly during migration. 
However, netting effort remained relatively constant between 1987 and 1998, except 
for periods of greater effort in spring 1996 and fall 1998. Captured Warbling Vireos 
were banded with standard U.S. Fish and Wildlife Service bands, measured, and 
released immediately. Aging of birds in the hand was achieved by noting the extent of 
skull pneumatization in the fall (Pyle 1997). 

Warbling Vireos do not breed at Coyote Creek, so our analyses were limited to 
examining trends for both migratory periods separately, fall (August-October) and 
spring (April-May). Our sample consisted of 335 individuals (202 in fall and 133 in 
spring). Using the program STATA (Stata Corp. 1997), we examined trends by linear 
regression. We used the number of individual birds captured per 1000 net-hours to 
standardize capture rate by netting effort. Capture totals were log-transformed in 
order to improve the normality of the model residuals (Zar 1996). We examined 
whether trends were nonlinear by testing for a significant quadratic coefficient for year 
in the presence of a linear term and found no trends to be nonlinear. Significance was 
assumed at a level of P < 0.05. 


Western Birds 32:173-176, 2001 


173 


NOTES 


Table 1 Fall Netting Effort and Captures of Warbling Vireos 
at Coyote Creek Field Station, 1987-1998 


Year 

Net hours 

Vireos captured 

Vireos per 1000 net hours 

1987 

2467.50 

17 

6.89 

1988 

2948.00 

24 

8.14 

1989 

2825.99 

56 

19.82 

1990 

3494.75 

19 

5.44 

1991 

2525.00 

6 

2.38 

1992 

2757.50 

26 

9.43 

1993 

2806.75 

13 

4.63 

1994 

2574.00 

14 

5.44 

1995 

2535.00 

9 

3.55 

1996 

2292.50 

13 

5.67 

1997 

3124.50 

8 

2.56 

1998 

4655.62 

6 

1.29 


The captures of fall migrants declined significantly over the course of our study ((3 
= -0.13, SE = 0.046, P = 0.02; Table 1; Figure 1); the average annual change was 
-12.2% per year. Fall captures were primarily of birds that hatched that year (86.8%). 
Capture rates of Warbling Vireos migrating through Coyote Creek in spring showed 
a negative but nonsignificant trend ((1 = -7.98, SE = 0.0003, P = 0.79). 

Over the course of this study there has been an increase in riparian habitat at the 
Coyote Creek Field Station due to restoration efforts. Capture rates of several 
species that occur at Coyote Creek only as migrants have increased over the study 
period (e.g., the Willow Flycatcher, Empidonax traillii ), perhaps in response to 
restoration efforts (San Francisco Bay Bird Observatory unpubl. data), while the 
Warbling Vireo has declined. However, whether capture rates during migration are 
responsive to habitat restoration remains unknown. Our spring capture data suggest 
a stable pattern for adult migrants, though our sample size may not have been great 
enough to detect a trend. In fall, however, we did detect a decline in abundance of 
Warbling Vireos, of which the majority were first-year birds. Hence, the decline in 
fall captures probably represents a decline in breeding productivity, as also reported 
by Gardali et al. (2000). 

Like this study and that of Gardali et al. (2000), the Breeding Bird Survey (BBS) has 
detected a decline in Warbling Vireo abundance in California (Sauer et al. 1999). We 
recorded a rate of decline, expressed as the slope of log-transformed data, similar to 
that of Gardali et al. (2000; p = -0.13 and p = -0.10, respectively). The BBS data 
showed a decline of 0.9% per year ( P = 0.06) from 1966 to 1998 compared to our 
-12.2%. When the analysis of BBS data was limited to 1980-1998 it showed a more 
severe annual decline of 2.3% per year (P = 0.01; Sauer et al. 1999). Total fall 
captures at Southeast Farallon Island demonstrated no trend from 1968 to 1992 (Pyle 
et al. 1994). The Farallon captures were of both adults and young birds, in contrast 
with mostly young birds at Coyote Creek. The Farallon study did report a significant 
negative trend in the ratio of young to adult birds, supporting our contention that there 
has been a decline in the Warbling Vireo’s breeding productivity. 

The mechanisms responsible for a decline in Warbling Vireo abundance in 
California are not known. The decline in abundance of this species at Coyote Creek 
in autumn, where first-year birds constitute a preponderance of Warbling Vireos 
captured, suggests that fewer young have been produced annually; the extent of the 


174 


NOTES 



Figure 1. Trend in fall captures of Warbling Vireos, 1987-1998, at Coyote Creek 
Field Station. Each circle represents datum for one year; line is least-square line of best 
fit (P = 0.02, r 2 = 0.39). 


breeding range sampled by our mist-nets is not known. Gardali et al. (2000) presented 
indirect evidence that linked poor reproductive success to population declines. Recent 
work in interior British Columbia suggests that local populations of western Warbling 
Vireos have the potential to be extirpated by high levels of parasitism by the Brown- 
headed Cowbird ( Molothrus ater) but that even at low levels of parasitism vireo 
nesting success is low (Ward and Smith 2000). More studies are needed to examine 
the specific factors limiting the reproductive success of Warbling Vireos. 

Our results, together with the BBS (Sauer et al. 1999) and Gardali et al. (2000), 
provide strong evidence that western Warbling Vireo populations are in a decline. 

None of these data would be available had it not been for the dedication of the 
Coyote Creek Field Station’s volunteer banders. Thank you for the commitment and 
the fine data. Part of this work was funded by the Santa Clara Valley Water District’s 
Coyote Creek Flood Control Project, Wildlife Monitoring Program. This manuscript 
was improved by the thoughtful reviews of Kimball Garrett, Tim Manolis, and Kathy 
Molina. Steve N. G. Howell provided valuable editorial comments. This is contribution 
769 of the Point Reyes Bird Observatory and 120 of the San Francisco Bay Bird 
Observatory. 

LITERATURE CITED 

Gardali, T., Ballard, G., Nur, N., and Geupel, G. R. 2000. Demography of a declining 
population of Warbling Vireos in coastal California, Condor 102:601-609. 

Otahal, C. D. 1995. Sexual differences in Wilson’s Warbler migration. J. Field 
Ornithol. 66:60-69. 


175 


NOTES 


Pyle, P. 1997. Identification Guide to North American Birds, part I, Columbidae to 
Ploceidae. Slate Creek Press, Bolinas, CA. 

Pyle, P., Nur, N., and DeSante, D. F. 1994. Trends in nocturnal migrant landbird 
populations at Southeast Farallon Island, California, 1968-1992. Studies Avian 
Biol. 15:58-74. 

Sauer, J. R., Hines, J. E., Thomas, I., Fallon, J., and Gough, G. 1999. The North 
American Breeding Bird Survey, results and analysis 1966-1998, version 98.1. 
U.S. Geol. Surv., Patuxent Wildlife Research Center, 12100 Beech Forest Rd., 
Suite 4039 Laurel, MD 20708-4039. 

Stata Corp, 1997. Stata Statistical Software: Release 5.0. Stata Corporation, 702 
University Dr., East College Station, TX 77840. 

Ward, D., and Smith, J. N. M. 2000. Brown-headed Cowbird parasitism results in a 
sink population in Warbling Vireos. Auk 117:337-344. 

Zar, J. H. 1996. Biostatistical Analysis, 3rd ed. Prentice Hall, Englewood Cliffs, N.J. 

Accepted 20 April 2001 


176 


NOTES 


BRANDT S CORMORANT SINKS AT SEA 

TERENCE R, WAHL, 3041 Eldridge, Bellingham, Washington 98225 

Cormorants’ limited degree of plumage waterproofing, the spread-winged posture 
attributed to the need to dry their feathers, their need to rest and roost out of the water, 
and their nearshore foraging distribution have been commented upon by many (e.g., 
Schneider and Hunt 1984, Boekelheide et al. 1990, Siegel-Causey and Litvinenko 
1993). I am unaware of descriptions of consequences for cormorants should they be 
unable leave the water before saturation. 

Counts of seabirds made on >330 one-day trips off Grays Harbor, Washington, 
between 1971 and 2000 (see Wahl and Tweit 2000) included a total of 26,340 
Double-crested ( Phalacrocorax auritus), Brandt’s (P. penicillatus ), Pelagic 
(P. pelagicus ), and unidentified cormorants. Their nearshore distribution was obvious: 
74% of the birds were in Grays Harbor and channel, 22% were between there and a 
depth of 20 m, and about 4% were in water 20-50 m deep, most of them just outside 
the 20-m contour. Thus essentially all birds were within 20 km of shore — only 97 were 
seen from there to the edge of the continental shelf. 

In addition to these one-day trips 1 spent several weeks aboard the R./V. Thomas 
G. Thompson , of the University of Washington Department of Oceanography, on 
research cruises from near shore to 126° 30' W, about 160 km offshore. On 18 
September 1976, during one such cruise, I first noted a Brandt’s Cormorant perched 
on the bow at 47° 07' N, 124° 54' W, about 138 km offshore. Over the next 5 days 
the bird (presumably the same individual) roosted aboard the ship and foraged nearby. 
The ship was under way much of the time and, even when it was stopped on station, 
people seldom visited the bow but remained on the afterdeck. The ship traveled east 
and west on a sampling track to 125° 01' W offshore and closest to land at 124° 41' 
W, about 35 km offshore. 

On 23 September at 0615 the ship stopped on station at 47° 07’ N, 124° 45’ W, 
when the sea was nearly flat. Shortly thereafter scientists and crew had scattered all 
over the ship. At 0830 I observed the cormorant in the water, about 50 m away. 1 do 
not know how long the bird had been in the water, but the ship had been collecting 
phytoplankton samples on a box pattern with stops about 0.7 to 2.2 km apart, and 
the bird could have been foraging for 2 to 3 hours with the slow-moving ship easily 
within sight. It was apparent that the bird was slowly settling lower in the water. When 
it was low in the water, with only its head and neck erect and exposed, two gulls ( Larus 
occidentalis or L. glaucescens) approached and pecked at it briefly, eliciting little 
response from the cormorant. Just before the ship moved off station at 0850 I noted 
only the bird’s head above the surface. My attention was diverted for a few seconds, 
and when I looked again the bird had sunk. I saw no evidence of a struggle resulting 
from underwater attack. It had appeared healthy the day before and flew normally. I 
concluded that the bird was either scared off the ship or, on return from foraging, had 
been intimidated by people, would not return to the ship, and became waterlogged 
and drowned. 

I thank D. Fix for useful review comments. 

LITERATURE CITED 

Boekelheide, R. J., Ainley, D. G., Morrell, S. H., and Lewis, T. J. 1990. Brandt’s 
Cormorant, in Seabirds of the Farallon Islands (D. G. Ainley and R. J. 
Boekelheide, eds.), pp. 163-194. Stanford Univ. Press. Stanford, CA. 

Schneider, D., and Hunt, G. L., Jr. 1984. A comparison of seabird diets and foraging 
distribution around the Pribilof Islands, Alaska, in Marine birds: Their feeding 


Western Birds 32:177-178, 2001 


177 


NOTES 


ecology and commercial fisheries relationships (D. N. Nettleship, G. A. Sanger, 
and P. F. Springer, eds.), pp. 86-95. Can. Wildlife Serv., Environment Canada, 
Ottawa, Ontario K1A 0H3. 

Siegel-Causey, D., and Litvinenko, N. M. 1993. Status, ecology and conservation of 
shags and cormorants of the temperate North Pacific, in The status, ecology and 
conservation of marine birds of the North Pacific (K. Vermeer, K. T. Briggs, K. H. 
Morgan, and D. Siegel-Causey, eds.), pp. 122-130. Can. Wildlife Serv., Envi- 
ronment Canada, Ottawa, Ontario K1A 0H3. 

Wahl, T. R., and Tweit, B. 2000. Seabird abundances off Washington, 1972-1998. 
W. Birds 31:69-88. 


Accepted 30 August 2001 


NOTES 


FIRST RECORD OF THE EUROPEAN 
GOLDEN-PLOVER ( PLUVIALIS APRICARIA) 

FROM THE PACIFIC 

ANDREW W. PISTON, P. O. Box 6553, Ketchikan, Alaska 99901 
STEVEN C. HEINL, P. O. Box 23101, Ketchikan, Alaska 99901 

The European Golden-Plover (Pluvialis opricaria) breeds from Iceland and the 
British Isles east to the base of the Taimyr Peninsula, Russia, at about 102° 30' E 
(Vaurie 1965). Virtually the entire population migrates to or through Europe to winter 
in the British Isles, western Europe, and throughout the Mediterranean Basin; small 
numbers winter east to the southern Caspian Sea and casually to eastern India, and 
small numbers winter on the Atlantic coast of Africa, casually south to Gambia (Vaurie 
1965, Cramp 1983). At the western periphery of this range, the European Golden- 
Plover is a regular vagrant to Greenland, where it is also a local breeder in the northeast 
(Boertmann 1994). It is a casual visitant to Newfoundland (including Labrador) and 
Saint Pierre et Miquelon, with nearly all records in April and May (Tuck 1968, 
Mactavish 1988, ABA 1996). There are reports from New Brunswick (Am. Birds [AB] 
42: 408, 1988), Nova Scotia (AB 43:56, 1989; AB 44:390 1990; Natl. Audubon 
Soc. Field Notes [NASFN] 49:222, 1995), and Quebec (AB 42:1271, 1988). 

On 14 January 2001 we collected a European Golden-Plover near the Ketchikan 
airport, Gravina Island, Alexander Archipelago, southeast Alaska (55° 17' N, 131° 
46' W). In a search of the literature, and contacts with shorebird specialists in Asia and 
the Pacific, we found no evidence of the prior occurrence of this species in the Pacific 
basin. 

Piston discovered the bird on 13 January 2001 and watched it for approximately 
one hour as it fed on a rocky gravel-covered beach with a flock of 35 Black Turnstones 
(Arenaria melanocephala ), three Rock Sandpipers ( Calidris ptilocnemis ), and two 
Surfbirds { Aphriza virgata). Knowing that the occurrence of a golden-plover in Alaska 
in the winter was unprecedented, he took photographs and notes in the field. Later 
that same day we looked over references and discussed the bird’s field marks. The bird 
was brightly colored, with gold speckles over the entire back and a golden wash on the 
head and breast. The bird also showed a faint white wing bar when it flew, a field mark 
of the European Golden-Plover. Other critical field marks were not noted, and we did 
not make much of the wing stripe at the time. We discussed the identification of the 
European Golden-Plover, but, quite naturally, did not seriously consider that species a 
possibility'. Instead we focused our discussion on the field identification of the 
American (P dominico ) and Pacific (P fulva) golden plovers. 

We determined to relocate the bird the next day and collect a voucher specimen of 
what we figured to be a Pacific Golden-Plover. We based this assumption simply on 
the fact that the bird was brightly colored, and also because that species would be the 
most likely golden-plover to occur in Alaska in the winter. The Pacific Golden-Plover 
winters locally in very small numbers in California (Garrett and Dunn 1981, Flarris 
1996), and it has been found casually in winter from Oregon (Gilligan et al. 1994, 
Contreras 1998) to the coast of southwestern British Columbia (Campbell et al. 
1990). More recently, the Pacific Golden-Plover has been reported in winter at the 
Queen Charlotte Islands, British Columbia, only 150 km southwest of Ketchikan. One 
at Massett 15 December 1991-February 1992 was reported as the first winter record 
for the Queen Charlotte Islands (AB 46: 304, 1992); two golden-plovers at Sandspit, 
27 December 1997, were considered ‘'about" the fourth winter record for the Queen 
Charlotte Islands (NASFN 52:245, 1998). Although the American Golden-Plover has 
been collected in mid-winter in the southeastern United States (Paulson and Lee 


Western Birds 32:179-181, 2001 


179 


NOTES 


1992), there are no substantiated midwinter records on the west coast of North 
America. No golden-plovers had previously been reported in Alaska in the winter. 

On 14 January' 2001 we relocated the bird with the same flock of shorebirds and 
again noted that it was relatively brightly colored. When it flew a short distance we saw 
that the bird indeed had a distinct narrow white wingstripe across the base of the flight 
feathers. At one point the bird flapped its wings and we were both stunned to see that 
it clearly had white axillaries and underwing coverts. We then noted that the bird 
appeared rather dumpy, with proportionately short legs, a chunky body and a 
neckless look, subtly different from the slimmer bodied, longer legged, and longer 
necked appearance that we are used to seeing in the Pacific Golden-Plover. We also 
noted that the bird had a shortish, slightly conical, deep-based, fine-tipped bill, and a 
uniform face pattern that lacked a strong supercilium (the supercilium looked to be the 
same dull yellowish color as the auriculars). All of these field marks led us to believe 
that the bird was probably a European Golden-Plover. (Excellent treatments of the 
field identification of the golden-plovers can be found in Hayman et al. 1986, Jonsson 
1992, Beaman and Madge 1998, and Svensson et al. 1999.) We saw the bird fly 
again three or four times, but the underwings did not look especially white; either the 
light was not good or we did not have the right angle. Each time it flew we noted the 
narrow white stripe on the upperwing, which could be seen clearly from at least 100 
meters. We agreed that we had never seen a Pacific or an American golden-plover 
with such a distinct wing stripe. We collected the bird and were again stunned to see 
that the bird had white axillaries and underwing coverts. 

We forwarded the specimen to the University of Alaska Museum (UAM), Fairbanks, 
where our identification was corroborated by Daniel D. Gibson. Gibson prepared the 
specimen as a study skin and preserved partial skeleton, frozen tissues, stomach 
contents, and guts (lower digestive tract for disease screening). The specimen (UAM 
12100) is a first-winter male, with mass 199.5 g (heavy fat), wing chord 181 mm, tail 
length 71.0 mm, bill length (from distal end of naris) 14.3 mm, bill depth 5.2 mm (at 
distal end of naris), bill width 4.9 mm (at distal end of naris), and tarsus length 45.0 mm 
(D. D. Gibson in litt.). The bird’s stomach contained six species of gastropod mollusks 
up to 3 mm in length — four snails ( Lacuna uinita, Margarites helicinus, Lirularia 
succincta, Littorina scutulata ) and two limpets ( Lottia sp. and Tectura sp.) — plus at 
least one crustacean fragment (Tanaidacea; Nora R. Foster, UAM, in litt.). 

We thank Daniel D. Gibson for providing the specimen data and references, and for 
helpful comments that improved the quality of this note. We also thank Kimball L. 
Garrett and Robert E. Gill for their reviews. 

LITERATURE CITED 

American Birding Association. 1996, ABA Checklist: Birds of the Continental United 
States and Canada, 5th ed. Am. Birding Assoc., Colorado Springs, CO. 

Beaman, M., and Madge, S. 1998. The Handbook of Bird Identification for Europe 
and the Western Palearctic. Princeton Univ. Press, Princeton, N.J. 

Boertmann, D. 1994. A annotated checklist to the birds of Greenland, Meddelelser 
om Gronland. Bioscience 38. 

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. Royal British 
Columbia Museum, Victoria, B. C. 

Contreras, A. 1998. Birds of Coos County, Oregon: Status and Distribution. Cape 
Arago Audubon Soc. and Ore. Field Ornithol. Spec. Publ. 12. 

Cramp, S. (ed.). 1983. Handbook of the Birds of Europe, the Middle East, and North 
Africa, vol. 3. Oxford Univ. Press, Oxford, England. 


180 


NOTES 


Garrett, K., and Dunn, J. 1981. Birds of Southern California. Los Angeles Audubon 
Soc., Los Angeles. 

Gilligan, J., Smith, M., Rogers, D., and Contreras, A. 1994. Birds of Oregon: Status 
and Distribution. Cinclus, McMinnville, OR. 

Harris, S. W. 1996. Northwestern California Birds. Humboldt State Univ. Press, 
Areata, CA. 

Hayman, P., Marchant, J., and Prater, T. 1986. Shorebirds: An Identification Guide. 
Houghton Mifflin, Boston. 

Jonsson, L. 1992. Birds of Europe with North Africa and the Middle East. Princeton 
Univ. Press, Princeton, N.J. 

Mactavish, B. 1988. Greater Golden-Plover invasion, 1988. Birding 20:242-249. 

Paulson, D. R., and Lee, D. S. 1992. Wintering of Lesser Golden-Plovers in eastern 
North America. J, Field Ornithol. 63:121-128. 

Svensson, L., Grant, P. J., Mullarney, K., and Zetterstrom, D. 1999. Birds of Europe. 
Princeton Univ. Press, Princeton, N.J. 

Tuck, L. M. 1968. Recent Newfoundland bird records. Auk 85:304-311. 

Vaurie, C. 1965. The Birds of the Palearctic Fauna. H. F. & G. Witherby, London. 

Accepted 1 7 August 2001 


181 


BOOK REVIEWS 


United States Shorebird Conservation Plan, by S. Brown, C. Hickey, B. 
Harrington, and R. Gill (eds.). Second edition. May 2001. Manomet Center for 
Conservation Sciences, Manomet, Massachusetts. Available on request from U.S. 
Fish and Wildlife Service, Division of Migratory Bird Management, 4401 North 
Fairfax Drive, Room 634, Arlington, VA 22203, or through the World Wide Web 
(with accompanying technical documents and regional plans) at http:// 
www.manomet.org/USSCP/files.htm. 

Shorebirds may not be as commercially valuable as waterfowl, or as widely 
appreciated by the general public as songbirds, yet they have long held a special 
fascination for birders and ornithologists. The spectacular migrations undertaken by 
some species, and the wild regions they often inhabit, stir both the soul and the mind. 
Until recently, however, most shorebirds have tended to slip through the cracks in 
conservation consciousness. Yet their highly migratory habits, their need to concen- 
trate at a few food-rich sites, and their use of habitats prone to human disturbance and 
development combine to make shorebirds vulnerable at many levels. 

The need to conserve these remarkable birds has taken on a new dimension with 
the publication of the U.S. Shorebird Conservation Plan (hereafter the plan). The plan 
is a partnership of state and federal agencies, nongovernmental organizations, 
academic institutions, and individuals committed to restoring and maintaining shore- 
bird populations in the U.S. and across the Western Hemisphere. The plan summa- 
rizes technical reports and recommendations produced by working groups that 
developed the plan and offers a wealth of information concerning what is needed for 
practical shorebird conservation in the United States. 

The plan has been designed to complement, and be integrated with, the continent- 
wide conservation initiatives of the North American Waterfowl Management Plan, 
Partners in Flight, and the North American Colonial Waterbird Conservation Plan, all 
of which share much common ground. International coordination is also underway 
with the Canadian Shorebird Conservation Plan, which shares responsibility for many 
of the same species. The need for information to be updated constantly and for goals 
to be reassessed means that the plan will be revised every five years over the next 15 
years, thereafter as necessary. 

Seven parts constitute the plan: an introduction to shorebirds’ general biology and 
conservation needs, a vision for shorebird conservation, population sizes and conser- 
vation status, national conservation strategies, regional goals and strategies, how the 
plan should be implemented, and a listing of associated technical reports (which 
should be consulted for greater detail and references). Five appendices give popula- 
tion estimates for shorebirds breeding in North America, map the 12 planning regions 
in the U.S., rank the relative importance of each species in each region, classify the 
level of risk for each species, and list species recorded rarely in North America and not 
treated by the plan. The five planning regions most relevant to the western U.S. are 
Alaska, Hawaii/Pacific Islands, Northern Pacific, Southern Pacific, and Intermoun- 
tain West. Each is based on bird-conservation regions identified by the North 
American Bird Conservation Initiative. 

The plan has three major goals at different scales: at a regional scale, to identify, 
protect, and manage important shorebird habitats; at the national level, to stabilize 
population levels of species suspected to be in decline, while keeping common species 
common; and at the hemispheric scale, to restore and maintain populations of all 
shorebird species in the Western Hemisphere through cooperative international 
efforts. 

The plan recognizes that effective shorebird conservation strategies should be 
based on sound science and that baseline population estimates of all species are a 
prerequisite to this. Estimates presented in the plan address geographically disjunct 


182 


Western Birds 32:182-185, 2001 


BOOK REVIEWS 


populations (e.g.. Pacific coast and interior Snowy Plovers) as well as formally 
recognized subspecies separately and, while crude in most cases, are a starting point 
for broad policy goals. The plan also sets population targets and estimates a level of 
risk (from “highly imperiled” to “not at risk”) for each population from variables such 
as population trend, relative abundance, and specific threats on the breeding and 
nonbreeding grounds. 

How the worthy aims of the plan should be accomplished is discussed in parts 4 
through 6. Part 4 advocates that the conservation strategies of greatest priority at the 
national level should be population and habitat monitoring (at an estimated cost of 
$1.5 million per year), the establishment of a national shorebird- research program to 
address shorebird biology (with annual funding of $2 million for national and $1.75 
million for regional research priorities), and concerted public education and outreach 
programs (no cost estimate). Part 5 summarizes regional goals and strategies, e.g., 
protecting important stopover sites along the Pacific coast. For the southern Pacific 
region, “regional priorities must include increasing populations of breeding species 
such as Snowy Plover, Killdeer, . . . Black-necked Stilt , and American Avocet. ” Although 
many birders and field ornithologists may question this statement for three of these 
species, can we prove otherwise? If nothing else, this should force us to recognize the 
need for reliable baseline data and make us question how much we take for granted. 
Part 6 proposes a model for the plan’s implementation, to be coordinated by the U.S. 
Shorebird Plan Council. Membership of the council is open to any organization, and 
current members include the American Bird Conservancy, Canadian Wildlife Service, 
Ducks Unlimited, the Nature Conservancy, and several branches of the U.S. govern- 
ment. 

A phenomenal amount of work went into coordinating and publishing this 
ambitious plan, which appears to have included all interested parties. Two observa- 
tions might be addressed in future editions. I appreciate that agendas as broad as 
nationwide shorebird conservation are difficult to communicate succinctly. With 
judicial editing and reorganization, however, the plan’s length could be cut signifi- 
cantly — with no loss of content. In this edition, anyone wishing to identify a focused 
strategy of action is likely to be lost in a maze of circumlocution. Second, I found no 
mention of the plan’s intended audience, although I assume there is one. Identifying 
an audience could help focus the plan’s content. The general public, unless fluent in 
jargon (e.g., “to enhance funding capabilities and delivery of habitat protection and 
restoration activities governed by provisions of...”) will likely “tune out” quickly. Those 
with a more scientific background also may be troubled by the nebulous content and 
lack of direct citations — but note that these are included in the technical documents 
accompanying the electronic version. Even if jargon-literate land managers and 
politicians are the intended audience, frequent repetition in the plan is inefficient of 
the reader’s time. 

The plan represents an overview of shorebird conservation issues, which now need 
to be addressed. The technical committees and regional working groups formed 
during the plan’s conception are working actively toward its fulfillment, and the best 
way for interested parties to become involved is to consult their regional plan and 
make contacts from there. Shorebirds need all the help they can get, and I am very 
glad that a growing constituency is dedicated to conserving what have long been 
among my favorite birds, 

Steve N. G. Howell 


183 


BOOK REVIEWS 


The Riparian Bird Conservation Plan, by California Partners in Flight and the 
Riparian Habitat Joint Venture, version 1.0. 2000. 88 pp. Available free (in pdf 
format) from the Point Reyes Bird Observatory (PRBO) website: http:// 
www. prbo . org/CPIF/ Consplan . html . 

Habitat loss may be the leading cause of population declines and range reductions 
among landbirds in western North America. To reverse this trend and maintain 
existing populations, remnant high-quality habitats must be protected and degraded 
habitats restored. Although efforts to protect and enhance riparian habitats are 
underway, land managers designing restoration programs face a high degree of 
uncertainty in deciding which management actions will be most effective. To reduce 
this uncertainty, the full range of knowledge and skills from the natural and social 
sciences should be brought to bear on the problem. In addition, dissemination of 
information among scientists, managers, and stakeholders should be rapid, recipro- 
cal, and continuous (Mangel et al. 1996). 

