ChapterPDF Available

Waterfowl and Wetland Birds

Authors:

Abstract and Figures

The future of wetland bird habitat and populations is intrinsically connected with the conservation of rangelands in North America. Many rangeland watersheds are source drainage for some of the highest functioning extant wetlands. The Central and Pacific Flyways have significant overlap with available rangelands in western North America. Within these flyways, the importance of rangeland management has become increasingly recognized by those involved in wetland bird conservation. Within the array of wetland bird species, seasonal habitat needs are highly variable. During the breeding period, nest survival is one of the most important drivers of population growth for many wetland bird species and rangelands often provide quality nesting cover. Throughout spring and fall, rangeland wetlands provide key forage resources that support energetic demands needed for migration. In some areas, stock ponds developed for livestock water provide migration stopover and wintering habitat, especially in times of water scarcity. In the Intermountain West, drought combined with water demands from agriculture and human population growth are likely headed to an ecological tipping point for wetland birds and their habitat in the region. In the Prairie Pothole Region, conversion of rangeland and draining of wetlands for increased crop production remains a significant conservation issue for wetland birds and other wildlife. In landscapes dominated by agricultural production, rangelands provide some of the highest value ecosystem services, including water quality and wetland function. Recent research has shown livestock grazing, if managed properly, is compatible and at times beneficial to wetland bird habitat needs. Either directly, or indirectly, wetland bird populations and their habitat needs are supported by healthy rangelands. In the future, rangeland and wetland bird managers will benefit from increased collaboration to aid in meeting ultimate conservation objectives.
Content may be subject to copyright.
Chapter 13
Waterfowl and Wetland Birds
Josh L. Vest, David A. Haukos, Neal D. Niemuth, Casey M. Setash,
James H. Gammonley, James H. Devries, and David K. Dahlgren
Abstract The future of wetland bird habitat and populations is intrinsically
connected with the conservation of rangelands in North America. Many rangeland
watersheds are source drainage for some of the highest functioning extant wetlands.
The Central and Pacific Flyways have significant overlap with available rangelands in
western North America. Within these flyways, the importance of rangeland manage-
ment has become increasingly recognized by those involved in wetland bird conser-
vation. Within the array of wetland bird species, seasonal habitat needs are highly
J. L. Vest (B)
U.S. Fish and Wildlife Service, Prairie Pothole Joint Venture, 922 Bootlegger Trail, Great Falls,
MT 59404, USA
e-mail: josh_vest@fws.gov
D. A. Haukos
U.S. Geological Survey, Kansas Cooperative Fish and Wildlife Research Unit, 1128 N. 7th Street,
Manhattan, KS 66506, USA
e-mail: dhaukos@ksu.edu
N. D. Niemuth
U.S. Fish and Wildlife Service, Habitat and Population Evaluation Team, 3425 Miriam Avenue,
Bismarck, ND 58501, USA
e-mail: neal_niemuth@fws.gov
C. M. Setash
Department of Fish, Wildlife, and Conservation Biology, Colorado State University, 1474 Campus
Delivery, Fort Collins, CO 80523, USA
e-mail: csetash@rams.colostate.edu
J. H. Gammonley
Colorado Parks and Wildlife, 317 West Prospect, Fort Collins, CO 80526, USA
e-mail: jim.gammonley@state.co.us
J. H. Devries
Ducks Unlimited Canada, Institute for Wetland and Waterfowl Research, P.O. Box 1160,
Stonewall, MB R0C 2Z0, Canada
e-mail: j_devries@ducks.ca
D. K. Dahlgren
Department of Wildland Resources, Utah State University, 5230 Old Main Hill, Logan,
UT 84322, USA
e-mail: dave.dahlgren@usu.edu
© The Author(s) 2023
L. B. McNew et al. (eds.), Rangeland Wildlife Ecology and Conservation,
https://doi.org/10.1007/978-3-031-34037-6_13
417
418 J. L. Vest et al.
variable. During the breeding period, nest survival is one of the most important drivers
of population growth for many wetland bird species and rangelands often provide
quality nesting cover. Throughout spring and fall, rangeland wetlands provide key
forage resources that support energetic demands needed for migration. In some areas,
stock ponds developed for livestock water provide migration stopover and wintering
habitat, especially in times of water scarcity. In the Intermountain West, drought
combined with water demands from agriculture and human population growth are
likely headed to an ecological tipping point for wetland birds and their habitat in
the region. In the Prairie Pothole Region, conversion of rangeland and draining of
wetlands for increased crop production remains a significant conservation issue for
wetland birds and other wildlife. In landscapes dominated by agricultural produc-
tion, rangelands provide some of the highest value ecosystem services, including
water quality and wetland function. Recent research has shown livestock grazing, if
managed properly, is compatible and at times beneficial to wetland bird habitat needs.
Either directly, or indirectly, wetland bird populations and their habitat needs are
supported by healthy rangelands. In the future, rangeland and wetland bird managers
will benefit from increased collaboration to aid in meeting ultimate conservation
objectives.
Keywords Conservation ·Livestock grazing ·Management ·Rangeland ·
Shorebirds ·Waterbirds ·Wat erfo wl ·Wetland birds
13.1 Introduction
Rangeland systems and the wetland birds using them vary across western North
America. This chapter addresses three groups of birds dependent on wetlands: water-
fowl, shorebirds, and waterbirds. Many wetland bird conservation plans recognize
the significant influence rangelands have on associated wetlands, including the North
American Waterfowl Management Plan (NAWMP 2018), U.S. Shorebird Conserva-
tion Plan (Brown et al. 2001), Canadian Shorebird Conservation Plan (Donaldson
et al. 2000), and North American Waterbird Conservation Plan (Kushlan et al. 2002).
Wetland birds typically exhibit large-scale mobility, including seasonal migration
across North America capitalizing on ecoregional resources to meet their annual cycle
needs. Wetland birds breeding in northern latitudes take advantage of primary produc-
tivity associated with extended summer daylight but as winter nears, they seek
resources at more southerly latitudes. Wetland bird migrations have heralded seasonal
change for societies over human history. Connected wetland networks sustain migra-
tions by providing rest and food resources and have demographic consequences for
populations.
Within seasonal home-ranges, wetlands birds can be highly mobile, and a single
wetland or wetland type can rarely meet daily, seasonal, or annual needs. Seasonal
wetlands tend to have high biological productivity, whereas wetlands with stable
water levels typically have reduced biological productivity. Because wetlands are
13 Waterfowl and Wetland Birds 419
dynamic, their availability and quality as habitat can be highly variable. Conse-
quently, wetland birds generally select landscapes with a diversity of wetlands
to maximize resources. Diversity within a complex of wetlands is a key strategy
for resource managers throughout North America (Baldassarre and Bolen 2006).
Wetland bird conservation has been coordinated across migration corridors (i.e.,
flyways) and regions. Rangelands cover significant areas of the Great Plains and
the West (Fig. 13.1; Table 13.1 [avian scientific names presented]). Throughout this
chapter ecoregional terminology is used consistent with wetland bird conservation
and management plans.
Previous reviews have provided important information on the ecology and
management of waterfowl (e.g., Smith et al. 1989;Battetal.
1992; Baldassarre and
Bolen 2006; Baldassarre 2014), shorebirds (Helmers 1992; Iglecia and Winn 2021),
Fig. 13.1 Major wetland bird ecoregions of rangeland systems in North America
420 J. L. Vest et al.
Table 13.1 Major wetland bird ecoregions within central and western North America with
subregions and regions of western rangelands
Major Wetland Bird
Region
Rangeland
Subregion Region
Prairie Pothole Aspen Parkland Great Plains
Northern Mixed Grass
Prairie
Northern Great
Plains
Sandhills of
Nebraska
Sand Hills of Nebraska
Rainwater Basin Tallgrass and Southern
Mixed Grass Prairie
Oaks and Prairie
High Plains & Playas Southern Mixed Grass
and Shortgrass Prairie
Gulf Coast Gulf Coast Prairies and
Marshes
Gulf Coastal Prairie
Intermountain West Rocky Mountains Rocky Mountains
Sierra Nevada
Mountains
Sierra Nevada Mountains
Cascade Mountains Western Deserts, Grasslands, Shrublands, and
Woodlands
Great Basin
Columbia Plateau
Colorado Plateau
Mojave Desert
Sonoran Desert
Chihuahuan Desert
Central Valley of
California
California Central
Valley
and other waterbirds (Beyersbergen et al. 2004; Ivey and Herziger 2006). Avail-
able research addressing rangeland management has focused on waterfowl, with less
empirical information for shorebirds and waterbirds. Therefore, this chapter relies
heavily on science addressing waterfowl and rangeland relationships. We provide
overviews of life history, regional variation, and population dynamics of wetland
birds that may be influenced by rangeland management and conservation.
13 Waterfowl and Wetland Birds 421
13.2 Wetland Systems
Wetlands occupy a relatively small footprint in many rangelands. However, wetlands,
riparian systems, and mesic habitats are often vital to the productivity, function,
and biodiversity of rangeland systems (Johnson 2019, Chap. 7). Wetlands provide
substantial ecosystem services and structure biological communities well beyond
their immediate footprint (Mitsch and Gosselink 2015; Donnelly et al. 2016; Johnson
2019). Wetlands are transitional areas with characteristics of both aquatic and terres-
trial ecosystems in addition to their own unique ecological conditions. Wetlands
typically occur where groundwater is at or near the surface or land is covered by
collection of water through runoff of surface water within a watershed (Cowardin
et al. 1979; Mitsch and Gosselink 2015). Wetlands have dynamic hydrology resulting
in conditions ranging from near-terrestrial to fully aquatic. Availability of habitat can
vary temporally and is subject to variation in response to climate patterns. Identi-
fying jurisdictional (i.e., subject to legal authority) wetlands includes a combination
of key factors: (1) presence of shallow water or moist soil for 14–21 days during
the growing season, (2) water-adapted plants (i.e., hydrophytic vegetation), and (3)
hydric soils influenced by anaerobic conditions of saturation (Cowardin et al. 1979;
Weller 1999). Not all wetlands are considered jurisdictional or subject to legal protec-
tions. For example, some wetlands in more arid environments are ephemeral, in some
cases inundated only once or twice over years.
Hydrology and water budget determine wetland type and associated ecolog-
ical processes (Mitsch and Gosselink 2015). Wetlands are commonly classified by
hydroperiod (Cowardin et al. 1979; Table 13.2). Hydrologic conditions such as water
depth, flow patterns, and flood frequency and duration (i.e., hydroperiod) influence
abiotic and biotic components. The hydroperiod is determined by water inflows and
outflows. Hydroperiod, largely dictates resource availability for wetland birds, other
wildlife, and livestock. For example, recharge wetlands are solely dependent upon
surface runoff linking hydroperiod to precipitation patterns. Conversely, discharge
wetlands have hydroperiods based on groundwater. Hydroperiod is more dynamic
in recharge versus discharge wetlands. Small hydrologic fluctuations can lead to
significant changes in plant and animal composition (Mitsch and Gosselink 2015).
Wetlands referenced in this chapter are either palustrine (i.e., marshy fresh or inland
saline waters or vegetated margins of large water bodies; Cowardin et al. 1979), or
lacustrine wetlands (i.e., relatively shallow, open, freshwater lakes or their sparsely
vegetated margins).
Wetland bird use tends to vary by water depth, vegetation characteristics, and
size (Laubhan and Gammonley 2000; Weller 1999;Maetal.
2010). There are five
general types of wetland plant associations: submerged plants, floating-leaved plants,
emergent plants, moist-soil plants, and woody plants. Submerged aquatic vegeta-
tion (SAV) communities provide important food sources for wetland birds—espe-
cially waterfowl—through their seeds, tubers, and leafy materials as well as asso-
ciated aquatic macroinvertebrates. Light penetration and turbidity affect subsurface
photosynthesis and influence establishment and productivity of SAV. Floating-leaved
422 J. L. Vest et al.
Table 13.2 Definitions of inland wetland hydroperiods
Permanently flooded—flooded throughout the year in all years
Intermittently exposed—flooded throughout the year except in years of extreme drought
Semipermanently flooded—flooded during the growing season in most years
Seasonally flooded—flooded for extended periods during the growing season, but usually no
surface water by end of the growing season
Saturated—substrate is saturated for extended periods during the growing season, but standing
water is rarely present
Temporarily flooded—flooded for brief periods during the growing season but the water table is
otherwise well below surface
Intermittently flooded—surface is usually exposed with surface water present for variable
periods without detectable seasonal pattern
Source Cowardin et al. (1979)
communities include both rooted and free-floating aquatic plants and provide little
value to most wetland birds. Function and productivity within rangeland wetlands
is primarily provided by emergent plants. These plants range from dense, robust
emergents such as cattail (Typha spp.), and bulrushes (Scirpus spp.), to relatively
shorter emergents with varying flood tolerances including sedges (Carex spp.), rushes
(Juncus spp.), spike-rushes (Eleocharis spp.), and water-tolerant grasses such as
cordgrass (Spartina spp.), panic grasses (Panicum spp.), and whitetop (Scholochloa
festucacea). Many emergent wetland species can be common livestock forages (Kirby
et al. 2002). Moist-soil plants include annuals or perennials that germinate following
drying events on exposed mudflats and provide abundant food via seeds and aquatic
invertebrates (Fredrickson and Taylor 1982; Haukos and Smith 1993; Anderson and
Smith 2000). Common moist-soil plants in western rangelands include smartweeds
(Polygnum spp.), barnyardgrass (Echinochloa crus-galli), spike-rushes, curly dock
(Rumex crispus), goosefoots and Lamb’s quarters (Chenopodium spp.), and alkali
bulrush (Scheonoplectus maritumus; Kadlec and Smith 1989; Haukos and Smith
1993; Dugger et al. 2007). Management of moist-soil habitats has been extensively
applied to wetland complexes providing forage for waterfowl (Fredrickson and Taylor
1982; Baldassarre and Bolen 2006). Periodic drying can temporarily reduce wetland
bird use, but is essential for cycling nutrients, succession of plant communities, and
maintaining productivity (Harris and Marshall 1963; Murkin et al. 1997).
13.2.1 Flyway Wetlands
Flyways including Atlantic, Central, Mississippi, and Pacific are useful constructs
for the administration of migratory bird management (Anderson et al. 2018; Roberts
et al. 2023), with rangelands primarily overlapping the Central and Pacific Flyways.
The Central Flyway includes prairie potholes, playas, and coastal marshes. Central
13 Waterfowl and Wetland Birds 423
Mixed-Grass Prairie and Tallgrass Prairie regions in the southern Central Flyway
provide key wetland habitats during migration (Smith et al. 1989;DU 2021; Hagy
et al. in review). Millions of pothole wetlands occur in the Northern Mixed-Grass
Prairie, northwestern Tallgrass Prairie, and Aspen Parklands within the Prairie
Pothole Region (PPR; Fig. 13.1). High wetland density with associated grasslands
makes the PPR unique and ecologically important in North America, and globally, for
breeding and migrating wetland birds (Baldassarre and Bolen 2006; Niemuth et al.
2010). The PPR is known as the “Duck Factory” producing between half to two-thirds
of all ducks in North America (Smith et al. 1964;Battetal. 1989; Baldassarre and
Bolen 2006) along with important water and forage resources for livestock (Johnson
2019). Playas are shallow, ephemeral, recharge wetlands abundant on the High Plains
of Central and Southern Shortgrass and Mixed-Grass prairies. Playas’ hydroperiods
are highly variable and inundation can range from days to years. Playas are drivers
of biodiversity in the region and the primary source of Ogallala Aquifer recharge
(Haukos and Smith 1994; Smith et al. 2012; Gitz and Brauer 2016). Millions of
wetland birds use playas during migration and winter (Haukos and Smith 1994; Moon
and Haukos 2008; Smith et al. 2012). Coastal marshes, tidal freshwater swamps, and
adjacent lagoons are defining features of the Gulf Coast. Freshwater and brackish
marshes generally support the most valuable habitats for wetland birds, particularly
waterfowl (Chabreck et al. 1989;Davis
2012). Coastal wetlands of Louisiana and
Texas are wintering grounds for millions of wetland birds (Baldassarre and Bolen
2006; Vermillion 2012; Henkel and Taylor 2015).
The Pacific Flyway includes the Intermountain West with a variety of wetlands
comprising < 10% of the area (McKinstry et al. 2004; Donnelly and Vest 2012). Many
seasonal wetlands have been converted to irrigated pastures and hay meadows for
production agriculture (McKinstry et al. 2004) and water management is generally
complex and controversial (Downard and Endter-Wada 2013; Donnelly et al. 2020;
Lovvorn and Crozier 2022). Within the region, wetlands are critical to sustaining
wetland birds, other wildlife, and agricultural-based economies (Sketch et al. 2020;
Donnelly et al. 2021, 2022; King et al. 2021). For most wildlife species, wetlands
are part of their annual life cycle (McKinstry et al. 2004). The region provides
migration, breeding, and wintering habitats for > 10 million wetland birds (Donnelly
and Vest 2012; IWJV 2013). Due t o precipitation patterns, wetlands experience high
annual variability in availability and productivity maintaining a network of functional
wetlands is critical to wetland bird conservation (Haig et al. 1998;Mackelletal.
2021;
Donnelly et al. 2020, 2021, 2022).
13.3 Life History, Annual Cycle, and Population Dynamics
The diverse taxa comprising wetland birds span a continuum of life-history strate-
gies that prioritize different fitness components (e.g., fecundity versus survival).
However, management occurs primarily at population levels and key vital rates that
shape population dynamics allow for some generalizations (Koons et al. 2014). Life
424 J. L. Vest et al.
histories vary from short-lived and high reproductive rates (i.e., more R-selected) to
long-lived and lower reproductive (i.e., more K-selected) strategies (Stearns 1992).
Accordingly, adult survival will have more influence on population growth rate for
species with moderate-to-long generation times, like geese, compared to species with
faster life histories, like teal, where reproductive success is more impactful (Koons
et al. 2014). Overall, both reproduction and survival of wetland bird populations are
influenced by environmental and habitat conditions. Sustaining functional wetland
networks, especially within rangelands, across flyways provides resiliency against
environmental stressors for wetland bird populations (Albanese and Haukos 2017;
Haig et al. 2019; Donnelly et al. 2020).
13.3.1 Nest and Female Survival
Nest survival, the probability that 1 egg hatches, is one of the primary drivers of
duck population growth rate and often the focus of management (Hoekman et al.
2002; Reynolds et al. 2006). Duck population growth rates can also be sensitive to
adult female survival with increased predation risk for nesting females (Hoekman
et al. 2002). Nest survival is generally higher for larger species like geese and swans
averaging 70%, whereas ducks average 15–20% (Hoekman et al. 2002; Baldassarre
and Bolen 2006; Baldassarre 2014). Clutch sizes range from 4 to 6 eggs for geese and
swans and 8–12 eggs for ducks, whereas shorebirds typically lay 4 eggs and some
other waterbird clutches may only have 1 egg (e.g., sandhill cranes). Waterfowl
and shorebirds that commonly nest in rangelands tend to be solitary nesters, but
semi-colonial behavior may occur where nest densities are high (e.g., islands). Nest
initiation starts in mid-April for early nesters like mallards and northern pintail, to
late June for late nesters like gadwall in high elevation systems (Baldassarre 2014).
Growing season interacts with environmental conditions dictating nesting phenology
and the propensity for renesting (Baldassarre 2014; Raquel et al. 2016).
Some wetland birds are generalists (e.g., mallards) that will nest in uplands, emer-
gent vegetation in wetland margins, artificial nest structures, or woody vegetation
along riparian areas (Baldassarre 2014). Others, like inland populations of snowy
plovers, nest exclusively in specialized habitat (e.g., unvegetated shorelines and
sandbars; Anteau et al. 2012). Agricultural lands can become ecological traps, such
as when northern pintail select cropland resulting in low nest survival (Buderman
et al. 2020). Waterfowl nesting habitat has three broad categories: (1) uplands
including grasslands, shrublands, and agriculture lands, (2) overwater vegetation
such as cattails and bulrushes or man-made platforms, and (3) cavities in trees or
nest boxes.
Ducks select nesting cover based on species, local conditions, and availability. For
example, mallards tend to select denser cover whereas northern pintail typically select
shorter, less dense vegetation (Baldassarre 2014). Proximity to wetlands is important
13 Waterfowl and Wetland Birds 425
for upland nesting ducks but varies by species. Blue-winged teal have relatively small
home ranges and nest closer to wetlands. Mallard and northern pintail can nest > 2 km
from a wetland (Reynolds et al. 2006). Lesser scaup have limited mobility in uplands
and nest very close to wetlands. When uplands lack cover, upland nesters tend to
seek cover in dry wetlands at the emergent fringe (Lovvorn and Crozier 2022).
Most adult female mortality (i.e., 65–80%) of ducks occurs during the breeding
season where nesting females are vulnerable to predators (Hoekman et al. 2002;
Arnold et al. 2012). Providing quality nesting habitat helps increase both nest and
female survival (Reynolds et al. 1995; Arnold et al. 2012). At the population level,
nest survival is impacted by large-scale environmental factors and local nest-site
characteristics; vegetation structure is more important than composition (Ringelman
et al. 2018; Sherfy et al. 2018; Bortolotti et al. 2022). Nest survival generally increases
with larger patch size and more perennial vegetation (Baldassarre and Bolen 2006;
Bortolotti et al. 2022). The relationship of habitat and nest survival is complex, varies
regionally, and difficult to differentiate among confounding factors like landscape
characteristics, environmental changes, and predator communities (Clark and Nudds
1991;Hornetal. 2005;Walkeretal. 2013a; Ringelman et al. 2018; Bortolotti et al.
