Next Article in Journal
Environmental Fragility Zoning Using GIS and AHP Modeling: Perspectives for the Conservation of Natural Ecosystems in Brazil
Previous Article in Journal
Agroforestry, Livelihood and Biodiversity Nexus: The Case of Madhupur Tract, Bangladesh
 
 
Font Type:
Arial Georgia Verdana
Font Size:
Aa Aa Aa
Line Spacing:
Column Width:
Background:
Article

Loss of Migratory Traditions Makes the Endangered Patagonian Huemul Deer a Year-Round Refugee in Its Summer Habitat

by
Werner T. Flueck
1,2,3,*,
Jo Anne M. Smith-Flueck
4,5,7,
Miguel E. Escobar
5,
Melina Zuliani
1,6,
Beat Fuchs
7,
Valerius Geist
8,†,
James R. Heffelfinger
9,
Patricia Black-Decima
10,
Zygmunt Gizejewski
11,
Fernando Vidal
12,13,
Javier Barrio
14,
Silvina M. Molinuevo
15,
Adrian J. Monjeau
1,7,
Stefan Hoby
16 and
Jaime E. Jiménez
17
1
National Council of Scientific and Technological Research (CONICET), Buenos Aires 1425, Argentina
2
Swiss Tropical and Public Health Institute, University of Basel, 4001 Basel, Switzerland
3
Argentine National Parks, Bariloche 8400, Argentina
4
Laboratorio de Teriogenología ‘Dr Héctor H. Morello’, Facultad de Ciencias Agrarias, Universidad Nacional del Comahue, Cinco Saltos 8303, Argentina
5
Fundación Shoonem, Parque Protegido Shoonem, Alto Río Senguer 9033, Argentina
6
Departamento de Análisis de Sistemas Complejos, Fundación Bariloche, Bariloche 8400, Argentina
7
DeerLab, Bariloche 8400, Argentina
8
Faculty of Environmental Design, University of Calgary, Calgary, AB TIN 1N4, Canada
9
Arizona Game and Fish Department, Phoenix, AZ 85006, USA
10
Facultad de Ciencias Naturales e Instituto Miguel Lillo, Universidad Nacional de Tucumán, San Miguel de Tucuman 4000, Argentina
11
Institute of Animal Reproduction and Food Research, Polish Academy of Sciences, Pl-10-747 Olsztyn, Poland
12
Departamento de Ciencias Basicas, Facultad de Ciencias, Univerdidad Santo Tomas, Villarrica 8370003, Chile
13
Fauna Andina, Centro de Conservacion y Manejo de Vida Silvestre, Villarrica 102, Chile
14
Centro de Ornitología y Biodiversidad, Lima 33, Peru
15
Laboratorio de Investigacion en Osteopatias y Metabolismo Mineral, Departamento de Ciencias Biologicas, Facultad de Ciencias Exactas, Universidad Nacional de La Plata, La Plata 1900, Argentina
16
Berne Animal Park, 3006 Bern, Switzerland
17
Department of Biological Sciences, Advanced Environmental Research Institute, University of North Texas, Denton, TX 76203, USA
*
Author to whom correspondence should be addressed.
Deceased.
Conservation 2022, 2(2), 322-348; https://doi.org/10.3390/conservation2020023
Submission received: 8 March 2022 / Revised: 13 April 2022 / Accepted: 29 April 2022 / Published: 31 May 2022

Abstract

:
The huemul (Hippocamelus bisulcus) is endangered, with 1500 deer split into >100 subpopulations along 2000 km of the Andes. Currently occupied areas are claimed-erroneously-to be critical prime habitats. We analyzed historical spatiotemporal behavior since current patterns represent only a fraction of pre-Columbian ones. Given the limited knowledge, the first group (n = 6) in Argentina was radio-marked to examine spatial behavior. Historically, huemul resided year-round in winter ranges, while some migrated seasonally, some using grasslands >200 km east of their current presence, reaching the Atlantic. Moreover, huemul anatomy is adapted to open unforested habitats, also corroborated by spotless fawns. Extreme naivety towards humans resulted in early extirpation on many winter ranges—preferentially occupied by humans, resulting in refugee huemul on surrounding mountain summer ranges. Radio-marked huemul remained in small ranges with minimal altitudinal movements, as known from other subpopulations. However, these resident areas documented here are typical summer ranges as evidenced by past migrations, and current usage for livestock. The huemul is the only cervid known to use mountain summer ranges year-round in reaction to anthropogenic activities. Losing migratory traditions is a major threat, and may explain their presently prevalent skeletal diseases, reduced longevity, and lacking recolonizations for most remaining huemul subpopulations.

1. Introduction

The Patagonian huemul (Hippocamelus bisulcus) is considered an endangered Odocoiline deer by the International Union for the Conservation of Nature [1], with Argentina having only an estimated 350–500 individuals left and split into 60 or more groups, and Chile having around 1000 remaining that are split into approximately 40 groups. Moreover, these groups are fragmented along some 2000 km of Andean mountains [2] and represent a numerical reduction of over 99% of the original population size [3]. Their social organization including social and sexual segregation, is very plastic as in other cervids, and likely highly influenced by population density [4]. The huemul have been negatively affected mainly by past overhunting, but also loss and fragmentation of habitat, malnutrition, diseases, dogs, and possibly by the introduction of alien wild and domestic ungulate species [1]. Unrestricted killing in the past to acquire valued products was one of the main factors that resulted in widespread population declines and the endangered status of this species (reviewed in Supplementary File S1). Extremely naive and tolerant of human presence [5], huemul can be easily killed at a close distance by the simple use of rocks, clubs or knives (Figure 1). This unique docile behavior towards humans has resulted in their local extirpation, especially in those areas used by indigenous people and early colonists (Supplementary File S1). However, past over-harvesting has not only resulted in local extirpation, but we hypothesize that it also eliminated their migratory traditions.
Although extant groups of huemul persist along the seasonal Andean Mountain range, none have been shown to still exhibit migratory behavior [6]. Seasonal migration allows a species to capitalize on spatiotemporal variation in resources and being a plastic behavior, also results in colonization of new areas. Among cervids, year-round resident populations on winter ranges in seasonal mountain environments are the norm, as well as the eventual development of a portion of that population that migrates altitudinally [7,8]. Migratory cultural behavior is transmitted vertically from mother to young, whereas dispersal is innate and by random diffusion that is predetermined genetically, but may also occur secondarily in response to environmental conditions [9]: thus, preserving cultural traits like migrating are considered important, especially for endangered species [10,11,12,13] (Supplementary File S2).
Most extant huemul subpopulations thus tend to exist today in remote mountain areas that have been unattractive for human settlement because of climatic conditions, topography, remoteness or having otherwise little value for agriculture or forestry [14,15]. Modern mainstream interpretations commonly consider areas of extant subpopulations as representing prime habitat in which the huemul evolved [2,16]. There is a common misconception that these areas represent their natural ecological niche, rather than recognizing that the huemul may have been isolated by anthropogenic activities to persist only in such suboptimal habitat that was less accessible to humans, and effectively has become a refugee species by being forced to contract its range and lose its migratory behavior [17,18,19,20]. The huemul continue to be described as “mountain deer adapted to Andes mountains”, “found only in the Andes”, “high-altitude species”, “mountain deer with unmistakable mountaineer anatomy”, “at 2000–3000 m asl of Andes”, “commonly in high-altitude regions that are also rocky and steep”, and “a forest deer” [6,21,22,23,24]. This unfounded assumption has led to the characterization of the huemul as predominately a browser and a non-migratory species. This has influenced the course of huemul conservation and the approaches taken over the past four decades [25]. Historical evidence of past huemul distribution is commonly depreciated by calling these “anecdotal accounts”, thus disregarding all such data, including those from reputable scientific explorers of the Patagonian region during the last two centuries (reviewed in Supplementary File S1).
Securing adequate sample sizes of field data is difficult with rare and endangered species that live in remote refuge areas as occurs with huemul [6]. The current pattern of habitat use by many isolated huemul subpopulations, including the few remaining huemul reported here from Shoonem Park, and the revealed differences with the documented historical patterns, highlights the central urgency to achieve a valid diagnosis of the causes and consequences of becoming a year-round refugee in areas which qualify as seasonal summer ranges [20].
For the present study, we (1) provide a summary of an extensive literature review of historical records of this species that we conducted in order to examine the evidence of huemul occurrences in the treeless landscape of Patagonia and their seasonal movement patterns; then (2), we provide new information on seasonal habitat use by the first group of huemul ever radio-collared in Argentina as a means to then compare current patterns with historical ones; (3) we evaluate the process of migration and occupation of new habitat areas by wild cervids to put into perspective our findings regarding the behavioral change between past and current movement patterns and habitat use by huemul; and finally, (4) we discuss the implications of huemul having lost migratory traditions and the consequences of being forced to live year-round in a refugia which formerly represented only a seasonal summer habitat, and the secondary problems that have arisen as a result.

2. Materials and Methods

2.1. Systematic Literature Review

To better understand the flexibility and phenotypic plasticity of huemul through the lens of their overall usage of habitats in Argentina, we gleaned the literature on historical habitat use by this species, and then evaluated processes involved in the migration, dispersal, and occupation of new habitat areas among other wild cervids. A comprehensive review (2021–2022) was based on the broad literature access provided by Swisscovery (https://slsp.ch, accessed on 4 April 2022), including the ISI Web-of-Knowledge and their 17 external databases, by searching for huemul and other related deer species to assemble actual and historical data on huemul occurrences, and compare this to distribution patterns of other cervids. The systematic search about huemul included Hippocamelus bisulcus and its synonymy [26,27], that is the various historic taxonomic terms used for this species, like Equus bisulcus (Molina, 1788), Equo bisulco (Leuckart, 1816), Cervequus andicus (Lesson, 1842), Camelus equinus (Treviranus, 1803), Lama bisulca (Fisher, 1829), Auchenia huemul (Smith, 1827), Auchenia huamel (Hamilton, 1842), Cervus andicus (Lesson, 1842), Cervus antisensis (Burmeister, 1879), Cervus chilensis (Gay and Gervais, 1846), Cervus leucotis (Giebel, 1855), Capreolus leucotis (Gray, 1849), Capreolus huamel (Gray, 1850), Furcifer huamel (Gray, 1850), Furcifer chilensis (Sclater, 1883), Furcifer andicus (Lesson, 1850), Furcifer antisensis (Wagner, 1855), Xenelaphus leucotis (Gray, 1872), Xenelaphus bisulcus (Prichard, 1902), Huamela leucotis (Gray, 1872), Creagroceros chilensis (Fitzinger, 1873), Cariacus chilensis (Brooke, 1879), Mazama bisulca (Lydekker, 1898), Odocoileus bisulcus (Trouessart, 1898), and Hippocamelus dubius (Leuckart, 1816). Additional older literature cited in historical accounts was accessed by visiting key libraries containing such old original documents. Given the absence of other cervids in Patagonia, except the extremely small pudu (Pudu puda), and clear physical differences to guanaco (Lama guanicoe), the past documentation of a cervid identified with any of the taxonomic synonyms for huemul were taken as valid data.
Moreover, assessing the validity of a given data point for representing evidence was based on the physical description of the observed animal, their drawings or their photographs, which basically prevented any biased data point. The key sources about historical habitat use-including spatiotemporal, are reviewed in Table 1, Supplementary File S1, and displayed in Section 4.

2.2. Radiotelemetry

We studied the huemul in the Protected Park Shoonem (44°51′ S, 71°48′ W; 167 km2 with elevations ranging from 850–2060 m asl), located on the eastern slopes of the southern Andes. The studied watershed contains Fontana and La Plata lakes, which are surrounded by tall mountains (Figure 2). Within the sub-Antarctic zone, the site containing huemul around lake La Plata is covered by old-growth, dense forests predominately of deciduous lenga beech trees (Nothofagus pumilio), which occur from the lake level (930 m) up to about 1300 m, with the uppermost lenga forming chaparral [28]. The seasons are defined as winter from June to August, and summer from December to February. The mean winter temperature is −3 °C with winter precipitation between 300–400 mm, principally as snow (Figure 2), while the total annual mean precipitation is 2000 mm [29]. Most Andean environments are characterized by harsh climatic conditions, extensive deep snow cover in winter, and contrasting altitudinal levels such that it results in guanaco migrating toward low altitudes when the snow cover is too deep to survive [30,31].
During the winter of 2017, adult huemul (3 females, 3 males) were immobilized by darting (permit Disp. Nr. 22/2017-DFyFS-MP, Province of Chubut). To minimize risks, a DanInject JM SP25 rifle allowing continuous pressure adjustment, and a mounted scope containing a laser range finder, were used to dart the animals. Medetomidine and ketamine, reversed by atipamezole are considered very safe for cervids [32] induction and reversal are fast (2 min or less), there is generally an ample tolerance for various concentrations, and highly concentrated drugs allowed the use of small darts, thus minimizing trauma. Moreover, changes in pulse and respiratory rates are minimal [33]. The time of induction was noted, the animal was kept in lateral recumbency, eyes were covered, the pulse, respiratory rates, and temperature were monitored, the general health condition was examined, and after blood collection and placing the radio collar, the reversal was applied.
Individuals were fitted with VHF radio transmitters (Sirtrack Ltd., Havelock North, NZ, USA; Telonics Inc., Mesa, AZ, USA) with several capable of transmitting signals for up to 10 years, a time window exceeding the average life span of deer in this area [34]. During the winter of 2021, the VHF radio collar was replaced with a satellite unit (Lotek, Newmarket, ON, Canada) on one male. We monitored the radio-collared individuals regularly to determine if they were alive or dead, and then located them using the telemetry equipment to allow us to make observations regarding their physical state, group composition and other biological data, and to record their general locations, or their precise GPS locations based on visual observations or triangulations [35]. The altitude above sea level (asl) of each location was determined via Google Earth. Given the behavioral responses of the animals and the reduced frequency of visual encounters, the method was considered to be acceptable regarding animal welfare. Confirmations via radio signals as to whether an individual was dead or alive were determined repeatedly (2 times per week usually), and covered every month of the study period (August 2017 to April 2022), and were accomplished more frequently than location determinations. Non-statistical techniques for range analysis were used to calculate home range sizes based on the determined locations, using a minimum concave polygon (MCP) approach. This choice was made in order to specifically exclude the lake surfaces, and thus to ensure that such areas not used by collared individuals were not included in home range calculations [35,36,37,38,39]. However, the present analysis emphasized only the documented maximal space use, independent of the frequency of usage as an indication of probability density surface [40]. Although a small number of precise locations may result in a reduced estimate of the home range size, the sampling over a large time period, as in this study, compensates by repeatedly covering all seasons [41]. Moreover, even crudely estimated home range sizes have led to insights into animal behavior and ecology [41], and this information is fundamentally important for managing and maintaining viable populations [6]. The home range sizes determined here may not allow comparisons to other studies, but among the present cases. The coverage of precise and general locations during all seasons of several years was considered sufficient to evaluate potential migratory movements. These were defined as follows: resident, the distance between centroids of seasonally used areas is less than 3 km; migrant, the distance between centroids of seasonally used areas is more than 3 km with repeated seasonal return [38,42], or the elevational separation is >500 m [43]. The perimeter lengths of the home ranges, the largest linear distance of displacements, and patterns of winter and summer locations—particularly elevational shifts (asl), were also calculated and compared regarding sex and seasons. Although the quantity of locations is limited and covers several years, inter-annual site fidelity is common among cervids and thus permits a description of seasonal habitat use patterns [44]. Descriptive statistics were used to summarize the data and to describe the samples [45]. Hence, means and standard errors were computed, and compared by sex and season using independent t-test. Lastly, the VHF data from a male collected over 48 months was compared to satellite data from his new collar, covering 8 months. Additionally, opportunistic observations of unmarked deer were recorded by date and location and used to document spatiotemporal habitat use patterns in this population.

