@article{weckworth_davis_dubovi_fountain-jones_packer_cleaveland_craft_eblate_schwartz_mills_et al._2020, title={Cross-species transmission and evolutionary dynamics of canine distemper virus during a spillover in African lions of Serengeti National Park}, volume={29}, ISSN={["1365-294X"]}, DOI={10.1111/mec.15449}, abstractNote={The outcome of pathogen spillover from a reservoir to a novel host population can range from a “dead‐end” when there is no onward transmission in the recipient population, to epidemic spread and even establishment in new hosts. Understanding the evolutionary epidemiology of spillover events leading to discrete outcomes in novel hosts is key to predicting risk and can lead to a better understanding of the mechanisms of emergence. Here we use a Bayesian phylodynamic approach to examine cross‐species transmission and evolutionary dynamics during a canine distemper virus (CDV) spillover event causing clinical disease and population decline in an African lion population (Panthera leo) in the Serengeti Ecological Region between 1993 and 1994. Using 21 near‐complete viral genomes from four species we found that this large‐scale outbreak was likely ignited by a single cross‐species spillover event from a canid reservoir to noncanid hosts <1 year before disease detection and explosive spread of CDV in lions. Cross‐species transmission from other noncanid species probably fuelled the high prevalence of CDV across spatially structured lion prides. Multiple lines of evidence suggest that spotted hyenas (Crocuta crocuta) could have acted as the proximate source of CDV exposure in lions. We report 13 nucleotide substitutions segregating CDV strains found in canids and noncanids. Our results are consistent with the hypothesis that virus evolution played a role in CDV emergence in noncanid hosts following spillover during the outbreak, suggest that host barriers to clinical infection can limit outcomes of CDV spillover in novel host species.}, number={22}, journal={MOLECULAR ECOLOGY}, author={Weckworth, Julie K. and Davis, Brian W. and Dubovi, Edward and Fountain-Jones, Nicholas and Packer, Craig and Cleaveland, Sarah and Craft, Meggan E. and Eblate, Ernest and Schwartz, Michael and Mills, L. Scott and et al.}, year={2020}, month={Nov}, pages={4308–4321} } @article{weckworth_davis_roelke-parker_wilkes_packer_eblate_schwartz_mills_2020, title={Identifying Candidate Genetic Markers of CDV Cross-Species Pathogenicity in African Lions}, volume={9}, ISSN={["2076-0817"]}, DOI={10.3390/pathogens9110872}, abstractNote={Canine distemper virus (CDV) is a multi-host pathogen with variable clinical outcomes of infection across and within species. We used whole-genome sequencing (WGS) to search for viral markers correlated with clinical distemper in African lions. To identify candidate markers, we first documented single-nucleotide polymorphisms (SNPs) differentiating CDV strains associated with different clinical outcomes in lions in East Africa. We then conducted evolutionary analyses on WGS from all global CDV lineages to identify loci subject to selection. SNPs that both differentiated East African strains and were under selection were mapped to a phylogenetic tree representing global CDV diversity to assess if candidate markers correlated with documented outbreaks of clinical distemper in lions (n = 3). Of 54 SNPs differentiating East African strains, ten were under positive or episodic diversifying selection and 20 occurred in the clinical strain despite strong purifying selection at those loci. Candidate markers were in functional domains of the RNP complex (n = 19), the matrix protein (n = 4), on CDV glycoproteins (n = 5), and on the V protein (n = 1). We found mutations at two loci in common between sequences from three CDV outbreaks of clinical distemper in African lions; one in the signaling lymphocytic activation molecule receptor (SLAM)-binding region of the hemagglutinin protein and another in the catalytic center of phosphodiester bond formation on the large polymerase protein. These results suggest convergent evolution at these sites may have a functional role in clinical distemper outbreaks in African lions and uncover potential novel barriers to pathogenicity in this species.}, number={11}, journal={PATHOGENS}, author={Weckworth, Julie K. and Davis, Brian W. and Roelke-Parker, Melody E. and Wilkes, Rebecca P. and Packer, Craig and Eblate, Ernest and Schwartz, Michael K. and Mills, L. Scott}, year={2020}, month={Nov} } @article{kumar_zimova_sparks_mills_2020, title={Snow-mediated plasticity does not prevent camouflage mismatch}, volume={194}, ISSN={["1432-1939"]}, DOI={10.1007/s00442-020-04680-2}, abstractNote={Global reduction in snow cover duration is one of the most consistent and widespread climate change outcomes. Declining snow duration has severe negative consequences for diverse taxa including seasonally color molting species, which rely on snow for camouflage. However, phenotypic plasticity may facilitate adaptation to reduced snow duration. Plastic responses could occur in the color molt phenology or through behavior that minimizes coat color mismatch or its consequences. We quantified molt phenology of 200 wild snowshoe hares (Lepus americanus), and measured microhabitat choice and local snow cover. Similar to other studies, we found that hares did not show behavioral plasticity to minimize coat color mismatch via background matching; instead they preferred colder, snow free areas regardless of their coat color. Furthermore, hares did not behaviorally mitigate the negative consequences of mismatch by choosing resting sites with denser vegetation cover when mismatched. Importantly, we demonstrated plasticity in the initiation and the rate of the molt and established the direct effect of snow on molt phenology; greater snow cover was associated with whiter hares and this association was not due to whiter hares preferring snowier areas. However, despite the observed snow-mediated plasticity in molt phenology, camouflage mismatch with white hares on brown snowless ground persisted and was more frequent during early snowmelt. Thus, we find no evidence that phenotypic plasticity in snowshoe hares is sufficient to facilitate adaptive rescue to camouflage mismatch under climate change.}, number={3}, journal={OECOLOGIA}, author={Kumar, Alexander V and Zimova, Marketa and Sparks, James R. and Mills, L. Scott}, year={2020}, month={Nov}, pages={301–310} } @article{davis_kumar_mills_2019, title={A camouflage conundrum: unexpected differences in winter coat color between sympatric species}, volume={10}, ISSN={["2150-8925"]}, DOI={10.1002/ecs2.2658}, abstractNote={Across the globe, more than 21 species undergo seasonal changes in coloration, molting white in winter to become camouflaged against snow. Given the adaptive value of seasonal camouflage against local snow duration, one might predict that sympatric coat color changing species would have similar winter coat color. This hypothesis, however, contrasts with anecdotal evidence and modeling results that predict sympatric winter white and winter brown species in some areas with transient snow cover. In one such area, West Virginia, we document coat color phenology between three sympatric species: snowshoe hares (Lepus americanus), long-tailed weasels (Mustela frenata), and least weasels (Mustela nivalis). Using a combination of field methods, we document and quantify each species’ winter coat color, illustrating an interspecific polymorphic response in winter coloration among sympatric winter white snowshoe hares and winter brown weasels. We then hypothesize what forces drive the interspecific differences between snowshoe hare and weasel winter coloration, highlighting areas of focus for future seasonal coat color}, number={3}, journal={ECOSPHERE}, author={Davis, Brandon M. and Kumar, Alexander V and Mills, L. Scott}, year={2019}, month={Mar} } @article{bragina_kays_hody_moorman_deperno_mills_2019, title={Effects on white‐tailed deer following eastern coyote colonization}, volume={83}, ISSN={0022-541X 1937-2817}, url={http://dx.doi.org/10.1002/JWMG.21651}, DOI={10.1002/jwmg.21651}, abstractNote={The expansion or recovery of predators can affect local prey populations. Since the 1940s, coyotes (Canis latrans) have expanded into eastern North America where they are now the largest predator and prey on white-tailed deer (Odocoileus virginianus). However, their effect on deer populations remains controversial. We tested the hypothesis that coyotes, as a novel predator, would affect deer population dynamics across large spatial scales, and the strongest effects would occur after a time lag following initial coyote colonization that allows for the predator populations to grow. We evaluated deer population trends from 1981 to 2014 in 384 counties of 6 eastern states in the United States with linear mixed models. We included deer harvest data as a proxy for deer relative abundance, years since coyote arrival in a county as a proxy of coyote abundance, and landscape and climate covariates to account for environmental effects. Overall, deer populations in all states experienced positive population growth following coyote arrival. Time since coyote arrival was not a significant predictor in any deer population models and our results indicate that coyotes are not controlling deer populations at a large spatial scale in eastern North America. 2019 The Wildlife Society.}, number={4}, journal={The Journal of Wildlife Management}, publisher={Wiley}, author={Bragina, Eugenia V. and Kays, Roland and Hody, Allison and Moorman, Christopher E. and Deperno, Christopher S. and Mills, L. Scott}, year={2019}, month={Mar}, pages={916–924} } @article{lafferty_zimova_clontz_hacklander_mills_2019, title={Noninvasive measures of physiological stress are confounded by exposure}, volume={9}, ISBN={2045-2322}, DOI={10.1038/s41598-019-55715-5}, abstractNote={Abstract Glucocorticoids and glucocorticoid metabolites are increasingly used to index physiological stress in wildlife. Although feces is often abundant and can be collected noninvasively, exposure to biotic and abiotic elements may influence fecal glucocorticoid metabolite (FGM) concentrations, leading to inaccurate conclusions regarding wildlife physiological stress. Using captive snowshoe hares ( Lepus americanus ) and simulated environmental conditions, we evaluated how different realistic field conditions and temporal sampling constraints might influence FGM concentrations using an 11-oxoetiocholanolone-enzyme immunoassay. We quantified how fecal pellet age (i.e., 0–6 days), variable summer temperatures, and precipitation affected FGM concentrations. Fecal pellet age had a strong effect on FGM concentrations (β Age = 0.395, s.d. = 0.085; β 2 Age = −0.061, s.d. = 0.012), which were lowest at the beginning and end of our exposure period (e.g., mean day6 = 37.7 ng/mg) and typically highest in the middle (mean day3 = 51.8 ng/mg). The effect of fecal pellet age on FGM concentrations varied across treatments with warm-dry and cool-wet conditions resulting in more variable FGM concentrations relative to control samples. Given the confounding effects of exposure and environmental conditions, if fresh fecal pellet collection is not an option, we encourage researchers to develop a temporally consistent sampling protocol to ensure all samples are exposed to similar environmental conditions.}, journal={SCIENTIFIC REPORTS}, author={Lafferty, Diana J. R. and Zimova, Marketa and Clontz, Lindsay and Hacklander, Klaus and Mills, L. Scott}, year={2019} } @article{bragina_kays_hody_moorman_deperno_mills_2019, title={White‐tailed deer and coyote colonization: a response to Kilgo et al. (2019)}, volume={83}, ISSN={0022-541X 1937-2817}, url={http://dx.doi.org/10.1002/jwmg.21766}, DOI={10.1002/jwmg.21766}, abstractNote={We appreciate the comments by Kilgo et al. (2019) on our recent paper and welcome the opportunity to expand the discussion of the effects of coyotes (Canis latrans) on white‐ tailed deer (Odocoileus virginianus) in eastern North America. This exchange allows us to reaffirm and elaborate on our finding that coyotes have not had large‐scale effects on white‐tailed deer population growth in eastern North America. A scientific evaluation of how any predator affects prey numbers is typically a complex challenge that requires a variety of approaches at a variety of scales. The evaluation becomes especially challenging for populations that interact across a complex and variable landscape because the dynamics of individual populations affect each other through immigration and emigration. Thus, for vagile wildlife species with a broad distribution, studies of single populations cannot fully answer how the predator affects the prey species across the larger connected landscape. Of course, intensive and detailed site‐ and time‐specific studies on local predator‐prey dynamics are critically important, and necessary to a mechanistic understanding of processes of predation. By themselves, however, a collection of such local studies over short time periods cannot determine how predator‐prey dynamics play out over large spatial and temporal scales. Coyotes and white‐tailed deer in the eastern United States are an example of a spatially complex predator‐prey system. Following the extinction of wolves (Canis lupus) and cougars (Puma concolor) in the region, coyotes colonized all eastern states over the last half‐century (Hody and Kays 2018). Coyotes now occupy all land cover types in the region, where they overlap, and prey on, white‐tailed deer. In Bragina et al. (2019), we noted site‐specific field studies conducted (Vreeland et al. 2004, Kilgo et al. 2012, Chitwood et al. 2015) on coyotes and deer in various parts of this system and agree that several of these studies show local effects of coyotes on white‐tailed deer vital rates or numbers (Vreeland et al. 2004, Howze et al. 2009, Chitwood et al. 2015); however, the results of these studies have not always been consistent, demonstrating the predator‐prey dynamic is variable over space or time. For example, Kilgo et al. (2016:747) concluded that “predation by coyotes on adult females was not important,” Kilgo et al. (2014:1261) stated that “despite an initial increase, the overall effect of coyote removal on neonate [deer] survival was modest,” and Gulsby et al. (2017) reported that coyote predation on fawns depended on habitat fragmentation. In short, the full collection of field studies reveal coyote‐deer interactions in certain places at certain times, but they were not designed to answer the question addressed in Bragina et al. (2019): How have coyotes affected deer population growth from 1981 to 2014 across multiple states in the eastern United States? In answering this question, nothing in Bragina et al. (2019) dismissed the valuable time‐ and place‐specific field studies of coyotes and white‐tailed deer to which Kilgo et al. (2019) referred. It is apparent the concerns of Kilgo et al. (2019) arise from an incomplete appreciation of the different strengths, scope of inference, and types of questions that can be addressed with local and time‐specific field studies, versus broader space and time studies such as ours. Faced with the overarching question of whether coyotes have affected white‐tailed deer population dynamics over 531,312 km and more than 3 decades, we would ideally use rigorously collected population estimates of both species from across the entire region. These estimates exist in vanishingly few cases. Therefore, with full awareness and appreciation of the promise and pitfalls of indices relative to rigorous estimates of abundance (Pollock et al. 2002, Mills et al. 2005, Johnson 2008), we used the only data available over time and space from 6 state agencies: indices of trends in white‐tailed Received: 27 August 2019; Accepted: 4 September 2019}, number={8}, journal={The Journal of Wildlife Management}, publisher={Wiley}, author={Bragina, Eugenia V. and Kays, Roland and Hody, Allison and Moorman, Christopher E. and DePerno, Christopher S. and Mills, L. Scott}, year={2019}, month={Oct}, pages={1641–1643} } @article{jones_mills_alves_callahan_alves_lafferty_jiggins_jensen_melo-ferreira_good_2018, title={Adaptive introgression underlies polymorphic seasonal camouflage in snowshoe hares}, volume={360}, ISSN={["1095-9203"]}, DOI={10.1126/science.aar5273}, abstractNote={Hybrid camouflage variation Snowshoe hares molt from a brown coat to a white coat in winter. In some populations, however, where winter snow is less extensive, hares molt from a brown coat to a brown coat. Jones et al. show that regulation of the pigmentation gene Agouti is responsible for the winter coat color change. Hybridization with jackrabbits has led to introgression around this gene that facilitates the brown winter morph. Hybridization appears to have provided important adaptive variation to the snowshoe hare. Science, this issue p. 1355 Exchange of genetic variants through hybridization can seed past and ongoing adaptation to rapidly changing environments. Snowshoe hares (Lepus americanus) maintain seasonal camouflage by molting to a white winter coat, but some hares remain brown during the winter in regions with low snow cover. We show that cis-regulatory variation controlling seasonal expression of the Agouti gene underlies this adaptive winter camouflage polymorphism. Genetic variation at Agouti clustered by winter coat color across multiple hare and jackrabbit species, revealing a history of recurrent interspecific gene flow. Brown winter coats in snowshoe hares likely originated from an introgressed black-tailed jackrabbit allele that has swept to high frequency in mild winter environments. These discoveries show that introgression of genetic variants that underlie key ecological traits can seed past and ongoing adaptation to rapidly changing environments.}, number={6395}, journal={SCIENCE}, author={Jones, Matthew R. and Mills, L. Scott and Alves, Paulo Celio and Callahan, Colin M. and Alves, Joel M. and Lafferty, Diana J. R. and Jiggins, Francis M. and Jensen, Jeffrey D. and Melo-Ferreira, Jose and Good, Jeffrey M.}, year={2018}, month={Jun}, pages={1355-+} } @article{kumar_sparks_mills_2018, title={Short-term response of snowshoe hares to western larch restoration and seasonal needle drop}, volume={26}, ISSN={["1526-100X"]}, DOI={10.1111/rec.12533}, abstractNote={Old‐growth western larch has been degraded throughout much of its historic range due to extensive timber harvest and fire suppression. We examined the effects of a restoration treatment of western larch on snowshoe hares, a denizen of the boreal forest serving as a focal animal species to indicate the health of the restored ecosystem. We implemented a restoration treatment using “doughnut thinning” to accelerate development of old‐growth attributes in larch stands and simultaneously examined the short‐term effects on snowshoe hare density, survival, and movement. Although typical forest management activities tend to have adverse effects on hares especially in the short term, we found that the restoration treatment did not affect hare density or survival in the short term. In addition, despite significant decreases in cover coinciding with the larch needle drop, we found evidence of year‐round immigration into larch stands by hares suggesting larch stands are suitable year‐round hare habitat. Taken together, our findings suggest that a larch restoration treatment designed to accelerate the development of old‐growth attributes can be implemented so as to have no measurable short‐term detrimental effects on hares.