Open Access
How to translate text using browser tools
1 May 2009 Landbird Migration in the American West: Recent Progress and Future Research Directions
Jay D. Carlisle, Susan K. Skagen, Barbara E. Kus, Charles Van Riper III, Kristina L. Paxtons, Jeff F. Kelly
Author Affiliations +
Abstract

Our knowledge of avian behaviors during the nonbreeding period still lags behind that of the breeding season, but the last decade has witnessed a proliferation in research that has yielded significant progress in understanding migration patterns of North American birds. And, although historically the great majority of migration research has been conducted in the eastern half of the continent, there has been much recent progress on aspects of avian migration in the West. In particular, expanded use of techniques such as radar, plasma metabolites, mist-netting, count surveys, stable isotopes, genetic data, and animal tracking, coupled with an increase in multi-investigator collaborations, have all contributed to this growth of knowledge. There is increasing recognition that migration is likely the most limiting time of year for migratory birds, increasing the importance of continuing to decipher patterns of stopover ecology, identifying critical stopover habitats, and documenting migration routes in the diverse and changing landscapes of the American West. Here, we review and briefly synthesize the latest findings and advances in avian migration and consider research needs to guide future research on migration in the West.

INTRODUCTION

Migration is likely the most dangerous time of the annual cycle for migratory birds, although only recently has mortality during migration been quantified. More than 85% of annual mortality of the Black-throated Blue Warbler (Dendroica caerulescens) is estimated to occur during spring and fall migration, with mortality rates during migration 15 times higher than those during the breeding and wintering periods when the bird is stationary (Sillett and Holmes 2002). Similarly, 64% of annual mortality of the southwestern subspecies of Willow Flycatcher (Empidonax traillii extimus) occurs during migration, although this period constitutes only a quarter of the annual cycle (Paxton et al. 2007a). Breeding and wintering birds may face less uncertainty than actively migrating birds, which must simultaneously cope with elevated nutritional demands, abiotic perturbations, navigation challenges, migration barriers (including water and deserts), annual variation in environmental conditions along migration routes, unfamiliarity with novel habitats, increased intra- and interspecific competition, and shifting predator assemblages.

En route, migrants face increasing uncertainty in habitat and resource availability as human-associated landscape changes continue through such activities as urbanization, agriculture, livestock grazing, and water management. In addition, projected changes in global climate will intensify the existing uncertainties faced by migrants as the resources on which they depend are redistributed in space and time. Increasingly, migrants face changes in the quality and quantity of stopover sites, experience increasing numbers of collisions with wind turbines and artificially lit man-made structures, and are threatened by rapidly spreading new and re-emerging avian pathogens and increased abundance of domestic and other human-associated predators (Mehlman et al. 2005, Kuvlesky et al. 2007). To what degree these anthropogenic challenges during migration have contributed to population declines of many migratory species (Robbins et al. 1989, DeSante and George 1994, Wang and Finch 1997, Sauer et al. 2008) is difficult to quantify. However, the high mortality experienced by actively migrating birds (Sillett and Holmes 2002, Paxton et al. 2007a) suggests that human activities have ample potential to influence migrants' population dynamics (Holmes 2007). Given that migration is a critical time period in the annual cycle of birds, that humans are rapidly altering the migration landscape, and that the quality of stopover sites is strongly tied to biogeographical factors as well as to habitat (Kelly et al. 2000), we urge that a high priority be placed on the development of a more quantitative view of future scenarios to be faced by western migrants (Petit 2000).

Our objective in this paper is to advance the study of migration in western North America by (1) reviewing and briefly synthesizing the latest information on landbird migration through this region and (2) identifying research needs to guide future research. We focus on the West because this is the geographic region encompassing our collective expertise, although certainly the issues we discuss apply to bird migration throughout the world. The ideas presented here are derived in part from a symposium on “Landbird migration across the diverse western landscape: recent progress, the importance of riparian and other habitats, and future directions” held at the annual meeting of the Cooper Ornithological Society in 2007. We do not outline the entire history of these ideas, many of which are not new, but focus here on how they have been explored, advanced, or extended in the past 10–20 years.

Western North America is topographically diverse, resulting in extreme contrasts among adjacent habitat types and distinctive patterns of land use. How migratory birds travel through and use this diverse landscape during spring and autumn migration depends on several aspects of their life history, including natal habitat, fledging date, molting patterns, wintering location, flight capacity, and migration timing (e.g., Kelly et al. 1999, DeLong et al. 2005). Location, food availability, and predation risk are some of the most important factors shaping migrants' choices of stopover habitat (DeLong et al. 2005). For many woodland birds, much of the western landscape (especially deserts, xeric shrublands, and alpine areas) may provide adequate habitat only when birds occur at low densities, whereas other habitats (i.e., deciduous forests) may support extraordinary densities of migrants during some parts of the year (Moore 2000). Human modifications of the landscape (including agriculture, livestock grazing, development, and modification of riparian systems) have likely increased the proportion of the western landscape unsuitable for some species, whereas many human-created oases (e.g., golf courses, cemeteries, irrigation ditches) likely provide migrant stopover habitat in novel areas. Thus, whereas western migrants (traveling between breeding grounds in western Canada and the western United States including Alaska and wintering areas that stretch from the southwestern United States through Mexico to northern Central America) are not faced with major overwater flights like many of their eastern counterparts (Kelly and Hutto 2005), the western landscape likely holds many natural and anthropogenic challenges for migrants.

We still do not know how birds negotiate the western landscape during migration. In particular, what routes do they follow, how far do they fly each night, and what habitat types do they prefer? Do migration routes and stopover habitats differ seasonally? Do migration strategies differ with species' affinities for wintering and breeding habitat? How much flexibility exists in migration strategies? And, what relationships carry over from periods of migration to breeding and wintering that affect the fitness of western birds? These questions highlight the need to decipher patterns of stopover ecology and identify critical stopover habitats and migration routes in the diverse and changing landscape of the American West.

RECENT PROGRESS

Over the past two decades, the number of scientific papers published annually on avian migration ecology has increased more than tenfold. In North America, the focus of migration research has evolved from early studies on orientation and navigation to a broad diversity of topics, most notably climate change, distribution and phenology, evolution and genetics, linkages in the annual cycle, and physiology ( www.isiwebofknowledge.com). Research on avian migration in western North America has increased greatly over the last 10–15 years (Moore 2000, George and Dobkin 2002), advancing our knowledge of the ecology of western migratory birds (Table 1). Below we discuss many important recent findings that are helping to reshape our thinking about future research needs and directions.

HABITAT ASSOCIATIONS OF MIGRANTS

In the West, riparian habitats have long been recognized for their disproportionate contribution, relative to area, to faunal richness, particularly for breeding populations of birds and other animals. Riparian habitats are also considered critical to migrants, both as routes for migratory movement and as stopover habitats for many migrants (Stevens et al. 1977, Skagen et al. 1998, 2005, Wang et al. 1998, Kelly et al. 2000, DeLong et al. 2005). In addition to the vegetation and food resources, part of the importance of riparian areas for passage migrants might simply be the availability of water (Sapir et al. 2004). Riparian obligates and non-obligates alike use this habitat during migration, expanding its importance to the annual cycle of birds beyond those that breed there. The majority of riparian habitats in the West has been greatly altered or destroyed (Webb et al. 2007); thus, the conservation and restoration of existing riparian systems is of critical importance. In particular, management of riparian-associated vegetation during the predicted future water shortage will be critical to maintaining these diverse habitats (IPCC 2007).

Whereas the spatial configuration of riparian habitat can influence the distribution of migrants over large geographic scales, less is known about stopover-habitat selection at finer scales. Migrants can occur in high densities in small isolated riparian oases as well as along continuous riparian corridors in southeast Arizona (Skagen et al. 1998). Migrantsapos; habitat selection with regard to exotic vegetation has been examined recently by Walker (2008), who compared the abundance and diversity of autumn migrants in native vegetation and tamarisk (Tamarix ramosissima) -invaded habitats that differed in floristics and physiognomy. Along the Rio Grande, bird abundance is related to the floristic composition of the vegetation whereas species diversity is more strongly linked to habitat structure. Walker (2008) speculated that floristic composition influences food availability (e.g., fruits, seeds, arthropods), which is, in turn, a stronger determinant of migrants' habitat selection than availability of diverse foraging sites in structurally complex habitats. In Walker's (2008) study area, migrants' abundance was higher at tamarisk-dominated than at native-dominated sites, and potential differences between the two habitat types in the composition, quality, and timing of availability of their arthropod communities may explain the difference in their use by autumn migrants. Along the lower Colorado River, van Riper et al. (2008) found a similar situation with spring migrants' use of tamarisk-dominated habitats. The use of native vegetation and tamarisk by Wilson's Warbler (Wilsonia pusilla) along the Colorado River during spring migration matched the plants' flowering phenology, corresponding to likely changes in associated resource availability (Paxton et al. 2008). Likewise, both Hudson (2000) and Carlisle (unpubl. data) have found that some migrants can be abundant and gain mass during stopover in habitats composed at least partially of the exotic Russian Olive (Elaeagnus angustifolia). These results caution against assuming that exotic vegetation is universally of low value to migrants and encourage further investigations of migrant—vegetation associations. Studies of the mechanisms underlying such associations (e.g., food availability, concealment from predators) and their effects on body condition, length of stay, etc., will be especially useful to our ability to understand the relative value of different habitats to migrating landbirds. On a related note, given that natural (e.g., Skagen et al. 1998) and artificial oases attract passage migrants, it would be worthwhile to examine how the size, plant composition, and location of various oases affect their use by migrants.