3. Results

3.1. Historical Distribution

Based on the broad literature review, a total of 130 historical records were found covering the years 1521 until 1915, and a further 190 records covering the years up to 1990. Publications since 1990 about huemul numbered 129, which is about 3.7% compared to the quantity of publications about red deer (Cervus elaphus), indicating the scant modern research activity concerning huemul.
Considering historical observations and records in the Protected Park Shoonem, huemul formerly occurred also much farther east of this watershed, following the water course and diminishing elevations. For example, numerous specimens were collected in a scientific expedition near the shore of Lake Fontana [46] some 35 km further east of currently extant huemul, and huemul were sighted in mountains some 180 km east of the present study population [47] (Section 4). The literature review resulted in numerous additional localities with historical evidence of huemul presence based on hunting collections, shed antlers, and archeological samples, as well as observations of residency and seasonal migratory behavior (n = 54, Table 1, Supplementary File S1). The historical distribution reached some 680 km further north of the currently northern-most and isolated population [48,49,50]. The historical distribution depicted in Section 4 is an approximation based on historical sites and at a scale that does not indicate potentially inhabitable areas [51]. However, most of this area contains guanaco [52] and allows livestock production [53], which thus serves to indicate that these areas also would sustain huemul. Historically used sites further east drop some 265 m in elevation for every 100 km towards the Atlantic coast, whereas annual precipitation drops from a maximum of 2000 mm at the continental divide to 400 mm at 100 km east, and down to 180 mm at another 150 km further east [5,54].

3.2. Extent of Altitudinal Movements

The six VHF radios of collared huemul resulted in 89 precise locations (mean = 14.83, SE = 2.6, range 6–24, Figure 3, Appendix A), whereas general surveys allowed additional confirmations of their seasonal presence (n = 935). The satellite unit placed on the male provided over 1675 additional location points over an 8-month period, corroborating the prior data based on his VHF radio. The coastline along the lake turned out to be the lowest elevation (930 m) used in the study area and thus represents the lower extent of the altitudinal gradient upon which movements were recorded, as none of the huemul moved further east and down the watershed during winter.
The elevations used during summer did not differ between males and females (t = 2.03, p = 0.056), and ranged between 930–1153 m asl (n = 37, mean = 1013.86 m, SE = 8.97). During winter, the elevations used also did not differ between males and females (t = −1.52, p = 0.101), and ranged between 930–1164 m asl (n = 29, mean = 969.03 m, SE = 10.27). The minimal elevational difference between the lowest summer and highest winter location for five huemul averaged a mere 36.2 m (SE = 21.06, range 11–119 m), with the highest elevations recorded in summer. One female though had a difference of 223 m, but with the highest elevation recorded during a mild winter, rather than during previous summers. However, these elevational usages are minimal and indicate residential behavior, as corroborated by general telemetry surveys every month, which numbered 37 to 239 observations per animal with a mean (±SE) of 150.17 (±40.06; n = 935). Year-round average elevations for males (mean = 979.53 m, SE = 7.36, range 966–992 m asl) were similar to those of females (mean = 1001.59 m, SE = 5.8, range 991–1011 m asl).
When comparing mean elevations in summer versus winter, these did not differ (t = −2.042, p = 0.055) for females (summer: 1050 m, SE = 31.82; winter: 976.42 m, SE = 16.87), whereas for males (summer: 1006.76 m, SE = 8.85; winter: 948.02 m, SE = 5.26) they differed slightly (t = −6.87, p = 0.02), although their absolute maximal altitudinal difference was only 159 m. Overall, the maximal annual elevational displacements by these six huemul across all months of the study period averaged only 149.5 m (range 74–229 m). Moreover, marked, and also many unmarked individuals, were located at the shore of the Lake La Plata (930 m asl) in every month of the year (Figure 3). Additional huemul signs (tracks, feces, bones) were recorded up to 1250 m asl, even though surveys were conducted up to 100 m above the treeline which is at 1300 m asl, corroborating earlier observations that there is little use above the treeline [3,28].

3.3. Extent of Horizontal Movements and Size of Annual Home Ranges (MCP)

Because individuals were marked in separate capture events in different areas, they all were considered to belong to different social groups. However, subsequent observations revealed that the home ranges of two females slightly overlapped, the home range of a male overlapped those of these two females (Figure 3a), and they were also sighted together. The longest horizontal displacement recorded within the habitually used yearly home ranges averaged 2133 m (SE = 448) among the six huemul. Two exceptional movement distances were not included in these data: a male with a home range not exceeding 1.6 km in dimension based on 14 months of data, suddenly left it in early spring, to move along the edge of the lake to a new site 4.7 km away (Figure 3b). There he died shortly after from starvation/malnutrition, and with severe bone pathology having negatively affected his foraging behavior [55]. One female (2–3 years old) moved 2.13 km after the capture event, but then returned shortly afterward to remain in a defined home range not exceeding 1.1 km in dimension. The perimeter length of yearly home ranges among 5 huemul averaged 5032 m (SE = 738), with an additional home range of a female having 11 km of perimeter. Since horizontal movements were recorded every month between August 2017 and April 2022, their magnitudes indicate the absence of migratory movements. This is in agreement with the absence of sightings of huemul (live, dead, shed antlers) further down the watershed for many decades.
The home range sizes reported here are a reflection of the sampling frequency of VHF radios allowing precise fixes, but covering all seasons and several years. The mean size of annual home ranges (MCP, representing minimal values) did not differ between males and females (t = −0.56674, p = 0.300589), averaging 159.33 ha (SE = 64.52). The home range size of one female was estimated at 464 ha and resulted from moving several kilometers to also use another flat area. In the one case, where the VHF collar was replaced with a satellite radio, the former technique determined a home range of 68 ha with a maximal displacement of 2147 m (n = 24), while the latter resulted in 190 ha with a maximal displacement of 2500 m (n = 1675). However, the 190 ha consisted of areas used at the end of winter and the beginning of summer, amounting to 163 ha and 147 ha, respectively, with an 80.5% overlap (Appendix A).

3.4. Processes of Migration and Geographic Expansion to New Habitat by Wild Cervids

Past over-harvesting not only resulted in the extirpation of local huemul subpopulations, but we hypothesize that it also eliminated their migratory traditions. Considering this loss, recognizing the processes involved in migratory traditions among cervids plays a key role in better understanding the consequence of losing such behavior [10].
Among cervids living in seasonal environments, including Odocoilines, a newly (re)colonized area initially has deer behaving as residents, whereby migratory behavior is non-existing. It also occurs even if the translocated animals stemmed from populations being migratory in their original site [12]. Only after multiple decades (up to 90 years for Alces) and increasing local population density, have translocated populations increase their propensity to start migrating again [12]. While deer movements are shaped by the distribution of resources for fine-scale foraging, this will eventually also include broad-scale migrations [56].
Regarding migration behavior, the multi-generational process to encounter and adopt movement corridors that allow green-wave surfing [56] plays an important role in the foraging strategy of Odocoilines, and the access to plant green-up along the migratory route is an additional key foraging benefit of migration [57]. Migration behavior thus not only refers to using fixed seasonal ranges, but also provides important foraging value while deer move along these corridors following the green wave (spring green-up), thereby enabling a prolonged exposure to high-quality forage and hence more energy [57].
Fundamental and primary mechanisms for ungulate migration evolvement are non-genetic processes of social learning and cultural transmission [12]. Moreover, spatial memory of the migration route had an extraordinary influence on migration, affecting movements manifold stronger than tracking spring green-up or autumn snow depth [56], and was characterized by strong fidelity [13,44,58]. Such spatial memory along with resource tracking allowed deer to repeatedly use the same migratory routes of 820 km round-trip [56,59]. Consequentially, the loss of migratory traditions will thus expunge generations of knowledge about the locations of high-quality forage and likely suppress population abundance [10,12], and leave pockets of potential habitat unoccupied because of the lost memory of viable migratory routes [13]. For instance, Odocoilines were shown to have little to no plasticity in terms of whether or where they migrated: resident deer remained residents, and migrant deer remained migrants, regardless of age, reproductive status or number of years monitored ([13,60]; Supplementary File S2). Certainly, some individual plasticity does occur and explains the development of new movement patterns including recolonizations [8,61].
For seasonal environments, Fretwell theorized in 1972 [17] how species would select habitats. Accordingly, under natural conditions, wild cervids tend to occupy all available habitats by doing best in source areas, so named for allowing positive population growth. Animals dispersing from source areas will also start to occupy suboptimal areas, including sink areas, so named because the local recruitment rate achieved there does not compensate for the local losses. There, populations are only maintained by replacement with newly arriving dispersers from source areas. Similarly, initial populations establishing themselves in source areas are year-round residents. Some dispersers, particularly in mountains at seasonal latitudes, will eventually move altitudinally to establish new summer ranges, and then return to their original winter area, thereby rejoining that resident population [62]. Thus, over several generations, basic plasticity becomes apparent, resulting in partial migration (coexisting resident and migratory individuals), changes in timing and routes, and also changes at the individual level [8]. Moreover, established migratory traditions can override signals of habitat quality and predation risks, such that deer can pass the best summer habitats to remain in the worst habitat at much further distance [13,58], or cross several mountain ranges to get to traditional winter-summer areas (Supplementary File S1). Similarly, when the culturally transmitted migratory behavior is interrupted after offspring lose their migratory mother, for instance, they will adopt resident behavior (Supplementary File S1). These well-documented processes of migratory behavior of cervids thus support the same hypothesized behavior among huemul, and the fact that it can be eradicated by over-killing, for instance.
Various observations show that the dispersal of adult or juvenile animals naturally connects source and sink areas [63]. Yet, source-sink population dynamics may change if dispersal is somehow constrained, e.g., by rapid anthropogenic changes in landscapes resulting in animals no longer making optimal habitat selection decisions as acquired by cultural transmission [64,65].
A comprehensive review by Xu et al., (2021) [8] revealed that many wild ungulates exhibit substantial migratory plasticity resulting in partial migration, and changes in migratory paths or localities. Their study revealed 127 migration change events in direct response to natural and human-induced environmental changes across 27 ungulate species. In addition to the suite of ecological processes playing a role which they described, we report here for the first time that the huemul is the only example of an ungulate in seasonal habitat having changed its behavior to become year-round residents in typical seasonal summer range habitat.

4. Discussion

4.1. Historical Spatial Habitat Use

The weight of evidence indicates that local extirpations of huemul resulted from overhunting by early humans and their dogs, which was exacerbated by huemul’s lack of anti-human behavior [66,67] (Supplementary File S1; Figure 1). Nonetheless, several historical accounts between the years 1521–1925 still mentioned huemul subpopulations — with some even considered numerous, extending from the Andean ecotonal foothills to the Patagonian mesas, and even reaching as far east as the Atlantic coast (Figure 4, Table 1). A huemul was harvested in 1904 at a site 270 km east of the continental divide, which is 225 km east of the eastern-most currently living huemul [25]. This hunt by the governor of Chubut was documented photographically [68]. Additionally, huemul were described to co-occur with guanaco in areas even reaching the Atlantic coast (Supplementary File S1).
The historical accounts of huemul also occurring far from forests are further corroborated by their osteology: analyses of rear limb ecomorphology indicate huemul are adapted to open habitat (unforested) areas [72,73]. The perceived short stature thus was not due to short limbs but to a thick coat of long hair (up to 19.5 cm long, Source: Shoonem Foundation collection) that concealed the leg length, thus creating a misperception [74]. Moreover, huemul limb morphology does not overlap with species considered mountain specialists, but falls within the range of other cervids, with some populations of Rangifer spp. and even Odocoileus virginianus having much shorter legs than huemul: these findings contradict the long-standing assertion that attributed the apparent short stature of huemul to be an adaptation to mountainous terrain [74,75]. Moreover, stable isotope analyses of archeological samples reveal that huemul’s diet from open environments cannot be differentiated from that of steppe guanaco [76]. Lastly, the cryptic pattern of fawns does not coincide with huemul having evolved in forested habitats. Camouflage appears to be the single most important evolutionary force in explaining why most cervids have spotted fawns: this crypsis provides the strongest protection in forests as a likely mechanism by which fawns could escape detection by predators [77]. Yet very few cervid species, including huemul, have non-spotted fawns, which is to be expected if natural selection acted on the species principally in open habitat areas [77]. Consequently, Webb (2000) proposed the cervid tribe Rangiferini, which includes the northern Rangifer and the southern Hippocamelus, the former using extensive open tundra and steppe [78]. Hence whereas historic data confirms a reduction in distributional range of huemul, together with anatomic data it also indicates the loss of migratory traditions (see below).

4.2. Historical Seasonal Habitat Use

The hypothesis that the huemul was once migratory like other cervids in seasonal environments, is corroborated by historical observations. Thus, in the past, huemul frequently were year-round residents in valleys and other low elevation winter ranges, while some migrated between summer and winter ranges, and formed large winter groups of more than 100 individuals [66,6979,80,81,82]. Current habitat use by huemul in the Protected Park Shoonem certainly represents only a small fraction of the watershed reported to have been used historically [46,47] (Figure 4). Additionally, old shed antlers, which are not related to a harvested animal, have been collected in historically used temperate grasslands, including a few sites near the Atlantic Ocean some 300 km east of Andean summer ranges. Analyses of aboriginal use of huemul showed that hunting occurred in grassland summer ranges, together with guanaco [82,83]. Moreover, it should be noted that even very early on (e.g., 1847, 1873, etc.) it was recognized that huemul already had regionally disappeared and remained mainly in high and inaccessible areas [31], which already were interpreted as being refugee areas (Supplementary File S1), though this perspective was lost in recent decades.
Similarly, bighorn sheep (Ovis canadensis) were overhunted, and lost their traditional seasonal migrations resulting in many sedentary herds and associated seasonal deficiencies due to low forage quality, which was considered the ultimate cause of declining herds, and one of the largest problems challenging their long-term persistence [84]. Sika deer (Cervus nippon) in Japan also respond strongly to hunting and disturbances, and their seasonal migrations aim toward safer areas to avoid hunting and culling, besides being warmer and less snowy in winter. Sika deer are avoiding hunting areas representing the most suitable foraging sites (e.g., pastures), to move to safer sites even with poor forage, like forested areas with hunting prohibited [85].

4.3. Contemporary Spatiotemporal Habitat Use in the Protected Park Shoonem

The pattern of seasonal habitat use was very similar among both sexes. Areas used during harsh winters occurred at the lowest possible local elevation (lake shores), but these are also regularly used during the remainder of the year. Conversely, during the mild winter of 2021, one female used an elevation even higher than in any summer since 2017. This explains the very reduced minimal yearly elevational displacement (mean = 36.2 m). Moreover, given that unmarked and some marked huemul were observed year-round at lake level, it appears most or all marked huemul used the shoreline during the summer as well. In comparison, huemul further north (Los Alerces National Park) [86] had a similar elevational displacement between average summer and winter locations of merely 200 m, which were classified as seasonal shifts. Moreover, these small seasonal elevational differences and distances, plus their presence at all elevations during the whole year, hence indicate that these huemul are non-migratory [6]. Huemul studied in Chile over several years in three different areas showed that winter and summer range usage largely overlapped, with an insignificant mean elevational displacement of about 200 m, and thus were considered non-migratory [87]. After reintroducing huemul around 1980 to Torres del Paine National Park, some family groups remained in low-elevation areas year-round, while other individuals eventually adopted a pattern of using areas somewhat elevated (up to 150 m higher) in summer, and descending to those lower areas mainly during winter, which was also considered as non-migratory [88,89,91]. Lastly, huemul in periglacial refuge areas by the Pacific coast also had limited elevational displacement as the treeline there is only at about 400 m asl. These huemul were non-migratory; they favored the flat and open grassland habitat, where twice as many fecal pellet groups were found as compared to those in forested hillsides [14].
Maximal horizontal movements were also very limited within the year-round habitually used home ranges of huemul in our study population at Shoonem Park. The longest movement (4.7 km) was made by a male in early spring, maybe induced by advanced disease, which terminated in his death from starvation shortly after (Figure 3b). In the study by Gill et al. [87], individuals considered non-migratory rarely moved more than 5 km, and the mean distance moved between summer and winter areas was 552 m (range 44–1219 m). In contrast, the reintroduction of seven huemul in the years since 2016 in the Los Rios region (Chile) revealed that exploratory movements of one male during the first month following his liberation included two excursions that reached 10 and 18 km in length, occurring in opposite directions, and that extended beyond his eventual home range (F. Vidal unpubl. data). The other animals moved less before establishing a home range. All released animals were born in the breeding center Huilo Huilo, they were radio-collared and released next to the center, and were permanently surveyed thereafter. It revealed that these adults and their fawns remained in the valley bottoms shared with the guanaco, and they never climbed the mountains that surround the center and their final home ranges (F. Vidal unpubl. data). The dispersal events and subsequent habitat use by these huemul so far are the first-ever documented cases, and they illustrate the movement potential of huemul, their all-year resident behavior in valley bottoms near riparian habitat with the best grass availability, which thus helps explain their historical distributions and movements, in concordance with the behavioral capacity of other Odocoilines.
The short movements displayed by huemul studied here explain the small year-round home ranges, averaging 167 ha (SE = 64.6), albeit possibly an underestimation. Year-round mean home range sizes in Tamango (Chile) were 318 ha, and similar between non-migratory females and males [87]. For huemul reintroduced to Torres del Paine National Park, year-round home range sizes as determined during a 10-year study varied between 269–336 ha [89]. In comparison, although similar-sized mule deer (Odocoileus hemionus) are typically migratory in the Rocky Mountains, resident deer on winter ranges utilized a continuous year-long home range, displacing only some 1300 m between seasons, and shifting to just slightly higher elevations in summer [60,90].
The huemul studied here clearly were year-round residents within a single and well-defined area, which was also corroborated by numerous antlers shed in that area. These resident huemul used an area that would be typical summer range habitat within this seasonal mountainous region, but is unsuitable for year-round inhabitancy due to nutritional limitations as evidenced by prevalent pathology [5,34,92] and lack of recovery of this population (Supplementary File S2). Moreover, solely between the early 1900s till the 1960s the snow line rose by about 100–200 m in many parts of the Andes [93], which likely plays a role in this study area and the concomitant performance of this study population.
One of the very few huemul subpopulations known to be recovering, that resulted from the reintroduction to Torres del Paine National Park, where valley bottoms function as source areas, has resulted in huemul spatially expanding to eastern grassland areas [88,91]. This suggests that the absence of recolonizations of additional areas by most other extant huemul groups is because their current habitats do not qualify as source areas. Thus, the absent population growth with simultaneous low densities, results in very few or no dispersers, and thus explains the recorded lack of recolonization. For instance, whereas initial colonists of low valleys reported unearthing old, shed antlers when first plowing [94], the very rare contemporary huemul disperser entering that valley usually ends up dying [95]. Additionally, the continuing extinction of numerous such groups localized in remote refugee areas have been documented [96]. This coincides with observations that whereas dispersing adults or juveniles naturally connect source and sink areas, this has not been registered among the remaining extant huemul groups.