}, number={1}, journal={RESTORATION ECOLOGY}, author={Kumar, Alexander V. and Sparks, James R. and Mills, L. Scott}, year={2018}, month={Jan}, pages={156–164} } @article{mills_bragina_kumar_zimova_lafferty_feltner_davis_hacklaender_alves_good_et al._2018, title={Winter color polymorphisms identify global hot spots for evolutionary rescue from climate change}, volume={359}, ISSN={["1095-9203"]}, DOI={10.1126/science.aan8097}, abstractNote={Changing coats with the season Many species of mammals and birds molt from summer brown to winter white coats to facilitate camouflage. Mills et al. mapped global patterns of seasonal coat color change across eight species including hares, weasels, and foxes. They found regions where individuals molt to white, brown, and both white and brown winter coats. Greater proportions of the populations molted to white in higher latitudes. Regions where seasonal coat changes are the most variable (molting to both brown and white) may provide resilience against the warming climate. Science, this issue p. 1033 Variation in winter coat color may protect species with winter camouflage—such as hares, weasels, and foxes—when the climate warms. Maintenance of biodiversity in a rapidly changing climate will depend on the efficacy of evolutionary rescue, whereby population declines due to abrupt environmental change are reversed by shifts in genetically driven adaptive traits. However, a lack of traits known to be under direct selection by anthropogenic climate change has limited the incorporation of evolutionary processes into global conservation efforts. In 21 vertebrate species, some individuals undergo a seasonal color molt from summer brown to winter white as camouflage against snow, whereas other individuals remain brown. Seasonal snow duration is decreasing globally, and fitness is lower for winter white animals on snowless backgrounds. Based on 2713 georeferenced samples of known winter coat color—from eight species across trophic levels—we identify environmentally driven clinal gradients in winter coat color, including polymorphic zones where winter brown and white morphs co-occur. These polymorphic zones, underrepresented by existing global protected area networks, indicate hot spots for evolutionary rescue in a changing climate.}, number={6379}, journal={SCIENCE}, author={Mills, L. Scott and Bragina, Eugenia V. and Kumar, Alexander V. and Zimova, Marketa and Lafferty, Diana J. R. and Feltner, Jennifer and Davis, Brandon M. and Hacklaender, Klaus and Alves, Paulo C. and Good, Jeffrey M. and et al.}, year={2018}, month={Mar}, pages={1033–1036} } @article{ferreira_alves_callahan_marques_mills_good_melo-ferreira_2017, title={The transcriptional landscape of seasonal coat colour moult in the snowshoe hare}, volume={26}, ISSN={["1365-294X"]}, DOI={10.1111/mec.14177}, abstractNote={Seasonal coat colour change is an important adaptation to seasonally changing environments but the evolution of this and other circannual traits remains poorly understood. In this study, we use gene expression to understand seasonal coat colour moulting in wild snowshoe hares (Lepus americanus). We used hair colour to follow the progression of the moult, simultaneously sampling skin from three moulting stages in hares collected during the peak of the spring moult from white winter to brown summer pelage. Using RNA sequencing, we tested whether patterns of expression were consistent with predictions based on the established phases of the hair growth cycle. We found functionally consistent clustering across skin types, with 766 genes differentially expressed between moult stages. “White” pelage showed more differentially expressed genes that were upregulated relative to other skin types, involved in the transition between late telogen (quiescent stage) and the onset of anagen (proliferative stage). Skin samples from transitional “intermediate” and “brown” pelage were transcriptionally similar and resembled the regressive transition to catagen (regressive stage). We also detected differential expression of several key circadian clock and pigmentation genes, providing important means to dissect the bases of alternate seasonal colour morphs. Our results reveal that pelage colour is a useful biomarker for seasonal change but that there is a consistent lag between the main gene expression waves and change in visible coat colour. These experiments establish that developmental sampling from natural populations of nonmodel organisms can provide a crucial resource to dissect the genetic basis and evolution of complex seasonally changing traits.}, number={16}, journal={MOLECULAR ECOLOGY}, author={Ferreira, Mafalda S. and Alves, Paulo C. and Callahan, Colin M. and Marques, Joao P. and Mills, L. Scott and Good, Jeffrey M. and Melo-Ferreira, Jose}, year={2017}, month={Aug}, pages={4173–4185} } @article{horne_hervert_woodruff_mills_2016, title={Evaluating the benefit of captive breeding and reintroductions to endangered Sonoran pronghorn}, volume={196}, ISSN={["1873-2917"]}, DOI={10.1016/j.biocon.2016.02.005}, abstractNote={Sonoran pronghorn (Antilocapra americana sonoriensis), an endangered subspecies of American pronghorn, are of great conservation concern in the southwestern U. S. Following a dramatic population decline in 2002, the U.S. Fish and Wildlife Service (USFWS) began a captive breeding program that has subsequently been used to supplement the wild population. Additionally, in 2009 the USFWS proposed to establish another, self-sustaining population outside of their range at that time. We modeled Sonoran pronghorn population dynamics based on time-series of abundance and conducted a population viability analysis (PVA) to evaluate the benefit of these management actions. We found that rates of change in the Sonoran pronghorn population were closely tied to the amount of precipitation in the area but that viability was greatly enhanced by conservation actions including the maintenance of a captive population, as well as the establishment of two additional populations outside the current range.}, journal={BIOLOGICAL CONSERVATION}, author={Horne, Jon S. and Hervert, John J. and Woodruff, Susannah P. and Mills, L. Scott}, year={2016}, month={Apr}, pages={133–146} } @article{zimova_mills_nowak_2016, title={High fitness costs of climate change-induced camouflage mismatch}, volume={19}, ISSN={["1461-0248"]}, DOI={10.1111/ele.12568}, abstractNote={Anthropogenic climate change has created myriad stressors that threaten to cause local extinctions if wild populations fail to adapt to novel conditions. We studied individual and population-level fitness costs of a climate change-induced stressor: camouflage mismatch in seasonally colour molting species confronting decreasing snow cover duration. Based on field measurements of radiocollared snowshoe hares, we found strong selection on coat colour molt phenology, such that animals mismatched with the colour of their background experienced weekly survival decreases up to 7%. In the absence of adaptive response, we show that these mortality costs would result in strong population-level declines by the end of the century. However, natural selection acting on wide individual variation in molt phenology might enable evolutionary adaptation to camouflage mismatch. We conclude that evolutionary rescue will be critical for hares and other colour molting species to keep up with climate change.}, number={3}, journal={ECOLOGY LETTERS}, author={Zimova, Marketa and Mills, L. Scott and Nowak, J. Joshua}, year={2016}, month={Mar}, pages={299–307} } @article{goldberg_tempa_norbu_hebblewhite_mills_wangchuk_lukacs_2015, title={Examining Temporal Sample Scale and Model Choice with Spatial Capture-Recapture Models in the Common Leopard Panthera pardus}, volume={10}, ISSN={["1932-6203"]}, DOI={10.1371/journal.pone.0140757}, abstractNote={Many large carnivores occupy a wide geographic distribution, and face threats from habitat loss and fragmentation, poaching, prey depletion, and human wildlife-conflicts. Conservation requires robust techniques for estimating population densities and trends, but the elusive nature and low densities of many large carnivores make them difficult to detect. Spatial capture-recapture (SCR) models provide a means for handling imperfect detectability, while linking population estimates to individual movement patterns to provide more accurate estimates than standard approaches. Within this framework, we investigate the effect of different sample interval lengths on density estimates, using simulations and a common leopard (Panthera pardus) model system. We apply Bayesian SCR methods to 89 simulated datasets and camera-trapping data from 22 leopards captured 82 times during winter 2010–2011 in Royal Manas National Park, Bhutan. We show that sample interval length from daily, weekly, monthly or quarterly periods did not appreciably affect median abundance or density, but did influence precision. We observed the largest gains in precision when moving from quarterly to shorter intervals. We therefore recommend daily sampling intervals for monitoring rare or elusive species where practicable, but note that monthly or quarterly sample periods can have similar informative value. We further develop a novel application of Bayes factors to select models where multiple ecological factors are integrated into density estimation. Our simulations demonstrate that these methods can help identify the “true” explanatory mechanisms underlying the data. Using this method, we found strong evidence for sex-specific movement distributions in leopards, suggesting that sexual patterns of space-use influence density. This model estimated a density of 10.0 leopards/100 km2 (95% credibility interval: 6.25–15.93), comparable to contemporary estimates in Asia. These SCR methods provide a guide to monitor and observe the effect of management interventions on leopards and other species of conservation interest.}, number={11}, journal={PLOS ONE}, author={Goldberg, Joshua F. and Tempa, Tshering and Norbu, Nawang and Hebblewhite, Mark and Mills, L. Scott and Wangchuk, Tshewang R. and Lukacs, Paul}, year={2015}, month={Nov} } @article{cheng_hodges_melo-ferreira_alves_mills_2014, title={Conservation implications of the evolutionary history and genetic diversity hotspots of the snowshoe hare}, volume={23}, ISSN={["1365-294X"]}, DOI={10.1111/mec.12790}, abstractNote={With climate warming, the ranges of many boreal species are expected to shift northward and to fragment in southern peripheral ranges. To understand the conservation implications of losing southern populations, we examined range‐wide genetic diversity of the snowshoe hare (Lepus americanus), an important prey species that drives boreal ecosystem dynamics. We analysed microsatellite (8 loci) and mitochondrial DNA sequence (cytochrome b and control region) variation in almost 1000 snowshoe hares. A hierarchical structure analysis of the microsatellite data suggests initial subdivision in two groups, Boreal and southwestern. The southwestern group further splits into Greater Pacific Northwest and U.S. Rockies. The genealogical information retrieved from mtDNA is congruent with the three highly differentiated and divergent groups of snowshoe hares. These groups can correspond with evolutionarily significant units that might have evolved in separate refugia south and east of the Pleistocene ice sheets. Genetic diversity was highest at mid‐latitudes of the species' range, and genetic uniqueness was greatest in southern populations, consistent with substructuring inferred from both mtDNA and microsatellite analyses at finer levels of analysis. Surprisingly, snowshoe hares in the Greater Pacific Northwest mtDNA lineage were more closely related to black‐tailed jackrabbits (Lepus californicus) than to other snowshoe hares, which may result from secondary introgression or shared ancestral polymorphism. Given the genetic distinctiveness of southern populations and minimal gene flow with their northern neighbours, fragmentation and loss of southern boreal habitats could mean loss of many unique alleles and reduced evolutionary potential.}, number={12}, journal={MOLECULAR ECOLOGY}, author={Cheng, Ellen and Hodges, Karen E. and Melo-Ferreira, Jose and Alves, Paulo C. and Mills, L. Scott}, year={2014}, month={Jun}, pages={2929–2942} } @article{melo-ferreira_seixas_cheng_mills_alves_2014, title={The hidden history of the snowshoe hare, Lepus americanus: extensive mitochondrial DNA introgression inferred from multilocus genetic variation}, volume={23}, ISSN={["1365-294X"]}, DOI={10.1111/mec.12886}, abstractNote={Hybridization drives the evolutionary trajectory of many species or local populations, and assessing the geographic extent and genetic impact of interspecific gene flow may provide invaluable clues to understand population divergence or the adaptive relevance of admixture. In North America, hares (Lepus spp.) are key species for ecosystem dynamics and their evolutionary history may have been affected by hybridization. Here we reconstructed the speciation history of the three most widespread hares in North America – the snowshoe hare (Lepus americanus), the white‐tailed jackrabbit (L. townsendii) and the black‐tailed jackrabbit (L. californicus) – by analysing sequence variation at eight nuclear markers and one mitochondrial DNA (mtDNA) locus (6240 bp; 94 specimens). A multilocus–multispecies coalescent‐based phylogeny suggests that L. americanus diverged ~2.7 Ma and that L. californicus and L. townsendii split more recently (~1.2 Ma). Within L. americanus, a deep history of cryptic divergence (~2.0 Ma) was inferred, which coincides with major speciation events in other North American species. While the isolation‐with‐migration model suggested that nuclear gene flow was generally rare or absent among species or major genetic groups, coalescent simulations of mtDNA divergence revealed historical mtDNA introgression from L. californicus into the Pacific Northwest populations of L. americanus. This finding marks a history of past reticulation between these species, which may have affected other parts of the genome and influence the adaptive potential of hares during climate change.}, number={18}, journal={MOLECULAR ECOLOGY}, author={Melo-Ferreira, Jose and Seixas, Fernando A. and Cheng, Ellen and Mills, L. Scott and Alves, Paulo C.}, year={2014}, month={Sep}, pages={4617–4630} } @article{mills_zimova_oyler_running_abatzoglou_lukacs_2013, title={Camouflage mismatch in seasonal coat color due to decreased snow duration}, volume={110}, ISSN={["0027-8424"]}, DOI={10.1073/pnas.1222724110}, abstractNote={Most examples of seasonal mismatches in phenology span multiple trophic levels, with timing of animal reproduction, hibernation, or migration becoming detached from peak food supply. The consequences of such mismatches are difficult to link to specific future climate change scenarios because the responses across trophic levels have complex underlying climate drivers often confounded by other stressors. In contrast, seasonal coat color polyphenism creating camouflage against snow is a direct and potentially severe type of seasonal mismatch if crypsis becomes compromised by the animal being white when snow is absent. It is unknown whether plasticity in the initiation or rate of coat color change will be able to reduce mismatch between the seasonal coat color and an increasingly snow-free background. We find that natural populations of snowshoe hares exposed to 3 y of widely varying snowpack have plasticity in the rate of the spring white-to-brown molt, but not in either the initiation dates of color change or the rate of the fall brown-to-white molt. Using an ensemble of locally downscaled climate projections, we also show that annual average duration of snowpack is forecast to decrease by 29–35 d by midcentury and 40–69 d by the end of the century. Without evolution in coat color phenology, the reduced snow duration will increase the number of days that white hares will be mismatched on a snowless background by four- to eightfold by the end of the century. This novel and visually compelling climate change-induced stressor likely applies to >9 widely distributed mammals with seasonal coat color.}, number={18}, journal={PROCEEDINGS OF THE NATIONAL ACADEMY OF SCIENCES OF THE UNITED STATES OF AMERICA}, author={Mills, L. Scott and Zimova, Marketa and Oyler, Jared and Running, Steven and Abatzoglou, John T. and Lukacs, Paul M.}, year={2013}, month={Apr}, pages={7360–7365} } @article{taylor_tack_naugle_mills_2013, title={Combined Effects of Energy Development and Disease on Greater Sage-Grouse}, volume={8}, ISSN={["1932-6203"]}, DOI={10.1371/journal.pone.0071256}, abstractNote={Species of conservation concern are increasingly threatened by multiple, anthropogenic stressors which are outside their evolutionary experience. Greater sage-grouse are highly susceptible to the impacts of two such stressors: oil and gas (energy) development and West Nile virus (WNv). However, the combined effects of these stressors and their potential interactions have not been quantified. We used lek (breeding ground) counts across a landscape encompassing extensive local and regional variation in the intensity of energy development to quantify the effects of energy development on lek counts, in years with widespread WNv outbreaks and in years without widespread outbreaks. We then predicted the effects of well density and WNv outbreak years on sage-grouse in northeast Wyoming. Absent an outbreak year, drilling an undeveloped landscape to a high permitting level (3.1 wells/km2) resulted in a 61% reduction in the total number of males counted in northeast Wyoming (total count). This was similar in magnitude to the 55% total count reduction that resulted from an outbreak year alone. However, energy-associated reductions in the total count resulted from a decrease in the mean count at active leks, whereas outbreak-associated reductions resulted from a near doubling of the lek inactivity rate (proportion of leks with a last count = 0). Lek inactivity quadrupled when 3.1 wells/km2 was combined with an outbreak year, compared to no energy development and no outbreak. Conservation measures should maintain sagebrush landscapes large and intact enough so that leks are not chronically reduced in size due to energy development, and therefore vulnerable to becoming inactive due to additional stressors.}, number={8}, journal={PLOS ONE}, author={Taylor, Rebecca L. and Tack, Jason D. and Naugle, David E. and Mills, L. Scott}, year={2013}, month={Aug} } @book{mills_2013, title={Conservation of wildlife populations : demography, genetics, and management (2nd Ed.)}, ISBN={9780470671504}, publisher={Chichester, West Sussex ; Hoboken, NJ : Wiley-Blackwell}, author={Mills, L. Scott}, year={2013} } @article{witczuk_pagacz_mills_2013, title={Disproportionate predation on endemic marmots by invasive coyotes}, volume={94}, ISSN={["1545-1542"]}, DOI={10.1644/12-mamm-a-199.1}, abstractNote={Abstract We investigated predation by nonnative coyotes (Canis latrans) on endemic Olympic marmots (Marmota olympus) in Olympic National Park, Washington, in 2005 and 2006. Although nearly the entire marmot range is protected within the park, declines and local extirpations of the species have been documented. Through analyses of carnivore scat across the range of the Olympic marmot we determined the distribution and relative density of coyotes and characterized the extent to which coyotes and native carnivores preyed on marmots. We used mitochondrial DNA analysis of scats to determine carnivore species, and microsatellite markers for individual coyote identification. Scat analysis indicated that invasive coyotes are widespread and the numerically dominant carnivore on sampled trails within the Olympic highlands—71% (301 of 426) of all scats verified to species arose from coyote. Out of all carnivore scats collected, 11.6% (111 of 958) contained marmot remains. For 85% of the samples with marmots, coyotes were confirmed as the predator. The remainder arose from bobcat (13%) and cougar (2%). Coyotes were the predominant marmot predator across all months and in most regions of the park. Twelve out of 13 coyote individuals identified with genetic markers preyed on marmots. Marmots ranked 5th in frequency of coyote diet items, after snowshoe hares (Lepus americanus), mountain beavers (Aplodontia rufa), voles, and cervids. Scat analysis indicated that in the Olympic Mountains, the coyote as an invasive generalist predator is subsidized by abundant multiple prey, and appears to be the primary terrestrial predator on the endemic Olympic marmot. We conclude that predation by coyotes on marmots is widespread and substantial across the marmot's species range, and therefore likely driving observed marmot declines and extinctions.