Recognizing these challenges, California Partners in Right (see Bonney et al. 2000 
for a discussion of the Partners-in-Right approach) teamed up in 1994 with a diverse 
coalition of federal, state, and nonprofit organizations and landowners to form the 
Riparian Habitat Joint Venture (RHJV). The RHJV was modeled after the highly 
successful joint ventures of the North American Waterfowl Management Plan, and to 
date 19 organizations throughout California have joined this initiative to protect 
biodiversity. Broadly stated, the RHJV’s goal is to conserve, increase, and improve 
riparian habitats to protect and enhance California’s native bird populations. 

As a major step toward achieving this goal, the RHJV, with PRBO as the leading 
organization, produced the first edition of the Riparian Bird Conservation Plan 
(hereafter “the plan”) in August 2000. The plan’s goal is to provide scientific 
information and technical guidance to help private landowners, land managers, 
agencies, and conservation organizations select, design, and implement the conserva- 
tion and land management projects of highest priority. More specifically, the plan 
synthesizes current scientific knowledge of the requirements of birds in riparian habitats 
and recommends strategies for habitat protection, restoration, management, monitor- 
ing, and policy. The plan’s guidelines are flexible, so that land-management practices 
designed to benefit wildlife do not conflict with resource-dependent economies. 

Although the plan has a California focus, many of its recommendations are 
relevant to other western states. In part this is because the plan presents separate 
conservation objectives for eight of California’s ten bioregions. To capture the 
conservation needs of these regional avifaunas, fourteen focal species were selected. 
It is hoped that the restoration strategies necessary to support these species represent 
a multispecies umbrella that will protect the riparian bird community at large 
(Lambeck 1997). The plan contains historical and current data on the distribution of 
these focal species from more than 350 sites throughout the state and provides 
population targets for each species in these eight bioregions. In documents separate 
from the plan (links on the PRBO website), each focal species is profiled in detail. 
These profiles present valuable information on life history and distribution but differ 
from other species accounts (e.g., The Birds of North America series) in that they 
focus on species-specific conservation priorities. 

The plan has tremendous potential to advance avian conservation efforts and, 
given the resources now being devoted to riparian restoration, its release could not be 
more timely. Although a landmark effort in its current form, there are aspects of the 
plan that could be improved. For example, it would help if there were greater 
acknowledgment of the degree to which particular conclusions or recommendations 
are based upon speculation versus empirical evidence. Recommendations that are 
based on best guesses and/or anecdotal observations could then be tested with 
rigorous scientific methods. In this manner the document has the potential to function 
as an important hypothesis-generating tool supporting adaptive management. In 


184 


BOOK REVIEWS 


adaptive management, science is used to evaluate current management practices, 
design tractable management experiments, monitor their effectiveness, and recom- 
mend midstream adjustments (Walters 1986). Although the plan does provide some 
research and monitoring recommendations, this section of the plan should be 
expanded. In its present form this section informs land managers of the need to 
integrate scientific investigation and management actions but offers less to the 
research ecologist interested in learning which uncertainties need focused investiga- 
tion. 

Ultimately the Riparian Bird Conservation Plan takes a heroic step forward in 
tightening the link between science and on-the-ground management, integration 
sorely needed if we are to meet today’s conservation challenges. The plan is a valuable 
resource that should be consulted by all those interested in managing riparian 
resources. Additional habitat-based bird-conservation plans (e.g., for oak woodlands) 
are being developed by California Partners in Flight, with drafts available at the PRBO 
website. 

LITERATURE CITED 

Bonney, R., Pashley, D. N., Cooper, R. J., and Niles, L.(eds.). 2000. Strategies for 
bird conservation: The Partners in Flight planning process. Proceedings of the 
3rd Partners in Right Workshop, 1-5 October 1995, Cape May, New Jersey. 
Proceedings RMRS-P-16. U.S. Dept, of Agriculture, Ogden, UT. 

Lambeck, R. J. 1997. Focal species: A multi-species umbrella for nature conserva- 
tion. Conserv, Biol. 11:849-856. 

Mangel M., and 41 others. 1996. Principles for the conservation of wild living 
resources. Ecol. Appi. 6:338-362. 

Walters, C. J. 1986. Adaptive Management of Renewable Resources. McMillan, New 
York. 


Gregory H. Golet 


185 


FEATURED PHOTO 


PARAPATRY IN WOODHOUSE’S AND 
CALIFORNIA SCRUB-JAYS REVISITED 

JON L. DUNN, RR2, Box 52R, Bishop, California 93514 

KIMBALL L. GARRETT, Section of Vertebrates, Natural History Museum of Los 
Angeles County, 900 Exposition Blvd., Los Angeles, California 90007 

Geographical variation in the scrub-jays (the Apheiocoma coerulescens species 
group) has intrigued and confounded ornithologists both before and since Pitelka’s 
(1951) extensive review. In part on the basis of the work of Peterson (1990, 1992), 
the American Ornithologists’ Union (1995) elevated the widespread continental 
scrub-jays (A. californica, the “Western Scrub-Jay”) and the endemic jay of Santa 
Cruz Island (A. insularis, the “Island Scrub-Jay”) to full species rank, restricting the 
name Apheiocoma coerulescens to the isolated Florida Scrub-Jay. Components of 
the Western Scrub-Jay have at times been given full species rank, in the form of a 
division between birds of the Pacific coast (the A. (c.) californica group of subspecies, 
hereafter “California Scrub-Jay”) and birds of interior western North America (the A, 
[c.] woodhouseii group, “Woodhouse’s Scrub-Jay”); an additional group of subspe- 
cies (the “Sumichrast’s Scrub-Jay,” A. [c.] sumichrasti ) occurs on the southern 
Mexican plateau. Woodhouse’s and California Scrub-Jays were treated as separate 
species most recently by Swarth (1918). 

Minor variation in plumage and measurements, differing identifications of type 
specimens, and vague type localities for old names have sparked taxonomic disagree- 
ments and, indeed, markedly conflicting nomenclature (e.g., Phillips 1986). Pitelka’s 
landmark study (1951) did establish limited intergradation between the two groups in 
a zone of parapatry around the Pine Nut Mountains of Douglas Co., Nevada, with 
intergrades encountered north to the Virginia Mountains and south to the California 
border area of northernmost Mono Co. Pitelka also documented intergradation in the 
Owens Valley region, Inyo Co., California. 

Larry Sansone took the photos on the back cover west of Wellington, Nevada, in 
the southernmost Pine Nut Mountains. The upper photo (taken 7 December 1996) 
shows a bird apparently typical of Woodhouse’s Scrub Jay in the western Great Basin, 
A. c. nevadae according to Pitelka’s (1951) nomenclature. The brighter bird in the 
lower photo (15 December 1996) shows characters of Pacific Coast birds (the nearest 
subspecies of the California Scrub-Jay being A. c. superciliosa of the 5th edition of 
the A.O.U. Check-List) but could represent an intergrade. 

Woodhouse’s Scrub-Jays of the western Great Basin are readily told from Califor- 
nia Scrub-Jays by several plumage and structural characters. The head, rump, and tail 
of nevadae are pale blue washed with gray, recalling a Pinyon Jay, and the underparts 
are gray with a blue wash on the undertail coverts; as a result, nevadae shows only a 
weak contrast between the bluish head and gray-brown back and between the dull 
bluish collar and the blended gray underparts. California Scrub-Jays are deeper blue 
dorsally, with strong contrast between the blue head and gray-brown back; they are 
mainly whitish below with a deep blue collar that is interrupted medially, and the 
undertail coverts only occasionally show a pale blue wash. California Scrub-Jays also 
show a blacker auricular patch with a more contrasting white supercilium (again, these 
areas appear more blended in nevadae). Structurally, the bill of nevadae has a 
distinctly thinner base and less decurved culmen than the stouter bill of coastal birds; 
nevadae has relatively longer wings than California Scrub-Jays. We caution here that 
more easterly Woodhouse’s Scrub-Jays of the subspecies woodhouseii and texana 
are somewhat brighter blue above than nevadae, though the other distinctions from 


186 


Western Birds 32:186-187, 2001 


FEATURED PHOTO 


the California group still hold; southward through Mexico, culminating in the 
Sumichrast’s Scrub-Jays, the plumage becomes even more “California-like.” 

We have found Woodhouse’s Scrub-Jays to be consistently much shyer than 
California birds; more easterly Woodhouse’s may tame down in a few picnic areas 
(e.g,, in the Davis Mtns. of Texas), but typically nevadae is quite skittish. Vocally, at 
least within the core range of nevadae, the common upslurred “jrr-eee?” call is 
relatively high pitched and almost two parted, so that the first syllable is level and the 
second rises abruptly. California Scrub-Jays have a somewhat lower, harsher and 
monosyllabic “shhrreee?” call. The rapid “shreek shreek shreek...” series of 
Woodhouse’s is, to our ears, higher pitched and squeakier. 

Woodhouse’s Scrub- Jays wander erratically away from their breeding range in fall 
and winter, with notable movements occurring every several years (most recently during 
the fall and winter of 2000-2001). In invasion years they may move as far as the 
southern Mojave Desert, the Imperial Valley, and the lowlands of southeastern Arizona. 
Exceptionally birds may wander west over the crest of the Sierra Nevada, e.g., in 
Yosemite National Park at Tuolumne Meadows (Gaines 1988) and McGurk Meadow, 
where Dunn observed one on 15 September 1996. There may be even more 
movement into the Pacific coast region, but vagrants there would be more difficult to 
detect than in areas where any scrub-jay is unusual. Eastward movements of California 
Scrub-Jays are far more limited, but birds of this group have occurred, for example, on 
the northern Mojave Desert at Galileo Hill, Kern Co., California (M. T. Heindel in lift.), 
and R. Higson collected one of the subspecies obscura in El Centro, Imperial Co., 
California 1 0 August 1 989 (SDNHM 45999). 

The California and Woodhouse's groups of scrub-jays were lumped (American 
Ornithologists’ Union 1931) in an era when the polytypic species concept first gained 
wide acceptance. Given the recognition of species status for the isolated Island and Honda 
scrub-jays, similar status for the Woodhouse’s and California groups is perhaps not 
unwarranted. Indeed, Woodhouse’s and California Scrub- Jays seem to us more distinct 
from one another than do the Island and California Scrub-Jays. Such a revision, however, 
requires further elucidation of the status of the sumichrasti group in south-central Mexico. 
Additional field studies of the behavior, vocalizations, ecology, and breeding biology of 
scrub-jays in the region of parapatry in western Nevada will also be enlightening. 

We thank Larry Sansone for obtaining these instructive photos, and Ed Harper for 
sharing some of his photos of Woodhouse’s Scrub-jays from Idaho. Philip Unitt made 
a number of very helpful editorial suggestions. 

LITERATURE CITED 

American Ornithologists’ Union. 1931. Check-list of North American Birds, 4th ed. 
Am. Ornithol. Union, Lancaster, PA. 

American Ornithologists’ Union. 1995. Fortieth supplement to the American Orni- 
thologists’ Union Check-list of North American Birds. Auk 112:819-830. 

Gaines, D. A. 1988. Birds of Yosemite and the east slope. Artemisia Press, Lee 
Vining, CA. 

Peterson, A. T. 1990. Evolutionary relationships of the Aphelocoma jays. Ph.D. 
dissertation, Univ. of Chicago. 

Peterson, A. T. 1992. Phylogeny and rates of molecular evolution in the Aphelocoma 
jays (Corvidae). Auk 109:133-147. 

Phillips, A. R. 1986. Known Birds of North and Middle America, part I. Denver Mus. 
Nat. Hist. Denver. 

Pitelka, F. A. 1951. Speciation and ecologic distribution in American jays of the genus 
Aphelocoma. Univ. Calif. Publ. Zool. 50:195-464. 

Swarth, H. S. 1918. The Pacific Coast jays of the genus Aphelocoma. Univ. Calif. 
Publ. Zool. 17:405-422. 


187 


WFO PUBLICATIONS — 
FUNDRAISING APPEAL 


Dear WFO members, 

We recently established a publication fund to enable us to increase the size 
of Western Birds, include more color photographs in its pages, and to 
initiate a new monograph series. Although the WFO board has already 
contributed to this fund directly and via various fundraising efforts (trip to 
Vera Cruz, raffle of spotting scope), we will not be successful without your 
support. We hope you all appreciate the high standards the journal has set 
and new additions, such as the regular book review section and the featured 
photo on each back cover. To further increase the quality of Western Birds , 
we ask you to consider a generous donation to the publication fund. Just 
think, if each member gave only $25 we could raise close to $25,000 in a 
short period! To support and improve the West’s finest journal of descriptive 
field ornithology, please consider contributing at least $25. 

Please send donations, payable to Western Field Ornithologists, and 
earmarked “Publication Fund,” to Dori Meyers, WFO Treasurer, 6011 
Saddletree Lane, Yorba Linda, CA 92886. Contributions are tax deductible. 

Thanks very much for your ongoing support of WFO and its contributions 
to field ornithology and conservation! 


Dave Shu ford 
Chair, WFO Publications Committee 

Mike San Miguel 
President, WFO 


188 


World Wide Web site: 
www.wfo-cbrc.org 

Quarterly Journal of Western Field Ornithologists 

President: Mike San Miguel, 2132 Highland Oaks Dr., Arcadia, CA 91006; 
sanmigbird@aol.com 

Vice-President: Daniel D. Gibson, University of Alaska Museum, 907 Yukon 
Dr., Fairbanks, AK 99775-6960 

Treasurer/Membership Secretary: Dori Myers, 6011 Saddletree Lane, Yorba 
Linda, CA 92S86 

Recording Secretary: Lucie Clark, 9889 Tahoe Blvd., #56, Incline Village, NV 
89451 

Directors .- Kimball Garrett, Daniel D. Gibson, Bob Gill, Gjon Hazard, Dave 
Krueper, Mike San Miguel, W. David Shuford, Mark K. Sogge, David Yee 

Editor: Philip Unitt, San Diego Natural History Museum, P.O. Box 121390, San 
Diego, CA 92112-1390; birds@sdnhm.org 

Associate Editors: Daniel D. Gibson, Robert A. Hamilton, Ronald R. LeValley, 
Tim Manolis, Kathy Molina, Mark K. Sogge 

Graphics Manager: Virginia P. Johnson, 4637 Del Mar Ave., San Diego, CA 92107 

Photo Editor: Peter La Tourrette, 1019 Loma Prieta Ct., Los Altos, CA 94024 

Featured Photo: Robert A. Hamilton, 34 Rivo Alto Canal, Long Beach, CA 90803 

Book Reviews: Steve N.G. Howell, Point Reyes Bird Observatory, 4990 Shoreline 
Highway, Stinson Beach, CA 94970 

Secretary, California Bird Records Committee: Guy McCaskie, P.O. Box 275, 
Imperial Beach, CA 91933-0275; guymcc@pacbell.net 

Chairman, California Bird Records Committee: Richard A. Erickson, LSA Associates, 
1 Park Plaza, Suite 500, Irvine, CA 92614; richard.erickson@lsa-assoc.com 


WESTERN BIRDS 


Membership dues, for individuals and institutions, including subscription to Western 
Birds: Patron, $1000.00; Life, $400.00 (payable in four equal annual installments); 
Supporting, $60 annually; Contributing, $34 annually; Family, $26; Regular U.S. 
$22 for one year, $41 for two years, $60 for three years, outside U.S. $27 for one 
year, $51 for two years, $73 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 orders for back issues and special publications to the Treasurer. Make 
checks payable to Western Field Ornithologists. 

Back issues of Western Birds within U.S. $24 per volume, $6.00 for single issues, 
plus $1.00 for postage. Outside the U.S. $30 per volume, $7.50 for single issues. 

The California Bird Records Committee of Western Field Ornithologists recently 
revised its 10-column Field List of California Birds (January 2000). The last list covered 
606 accepted species; the new list covers 613 species. Please send orders to WFO, 
c/o Dori Myers, Treasurer, 6011 Saddletree Lane, Yorba Linda, CA 92886. 
California addresses please add 7.75% sales tax. 

Quantity: 1-9, $1.50 each, includes shipping and handling. 10-39, $1.30 each, add $2.00 
for shipping and handling. 40 or more, $1.15 each, add $4.00 for shipping and handling. 


Published December 14, 2001 


ISSN 0045-3897 





Vol. 32, No. 4, 2001 




Volume 32, Number 4, 2001 


Breeding Status of the Black Tern in California 

IV. David Shuford, Joan M. Humphrey, and Nadav Nur 189 

Idaho Black Swifts: Nesting Habitat and a Spatial Analysis of Records 
R. Kasten Dumroese, Mark R. Mousseaux, 

Shirley Horning Sturts, Daniel A. Stephens, 

and Paul A. Holick 218 

NOTES 

Detections of California Black Rails in the Colorado River Delta, Mexico 


Osvel Hinojosa-Huerta, William W. Shaw, 

and Stephen DeStefano 228 

Book Reviews Steve N, G. Howell, Peter Pyle 233 

Featured Photo Kimball L. Garrett 237 

President’s Message Mike San Miguel 238 

Index Philip Unitt 240 


Cover photo by © Brian E. Small of Los Angeles, California: Sharp-tailed 
Sandpiper (Calidris acuminata), Bolsa Chica Ecological Preserve, Cali- 
fornia, March, 2000. 

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 include 
distribution, migration, status, identification, geographic variation, conservation, behavior, 
ecology, population dynamics, habitat requirements, the effects of pollution, and 
techniques for census ing, 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 the Rocky Mountain and Pacific 
states and provinces, including Alaska and Hawaii, western Texas, northwestern Mexico, 
and the northeastern Pacific Ocean. 

Send manuscripts to Kathy Molina, Section of Ornithology, Natural 
History Museum of Los Angeles County, 900 Exposition Blvd., Los Angeles, CA 90007. 
For matter of style consult the Suggestions to Contributors to Western Birds (8 pages 
available at no cost from the editor) and the Council of Biology Editors Style Manual 
(available for $24 from the Council of Biology Editors, Inc., 9650 Rockville Pike, 
Bethesda, MD 20814). 

Reprints can be ordered at author’s expense from the Editor when proof is returned or 
earlier. 

Good photographs of rare and unusual birds, unaccompanied by an article but 
with caption including species, date, locality and other pertinent 
information, are wanted for publication in Western Birds. Submit photos and 
captions to Photo Editor. Also needed are black and white pen and 
ink drawings of western birds. Please send these, with captions, to 
Graphics Manager. 


WESTERN BIRDS 


Volume 32, Number 4, 2001 



BREEDING STATUS OF THE BLACK TERN 
IN CALIFORNIA 

W. DAVID SHUFORD, JOAN M. HUMPHREY, and NADAV NUR, Point Reyes Bird 
Observatory, 4990 Shoreline Highway, Stinson Beach, California 94970 


The Black Tern is the familiar spirit of all fresh-water swamps in California 
north of the Tehachipe — Dawson (1923) 

ABSTRACT: We surveyed breeding Black Terns throughout California in 1997 
and 1998, following winters of very high runoff. We estimated the state’s nesting 
population at about 4150 pairs (±30%), of which 47% were in northeastern 
California and 53% in the Central Valley, The 1940 pairs in northeastern California 
were at 60 sites; 59% were at 10 sites and 70% were in Modoc County. State and 
federal wildlife refuges supported <4% of the regional population; the rest were 
mostly on U.S. Forest Service and private lands. Low emergents, primarily spikerush 
(Eleocharis spp.) and Juncus spp., dominated most nesting marshes in northeastern 
California. Percent cover of emergents (vs. open water) was >80% at 68% of breeding 
sites. About 90% of the Central Valley breeding population was in Sacramento Valley 
rice fields. The rest were in the San Joaquin Valley, primarily in flooded agricultural 
fields with residual crops or weeds and secondarily in rice fields. State, federal, or 
private refuges or reserves held <1% of Central Valley terns. 

Currently the Black Tern is extirpated locally at Lake Tahoe and in the Sacra- 
mento-San Joaquin River Delta. In the San Joaquin Valley, formerly a center of 
abundance, terns typically now breed mainly in two small areas of rice fields in the San 
Joaquin Basin. The Black Tern is quasi-extirpated in the Tulare Basin, where it nests 
irregularly and locally in ephemeral habitats, mainly in extremely wet years. The 
160,000 to 200,000 ha of rice currently planted annually in the Sacramento Valley 
may far exceed the average amount of natural shallow-water habitat available there 
before agriculture. 

We recommend a statewide survey of the California breeding population about 
once every 10 years, during typical climatic and habitat conditions, and monitoring for 
population trends annually. Conservation should focus on restoring, enhancing, and 
providing long-term protection for suitable wetlands and on maintaining isolation of 
colonies from humans and ground predators. Given the scarcity of water in the 
Central Valley in most summers, efforts to enhance tern habitat may be most fruitful 
in years of exceptional runoff. Research is needed on nesting and foraging ecology, 
habitat suitability, demography, limiting factors, population response to changing 
water conditions, and the value of rice fields versus wetlands as breeding habitat. 


Western Birds 32:189-217, 2001 


189 




BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


Although the New World subspecies of the Black Tern ( ChHdonias niger 
surinamensis ) breeds widely across southern Canada and the northern 
United States, concern has been expressed because the population declined 
across the contintent (Dunn and Agro 1995, Peterjohn and Sauer 1997, 
Shuford 1999) during a period of rapid wetland loss (Dahl et al. 1997). 
Breeding Bird Survey data imply that the Black Tern declined across the 
range surveyed, though not significantly, at an average rate of -1.6% (95% 
confidence interval -4. 4-1.1) annually (-41.3% overall) from 1966 to 1999 
(Sauer et al. 2000). Although extirpated from only two states, the Black Tern 
has declined in at least 14 of the 34 states, provinces, and territories where 
it currently breeds (Shuford 1999). Consequently, the Black Tern is listed as 
threatened or endangered in six states and variously designated of conserva- 
tion concern in 18 other states or provinces. For the United States overall, 
the Black Tern is listed as a “migratory nongame bird of management 
concern” (USFWS 1995), whereas in Canada it has no official status despite 
recommendations for listing as “threatened” by Gerson (1988) and “vulner- 
able” by Alvo and Dunn (1996). Knowledge of the Black Tern’s status is 
poor in the western United States, including California, where anecdotal 
information led to its listing as a species of special concern (CDFG 1992). 

To fill gaps in knowledge of breeding Black Terns in California, Point 
Reyes Bird Observatory surveyed populations statewide in 1997 and 1998 
as the focus of a project to assess the status of seven species of “inland- 
breeding seabirds” (Shuford 1998, Shuford et al. 1999). Here we report 
current statewide population estimates, breeding distribution, breeding 
phenology, and habitat associations of the Black Tern and compare them 
with the historical record. We also make recommendations for conservation, 
management, and long-term monitoring of the Black Tern in California. 

STUDY AREA AND METHODS 

Prior to field work, we searched the published and unpublished literature 
and contacted various field biologists to identify historic and potential 
breeding habitats of the Black Tern in California. We cite data from 
Audubon Field Notes (AFN) and American Birds (AB) by volume and page 
number and unpublished data from notebooks of the editors of the Middle 
Pacific Coast region of North American Birds as MPCR files. 

In the field, we contacted additional biologists for further information on 
potential breeding habitat. We later obtained egg-set data on Black Terns 
from the California Academy of Sciences (CAS), Los Angeles County 
Museum of Natural History (LACM), Moore Laboratory of Zoology (MLZ), 
Museum of Vertebrate Zoology (MVZ), San Bernardino County Museum 
(SBCM), San Diego Natural History Museum (SDNHM), Santa Barbara 
Museum of Natural History (SBMNH), and Western Foundation of Verte- 
brate Zoology (WFVZ). 

We varied field survey methods by region, to match local logistical 
constraints, and timed surveys to follow the passage of most migrants and 
begin with the initiation of nesting. 


190 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


Northeastern California 

The study area here included valleys of the Cascade Range, Klamath 
Mountains, and Sierra Nevada, the Modoc Plateau, and the Great Basin 
desert, that is, eastern Siskiyou, northeastern Trinity, eastern Shasta, 
Modoc, Lassen, Plumas, Sierra, and El Dorado counties. Potential Black 
Tern habitat in marshes, lakes, and reservoirs is scattered widely, primarily 
from 4000 to 6000 feet (1220-1830 m) elevation in intermountain valleys 
or in depressions in the Modoc Plateau. Precipitation, falling mostly from 
October through April as rain and snow, in the climate year (1 July-30 June) 
1996-97 was 114.3 cm in the Sacramento Drainage Division and 79.8 cm 
in the Northeast Interior Basins Division (results from weather stations 
throughout the region averaged). Combined, these divisions encompass 
most of the study area. As these figures represent 119% and 147%, 
respectively, of the long-term (n = 104) averages for these areas (Western 
Regional Climate Center; http://www.wrcc.dri.edu/divisional.html), wet- 
lands in the study area were well supplied with water in summer 1997. 

From 18 May to 19 July 1997, Shuford and colleagues surveyed most 
potential breeding habitat in northeastern California for Black Terns; all sites 
surveyed were listed by Shuford (1998). In addition, K. Laves and Shuford 
surveyed the south shore of Lake Tahoe, El Dorado County, on 14 June 
1998, and M. McVey surveyed most potential breeding wetlands in the 
Shasta Valley, Siskiyou County, in spring and summer 1998 and 1999. 
Shuford and colleagues also opportunistically resurveyed various sites in the 
summers of 1998 to 2001 , as indicated in the text or Table 1 . We conducted 
surveys mostly on foot and occasionally by kayak or canoe. We were unable 
to survey only a few areas with high potential for nesting terns. We did not 
survey Picnic Grove and Lakeshore reservoirs in the Devil’s Garden Ranger 
District of Modoc National Forest because of logistical difficulties, and we 
w r ere denied access to a few private holdings, the largest being Steele 
Swamp, Modoc County, and Dixie Valley, Lassen County. 

Early in the season it was possible at many sites to count both adult Black 
Terns using the wetland and all or most of their nests. We soon realized we 
would be unable to count all nests at all sites because of the time needed and 
our inability to count nests accurately once chicks began to leave their nests 
shortly after hatching. Thus, depending on circumstances, we obtained 
three types of counts and used three corresponding methods to estimate 
numbers of pairs of terns, presented here in order of their apparent 
reliability and annotated with respect to biases. When data are available to 
make more than one estimate, we present only the method of apparent 
highest reliability. 

(1) Total nests: obtained by systematically walking all of a marsh and 
locating all or most nests by visually scanning areas where terns were 
agitated, flushing adults from nests, or following terns back to nests. At sites 
where a thorough search was impractical, we made partial nest counts, 
which served only to document breeding. We estimated the number of 
breeding pairs as the total number of nests at the time of the survey. This 
method may underestimate the total because of the difficulty of finding all 


191 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


Table 1 Numbers of Adult Black Terns, Nests, and Estimated Pairs from 
Surveys of Wetlands in Northeastern California in 1997 


Site Survey date 

Number of adults' 1 
Disturbed Undisturbed 

Number of nests b 
Total Partial 

Estimated 

pairs c 

Siskiyou County 
Butte Valley WA 

14 July 

22 



2 

ll 2 

Butte Valley National 
Grasslands 

14 July 

0 

2 


1 

2 3 

Grass Lake d 

12 July 

— 

28 

— 

2+ 

22 3 

Orr Lake 

30 May 

— 

8 

— 

— 

6 3 

24, 26 June 

— 

6 

— 

— 

— 

Dry Lake (T44N, R1W, 
sect. 30, 31) 12 July 


4 


2 

3 3 

Lower Klamath NWR e 
Unit 4E 

18 June 

~73 

65 


3 

37 2 

Unit 4D 

18 June 

— 

18 

12+ 

— 

12 1 

Barnum Flat Reservoir 

1 July 

— 

68 

— 

2+ 

54 3 

Subtotal 

Modoc County? 