4.4. Implications of Having Lost Migratory Traditions

Areas used by extant resident huemul at high elevations are considered to represent summer ranges, thus constituting an ecological trap. This is substantiated by red deer introduced to former huemul winter ranges: initially, they behaved exclusively as resident deer, remaining on the winter range year-round. However, after several decades of population growth, a segment of the herd became migratory [5]. Importantly, guanaco are also known to use high mountains and forests, with corresponding seasonal migrations to low elevation winter ranges [30,97], with displacement distances reaching 70 km [98,99]. Moreover, guanaco were also drastically overhunted like huemul, but in contrast, have largely been eliminated from their prior mountainous distribution [31,52,100,101,102]. Notably, since colonial times, past and current livestock producers practice transhumance by herding their animals out of the Shoonem Protected Park before winter, as is the practice in other similar watersheds both in Argentina and Chile, in order to move them to areas considered appropriate winter ranges [103]. Odocoilines of similar body size were shown to avoid areas with >40 cm of snow [104], which may explain the use by huemul of the lake shores in the Protected Park Shoonem during peak winter, where snowpacks are considerably reduced along the beaches (Figure 2). However, this year-round residency in a seasonal area classified as a summer range can result in health problems due to dietary deficiencies.
Historical remarks already considered the use of summer and winter ranges as a determinant of huemul health. Given the low density of herbivores in most areas of extant huemul, protein and energy supplies are considered adequate and cannot explain the prevalent disease pattern or the lacking population recovery (Supplementary File S2). Health issues have now been corroborated by the high prevalence of skeletal pathologies in huemul spread over a large geographical region [95], including nearly 90% of individuals reported in this study, which qualifies these huemul as refugees (Supplementary File S2). Nutritional deficiencies were hypothesized to account for the high incidence of bone disease [106]. For one, valley bottoms tend to have soils enriched in minerals due to the topographic effect and accordingly, huemul reported here as residents in a summer range suffer from acute geochemical stress [105]. Moreover, huemul were shown to be deficient in essential micronutrients (Se, Cu, Mn) which coincides with their skeletal problems [55,107,108], and low average life span [34]. This is similar to situations in bighorn sheep [84] (Supplementary File S2), and likely explains the unusual reactions of huemul to other diseases due to their compromised metabolic and immune systems [55]. In contrast, migratory mammalian herbivores partially living in resource-poor environments travel farthest to fulfill their resource needs [109,110].
By eliminating sedentary subpopulations on winter ranges and consistently removing the last dispersing huemul, the remaining animals exhibit the aberrant behavior of becoming tied year-round to refugee areas that qualify as seasonal summer ranges. This artificial anthropogenic elimination of migratory traditions has resulted in most extant huemul remaining in suboptimal Anthropocene refugia [19,111]. Clearly, the resident behavior reported here for huemul taking place on a seasonal summer range is not the norm for cervids that use winter ranges either as residents or seasonal migrators.

4.5. Implications for Conservation

To base conservation strategies for huemul on its modern distribution is erroneous due to being an artifact, as has been recognized for other ungulates [18,84,108] (Supplementary File S2). Remarkably, Grzimek in 1973 [81] already recognized that huemul have been exterminated in most historical areas, such that they only survive in a few small mountain refuge areas (“bis auf wenige winzige Rückzugsgebiete ausgerottet”). Moreover, it is essential that the “shifting baseline syndrome” be overcome [112], that is repeating old, unfounded and outdated interpretations, like huemul being a “mountain deer”, being short-legged, non-migratory, etc., which qualify as stereotyping and compromising conservation efficacy [20]. As shown with published fake information, these are cited many times, over long periods, and have even caused an impact on human health [113]. The largest risk for refugee species occurs when the currently occupied suboptimal habitats are identified as the conservation priority areas for the species in question, as has been modeled for huemul based on the extant distribution (e.g., Riquelme et al.) [2]. This risk is especially large when the species has been limited to suboptimal habitat for numerous decades [17], even centuries, as has occurred with resilient huemul. Acknowledging historical species ranges is thus important for recovering endangered species [17,19,20,114,115,116,117] (Supplementary File S2). Most important for non-recovering subpopulations is the need to differentiate if extant subpopulations live in a marginal or natural sink area, or in an artificial ecological trap, since the latter two will drive a local subpopulation to become extirpated. Moreover, sedentariness on seasonal summer ranges by loss of migratory culture may be one of the largest problems challenging the long-term persistence of most huemul subpopulations, as has been determined for bighorn sheep populations (Supplementary File S2). Illustratively, the rare case of a growing huemul subpopulation after its reintroduction in Torres del Paine National Park, with resident groups in valley bottoms, is expanding into grassland areas where they overlap with guanaco [88,91]. Similarly, huemul reintroduced in the Los Rios region (Chile) became residents in valley bottoms together with guanaco (F. Vidal unpubl. data). To recover endangered huemul, Kauffman et al. [117] recently pointed out the importance to consider their historical distribution and migratory tradition. Moreover, it is critically important to recognize the length of time required to reestablish migratory behavior as shown in different cervids, which needed 12 or more generations to reinitiate migratory behavior, once a critical density among residents was attained (Supplementary File S2). A key requirement will be the conservation of “migratory routes”, a target essentially already projected in Argentina, by huemul being declared a Natural Monuments by federal and provincial laws [118]. An additional tool is declaring new areas containing migratory routes as a Natural Monument according to Category III of the IUCN. Importantly, the preservation of migratory routes also allows fundamental ecological processes to continue (food webs, nutrient cycling) [119], besides their function to assure the survival of species dependent on seasonal migration [10,117].
Differentiating between the proximate and ultimate causes of mortality is necessary to understand the population dynamics of ungulate populations. Particularly the interaction between predation and malnutrition as a cause of mortality is difficult to disentangle without manipulative experiments or other means of assessment [120]. Therefore, to experimentally test the refugee interpretation, it is highly recommended that huemul be reintroduced into habitats proposed to be critical source areas and with minimal modern anthropogenic threats, to monitor their habitat-specific fitness, while using animals in the currently inhabited refuge areas as controls [10,17,20,118]. Reverting the artificial situation would require creating resident subpopulations of huemul in formerly used winter ranges (Figure 4). Furthermore, instead of waiting until reaching densities that promote the natural emigration and re-establishment of migratory traditions, this process could be accelerated by training young animals via imprinting to acquire a migratory pattern, as has been done successfully with other ungulates [121,122]. Once winter ranges are repopulated, along with positive recruitment rates, the expansion to unoccupied ranges, i.e., neighboring winter and summer ranges, can occur. Available evidence supports the hypothesis that a major factor behind the current failure of many huemul subpopulations to recover numerically and spatially is the current absence of their members in suitable winter ranges. Repopulating such areas would in time also allow reconnections between the currently isolated subpopulations, concordant with the common pattern among other cervids in seasonal regions, which consists of mixed group compositions on both summer and winter ranges. In this way, a winter range frequently receives migratory members from several distinct summer ranges, while a summer range will receive members from distinct winter ranges [58,60,123,124,125], and thus contributes to gene pool diversity.
A lesson learned from this study, of general application to conservation biology, is that it can be a fatal mistake to define the “area of habitat” (AOH according to Brooks et al.) [126] for an endangered species on the basis of its current distribution. This distributional range is often not the same environmental space that was once occupied by the species under natural conditions (from source to sink habitats) but is instead a refuge where it was displaced by the human footprint, and frequently is nutritionally insufficient to sustain its populations. Unfortunately, the ‘protected area paradox’ [15,19,127], which is widespread and applies to huemul, has facilitated the provision of protection in less productive habitats and has resulted in ineffectual attempts to conserve huemul in suboptimal habitats (i.e., as refugee species) where the subpopulation barely persists at extremely low densities and with compromised health issues.

5. Conclusions

Making a leap towards conceptualizing what constitutes the fundamental factors preventing recovery of most subpopulations could release the huemul from its current imperilment. Many winter ranges historically used by all-year residents and also by migratory huemul apparently have turned from source to sink areas, mainly because of human predation in the past, and currently due to a lack of dispersers, due to the abundance of humans, dogs, automobile traffic, and agricultural land conversion. Among cervids in seasonal mountain areas, the huemul appears to be the only one that has mostly year-round resident subpopulations in what would be considered a typical summer range, and thus can be classified as an unfortunate refugee species, stuck in an ecological trap. To our knowledge, this is the first published account of a cervid species afflicted by these circumstances. Several huemul refugee subpopulations are known to be severely afflicted with disease resulting from concomitant micronutrient deficiencies, which explains their short life spans, and absence of both population growth and spatial expansion. Major steps towards reverting the prevailing absence of recovery over the past decades will be their reintroduction to historic winter source areas and the subsequent encouragement and fostering of reestablishing the migratory tradition. Additionally, a numerical and spatial recovery will also result in reconnecting the currently isolated subpopulations. The distributional retraction of the huemul and the extirpation of numerous local and isolated populations, including islands such as Tierra del Fuego and Chiloé, clearly show that without strong assistance from novel conservation technologies it will be difficult to prevent the extinction of this endemic deer in Patagonia.

Supplementary Materials

The following supporting information can be downloaded at: https://www.mdpi.com/article/10.3390/conservation2020023/s1. File S1: Compendium of the Review about the Huemul Distribution in Patagonia: Past and Present; File S2: Compendium: Review of Consequences for Ungulates when losing Migratory Traditions. References [129,130,131,132,133,134,135,136,137,138,139,140,141,142,143,144,145,146,147,148,149,150,151,152,153,154,155,156,157,158,159,160,161,162,163,164,165,166,167,168,169,170,171,172,173,174,175,176,177,178,179,180,181,182,183,184,185,186,187,188,189,190,191,192,193,194,195,196,197,198,199,200,201,202,203,204,205,206,207,208,209,210,211,212,213,214,215,216,217,218,219,220,221,222,223,224,225,226,227,228,229,230,231,232,233,234,235,236,237,238,239,240,241,242,243,244,245,246,247,248,249,250,251,252,253,254,255,256,257,258,259,260,261,262,263,264,265,266,267,268,269,270,271,272,273,274,275,276,277,278,279,280,281,282,283,284,285,286,287,288,289,290,291,292,293,294,295,296,297,298,299,300,301,302,303,304] are cited in Supplementary Materials File.

Author Contributions

Conceptualization, W.T.F., J.A.M.S.-F., Z.G. and F.V.; Data curation, W.T.F. and J.A.M.S.-F.; Formal analysis, W.T.F., J.A.M.S.-F., M.Z., V.G., J.R.H., P.B.-D., Z.G., A.J.M. and J.E.J.; Funding acquisition, W.T.F. and J.A.M.S.-F.; Investigation, W.T.F., J.A.M.S.-F., M.E.E., M.Z., B.F., F.V., J.B., S.M.M. and S.H.; Methodology, W.T.F.; Supervision, W.T.F.; Writing—original draft, W.T.F., J.A.M.S.-F., M.Z., B.F., V.G., J.R.H., P.B.-D., Z.G., F.V., J.B., S.M.M., A.J.M., S.H. and J.E.J.; Writing—review & editing, W.T.F., J.A.M.S.-F., M.E.E., M.Z., B.F., V.G., J.R.H., P.B.-D., Z.G., F.V., J.B., S.M.M., A.J.M., S.H. and J.E.J. All authors have read and agreed to the published version of the manuscript.

Funding

We would like to extend our gratitude to the Foundation Erlenmeyer, Basel, Switzerland who provided the main funding for this work, the Projekt Gabelhirsch.

Institutional Review Board Statement

All animals used in this study were managed in accordance with strict animal ethics protocols and according to the Ministry of Production, Province of Chubut in Argentina (permit Disp. Nr. 22/2017-DfyFS-M.P.).

Informed Consent Statement

Not applicable.

Data Availability Statement

Not applicable.

Acknowledgments

We recognize that this work has been supported mainly from a grant by the Erlenmeyer Stiftung, Switzerland. We thank Rudolf Roth, Martin Hug, and Kurt Aeschbacher for accompanying the huemul work during several years with a great understanding of all the challenges. We thank Enrique Verde and Renzo Mottino for outstanding field assistance, and the Dirección de Fauna y Flora Silvestre of the Province of Chubut for supporting, encouraging, and permitting the research and conservation management projects with huemul (permit Disp. Nr. 47/2019-DFyFS-M.P.). Much appreciated are the various assistants who helped with field- and other valuable work and input, and the Rewilding Argentina Foundation for the placement of a first satellite radio collar. We also thank Rory Putman and William Fagan for the detailed analyses and improvements made to the manuscript, as well as several reviewers for all their constructive comments. Finally, we dedicate this article in memory of our co-author Valerius Geist (1934–2021), a highly respected evolutionary and behavioral ecologist who was a pioneer in his field. He spent years towards understanding why the huemul fails to recover despite the millions of hectares available of so-called “prime habitat”, and he actively participated on the IUCN Huemul Task Force and in the last two assessments of this species for the IUCN Red list. Thank you, Val, for your invaluable input in making this paper a reality.

Conflicts of Interest

The authors declare no conflict of interest.

Appendix A

Table A1. Spatial behavior of huemul during summer and winter (precise locations based on VHF radio collars), in the Protected Park Shoonem, Chubut, Argentina, between August 2017 and April 2022.
Table A1. Spatial behavior of huemul during summer and winter (precise locations based on VHF radio collars), in the Protected Park Shoonem, Chubut, Argentina, between August 2017 and April 2022.
Total Data Points (N)Minimal Separation (m)Vital Status
CasesSummerWinterSpring/FallHorizontalAltitudinalCaptureDead
female 1364122682017unknown
female 2231214014320172018
female 3875002017alive
male 13354233820172018
male 2771897120172020
male 314460020172022
Table A2. Scale of the spatial behavior of Male 3 (satellite radio collar) at the end of winter, and the autumn of 2022.
Table A2. Scale of the spatial behavior of Male 3 (satellite radio collar) at the end of winter, and the autumn of 2022.
GPS Fix * (n)Area (ha)Perimeter (m)OverlapSeasonal Displacement
(a) Aug 13–Sept 30437163576076 %490 m more south than (b)
(b) Oct 1–Dec 10405147481085%235 m more north than (a)
(c) Dec 11–Mar 7 20228331585440 no changes
Total survey16751906610
* these are precise points with DOP of 2 or less.

