31.10.2012 Views

Habitat type % live fo rb cover - Sevilleta LTER

Habitat type % live fo rb cover - Sevilleta LTER

Habitat type % live fo rb cover - Sevilleta LTER

SHOW MORE
SHOW LESS

Create successful ePaper yourself

Turn your PDF publications into a flip-book with our unique Google optimized e-Paper software.

Consequences of woody plant encroachment <strong>fo</strong>r mammalian predators<br />

Herman H. Shugart<br />

Paolo D’Odorico<br />

R. Michael Erwin<br />

Henry M. Wilbur<br />

Stephen A. Macko<br />

Virginia Ann Seamster<br />

Santa Fe, New Mexico<br />

Bachelor of Science in Biology, University of Virginia, 2005<br />

A Dissertation presented to the Graduate Faculty<br />

of the University of Virginia in Candidacy <strong>fo</strong>r the Degree of<br />

Doctor of Philosophy<br />

Department of Environmental Sciences<br />

University of Virginia<br />

December, 2010


Abstract<br />

Woody plant encroachment is a widespread process of land <strong>cover</strong> change from<br />

grass- to woody plant-dominated habitats that has been observed in arid and semiarid<br />

ecosystems around the world. Woody plant encroachment affects ecosystem<br />

characteristics, including plant resource availability and microclimate, and it has the<br />

potential to have bottom-up effects on a wide range of animals through changes in habitat<br />

structure and the abundance and species richness of prey. Little is known about the<br />

impact that woody plant encroachment may have on the ecology of mammalian<br />

predators. This dissertation begins by assessing the consequences of woody plant<br />

encroachment <strong>fo</strong>r mammalian predator ecology globally and then <strong>fo</strong>cuses on the ecology<br />

of a widely distributed, North American predator, the coyote (Canis latrans).<br />

A dataset containing in<strong>fo</strong>rmation on the global distribution of mammals was used<br />

to generate a list of carnivores <strong>fo</strong>und in woody plant-encroached areas. Noninvasive<br />

genetic sampling and ca<strong>rb</strong>on isotope techniques were used to assess the individual-<br />

specific feeding ecology of coyotes at the <strong>Sevilleta</strong> National Wildlife Refuge (NWR) and<br />

Long Term Ecological Research (<strong>LTER</strong>) site in central New Mexico, USA. The specific<br />

objective was to determine whether woody plant encroachment, which has occurred over<br />

the past century at the <strong>Sevilleta</strong> NWR, has led to a shift in the base of the coyote <strong>fo</strong>od<br />

chain from native C4 grasses to encroaching C3 woody plants. <strong>Habitat</strong> characteristics<br />

associated with woody plant encroachment were assessed at multiple spatial scales to<br />

determine whether spatial scale has an impact on the observed relationship between<br />

coyote feeding ecology and woody plant encroachment.<br />

i


ii<br />

At least 97 mammalian carnivores are <strong>fo</strong>und in woody plant-encroached areas.<br />

Spatial scale affects the strength of the relationship between coyote feeding ecology and<br />

woody plant encroachment. A significant shift in the base of the coyote <strong>fo</strong>od chain from<br />

C4 grasses to C3 plants was observed only when habitat characteristics were evaluated at<br />

a small spatial scale. Further research regarding the effects of woody plant encroachment<br />

on mammalian predator ecology, especially on predator fitness and the ecology of<br />

specialized predators, is needed.


Table of contents<br />

Abstract…………………………………………………………………………………..i<br />

Table of contents………………………………………………………………………..iii<br />

List of figures………………………………………………………………………….....v<br />

List of tables…………………………………………………………………………….vii<br />

List of appendices……………………………………………………………………...viii<br />

Acknowledgements……………………………………………………………………...ix<br />

Overview of chapters…………………………………………………………………….1<br />

Chapter 1: Causes and consequences of woody plant encroachment………………...4<br />

Definition, importance, and extent of woody plant encroachment………………..4<br />

Causes of woody plant encroachment……………………………………………..8<br />

Abiotic consequences of woody plant encroachment……………………………11<br />

Biotic consequences of woody plant encroachment……………………………..13<br />

Conclusion……………………………………………………………………….21<br />

References………………………………………………………………………..23<br />

Chapter 2: Dissertation methods………………………………………………………34<br />

Study site…………………………………………………………………………34<br />

Field data collection……………………………………………………………...37<br />

Laboratory work………………………………………………………………….48<br />

GIS analysis……………………………………………………………………...60<br />

References………………………………………………………………………..63<br />

Chapter 3: Genetic results……………………………………………………………...67<br />

Results and discussion…………………………………………………………...67<br />

iii


iv<br />

Conclusion……………………………………………………………………….87<br />

References……………………………………………………………………….88<br />

Chapter 4: Bottom-up effects of seasonal and long-term habitat change on predator<br />

feeding ecology………………………………………………………………………….90<br />

Introduction………………………………………………………………………91<br />

Methods…………………………………………………………………………..96<br />

Results…………………………………………………………………………..102<br />

Discussion………………………………………………………………………107<br />

Conclusion……………………………………………………………………...116<br />

References………………………………………………………………………117<br />

Chapter 5: The impact of spatial scale on the relationship between coyote feeding<br />

ecology and local habitat characteristics…………………………………………….130<br />

Introduction……………………………………………………………………..131<br />

Methods…………………………………………………………………………135<br />

Results…………………………………………………………………………..141<br />

Discussion………………………………………………………………………146<br />

Conclusion……………………………………………………………………...152<br />

References………………………………………………………………………154<br />

Conclusion……………………………………………………………………………..165


List of figures<br />

Chapter 1<br />

Chapter 2<br />

Chapter 3<br />

Figure 1. Map of woody plant-encroached sites…………………………………..6<br />

Figure 2. Numbers of carnivores in woody plant-encroached areas…………….19<br />

Figure 1. Study site map…………………………………………………………35<br />

Figure 2. Scat transect map………………………………………………………37<br />

Figure 3. Seasonal variation in rainfall at the study site…………………………41<br />

Figure 4. Vegetation plot map…………………………………………………...44<br />

Figure 1. Species ID results……………………………………………………...67<br />

Figure 2. Scat samples per individual……………………………………………68<br />

Figure 3. Gimlet family group map……………………………………………...74<br />

Figure 4. ML-Relate family group map………………………………………….74<br />

Figure 5. Structure analysis results………………………………………………76<br />

Figure 6. BAPS analysis results………………………………………………….76<br />

Figure 7. Genetic groups generated by Structure………………………………...77<br />

Figure 8. Genetic groups generated by BAPS (K = 10)…………………………79<br />

Figure 9. Genetic groups generated by BAPS (K = 2)…………………………..80<br />

Figure 10. Map of individuals per scat transect………………………………….81<br />

Figure 11. Inter-habitat differences in number of individuals …………………..81<br />

Figure 12. Seasonal variation in number of individuals…………………………82<br />

Figure 13. Inter-habitat variation in home range size……………………………83<br />

v


Chapter 4<br />

Chapter 5<br />

vi<br />

Figure 14. Map of family groups relative to habitat use…………………………86<br />

Figure 1. Map of study site………………………………………………………96<br />

Figure 2. Inter-habitat and seasonal differences in <strong>live</strong> <strong>cover</strong> and coyote diet…105<br />

Figure 1. Maps of land <strong>cover</strong> at the study site………………………………….144<br />

Figure 2. Results of t-tests at multiple spatial scales…………………………...145


List of tables<br />

Chapter 1<br />

Chapter 2<br />

Chapter 3<br />

Chapter 4<br />

Table 1. Studies of woody plant encroachment…………………………………...7<br />

Table 2. Summary of carnivores in woody plant-encroached areas……………..20<br />

Table 1. Species ID primers……………………………………………………...49<br />

Table 2. Size ranges <strong>fo</strong>r mtDNA amplified from different carnivore species…...49<br />

Table 3. Microsatellite primer sequences………………………………………..51<br />

Table 4. Masses <strong>fo</strong>r stable isotope analysis……………………………………...59<br />

Table 5. Fractionation correction factors………………………………………...60<br />

Table 1. Microsatellite success and error rates…………………………………..69<br />

Table 2. Hardy-Weinberg test results……………………………………………70<br />

Table 3. Linkage equilibrium test results………………………………………...71<br />

Table 4. Observed and expected heterozygosity..………………………………..72<br />

Table 5. Data <strong>fo</strong>r chi square analysis of pairs of related individuals…………….84<br />

Table 6. Data <strong>fo</strong>r chi square analysis of trios of related individuals……………..85<br />

Table 1. Two-way ANOVA of percent <strong>live</strong> <strong>cover</strong>……………………………...103<br />

Table 2. Two-way ANOVA of percent <strong>live</strong> grass <strong>cover</strong>……………………….104<br />

Table 3. Repeated measures ANOVA of coyote diet…………………………..106<br />

vii


List of appendices<br />

Chapter 1<br />

Chapter 4<br />

Chapter 5<br />

Appendix 1. Woody plant encroachment references…………………………….32<br />

Appendix 1. Intra-individual variation in diet………………………………….126<br />

Appendix 2. δ 13 C values <strong>fo</strong>r vegetation samples and scat components………...127<br />

Appendix 3. Inter-habitat and seasonal differences in <strong>fo</strong><strong>rb</strong> <strong>cover</strong>……………...128<br />

Appendix 4. Composition and diet of the small mammal community…………129<br />

Appendix 1. Converting Landsat 7 ETM+ pixel values to reflectance values…162<br />

Appendix 2. Discriminant function analysis of vegetation variables…………..163<br />

Appendix 3. Discriminant function analysis of reflectance values from a Landsat<br />

7 ETM+ image………………………………………………………………….164<br />

Isotope Appendices<br />

Appendix 1. Seasonal variation in coyote diet………………………………….168<br />

Appendix 2. Assessing variation in vegetation end-member values……….......169<br />

Appendix 3. Ca<strong>rb</strong>on and nitrogen isotope data <strong>fo</strong>r scat samples……………….170<br />

Appendix 4. Ca<strong>rb</strong>on and nitrogen isotope data <strong>fo</strong>r vegetation samples………..178<br />

viii


Acknowledgements<br />

There are a large number of people without whom this dissertation would never<br />

have been written. First, I want to thank my advisor, Hank Shugart, and all of my<br />

committee members, especially Paolo D’Odorico, Mike Erwin, and Henry Wilbur, <strong>fo</strong>r<br />

their unfailing support and ever excellent advice. Thank you also to the <strong>fo</strong>llowing people<br />

at UVa <strong>fo</strong>r their assistance and advice: Katie Burke, Dave Carr, Eric Elton, Rachel<br />

Michaels, Tami Ransom, Dave Richardson, Keir Sode<strong>rb</strong>erg, and Mike Tuite.<br />

There are many people associated with the <strong>Sevilleta</strong> National Wildlife Refuge and<br />

<strong>LTER</strong> who provided field, logistical, financial, and/or moral support. Special thanks to:<br />

Amanda Boutz, Scott Collins, John Craig, John DeWitt, Michael Donovan, Jon Erz, Mike<br />

Friggens, Jennifer Johnson, Terri Koontz, Mike Parker, Dennis Prichard, and Matt<br />

Spinelli. There are several undergraduate students, Kelly Bowman, Cesar Coronado,<br />

Adrianna Foster, Jeffrey Freiberg, Damon Lowery, and Kelsey Wicks, who helped me in<br />

the field or did projects to which I referred when writing this dissertation.<br />

I am greatly indebted to Dr. Lisette Waits and her students at the University of<br />

Idaho. Without the training that they gave me and expertise that they shared, the genetics<br />

portion of my dissertation would never have been completed. Special thanks to the<br />

<strong>fo</strong>llowing students: Marta DeBa<strong>rb</strong>a, Kara Gebhardt, Caren Goldberg, Matt Mumma,<br />

Carisa Stansbury, and Claudia Wultsch. Thank you to Andrew Ouimette at the University<br />

of New Hampshire <strong>fo</strong>r running all of my samples <strong>fo</strong>r ca<strong>rb</strong>on isotope analysis.<br />

Finally, I owe my sanity and happiness over the past 5 and a half years to my<br />

parents, Tom and Teresa Seamster, and my friends, Almea, Erwin, Ian, Jackie, James,<br />

Jon, Karles, Nina, Peter, Sahu, Sujith, and Yo.<br />

ix


Overview of chapters<br />

Woody plant encroachment is characterized by a proliferation of woody plants in<br />

a grass-dominated environment. This process has been occurring over the last 50 to 200<br />

years in dryland areas on six of the seven continents. Very little is known about the<br />

impacts that this widespread shift in habitat, from grass- to woody plant-dominated, has<br />

on the ecology of predatory mammals. Mammalian predators often play a crucial role in<br />

local ecosystem structure and function and are also sensitive to changes that occur at the<br />

base of the <strong>fo</strong>od chain. The central question asked in this dissertation is: What are the<br />

consequences of woody plant encroachment <strong>fo</strong>r the ecology of mammalian predators?<br />

Chapter 1 is a review of the causes and consequences, especially the biotic<br />

consequences, of woody plant encroachment. This chapter also addresses the question:<br />

How many mammalian carnivores are present in areas affected by woody plant<br />

encroachment globally?<br />

The <strong>fo</strong>ur remaining chapters <strong>fo</strong>cus on assessing the impacts of woody plant<br />

encroachment on the feeding ecology of an abundant, generalist predator that is <strong>fo</strong>und<br />

throughout North America. Chapter 2 is a detailed description of the noninvasive genetic<br />

sampling and ca<strong>rb</strong>on isotope techniques used to assess the feeding ecology of the coyote<br />

(Canis latrans) population at the <strong>Sevilleta</strong> National Wildlife Refuge (NWR) and Long<br />

Term Ecological Research Site in New Mexico, USA. Creosote bush (Larrea tridentata)<br />

has been spreading into grama-(Bouteloua spp.) dominated grasslands at the <strong>Sevilleta</strong><br />

NWR over the past century. Noninvasive genetic sampling was used to identify<br />

individual coyotes and ca<strong>rb</strong>on isotope techniques were used to assess the feeding ecology<br />

of the identified individuals. In particular, the use of ca<strong>rb</strong>on isotopes allows <strong>fo</strong>r<br />

1


identification of the base of the coyote <strong>fo</strong>od chain as native C4 grasses or encroaching C3<br />

woody plants. Chapter 3 presents the results of the noninvasive genetic sampling scheme<br />

that is described in the second chapter and applied in Chapters 4 and 5. In<strong>fo</strong>rmation<br />

regarding coyote family groups, genetic structure of the coyote population, and habitat<br />

use patterns of related individuals are all included in this third chapter.<br />

Chapter 4 considers the impacts of both long-term habitat change, associated<br />

with the spread of C3 woody plants in a native C4 grassland, and seasonal shifts in habitat<br />

characteristics on coyote feeding ecology. The <strong>Sevilleta</strong> NWR is located in an arid<br />

environment where there is a strong seasonality to local rainfall patterns. Seasonal<br />

variation in habitat is characterized by pulses of C4 grass production in response to<br />

summer rains. Chapter 4 addresses the <strong>fo</strong>llowing two questions: 1) Is there a difference in<br />

the base of the coyote <strong>fo</strong>od chain between native grassland and woody plant-encroached<br />

shrubland habitats? and 2) Is there seasonal variation in the base of the coyote <strong>fo</strong>od chain<br />

in an area impacted by woody plant encroachment? It was expected that the base of the<br />

coyote <strong>fo</strong>od chain would differ between grassland and woody plant-encroached habitats<br />

such that the percent of the coyote diet coming indirectly from encroaching C3 woody<br />

plants would be higher in the woody plant-encroached habitat. It was also expected that<br />

there would be a seasonal shift in the base of the <strong>fo</strong>od chain with percent coyote diet<br />

coming indirectly from native C4 grasses increasing from the spring to the summer and<br />

fall in both grassland and woody plant-encroached habitats.<br />

Chapter 5 considers the effect of spatial scale on the strength of the relationship<br />

between coyote feeding ecology and habitat characteristics associated with woody plant<br />

encroachment. The spatial scale at which habitat is characterized can have a dramatic<br />

2


3<br />

effect on the observed relationship between various aspects of animal ecology and the<br />

local environment. A consideration of spatial scale is especially pertinent in studies of the<br />

ecology of wide-ranging animals such as the coyote. The <strong>fo</strong>llowing question is addressed:<br />

Does a) the size of the difference in the base of the coyote <strong>fo</strong>od chain between grassland<br />

and woody plant-encroached shrubland habitats, or b) the strength of the linear<br />

relationship between percent coyote diet from C3 woody plants and percent available<br />

shrubland habitat, change with the spatial scale at which the habitat variables are<br />

assessed? Given the wide-ranging nature of coyotes, it was expected that the difference in<br />

coyote feeding ecology between habitat <strong>type</strong>s would become larger, and the association<br />

between coyote diet and percent shrubland habitat would become stronger, as the spatial<br />

scale of the analysis increased and approached the size of an area that coyotes typically<br />

use (i.e., the size of a coyote home range). It was also expected that the strength of the<br />

relationship between coyote feeding ecology and habitat characteristics would decline at<br />

spatial scales larger than the area typically used by a coyote (i.e., larger than a coyote<br />

home range).


Chapter 1: Causes and consequences of woody plant encroachment<br />

Abstract<br />

Woody plant encroachment, the proliferation of woody plants in grassland or<br />

savanna areas, affects semiarid and arid landscapes around the world. Many studies have<br />

considered the potential drivers, as well as the abiotic and economic consequences, of<br />

this phenomenon. This widespread shift in habitat <strong>type</strong> also has pro<strong>fo</strong>und bottom-up<br />

effects on local faunal communities, which are less studied and in need of both review<br />

and further attention. The causes and consequences, especially the biotic consequences,<br />

of woody plant encroachment are reviewed here. Consequences include changes in<br />

resource availability, microclimate, habitat structure, species richness, and predator-prey<br />

interactions. Implications <strong>fo</strong>r the ecology and conservation of top predators are<br />

emphasized. These species often are threatened by environmental change and play a<br />

critical role in ecosystem function. The ecological impacts of woody plant encroachment<br />

are varied, wide reaching, and in need of further study.<br />

Definition, importance, and extent of woody plant encroachment<br />

Definition<br />

Woody plant encroachment is defined here as the increase in abundance, density,<br />

or <strong>cover</strong> of one or more native species of woody plants through time (based on Van<br />

Auken 2000, Roques et al. 2001). Woody plants include trees (> 5 m tall), shrubs (0.5 to<br />

5 m tall), and sub-shrubs (< 0.5 m tall; USDA, NRCS 2010). This review <strong>fo</strong>cuses on the<br />

spread of woody plants in grassland and savanna ecosystems located in dryland areas<br />

(i.e., arid, semiarid, dry sub-humid) that receive between 100 and 1200 mm of rain per<br />

year (D’Odorico and Porporato 2006). Of particular interest are habitat shifts from<br />

4


5<br />

grassland to shrubland (e.g., Gill and Burke 1999) or from savanna to woodland (e.g.,<br />

Archer 1989). In many cases, this habitat change takes place over the course of roughly<br />

100-200 years (Gill and Burke 1999, Archer 1995, Archer et al. 1995, van Auken 2000).<br />

However, analyses of historical aerial photographs and satellite imagery have detected<br />

increases in woody plant <strong>cover</strong> over much shorter time periods (e.g., 20-40 years; Hudak<br />

and Wessman 2001, Silva et al. 2001). There are a variety of terms, other than woody<br />

plant encroachment, that refer to the spread of woody plants in areas previously<br />

dominated by grasses and which, <strong>fo</strong>r the purposes of this paper, are interchangeable. The<br />

term used depends on the woody plant species being studied, as well as the geographic<br />

region of interest. These terms include shrub encroachment (e.g., Prosopis glandulosa<br />

encroachment in the southwestern United States; Goslee et al. 2003), and bush<br />

encroachment (e.g., Acacia spp. and Dichrostachys cinerea encroachment in southern<br />

Africa; Hudak and Wessman 2001, Muntifering et al. 2006).<br />

Importance and extent<br />

In some regions, woody plant encroachment is associated with a decline in<br />

<strong>live</strong>stock carrying capacity (e.g., Rappole et al. 1986, Bester 1996) or with desertification<br />

(e.g., Schlesinger et al. 1990, Huenneke et al. 2002, Li et al. 2006). Woody plant<br />

encroachment can reduce local cattle carrying capacities by up to 80% (Bester 1996).<br />

While it is possible to remove the undesirable woody vegetation, treatment may have to<br />

be repeated every 2 to 15 years (Rappole et al. 1986). Desertification is defined as<br />

degradation of land in dryland areas that is driven primarily by anthropogenic <strong>fo</strong>rcings<br />

(UNEP 1992, UNCCD 1994, Maestre et al. 2006). This process leads to a decline in <strong>fo</strong>od<br />

production in agricultural areas and has the potential to negatively impact roughly 16% of


6<br />

the global human population. Approximately 60,000 km 2 (0.1%) of dryland area is<br />

degraded each year and, while Asia contains the largest total area of degraded dryland,<br />

the continents in which the highest percentages of total dryland area have been degraded<br />

are North America and Africa (UNEP 1992).<br />

Woody plant encroachment is a global phenomenon. Drylands, specifically arid,<br />

semiarid and dry sub-humid areas, <strong>cover</strong> roughly 40% of the earth’s terrestrial surface<br />

(UNEP 1992), while grasslands and savannas, the ecosystems impacted by woody plant<br />

encroachment, account <strong>fo</strong>r approximately 20% (Ojima et al. 1996). Drylands are <strong>fo</strong>und in<br />

the western United States, the southern and western part of South America, the northern<br />

and southern-most parts of Africa, eastern Europe, western and central Asia and much of<br />

Australia (UNEP 1992). Grassland and savanna ecosystems on six of the seven<br />

continents have been affected by the encroachment of a variety of woody plant species<br />

(Archer 1995, Ravi et al. 2009; Figure 1 and Table 1).<br />

Figure 1. Map of woody plant-encroached sites. This map shows 34 locations (black<br />

circles; Table 1, Appendix 1) around the world where woody plant encroachment has<br />

been documented in grassland and savanna ecosystems that receive less than 1200 mm/yr<br />

of rain (dark gray regions; Olson et al. 2001). See Appendix 1 <strong>fo</strong>r further details.


Table 1. Studies of woody plant encroachment. Summary of 25 studies that document woody plant encroachment around the world.<br />

Average annual rainfall values (mm/yr; Appendix 1) were used to determine the dryland zone (arid = 100-250 mm/yr; semiarid = 250-<br />

600 mm/yr; dry sub-humid = 600-1200 mm/yr; D’Odorico and Porporato 2006). Encroaching species are woody plants that are<br />

classified as one of the <strong>fo</strong>llowing: sub-shrubs, shrubs or trees. For continents and countries/states, N = North, S = South. See<br />

Appendix 1 <strong>fo</strong>r further details.<br />

Continent Country/State Dryland zone Original habitat Encroaching species Reference<br />

Africa S. Africa arid/semiarid savanna Acacia spp., Boscia albitrunca, Grewia Tews et al. 2004, Palmer<br />

flava, Rhigozum trichotomum<br />

and van Rooyen 1998<br />

Swaziland dry sub-humid savanna Dichrostachys cinerea Roques et al. 2001<br />

Asia China arid/semiarid grassland/steppe Artemisia spp., Caragana spp., Li et al. 2006, Li et al. 2004,<br />

Ceratoides lateens<br />

Chen et al. 2005, Jin et al.<br />

2009<br />

Australia New S. Wales dry sub-humid grassland Acacia sophorae Costello et al. 2000<br />

Europe Greece semiarid grassland Quercus coccifera Zarovali et al. 2007<br />

N. America Alaska semiarid grassland Betula nana, Ledum palustre Knapp et al. 2008<br />

Arizona semiarid grassland Prosopis velutina Wheeler et al. 2007<br />

Cali<strong>fo</strong>rnia semiarid meadow Artemisia rothrockii Berlow et al. 2002<br />

Canada semiarid grassland Populus tremuloides Steinaker and Wilson 2008<br />

Kansas dry sub-humid grassland Cornus drummondii Knapp et al. 2008<br />

Mexico semiarid<br />

semiarid/dry<br />

grassland Ephedra trifurca, Prosopis glandulosa Ceballos et al. 2010<br />

Minnesota sub-humid prairie Juniperus virginiana Pierce and Reich 2010<br />

Goslee et al. 2003,<br />

Hochstrasser and Peters<br />

New Mexico arid grassland Larrea tridentata, Prosopis glandulosa<br />

2004<br />

Brown and Archer 1999,<br />

Ansley et al. 2001,<br />

Texas dry sub-humid savanna Juniperus ashei, Prosopis glandulosa Schwinning 2008<br />

Virginia dry sub-humid grassland Morella cerifera Knapp et al. 2008<br />

Wyoming semiarid grassland Artemisia tridentate Knapp et al. 2008<br />

S. America Argentina semiarid/dry grassland/steppe Austrocedrus chilensis, Chuquiraga Ghermandi et al. 2010,<br />

sub-humid<br />

avellanedae, Fabiana imbricata, Kitzberger et al. 2000, de<br />

Prosopis caldenia<br />

Villalobos et al. 2005,<br />

Beeskow et al. 1995<br />

Uruguay dry sub-humid grassland Baccharis spp., Eupatorium bunii<strong>fo</strong>lium Altesor et al. 2006<br />

7


8<br />

The areas affected by woody plant encroachment are very large. Various species of the<br />

woody plant mesquite (Prosopis spp.) are <strong>fo</strong>und on over 380,000 km 2 (9%) of the<br />

semiarid ecosystems of North America (UNEP 1992, Van Auken 2000). Approximately<br />

350,000 km 2 (29%) of a semiarid area in South America has been overgrazed and has<br />

turned into a dense shrubland (Abril and Bucher 2001). In Southern Africa, at least<br />

280,000 km 2 (~10%) of the land has been affected by woody plant encroachment (Bester<br />

1996, Moleele et al. 2002, Hagenah et al. 2009).<br />

Causes of woody plant encroachment<br />

The widespread nature and often negative implications of woody plant<br />

encroachment have lead to extensive study of the factors that drive this process. Proposed<br />

drivers <strong>fo</strong>r woody plant encroachment include: changes in fire frequency (Buffington and<br />

He<strong>rb</strong>el 1965, Archer 1995, Roques et al. 2001); fertilization associated with rising ca<strong>rb</strong>on<br />

dioxide concentrations (Archer et al. 1995, Bond and Midgley 2000); nitrogen deposition<br />

(Köchy and Wilson 2001, Wigley et al. 2010); overgrazing by cattle (Walker et al. 1981,<br />

Archer 1989, Archer 1995, Van Auken 2000); seed dispersal by <strong>live</strong>stock (Buffington<br />

and He<strong>rb</strong>el 1965, Brown and Carter 1998) and rodents (Buffington and He<strong>rb</strong>el 1965); soil<br />

erosion (Buffington and He<strong>rb</strong>el 1965, Walker et al. 1981, McGlynn and Okin 2006);<br />

decreased infiltration of water to the topsoil (Walker et al. 1981, Schlesinger et al. 1990);<br />

and climatic change, including changes in rainfall patterns and temperature (Buffington<br />

and He<strong>rb</strong>el 1965, Neilson 1986, Schlesinger et al. 1990, Fensham et al. 2005, He et al.<br />

2010). The mechanisms <strong>fo</strong>r several of these drivers relate to resource competition<br />

between grasses and woody plants. In particular, conditions that favor the growth of


9<br />

woody plants, or hinder the growth of grasses, will facilitate the process of woody plant<br />

encroachment.<br />

A reduction in fire frequency, above average rainfall, or the elevation of<br />

atmospheric ca<strong>rb</strong>on dioxide levels are examples of processes that can facilitate the<br />

growth and subsequent spread of shrubs and other woody vegetation (Buffington and<br />

He<strong>rb</strong>el 1965, Archer 1995, Archer et al. 1995, Van Auken 2000, Fensham et al. 2005).<br />

Shrub seedlings are prone to fire-caused mortality (Buffington and He<strong>rb</strong>el 1965) and, if<br />

fires occur too often, shrubs are not able to produce seeds and grow to a size at which<br />

they are more resistant to the effects of fire (Van Auken 2000). Multiple year periods of<br />

higher-than-average rainfall can lead to the spread of woody vegetation, especially when<br />

initial woody plant density is relatively low (Fensham et al. 2005). Elevated ca<strong>rb</strong>on<br />

dioxide concentrations can lead to a fertilization effect, whereby plants are able to fix<br />

more ca<strong>rb</strong>on, produce more photosynthate and thus grow faster than was previously<br />

possible (Bazzaz 1990, Shugart 1998). Due to differences in their photosynthetic<br />

pathways, many woody plant species (i.e., C3 plants) may benefit more from any such<br />

fertilization effect, and subsequently grow faster and accumulate more biomass, than<br />

many of the grasses (i.e., C4 species) <strong>fo</strong>und in semiarid regions (Idso 1992, Johnson et al.<br />

1993, Archer et al. 1995, Van Auken 2000). In particular, C3 plants are less efficient at<br />

taking up ca<strong>rb</strong>on dioxide and more prone to the loss of photosynthate via<br />

photorespiration than C4 plants, and their photosynthetic rates are not saturated at current<br />

atmospheric ca<strong>rb</strong>on dioxide concentrations (Johnson et al. 1993, Chapin et al. 2002). It<br />

has also been proposed that elevated ca<strong>rb</strong>on dioxide levels increase the probability of<br />

trees in savanna ecosystems growing to a size at which they are no longer prone to fire-


10<br />

induced mortality (Bond and Midgley 2000), thus facilitating an increase in woody plant<br />

biomass.<br />

Grazing, drought, erosion, and reduced water infiltration to the soil are examples<br />

of processes hindering grass growth (Buffington and He<strong>rb</strong>el 1965, Walker et al. 1981). In<br />

particular, heavy grazing by <strong>live</strong>stock and periods of low rainfall decrease local grass<br />

density and allows shrub species to become established (Buffington and He<strong>rb</strong>el 1965).<br />

Furthermore, the reduction in grass <strong>cover</strong> associated with overgrazing and drought leads<br />

to an increase in erosion of the exposed soil patches (Buffington and He<strong>rb</strong>el 1965,<br />

Walker et al. 1981, Schlesinger et al. 1990). This erosion negatively impacts the<br />

establishment and growth of grasses through the loss of topsoil and the deposition of soil<br />

on, and subsequent damage to, extant grasses (Buffington and He<strong>rb</strong>el 1965). The removal<br />

of grasses and creation of bare soil patches associated with overgrazing and drought also<br />

reduces water infiltration to the top soil layers (Walker et al. 1981), especially when<br />

combined with soil compaction associated with <strong>live</strong>stock movements (Schlesinger et al.<br />

1990). Since grasses tend to draw water from layers near the soil surface and grass<br />

biomass is positively associated with infiltration of water into the topsoil (Walker et al.<br />

1981), this reduction in infiltration can be detrimental to local grass populations.<br />

There is no single cause <strong>fo</strong>r woody plant encroachment. The drivers <strong>fo</strong>r this<br />

process work at different spatial scales, vary geographically, and often work together.<br />

Elevated ca<strong>rb</strong>on dioxide concentrations affect areas around the globe but overgrazing is a<br />

region-specific process (Wigley et al. 2010). Overgrazing and an associated change in<br />

fire frequency are cited as drivers in Southern Africa, where ranching plays an important<br />

part in the local economy (Bester 1996). Drought is one of the drivers of woody plant


11<br />

encroachment in the southwestern United States (Buffington and He<strong>rb</strong>el 1965) while the<br />

occurrence of higher-than-average rainfall in multiple years is reported as a driver of the<br />

spread of woody plants in Australian savannas (Bowman et al. 2001, Fensham et al.<br />

2005). Finally, there are cases where multiple woody plant encroachment drivers work in<br />

concert with one another (e.g., grazing, seed dispersal, soil erosion and drought in the<br />

southwestern United States; Buffington and He<strong>rb</strong>el 1965).<br />

Abiotic consequences of woody plant encroachment<br />

The abiotic consequences of the land <strong>cover</strong> shifts associated with woody plant<br />

encroachment are varied and include changes in patterns of resource availability<br />

(Schlesinger et al. 1990, Lett and Knapp 2003, Li et al. 2006). An increase in woody<br />

plant <strong>cover</strong> reduces the availability of light (Lett and Knapp 2003), nutrients (e.g.,<br />

nitrogen, phosphorous, and potassium), and water (Schlesinger et al. 1990, Li et al. 2006)<br />

<strong>fo</strong>r other plants, especially grasses. However, the abundance of resources, especially<br />

nitrogen and water, and the rates of nutrient fluxes, specifically of nitrogen<br />

mineralization, are often elevated under woody plants relative to the surrounding grassy<br />

or bare patches (Sánchez et al. 1997, Schlesinger et al. 1990).<br />

Woody plant encroachment can also have an impact on the local microclimate and<br />

hydrological cycle (Kidron 2009, Moran et al. 2009), and on the hydrological properties<br />

of the soil (Schlesinger et al. 1990, Li et al. 2006). In arid regions where shrubs are<br />

surrounded by bare ground, areas under shrubs tend to have reduced rates of evaporation<br />

and lower temperatures relative to more exposed areas between the shrubs (Kidron 2009).<br />

However, when comparing grassland and shrubland areas, grassland areas tend to have a<br />

higher ratio of transpiration to total evapotranspiration (Moran et al. 2009), which


12<br />

indicates that the proportion of water lost via evaporation is higher in shrubland areas.<br />

Water infiltration rates may be reduced by the conversion of grassland to shrubland,<br />

especially when grass <strong>cover</strong> is initially reduced by grazing and the exposed soil is<br />

compacted by the movements of the grazers. In this case, infiltration rates are high<br />

underneath encroaching woody plants but low in the areas that surround them<br />

(Schlesinger et al. 1990). In addition, a decline in soil water-holding capacity can occur<br />

when there is a transition from grass to shrub-dominated landscapes (Li et al. 2006).<br />

In summary, woody plant encroachment can have a significant impact on the<br />

distribution and cycling of important resources. Furthermore, the abiotic consequences of<br />

woody plant encroachment can combine with the drivers of this shift in habitat to<br />

generate a feedback between the environment and vegetation, which results in further<br />

spread of woody plants and, in some cases, further degradation of the local habitat<br />

(Buffington and He<strong>rb</strong>el 1965, Walker et al. 1981, Schlesinger et al. 1990, Maestre et al.<br />

2006). As an example of such a feedback, grazing and drought reduce grass <strong>cover</strong>, which<br />

leads to elevated soil erosion (Buffington and He<strong>rb</strong>el 1965, Walker et al. 1981). The<br />

reduction in grass biomass opens up space <strong>fo</strong>r shrubs to grow and reduces the probability<br />

of fires hot enough to kill shrub seedlings (Buffington and He<strong>rb</strong>el 1965). Soil erosion<br />

causes further grass death through the loss of topsoil and nutrients and soil deposition on<br />

previously established grass plants (Buffington and He<strong>rb</strong>el 1965, Schlesinger et al. 1990).<br />

Dust generated by erosion of bare soil patches leads to the <strong>fo</strong>rmation of clouds with a<br />

high percentage of small droplets, which results in a reduction in local rainfall (UNEP<br />

1992, Rosenfeld et al. 2001). This decline in rainfall, combined with reduced water<br />

infiltration rates associated with both trampling by <strong>live</strong>stock and lower grass <strong>cover</strong>,


13<br />

decreases water availability in the topsoil and further retards grass growth (Walker et al.<br />

1981, Schlesinger et al. 1990, Rosenfeld et al. 2001). Over time, nutrients and water<br />

accumulate under the encroaching shrubs and allow <strong>fo</strong>r shrub regeneration while a lack<br />

of suitable topsoil, nutrients and water in inter-shrub spaces prevents grass re-growth<br />

(Buffington and He<strong>rb</strong>el 1965, Schlesinger et al. 1990, Bhark and Small 2003).<br />

Biotic consequences of woody plant encroachment<br />

Changes in habitat characteristics<br />

The spread of woody plants, especially shrubs or trees, in a previously grass-<br />

dominated environment leads to changes in the structure, temporal stability, and quality<br />

of the habitat (Rappole et al. 1986, Brown et al. 1997, Hernández et al. 2005, Blaum et al.<br />

2006, Báez and Collins 2008). Woody plant encroachment implies the spread of plants<br />

that are taller, wider, and, in the case of shrubs, more impenetrable than the native<br />

grasses. In some cases, woody plants <strong>fo</strong>rm clusters that grow over time (Archer 1989).<br />

This introduction of woody vegetation leads to an increase in the complexity of the local<br />

habitat structure and transition from a single to multiple-stratum environment (Archer et<br />

al. 1988). In general, shrublands are likely to have a greater diversity of microhabitats<br />

than grasslands (Hernández et al. 2005). Furthermore, shrubs provide shade, <strong>cover</strong>, or<br />

ideal burrow sites <strong>fo</strong>r some organisms (Blaum et al. 2006, Blaum et al. 2007a) while<br />

presenting obstacles to the activities of others (Broomhall 2001, Mills et al. 2004, Blaum<br />

et al. 2007a). Yellow mongoose (Cynictis penicillata) burrows are often <strong>fo</strong>und<br />

underneath large Acacia shrubs as the shrubs help to moderate temperature fluctuations<br />

and provide protection against trampling by he<strong>rb</strong>ivores and predation by raptors (Blaum<br />

et al. 2007a). On the other hand, increased woody vegetation <strong>cover</strong> may hinder the high-


14<br />

speed hunting technique of the cheetah (Acinonyx jubatus, Bertram 1979, Broomhall<br />

2001, Mills et al. 2004).<br />

Woody plant encroachment has varied effects on the stability of the local habitat.<br />

In semiarid regions, primary productivity and resource availability tend to be more stable<br />

in shrubland than grassland habitats. In particular, grasses have short-term spikes in<br />

growth and seed production in response to rainfall events. These spikes are <strong>fo</strong>llowed by<br />

drying of the above-ground tissues and a decline in the availability of <strong>fo</strong>od resources,<br />

especially seeds, to various primary consumers. On the other hand, some shrub species<br />

are evergreen and, if there are a variety of species present, edible fruits may be available<br />

in all seasons (Hernández et al. 2005). As a result, trans<strong>fo</strong>rmation from grassland to<br />

shrubland via woody plant encroachment may increase the stability of <strong>fo</strong>od resources in<br />

the local environment. However, woody plant encroachment decreases the stability of the<br />

composition of the local plant community, with greater changes in community<br />

composition occurring over a decade in areas with higher percent shrub <strong>cover</strong> (Báez and<br />

Collins 2008).<br />

An increase in woody plant <strong>cover</strong> can reduce the quality of the local habitat <strong>fo</strong>r<br />

both domesticated and wild animal populations. The woody plant encroachment that has<br />

occurred on many ranches and commercial farms has led to a reduction in local <strong>live</strong>stock<br />

carrying capacity (e.g., Rappole et al. 1986, Bester 1996). In particular, the spread of<br />

shrubs leads to a decline in the production and nutritional value of grasses and legumes<br />

(Zarovali et al. 2007). Studies of wild animal populations have shown a decline in the<br />

abundance or density of a variety of organisms, including insects, rodents, ungulates, and<br />

carnivores, in areas affected by woody plant encroachment (Brown et al. 1997, Marker


15<br />

2002, Blaum et al. 2007b). This decline was driven by a reduction in the availability of<br />

various <strong>fo</strong>od resources (e.g., seeds, insects and rodents; Brown et al. 1997, Blaum et al.<br />

2007b).<br />

<strong>Habitat</strong> degradation has a variety of impacts on local fauna, especially on<br />

predators. Carnivores living in degraded habitats typically have larger home ranges and<br />

need to expend more energy in the course of caring <strong>fo</strong>r their offspring than organisms<br />

<strong>fo</strong>und in higher quality habitats (Sunquist and Sunquist 2001). This has been observed <strong>fo</strong>r<br />

predators living in areas impacted by woody plant encroachment. In particular, cheetahs<br />

in a woody plant-encroached area in southern Africa had home ranges that were at least<br />

three times larger than those of cheetahs inhabiting more open, grassland habitat (Marker<br />

2002). This provides further evidence that woody plant encroachment reduces local<br />

habitat quality.<br />

Changes in species richness and diversity<br />

Increasing woody plant <strong>cover</strong> is often associated with a decline in local species<br />

richness (Blaum et al. 2006, Blaum et al. 2007b, Báez and Collins 2008, Blaum et al.<br />

2009, Sirami et al. 2009). This decline has been observed across trophic levels, although<br />

several studies have <strong>fo</strong>cused on only one trophic level. One study reported a decline in<br />

plant species richness that coincided with an invasion by creosote bush (Larrea<br />

tridentata; Báez and Collins 2008). Another study documented a negative relationship<br />

between rodent species richness and percent shrub <strong>cover</strong> (Blaum et al. 2006).<br />

The relationship between woody plant <strong>cover</strong> and species richness or diversity<br />

becomes more complicated when multiple trophic levels are considered simultaneously<br />

(e.g., Blaum et al. 2007c, Ceballos et al. 2010). In South Africa, there is a nonlinear,


16<br />

parabolic relationship between percent shrub <strong>cover</strong> and the diversity, assessed using the<br />

Shannon index, of both small carnivores and their prey (Blaum et al. 2007c). The initial<br />

rise in species diversity is driven by the increase in niche diversity, which accompanies a<br />

small increase in shrub <strong>cover</strong>. However, as shrubs become more prevalent, there is a<br />

transition to a relatively homogenous, shrub-dominated environment and thus a decline in<br />

the diversity of both habitat niches and fauna. More specifically, small carnivore diversity<br />

peaks in areas with between 10 and 15% shrub <strong>cover</strong> and prey item diversity peaks in<br />

areas with 12.5 to 17.5% shrub <strong>cover</strong> (Blaum et al. 2007c). In Mexico, it was <strong>fo</strong>und that<br />

species richness of birds, reptiles and small mammals was higher, but the richness of<br />

carnivores lower, in shrubland than grassland habitat (Ceballos et al. 2010).<br />

Changes in animal ecology and predator-prey interactions<br />

Trophic relationships are also affected by an increase in woody plant <strong>cover</strong>.<br />

<strong>Habitat</strong> use patterns of organisms from different trophic levels have been studied in<br />

woody plant-encroached areas. Tree-dwelling lizards were <strong>fo</strong>und to use <strong>live</strong> trees in<br />

savanna areas and dead trees in bush-encroached sites as the dead trees provided hiding<br />

places and prey not available on <strong>live</strong> members of the encroaching woody plant<br />

populations (Meik et al. 2002). Cheetahs in a more open, plains habitat tend to locate<br />

their territories in areas that contain woody plant <strong>cover</strong> (e.g., woodland patches, wooded<br />

drainages; Caro 1994), while cheetahs in woody plant-encroached sites tend to use<br />

habitat patches characterized by longer distances of unobstructed vision and more grass<br />

<strong>cover</strong> (Muntifering et al. 2006). This suggests a shift in habitat use patterns in response to<br />

woody plant encroachment.


17<br />

Woody plant encroachment can have bottom-up effects on the diet of the local<br />

fauna. A study of coyote (Canis latrans) ecology in the southwestern United States<br />

provides an example of a shift in diet in response to an increase in woody plant <strong>cover</strong>.<br />

Over the course of two decades, two woody, fruit-bearing plants (persimmons, Diospyros<br />

texana and agarito ba<strong>rb</strong>erries, Be<strong>rb</strong>eris tri<strong>fo</strong>liolata) became more prevalent and, during<br />

the same time period, the percent of coyote diet composed of persimmons and ba<strong>rb</strong>erries<br />

increased (Andelt et al. 1987). These results indicate the potential <strong>fo</strong>r the diets of<br />

consumers, especially omnivores like the coyote, to change in response to the<br />

introduction of new <strong>fo</strong>od resources that accompanies the spread of woody plants.<br />

The spread of woody vegetation can also have an effect on predator-prey<br />

interactions. Cheetahs are able to hunt in the open and in more wooded areas (Eaton<br />

1974, Hamilton 1986). However, there is evidence that the percentage of attempted hunts<br />

that are successful is higher in more open areas (Mills et al. 2004) and that dense woody<br />

plant <strong>cover</strong> reduces the availability of prey to the local cheetah population (Marker 2002).<br />

There is also a change in the relationship between cheetahs, vegetation <strong>cover</strong> and prey<br />

between open and wooded habitats (Marker 2002). In particular, in open, plains habitat<br />

tall vegetation provides cheetahs with shade and <strong>cover</strong> when stalking prey (Caro 1994,<br />

Fitzgibbon 1990), while in bush-encroached areas, woody plants hinder cheetah<br />

movement and their ability to detect prey (Muntifering et al. 2006) and even cause injury<br />

to the cheetahs, especially to their eyes (Bauer 1998, Marshall 2006).<br />

There is evidence that the hunting strategies of predators that are less specialized<br />

than the cheetah are also affected by the presence of woody vegetation. In particular,<br />

although they are now <strong>fo</strong>und across the United States (Whitaker 2000) and in a variety of


18<br />

habitats (e.g., grassland with scattered shrubs, Hernández et al. 2002; <strong>fo</strong>rest, Dibello et al.<br />

1990), coyotes originated in open, grassland habitats (Young and Jackson 1951) and are<br />

not able to hunt as efficiently in thickly vegetated <strong>fo</strong>rests, in spite of relatively high prey<br />

abundance in these habitats. These factors lead to lower coyote densities and body<br />

reserves in <strong>fo</strong>rested landscapes than more open, agricultural areas (Richer et al. 2002).<br />

Given these observations, it is possible that the conversion of grassland to shrubland or<br />

savanna to <strong>fo</strong>rest via woody plant encroachment could have a detrimental impact on<br />

coyote hunting success, which could in turn impact coyote body condition and fitness.<br />

Carnivores impacted by woody plant encroachment<br />

Locations where woody plant encroachment has been observed (Figure 1) were<br />

combined with in<strong>fo</strong>rmation on the global distributions of carnivores in order to assess the<br />

magnitude of the impact that woody plant encroachment could have on animals,<br />

especially predators, worldwide (Figure 2, Table 2). In particular, a global mammal<br />

dataset compiled by the International Union <strong>fo</strong>r Conservation of Nature (IUCN 2008,<br />

Schipper et al. 2008) was used to obtain lists of species in the order Carnivora whose<br />

distributions overlapped 34 sites where woody plant encroachment has been documented<br />

(Figure 1). Predators are of particular interest as a result of their position at the top of the<br />

<strong>fo</strong>od chain and thus their tendency to be affected by changes in the local habitat, to<br />

impact the populations of animals in lower trophic levels, and to play an important role in<br />

the local ecosystem (Paine 1969, Estes and Duggins 1995, Crooks and Soule 1999,<br />

Gittleman et al. 2001, Smith and Smith 2003).


Figure 2. Numbers of carnivores in woody plant-encroached areas. Map showing<br />

number of species in the order Carnivora which are <strong>fo</strong>und in 34 different locations where<br />

woody plant encroachment has been observed (Figure 1).<br />

A total of 97 species from 9 different families in the order Carnivora are <strong>fo</strong>und in<br />

woody plant-encroached sites. The sites with the highest species richness are located in<br />

southern Africa (Figure 2). The continent with the largest number of species potentially<br />

affected by the habitat changes associated with the spread of woody plants is North<br />

America (34 species), though this is closely <strong>fo</strong>llowed by Africa (33 species; Table 2).<br />

These species include a variety of canids and felids. In addition to coyotes and cheetahs<br />

(discussed above), there are <strong>fo</strong>xes (e.g., red <strong>fo</strong>x, Vulpes vulpes and kit <strong>fo</strong>x, Vulpes<br />

macrotis), gray wolves (Canis lupus), caracals (Caracal caracal), leopards (Panthera<br />

pardus), lions (Panthera leo) and pumas (Puma concolor), to name only a few, present in<br />

woody plant-encroached sites in either North America or Africa. When considering the<br />

geographic distribution of the sites where woody plant encroachment has been observed,<br />

there is a definite bias towards North America (Figure 2). The studies included in this<br />

review are intended as a representative sample, not a comprehensive list, of all studies<br />

that document woody plant encroachment. As a result, there are many locations, and thus<br />

19


20<br />

many carnivores, that are affected by woody plant encroachment but were not considered<br />

here. In particular, in<strong>fo</strong>rmation is needed <strong>fo</strong>r areas <strong>cover</strong>ed by grassland or savanna in the<br />

northern and central portions of both Africa and South America, and in the western and<br />

central part of Asia (Figures 1 and 2).<br />

Table 2. Summary of carnivores in woody plant-encroached areas. This table<br />

provides a summary of carnivores whose distributions overlap at least one of 34 sites that<br />

have been affected by woody plant encroachment (Figures 1 and 2). See Appendix 1 <strong>fo</strong>r a<br />

complete list of sites. The number of species and genera in each family in the order<br />

Carnivora that is present in one or more of the sites affected by woody plant<br />

encroachment are given <strong>fo</strong>r 6 continents. The total number of studies from which<br />

in<strong>fo</strong>rmation on woody plant-encroached sites was drawn is also presented <strong>fo</strong>r each<br />

continent.<br />

Continent<br />

Number of<br />

species<br />

Number of<br />

genera Family<br />

Number of<br />

studies<br />

Africa 4 3 Canidae -<br />

7 5 Felidae -<br />

11 9 Herpestidae -<br />

3 3 Hyaenidae -<br />

5 5 Mustelidae -<br />

3 2 Viverridae -<br />

Total 33 27 6 3<br />

Asia 6 4 Canidae -<br />

5 5 Felidae -<br />

9 5 Mustelidae -<br />

1 1 Ursidae -<br />

Total 21 15 4 4<br />

Australia 0 0 - -<br />

Total 0 0 0 1<br />

Europe 3 2 Canidae -<br />

1 1 Felidae -<br />

5 4 Mustelidae -<br />

Total 9 7 3 1<br />

North America 6 4 Canidae -<br />

5 3 Felidae -<br />

6 3 Mephitidae -<br />

12 7 Mustelidae -<br />

3 3 Procyonidae -<br />

2 1 Ursidae -<br />

Total 34 21 6 12<br />

South America 3 1 Canidae -<br />

5 2 Felidae -<br />

2 1 Mephitidae -<br />

4 3 Mustelidae -<br />

Total 14 7 4 5


In spite of these limitations, these results indicate that the habitat shift associated<br />

with woody plant encroachment has strong conservation implications <strong>fo</strong>r carnivores<br />

globally (Table 2). A need <strong>fo</strong>r woody plant removal and associated improvement in local<br />

habitat quality has been recognized; a bush removal program has been implemented <strong>fo</strong>r a<br />

cheetah population in southern Africa (Marker 2002, Muntifering et al. 2006). Further<br />

study of carnivore populations in woody plant-encroached areas is likely to lead to the<br />

development of similar habitat restoration projects <strong>fo</strong>r other populations and predators.<br />

Conclusion<br />

Woody plant encroachment has been observed, and has led to significant<br />

environmental changes, in grassland and savanna ecosystems around the world. In some<br />

regions, this increase in woody vegetation is associated with habitat degradation and a<br />

decline in the production of <strong>fo</strong>od, especially beef, <strong>fo</strong>r human consumption. The factors<br />

that drive woody plant encroachment vary depending on the scale and geographic region<br />

of interest and many of them are related to human endeavor (e.g., overgrazing by<br />

<strong>live</strong>stock, reduction in fire frequency as a result of fire suppression). The effects of<br />

woody plant encroachment are diverse and impact organisms in all trophic levels. These<br />

effects include 1) a shift from a homogenous to a patchy distribution of water, nutrients<br />

and plant biomass across the landscape; 2) an increase in the structural complexity of the<br />

local habitat; 3) declines in plant and animal species richness; 4) a reduction in the<br />

quality and availability of <strong>fo</strong>od <strong>fo</strong>r local animal populations; and 5) changes in the diet,<br />

habitat use patterns and hunting success of local predators.<br />

21


22<br />

At least 97 carnivores are <strong>fo</strong>und in areas where woody plant encroachment has<br />

been observed. Evidence that the ecology of these animals is affected by woody plant<br />

encroachment, and that this habitat shift has implications <strong>fo</strong>r the conservation and<br />

management of predator populations, is mounting. <strong>Habitat</strong> restoration ef<strong>fo</strong>rts, including<br />

woody plant removal, would benefit some populations. However, detailed studies have<br />

been per<strong>fo</strong>rmed <strong>fo</strong>r only a small number of the mammalian predators <strong>fo</strong>und in woody<br />

plant-encroached areas. Much more research is needed regarding the species-specific and<br />

community-wide effects of woody plant encroachment and the effectiveness of habitat<br />

restoration ef<strong>fo</strong>rts <strong>fo</strong>r local predators.