TABLE 1.

A partial list of primary research findings since 2000 that have altered how we view migration in western North America.

t01_211.gif

In recent years, investigations of habitats used by western migrants have expanded to include habitats other than cottonwood- (Populus spp.) and willow- (Salix spp.) dominated riparian. In particular, recent studies have investigated migrants' use of mesquite woodlands, montane deciduous and coniferous habitats, xeric riparian scrub, desert shrublands, and other habitats (e.g., Puschock 1998, Carlisle et al. 2004, 2006, Hardy et al. 2004, DeLong et al. 2005, McGrath et al. 2008, van Riper et al. 2008). Along the Colorado River in Arizona spring migrants reach their greatest abundance as the flowering of honey mesquite (Prosopis glandulosa) peaks (McGrath et al. 2008). Within stopover habitats, migrating birds preferentially choose high flowering honey mesquite trees, which support the highest biomass of insects. Similarly, Hardy et al. (2004) found that in the Sonoran desert of southwestern Arizona xeric riparian scrub is especially important to spring migrants, particularly scrub containing mesquite and paloverde (Cercidium floridum, Parkinsonia aculeata) trees and catclaw acacia (Acacia greggii), willow, and creosote-bursage (Larrea tridentataAmbrosia spp.) vegetation. Puschock (1998) examined fall and spring use of pinyon—juniper (Pinus spp.—Juniperus spp.), mountain mahogany (Cercocarpus betuloides), mesquite scrub, and xeric riparian scrub (called “canyon bottom-arroyo” by Puschock) in New Mexico and found that whereas migrants were more common in the riparian (where capture rates were comparable to those along the Rio Grande), they did occur in all habitats examined. Thus, whereas cottonwood- and willow-dominated riparian habitats receive the greatest use, migrants use other habitats as well.

Several early studies found evidence for migrants' use of higher-elevation habitats, especially during autumn migration (Austin 1970, Greenberg et al. 1974, Blake 1984, Hutto 1985), and recent findings have extended the geographic scope over which this pattern occurs (Carlisle et al. 2004,2006, DeLong et al. 2005). In particular, montane deciduous and coniferous forests appear to support relatively high abundance and diversity of migrants during autumn migration. Both Carlisle et al. (2004) and DeLong et al. (2005) discussed the possibility that montane woodlands in and around the Great Basin may provide cooler microclimates and higher food availability during late summer and early autumn than surrounding lowlands and thus provide more favorable habitat to migrants at those times of year. Patten et al. (2003) suggested the same pattern for the Sonoran Desert region. In fact, Carlisle (2005) found that whereas arthropod and fruit availability at an autumn stopover site in southwestern Idaho differed significantly from year to year, most migrants were able to gain mass every year, indicating adequate food for refueling at this montane site. Recently, Wightman et al. (2007) showed that ponderosa pine forests with a deciduous component are also important for autumn migrants.

Findings of migrants using different habitats during autumn than during spring migration raise the question of whether there are important seasonal differences between the seasons in stopover-habitat requirements. For example, van Riper et al. (2008) found more birds during spring than fall migration along the lower Colorado River in southwestern Arizona, suggesting that some species migrate via different routes and/or use different habitats during autumn. Wethington et al. (2005) found a similar pattern in hummingbirds in western North America. As suggested above, migrants' choices of stopover habitat may ultimately be driven by patterns in food availability, which likely depends on variation in elevation, climate and weather, vegetation, and soil types.

Understanding the cues used by migrants to select stopover habitats requires knowledge of spatial patterning of habitats and how these patterns are perceived by migrants (Hilden 1965). The prevailing, and usually implicit, model for understanding migrants' habitat selection is based on a spatial hierarchy, in which the process of selecting a stopover site is thought to start with coarse-level topographical cues and progress toward finer-level distinctions among vegetation structure and floristics (Hilden 1965, Hutto 1980, Petit 2000, Deppe and Rottenberry 2008). This model implies a shift from a domain where factors extrinsic to local sites are prevalent at large spatial scales to one where factors intrinsic to local sites predominate at small spatial scales (Moore et al. 1995). In patchy landscapes of the interior West, understanding the spatial scale at which migrants shift from using extrinsic to intrinsic cues might be of greater importance to conservation of stopover sites than in more homogeneous landscapes.

MOLT MIGRATION

In North America, many species of migratory landbirds molt on the breeding grounds immediately after breeding, whereas others molt elsewhere (Pyle 1997). A higher proportion of western than eastern species and subspecies depart breeding areas before initiating flight-feather molt (∼50% vs. <10%; Rohwer et al. 2005), and this east—west difference is attributed to arid conditions across much of the West during late summer that do not support the energetic demands of molt (Rohwer et al. 2005). Although some species migrate to their winter grounds to molt, many recent studies have found that adults of several western North American species leave the breeding grounds in July or August, stop in the “Mexican monsoon region” in the southwestern United States and northwestern Mexico to molt, then resume migration to the winter grounds between western Mexico and Central America when molt is complete (e.g., Young 1991, Voelker and Rohwer 1998, Rohwer et al. 2005). Species that undertake molt migration to the desert Southwest include Bullock's Oriole (Icterus bullockii; Rohwer and Manning 1990), Lazuli Bunting (Passerina amoena; Young 1991), western populations of the Warbling Vireo (Vireo gilvus; Voelker and Rohwer 1998), and Western Tanager (Piranga ludoviciana; Butler et al. 2002). Many other species depart the breeding grounds before molting and, on the basis of feather condition and timing of migration documented for birds passing through western banding stations, are possible candidates for this list (e.g., Richardson et al. 2003, Carlisle et al. 2005a). Specifically, Rohwer et al. (2005) speculated that the Western Wood-Pewee (Contopus sordidulus), Dusky Flycatcher (Empidonax oberholseri), Cordilleran Flycatcher (E. occidentalis), Pacific-slope Flycatcher (E. difficilis), Ash-throated Flycatcher (Myiarchus cinerascens), Western Kingbird (Tyrannus verticalis), Hepatic Tanager (P. flava), Summer Tanager (P. rubra), Black-headed Grosbeak (Pheucticus melanocephalus), and Lesser Goldfinch (Carduelis psaltria) may also undertake molt migration to the desert Southwest. On the basis of observations at Southeast Farallon Island, P. Pyle (pers. comm.) speculates that some or all populations of the Gray Flycatcher (E. wrightii), Cassin's Vireo (V. cassinii), Nashville Warbler (Vermivora ruficapilla), Townsend's Warbler (Dendroica townsendi), Hermit Warbler (D. occidentalis), Chipping Sparrow (Spizella passerina), Lark Sparrow (Chondestes grammacus), Hooded Oriole (I. cucullatus), and Orchard Oriole (I. spurius) also undertake molt migration.

Because loss of habitat on molting grounds may affect adult survival adversely and, in turn, lead to population declines throughout the species' ranges, it is important to increase our knowledge of the stopover ecology of molt migrants in the desert Southwest. We need information about the species involved, habitat requirements, and duration of time spent in these molting areas (e.g., Leu and Thompson 2002). In areas that receive significant monsoon rainfall, riparian, mesquite, and montane (e.g., Hutto 1985) habitats and thorn forests in western Mexico (S. Rohwer, pers. comm.) may be especially important to molt migrants. Waterbirds also undertake molt migrations to specific habitats, such as waterfowl to large areas of protected wetlands and the Eared Grebe (Podiceps nigricollis) to Mono and Great Salt lakes. Much directed effort has been expended to conserve and manage these important areas to maintain population stability in these waterbirds (Jehl 1988, Hohman et al. 1992). Because some habitats used by molt-migrant landbirds may be threatened by development and livestock grazing in the desert Southwest, identification of critical molt-migration stopover habitats and regions will be an important first step toward establishing similar habitat protection and management for landbirds.

Two concurrent efforts to elucidate patterns of molt migration are underway, including studies by Sievert Rohwer (University of Washington) and Peter Pyle (Institute for Bird Populations) and their colleagues. These efforts are assessing the species composition, ranges, and habitat requirements of landbirds that make molt migrations to the southwestern United States and northwestern Mexico by using a combination of count surveys, mist-netting and banding, and collecting. Such studies will lead to a greater understanding of the stopover ecology of molt migrants and will contribute greatly to conservation and management strategies for critical habitats in the region.

