kevin l. monteith, vernon c. bleich, thomas r. stephenson...

51
POPULATION DYNAMICS OF MULE DEER IN THE EASTERN SIERRA NEVADA: IMPLICATIONS OF NUTRITIONAL CONDITION Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson, and Becky M. Pierce. California Department of Fish and Game, Bishop, California 93514 Abstract: Mule deer populations routinely fluctuate in size and managers are rarely afforded the opportunity to understand factors that determine such changes. Our goal was to understand the factors that limit population growth of mule deer in the Sierra Nevada, with an emphasis on developing our understanding of the nutritional constraints associated with reproduction and survival, and the application of those nutritional relationships to management of mule deer. We conducted research on a population of mule deer that spends winters in Round Valley, on the east slope of the Sierra Nevada, and spends summers on high elevation ranges on both sides of the Sierra crest. We monitored annual survival, reproduction, and nutritional condition of adult females during 1997-2009 and monitored survival of neonatal mule deer during 2006-2008. The population size of mule deer residing in Round Valley has been highly variable since 1985, largely in response to the highly variable climate in the Sierra Nevada. Post-winter nutritional condition of adult females was influenced by both intraspecific competition for forage (density dependence) and its association with forage production in response to winter precipitation. In addition to intrinsic habitat value, autumn body fat was influenced by reproductive status of each female and her summer residency. Females that spent summers west of the crest experienced lower reproductive success, but also lower somatic cost of reproduction compared with females that spent summers east of the Sierra crest. The major proximate cause of mortality for young mule deer was predation and most young west of the Sierra crest were lost to bear predation. Body fat of adult females following winter provided an encompassing measure of population health and predicted adult female survival, reproduction, recruitment of young, an index to abundance of males, and overall population trajectory during the following year. Nutritional limits on survival and recruitment of deer in Round Valley clearly indicate bottom-up limitation on deer dynamics and the importance of data on nutritional condition when interpreting dynamics of deer populations. Estimation of nutritional condition provided the greatest insight into multiple population parameters and predicted population rate of change (λ). Recent nutritional (March mean body fat = 2.6%) and demographic data suggests that Round Valley is experiencing a new, reduced carrying capacity (K). Less somatic cost of reproduction, but also lower recruitment for west-side females suggests that some predation experienced by west-side young is additive. Although a reduction in bear numbers might result in increased recruitment by west-side females, the increased number of animals on winter range would simply exacerbate the effects of an already forage

Upload: others

Post on 04-Jul-2020

0 views

Category:

Documents


0 download

TRANSCRIPT

Page 1: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

POPULATION DYNAMICS OF MULE DEER IN THE EASTERN SIERRA NEVADA: IMPLICATIONS OF NUTRITIONAL CONDITION

Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson, and Becky M. Pierce. California Department of Fish and Game, Bishop, California 93514

Abstract: Mule deer populations routinely fluctuate in size and managers are rarely afforded the opportunity to understand factors that determine such changes. Our goal was to understand the factors that limit population growth of mule deer in the Sierra Nevada, with an emphasis on developing our understanding of the nutritional constraints associated with reproduction and survival, and the application of those nutritional relationships to management of mule deer. We conducted research on a population of mule deer that spends winters in Round Valley, on the east slope of the Sierra Nevada, and spends summers on high elevation ranges on both sides of the Sierra crest. We monitored annual survival, reproduction, and nutritional condition of adult females during 1997-2009 and monitored survival of neonatal mule deer during 2006-2008. The population size of mule deer residing in Round Valley has been highly variable since 1985, largely in response to the highly variable climate in the Sierra Nevada. Post-winter nutritional condition of adult females was influenced by both intraspecific competition for forage (density dependence) and its association with forage production in response to winter precipitation. In addition to intrinsic habitat value, autumn body fat was influenced by reproductive status of each female and her summer residency. Females that spent summers west of the crest experienced lower reproductive success, but also lower somatic cost of reproduction compared with females that spent summers east of the Sierra crest. The major proximate cause of mortality for young mule deer was predation and most young west of the Sierra crest were lost to bear predation. Body fat of adult females following winter provided an encompassing measure of population health and predicted adult female survival, reproduction, recruitment of young, an index to abundance of males, and overall population trajectory during the following year. Nutritional limits on survival and recruitment of deer in Round Valley clearly indicate bottom-up limitation on deer dynamics and the importance of data on nutritional condition when interpreting dynamics of deer populations. Estimation of nutritional condition provided the greatest insight into multiple population parameters and predicted population rate of change (λ). Recent nutritional (March mean body fat = 2.6%) and demographic data suggests that Round Valley is experiencing a new, reduced carrying capacity (K). Less somatic cost of reproduction, but also lower recruitment for west-side females suggests that some predation experienced by west-side young is additive. Although a reduction in bear numbers might result in increased recruitment by west-side females, the increased number of animals on winter range would simply exacerbate the effects of an already forage

Page 2: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 2

limited winter range. Implementation of a female harvest on winter range in Round Valley would lower the population density with respect to K, which should result in improved nutritional condition of adult females and a population less susceptible to minor perturbations. Lack of fires on west side summer range and recent fires on winter range also may have altered the forage base and reduced K.

INTRODUCTION

Mule deer populations have experienced periodic declines over much of the last century and causes of the declines remain speculative and controversial (Connolly 1981; deVos et al. 2003). Cited causes of the declines include loss of habitat due to development, deterioration of nutritional forages, competition with other ungulates, predation, disease, increased hunting mortality, poaching, severe winter weather, and droughts (deVos et al. 2003). Similar widespread declines across regions suggest that large-scale environmental conditions may trigger those declines (Unsworth et al. 1999). Failure to identify factors that regulate populations, however, has precluded the detection of the underlying causes. Detailed and long-term investigations are necessary to identify the factors regulating populations of mule deer to improve their conservation and management (Caughley 1977, Connelly 1981, deVos et al. 2003, Mackie et al. 2003). Indeed, the scarcity of long-term investigations of large herbivores has limited the use of scientific approaches to manage and conserve those large mammals (Bleich et al. 2006, Lindstrom 1999).

Population parameters such as age-specific survival and reproduction are determinants of the dynamics of large mammal populations (Coulson et al. 2001, Gaillard et al. 2000, Lande et al. 2002) and those parameters are influenced by a combination of stochastic environmental variation and density dependence (Albon et al. 2000, Sæther 1997). Research has traditionally focused on the role of predation in limiting populations of ungulates with less focus given to resource availability (Sæther 1997). That predator-centric focus likely has hampered the understanding of predator-prey dynamics (Bowyer et al. 2005). Bottom-up versus top-down limitation of populations has remained a central theme of ecological studies and holds implications for both theoretical and applied ecology (Bowyer et al. 2005, Hairston et al. 1960, Sinclair and Krebs 2002). Controversy over whether populations of large mammals are top-down or bottom-up regulated, however, continues. Although often considered a dichotomy, top-down and bottom-up influences are rather a continuum and act simultaneously on populations (Bowyer et al. 2005). Bowyer et al. (2005) suggested that assessment of top-down versus bottom-up limitation may be most accurately and efficiently determined by understanding prey population dynamics.

Demographic (vital) rates of large herbivores are believed to respond to resource limitation in a predictable sequence from increased age of first reproduction, decreased survival of young, decreased reproduction by adults, and lastly, decreased survival of adults (Gaillard et al. 1998; 2000, Eberhardt

Page 3: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 3

2002). That variability in reproduction and survival is largely determined by animal condition (Cameron and Ver Hoef 1994, Testa and Adams 1998, Keech et al. 2000, Stewart et al. 2005, Bender et al. 2007). Moreover, the use of population vital rates and mortality factors to make inferences for top-down and bottom-up forces without knowledge of seasonal deficiencies in nutrition may be of limited value (Bowyer et al. 2005, Brown et al. 2007). Understanding patterns of nutritional condition and its relative impact on population dynamics will enhance our knowledge of life-history strategies and population regulation of ungulates (Stephenson et al. 2002, Cook et al. 2004, Parker et al. 2009); nutrition likely provides the basis for most life-history strategies (Parker et al. 2009). Measures of nutritional condition may ultimately function as the mechanism through which intraspecific competition for resources are mediated, and could provide the most direct and sensitive measure of resource limitation because effects of resource limitation on animal condition are realized before expression in reproduction and survival parameters (Bender et al. 2008).

Nutritional ecology is defined as the science relating an animal to its environment through nutritional interactions (Parker et al. 2009). Animal body condition is the integration of both nutrient intake and energetic costs. Knowledge of animal condition is critical to evaluate habitat conditions and predict population productivity because condition is related to nearly every parameter of productivity and survival (Short 1981). Animal condition reflects previous conditions encountered including precipitation and forage growth, deer population density, and energetic costs of reproduction (Cook et al. 2004, Stewart et al. 2005). Moreover, animal condition forms a predictable link between future survival and reproduction (Keech et al. 2000, Lomas and Bender 2006, Bender et al. 2007; 2008). Our overarching goal was to understand the factors that limit population growth of mule deer in the western Great Basin, with an emphasis on developing an understanding of the nutritional constraints associated with reproduction and survival, and the application of those nutritional relationships to management of mule deer. Indeed, knowledge of life-history characteristics and the inherent influence of nutritional constraints will provide insight into factors that regulate mule deer populations and allow for sound management decisions based on empirical data (Bowyer et al. 2005, Parker et al. 2009). Furthermore, understanding resource limitation will aid in developing harvest criteria for deer populations.

Additional objectives of this research included: (1) calculate annual population size using mark-resight estimates, (2) monitor annual patterns of survival and factors that influence survival of adult female mule deer, (3) determine survival and cause-specific mortality of neonatal mule deer born on either side of the Sierra crest, (4) determine the somatic costs of reproduction with respect to migratory patterns of female mule deer, (5) identify factors that underpin dynamics of mule deer in the western Great Basin and develop techniques to quantify carrying capacity for mule deer populations, and (6) apply those relationships to management of mule deer in Round Valley and the eastern Sierra Nevada. (7) Understanding deer population performance and trends is a critically important factor to be considered in the protection and, ultimately, the

Page 4: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 4

recovery of Sierra Nevada bighorn sheep. Further, deciphering and clarifying the “ecosystem link” between deer, lion, and sheep populations will result from this effort and ultimately will serve to solidify ecosystem conservation objectives in the eastern Sierra Nevada. Lastly, (8) preparation of manuscripts is currently underway and formal publication of the results obtained from the work in Round Valley will occur in the appropriate professional journals.

We conducted research on a population of mule deer that winters on the east side of the Sierra Nevada in Round Valley (37°24′, 118°34′W), Inyo and Mono Counties, California (Kucera 1998). Mule deer inhabited approximately 90 km2 of Round Valley during November-April, but the size of the area of use was dependent on snow depth (Kucera 1998). Most of those mule deer migrate northward and westward to high-elevation ranges in summer (Kucera 1992, Pierce et al. 1999); many migrate over passes to the west side of the Sierra Crest while others remain on the east side (Figure 1). As recently as 1981, 87% of the mule deer wintering in Round Valley migrated to occupy summer ranges west of the Sierra Crest (Kucera 1998). The habitats on either side of the Sierra crest contrast substantially, which afforded a unique opportunity to evaluate influences of summer range quality on patterns of recruitment, energetic cost of reproduction, and life-history tradeoffs associated with reproduction.

STUDY AREA The Sierra Nevada is a massive granite block that tilts to the western side with a gradual slope of only 2-6% encompassing 75 to 100 km before reaching a crest. Conversely, the spectacular eastern slope of the Sierra Nevada is characterized by steep slopes rising abruptly from the bordering valleys that merge with the western edge of the Great Basin. The Owens Valley area extending from the Sherwin Grade occurring north of the town of Bishop, south about 120km, is demarcated by elevation from 4200m at the mountain summits to 1220m at the valley floor which occur over horizontal distances of <10 km (Kucera 1988).

Within the Sierra, 95% of the precipitation including rain or snow accumulates between October and May (Storer et al. 2004). Winter storms from the Pacific Ocean deposit moisture as they move up the western slope, with a rain shadow effect leaving a more arid landscape on the eastern slope where the Great Basin Desert begins. At 1,676m in elevation, precipitation on the western slope severely contrasts the lack of moisture on the eastern slope creating a dramatically different flora and fauna on either side of the crest (Storer et al. 2004). The dense pine-fir stands and rivers commonly found on the west side contrast the sparse forests changing to sagebrush scrub below 2130m and few small streams found on the east side. The formidable Sierra Crest sharply delineates the western slope from the eastern slope of the Sierra Nevada and is traversable by a series of passes that increase in elevation from north to south (Kucera 1988).