References<br />

Abril, A. and E.H. Bucher. 2001. Overgrazing and soil ca<strong>rb</strong>on dynamics in the western<br />

Chaco of Argentina. Applied Soil Ecology. 16: 243-249.<br />

Altesor, A., G. Piñeiro, F. Lezama, R.B. Jackson, M. Sarasola, and J.M. Paruelo. 2006.<br />

Ecosystem changes associated with grazing in subhumid South American grasslands.<br />

Journal of Vegetation Science. 17: 323-332.<br />

Andelt, W.F., J.G. Kie, F.F. Knowlton, and K. Cardwell. 1987. Variation in coyote diets<br />

associated with season and successional changes in vegetation. Journal of Wildlife<br />

Management. 51(2): 273-277.<br />

Ansley, R.J., X.B., Wu, and B.A. Kramp. 2001. Observation: Long-term increases in<br />

mesquite canopy <strong>cover</strong> in a North Texas savanna. J. Range Manage. 54: 171-176.<br />

Archer, S. 1989. Have southern Texas savannas been converted to woodlands in recent<br />

history? The American Naturalist. 134(4): 545-561.<br />

Archer, S. 1995. Tree-grass dynamics in a Prosopis-thornscrub savanna parkland:<br />

Reconstructing the past and predicting the future. Ecoscience. 2(1): 83-99.<br />

Archer, S., D.S. Schimel, and E.A. Holland. 1995. Mechanisms of shrubland expansion:<br />

Land use, climate or CO2? Climatic Change. 29: 91-99.<br />

Archer, S., C. Scifres, C.R. Bassham, and R. Maggio. 1988. Autogenic succession in a<br />

subtropical savanna: Conversion of grassland to thorn woodland. Ecological<br />

Monographs. 58(2): 111-127.<br />

Báez, S. and S.L. Collins. 2008. Shrub invasion decreases diversity and alters community<br />

stability in northern Chihuahuan desert plant communities. PLoS ONE. 3(6): e2332.<br />

doi:10.1371/journal.pone.0002332.<br />

Bauer, G.A. 1998. Cheetah: Running Blind. Proceedings of a Symposium on Cheetahs as<br />

Game Ranch Animals. Onderstepoort, Pretoria, South Africa. October 23-24: 106-108.<br />

Bazzaz, F.A. 1990. The response of natural ecosystems to the rising global CO2 levels.<br />

Annual Review of Ecology and Systematics. 21: 167-196.<br />

Beeskow, A.M., N.O. Elissalde, and C.M. Rostagno. 1995. Ecosystem changes<br />

associated with grazing intensity on the Punta Ninfas rangelands of Patagonia, Argentina.<br />

Journal of Range Management. 48: 517-522.<br />

23


24<br />

Berlow, E.L., C.M. D’Antonio, and S.A. Reynolds. 2002. Shrub expansion in montane<br />

meadows: The interaction of local-scale distu<strong>rb</strong>ance and site aridity. Ecological<br />

Applications. 12(4): 1103-1118.<br />

Bertram, B.C.R. 1979. Serengeti predators and their social systems. In Sinclair, A.R.E<br />

and M. Norton-Griffiths (eds). Serengeti: Dynamics of an Ecosystem. Chicago:<br />

University of Chicago Press. Pps. 221-248.<br />

Bester, B. 1996. Bush encroachment: a thorny problem. Namibia Environ. 1: 175-177.<br />

Bhark, E.W. and E.E. Small. 2003. Association between plant canopies and the spatial<br />

patterns of infiltration in shrubland and grassland of the Chihuahuan Desert, New<br />

Mexico. Ecosystems. 6: 185-196.<br />

Blaum, N., E. Rossmanith, G. Fleissner, and F. Jeltsch. 2007a. The conflicting<br />

importance of shrubby landscape structures <strong>fo</strong>r the reproductive success of the yellow<br />

mongoose (Cynictis penicillata). Journal of Mammalogy. 88(1): 194-200.<br />

Blaum, N., E. Rossmanith, and F. Jeltsch. 2006. Land use affects communities in<br />

Kalahari savannah rangelands. African Journal of Ecology. 45: 189-195.<br />

Blaum, N., E. Rossmanith, A. Popp, and F. Jeltsch. 2007b. Shrub encroachment affects<br />

mammalian carnivore abundance and species richness in semiarid rangelands. Acta<br />

Oecologica. 31: 86-92.<br />

Blaum, N., E. Rossmanith, M. Schwager, and F. Jeltsch. 2007c. Responses of mammalian<br />

carnivores to land use in arid savanna rangelands. Basic and Applied Ecology. 8: 552-<br />

564.<br />

Blaum, N., C. Seymour, E. Rossmanith, M. Schwager, and F. Jeltsch. 2009. Changes in<br />

arthropod diversity along a land use driven gradient of shrub <strong>cover</strong> in savanna<br />

rangelands: identification of suitable indicators. Biodiversity Conservation. 18: 1187-<br />

1199.<br />

Bond, W.J. and G.F. Midgley. 2000. A proposed CO2-controlled mechanism of woody<br />

plant invasion in grasslands and savannas. Global Change Biology. 6: 865-869.<br />

Bowman, D.M.J.S., A. Walsh, and D.J. Milne. 2001. Forest expansion and grassland<br />

contraction within a Eucalyptus savanna matrix between 1941 and 1994 at Litchfield<br />

National Park in the Australian monsoon tropics. Global Ecology and Biogeography. 10:<br />

535-548.<br />

Broomhall, L.S. 2001. Cheetah Acinonyx jubatus ecology in the Kruger National Park: A<br />

comparison with other studies across the grassland-woodland gradient in African<br />

savannas. Unpublished M.Sc. thesis, University of Pretoria, South Africa. Pps. 51-78.


Brown, J.R. and S. Archer. 1999. Shrub invasion of grassland: Recruitment is continuous<br />

and not regulated by he<strong>rb</strong>aceous biomass or density. Ecology. 80(7): 2385-2396.<br />

Brown, J.R. and J. Carter. 1998. Spatial and temporal patterns of exotic shrub invasion in<br />

an Australian tropical grassland. Landscape Ecology. 13: 93-102.<br />

Brown, J.H., T.J. Valone, and C.G. Curtin. 1997. Reorganization of an arid ecosystem in<br />

response to recent climate change. Proceedings of the National Academy of Sciences of<br />

the United States of America. 94: 9729-9733.<br />

Buffington, L.C. and C.H. He<strong>rb</strong>el. 1965. Vegetational changes on a semidesert grassland<br />

range from 1858 to 1963. Ecological Monographs. 35(2): 139-164.<br />

Caro, TM. 1994. Cheetahs of the Serengeti plains: Group living in an asocial species.<br />

Chicago: University of Chicago Press. Pps.199-231.<br />

Ceballos, G., A. Davidson, R. List, J. Pacheco, P. Manzano-Fischer, G. Santo-Barrera,<br />

and J. Cruzado. 2010. Rapid decline of a grassland system and its ecological and<br />

conservation implications. PLoS ONE. 5(1): e8562. doi:10.1371/journal.pone.0008562.<br />

Chapin III, F.S., P.A. Matson, and H.A. Mooney. 2002. Principles of Terrestrial<br />

Ecosystem Ecology. New York: Springer. Pps. 97-122.<br />

Chen, S., Y. Bai, G. Lin, and X. Han. 2005. Variations in life-<strong>fo</strong>rm composition and<br />

<strong>fo</strong>liar ca<strong>rb</strong>on isotope discrimination among eight plant communities under different soil<br />

moisture conditions in the Xilin River Basin, Inner Mongolia, China. Ecological<br />

Research. 20: 167-176<br />

Costello, D.A., I.D. Lunt, and J.E. Williams. 2000. Effects of invasion by the indigenous<br />

shrub Acacia sophorae on plant composition of coastal grasslands in south-eastern<br />

Australia. Biological Conservation. 96: 113-121.<br />

Crooks, K.R. and M.E. Soule. 1999. Mesopredator release and avifaunal extinctions in a<br />

fragmented system. Nature. 400: 563-566.<br />

de Villalobos, A.E., D.V. Peláez, and O.R. Elia. 2005. Factors related to establishment of<br />

Prosopis caldenia seedlings in central rangelands of Argentina. Acta Oecologica. 27: 99-<br />

106.<br />

Dibello, F.J., S.M. Arthur, and W.B. Krohn. 1990. Food habits of sympatric coyotes,<br />

Canis latrans, red <strong>fo</strong>xes, Vulpes vulpes, and bobcats, Lynx rufus, in Maine. The Canadian<br />

Field-Naturalist. 104: 403-408.<br />

25


26<br />

D’Odorico, P. and A. Porporato. 2006. Ecohydrology of arid and semiarid ecosystems:<br />

An introduction. In P.D’Odroico and A. Porporato (eds). Dryland Ecohydrology. The<br />

Netherlands: Springer. Pps. 1-10.<br />

Eaton, R.L. 1974. The cheetah: Biology, ecology and behaviour of an endangered<br />

species. New York: Van Nostrand Reinhold. Pps. 41-87<br />

Estes, J.A. and D.O. Duggins. 1995. Sea otters and kelp <strong>fo</strong>rests in Alaska: Generality and<br />

variation in a community ecological paradigm. Ecological Monographs. 65 (1): 75-100.<br />

Fensham, R.J., R.J. Fairfax, and S.R. Archer. 2005. Rainfall, land use and woody<br />

vegetation <strong>cover</strong> change in semi-arid Australian savanna. Journal of Ecology. 68: 596-<br />

606.<br />

Fitzgibbon, C.D. 1990. Why do hunting cheetahs prefer male gazelles? Animal<br />

Behaviour. 40: 837-845.<br />

Ghermandi, L., M.I. de Torres Curth, J. Franzese, and S. Gonzalez. 2010. Non-linear<br />

ecological processes, fires, environmental heterogeneity and shrub invasion in<br />

northwestern Patagonia. Ecological Modelling. 221: 113-121.<br />

Gill, R.A. and I.C. Burke. 1999. Ecosystem consequences of plant life <strong>fo</strong>rm changes at<br />

three sites in the semiarid United States. Oecologia. 121: 551-563.<br />

Gittleman, J.L., S.M. Funk, D.W. Macdonald, and R.K. Wayne. 2001. Why ‘carnivore<br />

conservation’? In J.L.Gittleman, S.M. Funk, D.W. Macdonald, and R.K. Wayne (eds).<br />

Carnivore Conservation. Cambridge: Cambridge University Press. Pps. 1-7.<br />

Goslee, S.C., K.M. Havstad, D.P.C. Peters, A. Rango, and W.H. Schlesinger. 2003. Highresolution<br />

images reveal rate and pattern of shrub encroachment over six decades in New<br />

Mexico, U.S.A. Journal of Arid Environments. 54: 755-767.<br />

Hagenah, N., H. Munkert, K. Gerhardt, and H. Olff. 2009. Interacting effects of grass<br />

height and he<strong>rb</strong>ivores on the establishment of an encroaching savanna shrub. Plant<br />

Ecology. 201: 553-566.<br />

Hamilton, P.H. 1986. Status of the cheetah in Kenya, with reference to Sub-Saharan<br />

Africa. In Miller, S.D. and D.D. Everett (eds). Cats of the world: Biology, conservation<br />

and management. Washington, DC: National Wildlife Federation. Pgs. 65-76.<br />

He, Y., P. D’Odorico, S.F.J. De Wekker, J.D. Fuentes, and M. Litvak. 2010. On the<br />

impact of shrub encroachment on microclimate conditions in the northern Chihuahuan<br />

desert. Journal of Geophysical Research. 115: doi:10.1029/2009JD013529.


27<br />

Hernández, L., R.R. Parmenter, J.W. Dewitt, D.C. Light<strong>fo</strong>ot, and J.W. Laundré. 2002.<br />

Coyote diets in the Chihuahuan Desert, more evidence <strong>fo</strong>r optimal <strong>fo</strong>raging. Journal of<br />

Arid Environments. 51: 613-624.<br />

Hernández, L., A.G. Romero, J.W. Laundré, D. Light<strong>fo</strong>ot, E. Aragón, and J. López<br />

Portillo. 2005. Changes in rodent community structure in the Chihuahuan Desert Mexico:<br />

Comparisons between two habitats. Journal of Arid Environments. 60: 239-257.<br />

Hochstrasser, T. and D.P.C. Peters. 2004. Subdominant species distribution in microsites<br />

around two life <strong>fo</strong>rms at a desert grassland-shrubland transition zone. Journal of<br />

Vegetation Science. 15: 615-622.<br />

Hudak, A.T. and C.A. Wessman. 2001. Textural analysis of high resolution imagery to<br />

quantify bush encroachment in Madikwe Game Reserve, South Africa, 1955-1996.<br />

International Journal of Remote Sensing. 22(14): 2731-2740.<br />

Huenneke, L.F., J.P. Anderson, M. Remmenga, and W.H. Schlesinger. 2002.<br />

Desertification alters patterns of aboveground net primary production in Chihuahuan<br />

ecosystems. Global Change Biology. 8:247-264.<br />

Idso, S.B. 1992. Shrubland expansion in the American southwest. Climatic Change. 22:<br />

85-86.<br />

IUCN. 2008. 2008 IUCN Red List of Threatened Species. http://www.iucnredlist.org<br />

Jin, Z., Y. Qi, Y. Dong, and M. Domroes. 2009. Seasonal patterns of soil respiration in<br />

three <strong>type</strong>s of communities along grass-desert shrub transition in Inner Mongolia, China.<br />

Advances in Atmospheric Sciences. 26(3): 503-512.<br />

Johnson, H.B., H.W. Polley, and H.S. Mayeux. 1993. Increasing CO2 and plant-plant<br />

interactions: effects on natural vegetation. Vegetatio. 104/105: 157-170.<br />

Kidron, G.J. 2009. The effect of shrub canopy upon surface temperatures and evaporation<br />

in the Negev Desert. Earth Surface Processes and Land<strong>fo</strong>rms. 34: 123-132.<br />

Kitzberger, T., D.F. Steinaker, and T.T. Veblen. 2000. Effects of climatic variability on<br />

facilitation of tree establishment in northern Patagonia. Ecology. 81(7): 1914-1924.<br />

Knapp, A.K., J.M. Briggs, S.L. Collins, S.R. Archer, M.S. Bret-Harte, B.E. Ewers, D.P.<br />

Peters, D.R. Young, G.R. Shaver, E. Pendall, and M.B. Cleary. 2008. Shrub<br />

encroachment in North American grasslands: shifts in growth <strong>fo</strong>rm dominance rapidly<br />

alters control of ecosystem ca<strong>rb</strong>on inputs. Global Change Biology. 14: 615-623.<br />

Köchy, M. and S.D. Wilson. 2001. Nitrogen deposition and <strong>fo</strong>rest expansion in the<br />

northern Great Plains. Journal of Ecology. 89: 807-817.


Lett, M. and A.K. Knapp. 2003. Consequences of shrub expansion in mesic grassland:<br />

Resource alterations and graminoid responses. Journal of Vegetation Science. 14: 487-<br />

496.<br />

Li, X.R., X.H. Jia, and G.R. Dong. 2006. Influence of desertification on vegetation<br />

pattern variations in the cold semi-arid grasslands of Qinghai-Tibet Plateau, North-west<br />

China. Journal of Arid Environments. 64: 505-522.<br />

Li, X.R., Z.S. Zhang, J.G. Zhang, X.P. Wang, and X.H. Jia. 2004. Association between<br />

vegetation patterns and soil properties in the southeastern Tengger Desert, China. Arid<br />

Land Research and Management. 18: 369-383.<br />

Maestre, F.T., J.F. Reynolds, E. Huber-Sannwald, J. Herrick, and M. Staf<strong>fo</strong>rd Smith.<br />

2006. Understanding global desertification: Biophysical and socioeconomic dimensions<br />

of hydrology. In P.D’Odroico and A. Porporato (eds). Dryland Ecohydrology. The<br />

Netherlands: Springer. Pps. 315-332.<br />

Marker, L. 2002. Aspects of Namibian cheetah (Acinonyx jubatus) biology, ecology and<br />

conservation strategies. PhD. Thesis, Department of Zoology, University of Ox<strong>fo</strong>rd. Pps.<br />

199-242, 328-344, 397-417.<br />

Marshall, L. 2006. Cheetahs suffering more fatal crashes during chases, experts say.<br />

National Geographic News. <br />

(5 July 2010).<br />

McGlynn, I.O. and G.S. Okin. 2006. Characterization of shrub distribution using high<br />

spatial resolution remote sensing: Ecosystem implications <strong>fo</strong>r a <strong>fo</strong>rmer Chihuahuan<br />

Desert grassland. Remote Sensing of Environment. 101: 554-566.<br />

Meik, J.M., R.M. Jeo, J.R. Mendelson III, and K.E. Jenks. 2002. Effects of bush<br />

encroachment on an assemblage of diurnal lizard species in central Namibia. Biological<br />

Conservation. 106: 29-36.<br />

Mills, M.G.L., L.S. Broomhall, and J.T. du Toit. 2004. Cheetah Acinonyx jubatus feeding<br />

ecology in the Kruger National Park and a comparison across African savanna habitats: is<br />

the cheetah only a successful hunter on open grassland plains? Wildlife Biology. 10(3):<br />

177-186.<br />

Moleele, N.M., S. Ringrose, W. Matheson, and C. Vanderpost. 2002. More woody<br />

plants? the status of bush encroachment in Botwana’s grazing areas. Journal of<br />

Environmental Management. 64: 3-11.<br />

Moran, M.S., R.L. Scott, T.O. Keefer, W.E. Emmerich, M. Hernandez, G.S. Nearing,<br />

G.B. Paige, M.H. Cosh, and P.E. O’Neill. 2009. Partitioning evapotranspiration in<br />

28


29<br />

semiarid grassland and shrubland ecosystems using time series of soil surface<br />

temperature. Agricultural and Forest Meteorology. 149: 59-72.<br />

Muntifering, J.R., A.J. Dickman, L.M. Perlow, T. Hruska, P.G. Ryan, L.L. Marker, and<br />

R.M. Jeo. 2006. Managing the matrix <strong>fo</strong>r large carnivores: a novel approach and<br />

perspective from cheetah (Acinonyx jubatus) habitat suitability modeling. Animal<br />

Conservation. 9: 103-112.<br />

Neilson, R.P. 1986. High-resolution climatic analysis and southwest biogeography.<br />

Science. 232(4746):27-34.<br />

Ojima, D.S., et al. 1996. Impact of climate and atmospheric ca<strong>rb</strong>on dioxide changes on<br />

grasslands of the world. In A.I. Breymeyer, D.O. Hall, J.M. Melillo, and G.I. Agren<br />

(eds). Global Change: Effects on Coniferous Forests and Grasslands. Chichester: John<br />

Wiley & Sons Ltd. Pps. 271-311.<br />

Olson, D.M., E. Dinerstein, E.D. Wikramanayake, N.D. Burgess, G.V.N. Powell, E.C.<br />

Underwood, J.A. D’Amico, H.E. Strand, J.C. Morrison, C.J. Loucks, T.F. Allnutt, J.F.<br />

Lamoreux, T.H. Ricketts, I. Itoua, W.W. Wettengel, Y. Kura, P. Hedao, and K. Kassem.<br />

2001. Terrestrial ecoregions of the world: A new map of life on Earth. BioScience.<br />

51(11): 933-938. < http://www.worldwildlife.org/science/data/item1875.html > (29 July<br />

2010).<br />

Paine, R.T. 1969. A note on trophic complexity and community stability. The American<br />

Naturalist. 103 (929): 91-93.<br />

Palmer, A.R. and A.F. van Rooyen. 1998. Detecting vegetation change in the southern<br />

Kalahari using Landsat TM data. Journal of Arid Environments. 39: 143-153.<br />

Pierce, A.M. and P.B. Reich. 2010. The effects of eastern red cedar (Juniperus<br />

virginiana) invasion and removal on a dry bluff prairie ecosystem. Biological Invasions.<br />

12: 241-252.<br />

Rappole, J.H., C.E. Russell, J.R. Norwine, and T.E. Fulbright. 1986. Anthropogenic<br />

pressures and impacts on marginal, neotropical, semiarid ecosystems: The case of south<br />

Texas. The Science of the Total Environment. 55: 91-99.<br />

Ravi, S., P. D’Odorico, L. Wang, C.S. White, G.S. Okin, S.A. Macko, and S.L. Collins.<br />

2009. Post-fire resource redistribution in desert grasslands: A possible negative feedback<br />

on land degradation. Ecosystems. 12: 434-444.<br />

Richer, M.C., M. Crete, and J.P. Ouellet. 2002. The low per<strong>fo</strong>rmance of <strong>fo</strong>rest versus<br />

rural coyotes in northeastern North America: Inequality between presence and<br />

availability of prey. Ecoscience. 9(1): 44-54.


30<br />

Roques, K.G., T.G. O'Connor, and A.R. Watkinson. 2001. Dynamics of shrub<br />

encroachment in an African savanna: relative influences of fire, he<strong>rb</strong>ivory, rainfall and<br />

density dependence. Journal of Applied Ecology. 38:268-280.<br />

Rosenfeld, D., Y. Rudich, and R. Lahav. 2001. Desert dust suppressing precipitation: A<br />

possible desertification feedback loop. Proceedings of the National Academy of Sciences.<br />

98(11): 5975-5980.<br />

Sánchez, L.F., J. García-Miragaya, and N. Chacón. 1997. Nitrogen mineralization in soils<br />

under grasses and under trees in a protected Venezuelan savanna. Acta Ecologica. 18(1):<br />

27-37.<br />

Schipper, J. et al. 2008. The status of the world’s land and marine mammals: Diversity,<br />

threat and knowledge. Science. 322: 225-230.<br />

Schlesinger, W.H., J.F. Reynolds, G.L. Cunningham, L.F. Huenneke, W.M. Jarrell, R.A.<br />

Virginia, and W.G. Whit<strong>fo</strong>rd. 1990. Biological feedbacks in global desertification.<br />

Science, New Series. 247(4946): 1043-1048<br />

Schwinning, S. 2008. The water relations of two evergreen tree species in a karst<br />

savanna. Oecologia. 158: 373-383.<br />

Shugart, H.H. 1998. Terrestrial Ecosystems in Changing Environments. Cambridge:<br />

Cambridge University Press. Pps. 343-381.<br />

Silva, J.F., A. Zambrano, and M.R. Fariñas. 2001. Increase in the woody component of<br />

seasonal savannas under different fire regimes in Calabozo, Venezuela. Journal of<br />

Biogeography. 28: 977-983.<br />

Sirami, C., C. Seymour, G. Midgley, and P. Barnard. 2009. The impact of shrub<br />

encroachment on savanna bird diversity from local to regional scale. Diversity and<br />

Distributions. 15: 948-957.<br />

Smith, R.L. and T.M. Smith. 2003. Elements of Ecology. 5 th Ed. San Francisco: Benjamin<br />

Cummings. Pps. 289-308, 329-347.<br />

Steinaker, D.F. and S.D. Wilson. 2008. Phenology of fine roots and leaves in <strong>fo</strong>rest and<br />

grassland. Journal of Ecology. 96:1222-1229.<br />

Sunquist, M.E. and F. Sunquist. 2001. Changing landscapes: Consequences <strong>fo</strong>r<br />

carnivores. In J.L.Gittleman, S.M. Funk, D.W. Macdonald, and R.K. Wayne (eds).<br />

Carnivore Conservation. Cambridge: Cambridge University Press. Pps. 399-418.


31<br />

Tews, J., F. Schurr, and F. Jeltsch. 2004. Seed dispersal by cattle may cause shrub<br />

encroachment of Grewia flava on southern Kalahari rangelands. Applied Vegetation<br />

Science. 7: 89-102.<br />

UNCCD. 1994. United Nations Convention to Combat Desertification in Countries<br />

Experiencing Serious Drought and/or Desertification, Particularly in Africa. (Part I,<br />

Article 1; , 20 July<br />

2010).<br />

UNEP. 1992. World Atlas of Desertification. London: Edward Arnold. Pps. iv-ix, 1-25.<br />

USDA, NRCS. 2010. The PLANTS Database (, 14 July 2010).<br />

National Plant Data Center, Baton Rouge, LA 70874-4490 USA.<br />

Van Auken, O.W. 2000. Shrub invasions of North American semiarid grasslands. Annual<br />

Review of Ecology and Systematics. 31: 197-215.<br />

Walker, B.H., D. Ludwig, C. S. Holling, and R. M. Peterman. 1981. Stability of semi-arid<br />

savanna grazing systems. The Journal of Ecology. 69(2): 473-498.<br />

Wheeler, C.W., S.R. Archer, G.P. Asner, and C.R. McMurtry. 2007. Climatic/edaphic<br />

controls on soil ca<strong>rb</strong>on/nitrogen response to shrub encroachment in desert grassland.<br />

Ecological Applications. 17(7): 1911-1928.<br />

Whitaker, J.O. 2000. National Audubon Society Field Guide to North American<br />

Mammals. 2 nd Ed. Hong Kong: Alfred A. Knopf. Pps. 684-686<br />

Wigley, B.J., W.J. Bond, and M.T. Hoffman. 2010. Thicket expansion in a South African<br />

savanna under divergent land use: local vs. global drivers? Global Change Biology. 16:<br />

964-976.<br />

Young, S.P. and H.H.T. Jackson. 1951. The Clever Coyote. Harrisburg: The Stackpole<br />

Company. Pps. 11-46.<br />

Zarovali, M.P., M.D. Yiakoulaki, and V.P. Papanastasis. 2007. Effects of shrub<br />

encroachment on he<strong>rb</strong>age production and nutritive value in semi-arid Mediterranean<br />

grasslands. Grass and Forage Science. 62: 355-363.


Appendix 1. Woody plant encroachment references.<br />

Study Latitude Longitude Continent Country/State Rainfall (mm/yr) Original habitat<br />

Tews et al. 2004 -28.6167 24.8 AF South Africa 417 savanna<br />

Palmer and van Rooyen 1998 -26.125 20.125 AF South Africa 160-240 savanna<br />

Roques et al. 2001 -26.275 31.89167 AF Swaziland 675 savanna<br />

Li et al. 2006 35.7 100.83 AS China 367.9 grassland/steppe<br />

Li et al. 2004 37.483 104.767 AS China 180.2 grassland<br />

Chen et al. 2005 43.9583 116.367 AS China 250-400 grassland<br />

Jin et al. 2009 38.65 109.4167 AS China 345 grassland/steppe<br />

Costello et al. 2000 -36.0125 150.161 AU New South Wales 968 grassland<br />

Costello et al. 2000 -35.99583 150.161 AU New South Wales 968 grassland<br />

Zarovali et al. 2007 40.783 23.2 EU Greece 586 grassland<br />

Knapp et al. 2008 68.3 -106.8 NA AK 291 grassland<br />

Wheeler et al. 2007 31.9 -110.8 NA AZ 275-450 grassland<br />

Berlow et al. 2002 36 -118 NA CA 500 montane meadow<br />

Steinaker and Wilson 2008 50.467 -104.367 NA Canada 388 grassland<br />

Knapp et al. 2008 39.1 -96.4 NA KS 859 grassland<br />

van Auken 2000 18.5 -100 NA Mexico 230-600 grassland/savanna<br />

Ceballos et al. 2010 30.83 -108.4 NA Mexico 287 grassland<br />

Pierce and Reich 2010 44.505 -92.485 NA MN > 590 prairie<br />

Pierce and Reich 2010 44.185 -91.99 NA MN > 590 prairie<br />

Goslee et al. 2003 32.53 -106.84 NA NM 230 grassland<br />

Hochstrasser and Peters 2004 34.35 -106.883 NA NM 232 grassland<br />

Brown and Archer 1999 27.67 -98.2 NA TX 720 savanna<br />

Ansley et al. 2001 33.85 -99.43 NA TX 665 savanna<br />

Schwinning 2008 29.94167 -98.12 NA TX 800 savanna<br />

van Auken 2000 33 -104.5 NA TX,NM,AZ 230-600 grassland/savanna<br />

Knapp et al. 2008 37.3 -75.9 NA VA 1065 grassland<br />

Knapp et al. 2008 41.2 -107.2 NA WY 259 grassland<br />

Ghermandi et al. 2010 -41.05 -71.0167 SA Argentina 582 grassland<br />

Ghermandi et al. 2010 -41.167 -70.683 SA Argentina 308 grassland<br />

Kitzberger et al. 2000 -41.1 -71.23 SA Argentina 1100 steppe<br />

Kitzberger et al. 2000 -41.7167 -71 SA Argentina 900 steppe<br />

32


Appendix 1. Woody plant encroachment references.<br />

Study Latitude Longitude Continent Country/State Rainfall (mm/yr) Original habitat<br />

de Villalobos et al. 2005 -38.75 -63.75 SA Argentina 400 rangeland<br />

Beeskow et al. 1995 -43 -64.5 SA Argentina 254 steppe<br />

Altesor et al. 2006 -31.9 -58.25 SA Uruguay 1099 grassland<br />

Appendix 1. The references listed in this Appendix were the sources of the locations shown on the map in Figure 1 and a subset of<br />

these references (25 total; all but van Auken 2000) were used to generate the summary provided in Table 1. First, a list of 57<br />

references compiled by Sujith Ravi and utilized in Ravi et al. (2009) were obtained via web of science and checked <strong>fo</strong>r the<br />

requirements outlined below. Then a search was per<strong>fo</strong>rmed on web of science (key word “woody plant encroachment”, time period<br />

2008-2010, 43 studies returned, “study” shown in bold). Any study that was not available via web of science or that dealt with the<br />

spread of either exotic or non-native species, rather than species that were native and had been present in the region historically, were<br />

omitted. Studies that did not indicate that woody plants had been increasing in the area where the study was conducted were also<br />

omitted. Furthermore, studies had to indicate that the documented increase in woody plant <strong>cover</strong> was occurring in either a grassland or<br />

savanna habitat in an area characterized by one of the <strong>fo</strong>llowing climates: arid, semiarid, dry sub-humid. Studies that dealt with<br />

woodland thickening, succession, or re<strong>fo</strong>restation of pasture or that had been per<strong>fo</strong>rmed in an area with a climate that was wetter than<br />

dry sub-humid (> 1200 mm/yr) were omitted. Finally, studies that did not contain in<strong>fo</strong>rmation on any of the variables listed in Table 1,<br />

did not provide coordinates or in<strong>fo</strong>rmation on average annual rainfall <strong>fo</strong>r the study area, or were per<strong>fo</strong>rmed at the same site (e.g.,<br />

estate, ranch, refuge, research area/center, etc.) as another study that was already listed were omitted. Continent abbreviations are as<br />

<strong>fo</strong>llows: AF = Africa, AS = Asia, AU = Australia, EU = Europe, NA = North America, SA = South America. State abbreviations are<br />

as <strong>fo</strong>llows: AK = Alaska, AZ = Arizona, CA = Cali<strong>fo</strong>rnia, KS = Kansas, MN = Minnesota, NM = New Mexico, TX = Texas, VA =<br />

Virginia, WY = Wyoming. Rainfall (mm/yr) = average annual rainfall in millimeters. For Figure 1, please note that the grassland and<br />

savanna regions are based on a Terrestrial Ecoregions dataset from the World Wildlife Fund (Olson et al. 2001). In particular,<br />

ecoregions <strong>fo</strong>r which annual rainfall data were available, receive less than 1200 mm of rain per year, and are associated with one of<br />

<strong>fo</strong>ur biomes, each of which had the word “grassland” or “savanna” in their name, were included on the map (i.e., Tropical and<br />

subtropical grasslands, savannas and shrublands; Temperate grasslands, savannas and shrublands; Flooded grasslands and savannas;<br />

Montane grasslands and shrublands).<br />

33


Chapter 2: Dissertation methods<br />

This chapter provides a detailed description of the methods used to produce the<br />

results that are presented in later chapters of this dissertation. Given the diversity of field,<br />

laboratory and computer-based techniques employed in this dissertation, this chapter is a<br />

useful reference <strong>fo</strong>r subsequent chapters. One of the goals of this dissertation is to<br />

determine whether woody plant encroachment, or the transition from grassland to<br />

shrubland habitat, has led to a shift in the feeding ecology of one mammalian predator,<br />

the coyote (Canis latrans). The coyote is one of the top predators at a refuge in the<br />

southwestern US where shrubs have been moving northward into a grassland habitat over<br />

the past century. Techniques described in this chapter were used to assess and compare<br />

the base of the coyote <strong>fo</strong>od chain between grassland and shrubland sites at this refuge,<br />

thereby evaluating the impact that woody plant encroachment has had on coyote feeding<br />

ecology.<br />

Study site<br />

The fieldwork <strong>fo</strong>r this dissertation was carried out at the <strong>Sevilleta</strong> National<br />

Wildlife Refuge (NWR) and Long Term Ecological Research (<strong>LTER</strong>) site in central New<br />

Mexico. The refuge encompasses 1,000 km 2 (Hernández et al. 2002), has roughly 200<br />

miles of road, and contains grassland, shrubland and woodland habitat <strong>type</strong>s (Figure 1).<br />

The refuge is bounded by private, ranchland to the north and land owned by the Bureau<br />

of Land Management to the east. There are small communities to the north and south and<br />

a mixture of land ownership along the Rio Grande, which flows south, through the center<br />

of the refuge. The river divides the refuge into eastern and western parts that have<br />

distinct plant communities.<br />

34


Figure 1. Study site map. Map of the <strong>Sevilleta</strong> NWR and <strong>LTER</strong> study site in central<br />

New Mexico, USA (based on Muldavin et al. 1998). Black = shrubland, gray = grassland,<br />

white = other land <strong>cover</strong> <strong>type</strong>s. The path of the Rio Grande through the center of the<br />

refuge is indicated by a thick black line, while the boundary of the refuge is shown by a<br />

light gray line.<br />

On the eastern site of the refuge, and thus to the east of the Rio Grande, the grassland is<br />

dominated by grama grass (Bouteloua spp.), with threeawn (Aristida spp.), muhly<br />

(Muhlenbergia spp.), James’ galleta (Pleuraphis jamesii), burro (Scleropogon<br />

brevi<strong>fo</strong>lius), and dropseed (Sporobolus spp.) grasses also present. The grass <strong>cover</strong> on the<br />

western side is sparse relative to the eastern side. The grasses present in the grassland on<br />

the western side include low woollygrass (Dasyochloa pulchella) and representatives of<br />

all of the genera listed <strong>fo</strong>r the eastern side except Mulenbergia spp.<br />

The shrubland on the eastern side of the refuge is dominated by creosote (Larrea<br />

tridentata), while honey mesquite (Prosopis glandulosa) is dominant on the western side.<br />

Desert willow (Chilopsis linearis), Torrey’s jointfir (Ephedra torreyana), pale desert-<br />

35


36<br />

thorn (Lycium pallidum), and yucca (Yucca spp.) are also present on the eastern side,<br />

while <strong>fo</strong>urwing saltbush (Atriplex canescens), creosote, and broom dalea (Psorothamnus<br />

scoparius) are present on the western side. Broom snakeweed (Gutierrezia sarothrae),<br />

tree cholla (Cylindropuntia imbricata), and winterfat (Kraschenlnnikovia lanata) are<br />

present on both the eastern and western sides of the refuge.<br />

In general, grassland is more prevalent in the northern half of the refuge and<br />

shrubland in the southern half. Juniper (Juniperus spp.) savanna or pinyon-juniper (Pinus<br />

spp.-Juniperus spp.) woodland is present in higher elevation areas that include the<br />

mountain ranges running along the eastern and western-most edges of the refuge (Figure<br />

1). Shrubland areas on the eastern side of the refuge have grasses growing between and<br />

sometimes beneath the shrubs. Grass is very sparse on the southwestern side of the refuge<br />

and there are large, honey mesquite-<strong>cover</strong>ed sand dunes. In this way, the northeastern<br />

(dominated by C4 grasses) and southwestern (dominated by C3 woody plants) parts of the<br />

refuge are most distinct in terms of their species composition and habitat structure.<br />

Approximately 72% of the refuge is <strong>cover</strong>ed by either grassland or shrubland and<br />

there is an active transition zone between grama grassland and creosote shrubland areas<br />

on the eastern side of the refuge (Figure 1). More specifically, creosote shrubs have been<br />

moving into grama grassland areas over the past century (Gill and Burke 1999). These<br />

characteristics make the refuge an ideal setting <strong>fo</strong>r an assessment of the impact that<br />

woody plant encroachment has on coyote ecology.


Field data collection<br />

Field work <strong>fo</strong>r this study consists of two main parts: 1) coyote scat identification<br />

and collection along 1 mile long, road-based transects; and 2) surveys of habitat variables<br />

within circular (diameter = 30 m) vegetation plots.<br />

Scat surveys<br />

Three scat surveys were carried out at the <strong>Sevilleta</strong> NWR and <strong>LTER</strong> from June to<br />

July, 2008. Each survey involved the collection of carnivore scat along 20 road-based<br />

transects. Half of these transects were located in grassland habitat and half in shrubland<br />

habitat in order to allow <strong>fo</strong>r a comparison of coyote ecology between habitats (Figure 2).<br />

Figure 2. Scat transect map. Map of scat transect locations along roads (dotted light<br />

gray lines) at the <strong>Sevilleta</strong> NWR (refuge boundary in solid light gray). The two transects<br />

marked by stars were only surveyed in 2009. Half (n = 10) of the transects surveyed in<br />

2008 were located in grassland (dark gray) and half (n = 10) were located in shrubland<br />

(black) habitat.<br />

Each scat transect was one mile (1.6 km) long and was separated from all other transects<br />

by at least one mile (1.6 km) in an ef<strong>fo</strong>rt to ensure independence of scat samples<br />

collected on different transects (Roughton and Sweeny 1982). These transects <strong>cover</strong><br />

37


38<br />

roughly 10% of all of the roads on the refuge. The beginning and end of each scat<br />

transect were marked with a wooden stake to ensure that all surveys were per<strong>fo</strong>rmed<br />

along the same road segments. In an ef<strong>fo</strong>rt to standardize the period over which the scats<br />

were deposited on a given transect, old scats were removed from all transects be<strong>fo</strong>re the<br />

surveys began and surveys of a given transect were separated by 7 to 16 days.<br />

Surveys were carried out by driving along a given transect at slow speed (1-10<br />

mi/hr; 1.6-16.1 km/hr; Hernández et al. 2002, Parmenter 2004). When an item could not<br />

be identified from within the field vehicle, the observer would get out of the vehicle.<br />

Each scat sample encountered was identified and photographed and the latitude and<br />

longitude of the location of the sample were recorded. Scat was identified to species<br />

based on length, the average of two measurements of maximum diameter (based on Reed<br />

et al. 2004) and morphology. A sample was identified as being from a coyote if its<br />

average maximum diameter was between 1.80 and 3.30 cm, the length of the longest<br />

piece in the sample was between 6.4 and 22.9 cm and at least one end of one of the pieces<br />

in the sample was tapered (Green and Flinders 1981, Murie 1982, Halfpenny 2000). If a<br />

sample did not meet one of these three criteria, it was labelled as being “maybe coyote”<br />

and if it met only one or none of these criteria it was labelled as “not coyote.” Samples<br />

<strong>fo</strong>r which accurate length and diameter measurements could not be obtained, <strong>fo</strong>r example<br />

because all of the pieces in the sample had been run over by a vehicle or degraded by<br />

insects or the elements, were labelled as “unknown” and were photographed and their<br />

coordinates were recorded. Finally, a rough assessment of the freshness of each scat<br />

sample was made (Reed et al. 2004, Stenglein et al. 2010b). The assumption was that<br />

samples that were still soft and had little to no color variation were very fresh and


39<br />

samples that were mostly white and hardened were old. Samples of intermediate<br />

freshness had some color variation and were hardened, but were not mostly white.<br />

Once identified and photographed, a small subsample (roughly 0.4 mL) of the<br />

fecal material on the outside of each sample was placed in a 2 mL tube containing DETs<br />

(DMSO, EDTA, Tris, salt) buffer (Frantzen et al. 1998) in order to preserve it <strong>fo</strong>r genetic<br />

analysis (see genetics section below). Taking samples from the outside of the scats<br />

enhanced the probability of collecting material that contained epithelial cells, and<br />

there<strong>fo</strong>re DNA, from the lining of the gut of the predator that had deposited the scat and<br />

increased the likelihood that the genetic analysis of the samples would be successful<br />

(Stenglein et al. 2010a). DETs buffer was prepared and scat subsamples were collected in<br />

accordance with protocols developed by the Laboratory <strong>fo</strong>r Conservation and Ecological<br />

Genetics (LCEG) at the University of Idaho. For samples consisting of multiple pieces, a<br />

small amount of the fecal material from each piece was put into the 2 mL tube. This was<br />

done to increase the probability that mixed samples, or samples deposited by more than<br />

one individual, would be identified and eliminated from further analysis. All 2 mL tubes<br />

were placed in a light reflecting Styro<strong>fo</strong>am container in order to reduce the rate of DNA<br />

degradation. The tweezers used to collect the subsamples were cleaned with an alcohol<br />

swab and then flame sterilized in order to mitigate cross-sample contamination. The<br />

remainder of each scat sample was dried in a drying oven <strong>fo</strong>r 24 hrs at 70 o C and stored in<br />

a plastic bag with a desiccant until the stable ca<strong>rb</strong>on isotope signature of any bone, hair or<br />

vegetation in the scat could be determined (see ca<strong>rb</strong>on stable isotopes section below).<br />

A running tally was kept of the <strong>fo</strong>llowing: 1) all scats that were torn or very small<br />

relative to other coyote scat samples and were there<strong>fo</strong>re identified as “incomplete”, 2)


40<br />

scats that were worn down, often to the point that they were composed entirely of hair<br />

from a prey item, and appeared to be “old,” such that they had been missed when the<br />

transects were cleared at the beginning of the field season, 3) scats that could have been<br />

deposited during the survey period but did not appear to have any fecal material that<br />

might contain epithelial cells, and there<strong>fo</strong>re DNA, from the lining of the predators’ gut<br />

that could be collected <strong>fo</strong>r genetic analysis (i.e., “no DNA”), and 4) scats that were<br />

clearly from a predator other than a coyote, especially those containing insects and which<br />

appeared to be from a reptile. These <strong>fo</strong>ur <strong>type</strong>s of scat were removed from the transect<br />

but no further in<strong>fo</strong>rmation was recorded. It was assumed that the incomplete scats would<br />

not provide a fully representative sample of the coyote diet. Omitting the “old” scats from<br />

those sampled <strong>fo</strong>r genetic analysis improved the chances that the minimum counts of<br />

individuals obtained <strong>fo</strong>r each scat transect (see genetics section below) were based on<br />

samples that had accumulated over a known period of time (i.e., the time since the last<br />

survey) and were thus more comparable.<br />

In 2009, two surveys of all scat transects were carried out in each of three<br />

seasons: spring, summer and fall. This was done in order to collect further data on inter-<br />

habitat differences in the base of the coyote <strong>fo</strong>od chain and address the issue of seasonal<br />

variation in resource use by both coyotes and their prey. Previous studies have shown that<br />

coyote diet, in terms of frequency of occurrence of different prey items (e.g., rodents,<br />

arthropods), varies seasonally (e.g., spring to fall, Hernández et al. 2002) and that there is<br />

seasonal variation (pre monsoon to monsoon) in the use of C3 versus C4 plants by<br />

potential coyote prey species (Warne et al. 2010). This diet shift from C3 to C4 plants<br />

should be reflected in the stable ca<strong>rb</strong>on isotope signature of some of the components,


41<br />

especially vegetation and small mammal hair, that were taken from the coyote scat<br />

samples and run through a stable isotope analysis (see ca<strong>rb</strong>on stable isotopes section<br />

below).<br />

The monsoon, or rainy, season typically starts between late June and early August<br />

and ends in September or October. Scat samples were collected be<strong>fo</strong>re (April), during<br />

(July), and after (October) the peak of the monsoon season (Figure 3).<br />

Average precipitation (mm)<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Jan Feb Mar Apr May Jun Jul Aug Sept Oct Nov Dec<br />

Month<br />

Figure 3. Seasonal variation in rainfall at the study site. Average monthly rainfall at<br />

the <strong>Sevilleta</strong> NWR as recorded by a meteorological station located in the northeastern<br />

part of the NWR (data from Moore 2009). Error bars correspond to 95% confidence<br />

intervals. On average, August is the month with the most rainfall. The spring scat surveys<br />

corresponded to a low rainfall month (April), the summer surveys to a fairly high rainfall<br />

month (July), and the fall surveys to a month (October) after the peak of the monsoon<br />

(i.e., rainy) season.<br />

Scat surveys were carried out along the 20 road-based transects described previously as<br />

well as on two new transects located in the northwestern part of the refuge (Figure 2). In<br />

2008, the three transects on the western side were all located in shrubland habitat. The<br />

purpose of adding two new transects in 2009 was to allow <strong>fo</strong>r sampling of the coyote<br />

population in the grassland area on the western side of the refuge. This population was


42<br />

expected to contain different individuals from the population in the grassland on the<br />

eastern side since the two sides of the refuge are separated by a heavily travelled,<br />

interstate highway (I-25). As a result, sampling in the grassland on the western side was<br />

expected to increase sample size. Furthermore, given the floristic differences between the<br />

two sides of the refuge, it seemed important to be able to compare the ecology, especially<br />

the base of the <strong>fo</strong>od chain, of coyotes in the shrubland on the western side to that of<br />

coyotes in the grassland on the western side. End points <strong>fo</strong>r 3 of the transects surveyed in<br />

2008 were moved slightly in spring 2009 either because the wooden stakes that marked<br />

the end points were not <strong>fo</strong>und or because the transects were <strong>fo</strong>und to be slightly less than<br />

a mile in length.<br />

Old scats were cleared from all transects at the beginning of each season and<br />

surveys of a given transect were separated by 7 to 19 days. The results of the genetic<br />

analysis of samples collected in 2008 were used to change the criteria <strong>fo</strong>r identifying a<br />

sample as being from a coyote. In particular, the range of values <strong>fo</strong>r average maximum<br />

diameter and length used to identify coyote samples in 2009 were: 0.90 to 3.20 cm and<br />

2.5 to 19.0 cm, respectively. Another change that was made from the procedures used in<br />

2008 was that three rather than <strong>fo</strong>ur categories were used <strong>fo</strong>r scats that were tallied and<br />

not measured or sampled. In particular, scats were no longer categorized as “old”, only as<br />

“incomplete”, “no DNA”, or being from a predator other than a coyote. This meant that<br />

any scats that were old and worn down enough that they did not have any fecal material<br />

that contained epithelial cells from the predator that deposited them were now included in<br />

the “no DNA” category. The purpose of this change was to reduce the subjectivity of<br />

these categories while still omitting samples that would provide incomplete in<strong>fo</strong>rmation


43<br />

on coyote diet (“incomplete”) or did not appear to have material that could be sampled<br />

<strong>fo</strong>r genetic analysis (“no DNA”). It was also assumed that samples that were incomplete<br />

or had no DNA were likely to have been missed when the scat transects were cleared at<br />

the beginning of the season. This assumption was based on the fact that these samples<br />

were often smaller and had either a less distinct shape or lighter color which made them<br />

harder to detect and also made it more likely that these samples were old. As a result, the<br />

omission of these samples from the genetic analysis improved the likelihood that the<br />

minimum counts of individuals obtained <strong>fo</strong>r each transect (see genetics section below)<br />

were based on samples that had accumulated since the last scat survey and were thus<br />

comparable.<br />

Vegetation surveys<br />

Vegetation surveys provided a quantitative assessment of the habitat surrounding<br />

each scat transect and there<strong>fo</strong>re used by the coyotes being surveyed. This assessment<br />

augmented the qualitative classification of the habitat near the scat transects as grassland<br />

or shrubland that was used when the transect end points were first marked. In July and<br />

August 2008, vegetation was characterized in 40 circular vegetation plots, two per scat<br />

transect (Figure 4). Each plot was 30 m in diameter and was located 30 to 100 m from a<br />

scat transect. The distance from the beginning of the scat transect, the side of the road,<br />

and the distance from the road at which the plot was located were all randomized.<br />

Vegetation variables relevant to woody plant-encroached landscapes and coyote<br />

<strong>fo</strong>raging patterns were assessed in each plot. Specifically, in<strong>fo</strong>rmation was collected on:<br />

1) percent <strong>live</strong> woody plant <strong>cover</strong>, 2) average size of <strong>live</strong> woody plants ≥ 0.5 m tall, 3)<br />

average inter-plant distance <strong>fo</strong>r <strong>live</strong> woody plants that are at least 0.5 m tall, 4) dominant


44<br />

grass genera. The first three variables were measured <strong>fo</strong>r <strong>live</strong> woody plants (i.e., plants<br />

with green leaves on some part) that intersected two 30 m line intercept transects. These<br />

transects were perpendicular to one another and were oriented north-south and east-west<br />

such that they intersected at the center of the plot. The <strong>fo</strong>urth variable was determined<br />

using 5 randomly placed 1 m 2 quadrats per plot. These quadrats were located at a random<br />

distance and angle from the center of a given plot.<br />

Figure 4. Vegetation plot map. Map of vegetation plot locations <strong>fo</strong>r 2008 and 2009.<br />

Vegetation plots surveyed in both 2008 and 2009 are shown with asterisks and plots<br />

surveyed only in 2008 are shown with crosses. Two plots were only surveyed in 2009 and<br />

are indicated with a black circle. For 2008, 20 plots were located in grassland (gray)<br />

habitat and the other 20 plots were in shrubland (black) habitat. There were 12 plots in<br />

grassland areas (gray) and 10 plots in shrubland areas (black) in 2009.<br />

Each <strong>live</strong> woody plant that crossed the line intercept transects was identified to<br />

species, a height category (< 0.5 m or ≥ 0.5 m) was noted, and the beginning and ending<br />

points at which the plant intercepted the tape were recorded to the nearest 0.1 m. These<br />

interception points were later used to calculate values <strong>fo</strong>r total percent <strong>live</strong> woody plant<br />

<strong>cover</strong>. In this calculation, a correction was made <strong>fo</strong>r overlap between woody plants of


45<br />

different species or different size classes. This correction ensured that each segment of<br />

the line intercept transects was only counted once in the calculation of percent <strong>live</strong> woody<br />

plant <strong>cover</strong>.<br />

Woody plant size was determined by measuring the height, the longest axis and<br />

the axis perpendicular to the longest axis <strong>fo</strong>r each <strong>live</strong> woody plant that crossed one of<br />

the line intercept transects and was at least 0.5 m tall. All of these measurements were<br />

made to the nearest 0.1 m. When no woody plant 0.5 m or taller intersected the line<br />

intercept transects, these plant size measurements were recorded <strong>fo</strong>r any plants between<br />

which nearest neighbor distances (see below) were measured. When there was only one<br />

woody plant greater than 0.5 m tall within a plot (i.e., percent woody plant <strong>cover</strong> is zero<br />

and no interplant distances were measured), plant size measurements were taken <strong>fo</strong>r this<br />

plant.<br />

For the measurement of average interplant distance, up to <strong>fo</strong>ur <strong>live</strong> woody plants<br />

with a height of at least 0.5 m were selected in each plot. In most plots these selected<br />

plants intersected one of the line intercept transects and were the closest to the center of<br />

the plot. When no <strong>live</strong> woody plants with a height of at least 0.5 m crossed either line<br />

intercept transect, then the woody plant that was in the plot and was closest to the first 1<br />

m 2 quadrat that was surveyed was used instead. The interplant distances were measured<br />

from the centers of these selected plants to the centers of the 5 closest <strong>live</strong> woody plants<br />

that were also at least 0.5 m tall and were within the plot. Fewer than 5 distances were<br />

measured when there were fewer than 6 <strong>live</strong> woody plants (height ≥ 0.5 m) <strong>fo</strong>und within<br />

the plot boundaries.


