Human Ecological Integration in Subarctic Eastern

Item Type text; Electronic Dissertation

Authors Lanoe, Francois

Publisher The University of Arizona.

Rights Copyright © is held by the author. Digital access to this material is made possible by the University Libraries, University of Arizona. Further transmission, reproduction or presentation (such as public display or performance) of protected items is prohibited except with permission of the author.

Download date 29/09/2021 22:48:17

Link to Item http://hdl.handle.net/10150/624542 HUMAN ECOLOGICAL INTEGRATION IN SUBARCTIC EASTERN BERINGIA

By

François B. Lanoë

______Copyright © François Lanoë 2017

A Dissertation Submitted to the Faculty of the

SCHOOL OF ANTHROPOLOGY

In Partial Fulfillment of the Requirements

For the Degree of

DOCTOR OF PHILOSOPHY

In the Graduate College

THE UNIVERSITY OF ARIZONA

2017

1

THE UNIVERSITY OF ARIZONA GRADUATE COLLEGE As members of the Dissertation Committee, we certify that we have read the dissertation prepared by François B. Lanoë titled “Human Ecological Integration in Subarctic Eastern Beringia” and recommend that it be accepted as fulfilling the dissertation requirement for the Degree of Doctor of Philosophy.

Date: April 10th, 2017 Vance T. Holliday

Date: April 10th, 2017 Steven L. Kuhn

Date: April 10th, 2017 John W. Olsen

Date: April 10th, 2017 Joshua D. Reuther

Date: April 10th, 2017 Mary C. Stiner

Final approval and acceptance of this dissertation is contingent upon the candidate's submission of the final copies of the dissertation to the Graduate College. I hereby certify that I have read this dissertation prepared under my direction and recommend that it be accepted as fulfilling the dissertation requirement.

Date: April 10th, 2017 Dissertation Director: Joshua D. Reuther

Date: April 10th, 2017 Dissertation Director: Mary C. Stiner

2

STATEMENT BY AUTHOR

This dissertation has been submitted in partial fulfillment of the requirements for an advanced degree at the University of Arizona and is deposited in the University Library to be made available to borrowers under rules of the Library. Brief quotations from this dissertation are allowable without special permission, provided that an accurate acknowledgment of the source is made. Requests for permission for extended quotation from or reproduction of this manuscript in whole or in part may be granted by the head of the major department or the Dean of the Graduate College when in his or her judgment the proposed use of the material is in the interests of scholarship. In all other instances, however, permission must be obtained from the author.

SIGNED: François B. Lanoë

3

ACKNOWLEDGMENTS My dissertation co-advisors, Mary Stiner and Joshua Reuther, have been instrumental in this work and have provided invaluable guidance throughout the different stages of my graduate studies. I am also grateful to the other members of my doctoral committee, Vance Holliday, Steven Kuhn, and John Olsen, who have supported and helped me in many of my academic endeavors. Many people at the School of Anthropology of the University of Arizona have supported and assisted me. Veronica Peralta and Catherine Lehman spent a lot of time helping me understand some of the administrative aspects of being a Ph.D. student in Tucson, as did Diane Austin, Barbara Mills, Ben Beshaw, Norma Maynard, Scott Ellegood, and Rayshma Pereira. Archaeological fieldwork in was conducted in close collaboration with Charles Holmes and Jennifer Kielhofer. Several volunteers gracefully donated their time over the years, including Matthew Pailes, Amy Clark, Meredith Wismer, Ismael Sanchez, Benjamin Bellorado, and David Plaskett. Fieldwork was made possible by the generosity of Angela Linn, Grey Reuther, and Barbara Crass. I also enjoyed and learned a lot from time spent and conversations with Ben Potter, David Yesner, Jason Rogers, Evgenia Anichtchenko, Richard Vanderhoek, Bryan Wygal, Ted Goebel, Kathryn Krasinski, Cassidy Phillips, Holly McKinney, Norman Easton, and Carrin Halffman. Access to archaeological and paleontological collections for sampling was granted by Patrick Druckenmiller and Dale Guthrie (Department of Earth Sciences, University of Alaska Museum of the North), Ben Potter (University of Alaska Fairbanks), Jeff Rasic and Jillian Richie (National Park Service), Norman Easton (Yukon College), Jin Meng and Lindsay Jurgielewicz (American Museum of Natural History), Greg Hare and Christian Thomas (Archaeology Program, Government of Yukon), and Greg Hodgins and Rich Cruz (University of Arizona). Access to other archaeological collections was granted by Joshua Reuther and Scott Shirar (Department of Archaeology, University of Alaska Museum of the North), David Yesner (University of Alaska Anchorage), Charles Holmes and Richard Vanderhoek (Alaska State Historic Preservation Office), and Kristian Gregersen (Natural History Museum of Denmark). Access to comparative skeletal collections was granted by Barnett Pavao-Zuckerman (Stanley J. Olsen Laboratory of Zooarchaeology, Arizona State Museum), Mary Stiner (Hunter-Gatherer Laboratory, University of Arizona), Nieves Zedeño (Bureau of Applied Research in Anthropology, University of Arizona), Peter Reinthal (University of Arizona Museum of Natural History), Everett Lindsay (Department of Geosciences, University of Arizona), Grant Zazula (Paleontology Program, Government of Yukon), and Aren Gunderson and Link Olson (Department of Mammalogy, University of Alaska Museum of the North). Collagen extraction was conducted at the University of Arizona Accelerator Mass Spectrometry Laboratory directed by Greg Hodgins. In addition to Greg, Marcus Lee, Amanda Halawani, Mitzi Demartino, Rich Cruz, and particularly Rebecca Watson assisted me while working in the laboratory. Isotope analysis was conducted at the Environmental Isotope Laboratory of the University of Arizona, by David Dettman and Xiaoyu Zhang. One sample was gracefully analyzed by ZooMs by Michael Buckley at the Proteomics Laboratory of the University of Manchester.

4

Several folks have contributed in making the School of Anthropology an incredibly enjoyable learning and working environment. I am grateful to Nieves Zedeño, who has been a wonderful person to work with, as well as to Amy Clark, Marcy Reiser, Esther Echenique, Samantha Fladd, Amanda Hilton, Benjamin Bellorado, Luminița Mandache, Paula Ugalde, Nicole Mathwich, Matthew Pailes, Danielle Phelps, Brandi Bethke, Leslie Aragon, William Reitze, Emma Blake, Eleni Hasaki, David Killick, Takeshi Inomata, Sarah Renkert, Daniela Triadan, Anna Jansson, Ashleigh Thompson, William White, Arthur Jelinek, Ismael Sanchez, Megan Sheehan… many others. I have learned much of what I know about archaeology from interacting, in the past decade, with people in the field and in laboratories. In addition to people cited above, I am particularly indebted to James Enloe, Francine David, Jean-Laurent Monnier, Catherine Dupont, Jesse Ballenger, Marylène Patou-Mathis, Abdeljalil Bouzouggar, and Stéphane Péan. Lastly, friends and family have supported me emotionally over the years. I thank in particular my parents, Bertrand and Claudine Lanoë, my sister, Hélène Lanoë, and my cousins Lise Lejalé and Françoise Possémé; as well as Soumaya Belmecheri, Bethany Coulthard, Marcy Reiser, Amy Clark, Meredith Wismer, Morgann le Roch, Aurélie Manin, Julien Guilbert, and Alexis Licht. This doctoral research was funded by a Doctoral Dissertation Research Improvement Grant from the National Science Foundation (BCS-1504654), a Research and Exploration Grant from the National Geographic Society (#9708-15), the Lewis and Clark Fund for Exploration and Field Research from the American Philosophical Society, the Arizona State Museum, the Social and Behavioral Sciences Research Institute at the University of Arizona, and several grants from the Graduate and Professional Student Council and the School of Anthropology of the University of Arizona. The Department of Archaeology of the University of Alaska Museum of the North and the Department of Anthropology of the University of Alaska Fairbanks provided logistical support and excavation equipment. This research was conducted in the traditional lands of the Tanana River Athabaskan People, Selkirk First Nation, White River First Nation, and Tr’ondëk Hwëch’in First Nation.

5

To the people of Beringia, past and present.

6

CONTENTS

List of Figures ...... 10

List of Tables ...... 14

Abstract ...... 17

Introduction ...... 19

Literature Review...... 23

Paleoecology and Archaeology ...... 23

Human-Environment Interactions ...... 24

Human-Carnivore Coevolution ...... 25

Community Structure ...... 27

Economy & Mobility ...... 29

Late Glacial Beringia ...... 32

Paleoenvironments...... 32

Archaeology ...... 34

References ...... 37

Appendix A: as Raw Material in Beringia: Insights from the Site of Swan Point CZ4b,

Alaska ...... 60

Abstract ...... 61

Swan Point CZ4b: A Specialized Workshop ...... 64

7

Site Background ...... 64

Limited On-Site Megafauna Consumption for Food ...... 65

Organic Technology: Diversity of Materials and Techniques...... 67

Spatially Demarcated Activity Areas ...... 72

A Specialized Workshop ...... 77

Organic Technology in Beringia ...... 79

Supplemental Text 1: Radiocarbon Dating ...... 83

Supplemental Text 2: Patterns of Deposition and Overall Taphonomy...... 85

Supplemental Text 3: Faunal Remains...... 90

References ...... 101

Appendix B: Task-Specific Sites and Paleoindian Landscape Use in the Shaw Creek Flats of

Eastern Beringia ...... 123

Abstract ...... 124

Introduction ...... 125

Identifying the Function of Hunter-Gatherer Sites ...... 127

The Shaw Creek Flats ...... 130

Residential Sites ...... 136

Hunting Bison and Wapiti ...... 140

The Keystone Dune Site: A Wapiti Hunting Camp ...... 143

Swan Point CZ4b: A Short-Term Specialized Occupation ...... 150

8

Discussion ...... 153

Supplemental Text 1. Evidence for a Constructed Structure at Swan Point CZ4b...... 155

Supplemental Text 2. Estimation of the Length of Stay at Swan Point CZ4b...... 159

References ...... 163

Appendix C: The Lion’s Share: Human Paleoecological Integration in Subarctic Eastern

Beringia ...... 180

Abstract ...... 181

Introduction ...... 182

Background ...... 184

Approach ...... 187

Results ...... 189

Isotopic Values ...... 189

Zooarchaeological Results ...... 194

Discussion ...... 196

Herbivore Guilds ...... 196

Predator Guild...... 199

Conclusions ...... 201

Methods ...... 203

Supplemental Dataset 1: Table 15, Isotopic Values Obtained for this Study ...... 208

References ...... 210

9

LIST OF FIGURES

Figure 1. Location of Swan Point in relationship with sub-contemporaneous sites. A Digital

Terrain Model (USGS 2010) was used; contour line intervals are 200 m (top) and 5 m (bottom).

The shaded area corresponds to the area visible from the site at a distance of 10 km, at which a normal human visual acuity (one arc-minute [NDT Resource Center 2014]) can perceive a 3-m- large mammoth (d = 3/sin(1’) = 10,313 m). Thirty-nine percent of the landscape within this distance is visible from the site. At a radius of 1 km, 86.0 percent of the landscape is visible from the site...... 64

Figure 2. Shed right antler of an adult male caribou, with evidence of cleaving and percussion

(specimens #22244 and #22303). The two halves separate and were recorded about 2 m from each other (Figure 6)...... 68

Figure 3. Beveled mammoth calf tusk, medial and lateral views (specimen #20785). Its diameter varies between 23.3 and 23.5 mm for a chord length of 256.3 mm. The distal end is broken. .... 69

Figure 4. Left adult tusk from a ~15-year-old mammoth female, medial view, with evidence of technological work: percussion (above, medial view) and groove and splintering (below, ventral view) (specimen #21444)...... 70

Figure 5. Swan Point CZ4b refits: (a) microblades, microcores, and platform rejuvenation “ski” spalls; (b) burins and burin spalls. Left modified from Potter et al. (2013)...... 74

Figure 6. Spatial distribution of (a) bird and (b) remains, CZ4b. The two caribou antler pieces to the south refit. The horse cheek teeth most likely articulate together, as well as the mammoth calf cheek teeth and the mandible (see text)...... 76

Figure 7. Interpretation of the use of space at Swan Point CZ4b...... 78

10

Figure 8. Calibration of radiocarbon dates for CZ3 and CZ4. Calibration was conducted with

OxCal 4.2 and IntCal13 (Bronk Ramsey 1994, 2009, Reimer et al. 2013). Dates preceded by an asterisk are reported here for the first time; others can be found in Potter et al. (2013) and

Holmes (2011). Samples for β-190577, β-186682 and β-215328 were obtained from rodent burrows in CZ4b (see Supplemental Text 2)...... 83

Figure 9. Taphonomic processes affecting remains of Cricetid-Dipodid (in black), Sciurid (gray), and the rest of the faunal assemblage (stripes). SO: surface obstructed, no recording...... 88

Figure 10. North-south profile (facing west) of the lower layers of Swan Point. Scale is in meters, elevation is relative to the modern surface...... 89

Figure 11. Mammoth infant rib fragment, medial view (specimen #15941). Side is undetermined; the proximal side (towards the rib-head) is to the right. For comparison, a left rib (position 10 to

13, not to scale) is displayed on top. A notch characteristic of dynamic loading impact (Capaldo and Blumenschine 1994) is present on the mesio-proximal portion of the blade...... 94

Figure 12. Comparison of measurements of swan elements from Swan Point CZ4b with references for Cygnus columbianus and C. buccinator...... 99

Figure 13. Phasianidae specimens measurements from Swan Point CZ4b with references for

Lagopus lagopus, Lagopus leucurus, Tympanachus phasianellus and Dendragapus obscurus. 100

Figure 14. Shaw Creek Flats and sites mentioned in the text: (1) Swan Point, (2) Bachner, (3)

Broken Mammoth, (4) Cook, (5) Keystone Dune, (6) Mead, (7) Holzman. The red lines represent least-cost paths linking sites by the Tanana River to sites deep in the Flats. The paths follow

Shaw and Keystone Creeks...... 132

11

Figure 15. Distribution of large herbivore species (proportion of NISP) at Late Glacial occupations in the Shaw Creek Flats (computed from Yesner 2001, Potter et al. 2013, Lanoë and

Holmes 2016, Reuther et al. 2016, Lanoë et al. 2017)...... 141

Figure 16. Topographic and geologic variation in the Shaw Creek Flats. Areas in yellow have higher slope (> 10°) and are oriented to the south (135-225°) reflecting higher drainage and insolation. Hatched areas represent sand sheets deposited during the Late Glacial (data from

Weber et al. 1978). Legend as in Figure 14...... 143

Figure 17. Profile and map of the Keystone Dune site. Locations of artifacts in the 2014 test are computed from photographs...... 144

Figure 18. Taphonomic characteristics of the Keystone Dune faunal assemblage (proportion of

NISP). Left: breakage type. Right: burning stages following Stiner et al. (1995), with 0: unburned; 1-3: partly to wholly carbonized; and 4-6: partly to wholly calcined. Materials in or near the eastern hearth are in black, materials in other zones are in white. Differences between the subsamples are significant for burning (χ2 = 30.2; df = 4; p < 0.01) and breaking (χ2 = 45.6; df

= 3; p < 0.01)...... 149

Figure 19. Interpretation of the use of space at Swan Point CZ4b (reproduced from Lanoë and

Holmes [2016])...... 152

Figure 20. Ring and sector analysis, Swan Point CZ4b south and north hearths...... 158

Figure 21. Estimated minimum and maximum length of stay at CZ4b depending on group size.

Daily energy requirements, 2,100 Kcal, are the standard for an adult (World Health Organization

1985) but may have been slightly higher among mobile hunter-gatherers (Pontzer et al. 2012).

Maximum lengths of stay for one person and two people are not displayed for the sake of readability; they amount to 103 and 51 days, respectively...... 162

12

Figure 22. Known large mammal dates from subarctic eastern Beringia. Gulo presence has yet to be dated systematically, but is fairly common in Late eastern Beringian deposits and was probably present throughout the Ice Age (Youngman 1993; Tomasik and Cook 2005; Bryant

1987). Bison is continuous throughout the record but seems to undergo some population replacement (Heintzman et al. 2016). The single Puma date (Harington 2003) was excluded.

Dates are compiled from (Guthrie 2006; Barnes et al. 2002; Barnett et al. 2009; Shapiro et al.

2004; Zazula et al. 2014; Raghavan et al. 2014; Guthrie et al. 2001; Leonard et al. 2007; Fox-

Dobbs et al. 2008; Guthrie 2003; Harington 2011; Meiri et al. 2014; Weinstock et al. 2005;

Mann et al. 2013, 2015; Haile et al. 2009; Leonard et al. 2002; Sattler 1997; Endacott 2008;

Kurtén and Anderson 1980; Hundertmark et al. 2002; Stone and Cook 2000; Chambers et al.

2012; Potter 2008; Heintzman et al. 2016; Harington 2003) and calibrated with OxCal 4.2 and

IntCal 13 (Bronk Ramsey 1994; Reimer et al. 2013). Oxygen Isotopic Values are from Seierstad et al. (2014)...... 186

Figure 23. Map of the study area and geographic distribution of the samples used in the study.

Current glaciers are represented in black, and cross-hatching represents the limit of the

Cordilleran ice-sheet during the Last Glacial Maximum (following Manley and Kaufman 2002;

Geomatics Yukon 2015; Manley 2002; Dyke et al. 2003)...... 188

Figure 24. Recorded isotopic values for subarctic eastern Beringia megafauna...... 190

Figure 25. Habitat partitioning through time for large herbivores. Time intervals are defined following breaks in isotopic trends (Figure 24) and last between 2,300 and 2,700 years each. The algorithm used to create the ternary graphs allows for negative values (e.g. caribou)...... 192

Figure 26. Partitioning of the predator guild prior and after 14,300 cal B.P. Labels correspond to dietary tendencies discernable from stable isotopes...... 193

13

LIST OF TABLES

Table 1. Composition of identified spatial clusters...... 72

Table 2. Radiocarbon dates newly reported here...... 84

Table 3. Specimen identification for Swan Point CZ4b. Quantification units follow the definitions in Lyman (2008). Mammal and bird size classes are defined in Table S3...... 92

Table 4. Size classes used for this study...... 92

Table 5. Archaeological criteria used to differentiate residential and task-specific sites...... 128

Table 6. Known Late Glacial occupations in the Shaw Creek Flats. Previously unpublished date from the Cook site (AA-106445): 10,400 ± 60 14C B.P., on bone collagen, δ13C = -20.0 ‰; calibrated here with OxCal 4.2 and IntCal 13 (Bronk Ramsey 1994, 2009, Reimer et al. 2013).

...... 133

Table 7. Rates of sediment deposition during the Late Glacial at archaeological and geological localities of the Shaw Creek Flats (data computed from Dilley 1998, Potter et al. 2013, Reuther

2013, Reuther et al. 2016; rates calculated following Stein et al. [2003]). The ends of the date ranges are averages of intervals calibrated with OxCal 4.2 and IntCal 13 (Bronk Ramsey 1994,

2009, Reimer et al. 2013). For the Cook site, the date at 195 cm BS is referenced in Table 6; the date at 90 cm BS is as follows (UGAMS--18145): 8,320 ± 30 14C B.P., on bone collagen, δ13C =

-20.6 ‰, δ15N = 2.4 ‰...... 135

Table 8. Characteristics of several Late Glacial occupations in the Shaw Creek Flats. Densities were computed from Krasinski and Yesner (2008) for Broken Mammoth and from original data for Swan Point and Keystone Dune; estimated from figures and text in Potter et al. (2013) for

Mead...... 137

14

Table 9. Diversity of archaeological materials at several Late Glacial occupations in the Shaw

Creek Flats (computed from Yesner 1994, 1996, 2001, Holmes 1996, 2011, Yesner et al. 2000,

Potter et al. 2013, Goebel and Potter 2016, Lanoë and Holmes 2016, Reuther et al. 2016)...... 139

Table 10. Lithic materials recovered at the Keystone Dune Site. Specimens recorded near the hearth are significantly smaller than surface finds below the excavation block (χ2 = 80.9; df = 4; p < 0.01) or in other areas of the dune (χ2 = 131.1; df = 4; p < 0.01)...... 146

Table 11. Faunal materials recovered at the Keystone Dune Site. Levels of taxonomic identification are significantly higher (χ2 = 51.7; df = 3; p < 0.01) and specimens significantly larger (t = 2.2; df = 4.1; p = 0.09) in areas away from the eastern hearth...... 148

Table 12. Model of consumed body parts at Swan Point CZ4b and associated energetic returns.

The version displayed is the highest possible energetic return obtained from body parts recorded at CZ4b, with a total portion weight of 33.4 Kg and an associated energetic return of 215,000

Kcal. However, major energy contributors (e.g. swan, infant mammoth ribcage) include portions which were probably not (or only partially) consumed (Lanoë and Holmes 2016). With the assumptions that brains, hindfeet, and birds were acquired for technological purposes (tanning, sinew, and feathers, respectively), and that only 20% of the mammoth ribcage was consumed at the site, estimation of weight of consumed body parts falls to 5.0 Kg and associated energetic returns to 31,000 Kcal...... 161

Table 13. Multivariate analysis of variances in stable isotope ratios of megafauna prior and after

14,300 cal B.P. P-values under 0.05 (in bold) indicate significant differences in the recorded stable isotope ratios in the two periods...... 191

Table 14. Zooarchaeological data (NISP, when available) on animals consumed at subarctic eastern Beringian sites throughout the Late Pleistocene and early Holocene. Mammoth tooth

15 fragments as well as cervid antler (when indicated in the publication) are excluded, as well as faunal specimens likely to represent intrusions from other stratigraphic levels. Faunal frequencies for the Little John site are estimates based on published graphs. For the Broken Mammoth site, it is unclear from extant publications which components contain sheep, caribou or remains.

...... 196

Table 15. Isotopic values obtained for this study. …………………………………………...…..... 208

16

ABSTRACT

The human colonization of Beringia during the Late Glacial (14,500-11,700 years ago) is one of the most remarkable events in human history, as people coming from northeastern Eurasia permanently settled the subarctic for the first time and opened the way for the initial colonization of the New World, coincident with the extinction of mammalian megafauna.

This dissertation uses an interdisciplinary methodological and theoretical framework to investigate the trophic ecology of past Beringian hunter-gatherers, their place in predator guilds and in the broader mammal community. Methods of study include faunal and spatial analyses of existing archaeological collections, analyzing the function of the sites as well as their spatial relationship to resources, and documenting new sites in the region through excavations. In support of the archaeological work I conducted isotopic analyses (δ13C and δ15N) of megafauna from contemporaneous paleontological and archaeological sites in order to study habitat partitioning within the herbivore and predator guilds.

Results show that resources targeted by early Beringian people were concentrated in high biomass patches and that people exploited these patches through highly specialized, logistical sites. Specifically, the occupation at Swan Point CZ4b is interpreted as a specialized workshop dedicated to the production and maintenance of organic-based tools, providing evidence that

Beringian people relied on animals not only for food but also to a large extent for technological purposes. Isotopic data suggest that Beringian people are unlikely to have had profound negative effects on populations of large herbivores through their economic choices. On the other hand, humans do seem to have contributed to the extinction of Beringian large carnivores by competing with and ultimately displacing them at high trophic levels.

17

This dissertation provides new evidence of the impact that the integration of early Beringian people within predator communities had on material culture and economy as well as on larger- scale ecosystem processes.

18

INTRODUCTION

The colonization of Beringia and the Americas and the contemporaneous megafaunal extinctions remain hotly-debated issues among archaeologists and paleoecologists. To date, however, more research has been devoted to correlations with global climate change than to characterizing and understanding the finer-scale local ecological aspects of this process. The latter must have included complex interactions between soils, habitat, animals, and people, and these ecological dynamics were strongly influenced by disturbance and successional cycles. At the human level, the colonization of Beringia is one of the first incursions above a latitude of 60º north, and early

Beringians invaded mammalian communities with long co-evolved populations of predators and prey. Ecological issues such as the niche of early human settlers and their membership in predator guilds have yet to be adequately addressed, even though these fine-scale processes must have had a deep impact on (i) the tempo and geographical setting of human colonization, including their technologic and economic adaptations during and following colonization, and (ii) the organization and dynamics of mammalian paleocommunities starting in the Late Pleistocene.

The research presented in this dissertation focuses on the ecology and the economy of early

Beringian people using an interdisciplinary methodological and theoretical framework. The ecology and economy of hunter-gatherer populations are typically investigated by researchers from distinct disciplines. Most notably, zooarchaeologists, primarily anthropologists, reconstruct the diet of past societies from faunal remains found in the specific context of an archaeological site, while paleoecologists, primarily natural scientists, focus on the long-term trends of the environments and ecosystems of which people were a part. An explicitly interdisciplinary approach allows a focus on how people have integrated ecosystems over the long term, and how

19 economic choices may have translated into ecological relationships that influenced the structure of a large mammal community or an ecosystem from a human evolutionary perspective.

The first article, Animals as Raw Material in Beringia: Insights from the Site of Swan Point

CZ4b, Alaska, (Appendix A; Lanoë and Holmes 2016), explores a specific aspect of the relationship between Beringian people and megafauna. Megafauna were important for people not only as food, but also as a source of raw materials. The results of faunal and spatial analyses of the 14,000 years-old site of Swan Point CZ4b, in the Shaw Creek Flats of central Alaska, show that the site was used as a specialized workshop. Activities focused on the production of composite tools based on organic blanks of caribou antler and mammoth ivory. Access to and location of raw material in the landscape conditioned the location of the workshop. The importance of animals and their use as raw material for technology likely had a profound impact on the evolution of material culture of Beringian people.

The second article, Task-Specific Sites and Paleoindian Landscape Use in the Shaw Creek Flats of Eastern Beringia (Appendix B), focuses on the economic organization of Beringian people at the scale of the landscape. The Shaw Creek Flats in central Alaska provide an excellent context for preservation and discovery of Late Glacial archaeological occupations. Analysis of this record from published studies and data from new excavations allows us to recognize the function of sites used by hunter-gatherer groups. Residential sites used over the long term were located near critical but immovable resources such as water bodies and vegetation. Many food resources were instead acquired through the use of specialized hunting sites such as the 13,400 cal B.P.

Keystone Dune occupation. Late Glacial people targeted steppe bison (Bison priscus) and wapiti

( elaphus) in geologically unstable habitats dominated by grasses. The use of a logistical strategy of mobility maximized efficiency in harvesting these high-biomass ungulates.

20

The third article, The Lion’s Share: Human Paleoecological Integration in Subarctic Eastern

Beringia (Appendix C), explores ecological interactions between Beringian herbivores and large predators, including humans. This article provides a synthetic approach to the paleoecology of the Pleistocene-Holocene boundary in Beringia centering on habitat use by megafauna species.

The methods of investigation include stable isotope analysis of megafauna fossil collagen, as well as results from archaeological studies (see Appendices A and B). The results suggest that large herbivore population dynamics were related to variation in the heterogeneity and availability of favorable habitat, but also to competition for habitat between members of the herbivore guild. Competition for access to prey also helped structure the guild of large predators.

However, in this case humans’ efficient harvesting of high-return resources may have had a profound effect on other large predators, and people are likely to have contributed to the extinction or niche shift of other large predators in eastern Beringia at the end of the Pleistocene.

Biomes in Beringia changed radically at the end of the Pleistocene, and this turnover was sudden. In this dissertation, I show how this process may have played out: rather than simply being explained by top-down climate forcing, bottom-up effects resulting from ecological interactions between habitat, herbivores, predators, and humans were quite important. Beringia is not an isolated example by any means. Throughout the history of human evolution, the dispersal of people into new territories seems to have been favored by available niche space, and in many cases the arrival of people also severely impacted the organization of the predator guild (e.g.,

Stiner 2002).

In contrast to other biological and anthropological records, paleontological and archaeological records have a uniquely great time depth that provides the opportunity to observe significant patterns of system structural change over long periods. Eventually we can understand how

21 interactions between humans and other mammalian species have shaped biotic communities throughout the Quaternary, how hunter-gatherer societies were integrated in ecosystems, and whether people were factors of change, stability, or both. In turn, we can develop and refine robust hypotheses about the causes of intervals of stability and change, and, more globally, contribute to effective ecological and anthropological theory.

22

LITERATURE REVIEW

Paleoecology and Archaeology

Paleoecological and archaeological sciences are similar in many respects. Their focus is on records of past events, often using similar methods (e.g. stable isotope analysis, sedimentology, radiometric dating) and general theories (e.g. evolution, actualism). The single most important difference resides in their object of study. Archaeologists, as anthropologists, study human behavior (Binford 1962), whereas paleoecologists study ecosystems that include a variety of animal and plant species (Pianka 1998). Anthropological and ecological approaches are not always integrated as well as they could be, even though people cannot be understood outside their environment, and humans are animals with disproportionately large impacts on ecosystem structure and dynamics (e.g. Laland et al. 2000, Laland and O’Brien 2010).

Throughout their evolutionary history, humans did not disperse in isolation but were part of global mammalian turnovers that deeply influenced the structure of communities. Discussions about the colonization of the subarctic and the Americas should not be anthropocentric or centered on human behavior only, but rather should be understood as part of communities where humans may or may not have had important roles. Because of its geographic position, the biotic history of Beringia is complex; the region has been a major dispersal route for animals and plants since the Paleocene. Humans are only one of countless species that dispersed from Eurasia to

North America, or the reverse, throughout the Cenozoic (e.g., Van Valkenburgh 1999). As regards this research, ecological communities underwent major changes between 16,000 and

9,000 cal. BP. Ecological communities arguably are emergent patterns formed by interactions between biotic and abiotic components of an ecosystem including soil evapo-transpiration rates

23

(Guthrie 2001), patterns of landscape disturbance (Reuther 2013), or ecological engineering

(Zimov et al. 1995, 2012). Many models of the human colonization of North America emphasize the effects of global, top-down processes such as marine transgression, global warming, or ice sheet retreat. Human settlement of the arctic during the Late Glacial was definitely influenced by a period of major climatic change, but bottom-up processes also played a very significant role.

The subarctic and arctic archaeological records present a rich history of human adaptation and exploitation of different environments. Late Pleistocene people seem to have occupied terredstrial predator niches oriented to large mammal megafauna (references in text below).

Later people of the early Holocene shifted to a wider variety of niches, including coastal marine focused economies (Fitzhugh 2016). Human economic choices impact their environments, and subarctic and arctic ecosystems are particularly vulnerable to natural and human-induced perturbations (Post et al. 2009, Miller et al. 2010). In such contexts, the study of human societies cannot be thought of as separate from their ecosystems. The potential for combining archaeology and paleoecology therefore is great. Understanding the processes that link climate, habitat, large , and humans in the past will help us to understand and manage ecosystem dynamics in the future. Outcomes are also likely to significantly improve theories and methods used in both disciplines and, more widely, to contribute to theory-building in both anthropology and ecology.

Human-Environment Interactions

Environment conditions have a strong influence on several characteristics of hunter-gatherer groups (e.g. Binford 2001). Human behavior is expected to be rational or generally optimized in relation to the biotope the people occupy. Some aspects of this behavior, such as subsistence,

24 therefore can be modeled and predicted with relative accuracy (Bird and O’Connell 2006).

Nevertheless, a functional explanation of human behavior is only possible when detailed assessments of ecological community dynamics and variation are available, and archaeologists are often confronted with modes of human adaptation and environments that no longer exist.

Human influence on the environment and megafauna extinctions, on the other hand, are much less well understood. Many researchers tend to emphasize human’s role as a predator of large- bodied mammals (Lyons et al. 2004, Koch and Barnosky 2006). Typically, though, arguments for human-caused extinction are limited to discussions of contemporaneity between human dispersal events and species extinction (e.g. Surovell et al. 2016). The ecological background to this and other interpretations of human subsistence is often lacking. The extinction of an animal as the result of predation is generally limited to cases of carnivore invasion into predator-free environments where “naive” prey species did not evolve defense mechanisms, such as on islands. Overexploitation of a prey by a predator, or prey demographic decrease, is actually more likely to negatively affect the demography of a predator than of its prey (e.g. Sinclair 2003).

Human-Carnivore Coevolution

The issues of niche and guild membership of early Beringians and how the resource landscape was partitioned between humans and other large predators are essential considerations for improving our understanding of the paleoecology and the archaeology of Pleistocene-Holocene transition. Hominins’ invastion of the carnivore guilds seems to have shaped the evolutionary history of the Homininae for the past 2.6 million years, and competition with other predators

25 along a trophic continuum has led to processes of resource partitioning and character displacement.

Brantingham (1998a) shows that Plio-Pleistocene hominins at Olduvai invaded a niche located between top predators and confrontational scavengers. The invasion of this intermediate niche was arguably easier than becoming an apex predator or non-confrontational scavenger. In both cases, these other niches were already occupied by specialized species that were morphologically very well adapted. Humans would not have constituted efficient competitors for these niches.

Brantingham (1998b) also suggests that changes in behavior, in particular mobility strategies, conferred competitive advantages to humans over other predator species.

In Pleistocene Italy, taphonomic evidence suggests that Lower Paleolithic populations invaded the carnivore guild at different niche spaces than those occupied by the native large carnivores

(Stiner 1994, 2002, 2004). Humans (Homo heidelbergensis/neanderthalensis) invaded an unoccupied niche as they specialized on prime-aged ungulates from lowland and open habitats.

In contrast, wolves (Canis lupus) and cave hyenas (Crocuta spelaea) specialized on weak (young and old) individuals from woodland and highland habitats, and lowland and open habitats, respectively.

Discamps (2011, 2014) shows that levels of niche differentiation and inter-specific competition were highly dependent on ungulate biomass. Late Pleistocene humans (Homo neanderthalensis/sapiens) in southwestern France coexisted with hyenas when ungulate biomass was high and inter-specific competition low. In that particular region, as ungulate biomass decreased during Oxygen Isotopic Stage 3, hyenas were out-competed by humans.

26

Niche displacement and interaction with other members of the carnivore guild did not end with the dispersal of Homo sapiens out of Africa. It continued perhaps at an accelerated pace throughout the Late Pleistocene and Holocene. Humans encountered several species of large predators in Late Pleistocene Beringia, including brown bear (Ursus arctos) and short-faced bear

(Arctodus simus), American lion (Panthera leo atrox), wolf (Canis lupus), and saber-tooth cat

(Homotherium serum). The evolutionary history (Barnes et al. 2002, Sommer and Benecke 2005,

Leonard et al. 2007, Barnett et al. 2009, Davison et al. 2011) and diet (Matheus 1995, Sorkin

2006, Fox-Dobbs et al. 2008, Bocherens et al. 2011, Yeakel et al. 2013, Donohue et al. 2013,

Bocherens 2015) of these predators are fairly well-known. In contrast, changes in guild structure remain little studied. Carnivore guilds are likely to have undergone dramatic changes throughout the Late Glacial and early Holocene, evidenced for instance by the loss of a specialized scavenger niche, and demographic decrease as evidenced by decrease in rates of bone accumulation in karstic dens (Sattler et al. 2001).

Community Structure

Ecological literature related to dynamics of community change generally follows the lines of the old Gleason-Clements succession debate (McIntosh 1995). Clementsian perspectives emphasize communities as stable and predictable wholes comprised of coevolved species possessing well- defined niches. In contrast, Gleasonian perspectives emphasize individualistic responses of species to environmental dynamics and the potential prevalence of unpredictable, non-analogue communities for which equivalents no longer exist.

27

Rodríguez (2004, 2006) approaches Early and Middle Pleistocene megafauna communities in

Iberia by focusing on community structure as biome ecospace. An ecospace reflects the organization of different species in ecological niches defined from microhabitat, diet, locomotion, or body size. Comparisons between communities in Pleistocene Iberia and other fossil and modern biomes suggest that megafaunal species have responded to global climate forcing during the Pleistocene in an individualistic manner, resulting in multiple non-analogue communities over time. North American large mammals also responded to Pleistocene-Holocene climate change in an individualistic fashion, explaining the existence of Late Pleistocene non- analogue communities (Lyons 2003, 2005, Lyons et al. 2010), and supporting a more Gleasonian explanation of community change.

Rodríguez (2004, 2006) shows that despite species turnover, Iberia ecospace remained essentially unchanged throughout the Pleistocene, suggesting that community structure was resilient to deep time climatic change there even when individual species were not. In contrast, the Northern European ecospace changed drastically during the Pleistocene and community structure shifted between alternative stable states. While communities may have behaved as a

Gleasonian system at the species level, changes at the level of the community are best explained by a Clementsian perspective.

Networks concepts provide another approach to investigating the structure of ecological communities. Numerous researchers have investigated the relationship between trophic network

(food web) structure and community stability (May 1972, McCann 2000, Dunne et al. 2001, Li and Charnov 2001, Montoya and Solé 2002, Berlow et al. 2009, Gross et al. 2009, Thébault and

Fontaine 2010). The transition of trophic networks to different states depends on their resistance and resilience, that is their ability to resist or overcome exterior disturbances, and hence their

28 relative stability (Gunderson 2000, Folke 2006). Some trophic network parameters, such as nestedness or connectance, provide insights on the resilience and resistance of past communities.

These concepts can be applied successfully to some fossil communities (Dunne et al. 2008, 2014,

Maschner et al. 2009), and theoretical ecologists generally call for using this type of approach to records as a way to track the effect of long-term environmental changes on the structure of trophic and ecological networks (Ings et al. 2009). Most fossil datasets do not, however, provide the resolution needed to use network models, and these methods are not likely to be used commonly in paleoecology in the near future.

Economy & Mobility

“There is hardly a more romantic image in anthropology than that of a small band of hunter- gatherers setting off through the dunes and scrub, their few belongings on their back.”

(Kelly 2013)

Hunter-gatherers by definition move regularly to acquire resources, rather than producing them near their residence as farmers do. It is then no wonder that anthropologists traditionally associate hunter-gatherer societies with concepts surrounding the use of space, chief among these the ideas of mobility and territory, and that the study of mobility is tightly associated in anthropological archaeology with studies of subsistence and economy.

29

Simply moving around does not typically leave traces in the archaeological record and thus characterizing past hunter-gatherers mobility patterns can be quite difficult. Some of the best proxies of past human mobility come from indirect evidence related to raw material acquisition.

Hunter-gatherers are known to move long distances to acquire highly desirable raw materials

(Kelly and Todd 1988, Féblot-Augustins 2009, Reuther et al. 2011), even though it is not always possible to know from the archaeological record whether materials were directly extracted or acquired through trade with neighboring groups.

Archaeological sourcing studies emphasize the optimality and rationality of mobility associated with raw material procurement. Brantingham (2006), using ecological models of randomized mobility, suggests that the Aurignacian of southwestern France followed a highly planned and organized mobility strategy in regards to raw material procurement, one that differs from Lévy flight movement patterns (with parameter µ ~ 2) that are typically associated with foraging

(Bartumeus et al. 2005, Raichlen et al. 2014).

Wilson and Browne (Wilson 2007, Browne and Wilson 2011) also suggest that the Mousterian people of southeastern France exploited unevenly distributed lithic raw materials in an optimal and efficient fashion in regards to quality, quantity, ease of access and extraction. Féblot-

Augustins (2009) studying European Upper Paleolithic lithic assemblages suggest that differences in raw material provisioning strategies between groups may be determined more by environmental parameters, such as gradients of continentalism, rather than by cultural norms.

Hunter-gatherer mobility and economy are embodied in settlement strategies. Many archaeologists today continue to use Lewis Binford’s model (1980) for classifying settlement patterns along a continuum of residential to logistical mobility strategies (Perreault and

30

Brantingham 2011, Kelly 2013). Residential mobility (forager strategy) describes groups in which all members move from residential camp to residential camp and exploit resources in the immediate vicinity. In contrast, logistical mobility (collector strategy) describes groups in which task-specific teams conduct forays to bring resources from far-away areas to the rest of the group at a fairly stable residential camp.