References

  1. Black-Decima, P.A.; Corti, P.; Díaz, N.; Fernandez, R.; Geist, V.; Gill, R.; Gizejewski, Z.; Jiménez, J.; Pastore, H.; Saucedo, C.; et al. Hippocamelus bisulcus. In The IUCN Red List of Threatened Species; International Union for Conservation of Nature and Natural Resources: Gland, Switzerland, 2016. [Google Scholar]
  2. Riquelme, C.; Estay, S.A.; López, R.; Pastore, H.; Soto-Gamboa, M.; Corti, P. Protected areas’ effectiveness under climate change: A latitudinal distribution projection of an endangered mountain ungulate along the Andes Range. PeerJ 2018, 6, e5222. [Google Scholar] [CrossRef] [PubMed]
  3. Smith-Flueck, J.M.; Flueck, W.T. Natural mortality patterns in a population of southern Argentina huemul (Hippocamelus bisulcus), an endangered Andean cervid. Zeits. Jagdwiss. 2001, 47, 178–188. [Google Scholar] [CrossRef]
  4. Putman, R.; Flueck, W.T. Intraspecific variation in biology and ecology of deer: Magnitude and causation. Anim. Prod. Sci. 2011, 51, 277–291. [Google Scholar] [CrossRef] [Green Version]
  5. Flueck, W.T.; Smith-Flueck, J.M. Recent advances in the nutritional ecology of the Patagonian huemul: Implications for recovery. Anim. Prod. Sci. 2011, 51, 311–326. [Google Scholar] [CrossRef]
  6. Grotta-Neto, F.; Duarte, J.M.B. Movements of Neotropical Forest Deer: What Do We Know? In Movement Ecology of Neotropical Forest Mammals: Focus on Social Animals; Reyna-Hurtado, R., Chapman, C., Eds.; Springer Nature: Cham, Switzerland, 2019; pp. 95–109. [Google Scholar]
  7. Peters, W.; Hebblewhite, M.; Mysterud, A.; Spitz, D.; Focardi, S.; Urbano, F.; Morellet, N.; Heurich, M.; Kjellander, P.; Linnell, J.D.C.; et al. Migration in geographic and ecological space by a large herbivore. Ecol. Monogr. 2017, 87, 297–320. [Google Scholar] [CrossRef] [Green Version]
  8. Xu, W.; Barker, K.; Shawler, A.; Van Scoyoc, A.; Smith, J.A.; Mueller, T.; Sawyer, H.; Andreozzi, C.; Bidder, O.R.; Karandikar, H.; et al. The plasticity of ungulate migration in a changing world. Ecology 2021, 102, e03293. [Google Scholar] [CrossRef]
  9. Howard, W.E. Innate and environmental dispersal of individual vertebrates. Am. Midl. Nat. 1960, 63, 152–161. [Google Scholar] [CrossRef]
  10. Bolger, D.T.; Newmark, W.D.; Morrison, T.A.; Doak, D.F. The need for integrative approaches to understand and conserve migratory ungulates. Ecol. Lett. 2008, 11, 63–77. [Google Scholar] [CrossRef]
  11. Festa-Bianchet, M. Learning to migrate. Science 2018, 361, 972–973. [Google Scholar] [CrossRef]
  12. Jesmer, B.R.; Merkle, J.A.; Goheen, J.R.; Aikens, E.O.; Beck, J.L.; Courtemanch, A.B.; Hurley, M.A.; McWhirter, D.E.; Miyasaki, H.M.; Monteith, K.L.; et al. Is ungulate migration culturally transmitted? Evidence of social learning from translocated animals. Science 2018, 361, 1023–1025. [Google Scholar] [CrossRef] [Green Version]
  13. Sawyer, H.; Merkle, J.A.; Middleton, A.D.; Dwinnell, S.P.H.; Monteith, K.L. Migratory plasticity is not ubiquitous among large herbivores. J. Anim. Ecol. 2019, 88, 450–460. [Google Scholar] [CrossRef]
  14. Frid, A. Habitat use by the endangered huemul (Hippocamelus bisculcus): Cattle, snow, and the problem of multiple causes. Biol. Conserv. 2001, 100, 261–267. [Google Scholar] [CrossRef]
  15. Joppa, L.N.; Pfaff, A. High and Far: Biases in the Location of Protected Areas. PLoS ONE 2009, 4, e8273. [Google Scholar] [CrossRef]
  16. Quevedo, P.; Von Hardenberg, A.; Pastore, H.; Álvarez, J.; Corti, P. Predicting the potential distribution of the endangered huemul deer Hippocamelus bisulcus in North Patagonia. Oryx 2017, 51, 315–323. [Google Scholar] [CrossRef] [Green Version]
  17. Kerley, G.I.H.; Kowalczyk, R.; Cromsigt, J.P.G.M. Conservation implications of the refugee species concept and the European bison: King of the forest or refugee in a marginal habitat? Ecography 2012, 35, 519–529. [Google Scholar] [CrossRef]
  18. Faurby, S.; Araujo, M.B. Anthropogenic range contractions bias species climate change forecasts. Nat. Clim. Change 2018, 8, 252–256. [Google Scholar] [CrossRef]
  19. Nüchel, J.; Bocher, P.K.; Xiao, W.; Zhu, A.X.; Svenning, J.C. Snub-nosed monkeys (Rhinopithecus): Potential distribution and its implication for conservation. Biodiv. Cons. 2018, 27, 1517–1538. [Google Scholar] [CrossRef] [Green Version]
  20. Britnell, J.A.; Lewis, R.N.; Elsner-Gearing, F.; Harvey, N.; Stanbrook, E.; Shultz, S. Species stereotypes as a result of unconscious research biases compromise conservation efficacy. Biol. Cons. 2021, 261, 109275. [Google Scholar] [CrossRef]
  21. Redford, K.H.; Eisenberg, J.F. Mammals of the Neotropics: The Southern Cone, Volume 2, Chile, Argentina, Uruguay, Paraguay; The University of Chicago Press: Chicago, IL, USA; London, UK, 1992. [Google Scholar]
  22. Aldridge, D.; Lopez, R.; Saucedo, C.; Vila, A.R. Los Ultimos Senderos del Huemul; Fundacion Hiunay: Santiago, Chile, 2007; ISBN 9789567667062. [Google Scholar]
  23. Vila, A.R.; Saucedo, C.; Aldridge, D.; Ramilo, E.; Corti, P. South Andean Huemul Hippocamelus bisulcus (Molina 1782). In Neotropical Cervidology; Duarte, J.M., González, S., Eds.; FUNEP: Jaboticabal, Brazil, 2010; pp. 89–100. [Google Scholar]
  24. Gonzalez, S.; Barbanti Duarte, J.M. Speciation, evolutionary history and conservation trends of neotropical deer. Mastozool. Neotrop. 2020, 27, 37–47. [Google Scholar] [CrossRef]
  25. Escobar Ruiz, E.M.; Smith, J.M.; Flueck, W.T. El Huemul—Shoonem: Madera que se Mueve/re, 2nd ed.; Biblioteca Popular “Dr. Enrique Perea”: Alto Río Senguer, Argentina, 2020. [Google Scholar]
  26. Díaz, N.I.; Smith-Flueck, J. The Patagonian Huemul. A Mysterious Deer on the Brink of Extinction; Literature of Latin America: Buenos Aires, Argentina, 2000. [Google Scholar]
  27. Donoso, D.; Iriarte, A.; Segura, B.; Tirado, M. Antecedentes de Huemul (Capitulo 1). In El Huemul de Aysén y Otros Rincones; Iriarte, A., Donoso, D.S., Segura, B., Tirado, M., Eds.; Ediciones Secretaría Regional Ministerial de Agricultura de la Región de Aysén y Flora & Fauna Chile Ltd.: Aysen, Chile, 2017; pp. 13–61. [Google Scholar]
  28. Smith-Flueck, J.M. La Ecología del Huemul (Hippocamelus bisulcus) en la Patagonia Andina de Argentina y Consideraciones sobre su Conservación. Ph.D Thesis, Universidad Nacional del Comahue, Bariloche, Argentina, 2003; pp. 1–361. [Google Scholar]
  29. Sauter, T. Revisiting extreme precipitation amounts over southern South America and implications for the Patagonian Icefields. Hydrol. Earth Syst. Sci. 2020, 24, 2003–2020. [Google Scholar] [CrossRef] [Green Version]
  30. Puig, S.; Rosi, M.I.; Videla, F.; Mendez, E. Summer and winter diet of the guanaco and food availability for a High Andean migratory population (Mendoza, Argentina). Mammal. Biol. 2011, 76, 727–734. [Google Scholar] [CrossRef]
  31. Kölliker, A.; Kühn, F.; Reichert, F.; Tomsen, A.; Witte, L. Patagonia. Resultado de las expediciones en 1910 a 1916; Sociedad Científica Alemana: Buenos Aires, Argentina, 1917; Volumes 1 and 2. [Google Scholar]
  32. Citino, S.B.; Bush, M.; Grobler, D.; Lance, W. Anesthesia of boma-captured Lichtenstein’s hartebeest (Sigmocerus lichtensteinii) with a combination of thiafentanil, medetomidine and ketamine. J. Wildl. Dis. 2002, 38, 457–462. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  33. Muller, L.I.; Osborn, D.A.; Doherty, T.; Keel, M.K.; Miller, B.F.; Warren, R.J.; Mille, K.V. Optimal Medetomidine Dose When Combined with Ketamine and Tiletamine-zolazepam to Immobilize White-tailed Deer. J. Wildl. Dis. 2012, 48, 477–482. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  34. Flueck, W.T.; Smith-Flueck, J.M. Age-independent osteopathology in skeletons of a south American cervid, the Patagonian huemul (Hippocamelus bisulcus). J. Wildl. Dis. 2008, 44, 636–648. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  35. White, G.C.; Garrott, R.A. Analysis of Wildlife Radio-Tracking Data; Academic Press: San Diego, CA, USA, 1990. [Google Scholar]
  36. Harris, S.; Cresswell, W.J.; Forde, P.G.; Trewhella, W.J.; Woollard, T.; Wray, S. Home-range analysis using radio tacking data: A review of problems and techniques particularly as applied to the study of mammals. Mammal. Rev. 1990, 20, 97–123. [Google Scholar] [CrossRef]
  37. Börger, L.; Franconi, N.; de Michele, G.; Gantz, A.; Meschi, F.; Manica, A.; Lovari, S.; Coulson, T. Effects of sampling regime on the mean and variance of home range size estimates. J. Anim. Ecol. 2006, 75, 1393–1405. [Google Scholar] [CrossRef]
  38. Zweifel-Schielly, B.; Kreuzer, M.; Ewald, K.C.; Suter, W. Habitat selection by an Alpine ungulate: The significance of forage characteristics varies with scale and season. Ecography 2009, 32, 103–113. [Google Scholar] [CrossRef]
  39. Kropil, R.; Smolko, P.; Garaj, P. Home range and migration patterns of male red deer Cervus elaphus in Western Carpathians. Eur. J. Wildl. Res. 2015, 61, 63–72. [Google Scholar] [CrossRef] [Green Version]
  40. Kie, J.G.; Matthiopoulos, J.; Fieberg, J.; Powell, R.A.; Cagnacci, F.; Mitchell, M.S.; Gaillard, J.M.; Moorcroft, P.R. The home-range concept: Are traditional estimators still relevant with modern telemetry technology? Philos. Trans. R. Soc. B 2010, 365, 2221–2231. [Google Scholar] [CrossRef] [Green Version]
  41. Powell, R.A. Animal home ranges and territories and home range estimators. In Research Techniques in Animal Ecology: Controversies and Consequences; Boitani, L., Fuller, T., Eds.; Columbia University Press: New York, NY, USA, 2000. [Google Scholar]
  42. Bunnefeld, N.; Börger, L.; van Moorter, B.; Rolandsen, C.M.; Dettki, H.; Solberg, E.J.; Ericsson, G. A model-driven approach to quantify migration patterns: Individual, regional and yearly differences. J. Anim. Ecol. 2011, 80, 466–476. [Google Scholar] [CrossRef]
  43. Spitz, D.B.; Hebblewhite, M.; Stephenson, T.R.; German, D.W. How plastic is migratory behavior? Quantifying elevational movement in a partially migratory alpine ungulate, the Sierra Nevada bighorn sheep (Ovis canadensis sierrae). Can. J. Zool. 2018, 96, 1385–1394. [Google Scholar] [CrossRef] [Green Version]
  44. Morrison, T.A.; Merkle, J.A.; Hopcraft, J.G.C.; Aikens, E.O.; Beck, J.L.; Boone, R.B.; Courtemanch, A.B.; Dwinnell, S.P.; Fairbanks, W.S.; Griffith, B.; et al. Drivers of site fidelity in ungulates. J. Anim. Ecol. 2021, 90, 955–966. [Google Scholar] [CrossRef]
  45. Marshall, G.; Jonker, L. An introduction to descriptive statistics: A review and practical guide. Radiography 2010, 16, e1–e7. [Google Scholar] [CrossRef]
  46. Anchorena, A. Descripción Gráfica de la Patagonia y Valles Andinos; Compañia Sudamericana de Billetes de Banco: Buenos Aires, Argentina, 1902. [Google Scholar]
  47. Onelli, C. El huemul. Su patria: Su vida. Rev. Jardín Zool. Buenos Aires 1905, 1, 370–374. [Google Scholar]
  48. Bahre, C.J. Destruction of the Natural Vegetation of North-Central Chile; University of California Publications in Geography: Berkeley, CA, USA, 1979; Volume 23, pp. 1–117. ISBN 0-520-09594-4. [Google Scholar]
  49. Saavedra, B.; Simonetti, J.A. Archaeological evidence of Pudu pudu (Cervidae) in central Chile. Z. Saeugetierkunde 1991, 56, 252–253. [Google Scholar]
  50. Ale, A. A social economic formation of hunter-gatherers in the semiarid northern Chile: A revaluation of San Pedro Viejo of Pichasca site. Zaranda Ideas 2014, 11, 67–88. [Google Scholar]
  51. Merow, C.; Wilson, A.M.; Jetz, W. Integrating occurrence data and expert maps for improved species range predictions. Glob. Ecol. Biogeogr. 2017, 26, 243–258. [Google Scholar] [CrossRef]
  52. Travaini, A.; Zapata, S.C.; Bustamante, J.; Pedrana, J.; Zanón, J.I.; Rodríguez, A. Guanaco abundance and monitoring in Southern Patagonia: Distance sampling reveals substantially greater numbers than previously reported. Zool. Stud. 2015, 54, 23. [Google Scholar] [CrossRef] [Green Version]
  53. De Gea, G. El Ganado Lanar en la Argentina, 2nd ed.; Universidad Nacional de Río Cuarto: Río Cuarto, Argentina, 2007; ISBN 978-950-665-448-1. [Google Scholar]
  54. Uboldi, J.A.; Angeles, G.R.; Gentili, J.O.; Geraldi, A.M.; Melo, W.D.; Carbone, M.E. Geotecnologías del sur Argentino. Casos de Estudio; Departamento de Geografía y Turismo (TIG), Universidad Nacional del Sur: Bahía Blanca, Argentina, 2014. [Google Scholar]
  55. Flueck, W.T. Nutrition as an etiological factor causing diseases in endangered huemul deer. BMC Res. Notes 2020, 13, 276. [Google Scholar] [CrossRef]
  56. Merkle, J.A.; Sawyer, H.; Monteith, K.L.; Dwinnell, S.P.; Fralick, G.L.; Kauffman, M.J. Spatial memory shapes migration and its benefits: Evidence from a large herbivore. Ecol. Lett. 2019, 22, 1797–1805. [Google Scholar] [CrossRef]
  57. Aikens, E.O.; Kauffman, M.J.; Merkle, J.A.; Dwinnell, S.P.H.; Fralick, G.L.; Monteith, K.L. The greenscape shapes surfing of resource waves in a large migratory herbivore. Ecol. Lett. 2017, 20, 741–750. [Google Scholar] [CrossRef]
  58. Flueck, W.T. The effect of selenium on reproduction of black-tailed deer (Odocoileus hemionus columbianus) in Shasta County, California. Ph.D Thesis, University of California, Davis, CA, USA, 1989. [Google Scholar]
  59. Kauffman, M.J.; Copeland, H.E.; Berg, J.; Bergen, S.; Cole, E.; Cuzzocreo, M.; Dewey, S.; Fattebert, J.; Gagnon, J.; Gelzer, E.; et al. Ungulate Migrations of the Western United States; Report 2020–5101; U.S. Geological Survey Scientific Investigations: Reston, VA, USA, 2020; Volume 1. [Google Scholar]
  60. Gogan, P.J.P.; Klaver, R.W.; Olexa, E.M. Northern Yellowstone Mule Deer Seasonal Movement, Habitat Selection, and Survival Patterns. West. North Am. Nat. 2019, 79, 403–427. [Google Scholar] [CrossRef]
  61. Van de Kerk, M.; Larsen, R.T.; Olson, D.D.; Hersey, K.R.; McMillan, B.R. Variation in movement patterns of mule deer: Have we oversimplified migration? Mov. Ecol. 2021, 9, 44. [Google Scholar] [CrossRef] [PubMed]
  62. Haller, H. Der Rothirsch im Schweizerischen Nationalpark und dessen Umgebung. Eine alpine Population von Cervus elaphus zeitlich und räumlich dokumentiert. Natl. Forsch. Schweiz 2002, 91, 1–144. [Google Scholar]
  63. Howe, R.W.; Davis, G.J.; Mosca, V. The demographic significance of ‘sink’ populations. Biol. Conserv. 1990, 57, 239–255. [Google Scholar] [CrossRef]
  64. Remes, V. How can maladaptive habitat choice generate source-sink population dynamics? Oikos 2000, 91, 579–582. [Google Scholar] [CrossRef]
  65. Braunisch, V.; Bollmann, K.; Graf, R.F.; Hirzel, A.H. Living on the edge. Modelling habitat suitability for species at the edge of their fundamental niche. Ecol. Model. 2008, 214, 153–167. [Google Scholar] [CrossRef]
  66. Goss, R.J. Deer Antlers: Regeneration, Function and Evolution; Academic Press: New York, NY, USA, 1983. [Google Scholar]
  67. Tonko, J. Kawesqar travel narratives. Onomazein 2008, 18, 11–47. [Google Scholar]
  68. Anonymous. Excursión del gobernador del Chubut. Caras Caretas 1904, 7, 58. [Google Scholar]
  69. Moreno, F.P. Apuntes preliminares sobre una excursion a los territorios del Neuquén, Río Negro, Chubut y Santa Cruz. Rev. Mus. La Plata 1898, 8, 200–459. [Google Scholar]
  70. Steffen, H. Viajes de exploracion: Estudio en la Patagonia occidental 1892–1902. Anal. Univ. Chile 1910, 2, 1–419. [Google Scholar]
  71. Jiménez, J.; Guineo, G.; Corti, P.; Smith, J.A.; Flueck, W.T.; Vila, A.; Gizejewski, Z.; Gill, R.; McShea, B.; Geist, V. Hippocamelus bisulcus. In IUCN Red List of Threatened Species; IUCN: Gland, Switzerland, 2008. [Google Scholar]
  72. Curran, S.C. Expanding ecomorphological methods: Geometric morphometric analysis of Cervidae post-crania. J. Archaeol. Sci. 2012, 39, 1172–1182. [Google Scholar] [CrossRef]
  73. Curran, S.C. Exploring Eucladoceros ecomorphology using geometric morphometrics. Anat. Rec. 2015, 298, 291–313. [Google Scholar] [CrossRef]
  74. Flueck, W.T.; Smith-Flueck, J.M. Osteological comparisons of appendicular skeletons: A case study on Patagonian huemul deer and its implications for conservation. Anim. Prod. Sci. 2011, 51, 327–339. [Google Scholar] [CrossRef] [Green Version]
  75. Flueck, W.T. Functional limb anatomy in a refugee species: The endangered Patagonian huemul deer (Hippocamelus bisulcus). Anat. Histol. Embryol. 2021, 50, 411–416. [Google Scholar] [CrossRef]
  76. Barberena, R.; Méndez, C.; Mena, F.; Reyes, O. Endangered species, archaeology, and stable isotopes: Huemul (Hippocamelus bisulcus) isotopic ecology in central-western Patagonia (South America). J. Archaeol. Sci. 2011, 38, 2313–2323. [Google Scholar] [CrossRef]
  77. Caro, T. The adaptive significance of coloration in mammals. BioScience 2005, 55, 125–136. [Google Scholar] [CrossRef] [Green Version]
  78. Webb, S.D. Evolutionary history of New World Cervidae. In Antelopes, Deer, and Relatives; Vrba, E.S., Schaller, G.B., Eds.; Yale University Press: New York, NY, USA, 2000; pp. 38–64. [Google Scholar]