46<br />

For each 1 m 2 quadrat, the grass genus that had the highest percent <strong>cover</strong> was<br />

recorded as the “dominant” grass. When two genera had similar values <strong>fo</strong>r percent <strong>cover</strong>,<br />

the quadrat was considered to have “co-dominant” grasses and both genera were<br />

recorded. Samples of the dominant grass genera were collected from within the quadrats.<br />

Samples of all woody plant species (excluding cactus) that were <strong>fo</strong>und in the plot were<br />

also collected. These samples were then dried at 60 o C <strong>fo</strong>r 24 hours and stored in a plastic<br />

bag with desiccant in preparation <strong>fo</strong>r ca<strong>rb</strong>on stable isotope analysis (see ca<strong>rb</strong>on stable<br />

isotopes section below).<br />

To ensure that the two new scat transects (Figure 2) had habitat data comparable<br />

to that collected in 2008, <strong>fo</strong>ur vegetation plots, two per transect, were surveyed in<br />

summer, 2009. These <strong>fo</strong>ur plots were located 30 to 100 m from the road at randomly<br />

selected points along the new scat transects. The surveys of these plots included<br />

measurements of the <strong>fo</strong>llowing: percent <strong>live</strong> woody plant <strong>cover</strong> <strong>fo</strong>r plants with a height of<br />

at least 0.5 m; woody plant size (height, longest axis, perpendicular axis); and interplant<br />

distances. Dominant grass species were determined using five randomly placed 1 m 2<br />

quadrats per plot <strong>fo</strong>r the two plots that were also surveyed in spring and fall 2009 (see<br />

description below).<br />

During the spring, summer and fall of 2009, a subset of the habitat variables<br />

measured in 2008 was resurveyed, with some methodological changes, to assess seasonal<br />

variation in the percent <strong>live</strong> <strong>cover</strong> of, and there<strong>fo</strong>re availability of resources associated<br />

with, C3 (<strong>fo</strong><strong>rb</strong>s, woody plants) versus C4 (grasses) plants. Of particular interest was any<br />

shift in availability of <strong>live</strong> grasses as a <strong>fo</strong>od resource from spring to fall. Any such shift


47<br />

should be detected in the ca<strong>rb</strong>on isotope signatures of coyote prey items (see ca<strong>rb</strong>on<br />

stable isotopes section below).<br />

In 2009, vegetation measurements were per<strong>fo</strong>rmed in 22 plots (Figure 4), one per<br />

scat transect (Figure 2), in each season <strong>fo</strong>r a total of three surveys per plot. Specifically,<br />

one of the two vegetation plots surveyed along each of the scat transects during summer<br />

2008 was randomly selected and its center point was marked with a wooden stake to<br />

ensure that the line intercept transects intersected at the same point <strong>fo</strong>r each 2009 survey<br />

of the plot. Percent <strong>live</strong> woody vegetation, grass and <strong>fo</strong><strong>rb</strong> <strong>cover</strong> were measured, to the<br />

nearest 0.1m, along 2 x 30 m line intercept transects and height was recorded <strong>fo</strong>r each<br />

woody plant that intersected the transects and was at least 0.5 m tall. It is important to<br />

note that data on percent <strong>live</strong> grass and <strong>fo</strong><strong>rb</strong> <strong>cover</strong> were collected in 2009, while only<br />

percent <strong>live</strong> woody plant <strong>cover</strong> was assessed along the line intercept transects in 2008.<br />

Also, in 2009, an ef<strong>fo</strong>rt was made to measure percent <strong>cover</strong> <strong>fo</strong>r only the <strong>live</strong> portions of<br />

plants that intersected transects. More specifically, if a stalk or seed head or leaf were<br />

green, then it was considered to be "a<strong>live</strong>.” If only part of a given plant was green, then<br />

only that part was included in the calculation of percent <strong>live</strong> vegetation <strong>cover</strong>.<br />

Small samples of the dominant woody plant species (excluding cactus species,<br />

except in the fall surveys) and one of the dominant grass genera <strong>fo</strong>und within the plot<br />

were collected. Dominant species and genera were determined using the vegetation data<br />

collected in 2008 and the <strong>fo</strong>llowing criteria: the woody plant species that was at least 0.5<br />

m in height and had the highest value <strong>fo</strong>r percent <strong>cover</strong> within the plot; the grass genus<br />

with the highest percentage <strong>cover</strong> in at least one of five 1 m 2 quadrats that were surveyed<br />

in each plot. If a different species of woody plant was identified as being dominant, using


48<br />

the criteria outlined above, during one of the 2009 surveys, then an attempt was made to<br />

sample both the species that was dominant in 2008 and the species dominant in 2009. A<br />

representative sample of <strong>fo</strong><strong>rb</strong>s was also collected if any <strong>live</strong> <strong>fo</strong><strong>rb</strong>s intersected one or both<br />

of the line intercept transects <strong>fo</strong>r a given plot. All of these vegetation samples were dried<br />

<strong>fo</strong>r 24 hrs at 60 o C and stored in labeled plastic bags with desiccant <strong>fo</strong>r future ca<strong>rb</strong>on<br />

stable isotope analysis (see ca<strong>rb</strong>on stable isotopes section below).<br />

Laboratory work<br />

Laboratory work consisted primarily of the <strong>fo</strong>llowing: 1) genetic analysis of<br />

coyote scat collected at the <strong>Sevilleta</strong> NWR, 2) ca<strong>rb</strong>on stable isotope analysis of<br />

components separated from these scat samples.<br />

Genetics<br />

Genetic analyses were run at the Laboratory <strong>fo</strong>r Conservation and Ecological<br />

Genetics (LCEG) at the University of Idaho to confirm the field-based species<br />

identification of the scat samples collected at the <strong>Sevilleta</strong> NWR and determine the<br />

minimum number of individuals represented by all coyote scat samples. Segments of<br />

mitochondrial (mtDNA) and nuclear DNA in scat subsamples that were collected and<br />

preserved in DETs buffer in the field were extracted using the QIAamp DNA Stool Mini<br />

Kit from Qiagen Inc. Extractions were per<strong>fo</strong>rmed in a laboratory space that was separate<br />

from the area where DNA was amplified and a negative control was processed with each<br />

set of extracted samples. These procedures helped to reduce contamination of the low<br />

quality DNA from the scat samples with amplified DNA, and to identify cross sample<br />

contamination during the extraction process, respectively (Taberlet et al. 1999, Onorato et<br />

al. 2006).


49<br />

Once extracted, the mtDNA was amplified using polymerase chain reaction<br />

(PCR) techniques and primers designed <strong>fo</strong>r a species identification test. This test was<br />

per<strong>fo</strong>rmed primarily to confirm that the collected scat samples were deposited by coyotes<br />

and not other canids. In particular, two primers (Table 1, Murphy et al. 2000) that<br />

amplify a segment of the mtDNA control region (Onorato et al. 2006) were used. The<br />

length of this segment of the control region varies among some mammalian carnivore<br />

species. Specifically, the LCEG has identified size ranges <strong>fo</strong>r segments amplified from<br />

coyotes and red wolves, other canids (including gray wolves and domestic dogs), black<br />

bears, brown bears and lynx (Table 2).<br />

Table 1. Species ID primers. Names and sequences <strong>fo</strong>r primers used <strong>fo</strong>r species<br />

identification (Murphy et al. 2000). Please note the <strong>fo</strong>llowing two revisions from the<br />

in<strong>fo</strong>rmation presented by Murphy et al. (2000): 1) primer name is SIDL rather than IDL<br />

and 2) the H16145 sequence has been modified from 5’- AGG AAG AAG CAA CAG<br />

TCT C-3’. 6-FAM is a dye and A = adenine, C = cytosine, G = guanine, T = thymine.<br />

Primer name Primer sequence<br />

SIDL 5'-/6-FAM/TCT ATT TAA ACT ATT CCC TGG-3'<br />

H16145 5'-GGG CAC GCC ATT AAT GCA CG-3'<br />

Table 2. Size ranges <strong>fo</strong>r mtDNA amplified from different carnivore species. Sizes<br />

listed in terms of number of base pairs (bp). All values from LCEG.<br />

Species Latin name Fragment size range (bp)<br />

Coyote/Red wolf Canis latrans/Canis rufus 114.9 – 120.1<br />

Lynx Lynx canadensis<br />

Canis lupus/Canis lupus<br />

121.6 – 122.4<br />

Gray wolf/Domestic dog familiaris 123 – 128.2<br />

Brown bear Ursus arctos 146.9 – 153.5<br />

Black bear Ursus americanus 158.1 – 164.5<br />

For the PCR, 1.8 µL of extracted DNA was mixed with 3.5µL of the 2X Qiagen<br />

Master Mix, 0.7µL of 5X Q solution, 0.14µL of each 10 µM primer, and 0.72µL of sterile<br />

water. The reagents from Qiagen contain the enzymes and nucleotides necessary to


50<br />

replicate the DNA, as well as compounds that block PCR-inhibiting chemicals that could<br />

be present in the DNA extract. A segment of the control region of the mtDNA in this<br />

mixture was amplified using the <strong>fo</strong>llowing set of steps programmed into a DNA Engine<br />

Tetrad 2 Peltier Thermal Cycler: 1) 95°C <strong>fo</strong>r 15 minutes, 2) 94°C <strong>fo</strong>r 30 seconds, 3) 44°C<br />

<strong>fo</strong>r 1.5 minutes, 4) 72°C <strong>fo</strong>r 1 minute, 5) 60°C <strong>fo</strong>r 30 minutes. The second through <strong>fo</strong>urth<br />

steps were per<strong>fo</strong>rmed 40 times. 1µL of amplified mtDNA was then mixed with 0.33 µL<br />

of GeneScan TM 500 LIZ size standard (Applied Biosystems) and 9.67 µL of <strong>fo</strong>rmamide.<br />

In order <strong>fo</strong>r the DNA to denature, the mixture was placed on a thermocycler at 95°C <strong>fo</strong>r<br />

2-3 minutes, put on ice <strong>fo</strong>r 2 minutes, and then the DNA fragments were separated via<br />

electrophoresis on an Applied Biosystems 3130xl capillary machine.<br />

All DNA fragment length and quantity data produced by the capillary machine<br />

were analyzed using GeneMapper 3.7 and samples were categorized as <strong>fo</strong>llows: coyote,<br />

maybe coyote, not coyote, unknown. Coyote samples had a high quantity of DNA<br />

fragments in the <strong>fo</strong>llowing size range: 114.89 to 120.13bp. Samples that were “maybe<br />

coyote” had a smaller quantity of DNA fragments in the coyote size range or had DNA<br />

fragments that were at the lower end, or just outside, of this range. Samples that were<br />

“not coyote” had DNA fragments in size ranges used to identify other species (Table 2)<br />

and “unknown” samples had too little DNA in the coyote size range or were samples <strong>fo</strong>r<br />

which no DNA was amplified in the PCR.<br />

To determine the number of different coyotes that were sampled, nuclear DNA<br />

from all scat samples identified as “coyote” or “maybe coyote” was amplified using PCR<br />

techniques and canid-specific primers <strong>fo</strong>r 8 microsatellite loci (Ostrander et al. 1993,<br />

Ostrander et al. 1995, Francisco et al. 1996, Table 3). For the PCR, 1.4 µL of extracted


51<br />

DNA was mixed with 3.5µL of 2X Qiagen Master Mix, 0.7µL of 5X Q solution, 0.93µL<br />

of sterile water, and volumes of each primer needed to attain specific primer<br />

concentrations (Table 3). DNA at the 8 microsatellite loci of interest were amplified<br />

using the <strong>fo</strong>llowing set of steps programmed into a DNA Engine Tetrad 2 Peltier Thermal<br />

Cycler: 1) 95°C <strong>fo</strong>r 15 minutes, 2) 94°C <strong>fo</strong>r 30 seconds, 3) 63°C <strong>fo</strong>r 1.5 minutes, 4) 72°C<br />

<strong>fo</strong>r 1 minute, 5) 94°C <strong>fo</strong>r 30 seconds, 6) 55°C <strong>fo</strong>r 1.5 minutes, 7) 72°C <strong>fo</strong>r 1 minute, 8)<br />

60°C <strong>fo</strong>r 30 minutes. The second through <strong>fo</strong>urth steps were per<strong>fo</strong>rmed 16 times, with the<br />

temperature in step three (63°C) declining by 0.5°C in each cycle. Steps 5 through 7 were<br />

per<strong>fo</strong>rmed 31 times.<br />

Table 3. Microsatellite primer sequences. Sequences of and final PCR concentrations<br />

<strong>fo</strong>r the 8 primers used to amplify microsatellite loci as part of the process of identifying<br />

individual coyotes (Ostrander et al. 1993, Ostrander et al. 1995, Francisco et al. 1996). F<br />

= <strong>fo</strong>rward primer, R = reverse primer, T = thymine, C = cytosine, A = adenine, G =<br />

guanine.<br />

Primer<br />

Final<br />

concentration<br />

name Primer sequence<br />

F: 5'- TTGATTTCCCCTGTAGCTTA -3'<br />

<strong>fo</strong>r PCR (µM)<br />

CXX119 R: 5'- GATGTAAAGAATGAGAGAGG -3'<br />

F: 5’-ATCCAGGTCTGGAATACCCC-3’<br />

0.26<br />

CXX173 R: 5’-TCCTTTGAATTAGCACTTGGC-3’<br />

F: 5’-TTAGTTAACCCAGCTCCCCCA-3’<br />

0.05<br />

CXX250 R: 5’-TCACCCTGTTAGCTGCTCAA-3’<br />

F: 5’-ACGTGTTGATGTACATTCCTGC-3’<br />

0.06<br />

CXX377 R: 5’-CCACCCAGTCACACAATCAG-3’<br />

F: 5’-TCCTCCTCTTCTTTCCATTGG-3’<br />

0.04<br />

FH2001 R: 5’-TGAACAGAGTTAAGGATAGACACG-3’<br />

F: 5’-AAATGGAACAGTTGAGCATGC-3’<br />

0.06<br />

CXX2010 R: 5’-CCCCTTACAGCTTCATTTTCC-3’<br />

F: 5’-GCCTTATTCATTGCAGTTAGGG-3’<br />

0.04<br />

FH2054 R: 5’-ATGCTGAGTTTTGAACTTTCCC-3’<br />

F: 5’-CCCTCTGCCTACATCTCTGC-3’<br />

0.05<br />

FH2088 R: 5’-TAGGGCATGCATATAACCAGC-3’ 0.06


52<br />

The amplified nuclear DNA was then prepared <strong>fo</strong>r, and separated via<br />

electrophoresis on, a capillary machine using the same procedure described above <strong>fo</strong>r the<br />

mtDNA. GeneMapper 3.7 was used to analyze the resulting fragment size data and<br />

determine which alleles were present at each locus in each sample. Per locus<br />

amplification success rates were calculated based on this initial microsatellite screening<br />

run <strong>fo</strong>r all samples that were identified as “coyote” or “maybe coyote” in the species<br />

identification test. A locus was considered to be “successful” if any alleles were<br />

amplified in the PCR and identified using GeneMapper 3.7.<br />

The PCR and electrophoresis steps were repeated two more times <strong>fo</strong>r samples <strong>fo</strong>r<br />

which DNA was successfully amplified at 5 or more microsatellite loci in the initial<br />

nuclear DNA PCR. The 3 replicate PCRs <strong>fo</strong>r a given sample were compared in an<br />

attempt to obtain a consensus geno<strong>type</strong> <strong>fo</strong>r each locus. For a homozygous locus, one<br />

allele, and only that allele, had to be seen in all three replicate PCRs while, <strong>fo</strong>r a<br />

heterozygous locus, each allele had to be seen in two PCRs. Samples <strong>fo</strong>r which a<br />

consensus geno<strong>type</strong> was obtained at only 5 loci were screened two more times. Once<br />

these additional PCRs were assessed, all samples with consensus geno<strong>type</strong>s at 6 or more<br />

loci were analyzed in the programs Gimlet 1.3.3 (Valière 2002) and GenAlEx 6 (Peakall<br />

and Smouse 2006). These programs matched the consensus geno<strong>type</strong>s and determined the<br />

minimum number of individuals (i.e., different coyotes) from which samples had been<br />

collected. These programs also identified samples that had consensus geno<strong>type</strong>s which<br />

differed from those of one or more other samples by only 1 or 2 loci. If the geno<strong>type</strong>s of<br />

these samples were incomplete or were otherwise uncertain, then they were run through<br />

the microsatellite analysis two more times and the results of these reruns were compared


53<br />

to those of previous PCRs. The updated consensus geno<strong>type</strong>s were then reanalyzed in<br />

Gimlet 1.3.3 and GenAlEx 6. If the updated geno<strong>type</strong>s still differed from those of one or<br />

more samples at one or two loci and if the difference(s) could be the result of a single<br />

instance of allelic dropout or the presence of a single false allele (see definition below),<br />

then the updated geno<strong>type</strong>s were excluded from further analysis. Finally, the reliability of<br />

unique consensus geno<strong>type</strong>s that were observed only once was tested using the program<br />

RELIOTYPE (Miller et al. 2002) and a reliability criteria of 95%. If the reliability of a<br />

unique geno<strong>type</strong> was confirmed in the initial RELIOTYPE test, or <strong>fo</strong>llowing 2 reruns of<br />

the microsatellite analysis, it was retained in the dataset. Samples that were still identified<br />

as unreliable after a total of 7 microsatellite screenings, or, based on the RELIOTYPE<br />

results, would need more than a total of 7 screenings to be considered reliable, were<br />

excluded from further analysis.<br />

Per locus error (false allele and allelic dropout) rates and per sample probability<br />

of identity values were calculated in order to assess the quality of the microsatellite data.<br />

Error rates and probability of identity values were based on data from all samples that<br />

had consensus geno<strong>type</strong>s at 6 or more loci and which were not excluded from further<br />

analysis as described above. More specifically, error rates were determined using the first<br />

two successful PCRs, or PCRs <strong>fo</strong>r which one or more alleles were identified, <strong>fo</strong>r each<br />

sample at each locus. For each sample, loci that did not have at least two successful PCRs<br />

were excluded from this calculation. For a given sample, false alleles were alleles that<br />

appeared in one of the first two successful microsatellite screenings <strong>fo</strong>r a particular locus<br />

but did not appear in the consensus geno<strong>type</strong> <strong>fo</strong>r that locus and were thus seen in only<br />

one of the three or more replicate PCRs <strong>fo</strong>r that locus. Allelic dropout was noted <strong>fo</strong>r a


54<br />

sample when one of the two alleles in the consensus geno<strong>type</strong> <strong>fo</strong>r a heterozygous locus in<br />

that sample failed to amplify in one or more of the first two successful PCRs <strong>fo</strong>r that<br />

locus (Waits and Paetkau 2005). Per sample probability of identity (PID) <strong>fo</strong>r both<br />

unrelated individuals and siblings was calculated using unbiased per locus values <strong>fo</strong>r<br />

unrelated individuals and per locus values <strong>fo</strong>r siblings obtained via Gimlet 1.3.3 (Waits et<br />

al. 2001, Valière 2002). In particular, <strong>fo</strong>r each sample, the per locus PID values <strong>fo</strong>r either<br />

unrelated individuals or siblings <strong>fo</strong>r all loci <strong>fo</strong>r which consensus had been reached were<br />

multiplied together. PID values provide in<strong>fo</strong>rmation regarding the probability that either<br />

two unrelated individuals or two siblings in the population will have the same geno<strong>type</strong>.<br />

These values are affected by the number and allelic diversity of loci being used to<br />

differentiate individuals (Waits et al. 2001, Waits 2004).<br />

The consensus geno<strong>type</strong>s of unique individuals, as determined using Gimlet 1.3.3<br />

and GenAlEx 6 and confirmed with RELIOTYPE, were tested using the programs<br />

GENEPOP 4.0.10 (Raymond and Rousset 1995, Rousset 2008), Structure 2.3.1 (Pritchard<br />

et al. 2000), and BAPS 5 (Corander et al. 2008a, Corander et al. 2008b). GENEPOP<br />

4.0.10 was used to determine if the geno<strong>type</strong>s were in Hardy-Weinberg (HW)<br />

equilibrium and if they met the assumption of linkage equilibrium (LE). The null<br />

hypothesis <strong>fo</strong>r the LE test is that the geno<strong>type</strong> at one locus is independent of the<br />

geno<strong>type</strong>s at the 7 other loci. Bonferroni corrected p-values were used to assess the<br />

significance of the results of both the HW equilibrium (significant p-value = 0.006 or<br />

0.05 divided by the number of loci used) and LE (significant p-value = 0.002 or 0.05<br />

divided by the number of locus pairs) tests (based on Sacks et al. 2004). Significant<br />

deviations from HW equilibrium or rejection of the null hypothesis <strong>fo</strong>r the LE test may


55<br />

be an indication that errors were made in the process of determining consensus geno<strong>type</strong>s<br />

(Sacks et al. 2004).<br />

Structure 2.3.1 and BAPS 5 were used to determine whether the sampled<br />

individuals appeared to come from one or multiple populations or genetic groups. The<br />

presence of multiple genetic groups could indicate the presence of either more than one<br />

species, and thus errors in the species identification test, or a barrier to gene flow within<br />

the coyote population. The consensus geno<strong>type</strong>s were tested to see if there were between<br />

1 and 20 genetic groups present in the study area. For Structure 2.3.1, the burnin period<br />

was set to 100,000, the number of Markov Chain Monte Carlo repetitions used was<br />

1,000,000 (based on Pritchard et al. 2000, Wilson et al. 2009) and the program was run<br />

20 times <strong>fo</strong>r each value of K (i.e., number of genetic groups). Individuals were assigned<br />

to different genetic groups based on membership coefficients (i.e., Q-values; Pritchard et<br />

al. 2009) produced by the program. Genetic group membership was determined using the<br />

<strong>fo</strong>llowing two approaches: 1) individuals were assigned to the genetic group <strong>fo</strong>r which<br />

they obtained the largest Q-value; 2) individuals were assigned to a genetic group only if<br />

they obtained a Q-value larger than 0.8 (based on Sacks et al. 2004). For BAPS 5, the<br />

fixed K clustering function was used in order to run the program 20 times <strong>fo</strong>r each K<br />

value of interest (2 to 20) <strong>fo</strong>r the spatial clustering of individuals analysis. The program<br />

was also run 380 times (19 x 20) with the fixed K function disabled in order to obtain<br />

in<strong>fo</strong>rmation on the optimal number of genetic groups and optimal genetic group<br />

assignment <strong>fo</strong>r each individual. Each of the coordinates used <strong>fo</strong>r the spatial analysis in<br />

BAPS 5 represents the center point of all locations where a single individual was<br />

sampled.


56<br />

The “kinship” function in Gimlet 1.3.3, as well as the “relationship” function in<br />

ML-Relate (Kalinowski et al. 2006), were used to identify individuals that were related to<br />

one another, in particular individuals that had two parents among the animals that were<br />

sampled (Gimlet) or were in a parent-offspring relationship with another individual<br />

among those sampled (ML-Relate). The HW equilibrium and LE tests, as well as the<br />

analyses of genetic group number, were rerun once two out of every three individuals that<br />

were identified as being in a parent-offspring relationship by Gimlet 1.3.3. were removed.<br />

The presence of related individuals can influence the results of the HW and LE tests, as<br />

well as the analyses of genetic group number. The results of the Gimlet 1.3.3 kinship and<br />

ML-Relate relationship analyses were also used to determine which individuals were in<br />

the same “family” group. In particular, a family group was defined by starting with either<br />

2 (ML-Relate) or 3 (Gimlet) individuals identified as being in a parent-offspring<br />

relationship and then adding all relatives (i.e., mates, offspring, and parents) of 1) these<br />

2-3 individuals and 2) all mates, offspring, and parents of these 2-3 individuals.<br />

Once the reliability of the consensus geno<strong>type</strong>s had been evaluated, they were<br />

used to determine the minimum number of coyotes that were detected on each road-based<br />

transect in each of 9 scat surveys. These counts were then averaged 1) across all 9<br />

surveys and 2) across each pair of surveys that was conducted in a particular season. Both<br />

of these sets of average minimum counts were then averaged across all transects within<br />

the same habitat <strong>type</strong> (grassland versus shrubland) to obtain habitat-specific counts.<br />

These habitat-specific counts were analyzed to determine if there were any significant<br />

difference between habitats in the number of coyotes sampled 1) over the entire study


57<br />

period (Student’s t-test; PROC TTEST SAS 9.1), and 2) among seasons (two-way<br />

ANOVA and Duncan’s test; PROC GLM SAS 9.1).<br />

Ca<strong>rb</strong>on stable isotopes<br />

The purpose of per<strong>fo</strong>rming ca<strong>rb</strong>on stable isotope analyses was to assess the base<br />

of the <strong>fo</strong>od chain <strong>fo</strong>r individual coyotes sampled in grassland and shrubland habitats in<br />

different seasons. More specifically, the ca<strong>rb</strong>on isotope signatures of components of<br />

coyote scat samples provide in<strong>fo</strong>rmation on the percentage of the coyote diet that comes<br />

directly (seeds and fruits) or indirectly (rodents and other prey) from grasses (C4 plants)<br />

versus woody plants (C3 plants). The first scat sample collected <strong>fo</strong>r each individual<br />

coyote in each of <strong>fo</strong>ur field seasons (summer 2008, or spring, summer and fall 2009) was<br />

prepared <strong>fo</strong>r ca<strong>rb</strong>on isotope analysis (see description below). This was done in order to<br />

maximize sample size, since samples from the same individual are not independent of<br />

one another, while minimizing the cost of the ca<strong>rb</strong>on stable isotope analysis.<br />

To prepare samples <strong>fo</strong>r ca<strong>rb</strong>on stable isotope analysis, scat samples that were<br />

dried at 70°C <strong>fo</strong>r 24 hours were first flattened and spread out using a mortar and pestle<br />

and <strong>fo</strong>rceps. For individuals that were just sampled in one season, small quantities of<br />

bone were then separated from the rest of the scat components and cleaned with ethanol.<br />

For individuals that were sampled in more than one season in 2009, small amounts of<br />

bone, hair and vegetation (woody plant or <strong>fo</strong><strong>rb</strong> berries or seeds) were separated from the<br />

rest of the scat sample and cleaned with ethanol. The bones and berries or seeds were<br />

then crushed into a powder. The bones were expected to pick up long-term shifts (i.e.,<br />

those related to habitat changes associated with woody plant encroachment; Ambrose and<br />

DeNiro 1986) in the base of the coyote <strong>fo</strong>od chain while hair and vegetation were


58<br />

expected to pick up short-term (i.e., seasonal) variation (Tieszen et al. 1983). Samples of<br />

woody plants, <strong>fo</strong><strong>rb</strong>s and grasses that had been collected in the field and then dried at 60 o C<br />

<strong>fo</strong>r 24 hours were ground using a Wiley Mill until they passed through a 20 mesh filter.<br />

These ground samples included grasses and woody plants <strong>fo</strong>und to be dominant in one or<br />

more of the vegetation plots and <strong>fo</strong><strong>rb</strong>s that were collected in plots located in different<br />

parts of the refuge (northeast, southeast, northwest, southwest). They also included two<br />

woody plants that produce seeds or fruits that coyotes eat (V. Seamster, unpublished<br />

data).<br />

Scat samples obtained from a group of captive Mexican wolves (Canis lupus<br />

baileyi) that are housed at the <strong>Sevilleta</strong> NWR, as well as small samples of the meat and<br />

kibble that these wolves are fed, were also prepared <strong>fo</strong>r ca<strong>rb</strong>on stable isotope analysis.<br />

The purpose of analyzing these samples was to obtain a correction factor that could be<br />

used to correct ca<strong>rb</strong>on isotope signatures of canid scat components (i.e., bone or hair) <strong>fo</strong>r<br />

diet to scat fractionation (see below). This correction factor was calculated by subtracting<br />

the average ca<strong>rb</strong>on stable isotope signature of the <strong>fo</strong>od samples from the signature of<br />

each of the scats and then averaging the differences. Scats and <strong>fo</strong>od samples were dried at<br />

70°C <strong>fo</strong>r 24 hours and ground to a powder using a mortar and pestle.<br />

Appropriate masses (Table 4) of the cleaned scat components (i.e., bone or<br />

vegetion powder and hair), the milled vegetation samples, and the crushed wolf scat and<br />

<strong>fo</strong>od samples were then placed in small tin cups in preparation <strong>fo</strong>r analysis on a coupled<br />

elemental analyzer (Costech 4010) and isotope ratio mass spectrometer (DeltaPlux XP) at<br />

the Stable Isotope Laboratory at the University of New Hampshire (UNH).


59<br />

Table 4. Masses <strong>fo</strong>r stable isotope analysis. Masses used <strong>fo</strong>r weighing out samples of<br />

different scat components and standards <strong>fo</strong>r stable isotope analysis.<br />

Component Mass (mg)<br />

bolete (known unknown) 1.8-2.2<br />

bone 2.8-3.2<br />

citrus (known unknown) 3.8-4.2<br />

hair 0.8-1.2<br />

NIST 1515/1575a (standards) 4.0-4.5<br />

tuna (standard) 0.8-1.2<br />

vegetation 3.8-4.2<br />

wolf scat/<strong>fo</strong>od 3.8-4.2<br />

4-5 out of every 58-65 samples was replicated. In addition, 3 standards (NIST 1515 and<br />

1575a, tuna) and one known unknown (bolete; Table 4) were run <strong>fo</strong>r every 10-12<br />

samples. 20 tins containing a second known unknown (citrus; Table 4) were mixed in<br />

with the first 277 samples run.<br />

The ca<strong>rb</strong>on isotope signatures (i.e., δ 13 C values) derived from the mass<br />

spectrometer were corrected <strong>fo</strong>r shifts both over time and in response to variation in<br />

sample weight (A. Ouimette, UNH, 2010, personal communication). The signatures <strong>fo</strong>r<br />

scat components, specifically bone and hair, were also corrected <strong>fo</strong>r diet to tissue and diet<br />

to scat fractionation (Table 5) and used to calculate the percentage of coyote diet that<br />

came indirectly from grasses (C4) versus woody plants (C3) in grassland versus shrubland<br />

areas (Equation 1, based on Faure and Mensing 2005).<br />

Where<br />

δ 13 Cscat = δ 13 CC 3 (fC 3 ) + δ 13 CC 4 (1-fC 3 ) (1)<br />

δ 13 Cscat = the ca<strong>rb</strong>on isotope signature of a coyote scat component (i.e., bone or hair);<br />

δ 13 CC 3 = the average ca<strong>rb</strong>on isotope signature of C3 woody plants; fC 3 = the fraction of


coyote diet that comes indirectly from C3 woody plants; δ 13 CC 4 = the average ca<strong>rb</strong>on<br />

isotope signature of C4 grasses.<br />

The average δ 13 C values calculated from the woody plant and grass samples that<br />

were collected in the field represented the “end members” (i.e., δ 13 CC 3 , δ 13 CC 4 ) <strong>fo</strong>r this<br />

calculation such that the ca<strong>rb</strong>on isotope signatures of the coyote scat components were<br />

expected to fall between these values and represent a mixture of grass and woody plant<br />

<strong>fo</strong>od resources (Faure and Mensing 2005; Equation 1).<br />

Table 5. Fractionation correction factors. Values used to correct <strong>fo</strong>r diet to tissue and<br />

diet to scat fractionation. Please note that these values were subtracted from the δ 13 C<br />

values <strong>fo</strong>r the appropriate scat components and that no correction <strong>fo</strong>r fractionation was<br />

applied to any of the standards or vegetation samples.<br />

Component<br />

Correction<br />

factor ( o /oo) Correction <strong>type</strong> Source<br />

bone 1* diet to tissue DeNiro and Epstein 1978<br />

hair 1.0 diet to tissue Tieszen et al. 1983<br />

bone/hair 0 ± 2.0** diet to scat This study<br />

* This value was chosen over species-specific correction factors <strong>fo</strong>r bone collagen<br />

because whole bone was used in this study and collagen extraction was not per<strong>fo</strong>rmed.<br />

As a result, it is not known whether, or how much, collagen is present in the bone<br />

samples used.<br />

** Value presented as average ± 1 standard deviation. The average value was used as the<br />

correction factor.<br />

GIS analysis<br />

GIS techniques were used to determine whether the relationship between coyote<br />

feeding ecology and the local habitat changed as the spatial scale at which the habitat was<br />

characterized increased (see Chapter 5). The results of the noninvasive genetic sampling<br />

techniques described above, as well as previously published in<strong>fo</strong>rmation on coyote home<br />

range size (Windberg et al. 1997), was used in determining the range of spatial scales<br />

considered.<br />

60


Assessing coyote home range size and patterns of habitat use<br />

The location in<strong>fo</strong>rmation <strong>fo</strong>r scat samples collected from unique individuals at the<br />

<strong>Sevilleta</strong> NWR was used to determine the average size of coyote home ranges. In<br />

particular, the location in<strong>fo</strong>rmation was analyzed using the freeware program ABODE v.<br />

2 (Laver 2005) and ArcGIS 9.2 (ESRI). A minimum convex polygon was generated to<br />

represent the movements of each individual and the size of each polygon was determined.<br />

The average size of these polygons was taken to represent the average home range size<br />

<strong>fo</strong>r a coyote at the <strong>Sevilleta</strong> NWR. This value was compared to a previously published<br />

value <strong>fo</strong>r coyote home range size which was based on radiotelemetry data (Windberg et<br />

al. 1997). Both the noninvasive and previously published datasets were collected in arid<br />

environments (average annual rainfall 230-232mm; Goslee et al. 2003, Hochstrasser and<br />

Peters 2004) and the two study sites (<strong>Sevilleta</strong> NWR and <strong>LTER</strong> and Jornada <strong>LTER</strong>) are<br />

within 200 miles of one another.<br />

In<strong>fo</strong>rmation on the habitat in which the scat samples were collected was used to<br />

determine whether individuals had moved between habitats or stayed in one habitat <strong>type</strong>.<br />

The home range sizes were naturally log trans<strong>fo</strong>rmed and an ANOVA and Duncan’s test<br />

(PROC ANOVA, SAS 9.1) were per<strong>fo</strong>rmed to determine whether average home range<br />

size differed among individuals that used one (grassland or shrubland) or both habitat<br />

<strong>type</strong>s. Finally, two chi square analyses were used to determine whether individuals that<br />

stayed within one habitat <strong>type</strong> used the same habitat as, or a different habitat than, either<br />

one or two individuals to which they were related. The chi square analysis was per<strong>fo</strong>rmed<br />

twice <strong>fo</strong>r pairs of related individuals and twice <strong>fo</strong>r trios. The first analysis of both pairs<br />

and trios of related individuals was based on relationship in<strong>fo</strong>rmation provided by Gimlet<br />

61


62<br />

1.3.3 and the second was based on relationships identified by ML-Relate. The probability<br />

that an individual or their relative was sampled in either the grassland or the shrubland<br />

habitat was calculated based on the number of pair members (<strong>fo</strong>r pairs) or pairs and third<br />

individuals (<strong>fo</strong>r trios) that were sampled in either habitat <strong>type</strong>.


References<br />

Ambrose, S.H. and M.J. DeNiro. 1986. The isotopic ecology of East African mammals.<br />

Oecologia. 69: 395-406.<br />

Corander, J., P. Marttinen, J. Sirén, and J. Tang. 2008a. Enhanced Bayesian modelling in<br />

BAPS software <strong>fo</strong>r learning genetic structures of populations. BMC Bioin<strong>fo</strong>rmatics.<br />

9:539.doi:10.1186/1471-2105-9-539.<br />

Corander, J., J. Sirén, and E. Arjas. 2008b. Bayesian Spatial Modelling of Genetic<br />

Population Structure. Computational Statistics. 23: 111-129.<br />

DeNiro, M.J. and S. Epstein. 1978. Influence of diet on the distribution of ca<strong>rb</strong>on<br />

isotopes in animals. Geochimica et Cosmochimica Acta. 42: 495-506.<br />

Faure, G. and T.M. Mensing. 2005. Isotopes: Principles and Applications. Hoboken,<br />

New Jersey: John Wiley & Sons. Pps. 693-752.<br />

Francisco, L.V., A.A. Langston, C.S. Mellersh, C.L. Neal, and E.A. Ostrander. 1996. A<br />

class of highly polymorphic tetranucleotide repeats <strong>fo</strong>r canine genetic mapping.<br />

Mammalian Genome. 7: 359-362.<br />

Frantzen, M.A.J., J.B. Silk, J.W.H. Ferguson, R.K. Wayne, and M.H. Kohn. 1998.<br />

Empirical evaluation of preservation methods <strong>fo</strong>r fecal DNA. Molecular Ecology. 7:<br />

1423–1428.<br />

Gill, R.A. and I.C. Burke. 1999. Ecosystem consequences of plant life <strong>fo</strong>rm changes at<br />

three sites in the semiarid United States. Oecologia. 121: 551-563.<br />

Green, J.S. and J.T. Flinders. 1981. Diameter and pH comparisons of coyote and red <strong>fo</strong>x<br />

scats. Journal of Wildlife Management. 45(3): 1981.<br />

Goslee, S.C., K.M. Havstad, D.P.C. Peters, A. Rango, and W.H. Schlesinger. 2003. Highresolution<br />

images reveal rate and pattern of shrub encroachment over six decades in New<br />

Mexico, U.S.A. Journal of Arid Environments. 54: 755-767.<br />

Halfpenny, J.C. 2000. Scat and tracks of the desert southwest. Guil<strong>fo</strong>rd, Connecticut:<br />

Morris Book Publishing, LLC. 144 p.<br />

Hernández, L., R.R. Parmenter, J.W. Dewitt, D.C. Light<strong>fo</strong>ot, and J.W. Laundré. 2002.<br />

Coyote diets in the Chihuahuan Desert, more evidence <strong>fo</strong>r optimal <strong>fo</strong>raging. Journal of<br />

Arid Environments. 51: 613-624.<br />

63


64<br />

Hochstrasser, T. and D.P.C. Peters. 2004. Subdominant species distribution in microsites<br />

around two life <strong>fo</strong>rms at a desert grassland-shrubland transition zone. Journal of<br />

Vegetation Science. 15: 615-622.<br />

Kalinowski, S.T., A.P. Wagner, and M.L. Taper. 2006. ML-Relate: a computer program<br />

<strong>fo</strong>r maximum likelihood estimation of relatedness and relationship. Molecular Ecology<br />

Notes 6:576-579.<br />

Laver, P.N. 2005. Abode: Kernel Home Range Estimation <strong>fo</strong>r ArcGIS, using VBA and<br />

ArcObjects. Department of Fisheries and Wildlife Sciences, Virginia Tech<br />

Miller, C.R., P. Joyce, and L.P. Waits. 2002. Assessing allelic dropout and geno<strong>type</strong><br />

reliability using maximum likelihood. Genetics. 160: 357-366.<br />

Moore, D. 2009. ClimDB Monthly Data. Albuquerque, NM: <strong>Sevilleta</strong> Long Term<br />

Ecological Research Site. <br />

(8 February 2010).<br />

Muldavin, E., G. Shore, K. Taugher, and B. Milne. 1998. A vegetation classification and<br />

map <strong>fo</strong>r the <strong>Sevilleta</strong> National Wildlife Refuge, New Mexico. New Mexico Natural<br />

Heritage Program and <strong>Sevilleta</strong> Long Term Ecological Research Program. Biology<br />

Department, University of New Mexico, Albuquerque, NM 87131.<br />

Murie, O.J. 1982. A field guide to animal tracks. New York: Houghton Mifflin Company.<br />

375 p.<br />

Murphy, M.A., L.P. Waits, and K.C. Kendall. 2000. Quantitative evaluation of fecal<br />

drying methods <strong>fo</strong>r brown bear DNA analysis. Wildlife Society Bulletin. 28: 951–957.<br />

Onorato, D., C. White, P. Zager, and L.P. Waits. 2006. Detection of predator presence at<br />

elk mortality sites using mtDNA analysis of hair and scat samples. Wildlife Society<br />

Bulletin. 34(3): 815-820.<br />

Ostrander, E.A., F.A. Mapa, M. Yee, and J. Rine. 1995. One hundred and one new simple<br />

sequence repeat-based markers <strong>fo</strong>r the canine genome. Mammalian Genome. 6: 192-195.<br />

Ostrander, E.A., G.F. Sprague, and J. Rine. 1993. Identification and characterization of<br />

dinucleotide repeat (CA)n markers <strong>fo</strong>r genetic mapping in dog. Genomics. 16: 207-213.<br />

Parmenter, B. 2004. Coyote Scat Survey. Albuquerque, NM: <strong>Sevilleta</strong> Long Term<br />

Ecological Research Site Database: SEV049.<br />

(8 May 2008).<br />

Peakall, R. and P.E. Smouse. 2006. GENALEX 6: genetic analysis in Excel. Population<br />

genetic software <strong>fo</strong>r teaching and research. Molecular Ecology Notes. 6: 288-295.


Pritchard, J. K., M. Stephens, and P. Donnelly. 2000. Inference of population structure<br />

using multilocus geno<strong>type</strong> data. Genetics. 155: 945-959.<br />

Pritchard, J.K., X. Wen, and D. Falush. 2009. Documentation <strong>fo</strong>r structure software:<br />

Version 2.3.<br />

(4 May 2010)<br />

Raymond, M. and F. Rousset. 1995. GENEPOP (version 1.2): population genetics<br />

software <strong>fo</strong>r exact tests and ecumenicism. Journal of Heredity. 86:248-249.<br />

Reed, J.E., R.J. Baker, W.B. Ballard, and B.T. Kelly. 2004. Differentiating Mexican gray<br />

wolf and coyote scats using DNA analysis. Wildlife Society Bulletin. 32(3): 685-692.<br />

Roughton, R.D. and M.W. Sweeny. 1982. Refinements in scent-station methodology <strong>fo</strong>r<br />

assessing trends in carnivore populations. Journal of Wildlife Management. 46(1): 217-<br />

229.<br />

Rousset, F. 2008. Genepop'007: a complete reimplementation of the Genepop software<br />

<strong>fo</strong>r Windows and Linux. Molecular Ecology Resources. 8:103-106.<br />

Sacks, B.N., S.K. Brown, and H.B. Ernest. 2004. Population structure of Cali<strong>fo</strong>rnia<br />

coyotes corresponds to habitat-specific breaks and illuminates species history. Molecular<br />

Ecology. 13: 1265-1275.<br />

Stenglein, J.L., M. De Ba<strong>rb</strong>a, D.E. Ausband, and L.P. Waits. 2010a. Impact of sampling<br />

location within a faeces on DNA quality in two carnivore species. Molecular Ecology<br />

Resources. 10: 109-114.<br />

Stenglein, J.L., L.P. Waits, D.E. Ausband, P. Zager, and C.M. Mack. 2010b. Efficient,<br />

noninvasive genetic sampling <strong>fo</strong>r monitoring introduced wolves. Journal of Wildlife<br />

Management. 74(5): 1050-1058.<br />

Taberlet, P., L.P. Waits, and G. Luikart. 1999. Noninvasive genetic sampling: look be<strong>fo</strong>re<br />

you leap. Trends in Ecology and Evolution. 14(8): 323-327.<br />

Tieszen, L.L., T.W. Boutton, K.G. Tesdahl, and N.A. Slade. 1983. Fractionation and<br />

turnover of stable ca<strong>rb</strong>on isotopes in animal tissues: Implications <strong>fo</strong>r δ 13 C analysis of<br />

diet. Oecologia. 57: 32-37.<br />

Valière, N. 2002. GIMLET: a computer program <strong>fo</strong>r analyzing genetic individual<br />

identification data. Molecular Ecology Notes. 10.1046/j.1471-8278 .2002.00228.x.<br />

65


66<br />

Waits, L.P. 2004. Using noninvasive genetic sampling to detect and estimate abundance<br />

of rare wildlife species. In W.L. Thompson (ed). Sampling Rare or Elusive Species:<br />

Concepts, Designs, and Techniques <strong>fo</strong>r Estimating Population Parameters. Washington,<br />

D.C.: Island Press. Pps. 211-228.<br />

Waits, L.P., G. Luikart, and P. Taberlet. 2001. Estimating the probability of identity<br />

among geno<strong>type</strong>s in natural populations: cautions and guidelines. Molecular Ecology. 10:<br />

249-256.<br />

Waits, L.P. and D. Paetkau. 2005. Noninvasive genetic sampling tools <strong>fo</strong>r wildlife<br />

biologists: A review of applications and recommendations <strong>fo</strong>r accurate data collection.<br />

Journal of Wildlife Management. 69(4): 1419-1433.<br />

Warne, R.W., A.D. Pershall, and B.O. Wolf. 2010. Linking precipitation and C3-C4 plant<br />

production to resource dynamics in higher-trophic-level consumers. Ecology. 91(6):<br />

1628-1638.<br />

Wilson, P.J., S.K. Grewal, F.F. Mallory, and B.N. White. 2009. Genetic characterization<br />

of hybrid wolves across Ontario. Journal of Heredity. 100 (Supplement I): S80-S89.<br />

Windberg, L.A., S.M. Ebbert, and B.T. Kelly. 1997. Population characteristics of coyotes<br />

(Canis latrans) in the Northern Chihuahuan Desert of New Mexico. American Midland<br />

Naturalist. 138(1): 197-207.