STOPOVER ECOLOGY AND ENERGETIC CONDITION

A host of recent studies have enhanced our knowledge of migrants' stopover behavior in the West. In particular, applications of established techniques (mist-netting and associated analyses) to previously unstudied areas, and use of an important and novel technique (plasma metabolite sampling), have contributed to a growing body of literature pertaining to stopover ecology and habitat suitability (Wang et al. 1998, Kelly et al. 2002a, Cerasale 2004, Carlisle et al. 2005b, Guglielmo et al. 2005, Paxton et al. 2007b, 2008).

Recent studies by Wang et al. (1998), Kelly et al. (2002b), and Paxton et al. (2008) of Wilson's Warbler migration in the West highlight important age and sex differences in stopover ecology as well as the role that competition may play in impacting stopover success. In particular, Wang et al. (1998) found that along the Rio Grande in New Mexico northbound males arrived earlier and with more fat than females and also had shorter stopover durations and faster mass gains. During autumn, southbound immature birds had lower fat stores, slower mass gain rates, and longer stopovers than adults. Among northbound spring migrants at a stopover site on the lower Colorado River, second-year birds moved faster and more meanderingly than older adults, suggesting age-related differences in foraging efficiency (Paxton et al. 2008). In contrast, Carlisle and colleagues (unpubl. data) found no differences in stopover behavior between males and females in Wilson's Warbler or other species during spring or fall migration in eastern Idaho and found that immature Wilson's Warblers gained significantly more mass than adults (0.19 g vs. 0.05 g), albeit over a longer stopover period (∼2.1 days vs. ∼1.6 days). Interestingly, the arrival-condition indices were significantly lower for immature birds than for adults, suggesting that their need for mass gain was greater. Also, Carlisle et al. (2005b) investigated autumn stopover ecology at a southwestern Idaho montane site with deciduous shrubs adjacent to conifer forest and found that most species were able to gain mass, suggesting the importance of montane habitats to southbound migrants (see above).

Growing evidence suggests that competition may have a considerable effect on migrants' refueling rates and thus warrants further attention (Hansson and Pettersson 1989, Moore and Yong 1991, Kelly et al. 2002b, Cerasale 2004). Competition among migrants is supported by several lines of evidence: (1) concentrations of intercontinental migrants can depress local food abundance during stopover (Moore and Yong 1991), (2) southbound migrants tend to have larger initial mass decreases when at high densities (Hansson and Pettersson 1989), and (3) migrants gain mass more slowly during refueling when densities of migrants, especially potential competitors, are highest (Moore and Yong 1991, Kelly et al. 2002b, Cerasale 2004). Competitive interactions may also occur between migrant and resident (either year-round or seasonally resident) species, as suggested by strongly negatively correlated abundances of breeding-season residents and northbound migrants of shrub/canopy-foraging species at many riparian sites in southeastern Arizona (Skagen and colleagues, unpubl. data). Collectively, these findings underscore the importance of assessing what effect the densities of both residents and passage migrants, in combination with the amount of available habitat at a stopover site, have on migrants' abundance and the site's suitability.

Although mass changes of recaptured birds provide valuable insights into the suitability of stopover sites, such data require intensive (ideally daily) sampling, and recapture rates are low, often <10% (e.g., Carlisle et al. 2005b). An emerging technique, plasma-metabolite profiling, enables investigators to compare birds' fuel-deposition rate at various sites and seasons with reduced sample sizes (Guglielmo et al. 2005). With this technique, circulating levels of blood metabolites such as triglycerides, glycerol, and β-hydroxybutyrate together reflect rates of mass change in individuals at the time of collection of the blood sample. Using this technique to assess the suitability of non-native tamarisk vegetation relative to native riparian vegetation for migrating Wilson's Warblers, Cerasale (2004) found higher triglyceride levels at tamarisk-invaded sites, suggesting that birds there were gaining mass more quickly than in native vegetation. Cerasale (2004) concluded that lack of competition from other migrant species (significantly more abundant at native sites) was the most likely explanation. Smucker (2007) found that Wilson's Warbler's triglyceride levels were highest in riparian areas with dense understory vegetation and declined with decreasing vegetation structure, suggesting faster rates of mass gain in areas with more subcanopy vegetation. Wider application of this tool and other measures of individual condition (e.g., immune function, hormonal state, hematological state, polychromatia, oxidative stress, uric acid, carotenoids) to address differences among individuals, habitats, and seasons holds great potential for revealing patterns of stopover-site use and the factors influencing migrants' patterns of habitat use.

MIGRATORY CONNECTIVITY

Since the late 1980s, a major concern of ornithologists has been the downward population trajectories of many species of migratory birds (Robbins et al. 1989). It now appears that the dynamics of most migrants' population trends are characterized spatially by regions of decline and regions of increase (Sauer et al. 2008) and in the West some species are increasing, others decreasing (DeSante and George 1994). Emerging from the recent debates and research on migratory birds' population trends is the concept of population connectivity (Webster et al. 2002). Although nascent versions of this idea have been traced to Salomonsen (1955), the degree to which migration connects breeding and wintering populations has only recently been recognized as a driver of migrants' population dynamics. The core idea is that at one extreme, specific wintering and breeding populations are tightly linked, without any seasonal mixing of individuals from other populations (strong connectivity). At the other extreme, the population of birds at a particular wintering site are a random mix of individuals from throughout the breeding range, or vice versa (weak connectivity). If connectivity is strong, the effects of events in the winter range will be limited to specific breeding populations and are likely to be more noticeable. In contrast, if connectivity is weak, events in the winter range will have only diffuse effects on breeding populations across a broader area.

Several good examples of tracing wintering individuals to their population of origin (breeding) have recently emerged. Koronkiewicz et al. (2006) linked Willow Flycatchers wintering in Costa Rica to a breeding population at Roosevelt Lake in Arizona. On the basis of subspecies' distributions, Kelly and Hutto (2005) concluded that connectivity at a regional scale on the wintering grounds is weak within the Orangecrowned Warbler (Vermivora celata). Using mtDNA, Smith et al. (2005) assigned overwintering individuals of five species of neotropical migrants to either eastern or western lineages and then determined the degree of migratory connectivity between breeding and wintering populations. Their results varied, showing strong connectivity for the Yellow-breasted Chat (Icteria virens) and Swainson's Thrush (Catharus ustulatus) and weaker connectivity for Common Yellowthroat (Geothlypis trichas), Nashville Warbler, and Wilson's Warbler. These examples, even at broad scales, indicate that understanding the degree of connectivity can have serious implications for the conservation of avian populations.

Whereas most research on migratory connectivity has focused on linking breeding and wintering populations, determining the degree of migratory connectivity while birds are en route is also critical to understanding population dynamics of migratory birds, especially given the growing evidence of high mortality during migration (Sillett and Holmes 2002, Paxton et al. 2007a). Our understanding of many aspects of stopover ecology has been limited because departure and ultimate destination points are often unknown. The advent of more quantifiable ecological tracers (genetic and isotopic markers), however, has enhanced our ability to discern at broad scales how breeding populations are temporally and spatially distributed during migration. Kelly et al. (2002a) and Clegg et al. (2003) showed relatively strong connectivity for western breeding populations of Wilson's Warblers across their wintering range and revealed that Wilson's Warblers engage in leapfrog migration (i.e., birds breeding at higher latitudes winter farther south than conspecifics breeding at lower latitudes). These patterns have since been corroborated at additional sites in the West (Paxton et al. 2007b) and demonstrated for MacGillivray's Warbler (Oporornis tolmiei) as well (Paxton 2004, Paxton and van Riper 2006). Strong relationships between timing of migration and breeding location have been documented for western populations of the Orange-crowned Warbler, Common Yellowthroat, and Yellow Warbler (Dendroica petechia), although the direction of the relationship varied (Kelly 2006). There is also evidence of strong temporal patterns and chain migration (i.e., birds originating at higher latitudes wintering farther north than those from lower latitudes) of populations of immature western Sharp-shinned Hawks (Accipiter striatus) during autumn migration (Smith et al. 2003). Additional information about the distribution of breeding populations on the wintering grounds is needed to ascertain the relative prevalence of leapfrog versus chain migration among western species.

Knowing the degree of temporal and spatial segregation of breeding populations during migration provides the foundation for in-depth questions about stopover ecology that were previously not possible. For example, we can now begin to address questions about how habitat alteration or loss of habitat along migration routes affects different breeding populations. Whether a population migrates on a broad front or uses distinct migration pathways affects management strategies for stopover habitat along migration routes greatly. For example, Paxton et al. (2007b) found that Wilson's Warblers breeding in different geographic regions use different migration routes through the southwestern U.S. As a result, local habitat alterations could affect different populations differently. Although our knowledge of migratory connectivity has greatly increased and its applications are and will be critical to the future of migratory birds' conservation, the field is still new and there is an urgent need to improve the resolution of existing information and to expand the number of species for which data are available.