Radiating from a residential camp, hunter-gatherers’ economic landscape may be divided into a series of nested areas (Binford 1982, Morgan 2008): (i) residential camp; (ii) foraging radius, in which resources are exploited by daily work parties; (iii) logistical radius, in which resources are exploited by specialized multi-day task groups; and (iv) visiting radius, the area known but not exploited and, generally occupied by relatives or economic partners.

The forager/collector model is not a simple dichotomy. Variation in settlement strategies among ethnographically- and archaeologically-documented hunter-gatherer groups generally represents a continuum. Factors explaining this variation can be tied to environmental parameters, among other factors (Kelly 1983, Grove 2009).

Intuitively, settlement strategies depend on the homogeneity of an economic landscape.

Homogeneous landscapes would be characterized by a low β diversity (species variation across space) and low variation in resource seasonality. In these landscapes the resources targeted by hunter-gatherers would occur everywhere and at any given time, whether in small or large quantities. In contrast, heterogeneous environments would be characterized by a high β diversity, and targeted resources would only be found at key locations on the landscape, or at specific times of the year, but possibly in predictable locations and times, and in higher, “bulk” quantities.

31

In homogeneous landscapes, a hunter-gatherer group will move residence in response to local depletion of resources. Patterns of residential movement can be classified, depending on the extent of relocation of the foraging radius, as a half foraging radius move, a complete radius

“leap frog” move, or a point-to-point move (Binford 1982). Following the marginal value theorem, areas outside the foraging radius will become economically attractive when the expected returns outweigh those provided by resources within the foraging radius, taking in account the cost of moving (Bird and O’Connell 2006). In heterogeneous landscapes with resources that are available at different seasons, optimal exploitation by humans is likely to follow a logistical organization centered around a residential base. The location and duration of occupation will be weighted to search, harvesting, and transport costs of each targeted resource.

Models of hunter-gatherer mobility provide a conceptual framework to understand the economic behavior of prehistoric groups at the scale of the landscape, particularly how human behavior may be constrained by the distribution of adequate resources and habitat. Similarly, these models outline how ecological interactions between humans and resources result from economic strategies.

Late Glacial Beringia

Paleoenvironments

Prior to the Last Glacial Maximum (LGM; about 26,500-19,000 cal. BP), eastern Beringia supported a relatively high biomass, diverse community of large herbivores (woolly mammoth

Mammuthus primigenius, Yukon horse Equus lambei, stilt-legged horse Equus cf. kiang, steppe

32 bison Bison priscus, Western camel Camelops hesternus, caribou Rangifer tarandus, broad- fronted moose latifrons, Dall sheep Ovis dalli/nivicola, helmeted muskox Bootherium bombifrons, modern muskox Ovibos moschatus, saiga antelope Saiga tatarica) and carnivores

(American lion Panthera atrox, wolf Canis lupus, short-faced bear Arctodus simus, brown bear

Ursus arctos) (Harington 2003). This community thrived in a fertile landscape dominated by productive graminoids and forbs. A large proportion of the plant biomass was edible to large herbivores rather than being locked in tough cellulose structures such as wood (Guthrie 1984,

Zazula et al. 2003, Blinnikov et al. 2011).

The composition of the mammalian and plant communities changed after the LGM. Shrub

(Betula nana) tundra replaced the graminoid-steppe over most of eastern Beringia by 16,000 to

14,000 cal. BP (Bigelow and Powers 2001, Anderson et al. 2004, Viau et al. 2008) in association with a process of general paludification and formation of wetlands (Mann et al. 2001, Jones and

Yu 2010, Reuther 2013). Poplar (Populus spp.) colonized central and northern Alaska by 13,000-

10,000 cal. BP, favored by a warmer and wetter climate (Mann et al. 2001, Anderson et al. 2004,

Llyod et al. 2006). Mammoth, horse, and lion, went extinct by 14,000-13,000 cal. BP (Guthrie

2006), although remnant populations of mammoth and horse may have persisted as late as

10,000 cal. BP (Haile et al. 2009). Wapiti (Cervus elaphus) and humans colonized eastern

Beringia from the west by 15,000 cal. BP (Meiri et al. 2014), followed by moose (Alces alces) around 14,000-13,000 cal. BP (Guthrie 2006).

From 10,000 to 5,000 cal. BP, generally cooler and wetter conditions accompanied stabilization of the landscape (Reuther 2013) and the development of white spruce (Picea glauca) forests in central Alaska (Edwards et al. 2001, Llyod et al. 2006). Modern tussock tundra and peaty soils spread to northern Alaska by 10,000 cal. BP, and caribou and modern muskoxen replaced the

33 bison and horse that were once widespread there during the Late Glacial (Mann et al. 2001,

2013).

Alaskan ecosystems experienced comparatively little change over the rest of the Holocene and remain characterized by boreal forest in the subarctic and tundra in the arctic and high elevations.

In the subarctic, poplar and white spruce now form intermediate successional stages of the black spruce (Picea mariana) climax forest (Chapin III et al. 2006b, Llyod et al. 2006), and steppic refugia persist on well-drained insulated slopes (Guthrie 2001, Chapin III et al. 2006a). Wapiti and bison remained in the region until the Late Holocene (Stephenson et al. 2001, Potter 2008a).

Archaeology

Long-term human settlement of Eurasian regions north of 60° N does not seem to have occurred before the Last Glacial Maximum (Goebel 1999, Graf 2009), despite sporadic evidence for presence of humans as early as 30,000 years ago (Pitulko et al. 2004). Human dispersal in

Beringia during the Late Glacial likely originated in southern where numerous sites with microblade industries dating to 21,000-19,000 cal B.P. have been found in the Upper Yenisei basin and the Transbaikal region (Goebel et al. 2000, Vasil’ev 2001, Graf 2010, 2011, Buvit and

Terry 2011). These populations reached western Beringia by 15,000 to 14,000 cal B.P. based on evidence from the sites of Dyuktai Cave and Berelekh, Yakutia; and Ushki Lake, Kamchatka

(Dikov 1979, Yi et al. 1985, Mochanov and Fedoseeva 1996, Slobodin 1999, Goebel et al. 2003,

2010, Pitulko 2011).

The earliest evidence for the peopling of eastern Beringia dates to 14,000 cal B.P. at the site of

Swan Point CZ4b in Alaska (Holmes 2011). Sites in subarctic Alaska and Yukon between

34

14,000 and 12,000 cal B.P. show a relatively diverse set of technological strategies. In the

Tanana valley, industries tend to include microblade and microcore technologies, but also triangular or tear-drop shaped bifacial points named “Chindadn points” (Potter et al. 2013). At the oldest site of Swan Point CZ4b, microblades were produced using the Yubetsu technique characteristic of western Beringian sites, while at later sites microblades were produced using the

Campus method endemic to eastern Beringia (Holmes 2011, Gómez Coutouly 2012). In contrast, in the Nenana valley, sites older than 13,000 cal B.P. are dominated by macroblade technology, while also including triangular points similar to Chindadn points (Hoffecker 2001). After 13,000 cal B.P. most or all sites are dominated by microblade and microcore technologies and grouped into the Denali culture (Potter 2008a, 2011) that continued well into the Holocene. Similarly, post-13,000 cal B.P. material culture in western Beringia is dominated by microblades, though produced with a different method (Slobodin 1999, Pitulko 2001).

Some authors have argued that technological variation before 13,000 cal B.P. reflects locally distinct cultures such as Dyuktai, Nenana, Chindadn, and Denali (reviewed in Bever 2006).

Other authors have grouped some or all of these traditions into a few Beringian complexes (West

1981, Holmes 2011), arguing that the observed typological differences reflect behavioral organization (Potter et al. 2013) or adaptation to local topography and environmental conditions

(Wygal 2017) more than being distinct cultural markers. Composite tools made with microblades are fundamentally different from bifacial tools as examples of additive versus reductive technology, respectively (Dixon 2016). Nevertheless, it is generally assumed that they fulfilled a similar function, as projectile points and probably for atlatl darts rather than arrows (Dixon 1999;

Potter 2008b; Rasic and Slobodina 2008; Hoffecker 2011; although see Ackerman 2011).

Composite projectile points with inset microblades may have been more efficient where and

35 when caribou antler was available (Guthrie 1983) and at times when a reliable and maintainable hunting technology was necessary (Elston and Brantingham 2002).

In northern and southwestern Alaska, sites dating to the Younger Dryas (12,900-11,700 cal B.P.) are dominated by bifacial projectile points nearly identical to those found in Paleoindian sites of the North American Great Plains. These include lanceolate points of the Agate Basin or Hell Gap type, named Mesa points in Alaska (Kunz and Reanier 1994, Ackerman 2001, Bever 2001), and

Folsom-like fluted points (Goebel et al. 2013). In contrast to sites in central Alaska that show direct technological continuity with sites in western Beringia, sites of the Mesa and northern fluted point complexes represent a northward migration of Paleoindian populations after the opening of the ice-free corridor (Kunz and Reanier 1994, Mann et al. 2001).

Overall, the archaeological record of Beringia during the Late Glacial is characterized by a high level of diversity and variability. Some of this diversity may just reflect our lack of understanding of how technological variability affected the typology of archaeological materials found at different sites. Alternatively, material diversity may reflect different populations moving within and across Beringia during the Late Glacial. Most of the archaeological assemblages known today cluster into specific regionsm and it is unlikely that the “diversity debate” will be resolved before a more spatially homogeneous archaeological record is documented.

36

REFERENCES

Ackerman, Robert E.

2001 Spein Mountain: A Mesa Complex Site in Southwestern Alaska. Arctic

Anthropology 38(2): 81–97.

2011 Microblade Assemblages in Southwestern Alaska: An Early Holocene

Adaptation. In From the Yenisei to the Yukon: Interpreting Lithic Assemblages

Variability in Late Pleistocene / Early Holocene Beringia, edited by Ted Goebel

and Ian Buvit, pp. 255-269, Texas A & M University Press, College Station.

Anderson, Patricia M., Mary E Edwards, and Linda B Brubaker

2004 Results and Paleoclimate Implications of 35 Years of Paleoecological Research in

Alaska. Developments in Quaternary Science 1: 427–440.

Barnes, I, P Matheus, Beth Shapiro, D Jensen, and A Cooper

2002 Dynamics of Pleistocene population extinctions in Beringian brown bears.

Science 295: 2267–70.

Barnett, Ross, Beth Shapiro, Ian Barnes, Simon Y. W. Ho, Joachim Burger, Nobuyuki

Yamaguchi, Thomas F. G. Higham, H. Todd Wheeler, Wilfried Rosendahl, Andrei V. Sher,

Marina Sotnikova, Tatiana Kuznetsova, Gennady F. Baryshnikov, Larry D. Martin, C.

Richard Harington, James A Burns, and Alan Cooper

2009 Phylogeography of lions (Panthera leo ssp.) reveals three distinct taxa and a late

37

Pleistocene reduction in genetic diversity. Molecular Ecology 18(8): 1668–77.

Bartumeus, Frederic, M. G. E. Da Luz, G. M. Viswanatham, and J. Catalan

2005 Animal Search Strategies: A Quantitative Random Walk Analysis. Ecology

86(11): 3078–3087.

Berlow, Eric L., Jennifer A. Dunne, Neo D. Martinez, Philip B. Stark, Richard J. Williams, and

Ulrich Brose

2009 Simple Prediction of Interaction Strengths in Complex Food Webs. Proceedings

of the National Academy of Sciences 106(1): 187–91.

Bever, Michael R.

2001 Technology and the Mesa Complex: Developing a Framework of

Alaskan Paleoindian Prehistory. Arctic Anthropology 38(2): 98–118.

2006 Too Little, Too Late ? The Radiocarbon Chronology of Alaska and the Peopling

of the New World. American Antiquity 71(4): 595–620.

Bigelow, Nancy H.

1997 Late Quaternary Vegetation and Lake Level Changes in Central Alaska. PhD

Dissertation, University of Alaska Fairbanks.

Bigelow, Nancy H., and W. Roger Powers

2001 Climate, Vegetation, and Archaeology 14,000-9,000 Cal Yr B.P. in Central

Alaska. Arctic Anthropology 38(2): 171–195.

38

Binford, Lewis R.

1962 Archaeology as Anthropology. American Antiquity 28(2): 217–225.

1980 Smoke and Dogs’ Tails: Hunter-Gatherer Settlement Systems and

Archaeological Site Formation. American Antiquity 45: 4–20.

1982 The Archaeology of Place. Journal of Anthropological Archaeology 1(1): 5–31.

2001 Constructing Frames of Reference: An Analytical Method for Archaeological

Theory Building Using Ethnographic and Environmental Data Sets. University of

California Press, Berkeley.

Bird, Douglas W., and James F. O’Connell

2006 Behavioral Ecology and Archaeology. Journal of Archaeological Research 14(2):

143–188.

Blinnikov, Mikhail S., Benjamin V. Gaglioti, Donald A. Walker, Matthew J. Wooller, and Grant

D. Zazula

2011 Pleistocene graminoid-dominated ecosystems in the Arctic. Quaternary Science

Reviews 30(21–22): 2906–2929.

Bocherens, Hervé

2015 Isotopic tracking of large carnivore palaeoecology in the mammoth steppe.

Quaternary Science Reviews 117: 42–71.

Bocherens, Hervé, Dorothée G. Drucker, Dominique Bonjean, Anne Bridault, Nicholas J.

39

Conard, Christophe Cupillard, Mietje Germonpré, Markus Höneisen, Susanne C. Münzel,

Hannes Napierala, Marylène Patou-Mathis, Elisabeth Stephan, Hans-Peter Uerpmann, and

Reinhard Ziegler

2011 Isotopic evidence for dietary ecology of cave lion (Panthera spelaea) in North-

Western Europe: Prey choice, competition and implications for extinction.

Quaternary International 245: 249–261.

Bodu, Pierre, Michèle Julien, Borix Valentin, and Grégory Debout (editors).

2006 Un dernier hiver à Pincevent : les Magdaléniens du niveau IV0. Gallia

Préhistoire.

Brantingham, P. Jeffrey

1998a Mobility, Competition, and Plio-Pleistocene Hominid Foraging Groups. Journal

of Archaeological Method and Theory 5(1): 57–98.

1998b Hominid–Carnivore Coevolution and Invasion of the Predatory Guild. Journal of

Anthropological Archaeology 17: 327–353.

2006 Measuring Forager Mobility. Current Anthropology 47(3): 435–459.

Bronk Ramsey, Christopher

1994 Analysis of Chronological Information and Radiocarbon Calibration : The

Program OxCal. Archaeological Computing Newsletter 41: 11–16.

2009 Bayesian Analysis of Radiocarbon Dates. Radiocarbon 51(1): 337–360.

40

Browne, Constance L., and Lucy Wilson

2011 Resource selection of lithic raw materials in the Middle Paleolithic in southern

France. Journal of Human Evolution 61(5): 597–608.

Buvit, Ian, and Karisa Terry

2011 The twilight of Paleolithic Siberia: Humans and their environments east of Lake

Baikal at the late-glacial/Holocene transition. Quaternary International 242(2):

379–400.

Chapin III, F.S., T Hollingsworth, D F Murray, L A Viereck, and M D Walker

2006 Floristic Diversity and Vegetation Distribution in the Alaskan Boreal Forest. In

Alaska’s Changing Boreal Forest, edited by F. S. Chapin III, M. W. Oswood, K.

Van Cleve, L. A. Viereck, and D. L. Verbyla, pp. 81–99. Oxford University Press,

New York.

Chapin III, F.S., L. A. Viereck, P. C. Adams, K. Van Cleve, C. L. Fastie, R. A. Ott, Daniel H.

Mann, and J. F. Johnstone

2006 Successional Processes in the Alaskan Boreal Forest. In Alaska’s Changing

Boreal Forest, edited by F.S. Chapin III, M. W. Oswood, K. Van Cleve, L. A.

Viereck, and D. L. Verbyla, pp. 100–120. Oxford University Press, New York.

Davison, John, Simon Y.W. Ho, Sarah C. Bray, Marju Korsten, Egle Tammeleht, Maris

Hindrikson, Kjartan Østbye, Eivind Østbye, Stein-Erik Lauritzen, and Jeremy Austin

2011 Late-Quaternary biogeographic scenarios for the brown bear (Ursus arctos), a

41

wild mammal model species. Quaternary Science Reviews 30(3–4): 418–430.

Dikov, Nikolai N.

1979 Early Cultures of Northeastern Asia. Nauka, Moscow.

Discamps, Emmanuel

2011 Hommes et hyènes face aux recompositions des communautés d’Ongulés

(MIS 5-3) : Éléments pour un cadre paléoécologique des sociétés du

Paléolithique moyen et supérieur ancien d’Europe de l’Ouest. PhD Dissertation,

Université de Bordeaux 1.

2014 Ungulate biomass fluctuations endured by Middle and Early Upper Paleolithic

societies (SW France, MIS 5-3): The contributions of modern analogs and cave

hyena paleodemography. Quaternary International 337: 64–79.

Dixon, E. James

1999 Bones, boats, and bison: archaeology and the first colonization of western North

America. University of New Mexico Press, Albuquerque.

2016 Arrows and Atl Atls. National Park Service.

Donohue, Shelly L., Larisa R. G. DeSantis, Blaine W. Schubert, and Peter S. Ungar

2013 Was the Giant Short-Faced Bear a Hyper-Scavenger? A New Approach to the

Dietary Study of Ursids Using Dental Microwear Textures. Ed. Laurent Viriot.

PLoS ONE 8(10): e77531.

42

Dunne, Jennifer A, Ulrich Brose, Richard J Williams, and Neo D. Martinez

2001 Modeling food-web dynamics : complexity – stability implications. In Aquatic

Food Webs: An Ecosystem Approach, edited by Andrea Belgrano, Ursula M

Scharler, Jennifer Dunne, and Robert E Ulanowicz, pp. 117–129. Oxford

University Press.

Dunne, Jennifer A., Conrad C. Labandeira, and Richard J. Williams

2014 Highly resolved early Eocene food webs show development of modern trophic

structure after the end-Cretaceous extinction. Proceedings of the Royal Society B

281:2013280.

Dunne, Jennifer A., Richard J. Williams, Neo D. Martinez, Rachel A. Wood, and Douglas H.

Erwin

2008 Compilation and Network Analyses of Cambrian Food Webs. PLoS Biology 6(4):

e102.

Edwards, M. E., C. J. Mock, B. P. Finney, Valerie A. Barber, and P. J. Bartlein

2001 Potential analogues for paleoclimatic variations in eastern interior Alaska during

the past 14,000 yr: atmospheric-circulation controls of regional temperature and

moisture responses. Quaternary Science Reviews 20(1–3): 189–202.

Elston, Robert G., and P. Jeffrey Brantingham

2002 Microlithic Technology in Northern Asia: A Risk-Minimizing Strategy of the

Late Paleolithic and Early Holocene. In Thinking Small: Global Perspectives on

43

Microlithization, edited by Robert G. Elston and Steven L. Kuhn, pp. 101-116.

American Anthropological Association, Arlington, Virginia.

Féblot-Augustins, Jehanne

2009 Revisiting European Upper Paleolithic Raw Material Transfers: The Demise of

the Cultural Ecological Paradigm? In Lithic Materials and Paleolithic Societies,

edited by B Adams and B S Blades, pp. 25–46. Blackwell Publishing.

Fitzhugh, Ben

2016 The Origins and Development of Arctic Maritime Adaptations in the Subarctic

and Arctic Pacific. In The Oxford Handbook of the Prehistoric Arctic, edited by T

Max Friesen and Owen K Mason, pp. 253-278, Oxford University Press, New

York.

Folke, Carl

2006 Resilience: The emergence of a perspective for social–ecological systems

analyses. Global Environmental Change 16(3): 253–267.

Fox-Dobbs, Kena, Jennifer A Leonard, and Paul L. Koch

2008 Pleistocene megafauna from eastern Beringia: Paleoecological and

paleoenvironmental interpretations of stable carbon and nitrogen isotope and

radiocarbon records. Palaeogeography, Palaeoclimatology, Palaeoecology

261(1–2): 30–46.

Goebel, Ted

44

1999 Pleistocene human colonization of Siberia and peopling of the Americas: An

ecological approach. Evolutionary Anthropology 8(6): 208–227.

Goebel, Ted, Sergei B. Slobodin, and Michael R. Waters

2010 New dates from Ushki-1, Kamchatka, confirm 13,000 cal BP age for earliest

Paleolithic occupation. Journal of Archaeological Science 37(10): 2640–2649.

Goebel, Ted, Heather L. Smith, Lyndsay DiPietro, Michael R. Waters, Bryan Hockett, Kelly E.

Graf, Robert Gal, Sergei B. Slobodin, Robert J. Speakman, Steven G. Driese, and David

Rhode

2013 Serpentine Hot Springs, Alaska: results of excavations and implications for the

age and significance of northern fluted points. Journal of Archaeological Science

40(12): 4222–4233.

Goebel, Ted, Michael R. Waters, Ian Buvit, Mikhail V. Konstantinov, and Aleksander V.

Konstantinov

2000 Studenoe-2 and the origins of microblade technologies in the Transbaikal, Siberia.

Antiquity 74.

Goebel, Ted, Michael R. Waters, and Margarita Dikova

2003 The archaeology of Ushki Lake, Kamchatka, and the Pleistocene peopling of the

Americas. Science 301(5632): 501–5.

Gómez Coutouly, Yan Axel

2012 Pressure Microblade Industries in Pleistocene-Holocene Interior Alaska: Current

45

Data and Discussions. In The Emergence of Pressure Blade Making: From Origin

to Modern Experimentation, edited by Pierre M. Desrosiers. Springer US, Boston,

MA.

Graf, Kelly E.

2009 Modern Human Colonization of the Siberian Mammoth Steppe: A View from

South-Central Siberia. In Sourcebook of Paleolithic Transitions, edited by Marta

Camps and Parth Chauhan, pp. 479–501. Springer New York, New York, NY.

2010 Hunter–gatherer dispersals in the mammoth-steppe: technological provisioning

and land-use in the Enisei River valley, south-central Siberia. Journal of

Archaeological Science 37(1): 210–223.

2011 On Late Upper Paleolithic Variability in South-Central Siberia: Rethinking the

Afontova and Kokorevo Cultures. In From the Yenisei to the Yukon: Interpreting

Lithic Assemblage Variability in Late Pleistocene/Early Holocene Beringia,

edited by Ted Goebel and Ian Buvit, pp. 33–46. Texas A&M University Press,

College Station.

Gross, Thilo, Lars Rudolf, Simon A. Levin, and Ulf Dieckmann

2009 Generalized models reveal stabilizing factors in food webs. Science 325(5941):

747–50.

Grove, Matt

2009 Hunter-Gatherer Movement Patterns: Causes and Constraints. Journal of

Anthropological Archaeology 28: 222–233.

46

Gunderson, Lance H.

2000 Ecological Resilience - In Theory and Application. Annual Review of Ecology,

Evolution, and Systematics 31: 425–439.

Guthrie, R. Dale

1983 Osseous Projectile Points: Biological Considerations Affecting Raw Material

Selection and Design among Paleolithic and Paleoindian People. In Animals and

Archaeology: 1. Hunters and their Prey, edited by Juliet Clutton-Brock and

Caroline Grigson, pp. 273-294. Archaeopress, Oxford.

1984 Mosaics, Allelochemics, and Nutrients: An Ecological Theory of Late Pleistocene

Megafaunal Extinctions. In Quaternary Extinctions: A Prehistoric Revolution,

edited by Paul S Martin and Richard G Klein, pp. 259–298. University of Arizona

Press, Tucson.

2001 Origin and Causes of the Mammoth Steppe: A Story of Cloud Cover, Woolly

Mammal Tooth Pits, Buckles, and Inside-Out Beringia. Quaternary Science

Reviews 20: 549–574.

2006 New Carbon Dates Link Climatic Change with Human Colonization and

Pleistocene Extinctions. Nature 441: 207–9.

Haile, James, Duane G. Froese, Ross D.E. MacPhee, Richard G. Roberts, Lee J. Arnold, Alberto

V. Reyes, Morten Rasmussen, Rasmus Nielsen, Barry W. Brook, Simon Robinson, Martina

Demuro, M. Thomas P. Gilbert, Kasper Munch, Jeremy J. Austin, Alan Cooper, Ian Barnes,

Per Möller, and Eske Willerslev

47

2009 Ancient DNA reveals late survival of mammoth and horse in interior Alaska.

Proceedings of the National Academy of Sciences 106(52): 22352–7.

Harington, C. R.

2003 Annotated Bibliography of Quaternary Vertebrates of North America. University

of Toronto Press.

Hoffecker, John F.

2001 Late Pleistocene and Early Holocene Sites in the Nenana River Valley, Central

Alaska. Arctic Anthropology 38(2): 139–153.

2011 Assemblage Variability in Beringia: The Mesa Factor. In From the Yenisei to the

Yukon: Interpreting Lithic Assemblage Variability in Late Pleistocene/Early

Holocene Beringia, edited by Ted Goebel and Ian Buvit, pp. 165-178. Texas

A&M University Press, College Station, TX.

Holmes, Charles E.

2011 The Beringian and Transitional Periods in Alaska: Technology of the East

Beringian Tradition as Viewed from Swan Point. In From the Yenisei to the

Yukon: Interpreting Lithic Assemblage Variability in Late Pleistocene/Early

Holocene Beringia, edited by Ted Goebel and Ian Buvit, pp. 179–191. Texas

A&M University Press, College Station, TX.

Jones, Miriam C., and Zicheng Yu

2010 Rapid deglacial and early Holocene expansion of peatlands in Alaska.

48

Proceedings of the National Academy of Sciences 107(16): 7347–52.

Kelly, Robert L.

1983 Hunter-Gatherer Mobility Strategies. Journal of Anthropological Research 39(3):

277–306.

2013 The Lifeways of Hunter-Gatherers: The Foraging Spectrum. Cambridge

University Press.

Kelly, Robert L., and Lawrence C. Todd

1988 Coming into the Country: Early Paleoindian Hunting and Mobility. American

Antiquity 53(2): 231–244.

Koch, Paul L., and Anthony D. Barnosky

2006 Late Quaternary Extinctions: State of the Debate. Annual Review of Ecology,

Evolution, and Systematics 37(1): 215–250.

Kunz, Michael L., and Richard E. Reanier

1994 Paleoindians in Beringia: Evidence from Arctic Alaska. Science 263(5147): 660–

2.

Laland, K. N., J. Odling-Smee, and M. W. Feldman

2000 Niche construction, biological evolution, and cultural change. Behavioral and

Brain Sciences 23(1): 131–75.

Laland, Kevin N., and Michael J. O’Brien

49

2010 Niche Construction Theory and Archaeology. Journal of Archaeological Method

and Theory 17(4): 303–322.

Lanoë, François B., and Charles E. Holmes

2016 Animals as Raw Materials in Beringia: Insights from the Site of Swan Point

CZ4b, Alaska. American Antiquity 81(4): 682–696.

Leonard, Jennifer A, Carles Vilà, Kena Fox-Dobbs, Paul L. Koch, Robert K. Wayne, and Blaire

Van Valkenburgh

2007 Megafaunal extinctions and the disappearance of a specialized wolf ecomorph.

Current Biology 17(13): 1146–50.

Li, Bai-Lian, and Eric L. Charnov

2001 Diversity–stability relationships revisited: scaling rules for biological

communities near equilibrium. Ecological Modelling 140(3): 247–254.

Llyod, A. H., M. E. Edwards, B. P. Finney, J. A. Lynch, Valerie A. Barber, and Nancy H.

Bigelow

2006 Holocene Development of the Alaskan Boreal Forest. In Alaska’s Changing

Boreal Forest, edited by F.S. Chapin III, M. W. Oswood, K. Van Cleve, L. A.

Viereck, and D. L. Verbyla, pp. 62–80. Oxford University Press, New York.

Lyons, S. Kathleen

2003 A Quantitative Assessment of the Range Shifts of Pleistocene Mammals. Journal

of Mammalogy 84(2): 385–402.

50

2005 A quantitative model for assessing community dynamics of pleistocene mammals.

The American Naturalist 165(6): E168–E185.

Lyons, S. Kathleen, Felisa A. Smith, and James H. Brown

2004 Of mice , mastodons and men : human-mediated extinctions on four continents.

Evolutionary Ecology Research 6: 339–358.

Lyons, S. Kathleen, Peter J. Wagner, and Katherine Dzikiewicz

2010 Ecological correlates of range shifts of Late Pleistocene mammals. Philosophical

Transactions of the Royal Society 365(1558): 3681–3693.

Mann, Daniel H., Pamela Groves, Michael L. Kunz, Richard E. Reanier, and Benjamin V.

Gaglioti

2013 Ice-age megafauna in Arctic Alaska: extinction, invasion, survival. Quaternary

Science Reviews 70: 91–108.

Mann, Daniel H., Richard E. Reanier, Dorothy M. Peteet, Michael L. Kunz, and Mark Johnson

2001 Environmental Change and Arctic Paleoindians. Arctic Anthropology 38(2): 119–

138.

Maschner, Herbert D. G., Matthew W. Betts, Joseph Cornell, Jennifer A. Dunne, Bruce Finney,

Nancy Huntly, James W. Jordan, Aaron A. King, Nicole Misarti, Katherine L. Reedy-

maschner, Roland Russell, Amber Tews, Spencer A. Wood, and Buck Benson

2009 An Introduction to the Biocomplexity of Sanak Island, Western Gulf of Alaska.

Pacific Science 63(4): 673–709.

51

Matheus, Paul E.

1995 Diet and Co-ecology of Pleistocene Short-Faced Bears and Brown Bears in

Eastern Beringia. Quaternary Research 44: 447–453.

May, Robert

1972 Will a Large Complex System be Stable? Nature 238: 413–414.

McCann, K. S.

2000 The diversity-stability debate. Nature 405(6783): 228–33.

McIntosh, Robert P.

1995 H.A. Gleason’s “Individualistic Concept” and Theory of Animal Communities: A

Continuing Controversy. Biological Reviews 70: 317–357.

Meiri, Meirav, Adrian M. Lister, Matthew J. Collins, Noreen Tuross, Ted Goebel, Simon

Blockley, Grant D. Zazula, Nienke Van Doorn, R. Dale Guthrie, Gennady G. Boeskorov,

Gennady F Baryshnikov, Andrei Sher, and Ian Barnes

2014 Faunal record identifies Bering isthmus conditions as constraint to end-

Pleistocene migration to the New World. Proceedings of the Royal Society 281:

20132167.

Miller, Gifford H., Richard B. Alley, Julie Brigham-Grette, Joan J. Fitzpatrick, Leonid Polyak,

Mark C. Serreze, and James W.C. White

2010 Arctic amplification: can the past constrain the future? Quaternary Science

52

Reviews 29(15–16): 1779–1790.

Mochanov, Yuri A., and Svetlana A. Fedoseeva

1996 Dyuktai Cave. In American Beginnings: The Prehistory and Palaeoecology of

Beringia, edited by F. H. West, pp. 164–173. University of Chicago Press.

Montoya, Jose M., and Ricard V. Solé

2002 Small world patterns in food webs. Journal of Theoretical Biology 214(3): 405–12.

Morgan, Christopher

2008 Reconstructing prehistoric hunter–gatherer foraging radii: a case study from

California’s southern Sierra Nevada. Journal of Archaeological Science 35(2):

247–258.

Perreault, Charles, and P. Jeffrey Brantingham

2011 Mobility-driven cultural transmission along the forager–collector continuum.

Journal of Anthropological Archaeology 30(1): 62–68.

Pianka, Eric R.

1998 Evolutionary Ecology. Benjamin / Cummins, San Francisco

Pitulko, Vladimir V.

2001 Terminal Pleistocene - Early Holocene occupation in northeast Asia and the

Zhokhov assemblage. Quaternary Science Reviews 20(1–3): 267–275.

2011 The Berelekh quest: A review of forty years of research in the mammoth

53

graveyard in northeast Siberia. Geoarchaeology 26(1): 5–32.

Pitulko, Vladimir V., P. A. Nikolsky, E. Yu. Girya, A. E. Basilyan, V. E. Tumskoy, S. A.

Koulakov, S. N. Astakhov, Elena Yu. Pavlova, and M. A. Anisimov

2004 The Yana RHS site: humans in the Arctic before the last glacial maximum.

Science 303(5654): 52–6.

Post, Eric, Mads C. Forchhammer, M. Syndonia Bret-Harte, Terry V. Callaghan, Torben R.

Christensen, Bo Elberling, Anthony D. Fox, Olivier Gilg, David S. Hik, Toke T. Høye, Rolf

A. Ims, Erik Jeppesen, David R. Klein, Jesper Madsen, David McGuire, Søren Rysgaard,

Daniel E Schindler, Ian Stirling, Mikkel P. Tamstorf, Nicholas J. C. Tyler, Rene van der

Wal, Jeffrey Welker, Philip A. Wookey, Niels Martin Schmidt, and Peter Aastrup

2009 Ecological dynamics across the Arctic associated with recent climate change.

Science 325(5946): 1355–8.

Potter, Ben A.

2008a Radiocarbon Chronology of Central Alaska: Technological Continuity and

Economic Change. Radiocarbon 50(2): 181–204.

2008b A First Approximation of Holocene Inter-Assemblage Variability in Central

Alaska. Arctic Anthropology 45: 89-113.

2011 Late Pleistocene and Early Holocene Assemblage Variability in Central Alaska.

In From the Yenisei to the Yukon: Interpreting Lithic Assemblage Variability in

Late Pleistocene/Early Holocene Beringia, edited by T. Goebel and I. Buvit, pp.

215–233. Center for the Study of the First Americans, College Station, TX.

54

Potter, Ben A., Charles E. Holmes, and David R. Yesner

2013 Technology and Economy among the Earliest Prehistoric Foragers in Interior

Eastern Beringia. In Paleoamerican Odyssey, edited by Kelly E. Graf, C. V.

Ketron, and M. R. Waters, pp. 81–103. Center for the Study of the First

Americans, College Station, TX.

Raichlen, David A., Brian M. Wood, Adam D. Gordon, Audax Z. P. Mabulla, and Frank W.

Marlowe

2014 Evidence of Lévy walk foraging patterns in human hunter – gatherers.

Proceedings of the National Academy of Sciences 111(2): 728–733.

Rasic, Jeffery T., and Natalia S. Slobodina

2008 Weapon Systems and Assemblage Variability During the Northern Archaic

Period in Northern Alaska. Arctic Anthropology 45: 71-88.

Reuther, Joshua D.

2013 Late Glacial and Early Holocene Geoarchaeology and Terrestrial Paleoecology

in the Lowlands of the Middle Tanana Valley, Subarctic Alaska. PhD Dissertation,

University of Arizona.

Reuther, Joshua D., N. S. Slobodina, Jeffrey T. Rasic, John P. Cook, and Robert J. Speakman

2011 Gaining Momentum: Late Pleistocene and Early Holocene Archaeological

Obsidian Source Studies in Interior and Northeastern Beringia. In From the

Yenisei to the Yukon: Interpreting Lithic Assemblage Variability in Late

55

Pleistocene/Early Holocene Beringia, edited by Ted Goebel and Ian Buvit, pp.

270–286. Texas A&M University Press, College Station.

Rodríguez, Jesús

2004 Stability in Pleistocene Mediterranean mammalian communities.

Palaeogeography, Palaeoclimatology, Palaeoecology 207(1–2): 1–22.

2006 Structural continuity and multiple alternative stable States in Middle Pleistocene

European mammalian communities. Palaeogeography, Palaeoclimatology,

Palaeoecology 239(3–4): 355–373.

Sattler, Robert A., Vinson, and T. E. Gillispie

2001 Calibrated Radiocarbon Ages and Taphonomic Factors in Beringian Cave Faunas

at the End of the Pleistocene. In People and Wildlife in Northern North America,

edited by S. C. Gerlach and M. S. Murray, pp. 112–123. Archaeopress: BAR

International Series 944, Oxford.

Slobodin, Sergei B.

1999 Northeast Asia in the Late Pleistocene and Early Holocene. World Archaeology

30(3): 484–502.

Sommer, Robert S., and Norbert Benecke

2005 The recolonization of Europe by brown bears Ursus arctos Linnaeus, 1758 after

the Last Glacial Maximum. Mammal Review 35(2): 156–164.

Sorkin, B.

56

2006 Ecomorphology of the giant short-faced bears Agriotherium and Arctodus.

Historical Biology 18(1): 1–20.

Stephenson, Robert O., S. C. Gerlach, R. Dale Guthrie, C. R. Harington, R. O. Mills, and G.

Hare

2001 Wood Bison in Late Holocene Alaska and Adjacent Canada: Paleontological,

Archaeological and Historical Records. In People and Wildlife in Northern North

America, edited by S. C. Gerlach and M. S. Murray, pp. 125–133. Archaeopress:

BAR International Series, Oxford.

Stiner, Mary C.

1994 Honor among Thieves: A Zooarchaeological Study of Neandertal Ecology.

Princeton University Press.

2002 Carnivory, Coevolution, and the Geographic Spread of the Genus Homo. Journal

of Archaeological Research 10(1): 1–63.

2004 Comparative ecology and taphonomy of spotted hyenas, humans, and wolves in

Pleistocene Italy. Revue de Paléobiologie 23(2): 771–785.

Surovell, Todd A., Spencer R. Pelton, Richard Anderson-Sprecher, and Adam D. Myers

2016 Test of Martin’s overkill hypothesis using radiocarbon dates on extinct

megafauna. Proceedings of the National Academy of Sciences 113: 886–91.

Thébault, Elisa, and Colin Fontaine

2010 Stability of ecological communities and the architecture of mutualistic and trophic

57

networks. Science 329(5993): 853–6.

Van Valkenburgh, Blaire

1999 Major Patterns in the History of Carnivorous Mammals. Annual Review of Earth

and Planetary Sciences 27.

Vasil’ev, Sergey A.

2001 The Final Paleolithic in Northern Asia: Lithic Assemblage Diversity and

Explanatory Models. Arctic Anthropology 38(2): 3–30.

Viau, A. E., K. Gajewski, M. C. Sawada, and J. Bunbury

2008 Low- and high-frequency climate variability in eastern Beringia during the past

25 000 years. Canadian Journal of Earth Sciences 45(11): 1435–1453.

West, F. H. (editor).

1981 The Archaeology of Beringia. Columbia University Press, New York.

Wilson, Lucy

2007 Understanding Prehistoric Lithic Raw Material Selection: Application of a

Gravity Model. Journal of Archaeological Method and Theory 14(4): 388–411.

Wygal, Brian

2017 The peopling of eastern Beringia and its archaeological complexities. Quaternary

International. DOI 10.1016/j.quaint.2016.09.024

Yeakel, Justin D., Paulo R Guimarães, Hervé Bocherens, and Paul L. Koch

58

2013 The impact of climate change on the structure of Pleistocene food webs across the

mammoth steppe. Proceedings of the Royal Society 280: 2013.0239.

Yi, Seonbok, Geoffrey Clark, Jean S. Aigner, S Bhaskar, Alexander B. Dolitsky, Kathleen F.

Galvin, Fumiko Ikawa-smith, Shimpei Kato, Philip L. Kohl, Richard B. Stamps, and

William B. Workman

1985 The “ Dyuktai Culture ” and New World Origins. Current Anthropology 26(1): 1–

20.