  79. Prichard, H.H. Field notes upon some of the larger mammals of Patagonia made between September 1900 and June 1901. Proc. Zool. Soc. Lond. 1902, 1, 272–277. [Google Scholar]
  80. Giai, A.G. Huemul, inofensivo venado de las soledades cordilleranas de la Patagonia. Chacra 1936, 6, 99–101. [Google Scholar]
  81. Grzimek, B. Grzimeks Tierleben: Enzyklopädie des Tierreichs—Säugetiere 3; Neue Schweizer Bibliothek: Zurich, Switzerland, 1973. [Google Scholar]
  82. Bürger, O. Aus der Wildnis des Huemul. Erlebnisse und Abenteuer unter den Kolonisten und Indianern Chiles; Verlag Deutsche Buchwerkstätten: Dresden, Germany, 1924. [Google Scholar]
  83. Goni, R.A.; Belardi, J.B.; Re, A.; Nuevo Delaunay, A.; Molinari, R.L.; Ferraro, L. Los grabados de la meseta del lago Strobel (Patagonia argentina) desde una perspectiva regional. In Actas del Primer Simposio Nacional de Arte Rupestre; Cusco Nov. 2004; Institut Francais d’Etudes Andines: Lima, Peru, 2007; pp. 427–438. [Google Scholar]
  84. Risenhoover, K.L.; Bailey, J.A.; Wakelyn, L.A. Assessing the Rocky Mountain Bighorn Sheep Management Problem. Wildl. Soc. Bull. 1988, 16, 346–352. [Google Scholar]
  85. Nagaike, T. Bark Stripping by Deer Was More Intensive on New Recruits than on Advanced Regenerants in a Subalpine Forest. Forests 2020, 11, 490. [Google Scholar] [CrossRef]
  86. Diaz, P.; Marqués, B.I.; Vila, A.R. Seasonal habitat use and selection of the endangered huemul deer (Hippocamelus bisulcus) in Patagonian Andes. Mammalia 2013, 77, 371–380. [Google Scholar] [CrossRef]
  87. Gill, R.; Saucedo, C.; Aldridge, D.; Morgan, G. Ranging behavior of huemul in relation to habitat and landscape. J. Zool. 2008, 274, 254–260. [Google Scholar] [CrossRef]
  88. Guineo, O.; Guineo Garay, R.; Garay, G. Conociendo al Huemul de Torres del Paine; La Prensa Austral: Punta Arenas, Chile, 2008. [Google Scholar]
  89. Garay, G.; Ortega, I.M.; Guineo, O. Social ecology of the huemul at Torres Del Paine National Park, Chile. Anal. Inst. Patagon. 2016, 44, 25–38. [Google Scholar] [CrossRef] [Green Version]
  90. Webb, S.L.; Dzialak, M.R.; Houchen, D.; Kosciuch, K.L.; Winstead, J.B. Spatial ecology of female Mule deer in an area proposed for wind energy development. West. N. Am. Nat. 2013, 73, 347–356. [Google Scholar] [CrossRef] [Green Version]
  91. Rau, J.A. Crecimiento poblacional de huemules del sur nativos y reintroducidos en la zona austral de Chile. In 4ta Reunión Chileno-Argentina sobre Estrategias de Conservación del Huemul; Acosta-Jamett, G., Ed.; CONAF and CODEFF: Santiago, Chile, 2003; pp. 43–45. [Google Scholar]
  92. Flueck, W.T.; Smith-Flueck, J.M. Troubling disease syndrome in endangered live Patagonian huemul deer (Hippocamelus bisulcus) from the Protected Park Shoonem: Unusually high prevalence of osteopathology. BMC Res. Notes 2017, 10, 739. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  93. Clapperton, C.M. The glaciation of the Andes. Quat. Sci. Rev. 1983, 2, 83–155. [Google Scholar] [CrossRef]
  94. Flueck, W.T.; Smith-Flueck, J.M. Huemul heresies: Beliefs in search of supporting data. 2. Biological and ecological considerations. Anim. Prod. Sci. 2012, 52, 694–706. [Google Scholar] [CrossRef]
  95. Flueck, W.T. Elusive cranial lesions severely afflicting young endangered Patagonian huemul deer. BMC Res. Notes 2018, 11, 638. [Google Scholar] [CrossRef]
  96. Smith-Flueck, J.M.; Barrio, J.; Ferreyra, N.; Nuñez, A.; Tomas, N.; Guzman, J.; Flueck, W.T.; Hinojosa, A.; Vidal, F.; Garay, G.; et al. Advances in ecology and conservation of Hippocamelus species in South America. Anim. Prod. Sci. 2011, 51, 378–383. [Google Scholar] [CrossRef] [Green Version]
  97. Franklin, W.L. Biology, ecology, and relationship to man of the South American camelids. In Mammalian Biology in South America; Mares, M.A., Genoways, H.H., Eds.; Pymatuning Laboratory of Ecology, University Pittsburgh: Linesville, PA, USA, 1982; Volume 6, pp. 457–489. [Google Scholar]
  98. Mueller, T.; Olson, K.A.; Dressler, G.; Leimgruber, P.; Fuller, T.K.; Nicolson, C.; Novaro, A.J.; Bolgeri, M.J.; Wattles, D.; DeStefano, S.; et al. How landscape dynamics link individual- to population-level movement patterns: A multispecies comparison of ungulate relocation data. Glob. Ecol. Biogeogr. 2011, 20, 683–694. [Google Scholar] [CrossRef]
  99. Gelin, M.L.; Branch, L.C.; Thornton, D.H.; Novaro, A.J.; Gould, M.J.; Caragiulo, A. Response of pumas (Puma concolor) to migration of their primary prey in Patagonia. PLoS ONE 2017, 12, e0188877. [Google Scholar] [CrossRef] [Green Version]
  100. Franklin, W.L.; Bass, F.; Bonacic, C.F.; Cunazza, C.; Soto, N. Management of Patagonian guanaco in the grazing agroecosystem of southern Chile. Wildl. Soc. Bull. 1997, 25, 65–73. [Google Scholar]
  101. Bonavia, D. The South American Camelids: An Expanded and Corrected Edition; Monograph 64; Cotsen Institute of Archaeology Press, UCLA: Los Angeles, CA, USA, 2009; Available online: https://escholarship.org/uc/item/7xs9j2zs (accessed on 4 April 2022).
  102. Schroeder, N.M.; Matteucci, S.D.; Moreno, P.G.; Gregorio, P.; Ovejero, R.; Taraborelli, P.; Carmanchahi, P.D. Spatial and Seasonal Dynamic of Abundance and Distribution of Guanaco and Livestock: Insights from Using Density Surface and Null Models. PLoS ONE 2014, 9, e85960. [Google Scholar]
  103. Ladio, A.H.; Lozada, M. Summer cattle transhumance and wild edible plant gathering in a Mapuche community of northwestern Patagonia. Hum. Ecol. 2004, 32, 225–240. [Google Scholar] [CrossRef]
  104. Poole, K.G.; Mowat, G. Winter habitat relationships of deer and elk in the temperate interior mountains of British Columbia. Wildl. Soc. Bull. 2005, 33, 1288–1302. [Google Scholar] [CrossRef]
  105. Myburgh, J.; McGowan, K.; Davis, A. Veterinary Geology. In Practical Applications of Medical Geology; Siegel, M., Selinus, O., Finkelman, R., Eds.; Springer: Cham, Switzerland, 2021. [Google Scholar]
  106. Flueck, W.T. Consideraciones acerca de la calidad nutritiva de hábitat, hábitat óptimo, y evaluación de hábitat para huemul. In 4ta Reunión Chileno-Argentina Sobre Estrategias de Conservación del Huemul; Acosta-Jamett, G., Ed.; CONAF and CODEFF: Santiago, Chile, 2003; pp. 30–34. [Google Scholar]
  107. Landete-Castillejos, T.; Molina-Quilez, I.; Estevez, J.A.; Ceacero, F.; Garcia, A.J.; Gallego, L. Alternative hypothesis for the origin of osteoporosis: The role of Mn. Front. Biosci. Elite Ed. 2012, 4, 1385–1390. [Google Scholar] [CrossRef]
  108. Gambín, P.; Serrano, M.P.; Gallego, L.; García, A.; Cappelli, J.; Ceacero, F.; Landete-Castillejos, T. Does Cu supplementation affect the mechanical and structural properties and mineral content of red deer antler bone tissue? Animal 2017, 11, 1312–1320. [Google Scholar] [CrossRef] [Green Version]
  109. Bartlam-Brooks, H.L.; Beck, P.S.; Bohrer, G.; Harris, S. In search of greener pastures: Using satellite images to predict the effects of environmental change on zebra migration. J. Geophys. Res. Biogeosci. 2013, 118, 1427–1437. [Google Scholar] [CrossRef] [Green Version]
  110. Teitelbaum, C.S.; Fagan, W.F.; Fleming, C.H.; Dressler, G.; Calabrese, J.M.; Leimgruber, P.; Mueller, T. How far to go? Determinants of migration distance in land mammals. Ecol. Lett. 2015, 18, 545–552. [Google Scholar] [CrossRef]
  111. Monsarrat, S.; Jarvie, S.; Svenning, J.C. Anthropocene refugia: Integrating history and predictive modelling to assess the space available for biodiversity in a human-dominated world. Philos. Trans. R. Soc. B 2019, 374, 20190219. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  112. Soga, M.; Gaston, K.J. Shifting baseline syndrome: Causes, consequences and implications. Front. Ecol. Environ. 2018, 16, 222–230. [Google Scholar] [CrossRef] [Green Version]
  113. Bar-Ilan, J.; Halevi, G. Retracted articles—The scientific version of fake news. In The Psychology of Fake News: Accepting, Sharing, and Correcting Misinformation; Greifeneder, R., Jaffé, M.E., Newman, E.J., Schwarz, N., Eds.; Taylor & Francis Group: New York, NY, USA, 2021. [Google Scholar]
  114. Novillo, A.; Ovejero, A.J.A.; Cristobal, L.; Ojeda, R.A. Alpine Mammals of South America. In Encyclopedia of the World’s Biomes; Goldstein, M.I., DellaSala, D.A., Eds.; Elsevier: Amsterdam, The Netherlands, 2020; pp. 441–460. [Google Scholar]
  115. Cromsigt, J.P.G.M.; Kerley, G.I.H.; Kowalczyk, R. The difficulty of using species distribution modelling for the conservation of refugee species—The example of European bison. Div. Distrib. 2012, 18, 1253–1257. [Google Scholar] [CrossRef] [Green Version]
  116. Lea, J.M.D.; Kerley, G.I.H.; Hrabar, H.; Barry, T.J.; Shultz, S. Recognition and management of ecological refugees: A case study of the Cape mountain zebra. Biol. Conserv. 2016, 203, 207–215. [Google Scholar] [CrossRef]
  117. Kauffman, M.J.; Cagnacci, F.; Chamaillé-Jammes, S.; Hebblewhite, M.; Hopcraft, J.G.C.; Merkle, J.A.; Mueller, T.; Mysterud, A.; Peters, W.; Roettger, C.; et al. Mapping out a future for ungulate migrations. Limited mapping of migrations hampers conservation. Science 2021, 372, 566–569. [Google Scholar] [CrossRef]
  118. Lopez Alfonsin, M.A.; Bucetto, M.S. Endangered species and mechanisms for the recovery and conservation of biodiversity: A study on the viability of mechanisms and bureaucratic obstacles. LEX 2019, 17, 297–324. [Google Scholar]
  119. Doughty, C.E.; Roman, J.; Faurby, S.; Wolf, A.; Haque, A.; Bakker, E.S.; Malhi, Y.; Dunning, J.B.; Svenning, J.C. Global nutrient transport in a world of giants. Proc. Nat. Acad. Sci. USA 2018, 113, 868–873. [Google Scholar] [CrossRef] [Green Version]
  120. Monteith, K.L.; Bleich, V.C.; Stephenson, T.R.; Pierce, B.M.; Conner, M.M.; Kie, J.G.; Bowyer, R.T. Life-history characteristics of mule deer: Effects of nutrition in a variable environment. Wildl. Monogr. 2014, 186, 1–62. [Google Scholar] [CrossRef]
  121. Allred, W.J. Re-establishment of seasonal elk migration through transplanting. N. Am. Wildl. Conf. 1950, 15, 597–611. [Google Scholar]
  122. Stüwe, M.; Nievergelt, B. Recovery of Alpine ibex from near extinction: The result of effective protection, captive breeding, and reintroductions. Appl. Anim. Behav. Sci. 1991, 29, 379–387. [Google Scholar] [CrossRef]
  123. Gruell, G.E.; Papez, N.J. Movements of mule deer in northeastern Nevada. J. Wildl. Manag. 1963, 27, 414–427. [Google Scholar] [CrossRef]
  124. Haywood, D.D.; Brown, R.L.; Smith, R.H.; McCulloch, C.Y. Migration Patterns and Habitat Utilization by Kaibab Mule Deer; Arizona Game & Fish Department Research Branch: Phoenix, AZ, USA, 1987; pp. 1–44. [Google Scholar]
  125. Brown, C.G. Movement and migration patterns of mule deer in southeastern Idaho. J. Wildl. Manag. 1992, 56, 246–253. [Google Scholar] [CrossRef]
  126. Brooks, T.M.; Pimm, S.L.; Akçakaya, H.R.; Buchanan, G.M.; Buthchart, S.H.M.; Foden, W.; Hilton-Taylor, C.; Hoffmann, M.; Jenkins, C.N.; Joppa, L.; et al. Measuring terrestrial area of habitat (AOH) and its utility for the IUCN Red List. Trends Ecol. Evol. 2019, 24, 977–986. [Google Scholar] [CrossRef] [Green Version]
  127. Kerley, G.I.H.; te Beest, M.; Cromsigt, J.P.G.M.; Pauly, D.; Shultz, S. The Protected Area Paradox and refugee species: The giant panda and baselines shifted towards conserving species in marginal habitats. Conserv. Sci. Pract. 2020, 2, e203. [Google Scholar] [CrossRef]
  128. Adams, A.W. Migration. In Elk of North America. Ecology and Management; Thomas, J.W., Toweill, D.E., Eds.; Stackpole Books: Harrisburg, PA, USA, 1982; pp. 301–321. [Google Scholar]
  129. Agassiz, L. The 1871-1872 Hassler expedition. In Museum of Comparative Zoology; Occurrence Dataset, Version 162.229; Morris, P.J., Ed.; Harvard University: Cambridge, MA, USA, 1872. [Google Scholar]
  130. Aldenderfer, M.S. Montane Foragers: Asana and the South-Central Andean Archaic; University of Iowa Press: Iowa, IA, USA, 1998. [Google Scholar]
  131. Andersen, R. Habitat deterioration and the migratory behavior of moose (Alces alces L.) in Norway. J. Appl. Ecol. 1991, 28, 102–108. [Google Scholar] [CrossRef]
  132. Agencia CTyS-UNLaM/DICYT. Descubren los Restos Fósiles de Seis Ciervos Prehistóricos. 2021. Available online: www.dicyt.com (accessed on 7 October 2021).
  133. Armesto, J.J.; Manuschevich, D.; Mora, A.; Smith-Ramireza, C.; Rozzi, R.; Abarzúa, A.M.; Marquet, P.A. From the Holocene to the Anthropocene: A historical framework for land cover change in southwestern South America in the past 15,000 years. Land Use Policy 2010, 27, 148–160. [Google Scholar] [CrossRef]
  134. Aschero, C.A. Las escenas de caza en Cueva de las Manos: Una perspectiva regional (Santa Cruz, Argentina). In IFRAO Congress—Symposium: Pleistocene Art of the Americas (Pre-Acts) Austral; Talleres Gráficos Guillermo Kraft Ltda: Buenos Aires, Argentina, 2010. [Google Scholar]
  135. Avital, E.; Jablonka, E. Animal Traditions; Cambridge University Press: Cambridge, UK, 2000. [Google Scholar]
  136. Battin, J. When good animals love bad habitats: Ecological traps and the conservation of animal populations. Cons. Biol. 2004, 18, 1482–1491. [Google Scholar] [CrossRef]
  137. Baumann, M.; Babota, C.; Schibler, J. Native or naturalized? Validating alpine chamois habitat models with archaeozoological data. Ecol. Appl. 2005, 15, 1096–1110. [Google Scholar] [CrossRef] [Green Version]
  138. Behm, E. Reise im südwestlichen Patagonien von J.T. Rogers und E. Ibar, 1877, nebst den Tagebüchern von A. de Viedma 1782 und J.H. Gardiner 1867. Petermanns Geogr. Mitteilungen 1880, 26, 47–64. [Google Scholar]
  139. Berger, J.; Wangchuk, T.; Briceno, C.; Vila, A.; Lambert, J.E. Disassembled Food Webs and Messy Projections: Modern Ungulate Communities in the Face of Unabating Human Population Growth. Front. Ecol. Evol. 2020, 8, 128. [Google Scholar] [CrossRef]
  140. Berger, J.; Steven, L.; Cain, S.L.; Berger, K.M. Connecting the dots: An invariant migration corridor links the Holocene to the present. Biol. Lett. 2006, 2, 528–531. [Google Scholar] [CrossRef]
  141. Brandborg, S.M. Life History and Management of the Mountain Goat in Idaho; Wildlife Bulletin No. 2; Department of Fish and Game: Boise, ID, USA, 1955; 142p.
  142. Bubenik, G.A.; Bubenik, A.B. Horns, Pronghorns, and Antlers; Springer: New York, NY, USA, 1990. [Google Scholar]
  143. Burmeister, C.V. Memoria Sobre el Territorio de Santa Cruz; Imprenta La Nación, Ministerio de Agricultura de la Republica Argentina: Buenos Aires, Argentina, 1901. [Google Scholar]
  144. Burmeister, H. The huemul. Nature 1873, 9, 82. [Google Scholar] [CrossRef]
  145. Burmeister, C.V. Nuevos datos sobre el territorio Patagonico de Santa Cruz. Rev. Museo Plata 1893, 4, 227–256, 338–352. [Google Scholar]
  146. Cabrera, A.; Yepes, J. Mamiferos Sudamericanos, 1st ed.; Compañia Argentina de Editores: Buenos Aires, Argentina, 1940; 370p. [Google Scholar]
  147. Carballo Marina, F.; Manzi, L.M.; Campan, P.A.; Belardi, J.B.; Tiberi, P.; Manero, A.; Saenz, J.L. Distribución del registro arqueológico en la cuenca del rio Gallegos (Santa Cruz): Linea de base y aporte a la preservación del patrimonio. In Arqueologia del Extremo sur del Continente Americano; Borrero, L.A., Franco, N., Eds.; Editorial Dunken: Buenos Aires, Argentina, 2008; pp. 175–225. [Google Scholar]
  148. Cardich, A.; Miotti, L. Recursos Faunisticos en la Economia de los Cazadores-Recolectores de Los Toldos (Provincia de Santa Cruz). Relaciones Soc. Argent. Antropol. 1983, 16, 269–273. [Google Scholar]
  149. Castellanos, A. Paleontología estratigráfica de los sedimentos neógenos de la Provincia de Córdoba. Publ Inst. Fisiogr. Geol. 1944, 23, 1–47. [Google Scholar]
  150. Church, G.E. A traveller in Patagonia. Nature 1903, 67, 321–322. [Google Scholar] [CrossRef]
  151. Clapperton, C.M. Nature of environmental changes in South America at the Last Glacial Maximum. Palaeogeogr. Paleoclimatol. Palaeoecol. 1993, 101, 189–208. [Google Scholar] [CrossRef]
  152. Claraz, M.G. Sur l’Equus bisulcus, de Molina. Revue et Magasin de Zoologie Pure et Apliquee 1864, 241–248. [Google Scholar]
  153. Coltorti, M.; Abbazzi, L.; Ferretti, M.P.; Iacumin, P.; Paredes Rios, F.; Pellegrini, M.; Pieruccini, P.; Rustioni, M.; Tito, G.; Rook, L. Last Glacial mammals in South America: A new scenario from the Tarija Basin (Bolivia). Naturwissenschaften 2007, 94, 288–299. [Google Scholar] [CrossRef]
  154. Conway, W. Act III in Patagonia: People and Wildlife; Island Press: Washington, DC, USA, 2005. [Google Scholar]
  155. Cox, G.E. Viaje a las Regiones Septentrionales de la Patagonia: 1862–1863; Imprenta Nacional: Santiago, Chile, 1863; 266p. [Google Scholar]
  156. Cruz, I.; Munoz, A.S.; Caracotche, M. A huemul (Hippocamelus bisulcus) antler artefact in archaeological deposits of the Atlantic coast. Implications for human mobility and species distribution. Magallania 2010, 38, 287–294. [Google Scholar] [CrossRef] [Green Version]