2022; Pearse et al. 2022). Rangelands, with associated wetlands, generally provide
extensive areas of perennial cover and reliably have high duck nest survival (Stephens
et al. 2005;Walkeretal.
2013a; Bortolotti et al. 2022). Increased nest survival
in rangelands, compared to cropland landscapes, is likely due to reduced predator
efficiency within large intact habitat and/or lower predator densities (Ball et al. 1995;
Phillips et al. 2003;Hornetal.
2005). Large areas of intact rangelands may also
support a greater abundance and diversity of other prey, reducing predation pressure
on duck nests (Ackerman 2002). Although not fully understood at continental and
population scales, intact rangelands are likely important in sustaining waterfowl in
North America due to the potential for high nesting productivity (Higgins et al. 2002;
PHJV 2021; PPJV 2017).
Nearly all shorebirds are ground nesters, but habitats and breeding behavior
vary widely by species (Iglecia and Winn 2021). Before Euro-American settlement,
breeding shorebirds in the Great Plains specialized in exploiting the diverse grass-
land mosaics left by bison (Bison bison) and fire (Eldridge 1992). Shorebird breeding
habitat includes unvegetated beaches and salt/alkali flats to moderately tall and dense
grasslands (Eldridge 1992; Iglecia and Winn 2021). Long-billed curlew, marbled
godwit, willet, killdeer, and mountain plover all nest and forage in short (< 15 cm)
grassland vegetation often far from wetlands. Wilson’s phalarope and upland sand-
piper typically use taller (10–30 cm) and denser vegetation (Eldridge 1992). For
species that rely on wetland invertebrates, proximity to wetlands is important when
selecting nesting habitat (e.g., Wilson’s phalarope, American avocet, piping plover,
snowy plover, marbled godwit, willet; Eldridge 1992; Specht et al. 2020). Drivers
of shorebird nest survival may be similar to those of waterfowl due to shared nest
predators (Specht et al. 2020).
Diving ducks and swans (Table 13.3) primarily build overwater nests from emer-
gent vegetation such as bulrush, cattail, and sedges (Baldassarre 2014). These over-
water nesters often have limited available nesting cover and are generally associated
426 J. L. Vest et al.
with semi-permanent and permanent wetlands (Baldassarre 2014). Overwater nests
are more protected, and survival tends to be higher than upland nests, although
predation rates can increase with decreasing water levels (Baldassarre and Bolen
2006). Across the PPR, mallards nest in emergent wetland vegetation and experi-
ence relatively higher nest survival rates compared to upland nests (Baldassarre and
Bolen 2006; Baldassarre 2014). Other waterbird species also nest over water either
in dense emergent vegetation (e.g., sandhill crane) or on floating mats of vegetation
(e.g., grebes). Some waterbirds nest on islands (e.g., pelicans) and in trees (e.g.,
herons; Beyersbergen et al. 2004).
13.3.2 Juvenile Survival
Juvenile survival can also strongly influence population growth rate for wetland birds,
especially dabbling ducks (Hoekman et al. 2002). Chick survival is lowest within the
first two weeks post-hatch. Small size and lack of thermoregulation during this time
makes chicks vulnerable to exposure (Bloom et al. 2012; Iglecia and Winn 2021)
and a wide range of predators (Sargeant and Raveling 1992; Baldassarre and Bolen
2006). Females can move their brood long distances to find quality habitat, which
includes abundant invertebrates for food and security cover. Brood occurrence and
survival has been shown to correlate with the availability of perennial herbaceous
vegetation and wetland area (Krapu et al. 2000;Walkeretal.
2013b). Rangelands
with abundant and diverse wetlands in both size and hydroperiod are essential to
sustaining wetland bird populations in North America (Helmers 1992; Beyersbergen
et al. 2004;Walkeretal. 2013b).
Waterbird species have chicks that range from precocial to altricial. Sandhill
crane colts leave the nest directly after hatching whereas loons, grebes, most rails,
and coots rely on parental feeding at the nest for several days. Gull and tern chicks
may quickly leave the nest but remain close to the nest site for several days. Ibis,
pelicans, cormorants, and herons feed chicks in nests until mobility develops, which
varies from 2 to 11 weeks (Weller 1999). Sandhill crane parents feed young for
the first few weeks and colt mortality can be high at this time (Gerber et al. 2015).
Sandhill cranes have the lowest recruitment of hunted avian species in North America
(Drewien et al. 1995).
13.3.3 Post-breeding Survival and Migration
Post-breeding is bracketed by the reproductive and fall migration periods (Hohman
et al. 1992). Most waterfowl molt flight feathers rendering birds flightless for 3–
5 weeks (Baldassarre and Bolen 2006; Fox et al. 2014). Post-breeding waterfowl are
vulnerable to habitat changes (e.g., drying, or de-watering of wetlands) that increase
predation risk or limit access to food resources (Hohman et al. 1992). Molting has
13 Waterfowl and Wetland Birds 427
Table 13.3 Common waterfowl species in North America and their primary occurrence in
rangelands, population size, trend, and conservation or management status in the United States
Common name Scientific
name
Rangeland
overlapa
Population
Estimate:
LTA-T S A b
Estimate:
PIF (US,
CA)c
Trend
(%/
yr)d
Statuse
Northern PintailDA Anas acuta B, NB 3,866,300 3,200,000 1.2 BMC
Green-winged
TealDA
Anas crecca B, NB 2,179,200 3,900,000 1.7 BMC
Mexican Duck*DA Anas diazi B, NB 55,000
Mottled DuckDA Anas fulvigula B, NB 180,000 2.5 BMC,
BCC
MallardDA Anas
platyrhynchos
B, NB 7,930,400 11,000,000 0.7 BMC
American Black
DuckDA
Anas rubripes 700,000 1BMC
Muscovy DuckDA Cairina
moschata
B, NB
American
WigeonDA
Mareca
americana
B, NB 2,618,100 2,700,000 0.2 BMC
GadwallDA Mareca
strepera
B, NB 2,057,300 3,400,000 2.4 BMC
Northern
ShovelerDA
Spatula
clypeata
B, NB 2,643,900 4,400,000 2.3 BMC
Cinnamon TealDA Spatula
cyanoptera
B, NB 440,000 2.2 BMC,
BCC
Blue-winged
TealDA
Spatula
discors
B, NB 5,127,600 7,800,000 1.5 BMC
Wood DuckDC Aix sponsa B, NB 4,600,000 1.8 BMC
Fulvous
Whistling-DuckWD
Dendrocygna
bicolor
B, NB 120,000
Black-bellied
Whistling-DuckWD
Dendrocygna
autumnalis
B, NB 7.4
Lesser ScaupDI Aythya affinis B, NB 4,947,300c 3,700,000 1.2 BMC
RedheadDI Aythya
americana
B, NB 732,700 1,200,000 1.6 BMC
Ring-necked
DuckDI
Aythya
collaris
B, NB 2,000,000 3.3 BMC
Greater ScaupDI Aythya marila NB c720,000 1.5 BMC
CanvasbackDI Aythya
valisineria
B, NB 591,300 690,000 0.8 BMC
(continued)
428 J. L. Vest et al.
Table 13.3 (continued)
Common name Scientific
name
Rangeland
overlapa
Population
Estimate:
LTA-T S A b
Estimate:
PIF (US,
CA)c
Trend
(%/
yr)d
Statuse
Ruddy DuckDO Oxyura
jamaicensis
B, NB 1,300,000 1.7 BMC
BuffleheadSM Bucephala
albeola
b, NB 1,300,000 3.5
Common
GoldeneyeSM
Bucephala
clangula
B, NB 1,200,000 0.7 BMC
Barrow’s
GoldeneyeSM
Bucephala
islandica
B, NB 180,000 1.4
Long-tailed
DuckSM
Clangula
hyemalis
1,000,000 4.8 BMC
Harlequin DuckSM Histrionicus
histrionicus
b170,000 0.5 BMC
Hooded
MerganserSM
Lophodytes
cucullatus
B, NB 1,100,000 4.7
Black ScoterSM Melanitta
americana
500,000 2.3 BMC
White-winged
ScoterSM
Melanitta
deglandi
b, nb 400,000 0.6 BMC
Surf ScoterSM Melanitta
perspicillata
nb 470,000 0.2 BMC
Common
MerganserSM
Mergus
merganser
B, NB 1,200,000 0.4
Red-breasted
MerganserSM
Mergus
serrator
NB 400,000 3.3
Steller’s EiderSM Polysticta
stelleri
660 4ESA
Spectacled EiderSM Somateria
fischeri
20,000 ESA
Common EiderSM Somateria
mollissima
750,000 1BMC
King EiderSM Somateria
spectabilis
600,000 6.4 BMC
Greater
White-fronted
GooseGA
Anser
albifrons
b, NB 4,300,000 4.9 BMC
Snow GooseGA Anser
caerulescens
NB 15,000,000 6.1 BMC
Emperor GooseGA Anser
canagicus
98,000 0.4 BMC,
BCC
Ross’s GooseGA Anser rossii NB 1,600,000 11.7 BMC
(continued)
13 Waterfowl and Wetland Birds 429
Table 13.3 (continued)
Common name Scientific
name
Rangeland
overlapa
Population
Estimate:
LTA-T S A b
Estimate:
PIF (US,
CA)c
Trend
(%/
yr)d
Statuse
BrantGA Branta
bernicla
nb 340,000 0.2 BMC,
BCC
Canada GooseGA Branta
canadensis
B, NB 7,500,000 10.3 BMC
Cackling GooseGA Branta
hutchinsii
nb 4,100,000 6.2 BMC
Trumpeter SwanSC Cygnus
buccinator
B, NB 63,000 6.6 BMC
Tundra SwanSC Cygnus
columbianus
NB 190,000 0BMC
Mute SwanSC Cygnus olor 31,000 3.6
aSpecies occurrence in central and western rangeland regions of North America and annual cycle
importance. B (b) =breeding, NB (nb) =non-breeding; capital letters indicate common or abundant,
lowercase letters indicate uncommon, rare, or minimal rangeland overlap
bPopulation estimate based on the long-term average (LTA[1955–2022]) from the traditional survey
area (TSA) of the Breeding Waterfowl and Habitat Survey conducted by U.S. Fish & Wildlife Service
and Canadian Wildlife Service (USFWS 2022). Population estimates for lesser and greater scaup
combined
cPopulation estimate from Partners in Flight (2021;PIF)
dPopulation trend (% change per year) from Partners in Flight (2021)
eConservation and management status identified by U.S. Fish and Wildlife Service. BMC = birds
of management concern, BCC = birds of conservation concern, ESA = threatened or endangered
status under the Endangered Species Act
Guild and (Tribe): DA Dabbler (Anatani), DC Dabbler (Cairinini), WD Whistling Duck (Dendro-
cygnini), DIDiver (Aythini), DO Diver (Oxyurini), SMSea Duck (Mergini), GA Goose (Anserini),
SCSwan (Cygnini)
high nutrient demands like protein-rich foods (e.g., aquatic insects). Post-breeding
waterfowl select habitats that lower predation risk and offer abundant food resources
(Fox et al. 2014). Semi-permanent or permanent wetlands with emergent vegetation
and open water are often selected post-breeding (Hohman et al. 1992; Fleskes et al.
2010). Such habitats also offer key migratory stopover areas when energetic demands
increase and wetland bird diets transition to more carbohydrate-rich food sources
such as wetland plant seeds, tubers, rhizomes, and agricultural grains (Baldasarre
and Bolen 2006; Donnelly et al. 2021). Shorebirds will consume small amounts of
plant material, but they primarily consume invertebrates for energy and some species
may double their body mass prior to migration (Baker et al. 2014; Iglecia and Winn
2021).
Across rangelands, wetland availability is lowest during late summer and early
fall (Johnson et al. 2010; Donnelly et al. 2019). Habitat availability is typically
lowest during the post-breeding period when birds have high nutrient demands. Low
430 J. L. Vest et al.
nutrient reserves may negatively affect autumn survival (Sedinger and Alisauskas
2014). Additionally, diseases such as botulism, avian cholera, and avian influenza
virus increase mortality risk, particularly for waterfowl, especially with decreased
wetland availability (Friend et al. 2001; Baldassarre and Bolen 2006; Kent et al.
2022). Changes in land and water use, often in combination with drought, decrease
wetland availability resulting in bird concentrations and recurring disease issues
(Fleskes et al. 2010; Donnelly et al. 2022; Kahara et al. 2021). Similar to the breeding
period, rangelands that provide wetland habitat during the post-breeding period are
vital to wetland birds (Johnson et al. 2010; Gerber et al. 2015; Kemink et al. 2021;
Donnelly et al. 2022).
Ideally, migration and wintering habitat provide key nutrients and energy (i.e.,
lipids) sources during migration and highlight the importance of available wetland
complexes (Moon and Haukos 2006, 2009; Davis et al. 2014; Yetter et al. 2018).
Selected food resources may change based on physiology and behavior and in
response to environmental conditions. Narrow migration windows may or may not
align with food availability. Donnelly et al. (2019) found that most seasonal wetlands
were available during spring migration, whereas 20% were available for fall migra-
tion. In winter, freezing and snow accumulation can decrease food (e.g., grains)
availability inhibiting migration.
Some wetland birds (e.g., waterfowl, coots, sandhill cranes and other rails) are
hunted during fall and winter. Hunters, through harvest reporting (e.g., band returns,
wing collections, surveys) and funding (e.g., duck stamp), have increased our under-
standing of population dynamics, movements, and conservation (Anderson et al.
2018). For example, adult female mallard survival during the non-breeding season
has little impact on population growth rates relative to the breeding period and males
have low natural mortality making them even more available for sustainable harvest
(Hoekman et al. 2002). Consequently, hunting harvest is the primary mortality cause
for male ducks (Hoekman et al. 2002; Riecke et al. 2022a). Female ducks gener-
ally experience lower harvest rates t han males (Riecke et al. 2022a, b). Waterfowl
harvest is carefully managed across flyways and represents one of the most successful
examples of adaptive management in the world (Nichols et al. 2019).
Non-breeding habitat conditions can have carry-over effects to breeding success
(Sedinger and Alisauskas 2014; Swift et al. 2020). Generally, birds in better nutri-
tional state (i.e., body condition) during winter and spring may arrive in breeding
areas earlier, nest earlier, and experience greater breeding success (Devries et al.
2008; Sedinger and Alisauskas 2014; Swift et al. 2020). Management that enhances
nutritive resources in non-breeding habitats can also increase vital rates (Davis et al.
2014; Stafford et al. 2014). More information is needed to better understand shore-
bird vital rates and population dynamics, along with the impacts of migration and
winter habitat. However, adult annual survival sustains populations for several arctic-
nesting shorebirds during migration (Weiser et al. 2020). Like other wetland birds,
wetland networks are critical (Albanese and Davis 2015). Wetlands within range-
lands generally have less functional impairment than in croplands and are critical
to wetland bird survival (Tsai et al. 2012; Collins et al. 2014; Albanese and Davis
2015; McCauley et al. 2015; Tangen et al. 2022).
13 Waterfowl and Wetland Birds 431
13.3.4 Spring Migration
Spring migration is another critical time for wetland birds and includes additional
energetic demands, like molt and courtship (Anteau et al. 2011; Stafford et al. 2014).
Survival is usually high (Moon and Haukos 2006; Osnas et al. 2021), and habitat avail-
ability remains important (Anteau and Afton 2011; Sedinger and Alisauskas 2014).
For many species, early arrival to breeding areas correlates with increased repro-
ductive success. Shallow flooded wetlands are often the first to thaw and provide
important food resources in spring. Overall, the timing, stop-over frequency, and
duration of spring migration is influenced by weather conditions, habitat availability
(i.e., food abundance), and initial body condition (Miller et al. 2005; Haukos et al.
2006; Stafford et al. 2014). Wetland networks are therefore needed to support migra-
tion survival and breeding success (Devries et al. 2008; Zarzycki 2017; Osnas et al.
2021).
13.4 Current Species Population Status and Monitoring
More than 200 million individuals of 280 species of wetland birds occur in North
America (PIF 2021). Over half of these species have seasonal distributions that
overlap with rangelands, comprising > 160 million wetland birds (Tables 13.4, 13.5
and 13.6). Wetland bird populations have increased between 1970 and 2017, primarily
from waterfow and geese, but other wetland birds have declined (Rosenberg et al.
2019).
13.4.1 Monitoring Programs
Large-scale programs have been developed to monitor population status. Several
are agency-led, particularly for hunted species, while some rely on citizen science
efforts. Since 1955, the U.S. Fish and Wildlife Service (USFWS) and Canadian
Wildlife Service (CWS) have conducted the Waterfowl Breeding Population and
Habitat Survey (WBPHS) to estimate breeding populations in Alaska, Canada, and
north-central United States (USFWS 2022). The WBPHS is used for estimates of
multiple waterfowl species populations and wetland abundance. The USFWS, in
coordination with state wildlife agencies, conducts an annual mid-winter waterfowl
survey within each flyway to index waterfowl populations (USFWS 2023). Large-
scale monitoring for shorebirds has been proposed, with implementation of some
periodic, regional surveys (Cavitt et al. 2014). Secretive marsh bird surveys have
been implemented in multiple regions (Johnson et al. 2009).
Publicly-sourced data collection has become increasingly important. The
Breeding Bird Survey (BBS) is the main source of avian population status in
North America and provides representative sampling of wetlands (Sauer et al. 2003;
432 J. L. Vest et al.
Table 13.4 Common shorebird species in North America and their primary occurrence in
rangelands, population size, trend, and conservation or management status in the United States
Common name Scientific name Rangeland
overlapa
Population
Estimate: PIF
(US, CA)b
Trend (%/
yr)c
Statusd
Piping PloverCCharadrius
melodus
B, NB 8400 1.9 ESA
Mountain
PloverC
Charadrius
montanus
B, NB 20,000 3.1 BMC, BCC
Snowy PloverCCharadrius
nivosus
B, NB 24,000 0.4 BMC, BCC
Semipalmated
PloverC
Charadrius
semipalmatus
NB 200,000 0.4
KilldeerCCharadrius
vociferus
B, NB 1,800,000 1
Wilson’s PloverC Charadrius
wilsonia
B8600 1.9 BMC, BCC
American
Golden-PloverC
Pluvialis
dominica
NB 500,000 1.9 BCC
Pacific
Golden-PloverC
Pluvialis fulva 43,000 1.7
Black-bellied
PloverC
Pluvialis
squatarola
NB 360,000 1.6
Black
OystercatcherH
Haematopus
bachmani
10,000 3.5 BMC, BCC
American
OystercatcherH
Haematopus
palliatus
B12,000 1BMC, BCC
Northern JacanaJ Jacana spinosa B, NB
Black-necked
StiltR
Himantopus
mexicanus
B, NB 180,000 2.4
American
AvocetR
Recurvirostra
americana
B, NB 450,000 0.5 BCC
Spotted
SandpiperS
Actitis
macularius
B, NB 660,000 1.4
Ruddy
TurnstoneS
Arenaria
interpres
NB 250,000 4.7 BCC
Black TurnstoneS Arenaria
melanocephala
NB 95,000 0.4 BCC
Upland
SandpiperS
Bartramia
longicauda
B, NB 750,000 0.5 BMC, BCC
SanderlingSCalidris alba NB 300,000 3.3
(continued)
13 Waterfowl and Wetland Birds 433
Table 13.4 (continued)
Common name Scientific name Rangeland
overlapa
Population
Estimate: PIF
(US, CA)b
Trend (%/
yr)c
Statusd
DunlinSCalidris alpina nb 1,500,000 2.7 BMC, BCC
Baird’s
SandpiperS
Calidris bairdii NB 280,000 1.3
Red KnotSCalidris canutus MB 140,000 5.7 BMC, BCC
White-rumped
SandpiperS
Calidris
fuscicollis
NB 1,700,000 1.3
Stilt SandpiperSCalidris
himantopus
NB 1,200,000 1.5
Purple
SandpiperS
Calidris
maritima
25,000 1.2 BMC, BCC
Western
SandpiperS
Calidris mauri NB 3,500,000 0.4
Pectoral
SandpiperS
Calidris
melanotos
1,500,000 2BCC
Least SandpiperS Calidris
minutilla
NB 700,000 0.2
Rock SandpiperS Calidris
ptilocnemis
140,000 2.8 BMC, BCC
Semipalmated
SandpiperS
Calidris pusilla NB 2,300,000 3.1 BMC, BCC
Buff-breasted
SandpiperS
Calidris
subruficollis
NB 56,000 1.8 BMC, BCC
SurfbirdSCalidris virgata NB 70,000 1.7
Wilson’s SnipeSGallinago
delicata
B, NB 2,000,000 0.3 BMC
Short-billed
DowitcherS
Limnodromus
griseus
NB 150,000 2.9 BMC, BCC
Long-billed
DowitcherS
Limnodromus
scolopaceus
NB 520,000 0.3
Marbled
GodwitS
Limosa fedoa B, NB 170,000 0.7 BCC
Hudsonian
GodwitS
Limosa
haemastica
NB 77,000 3.4 BMC, BCC
Bar-tailed
GodwitS
Limosa
lapponica
90,000 BMC, BCC
Long-billed
CurlewS
Numenius
americanus
B, NB 140,000 0BMC, BCC
(continued)