MIGRATION ROUTES AND BROAD-SCALE FLIGHT PATTERNS

Compared to that for the eastern U.S., our knowledge of migration routes through the West is in its infancy. That many migrants move in a broad front at regional and continental scales is suggested by the influence of biogeographical factors (size of the breeding range, breeding bird density, and distance from the breeding range) on migrants' abundance at stopover sites (Kelly et al. 1999). Riparian-associated birds migrating across the southwestern U.S. are less strongly influenced by biogeographical factors, however, and may adjust their migration routes to maximize time spent in riparian zones (Skagen et al. 2005). Moreover, specific populations may use specific migration routes, as evidenced by connectivity patterns of breeding and wintering populations of species with extensive ranges, such as Wilson's (Paxton et al. 2007b) and MacGillivray's warblers (Paxton 2004). Some species with a limited breeding distribution in the West, such as the Nashville Warbler, migrate within a small geographic region (Paxton 2004). Thus alterations to particular habitats could have different effects on different species.

Many species and populations may use different routes and/or habitat types during northbound and southbound migrations. For example, a loop migration pattern has been well documented in western hummingbirds, many of which migrate north along the coast and at lower elevations during spring, then south in autumn along several mountain ranges of the West—feeding on flowers in mountain meadows along the way (Phillips 1975, Wethington et al. 2005). Similarly, Patten et al. (2003) observed that in southern California migrants use the Salton Sink but largely avoid the area during fall, when they are found in the Peninsular Ranges or farther east into Arizona. In addition, several studies based on capture data during spring and fall migration at the same sites have found abundances in the two seasons to differ (Finch and Wang 2000, van Riper et al. 2008, Carlisle, unpubl. data). For example, in eastern Idaho, Wilson's Warblers were captured over than six times more frequently in fall than in spring, whereas Swainson's Thrushes (Catharus ustulatus) were captured over 10 times more frequently in spring than in fall (Carlisle, unpubl. data). Together, these data suggest that use of migration routes and/or stopover habitats varies by season, and that patterns vary by species.

The application of NEXRAD (NEXt Generation RADar, a nationwide network of over 150 large-scale Doppler weathersurveillance towers) data to the study of bird migration in the southwestern U.S. has recently provided valuable insights on flight patterns of migrants (Felix et al. 2008). NEXRAD stations detect birds during migration and return data that can be analyzed to determine migrants' velocity, altitude, and direction. Using data from seven towers along the border in Texas, New Mexico, and Arizona, Felix et al. (2008) determined that spring migrants fly faster and lower than fall migrants and that seasonal differences are partially explained by the altitudes of favorable winds. Future objectives of this effort are to estimate migrants' density and to describe their habitat associations across the Southwest. Once this study refines applications of this technique in the topographically diverse desert Southwest, NEXRAD data can be applied to study migration across the American West. Moreover, as radar technology continues to improve with the scheduled deployment of dual-polarimetric technology in the next three years, the capacity to monitor migration will also improve (Bachmann and Zrnic 2007).

SEASONAL INTERACTIONS AND CARRY-OVER EFFECTS

A conceptual framework likely to revolutionize how we view and study migratory landbirds has at its core the notion that conditions during one season in the life cycle of migratory birds can affect individual fitness and population size in a subsequent season (Webster and Marra 2005, Norris and Marra 2007). There is now a growing body of information, formerly limited to waterbirds, on how conditions at staging areas or stopover sites affect breeding success and winter population size, as well as how breeding success can affect arrival at autumn stopover sites (Newton 2006, Norris and Marra 2007). In migratory passerines, several patterns have emerged: conditions during winter or spring migration can affect the timing of arrival and population densities on the breeding grounds, individual condition, clutch size, and breeding success, conditions on the breeding grounds (including time spent and reproductive effort) can affect fall departure date as well as future physical condition, and the strength of seasonal interactions depends on migratory connectivity (e.g., Norris et al. 2004, Norris and Marra 2007). Using techniques described under Migratory Connectivity (above), scientists can now study these “carry-over effects” by asking, for example, if early and late migrants from the same breeding populations differ in time spent at stopovers or physiological state at arrival and departure, all factors that can influence migration success, survival, and reproductive success (Sandberg and Moore 1996, Sillett and Holmes 2002, Smith and Moore 2003, Moore et al. 2005).

Norris (2005) incorporated carry-over effects into models of the response of populations to habitat loss and found relationships between habitat loss and reproduction that were not detectable with traditional single-season approaches. Population response differed with the quality as well as quantity of habitat lost, and carry-over effects may vary by specific demographic factors and season. Given the magnitude of the threat of habitat loss facing western birds, modeling exercises such as this should play a prominent role in the conservation of migratory species, and amassing the data needed to develop such models (e.g., strength of density dependence on breeding and wintering grounds, magnitude and generality of individual carry-over effects, extent to which carry-over effects interact with population density) should be a high priority. Identification of critical demographic factors and mechanisms driving migration behavior can help to prioritize potential conservation actions and help determine which drivers can be managed (e.g., habitat loss, degradation) and which cannot (e.g., weather).

EVOLUTIONARY CONSTRAINTS, PHENOTYPIC PLASTICITY, AND FLEXIBILITY OF MIGRANTS

Migration is a process that has been shaped over time by genotypic and phenotypic responses of individuals to environmental variability. Modern species of birds arose mainly in the late Pliocene and Pleistocene (0.08–5.3 mya; James 2005), yet their current migratory pathways across North America probably date back only 18 000 years, to the retraction of the last glacial ice sheets during the late Pleistocene (Joseph 2005). Long before European settlement of North America, migrants coped with intensive and extensive perturbations such as glaciation and decade-long droughts. The “great evolutionary flexibility” associated with long-distance migration (Alerstam et al. 2003) allows for the continuing evolution of migration patterns in response to environmental changes. Changes can occur through microevolution (Bairlein and Hüppop 2006, Pulido and Berthold 2006) or through the expression of ecological flexibility of individuals. Some aspects of migration, such as migration routes of species in which parental care is relatively brief, are believed to be determined genetically. Evidence of substantial change of migration routes in historic times is documented for 26 bird species (Sutherland 1998), including the Blackcap (Sylvia atricapilla). Through experiments in captive rearing and orientation, Berthold et al. (1992) established the genetic basis of the recently evolved and novel westerly migration route of the Blackcap in continental Europe, and Bearhop et al. (2005) demonstrated that assortative mating of Blackcaps with different migration strategies is contributing to this rapid evolution of a new migratory route. Whether bird populations can alter migration routes in response to loss of breeding or wintering habitats in western North America has strong implications for future population viability (Dolman and Sutherland 1995).

Phenotypic plasticity “indicates the general capacity for change or transformation within genotypes in response to different environmental conditions” (Piersma and Drent 2003). Although historically used in the context of character states and reaction norms that do not pertain to reversible variation of an individual, the term “phenotypic plasticity” recently has been broadened to encompass both irreversible (developmental plasticity) and reversible (phenotypic flexibility) forms (Piersma and Drent 2003, Charmantier et al. 2008). Reversible plasticity within individuals, or flexibility, enables individuals to cope with unfamiliar and highly unpredictable environments they encounter throughout their annual cycles. Long-distance migrants exhibit several forms of flexibility, including structural and physiological plasticity, illustrated by the reversible structural changes in body composition (body fat, muscle mass, reproductive and digestive organs) associated with migration (Weber and Hedenstrom 2001, Piersma and Drent 2003), and ecological and dietary flexibility, behavioral responses that allow for the exploitation of the variable resources birds encounter throughout the year (Skagen and Oman 1996, Parrish 2000, Mettke-Hofmann and Greenberg 2005).

Although migrants have clearly coped with profound change through time, rates of land-use change in North America in the past several decades and the expected rate of global climate change are more rapid than the changes birds have generally experienced in their evolutionary pasts. The capacity of migrants to adjust to these rapidly changing conditions is essentially unknown (van Noordwijk et al. 2006). Whereas aspects of migration assumed to be genetically determined, including timing of the onset of migratory activity (Pulido et al. 2001) and migration route (Sutherland 1998, Pulido and Berthold 2006), have evolved in a matter of decades, the potential for widespread genetic adaptation may be constrained by the speed of anticipated environmental change. Phenotypic plasticity, the extent of which varies by species, enables migrants to track localized changes in resource quality and availability and to alter the speed of migration in response to weather conditions en route (Both et al. 2005, Marra et al. 2005). We hypothesize, however, that plasticity may be insufficient to enable some species to cope with the large-scale shifts in the timing and distribution of resources and habitats in breeding, wintering, and migratory ranges predicted by forecasts of climate change. Moreover, we believe that shortdistance migrants, especially those with a known capacity for facultative movements (e.g., Yellow-rumped Warbler, Dendroica coronata; Terrill and Ohmart 1984), will probably be better able to respond to a changing environment than will long-distance migrants with more fixed movement patterns. The potential for, and constraints on, genetic and phenotypic responses by migrating birds to changing environments are in great need of research attention so options for conservation can be identified.