Zazula, Grant D., Duane G. Froese, Charles E. Schweger, Rolf W. Mathewes, Alwynne B.

Beaudouin, Alice M. Telka, C. Richard Harington, and John A. Westgate

2003 Ice-age steppe vegetation in east Beringia. Nature 423: 603.

Zimov, S.A., N.S. Zimov, A.N. Tikhonov, and F.S. Chapin III

2012 Mammoth steppe: a high-productivity phenomenon. Quaternary Science Reviews

57: 26–45.

Zimov, S. A., V. I. Chuprynin, A. P. Oreshko, F.S. Chapin III, J. F. Reynolds, and C. Chapin

1995 Steppe-Tundra Transition : A Herbivore-Driven Biome Shift at the End of the

Pleistocene. The American Naturalist 146(5): 765–794.

59

APPENDIX A:

ANIMALS AS RAW MATERIAL IN BERINGIA: INSIGHTS FROM THE SITE OF SWAN

POINT CZ4B, ALASKA

François B. Lanoë1 and Charles E. Holmes2

1School of Anthropology, University of Arizona

2Department of Anthropology, University of Alaska Fairbanks

60

Abstract

We document the use of organic raw material in late Pleistocene eastern Beringia through the study of the site of Swan Point CZ4b, in central Alaska. CZ4b is attributed to the Dyuktai culture and dates to about 14,000 cal B.P. We interpret the occupation as a specialized workshop dedicated to the production and maintenance of organic-based tools following three lines of evidence: (1) limited on-site consumption of megafauna; (2) diversity of organic raw materials and techniques used in processing them; and (3) spatial demarcation of specialized activity areas.

Specialized workshops are located in the vicinity of naturally occurring accumulations of mammoth bones in both western and eastern Beringia and suggest similarities in animal resource use across Beringia for the Dyuktai culture. Organic technology was a major portion of Dyuktai technology in eastern Beringia and its lack of visibility in archaeological assemblages is probably due to taphonomic reasons. Changes in the availability of organic raw material throughout the Late Pleistocene offer some implications on the evolution of lithic technology and material culture.

61

The colonization of Beringia is one of the most remarkable events in human history as people permanently settled the subarctic, coincidently with a vanishing megafauna, and opened the way for the initial colonization of the New World. How and when exactly Beringia was colonized remains debated (Hoffecker et al. 2016). Late Glacial archaeological sites show a diversity of lithic technology including both microblades and bifacial points which reflect different strategies of projectile point manufacture (additive vs. reductive) and are typically associated with the

Eurasian Late Paleolithic and the North American Paleoindian, respectively (Dixon 2013). One of the oldest recognized culture or tradition is Dyuktai, characterized by microblades produced with the Yubetsu technique (Dikov 2003; Holmes 2011; Mochanov 2009), and which originates in the Upper Paleolithic of northeastern Eurasia (Goebel 1999; Goebel and Buvit 2011; Gómez

Coutouly 2011). Other traditions of the Bølling-Allerød are characterized by bifacial projectile points and their significance and relationship with Paleolithic and Paleoindian cultures remain unclear; these include Nenana (Graf et al. 2015; Hoffecker 2001) and Ushki VII (Dikov 2003).

Dyuktai arguably transitioned in eastern Beringia and throughout the Bølling-Allerød and

Younger Dryas into the Chindadn then Denali complexes (Easton et al. 2011; Holmes 2011).

Northern Paleoindian traditions dominated by large lanceolate or fluted bifacial projectile points are also present in eastern Beringia during the Younger Dryas with the Mesa (Hoffecker 2011) and Northern Fluted Points complexes (Goebel et al. 2013).

How eastern Beringian people and animals interacted has been at the core of archaeological research for several decades. Human colonization is contemporaneous of a megafauna turnover; the scale to which humans affected Pleistocene megafauna remains widely debated, particularly when dealing with the extinction of horse (Equus lambei) and mammoth (Mammuthus primigenius) (Buck and Bard 2007; Guthrie 2003, 2006; Haile et al. 2009; MacDonald et al.

62

2012; Solow et al. 2006; Surovell et al. 2016). It has for instance been proposed that eastern

Beringian people were large megafauna specialists gradually diversifying to lesser-ranked prey as mammoth and horse populations dwindled and disappeared (e.g. Potter et al. 2013).

In contrast, the use of animals as raw material is rarely mentioned in discussions about technological variability and functionality (e.g. Goebel and Buvit 2011). Technology based on organic raw materials plays a large part in most hunter-gatherer societies, but recovered organic artifacts in eastern Beringia are rare, limited to a few wapiti (Cervus elaphus) antler, mammoth ivory, and bone tools (Holmes 1996, 2001; Potter 2001; Yesner 2001; Yesner et al. 2000; Potter et al. 2013, 2014; Lyman 2015). In comparison, osseous and wood materials are quite common in western Beringia sites (Giria and Pitul’ko 1994; Khlopachev and Girya 2010; Pitul’ko 1993,

2013; Pitul’ko and Kasparov 1996; Pitul’ko et al. 2004, 2015). Based on the similarity of environments and, arguably, a historical continuity of population and culture, one can expect that organic technology also had an important role among eastern Beringian hunter-gatherers and that our current knowledge of eastern Beringian technology is heavily biased towards lithic materials.

In this article we document the use of animal raw material at the site of Swan Point Cultural

Zone 4b (hereafter CZ4b), in central Alaska. We propose that CZ4b was a specialized workshop dedicated to the production and maintenance of organic-based tools following three lines of evidence: (1) limited on-site consumption of megafauna for food; (2) diversity of organic raw materials and techniques used in processing them; and (3) spatial demarcation of specialized activity areas. We discuss the implications that this site holds for our understanding of the evolution of technology and material culture in eastern Beringia.

63

Swan Point CZ4b: A Specialized Workshop

Site Background

Swan Point (XBD-146) is a multi-component site that dates from the terminal Pleistocene and throughout the Holocene. It is located on top of a ridge (64°18’N, 146°02’W, 322 m asl), a strategic location overlooking up to 86% of the surrounding Shaw Creek Flats and Tanana-

Yukon Uplands (Figure 1).

Figure 1. Location of Swan Point in relationship with sub-contemporaneous sites. A Digital

Terrain Model (USGS 2010) was used; contour line intervals are 200 m (top) and 5 m (bottom).

The shaded area corresponds to the area visible from the site at a distance of 10 km, at which a normal human visual acuity (one arc-minute [NDT Resource Center 2014]) can perceive a 3-m- large mammoth (d = 3/sin(1’) = 10,313 m). Thirty-nine percent of the landscape within this

64 distance is visible from the site. At a radius of 1 km, 86.0 percent of the landscape is visible from the site.

Radiocarbon dates suggest that the CZ4b occupation was restricted to a short timeframe around

14,000 cal B.P., during the Bølling-Allerød interstadial chronozone (Supplemental Text 1).

Archaeological materials were deposited during a time of intense regional sand dune buildup and calcareous loess deposition (Dilley 1998). The spatial distribution and preservation of the fauna suggest a deposition by humans and a rapid post-abandonment burial that preserved the integrity of the site; refits of both fauna and lithic artifacts indicate a single occupation that can easily be separated from components above it (Supplemental Text 2).

CZ4b is arguably the oldest recorded archaeological occupation in eastern Beringia, and also the only component to include microblades produced with the Yubetsu technique; this technique is characteristic of the Dyuktai tradition and is typically found in western Beringia (Gómez

Coutouly 2012; Holmes 2011).

Limited On-Site Megafauna Consumption for Food

Faunal remains at CZ4b include large herbivores, lagomorphs, rodents, and birds (Supplemental

Text 3). Megafauna is dominated by mammoth, followed by horse and caribou (Rangifer tarandus); isolated remains of moose (Alces alces) and bison (Bison cf. priscus) were also identified. Most (98.6%) mammoth remains consist of tusk and cheek teeth fragments from juvenile and adult individuals that were probably acquired from sub-fossil skeletons (see below), but no post-cranial material was recovered for these age classes. The only mammoth remains that present evidence of human consumption for food are a few neonate ribs accounting for a single

65 individual; one of the ribs bears a notch characteristic of dynamic loading impact (Supplemental

Text 3).

Recorded elements for other ungulates are limited to skulls or lower limbs (autopod). While surface preservation does not enable recognition of butchery cutmarks, overall these anatomical parts have little to no nutritional value and these specimens may relate to activities other than food extraction, such as skin or sinew extraction. By contrast, elements typically abandoned at hunter-gatherer sites after episodes of meat or marrow consumption, such as long bone fragments with green breaks, or vertebrae fragments, are absent (Supplemental Text 3).

Birds are dominated by waterfowl, in particular geese (Anserini) and swan (Cygnus cf. columbianus). Element representation for geese suggests they were brought to the site as complete carcasses; on the other hand, for swan, wings, or legless bodies, were likely transported to the site after field processing (Supplemental Text 3).

CZ4b does provide evidence that Dyuktai people hunted or scavenged megafauna, including horse, mammoth neonate, and caribou, and also acquired smaller prey items such as waterfowl and hares; but there is no evidence for hunting or scavenging of juvenile and adult mammoth.

Faunal remains suggest that food preparation and consumption was not a major activity at CZ4b, and that length of stay at the site was limited. In contrast to other Beringian sites (e.g. Potter

2005), there is no evidence at CZ4b of primary butchery and large-scale consumption of ungulates. Nutritious body parts may have been transported away for eating, or possibly processed at another location or in a non-excavated area of the site.

66

Organic Technology: Diversity of Materials and Techniques

Differential preservation of osseous specimens at CZ4b generally hinders recognition of human modifications; for most specimens little to no portion of the original surfaces was preserved

(Supplemental Text 2). Evidence for organic technology is limited to inferences and a few direct observations.

Lithic artifacts provide indirect insights about organic technology (Holmes 2011, also Figure 5).

Thirty-two burins were recorded in CZ4b, and 34% were refitted to burin spalls, demonstrating that burins were used and rejuvenated at the site. Burins can be used for various tasks, but are generally more fit for engraving, carving, and grooving organic materials (Barton et al. 1996).

Numerous Yubetsu microcores, microblades, and ski spalls (some of which were refitted to microcores) were also recorded, indicating that large quantities of microblades were produced at

CZ4b. They were most likely used at the site in the production and maintenance of composite tools. Indeed, Yubetsu microcores are an efficient way of transporting future microblades (Elston and Brantingham 2002) and there would be little incentive to produce them before their immediate use.

Evidence for antler work includes the beam of a shed adult male caribou antler from which cortical flakes were extracted (Figure 2). The two halves were separated along a fresh surface and recovered about 2 meters from each other, suggesting the beam was split, probably by inserting a wedge in the spongious portion at the base (i.e., cleaving). While surface impacts are not preserved, green break scars suggest repeated impacts using the newly created surface as a platform (Figure 2). Fracture of caribou antler by cleaving and direct percussion, using an anvil as support, was widespread in the Upper Paleolithic (Baumann and Maury 2013). Typically, the desired end-products were not the short and wide flakes created by percussion, but rather the

67 straight, elongated pieces of the beam or second spike from which flakes were extracted - shaping/façonnage rather than débitage (Pétillon and Ducasse 2012). During the Upper

Paleolithic, other methods have been used for producing elongated antler pieces, including splitting (using a wedge longitudinally, Baumann and Maury 2013) and “groove and splinter” techniques (Goutas 2009). The advantage of direct percussion in regards to other techniques is speed, but it is less efficient in terms of raw material conservation (Pétillon and Ducasse 2012).

Figure 2. Shed right antler of an adult male caribou, with evidence of cleaving and percussion

(specimens #22244 and #22303). The two halves separate and were recorded about 2 m from each other (Figure 6).

Few mammoth tusk specimens display explicit evidence of technological work, even as they hold a large place in the faunal assemblage (75% of non-rodents identified specimens). The lack 68 of visibility of technological traces is most likely due to poor bone surface preservation

(Supplemental Text 2).

A mammoth calf tusk is beveled on the proximal end (Figure 3) - the bevel may have been produced by scraping/abrading in order to facilitate the removal of the pulp cavity (similar to specimens in Pitul’ko et al. 2015). Beveled cylindrical elongated objects (rods) of similar dimensions, and made of ivory, antler, or bone, have been recovered in both Paleolithic and

Paleoindian contexts; functional interpretations have varied between projectile point, foreshaft, wedge or pressure flaker (Derev’anko et al. 1998; Hemmings 2004; Lyman et al. 1998; Pitul’ko et al. 2004, 2015; Potter 2005, Potter et al. 2014; Yesner 1994, 2001; Yesner et al. 2000).

Without further evidence it is unclear what the CZ4b artifact was intended to be.

Figure 3. Beveled mammoth calf tusk, medial and lateral views (specimen #20785). Its diameter varies between 23.3 and 23.5 mm for a chord length of 256.3 mm. The distal end is broken.

An adult mammoth tusk displays evidence of percussion (Figure 4). Rather than a failed attempt to produce flakes, percussion probably denotes the use of the tusk as an anvil. This concurs with

69 evidence, from antler specimens, of anvil use (see above), and the absence of recovery of large rocks at the site. On the other surface of the same specimen, a large (50 by 3.3-4.0 cm) ivory splinter was extracted by the groove and splinter method: burins were used to groove through the soft outer layers of dentine, before levering the desired blank with a wedge or chisel (Christensen

1999). The use of this method on ivory was widespread in the Upper Paleolithic (Christensen

1999; Hahn et al. 1995), including examples from northeastern Eurasia (Pitul’ko et al. 2015;

Semenov 1964), and probably Alaska (Gelvin-Reymiller et al. 2006). The end-product was an elongated splinter similar to those recovered at Yana (Pitul’ko et al. 2015), later modified for various purposes: ornament, projectile point, or foreshaft (Hahn 1995; Pitul’ko et al. 2015; White

1995).

Figure 4. Left adult tusk from a ~15-year-old mammoth female, medial view, with evidence of technological work: percussion (above, medial view) and groove and splintering (below, ventral view) (specimen #21444).

70

Beyond these large artifacts, it is likely that most recorded ivory fragments represent débitage debris. Additional ivory artifacts must have included flakes obtained by percussion (Heckel and

Wolf 2014; Khlopachev and Girya 2010), as is known in northeastern Eurasia (Pitul’ko and

Kasparov 1996; Pitul’ko et al. 2015). At CZ4b the poor bone surface preservation generally hinders the recognition of such artifacts.

Wings of swans were valued enough to be selectively transported to the site (see above). Little nutritional value is offered by wings; rather, this selective representation suggests a focus on the feathers (Bovy 2002, 2012; Finlayson et al. 2012) and the long, straight ulnae and radii that have been used as awls and tubes throughout prehistoric times (Bovy 2002, 2012). Similarly, spatial distribution of geese elements suggests wings were given a special treatment (see below).

Waterfowl feathers used for fletching, along with those of other birds, have been recovered from prehistoric contexts in the Alaska-Yukon ice patches (Dove et al. 2005; VanderHoek et al. 2007); today, goose feathers are still a favorite among traditional bowyers (Allely et al. 1992).

Overall, combined information from lithic and organic materials at CZ4b indicate a specialized focus on the production and/or repair of composite tools. Grooved organic blanks as well as microblades were manufactured on site. Most likely, shafts, fletching and hafting implements

(animal/vegetal glue, sinew) were produced as well. The composite end-products, whether they were knives or projectiles, must have been similar to those made of bone, antler and ivory, recovered from Chernooz’or’ye II, Kurla II/III (Derev’anko et al. 1998), and Zhokhov Island

(Giria and Pitul’ko 1994).

71

Spatially Demarcated Activity Areas

There are three major concentrations of remains with densities of the same magnitude (Table 1,

Figure 5, Figure 6): to the south (N 91-95, E 97-101), the north (N 97.5-101, E 96.5-99.5), and the east (N 95.5-98, E 100.5-102). The south and north clusters are similar in their material composition (Table 1; χ2 = 3.8; df = 3; p = .29) and are both organized around dispersed patches of burned organic residues. Burned patches are described in detail elsewhere (Crass et al. 2011;

Holmes 2011; Kedrowski et al. 2009). Briefly, these patches include fatty residues from animals

(ruminant and monogastric herbivores) as well as plants, and comparatively little charcoal. Based on experimental work they are most consistent with hearths fueled by bone and other materials such as herbivore dung or graminoids. The east cluster is nearly entirely composed of mammoth tusk fragments (Table 1). The rest of the site is composed of low-density areas with scattered remains.

Cluster Surface N Density Lithics Bone Ivory Charcoal &

(m2) (N/m2) (%) (%) (%) Burnt Residue

(%)

North 10.5 1,430 136.2 76.6 12.4 3.4 6.9

South 16.0 4,012 250.8 74.8 19.4 3.0 2.4

East 3.8 455 121.3 13.4 1.8 83.1 1.8

Table 1. Composition of identified spatial clusters.

72

Ethnoarchaeological and archaeological research in the past 40 years has helped define patterns of space use that are universal among hunter-gatherer societies. In particular, activity areas are defined by the presence and/or absence of particular types of artifacts. Activities such as flint- knapping are best defined by the presence of debris related to the activity conducted, rather than by the tools or cores themselves which are generally removed from the activity loci through recycling or cleaning (Binford 1978, 1983; Clark 2015; Surovell and Waguespack 2007;

Vaquero and Pastó 2001; Waguespack and Surovell 2014). The distribution of different types of archaeological materials at CZ4b (Figure 5, Figure 6) informs also on the types and location of activities conducted.

Microcore rejuvenation flakes and burin spalls relate to the production of microblades and the use and rejuvenation of burins in both north and south hearth areas, while microcores and burins are preferentially located further away (Figure 5). Composite tool production and/or maintenance were thus located in the immediate vicinity of hearths. Production and maintenance of composite tools require not only skill but also extensive time investment (Elston and Brantingham 2002) and are best executed near light and warmth, as has been documented in other archaeological contexts (Bodu et al. 2006; Leroi-Gourhan and Brézillon 1966, 1972; Olive and Pigeot 2006;

Pigeot 2010; Potter 2005; Wenzel 2011).

73

Figure 5. Swan Point CZ4b refits: (a) microblades, microcores, and platform rejuvenation “ski” spalls; (b) burins and burin spalls. Left modified from Potter et al. (2013).

Fragments from highly nutritious anatomical parts, such as mammoth ribs and horse hyoid, indicate a food consumption activity, and are preferentially located near the hearths (Figure 6).

Whereas, “bulky” elements, such as vertebra, or elements that have little to no nutritional value, such as phalanges and metatarsals, are absent or located further from the hearth areas in relatively empty zones. “Empty zones” are characteristically locations where activities that need

74 a lot of space but generate little debris are conducted, such as primary butchering, skinning and hide preparation, or sinew extraction (Audouze 2010; Binford 1983; Olive and Pigeot 2006). At

CZ4b such areas may have been located away from hearth areas; for instance, metatarsals and phalanges are consistent with remains from skin preparation and/or sinew extraction (Figure 6).

In the case of primary butchering, assuming such areas existed at CZ4b and that low-density elements such as vertebra were preserved, they are probably located in non-excavated parts of the site.

The majority of the recovered mammoth ivory fragments occur in the east cluster (Figure 6).

This cluster is at the edge of the excavated area, and future finds will probably provide additional elements of interpretation. Nevertheless, a concentration of small ivory fragments suggests a focus on ivory knapping, percussion, or other types of technological methods recognized on the faunal remains (see above). At CZ4b, large ivory fragments may have been moved away from this zone for technological use, if end-products; or for recycling or cleaning purposes, if by- products.

Bird bones are differentially located across the site. Swan and geese bones are segregated between north and south hearth areas, respectively (Figure 6). Within the south hearth area, geese hindlimb elements are preferentially located towards the center of the hearth while forelimb elements are located 1-2 m farther away (Figure 6), suggesting that legs and wings underwent different treatment. Legs could have been used primarily for food consumption, and dropped in the fire; while wings would go through further processing for feather extraction and/or bone technology.

75

Figure 6. Spatial distribution of (a) bird and (b) mammal remains, CZ4b. The two caribou antler pieces to the south refit. The horse cheek teeth most likely articulate together, as well as the mammoth calf cheek teeth and the mandible (see text).

76

A Specialized Workshop

A major function of CZ4b was as a workshop for producing composite tools and/or weapons.

Manufacturing activities were paramount and well demarcated in space. Other activities such as acquiring and processing food may also have been conducted at the site but are less visible archaeologically.

Whether mammoth tusks were used fresh or scavenged from subfossil individuals has important implications on the interpretation of CZ4b. Mammoth tusks are much easier to extract (from the jaw) and modify when they are subfossilized, that is when the structural collagen has partially decayed (Christensen 1999, Pitul’ko et al. 2015). Experiments have shown that while producing conchoidal flakes is easiest on frozen fresh ivory (Khlopachev and Girya 2010), grooving is up to 20 times faster with degraded rather than with fresh ivory; regardless of any prior chemical

(e.g. vegetal acids, urine) or physical (e.g. soaking, steaming, freezing) treatment (Christensen

1999; Steguweit 2015). The 50 cm long splinter from CZ4b would have been extracted in as little as 2 hours on degraded ivory, in contrast to about 40 hours on fresh ivory (Christensen

1999).

Working fossil ivory was common in the Paleolithic. At the nearby sites of Mead and Broken

Mammoth, ivory several thousands of years older than the human occupation was used (Yesner

2007). For other sites, such as CZ4b, where dated ivory is contemporaneous with human occupation, people may have deliberately come to places on the landscape where they knew ivory was to be found: whether that be where mammoth were likely to die, e.g., “mammoth cemeteries”, or where people had killed a mammoth some time before and kept the ivory “in storage” (Pitul’ko et al. 2015; Steguweit 2015). Regardless of the scenario, at CZ4b, the absence of post-cranial skeletal material assigned to the same age categories as those for which tusks

77 were recovered (see faunal remains) concords with the idea that at least some tusks were acquired as subfossils. This is particularly true of the large adult tusk (Figure 4), on which a direct radiocarbon date is slightly older than those for the rest of the component (AA-98488,

Supplemental Text 1).

Figure 7. Interpretation of the use of space at Swan Point CZ4b.

78

Swan Point must have been in close proximity to an ivory source as people probably did not carry large and heavy tusks for long distances. The site’s position on the landscape is striking on several accounts; it lies at the interface of uplands and lowlands, with high visual control of the surrounding landscape (Figure 1). If mammoth bones also occurred in the vicinity of the site,

Swan Point was the ideal location for a workshop. As such CZ4b is very similar to the sites of

Yana and Berelekh in northeastern Eurasia, archaeological sites located in the vicinity of massive, naturally occurring accumulations of mammoth bones, which people used for raw material as “ivory mines” (Basilyan et al. 2011; Pitul’ko 2011; Pitul’ko et al. 2014, 2015). The

Yana site also includes areas with high concentrations of ivory artifacts and debris suggesting specialized manufacturing areas (Pitul’ko et al. 2015).

In contrast, later sites in eastern Beringia are different in their artifact composition and spatial patterning, and most have been interpreted as more residential camps where a diverse range of activities were conducted (e.g. Potter 2005). CZ4b can probably be considered as a logistical foray, dedicated to resupplying on composite tools, within a larger settlement strategy at the landscape level.

Organic Technology in Beringia

Combined faunal and spatial analyses at CZ4b confirm that organic materials were part of eastern Beringian technology, and that their low visibility in archaeological sites in comparison to assemblages from western Beringia can probably be attributed to taphonomic factors.

Based on the limited evidence available, Dyuktai people in eastern Beringia had a similar relationship with animals than their Eurasian counterparts. Dyuktai people preyed on the same

79 animal species on either side of the Bering Strait. For instance, prey at both CZ4b and Berelekh includes horse, caribou, hare (Lepus spp.), geese (Anserini), ptarmigan (Lagopus spp.), and, at least at CZ4b, neonate mammoth (Lanoë and Holmes 2016; Pitul’ko 2011). Dyuktai people thus relied on a diversity of animals including high-ranking prey such as neonate mammoth and other megafauna, but also smaller prey such as birds. In addition, both Berelekh and CZ4b may have been used at least in part for processing materials derived from sub-fossilized skeletons of juvenile and adult mammoths.

The diversity of techniques for working organic raw materials is obvious at CZ4b despite the small sample size. Methods of extraction included cleaving, direct percussion on anvil, and groove-splintering. Methods of shaping included scraping/abrading and most likely grooving.

Given that Dyuktai people possessed a great breadth and depth of technological skills, their choice of methods in a given circumstance must have depended on which currency was favored.

If time was the most important limitation, percussion would likely have been preferred; if raw material quantity was a limiting factor, a method such as groove and splinter would have been more suitable. Dyuktai people at CZ4b also took full advantage of the different mechanical and physical properties of the variety of osseous raw materials available to them. Antler is elastic and tough (Currey 1979; Margaris 2014) and was probably used for producing projectile parts. Ivory is massive, compact and homogeneous (Christensen 1999), its shape is ideal for producing elongated objects or objects with complex shapes. Bone is strong and abundant but somewhat more brittle (Currey 1979; Margaris 2014) and may have been favored for tools used by pressure rather than percussion, such as large mammal bones for composite knives (e.g. Pitul’ko and

Kasparov 1996), and elongate bird bones for needles, tubes and awls (Bovy 2002, 2012).

80

The evidence for organic technology at CZ4b holds some important implications on the evolution of material culture in eastern Beringia. After about 14,000 cal B.P., changes occurred both in the environment and the economy and technology of eastern Beringian people. With the development of more forested habitats (Bigelow and Powers 2001; Llyod et al. 2006; Potter

2008, 2011), mammoth, horse and caribou became locally and/or regionally extinct, and human diet shifted towards a heavy emphasis on bison and wapiti (Lanoë 2015; Lanoë et al. 2016;

Potter 2007; Potter et al. 2013; Yesner et al. 2011), two species more tolerant of more closed habitats. While microblades were still present in Chindadn and Denali toolkits, they were produced primarily with the Campus method rather than the Yubetsu method characteristic of

Dyuktai (Gómez Coutouly 2012). Mammoth artifacts made from fossil ivory, particularly rods, were still in use after 14,000 cal B.P., but were quite rare (Holmes 1996, 2001; Potter 2001;

Yesner 2001; Yesner et al. 2000). Bifacial projectile points first appeared and were used, at least in some cases, in association with foreshafts made of wapiti antler (Potter et al. 2014). Projectile bifacial points (reductive technology) are fundamentally different than projectile points based on microblades (additive technology [Dixon 2013]) by the amount of skills involved in making them, but also in terms of strategy of resource acquisition (Elston and Brantingham 2002).

At least some changes in technology and material culture may be related to renewal of cervid species. Caribou antler has especially thick cortical walls and is ideal for producing projectile points that can resist impact stress (West 1981, Guthrie 1983), therefore is best suited for composite projectile points with microblade insets. Indeed, the vast majority of recorded composite projectile points in both Eurasia and North America were made on antler; when the species is mentioned, it is caribou (Gómez Coutouly 2011; Pétillon et al. 2011). Wapiti antler is less tough but straighter and may have been preferred for elongated objects that were subjected

81 to less stress, such as projectile foreshafts (Guthrie 1983; Lyman et al. 1998). As such, evidence from CZ4b suggests that environmental change may have indirectly influenced the evolution of material culture, particularly lithics, by changing the availability of raw materials and specific megafauna species on the landscape.

82

Supplemental Text 1: Radiocarbon Dating

Figure 8. Calibration of radiocarbon dates for CZ3 and CZ4. Calibration was conducted with

OxCal 4.2 and IntCal13 (Bronk Ramsey 1994, 2009, Reimer et al. 2013). Dates preceded by an asterisk are reported here for the first time; others can be found in Potter et al. (2013) and

Holmes (2011). Samples for β-190577, β-186682 and β-215328 were obtained from rodent burrows in CZ4b (see Supplemental Text 2).

83

Lab Material Age BP SD Calibrated Age BP at 2σ

Number

β-175489 burned bone collagen 11,360 50 13,300-13,090

β-186682 bone collagen 4,760 40 5,550-5,420

β-190577 charcoal 4,670 40 5,480-5,400

β-215328 charcoal 8,460 60 9,550-9,390

β-355852 charcoal 11,350 50 13,290-13,090

β-355867 charcoal 10,160 50 12,050-11,510

β-357816 charcoal 10,070 40 11,920-11,360

β-365062 dentine collagen 12,170 50 14,210-13,860

β-401126 charcoal 10,620 40 12,700-12,530

Table 2. Radiocarbon dates newly reported here.

84

Supplemental Text 2: Patterns of Deposition and Overall Taphonomy

There is little doubt as to the anthropogenic nature of the accumulation of the faunal material.

Most specimens are spatially associated with lithic artifacts and hearths (Figures 5, 6); indeed, recorded fauna and lithic artifacts spatially co-vary both at a resolution of .25 m2 (Spearman’s ρ

= .62; p < .01) and 1 m2 (ρ = .72; p < .01). While the exterior surface of most faunal specimens is very poorly preserved (e.g. Figures 2, 3, 4) and obscures any traces of butchery marks or carnivore ravaging, there is ample evidence of osseous toolmaking and one probable human impact of percussion recorded on a mammoth rib (Supplemental Text 3).

The origin of accumulation is less straightforward for rodents (ground squirrels, voles and jumping mice). Cricetid and Dipodid remains weakly co-vary (i.e., are independent) with other recorded archaeological remains both at a resolution of .25 m2 (Spearman’s ρ = .19; p < .01) and

1 m2 (ρ = .33; p < .01). The taphonomic processes recorded on the Cricetid-Dipodid sample are different than for the rest of the assemblage (Figure 9): no remains were found isolated, 83% were recorded unbroken, and with weak modifications due to weathering and dissolution. In addition, radiocarbon dating of a Dipodid bone and a charcoal associated with a Dipodid skeleton yielded dates of 5,550-5,420 and 5,480-5,400 cal B.P., respectively (Figure 8).

Altogether, it is most likely that Cricetid-Dipodid specimens were deposited as intact bodies under the surface (i.e. in burrows or nests) or in a context of rapid deposition, in a manner unrelated to predation.

Ground squirrel remains weakly co-vary with other recorded archaeological remains both at a resolution of .25 m2 (Spearman’s ρ = .08; p = .15) and 1 m2 (ρ = .18; p = .08). Recorded taphonomic processes are generally intermediate between those observed for Cricetid-Dipodid remains and for the rest of the assemblage (Figure 9). Some of the remains cluster in areas of the

85 site where otherwise few archaeological remains were recovered, while others are associated with dense artifact concentrations and human features (Figure 6). A charcoal sample associated with a ground squirrel-sized burrow was dated to 9,550-9,390 cal B.P. (Figure 8). Overall it is likely that some or all of the squirrel remains were deposited after human abandonment of the site, within krotovinas/burrows where the animals died during hibernation from thermal or nutritional stress (Morrison and Galster 1975). The location of burrow excavating may have been favorably influenced by the accumulation of artifacts and/or sediment disturbance by humans; alternatively, a few individuals may have been preyed upon such as at the nearby site of Upward

Sun River (Potter et al. 2011, 2014).

Burrowing by ground squirrel and Cricetid-Dipodid may have caused some bioturbation within

CZ4b and between CZ4b and the components above; this perturbation was on a limited scale though, as it did not affect the 845 recorded mammoth teeth fragments, which were all recovered in CZ4b (Figure 10).

Weathering had an overall low impact on the assemblage (Figure S2), as expected in the context of rapid sediment deposition that occurred between 14,000 and 8,000 cal B.P. throughout the

Shaw Creek basin (Dilley 1998; Reuther 2013). The extent of surface dissolution is high for the non-rodent specimens (Figure 9); on most specimens little to no portion of the original bone surface was preserved. For both birds and mammals, dissolution of the outer surfaces must have erased any potential butchery marks.

Bone dissolution is caused by reaction of the structural hydroxylapatite with acids (Andrews

1990; Lyman 1994; Stiner et al. 2001). At CZ4b it potentially derived from three types of agents: carnivore digestion/salivary enzymes; plants obtaining nutrient through their root system; or

86 acids in solution in the sediments. On many specimens, manganese oxide surface coating is interrupted by dissolution patterns; thus dissolution occurred after oxide coating, when the bones were already buried, ruling out carnivore damage. While some specimens still have roots of modern vegetation attached to them, overall specimens lack the characteristic dissolution pattern expected from root-etching, which is dendritic and preferentially located on one side of the bone.

Rather, specimens show a general surface pitting more characteristic of dissolution from acids in solution in the sediment (Andrews 1990), a pattern that can be found at other sites in the region

(e.g. Potter 2005).

Today the sediments in which CZ4b is located are slightly alkaline (pH 7.5-7.7, Dilley 1998); therefore bone dissolution is not due to acid leaching in the current inceptisol, but may rather be related to late Pleistocene or early Holocene periods of soil development. Regardless of this exact timing, bone dissolution at CZ4b is post-depositional, and the lack of biotic modification is consistent with rapid burial of archaeological materials after human abandonment of the site.

87

1 1 1 0,8 0,8 0,8 0,6 0,6 0,6 0,4 0,4 0,4 0,2 0,2 0,2

0 0 0 0 25 50 75 100 SO 0 1 2 3 4 5 6 SO 0 25 50 75 100 SO

Extent of manganese oxide Weathering stages Extent of surface dissolution

coating (increments of 25%) (Behrensmeyer 1978) (increments of 25%)

Fraction of total NISP total of Fraction 1 1 400 0,8 0,8 300 0,6 0,6 200 0,4 0,4 0,2 0,2 100 0 0 0 0 1 2 3 4 5 6 00-10-1 .1 100-1011 10 10102-1030 1000

3 Breakage patterns Combustion stages Specimen weight frequency,

(Stiner et al. 1995) whole assemblage (in g)

Figure 9. Taphonomic processes affecting remains of Cricetid-Dipodid (in black), Sciurid (gray), and the rest of the faunal assemblage (stripes). SO: surface obstructed, no recording.

88

Figure 10. North-south profile (facing west) of the lower layers of Swan Point. Scale is in meters, elevation is relative to the modern surface.

89

Supplemental Text 3: Faunal Remains

NISP MNE MNI

Mammalia

Artiodactyla

Bovidae

Bison sp. Bison 1 1 1

Cervidae

Alces alces Moose 1 1 1

Rangifer tarandus Caribou 3 2 2

Unidentified Cervid (Antler) 1 - -

Perissodactyla

Equidae

Equus lambei Yukon Horse 10 4 1

Proboscidea

Elephantidae

Mammuthus primigenius Woolly Mammoth

Skull 1 1

Cheek teeth 127 2 4 Incisors 718 3

Ribs 12 8

Lagomorpha

Leporidae

Lepus (Lepus) sp. Unidentified Arctic Hare 2 2 1

Lepus (Poecilolagus) americanus Snowshoe Hare 2 2 1

Rodentia

Cricetidae

Myodes rutilus Northern Red-Backed Vole 3 3 1

90

Microtus oeconomus Tundra Vole 3 3 2

Microtus xanthognathus Taiga Vole 3 3 1

Microtus/Myodes Unidentified Vole 2 2 -

Dipodidae

Zapus hudsonius Jumping Mouse 215 210 10

Sciuridae

Urocitellus parryi Arctic Ground Squirrel 79 69 4

Undetermined Order

Mammal Size 2 1 1 -

Mammal Size 5 7 5 -

Aves

Anseriformes

Anatidae

Anas sp. Teal 2 2 1

Anas sp. Pintail, or Wigeon 1 1 1

Anatinae Unidentified Large Duck 1 1 -

Anser fabalis / Branta canadensis Unidentified Large Goose 1 1 -

Anserini Unidentified Goose 33 31 4

Anatidae Unidentified Goose or Duck 8 8 -

Cygnus columbianus Tundra Swan 3 3 -

Cygnus sp. Unidentified Swan 5 5 2

Galliformes

Phasianidae

Lagopus lagopus Willow Ptarmigan 5 5 1

Lagopus leucurus/mutus Rock or White-tailed Ptarmigan 1 1 1

Lagopus sp. Ptarmigan 2 2 -

91

Undetermined Order

Bird Size 2 2 2 -

Bird Size 4 1 1 -

Bird Size 5 1 1 -

Total 1,257 383 38

Table 3. Specimen identification for Swan Point CZ4b. Quantification units follow the definitions in Lyman (2008). Mammal and bird size classes are defined in Table 4.

Size Class Possible Genera (Non-Exhaustive)

Mammal 1 Microtus, Myodes, Zapus

Mammal 2 Urocitellus, Marmota, Tamiasciurus

Mammal 3 Lepus, Vulpes, Castor

Mammal 4 Canis, Ovis, Saiga

Mammal 5 Cervus, Bison, Alces, Equus

Mammal 6 Mammuthus

Bird 1 (not assigned)

Bird 2 Lagopus, Tympanuchus, Dendragapus, Bonasa, Anas, Aythya, Bucephala

Bird 3 Anas, Aythya, Bucephala, Clangula, Histrionicus, Melanitta, Mergus

Bird 4 Branta, Anser, Chen, Pandion, Buteo, Bubo

Bird 5 Cygnus, Aquila, Haliaeetus, Grus, Pandion

Table 4. Size classes used for this study.

92

Mammoth

Three types of mammoth remains were identified, including incisors (i.e. ivory, or tusk), cheek teeth, and bones. Cheek teeth fragments account for a MNE of 2; two well-preserved fragments yielded widths of 51.30 and 51.66 mm, falling in the size range of the fourth premolar

(dp4/M3) and the first molar (m1/M4) (Maschenko 2002; Metcalfe et al. 2010; Roth and

Shoshani 1988). No occlusal wear was present; based on eruption and wear stages, the individual represented must have been around two (if a dp4) or five (m1) years old (Haynes 1991;

Maschenko 2002). One mandible fragment can be assigned to a three to five year-old calf based on the estimated length of the mandible (30 to 40 cm, to compare with specimens in Maschenko

[2002]).

A large left tusk (Figure 4) was previously attributed to a 15-year-old female based on growth rings (Holmes et al. 2012), and another fragment of similar diameter could be from the same individual. A small broken tusk missing the pulp cavity (Figure 3), with a chord length of 256 mm and a maximal diameter estimated to 26-30 mm can be assigned to an individual between six and eight years old (Maschenko et al. 2006). Other tusk fragments consist in numerous, small and non-descript splinters.

Eight ribs were recorded, of which at least four were attributed to a very young calf, one or a few months old (Holmes et al. 2012). Green breaks were recorded on two ribs, and one of them displays a notch characteristic of dynamic loading impact (Figure 11); it is likely that part of the ribcage was butchered by humans and consumed.

93

Overall at least four individuals are represented: a young prime adult of about 15 years old, a calf about five years old, a calf older than six years old, and a neonate. Evidence of human procurement for consumption, whether hunting or scavenging, is only present for the neonate.

Figure 11. Mammoth infant rib fragment, medial view (specimen #15941). Side is undetermined; the proximal side (towards the rib-head) is to the right. For comparison, a left rib (position 10 to

13, not to scale) is displayed on top. A notch characteristic of dynamic loading impact (Capaldo and Blumenschine 1994) is present on the mesio-proximal portion of the blade.

Other Ungulates

Bison (either Bison priscus, B. antiquus, or B. occidentalis, see Shapiro et al. 2004; Wilson et al.

2008) is represented by a single cranial fragment. Caribou is represented by a juvenile metatarsal

94 and a male adult shed antler (Figure 2). Yukon horse (Equus lambei) was identified from cheek tooth measurements (Burke and Cinq-Mars 1996) and is represented by at least two opposite M3 of similar dimensions and wear, which can be assigned to an individual 3-5 years old (following

Guadelli 1998), and a stylohyoid. Moose (Alces alces) is represented by a distal phalanx.