434 J. L. Vest et al.
Table 13.4 (continued)
Common name Scientific name Rangeland
overlapa
Population
Estimate: PIF
(US, CA)b
Trend (%/
yr)c
Statusd
Eskimo CurlewSNumenius
borealis
50 ESA
WhimbrelSNumenius
phaeopus
MNB 80,000 2.1 BMC
Bristle-thighed
CurlewS
Numenius
tahitiensis
10,000 BMC, BCC
Red PhalaropeSPhalaropus
fulicarius
1,600,000
Red-necked
PhalaropeS
Phalaropus
lobatus
NB 2,500,000
Wilson’s
PhalaropeS
Phalaropus
tricolor
B, NB 1,500,000 0.2
American
WoodcockS
Scolopax minor 3,500,000 0.8 BMC
Lesser
YellowlegsS
Tringa flavipes NB 660,000 2.8 BMC, BCC
Wandering
TattlerS
Tringa incana NB 16,000 4.3 BCC
Greater
YellowlegsS
Tringa
melanoleuca
NB 140,000 0.5
Will etSTringa
semipalmata
B, NB 250,000 0.6 BCC
Solitary
SandpiperS
Tringa solitaria B, NB 190,000 0.7 BMC, BCC
aSpecies occurrence in central and western rangeland regions of North America and annual cycle
importance. B (b) =breeding, NB (nb) =non-breeding; capital letters indicate common or abundant,
lowercase letters indicate uncommon, rare, or minimal rangeland overlap
bPopulation estimate from Partners in Flight (2021;PIF)
cPopulation trend (% change per year) from Partners in Flight (2021)
dConservation and management status identified by U.S. Fish and Wildlife Service. BMC = birds
of management concern, BCC = birds of conservation concern, ESA = threatened or endangered
status under the Endangered Species Act
Family names: CCharadriidae, HHaematopodidae, JJacanidae, RRecurvirostridae, SScolopacidae
Niemuth et al. 2007; Veech et al. 2017). The BBS may not suffice for all species and
formal evaluations are needed concerning wetland birds (Hudson et al. 2017). For
many wetland birds, BBS data could be more useful if wetland habitat availability
were included (Niemuth and Solberg 2003; Niemuth et al. 2009). For other species,
targeted monitoring may be necessary. eBird, a global online database launched in
2002 (Cornell Lab of Ornithology), compiles public records of avian species detec-
tions with location and date (Sullivan et al. 2014). Biologists have used eBird data to
13 Waterfowl and Wetland Birds 435
Table 13.5 Common waterbird species in North America and their primary occurrence in
rangelands, population size, trend, and conservation or management status in the United States
Common name Scientific name Rangeland
overlapa
Population
Estimate:
PIF (US,
CA)b
Trend
(%/yr)c
Statusf
Black TernCL Chlidonias niger B, NB 2,300,000 1.9 BMC, BCC
Bonaparte’s
GullCL
Chroicocephalus
philadelphia
NB 790,000 1.9
Gull-billed TernCL Gelochelidon
nilotica
B, NB 8,000 1.3 BMC, BCC
Caspian TernCL Hydroprogne
caspia
B, NB 78,000 0.9
Herring GullCL Larus argentatus B, NB 2,900,000 3.9
California GullCL Larus californicus B, NB 1,100,000 1.6 BCC
Ring-billed GullCL Larus
delawarensis
B, NB 3,700,000 1.5
Glaucous-winged
GullCL
Larus glaucescens B, NB 440,000 0.6
Iceland GullCL Larus glaucoides NB 84,000
Heermann’s
GullCL
Larus heermanni NB BCC
Yellow-footed
GullCL
Larus livens B, NB BCC
Laughing GullCL Leucophaeus
atricilla
B, NB 680,000 2
Franklin’s GullCL Leucophaeus
pipixcan
B, NB 2,300,000 1.9 BCC
Sooty TernCL Onychoprion
fuscatus
B
Black SkimmerCL Rynchops niger B, NB 60,000 3.1 BCC
Forster’s TernCL Sterna forsteri B, NB 130,000 1.4 BCC
Common TernCL Sterna hirundo B470,000 2.1
Least TernCL Sternula
antillarum
B52,000 3.6 ESA
Elegant TernCL Thalasseus
elegans
B, NB BCC
Royal TernCL Thalasseus
maximus
B35,000 0.5
Sandwich TernCL Thalasseus
sandvicensis
B, NB 94,000 1.4 BMC, BCC
Wood StorkCC Mycteria
americana
NB 16,000 1.6 ESA (SE pop)
(continued)
436 J. L. Vest et al.
Table 13.5 (continued)
Common name Scientific name Rangeland
overlapa
Population
Estimate:
PIF (US,
CA)b
Trend
(%/yr)c
Statusf
Common LoonGA Gavia immer B, NB 1,100,000 0.8
Sandhill CraneGU Antigone
canadensis
B, NB 500,000 5.1 BMC
Whooping
CraneGU
Grus americana B, NB 370 ESA
Yellow RailGR Coturnicops
noveboracensis
B, NB 12,000 BMC, BCC
American CootGR Fulica americana B, NB 5,500,000 0.2 BMC
Common
GallinuleGR
Gallinula galeata B, NB 500,000 1.1
Black RailGR Laterallus
jamaicensis
B, NB BMC
Purple GallinuleGR Porphyrio
martinicus
B, NB 20,000 1.9
SoraGR Porzana carolina B, NB 4,400,000 0.5 BMC
Clapper RailGR Rallus crepitans B, NB 170,000 0.8 BMC
King RailGR Rallus elegans B, NB 63,000 4.5 BMC, BCC
Virginia RailGR Rallus limicola B, NB 230,000 1.5 BMC
Great EgretPA Ardea alba B, NB 710,000 2.5
Great Blue
HeronPA
Ardea herodias B, NB 620,000 0.7
American
BitternPA
Botaurus
lentiginosus
B, NB 2,500,000 0.6 BMC
Cattle EgretPA Bubulcus ibis B, NB 2,800,000 1.4
Green HeronPA Butorides
virescens
B, NB 770,000 1.9
Little Blue
HeronPA
Egretta caerulea B, NB 270,000 1.4 BMC, BCC
Reddish EgretPA Egretta rufescens B, NB 2400 1.3 BMC
Snowy EgretPA Egretta thula B, NB 220,000 2.2
Tricolored
HeronPA
Egretta tricolor B, NB 58,000 0.5
Least BitternPA Ixobrychus exilis B, NB 130,000 0.7 BMC
Yellow-crowned
Night-HeronPA
Nyctanassa
violacea
B, NB 130,000 0.5
(continued)
13 Waterfowl and Wetland Birds 437
Table 13.5 (continued)
Common name Scientific name Rangeland
overlapa
Population
Estimate:
PIF (US,
CA)b
Trend
(%/yr)c
Statusf
Black-crowned
Night-HeronPA
Nycticorax
nycticorax
B, NB 420,000 0.4 BMC
American White
PelicanPP
Pelecanus
erythrorhynchos
B, NB 410,000 6.3 BCC
Brown PelicanPP Pelecanus
occidentalis
B, NB 100,000 3.6
White IbisPT Eudocimus albus B, NB 1,200,000 4.4
Roseate
SpoonbillPT
Platalea ajaja B, NB 11,000 7.3
White-faced IbisPT Plegadis chihi B, NB 1,300,000 3.5
Glossy IbisPT Plegadis
falcinellus
36,000 5.3
Clark’s GrebePO Aechmophorus
clarkii
B, NB 72,000 2.9 BCC
Wes t ern G reb ePO Aechmophorus
occidentalis
B, NB 990,000 3.6 BMC, BCC
Horned GrebePO Podiceps auritus B, NB 250,000 1.4 BMC
Red-necked
GrebePO
Podiceps
grisegena
B740,000 0.9
Eared GrebePO Podiceps
nigricollis
B, NB 2,000,000 1.1 BMC
Pied-billed
GrebePO
Podilymbus
podiceps
B, NB 1,100,000 0.9
Least GrebePO Tachybaptus
dominicus
B, NB
AnhingaSA Anhinga anhinga B, NB 27,000 1.6
Double-crested
CormorantSP
Phalacrocorax
auritus
B, NB 560,000 4BMC (OA)
Neotropic
CormorantSP
Phalacrocorax
brasilianus
B, NB 7.6
aSpecies occurrence in central and western rangeland regions of North America and annual cycle
importance. B (b) =breeding, NB (nb) =non-breeding; capital letters indicate common or abundant,
lowercase letters indicate uncommon, rare, or minimal rangeland overlap
bPopulation estimate from Partners in Flight (2021;PIF)
CPopulation trend (% change per year) Partners in Flight (2021)
dConservation and management status identified by U.S. Fish and Wildlife Service. BMC = birds
of management concern, BCC = birds of conservation concern, ESA = threatened or endangered
status under the Endangered Species Act 1973
Order and Family: CL Ciconiiformes Laridae, CC Ciconiiformes Ciconiidae, GAGaviiformes
Gaviidae, GUGruiformes Gruidae, GR Gruiformes Rallidae, PA Pelecaniformes Ardeidae,
PPPelecaniformes Pelecanidae, PT Pelecaniformes Threskiornithidae, POPodicipediformes
Podidipedidae, SASuliformes Anhingidae, SPSuliformes Phalacrocoracidae
438 J. L. Vest et al.
assess migration chronology, distribution, abundance, and population trends (Walker
and Taylor 2017; Horns et al. 2018; Fink et al. 2020).
International banding programs provide information on movements and demo-
graphics of many wetland bird populations. The U.S. Geological Survey Bird
Banding Laboratory distributes about one million aluminum leg bands to managers
and researchers in the U.S. and Canada each year and manages an archive of over
77 million banding records and 5 million band encounters (U.S. Geological Survey,
Bird Banding Laboratory 2020). Hunter participation in waterfowl band reporting
has been one of the longest and most significant information sources for waterfowl
research and conservation. Mark-recovery methods (Brownie et al. 1985; Williams
et al. 2002) are used to identify harvest distribution and associated breeding areas,
estimate harvest rates, and survival rates for species, age, and sex (Smith et al. 1989).
The USFWS and CWS conduct annual hunter surveys for hunted wetland birds
(Martin and Carney 1977; Cooch et al. 1978; Martin et al. 1979). Harvest estimates
can provide an index to population trends when other data sources are limited for some
species. Banding and harvest survey data can be combined to estimate population
abundance in some cases (Lincoln 1930; Alisauskas et al. 2009, 2014).
13.4.2 Waterfowl
As a group, there are fewer waterfowl species (n = 46) than either shorebird (n = 51)
or waterbirds (n = 62), but waterfowl populations are more abundant. Waterfowl have
also had more monitoring due to their gamebird status and associated socio-economic
values (Anderson et al. 2018). Indices of breeding ducks from the WBPHS have
fluctuated from lows of 25 million in the early 1960s and 1990s to nearly 50 million
in 2014–2015. Duck populations show a cyclical pattern over time influenced largely
by conditions in the PPR (Fig. 13.2; Baldassarre and Bolen 2006; USFWS 2022).
Duck species that rely primarily on rangeland wetlands tend to have small popu-
lations or be in decline. Cinnamon teal are widely distributed across western range-
lands. Mottled ducks occur primarily along the Gulf Coast. Cinnamon teal, mottled
duck, and Mexican duck are among the least studied species with lower abundance
and are identified as Species of Conservation Concern (Table 13.3; Baldassarre 2014).
Northern pintails have declined since the early 1970s and remain below popula-
tion objectives (NAWMP 2018; USFWS 2022). Rangeland conversion to row-crop
production, especially in the PPR, has contributed to pintail declines (Baldassarre
2014; Buderman et al. 2020). Lesser and greater scaup have had similar population
declines and status as pintails (NAWMP 2018; USFWS 2022). Most lesser scaup
breed in the Western Boreal Forest, but at least 25% breed in rangeland wetlands
where livestock grazing is a prominent land-use (Baldassarre 2014).
Goose and swan populations have generally increased since the 1970s, with
overabundance of some goose populations (USFWS BMC 2011; Baldassarre 2014;
USFWS 2022). The dramatic increase in snow geese is largely due to 4 factors:
(1) increased food availability due to crop-conversion and enhanced fertilizer-based
13 Waterfowl and Wetland Birds 439
Fig. 13.2 Total duck population change 1955–2022 from the Traditional Survey Area of the
Waterfowl Breeding Population and Habitat Survey (USFWS 2022)
yields, (2) establishment of staging and wintering areas on refuges, (3) declines in
harvest rates, and (4) climate change (Jefferies et al. 2003; Baldassarre 2014). Several
Canada goose populations breed extensively in rangelands (Baldassarre 2014). The
Hi-Line population in north-central Montana increased tenfold following their 1960s
reintroductions concomitant with reservoir and stock pond development (Nieman
et al. 2000; Baldassarre 2014).
13.4.3 Shorebirds
Shorebirds have experienced significant declines (i.e., 37%) since the 1970s
(Fig. 13.2; Rosenberg et al. 2019; Smith et al. 2023). Most shorebirds are considered
species of high conservation concern with 5 listed under the Endangered Species
Act (1973; Table 13.4; U.S. Shorebird Conservation Plan Partnership 2016). Over
80% of breeding shorebirds migrate to Mexico, Central, and South America (Iglecia
and Winn 2021). At least 16 species have significant breeding range overlap with
rangelands and ~ 50% exhibit declining population trends (Table 13.4). Causes of
shorebird declines are poorly understood.
13.4.4 Waterbirds
Waterbirds include > 180 species across 7 taxonomic orders that use marine and
inland aquatic habitats (PIF 2021). More than 60 waterbird species inhabit range-
lands (Table 13.5). Status of some waterbirds, especially secretive species, is poorly
440 J. L. Vest et al.
understood (Sauer 1999; Johnson et al. 2009), but survey information suggests vari-
ation in trends. Rosenberg et al. (2019) estimated a 22% decrease across 77 species.
Sandhill crane populations have increased in recent decades (Seamans 2022).
13.5 Habitat Associations
A broad overview of functional habitat relationships across groups of wetland birds
is provided herein. Habitat use varies by species, season, and time of day and
habitat associations are available for most species (e.g., eBird, Birds of the World).
Breeding and foraging characteristics are highly varied among waterbirds, shore-
birds, and waterfowl. Heterogeneous habitat with assorted wetland types, water
depths, vegetation density, and food support a diversity of wetland birds (Ma et al.
2010).
13.5.1 Waterfowl
Puddle ducks, or dabblers, are associated with shallow wetlands foraging near the
water surface by “tipping-up” to reach food items (Table 13.3;Fig. 13.3). However,
they can perform shallow dives to avoid predators or reach food. Dabblers use
seasonal and perennial wetlands with emergent vegetation for foraging and escape
cover, particularly important during the brooding period (Walker et al. 2013b;
Fig. 13.3). Seasonal wetlands are in overall decline (Collins et al. 2014; McCauley
et al. 2015; Donnelly et al. 2022). In the Intermountain West, rangeland seasonal
wetlands have been converted to flood-irrigated fields but can still provide important
habitat (Fleskes and Gregory 2010; Donnelly et al. 2019; Mackell et al. 2021). Semi-
permanent wetlands are habitat for dabblers and may be especially important during
periods of water scarcity (McCauley et al. 2015; Donnelly et al. 2022). Dabblers
also use open water as roosting habitat, especially when foraging habitat is nearby.
During the breeding period, grass-dominated upland habitats are vital for dabbler
nesting habitat. For example, nearly 90% of waterfowl in the PPR nest in uplands
(PPJV 2017).
Pochards, or diving ducks, are adapted to deeper aquatic systems where they forage
in the water column or benthic substrate (Figs. 13.3 and 13.4; Table 13.3). Common
benthic forage includes bivalves, worms, and insect larvae (Baldassarre and Bolen
2006). Divers often use wetlands with SAV for foraging (e.g., pondweeds [Potamoge-
tonaceae]). Sea ducks generally have high salinity tolerance, forage deeper, and are
uncommon in rangelands (Table 13.3), though bufflehead and common goldeneye
use rangeland wetlands (Baldassarre and Bolen 2006). Mergansers also use rangeland
aquatic habitats and forage on small fish, often in deeper water systems.
Swans use wetland habitat similar to diving ducks (Fig. 13.4). Swans use their long
necks to access SAV. Breeding trumpeter swans use freshwater marshes, ponds, lakes,
13 Waterfowl and Wetland Birds 441
Table 13.6 General waterbird habitat associations based on amount of emergent vegetation, open
water, and nesting habitat
Group A Group B Group C Group D Group E
Wetland with: Wetland with: Wetland with: Wetland with: Lake or River
with:
Substantial
emergent
vegetation
Emergent
vegetation
Emergent
vegetation
Emergent
vegetation
Open water
Variable open
water
Partial open
water
Extensive open
water
Open water •Barren
ground
Nesting trees Islands
American Bittern Sandhill Crane Common Loon Great Blue
Heron
American
White Pelican
Least Bittern White-faced Ibis Pied-billed Grebe Great Egret Double-crested
Cormorant
Black-crowned
Night-Heron
Franklin’s Gull Horned Grebe Snowy Egret Ring-billed
Gull
Yellow Rail Bonaparte’s Gull Red-necked Grebe Tricolored
Heron
California Gull
Black Rail Forster’s Tern Eared Grebe Little Blue
Heron
Herring Gull
King Rail Black Tern Wes t ern G reb e Cattle Egret Caspian Tern
Virgina Ra il Clark’s Grebe Green Heron Common Tern
Sora White-faced Ibis Yellow-crowned
Night Heron
Least Tern
American Coot
Common
Moorhen
Adapted from Beyersbergen et al. (2004)
and occasionally slowly moving streams. Basic breeding habitat features include
sufficient open water to take flight (about 100 m), SAV, stable water levels, structure
for nest sites, and low human disturbance (Baldassarre 2014; Mitchell and Eicholz
2020). Both tundra and trumpeter swans can forage in upland agricultural areas during
the non-breeding season. Migrating tundra swans show strong selection for wetlands
with sago pondweed (Stuckenia pectinata) while nonforaging swans selected large
open water areas (Earnst 1994).
Canada geese commonly occur in rangelands (Baldassarre 2014) and are primarily
grazers of grasses and sedges, though non-breeding geese can be dependent on crops.
Canada geese use a greater diversity of nest sites than other waterfowl (Baldassarre
2014). Common brood-rearing habitat includes gradually sloping ponds or river
shorelines, abundant graminoids, and mudflats (Mowbray et al. 2020).
442 J. L. Vest et al.
Fig. 13.3 Preferred foraging depths of select wetland birds. Modified from Tori et al. (2002),
Helmers (1992), and Richmond et al. (2012)
Fig. 13.4 Principal foraging habitats of various waterfowl groups with respect to water depth, plant
communities, and wetland hydroperiod. Modified from Krapu and Reinecke (1992)
13.5.2 Shorebirds
Shorebirds use a variety of wetland and upland habitats throughout the year. Most
shorebirds select shallow wetlands, wet meadows, shorelines, and open mud flats
for foraging and avoid tall and dense vegetation (Iglecia and Winn 2021). For
example, marbled godwits and willets select short sparse upland vegetation and
13 Waterfowl and Wetland Birds 443
wetland complexes for nesting and foraging (Niemuth et al. 2012; Shaffer et al.
2019a, b; Specht et al. 2020). However, these species can use taller and denser
vegetation when brooding (Shaffer et al. 2019a, b). Some shorebirds use uplands
for breeding, but s hift to wetlands later (Shaffer et al. 2019a, c; Niemuth et al.
2012). Shorebirds exhibit varied wetland salinity tolerances. Some breeding shore-
birds solely use uplands (Shaffer et al. 2019d, e, f; Iglecia and Winn 2021). During
migration, shorebirds select shallow, sparsely vegetated wetlands often with mudflats.
For example, shorebirds in the fall correlate positively with grazing pressure, and
negatively with denser vegetation (Albanese and Davis 2015). Aquatic and terres-
trial invertebrates are common shorebird foods, although seeds, vegetation, algae,
and small fish are consumed opportunistically. Dominant invertebrate prey items
include chironomids, flies (Diptera), beetles (Coleoptera), true bugs (Hemiptera),
amphipods (Amphipoda), snails (Gastropoda), and clams and mussels (Bivalvia).
Water depth, in combination with leg and bill length, determines food availability
and habitat types used by different shorebirds (Fig. 13.3).
13.5.3 Waterbirds
Waterbirds exhibit diversity in morphology, life history, and habitat use. Species
range from large and conspicuous cranes to secretive marsh birds such as bitterns
and rails (Table 13.5). Waterbirds use various wetland types with assorted amounts of
emergent vegetation, open water, water depth, and woody vegetation (Fig. 13.3; Table
13.6; Beyersbergen et al. 2004). Species use different areas within a wetland. For
example, white-faced ibis nest on emergent vegetation in colonies and use shallow
flooded areas to forage (Coons 2021; Moulton et al. 2022). Similarly, sandhill cranes
nest on mounds in shallow water and use adjacent uplands for foraging (Austin et al.
2007; Ivey and Dugger 2008). Many other waterbirds use flooded areas in rangelands
where management often mimics natural hydroperiods (Ivey and Herziger 2006).
13.6 Rangeland Management
Livestock production and wetland bird populations are linked by their dependence
on rangelands and surface water (Bue et al. 1964; Richmond et al. 2012; Brasher
et al. 2019). Grazing, burning, haying, and water management in wetlands and
uplands often increase resources for wetland birds (Kadlec and Smith 1992; Naugle
et al. 2000; Baldassarre and Bolen 2006). Response to management practices vary
among species, spatial scales, biological parameters, season, and locale. Managers
should consider objectives, seasonal habitat needs, and potential tradeoffs. Generally,
management that provides a mosaic of upland and wetland habitat is best (Naugle
et al. 2000; Baldassarre and Bolen 2006; Krausman et al. 2009;Maetal. 2010).