Page 5: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 5

Winter range Round Valley is bounded to the west by the Sierra Nevada, particularly

Mount Tom (4,161 m) and Wheeler Ridge (3,640 m), to the south by large boulders and granite ridges of the Tungsten Hills and Buttermilk’s, and to the east by Highway 395 connecting Reno, Nevada to the Los Angeles Basin, California (Figure 2). The northern end of Round Valley gradually rises from the valley floor at 1,375 m to the top of the Sherwin Grade at 2,135 m. Open pasture land comprised about 18.3 km2 of the eastern portion of the valley and 3.2 km2 of Round Valley was developed residential housing. Deer only used pasture when heavy snows forced them from higher elevation areas dominated by bitterbrush (Pierce et al. 2004).

Vegetation of Round Valley was characteristic of that portraying the western Great Basin and the sagebrush belt (Storer et al. 2004). Typical vegetation composing habitats used by mule deer in Round Valley included bitter brush (Purshia glandulosa), sage brush (Artemesia tridentata), blackbrush (Coleogyne ramosissima), desert peach (Prunus andersonii), rabbitbrush (Chrysothamnus nauseosus),and Mormon tea (Ephedra nevadensis). Riparian areas consisted of willow (Salix sp.), Rose (Rosa sp.), and water birch (Betula occidentalis), although forbs and graminoids were uncommon in Round Valley during winter (Kucera 1988, Kucera 1992, Kucera 1997, Pierce et al. 2000; 2004).

Summer range Summer range for mule deer that overwinter in Round Valley occurs on

both sides of the Sierra crest at elevations ranging from 2,200 to >3,600m (Kucera 1988; Figure 2). Summer ranges west of the Sierra crest are substantially more mesic and forested than that east of the crest (Bleich et al. 2006). The western slope of the summer range is dominated by the upper montane and mixed conifer vegetation zones (Storer et al. 2004) consisting of conifer stands with little understory including red fir (Abies magnifica), white fir (Abies concolor), lodgepole pine (Pinus contorta), western white pine (Pinus monticola), Jeffery pine (Pinus jeffreyi), and quaking aspen (Populus tremuloides). Nevertheless, montane chaparral consisting of dense stands of manzanitas (Arctostaphylos sp.), Ceanothus (Ceanothus sp.), and bush chinquapin (Chrysolepis sempervirens) often exists at lower elevations within drainages on the western slope (Storer et al. 2004). Mountain springs and streams, including large rivers, are common within the western Sierra.

The eastern slope of the Sierra Nevada up to approximately 2,130 m is dominated by the sagebrush vegetation zone (Storer et al. 2004). This zone is dominated by sagebrush, but also includes other shrub species such as bitterbrush, ceanothus, manzanitas, rabbitbrush, mountain mahogany (Cercocarpus betuloides), and contains pure stands of Jeffrey pine in some areas (Storer et al. 2004).

Grizzly bears (Ursos arctos) formerly existed west of the Sierra crest, but have since been extirpated (Storer and Tevis 1955). Nonetheless, areas occupied by migratory mule deer include a full complement of predators of both winter and summer ranges consisting of mountain lions (Puma concolor),

Page 6: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 6

coyotes (Canis latrans), and bobcats (Lynx rufus; Pierce et al. 2000). Black bears (Ursus americanus) are frequently encountered on summer ranges west of the Sierra crest, but are uncommon in the eastern Sierra Nevada (Wildlife Programs Branch 2006).

METHODS

Adult Capture and Handling During March 1997-2009 and November 2002-2008, we captured adult

female mule deer on winter range in Round Valley using a hand-held net gun fired from a helicopter (Barrett et al. 1982, Krausman et al. 1985). We hobbled and blindfolded each animal prior to being ferried by helicopter to a central processing station with shelter and processing crews. During processing, we frequently monitored temperature, pulse, and respiration of each animal. When necessary, we placed cold water on the animals head, chest, and abdomen to facilitate cooling. In addition, if animals with elevated body temperatures were reluctant to return to a safe level, we administered Banamine. Captured animals were administered prophylactics including antibiotics, vitamin E/selenium, and vitamin E. We measured chest circumference, length of the metataurus, body mass, and collected blood and a fecal sample from each individual. To allow aging by cementum annuli technique (Matson’s Laboratory), we removed one incisiform canine using techniques previously described by Swift et al. (2002), which have been verified to have no effect on body mass, percent body fat, pregnancy rate, or fetal rate (Bleich et al. 2003). We fitted each animal with a radiocollar equipped with mortality sensor and finished with orange tape or plastic to ensure visibility during mark/resight population surveys.

During captures, we employed ultrasonography using an Aloka 210 portable ultrasound device (Aloka, Wallinford, CT) with a 5-MHz transducer to determine body fat as a measure of nutritional condition in mule deer. We used ultrasound to measure maximum thickness of subcutaneous fat deposition at the thickest point cranial to the cranial process of the tuber ischium to the nearest 0.1 cm (Stephenson et al. 2002). We also measured the thickness of the bicep femoralis and longissimus dorsi to provide a measure of protein catabolism for animals in poor condition (Cook et al. 2001, Stephenson et al. 2002). We complemented ultrasonography with palpation to achieve a body condition score validated for mule deer (Cook et al. 2007) to aid in evaluating nutritional condition of animals that have mobilized subcutaneous fat reserves (<5.6% ingesta-free body fat). We calculated rLIVINDEX as subcutaneous rump fat thickness plus rump body condition score (Cook et al. 2001). We used rLIVINDEX to calculate ingesta-free body fat where, ingesta-free body fat = 2.920*rLIVINDEX-0.496 (Cook et al. 2007). We incorporated data on nutritional condition with that reported by Kucera (1988) and Pierce (1999). We calculated ingesta-free body fat from kidney fat indices collected 1984-1996 (Kucera 1988, Pierce 1999) following equations provided by (Stephenson et al. 2002).

During March captures, we used ultrasonography with an Aloka 210 portable ultrasound unit (Aloka, Inc., Wallinford, Connecticut USA) with a 3-MHz

Page 7: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 7

linear transducer to determine pregnancy and fetal rates of captured females during the second trimester of pregnancy (Stephenson et al. 1995). We shaved the left caudal abdomen behind the last rib and applied lubricant to facilitate transabdominal scanning. Presence and number of fetuses were determined, as well as fetal-eye diameter when fetus orientation was suitable for measurement. Upon completion of ultrasonography, numerous pregnant females were fitted with vaginal implant transmitters (VITs) during 2006-2008. Chemical immobilization was not necessary to fit females with vaginal implants (Bishop et al. 2006), thus all animals were physically restrained during processing. In pregnant females, we inserted VITs using a technique similar to that described in detail by Bishop et al. (2006). We placed VITs approximately 20 cm into the vaginal canal or until the silicone wings were pressed firmly against the cervix. We used VITs (M3930, Advanced Telemetry Systems, Isanti, Minnesota), which have been sufficiently described in detail elsewhere (Bowman and Jacobson 1998, Carstensen et al. 2003, Johnstone-Yellin et al. 2006) to identify and characterize birthing habitat and facilitate locating and capturing neonatal mule deer. Antennas of the VITs were approximately 9.5 cm in length and the tips were encapsulated in a resin bead to avoid fraying of the antenna. We used a temperature-sensitive switch (Carstenson et al. 2003, Johnstone-Yellin et al. 2006, Bishop et al. 2006) that increased pulse rate from 40 pulses to 80 pulses per minute when the temperature decreased below 32° C representative of the VIT being expelled by the deer. Vaginal implant transmitters have been employed without any reproductive problems or effects on female survival (Bowman and Jacobson 1998, Carstensen et al. 2003, Bishop et al. 2006) and are likely a practical technique for locating birth sites and subsequently, the capture of neonates (Carstensen 2003, Johnstone-Yellin 2006, Bishop et al. 2007).

Neonate capture and handling During 3 years, 2006-2008, we located and captured neonatal mule deer

from 15 June-20 July by searching for and observing females that exhibited postpartum behavior and by the use of VITs (White et al. 1972, Bishop et al. 2006). We located radiocollared females and monitored VITs for evidence of parturition at first daylight on a daily basis during the typical parturition period (i.e., 15 June – 31 July) using a Cessna 180 fitted with 2 2-element Yagi antennas. When a VIT exhibited a fast pulse (i.e., postpartum), the fixed-wing pilot acquired the approximate location of the implant and ground telemetry was employed to immediately locate the VIT. We used the location of the VIT and the location and behavior of the female to identify the search area. When searches failed to produce neonates, we evaluated whether the location of VIT was an actual birthsite and confirmed that supposition by observing pregnancy status and behavior of female. If the female appeared to have undergone parturition and if personnel were available, we attempted to observe the female from a safe distance to avoid disturbing the female and utilized her postpartum behavior to locate neonates. We also opportunistically observed random females at first light with binoculars or spotting scopes and focused on those that exhibited maternal behavior to locate neonatal mule deer (Carstenson Powell et al. 2003). When a

Page 8: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 8

neonate was located, we hiked to the area the neonate had bedded and conducted ground searches to locate and capture the neonate.

When we located neonatal mule deer, we captured them by hand and placed them in a cloth bag containing sage brush to minimize scent transfer, although that likely would have had little influence on potential abandonment (Carstensen Powell et al. 2005). We determined sex of each neonate and approximate age by appearance of hooves, umbilicus, and passiveness (Haugen and Speake 1958). We also acquired a measurement of new hoof growth using a dial calipers to calculate age of neonates >2 days old (Brinkman et al. 2004, Robinette et al. 1973). We determined the weight of each fawn to the nearest 0.10 kg within the cloth bag using a hand-held spring scale. We characterized and recorded the global positioning system (GPS) location of each capture site and processed all fawns as quickly as possible to minimize the potential for abandonment or attraction of predators (Livezey 1990). We fit all fawns with an expandable radiocollar (Advanced Telemetry Systems, Inc., Isanti, MN; Telemetry Solutions, Walnut Creek, CA) with a 4-hour mortality delay.

Survival and cause-specific mortality We monitored all radiocollared animals with ground telemetry on winter

range approximately 3 days per week from October through April. We monitored radiocollared neonates from a fixed-wing aircraft and ground-based radiotelemetry daily from capture until 31 August, when risk of mortality is relatively high, and approximately 3 days per week thereafter. Frequent monitoring of all animals on winter range and of neonates during summer allowed us to determine cause of most deaths. When mortality was detected, ground telemetry was used to locate the carcass as quickly as possible. When located, we examined carcass to estimate time of death and condition of the animal. We evaluated and recorded the location and arrangement of the carcass, presence and position of tooth marks, ante- and postmortem bleeding or bruising, fractures and remaining organs. We identified other physical evidence of predation including tracks and scat (Elbroch 2003) and collected hair for further confirmation (Moore et al. 1997). When the cause of death could not be ascertained, the carcass was either taken from the field to the laboratory to be necropsied or field necropsies were performed when animal location precluded removal from field.

We classified proximate causes of death as: 1) predation, if an animal died due to predation regardless of whether exact predator was identified; 2) malnutrition, included small and weak neonates and adults with <12% femur marrow fat; 3) disease, if carcass was intact, did not show signs of predation, starvation, or trauma, and postmortem examination indicated infection or disease; 4) accident, if a carcass was located mainly intact with broken bones or other premortem physical trauma and included deer-vehicle collisions and drowning; 5) other natural, if cause of death could not be ascertained, but carcass was primarily intact with no signs of predation; and 6) undetermined, if cause of death could not be categorized in the aforementioned categories or lack of evidence precluded determination of cause of death. For predation related mortalities, we attempted to identify the predator responsible for the death. For

Page 9: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 9

neonates we also included abandonment and stillborns as a proximate cause of mortality. We considered a neonate as abandoned if it was apparently healthy with postmortem necropsy revealing an empty omasum, indicative of an absence of nursing (Church 1988). We considered the proximate cause of mortality as the ultimate cause except for those adult animals that had femur marrow fat <12% (Neiland 1970). Femur marrow fat <12% is indicative of acute starvation, thus we considered the cause of mortality as malnutrition, regardless of proximate indicators (Depperschmidt et al. 1987, Ratcliffe 1980).

During summer, we only opportunistically monitored adult females for survival on summer range because of logistical constraints and the vast expanse of summer range. Infrequent monitoring often precluded determination of cause of death, but provided an estimate of survival. We attempted to locate each animal at least once during 15 June – 30 September to determine summer occupancy and migratory status. We grouped animals by their migratory status as an “east” or “west” side animal by its use of summer range on the east or west side of the Sierra crest.