Unidentified ungulates are represented by rib and tooth fragments.

Apart from the ribs, the elements recorded have limited to no nutritional value (Outram and

Rowley-Conwy 1998). Bison and horse brains and tongues, as well as cervid feet, may have been used primarily for other purposes than food, such as hide tanning or sinew extraction. In contrast to other Beringian sites (Potter 2005), there is no evidence at CZ4b of primary butchery and large-scale consumption of ungulates; nutritious body parts may have been transported away for consumption and/or processed at another location or in a non-excavated area of the site.

Moose and bison are only represented by one specimen each. Because both species are present in

CZ3 (Lanoë 2015), and in the absence of direct radiocarbon dating, it cannot be excluded that their presence in CZ4b is due to contamination from the components above. However this issue has no bearing on the arguments presented throughout this article.

The mammoth and horse remains from CZ4b are among the most recent occurrences of their genus in eastern Beringia, while the moose, if in situ, is among the earliest (Guthrie 2006).

Lagomorphs and Rodents

Hares are represented by four elements. Two subgenera were likely to be present in eastern

Beringia at the Pleistocene-Holocene transition: the arctic hare s.l., Lepus (Lepus), for which possible species include Lepus othus and Lepus timidus (Waltari and Cook 2005), and the

95 snowshoe hare, Lepus (Poecilolagus) americanus. Based on measurements both are present in

CZ4b.

Rodents are represented by the Cricetid, Sciurid, and Dipodid families (Table 3). Cricetid species identification followed Semken and Wallace (2002). Human predation is at best limited to the arctic ground squirrel (Supplemental Text 2). Skeletal part distribution shows a relatively even representation of body parts with high-density elements suggesting intact bodies were brought to the site.

Birds

At least six species of birds were identified in the assemblage including tundra swan, a large goose, a large duck, a teal, the willow ptarmigan and a smaller ptarmigan (Table 3). Anatidae elements are notoriously difficult to identify to species, and this is particularly true for eastern

Beringia. While the American central and Pacific migratory flyways might have been in place by the late Pleistocene (Yesner 1996), Palearctic species may also have held breeding grounds in eastern Beringia at a time when the Bering Strait was not a major biogeographic barrier (Buehler et al. 2006; Salomonsen 1972), and when Nearctic species could have established preferential breeding grounds south of the Laurentide icesheet (Moss and Erlandson 2013). To the potential

27 species of terrestrial Anatidae that occur today in Alaska can be added six strictly Palearctic species; as a result, identification of Beringian Anatid is difficult, particularly in the absence of dated well-preserved assemblages (see Fitzgerald 1991; and Harington and Cinq-Mars 2008 for non-dated assemblages). A detailed rationale for CZ4b specimen identification is presented in the next section.

96

Goose is the most common kind of bird with at least four individuals (Table 3). Skeletal part distribution is relatively even for high-density elements suggesting geese were brought to the site as complete carcasses. On the other hand, swan is only represented by forelimb elements, including humeri, radii, and ulnae, accounting for at least one adult and one juvenile. The swan skeletal part pattern suggests that wings, or possibly legless bodies, were transported to the site after field processing. Considerations of weight may have been more important for the comparatively heavy swan (up to 9.6-12.7 kg, Brazil 2009) than for geese (up to 3.0-5.1 kg, ibid.). Alternatively, the difference in skeletal pattern may be related to organic technology (see below).

Other waterfowl specimens include large duck, teal, and unidentified Anatid (Table 3). The dissolution of extremities hinders recognition of juvenile individuals; nevertheless, five, comparatively small, geese elements, and eight elements that could not be identified to tribe level probably belong to juvenile geese.

In addition to waterfowl, specimens of at least two individuals of two different ptarmigan species were recovered, and two bird specimens could not be identified to family (Table 3). A total of

489 small rounded pebbles, arguably gastroliths (Hoskin et al. 1970), were recorded; but no eggshell.

90% of the articular ends of birds long bones were absent, leaving only diaphyses and hindering identification to species level. This preservation pattern is most likely due to the bones light build

(Serjeantson 2009) making them vulnerable to dissolution, rather than human modification (only

17% show green breakage).

97

Once identified to the Anatidae family, specimens were assigned to classes, including swan, large duck (Anas acuta, Anas americana, Anas clypeata, Anas penelope, Anas platyrhynchos,

Aythya americana, Aythya fuligula, Aythya marila, Aythya vasilineria, Bucephala clangula,

Bucephala islandica, Clangula hyemalis, Histrionicus histrionicus, Melanitta spp., Mergus spp.), and small duck (Aythya affinis, Aythya collaris, Anas americana, Anas carolinensis, Anas clypeata, Anas discors, Anas formosa, Bucephala hyemalis, Mergellus albellus). Small or young geese can be differentiated from large ducks on most long bones specimens based on the ratio length/width of the bone, higher in geese.

Specimens identified to swan were measured when possible (non-standard measurements) and compared with comparative specimens of Cygnus columbianus and C. buccinator. The three possible species of swan range in size from C. columbianus (length 120-150 cm, wingspan 167-

225 cm) to C. cygnus (140-165, 205-243) and C. buccinator (150-180, 230-260) (Brazil 2009).

Figure 12 suggests at least two individuals are represented here. Three specimens (1, 2, 3) fall close to the measurements for C. columbianus. The fourth measurement cannot be attributed to either species. The fifth is particularly small, and probably represents a juvenile. Overall, attribution to C. columbianus is likely for at least three specimens.

Specimens identified to goose were compared to comparative specimens from Anser albifrons,

Chen caerulescens, and Branta canadensis (no comparative specimen was available for Anser fabalis, A. erythropus, and Chen rossii). Geese bones are notoriously difficult to identify to species and even genus (Bacher 1967), even more so when fragmented. Only one ulna was large enough (element length estimated to 165 mm) that it could be attributed to a large goose, either

Branta canadensis or Anser fabalis (Gilbert et al. 1981).

98

1,4

1,3 Cygnus

1,2 )

1,1 Swan Point CZ4 Cygnus columbianus 1 columbianus Cygnus buccinator 0,9

0,8 Measurement(Standardized to 1 2 3 4 5 Specimens

Figure 12. Comparison of measurements of swan elements from Swan Point CZ4b with references for Cygnus columbianus and C. buccinator.

Eight specimens could be identified to the Anatidae but it was not possible to differentiate between a small (or juvenile) goose and a large (adult) duck.

Three adult duck specimens were identified to the genus Anas following Woelfle (1967). One was in the size range of the larger Anas ducks (Anas (Dafila) acuta, A. (Mareca) americana/penelope, A. (Anas) platyrhynchos), and two were in the size range of the teals (Anas

(Tettion) crecca, A. (Spatula) discors, A. (Sibirionetta) formosa).

Another specimen could be identified as a large duck but not at the genus level.

Once identified to the Phasianidae family, specimens were compared to comparative specimens of Lagopus leucurus (also used as a substitute for the similar-sized Lagopus mutus), Lagopus lagopus, Tympanuchus phasianellus, and Dendragapus obscurus (used as a substitute for the

99 similar-sized Dendragapus canadensis). I did not have access to specimens of Bonasa umellus which is similar in size to Tympanachus phasianellus (Alderfer 2005).

There is a size graduation between the six species of Phasianidae likely to occur in central

Alaska, in particular ptarmigans are significantly smaller than grouses. For all specimens recorded a measurement was taken and compared to comparative specimens. Four measurements

(numbers 1, 2, 3 and 5) were in the size range closest to Lagopus lagopus, and 1 (number 4) was closest to Lagopus leucurus (Figure 13). Number 6 is closest to Lagopus leucurus but is from an immature element. In addition, two adult tarsometatarsals provided standardized measurements:

39.64 mm (number 3 on Figure 13) and 30.96 mm (number 4). Those fall neatly in the published range for Lagopus lagopus and Lagopus leucurus/mutus, respectively (Kraft 1972, Bochenski

1985), for an element in which size is characteristic to species (Bochenski 1985).

1,3

1,2 Lagopus

1,1

) Swan Point 1 L. lagopus

lagopus L. leucurus 0,9 Tympanachus Dendragapus

0,8 Measurement(Standardized to 0,7 1 2 3 4 5 6 Specimens

Figure 13. Phasianidae specimens measurements from Swan Point CZ4b with references for

Lagopus lagopus, Lagopus leucurus, Tympanachus phasianellus and Dendragapus obscurus.

100

References

Alderfer, Jonathan

2005 National Geographic Complete Birds of North America. National Geographic

Society, Washington, D.C.

Allely, Steve, Tim Baker, Paul Comstock, Jim Hamm, Ron Hardcastle, Jay Massey, and John

Strunk

1992 The Traditional Bowyer’s Bible: Volume One. Bois d’Arc Press, Fort Worth,

Texas.

Andrews, Peter

1990 Owls, Caves and Fossils. University of Chicago Press.

Audouze, Françoise

2010 Domesticity and Spatial Organization at Verberie. In The Magdalenian

Household: Unraveling Domesticity, edited by Ezra Zubrow, Françoise Audouze,

and James G. Enloe, pp. 145–175. State University of New York Press, Albany.

Bacher, Alois

1967 Vergleichend Morphologische Untersuchungen an Einzelknochen des

Postcranialen Skeletts in Mitteleuropa Vorkommender Schwäme und Gänse.

Veterinary Medicine Dissertation, Ludwig-Maximilian-Universtät München.

Barton, C. M., Deborah I. Olszewski, and Nancy R. Coinman

1996 Beyond the Graver: Reconsidering Burin Function. Journal of Field Archaeology

23(1): 111–125.

Basilyan, A., M.A. Anisimov, P.A. Nikolskiy, and V.V. Pitulko

101

2011 Wooly Mammoth Mass Accumulation Next to the Paleolithic Yana RHS Site,

Arctic Siberia: Its Geology, Age, and Relation to Past Human Activity. Journal of

Archaeological Science 38(9): 2461-74.

Baumann, Malvina, and Serge Maury

2013 Ideas no Longer Written in Antler. Journal of Archaeological Science 40(1): 601–

614.

Behrensmeyer, Anna K.

1978 Taphonomic and Ecologic Information from Bone Weathering. Paleobiology

4(2): 150–162.

Bigelow, Nancy H, and W. Roger Powers

2001 Climate, Vegetation, and Archaeology 14,000-9,000 Cal Yr B.P. in Central

Alaska. Arctic Anthropology 38(2): 171–195.

Binford, Lewis R.

1978 Dimensional Analysis of Behavior and Site Structure: Learning from an Eskimo

Hunting Stand. American Antiquity 43(3): 330–361.

1983 In Pursuit of the Past. Thames and Hudson, New York.

Bochenski, Z.

1985 Osteological Differentiation in Willow Grouse. In Functional Morphology in

Vertebrates, Proceedings of the 1st International Symposium on Vertebrate

Morphology, Giessen, 1983, edited by H-R Duncker and G Fleischer, pp. 69–72.

Gustav Fischer Verlag, Stuttgart.

Bodu, Pierre, Michèle Julien, Borix Valentin, and Grégory Debout (editors).

102

2006 Un dernier hiver à Pincevent : les Magdaléniens du niveau IV0. Gallia

Préhistoire, Paris.

Bovy, Kristine M.

2002 Differential Avian Skeletal Part Distribution: Explaining the Abundance of

Wings. Journal of Archaeological Science 29(9): 965–978.

2012 Why So Many Wings? A Re-Examination of Avian Skeletal Part Representation

in the South-Central Northwest Coast, USA. Journal of Archaeological Science

39: 2049-2059.

Brazil, Mark

2009 Field Guide to the Birds of East Asia. Helm Field Guide, London.

Bronk Ramsey, Christopher

1994 Analysis of Chronological Information and Radiocarbon Calibration : The

Program OxCal. Archaeological Computing Newsletter 41: 11–16.

2009 Bayesian Analysis of Radiocarbon Dates. Radiocarbon 51(1): 337–360.

Buck, Caitlin E., and Edouard Bard

2007 A Calendar Chronology for Pleistocene Mammoth and Horse Extinction in North

America Based on Bayesian Radiocarbon Calibration. Quaternary Science

Reviews 26(17-18): 2031-5.

Buehler, Deborah M., Allan J. Baker, and Theunis Piersma

2006 Reconstructing Palaeoflyways of the Late Pleistocene and Early Holocene Red

Knot (Calidris canutus). Ardea 94(3): 485–498.

Burke, Ariane, and Jacques Cinq-Mars

103

1996 Dental Characteristics of Late Pleistocene Equus lambei from the Bluefish Caves,

Yukon Territory, and their Comparison with Eurasian Horses. Géographie

Physique et Quaternaire 50(1): 81.

Capaldo, Salvatore D., and Robert J. Blumenschine

1994 A Quantitative Diagnosis of Notches Made by Hammerstone Percussion and

Carnivore Gnawing on Bovid Long Bones. American Antiquity 59(4): 724–748.

Christensen, Marianne

1999 Technologie de l’ivoire au Paléolithique supérieur: caractérisation physico-

chimique du matériau et analyse fonctionnelle des outils de transformation.

Archaeopress, BAR-IS 751, Oxford.

Clark, Amy E.

2015 Spatial Structure and the Temporality of Assemblage Formation: a Comparative

Study of Seven Open Air Middle Paleolithic Sites in France. PhD Dissertation,

University of Arizona, Tucson.

Crass, Barbara A., Brant L Kedrowski, Jacob Baus, and Jeffery A. Behm

2011 Residue Analysis of Bone-Fueled Pleistocene Hearths. In From the Yenisei to the

Yukon: Interpreting Lithic Assemblage Variability in Late Pleistocene/Early

Holocene Beringia, edited by Ted Goebel and Ian Buvit, pp. 192–198. Texas

A&M University Press, College Station.

Currey, J. D.

1979 Mechanical Properties of Bone Tissues with Greatly Differing Functions. Journal

of Biomechanics 12(4): 313–319.

Derev’anko, Anatoliy P., Dimitri B. Shimkin, and W. Roger Powers

104

1998 The Paleolithic of Siberia. University of Illinois Press, Urbana.

Dikov, Nikolai N.

2003 Archaeological Sites of Kamchatka, Chukotka, and the Upper Kolyma. Translated

by Richard L. Bland. National Park Service, Anchorage, Alaska.

Dilley, T. E.

1998 Late Quaternary Loess Stratigraphy, Soils, and Environments of the Shaw Creek

Flats Paleoindian Sites, Tanana Valley, Alaska. PhD Dissertation, University of

Arizona, Tucson.

Dixon, E. James

2013 Arrows and Atl Atls: A Guide to the Archaeology of Beringia. National Park `

Service, Anchorage, Alaska.

Dove, Carla J., P. Gregory Hare, and Marcy Heacker

2005 Identification of Ancient Feather Fragments Found in Melting Alpine Ice Patches

in Southern Yukon. Arctic 58(1): 38–43.

Easton, Norman A., Glen R. Mackay, Patricia B. Young, Peter Schnurr, and David R. Yesner

2011 Chindadn in Canada? Emergent Evidence of the Pleistocene Transition in

Southeast Beringia as Revealed by the Little John Site, Yukon. In From the

Yenisei to the Yukon: Interpreting Lithic Assemblage Variability in Late

Pleistocene/Early Holocene Beringia, edited by Ted Goebel and Ian Buvit, pp.

289–307. Texas A&M University Press, College Station.

Elston, Robert G., and P. Jeffrey Brantingham

2002 Microlithic Technology in Northern Asia: A Risk-Minimizing Strategy of the

Late Paleolithic and Early Holocene. In Thinking Small: Global Perspectives on

105

Microlithization, edited by R. G. Elston and Steven L. Kuhn, pp. 103–116.

American Anthropological Association, Washington, D.C.

Finlayson, Clive, Kimberly Brown, Ruth Blasco, Jordi Rosell, Juan José Negro, Gary R.

Bortolotti, Geraldine Finlayson, Antonio Sánchez Marco, Francisco Giles Pacheco, Joaquín

Rodríguez Vidal, José S. Carrión, Darren A. Fa, and José M. Rodríguez Llanes

2012 Birds of a Feather: Neanderthal Exploitation of Raptors and Corvids. PLoS ONE

7(9): e45927.

Fitzgerald, G. R.

1991 Pleistocene Ducks of the Old Crow Basin, Yukon Territory, Canada. Canadian

Journal of Earth Sciences 28: 1561–1571.

Gelvin-Reymiller, Carol, Joshua D. Reuther, Ben A. Potter, and Peter M. Bowers

2006 Technical Aspects of a Worked Proboscidean Tusk from Inmachuk River, Seward

Peninsula, Alaska. Journal of Archaeological Science 33(8): 1088–1094.

Gilbert, B. Miles, L. D. Martin, and H. G. Savage

1981 Avian Osteology. Missouri Archaeological Society, Columbia.

Giria, Evgenij Y., and Vladimir V. Pitul’ko

1994 A High Arctic Mesolithic Industry on Zhokov Island: Inset Tools and Knapping

Technology. Arctic Anthropology 31(2): 31–44.

Goebel, Ted

1999 Pleistocene Human Colonization of Siberia and Peopling of the Americas: An

Ecological Approach. Evolutionary Anthropology 8(6): 208–227.

Goebel, Ted, and Ian Buvit (editors).

106

2011 From the Yenisei to the Yukon: Interpreting Lithic Assemblage Variability in Late

Pleistocene/Early Holocene Beringia. Texas A&M University Press, College

Station.

Goebel, Ted, Heather L. Smith, Lyndsay DiPietro, Michael R. Waters, Bryan Hockett, Kelly E.

Graf, Robert Gal, Sergei B. Slobodin, Robert J. Speakman, Steven G. Driese, and David Rhode

2013 Serpentine Hot Springs, Alaska: Results of Excavations and Implications for the

Age and Significance of Northern Fluted Points. Journal of Archaeological

Science 40(12): 4222–4233.

Gómez Coutouly, Yan Axel

2011 Industries lithiques à composante lamellaire par pression du nord Pacifique de la

fin du Pléistocène au début de l’Holocène: de la diffusion d’une technique en

Extrême-Orient au peuplement initial du Nouveau Monde. PhD Dissertation,

Université Paris-Ouest Nanterre, France.

2012 Pressure Microblade Industries in Pleistocene-Holocene Interior Alaska: Current

Data and Discussions. In The Emergence of Pressure Blade Making: From Origin

to Modern Experimentation, edited by Pierre M. Desrosiers, pp. 347-374.

Springer, New York.

Goutas, Nejma

2009 Réflexions sur une innovaion technique gravettienne importante : le double

rainurage longitudinal. Bulletin de la Société Préhistorique Française 106(3):

437–456.

Graf, Kelly E., Lyndsay M. DiPietro, Kathryn E. Krasinski, Angela K. Gore, Heather L. Smith,

Brendan J. Culleton, Douglas J. Kennett, and David Rhode

107

2015 Dry Creek Revisited: New Excavations, Radiocarbon Dates, and Site Formation

Inform on the Peopling of Eastern Beringia. American Antiquity 80(4): 671–694.

Guadelli, Jean Luc

1998 Détermination de l’âge des chevaux fossiles et établissement des classes d'âge.

Paléo 10: 87–93.

Guthrie, R. Dale

1983 Osseous Projectile Points: Biological Considerations Affecting Raw Material

Selection and Design among Paleolithic and Paleoindian People. In Animals and

Archaeology: 1. Hunters and their Prey, edited by J. Clutton-Brock and C.

Grigson, pp. 273–294. Archaeopress, BAR-IS 163, Oxford.

2003 Rapid Body Size Decline in Alaskan Pleistocene Horses Before Extinction.

Nature 426: 169-171.

2006 New Carbon Dates Link Climatic Change with Human Colonization and

Pleistocene Extinctions. Nature 441(7090): 207–9.

Hahn, J.

1995 Les ivoires en Allemagne : débitage, façonnage et utilisation au Paléolithique

supérieur. In Le travail et l’usage de l'ivoire au Paléolithique supérieur, Actes

Table Ronde de Ravello 1992, edited by J. Hahn, M. Menu, Yvette Taborin, P.

Walter, and F. Widemann, pp. 115–135. Centro Universitario Europeo per i Beni

Culturali, Ravello, Italy.

Hahn, J, M. Menu, Yvette Taborin, P. Walter, and F. Widemann (editors).

1995 Le travail et l’usage de l'ivoire au Paléolithique supérieur. Centro Universitario

Europeo per i Beni Culturali, Ravello, Italy.

108

Haile, James, Duane G. Froese, Ross D.E. MacPhee, Richard G. Roberts, Lee J. Arnold, Alberto

V. Reyes, Morten Rasmussen, Rasmus Nielsen, Barry W. Brook, Simon Robinson, Martina

Demuro, M. Thomas P. Gilbert, Kasper Munch, Jeremy J. Austin, Alan Cooper, Ian Barnes, Per

Möller, and Eske Willerslev

2009 Ancient DNA Reveals Late Survival of Mammoth and Horse in Interior Alaska.

Proceedings of the National Academy of Sciences 106(52): 22352-7.

Harington, C. R., and Jacques Cinq-Mars

2008 Bluefish Caves - Fauna and Context. Beringian Research Notes 19.

Haynes, Gary

1991 Mammoths, Mastodonts, and Elephants: Biology, Behavior, and the Fossil

Record. Cambridge University Press.

Heckel, Claire E., and Sibylle Wolf

2014 Ivory Debitage by Fracture in the Aurignacian: Experimental and Archaeological

Examples. Journal of Archaeological Science 42(1): 1–14.

Hemmings, Christopher A.

2004 The Organic Clovis: A Single Continent-Wide Cultural Adapt. PhD Dissertation,

University of Florida, Gainesville.

Hoffecker, John F.

2001 Late Pleistocene and Early Holocene Sites in the Nenana River Valley, Central

Alaska. Arctic Anthropology 38(2): 139–153.

2011 Assemblage Variability in Beringia: The Mesa Factor. In From the Yenisei to the

Yukon: Interpreting Lithic Assemblage Variability in Late Pleistocene/Early

109

Holocene Beringia, edited by Ted Goebel and Ian Buvit, pp. 165–178. Texas

A&M University Press, College Station.

Hoffecker, John F., Scott A. Elias, Dennis H. O’Rourke, G. Richard Scott, and Nancy H.

Bigelow

2016 Beringian and the Global Dispersal of Modern Humans. Evolutionary

Anthropology 25: 64-78.

Holmes, Charles E.

1996 Broken Mammoth. In American Beginnings: The Prehistory and Paleoecology of

Beringia, edited by Frederick H. West, pp. 312–319. University of Chicago Press.

2001 Tanana River Valley Archaeology circa 14,000 to 9,000 BP. Arctic Anthropology

38(2): 154–170.

2011 The Beringian and Transitional Periods in Alaska: Technology of the East

Beringian Tradition as Viewed from Swan Point. In From the Yenisei to the

Yukon: Interpreting Lithic Assemblage Variability in Late Pleistocene/Early

Holocene Beringia, edited by Ted Goebel and Ian Buvit, pp. 179–191. Texas

A&M University Press, College Station.

Holmes, Charles E, Daniel C Fisher, and Adam Rountrey

2012 Dental and Skeletal Evidence Point to the Hunting of Mammoths in Central

Alaska. Poster Presented at the 39th Annual Meeting of the Alaska

Anthropological Association, Seattle, WA, February 29 - March 3.

Hoskin, C. M., R. Dale Guthrie, and B. L. P. Hoffman

1970 Pleistocene, Holocene and Recent Bird Gastroliths from Interior Alaska. Arctic

23(1): 14–23.

110

Kedrowski, Brant L, Barbara A. Crass, Jeffery A Behm, J. C. Luetke, A. L. Nichols, A. M.

Moreck, and Charles E Holmes

2009 Gc/Ms Analysis of Fatty Acids From Ancient Hearth Residues at the Swan Point

Archaeological Site. Archaeometry 51(1): 110–122.

Khlopachev, G. A., and E. Y. Girya

2010 Sekrety Drevnikh Kostorezov Vostotchnoi Euvropy i Sibiri: Priemy Obrabotki

Bivnia Mamonta i Roga Severnogo Olenia v Kamennom Veke. Nauka, Saint-

Petersburg (Secrets of Ancient Bone Carvers in Eastern Europe and Siberia: The

Processing Methods of Mammoth Ivory and Antler in the Paleolithic,

Archaeological and Experimental Data).

Kraft, E.

1972 Verleichend Morphologische Untersuchungen an Einzelknoachen Nord- Und

Mitteleuropäischer Kleinerer Hühnervögel. Veterinary Medicine Dissertation,

Ludwig-Maximilian-Universtät München.

Lanoë, François B.

2015 Faunal Analysis of the Swan Point Site (XBD-156), AK. Report on file at the

Alaska State Historic Preservation Office.

Lanoë, François B., Joshua D. Reuther, Charles E. Holmes, Jennifer R. Kielhofer, and David

Plaskett

2016 The Keystone Dune Site: Identifying a Bølling-Allerød Short-Term Hunting Camp.

Paper presented at the 43rd Annual Meeting of the Alaska Anthropological

Association, Sitka, Alaska, March 2-6.

111

Leroi-Gourhan, André, and M Brézillon

1966 L’habitation magdalénienne n° 1 de Pincevent près Montereau (Seine-et-Marne).

Gallia Préhistoire IX(2): 263–385.

1972 Fouilles de Pincevent - Essai d’analyse ethnographique d'un habitat magdalénien

(la section 36). CNRS, Paris.

Llyod, A. H., M. E. Edwards, B. P. Finney, J. A. Lynch, V. Barber, and Nancy H. Bigelow

2006 Holocene Development of the Alaskan Boreal Forest. In Alaska’s Changing

Boreal Forest, edited by F. Stuart Chapin III, M. W. Oswood, K. Van Cleve, L.

A. Viereck, and D. L. Verbyla, pp. 62–80. Oxford University Press.

Lyman, R. Lee

1994 Vertebrate Taphonomy. Cambridge University Press.

2008 Quantitative Paleozoology. Cambridge University Press.

2015 North American Paleoindian Eyed Bone Needles: Morphometrics, Sewing, and

Site Structure. American Antiquity 80: 146-160.

Lyman, R. Lee, Michael J. O’Brien, and Virgil Hayes

1998 A Mechanical and Functional Study of Bone Rods from the Richey–Roberts

Clovis Cache, Washington, U.S.A. Journal of Archaeological Science 25: 887–

906.

MacDonald, G. M., D.W. Beilman, Y.V. Kuzmin, L.A. Orlova, K.V. Kremenetski, B. Shapiro,

R.K. Wayne, and B. Van Valkenburgh

2012 Pattern of Extinction of the Woolly Mammoth in Beringia. Nature

Communications 3:893.

Margaris, Amy V.

112

2014 Reconsidering Raw Material Selection: Skeletal Technologies and Design for

Durability in Subarctic Alaska. Journal of Archaeological Method and Theory

21(3): 669–695.

Maschenko, Evgeny N.

2002 Individual Development, Biology and Evolution of the Woolly Mammoth.

Cranium 19: 4–120.

Maschenko, Evgeny N., Svetlana S. Gablina, Alexey S. Tesakov, and Alexandra N. Simakova

2006 The Sevsk Woolly Mammoth (Mammuthus primigenius) Site in Russia:

Taphonomic, Biological and Behavioral Interpretations. Quaternary International

142-143: 147–165.

Metcalfe, Jessica Z., Fred J. Longstaffe, and Grant D. Zazula

2010 Nursing, Weaning, and Tooth Development in Woolly Mammoths from Old

Crow, Yukon, Canada: Implications for Pleistocene Extinctions.

Palaeogeography, Palaeoclimatology, Palaeoecology 298(3-4): 257–270.

Mochanov, Yuri A.

2009 The Earliest Stages of Settlement by People of Northeast Asia. Translated by

Richard L. Bland. National Park Service, Anchorage, Alaska.

Morrison, P., and W. Galster

1975 Patterns of Hibernation in the Arctic Ground Squirrel. Canadian Journal of

Zoology 53(9): 1345–1355.

Moss, Madonna L., and Jon M. Erlandson

2013 Waterfowl and Lunate Crescents in Western North America: The Archaeology of

the Pacific Flyway. Journal of World Prehistory 26(3): 173–211.

113

NDT Resource Center

2014 Visual Acuity of the Human Eye. Electronic document, https://www.nde-ed.org,

accessed January 28, 2016.

Olive, Monique, and Nicole Pigeot

2006 Réflexions sur le temps d’un séjour à Étiolles (Essonne). Bulletin de la Société

Préhistorique Française 103(4): 673–682.

Outram, Alan, and Peter Rowley-Conwy

1998 Meat and Marrow Utility Indices for Horse (Equus). Journal of Archaeological

Science 25: 839–849.

Pétillon, Jean-Marc, and Sylvain Ducasse

2012 From Flakes to Grooves: a Technical Shift in Antlerworking During the Last

Glacial Maximum in Southwest France. Journal of Human Evolution 62(4): 435–

65.

Pétillon, Jean-Marc, Olivier Bignon, Pierre Bodu, Pierre Cattelain, Grégory Debout, Mathieu

Langlais, Véronique Laroulandie, Hugues Plisson, and Boris Valentin

2011 Hard Core and Cutting Edge: Experimental Manufacture and Use of Magdalenian

Composite Projectile Points. Journal of Archaeological Science 38: 1266-1283.

Pigeot, Nicole

2010 Éléments d’une organisation sociale magdaénienne à Étiolles du savoir-faire au

statut social des personnes. In The Magdalenian Household: Unraveling

Domesticity, edited by Ezra Zubrow, Françoise Audouze, and James G. Enloe, pp.

198–212. State University of New York Press, Albany.

Pitul’ko, Vladimir V.

114

1993 An Early Holocene Site in the Siberian High Arctic. Arctic Anthropology 30(1):

13–21.

2011 The Berelekh Quest: A Review of Forty Years of Research in the Mammoth

Graveyard in Northeast Siberia. Geoarchaeology 26: 5-32.

2013 The Zhokhov Island Site and Ancient Habitation in the Arctic. Archaeology Press,

Burnaby, British Columbia.

Pitul’ko, Vladimir V., and Aleksey K Kasparov

1996 Ancient Arctic Hunters: Material Culture and Survival Strategy. Arctic

Anthropology 33(1): 1–36.

Pitul’ko, Vladimir V., Pavel. A. Nikolsky, E. Yu. Girya, A. E. Basilyan, V. E. Tumskoy, S. A.

Koulakov, S. N. Astakhov, Elena Y. Pavlova, and M. A. Anisimov

2004 The Yana RHS Site: Humans in the Arctic before the Last Glacial Maximum.

Science 303(5654): 52–6.

Pitulko, Vladimir V., Aleksandr E. Basilyan, and Elena Y. Pavlova

2014 The Berelekh Mammoth “Graveyard”: New Chronological and Stratigraphical

Data from the 2009 Field Season. Geoarchaeology 29(4): 277-299.

Pitul’ko, Vladimir V., Elena Y. Pavlova, and Pavel A. Nikolskiy

2015 Mammoth Ivory Technologies in the Upper Paleolithic: A Case Study Based on

the Materials from Yana RHS, Northern Yana-Indighirka Lowland, Arctic

Siberia. World Archaeology 47: 1–57.

Potter, Ben A.

2001 Recent Investigations at the Gerstle River Site, a Multicomponent Site in Central

Alaska. Current Research in the Pleistocene 18: 52-54.

115

2005 Site Structure and Organization in Central Alaska: Archaeological Investigations

at Gerstle River. PhD Dissertation, University of Alaska Fairbanks.

2007 Models of Faunal Processing and Economy in Early Holocene Interior Alaska.

Environmental Archaeology 12(1): 3–23.

2008 Radiocarbon Chronology of Central Alaska: Technological Continuity and

Economic Change. Radiocarbon 50(2): 181–204.

2011 Late Pleistocene and Early Holocene Assemblage Variability in Central Alaska.

In From the Yenisei to the Yukon: Interpreting Lithic Assemblage Variability in

Late Pleistocene/Early Holocene Beringia, edited by Ted Goebel and Ian Buvit,

pp. 215–233. Texas A&M University Press, College Station.

Potter, Ben A., Charles E. Holmes, and David R. Yesner

2013 Technology and Economy among the Earliest Prehistoric Foragers in Interior

Eastern Beringia. In Paleoamerican Odyssey, edited by Kelly E. Graf, C. V.

Ketron, and M. R .Waters, pp. 81–103. Texas A&M University Press, College

Station.

Potter, Ben A., Joel D. Irish, Joshua D. Reuther, Carol Gelvin-Reymiller, and Vance T. Holliday

2011 A Terminal Pleistocene Child Cremation and Residential Structure from Eastern

Beringia. Science 331(6020): 1058–62.

Potter, Ben A., Joel D. Irish, Joshua D. Reuther, and Holly J. McKinney

2014 New Insights into Eastern Beringian Mortuary Behavior: A Terminal Pleistocene

Double Infant Burial at Upward Sun River. Proceedings of the National Academy

of Sciences 111(48): 17060–17065.

116

Reimer, P. J., E. Bard, A. Bayliss, J. Beck, P. G. Blackwell, C. Bronk Ramsey, P. M. Grootes, T

P. Guilderson, H. Hadlifason, I. Hajdas, C. Hatte, T. J. Heaton, D. L. Hoffmann, A. G. Hogg, K.

A. Hughen, K. F. Kaiser, B. Kromer, S. W. Manning, M. Niu, R .W. Reimer, D. A. Richards, E.

M. Scott, J. R. Southon, R. A. Staff, C. S. M. Turney, and J. van der Plicht

2013 IntCal13 and Marine13 Radiocarbon Age Calibration Curves 0-50,000 Years cal

BP. Radiocarbon 55(4): 1869–1887.

Reuther, Joshua D.

2013 Late Glacial and Early Holocene Geoarchaeology and Terrestrial Paleoecology in

the Lowlands of the Middle Tanana Valley, Subarctic Alaska. PhD Dissertation,

University of Arizona, Tucson.

Roth, V.L., and J. Shoshani

1988 Dental Identification and Age Determination in Elephas maximus. Journal of

Zoology 214: 567–588.

Salomonsen, F .

1972 Presidential Address: Zoogeographical and Ecological Problems in Arctic Birds.

In Proceedings of the 15th International Ornithological Congress, edited by K. H.

Voous, pp. 25–77. E. J. Brill, Leiden.

Semenov, S. A.

1964 Prehistoric Technology. Cory, Adams & Mackay Ltd, London.

Semken, H.A., and S.C. Wallace

2002 Key to Arvicoline (“Microtine” Rodents) and Arvicoline-Like Lower First Molars

Recovered from Late Wisconsinan and Holocene Archaeological and

117

Palaeontological Sites in Eastern North America. Journal of Archaeological

Science 29: 23–31.

Serjeantson, Dale

2009 Birds. Cambridge University Press.

Shapiro, Beth, Alexei J. Drummond, Andrew Rambaut, Michael C. Wilson, Paul E. Matheus,

Andrei V. Sher, Oliver G. Pybus, M. Thomas P. Gilbert, Ian Barnes, Jonas Binladen, Eske

Willerslev, Anders J. Hansen, Gennady F. Baryshnikov, James A. Burns, Sergei Davydov,

Jonathan C. Driver, Duane G. Froese, C. Richard Harington, Grant Keddie, Pavel Kosintsev,

Michael L. Kunz, Larry D. Martin, Robert O. Stephenson, John Storer, Richard Tedford, Sergei

Zimov, and Alan Cooper

2004 Rise and Fall of the Beringian Steppe Bison. Science 306: 1561–1565.

Solow, Andrew R., David L. Roberts, and Karen M. Robbirt

2006 On the Pleistocene Extinctions of Alaskan Mammoths and Horses. Proceedings of

the National Academy of Sciences 103(19): 7351-3.

Steguweit, Leif

2015 Rotten Ivory as Raw Material Source in European Upper Paleolithic. Quaternary

International 361: 1–10.

Stiner, Mary C., Steven L. Kuhn, Stephen Weiner, and Ofer Bar-Yosef

1995 Differential Burning, Recrystallization, and Fragmentation of Archaeological

Bone. Journal of Archaeological Science: 223–237.

Stiner, Mary C., Steven L. Kuhn, Todd A. Surovell, Paul Goldberg, Liliane Meignen, Stephen

Weiner, and Ofer Bar-Yosef

118

2001 Bone Preservation in Hayonim Cave (Israel): a Macroscopic and Mineralogical

Study. Journal of Archaeological Science 28(6): 643–659.

Surovell, Todd A., and Nicole Waguespack

2007 Folsom Hearth-Centered Use of Space at Barger Gulch, Locality B. In Emerging

Frontiers in Colorado Paleoindian Archaeology, edited by R. H. Brunswig and B.

L. Pitblado, pp. 219–259. University of Colorado Press, Boulder.

Surovell, Todd A., Spencer R. Pelton, Richard Anderson-Sprecher, and Adam D. Myers

2016 Test of Martin’s Overkill Hypothesis Using Radiocarbon Dates on Extinct

Megafauna. Proceedings of the National Academy of Sciences 113: 886-891.

United States Geological Survey

2010 Interferometric Synthetic Aperture Radar (IfSAR) Digital Terrain Model (DTM)

for Alaska. Electronic document, http://ifsar.gina.alaska.edu/, accessed January

28, 2016.

VanderHoek, Richard, R. M. Tedor, and J. D. McMahn

2007 Cultural Materials Recovered from Ice Patches in the Denali Highway Region,

Central Alaska, 2003-2005. Alaska Journal of Anthropology 5(2): 185–200.

Vaquero, Manuel, and Ignasi Pastó

2001 The Definition of Spatial Units in Middle Paleolithic Sites: The Hearth-Related

Assemblages. Journal of Archaeological Science 28(11): 1209–1220.

Waguespack, Nicole M., and Todd A. Surovell

2014 A Simple Method for Identifying Households Using Lithic Assemblages: A Case

Study from a Folsom Campsite in Middle Park, Colorado. In Lithics in the

West:Using Lithic Analysis to Solve Archaeological Problems in Western North

119

America, edited by Douglas H MacDonald, William Jr. Andrefsky, and Pei-Lin

Yu, pp. 35–49. University of Montana Press, Missoula.

Waltari, Eric, and Joseph A. Cook

2005 Hares on Ice: Phylogeography and Historical Demographics of Lepus arcticus, L.

othus, and L. timidus (Mammalia: Lagomorpha). Molecular Ecology 14(10):

3005–3016.

Wenzel, Stefan

2011 The Magdalenian Dwelling of Orp East (Belgium) and its Spatial Organization. In

Site-Internal Spatial Organization of Hunter-Gatherer Societies: Case Studies

from the European Paleolithic and Mesolithic, edited by Sabine Gaudzinski-

Windheuser, Olaf Jöris, Martina Sensburg, Martin Street, and Elaine Turner, pp.

141–158. Römisch-Germanischen Zentralmuseums, Mainz, Germany.

West, Frederick H.

1981 Reflections on Beringian Prehistory and Early Man in America. Geoscience and

Man 22: 103–114.

White, Randall

1995 Ivory Personal Ornaments of Aurignacian Age: Technological, Social and

Symbolic Perspectives. In Le travail et l’usage de l'ivoire au Paléolithique

supérieur, edited by J. Hahn, M. Menu, Yvette Taborin, P. Walter, and F.

Widemann, pp. 29–62. Istituto Poligrafico E Zecca Dello Stato, Ravello, Italy.

Wilson, Michael C., Leonard V. Hills, and Beth Shapiro

120

2008 Late Pleistocene Northward-Dispersing Bison antiquus from the Bighill Creek

Formation, Gallelli Gravel Pit, Alberta, Canada, and the Fate of Bison

occidentalis. Canadian Journal of Earth Sciences 45(7): 827–859.