Dry Lake (T44N, R6E, 
sect. 4, 5) 

20 June 


12 



147 

9 3 

Fourmile Valley 

27 May 

38 

27 

27 

— 

27 1 

Wild Horse Valley 

28 May 

6 

8 

3 

— 

3 1 

Buchanan Flat 26-27 May 

36 

29 

21 

— 

21 1 

Weed Valley 

3 June 

— 

203 

— 

6 

160 3 

Baseball Reservoir 

26 May 

47 

47 

42 

— 

42 1 

Dry Valley Reservoir 

25 May 

58 

— 

30 

— 

30 1 

Hager Basin (North) 

24 May 

22 

13 

14 

— 

14 1 

Hager Basin (South) 

24 May 

51 

21 

18 

— 

18 1 

Telephone Flat 
Reservoir 

31 May 

23 


7 


7 1 

South Mountain 

Reservoir 31 May-1 June 

6 


2 


2 1 

Pease Flat 9 

21 May 

— 

1 

0 

— 

— 


17 July 

— 

~60 

— 

— 

47 3 


18 July 

— 

19 

2+ 

— 

— 

Mud Lake (T46N, 
R12E, sect. 16) 

22 May 

26 

8-10 

16 


16 1 

Crowder Mt. 
Reservoir 

1 June 


41+ 

40 


40 1 

Whitney Reservoir 

20 June 

10 

— 

— 

— 

5 2 

Hackamore Reservoir 

20 June 

20 

— 

— 

4 

10 2 

Spaulding Reservoir 

21 June 

40 

— 

— 

10 

20 2 

Beeler Reservoir 

22 June 

26 

— 

— 

10 

13 2 

Pinky’s Pond 

22 June 

14 

— 

— 

3 

7 2 

Widow Valley 

22 June 

— 

82 

— 

1 

64 3 

Bucher Swamp 

22 June 

— 

122 

— 

5 

96 3 

Six Shooter Tank 

23 June 

18 

12 

— 

1 

9 2 


(continued) 


192 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


Table 1 ( continued ) 


Site 

Survey date 

Number of adults 0 
Disturbed Undisturbed 

Number of nests b 
Total Partial 

Estimated 

pairs 0 

Deadhorse Hat 
Reservoir 

23 June 


45 


1 

35 3 

Surveyors Valley 

23 June 

— 

35 

— 

1 

28 3 

Boles Meadow 

7 June 

— 

211 

— 

15 

166 3 

Hetcher Creek 
Reservoir 

16-17 June 


48 

31 


31 1 

Jack’s Swamp 

5 June 

— 

64 

26 

— 

26 1 

Dead Horse Reservoir 29 May 

— 

7+ 

11 

— 

ll 1 

Jesse Valley 

26 June 

— 

13 

— 

4 

10 3 

Whitehorse Flat 
Reservoir 

1 July 


37 


4+ 

29 3 

Egg Lake 30 June-1 July 

— 

343 

— 

1+ 

270 3 

Taylor Cr. wetlands 

30 June 

— 

128 

— 

2 

101 3 

Subtotal 

Lassen County 
Muck Valley 

2 July 


53 


5 

1367 
42 3 

Hoover Hat Reservoir 3 July 

— 

7 

— 

— 

6 3 

Moll Reservoir 

27 June 

34 

20 

— 

3+ 

17 2 


16 July 

13 

— 

— 

— 

— 

Okendine’s Spring 

27 June 

9 

5 

— 

— 

5 2 


16 July 

— 

0 

— 

— 

— 

Ash Valley (main) 

27 June 

— 

66 

— 

— 

52 3 

Ash Valley (SE) 

19 July 

— 

9 

— 

— 

7 3 

Red Rock Lakes 
complex 

26-27 June 


72 


2+ 

57 3 

Boot Lake 

25-26 June 

— 

15 

— 

8 

12 3 

Poison Lake h 

5 July 

76 

43 

— 

2 

38 2 

Dry Lake 
(Grass Valley) 

10 June 



6 

_ 

_ 

5 3 


5 July 

— 

0 

— 

— 

— 

Straylor Lake 

26 May 

— 

11 

— 

— 

9 3 


11 July 

— 

1 

— 

— 

— 

Long Lake (T34N, 
R8E, sect. 22) 

26 May 


6 



5 3 


1 1 July 

— 

0 

— 

— 

— 

Ashurst Lake 

26 May 

— 

7 

— 

— 

? 


13 June 

— 

2 

— 

— 

2 3 


10 July 

— 

2 

— 

— 

— 

Gordon Lake 

9 June 

— 

12 

— 

— 

9 3 


10 July 

— 

10 

— 

— 

— 

Pine Creek wetlands 
(T32N, R9E, 
sect. 28) 

10 June 


9 



7 3 


10 July 

— 

5 

— 

— 

— 

McCoy waterpit 

9 June 

— 

12 

— 

— 

9 3 


10 July 

— 

0 

— 

— 

— 


(continued) 


193 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


Table 1 ( continued ) 


Site 

Survey date 

Number of adults 0 
Disturbed Undisturbed 

Number of nests b 
Total Partial 

Estimated 

pairs c 

Eagle Lake' 

8-9 July 



142 

- 3+ 

112 3 

Willow Creek WA 

10 June 

— 

13 

— — 

10 3 

Horse Lake 
Mountain Meadow 

8 July 

15 

15 

— 1+ 

8 2 

Reservoir 
Honey Lake N 

7 July 

22 

20 



ll 2 

(private) 

Subtotal 

Total 

15 June 

5 

5 

— 1 

3 2 

426 

1940 


°Numbers of adults from either disturbed or undisturbed counts (see Methods). 

,; 'N umbers of nests from either total or partial counts (see Methods). 

c Numbers of pairs estimated by three methods, listed here in order of apparent reliability, on the 
basis of ’ numbers of total nests, 2 counts of total disturbed adults, and 3 counts of total undisturbed 
adults (see Methods). When data enable more than one type of estimate, the estimate presented 
is from the method of highest apparent reliability 

d A count of 13 undisturbed adults at Grass Lake on 24 June 1999 yields an estimate of 10 
breeding pairs that year; no terns were seen there on 28 May during the drought year of 2001. 

e Counts of undisturbed adults of 54 in Unit 6B, 220± in Unit 6C, 10 in Unit 10A, and 146 in 
Unit 12C on 21 June 2001 yield estimates of 42, 173, 8, and 115 breeding pairs in those units, 
respectively, that year. 

J 'A count of 57 undisturbed adults on 21 June 1999 yields an estimate of 45 breeding pairs at 
Lost Valley, which was mostly dry and devoid of waterbirds on 22 June 1997. 

3 A count of 23 undisturbed adults at Pease Flat on 20 June 1999 yields an estimate of 18 
breeding pairs that year. 

h Five adults at Poison Lake on 23 June 1999 showed no signs of site attachment or other 
evidence of breeding. 

'A count of 160 undisturbed adults at Eagle Lake on 23 June 1999 yields an estimate of 126 
breeding pairs that year. 


nests, particularly in large marshes, and, because of asynchronous egg 
laying among colonies or subcolonies, some birds may not have initiated or 
completed laying at the time of surveys. 

(2) Total disturbed adults: taken from within the colony when the observer 
(or a predator) disturbed birds, and all or most terns, including adults 
attending nests, joined a mobbing flock around the intruder. We estimated 
the number of pairs as the best count of total disturbed adults rounded to the 
nearest even number and divided by two. This method does not account for 
adults foraging far from the colony, hence not attracted to mobbing flocks, 
adults not joining the mobbing flock, or failed breeders having left the 
colony. We did not use this method at large wetlands, where we were unable 
to obtain accurate counts because of many adults swirling rapidly around the 
observer and terns continuously joining or leaving the mobbing flock as they 


194 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


flushed from, or returned to, nests as the intruder approached or left their 
“zone of concern.” 

(3) Total visible undisturbed adults: taken from the edge of the wetland or 
from a vantage point within where the observer did not attract mobbing 
adults. We estimated the number of breeding pairs as the best count divided 
by 1.27 {standard error 0. 16), the mean ratio of undisturbed adults counted 
to nests at the 10 sites where we collected both types of data (317 total 
adults, 247 total nests) during the incubation period. The method’s primary 
biases, adjusted by a correction factor, are that it underestimates total adults 
or pairs because of the difficulty of seeing many incubating and roosting 
terns obscured by vegetation or other visual obstructions and does not 
account for adults foraging away from the colony. Also, the number of visible 
adults may increase as nests hatch and adults spend more time foraging, or, 
conversely, may decrease as nests fail and adults disperse. 

To characterize habitat at each breeding site, observers recorded the 
dominant species of emergent vegetation and visually estimated the percent 
cover of both emergent vegetation and open water. We estimated these 
variables for the entire wetland, except at managed refuges where we 
estimated them for just the diked wetland units in which terns were breeding 
rather than for the entire complex of units. 

Central Valley 

The Central Valley, surrounded by mountains except at its western outlet 
into the San Francisco Bay estuary, averages about 644 km long and 64 km 
wide. It is divided into the Sacramento Valley, draining south, the San 
Joaquin Valley, draining north, and the Sacramento-San Joaquin River 
Delta where these rivers converge. The Sacramento Valley is further divided 
into the Colusa, Butte, Sutter, American, and Yolo drainage basins, the San 
Joaquin Valley into the San Joaquin Basin and the (usually closed) Tulare 
Basin. 

Over 90% of the Central Valley’s presettlement wetlands have been lost 
(Frayer et al. 1989, Kempka et al. 1991), and the dominant land use is 
agriculture. Hence, breeding habitat for waterbirds typically is scarce. 
Precipitation, falling mainly from October through April (as rain, or snow in 
adjacent mountains), is highly variable. Despite a massive reservoir storage 
and drainage system and high summer temperatures, in the wettest years 
extensive shallow water can persist through the breeding season. Precipita- 
tion in the climate year 1997-98, during El Nino, was 153.7 cm in the 
Sacramento Drainage Division and 86.9 cm in the San Joaquin Drainage 
Division, representing 161% and 169%, respectively, of the long-term (n = 
104) averages for these regions (Western Regional Climate Center; http:// 
www.wrcc.dri.edu/divisional.html). Hence the breeding season of 1998 
provided some of the best conditions for nesting waterbirds in the Central 
Valley since the 1950s. Shallow-water breeding habitat increased primarily 
in the Tulare Basin, where large areas of agricultural land were flooded, 
intentionally or unintentionally, and secondarily near Los Banos, Merced 
County, on refuges and in flood-control bypasses. 

Large areas of cultivated rice fields in the Sacramento Valley, and smaller 
areas in the delta and San Joaquin Basin, typically provide potential nesting 


195 


BREEDING STATUS OF THE BLACK TERN IN CAUFORNIA 


habitat for the Black Tern. In 1998, the intense and extended rainy season 
delayed rice planting in the Sacramento Valley by about three weeks, and 
only about 75% of the crop had been planted at the time of our surveys (60% 
by 31 May, 90% by 7 June; 9 June 1998 “Weekly Weather and Crop 
Bulletin,” Natl. Agric. Statistics Serv., Agric. Statistics Board, U.S. Dept. 
Agric.). Other habitats in the Central Valley sometimes suitable for breeding 
terns include managed wetlands on refuges and duck clubs (limited summer 
water) and floodwater storage or recharge facilities (e.g., South Wilbur Flood 
Area, Kern Fan Element Water Bank). The average May to July tempera- 
tures of 62.5° F (16.9° C) and 66.5° F (19.2° C) for the Sacramento and San 
Joaquin drainage divisions, respectively, were the second lowest and lowest 
on record (Western Regional Climate Center, http://www.wrcc.dri.edu/ 
divisional.html; n = 105). These were ideal conditions for both surveying in 
this typically very hot climate and delaying desiccation of the tern’s breeding 
habitats. 

Because of the 187,000 ha of rice planted in the Sacramento Valley in 
1998, and limited access to private lands, we were unable to survey all 
potential breeding habitat. Instead, from 29 May to 10 June (also 18 June), 
seven observers conducted roadside transect surveys along most lightly 
traveled roads in the Sacramento Valley rice country (Glenn and Butte 
counties south to Yolo County) to estimate densities of Black Terns breeding 
there. Single observers covered routes by driving roads at 24 to 32 km/hr 
and counting terns seen within the primary census zone of 0. 1 mile (160 m) 
on each side of the road. We surveyed without the aid of binoculars, except 
when needed to confirm identifications or estimate numbers accurately. We 
surveyed from 0600 to 1000 hours, later if temperatures were under 29° C; 
as temperatures often were below normal, this meant sometimes all day. We 
halted during strong winds (>24 km/hr) or persistent rain. 

Observers recorded weather conditions, start and stop times, route cov- 
ered, miles driven, distance surveyed (each side of the road tallied separately), 
number and age of terns, location(s) and habitat type where terns were 
observed, and any breeding evidence, including details of nest locations. 
Observers also recorded any terns seen beyond the primary census zone or 
off survey routes, but we did not use these data to calculate densities of 
breeding terns in rice fields. Observers recorded all observations of terns on 
maps in the field for later use in mapping patterns of breeding distribution. 
Observers were asked to try to confirm nesting by returning to make 
observations after finishing a survey or on a subsequent visit. We considered 
confirmed nesting all observations of nests with eggs, adults sitting in 
incubation posture on an apparent nest, adults feeding non-flying young, 
adults repeatedly carrying food to the same spot (presumably to an unseen 
chick), or nonflying or very weakly flying young. Because of delayed planting, 
at the time of our surveys little growing rice had emerged above water (15% 
emerged on 31 May, 35% on 7 June; 9 June 1998 “Weekly Weather and 
Crop Bulletin,” Natl. Agric. Statistics Serv., Agric. Statistics Board, U.S. 
Dept. Agric.), and hence most terns sitting on nests were still visible. 

We calculated densities of Black Terns in rice fields by first multiplying the 
distance surveyed on each route by 160 m, the width of the primary census 
zone, then converting this to hectares of habitat surveyed. We next deter- 


196 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


mined the mean density of terns per 100 hectares for each county (or 
grouping of counties) by calculating the mean density for all of the county’s 
routes weighted by distance surveyed. We estimated the total number of 
breeding terns in each county by multiplying tern density per county times 
the number of hectares of planted rice per county (M. Leighton, Calif. Agric. 
Statistics Serv. in litt . ; National Agric. Statistics Serv., http:// 
www.nass.usda.gov: 100/ipedb/), adjusted by a correction factor of 0.75, 
the estimated proportion of rice planted at the time of our surveys. Field 
observations did not suggest any evidence of avoidance of, or attraction to, 
roads by nesting terns, which might have biased our estimates. 

By contrast, in the delta, San Joaquin Valley, and in habitats in the 
Sacramento Valley other than rice fields, we surveyed from the ground or by 
boat all known potential breeding habitat for Black Terns. See Shuford et al. 
(1999) for a list of sites surveyed. In 1998, we surveyed the entire 807, 
1817, 2220, and 1211 ha of planted rice in Stanislaus, San Joaquin, 
Merced, and Fresno counties, respectively, rather than sampling them as in 
the Sacramento Valley. We counted mainly visible undisturbed adults and, 
rarely, total nests via thorough nest searches. We did not count total 
disturbed adults or total nests at most sites because of the potential to 
damage crops by doing so. Partial nest counts at many sites served only to 
document breeding. Hence, depending on available data, we estimated 
numbers of pairs of Black Terns by either the “total nests” or “undisturbed 
adults” methods described above for northeastern California. In the latter 
case, the correction factor used for the Central Valley was that derived in 
northeastern California in 1997. 

RESULTS 

Population Size and Distribution 

We estimated about 4153 pairs of Black Terns nested in the state in 1997 
and 1998, 46.7% in northeastern California and 53.3% in the Central 
Valley. 

Northeastern California. An estimated 1940 pairs nested at 60 widely 
scattered sites in this region (Table 1, Figure 1). About 70.5%, 22.0%, and 
7.6% of that population was located in Modoc, Lassen, and Siskiyou 
counties, respectively. The 10 sites with >50 pairs of terns, which combined 
held 58.7% of the regional population, were Barnum Flat Reservoir, 
Siskiyou County; Weed Valley, Widow Valley, Bucher Swamp, Boles 
Meadow, Egg Lake, and Taylor Creek wetlands, Modoc County; and Ash 
Valley (main), Red Rock Lakes complex, and Eagle Lake, Lassen County. 

Central Va//ey. Of the estimated 2213 pairs of Black Terns that bred in 
the Central Valley in 1998, 89.8% were in the Sacramento Valley and 
10.2% in the San Joaquin Valley (Tables 2 and 3, Figures 2 and 3). From 
roadside surveys, we estimated that about 2523 ± 754 (1769-3277) adult 
terns, or about 1987 (1393-2581) pairs, bred in Sacramento Valley rice 
fields (Table 2). Although the birds were spread widely, the largest numbers 
were in the northern Colusa Basin (Table 2, Figure 2). In the San Joaquin 
Valley, about 75 pairs bred at five sites in the San Joaquin Basin and 151 
pairs bred at six sites in the Tulare Basin (Table 3). 


197 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 



Figure 1 . Distribution and size (number of pairs) of Black Tern colonies in northeastern 
California in 1997 (see Table 1), plotted with historical (1899-1960) and other recent 
(1961-1998) records of confirmed breeding (see Appendices 1 and 2). 


198 





BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


Table 2 Estimated Numbers of Black Terns Breeding in the Sacramento 
Valley from Roadside Surveys of Rice Fields, 29 May-10 June 1998 


County 

Hectares 
planted rice a 

Survey 
routes (n) 

Distance 
surveyed (km) b 

Terns per 
100 ha (± SE) C 

Terns estimated 
(± SE) d 

Colusa 

Sutter-Yolo- 

36,637 

38 

370.2 

2.67 ± 0.67 

978 ± 245 

Sacramento 6 

36,485/ 

26 

284.5 

0.70 ± 0.23 

255 ± 84 

Butte 

26,645 

10 

234.5 

0.85 ± 0.31 

226 ± 82 

Glenn 

25.131 

44 

352.8 

3.68 ± 1.56 

925 ± 392 

Yuba 

11,294 

16 

122.1 

1.22 ± 0.44 

138 ± 50 

Placer 

4239 

4 

47.0 

0.00 ± 0.00 

0 

Tehama 3 

363 

0 

0 

— 

0 

Totals 

140,794 

138 

1411.11 

1.80 ± 0.54” 

2523 ± 754 


0 Planted rice acreage adjusted to account for estimate that only 75% of the total for the year had 
been planted at the time of our surveys (see Methods). 

b Each side of road tallied separately. 

c Density estimates for each county are means of survey routes, weighted by distance surveyed. 
SE, standard error. 

a Tern numbers estimated by multiplying densities on roadside surveys times acreage of available 
rice fields, Standard errors represent variation in densities of terns on survey routes but do not 
account for possible error in the estimate of the amount of planted rice at the time of tern 
surveys. 

e Data for these counties pooled because of small sample sizes for Yolo and Sacramento counties. 
Number of survey routes and distance surveyed, respectively, per county: Sutter, 15, 204.0; 
Yolo, 10, 69.4; Sacramento, 1, 11.1. 

/Numbers of hectares planted rice per county at time of survey: Sutter, 27,553; Yolo, 6177; 
Sacramento, 2755 

9 Although we surveyed no routes in Tehama Co. in 1998, coverage since the 1970s there has 
shown no evidence of terns there in the breeding season (S. Laymon in litt). If terns breed there 
now the number would be small: 7 or 13 if densities were the same as for the entire Sacramento 
Valley or for Glenn County, respectively. 

h Mean of county density estimates, weighted by hectares of rice. 


Nesting Phenology 

In northeastern California in 1997, we observed the first nests with eggs 
at Mud Lake, Modoc County, on 22 May and the first hatched young at 
Fletcher Creek Reservoir on 17 June. On the basis of the species’ 19-21 
day incubation period (Dunn and Agro 1995), eggs at Mud Lake likely 
hatched by at least 12 June. Collection dates of egg sets for the region 
extend from 23 May to 30 June and reach a peak in early June (Figure 4). 
Observations in the Central Valley in 1998 were inadequate for assessing 
breeding phenology; dates of egg sets collected there extend from 3 May to 
5 July and reach a peak from late May to early June (Figure 4). 

Habitat Associations 

Northeastern California. Of 60 breeding sites in northeastern Califor- 
nia, 52 (86.7%) were marshes dominated by low (<1 m) emergents, six 


199 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


Table 3 Numbers of Adult Black Terns, Nests, and Estimated Pairs in the 
San Joaquin Valley in 1998 

Number of nests fa 

Estimated 

Site Survey date Number of Adults' 3 Total Partial Pairs' 3 


Merced County 
Rice fields SW of Merced 

22 June 30 — — 

3 July 25 — 2 

Raccoon Marsh, West Bear Creek Unit, San Luis NWR 

22 June 4 2 — 

Cinnamon Slough, Merced NWR 

23 June 4 2 — 

Fresno County 

Rice fields S of Dos Palos, Merced Co. 

22-23 June 58 

James Bypass S of James Rd. 

1 July 2 

Kings County 

S of Hacienda Ranch Hood Basin (T24S, 

19 June 69 

S of Hacienda Ranch Hood Basin (T24S, 

19 June 28+ 

13 July — 

Tulare County 

Vicinity jet. Hwy. 43 and Virginia Ave. 

25 June 35+ — 2 

2 mi W of Rd. 40 about 3 mi S of Alpaugh 

23 June 21 — 1 

Just W of Rd. 40 about 4 mi S of Alpaugh 

22 June 32 — 3 

Kern County 

Kern Fan Element Water Bank (pond W-2), W of 1-5 

20 June 7 — 1 


— 5 

1 — 

R21E, sect. 31, 32) 

— 7 
R21E. sect. 28, 33) 

— 3 

— 3-4 


24 2 


2 1 

2 1 


46 2 

l 1 


54 2 

22 2 


28 2 

16 2 

25 2 


6 2 


Total 


226 


“Numbers of adults from counts of undisturbed birds {see Methods). 

^Numbers of nests from either total or partial nest counts (see Methods). 

“Numbers of pairs estimated by two methods, listed here in order of apparent reliability, on the 
basis of founts of total nests or 2 c.ounts of total undisturbed adults (see Methods). When data 
enable more than one type of estimate, the estimate presented is from the method of highest 
apparent reliability. 


200 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 



Figure 2. Distribution of breeding Black Terns in California’s Sacramento Valley and 
Delta in 1998 (see Table 2), plotted with historical (1886-1960) and other recent 
(1961-1999) records of confirmed breeding (see Appendices 1 and 2). Stippling 
denotes areas where rice currently is grown. 


201 



BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 



Figure 3. Distribution and size (number of pairs) of Black Tern colonies in California’s 
San Joaquin Valley in 1998 (see Table 3) plotted with historical (1893-1960) and 
other recent (1961-1997) records of confirmed breeding (see Appendices 1 and 2). 
Stippling denotes areas where rice currently is grown. 


202 




BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


100 


CENTRAL VALLEY 
NORTHEASTERN CALIFORNIA 


80 - 

V) 

h 

111 

w 


§ »- 

in 



>. 

>, 

>v 

>. 

>. 

d> 

<12 

03 

CD 

d) 

d) 

J>x 

(0 

CO 

(0 

CO 

(0 

c 

C 

C 

C 

c 

C 

3 

2 

2 

5 

5 

2 

3 

—> 

3 

“3 

3 

“3 

3 

—> 

3 

—> 

3 

“3 

“3 

o 

in 

o 

CM 

in 

CM 

o 

ro 

1 

03 

i 


<y> 

tJ- 

CN 

03 

CM 

Q) 

CD 

T 

CO 

CM 

CO 

CM 

CO 

2 

LO 

i 

o 

■ 

LO 

t — 

i 

O 

CM 

i 

LO 

CM 

C 

3 

3 












O 






co 






CO 


5-DAY PERIODS 


Figure 4. Temporal distribution of egg-set records of the Black Tern in California, 
1886-1960. Data from major California museums (see Methods, Appendix 1). 


(10%) by a mixture of tall (>1 m) and low emergents. At Lower Klamath 
National Wildlife Refuge (NWR), Black Terns nested in shallowly flooded 
basins dominated by residual barley stubble and algal mats and lacking much 
live emergent vegetation. At Boot Lake, Lassen County, in the Warner 
Mountains at 6560 feet (2000 m), the highest colony, breeding habitat was 
dominated by the floating yellow pond-lily ( Nuphar luteum ssp. 
polpsepalum). Of 58 sites with emergent vegetation, 50 (86.2%) were 
dominated or co-dominated by low emergent spikerush ( Eleocharis spp.) or 
Jurtcus spp., seven by a mixture of tall emergents such as tules {Scirpus 
spp.) or cattails ( Typha spp.) and low emergents, and one by a low emergent 
composite ( Arnica sp.). 

Percent cover of emergents was >80% at 41 sites (68.3%), between 60% 
and 80% at nine sites (15%), between 40% and 60% at three sites (5%), 
between 20% and 40% at no sites, and between 0% and 20% at seven sites 
(11.7%). All sites with <20% emergent cover, except Lower Klamath NWR, 
were lakes or reservoirs with mostly open water fringed by marsh vegetation. 
If we had limited estimates of vegetative cover to actual breeding sites, rather 


203 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


than the entire wetland, the proportion of total sites with >80% cover would 
have been higher. 

Central Valley. Of the valleywide total, about 2057 pairs (93.0%) bred in 
cultivated rice fields, 151 (6.8%) in agricultural fields with residual crops and 
weeds and shallow water remaining from winter floods, and 5 (0.2%) in 
emergent wetlands of low stature. Of the four pairs that bred in protected 
areas, two each were at Merced and San Luis NWRs in the San Joaquin 
Basin. All breeding evidence in the Sacramento Valley was from rice fields, 
though one colony in Glenn County was in sedges in the corner of a field 
rather than in the rice itself (Shuford pers. obs.). Of the 226 pairs in the San 
Joaquin Valley, 66.8% were in flooded agricultural fields with residual crops 
or weeds, 31.0% in rice fields, and 2.2% in emergent wetlands of low 
stature. 

DISCUSSION 

Accuracy of 1997-1998 Surveys 

Although unable to estimate the precision of all methods used to survey 
terns in 1997-1998, we suspect our overall population estimate for the 
state was within 30% of the actual number. The 95% confidence interval was 
±60% for surveys in rice fields of the Sacramento Valley (Table 2) versus 
±25% for the “undisturbed adults” method used in northeastern California 
and the San Joaquin Valley (Tables 1 and 3). The precision estimate for the 
latter method, though, is applicable only for the incubation period, when it 
was derived, even though this method was used for data collected through- 
out the breeding season. Although the “total nests” and "disturbed adults” 
methods, also used widely in northeastern California and the San Joaquin 
Valley (Tables 1 and 3), lack estimates of precision both underestimated 
numbers of nesting pairs (see Methods). Another source of underestimation 
was our inability to cover some potential tern nesting habitat in northeastern 
California. Hence our overall estimate is more likely an under- than an 
overestimate of the statewide nesting population during the survey period. 

To have conducted the statewide survey in one rather than two years 
would have been desirable also. We suspect, though, that numbers from the 
1997-98 survey were representative of those statewide in 1998, when we 
surveyed the entire Central Valley and water conditions in northeastern 
California were similar to those in 1997. 