  157. da Silva, F.M.; da Silva Alves, R.; Franca Barreto, A.M.; Bezerra de Sá, F.; Borges Lins e Silva, A.C. A megafauna pleistocnica do estado de Pernambuco. Estudos Geologicos 2006, 16, 55–66. [Google Scholar]
  158. Dabbene, R. Sobre la existencia del huemul de Bolivia y Perú, Odocoileus (Hippocamelus) antisensis (Orb.) y del avestruz petiso, Rhea darwini Gould en el N.W. de la República Argentina. Anales del Museo Nacional Buenos Aires 1911, 14, 293–307. [Google Scholar]
  159. Dawilov. Coihue; Gmo. van Woerden & Cia: Buenos Aires, Argentina, 1926; 103p. [Google Scholar]
  160. de la Cruz, L. Descripción de la naturaleza de los terrenos que se comprenden en los Andes, poseídos por los Peguenches; y los demás espacios hasta el rio de Chadileubu. In Colección de Obras y Documentos Relativos a la Historia Antigua y Moderna de las Provincias del Río de la Plata; de Angelis, P., Ed.; Imprenta del Estado: Buenos Aires, Argentina, 1836; pp. 1–67. [Google Scholar]
  161. De Agostini, A.M. Andes Patagónicos: Viajes de Exploración a la Cordillera Patagónica, 1era Versión; Imprenta Gotelli: Buenos Aires, Argentina, 1941. [Google Scholar]
  162. De Agostini, A.M. Andes Patagónicos: Viajes de Exploración a la Cordillera Patagónica Austral; Talleres Gráficos Guillermo Kraft Ltda: Buenos Aires, Argentina, 1945; Volume 1, pp. 1–409. [Google Scholar]
  163. Diaz, N.I. Antecedentes sobre la historia natural de la taruca (Hippocamelus antisensis) y su rol en la economia Andina. Chungara 1995, 27, 45–55. [Google Scholar]
  164. Diaz, N.I. Changes in the range distribution of Hippocamelus bisulcus in Patagonia. Z. Säugetierkunde 1993, 58, 344–351. [Google Scholar]
  165. Díaz, N.I. The huemul (Hippocamelus bisulcus Molina, 1782): A historical perspective. In The Patagonian Huemul, a Mysterious Deer on the Brink of Extinction; Díaz, N.I., Smith-Flueck, J., Eds.; L.O.L.A.: Buenos Aires, Argentina, 2000; pp. 1–31. [Google Scholar]
  166. Díaz, N.I.; Prieto, A.; Bahamonde, G. Guanacos tímidos, huemules confiados: El límite occidental de los cazadores terrestres australes. Magallania 2007, 35, 133–138. [Google Scholar] [CrossRef] [Green Version]
  167. Eastman, C.R. Beginnings of American natural history. Amer. Museum J. 1915, 15, 349–355. [Google Scholar]
  168. Eisenberg, J.F. The contemporary Cervidae of Central and South America. In Antelopes, Deer, and Relatives; Vrba, E.S., Schaller, G.B., Eds.; Yale University Press: New York, NY, USA, 2000; pp. 189–202. [Google Scholar]
  169. Falkner, T. A Description of Patagonia and the Adjoining Parts of South America: Containing an Account of the Soil, Produce, the Religion, Government, and Some Particulars Relating to Falkland Islands; Hereford: London, UK, 1774; 144p. [Google Scholar]
  170. Fernandez, P.M.; Cruz, I.; Bautista Belardi, J.; de Nigris, M.; Muñoz, S. La explotación del huemul (Hippocamelus bisulcus, Molina 1782) en la Patagonia a lo largo del holoceno. Magallania 2016, 44, 187–209. [Google Scholar] [CrossRef] [Green Version]
  171. Fernández, O.A.; Busso, C.A. Arid and semi-arid rangelands: Two thirds of Argentina. In Proceedings from an International Workshop in Iceland; Arnalds, O., Archer, S., Eds.; Rala Report no. 200; Agricultural Research Institute: Reykjavik, Iceland, 1997; pp. 41–60. [Google Scholar]
  172. Fielder, P.C. Implications of selenium levels in Washington mountain goats, mule deer, and Rocky Mountain elk. Northwest Sci. 1986, 60, 15–20. [Google Scholar]
  173. Flint, R.F.; Fidalgo, F. Glacial Drift in the Eastern Argentine Andes between Latitude 41°10′ S. and Latitude 43°10′ S. Bull. Geol. Soc. America 1969, 80, 1043–1052. [Google Scholar] [CrossRef]
  174. Flueck, W.T. Spatio-temporal movements among red deer males, Cervus elaphus, introduced to Patagonia. In XXVIIth Congress of the International Union of Game Biologists; Pohlmeyer, K., Ed.; Hannover DSV-Verlag: Hamburg, Germany, 2005; pp. 330–332. [Google Scholar]
  175. Flueck, W.T. Exotic deer in southern Latin America: What do we know about impacts on native deer and on ecosystems? Biol. Invasions 2010, 12, 1909–1922. [Google Scholar] [CrossRef] [Green Version]
  176. Flueck, W.T. Osteopathology and selenium deficiency co-occurring in a population of endangered Patagonian huemul (Hippocamelus bisulcus). BMC Res. Notes 2015, 8, 330. [Google Scholar] [CrossRef] [PubMed] [Green Version]
  177. Flueck, W.T.; Smith-Flueck, J.M. Über das in Argentinien angesiedelte Rotwild (Cervus elaphus L., 1758): Verbreitung und Tendenzen. Zeits. Jagdwiss. 1993, 39, 153–160. [Google Scholar] [CrossRef]
  178. Flueck, W.T.; Smith-Flueck, J.M. Predicaments of endangered huemul deer, Hippocamelus bisulcus, in Argentina: A review. Europ. J. Wildl. Res. 2006, 52, 69–80. [Google Scholar] [CrossRef]
  179. Flueck, W.T.; Smith-Flueck, J.M. Huemul heresies: Beliefs in search of supporting data. 1. Historical and zooarcheological considerations. Anim. Prod. Sci. 2012, 52, 685–693. [Google Scholar] [CrossRef] [Green Version]
  180. Flueck, W.T.; Smith-Flueck, J.M. Radio marking the first group of endangered Patagonian huemul deer in Argentina. J. Neotrop. Mammal. 2018, 25, 461–465. [Google Scholar] [CrossRef]
  181. Flueck, W.T.; Smith-Flueck, J.M.; Mincher, B.J.; Winkel, L.H.E. Soil selenium levels corroborate direct evidence of selenium deficiency in endangered Patagonian huemul deer (Hippocamelus bisulcus). In Proceedings of the 8th International Deer Biology Congress; Ma, J., Zhang, M., Halbrook, R., Liu, B., Zhang, W., Eds.; Northeast Forestry University: Harbin, China, 2014; pp. 52–53. [Google Scholar]
  182. Frailey, D.; Campbell, K.E.; Wolff, R.G. Additions to the knowledge of Hippocamelus, Ctenomys, and Myocastor from the middle Pleistocene of the Tarija basin, Bolivia. Occas. Papers Museum Nat. Hist. Univ. Kansas 1980, 85, 1–14. [Google Scholar]
  183. Gay, C. Historia Fisica y Politica de Chile: Zoologia; Museo de Historia Natural de Santiago: Santiago, Chile, 1847; 495p. [Google Scholar]
  184. Gazzolo, C. Botanical composition of taruka (Hippocamelus antisensis) diet during rainy season in Huascaran national park, Peru. In Advances in Deer Biology; Bartos, L., Dusek, A., Kotrba, R., Bartosova, J., Eds.; Research Institute of Animal Production: Praha, Czech Republic, 2006; p. 216. [Google Scholar]
  185. Geist, V. Deer of the World; Stackpole Books: Pennsylvania, PA, USA, 1998; 421p. [Google Scholar]
  186. Gigoux, E.E. El huemul. Rev. Chilena Hist. Nat. 1929, 23, 573–582. [Google Scholar]
  187. Goni, R.A. Arqueologia de momentos tardios en el Parque Nacional Perito Moreno (Santa Cruz, Argentina). In Precirculados del IX Congreso Nacional de Arqueologia Argentina; Universidad de Buenos Aires: Buenos Aires, Argentina, 1988; pp. 140–151. [Google Scholar]
  188. Gonzalez, V.; Tapia, V. Manual bovino de carne. Boletín INIA 2017, 4, 1–173. [Google Scholar]
  189. Grosse, A. El huemul—Ciervo de los Andes y emblema del escudo Chileno. Condor 1949, 12, 10–12. [Google Scholar]
  190. Guérin, C.; Faure, M. The Cervidae (Mammalia, Artiodactyla) of the Upper Pleistocene/Lower Holocene deposits of the Serra da Capivara National Park Region (Piauý, Brazil). Geobios 2009, 42, 169–195. [Google Scholar] [CrossRef]
  191. Günther, A. Comments about Sclater and Cervus chilensis. In Proceedings of the Scientific Meetings of the Zoological Society of London; Messrs. Longmans, Green, Reader and Dyer: London, UK, 1875; pp. 44–46. [Google Scholar]
  192. Hatcher, J.B. Reports of the Princeton University Expeditions to Patagonia, 1896–1899. Vol. I: Narrative of the Expeditions. Geography of Southern Patagonia; E. Schweizerbart’sche Verlagshandlung: Stuttgart, Germany, 1903; 314p. [Google Scholar]
  193. Hauman, L. Étude phytogéographique de la Patagonie. Bulletin de la Société Royale de Botanique de Belgique 1926, 58, 105–179. [Google Scholar]
  194. Hershkovitz, P. The recent mammals of the neotropical region: A zoogeographic and ecological review. In Evolution, Mammals, and Southern Continents; Keast, A., Erk, F.C., Glass, B., Eds.; State University New York Press: New York, NY, USA, 1972; pp. 311–431. [Google Scholar]
  195. Hershkovitz, P. The recent mammals of the Neotropical region: A zoogeographic and ecological review. Quart. Rev. Biol. 1969, 44, 1–70. [Google Scholar] [CrossRef]
  196. Hoffstetter, R. La faune pleistocene de Tarija (Bolivie). Note preliminaire. Bulletin Muséum National d’Histoire Naturelle 1963, 35, 194–203. [Google Scholar]
  197. Honess, R.F.; Frost, N.M. A Wyoming bighorn sheep study. Wyoming Game Fish Depart. Bull. 1942, 1, 1–127. [Google Scholar]
  198. Horkheimer, H. Nahrung und Nahrungsgewinnung im Vorspanischen Peru; Colloquim Verlag: Berlin, Germany, 1960. [Google Scholar]
  199. Housse, P.R. Animales Salvajes de Chile en su Clasificación Moderna: Su Vida y Sus Costumbres; Ediciones de la Universidad de Chile: Santiago, Chile, 1953; 189p. [Google Scholar]
  200. Ibar Bruce, J. Aisen, Hombres y Naturaleza; Imprenta de la Armada: Valparaiso, Chile, 1973; pp. 1–164. [Google Scholar]
  201. Iglesias, R.E. El huemul. Montaña 1965, 7, 26–28. [Google Scholar]
  202. Jakopak, R.P.; LaSharr, T.N.; Dwinnell, S.P.H.; Fralick, G.L.; Monteith, K.L. Rapid acquisition of memory in a complex landscape by a mule deer. Ecology 2019, 100, e02854. [Google Scholar] [CrossRef]
  203. Johnson, L. Informe sobre una prospección arqueológica en magallanes. Ans. Inst. Pat. Punta Arenas 1976, 7, 87–94. [Google Scholar]
  204. Kolliker Frers, A. Das Waidwerk und die autochthonen Cerviden in Argentinien. In Parque Diana; Vogel, C.A., Ed.; Stefan Schwarz Verlag: München, Germany, 1969; pp. 25–31. [Google Scholar]
  205. Koprowski, J.L.; Krausman, P.R. International Wildlife Management: Conservation Challenges in a Changing World; Johns Hopkins University Press: Baltimore, MD, USA, 2019; 248p. [Google Scholar]
  206. Krieg, H. Biologische Reisestudien in Südamerika. V. Die chilenischen Hirsche. Zeits. Morphol. Ökol. Tiere 1925, 4, 585–597. [Google Scholar] [CrossRef]
  207. Krieg, H. Als Zoologe in Steppen und Wäldern Patagoniens; Bayerischer Landwirtschaftsverlag: Muenchen, Germany, 1940; 197p. [Google Scholar]
  208. Lacroix, F. Historia de la Patagonia, Tierra de Fuego, e Islas Malvinas; Imprenta del Liberal Barcelones: Barcelona, Spain, 1841. [Google Scholar]
  209. Laliberte, A.S.; Ripple, W.J. Range contractions of North American carnivores and ungulates. BioScience 2004, 54, 123–138. [Google Scholar] [CrossRef] [Green Version]
  210. Laming-Emperaire, A.; Lavallée, D.; Humbert, R. Le site de Marazzi en Terre de Feu. Objets et Mondes 1972, 12, 225–244. [Google Scholar]
  211. Latcham, R.E. Expedicion cientifica Macqueen al Aysen. Boletin del Museo Nacional 1935, 14, 7–31. [Google Scholar]
  212. Leopold, A.S.; Cain, S.A.; Cottam, C.; Gabrielson, I.N.; Kimball, T.L. Wildlife Management in the National Parks. Am. For. 1963, 4, 32–35, 61–63. [Google Scholar]
  213. Liebermann, J. Sobre la historia natural del huemul. Anal. Acad. Argent. Geogr. 1962, 6, 157–168. [Google Scholar]
  214. Lingle, S.; Wilson, W.F. Detection and avoidance of predators in white-tailed deer (Odocoileus virginianus) and mule deer (O. hemionus). Ethology 2001, 107, 125–147. [Google Scholar] [CrossRef] [Green Version]
  215. Lista, R. La Tierra del Fuego y sus habitantes. Boletin del Instituto Geografico Argentino 1881, 2, 109–114. [Google Scholar]
  216. Lydekker, R. The Deer of all Lands: A History of the Family Cervidae, Living and Extinct; R. Ward: London, UK, 1898; pp. 1–329. [Google Scholar]
  217. Lyman, R.L. Paleozoology in the service of conservation biology. Evol. Anthropol. 2006, 15, 11–19. [Google Scholar] [CrossRef]
  218. Lyras, G.A.; Giannakopoulou, A.; Lillis, T.; van der Geer, A.A.E. Paradise lost: Evidence for a devastating metabolic bone disease in an insular Pleistocene deer. Intern. J. Paleopathol. 2019, 24, 213–226. [Google Scholar] [CrossRef] [PubMed]
  219. MacDouall, J. Narratives of a Voyage to Patagonia and Terra del Fuego; Renshaw and Rush: London, UK, 1833; 320p. [Google Scholar]
  220. Machon, J.F.; Juarez, F.N. Patagonia 1892: Diario del Explorador Suizo Dr. Francisco Machón; Editorial Dunken: Buenos Aires, Argentina, 2013. [Google Scholar]
  221. Magalhaes, R.M.; Mello, M.G.; Bergqvist, L.P. Os cervidas pleistocenicos da regiao nordeste Brasileira. Anais da Academia Brasileira de Ciencias 1992, 64, 149–154. [Google Scholar]
  222. Magne de la Croix, P. El huemul. Caras y Caretas 1937, 40, 117. [Google Scholar]
  223. Mansur, M.E.; Piqué, R. Between the Forest and the Sea: Hunter-Gatherer Occupations in the Subantarctic Forests in Tierra del Fuego, Argentina. Arctic Anthropol. 2009, 46, 144–157. [Google Scholar] [CrossRef]
  224. Markgraf, V.; Kenny, R. Character of rapid vegetation and climate change during the late-Glacial in southernmost South America. In Past and Future Rapid Environmental Changes: Spatial and Evolutionary Responses to terrestrial Biota; Huntley, B., Ed.; Springer: Berlin, Germany, 1997; pp. 81–90. [Google Scholar]
  225. Marshall, L.G. Land Mammals and the Great American Interchange. Am. Sci. 1988, 76, 380–388. [Google Scholar]
  226. Massara Paletto, V.; Buono, G. Métodos de Evaluación de Pastizales en Patagonia Sur; Centro Regional Patagonia, Ediciones INTA: Buenos Aires, Argentina, 2020; 288p. [Google Scholar]
  227. Massone, M. Los paraderos tehuelches y proto-tehuelches en la costa del Estrecho de Magallanes. Anales del Instituto de la Patagonia 1984, 15, 27–42. [Google Scholar]
  228. McClure, M.F.; Bissonette, J.A.; Conover, M.R. Migratory strategies, fawn recruitment, and winter habitat use by urban and rural mule deer (Odocoileus hemionus). Europ. J. Wildl. Res. 2005, 51, 170–177. [Google Scholar] [CrossRef]
  229. Miller, S.; Rottman, J.; Taber, R.D. Dwindling and endangered ungulates of Chile: Vicugna, lama, Hippocamelus, and Pudu. Trans. N. Am. Wildl. Natural Res. Conf. 1973, 38, 55–67. [Google Scholar]
  230. Mincher, B.J.; Mionczynski, J.; Hnilicka, P.P.; Ball, R.D.; Houghton, T.X. Some aspects of geophagia in Wyoming big-horn sheep (Ovis canadensis). Europ. J. Wildl. Res. 2008, 54, 192–198. [Google Scholar] [CrossRef]
  231. Molina, J.I. The Geographical, Natural, and Civil History of Chili; Longman, Hurst, Rees, and Orme: London, UK, 1809; Volume 1. [Google Scholar]
  232. Moreira-Arce, D.; Pefiaranda, D.A.; Lopéz, R.; Stipicic, G.J.; Hidalgo-Hermoso, E.; Simonetti, J.A. Observations of a coastal population of huemul, Hippocamelus bisulcus (Artiodactyla: Cervidae) in Riesco lsland, Magallanes Region, Chile: A conservation opportunity. Mammalia 2021, 85, 291–295. [Google Scholar] [CrossRef]
  233. Morejohn, G.V.; Dailey, D.C. The identity and postcranial osteology of Odocoileus lucasi (Hay) 1927. Sierra Coll. Nat. Hist. Museum Bull. 2004, 1, 1–54. [Google Scholar]
  234. Moreno, F.P. Explorations in Patagonia. Geogr. J. 1899, 14, 241–269. [Google Scholar] [CrossRef] [Green Version]
  235. Moreno, P.I.; Villagran, C.; Marquet, P.A.; Marshall, L.G. Quaternary paleobiogeography of northern and central Chile. Rev. Chilena Hist. Nat. 1994, 67, 487–502. [Google Scholar]
  236. Moser, C.A. The Bighorn Sheep of Colorado: A Review of Colorado’s Bighorn Sheep Studies; Technical Publication No. 10; The Colorado Game and Fish Department: Denver, CO, USA, 1962; 49p. [Google Scholar]
  237. Musters, R.N. A year in Patagonia. J. Royal Geogr. Soc. London 1871, 41, 59–77. [Google Scholar] [CrossRef]
  238. Mysterud, A.; Loe, L.E.; Zimmermann, B.; Bischof, R.; Veiberg, V.; Meisingset, E. Partial migration in expanding red deer populations at northern latitudes—A role for density dependence? Oikos 2011, 120, 1817–1825. [Google Scholar] [CrossRef]
  239. Nelson, M.E.; Mech, L.D. Twenty-year home-range dynamics of a white-tailed deer matriline. Can. J. Zool. 1999, 77, 1128–1135. [Google Scholar] [CrossRef]
  240. Neveu-Lemaire, M.; Grandidier, G. Notes sur les Mammiféres des Hauts Plateaux de l’Amérique du Sud; Imprimerie Nationale: Paris, France, 1911; 127p. [Google Scholar]
  241. Onelli, C. Trepando los Andes; Compania Sud-Americana de Billetes de Banco: Buenos Aires, Argentina, 1904; 297p. [Google Scholar]
  242. Osgood, W.H. The journal of Wilfred Osgood: The Marshall Field Chilean Expedition of 1922-23. In Patterson, B.D. Field Museum of Natural History Bulletin 1983, 54, 28–33. [Google Scholar]
  243. Packard, F.M. An ecological study of the Bighorn sheep in Rocky Mountain National Park, Colorado. J. Mammal. 1946, 27, 3–28. [Google Scholar] [CrossRef]
  244. Paillan, J.T.; Tello, G.E. Los recursos naturales y culturales, 28 de Noviembre, Guer Aaike. In Santa Cruz: Su Importancia Turistica y Patrimonial; Informe Científico Técnico UNPA 4, ICT-UNPA-35-2012; Universidad Nacional de la Patagonia Austral: Rio Turbio, Argentina, 2012. [Google Scholar]
  245. Paula Couto, C. Paleontologia Brasileira (Mamíferos); Instituto Nacional do Livro: Rio de Janeiro, Brasil, 1953. [Google Scholar]