FUTURE RESEARCH NEEDS

Although many prior migration studies have focused on basic aspects of life history and migration ecology, extensive discussions and questionnaire responses at recent venues have explicitly or implicitly identified conservation issues, such as land-use change and climate change, as driving forces in determining today's highest-priority research needs. The discussions also illuminated a need for collaborative self-organizing approaches to studying highly complex migration systems. In fact, collaborative study of landbirds' migration ecology may now be necessary to achieve the rapid scientific advances needed to confront urgent migration-related conservation issues (Heglund and Skagen 2005, Kelly and Hutto 2005). One example of an emerging collaboration related to the study of migratory birds is the Landbird Monitoring Network of the Americas (LaMNA;  http://www.klamathbird.org/lamna/).

Emerging questions for migration research have been identified by research ecologists and managers attending recent migration workshops, symposia, and meetings of MIGRATE (Migration Interest Group: Research Applied Toward Education;  http://migrate.ou.edu), a National Science Foundation research-coordination network pursuing an integrated understanding of animal migration. Here we synthesize questions defined at four venues: (A) a Southwest-focused migration workshop in 2004 (Skagen et al. 2004), (B) the inaugural meeting of MIGRATE in 2007, (C) questionnaires associated with a symposium on migration at the Cooper Ornithological Society 2007 meeting (28 respondents), and (D) the Partners in Flight Research and Monitoring Needs 2008 database ( http://www.partnersinflight.org/pifneeds; Table 2). Figure 1 portrays the relationships between these perceived research needs, the review information presented earlier in this document, and potential conservation and management actions.

Several common themes emerged from the venues and are summarized in Table 2, which presents primary questions generated at two or more of the four venues. The full suite of proposed topics covers a broad range of disciplines including behavior, ecology, physiology, genetics, climate change, geology, and information management. As an example of questions generated by such efforts, participants in MIGRATE defined high-priority research questions in which today's major conservation challenges were clearly identified. For example: What is the impact of environmental change on migratory life histories? In particular, what are the effects of land-use and climate change? Two of MIGRATE's highpriority questions focus on connections between migration and the remainder of the annual cycle: What are the determinants of individual fitness? What are the drivers of population dynamics in migratory animals? MIGRATE participants also prioritized the need to know more about mechanisms of coping with change: What are the determinants of behavioral plasticity in migrants, and what are the constraints on behavioral adaptation?

The answers to these and other high-priority questions will be essential for guiding management to prepare for and respond to environmental change effectively (Figure 1). In particular, coordinated efforts to protect migrants will rely on an understanding of the roles that specific habitats throughout the Western Hemisphere play in the annual cycles of migratory species. Targeting species, habitats, or sites for conservation action will be most effective when informed by research allowing prioritization of such actions on the basis of relative need and expected benefits to birds. As our understanding of the interconnectedness among the various stages of the annual cycle grows, so too will the need for international partnerships to protect migratory birds throughout their ranges. Knowledge of the capacity of birds to respond phenotypically or genetically to change will be another useful element to ranking species in terms of their need for management.

New collaborations are required to take full advantage of new technologies and conceptual frameworks to pursue coordinated investigations at multiple spatial and temporal scales over multiple continents (e.g., Bairlein 2003). Descriptions of the types of tools and approaches to address these questions are presented above; below, we offer brief suggestions for two fruitful lines of future investigation, on stopover-habitat associations and migration routes, based on our own work.

PATTERNS OF STOPOVER-HABITAT USE BY SEASON

In general, riparian areas are relatively productive during spring when conditions in much of the West are relatively cool and moist (Anderson et al. 2004, Hardy et al. 2004). In contrast, the productivity of riparian areas during the warmer and drier late summer likely depends on habitat structure and water availability. Larger streams and rivers that flow year round likely remain more productive than smaller streams that are dry during summer. Diversions on streams or small rivers can reduce water availability, which could affect migrants' stopover by reducing cover and food availability. Higher-elevation habitats may be less productive during spring migration if plant and arthropod phenology does not coincide with that of migration but may remain productive longer into the summer and autumn because of their cooler temperatures and retention of moisture.

TABLE 2.

Emerging questions and research needs for migration research as identified by research ecologiste and managers, 2004–2008.

t02_211.gif

FIGURE 1.

Topics for priority research questions on bird migration (left), as defined by research ecologists and conservation practitioners at recent meetings addressing migration, and their relationship to potential conservation and management practices (right). All topics relate to understanding migratory bird responses to land-use and climate change.

f01_211.eps

Therefore, we predict that low- to mid-elevation riparian habitats, especially in the Great Basin and Intermountain West (more xeric areas), will be especially important to migrants during spring migration. Although larger rivers (that retain water year round) likely remain important during autumn migration as well (e.g., Rio Grande), we predict that autumn migrants also rely on mid- to high-elevation deciduous shrublands, mixed woodlands, and coniferous forests because of moisture retention, fruit availability, and a delayed plant and arthropod phenology that provides more food during late summer and autumn. Reliance on montane habitats in the northern Great Basin and Intermountain West should be most pronounced during early migration (late July into August) because the hot, dry conditions at lower elevations are alleviated by cooling temperatures and increasingly frequent storms in September and October. One relatively simple and testable approach is to use remotely sensed data on net primary productivity (NPP) on a biweekly basis to predict the number of migrants on a regional scale. These NPP data integrate the existing vegetation potential as well as current climatic conditions and may reflect the types of cues that migrants are using to optimize survivorship.

MIGRATION ROUTES

Identifying species- and population-level patterns in migratory routes for western birds has important implications for how we approach conservation of habitats on which birds depend, as well as how we predict the effects of events in particular habitats. On the basis of our present knowledge, we suggest that most species of landbirds migrate through the West on a relatively broad front but that, within species, routes may vary seasonally, and specific populations are likely restricted to narrower migration routes. It is difficult to know whether the seasonal differences that are routinely documented reflect relatively local differences in habitat use in spring vs. fall or whether these observations suggest that loop migrations are common in the interior West. It seems likely that these patterns are tied both to potential for local productivity of habitats and regional drivers of that productivity (e.g., monsoons). Ultimately, understanding seasonal differences in the temporal and spatial dynamics of migration requires tracking of individual migrants.

Identifying the relevant scale at which to approach this issue for a given species might be accomplished by examining trends in populations across the species' breeding range. Species in which trends vary by population (i.e., some increasing, others declining) might be yielding evidence of strong connectivity with wintering populations and be best studied at the population level. In contrast, correlation of trends of breeding populations in species with weak connectivity would be predicted to be low, and these species would thus be appropriate for broader-scale examinations of migratory routes.

Whereas some details of migration routes and connectivity have come from banding recoveries, more recently, natural markers such as isotope and genetic data have greatly increased our understanding of migration routes and connectivity. These natural markers become even more powerful as tools when combined with radio-telemetry data. For example, the combined information from stable isotopes and radio telemetry indicates that different breeding populations of Wilson's Warbler use different migration routes when departing from a stopover site on the lower Colorado River (Paxton and van Riper unpubl. data). Considering the importance of route and connectivity information to conservation of migratory birds, a large-scale, collaborative effort using these natural markers to map ranges and migration routes of all migratory species (and populations) is warranted. Further refinement of our knowledge of routes and connectivity will be enhanced by further miniaturization of satellite GPS and other technology to enable tracking of individual birds throughout the annual cycle (e.g., Stutchbury et al. 2009). The recent advent of radios weighing as little as 0.2 g (Naef-Denzler 2006) and the availability of archival tags have pushed the envelope for extrinsic tracking technology down to and below birds of 20 g (Wikelski et al. 2007). In the next decade it will be possible to track the movements of individuals that, heretofore, have been too small to sample.

CONCLUSIONS

Tens of millions of migratory birds depend on stopover habitats in western North America for refueling during migration. On this journey, birds depend on spatially and temporally restricted resources for water, habitat, and food in otherwise arid or semi-arid landscapes. It is this reliance on habitats of limited extent, particularly riparian habitats, that makes western migratory birds highly vulnerable to environmental change and creates uncertainty for their future. Throughout the West, continued growth and expansion of human populations place increased demands on floodplains for water and land to support urbanization, agriculture, recreation, and other uses. The result is that regions with high human population densities, such as the Southwest, have suffered the loss or degradation of much riparian habitat during the last century (Webb et al. 2007). In addition to imperiling numerous obligate riparian breeders, this loss threatens birds during migration by disrupting the habitats upon which migrants depend en route. Human activities may be affecting birds in nonriparian habitats as well, such as montane deciduous woodlands and other habitat types and areas that we have yet to fully appreciate in terms of their importance to migrants.