Chapter 3: Genetic results<br />

The purpose of this chapter is to provide detailed in<strong>fo</strong>rmation regarding the results<br />

of the genetic analyses of noninvasively collected scat samples that were described in<br />

Chapter 2. This chapter also serves as a reference <strong>fo</strong>r Chapters 4 and 5, which present the<br />

results of these genetic analyses in abbreviated <strong>fo</strong>rm. Several calculations were<br />

per<strong>fo</strong>rmed and maps were generated to assess the quality of the genetic data and thus the<br />

reliability of the individual identification in<strong>fo</strong>rmation that was derived from these data.<br />

This individual identification in<strong>fo</strong>rmation is presented and applied in later chapters.<br />

Results and discussion<br />

Species identification<br />

More than two-thirds of collected scat samples came from coyotes. The use of a<br />

genetic-based species identification test increased the number of samples identified as<br />

being from coyotes and decreased the number of samples with an “unknown” identity<br />

relative to the field-based identification technique (Figure 1).<br />

Number of samples<br />

700<br />

600<br />

500<br />

400<br />

300<br />

200<br />

100<br />

0<br />

Coyote Maybe<br />

Coyote<br />

Sample identification<br />

Not Coyote Unknown<br />

Field<br />

Genetics<br />

Figure 1. Species ID results. Number of samples identified as coyote, maybe coyote, not<br />

coyote or unknown either in the field (black) or based on a genetic analysis of species<br />

(gray).<br />

67


Individual identification<br />

There were a total of 81 individuals, 17 of which were only detected once. The<br />

consensus geno<strong>type</strong>s of all individuals that were only detected once were analyzed using<br />

RELIOTYPE and had estimated reliabilities between 0.99 and 1. The largest number of<br />

samples collected <strong>fo</strong>r a single individual was 25 (Figure 2). The average nearest distance<br />

between scat samples collected from a single individual ranged from 1 to 166 m.<br />

Frequency<br />

18<br />

16<br />

14<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

1 3 5 7 9 11 13 15 17 19 21 23 25<br />

Number of scats/individual<br />

Figure 2. Scat samples per individual. Distribution of number of samples per<br />

individual, with a total of 520 samples and 81 individuals.<br />

Per locus amplification success rates were all above 75 percent and ranged from<br />

78 to 88 percent. With one exception (allelic dropout at locus 2010), error rates were all<br />

below 10% (Table 1). These success rates are as high and error rates as low as those<br />

<strong>fo</strong>und in other, recent studies that have utilized microsatellite markers (e.g., Murphy et al.<br />

2003, Adams et al. 2007, Adams and Waits 2007, De Ba<strong>rb</strong>a and Waits 2010, Stenglein et<br />

68


69<br />

al. 2010a, Stenglein et al. 2010b). Genotyping success rate, calculated as the number of<br />

scat samples <strong>fo</strong>r which a consensus geno<strong>type</strong> was obtained at 6 or more loci (n = 520)<br />

divided by the total number of samples <strong>fo</strong>r which individual identification was attempted<br />

(n = 795), was roughly 65%.<br />

Table 1. Microsatellite success and error rates. Per locus success and error (allelic<br />

dropout, false allele) rates. Success rate indicates success of amplification via polymerase<br />

chain reaction (PCR) and is based on the first microsatellite PCR of 761 samples<br />

previously identified as “coyote” or “maybe coyote” via a genetic-based species<br />

identification test (Figure 1). Error rates are based on the first two successful PCRs <strong>fo</strong>r<br />

520 samples <strong>fo</strong>r which consensus was reached at 6 or more loci. The largest allelic<br />

dropout and false allele rates are shown in bold.<br />

Amplification Allelic<br />

Locus Success rate dropout False alleles<br />

CXX119 0.88 0.05 0.01<br />

CXX173 0.85 0.04 0.01<br />

CXX250 0.83 0.08 0.01<br />

CXX377 0.83 0.05 0.03<br />

FH2001 0.79 0.10 0.04<br />

CXX2010 0.78 0.13 0.01<br />

FH2054 0.80 0.09 0.05<br />

FH2088 0.86 0.05 0.01<br />

Per sample values <strong>fo</strong>r probability of identity (PID) <strong>fo</strong>r unrelated individuals<br />

ranged from 1.5 x 10 -11 to 2.9 x 10 -8 with an average of 5.4 x 10 -10 <strong>fo</strong>r 520 samples <strong>fo</strong>r<br />

which a consensus geno<strong>type</strong> was obtained <strong>fo</strong>r 6 to 8 loci. Per sample PID values <strong>fo</strong>r<br />

siblings <strong>fo</strong>r these same 520 samples ranged from 2.4 x 10 -4 to 2.3 x 10 -3 with an average<br />

of 4.5 x 10 -4 . Per sample PID values <strong>fo</strong>r siblings ranged from 1.4 x 10 -3 to 2.3 x 10 -3 <strong>fo</strong>r<br />

41 samples <strong>fo</strong>r which consensus was obtained at 6 loci (i.e., the minimum number of loci<br />

required <strong>fo</strong>r individual identification). All PID values, both <strong>fo</strong>r unrelated individuals and<br />

siblings, are within or lower than the recommended range of values (1 x 10 -2 to 1 x 10 -4 ).<br />

This indicates that the probability of two siblings or two unrelated individuals having the<br />

same consensus geno<strong>type</strong> is low (Waits et al. 2001). Three out of eight loci have


70<br />

significant p-values (below 0.05) <strong>fo</strong>r the Hardy-Weinberg (HW) exact test and are thus<br />

out of HW equilibrium. If a Bonferroni corrected p-value is used (0.006), then only one<br />

locus (CXX173) is out of HW equilibrium (Table 2A). When 31 related individuals, as<br />

identified using the kinship analysis in Gimlet 1.3.3., were removed from the population,<br />

either one locus (FH2088; p < 0.05) was, or no loci (p < 0.006) were, out of equilibrium<br />

(Table 2B). This indicates that a high degree of relatedness among some of the sampled<br />

individuals, rather than genotyping errors or locus-specific selection (Sacks et al. 2004),<br />

was likely causing the CXX173 locus to be out of HW equilibrium.<br />

Table 2. Hardy-Weinberg test results. P- and S.E. values <strong>fo</strong>r the Hardy-Weinberg<br />

(HW) exact test in GENEPOP 4.0.10. The p-values shown in bold are less than 0.05. An<br />

asterisk denotes a value below the Bonferroni corrected p-value of 0.006 (Sacks et al.<br />

2004). A) These are the values obtained when all sampled coyotes are included in the<br />

analysis (n = 81; 700 batches, S.E. < 0.01). B) These are the values obtained after related<br />

individuals were removed (n = 50, 600 batches, S.E. < 0.01). Related individuals were<br />

identified using the kinship analysis in Gimlet 1.3.3. Only one of every three individuals<br />

identified as being in a parent/offspring relationship by Gimlet was retained in the HW<br />

analysis.<br />

A) B)<br />

Locus p-value S.E.<br />

CXX119 0.562 0.008<br />

CXX173 0.000* 0.000<br />

CXX250 0.551 0.009<br />

CXX377 0.179 0.008<br />

FH2001 0.086 0.005<br />

CXX2010 0.017 0.001<br />

FH2054 0.099 0.005<br />

FH2088 0.022 0.002<br />

Locus p-value S.E.<br />

CXX119 0.266 0.008<br />

CXX173 0.085 0.006<br />

CXX250 0.673 0.009<br />

CXX377 0.178 0.009<br />

FH2001 0.450 0.009<br />

CXX2010 0.182 0.003<br />

FH2054 0.302 0.008<br />

FH2088 0.021 0.002<br />

Results of a linkage equilibrium (LE) test showed that, <strong>fo</strong>r eight out of 28 inter-<br />

locus comparisons (29%), the geno<strong>type</strong>s of the two loci being considered were not<br />

independent of one another (p < 0.05; Table 3A). Most of these significant comparisons<br />

were associated with at least one of the three loci <strong>fo</strong>und to be out of HW equilibrium


71<br />

(Table 2A). However, when a Bonferroni corrected p-value (0.002) was used, the null<br />

hypothesis of the LE test was rejected <strong>fo</strong>r only two comparisons, one of which was<br />

associated with the locus <strong>fo</strong>und to be out of HW equilibrium when a Bonferroni corrected<br />

p-value was used (CXX173; Tables 2A and 3A). When the same 31 related individuals<br />

that were removed from the HW equilibrium test were removed from the LE test, there<br />

was one significant comparison (p < 0.05) or zero significant comparisons (p < 0.002;<br />

Table 3B). The one significant comparison is not associated with the locus <strong>fo</strong>und to be<br />

out of HW equilibrium after the related individuals were removed (FH2088; Table 2B).<br />

Table 3. Linkage equilibrium test results. P-values <strong>fo</strong>r linkage equilibrium (LE)<br />

analysis in GENEPOP 4.0.10. The p-values shown in bold are less than 0.05. An asterisk<br />

denotes a value below the Bonferroni corrected p-value of 0.002 (Sacks et al. 2004). A)<br />

P-values obtained when all sampled coyotes are retained in the LE analysis (n = 81; 600<br />

batches; S.E. < 0.02). B) P-values obtained when related individuals, as identified using<br />

Gimlet 1.3.3 (see Table 2 <strong>fo</strong>r details), are removed from the analysis (n = 50, 600<br />

batches, S.E. < 0.02).<br />

A)<br />

Locus CXX119 CXX173 CXX250 CXX377 FH2001 CXX2010 FH2054 FH2088<br />

CXX119 x 0.046 0.238 0.156 0.059 0.229 0.673 0.001*<br />

CXX173 x 0.04 0.109 0.007 0.341 0.007 0.000*<br />

CXX250 x 0.1 0.125 0.276 0.126 0.053<br />

CXX377 x 0.196 0.303 0.609 0.119<br />

FH2001 x 0.014 0.045 0.238<br />

CXX2010 x 0.194 0.243<br />

FH2054 x 0.105<br />

FH2088<br />

B)<br />

x<br />

Locus CXX119 CXX173 CXX250 CXX377 FH2001 CXX2010 FH2054 FH2088<br />

CXX119 x 0.357 0.050 0.540 0.503 0.729 0.963 0.605<br />

CXX173 x 0.898 0.708 0.389 0.289 0.188 0.389<br />

CXX250 x 0.025 0.641 0.060 0.748 0.519<br />

CXX377 x 0.446 0.500 0.899 0.325<br />

FH2001 x 0.101 0.078 0.866<br />

CXX2010 x 0.291 0.163<br />

FH2054 x 0.648<br />

FH2088 x


72<br />

As was seen <strong>fo</strong>r the HW equilibrium test, the high degree of relatedness among some of<br />

the sampled individuals, rather than genotyping errors or physical linkage between one or<br />

more pairs of loci (Sacks et al. 2004), appears to be the primary factor leading to a<br />

rejection of the null hypothesis of the LE test <strong>fo</strong>r some loci. A comparison of values <strong>fo</strong>r<br />

observed and expected heterozygosity indicated that the deviation from HW equilibrium<br />

(Tables 2A and 2B) might be due to the presence of null alleles at one locus (FH2088;<br />

Tables 4A and 4B).<br />

Table 4. Observed and expected heterozygosity. Observed and expected numbers of<br />

heterozygous individuals (generated by GENEPOP 4.0.10) were divided by the number<br />

of individuals geno<strong>type</strong>d at a particular locus to determine observed (Hobs) and expected<br />

(Hexp) heterozygosity. Hexp > Hobs and p-value <strong>fo</strong>r HW exact test < 0.05 (Table 2) <strong>fo</strong>r<br />

locus shown in bold. A) Values <strong>fo</strong>r all coyotes sampled (n = 81); B) values once related<br />

individuals are removed (n = 50).<br />

A)<br />

B)<br />

Locus<br />

Expected #<br />

heterozygotes<br />

Observed #<br />

heterozygotes<br />

# of individuals<br />

geno<strong>type</strong>d Hexp Hobs<br />

CXX119 69.956 68 80 0.87 0.85<br />

CXX173 70.559 73 81 0.87 0.9<br />

CXX250 69.3665 68 81 0.86 0.84<br />

CXX377 71.677 70 81 0.89 0.86<br />

FH2001 65.6273 60 81 0.81 0.74<br />

CXX2010 57.7453 66 81 0.71 0.82<br />

FH2054 62.2795 61 81 0.77 0.75<br />

FH2088 66.4969 60 81 0.82 0.74<br />

Locus<br />

Expected #<br />

heterozygotes<br />

Observed #<br />

heterozygotes<br />

# of individuals<br />

geno<strong>type</strong>d Hexp Hobs<br />

CXX119 42.9175 41 49 0.88 0.84<br />

CXX173 43.7576 44 50 0.88 0.88<br />

CXX250 42.8182 43 50 0.86 0.86<br />

CXX377 44.9293 43 50 0.9 0.86<br />

FH2001 40.7273 38 50 0.81 0.76<br />

CXX2010 35.7576 40 50 0.72 0.8<br />

FH2054 39.5051 44 50 0.79 0.88<br />

FH2088 40.798 33 50 0.82 0.66


73<br />

More specifically, expected was larger than observed heterozygosity <strong>fo</strong>r this locus,<br />

indicating that there were many more samples with a homozygous geno<strong>type</strong> at this locus<br />

than expected. Since the occurrence of allelic dropout is fairly low <strong>fo</strong>r the FH2088 locus<br />

(Table 1), it is possible that this high level of homozygosity is due to the presence of one<br />

or more null alleles (i.e., alleles that can no longer be amplified due to the occurrence of a<br />

mutation). However, if Bonferroni corrected p-values are used to assess the significance<br />

of the HW exact test, then this locus is no longer <strong>fo</strong>und to be out of HW equilibrium.<br />

Furthermore, to my knowledge, no other studies have reported the presence of null alleles<br />

at this locus in coyotes, so it is possible that my observations regarding this locus are due<br />

to another cause. For example, this locus (FH2088) may be linked to another locus that is<br />

under selection (L.Waits, University of Idaho, personal communication).<br />

Related individuals and population structure<br />

A qualitative assessment of the spatial organization of groups of related<br />

individuals, or family groups, at the <strong>Sevilleta</strong> National Wildlife Refuge (NWR) shows<br />

that related individuals were often sampled either on only one scat transect, and thus on a<br />

mile-long stretch of road, or on multiple, contiguous transects. There is one notable<br />

exception; one family was <strong>fo</strong>und on transects in both the northeastern and southeastern<br />

parts of the refuge (family 3; Figure 3) and even extended to the western side of the<br />

refuge (family 3; Figure 4). There was a good deal of similarity in the number and spatial<br />

distribution of families identified using the results of the kinship analysis in Gimlet 1.3.3<br />

(Figure 3) and the relationship analysis in ML-Relate (Figure 4).


Figure 3. Gimlet family group map. Map of family groups based on the results of a<br />

kinship analysis in Gimlet 1.3.3. The members of each family group were determined by<br />

starting with a single offspring and its two parents and adding in all mates, parents, and<br />

offspring of 1) each of these three individuals, 2) each of the mates, parents and offspring<br />

of these three individuals.<br />

Figure 4. ML-Relate family group map. Map of family groups based on the results of a<br />

relationship analysis in ML-Relate. The members of each family group were determined<br />

by starting with a single offspring-parent pair and adding in all mates, parents, and<br />

offspring of 1) both of these individuals, 2) each of the mates, parents, and offspring of<br />

these two individuals. Families that share at least one individual with a family group<br />

determined using the kinship analysis in Gimlet 1.3.3 are shown with the same symbols<br />

as in Figure 3.<br />

74


In particular, the results of one program identified 8 and the other one 9 families and the<br />

membership and spatial distribution of several families overlap and are thus shown with<br />

the same symbols in Figures 3 and 4. There are however some differences between the<br />

results of the two programs. In particular, one family group (family 3) identified based on<br />

ML-Relate results has a much larger spatial distribution than the comparable family group<br />

identified using Gimlet (Figures 3 and 4). Furthermore, there are 5 families that either do<br />

not match any family group, or are subsumed by a larger family group, identified by the<br />

other program (Gimlet families 6, 7, and 8 Figure 3; ML-Relate families 6 and 7 Figure<br />

4).<br />

An analysis of the population structure of all 81 individuals shows that there are<br />

either 7 (Figure 5 Structure 2.3.1; aspatial) or 10 (Figure 6; BAPS 5; spatial) different<br />

genetic groups present at the study site. For Structure 2.3.1, 7 was the number of genetic<br />

groups with the largest average value <strong>fo</strong>r natural log probability of data (Figure 5). For<br />

BAPS 5, 10 was the optimal number of genetic groups identified in each of 20<br />

independent runs and had the largest log (ml) value (Figure 6). When only one of every<br />

three individuals identified as being related by the kinship analysis in Gimlet 1.3.3 are<br />

retained in the population structure analysis, the number of genetic groups declines to 1<br />

(Figure 5; Structure 2.3.1) or 2 (Figure 6; BAPS 5). The number of genetic groups<br />

identified in the population structure analyses (7 to 10) is similar to the number of family<br />

groups (8 to 9) present at the study site. Furthermore, the number of genetic groups<br />

identified declines significantly once related individuals are removed from the population<br />

structure analyses.<br />

75


Ln P(D)<br />

-1250<br />

-1750<br />

-2250<br />

-2750<br />

-3250<br />

-3750<br />

K<br />

1 2 3 4 5 6 7 8 9 10 11 12 13 14 15 16 17 18 19 20<br />

Structure_n=81<br />

Structure_n=50<br />

Figure 5. Structure analysis results. Average natural log probability of data as<br />

determined using Structure 2.3.1 <strong>fo</strong>r all individuals (n = 81) and once related individuals<br />

have been removed from the analysis (n = 50). Related individuals were removed by<br />

taking out two of every three individuals identified as being in a parent-offspring<br />

relationship by the kinship analysis in Gimlet 1.3.3. K = the number of genetic groups<br />

and error bars represent 95% confidence intervals across 20 independent runs <strong>fo</strong>r each K,<br />

except where outliers were removed.<br />

Log(ml)_n = 81<br />

-2620<br />

-2630<br />

-2640<br />

-2650<br />

-2660<br />

-2670<br />

-2680<br />

-2690<br />

-2700<br />

-2710<br />

-2720<br />

K<br />

1 2 3 4 5 6 7 8 9 1011121314151617181920<br />

-1660<br />

-1680<br />

-1700<br />

-1720<br />

-1740<br />

-1760<br />

-1780<br />

-1800<br />

-1820<br />

Log(ml)_n = 50<br />

BAPS_n=81_spatial<br />

BAPS_n=50_spatial<br />

Figure 6. BAPS analysis results. Average log(ml) values as determined using the spatial<br />

clustering of individuals analysis in BAPS 5 <strong>fo</strong>r all individuals (n = 81) and once related<br />

individuals have been removed (n = 50). Coordinates used <strong>fo</strong>r this analysis are the midpoints<br />

of all locations obtained <strong>fo</strong>r a particular individual. K = the number of genetic<br />

groups and error bars represent 95% confidence intervals across 10 independent runs <strong>fo</strong>r<br />

every K except K = 2.<br />

76


77<br />

These results indicate that relatedness among individuals accounts <strong>fo</strong>r most of the<br />

population structure identified by both an aspatial (Structure 2.3.1) and a spatial (BAPS<br />

5) analysis.<br />

A)<br />

B)<br />

Figure 7. Genetic groups generated by Structure. Genetic groups generated by<br />

Structure 2.3.1 (K = 7). A) Individuals were assigned to genetic groups based on the<br />

highest average membership coefficient (Q-value). B) Individuals were assigned to<br />

genetic groups only if they had an average Q-value of 0.8 or higher. For A and B, Qvalues<br />

were averaged across 20 independent runs and genetic groups that share multiple<br />

individuals with family groups determined by both Gimlet 1.3.3 and ML-Relate are<br />

shown with the same symbols as in Figures 3 and 4.


78<br />

There are some similarities between genetic groups identified using both Structure<br />

2.3.1 (Figure 7) and BAPS 5 (Figure 8) and the family groups based on results from<br />

Gimlet and ML-Relate (Figures 3 and 4). In particular, when individuals are assigned to<br />

genetic groups based on the largest average membership coefficient (Figure 7A), the<br />

composition of each of 4 out of 7 genetic groups identified by Structure overlaps that of<br />

one of the families identified by both Gimlet and ML-Relate and a fifth genetic group is<br />

similar to a family group identified by ML-Relate. One of the remaining genetic groups<br />

(genetic group 3) is primarily composed of individuals that were not assigned to a family<br />

group. When only large (0.8 or above) membership coefficients are considered (Figure<br />

7B), the compositions of 2 of the 3 genetic groups to which individuals are assigned<br />

correspond well with those of 2 family groups identified by Gimlet and ML-Relate.<br />

For the results of the BAPS 5 population structure analysis (Figure 8), the<br />

composition of each of 5 out of 10 genetic groups overlaps that of one of the family<br />

groups identified by both Gimlet and ML-Relate and a sixth genetic group matches a<br />

family identified by Gimlet. The 4 remaining genetic groups largely contain individuals<br />

that were not assigned to any family group. All of these similarities between genetic<br />

groups and family groups provide further support to the conclusion that most of the<br />

population structure identified using the programs Structure and BAPS is the result of<br />

relatedness among individuals rather than actual barriers to gene flow or the presence of<br />

more than one species in the dataset. This observed lack of barriers to gene flow is<br />

slightly surprising given the presence of both an interstate highway and a river that run<br />

north-south and divide the study site roughly in half. Other studies have <strong>fo</strong>und that roads<br />

can act as gene flow barriers <strong>fo</strong>r large mammals (Epps et al. 2005, bighorn sheep; Riley


79<br />

et al. 2006, bobcats and coyotes; Dixon et al. 2007, black bear; Pérez-Espona et al. 2008,<br />

red deer; Rashleigh et al. 2008, coyotes; but see Gula et al. 2009).<br />

Figure 8. Genetic groups generated by BAPS (K = 10). Genetic groups generated by<br />

spatial clustering of individuals analysis in BAPS 5 (K = 10). BAPS analysis included all<br />

81 individuals sampled at the study site. Genetic groups which contain several of the<br />

same individuals as family groups identified using Gimlet and ML-Relate are shown with<br />

the same symbols as in Figures 3 and 4.<br />

Once related individuals are removed from the BAPS 5 population structure<br />

analysis, most individuals are placed in the same genetic group (Figure 9). One individual<br />

was identified as being in a separate group in both BAPS 5 analyses (Figures 8 and 9).<br />

This individual was placed in a genetic group with other individuals in the Structure 2.3.1<br />

analysis (genetic group 3, Figure 7A), so it is unlikely that this individual is a member of<br />

a non-target species.


Figure 9. Genetic groups generated by BAPS (K = 2). Genetic groups generated by<br />

spatial clustering of individuals analysis in BAPS 5 (K = 2). This BAPS analysis was<br />

based on 50 individuals not identified by Gimlet 1.3.3 as being in a parent-offspring<br />

relationship with other individuals included in the analysis. The genetic group which<br />

contains the same individual as a genetic group identified by BAPS 5 when related<br />

individuals were included in the analysis (n = 81) is shown with the same symbol as in<br />

Figure 8.<br />

Coyote count variation between habitats and among transects and seasons<br />

The largest number of individuals were sampled on two shrubland scat transects,<br />

one on the eastern side and one on the western side of the <strong>Sevilleta</strong> NWR. There are<br />

many individuals in the north-eastern part of the study area, which is characterized by a<br />

grassland habitat, and the south-western part, which is <strong>cover</strong>ed by a honey mesquite<br />

(Prosopis glandulosa) dominated shrubland (Figure 10). There was no significant<br />

difference between grassland and shrubland habitat <strong>type</strong>s in the number of individuals<br />

sampled (t = 0.67, d.f. = 20, p = 0.51; Figure 11 and F = 3.59, d.f. = 1, 60, p = 0.06;<br />

Figure 12) and the interaction between habitat <strong>type</strong> and season did not represent a<br />

significant source of variation in the number of individuals sampled (F = 0.21, d.f. = 2,<br />

60, p = 0.81; Figure 12).<br />

80


Figure 10. Map of individuals per scat transect. Total number of individuals sampled<br />

on each scat transect during summer 2008 and spring, summer, and fall 2009. The<br />

boundary of the <strong>Sevilleta</strong> NWR is shown in solid light gray, the roads in dotted light<br />

gray, the scat transects in grassland areas in dark gray and the transects in shrubland areas<br />

in black.<br />

Average number of individuals<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

Grassland Shrubland<br />

<strong>Habitat</strong> <strong>type</strong><br />

Figure 11. Inter-habitat differences in number of individuals. Average number of<br />

individuals sampled along scat transects in grassland (n = 12) and shrubland (n = 10)<br />

habitats. <strong>Habitat</strong>-specific averages are based on counts averaged across 6-9 surveys of<br />

each of twenty-two scat transects.<br />

More coyotes were sampled in the spring than in the summer and fall (F = 9.06,<br />

d.f. = 2, 60, p = 0.0004; Figure 12). The decline in the number of individuals sampled in<br />

the summer and fall may be due to the increased rainfall during these seasons and thus an<br />

81


82<br />

increased likelihood of samples being washed away or degraded past recognition<br />

(Cavallini 1994). Dispersal of juvenile coyotes in the spring or fall may also be a<br />

contributing factor (Berg and Chesness 1978, Gese et al. 1996, Larrucea et al. 2006).<br />

Average number of individuals<br />

4<br />

3.5<br />

3<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

Spring Summer Fall<br />

Season<br />

Grassland<br />

Shrubland<br />

Figure 12. Seasonal variation in number of individuals. Average number of<br />

individuals sampled along scat transects in grassland (n = 12) versus shrubland (n = 10)<br />

habitat in each of three seasons (spring, summer, fall) in 2009. <strong>Habitat</strong>-specific averages<br />

are based on counts averaged across two surveys that were per<strong>fo</strong>rmed within each season<br />

<strong>fo</strong>r each scat transect. Significantly more coyotes were sampled in the spring (a) than in<br />

summer and fall (b; F = 9.06, d.f. = 2, 60, p = 0.0004).<br />

Home range analysis<br />

a<br />

b<br />

The average home range size <strong>fo</strong>r coyotes sampled 3 or more times among all<br />

surveys of the road-based scat transects is 1.7 km 2 (n = 58; Figure 13). This value is<br />

based on the average size of minimum convex polygons generated using scat sample<br />

collection locations (see Chapter 2 <strong>fo</strong>r more details). Roughly 9% of all individuals<br />

sampled (n = 81), and 11% of all individuals sampled two or more times (n = 64), used<br />

road-based scat transects both grassland and shrubland habitat <strong>type</strong>s, indicating that most<br />

b


83<br />

individuals stayed in only one of these two habitat <strong>type</strong>s. There is a significant difference<br />

in average home range size between coyotes that used one habitat <strong>type</strong> versus both<br />

grassland and shrubland habitat (F = 7.17, d.f. = 2, 55, p = 0.002; Figure 13). In<br />

particular, coyotes that used both habitat <strong>type</strong>s had, on average, larger home ranges (6.7<br />

km 2 ) than coyotes that used only grassland (1.3 km 2 ) or only shrubland (0.7 km 2 ) habitat.<br />

Roughly 51% of all individuals sampled (n = 81), and 64% of all individuals sampled 2<br />

or more times (n = 64), moved between different scat transects at least once during the<br />

time period over which scat samples were collected.<br />

Area (km 2 )<br />

12<br />

10<br />

8<br />

6<br />

4<br />

2<br />

0<br />

a<br />

Both Grassland Shrubland<br />

<strong>Habitat</strong> used<br />

Figure 13. Inter-habitat variation in home range size. Average size of coyote home<br />

ranges, as defined by minimum convex polygons, <strong>fo</strong>r individuals that stayed in only one<br />

habitat <strong>type</strong> (grassland versus shrubland, n = 51) or moved between habitats (both, n = 7).<br />

Error bars correspond to 95% confidence intervals. There is a significant difference in<br />

average home range size between coyotes that used both habitat <strong>type</strong>s (a) and those that<br />

only used one habitat <strong>type</strong> (b; F = 7.17, d.f. = 2, 55, p = 0.002). Please note that the<br />

ANOVA was per<strong>fo</strong>rmed on home range areas that were natural log trans<strong>fo</strong>rmed..<br />

b<br />

b


<strong>Habitat</strong> selection patterns of related individuals<br />

There is evidence that the <strong>type</strong> of habitat used by a given individual is not<br />

independent of the <strong>type</strong> of habitat use by related individuals. In particular, based on the<br />

results of the kinship analysis in Gimlet 1.3.3, the relationship analysis in ML-Relate<br />

(Figure 4), and in<strong>fo</strong>rmation on the habitat use patterns of individual coyotes, more pairs<br />

of related individuals than expected consisted of individuals that used the same habitat<br />

<strong>type</strong> and fewer pairs of related individuals than expected consisted of individuals that<br />

used different habitat <strong>type</strong>s (Table 5; Χ 2 = 33.95, d.f. = 2, p < 0.05, Gimlet; Χ 2 = 25.40,<br />

d.f. = 2, p < 0.05, ML-Relate).<br />

Table 5. Data <strong>fo</strong>r chi square analysis of pairs of related individuals. Observed and<br />

expected counts of pairs of individuals <strong>fo</strong>r which both individuals use the same habitat<br />

(Both G or Both S) or each individual uses a different habitat <strong>type</strong> (One G, One S; G =<br />

Grassland, S = Shrubland). Probability of occurrence was calculated based on the number<br />

of pair members that used a particular habitat <strong>type</strong> and the total number of pair members<br />

(i.e., the number of pairs multiplied by 2). Observed counts are higher than expected <strong>fo</strong>r<br />

pairs of individuals <strong>fo</strong>r which both individuals used the same <strong>type</strong> of habitat. Observed<br />

counts are lower than expected <strong>fo</strong>r pairs of individuals <strong>fo</strong>r which each individual used a<br />

different habitat <strong>type</strong>.<br />

Category<br />

Observed<br />

count<br />

Probability of<br />

occurrence<br />

Expected<br />

count<br />

Program used to<br />

assess family<br />

Both G 20 0.17 9 Gimlet<br />

One G, One S 6 0.48 27 Gimlet<br />

Both S 30 0.35 19 Gimlet<br />

Both G 20 0.19 11 ML-Relate<br />

One G, One S 9 0.49 28 ML-Relate<br />

Both S 27 0.32 18 ML-Relate<br />

A similar analysis of trios of related individuals showed that more trios than expected<br />

consisted of individuals that all used the same habitat <strong>type</strong> (Table 6, Χ 2 = 21.0, d.f. = 1, p<br />

< 0.05, Gimlet; Χ 2 = 9.14, d.f. = 1, p < 0.05, ML-Relate). Furthermore, all individuals<br />

used the same habitat <strong>type</strong> as all of their relatives in at least 50% of the family groups<br />

84


85<br />

identified using both Gimlet 1.3.3 (6 out of 9 family groups; Figures 3 and 14A) and ML-<br />

Relate (4 out of 8 family groups; Figures 4 and 14B). Regardless of the program used to<br />

determine family group (Gimlet 1.3.3 versus ML-Relate), most of the individuals that<br />

used the same habitat as all the relatives in their family group were <strong>fo</strong>und either on the<br />

western side of the refuge or in the central part of the eastern side of the refuge. Overall,<br />

these results are similar to those obtained in a study of coyotes in Cali<strong>fo</strong>rnia which<br />

showed that coyote population structure was related to different habitat “bioregions,”<br />

such that individuals that were more closely related were <strong>fo</strong>und in the same habitat<br />

(Sacks et al. 2004).<br />

Table 6. Data <strong>fo</strong>r chi square analysis of trios of related individuals. Observed and<br />

expected counts of trios of related individuals <strong>fo</strong>r which all three individuals use the same<br />

habitat (All 3 G or All 3 S) or the third individual uses a different habitat <strong>type</strong> from the<br />

other two (2 G 1 S, 2 S 1 G; G = Grassland, S = Shrubland). Probability of occurrence<br />

was calculated based on the number of pairs of individuals that both used the same<br />

habitat <strong>type</strong> and the number of “third” individuals that used either grassland or shrubland<br />

habitats. Observed counts are higher than expected <strong>fo</strong>r trios of individuals <strong>fo</strong>r which all<br />

three individuals used the same <strong>type</strong> of habitat. Observed counts are lower than expected<br />

<strong>fo</strong>r trios of individuals <strong>fo</strong>r which the third individual used a different habitat <strong>type</strong> than the<br />

other two individuals.<br />

Category<br />

Observed<br />

count<br />

Probability of<br />

occurrence<br />

Expected<br />

count<br />

Program used to<br />

assess family<br />

All 3 G 21 0.17 10 Gimlet<br />

2 G 1 S 11 0.23 14 Gimlet<br />

All 3 S 23 0.35 20 Gimlet<br />

2 S 1 G 4 0.25 15 Gimlet<br />

All 3 G 16 0.13 10 ML-Relate<br />

2 G 1 S 16 0.29 21 ML-Relate<br />

All 3 S 34 0.39 29 ML-Relate<br />

2 S 1 G 7 0.18 13 ML-Relate


A)<br />

B)<br />

Figure 14. Map of family groups relative to habitat use. Maps of family groups<br />

determined by A) Gimlet and B) ML-Relate. Families in which all individuals used the<br />

same habitat <strong>type</strong> are shown with dark symbols (circles, A; diamonds, B). Families in<br />

which individuals used different habitat <strong>type</strong>s from some of their relatives are shown with<br />

white symbols (squares, A; triangles, B).<br />

86


Conclusion<br />

A high percentage of collected scat samples were from coyotes and a minimum of<br />

81 different coyotes were sampled. Success rates associated with the individual<br />

identification process were high and error rates were low. Relatedness among individuals<br />

appears to account <strong>fo</strong>r most other irregularities in the data and <strong>fo</strong>r most of the population<br />

structure identified by two different programs. There was no difference in the number of<br />

coyotes sampled in grassland versus shrubland habitat, but more individuals were<br />

sampled in the spring than in either the summer or fall seasons. Most coyotes that were<br />

sampled stayed in either the grassland or the shrubland habitat and did not appear to<br />

move often between the two habitat <strong>type</strong>s. On average, the few coyotes that moved<br />

between habitat <strong>type</strong>s used a larger area than the coyotes that used only one habitat <strong>type</strong>.<br />

There is some evidence that both pairs and trios of related individuals tend to use the<br />

same habitat <strong>type</strong> and thus there appears to be habitat fidelity within family groups.<br />

87


References<br />

Adams, J.R., C. Lucash, L. Schutte, and L.P. Waits. 2007. Locating hybrid individuals in<br />

the red wolf (Canis rufus) experimental population area using a spatially targeted<br />

sampling strategy and faecal DNA genotyping. Molecular Ecology. 16: 1823-1834.<br />

Adams, J.R. and L.P. Waits. 2007. An efficient method <strong>fo</strong>r screening faecal DNA<br />

geno<strong>type</strong>s and detecting new individuals and hybrids in the red wolf (Canis rufus)<br />

experimental population area. Conservation Genetics. 8: 123-131.<br />

Berg, W.E. and R.A. Chesness. 1978. Ecology of coyotes in Northern Minnesota. In M.<br />

Bekoff (ed). Coyotes: Biology, Behavior, and Management. New York: Academic Press,<br />

Inc. Pps. 229-248.<br />

Cavallini, P. 1994. Faeces count as an index of <strong>fo</strong>x abundance. Acta Theriologica. 39(4):<br />

417-424.<br />

De Ba<strong>rb</strong>a, M. and L.P. Waits. 2010. Multiplex pre-amplification <strong>fo</strong>r noninvasive genetic<br />

sampling: is the extra ef<strong>fo</strong>rt worth it? Molecular Ecology Resources. 10: 659-665.<br />

Dixon, J.D., M.K. Oli, M.C. Wooten, T.H. Eason, J.W. McCown, and M.W.<br />

Cunningham. 2007. Genetic consequences of habitat fragmentation and loss: the case of<br />

the Florida black bear (Ursus americanus floridanus). Conservation Genetics. 8: 455-<br />

464.<br />

Epps, C.W., P.J. Palsboll, J.D. Wehausen, G.K. Roderick, R.R. Ramey II, and D.R.<br />

McCullough. 2005. Highways block gene flow and cause a rapid decline in genetic<br />

diversity of desert bighorn sheep. Ecology Letters. 8: 1029-1038.<br />

Gese, E.M., R.L. Ruff, and R.L. Crabtree. 1996. Social and nutritional factors influencing<br />

the dispersal of resident coyotes. Animal Behaviour. 52: 1025-1043.<br />

Gula, R., R. Hausknecht, and R. Kuehn. 2009. Evidence of wolf dispersal in<br />

anthropogenic habitats of the Polish Carpathian mountains. Biodiversity and<br />

Conservation. 18: 2173-2184.<br />

Larrucea, E.S., P.F. Brussard, M.M. Jaeger, and R.H. Barrett. 2007. Cameras, coyotes,<br />

and the assumption of equal detectability. Journal of Wildlife Management. 71(5): 1682-<br />

1689.<br />

Murphy, M.A., L.P. Waits, and K.C. Kendall. 2003. The influence of diet on faecal DNA<br />

amplification and sex identification in brown bears (Ursus arctos). Molecular Ecology.<br />

12: 2261-2265.<br />

88


89<br />

Pérez-Espona, S., F.J. Pérez-Ba<strong>rb</strong>eria, J.E. McLeod, C.D. Jiggins, I.J. Gordon, and J.M.<br />

Pemberton. 2008. Landscape features affect gene flow of Scottish Highland red deer<br />

(Cervus elaphus). Molecular Ecology.17: 981-996.<br />

Rashleigh, R.M., R.A. Krebs, and H. Van Keulen. 2008. Population structure of coyote<br />

(Canis latrans) in the u<strong>rb</strong>an landscape of the Cleveland, Ohio area. Ohio Journal of<br />

Science. 108(4): 54-59.<br />

Riley, S.P.D., J.P. Pollinger, R.M. Sauvajot, E.C. York, C. Bromley, T.K. Fuller, and<br />

R.K. Wayne. 2006. A southern Cali<strong>fo</strong>rnia freeway is a physical and social barrier to gene<br />

flow in carnivores. Molecular Ecology. 15: 1733-1741.<br />

Sacks, B.N., S.K. Brown, and H.B. Ernest. 2004. Population structure of Cali<strong>fo</strong>rnia<br />

coyotes corresponds to habitat-specific breaks and illuminates species history. Molecular<br />

Ecology. 13: 1265-1275.<br />

Stenglein, J.L., M. De Ba<strong>rb</strong>a, D.E. Ausband, and L.P. Waits. 2010a. Impacts of sampling<br />

location within a faeces on DNA quality in two carnivore species. Molecular Ecology<br />

Resources. 10: 109-114.<br />

Stenglein, J.L., L.P. Waits, D.E. Ausband, P. Zager, and C.M. Mack. 2010b. Efficient,<br />

noninvasive genetic sampling <strong>fo</strong>r monitoring introduced wolves. Journal of Wildlife<br />

Management. 74(5): 1050-1058.<br />

Waits, L.P., G. Luikart, and P. Taberlet. 2001. Estimating the probability of identity<br />

among geno<strong>type</strong>s in natural populations: cautions and guidelines. Molecular Ecology. 10:<br />

249-256.


90<br />

Chapter 4: Bottom-up effects of seasonal and long-term habitat change<br />

on predator feeding ecology<br />

Abstract<br />

Mammalian carnivores are susceptible to bottom-up effects associated with<br />

changes in the local environment. Woody plant encroachment, the spread of woody<br />

plants in grass-dominated landscapes, is one of many processes leading to habitat change<br />

on a global scale. This process occurs in dryland areas which tend to have strong seasonal<br />

variation in rainfall and primary productivity. This study examines the effects of both<br />

woody plant encroachment and seasonal change in grass-derived resource availability on<br />

the feeding ecology, specifically the base of the <strong>fo</strong>od chain, of a top predator. Seasonal<br />

changes in <strong>live</strong> grass <strong>cover</strong> were measured in grassland and woody plant-encroached<br />

shrubland habitats at the <strong>Sevilleta</strong> National Wildlife Refuge in central New Mexico,<br />

USA. Non-invasive genetic sampling of feces was used to identify individuals in the local<br />

coyote (Canis latrans) population. Stable ca<strong>rb</strong>on isotope analyses of pieces of bone and<br />

hair removed from the coyote scats were used to assess 1) the impacts of woody plant<br />

encroachment on coyote feeding ecology by comparing the percent of the coyote diet that<br />

came indirectly from C3 woody plants between grassland and shrubland habitats and 2)<br />

seasonal variation in the percent of the coyote diet that came indirectly from C4 grasses.<br />

Percent <strong>live</strong> grass <strong>cover</strong> increased from the spring to the summer and fall. A total of 81<br />

coyotes were sampled and there were no significant differences in percent coyote diet<br />

from C3 woody plants between grassland and shrubland habitats or in percent diet from<br />

C4 grasses among 3 seasons. However, percent coyote diet from C3 woody plants was<br />

large relative to the percentage of total <strong>live</strong> <strong>cover</strong> accounted <strong>fo</strong>r by woody plants in


91<br />

grassland areas. Additionally, percent coyote diet from C4 grasses <strong>fo</strong>llowed different<br />

trends in grassland versus shrubland habitats; percentages declined slightly from the<br />

spring to the summer in grassland areas but increased slightly in shrubland areas. Based<br />

on these results, woody plant encroachment and seasonal variation in <strong>live</strong> grass <strong>cover</strong> do<br />

not appear to lead to significant shifts in the base of the coyote <strong>fo</strong>od chain. However, C3<br />

plants are an important <strong>fo</strong>od resource <strong>fo</strong>r coyote prey in grassland areas in this arid<br />

ecosystem. Additionally, in the shrubland habitat, percent coyote diet from C4 grasses<br />

<strong>fo</strong>llows a trend somewhat similar to that of percent <strong>live</strong> grass <strong>cover</strong> among seasons. This<br />

study highlights the need <strong>fo</strong>r further investigation of the bottom-up effects of woody<br />

plant encroachment, and variation in the availability of C3 versus C4 plants at the base of<br />

the <strong>fo</strong>od chain, on the ecology and fitness of top predators, especially specialized<br />

predators, and their prey.<br />

Introduction<br />

Predators play an important role in many ecosystems (e.g., Paine 1969, McLaren<br />

and Peterson 1994, Estes and Duggins 1995, Crooks and Soule 1999, Schmitz et al. 2000,<br />

Ripple and Beschta 2004) and are also prone to bottom-up effects associated with<br />

changes in the abundance or productivity of organisms in lower trophic levels (Brand et<br />

al. 1976, White 1978, Todd et al. 1981, King 1983, Jaksic et al. 1997, Stenseth et al.<br />

1997, Previtali et al. 2009). Pulses in primary production and plant-derived <strong>fo</strong>od<br />

resources associated with either rainfall in an arid environment (Jaksic et al. 1997) or tree<br />

masting in a <strong>fo</strong>rested environment (King 1983, Ostfeld and Keesing 2000) lead to spikes<br />

in both primary and secondary consumer populations (King 1983, Jaksic et al. 1997). The<br />

populations of some predators, including coyotes and lynx (Lynx canadensis), have been


92<br />

shown to closely track the abundance of prey species, with peaks in prey population size<br />

leading to peaks in predator abundance and population growth (Brand et al. 1976, Todd et<br />

al. 1981, O’Donoghue et al. 1997, Hone et al. 2007). Overall, the decline of a top<br />

predator can have significant effects on a local community (e.g., Paine 1969, Estes and<br />

Duggins 1995, Crooks and Soule 1999, Ripple and Beschta 2004), and top predators are<br />

particularly sensitive to changes that occur at the base of the <strong>fo</strong>od chain. These factors<br />

make top predators ideal <strong>fo</strong>cal organisms <strong>fo</strong>r investigations of the impacts of habitat<br />

change on biotic communities.<br />

Woody plant encroachment is a widespread process of habitat change occurring in<br />

dryland areas around the world (Archer 1995, Ravi et al. 2009), yet relatively little is<br />

known about its effects on local, top predators (Blaum et al. 2007a). This habitat change<br />

is characterized by the local proliferation of one or more native woody plant species in a<br />

grass-dominated habitat that results in a shift from grassland to shrubland or savanna to<br />

woodland habitat over a period of 50-200 years (Archer 1989, Archer 1995, Gill and<br />

Burke 1999, Van Auken 2000). Drivers of woody plant encroachment include:<br />

overgrazing by <strong>live</strong>stock (Bester 1996, Van Auken 2000), changes in local fire frequency<br />

(Archer 1995, Roques et al. 2001), elevated levels of ca<strong>rb</strong>on dioxide (Bond and Midgley<br />

2000), and nitrogen deposition (Köchy and Wilson 2001, Wigley et al. 2010). This<br />

habitat shift has many implications <strong>fo</strong>r the local biological community, especially<br />

predators. Woody plant encroachment leads to changes in habitat structure that can<br />

hinder predator movement and the ability of predators to detect prey (Muntifering et al.<br />

2006) and reduce both their hunting success (Mills et al. 2004) and prey availability<br />

(Marker 2002, Blaum et al. 2007b). A large number and diversity of mammalian


93<br />

carnivores are present in areas affected by woody plant encroachment, and more research<br />

regarding the bottom-up effects of this habitat shift on predators is needed (Blaum et al.<br />

2007a, Chapter 1).<br />

Woody plant encroachment occurs in grassland and savanna habitats in dryland<br />

areas globally (i.e., arid, semi-arid and dry-subhumid regions; Archer 1995, Van Auken<br />

2000, D’Odorico and Porporato 2006, Ravi et al. 2009, Chapter 1). These areas have low<br />

annual rates of precipitation (100-1200 mm/yr; D’Odorico and Porporato 2006). In many<br />

dryland areas, droughts are common and rainfall events are concentrated during a<br />

relatively short winter or summer rainy season (Noy-Meir 1973, D’Odorico and<br />

Porporato 2006). This seasonal climatic variation has bottom-up effects on the local<br />

biotic community as a result of the tight coupling between water availability and various<br />

ecological processes, including grass primary production (Noy-Meir 1973, Ernest et al.<br />

2000, Schwinning and Sala 2004, Muldavin et al. 2008, Warne et al. 2010b). Seasonal<br />

variation in both prey availability and consumer diet has been observed in dryland areas<br />

(Andelt et al. 1987, Hernández et al. 2002, Hernández et al. 2005, Warne et al. 2010b).<br />

Of particular relevance is the observation of a sharp increase in net primary production of<br />

grasses and a corresponding increase in their use during the summer monsoon by<br />

consumers in a woody plant-encroached area (Báez and Collins 2008, Muldavin et al.<br />

2008, Warne et al. 2010b). This highlights the impact that seasonal variation in primary<br />

productivity can have on the feeding ecology of organisms in higher trophic levels.<br />

The coyote is an ideal <strong>fo</strong>cal species <strong>fo</strong>r a study on the impacts of both woody<br />

plant encroachment and seasonal habitat change on mammalian predator feeding ecology.<br />

Woody plant encroachment has been documented at many sites in North America


94<br />

(Archer 1995, Ravi et al. 2009, Chapter 1) and the distribution of the coyote overlaps the<br />

majority of these sites as it encompasses most of the United States, Canada and Mexico<br />

(IUCN 2010). The coyote is a common, generalist species (IUCN 2010) that can also be a<br />

top predator and play an important role in the dynamics of the local community (Crooks<br />

and Soule 1999). Coyote diet varies geographically (e.g., Hidalgo-Mihart et al. 2001,<br />

Samson and Crete 1997) and in response to both short- and long-term changes in <strong>fo</strong>od<br />

resource availability (Todd et al. 1981, Hamlin et al. 1984, Andelt et al. 1987, Windberg<br />

and Mitchell 1990, Young et al. 2006). Of particular interest is the observation that<br />

coyote diet tracks changes in the availability of fruits from different woody plants both<br />

seasonally and over a period of nearly twenty years (Andelt et al. 1987). Additionally,<br />

two potential prey species of the coyote have been <strong>fo</strong>und to shift their diet in response to<br />

an increase in grass primary production during the summer monsoon season (Muldavin et<br />

al. 2008, Warne et al. 2010b). All of these factors indicate that the feeding ecology of<br />

coyotes <strong>fo</strong>und in an arid, woody plant-encroached area is likely to be affected both by the<br />

increase in woody plant-derived resources associated with woody plant encroachment<br />

and the spike in grass-derived resources that is driven by the seasonal variation in rainfall<br />

typical of arid environments.<br />

Shifts in consumer feeding ecology in response to habitat changes associated with<br />

both woody plant encroachment and seasonal variation in rainfall can be analyzed using<br />

stable ca<strong>rb</strong>on isotope techniques. C3 and C4 plants have distinct ca<strong>rb</strong>on isotope signatures<br />

as a result of differences in their photosynthetic pathways (Bender 1971, Farquhar 1983,<br />

Sternberg et al. 1984, Marshall et al. 2007, Warne et al. 2010b). Encroaching, woody<br />

plants use the C3 photosynthetic pathway while grasses at the same sites are often C4


95<br />

plants (e.g., Boutton et al. 1998, Gill and Burke 1999, McKinley and Blair 2008,<br />

Muldavin et al. 2008). Ca<strong>rb</strong>on isotope techniques have been used to determine whether<br />

various he<strong>rb</strong>ivores are browsers that eat C3 plants, or grazers that <strong>fo</strong>rage on C4 grasses<br />

(Ambrose and DeNiro 1986, Codron et al. 2005, Codron et al. 2007, Wallington et al.<br />

2007). These techniques have also shown that both primary and secondary consumers in<br />

a woody plant-encroached area shift their diet from C3- to C4-derived resources during<br />

the summer monsoon season (Muldavin et al. 2008, Warne et al. 2010b).<br />

My primary objective is to use stable ca<strong>rb</strong>on isotope techniques to assess the<br />

impact that long-term habitat change associated with woody plant encroachment and<br />

short-term habitat change driven by seasonal climatic variation has on the feeding<br />

ecology of a common, generalist predator. The facet of coyote feeding ecology that the<br />

use of stable ca<strong>rb</strong>on isotopes enables me to assess is the base of the coyote <strong>fo</strong>od chain<br />

and thus whether coyote prey are feeding on C4 grasses or C3 plants. My specific<br />

questions are: 1) Is there a difference in the base of the coyote <strong>fo</strong>od chain between native<br />

grassland and woody plant-encroached shrubland habitats? and 2) Is there seasonal<br />

variation in the base of the coyote <strong>fo</strong>od chain in an area impacted by woody plant<br />

encroachment? I expected that the base of the coyote <strong>fo</strong>od chain would differ between<br />

grassland and shrubland habitats such that the percent of the coyote diet coming<br />

indirectly from C3 woody plants would be higher <strong>fo</strong>r coyotes sampled in the shrubland<br />

habitat. I also expected that there would be a seasonal shift in the base of the <strong>fo</strong>od chain<br />

with percent coyote diet coming indirectly from C4 grasses increasing from the spring to<br />

the summer and fall in both grassland and shrubland habitats.