Woelfle, E.

1967 Vergleichend Morphologische Untersuchungen an Einzelknochen des

Postcranialen Skelettes in Mitteleuropa Vorkommender Enten, Halbgänse und

Säger. Veterinary Medicine Dissertation, Ludwig-Maximilian-Universtät

München.

Yesner, David R.

1994 Subsistence Diversity and Hunter-Gatherer Strategies in Late Pleistocene/ Early

Holocene Beringia: Evidence from the Broken Mammoth Site, Big Delta, Alaska.

Current Research in the Pleistocene 11: 154–156.

1996 Human Adaptation at the Pleistocene-Holocene (circa 13,000 to 8,000 BP)

Boundary in Eastern Beringia. In Humans at the End of the Ice Age, edited by

L.G. Straus, B.V. Eriksen, J.M. Erlandson, and D.R. Yesner, pp. 255–276.

Plenum Press, New York.

2001 Human Colonization of Eastern Beringia and the Question of Mammoth Hunting.

In Proceedings of the International Conference on Mammoth Site Studies, edited

by Dixie West, pp. 69-84. University of Kansas Press, Lawrence.

2007 Faunal Extinction, Hunter-Gatherer Foraging Strategies, and Subsistence

Diversity among Eastern Beringian Paleoindians. In Foragers of the Terminal

Pleistocene in North America, edited by R. B. Walker and B. N. Driskell, pp. 15–

31. University of Nebraska Press, Lincoln.

121

Yesner, David R., G. A. Pearson, and S. E. Stone

2000 Additional Organic Artifacts from the Broken Mammoth Site, Big Delta, Alaska.

Current Research in the Pleistocene 17: 87–90.

Yesner, David R.., K. J. Crossen, and Norman A Easton

2011 Gearchaeological and Zooarchaeological Correlates of Early Beringian

Assemblages: Insights from the Little John Site, Yukon. In From the Yenisei to

the Yukon: Interpreting Lithic Assemblage Variability in Late Pleistocene/Early

Holocene Beringia, edited by Ted Goebel and Ian Buvit, pp. 308–322. Texas

A&M University Press, College Station.

122

APPENDIX B:

TASK-SPECIFIC SITES AND PALEOINDIAN LANDSCAPE USE IN THE SHAW CREEK

FLATS OF EASTERN BERINGIA

François B. Lanoë1, Joshua D. Reuther2,3, Charles E. Holmes3

1 School of Anthropology, University of Arizona

2 University of Alaska Museum of the North

3 Department of Anthropology, University of Alaska Fairbanks

123

Abstract

The Shaw Creek Flats and nearby middle Tanana river, in central Alaska, constitute one of the areas in the Americas with the densest known distribution of Late Glacial (about 14,500-11,700 cal. B.P.) archaeological sites. Local high rates of sediment deposition and low post-depositional disturbance allow for the interpretation of the function of archaeological occupations within larger economic and mobility strategies. Residential sites used over the long term seem to be located near critical but immovable resources such as clear water and vegetation. The spatial association of artifacts and faunal remains at other sites in the Flats suggest that they were specialized, short-lived locations dedicated to a single or few activities. The site of Swan Point

CZ4b is interpreted as a workshop related to the production of composite tools, particularly on mammoth ivory; and the site of Keystone Dune is interpreted as a camp related to wapiti (Cervus elaphus) hunting. Task-specific sites were probably used as part of a predominantly logistical mobility and economy strategy, which maximized efficiency in harvesting and processing resources that were distributed heterogeneously on the landscape.

124

Introduction

How groups of hunter-gatherers behave within an economically-defined landscape is a central focus of archaeological research. Economic organization is tightly intertwined with diet, technology and other aspects of material culture, and also bears heavily on the larger ecological role of people within ecosystems. As a result, variation in how people moved around and exploited the resources in their territories has shaped many major prehistoric socio-economic transitions. During the Late Pleistocene, characterizing hunter-gatherer economic organization is central to understanding issues related to their dispersal into un-colonized landscapes, throughout subarctic Eurasia and the Americas, as well as their possible ecological role in the turnover of large mammalian fauna.

Ethnographically documented hunter-gatherers exhibit a wide spectrum of economic and mobility strategies (Binford 1980, 1982, Kelly 2013). Nevertheless, how people acquire resources generally depends on the scale at which these resources are distributed across the landscape. In environments where the resources targeted by hunter-gatherers are distributed homogeneously in space and time, people tend to favor a strategy incorporating frequent residential mobility (Kelly 1983, Grove 2009). They exploit resources in the immediate vicinity, generally a daily foraging radius, of a residential camp. The camp is moved to another location when resources in the patch are diminished to the point that other patches become more attractive (Bird and O’Connell 2006, Venkataraman et al. 2017). Typically, the frequency of residential moves correlates negatively with the resource density in the environment (Kelly

2013).

In environments where targeted resources are distributed heterogeneously over the landscape, no patch contains all of the resources needed by a group of hunter-gatherers, and there are often

125 gaps between concentrations of key resources. People then rely more on logistical mobility strategies. The location of residential camps depends on the search and transport costs associated with some key resources, and some group members conduct specialized trips to supply the residential camp with resources that are located farther away. Resources acquired during these trips can then be stored at the residential camp for use by the entire group. The location and duration of occupation of a residential camp in the case of a logistical strategy depends on a complex interplay between the costs associated with the acquisition, transport, and processing of each resource. As a rule, though, groups that emphasize logistical mobility move their residential camps less often and use them for longer periods of time (Kelly 2013).

Characterizing resource distribution on a landscape depends on the scale of consideration. Kelly and Todd (1988) famously proposed that Late Pleistocene hunter-gatherers dispersing for the first time into the Americas used a highly mobile residential strategy based on the exploitation of megafauna. Having little knowledge of the distribution of resources, or “resource geography,” of the landscapes they colonized, people would have targeted high-visibility and high-return large- bodied mammals. These resources are assumed to have been “regionally abundant, but locally unpredictable” (Kelly and Todd 1988 p. 235); in other words, megafauna would have been homogeneously distributed at a regional scale but heterogeneously at a local scale. In conditions of low population density and low inter-group competition, Late Pleistocene people would have moved over larger distances, creating sites with redundant assemblages, and possibly depleting megafauna populations as they went. In this scenario, hunter-gatherers would have dramatically disturbed otherwise well-established large mammal communities and contributed to the extinction of Rancholabrean fauna (Kelly and Todd 1988).

126

This model of Late Pleistocene mobility and economy is attractive considering some aspects of the archaeological record in the mid-latitudes of North America, particularly the evidence for proboscidean hunting (Surovell and Waguespack 2008) long-distance raw material procurement, and the homogeneity of material culture (Kelly and Todd 1988). Other researchers, though, argue that the archaeological record also offers evidence for extensive use of diverse resources and that mobility and economic systems varied by region (e.g. Chatters et al. 2012, Jones and

Beck 2012).

The Late Pleistocene archaeological record in most regions of North America does not necessarily offer the resolution to explore hunter-gatherer economic and mobility systems, and in most cases it remains difficult to test our interpretive models. In contrast, the record from the

Shaw Creek Flats, in subarctic eastern Beringia, offers a unique potential for research because of the number of documented occupations and their generally good preservation context. Here we present results from a multiscale spatial analysis, ranging from individual site to the collective landscape, to discuss evidence for the economic organization of Late Glacial hunter-gatherers in the Shaw Creek Flats and its implications for human dispersal in the subarctic and the Americas at the end of the Pleistocene.

Identifying the Function of Hunter-Gatherer Sites

Archaeologists generally recognize two broad types of hunter-gatherer sites (Binford 1980,

1982): residential and task-specific sites. Residential sites are used over longer periods of time by one to several households that include adult men and women, children, and the elderly.

Residential sites, or base camps, function as central places to which resources are brought to be

127 consumed or processed by the group, and where most technological and subsistence activities are performed (Jochim 2006, Kelly 2013).

The location of resources in the landscape conditions the suitable areas for establishing residential camps. Hunter-gatherers are interested by a diversity of resources; in higher latitudes these include food, water, fuel, and raw material, the first three of which are consumed on a daily basis. Water and fuel are comparatively costly to transport and are likely to have been of first importance for choosing the location of a residential camp (Table 5).

Criteria Residential Site Task-Specific Site

near water, fuel; Location near resource patches near rivers

Orientation south variable

Size larger smaller

lower; Material density higher and consistent or higher but clustered

Material diversity high low

Activity areas indistinct well-defined

Habitation more elaborate less elaborate structures

Table 5. Archaeological criteria used to differentiate residential and task-specific sites.

128

The longer duration of occupation and complexity of activities undertaken at a residential site should be reflected in the location of the site within a particular landform. It should also be reflected by the composition and spatial distribution of its archaeological assemblage (Table 5).

In the subarctic, for reasons of comfort, people are likely to have favored southern exposures and invested in more elaborate habitation structures (e.g. Binford 1978, Surovell and O’Brien 2016).

By-products from consumption and technological activities are likely to become mixed over the time of occupation of the site due to re-using the same locations for different activities, or through periodic cleaning and disposal (Binford 1983, Waguespack and Surovell 2014).

Residential sites centralize the acquisition and processing of resources, regardless of whether hunter-gatherers employ mobility strategies that are dominantly residential or dominantly logistical. Differences between residential sites that are part of a dominantly residential versus logistical mobility strategy essentially boil down to how far away resources were acquired.

Procedures implemented to reduce transport costs can inform how the residential site functioned in a larger economic and mobility system (Kelly 2013). Where resources are dispersed and people use logistical trips to acquire them, transport costs are most important. Costs can be reduced by conducting primary processing of resources such as large mammals or toolstone at the place of extraction. Other strategies to reduce transport costs include prolonging the use-lives of the artifacts that are made on costly raw material, and establishing residential camps near routes of travel. In the subarctic lowlands, rivers provide natural low-slope routes for movement, and establishing residential camps near rivers would also have made the relocation of a residential camp and communication with other groups easier.

In contrast to residential sites, task-specific sites are the result of acquiring and/or processing a specific resource by a subset of specialists within the group and may be used for a relatively

129 short period of time (Binford 1980). Despite their short duration of occupation, some task- specific sites will accumulate a large quantity of archaeological material because the activity generates a lot of waste (e.g. quarry, kill-site) or because a location is re-used repeatedly for the same purpose over a long period of time (Surovell 2009). Nevertheless the low level of behavioral complexity at such sites should be reflected in the composition and spatial distribution of their archaeological assemblage (Table 5).

In practice, most sites represent palimpsests of several occupations for which quantity and duration are difficult to estimate. Good locations may be used again and again for the same or different purposes, and a high diversity and density of archaeological materials at a site can result from a single, long-term residential occupation, or from many short-term logistical occupations

(Surovell 2009, Clark 2016). Task-specific sites typically exist at lower densities than residential sites, and may go unrecognized in many geological contexts in Beringia (Rasic 2011) or are blurred within palimpsests representing several distinct occupations (Surovell 2009, Hoffecker

2011).

The Shaw Creek Flats

The Shaw Creek Flats in central Alaska is an alluvial plain or terrace created by the aggradation of the Tanana River during the Last Glacial Maximum (Reger et al. 2008). The Flats and the surrounding Yukon-Tanana Uplands are drained by Shaw Creek and its tributaries, which eventually feed the Tanana River (Figure 14). The modern landscape is dominated by wetlands, muskeg and black spruce (Picea mariana) forests in the lowlands, and by closed canopy forests

130 of white spruce (Picea alba) and (Populus tremuloides) on the hills (Chapin III et al.

2006a).

During the Late Glacial (14,500-11,700 cal B.P.), when precipitation was 35-75% lower than present (Barber and Finney 2000), the Flats were drier and more unstable than today. Extant lakes were absent or at low levels until about 11,000 cal B.P. (Bigelow 1997, Wooller et al.

2012). High-energy aeolian deposition prevented extensive soil succession (Dilley 1998, Reuther et al. 2016). Vegetation was dominated by shrub birch (Betula nana) and graminoids (Poaceae and Cyperaceae) until about 12,000 cal B.P. (Bigelow and Powers 2001). Poplars (Populus spp.) appeared around 13,000 cal B.P. and began to dominate the record by 12,000 cal B.P. (Graf and

Bigelow 2011, Reuther et al. 2016), followed by white spruce by 10,000 cal B.P. (Anderson et al. 2004).

Large herbivore guilds were dominated by woolly mammoth (Mammuthus primigenius), Yukon horse (Equus lambei) and steppe bison (Bison priscus) until about 15,000-14,000 cal B.P.

(Guthrie 1990, 2006). Wapiti (Cervus elaphus) and moose (Alces alces), along with bison (Bison cf. priscus) become dominant in the record after the extinction of mammoth and horse (Guthrie

2006, Lanoë et al. 2017). Migratory waterfowl (Anatidae) are first recorded by 14,000 cal B.P.

(Yesner 2001, Lanoë and Holmes 2016) and anadromous fish (Salmonidae) by 11,800 cal B.P.

(Halffman et al. 2015, Choy et al. 2016), following changes in the location and extant of major regional biogeographic barriers in the Late Glacial.

131

Figure 14. Region of the Shaw Creek Flats and sites mentioned in the text: (1) Swan Point, (2)

Bachner, (3) Broken Mammoth, (4) Cook, (5) Keystone Dune, (6) Mead, (7) Holzman. The red lines represent least-cost paths linking sites by the Tanana River to sites deep in the Flats. The paths follow Shaw and Keystone Creeks.

The Shaw Creek Flats include one of the densest known concentrations of buried, in situ

Pleistocene archaeological sites in the Americas with 12 to 16 documented occupations in an area roughly 20 km by 8 km (Table 6, Figure 14). Diagnostic archaeological material recovered at Late Glacial sites in the Flats are dominated by bifacial points and microblades produced by the Campus method and these assemblages have been assigned to the Chindadn culture of eastern Beringia (Holmes 2011). One exception is the occupation of Swan Point Cultural Zone

4b (hereafter CZ 4b) (Holmes et al. 1996). Arguably the oldest documented site in northern

132

North America, it is dominated by burins and microblades produced by the Yubetsu method

(Gómez Coutouly 2012) and has been assigned to the Dyuktai culture recorded in the whole of

Beringia (Holmes 2011).

Occupation Date (cal B.P.) References

Mead CZ3b 12,120-11,850 Potter et al. 2013

Broken Mammoth CZ3 12,386-11,769 Yesner 2001

Swan Point CZ3 (a?, b) 12,830-11,390 Holmes 2014

Cook, Lower Component 12,527-12,046 this paper

Holzman (a/b?) 13,250-11,620 Wygal et al. 2016

Bachner, Component 1 13,100-12,700 Wooller et al. 2012

Mead CZ4 13,110-12,790 Potter et al. 2013

Swan Point CZ4a 13,420-13,080 Holmes 2014

Keystone Dune 13,400-13,220 Reuther et al. 2016

Mead CZ5 13,440-13,200 Potter et al. 2013

Broken Mammoth CZ4 (a/b/c?) 13,647-13,152 Yesner 2001

Swan Point CZ4b 14,150-13,870 Lanoë and Holmes 2016

Table 6. Known Late Glacial occupations in the Shaw Creek Flats. Previously unpublished date from the Cook site (AA-106445): 10,400 ± 60 14C B.P., on bone collagen, δ13C = -20.0 ‰; calibrated here with OxCal 4.2 and IntCal 13 (Bronk Ramsey 1994, 2009, Reimer et al. 2013).

133

Landforms in the Flats are dominated by aeolian sediments deposited by persistent winds with abundant sediment supply during the Late Glacial (Dilley 1998, Reuther et al. 2016). These sediments are carbonaceaous and slightly alkaline favoring the preservation of biogenic materials in Late Glacial archaeological occupations. These are generally associated with thin buried dark brown sandy and silty loam soil (Ab) horizons that provide stratigraphic landmarks allowing archaeologists to identify past surfaces and the extent of cryo-, bio- or tectonoturbation.

Sediment deposition rates are high during the Late Glacial compared to later periods. They range from about 1 cm per century at Mead and Swan Point sites, up to 5 to 7 cm per century at the

Bachner, Tannenbaum, and Keystone localities, and generally buried stable surfaces and helped protect human occupations from post-depositional disturbance (Table 7). In several sites, or portions of sites, archaeologically sterile horizons separate non-contemporaneous occupations. In this case, high chronological resolution and preserved spatial organization allow for an accurate interpretation of the function of the occupations. Other occupations, such as Broken Mammoth

CZ4 a/b/c or Swan Point CZ3 a/b, are more difficult to interpret because of a lower stratigraphic resolution and the formation of palimpsests (Table 6). Sites such as Bachner and Cook require more extensive excavations to provide reliable interpretations of human behavior. Functional interpretations in this paper focus on the best preserved and documented Late Glacial occupations of the Flats.

134

Section Depth Date Range Deposition Rate Site (cm below Sediment pH Paleosols (cal B.P.) (cm/century) surface)

Bachner Site 139-115 12,370-11,860 Loess 7.8-8.1 5 4.7

Camp Section 158-127 12,584-11,618 Loess 8.3-8.6 3 to 5 3.2

Tannenbaum 168-122 12,698-11,803 Loess - 4 5.1 Section

288 Bluff Section 142-102 12,864-12,045 Loess 8.1-8.6 5 to 11 4.9

Broken Mammoth 150-127 13,295-12,090 Loess 8.3-8.5 4 to 6 1.9 Site

Cook Site 195-90 12,287-9,349 Loess 6.8-8.8 6 to 8 3.6

Keystone Dune 470-315 13,300-11,160 Sand - > 18 7.2

Mead Site 137-125 13,320-11,985 Loess 8.5-8.6 4 to 6 0.9

Swan Point Site 78-54 14,010-11,680 Loess 6.8-7.7 1 to 3 1.0

Table 7. Rates of sediment deposition during the Late Glacial at archaeological and geological localities of the Shaw Creek Flats (data computed from Dilley 1998, Potter et al. 2013, Reuther

2013, Reuther et al. 2016; rates calculated following Stein et al. [2003]). The ends of the date ranges are averages of intervals calibrated with OxCal 4.2 and IntCal 13 (Bronk Ramsey 1994,

2009, Reimer et al. 2013). For the Cook site, the date at 195 cm BS is referenced in Table 6; the date at 90 cm BS is as follows (UGAMS--18145): 8,320 ± 30 14C B.P., on bone collagen, δ13C =

-20.6 ‰, δ15N = 2.4 ‰.

135

Residential Sites

Broken Mammoth CZ3 and Mead CZ4 are well-documented occupations located 1.2 km apart on a bluff overlooking the confluence of Shaw Creek and the Tanana River. They meet the expectations for residential sites outlined in Table 5, confirming the interpretations made by the original excavators (Potter et al. 2013).

The Broken Mammoth and Mead sites are oriented to the south along the bluff edge as would be expected for sites used over long durations. The sites are also located in the immediate vicinity of the Tanana River (Table 8). The recovery of obsidian at Swan Point (CZ3) and Broken

Mammoth (CZ4) from Batza Téna and Wiki Peak distant 420 km NW and 360 km SE from the sites, respectively (Reuther et al. 2011), indicates that Late Glacial people of the Shaw Creek

Flats moved or communicated over long distances. The Tanana River probably channeled long- distance mobility and/or communication in the past much as it does today; in addition, proximity to the Tanana River probably facilitated residential camp relocation throughout its drainage.

Distance Distance to Nearest Hearth Fauna Lithics Excavated Occupation Aspect to Tanana Shaw Creek Density Density Density Area (m2) (km) Creek (km) (km) (/m2) (/m2) (/m2)

Broken

Mammoth SSE < 0.1 0.4 0.4 408 .015 13.1 9.8

CZ3

Mead CZ4 SE 1.0 0.5 0.5 128 .023 ~1.6 11.7

Swan Point 360° 6.0 3.3 1.6 80 .025 18.9 63.9 CZ4b

Keystone SSW 7.4 5.6 0.5 22 .091 5.7 8.0 Dune

136

Table 8. Characteristics of several Late Glacial occupations in the Shaw Creek Flats. Densities were computed from Krasinski and Yesner (2008) for Broken Mammoth and from original data for Swan Point and Keystone Dune; estimated from figures and text in Potter et al. (2013) for

Mead.

The Broken Mammoth and Mead sites are located in the immediate vicinity of Shaw Creek

(Table 8). Shaw Creek and its tributaries provided the least costly path to the Flats from the

Tanana River and would have facilitated mobility in that area (Figure 14). Shaw Creek is much clearer than the turbid, sediment-loaded Tanana River: today RGB values from Landsat photographs near the confluence range between 0.68-0.74 for the Tanana River and 0.19-0.25 for

Shaw Creek. Shaw Creek was probably favored for drinking water as these two water courses must have been the only perennial water sources during the drier Late Glacial. Wood fuel was probably present along the Tanana River and Shaw Creek during most of the Late Glacial in the form of gallery forests of Populus spp. and Salix spp. (Anderson and Brubaker 1994, Bigelow

1997). In contrast, at Swan Point CZ4b further from the Tanana River people used bones in addition to charcoal to sustain their campfires (Kedrowski et al. 2009, Holmes 2011) even though the thermic properties of animal tissues are not as good as those of woody vegetation

(Théry-Parisot and Costamagno 2005). Lastly, local lithic raw materials were available in the bed of the Tanana River and on the bluff in the form of secondary cobbles and quartz nodules. These were used heavily at both Mead CZ4 and Broken Mammoth CZ3 (Potter et al.

2013).

137

Densities of archaeological materials in the Broken Mammoth CZ3 and Mead CZ4 occupations are comparable or even lower than other occupations in the Flats (Table 8). Nevertheless, materials were recorded at consistent densities over very large surface areas (Krasinski and

Yesner 2008, Potter et al. 2013). In contrast, materials at Swan Point CZ4b and Keystone Dune were recorded in restricted areas with high densities of archaeological materials (Lanoë and

Holmes 2016). The Keystone Dune occupation, though partly eroded along the dune edge, does not seem to have extended over more than ~15 m2 (see below). In addition, Swan Point CZ4b is atypical as a short-term occupation. Evidence from the analysis of faunal and lithic materials and their spatial layout suggest that people at this location focused on producing microblade-based composite tools, and created disproportionately large amounts of waste from this activity (Lanoë and Holmes 2016).

Archaeological materials recovered at Broken Mammoth CZ3 and Mead CZ4 are diverse (Table

9

Table 9). Tool forms in particular suggest a diversity of activities, such as hide tanning, tool and textile making, in addition to primary reduction of local chert cobbles and quartz nodules (Potter et al. 2013). In contrast, shaped tools are mostly absent from the Keystone Dune occupation, and the artifact diversity observed at Swan Point CZ4b is consistent with the production of composite tools only (Lanoë and Holmes 2016).

The faunal assemblages in Broken Mammoth CZ3 and Mead CZ4 are dominated by bison and wapiti but a wider array of animal species were also recovered, particularly in Broken Mammoth

CZ3 (Table 9). The faunal composition suggests that the sites supported both short and long- distance forays that targeted different animal resources. It is not known whether bison or wapiti

138 primary or secondary butchering occurred on site; nevertheless, the faunal assemblages are strikingly different than at a specialized field butchering site such as the nearby Early Holocene occupation of Gerstle River C3 (Potter 2007), or the monospecific assemblage of Keystone Dune

(Table 9).

Occupation Lithic Tools Bone Tools & Art Animal Remains

bifaces ivory points Bison, Cervus, Rangifer, Ovis, Canis, Alopex, Broken Mammoth CZ3 points decorated bone rod Lepus, Marmota, Anatidae, Phasianidae scrapers bone needle

burins antler billet Mead CZ4 scrapers Bison, Cervus, Anatidae worked ivory? bifaces

ivory rod

microblades ivory blanks Equus, Mammuthus, Rangifer, Swan Point CZ4b burins antler blanks Lepus, Anatidae, Phasianidae

bird bone tools?

Keystone Dune scraper - Cervus

Table 9. Diversity of archaeological materials at several Late Glacial occupations in the Shaw

Creek Flats (computed from Yesner 1994, 1996, 2001, Holmes 1996, 2011, Yesner et al. 2000,

Potter et al. 2013, Goebel and Potter 2016, Lanoë and Holmes 2016, Reuther et al. 2016).

Some degree of spatial structure can be recognized in the Late Glacial occupations in the Shaw

Creek Flats. Associations of hearths with drop zones related to lithic reduction or fauna 139 consumption are typical of ethnographic and archaeological contexts (e.g. Leroi-Gourhan and

Brézillon 1972, Binford 1983) and are also found in Mead CZ4 (Potter et al. 2013), Broken

Mammoth CZ3 (Krasinski and Yesner 2008), Swan Point CZ4b (Lanoë and Holmes 2016), and at Keystone Dune (see below). Constraints on the distribution of artifacts that are characteristic of constructed habitation structures (Stapert 2003, Leesch and Bullinger 2012) are also found at all four occupations.

The occupations differ at other levels of spatial organization. At Mead CZ4, archaeological materials show a bimodal distribution around hearth F2011-6. High density areas are located within 1 meter and more than 3 meters from the hearth, while the area in between is empty of materials (Potter et al. 2013 fig. 5.2). In contrast, an area up to 3 m around the south hearth at

Swan Point CZ4b is littered with waste (Lanoë and Holmes 2016). The spatial distribution of materials at Mead CZ4 is typical of a habitation structure in which the living area was regularly cleaned and relocated to dumps near the door. This pattern suggests that people stayed at Mead long enough to be bothered by the accumulation of waste indoors.

Hunting Bison and Wapiti

Bison and wapiti dominate the Late Glacial archeofaunal record in the Shaw Creek Flats (Figure

15), even though other large herbivores, such as caribou (Rangifer tarandus), Dall sheep (Ovis dalli), and moose, were present in the landscape (Yesner 2001). The exception to the dominance of bison and wapiti is Swan Point CZ4b, but where meat consumption was minimal in comparison to the use of animal tissues for raw material (Lanoë and Holmes 2016). Bison and wapiti probably formed a staple of the economy until the Early Holocene throughout subarctic

140 eastern Beringia (Potter 2008, 2011), and emphasis on these two ungulate species must have played a large part in determining the economic and mobility strategies of Late Glacial people.

100%

80%

60% Bison 40% Wapiti Other large herbivores 20%

0% Mead CZ4 Broken Broken Swan Point Swan Point Keystone Mammoth Mammoth CZ4b CZ3 Dune CZ4 CZ3

Figure 15. Distribution of large herbivore species (proportion of NISP) at Late Glacial occupations in the Shaw Creek Flats (computed from Yesner 2001, Potter et al. 2013, Lanoë and

Holmes 2016, Reuther et al. 2016, Lanoë et al. 2017).

Bison and wapiti prefer grasslands or early successional habitat dominated by graminoids, forbs, and a limited amount of leafy vegetation (Guthrie 1983, 1990, Stephenson et al. 2001, McCabe

2002). In eastern Beringia, prior to 14,000 cal B.P., they occupied a niche intermediate to specialized grazers such as horse or mammoth, and specialized browsers such as moose, but they would have used grasslands more after the extinction of mammoth and horse (Lanoë et al. 2017).

141

Grassland patches in the Late Glacial were more extensive than today. Locally favorable conditions included xeric areas with higher drainage, insulation, and aeolian activity (Guthrie

1983, 2001), as well as areas with cyclic disturbance from high-energy alluvial and aeolian sedimentation which would have maintained soils and vegetation in an early successional stage

(Stephenson et al. 2001, Reuther et al. 2016). Early successional habitat on gravel bars and alluvial terraces is used today by introduced plains bison (Bison bison) near the confluence of the

Tanana and Delta rivers ~25 km south of the Shaw Creek Flats (Glassburn 2015). Because of seasonal changes in forage quality and availability bison and wapiti probably moved between these different types of habitats throughout the year, generally favoring uplands in the summer and lowlands in the winter (Guthrie 1983, Rasic 2011, Glassburn 2015).

Figure 16 models the distribution of favorable grassland habitat during the Late Glacial in the

Shaw Creek Flats, assuming that steep and south-oriented slopes of the Uplands provided dry conditions and that the depositing sand sheet provided enough disturbance to maintain soil and vegetation in an early successional stage. Following this model, favorable bison and wapiti habitat would have centered on the southern slopes of the Yukon-Tanana Uplands on the northern edge of the Flats, where the Keystone Dune site is located (Figure 16). These patches were probably favored in the summer, while suitable winter habitat may have been located along the Tanana River. Nevertheless, most of the lowlands are likely to have concentrated gallery forests, shrub birch, and nascent wetlands instead, creating a habitat more favorable for the specialized browser and wetland-adapted moose.

142

Figure 16. Topographic and geologic variation in the Shaw Creek Flats. Areas in yellow have higher slope (> 10°) and are oriented to the south (135-225°) reflecting higher drainage and insolation. Hatched areas represent sand sheets deposited during the Late Glacial (data from

Weber et al. 1978). Legend as in Figure 14.

The Keystone Dune Site: A Wapiti Hunting Camp

The Keystone Dune site (XBD-363) is situated in the Rosa-Keystone dunefield, at the northeastern horn of a parabolic dune, and overlooks Keystone Creek and its valley. The dune is composed of horizontal beds of coarse to fine sands, in which a cut up to 15 m deep was created by a local mining road (Reuther et al. 2016). The single archaeological occupation at the site did not yield diagnostic artifacts but may be assigned to Phase II of the Beringian tradition (Holmes

143

2011) based on its age, which is similar to other sites in the Shaw Creek Flats (Table 6). All in- situ archaeological materials were found on the same surface, associated with a charred Ab horizon that show little to no evidence of post-depositional disturbance (Reuther et al. 2016).

Rapid sediment deposition (Table 7) favored the spatial preservation of the archaeological occupation, but a portion of it was removed by the erosion of the dune along the roadcut.

Figure 17. Profile and map of the Keystone Dune site. Locations of artifacts in the 2014 test are computed from photographs.

144

Archaeological materials are distributed in relation to two hearths (Figure 17). Lithic artifacts recovered in-situ consist almost exclusively of small reduction by-products (95.0% smaller than

10 mm; Table 10). They are organized in two clusters centered 0.41 and 1.07 m east from the eastern hearth and represent a drop zone around the hearth characteristic of hunter-gatherer camps (e.g. Leroi-Gourhan and Brézillon 1972, Binford 1983). Lithic specimens were also recovered at the bottom of the dune during survey (Table 10), corresponding to artifacts that were most likely located south of the hearths and later eroded away with this portion of the dune.

While suffering from an obvious size taphonomic bias, they are nevertheless larger (χ2 = 80.9; df

= 4; p < 0.01) and may relate to artifacts removed from the main activity area by cleaning, tossing, or recycling (see Binford 1983, Stiger 2006, Waguespack and Surovell 2014).

145

Surface finds Materials Unprovenienced below excavation in situ surface finds block

Size range (mm)

0-5 48 2

5-10 104 11 70

10-15 16 15 65

15-20 5 11 25

> 20 3 12 43

Typology

Cores 0 1

Unretouched flake 175 50 198

[biface thinning flake] [2] [4] [6]

Unifacial tools 1 1

Bifacial tools 4

Raw material

Basalt 157 32 53

Rhyolite 1 7 41

Chert 18 12 109

Table 10. Lithic materials recovered at the Keystone Dune Site. Specimens recorded near the hearth are significantly smaller than surface finds below the excavation block (χ2 = 80.9; df = 4; p < 0.01) or in other areas of the dune (χ2 = 131.1; df = 4; p < 0.01).

146

The faunal materials located immediately east of the eastern hearth (Figure 17) consist of small unidentifiable bone fragments with generally poor preservation causing them to crumble upon contact (Table 11, Figure 18). Similar to the situation for the lithic débitage, this concentration of small bone fragments can be interpreted as a drop zone. People broke bones, likely during extraction of medullary marrow from long bones. Larger bone specimens are scattered in other areas of the site and generally are better preserved; those items were removed deliberately from the activity area. Overall, the faunal specimens are consistent with the consumption of tissues from two limbs of a single wapiti. Most of the portions of the wapiti carcass did not reach the site or were taken away subsequently.

147

Zone in/east Other zones

of hearth

Average size (mm) 13.6 ± 17.3 65.2 ± 52.1

[number measured] [14] [5]

Unidentified NSP not recovered 6 mostly cancellous bone 18 5 mostly compact bone 42 3

Ungulate NISP tooth fragment 39 2 long bone 1 rib 1

Cervid NISP femur 3 radius 1

Wapiti NISP proximal phalanx 2 mesial phalanx 2 distal phalanx 1

Total 106 20

Table 11. Faunal materials recovered at the Keystone Dune Site. Levels of taxonomic identification are significantly higher (χ2 = 51.7; df = 3; p < 0.01) and specimens significantly larger (t = 2.2; df = 4.1; p = 0.09) in areas away from the eastern hearth.

148

Figure 18. Taphonomic characteristics of the Keystone Dune faunal assemblage (proportion of

NISP). Left: breakage type. Right: burning stages following Stiner et al. (1995), with 0: unburned; 1-3: partly to wholly carbonized; and 4-6: partly to wholly calcined. Materials in or near the eastern hearth are in black, materials in other zones are in white. Differences between the subsamples are significant for burning (χ2 = 30.2; df = 4; p < 0.01) and breaking (χ2 = 45.6; df

= 3; p < 0.01).

Clusters of materials are well defined: the mean distance between specimens in each cluster ranges between 3.6 cm (western lithic drop zone), 5.8 cm (fauna drop zone), and 6.1 cm (eastern lithic drop zone). The rest of the area excavated is otherwise void of materials (Figure 17, Table

10, Table 11). This sharp spatial definition suggests limited horizontal disturbance from trampling, cleaning or reuse of space for multiple activities. Strict delimitation of the dense material zone to the east of the eastern hearth also suggests a spatial constraint on the circulation of artifacts, an effect typically created by a physical barrier (Stapert 2003, Surovell and

Waguespack 2007). Likewise, charcoal and ochre specks are limited to this same area and indicate scattering and/or trampling within a confined space. Assuming this area constraining 149 artifact spatial distribution relates to a constructed structure, the eastern hearth would have been located at its western entrance. Drop zones related to maintenance and consumption activities would have been located between the hearth and a wall. A waste zone resulting from rough cleaning would have been located by the opening, as is typical of the hunter-gatherer ethnographic and archaeological records (Binford 1983, Bodu et al. 2006, Waguespack and

Surovell 2014, Surovell and O’Brien 2016).

The Keystone Dune site meets the expectations of a short-term task-specific site (Table 5) by its location on the landscape as well as by its composition and spatial layout. The absence of complex tools or reduction processes suggests that technological activities were limited to repair or maintenance of tools brought into the site. The character of the faunal assemblage and the lack of explicit spatial maintenance or elaborate structures suggest that only a few people stayed at the site for a limited amount of time. The site was probably used specifically as a hunting camp; that is, a base used by hunters for rest and preparation before and after hunting expeditions in the nearby wapiti hunting grounds.

Swan Point CZ4b: A Short-Term Specialized Occupation

Swan Point CZ4b is located on top of a knoll overlooking the Shaw Creek Flats and the southern slopes of the Yukon-Tanana Uplands (Dilley 1998, Holmes 2011, Lanoë and Holmes 2016). The location of the site as well as the material composition and spatial layout of the CZ4b occupation indicate that it was used specifically for the acquisition and transformation of organic raw materials.

150

Combined information from lithic and organic materials indicates a focus on the production or maintenance of composite tools. Microblades and organic blanks were produced and probably assembled at the site in well-delimited manufacturing areas, including several hearth - drop zones associations (Lanoë and Holmes 2016; Figure 19). The artifacts recovered, although numerous (Table 8), are not very diverse (Table 9) and all relate to the different steps of the complex chaîne opératoire to produce composite microblade organic tools.

151

Figure 19. Interpretation of the use of space at Swan Point CZ4b (reproduced from Lanoë and

Holmes [2016]).

152

Despite a high density of archaeological materials (Table 8), Swan Point CZ4b was likely used by only a few people and for a short time. The spatial distribution of the materials at Swan Point

CZ4b suggests the existence of a constructed structure centered on one of the hearths. Different types of materials form delimited clusters suggesting an absence of cleaning or extensive trampling during the occupation. These characteristics of the site do not indicate use over a long duration (Supplemental Text 1). Moreover, there is no evidence of primary butchery and large- scale consumption of ungulates. Faunal remains, however diverse (Table 9), relate closely to the technological activities (Lanoë and Holmes 2016). The other faunal specimens, which can reasonably be interpreted as resulting from a food consumption activity, represent less than

50,000 Kcal worth, even while assuming poor skeletal preservation. This limited amount of food would have, at best, sustained a small party for less than two weeks (Supplemental Text 2).

Swan Point was an ideal location for a workshop given this proximity to raw material sources. It is located in the middle of wetlands that provided favorable habitat for waterfowl. The site was also probably located in the vicinity of natural accumulations of mammoth ivory. People brought numerous pieces of fossil or subfossil ivory to the site, some of them so large that it would have been unpractical to transport them over long distances (Lanoë and Holmes 2016), suggesting that they took advantage of a nearby ivory “mine,” as has been argued for other Upper Paleolithic sites rich in mammoth ivory specimens (Pitulko et al. 2015, Steguweit 2015).

Discussion

The Late Glacial archaeological record of the Shaw Creek Flats is exceptionally dense and well- preserved in comparison to most regions of northeastern Eurasia and the Americas.

153

Archaeological occupations are well-preserved both in terms of their archaeological assemblages and their spatial layouts, thanks to the high rates of fine sediment deposition during the Late

Glacial. Hunter-gatherer task-specific sites characterized by low-densities typically become blurred into palimpsests because the location was re-used for different purposes and/or by different groups (Hoffecker 2011, Rasic 2011). In contrast, the sites in the Shaw Creek Flats have high chronological resolution that provides synchronic “snapshots” in time.

Late Glacial people in the Shaw Creek Flats exploited resources in a complex logistical fashion.

Sites used for residential, longer-term purposes were strategically located to access everyday resources while also accommodating residential mobility. Residential sites centralized the acquisition of resources within the larger landscape. Some of these resources, including food items and raw materials, were acquired and processed at specialized sites before being transported to residential sites. This heterogeneity in site types explains much of the lithic variability observed for the period (Potter 2011, Rasic 2011). The short-term sites in particular attest to the importance of logistical mobility in the economic system in the Shaw Creek Flats.

Archaeological research so far has focused on landforms overlooking the Tanana River where residential sites are likely to be found. We expect that future surveys deeper in the Flats will increase the number of Late Glacial task-specific occupations used as part of logistical economic strategies.

Potential resources for Late Glacial people in the Shaw Creek Flats included those from swamps and forested areas that were dominated by moose, grassland dominated by bison and wapiti, and uplands dominated by caribou and sheep. The resources that people favored from grassland habitats were heterogeneously distributed, while those from forested areas were probably more homogeneously available in time and space. Logistical mobility must have allowed people to

154 reduce the search and transport costs associated with the acquisition of grassland resources and maintain their economic attractiveness. In contrast, had people focused mainly on wetland resources, they could have followed a purely residential mobility strategy, moving and exploiting moose populations along a geographic gradient within the lowlands.