Western migrants will persist only as long as the integrity of the chains of stopover sites connecting their breeding and wintering sites is preserved. This preservation is a formidable challenge under current conditions, made even more daunting by projections of models of global climate change (IPCC 2007) predicting changes, including higher temperatures and less precipitation in the North American Southwest, likely to alter the suitability of existing stopover sites. Moreover, anticipated changes in bird distributions (northward, upward in elevation) in response to global warming may necessitate new migratory routes, rendering existing stopover sites less important. How migratory birds will respond to the rapid rate of current and future climate change poses an unprecedented challenge to our ability to protect birds and the habitats on which they depend. The use of climate models to understand the scope and nature (e.g., negative, positive, or neutral) of effects of climate change on bird populations and to delineate ways to mitigate for and adapt to this far-reaching challenge is gaining in popularity (Seavy et al. 2008), as are interdisciplinary collaborative approaches that create opportunities to understand migratory birds within the context of global systems. We encourage these approaches and the advancement they are likely to achieve in securing the future of birds migrating through western North America.

ACKNOWLEDGMENTS

We thank the many researchers whose contributions have helped to advance the field of migration ecology—both worldwide and in the western United States. Thanks also to K. Gutzwiller, T. Carr, M. Mueller, M. Patten, and two anonymous reviewers for their helpful comments on this manuscript. The use of trade, product, or firm names in this publication is for descriptive purposes only and does not imply endorsement by the U.S. government.

LITERATURE CITED

1.

T. Alerstam , A. Hedenström , and S. Åkesson . 2003. Longdistance migration: evolution and determinants. Oikos 103:247– 260. Google Scholar

2.

B. W. Anderson, P. E. Russell, and R. D. Ohmart. 2004. Riparian revegetation: an account of two decades of experience in the arid Southwest. Avvar Books, Blythe, CA. Google Scholar

3.

G. T. Austin 1970. Migration of warblers in southern Nevada. Southwestern Naturalist 15:231–237. Google Scholar

4.

S. Bachmann , and D. Zrnic . 2007. Spectral density of polarimetric variables separating biological scatters in the VAD display. Journal of Atmospheric and Oceanic Technology 24:1186–1198. Google Scholar

5.

F. Bairlein 2003. The study of bird migrations—some future perspectives. Bird Study 50:243–253. Google Scholar

6.

F. Bairlein, and O. Hüppop. 2006. Migratory fuelling and global climate change, p. 33–47. In A. P. Møller, W. Fiedler, and P. Berthold [EDS.], Birds and climate change. Elsevier, Oxford, UK. Google Scholar

7.

S. Bearhop , W. Fiedler , R. W. Furness , S. C. Votier , S. Waldron , J. Newton , G. J. Bowen , P. Berthold , and K. Farnsworth . 2005. Assortative mating as a mechanism for rapid evolution of a migratory divide. Science 310:502–504. Google Scholar

8.

P. Berthold , A. J. Helbig , G. Mohr , and U. Querner . 1992. Rapid microevolution of migratory behaviour in a wild bird species. Nature 360:668–670. Google Scholar

9.

J. G. Blake 1984. A seasonal analysis of bird communities in southern Nevada. Southwestern Naturalist 29:463–474. Google Scholar

10.

C. Both , R. G. Bijlsma , and M. E. Visser . 2005. Climatic effects on spring migration and breeding in a long-distance migrant, the Pied Flycatcher Ficedula hypoleuca . Journal of Avian Biology 36:368–373. Google Scholar

11.

L. K. Butler , M. G. Donahue , and S. Rohwer . 2002. Moltmigration in Western Tanagers ( Piranga ludoviciana ): age effects, aerodynamics, and conservation implications. Auk 119:1010–1023. Google Scholar

12.

J. D. Carlisle 2005. Habitat use and stopover ecology of autumn landbird migrants in the Boise Foothills of southwestern Idaho. Ph.D. dissertation. University of South Dakota, Vermillion, SD. Google Scholar

13.

J. D. Carlisle, G. S. Kaltenecker, and D. L. Swanson. 2005a. Molt strategies and age differences in migration timing among autumn landbird migrants in southwestern Idaho. Auk 122:1070–1085. Google Scholar

14.

J. D. Carlisle, G. S. Kaltenecker, and D. L. Swanson. 2005b. Stopover ecology of autumn landbird migrants in the Boise Foothills of southwestern Idaho. Condor 107:244–258. Google Scholar

15.

J. D. Carlisle , S. L. Stock , G. S. Kaltenecker , and D. L. Swanson . 2004. Habitat associations, relative abundance, and species richness of autumn landbird migrants in southwestern Idaho. Condor 106:549–566. Google Scholar

16.

J. D. Carlisle , C. H. Trost , S. L. Stock , and G. S. Kaltenecker . 2006. Autumn landbird communities in the Boise Foothills and Owyhee Mountains of southwestern Idaho. Western Birds 37:215–227. Google Scholar

17.

D. J. Cerasale 2004. Physiological assessment of avian refueling performance. M. Sc. thesis, University of Montana, Missoula, MT. Google Scholar

18.

A. Charmantier , R. H. Mccleery , L. R. Cole , C. Perrins , L. E. B. Kruuk , and B. C. Sheldon . 2008. Adaptive phenotypic plasticity in response to climate change in a wild bird population. Science 320:800–803. Google Scholar

19.

S. M. Clegg , J. F. Kelly , M. Kimura , and T. B. Smith . 2003. Combining genetic markers and stable isotopes to reveal population connectivity and migration patterns in a neotropical migrant, Wilson's Warbler ( Wilsoniapusilla ). Molecular Ecology 12:819–830. Google Scholar

20.

J. P. Delong , S. W. Cox , and N. S. Cox . 2005. A comparison of avian use of high- and low-elevation sites during autumn migration in central New Mexico. Journal of Field Ornithology 76:326–333. Google Scholar

21.

J. L. Deppe , and J. T. Rotenberry . 2008. Scale-dependent habitat use by fall migratory birds: vegetation structure, floristics, and geography. Ecological Monographs 78:461–487. Google Scholar

22.

D. F. Desante , and T. L. George . 1994. Population trends in the landbirds of western North America. Studies in Avian Biology 15:173–190. Google Scholar

23.

P. M. Dolman , and W. J. Sutherland . 1995. The response of bird populations to habitat loss. Ibis 137:838–846. Google Scholar

24.

R. K. Felix Jr. , R. H. Diehl , and J. M. Ruth . 2008. Seasonal passerine migratory movements over the arid southwest. Studies in Avian Biology 37:126–137. Google Scholar

25.

D. M. Finch , and Y. Wang . 2000. Landbird migration in riparian habitats of the middle Rio Grande: A case study. Studies in Avian Biology 20:88–98. Google Scholar

26.

T. L. George, and D. S. Dobkin [Eds.]. 2002. Effects of habitat fragmentation on birds in western landscapes: contrasts with paradigms from the eastern United States. Studies in Avian Biology 25. Google Scholar

27.

R. Greenberg , T. Keeler-Wolf , and V. Keeler-Wolf . 1974. Wood warbler populations in the Yolla Bolly Mountains of California. Western Birds 5:81–90. Google Scholar

28.

C. G. Guglielmo , D. J. Cerasale , and C. Eldermire . 2005. A field validation of plasma metabolite profiling to assess refueling performance of migratory birds. Physiological and Biochemical Zoology 78:116–125. Google Scholar

29.

M. Hansson , and J. Pettersson . 1989. Competition and fat deposition in Goldcrests Regulus regulus at a migration stop-over site. Vogelwarte 35:21–31. Google Scholar

30.

P. C. Hardy , D. J. Griffin , A. J. Kuenzi , and M. L. Morrison . 2004. Occurrence and habitat use of passage neotropical migrants in the Sonoran desert. Western North American Naturalist 64:59–71. Google Scholar

31.

P. J. Heglund , and S. K. Skagen . 2005. Ecology and physiology of en route nearctic—neotropical migrat—ory birds: a call for collaboration. Condor 107:193–196. Google Scholar

32.

W L. Hohman, C. D. Ankney, and D. H. Gordon. 1992. Ecology and management of postbreeding waterfowl, p. 128–189. In B. D. J. Batt, A. D. Afton, M. G. Anderson, C. D. Ankney, D. H. Johnson, J. A. Kadlec, and G. L. Krapu [EDS.], Ecology and management of breeding waterfowl. University of Minnesota Press, Minneapolis, MN. Google Scholar

33.

R. T. Holmes 2007. Understanding population change in migratory songbirds: long-term and experimental studies of neotropical migrants in breeding and wintering areas. Ibis 149:2–13. Google Scholar

34.

S. E. Hudson 2000. Avian use of riparian areas in the mid-Columbia River basin during fall migration. M.Sc. thesis, University of Idaho, Moscow, ID. Google Scholar

35.

R. L. Hutto 1985. Seasonal changes in the habitat distribution of transient insectivorous birds in southeastern Arizona: competition mediated? Auk 102:120–132. Google Scholar

36.

Intergovermental Panel On Climate Change (IPCC). 2007. The physical science basis of climate change, < http://ipcc-wg1.ucar.edu/wg1/wg1-report.html> (20 June 2008). Google Scholar

37.