Historical Patterns of Distribution and Abundance 

Past data on the distribution and abundance of breeding Black Terns in 
California are mostly anecdotal (Grinnell and Miller 1944, Cogswell 1977, 
Shuford 1999). Breeding populations were restricted to two distinct areas: 

(1) the Modoc Plateau and mountain valleys of northeastern California and 

(2) the lowlands of the Central Valley. Today the outlines of the breeding 
range remain much the same (Figures 1-3). In northeastern California, 
however, the species is extirpated at Lake Tahoe. In the Central Valley it is 
extirpated in the delta, and in much of the San Joaquin Valley it is either 
extirpated or breeds irregularly, only in exceptionally wet years. Grinnell and 


204 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


Miller (1944) considered the Black Tern a “locally common” breeder in 
California. Evaluating subsequent population trends is very difficult given the 
anecdotal nature of early accounts and the lack of recent Breeding Bird 
Survey data for California sufficient for trend analysis (Peterjohn and Sauer 
1997, Sauer et al. 2000). 

Northeastern California. Historic locations of confirmed breeding for 
this region include Tule Lake and Alturas Meadow, Modoc County; Grass- 
hopper Meadows/Lake and Eagle Lake, Lassen County; and Lake Tahoe, El 
Dorado County (Grinnell and Miller 1944; Appendix 1). Assessing popula- 
tion trends is problematic given the few known historic breeding areas, the 
large number of recent breeding sites (Table 1), few of which have a long 
record of occupancy, and the species’ propensity to shift from site to site 
with fluctuating environmental conditions. Extensive wetland loss, particu- 
larly in the Klamath Basin, may have been partially offset on the Modoc 
Plateau by the creation of shallow reservoirs for livestock grazing and recent 
enhancement for waterfowl (T. Ratcliff, G. Studinski pers. comm.). 

Black Terns have bred at Eagle Lake since at least 1918 (Appendix 1). 
Using a combination of nest counts in habitat visited, the behavior and 
number of terns in suitable habitat not well surveyed for nests, and the extent 
of suitable habitat not visited, Gould (1974, in litt.) estimated 300 breeding 
adults at Eagle Lake in 1970 and 150 in 1971. From nest counts, Lederer 
(1976) estimated 46 breeding adults in 1974; Shaw (1998) estimated 78 in 
1996 and 64 in 1997. Our independent estimate in 1997, based on counts 
of undisturbed adults, was 224 breeding adults (112 pairs; Table 1). A count 
of 160 undisturbed adults on 23 June 1999 yielded an estimate of 252 
breeding adults (126 pairs; Table 1, footnote). Together these numbers may 
reflect year-to-year variation in nesting population size, perhaps mirroring 
changing patterns of emergent vegetation in response to lake levels (G. 
Gould pers. comm.) rather than a population decline followed by recovery, 
or, in part, the variation in survey techniques among observers. 

Black Terns formerly reached their southeastern breeding limit in the 
region at Lake Tahoe, where they nested mainly at Rowlands Marsh near the 
mouth of the Upper Truckee River, El Dorado County (Orr and Moffitt 
1971). That colony once held over 100 pairs, and, prior to 1920, colonies 
of four or five pairs bred near the mouth of Emerald Bay, at Meeks Bay, and 
near Tahoe Vista; “a few pairs” also formerly nested annually west of Tallac 
(Orr and Moffitt 1971; Appendix 1). Habitat loss and degradation from 
development and lowering of water levels eliminated breeding Black Terns at 
Lake Tahoe (Orr and Moffitt 1971, Cogswell 1977, K. Laves pers. comm., 
Shuford pers. obs. in 1998). Today the species reaches its southern limit in 
the Sierra Nevada at Sierra Valley, Plumas and Sierra counties, and at 
Kyburz Hat, Sierra County, where breeding is irregular, particularly at 
Kyburz (Appendix 2). 

Black Terns have been reported breeding in the extensive marshes of 
Sierra Valley since at least the early 197 0s (Appendix 2), primarily by birders 
looking off Marble Hot Springs Road (Dyson Lane), Plumas County. 
Numbers there fluctuate annually (up to 50 on 27 May 1989, J. McCormick 
in litt.), and terns are absent in some drought years, such as 2001 (Shuford 


205 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


pers. obs. , 12-16 June) The species also occurs in the Sierra County portion 
of the marsh (e.g. , up to nine adults ~4 miles north of Sierraville 15-20 June 
1999; J. McCormick, Shuford pers. obs.), but this area gets little coverage. 

Small’s (1994) report of ‘the largest regular [breeding] concentration 
(1000+) in northern California” in the Klamath Basin is unsubstantiated. 
Although Zeiner et al. (1990) mapped the summer range as including the 
Shasta Valley, Siskiyou County, and Small (1994) cited Cedar Lakes, Shasta 
Valley, as a recent breeding area, we know of no documented breeding 
records for that valley (Appendices 1 and 2, R. Ekstrom in litt). M. McVey 
(in litt.) did not find evidence of Black Terns breeding at any of numerous 
wetlands in the Shasta Valley he surveyed in 1998 and 1999. 

Central Valley. Grinnell and Miller (1944) reported Black Terns nesting in 
the Central Valley along the Sacramento and San Joaquin rivers (latter near 
Merced) and at Los Banos, Merced County, Laton and Firebaugh, Fresno 
County, and Buena Vista Lake, Kern County. They also noted the species 
had colonized rice fields. Egg-set data provide a minimum of two nesting 
records for the Sacramento Valley, 39 for the delta (none reported by 
Grinnell and Miller), and 399 for the San Joaquin Valley (Appendix 1). 
Although terns were widely scattered in the latter valley, many egg sets were 
collected from the Los Banos area of the San Joaquin Basin, perhaps 
reflecting local abundance, ease of observer access, or collector bias. In the 
early 20th century the San Joaquin Valley was of great importance to 
breeding Black Terns; Ray (1906), Chapman (1908), Tyler (1913), and van 
Rossem (1933) described the species as very numerous there. Among the 
few early quantitative estimates were over 100 nests near Los Banos and 
South Dos Palos from 19 to 22 May 1919 (J. G. Tyler et al, Appendix 2), 
“many hundreds in sight in all directions” in overflow lands of the San 
Joaquin River near Los Banos on 29 May 1941 (W. B. Minturn field notes), 
and a colony of “about 200 pairs” at Buena Vista Lake on 21 June 1921 (A. 
J. van Rossem egg data slip, WFVZ 2470). Sacramento Valley and delta 
records being so few may in part reflect limited egg collecting there. 

From the cessation of most egg collecting in the mid-1940s until our 
1997-1998 surveys, the limited information on the Black Tern’s status in 
the Centra] Valley was primarily anecdotal. Although Grinnell and Miller 
(1944) noted the at least partial shift in breeding from reclaimed marshes to 
cultivated rice, it is unclear how widespread or numerous terns were in rice 
fields, which in 1943 totaled 96,000 ha in California (National Agric. 
Statistics Serv.; http://www.nass.usda.gov: 100/ipedb/). The estimate of 
1000+ migrant Black Terns at the Woodland Sugar Ponds, Yolo County, 8- 
21 May 1955, though notable at the time (AFN 9:355), far exceeds any 
such recent estimate for the Sacramento Valley. Black Terns bred at 
Sacramento NWR in 1958 (Appendix 2), when rice was regularly grown on 
the refuge, but no terns are known to have bred on any federal refuge in the 
Sacramento Valley since at least the early 1980s (J. Silveira pers. comm.). 
Greenberg (1972) reported up to 48 Black Terns in mid- June on two 6-mile 
(9.6-km) road transects in Sutter and Sacramento counties from 1969 to 
1971, but data are too few for trend analysis. Cogswell (1977) concluded 
that Black Tern numbers declined in the Central Valley with loss of marshes, 


206 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


increased with expansion of rice growing, and declined again “recently,” 
perhaps from pesticide accumulation. The anecdotal nature of his and 
others’ claims of declines (AFN 24:638, AB 32:1205, AB 39:98) or 
upswings (AB 31:1185) in tern numbers in the Sacramento Valley in the 
1970s and 1980s make them hard to evaluate. Lee (1984) found six 
colonies (one in Sacramento Co., five in Sutter Co.) while studying Black 
Tern nesting biology in rice fields in 1976 and 1977 but made no population 
estimates for the area. 

Numbers of Black Terns recorded during surveys of pheasant broods in 
Butte County in late June and early July, 1976-1992 (J. Snowden in litt.), 
did not show any significant temporal trend but appeared to track the 
county’s rice acreage (Figure 5). Similarly, the only Breeding Bird Survey 
route in California (no. 148) with moderate numbers of Black Terns (median 
9, range 0-54), in Glenn and Colusa counties in the Sacramento Valley, 
showed substantial variability in numbers and no clear trend from 1971 to 
1999 (USGS Patuxent Wildlife Research Center 2000; http://www.mp2- 
pwrc . usgs . gov/bbs/ retrieval/) . 

W. B. Minturn (field notes) was still observing up to 500+ at various sites in 
the San Joaquin Valley in May through at least 1950. In the Tulare Basin, 
Black Terns probably bred regularly at Tulare and Buena Vista lakes until the 
1940s and 1950s (Appendix 1), when dams constructed on the rivers feeding 
them cut off most runoff. Since then breeding in that basin has been irregular, 



YEAR 

Figure 5. Numbers of Black Terns on California Department of Fish and Game survey 
routes for pheasant broods in Butte County, 1976 to 1992 (see Methods), relative to 
yearly totals of rice acreage for that county (National Agricultural Statistics Service 
1999; http://www. nass.usda.gov: 100/ipedb/). 


207 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 

mainly in extremely wet years. On 22 July 1983, following an El Nino winter, 
R. Hansen (MPCR files) found “many” nests and up to 70 Black Terns, mostly 
fledged young, at the Hacienda Ranch Flood Basin and South Wilbur Flood 
Area, Kings County. Also in the early 1980s, Black Terns nested in thick 
ditch-grass ( Ruppia maritima ) at J & W Farms agricultural evaporation 
ponds, 18 km southwest of Corcoran, Kings County (G. Gerstenberg pers. 
comm.). These ponds are no longer active, though, and other evaporation 
ponds in the basin currently do not support much emergent growth or any 
breeding Black Terns (R. Hansen, J. Seay pers. comm.). 

Similar local extirpations have also occurred in the San Joaquin Basin. In 
the early to mid-1950s, Black Terns nested where a slough flooded areas of 
spiny saltbush ( Atriplex spinifera) interspersed with vernal-pool-like habitat 
at Volta Wildlife Area (WA) near Los Banos (R. Wilbur pers. comm.). Black 
Terns appear to have been still widespread around Los Banos at that time, 
as R. D. Ross (MPCR files) estimated 200-250 terns on a drive across 
Highway 152 on 15 June 1956. Subsequently, P. J. Metropulos (in litt.) 
found the species “common” in the Los Banos area on 20 June 1970, V. 
Remsen and P. Myers (AB 27:914) counted 43 south of Los Banos on 16 
June 1973, and R. J. Bacon (MPCR files) sighted two or three nesting pairs 
at Merced NWR in summer 1983. Black Terns apparently were nesting in 
the Traction Ranch section of Mendota WA, Fresno County, when ponds 
were first developed there about 1.990 (S. Bruggemann pers. comm.). Today 
terns do not nest regularly at any of the state or federal refuges in the San 
Joaquin Valley (J. Allen, S. Bruggemann, R. Wilbur, D. Woolington pers. 
comm.). They likely breed regularly, though, in the limited rice acreage near 
Merced and South Dos Palos, though this remains to be documented. The 
current tenuous status of breeding Black Terns in the San Joaquin Valley 
documents a major population decline there over the last 100 years and an 
apparent shift of abundance to the Sacramento Valley. It is possible, though, 
that the Sacramento Valley has always been an important, though poorly 
documented, breeding area. 

Extralimital breeding. Apparent nesting at Merritt Lake, near Castroville, 
Monterey County (Silliman 1915), likely represents an extralimital attempt, 
as the species has not bred elsewhere along the coast of California. 

Historical Habitat Shifts 

In the past, Black Terns nested in the Central Valley in ephemeral early 
successional wetlands created by natural overflow of rivers and lakes 
(Mailliard 1904, Tyler 1913, van Rossem 1933) or by flood irrigation of 
pasturelands (Chapman 1908). It is hard to estimate the extent of tern 
breeding habitat, particularly ephemeral overflow lands, available prior to 
the massive alteration of the Central Valley’s natural hydrology. Hall (1880) 
estimated 324,000 ha of the Sacramento Valley were subject to inundation 
from annual overflow and an additional 117,000 ha by “occasional tempo- 
rary overflow.” In the San Joaquin Valley, he estimated 253,000 ha of 
swamp land were subject to periodic inundation. In the Tulare Basin alone 
the fluctuating margins of Tulare Lake could engulf many thousands of 
additional hectares after a series of wet winters. 


208 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


Although it is unclear how much ephemeral habitat remained through the 
terns’ breeding season, the vast flood plains and natural flood basins delayed 
transmission of flood flows, reduced peak flows and velocities, and increased 
summer river flows, as the flood waters slowly drained back into the rivers, 
sometimes through July, or evaporated (Bay Institute 1998). The buffering 
effect of the flood basins shifted high upstream flows of January to May to a 
period of high river outflow from March to June. Rainfall-induced floods 
(December-March) predominated in the Sacramento Valley, whereas pro- 
longed snowmelt floods (April-June) were the norm in the San Joaquin 
Valley, particularly in the Tulare Basin (Bay Institute 1998). Hence, the latter 
region likely had the most ephemeral habitat for breeding terns. 

Today’s water-management infrastructure keeps rivers within their banks, 
except during extreme floods, after which water usually rapidly drains or is 
pumped back into river and bypass channels, leaving few areas of shallow 
water where Black Terns could breed. The exception is the closed Tulare 
Basin where in extreme winters flood waters are diverted into shallow 
storage basins or run unchecked into fields. Flood frequency has decreased 
such that floods in the Sacramento Valley that formerly occurred about every 
2 years now occur once every 7 to 13 years, 10-year floods every 100 years 
(Bay Institute 1998). Valleywide, the volumes of large floods remain largely 
unchanged, but only in years of a very heavy snowpack in the Sierra Nevada 
do flood flows in the San Joaquin Valley approach historic levels. 

The great loss of wetlands was mitigated in the Sacramento Valley by the 
expansion of rice to the current annual level of 160,000 to 200,000 ha 
(Figure 6), which may far exceed the average extent of shallow-water habitat 
available there previously in summer. By contrast, wetlands lost in the San 
Joaquin Valley have been replaced to only a tiny degree by rice, which has 
declined there since the mid-1950s (Figure 6). Terns formerly bred in rice 
fields as far south as Kern County (Appendix 1) but no longer do so. 

Migratory Stopovers 

From 1949 to 1977, estimated peak counts of Black Terns at Tule Lake 
NWR, Siskiyou and Modoc counties, from July to early September ranged 
from 2000 to 19,000 (n = 17 years, median 5000, Klamath Basin NWR 
files), documenting it as an important postbreeding or migratory stopover 
for the species. Estimates of tern numbers at Tule Lake, 15 July-4 August 
1997, ranged from 1000 to 6000 (J. Beckstrand, R. Ryno, R. Ekstrom in 
Shuford 1998). In five years from 1958 to 1972, peak counts at Lower 
Klamath NWR in August exceeded 1000 (maximum 9000, Klamath Basin 
NWR files); large numbers have not been reported there in recent years. The 
only other key stopover site known in the state is the Salton Sea, Riverside 
and Imperial counties, outside the breeding range. Up to 15,000 have been 
estimated there in early August (Patten et al. in press), but the only census, 
13-16 August 1999, tallied 4011 individuals (Shuford et al. in press). Small 
(1994) implied that numbers at the Salton Sea have declined since 1987, but 
M. A. Patten (in litt.) has no evidence of this. Numbers of migrants on the 
southern California coast have declined greatly since the early 20th century 
(Garrett and Dunn 1981). 


209 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 



Figure 6. Historic changes in acreage of planted rice in California, 1912 to 1998, and 
in the Sacramento Valley and Delta and San Joaquin Valley, 1953 to 1998. Data from 
National Agricultural Statistics Service (1999; http://www.nass.usda.gov: 100/ 
ipedb/); county and district breakdowns available only since 1953. 


Current Threats 

Central Valley. Agricultural practices that rapidly draw down water levels 
in rice fields have exposed tern nests to rat predation only to destroy 
renesting attempts later when fields were reflooded above original levels (Lee 
1984). In 1998, Shuford saw terns sitting on nests in muddy fields from 
which the water had been temporarily drained, and later some of these nests 
were abandoned. The rapid increase in rice cultivation coincided with the 
post- World War II boom in chemical use in agriculture. Three egg yolks 
collected from a Black Tern colony in rice fields in the Sacramento Valley in 
1969 had 8.0, 9.1, and 11.8 ppm DDE (Greenberg 1972), but there is no 
evidence of deleterious effects of pesticides or other agricultural chemicals 
on terns breeding in rice fields. Dunn and Agro (1995) and Weseloh et al. 
(1997) reviewed the literature on contaminants in Black Tern eggs but found 
no evidence of reproductive effects. They concluded that direct chemical 
toxicity is generally not a problem with these terns, but pesticides may 
reduce favored insect foods. Loss of insect diversity or biomass could lead to 
chick starvation. 


210 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


Conservation Priorities 

Black Tern conservation should focus on protecting and enhancing high- 
quality nesting habitat. Because of the variable and generally arid climate in 
much of the California breeding range, the value of management efforts in 
years of short versus abundant water supply should be assessed carefully. 
Water levels should be regulated to protect important nesting and foraging 
habitats from ground predators and human disturbance, and management 
for terns should be coordinated with activities to conserve other species 
groups via other multi-partner conservation efforts (e.g., Central Valley 
Habitat Joint Venture, U.S. Shorebird Conservation Plan). Protection of key 
migratory stopovers, such as Tule Lake and the Salton Sea, is also vital. 

Northeastern California. Because few Black Terns currently nest on 
refuges, it would be valuable to see if spikerush-dominated marshes, the 
species’ main breeding habitat in the region, could be established or 
expanded in these areas. 

Central Valley. Because summer water typically is scarce in the Central 
Valley, and because seabirds’ productivity often runs in cycles of boom and 
bust, efforts to enhance tern habitat may be most fruitful in years of 
exceptional runoff. Recent evidence of extensive, irregular nesting in flooded 
fields with residual vegetation or crop stubble in the Tulare Basin indicates 
the species would benefit from more of this habitat. Perhaps fields with 
marginal crop yields could be retired from production and put in a conserva- 
tion bank to be flooded when excess water is available. Such flooding should 
be weighed against possible waterbird mortality from botulism outbreaks, 
which might be reduced by rotating the fields to be flooded and choosing 
areas with no prior evidence of disease. Scant breeding on newly restored 
wetlands on refuges near Los Banos perhaps could be increased in the 
future, particularly in years of high runoff. In such years, infrastructure 
improvements likely could spread water over larger areas within or adjacent 
to bypasses, such as the Eastside Bypass near Los Banos and the James 
Bypass/Fresno Slough south of Mendota WA. Study also is needed of 
whether suitable habitat on refuges adjoining bypasses could be increased 
during years of high flow by drawing upstream water, circulating it through 
refuge ponds, and draining it back into the bypass downstream. Maintaining 
a slow, steady flow likely would reduce the chances of botulism. 

Monitoring and Research Needs 

Shuford (1999) summarized monitoring and research needs for the Black 
Tern across North America. Efforts should be made to coordinate California 
surveys and research with those in other states and provinces to establish a 
broad perspective on population trends and ecology. Research is particularly 
needed on the foraging and nesting ecology of Black Terns in California. 
Studies of habitat suitability at both the local and landscape level are needed 
to guide wetland protection and acquisition efforts (Naugle et al. 2000). 
Banding studies are needed to reveal how terns shift with changing water 
conditions, as are demographic studies to identify which breeding habitats 
are sources or sinks for the overall population. Researchers should focus on 
variation in aspects of reproduction most likely to influence population 


211 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


trends and on understanding factors limiting nest success (Servello 2000), 
using methods that minimize disturbance to nesting terns (Shealer and 
Haverland 2000). 

We recommend a statewide survey approximately every 10 years, during 
typical climatic and habitat conditions, to document potential range shifts 
and calibrate long-term monitoring data. We recommend annual monitoring 
by trained observers using sampling protocols incorporating precision 
estimates; methods should be suitable for local conditions and responsive to 
the shifting of breeding locations. 

Northeastern California. Monitoring should be conducted in mid-June 
by counting undisturbed adults from points where observers do not attract 
mobbing terns. Surveys should be based on a random or stratified sampling 
of a subset of potential breeding sites, accounting for the difficulty of 
reaching some. 

Central Valley. Monitoring should be conducted via standardized road- 
side transects in rice fields in the Sacramento Valley. Concern about the 
potential effects of agricultural pesticides and cultivation practices on Black 
Terns (Lee 1984) begs for expansion of research on these topics. Studies are 
needed to assess whether the value of rice fields to Black Terns equals that 
of ephemeral overflow lands or natural marshes. 

ACKNOWLEDGMENTS 

Funding was supplied by the California Department of Fish and Game (CDFG), 
Chevron USA, Inc., the Richard Grand Foundation, Klamath Basin NWR Complex, 
the Weeden Foundation, U.S. Fish and Wildlife Service’s Nongame Migratory Bird 
Program, and individual donors to Point Reyes Bird Observatory. We are very grateful 
to the many individuals who provided valuable information, advice, access to private 
or restricted public lands, or help in the field, without which this study would not have 
been possible. We especially thank Lynne Stenzel for insight on sampling design for 
roadside surveys in the Sacramento Valley; A1 DeMartini, Tim Manolis, Jim Snowden, 
Brad Stovall, and Bruce Webb for help on roadside surveys; Grant Ballard and Diana 
Stralberg for preparing distribution maps; Jim Snowden for sharing long-term data on 
Black Terns from CDFG surveys of pheasant broods in Butte County; Martha 
Leighton of the California Agricultural Statistics Service for data on acreage by county 
of planted rice in the Central Valley; Jay Dee Garr, Cass Mutters, and Jack Williams 
for insight on timing of rice planting in 1998 and methods of rice cultivation; Dawn 
Harris for unpublished data on Black Terns from surveys of duck ponds in the 
Sacramento Valley; Rob Hansen and associates for arranging access and helping on 
surveys of the Hacienda Ranch Rood Basin and the South Wilbur Rood Area and for 
supplying unpublished field notes of Ward B. Minturn and John G. Tyler; Frances 
Bidstrup, Danny Cluck, Mike McVey, Bobby Tatman, and David Wimpfheimer for 
surveying remote wetlands in northeastern California; the Devil’s Garden Ranger 
District of Modoc National Forest, and particularly George Studinski, for logistical 
support and sharing information about wetlands there; various other state and federal 
land management agencies statewide for providing access to lands and personnel to 
help on surveys; Ken Blum of Tamal Saka, Tomales Bay Kayaking, for kayaking 
equipment; Helen Green for providing data from the editors of the Middle Pacific 
Coast region of North American Birds on file with Golden Gate Audubon Society; 
Karen Cebra, Carla Cicero, Rene Corado, Krista Fahy, Kimball Garrett, Ned Johnson, 
Freda Kinoshita, Douglas Long, Bob McKernan, James Northern, and Philip Unitt for 
Black Tern egg-set data, or confirmation of a lack thereof, from their respective 


212 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


institutions; and Paul Fonteyn, Kevin Hunting, Dave Mauser, Gary Page, Tom Smith, 
and Denise Wight for support and encouragement for the project. John Cooper, 
Gordon Gould, Kevin Hunting, Tim Manolis, Bruce Peterjohn, and Tara Zimmerman 
made helpful comments on drafts of the manuscript. This is contribution 718 of Point 
Reyes Bird Observatory. 

LITERATURE CITED 

Alvo, R., and Dunn, E. 1996. Updated report on the Black Tern Chlidonias niger in 
Canada. Unpubl. report, Committee on the Status of Endangered Wildlife in 
Canada, c/o Canadian Wildlife Serv., Environment Canada, Ottawa, Ontario, 
Canada K1A 0H3. 

Bailey, V. 1902. Unprotected breeding grounds. Condor 4:62-64. 

Bay Institute. 1998. From the Sierra to the sea: The ecological history of the San 
Francisco Bay-Delta watershed. The Bay Institute of San Francisco, 55 Shaver 
St., Suite 330, San Rafael, CA 94901. 

California Department of Fish and Game. 1992. Bird species of special concern. 
Calif. Dept. Fish & Game, 1416 Ninth St., Sacramento, CA 95814. 

Chapman, F. M. 1908. Camps and Cruises of an Ornithologist. Appleton, New York. 

Cogswell, H. L. 1977. Water Birds of California. Univ. Calif. Press, Berkeley. 

Dahl, T. E., Young, R. D., and Caldwell, M. C. 1997. Status and trends of wetlands 
in the conterminous United States: Projected trends 1985 to 1995. U.S. Dept. 
Interior, Fish & Wildlife Serv., Washington, D.C. 

Dawson, W. L. 1923. The Birds of California. South Moulton, San Diego. 

Dunn, E. H., and Agro, D. J. 1995. Black Tern ( Chlidonias niger), in the Birds of 
North America {A. Poole and F. Gill, eds.), no. 147. Acad. Nat. Sci., Philadel- 
phia. 

Frayer, W. E., Peters, D. D., and Pywell, H. R. 1989. Wetlands of the California 
Central Valley: Status and trends. Unpubl. report, U.S. Fish & Wildlife Serv. , 911 
NE 11th Ave., Portland, OR 97232-4181. 

Garrett, K., and Dunn, J. 1981. Birds of Southern California: Status and Distribution. 
Los Angeles Audubon Soc., Los Angeles. 

Gerson, H. 1988. Status report on the Black Tern Chlidonias niger. Unpubl. report, 
Committee on the Status of Endangered Wildlife in Canada, c/o Canadian 
Wildlife Serv., Environment Canada, Ottawa, Ontario, Canada K1A 0H3. 

Gould, G. I., Jr. 1974. Breeding success of piscivorous birds at Eagle Lake, California. 
Master’s thesis, Humboldt State Univ., Areata, CA. 

Greenberg, D. 1972. Black Tern survey Sutter County, California, 1969-71. 
Unpubl. report, Wildlife Mgmt. Branch Special Wildlife Invest. Calif. Dept. Fish 
& Game, 1416 Ninth St., Sacramento, CA 95814. 

Grinnell, J., Dixon, J., and Linsdale, J. M. 1930. Vertebrate natural history of a 
section of northern California through the Lassen Peak region. Univ. Calif. Publ. 
ZooL 35:1-594. 

Grinnell, J., and Miller, A. H. 1944. The distribution of the birds of California. Pac. 
Coast Avifauna 27. 

Hall, W. H. 1880. Drainage of the valleys and the improvement of the navigation of 
rivers. Report of the State Engineer to the Legislature of the State of California, 
Session 1880, in three parts. 


213 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


Kempka, R. G., Kollasch, R. R, and Olson, J. D. 1991. Aerial techniques measure 
shrinking waterfowl habitat. Geo Info Systems (Nov/Dec), pp. 48-52. 

Lederer, R. J. 1976. The breeding populations of piscivorous birds of Eagle Lake. 
Am. Birds 30:771-772. 

Lee, R. C., Jr. 1984, Nesting biology of the Black Tern {Chlidonias niger ) in rice fields 
of the Central Valley, California. Master’s thesis, Calif. State Univ., Sacramento. 

Mailliard, J. 1904. A few records supplementary to Grinnell’s check-list of California 
birds. Condor 6:14-16. 

Naugle, D. E., Higgins, K. F., Estey, M. E., Johnson, R. R., and Nusser, S. M. 2000. 
Local and landscape-level factors influencing Black Tern habitat suitability. J. 
Wildlife Mgmt. 64:253-260. 