  246. Paula Couto, C. Tratado de Paleomastozoologia; Academia Brasileira de Ciencias: Rio de Janeiro, Brasil, 1979. [Google Scholar]
  247. Pefaur, J.; Hermosilla, W.; DiCastri, F.; Gonzalez, R.; Salinas, F. Estudio preliminar de mamíferos silvestres chilenos: Su distribución, valor económico e importancia zoonótica. Rev. Soc. Med. Vet. 1968, 18, 3–15. [Google Scholar]
  248. Pennant, T. History of Quadrupeds, 3rd ed.; B & J White: London, UK, 1793. [Google Scholar]
  249. Perez, A.E.; Batres, D.A. Los otros cazadores. Explotación de cérvidos en la Localidad Arqueológica Meliquina, Parque Nacional Lanin, República Argentina. In Zooarqueología hoy. Encuentros Hispano-Argentinos; Díez, J.C., Ed.; Universidad de Burgos: Burgos, Spain, 2008; pp. 89–107. [Google Scholar]
  250. Philippi, R.A. Über den Guemul von Molina. Archiv für Naturgeschichte 1857, 23, 135–136. [Google Scholar]
  251. Philippi, R.A. Zoología: Sinonimia del huemul. Anales de la Universidad de Chile 1873, 717–722. [Google Scholar]
  252. Philippi, R.A. El guemul de Chile. Anal. Museo Nac. Chile Primer Seccion Zool. 1892, 2, 1–9. [Google Scholar]
  253. Phoca-Cosmetatou, N. Site function and the ’ibex-site phenomenon’: Myth or reality? Oxford J. Archaeol. 2004, 23, 217–242. [Google Scholar] [CrossRef]
  254. Prichard, H.H. Through the Heart of Patagonia; D. Appleton and Co.: New York, NY, USA, 1902; 346p. [Google Scholar]
  255. Prichard, H.H. Hunting Camps in Wood and Wilderness; William Heinemann: London, UK, 1910; 274p. [Google Scholar]
  256. Prothero, D.R.; Foss, S.E. The Evolution of Artiodactyls; JHU Press: Baltimore, MD, USA, 2007. [Google Scholar]
  257. Pulliam, H.R. Sources, sinks, and population regulation. Am. Naturalist 1988, 132, 652–661. [Google Scholar] [CrossRef]
  258. Rabassa, J.; Coronato, A. Glaciations in Patagonia and Tierra del Fuego during the Ensenadan Stage/Age (Early Pleistocene-earliest Middle Pleistocene). Quaternary Intern. 2009, 210, 18–36. [Google Scholar] [CrossRef]
  259. Rabassa, J.; Coronato, A.; Martínez, O. Late Cenozoic glaciations in Patagonia and Tierra del Fuego: An updated review. Biol. J. Linnean Soc. 2011, 103, 316–335. [Google Scholar] [CrossRef] [Green Version]
  260. Ramirez Morales, F. Apuntes para una historia ecológica de Chile. Cuadernos de Historia 1991, 11, 149–196. [Google Scholar]
  261. Rasmussen, P.C. Geographic variation in morphology and allozymes of south american imperial shags. The Auk 1994, 111, 143–161. [Google Scholar] [CrossRef]
  262. Re, A.; Delaunay, A.N.; Ferraro, L. Grabados en la meseta del lago Strobel (provincia de Santa Cruz, Argentina), el sitio laguna del Faldeo Verde. Relaciones Soc. Argentina Antropol. 2005, 30, 245–256. [Google Scholar]
  263. Reichlen, H. Huemul in Fell’s Cave, Chile: Specimen MNHN-2M-MO-1988-211; Museum National d’Histoire Naturelle: Paris, France, 1959. [Google Scholar]
  264. Ren, J.Z.; Zhou, Z.Y.; Pan, B.; Chen, W. Selenium distribution in four grassland classes of China. In Selenium in Biology and Medicine; Comb, G.F., Spall-holz, J.E., Levander, O.A., Oldfield, J.E., Eds.; AVI Books: New York, NY, USA, 1987; pp. 769–774. [Google Scholar]
  265. Ringuelet, R.A. Temas de Ciencia Naturales; Museo de La Plata: La Plata, Argentina, 1946. [Google Scholar]
  266. Rosas, Y.M.; Peri, P.L.; Herrera, A.H.; Pastore, H.; Pastur, G.M. Modeling of potential habitat suitability of Hippocamelus bisulcus: Effectiveness of a protected areas network in Southern Patagonia. Ecol. Processes 2017, 6. [Google Scholar] [CrossRef]
  267. Roulin, M. Mémoire pour servir a l’histoire du tapir: Et description d’une espece nouvelle (le tapir pinchaque) appartenant aux hautes régions de la Cordillere des Andes. Mémoires des Savans Étrangers 1835, 6, 5–112. [Google Scholar]
  268. Rusconi, C. Animales Extinguidos de Mendoza y de la Argentina; Imprenta Oficial: Mendoza, Argentina, 1967. [Google Scholar]
  269. Ryan, S.J. The role of culture in conservation planning for small or endangered populations. Conserv. Biol. 2006, 20, 1321–1324. [Google Scholar] [CrossRef]
  270. Santos Gollan, J. Contribución al Conocimiento de los Mamíferos del Parque Nacional de Nahuel Huapi; Facultad de Ciencias Exactas y Naturales, Universidad de Buenos Aires: Buenos Aires, Argentina, 1946; 66p. [Google Scholar]
  271. Sawyer, H.; Middleton, A.D.; Hayes, M.M.; Kauffman, M.J.; Monteith, K.L. The extra mile: Ungulate migration distance alters the use of seasonal range and exposure to anthropogenic risk. Ecosphere 2016, 7, e01534. [Google Scholar] [CrossRef]
  272. Sclater, P.L. Remarks on Cervus chilensis and Cervus antisiensis. J. Nat. History Ser. 4 1873, 11, 213–214. [Google Scholar] [CrossRef] [Green Version]
  273. Sclater, P.L. On Cervus chilensis. Proc. Zool. Soc. Lond. 1875, 2, 44–47. [Google Scholar]
  274. Serret, A. Observaciones Preliminares de Huemul, Hippocamelus bisulcus, en el Lago Nansen del Parque Nacional Perito Moreno, Provincia Santa Cruz; Fundacion Vida Silvestre: Buenos Aires, Argentina, 1990; 23p. [Google Scholar]
  275. Sierralta, D. La microhistología de fecas para el estudio de dieta del huemul. In Huemul Ecology Research for Conservation Planning, Darwin Initiative ed.; Darwin Initiative: Cochrane, Chile, 2003. [Google Scholar]
  276. Siewert, C. Un viaje a Patagonia. Boletin Inst. Geogr. Argentino 1896, 17, 363–391. [Google Scholar]
  277. Silveira, M.J. Analisis e Interpretacion de los Restos Faunisticos de la Cueva Grande del Arroyo Feo. Relaciones Soc. Argent. Antropol. 1979, 13, 229–253. [Google Scholar]
  278. Simmonds, P.L. A Dictionary of Useful Animals and Their Products; E. & F.N. Spon: London, UK, 1883; 136p. [Google Scholar]
  279. Sinclaire, C. Daily life among the fishermen of the fog. In Fishermen of the fog: The Changos and Their Ancestors; Museo Chileno de Arte Precolombino: Santiago, Chile, 2009; pp. 41–48. [Google Scholar]
  280. Skottsberg, C. The wilds of Patagonia; Edward Arnold: London, UK, 1911; 336p. [Google Scholar]
  281. Stankowich, T. Tail-Flicking, Tail-Flagging, and Tail Position in Ungulates with Special Reference to Black-Tailed Deer. Ethology 2008, 114, 875–885. [Google Scholar] [CrossRef]
  282. Stankowich, T.; Coss, R.G. Effects of predator behavior and proximity on risk assessment by Columbian black-tailed deer. Behav. Ecol. 2006, 17, 246–254. [Google Scholar] [CrossRef]
  283. Steffen, H. Die Erforschung de Rio Puelo. Petermanns Geographischen Mitteilungen 1895, 41, 190–193. [Google Scholar]
  284. Steffen, H. Die chilenische Aisen Expedition. Verhandlungen der Gesellschaft für Erdkunde zu Berlin 1897, 24, 461–474. [Google Scholar]
  285. Steffen, H. Reisen in den Patagonischen Anden. Verhandlungen der Gesellschaft für Erdkunde zu Berlin 1900, 27, 194–220. [Google Scholar]
  286. Steward, J.H. Handbook of South American Indians. Volume 1. The Marginal Tribes; Smithsonian Institution: Washington, DC, USA, 1946. [Google Scholar]
  287. Stringham, S.F.; Rogers, L.L. Fear of Humans by Bears and Other Animals (Anthropophobia): How Much is Natural? J. Behav. 2017, 2, 1009. [Google Scholar]
  288. Tarifa, T.; Yensen, E. Mammals of Bolivian Polylepis woodlands. Revista Boliviana de Ecología y Conservación Ambiental 2001, 9, 29–44. [Google Scholar]
  289. Tatura, A.; del Valle, R.; Bianchi, M.; Outes, V.; Villarosa, G.; Niegodzisz, J.; Debaene, G. Late Pleistocene palaeolakes in the Andean and Extra-Andean Patagonia at mid-latitudes of South America. Quaternary Intern. 2002, 89, 135–150. [Google Scholar] [CrossRef]
  290. Teta, P.; Rodríguez, D. Mammalogy National Collection (MACNMa); Occurrence Dataset; Museo Argentino de Ciencias Naturales ’Bernardino Rivadavia’ (MACN): Caba, Argentina, 2020. [Google Scholar]
  291. Thirgood, S.J. The effect of sex, season and habitat availability on patterns of habitat use in fallow deer. J. Zool. Lond. 1995, 235, 645–659. [Google Scholar] [CrossRef]
  292. Torrejon, F. Variables geohistoricos en la evolucion del sistema economico Pehuenche durante el periodo colonial. Revista Universum 2001, 16, 219–236. [Google Scholar]
  293. Van Soest, P.J. Nutritional Ecology of the Ruminant; O & B Books, Inc.: Corvallis, OR, USA, 1982; 373p. [Google Scholar]
  294. Via, S.; Gomulkiewicz, R.; DeJong, G.; Scheiner, C.D.; Schlichting, S.M.; Van Tienderen, P.H. Adaptive phenotypic plasticity: Consensus and controversy. TREE 1995, 10, 212–217. [Google Scholar] [CrossRef]
  295. Vidaurre, F.G. Des Herrn Abts Vidaure Kurzgefasste, Geographische, Natürliche und Bürgerliche Geschichte des Königreichs Chile; Carl Ernst Bohn: Hamburg, Germany, 1782; 208p. [Google Scholar]
  296. von Colditz, R. Im Reiche des Kondor; Paul Parey: Berlin, Germany, 1925; 415p. [Google Scholar]
  297. von Thüngen, J.; Lanari, M.R. Profitability of sheep farming and wildlife management in Patagonia. Pastoralism 2010, 1, 274–290. [Google Scholar]
  298. Wagner, J.U. Die Säugthiere in Abbildungen nach der Natur; L.D. Weigel: Leipzig, Germany, 1855. [Google Scholar]
  299. Waterhouse, G.R. Mammalia. In The Zoology of the Voyage of H.M.S. Beagle, Under the Command of Captain Fitzroy, During the years 1832-1836; Smith, Elder and Co.: London, UK, 1839. [Google Scholar]
  300. Webb, S.D. A history of savanna vertebrates in the New World. Part II: South America and the Great Interchange. Annual Rev. Ecol. Evol. System. 1978, 9, 393–426. [Google Scholar] [CrossRef]
  301. Weber, A. Chiloe: Su Estado Actual, su Colonizacion, su Porvenir; Imprenta Mejia: Santiago, Chile, 1903; 194p. [Google Scholar]
  302. Wells, K.W.; Stangl, F.B. Superior size and antler development in populations of white-tailed deer (Odocoileus virginianus) from the North Texas rolling plains. Texas J. Sci. 2003, 55, 337–346. [Google Scholar]
  303. Wolffsohn, J.W. Notas sobre el huemul. Revista Chilena de Historia Natural 1910, 14, 227–234. [Google Scholar]
  304. Yockney, I.J.; Hickling, G.J. Distribution and diet of chamois (Rupicapra rupicapra) in Westland forests, South Island, New Zealand. N. Z. J. Ecol. 2000, 24, 31–38. [Google Scholar]
Figure 1. Huemul are well-known for their use of steppe grassland and lack of fear of humans. (A) Gaucho with knife hunting a huemul, photographed by Onelli in 1904, see Supplementary File S1, (B) huemul walking away after having sniffed the leg of Prichard in 1902, see Supplementary File S1, (C) a more recent close encounter near Cochrane, Chile.
Figure 1. Huemul are well-known for their use of steppe grassland and lack of fear of humans. (A) Gaucho with knife hunting a huemul, photographed by Onelli in 1904, see Supplementary File S1, (B) huemul walking away after having sniffed the leg of Prichard in 1902, see Supplementary File S1, (C) a more recent close encounter near Cochrane, Chile.
Conservation 02 00023 g001
Figure 2. (A) Protected Park Shoonem (only its western part: 44°51′ S, 71°48′ W), owned by the village of Alto Rio Senguer, province of Chubut. The view is of the western portion of Lake La Plata, with the surrounding mountains physically defining the watershed of the river Senguer and serving as a refuge area for huemul. The edge of the lake provides the lowest elevation accessed by huemul during winter, (B) snow conditions frequently result in huemul using open water for their displacements, (C) photos taken in September 2017.
Figure 2. (A) Protected Park Shoonem (only its western part: 44°51′ S, 71°48′ W), owned by the village of Alto Rio Senguer, province of Chubut. The view is of the western portion of Lake La Plata, with the surrounding mountains physically defining the watershed of the river Senguer and serving as a refuge area for huemul. The edge of the lake provides the lowest elevation accessed by huemul during winter, (B) snow conditions frequently result in huemul using open water for their displacements, (C) photos taken in September 2017.
Conservation 02 00023 g002
Figure 3. (A) Annual home ranges of 3 male (red) and 3 female (white) huemul in the Protected Park Shoonem. Precise locations in winter (yellow) and summer (red) of 3 male (B), and of 3 female huemul (C). Arrow indicates one unusual displacement in late winter, which ended in death by starvation.
Figure 3. (A) Annual home ranges of 3 male (red) and 3 female (white) huemul in the Protected Park Shoonem. Precise locations in winter (yellow) and summer (red) of 3 male (B), and of 3 female huemul (C). Arrow indicates one unusual displacement in late winter, which ended in death by starvation.
Conservation 02 00023 g003
Figure 4. (A) Historical and current spatial distribution of huemul in the Protected Park Shoonem. The blue polygons show lakes Fontana and La Plata. The current year-round range (grey polygon outlined in white) is at the upper elevational margin of summer ranges and near the Andean continental divide, compared to historically used areas (indicated by yellow tacks) [47,69,70]. Distances traveled to summer ranges in the past are well within common seasonal movement distances for similar cervids like Odocoileus. (B) Historical distribution (yellow); current distribution (purple: Jimenez et al. [71]), unsuitable areas (internal white zones indicating bare ice fields, rocky slopes without vegetation, or lakes), and historical locations (green dots), based on reports by naturalists, shed antlers, and archeological samples (Supplementary File S1).
Figure 4. (A) Historical and current spatial distribution of huemul in the Protected Park Shoonem. The blue polygons show lakes Fontana and La Plata. The current year-round range (grey polygon outlined in white) is at the upper elevational margin of summer ranges and near the Andean continental divide, compared to historically used areas (indicated by yellow tacks) [47,69,70]. Distances traveled to summer ranges in the past are well within common seasonal movement distances for similar cervids like Odocoileus. (B) Historical distribution (yellow); current distribution (purple: Jimenez et al. [71]), unsuitable areas (internal white zones indicating bare ice fields, rocky slopes without vegetation, or lakes), and historical locations (green dots), based on reports by naturalists, shed antlers, and archeological samples (Supplementary File S1).
Conservation 02 00023 g004
Table 1. Selection of historical reports about huemul: distribution between the Andean Mountains and the Atlantic Ocean, coexistence with guanaco, group sizes, migratory-resident behavior, and hunting.
Table 1. Selection of historical reports about huemul: distribution between the Andean Mountains and the Atlantic Ocean, coexistence with guanaco, group sizes, migratory-resident behavior, and hunting.
Date Author(s)Observations of Presence
1521 and 1598 after Eastman 1915By Atlantic ports of San Julian and Desire, Atlantic
1793 PennantBy Port Desire, Atlantic Coast
1833 MacDouallBy Gregory Bay, open area, 100 km from the nearest forest
1835 RoulinBy Port San Julian, Atlantic Coast
1863 CoxYear-round resident populations on winter ranges, coexisting with guanaco
1864 ClarazMany guanacos coexisting with equally as many huemul in lowlands
1871 MustersHarvested huemul in open, treeless areas
1875 GüntherBetween Andean Mountain foothills and Patagonian mesas, reaching the Atlantic Coast
1880 BehmSeen in open area far from forests while hunting 45 km east of Chilean border
1892 PhilippiLarge groups during seasonal migration to lower areas
1898 LydekkerMigration down to the open grassland plains where they remain during winter
1900 SteffenHis team ate huemul for weeks, working in open grasslands of foothills:
coexistence with guanaco further east
1901 BurmeisterEighteen huemul (2 groups) in open grasslands, 220 km from nearest forest
1902 PrichardGroups with more than 100 huemul coexisting with guanaco during winter in valleys
1903 ChurchGrassland plains were the home of guanaco, huemul and ostriches
1903 HatcherHarvesting huemul in the open 100 km from the nearest forest
1904 AnonymousThe governor hunted huemul far east of continental divide 270 km from forests (includes photo)
1905 OnelliNear Choiquenilahue and between Senguer and Chubut rivers, 120 km from forests
1911 Neveu-Lemaire Reported winter migration down to valleys
1923 OsgoodHarvesting huemul in steppe grasslands far from forest, coexisting with guanaco
1925 von ColditzHarvesting several huemul in steppe grasslands far from forest
1929 GigouxReported seasonal migration and formation of large herds in winter
1936 GiaiReported seasonal migration and formation of 50 or more huemul in winter
1940 KriegYear-round resident huemul in low valleys
1945 AgostiniMany guanacos and equally as many huemul in open grasslands
1949 GrosseSeasonal migrations and large herds in low valleys
1962 LiebermannSeasonal migrations in winter down to protected valleys and foothill areas
1969 Kolliker FrersReported huemul still occurred in Patagonian open grasslands until 1850s
Publisher’s Note: MDPI stays neutral with regard to jurisdictional claims in published maps and institutional affiliations.