}, number={3}, journal={JOURNAL OF MAMMALOGY}, author={Witczuk, Julia and Pagacz, Stanislaw and Mills, L. Scott}, year={2013}, month={Jun}, pages={702–713} } @article{newby_mills_ruth_pletscher_mitchell_quigley_murphy_desimone_2013, title={Human-caused mortality influences spatial population dynamics: Pumas in landscapes with varying mortality risks}, volume={159}, ISSN={["1873-2917"]}, DOI={10.1016/j.biocon.2012.10.018}, abstractNote={An understanding of how stressors affect dispersal attributes and the contribution of local populations to multi-population dynamics are of immediate value to basic and applied ecology. Puma (Puma concolor) populations are expected to be influenced by inter-population movements and susceptible to human-induced source–sink dynamics. Using long-term datasets we quantified the contribution of two puma populations to operationally define them as sources or sinks. The puma population in the Northern Greater Yellowstone Ecosystem (NGYE) was largely insulated from human-induced mortality by Yellowstone National Park. Pumas in the western Montana Garnet Mountain system were exposed to greater human-induced mortality, which changed over the study due to the closure of a 915 km2 area to hunting. The NGYE’s population growth depended on inter-population movements, as did its ability to act as a source to the larger region. The heavily hunted Garnet area was a sink with a declining population until the hunting closure, after which it became a source with positive intrinsic growth and a 16× increase in emigration. We also examined the spatial and temporal characteristics of individual dispersal attributes (emigration, dispersal distance, establishment success) of subadult pumas (N = 126). Human-caused mortality was found to negatively impact all three dispersal components. Our results demonstrate the influence of human-induced mortality on not only within population vital rates, but also inter-population vital rates, affecting the magnitude and mechanisms of local population’s contribution to the larger metapopulation.}, journal={BIOLOGICAL CONSERVATION}, author={Newby, Jesse R. and Mills, L. Scott and Ruth, Toni K. and Pletscher, Daniel H. and Mitchell, Michael S. and Quigley, Howard B. and Murphy, Kerry M. and DeSimone, Rich}, year={2013}, month={Mar}, pages={230–239} } @article{tempa_hebblewhite_mills_wangchuk_norbu_wangchuk_nidup_dendup_wangchuk_wangdi_et al._2013, title={Royal Manas National Park, Bhutan: a hot spot for wild felids}, volume={47}, ISSN={["1365-3008"]}, DOI={10.1017/s0030605312001317}, abstractNote={Abstract The non-uniformity of the distribution of biodiversity makes allocation of the limited resources available for conservation of biodiversity a difficult task. Approaches such as biodiversity hotspot identification, endemic bird areas, crisis ecoregions, global 200 ecoregions, and the Last of the Wild are used by scientists and international conservation agencies to prioritize conservation efforts. As part of the biodiverse Eastern Himalayan region, Bhutan has been identified as a conservation priority area by all these different approaches, yet data validating these assessments are limited. To examine whether Bhutan is a biodiversity hot spot for a key taxonomic group, we conducted camera trapping in the lower foothills of Bhutan, in Royal Manas National Park, from November 2010 to February 2011. We recorded six species of wild felids of which five are listed on the IUCN Red List: tiger Panthera tigris, golden cat Pardofelis temminckii, marbled cat Pardofelis marmorata, leopard cat Prionailurus bengalensis, clouded leopard Neofelis nebulosa and common leopard Panthera pardus. Our study area of 74 km2 has c. 16% of felid species, confirming Bhutan as a biodiversity hot spot for this group.}, number={2}, journal={ORYX}, author={Tempa, Tshering and Hebblewhite, Mark and Mills, L. Scott and Wangchuk, Tshewang R. and Norbu, Nawang and Wangchuk, Tenzin and Nidup, Tshering and Dendup, Pema and Wangchuk, Dorji and Wangdi, Yeshi and et al.}, year={2013}, month={Apr}, pages={207–210} } @inbook{mills_johnson_2013, title={Wildlife population dynamics}, ISBN={9781421409863}, booktitle={Wildlife management: Contemporary principles and practices}, publisher={Baltimore : John Hopkins University Press}, author={Mills, L. S. and Johnson, H. E.}, year={2013} } @article{hartway_mills_2012, title={A Meta-Analysis of the Effects of Common Management Actions on the Nest Success of North American Birds}, volume={26}, ISSN={["1523-1739"]}, DOI={10.1111/j.1523-1739.2012.01883.x}, abstractNote={Abstract:  Management strategies for the recovery of declining bird populations often must be made without sufficient data to predict the outcome of proposed actions or sufficient time and resources necessary to collect these data. We quantitatively reviewed studies of bird management in Canada and the United States to evaluate the relative efficacy of 4 common management interventions and to determine variables associated with their success. We compared how livestock exclusion, prescribed burning, removal of predators, and removal of cowbirds (Molothrus ater) affect bird nest success and used meta‐regression to evaluate the influence of species and study‐specific covariates on management outcomes. On average, all 4 management interventions increased nest success. When common species and threatened, endangered, or declining species (as defined by long‐term trend data from the North American Breeding Bird Survey) were analyzed together, predator removal was the most effective management option. The difference in mean nest success between treatment and control plots in predator‐removal experiments was more than twice that of either livestock exclusion or prescribed burning. However, when we considered management outcomes from only threatened, endangered, or declining species, livestock exclusions resulted in the greatest mean increase in nest success, more than twice that of the 3 other treatments. Our meta‐regression results indicated that between‐species variation accounted for approximately 86%, 40%, 35%, and 7% of the overall variation in the results of livestock‐exclusion, prescribed‐burn, predator‐removal, and cowbird‐removal studies, respectively. However, the covariates we tested explained significant variation only in outcomes among prescribed‐burn studies. The difference in nest success between burned and unburned plots displayed a significant, positive trend in association with time since fire and was significantly larger in grasslands than in woodlands. Our results highlight the importance of comparative studies on management effects in developing efficient and effective conservation strategies.}, number={4}, journal={CONSERVATION BIOLOGY}, author={Hartway, Cynthia and Mills, L. Scott}, year={2012}, month={Aug}, pages={657–666} } @article{taylor_walker_naugle_mills_2012, title={Managing multiple vital rates to maximize greater sage-grouse population growth}, volume={76}, ISSN={["1937-2817"]}, DOI={10.1002/jwmg.267}, abstractNote={Despite decades of greater sage grouse (Centrocercus urophasianus) field research, the resulting range-wide demographic data has yet to be synthesized into sensitivity analyses to guide management actions. We summarized range-wide demographic rates from 71 studies from 1938-2008 to better understand greater sage-grouse population dynamics. We used data from 38 of these studies with suitable data to parameterize a two-stage, female-based population matrix model. We conducted analytical sensitivity, elasticity, and variancestabilized sensitivity analyses to identify the contribution of each vital rate to population growth rate (λ) and life-stage simulation analysis (LSA) to determine the proportion of variation in λ accounted for by each vital rate. Greater sage grouse showed marked annual and geographic variation in multiple vital rates. Sensitivity analyses suggest that, in contrast to most other North American galliforms, female survival is as important for population growth as chick survival and more important than nest success. In lieu of quantitative data on factors driving local populations, we recommend that management efforts for sage grouse focus on increasing juvenile, yearling, and adult female survival by restoring intact sagebrush landscapes, reducing persistent sources of mortality, and eliminating anthropogenic habitat features that subsidize predators. Our analysis also supports efforts to increase chick survival and nest success by managing shrub, forb, and grass cover and height to meet published brood-rearing and nesting habitat guidelines, but not at the expense of reducing shrub cover and height below that required for survival in fall and winter.}, number={2}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={Taylor, Rebecca L. and Walker, Brett L. and Naugle, David E. and Mills, L. Scott}, year={2012}, month={Feb}, pages={336–347} } @inbook{kelly_betsch_wultsch_mesa_mills_2012, title={Noninvasive Sampling for carnivores}, ISBN={9780191625336}, DOI={10.1093/acprof:oso/9780199558520.003.0004}, abstractNote={Carnivores are notoriously difficult to study, yet recently, noninvasive carnivore studies have increased dramatically due to technological and methodological advances, and new analytical analysis techniques. For example, the ability to extract DNA from very small and/or poor quality samples, such as single hairs, and other techniques for assessing hormones from faecal samples have increased the ability to make individual and population-level inferences about carnivores. Increased availability of infrared remote cameras has facilitated large-scale surveys across vast areas, targeting multiple carnivores simultaneously. Today, researchers can estimate population size and survival, as well as historic and current rates of movement across fragmented landscapes, and can measure carnivore stress loads without ever catching, handling, or seeing a single individual animal first-hand. Noninvasive approaches draw from cutting-edge advances in genetics, population biology, biostatistics, endocrinology, and epidemiology. It is an exciting time to study carnivores using noninvasive sampling techniques.}, booktitle={Carnivore ecology and conservation: a handbook of techniques}, publisher={Oxford : Oxford University Press}, author={Kelly, M. J. and Betsch, J. and Wultsch, C. and Mesa, B. and Mills, L. S.}, editor={Boitani, L. and Powell, R.Editors}, year={2012} } @book{wildlife research techniques in rugged mountainous asian landscape_2012, year={2012} } @article{lewis_hodges_koehler_mills_2011, title={Influence of stand and landscape features on snowshoe hare abundance in fragmented forests}, volume={92}, ISSN={["1545-1542"]}, DOI={10.1644/10-mamm-a-095.1}, abstractNote={Abstract Habitat fragmentation often separates and reduces populations of vertebrates, but the relative effects of habitat attributes within remnant patches versus the matrix surrounding the patches are less clear. For snowshoe hares (Lepus americanus) lower densities and disrupted cycles in their southern range have been ascribed to habitat fragmentation, although relevant scales of landscape influence remain unknown. In a fragmented forest in north-central Washington we counted fecal pellets of snowshoe hares to examine the extent to which relative snowshoe hare densities within stands of suitable habitat changed with the composition of surrounding habitats. Pellet densities were associated primarily with density of large shrubs and saplings and medium trees within a stand. Pellet densities also were correlated positively with the amount of moist forest (dominated by Engelmann spruce [Picea engelmannii] and subalpine fir [Abies lasiocarpa]) and correlated negatively with the amount of open-structured habitat within 300 m of the stand perimeter. These results suggest that forest managers will have positive impacts on hare densities by managing both focal stands and the surrounding stands for the higher densities of large shrubs and saplings and medium trees that hares select.}, number={3}, journal={JOURNAL OF MAMMALOGY}, author={Lewis, C. W. and Hodges, K. E. and Koehler, G. M. and Mills, L. S.}, year={2011}, month={Jun}, pages={561–567} } @article{johnson_mills_wehausen_stephenson_luikart_2011, title={Translating Effects of Inbreeding Depression on Component Vital Rates to Overall Population Growth in Endangered Bighorn Sheep}, volume={25}, ISSN={["0888-8892"]}, DOI={10.1111/j.1523-1739.2011.01739.x}, abstractNote={Abstract:  Evidence of inbreeding depression is commonly detected from the fitness traits of animals, yet its effects on population growth rates of endangered species are rarely assessed. We examined whether inbreeding depression was affecting Sierra Nevada bighorn sheep (Ovis canadensis sierrae), a subspecies listed as endangered under the U.S. Endangered Species Act. Our objectives were to characterize genetic variation in this subspecies; test whether inbreeding depression affects bighorn sheep vital rates (adult survival and female fecundity); evaluate whether inbreeding depression may limit subspecies recovery; and examine the potential for genetic management to increase population growth rates. Genetic variation in 4 populations of Sierra Nevada bighorn sheep was among the lowest reported for any wild bighorn sheep population, and our results suggest that inbreeding depression has reduced adult female fecundity. Despite this population sizes and growth rates predicted from matrix‐based projection models demonstrated that inbreeding depression would not substantially inhibit the recovery of Sierra Nevada bighorn sheep populations in the next approximately 8 bighorn sheep generations (48 years). Furthermore, simulations of genetic rescue within the subspecies did not suggest that such activities would appreciably increase population sizes or growth rates during the period we modeled (10 bighorn sheep generations, 60 years). Only simulations that augmented the Mono Basin population with genetic variation from other subspecies, which is not currently a management option, predicted significant increases in population size. Although we recommend that recovery activities should minimize future losses of genetic variation, genetic effects within these endangered populations—either negative (inbreeding depression) or positive (within subspecies genetic rescue)—appear unlikely to dramatically compromise or stimulate short‐term conservation efforts. The distinction between detecting the effects of inbreeding depression on a component vital rate (e.g., fecundity) and the effects of inbreeding depression on population growth underscores the importance of quantifying inbreeding costs relative to population dynamics to effectively manage endangered populations.}, number={6}, journal={CONSERVATION BIOLOGY}, author={Johnson, Heather E. and Mills, L. Scott and Wehausen, John D. and Stephenson, Thomas R. and Luikart, Gordon}, year={2011}, month={Dec}, pages={1240–1249} } @article{johnson_mills_wehausen_stephenson_2010, title={Combining ground count, telemetry, and mark-resight data to infer population dynamics in an endangered species}, volume={47}, ISSN={["0021-8901"]}, DOI={10.1111/j.1365-2664.2010.01846.x}, abstractNote={Summary 1. To successfully manipulate populations for management and conservation purposes, managers must be able to track changes in demographic rates and determine the factors driving spatial and temporal variation in those rates. For populations of management concern, however, data deficiencies frequently limit the use of traditional statistical methods for such analyses. Long-term demographic data are often piecemeal, having small sample sizes, inconsistent methodologies, intermittent data, and information on only a subset of important parameters and covariates. 2. We evaluated the effectiveness of Bayesian state-space models for meeting these data limitations in elucidating dynamics of federally endangered Sierra Nevada bighorn sheep Ovis canadensis sierrae. We combined ground count, telemetry, and mark–resight data to: (1) estimate demographic parameters in three populations (including stage-specific abundances and vital rates); and (2) determine whether density, summer precipitation, or winter severity were driving variation in key demographic rates. 3. Models combining all existing data types increased the precision and accuracy in parameter estimates and fit covariates to vital rates driving population performance. They also provided estimates for all years of interest (including years in which field data were not collected) and standardized the error structure across data types. 4. Demographic rates indicated that recovery efforts should focus on increasing adult and yearling survival in the smallest bighorn sheep population. In evaluating covariates we found evidence of negative density dependence in the larger herds, but a trend of positive density dependence in the smallest herd suggesting that an augmentation may be needed to boost performance. We also found that vital rates in all populations were positively associated with summer precipitation, but that winter severity only had a negative effect on the smallest herd, the herd most strongly impacted by environmental stochasticity. 5. Synthesis and applications. For populations with piecemeal data, a problem common to both endangered and harvested species, obtaining precise demographic parameter estimates is one of the greatest challenges in detecting population trends, diagnosing the causes of decline, and directing management. Data on Sierra Nevada bighorn sheep provide an example of the application of Bayesian state-space models for combining all existing data to meet these objectives and better inform important management and conservation decisions.}, number={5}, journal={JOURNAL OF APPLIED ECOLOGY}, author={Johnson, Heather E. and Mills, L. Scott and Wehausen, John D. and Stephenson, Thomas R.}, year={2010}, month={Oct}, pages={1083–1093} } @article{pierson_mills_christian_2010, title={Foraging patterns of cavity-nesting birds in fire-suppressed and prescribe-burned ponderosa pine forests in Montana}, volume={41}, DOI={10.2174/1876325101004010041}, abstractNote={Fuel-reduction/forest restoration treatments that consist of thinning followed by prescribed burning are becoming increasingly important land management actions that likely affect various wildlife species. To assess potential effects on bark-gleaning birds, we compared the foraging patterns of five cavity-nesting species in thinned and burned ponderosa pine (Pinus ponderosa) forest sites and control sites. We recorded foraging behavior, location on forage tree, and tree characteristics that may be important in the selection of foraging substrates. Foraging surveys were conducted on three replicate 20-ha thinned/burned plots located within larger treatments that ranged from 60 - 250 ha, paired with three replicate control plots. Red-breasted Nuthatches (Sitta canadensis) foraged more often in control sites. Mountain Chickadees (Poecile gambeli) foraged at similar rates on both treatment types. Black-backed Woodpeckers (Picoides arcticus), Hairy Woodpeckers (P. villosus) and White-breasted Nuthatches (Sitta carolinensis) foraged almost exclusively in thinned/burned sites. Overall, all species selectively foraged on larger diameter trees. In control sites, Red-breasted Nuthatches selected larger ponderosa pine trees and Mountain Chickadees selected larger, live trees. In thinned/burned sites, Red-breasted Nuthatches selected larger, live trees, Mountain Chickadees selected larger trees with more canopy connections, Black-backed Woodpeckers selected trees with beetle evidence present and Hairy Woodpeckers selected recently dead trees. These results suggest fuel reduction/forest restoration treatments in dry ponderosa pine forests may be compatible with providing foraging substrates for cavity-nesting species often present in post-fire habitats.