444 J. L. Vest et al.
13.6.1 Grazing
Upland nesting ducks generally favor dense cover within 4 km (~ 2.5 miles) of
wetlands (Reynolds et al. 2006). Nest survival correlates positively with vegetation
height (Baldassarre and Bolen 2006; Bloom et al. 2013) and the amount of adjacent
grassland (Greenwood et al. 1995; Reynolds et al. 2001; Stephens et al. 2005), rein-
forcing the need to conserve rangelands. Lack of disturbance can negatively impact
grassland, and duck productivity (Naugle et al. 2000; Dixon et al. 2019; Grant et al.
2009). Recent literature has demonstrated the compatibility of livestock grazing with
waterfowl habitat (Naugle et al. 2000; Ignatiuk and Duncan 2001;Warrenetal. 2008;
Bloom et al. 2013; Rischette et al. 2021). Livestock grazing is a land-use that can
ultimately support wetland birds (PPJV 2017; Brasher et al. 2019;PHJV 2021).
However, localized impacts of grazing can depend on timing, intensity, duration,
bird species, and demographics (Briske et al. 2011; Lipsey and Naugle 2017).
Managing grazing for residual cover (> 28 cm; Bloom et al. 2013) will enhance
waterfowl nest survival (Warren et al. 2008; Rischette et al. 2021), which is highest
when cover provides a physical barrier to predators. Grazing timing and intensity has
complex interactions with nest density and survival, along with local and landscape
conditions such as precipitation, site quality, and predator dynamics (Herkert et al.
2003; Stephens et al. 2005; Warren et al. 2008; Bloom et al. 2013; Ringelman et al.
2018). Mismanagement leading to overgrazing is detrimental to wetland birds and
rangeland health (Kadlec and Smith 1992; Krausman et al. 2009). To maximize
productivity, disturbances (e.g., grazing) should occur after or late in the nesting
period (Barker et al. 1990; Naugle et al. 2000). From an operational viewpoint, when
areas must be grazed, moderate to low stocking rates are preferred for waterfowl
nesting cover (Bloom et al. 2013; Rischette et al. 2021).
Multiple grazing systems can support wetland birds while meeting rangeland
health and producer objectives. Generally, systems that emphasize residual and
dense grass cover are beneficial for waterfowl nesting habitat (Chap. 4; Table 4.2;
Holechek et al. 1982;Barkeretal. 1990; Ignatiuk and Duncan 2001; Murphy et al.
2004; West and Messmer 2006; Krausman et al. 2009). Studies have indicated that
grazing systems with deferment, rotation, and rest (e.g., deferred rotation, rest rota-
tion, deferred rest rotation, and high-intensity low-frequency) can increase residual
cover and support wetland bird productivity (Gjersin 1975; Mundinger 1976;Barker
et al. 1990; Ignatiuk and Duncan 2001; Murphy et al. 2004; Carroll et al. 2007;
Emery et al. 2005; Shaffer et al. 2019a, b, d, e). Resting or deferring grazing in
wetlands during the non-breeding season can maintain plant-based foods for water-
fowl. Conversely, for many shorebird species, abundance correlates positively with
increased grazing pressure, particularly in the non-breeding season (Holechek et al.
1982; Powers and Glimp 1996; Albanese and Davis 2015). In areas with longer
growing seasons (e.g., Central Valley of California), grazing July–October supported
forage for wintering geese and cranes along with nesting cover (Carroll et al. 2007).
However, fall and winter grazing within shorter growing seasons may reduce initial
13 Waterfowl and Wetland Birds 445
residual cover, albeit with less influence on later nests due to vegetation growth. Eval-
uating contributions of local-scale management over the short-term (2–3 years) is
challenging because productivity can also be influenced by large-scale and carry-over
effects (Ringelman et al. 2018; Bortolotti et al. 2022).
Maintaining wetland vegetation structure and availability (Murkin et al. 1997;
Masto et al. 2022) is key to nesting, foraging, and brood-rearing habitat for most
species (Harrison et al. 2017). In wetlands dominated by robust and monotypic
perennials ( e.g., cattail) or invasives (e.g., reed canary grass [Phalaris arundinacea]),
grazing can improve habitat structural diversity, especially in conjunction with prac-
tices such as fire, herbicides, and water-level manipulation (Stutzenbaker and Weller
1989; Schultz et al. 1994; Anderson et al. 2019; Bansal et al. 2019; Hillhouse 2019).
Maintaining emergent vegetation is important for escape cover and food (Walker
et al. 2013b). While reducing vegetation structure along shorelines may be better for
shorebirds, excessive grazing can reduce habitat quality for other species (Hoffman
and Stanley 1978; Harrison et al. 2017; Iglecia and Winn 2021). For nests along
shorelines (e.g., snowy plover), restricting livestock access, or delaying grazing, can
increase productivity (Iglecia and Winn 2021). Many wetland plants have high nutri-
tion value and forage production generally exceeds uplands sites (Johnson 2019).
Graminoids in mesic areas usually provide high forage quality for livestock (Hubbard
1988; Kirby et al. 2002).
In regions where available water is limited, livestock disproportionately select wet
areas increasing the risk of habitat degradation. Historically, improper grazing has led
to deterioration of wetlands and negatively impacted wetland birds (Tessman 2004).
However, there is a paucity of research concerning grazing impacts on wetland bird
survival and productivity, especially in the Intermountain West (Gilbert et al. 1996;
Powers and Glimp 1996; Ivey and Dugger 2008; McWethy and Austin 2009). Risk
of nest failure due to predation or trampling is generally associated with increased
stocking rates (Littlefield and Paullin 1990; Bleho et al. 2014; Harrison et al. 2017;
Shaffer et al. 2019d, e). Increases in water scarcity will likely exacerbate grazing
impacts on wetland birds.
During the non-breeding period, moist-soil vegetation and seasonally flooded
areas should be the focus of resource managers (Fredrickson and Taylor 1982; Smith
et al. 1989; Haukos and Smith 1993; Hillhouse 2019). Moist-soil communities domi-
nated by annual plants such as smartweed, common ragweed (Ambrosia artemisi-
ifolia), and barnyardgrass, as well as perennials such as sedges, spike-rushes, giant
bur-reed (Sparganium eurycarpum), and dock (Rumex spp.) offer high quality forage
(Chabreck et al. 1989; Haukos and Smith 1993; Anderson et al. 2019). Consequently,
factors that decrease seed production reduce food availability and carrying capacity.
Grazing late summer can reduce seed production, whereas grazing until mid-summer
may allow plants and seed production to recover (Chabreck et al. 1989; Anderson
et al. 2019; Hillhouse 2019).
At landscape scales, livestock grazing helps maintain rangeland and wetland
habitat, but negative effects can occur at smaller scales, although most issues can
be addressed through management. For example, rotation and cross fencing can be
used to control when and where grazing occurs and help maintain economic and
446 J. L. Vest et al.
ecological viability (Fynn and Jackson 2022). However, fencing can facilitate meso-
predator movements and cause collisions for wetland birds, especially for species
that fly close to the water surface or take flight by running across the water surface
(Cornwell and Hochbaum 1971; Allen and Ramirez 1990).
13.6.2 Haying/mowing
Delaying haying until late nesting season helps minimize adult mortality and nest
failure. However, optimal hay quality in some areas may occur earlier creating a chal-
lenge for livestock operations (Epperson et al. 1999; Gruntorad et al. 2021). Flushing
bars mounted to haying equipment may help prevent adult mortality, but nests are still
destroyed. Haying patterns that move concentrically out from the middle of the field
may provide more opportunity for young birds to escape (Ivey 2011). Haying reduces
residual vegetation the following nesting season and generally results in lower nest
densities and productivity (Renner et al. 1995; Naugle et al. 2000; Rischette et al.
2021). Early nesting species (e.g., mallard, northern pintail) are impacted more by
haying than later nesting species (Luttschwager et al. 1994; Renner et al. 1995).
Ideally, haying should be late enough to minimize disturbance to nesting birds but
early enough for precipitation and regrowth late in the growing season (Rischette
et al. 2021).
Many wetland resources depend on irrigation with haying and grazing (Lovvorn
and Hart 2004; Copeland et al. 2010; Donnelly et al. 2021). Early haying (e.g.,
mid-June) may cause nest failure and reduced foraging as well as mortality of
unfledged waterbirds (Littlefield 1999; Ivey and Herziger 2006). Concomitantly, irri-
gated hayfields provide productive breeding habitat for species that select shorter and
sparse vegetation (Hartman and Oring 2009; Shaffer et al. 2019d). The short-stature
vegetation from haying (or grazing) can provide foraging habitat the following spring
and summer when these areas are flooded (Fleskes and Gregory 2010; Donnelly et al.
2019).
13.6.3 Fire
Historically, fire was a principal driver of ecosystem structure throughout the Great
Plains (Chap. 6). Burns reset succession to more productive states providing improved
nesting and foraging habitat. Prescribed burns can be used to provide desired plant
communities for wetland birds (Smith et al. 1989; Kadlec and Smith 1992; Anderson
et al. 2019). Fire in marshes and prairie wetlands can reduce dense vegetation,
increase food resources, promote desirable plants, provide new growth, and increase
plant nutrition (Smith and Kadlec 1985, 1992; Chabreck et al. 1989; Stutzenbaker
and Weller 1989; Naugle et al. 2000; Brennan et al. 2005; Venne and Frederick
2013; Anderson et al. 2019). Fire effects vary by location, season, and species needs.
13 Waterfowl and Wetland Birds 447
Because seasonal habitat requirements vary widely across wetland bird species,
providing a mosaic of burned and unburned areas at multiple scales is likely ideal
(Gray et al. 2013). If well-managed, fire can support broad ecological functioning
(Hovick et al. 2017).
13.6.4 Water Management
Livestock operations in semi-arid rangelands have long used surface water develop-
ments. Inadequate water can lead to poor livestock distribution and utilization issues
(Bue et al. 1964; Holechek et al. 2011). Water developments (e.g., stock ponds)
for livestock can provide habitat for wetland birds (Forman et al. 1996; Pederson
et al. 1989; May et al. 2002; Baldassarre and Bolen 2006). Stock ponds are dammed
watercourses, excavated areas, or a combination of both. Excavated stock ponds in
seasonal wetlands provide additional water accumulation, causing altered hydrope-
riods and less-preferred vegetation (Gray et al. 2013; Smith 2003; Baldassarre and
Bolen 2006). Constructing terraces can provide shallow water and emergent vege-
tation (Gray and Bolen 1987). Selection of stock ponds is influenced by multiple
factors including size, water depth, emergent and submergent vegetation, proximity
to other wetlands, and adjacent nesting cover (Austin and Buhl 2009).
Stock ponds that provide various water depths and diverse vegetation will be
attractive to multiple species (Ma et al. 2010). Surface area, shoreline complexity,
and vegetation composition are key characteristics for breeding season selection
(Flake et al. 1977; Austin and Buhl 2009). Shorebirds may benefit from grazed pond
margins and adjacent uplands (Laubhan and Gammonley 2000; May et al. 2002).
Irregular shorelines, improved water quality, and SAV are attractive to breeding
ducks (Hudson 1983; Svingen and Anderson 1998; Austin and Buhl 2009). Ponds,
and natural wetlands, that approximate a 50:50 ratio of emergent vegetation and open
water (i.e., hemi-marsh) provide ideal conditions for many wetland birds, particularly
waterfowl (Murkin et al. 1997; Smith et al. 2004). Stock ponds are common in areas
with limited water availability. Rumble and Flake (1983) recommend ponds for
waterfowl broods that have: (1) larger surface area, (2) shallow water supporting
submersed and emergent vegetation, (3) grazing management fostering emergent
vegetation, (4) adjacent upland cover, and (5) undrained nearby wetlands. Exclusion
fencing in shallows may promote emergent and moist-soil plants for food and cover.
Water developments for livestock are also used during non-breeding periods.
Approximately half of the ducks during 1997–2014 mid-winter surveys in Texas were
detected on stock ponds (Texas Parks and Wildlife Department unpublished data;
DU 2021). Medium-sized ponds (0.81–16.2 ha) had higher occupancy (32–51%)
compared to smaller ponds (< 0.81 ha; 11–26%; Texas Parks and Wildlife Department
unpublished data; Mason et al. 2013). Evidence suggests stock ponds may help offset
reduced habitat availability during drought (DU 2021). Along the Gulf Coast, stock
ponds provided wintering habitat and freshwater sources for waterfowl, shaping
distribution, abundance, and foraging patterns (Adair et al. 1996; Ballard et al. 2010).
448 J. L. Vest et al.
Forage availability in stock ponds is currently not well understood, but likely highly
variable (Kraai 2003;Clark 2016). Stock ponds may provide important refugia during
non-breeding periods (Kraai 2003; K. Kraai, Texas Parks and Wildlife Department,
personal communication).
The relationship between irrigation and wetlands is complex (Bolen et al. 1989;
Lovvorn and Hart 2004; Bishop and Vrtiska 2008; Moore 2016; Donnelly et al.
2020; King et al. 2021). Donnelly et al. (2022) indicate rapid wetland decline in
western North America may be approaching an ecological tipping point for wetland
bird populations. In the West, most surface water rights are agricultural and used in
irrigation systems (Kendy 2006; Downard and Endter-Wada 2013; Donnelly et al.
2020; King et al. 2021). In many areas, availability of wetland habitat follows irri-
gation schedules and further research is needed to better understand benefits and
relative tradeoffs for wetland birds throughout the annual cycle (Copeland et al.
2010; Donnelly et al. 2019, 2020, 2021; Lovvorn and Crozier 2022). Water scarcity
is intensifying socio-political pressures, including “use it or lose it” policies, to
improve efficiency (Grafton et al. 2018; Sketch et al. 2020). However, more efficient
irrigation practices (e.g., pressurized sprinklers) could lead to significant declines
in flood irrigation and negatively impact wetland bird habitat and other ecosystem
services (Baker et al. 2014; Moulton et al. 2016; Donnelly et al. 2020, 2021). Rapid
wetland declines in the West may be approaching an ecological tipping point for
wetland bird populations (Donnelly et al. 2020, 2022).
13.7 Ecosystem Threats
Wetland loss has been extensive, with > 50% declines in the western U.S. and
Great Plains (Dahl 1990, 2014) and comparable losses (40–70%) in western Canada
(Doherty et al. 2013), and Mexico (25–98%; Landgrave and Mereno-Casasola 2012).
In the Great Plains, the most significant driver has been wetland drainage (e.g., tiling)
tied to row-crop expansion, and loss of wetland legal protections (Dahl 1990, 2011;
Lark et al. 2020). In the West, agricultural development and large-scale overexploita-
tion of beavers in the 1800s led to widespread wetland losses (Dahl 1990;Lemly
et al. 2000; McKinstry et al. 2001; Chap. 7). Intact wetlands and rangelands tend to
be associated with livestock production and land owned by public agencies. Wetland
bird conservation is therefore intrinsically linked to livestock production (Higgins
et al. 2002; Anderson et al. 2018; Brasher et al. 2019). Climate change is predicted to
exacerbate threats (Niemuth et al. 2014;Haigetal.
2019;Larketal.
2020; Donnelly
et al. 2021; Moon et al. 2021). Conservation of remaining wetlands, especially in
rangelands, will be important to sustain wetland birds (Bartuszevige et al. 2012;Tsai
et al. 2012; PPJV 2017;PHJV 2021; Donnelly et al. 2021).
13 Waterfowl and Wetland Birds 449
13.7.1 Habitat Conversion and Alteration
Recent changes in row-crop agriculture, such as the development of drought-resistant
crop varieties and increased farming efficiencies, provide incentives to convert range-
land and other marginal areas into crop production (Higgins et al. 2002; Doherty et al.
2013;Larketal. 2020). Recently, the most extensive conversion has occurred in the
PPR and High Plains (RWBJV 2013; Fields and Barnes 2019;Larketal. 2020).
Lark et al. (2020) found recently converted grasslands and wetlands in the PPR had
37% less nesting accessibility for ducks than non-converted areas, demonstrating
the significant risk of agricultural conversion to wetland bird productivity. Along
the Gulf Coast, human development, crop conversion, non-native grass pastures, and
wetland draining has led to < 1% of native prairie remaining and significant loss to
wetland bird nesting habitat (Smeins et al. 1991; Wilson and Esslinger 2002; Vermil-
lion et al. 2008). The loss of ranching operations and subdivision of land ownership
has contributed to habitat declines. Future development is expected to increase 72%
over the next 80 years putting remaining rangeland and wetlands at further risk (Moon
et al. 2021). In the Intermountain West, human population growth and water scarcity
have intensified competition for water resources driving substantial land-use changes
that impact wetland bird habitat (Hansen et al. 2002; Baker et al. 2014; Donnelly
et al. 2021; King et al. 2021). Water-use is increasingly transferred from agricultural
to municipal holdings for growing urban water demands, increasing the challenge
of maintaining regional wetland networks (Brewer et al. 2007; Dilling et al. 2019;
Donnelly et al. 2021). The accumulation of increasing threats within the Intermoun-
tain West has potential negative population-level impacts (Haig et al. 1998, 2019;
Donnelly et al. 2020, 2021; Mackell et al. 2021).
13.7.2 Energy Development
Energy development continues to increase across rangelands (Ott et al. 2021). Colli-
sions, habitat loss and degradation, and displacement are common impacts from
energy development that threaten wetland bird populations (Shaffer et al. 2019g).
Oil field wastewater developments in semi-arid rangelands are commonly mistaken as
habitat by wetland birds resulting in mortality (Flickinger 1981; Flickinger and Bunck
1987;Trail 2006; Ramirez 2010). Oil spills and flowback water from fracking occur
regularly and can contaminate wetlands. Brine contamination has been frequently
reported in wetlands in the Bakken Formation and can negatively affect local aquatic
invertebrates (Preston and Ray 2017; Blewett et al. 2017). The demand for biofuel,
particularly corn ethanol, has accelerated grassland and wetland conversion of >
400,000 ha per year (Wright and Wimberly 2013;Larketal.
2015, 2020). This
conversion leads to increases in land values, affecting livestock operation sustain-
ability, and portends challenges for ranching economies and associated ecosystem
services (Johnson and Stephens 2011).
450 J. L. Vest et al.
Indirect losses from energy development include fragmentation and displacement,
which significantly increases the footprint of habitat loss (Johnson and Stephens
2011; Loesch et al. 2013). Indirect effects vary by species, seasons, and spatial scale
of habitat (Shaffer et al. 2019g; Pearse et al. 2021). Lower breeding (Loesch et al.
2013) and wintering abundance (Lange et al. 2018) of ducks have been documented
near wind energy facilities as well as avoidance during migration by whooping cranes
(Pearse et al. 2021). Fragmentation and displacement from wind development are
of greater conservation concern compared to direct mortality (Shaffer et al. 2019g;
Hise et al. 2020). Larger wetland birds, such as sandhill cranes, are at greater risk of
collision (Brown and Drewien 1995; Navarrete and Griffis-Kyle 2014; Murphy et al.
2016; Pearse et al. 2016; Hays et al. 2021).
13.7.3 Invasive Species
Invasive flora and fauna affect wetland birds in rangelands. Invasive aquatic plants
reduce overall biodiversity and habitat quality for waterbirds. Native and non-native
plant species such as cattail, common reed, reed canary grass, and creeping foxtail
(Alopecurus arundinaceus) form dense monotypic stands that outcompete more
desirable vegetation (Baldassarre and Bolen 2006; Hillhouse 2019; Johnson 2019).
Cattail species have proliferated in the absence of natural disturbances (e.g., grazing
and fire) and row-crop agriculture provides conditions that promote cattail establish-
ment and vigor (i.e., nutrient runoff, sediment accumulation; Bansal et al. 2019).
Dense stands of cattail can dominate wetlands, eliminate open-water, replace emer-
gents and SAV, and preclude wetland bird species (Bansal et al. 2019). Similarly,
common reed (i.e., phragmites) is a growing problem in the Intermountain West
(Duncan et al. 2019; Rohal et al. 2019). Reed canary grass is widely used as live-
stock forage but can quickly form dense unproductive monotypic stands (Paveglio
and Kilbride 2000; Evans-Peters et al. 2012; Hillhouse 2019).
Invasive animals also negatively affect waterfowl either directly through preda-
tion (e.g., northern pike (Esox lucius), or indirectly through habitat degradation.
Common carp (Cyprinus carpio) are pervasive and degrade habitat quality and water-
fowl productivity by consuming SAV and increasing turbidity that reduces forage
availability (Ivey et al. 1998; Bajer et al. 2009). However, carp control is challenging,
and success is often short-lived (Pearson et al. 2019). PPR wetlands evolved under
isolated and intermittent drying conditions with only temporary surface-hydrologic
connections. Wetland drainage has resulted in deeper, more stabilized hydrology,
with interconnected basins that permit fish to persist (McLean et al. 2022). Simi-
larly, fish are not endemic to playas in the High Plains but excavated ponds for
irrigation support introduced fish, causing similar issues as above (Bolen et al. 1989;
Smith et al. 2012).
13 Waterfowl and Wetland Birds 451
13.7.4 Climate Change
The availability and function of wetlands are balanced by precipitation and evapo-
transpiration, making them sensitive to changes in climate (McKenna et al. 2021a).
Climate change will likely have variable effects on wetland function and produc-
tivity throughout North American rangelands. Indirect climate change impacts on
land-use are also conservation concerns for wetland birds (McKenna et al. 2019). In
response to climate change, water availability and land-use patterns will increasingly
challenge agricultural-based economies and wetland bird populations.