Share and Cite

MDPI and ACS Style

Flueck, W.T.; Smith-Flueck, J.A.M.; Escobar, M.E.; Zuliani, M.; Fuchs, B.; Geist, V.; Heffelfinger, J.R.; Black-Decima, P.; Gizejewski, Z.; Vidal, F.; et al. Loss of Migratory Traditions Makes the Endangered Patagonian Huemul Deer a Year-Round Refugee in Its Summer Habitat. Conservation 2022, 2, 322-348. https://doi.org/10.3390/conservation2020023

AMA Style

Flueck WT, Smith-Flueck JAM, Escobar ME, Zuliani M, Fuchs B, Geist V, Heffelfinger JR, Black-Decima P, Gizejewski Z, Vidal F, et al. Loss of Migratory Traditions Makes the Endangered Patagonian Huemul Deer a Year-Round Refugee in Its Summer Habitat. Conservation. 2022; 2(2):322-348. https://doi.org/10.3390/conservation2020023

Chicago/Turabian Style

Flueck, Werner T., Jo Anne M. Smith-Flueck, Miguel E. Escobar, Melina Zuliani, Beat Fuchs, Valerius Geist, James R. Heffelfinger, Patricia Black-Decima, Zygmunt Gizejewski, Fernando Vidal, and et al. 2022. "Loss of Migratory Traditions Makes the Endangered Patagonian Huemul Deer a Year-Round Refugee in Its Summer Habitat" Conservation 2, no. 2: 322-348. https://doi.org/10.3390/conservation2020023

Article Metrics

Back to TopTop