During each autumn, we determined recruitment status of each marked female as they arrived to winter range in late-Oct through Nov. We located each radiocollared female using ground-based telemetry and stalked within visible range. We observed each female using binoculars and/or spotting scopes until we could confidently determine the number of young-at-heel. We identified the number of young-at-heel by observing nursing and other maternal behavior (Geist 1981). We defined recruitment status of radiocollared females as the number of young-at-heel identified each autumn.

Population Survey We conducted 2 helicopter surveys during each January to (1) estimate

the number of deer wintering in Round Valley and (2) the proportion of adult males, adult females, and young. We completed surveys in a Bell Jet Ranger 206 BIII with 3 observers and the doors removed to improve visibility (Clancy 1999). Two observers in the rear of the helicopter determined the number of deer in each group on either side of the helicopter and a third member of the crew seated alongside the pilot recorded deer on the centerline that were not noted by the other observers. We noted the number of marked deer in each group, but did not classify deer with respect to age class or gender, thereby alleviating the need for the pilot to deviate from and then attempt to return to the transect line. The pilot made every effort to maintain an elevation of 25 m AGL and a ground speed of approximately 40 knots. For mark-resight surveys, we flew a series of parellel transects that were perpendicular to a baseline that bisected the study area (Norton-Griffiths 1978). Location of the initial transect was established randomly, but subsequent transects were parallel to the initial transect and spaced at intervals of approximately 0.5 km. Sampling was considered systematic without replacement. Observers recorded the group size and number of marked animals in each group observed within 200m of centerline along transects.

For population composition surveys, we flew aerial transects and classified deer as encountered with 3 observers and thus, sampling was

Page 10: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 10

systematic without replacement. For each group observed, we carefully identified the size and composition of each group, which included adult males, adult females, and young.

Analyses For some analyses, we combined our results with those collected by

Kucera (1988) and Pierce (1999) to expand our available data to approximately 1985 for many population characteristics including, population size, body fat, and fetal and pregnancy rates. We used Chapman’s (1951) modification of the Lincoln-Peterson estimator to calculate an unbiased estimate of population size from the mark-resight data collected during the annual helicopter surveys. We used one- and two- way analysis of covariance (ANCOVA) to compare levels of ingesta-free body fat between years, fetal rates, recruitment status, and summer residency of adult females (Zar 1999). Age was included as the covariate in all ANCOVA’s to control for the effect of age on various reproductive variables. To determine the influence of summer residency and year on autumn recruitment status of adult females we used multi-way, log-linear models (Zar 1999). We also used log-linear models to determine the influence of year, sex, and summer residency on the frequency of causes of mortality of neonatal mule deer that were radiomarked during 2006-2008.

We calculated monthly survival up to 5 months for neonates and annual and seasonal (i.e., winter and summer) survival for adult female mule deer using the Kaplan-Meyer procedure modified to allow staggered entry and exit of marked animals (Pollock et al. 1989). We censored all mortalities of adult females that occurred within 14 days of capture by helicopter, regardless of proximate cause. We evaluated differences in survival rates among years and specific animal groups using a generalized Chi-square test in program CONTRAST (Sauer and Williams 1989) and maintained experiment-wise error rate for multiple comparisons using Bonferroni corrections (Neter et al. 1996). We used simple linear regression to evaluate the relationship between estimated population parameters, λ, and environmental variation (Neter et al. 1996).

RESULTS AND DISCUSSION

Mule deer inhabiting winter range in Round Valley have been subjected to the vagaries of climate during the last 20 years with further influences of density dependence (Kucera 1988) and thus have exhibited great variation in population size. Kucera (1988) observed a marked decline in the population of deer wintering in Round Valley during a severe drought. Total numbers declined from 5,978 animals (66/km2) in 1985 to a low of 939 (10/km2) in 1991 (Kucera 1988, Pierce et al. 2000). During the population decline, pregnancy rates, fetal rates, fetal sizes, adult weights and kidney fat varied with precipitation and forage growth on winter range (Kucera 1988). Kucera (1988) also concluded that influences of climate conditions and density dependence contributed to poor animal condition and population decline. Following the prolonged drought and population low in 1991, numbers increased to 1,931 (21/km2) in 1997 (Pierce et al. 2000) when we began to estimate body fat of live animals. Since then, we

Page 11: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 11

have continued to acquire population estimates using mark-resight techniques. The population of mule deer in Round Valley has slowly recovered and increased to near 3,000 animals (33/km2), but has leveled off and initiated a decline during recent years (Figure 3). Those patterns in population growth were associated with a concomitant decline in nutritional condition of females that spent winters in Round Valley (Figure 4). Following the population crash in 1991, nutritional condition of females improved and the population increased to about 2,000 animals. As the population size reached an asymptote near 2,000 animals in 1997-1998, nutritional condition began to decline and continued to exhibit a declining trend through 2009. Body fat (mean = 2.6%) of females in Round Valley during March 2009 was the lowest observed during the past 15 years and substantially less than peak values (mean = 9.9%) observed during March 1996.

Annual survival (Jan-Dec) of adult females has been variable from 1998 to 2008 (Figure 5), has ranged from 80% to 100%, and differed among years (χ2=175.7, df=10, P<0.001). Seasonal survival also exhibited similar variability (Figure 6,7). Winter survival (16 Oct – 15 Apr) was typically >90% (Figure 8), whereas summer survival (16 Apr – 15 Oct) of adult females was <90% during 2005-2008 (Figure 7). Furthermore, winter (χ2=73.0, df=9, P<0.001) and summer (χ2=69.5, df=10, P<0.001) survival differed among years. During 1998-2009, mean summer and winter survival, however, was not different (93.6%; (χ2=0.0009, df=1, P=1.0). Annual survival of adult female mule deer in 3 western states was similar among states and among years, despite annual variation in climatic conditions within the Intermountain West (Unsworth et al. 1999). The increased variability in adult survival that we observed was partially related to the substantial changes in population size of mule deer in Round Valley as they recovered from the population crash after 1991. Survival was high in 1998, declined until 2001, and has been variable until 2009 as the population has likely approached K because nutritional condition of adult females has continued to decline, predisposing them to mortality (Bender et al. 2007).

We also considered annual and seasonal survival of adult females separately based on their migratory designation, whether they resided east or west of the Sierra crest during the summer (Figure 8). Overall annual survival for east-side females (87.8%) was lower compared with west-side females (89.0%), but was statistically similar (χ2=0.44, df=1, P=0.51). Likewise, both summer and winter survival for east-side females (93.6% and 93.8%; respectively) was lower than west-side females (94.3% and 96.1%; respectively), but summer and winter survival between side of crest did not differ (χ2=0.17, df=1, P=0.69; χ2=2.5, df=1, P=0.22).

Recruitment rate (i.e., number of young-at-heel in autumn) of radio-marked females ranged from 0.35–0.66 young-at-heel during 1998-2008 (Figure 9) and varied significantly by year (χ2=54.6, df=36, P=0.024). Nevertheless, number of young recruited per adult female was dependent on both year and side-of-crest (χ2=112.0, df=87, P=0.037). During most years, number of recruited young was greater for females that summered east of the Sierra crest than those that resided west of the crest (Figure 10). Indeed, east-side females recruited significantly more young than west-side females (χ2=17.72, df=3, P<0.001).

Page 12: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 12

Body fat in November from 2002-2008 was highly variable (Figure 11) and differed among years (F6,335=5.1, P<0.001). Depending upon recruitment status, mean body fat in November varied from 1 – 13%; biological maxima for mule deer exceeds 20% (Cook et al. 2007). Moreover, females that resided on the west side of the Sierra crest during the summer on average had 67% more body fat (Figure 12) and differed from east-side females (F6,335=74.6, P<0.001). Nevertheless, body fat of females in November was dependent on the number of young recruited to winter range in Round Valley. Females that were unsuccessful in recruiting young were consistently in the best nutritional condition in November, followed by those that recruited 1 or more young (Figure 13). Indeed, body fat differed by recruitment status (F6,269=23.8, P<0.001), but also by side of crest (F6,269=34.1, P<0.001). Females that summered west of the Sierra crest had more body fat with respect to recruitment status compared with east-side females (Figure 14). Moreover, recruitment status of each female in November held residual influence through the winter, with increasing levels of recruitment resulting in decreased body fat the following March (F3,437=4.04, P=0.007). Accordingly, west-side females were in marginally better nutritional condition with respect to recruitment status (Figure 15; F1,437=3.3, P=0.070). Considering body condition of large herbivores has an influence on probability of survival (Bender et al. 2007, 2008), somatic costs associated with reproduction inherently holds a trade-off associated with survival of the female. That tradeoff may explain the slightly lower survival exhibited by east-side females because they typically experienced higher costs of reproduction compared with west-side females. Forage quality and quantity are greatest during the growing season and thus, fat and protein deposition during summer are typically greatest and prepare animals for the food limited winter months. Nonetheless, nutritional requirements for late gestation, lactation, and growth of young are taxing for large herbivores (Moen 1978, Sadlier 1982); consequently, reproductive status during summer had a direct influence on body condition of females prior to winter (Figure 14). When forage resources are sufficient, however, reproductive females may be capable of attaining similar fat reserves by mid-autumn compared with non-lactating females (Cook et al. 2004, Piasecke and Bender 2009). Generally, females are thought to be in best condition with the onset of winter (Mautz 1978, Parker et al. 2009). The influence of the effects of summer range and reproductive status on body condition of females prior to the onset of winter underscores the importance of summer range in providing adequate forage for reproduction and accumulation of fat reserves for winter (Mautz 1978, Cook et al. 2004, Parker et al. 2009).

Seasonal patterns of nutritional condition followed consistent changes from autumn to late-winter. Adult females were had consistently less body fat following winter compared with autumn body fat; on average, they lost from 20 – 80% of their fat reserves. (Figure 16). Body condition of adult females in March varied by year (F12,836=23.7, P<0.001) and on average, body fat of west-side females differed and was 11.7% greater than east-side females (F3,438=29.5, P=0.002). Both temperate and northern herbivores track seasonal cycles in nutrient intake following seasonal changes in quality and quantity of food

Page 13: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 13

resources, with marked declines during the dormant season. Seasonal changes in food availability and energetic costs lead to concurrent cycles in deposition and catabolism of fat (Mautz 1978), the primary energy reserve of the body. Conditions on winter range are typically characterized by a decline in quantity and quality of forage; consequently, ungulates have adapted physiologically and behaviorally to decrease metabolic rate, nutrient demand, and activity (Moen 1978, Parker et al. 1993, Taillon et al. 2006). Nevertheless, catabolism of body fat for maintenance normally occurs during winter and may be exacerbated when foraging conditions are poor (Short et al. 1966, Torbit et al. 1985).

Fetal rate of adult females wintering in Round Valley was dependent upon year and residency status of female (χ2=112.0, df=87, P=0.037). Overall, fetal rate was greater for females that resided west of the Sierra crest during the summer (Figure 17; χ2=17.7, df=3, P<0.001). West-side females were consistently in better nutritional condition in autumn compared with east-side females because west-side females incurred lower somatic costs of reproduction (Figure 14). Nutritional status plays a considerable role in litter size of adult female cervids (Swihart et al. 1998, Testa and Adams 1998). Nutritional condition of adult females in November affected fetal rate the following March (F3,260=4.8, P=0.003), but pairwise comparisons indicated that differences only occurred between females pregnant with singletons compared with twins (P=0.053). Females pregnant with twins had greater body fat compared with females pregnant with only singletons (Figure 18). Although, nutritional condition for females that were not pregnant and those pregnant with triplets was on average greater than females with singetons or twins, sample size in those 2 groups was insufficient to interpret those results (not pregnant: n=4, triplets: n=5).

We measured neonatal survival during 2006-2008 on both sides of the Sierra crest, but sample size was greater for east-side neonates (Table 1). Prior to 150 days-of-age, we censored 1 animal in 2006 and 4 animals in 2007 because they prematurely shed their collars. We assumed the fate of those animals was independent of the shedding of the collars. Monthly survival of neonates varied by year (χ2=24.0, df=2, P<0.001), and ranged from 31 – 47% with survival to 5 months being greatest in 2006 (Figure 19). Average monthly survival to 5 months was 40.4% (±0.05). Male and female neonates exhibited similar patterns of monthly survival (Figure 20; χ2=1.12, df=1, P=0.28). Monthly survival of young born on the west side of the Sierra crest, however, was approximately 40% lower and differed from young born east of the Sierra crest (Figure 21; χ2=79.0, df=1, P<0.001).