The PPR has received considerable attention for evaluating potential climate
change effects on wetland birds. Recommendations for waterfowl conservation
strategies have shifted as climate change has been increasingly understood. Areas that
currently support the largest densities of intact wetlands and breeding populations
will likely be most critical to future continental waterfowl populations (Loesch et al.
2012; Niemuth et al. 2014; Sofaer et al. 2016; McKenna et al. 2021a). Many of these
wetlands overlap rangeland areas with ranch and livestock-based economies (PPJV
2017). In the southern PPR, a shift from winter to summer and fall precipitation-
driven hydrology has occurred in recent decades (McKenna et al. 2017). More
precipitation may initially seem beneficial, but wetland productivity and function
can decline with less periodic drying (Euliss et al. 2004; McCauley et al. 2015).
Under wetter conditions, wetlands would deepen and have more stable water levels
promoting fish persistence and cattail domination (Anteau et al. 2016). Shorebirds
that require exposed shorelines and mudflats would be less likely to find habitat
(Anteau et al. 2016). Alternatively, prolonged dry periods can result in loss of seasonal
wetlands and shrinking wetlands alter plant and invertebrate communities. Upland
management, such as grazing and burning, adjacent to wetlands can help increase
runoff into wetlands and reduce ponding loss during the breeding season (McKenna
et al. 2021b).
In the Southern Great Plains, spring and summer are expected to become hotter and
drier with fewer, but more intense and unpredictable, precipitation events (Londe et al.
2022). Recent models indicate a high likelihood that wetland networks will exhibit
reduced connectivity, with playas especially at risk (Uden et al. 2015; Albanese
and Haukos 2017; McIntyre et al. 2018; Verheijen et al. 2020; Londe et al. 2022).
Loss of stopover habitat and forage can reduce survival during the non-breeding
period (Moon and Haukos 2006) and subsequent reproductive success (Sedinger and
Alisauskas 2014). Reduced summer wetland inundation also means less available
water for livestock. Opportunities to introduce rangeland management practices,
such as fire and/or grazing in Conservation Reserve Program lands, (Cariveau et al.
2011; Smith et al. 2011) may become increasingly important to address climate
impacts.
The West is experiencing rising temperatures, reduced snowpack, and earlier
runoff resulting in water scarcity (Kapnick and Hall 2012;Moteetal.
2018; Snyder
et al. 2019). Snowpack runoff drives availability and function for most western
452 J. L. Vest et al.
wetlands. In recent decades, water surface area in wetlands have declined by 47%
or more while important aquatic systems like the Great Salt Lake have declined by
27% (Donnelly et al. 2020). Terminal basins and lower portions of watersheds in
the Great Basin (Kadlec and Smith 1989; Donnelly et al. 2020) are strongly influ-
enced by upstream water management decisions (Moore 2016; Null and Wurtsbaugh
2020; King et al. 2021; Donnelly et al. 2022). Climate change brings increasing
temperatures and evapotranspiration rates intensifying water scarcity and ultimately
impacting wetland bird habitat in the region (Downard and Endter-wada 2013; Moore
2016;Haigetal. 2019; Donnelly et al. 2020, 2021, 2022).
Climate change has potential to affect Gulf Coast habitat through sea-level rise
and intensification of tropical storms. Coastal wetlands are vulnerable to increasing
salinity, which decreases primary production, altering habitat quality (Battaglia et al.
2012; Moon et al. 2021). Freshwater and irregularly flooded marshes (Chabreck et al.
1989; Wilson and Esslinger 2002), are projected to dramatically decrease (Moon et al.
2021). Inland prairie and agricultural wetlands are also at risk (Battaglia et al. 2012;
Moon et al. 2021) but may continue to provide vital habitat to species like mottled
duck (Moon et al. 2021).
13.8 Conservation and Management Actions
13.8.1 Addressing Loss and Fragmentation of Wetlands
and Rangelands
Minimizing the conversion of wetlands and rangelands to cultivated agricultural
production is one of the greatest conservation challenges and priorities for wetland
birds. Unfortunately, increases in commodity prices and the slow pace of conser-
vation actions are unlikely to reverse wetland bird habitat losses in rangelands or
offset anticipated future losses (Higgins et al. 2002; Doherty et al. 2013;Larketal.
2020). However, maintaining livestock production on rangelands decreases the like-
lihood of cropland conversion and other land use changes (Higgins et al. 2002).
Therefore, sustaining grazing as part of the region’s socio-economic fabric will be
vital for conserving wetland bird habitats (Higgins et al. 2002). Where grasslands
have been lost, the maintenance and conservation of wetland basins supports wetland
bird persistence (Reynolds et al. 2006; Niemuth et al. 2009). Nevertheless, keeping
rangelands “green side up” and wetlands intact are primary conservation goals to
sustain wetland bird populations.
Flood irrigation, beaver restoration, and low-tech riparian and wet meadow
restoration (e.g., beaver dam mimicry or analogs, Zeedyk structures) offer opportu-
nities to enhance natural water storage (Blevins 2015; Silverman et al. 2018; Moore
and McEvoy 2022). Enhanced soil water storage capacity from such practices can
increase watershed resilience to climate drivers, enhance wetland wildlife habitat,
13 Waterfowl and Wetland Birds 453
and increase livestock forage production (Silverman et al. 2018). Financial incen-
tives, access to technical assistance, and local partnerships help managers implement
restoration as well as maintain or upgrade flood irrigation infrastructure (Sketch
et al. 2020; Donnelly et al. 2021; Moore and McEvoy 2022). Watershed and state-
based partnerships will help managers navigate water management, water rights, and
restoration techniques within the social-ecological systems of western watersheds
(Downard and Endter-Wada 2013; Moore and McEvoy 2022).
13.8.2 Partnerships and Programs
Conserving wetland birds requires effective public–private partnerships at local,
regional, and international scales (Anderson et al. 2018; Brasher et al. 2019). For
example, the NAWMP acknowledges sustaining waterfowl populations is impossible
without conservation on private lands and no single entity can solely address habitat
loss. Conservation partnerships for wetlands and grasslands have historically focused
on voluntary incentive programs such as those available through the federal Farm
Bill (Hohman et al. 2014) for sustaining and growing wetland bird populations (e.g.,
Gray and Teels 2006; Reynolds et al. 2006; Bishop and Vrtiska 2008;Drumetal.
2015). More recent partnerships have focused on adaptive conservation projects in
working rangelands, including the creation and maintenance of water sources that
concurrently improve livestock grazing management and wildlife habitat. Effective
wetland bird conservation includes a broad suite of short-term and long-term steward-
ship programs and incentives for livestock operations (Higgins et al. 2002; Brasher
et al. 2019).
Numerous agency programs are available to assist with range improvements,
grazing infrastructure, and wetland restoration and protection (Brasher et al. 2019).
Prominent federal examples include Natural Resource Conservation Service (e.g.,
EQIP, WRE—including reserved grazing rights option) and the U.S. Fish and
Wildlife Service Partners for Fish and Wildlife Program which can provide technical
expertise and funding for wetland conservation projects that align with supporting
producer objectives. Community-based conservation efforts can foster productive
dialogue among stakeholders for meaningful conservation actions (Neudecker et al.
2011; Bennett et al. 2021). Voluntary conservation easements, including NRCS’s
Agricultural Conservation Easement program, limit sub-division, development, and
conversion of rangelands to other land-uses (Brasher et al. 2019; Bennett et al. 2021).
Prioritization is needed to help distribute limited resources (Niemuth et al. 2022).
Facilitating land-use changes, like the transition of expiring Conservation Reserve
Program lands into grazed rangeland will sustain or improve habitat conditions for
wetland birds, expand grazing opportunities, and improve landscape resilience by
supporting sustainable ranching economies that keep grasslands and wetlands on the
landscape (Higgins et al. 2002; PPJV 2017; NRCS 2021).
The growing awareness of ecosystem services provided to society through
wetlands and rangelands are likely to generate additional public–private partnership
454 J. L. Vest et al.
opportunities and funding sources for conservation. Ecosystem services from range-
lands and wetlands include flood control, water quality, groundwater recharge and
discharge, carbon storage, and ecological resilience (Mitsch and Gosselink 2015).
Focus on improved wildlife resources has been a primary message for conservation
groups to date; however, helping people understand the life-sustaining ecosystem
services provided by rangeland and wetlands may increase stakeholder interest and
funding available for conservation (Bartuszevige et al. 2016; Humburg et al. 2018;
Brasher et al. 2019).
References
Ackerman JT (2002) Of mice and mallards: positive indirect effects of coexisting prey on waterfowl
nest success. Oikos 99:469–480
Adair SE, Moore JL, Kiel WH (1996) Wintering diving duck use of coastal ponds: an analysis of
alternative hypotheses. J Wildl Manage 60:83–93
Albanese G, Davis CA (2015) Characteristics within and around stopover wetlands used by
migratory shorebirds: is the neighborhood important? Condor 117:328–340
Albanese G, Haukos DA (2017) A network model framework for prioritizing wetland conservation
in the Great Plains. Landsc Ecol 32:115–130
Alisauskas RT, Drake KL, Nichols JD (2009) Filling a void: abundance estimation of North
American populations of Arctic geese using hunter recoveries. In: Thomson DL, Cooch EG,
Conroy MJ (eds) Modeling demographic processes in marked populations. Environmental and
ecological statistics vol 3. Springer Science, New York, pp 463–489
Alisauskas RT, Arnold TW, Leafloor JO, Otis DL, Sedinger JS (2014) Lincoln estimates of mallard
(Anas platyrhynchos) abundance in North America. Ecol Evol 4:132–143
Allen GT, Ramirez P (1990) A review of bird deaths on barbed-wire fences. Wilson Bull 102:553–
558
Anderson JT, Smith LM (2000) Invertebrate response to moist-soil management of playa wetlands.
Ecol Appl 550–558
Anderson MG, Alisauskas RT, Batt BDJ et al (2018) The migratory bird treaty and a century of
waterfowl conservation. J Wild Manage 82:247–259
Anderson BE, Hillhouse HL, Bishop AA et al (2019) Grazing Rainwater Basin wetlands. Exten-
sion circular EC3040, Institute of Agriculture and Natural Resources. University of Nebraska,
Lincoln, NE, USA. https://extensionpubs.unl.edu/publication/9000020939194/grazing-rainwa
ter-basin-wetlands/
Anteau MH, Afton AD (2011) Lipid catabolism of invertebrate predator indicates widespread
wetland ecosystem degradation. PLoS ONE 6:e16029
Anteau MJ, Afton AD, Anteau ACE et al (2011) Fish and land use influence Gammarus lacus-
tris and Hyalella Azteca (Amphipoda) densities in large wetlands across the upper Midwest.
Hydrobiologia 664:69–80
Anteau MJ, Shaffer TL, Sherfy MH et al (2012) Nest survival of piping plovers at a dynamic
reservoir indicates an ecological trap for a threatened population. Oecologia 170:1167–1179
Anteau MJ, Wiltermuth MT, van der Burg MP et al (2016) Prerequisites for understanding climate-
change impacts on northern prairie wetlands. Wetlands 36:299–307
Arnold TW, Roche EA, Devries JH et al (2012) Costs of reproduction in breeding female mallards:
predation risk during incubation drives annual mortality. Avian Conserv Ecol 7:1. https://doi.
org/10.5751/ACE-00504-070101
Austin JE, Buhl DA (2009) Factors associated with duck use of impounded and natural wetlands
in western South Dakota. Prairie Nat 41:1–27
13 Waterfowl and Wetland Birds 455
Austin JE, Henry AR, Ball IJ (2007) Sandhill crane abundance and nesting ecology at Grays Lake,
Idaho. J Wild Manage 71:1067–1079
Bajer PG, Sullivan G, Sorensen PW (2009) Effects of a rapidly increasing population of common
carp on vegetative cover and waterfowl in a recently restored Midwestern shallow lake.
Hydrobiologia 632:235–245
Baker JM, Everett Y, Liegel L, Van Kirk R (2014) Pattern of irrigated agricultural land conversion
in a Western U.S. watershed: implication for landscape-level water management and land-use
planning. Soc Nat Resour 27(11):1145–1160. https://doi.org/10.1080/08941920.2014.918231
Baldassarre GA (ed) (2014) Ducks, geese, and swans of North America. Johns Hopkins University
Press, Baltimore, Maryland
Baldassarre GA, Bolen EG (2006) Waterfowl ecology and management, 2nd edn. Krieger
Publishing, Malabar, Florida, USA
Ball IJ, Eng RL, Ball SK (1995) Population density and productivity of ducks on large grassland
tracts in northcentral Montana. Wild Soc Bull 23:767–773
Ballard BM, James JD, Bingham RL et al (2010) Coastal pond use by redhead wintering in the
Laguna madre, Texas. Wetlands 30:669–674
Bansal S, Lishawa SC, Newman S et al (2019) Typha (Cattail) invasion in North American wetlands:
biology, regional problems, impacts, ecosystem services, and management. Wetlands 39:645–
684
Barker WT, Sedivec KK, Messmer TA et al (1990) Effects of specialized grazing systems on
waterfowl production in south-central North Dakota. Trans North Am Wildl Nat Resour Conf
55:462–474
Bartuszevige AM, Pavlacky DC Jr, Burris L et al (2012) Inundation of playa wetlands in the western
Great Plains relative to landcover context. Wetlands 32:1103–1113
Bartuszevige AM, Taylor K, Daniels A et al (2016) Landscape design: integrating ecological, social,
and economic considerations into conservation planning. Wildl Soc Bull 40:411–422
Batt BDJ, Anderson MG, Anderson CD et al (1989) The use of prairie potholes by North American
ducks. In: van der valk AG (ed) Northern Prairie wetlands. Iowa State University Press, Ames,
Iowa, pp 2–14
Batt BD, Afton AD, Anderson MG et al (eds) (1992) Ecology and management of breeding
waterfowl. University of Minnesota Press, Minneapolis, Minnesota
Battaglia LL, Woodrey MS, Peterson MS et al (2012) Wetlands of the northern Gulf Coast. In:
Batzer DP, Baldwin AH (eds) Wetland habitats of North America. University of California
Press, Berkely and Los Angeles, California, pp 75–88
Bennett DE, Knapp CN, Knight RL (2021) The evolution of rangeland trusts network as a catalyst
for community-based conservation in the American West. Conserv Sci Pract 3:e257
Beyersbergen GW, Niemuth ND, Norton MR (eds) (2004) Northern Prairie and Parkland Waterbird
conservation plan. Waterbird Conservation for the Americas Initiative, Prairie Pothole Joint
Venture, Denver, Colorado, USA
Bishop AA, Vrtiska M (2008) Effects of the wetland reserve program on waterfowl carrying
capacity in the Rainwater Basin Region of south-central Nebraska. A Conservation Effects
Assessment Project, Wildlife component assessment. U.S. Department of Agriculture Natural
Resources Conservation Service, Washington, D.C. 51 pp. https://nrcs.usda.gov/publications/
ceap-wildlife-2008-wrp-waterfowl-rainwater-nebraska.pdf
Bleho BI, Koper N, Machtans CS (2014) Direct effects of cattle on grassland birds in Canada.
Conserv Biol 28:724–734
Blevins SC (2015) Valuing the non-agricultural benefits of flood irrigation in the Upper Green
River Basin. Master’s Thesis, Department of Agriculture and Applied Economics, University
of Wyoming, Laramie, Wyoming, USA, 79 pp
Blewett TA, Delompré PLM, He Y et al (2017) Sublethal and reproductive effects of acute and
chronic exposure to flowback and produced water from hydraulic fracturing on the water flea
Daphnia magna. Environ Sci Technol 51:3032–3039. https://doi.org/10.1021/acs.est.6b05179
456 J. L. Vest et al.
Bloom PM, Clark RG, Howerter DW (2012) Landscape-level correlates of mallard duckling
survival: implications for conservation programs. J Wildl Manage 76:813–823
Bloom PM, Howerter DW, Emery RE et al (2013) Relationships between grazing and waterfowl
production in the Canadian prairies. J Wildl Manage 77:534–544
Bolen EG, Smith LM, Schramm HL Jr (1989) Playa lakes: prairie wetlands of the Southern high
plains. Bioscience 39:615–623
Bortolotti LE, Emery RB, Armstrong LM et al (2022) Landscape composition, climate variability,
and their interaction drive waterfowl nest survival in the Canadian prairies. Ecosphere 13:e3908
Brasher MG, Giocomo JJ, Azure DA et al (2019) The history and importance of private lands for
North American waterfowl conservation. Wildl Soc Bull 43:338–354
Brennan EK, Smith LM, Haukos DA et al (2005) Short-term response of wetland birds to prescribed
burning in Rainwater Basin wetlands. Wetlands 25:667–674
Brewer J, Glennon R, Ker A et al (2007) Water markets in the West: prices, trading, and contractual
forms. NBER Working Paper No. 13002. National Bureau of Economic Research, Cambridge,
MA, USA. http://www.nber.org/papers/w13002
Briske DD, Derner JD, Milchunas DG et al (2011) An evidence-based assessment of prescribed
grazing practices. Rotational grazing on rangelands: reconciliation of perception and experi-
mental evidence. Rangel Ecol Manag 61:3–17
Brown WM, Drewien RC (1995) Evaluation of two power line markers to reduce crane and
waterfowl collision mortality. Wildl Soc Bull 23(2):217–227. https://www.jstor.org/stable/378
2794
Brown S, Hickey C, Harrington B, Gill R (eds) (2001) The U.S. shorebird conservation plan, 2nd
edn. Manomet Center for Conservation Sciences, Manomet, MA. https://www.shorebirdplan.
org/
Brownie C, Anderson Dr, Burnham KP et al (1985) Statistical inference from band recovery data—a
handbook. U. S. Department of the Interior, Fish and Wildlife Service, Resource Publication
156
Buderman FE, Devries JH, Koons DN (2020) Changes in climate and land use interact to create an
ecological trap in a migratory species. J Anim Ecol 89:1961–1977
Bue IG, Uhlig HG, Smith JD (1964) Stock ponds and dugouts. In: Linduska JP, Nelson AL (eds)
Waterfowl tomorrow. US Department of Interior, Washington, DC, pp 391–398
Cariveau AB, Pavlacky DC Jr, Bishop AA et al (2011) Effects of surrounding land use on playa
inundation following intense rainfall. Wetlands 31:65–73
Carroll LC, Arnold TW, Beam JA (2007) Effects of rotational grazing on nesting ducks in California.
J Wildl Manage 71:681–1021
Cavitt JF, Jones SL, Wilson NM et al (2014) Atlas of breeding colonial waterbirds in the interior
western United States. Research Report, U.S. Department of the Interior, Fish and Wildlife
Service, Denver, Colorado
Chabreck RH, Joanen T, Paulus SL (1989) Southern coastal marshes and lakes. In: Smith LM,
Pederson RL, Kaminski RM (eds) Habitat management for migrating and wintering waterfowl
in North America. Texas Tech University Press, Lubbock, Texas, pp 249–277
Clark L (2016) Stock pond forage resource relationships for nonbreeding ducks in the Rolling Plains
of Texas. Thesis, Texas Tech University, Lubbock, Texas
Clark RG, Nudds TD (1991) Habitat patch size and duck nest success: the crucial experiments have
not been performed. Wildl Soc Bull 19:534–543
Collins SD, Heintzman LJ, Starr SM et al (2014) Hydrologic dynamics of temporary wetlands in
the southern Great Plains as a function of surrounding land use. J Arid Environ 109:6–14
Cooch FG, Wendt S, Smith GEJ et al (1978) The Canada migratory game bird hunting permit and
associated surveys. In: Boyd H, Finney GH (eds) Migratory game bird hunters and hunting in
Canada. Canadian Wildlife Service Report Series vol 43, pp 8–39
Coons SP (2021) Monitoring the wetland landscape: white-faced ibis (Plegadis chihi) breeding
habitat as a model assemblage. Graduate student theses, dissertations, and professional papers.
11836. University of Montana, Missoula, Montana. https://scholarworks.umt.edu/etd/11838
13 Waterfowl and Wetland Birds 457
Copeland HE, Tessman SA, Girvetz EH et al (2010) A geospatial assessment on the distribution,
condition, and vulnerability of Wyoming’s wetlands. Ecol Indic 10:869–879
Cornwell G, Hochbaum HA (1971) Collisions with wires: a source of Anatid mortality. Wilson Bull
83:305–306
Cowardin LM, Carter V, Golet FC et al (1979) Classification of wetlands and deepwater habitats of
the United States. US Fish and Wildlife Service, Washington, DC, USA. FWS/OBS-79/31
Dahl TE (1990) Wetland losses in the United States, 1780s to 1980s. U.S. Fish and Wildlife
Service, Washington, D.C. USA. https://www.fws.gov/wetlands/documents/Wetlands-Status-
and-Trends-in-the-Conterminous-United-States-Mid-1970s-to-Mid-1980s.pdf
Dahl TE (2011) Status and trends of wetlands in the conterminous United States 2004–2009. U.S.