Paleoindian/Paleoarctic people in the subarctic eastern Beringia do not seem to have behaved as free-wandering groups exhausting megafauna along long-distance travels as once modeled by

Kelly and Todd (1988). Instead, they had an excellent knowledge of the resource geography in a patchy landscape, probably occupied well-defined territories (Rasic 2011), and perceived the local landscape and its components at a fine temporal and spatial scale. In that aspect they were similar to other Paleoindian groups documented in the Great Basin and Pacific Northwest (Jones et al. 2003, Chatters et al. 2012, Jones and Beck 2012). In other aspects related to the importance of residential versus logistical mobility, people of the Shaw Creek Flats seem to add to the diversity of economic strategies observed among Late Pleistocene and Early Holocene

Paleoindians. Major changes in economic and mobility strategies in eastern Beringia are likely to have occurred later in the Early to Middle Holocene when people had to reorient their economy towards the exploitation of boreal forest and tundra habitats (Esdale 2008, Potter 2008, 2016) due to the establishment of the modern biomes of central Alaska (Kaufman et al. 2016).

Supplemental Text 1. Evidence for a Constructed Structure at Swan Point CZ4b.

In ethnographic and archaeological cases hearth-associated shelters can be identified through indirect evidence of walls: stones (e.g. Pigeot 2010), arcs of debris (e.g. Leroi-Gourhan and

Brézillon 1966), or dramatic change in artifact density (e.g. Gelhausen et al. 2009); but in most

155 cases the existence of a shelter, its emplacement and morphology rely on the subjectivity of the researcher (Leesch and Bullinger 2012).

Here we use a ring and sector analysis to provide a quantitative interpretation of spatial patterns associated with Swan Point CZ4b hearths (modified from Stapert 1989, 2003, Surovell and

Waguespack 2007). Artifact density is computed in .2 m-wide rings around the centroid of each hearth (Lanoë and Holmes 2016). The modality of density distribution observed in ring-and- sector analysis provides indications on the presence or absence of an enclosed structure. One mode of strong densities near the center of the hearth that gradually decrease with distance is typical of open-air hearths; whereas additional modes of high-density further from the center of the hearth is typical of hearths associated with a closed space, where dense zones are produced by “wall effects” stopping the centrifugal dispersion of artifacts and/or zones of secondary discard associated with cleaning of habitat.

The north hearth has two zones of high density immediately west and east of the hearth centroid; from there, densities decrease as a function of distance (Figure 20). This pattern is most consistent with an open-air hearth. In contrast, the south hearth does not follow the unimodal pattern expected for an open-air hearth. High-density rings immediately west of the hearth centroid suggest a drop zone related to a dense activity area (Figure 20). Density peaks again between 1.6 and 2.2 m east of the hearth centroid and is consistent with a “wall effect.” Shape and type of shelter (e.g. tent, windbreak, etc.) are unknown, but delimiting walls following peak density rings encompasses the whole diffuse burned stain of the hearth (Lanoë and Holmes

2016). Shelter position is supported by the spatial distribution of faunal remains; this is most noticeable for the mammoth molar fragments that are limited to this area and were probably dispersed, by trampling, from two original elements (Lanoë and Holmes 2016). A large portion

156 of this delimited space lacks burned residues and has proportionally low densities of remains; it may have been reserved for an indoor activity that produced little to no debris, for instance resting.

The area located between the two hearths contains remains of mixed nature: bulky (horse skull, cores) and nutritious elements (mammoth ribs), as well as medium (geese bones) and small fragments (ivory, microblades). A mixed assemblage is most consistent with secondary refuse resulting from cleaning of various activity areas. Accordingly, and given the similarity in types of materials represented, these items likely originated from the closed space around the south hearth. Its proximity with the south hearth suggests a discard zone similar to a “door dump”

(Leroi-Gourhan and Brézillon 1972, Binford 1983, Julien 2006, Waguespack and Surovell

2014).

157

Figure 20. Ring and sector analysis, Swan Point CZ4b south and north hearths.

158

Supplemental Text 2. Estimation of the Length of Stay at Swan Point CZ4b.

Inferred Portions Total Mode of Tissue Mode of Total Edible Animal Anatomical N Weight Weight Calculation of References Value Calculation of References Energy Portions Part (g) (g) Weight (Kcal/g) Value (Kcal)

brain: bison; Brain, tongue: beef; (Meagher 1986); composite (Carmody and bison head 1 tongue, 1,735 1,735 6.2 10,756 marrow: 75% of grasslandbeef.com meat/fat of beef Wrangham 2009) marrow horse

200% of adult (Madrigal and Capaldo marrow of white- (McCullough and moose hindfeet 1 Marrow 149 149 white tailed 9.4 1,400 1999) tailed deer Ullrey 1983) (wet weight)

caribou adult white-tailed Madrigal and Capaldo marrow of white- McCullough and hindfeet 1 Marrow 75 75 9.4 700 (juvenile) deer, wet weight 1999 tailed deer Ullrey 1983

brain and marrow: Brain, (Outram and Rowley- horse; composite Carmody and horse head 1 tongue, 2,196 2,196 Conwy 1998, Cozzi et 6.2 13,614 tongue: 125% of meat/fat of beef Wrangham 2009 marrow al. 2014) beef mammoth ivory 3 None - - - - 0.0 - - 0 mammoth Brain, (Hanks 1972); composite Carmody and (5- years- head 1 tongue, 2,391 2,391 see (1) below 6.2 14,824 faculty.washington.edu meat/fat of beef Wrangham 2009 old) marrow mammoth animals.nationalgeogra composite Carmody and (1-month- ribcage 1 Meat 9,100 9,100 see (2) below 6.2 56,420

phic.com meat/fat of beef Wrangham 2009 old)

159 snowshoe composite Carmody and complete 1 Meat 900 900 snowshoe hare (Ashley 2002) 6.2 5,580 hare meat/fat of beef Wrangham 2009

composite Carmody and arctic hare complete 1 Meat 2,300 2,300 arctic hare Ashley 2002 6.2 14,260 meat/fat of beef Wrangham 2009

Meat, Carmody and teal complete 1 400 400 80% of widgeon Ashley 2002 7.3 mallard meat 2,930 skin Wrangham 2009

Meat, Carmody and duck complete 1 800 800 mallard Ashley 2002 7.3 mallard meat 5,860 skin Wrangham 2009

Meat, Carmody and goose complete 4 220 880 Canada goose Ashley 2002 6.7 goose meat 5,927 skin Wrangham 2009

swan (species

tundra Meat, unspecified; Carmody and complete 2 5,800 11,600 Ashley 2002 6.7 goose meat 78,126 swan skin probably tundra Wrangham 2009

swan)

willow Meat, Carmody and complete 1 500 500 willow ptarmigan Ashley 2002 5.8 chicken meat 2,908 ptarmigan skin Wrangham 2009

small Meat, Carmody and complete 1 400 400 rock ptarmigan Ashley 2002 5.8 chicken meat 2,326 ptarmigan skin Wrangham 2009

Total 33,426 215,630

(1) Few portions weight values were found in the literature for proboscideans. The brain weighs 30.3% of the total body weight in an adult elephant (Chudler 2015); for a 5-years-old woolly mammoth, this proportion was arbitrarily increased to 36.4 (120% of the latter) because mammalian juvenile brains are proportionally heavier. For the tongue and marrow, we followed the ratios tongue/brain weight and marrow/brain weight published for bison (Meagher 1986).

(2) We followed the ratio thorax/body weight published for horse (Outram and Rowley-Conwy 1998), using the weight of a newborn elephant calf (Hanks 1972).

(3) Carmody and Wrangham 2009: reported values average raw and cooked values

160

Table 12. Model of consumed body parts at Swan Point CZ4b and associated energetic returns. The version displayed is the highest possible energetic return obtained from body parts recorded at CZ4b, with a total portion weight of 33.4 Kg and an associated energetic return of 215,000 Kcal. However, major energy contributors (e.g. swan, infant mammoth ribcage) include portions which were probably not (or only partially) consumed (Lanoë and Holmes 2016). With the assumptions that brains, hindfeet, and birds were acquired for technological purposes (tanning, sinew, and feathers, respectively), and that only 20% of the mammoth ribcage was consumed at the site, estimation of weight of consumed body parts falls to 5.0 Kg and associated energetic returns to 31,000 Kcal.

161

40 35 30 25 20 15 10 Lengthof (days) stay 5 0 1 2 3 4 5 6 7 8 9 10 Group size

Figure 21. Estimated minimum and maximum length of stay at CZ4b depending on group size. Daily energy requirements, 2,100

Kcal, are the standard for an adult (World Health Organization 1985) but may have been slightly higher among mobile hunter- gatherers (Pontzer et al. 2012). Maximum lengths of stay for one person and two people are not displayed for the sake of readability; they amount to 103 and 51 days, respectively.

162

References

Anderson, Patricia M., and Linda B Brubaker

1994 Vegetation History of Northcentral Alaska: A Mapped Summary of Late-

Quaternary Data. Quaternary Science Reviews 13: 71–92.

Anderson, Patricia M., Mary E Edwards, and Linda B Brubaker

2004 Results and Paleoclimate Implications of 35 Years of Paleoecological Research in

Alaska. Developments in Quaternary Science 1: 427–440.

Ashley, B.

2002 Edible Weights of Wildlife Species used for Country Food in the Northwest

Territories and Nunavut. Government of the Northwest Territories, Yellowknife,

NWT.

Barber, Valerie A, and Bruce P Finney

2000 Late Quaternary Paleoclimatic Reconstructions for Interior Alaska Based on

Paleolake-Level Data and Hydrologic Models. Journal of Paleolimnology 24: 29–

41.

Bigelow, Nancy H.

1997 Late Quaternary Vegetation and Lake Level Changes in Central Alaska. PhD

Dissertation, University of Alaska Fairbanks.

Bigelow, Nancy H., and W. Roger Powers

163

2001 Climate, Vegetation, and Archaeology 14,000-9,000 Cal Yr B.P. in Central

Alaska. Arctic Anthropology 38(2): 171–195.

Binford, Lewis R.

1978 Nunamiut Ethnoarchaeology. Academic Press, New York.

1980 Willow Smoke and Dogs’ Tails: Hunter-Gatherer Settlement Systems and

Archaeological Site Formation. American Antiquity 45: 4–20.

1982 The Archaeology of Place. Journal of Anthropological Archaeology 1(1): 5–31.

1983 In Pursuit of the Past. Thames and Hudson, New York.

Bird, Douglas W., and James F. O’Connell

2006 Behavioral Ecology and Archaeology. Journal of Archaeological Research 14(2):

143–188.

Bodu, Pierre, Michèle Julien, Borix Valentin, and Grégory Debout (editors).

2006 Un dernier hiver à Pincevent : les Magdaléniens du niveau IV0. Gallia

Préhistoire.

Bronk Ramsey, Christopher

1994 Analysis of Chronological Information and Radiocarbon Calibration : The

Program OxCal. Archaeological Computing Newsletter 41: 11–16.

2009 Bayesian Analysis of Radiocarbon Dates. Radiocarbon 51(1): 337–360.

Carmody, R. N., and R. W. Wrangham.

164

2009 The Energetic Significance of Cooking. Journal of Human Evolution 57: 379–91.

Chapin III, F.S., T. Hollingsworth, D. F. Murray, L. A. Viereck, and M. D. Walker

2006 Floristic Diversity and Vegetation Distribution in the Alaskan Boreal Forest. In

Alaska’s Changing Boreal Forest, edited by F. S. Chapin III, M. W. Oswood, K.

Van Cleve, L. A. Viereck, and D. L. Verbyla, pp. 81–99. Oxford University Press,

New York.

Chatters, James C., S. Hackenberger, Anna Marie Prentiss, and J.-L. Thomas

2012 The Paleoindian indian to Archaic Transition in the Pacific Northwest: In Situ

Development or Ethnic Replacement? In From the Pleistocene to the Holocene,

edited by C. Britt Bousman, Bradley Vierra, and George T. Jones, pp. 100–173.

Texas A&M University Press, College Station.

Choy, Kyungcheol, Ben A. Potter, Holly J. McKinney, Joshua D. Reuther, Shiway W. Wang, and Matthew J. Wooller

2016 Chemical profiling of ancient hearths reveals recurrent salmon use in Ice Age

Beringia. Proceedings of the National Academy of Sciences 113(35): 9757–9762.

Chudler, E. H.

2015 Brain Facts and Figures. https://faculty.washington.edu/chudler/facts.html.

Clark, Amy Elizabeth

2016 Time and Space in the Middle Paleolithic: Spatial Structure and Occupation

Dynamics of Seven Open-Air Sites. Evolutionary Anthropology 25(3): 153–163.

165

Cozzi, B., M. Povinelli, C. Ballarin, and A. Granato

2014 The Brain of the Horse: Weight and Cephalization Quotients. Brain, Behavior

and Evolution 83: 9–16.

Dilley, T. E.

1998 Late Quaternary Loess Stratigraphy, Soils, and Environments of the Shaw Creek

Flats Paleoindian Sites, Tanana Valley, Alaska. PhD Dissertation, University of

Arizona.

Esdale, Julie A.

2008 A Current Synthesis of the Northern Archaic. Arctic Anthropology 45(2): 3–38.

Gelhausen, F., J. F. Kegler, and S. Wenzel

2009 Find Concentrations and Dwelling Structures. The Interpretation of Final

Paleolithic Find Scatters. Pages 450–457in S. McCartan, R. Schulting, G. Warren,

and P. Woodman, editors. Mesolithic Horizons : Papers Presented at the Seventh

International Conference on the Mesolithic in Europe, Belfast 2005. Oxbow

Books, Oxford.

Glassburn, Crystal L.

2015 A Reconstruction of Steppe Bison Mobility in the Yukon-Tanana Uplands and

Implications for Prehistoric Human Behavior. Master’s Thesis, University of

Alaska Fairbanks.

Goebel, Ted, and Ben A. Potter

166

2016 First Traces: Late Pleistocene Human Settlement of the Arctic. In The Oxford

Handbook of the Prehistoric Arctic, edited by T. M. Friesen and Owen K. Mason,

pp. 223–252. Oxford University Press.

Gómez Coutouly, Yan Axel

2012 Pressure Microblade Industries in Pleistocene-Holocene Interior Alaska: Current

Data and Discussions. In The Emergence of Pressure Blade Making: From Origin

to Modern Experimentation, edited by Pierre M. Desrosiers. Springer US, Boston,

MA.

Graf, Kelly E., and Nancy H. Bigelow

2011 Human Response to Climate During the Younger Dryas Chronozone in Central

Alaska. Quaternary International 242(2): 434–451.

Grove, Matt

2009 Hunter-Gatherer Movement Patterns: Causes and Constraints. Journal of

Anthropological Archaeology 28: 222–233.

Guthrie, R. Dale

1983 Paleoecology of the Site and Its Implications for Early Hunters. In Dry Creek:

Archaeology and Paleoecology of a Late Pleistocene Alaskan Hunting Camp,

edited by W. Roger Powers, R. Dale Guthrie, and John F. Hoffecker, pp. 209–

287. Report Filed with the National Park Service.

1990 Frozen Fauna of the Mammoth Steppe. University of Chicago Press.

167

2001 Origin and Causes of the Mammoth Steppe: A Story of Cloud Cover, Woolly

Mammal Tooth Pits, Buckles, and Inside-Out Beringia. Quaternary Science

Reviews 20: 549–574.

2006 New Carbon Dates Link Climatic Change with Human Colonization and

Pleistocene Extinctions. Nature 441: 207–9.

Halffman, Carrin M., Ben A. Potter, Holly J. McKinney, Bruce P. Finney, Antonia T. Rodrigues,

Dongya Y. Yang, and Brian M. Kemp

2015 Early Human Use of Anadromous Salmon in North America at 11,500 y ago.

Proceedings of the National Academy of Sciences 112(40): 12344–12348.

Hanks, J.

1972 Growth of the African Elephant (Loxodonta africana). East African Wildlife

Journal 10: 251–272.

Hoffecker, John F.

2011 Assemblage variability in Beringia: The Mesa factor. In From the Yenisei to the

Yukon: Interpreting Lithic Assemblage Variability in Late Pleistocene/Early

Holocene Beringia, edited by Ted Goebel and Ian Buvit, pp. 165–178. Texas

A&M University Press, College Station.

Holmes, Charles E.

168

1996 Broken Mammoth. In American Beginnings: The Prehistory and Palaeoecology

of Beringia, edited by Frederick H West, pp. 312–319. University of Chicago

Press.

2011 The Beringian and Transitional Periods in Alaska: Technology of the East

Beringian Tradition as Viewed from Swan Point. In From the Yenisei to the

Yukon: Interpreting Lithic Assemblage Variability in Late Pleistocene/Early

Holocene Beringia, edited by Ted Goebel and Ian Buvit, pp. 179–191. Texas

A&M University Press, College Station, TX.

2014 New Evidence Pertaining to the Early Archaeological Sequence at Swan Point ,

Central Alaska. Poster Presented at the 41st Annual Meeting of the Alaska

Anthropological Association, Fairbanks, AK, March 5-8.

Holmes, Charles E., Richard VanderHoek, and Thomas E. Dilley

1996 Swan Point. In American Beginnings: The Prehistory and Palaeoecology of

Beringia, edited by F. H. West, pp. 319–322. University of Chicago Press.

Jochim, Michael A.

2006 Regional Perspectives on Early Mesolithic Land Use in Southwestern Germany.

Journal of Anthropological Archaeology 25(2): 204–212.

Jones, George T, and Charlotte Beck

2012 The Emergence of the Desert Archaic in the Great Basin. In From the Pleistocene

to the Holocene, edited by C. Britt Bousman, Bradley Vierra, and George T.

Jones, pp. 267–317. Texas A&M University Press, College Station.

169

Jones, George T., Charlotte Beck, Eric E. Jones, and Richard E. Hughes

2003 Lithic Source Use and Paleoarchaic Foraging Territories in the Great Basin.

American Antiquity 68(1): 5–38.

Julien, Michèle

2006 À la recherche des campements d’hiver dans le Magdalénien du Bassin parisien.

Bulletin de la Société Préhistorique Française 103: 695–709.

Kaufman, Darrell S., Yarrow L. Axford, Andrew C G Henderson, Nicholas P. McKay, W. Wyatt

Oswald, Casey Saenger, R. Scott Anderson, Hannah L. Bailey, Benjamin Clegg, Konrad

Gajewski, Feng Sheng Hu, Miriam C. Jones, Charly Massa, Cody C. Routson, Al Werner,

Matthew J. Wooller, and Zicheng Yu

2016 Holocene Climate Changes in Eastern Beringia (NW North America) - A

Systematic Review of Multi-Proxy Evidence. Quaternary Science Reviews 147:

312–339.

Kedrowski, Brant L, Barbara A. Crass, Jeffery A Behm, J. C. Luetke, A. L. Nichols, A. M.

Moreck, and Charles E. Holmes

2009 Gc/Ms Analysis of Fatty Acids from Ancient Hearth Residues at the Swan Point

Archaeological Site. Archaeometry 51(1): 110–122.

Kelly, Robert L.

1983 Hunter-Gatherer Mobility Strategies. Journal of Anthropological Research 39(3):

277–306.

170

2013 The Lifeways of Hunter-Gatherers: The Foraging Spectrum. Cambridge

University Press.

Kelly, Robert L., and Lawrence C. Todd

1988 Coming into the Country: Early Paleoindian Hunting and Mobility. American

Antiquity 53(2): 231–244.

Krasinski, and David R. Yesner

2008 Late Pleistocene-Early Holocene Site Structure in Beringia. Alaska Journal of

Anthropology 6: 27–41.

Lanoë, François B., and Charles E. Holmes

2016 Animals as Raw Materials in Beringia: Insights from the Site of Swan Point

CZ4b, Alaska. American Antiquity 81(4): 682–696.

Lanoë, François B., Joshua D. Reuther, Charles E. Holmes, and Gregory Hodgins

2017 The Lion’s Share: Human Paleoecological Integration in Eastern Beringia.

Submitted to Quaternary Science Reviews.

Leesch, Denise, and Jérôme Bullinger

2012 Identifying Dwellings in Upper Paleolithic Open-Air Sites – The Magdalenian

Site at Monruz and its Contribution to Analysing Palimpsests. In A Mind Set on

Flint. Studies in Honour of Dick Stapert, edited by M. Niekus, R. Barton, Martin

Street, and T. Terberger, pp. 165–181. Groningen University.

171

Leroi-Gourhan, André, and Michel Brézillon.

1966 L’habitation magdalénienne n° 1 de Pincevent près Montereau (Seine-et-Marne).

Gallia Préhistoire IX: 263–385.

1972 Fouilles de Pincevent - Essai d’analyse ethnographique d’un habitat magdalénien

(la section 36). CNRS, VII supplement to Gallia Préhistoire, Paris.

Madrigal, T. C., and S. D. Capaldo

1999 White-Tailed Deer Marrow Yields and Late Archaic Hunter-Gatherers. Journal of

Archaeological Science 26: 241–249.

McCabe, Richard E. (editor)

2002 North American : Ecology and Management. Smithsonian Institution Press,

Washington, D.C.

McCullough, D. R., and D. E. Ullrey

1983 Mineral and Gross Energy Composition of White-Tailed Deer. Journal of Wildlife

Management 47: 430–441.

Meagher, M.

1986 Bison bison. Mammalian Species 266: 1–8.

Outram, A., and P. Rowley-Conwy

1998 Meat and Marrow Utility Indices for Horse (Equus). Journal of Archaeological

Science 25:839–849.

172

Pigeot, Nicole

2010 Éléments d’une organisation sociale magdaénienne à Étiolles du savoir-faire au

statut social des personnes. Pages 198–212in E. Zubrow, F. Audouze, and J. G.

Enloe, editors.The Magdalenian Household: Unraveling Domesticity. State

University of New York Press, Albany.

Pitulko, Vladimir V., Elena Yu Pavlova, and Pavel A. Nikolskiy

2015 Mammoth Ivory Technologies in the Upper Paleolithic: A Case Study Based on

the Materials from Yana RHS, Northern Yana-Indighirka Lowland, Arctic

Siberia. World Archaeology 47: 333–389.

Pontzer, H., D. A. Raichlen, B. M. Wood, A. Z. P. Mabulla, S. B. Racette, and F. W. Marlowe

2012 Hunter-Gatherer Energetics and Human Obesity. PLoS ONE 7: e40503.

Potter, Ben A.

2007 Models of Faunal Processing and Economy in Early Holocene Interior Alaska.

Environmental Archaeology 12(1): 3–23.

2008 Radiocarbon Chronology of Central Alaska: Technological Continuity and

Economic Change. Radiocarbon 50(2): 181–204.

2011 Late Pleistocene and Early Holocene Assemblage Variability in Central Alaska.

In From the Yenisei to the Yukon: Interpreting Lithic Assemblage Variability in

Late Pleistocene/Early Holocene Beringia, edited by T Goebel and I Buvit, pp.

215–233. Center for the Study of the First Americans, College Station, TX.

173

2016 Holocene Prehistory of the Northwestern Subarctic. In The Oxford Handbook of

the Prehistoric Arctic, edited by T. Max Friesen and Owen K. Mason, pp. 537–

562. Oxford University Press.

Potter, Ben A., Charles E. Holmes, and David R. Yesner

2013 Technology and Economy among the Earliest Prehistoric Foragers in Interior

Eastern Beringia. In Paleoamerican Odyssey, edited by Kelly E Graf, C V Ketron,

and M R Waters, pp. 81–103. Center for the Study of the First Americans,

College Station, TX.

Rasic, Jeffrey T.

2011 Functional Variability in the Late Pleistocene Archaeological Record of Eastern

Beringia: A Model of Late Pleistocene Land Use and Technology from Northwest

Alaska. In From the Yenisei to the Yukon: Interpreting Lithic Assemblage

Variability in Late Pleistocene/Early Holocene Beringia, edited by Ted Goebel

and Ian Buvit, pp. 128–164. Texas A&M University Press, College Station.

Reger, R. D., D. S. P. Stevens, and D. N. Solie

2008 Surficial Geology of the Alaska Highway Corridor, Delta Junction to Dot Lake,

Alaska. Alaska Division of Geologic and Geophysical Surveys.

Reimer, P. J., E .Bard, A. Bayliss, J. Beck, P. G. Blackwell, Christopher Bronk Ramsey, P. M.

Grootes, T. P. Guilderson, H. Hadlifason, I. Hajdas, C. Hatte, T. J. Heaton, D. L. Hoffmann, A.

G. Hogg, K. A. Hughen, K. F. Kaiser, B. Kromer, S. W. Manning, M. Niu, R. W. Reimer, D. A.

Richards, E. M. Scott, J. R. Southon, R. A. Staff, C. S. M. Turney, and J. van der Plicht

174

2013 IntCal13 and Marine13 Radiocarbon Age Calibration Curves 0-50,000 Years cal

BP. Radiocarbon 55(4): 1869–1887.

Reuther, Joshua D.

2013 Late Glacial and Early Holocene Geoarchaeology and Terrestrial Paleoecology

in the Lowlands of the Middle Tanana Valley, Subarctic Alaska. PhD Dissertation,

University of Arizona.

Reuther, Joshua D., Ben A. Potter, Charles E. Holmes, J. K. Feathers, François Lanoë, and

Jennifer Kielhofer

2016 The Rosa-Keystone Dunes Field: The Geoarchaeology and Paleoecology of a

Late Quaternary Stabilized Dune Field in Eastern Beringia. The Holocene 26(12):

1939–1953.

Reuther, Joshua D., N. S. Slobodina, Jeffrey T. Rasic, John P. Cook, and Robert J. Speakman

2011 Gaining Momentum: Late Pleistocene and Early Holocene Archaeological

Obsidian Source Studies in Interior and Northeastern Beringia. In From the

Yenisei to the Yukon: Interpreting Lithic Assemblage Variability in Late

Pleistocene/Early Holocene Beringia, edited by Ted Goebel and Ian Buvit, pp.

270–286. Texas A&M University Press, College Station.

Stapert, D.

1989 The Ring and Sector Method: Intrasite Spatial Analysis of Stone Age Sites, with

Special Reference to Pincevent. Palaeohistoria 31: 1–57.

175

2003 Towards Dynamic Models of Stone Age Settlements. In Perceived Landscapes

and Built Environments: The Cultural Geography of Late Paleolithic Eurasia,

edited by Sergey A. Vasil’ev, O Soffer, and Janusz K Kozlowski, pp. 5–16. BAR

International Series 1122, Archaeopress, Oxford.

Steguweit, Leif

2015 Rotten Ivory as Raw Material Source in European Upper Paleolithic. Quaternary

International 361: 313–318.

Stein, Julie K., Jennie N. Deo, and Laura S. Phillips

2003 Big Sites—Short Time: Accumulation Rates in Archaeological Sites. Journal of

Archaeological Science 30(3): 297–316.

Stephenson, Robert O., S. C. Gerlach, R. Dale Guthrie, C. R. Harington, R. O. Mills, and G.

Hare

2001 Wood Bison in Late Holocene Alaska and Adjacent Canada: Paleontological,

Archaeological and Historical Records. In People and Wildlife in Northern North

America, edited by S C Gerlach and M S Murray, pp. 125–133. Archaeopress:

BAR International Series, Oxford.

Stiger, Mark

2006 A Folsom Structure in the Colorado Mountains. American Antiquity 71(2): 321–

351.

Stiner, Mary C., Steven L. Kuhn, Stephen Weiner, and Ofer Bar-Yosef

176

1995 Differential Burning, Recrystallization, and Fragmentation of Archaeological

Bone. Journal of Archaeological Science 22(2): 223–237.

Surovell, Todd A.

2009 Toward a Behavioral Ecology of Lithic Technology: Cases from Paleoindian

Archaeology. University of Arizona Press, Tucson.

Surovell, Todd A., and Matthew O’Brien

2016 Mobility at the Scale of Meters. Evolutionary Anthropology 25(3): 142–152.

Surovell, Todd A., and Nicole M. Waguespack

2007 Folsom Hearth-Centered Use of Space at Barger Gulch, Locality B. In Emerging

Frontiers in Colorado Paleoindian Archaeology, edited by R. H. Brunswig and B.

L. Pitblado, pp. 219–259. University of Colorado Press, Boulder.

2008 How Many Elephant Kills Are 14? Clovis Mammoth and Mastodon Kills in

Context. Quaternary International 191(1): 82–97.

Théry-Parisot, Isabelle, and Sandrine Costamagno

2005 Propriétés combustibles des ossements : données expérimentales et réflexions

archéologiques sur leur emploi dans les sites paléolithiques. Gallia Préhistoire

47: 235–254.

Venkataraman, Vivek V., Thomas Kraft, Nathaniel J. Dominy, and Kirk M. Endicott

177

2017 Hunter-gatherer residential mobility and the marginal value of rainforest patches.

Proceedings of the National Academy of Sciences 114(12): 3097–3102.

Waguespack, Nicole M., and Todd A. Surovell

2014 A Simple Method for Identifying Households Using Lithic Assemblages: A Case

Study from a Folsom Campsite in Middle Park, Colorado. In Lithics in the West:

Using Lithic Analysis to Solve Archaeological Problems in Western North

America, edited by Douglas H. MacDonald, William Jr. Andrefsky, and Pei-Lin

Yu, pp. 35–49. University of Montana Press, Missoula.

Weber, F. R., H. L. Foster, T. E. C. Keith, and C. Dusel-Bacon

1978 Preliminary Geologic Map of the Big Delta Quadrangle, Alaska.

Wooller, Matthew J., Joshua Kurek, Benjamin V. Gaglioti, Les C. Cwynar, Nancy H. Bigelow,

Joshua D. Reuther, Carol Gelvin-Reymiller, and John P. Smol

2012 An ~11,200 Year Paleolimnological Perspective for Emerging Archaeological

Findings at Quartz Lake, Alaska. Journal of Paleolimnology 48(1): 83–99.

World Health Organization

1985 Energy and Protein Requirement. World Health Organization, Geneva.

Wygal, Brian T., Kathryn E. Krasinski, Charles E. Holmes, and Barbara A. Crass

2016 Holzman : A Newly Discovered Late Pleistocene Occupation in the Tanana

Valley, Alaska. Poster Presented at the 43rd Alaska Anthropological Association

Meeting, Sitka, AK, March 2-6.

178

Yesner, David R.

1994 Subsistence Diversity and Hunter-Gatherer Strategies in late Pleistocene/ early

Holocene Beringia: Evidence from the Broken Mammoth Site, Big Delta, Alaska.

Current Research in the Pleistocene 11: 154–156.

1996 Human Adaptation at the Pleistocene-Holocene (circa 13,000 to 8,000 BP)

Boundary in Eastern Beringia. In Humans at the End of the Ice Age, edited by L.

G. Straus, B. V. Eriksen, J. M. Erlandson, and D. R. Yesner, pp. 255–276. Plenum

Press, New York.

2001 Human Dispersal into Interior Alaska: Antecedent Conditions, Mode of

Colonization, and Adaptations. Quaternary Science Reviews 20(1–3): 315–327.

Yesner, David R., G. A. Pearson, and S. E. Stone

2000 Additional Organic Artifacts from the Broken Mammoth Site, Big Delta, Alaska.

Current Research in the Pleistocene 17: 87–90.

179

APPENDIX C:

THE LION’S SHARE: HUMAN PALEOECOLOGICAL INTEGRATION IN SUBARCTIC

EASTERN BERINGIA

François B. Lanoë1, Joshua D. Reuther2,3, Charles E. Holmes3, Gregory W.L. Hodgins1,4

1 School of Anthropology, University of Arizona

2 University of Alaska Museum of the North

3 Department of Anthropology, University of Alaska Fairbanks

4 National Science Foundation - Arizona Accelerator Mass Spectrometry Laboratory

180

Abstract

The extinction of Pleistocene megafauna is central to our interpretative models of past and present ecosystem dynamics and of the sustainable integration of hunter-gatherer populations within their ecosystems. Here we contribute to the understanding of megafauna extinction and human dispersal in subarctic eastern Beringia by focusing on changes in the trophic dynamics of large mammal community as well as the ecological role of humans as a predator and competitor.

We reconstruct use of habitat by megafauna and humans throughout the Pleistocene-Holocene boundary based on zooarchaeological data and extensive isotopic sampling. Our results suggest that large herbivore population dynamics can be related to variation in the heterogeneity and availability of favorable habitat, but also to competition for habitat between members of the herbivore guild. Human dispersal in the region seems to have been favored by an increase in herbivore diversity and biomass at the beginning of the Bølling-Allerød interstadial and comparative lack of competitors. As resources dwindled later in the Late Glacial, people increasingly relied on high returns patches of bison and wapiti to the detriment of large carnivores, and are likely to have contributed to the extinction or niche shift of other Beringian predators. The patterns of biotic turnover that we observe in subarctic eastern Beringia corroborate the idea that communities are not always regulated by global, top-down processes, but are also complex systems which evolution arises from the interactions between their different agents.

181

Introduction

The fossil record of eastern Beringia at the Pleistocene-Holocene boundary has drawn much attention from archaeologists and paleoecologists, principally because of its relevance to the extinction of Late Pleistocene mammalian megafauna* and to the long-term human settlement of high latitudes of Northeastern Eurasia and North America. These issues are at the core of both disciplines in that they bear strongly on our interpretative models of the functioning of past and present ecosystems and of the sustainable integration of hunter-gatherer populations within communities of large mammals.

Processes and causes leading up to megafauna turnover have been the subject of intense discussion for decades. Some have considered megafauna extinction to be the end-result of protracted population declines, in response to habitat degradation/reduction from climatic deterioration or instability throughout the MIS 4-2 (e.g. Szpak et al. 2010). This “long-term hypothesis” is supported by genetic evidence of population stress and decline among steppe bison (Shapiro et al. 2004), brown bear (Barnes et al. 2002; Leonard et al. 2000), lion (Barnett et al. 2009), mammoth (Debruyne et al. 2008), and, from bone measurements, horse (Guthrie

2003). These climate-based hypotheses however rely for the most part on the coincidence between climate change and population decline and tend to overlook the specific processes by which megafauna would have gone extinct. Species’ tolerances depend on a variety of physical conditions and pattern of coexistence, and few of these studies have discussed how complex ecological interactions may have contributed to population dynamics (for exceptions see Guthrie

2004; Mann et al. 2013, 2015).

On the other side, hypotheses based on ecological interactions have been limited to models of extinction resulting from human predation. In eastern Beringia, these have centered on horse and

182 mammoth, which are described by some as ecologically naïve prey that human hunters would have over-harvested, surpassing their natural population replacement rate (e.g. Surovell et al.

2016). In turn, the loss of these large ecosystem engineers and of their structural influence on soil formation and vegetation succession would have triggered a larger biome shift (Zimov et al.

1995). However, this “overkill hypothesis” is, also, based foremost on the temporal coincidence between human colonization and horse and mammoth extinction (Solow et al. 2006; Buck and

Bard 2007; Haile et al. 2009; Guthrie 2006, 2003; Bradshaw et al. 2012; MacDonald et al. 2012;

Nogués-Bravo et al. 2008; Stuart et al. 2004; Surovell et al. 2016). It overlooks the facts that (i) with the exception of insular ecosystems and post-industrial human societies, large herbivore populations tend to be regulated by bottom-up (habitat) rather than top-down (predation) processes (Sinclair et al. 2003; Sinclair 2003; Hopcraft et al. 2010; Fritz et al. 2011; Owen-Smith and Mills 2008); (ii) current archaeological data suggest that eastern Beringian hunters were mostly focusing on bison and wapiti (Table 14 further in text) even when and where horse were also available (Mann et al. 2001, 2013); and (iii) some species that are particularly vulnerable to human predation due to their behavior, such as the muskox, survived into the Late Holocene.

Rather than concentrating on a few selected herbivore species, human hunters may have had more impact on the structure of the predator guild† and the overall food web. Humans have a long history of competition and co-evolution with other large predators (Brantingham 1998a,

1998b, Stiner 2002, 2004; Discamps et al. 2011; Van Valkenburgh 2001; Rodríguez-Gómez et al. 2013, 2016; Stiner 1994). As such the ecological role and impact of people in the eastern

Beringia megafauna communities may have depended on how resources were partitioned among human and non-human predators and the extent of competitive exclusion on the breadth of their respective niches. Humans as a species have a flexible diet, varying in recent times from a

183 specialized foraging to broad omnivory (e.g. Stiner et al. 2000). This versatility generally allows humans to interact with other species at multiple trophic levels and spatial scales (Brose et al.

2005) and, depending on diet choice along a continuum of specialization to generalism, people may impact the whole trophic network of large mammals, or alternatively contribute to its stabilization.

In this article we contribute to the paleoecological and archaeological issues related to megafauna turnover at the Pleistocene-Holocene boundary in subarctic eastern Beringia. In contrast to previous studies, we use an approach centered on the community-level ecological dynamics, with a particular focus on humans and their role as a predator and competitor. Data and interpretive models suggest that people had a major ecological role in the megafauna turnover, but probably more as a competitor than as a predator.

Background

Prior to the Last Glacial Maximum (LGM; about 24,000-19,000 cal B.P) eastern Beringia hosted a high biomass megafauna community which thrived in a dry, nutrient-rich vegetation landscape dominated by graminoids and high-protein forbs (Blinnikov et al. 2011; Willerslev et al. 2014;

Zazula et al. 2003; Zimov et al. 2012; Goetcheus and Birks 2001; Mann et al. 2013). This

“mammoth-steppe” biome supported a range of sympatric species that do not co-exist today, and for this reason is often considered to lack modern analogues (Guthrie 2001; Goetcheus and Birks

2001).

While the mammoth steppe ecosystem had remained essentially unchanged throughout much of the Late Pleistocene (MIS 5-2), it abruptly came to an end during the Late Glacial (about 19,000-

184

11,700 cal B.P.). The dominant vegetation type in eastern Beringia shifted to shrub tundra with dwarf birch (Betula nana) by 16,000 to 14,000 cal. BP. In the subarctic, it then transitioned to a more closed environment by 13,000 to 10,000 cal. BP, including poplar (Populus balsamifera and P. tremuloides) and white spruce (Picea glauca) (Bigelow and Powers 2001; Anderson et al.

2004; Viau et al. 2008; Edwards et al. 2001; Llyod et al. 2006), while at the same time tussock tundra became dominant in the arctic (Mann et al. 2010; Oswald et al. 2014). In both regions vegetation change was accompanied by the spread of wetlands and peatlands and a gradual stabilization of the soils (Reuther 2013; Mann et al. 2010; Jones and Yu 2010).

Dramatic changes in vegetation and soils appear synchronous with mammalian megafauna turnover (Figure 22). In the subarctic up to six species or genetically distinct populations seemingly became extinct during a short interval ca. 15,000-14,000 cal B.P. The appearance of new migrants is not as well dated; nevertheless, up to seven species spread into eastern Beringia from either western Beringia or North America. Megafaunal diversity peaked at 15-18 species around 14,000 cal. BP forming a non-analogue community whose diversity was unmatched prior to and after that time.

185

Figure 22. Known large mammal dates from subarctic eastern Beringia. Gulo presence has yet to be dated systematically, but is fairly common in Late Pleistocene eastern Beringian deposits and was probably present throughout the Ice Age (Youngman 1993; Tomasik and Cook 2005; Bryant

1987). Bison is continuous throughout the record but seems to undergo some population replacement (Heintzman et al. 2016). The single Puma date (Harington 2003) was excluded.

Dates are compiled from (Guthrie 2006; Barnes et al. 2002; Barnett et al. 2009; Shapiro et al.

2004; Zazula et al. 2014; Raghavan et al. 2014; Guthrie et al. 2001; Leonard et al. 2007; Fox-

Dobbs et al. 2008; Guthrie 2003; Harington 2011; Meiri et al. 2014; Weinstock et al. 2005;

Mann et al. 2013, 2015; Haile et al. 2009; Leonard et al. 2002; Sattler 1997; Endacott 2008;

Kurtén and Anderson 1980; Hundertmark et al. 2002; Stone and Cook 2000; Chambers et al.