Orr, R. T., and Moffitt, J. 1971. Birds of the Lake Tahoe Region. Calif. Acad. Sci., 
San Francisco. 

Patten, M. A., McCaskie, G., and Unitt, P. In press. Birds of the Salton Sea: Status, 
Biogeography, and Ecology. Univ. Calif. Press, Berkeley. 

Peterjohn, B, G,, and Sauer, J. R. 1997. Population trends of the Black Tern from the 
North American Breeding Bird Survey, 1966-1996. Colonial Waterbirds 
20:566-573. 

Ray, M. S. 1906. A-birding in an auto. Auk 23:400-418. 

Ray, M. S. 1913. Some further notes from the Tahoe region. Condor 15:111-115. 

Sauer, J, R., Hines, J. E., Thomas, I., Fallon, J., and Gough, G. 2000. The North 
American Breeding Bird Survey, results and analysis 1966-1999. Version 98.1, 
USGS Patuxent Wildlife Res. Ctr., Laurel MD (http://www.mbr-pwrc.usgs.gov/ 
bbs/bbs.html). 

Servello, F. A. 2000. Population research priorities for Black Terns developed from 
modeling analyses. Waterbirds 23:440-448. 

Shaw, D. W. H. 1998. Changes in population size and colony location of breeding 
waterbirds at Eagle Lake, California, between 1970 and 1997. M.S. thesis, Calif. 
State Univ., Chico. 

Shealer, D. A., and Haverland, J. A. 2000. Effects of investigator disturbance on the 
reproductive behavior and success of Black Terns. Waterbirds 23:15-23. 

Shuford, W. D. 1998. Surveys of Black Terns and other inland-breeding seabirds in 
northeastern California in 1997. Report 98-03, Bird and Mammal Conservation 
Program, Calif. Dept. Fish & Game, 1416 Ninth St., Sacramento, CA 95814. 

Shuford, W. D. 1999. Status assessment and conservation plan for the Black Tern 
( Chlidonias niger surinamensis) in North America. U.S. Dept. Interior, Fish & 
Wildlife Serv., Denver Federal Center, Denver, CO 80225-0486. 

Shuford, W. D., Humphrey, J. M., and Nur, N. 1999. Surveys of terns and 
cormorants in California’s Central Valley in 1998. Point Reyes Bird Observatory 
(Contr. 772), 4990 Shoreline Hwy., Stinson Beach, CA 94970. 

Shuford, W. D., Warnock, N., Molina, K. C., and Sturm, K. K. In press. The Salton 
Sea as critical habitat to migrating and resident waterbirds. Hydrobiologia (Dev. 
HydrobioL). 

Silliman, O. P 1915. Another Mexican Ground Dove for California, and other notes. 
Condor 17:207. 

Small, A. 1994. California Birds: Their Status and Distribution. Ibis Publ., Vista, CA. 

Tyler, J. G. 1913. Some birds of the Fresno district, California. Pac. Coast Avifauna 9. 


214 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


U.S. Fish and Wildlife Service. 1995. Migratory nongame birds of management 
concern in the United States: The 1995 list. Office of Migratory Bird Manage- 
ment, U.S. Fish & Wildlife Serv., Washington, D C. 

Van Rossem, A. J. 1933. Terns as destroyers of birds’ eggs. Condor 35:49-51. 

Weseloh, D. V. C., Rodrigue, J., Blokpoel, H., and Ewins, P. J. 1997. Contaminant 
concentrations in eggs of Black Terns ( Chlidonias niger) from southern Ontario 
and southern Quebec, 1989-1996. Colonial Waterbirds 20:604-616. 

Zeiner, D. C., Laudenslayer, W. F., Jr., Mayer, K. E., and White, M. 1990. California’s 
Wildlife, vol. II, Birds. Calif. Statewide Wildlife Habitat Relationship System, 
Calif. Dept. Fish & Game, 1416 Ninth St., Sacramento, CA 95814. 

Accepted 14 September 2001 


Appendix 1. Numbers of egg sets of the Black Tern in California, 1886- 
1960, from major California museums. 

Northeastern California 

Modoc County: Alturas Meadow, 9 June 1918 (2); 3.7 mi W of Alturas, 9 June 
1918 (4). 

Lassen County: Grasshopper Meadows/Lake, 2-22 June 1918 (20); Spaulding’s, 
Eagle Lake, 3 June 1918 (7); Eagle Lake, 3-6 June 1918 (5), 22 June 1928 (1); nr. 
Truxell’s, east shore of Eagle Lake, 23 May 1923 (1); Upper Ragar Meadow, 1 June 
1935 (1). 

El Dorado County: nr. Bijou, Lake Tahoe, 19 June 1899 (1), 9 June 1911 (5), 10 
June 1912(2), 6 June 1918(9); Lake Tahoe, 6 June 1910(1); Rowland’s Marsh (i.e., 
Al-Tahoe), Lake Tahoe, 22 June 1902 (1), 10 June 1909 (3), 23 May-15 June 1910 
(7), 30 May-9 June 1914 (18), 30 June 1918 (1), 5 June 1919 (1), 30 May 1920 (1), 
14 June 1928 (1), 21 June 1930 (1), 15 June 1939 (2); nr. Tallac, Lake Tahoe, 22 
June 1911 (1). 

Central Valley: Sacramento Valley 

Colusa County: Maxwell, 23 June 1939 (1). 

Yolo County: Woodland, 11 May 1886 (1). 

Central Valley: Sacramento-San Joaquin River Delta 

Sacramento County: 0.5 mi S of Freeport, 15 June 1899 (2); Bear Lake, 27 May 
1923 (5); vie. Sacramento (Pcounty, ?delta), 7 June 1902 (1), 13 May 1906 (2); Stone 
Lake, 15-29 May 1921 (23), 4 June 1922 (1), 13-30 May 1923 (4). 

San Joaquin County: White Ranch, 9 mi N of Stockton, 3 June 1921 (1); 
Kettleman Swamp, 9.5 mi NW of Stockton, 1 June 1947 (3). 

Central Valley: San Joaquin Valley 

Madera County: Chowchilla (egg record says “Merced Co.”), 23 June 1900 (5); 15 
mi W of Madera, 30 June 1923 (1); “Madera Co.,’ nr. Firebaugh, Fresno Co., 16-17 
May 1927 (8); “Madera Co.,” 10 mi E of Firebaugh, Fresno Co., 26 May 1927 (4); 
10 mi from Firebaugh (?county), 5 June 1927 (1); “Madera Co.,” 28 May 1928 (1), 
9 June 1930 (3). 

Merced County: nr. Brito, 21 May 1919 (1); Dos Palos, 17-22 May 1912 (5), 8 
June 1927 (1); Gadwall, 16 May 1914 (1), 1-2 July 1917 (6), 12 May-4 June 1918 
(14); Gustine, 14 May 1931 (1), 5 June 1932 (1), 12 June 1934 (1), 7 June 1937 (5); 


215 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


Los Banos, 17 May 1898 (1), 8 June 1901 (1), 5 June 1905 (3), 2 June 1908 (2), 26 
May 1910 (3), 2 July 1913 (1), 7 June 1914 (2), 29 May-25 June 1916 (12), 3-4 
June 1918 (4), 26 June 1919 (1). 30 May 1920 (1), 28 May 1921 (1), 3-4 June 1922 
(3), 1-18 June 1923 (15), 21 May-21 June 1925 (9), 23 May-20 June 1926 (8), 3 
June 1928 (1), 30 Ma^l July 1930 (23), 14 June 1931 (1), 13 May-5 June 1932 
(10), 12 May-12 June 1934 (4), 11-13 May 1935 (5), 25 May 1936 (4), 14 May-1 
June 1937 (1 1), 4 June 1938 (1), 10 May-7 June 1939 (8), 19 May 1940 (4), 9 June 
1941 (1), 9 June 1942 (2); 4-6 mi S of Los Banos, 21-23 May 1919 (2), 30 May 
1920 (1), 12-21 June 1931 (5); 8-10 mi E of Los Banos, 7 May 1927 (1), 13 May- 
1 June 1935 (5); .5-10 mi NE of Los Banos, 9 May 1936 (1), 14 May 1937 (3), 31 
May 1939 (1); “Merced Co.,” 20 May 1899 (1), 25 May-1 July 1908 (16), 26 May 
1909 (9), 15-28 May 1926 (2), 28 May 1928 (1), 9 June 1930 (2), 25 June 1931 
(1), 12-19 May 1935 (4), 3 May-9 June 1936 (20), 10 May 1939 (1), 5 June 1946 
(1), 22-23 June 1948 (4), 22-23 May 1949 (6); San Joaquin R. at Los Banos 
Crossing, 15 May 1897 (2). 

Fresno County: Columbia Ranch, 24-25 June 1919 (4), 8-9 June 1920 (8), 22- 
23 June 1921 (5); Firebaugh, 28 May 1916 (2); “Fresno Co.,” nr. South Dos Palos, 
Merced Co., 20 May 1919 (3); nr. Laton, 21 June 1919 (2), 3 June 1922 (1); 
McNeil’s Ranch SW of Fresno, 7 June 1920 (1); Mendota, 26 May 1915 (1), 7-21 
June 1930 (5); Riverdale, 25 May 1917 (1), 24 May 1919 (1). 

Kings County: 12 mi from Corcoran (egg record says “Kern Co.”), 24 May 1940 
(1); Gernsey Slough, 3 June 1946 (2); 3 mi E of Hanford, 24 May 1922 (1); nr. 
Stratford, 23-24 May 1936 (10); border of Tulare Lake, 4 mi W of Waukena, Tulare 
Co., 6 June 1893 (3); Tulare Lake, 8 June 1941 (11); Tulare Lake (Kings Co.), 24 
May-8 June 1941 (18); 14 mi NW of Tulare, 6 June 1893 (1). 

Kern County: Buena Vista Lake, 10 June 1907 (1), 19-20 June 1914(2), 18 June 
1916 (1), 19-21 June 1921 (3), 11 June-5 July 1922 (8), 4 July 1937 (1), 5 June 
1938 (4), 6 June 1948 (1), 20 June 1954 (2), 24 June 1956 (1); rice field between 
Wheeler Ridge and Buena Vista Lake, 12 June 1960 (2); Kern River, 5 June 1938 (2). 

Appendix 2 . Sight records of confirmed breeding of Black Terns in Califor- 
nia, 1899 to 1999, other than those listed or summarized in the text. 

Northeastern California 

Siskiyou County: Lower Klamath NWR, Unit 13A, mid-July 1995 (22 nests, 
Klamath Basin NWRs files). 

Siskiyou/Modoc County: Tule Lake, early July 1899 (“nests,” Bailey 1902). 

Modoc County: Beeler Reservoir, 19 June 1976 (nest, B. E. Deuel). 

Lassen County: Eagle Lake, 22 June 1921 (15-20 nests, Grinnell et al. 1930), 
summer 1974 (23 nests, Lederer 1976); Delta Bay, Eagle Lake, late May-mid July 
1971 (1 nest, Gould 1974); North Basin, Eagle Lake, late May-late June 1996 (29 
nests, Shaw 1998), early June-mid July 1997 (21 nests, Shaw 1998); southwest 
shore, Eagle Lake, late May-mid July 1971 (6 nests, Gould 1974); nr. Spaulding’s, 
Eagle Lake, 9 June 1925 (“many nests,” Grinnell et al. 1930), late May-mid July 
1970 (30 nests, Gould 1974), late May-mid July 1971 (33 nests, Gould 1974), early 
June-mid July 1997 (11 nests, Shaw 1998); eastside bays (Troxel and Duck Island 
bays), Eagle Lake, late May-mid July 1971 (1 1 nests, Gould 1974), late May-late June 
1996 (10 nests, Shaw 1998). 

Plumas County: Sierra Valley, 23 July 1973 (nest, G. Zamzow), 14 June 1989 
(nest, D. Shuford et al.), 13 June 1998 (nest, D. Shuford, J. McCormick). 

Sierra County: Kyburz Flat, 28 June 1973 (nest, G. Zamzow), 19 July 1973 (nest, 
G. Zamzow). 

El Dorado County: Emerald Bay, Lake Tahoe, 10 August 1918 (“parents feeding 


216 


BREEDING STATUS OF THE BLACK TERN IN CALIFORNIA 


young,” J. W. Mailliard in Orr and Moffitt 1971); Rowland’s Marsh (i.e., Al-Tahoe), 
Lake Tahoe, 1 June 1909 (“scores of nests,” Ray 1913). 

Central Valley: Sacramento Valley 

Butte County: W of Biggs, 6 July 1987 (2 nests, J. Snowden in litt.); 3 mi S of 
Durham, 1 June 1985 (2 nests, J. Hornstein); 2 mi NE of Richvale, 3 July 1984 (7 
nests, J. Snowden in litt.). 

Glenn County: Sacramento NWR, 9 June 1958 (2 chicks banded; refuge files). 

Colusa County: S side of White Rd. 0.7 mi W of Browning Rd., 26 June 1999 (2+ 
nests, B. Williams in litt.). 

Sutter County: S of Kirkville Rd. adjacent to Sutter Bypass, June-July 1976 (3 
nests, Lee 1984); jet. Hwy. 113 and Varney Rd., June-July 1976 (10 nests, Lee 
1984); E of Armour Rd. between Kirkville and Varney roads, May-July 1977 (2 nests, 
Lee 1984); jet. Hwy. 113 and Kirkville Rd., June-July 1977 (8 nests, Lee 1984); N 
of Kirkville Rd. adjacent to Sutter Bypass, June-July 1977 (11 nests, Lee 1984); N of 
Robbins, June 1969 (“colony of 12 terns [with] nests,” Greenberg 1972). 

Sacramento County: jet. Hwy. 99 and Elkhorn Blvd., 24 May-22 June 1976 (13 
nests, Lee 1984). 

Central Valley: San Joaquin Valley 

Merced County: nr. Los Banos, 16 June 1903 (young of year just beginning to fly, 
Chapman 1908), prior to 1923 (photo of chicks, Dawson 1923); San Joaquin River 
nr. Merced, prior to 1904 (“number of nests recorded,” Mailliard 1904). 

Merced/Fresno County: vie. Los Banos and South Dos Palos, 19-22 May 1919 
(>100 nests examined, J. G. Tyler et al.). 

Fresno County: nr. Laton, 31 May 1910 (“set of 3 eggs,” C. Lamb in Tyler 1913), 
27 May 1917 (colony of about 30 pairs, 8 egg sets, N. K, Carpenter, A. M, Ingersoll 
fide J. G. Tyler); Firebaugh, 30 May 1912 (several birds “sitting on nests,” Tyler 
1913); Riverdale, 25 May 1917 (colony of 20-25 pairs, 13 egg sets, J. G, Tyler, N. 
Carpenter); pond S of Fowler, 30 May 1918 (nest, J. G. Tyler); Mendota, 30 May 
1928 (3 nests, W. B. Minturn, J. G. Tyler); White’s Bridge Rd., Mendota, 17 May 
1930 (nest, W. B. Minturn), 7 June 1930 (about 20 nests, W. B. Minturn, J. G. Tyler), 
1 May 1937 (7 nests being built, W. B. Minturn), 22 May 1937 (8 nests, W. B. 
Minturn, C. Chandler), 14 May 1941 (partly completed nest, W. B. Minturn), 7 June 
1941 (nest, W. B. Minturn). 

Fresno/Madera County: Mendota Dam (i.e., Mendota Pool), 3 June 1933 (8 nests, 
W. B. Minturn, J. G. Tyler), 23-24 June 1933 (7 nests, W. B. Minturn, J. G. Tyler). 

Madera County: 12 mi W of Madera, 9 June 1934 (2 nests, J. G. Tyler). 

Kings County: Hacienda Ranch Flood Basin and South Wilbur Rood Area, 22 July 
1983 (“many nests”" one photographed, R. Hansen); East Hacienda Ranch Flood 
Basin, 29 June 1997 (5 nests, R. Hansen in litt.) 


217 


IDAHO BLACK SWIFTS: NESTING HABITAT AND 
A SPATIAL ANALYSIS OF RECORDS 

R. KASTEN DUMROESE, USDA Forest Service, SRS, 1221 South Main Street, 
Moscow, Idaho 83S43 

MARK R. MOUSSEAUX, IJSDI Bureau of Land Management, Medford District 
Office, 3040 Biddle Road, Medford, Oregon 97504 

SHIRLEY HORNING STURTS, East 4615 Fernan Lake Road, Coeur d’Alene, Idaho 
83814 

DANIEL A. STEPHENS, Department of Biology, Wenatchee Valley College, 1300 
Fifth Street, Wenatchee, Washington, 98801 

PAUL A. HOLICK, 1037 Showalter Road, Moscow, Idaho, 83843 

ABSTRACT: The Black Swift (Cypseloides niger) was first confirmed breeding in 
Idaho in 1997 and 1998 when four and five pairs, respectively, nested near Shadow 
and Fern falls along the North Fork Coeur d’Alene River, Shoshone County. Nest sites 
were on cliffs composed of argillite within the large Precambrian Belt Supergroup 
geologic formation and associated with a narrow riparian strip of western redcedar 
and devil’s club. The microcommunity along cliff faces consisted of a variety of 
mosses, liverworts, and ferns. We analyzed all Black Swift sight records for Idaho, 
finding that 7 8% were from the breeding season and most breeding-season records 
(96%) were associated with the Precambrian Belt Supergroup. 

The Black Swift ( Cypseloides niger) breeds locally in the mountains and 
along the coasts of western North America from southeast Alaska south to 
Mexico and east to central Colorado (AOU 1983). Nests are associated with 
coastal cliffs (Vrooman 1901), waterfalls (Smith 1928), and caves (Davis 
1964). Although the Black Swift breeds in adjacent areas in the northern 
Rocky Mountains (Kondla 1973, Holroyd and Holroyd 1987, Hunter and 
Baldwin 1972), details of its status in Idaho are lacking. Larrison et al. (1967) 
considered the Black Swift a rare migrant and possible breeder, while 
Burleigh (1972) failed to mention the species. Since 1985, however, reports 
of summering Black Swifts have increased, and discovery of four nests in 
1997 (Svingen and Trochlell 1998) along the North Fork Coeur d’Alene 
River, Shoshone Co., provided the first confirmation of breeding in Idaho 
(Stephens and Sturts 1998). 

Here we describe the known nesting habitat of Idaho Black Swifts, explore 
the relationship of a prominent geologic formation to sight records, and 
suggest other potential suitable areas in Idaho for breeding swifts. 


STUDY AREA AND METHODS 

Black Swifts nested at two waterfalls, Shadow and Fern, on Falls Branch 
of Yellow Dog Creek in the North Fork Coeur d’Alene River watershed, 
Idaho Panhandle National Forest (47° 45' 13" N, 116° 06' 19" W). The 
region, characterized by steeply dissected mountains between 600 and 
2570 m elevation, lies over the Precambrian Belt Supergroup, a large 
formation of sedimentary rock (Hobbs et al. 1965). Annual precipitation 
ranges from 648 to 1415 mm, with 70% falling as snow (R. Kasun pers. 


218 


Western Birds 32:218-227, 2001 


IDAHO BLACK SWIFTS: NESTING HABITAT; SPATIAL ANALYSIS OF RECORDS 


comm.). The temperature at nearby Kellogg, Idaho, averages about 8° C, 
and the climate receives maritime influence. Winters are relatively mild and 
summers are dry (see ecoregion M333Ba, Nesser et al. 1997). 

We visited the nesting site on 8, 23, and 24 August 1998 and character- 
ized the waterfalls and associated macro- and micro-habitats. We compiled 
all known Idaho Black Swift sight records from 1929 through 2001 from 
several sources: Stephens and Sturts (1998); Audubon Field Notes , Ameri- 
can Birds , and their successors; U.S. Fish and Wildlife Service (USFWS) 
Breeding Bird Surveys; unpublished records of M. T. Jollie (Univ. of Idaho); 
the database of Thomas Rogers, long-time regional editor for Audubon 
Field Notes and American Birds ; and individual birders. Using Black Swift 
breeding phenology (Hunter and Baldwin 1962; Marin 1997, 1999) and 
our observations, we classified as summer (breeding) sight records from 7 
June through 31 August; all others were migration records. Finally, we 
mapped records in the context of a USFWS latilong grid over the distribution 
of the geology found at the nesting site. 

RESULTS 
Nest-Site. Habitat 

Shadow and Fern falls face east and lie in the bottom of a steeply sloped 
(45°) valley. Elevation of the precipices of Shadow and Fern falls is 998 and 
980 m, respectively, with Shadow Falls about 100 m upstream of Fern Falls. 
Although the creek flows through a narrow riparian macro-community 
dominated by western redcedar ( Thuja plicata ) and devil’s club ( Oplopanax 
horridum ) (Cooper et al. 1991), this community is in a mostly later serai 
state dominated by western hemlock ( Tsuga heterophplla) and lady fern 
( Athprium felix-femina). It contains the following key plants: twisted stalk 
( Stetopus stretopoides), northern wood fern ( Drpopteris expansa), 
Anderson’s swordfern ( Polpstichum andersonii ), Dewey’s sedge ( Carex 
dewepana), fool’s huckleberry ( Menziesia ferruginea), and smooth willow- 
weed ( Epilobium glaberrimum) . Upslope of the narrow, later successional 
riparian zone the plant community quickly grades to a moist western 
hemlock/oak fern ( Gpmnocarpium dr pop ter is) community and then to a 
drier and younger western hemlock/queencup beadlily ( Clintonia uni flora) 
community (Cooper et al. 1991) with few or no species associated with 
moist riparian areas. We observed a few old stumps, likely of the western 
white pine (Pinus monticola), indicating some historic logging and associ- 
ated burning of the residual logging debris that was scattered across the site 
(probably before 1940s), but there has been no recent logging nearby. 

Rock outcrops at both falls are green or bluish-gray thinly layered (0.2 to 
10 cm) argillite, with occasional thicker layers (10 to 15 cm) of impure 
quartzite. Argillite layers generally contain abundant mud cracks and erode 
more readily than do the quartzites. Thicker, more resistant quartzites 
generally form tops of precipices over which water descends, with less 
resistant, thinly layered argillite [which dips back (8° to 12°) into the face of 
the cliff] forming undercut faces of cliffs. Rocks on undercut faces are 
moderately fractured, slumping forward in places until they are almost 
horizontal. On the basis of correlations with previously mapped areas, rock 


219 


IDAHO BLACK SWIFTS: NESTING HABITAT; SPATIAL ANALYSIS OF RECORDS 


outcrops most likely represent the upper Wallace Formation within the 
Precambrian Belt Supergroup (Hobbs et al. 1965). 

Shadow Falls is about 8 m tall, and the curtain (free falling water no longer 
in contact with the cliff) is 5 m high, with the back of the falls cut at about a 
45° angle and approximately 2 m deep at the base. The main fall is 4 m 
wide, with a more fragmented curtain to 5 m, and the total width of dripping 
water is 8 m. Fern Falls is approximately 5 m tall; its curtain is 3 m high with 
a backcut similar to Shadow Falls’, and the drip zone is about 12 m wide. 

We found four nests, each with one nestling, at Shadow Falls in 1998 
(Figure 1). The lowest nest was near the southern edge of the curtain, behind 
falling water, and about 3.5 m above the bottom of the falls. The second nest 
was about 1 m directly above the first nest, behind the curtain, and about 0.3 
m below where the curtain began. A third nestling was 0.75 m due north of 
the second nest along a nearly horizontal ledge. The fourth nest was not 
behind falling water but about 1.5 m due south of the second nest and 
underneath an overhang 5 m deep that extended about 8 m to the 
southeast. The single nest and nestling at Fern Falls was only 2 m above the 
bottom of the falls near the northern edge of the main curtain (Figure 2). 

The dominant micro-community on the wet cliff faces included broom 
moss ( Dicranium scoparium), eurhynchium moss (Eurhynchium praelon- 
gum), stair step moss ( Hylocomium splendens), leucolepis moss ( Leucolepis 
menziesii ), platydictya moss ( Platydictya jungermannioides), big redstem 
moss ( Pleurozium schreiberii ), bentleaf moss ( Rhytidiadelpus squarrosus), 
snake liverwort ( Conocephalum conicum ), lung liverwort ( Marchantia 
polymorpha ), yellow-ladle liverwort ( Scapania bolanderi), lady fern, fragile 
fern ( Cystopteris fragilis), western polypody ( Polypodium hesperium), and 
oak fern. Nests appeared to be made exclusively from some or all of these 
bryophytes. 

Currently the site is indirectly protected under the USDA Forest Service’s 
conservation protocols that require maintaining the riparian community for 
50 m on either side of the creek. The falls are a popular tourist spot, with a 
well-maintained trail leading to Shadow Falls. Despite heavy visitation, most 
persons we talked with were unaware of nests. Black Swifts have high fidelity 
to nest sites (Collins and Foerster 1995, Marin 1997), and our observation 
that nearby human activity had no impact on breeding concurs with Foerster 
and Collins (1990), although Hirshman (1998) found that nestlings closer to 
human activity fledged later than those in less disturbed areas. 

Knorr (1961) listed five characteristics of Black Swift nest sites: (1) presence 
of moving water; (2) high relief so swifts leaving nests are automatically at 
potential foraging altitude above surrounding terrain; (3) inaccessibility to 
terrestrial predators; (4) sunlight not reaching occupied nests; and (5) unob- 
structed flyways immediately in front of nests. After additional work, Knorr 
(1994) concluded that high relief is not a requirement, although the character- 
istic is almost invariably present, and added a sixth criterion, the presence of 
niches in rocks for nest sites. 

Flowing water, inaccessible nests, and unobstructed flyways were present 
at both Shadow and Fern falls. Nests at Shadow Falls never received direct 
sunlight, but the nest at Fern Falls received about 20 minutes of direct 
sunlight around 08:30 the last days of August (J. Acton pers. comm.). Smith 


220 


IDAHO BlY\CK SWIFTS: NESTING HABITAT: SPATIAL ANALYSIS OF RECORDS 



Figure 1. Three Black Swift nestlings behind Shadow Falls. Shoshone Co.. Idaho. 

Photo by Dave Holick 


221 


IDAHO BLACK SWIFTS: NESTING HABITAT; SPATIAL ANALYSIS OF RECORDS 



Figure 2. A Black Swift nestling at Fern Falls, Shoshone Co., Idaho. 

Photo by Dave Holick 


(1928) reported sunlight to shine directly on a California nest site for an hour 
in early morning. Because of their location in the bottom of a narrow, steep 
valley, neither Shadow nor Fern falls offers immediate high-altitude foraging 
for adults. Like some Black Swifts in southern California (Foerster and 
Collins 1990), birds at Shadow Falls also flew through a maze of tree 
branches when exiting the nesting site (Figure 3; D. Johnson pers. comm.). 
The area immediately in front of the nests, however, was unobstructed as 
defined by Knorr (1961). Farther upstream of Shadow and Fern falls at 
1054 m elevation, we found three additional falls, heights 2, 3.5, and 6 m. 
that were characterized by the thicker, more resistant quartzites and lacked 
layering and nests. Marin (1997) concluded cypseloidine swifts breed by 
water to ameliorate daily temperature changes around the nest and that 
Knorr's (1961) other criteria were secondary consequences of nesting near 
waterfalls. Our observations imply that high relief and lack of direct sunlight 
striking an occupied nest are less important ecological characteristics for 
successful Black Swift breeding than the presence of moving water, protec- 
tion from terrestrial predators, available niches for nest construction, and 
unobstructed fly ways. 