Department of the Interior, Fish and Wildlife Service, Washington, D.C. 108 pp
Dahl TE (2014) Status and trends of prairie wetlands in the United States 1997–2009. U.S. Depart-
ment of the Interior, Fish and Wildlife Service, Ecological Services, Washington, D.C. USA,
67pp
Davis BE (2012) Habitat use, movements, and ecology of female mottled ducks in the Gulf Coast
of Louisiana and Texas. Dissertation, Louisiana State University, Baton Rouge, Louisiana
Davis JB, Guillemain M, Kaminski RM (2014) Habitat and resource use by waterfowl in the northern
hemisphere in autumn and winter. Wildfowl (2014 Special Issue) 4:17–69
Devries JH, Brook RW, Howerter DW et al (2008) Effects of spring body condition and age on
reproduction in mallards (Anas platyrhynchos). Auk 125:618–628
Dilling L, Berggren J, Henderson J, Kenney D (2019) Savior of rural landscapes or Solomon’s
choice? Colorado’s experiment with alternative transfer methods for water (ATMs). Water Secur
6:100027. https://doi.org/10.1016/j.wasec.2019.100027
Dixon C, Vacek S, Grant T (2019) Evolving management paradigms on the U.S. Fish and Wildlife
Service lands in the Prairie Pothole Region. Rangelands 41:36–43
Doherty KE, Ryba AJ, Stemler CL, Niemuth ND, Meeks WA (2013) Conservation planning in an
era of change: state of the U.S. Prairie Pothole Region. Wildl Soc Bull 37(3):546–563. https://
doi.org/10.1002/wsb.284
Donaldson G, Hyslop C, Morrison G, Dickson L, Davidson I (eds) (2000) Canadian shorebird
conservation plan. Canadian Wildlife Service, Environment Canada, Ottaway, Ontario, Canada
Donnelly JP, Vest JL (2012) Identifying science priorities: 2013–2018 wetland focal strategies.
Intermountain West Joint Venture technical series 2012-3. Intermountain West Joint Venture,
Missoula, Montana. https://iwjv.org/wp-content/uploads/2019/08/iwjv_3_science_wetlands_2
013-2018.pdf
Donnelly JP, Naugle DE, Hagen CE et al (2016) Public lands and private waters: scarce mesic
resources structure land tenure and sage-grouse distributions. Ecosphere 7:1. https://doi.org/10.
1002/ecs2.1208
Donnelly JP, Naugle DE, Collins DP et al (2019) Synchronizing conservation to seasonal wetland
hydrology and waterbird migration in semi-arid landscapes. Ecosphere 10(6):e02758. https://
doi.org/10.1002/ecs2.2758
Donnelly JP, King S, Silverman NL et al (2020) Climate and human water use diminish wetland
networks supporting continental waterbird migration. Glob Chang Biol 26(4):2042–2059. https:/
/doi.org/10.1111/gcb.15010
Donnelly JP, King SL, Knetter J et al (2021) Migration efficiency sustains connectivity across
agroecological networks supporting sandhill crane migration. Ecosphere 12:6. https://doi.org/
10.1002/ecs2.3543
Donnelly JP, Moore JN, Casazza ML et al (2022) Functional wetland loss drives emerging risks to
waterbird migration networks. Front Ecol Evol 10:844278. https://doi.org/10.3389/fevo.2022.
844278
Downard R, Endter-Wada J (2013) Keeping wetlands wet in the western United States: adaptations
to drought in agriculture-dominated human-natural systems. J Environ Manag 131:394–406.
https://doi.org/10.1016/j.jenvman.2013.10.008
458 J. L. Vest et al.
Drewien RC, Brown WM, Kendall WL (1995) Recruitment of rocky mountain greater sandhill
cranes and comparison with other crane populations. J Wildl Manage 59:339–356
Drum RG, Loesch CR, Carrlson K et al (2015) Assessing the biological benefits of the USDA-
Conservation Reserve Program (CRP) for waterfowl and grassland passerines in the Prairie
Pothole Region of the United States: spatial analyses for targeting CRP to maximize benefits
for migratory birds. Final Report for USDA-FSA Agreement: 12-IA-MRE-CRP-TA
DU (2021) Ducks unlimited conservation priorities. Ducks Unlimited, Inc., Memphis, Tennessee
Dugger BD, Moore ML, Finger RS e t al (2007) True metabolizable energy for seeds of common
moist-soil plant species. J Wildl Manage 71:1964–1967
Duncan BL, Hansen R, Hambrecth K et al (2019) Cattle grazing for invasive Phragmites australis
(common reed) management in Northern Utah wetlands. Utah State University Extension Fact
Sheet NR/Wildlands/2019-01pr, Logan, Utah, USA
Earnst SL (1994) Tundra swan habitat preferences during migration in North Dakota. J Wildl
Manage 58:546–551
Eldridge J (1992) 13.2.14 management of habitat for breeding and migrating shorebirds in the
Midwest. Waterfowl Management Handbook, US Fish and Wildlife Service Leaflet vol 11, p 7
Emery RB, Howerter DW, Armstrong LM et al (2005) Seasonal variation in waterfowl nesting
success and its relation to cover management in the Canadian prairies. J Wildl Manage 69:1181–
1193
Epperson WL, Eadie JM, Marcum DB et al (1999) Late season hay harvest provides habitat for
marshland birds. Calif Agric 53:12–17
Euliss NH, LaBaugh JW, Fredrickson LH et al (2004) The wetland continuum: a conceptual
framework for interpreting biological studies. Wetlands 24:448–458
Evans-Peters GR, Dugger BD, Petrie MJ (2012) Plant community composition and waterfowl food
production on wetland reserve program easements compared to those on managed public lands
in western Oregon and Washington. Wetlands 32:391–399
Fields S, Barnes K (2019) Grassland assessment of North American Great Plains Migratory Bird
Joint Ventures. Unpublished report. Prairie Pothole Joint Venture, Great Falls, MT USA. http:/
/ppjv.org/assets/docs/Great_Plains_Grassland_Assessment_Final_Report.pdf
Fink D, Auer T, Johnston A et al (2020) Modeling avian full annual cycle distribution and population
trends with citizen science data. Ecol Appl 30:02056. https://doi.org/10.1002/eap.2056
Flake LD, Petersen GL, Tucker WL (1977) Habitat relationships of breeding waterfowl on stock
ponds in northwestern South Dakota. Proc South Dakota Acad Sci 56:135–151
Fleskes JP, Gregory CJ (2010) Distribution and dynamic of waterbird habitat during spring in
Southern Oregon-Northeastern California. West North Am Nat 70(1):26–38. https://doi.org/10.
3398/064.070.0104
Fleskes JP, Mauser DM, Yee JL, Blehert D, Yarris GS (2010) Flightless and post-molt survival and
movements of female mallards molting in Klamath Basin. Waterbirds 33(2):208–220. https://
doi.org/10.1675/063.033.0209
Flickinger EL (1981) Wildlife mortality at petroleum pits in Texas. J Wildl Manage 45:560–564
Flickinger EL, Bunck CM (1987) Number of oil-killed birds and fate of bird carcasses at crude oil
pits in Texas. Southwestern Nat 32(3):377–381. https://doi.org/10.2307/3671456
Forman KJ, Madsen CR, Hogan MJ (1996) Creating multiple purpose wetlands to enhance livestock
grazing distribution, range condition, and waterfowl production in western South Dakota. In:
Schaack J, Anderson SS (eds) Water for agriculture and wildlife and the environment: win-win
opportunities. U.S. Committee on Irrigation and Drainage, Denver, CO, USA, pp185–192
Fox AD, Flint PL, Hohman WL et al (2014) Waterfowl habitat use and selection during the remigial
moult period in the northern hemisphere. Wildfowl 4:131–168
Fredrickson LH, Taylor TS (1982) Management of seasonally flooded impoundments for wildlife.
Resource Publication 148. US Department of Interior, Fish and Wildlife Service, Washington,
DC, USA
Friend M, McLean RG, Dein FJ (2001) Disease emergence in birds: challenges of the twenty-first
century. Auk 118:290–303
13 Waterfowl and Wetland Birds 459
Fynn R, Jackson J (2022) Grazing management on commercial cattle ranches: incorporating
foraging ecology and biodiversity conservation principles. Rangelands 44:136–147
Gerber BD, Kendall WL, Hooten MB et al (2015) Optimal population prediction of sandhill crane
recruitment based on climate-mediated habitat limitations. J Anim Ecol 84:1299–1310
Gilbert DW, Anderson DR, Ringleman JK et al (1996) Response of nesting ducks to habitat
management on the Monte Vista National Wildlife Refuge, Colorado. Wildl Monogr 131
Gitz D, Brauer D (2016) Trends in playa inundation and water storage in the Ogallala Aquifer on
the Texas High Plains. Hydrology 3:31. https://doi.org/10.3390/hydrology3030031
Gjersin FM (1975) Waterfowl production in relation to rest-rotation grazing. Rangel Ecol Manag
28:37–42
Grafton RQ, Williams J, Perry CJ et al (2018) The paradox of irrigation efficiency. Science 361:748–
750
Grant TA, Flanders-Wanner B, Shaffer TL et al (2009) An emerging crisis across Northern Prairie
refuges: prevalence of invasive plants and a plan for adaptive management. Ecol Restor 27:58–65
Gray PN, Bolen EG (1987) Seed reserves in tailwater pits of playa lakes in relation to waterfowl
management. Wetlands 7:11–23
Gray RL, Teels BM (2006) Wildlife and fish conservation through the farm bill. Wildl Soc Bull
34:906–913
Gray MJ, Hagy HM, Nyman JA et al (2013) Management of wetlands for wildlife. In: Anderson
JT, Davis CA (eds) Wetland techniques. Springer, Dordrecht, pp 121–180
Greenwood RJ, Sargent AB, Johnson DG et al (1995) Factors associated with duck nest success in
the Prairie Pothole Regin of Canada. Wildl Monogr 128:57
Gruntorad MP, Graham KA, Arcilla N et al (2021) Is hay for the birds? Investigating landowner
willingness to time hay harvests for grassland bird conservation. Animals 11:1030. https://doi.
org/10.3390/ani11041030
Hagy HM, Brasher MG, Flesks JP et al (in review) Important geographies for migrating and
wintering waterfowl. In: Ballard, BM, Brasher MG, Fleskes JP (eds) Waterfowl in winter. Texas
A&M University Press, College Station, USA
Haig SM, Mehlman DW, Oring LW (1998) Avian movements and wetland connectivity in landscape
conservation. Conserv Biol 12:749–758
Haig SM, Murphy SP, Matthews JH et al (2019) Climate-altered wetlands challenge waterbird use
and migratory connectivity in arid landscapes. Sci Rep 9:4666. https://doi.org/10.1038/s41598-
019-41135-y
Hansen AJ, Rasker R, Maxwell B et al (2002) Ecological causes and consequences of demographic
change in the New West. Bioscience 52:151–162. https://doi.org/10.1641/0006-3568(2002)052
[0151:ECACOD]2.0.CO;2
Harris SW, Marshall WH (1963) Ecology of water-level manipulations on a northern marsh. Ecology
44:331–343
Harrison RB, Jones WM, Clark D et al (2017) Livestock grazing in intermountain depressional
wetlands: effects on breeding waterfowl. Wetl Ecol Manag 25:471–484
Hartman CA, Oring LW (2009) Reproductive success of long-billed curlews (Numenius americanus)
in northeastern Nevada hay fields. Auk 126:420–430
Haukos DA, Smith LM (1993) Moist-soil management of playa lakes for migrating and wintering
ducks. Wildl Soc Bull 21:288–298
Haukos DA, Smith LM (1994) The importance of playa wetlands to biodiversity of the Southern
High Plains. Landsc Urban Plan 28:83–98
Haukos DA, Miller MR, Orthmeyer DL et al (2006) Spring migration of northern pintails from
Texas and New Mexico, USA. Waterbirds 29:127–241
Hays QR, Tredennick AT, Carlisle JD et al (2021) Spatially explicit assessment of sandhill crane
exposure to potential transmission line collision risk. J Wildl Manage 85(7):1440–1449. https:/
/doi.org/10.1002/jwmg.22100
Helmers DL (1992) Shorebird management manual. Western Hemisphere Shorebird Reserve
Network, Manomet, MA, 58pp
460 J. L. Vest et al.
Henkel JR, Taylor CM (2015) Migration strategy predicts stopover ecology in shorebird on the
northern Gulf of Mexico. Anim Migr 2:63–75
Herkert JR, Reinking DL, Wiedenfeld DA et al (2003) Effects of prairie fragmentation on the nest
success of breeding birds in the midcontinental United States. Conserv Biol 17:587–594
Higgins KF, Naugle DE, Forman, KJ (2002) A case study of changing land use practices in
the Northern Great Plains, U.S.A.: an uncertain future for waterbird conservation. Waterbirds
25(Special Publication 2):42–50. https://www.jstor.org/stable/1522450
Hillhouse HL (2019) Impacts of cattle grazing on seed production in Rainwater Basin wetlands.
Wetl Ecol Manag 27:141–147
Hise C, Obermeyer B, Ahlering M et al (2020) Site wind right: identifying low-impact wind devel-
opment areas in the central United States. Unpublished manuscript. https://doi.org/10.1101/
2020.02.11.943613
Hoekman ST, Mills LS, Howerter DW et al (2002) Sensitivity analyses of the life cycle of
midcontinent mallards. J Wildl Manage 66:883–900
Hoffman GR, Stanley LD (1978) Effects of cattle grazing on shore vegetation of fluctuating water
level reservoirs. J Range Manag 31:412–416
Hohman WL, Ankney CD, Gordon DH (1992) Ecology and management of postbreeding water-
fowl. In: Batt BDJ, Afton AD, Anderson MG et al (eds) Ecology and management of breeding
waterfowl. University of Minnesota Press Minneapolis, Minnesota, pp 128–189
Hohman WL, Lindstrom EB, Rashford BS et al (2014) Opportunities and challenges to waterfowl
habitat conservation on private land. Wildfowl (special Issue) 4:368–406
Holechek JL, Pieper RD, Herbel CH (2011) Range management: principles and practices, 6th
edition. Pearson Education, Inc. New York, NewYork, USA
Holechek JL, Valdez R, Schemnitz SD et al (1982) Manipulation of grazing to improve or maintain
wildlife habitat. Wildl Soc Bull 10:204–210
Horn DJ, Phillips ML, Koford RR et al (2005) Landscape composition, patch size, and distance to
edges: interactions affecting duck reproductive success. Ecol Appl 15:1367–1376
Horns JJ, Adler FR, ¸Sekercio ˘glu CH (2018) Using opportunistic citizen science data to estimate
avian population trends. Biol Conserv 221:151–159
Hovick TJ, Carroll JM, Elmore RD et al (2017) Restoring fire to grasslands is critical for migrating
shorebird populations. Ecol Appl 27:1805–1814
Hubbard DE (1988) Using your wetland as forage. South Dakota State University, Cooperative
Extension, Extension Fact Sheets 1023. https://openprairie.sdstate.edu/extension_fact/1023
Hudson M (1983) Waterfowl production on three age-classes of stock ponds in Montana. J Wildl
Manage 47:112–117
Hudson M, Francis CM, Campbell KJ et al (2017) The role of the North American breeding bird
survey in conservation. Condor 119:526–545
Humburg DD, Anderson MG, Brasher MG (2018) Implementing the 2012 North American
Waterfowl Management Plan revision: populations, habitat, and people. J Wildl Manage
82:275–286
Iglecia M, Winn B (2021) A shorebird management manual. Manomet, Massachusetts, USA.
https://www.manomet.org/wp-content/uploads/2021/01/Iglecia_and_Winn_2021_AShorebir
dManagementManual-012021-web.pdf
Ignatiuk JB, Duncan DC (2001) Nest success of ducks on rotational and season-long grazing systems
in Saskatchewan. Wildl Soc Bull 29:211–217
Ivey G (2011) Making a home for Greater Sandhill Cranes on private lands. Audubon
California. Working Land Series, Summer 2011. Audubon California, Oakland, CA,
USA. https://www.academia.edu/535531/Audubon_Californias_Working_Lands_Series_Mak
ing_a_home_for_Greater_Sandhill_Cranes_on_private_lands
Ivey GL, Dugger BD (2008) Factors influencing nest success of Greater Sandhill Cranes at Malheur
National Wildlife Refuge, Oregon. Waterbirds 31:52–61
13 Waterfowl and Wetland Birds 461
Ivey GL, Herziger CP (2006) Intermountain West waterbird conservation plan, version 1.2. Water-
bird Conservation for the Americas Initiative, US Fish and Wildlife Service, Portland, Oregon,
USA
Ivey GL, Cornely JE, Ehlers BD (1998) Carp impacts on waterfowl at Malheur National Wildlife
Refuge, Oregon. In: Transactions of the 63rd North American wildlife and natural resources
conference vol. 63. pp 66–74
IWJV (2013) Intermountain West Joint Venture 2013 Implementation plan. Intermountain West
Joint Venture, Missoula, Montana. https://iwjv.org/resource/iwjv-2013-implementation-plan-
entire-plan/
Jefferies RL, Rockwell RF, Abraham KF (2003) The embarrassment of riches: agricultural food
subsidies, high goose numbers, and loss of arctic wetlands—continuing saga. Environ Rev
11:193–232
Johnson WC (2019) Ecosystem services provided by prairie wetlands in northern rangelands.
Rangeland 41:44–48
Johnson GD, Stephens SE (2011) Wind power and biofuels: a green dilemma for wildlife conser-
vation. In: Naugle DE (ed) Energy development and wildlife conservation in Western North
America. Island Press, Washington, D.C., USA, pp 131–155
Johnson DH, Gibbs JP, Herzog M (2009) A sampling design framework for monitoring secretive
marshbirds. Waterbirds 32:203–215
Johnson WC, Werner B, Guntenspergen GR et al (2010) Prairie wetland complexes as landscape
functional units in a changing climate. Bioscience 60:128–140
Kadlec JA, Smith LM (1989) The Great Basin marshes. In: Smith LM, Pederson RL, Kaminski
RM (eds) Habitat management for migrating and wintering waterfowl in North America. Texas
Tech University Press, Lubbock, Texas, pp 451–474
Kadlec JA, Smith LM (1992) Habitat management for breeding areas. In: Batt BDJ, Afton AD,
Anderson MG et al (eds) Ecology and management of breeding waterfowl. University of
Minnesota Press, Minneapolis, MN, USA
Kahara SN, Skalos D, Madurapperuma B, Hernandez K (2021) Habitat quality and drought effects
on breeding mallard and other water populations in California. J Wildl Manage. https://doi.org/
10.1002/jwmg.22133
Kapnick S, Hall A (2012) Causes of recent changes in western North American snowpack. Clim
Dyn 38:1885–1899
Kemink KM, Adams VM, Pressey RL (2021) Integrating dynamic processes into waterfowl
conservation prioritization tools. Divers Distrib 27:585–601
Kendy E (2006) Impacts of changing land use and irrigation practices on western wetlands. Nat
Wetlands Newsletter 28:12–27
Kent CM, Ramey AM, Ackerman JT (2022) Spatiotemporal changes in influenza A virus prevalence
among wild waterfowl inhabiting the continental United States throughout the annual cycle. Sci
Rep 12:13083. https://doi.org/10.1038/s41598-022-17396-5
King SL, Laubhan MK, Tashjian P et al (2021) Wetland conservation: challenges related to water
law and farm policy. Wetlands 41:54. https://doi.org/10.1007/s13157-021-01449-y
Kirby DR, Krabbenhoft KD, Sedivec KK et al (2002) Wetlands in Northern Plains prairies:
benefitting wildlife and livestock. Rangelands 24:22–25
Koons DN, Gunnarson G, Schmutz JA et al (2014) Drivers of waterfowl population dynamics: from
teal to swans. Wildfowl (2014 Special Issue) 4:169–191
Kraai KJ (2003) Late winter feeding habits, body condition, and feather molt intensity of
female mallards utilizing livestock ponds in northeast Texas. Thesis, Texas A&M University,
Commerce, Texas
Krapu GL, Pietz PJ, Brandt DA et al (2000) Factors limiting mallard brood survival in Prairie
Pothole landscapes. J Wildl Manage 64:553–561
Krapu GL, Reinecke KJ (1989) Foraging ecology and nutrition. In: Batt BDJ, Afton AD, Anderson
MG et al (ed) Ecology and Management of Breeding Waterfowl. University of Minnesota Press,
Minneapolis, Minnesota pp 1–30
462 J. L. Vest et al.
Krausman PR, Naugle DE, Frisina MR et al (2009) Livestock grazing, wildlife habitat, and rangeland
values. Rangelands 31:15–19
Kushlan JA, Steinkamp MJ, Parson KC et al (2002) Waterbird conservation for the Americas:
the Northern American waterbird conservation plan, version 1. Waterbird Conservation for the
Americas, Washington, DC, USA, 78 pp. https://www.fws.gov/partner/north-american-waterb
ird-conservation-plan
Landgrave R, Moreno-Casasola P (2012) Evaluación cuantitiva de la pérdida de humedales en
México. Inv Ambiental 4(1):19–35. https://proyectopuente.com.mx/wp-content/uploads/2019/
05/121-707-1-pb.pdf
Lange CJ, Ballard BM, Collins DP (2018) Impacts of wind turbines on redheads in the Laguna
Madre. J Wildl Manage 82(3):531–537. https://doi.org/10.1002/jwmg.21415
Lark TJ, Salmon M, Gibbs HK (2015) Cropland expansion outpaces agricultural and biofuel
polices in the United States. Environ Res Lett 10:044003. https://doi.org/10.1088/1748-9326/
10/4/044003
Lark TJ, Spawn SA, Bougie M, Gibbs HK (2020) Cropland expansion in the United states produces
marginal yields at high costs to wildlife. Nat Commun 11:4295. https://doi.org/10.1038/s41467-
020-18045-z
Laubhan MK, Gammonley JH (2000) Density and foraging habitat selection of waterbirds breeding
in the San Luis Valley of Colorado. J Wildl Manage 64:808–819
Lemly AD, Kingsford RT, Thompson JR (2000) Irrigated agriculture and wildlife conservation:
conflict on a global sale. Environ Manag 25(5):485–512. https://doi.org/10.1007/s00267991
0039.pdf
Lincoln FC (1930) Calculating waterfowl abundance on the basis of banding returns. U.S.