}, journal={Open Environmental Sciences}, author={Pierson, J. C. and Mills, L. S. and Christian, D. P}, year={2010}, pages={41–52} } @article{johnson_mills_stephenson_wehausen_2010, title={Population-specific vital rate contributions influence management of an endangered ungulate}, volume={20}, ISSN={["1939-5582"]}, DOI={10.1890/09-1107.1}, abstractNote={To develop effective management strategies for the recovery of threatened and endangered species, it is critical to identify those vital rates (survival and reproductive parameters) responsible for poor population performance and those whose increase will most efficiently change a population's trajectory. In actual application, however, approaches identifying key vital rates are often limited by inadequate demographic data, by unrealistic assumptions of asymptotic population dynamics, and of equal, infinitesimal changes in mean vital rates. We evaluated the consequences of these limitations in an analysis of vital rates most important in the dynamics of federally endangered Sierra Nevada bighorn sheep (Ovis canadensis sierrae). Based on data collected from 1980 to 2007, we estimated vital rates in three isolated populations, accounting for sampling error, variance, and covariance. We used analytical sensitivity analysis, life-stage simulation analysis, and a novel non-asymptotic simulation approach to (1) identify vital rates that should be targeted for subspecies recovery; (2) assess vital rate patterns of endangered bighorn sheep relative to other ungulate populations; (3) evaluate the performance of asymptotic vs. non-asymptotic models for meeting short-term management objectives; and (4) simulate management scenarios for boosting bighorn sheep population growth rates. We found wide spatial and temporal variation in bighorn sheep vital rates, causing rates to vary in their importance to different populations. As a result, Sierra Nevada bighorn sheep exhibited population-specific dynamics that did not follow theoretical expectations or those observed in other ungulates. Our study suggests that vital rate inferences from large, increasing, or healthy populations may not be applicable to those that are small, declining, or endangered. We also found that, while asymptotic approaches were generally applicable to bighorn sheep conservation planning; our non-asymptotic population models yielded unexpected results of importance to managers. Finally, extreme differences in the dynamics of individual bighorn sheep populations imply that effective management strategies for endangered species recovery may often need to be population-specific.}, number={6}, journal={ECOLOGICAL APPLICATIONS}, author={Johnson, Heather E. and Mills, L. Scott and Stephenson, Thomas R. and Wehausen, John D.}, year={2010}, month={Sep}, pages={1753–1765} } @book{foresman_mills_phurba_2010, title={Procedures for implementing small mammal inventories in Bhutan}, publisher={Bhutan: Ugyen Wangchuck Institute for Conservation and the Environment}, author={Foresman, K. R. and Mills, L.S. and Phurba}, year={2010} } @article{griffin_taper_hoffman_mills_2010, title={Ranking Mahalanobis Distance Models for Predictions of Occupancy From Presence-Only Data}, volume={74}, ISSN={["0022-541X"]}, DOI={10.2193/2009-002}, abstractNote={Abstract The Mahalanobis distance statistic (D2) has emerged as an effective tool to identify suitable habitat from presence data alone, but there has been no mechanism to select among potential habitat covariates. We propose that the best combination of explanatory variables for a D2 model can be identified by ranking potential models based on the proportion of the entire study area that is classified as potentially suitable habitat given that a predetermined proportion of occupied locations are correctly classified. In effect, our approach seeks to minimize errors of commission, or maximize specificity, while holding the omission error rate constant. We used this approach to identify potentially suitable habitat for the Olympic marmot (Marmota olympus), a declining species endemic to Olympic National Park, Washington, USA. We compared models built with all combinations of 11 habitat variables. A 7-variable model identified 21,143 ha within the park as potentially suitable for marmots, correctly classifying 80% of occupied locations. Additional refinements to the 7-variable model (e.g., eliminating small patches) further reduced the predicted area to 18,579 ha with little reduction in predictive power. Although we sought a model that would allow field workers to find 80% of Olympic marmot locations, in fact, <3% of 376 occupied locations and <9% of abandoned locations were >100 m from habitat predicted by the final model, suggesting that >90% of occupied marmot habitat could be found by observant workers surveying predicted habitat. The model comparison procedure allowed us to identify the suite of covariates that maximized specificity of our model and, thus, limited the amount of less favorable habitat included in the final prediction area. We expect that by maximizing specificity of models built from presence-only data, our model comparison procedure will be useful to conservation practitioners planning reintroductions, searching for rare species, or identifying habitat for protection.}, number={5}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={Griffin, Suzanne C. and Taper, Mark L. and Hoffman, Roger and Mills, L. Scott}, year={2010}, month={Jul}, pages={1112–1121} } @article{hebblewhite_musiani_mills_2010, title={Restoration of genetic connectivity among Northern Rockies wolf populations}, volume={19}, ISSN={["1365-294X"]}, DOI={10.1111/j.1365-294x.2010.04770.x}, abstractNote={Probably no conservation genetics issue is currently more controversial than the question of whether grey wolves (Canis lupus) in the Northern Rockies have recovered to genetically effective levels. Following the dispersal‐based recolonization of Northwestern Montana from Canada, and reintroductions to Yellowstone and Central Idaho, wolves have vastly exceeded population recovery goals of 300 wolves distributed in at least 10 breeding pairs in each of Wyoming, Idaho and Montana. With >1700 wolves currently, efforts to delist wolves from endangered status have become mired in legal battles over the distinct population segment (DPS) clause of the Endangered Species Act (ESA), and whether subpopulations within the DPS were genetically isolated. An earlier study by vonHoldt et al. (2008) suggested Yellowstone National Park wolves were indeed isolated and was used against delisting in 2008. Since then, wolves were temporarily delisted, and a first controversial hunting season occurred in fall of 2009. Yet, concerns over the genetic recovery of wolves in the Northern Rockies remain, and upcoming District court rulings in the summer of 2010 will probably include consideration of gene flow between subpopulations. In this issue of Molecular Ecology, vonHoldt et al. (2010) conduct the largest analysis of gene flow and population structure of the Northern Rockies wolves to date. Using an impressive sampling design and novel analytic methods, vonHoldt et al. (2010) show substantial levels of gene flow between three identified subpopulations of wolves within the Northern Rockies, clarifying previous analyses and convincingly showing genetic recovery.}, number={20}, journal={MOLECULAR ECOLOGY}, author={Hebblewhite, Mark and Musiani, Marco and Mills, L. Scott}, year={2010}, month={Oct}, pages={4383–4385} } @article{humbert_mills_horne_dennis_2009, title={A better way to estimate population trends}, volume={118}, ISSN={["1600-0706"]}, DOI={10.1111/j.1600-0706.2009.17839.x}, abstractNote={Estimation of a population trend from a time series of abundance data is an important task in ecology, yet such estimation remains logistically and conceptually challenging in practice. First, the extent to which unequal intervals in the time series, due to missing observations or irregular sampling, compromise trend estimation is not well-known. Furthermore, the predominant trend estimation method (loglinear regression of abundance data against time) ignores the possibility of process noise, while an alternative method (the ‘diffusion approximation’) ignores observation error in the abundance data. State-space models that account for both process noise and observation error exist but have been little used. We study an adaptation of the exponential growth state-space (EGSS) model for use with missing data in the time series, and we compare its trend estimation to the status quo methods. The EGSS model provides superior estimates of trend across wide ranges of time series length and sources of variation. The performance of the EGSS model even with half of the counts in the time series missing implies that trend estimates may be improved by diverting effort away from annual monitoring and towards increasing time series length or improving precision of the abundance estimates for years that data are collected.}, number={12}, journal={OIKOS}, author={Humbert, Jean-Yves and Mills, L. Scott and Horne, Jon S. and Dennis, Brian}, year={2009}, month={Dec}, pages={1940–1946} } @article{hodges_mills_murphy_2009, title={DISTRIBUTION AND ABUNDANCE OF SNOWSHOE HARES IN YELLOWSTONE NATIONAL PARK}, volume={90}, ISSN={["0022-2372"]}, DOI={10.1644/08-mamm-a-303.1}, abstractNote={Abstract Snowshoe hares (Lepus americanus) are widespread in boreal and montane forests of North America, vary in their temporal dynamics, and are major drivers in their food webs. In Yellowstone National Park, Wyoming, hare abundance, distribution, and temporal dynamics are unknown, yet Yellowstone contains a large area within their southern range that is relatively unfragmented by human activities. The 1988 Yellowstone fires have led to extensive regenerating stands, a seral condition that elsewhere supports relatively high numbers of hares. To examine snowshoe hare dynamics in the park from 2002 to 2007, we surveyed stands within 7 cover types and estimated abundance for a subset of sites. Both livetrapping data and fecal pellet count surveys showed that snowshoe hares are rare in Yellowstone. More than 36% of surveyed stands did not support any hares. Mature forest cover types were more likely to have hares than were stands regenerating after the 1988 fires, but very few stands supported high numbers; 96% of stands had <0.5 hares/ha. Three stands that burned in 2003 had hares before the fire, but none afterward. Hare numbers fluctuated modestly over time, but patterns were not indicative of a cycle. Taken altogether, our results indicate that snowshoe hares in Yellowstone are rare, patchily distributed, and apparently acyclic, important findings both for understanding hare dynamics and for implications for the Yellowstone food web that includes the federally Threatened Canada lynx (Lynx canadensis).}, number={4}, journal={JOURNAL OF MAMMALOGY}, author={Hodges, Karen E. and Mills, L. Scott and Murphy, Kerry M.}, year={2009}, month={Aug}, pages={870–878} } @article{griffin_griffin_taper_mills_2009, title={MARMOTS ON THE MOVE? DISPERSAL IN A DECLINING MONTANE MAMMAL}, volume={90}, ISSN={["1545-1542"]}, DOI={10.1644/08-mamm-a-159r1.1}, abstractNote={Abstract Olympic marmots (Marmota olympus) are large, burrowing rodents inhabiting scattered subalpine meadows on the Olympic Peninsula, Washington. Recently, the population has declined and become increasingly fragmented. The ability of Olympic marmots to recolonize abandoned habitat and to maintain gene flow among extant populations will depend on the number and success of dispersers and the distances that they travel. We monitored 84 radiotagged Olympic marmots to determine dispersal rates, distances, and success. Contrary to previous observations, 3-year-olds were most likely to disperse, although some 2-year-olds and even some older animals, particularly males, moved as well. Of marmots known to be still on their natal home range in the spring of a given year, 16% of 2-year-old males, 50% of 3-year-old males, 17% of 2-year-old females, and 29% of 3-year-old females subsequently dispersed. Dispersal rates for 3-year-olds were slightly lower when all animals were included in the analysis regardless of whether their dispersal history was known. Males dispersed farther than females (median  =  984 m, n  =  14 versus median  =  267 m, n  = 13) and 69% of females settled within 500 m of their original home range. If the observed dispersal patterns are representative of range-wide patterns and if Olympic marmot densities remain low, successful dispersal may be too infrequent to sustain reliable recolonization of vacant habitats or even genetic or demographic rescue of isolated marmot groups.}, number={3}, journal={JOURNAL OF MAMMALOGY}, author={Griffin, Suzanne C. and Griffin, Paul C. and Taper, Mark L. and Mills, L. Scot}, year={2009}, month={Jun}, pages={686–695} } @article{griffin_mills_2009, title={Sinks without borders: snowshoe hare dynamics in a complex landscape}, volume={118}, ISSN={["1600-0706"]}, DOI={10.1111/j.1600-0706.2009.17621.x}, abstractNote={A full understanding of population dynamics of wide-ranging animals should account for the effects that movement and habitat use have on individual contributions to population growth or decline. Quantifying the per-capita, habitat-specific contribution to population growth can clarify the value of different patch types, and help to differentiate population sources from population sinks. Snowshoe hares, Lepus americanus, routinely use various habitat types in the landscapes they inhabit in the contiguous US, where managing forests for high snowshoe hare density is a priority for conservation of Canada lynx, Lynx canadensis. We estimated density and demographic rates via mark-recapture live trapping and radio-telemetry within four forest stand structure (FSS) types at three study areas within heterogeneous managed forests in western Montana. We found support for known fate survival models with time-varying individual covariates representing the proportion of locations in each of the FSS types, with survival rates decreasing as use of open young and open mature FSS types increased. The per-capita contribution to overall population growth increased with use of the dense mature or dense young FSS types and decreased with use of the open young or open mature FSS types, and relatively high levels of immigration appear to be necessary to sustain hares in the open FSS types. Our results support a conceptual model for snowshoe hares in the southern range in which sink habitats (open areas) prevent the buildup of high hare densities. More broadly, we use this system to develop a novel approach to quantify demographic sources and sinks for animals making routine movements through complex fragmented landscapes.}, number={10}, journal={OIKOS}, author={Griffin, Paul C. and Mills, L. Scott}, year={2009}, month={Oct}, pages={1487–1498} } @misc{mills_2008, title={Crossing disciplines for endangered Species (review of Scott, J. Michael, Dale D. Goble, and Frank W. Davis, editors. 2006. The Endangered Species Act at thirty. Conserving Biodiversity in Human-dominated Landscape)}, volume={89}, DOI={10.1890/br08-11.1}, abstractNote={EcologyVolume 89, Issue 2 p. 592-593 Book Review Scott, Goble, and Davis — The Endangered Species Act at thirty. Conserving Biodiversity in Human-dominated Landscapes. Volume 2 CROSSING DISCIPLINES FOR ENDANGERED SPECIES L. Scott Mills, L. Scott Mills University of Montana, Wildlife Biology Program, Department of Ecosystem and Conservation Sciences, College of Forestry and Conservation, Missoula, Montana 59812 E-mail: lscott.mills@montana.eduSearch for more papers by this author L. Scott Mills, L. Scott Mills University of Montana, Wildlife Biology Program, Department of Ecosystem and Conservation Sciences, College of Forestry and Conservation, Missoula, Montana 59812 E-mail: lscott.mills@montana.eduSearch for more papers by this author First published: 01 February 2008 https://doi.org/10.1890/BR08-11.1Read the full textAboutPDF ToolsRequest permissionExport citationAdd to favoritesTrack citation ShareShare Give accessShare full text accessShare full-text accessPlease review our Terms and Conditions of Use and check box below to share full-text version of article.I have read and accept the Wiley Online Library Terms and Conditions of UseShareable LinkUse the link below to share a full-text version of this article with your friends and colleagues. Learn more.Copy URL Share a linkShare onFacebookTwitterLinked InRedditWechat Volume89, Issue2February 2008Pages 592-593 RelatedInformation}, journal={Ecology}, author={Mills, L. S.}, year={2008}, pages={592–593} } @article{hodges_mills_2008, title={Designing fecal pellet surveys for snowshoe hares}, volume={256}, ISSN={["1872-7042"]}, DOI={10.1016/j.foreco.2008.07.015}, abstractNote={Index methods can be valuable for monitoring forest-dwelling vertebrates over broad spatial or temporal scales. Fecal pellet counts are often used as an index of density or habitat use of snowshoe hares, Lepus americanus, but previous surveys have used different plot types and sample sizes, leading to problems comparing results from different studies and questions about the inferential power of each study. In this paper, we use field data and simulations to examine how the precision, bias, and efficiency of four commonly used plot types vary with plot type, pellet density, and sample size. Although no one plot type was consistently superior, we recommend thin rectangles (5.08 cm × 305 cm (2 in. × 10 ft), 0.155 m2) or 1 m2 circles over 0.155 m2 circles or 10 cm × 10 m (1 m2) rectangles. We recommend that researchers explicitly address the power of their survey design to detect different pellet densities, because much larger sample sizes are needed at low pellet densities than at high pellet densities to obtain similar precision. Small sample sizes are also much more likely to be biased, which could lead to incorrect inferences about management of snowshoe hare populations. Both uncleared and cleared plots performed well and will have value in different research contexts.}, number={11}, journal={FOREST ECOLOGY AND MANAGEMENT}, author={Hodges, K. E. and Mills, L. S.}, year={2008}, month={Nov}, pages={1918–1926} } @article{harris_kauffman_mills_2008, title={Inferences about ungulate population dynamics derived from age ratios}, volume={72}, ISSN={["1937-2817"]}, DOI={10.2193/2007-277}, abstractNote={Abstract: Age ratios (e.g., calf:cow for elk and fawn:doe for deer) are used regularly to monitor ungulate populations. However, it remains unclear what inferences are appropriate from this index because multiple vital rate changes can influence the observed ratio. We used modeling based on elk (Cervus elaphus) life‐history to evaluate both how age ratios are influenced by stage‐specific fecundity and survival and how well age ratios track population dynamics. Although all vital rates have the potential to influence calf:adult female ratios (i.e., calf:cow ratios), calf survival explained the vast majority of variation in calf:adult female ratios due to its temporal variation compared to other vital rates. Calf:adult female ratios were positively correlated with population growth rate (Λ) and often successfully indicated population trajectories. However, calf:adult female ratios performed poorly at detecting imposed declines in calf survival, suggesting that only the most severe declines would be rapidly detected. Our analyses clarify that managers can use accurate, unbiased age ratios to monitor arguably the most important components contributing to sustainable ungulate populations, survival rate of young and Λ. However, age ratios are not useful for detecting gradual declines in survival of young or making inferences about fecundity or adult survival in ungulate populations. Therefore, age ratios coupled with independent estimates of population growth or population size are necessary to monitor ungulate population demography and dynamics closely through time.