Statewide Distribution 

As annotated in the appendix, Idaho’s 82 sight records of the Black Swift 
are from 34 general locations, shown in Figure 4 against the range of 
sedimentary rocks of the Precambrian Belt Supergroup. The 1929 sighting 
in latilong 5 probably formed the basis for the species being listed by 
Larrison et al. (1967:128) and Larrison (1981 : 171). Except for that sighting 


222 



IDAHO BLACK SWIFTS: NESTING HABITAT; SPATIAL ANALYSIS OF RECORDS 



Figure 3. View directly downstream from Shadow Falls. An adult Black Swift was 
observed flying through this forest when leaving the nest, 

Photo by Dave Holick 


and one in 1973 (latilong 1), all records have been since 1985. Most records 
(64, 78%) are from the breeding season. Most breeding-season records 
(96%) are from Idaho’s panhandle (latilongs 1 through 5) and associated with 
the Precambrian Belt Supergroup (Figure 4). 

The three observations from latilongs 7,8, 11, and 13 (15 July, 24 June, 
12 June, and 8 July, respectively) probably represent widely dispersed 


223 


IDAHO BLACK SWIFTS: NESTING HABITAT; SPATIAL ANALYSIS OF RECORDS 



Figure 4. Latilongs, Black Swift sighting locations, and distribution of Precambrian 
Belt Supergroup sedimentary rocks (shaded areas) in Idaho. If the species has been 
sighted in both summer and migration at the same location, only a summer sighting 
is indicated. Adapted from our appendix, Stephens and Sturts (1998), and Idaho 
Geologic Survey (1978). 


224 


IDAHO BLACK SWIFTS: NESTING HABITAT; SPATIAL ANALYSIS OF RECORDS 


foraging adults. Although the Belt Supergroup is a large formation, south of 
the Lochsa River (generally latilongs 7 +) the more highly metamorphosed 
Precambrian Belt rocks loose some of their layering as they change into 
schist, probably reducing the availability of nesting ledges like those at 
Shadow and Fern falls. Spring records from latilongs 17, 19, and 23 are too 
early for breeding, and fall records in latilong 17 are probably of migrating 
birds since adults and fledglings apparently begin migration immediately 
(Marin 1997). 

Our observations of nest-site habitat at Shadow and Fern falls and the 
prevalence of summer sight records in latilongs 1 to 5 suggest that any 
northern Idaho waterfall on sedimentary rock may meet the requirements of 
nesting Black Swifts and should be investigated. Additional field work should 
enhance our knowledge of the distribution of Black Swifts in Idaho. 

ACKNOWLEDGMENTS 

We thank Debbie and Niklaas Dumroese, Dave Holick, and Kenneth Quick for their 
assistance in describing the nesting site, Charles Collins, Manuel Marin, Owen Knorr, 
and Kathy Molina for their editorial suggestions, and the following observers who 
shared records: Jim Acton, Nancy Aley, Dan Audet, Vivian Bellemere, Kris Buchler, 
Roger Burwell, Jay Carlisle, Bev & Earl Chapin, Rich Del Carlo, Pat Cole, Marty 
Collar, Mark Collie, Pam & Gordon Comrie, Brian Cooper, Ellen Dietal, Elise and 
Dave Faike, Aubrey Fautheree, Dave Finkelnburg, John Gatchet, Dale Goble, Sarah 
Hamilton, R. L. Hand, Fran and Brad Haywood, Winifred Hepburn, Dave Holick, 
Don Johnson, Marlin Jones, Barry Kendall, Florence Knoll, Merlene Koliner, Cynthia 
Langlitz, Louise LaVoie, Joe Lipar, Matthew Moskwik, Barbara North, Eleanor 
Pnaett, Jimmy Reynolds, John Shipley, Paul Sieracki, Jim Spohn, Brian Sturges, 11a 
and Dan Svingen, Sue and Peder Svingen, Colleen Sweeney, Charles Swift, Dave 
Trochlell, Chuck Trost, Carole Vande Vorde, Fred Vanhove, Wendy Warren, John 
Weber, and Susan Weller. When this manuscript was submitted, Kas Dumroese was 
with the Department of Forest Resources, University of Idaho, Moscow, and Mark 
Mousseaux was botanist on the Idaho Panhandle National Forest. 

LITERATURE CITED 

American Ornithologists’ Union. 1983. Check-list of North American Birds, 6th ed. 
Am. Ornithol. Union, Washington, D.C. 

Burleigh, T. D. 1972. Birds of Idaho. Caxton, Caldwell, ID. 

Collins, C. T., and Foerster, K. S. 1995. Nest-site fidelity and adult longevity in the 
Black Swift (CypseJoides niger). North Am. Bird Bander 20:11-14. 

Cooper, S. V., Neiman, K. E. and Roberts D. W. 1991. Forest habitat types of 
northern Idaho: A second approximation. U.S. Dept. Agric. Forest Serv. Gen. 
Tech Rep. INT-236. 

Davis, D. G. 1964. Black Swifts nesting in a limestone cave in Colorado. Wilson Bull. 
76:295-296. 

Foerster, K, S., and Collins, C. T. 1990. Breeding distribution of the Black Swift in 
southern California. W. Birds 21:1-7. 

Hirshman, S. 1998. Black Swifts (Cypseloides niger) in Box Canyon, Ouray, 
Colorado. J. Colo. Field Ornithol. 32 (2):53-60. 

Hobbs, S. W., Griggs, A. B, Wallace, R. E., and Campbell, A. B. 1965. Geology of the 


225 


IDAHO BLACK SWIFTS: NESTING HABITAT; SPATIAL ANALYSIS OF RECORDS 


CoeurdAlene District, Shoshone County, Idaho. U.S. Dept. Interior Geol. Surv. 
Prof. Paper 478. 

Holroyd, G. L., and Holroyd, W. M. 1987. The Black Swift nest at Maligne Canyon, 
Jasper National Park. Alberta Nat. 17:46-48. 

Hunter, W. F., and Baldwin, P. C. 1962. Nesting of the Black Swift in Montana. 
Wilson Bull. 74:409-416. 

Hunter, W. F., and Baldwin, P. C. 1972. Black Swift nest in Glacier National Park. 
Murrelet 53:50-51. 

Idaho Geologic Survey. 1978. Geologic Map of Idaho (1:500,000). 

Knorr, O. A. 1961 . The geographical and ecological distribution of the Black Swift in 
Colorado. Wilson Bull. 73:155-170. 

Knorr, O. A. 1994. Black swift nesting site characteristics: Some new insights. 
Avocetta 17:139-140. 

Kondla, N. G. 1973. Nesting of the Black Swift at Johnston’s Canyon, Alberta. Can. 
Field-Nat. 87:64-65. 

Larrison, E. J. 1981. Birds of the Pacific Northwest. Univ. Idaho Press, Moscow, ID. 

Larrison, E. J., Tucker, J. L., and Jollie, M. T. 1967. Guide to Idaho Birds. J. Idaho 
Acad. Sci. 5. 

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

Marin, M. 1999. Food, foraging, and timing of breeding of the Black Swift in 
California. Wilson Bull. 111:30-37. 

Nesser, J. A., Ford, G. L., Maynard, C. L. and Page-Dumroese, D. S. 1997. 
Ecological units of the northern region: Subsections. U.S. Dept. Agric. Forest 
Serv. Gen. Tech. Rep. INT-GTR-369. 

Smith, E. 1928. Black Swifts nest behind a waterfall. Condor 30:136-138. 

Stephens, D. A., and Sturts, S. H. 1998. Idaho bird distribution. Idaho Mus. Nat. Hist. 
Spec. Publ. 13. 

Svingen, D., and Trochlell, D. 1998. Idaho-western Montana region. Field Notes 
52:93-95. 

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


Accepted 12 January 2000 


APPENDIX — Idaho Black Swift Sightings 

See Figure 1 for latilongs. Abbreviations: American Birds (AB), National Audubon 
Society Field Notes (FN), North American Birds (NAB), National Wildlife Refuge 
(NWR), breeding bird survey (BBS). 

Latilong 1 — Boundary Co. Smith Creek drainage (T64N R3W Sec. 4): 1 on 6 
Jun 1998. Kootenai NWR: 1 on 20 Jun 1985, 20 during summer 1985 (AB 39:940, 
1985), 1 on 16 May 1986, 12 on 20 Jul 1986, 1 on 1 Jun 1988, 20 on 13 Jun 
1998, 5 on 24 Jun 2001. Along Farnham Creek (T64N R1W Sec. 33 & 34): 1 seen 
often during Jun 1998. Westside Road near Copeland Bridge: 1 with 200 Vaux’s 
Swifts 29 Aug 1997. Myrtle Creek Road (T62N R1W): 4 on 15 Jul 1998. Naples 
BBS (IDA-001) following Westside Road between Myrtle and Trout creeks: 3 on 11 
Jun 1973, 7 on 12 Jun 1988, 6 on 11 Jun 1989, 1 on 23 Jun 1990, 2 on 30 Jun 


226 


IDAHO BLACK SWIFTS: NESTING HABITAT; SPATIAL ANALYSIS OF RECORDS 

1994, 7 on 15 Jun 1997. Pack River Road at Upper Pack River Bridge: 2 with 4 
Vaux’s Swifts on 7 Aug 1993. Bonner’s Ferry: 1 on 19 Jun 1989. McArthur Wildlife 
Management Area: 1+ on 27 Jul 2000. Bonner Co. Upper Grouse Creek Road: 4 
to 6 above 1220 m elevation during Jun and Jul from 1988 through 1994. Clark Fork 
River between Clark Fork and the Cabinet Gorge Dam on the Montana border: 25+ 
on 6 and 24 Jun 1985 (AB 39:940, 1985), 2 on 10 Aug 1985, 1 on 13 Aug 1985 
(Idaho Rare Bird Committee 3-88), 2 on 23 Jul 1986 (AB 40:1231, 1986), 4 to 6 
each summer from 1988 through 1996. North of Clark Fork along Lightning Creek 
(T56N R2E Sec. 13): 11 with Vaux’s Swifts on 16 Jun 1989 (AB 43: 1344," 1989), 
“many” on 13 Jun 1991. Schweitzer ski area, Sandpoint: 2 soaring on 18 Jul 1997. 
Along the Priest River north of Priest River: 6 Jun 2001. 

Latilong 2 — Kootenai Co. North shore Lake Coeur d’Alene in Coeur d’Alene: 1 
dead female on 28 May 1998 (Idaho Museum of Natural History 2643), Shoshone 
Co. North Fork Coeur d’Alene River near Cinnamon Creek (T52N R3E): 2 on 31 
May 1989, 1 with Vaux’s Swift on 1 Jun 1997 (FN 51:1025, 1997), 1 on 30 Aug 

1997. North Fork Coeur d’Alene River, Shadow and Fern falls (T51N R3E Sec. 20): 
5 including 3 and 1 nestlings, respectively, on 28 and 30 Aug 1997 (FN 52:95, 
1998), Fern Falls nestling photographed clinging vertically to its nest was erroneously 
published horizontally (FN 52:143, 1998), 1 adult on a Shadow Falls nest on 8 Jun 

1998, 1 adult on 22 and 23 Jun 1998, 5 nestlings (4 at Shadow Falls and 1 at Fern 
Falls) 26 Aug 1998 (NAB 53:80, 1999), 4 nestlings (3 at Shadow Falls and 1 at Fern 
Falls) on 27 Aug 1999; 3 nestlings at Shadow Falls on 9 Aug 2000; 4 nestlings (3 at 
Shadow Falls and 1 at Fern Falls) on 24 Aug 2001. 

Latilong 3 — Shoshone Co. Middle Sister Lookout: 1 flew by at 2040 m on 4 Aug 

2000 . 

Latilong 4 — Nez Perce Co. Lewiston Orchards: 5 or 6 on 27 May 1980. 

Latilong 5 — Idaho Co. Lochsa River, Wilderness Gateway Campground: 1 on 14 
Jul 1984, 3 on 12 Jun 1987, 5 on 17 Jun 1989 (Boulder Creek, T35N R9ESec. 34; 
AB 43:1344, 1989), 5 on 20 Aug 1990, 1 on 5 and 6 Jun 1993, 3 on 27 Jul 1997 
(FN 51: 1025, 1997). Selway River 64 km west of Lolo Pass: 1 with 8+ Vaux’s Swifts 
on 30 Jul 1993. Confluence of Lochsa and Selway rivers: mouth of Pete King Creek: 
1 foraging low with Vaux’s Swifts during stormy weather on 17 Jun 1929; Johnson 
Bar: 1 with 10 Vaux’s Swifts on 28 Jul 1993. 

Latilong 7 — Idaho Co. Joseph Plain along Billy Creek Road (about T29N & T30N 
R1W): 1 on 15 Jul 1989 (AB 43:1344, 1989). 

Latilong 8 — Idaho Co. Warren, Burgdorf BBS (IDA-208): 4 on 24 Jun 1989. 

Latilong 10 — Although Stephen and Sturts (1998) indicated a migration record, no 
data confirm it, and this is most likely a typographical error. 

Latilong 11 — Valley Co. Bear Basin, McCall: 2 on 12 Jun 2000 (NAB 54:402, 

2000 ). 

Latilong 13 — Custer Co. Spar Canyon, 25 km south of Challis: 1 with a few 
White-throated Swifts flying around a 120-m east-oriented cliff on a south-facing 
slope at 1850 m elevation 8 Jul 1984. 

Latilong 17 — Canyon Co. Deer Flat NWR: 1 flew under low clouds and over the 
New York Canal at Lake Lowell on 5 Sep 1992 (AB 47:122, 1993). Ada Co. Lucky 
Peak Bird Obseiwatory (T3N R3E Sec. 11): 30 to 35 on 25 Sep 1997 (FN 52:95, 
1998); 3 on 10 Sep 2001. Boise: 3 on 4 Jun 1999 (NAB 53:412, 1999); 2 on 5 Sep 
2000 (NAB 55:76, 2001). Owyhee Co. Near C.J. Strike Dam: 1 on 13 Oct 2000 
(erroneously published as Elmore Co. in NAB 55:76, 2001). 

Latilong 19 — Blaine Co. Hailey: 1 about 8 m above the ground in a mixed flock 
of swallows at dusk on 2 May 1998 (erroneously published as 24 May in FN 52:361, 
1998). 

Latilong 23 — Owyhee Co. Owyhee River: 2 on 1 Jun 1984. 


227 


NOTES 


DETECTIONS OF CALIFORNIA BLACK RAILS IN 
THE COLORADO RIVER DELTA, MEXICO 

OSVEL HINOJOSA-HUERTA and WILLIAM W. SHAW, 104 Biological Sciences 
East, School of Renewable Natural Resources, University of Arizona, Tucson, Arizona 
85721 

STEPHEN DeSTEFANO, U. S. Geological Survey, Massachusetts Cooperative Fish 
and Wildlife Research Unit, Holdsworth Natural Resource Center, University of 
Massachusetts, Amherst, Massachusetts 01003 


Populations of California Black Rails ( Laterallus jamaicensis coturniculus ) have 
been drastically reduced in western North America over the last several decades 
(Repking and Ohmart 1977, Evens et al. 1991). The California Black Rail is listed as 
threatened by the California Department of Fish and Game and is considered a 
“species of concern” by the U.S. Fish and Wildlife Service (www.dfg.ca.gov/endan- 
gered/birds.html). In Mexico, the California Black Rail is listed as endangered (Diario 
Oficial de la Federation 2000). Current management of the lower Colorado River, 
including the difficulty maintaining shallow and stable water levels at potentially 
suitable wetlands, is the critical factor limiting the distribution of Black Rails inland 
(Eddleman et al. 1994). Black Rails were first detected in the Colorado River delta in 
1998, at the Cienega de Santa Clara (Piest and Campoy 1999), despite rail surveys 
having been conducted there in the 1980s and early 1990s (Abarca et al. 1993, 
Eddleman et al 1994). Evens et al. (1991) speculated that Black Rails were absent 
from these wetlands because of massive habitat degradation. The purpose of our study 
was to assess the presence or absence of Black Rails in the remnant wetlands of the 
Colorado River delta, Baja California and Sonora, Mexico. 

Our surveys were focused in the Cienega de Santa Clara, at about 5800 ha the 
largest marsh in the delta. This wetland lies within the Upper Gulf of California and 
Colorado River Delta Biosphere Reserve. Other surveyed areas were Hardy River, El 
Mayor River, Pescaderos River, Riito Drain, Cucapa Complex, El Indio Lagoon, El 
Doctor wetlands, and Eastern Drains (Figure 1). The surveyed wetlands included 
emergent marshes fed by agricultural runoff and natural springs. Vegetation was 
dominated mostly by cattail ( Typha domingensis), with common reed (Phragmites 
australis ), bulrush ( Scirpus spp.), saltcedar ( Tamarix ramosissima), and salt grasses 
( Distichlis spp.) present also. Glenn et al. (2001) described each wetland. 

We conducted call-response surveys from March to June 2000 during Yuma 
Clapper Rail ( Rallus longirostris yumanensis ) surveys (Hinojosa-Huerta et al. 2001). 
After completing the Yuma Clapper Rail survey protocol, we played taped vocaliza- 
tions of California Black Rails for 2 minutes, followed by a silent period of 2 minutes 
in which we recorded all responding individuals. The tape included two types of calls 
(starting with kickee-doo, followed by grunt vocalization) arranged in a 1-minute loop. 
Surveys started at sunrise and continued no later than 1030. Survey stations were 
circular plots (30-m radius) located at least 200 m apart and grouped into routes. We 
visited each station once during the early breeding season (11-17 March) and again 
during late breeding season (18-22 May). We selected the location of the routes in the 
Cienega de Santa Clara randomly (16 routes, 173 stations, covering about 9% of the 
total area of the cienega). Other wetlands considered potential habitat were much 
smaller, and we placed routes in them nonrandomly to maximize coverage of these 
areas (8 routes, 55 stations). 


228 


Western Birds 32:228-232, 2001 


NOTES 





Key to Features 

a 

Colorado River 

b 

Handy River 

c 

El Mayor River 

d 

Laguna del Indio 

e 

Cienega de Santa Clara 

f 

El Doctor 

• 

Detections of Black Rails 

f\J Streams 

/\, 

' Levee 

LZ 

] Wetland areas 



Wetlands 

1 - Riparian corridor 

2 - Hardy - El Mayor 

3 - Floodplain 

4 - Laguna del Indio - Eastern Drains 

5 - Cienega de Santa Clara - El Doctor 


Figure 1. Wetland areas where California Black Rails were detected in the Colorado 
River delta, Baja California and Sonora, Mexico. 


We detected a total of 19 Black Rails at 18 survey stations (3.94 % of the surveyed 
stations) during the two survey periods (Figure 2). Most were located in the Cienega de 
Santa Clara, where we detected 10 Black Rails at nine stations (5.20% of the surveyed 
stations) during early breeding season and six at six stations (3.46% of the surveyed 
stations) during late breeding season. All of the detections in the late breeding season 
were >500 m away from any detection in the early breeding season (Figure 2). Cattail 
was the dominant vegetation (>50% of vegetation cover) at all stations where Black 
Rails were detected in the cienega. Bulrush was an important vegetation feature 


229 


NOTES 


^ Detections of black rails 

during early breeding season 

# Detections of black rails 

during late breeding season 

Open water 

Marsh 



Figure 2. Survey stations where California Black Rails were detected in the Cienega 
de Santa Clara, Sonora, Mexico during 2000. 


(>15% of the vegetation coverage) at five of these stations, saltcedar only at two. 

We detected three Black Rails at other wetlands in the delta during March (early 
breeding season) (present at 5.45% of the surveyed stations). One individual was along 
El Mayor River in a shallow (12-30 cm) agricultural drain dominated by cattail. A 
second Black Rail was at El Doctor wetlands in Sonora, which are maintained by 
natural springs. This area has constant shallow water throughout the year, and its 
vegetation is dominated by bulrush. The third individual was at El Indio Lagoon, Baja 
California, a recently restored wetland supported by agricultural drainage water, 
dominated by saltcedar but with substantial cattail. We detected no Black Rails at 


230 


NOTES 


Hardy River, Pescaderos River, Riito Drain, Cucapa Complex, or Eastern Drains. This 
absence may be related to the dominance of saltcedar and the lack of a constant 
source of water at these wetlands (Hinojosa-Huerta 2000). 

California Black Rails were scarce and located at only a few sites in the Colorado 
River delta during our surveys in 2000. However, these detections document 
California Black Rails occurring in these wetlands during the breeding season. These 
sites should be monitored for changes in the population, with a survey design specific 
for the subspecies, to avoid potential disruption caused by other procedures (e.g., 
eliciting taped calls of Yuma Clapper Rails). An optimum protocol to survey California 
Black Rails should follow recommendations by Spear et al. (1999), although further 
research is needed. The sequence of calls in the survey tape should be reassessed, as 
the kickee-doo call may interfere with the grunt vocalization (M. Legare pers. comm.). 

The subspecies should be included in management plans for the biosphere reserve, 
with specific practices to enhance or restore Black Rail habitat. Our results indicate 
that there is potential in agricultural drainage water and remnant wetlands in the delta 
for the conservation of the subspecies. This information should be considered in the 
development of conservation and recovery plans for inland populations. 

This research was funded by the U.S. Fish and Wildlife Service, Region 2, 
(agreement number 1448-20181-98-G942), through the Arizona Cooperative Fish 
and Wildlife Research Unit. We thank Lin Piest, Jose Campoy, and Martha Roman for 
their continuous help throughout the project. Jules Evens, Robert A. Hamilton, 
Michael Legare, and Philip Unitt made helpful comments on an earlier draft. 


LITERATURE CITED 

Abarca, F. J,, Ingraldi, M. F., and Varela-Romero, A. 1993. Observations on the 
desert pupfish (Cyprinodon macutarius), Yuma Clapper Rail (Ratlus iongirostris 
yumanensis), and shorebird communities in the Cienega de Santa Clara, 
Sonora, Mexico. Nongame and Endangered Wildlife Program Tech. Rep., Ariz. 
Game and Fish Dept., 2221 West Greenway Road, Phoenix, AZ 85023-4312. 

Diario Oficial de la Federacion. 2000. PROY-NOM-059-ECOL-2000 16/OCT/ 
2000 Protection ambiental — Especies de flora y fauna silvestres de Mexico. — 
Categorias de riesgo y especificaciones para su inclusion, exclusion o cambio. — 
Lista de especies en riesgo. Secretaria de Gobernacion, Mexico, D.F. 

Eddleman, W. R,, Flores, R. E., and Legare, M. L. 1994. Black Rail ( Laterallus 
jamaicensis), in The Birds of North America (A. Poole and F. Gill, eds.), no 123. 
Acad. Nat. Sen, Philadelphia. 

Evens, J. G., Page, G. W., Laymon, S. A., and Stallcup, R. W. 1991. Distribution, 
relative abundance, and status of the California Black Rail in western North 
America. Condor 93:952-966. 

Flores, R. E., and Eddleman, W. R. 1995. California Black Rail use of habitat in 
southwestern Arizona. J. Wildlife Mgmt. 59:357-363. 

Glenn, E. P., Zamora- Arroyo, F., Nagler, P. L., Briggs, M., Shaw, W., and Flessa, K. 
2001. Ecology and conservation biology of the Colorado River delta, Mexico. J. 
Arid Env. 49:5-15. 

Hinojosa-Huerta, O. 2000. Distribution, abundance, and habitat use of the Yuma 
Clapper Rail ( Rallus Iongirostris yumanensis ) in the Colorado River delta, 
Mexico. M.S. thesis, Univ. of Ariz., Tucson. 

Hinojosa-Huerta, O., DeStefano, S., and Shaw, W. 2001. Abundance and distribu- 
tion of the Yuma Clapper Rail (Rallus Iongirostris yumanensis) in the Colorado 
River delta, Mexico. J. Arid Env. 49:171-182. 


231 


NOTES 


Piest, L., and Campoy, J. 1999. Report of Yuma Clapper Rail surveys at the Cienega 
de Santa Clara, Sonora, 1998. Ariz. Game and Fish Dept., Yuma Regional 
Office, 9140 E. County IOV 2 St., Yuma, AZ 85365. 

Repking, C. F., and Ohmart, R. D. 1977. Distribution and density of Black Rail 
populations along the lower Colorado River. Condor 79:486-489. 

Spear, L. B., Terrill, S. B., Lenihan, C., and Delevoryas, P. 1999. Effects of temporal 
and environmental factors on the probability of detecting California Black Rails. 
J. Field Ornithol. 70:465-480. 


Accepted 27 September 2001 



232 




BOOK REVIEWS 


Birds of North America, Western Region, by Fred J. Alsop, III. Smithsonian 
Handbooks, DK Publishing, New York. Paperback, $24.95. ISBN 0-7894-7157-4. 

“Until now, no tool for identifying birds has also provided access to information on 
behavior, nesting, flight patterns, and similar birds in a compact and user-friendly 
format.” So proclaims the introduction to Birds of North America, Western Region 
(hereafter BNAW), and that this book is written for novices and for experienced 
birders. How does it uphold these claims? 

The subject matter of BNAW is birds recorded west of the 100th meridian, north of 
Mexico. A brief introduction explains how to use the book (e.g., the numerous icons), 
including sections on watching birds and conservation. The introduction is not free of 
miscues: the “Ruby-throated Hummingbird” photo (p. 13) is of a Plain-capped 
Starthroat, and in the abundance and distribution terms (pp. 24-25) the European 
Starling is “abundant,” while the Cattle Egret is “exotic,” and the Antillean Nighthawk 
(unknown in the West) is “rare.” Single-page accounts follow for 696 species, in the 
latest AOU sequence. Each has one or two color photos (or paintings), a small color 
map, and various other information conveyed by text (song, behavior, breeding, 
nesting, population) or icons (flight pattern, nest identification, habitat). A small box 
covers similar species. The book concludes with a list of 80 “accidental” species, a 
glossary, and an index. 

I have yet to see an identification guide that works well for all levels. In regard to 
BNAW, I doubt a beginner wants to wade through many pages of species he will never 
see in an attempt to identify a bird in his yard. Advanced birders will be frustrated (or 
amused) by the concept and execution: e.g., criteria for which species are addressed 
in accounts appear purely whimsical: included are the Streaked (but not Manx) 
Shearwater, Red-tailed (but not Red-billed) Tropicbird, Great Knot (but not Ruff!), and 
so on. Illustrations are the crux of any identification guide, yet one full-color photo of 
a Red-throated Loon or Sanderling in breeding plumage is unlikely to help beginners 
identify these species during most of the year. Photos generally have been trimmed to 
the bird’s outline, but some retain an unappealing and distracting color outline (e.g., 
the Great Egret and Baird’s Sandpiper), and many are of birds in poses unhelpful for 
identification. The paintings appear to have come from “professional” design people 
unfamiliar with avian anatomy or live birds. I didn’t search for misidentified photos but 
the female “Rufous” Hummingbird is a Lucifer, and the White Wagtail is not of the 
Alaskan subspecies ocularis. 

The text is of little substantive value for field identification. For other topics (such as 
behavior and nesting) it contains much that is accurate, but little reflects first-hand 
experience. An eastern bias is pervasive, and the editing is poor. Equally good or 
better life-history summaries can be found in The Birder’s Handbook by Paul Ehrlich 
et al, (Simon and Schuster, 1988, whence much information in BNAW appears to 
have been copied directly) or Kenn Kaufman’s Lives of North American Birds 
(Houghton Mifflin, 1986). For field-identification use the Sibley, National Geographic 
Society, or Kaufman guides, which are all excellent, well-conceived, and together 
cater to the full spectrum of birders. You’d be unlikely to take BNAW afield (it’s heavy, 
at 8.5" x 5.5" x 1.75”), so having two or more useful books at home would be better 
than having this one. 

BNAW appears to be a well-intentioned but poorly executed attempt to combine 
life-history information with identification. Usually I donate review copies to a library, 
but in this case I’ll recycle my copy when this review is finished. 