Department of Agriculture Circular No. 118
Lipsey MK, Naugle DE (2017) Precipitation and soil productivity explain effects of grazing on
grassland songbirds. Rangel E col Manag 70:331–340
Littlefield CD (1999) Greater Sandhill Crane productivity on privately owned wetlands in Eastern
Oregon. Western Birds 30:206–210
Littlefield CD, Paullin DG (1990) Effects of land management on nesting success of Sandhill Cranes
in Oregon. Wildl Soc Bull 18:63–65
Loesch CR, Reynolds RE, Hansen LT (2012) An assessment of re-directing breeding waterfowl
conservation relative to predictions of climate change. J Fish Wildl Manag 3:1–22
Loesch CR, Walker JA, Reynolds RE et al (2013) Effect of wind energy development on breeding
duck densities in the Prairie Pothole Region. J Wildl Manage 77(3):587–598
Londe DW, Dvorett D, Davis CA et al (2022) Inundation of depressional wetland declines under a
changing climate. Clim Change 172:27
Lovvorn JR, Crozier ML (2022) Duck use of saline wetlands created by irrigation in a semiarid
landscape. Wetlands 42:4. https://doi.org/10.1007/s13157-021-01525-3
Lovvorn JR, Hart EA (2004) Irrigation, salinity, and landscape patterns of natural palustrine
wetlands. In: McKinstry MC, Hubert WA, Anderson SH (eds) Wetland and riparian areas of the
Intermountain West: ecology and management. University of Texas Press, Austin, pp 105–129
Luttschwager KA, Higgins KF, Jenks JA (1994) Effects of emergency haying on duck nesting in
conservation reserve program fields, South Dakota. Wildl Soc Bull 22:403–408
Ma Z, Cai Y, Li B et al (2010) Managing wetland habitats for waterbirds: an international perspective.
Wetlands 30:15–27
Mackell DA, Casazza ML, Overton CT et al (2021) Migration stopover ecology of Cinnamon Teal
in western North America. Ecol Evol 1–14.https://doi.org/10.1002/ece3.8115
Martin EM, Carney SM (1977) Population ecology of the mallard: IV. A review of duck hunting
regulations, activity, and success, with special reference to the mallard. U.S. Fish and Wildlife
Service, Resource Publication 130, Washington, D.C., USA
13 Waterfowl and Wetland Birds 463
Martin FW, Pospahala RS, Nichols JD (1979) Assessment and population management of North
American migratory birds. In: Cairns J, Patil GP, Walters WE (eds) Environmental biomon-
itoring, assessment, prediction, and management-certain case studies and related quantitative
issues. International Cooperative Publishing House, Fairland, Maryland, pp 187–239
Mason CD, Whiting RM Jr, Conway WC (2013) Time-activity budgets of waterfowl wintering on
livestock ponds in northeast Texas. Southeast Nat 12:757–768
Masto NM, Kaminski RM, Prince HH (2022) Hemi-marsh concept prevails? Kaminski and Prince
(1981) revisited. J Wildl Manage 86:e22301
May SM, Naugle DE, Higgins KF (2002) Effects of land use on nongame wetland birds in western
South Dakota stock ponds, U.S.A. Waterbirds 25(Special Publication):51–55
McCauley LA, Anteau MH, van der Burg MP et al (2015) Land use and wetland drainage affect water
levels and dynamics of remaining wetlands. Ecosphere 6:92. https://doi.org/10.1890/ES14-004
94.1
Mckenna OP, Kucia SR, Mushet DM et al. (2019) Synergistic interaction of climate and land-use
drivers alter function of North American, Prairie Pothole Wetlands. Sustain 11:6581. https://
doi.org/10.3390/su11236581
McLean K, Mushet D, Sweetman J (2022) Climate and land use driven ecosystem homogenization
in the Prairie Pothole Region. Water 14:3106. https://doi.org/10.3390/w14193106
McIntyre NE, Collins SD, Heintzman LJ et al (2018) The challenge of assaying landscape connec-
tivity in a changing world: a 27-year case study in the southern Great Plains (USA) playa
network. Ecol Indic 91:607–616
McKenna OP, Mushet DM, Rosenberry DO et al (2017) Evidence for a climate-induced ecohy-
drological state shift in wetland systems of the southern Prairie Pothole Region. Clim Change
145:273–287
McKenna OP, Mushet DM, Kucia SR et al (2021a) Limited shifts in the distribution of migratory
bird breeding habitat density in response to future changes in climate. Ecol Appl 31:e02428
McKenna OP, Renton DA, Mushet DM et al (2021b) Upland burning and grazing as strategies to
offset climate-change effects on wetlands. Wetl Ecol Manag 29:193–208
McKinstry MC, Caffrey P, Anderson SH (2001) The importance of beaver to wetland habitats and
waterfowl in Wyoming. J Am Water Resour Assoc 37(6):1571–1577
McKinstry MC, Hubert WA, Anderson SH (eds) (2004) Wetland and riparian areas of the
Intermountain West. University of Texas Press, Austin, Texas, USA
McWethy DB, Austin JE (2009) Nesting ecology of Greater Sandhill Cranes (Grus canadensis
tabida) in riparian and palustrine wetlands of Eastern Idaho. Waterbirds 32:106–115
Miller MR, Takekawa JY, Fleskes JP et al (2005) Spring migration of northern pintails from
California’s Central Valley wintering area tracked with satellite telemetry: routes, timing, and
destinations. Can J Zool 83:1314–1332
Mitchell CD, Eicholz MW (2020) Trumpeter swan (Cygnus buccinator), version 1.0. In: Rodewald
PG (ed) Birds of the world. Cornell Lab of Ornithology, Ithaca, NY, USA. https://doi.org/10.
2173/bow.truswa.01
Mitsch WJ, Gosselink JG (2015) Wetlands, 5th edn. Wiley, Hoboken, New Jersey, USA
Moon JA, Haukos DA (2006) Survival of female northern pintails wintering in the Playa Lakes
Region of northwestern Texas. J Wildl Manage 70:777–783
Moon JA, Haukos DA (2008) Habitat use by northern pintails wintering in the Playa Lakes Region.
Proc Southeast Assoc Fish Wildlife Agencies 62:82–87
Moon JA, Haukos DA (2009) Factors affecting body condition of Northern Pintails wintering in
the Playa Lakes Region. Waterbirds 32:87–95
Moon JA, Lehnen SE, Metzger KL et al (2021) Projected impact of sea-level rise and urbanization
on mottled duck (Anas fulvigula) habitat along the Gulf Coast of Louisiana and Texas through
2100. Ecol Ind 132:108276. https://doi.org/10.1016/j.ecolind.2021.108276
Moore JM (2016) Recent desiccation of western Great Basin saline lakes: lessons from Lake Abert,
Oregon, U.S.A. Sci Total Environ 554–555:142–154
464 J. L. Vest et al.
Moore MA, McEvoy J (2022) “In Montana, you’re only a week away from a drought”: Rancher’s
perspectives on flood irrigation and beaver mimicry as drought mitigation strategies. Rangelands
44:258–269
Mote PW, Li S, Lettenmaier DP et al (2018) Dramatic declines in snowpack in the western US. NPJ
Clim Atmos Sci 1:2
Moulton CE, Carlisle JD, Knetter SJ et al (2022) Importance of flood irrigation for foraging colonial
waterbirds. J Wildl Manage 86:e22288
Mowbray TB, Ely CR, Sedinger JS et al (2020) Canada goose (Branta canadensis), version 1.0. In
Rodewald PG (ed) Birds of the world. Cornell Lab of Ornithology, NY, USA. https://doi.org/
10.2173/bow.cangoo.01
Mundinger JG (1976) Waterfowl response to rest-rotation grazing. J Wildl Manage 40:60–68
Murkin HR, Murkin EJ, Ball JP (1997) Avian habitat selection and prairie wetland dynamics: a
10-year experiment. Ecol Appl 7:1144–1159
Murphy RK, Schindler DJ, Crawford RD (2004) Duck nesting on rotational and continuous grazed
pastures in North Dakota. Prairie Nat 36:83–94
Murphy RK, Dwyer JF, Mojica EK et al (2016) Reactions of sandhill cranes approaching a marked
transmission power line. J Fish Wildl Manag 7(2):480–489. https://doi.org/10.3996/052016-
JFWM-037
Naugle DE, Higgins KF, Bakker KK (2000) A synthesis of the effects of upland management
practices on waterfowl and other birds in the Northern Great Plains of the U.S. and Canada.
College of Natural Resources, University of Wisconsin-Stevens Point, WI. Wildlife Technical
Report 1, 28p
Navarrete L, Griffis-Kyle KL (2014) Sandhill crane collisions with wind turbines in Texas. In:
Proceedings of the North Americna Crane workshop vol 12. pp 65–67. https://digitalcommons.
unl.edu/nacwgproc/380/
NAWMP (2018) North American Waterfowl Management Plan (NAWMP) Update: connecting
people, waterfowl, and wetlands. Canadian Wildlife Service, U.S. Fish and Wildlife Service,
and Secretaria de Medio Ambiente y Recursos Naturales. https://www.fws.gov/partner/north-
american-waterfowl-management-plan
Neudecker GA, Duvall AL, Stutzman JW (2011) Community-based landscape conservation: a
roadmap for the future. In: Naugle DE (ed) Energy development and wildlife conservation in
western North America. Island Press, Washington, DC, USA
Nichols JD, Kendall WL, Boomer GS (2019) Accumulating evidence in ecology: once is not enough.
Ecol Evol 9:13991–14004
Nieman DJ, Didiuk AB, Smith JR (2000) Status of Canada Geese of the Canadian prairies. In:
Dickson KM (ed) Towards conservation of the diversity of Canada geese (Branta canadensis).
Occasional Paper 103. Canada Wildlife Service, Ottawa, Ontario, pp 139–150
Niemuth ND, Solberg SW (2003) Response of waterbirds to number of wetlands in the Prairie
Pothole Region of North Dakota, U.S.A. Waterbirds 26:233–238
Niemuth ND, Dahl AL, Estey ME et al (2007) Representation of landcover along breeding bird
survey routes in the Northern Plains. J Wildl Manage 71:2258–2265
Niemuth ND, Reynolds RE, Granfors DE et al (2009) Landscape-level planning for conservation
of wetland birds in the U.S. Prairie Pothole Region. In: Millspaugh JJ, Thompson FR III (eds)
Models for planning wildlife conservation in large landscapes. Elsevier Science, pp 533–560
Niemuth ND, Wangler B, Reynolds RE (2010) Spatial and temporal variation in wet area of wetlands
in the Prairie Pothole Region of North Dakota and South Dakota. Wetlands 30:1053–1064
Niemuth ND, Estey ME, Reynolds RE (2012) Factors influencing presence and detection of breeding
shorebirds in the Prairie Pothole Region of North Dakota, South Dakota, and Montana, USA
Niemuth ND, Fleming KK, Reynolds RE (2014) Waterfowl conservation in the US Prairie Pothole
Region: confronting the complexities of climate change. PLoS ONE 9(6):e100034. https://doi.
org/10.1371/journal.pone.0100034
Niemuth ND, Barnes KW, Tack JD et al (2022) Past is prologue: historic landcover patterns predict
contemporary grassland loss in the U.S Northern Great Plains. Landsc Ecol 37:3011–3027
13 Waterfowl and Wetland Birds 465
NRCS (2021) A framework for conservation action in the Great Plains grassland biome. USDA
Natural Resources Conservation Service, Working Lands for Wildlife
Null SE, Wurtsbaugh WA (2020) Water development, consumptive water uses, and Great Salt Lake.
In: Baxter B, Butler J (eds) Great Salt Lake biology a terminal lake in a time of change. Springer,
Cham, Switzerland, pp 1–21
Osnas EE, Boomer GS, Devries JH et al (2021) Decision-support framework for linking regional-
scale management actions to continental-scale conservation of wide-ranging species. US
Geological Survey Open-File Report 2020–1084.https://doi.org/10.3133/ofr20201084
Ott JP, Hanberry BB, Khalil M, Paschke MW, van der Burg MP, Prenni AJ (2021) Energy devel-
opment and production in the Great Plains: implications and mitigation opportunities. Rangel
Ecol Manag 78:257–272. https://doi.org/10.1016/j.rama.2020.05.003
Paveglio FL, Kilbride K (2000) Response of vegetation to control of reed canarygrass in seasonally
managed wetlands of southwestern Washington. Wildl Soc Bull 28:730–740
Pearse AT, Brandt DA, Krapu GL (2016) Wintering sandhill crane exposure to wind energy devel-
opment in the central and southern Great Plains, USA. Condor 118:391–401. https://doi.org/10.
1650/CONDOR-15-99.1
Pearse AT, Metzger KL, Brandt DA et al (2021) Migrating whooping cranes avoid wind-energy
infrastructure when selecting stopover habitat. Ecol Appl 30(5):e02324. https://doi.org/10.1002/
eap.2324
Pearse AT, Anteau MJ, van der Burg MP et al (2022) Reassessing perennial cover as a driver of
duck nest survival in the Prairie Pothole Region. J Wildl Manage 86:e22227. https://doi.org/10.
1002/jwmg.22227
Pearson J, Dunham J, Bellmore JR et al (2019) Modeling control of common carp (Cyprinus carpio)
in a shallow lake-wetland system. Wetl Ecol Manag 27:663–682
Pederson RL, Jorde DG, Simpson SG (1989) Northern Great Plains. In: Smith LM, Pederson
RL, Kaminski RM (eds) Habitat management for migrating and wintering waterfowl in North
America. Texas Tech University Press, Lubbock, Texas
Phillips ML, Clark WR, Sovada MA et al (2003) Predator selection of prairie landscape features
and its relation to duck nest success. J Wildl Manage 67:104–114
PHJV (2021) Prairie Habitat Joint Venture implementation plan 2021–2025: the Prairie Parklands.
Report to the Prairie Habitat Joint Venture. Environment Canada, Edmonton, Alberta
PIF (2021) Avian conservation assessment database, version 2021. Partners in Flight, Available at
http://pif.birdconservancy.org/ACAD
Powers LC, Glimp HA (1996) Impacts of livestock on shorebirds: a review and application to
shorebirds of the western Great Basin. Int Wader Stud 9:55–63
PPJV (2017) Implementation plan. Prairie Pothole Joint Venture, Lakewood, Colorado. https://ppjv.
org/assets/pdf/2017-PPJV-implementation-plan.zip
Preston TM, Ray AM (2017) Effects of energy development on wetland plants and macroinvertebrate
communities in Prairie Pothole Region wetlands. J Freshw Ecol 32(1):29–34. https://doi.org/
10.1080/02705060.2016.1231137
Ramirez P Jr (2010) Bird mortality in oil field wastewater disposal facilities. Environ Manage
46:820–826. https://doi.org/10.1007/s00267-010-9557-4
Raquel AJ, Devries JH, Howerter DW et al (2016) Timing of nesting of upland-nesting ducks in
the Canadian prairies and its relation to spring wetland conditions. Can J Zool 94:575–581
Renner RW, Reynolds RE, Batt BDJ (1995) The impact of haying conservation reserve program
lands on productivity of ducks nesting in the Prairie Pothole Region of North Dakota and South
Dakota. In: Transactions of the 60th North American wildlife and natural resources conference,
vol 60. pp 221–229
Reynolds RE, Blohm RJ, Nichols JD et al (1995) Spring-summer survival rates of yearlings versus
adult mallard females. J Wildl Manage 59:691–696
Reynolds RE, Shaffer TL, Renner RW et al (2001) Impact of the conservation reserve program on
duck recruitment in the U.S. Prairie Pothole Region. J Wildl Manage 65:765–780
466 J. L. Vest et al.
Reynolds RE, Shaffer TL, Loesch CR (2006) The farm bill and duck production in the Prairie
Pothole Region: increasing the benefits. Wildl Soc Bull 34:963–974
Richmond OMW, Tecklin J, Beissinger SR (2012) Impact of cattle grazing in the occupancy of a
cryptic, threatened rail. Ecol Appl 22:1655–1664
Riecke TV, Sedinger BS, Arnold TW et al (2022a) A hierarchical model for jointly assessing
ecological and anthropogenic i mpacts on animal demography. J Anim Ecol 91:1612–1626
Riecke TV, Lohman MG, Sedinger BS (2022b) Density-dependence produces spurious relationships
among demographic parameters in a harvested species. J Anim Ecol 00:1–12. https://doi.org/
10.1111/1365-2656.13807
Ringelman KM, Walker J, Ringelman JK et al (2018) Temporal and multi-spatial environmental
drivers of duck nest survival. Auk 135:486–494
Rischette AC, Geaumont BA, Elmore RD et al (2021) Duck nest density and survival in post-
conservation reserve program lands. Wildl Soc Bull 45:630–637
Roberts A, Scarpignato AL, Huysman A et al (2023) Migratory connectivity of North American
waterfowl across administrative flyways. Ecol Appl 2023:e2788. https://doi.org/10.1002/eap.
2788
Rohal CB, Cranney C, Hazelton ELG et al (2019) Invasive Phragmites australis management
outcomes and native plant recovery are context dependent. Ecol Evol 9:13835–13849
Rosenberg KV, Dokter AM, Blancher PJ et al (2019) Decline of North American avifauna. Science
366:120–124
Rumble MA, Flake LD (1983) Management considerations to enhance use of stock ponds by
waterfowl broods. Rangel Ecol Manag 36:691–694
RWBJV (2013) Rainwater Basin Joint Venture waterfowl plan: a regional contribution to the North
American Waterfowl Management Plan and the Rainwater Basin Joint Venture Implementation
Plan. Rainwater Basin Joint Venture, Grand Island, Nebraska, USA. https://www.rwbjv.org/wp-
content/uploads/Rainwater-Basin-Joint-Venture-Waterfowl-Plan-2013.pdf
Sargeant AB, Raveling DG (1992) Mortality during the breeding season. In: Batt BDJ, Afton
AD, Anderson MG et al (eds) Ecology and management of breeding waterfowl. University of
Minnesota Press Minneapolis, Minnesota, pp 396–422
Sauer JR (1999) Marsh birds and the North American breeding bird survey: judging the value of
a landscape level survey for habitat specialist species with low detection rates. In: Ribic CA,
Lewis SJ, Melvin S et al (eds) Proceedings of the marsh bird monitoring workshop. U.S. Fish
and Wildlife Service and U.S. Geological Survey, Laurel, MD, USA. p 43
Sauer JR, Fallon JE, Johnson R (2003) Use of North American breeding bird survey data t o estimate
population change for bird conservation regions. J Wildl Manage 67:372–389
Schultz BD, Hubbard DE, Jens JA et al (1994) Plant and waterfowl responses to cattle grazing in
two South Dakota semipermanent wetlands. Proc South Dakota Acad Sci 73:121–134
Seamans ME (2022) Status and harvests of sandhill cranes: mid-continent, Rocky Mountain, Lower
Colorado River Valley and Eastern Populations. Administrative Report, U.S. Fish and Wildlife
Service, Lakewood, Colorado
Sedinger JS, Alisauskas RT (2014) Cross-seasonal effects and the dynamics of waterfowl
populations. Wildfowl 4:277–304
Shaffer JA, Igl LD, Johnson DH et al (2019a) The effects of management practices on grass-
land birds–Marbled Godwit (Limosa fedora), Chapter H. In: Johnson DH, Igl LD, Shaffer JA
et al (eds) The effects of management practices on grassland birds. U.S. Geological Survey
Professional Paper 1842, 9p. https://doi.org/10.3133/pp1842H
Shaffer JA, Igl LD, Johnson DH et al (2019b) The effects of management practices on grassland
birds–Willet (Tringa semipalmata inornata), Chapter I. In: Johnson DH, Igl L D, Shaffer JA
et al (eds) The effects of management practices on grassland birds. U.S. Geological Survey
Professional Paper 1842, 9p. https://doi.org/10.3133/pp1842I
Shaffer JA, Igl LD, Johnson DH et al (2019c) The effects of management practices on grassland
birds–Wilson’s Phalarope (Phaloropus tricolor), Chapter J. In: Johnson DH, Igl LD, Shaffer
13 Waterfowl and Wetland Birds 467
JA et al (eds) The effects of management practices on grassland birds. U.S. Geological Survey
Professional Paper 1842, 10p. https://doi.org/10.3133/pp1842J
Shaffer JA, Igl LD, Johnson DH et al (2019d) The effects of management practices on grassland
birds–Long-billed Curlew (Numenius americanus), Chapter G. In: Johnson DH, Igl LD, Shaffer
JA et al (eds) The effects of management practices on grassland birds. U.S. Geological Survey
Professional Paper 1842, 12p. https://doi.org/10.3133/pp1842G
Shaffer JA, Igl LD, Johnson DH et al (2019e) The effects of management practices on grassland
birds–Upland Sandpiper (Bartramia longicauda), Chapter F. In: Johnson DH, Igl LD, Shaffer
JA et al (eds) The effects of management practices on grassland birds. U.S. Geological Survey
Professional Paper 1842, 20p. https://doi.org/10.3133/pp1842F
Shaffer JA, Igl LD, Johnson DH et al (2019f) The effects of management practices on grassland
birds–Mountain Plover (Charadrius montanus), Chapter E. In: Johnson DH, Igl LD, Shaffer
JA et al (eds) The effects of management practices on grassland birds. U.S. Geological Survey
Professional Paper 1842, 10p. https://doi.org/10.3133/pp1842E
Shaffer JA, Loesch CR, Buhl DA (2019g) Estimating offsets for avian displacement effects of
anthropogenic impacts. Ecol Appl 28(8):e01983. https://doi.org/10.1002/eap.1983
Sherfy MH, Anteau MJ, Shaffer TL et al (2018) Density and success of upland duck nests in native-
and tame-seeded conservation fields. Wildl Soc Bull 42:204–212
Silverman NJ, Allred BW, Donnelly JP (2018) Low-tech riparian and wet meadow restoration
increases vegetation productivity and resilience across semiarid rangelands. Restor Ecol 27:269–
278
Sketch M, Dayer AA, Metcalf AL (2020) Western ranchers’ perspectives on enablers and constraints
to flood irrigation. Rangel Ecol Manag 73(2):285–296
Smeins FE, Diamond DD, Hanselka CW (1991) Coastal prairie. In: Coupland RT (ed) Ecosystems
of the world 8. A natural grasslands-introduction and western hemisphere. Elsevier Press, New
York. pp 269–290
Smith LM (2003) Playas of the Great Plains. University of Texas Press, Austin, Texas, USA
Smith LM, Kadlec JA (1985) Fire and herbivory in a Great Salt Lake marsh. Ecology 66:259–265
Smith AG, Stoudt JH, Gollp JB (1964) Prairie potholes and marshes. In: Linduska JP, Nelson AL
(eds) Waterfowl tomorrow. US Department of Interior, Washington, DC, pp 39–50
Smith LM, Pederson RL, Kadlec JA (1989) Habitat management for migrating and wintering
waterfowl. Texas Tech University Press, Lubbock, Texas, USA
Smith LM, Haukos DA, Prather RM (2004) Avian response to vegetative pattern in playa wetlands
during winter. Wildl Soc Bull 32:474–480
Smith LM, Haukos DA, McMurry ST et al (2011) Ecosystem services provided by playas in the
High Plains; potential influences of USDA conservation programs. Ecol Appl 21:82–92
Smith LM, Haukos DA, McMurry ST (2012) High Plains playas. In: Batzer DP, Baldwin AH (eds)
Wetland habitats of North America: ecology and conservation concerns. University of California
Press, Berkely and Los Angeles, California, pp 299–311
Smith PA, Smith AC, Andres B et al (2023) Accelerating declines of North America’s shorebirds
signal the need for urgent conservation action. Ornithol Appl 125:1–14
Snyder KA, Evers L, Chambers JC et al (2019) Effects of changing climate on the hydrological
cycle in cold desert ecosystems of the Great Basin and Columbia Plateau. Rangel Ecol Manag
72:1–12
Sofaer HR, Skagen SK, Barsugli JJ et al (2016) Projected wetland densities under climate change:
habitat loss but little geographic shift in conservation strategy. Ecol Appl 26:1677–1692
Specht H, St-Louis V, Gratto-Trevor CL et al (2020) Habitat selection and nest survival in two Great
Plains shorebirds. Avian Conserv Ecol 15:3. https://doi.org/10.5751/ACE-01487-150103
Stafford JD, Janke AK, Anteau MJ et al (2014) Spring migration of waterfowl in the northern
hemisphere: a conservation perspective. Wildfowl 4:70–85
Stearns SC (1992) The evolution of life histories. Oxford University Press, Oxford, England
Stephens SE, Rotella JJ, Lindberg MS et al (2005) Duck nest survival in the Missouri Coteau of
North Dakota: landscape effects at multiple spatial scales. Ecol Appl 15:2137–2149