Methods<br />

Study site<br />

The fieldwork <strong>fo</strong>r this study was per<strong>fo</strong>rmed at the <strong>Sevilleta</strong> National Wildlife<br />

Refuge (NWR) and Long Term Ecological Research (<strong>LTER</strong>) site in central New Mexico<br />

(34.32°N, 106.81°W). The refuge encompasses 1,000 km 2 (Hernández et al. 2002), has<br />

roughly 200 miles of road, and contains grassland, shrubland and woodland habitat <strong>type</strong>s<br />

(Figure 1). The average annual rainfall from 1988-2008 was 235 mm (based on Moore<br />

2009). The Rio Grande divides the refuge into eastern and western parts that have<br />

distinct plant communities. The grassland on the eastern side of the refuge, and thus to<br />

the east of the Rio Grande, is dominated by grama grass (Bouteloua spp.) and the<br />

shrubland is dominated by creosote bush (Larrea tridentata), which has grasses growing<br />

in gaps among the shrubs. The grass <strong>cover</strong> on the western side is sparse relative to the<br />

eastern side. Honey mesquite (Prosopis glandulosa), which often grows on top of sand<br />

dunes, dominates the shrubland on the western side.<br />

Figure 1. Map of study site. Map of the <strong>Sevilleta</strong> NWR and <strong>LTER</strong> study site in central<br />

New Mexico, USA. Black = shrubland, gray = grassland, white = other land <strong>cover</strong> <strong>type</strong>s.<br />

The path of the Rio Grande through the center of the refuge is indicated by a thick black<br />

line, while the boundary of the refuge is shown by a light gray line.<br />

96


97<br />

Approximately 72% of the refuge is <strong>cover</strong>ed by either grassland or shrubland.<br />

Grassland is more prevalent in the northern half of the refuge and shrubland in the<br />

southern half. The eastern side of the refuge contains an active transition zone between<br />

grama grassland and creosote shrubland. More specifically, creosote shrubs have been<br />

moving into grama grassland areas at the refuge over the past century (Gill and Burke<br />

1999). These characteristics make the refuge an ideal setting <strong>fo</strong>r an assessment of the<br />

impact that woody plant encroachment has on coyote ecology.<br />

Scat surveys<br />

All field and laboratory methods outlined below are described in more detail in<br />

Chapter 2. Carnivore scat samples were collected between June 2008 and November<br />

2009. 3 surveys of 20 road-based transects (10 in grassland, 10 in shrubland) were<br />

per<strong>fo</strong>rmed in the summer (June-July) of 2008. 6 surveys of 22 transects (12 in grassland,<br />

10 in shrubland) were conducted in 2009, two in each of three seasons: spring (April-<br />

May), summer (July), and fall (October-November). Each transect was 1 mile (1.6 km)<br />

long and was separated from all other transects by at least a mile (1.6 km). Transects<br />

were driven at slow speed (< 16.1 km/hr) and each carnivore scat that was encountered<br />

was measured and sub-sampled <strong>fo</strong>r genetic analysis (see below). Measurements included<br />

length and maximum diameter. Roughly 0.4 mL of the fecal material on the outside of<br />

each sample was placed in a 2 mL tube containing DETs (DMSO, EDTA, Tris, salt)<br />

buffer (Frantzen et al. 1998). A GPS coordinate was recorded be<strong>fo</strong>re the remainder of the<br />

sample was collected <strong>fo</strong>r ca<strong>rb</strong>on isotope analysis (see below).


Genetic analyses<br />

All sub-samples stored in DETs buffer were extracted using QIAamp DNA Stool<br />

Mini Kits from Qiagen Inc. Extractions were per<strong>fo</strong>rmed in a laboratory space that was<br />

separate from the area where DNA was amplified and a negative control was processed<br />

with each set of extracted samples (Onorato et al. 2006). Primers (Murphy et al. 2000)<br />

that amplify a segment of the mitochondrial DNA (mtDNA) control region (Onorato et<br />

al. 2006) and polymerase chain reaction (PCR) techniques (see Chapter 2 <strong>fo</strong>r PCR<br />

profiles) were used to identify the species that had deposited each scat sample and screen<br />

out any samples deposited by canids other than the coyote. All samples <strong>fo</strong>und to be from<br />

coyotes were run through a microsatellite analysis in order to determine the minimum<br />

number of individual coyotes that had been sampled. In particular, coyote samples were<br />

screened using primers <strong>fo</strong>r 8 canid-specific microsatellite loci (CXX173 and CXX250,<br />

Ostrander et al. 1993; CXX377, Ostrander et al. 1995; FH2001, CXX2010, FH2054, and<br />

FH2088, Francisco et al. 1996; CXX119) and per locus success rates were calculated<br />

based on the results. Samples <strong>fo</strong>r which 5 or more loci were successfully amplified were<br />

screened 2-6 more times with all 8 primers. Replicate microsatellite PCR results were<br />

compared in order to obtain a consensus geno<strong>type</strong> <strong>fo</strong>r each sample. For a homozygous<br />

locus, one allele, and only that allele, had to be seen in all replicate PCRs while, <strong>fo</strong>r a<br />

heterozygous locus, each allele had to be seen at least twice. All PCR products were<br />

separated via electrophoresis on an Applied Biosystems 3130xl capillary machine and all<br />

microsatellite data was viewed using GeneMapper 3.7.<br />

All samples <strong>fo</strong>r which a consensus geno<strong>type</strong> was obtained at 6 or more loci were<br />

analyzed using Gimlet 1.3.3. (Valière 2002) in order to determine the minimum number<br />

98


99<br />

of individuals from which samples had been collected. The reliability of geno<strong>type</strong>s that<br />

were unique in the population was tested using RELIOTYPE (Miller et al. 2002) and a<br />

reliability criteria of 95%. Per locus error rates, <strong>fo</strong>r both allelic dropout and false alleles,<br />

were determined using the first two successful microsatellite PCRs <strong>fo</strong>r samples that were<br />

included in the Gimlet analysis and were not excluded <strong>fo</strong>llowing analysis with<br />

RELIOTYPE. GENEPOP 4.0.10 (Raymond and Rousset 1995, Rousset 2008) was used<br />

to determine if the geno<strong>type</strong>s were in Hardy-Weinberg (HW) equilibrium and if they met<br />

the assumption of linkage equilibrium (LE). Deviation from HW equilibirum or rejection<br />

of the null hypothesis of the LE test could indicate that there were errors in the consensus<br />

geno<strong>type</strong>s used to determine the minimum number of individual coyotes sampled (Sacks<br />

et al. 2004). Structure 2.3.1 (Pritchard et al. 2000) and BAPS 5 (Corander et al. 2008a,<br />

Corander et al. 2008b) and the kinship analysis in Gimlet 1.3.3 were used to determine<br />

whether there was any population structure that could not be explained in terms of<br />

relatedness among individuals.<br />

Vegetation surveys<br />

Data on percent <strong>live</strong> grass and woody plant <strong>cover</strong> were collected in 22 circular<br />

plots, one per scat transect. Each plot was 30 m in diameter and was located at a<br />

randomly selected site within 100 m of one of the road-based scat transects. Vegetation<br />

surveys were per<strong>fo</strong>rmed in each of the three seasons in 2009 in which scat surveys were<br />

conducted (spring, summer, fall). For each survey, percent <strong>live</strong> <strong>cover</strong> data were collected<br />

in each circular plot along 2, orthogonal line intercept transects that were each 30 m long.<br />

Points at which <strong>live</strong> vegetation intersected transects were recorded to the nearest 0.1 m.<br />

Vegetation was considered to be a<strong>live</strong> if its leaves (woody plants and grasses) or stems


100<br />

(grasses) were green. Small samples of the dominant woody plant species and one of the<br />

dominant grass genera <strong>fo</strong>und within each plot were collected in each season <strong>fo</strong>r ca<strong>rb</strong>on<br />

isotope analysis (see below).<br />

Ca<strong>rb</strong>on isotope analyses<br />

Prior to per<strong>fo</strong>rming ca<strong>rb</strong>on isotope analysis, all scat (70 o C) and vegetation (60 o C)<br />

samples were oven-dried <strong>fo</strong>r 24 hours. Vegetation samples were ground with a Wiley<br />

Mill until they passed through a 20 mesh filter. Small pieces of bone and hair were<br />

removed from the first scat sample collected from each individual in the summer of 2008<br />

and in each of the three seasons that surveys were per<strong>fo</strong>rmed in 2009. Individuals that<br />

moved between scat transects that were in different habitat <strong>type</strong>s at any point during the<br />

study were excluded from this analysis. Hair was only removed if a given individual had<br />

been sampled in more than one season in 2009. To test the validity of this approach and<br />

ensure that it did not overlook any impact of intra-individual variation in diet, bone and<br />

hair were removed from all samples collected within a single season from 6 (bone) or 7<br />

(hair) individuals sampled in different parts of the study area (Appendix 1). Each of the<br />

individuals used only one habitat <strong>type</strong> (grassland or shrubland).<br />

Bone and hair pieces were cleaned with ethanol and bone was crushed to a<br />

powder. Small masses of each item (vegetation, 4mg; hair, 1mg; bone, 3mg) were<br />

weighed into 5x9mm tin capsules. These weighed samples were analyzed on a coupled<br />

elemental analyzer (Costech 4010) and isotope ratio mass spectrometer (DeltaPlux XP) at<br />

the Stable Isotope Laboratory at the University of New Hampshire (UNH). In general, 4<br />

out of every 58 samples were replicated and 3 standards (NIST 1515 and 1575a, tuna)<br />

and one known “unknown” (bolete) were run <strong>fo</strong>r every 10 samples. 20 tins containing a


101<br />

second known unknown (citrus) were mixed in with the first 277 samples run. The<br />

ca<strong>rb</strong>on isotope signatures (i.e., δ 13 C values) derived from the mass spectrometer were<br />

corrected <strong>fo</strong>r shifts both over time and in response to variation in sample weight (A.<br />

Ouimette, UNH, 2010, personal communication). The signatures <strong>fo</strong>r bone and hair<br />

samples were also corrected <strong>fo</strong>r diet to tissue (1 o /oo, bone, DeNiro and Epstein 1978; 1.0<br />

o /oo, hair, Tieszen et al. 1983) and diet to scat (0 o /oo, both bone and hair, Chapter 2)<br />

fractionation. The corrected signatures were used to calculate the percentage of coyote<br />

diet of each individual in each season that came indirectly from C3 woody plants<br />

(Equation 1, based on Faure and Mensing 2005). These percentages were subtracted from<br />

100% to obtain the percentage of coyote diet that came indirectly from C4 grasses.<br />

Where<br />

δ 13 Cscat = δ 13 CC 3 (fC 3 ) + δ 13 CC 4 (1-fC 3 ) (1)<br />

δ 13 Cscat = the ca<strong>rb</strong>on isotope signature of a coyote scat component (i.e., bone, hair);<br />

δ 13 CC 3 = the average ca<strong>rb</strong>on isotope signature of C3 woody plants; fC 3 = the fraction of<br />

coyote diet that comes indirectly from C3 woody plants; δ 13 CC 4 = the average ca<strong>rb</strong>on<br />

isotope signature of C4 grasses.<br />

Statistical analyses<br />

Vegetation survey data was used to assess differences in habitat characteristics<br />

between grassland and shrubland habitats both on average and among seasons. Percent<br />

<strong>live</strong> grass and woody plant <strong>cover</strong> values were averaged across seasons, arcsine<br />

trans<strong>fo</strong>rmed to meet the assumption of normality, and analyzed using a two-way<br />

ANOVA (PROC GLM, SAS 9.1), with habitat and <strong>cover</strong> <strong>type</strong> as independent variables.<br />

Two a-priori contrasts were per<strong>fo</strong>rmed as part of this two-way ANOVA. These contrasts


102<br />

compared percent <strong>live</strong> grass and percent <strong>live</strong> woody plant <strong>cover</strong> between grassland and<br />

shrubland habitats. Percent <strong>live</strong> grass values were analyzed using a second two-way<br />

ANOVA with season and habitat as independent variables. A Student’s t-test (PROC<br />

TTEST, SAS 9.1) was used to assess differences in the percent of coyote diet coming<br />

indirectly from C3 woody plants <strong>fo</strong>r individuals sampled in native grassland versus<br />

woody plant-encroached shrubland habitats. The coyote diet values used <strong>fo</strong>r this analysis<br />

were derived from the stable ca<strong>rb</strong>on isotope signatures of bones. Bones represent a<br />

dietary record that can span the life of an organism, while hair integrates dietary<br />

in<strong>fo</strong>rmation <strong>fo</strong>r shorter periods of time (Ambrose and DeNiro 1986, Craw<strong>fo</strong>rd et al.<br />

2008). The stable ca<strong>rb</strong>on isotope signatures of bone are thus more likely than those of<br />

hair to reflect any long-term differences in <strong>fo</strong>raging pattern between grassland and<br />

shrubland habitats, while hair is likely to pick up shorter term variation. A repeated<br />

measures ANOVA (PROC MIXED, SAS 9.1) was used to assess seasonal variation in<br />

the percent of the coyote diet coming indirectly from C4 grasses <strong>fo</strong>r individuals sampled<br />

in grassland versus shrubland habitats. Ca<strong>rb</strong>on isotope signatures of hair samples were<br />

used to calculate the diet values <strong>fo</strong>r this analysis.<br />

Results<br />

Scat surveys and genetic analyses<br />

A total of 935 scat samples were sub-sampled <strong>fo</strong>r genetic analysis. The mtDNA<br />

species identification test indicated that just over two thirds of these samples had been<br />

deposited by coyotes. Analysis of the 520 scats (65% of 795 samples <strong>fo</strong>r which<br />

individual identification was attempted) <strong>fo</strong>r which consensus geno<strong>type</strong>s were obtained at<br />

6 or more microsatellite loci in Gimlet 1.3.3 indicated that a minimum of 81 individuals


103<br />

had been sampled. Per locus success rates <strong>fo</strong>r the microsatellite analysis ranged from 0.78<br />

to 0.88. Per locus error rates ranged from 0.04 to 0.13 <strong>fo</strong>r allelic dropout and from 0.01 to<br />

0.05 <strong>fo</strong>r false alleles. Results of the Hardy-Weinberg (HW) exact test and linkage<br />

equilibrium analysis in GENEPOP 4.0.10 indicated that, once relatedness among<br />

individuals was accounted <strong>fo</strong>r, one locus was out of HW equilibrium (FH2088, p < 0.05)<br />

and geno<strong>type</strong>s at one pair of loci (CXX250, CXX377, p < 0.05) were not independent of<br />

one another. If Bonferroni corrected p-values are used (0.05/number of comparisons;<br />

based on Sacks et al. 2004), then no loci were out of HW equilibrium and geno<strong>type</strong>s at all<br />

loci were independent of one another. Results from Structure 2.3.1 and BAPS 5 indicated<br />

that, when all 81 individuals were included, there were 7-10 genetic groups at the study<br />

site. When 31 related individuals were removed from the Structure 2.3.1 and BAPS 5<br />

analyses, the number of genetic groups declined to only 1 or 2 (see Chapter 3 <strong>fo</strong>r further<br />

details).<br />

Vegetation surveys and stable ca<strong>rb</strong>on isotope analyses<br />

The two-way ANOVA analysis of arcsine trans<strong>fo</strong>rmed values of <strong>live</strong> percent<br />

grass and woody plant <strong>cover</strong> between habitat <strong>type</strong>s indicated that, while there were no<br />

significant differences between habitat <strong>type</strong>s, there were significant differences between<br />

<strong>cover</strong> <strong>type</strong>s and there was a habitat*<strong>cover</strong> interaction (Table 1, Figure 2A).<br />

Table 1. Two-way ANOVA of percent <strong>live</strong> <strong>cover</strong>. Results of a two-way ANOVA <strong>fo</strong>r<br />

arcsine trans<strong>fo</strong>rmed values of percent <strong>live</strong> <strong>cover</strong> with habitat and <strong>cover</strong> <strong>type</strong> as<br />

independent variables (d.f. = degrees of freedom).<br />

Source d.f. F-value p-value<br />

<strong>Habitat</strong> 1 0.17 0.69<br />

Cover 1 60.95 < 0.01<br />

<strong>Habitat</strong>*Cover 1 11.37 < 0.01<br />

Error 40 . .


The results of two a-priori contrasts comparing percent <strong>live</strong> grass and percent <strong>live</strong> woody<br />

plant <strong>cover</strong> between habitat <strong>type</strong>s were evaluated at the Bonferroni corrected p-value of<br />

0.025. These contrasts indicated that percent <strong>live</strong> grass <strong>cover</strong> did not differ significantly<br />

between habitats (F = 4.40, p = 0.04, d.f. = 1, 40) but that percent woody plant <strong>cover</strong> was<br />

higher in shrubland areas (F = 7.14, p = 0.01, d.f. = 1, 40; Figure 2A). Results of the two-<br />

way ANOVA and Duncan’s test of percent <strong>live</strong> grass <strong>cover</strong> indicates that there are<br />

differences between habitat <strong>type</strong>s and between the spring and the summer and fall<br />

seasons but that there is no habitat*season interaction (Table 2, Figure 2B).<br />

104<br />

Table 2. Two-way ANOVA of percent <strong>live</strong> grass <strong>cover</strong>. Results of a two-way ANOVA<br />

<strong>fo</strong>r untrans<strong>fo</strong>rmed values of percent <strong>live</strong> grass <strong>cover</strong> with habitat and season as<br />

independent variables (d.f. = degrees of freedom).<br />

Source d.f. F-value p-value<br />

<strong>Habitat</strong> 1 4.59 0.04<br />

Season 2 9.73 < 0.01<br />

<strong>Habitat</strong>*Season 2 0.54 0.58<br />

Error 58 . .<br />

Average differences (± 1 standard deviation, i.e., s.d.) between observed and true<br />

values <strong>fo</strong>r the two known unknowns were small (0 ± 0.1 o /oo, bolete; 0.3 ± 0.1 o /oo, citrus).<br />

The standard deviation of all ca<strong>rb</strong>on isotope signatures obtained was 0.1 o /oo <strong>fo</strong>r each of<br />

the three standards (NIST 1515 and 1575a, tuna). The δ 13 C values <strong>fo</strong>r 18 samples of<br />

dominant grasses, representing species from 6 genera, were used to calculate an average<br />

(± 1 s.d.) ca<strong>rb</strong>on isotope signature <strong>fo</strong>r C4 grasses (-14.8 o /oo ± 0.3, Appendix 2). The δ 13 C<br />

values <strong>fo</strong>r 14 samples of 5 dominant woody plant species, as well as 4 samples of two<br />

woody plant species which produce fruit or seeds eaten by coyotes (V. Seamster,


105<br />

unpublished data), were used to calculate an average (± 1 s.d.) ca<strong>rb</strong>on isotope signature<br />

<strong>fo</strong>r C3 woody plants (-24.9 o /oo ± 0.7, Appendix 2).<br />

A) B)<br />

% <strong>live</strong> <strong>cover</strong><br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Grassland Shrubland<br />

<strong>Habitat</strong> <strong>type</strong><br />

Grass<br />

Woody<br />

C) D)<br />

% coyote diet<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Grassland Shrubland<br />

<strong>Habitat</strong> <strong>type</strong><br />

Grass<br />

Woody<br />

% <strong>live</strong> grass <strong>cover</strong><br />

% coyote diet from grass<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

0<br />

0<br />

20<br />

10<br />

Spring Summer Fall<br />

Season<br />

Spring Summer Fall<br />

Season<br />

Grassland<br />

Shrubland<br />

Grassland<br />

Shrubland<br />

Figure 2. Inter-habitat and seasonal differences in <strong>live</strong> <strong>cover</strong> and coyote diet. A)<br />

There are significant differences between percent <strong>live</strong> grass and woody plant <strong>cover</strong> <strong>type</strong>s<br />

and there is a significant habitat*<strong>cover</strong> interaction (Table 1). Please note that statistical<br />

analyses were per<strong>fo</strong>rmed on arcsine trans<strong>fo</strong>rmed values and untrans<strong>fo</strong>rmed percentages<br />

are shown here. B) Percent <strong>live</strong> grass <strong>cover</strong> is significantly different between habitats and<br />

between the spring (a) and the summer and fall (b; Table 2). C) There is no significant<br />

difference in percent coyote diet obtained indirectly from C3 woody plants or C4 grasses<br />

between habitat <strong>type</strong>s (see text). D) Percent of the coyote diet from C4 grasses does not<br />

vary between habitats or among seasons, but there is a significant habitat*season effect at<br />

α = 0.1 (Table 3). For all graphs, error bars represent 95% confidence intervals.<br />

a<br />

b b


106<br />

Ca<strong>rb</strong>on isotope signatures of bone pieces taken from 62 scat samples, each<br />

obtained from a different individual, were used to calculate values <strong>fo</strong>r percent coyote diet<br />

coming indirectly from C3 woody plants in grassland versus shrubland areas (Figure 2C,<br />

Appendix 2). Bone pieces were taken from scat samples collected in 2009 <strong>fo</strong>r 58 of the<br />

62 individuals. There was no significant difference in percent coyote diet from C3 woody<br />

plants between grassland and shrubland habitats (t = 0.94, d.f. = 60, p = 0.35; Figure 2C).<br />

However, average percent coyote diet from C3 woody plants was 32% or higher in both<br />

habitat <strong>type</strong>s, while woody plants, on average, represented 30% of total <strong>live</strong> <strong>cover</strong> in<br />

shrubland habitats and less than 10% in grassland areas (Figure 2A and C). The average<br />

(± 1 s.d.) difference between δ 13 C values of replicate samples of bone pieces taken from a<br />

single scat was small (0 ± 0.3 o /oo).<br />

Ca<strong>rb</strong>on isotope signatures of hair taken from 70 scat samples obtained from 30<br />

individuals were used to calculate values <strong>fo</strong>r percent coyote diet coming indirectly from<br />

C4 grasses (Figure 2D, Appendix 2). There were no significant differences between<br />

habitat <strong>type</strong>s or among seasons in percent coyote diet from C4 grasses, but the<br />

habitat*season interaction was significant at α = 0.1 (Table 3).<br />

Table 3. Repeated measures ANOVA of coyote diet. Results of a repeated measures<br />

ANOVA <strong>fo</strong>r percent of the coyote diet from C4 grasses with habitat and season as<br />

independent variables (d.f. = degrees of freedom).<br />

Effect Numerator d.f. Denominator d.f. F p<br />

<strong>Habitat</strong> 1 28 0.08 0.78<br />

Season 2 36 0.06 0.94<br />

<strong>Habitat</strong>*Season 2 36 2.56 0.09<br />

With the exception of diet values <strong>fo</strong>r coyotes sampled in grassland habitat in the summer,<br />

values <strong>fo</strong>r percent coyote diet from C4 grasses <strong>fo</strong>llowed a somewhat similar pattern as


107<br />

percent <strong>live</strong> grass <strong>cover</strong> values both between habitats and among seasons (Figure 2B and<br />

D). Additionally, average percent coyote diet from C4 grasses was 53% or higher in both<br />

habitats and all three seasons (Figure 2D). The average (± 1 s.d.) difference between δ 13 C<br />

values of replicate samples of hair pieces taken from a single scat was small (0 ± 0.5 o /oo).<br />

Discussion<br />

Scat surveys and genetic analyses<br />

The number of individuals sampled in this study is within the range of sample<br />

sizes (30-122) obtained in other studies of canid species that used noninvasive genetic<br />

sampling techniques to identify individuals (e.g., Kohn et al. 1999, Smith et al. 2006,<br />

Stenglein et al. 2010b). There are several studies of canid populations and canid ecology<br />

that utilized invasive techniques (e.g., radio-telemetry) and were based on much smaller<br />

numbers of individuals (n = 12 to 65; Koehler and Hornocker 1991, Windberg et al.<br />

1997, Kitchen et al. 2000, Kamler et al. 2005, Young et al. 2006, Boisjoly et al. 2010).<br />

The per locus success rates obtained in this study are as high, and error rates as low, as<br />

those <strong>fo</strong>und in multiple, recent studies that have utilized microsatellite primers to amplify<br />

DNA extracted from fecal samples (Murphy et al. 2003, De Ba<strong>rb</strong>a and Waits 2010,<br />

Stenglein et al. 2010a, Stenglein et al. 2010b). The majority of the population structure<br />

identified by the programs Structure 2.3.1 and BAPS 5 was explained by the presence of<br />

closely related individuals. This conclusion is supported by the observation that Structure<br />

can overestimate the number of populations or genetic groups when related individuals<br />

are included in the analysis (Anderson and Dunham 2008). It is thus likely that all 81<br />

individuals sampled in this study are part of a panmictic coyote population with no<br />

physical barriers to gene flow.


Vegetation surveys and stable ca<strong>rb</strong>on isotope analyses<br />

108<br />

Percent <strong>live</strong> grass <strong>cover</strong> is higher than percent <strong>live</strong> woody plant <strong>cover</strong> in both<br />

grassland and shrubland habitats. This reflects field-based observations of grass patches<br />

both in the inter-shrub spaces and even beneath shrubs in the woody plant-encroached<br />

shrubland habitat. The difference between the percentages <strong>fo</strong>r these two <strong>cover</strong> <strong>type</strong>s is<br />

larger in grassland than in shrubland areas. This indicates that grass is the dominant life<br />

<strong>fo</strong>rm in grassland areas but that woody plants and grasses are closer to being co-dominant<br />

in woody plant-encroached shrubland habitats. Additionally, though there is no<br />

significant difference in percent <strong>live</strong> <strong>cover</strong> between habitat <strong>type</strong>s, percent <strong>live</strong> woody<br />

plant <strong>cover</strong> is significantly higher in the shrubland habitat. These inter-habitat differences<br />

are to be expected and indicate that the landscape surrounding the road-based,<br />

“grassland” scat transects has different habitat characteristics from that surrounding the<br />

“shrubland” transects. Seasonal variation in percent <strong>live</strong> grass <strong>cover</strong> <strong>fo</strong>llows an expected<br />

trend in both habitat <strong>type</strong>s, with percent <strong>cover</strong> values increasing significantly between the<br />

spring and the summer and remaining high in the fall. This matches other observations at<br />

the <strong>Sevilleta</strong> NWR of increased net primary production of C4 plants between the spring<br />

and summer (Warne et al. 2010b).<br />

Contrary to expectations, there was no significant shift in the base of the coyote<br />

<strong>fo</strong>od chain between grassland and woody plant-encroached shrubland habitats and thus in<br />

response to woody plant encroachment. More specifically, there was not a statistically<br />

significant increase in percent coyote diet from C3 woody plants between grassland and<br />

shrubland areas and thus the null hypothesis of no difference in the base of the <strong>fo</strong>od chain<br />

between habitat <strong>type</strong>s could not be rejected. However, C3 plants accounted <strong>fo</strong>r a


109<br />

surprisingly large percentage of coyote diet in grassland areas (32% on average) when the<br />

availability of woody plants, quantified in terms of percent of total <strong>live</strong> <strong>cover</strong> (8% on<br />

average in grassland habitat), is considered. This highlights the importance of C3 plants<br />

as a <strong>fo</strong>od resource <strong>fo</strong>r coyote prey that use grassland areas in this ecosystem. Other<br />

researchers have <strong>fo</strong>und that C3 plants are an important source of energy <strong>fo</strong>r consumers in<br />

this arid ecosystem (Warne et al. 2010b) and indicated that C3 plants tend to be a higher<br />

quality <strong>fo</strong>od resource than C4 plants (Caswell et al. 1973, Ehleringer et al. 2002,<br />

Ba<strong>rb</strong>ehenn et al. 2004a). In particular, C3 plants tend to have higher protein, nonstructural<br />

ca<strong>rb</strong>ohydrate, and water content while having lower ca<strong>rb</strong>on to nitrogen ratios, less fiber,<br />

and being overall less tough than C4 plants (Ehleringer et al. 2002, Ba<strong>rb</strong>ehenn et al.<br />

2004a). It is important to note that woody plants are not the only C3 plants present at the<br />

<strong>Sevilleta</strong> NWR. Several small mammal species <strong>fo</strong>und at the study site eat the seeds of<br />

various <strong>fo</strong><strong>rb</strong>s (Hope and Parmenter 2007). Many <strong>fo</strong><strong>rb</strong>s have ca<strong>rb</strong>on isotopic signatures<br />

that are very close in value to those of woody plants (<strong>fo</strong><strong>rb</strong>s = -26.1 o /oo ± 1.1; woody<br />

plants = 24.9 o /oo ± 0.7; values shown as average ± 1 s.d.; Appendix 2) and average<br />

percent <strong>live</strong> <strong>fo</strong><strong>rb</strong> <strong>cover</strong> (4%) is comparable to average percent <strong>live</strong> woody plant <strong>cover</strong><br />

(3%) in grassland areas (Figure 2A and Appendix 3).<br />

My expectation that there would be a shift in the base of the <strong>fo</strong>od chain from C3<br />

woody plants to C4 grasses from spring to summer and fall in both grassland and<br />

shrubland habitats was not met as there was no significant seasonal variation in percent<br />

coyote diet from C4 grasses. Furthermore, percent coyote diet from C4 grasses <strong>fo</strong>llowed<br />

different trends between the grassland and shrubland habitats. However, grass did<br />

account <strong>fo</strong>r a fairly high percentage of coyote diet (53-71%) and total <strong>live</strong> <strong>cover</strong> (52-94


110<br />

%) in both habitats and all seasons. Additionally, percent coyote diet from C4 grasses did<br />

<strong>fo</strong>llow a similar trend as percent <strong>live</strong> grass <strong>cover</strong> in the shrubland habitat. These results<br />

indicate that seasonal changes in habitat characteristics do not appear to lead to<br />

significant changes in coyote feeding ecology, specifically in the base of the coyote <strong>fo</strong>od<br />

chain. However, grass is a consistent <strong>fo</strong>od resource <strong>fo</strong>r coyote prey regardless of habitat<br />

<strong>type</strong> and season, though its importance is not always proportional to its availability in the<br />

ecosystem. These results do not mirror the definitive increase in C4-derived <strong>fo</strong>od resource<br />

use from spring to summer documented by other researchers <strong>fo</strong>r both primary<br />

(arthropods) and secondary (lizards) consumers at the <strong>Sevilleta</strong> NWR (Warne et al.<br />

2010b). However, Warne et al. (2010b) did observe that, in a year of high C4 plant<br />

production, percent of C4-derived ca<strong>rb</strong>on in the blood of secondary consumers (40%) was<br />

low relative to C4 plant availability (87% of total annual net primary production).<br />

There are a variety of factors that may account <strong>fo</strong>r the observed lack of significant<br />

variation in coyote feeding ecology, specifically the base of the coyote <strong>fo</strong>od chain,<br />

between habitat <strong>type</strong>s and the high percentage of C3 plants in the coyote diet in grassland<br />

habitats. The habitat at the <strong>Sevilleta</strong> NWR is quite heterogeneous, especially in the<br />

southern and western parts of the refuge (Figure 1) and there are many places where<br />

patches of grassland and shrubland habitat interdigitate. Even though individuals that<br />

moved between road-based scat transects located in different habitat <strong>type</strong>s were excluded<br />

from the statistical analyses per<strong>fo</strong>rmed in this study, coyotes are wide ranging animals<br />

(e.g., Roy and Dorrance 1985, Windberg et al. 1997, Kamler et al. 2005, Boisjoly et al.<br />

2010) and it is likely that many of them passed between patches of different habitat <strong>type</strong>s<br />

while <strong>fo</strong>raging. This could account <strong>fo</strong>r the high percentage of C3 plants in the diets of


111<br />

coyotes sampled in grassland areas. Furthermore, many of the bone pieces used to assess<br />

percent coyote diet in different habitat <strong>type</strong>s were obtained from scat samples collected in<br />

the spring season. Though bone is thought to integrate in<strong>fo</strong>rmation on diet over the life of<br />

an animal (Ambrose and DeNiro 1986, Hobson and Clark 1992), it is possible that the<br />

coyotes preyed on juvenile rodents or other small mammals that had only been a<strong>live</strong> <strong>fo</strong>r a<br />

few weeks or months. Various small mammals <strong>fo</strong>und at the <strong>Sevilleta</strong> NWR, including<br />

Dipodomys merriami, Perognathus flavus, and Neotoma albigula, can breed during the<br />

late winter or early spring months (Whitaker 1996, Friggens 2008). In that case, the<br />

signature of the bones would represent resource use over a much shorter time period,<br />

especially the period during which the bones were growing (Hobson and Clark 1992).<br />

Given the greater availability of C3 relative to C4 plants in the spring (Warne et al.<br />

2010b), this use of a large number of scat samples collected during the spring could<br />

partially account <strong>fo</strong>r the high percentage of C3 plants in the coyote diet observed in the<br />

grassland habitat. Additionally, there is a good deal of intra-individual variation in coyote<br />

diet within a single season (Appendix 1). The analysis of inter-habitat differences in<br />

coyote diet presented here is based on the ca<strong>rb</strong>on isotope signatures of bone pieces taken<br />

from one scat sample collected from each individual. It is possible that inclusion of more<br />

samples per individual would provide different results that better represent the total<br />

variation in coyote diet.<br />

There are several possible explanations <strong>fo</strong>r the lack of significant seasonal<br />

variation in the base of the coyote <strong>fo</strong>od chain and the surprising, marginally significant<br />

interaction between habitat <strong>type</strong> and season <strong>fo</strong>r values of percent coyote diet from C4<br />

grasses observed in this study. It is possible that the analysis of hair pieces obtained from


112<br />

one scat sample from each individual in each season does not account <strong>fo</strong>r all of the<br />

variation in coyote diet (Appendix 1). It is also possible that the hair of the rodents and<br />

other mammals consumed by coyotes at the <strong>Sevilleta</strong> NWR does not turn over rapidly<br />

enough <strong>fo</strong>r inter-seasonal variation in diet to be detected. The half-life of ca<strong>rb</strong>on in hair<br />

ranges from 47.5 days (ge<strong>rb</strong>ils, Tieszen et al. 1983) up to 136 days (horses, Ayliffe et al.<br />

2004), and can be as high as 537 days (bats, Voigt et al. 2003) <strong>fo</strong>r some species. If the<br />

half-life of ca<strong>rb</strong>on in the hair of mammals at the <strong>Sevilleta</strong> NWR is close to 47.5 days,<br />

then any diet shifts should be detectable between the spring and the fall (April to<br />

October) and possibly from the spring to the summer (April to July) and the summer to<br />

the fall (July to October). However, if the half life is closer to 136 days, or longer, diet<br />

shifts from the spring to the fall might be detectable but the difference would probably be<br />

small as only about half of the ca<strong>rb</strong>on in the hair would have turned over between<br />

sampling events. The study that detected a shift in the use of C4-derived ca<strong>rb</strong>on by<br />

secondary consumers between spring and summer at the <strong>Sevilleta</strong> NWR utilized samples<br />

of blood plasma (Warne et al. 2010b) and blood plasma has a faster turnover time (25-44<br />

days, Warne et al. 2010a) than hair. The moderately significant habitat <strong>type</strong> by season<br />

interaction is driven by opposing trends in percent coyote diet from C4 grasses between<br />

the spring and the summer in the two different habitat <strong>type</strong>s. In grassland areas, percent<br />

coyote diet from C4 grasses declines from the spring to the summer while in shrubland<br />

areas, it increases. The increase from spring to summer in shrubland areas matches my<br />

expectations. The decline in percent C4 grasses in coyote diet from spring to summer in<br />

grassland areas may be due to a shift in diet from C4 grass to C3 <strong>fo</strong><strong>rb</strong> seeds or other<br />

tissues on the part of the coyote prey. For some small mammal species at the <strong>Sevilleta</strong>


113<br />

NWR, the consumption of <strong>fo</strong><strong>rb</strong> seeds increases between the spring and the summer in<br />

grassland areas (Hope and Parmenter 2007). There is a similar decline in percent C4 grass<br />

in coyote diet from summer to fall in shrubland areas. This could also be driven by an<br />

increase in the consumption of C3 <strong>fo</strong><strong>rb</strong> seeds by coyote prey as there is an increase in the<br />

availability of <strong>fo</strong><strong>rb</strong>s from the summer to the fall in shrubland areas (Appendix 3).<br />

Management implications and future work<br />

The impact that continued woody plant encroachment may have on the ecology<br />

and fitness of top predators, such as the coyote, and their prey is uncertain. There is no<br />

significant difference in the base of the coyote <strong>fo</strong>od chain between habitats and the results<br />

presented here indicate that woody plant encroachment has had limited impact on coyote<br />

feeding ecology. Given the apparent importance of C3 plants as a <strong>fo</strong>od resource in<br />

grassland areas, woody plant encroachment may be beneficial <strong>fo</strong>r coyote prey. However,<br />

these observations are tempered by several observations. First of all, grass <strong>cover</strong> is<br />

relatively high in the woody plant-encroached habitats considered in this study and there<br />

are woody plants present in the grassland areas, as well as patches of shrubland within<br />

grassland habitats and vice versa. It would be necessary to compare grassland and woody<br />

plant-encroached habitats with a greater difference in grass availability and to survey<br />

predator populations in grasslands that are either devoid of woody vegetation or are<br />

further removed from shrubland areas to be certain that this habitat shift has no impact on<br />

predator feeding ecology. Second, the use of C4 grasses as a <strong>fo</strong>od resource is both high<br />

and consistent between habitats and among seasons and it is not possible to differentiate<br />

the use of C3 woody plants and <strong>fo</strong><strong>rb</strong>s with the techniques used in this study. Third, the<br />

coyote is a generalist species capable of using a variety of <strong>fo</strong>od resources. An evaluation


114<br />

of the feeding ecology of a suite of predators, including more specialized feeders, would<br />

provide more in<strong>fo</strong>rmation regarding the importance of C3 versus C4 plants in the local<br />

<strong>fo</strong>od web and the probable impact of continued woody plant encroachment on local<br />

predator populations.<br />

Another factor to consider is that only one facet of coyote feeding ecology was<br />

evaluated in this study. The use of stable ca<strong>rb</strong>on isotopes allowed <strong>fo</strong>r an assessment of<br />

the base of the coyote <strong>fo</strong>od chain and differentiation between native C4 grasses and<br />

encroaching C3 woody plants. However, this approach did not account <strong>fo</strong>r other aspects<br />

of coyote feeding ecology that may have been affected by woody plant encroachment. In<br />

particular, it did not account <strong>fo</strong>r habitat-related shifts in prey availability or in the use of<br />

different prey species by the coyote. Surveys conducted in spring and fall of 2008 in<br />

grassland and shrubland habitats at the <strong>Sevilleta</strong> NWR indicate that there are inter-habitat<br />

differences in the species composition of the small mammal community (Friggens 2008,<br />

Appendix 4) and thus in the coyote’s potential prey base. The 4 species <strong>fo</strong>und to be most<br />

abundant in these surveys were: Dipodomys merriami, Dipodomys ordii, Dipodomys<br />

spectabilis, and Perognathus flavus (Friggens 2008, Appendix 4). The typical diet of<br />

these <strong>fo</strong>ur species is quite varied and includes seeds, green vegetation, and insects. The<br />

seeds that these small mammals eat come from C3 (<strong>fo</strong><strong>rb</strong>s, shrubs), C4 (grasses), and CAM<br />

(cacti) plants (Hope and Parmenter 2007, Appendices 2 and 4). It is important to note that<br />

the average (± 1 s.d.) ca<strong>rb</strong>on isotope signature of CAM plants (-12.0 ± 0.1 o /oo) was<br />

indistinguishable from that of C4 grasses (-14.8 o /oo ± 0.3; Appendix 2) in this study.<br />

These observations lend support to my finding that coyote prey consume both C3 and C4


115<br />

plants and rein<strong>fo</strong>rce my previous statements regarding the likelihood that coyote prey are<br />

using both C3 <strong>fo</strong><strong>rb</strong>s and C3 woody plants.<br />

In<strong>fo</strong>rmation on the body condition and fitness of both primary and secondary<br />

consumers with plants of different photosynthetic pathways and functional <strong>type</strong>s at the<br />

base of their <strong>fo</strong>od chain (C4 versus C3 and C3 woody plant versus C3 <strong>fo</strong><strong>rb</strong>) would help to<br />

clarify the impact that the shift in plant dominance from C4 grasses to C3 woody plants<br />

associated with woody plant encroachment has on local mammalian predator populations.<br />

Several studies have assessed the use of C3- versus C4-derived <strong>fo</strong>od resources by primary<br />

and secondary consumers (e.g., Fry et al. 1978, Ambrose and DeNiro 1986, Magnusson<br />

et al. 1999, Smith et al. 2002, Codron et al. 2007, Warne et al. 2010b) and considered the<br />

nutritional quality of the diets of he<strong>rb</strong>ivores that eat C3 versus C4 plants (Codron et al.<br />

2005). In contrast and to my knowledge, very little is known regarding the effect that<br />

differences in the quality of these two <strong>fo</strong>od resources can have on the growth and<br />

reproductive rates of mammalian predators (but see Ba<strong>rb</strong>ehenn et al. 2004b <strong>fo</strong>r<br />

in<strong>fo</strong>rmation on insects). In addition, studies of the impact of woody plant encroachment<br />

on the availability and nutritional quality of C3 <strong>fo</strong><strong>rb</strong>s (e.g., Zarovali et al. 2007, Báez and<br />

Collins 2008), when combined with extant in<strong>fo</strong>rmation on the <strong>fo</strong>raging patterns of<br />

various primary consumers (e.g., Bradley and Mauer 1971, Alcoze and Zimmerman<br />

1973, Flake 1973, Soholt 1973, Dial 1988, Hope and Parmenter 2007, Appendix 4),<br />

would further elucidate the likely impact of woody plant encroachment on key prey<br />

species and thus on mammalian predator populations.


Conclusion<br />

116<br />

Overall, long term (woody plant encroachment) and seasonal changes in habitat<br />

characteristics do not appear to have strong bottom-up effects on the coyote population at<br />

the <strong>Sevilleta</strong> NWR. However, C3 plants, which include both <strong>fo</strong><strong>rb</strong>s and woody plants,<br />

appear to be an important <strong>fo</strong>od resource in grassland areas in this arid environment. This<br />

is likely the result of these plants being of higher nutritional quality <strong>fo</strong>r coyote prey than<br />

C4 grasses. Additionally, the base of the coyote <strong>fo</strong>od chain shows different patterns of<br />

seasonal variation between grassland and shrubland habitat <strong>type</strong>s. This difference is<br />

likely driven by the summer use of C3 <strong>fo</strong><strong>rb</strong>s by coyote prey in grassland areas. Woody<br />

plant encroachment does not appear to have had a significant impact on coyote feeding<br />

ecology, specifically on the base of the coyote <strong>fo</strong>od chain, and may even be beneficial<br />

given the importance of C3 <strong>fo</strong>od resources in this ecosystem. However, more in<strong>fo</strong>rmation<br />

regarding the effects of woody plant encroachment on different facets of coyote feeding<br />

ecology, and the effects of different <strong>fo</strong>od resources (C4 versus C3 <strong>fo</strong><strong>rb</strong> versus C3 woody<br />

plant) on the fitness and ecology of both generalist and more specialized primary and<br />

secondary consumers, is needed.


References<br />

117<br />

Alcoze, T.M. and E.G. Zimmerman. 1973. Food habits and dietary overlap of two<br />

heteromyid rodents from the mesquite plains of texas. Journal of Mammalogy. 54(4):<br />

900-908.<br />

Ambrose, S.H. and M.J. DeNiro. 1986. The isotopic ecology of East African mammals.<br />

Oecologia. 69: 395-406.<br />

Andelt, W.F., J.G. Kie, F.F. Knowlton, and K. Cardwell. 1987. Variation in coyote diets<br />

associated with season and successional changes in vegetation. Journal of Wildlife<br />

Management. 51(2): 273-277.<br />

Anderson, E.C. and K.K. Dunham. 2008. The influence of family groups on inferences<br />

made with the program Structure. Molecular Ecology Resources. 8: 1219-1229.<br />

Archer, S. 1989. Have southern Texas savannas been converted to woodlands in recent<br />

history? The American Naturalist. 134(4): 545-561.<br />

Archer, S. 1995. Tree-grass dynamics in a Prosopis-thornscrub savanna parkland:<br />

Reconstructing the past and predicting the future. Ecoscience. 2(1): 83-99.<br />

Ayliffe, L.K., T.E. Cerling, T. Robinson, A.G. West, M. Sponheimer, B.H. Passey, J.<br />

Hammer, B. Roeder, M.D. Dearing, and J.R. Ehleringer. 2004. Turnover of ca<strong>rb</strong>on<br />

isotopes in tail hair and breath CO2 of horses fed an isotopically varied diet. Oecologia.<br />

139: 11-22.<br />

Báez, S. and S.L. Collins. 2008. Shrub invasion decreases diversity and alters community<br />

stability in northern Chihuahuan desert plant communities. PLoS ONE. 3(6): e2332.<br />

doi:10.1371/journal.pone.0002332.<br />

Ba<strong>rb</strong>ehenn, R.V., Z. Chen, D.N. Karowe, and A. Spickard. 2004a. C3 grasses have higher<br />

nutritional quality than C4 grasses under ambient and elevated atmospheric CO2. Global<br />

Change Biology. 10: 1565-1575.<br />

Ba<strong>rb</strong>ehenn, R.V., D.N. Karowe, and A. Spickard. 2004b. Effects of elevated atmospheric<br />

CO2 on the nutritional ecology of C3 and C4 grass-feeding caterpillars. Oecologia. 140:<br />

86-95.<br />

Bender, M.M. 1971. Variations in the 13 C/ 12 C ratios of plants in relation to the pathway<br />

of photosynthetic ca<strong>rb</strong>on dioxide fixation. Phytochemistry. 10:1239-1244.<br />

Bester, B. 1996. Bush encroachment: a thorny problem. Namibia Environ. 1: 175-177.


118<br />

Blaum, N., E. Rossmanith, A. Popp, and F. Jeltsch. 2007a. Shrub encroachment affects<br />

mammalian carnivore abundance and species richness in semiarid rangelands. Acta<br />

Oecologica. 31: 86-92.<br />

Blaum, N., E. Rossmanith, M. Schwager, and F. Jeltsch. 2007b. Responses of<br />

mammalian carnivores to land use in arid savanna rangelands. Basic and Applied<br />

Ecology. 8: 552-564.<br />

Boisjoly, D., J.P. Ouellet, and R. Courtois. 2010. Coyote habitat selection and<br />

management implications <strong>fo</strong>r the Gaspesie caribou. Journal of Wildlife Management.<br />

74(1): 3-11.<br />

Bond, W.J. and G.F. Midgley. 2000. A proposed CO2-controlled mechanism of woody<br />

plant invasion in grasslands and savannas. Global Change Biology. 6: 865-869.<br />

Boutton, T.W., S.R. Archer, A.J. Midwood, S.F. Zitzer, and R. Bol. 1998. δ 13 C values of<br />

soil organic ca<strong>rb</strong>on and their uses in documenting vegetation change in a subtropical<br />

savanna ecosystem. Geoderma. 82: 5-41.<br />

Bradley, W.G. and R.A. Mauer. 1971. Reproduction and <strong>fo</strong>od habits of Merriam’s<br />

kangaroo rat, Dipodomys merriami. Journal of Mammalogy. 52(3): 497-507.<br />

Brand, C.J., L.B. Keith, and C.A. Fischer. 1976. Lynx responses to changing snowshoe<br />

hare densities in central Alberta. Journal of Wildlife Management. 40(3): 416-428.<br />

Caswell, H., F. Reed, S.N. Stephenson, and P.A. Werner. 1973. Photosynthetic pathways<br />

and selective he<strong>rb</strong>ivory: A hypothesis. The American Naturalist. 107(956): 465-480.<br />

Codron, D., J. Codron, J.A. Lee-Thorp, M. Sponheimer, and D. de Ruiter. 2005. Animal<br />

diets in the Wate<strong>rb</strong>erg based on stable isotopic composition of faeces. South African<br />

Journal of Wildlife Research. 35(1): 43-52.<br />

Codron, D., J. Codron, J.A. Lee-Thorp, M. Sponheimer, D. de Ruiter, J. Sealy, R. Grant,<br />

and N. Fourie. 2007. Diets of savanna ungulates from stable ca<strong>rb</strong>on isotope composition<br />

of faeces. Journal of Zoology. 273: 21-29.<br />

Corander, J., P. Marttinen, J. Sirén, and J. Tang. 2008a. Enhanced Bayesian modelling in<br />

BAPS software <strong>fo</strong>r learning genetic structures of populations. BMC Bioin<strong>fo</strong>rmatics.<br />

9:539.doi:10.1186/1471-2105-9-539.<br />

Corander, J., J. Sirén, and E. Arjas. 2008b. Bayesian Spatial Modelling of Genetic<br />

Population Structure. Computational Statistics. 23: 111-129.<br />

Craw<strong>fo</strong>rd, K., R.A. Mcdonald, and S. Bearhop. 2008. Applications of stable isotope<br />

techniques to the ecology of mammals. Mammal Rev. 38(1): 87-107.


Crooks, K.R. and M.E. Soule. 1999. Mesopredator release and avifaunal extinctions in a<br />

fragmented system. Nature. 400: 563-566.<br />

De Ba<strong>rb</strong>a, M. and L.P. Waits. 2010. Multiplex pre-amplification <strong>fo</strong>r noninvasive genetic<br />

sampling: is the extra ef<strong>fo</strong>rt worth it? Molecular Ecology Resources. 10: 659-665.<br />

DeNiro, M.J. and S. Epstein. 1978. Influence of diet on the distribution of ca<strong>rb</strong>on<br />

isotopes in animals. Geochimica et Cosmochimica Acta. 42: 495-506.<br />

Dial, K.P. 1988. Three sympatric species of Neotoma: dietary specialization and<br />

coexistence. Oecologia. 76: 531-537.<br />

119<br />

D’Odorico, P. and A. Porporato. 2006. Ecohydrology of arid and semiarid ecosystems:<br />

An introduction. In P.D’Odroico and A. Porporato (eds). Dryland Ecohydrology. The<br />

Netherlands: Springer. Pps. 1-10.<br />

Ehleringer, J.R., T.F. Cerling, and M.D. Dearing. 2002. Atmospheric CO2 as a global<br />

change driver influencing plant-animal interactions. Integrative and Comparative<br />

Biology. 42: 424-430.<br />

Ernest, S.K.M., J.H. Brown, and R.R. Parmenter. 2000. Rodents, plants, and<br />

precipitation: spatial and temporal dynamics of consumers and resources. Oikos. 88: 470-<br />

482.<br />

Estes, J.A. and D.O. Duggins. 1995. Sea otters and kelp <strong>fo</strong>rests in Alaska: Generality and<br />

variation in a community ecological paradigm. Ecological Monographs. 65 (1): 75-100.<br />

Farquhar, G.D. 1983. On the nature of ca<strong>rb</strong>on isotope discrimination in C4 species.<br />

Australian Journal of Plant Physiology. 10: 205-226.<br />

Faure, G. and T.M. Mensing. 2005. Isotopes: Principles and Applications. Hoboken,<br />

New Jersey: John Wiley & Sons. Pps. 693-752.<br />

Flake, L.D. 1973. Food habits of <strong>fo</strong>ur species of rodents on a short-grass prairie in<br />

Colorado. Journal of Mammalogy. 54(3): 636-647.<br />

Francisco, L.V., A.A. Langston, C.S. Mellersh, C.L. Neal, and E.A. Ostrander. 1996. A<br />

class of highly polymorphic tetranucleotide repeats <strong>fo</strong>r canine genetic mapping.<br />

Mammalian Genome. 7: 359-362.<br />

Frantzen, M.A.J., J.B. Silk, J.W.H. Ferguson, R.K. Wayne, and M.H. Kohn. 1998.<br />

Empirical evaluation of preservation methods <strong>fo</strong>r fecal DNA. Molecular Ecology. 7:<br />

1423–1428.