Causes of mortality were similar among the 3 years of study (χ2=17.5, df=14, P=0.231), with predation comprising the greatest amount of neonatal loss in all years (60% of mortality; Table 2). Within 5 months of age, predators had taken 30% or more of young mule deer during the 3 years of this study (Table 2). Although cause-specific mortality was similar between neonates on either side of the Sierra crest (χ2=7.9, df=7, P=0.34), 77% of the deaths of neonates born on the west side of the Sierra crest were caused by predation, whereas, predation was responsible for 48% of neonatal deaths east of the Sierra crest. Of the predation caused mortality, black bears were responsible for 87% west of the

Page 14: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 14

Sierra crest and only 16% on the east-side. Coyotes were responsible for 53% of predation caused mortality east of the Sierra crest compared with only 4% west of the Sierra crest (Table 3).

The ratio of young:adult females and males: adult females has varied considerably since the 1950’s. The young:adult female ratio reached low’s of near 20 young per 100 females in the early 1990’s during the population crash, but then increased to near 60 following the population crash and has declined more recently until 2009 (Figure 22). The male:female ratio has followed a similar pattern with a low of approximately 10 males per 100 females in the early 1990’s, which then increased to 30-40 following the population crash (Figure 23). More recently however, the male:female ratio has decreased to under 20 males per 100 females (Figure 23).

Climate in the Sierra Nevada is highly variable, with April snowpack varying from 0.5 cm to 122.7 cm during 1950-2008 (CV = 77%). In other variable and arid environments, forage abundance and its nutritional quality is inextricably linked to patterns of precipitation (Marshal et al. 2005a, 2005b). Furthermore, timing of seasonal precipitation is important for forage growth, particularly in the Sierra Nevada where the majority of the annual precipitation falls as snow between October and April (Storer et al. 2004, Concilio et al. 2009). We acquired data on leader growth of bitterbrush in Round Valley from the Bureau of Land Management (BLM) for most years. Bureau of Land Management measured new leader growth of bitterbrush during most autumns in Round Valley. Annual growth of bitterbrush, the primary winter forage for mule deer in Round Valley (Kucera 1988), was predicted by water content of April snowpack (r2=0.65, P=0.001; Figure 24). Accordingly, extreme variation in winter precipitation had significant effects on nutritional condition of adult females during winter. April snowpack had a strong positive influence on nutritional condition of adult females in Round Valley the following March (r2=0.53, P<0.001; Figure 25). Similarly, a coarse index to body condition in the Sonoran Desert yielded significant relationships with the accumulated rainfall 6-12 months prior to sampling (Marshal et al. 2008). Subcutaneous fat reserves are mobilized to support metabolic need during periods of forage scarcity (Stephenson et al. 1998, Cook et al. 2001). Forage availability however, is influenced by both forage growth, which is clearly influenced by winter precipitation in the Sierra Nevada, and by intraspecific competition for forage (density dependence).

Prior to the crash in 1991, there was no relationship between population size and nutritional condition (P=0.21) of adult females post-winter; however, following the crash, nutritional condition of mule deer in Round Valley tended towards a density dependent effect (r2=0.17, P=0.088; Figure 26). Within that analysis however, 1993 was a severe outlier. That year was characterized by a wintering deer population at relatively low density, but low winter snowfall the previous year resulted in little new growth of forage in Round Valley, thereby driving the relationship between forage growth and deer condition. With 1993 removed, nutritional condition of mule deer in Round Valley exhibited strong density dependence, with population size having a significant negative effect on nutritional condition (r2=0.38, P=0.009; Figure 26). As a population progresses

Page 15: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 15

towards K, intraspecific competition for forage causes a decline in per capita availability of forage resources, that may include both quality and quantity. Consequently, lower per capita availability of forage resources is reflected in declining body condition as the population nears K (Kie et al. 2003). Nevertheless, this arid system in the western Great Basin is clearly influenced by annual precipitation, which may cause annual deviations from long-term trends in K. Range conditions in the Sierra Nevada are largely influenced by winter snowfall and the corresponding forage growth that results in annual fluctuations in K, a pattern identified in other arid systems (Marshal et al. 2002, 2005a).

Nutritional condition of adult females in March explained much of the variation in both young:adult female and male:adult female ratios the following January. Following the crash in 1991, the male:adult female ratio lagged behind the population response to improved conditions and low density, which resulted in 2 outliers in the relationship between ingesta-free body fat of adult females in March and the male:adult female ratio the following January. With those outliers removed, ingesta-free body fat of adult females in March explained >70% of the variation in the male:adult female ratio the following January (Figure 27). We do not imply cause and effect in this relationship. Clearly, there is no biological mechanism where a declining nutritional condition of females would cause a decrease in the number of adult males. We believe, however, that this relationship suggests the importance of understanding nutritional condition as a collective measure of population health. Nutritional condition of adult females in March represents the physical condition of females following winter, when animals typically experience a negative energy balance and poor foraging conditions (Mautz 1978, Barboza et al. 2009). Adult males follow a slightly different fat cycle compared with females. Females presumably will have the opportunity to replenish some fat reserves that were utilized to support reproduction during the summer before poor foraging conditions prevail on winter range (Cook et al. 2004, Parker et al. 2009). In contrast, adult males expend nutritional reserves to support breeding activities during late-November and early-December. Therefore, it follows that males enter the winter season in poorer physical condition than females and males would, therefore, be most susceptible to factors of mortality on winter range. We believe that nutritional condition of adult females in March serves as a reliable predictor of the number of males relative to females because nutritional condition of females post-winter is a reliable indicator of herd health. As nutritional condition of adult females decline in response to increasing population density or decreasing winter precipitation, we expect that the number of males relative to females will accordingly decline. Although this prediction remains largely untested, the small amount of research on over-winter survival of adult males is suggestive of their higher level of susceptibility to natural mortality factors following the breeding season compared with females (Ditchkoff et al. 2001, Webb et al. 2006). In polygynous mating systems, males are probably more likely than females to succumb to resource limitation (Clutton-Brock et al. 1985). Lower survival of males in ungulate populations is common (Loison et al. 1999) and studies that failed to document intersexual differences in survival were probably expanding

Page 16: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 16

populations that were not food limited (Fancy et al. 1994, Toïgo et al. 1997). Although often considered negligible, the presence and proportion of adult males in ungulate populations can have cascading effects on population demographics (Mysterud et al. 2002). Presence of mature males can stimulate estrus in females and therefore, influences length of breeding season and subsequently, timing of parturition (Mysterud et al. 2002). In turn, timing of parturition can influence forage quality (Bowyer 1991, Côté and Festa-Bianchet 2001), neonatal survival (Bowyer 1991, Feder et al. 2008), and over-winter survival by affecting the time available for mass gain of young prior to the onset of winter (Loison et al. 1999, Holand et al. 2003).

Physical condition of adult females in March was positively associated with the number of young per female recruited to winter range the following January (Figure 28). Maternal condition during the last trimester of gestation affects birth mass and subsequent investment in young (Sams et al. 1996, Pekins et al. 1998, Testa and Adams 1998), which directly influence probability of survival and susceptibility of young to mortality (Julander et al. 1961, Pederson and Harper 1987, Lomas and Bender 2006, Bender et al. 2007). Growth and subsequent production from entire cohorts of animals may be impacted by maternal condition during gestation (Albon et al. 1987, Anderson and Linnell 1997, Rose et al. 1998, Forchhammer et al. 2001). Poor conditions during gestation can result in the transmission of negative maternal effects on offspring that can persist through adulthood and into subsequent generations resulting in time lags in population response to improved conditions (Monteith et al. 2009). Physical condition of maternal females during gestation provides insight into expected reproductive performance the following year, despite confounding influences of summer range conditions (Julander et al. 1961).

Physical condition of adult females in March also explained 50% of the variation in annual survival of those females (Figure 29). Poor nutritional condition of adult females in March provided inference into annual survival due to the influence of nutritional condition on an animals susceptibility to mortality. Previously, poor physical condition was thought to only influence the probability of death of an animal near starvation (<12% bone marrow fat). Nevertheless, recent evidence suggests that animals may be susceptible to mortality factors across a range of physical condition (Bender et al. 2008), which corroborates the relationship between nutritional condition and annual survival.

Adult survival is consistently identified as the vital rate with the greatest elasticity on population growth with λ being relatively insensitive to changes in survival of young (Gaillard et al. 2000, Eberhardt 2002, Garrott et al. 2003). The impact of a demographic rate on population growth is a function of both its influence (elasticity) and its variability (Gaillard et al. 1998, Raithel et al. 2007). Nevertheless, the variability in demographic rates actually determine the temporal variation in λ, not its elasticity (Raithel et al. 2007). We regressed population vital rates, winter precipitation, and physical condition of females to determine their relative influence on λ. Although the relationship between winter precipitation and λ was significant (P = 0.04), winter precipitation explained only a small amount of the variation in λ the following year (r2 = 0.18). Annual survival

Page 17: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 17

of adult females exhibited no relationship with λ (Figure 30). Indeed, mean annual survival of adult females was 89% with a CV of only 8.7%. Small temporal variation of adult survival likely plays a minimal role in influencing annual changes in λ (Raithel et al. 2007). Conversely, with the exception of 1991, recruitment of young (young:female) had a CV of 29%, was positively associated with λ, and explained 45% of the annual variation in λ (Figure 31). Following the population crash in 1991, ingesta-free body fat of adult females in March also was positively associated with λ during the same year (Figure 32). Body fat of adult females in March explained 44% of the variation in lambda during the same year, probably because of the influence of nutritional condition on survival of the female (Figure 29) and their subsequent reproductive investment and probability of survival of young (Figure 28). Body fat integrates energetic conditions experienced by an individual including nutrition as determined by intrinsic habitat value and density dependent competition for forage resources. Consequently, understanding the nutritional status of individuals within a population isolates the role of bottom-up vs. top-down influences on population limitation.

MANAGEMENT IMPLICATIONS

Nutritional condition of adult females in autumn was influenced strongly by reproductive effort expended by females during the summer. Their nutritional condition in autumn then influenced fetal rates as measured the following March. Lactation is the most energetically demanding time of year for female cervids (Moen 1985); therefore, we expected reproductive status in autumn to influence physical condition of females. Increasing recruitment of young resulted in decreasing nutritional condition in November. Nevertheless, the cost of reproduction was greater for females that resided on the east-side of the Sierra crest compared with those on the west side. Habitats and annual moisture regimes differ considerably between either side of the Sierra crest (Storer et al. 2004). The more mesic environment and lower deer densities on the west-side of the Sierra crest probably resulted in better foraging conditions for deer, thereby, lessening the somatic costs of reproduction. Less costly reproduction might suggest improved recruitment for those west-side females. Nonetheless, recruitment of young was consistently higher for east-side females since 1997. Higher recruitment by east-side females despite concomitantly higher somatic costs to reproduction suggests that a portion of the mortality of young on the west-side of the Sierra crest had an additive effect on the population. Disparity in recruitment in the absence of a difference in survival between east- and west-side females, was likely responsible for the shift in migratory components of the population. In 1987, Kucera (1988) determined that >80% of the population migrated to the west-side of the Sierra crest during summers. Currently, only about 55% of the marked segment of adult females migrates to west-side summer ranges. Apparently selective pressures in Round Valley mule deer have shifted during the most recent decade in favor of animals that spend summers on the east-side of the Sierra crest. Predation pressures on young born on the

Page 18: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 18

west-side were nearly double that experienced by east-side animals (Table 2) and 87% of that predation was caused by black bears (Table 3), whereas, only 16% of predation was caused by black bears on the east-side of the Sierra crest. The increasing abundance of black bears on the west-side of the Sierra crest has likely played a significant role in influencing the shift in demographics between the east and west-side segments of the Round Valley deer population.

Although removal of predators, namely bears, likely would result in an immediate response in recruitment of young by west-side females, it also would exacerbate the impact of an already limited winter range. Nutritional condition of females during March 2009 indicated that the deer population was at or beyond nutritional K and to expect poor recruitment, poor adult survival, and a population decline during 2009. Increased recruitment by west-side females would simply result in an increase in the density-dependent constraints of winter range. Nevertheless, if managers are concerned of the shifting demographics within the deer population, bear control and habitat manipulations on the west-side of the Sierra crest should be accompanied by either habitat improvements or density reductions on winter range in Round Valley. Considering that habitat manipulations in sage-brush steppe are often very costly and ineffective, however, we suggest that density reduction be employed to lessen competition for limited forage on winter range and lessen the magnitude of another decline (Berryman and Lima 2006, Simard et al. 2008). Intersexual differences in body size, metabolic demand, and digestive function commonly results in partitioning of resources between sexes of dimorphic ruminants (Kie et al. 1999, McCullough 1999, Barboza and Bowyer 2000). Therefore, density reductions should be focused on females to minimize density-dependent effects on nutritional condition and productivity (McCullough 1979, 1999).