}, number={5}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={Harris, Nyeema C. and Kauffman, Matthew J. and Mills, L. Scott}, year={2008}, month={Jul}, pages={1143–1151} } @article{witczuk_pagacz_mills_2008, title={Optimising methods for monitoring programs: Olympic marmots as a case study}, volume={35}, ISSN={["1035-3712"]}, DOI={10.1071/wr07187}, abstractNote={Monitoring of rare and declining species is one of the most important tasks of wildlife managers. Here we present a large-scale, long-term monitoring program for Olympic marmot (Marmota olympus) throughout its range across a logistically challenging mountainous park. Our multiple-stage process of survey design accounts for the difficulty imposed by access to remote habitats and funding constraints. The Olympic marmot is endemic to the Olympic Mountains, Washington State, USA. Although nearly all of its range is enclosed within Olympic National Park, declines and local extirpations of the species have been documented. We considered several possible alternative survey approaches, and propose a monitoring program designed to reflect extinction–recolonisation dynamics using presence–absence data. The sampling design is based on annual surveys of a set of at least 25 randomly selected clusters (closely located groups of sites with record of current or historical occupancy by marmots), and supplemented by sampling 15 never-occupied sites to test for new colonisations. The monitoring plan provides a framework that park managers can use for assessing changes over time in Olympic marmot distribution across the range of the species. Our sampling design may serve as a useful case study for establishing monitoring programs for other species with clumped distributions.}, number={8}, journal={WILDLIFE RESEARCH}, author={Witczuk, Julia and Pagacz, Stanislaw and Mills, L. Scott}, year={2008}, pages={788–797} } @article{griffin_taper_hoffman_mills_2008, title={The case of the missing marmots: Are metapopulation dynamics or range-wide declines responsible?}, volume={141}, ISSN={["1873-2917"]}, DOI={10.1016/j.biocon.2008.03.001}, abstractNote={In the mid-1990s, anecdotal reports of Olympic marmot (Marmota olympus) disappearances from historically occupied locations suggested that the species might be declining. Concern was heightened by the precipitous decline of the Vancouver Island marmot (Marmota vancouverensis), coupled with reports that climate change was affecting other high-elevation species. However, it was unclear whether the Olympic marmot was declining or undergoing natural extinctions and recolonizations; distinguishing between normal metapopulation processes and population declines in naturally fragmented species can be difficult. From 2002–2006, we used multiple approaches to evaluate the population status of the Olympic marmot. We surveyed sites for which there were records indicating regular occupancy in the later half of the 20th century and we conducted range-wide surveys of open high-elevation habitat to establish current and recent distribution. We used these targeted and general habitat surveys to identify locations and regions that have undergone extinctions or colonizations in the past 1–4 decades. Simultaneously, we conducted detailed demographic studies, using marked and radio-tagged marmots, to estimate the observed and projected current population growth rate at nine locations. The habitat surveys indicate that local extinctions have been wide-spread, while no recolonizations were detected. Abundance at most intensive study sites declined from 2002–2006 and the demographic data indicate that these local declines are ongoing. Adult female survival in particular is considerably lower than it was historically. The spatial pattern of the extinctions is inconsistent with observed metapopulation dynamics in other marmot species and, together with very low observed dispersal rates, indicates that population is not at equilibrium.}, number={5}, journal={BIOLOGICAL CONSERVATION}, author={Griffin, Suzanne C. and Taper, Mark L. and Hoffman, Roger and Mills, L. Scott}, year={2008}, month={May}, pages={1293–1309} } @inbook{mills_2007, title={Chapter 14: North America}, ISBN={9780521705974}, booktitle={Climate change 2007 - impacts, adaptation and vulnerability working group ii contribution to the fourth assessment report of the Intergovernmental Panel on Climate Change}, publisher={New York : Cambridge University Press}, author={Mills, L. S. et al.}, editor={Field, C. B. and L. D. Mortsch, M. Brklacich and D. L. Forbes, P. Kovaks and J. A. Patz, S. W. Running and Scott, M. J.Editors}, year={2007} } @book{mills_2007, title={Conservation of wildlife populations : demography, genetics, and management (1st Ed.)}, ISBN={9781405121460}, publisher={Malden, MA : Blackwell Pub}, author={Mills, L. Scott}, year={2007} } @article{griffin_valois_taper_mills_2007, title={Effects of tourists on behavior and demography of olympic marmots}, volume={21}, ISSN={["1523-1739"]}, DOI={10.1111/j.1523-1739.2007.00688.x}, abstractNote={Abstract:  If changes in animal behavior resulting from direct human disturbance negatively affect the persistence of a given species or population, then these behavioral changes must necessarily lead to reduced demographic performance. We tested for the effects of human disturbance on Olympic marmots (Marmota olympus), a large ground‐dwelling squirrel that has disappeared from several areas where recreation levels are high. We assessed the degree to which antipredator and foraging behavior and demographic rates (survival and reproduction) differed between sites with high recreation levels (high use) and those with little or no recreation (low use). Compared with the marmots at low‐use sites, marmots at high‐use sites displayed significantly reduced responses to human approach, which could be construed as successful accommodation of disturbance or as a decrease in predator awareness. The marmots at high‐use sites also looked up more often while foraging, which suggests an increased wariness. Marmots at both types of sites had comparable reproductive and survival rates and were in similar body condition. Until now, the supposition that marmots can adjust their behavior to avoid negative demographic consequences when confronted with heavy tourism has been based on potentially ambiguous behavioral data. Our results support this hypothesis in the case of Olympic marmots and demonstrate the importance of considering demographic data when evaluating the impacts of recreation on animal populations.}, number={4}, journal={CONSERVATION BIOLOGY}, author={Griffin, Suzanne C. and Valois, Tanguy and Taper, Mark L. and Mills, L. Scott}, year={2007}, month={Aug}, pages={1070–1081} } @article{griffin_taper_mills_2007, title={Female olympic marmots (Marinota olympus) reproduce in consecutive years}, volume={158}, ISSN={["1938-4238"]}, DOI={10.1674/0003-0031(2007)158[221:fommor]2.0.co;2}, abstractNote={ABSTRACT Olympic marmots (Marmota olympus) are reported to skip at least one year between reproductive efforts. We observed several female marmots weaning infants in consecutive years. There was no evidence that reproductive skipping was more common than annual reproduction. High spring food availability resulting from climate change may allow females to wean consecutive litters regularly.}, number={1}, journal={AMERICAN MIDLAND NATURALIST}, author={Griffin, Suzanne Cox and Taper, Mark L. and Mills, L. Scott}, year={2007}, month={Jul}, pages={221–225} } @article{griffin_mills_2007, title={Precommercial thinning reduces snowshoe hare abundance in the short tenn}, volume={71}, ISSN={["1937-2817"]}, DOI={10.2193/2004-007}, abstractNote={Abstract Management of young forests is not often considered in conservation plans, but young forests provide habitat for some species of conservation concern. Snowshoe hares (Lepus americanus), critical prey of forest carnivores including the United States federally threatened Canada lynx (Lynx canadensis), can be abundant in young montane and subalpine forests with densely spaced saplings and shrub cover. Precommercial thinning (PCT) is a silvicultural technique that reduces sapling and shrub density on young forest stands. We tested for effects of PCT on snowshoe hare abundance for 2 years after experimental treatment at 3 replicate study areas. We also tested the effectiveness of a precommercial thinning with reserves (PCT-R) prescription, where 20% of the total stand was retained in uncut quarter-hectare patches. All stands were in montane–subalpine coniferous forests of western Montana, USA, where there is a persistent population of Canada lynx. Posttreatment changes in abundance were strongly negative on stands treated with standard PCT prescriptions (100% of the stand was treated), relative to both controls and stands treated with PCT-R. Trapping, snowtrack, and winter fecal-pellet indices indicated that snowshoe hares used the quarter-ha retention patches more than thinned portions of the PCT-R-treated stands in winter. We suggest that managing forest landscapes for high snowshoe hare abundance will require adoption of silvicultural techniques like PCT-R for stands that will be thinned, in addition to conservation of structurally valuable early and late-successional forest stands.}, number={2}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={Griffin, Paul C. and Mills, L. Scott}, year={2007}, month={Apr}, pages={559–564} } @article{hard_mills_peek_2006, title={Genetic implications of reduced survival of male red deer Cervus elaphus under harvest}, volume={12}, ISSN={["1903-220X"]}, DOI={10.2981/0909-6396(2006)12[427:giorso]2.0.co;2}, abstractNote={Abstract We use simple, multivariate evolutionary models to evaluate the short-term potential for size-selective harvest to reduce genetic variability and alter life history in cervids. These genetic effects limit sustainable levels of harvest of the animals because they determine how changes in sex ratio, generation length and traits contributing to fitness influence population growth rate and local adaptation. Our analysis of harvest-mediated adaptive evolution employs a genetic approach that parameterizes models with empirical data obtained from European red deer Cervus elaphus. The analysis indicates that harvest, if sufficiently high to reduce the breeding ratio of males to females to below about 15:100, can reduce effective population size to a level that threatens adaptive potential. The reduction in effective size is realized through decreases in both sex ratio of breeders and the age of breeding males. Predicted selective effects of harvest on body size indicates a weak potential to alter most life-history traits over 10 generations under two harvest scenarios; the patterns suggest that current modes of harvest are unlikely to produce substantial life-history changes in red deer over 10 or fewer generations unless the genetic influences on red deer traits are considerably higher than those predicted here. Nevertheless, male reproductive success is expected to decline detectably if male harvest rate is sufficiently high (> 30%). Collectively, our results imply that harvest methods should permit higher post-hunt male:female ratios (18:100 or higher) and ensure that a sufficient number of larger, older males survive the breeding season. The capacity of selective harvest to alter demography and life history depends heavily on the genetic covariance structure underlying variation in these traits, information that is unknown for many red deer populations. Prudent harvest management should therefore implement and monitor approaches to hunting that aim to conserve life-history variation; meanwhile, use of less selective methods can reduce the risk to long-term adaptive potential and may permit higher sustainable harvest rates.}, number={4}, journal={WILDLIFE BIOLOGY}, author={Hard, Jeffrey J. and Mills, L. Scott and Peek, James M.}, year={2006}, month={Dec}, pages={427–441} } @inbook{mills_scott_strickler_temple_2005, title={Ecology and management of small populations}, ISBN={9780933564152}, booktitle={Techniques for wildlife investigations and management (6th ed.)}, publisher={Bethesda, Md. : Wildlife Society}, author={Mills, L. S. and Scott, J. M. and Strickler, K. M. and Temple, S. A.}, year={2005}, pages={691–713} } @article{schwartz_mills_2005, title={Gene flow after inbreeding leads to higher survival in deer mice}, volume={123}, ISSN={["1873-2917"]}, DOI={10.1016/j.biocon.2004.11.016}, abstractNote={We test the ability of gene flow to alleviate the deleterious effects of inbreeding in a small mammal, the deer mouse (Peromyscus maniculatus). After three generations of sib–sib mating, individuals from three lines of mice were either subject to further inbreeding or were mated with an outbred individual. Subsequently, these mice, plus a control line, which were first generation (F1) mice from unrelated individuals kept in captivity for the same duration as the treatment lines, were released into isolated pens in a forest in western Montana. Survival of individual mice was recorded. Survival models that allowed variation in breeding treatments were well supported, whereas models explaining variation in line, or release location were not well supported. Survival was highest for offspring of the outcross group, intermediate for the inbred animals, and lowest for the control group. This suggests that the introduction of migrants can reduce inbreeding depression, as theory predicts. We also show limited evidence for purging of deleterious recessive alleles that can cause inbreeding depression. While purging may have occurred, the demographic cost was non-trivial as 5 of 8 of our inbred mouse lines went extinct during the inbreeding process.}, number={4}, journal={BIOLOGICAL CONSERVATION}, author={Schwartz, MK and Mills, LS}, year={2005}, month={Jun}, pages={413–420} } @article{mildenstein_stier_nuevo-diego_mills_2005, title={Habitat selection of endangered and endemic large flying-foxes in Subic Bay, Philippines}, volume={126}, ISSN={["1873-2917"]}, DOI={10.1016/j.biocon.2005.05.001}, abstractNote={Large flying-foxes in insular Southeast Asia are the most threatened of the Old World fruit bats due to high levels of deforestation and hunting and effectively little local conservation commitment. The forest at Subic Bay, Philippines, supports a rare, large colony of vulnerable Philippine giant fruit bats (Pteropus vampyrus lanensis) and endangered and endemic golden-crowned flying-foxes (Acerodon jubatus). These large flying-foxes are optimal for conservation focus, because in addition to being keystone, flagship, and umbrella species, the bats are important to Subic Bay’s economy and its indigenous cultures. Habitat selection information streamlines management’s efforts to protect and conserve these popular but threatened animals. We used radio telemetry to describe the bats’ nighttime use of habitat on two ecological scales: vegetation and microhabitat. The fruit bats used the entire 14,000 ha study area, including all of Subic Bay Watershed Reserve, as well as neighboring forests just outside the protected area boundaries. Their recorded foraging locations ranged between 0.4 and 12 km from the roost. We compared the bats’ use to the availability of vegetative habitat types, riparian areas, and bat trees. The fruit bats’ locations showed a preference for undisturbed forest types and selection against disturbed and agricultural areas. Bat locations also showed selection for particular fruiting/flowering bat trees. The bats showed strong preference for riparian areas; locations were in riparian areas over four times more than expected. From these results we recommend that management focus flying-fox conservation efforts on undisturbed forest and riparian areas.}, number={1}, journal={BIOLOGICAL CONSERVATION}, author={Mildenstein, TL and Stier, SC and Nuevo-Diego, CE and Mills, LS}, year={2005}, month={Nov}, pages={93–102} } @article{griffin_griffin_waroquiers_mills_2005, title={Mortality by moonlight: predation risk and the snowshoe hare}, volume={16}, ISSN={["1465-7279"]}, DOI={10.1093/beheco/ari074}, abstractNote={Optimal behavior theory suggests that prey animals will reduce activity during intermittent periods when elevated predation risk outweighs the fitness benefits of activity. Specifically, the predation risk allocation hypothesis predicts that prey activity should decrease dramatically at times of high predation risk if there is high temporal variation in predation risk but should remain relatively uniform when temporal variation in predation risk is low. To test these predictions we examined the seasonably variable response of snowshoe hares to moonlight and predation risk. Unlike studies finding uniform avoidance of moonlight in small mammals, we find that moonlight avoidance is seasonal and corresponds to seasonal variation in moonlight intensity. We radio-collared 177 wild snowshoe hares to estimate predation rates as a measure of risk and used movement distances from a sample of those animals as a measure of activity. In the snowy season, 5-day periods around full moons had 2.5 times more predation than around new moons, but that ratio of the increased predation rate was only 1.8 in the snow-free season. There was no significant increase in use of habitats with more hiding cover during full moons. Snowshoe hares' nightly movement distances decreased during high-risk full-moon periods in the snowy season but did not change according to moon phase in the snow-free season. These results are consistent with the predation risk allocation hypothesis. Copyright 2005.}, number={5}, journal={BEHAVIORAL ECOLOGY}, author={Griffin, PC and Griffin, SC and Waroquiers, C and Mills, LS}, year={2005}, pages={938–944} } @article{mills_griffin_hodges_mckelvey_ruggiero_ulizio_2005, title={Pellet count indices compared to mark-recapture estimates for evaluating snowshoe hare density}, volume={69}, ISSN={["1937-2817"]}, DOI={10.2193/0022-541x(2005)069[1053:pcictm]2.0.co;2}, abstractNote={Abstract Snowshoe hares (Lepus americanus) undergo remarkable cycles and are the primary prey base of Canada lynx (Lynx canadensis), a carnivore recently listed as threatened in the contiguous United States. Efforts to evaluate hare densities using pellets have traditionally been based on regression equations developed in the Yukon, Canada. In western Montana, we evaluated whether or not local regression equations performed better than the most recent Yukon equation and assessed whether there was concordance between pellet-based predictions and mark–recapture density estimates of hares. We developed local Montana regression equations based on 224 data points consisting of mark-recapture estimates and pellet counts, derived from 38 sites in 2 different areas sampled for 1 to 5 years using 2 different pellet plot shapes. We evaluated concordance between estimated density and predicted density based on pellet counts coupled with regression equations at 436 site-area-season combinations different from those used to develop the regression equations. At densities below 0.3 hares/ha, predicted density based on pellets tended to be greater than for mark–recapture; the difference was usually <1 hare per ha on an absolute scale, but at low densities this translated to proportional differences of 1,000% or greater. At densities above 0.7 hares/ha, pellet regressions tended to predict lower density than mark–recapture. Because local regression equations did not outperform the Yukon equation, we see little merit in further development of local regression equations unless a study is to be conducted in a formal double-sampling framework. We recommend that widespread pellet sampling be used to identify areas with very low hare densities; subsequent surveys using mark–recapture methodology can then focus on higher density areas where density inferences are more reliable.}, number={3}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={Mills, LS and Griffin, PC and Hodges, KE and McKelvey, K and Ruggiero, L and Ulizio, T}, year={2005}, month={Jul}, pages={1053–1062} } @article{hodges_mills_2005, title={Snowshoe hares in Yellowstone}, volume={13}, journal={Yellowstone Science}, author={Hodges, K. E. and Mills, L. S.