120<br />

Friggens, M. 2008. Small Mammal Mark-Recapture Population Dynamics at SEV <strong>LTER</strong><br />

Core Research Sites. Albuquerque, NM: <strong>Sevilleta</strong> Long Term Ecological Research<br />

Database: SEV008. (20<br />

November 2010).<br />

Fry, B., A. Joern, and P.L. Parker. 1978. Grasshopper <strong>fo</strong>od web analysis: Use of ca<strong>rb</strong>on<br />

isotope ratios to examine feeding relationships among terrestrial he<strong>rb</strong>ivores. Ecology.<br />

59(3): 498-506.<br />

Gill, R.A. and I.C. Burke. 1999. Ecosystem consequences of plant life <strong>fo</strong>rm changes at<br />

three sites in the semiarid United States. Oecologia. 121: 551-563.<br />

Hamlin, K.L., S.J. Riley, D. Pyrah, A.R. Dood, and R.J. Mackie. 1984. Relationships<br />

among mule deer fawn mortality, coyotes, and alternate prey species during summer.<br />

Journal of Wildlife Management. 48(2): 489-499.<br />

Hernández, L., R.R. Parmenter, J.W. Dewitt, D.C. Light<strong>fo</strong>ot, and J.W. Laundré. 2002.<br />

Coyote diets in the Chihuahuan Desert, more evidence <strong>fo</strong>r optimal <strong>fo</strong>raging. Journal of<br />

Arid Environments. 51: 613-624.<br />

Hernández, L., A.G. Romero, J.W. Laundré, D. Light<strong>fo</strong>ot, E. Aragón, and J. López-<br />

Portillo. 2005. Changes in rodent community structure in the Chihuahuan Desert México:<br />

comparisons between two habitats. Journal of Arid Environments. 60: 239-257.<br />

Hidalgo-Mihart, M.G., L. Cantú-Salazar, C.A. López-González, E. Martínez-Meyer, and<br />

A. González-Romero. 2001. Coyote (Canis latrans) <strong>fo</strong>od habits in a tropical deciduous<br />

<strong>fo</strong>rest of western Mexico. The American Midland Naturalist. 146: 210-216.<br />

Hobson, K.A. and R.G. Clark. 1992. Assessing avian diets using stable isotopes I:<br />

Turnover of 13 C in tissues. The Condor. 94: 181-188.<br />

Hone, J., C. Krebs, M. O’Donoghue, and S. Boutin. 2007. Evaluation of predator<br />

numerical responses. Wildlife Research. 34: 335-341.<br />

Hope, A.G. and R.R. Parmenter. 2007. Food habits of rodents inhabiting arid and semiarid<br />

ecosystems of central New Mexico. Special Publication of the Museum of<br />

Southwestern Biology. 9: 1-75.<br />

IUCN 2010. IUCN Red List of Threatened Species. Version 2010.2.<br />

. (16 August 2010).<br />

Jaksic, F.M., S.I. Silva, P.L. Meserve, and J.R. Gutiérrez. 1997. A long-term study of<br />

vertebrate predator responses to an El Niño (ENSO) distu<strong>rb</strong>ance in western South<br />

America. Oikos. 78(2): 341-354.


121<br />

Kamler, J.F., W.B. Ballard, and P.R. Lemons. 2005. Home range and habitat use of<br />

coyotes in an area of native prairie, farmland, and CRP fields. American Midland<br />

Naturalist. 153: 396-404.<br />

King, C.M. 1983. The relationship between beech (Nothofagus sp.) seedfall and<br />

populations of mice (Mus musculus) and the demographic and dietary responses of stoats<br />

(Mustela erminea), in three New Zealand <strong>fo</strong>rests. Journal of Animal Ecology. 52(1): 141-<br />

166.<br />

Kitchen, A.M., E.M. Gese, and E.R. Schauster. 2000. Long-term spatial stability of<br />

coyote (Canis latrans) home ranges in southeastern Colorado. Canadian Journal of<br />

Zoology. 78: 458-464.<br />

Köchy, M. and S.D. Wilson. 2001. Nitrogen deposition and <strong>fo</strong>rest expansion in the<br />

northern Great Plains. Journal of Ecology. 89: 807-817.<br />

Koehler, G.M. and M.G. Hornocker. 1991. Seasonal resource use among mountain lions,<br />

bobcats, and coyotes. Journal of Mammalogy. 72(2): 391-396.<br />

Kohn, M.H., E.C. York, D.A. Kamradt, G. Haught, R.M. Sauvajot, and R.K. Wayne.<br />

1999. Estimating population size by genotyping faeces. Proceedings of the Royal Society,<br />

London B. 266: 657-663.<br />

Magnusson, W.E., M. Carmozina de Araujo, R. Cintra, A.P. Lima, L.A. Martinelli, T.M.<br />

Sanaiotti, H.L. Vasconcelos, and R.L. Victoria. 1999. Contributions of C3 and C4 plants<br />

to higher trophic levels in an Amazonian savanna. Oecologia. 119: 91-96.<br />

Marker, L. 2002. Aspects of Namibian cheetah (Acinonyx jubatus) biology, ecology and<br />

conservation strategies. PhD. Thesis, Department of Zoology, University of Ox<strong>fo</strong>rd. Pps.<br />

199-242.<br />

Marshall, J.D., J.R. Brooks, and K. Lajtha. 2007. Sources of variation in the stable<br />

isotopic composition of plants. In R. Michener and K. Lajtha (eds). Stable isotopes in<br />

ecology and environmental science. Ox<strong>fo</strong>rd: Blackwell Publishing. Pps. 22-60.<br />

McKinley, D.C. and J.M. Blair. 2008. Woody plant encroachment by Juniperus<br />

virginiana in a mesic native grassland promotes rapid ca<strong>rb</strong>on and nitrogen accrual.<br />

Ecosystems. 11: 454-468.<br />

McLaren, B.E. and R.O. Peterson. 1994. Wolves, moose, and tree rings on Isle Royale.<br />

Science, New Series. 266(5190):1555-1558.<br />

Miller, C.R., P. Joyce, and L.P. Waits. 2002. Assessing allelic dropout and geno<strong>type</strong><br />

reliability using maximum likelihood. Genetics. 160: 357-366.


122<br />

Mills, M.G.L., L.S. Broomhall, and J.T. du Toit. 2004. Cheetah Acinonyx jubatus feeding<br />

ecology in the Kruger National Park and a comparison across African savanna habitats: is<br />

the cheetah only a successful hunter on open grassland plains? Wildlife Biology. 10(3):<br />

177-186.<br />

Moore, D. 2009. ClimDB Monthly Data. Albuquerque, NM: <strong>Sevilleta</strong> Long Term<br />

Ecological Research Site. <br />

(8 February 2010).<br />

Muldavin, E.H., D.I. Moore, S.L. Collins, K.R. Wetherill, and D.C. Light<strong>fo</strong>ot. 2008.<br />

Aboveground net primary production dynamics in a northern Chihuahuan Desert<br />

Ecosystem. Oecologia. 155: 123-132.<br />

Muntifering, J.R., A.J. Dickman, L.M. Perlow, T. Hruska, P.G. Ryan, L.L. Marker, and<br />

R.M. Jeo. 2006. Managing the matrix <strong>fo</strong>r large carnivores: a novel approach and<br />

perspective from cheetah (Acinonyx jubatus) habitat suitability modeling. Animal<br />

Conservation. 9: 103-112.<br />

Murphy, M.A., L.P. Waits, and K.C. Kendall. 2000. Quantitative evaluation of fecal<br />

drying methods <strong>fo</strong>r brown bear DNA analysis. Wildlife Society Bulletin. 28: 951–957.<br />

Murphy, M.A., L.P. Waits, and K.C. Kendall. 2003. The influence of diet on faecal DNA<br />

amplification and sex identification in brown bears (Ursus arctos). Molecular Ecology.<br />

12: 2261-2265.<br />

Noy-Meir, I. 1973. Desert Ecosystems: Environment and Producers. Annual Review of<br />

Ecology and Systematics. 4: 25-51.<br />

O’Donoghue, M., S. Boutin, C.J. Krebs, and E.J. Hofer. 1997. Numerical responses of<br />

coyotes and lynx to the snowshoe hare cycle. Oikos. 80: 150-162.<br />

Onorato, D., C. White, P. Zager, and L.P. Waits. 2006. Detection of predator presence at<br />

elk mortality sites using mtDNA analysis of hair and scat samples. Wildlife Society<br />

Bulletin. 34(3): 815-820.<br />

Ostfeld, R.S. and F. Keesing. 2000. Pulsed resources and community dynamics of<br />

consumers in terrestrial ecosystems. Trends in Ecology and Evolution. 15(6): 232-237.<br />

Ostrander, E.A., F.A. Mapa, M. Yee, and J. Rine. 1995. One hundred and one new simple<br />

sequence repeat-based markers <strong>fo</strong>r the canine genome. Mammalian Genome. 6: 192-195.<br />

Ostrander, E.A., G.F. Sprague, and J. Rine. 1993. Identification and characterization of<br />

dinucleotide repeat (CA)n markers <strong>fo</strong>r genetic mapping in dog. Genomics. 16: 207-213.


123<br />

Paine, R.T. 1969. A note on trophic complexity and community stability. The American<br />

Naturalist. 103 (929): 91-93.<br />

Previtali, A.M., M. Lima, P.L. Meserve, D.A. Kelt, and J.R. Gutiérrez. 2009. Population<br />

dynamics of two sympatric rodents in a variable environment: rainfall, resource<br />

availability, and predation. Ecology. 90(7): 1996-2006.<br />

Pritchard, J. K., M. Stephens, and P. Donnelly. 2000. Inference of population structure<br />

using multilocus geno<strong>type</strong> data. Genetics. 155: 945-959.<br />

Ravi, S., P. D’Odorico, L. Wang, C.S. White, G.S. Okin, S.A. Macko, and S.L. Collins.<br />

2009. Post-fire resource redistribution in desert grasslands: A possible negative feedback<br />

on land degradation. Ecosystems. 12: 434-444.<br />

Raymond, M. and F. Rousset. 1995. GENEPOP (version 1.2): population genetics<br />

software <strong>fo</strong>r exact tests and ecumenicism. Journal of Heredity. 86:248-249.<br />

Ripple, W.J. and R.L. Beschta. 2004. Wolves, elk, willows, and trophic cascades in the<br />

upper Gallatin range of southwestern Montana, USA. Forest Ecology and Management.<br />

200: 161-181.<br />

Roques, K.G., T.G. O'Connor, and A.R. Watkinson. 2001. Dynamics of shrub<br />

encroachment in an African savanna: relative influences of fire, he<strong>rb</strong>ivory, rainfall and<br />

density dependence. Journal of Applied Ecology. 38:268-280.<br />

Rousset, F. 2008. Genepop'007: a complete reimplementation of the Genepop software<br />

<strong>fo</strong>r Windows and Linux. Molecular Ecology Resources. 8:103-106.<br />

Roy, L.D. and M.J. Dorrance. 1985. Coyote movements, habitat use, and vulnerability in<br />

central Alberta. Journal of Wildlife Management. 49(2): 307-313.<br />

Sacks, B.N., S.K. Brown, and H.B. Ernest. 2004. Population structure of Cali<strong>fo</strong>rnia<br />

coyotes corresponds to habitat-specific breaks and illuminates species history. Molecular<br />

Ecology. 13: 1265-1275.<br />

Samson, C. and M. Créte. 1997. Summer <strong>fo</strong>od habits and population density of coyotes,<br />

Canis latrans, in boreal <strong>fo</strong>rests of southeastern Québec. The Canadian Field-Naturalist.<br />

111: 227-233.<br />

Schmitz, O.J., P.A. Hamback, and A.P. Beckerman. 2000. Trophic cascades in terrestrial<br />

systems: A review of the effects of carnivore removals on plants. The American<br />

Naturalist. 155(2): 141-153.<br />

Schwinning, S. and O.E. Sala. 2004. Hierarchy of responses to resource pulses in arid and<br />

semi-arid ecosystems. Oecologia. 141: 211-220.


124<br />

Smith, D.A., K. Ralls, A. Hurt, B. Adams, M. Parker, and J.E. Maldonado. 2006.<br />

Assessing reliability of microsatellite geno<strong>type</strong>s from kit <strong>fo</strong>x faecal samples using genetic<br />

and GIS analyses. Molecular Ecology. 15: 387-406.<br />

Smith, K.F., Z.D. Sharp, and J.H. Brown. 2002. Isotopic composition of ca<strong>rb</strong>on and<br />

oxygen in desert fauna: investigations into the effects of diet, physiology, and<br />

seasonality. Journal of Arid Environments. 52: 419-430.<br />

Soholt, L.F. 1973. Consumption of primary production by a population of kangaroo rats<br />

(Dipodomys merriami) in the Mojave Desert. Ecological Monographs. 43(3): 357-376.<br />

Stenglein, J.L., M. De Ba<strong>rb</strong>a, D.E. Ausband, and L.P. Waits. 2010a. Impacts of sampling<br />

location within a faeces on DNA quality in two carnivore species. Molecular Ecology<br />

Resources. 10: 109-114.<br />

Stenglein, J.L., L.P. Waits, D.E. Ausband, P. Zager, and C.M. Mack. 2010b. Efficient,<br />

noninvasive genetic sampling <strong>fo</strong>r monitoring introduced wolves. Journal of Wildlife<br />

Management. 74(5): 1050-1058.<br />

Stenseth, N.C., W. Falck, O.N. Bjornstad, and C.J. Krebs. 1997. Population regulation in<br />

snowshoe hare and Canadian lynx: Asymmetric <strong>fo</strong>od web configurations between hare<br />

and lynx. Proceedings of the National Academy of Sciences, USA. 94: 5147-5152.<br />

Sternberg, L.O., M.J. DeNiro, and H.B. Johnson. 1984. Isotope ratios of cellulose from<br />

plants having different photosynthetic pathways. Plant Physiology. 74: 557-561.<br />

Tieszen, L.L., T.W. Boutton, K.G. Tesdahl, and N.A. Slade. 1983. Fractionation and<br />

turnover of stable ca<strong>rb</strong>on isotopes in animal tissues: Implications <strong>fo</strong>r δ 13 C analysis of<br />

diet. Oecologia. 57: 32-37.<br />

Todd, A.W., L.B. Keith, and C.A. Fischer. 1981. Population ecology of coyotes during a<br />

fluctuation of snowshoe hares. Journal of Wildlife Management. 45(3): 629-640.<br />

Valière, N. 2002. GIMLET: a computer program <strong>fo</strong>r analyzing genetic individual<br />

identification data. Molecular Ecology Notes. 10.1046/j.1471-8278 .2002.00228.x.<br />

Van Auken, O.W. 2000. Shrub invasions of North American semiarid grasslands. Annual<br />

Review of Ecology and Systematics. 31: 197-215.<br />

Voigt, C.C., F. Matt, R. Michener, and T.H. Kunz. 2003. Low turnover rates of ca<strong>rb</strong>on<br />

isotopes in tissues of two nectar-feeding bat species. The Journal of Experimental<br />

Biology. 206: 1419-1427.


125<br />

Wallington, B.P., A.E. McKechnie, N. Owen-Smith, and S. Woodborne. 2007. Stable<br />

ca<strong>rb</strong>on isotope analysis of eland (Taurotragus oryx) diet in the Suike<strong>rb</strong>osrand Nature<br />

Reserve. South African Journal of Wildlife Research. 37(2): 127-131.<br />

Warne, R.W., C.A. Gilman, and B.O. Wolfe. 2010a. Tissue-ca<strong>rb</strong>on incorporation rates in<br />

lizards: Implications <strong>fo</strong>r ecological studies using stable isotopes in terrestrial ectotherms.<br />

Physiological and Biochemical Zoology. 83(4): 608-617.<br />

Warne, R.W., A.D. Pershall, and B.O. Wolf. 2010b. Linking precipitation and C3-C4 plant<br />

production to resource dynamics in higher-trophic-level consumers. Ecology. 91(6):<br />

1628-1638.<br />

Whitaker, J.O. 1996. National Audubon Society: Field Guide to North American<br />

Mammals. New York: Alfred A. Knopf, Inc. 937 p.<br />

White, T.C.R. 1978. The importance of a relative shortage of <strong>fo</strong>od in animal ecology.<br />

Oecologia. 33: 71-86.<br />

Wigley, B.J., W.J. Bond, and M.T. Hoffman. 2010. Thicket expansion in a South African<br />

savanna under divergent land use: local vs. global drivers? Global Change Biology. 16:<br />

964-976.<br />

Windberg, L.A., S.M. Ebbert, and B.T. Kelly. 1997. Population characteristics of coyotes<br />

(Canis latrans) in the Northern Chihuahuan Desert of New Mexico. American Midland<br />

Naturalist. 138(1): 197-207.<br />

Windberg, L.A. and C.D. Mitchell. 1990. Winter diets of coyotes in relation to prey<br />

abundance in southern Texas. Journal of Mammalogy. 71(3): 439-447.<br />

Young, J.K., W.F. Andelt, P.A. Terletzky, and J.A. Shivik. 2006. A comparison of coyote<br />

ecology after 25 years: 1978 versus 2003. Canadian Journal of Zoology. 84: 573-582.<br />

Zarovali, M.P., M.D. Yiakoulaki, and V.P. Papanastasis. 2007. Effects of shrub<br />

encroachment on he<strong>rb</strong>age production and nutritive value in semi-arid Mediterranean<br />

grasslands. Grass and Forage Science. 62: 355-363.


Appendix 1. Intra-individual variation in diet.<br />

A)<br />

B)<br />

% coyote diet from woody plants<br />

% coyote diet from grass<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

1 15 16 30 38 55<br />

Individual #<br />

1 14 15 16 30 38 55<br />

Individual #<br />

126<br />

Appendix 1. Percent diet values determined via the ca<strong>rb</strong>on isotope analysis of A) bone<br />

and B) hair pieces removed from scat samples deposited by 6 (bone) or 7 (hair)<br />

individuals. All samples were collected during spring 2009. Each individual deposited 3-<br />

8 samples along transects located in either grassland or shrubland habitat. Error bars<br />

represent 95% confidence intervals.


Appendix 2. δ 13 C values <strong>fo</strong>r vegetation samples and scat components.<br />

-30 -25 -20 -15 -10<br />

delta 13 C<br />

Grass<br />

Woody<br />

Fo<strong>rb</strong><br />

Cactus<br />

Bone<br />

Hair<br />

127<br />

Appendix 2. δ 13 C values <strong>fo</strong>r plant samples and hair and bone components of scat samples<br />

collected at the <strong>Sevilleta</strong> NWR. The δ 13 C values <strong>fo</strong>r grasses and woody plants shown here<br />

were used to calculate average ca<strong>rb</strong>on isotope signatures <strong>fo</strong>r C3 woody plants and C4<br />

grasses (Equation 1). Each point on this graph represents either a single sample or the<br />

average of 2 replicates of a sample (scat components, <strong>fo</strong><strong>rb</strong>s, cacti) or the average of 1-3<br />

samples <strong>fo</strong>r a single genus (grasses) or species (woody plants). For the <strong>fo</strong><strong>rb</strong>s and cacti,<br />

each sample is from a different location. For the bones, each sample is from a scat from a<br />

different individual. The bone values shown here were used to calculate average values<br />

<strong>fo</strong>r percent coyote diet from C4 grasses and C3 woody plants in grassland versus<br />

shrubland habitats (Figure 2C). For the hair, each sample is from either a different<br />

individual or from the same individual but a different season. Hair values shown here<br />

were used to calculate average values <strong>fo</strong>r percent coyote diet from C4 grasses in grassland<br />

versus shrubland areas in spring, summer, and fall (Figure 2D). The values <strong>fo</strong>r bone and<br />

hair samples shown here have not been corrected <strong>fo</strong>r either diet to tissue or diet to scat<br />

fractionation.


Appendix 3. Inter-habitat and seasonal differences in <strong>fo</strong><strong>rb</strong> <strong>cover</strong>.<br />

A)<br />

B)<br />

% <strong>live</strong> <strong>fo</strong><strong>rb</strong> <strong>cover</strong><br />

% <strong>live</strong> <strong>fo</strong><strong>rb</strong> <strong>cover</strong><br />

100<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Grassland Shrubland<br />

<strong>Habitat</strong> <strong>type</strong><br />

Spring Summer Fall<br />

Season<br />

Grassland<br />

Shrubland<br />

128<br />

Appendix 3. A) Inter-habitat differences and B) seasonal variation in percent <strong>live</strong> <strong>cover</strong><br />

of <strong>fo</strong><strong>rb</strong>s. Percent <strong>live</strong> <strong>fo</strong><strong>rb</strong> <strong>cover</strong> was measured using the same techniques and in the same<br />

circular vegetation plots (n = 22) in which percent <strong>live</strong> grass and woody plant <strong>cover</strong> were<br />

assessed in spring, summer, and fall of 2009.


Appendix 4. Composition and diet of the small mammal community.<br />

A)<br />

<strong>Habitat</strong> <strong>type</strong> Latin name Common name Count<br />

Grassland Dipodomys ordii Ord’s kangaroo rat 120<br />

Perognathus flavus Silky pocket mouse 48<br />

Dipodomys spectabilis Banner-tailed kangaroo rat 15<br />

Peromyscus boylii Brush mouse 7<br />

Neotoma albigula White-throated woodrat 3<br />

Onychomys arenicola Mearn’s grasshopper mouse 3<br />

Shrubland Dipodomys merriami Merriam’s kangaroo rat 198<br />

Dipodomys spectabilis Banner-tailed kangaroo rat 94<br />

Dipodomys ordii Ord’s kangaroo rat 54<br />

Perognathus flavus Silky pocket mouse 48<br />

Peromyscus eremicus Cactus mouse 3<br />

Peromyscus leucopus White-<strong>fo</strong>oted mouse 3<br />

Reithrodontomys megalotis Western harvest mouse 2<br />

Neotoma albigula White-throated woodrat 1<br />

Onychomys arenicola Mearn’s grasshopper mouse 1<br />

B)<br />

Species Item in diet<br />

Average %<br />

volume <strong>Habitat</strong> <strong>type</strong><br />

Dipodomys merriami Seed (F,G) 69 Shrubland<br />

Plant 11<br />

Arthropod 21<br />

Dipodomys ordii Seed (C,F,G,S) 60 Grassland<br />

Plant 33<br />

Arthropod 7<br />

Dipodomys spectabilis Seed (C,F) 7 Grassland/Shrubland<br />

Plant 60<br />

Arthropod 3<br />

Unidentified 30<br />

Perognathus flavus Seed (C,F,G,S) 96 Grassland<br />

Arthropod 5<br />

Appendix 4. A) Abundance of small mammals surveyed at the <strong>Sevilleta</strong> NWR during the<br />

spring and fall of 2008 (based on Hope and Parmenter 2007, Friggens 2008). B) Diet<br />

(based on Hope and Parmenter 2007) of the 4 most abundant species of small mammals<br />

in grassland and shrubland habitats at the <strong>Sevilleta</strong> NWR (based on Appendix 4A,<br />

Friggens 2008). Diet data was collected in 1998 and was averaged across seasons. C =<br />

cactus, F = <strong>fo</strong><strong>rb</strong>, G = grass, S = shrub; Plant = green vegetation/plant matter.<br />

129


130<br />

Chapter 5: The impact of spatial scale on the relationship between<br />

coyote feeding ecology and local habitat characteristics<br />

Abstract<br />

Many ecological patterns are scale-dependent and consideration of spatial scale is<br />

important in studies of the ecology of wide-ranging animals. Little is known about the<br />

impacts of woody plant encroachment, a shift in habitat <strong>type</strong> from grass- to woody plant-<br />

dominated habitats, on the ecology of local predators. The purpose of this study is to<br />

determine whether the strength of the relationship between predator feeding ecology and<br />

local habitat characteristics in a woody plant-encroached area varies with spatial scale.<br />

The coyote (Canis latrans) is an abundant, top predator at the <strong>Sevilleta</strong> National Wildlife<br />

Refuge in central New Mexico, USA. Individual coyotes were identified using<br />

noninvasive genetic sampling of feces and stable ca<strong>rb</strong>on isotope techniques were used to<br />

assess coyote feeding ecology in native grassland and woody plant-encroached shrubland<br />

areas. <strong>Habitat</strong> characteristics were evaluated at 30 different spatial scales using circular<br />

buffers centered on scat collection sites and two different land <strong>cover</strong> maps. Buffer size<br />

ranged from 0.03 to 28.3 km 2 . The difference in percent coyote diet that came indirectly<br />

from C3 woody plants between grassland and shrubland habitats and the linear<br />

relationship between coyote diet and percent shrubland habitat were evaluated at each<br />

spatial scale. A significant difference in coyote diet between habitat <strong>type</strong>s was observed<br />

at a small spatial scale (buffer area = 0.03 km 2 ). There was no significant or strong linear<br />

relationship between coyote diet and percent shrubland habitat at any spatial scale. There<br />

were scale-driven break points in the p-values associated with coyote diet differences<br />

between habitat <strong>type</strong>s. These break points occurred at transitions between the <strong>fo</strong>llowing


131<br />

buffer sizes: 0.03 to 0.1 km 2 , 2.0 to 2.5 km 2 , 4.5 to 5.3 km 2 , 10.2 to 11.3 km 2 , and 15.2 to<br />

16.6 km 2 . Woody plant encroachment appears to have a significant impact on coyote<br />

feeding ecology, specifically on the base of the coyote <strong>fo</strong>od chain, but only when habitat<br />

characteristics are evaluated at a small spatial scale. The size of the difference in coyote<br />

diet between habitat <strong>type</strong>s declines rapidly when the spatial scale of the analysis exceeds<br />

previously published values <strong>fo</strong>r the sizes of both estimated core and total coyote home<br />

ranges in an arid environment. Overall, these results indicate that spatial scale impacts the<br />

strength of the relationship between coyote ecology and local habitat characteristics, and<br />

that coyotes may be <strong>fo</strong>raging in an area that is quite small relative to the size of a typical<br />

coyote home range.<br />

Introduction<br />

Spatial scale can play an important role in assessments of ecological patterns<br />

(Wiens 1989, Levin 1992, Ludwig et al. 2000, Trani (Griep) 2002), particularly in studies<br />

of animal ecology (Senft et al. 1987, McLoughlin et al. 2004, Bowyer and Kie 2006,<br />

Mayor et al. 2009, van Beest et al. 2010). Results of studies of mammalian ecology,<br />

especially of habitat use and selection patterns, can vary greatly with the spatial scale at<br />

which habitat variables are evaluated (e.g., Pereira and Itami 1991, Powell 1994, Carr et<br />

al. 2002, Apps et al. 2004, Jiang et al. 2010, Pedersen et al. 2010, van Beest et al. 2010,<br />

but see Vanak and Gompper 2010). Furthermore, the significance of the relationship<br />

between animal ecology and habitat characteristics can vary with spatial scale such that<br />

strong relationships are observed at one spatial scale of habitat assessment and not at<br />

another (Quinn 1997, Kie et al. 2002, McLoughlin et al. 2004, Wilson and Nielsen 2007).<br />

The concept of scale-dependency in animal-habitat relationships is particularly relevant


132<br />

in studies of wide-ranging species. Many studies that consider the importance of scale<br />

have <strong>fo</strong>cused on large he<strong>rb</strong>ivores. Far fewer studies have dealt with mammalian<br />

carnivores (Hobbs 2003, Bowyer and Kie 2006). Many mammalian carnivores are wide<br />

ranging (Sunquist and Sunquist 2001) and their home ranges can be as large as several<br />

hundred (several species, Lindstedt et al. 1986, Woodroffe and Ginsberg 1998; wild dog,<br />

Claridge et al. 2009) to over a thousand square kilometers (grizzly bear, Blanchard and<br />

Knight 1991; cheetah, Marker et al. 2008). These observations highlight the importance<br />

of considering multiple spatial scales when assessing the relationship between habitat<br />

characteristics and the ecology of mammalian predators.<br />

A large number and diversity of mammalian carnivores are present in areas<br />

affected by woody plant encroachment (Chapter 1), a widespread process of habitat<br />

change that is occurring on six of the seven continents (Archer 1995, Ravi et al. 2009).<br />

Relatively little is known about the effects of woody plant encroachment on local<br />

mammalian predators (Blaum et al. 2007a). This habitat change is characterized by a shift<br />

from grassland to shrubland, or savanna to woodland, habitat over a period of 50-200<br />

years (Archer 1989, Archer 1995, Gill and Burke 1999, Van Auken 2000). Drivers of<br />

woody plant encroachment include: overgrazing by cattle (Bester 1996, Van Auken<br />

2000), changes in local fire frequency (Archer 1995, Roques et al. 2001), elevated levels<br />

of ca<strong>rb</strong>on dioxide (Bond and Midgley 2000), and nitrogen deposition (Köchy and Wilson<br />

2001, Wigley et al. 2010). This habitat shift has many implications <strong>fo</strong>r the ecology of<br />

local predator populations. Woody plant encroachment leads to changes in habitat<br />

structure that can hinder predator movement and the ability of predators to detect prey<br />

(Muntifering et al. 2006) and reduce both hunting success (Mills et al. 2004) and prey


133<br />

availability (Marker 2002, Blaum et al. 2007b). More research regarding the effects of<br />

this habitat shift on predator ecology is needed (Blaum et al. 2007a, Chapter 1).<br />

Secondary consumer diet and feeding ecology varies in response to changes in<br />

lower trophic levels (e.g., Brand et al. 1976, Todd et al. 1981, Andelt et al. 1987, Dibello<br />

et al. 1990, Genovesi et al. 1996, Warne et al. 2010). Stable isotope techniques have been<br />

used to evaluate the feeding ecology of he<strong>rb</strong>ivores and carnivores (Botha and Stock 2005,<br />

Codron et al. 2005, Codron et al. 2007a, Codron et al. 2007b, Bowman et al. 2010). C3<br />

and C4 plants have distinct ca<strong>rb</strong>on stable isotope signatures as a result of differences in<br />

their photosynthetic pathways (Bender 1971, Farquhar 1983, Sternberg et al. 1984,<br />

Marshall et al. 2007, Warne et al. 2010). Encroaching, woody plants use the C3<br />

photosynthetic pathway while grasses at woody plant-encroached sites are often C4 plants<br />

(e.g., Boutton et al. 1998, Gill and Burke 1999, McKinley and Blair 2008, Muldavin et al.<br />

2008). As a result, stable ca<strong>rb</strong>on isotope techniques are particularly useful <strong>fo</strong>r assessing<br />

the effects of the encroachment of C3 woody plants into C4 native grassland areas on the<br />

feeding ecology of predators.<br />

The coyote (Canis latrans) is a wide ranging (Boisjoly et al. 2010) and abundant<br />

generalist predator (IUCN 2010). The size of home ranges of resident coyotes varies<br />

widely (Windberg et al. 1997, Young et al. 2006, Gehrt et al. 2009, Boisjoly et al. 2010)<br />

but can be as large as 121 km 2 (Canada; Boisjoly et al. 2010). Average home range size<br />

<strong>fo</strong>r resident coyotes in an arid environment comparable to that considered in this study is<br />

12.6 km 2 (Windberg et al. 1997). Even areas of high use, or coyote core home ranges,<br />

may be relatively large (e.g., 1.73-5.6 km 2 ; New Mexico, Windberg et al. 1997;<br />

Mississippi, Chamberlain et al. 2000; Florida, Thornton et al. 2004). However, some


134<br />

studies in the southwestern United States report small core areas (0.04-0.05 km 2 ; Young<br />

et al. 2006, Young et al. 2008). Coyote diet varies geographically (e.g., Hidalgo-Mihart et<br />

al. 2001, Samson and Crete 1997) and over time (Andelt et al. 1987). Of particular<br />

interest is a study showing that coyote diet tracked changes in the availability of fruits<br />

from different woody plants over a period of nearly twenty years (Andelt et al. 1987).<br />

Woody plant encroachment has been documented at many sites in North America<br />

(Archer 1995, Ravi et al. 2009, Chapter 1), and the distribution of the coyote overlaps<br />

most of these sites (IUCN 2010). These factors make coyotes an ideal <strong>fo</strong>cal species <strong>fo</strong>r a<br />

study of the impacts of woody plant encroachment on predator feeding ecology. Coyotes<br />

are capable of shifting their <strong>fo</strong>raging habits and diet in response to a change in habitat<br />

characteristics (i.e., woody plant encroachment), and they use large enough areas that an<br />

assessment of this animal ecology-habitat relationship at multiple spatial scales is both<br />

appropriate and advisable.<br />

My primary objective is to determine whether the relationship between predator<br />

feeding ecology and local habitat characteristics changes with the spatial scale of the<br />

analysis. My emphasis is on determining whether there is a shift in the base of the coyote<br />

<strong>fo</strong>od chain between native grassland and woody plant-encroached habitats and whether<br />

the spatial scale at which habitat characteristics are assessed impacts the strength of this<br />

shift. In particular, I address the <strong>fo</strong>llowing question: Does a) the size of the difference in<br />

the base of the coyote <strong>fo</strong>od chain between grassland and woody plant-encroached<br />

shrubland habitats, or b) the strength of the linear relationship between percent coyote<br />

diet from C3 woody plants and percent available shrubland habitat, change with the<br />

spatial scale at which the habitat variables are assessed? The stable ca<strong>rb</strong>on isotope data


135<br />

used in this study has been utilized previously to assess the relationship between coyote<br />

feeding ecology and woody plant encroachment (Chapter 4). This previous assessment<br />

was based on an evaluation of habitat characteristics at a small spatial scale (i.e., 30 m<br />

diameter vegetation plots). Given the wide-ranging nature of coyotes, I expected that the<br />

difference in coyote feeding ecology between habitat <strong>type</strong>s would become larger, and the<br />

association between coyote diet from C3 woody plants and percent shrubland habitat<br />

would become stronger, as the spatial scale of the analysis increased and approached the<br />

size of an area that coyotes typically use (i.e., the size of a coyote home range). I also<br />

expected that the strength of the relationship between coyote feeding ecology and habitat<br />

characteristics would decline at spatial scales larger than the area typically used by a<br />

coyote (i.e., larger than a coyote home range).<br />

Methods<br />

Study site<br />

The study site, field data collection and laboratory analyses described here are<br />

presented in much greater detail in Chapter 2. Briefly, the fieldwork <strong>fo</strong>r this study was<br />

per<strong>fo</strong>rmed at the <strong>Sevilleta</strong> National Wildlife Refuge (NWR) and Long Term Ecological<br />

Research (<strong>LTER</strong>) site in central New Mexico (34.32°N, 106.81°W). The refuge<br />

encompasses 1,000 km 2 (Hernández et al. 2002) and has roughly 200 miles of road. The<br />

average annual rainfall from 1988-2008 was 235 mm (based on Moore 2009).<br />

Approximately 72% of the refuge is <strong>cover</strong>ed by either grassland or shrubland habitat.<br />

Grassland is more prevalent in the northern half of the refuge and shrubland in the<br />

southern half. The eastern side of the refuge contains an active transition zone between<br />

grama (Bouteloua spp.) grassland and creosote (Larrea tridentata) shrubland. More


136<br />

specifically, creosote shrubs have been moving into grama grassland areas over the past<br />

century (Gill and Burke 1999). The coyote is one of several predators present at the<br />

<strong>Sevilleta</strong> NWR and the <strong>fo</strong>raging strategy of the local coyote population has been studied<br />

previously (Hernández et al. 2002). These characteristics make the refuge an ideal setting<br />

<strong>fo</strong>r an assessment of the impact that woody plant encroachment has on coyote feeding<br />

ecology.<br />

Field data collection<br />

Carnivore scat samples were collected between June 2008 and November 2009. 3<br />

surveys of 20 road-based transects (10 in grassland, 10 in shrubland) were per<strong>fo</strong>rmed in<br />

the summer (June-July) of 2008. 6 surveys of 22 transects (12 in grassland, 10 in<br />

shrubland) were conducted in 2009, two in each of three seasons: spring (April-May),<br />

summer (July), and fall (October-November). Each transect was 1 mile (1.6 km) long and<br />

was separated from all other transects by at least a mile (1.6 km). Transects were driven<br />

at slow speed (< 16.1 km/hr) and each carnivore scat that was encountered was sub-<br />

sampled <strong>fo</strong>r genetic analysis (see below). Roughly 0.4 mL of the fecal material on the<br />

outside of each sample was placed in a 2 mL tube containing DETs (DMSO, EDTA, Tris,<br />

salt) buffer (Frantzen et al. 1998). A GPS coordinate was recorded be<strong>fo</strong>re the remainder<br />

of the sample was collected <strong>fo</strong>r ca<strong>rb</strong>on isotope analysis (see below).<br />

In July and August 2008, vegetation was characterized in a total of 40 circular<br />

vegetation plots, two per scat transect. Each plot was 30 m in diameter and was located a<br />

randomly selected distance, between 30 and 100 m, from a scat transect. Vegetation<br />

variables relevant to woody plant-encroached landscapes were assessed in each plot.<br />

Specifically, data were collected on: 1) percent woody plant <strong>cover</strong>, 2) woody plant size


137<br />

and 3) inter-woody plant distance. These variables were measured to the nearest 0.1 m <strong>fo</strong>r<br />

<strong>live</strong> woody plants (i.e., plants with green leaves on some part) that intersected two,<br />

orthogonal, 30 m line intercept transects. Woody plant size was determined by measuring<br />

the height, the longest axis and the axis perpendicular to the longest axis <strong>fo</strong>r each woody<br />

plant that was at least 0.5 m tall. For the measurement of average inter-plant distance, up<br />

to <strong>fo</strong>ur woody plants with a height of at least 0.5 m were selected in each plot. The inter-<br />

plant distances were measured from the centers of these selected plants to the centers of<br />

the 5 closest woody plants that were also at least 0.5 m tall and were within the plot. In<br />

2009, small samples of the dominant woody plant species and one of the dominant grass<br />

genera <strong>fo</strong>und within each of 22 circular, 30 m diameter plots, one per scat transect, were<br />

collected in each season that scat surveys were conducted. These vegetation samples<br />

were then prepared <strong>fo</strong>r ca<strong>rb</strong>on isotope analysis (see below).<br />

Laboratory analyses<br />

All scat sub-samples stored in DETs buffer were extracted using QIAamp DNA<br />

Stool Mini Kits from Qiagen Inc. Primers (Murphy et al. 2000) that amplify a segment of<br />

the mitochondrial DNA (mtDNA) control region (Onorato et al. 2006) and polymerase<br />

chain reaction (PCR) techniques (see Chapter 2 <strong>fo</strong>r PCR profiles) were used to identify<br />

the species that had deposited each scat sample. This species identification test was done<br />

to screen out any samples that had been deposited by canids other than the coyote. All<br />

samples <strong>fo</strong>und to be from coyotes were run through a microsatellite analysis to determine<br />

the minimum number of different individuals that had been sampled. In particular, coyote<br />

samples were screened using primers <strong>fo</strong>r 8 canid-specific microsatellite loci (CXX173<br />

and CXX250, Ostrander et al. 1993; CXX377, Ostrander et al. 1995; FH2001, CXX2010,


138<br />

FH2054, and FH2088, Francisco et al. 1996; CXX119) All PCR products were separated<br />

via electrophoresis on an Applied Biosystems 3130xl capillary machine and all<br />

microsatellite data was viewed using GeneMapper 3.7. Samples <strong>fo</strong>r which 5 or more loci<br />

were successfully amplified were screened 2-6 more times with all 8 primers. Replicate<br />

microsatellite PCR results were compared to obtain a consensus geno<strong>type</strong> <strong>fo</strong>r each<br />

sample. All samples <strong>fo</strong>r which a consensus geno<strong>type</strong> was obtained at 6 or more loci were<br />

analyzed using Gimlet 1.3.3. (Valière 2002) to determine the minimum number of<br />

individuals from which samples had been collected.<br />

Prior to per<strong>fo</strong>rming ca<strong>rb</strong>on isotope analysis, all scat (70 o C) and vegetation (60 o C)<br />

samples were oven-dried <strong>fo</strong>r 24 hours. Vegetation samples were ground with a Wiley<br />

Mill until they passed through a 20 mesh filter. Small pieces of bone were removed from<br />

the first scat sample collected from each individual in the summer of 2008 and in each of<br />

the three seasons that surveys were per<strong>fo</strong>rmed in 2009. Individuals that moved between<br />

scat transects that were in different habitat <strong>type</strong>s at any point during the study were<br />

excluded from this analysis. Bone pieces were cleaned with ethanol and crushed to a<br />

powder. Small masses of each item (vegetation, 4mg; bone, 3mg) were weighed into<br />

5x9mm tin capsules. These weighed samples were analyzed on a coupled elemental<br />

analyzer (Costech 4010) and isotope ratio mass spectrometer (DeltaPlux XP) at the Stable<br />

Isotope Laboratory at the University of New Hampshire (UNH). The ca<strong>rb</strong>on isotope<br />

signatures (i.e., δ 13 C values) derived from the mass spectrometer <strong>fo</strong>r the bone samples<br />

were corrected <strong>fo</strong>r diet to tissue (1 o /oo, DeNiro and Epstein 1978) and diet to scat (0 o /oo,<br />

Chapter 2) fractionation. The corrected signatures were used to calculate the percentage


139<br />

of the diet of each individual in each season that came indirectly from C3 woody plants<br />

(Equation 1, based on Faure and Mensing 2005).<br />

Where<br />

δ 13 Cbone = δ 13 CC 3 (fC 3 ) + δ 13 CC 4 (1-fC 3 ) (1)<br />

δ 13 Cbone = the ca<strong>rb</strong>on isotope signature of bone taken from a coyote scat; δ 13 CC 3 = the<br />

average ca<strong>rb</strong>on isotope signature of C3 woody plants; fC 3 = the fraction of coyote diet that<br />

comes indirectly from C3 woody plants; δ 13 CC 4 = the average ca<strong>rb</strong>on isotope signature of<br />

C4 grasses.<br />

Multiple spatial scale analysis<br />

Field collected vegetation data on percent <strong>live</strong> woody plant <strong>cover</strong>, woody plant<br />

size, and inter-woody plant distance were used to per<strong>fo</strong>rm a discriminant function<br />

analysis (PROC DISCRIM, SAS 9.1) and identify key habitat variables that differentiated<br />

grassland and shrubland habitats. For this analysis, vegetation plots, and thus the<br />

vegetation data collected in those plots, were initially assigned to the same habitat <strong>type</strong> as<br />

the nearest scat transect. Scat transects were assigned to habitat <strong>type</strong>s based on a visual<br />

assessment of the vegetation visible from the road. Values were extracted from 6 Landsat<br />

7 ETM+ bands (1-5 and 7) <strong>fo</strong>r pixels (30 m resolution) corresponding to the locations of<br />

the field surveyed vegetation plots (30 m diameter) which were correctly classified as<br />

grassland or shrubland habitat in the discriminant function analysis. These raw digital<br />

numbers were converted to reflectance values (Appendix 1) and run through a second<br />

discriminant function analysis to determine which Landsat 7 ETM+ bands were most<br />

useful in differentiating grassland and shrubland habitats. The results of this second<br />

discriminant function analysis and the Landsat 7 ETM+ image, which was taken on


140<br />

March 31 st , 2009 and obtained from the U.S. Geological Survey<br />

(), were used to generate a map of grassland and shrubland<br />

habitat across the <strong>Sevilleta</strong> NWR. An image taken in the spring was considered more<br />

appropriate <strong>fo</strong>r distinguishing grass- and shrub-dominated habitats as some shrub species<br />

at the field site are evergreen and most of the grass biomass does not green up until the<br />

beginning of the summer rainy season.<br />

Both the Landsat 7 ETM+-derived map of grassland and shrubland habitats<br />

(Figure 1B) and a previously developed land <strong>cover</strong> map (Muldavin et al. 1998; Figure<br />

1A) were used to assess habitat characteristics at various spatial scales. In particular, data<br />

on the land <strong>cover</strong> <strong>type</strong>s of the areas encompassed by circles (radius = 100 m, area = 0.03<br />

km 2 ) that were centered on scat sample collection sites were extracted using ArcGIS 9.2.<br />

Only scats <strong>fo</strong>r which data on percent coyote diet from C3 woody plants had been obtained<br />

were included in this analysis. Percent area <strong>cover</strong>ed by each land <strong>cover</strong> <strong>type</strong> was<br />

calculated <strong>fo</strong>r each circle and thus <strong>fo</strong>r the habitat within a particular distance of each scat<br />

sample. Each scat sample, and thus each individual coyote, was assigned to the land<br />

<strong>cover</strong> <strong>type</strong> (grassland versus shrubland) with the highest percent area within the circle<br />

centered on the scat’s collection site. This analysis was repeated <strong>fo</strong>r circles with radii that<br />

increased to 3000 m (buffer area = 28.3 km 2 ) in increments of 100 m. The <strong>fo</strong>llowing<br />

analyses were per<strong>fo</strong>rmed at each spatial scale (i.e., circle size) considered: 1) a Student’s<br />

t-test (PROC TTEST, SAS 9.1) that assessed the difference in percent diet from C3<br />

woody plants between coyotes assigned to grassland versus shrubland areas, and 2) a<br />

linear regression analysis that evaluated the relationship between the percent of the


141<br />

coyote diet that comes indirectly from C3 woody plants and percent area <strong>cover</strong>ed by<br />

shrubland habitat (PROC REG, SAS 9.1).<br />

The upper bound of 28.3 km 2 <strong>fo</strong>r the size of the circular buffers, and thus the area<br />

in which habitat characteristics were assessed, was based on in<strong>fo</strong>rmation regarding<br />

average coyote home range size. The average size of a coyote home range as determined<br />

using noninvasive genetic sampling techniques and scat samples collected at the <strong>Sevilleta</strong><br />

NWR and <strong>LTER</strong> (Chapters 2 and 3) is 1.7 km 2 . Another study, which used<br />

radiotelemetry data to evaluate home range size and was per<strong>fo</strong>rmed in an arid<br />

environment comparable to the <strong>Sevilleta</strong> NWR, indicates that, <strong>fo</strong>r territorial coyotes, the<br />

average size of a core home range is 5.6 km 2 , while the average size of a total home range<br />

is 12.6 km 2 (Windberg et al. 1997). By using circles that range in size from 0.03 to 28.3<br />

km 2 , habitat characteristics were assessed in areas that are much smaller and notably<br />

larger than an average coyote home range <strong>fo</strong>r an arid environment.<br />

Results<br />

Field data collection and laboratory analyses<br />

A total of 935 carnivore scat samples were sub-sampled <strong>fo</strong>r genetic analysis. The<br />

mtDNA species identification test indicated that 69% of these samples had been<br />

deposited by coyotes. Analysis in Gimlet 1.3.3 of the 520 scats <strong>fo</strong>r which consensus<br />

geno<strong>type</strong>s were obtained at 6 or more microsatellite loci indicated that a minimum of 81<br />

individuals had been sampled (see Chapter 3 <strong>fo</strong>r further details). The δ 13 C values <strong>fo</strong>r<br />

samples (n = 14) of 5 dominant woody plant species, as well as 4 samples of two woody<br />

plant species which produce fruit or seeds eaten by coyotes (V. Seamster, unpublished<br />

data), were used to calculate an average (± 1 standard deviation, i.e., s.d.) ca<strong>rb</strong>on isotope


142<br />

signature <strong>fo</strong>r C3 woody plants (-24.9 o /oo ± 0.7). The δ 13 C values <strong>fo</strong>r samples (n = 18) of<br />

grass species from 6 genera were used to calculate an average ca<strong>rb</strong>on isotope signature<br />

<strong>fo</strong>r C4 grasses (-14.8 o /oo ± 0.3). Ca<strong>rb</strong>on isotope signatures of bone pieces taken from 62<br />

scat samples, each obtained from a different individual, were used to calculate values <strong>fo</strong>r<br />

percent coyote diet coming indirectly from C3 woody plants. These percent diet values<br />

were used in all t-test and regression analyses described below (see Chapter 4 <strong>fo</strong>r further<br />

details).<br />

Multiple spatial scale analysis<br />

The discriminant function analysis of field collected vegetation measurements<br />

was based on data from 37 plots and the association between different habitat <strong>type</strong>s<br />

(grassland versus shrubland) and the measured vegetation variables was significant (F =<br />

5.83, d.f. = 5, 31, p = 0.0007). The squared canonical correlation was 0.48, and 6 of 37<br />

plots (16%) were misclassified. The discriminant function was positively associated with<br />

percent woody plant <strong>cover</strong> and plant size and negatively associated with inter-plant<br />

distance (Appendix 2). The mean discriminant function value <strong>fo</strong>r grassland plots was<br />

-1.02 and <strong>fo</strong>r shrubland plots was 0.87. The discriminant function analysis of 6 Landsat 7<br />

ETM+ bands was based on reflectance values <strong>fo</strong>r 24 pixels corresponding to vegetation<br />

plot locations. The 6 plot locations misclassified in the discriminant function analysis of<br />

vegetation measurements were excluded from the Landsat analysis. Locations that fell in<br />

areas with missing data (Figure 1B) were also excluded. The association between<br />

different habitat <strong>type</strong>s (grassland versus shrubland) and reflectance values from Landsat 7<br />

ETM+ bands was significant (F = 7.04, d.f. = 6, 17, p = 0.0007). The squared canonical<br />

correlation was 0.71 and 3 of 24 plots (13%) were misclassified. The discriminant


143<br />

function was negatively associated with all Landsat 7 ETM+ bands except bands 3 (red)<br />

and 4 (near infrared) and was most strongly associated with reflectance values from<br />

bands 4 and 5 (middle infrared, Jensen 2007; Appendix 3). The mean discriminant<br />

function value <strong>fo</strong>r grassland locations was -1.51 and <strong>fo</strong>r shrubland locations was 1.51.<br />

The results of the linear discriminant function analysis (Appendix 3) were used to<br />

generate a map of grassland and shrubland habitat across the study site (Figure 1B).<br />

There was no significant difference (p > 0.05) in percent coyote diet from C3<br />

woody plants between grassland and shrubland areas at any spatial scale when habitat<br />

was evaluated using a previously developed map of the study site (Figures 1A and 2A).<br />

However, there was a significant difference in coyote diet between habitat <strong>type</strong>s at small<br />

spatial scales (p < 0.05, buffer size = 0.03 km 2 ) when a Landsat 7 ETM+-derived land<br />

<strong>cover</strong> map was used to evaluate habitat characteristics (Figures 1B and 2B). There were<br />

multiple sharp increases in the p-values at transitions between different buffer sizes. For<br />

the previously developed map, these sharp increases occurred when the spatial scale of<br />

the analysis increased from a buffer area of 0.03 to 0.1 km 2 and from 4.5 to 5.3 km 2<br />

(Figure 2A). For the Landsat 7 ETM+-derived map, sharp increases occurred when the<br />

buffer area increased from 2.0 to 2.5 km 2 , 10.2 to 11.3 km 2 , and from 15.2 to 16.6 km 2<br />

(Figure 2B). Regardless of the land <strong>cover</strong> map used to evaluate habitat characteristics,<br />

there were no strong or statistically significant relationships between percent coyote diet<br />

from C3 woody plants (dependent variable) and percent shrubland habitat (independent<br />

variable) at any spatial scale (R 2 < 0.05; d.f. = 1, 60; p > 0.05 <strong>fo</strong>r all spatial scales except<br />

buffer area 0.03 km 2 where d.f. = 1, 59).