Release from major density-dependent influences should result in a more stable and predictable system. With density-dependent effects interacting with environmental variation, minor perturbations in the system can have profound effects on overwinter survival and subsequent reproduction (Kie et al. 2003; Figure 33). Resource limitation has already constrained nutritional condition of mule deer and minor deviations from normal precipitation regimes will cause extreme fluctuations in population growth. Lower population size with respect to K would result in a deer population less susceptible to annual variation in environmental conditions. Nevertheless, severe drought conditions will continue to influence population parameters even when the population is at lower density (1993 for example), but patterns of nutritional condition, survival, and reproduction should be less susceptible to minor deviations in environmental conditions. In populations held below K, harvest may be maintained if managed for maximum sustained yield, but this assumes that predation does not unduly limit recruitment required to support such harvest.

In June 1995, a fire burned 22 km2 of the winter range in an area dominated by bitterbrush and sagebrush. Due to the intensity of the fire, little regrowth of bitterbrush occurred in subsequent years and has become dominated by desert peach and cheat grass (Bromus tectorum), both of which offer little forage value to deer (Pierce et al. 2004). Coincident with the fire was

Page 19: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 19

the closure of an alfalfa ranch that was frequented by hundred’s of deer on a daily basis. The devastating fire and closure of the alfalfa ranch clearly resulted in a decline in K of winter range because of the loss in quantity of critical winter forage, which likely prevented recovery of the population to the 6000 animals present in 1985 (Figure 3). Furthermore, acute overbrowsing of forage species on winter range may have influenced subsequent growth and recovery of habitat in subsequent years (Kie et al. 2003, Simard et al. 2008). The marked changes in density that the mule deer population in Round Valley has experienced may result in continued transient dynamics caused by the effects of severe resource limitation on growth and development of young, age at first reproduction, entire cohorts, and causing negative maternal effects (Eberhardt 2002, Coulson et al. 2004, Monteith et al. 2009). Lagged effects from severe nutritional limitation caused by the increased deer density in Round Valley and the corresponding changes in habitat condition may continue to persist for decades (Coulson et al. 2004, Monteith et al. 2009). Current body condition of adult females following winter in Round Valley clearly indicates continued resource limitation through density dependent feedbacks and variation in winter precipitation.

Nutritional condition has an influence on nearly every component of survival and reproduction of large herbivores (Cameron and Ver Hoef 1994, Testa and Adams 1998, Keech et al. 2000, Stewart et al. 2005, Bender et al. 2007). Furthermore, others have identified animal condition as a valuable tool to collate population health (Parker et al. 2009, Barboza et al. 2009). Nutritional condition of adult female mule deer in Round Valley following winter was sensitive to conditions of winter habitat because it was influenced by intraspecific competition for forage (Figure 26) and the availability of new growth of the bitterbrush as influenced by water content of winter snowpack the previous April (Figure 25). Moreover, nutritional condition in March was related to male:female and young:female ratios the following autumn (Figure 27,28), annual survival of adult females (Figure 29), and itself, held predictive value for population growth during the same year (Figure 30). Physical condition of adult females following winter provided an encompassing measure of population health and expected measures of adult female survival, recruitment of young, an index to abundance of males, and overall population trajectory during the following year. Nutritional limitations on survival and recruitment of deer in Round Valley clearly indicate bottom-up limitation on deer dynamics and the importance of data on nutritional condition when interpreting dynamics of deer populations. Consequently, levels of nutritional condition within a population may be used to determine carrying capacity (Piasecke and Bender 2009).

The patterns of population demographics collected from research on mule deer in Round Valley also should be applicable to other deer populations in California, particularly those with winter ranges that reside in the western Great Basin, which includes the Gooddale, White Mountains, Walker, Mono Lake, Casa Diablo, and other mule deer populations. Climatic variation has an influence on mule deer in arid regions (Marshal et al. 2005, 2005a; this study). Patterns of precipitation have an indirect influence on body condition through its direct influence on forage quality and quantity. Nutritional condition then influences

Page 20: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 20

nearly all population parameters. Region-wide variability in climatic conditions should be similar in the western Great Basin, therefore, dynamics of deer populations in that area are likely correlated. For example, across Idaho, Colorado, and Montana, Unsworth et al. (1999) determined that general patterns in dynamics of mule deer populations were similar despite some difference in climate and geography. Mule deer populations likely respond to large-scale climatic variation across wide geographical areas.

Overall, the research that has been conducted on the population of mule deer in Round Valley has provided new and valuable insight into research methodology, reproductive ecology, population dynamics, and management of mule deer in a highly variable environment (Table 4). The data collected from this population will continue to advance wildlife science during the next few years as the current dataset is more rigorously analyzed and interpreted for peer-reviewed publication, and to provide the Department with practical management recommendations. Numerous peer-reviewed publications are in progress and are forthcoming in the next few years (Table 5).

ACKNOWLEDGMENTS An effort of this magnitude and for this duration is exceedingly rare in wildlife studies, particularly those on large mammals, and would not have been possible without the dedication of countless biologists, seasonal employees, and volunteers. We are thankful for the assistance of fixed-wing pilots for CDFG T. Evans and G. Schales, and also for the professionalism and dedication of G. Pope of Black Mountain Air Service and R. Nixon. Landell’s Aviation was largely responsible for the helicopter capture and survey work, especially, S. DeJesus and his support crew, C. Pickrell. We are also very appreciative to have worked with numerous dedicated field assistants, including P. McGrath, J. Augustine, D. Spitz, A. Howell, T. Swearingen, W. Allsup, A. Stephenson, M. Kiner, T. Branston, C. Schroeder, H. Johnson, M. Leonard-Cahn, K. Monteith, R. Long, and S. Monteith. We are also indebted to volunteers N. Partridge, P. Partridge, and D. DeJesus for their selfless dedication to Round Valley deer research. Support for capture work was provided by numerous individuals including, J. Villepique, J. McKeever, T. Taylor, B. Clark, and B. Adams. Also, capture support was provided by the Wildlife Investigations Lab including, S. Torres, K. Jones, A. Hunter, B. Gonzales, T. Glenner, P. Swift, J. Schultz, and H. Zurawka. Funding was provided by the California Deer Association, Mule Deer Foundation, and Granite Bay Safari Club International. The bulk of the funding and logistical support was provided by the California Department of Fish and Game.

Page 21: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 21

LITERATURE CITED Albon, S. D., T. H. Clutton-Brock, F. E. Guinness. 1987. Early development and

population dynamics in red deer. II. Density-independent effects and cohort variation. Journal of Animal Ecology 56:69-81.

Andersen, R., and J. D. C. Linnell. 1997. Variation in maternal investment in a small cervid; the effects of cohort, sex, litter size and time of birth in roe deer (Capreolus capreolus) fawns. Oecologia 109:74-79.

Anderson, R., and J. D. C. Linnell. 1997. Variation in maternal investment in a small cervid; the effects of cohort, sex, litter size and time of birth in roe deer (Capreolus capreolus) fawns. Oecologia 109:74-79.

Barboza, P. S., and R. T. Bowyer. 2000. Sexual segregation in dimorphic deer: a new gastrocentric hypothesis. Journal of Mammalogy 81(2):473-489.

Barboza, P. S., K. L. Parker, and I. D. Hume. 2009. Integrative wildlife nutrition. Springer, Verlag Berlin Heidelberg, Germany.

Barrett, M. W., J. W. Nolan, and L. D. Roy. 1982. Evaluation of a hand-held net-gun to capture large mammals. Wildlife Society Bulletin 10:108-114.

Bender, L. C., J. G. Cook, R. C. Cook, and P. B. Hall. 2008. Relations between nutritional condition and survival of North American elk Cervus elaphus. Wildlife Biology 14:70-80.

Bender, L. C., L. A. Lomas, and J. Browning. 2007. Condition, survival, and cause-specific mortality of adult female mule deer in north-central New Mexico. Journal of Wildlife Management 71:1118-1124.

Berryman, A., and L. Mauricio. 2006. Deciphering the effects of climate on animal populations: diagnostic analysis provided new interpretation of soay sheep dynamics. The American Naturalist 168:784-795.

Bishop, C. J., D. J. Freddy, G. C. White, B. E. Watkins, T. R. Stephenson, and L. L. Wolfe. 2007. Using vaginal implant transmitters to aid in capture of mule deer neonates. Journal of Wildlife Management 71:945-954.

Bleich, V. C., T. R. Stephenson, B. M. Pierce, and M. J. Warner. 2007. Body condition of radio-collared mule deer while injured and following recovery. Southwestern Naturalist 52:164-167.

Bleich, V. C., T. R. Stephenson, N. J. Holste, I. C. Snyder, J. P. Marshal, P. W. McGrath, and B. M. Pierce. 2003. Effects of tooth extraction on body condition and reproduction of mule deer. Wildlife Society Bulletin 31:233-236.

Bowden, D. C., G. C. White, and R. M. Bartmann. 2000. Optimal allocation of sampling effort for monitoring a harvested mule deer population. Journal of Wildlife Management 64:1013-1024.

Bowman, J. L., and H. A. Jacobson. 1998. An improved vaginal-implant transmitter for locating white-tailed deer birth sites and fawns. Wildlife Society Bulletin 26:295-298.

Bowyer, R. T. 1991. Timing of parturition and lactation in southern mule deer. Journal of Mammalogy 72:138-145.

Bowyer, R. T., D. K. Person, and B. M. Pierce. 2005. Detecting top-down versus bottom-up regulation of ungulates by large carnivores: implications for

Page 22: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 22

biodiversity. Pages 342-361 in J. C. Ray, K. H. Redford, R. S. Steneck, and J. Berger, editors. Large carnivores and the conservation of biodiversity. Island Press, Washington DC, USA.

Bowyer, R. T., D. M. Leslie, Jr., and R. L. Rachlow. 2000. Dall’s and Stone’s sheep. Pages 491–516 in S. Demarais, and P.R. Krausman, editors. Ecology and Management of Large Mammals in North America. PrenticeHall, Upper Saddle River, New Jersey, USA.

Brinkman, T. J., K. L. Monteith, J. A. Jenks, and C. S. DePerno. 2004. Predicting neonatal age of white-tailed deer in the Northern Great Plains. The Prairie Naturalist 36: 75-81.

Brown, G. S., L. Landriault, D. J. H. Sleep, and F. F. Mallory. 2007. Comment arising from a paper by Wittmer et al.: hypothesis testing for top-down and bottom-up effects in woodland caribou population dynamics. Oecologia 154:485-492.

Cameron, R. D., and J. M. Ver Hoef. 1994. Predicting parturition rate of caribou from autumn body mass. Journal of Wildlife Management 58:674-679.

Carstensen Powell, M., G. D. DelGiudice, and B. A. Sampson. 2003. Using doe behavior and vaginal-implant transmitters to capture neonate white-tailed deer in north-central Minnesota. Wildlife Society Bulletin 31:634–641.

Carstensen Powell, M., G. D. DelGiudice, and B. A. Sampson. 2005. Low risk of marking-induced abandonment in free-ranging white-tailed deer neonates. Wildlife Society Bulletin 33:643–655.

Caughley, G. 1977. Analysis of vertebrate populations. The Blackburn Press, Caldwell, New Jersey, USA.

Chapman, D. G. 1951. Some properties of the hypergeometric distribution with application to zoological sample censuses. University of California Publications in Statistics 1:131-160.

Church, D. C. 1988. The ruminant animal. Prentice-Hall, Inc., Englewood Cliffs, New Jersey, USA.

Clancy, T. F. 1999. Choice of survey platforms and technique for broad-scale monitoring of kangaroo populations. Australian Zoologist 31:267-274.

Clutton-Brock, T. H., M. Major, and F. E. Guinness. 1985. Population regulation in male and female red deer. Journal of Animal Ecology 54:831-846.

Compton, B. B., P. Zager, and G. Servheen. 1995. Survival and mortality of translocated woodland caribou. Wildlife Society Bulletin 23:490-496.

Connolly, G. E., 1981. Assessing populations. Pages 287-245 in O. C. Wallmo, editor. Mule deer and black-tailed deer of North America. University of Nebraska Press, Lincoln, Nebreaska, USA.