2012; Potter 2008; Heintzman et al. 2016; Harington 2003) and calibrated with OxCal 4.2 and

IntCal 13 (Bronk Ramsey 1994; Reimer et al. 2013). Oxygen Isotopic Values are from Seierstad et al. (2014).

186

Approach

Paleoecological studies are often based on isotopic data with carbon and nitrogen isotope values reflecting the range of habitats and resources used by the animal during its life - its

Hutchinsonian niche (Newsome et al. 2007; Flaherty and Ben-David 2010). Niche is constrained to some extent by an animal’s morphology and is generally conserved through time in the fossil record (Martinez-Meyer et al. 2004); however, there is considerable flexibility in dietary and habitat use strategies among clades such as canids (Meachen et al. 2014; Pardi and Smith 2015;

Newsome et al. 2016), ursids (Bocherens 2015; Milakovic and Parker 2013), cervids (Drucker et al. 2011; Rivals and Semprebon 2016) and bovids (Rivals and Semprebon 2011; Guthrie 1980), enabling members of a same guild to occupy significantly different niches (Kartzinel et al. 2015).

There is an apparent abundance of isotopic data from Beringian megafauna (Bocherens et al.

1996; Iacumin et al. 2000; Bocherens 2003, 2015; Fox-Dobbs et al. 2008; Yeakel et al. 2013;

Guthrie 2001; Rivals et al. 2010; Mann et al. 2013; Leonard et al. 2007). However, these studies

(with the exceptions of Mann and colleagues’ work in arctic eastern Beringia [2013, 2015]) lack spatial and temporal resolution, sometimes retaining scales as large as the whole mammoth steppe (Iberia to the Yukon) and/or the entire Late Pleistocene. In contrast, we considered for this study a restricted region and period with a megafauna community formed of sympatric, co- extant species. We conducted additional isotopic sampling to obtain an adequate sample of specimens, increasing by more than three folds the existing dataset.

187

Data on human diets were derived from the archaeological literature and unpublished data from the authors, acknowledging that reconstructing diet from animal bones left at archaeological sites is biased to some extent by taphonomic processes.

Figure 23. Map of the study area and geographic distribution of the samples used in the study.

Current glaciers are represented in black, and cross-hatching represents the limit of the

Cordilleran ice-sheet during the Last Glacial Maximum (following Manley and Kaufman 2002;

Geomatics Yukon 2015; Manley 2002; Dyke et al. 2003).

188

Results

Isotopic Values

Isotopic values for primary consumers are typical of the mammoth-steppe biome overall and similar to those reported for earlier periods, with herbivores δ13C values ranging from -19 to -22

‰ and δ15N values from ~0 to 8 ‰ (Bocherens 2015).These herbivore δ13C values are markedly higher than those from Middle to Late Holocene Alaskan environments, where primary consumer δ13C values typically range from -23.0 to -26.5 ‰ in boreal forests (Drucker et al.

2010; Fox-Dobbs et al. 2007; Urton and Hobson 2005; Ben-David et al. 2001) and -21.5 to -25.0

‰ in more open parklands (Fox-Dobbs et al. 2012) or shrub tundra (Ben-David et al. 2001;

Dalerum et al. 2009). Moose δ13C values in Alaska decrease significantly by 9,000 cal B.P. only

(Guthrie 2006), corroborating pollen studies that indicate a more forested environment with a rise in Picea at this time (Bigelow and Powers 2001). δ15N values decrease for most of the species sampled by around 14,500-13,500 cal B.P. (Figure 24); see also (Fox-Dobbs et al. 2008), suggesting environmental change at the community scale (Drucker et al. 2011).

189

Figure 24. Recorded isotopic values for subarctic eastern Beringia megafauna.

Prior to 14,300 cal. B.P., large herbivores partitioned the isoscape with little or no overlap

(Figure 25). Mammoth were at one extreme of the grazing continuum, followed by horse; bison and wapiti used habitats with less graminoids and more forbs; while small bovids seem to have specialized on forbs. From 14,300 to 11,700 cal B.P., while the mammoth niche remained unchanged, bison-wapiti and horse shifted to wetter habitats (Table 13) with less graminoids

190 where they seem to have overlapped partially with the new migrant moose, while caribou shifted to extreme niches with very low nitrogen values. After 11,700 cal B.P., bison-wapiti and moose occuped alone a more generalist niche, with bison-wapiti possibly using more graminoid- dominated habitats after 10,500 cal B.P.

Wilks' Λ F d.f. p-value

Bison 0.47 5.67 2, 10 0.02

Canis 0.22 7.19 2, 4 0.05

Cervus 0.92 0.66 2, 15 0.53

Equus 0.54 6.30 2, 15 0.01

Mammuthus 0.91 0.84 2, 16 0.45

Panthera 0.93 0.25 2, 7 0.79

Rangifer 0.22 3.53 2, 2 0.22

Ursus 0.48 6.03 2, 11 0.02

Table 13. Multivariate analysis of variances in stable isotope ratios of megafauna prior and after

14,300 cal B.P. P-values under 0.05 (in bold) indicate significant differences in the recorded stable isotope ratios in the two periods.

191

Figure 25. Habitat partitioning through time for large herbivores. Time intervals are defined following breaks in isotopic trends (Figure 24) and last between 2,300 and 2,700 years each. The algorithm used to create the ternary graphs allows for negative values (e.g. caribou).

192

Figure 26. Partitioning of the predator guild prior and after 14,300 cal B.P. Labels correspond to dietary tendencies discernable from stable isotopes.

Large carnivores prior to 14,300 cal B.P. have similar isotopic values with all three species behaving like meso- or hypercarnivores (Figure 26). Stable isotopes alone are not sufficient to distinguish specific strategies of resource use; nevertheless, isotopic values show overall variation across species that suggest it is unlikely that they shared the same niches. Lion may have been specialized on drier habitats and/or caribou; wolves more generalized or, alternatively,

193 specialized on resources with “average” isotopic values; and bears probably incorporated more vegetation in their diet. After 14,300 cal B.P., bear and wolf seem to move away from the region of overlap. Bears show an increase in δ13C values and a very diverse range of δ15N values suggesting greater use of vegetation and of freshwater or anadromous fish (Table 13). The single wolf specimen for this period shows a marked decrease in δ15N (Table 13), suggesting a shift to a hypocarnivorous diet or to a specialized predation on moose, and generally corroborating morphological and osteometric data that indicate that Holocene wolves were less carnivorous than their Pleistocene counterparts (Leonard et al. 2007). On the other hand, values for lion specimens remain in the variability observed in the previous period (Table 13).

Zooarchaeological Results

Faunal remains abandoned after consumption at archaeological sites in subarctic eastern Beringia provide information on human dietary strategies (Table 14). At the oldest Swan Point CZ4b occupation (about 14,000 cal B.P.), people consumed horse and neonatal mammoth (Lanoë and

Holmes 2016). After about 14,000 cal B.P., human dietary strategies shifted to bison and wapiti with occasional use of sheep, caribou, and moose. Small game (lagomorph, rodents, Anatidae,

Phasianidae, and Salmonidae) are present throughout the archaeological record but, because of their smaller body weight, were probably secondary in terms of calories consumed or only seasonally important (Choy et al. 2016).

194

Date

Site (Cal B.P.) References

Horse Mammoth* Bison Wapiti Moose Sheep Caribou Lagomorph/ Rodent* Bird Fish Hollembaek Hill 8,100-7,800 x x x x this paper

Potter 2008; Holmes et al. Swan Point CZ2 8,370-8,040 x 1996

Broken Mammoth CZ2 8,590-8,220 x x x x Potter 2008; Holmes 1996

Cook CZ2 8,970-8,490 ? ? this paper

Gerstle River C3 10,160-9,870 33 72 Potter 2007

Easton et al. 2011 ; Yesner

Little John Paleosol 11,070-10,700 76 3 4 27 x x et al. 2011 ; Potter et al.

2013

Potter et al. 2014; Choy et Upward Sun River C3 11,610-11,280 x x al. 2016

Swan Point CZ3 11,960-11,400 5 3 4 1 2 Potter et al. 2013; this paper

Upward Sun River C2 11,990-11,510 x Choy et al. 2016

Mead CZ3b 12,120-11,850 10 1 2 4 1 Potter et al. 2013

Broken Mammoth CZ3 12,390-11,770 133 87 ? ? ? 33 97 ? Yesner 2001

Cook CZ3 12,530-12,050 x this paper

Mead CZ4 13,110-12,790 5 2 Potter et al. 2013

Upward Sun River C1 13,300-13,120 x x x Choy et al. 2016

Altamira Consulting 2014; Britannia Creek 13,930-12,730 2 1 x this paper

Reuther et al. 2016; this Keystone Dune 13,400-13,220 6 paper

Mead CZ5 13,440-13,200 x Potter et al. 2013

Broken Mammoth CZ4 13,650-13,150 21 44 44 801 ? Yesner 2001

Easton et al. 2011 ; Yesner Little John Loess 14,050-13,720 1 1 et al. 2011

Potter et al. 2013; Lanoë and Swan Point CZ4b 14,150-13,870 10 12 ? ? 1 4 66 Holmes 2016

195

Table 14. Zooarchaeological data (NISP, when available) on animals consumed at subarctic eastern Beringian sites throughout the Late Pleistocene and early Holocene. Mammoth tooth fragments as well as cervid antler (when indicated in the publication) are excluded, as well as faunal specimens likely to represent intrusions from other stratigraphic levels. Faunal frequencies for the Little John site are estimates based on published graphs. For the Broken Mammoth site, it is unclear from extant publications which components contain sheep, caribou or moose remains.

Discussion

Herbivore Guilds

Immediately following the LGM, megafauna biomass in subarctic eastern Beringia was dominated by the same three species as before the LGM, mammoth, horse, and bison (Guthrie

1990), in what must have been a typical glacial steppe. Large herbivores partitioned resources with little overlap and were distributed along a continuum from exclusive grazer to more selective feeder. Overall, herbivore megafauna probably partitioned resources following the ecological rule that herbivores forage at different spatial scales depending on their body size and physiology, with larger herbivores, particularly cecalids (hindgut fermenters), eating lower quality food but in higher quantities, therefore over a larger spatial scale. (Cromsigt and Olff

2006; Kleynhans et al. 2011; Ritchie and Olff 2004), and at the other end of the spectrum smaller herbivores and ruminants selecting high quality food at a finer scale (Riginos and Grace 2008;

Guthrie 1984). Mammoth grazed habitats over large ranges enriched in 15N, resulting from dung-fertilization or higher aridity (Schwartz-Narbonne et al. 2015; Kirillova et al. 2015). In contrast, horse and bison must have used a more diverse range of smaller habitats, or of different

196 vegetation succession stages, while smaller ruminants must have been relegated to marginal habitats in terms of available biomass.

Large herbivore diversity increased around 15,000 cal B.P. with the immigration of wapiti and saiga antelope in subarctic eastern Beringia. All herbivore species present on the landscape subsisted on graminoid-rich habitats, suggesting that available niche space expanded at that time.

Work by Mann et al. (2013, 2015) in arctic eastern Beringia shows that the period around

15,000-14,000 cal B.P. favored a higher graminoid biomass because of a disequilibrium between climatic conditions and pace of plant dispersal. Temperature was warmer but vegetational and soil succession had not had enough time to occur, briefly providing optimal growing conditions for graminoids and in turn optimal grazing conditions for numerous sympatric herbivores on different micro-habitats or different successional stages of the same plants. Immigration of wapiti and saiga antelope must therefore have been favored by reduced competition for graminoid habitat within the herbivore guild.

Rise of the water table, peatland development, soils stabilization, and vegetational succession by

14,500-13,000 cal B.P. contributed to the fragmentation of habitat and the buildup of an increasingly heterogeneous landscape. Depending on local conditions, patches must have been dominated either by wetlands, shrub Betula, Populus trees, or graminoids. Moose migrated into subarctic eastern Beringia at that time to exploit a newly available, more hydric niche. Persisting graminoids patches were maintained locally by continuing aeolian and/or alluvial deposition

(Reuther et al. 2016; Reuther 2013), but their size and productivity must have decreased down to a threshold below which they could not support a large diversity of large herbivores. Indeed, after 14,300 cal B.P. most large herbivores shifted towards wetter habitats. Some horse populations may also have relocated northwards in arctic eastern Beringia, where they persist

197 later in more favorable arid habitats (Mann et al. 2013, 2015). Mammoth, unlike other large herbivores, continued to use shrinking graminoid habitats. As specialized grazers (Schwartz-

Narbonne et al. 2015), mammoth may have lacked the ecological adaptability to exploit another niche; alternatively, they may have monopolized shrinking graminoid habitat and out-competed other large herbivores.

Constant graminoid habitat reduction and fragmentation in the face of the development of forested habitats must have led to the eventual demise of the mammoth and horse shortly before

13,000 cal B.P. Graminoid habitat size must have reached a threshold at which a very large cecalid such as mammoth could not be sustained anymore, as has been found in other settings

(Graham et al. 2016). The lack of adequately-sized grazer habitat must also have provoked the demise of horses. While they can, to some extent, subsist on some foliate vegetation in addition to graminoids, horses must eventually have been outcompeted for “edge” niches by wapiti and bison which, as ruminants, are more efficient at processing limited amounts of resources. Both mammoth and horse persisted later in the arctic, but in treeless habitats with fewer competitors

(Mann et al. 2015, 2013).

The landscape in subarctic eastern Beringia became more homogeneously dominated by wetlands and trees after 9,000 cal B.P., an environment in which a specialized browser such as moose must have thrived. Bison and wapiti nevertheless persisted as recently as 2,000 cal B.P.

(Stephenson et al. 2001; Guthrie 2006; Heintzman et al. 2016). In the absence of large-bodied competitors after the demise of horse and mammoth, bison and wapiti must have heavily used graminoid habitats as neither species can subsist on a specialized browser diet and grasslands must have been particularly attractive in providing high-productivity resources. Bison and wapiti must also have helped maintaining grasslands by inhibiting shrub encroachment, creating

198 constant disturbance through trampling and browsing (Asner et al. 2009; Bakker et al. 2015), akin to how moose maintain spruce parkland environments optimal for browsing (Pastor et al.

1993; McInnes et al. 1992). This type of positive feedback must have ensured the persistence of wapiti, bison, and graminoid habitats until very late in the Holocene. Indeed, the steppe bison

Bison priscus persisted in central Alaska as late as 2,000 cal B.P. even as the more browsing wood bison Bison (bison) athabascae had taken hold further east in the Yukon (Heintzman et al.

2016).

Predator Guild

Populations of large predators tend to be regulated by bottom-up processes, specifically by the availability of large herbivore biomass and by competition within the predator guild (Van

Valkenburgh et al. 2015; Sinclair 2003). Prior to the Late Glacial, a relatively high diversity and biomass of herbivores allowed the co-existence of several carnivores in subarctic eastern

Beringia, including lion, wolf and brown bear, but probably also wolverine as a specialized low- productivity scavenger (Dalerum et al. 2009; Inman et al. 2012). In contrast to earlier periods of the Late Pleistocene though, herbivore biomass levels did not allow the presence of a specialized active scavenger such as the short-faced bear (Arctodus simus) (Bocherens 2015; Matheus 1995), ecological equivalents of which can be found in the Eurasian mammoth steppe (Turner et al.

2008).

Some degree of hierarchical dominance within the predator guild must have regulated access to the highest-ranked resources. Predator guild dominance tends to be correlated with body size.

When present, lions dominate other predators for access to resources (Donadio and Buskirk

199

2006; Cozzi et al. 2012; Vanak et al. 2013), and eastern Beringia lion was probably no exception.

No evidence points to frequent predation of mammoth by either species of large carnivore

(Schwartz-Narbonne et al. 2015), even as osteometric data indicates that they could have potentially preyed upon juvenile megaherbivores (Van Valkenburgh et al. 2015). Lion, which as a species tend to be generalist meat-eaters (Pérez-Claros and Palmqvist 2008), seem to have specialized somewhat on drier habitats or on caribou in Beringia (Bocherens 2015). Wolves tend to be more specialized carnivores (Pérez-Claros and Palmqvist 2008; Newsome et al. 2016), and the Beringian hypercarnivorous wolves (Leonard et al. 2007) may have specialized on herbivores with “average” isotopic values, such as bison. As for brown bears, while they were more carnivorous in Beringia than typically observed after benefitting from a competitor release after the extinction of the short-faced bear (Bocherens 2015), they were nevertheless less carnivorous than co-extant lions and wolves.

Opening of additional niche space within the predator guild around 15,000-14,000 cal B.P., as resource productivity and herbivore biomass increased, must have favored human colonization in subarctic eastern Beringia. The limited archaeological data pertaining to the earliest dispersal of people suggest some degree of historical continuity in economic strategies: though we don’t known if these dietary items, people at Swan Point CZ4b consumed horse, neonatal mammoth, and hare, the same preys than found at the site of Berelekh, in western Beringia (Pitulko 2011).

Changes in the partitioning of the predator guild can be observed after about 14,300 cal B.P., each species showing a different response to the decrease in herbivore diversity and biomass at that time. Brown bears seemingly abandoned hypercarnivory for dietary strategies ranging on a continuum of hypocarnivory to piscivory. This shift in diet is roughly contemporaneous with the proposed first appearance of anadromous fish in the region (Halffman et al. 2015), and probably

200 relates to biogeographical in addition to ecological factors. The Pleistocene hypercarnivorous wolf was probably replaced by a more hypocarnivorous population (Leonard et al. 2007). Lion use of habitat remains seemingly unchanged, until their sudden extinction around 14,000 cal B.P.

Bear and/or wolf population also go down at that time with a decrease of bone deposition rates in karstic cavities (Sattler et al. 2001).

Humans began to intensively focus on wapiti and bison starting about 14,000 cal B.P. Wapiti and bison must have been the highest-ranked resources for predators at that time. They most likely provided the most in terms of resource quantity and availability by being present at high densities in restricted and unstable patches on the landscape dominated by early vegetation and soil successional stages, in contrast to solitary moose scattered throughout more forested environments. People specializing on these unstable habitats, as they did later in arctic eastern

Beringia (Mann et al. 2001), became the dominant predator. Rather than benefitting from the demise of other predators, they probably displaced them to more mesocarnivorous diets and/or less productive, marginal niches. In particular wolves shifted to lower trophic levels even though they could have benefitted instead of a mesopredator release after the extinction of the lion, mimicking in this wolf populations in other regions of North America at the time of human arrival (Pardi and Smith 2015; Meachen et al. 2014). People seem to have taken the lion’s share, quite literally.

Conclusions

Patterns of biotic turnover in subarctic eastern Beringia at the Pleistocene-Holocene transition suggest that communities are not necessarily regulated by global, top-down processes most of the

201 time; they are also complex systems whose patterning arises from the interaction between their agents (Ings et al. 2009; Morin 2011). The processes leading to the extinction of megafauna in eastern Beringia seem to have resulted at least partly from complex feedbacks. Ultimately, a top- down process, climate change, triggered biome change from the Ice-Age steppe towards a new attractor more typical of the Holocene boreal forest biome. However this turnover functioned more like a sudden system shift rather than a gradual process accompanying changes in temperature/humidity, and this pace of change can be related to nonlinear interactions amongst biota.

Bottom-up processes resulting from ecological interactions amongst herbivores and between herbivores and plant communities play a large role in ecosystem dynamics (Sinclair 2003;

Sinclair et al. 2003; Hopcraft et al. 2010; Fritz et al. 2011; Owen-Smith and Mills 2008).

Ecosystems in which climate could allow the dominance of forests are not necessarily forested

(Sankaran et al. 2005; Bond 2008) because megaherbivores maintain open habitat through constant disturbance slowing vegetation succession (Johnson 2009; Bakker et al. 2015), as well as with lateral transport of nutrients (Olofsson et al. 2009, 2004; Post and Pedersen 2008;

Doughty et al. 2015; Wolf et al. 2013); this is particularly true for diverse herbivore guilds

(Bakker et al. 2015) and for proboscideans (Haynes 2012). In subarctic eastern Beringia, the mammoth steppe biome remained remarkably stable until the Late Glacial as short-term instabilities, albeit frequent (Mann et al. 2015), did not overcome the built-in resistance and resilience of the ecosystem (Gunderson 2000; Folke 2006). However, the eventual loss of major ecological engineers such as mammoth and horse must have released plant communities from predation and led the pace to vegetational and soils succession and paludification. Predation release for plants must have triggered an increase in fire cycle frequency (Bond and Keeley

202

2005), visible in the geological record between 13,000 and 10,000 cal. B.P. in central Alaska

(Reuther et al. 2016) and elsewhere in North America (Gill et al. 2009, 2012).

People were actors in the eastern Beringian biotic renewal. Abundance of herbivore resources and comparative lack of predators seem to have favored human dispersal in subarctic eastern

Beringia, as it did at several stages of the human evolutionary history (e.g. Palombo and Mussi

2006; Croitor and Brugal 2010). People do not seem to have any major adverse effect on large herbivore population dynamics, which can be instead related to variation in the heterogeneity and availability of favorable habitat. Habitats did not change radically during that time; but the scale to which they were expressed in the landscape did, forcing new modes of habitat partitioning in the herbivore guild and contributing to a species turnover. On the other hand, people probably contributed to the extinction of large carnivore species or populations. Passed the initial colonization stage, people increasingly focused on high returns resource patches, and this to the probable detriment of other large predators. While a negative impact on large predator populations seem to have been a recurrent process accompanying human dispersal throughout the globe (Pardi and Smith 2015; Discamps et al. 2011), Beringian hypercarnivores would probably have gone extinct regardless of human action, but later in the Holocene as bison and wapiti populations declined and eventually disappeared from the region (Stephenson et al. 2001).

Methods

This study provides 190 new isotopic values and 11 radiocarbon dates; the other 76 isotopic values were obtained from the literature. All radiocarbon dates were calibrated with OxCal 4.2 and IntCal13 (Bronk Ramsey 2009; Reimer et al. 2013); when discrete ages are used in this

203 article (Figure 24), they correspond to the average of the ends of the confidence interval at 2σ.

The complete list of isotope samples and values, along with anatomical element, provenience, curation facility, radiocarbon date, and references, is provided as supplemental material (Table

15).

Materials for the present study include specimens of the twelve megafauna species dated between 19,000 and 10,000 cal. BP in central eastern Beringia, considering the Tanana and upper

Yukon basins from the Alaska Range to about 65° north of latitude, and west of the Cordilleran ice sheet to about 150° of longitude (Table 15). This area includes the fossil-rich regions of

Fairbanks, Chicken Creek, and Klondike, as well as several well-dated and clearly stratified archaeological sites that provide a reliable record of early human presence in eastern Beringia.

Specimens sampled include individually dated paleontological specimens collected in mining placers as well as archaeological specimens dated by stratigraphic association. The study area spans several modern ecoregions (Yukon-Tanana Uplands, Tanana Lowlands, Yukon Plateau and Flats) but a similar geographic context overall: the transitional area between the lowlands and the uplands or in low drainages within the uplands, about 150-700 m above mean sea level, and generally intermediate between the lowlands and the uplands (Figure 23).

Carbon and nitrogen isotope values vary significantly among subarctic C3 plants (Wooller et al.

2007; Valentine et al. 2006; Ben-David et al. 2001) because of differential isotopic fractionation.

Variation in δ13C values reflect differences between vascular plants and lichen/fungi (Drucker et al. 2001), as well as habitat differences among vascular plants including amount of vegetation cover (Drucker et al. 2008) and overall aridity (Fox-Dobbs et al. 2008). Variation in δ15N values reflects different modes of nitrogen fixation (Szpak 2014) and overall temperature and aridity

(Stevens and Hedges 2004). Tissues of secondary and tertiary consumers in turn reflect the

204 isotopic values of their diet, along with a measurable enrichment in heavier isotopes at each trophic level averaging 1.1 ± 0.2‰ for δ13C and 3.8 ± 1.1‰ for δ15N (Bocherens 2015). The identification of specific dietary items is generally not possible in an ecosystem where the isotopic signature of primary producers overlaps widely (Bocherens 2015; Fox-Dobbs et al.

2008; Ben-David et al. 2001) and where isotope values can also reflect abiotic conditions

(Stevens and Hedges 2004). Variation in isotopic values among sympatric species of herbivores and carnivores nevertheless reflects the relative partitioning of habitat and resources at the guild level.

In order to relate isotopic values to habitat and resources partitioning among sympatric herbivores, carbon/nitrogen bi-dimensional isoscapes were transformed to a tri-dimensional space in which the three poles are defined from extreme habitat use strategies that can be discerned from carbon and nitrogen isotopes (Fox-Dobbs et al. 2008; Bocherens 2015; Ben-

David et al. 2001; Schwartz-Narbonne et al. 2015): (i.) habitat with more graminoids, or grazer

(higher δ15N, lower δ13C; here, values of 10 and -23 ‰, respectively); (ii.) wetter habitat with more woody plants, or browser (lower δ15N and δ13C; -3 and -23 ‰); and (iii.) habitat with more forbs and/or lichen, or selective feeder (average δ15N, higher δ13C; 4 and -17 ‰).

Equations (1) to (3) describe the system used to create ternary graphs, with δ13C and δ15N the isotope values of a specimen i; pW, pG and pFL the proportions on the ternary graph of wetter- woody, graminoids, and forb-lichen habitats, respectively, for specimen i; and δ13Cx and δ15Nx the isotope values of the corresponding habitat extreme x.

13 13 13 13 훿 퐶 = 푝푊푖. 훿 퐶푊 + 푝퐺푖. 훿 퐶퐺 + 푝퐹퐿푖. 훿 퐶퐹퐿 (1)

15 15 15 15 훿 푁 = 푝푊푖. 훿 푁푊 + 푝퐺푖. 훿 푁퐺 + 푝퐹퐿푖. 훿 푁퐹퐿 (2)

205

푝푊푖 + 푝퐺푖 + 푝퐹퐿푖 = 1 (3)

Reconstructing carnivore diet from isotopic values can also be ambiguous. While strategies such as piscivory or hypocarnivory leave recognizable isotopic signatures, it is difficult to differentiate, with isotopic values alone, diets of a mesocarnivore, a generalist hypercarnivore, or a hypercarnivore specialized on an isotopically “average” herbivore (Bocherens 2015). As in the case of herbivores, isotope values for carnivores reflect relative resource partitioning within guilds rather than specifying the details of diet.

Most specimens sampled are cortical bone fragments of adult individuals for which isotopic values reflect nutrient intake over a period of several months or years prior to the death of the animal (Bocherens 2015). Other skeletal materials were assayed in some cases: cervid antler (N

= 4) which reflect nutrient intake during the summer season of antler growth; carnivore brachyodont teeth (N = 1) in which growth stops early in life; and bone from juvenile individuals

(N = 1). For the carnivore teeth and juvenile bone samples, isotopic values were adjusted to correct for metabolic fractionation between mother and nursing offspring at the time of tooth/bone formation (following Jenkins et al. 2001; Bocherens 2015).

Collagen extraction for this study was conducted at the University of Arizona AMS Laboratory using the modified Longin method. A 69 to 1,634 mg sample (depending on the specimen's state of preservation) was extracted with a diamond-coated rotary wheel and exogenous sediments and materials adhering to the surface were cleaned off with a milling bit. The sample was then crushed into a powder retaining a grain size between 0 or 0.5 and 1.0 mm and washed using standard acid-base-acid (ultrapure water, 0.5 M HCl, 0.1 M NaOH, and 0.01 M HCl). One

206 collagen specimen was obtained from the National Ocean Sciences AMS Laboratory where collagen was extracted with EDTA (Tuross 2012) rather than HCl demineralization. The collagen fraction was then washed into test tubes, heated at 70°C for 24 hours to gelatinize, filtered, and freeze-dried. Isotopic ratios of carbon and nitrogen as well as carbon and nitrogen content were measured at the Environmental Isotope Laboratory of the University of Arizona on a continuous-flow gas-ratio mass spectrometer (Finnigan Delta PlusXL). Samples of 1.0 mg of collagen were combusted in an elemental analyzer (Costech). Standardization is based on acetanilide for elemental concentration, NBS-22 and USGS-24 for δ13C, and IAEA-N-1 and

IAEA-N-2 for δ15N. Precision is better than ± 0.08 for δ13C and ± 0.2 for δ15N (1σ), based on repeated internal standards. Stable isotope compositions are reported in ‰ as δ (e.g. δ13C =

[(13C/12C sample) / (13C/12C standard) - 1] * 1,000) and are referenced to Vienna PeeDee

Belemnite for carbon, and air for nitrogen. All samples reported have C/N atomic ratios ranging from 2.2 to 3.6 and nitrogen contents > 5 % known for viable collagen (Ambrose 1990; Fox-

Dobbs et al. 2008). Two outliers with small collagen yields (< 2.5 %) were excluded from further analysis (Table 15).

207

Supplemental Dataset 1: Table 15, Isotopic Values Obtained for this Study

14C Average Collagen Repository Catalog Alt. Catalog Species Locality Description +/- Cal. Date B.P. Laboratory Code Reference for Date δ15N δ13C % C % N C:N Reference for Isotope ValuesNotes Date B.P. Cal. Date Yield (%) AMNH F:AM 33588 Alces alces Fox not found in museum - probably typo for 34588, mandible (bone) 8,839 83 10,191 9,631 9,911 AA-14245 Guthrie 2006 2.4 -20.3 6.6 43.0 15.4 2.8 this paper AMNH F:AM 34602-E Alces alces Fox metatarsal 9,000 80 10,374 9,796 10,085 AA-14340 Guthrie 2006 2.2 -20.1 7.0 45.1 16.1 2.8 this paper AMNH F:AM 144728 A-276-5284 Alces alces Fairbanks Creek mandible (bone) 9,114 81 10,521 9,975 10,248 AA-14242 Guthrie 2006 2.6 -20.0 13.6 44.2 15.7 2.8 this paper AMNH F:AM 3459A Alces alces Cleary not found in museum 9,544 80 11,166 10,606 10,886 AA-14272 Guthrie 2006 4.7 -19.9 12.6 30.6 11.1 2.8 this paper AMNH F:AM 5312 Alces alces Fairbanks Creek metatarsal 9,650 77 11,208 10,758 10,983 AA-14338 Guthrie 2006 2.3 -20.1 15.8 31.8 10.8 2.9 this paper AMNH F:AM 2651 Alces alces Cripple Creek mandible (tooth) (crown & root) 9,927 67 11,695 11,214 11,455 AA-14238 Guthrie 2006 2.6 -19.8 1.8 37.8 13.8 2.7 this paper AMNH F:AM 4734 Alces alces Hunter Creek cranium (tooth) (crown & root) 10,062 77 11,969 11,286 11,628 AA-14277 Guthrie 2006 2.7 -19.9 14.0 43.7 15.9 2.7 this paper AMNH F:AM 144733 A 570-5174 Alces alces Fairbanks Creek metatarsal 10,756 64 12,750 12,578 12,664 AA-14326 Guthrie 2006 0.8 -20.0 9.9 43.3 15.5 2.8 this paper AMNH F:AM 144732 A 201-6776 Alces alces Lower Goldstream mandible (bone) 11,000 80 13,041 12,726 12,884 AA-14274 Guthrie 2006 1.2 -21.1 9.6 44.8 15.2 2.9 this paper AMNH F:AM 34590 Alces alces Cleary Creek mandible (tooth) (root) 11,210 81 13,257 12,861 13,059 AA-14250 Guthrie 2006 2.9 -19.8 4.7 40.1 14.6 2.7 this paper AMNH F:AM 144730 A 108-7779 Alces alces Ester Creek mandible (bone) 11,267 95 13,322 12,913 13,118 AA-14258 Guthrie 2006 1.6 -20.1 14.7 41.5 14.3 2.9 this paper AMNH F:AM 1387 Alces alces Fairbanks Creek not found in museum, probably metapodial 11,289 72 13,291 13,045 13,168 AA-14335 Guthrie 2006 1.7 -20.2 14.5 55.2 20.0 2.8 this paper AMNH F:AM 144735 A 575-5172 Alces alces Fairbanks Creek metatarsal 11,671 76 13,715 13,332 13,524 AA-14334 Guthrie 2006 1.2 -19.6 9.4 43.6 15.6 2.8 this paper AMNH F:AM 34592-C Alces alces Cleary Creek mandible (tooth) (crown & root) 11,678 81 13,721 13,345 13,533 AA-14291 Guthrie 2006 2.8 -20.0 4.8 31.4 11.4 2.8 this paper AMNH F:AM 144729 A-276-5282 Alces alces Fairbanks Creek mandible (bone) 11,790 117 13,946 13,391 13,669 AA-14244 Guthrie 2006 1.6 -19.9 12.6 43.1 15.3 2.8 this paper AMNH F:AM 144734 A 417-1121 Alces alces Fairbanks Creek metatarsal 11,947 89 14,031 13,566 13,799 AA-14327 Guthrie 2006 1.3 -20.1 12.3 44.7 15.9 2.8 this paper AMNH F:AM 3228 Alces alces Dome Creek mandible (tooth) (crown & root) 11,988 82 14,063 13,602 13,833 AA-14268 Guthrie 2006 2.7 -20.6 3.6 45.0 16.4 2.7 this paper AMNH F:AM 2165 Alces alces Fairbanks Creek cranium (bone) 12,022 153 14,336 13,474 13,905 AA-14260 Guthrie 2006 2.2 -19.8 10.6 44.2 16.0 2.8 this paper AMNH F:AM 144731 A 328-5244 Alces alces Fairbanks Creek mandible (bone) 12,201 85 14,476 13,781 14,129 AA-14266 Guthrie 2006 0.6 -20.5 7.7 43.9 15.2 2.9 this paper SHPO 2478 Alces alces Swan Point CZ3 (XBD-146) incisor 10,100 11,700 dated by strat. Association Lanoë and Holmes 2016 0.7 -20.2 6.3 35.1 11.7 3.0 this paper YC 4320-02 Bison priscus Little John CZ7 (KdVo-6) - 9,983 60 11,710 11,250 11,480 AA-108777 this paper -0.5 -20.4 0.5 41.4 12.9 3.2 this paper excluded from analysis (see text) YC Fa14-10-01 Bison priscus Little John CZ7 (KdVo-6) tibia 9,996 70 11,760 11,250 11,505 AA-108779 this paper -0.6 -20.3 2.3 42.2 13.2 3.2 this paper excluded from analysis (see text) YC Fa13-051 Bison priscus Little John CZ7 (KdVo-6) naviculocuboid 10,110 30 11,978 11,414 11,696 Beta-382339 Easton, pers. com. 2016 2.9 -19.8 - - - - Easton, pers. com. 2016 YC Fa03-14 Bison priscus Little John CZ8 (KdVo-6) cranium (tooth) 11,103 63 13,089 12,802 12,946 UOC-3255 Easton, pers. com. 2016 2.7 -20.1 - 36.1 12.8 2.8 Easton, pers. com. 2016 UAF UAM:ES:5936 V-54-17 Bison priscus Lost Chicken Creek humerus 12,340 65 14,743 14,082 14,413 OxA-10679 Shapiro et al. 2004 3.5 -19.8 11.4 40.2 14.1 2.9 this paper UAF UAM:ES:5848 V-54-365 Bison priscus Lost Chicken Creek tibia 12,380 60 14,806 14,124 14,465 OxA-11226 Shapiro et al. 2004 3.7 -19.8 8.8 37.2 13.1 2.8 this paper YC Fa06-131-01 Bison priscus Little John CZ8 (KdVo-6) vertebra 12,433 90 15,020 14,160 14,590 AA-108778 this paper 3.6 -20.4 5.4 41.2 13.5 3.1 this paper UAF UAM:ES:6244 V-54-320 Bison priscus Lost Chicken Creek humerus 12,465 75 15,060 14,208 14,634 OxA-11245 Shapiro et al. 2004 4.1 -20.1 13.9 42.0 14.8 2.8 this paper UAF UAM:ES:6137 V-54-1105 Bison priscus Lost Chicken Creek radio-ulna 12,525 70 15,116 14,321 14,719 OxA-10541 Shapiro et al. 2004 2.7 -20.0 5.5 40.2 14.0 2.9 this paper UAF UAM:ES:6846 V-54-677 Bison priscus Lost Chicken Creek humerus 12,665 65 15,279 14,760 15,020 OxA-10855 Shapiro et al. 2004 3.5 -20.0 2.4 40.9 14.2 2.9 this paper UAF UAM:ES:6254 V-54-226 Bison priscus Lost Chicken Creek humerus 12,960 70 15,755 15,250 15,503 OxA-10538 Shapiro et al. 2004 3.5 -20.0 7.6 35.8 12.4 2.9 this paper UAF UAM:ES:6136 V-54-1088 Bison priscus Lost Chicken Creek radio-ulna 13,160 70 16,068 15,554 15,811 OxA-10540 Shapiro et al. 2004 3.2 -20.0 7.1 41.9 14.6 2.9 this paper NA (lost AMNH F:AM 144743 A-556-4160 Bison priscus Fairbanks Creek metacarpal 13,235 65 16,129 15,695 OxA-11617 Shapiro et al. 2004 3.6 -19.7 44.6 15.9 2.8 this paper 15,912 about half) UAF UAM:ES:6247 V-54-1157 Bison priscus Lost Chicken Creek femur 13,240 75 16,163 15,681 15,922 OxA-10543 Shapiro et al. 2004 3.3 -20.2 7.8 40.8 14.4 2.8 this paper SHPO 26064 Bison priscus Swan Point CZ3 (XBD-146) humerus 10,100 11,700 dated by strat. Association Lanoë and Holmes 2016 2.8 -19.3 11.6 40.9 15.0 2.7 this paper AMNH F:AM 30450 Canis lupus Lower Goldstream cranium 7,751 64 8,645 8,406 8,526 AA-48695 Leonard et al. 2007 5.6 -19.4 - - - 2.2 Leonard et al. 2007 AMNH F:AM 67165 Canis lupus Gilmore Dragline cranium 12,600 150 15,358 14,188 14,773 AA-42317 Leonard et al. 2007 8.6 -19.2 - - - 2.2 Leonard et al. 2007 AMNH F:AM 67157 Canis lupus Goldstream Yukon-Koyukuk cranium 14,690 190 18,361 17,436 17,899 AA-42315 Leonard et al. 2007 7.9 -18.7 - - - 2.3 Leonard et al. 2007 AMNH F:AM 70944 Canis lupus Gold Hill mandible (bone) 15,268 169 18,871 18,121 18,496 AA-38449 Leonard et al. 2007 7.3 -19.2 - - - 2.7 Leonard et al. 2007 AMNH F:AM 30447 Canis lupus Upper Cleary Creek femur 15,580 190 19,320 18,438 18,879 AA-35223 Leonard et al. 2007 7.9 -19.7 - - - 2.2 Leonard et al. 2007 AMNH F:AM 30451 Canis lupus Upper Cleary Creek cranium 15,800 90 19,322 18,846 19,084 UCR-3761 Leonard et al. 2007 8.8 -18.6 - - - 2.4 Leonard et al. 2007 AMNH F:AM 67227 Canis lupus Cripple Creek mandible (bone) 15,870 190 19,614 18,760 19,187 AA-35231 Leonard et al. 2007 7.9 -18.9 - - - 2.3 Leonard et al. 2007 AMNH F:AM 7738 Cervus elaphus Lower Goldstream cranium (bone) 8,650 90 10,112 9,478 9,795 AA-25960 Guthrie 2006 5.7 -20.7 3.4 31.4 10.9 2.9 this paper AMNH F:AM 2160 Cervus elaphus Chicken Creek not found in museum, probably antler 9,793 99 11,605 10,789 11,197 AA-14315 Guthrie 2006 3.2 -19.8 22.5 36.9 13.4 2.8 this paper AMNH F:AM 6330 Cervus elaphus Ester Creek metatarsal 10,292 111 12,530 11,626 12,078 AA-25975 Guthrie 2006 2.9 -19.7 7.8 43.4 15.4 2.8 this paper AMNH F:AM 3380 Cervus elaphus Forty Mile Mining antler (adult) 10,533 83 12,688 12,146 12,417 AA-14304 Guthrie 2006 1.7 -19.8 19.0 44.7 16.2 2.8 this paper UAF KD-14-C-02 Cervus elaphus Keystone Dune Site (XBD-363) phalanx 11,470 40 13,425 13,219 13,322 UGAMS-18143 Reuther et al. 2016 1.9 -19.4 - - - - this paper AMNH F:AM 1572 Cervus elaphus Fairbanks Area not found in museum, probably antler 11,664 96 13,714 13,290 13,502 AA-14302 Guthrie 2006 2.5 -19.3 15.8 47.5 17.2 2.8 this paper AMNH F:AM 34669 Cervus elaphus Cleary Creek mandible (bone) 11,837 83 13,823 13,462 13,643 AA-25964 Guthrie 2006 2.6 -19.7 11.0 43.6 15.7 2.8 this paper AMNH F:AM 6233 Cervus elaphus Lower Goldstream metacarpal 11,980 152 14,215 13,467 13,841 AA-25990 Guthrie 2006 2.7 -19.5 15.5 44.6 15.4 2.9 this paper AMNH F:AM 1547 Cervus elaphus Fairbanks Creek metacarpal 12,110 133 14,500 13,589 14,045 AA-25987 Guthrie 2006 3.0 -20.0 9.6 43.7 14.9 2.9 this paper AMNH F:AM 9190 Cervus elaphus Ester Creek metatarsal 12,201 135 14,753 13,751 14,252 AA-25976 Guthrie 2006 3.4 -19.8 10.5 37.5 13.3 2.8 this paper AMNH F:AM 34674C Cervus elaphus Lower Goldstream metacarpal 12,237 208 15,095 13,720 14,408 AA-25994 Guthrie 2006 2.9 -20.0 5.6 41.1 14.2 2.9 this paper AMNH F:AM 34672D Cervus elaphus Upper Cleary Creek metacarpal 12,321 138 15,038 13,932 14,485 AA-25992 Guthrie 2006 2.6 -19.8 7.9 44.4 15.1 2.9 this paper AMNH F:AM 34672C Cervus elaphus Fox metacarpal 12,480 139 15,159 14,131 14,645 AA-25988 Guthrie 2006 3.2 -19.7 11.7 41.9 14.4 2.9 this paper AMNH F:AM 6637 Cervus elaphus Lower Goldstream cranium (bone) 12,562 94 15,191 14,306 14,749 AA-14297 Guthrie 2006 3.7 -19.7 3.2 33.3 11.2 3.0 this paper AMNH F:AM 6606 Cervus elaphus Lower Goldstream mandible (bone) 12,640 140 15,434 14,257 14,846 AA-25957 Guthrie 2006 4.6 -19.7 6.0 32.3 11.4 2.8 this paper AMNH F:AM 4008 Cervus elaphus Lower Goldstream not found in museum 12,657 89 15,339 14,578 14,959 AA-14323 Guthrie 2006 3.2 -19.3 11.4 43.7 15.8 2.8 this paper SHPO 2195 Cervus elaphus Swan Point CZ3 (XBD-146) radius 10,100 11,700 dated by strat. Association Lanoë and Holmes 2016 3.2 -19.2 3.7 41.3 15.0 2.7 this paper