}, year={2005}, pages={3–6} } @article{tallmon_mills_2004, title={Edge effects and isolation: Red-backed voles revisited}, volume={18}, ISSN={["1523-1739"]}, DOI={10.1111/j.1523-1739.2004.00439.x}, abstractNote={Abstract:  We examined demographic responses of California red‐backed voles (Clethrionomys californicus) to forest fragmentation in southwestern Oregon at sites where this species has previously shown negative responses to fragmentation. Voles were captured in live traps and released. Voles were rarely caught in clearcuts surrounding 11 forest fragments, but relative vole density did not decrease from the forest‐fragment interiors to edges. The first result agrees with previous findings at these sites 6 years earlier, but the latter result does not. There was no evidence that vole response to edge changes with fragment age. Two years of intensive mark‐recapture efforts at two forest‐fragment sites and two unfragmented (control) sites did not show negative effects of fragmentation on vole survival, an important demographic rate. Vole capture probabilities varied greatly across space and time on these four sites, which may explain the differences in vole responses to edge seen between this and the previous study. These results suggest that reliable appraisal of edge effects may be difficult for many species on small fragments because the data necessary to apply population estimators require great efforts to obtain and the use of indices leads to a confounding of detection probabilities with demographic change.}, number={6}, journal={CONSERVATION BIOLOGY}, author={Tallmon, DA and Mills, LS}, year={2004}, month={Dec}, pages={1658–1664} } @article{leberg_carloss_dugas_pilgrim_mills_green_scognamillo_2004, title={Recent record of a cougar (Puma concolor) in Louisiana, with notes on diet, based on analysis of fecal materials}, volume={3}, ISSN={["1938-5412"]}, DOI={10.1656/1528-7092(2004)003[0653:rroacp]2.0.co;2}, abstractNote={Abstract We report a sighting, supported by DNA evidence from a scat, of a cougar (Puma concolor) in southeastern Louisiana. The 16S-rRNA genotype obtained from mtDNA is one that is common throughout North America, making it difficult to determine the origin of the individual. Based on DNA and hair scale analysis, the scat contained the partially digested remains of a dog (Canis familiaris) and an eastern cottontail (Sylvilagus floridanus), indicating that the individual was successfully foraging on locally occurring prey.}, number={4}, journal={SOUTHEASTERN NATURALIST}, author={Leberg, PL and Carloss, MR and Dugas, LJ and Pilgrim, KL and Mills, LS and Green, MC and Scognamillo, D}, year={2004}, pages={653–658} } @inbook{griffin_mills_2004, title={Snowshoe hares in a dynamic managed landscape}, ISBN={9780195166460}, booktitle={Species conservation and management : case studies}, publisher={New York : Oxford University Press}, author={Griffin, P. C. and Mills, L. S.}, editor={Editors H. R. Akcakaya, M. A. Burgman and O. Kindvall, C. Wood and P. Sjogren-Gulve, J. Hatfield and McCarthy, M. A.Editors}, year={2004} } @article{griffin_bienen_gillin_mills_2003, title={Estimating pregnancy rates and litter size in snowshoe hares using ultrasound}, volume={31}, number={4}, journal={Wildlife Society Bulletin}, author={Griffin, P. C. and Bienen, L. and Gillin, C. M. and Mills, L. S.}, year={2003}, pages={1066–1072} } @article{riddle_pilgrim_mills_mckelvey_ruggiero_2003, title={Identification of mustelids using mitochondrial DNA and non-invasive sampling}, volume={4}, ISSN={["1566-0621"]}, DOI={10.1023/a:1023338622905}, number={2}, journal={CONSERVATION GENETICS}, author={Riddle, AE and Pilgrim, KL and Mills, LS and McKelvey, KS and Ruggiero, LF}, year={2003}, pages={241–243} } @article{schwartz_mills_ortega_ruggiero_allendorf_2003, title={Landscape location affects genetic variation of Canada lynx (Lynx canadensis)}, volume={12}, ISSN={["1365-294X"]}, DOI={10.1046/j.1365-294x.2003.01878.x}, abstractNote={The effect of a population's location on the landscape on genetic variation has been of interest to population genetics for more than half a century. However, most studies do not consider broadscale biogeography when interpreting genetic data. In this study, we propose an operational definition of a peripheral population, and then explore whether peripheral populations of Canada lynx (Lynx canadensis) have less genetic variation than core populations at nine microsatellite loci. We show that peripheral populations of lynx have fewer mean numbers of alleles per population and lower expected heterozygosity. This is surprising, given the lynx's capacity to move long distances, but can be explained by the fact that peripheral populations often have smaller population sizes, limited opportunities for genetic exchange and may be disproportionately affected by ebbs and flows of species’ geographical range.}, number={7}, journal={MOLECULAR ECOLOGY}, author={Schwartz, MK and Mills, LS and Ortega, Y and Ruggiero, LF and Allendorf, FW}, year={2003}, month={Jul}, pages={1807–1816} } @inbook{mills_schwartz_tallmon_lair_2003, title={Measuring and interpreting changes in connectivity for mammals in coniferous forests}, ISBN={9780521810432}, DOI={10.1017/cbo9780511615757.018}, abstractNote={Western coniferous forests have a history of natural disturbance due to fire, disease, and other factors (Agee 1993), but during the past century late-seral forests have been increasingly fragmented due to logging and development. For example, in the Pacific Northwest, less than half of pre-settlement, old-growth Douglas-fir (Pseudotsuga menziesii) forest remains, often in relatively small remnants of 100 ha or less in a matrix of clear-cuts and regenerating forest (Booth 1991, Garmon et al. 1999, Jules et al. 1999). Road building has also impacted wildlife habitat, with an average of 3.4 miles of road per square mile on United States Forest Service roaded-lands and approximately twice that on private lands (Federal Budget Consulting Group and Price-Waterhouse LLP 1997, Coghlan and Sowa 1998, Federal Register 2001, USDA 2001).}, booktitle={Mammal community dynamics : management and conservation in the coniferous forests of western North America}, publisher={New York, NY, USA : Cambridge University Press}, author={Mills, L. S. and Schwartz, M. K. and Tallmon, D. A. and Lair, K. P.}, editor={Cynthia J. Zabel, Robert G. AnthonyEditor}, year={2003} } @article{tallmon_jules_radke_mills_2003, title={Of mice and men and trillium: Cascading effects of forest fragmentation}, volume={13}, ISSN={["1051-0761"]}, DOI={10.1890/02-5111}, abstractNote={Cascading ecological effects of anthropogenic habitat fragmentation have been studied primarily in extreme cases (e.g., the isolation of habitat fragments in a novel habitat matrix such as suburban developments, reservoirs, or agricultural fields), with less attention to more subtle and widespread cases, such as habitat fragmentation due to timber harvest. Few studies have used rigorous demographic data to demonstrate the direct and indirect effects of habitat fragmentation. We trapped deer mice (Peromyscus maniculatus) at five sites over two years in southwest Oregon, USA, and used multi-state capture- recapture models to estimate deer mouse survival and movement in clearcuts, forest-frag- ment edges, forest-fragment interiors, and contiguous forests. We also estimated deer mouse densities in fragmented and unfragmented forests and combined deer mouse demographic studies with trillium (Trillium ovatum) seed predation trials to link deer mouse changes to reduced trillium recruitment previously observed at the same study sites. Mouse survival was highest in clearcuts, intermediate in forest fragments, and lowest in unfragmented (control) forests. Mouse movement among clearcuts, forest edges, and forest interiors was common over short time intervals. Collectively, demographic rates led to mouse densities that were 3-4 times higher at forest-fragment sites than at unfragmented sites. Trillium seeds were ;3 times more likely to be depredated in areas of elevated relative mouse abundance than in areas of lower relative abundance. Forest fragmentation has favored mouse populations, resulting in increased seed predation that may decrease recruitment rates and increase local extinction risks for trillium.}, number={5}, journal={ECOLOGICAL APPLICATIONS}, author={Tallmon, DA and Jules, ES and Radke, NJ and Mills, LS}, year={2003}, month={Oct}, pages={1193–1203} } @article{funk_mills_2003, title={Potential causes of population declines in forest fragments in an Amazonian frog}, volume={111}, ISSN={["0006-3207"]}, DOI={10.1016/s0006-3207(02)00274-4}, abstractNote={Forest fragmentation results in population declines and extinctions for many forest vertebrates, but little is known about the mechanisms causing declines in fragments. We investigated potential causes of declines in forest fragments for an Amazonian forest frog (Colostethus stepheni) at an experimental fragmentation study site in central Amazonian Brazil using field estimates of abundance and vital rates coupled with population simulations. Although adult male survival was not reduced by fragmentation, mean clutch size was reduced by 17%. Population simulations demonstrate that a reduction in clutch size of this magnitude is sufficient to cause the observed magnitude of population declines in fragments. Female snout-vent length was also reduced in fragments and may be related to the observed reduction in clutch size.}, number={2}, journal={BIOLOGICAL CONSERVATION}, author={Funk, WC and Mills, LS}, year={2003}, month={Jun}, pages={205–214} } @article{schwartz_mills_mckelvey_ruggiero_allendorf_2002, title={DNA reveals high dispersal synchronizing the population dynamics of Canada lynx}, volume={415}, ISSN={["0028-0836"]}, DOI={10.1038/415520a}, abstractNote={Population dynamics of Canada lynx (Lynx canadensis) have been of interest to ecologists for nearly sixty years. Two competing hypotheses concerning lynx population dynamics and large-scale spatial synchrony are currently debated. The first suggests that dispersal is substantial among lynx populations, and the second proposes that lynx at the periphery of their range exist in small, isolated patches that maintain cycle synchrony via correlation with extrinsic environmental factors. Resolving the nature of lynx population dynamics and dispersal is important both to ecological theory and to the conservation of threatened lynx populations: the lack of knowledge about connectivity between populations at the southern periphery of the lynx's geographic range delayed their legal listing in the United States. We test these competing hypotheses using microsatellite DNA markers and lynx samples from 17 collection sites in the core and periphery of the lynx's geographic range. Here we show high gene flow despite separation by distances greater than 3,100 km, supporting the dispersal hypothesis. We therefore suggest that management actions in the contiguous United States should focus on maintaining connectivity with the core of the lynx's geographic range.}, number={6871}, journal={NATURE}, author={Schwartz, MK and Mills, LS and McKelvey, KS and Ruggiero, LF and Allendorf, FW}, year={2002}, month={Jan}, pages={520–522} } @article{mckelvey_mcdaniel_mills_griffin_2002, title={Effects of plot size and shape on pellet density estimates for snowshoe hares}, volume={30}, number={3}, journal={Wildlife Society Bulletin}, author={McKelvey, K. S. and McDaniel, G. W. and Mills, L. S. and Griffin, P. C.}, year={2002}, pages={751–755} } @misc{reed_mills_dunning_menges_mckelvey_frye_beissinger_anstett_miller_2002, title={Emerging issues in population viability analysis}, volume={16}, ISSN={["1523-1739"]}, DOI={10.1046/j.1523-1739.2002.99419.x}, abstractNote={Abstract: Population viability analysis ( PVA) has become a commonly used tool in endangered species management. There is no single process that constitutes PVA, but all approaches have in common an assessment of a population's risk of extinction (or quasi extinction) or its projected population growth either under current conditions or expected from proposed management. As model sophistication increases, and software programs that facilitate PVA without the need for modeling expertise become more available, there is greater potential for the misuse of models and increased confusion over interpreting their results. Consequently, we discuss the practical use and limitations of PVA in conservation planning, and we discuss some emerging issues of PVA. We review extant issues that have become prominent in PVA, including spatially explicit modeling, sensitivity analysis, incorporating genetics into PVA, PVA in plants, and PVA software packages, but our coverage of emerging issues is not comprehensive. We conclude that PVA is a powerful tool in conservation biology for comparing alternative research plans and relative extinction risks among species, but we suggest caution in its use: (1) because PVA is a model, its validity depends on the appropriateness of the model's structure and data quality; (2) results should be presented with appropriate assessment of confidence; (3) model construction and results should be subject to external review, and (4) model structure, input, and results should be treated as hypotheses to be tested. We also suggest (5) restricting the definition of PVA to development of a formal quantitative model, (6) focusing more research on determining how pervasive density‐dependence feedback is across species, and (7) not using PVA to determine minimum population size or (8) the specific probability of reaching extinction. The most appropriate use of PVA may be for comparing the relative effects of potential management actions on population growth or persistence.}, number={1}, journal={CONSERVATION BIOLOGY}, author={Reed, JM and Mills, LS and Dunning, JB and Menges, ES and McKelvey, KS and Frye, R and Beissinger, SR and Anstett, MC and Miller, P}, year={2002}, month={Feb}, pages={7–19} } @misc{mills_2002, title={False samples are not the same as blind controls - Informal efforts to "test" a laboratory corrupt the data stream, where integrity is crucial.}, volume={415}, ISSN={["1476-4687"]}, DOI={10.1038/415471a}, abstractNote={Informal efforts to “test” a laboratory corrupt the data stream, where integrity is crucial.}, number={6871}, journal={NATURE}, author={Mills, LS}, year={2002}, month={Jan}, pages={471–471} } @misc{mills_2002, title={Genetics, Demography, and Viability of Fragmented Populations}, volume={77}, DOI={10.1086/343996}, abstractNote={Previous articleNext article No AccessEnvironmental SciencesGenetics, Demography and Viability of Fragmented Populations. Based on a symposium held in Sydney, Australia, 1998. Conservation Biology, Volume 4. Edited by Andrew G Young and , Geoffrey M Clarke. Cambridge and New York: Cambridge University Press. $110.00 (hardcover); $39.95 (paper). xviii + 438 p; ill.; index. ISBN: 0–521–78207–4 (hc); 0–521–79421–8 (pb). 2000.L Scott MillsL Scott MillsWildlife Biology Program, University of Montana, Missoula, Montana Search for more articles by this author Wildlife Biology Program, University of Montana, Missoula, MontanaPDFPDF PLUSFull Text Add to favoritesDownload CitationTrack CitationsPermissionsReprints Share onFacebookTwitterLinkedInRedditEmail SectionsMoreDetailsFiguresReferencesCited by The Quarterly Review of Biology Volume 77, Number 2June 2002 Published in association with Stony Brook University Article DOIhttps://doi.org/10.1086/343996 Views: 8Total views on this site PDF download Crossref reports no articles citing this article.}, journal={Quarterly Review of Biology}, author={Mills, L. S.}, year={2002}, pages={222–223} } @article{tallmon_draheim_mills_allendorf_2002, title={Insights into recently fragmented vole populations from combined genetic and demographic data}, volume={11}, ISSN={["1365-294X"]}, DOI={10.1046/j.1365-294x.2002.01480.x}, abstractNote={We combined demographic and genetic data to evaluate the effects of habitat fragmentation on the population structure of the California red‐backed vole (Clethrionomys californicus). We analysed variation in the mitochondrial DNA (mtDNA) control region and five nuclear microsatellite loci in small samples collected from two forest fragments and an unfragmented control site in 1990–91. We intensively sampled the same forest fragments and two different control sites in 1998 and 1999. Vole abundances fluctuated greatly at sizes below 50 individuals per fragment. Fragment populations had significantly lower mtDNA allelic diversity than controls, but not nuclear heterozygosity or numbers of alleles. The use of only trapping and/or mtDNA marker data would imply that fragment populations are at least partially isolated and vulnerable to inbreeding depression. In contrast, the abundance estimates combined with microsatellite data show that small fragment populations must be linked to nearby forests by high rates of migration. These results provide evidence for the usefulness of combining genetic and demographic data to understand nonequilibrium population structure in recently fragmented habitats.}, number={4}, journal={MOLECULAR ECOLOGY}, author={Tallmon, DA and Draheim, HM and Mills, LS and Allendorf, FW}, year={2002}, month={Apr}, pages={699–709} } @misc{hoekman_mills_howerter_devries_ball_2002, title={Sensitivity analyses of the life cycle of midcontinent mallards}, volume={66}, ISSN={["1937-2817"]}, DOI={10.2307/3803153}, abstractNote={Relationships between vital rates and population growth rate (λ) are critical to understanding and managing population dynamics. Considerable study of the midcontinent mallard (Anas platyrhynchos) population has been directed to understanding how vital rates respond to environmental fluctuations and management, but inference to the relative importance of specific vital rates to λ remains weak. We used analytic and simulation-based sensitivity analyses of a stage-based matrix model of female midcontinent mallards to compare the relative importance of vital rates to λ. For each vital rate, we estimated mean values and process variation (biological variation across space and time) for females breeding on sites of approximately 70 km 2 in the Prairie Pothole Region (PPR) of the United States (Montana, North Dakota, South Dakota, Minnesota) and Canada (Saskatchewan, Manitoba, Alberta). We conducted perturbation analyses (i.e., analytic sensitivities and elasticities) to predict the relative influence of changes in vital rates on λ. We conducted variance decomposition analyses to assess the proportion of spatial and temporal variation in λ explained by process variation in each vital rate. At mean values of vital rates, analytic sensitivity of λ was highest to nest success and survival of adult females during the breeding season and non-breeding season; hence, equal absolute changes in these vital rates would be predicted to result in the largest Δλ, relative to other vital rates. Variation in sensitivities and elasticities across process variation in vital rates was primarily explained by variation in nest success and survival of ducklings. Process variation in breeding parameters was driving variation in λ: vital rates explaining the most variation were nest success (43%), survival of adult females during the breeding season (19%), and survival of ducklings (14%). Survival of adult females outside the breeding season accounted for only 9% of variation in λ. Our analyses suggested that predation processes on the breeding grounds were the primary proximate factors limiting population growth.}, number={3}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={Hoekman, ST and Mills, LS and Howerter, DW and Devries, JH and Ball, IJ}, year={2002}, month={Jul}, pages={883–900} } @inbook{mills_lindberg_2002, title={Sensitivity analysis to evaluate the consequences of conservation actions}, ISBN={9780226041773}, booktitle={Population viability analysis}, publisher={Chicago : University of Chicago Press}, author={Mills, L. S. and Lindberg, M. S.}, editor={Beissinger, Steven R. and McCullough, Dale R.Editors}, year={2002} } @article{biek_mills_bury_2002, title={Terrestrial and stream amphibians across clearcut-forest interfaces in the Siskiyou Mountains, Oregon}, volume={76}, number={2}, journal={Northwest Science}, author={Biek, R. and Mills, L. S. and Bury, R. B.