Cook, J. C., B. K. Johnson, R. C. Cook, R. A. Riggs, T. Delcurto, L. D. Bryant, and L. L. Irwin. 2004. Effects of summer-autumn nutrition and parturition date on reproduction and survival of elk. Wildlife Monographs No. 155.

Cook, R. C., J. G. Cook, D. L. Murray, P. Zager, B. K. Johnson, and M. W. Gratson. 2001. Development of predictive models of nutritional condition for rocky mountain elk. Journal of Wildlife Management 65:973-987.

Page 23: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 23

Cook, R. C., T. R. Stephenson, W. L. Myers, J. G. Cook, and L. A. Shipley. 2007. Validating predictive models of nutritional condition for mule deer. Journal of Wildlife Management 71:1934-1943.

Côté, S. D., and M. Festa-Bianchet. 2001. Birthdate, mass and survival in mountain goat kids: effects of maternal characteristics and forage quality. Oecologia 127:230-238.

Côté, S.D., and M. Festa-Bianchet. 2001. Birthdate, mass and survival in mountain goat kids: effects of maternal characteristics and forage quality. Oecologia 127:230-238.

Coulson, T., E. A. Catchpole, S. D. Albon, B. J. T. Morgan, J. M. Pemberton, T. H. Clutton-Brock, M. J. Crawley, and B. T. Grenfell. 2001. Age, sex, density, winter weather, and population crashes in Soay sheep. Science 292:1528-1531.

Coulson, T., F. Giunness, J. Pemberton, and T. Clutton-Brock. 2004. The demographic consequences of releasing a population of red deer from culling. Ecology 85:411-422.

Cransac, N., A. J. M. Hewison, J.-M. Gaillard, J.-M. Cugnasse, and M. L. Maublanc. 1997. Patterns of mouflon (Ovis gmelini) survival under moderate environmental conditions: effects of sex, age and epizootics. Canadian Journal of Zoology 75:1867-1875.

de Vos, Jr. J. C., M. R. Conover, and N. E. Headrick. 2003. Mule deer conservation: issues and management strategies. Berryman Institute Press, Utah State University, Logan, Utah, USA.

Depperschmidt, J. D., S. C. Torbit, A. W. Alldredge, and R. D. Deblinger. 1987. Body condition indices for starved pronghorn. Journal of Wildlife Management 51:675-678.

Ditchkoff, S. S., E. R. Welch, Jr., R. L. Lochmiller, R. E. Masters, and W. R. Starry. 2001. Age-specific causes of mortality among male white-tailed deer support mate-competition theory. Journal of Wildlife Management 63:552-559.

Eberhardt, L. L. 2002. A paradigm for population analysis of long-lived vertebrates. Ecology 83:2841-2854.

Elbroch, M. 2003. Mammal tracks and sign. Stackpole Books, Mechanicsburg, Pennsylvania.

Fancy, S. G., K. R. Whitten, and D. E. Russell. 1994. Demography of the Porcupine caribou herd 1983-1992. Canadian Journal of Zoology 72:840-846.

Feder, C., J. G. A. Martin, M. Festa-Bianchet, C. Bérubé, and J. Jorgenson. 2008. Never too late? Consequences of late birthdate for mass and survival of bighorn lambs. Oecologia 156:773-781.

Forchhammer, M. C., T. H. Clutton-Brock, J. Lindstrom, and S. D. Albon. 2001. Climate and population density induce long-term cohort variation in a northern ungulate. Journal of Animal Ecology 70:721-729.

Gaillard, J. -M., M. Festa-Bianchet, N. G. Yoccoz, A. Loison, and C. Toigo. 2000. Temporal variation in fitness components and population dynamics of large herbivores. Annual Review of Ecology and Systematics 31:367-393.

Page 24: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 24

Gaillard, J.-M., M. Festa-Bianchet, and N. G. Yoccoz. 1998. Population dynamics of large herbivores: variable recruitment with constant adult survival. Trends in Ecology and Evolution 13:58-63.

Garrott, R. A., L. L. Eberhardt, P. J. White, and J. Rotella. 2003. Climate-induced variation in vital rates of an unharvested large-herbivore population. Canadian Journal of Zoology 81:33-45.

Giest, V. 1981. Behavior: adaptive strategies in mule deer. Pages 157-224 in O. C. Wallmo, editor. Mule and black-tailed deer of North America. Univeristy of Nebraska Press, Lincoln, Nebraska, USA.

Hairston, N. G., F. E. Smith, and L. B. Slobodkin. 1960. Community structure, population control, and competition. American Naturalist 94:421-425.

Haugen, A. O., and D. W. Speake. 1958. Determining age of young fawn white-tailed deer. The Journal of Wildlife Management 22:319-321.

Holand, Ø, K. H. Røed, A. Mysterud, J. Kumpula, M. Nieminen, and M. E. Smith. 2003. The effect of sex ratio and male age structure on reindeer calving. The Journal of Wildlife Management 67:25-33.

Holand, Ø., K. H. Røed, A. Mysterud, J. M. Kumpula, M. Nieminen, and M. E. Smith. 2003. The effect of sex ratio and male age structure on reindeer calving. Journal of Wildlife Management 67:25-33.

Johnstone-Yellin, T. L., L. A. Shipley, and W. L. Myers. 2006. Effectiveness of vaginal implant transmitters for locating neonatal mule deer fawns. Wildlife Society Bulletin 34:338-344.

Julander, O., W. L. Robinette, and D. A. Jones. 1961. Relation of summer range condition to mule deer herd productivity. Journal of Wildlife Management 25:54-60.

Keech, M. A., R. T. Bowyer, J. M. Ver Hoef, R. D. Boertje, B. W. Dale, and T. R. Stephenson. 2000. Life history consequences of maternal condition in Alaskan moose. Journal of Wildlife Management 64:450-462.

Kie, J. G., and R. T. Bowyer. 1999. Sexual segregation in white-tailed deer: density-dependent changes in use of space, habitat selection, and dietary niche. Journal of Mammalogy 80:1004-1020.

Kie, J. G., R. T. Bowyer, and K. M. Stewart. 2003. Ungulates in western forests: habitat relationships, population dynamics, and ecosystem processes. Pages 296-340 in C. Zabel and R. Anthony, editors. Mammal community dynamics in western coniferous forests: management and conservation. The Johns Hopkins University Press, Baltimore, Maryland.

Krausman, P. R., J. J. Hervert, and L. L. Ordway. 1985. Capturing deer and mountain sheep with a net-gun. Wildlife Society Bulletin 13:71-73.

Kucera, T. E. 1988. Ecology and population dynamics of mule deer in the Eastern Sierra Nevada, California. Ph.D. Dissertation. University of California, Berkely, California, USA.

Kucera, T. E. 1991. Adaptive variation in sex ratios of offspring in nutritionally stressed mule deer. Journal of Mammalogy 72:745-749.

Kucera, T. E. 1992. Influences of sex and weather on migration of mule deer in California. Great Basin Naturalist 52:122-130.

Page 25: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 25

Kucera, T. E. 1997. Fecal indicators, diet, and population parameters in mule deer. Journal of Wildlife Management 61:550-560.

Lande, R., S. Engen, B.-E. Sæther, R. Filli, E. Matthysen, and H. Weimerskirch. 2002. Estimating density dependence from population time series using demographic theory and life-history data. American Naturalist 159:321-337.

Livezey, K. B. 1990. Toward the reduction of marking-induced abandonment of newborn ungulates. Wildlife Society Bulletin 19:193-203.

Loison, A., R. Langvatn, and E. J. Solberg. 1999. Body mass and winter mortality in red deer calves: disentangling sex and climate effects. Ecography 22:20-30.

Lomas, L. A., and L. C. Bender. 2007. Survival and cause-specific mortality of neonatal mule deer fawns, north-central New Mexico. Journal of Wildlife Management 71:884-894.

Marshal, J. P., P. R. Krausman, and V. C. Bleich. 2005a. Dynamics of mule deer forage in the Sonoran Desert. Journal of Arid Environments 60:593-609.

Marshal, J. P., P. R. Krausman, and V. C. Bleich. 2005b. Rainfall, temperature, and forage dynamics affect nutritional quality of desert mule deer forage. Rangeland Ecology and Management 58:360-365.

Marshal, J. P., P. R. Krausman, and V. C. Bleich. 2008. Body condition of mule deer in the Sonoran Desert is related to rainfall. The Southwestern Naturalist 53:311-318.

Marshal, J. P., P. R. Krausman, V. C. Bleich, W. B. Ballard, and J. S. McKeever. 2002. Rainfal, El Niño, and dynamics of mule deer in the Sonoran Desert, California. Journal of Wildlife Management 66:1283-1289.

Mautz, W. W. 1978. Sledding on a bushy hillside: the fat cycle in deer. Wildlife Society Bulletin 6:88-90.

McCullough, D. R. 1979. The George Reserve deer herd: population ecology of a K-selected species. University of Michigan Press, Ann Arbor.

McCullough, D. R. 1999. Density dependence and life-history strategies of ungulates. Journal of Mammalogy 80:1130-1146.

Moen, A. N. 1978. Seasonal changes in heart rates, activity, metabolism, and forage intake of white-tailed deer. Journal of Wildlife Management 42:715-738.

Monteith, K. L., L. E. Schmitz, J. A. Jenks, J. A. Delger, and R. T. Bowyer. 2009. Growth of male white-tailed deer: consequences of maternal effects. Journal of Mammalogy 90:651-660.

Moore, T. D., L. E. Spence, and C. E. Dugnolle. 1997. Identification of the dorsal guard hairs of some mammals of Wyoming. Wyoming Game and Fish Department, Cheyenne, WY, USA.

Neiland, K. A. 1970. Weight of dried marrow as indicator of fat in caribou femurs. Journal of Wildlife Management 34:904-907.

Neter, J., M. H. Kutner, C. J. Nachtsheim, and W. Wasserman. 1996. Applied Linear Statistical Models. Fourth Edition. McGraw-Hill. Boston, Massachusetts.

Page 26: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 26

Norton-Griffiths, M. 1978. Counting animals. 2nd edition. Serengeti Ecological Monitoring Programme, African Wildlife Leadership Foundation, Nairobi, Kenya.

Parker, K. L., P. S. Barboza, and M. P. Gillingham. 2009. Nutrition integrates environmental responses of ungulates. Functional Ecology 23:57-69.

Pederson, J. C., and K. T. Harper. 1978. Factors influencing productivity of two mule deer herds in Utah. Journal of Range Management 31:105-110.

Pederson, J. C., and K. T. Harper. 1984. Does summer range quality influence sex ratios among mule deer fawns in Utah. Journal of Range Management 37:64-66.

Pekins, P. J., K. S. Smith, and W. W. Mautz. 1998. The energy cost of gestation in white-tailed deer. Canadian Journal of Zoology 76:1091-1097.

Piasecke, J. R., and L. C. Bender. 2009. Relationships between condition of adult females and relative carrying capacity for Rocky Mountain Elk. Rangeland Ecology and Management 62:145-152.

Pierce, B. M. 1999. Predator-prey dynamics between mountain lions and mule deer: effects on distribution, population regulation, habitat selection and prey selection. PhD Dissertation, University of Alaska Fairbanks, Fairbanks, Alaska.

Pierce, B. M., R. T. Bowyer, and V. C. Bleich. 2004. Habitat selection by mule deer: forage benefits or risk of predation? Journal of Wildlife Management 68:533-541.

Pierce, B. M., V. C. Bleich, and R. T. Bowyer. 2000. Social organization of mountain lions: does a land-tenure system regulate population size? Ecology 81:1533-1543.

Pierce, B. M., V. C. Bleich, J. D. Wehausen, and R. Terry Bowyer. 1999. Migratory patterns of mountain lions: implications for social regulation and conservation. Journal of Mammalogy 80:986-992.

Pollock, K. H., S. R. Winterstein, C. M. Bunck, and P. D. Curtis. 1989. Survival analysis in telemetry studies: the staggered entry design. Journal of Wildlife Management 53:7-15.

Raithel, J. D., M. J. Kauffman, and D. H. Pletscher. 2007. Impact of spatial and temporal variation in calf survival on the growth of elk populations. Journal of Wildlife Management 71:795-803.

Ratcliffe, P. R. 1980. Bone marrow fat as an indicator of condition in Roe Deer. Acta Theriologica 25:333-340.

Robbinette, W. L., C. H. Baer, R. E. Pillmore, and C. E. Knittle. 1973. Effects of nutritional change on captive mule deer. Journal of Wildlife Management 37:312-326.

Rose, K. E., T. H. Clutton-Brock, and F. E. Guinness. 1998. Cohort variation in male survival and lifetime breeding success in red deer. Journal of Animal Ecology 67:979-986.