468 J. L. Vest et al.
Stutzenbaker CD, Weller MW (1989) The Texas Coast. In: Smith LM, Pederson RL, Kaminski RM
(eds) Habitat management for migrating and wintering waterfowl in North America. Texas Tech
University Press, Lubbock, Texas, USA
Sullivan BL, Aycrigg JL, Barry JH et al (2014) The eBird enterprise: an integrated approach to
development and application of citizen science. Biol Conserv 169:31–40
Svingen D, Anderson SH (1998) Waterfowl management on grass-sage stock ponds. Wetlands
18:84–89
Swift RJ, Rodewald AD, Johnson JA, Andres BA, Senner NR (2020) Seasonal survival and reversible
state effects in a long-distance migratory shorebird. J Anim Ecol 89:2043–2055
Tangen BA, Bansal S, Jones S et al (2022) Using a vegetation index to assess wetland condition in
the Prairie Pothole Region of North America. Front Environ Sci 10:889170. https://doi.org/10.
3389/fenvs.2022.889170
Tessman SA (2004) Management of created palustrine wetlands. In: McKinstry MC, Hubert WA,
Anderson SH (eds) Wetland and riparian areas of the Intermountain West. University of Texas
Press, Austin, Texas, pp 154–184
Tori GM, McLeod S, McKnight K et al (2002) Wetland conservation and Ducks Unlimited: real
world approaches to multispecies management. Waterbirds 25 (Special Publication 2):115–121
Trail PW (2006) Avian mortality at oil pits in the United States: a review of the problem and efforts
for its solution. Environ Manage 38:532–544. https://doi.org/10.1007/s00267-005-0201-7
Tsai J, Venne LS, Smith LM et al (2012) Influence of local and landscape characteristics on avian
richness and density of wet playas of the Southern Great Plains, USA. Wetlands 32:605–618
U.S. Geological Survey Bird Banding Laboratory (2020) USGS celebrates 100 years of bird banding
lab: A century of advancing bird conservation science. https://www.usgs.gov/news/featured-
story/
U.S. Shorebird Conservation Plan Partnership (2016) U.S. shorebirds of conservation concern–
2016. U.S. Fish and Wildlife Service. http://www.shorebirdplan.org/science/assessment-conser
vation-status-shorebirds/
Uden DR, Allen CR, Bishop AA et al (2015) Predictions of future ephemeral springtime waterbird
stopover habitat availability under global change. Ecosphere 6:1–26
USFWS (2011) Birds of management concern and focal species. U.S. Fish and Wildlife Service,
Migratory Bird Program. U.S. Department of the Interior, Washington, DC, USA
USFWS (2022) Waterfowl population status, 2022. U.S. Fish and Wildlife Service, U.S. Department
of the Interior, Washington, DC, USA
USFWS (2023) Fish and Wildlife Service. Mid-winter waterfowl survey. U.S. Department of the
Interior, Washington, DC, USA. https://migbirdapps.fws.gov/mbdc/databases/mwi/mwidb.asp
Veech JA, Pardieck KL, Ziolkowski DJ Jr (2017) How well do route survey areas represent land-
scapes at larger spatial extents? An a nalysis of land cover composition along Breeding Bird
Survey routes. Condor 119:607–615
Venne LS, Frederick PC (2013) Foraging wading birds (CICONIIFORMES) attraction to prescribed
burns in an oligotrophic wetland. Fire Ecol 9:78–95
Verheijen BHF, Varner DM, Haukos DA (2020) Future loss of playa wetlands decrease network
structure and connectivity of the Rainwater Basin, Nebraska. Landsc Ecol 35:453–467
Vermillion WG (2012) Fall habitat objectives for priority Gulf Coast Joint Venture shorebird species
using managed wetlands and grasslands, version 4.0. Gulf Coast Joint Venture, Lafayette,
Louisiana
Vermillion W, Eley JW, Wilson B et al (2008) Partners in Flight bird conservation plan, Gulf Coastal
Prairie, bird conservation region 37, version 1.3. Gulf Coast Bird Observatory, Lake Jackson,
Texas, USA
Walker J, Taylor PD (2017) Using eBird data to model population change of migratory bird species.
Avian Conserv Ecol 12(1):4. https://doi.org/10.5751/ACE-00960-120104
Walker J, Rotella JJ, Stephens SE et al (2013a) Time-lagged variation in pond density and primary
productivity affects duck nest survival in the Prairie Pothole Region. Ecol Appl 23:1061–1074
13 Waterfowl and Wetland Birds 469
Walker J, Rotella JJ, Schmidt JH et al (2013b) Distribution of duck broods relative to habitat
characteristics in the Prairie Pothole Region. J Wildl Manage 77:392–404
Warren J, Rotella J, Thompson J (2008) Contrasting effects of cattle grazing intensity on upland-
nesting duck production at nest and field scales in the Aspen Parkland. Avian Conserv Ecol
3:6
Weiser EL, Lanctot RB, Brown SC et al (2020) Annual adult survival drives trends in Arctic-breeding
shorebirds but knowledge gaps in other vital rates remain. Condor 122:1–14
Weller M (1999) Wetland birds: habitat resources and conservation implications. Cambridge
University Press, Cambridge. https://doi.org/10.1017/CBO9780511541919
West BC, Messmer TA (2006) Effects of livestock grazing on duck nesting habitat in Utah. Rangel
Ecol Manag 59:208–211
Williams BK, Nichols JD, Conroy MJ (2002) Analysis and management of animal populations.
Academic Press, San Diego, California, USA
Wilson BC, Esslinger CG (2002) North American waterfowl management plan, Gulf Coast
Joint Venture: Texas Mid-Coast initiative. North American Waterfowl Management Plan,
Albuquerque, New Mexico, USA, 28pp
Wright CK, Wimberly MC (2013) Recent land use change in the Western Corn Belt threatens
grasslands and wetlands. PNAS 110(10):4134–4139. https://doi.org/10.1073/pnas.1215404110
Yetter AP, Hagy HM, Horath MM et al (2018) Mallard survival, movements, and habitat use during
autumn in Illinois. J Wildl Manage 82:182–191
Zarzycki M (2017) Evidence for cross-seasonal effects: insights from long-term data on northern
pintail. Thesis, Oregon State University, Corvallis, Oregon, USA
Open Access This chapter is licensed under the terms of the Creative Commons Attribution 4.0
International License (http://creativecommons.org/licenses/by/4.0/), which permits use, sharing,
adaptation, distribution and reproduction in any medium or format, as long as you give appropriate
credit to the original author(s) and the source, provide a link to the Creative Commons license and
indicate if changes were made.
The images or other third party material in this chapter are included in the chapter’s Creative
Commons license, unless indicated otherwise in a credit line to the material. If material is not
included in the chapter’s Creative Commons license and your intended use is not permitted by
statutory regulation or exceeds the permitted use, you will need to obtain permission directly from
the copyright holder.
... Semi-permanent basin wetlands and reservoirs, on the other hand, have hydrologies that encourage the growth of submerged aquatic vegetation (SAV), which provide growth and reproduction substrates for macroinvertebrates (Batzer, 2013;Bauer et al., 2020;Fredrickson & Reed, 1988;Hagy et al., 2011;Schad et al., 2020). Further research on both the physical and chemical properties of working wetlands and how they might further impact patterns of invertebrate distribution are warranted (Arzel et al., 2020;Kantrud, 1986;Longcore et al., 2006;Swanson et al., 1988;Vargas et al., 2022;Vest et al., 2023). ...
Article
Full-text available
Food availability varies considerably over space and time in wetland systems, and consumers must be able to track those changes during energetically‐demanding points in the life cycle like breeding. Resource tracking has been studied frequently among herbivores, but receives less attention among consumers of macroinvertebrates. We evaluated the change in resource availability across habitat types and time and the simultaneous density of waterfowl consumers throughout their breeding season in a high‐elevation, flood‐irrigated system. We also assessed whether the macroinvertebrate resource density better predicted waterfowl density across habitats, compared to consistency (i.e., temporal evenness) of the invertebrate resource or taxonomic richness. Resource density varied marginally across wetland types but was highest in basin wetlands (i.e., ponds) and peaked early in the breeding season, whereas it remained relatively low and stable in other wetland habitats. Breeding duck density was positively related to resource density, more so than temporal resource stability, for all species. Resource density was negatively related to duckling density, however. These results have the potential to not only elucidate mechanisms of habitat selection among breeding ducks in flood‐irrigated landscapes but also suggest there is not a consequential trade‐off to selecting wetland sites based on energy density versus temporal resource stability and that good‐quality wetland sites provide both.
Book
Full-text available
From 2009 through 2011, U.S. Fish and Wildlife Service Regions 1 (Pacific), 2 (Southwest), 6 (Mountain-Prairie), 8 (Pacific-Southwest), and 9 (National Headquarters), together with biologists from the interior western states designed and implemented a survey to inventory and document the status of colonial waterbirds. The Western Colonial Waterbird Survey (WCWS) and resulting Atlas of Breeding Colonial Waterbirds in the Interior Western United States were designed primarily to provide biologists, resource managers, regulatory agencies, and researchers with the best available data on the size and location of inland western waterbird colonies, in a standardized format. This Atlas presents summary results of surveys conducted in the eight participating interior western states. The information compiled here serves as a baseline of recent and available historical distribution and abundance. Atlas maps and individual state information can be found at: https://departments.weber.edu/avianecologylab/WesternWaterbirdAtlas/atlas.html
Article
Full-text available
Shorebirds are declining to a greater extent than many other avian taxa around the world. In North America, shorebirds, along with aerial insectivores and grassland birds, have some of the highest proportions of declining species of any group. Here, we apply a new hierarchical Bayesian model to analyze shorebird migration monitoring data from across North America, from 1980 to 2019, and present the most recent available estimates of trends for 28 species. Point estimates for survey-wide trends in abundance were negative for 26 of 28 species (93%). Despite challenges with low precision associated with migration count data, trends for 19 species had 95% credible intervals that were entirely negative. More than half of the species were estimated to have lost >50% of their abundance. Furthermore, estimated rates of decline have accelerated during the last three generations for most species. Point estimates of trend were more negative for 18 species (64%) during the most recent three-generation period in comparison to the previous three-generation period. Many species now exceed international criteria for threatened species listing. The analytic approach used here allows us to model regional variation in trends, although survey coverage and strength of inference were greatest in the eastern portions of North America (east of 100°W). We found the greatest declines at staging sites along the Atlantic Coast from North Carolina to Nova Scotia, and lesser declines along the Gulf Coast and in the midcontinental United States. The declines in shorebird populations reported here are worrisome and signal the urgent need for conservation action. In addition, it would be beneficial to validate these results through collection and analysis of complementary data, and to initiate demographic studies throughout the annual cycle to determine where and when declines are most likely to originate. This improved information will allow for development of more targeted efforts to reverse declines through conservation action.
Article
Full-text available
Management of waterfowl that migrate seasonally across North America occurs within four flyways that were delineated in the early 1900s to include the annual movements of populations. Movements may have changed over the past century since the administrative flyways were established, and may do so while management plans are in use, so information about transitions among flyways through time can illustrate how management assumptions may change. Today there are more than 12 million records from 60 years of migratory waterfowl band recoveries to assess adaptive management approaches that will be most effective when they account for movements within and between flyways. We examined how much the movement of North American waterfowl occurs between flyways, whether those movements have changed through time, and whether movements of mallards are representative of multiple species, as suggested by current harvest management strategies. We estimated the probability a duck would transition from one flyway to another and the strength of migratory connectivity (MC) for each species within and among flyways. We used capture–mark–recovery models to estimate population‐specific movement within and among flyways (transition probabilities) for 15 migratory waterfowl species that were banded during breeding and recovered during winter. We developed new functionality in the R package MigConnectivity to estimate the species‐specific strength of MC using transition probability samples from the capture–mark–recovery models. We found the regular movement of duck populations among flyways, overall weak MC, and no consistent change in migratory movements through time. Mallard movements were median among all duck species, but significantly different from many species, particularly diving ducks. Despite the significant movement between flyways, our work suggests flyway management of waterfowl matches many of the seasonal movements of these species when considering mid‐continent flyway management. We recommend models accounting for all transition probabilities between populations and regularly estimating harvest derivations, transition probabilities, and MC metrics to verify that the current movements match model assumptions.
Article
Full-text available
Context Grasslands of the North American Great Plains are among the world’s most imperiled ecosystems. Determining landscapes at risk of grassland loss will benefit grassland conservation programs by enabling prioritization of parcels for acquisition. Objectives We hypothesized that gradients in the amount of grass in local landscapes resulting from past conversion would be a useful predictor of future conversion. We developed simple, intuitive models predicting grassland conversion across the United States northern Great Plains. Methods We used the grassland/herbaceous, pasture/hay, and emergent herbaceous wetland classes from the National Land Cover Database to evaluate grassland loss from 2001 to 2016. We classified the amount of grass in 13.2-km² landscapes in 2001 into percentage bins and used generalized additive models to relate these values to the total and proportion of grassland lost, by percentage bin. We applied models to landcover data to create spatial surfaces predicting conversion. Results Area of grass lost was highest in 40–60% grass bins, except for heavily cropped states, where highest losses occurred in 10–20% grass bins. Percentage of grass in local landscapes was generally a strong predictor of state-level total grassland loss and proportion of grassland lost. Predicted conversion for remaining grasslands varied within and among states. Conclusion The amount of grass in local landscapes can be a useful indicator of grassland conversion. Our simple models complement species distribution models used to guide conservation in the region. Mechanistic models of conversion can be improved by including amount of grass in local landscapes.
Article
Full-text available
The homogenization of freshwater ecosystems and their biological communities has emerged as a prevalent and concerning phenomenon because of the loss of ecosystem multifunctionality. The millions of prairie-pothole wetlands scattered across the Prairie Pothole Region (hereafter PPR) provide critical ecosystem functions at local, regional, and continental scales. However, an estimated loss of 50% of historical wetlands and the widespread conversion of grasslands to cropland make the PPR a heavily modified landscape. Therefore, it is essential to understand the current and potential future stressors affecting prairie-pothole wetland ecosystems in order to conserve and restore their functions. Here, we describe a conceptual model that illustrates how (a) historical wetland losses, (b) anthropogenic landscape modifications, and (c) climate change interact and have altered the variability among remaining depressional wetland ecosystems (i.e., ecosystem homogenization) in the PPR. We reviewed the existing literature to provide examples of wetland ecosystem homogenization, provide implications for wetland management, and identify informational gaps that require further study. We found evidence for spatial, hydrological, chemical, and biological homogenization of prairie-pothole wetlands. Our findings indicate that the maintenance of wetland ecosystem multifunctionality is dependent on the preservation and restoration of heterogenous wetland complexes, especially the restoration of small wetland basins.
Article
Full-text available
Wetlands deliver a suite of ecosystem services to society. Anthropogenic activities, such as wetland drainage, have resulted in considerable wetland loss and degradation, diminishing the intrinsic value of wetland ecosystems worldwide. Protecting remaining wetlands and restoring degraded wetlands are common management practices to preserve and reclaim wetland benefits to society. Accordingly, methods for monitoring and assessing wetlands are required to evaluate their ecologic condition and outcomes of restoration activities. We used an established methodology for conducting vegetation-based assessments and describe a case study consisting of a wetland condition assessment in the Prairie Pothole Region of the North American Great Plains. We provide an overview of an existing method for selecting wetlands to sample across broad geographic distributions using a spatially balanced statistical design. We also describe site assessment protocols, including vegetation survey methods, and how field data were applied to a vegetation index that categorized wetlands according to ecologic condition. Results of the case study indicated that vegetation communities in nearly 50% of the surveyed wetlands were in very poor or poor condition, while only about 25% were considered good or very good. Approximately 70% of wetlands in native grasslands were categorized as good or very good compared to only 12% of those in reseeded grasslands (formerly cropland). In terms of informing restoration and management activities, results indicated that improved restoration practices could include a greater focus on establishing natural vegetation communities, and both restored and native prairie wetlands would benefit from enhanced management of invasive species.
Article
Full-text available
Harvest of wild organisms is an important component of human culture, economy, and recreation, but can also put species at risk of extinction. Decisions that guide successful management actions therefore rely on the ability of researchers to link changes in demographic processes to the anthropogenic actions or environmental changes that underlie variation in demographic parameters. Ecologists often use population models or maximum sustained yield curves to estimate the impacts of harvest on wildlife and fish populations. Applications of these models usually focus exclusively on the impact of harvest and often fail to consider adequately other potential, often collinear, mechanistic drivers of the observed relationships between harvest and demographic rates. In this study, we used an integrated population model and long‐term data (1973–2016) to examine the relationships among hunting and natural mortality, the number of hunters, habitat conditions, and population size of blue‐winged teal Spatula discors, an abundant North American dabbling duck with a relatively fast‐paced life history strategy. Over the last two and a half decades of the study, teal abundance tripled, hunting mortality probability increased slightly (<0.02), and natural mortality probability increased substantially (>0.1) at greater population densities. We demonstrate strong density‐dependent effects on natural mortality and fecundity as population density increased, indicative of compensatory harvest mortality and compensatory natality. Critically, an analysis that only assessed the relationship between survival and hunting mortality would spuriously indicate depensatory mortality due to multicollinearity between abundance, natural mortality and hunting mortality. Our findings demonstrate that models that only consider the direct effect of hunting on survival or natural mortality can fail to accurately assess the mechanistic impact of hunting on population dynamics due to multicollinearity among demographic drivers. This multicollinearity limits inference and may have strong impacts on applied management actions globally.
Book
This book introduces life history evolution to postgraduate students just beginning their research in population biology, ecology, or evolutionary biology. It discusses major analytical tools, gives examples of their applications, and provides problems for discussion at the end of each chapter. It will interest all biologists wishing to understand the evolution of the life cycle and the causes of phenotypic variation in fitness, and it contains the seeds of applications of life history theory to population dynamics, behaviour, and community ecology. Care is taken in Part I to build up the tools needed for a well-rounded evolutionary explanation: demography, quantitative genetics, reaction norms, trade offs and phylogenetic/comparative analysis. Part II discusses the evolution of major life history traits. This is a comprehensive, up-to-date text in a field that holds a central position in modern ecology and evolution.