}, year={2002}, pages={129–140} } @article{biek_funk_maxell_mills_2002, title={What is missing in amphibian decline research: Insights from ecological sensitivity analysis}, volume={16}, ISSN={["1523-1739"]}, DOI={10.1046/j.1523-1739.2002.00433.x}, abstractNote={Abstract: Inventory, monitoring, and experimental studies have been the primary approaches for documenting and understanding the problem of amphibian declines. However, little attention has been given to placing human‐caused perturbations affecting one or more life‐history stages in the context of the overall population dynamics of particular species. We used two types of ecological sensitivity analysis to determine which vital rates have the strongest influence on the population dynamics of western toads (   Bufo boreas), red‐legged frogs (  Rana aurora), and common frogs (  Rana temporaria), pond‐breeding amphibians that have declined in all or portions of their ranges. Our results suggest that post‐metamorphic vital rates and highly variable vital rates both have a strong influence on the population dynamics of these species and therefore deserve more research and management attention. Ecological sensitivity analysis should be more widely applied to the issue of amphibian declines in order to identify the most plausible mechanisms of decline and prioritize which life‐history stages should be the focus of research and management efforts. Future experimental studies of perturbations in one or more life‐history stage should attempt to link the magnitude of the perturbation measured with the overall population‐level consequences. Finally, current research, inventory, and monitoring efforts should be supplemented with demographic studies so that quantitative analyses can be applied to a wider range of species and life‐history groups.}, number={3}, journal={CONSERVATION BIOLOGY}, author={Biek, R and Funk, WC and Maxell, BA and Mills, LS}, year={2002}, month={Jun}, pages={728–734} } @article{mills_doak_wisdom_2001, title={Elasticity analysis for conservation decision making: Reply to Ehrlen et al.}, volume={15}, DOI={10.1111/j.1523-1739.2001.00300.x}, abstractNote={Conservation BiologyVolume 15, Issue 1 p. 281-283 Elasticity Analysis for Conservation Decision Making: Reply to Ehrlén et al. L. Scott Mills, L. Scott Mills Wildlife Biology Program, School of Forestry, University of Montana, Missoula, MT 59812, U.S.A., email smills@forestry.umt.eduSearch for more papers by this authorDaniel F. Doak, Daniel F. Doak Department of Biology, University of California, Santa Cruz, CA 95064, U.S.A.Search for more papers by this authorMichael J. Wisdom, Michael J. Wisdom U.S. Forest Service, Forestry and Range Sciences Laboratory, 1401 Gekeler Lane, La Grande, OR 97850, U.S.A.Search for more papers by this author L. Scott Mills, L. Scott Mills Wildlife Biology Program, School of Forestry, University of Montana, Missoula, MT 59812, U.S.A., email smills@forestry.umt.eduSearch for more papers by this authorDaniel F. Doak, Daniel F. Doak Department of Biology, University of California, Santa Cruz, CA 95064, U.S.A.Search for more papers by this authorMichael J. Wisdom, Michael J. Wisdom U.S. Forest Service, Forestry and Range Sciences Laboratory, 1401 Gekeler Lane, La Grande, OR 97850, U.S.A.Search for more papers by this author First published: 18 July 2008 https://doi.org/10.1111/j.1523-1739.2001.00300.xCitations: 11AboutPDF ToolsRequest permissionExport citationAdd to favoritesTrack citation ShareShare Give accessShare full text accessShare full-text accessPlease review our Terms and Conditions of Use and check box below to share full-text version of article.I have read and accept the Wiley Online Library Terms and Conditions of UseShareable LinkUse the link below to share a full-text version of this article with your friends and colleagues. Learn more.Copy URL Share a linkShare onFacebookTwitterLinkedInRedditWechat No abstract is available for this article.Citing Literature Volume15, Issue1February 2001Pages 281-283 RelatedInformation}, number={1}, journal={Conservation Biology}, author={Mills, L. S. and Doak, D. F. and Wisdom, M. J.}, year={2001}, pages={281–283} } @inproceedings{yale_mills_2000, title={Do highways fragment small mammal populations?}, booktitle={Proceedings of the International Conference on Ecology and Transportation}, author={Yale, R. and Mills, L. S.}, year={2000} } @article{mills_citta_lair_schwartz_tallmon_2000, title={Estimating animal abundance using noninvasive DNA sampling: Promise and pitfalls}, volume={10}, DOI={10.2307/2641002}, number={1}, journal={Ecological Applications}, author={Mills, L. S. and Citta, J. J. and Lair, K. P. and Schwartz, M. K. and Tallmon, D. A.}, year={2000}, pages={283–294} } @article{mills_pilgrim_schwartz_mckelvey_2000, title={Identifying lynx and other North American felids based on mtDNA analysis}, volume={1}, DOI={10.1023/a:1011574209558}, journal={Conservation Genetics}, author={Mills, L. S. and Pilgrim, K. L. and Schwartz, M. K. and McKelvey, K.}, year={2000}, pages={285–288} } @article{wisdom_mills_doak_2000, title={Life stage simulation analysis: Estimating vital-rate effects on population growth for conservation}, volume={81}, ISSN={["1939-9170"]}, DOI={10.2307/177365}, abstractNote={We developed a simulation method, known as life-stage simulation analysis (LSA) to measure potential effects of uncertainty and variation in vital rates on population growth (λ) for purposes of species conservation planning. Under LSA, we specify plausible or hypothesized levels of uncertainty, variation, and covariation in vital rates for a given population. We use these data under resampling simulations to establish random combinations of vital rates for a large number of matrix replicates and finally summarize results from the matrix replicates to estimate potential effects of each vital rate on λ in a probability-based context. Estimates of potential effects are based on a variety of summary statistics, such as frequency of replicates having the same vital rate of highest elasticity, difference in elasticity values calculated under simulated conditions vs. elasticities calculated using mean invariant vital rates, percentage of replicates having positive population growth, and variation in λ explained by variation in each vital rate. To illustrate, we applied LSA to vital rates for two vertebrates: desert tortoise (Gopherus agassizii) and Greater Prairie Chicken (Tympanuchus cupido). Results for the prairie chicken indicated that a single vital rate consistently had greatest effect on population growth. Results for desert tortoise, however, suggested that a variety of life stages could have strong effects on population growth. Additional simulations for the Greater Prairie Chicken under a hypothetical conservation plan also demonstrated that a variety of vital rates could be manipulated to achieve desired population growth. To improve the reliability of inference, we recommend that potential effects of vital rates on λ be evaluated using a probability-based approach like LSA. LSA is an important complement to other methods that evaluate vital-rate effects on λ, including classical elasticity analysis, retrospective methods of variance decomposition, and simulation of the effects of environmental stochasticity.}, number={3}, journal={ECOLOGY}, author={Wisdom, MJ and Mills, LS and Doak, DF}, year={2000}, month={Mar}, pages={628–641} } @inbook{groom_jensen_knight_gatewood_mills_boyd-heger_mills_soule_1999, title={Buffer zones: Benefits and dangers of compatible stewardship}, ISBN={9781559636971}, booktitle={Continental conservation : scientific foundations of regional reserve networks}, publisher={Washington, D.C. : Island Press}, author={Groom, M. and Jensen, D. B. and Knight, R. L. and Gatewood, S. and Mills, L. and Boyd-Heger, D. and Mills, L. S. and Soule, M. E.}, editor={M. E. Soule and Terborgh, J.Editors}, year={1999}, pages={191–198} } @inbook{dobson_ralls_foster_soule_simberloff_d._j. a._l. s._d._r._et al._1999, title={Connectivity: Maintaining flows in fragmented landscapes}, ISBN={9781559636988}, booktitle={Continental conservation : scientific foundations of regional reserve networks}, publisher={Washington, D.C. : Island Press}, author={Dobson, A. and Ralls, K. and Foster, M. and Soule, M. E. and Simberloff, D. Doak and D., Estes and J. A., Mills and L. S., Mattson and D., Dirzo and R., Arita and et al.}, editor={Soule?, Michael E. and Terborgh, JohnEditors}, year={1999} } @article{jules_frost_mills_tallmon_1999, title={Ecological consequences of forest fragmentation in the Klamath region}, volume={19}, number={4}, journal={Natural Areas Journal}, author={Jules, E. S. and Frost, E. J. and Mills, L. S. and Tallmon, D. A.}, year={1999}, pages={368–378} } @article{mills_doak_wisdom_1999, title={Reliability of conservation actions based on elasticity analysis of matrix models}, volume={13}, ISSN={["1523-1739"]}, DOI={10.1046/j.1523-1739.1999.98232.x}, abstractNote={Abstract: Matrix population models have entered the mainstream of conservation biology, with analysis of proportional sensitivities (elasticity analysis) of demographic rates becoming important components of conservation decision making. We identify areas where management applications using elasticity analysis potentially conflict with the mathematical basis of the technique, and we use a hypothetical example and three real data sets (Prairie Chicken [  Tympanuchus cupido], desert tortoise [Gopherus agassizii], and killer whale [Orcinus orca]) to evaluate the extent to which conservation recommendations based on elasticities might be misleading. First, changes in one demographic rate can change the qualitative ranking of the elasticity values calculated from a population matrix, a result that dampens enthusiasm for ranking conservation actions based solely on which rates have the highest elasticity values. Second, although elasticities often provide accurate predictions of future changes in population growth rate under management perturbations that are large or that affect more than one rate concurrently, concordance frequently fails when different rates vary by different amounts. In particular, when vital rates change to their high or low values observed in nature, predictions of future growth rate based on elasticities of a mean matrix can be misleading, even predicting population increase when the population growth rate actually declines following a perturbation. Elasticity measures will continue to be useful tools for applied ecologists, but they should be interpreted with considerable care. We suggest that studies using analytical elasticity analysis explicitly consider the range of variation possible for different rates and that simulation methods are a useful tool to this end.}, number={4}, journal={CONSERVATION BIOLOGY}, author={Mills, LS and Doak, DF and Wisdom, MJ}, year={1999}, month={Aug}, pages={815–829} } @inbook{mills_tallmon_1999, title={The role of genetics in understanding forest fragmentation}, ISBN={9789004113886}, booktitle={Forest fragmentation : wildlife and management implications}, publisher={Leiden : Brill}, author={Mills, L. S. and Tallmon, D.}, editor={James A. Rochelle, Leslie A. Lehmann and Wisniewski, JoeEditors}, year={1999} } @inproceedings{citta_mills_1999, title={What do demographic sensitivity analyses tell us about controlling Brown-headed Cowbirds?}, number={18}, booktitle={Research and management of the brown-headed cowbird in western landscapes}, author={Citta, J. J. and Mills, L. S.}, year={1999}, pages={121–134} } @article{soule_mills_1998, title={No need to Isolate Genetics}, volume={282}, journal={Science}, author={Soule, M. E. and Mills, L. S.}, year={1998}, pages={1658–1659} } @article{soule_mills_1998, title={Population genetics - No need to isolate genetics}, volume={282}, ISSN={["0036-8075"]}, DOI={10.1126/science.282.5394.1658}, abstractNote={Does the loss of genetic variability in small populations contribute to their extinction? SoulA© comments on this question in his Perspective, referring to new data from a long-term study of the declining population of the Illinois prairie chicken (Westemeier et al .) on p. [1695][1] of this issue. Results from this study indicate that inbreeding depression does in fact exacerbate the likelihood of extinction. [1]: http://www.sciencemag.org/cgi/content/short/282/5394/1695}, number={5394}, journal={SCIENCE}, author={Soule, ME and Mills, LS}, year={1998}, month={Nov}, pages={1658–1659} } @misc{mills_1997, title={Book review: Population management for survival and recovery: analytical methods and strategies in small population conservation}, volume={61}, DOI={10.2307/3802439}, journal={Journal of Wildlife Management}, author={Mills, L. S.}, year={1997}, pages={251–252} } @article{wisdom_mills_1997, title={Sensitivity analysis to guide population recovery: Prairie-chickens as an example}, volume={61}, ISSN={["1937-2817"]}, DOI={10.2307/3802585}, abstractNote={Calculation of elasticities in matrix population models is a formal type of sensitivity analysis that is used increasingly to guide recovery of declining populations. Results presumably allow recovery efforts to focus on the life stage most responsible for change in population growth, as indexed by the highest elasticity. Specifically, the highest elasticity denotes the vital rate whose proportionate change exerts the largest proportionate effect on the finite rate of increase (λ). We examined the utility of this analysis given uncertainty in parameter estimates and random variation in vital rates. We modeled these conditions to test the hypothesis that nest success and brood survival exert the greatest effect on population growth of greater prairie-chickens (Tympanuchus cupido pinnatus). We calculated elasticity associated with each age-specific vital rate contained in 1,000 randomly-generated replicates of a Leslie matrix model, and regressed λ on each randomly-varying rate. Age 0 survival (S o ) was associated with highest elasticity for 100% of the replicates and accounted for most of the variation in λ (r 2 = 0.95). Within S o , nest success and brood survival accounted for more variation in λ than other life stage combinations. These results demonstrate the utility of sensitivity analysis, but additional results point to its limitations. For example, the vital rate consistently associated with the second highest elasticity (S 1 ) accounted for minuscule variation in λ (r 2 = 0.0009), implying that rank of elasticities can fail to index the magnitude of a vital rate's effect on λ when vital rates vary simultaneously and disproportionately. To ensure that results are reliable, we recommend that sensitivity analysis be performed across the range of plausible vital rates, that simulations involve randomization of values within these ranges, and that elasticities be calculated in tandem with regression analysis to fully illuminate potential relations of vital rates with λ. A critical assumption is that variance of vital rates is estimated accurately.}, number={2}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={Wisdom, MJ and Mills, LS}, year={1997}, month={Apr}, pages={302–312} } @misc{power_tilman_estes_menge_bond_mills_daily_castilla_lubchenco_paine_1996, title={Challenges in the quest for keystones}, volume={46}, ISSN={["0006-3568"]}, DOI={10.2307/1312990}, abstractNote={Mary E. Power is a professor in the Department of Integrative Biology, University of California, Berkeley, CA 94720. David Tilman is a professor in the Department of Ecology, Evolution, and Behavior, University of Minnesota, St. Paul, MN 55108. James A. Estes is a wildlife biologist in the National Biological Service, Institute of Marine Science, University of California, Santa Cruz, CA 95064. Bruce A. Menge is a professor in the Department of Zoology, Oregon State University, Corvallis, OR 97331. William J. Bond is a professor doctor in the Department of Botany, University of Cape Town, Rondebosch 7700 South Africa. L. Scott Mills is an assistant professor in the Wildlife Biology Program, School of Forestry, University of Montana, Missoula, MT 59812. Gretchen Daily is Bing Interdisciplinary Research Scientist, Department of Biological Science, Stanford University, Stanford, CA 94305. Juan Carlos Castilla is a full professor and marine biology head in Facultad de Ciencias Biologicas, Pontificia Universidad Catolica de Chile, Casilla 114-D, Santiago, Chile. Jane Lubchenco is a distinguished professor in the Department of Zoology, Oregon State University, Corvallis, OR 97331. Robert T. Paine is a professor in the Department of Zoology, NJ-15, University of Washington, Seattle, WA 98195. ? 1996 American Institute of Biological Sciences. A keystone species is}, number={8}, journal={BIOSCIENCE}, author={Power, ME and Tilman, D and Estes, JA and Menge, BA and Bond, WJ and Mills, LS and Daily, G and Castilla, JC and Lubchenco, J and Paine, RT}, year={1996}, month={Sep}, pages={609–620} } @misc{mills_1996, title={Cheetah extinction: Genetics or extrinsic factors?}, volume={10}, ISSN={["0888-8892"]}, DOI={10.1046/j.1523-1739.1996.10020313-2.x}, abstractNote={Conservation BiologyVolume 10, Issue 2 p. 315-315 Cheetah Extinction: Genetics or Extrinsic Factors? L. Scott Mills, L. Scott Mills Wildlife Biology Program, School of Forestry, University of Montana, Missoula, MT 59812, U.S.A., email [email protected]Search for more papers by this author L. Scott Mills, L. Scott Mills Wildlife Biology Program, School of Forestry, University of Montana, Missoula, MT 59812, U.S.A., email [email protected]Search for more papers by this author First published: April 1996 https://doi.org/10.1046/j.1523-1739.1996.10020313-2.xCitations: 5AboutPDF ToolsRequest permissionExport citationAdd to favoritesTrack citation ShareShare Give accessShare full text accessShare full-text accessPlease review our Terms and Conditions of Use and check box below to share full-text version of article.I have read and accept the Wiley Online Library Terms and Conditions of UseShareable LinkUse the link below to share a full-text version of this article with your friends and colleagues. Learn more.Copy URL Citing Literature Volume10, Issue2April 1996Pages 315-315 RelatedInformation}, number={2}, journal={CONSERVATION BIOLOGY}, author={Mills, LS}, year={1996}, month={Apr}, pages={315–315} } @article{mills_hayes_baldwin_wisdom_citta_mattson_murphy_1996, title={Factors leading to different viability predictions for a grizzly bear data set}, volume={10}, ISSN={["0888-8892"]}, DOI={10.1046/j.1523-1739.1996.10030863.x}, abstractNote={Population viability analysis programs are being used increasingly in research and management applications, but there has not been a systematic study of the congruence of different program predictions based on a single data set. We performed such an analysis using four population viability analysis computer programs: GAPPS, INMAT, RAMAS/AGE, and VORTEX. The standardized demographic rates used in all programs were generalized from hypothetical increasing and decreasing grizzly bear (Ursus arctos horribilis) populations. Idiosyncracies of input format for each program led to minor differences in intrinsic growth rates that translated into striking differences in estimates of extinction rates and expected population size. In contrast, the addition of demographic stochasticity, environmental stochasticity, and inbreeding costs caused only a small divergence in viability predictions. But, the addition of density dependence caused large deviations between the programs despite our best attempts to use the same density-dependent functions. Population viability programs differ in how density dependence is incorporated, and the necessary functions are difficult to parameterize accurately. Thus, we recommend that unless data clearly suggest a particular density-dependent model, predictions based on population viability analysis should include at least one scenario without density dependence. Further, we describe output metrics that may differ between programs; development of future software could benefit from standardized input and output formats across different programs. Los programas de analisis de viabilidad de poblaciones estan siendo usados coda vez mas en investigacion y aplicaciones de manejo. Sin embargo, no existe un estudio sistematico de la congruencia de predicciones de diferentes programas basadas en un solo conjunto de datos. Realizamos un analisis usando cuatro programas de computadora para analisis de viabilidad poblacional (GAPPS, INMAT, RAMAS/AGE y VORTEX). Las tasas demograficas estandarizadas usadas en todos los programas fueron generalizadas para poblaciones de osos que incrementan y disminuyen hipoteticamente. La idiosincrasia del formato de entrada para cada programa condujo a diferencias menores en la tasa de crecimiento intrinseco, mismas que se tradujeron como diferencias notables en las estimaciones de tasas de extincion y tamano poblacional esperado. En contraste, la adicion de aleatoriedad demografica, aleatoriedad ambiental y consanguinidad causaron unicamente pequenas divergencias en tas predicciones de la viabilidad de esta especie. Sin embargo, la adicion de dependencia de la densidad ocasiono grandes desviaciones entre los programas a pesar de nuestros mejores esfuerzos por usar las mismas funciones densodependendientes. Los programas de viabilidad de poblaciones difieren en como se incorpora la dependencia a la densidad y las funciones necesarias son dificiles de parametrizar con precision. Espar ello que recomendamos que, a menas que los datos sugieran claramente un modela particular de dependencia de la densidad, las predicciones de analisis de viabilidad de poblaciones (PVA) deberan incluir por lo menos un escenario sin dependencia de la densidad. Mas aun, se describen datos de salida que pueden diferir entre programas. El desarrollo de paqueteria de computo podria beneficiarce si se estandarizaran los formatas de registre y resultados arrojados por diferentes programas.