Rose, K. E., T. H. Clutton-Brock, and F. E. Guinness. 1998. Cohort variation in male survival and lifetime breeding success in red deer. The Journal of Animal Ecology 67:979-986.

Page 27: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 27

Sadlier, R. M. F. S. 1980. Milk yield of black-tailed deer. Journal of Wildlife Management 44:472-478.

Sadlier, R. M. F. S. 1982. Energy consumption and subsequent partitioning in lactating black-tailed deer. Canadian Journal of Zoology 60:382-386.

Sæther, B. E. 1997. Environmental stochasticity and population dynamics of large herbivores: a search for mechanisms. Trends in Ecology and Evolution 12:143-149.

Sams, M. G., R. L. Lochmiller, C. W. Qualls, Jr., D. M. Leslie, Jr., and M. E. Payton. 1996. Physiological correlates of neonatal mortality in an overpopulated herd of white-tailed deer. Journal of Mammalogy 77:179-190.

Sauer, J. R., and B. K. Williams. 1989. Generalized procedures for testing hypotheses about survival or recovery rates. Journal of Wildlife Management 53:137-142.

Short, H. L. 1981. Nutrition and metabolism. Pages 99-127 in O. C. Wallmo, editor. Mule and black-tailed deer of North America. Univeristy of Nebraska Press, Lincoln, Nebraska, USA.

Short, H. L., D. E. Medin, and A. E. Anderson. 1966. Seasonal variations in volatile fatty acids in the rumen of mule deer. Journal of Wildlife Management 30:466-470.

Simard, M. A., S. D. Côté, R. B. Weladji, and Jean Huot. 2008. Feedback effects of chronic browsing on life-history traits of a large herbivore. Journal of Animal Ecology 77:678-686.

Sinclair, A. R. E., and C. J. Krebs. 2002. Complex numerical responses to top-down and bottom-up processes in vertebrate populations. Philosophical Transactions of the Royal Society of London Series B 357:1221-1231.

Stephenson, T. R., B. C. Bleich, B. M. Pierce, and G. P. Mulcahy. 2002. Validation of mule deer body composition using in vivo and post-mortem indices of nutritional condition. Wildlife Society Bulletin 30:557-564.

Stephenson, T. R., J. W. Testa, G. P. Adams, R. G. Sasser, C. C. Schwartz, and K. J. Hundertmark. 1995. Diagnosis of pregnancy and twinning in moose by ultrasonography and serum assay. Alces 31:167-172.

Stewart, K. M., R. T. Bowyer, B. L. Dick, B. K. Johnson, and J. G. Kie. 2005. Density-dependent effects on physical condition and reproduction in North American elk: an experimental test. Oecologia 143:85-93.

Storer, T. I., and L. P. Tevis, Jr. 1955. The California grizzly. University of California Press, Berkeley, California, USA.

Storer, T. I., R. L. Usinger, and D. Lukas. 2004. Sierra Nevada Natural History. University of California Press, Berkeley, California, USA.

Swift, P. K., V. C. Bleich, T. R. Stephenson, A. E. Adams, B. J. Gonzales, B. M. Pierce, and J. P. Marshal. 2002. Tooth extraction from live-captured mule deer in the absence of chemical immobilization. Wildlife Society Bulletin 30:253-255.

Swihart, R. K., H. P. Weeks, Jr., A. L. Easter-Pilcher, and A. J. DeNicola. 1998. Nutritional condition and fertility of white-tailed deer (Odocoileus

Page 28: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 28

virginianus) from areas with contrasting histories of hunting. Canadian Journal of Zoology 76:1932-1941.

Taillon, J., D. G. Sauvé, and S. D. Côté. 2006. The effects of decreasing winter diet quality on foraging behavior and life-history traits of white-tailed deer fawns. Journal of Wildlife Management 70:1445-1454.

Testa, J. W., and G. P. Adams. 1998. Body condition and adjustments to reproductive effort in female moose (Alces alces). Journal of Mammalogy 79:1345-1354.

Toïgo, C., J.-M. Gaillard, and J. Michallet. 1997. Adult survival of the sexually dimorphic Alpine ibex (Capra ibex ibex). Canadian Journal of Zoology 75:75-79.

Torbit, S. C., L. H. Carpenter, D. M. Swift, and A. W. Alldredge. 1985. Differential loss of fat and protein by mule deer during winter. Journal of Wildlife Management 49:80-85.

Unsworth, J. W., D. F. Pac, G. C. White, and R. M. Bartmann. 1999. Mule deer survival in Colorado, Idaho, and Montana. Journal of Wildlife Management 63:315-326.

Webb, S. L., D. G. Hewitt, and M. W. Hellickson. 2007. Survival and cause-specific mortality of mature male white-tailed deer. Journal of Wildlife Management 71:555-558.

Zar, J. H. 1999. Biostatistical Analysis. 4th Edition. Prentice Hall. Upper Saddle River NewJersey.

Page 29: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 29

Table 1. Number of neonatal mule deer captured and monitored on summer range each year in the Sierra Nevada.

Year East West Censored Total

2006 32 10 1 41

2007 20 14 4 30

2008 29 14 0 43 Table 2. Survival rate (5 months) and percent of cause-specific mortality of neonatal mule deer captured and monitored on summer range each year and for 2006-2007 combined with respect to their natal range occurring on either side of the Sierra crest. Year & side crest

% Survival

% Abandon-

ment %

Accident %

Malnutrition% Other natural

% Predation

% Starvation

% Stillborn

% Unknown

2006 47.8 0.0 2.4 7.3 7.3 29.3 0.0 0.0 7.3 2007 30.7 3.0 6.1 0.0 0.0 45.5 0.0 6.1 9.1 2008 41.8 0.0 4.7 7.0 2.3 34.9 4.7 0.0 4.7 East 49.7 1.3 5.1 3.8 3.8 24.4 2.6 1.3 9.0 West 21.3 0.0 2.7 8.1 2.7 62.2 0.0 2.7 2.7

Table 3. Percent of total mortality caused by predation and the percent of that predation contributed to each predator of neonatal mule deer captured and monitored on summer range during 2006-2007 in the Sierra Nevada with respect to their natal range occurring on either side of the Sierra crest. Percents are total percent predation and of that, the predators responsible. Side of crest

% Predation

% Bear

% Bobcat

% Coyote

% Eagle

% Lion

East 48 16 10 53 5 16 West 77 87 9 4 0 0

Page 30: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 30

Table 4. Citations from peer-reviewed literature published involving research on mule deer in Round Valley, Bishop, California. Bleich, V. C., and B. M. Pierce. 1999. Expandable and economical long-term

collars for juvenile mule deer. California Fish and Game 85:56-62. Bleich, V. C., and B. M. Pierce. 2001. Accidental mass mortality of migrating

mule deer. Western North American Naturalist 61:124-125. Bleich, V. C., and T. J. Taylor. 1998. Survivorship and cause-specific mortality in

five populations of mule deer. Great Basin Naturalist 58:265-272. Bleich, V. C., B. M. Pierce, J. L. Jones, and R. T. Bowyer. 2006. Variance in

survival of young mule deer in the Sierra Nevada, California. California Fish and Game 92:24-38.

Bleich, V. C., J. T. Villepique, T. R. Stephenson, B. M. Pierce, and G. M. Kutiyev.

2005. From the field: efficacy of aerial telemetry as an aid to capturing specific individuals – a comparison of 2 techniques. Wildlife Society Bulletin 33:332-336.

Bleich, V. C., T. R. Stephenson, B. M. Pierce, and M. J. Warner. 2007. Body

condition of radio-collared mule deer while injured and following recovery. Southwestern Naturalist 52:164-167.

Bleich, V. C., T. R. Stephenson, N. J. Holste, I. C. Snyder, J. P. Marshal, P. W.

McGrath, and B. M. Pierce. 2003. Effects of tooth extraction on body condition and reproduction of mule deer. Wildlife Society Bulletin 31:233-236.

Bowyer, R. T., D. K. Person, and B. M. Pierce. 2005. Detecting top-down versus

bottom-up regulation of ungulates by large carnivores: implications for biodiversity. Pages 342-361 in J. C. Ray, K. H. Redford, R. S. Steneck, and J. Berger, editors. Large carnivores and the conservation of biodiversity. Island Press, Washington DC, USA.

Cook, R. C., T. R. Stephenson, W. L. Myers, J. G. Cook, and L. A. Shipley. 2006.

Validating predictive models of nutritional condition for mule deer. Journal of Wildlife Management 71:1934-1943.

Kucera, T. E. 1988. Ecology and population dynamics of mule deer in the

Eastern Sierra Nevada, California. Ph.D. Dissertation. University of California, Berkely, California, USA.

Page 31: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 31

Kucera, T. E. 1991. Adaptive variation in sex ratios of offspring in nutritionally stressed mule deer. Journal of Mammalogy 72:745-749.

Kucera, T. E. 1992. Influences of sex and weather on migration of mule deer in

California. Great Basin Naturalist 52:122-130. Kucera, T. E. 1997. Fecal indicators, diet, and population parameters in mule

deer. Journal of Wildlife Management 61:550-560. Pease, K. M., A. H. Freedman, J. P. Pollinger, J. E. McCormack, W. Buermann,

J. Rodzen, J. Banks, E. Meredith, V. C. Bleich, R. J. Schaefer, K. Jones, and R. K.. Wayne. 2009. Landscape genetics of California mule deer (Odocoileus hemionus): the roles of ecological and historical factors in generating differentiation. Molecular Ecology 18:1848-1862.

Pierce, B. M. 1999. Predator-prey dynamics between mountain lions and mule

deer: effects on distribution, population regulation, habitat selection and prey selection. PhD Dissertation, University of Alaska Fairbanks, Fairbanks, Alaska.

Pierce, B. M., R. T. Bowyer, and V. C. Bleich. 2004. Habitat selection by mule

deer: forage benefits or risk of predation? Journal of Wildlife Management 68:533-541.

Pierce, B. M., V. C. Bleich, and R. T. Bowyer. 2000. Social organization of

mountain lions: does a land-tenure system regulate population size? Ecology 81:1533-1543.

Pierce, B. M., V. C. Bleich, and R. T. Bowyer. 2000a. Selection of mule deer by

mountain lions and coyotes, effects of hunting style, body size, and reproductive status. Journal of Mammalogy 81:462-472.

Pierce, B. M., V. C. Bleich, C. L. B. Chetkiewicz, and J. D. Wehausen. 1998.

Timing of feeding bouts of mountain lions. Journal of Mammalogy 79:222-226.

Pierce, B. M., V. C. Bleich, J. D. Wehausen, and R. Terry Bowyer. 1999.

Migratory patterns of mountain lions: implications for social regulation and conservation. Journal of Mammalogy 80:986-992.

Stephenson, T. R., B. C. Bleich, B. M. Pierce, and G. P. Mulcahy. 2002.

Validation of mule deer body composition using in vivo and post-mortem indices of nutritional condition. Wildlife Society Bulletin 30:557-564.

Swift, P. K., V. C. Bleich, T. R. Stephenson, A. E. Adams, B. J. Gonzales, B. M.

Pierce, and J. P. Marshal. 2002. Tooth extraction from live-captured mule

Page 32: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 32

deer in the absence of chemical immobilization. Wildlife Society Bulletin 30:253-255.

Villepique, J. T., V. C. Bleich, B. M. Pierce, T. R. Stephenson, R. A. Botta, and R.

T. Bowyer. 2009. Evaluating GPS collar errors: a critical evaluation of Televilt Posrec-Science ™ collars and a method for screening location data. California Fish and Game 94:155-168.

Page 33: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 33

Table 5. Topics of forthcoming manuscripts from research conducted on mule deer in Round Valley, Bishop, California.

1) Population dynamics and reproductive ecology of mule deer in the Sierra

Nevada. 2) Timing of migration of mule deer in the Sierra Nevada: effects of a changing

climate 3) Estimating pregnancy, fetal rate, and timing of parturition in mule deer 4) Estimating time and location of parturition using patterns of movement in mule

deer 5) Capture and handling techniques of ungulates: animal care and censor

interval 6) Resource use and reproductive status: implications for estimating habitat use

and selection of ungulates 7) Senescence in reproduction and survival in mule deer 8) Dynamics of an eastern Sierra Nevada mule deer population: top-down or

bottom-up regulation?

Page 34: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 34

Figure 1. Map of study area displaying some of the primary migration routes utilized by adult female mule deer fitted with GPS collars.