Steve N. G. Howell 


Western Birds 32:233-236, 2001 


233 


BOOK REVIEWS 


Isles of Refuge, by Mark J. Rauzon. 2001 . University of Hawaii Press, Honolulu. 
206 pp., 97 color and 46 black-and-white photos, 21 illustrations, 11 maps. 
Paperback, $29.95. ISBN 0-8248-2330-3. 

You are on a May pelagic trip off Bodega Bay. You reach the continental shelf, and 
suddenly the chummed-in fray is joined by one, two, 25, 50, perhaps even 500 
incoming bombers: majestic, graceful Black-footed Albatrosses splay their feet and 
eagerly scatter the screeching gulls in search of a morsel of popcorn or suet. Some of 
them have bleached heads, some are in fresh plumage, some have begun their molt, 
others have tattered wings. A few wear metal or color bands around one or both legs. 
You wonder whence they have traveled that day or that week. As was discovered 
recently through satellite transmitters, the answer is quite astounding: the bleach- 
headed birds are tending chicks over 5000 km away in the northwestern Hawaiian 
Islands. They have ridden the winds on a quick grocery run to the California coast. You 
wonder what kind of place can produce such marvels of the sea. 

Aptly named, Isles of Refuge takes you on a sun-drenched cruise through the atolls 
and pinnacles of the Black-footed Albatross and many other tropical seabirds. 
Although making up less than 0. 1% of Hawaii’s land area, the northwestern Hawaiian 
Islands span a distance of more than 1600 km, over twice that of the populated or 
“main” Hawaiian Islands. As the Pacific plate moves over Hawaii’s volcanic hotspot, 
lands are created, eroded, submerged, maintained by surface-reaching coral, and 
finally extinguished as they reach “Darwin’s Line” at 29° N. The northwestern islands 
consist of five remnant rocky volcano tops (e.g., Nihoa, Laysan, and Gardner 
Pinnacles) and six coral atolls (e.g., French Frigate Shoals, Midway, and Kure), the 
latter graced with shifting snow-white sands, sparse native vegetation, and impossibly 
blue lagoons. These remote islands have a rich and storied human history, culminating 
in their protection as national fish and wildlife refuges and Hawaii state refuges. But 
it is the stunning images and insights into the spectacular natural history of Isles of 
Refuge that will be of interest to the readers of Western Birds. 

The book is divided into 22 chapters. Those on each of the island groups are 
interspersed with others covering the biology of keystone species (such as the 
“goonies,” Monk Seal, and Green Sea Turtle), vertebrate and subtidal ecology, 
conservation concerns and initiatives, personal stories, and human campaigns. The 
last range the spectrum from a voyage of the native Hawaiian canoe Hokule’a to 
Nihoa in search of sacred Polynesian sites to the battle of Midway. Being a writer, 
photographer, and ornithologist, Rauzon shares with the reader a unique understand- 
ing and perspective of this important biological reserve. You learn how it is to be a 
biologist “stranded” for months at a time with little but the wind, waves, and seabirds 
to keep you company: “It was OK when the goonies talked to you, but when you 
understood what they said, it was time to leave.” You gain an appreciation for the 
spiritual ways and poetic views of the kamaaina, or peoples of Hawaii. You contrast 
this with the bungling, shipwrecking, murderous, warring, and ecologically apocalyp- 
tic exploitations of the haoles and other foreign visitors of the nineteenth and early 
twentieth centuries. Most important, you see hope in recent efforts to study the 
resident species, eradicate alien flora and fauna, and restore the islands to their native 
constituents. 

Beautifully written, entertaining, and with few miscues (fans of Pink Floyd will 
cringe at the reference to “Dark Side of the Sun”), I fully recommend Isles of Refuge 
to those with an interest in the seabirds and ecology of the tropical Pacific. 

Peter Pyle 


234 


BOOK REVIEWS 


Bird Songs of Southeastern Arizona and Sonora, Mexico by Geoffrey A. 
Keller. 2001. Library of Natural Sounds, Cornell Laboratory of Ornithology. Set of 2 
compact disks; $24.95. ISBN 0-938027-58-1. 

This fine set of recordings appears to be part of a series from Cornell that, to date, 
has covered Alaska (1999), the Rocky Mountain states and provinces (1999; reviewed 
in W. Birds 31:64), and the lower Rio Grande valley and southwestern Texas (2000). 
The compilation reviewed here contains vocalizations from 151 “Arizona species” 
(those with “at least one confirmed breeding record in the state” — although, appar- 
ently, a nest with two eggs of the Rufous-capped Warbler [Rosenberg and Witzeman 
1999. W. Birds 30:94-120] does not qualify) and a further 51 “Sonora species,” 
including some that stray rarely to southeastern Arizona (e.g., the Plain-capped 
Starthroat and Yellow Grosbeak). As with other works in this series, the selection of 
species is somewhat eclectic with no stated reason for inclusion or omission. For 
example, included are the Sharp-shinned Hawk but not the Red-tailed, the Eastern 
Meadowlark but not the Western. Nonetheless, apparently all of the “southeastern 
Arizona specialties” are included, which may be the main focus for most potential 
buyers. Species are arranged in AOU (1998 checklist) sequence, but the 42nd 
checklist supplement (2000) was not followed, e.g., the Arizona Woodpecker ( P. 
arizonae ) remains as Strickland’s Woodpecker ( P. stricklandi). An accompanying 
booklet provides context for most vocalizations and also gives the location of each 
recording (at the state level only, see below). The booklet includes many written 
transcriptions and often articulates helpful pointers for distinguishing vocalizations of 
similar-sounding species. Regional or subspecific variation in song is mentioned for 
some taxa, e.g., the Southwestern Willow Flycatcher ( E . t. extimus) and Eastern 
Meadowlark (S. m. lilianae), but not for others, e.g., the Northern Pygmy-Owl or 
White-breasted Nuthatch. 

The recordings are, almost without exception, of superb quality which, ironically, 
makes some sound “unnatural,” because ambient noise we usually hear in the field is 
lacking. Among many possibilities, one can make useful comparisons of the songs of 
the Bendire’s, Curve-billed, Crissal, and Le Conte’s thrashers, or of the calls of the 
Ladder-backed, Hairy, and Arizona/Strickland’s woodpeckers. The three southwest- 
ern Myiarchus flycatchers are included, as is the Nutting’s Flycatcher from Sonora, 
and other notable species included are the White-throated and Pine flycatchers and 
Black-capped Gnatcatcher. Other species are sometimes audible in the background 
but, although not identified, are unlikely to be confused with the featured species. 

The following observations are intended as constructive comments to be consid- 
ered for future editions and other works in this series and, I acknowledge, are for the 
most part niggling relative to the overall scope and quality of this undertaking. My 
main complaint is that the date and location are not provided for each cut, although, 
according to the booklet’s introductory material, such information can be requested. 
Nonetheless, in an age when appreciation of geographic variation with respect to 
subspecies (or even local dialects) is growing, the addition of date and location data 
should be mandatory and, with forethought, would not take up much space in the 
booklet. Instead, locations are given only at the state or province level, and without 
date. More than one subspecies can breed in a state, however, and subspecies not 
breeding in a region can sing during migration. Related to this, 25% of the “Arizona” 
species were recorded outside that state, as were many additional cuts for other 
species. Consequently the title is somewhat misleading and might more accurately 
read “Songs and calls of birds that occur in southeastern Arizona....” Surely there are 
recordings from Arizona of the Common Raven or Brown Creeper? The flight call of 
the male Brown-headed Cowbird is renowned for its geographic variation (the cut 
included is from Oregon), and the vocalizations of Berylline Hummingbird (from 
Chiapas, southern Mexico) pertain to a very different subspecies and are unlikely to be 
of much help in Arizona. 


235 


BOOK REVIEWS 


I recognize that the vocal repertoires of very few species are completely known, let 
alone represented by recordings. In particular, this might pertain to the Mexican 
species on these CDs — and can highlight where field work needs to be directed — but 
for other species I would hope that fuller ranges of sounds could be found, and of the 
relevant subspecies. For example, no drum is included for the Ladder-backed 
Woodpecker, and the peek call of the Hairy Woodpecker is from an Oregon 
subspecies. Some calls I hear commonly from several species are not included in the 
cuts, e.g,, the chup call of a Hermit Thmsh or the Berylline Hummingbird’s diagnostic 
buzz call. Several vocalizations sound as if birds were agitated (perhaps by playback or 
near the nest?), e.g., the unusually persistent (and “atypical”) Flame-colored Tanager 
calls. If birds were responding to playback this could be noted; the resulting vocaliza- 
tions are still “natural” but may not be those heard under most field conditions. 

The distinction between a “song” and a “call” is, of course, rather anthropomor- 
phic and often somewhat subjective, but the Gray-collared Becard’s so-called “song” 
sounds like an infrequently heard “agitation call,” while the second cut is an attenuated 
version of this species’ typical “territorial song.” Users should also be aware that the 
booklet’s text sometimes compares non-analogous vocalizations. This could lead to 
mistaken beliefs that some species’ calls are quite different, e.g., the calls provided for 
Northern Cardinal (from Honda) and Pyrrhuloxia. Alternatively, the Eared Trogon’s 
voice is reportedly “very different from that of the Elegant Trogon.” True enough, but 
the relevant cuts compare Elegant Trogon “songs” with non-analogous “calls” of the 
Eared. 

My final observation, having been involved in some of the in-progress review work, 
is that the completed CDs and booklet appear not to have been subjected to final 
review — an important step not to overlook. Thus, although I am credited with 
reviewing the Mexican species for authenticity I was surprised to see the inclusion of 
the Maroon-fronted Parrot (endemic to northeast Mexico!) under the Thick-billed 
Parrot (the Nuevo Leon cut). Also note the “Common Black Hawk” from “Costa Rica: 
Puntarenas,” which, on geographic grounds, seems more likely to have been a 
Mangrove Black Hawk (split by the AOU if not by Middle American authorities, e.g., 
see Stiles and Skutch, 1989, A Guide to the Birds of Costa Rica). The booklet also 
could have benefited from a copy-editing check for grammar, but, all in all, my 
comments and complaints are minor set against the fine resource that Geoff Keller 
and Cornell have produced. Anyone interested in the songs and calls of North 
American and Mexican birds should own these CDs and, if possible, help fill in gaps 
to make future editions more comprehensive. 

Steve N. G. Howell 


236 


FEATURED PHOTO 


AN UNUSUAL PLUMAGE VARIANT 
OF THE HEERMANN’S GULL 

KIMBALL L. GARRETT, Section of Vertebrates, Natural History Museum of Los 
Angeles County, 900 Exposition Blvd., Los Angeles, California 90007 


Minor individual plumage variation in gulls is one of many factors confounding field 
identification of this notoriously difficult group. Plumage variation can sometimes take 
the form of more extreme anomalies such as leucism and albinism (Grant 1986). A 
well-known variant of the Heermann’s Gull, Larus heermarmi, shows patches of 
white on the greater primary-coverts (Hubbs and Bartholomew 1951), as depicted, 
for example, by Sibley (2000); some individuals have additional white on other wing 
coverts, scapulars, or remiges (pers. obs.). White primary-covert patches occur in 
0.01% (Hubbs and Bartholomew 1951) to 0.5% (Sibley 2000) of Heermann’s Gulls. 
I photographed a different Heermann’s Gull variant in Santa Barbara, Santa Barbara 
County, California, on 16 January 2000 (see back cover). 

Typical of leucistic or albinistic gulls (Grant 1986), the Santa Barbara bird shows 
bare parts (bright red basal four-fifths of bill and thin red orbital ring) that are normally 
pigmented and in this case suggest that the bird was an adult in definitive plumage. 
The normal dark grays of the body plumage have been replaced by paler gray (many 
upperwing coverts and scapulars) or creamy white (underparts, center of back, greater 
secondary coverts, and tertials). Other photographs not reproduced here show pale 
gray underwing coverts and confirm that the normally blackish areas of the remiges 
and rectrices were dusky brown (secondaries, incoming inner two primaries, dark 
areas of rectrices) to pale brown (outer eight primaries, which are presumably faded). 
Active primary molt in Heermann’s Gulls generally commences in May (S. N. G. 
Howell pers. comm.), so this individual is anomalous in this regard, as well as in its 
pigmentation. The pure white head indicates definitive alternate plumage, which is 
typically acquired between November and January (S. N. G. Howell pers. comm., 
contra Dwight 1925). 

The bird is at least 2.5 to 3.5 years old and thus may have survived into adulthood 
in this leucistic plumage; alternatively, it bore normal subadult plumages and exhibited 
leucism only in its first (or a subsequent) definitive plumage. In this case a suite of 
characters including size, shape, and, most importantly, bare-part coloration makes 
the identification of this unusual variant simple. However, varying degrees of leucism 
shown by other gulls, in particular the large white-headed gulls of the northern 
hemisphere, can lead to difficult or intractable field-identification problems. 

I thank Steve N. G. Howell and Philip Unitt for comments that improved the 
manuscript. 

LITERATURE CITED 

Dwight, J. 1925. The gulls (Laridae) of the world; their plumages, moults, variations, 
relationships and distribution. Bull. Am. Mus. Nat. Hist. 52:63-401. 

Grant. P. J. 1986. Gulls, A Guide to Identification, 2nd ed. Buteo Books, Vermillion, 
■SD. 

Hubbs, C. L., and Bartholomew, G. A., Jr. 1951. Persistence of a rare color 
aberration in the Heermann Gull. Condor 53:221-227. 

Sibley, D. A. 2000. The Sibley Guide to Birds. Knopf, New York. 


Western Birds 32:237, 2001 


237 


PRESIDENT S MESSAGE 


At our annual meeting in Reno last fall your WFO board of directors made several 
important decisions aimed at improving the quality of our journal, increasing our 
appeal to a broader range of field ornithologists and ensuring the continued financial 
health of your organization, Western Birds has and always will be the primary focus 
of WFO, and we continue to work for its improvement. The publication fund we 
established about a year ago is the primary tool for this objective, and to that end we 
have substantially increased the amount in the fund. In the most recent issue of 
Western Birds we appealed for donations to the publication fund, and you have been 
both quick to respond and generous with your gifts. For those of you who have already 
responded we are pleased and very grateful. If you have not yet sent us your 
contribution please consider doing so. The recent WFO-sponsored trip to Yucatan, 
Mexico, led by Steve Howell and David Yee was a great success, and the tour 
participants also contributed generously to the publication fund. Because of your 
generosity we expect in the near future to publish more color photographs in the 
journal and continue our series of monographs. Thank you all for your support! 

Since its inception 31 years ago the WFO board of directors has been made up of 
nine members. Both to improve the balance of the board and to increase our appeal 
to a broader constituency we changed the WFO by-laws to increase the membership 
to 12. The board is currently well represented by top field ornithologists and experts 
from wildlife agencies. While maintaining our focus on field ornithology, we hope to 
improve the board by filling the new positions with members who have experience 
and expertise in the areas of law, marketing, and fund raising. If any of you have some 
of these qualifications I invite you to join the board. If you are interested you should 
submit your name and biography to the nominations committee, chaired by Dave 
Krueper, or to Dave Shuford or Gjon Hazard who also serve on the committee. 

In the past few months there have been several changes among key positions in 
WFO. For nearly nine years Dori Myers has served as your treasurer and membership 
chair. This is a very important responsibility in WFO, and Dori has served us very well 
for all those years. It was with great reluctance that we accepted her resignation at the 
last board meeting. Fortunately, Robbie Fisher has stepped in as her replacement, and 
recently the transition from Dori to Robbie was completed in a smooth and timely 
manner. At about the same time Lucie Clark resigned as recording secretary, but we 
were very fortunate that Kei Sochi has stepped forward to fill that position. To 
represent the Great Basin region better the newest member elected to the Board is 
Ted Floyd from the Great Basin Bird Observatory, who replaces Mark Sogge from 
Arizona. Thanks to all who have served us so well, and welcome to the new officers 
and board members. 

The annual WFO meetings continue to improve and grow in appeal to our 
membership. The upcoming meeting in Orange County should prove to be no 
exception. Under the fast-paced and very capable leadership of Catherine Waters 
from Sea and Sage Audubon, our co-sponsor, the meeting is shaping up to be one of 
the best ever. It will include a Friday evening barbecue and feature the great seabird 
expert Bob Pitman as our banquet speaker. The field trips should appeal to every level 
of interest among birders and include a wide variety of destinations. For the meeting, 
we have booked a first-class facility, the Ayers Country Inn and Suites in Costa Mesa, 
so mark your calendars for 10-13 October 2002. 

Look for details on our web site, www.wfo-cbrc.org. We hope to see you there. 

Mike San Miguel 
President, WFO 


27th ANNUAL MEETING 
OF WESTERN FIELD ORNITHOLOGISTS 


will be held 10-13 October 2002 at the San Joaquin Wildlife Sanctuary in 
Irvine, California, and at Ayres Country Inn in nearby Costa Mesa. The 
meeting is hosted and co-sponsored by the Sea and Sage Audubon Society. 

Our featured speaker on Saturday evening, 12 October, will be famed 
marine biologist Robert L. Pitman, addressing the seabirds of the eastern 
Pacific. Well enjoy also day and evening field trips, a pelagic field trip, 
afternoon scientific research presentations, and break-out sessions focusing 
on documentation for the California Bird Records Committee, an update by 
Philip Unitt on the surprises revealed by the San Diego County Bird Atlas, 
and social activities. 

Membership in Western Field Ornithologists, subscription to Western 
Birds, and the afternoon sessions are included in the meeting registration 
fee. For details see the WFO website (www.wfo-cbrc.org) or the Sea and 
Sage Audubon website (www.seaandsageaudubon.org) or contact 
Catherine Waters at robcatwaters@earthlink.net or 562-869-671 8. 


Call for Papers 

Oral and poster presentations should reflect original research, or summa- 
rize existing unpublished information, and be presented in a manner that will 
be of interest to serious amateur field ornithologists. We welcome talks on a 
wide variety of topics relevant to the ornithology of western North America, 
such as distribution, migration, identification, geographic variation, behav- 
ior, ecology, conservation, and field techniques. Plan on 20 minutes per oral 
presentation, which should include 5 minutes for questions and discussion. 
Slide and overhead projectors will be available. Should you wish to make a 
presentation by means of Power Point, you must bring the equipment 
necessary and test it before your presentation. Posters should fit within a 
width of 6 feet. 

An abstract of your talk in the following format should be submitted no 
later than 20 June to Philip Unitt at birds@sdnhm.org or San Diego Natural 
History Museum, P. O. Box 121390, San Diego, CA 92112-1390. 

LAST NAME, FIRST NAME. Your affiliation (if any), complete mailing 
address, and (optional) e-mail address. Title of Your Talk. Brief (300-word 
maximum) summary of the goals, results, and conclusions of your study. 

Procedures for submission can also be downloaded from the Sea and Sage 
Audubon website (www.seaandsageaudubon.org) or the WFO website 
(www.wfo-cbrc.org), or contact Catherine Waters at 562-869-6718 or 
robcatwaters@earthlink.net. 


239 


WESTERN BIRDS, INDEX, VOLUME 32, 2001 

Compiled by Philip Unitt 


Accipiter stria tus, 101, 107 
Acevedo, Marcos, see Arnaud, G. 
Actitis macularia, 108, 150, 152, 155 
Aegolius acadicus, 110 
Aimophila cassinii, 33 
Aix sponsa, 105 
Ajaia ajaja, 53 
Albatross, Laysan, 72 
Short-tailed, 14, 16 
Shy, 13 

Amador, Edgar, see Arnaud, G. 
Amazilia beryllina, 59, 65 
Ammodramus sauannarum , 115 
Amphispiza bilineata, 115 
Anas cyanoptera, 105 
discors, 105 
falcata, 13, 35, 44 
penelope, 68, 105 
platyrhynchos, 105 
querquedula, 18 
rubripes, 35 

Ani, Groove-billed, 26, 59, 65, 109 
Anser albifrons , 103, 105 
erythropus, 13, 35 
Anthus cervinus, 67 
hodgsoni, 13, 32 
spragueii, 32 

Aphelocoma californica, 186-187 
coerulescens , 186 
insular is, 186 
Aphriza virgata, 151 
Archilochus alexandri, 26, 127, 132 
colubris, 14, 26, 36, 132 
Ardea cinerea, 88-90 
herodias, 88, 89, 104 
Ardeola bacchus, 89 
Arenaria interpres, 151 

melanocephala , 149, 151, 152, 
153, 154, 155, 157 
Arnaud, Gustavo, Edgar Amador, and 
Marcos Acevedo, A potential threat 
to Bald Eagles in Baja California 
Sur, Mexico, 136 
Asio otus, 110 
Athene cunicularia, 73 
Avocet, American, 1-12, 150 
Aythya ferina, 14, 18, 20 
mania, 105 

Barrowclough, George F., see Sweet, P. R. 
Bartramia longicauda, 56 


Basileuterus rufifrons, 50, 62 
Bishop, Orange, 81-82 
Bittern, Least, 103, 104 
Blackbird, Brewer’s, 41 
Rusty, 63 

Black-Hawk, Common, 107 
Bluebird, Eastern, 13 
Bobolink, 103, 116 
Bobwhite, Northern, 13 
Bombycilla garrulus, 61, 114 
Booby, Blue-footed, 17, 34, 135 
Brown, 17-18, 20, 41, 134, 135 
Masked, 17 
Nazca, 17 

Red-footed, 17-18, 20 
Brachyramphus perdix , 14, 26 
Brant, 11, 53 
Branta bernicla, 11, 53 
leucopsis, 13, 41 

Brown, Nikolle L„, see LaRue, C. T. 
Bubo uirginianus, 72-73 
Bubulcus ibis, 101, 104 
Bucephala albeola , 106 
clangula, 101, 106 
islandica, 106 
Bufflehead, 106 
Bulweria buluierii, 13, 16 
fallax, 16 

Bunting, McKay’s, 13, 40 
Painted, 14, 34, 41, 42-44 
Snow, 34 
Varied, 40-41 

Burton, Kenneth M., and Sean D, 
Smith, First report of the Gray 
Heron in the United States, 88-90 
Buteo albicaudatus, 64 
albonotatus, 18, 35, 107 
brachyurus, 50, 55 
Iineatus, 53, 55, 64, 68, 107 
nitidus, 55 

platypterus , 55, 64, 68 
Buteogallus anthracinus, 107 

Calcarius lapponicus, 63, 68 
pictus, 40 

Calidris alba, 103, 108, 145, 151, 
152, 153, 153, 155, 157, 159 
alpina, 148, 149, 151, 152, 153, 
154, 155, 156, 157, 160, 161 
canutus, 151, 152, 155 
fuscicollis, 36 


240 


Western Birds 32:240-248, 2001 


INDEX 


mauri, 151, 152, 153, 156, 157, 
160, 161 
minuta, 36 

minutilla , 149, 151, 152, 153, 

154, 155, 156, 157 
pusilla, 108 
ruficoHis , 14, 35-36 
Callipepla douglasii, 64 
Calonectris leucomelas, 34 
Calothorax lucifer, 59 
Calypte anna, 73, 91, 101, 110, 127, 
131-133 
costae, 110, 132 

Campylorhynchus brunneicapillus, 
112 

Caprimulgus ridgwayi, 65 
vociferus, 65, 110 
Caracara, Crested, 44 
Caracara cheriway, 44 
Cardellina rubrifrons, 33 
Cardinalis sinuatus , 41-42 
Carduelis sinica, 44 
Carpodacus mexicanus, 74 
purpureu s, 68 

Catbird, Gray, 14, 15, 29-30, 31 
Catharus fuscescens, 14, 29, 38, 66 
guttatus, 113 
minimus, 29, 30 
ustulatus, 71, 73 

Catoptrophorus semipalmatus, 150, 
152, 153, 154, 155, 157, 161 
Centrocercus urophasianus, 64 
Cervantes-Sanchez, Juan, and Eric 
Mellink, Nesting of Brandt’s 
Cormorants in the northern Gulf of 
California, 134-135 
Charadrius alexandrinus, 150, 152, 
154, 161 
mongolus, 14, 19 
semipalmatus, 149, 150, 152, 

153, 154, 155, 157 
vociferus, 150, 152, 153, 154, 157 
l vilsonia, 14, 19, 21 
Chen rossii , 53, 101, 103, 105 
Chlidonias niger, 189-217 
Clangula hyemalis, 106 
Coccyzus americanus, 109 
erythropthalmus, 50, 59, 65 
Cogswell, Howard L., Book review: 

The California Condor, 173-141 
Colaptes auratus, 71, 73, 74-79 
Colinus virginianus , 13 
Columbfna irtca, 109 
minuta , 26, 36 


passerina, 109 
talpacoti, 26, 36 

Colwell, Mark A., Tamar Danufsky, 
Ryan L. Mathis, and Stanley W. 
Harris, Historical changes in the 
abundance and distribution of the 
American Avocet at the northern 
limit of its winter range, 1-12 
Condor, California, 137-140 
Contopus pertinax, 26, 36-37, 38 
sordid ulus, 26 

virens, 14, 26-27 , 37, 50, 59 
Cormorant. Brandt’s, 72, 134-135, 
177-178 

Double-crested, 101, 104, 134, 177 
Neotropic, 14, 18 
Pelagic, 177 

Coturnicops noveboracensis, 19, 35 
Cowbird, Brown-headed, 168, 175 
Crossbill, Red, 116 
Crotophaga sulcirostris, 26, 59, 65, 
109 

Cuckoo, Black-billed, 50, 59, 65 
Yellow-billed, 109 
Curlew, Bristle-thighed, 13, 21, 35 
Long-billed, 150 

Cyanocitta cristata, 60, 103, 112 
Cygnus buccinator, 18, 35, 44, 103, 
105 

cygnus, 14, 18 
C'ynanthus latirostris, 26 
Cypseloides niger, 50, 218-227 

Danufsky, Tamar, see Colwell, M. A. 
Dendroica castanea, 39, 62 
discolor, 50, 57, 62, 103, 114 
dominica, 32, 62 
fusca, 61, 67 
graciae, 32 
magnolia, 61 
pinus , 32, 39, 62 
striata, 39, 62 
tigrina, 61 
virens, 61, 67 

DeStefano, Stephen, see Hinojosa-H., 

°. 

Dickson, Lara L., see LaRue, C. T. 
Dolichonyx oryzivorus, 103, 116 
Dove, Common Ground, 109 
Inca, 109 
Mourning, 72 
Plain-breasted Ground, 26 
Ruddy Ground, 26, 36 
White-winged, 101, 103, 108-109 


241 


INDEX 


Dowitcher sp., 151, 152, 153, 156, 
157, 160 

Duck, American Black, 35 
Falcated, 13, 35, 44 
Long-tailed, 106 
Wood, 105 

Dumetella carolinensis, 14, 15, 29- 
30, 31 

Dumroese, R. Kasten, Mark R. 