A)<br />

B)<br />

144<br />

Figure 1. Maps of land <strong>cover</strong> at the study site. Two land <strong>cover</strong> maps used to assess<br />

habitat characteristics at the <strong>Sevilleta</strong> NWR. A) Previously developed map described by<br />

Muldavin et al. (1998). The original map was based on an analysis of Landsat Thematic<br />

Mapper images taken between 1987 and 1993 and had 13 land <strong>cover</strong> <strong>type</strong>s which were<br />

simplified to 3 <strong>fo</strong>r the purposes of this study. Black = shrubland, gray = grassland, white<br />

= other land <strong>cover</strong> <strong>type</strong>s including savanna, woodland, barren areas or water. B) Map<br />

developed <strong>fo</strong>r this study using a Landsat 7 ETM+ image taken on March 31 st , 2009.<br />

Black = shrubland, gray = grassland. White lines are the result of missing data. Light<br />

gray line represents the study site boundary <strong>fo</strong>r both maps.


A)<br />

B)<br />

T-value<br />

T-value<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

2.5<br />

2<br />

1.5<br />

1<br />

0.5<br />

0<br />

0 1 5 10 18 28<br />

Buffer area (km 2 )<br />

0 1 5 10 18 28<br />

Buffer area (km 2 )<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

1<br />

0.9<br />

0.8<br />

0.7<br />

0.6<br />

0.5<br />

0.4<br />

0.3<br />

0.2<br />

0.1<br />

0<br />

p -value<br />

p -value<br />

T-value<br />

p-value<br />

T-value<br />

p-value<br />

145<br />

Figure 2. Results of t-tests at multiple spatial scales. Results of t-tests of the difference<br />

in percent coyote diet from C3 woody plants <strong>fo</strong>r individuals sampled in grassland versus<br />

shrubland areas. <strong>Habitat</strong> <strong>type</strong> was evaluated using circular buffers centered on the<br />

locations at which coyote scats were sampled and one of two different land <strong>cover</strong> maps.<br />

The two land <strong>cover</strong> maps used were: A) a previously developed map of the study site<br />

(Figure 1A, Muldavin et al. 1998), B) a map of grassland and shrubland habitat generated<br />

from a Landsat 7 ETM+ image taken in March, 2009 (Figure 1B). d.f. = 60 <strong>fo</strong>r all tests<br />

except <strong>fo</strong>r buffer size of 0.03 km 2 in B, where d.f. = 59. Gray line indicates p = 0.05.


Discussion<br />

Multiple spatial scale analysis<br />

146<br />

The results of the discriminant function analysis of field measured vegetation<br />

variables indicate that woody plant-encroached shrubland areas had higher percent<br />

woody plant <strong>cover</strong> and larger woody plants while grassland areas were characterized by<br />

larger inter-woody plant distances. These results show that the vegetation plots, and thus<br />

the scat transects with which they were associated, were located in areas with structurally<br />

different habitats. The discriminant function analysis of 6 Landsat 7 ETM+ bands<br />

indicates that the reflectance of all bands, except bands 3 (red) and 4 (near infrared), was<br />

lower in shrubland than grassland areas. This is not surprising given the tendency of<br />

plants to reflect near infrared light (band 4), but is somewhat surprising given the<br />

tendency of plants to abso<strong>rb</strong> red light (band 3). Many vegetation indices developed <strong>fo</strong>r<br />

the analysis of satellite images depend on the difference in reflectance of red and near<br />

infrared radiation to detect variation in green, photosynthetically active plant biomass or<br />

in the leaf area indices of local vegetation (Tucker 1979, Running et al. 1994, Jensen<br />

2007). It is however important to note that the strength of the positive relationship<br />

between reflectance and the discriminant function was stronger <strong>fo</strong>r band 4 than band 3, so<br />

it is likely that reflectance of red light did not increase very much between grassland and<br />

shrubland areas.<br />

The vegetation map developed <strong>fo</strong>r this study differs from the land <strong>cover</strong> map<br />

developed by other researchers <strong>fo</strong>r the <strong>Sevilleta</strong> NWR (Muldavin et al. 1998). Though the<br />

general vegetation patterns are similar, with grassland prevalent in the northeastern part<br />

of the study area, the map developed <strong>fo</strong>r this study appears to overestimate the


147<br />

occurrence of shrubland habitat, especially on the western side of the refuge. Part of this<br />

discrepancy between the two land <strong>cover</strong> maps can be explained by differences in the<br />

number of land <strong>cover</strong> categories considered. The map developed by Muldavin et al.<br />

(1998) had a total of 13 land <strong>cover</strong> <strong>type</strong>s and was thus more sensitive to local habitat<br />

heterogeneity. Approximately 19% of the areas classified as shrubland in the Landsat 7<br />

ETM+-derived map were classified as woodland or savanna by Muldavin et al. (1998).<br />

This indicates that the approach taken in this study is more appropriate <strong>fo</strong>r differentiating<br />

grass- and woody plant-dominated areas rather than grassland and shrubland habitats and<br />

does not account <strong>fo</strong>r finer distinctions between habitat <strong>type</strong>s or fully represent the<br />

diversity of habitats present at the <strong>Sevilleta</strong> NWR. Differences between the images and<br />

procedures used to construct the two land <strong>cover</strong> maps should also be considered. Images<br />

used to create the map developed by Muldavin et al. (1998) were taken between<br />

September 1 st , 1987 and September 30 th , 1993 during the spring, summer, and fall<br />

seasons while the map generated <strong>fo</strong>r this study is based on a single image acquired on<br />

March 31 st , 2009. It is probable that some habitat changes have occurred in the 15+ years<br />

between the acquisition dates of these two sets of images. Given that creosote shrubs<br />

have been spreading into grassland areas at the <strong>Sevilleta</strong> NWR over the past century (Gill<br />

and Burke 1999, Báez and Collins 2008), it is likely that some areas classified as<br />

grassland by Muldavin et al. (1998) have since turned into shrubland. Furthermore,<br />

Muldavin et al. (1998) used a much more complicated procedure, which involved<br />

unsupervised image classification techniques <strong>fo</strong>llowed by extensive field surveys <strong>fo</strong>r the<br />

purpose of ground truthing, to develop their map. If the image classification approach<br />

utilized in this study were more similar to that used by Muldavin et al. (1998), then the


148<br />

two land <strong>cover</strong> maps (Muldavin et al. 1998 and this study) would likely be more<br />

comparable.<br />

Contrary to expectations, the size of the difference in coyote diet between habitat<br />

<strong>type</strong>s and the strength of the linear relationship between coyote diet and percent<br />

shrubland habitat neither increased consistently with spatial scale nor peaked at a spatial<br />

scale that approximates the size of an average coyote home range. The only significant<br />

relationship between coyote feeding ecology and habitat characteristics was observed at<br />

the smallest spatial scale when a Landsat 7 ETM+-derived land <strong>cover</strong> map was used to<br />

evaluate habitat and a t-test was used to assess the size of the difference in diet between<br />

habitat <strong>type</strong>s. There were no strong or significant linear relationships between coyote diet<br />

and percent shrubland habitat at any spatial scale <strong>fo</strong>r either land <strong>cover</strong> map. Regardless of<br />

the land <strong>cover</strong> map used, the size of all differences in coyote diet between habitat <strong>type</strong>s<br />

declined as the size of the area in which habitat was evaluated increased. The results of<br />

the habitat analysis based on a previously developed land <strong>cover</strong> map (Muldavin et al.<br />

1998) provide further support <strong>fo</strong>r the conclusion presented in Chapter 4 that woody plant<br />

encroachment does not appear to have a significant impact on coyote feeding ecology,<br />

specifically on the base of the coyote <strong>fo</strong>od chain. However, the results of the habitat<br />

analysis based on the Landsat 7 ETM+-derived map developed <strong>fo</strong>r this study indicate that<br />

woody plant encroachment does have a significant impact on coyote feeding ecology<br />

when habitat is evaluated at a small spatial scale (roughly 0.03 km 2 ). This could indicate<br />

that coyotes are <strong>fo</strong>raging in relatively small areas close to the locations where scat<br />

samples were collected. This observation of the importance of habitat characteristics<br />

measured on a small spatial scale mirrors findings of other studies of carnivore ecology


149<br />

(Ray 1998, Wilson and Nielsen 2007). For example, microhabitat characteristics have<br />

been <strong>fo</strong>und to be important in predicting the location of raccoon (Procyon lotor) daytime<br />

resting sites (Wilson and Nielsen 2007).<br />

There were five sharp increases or “break points” in p-values <strong>fo</strong>r the tests of<br />

differences in coyote diet between habitat <strong>type</strong>s, two <strong>fo</strong>r the analyses based on the land<br />

<strong>cover</strong> map developed by Muldavin et al. (1998) and 3 <strong>fo</strong>r the land <strong>cover</strong> map developed<br />

<strong>fo</strong>r this study. The first break point <strong>fo</strong>r the analysis that utilized the map developed by<br />

Muldavin et al. (1998) provides further support <strong>fo</strong>r the conclusion that woody plant<br />

encroachment appears to affect coyote feeding ecology only when habitat characteristics<br />

are evaluated at small spatial scales. For this break point, the size of the difference in<br />

average coyote diet between habitat <strong>type</strong>s declined as the spatial scale increased from a<br />

buffer size of 0.03 to 0.1 km 2 . The other <strong>fo</strong>ur break points can be partially explained in<br />

terms of in<strong>fo</strong>rmation on average coyote home range size presented in this dissertation and<br />

in previously published studies per<strong>fo</strong>rmed in arid environments comparable to my study<br />

site (i.e., Windberg et al. 1997). The first break point in the analysis based on the land<br />

<strong>cover</strong> map developed <strong>fo</strong>r this study occurred when the spatial scale of the analysis<br />

exceeded the average size of a coyote home range presented in this dissertation. In<br />

particular, the break point occurred when buffer size increased from 2.0 to 2.5 km 2 and<br />

the average size of a coyote home range is 1.7 km 2 (Chapters 2 and 3). The second break<br />

point <strong>fo</strong>r the analysis based on a land <strong>cover</strong> map developed by Muldavin et al. (1998)<br />

occurred when the evaluation of habitat approached the spatial scale of a core home range<br />

of a territorial coyote. In particular, the break point occurred when buffer size increased<br />

from 4.5 to 5.3 km 2 , and the average size of a core home range <strong>fo</strong>r territorial coyotes


150<br />

<strong>fo</strong>und in an arid environment is 5.6 km 2 (Windberg et al. 1997). The second and third<br />

break points <strong>fo</strong>r the Landsat 7 ETM+-based analysis occurred when the spatial scale of<br />

the habitat assessment approached and exceeded the size of a resident coyote home range,<br />

respectively. More specifically, the break points occurred at the transitions from buffers<br />

of size 10.2 to 11.3 km 2 and 15.2 to 16.6 km 2 . The average size of a home range of a<br />

territorial coyote in an arid environment is 12.6 km 2 (Windberg et al. 1997). These break<br />

points match my expectation that the strength of the relationship between coyote feeding<br />

ecology and local habitat characteristics would decline when habitat is evaluated across<br />

areas larger than those that coyotes typically use.<br />

There are several possible reasons why a statistically significant relationship<br />

between coyote feeding ecology and local habitat characteristics was observed only at a<br />

small spatial scale (buffer area = 0.03 km 2 ). Other studies have <strong>fo</strong>und that the specific<br />

habitat variables associated with carnivore movement patterns vary with spatial scale<br />

(e.g., Powell 1994, Carr et al. 2002, Constible et al. 2006, Pedersen et al. 2010). For<br />

example, a study of Asiatic black bears (Ursus thibetanus) showed that, at a smaller<br />

spatial scale, variables related to <strong>fo</strong>od availability were important in explaining habitat<br />

use patterns while, at the larger spatial scale, bears selected areas that had minimal<br />

agricultural development (Carr et al. 2002). This emphasizes the possibility that, at larger<br />

spatial scales (buffer area > 0.03 km 2 ), there is a relationship between coyote feeding<br />

ecology and habitat characteristics other than the ones considered here. Since I evaluated<br />

the same habitat characteristic at each spatial scale in each analysis per<strong>fo</strong>rmed (i.e.,<br />

habitat <strong>type</strong> <strong>fo</strong>r the t-test and percent shrubland habitat <strong>fo</strong>r the linear regression), I<br />

precluded the possibility of finding significant relationships between feeding ecology and


151<br />

different habitat characteristics at different spatial scales. Additionally, the results of<br />

various studies provide evidence that coyotes may <strong>fo</strong>rage on a relatively small spatial<br />

scale (Laundré and Keller 1981, Reichel 1991). In particular, coyote movement rates are<br />

lower when they are hunting as opposed to just moving through their home range<br />

(Laundré and Keller 1981) and there is evidence that coyotes seek out small patches with<br />

high prey densities (Reichel 1991). Given my emphasis on feeding ecology, my results<br />

are strongly affected by the spatial scale at which coyotes <strong>fo</strong>rage, and thus by a process<br />

that may occur at a relatively fine scale. Furthermore, a high percentage of coyote<br />

movements may be unrelated to <strong>fo</strong>raging and coyotes may hunt primarily when they<br />

encounter a prey item in the course of per<strong>fo</strong>rming other activities (Sacks and Neale<br />

2002). Given the utility of trails and roads as movement corridors <strong>fo</strong>r carnivores<br />

(Macdonald 1980, Mahon et al. 1998, Harmsen et al. 2010) and my use of road-based<br />

scat transects, it is possible that many of the scats I collected had been deposited by<br />

coyotes that traveled along roads and <strong>fo</strong>raged whenever they encountered prey on or in<br />

the vicinity of the road. Finally, my use of circular buffers centered on scat collection<br />

locations to characterize the habitat in which the scats were deposited, and thus the areas<br />

potentially used by coyotes while <strong>fo</strong>raging, is very simple. This approach does not<br />

account <strong>fo</strong>r the possibility that scat samples were deposited at the periphery, rather than<br />

the center, of a coyote’s movements (Gese and Ruff 1997), nor does it account <strong>fo</strong>r the<br />

asymmetrical shape of the areas used by a coyote (e.g., a coyote home range;<br />

Chamberlain et al. 2000, Young et al. 2008).<br />

The conclusions presented in Chapter 4 highlight another important factor<br />

contributing to the paucity of statistically significant results obtained in this study; woody


152<br />

plants are not the only C3 plants present at the study site. Fo<strong>rb</strong>s are present in both<br />

grassland and shrubland habitats at the <strong>Sevilleta</strong> NWR (Báez and Collins 2008, Chapter<br />

4) and several species of small mammals <strong>fo</strong>und at the refuge are known to eat <strong>fo</strong><strong>rb</strong> seeds<br />

(Hope and Parmenter 2007). The techniques used in this study do not allow me to<br />

differentiate between C3 woody plants and C3 <strong>fo</strong><strong>rb</strong>s in my evaluation of the base of the<br />

coyote <strong>fo</strong>od chain. However, my analysis of habitat characteristics only provides<br />

in<strong>fo</strong>rmation on the availability of C3 woody plants. If I had been able to consider the<br />

availability of both C3 <strong>fo</strong><strong>rb</strong>s and C3 woody plants in my evaluation of the relationship<br />

between percent coyote diet from C3 plants and local habitat characteristics, it is likely<br />

that I would have observed stronger relationships at several spatial scales.<br />

Conclusion<br />

Overall, the magnitude of the observed shift in coyote feeding ecology between<br />

native grassland and woody plant-encroached areas does vary with spatial scale. A<br />

significant shift is only observed when habitat characteristics are assessed at a small<br />

spatial scale (buffer area = 0.03 km 2 ). This could indicate that coyotes are <strong>fo</strong>raging in<br />

fairly small areas that are close to where their scats were sampled along road-based<br />

transects. Despite the lack of statistically significant relationships at larger spatial scales,<br />

the results presented here do demonstrate an expected decline in the size of the difference<br />

in coyote diet between habitat <strong>type</strong>s as the scale of the habitat analysis exceeds the size of<br />

areas typically used by coyotes (buffer area = 12.6 km 2 ). An evaluation of habitat<br />

variables other than the ones considered here might lead to the observation of a<br />

significant relationship between coyote feeding ecology and habitat characteristics at<br />

intermediate spatial scales (0.03 km 2 < buffer area < 12.6 km 2 ). A habitat assessment that


153<br />

provides in<strong>fo</strong>rmation on the availability of all C3 plants, which include both woody plants<br />

and <strong>fo</strong><strong>rb</strong>s, or that addresses the non-circular shape of the areas typically used by coyotes,<br />

might also lead to the observation of significant relationships across a wider range of<br />

spatial scales.


References<br />

154<br />

Andelt, W.F., J.G. Kie, F.F. Knowlton, and K. Cardwell. 1987. Variation in coyote diets<br />

associated with season and successional changes in vegetation. Journal of Wildlife<br />

Management. 51(2): 273-277.<br />

Apps, C.D., B.N. McLellan, J.G. Woods, and M.F. Proctor. 2004. Estimating grizzly bear<br />

distribution and abundance relative to habitat and human influence. Journal of Wildlife<br />

Management. 68(1): 138-152.<br />

Archer, S. 1989. Have southern Texas savannas been converted to woodlands in recent<br />

history? The American Naturalist. 134(4): 545-561.<br />

Archer, S. 1995. Tree-grass dynamics in a Prosopis-thornscrub savanna parkland:<br />

Reconstructing the past and predicting the future. Ecoscience. 2(1): 83-99.<br />

Báez, S. and S.L. Collins. 2008. Shrub invasion decreases diversity and alters community<br />

stability in northern Chihuahuan desert plant communities. PLoS ONE. 3(6): e2332.<br />

doi:10.1371/journal.pone.0002332.<br />

Bender, M.M. 1971. Variations in the 13 C/ 12 C ratios of plants in relation to the pathway<br />

of photosynthetic ca<strong>rb</strong>on dioxide fixation. Phytochemistry. 10:1239-1244.<br />

Bester, B. 1996. Bush encroachment: a thorny problem. Namibia Environ. 1: 175-177.<br />

Blanchard, B.M. and R.R. Knight. 1991. Movements of Yellowstone grizzly bears.<br />

Biological Conservation. 58: 41-67.<br />

Blaum, N., E. Rossmanith, A. Popp, and F. Jeltsch. 2007a. Shrub encroachment affects<br />

mammalian carnivore abundance and species richness in semiarid rangelands. Acta<br />

Oecologica. 31: 86-92.<br />

Blaum, N., E. Rossmanith, M. Schwager, and F. Jeltsch. 2007b. Responses of<br />

mammalian carnivores to land use in arid savanna rangelands. Basic and Applied<br />

Ecology. 8: 552-564.<br />

Boisjoly, D., J.P. Ouellet, and R. Courtois. 2010. Coyote habitat selection and<br />

management implications <strong>fo</strong>r the Gaspesie caribou. Journal of Wildlife Management.<br />

74(1): 3-11.<br />

Bond, W.J. and G.F. Midgley. 2000. A proposed CO2-controlled mechanism of woody<br />

plant invasion in grasslands and savannas. Global Change Biology. 6: 865-869.


155<br />

Botha, M.S. and W.D. Stock. 2005. Stable isotope composition of faeces as an indicator<br />

of seasonal diet selection of wild he<strong>rb</strong>ivores in southern Africa. South African Journal of<br />

Science. 101: 371-374.<br />

Boutton, T.W., S.R. Archer, A.J. Midwood, S.F. Zitzer, and R. Bol. 1998. δ 13 C values of<br />

soil organic ca<strong>rb</strong>on and their uses in documenting vegetation change in a subtropical<br />

savanna ecosystem. Geoderma. 82: 5-41.<br />

Bowman, D.M.J.S., B.P. Murphy, and C.R. McMahon. 2010. Using ca<strong>rb</strong>on isotope<br />

analysis of the diet of two introduced Australian megahe<strong>rb</strong>ivores to understand<br />

Pleistocene megafaunal extinctions. Journal of Biogeography. 37: 499-505.<br />

Bowyer, R.T. and J.G. Kie. 2006. Effects of scale on integrating life-history<br />

characteristics of ungulates and carnivores. Diversity and Distributions. 12: 244-257.<br />

Brand, C.J., L.B. Keith, and C.A. Fischer. 1976. Lynx responses to changing snowshoe<br />

hare densities in central Alberta. Journal of Wildlife Management. 40(3): 416-428.<br />

Carr, M.M., J. Yoshizaki, F.T. Van Manen, M.R. Pelton, O.C. Huygens, H. Hayashi, and<br />

M. Maekawa. 2002. A multi-scale assessment of habitat use by Asiatic black bears in<br />

central Japan. Ursus. 13: 1-9.<br />

Chamberlain, M.J., C.D. Lovell, and B.D. Leopold. 2000. Spatial-use patterns,<br />

movements, and interactions among adult coyotes in central Mississippi. Canadian<br />

Journal of Zoology. 78: 2087-2095.<br />

Claridge, A.W., D.J. Mills, R. Hunt, D.J. Jenkins, and J. Bean. 2009. Satellite tracking of<br />

wild dogs in south-eastern mainland Australian <strong>fo</strong>rests: Implications <strong>fo</strong>r management of a<br />

problematic top-order carnivore. Forest Ecology and Management. 258: 814-822.<br />

Codron, D., J. Codron, J.A. Lee-Thorp, M. Sponheimer, and D. de Ruiter. 2005. Animal<br />

diets in the Wate<strong>rb</strong>erg based on stable isotopic composition of faeces. South African<br />

Journal of Wildlife Research. 35(1): 43-52.<br />

Codron, D., J. Codron, J.A Lee-Thorp, M. Sponheimer, D. de Ruiter, and J.S. Brink.<br />

2007a. Stable isotope characterization of mammalian predator-prey relationships in a<br />

South African savanna. European Journal of Wildlife Research. 53: 161-170.<br />

Codron, D., J. Codron, J.A. Lee-Thorp, M. Sponheimer, D. de Ruiter, J. Sealy, R. Grant,<br />

and N. Fourie. 2007b. Diets of savanna ungulates from stable ca<strong>rb</strong>on isotope composition<br />

of faeces. Journal of Zoology. 273: 21-29.<br />

Constible, J.M., M.J. Chamberlain, and B.D. Leopold. 2006. Relationships between<br />

landscape pattern and space use of three mammalian carnivores in central Mississippi.<br />

American Midland Naturalist. 155(2): 352-362.


DeNiro, M.J. and S. Epstein. 1978. Influence of diet on the distribution of ca<strong>rb</strong>on<br />

isotopes in animals. Geochimica et Cosmochimica Acta. 42: 495-506.<br />

156<br />

Dibello, F.J., S.M. Arthur, and W.B. Krohn. 1990. Food habits of sympatric coyotes,<br />

Canis latrans, red <strong>fo</strong>xes, Vulpes vulpes, and bobcats, Lynx rufus, in Maine. The Canadian<br />

Field-Naturalist. 104: 403-408.<br />

Farquhar, G.D. 1983. On the nature of ca<strong>rb</strong>on isotope discrimination in C4 species.<br />

Australian Journal of Plant Physiology. 10: 205-226.<br />

Faure, G. and T.M. Mensing. 2005. Isotopes: Principles and Applications. Hoboken,<br />

New Jersey: John Wiley & Sons. Pps. 693-752.<br />

Francisco, L.V., A.A. Langston, C.S. Mellersh, C.L. Neal, and E.A. Ostrander. 1996. A<br />

class of highly polymorphic tetranucleotide repeats <strong>fo</strong>r canine genetic mapping.<br />

Mammalian Genome. 7: 359-362.<br />

Frantzen, M.A.J., J.B. Silk, J.W.H. Ferguson, R.K. Wayne, and M.H. Kohn. 1998.<br />

Empirical evaluation of preservation methods <strong>fo</strong>r fecal DNA. Molecular Ecology. 7:<br />

1423–1428.<br />

Gehrt, S.D., C. Anchor, and L.A. White. 2009. Home range and landscape use of coyotes<br />

in a metropolitan landscape: Conflict or coexistence? Journal of Mammalogy. 90(5):<br />

1045-1057.<br />

Genovesi, P., M. Secchi, and L. Boitani. 1996. Diet of stone martens: an example of<br />

ecological flexibility. Journal of Zoology. 238: 545-555.<br />

Gese, E.M. and R.L. Ruff. 1997. Scent-marking by coyotes, Canis latrans: the influence<br />

of social and ecological factors. Animal Behaviour. 54: 1155-1166.<br />

Gill, R.A. and I.C. Burke. 1999. Ecosystem consequences of plant life <strong>fo</strong>rm changes at<br />

three sites in the semiarid United States. Oecologia. 121: 551-563.<br />

Harmsen, B.J., R.J. Foster, S. Silver, L. Ostro, and C.P. Doncaster. 2010. Differential use<br />

of trails by <strong>fo</strong>rest mammals and the implications <strong>fo</strong>r camera-trap studies: A case study<br />

from Belize. Biotropica. 42(1): 126-133.<br />

Hernández, L., R.R. Parmenter, J.W. Dewitt, D.C. Light<strong>fo</strong>ot, and J.W. Laundré. 2002.<br />

Coyote diets in the Chihuahuan Desert, more evidence <strong>fo</strong>r optimal <strong>fo</strong>raging. Journal of<br />

Arid Environments. 51: 613-624.


157<br />

Hidalgo-Mihart, M.G., L. Cantú-Salazar, C.A. López-González, E. Martínez-Meyer, and<br />

A. González-Romero. 2001. Coyote (Canis latrans) <strong>fo</strong>od habits in a tropical deciduous<br />

<strong>fo</strong>rest of western Mexico. The American Midland Naturalist. 146: 210-216.<br />

Hobbs, N.T. 2003. Challenges and opportunities in integrating ecological knowledge<br />

across scales. Forest Ecology and Management. 181: 223-238.<br />

Hope, A.G. and R.R. Parmenter. 2007. Food habits of rodents inhabiting arid and semiarid<br />

ecosystems of central New Mexico. Special Publication of the Museum of<br />

Southwestern Biology. 9: 1-75.<br />

IUCN 2010. IUCN Red List of Threatened Species. Version 2010.3.<br />

. (16 October 2010).<br />

Jensen, J.R. 2007. Remote Sensing of the Environment. 2 nd Ed. Upper Saddle River, NJ:<br />

Pearson Prentice Hall. Pps. 193-248, 355-408.<br />

Jiang, G., J. Ma, M. Zhang, and P. Stott. 2010. Multi-scale <strong>fo</strong>raging habitat use and<br />

interactions by sympatric cervids in northeastern China. Journal of Wildlife Management.<br />

74(4): 678-689.<br />

Kie, J.G., R.T. Bowyer, M.C. Nicholson, B.B. Boroski, and E.R. Loft. 2002. Landscape<br />

heterogeneity at differing scales: Effects on spatial distribution of mule deer. Ecology.<br />

83(2): 530-544.<br />

Köchy, M. and S.D. Wilson. 2001. Nitrogen deposition and <strong>fo</strong>rest expansion in the<br />

northern Great Plains. Journal of Ecology. 89: 807-817.<br />

Laundré, J.W. and B.L. Keller. 1981. Home-range use by coyotes in Idaho. Animal<br />

Behaviour. 29:449-461.<br />

Levin, S.A. 1992. The problem of pattern and scale in ecology: The Robert H. MacArthur<br />

Award lecture. Ecology. 73(6): 1943-1967.<br />

Lindstedt, S.L., B.J. Miller, and S.W. Buskirk. 1986. Home range, time and body size in<br />

mammals. Ecology. 67(2): 413-418.<br />

Ludwig, J.A., J.A. Wiens, and D.J. Tongway. 2000. A scaling rule <strong>fo</strong>r landscape patches<br />

and how it applies to conserving soil resources in savannas. Ecosystems. 3: 84-97.<br />

Macdonald, D.W. 1980. Patterns of scent marking with urine and faeces amongst<br />

carnivore communities. Symposium of the Zoological Society of London. 45: 107-139.


158<br />

Mahon, P.S., P.B. Banks, and C.R. Dickman. 1998. Population indices <strong>fo</strong>r wild<br />

carnivores: a critical study in sand-dune habitat, south-western Queensland. Wildlife<br />

Research. 25: 11-22.<br />

Marker, L. 2002. Aspects of Namibian cheetah (Acinonyx jubatus) biology, ecology and<br />

conservation strategies. PhD. Thesis, Department of Zoology, University of Ox<strong>fo</strong>rd. Pps.<br />

199-242.<br />

Marker, L.L., A.J. Dickman, M.G.L. Mills, R.M. Jeo, and D.W. Macdonald. 2008.<br />

Spatial ecology of cheetahs on north-central Namibian farmlands. Journal of Zoology.<br />

274: 226-238.<br />

Marshall, J.D., J.R. Brooks, and K. Lajtha. 2007. Sources of variation in the stable<br />

isotopic composition of plants. In R. Michener and K. Lajtha (eds). Stable isotopes in<br />

ecology and environmental science. Ox<strong>fo</strong>rd: Blackwell Publishing. Pps. 22-60.<br />

Mayor, S.J., D.C. Schneider, J.A. Schaefer, and S.P. Mahoney. 2009. <strong>Habitat</strong> selection at<br />

multiple spatial scales. Ecoscience. 16(2): 238-247.<br />

McKinley, D.C. and J.M. Blair. 2008. Woody plant encroachment by Juniperus<br />

virginiana in a mesic native grassland promotes rapid ca<strong>rb</strong>on and nitrogen accrual.<br />

Ecosystems. 11: 454-468.<br />

McLoughlin, P.D., L.R. Walton, H.D. Cluff, P.C. Paquet, and M.A. Ramsay. 2004.<br />

Hierarchical habitat selection by tundra wolves. Journal of Mammalogy. 85(3): 576-580.<br />

Mills, M.G.L., L.S. Broomhall, and J.T. du Toit. 2004. Cheetah Acinonyx jubatus feeding<br />

ecology in the Kruger National Park and a comparison across African savanna habitats: is<br />

the cheetah only a successful hunter on open grassland plains? Wildlife Biology. 10(3):<br />

177-186.<br />

Moore, D. 2009. ClimDB Monthly Data. Albuquerque, NM: <strong>Sevilleta</strong> Long Term<br />

Ecological Research Site. <br />

(8 February 2010).<br />

Muldavin, E.H., D.I. Moore, S.L. Collins, K.R. Wetherill, and D.C. Light<strong>fo</strong>ot. 2008.<br />

Aboveground net primary production dynamics in a northern Chihuahuan Desert<br />

Ecosystem. Oecologia. 155: 123-132.<br />

Muldavin, E., G. Shore, K. Taugher, and B. Milne. 1998. A vegetation classification and<br />

map <strong>fo</strong>r the <strong>Sevilleta</strong> National Wildlife Refuge, New Mexico. New Mexico Natural<br />

Heritage Program and <strong>Sevilleta</strong> Long Term Ecological Research Program. Biology<br />

Department, University of New Mexico, Albuquerque, NM 87131.


159<br />

Muntifering, J.R., A.J. Dickman, L.M. Perlow, T. Hruska, P.G. Ryan, L.L. Marker, and<br />

R.M. Jeo. 2006. Managing the matrix <strong>fo</strong>r large carnivores: a novel approach and<br />

perspective from cheetah (Acinonyx jubatus) habitat suitability modeling. Animal<br />

Conservation. 9: 103-112.<br />

Murphy, M.A., L.P. Waits, and K.C. Kendall. 2000. Quantitative evaluation of fecal<br />

drying methods <strong>fo</strong>r brown bear DNA analysis. Wildlife Society Bulletin. 28: 951–957.<br />

NASA. 2009. Chapter 11: Data Products. In Landsat 7: Science Data Users Handbook.<br />

(20 September 2010).<br />

NOAA. 2010. Solar Position Calendar.<br />

(20 September 2010).<br />

Onorato, D., C. White, P. Zager, and L.P. Waits. 2006. Detection of predator presence at<br />

elk mortality sites using mtDNA analysis of hair and scat samples. Wildlife Society<br />

Bulletin. 34(3): 815-820.<br />

Ostrander, E.A., F.A. Mapa, M. Yee, and J. Rine. 1995. One hundred and one new simple<br />

sequence repeat-based markers <strong>fo</strong>r the canine genome. Mammalian Genome. 6: 192-195.<br />

Ostrander, E.A., G.F. Sprague, and J. Rine. 1993. Identification and characterization of<br />

dinucleotide repeat (CA)n markers <strong>fo</strong>r genetic mapping in dog. Genomics. 16: 207-213.<br />

Pedersen, A., R.A. Ims, N.G. Yoccoz, and V.H. Hausner. 2010. Scale-dependent<br />

responses of predators and their prey to spruce plantations in subarctic birch <strong>fo</strong>rests in<br />

winter. Ecoscience. 17(2): 123-136.<br />

Pereira, J.M.C. and R.M. Itami. 1991. GIS-Based habitat modeling using logistic multiple<br />

regression: A study of Mt. Graham Red Squirrel. Photogrammetric Engineering &<br />

Remote Sensing. 57(11): 1475-1486.<br />

Powell, R.A. 1994. Effects of scale on habitat selection and <strong>fo</strong>raging behavior of fishers<br />

in winter. Journal of Mammalogy. 75(2): 349-356.<br />

Quinn, T. 1997. Coyote (Canis latrans) habitat selection in u<strong>rb</strong>an areas of western<br />

Washington via analysis of routine movements. Northwest Science. 71(4): 289-297.<br />

Ravi, S., P. D’Odorico, L. Wang, C.S. White, G.S. Okin, S.A. Macko, and S.L. Collins.<br />

2009. Post-fire resource redistribution in desert grasslands: A possible negative feedback<br />

on land degradation. Ecosystems. 12: 434-444.<br />

Ray, J.C. 1998. Temporal variation of predation on rodents and shrews by small African<br />

<strong>fo</strong>rest carnivores. Journal of Zoology. 244: 363-370.


Reichel, J.D. 1991. Relationships among coyote <strong>fo</strong>od habits, prey populations, and<br />

habitat use. Northwest Science. 65(3): 1991.<br />

160<br />

Roques, K.G., T.G. O'Connor, and A.R. Watkinson. 2001. Dynamics of shrub<br />

encroachment in an African savanna: relative influences of fire, he<strong>rb</strong>ivory, rainfall and<br />

density dependence. Journal of Applied Ecology. 38:268-280.<br />

Running, S.W., C.O. Justice, V. Salomonson, D. Hall, J. Barker, Y.J. Kaufmann, A.H.<br />

Strahler, A.R. Huete, J.-P. Muller, V. Vande<strong>rb</strong>ilt, Z.M. Wan, P. Teillet, and D.<br />

Carneggie. 1994. Terrestrial remote sensing science and algorithms planned <strong>fo</strong>r<br />

EOS/MODIS. International Journal of Remote Sensing. 15(17): 3587-3620.<br />

Sacks, B.N. and J.C.C. Neale. 2002. Foraging strategy of a generalist predator toward a<br />

special prey: coyote predation on sheep. Ecological Applications. 12(1): 299-306.<br />

Samson, C. and M. Créte. 1997. Summer <strong>fo</strong>od habits and population density of coyotes,<br />

Canis latrans, in boreal <strong>fo</strong>rests of southeastern Québec. The Canadian Field-Naturalist.<br />

111: 227-233.<br />

Senft, R.L., M.B. Coughenour, D.W. Bailey, R.L. Rittenhouse, O.E. Sala, and D.M.<br />

Swift. 1987. Large he<strong>rb</strong>ivore <strong>fo</strong>raging and ecological hierarchies. BioScience. 37(11):<br />

789-799.<br />

Sternberg, L.O., M.J. DeNiro, and H.B. Johnson. 1984. Isotope ratios of cellulose from<br />

plants having different photosynthetic pathways. Plant Physiology. 74: 557-561.<br />

Sunquist, M.E. and F. Sunquist. 2001. Changing landscapes: Consequences <strong>fo</strong>r<br />

carnivores. In J.L. Gittleman, S.M. Funk, D. Macdonald, and R.K. Wayne (eds).<br />

Carnivore Conservation. Cambridge: Cambridge University Press. Pps. 399-418.<br />

Thornton, D.H., M.E. Sunquist, and M.B. Main. 2004. Ecological separation within<br />

newly sympatric populations of coyotes and bobcats in south-central Florida. Journal of<br />

Mammalogy. 85(5): 973-982.<br />

Todd, A.W., L.B. Keith, and C.A. Fischer. 1981. Population ecology of coyotes during a<br />

fluctuation of snowshoe hares. Journal of Wildlife Management. 45(3): 629-640.<br />

Trani (Griep), M.K. 2002. The influence of spatial scale on landscape pattern description<br />

and wildlife habitat assessment. In J.M. Scott, P.J. Heglund, M.L. Morrison, J.B. Haufler,<br />

M.G. Raphael, W.A. Wall, and R.B. Samson (eds). Predicting Species Occurrences:<br />

Issues of Accuracy and Scale. Washington, D.C.: Island Press. Pps. 141-155.<br />

Tucker, C.J. 1979. Red and photographic infrared linear combinations <strong>fo</strong>r monitoring<br />

vegetation. Remote Sensing of Environment. 8: 127-150.


Valière, N. 2002. GIMLET: a computer program <strong>fo</strong>r analyzing genetic individual<br />

identification data. Molecular Ecology Notes. 10.1046/j.1471-8278 .2002.00228.x.<br />

161<br />

Vanak, A.T. and M.E. Gompper. 2010. Multi-scale resource selection and spatial ecology<br />

of the Indian <strong>fo</strong>x in a human-dominated dry grassland ecosystem. Journal of Zoology.<br />

281: 140-148.<br />

Van Auken, O.W. 2000. Shrub invasions of North American semiarid grasslands. Annual<br />

Review of Ecology and Systematics. 31: 197-215.<br />

van Beest, F.M., A. Mysterud, L.E. Loe, and J.M. Milner. 2010. Forage quantity, quality,<br />

and depletion as scale-dependent mechanisms driving habitat selection of a large<br />

browsing he<strong>rb</strong>ivore. Journal of Animal Ecology. 79: 910-922.<br />

Warne, R.W., A.D. Pershall, and B.O. Wolf. 2010. Linking precipitation and C3-C4 plant<br />

production to resource dynamics in higher-trophic-level consumers. Ecology. 91(6):<br />

1628-1638.<br />

Wiens, J.A. 1989. Spatial scaling in ecology. Functional Ecology. 3(4):385-397.<br />

Wigley, B.J., W.J. Bond, and M.T. Hoffman. 2010. Thicket expansion in a South African<br />

savanna under divergent land use: local vs. global drivers? Global Change Biology. 16:<br />

964-976.<br />

Wilson, S.E. and C.K. Nielsen. 2007. <strong>Habitat</strong> characteristics of raccoon daytime resting<br />

sites in southern Illinois. American Midland Naturalist. 157(1):175-186.<br />

Windberg, L.A., S.M. Ebbert, and B.T. Kelly. 1997. Population characteristics of coyotes<br />

(Canis latrans) in the Northern Chihuahuan Desert of New Mexico. American Midland<br />

Naturalist. 138(1): 197-207.<br />

Woodroffe, R. and J.R. Ginsberg. 1998. Edge effects and the extinction of populations<br />

inside protected areas. Science. 280(5372): 2126-2128.<br />

Young, J.K., W.F. Andelt, P.A. Terletzky, and J.A. Shivik. 2006. A comparison of coyote<br />

ecology after 25 years: 1978 versus 2003. Canadian Journal of Zoology. 84: 573-582.<br />

Young, J.K., S.N. Glasscock, and J.A. Shivik. 2008. Does spatial structure persist despite<br />

resource and population changes? Effects of experimental manipulations on coyotes.<br />

Journal of Mammalogy. 89(5): 1094-1104.


Appendix 1. Converting Landsat 7 ETM+ pixel values to reflectance values.<br />

162<br />

All of the equations and values in this Appendix are based on NASA (2009) or NOAA<br />

(2010). Pixel values, in digital number (DN), were first converted to values <strong>fo</strong>r spectral<br />

radiance (Equation 2) and then to values <strong>fo</strong>r planetary reflectance (Equation 3).<br />

Lλ = ((LMAXλ-LMINλ)/(QCALMAX-QCALMIN))*(QCAL-QCALMIN)+LMINλ (2)<br />

Where<br />

Lλ = spectral radiance (watts/m 2 *sr*µm)<br />

LMAXλ = spectral radiance scaled to QCALMAX (watts/m 2 *sr*µm; Appendix 1A)<br />

LMINλ = spectral radiance scaled to QCALMIN (watts/m 2 *sr*µm; Appendix 1A)<br />

QCALMAX = maximum Landsat 7 ETM+ pixel value (255 DN)<br />

QCALMIN = minimum Landsat 7 ETM+ pixel value (1 DN)<br />

QCAL = Landsat 7 ETM+ pixel value<br />

Where<br />

ρp = (π*Lλ*d 2 )/(ESUNλ*cosθs) (3)<br />

ρp = planetary reflectance (unitless)<br />

Lλ = spectral radiance (watts/m 2 *sr*µm)<br />

d = earth-sun distance (0.99897 astronomical units)<br />

ESUNλ = mean solar exoatmospheric irradiances (watts/m 2 *µm; Appendix 1B)<br />

cosθs = cosine of solar zenith angle (0.3977; NOAA 2010)<br />

A)<br />

Landsat band LMINλ LMAXλ<br />

B)<br />

1 -6.2 293.7<br />

2 -6.4 300.9<br />

3 -5 234.4<br />

4 -5.1 241.1<br />

5 -1 47.57<br />

7 -0.35 16.54<br />

Landsat band ESUNλ<br />

1 1997<br />

2 1812<br />

3 1533<br />

4 1039<br />

5 230.8<br />

7 84.9<br />

Appendix 1. A) Values <strong>fo</strong>r LMAXλ and LMINλ in watts/m 2 *sr*µm used <strong>fo</strong>r Equation 2.<br />

B) Values <strong>fo</strong>r ESUNλ in watts/m 2 *µm used <strong>fo</strong>r Equation 3.


163<br />

Appendix 2. Discriminant function analysis of vegetation variables.<br />

A)<br />

Vegetation variable Correlation coefficient<br />

Percent woody plant <strong>cover</strong> 0.75<br />

Inter-plant distance -0.61<br />

Long axis 0.45<br />

Short axis 0.40<br />

Height 0.50<br />

B)<br />

Variable Grassland Shrubland<br />

Constant -5.59 -8.01<br />

Percent woody plant <strong>cover</strong> 0.12 0.20<br />

Inter-plant distance 0.92 0.47<br />

Long axis 9.16 6.96<br />

Short axis -10.99 -8.17<br />

Height 7.19 11.27<br />

Appendix 2. A) Correlation coefficients showing the strength of the association between<br />

field-measured vegetation variables and the discriminant function. B) Results of the<br />

linear discriminant function analysis.


164<br />

Appendix 3. Discriminant function analysis of reflectance values from a Landsat 7<br />

ETM+ image.<br />

A)<br />

Landsat band Correlation coefficient<br />

1 (blue) -0.07<br />

2 (green) -0.05<br />

3 (red) 0.03<br />

4 (near infrared) 0.13<br />

5 (middle infrared) -0.24<br />

7 (middle infrared) -0.12<br />

B)<br />

Variable Grassland Shrubland<br />

Constant -849.36 -883.48<br />

Band 1 11597 11767<br />

Band 2 -9171 -9238<br />

Band 3 -2419 -2666<br />

Band 4 2479 2779<br />

Band 5 -1347 -1523<br />

Band 7 3588 3710<br />

Appendix 3. A) Correlation coefficients showing the strength of the association between<br />

reflectance values <strong>fo</strong>r each Landsat 7 ETM+ band and the discriminant function.<br />

In<strong>fo</strong>rmation on Landsat band wavelengths taken from Jensen (2007). B) Results of the<br />

linear discriminant function analysis. These coefficients were used to classify the Landsat<br />

7 ETM+ image and generate a map of grassland and shrubland land <strong>cover</strong> <strong>type</strong>s across<br />

the <strong>Sevilleta</strong> NWR (Figure 1B).


Conclusion<br />

165<br />

This dissertation addressed several questions relating to the consequences of<br />

woody plant encroachment <strong>fo</strong>r mammalian predators. The first question posed was: How<br />

many mammalian carnivores are present in areas affected by woody plant encroachment<br />

globally? According to a dataset on the global distribution of mammalian carnivores and<br />

a map of woody plant-encroached sites, there are at least 97 different carnivores <strong>fo</strong>und in<br />

areas affected by woody plant encroachment. Given what is currently known about how<br />

carnivores respond to woody plant encroachment, many of these species could be<br />

negatively affected by continued spread of woody plants into native grassland areas. In<br />

particular, some carnivore populations appear to decline in abundance once a critical<br />

threshold of woody plant <strong>cover</strong> is reached. Other carnivores are physically hindered by<br />

the presence of woody plants; they are less successful at hunting, experience a reduction<br />

of prey availability, and are even injured by contact with woody plants as they move<br />

through the landscape. All of these observations highlight the need <strong>fo</strong>r further<br />

investigation of the impact of woody plant encroachment on the ecology of mammalian<br />

predators.<br />

The remaining three questions posed in this dissertation dealt with the ecology of<br />

one predator, the coyote (Canis latrans), that is abundant throughout North America and<br />

is a top predator at the <strong>Sevilleta</strong> National Wildlife Refuge (NWR) in New Mexico, USA.<br />

In particular, the <strong>fo</strong>llowing questions were addressed: 1) Is there a difference in the base<br />

of the coyote <strong>fo</strong>od chain between native grassland and woody plant-encroached<br />

shrubland habitats?; 2) Is there seasonal variation in the base of the coyote <strong>fo</strong>od chain in<br />

response to pulses of grass productivity in an area impacted by woody plant


166<br />

encroachment?; 3) Does a) the size of the difference in the base of the coyote <strong>fo</strong>od chain<br />

between grassland and woody plant-encroached shrubland habitats, or b) the strength of<br />

the linear relationship between percent coyote diet from C3 woody plants and percent<br />

available shrubland habitat, change with the spatial scale at which the habitat variables<br />

are assessed? The answer to the first two questions is no; woody plant encroachment and<br />

seasonal variation in grass productivity in a woody plant-encroached area do not lead to<br />

significant shifts in the base of the coyote <strong>fo</strong>od chain. However, the answer to the first<br />

question is tempered by the answer to the third question; a statistically significant shift in<br />

the base of the coyote <strong>fo</strong>od chain between habitat <strong>type</strong>s is observed when habitat<br />

characteristics are assessed at a small spatial scale. Thus, the assessment of the<br />

relationship between coyote feeding ecology and habitat characteristics is a scale-<br />

dependent process.<br />

There are several additional conclusions regarding the feeding ecology of coyotes<br />

in a woody plant-encroached environment. First, C3 plants, which include woody plants<br />

and <strong>fo</strong><strong>rb</strong>s, appear to be an important <strong>fo</strong>od resource in the local <strong>fo</strong>od web in grassland<br />

areas relative to their availability across the landscape. This matches observations by<br />

other researchers that these plants are of higher nutritional quality than C4 plants. Percent<br />

coyote diet that comes indirectly from C4 plants <strong>fo</strong>llows different seasonal trends in<br />

grassland versus shrubland habitats, especially between the spring and summer seasons.<br />

In shrubland areas, coyote diet from C4 grasses <strong>fo</strong>llows a trend that matches expectations<br />

and mirrors the spike in grass productivity in response to summer rains. In grassland<br />

areas, it is likely that consumption of C3 <strong>fo</strong><strong>rb</strong>s by coyote prey increases from the spring to


167<br />

the summer. Finally, coyotes appear to <strong>fo</strong>rage in areas that are small relative to the size of<br />

a typical coyote home range.<br />

Overall, the consequences of woody plant encroachment <strong>fo</strong>r mammalian predators<br />

are varied and significant. Further research regarding the impacts of woody plant<br />

encroachment on the ecology and fitness of mammalian predators is needed, especially in<br />

areas where the decline of the local grass population in response to woody plant<br />

encroachment is more severe than it has been to date at the <strong>Sevilleta</strong> NWR. Of particular<br />

interest are studies of the nutritional quality of C3 versus C4 plants and the impacts of a<br />

shift from C4 grasses to C3 woody plants at the base of the <strong>fo</strong>od chain on the survival and<br />

fitness of both primary and secondary consumers. Of equal importance are studies of<br />

facets of predator feeding ecology other than the base of the <strong>fo</strong>od chain, and of the<br />

ecology of more specialized mammalian predators <strong>fo</strong>und in woody plant-encroached<br />

areas. The predator emphasized in this dissertation is a generalist in terms of its diet and<br />

patterns of habitat use. This ecological flexibility is ideal <strong>fo</strong>r detecting shifts in ecology<br />

between habitat <strong>type</strong>s and in response to seasonal variation. However, other species with<br />

more specialized habitat and diet requirements are less likely to be able to shift their<br />

resource use patterns in response to woody plant encroachment. Populations of these<br />

more specialized species are likely to decline and may eventually go locally extinct in<br />

woody plant-encroached areas.