H. F. James 2005. Paleogene fossils and the radiation of modern birds. Auk 122:1049–1054. Google Scholar

38.

J. R. Jehl Jr. 1988. Biology of the Eared Grebe and Wilson's Phalarope in the nonbreeding season: a study of adaptations to saline lakes. Studies in Avian Biology 12. Google Scholar

39.

L. Joseph 2005. Molecular approaches to the evolution and ecology of migration, p. 18–26. In R. Greenberg and P. P. Marra [EDS.]. Birds of two worlds: the ecology and evolution of migration. Johns Hopkins University Press, Baltimore. Google Scholar

40.

J. F. Kelly 2006. Stable isotope evidence links breeding geography and migration timing in wood warblers (Parulidae). Auk 123:431–437. Google Scholar

41.

J. F. Kelly, V. Atudorei, Z. D. Sharp, and D. M. Finch. 2002b. Insights into Wilson's Warbler migration from analyses of hydrogen stable-isotope ratios. Oecologia 130:216–221. Google Scholar

42.

J. F. Kelly, L. S. Delay, and D. M. Finch. 2002a. Densitydependent mass gain by Wilson's Warblers during stopover. Auk 119:210–213. Google Scholar

43.

J. F. Kelly , D. M. Finch , and Y. Wang . 2000. Vegetative associations of wood warblers migrating along the middle Rio Grande Valley, New Mexico. Southwestern Naturalist 45:159–168. Google Scholar

44.

J. F. Kelly , and R. L. Hutto . 2005. An east—west comparison of migration in North American wood warblers. Condor 107:197–211. Google Scholar

45.

J. F. Kelly , R. Smith , D. M. Finch , F. R. Moore , and W. Yong . 1999. Effects of summer biogeography on the stopover abundance of Wood Warblers. Condor 101:76–85. Google Scholar

46.

T. J. Koronkiewicz , M. K. Sogge , C. Van Riper III , and E. H. Paxton . 2006. Territoriality, site fidelity, and survivorship of Willow Flycatchers in Costa Rica. Condor 108:558–570. Google Scholar

47.

W. P. Kuvlesky Jr. , L. A. Brennan , M. L. Morrison , K. K. Boydston , B. M. Ballard , and F. C. Bryant . 2007. Wind energy development and wildlife conservation: challenges and opportunities. Journal of Wildlife Management 71:2487–2498. Google Scholar

48.

M. Leu , and C. W. Thompson . 2002. The potential importance of migratory stopover sites as flight feather molt staging areas: a review for neotropical migrants. Biological Conservation 106:45-56. Google Scholar

49.

P. P. Marra , C. M. Francis , R. S. Mulvihill , and F. R. Moore . 2005. The influence of climate on the timing and rate of spring bird migration. Oecologia 142:307–315. Google Scholar

50.

L. J. Mcgrath , C. Van Riper III , and J. J. Fontaine . 2009. Flower power: tree flowering phenology as a settlement cue for migrating birds. Journal of Animal Ecology 78:22–30. Google Scholar

51.

D. W. Mehlman , S. E. Mabey , D. N. Ewert , C. Duncan , B. Abel , D. Cimprich , R. D. Sutter , and M. Woodrey . 2005. Conserving stopover sites for forest-dwelling migratory landbirds. Auk 122:1281–1290. Google Scholar

52.

C. Mettke-Hofmann, and R. Greenberg. 2005. Behavioral and cognitive adaptations to long-distance migration, p. 114–126. In R. Greenberg and P. P. Marra [EDS.], Birds of two worlds: the ecology and evolution of migration. Johns Hopkins University Press, Baltimore. Google Scholar

53.

F. R. Moore [Ed.]. 2000. Stopover ecology of neartic-neotropical landbird migrants: habitat relations and conservation implications. Studies in Avian Biology 20. Google Scholar

54.

F. R. Moore, R. J. Smith, and R. Sandberg. 2005. Stopover ecology of intercontinental migrants: en route problems and consequences for reproductive performance, p. 251–261. In R. Greenberg and P. P. Marra [EDS.], Birds of two worlds: the ecology and evolution of migration. Johns Hopkins University Press, Baltimore. Google Scholar

55.

F. R. Moore , and Y. Wang . 1991. Evidence of food-based competition among passerine migrants during stopover. Behavioral Ecology and Sociobiology 28:85–90. Google Scholar

56.

B. Naef-Daenzer , D. Fruh , M. Stalder , P. Wetli , and E. Weise . 2006. Miniaturization (0.2 g) and evaluation of attachment techniques of telemetry transmitters. Journal of Experimental Biology 208:4063–4068. Google Scholar

57.

I. Newton 2006. Can conditions experienced during migration limit the population levels of birds? Journal of Ornithology 147:146–166. Google Scholar

58.

D. R. Norris 2005. Carry-over effects and habitat quality in migratory populations. Oikos 109:176–186. Google Scholar

59.

D. R. Norris , P. P. Marra , T. K. Kyser , T. W. Sherry , and L. M. Ratcliffe . 2004. Tropical winter habitat limits reproductive success on the temperate breeding grounds in a migratory bird. Proceedings of the Royal Society London Series B 271:59–64. Google Scholar

60.

D. R. Norris , and P. P. Marra . 2007. Seasonal interactions, habitat quality, and population dynamics in migratory birds. Condor 109:535–547. Google Scholar

61.

J. D. Parrish 2000. Behavioral, energetic, and conservation implications of foraging plasticity during migration. Studies in Avian Biology 20:53–70. Google Scholar

62.

M. A. Patten, G. Mccaskie, and P. Unitt. 2006. Birds of the Salton Sea: status, biogeography, and ecology. University of California Press, Berkeley. Google Scholar

63.

E. H. Paxton, M. K. Sogge, S. L. Durst, T. C. Theimer, and J. R. Hatten. 2007a. The ecology of the Southwestern Willow Flycatcher in central Arizona: a 10-year synthesis report. U.S. Geological Survey Open-File Report 2007–1381. Google Scholar

64.

K. L. Paxton 2004. Spatial and temporal migration patterns of neotropical migrants in the southwest revealed by stable isotopes. M. Sc. thesis, Northern Arizona University, Flagstaff, AZ. Google Scholar

65.

K. L. Paxton, and C. Van Riper III. 2006. Spatial and temporal migration patterns of neotropical migrants in the Southwest revealed by stable isotopes. USGS OFR 2005–1298. U.S. Geological Survey, Southwest Biological Science Center, Sonoran Desert Research Station, University of Arizona, Tucson, AZ. Google Scholar

66.

K. L. Paxton, C. Van Riper III, T. C. Theimer, and E. H. Paxton. 2007b. Spatial and temporal migration patterns of Wilson's Warbler ( Wilsonia pusilla ) in the southwest as revealed by stable isotopes. Auk 124:162–175. Google Scholar

67.

K. L. Paxton , C. Van Riper III , and C. O'brien . 2008. Movement patterns and stopover ecology of Wilson's Warblers during spring migration on the lower Colorado River in southwestern Arizona. Condor 110:672–681. Google Scholar

68.

D. R. Petit 2000. Habitat use by landbirds along nearctic-neotropical migration routes: implications for conservation of stopover habitats. Studies in Avian Biology 20:15–33. Google Scholar

69.

A. R. Phillips 1975. The migrations of Alien's and other hummingbirds. Condor 77:196–205. Google Scholar

70.

T. Piersma , and J. Drent . 2003. Phenotypic flexibility and the evolution of organismal design. Trends in Ecology and Evolution 18:228–233. Google Scholar

71.

F. Pulido , P. Berthold , G. Mohr , and U. Querner . 2001. Heritability of the timing of autumn migration in a natural bird population. Proceedings of the Royal Society Series B 268:953–959. Google Scholar

72.

F. Pulido, and P. Berthold. 2006. Microevolutionary response to climate change, p. 151–183. In A. P. Møller, W. Fiedler, and P. Berthold [EDS.], Birds and climate change. Elsevier, Oxford, UK. Google Scholar

73.

J. E. Puschock 1998. Route selection and stopover behavior of selected passerine species in south-central New Mexico. M.Sc. thesis, New Mexico State University, Las Cruces, NM. Google Scholar

74.

P. Pyle 1997. Identification guide to North American birds, Part 1. Slate Creek Press, Bolinas, CA. Google Scholar

75.

T. W. Richardson , P. Pyle , R. Burnett , and P. Capitolo . 2003. The occurrence and seasonal distribution of migratory birds on Southeast Farallon Island, 1968–1999. Western Birds 34:58–96. Google Scholar

76.

C. S. Robbins , J. R. Sauer , R. S. Greenberg , and S. Droege . 1989. Population declines in North American birds that migrate to the neotropics. Proceedings of the National Academy of Sciences 86:7658–7662. Google Scholar

77.

S. Rohwer, L. K. Butler, and D. R. Froehlich. 2005. Ecology and demography of east—west differences in molt scheduling of neotropical migrant passerines, p. 87–105. In R. Greenberg and P. P. Marra [EDS.], Birds of two worlds: the ecology and evolution of migration. Johns Hopkins University Press, Baltimore. Google Scholar

78.