Mousseaux, Shirley Horning Sturts, 
Daniel A. Stephens, and Paul A. 
Holick, Idaho Black Swifts: Nesting 
habitat and a spatial analysis of 
records, 218-227 
Dunlin, 148, 149, 151, 152, 153, 

154, 155, 156, 157, 160, 161 
Dunn, Jon L,, and Kimball L, Garrett, 
Featured photo: Parapatry in 
Woodhouse’s and California Scrub- 
Jays revisited, 186-187 
Dunnock, 91 

Eagle, Bald, 136 

Earnheart-Gold, Sasha, and Peter Pyle, 
Occurrence patterns of Peregrine 
Falcons on Southeast Farallon 
Island, California, by subspecies, 
age, and sex, 119-126 
Eider, King, 18, 35 
Egret, Cattle, 101, 104 
Reddish, 18, 52 
Egretto rufescens, 18, 52 
tricolor, 18 

Elanoides forficatus, 50 
Empidonax alnorum, 37, 39, 43, 171 
difficilis, 27 
flaviventris, 27, 65 
hammondii. 111 
minimus, 59, 65 
oberholseri, 111, 171 
occidentals, 171 

traillii, 37, 39, 43, 110, 167-172, 
174 

virescens, 27 
wrightii, 111 

Erickson, Richard A., and Hamilton, 
Robert A., Report of the California 
Bird Records Committee: 1998 
records, 13-49 

Escalante-Pliego, Patricia, see Sweet, P. R. 
Eudocimus albus, 44, 52, 64 
Eugenes fulgens, 59 
Euphagus carolinus, 63 
cyanocephalus, 41 


Euplectes franciscanus, 81-82 
Euptilotis neoxenus, 59, 65 
Euthlypis lachrymosa, 50, 62 

Falco fernoralis, 64 

peregrinus, 107, 119-126 
rusticolus, 14, 19 
sparuerius, 72 
Falcon, Aplomado, 64 
Peregrine, 107, 119-126 
Finch, House, 74 
Purple, 68 

Finlay, J. Cam, see Scarfe, A. 

Flicker, Northern, 71, 73, 74-79 
Flycatcher, Acadian, 27 
Alder, 37, 39, 43, 171 
Brown-crested, 27, 101, 111 
Cordilleran, 171 
Dusky, 111, 171 
Dusky-capped, 14, 27 
Fork- tailed, 67 
Gray, 111 

Gray Silky, 13-14, 44, 67 
Great Crested, 27, 65 
Hammond’s, 111 
Least, 59, 65 

Nutting’s, 50, 55, 59-60, 65 
Scissor-tailed, 103, 111 
Streaked, 28 
Sulphur-bellied, 14, 28 
Vermilion, 111 
Western, 27 

Willow, 37, 39, 43, 110, 167-172, 
174 

Yellow-bellied, 27, 65 

Gallinago gallinago, 151 
Gallinula chloropus, 107 
Gallinule, Purple, 68 
Gardali, Thomas, and Alvaro Jaramillo, 
Further evidence for a population 
decline in the Western Warbling 
Vireo, 173-176 
Garganey, 18 

Garrett, Kimball L., Featured photo: An 
unusual plumage variant of the 
Heermann’s Gull, 237; see also 
Dunn, J. L. 

Gatz, Thomas A, , Orange Bishops breed- 
ing in Phoenix, Arizona, 81-82 
Gavia adamsii, 16 
immer, 103-104 
pacifica, 104 
stellata, 51, 63 


242 


INDEX 


Geothlypis trie has, 40, 114 
Gnatcatcher, Black-capped, 50, 51, 

56, 60, 66 
Black-tailed, 60, 66 
Blue-gray, 113 
Godwit, Bar-tailed, 22 
Marbled, 145, 151, 152, 153, 155, 
157, 159 

Goldeneye, Barrow’s, 106 
Common, 101, 106 
Golden-Plover, American, 55-56, 64, 
68, 150, 179 
European, 179-181 
Pacific, 50, 53, 56, 150, 152, 154, 
179, 180 

Golet, Gregory H., Book review: The 
Riparian Bird Conservation Plan, 
184-185 

Goose, Barnacle, 13, 41 

Greater White-fronted, 103, 105 
Lesser White-fronted, 13, 35 
Ross’, 53, 101, 103, 105 
Grackle, Common, 14, 34. 41, 63, 

68, 101, 103, 116 
Great-tailed, 41, 141-143 
Grebe, Horned, 104 
Least, 34, 51, 63 
Greenfinch, Oriental, 44 
Grosbeak, Pine, 68 
Ground-Dove, Common, 109 
Plain-breasted, 26, 36 
Ruddy, 26, 36 
Grouse, Sage, 64 
Gull, Black-headed, 22 
Glaucous, 58 
Glaucous-winged, 177 
Heermann’s, 237 
Iceland, 13, 15, 22-24, 36 
Laughing, 58 

Lesser Black-backed, 14, 23, 24-25 
Little, 22 
Mew, 58 
Ross’, 13, 36 
■ Slaty-backed, 13 
Thayer’s, 23 
Western, 65, 72, 177 
Yellow-footed, 50, 54, 58, 64-65 
Gutierrez, R, J., see Zimmerman, 

G. S.’ 

Gymnogyps ca/ifornianus, 137-140 
Gyrfalcon, 14, 19 

Haematopus palliatus, 21 
Haliaeetus leucocephalus, 136 


Hamilton, Robert A., Book review: The 
Sibley Guide to Birds, 95-96; see 
also Erickson, R. A. 

Harris, Stanley W., see Colwell, M. A. 
Hawk, Broad-winged, 55, 64, 68 
Common Black, 107 
Gray, 55 

Harris’, 13, 15, 19, 21 
Red-shouldered, 53, 55, 64, 68, 

107 

Sharp-shinned, 101, 107 
Short-tailed, 50, 55 
White-tailed, 64 
Zone-tailed, 19, 35, 107 
Heliomaster constantii, 59 
Helmitheros vermiuorus, 33, 39-40, 67 
Heron, Black-crowned Night, 89 
Chinese Pond, 89 
Gray, 88-90 
Great Blue, 88, 89, 104 
Tricolored, 18 

Yellow-crowned Night, 18, 64 
Heteroscelus breuipes, 35 
irtcanus, 50, 56 

Hinojosa-Huerta, Osvel, William W. 
Shaw, and Stephen DeStefano, 
Detections of California Black Rails 
in the Colorado River delta, 

Mexico, 228-232 

Holick, Paul A., see Dumroese, R. K. 
Howell, Steve N. G., Book review: 

Birds of North America, 93-94; 
Featured photo: Field identification 
of female Allen’s and Rufous 
Hummingbirds, 97-98; Book 
review: United States Shorebird 
Conservation Plan, 182-183; 

Book review: Birds of North 
America, Western Region, 233; 
Book review: Bird Songs of 
Southeastern Arizona and Sonora, 
Mexico, 235-236 

Hummingbird, Allen’s, 91-92, 97-98, 
132 

Anna’s, 73, 91, 101, 110, 127, 
131-133 
Berylline, 59, 65 
Black-chinned, 26, 127, 132 
Broad-billed, 26 
Broad-tailed, 132 
Calliope, 127-130 
Costa’s, 110, 132 
Lucifer, 59 
Magnificent, 59 


243 


INDEX 


Ruby- throated, 14, 26, 36, 132 
Rufous, 91, 97-98, 132 
Humphrey, Joan M., see Shuford, W. D. 
Hylocichla mustelina, 29, 61, 66 

Ibis, Glossy, 13 
Sacred, 64 
White, 44, 52, 64 
Icterus bullockii, 63 
cucullatus, 71, 74 
galbula , 63, 68 
pustulatus, 63, 103, 116 
spurius, 63, 68 
Ictinia mississippiensis, 19 
Ixobrychus exilis, 103, 104 

Jagana, Northern, 54, 56 
Jacana spinosa, 54, 56 
Jaeger, Parasitic, 56, 58 
Pornarine, 50, 56 
Jaramillo, Alvaro, see Gardali, T. 

Jay, Blue, 60, 103, 112 
Florida Scrub, 186 
Island Scrub, 186 
Western Scrub, 186-187 
Junco, Dark-eyed, 40 
Guadalupe, 73-74 
Junco hyemalis, 40 
insu laris, 73-74 

Kelly, John P, Distribution and 

abundance of winter shorebirds on 
Tomales Bay, California: Implica- 
tions for conservation, 145-166 
Kestrel, American, 72 
Killdeer, 150, 152, 153, 154, 157 
Kingbird, Eastern, 111 
Thick-billed, 28 
Kiskadee, Great, 66 
Kite, Mississippi, 19 
Swallow-tailed, 50 
Kittiwake, Black-legged, 59 
Klicka, John T., see Sweet, P. R. 

Knot, Red, 151, 152, 155 
Koronkiewicz, Thomas J., see Sogge, 
M. K. 

LaHave, William S., see Zimmerman, 
G, S. 

Langridge, Suzanne M., see Sogge, M. 
K. 

Lanius excubiior, 111-112 
LaRue, Charles T., Lara L. Dickson, 
Nikolle L. Brown, John R. Spence, 


and Lawrence E. Stevens, Recent 
bird records from the Grand 
Canyon region, 1974-2000, 101- 
118 

Larus atricilla, 58 
canus, 58 

fuscus, 14, 23, 24-25 
glaucescens, 177 
glaucoides, 13, 15, 22-24, 36 
heermanni , 237 
hyperboreus, 58 
livens, 50, 54, 58, 64—65 
minutus, 22 

occidentalis, 65, 72, 177 
ridibundus, 22 
schistisagus, 13 
thayeri, 23 

Laterallus jamaicensis, 228-232 
Leonard, Janet L., Cloacal inspection 
of pecking in Allen’s Hummingbird, 
91-92 

Limnodromus sp., 151, 152, 153, 

156, 157, 160 

Limosa fedoa, 145, 151, 152, 153, 
155, 157, 159 
lapponica, 22 
Longspur, Lapland, 63, 68 
Smith’s, 40 

Loon, Common, 103-104 
Pacific, 104 
Red-throated, 51, 63 
Yellow-billed, 16 
Lophodytes cucullatus, 106 
Loxia curvirostra, 116 
Lymnocryptes minimus, 36 

Mallard, 105 

Mathis, Ryan L., see Colwell, M. A. 
Meadowlark, Western, 71, 74 
Melanerpes erythrocephalus, 59 
Melanitta fusca, 53, 106 
nigra, 53 

perspicillata, 105-106 
Melanotis caerulescens, 66-67 
Mellink, Eric, see Cervantes-Sanchez, J. 
Melospiza georgiana, 115 
lincolnii, 115 
Merganser, Hooded, 106 
Red-breasted, 107 
Mergus senator, 107 
Mimus polygiottos, 73, 113 
Mockingbird, Blue, 66-67 
Northern, 73, 113 
Molothrus ater , 168, 175 


244 


INDEX 


Montanez-Godoy, Liliana, see Sweet, P. R. 
Moorhen, Common, 107 
Motacilla alba , 15, 30-32 
lugens, 15, 30-32 

Mousseaux, Mark R., see Dumroese, R. K. 
Murrelet, Long-billed, 14, 26 
Myiarchus crinitus, 27, 65 
nuttingi , 50, 55, 59-60, 65 
tuberculifer , 14, 27 
tyrannulus, 27, 101, 111 
Myioborus miniatus, 67 
pictus, 67, 114 

Myiodynastes luieiventris, 14, 28 
maculatus, 28 

Night-Heron, Black-crowned, 89 
Yellow-crowned, 18, 64 
Nightjar, Buff-collared, 65 
Numenius americanus, 150 
phaeopus , 150 
tahitiensis, 13, 21, 35 
Nur, Nadav, see Shuford, W. D. 
Nyctanassa violacea, 18, 64 
Nycticorax nycticorax, 89 

Oceanodroma leucorhoa, 50, 52, 64 
melania, 50, 52, 64 
microsoma, 50, 52 
tethys, 44 
Opororrtis agilis, 40 
formosus, 62, 67 
Philadelphia, 33, 40, 62, 67 
Oreoscoptes montanus , 113 
Oriole, Baltimore, 63, 68 
Bullock’s, 63 
Hooded, 71, 74 
Orchard, 63, 68 
Streak-backed, 63, 103, 116 
Osprey, 107, 136 
Otus flammeolus, 109-110 
Owen, Jennifer C., see Sogge, M. K. 
Owl, Barn, 71, 72-73 
Burrowing, 73 
Flammulated, 109-110 
Great Horned, 72-73 
Long-eared, 110 
Northern Saw-whet, 110 
Spotted, 83-87 
Oystercatcher, American, 21 

Pandion haliaetus, 107, 136 
Parabuteo unicinctus, 13, 15, 19, 21 
Parula, Northern, 103, 114 
Parula americana, 103, 114 


Passerella iliaca, 115 
Passerina ciris, 14, 34, 41, 42-44 
versicolor , 40 

Paxton, Eben H., see Sogge, M. K. 
Pelecanus erythrorhynchos, 104 
occidentalis, 104 
Pelican, American White, 104 
Brown, 104 

Petrel, Black Storm, 50, 52, 64 
Bulwer’s, 13, 16 
Jouanin’s, 16 
Leach’s Storm, 50, 52, 64 
Least Storm, 50, 52 
Wedge-rumped Storm, 44 
Pewee, Eastern Wood, 14, 26-27, 37, 
50, 59 

Greater, 26, 37, 38 
Western Wood, 26-27 
Phaethon rubricauda, 34 
Phalacrocorax auritus, 101, 104, 

134, 177 

brasilianus, 14, 18 
pelagicus, 177 

penicillatus, 72, 134-135, 177-178 
Phalaenoptilus nuttallii, 110 
Phalarope, Red, 151 
Red-necked, 151 
Phalaropus fulicaria, 151 
lobatus, 151 

Phoebastria albatrus, 14, 16 
immutabilis, 72 
Phoebe, Black, 101, 111 
Eastern, 103, 111 
Pinicola enucleator, 68 
Pipit, Olive-backed, 13, 32 
Red-throated, 67 
Sprague’s, 32 
Piranga bidentata, 63 
erythrocephala , 67-68 
ludouiciana, 63 
ohvacea, 33, 40, 62-63, 67 
rubra, 101, 114-115 
Pitangus sulphuratus, 66 
Plectrophenax hyperboreus, 13, 40 
nivalis, 34 

Plegadis falcinellus, 13 
Plover, American Golden, 55-56, 64, 
68, 150, 179 

Black-bellied, 150, 152, 153, 154, 
155, 157 

European Golden, 179-181 
Mongolian, 14, 19 
Pacific Golden, 50, 53, 56, 150, 
152, 154, 179, 1890 


245 


INDEX 


Semipalmated, 149, 150, 152, 153, 

154, 155, 157 
Snowy, 150, 152, 154, 161 
Wilson’s, 14, 19, 21 

Pluuialis dominica , 55-56, 64, 68, 150 
fulva, 50, 53, 56, 150, 152, 154 
squatarola, 150, 152, 153, 154, 

155, 157 

Pochard, Common, 14, 18, 20 
Podiceps auritus, 104 
Polioptila caerulea, 113 
meianura, 60, 66 
nigriceps, 50, 51, 56, 60, 66 
Pond-Heron, Chinese, 89 
Pooecetes gramineus, 40 
Poorwill, Common, 110 
Porphgrula martinica, 68 
Porzona Carolina, 107 
Protonotaria citrea, 103, 114 
Prunella modularis, 91 
Ptilogongs cinereus, 13-14, 44, 67 
Puffinus pacificus, 14, 16-17 
puffinus , 17, 34 
Pyle, Peter, Book review: Isles of 
Refuge, 234; see also Earnheart- 
Gold, S. 

Pyrocephalus rubinus, 111 
Pyrrhuloxia, 41-42 

Quail, Elegant, 64 

Quiscalus mexicanus, 41, 141-143 
quiscula , 14, 34, 63, 68, 101, 103, 

116 

Rail, Black, 228-232 
Clapper, 228, 231 
Virginia, 107 
Yellow, 19, 35 
Rallus Iimicola, 107 
lorigirostris, 228, 231 
Recurvirostra americana, 1-12, 150 
Redstart, American, 114 
Painted, 67, 114 
Slate-throated, 67 
Rhodostethia rosea, 13, 36 
Rissa tridactyla, 59 
Robin, Rufous-backed, 61, 66 
Rosenberg, Gary H., Arizona Bird 
Committee report: 1996-1999 
records, 50-70 

Salpinctes obsoletus, 73 
Sanderling, 103, 108, 145, 151, 152, 
153, 155, 157, 159 


Sandpiper, Least, 149, 151, 152, 153, 
154, 155, 156, 157 
Semipalmated, 108 
Solitary, 108 

Spotted, 108, 150, 152, 155 
Upland, 56 

Western, 151, 152, 153, 156, 157, 
160, 161 
White-rumped. 36 

San Miguel, Mike, President’s message, 
238 

Sapsucker, Yellow-bellied, 103, 110 
Sagornis nigricans , 101, 111 
phoebe, 103, 111 

Scarfe, Ann, and J. Cam Finlay, Rapid 
second nesting by Anna’s Hum- 
mingbird near its northern breeding 
limits, 131-133 
Scaup, Greater, 105 
Scolopax minor, 13, 22 
Scoter, Black, 53 
Surf, 105-106 
White- winged, 53, 106 
Scrub-Jay, Florida, 186 
Island, 186 
Western, 186-187 
Seiurus motacilla, 62 
Selasphorus platgcercus, 132 
rufus, 91, 97-98, 132 
sasin, 91-92, 97-98, 132 
Setophaga ruticilla, 114 
Shaw, William W., see Hinojosa-H., O. 
Shearwater, Manx, 17, 34 
Streaked, 34 
Wedge-tailed, 14, 16-17 
Shrike, Northern, 111-112 
Shuford, W. David, Joan M. 

Humphrey, and Nadav Nur, 
Breeding status of the Black Tern in 
California, 189-217 
Sialia sialis, 13 

Silky-flycatcher, Gray 13-14, 44, 67 
Smith, Sean D., see Burton, K. M. 
Snipe, Common, 151 
Jack, 36 

Sogge, Mark K., Jennifer C. Owen, 
Eben H. Paxton, Suzanne M, 
Langridge, and Thomas J. 
Koronkiewicz, A targeted mist net 
capture technique for the Willow 
Flycatcher, 167-172 
Somateria spectabilis , 18, 35 
Sora, 107 

Sparrow, American Tree, 68 


246 


INDEX 


Black-throated, 115 
Brewer’s, 115 
Cassin’s, 33 
Chipping, 115 
Clay-colored, 115 
Field, 14, 33, 50, 63, 68 
Fox, 115 

Golden-crowned, 116 
Grasshopper, 115 
Lincoln’s, 115 
Swamp, 115 
Vesper, 40 

White-throated , 115-116 
Spence, John R., see LaRue, C. T. 
Sphyrapicus varius, 103, 110 
Spizella arborea, 68 
breweri, 115 
pallida, 115 
passer ina, 115 
pusilla , 14, 33, 50, 63, 68 
Spoonbill, Roseate, 53 
Starling, European, 71, 73, 113-114 
Starthroat, Plain-capped, 59 
Stelgidopteryx serripennis , 112 
Stellula calliope, 127-130 
Stephens, Daniel A., see Dumroese, R. K. 
Stercorarius parasiticus, 56, 58 
pomarinus, 50, 56 
Sterna anaethetus, 13, 25 
caspia, 103, 108 
fuscata, 14, 26 
lunata, 25 
paradisaea, 59 

Stevens, Lawrence E., see LaRue, C. T. 
Stint, Little, 36 
Red-necked, 14, 35-36 
Storm-Petrel, Black, 50, 52, 64 
Leach’s, 50, 52, 64 
Least, 50, 52 
Wedge-rumped, 44 
Strix occidentals, 83-87 
Sturnella neglecta, 71, 74 
Sturnus vulgaris, 71, 73, 113-114 
Sturts, Shirley Horning, see Dumroese, 
R. K. 

Su/o dactylatra, 17 
grand, 17 

leucogaster, 17-18, 20, 41, 134, 
135 

nebouxii, 17, 34, 135 
sula, 17-18, 20 
Surfbird, 151 

Swallow, Northern Rough-winged, 112 
Tree, 112 


Swan, Trumpeter, 18, 35, 44, 103, 
105 

Whooper, 14, 18 
Sweet, Paul R., George F. 

Barrowclough, John T. Klicka, 
Liliana Montanez-Godoy, and 
Patricia Escalante-Pliego, 
Recolonization of the flicker and 
other notes from Isla Guadalupe, 
Mexico, 71-80 
Swift, Black, 50, 218-227 

Tachybaptus dominicus, 34, 51, 63 
Tachycineta bicolor, 112 
Tanager, Flame-colored, 63 
Red-headed, 67 
Scarlet, 33, 40, 62-63, 67 
Summer, 101, 114-115 
Western, 63 
Tattler, Gray-tailed, 35 
Wandering, 50, 56 
Teal, Blue-winged, 105 
Cinnamon, 105 
Tern, Arctic, 59 
Black, 189-217 
Bridled, 13, 25 
Caspian, 103, 108 
Gray-backed, 25, 25 
Sooty, 14, 26 
Thalassarche cauta, 13 
Thrasher, Crissal, 101, 113 
Sage, 113 

Threskiornis aethiopicus, 64 
Thrush, Gray-cheeked, 29, 30 
Hermit, 113 
Swainson’s, 71, 73 
Wood, 29, 61, 66 
Thryothorus ludovicianus, 50, 60 
Toxostoma crissale, 101, 103 
Tringa melanoleuca, 150, 152, 154 
solitaria, 108 

Troglodytes aedon, 112-113 
troglodytes , 50, 60, 113 
Trogon, Eared, 59, 65 
Tropicbird, Red-tailed, 34 
Turdus rufopatliatus, 61, 66 
Turnstone, Black, 149, 151, 152, 153, 
154, 155, 157 
Ruddy, 151 

Tyrannus crassirostris, 28 
forficatus, 103, 111 
saoana, 66 
tyrannus, 111 
Tyto alba, 71, 72-73 


247 


INDEX 


Veery, 14, 29, 38, 66 
Vermiuora celata, 114 
chrysoptera, 32, 61 
peregrina, 61, 67, 68 
pinus, 32, 38-39, 50, 57, 61, 67 
Vireo, Bell’s, 60, 101, 112 

Blue-headed, 14, 25, 28-29, 37, 
50, 66 

Cassin’s, 25, 28, 112 
Hutton’s, 103, 112 
Philadelphia, 14, 28-29, 38, 60 
Red-eyed, 38, 60, 66, 112 
Warbling, 29, 173-176 
White-eyed, 28, 60, 66 
Yellow-green, 29, 38, 60, 66 
Yellow-throated, 28, 66 
Vireo bellii , 60, 101, 112 
cassinii, 25, 28, 112 
flavifrons, 28, 66 
flavoviridis, 29, 38, 60, 66 
gilvu s, 29, 173-176 
griseus , 28, 60, 66 
huttoni, 103, 112 
oliuaceus, 38, 60, 66, 112 
philadelphicus, 14, 28-29, 38, 60 
s olitarius, 14, 25, 28, 37, 50, 66 

Wagtail, Black-backed, 15, 30-32 
White, 15, 30-32 

Wahl, Terence R., Brandt’s Cormorant 
sinks at sea, 177-178 
Warbler, Bay-breasted, 39, 62 
Blackburnian, 61, 67 
Blackpoll, 39, 62 
Black-throated Green, 61, 67 
Blue-winged, 32, 38-39, 50, 57, 
61, 67 

Cape May, 61 
Connecticut, 40 
Fan-tailed, 50, 62 
Golden-winged, 32, 61 
Grace’s, 32 
Kentucky, 62, 67 
Magnolia, 61 
Mourning, 33, 40, 62, 67 
Orange-crowned, 114 


Pine, 33, 39, 62 
Prairie, 50, 57, 62, 103, 114 
Prothonotary, 103, 114 
Red-faced, 33 
Rufous-capped, 50, 62 
Tennessee, 61, 67, 68 
Worm-eating, 33, 39-40, 67 
Yellow-throated, 32, 62 
Waterthrush, Louisiana, 62 
Waxwing, Bohemian, 61, 114 
Wehtje, Walter, Featured photo: Range 
expansion of the Great-tailed 
Grackle in western North America, 
141-143 
Whimbrel, 150 
Whip-poor-will, 65, 110 
Wigeon, Eurasian, 68, 105 
Willet, 150, 152, 153, 154, 155, 157, 
161 

Williams, Brian D. C., Low-elevation 
nesting by Calliope Hummingbirds 
in the western Sierra Nevada 
foothills, 127-130 
Woodcock, American, 13, 22 
Woodpecker, Red-headed, 59 
Wood-Pewee, Eastern, 14, 26-27, 37, 
50, 59 

Western, 26-27 
Wren, Cactus, 112 
Carolina, 50, 60 
House, 112-113 
Rock, 73 

Winter, 50, 60, 113 

Yellowlegs, Greater, 150, 152, 154 
Yellowthroat, Common, 40, 114 

Zenaida asiatiea, 101, 103, 108-109 
macroura, 72 

Zimmerman, Guthrie S., William S. 
LaHaye, and R. J. Gutierrez, 
Breeding-season home ranges of 
Spotted Owls in the San Bernar- 
dino Mountains, California, 83-87 
Zonotrichia albicollis , 115-116 
atricapilla , 116 


248 


World Wide Web site: 

WESTERN BIRDS www.wfo-cbrc.org 

Quarterly Journal of Western Field Ornithologists 

President: Mike San Miguel, 2132 Highland Oaks Dr., Arcadia, CA 91006; 
sanmigbird@aol.com 

Vice-President: Daniel D. Gibson, University of Alaska Museum, 907 Yukon 
Dr., Fairbanks, AK 99775-6960 

Treasurer/Membership Secretary: Dori Myers, 6011 Saddletree Lane, Yorba 
Linda, CA 92886 

Recording Secretary: Lucie Clark, 9889 Tahoe Blvd., #56, Incline Village, NV 
89451 

Directors: Kimball Garrett, Daniel D. Gibson, Bob Gill, Gjon Hazard, Dave 
Krueper, Mike San Miguel, W. David Shuford, Mark K. Sogge, David Yee 

Editor.- Philip Unitt, San Diego Natural History Museum, P.0. Box 121390, San 
Diego, CA 92112-1390; birds@sdnhm.org 

Associate Editors: Daniel D. Gibson, Robert A. Hamilton, Ronald R. LeValley, 
Tim Manolis, Kathy Molina, Mark K. Sogge 

Graphics Manager: Virginia P. Johnson, 4637 Del Mar Ave., San Diego, CA 92107 

Photo Editor: Peter La Tourrette, 1019 Loma Prieta Ct., Los Altos, CA 94024 

Featured Photo: Robert A. Hamilton, 34 Rivo Alto Canal, Long Beach, CA 90803 

Book Reuiews: Steve N.G. Howell, Point Reyes Bird Observatory, 4990 Shoreline 
Highway, Stinson Beach, CA 94970 

Secretary, California Bird Records Committee: Guy McCaskie, P.O. Box 275, 
Imperial Beach, CA 91933-0275; guymcc@pacbell.net 

Chairman, California Bird Records Committee: Richard A. Erickson, LSA Associates, 
1 Park Plaza, Suite 500, Irvine, CA 92614; richard.erickson@lsa-assoc.com 


Membership dues, for individuals and institutions, including subscription to Western 
Birds: Patron, $1000.00; Life, $400.00 (payable in four equal annual installments); 
Supporting, $60 annually; Contributing, $34 annually; Family, $26; Regular U.S. 
$22 for one year, $41 for two years, $60 for three years, outside U.S, $27 for one 
year, $51 for two years, $73 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 orders for back issues and special publications to the Treasurer. Make 
checks payable to Western Field Ornithologists. 

Back issues of Western Birds within U.S. $24 per volume, $6.00 for single issues, 
plus $1.00 for postage. Outside the U.S. $30 per volume, $7.50 for single issues. 

The California Bird Records Committee of Western Field Ornithologists recently 
revised its 10-column Field List of California Birds (January 2000). The last list covered 
606 accepted species; the new list covers 613 species. Please send orders to WFO, 
c/o Dori Myers, Treasurer, 6011 Saddletree Lane, Yorba Linda, CA 92886. 
California addresses please add 7.75% sales tax. 

Quantity: 1-9, $1.50 each, includes shipping and handling. 10-39, $1.30 each, add $2.00 
for shipping and handling. 40 or more, $1.15 each, add $4.00 for shipping and handling. 


Published March 15, 2002 


ISSN 0045-3897