Page 35: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 35

Figure 2. Map of the study area (Madera, Fresno, Mono, and Inyo counties, CA) with west-side summer ranges occurring in the San Joaquin River drainage and east-side summer ranges occurring along the eastern side of the Sierra crest.

Page 36: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 36

Year1982

19841986

19881990

19921994

19961998

20002002

20042006

20082010

2012

Tota

l Ani

mal

s

0

1000

2000

3000

4000

5000

6000

7000

Figure 3. Annual population estimates (±95% CI)of mule deer as determined from total counts prior to 1994 and from mark-resight surveys thereafter, Round Valley, Bishop, California.

Year

19821984

19861988

19901992

19941996

19982000

20022004

20062008

20102012

Tota

l Ani

mal

s

0

1000

2000

3000

4000

5000

6000

7000

Mar

ch in

gest

a-fre

e bo

dy fa

t (%

)

0

2

4

6

8

10

12Population sizeBody fat

Figure 4. Annual population estimates (±95% CI)of mule deer and March ingesta-free body fat (±SE) of adult female mule deer wintering in Round Valley, Bishop, California.

Page 37: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 37

Year

1997

1998

1999

2000

2001

2002

2003

2004

2005

2006

2007

2008

2009

Adu

lt fe

mal

e an

nual

sur

viva

l (%

)

70

75

80

85

90

95

100

Figure 5. Annual survival (±SE; Jan-Dec) of adult (>1 yr old) female mule deer that overwinter in Round Valley, Bishop, California and summer in the Sierra Nevada, 1998-2008.

Winter Year

1998

/1999

1999

/2000

2000

/2001

2001

/2002

2002

/2003

2003

/2004

2004

/2005

2005

/2006

2006

/2007

2007

/2008

Adu

lt fe

mal

e w

inte

r sur

viva

l (%

)

70

75

80

85

90

95

100

Figure 6. Winter survival (±SE; 16 Oct – 15 April) of adult (>1 yr old) female mule deer in Round Valley, Bishop, California, 1998-2008.

Page 38: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 38

Year

1997

1998

1999

2000

2001

2002

2003

2004

2005

2006

2007

2008

2009

Adu

lt fe

mal

e su

mm

er s

urvi

val (

%)

70

75

80

85

90

95

100

Figure 7. Summer survival (±SE; 16 April – 15 Oct) of adult (>1 yr old) female mule deer in the Sierra Nevada, 1998-2008.

Year

1998

1999

2000

2001

2002

2003

2004

2005

2006

2007

2008

Ann

ual s

urvi

val (

%)

60

70

80

90

100East-side femalesWest-side females

Figure 8. Annual survival (±SE) of adult (>1 yr old) female mule deer with respect to summer residency status in the Sierra Nevada, 1998-2008.

Page 39: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 39

Year

19961998

20002002

20042006

20082010

Rec

ruitm

ent r

ate

0.0

0.2

0.4

0.6

0.8

1.0

Figure 9. Annual recruitment rate (±SE) in autumn of adult (>1 yr old) female mule deer that were radiocollared, Round Valley, Bishop, California, 1997-2008.

Year

1996 1998 2000 2002 2004 2006 2008 2010

Rec

ruitm

ent r

ate

0.0

0.2

0.4

0.6

0.8

1.0

1.2 East-side femalesWest-side females

Figure 10. Annual recruitment rate (±SE) of adult (>1 yr old) female mule deer with respect to summer residency (side of Sierra crest) that were radiocollared, Round Valley, Bishop, California, 1997-2008.

Page 40: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 40

Year

2000 2002 2004 2006 2008 2010

Nov

embe

r ing

esta

-free

bod

y fa

t (%

)

5

6

7

8

9

10

11

12

Figure 11. Average ingesta-free body fat (±SE) in November of adult (>1yr old) female mule deer per year, Round Valley, Bishop, California.

Year

2000 2002 2004 2006 2008 2010

Nov

embe

r ing

esta

-free

bod

y fa

t (%

)

2

4

6

8

10

12

14

16 East-side femalesWest-side females

Figure 12. Average ingesta-free body fat (±SE) in November of adult (>1yr old) female mule deer relative to summer residency status (i.e., side of Sierra crest), Round Valley, Bishop, California.

Page 41: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 41

Number recruited young

Year

2001 2002 2003 2004 2005 2006 2007 2008 2009

Nov

embe

r ing

esta

-free

bod

y fa

t (%

)

0

2

4

6

8

10

12

14

160123

Figure 13. Ingesta-free body fat (%±SE) in November of adult female (>1yr old) mule deer relative to the number of young recruited per year, Round Valley, Bishop, California.

Number young recruited

0 1 2 3

Nov

embe

r ing

esta

-free

bod

y fa

t (%

)

0

2

4

6

8

10

12

14East-side femalesWest-side females

Figure 14. Ingesta-free body fat (%±SE) in November of adult female (>1yr old) mule deer relative to the number of young recruited and summer residency (i.e., side of Sierra crest), Round Valley, Bishop, California.

Page 42: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 42

Number of young recruited

0 1 2 3

Mar

ch in

gest

a-fre

e bo

dy fa

t (%

)

0

1

2

3

4

5

6East-side females West-side females

Figure 15. Average ingesta-free body fat (±SE) of adult (>1yr old) female mule deer relative to recruitment status the previous autumn with respect to summer residency status (i.e., side of Sierra crest), Round Valley, Bishop, California.

Year

2002

2003

2004

2005

2006

2007

2008

2009

Inge

sta

free-

body

fat (

%)

0

2

4

6

8

10

12 March November

Figure 16. Average ingesta-free body fat (±SE) of adult (>1yr old) female mule deer during March and November 2002-2009, Round Valley, Bishop, California.

Page 43: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 43

Year

1996 1998 2000 2002 2004 2006 2008 2010

Feta

l rat

e

1.0

1.2

1.4

1.6

1.8

2.0

2.2East-side femalesWest-side females

Figure 17. Average number of fetuses per adult (>1yr old) female mule deer during 1997-2009 with respect to summer residency status (i.e., side of Sierra crest), Round Valley, Bishop, California.

Number of fetuses

0 1 2 3

Nov

embe

r ing

esta

-free

bod

y fa

t (%

)

0

2

4

6

8

10

12

14

16 (n=4)

(n=5)

(n=97)(n=185)

Figure 18. Average ingesta-free body fat (±SE) in November of adult (>1yr old) female mule deer relative to number of fetuses present the following March, Round Valley, Bishop, California. Sample size per group is shown parenthetically.

Page 44: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 44

Age of neonate (days)

0-30 31-60 61-90 91-120 121-150

Neo

nata

l sur

viva

l (%

)

10

20

30

40

50

60

70

80

90

2006 2007 2008

Figure 19. Monthly survival (±SE) of neonatal mule deer from birth to 150 days-of-age in the central Sierra Nevada, 2006-2008.

Age of neonate (days)

0-30 31-60 61-90 91-120 121-150

Neo

nata

l sur

viva

l (%

)

20

30

40

50

60

70

80

FemaleMale

Figure 20. Monthly survival (±SE) of neonatal mule deer by sex from birth to 150 days-of-age in the central Sierra Nevada, 2006-2008.

Page 45: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 45

Age of neonate (days)

0-30 31-60 61-90 91-120 121-150

Neo

nata

l sur

viva

l (%

)

10

20

30

40

50

60

70

80

East of Sierra crestWest of Sierra crest

Figure 21. Monthly survival (±SE) of neonatal mule deer from birth to 150 days-of-age in the central Sierra Nevada with respect to which side of the Sierra crest each neonate was born, 2006-2008.

Year

1950 1960 1970 1980 1990 2000 2010

You

ng:a

dult

fem

ales

ratio

0

20

40

60

80

100

Figure 22. Number of young per 100 females (±95% CI) observed during composition surveys conducted each January in Round Valley, Bishop, California during 1950-2008. Data are reported as the post-season composition of the previous year.

Page 46: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 46

Year

1950 1960 1970 1980 1990 2000 2010

Adu

lt m

ales

:adu

lt fe

mal

es ra

tio

0

10

20

30

40

50

60

Figure 23. Number of adult males (>1 yr old) per 100 adult females (>1 yr old; ±95% CI) observed during composition surveys conducted each January in Round Valley, Bishop, California during 1950-2008. Data are reported as the post-season composition of the previous year.

Water content of April snowpack (cm)t

0 10 20 30 40 50 60 70

Bitte

rbru

sh le

ader

gro

wth

(cm

) t

0

5

10

15

20

25r2 = 0.65, P = 0.001y = 0.237x + 1.19

Figure 24. Length of annual growth of bitterbrush leaders relative to water content of snowpack measured in April, 1984-2008, Round Valley, Bishop, California.

Page 47: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 47

Water content of snow pack (cm)t

0 10 20 30 40 50 60 70

Mar

ch in

gest

a-fre

e bo

dy fa

t (%

) t+1

0

2

4

6

8

10

12

14(r2 = 0.53, P < 0.001)y = 0.12x + 2.55

Figure 25. Influence of water content of snow pack on percent of ingesta-free body fat of adult female mule deer the following March in Round Valley, Bishop, California, 1985-2009.

Total animalst

1000 1500 2000 2500 3000 3500

Mar

ch in

gest

a-fre

e bo

dy fa

t (%

) t+1

2

4

6

8

10

12r2 = 0.38, P = 0.009y = -0.002x + 11.1

(1993)

Figure 26. Relationship between population size of mule deer and average percent of ingesta-free body fat of adult females wintering in Round Valley, Bishop, California, following the population crash 1992-2009.

Page 48: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 48

March ingesta-free body fat (%)t

0 2 4 6 8 10 12

Adu

lt m

ales

:adu

lt fe

mal

es ra

tiot+

1

0

10

20

30

40

50r2 = 0.65, P = 0.001y = 2.85x + 14.5

(1992)

(1994)

Figure 27. Relationship between ingesta-free body fat of adult female mule deer in March and the number of adult males:adult females the following January, Round Valley, Bishop, California, 1991-2008. Linear regression results are reported with outliers removed (i.e., 1992, 1994).

March ingesta-free body fat (%)t

0 2 4 6 8 10 12

You

ng:a

dult

fem

ales

ratio

t+1

0

10

20

30

40

50

60

70r2 = 0.45, P = 0.002y = 3.2x + 23.2

Figure 28. Relationship between ingesta-free body fat of adult female mule deer in March and the number of young:100 adult females the following January, Round Valley, Bishop, California, 1991-2008.

Page 49: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 49

March ingesta-free body fat (%)t

0 2 4 6 8 10

Adu

lt fe

mal

e su

rviv

al (%

) t

75

80

85

90

95

100r2 = 0.50, P = 0.014y = 4.03x + 65.9

Figure 29. Relationship between ingesta-free body fat of adult female mule deer in March and annual survival of adult females, Round Valley, Bishop, California, 1991-2008.

Annual survival of adult females (%)

75 80 85 90 95 100

Lam

bda

0.7

0.8

0.9

1.0

1.1

1.2

1.3r2 = 0.09, P = 0.38

Figure 30. Relationship between annual survival of adult female mule deer (%) and lambda, Round Valley, Bishop, California, 1991-2008.

Page 50: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 50

Young / 100 adult females

0 10 20 30 40 50 60 70

Lam

bda

0.4

0.6

0.8

1.0

1.2

1.4

r2 = 0.45, P < 0.001y = 0.012x + 0.49

(1991)

Figure 31. Relationship between number of young:100 females and lambda during the current year, Round Valley, Bishop, California, 1985-2009. Linear regression results are reported with outlier (1991) removed.

March ingesta-free body fat (%)

0 2 4 6 8 10 12

Lam

bda

0.7

0.8

0.9

1.0

1.1

1.2

1.3 r2 = 0.44, P = 0.004y = 0.051x + 0.72

Figure 32. Relationship between ingesta-free body fat of adult female mule deer in March and lambda during the current year, Round Valley, Bishop, California, 1991-2008.

Page 51: Kevin L. Monteith, Vernon C. Bleich, Thomas R. Stephenson ...mcbadeer.com/DFG_ROUND_VALLEY_STUDY.pdf · condition is the integration of both nutrient intake and energetic costs

Monteith et al. • Mule Deer Population Dynamics 51

Figure 33. A conceptual model depicting relationships between population density, winter severity, and rate of mortality. Representative curves are provided for (a) density independent, (b) large, and (c) moderate density-dependent effects interacting with winter severity. The lines around the inset graphs show the area on the growth curve to which each inset corresponds. Note that the shape of the population-growth curve need not be symmetrical for the proposed relationships to hold (adapted from Bowyer et al. 2000).