}, number={3}, journal={CONSERVATION BIOLOGY}, author={Mills, LS and Hayes, SG and Baldwin, C and Wisdom, MJ and Citta, J and Mattson, DJ and Murphy, K}, year={1996}, month={Jun}, pages={863–873} } @inbook{mills_1996, title={Fragmentation of a natural area: Dynamics of isolation for small mammals on forest remnants}, ISBN={9780865424968}, booktitle={National parks and protected areas : their role in environmental protection}, publisher={Cambridge, Mass. : Blackwell Science}, author={Mills, L. S.}, year={1996}, pages={199–219} } @article{morrison_mills_kuenzi_1996, title={Study and management of an isolated, rare population: The Fresno kangaroo rat}, volume={24}, number={4}, journal={Wildlife Society Bulletin}, author={Morrison, M. L. and Mills, L. S. and Kuenzi, A. J.}, year={1996}, pages={602–606} } @article{mills_allendorf_1996, title={The one-migrant-per-generation rule in conservation and management}, volume={10}, ISSN={["1523-1739"]}, DOI={10.1046/j.1523-1739.1996.10061509.x}, abstractNote={In the face of continuing habitat fragmentation and isolation, the optimal level of connectivity be- tween populations has become a central issue in conservation biology. A common rule of thumb holds that one migrant per generation into a subpopulation is sufficient to minimize the loss of polymorphism and het- erozygosity within subpopulations while allowing for divergence in allele frequencies among subpopulations. The one-migrant-per-generation rule is based on numerous simplifying assumptions that may not hold in nat- ural populations. We examine the conceptual and theoretical basis of the rule and consider both genetic and nongenetic factors that influence the desired level qf connectivity among subpopulations. We conclude that one migrant per generation is a desirable minimum, but it may be inadequaite for many natural popula- tions. We suggest that a minimum of I and a maximum of 10( migrants per generation would be an appro- priate general rule of thumb for genetic purposes, bearing in mind that fctctors other than genetics maly fur- ther influence the ideal level of connectivity.}, number={6}, journal={CONSERVATION BIOLOGY}, author={Mills, LS and Allendorf, FW}, year={1996}, month={Dec}, pages={1509–1518} } @article{mills_1995, title={EDGE EFFECTS AND ISOLATION - RED-BACKED VOLES ON FOREST REMNANTS}, volume={9}, ISSN={["0888-8892"]}, DOI={10.1046/j.1523-1739.1995.9020395.x}, abstractNote={Negative effects of habitat edge have been advanced as an important proximate cause of extinction, and a growing literature calls attention to the matrix surrounding habitat remnants as a critical factor determining population persistence. I examined spatial distribution of California red-backed voles (Clethrionomys californicus) on 13 forest remnants and five control sites in southwestern Oregon. The species was virtually isolated on remnants, making little use of the regenerating clearcuts surrounding the remnants. The effects of the clearcut also impinged on the remnants as edge effects: six times more voles were captured per trap in the interior of remnants than on the edge. Consequently, the density of voles per unit area on remnants increased with remnant size, despite the potential buildup of population density in small isolates due to limited emigration. I explored potential mechanisms of the negative edge effect on voles and found that the biomass of coarse woody debris, per se, did not explain the vole distribution because both number and volume of logs increased from the interior to the edge of remnants. However, the distribution of the vole's primary food item, hypogeous sporocarps of mycorrhizal fungi, did correspond to the vole edge effect}, number={2}, journal={CONSERVATION BIOLOGY}, author={MILLS, LS}, year={1995}, month={Apr}, pages={395–402} } @inbook{mills_1995, title={Keystone species}, ISBN={9780122267307}, booktitle={Encyclopedia of environmental biology}, publisher={San Diego : Academic Press}, author={Mills, L. S.}, year={1995}, pages={381–387} } @article{scott_tear_mills_1995, title={SOCIOECONOMICS AND THE RECOVERY OF ENDANGERED SPECIES - BIOLOGICAL ASSESSMENT IN A POLITICAL WORLD}, volume={9}, ISSN={["0888-8892"]}, DOI={10.1046/j.1523-1739.1995.09010214.x}, abstractNote={Murphy et al. (1994) recently articulated 12 reasons for a strong and effective Endangered Species Act (ESA). At the same time, they pointed to the threats facing reauthorization of the ESA. These conflicts between conservation mandates and the political climate bring us to the sticky question: What role should politics play in endangered species management? The ESA states that the determination of a species' status as threatened or endangered is to be made "solely on basis of the best scientific and commercial data available .. . " (ESA ? 4(b)( 1 )(A), emphasis added). In addition, recovery plans are to provide "objective, measurable criteria" by which a species could be delisted (ESA ? 4(f)(1)(B)(ii)). However, the strict emphasis on biological criteria to establish recovery goals can result in goals that are not necessarily achievable, practical, politically acceptable, or even expedient. While it is inevitable that politics, economics, psychology, and sociology also play a role in establishing and implementing recovery plan goals for endangered species, it is not clear how these "nonbiological" concerns should be incorporated into the biological decision making process. A conflict of opinion has emerged from this uncertainty, whereby some argue for incorporating socioeconomic and political realities into recovery goals, while others urge species recovery based strictly on biological criteria. In addition, lack of distinction between "political" and biological goals has been suggested as a reason for setting low recovery goals (Tear et al. 1993). Similar debates have surfaced with such high-profile species as the Northern Spotted Owl (Thomas & Verner 1992; Yaffee 1994), the Red-cockaded Woodpecker (McFarlane 1992), and the Dusky Seaside Sparrow (Walters 1992). This issue over how to incorporate biological and nonbiological factors may also lie at the center of the current debate over whether to accept or reject the 1993 revision of the grizzly bear recovery plan. All involved in the conservation of endangered species would agree on the most basic of points: recover the species rather than compromise its chances for survival. Common ground must be sought between the opposing points of view that pit biological estimates of viability against the constraints of social, political, and economic realities. Toward that end improvements have been made in the recovery process evident in the 1988 "recovery plan amendments" (Fitzgerald 1989) and policy guidelines (U.S. Fish and Wildlife Service 1990a). The following suggestions are based on the premise that the separate but relative influences of biological and socioeconomic factors should be explicitly stated when a species' probable path to recovery is estimated. Recovery plans are supposed to provide estimates of the time and expense of achieving recovery (USFWS 1990a). We recognize that forecasting the future of any species is a difficult task. However, some guidelines for making informed predictions have emerged from population viability analysis (PVA) which may help improve the process. We start with the suggestion that recovery goals be considered for both the shortand long-term. Establishing a specific time frame for each of these levels will vary among and within taxonomic groups. For example, large mammal recovery efforts might target 1020 years for the short-term and 100 years or so for long-term goals, while much shorter intervals might be more applicable for invertebrates. Second, the recovery team will need to agree on some acceptable probability of persistence for each time period in order to evaluate and compare recovery options. In addition to the traditional extinction threshold of zero individuals, we recommend that other threshPaper submitted May 21, 1994; revised manuscript accepted August 26, 1994.}, number={1}, journal={CONSERVATION BIOLOGY}, author={SCOTT, JM and TEAR, TH and MILLS, LS}, year={1995}, month={Feb}, pages={214–216} } @article{power_mills_1995, title={THE KEYSTONE COPS MEET IN HILO}, volume={10}, ISSN={["0169-5347"]}, DOI={10.1016/s0169-5347(00)89047-3}, abstractNote={Despite criticisms, the umbrella species concept remains a fundamental conservation tool for protecting biodiversity in the face of global change, yet it is rarely tested. Food web theory provides a tool to test both umbrella-species' suitability and their ecological function, which we investigate in a large-mammal food web. Using data from 698 camera trap locations in the Canadian Rockies, we develop hierarchical occupancy models to predict the co-occurrence of 16 large mammal species. We draw upon previous diet studies in the Canadian Rockies to describe the meta food-web (meta-web) for these species. Next, we filtered the meta-web using predicted occupancy to estimate realized food webs at each camera location. We tested the umbrella species concept using predicted occupancy across all 698 camera sites. We then tested for carnivore effects using realized food webs on 5 food-web properties: species richness, links, connectance, nestedness and modularity using generalized linear models while accounting for landscape covariates known to affect food web dynamics. Our multispecies occupancy models reflected factors previously demonstrated to affect large mammal occurrence. Our results also demonstrated that grizzly bear (Ursus horribilis), a generalist carnivore, was the best umbrella carivore species, and explained species richness the best. When considering food web properties, however, wolves (Canis lupus) and cougars (Felis concolor) served as better umbrellas that also captured food web properties such as connectance, links and nestedness that better reflect ecological interactions. Our results support the role of large carnivores as umbrella and ecologically interactive species in conservation planning.}, number={5}, journal={TRENDS IN ECOLOGY & EVOLUTION}, author={POWER, ME and MILLS, LS}, year={1995}, month={May}, pages={182–184} } @article{zager_mills_wakkinen_d._1995, title={Woodland caribou: A conservation dilemma}, volume={12}, number={10/11}, journal={Endangered Species Update}, author={Zager, P. and Mills, L. S. and Wakkinen, W. and D., Tallmon}, year={1995} } @article{doak_mills_1994, title={A USEFUL ROLE FOR THEORY IN CONSERVATION}, volume={75}, ISSN={["0012-9658"]}, DOI={10.2307/1941720}, abstractNote={Conservation biology has frequently been designated a "crisis discipline" (Soule 1985, Maguire 1991). This moniker refers both to the urgency of the issues that conservation biologists seek to address and to the fact that during crises there is rarely time to assess a situation carefully before one must act. Indeed, the paucity of data available for most endangered species, communities, and ecosystems often forces biologists and policy makers to make management decisions without any quantitative information. Further, this lack of data may be insoluble: many management decisions have to be made hurriedly, and rarity itself precludes the quick or easy collection of data on threatened species or communities.}, number={3}, journal={ECOLOGY}, author={DOAK, DF and MILLS, LS}, year={1994}, month={Apr}, pages={615–626} } @misc{mills_1994, title={Book review: Principles of conservation biology}, volume={68}, DOI={10.1016/s0277-9536(99)00265-8}, journal={Northwest Science}, author={Mills, L. S.}, year={1994}, pages={303–304} } @article{mills_smouse_1994, title={DEMOGRAPHIC CONSEQUENCES OF INBREEDING IN REMNANT POPULATIONS}, volume={144}, ISSN={["1537-5323"]}, DOI={10.1086/285684}, abstractNote={Although traditional population fragmentation theory and management has been strongly oriented toward concerns arising from inbreeding depression, recent papers suggest that small populations will be eliminated by demographic and/or environmental events before inbreeding becomes a problem. We explore the interaction between these factors by developing a stochastic, discrete time Leslie model that incorporates inbreeding depression. We model small population dynamics with three realistic demographic schedules: low growth rate "ungulates," medium growth rate "felids," and high growth rate "rodents," examining the impact of survival and fertility depression commensurate with inbreeding effects reported in the literature. Focusing on the first few generations after habitat fragmentation and isolation, we find that (a) high growth rate populations are affected only by strong inbreeding depression, but low growth rate populations are extremely vulnerable to even minor inbreeding depression; (b) vulnerability to extinction is affected more by survival depression than by fecundity depression; and (c) reductions in the sex ratio exacerbate inbreeding accumulation and hence extinction rate. Counter to the current fashion, which downplays the importance of inbreeding in stochastic environments, we conclude that, while inbreeding depression is not necessarily the primary cause of extinction, it can be critical.}, number={3}, journal={AMERICAN NATURALIST}, author={MILLS, LS and SMOUSE, PE}, year={1994}, month={Sep}, pages={412–431} } @article{clarkson_mills_1994, title={Hypogeous sporocarps in forest remnants and clearcuts in Southwest Oregon}, volume={68}, number={4}, journal={Northwest Science}, author={Clarkson, D. A. and Mills, L. S.}, year={1994}, pages={259–265} } @article{tallmon_mills_1994, title={USE OF LOGS WITHIN HOME RANGES OF CALIFORNIA RED-BACKED VOLES ON A REMNANT OF FOREST}, volume={75}, ISSN={["0022-2372"]}, DOI={10.2307/1382240}, abstractNote={We used radiotelemetry to investigate patterns of space use of the California red-backed vole ( Clethrionomys californicus ) on a remnant of forest in southwestern Oregon. We radiotracked four voles and mapped the locations of logs of varying decay classes within the home range of each vole. Of the collective locations of voles, 98% coincided with downed logs even though logs covered only 7% of the areas of estimated home ranges. Furthermore, voles used logs in later stages of decay significantly more often than logs in earlier stages of decay. This high use of decayed logs suggests that decayed logs are a critical component of suitable habitat for voles.}, number={1}, journal={JOURNAL OF MAMMALOGY}, author={TALLMON, D and MILLS, LS}, year={1994}, month={Feb}, pages={97–101} } @article{mills_fredrickson_moorhead_1993, title={CHARACTERISTICS OF OLD-GROWTH FORESTS ASSOCIATED WITH NORTHERN SPOTTED OWLS IN OLYMPIC-NATIONAL-PARK}, volume={57}, ISSN={["0022-541X"]}, DOI={10.2307/3809428}, abstractNote={The relationship between northern spotted owls (Strix occidentalis caurina) and old-growth fores is well established, but there are few studies indicating which particular characteristics of these forests are most important to spotted owls. Thus, we compared characteristics of forests at owl response and non-response sites across Olympic National Park to identify the specific features which best predict daytime presence of northern spotted owls in relatively pristine forest. At 32 owl response sites and 230 non-response sites we measured 9 structural and species composition variables thought to be important to spotted owls. Based on these variables, we used stepwise logistic regression to develop models predicting the presence or absence of spotted owls}, number={2}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={MILLS, LS and FREDRICKSON, RJ and MOORHEAD, BB}, year={1993}, month={Apr}, pages={315–321} } @article{mills_soule_doak_1993, title={THE KEYSTONE-SPECIES CONCEPT IN ECOLOGY AND CONSERVATION}, volume={43}, ISSN={["0006-3568"]}, DOI={10.2307/1312122}, abstractNote={1Will the extinction of a single species in a community cause the loss of many others? Can we identify a set of species that are so important in determining the ecological functioning of a community that they warrant special conservation efforts? The answer to these questions hinges on the existence of a limited number of species whose loss would precipitate many further extinctions; these species have often been labeled keystone species. The term keystone species has enjoyed an enduring popularity in the ecological literature since its introduction by Robert T. Paine in 1969: Paine (1969) was cited in more than 92 publications from 1970 to 1989; an earlier paper (Paine 1966), which introduced the phenomenon of keystone species in intertidal systems but did not use the term, was cited more than 850 times during the same period. As used by Paine and other ecologists, there are two hallmarks of keystone species. First, their presence is crucial in maintaining the organization and diversity of their ecological communities. Second, it is implicit that these species are exceptional, relative to the rest of the community, in their importance.}, number={4}, journal={BIOSCIENCE}, author={MILLS, LS and SOULE, ME and DOAK, DF}, year={1993}, month={Apr}, pages={219–224} } @inbook{soule_mills_1992, title={Conservation genetics and conservation biology: a troubled marriage}, ISBN={9788200215080}, booktitle={Conservation of biodiversity for sustainable development}, publisher={Oslo, Norway : Scandinavian University Press}, author={Soule, M.E. and Mills, L. S.}, editor={O. T. Sandlund, K. Hindar and A. H. D. BrownEditors}, year={1992}, pages={55–69} } @article{mills_knowlton_1991, title={COYOTE SPACE USE IN RELATION TO PREY ABUNDANCE}, volume={69}, ISSN={["0008-4301"]}, DOI={10.1139/z91-212}, abstractNote={Food abundance is an important factor determining space use in many species, but its effect on carnivore home range and territory size has rarely been investigated. We explored the relationship between food abundance for the coyote (Canis latrans) and space use in two study areas in the northern Great Basin, where the primary prey, the black-tailed jackrabbit (Lepus californicus), fluctuates dramatically in abundance. At one site, home ranges and territories were significantly larger during a time of prey-scarcity than when prey was abundant. Coyotes on the second site had similar-size home ranges and territories at low and high prey abundance, but a higher proportion and probably a higher number of individuals were transients during the prey-scarcity period. We propose mortality rates of coyotes as an important factor mediating adjustments in space use to food abundance, and suggest two mechanisms by which mortality might interact with food abundance. Higher mortality rates may simply permit more rapid a...}, number={6}, journal={CANADIAN JOURNAL OF ZOOLOGY-REVUE CANADIENNE DE ZOOLOGIE}, author={MILLS, LS and KNOWLTON, FF}, year={1991}, month={Jun}, pages={1516–1521} } @article{mills_knowlton_1989, title={OBSERVER PERFORMANCE IN KNOWN AND BLIND RADIO-TELEMETRY ACCURACY TESTS}, volume={53}, ISSN={["0022-541X"]}, DOI={10.2307/3801134}, number={2}, journal={JOURNAL OF WILDLIFE MANAGEMENT}, author={MILLS, LS and KNOWLTON, FF}, year={1989}, month={Apr}, pages={340–342} }