Isotope appendices<br />

Appendix 1. Seasonal variation in coyote diet.<br />

A) B)<br />

% coyote diet from grass<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Spring Summer Fall<br />

Season<br />

C) D)<br />

% coyote diet from grass<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Spring Summer Fall<br />

Season<br />

% coyote diet from grass<br />

100<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

0<br />

Spring Summer Fall<br />

Season<br />

Spring Summer Fall<br />

Season<br />

Appendix 1. Seasonal variation in percent coyote diet from grass <strong>fo</strong>r individuals sampled<br />

in A) grassland habitat in 3 seasons (n = 3); B) shrubland habitat in 3 seasons (n = 7); C)<br />

grassland habitat in 2 seasons (n = 13); D) shrubland habitat in 2 seasons (n = 7). The<br />

data from all of these individuals was used to assess average percent coyote diet from C4<br />

grasses in both grassland and shrubland habitats in the spring, summer, and fall (Figure<br />

2D; Chapter 4). The individuals <strong>fo</strong>r which intra-seasonal variation is shown in Appendix<br />

1, Chapter 4 are shown in this appendix as <strong>fo</strong>llows: Individual 1, Appendix 1C, white<br />

diamonds; Individual 14, Appendix 1C, gray triangles; Individual 15, Appendix 1B, gray<br />

diamonds; Individual 16, Appendix 1C, black x’s; Individual 30, Appendix 1D, gray<br />

diamonds; Individual 55, Appendix 1B, gray crosses.<br />

% coyote diet from grass<br />

168


Appendix 2. Assessing variation in vegetation end-member values.<br />

A) B)<br />

% coyote diet from woody plants<br />

C)<br />

% coyote diet from grass<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Grassland Shrubland<br />

<strong>Habitat</strong> <strong>type</strong><br />

Spring Summer Fall<br />

Season<br />

Average<br />

Narrow range<br />

Wide range<br />

Average<br />

Narrow range<br />

Wide range<br />

% coyote diet from grass<br />

100<br />

90<br />

80<br />

70<br />

60<br />

50<br />

40<br />

30<br />

20<br />

10<br />

0<br />

Spring Summer Fall<br />

Season<br />

169<br />

Average<br />

Narrow range<br />

Wide range<br />

Appendix 2. Average values <strong>fo</strong>r percent coyote diet from: A) woody plants in grassland<br />

and shrubland habitats, B) grasses in grassland areas, and C) grasses in shrubland areas.<br />

Values <strong>fo</strong>r coyote diet calculated based on: average end member values <strong>fo</strong>r C3 woody<br />

plants and C4 grasses; values calculated by adding or subtracting a standard deviation<br />

from the average end member values and using the C3 and C4 end member values with<br />

the smallest (narrow range) or largest (wide range) difference between them. Error bars<br />

represent 95% confidence intervals. The same set of A) bone or B) and C) hair samples<br />

was used <strong>fo</strong>r the calculations <strong>fo</strong>r each series (average, narrow range, wide range). The<br />

values in the “average” series in each of these graphs were used to generate figures<br />

(Figure 2C and D) and run statistical analyses (Table 3) in Chapter 4.


170<br />

Appendix 3. Ca<strong>rb</strong>on and nitrogen isotope data <strong>fo</strong>r scat samples.<br />

Month Day Year <strong>Habitat</strong> Lab_ID Individual Component Wt (mg)<br />

15<br />

N N%<br />

13<br />

C C%<br />

6 24 2008 G 1 1 hair 1.023 8.6 14.8 -14.2 43.0<br />

6 24 2008 G 1 1 bone 3.141 6.3 3.9 -17.5 15.6<br />

6 24 2008 G 12 3 hair 0.893 5.4 14.5 -20.3 43.3<br />

6 24 2008 G 12 3 bone_rep 2.049 4.4 2.6 -20.6 8.0<br />

6 24 2008 G 12 3 bone 2.09 4.5 2.9 -20.4 9.7<br />

6 24 2008 G 16 2 bone 2.214 4.7 3.3 -14.2 10.9<br />

6 24 2008 G 31 5 bone 2.965 2.1 3.7 -17.8 12.6<br />

6 24 2008 G 32 6 hair 0.936 4.2 14.0 -18.5 41.2<br />

6 24 2008 S 38 7 bone 3.118 7.0 3.4 -15.9 10.9<br />

6 24 2008 S 40 8 bone 3.076 4.1 3.4 -17.0 12.1<br />

6 26 2008 G 42 9 hair 1.023 6.6 14.4 -17.9 41.9<br />

6 26 2008 G 42 9 bone_rep 2.173 7.9 4.0 -18.2 13.5<br />

6 26 2008 G 42 9 bone 2.14 8.3 3.6 -18.3 12.5<br />

6 27 2008 S 44 10 bone 3.157 5.2 3.4 -15.7 11.1<br />

7 8 2008 G 47 12 hair 1.081 5.1 14.0 -20.4 43.1<br />

7 8 2008 G 47 12 bone 3.125 6.3 11.1 -20.6 41.8<br />

7 23 2008 G 51 14 hair 1.064 8.8 13.6 -17.1 39.8<br />

7 23 2008 G 51 14 bone 3.094 7.4 2.9 -17.4 9.5<br />

7 7 2008 S 52 15 hair 1.112 4.1 14.5 -20.5 42.4<br />

7 7 2008 S 52 15 bone 2.922 4.7 3.8 -18.6 13.5<br />

7 7 2008 G 53 16 hair 0.938 3.4 13.9 -18.8 42.7<br />

7 7 2008 G 53 16 bone 1.986 3.5 4.0 -19.3 15.3<br />

7 8 2008 G 54 17 hair 0.959 6.2 13.7 -17.4 41.4<br />

7 8 2008 G 54 17 bone 2.223 7.2 4.1 -16.3 15.1<br />

7 8 2008 G 57 2 bone 2.995 5.2 4.1 -15.2 13.3<br />

6 25 2008 S 63 19 hair 0.89 3.9 13.9 -17.3 41.5<br />

6 25 2008 S 63 19 bone 1.949 4.6 3.3 -18.8 10.5<br />

6 25 2008 S 64 20 bone 3.062 8.9 3.5 -13.8 11.2<br />

6 25 2008 G 67 21 hair 0.633 7.2 13.4 -20.1 40.9<br />

6 25 2008 G 73 23 hair 1.05 5.0 14.5 -19.7 42.9<br />

6 25 2008 G 73 23 bone 1.9 4.4 3.8 -18.6 14.8<br />

6 25 2008 G 75 24 hair 0.955 6.1 15.0 -10.3 44.1<br />

6 25 2008 G 75 24 bone 2.869 6.5 3.8 -13.5 13.0<br />

6 25 2008 G 76 25 bone 2.852 4.4 3.3 -21.1 11.4<br />

6 25 2008 G 76 25 hair 0.113 8.3 4.2 -16.9 29.7<br />

6 25 2008 G 78 26 hair 0.855 7.9 14.9 -17.6 50.9<br />

6 25 2008 G 78 26 bone 3.08 9.1 3.5 -15.2 12.6<br />

6 26 2008 G 92 27 hair 1.103 2.8 13.8 -18.3 40.8<br />

6 26 2008 G 92 27 bone 2.852 3.1 3.3 -17.8 13.8<br />

6 26 2008 G 93 27 bone 2.822 3.5 3.3 -14.9 10.9<br />

6 26 2008 G 94 28 hair 0.328 6.5 1.9 -24.1 44.1<br />

6 26 2008 G 95 2 bone 1.656 3.2 2.5 -17.8 8.9<br />

6 26 2008 G 96 29 hair 0.875 4.3 14.5 -17.7 43.6<br />

6 26 2008 G 96 29 bone 3.073 3.9 3.4 -15.5 11.4<br />

6 27 2008 S 106 30 hair 0.164 6.2 13.6 -18.7 34.3<br />

6 27 2008 G 115 31 hair 0.969 2.3 12.5 -22.9 42.9<br />

6 27 2008 G 115 31 bone_rep 2.887 3.6 3.8 -21.7 14.6<br />

6 27 2008 G 115 31 bone 3.112 3.6 3.9 -21.7 15.2


Month Day Year <strong>Habitat</strong> Lab_ID Individual Component Wt (mg)<br />

15<br />

N N%<br />

13<br />

C<br />

171<br />

C%<br />

6 28 2008 S 123 33 hair 0.918 3.1 14.2 -19.4 43.3<br />

6 28 2008 S 123 33 bone 2.988 3.3 3.9 -18.7 15.1<br />

6 28 2008 S 124 10 bone 3.012 2.5 3.8 -18.6 13.4<br />

6 28 2008 S 128 34 hair 1.103 7.4 14.2 -22.0 42.7<br />

6 28 2008 S 128 34 bone 2.917 5.1 3.5 -20.6 11.6<br />

6 28 2008 S 133 35 hair 1.201 3.5 14.3 -16.5 42.7<br />

6 28 2008 S 133 35 bone 2.884 4.1 3.7 -18.5 12.8<br />

6 28 2008 S 138 36 hair 0.87 5.1 14.3 -19.4 43.9<br />

6 28 2008 S 138 36 bone 2.984 4.1 2.1 -19.7 7.8<br />

6 28 2008 S 139 37 hair_rep 1.041 3.8 18.9 -21.4 55.0<br />

6 28 2008 S 139 37 hair 0.935 3.9 14.4 -21.1 42.9<br />

6 28 2008 S 139 37 bone 2.877 3.0 3.1 -20.0 11.2<br />

6 28 2008 S 146 38 hair_rep 0.898 5.3 13.6 -20.5 42.5<br />

6 28 2008 S 146 38 hair 0.933 5.3 13.3 -20.5 41.9<br />

6 28 2008 S 146 38 bone 2.922 6.8 3.4 -17.3 10.6<br />

6 28 2008 S 147 37 bone 2.93 6.2 4.1 -19.7 13.7<br />

7 7 2008 G 161 2 bone 2.86 5.7 3.5 -20.3 13.9<br />

7 7 2008 G 162 39 hair 1.019 4.1 12.0 -20.6 37.7<br />

7 7 2008 G 162 39 bone 3.033 4.0 3.2 -17.2 11.5<br />

7 7 2008 G 168 27 bone 1.942 3.6 4.2 -18.9 14.6<br />

7 7 2008 S 170 40 bone 2.832 5.2 2.6 -19.0 8.7<br />

7 7 2008 S 170 40 hair 1.13 6.7 3.5 -20.4 22.9<br />

7 7 2008 S 177 41 hair 0.998 4.0 13.6 -17.9 40.8<br />

7 7 2008 S 177 41 bone 1.901 2.8 3.5 -17.4 11.3<br />

7 8 2008 G 199 5 bone 2.806 4.7 1.3 -17.4 4.0<br />

7 8 2008 G 201 5 bone 2.916 6.7 3.5 -18.1 11.9<br />

7 8 2008 G 204 42 bone 2.907 5.0 3.8 -17.0 13.4<br />

7 8 2008 G 207 27 bone 2.942 5.5 4.2 -21.7 15.7<br />

7 8 2008 G 211 2 bone 2.926 4.5 3.6 -17.8 12.8<br />

7 8 2008 G 214 2 bone 3.151 5.3 4.1 -15.7 13.7<br />

7 9 2008 S 216 30 bone 2.964 6.5 3.8 -20.4 13.4<br />

7 9 2008 S 217 30 bone 3.172 7.0 3.9 -18.7 12.5<br />

7 9 2008 S 217 30 bone_rep 2.87 7.1 3.3 -19.3 10.6<br />

7 20 2008 G 248 2 bone 2.156 5.3 3.9 -18.5 15.0<br />

7 21 2008 S 255 43 hair 1.117 2.5 14.5 -18.7 43.9<br />

7 21 2008 S 255 43 bone 3.131 5.8 3.2 -15.6 10.7<br />

7 21 2008 G 267 17 bone 2.065 6.0 3.9 -21.2 14.5<br />

7 22 2008 G 273 5 bone 1.189 6.6 9.6 -18.0 31.8<br />

7 23 2008 S 279 44 hair 1.172 4.1 12.8 -19.2 38.3<br />

7 23 2008 S 279 44 bone 2.801 2.8 3.3 -16.9 10.4<br />

7 23 2008 S 280 30 bone 2.991 7.1 3.1 -19.3 10.1<br />

7 23 2008 G 282 45 hair 0.984 5.8 14.5 -17.6 42.6<br />

7 23 2008 G 282 45 bone 3.017 5.0 3.5 -19.3 12.7<br />

7 24 2008 S 289 37 bone 3.057 3.4 3.4 -15.4 11.5<br />

7 24 2008 S 295 37 bone 2.897 5.7 4.5 -14.5 15.0<br />

4 13 2009 G 298 1 hair 0.955 4.9 13.9 -18.4 40.9<br />

4 13 2009 G 298 1 bone 2.123 4.9 3.8 -17.3 14.4<br />

4 13 2009 G 300 1 hair 1.002 5.7 13.0 -16.2 40.8<br />

4 13 2009 G 300 1 bone_rep 2.887 5.6 3.8 -15.5 13.3


Month Day Year <strong>Habitat</strong> Lab_ID Individual Component Wt (mg)<br />

15<br />

N N%<br />

13<br />

C<br />

172<br />

C%<br />

4 13 2009 G 300 1 bone 2.853 5.2 3.8 -15.7 13.6<br />

4 13 2009 G 301 46 hair 1.051 4.8 13.9 -19.7 41.5<br />

4 13 2009 G 301 46 bone 1.941 3.9 3.6 -16.2 12.2<br />

4 13 2009 G 303 2 bone 2.2 4.1 3.1 -14.9 10.3<br />

4 13 2009 G 304 3 bone 0.352 6.8 2.9 -13.4 9.4<br />

4 13 2009 G 307 39 bone 2.872 8.6 3.7 -16.5 12.0<br />

4 13 2009 G 309 29 bone 3.025 8.0 2.4 -11.5 8.3<br />

4 13 2009 G 309 29 hair 0.082 6.4 4.6 -20.7 48.2<br />

4 13 2009 S 312 19 hair 1.076 5.8 13.6 -20.6 40.5<br />

4 13 2009 S 312 19 bone 2.912 4.3 3.9 -20.0 16.2<br />

4 13 2009 S 313 47 bone 2.097 6.0 4.1 -17.0 14.6<br />

4 13 2009 S 315 43 hair 0.862 5.0 14.4 -19.8 42.7<br />

4 13 2009 S 315 43 bone 2.113 4.0 3.7 -18.7 12.8<br />

4 13 2009 S 317 41 hair 0.41 8.1 14.4 -15.4 39.1<br />

4 13 2009 S 317 41 bone 2.198 9.0 2.4 -13.8 8.2<br />

4 13 2009 S 324 15 hair 0.903 7.8 14.0 -13.1 42.3<br />

4 13 2009 S 324 15 bone 2.044 6.7 5.5 -17.6 20.6<br />

4 13 2009 S 325 15 hair 0.89 5.9 14.6 -19.2 47.7<br />

4 13 2009 S 325 15 bone 3.067 4.6 2.5 -20.3 8.9<br />

4 13 2009 S 331 15 hair 0.872 6.1 12.9 -14.8 39.1<br />

4 13 2009 S 331 15 bone 3.028 5.8 3.0 -15.8 10.6<br />

4 14 2009 S 333 15 hair 0.882 7.1 17.7 -17.2 52.7<br />

4 14 2009 S 333 15 bone 2.98 5.3 3.8 -18.3 14.0<br />

4 14 2009 G 338 16 hair 1.154 4.7 13.8 -16.3 42.7<br />

4 14 2009 G 338 16 bone_rep 2.906 3.4 3.2 -16.6 11.1<br />

4 14 2009 G 338 16 bone 3.274 2.8 1.2 -16.3 4.2<br />

4 14 2009 G 341 16 hair_rep 1.003 5.0 14.5 -15.9 41.4<br />

4 14 2009 G 341 16 hair 0.899 5.0 14.4 -16.6 43.3<br />

4 14 2009 G 341 16 bone 2.884 5.0 3.3 -18.3 12.3<br />

4 14 2009 G 343 17 hair 1.173 8.2 13.9 -12.8 43.1<br />

4 14 2009 G 343 17 bone 2.917 5.0 2.2 -17.9 8.9<br />

4 14 2009 G 344 16 hair 0.987 5.0 14.0 -15.9 44.0<br />

4 14 2009 G 344 16 bone 2.846 4.1 3.7 -17.4 14.1<br />

4 14 2009 G 345 16 hair 1.03 4.2 14.4 -17.1 43.2<br />

4 14 2009 G 345 16 bone 3.204 2.8 3.0 -14.7 10.1<br />

4 14 2009 G 346 16 hair 1.075 8.7 14.4 -17.2 43.0<br />

4 14 2009 G 346 16 bone 0.36 4.9 4.4 -15.7 20.0<br />

4 14 2009 G 347 16 bone 2.896 3.9 2.1 -13.6 7.5<br />

4 14 2009 G 351 24 hair 1.037 5.9 14.5 -17.0 42.4<br />

4 14 2009 G 351 24 bone_rep 3.142 6.2 4.3 -16.0 14.8<br />

4 14 2009 G 351 24 bone 2.912 5.9 3.5 -16.2 12.3<br />

4 14 2009 G 352 48 bone 2.916 6.3 3.4 -17.5 13.1<br />

4 14 2009 G 355 25 hair 0.915 7.0 14.8 -16.1 45.0<br />

4 14 2009 G 357 12 bone 2.861 5.2 3.7 -17.4 11.8<br />

4 15 2009 S 360 30 hair 1.111 4.7 13.8 -18.2 42.9<br />

4 15 2009 S 360 30 bone 2.92 4.8 3.3 -17.4 12.5<br />

4 15 2009 S 361 30 hair 1.016 2.1 14.2 -19.4 43.0<br />

4 15 2009 S 361 30 bone 2.834 1.8 4.5 -14.6 14.7<br />

4 15 2009 S 363 49 hair 0.997 4.3 14.1 -21.2 44.2


Month Day Year <strong>Habitat</strong> Lab_ID Individual Component Wt (mg)<br />

15<br />

N N%<br />

13<br />

C<br />

173<br />

C%<br />

4 15 2009 S 363 49 bone 3.198 3.8 3.2 -20.1 11.3<br />

4 15 2009 S 364 30 hair 1.076 5.6 13.7 -22.2 42.8<br />

4 15 2009 S 364 30 bone 3.077 5.6 3.2 -18.5 10.9<br />

4 15 2009 S 366 44 hair 1.11 5.2 13.9 -19.2 43.6<br />

4 15 2009 S 366 44 bone 3.144 5.3 2.9 -17.1 10.3<br />

4 15 2009 G 367 45 hair 0.869 4.5 14.0 -16.4 43.8<br />

4 15 2009 G 367 45 bone 2.995 3.9 3.3 -12.2 11.2<br />

4 15 2009 G 368 14 hair 0.847 7.5 14.0 -17.4 43.2<br />

4 15 2009 G 368 14 bone 2.884 5.7 3.6 -15.5 13.0<br />

4 15 2009 G 371 31 hair 0.867 5.8 13.4 -14.3 41.7<br />

4 15 2009 G 371 31 bone 3.06 4.5 2.6 -15.2 8.6<br />

4 15 2009 G 372 50 hair_rep 1.154 3.6 13.8 -21.6 43.2<br />

4 15 2009 G 372 50 hair 0.858 3.9 13.4 -20.9 43.8<br />

4 16 2009 S 382 52 hair 0.831 6.9 14.5 -12.5 42.8<br />

4 16 2009 S 387 55 hair 0.958 4.7 13.8 -21.4 41.7<br />

4 16 2009 S 387 55 bone 2.094 3.8 3.3 -18.6 12.2<br />

4 16 2009 S 388 55 hair 0.939 6.1 14.6 -17.0 42.9<br />

4 16 2009 S 388 55 bone 3.183 9.7 3.6 -15.1 11.4<br />

4 16 2009 S 393 55 hair 1.168 7.0 14.8 -16.2 43.1<br />

4 16 2009 S 393 55 bone 3.2 5.6 3.5 -16.6 12.8<br />

4 16 2009 S 394 55 hair 1.136 4.8 14.8 -22.4 43.8<br />

4 16 2009 S 394 55 bone 3.162 5.5 3.7 -17.3 12.3<br />

4 17 2009 G 396 56 bone 1.964 10.6 6.6 -20.2 21.3<br />

4 17 2009 G 397 27 hair 0.996 6.0 13.9 -15.5 41.2<br />

4 17 2009 G 397 27 bone 2.02 5.3 4.2 -13.8 13.9<br />

4 17 2009 G 398 57 hair 1.147 6.1 13.3 -15.0 43.0<br />

4 17 2009 G 398 57 bone 1.933 7.7 4.1 -16.3 15.5<br />

4 17 2009 G 399 27 bone 3.106 6.4 3.5 -18.1 12.4<br />

4 17 2009 G 402 1 hair 0.957 5.8 13.0 -16.2 39.4<br />

4 17 2009 G 402 1 bone 2.915 5.7 3.4 -13.1 11.7<br />

4 17 2009 G 404 6 bone 2.167 6.1 3.6 -17.9 13.0<br />

4 20 2009 G 407 1 hair 0.816 6.5 14.6 -14.8 43.9<br />

4 20 2009 G 407 1 bone 2.883 4.6 3.2 -17.0 11.6<br />

4 20 2009 G 408 1 hair 1.111 7.6 14.7 -16.4 43.1<br />

4 20 2009 G 408 1 bone 3.071 6.7 2.4 -17.4 9.0<br />

4 20 2009 G 414 1 hair 1.019 4.6 14.0 -16.2 42.0<br />

4 20 2009 G 414 1 bone 2.981 4.3 4.1 -15.5 13.2<br />

4 20 2009 G 415 13 bone 2.091 3.9 2.7 -17.8 10.2<br />

4 20 2009 G 416 28 hair 0.817 4.2 14.3 -14.9 41.7<br />

4 20 2009 G 416 28 bone 1.964 3.3 3.8 -19.0 14.3<br />

4 20 2009 G 417 1 hair 0.967 5.7 14.2 -15.0 43.7<br />

4 20 2009 G 417 1 hair_rep 0.997 5.7 14.1 -15.3 44.0<br />

4 20 2009 G 417 1 bone 2.899 5.3 3.1 -15.9 10.2<br />

4 20 2009 S 422 35 hair 1.052 6.7 13.7 -19.0 42.8<br />

4 20 2009 S 422 35 bone 2.216 7.2 4.1 -16.8 14.3<br />

4 20 2009 S 423 58 bone 2.956 6.1 3.5 -22.4 12.9<br />

4 20 2009 S 425 59 bone 2.267 3.7 3.2 -14.8 10.7<br />

4 20 2009 S 427 60 bone 1.386 5.5 3.5 -20.7 12.2<br />

4 20 2009 S 429 36 hair 1.181 3.2 14.0 -16.1 43.2


Month Day Year <strong>Habitat</strong> Lab_ID Individual Component Wt (mg)<br />

15<br />

N N%<br />

13<br />

C<br />

174<br />

C%<br />

4 20 2009 S 429 36 bone 1.994 2.8 3.7 -18.3 14.5<br />

4 20 2009 S 430 61 bone 2.889 3.7 3.6 -14.7 12.0<br />

4 21 2009 S 434 37 hair 1.123 5.5 13.4 -21.3 41.9<br />

4 21 2009 S 434 37 bone 2.044 6.4 4.4 -14.9 14.4<br />

4 21 2009 S 436 37 bone 2.921 7.8 2.6 -16.9 8.2<br />

4 21 2009 G 440 62 hair 1.192 5.7 13.7 -23.1 43.2<br />

4 22 2009 G 447 2 bone 3.124 5.0 2.8 -16.3 9.6<br />

4 22 2009 G 447 2 bone_rep 3.119 5.0 2.5 -16.7 8.8<br />

4 22 2009 G 448 2 bone 3.153 3.0 3.4 -15.4 11.0<br />

4 22 2009 S 459 40 hair 1.068 6.0 14.6 -17.8 42.8<br />

4 22 2009 S 459 40 bone_rep 2.096 4.7 2.8 -21.0 8.5<br />

4 22 2009 S 459 40 bone 2.152 4.6 2.6 -21.3 9.0<br />

4 22 2009 S 467 15 hair 0.893 5.4 13.2 -16.7 43.2<br />

4 22 2009 S 467 15 bone 2.851 6.2 3.1 -16.0 10.2<br />

4 22 2009 S 468 15 hair 1.016 5.3 13.7 -16.2 43.6<br />

4 23 2009 S 473 7 bone 1.916 6.3 3.0 -17.2 9.3<br />

4 23 2009 S 475 43 bone 3.171 4.8 3.7 -19.8 12.3<br />

4 23 2009 S 478 15 hair 0.916 6.0 14.0 -15.8 42.6<br />

4 23 2009 S 478 15 bone 2.921 5.6 3.3 -16.1 10.8<br />

4 23 2009 S 479 43 bone 2.929 3.7 4.8 -17.2 15.7<br />

4 23 2009 S 481 15 hair 1.163 7.2 13.3 -14.7 43.2<br />

4 23 2009 S 481 15 bone 2.91 7.1 3.8 -14.9 12.6<br />

4 23 2009 G 483 16 hair 0.934 5.1 15.0 -17.5 44.6<br />

4 23 2009 G 483 16 bone 2.985 5.9 2.0 -14.5 6.5<br />

4 23 2009 G 484 16 hair 1.178 6.3 14.8 -16.6 44.0<br />

4 23 2009 G 484 16 bone 3.052 6.3 3.5 -13.9 11.4<br />

4 23 2009 G 485 22 hair 0.845 9.4 13.0 -16.2 44.1<br />

4 23 2009 G 485 22 bone 3.165 9.4 3.3 -17.0 11.6<br />

4 23 2009 G 489 23 hair 0.843 8.5 14.0 -15.8 44.8<br />

4 23 2009 G 490 21 bone 2.949 2.8 3.3 -16.5 11.6<br />

4 24 2009 G 497 26 bone 1.389 9.1 4.3 -14.4 15.7<br />

4 24 2009 S 502 30 hair 0.837 3.0 14.9 -19.5 43.6<br />

4 24 2009 S 502 30 bone 2.988 2.4 4.1 -19.2 15.9<br />

4 28 2009 G 506 14 hair 0.995 10.7 13.7 -17.1 43.5<br />

4 28 2009 G 507 14 hair 1.128 7.4 14.2 -14.0 42.5<br />

4 28 2009 G 507 14 bone 0.968 5.9 4.0 -12.5 12.9<br />

4 28 2009 G 517 10 bone 3.151 4.4 3.1 -18.6 10.9<br />

4 28 2009 G 521 10 bone 2.987 3.9 3.8 -18.5 14.9<br />

4 29 2009 S 528 10 bone 3.168 4.1 3.7 -18.3 13.6<br />

4 29 2009 S 530 65 hair 1.09 6.5 12.6 -16.8 43.3<br />

4 29 2009 S 530 65 bone_rep 3.16 5.6 3.8 -16.1 13.6<br />

4 29 2009 S 530 65 bone 2.951 5.1 3.6 -16.1 13.2<br />

4 29 2009 S 537 66 bone 2.904 6.7 2.3 -15.8 7.5<br />

4 29 2009 S 542 55 hair 1.079 5.9 14.5 -16.1 43.8<br />

4 29 2009 S 542 55 bone 2.817 5.8 4.2 -17.8 16.7<br />

4 29 2009 S 543 33 hair 1.184 2.8 14.7 -16.5 43.3<br />

4 29 2009 S 543 33 bone 3.08 3.3 3.5 -15.1 11.4<br />

4 29 2009 S 545 55 hair 0.918 3.7 14.6 -17.9 43.3<br />

4 29 2009 S 545 55 bone 3.08 4.3 4.0 -17.9 13.9


Month Day Year <strong>Habitat</strong> Lab_ID Individual Component Wt (mg)<br />

15<br />

N N%<br />

13<br />

C<br />

175<br />

C%<br />

4 29 2009 S 545 55 bone_rep 2.916 4.5 4.0 -18.3 14.2<br />

5 6 2009 G 569 27 bone 3.094 3.9 3.1 -16.9 10.3<br />

5 6 2009 G 571 1 hair 1.061 6.0 14.7 -17.4 44.1<br />

5 6 2009 G 571 1 bone 2.823 8.0 1.7 -16.1 5.4<br />

5 6 2009 G 573 67 bone 3.087 5.3 3.7 -18.4 12.3<br />

5 6 2009 G 576 27 bone 3.109 4.9 3.3 -16.9 11.3<br />

5 7 2009 S 589 38 hair 1.036 4.5 15.2 -17.4 42.4<br />

5 7 2009 S 589 38 bone 2.955 6.5 2.8 -16.2 9.2<br />

5 7 2009 S 592 70 hair 1.225 4.9 13.3 -18.8 43.8<br />

5 7 2009 S 593 38 hair 1.105 4.8 14.6 -18.4 44.4<br />

5 7 2009 S 593 38 bone 2.888 4.5 3.3 -16.6 11.8<br />

5 7 2009 S 595 38 hair 0.887 5.8 12.7 -18.5 47.7<br />

5 7 2009 S 595 38 bone 2.97 6.2 4.9 -18.9 22.3<br />

5 7 2009 S 596 34 hair 1.209 9.9 14.4 -17.5 43.9<br />

5 7 2009 S 596 34 bone 1.875 9.9 3.5 -17.0 11.7<br />

5 8 2009 G 605 71 bone 3.046 4.1 3.9 -20.4 13.3<br />

7 6 2009 G 612 46 hair 0.894 7.8 14.6 -15.2 44.2<br />

7 6 2009 G 612 46 hair_rep 0.812 7.6 14.1 -15.4 42.5<br />

7 6 2009 G 612 46 bone 3.177 8.5 2.5 -15.1 8.3<br />

7 6 2009 S 614 72 hair 1.155 4.5 14.5 -18.2 42.5<br />

7 6 2009 S 614 72 bone 2.931 4.4 3.3 -19.2 11.0<br />

7 6 2009 S 615 73 hair 1.144 3.9 15.0 -16.2 44.0<br />

7 6 2009 S 615 73 bone 3.186 4.3 3.5 -16.6 11.8<br />

7 6 2009 S 616 43 hair 0.131 9.5 12.4 -13.1 37.2<br />

7 6 2009 S 616 43 bone 1.426 9.5 3.4 -15.3 13.0<br />

7 7 2009 S 620 15 hair 1.03 5.5 14.8 -17.1 43.7<br />

7 7 2009 S 620 15 bone 3.081 6.9 3.9 -20.5 14.0<br />

7 7 2009 S 621 43 bone_rep 2.988 4.6 3.4 -16.3 10.9<br />

7 7 2009 S 621 43 bone 3.184 4.6 3.3 -16.4 10.9<br />

7 7 2009 G 626 16 hair 0.888 5.0 13.8 -16.2 42.7<br />

7 7 2009 G 626 16 bone_rep 3.035 5.2 3.7 -18.3 14.7<br />

7 7 2009 G 626 16 bone 3.021 5.3 3.1 -19.3 14.8<br />

7 7 2009 G 628 23 bone 2.945 4.6 3.5 -16.0 11.7<br />

7 7 2009 G 629 25 hair 0.844 5.6 13.2 -19.8 40.5<br />

7 7 2009 G 629 25 bone 3.058 4.9 3.1 -15.1 9.8<br />

7 8 2009 G 631 57 bone 0.634 10.2 2.7 -13.5 7.7<br />

7 8 2009 G 631 57 hair 1.026 6.7 14.3 -18.2 44.0<br />

7 13 2009 G 641 14 hair 0.84 8.2 13.6 -19.7 44.6<br />

7 13 2009 G 641 14 bone 2.906 7.8 3.6 -14.9 11.7<br />

7 13 2009 G 643 50 hair 0.835 4.9 14.2 -21.9 43.8<br />

7 13 2009 G 643 50 bone 3.19 2.8 3.0 -20.5 10.1<br />

7 13 2009 G 646 31 hair 0.932 10.6 13.5 -16.0 42.8<br />

7 13 2009 G 646 31 bone 3.008 3.7 3.3 -19.1 11.5<br />

7 14 2009 S 657 65 hair 0.835 5.6 14.3 -14.8 45.1<br />

7 14 2009 S 657 65 bone 3.079 6.2 3.5 -14.2 12.7<br />

7 14 2009 S 660 52 hair 0.994 6.1 14.8 -16.2 43.9<br />

7 14 2009 S 660 52 bone 3.144 6.6 2.6 -17.4 8.8<br />

7 15 2009 S 666 55 hair 1.244 4.4 14.3 -20.5 44.0<br />

7 15 2009 S 666 55 bone 3.137 3.1 3.3 -20.0 10.7


Month Day Year <strong>Habitat</strong> Lab_ID Individual Component Wt (mg)<br />

15<br />

N N%<br />

13<br />

C<br />

176<br />

C%<br />

7 15 2009 S 670 37 hair 0.933 7.2 14.3 -15.8 42.6<br />

7 15 2009 S 670 37 bone 3.049 7.4 4.1 -17.1 15.1<br />

7 15 2009 S 673 35 hair 0.854 4.1 12.9 -17.8 42.5<br />

7 15 2009 S 673 35 bone 2.822 4.6 4.3 -18.4 17.1<br />

7 16 2009 S 686 36 hair 0.395 4.3 13.4 -15.0 41.4<br />

7 16 2009 S 686 36 bone 2.953 3.5 3.5 -15.0 11.4<br />

7 17 2009 G 691 76 bone 3.162 4.0 4.3 -17.6 16.1<br />

7 20 2009 G 693 1 hair 0.908 4.4 15.1 -18.6 44.4<br />

7 20 2009 G 693 1 bone 3.091 5.2 2.3 -17.0 7.7<br />

7 21 2009 S 705 40 bone 3.088 7.0 3.3 -18.3 10.6<br />

7 21 2009 G 719 17 hair 1.054 4.8 14.2 -15.7 42.5<br />

7 21 2009 G 719 17 bone 3.148 5.4 2.3 -18.5 8.0<br />

7 22 2009 G 721 17 bone 3.032 5.0 2.0 -13.1 6.6<br />

7 23 2009 G 723 27 hair 0.851 6.2 14.0 -16.6 44.4<br />

7 23 2009 G 723 27 bone 2.926 4.3 3.4 -16.1 11.6<br />

7 27 2009 G 734 45 hair 0.822 7.0 14.2 -21.5 44.4<br />

7 27 2009 G 734 45 bone 3.167 7.2 2.9 -19.5 9.6<br />

7 27 2009 G 747 10 bone 3.18 5.4 3.6 -18.2 11.3<br />

7 28 2009 S 755 77 bone 3.168 7.2 3.5 -17.8 11.8<br />

7 29 2009 S 759 33 hair 0.809 4.4 14.0 -16.0 43.0<br />

7 29 2009 S 759 33 bone 3.243 1.6 2.9 -15.6 9.5<br />

7 30 2009 S 767 34 hair_rep 0.923 5.0 14.1 -16.2 41.2<br />

7 30 2009 S 767 34 hair 0.876 4.9 13.8 -16.1 41.9<br />

7 30 2009 S 767 34 bone 3.24 4.1 3.1 -19.7 10.3<br />

7 31 2009 S 769 37 bone 2.839 5.0 3.7 -17.0 12.2<br />

10 13 2009 G 770 39 bone 2.82 5.5 4.4 -18.2 14.9<br />

10 13 2009 G 773 3 hair 1.163 4.6 14.1 -15.8 44.6<br />

10 13 2009 G 773 3 bone 2.948 4.2 3.0 -15.9 11.1<br />

10 13 2009 G 773 3 bone_rep 2.933 4.4 2.4 -16.0 8.9<br />

10 13 2009 G 775 29 hair 0.86 6.2 14.0 -16.9 43.7<br />

10 14 2009 S 779 5 bone 3.103 5.3 3.4 -13.6 11.7<br />

10 14 2009 S 780 73 hair_rep 0.97 4.5 14.2 -20.3 44.0<br />

10 14 2009 S 780 73 hair 1.193 4.3 13.9 -21.2 44.0<br />

10 14 2009 S 783 5 bone 2.92 6.3 3.6 -16.3 13.4<br />

10 14 2009 S 784 43 bone 3.05 6.9 3.8 -14.6 13.3<br />

10 14 2009 S 785 15 hair 0.924 4.5 14.0 -14.4 42.8<br />

10 14 2009 S 785 15 bone 2.945 6.2 3.7 -13.6 12.1<br />

10 14 2009 G 792 16 hair 1.05 3.9 14.4 -13.1 43.2<br />

10 14 2009 G 792 16 bone 2.945 3.4 3.4 -13.0 11.1<br />

10 17 2009 G 802 57 bone 3.109 4.3 4.6 -18.7 14.1<br />

10 17 2009 G 802 57 hair 1.023 4.6 14.0 -21.7 43.7<br />

10 17 2009 G 806 17 bone 3.063 3.4 3.5 -15.2 13.6<br />

10 19 2009 G 808 27 hair 1.219 7.6 14.1 -19.0 43.6<br />

10 19 2009 G 808 27 bone 3.092 7.1 5.1 -16.8 16.4<br />

10 19 2009 G 809 27 bone 0.182 8.8 5.5 -17.8 17.5<br />

10 19 2009 G 810 28 hair 0.722 6.3 13.5 -16.5 44.0<br />

10 19 2009 G 810 28 bone 3.127 9.3 1.6 -12.7 28.5<br />

10 19 2009 G 815 27 bone 2.82 7.3 4.7 -16.1 14.7<br />

10 19 2009 S 816 30 hair 0.816 5.3 14.3 -20.3 43.7


Month Day Year <strong>Habitat</strong> Lab_ID Individual Component Wt (mg)<br />

15<br />

N N%<br />

13<br />

C<br />

177<br />

C%<br />

10 22 2009 S 829 35 hair 1.015 4.8 13.7 -22.0 42.1<br />

10 22 2009 S 829 35 bone 2.949 5.1 5.0 -21.0 15.7<br />

10 22 2009 S 829 35 bone_rep 2.956 4.7 4.5 -20.9 14.7<br />

10 22 2009 S 830 36 hair 1.172 4.2 14.5 -16.9 43.4<br />

10 22 2009 S 830 36 bone 3.159 3.5 3.8 -17.6 14.8<br />

10 23 2009 S 833 34 hair 0.707 8.2 14.3 -20.8 41.8<br />

10 23 2009 S 833 34 bone 3.121 8.4 3.2 -19.7 10.6<br />

10 23 2009 S 834 38 bone 2.899 6.2 2.7 -17.5 9.3<br />

10 23 2009 S 835 37 hair 1.02 6.6 14.0 -15.4 42.9<br />

10 23 2009 S 835 37 bone 2.903 6.5 4.8 -13.8 14.8<br />

10 23 2009 S 839 37 bone 3.02 3.4 3.0 -19.0 9.5<br />

10 23 2009 G 840 62 hair 1.033 5.2 14.2 -17.4 43.1<br />

10 23 2009 G 840 62 bone 3.101 4.3 3.7 -17.5 12.4<br />

10 26 2009 S 845 10 bone 3.139 4.9 2.7 -20.0 9.5<br />

10 26 2009 S 857 55 hair 1.047 4.8 13.9 -17.6 41.6<br />

10 26 2009 S 857 55 bone 3.075 4.3 3.7 -17.9 15.5<br />

10 27 2009 S 863 43 bone 3.105 6.1 4.0 -14.0 13.0<br />

10 27 2009 S 867 43 bone 2.975 4.5 3.2 -17.4 11.8<br />

10 27 2009 S 871 41 hair 1.141 5.8 14.2 -18.1 44.3<br />

10 27 2009 S 871 41 bone 3.01 2.9 3.5 -19.3 14.3<br />

10 28 2009 S 873 43 bone 3.176 4.8 3.1 -16.1 10.7<br />

10 28 2009 S 874 43 bone 3.111 4.4 3.3 -16.7 11.1<br />

10 28 2009 S 875 43 bone 3.138 4.0 3.1 -16.8 10.4<br />

10 28 2009 S 876 72 hair 1.024 5.2 14.5 -16.3 45.2<br />

10 28 2009 S 876 72 bone 2.837 5.1 4.4 -14.7 14.1<br />

10 29 2009 G 892 24 hair 0.864 6.2 14.5 -14.9 43.0<br />

10 29 2009 G 892 24 bone 3.133 5.2 4.0 -16.5 12.9<br />

10 30 2009 S 899 74 hair 1.103 6.4 14.4 -22.5 43.9<br />

10 30 2009 S 899 74 bone 2.919 3.1 3.0 -16.8 10.9<br />

10 30 2009 S 903 70 hair 0.96 3.7 14.1 -17.8 43.7<br />

10 30 2009 S 903 70 bone 3.023 2.4 3.7 -18.6 14.0<br />

11 2 2009 G 910 48 hair 1.168 7.7 14.4 -11.7 44.4<br />

11 2 2009 G 910 48 bone 2.804 7.7 3.9 -10.9 12.6<br />

11 3 2009 G 915 27 bone_rep 2.924 4.9 4.6 -14.2 14.2<br />

11 3 2009 G 915 27 bone 3.072 5.0 4.5 -14.3 14.3<br />

11 3 2009 S 916 30 bone 2.986 3.2 3.8 -16.2 13.6<br />

11 3 2009 S 917 44 hair 1.144 5.8 14.1 -16.2 43.7<br />

11 3 2009 S 917 44 bone 3.148 6.3 3.8 -14.2 12.4<br />

11 4 2009 G 920 50 bone 4.078 2.0 0.5 -21.4 48.4<br />

11 6 2009 S 931 10 bone 2.829 5.0 3.0 -18.4 10.3<br />

11 6 2009 S 934 33 hair 1.224 2.1 14.4 -17.5 43.0<br />

11 6 2009 S 934 33 bone 3.093 2.4 3.6 -18.9 13.5<br />

Appendix 3. δ 13 C and δ 15 N values <strong>fo</strong>r hair and bone samples taken from coyote scats<br />

collected at the <strong>Sevilleta</strong> National Wildlife Refuge. Each sample has a unique lab<br />

identification number (Lab_ID) and each individual has a unique individual identification<br />

number (Individual). G = grassland habitat, S = shrubland habitat, rep = replicate sample<br />

of pieces of hair or bone taken from a single scat sample. Month, day and year<br />

correspond to the date that the scat sample was collected in the field.


Appendix 4. Ca<strong>rb</strong>on and nitrogen isotope data <strong>fo</strong>r vegetation samples<br />

Month Day Year Identifier Weight (mg)<br />

15<br />

N N%<br />

13<br />

C C%<br />

4 19 2009 Aristida 4.153 0.18 0.87 -13.51 44.12<br />

7 12 2009 Aristida 3.817 1.70 1.36 -14.36 44.19<br />

10 16 2009 Aristida 3.806 1.05 1.27 -14.39 44.70<br />

4 25 2009 ATCA1 4.012 7.30 2.24 -14.85 41.89<br />

7 19 2009 ATCA1 4.126 6.85 3.20 -16.50 41.19<br />

10 24 2009 ATCA2 4.057 4.47 1.11 -14.97 42.10<br />

4 4 2009 Bouteloua 4.118 0.36 0.90 -14.95 46.26<br />

7 11 2009 Bouteloua 4.19 -0.12 0.76 -15.03 44.22<br />

10 11 2009 Bouteloua 4.11 -1.99 1.45 -14.46 44.61<br />

4 4 2009 Bouteloua_rep 4.066 -0.06 0.73 -14.99 46.28<br />

10 16 2009 CYIMI1_F 3.899 5.58 2.79 -12.05 50.86<br />

10 16 2009 CYIMI1_F 3.858 4.61 2.22 -11.91 49.74<br />

4 12 2009 Dasyochloa 4.042 2.44 1.24 -14.79 42.69<br />

7 18 2009 Dasyochloa 3.971 2.14 1.67 -15.37 43.63<br />

10 10 2009 Dasyochloa 3.985 2.04 1.89 -15.07 42.55<br />

7 18 2009 EPTO1 4.157 2.34 1.49 -25.33 47.28<br />

10 24 2009 EPTO2 4.135 1.06 0.91 -25.01 46.34<br />

7 11 2009 Fo<strong>rb</strong>1 4.044 2.49 2.84 -26.10 42.20<br />

10 11 2009 Fo<strong>rb</strong>1 3.809 3.83 2.34 -20.62 41.32<br />

7 11 2009 Fo<strong>rb</strong>2 4.009 1.39 1.71 -26.45 45.81<br />

10 11 2009 Fo<strong>rb</strong>2 3.969 0.82 1.56 -26.31 44.51<br />

10 11 2009 Fo<strong>rb</strong>3 3.94 2.25 3.36 -26.80 41.32<br />

10 11 2009 Fo<strong>rb</strong>3_rep 4.139 2.18 3.22 -26.34 40.97<br />

7 19 2009 Fo<strong>rb</strong>4 3.912 6.18 4.05 -14.02 38.24<br />

10 24 2009 Fo<strong>rb</strong>5 4.184 -1.48 2.11 -26.32 42.93<br />

4 25 2009 Fo<strong>rb</strong>6 3.951 0.43 1.39 -26.21 45.05<br />

4 25 2009 Fo<strong>rb</strong>6_rep 3.876 0.40 1.33 -25.76 44.01<br />

4 19 2009 Fo<strong>rb</strong>7 3.81 0.55 1.70 -27.37 43.07<br />

7 25 2009 Fo<strong>rb</strong>7 4.092 -0.03 2.49 -26.65 44.79<br />

10 31 2009 Fo<strong>rb</strong>7 3.869 0.86 2.73 -27.40 42.07<br />

4 19 2009 Fo<strong>rb</strong>8 3.947 2.52 2.39 -25.32 42.64<br />

4 18 2009 Fo<strong>rb</strong>9 4.156 -0.33 2.43 -26.94 43.45<br />

7 18 2009 Fo<strong>rb</strong>10 3.939 0.68 2.67 -26.89 46.08<br />

4 26 2009 Fo<strong>rb</strong>11 3.878 2.42 2.89 -25.10 39.68<br />

7 26 2009 Fo<strong>rb</strong>11 4.124 1.52 3.40 -25.96 41.07<br />

10 25 2009 Fo<strong>rb</strong>12 3.818 -0.73 3.19 -27.11 44.81<br />

4 25 2009 GUSA 4.196 0.61 1.97 -25.87 47.71<br />

7 12 2009 GUSA 4.035 0.10 1.20 -25.65 46.69<br />

10 25 2009 GUSA 4.035 1.32 2.16 -25.15 49.79<br />

10 10 2009 JUMO 3.969 -1.12 1.20 -23.61 55.79<br />

10 10 2009 JUMO_F 3.999 -0.26 1.89 -20.25 52.02<br />

10 10 2009 JUMO_rep 3.838 -1.15 1.23 -23.43 53.52<br />

4 19 2009 KRLA 3.836 5.42 2.14 -25.30 42.28<br />

7 12 2009 KRLA 4.073 4.96 2.05 -25.34 42.17<br />

10 16 2009 KRLA 4.117 5.87 2.27 -24.20 42.46<br />

4 25 2009 LATR 4.101 4.19 1.64 -24.89 48.40<br />

7 19 2009 LATR 4.09 5.87 1.79 -24.36 48.92<br />

178


Month Day Year Identifier Weight (mg)<br />

15<br />

N N%<br />

13<br />

C C%<br />

10 24 2009 LATR 3.936 4.79 1.72 -25.10 49.00<br />

4 19 2009 Pleuraphis 3.802 1.83 0.70 -14.43 44.31<br />

7 25 2009 Pleuraphis 3.95 1.45 1.18 -15.02 43.29<br />

10 31 2009 Pleuraphis 4.147 0.81 1.46 -15.10 42.85<br />

10 24 2009 PRGL1_F 4.026 6.34 7.88 -23.98 44.20<br />

4 25 2009 PRGL2 4.125 1.27 2.88 -24.30 46.39<br />

7 19 2009 PRGL2 3.873 1.09 2.77 -24.72 46.65<br />

10 24 2009 PRGL2 4.121 1.57 2.52 -25.21 46.30<br />

7 19 2009 PRGL2_F 4.113 3.35 4.30 -23.19 44.76<br />

10 24 2009 PRGL2_F 3.911 3.81 4.75 -23.96 45.25<br />

10 10 2009 PRGL3_F 3.952 3.23 3.74 -25.69 44.60<br />

4 18 2009 Scleropogon 3.902 5.36 1.57 -14.74 43.19<br />

7 25 2009 Scleropogon 4.078 5.48 1.89 -14.63 43.59<br />

10 31 2009 Scleropogon 3.833 5.15 2.32 -14.94 42.71<br />

4 25 2009 Sporobolus 3.801 0.27 0.75 -13.90 42.83<br />

7 19 2009 Sporobolus 4.175 2.40 1.33 -15.27 43.15<br />

10 24 2009 Sporobolus 3.907 1.43 1.84 -15.62 42.71<br />

4 4 2009 YUGL 3.833 2.28 1.68 -24.52 50.56<br />

7 11 2009 YUGL 4.152 2.24 1.37 -26.87 51.30<br />

10 11 2009 YUGL 4.021 1.43 1.58 -25.63 51.57<br />

Appendix 4. δ 13 C and δ 15 N values <strong>fo</strong>r vegetation samples collected at the <strong>Sevilleta</strong><br />

National Wildlife Refuge. Month, day, and year correspond to dates that the samples<br />

were collected in the field. Identifiers are species codes <strong>fo</strong>r woody plants, genera names<br />

<strong>fo</strong>r grasses, functional <strong>type</strong> <strong>fo</strong>r <strong>fo</strong><strong>rb</strong>s. F = a sample of a plant that coyotes eat (<strong>fo</strong>od) and<br />

that was cleaned with ethanol prior to being prepared <strong>fo</strong>r isotope analysis; rep = replicate<br />

sample; different numbers within a given identifier (e.g., Fo<strong>rb</strong>1, Fo<strong>rb</strong>2, etc.) indicate<br />

samples that were collected at different locations. ATCA = Atriplex canescens, CYIMI =<br />

Cylindropuntia imbricata, EPTO = Ephedra torreyana, GUSA = Gutierrezia sarothrae,<br />

JUMO = Juniperus monosperma, KRLA = Krascheninnikovia lanata, LATR = Larrea<br />

tridentata, PRGL = Prosopis glandulosa, YUGL = Yucca glauca.<br />

179

Hooray! Your file is uploaded and ready to be published.

Saved successfully!

Ooh no, something went wrong!