S. Rohwer , and J. Manning . 1990. Differences in timing and number of molts for Baltimore and Bullock's Orioles: implications to hybrid fitness and theories of delayed plumage maturation. Condor 92:125–140. Google Scholar

79.

F. Salomonsen 1955. The evolutionary significance of bird migration. Biologiske Meddelelser 22:1–62. Google Scholar

80.

R. Sandberg , and F. R. Moore . 1996. Fat stores and arrival on the breeding grounds: reproductive consequences for passerine migrants. Oikos 77:577–581. Google Scholar

81.

N. Sapir , I. Tsurim , B. Gal , and Z. Abramsky . 2004. The effect of water availability on fuel deposition of two staging Sylvia warblers. Journal of Avian Biology 35:25–32. Google Scholar

82.

J. R. Sauer, J. E. Hines, and J. Fallon. 2008. The North American Breeding Bird Survey, results and analysis 1966–2007. Version 5.15.2008. USGS Patuxent Wildlife Research Center, Laurel, MD. < http://www.mbr-pwrc.usgs.gov/bbs> (26 September 2008). Google Scholar

83.

N. E. Seavy , K. E. Dybala , and M. A. Snyder . 2008. Climate models and ornithology. Auk 125:1–10. Google Scholar

84.

T. S. Sillett , and R. T. Holmes . 2002. Variation in survivorship of a migratory songbird throughout its annual cycle. Journal of Animal Ecology 71:296–308. Google Scholar

85.

S. K. Skagen , D. A. Granfors , and C. P. Melcher . 2008. On determining the significance of ephemeral continental wetlands to North American migratory birds. Auk 125:20–29. Google Scholar

86.

S. K. Skagen , J. F. Kelly , C. Van Riper III , R. L. Hutto , D. M. Finch , D. J. Krueper , and C. P. Melcher . 2005. Geography of spring landbird migration through riparian habitats in southwestern North America. Condor 107:212–227. Google Scholar

87.

S. K. Skagen, C. P. Melcher, and R. Hazlewood. 2004. Migration stopover ecology of western avian populations: southwestern migration workshop. U.S. Geological Survey Open-File Report 2004–1452. Google Scholar

88.

S. K. Skagen , C. P. Melcher , W. H. Howe , and F. L. Knopf . 1998. Comparative use of riparian corridors and oases by migrating birds in southeast Arizona. Conservation Biology 12:896–909. Google Scholar

89.

S. K. Skagen , and H. D. Oman . 1996. Dietary flexibility of shorebirds in the Western Hemisphere. Canadian Field-Naturalist 110:419–444. Google Scholar

90.

R. B. Smith , T. D. Meehan , and B. O. Wolf . 2003. Assessing migration patterns of Sharp-shinned Hawks Accipiter striatus using stable-isotope and band encounter analysis. Journal of Avian Biology 34:387–392. Google Scholar

91.

R. J. Smith , and F. R. Moore . 2003. Arrival fat and reproductive performance in a long-distance passerine migrant. Oecologia 134:325–331. Google Scholar

92.

T. S. Smith, S. M. Clegg, M. Kimura, K. C. Ruegg, B. Mila, and I. J. Lovette. 2005. Molecular genetic approaches to linking breeding and overwintering areas in five neotropical migrant passerines, p. 222–234. In R. Greenberg and P. P. Marra [EDS.], Birds of two worlds: the ecology and evolution of migration. Johns Hopkins University Press, Baltimore. Google Scholar

93.

T. D. Smucker 2007. Patterns of nest success and migratory refueling performance in relation to cottonwood riparian habitat structure. M. Sc. thesis, University of Montana, Missoula, MT. Google Scholar

94.

L. E. Stevens, B. T. Brown, J. M. Simpson, and R. R. Johnson. 1977. The importance of riparian habitat to migrating birds, p. 156–164. In R. R. Johnson, and D. A. Jones [EDS.], Importance, preservation and management of riparian habitat. USDA Forest Service General Technical Report RM- 43. Google Scholar

95.

B. J. M. Stutchbury , S. A. Tarof , T. Done , E. Gow , P. M. Kramer , J. Tautin , J. W Fox , and V. Afanasyev . 2009. Tracking longdistance songbird migration by using geolocators. Science 323:896. Google Scholar

96.

W. J. Sutherland 1998. Evidence for flexibility and constraint in migration systems. Journal of Avian Biology 29:441–446. Google Scholar

97.

S. B. Terrill , and R. D. Ohmart . 1984. Facultative extension of fall migration by Yellow-rumped Warblers ( Dendroica coronata ). Auk 101:427–438. Google Scholar

98.

A. J. Van Noordwijk , F. Pulido , B. Helm , T. Coppack , J. Delingat , H. Dingle , A. Hedenström , H. Van Der Jeugd , C. Marchetti , A. Nilsson , and J. Pérez-Tris . 2006. A framework for the study of genetic variation in migratory behavior. Journal of Ornithology 147:221–233. Google Scholar

99.

C. Van Riper III, K. L. Paxton, C. J. Van Riper, K. A. Van Riper, L. J. Mcgrath, and J. J. Fontaine. 2008. The role of protected areas as bird stop-over habitat: ecology and habitat utilization by migrating land birds within Colorado River riparian forests of southwestern North America, p. 312–320. In D. Harmon [ED.], Protected areas in park management, proceedings of the 12th conference on research and resource management in parks and public lands. George Wright Society, Hancock, MI. Google Scholar

100.

G. Voelker , and S. Rohwer . 1998. Contrasts in scheduling of molt and migration in eastern and western Warbling Vireos. Auk 115:142–155. Google Scholar

101.

H. A. Walker 2008. Floristics and physiognomy determine migrant landbird response to tamarisk ( Tamarix ramosissima ) invasion in riparian areas. Auk 125:520–531. Google Scholar

102.

Y. Wang , and D. M. Finch . 1997. Population trends of migratory landbirds along the middle Rio Grande. Southwestern Naturalist 42:137–147. Google Scholar

103.

Y. Wang , D. M. Finch , F. R. Moore , and J. F. Kelly . 1998. Stopover ecology and habitat use of migratory Wilson's Warblers. Auk 115:829–842. Google Scholar

104.

R. H. Webb, S. A. Leake, and R. M. Turner. 2007. The ribbon of green: change in riparian vegetation in the southwestern United States. University of Arizona Press, Tucson, AZ. Google Scholar

105.

T. P. Weber , and A. Hedenström . 2001. Long-distance migrants as a model system of structural and physiological plasticity. Evolutionary Ecology Research 3:255–271. Google Scholar

106.

M. S. Webster, and P. P. Marra. 2005. The importance of understanding migratory connectivity and seasonal interactions, p. 199–209. In R. Greenberg and P. P. Marra [EDS.], Birds of two worlds: the ecology and evolution of migration. Johns Hopkins University Press, Baltimore Google Scholar

107.

M. S. Webster , P. P. Marra , S. H. Haig , S. Bensch , And R. T. Holmes . 2002. Links between worlds: unraveling migratory connectivity. Trends in Ecology and Evolution 17:76–83. Google Scholar

108.

S. M. Wethington, S. M. Russell, and G. C. West. 2005. Timing of hummingbird migration in southeastern Arizona: implications for conservation, p. 646–651. In C. J. Ralph and T. D. Rich [EDS.], Bird conservation implementation and integration in the Americas. USDA Forest Service General Technical Report, vol. 1. PSW-GTR-191. Google Scholar

109.

C. S. Wightman , S. S. Germaine , and P. Beier . 2007. Landbird community composition varies among seasons in a heterogeneous ponderosa pine forest. Journal of Field Ornithology 78:184–194. Google Scholar

110.

M. Wikelski , R. W. Kays , N. J. Kasdin , K. Thorup , J. A. Smith , and G. W. Swenson Jr. 2007. Going wild: what a global smallanimal tracking system could do for experimental biologists. Journal of Experimental Biology 210:181–186. Google Scholar

111.

B. E. Young 1991. Annual molts and interruption of the fall migration for molting in Lazuli Buntings. Condor 93:236–250. Google Scholar
© 2009 by The Cooper Ornithological Society. All rights reserved. Please direct all requests for permission to photocopy or reproduce article content through the University of California Press's Rights and Permissions website, http://www.ucpressjournals.com/reprintInfo.asp.
Jay D. Carlisle, Susan K. Skagen, Barbara E. Kus, Charles Van Riper III, Kristina L. Paxtons, and Jeff F. Kelly "Landbird Migration in the American West: Recent Progress and Future Research Directions," The Condor 111(2), 211-225, (1 May 2009). https://doi.org/10.1525/cond.2009.080096
Received: 25 November 2008; Accepted: 1 February 2009; Published: 1 May 2009
KEYWORDS
landbird migration
stopover ecology
western United States
Back to Top