208

SHPO 5357 Cervus elaphus Swan Point CZ3 (XBD-146) radius 10,100 11,700 dated by strat. Association Lanoë and Holmes 2016 2.0 -19.3 13.7 41.0 15.0 2.7 this paper SHPO 15791 Equus lambei Swan Point CZ4b (XBD-146) cheektooth 11,950 100 14,057 13,561 13,809 AA-74249 Lanoë and Holmes 2016 1.3 -20.3 7.3 42.1 13.8 3.0 this paper AMNH F:AM 142429 Equus lambei Goldstream 12,310 45 14,601 14,071 14,336 CAMS-119982 Leonard et al. 2007 4.3 -21.3 - - - 2.4 Leonard et al. 2007 AMNH F:AM 144736 A-46-34 Equus lambei Upper Cleary Creek metacarpal 12,482 80 15,076 14,139 14,608 AA-26830 Guthrie 2006 4.5 -20.7 12.0 44.8 15.9 2.8 this paper AMNH F:AM 144737 A-144-9422 Equus lambei Ester Creek metacarpal 12,510 130 15,177 14,171 14,674 AA-26819 Guthrie 2006 4.0 -20.4 7.8 43.8 15.6 2.8 this paper AMNH F:AM 144738 A-144-9399 Equus lambei Ester Creek metacarpal 12,580 140 15,297 14,210 14,754 AA-26828 Guthrie 2006 4.1 -20.7 7.0 41.8 14.9 2.8 this paper AMNH F:AM 142423 Equus lambei Goldstream 12,560 50 15,143 14,530 14,837 CAMS-119976 Leonard et al. 2007 3.9 -20.9 - - - 2.4 Leonard et al. 2007 AMNH F:AM 144739 A-216-6846 Equus lambei Goldstream metacarpal 12,710 170 15,675 14,305 14,990 AA-26822 Guthrie 2006 4.3 -21.1 8.6 43.3 15.4 2.8 this paper AMNH F:AM 144740 A 557-4339 Equus lambei Fairbanks Creek metacarpal 12,860 140 15,841 14,895 15,368 AA-26829 Guthrie 2006 3.8 -21.0 8.5 43.6 15.6 2.8 this paper AMNH F:AM 60025 Equus lambei Fairbanks Area 13,710 60 16,052 15,610 15,831 CAMS-120061 Leonard et al. 2007 4.7 -21.2 - - - 2.3 Leonard et al. 2007 AMNH F:AM 144741 A-276 Equus lambei Cleary Creek metacarpal 13,940 160 17,407 16,398 16,903 AA-26809 Guthrie 2006 3.8 -20.5 10.3 44.2 15.9 2.8 this paper AMNH F:AM 144742 A-558-2402 Equus lambei Cripple Creek metacarpal 14,260 160 17,834 16,908 17,371 AA-26840 Guthrie 2006 4.1 -20.2 15.2 43.7 15.7 2.8 this paper AMNH F:AM 60005 Equus lambei Bear Creek 14,630 60 17,996 17,622 17,809 CAMS-119969 Leonard et al. 2007 1.6 -21.3 - - - 2.3 Leonard et al. 2007 AMNH F:AM 142421 Equus lambei Fairbanks Area 14,860 60 18,271 17,887 18,079 CAMS-119974 Leonard et al. 2007 4.8 -21.0 - - - 2.4 Leonard et al. 2007 UAF UA-97-061-221 Equus lambei Gerstle River (XMH-246) radius 15,090 80 18,568 18,092 18,330 Beta-109267 Holmes 1998 2.2 -20.3 11.8 44.4 15.8 2.8 this paper AMNH F:AM 142424 Equus lambei Fairbanks Area 15,460 70 18,872 18,572 18,722 CAMS-119977 Leonard et al. 2007 2.9 -20.7 - - - 2.4 Leonard et al. 2007 AMNH F:AM 60032 Equus lambei Fairbanks Area 15,850 70 19,339 18,913 19,126 CAMS-120068 Leonard et al. 2007 2.5 -20.9 - - - 2.4 Leonard et al. 2007 SHPO 22193 Equus lambei Swan Point CZ4b (XBD-146) hyoid 12,000 14,000 dated by strat. Association Lanoë and Holmes 2016 2.2 -21.0 16.4 42.6 14.6 2.9 this paper SHPO 12708 Equus lambei Swan Point CZ4b (XBD-146) unknown (identified by collagen fingerprinting) 12,000 14,000 dated by strat. Association Lanoë and Holmes 2016 1.5 -21.7 11.0 43.7 14.1 3.1 this paper CMN CMN 36259 Mammuthus primigenius Dawson Area 11,860 120 13,998 13,454 13,726 AA-17559 Guthrie 2006 5.5 -20.4 9.0 44.8 16.3 2.8 this paper SHPO 15917 Mammuthus primigenius Swan Point CZ4b (XBD-146) ivory 12,050 120 14,206 13,580 13,893 AA-74251 Lanoë and Holmes 2016 7.4 -21.3 18.3 44.8 15.8 2.8 this paper AMNH F:AM 6675 Mammuthus primigenius Goldstream molar M6 (root) 12,123 88 14,213 13,756 13,985 AA-14940 Guthrie 2006 6.3 -20.6 6.8 45.5 15.4 3.0 this paper SHPO 19199 Mammuthus primigenius Swan Point CZ4b (XBD-146) ivory 12,110 120 14,441 13,626 14,034 AA-74250 Lanoë and Holmes 2016 6.9 -20.3 5.0 37.6 13.3 2.8 this paper AMNH F:AM 11 Mammuthus primigenius Cleary Creek molar M6 (root) 12,337 108 14,943 14,014 14,479 AA-14938 Guthrie 2006 6.9 -21.4 6.4 44.4 15.3 2.9 this paper AMNH F:AM 1375 Mammuthus primigenius Engineer Creek not found in museum 12,429 178 15,230 14,020 14,625 AA-14860 Guthrie 2006 7.2 -20.7 16.2 44.5 16.0 2.8 this paper AMNH F:AM 470 Mammuthus primigenius Cleary Creek not found in museum 12,476 81 15,070 14,228 14,649 AA-14916 Guthrie 2006 6.7 -21.0 8.8 44.2 15.2 2.9 this paper SHPO 21444 Mammuthus primigenius Swan Point CZ4b (XBD-146) ivory 12,500 150 15,208 14,129 14,669 AA-98488 Lanoë and Holmes 2016 8.0 -20.9 6.0 48.1 17.0 2.8 this paper AMNH F:AM 6677 Mammuthus primigenius Goldstream not found in museum 12,576 147 15,306 14,186 14,746 AA-14880 Guthrie 2006 8.1 -21.2 12.2 44.5 15.3 2.9 this paper AMNH F:AM 1208 Mammuthus primigenius Sullivan Creek not found in museum 12,677 142 15,556 14,331 14,944 AA-14888 Guthrie 2006 5.9 -20.8 - - - 3.3 Szpak et al. 2010 AMNH F:AM 10412 Mammuthus primigenius Goldstream tooth 12,884 124 15,817 15,044 15,431 AA-14909 Guthrie 2006 5.9 -21.2 5.2 39.9 14.4 2.8 this paper AMNH F:AM 1287 Mammuthus primigenius Forty Mile mandible (bone) (adult or subadult) 13,226 85 16,176 15,630 15,903 AA-14925 Guthrie 2006 5.8 -20.6 12.9 44.5 15.2 2.9 this paper AMNH F:AM 4953 Mammuthus primigenius Goldstream not found in museum 13,339 150 16,516 15,613 16,065 AA-14867 Guthrie 2006 7.2 -20.4 13.5 41.0 14.8 2.8 this paper AMNH F:AM 1021 Mammuthus primigenius Ester Creek not found in museum 13,410 152 16,617 15,709 16,163 AA-14883 Guthrie 2006 6.2 -20.6 15.4 43.6 15.0 2.9 this paper AMNH F:AM 24 Mammuthus primigenius Cleary Creek molar M6 (root) 13,436 87 16,464 15,890 16,177 AA-14934 Guthrie 2006 7.0 -21.5 9.7 44.9 15.4 2.9 this paper AMNH F:AM 500 Mammuthus primigenius Fairbanks Creek not found in museum 14,023 98 17,389 16,670 17,030 AA-14895 Guthrie 2006 6.7 -20.7 8.0 41.3 14.8 2.8 this paper AMNH F:AM 508 Mammuthus primigenius Cleary Creek not found in museum 14,390 92 17,840 17,246 17,543 AA-14923 Guthrie 2006 8.0 -21.3 10.0 37.5 13.0 2.9 this paper AMNH F:AM 527 Mammuthus primigenius Banner Creek not found in museum 14,679 174 18,313 17,461 17,887 AA-14882 Guthrie 2006 7.0 -20.4 7.1 38.3 13.4 2.9 this paper AMNH F:AM 503 Mammuthus primigenius Cleary Creek not found in museum 15,102 135 18,660 18,004 18,332 AA-14912 Guthrie 2006 7.8 -20.7 13.1 41.4 14.9 2.8 this paper YHR 276.1 Ovibos moschatus Klondike 9,280 60 10,650 10,265 10,458 AA-11743 Raghavan et al. 2010, 2014 3.2 -18.8 - 44.8 16.5 3.2 Raghavan et al. 2010, 2014 NPSF 20150929.01 Ovis dalli Cyclops Cave calcaneus 12,900 60 15,658 15,190 15,424 OS-122686 Richie, pers. com. 2016 3.1 -20.7 - 43.0 14.8 2.9 this paper UAF UAF-EVA-15-02 Ovis dalli Gerstle River (XMH-246) metatarsal 13,131 80 16,042 15,434 15,738 AA-107343 this paper 3.2 -19.1 14.3 45.9 15.9 2.9 this paper UAF UA-97-061-254 Ovis dalli Gerstle River (XMH-246) innominate 13,451 80 16,472 15,927 16,200 AA-107344 this paper 2.9 -19.7 7.7 44.1 15.4 2.9 this paper UAF UA-77-055-0085 Ovis dalli Gerstle River (XMH-246) radio-ulna 14,946 100 18,428 17,913 18,171 AA-106802 this paper 3.4 -18.9 6.6 43.2 14.9 2.9 this paper AMNH - A-780-1435 Panthera leo Fairbanks Creek humerus 11,925 70 13,975 13,568 13,772 OxA-10080 Barnett et al. 2009 7.0 -17.8 - - - - Barnett et al. 2009 AMNH F:AM 69167 Panthera leo Ester Creek tibia 12,090 80 14,143 13,756 13,950 OxA-13451 Barnett et al. 2009 6.7 -18.6 - - - - Barnett et al. 2009 AMNH F:AM 69094 Panthera leo Lower Goldstream humerus 12,540 75 15,138 14,331 14,735 OxA-10081 Barnett et al. 2009 7.5 -18.3 - - - - Barnett et al. 2009 CMN CMN 47294 Panthera leo Gold Run Creek, Yukon humerus 12,640 75 15,290 14,663 14,977 OxA-10083 Barnett et al. 2009 5.9 -17.6 - - - - Barnett et al. 2009 AMNH F:AM 69080 Panthera leo Fairbanks Area metatarsal 12,990 70 15,791 15,276 15,534 CAMS-131350 Leonard et al. 2007 6.7 -18.8 - - - 2.4 Leonard et al. 2007 AMNH F:AM 69053 Panthera leo Fairbanks Area metacarpal 13,040 70 15,869 15,319 15,594 CAMS-131348 Leonard et al. 2007 6.7 -18.8 - - - 2.3 Leonard et al. 2007 AMNH F:AM 69173 Panthera leo Fairbanks Area phalanx 14,050 80 17,391 16,768 17,080 CAMS-131347 Leonard et al. 2007 7.5 -19.0 - - - 2.3 Leonard et al. 2007 CMN CMN 35609 Panthera leo Hunker Creek, Dawson femur 15,550 90 18,986 18,612 18,799 OxA-10086 Barnett et al. 2009 8.7 -18.1 - - - - Barnett et al. 2009 AMNH F:AM 30757 Panthera leo Banner Creek mandible (bone) 15,978 65 19,509 19,056 19,283 OxA-13832 Barnett et al. 2009 8.3 -18.0 - - - - Barnett et al. 2009 AMNH F:AM 69139 Panthera leo Fairbanks Creek ulna 16,005 65 19,538 19,098 19,318 OxA-13834 Barnett et al. 2009 7.2 -17.9 - - - - Barnett et al. 2009 UAF UA-2000-054-808 Rangifer tarandus Gerstle River (XMH-246) femur 9,390 60 10,768 10,427 10,598 AA-106447 this paper 2.9 -20.6 8.5 45.8 15.9 2.9 this paper YC - Rangifer tarandus Little John CZ7/CZ6 (KdVo-6) - 9,530 40 11,084 10,694 10,889 Beta-217279 Easton, pers. com. 2016 -1.0 -19.8 - - - - Easton, pers. com. 2016 AMNH F:AM 142443 Rangifer tarandus Lower Goldstream mandible (bone) 16,000 190 19,791 18,870 19,331 AA-48686 Leonard et al. 2007 3.3 -19.2 - - - 2.4 Leonard et al. 2007 SHPO 18038 Rangifer tarandus Swan Point CZ4b (XBD-146) metatarsal (juvenile) 12,000 14,000 dated by strat. Association Lanoë and Holmes 2016 -2.5 -19.7 11.6 42.2 15.1 2.8 this paper SHPO 22303 Rangifer tarandus Swan Point CZ4b (XBD-146) antler (adult) 12,000 14,000 dated by strat. Association Lanoë and Holmes 2016 -2.7 -18.5 7.9 41.3 14.8 2.8 this paper UAF UAM:ES:4584 UA-V-54-970 Saiga tatarica Lost Chicken Creek cranium 12,390 120 15,043 14,076 14,560 AA-3075 Guthrie et al. 2001 4.4 -19.5 10.2 40.7 14.1 2.9 this paper UAF UAM:ES:4586 UA-V-54-196 Saiga tatarica Lost Chicken Creek cranium 12,620 110 15,306 14,338 14,822 AA-3074 Guthrie et al. 2001 4.8 -18.6 15.0 41.3 14.5 2.9 this paper UAF UA-97-061-222 Saiga tatarica Gerstle River (XMH-246) humerus 12,700 80 15,361 14,746 15,054 AA-106446 this paper 3.7 -19.1 3.7 41.7 14.2 2.9 this paper AMNH F:AM 95653 Ursus arctos Engineer Creek ulna 9,535 75 11,145 10,600 10,873 OxA-9800 Barnes et al. 2002 8.0 -17.8 - - - - Barnes et al. 2002 AMNH F:AM 95612 A-199-8654 Ursus arctos Ester Creek tooth 10,015 62 11,766 11,266 11,516 AA-17506 Barnes et al. 2002 11.5 -17.5 Matheus 1995 AMNH F:AM 95641 A-203-8241 Ursus arctos Engineer Creek tibia 11,940 100 14,053 13,556 13,805 OxA-9798 Barnes et al. 2002 4.8 -18.4 - - - - Barnes et al. 2002 AMNH F:AM 95628 A-210-6359 Ursus arctos Lower Goldstream humerus 12,310 65 14,691 14,051 14,371 OxA-9828 Barnes et al. 2002 7.5 -18.5 - - - - Barnes et al. 2002 AMNH F:AM 95598 A-197-2966 Ursus arctos Cripple Creek mandible (bone) 12,320 90 14,825 14,012 14,419 OxA-9799 Barnes et al. 2002 7.1 -18.2 - - - - Barnes et al. 2002 AMNH F:AM 95595 Ursus arctos Goldstream Creek mandible (bone) 12,441 101 15,055 14,156 14,606 AA-17508 Barnes et al. 2002 6.7 -18.9 Matheus 1995 AMNH F:AM 95659 #1602 Ursus arctos Goldstream ulna 13,415 70 16,366 15,893 16,130 OxA-9709 Barnes et al. 2002 8.0 -18.5 - - - - Barnes et al. 2002 AMNH F:AM 95665 Ursus arctos Gold Hill radius 13,760 50 16,890 16,382 16,636 UCR-3743/CAMS-54128 Leonard et al. 2000 8.1 -19.5 13.1 44.2 15.6 2.8 this paper AMNH F:AM 95642 A-425 Ursus arctos Ester Creek femur 14,150 90 17,515 16,936 17,226 OxA-9262 Barnes et al. 2002 8.3 -18.5 - - - - Barnes et al. 2002 AMNH F:AM 95599 A-255-6945 Ursus arctos Goldstream mandible (bone) 14,310 100 17,726 17,109 17,418 OxA-9801 Barnes et al. 2002 6.5 -19.1 - - - - Barnes et al. 2002 AMNH F:AM 95632 A-340-1029 Ursus arctos Rosie Creek femur 14,810 80 18,267 17,807 18,037 OxA-9830 Barnes et al. 2002 8.4 -18.6 - - - - Barnes et al. 2002 AMNH F:AM 95630 Ursus arctos Lower Goldstream humerus 14,980 60 18,386 17,991 18,189 UCR-3741/CAMS-51805 Leonard et al. 2000 7.0 -19.3 18.9 43.2 15.5 2.8 this paper AMNH F:AM 95679 Ursus arctos Lower Goldstream ulna 15,370 60 18,786 18,501 18,644 UCR-3742/CAMS-51806 Leonard et al. 2000 7.5 -19.2 7.6 43.1 15.4 2.8 this paper AMNH F:AM 95670 A-658-1545 Ursus arctos Gold Hill ulna 15,830 100 19,398 18,865 19,132 OxA-9263 Barnes et al. 2002 5.4 -19.4 - - - - Barnes et al. 2002

209

References

Altamira Consulting Ltd.

2014 Western Copper and Gold Corporation Casino Project Archaeological Resources

Mitigation. Report On File at the Yukon Cultural Services Branch, Whitehorse,

Canada.

Ambrose, Stanley H.

1990 Preparation and Characterization of Bone and Tooth Collagen for Isotopic

Analysis. Journal of Archaeological Science 17: 431–451.

Anderson, Patricia M., Mary E. Edwards, and Linda B. Brubaker

2004 Results and paleoclimate implications of 35 years of paleoecological research in

Alaska. Developments in Quaternary Science 1: 427–440.

Asner, Gregory P., Shaun R. Levick, Ty Kennedy-Bowdoin, David E. Knapp, Ruth Emerson,

James Jacobson, Matthew S. Colgan, and Roberta E. Martin

2009 Large-scale impacts of herbivores on the structural diversity of African savannas.

Proceedings of the National Academy of Sciences 106(12): 4947–4952.

Bakker, Elisabeth S., Jacquelyn L. Gill, Christopher N. Johnson, Frans W. M. Vera, Christopher

J. Sandom, Gregory P. Asner, and Jens-Christian Svenning

2015 Combining paleo-data and modern exclosure experiments to assess the impact of

megafauna extinctions on woody vegetation. Proceedings of the National

Academy of Sciences 113(4): 845–855.

Barnes, I., P. Matheus, Beth Shapiro, D. Jensen, and A. Cooper

2002 Dynamics of Pleistocene population extinctions in Beringian brown bears.

Science 295: 2267–70. 210

Barnett, Ross, Beth Shapiro, Ian Barnes, Simon Y. W. Ho, Joachim Burger, Nobuyuki

Yamaguchi, Thomas F. G. Higham, H. Todd Wheeler, Wilfried Rosendahl, Andrei V. Sher,

Marina Sotnikova, Tatiana Kuznetsova, Gennady F. Baryshnikov, Larry D. Martin, C. Richard

Harington, James A. Burns, and Alan Cooper

2009 Phylogeography of lions (Panthera leo ssp.) reveals three distinct taxa and a late

Pleistocene reduction in genetic diversity. Molecular Ecology 18(8): 1668–77.

Ben-David, Merav, Einav Shochat, and Layne G. Adams

2001 Utility of Stable Isotope Analysis in Studying Foraging Ecology of Herbivores:

Examples from Moose and Caribou. Alces 37(2): 421–434.

Bigelow, Nancy H, and W. Roger Powers

2001 Climate, Vegetation, and Archaeology 14,000-9,000 Cal Yr B.P. in Central

Alaska. Arctic Anthropology 38(2): 171–195.

Blinnikov, Mikhail S., Benjamin V. Gaglioti, Donald A. Walker, Matthew J. Wooller, and Grant

D. Zazula

2011 Pleistocene graminoid-dominated ecosystems in the Arctic. Quaternary Science

Reviews 30(21–22): 2906–2929.

Bocherens, Hervé

2003 Isotopic biogeochemistry and the paleoecology of the mammoth steppe fauna.

Deinsea 9: 57–76.

2015 Isotopic tracking of large carnivore palaeoecology in the mammoth steppe.

Quaternary Science Reviews 117: 42–71.

Bocherens, Hervé, Gilles Pacaud, Pyotr A. Lazarev, and Andr Mariotti

211

1996 Stable isotope abundances (13C, 15N) in collagen and soft tissues from

Pleistocene mammals from Yakutia: Implications for the palaeobiology of the

mammoth steppe. Palaeogeography, Palaeoclimatology, Palaeoecology 126: 31–

44.

Bond, William J.

2008 What Limits Trees in C4 Grasslands and Savannas? Annu. Rev. Ecol. Evol. Syst

39(2008): 641–59.

Bond, William J., and Jon E. Keeley

2005 Fire as a global “herbivore”: The ecology and evolution of flammable ecosystems.

Trends in Ecology and Evolution 20(7): 387–394.

Bradshaw, C.J.A., A. Cooper, C.S.M. Turney, and B.W. Brook

2012 Robust estimates of extinction time in the geological record. Quaternary Science

Reviews 33: 14–19.

Brantingham, P. Jeffrey

1998a Mobility, Competition, and Plio-Pleistocene Hominid Foraging Groups. Journal

of Archaeological Method and Theory 5(1): 57–98.

1998b Hominid–Carnivore Coevolution and Invasion of the Predatory Guild. Journal of

Anthropological Archaeology 17: 327–353.

Bronk Ramsey, Christopher

1994 Analysis of Chronological Information and Radiocarbon Calibration : The

Program OxCal. Archaeological Computing Newsletter 41: 11–16.

2009 Bayesian Analysis of Radiocarbon Dates. Radiocarbon 51(1): 337–360.

212

Brose, Ulrich, M. Pavao-Zuckerman, A. Eklof, J. Bengtsson, M. P. Berg, S. H. Cousins, C.

Mulder, H. A. Verhoef, and V. Wolters

2005 Spatial Aspects of Food Webs. In Dynamic Food Webs: Multispecies

Assemblages, Ecosystem Development, and Environmental Change, edited by

Peter C. de Ruiter, V. Wolters, and J. Moore, pp. 463–469. Springer, New York.

Bryant, H. N.

1987 Wolverine from the Pleistocene of the Yukon: evolutionary trends and

of Gulo (Carnivora: Mustelidae). Canadian Journal of Earth Sciences 24: 654–

663.

Buck, Caitlin E., and Edouard Bard

2007 A calendar chronology for Pleistocene mammoth and horse extinction in North

America based on Bayesian radiocarbon calibration. Quaternary Science Reviews

26(17–18): 2031–2035.

Carbone, Chris, G. M. Mace, S. C. Roberts, and David W. Macdonald

1999 Energetic constraints on the diet of terrestrial carnivores. Nature 402: 286–288.

Chambers, Steven M., Steven R. Fain, Bud Fazio, and Michael Amaral

2012 An account of the taxonomy of North American wolves from morphological and

genetic analyses. North American Fauna 77: 1–67.

Choy, Kyungcheol, Ben A. Potter, Holly J. McKinney, Joshua D. Reuther, Shiway W. Wang, and Matthew J. Wooller

2016 Chemical profiling of ancient hearths reveals recurrent salmon use in Ice Age

Beringia. Proceedings of the National Academy of Sciences 113(35): 9757–9762.

213

Cozzi, Gabriele, Femke Broekhuis, John W. Mcnutt, Lindsay A. Turnbull, David W. Macdonald, and Bernhard Schmid

2012 Fear of the dark or dinner by moonlight? Reduced temporal partitioning among

Africa’s large carnivores. Ecology 93(12): 2590–2599.

Croitor, Roman, and Jean Philippe Brugal

2010 Ecological and evolutionary dynamics of the carnivore community in Europe

during the last 3 million years. Quaternary International 212: 98–108.

Cromsigt, J. P. G. M., and Han Olff

2006 Resource partitioning among savanna grazers mediated by local heterogeneity:

An experimental approach. Ecology 87(6): 1532–1541.

Dalerum, F., A Angerbjörn, K. Kunkel, and B. S. Shults

2009 Patterns of d13C and d15N in wolverine Gulo gulo tissues from the Brooks

Range, Alaska. Current Zoology 55(3): 188–192.

Debruyne, Regis, Genevieve Chu, Christine E. King, Kirsti Bos, Melanie Kuch, Carsten

Schwarz, Paul Szpak, Darren R. Gröcke, Paul Matheus, Grant D. Zazula, R. Dale Guthrie, Duane

Froese, Bernard Buigues, Christian de Marliave, Clare Flemming, Debi Poinar, Daniel Fisher,

John Southon, Alexei N. Tikhonov, Ross D. E. MacPhee, and Hendrik N. Poinar

2008 Out of America: Ancient DNA Evidence for a New World Origin of Late

Quaternary Woolly Mammoths. Current Biology 18: 1320–1326.

Discamps, Emmanuel, Jacques Jaubert, and François Bachellerie

2011 Human choices and environmental constraints: deciphering the variability of large

game procurement from Mousterian to Aurignacian times (MIS 5-3) in

southwestern France. Quaternary Science Reviews 30(19–20): 2755–2775.

214

Donadio, Emiliano, and Steven W. Buskirk

2006 Diet, morphology, and interspecific killing in carnivora. The American Naturalist

167(4): 524–536.

Doughty, Christopher E., Joe Roman, Søren Faurby, Adam Wolf, Alifa Haque, Elisabeth S.

Bakker, Yadvinder Malhi, John B Dunning, and Jens-Christian Svenning

2015 Global nutrient transport in a world of giants. Proceedings Of The National

Academy Of Sciences 113(4): 868–873.

Drucker, Dorothée G., Hervé Bocherens, Anne Pike-Tay, and André Mariotti

2001 Isotopic tracking of seasonal dietary change in dentine collagen: preliminary data

from modern caribou. Comptes Rendus de l’Académie des Sciences 333(5): 303–

309.

Drucker, Dorothée G., A. Bridault, C. Cupillard, A. Hujic, and Hervé Bocherens

2011 Evolution of habitat and environment of red deer (Cervus elaphus) during the

Late-glacial and early Holocene in eastern France (French Jura and the western

Alps) using multi-isotope analysis (d13C, d15N, d18O, d34S) of archaeological

remains. Quaternary International 245(2): 268–278.

Drucker, Dorothée G., Anne Bridault, Keith A. Hobson, Elwira Szuma, and Hervé Bocherens

2008 Can carbon-13 in large herbivores reflect the canopy effect in temperate and

boreal ecosystems? Evidence from modern and ancient ungulates.

Palaeogeography, Palaeoclimatology, Palaeoecology 266: 69–82.

Drucker, Dorothée G., Keith A. Hobson, Jean-Pierre Ouellet, and Rehaume Courtois

215

2010 Influence of forage preferences and habitat use on 13C and 15N abundance in

wild caribou (Rangifer tarandus caribou) and moose (Alces alces) from Canada.

Isotopes in Environmental and Health Studies 46(1): 107–121.

Dyke, A. S., A. Moore, and L. Robertson

2003 Thirty-two digital maps at 1:7,000,000 scale with accompanying digital

chronological database and one poster (two sheets) with full map series.

Geological Survey of Canada 1574.

Easton, Norman A., Glen R. Mackay, Patricia B. Young, Peter Schnurr, and David R. Yesner

2011 Chindadn in Canada? Emergent Evidence of the Pleistocene Transition in

Southeast Beringia as Revealed by the Little John Site, Yukon. In From the

Yenisei to the Yukon: Interpreting Lithic Assemblage Variability in Late

Pleistocene/Early Holocene Beringia, edited by Ted Goebel and Ian Buvit, pp.

289–307. Texas A&M University Press, College Station.

Edwards, M. E., C. J. Mock, B. P. Finney, V. A. Barber, and P. J. Bartlein

2001 Potential analogues for paleoclimatic variations in eastern interior Alaska during

the past 14,000 yr: atmospheric-circulation controls of regional temperature and

moisture responses. Quaternary Science Reviews 20(1–3): 189–202.

Endacott, Neal Andrew

2008 The Zooarchaeology of Lime Hills Cave: Paleoecological and Taphonomic

Insights. PhD Dissertation, Washington State University.

Flaherty, Elizabeth A., and Merav Ben-David

2010 Overlap and partitioning of the ecological and isotopic niches. Oikos 119(9):

1409–1416.

216

Folke, Carl

2006 Resilience: The emergence of a perspective for social–ecological systems

analyses. Global Environmental Change 16(3): 253–267.

Fox-Dobbs, Kena, J. K. Bump, R. O. Peterson, D. L. Fox, and P. L. Koch

2007 Carnivore-specific stable isotope variables and variation in the foraging ecology

of modern and ancient wolf populations: case studies from Isle Royale,

Minnesota, and La Brea. Canadian Journal of Zoology 85(4): 458–471.

Fox-Dobbs, Kena, Jennifer A. Leonard, and Paul L. Koch

2008 Pleistocene megafauna from eastern Beringia: Paleoecological and

paleoenvironmental interpretations of stable carbon and nitrogen isotope and

radiocarbon records. Palaeogeography, Palaeoclimatology, Palaeoecology

261(1–2): 30–46.

Fox-Dobbs, Kena, Abigail A. Nelson, Paul L. Koch, and Jennifer A Leonard

2012 Faunal isotope records reveal trophic and nutrient dynamics in twentieth century

Yellowstone grasslands. Biology Letters 8(5): 838–41.

Fritz, Hervé, Michel Loreau, Simon Chamaillé-Jammes, Marion Valeix, and Jean Clobert

2011 A food web perspective on large herbivore community limitation. Ecography

34(2): 196–202.

Geomatics Yukon

2015 GeoYukon. Government of Yukon: http://mapservices.gov.yk.ca/GeoYukon/.

Gill, Jacquelyn L., J. W. Williams, S. T. Jackson, K. B. Lininger, and G. S. Robinson

2009 Pleistocene Megafaunal Collapse, Novel Plant Communities, and Enhanced Fire

Regimes in North America. Science 326: 1100–1103.

217

Gill, Jacquelyn L., John W. Williams, Stephen T. Jackson, Jeffrey P. Donnelly, and Grace C.

Schellinger

2012 Climatic and megaherbivory controls on late-glacial vegetation dynamics: A new,

high-resolution, multi-proxy record from Silver Lake, Ohio. Quaternary Science

Reviews 34: 66–80.

Goetcheus, Victoria G., and Hilary H. Birks

2001 Full-glacial upland tundra vegetation preserved under tephra in the Beringia

National Park, Seward Peninsula, Alaska. Quaternary Science Reviews 20(1–3):

135–147.

Graham, Russell W., Soumaya Belmecheri, Kyungcheol Choy, Brendan J. Culleton, Lauren J.

Davies, Duane Froese, Peter D. Heintzman, Carrie Hritz, Joshua D. Kapp, Lee A. Newsom, Ruth

Rawcliffe, Émilie Saulnier-Talbot, Beth Shapiro, Yue Wang, John W. Williams, and Matthew J.

Wooller

2016 Timing and causes of mid-Holocene mammoth extinction on St. Paul Island,

Alaska. Proceedings of the National Academy of Sciences 113(33): 201604903.

Gunderson, Lance H.

2000 Ecological Resilience - In Theory and Application. Annual Review of Ecology,

Evolution, and Systematics 31: 425–439.

Guthrie, R. Dale

1980 Bison and man in North America. Canadian Journal of Anthropology 1(1): 55–

73.

1984 Mosaics, Allelochemics, and Nutrients: An Ecological Theory of Late Pleistocene

Megafaunal Extinctions. In Quaternary Extinctions: A Prehistoric Revolution,

218

edited by Paul S. Martin and Richard G. Klein, pp. 259–298. University of

Arizona Press, Tucson.

1990 Frozen Fauna of the Mammoth Steppe. University of Chicago Press.

2001 Origin and causes of the mammoth steppe: a story of cloud cover, woolly

mammal tooth pits, buckles, and inside-out Beringia. Quaternary Science Reviews

20: 549–574.

2003 Rapid body size decline in Alaskan Pleistocene horses before extinction. Nature

426: 169–171.

2006 New carbon dates link climatic change with human colonization and Pleistocene

extinctions. Nature 441: 207–9.

Guthrie, R. Dale, A. V. Sher, and C. R. Harington

2001 New Radiocarbon Dates on Saiga Antelopes (Saiga tatarica) from Alaska,

Canada, and Siberia: their Paleoecological Significance. In People and Wildlife in

Northern North America, edited by S. C. Gerlach and M. S. Murray, pp. 50–57.

Archhaeopress: BAR International Series 944, Oxford.

Haile, James, Duane G. Froese, Ross D.E. MacPhee, Richard G. Roberts, Lee J. Arnold, Alberto

V. Reyes, Morten Rasmussen, Rasmus Nielsen, Barry W Brook, Simon Robinson, Martina

Demuro, M. Thomas P. Gilbert, Kasper Munch, Jeremy J. Austin, Alan Cooper, Ian Barnes, Per

Möller, and Eske Willerslev

2009 Ancient DNA reveals late survival of mammoth and horse in interior Alaska.

Proceedings of the National Academy of Sciences 106(52): 22352–7.

Halffman, Carrin M., Ben a. Potter, Holly J. McKinney, Bruce P. Finney, Antonia T. Rodrigues,

Dongya Y. Yang, and Brian M. Kemp

219

2015 Early human use of anadromous salmon in North America at 11,500 y ago.

Proceedings of the National Academy of Sciences 112(40): 12344–12348.

Harington, C.R.

2011 Pleistocene vertebrates of the Yukon Territory. Quaternary Science Reviews

30(17–18): 2341–2354.

2003 Annotated Bibliography of Quaternary Vertebrates of North America. University

of Toronto Press.

Haynes, Gary

2012 Elephants (and extinct relatives) as earth-movers and ecosystem engineers.

Geomorphology 157–158: 99–107.

Heintzman, Peter D., Duane Froese, John W. Ives, André E. R. Soares, Grant D. Zazula,

Brandon Letts, Thomas D. Andrews, Jonathan C. Driver, Elizabeth Hall, P. Gregory Hare,

Christopher N. Jass, Glen Mackay, John R. Southon, Mathias Stiller, Robin Woywitka, Marc A.

Suchard, and Beth Shapiro

2016 Bison phylogeography constrains dispersal and viability of the Ice Free Corridor

in western Canada. Proceedings of the National Academy of Sciences 113(29):

8057–8063.

Holmes, Charles E.

1996 Broken Mammoth. In American Beginnings: The Prehistory and Palaeoecology

of Beringia, edited by Frederick H West, pp. 312–319. University of Chicago

Press.

1998 Archaeological testing and evaluation of the Gerstle River Quarry, east-central

Alaska, 1996 (Report #65 on File at the Alaska Office of History and

220

Archaeology).

Holmes, Charles E., R. VanderHoek, and T. E .Dilley

1996 Swan Point. In American Beginnings: The Prehistory and Palaeoecology of

Beringia, edited by F H West, pp. 319–322. University of Chicago Press.

Hopcraft, J. G. C., Han Olff, and A. R. E. Sinclair

2010 Herbivores, resources and risks: alternating regulation along primary

environmental gradients in savannas. Trends in Ecology and Evolution 25(2):

119–128.

Hundertmark, Kris J., Gerald F. Shields, Irina G. Udina, R. Terry Bowyer, Alexei A. Danilkin, and Charles C. Schwartz

2002 Mitochondrial phylogeography of moose (Alces alces): late pleistocene

divergence and population expansion. Molecular phylogenetics and evolution

22(3): 375–387.

Iacumin, P., V. Nikolaev, and M. Ramigni

2000 C and N stable isotope measurements on Eurasian fossil mammals, 40000 to

10000 years BP: Herbivore physiologies and palaeoenvironmental reconstruction.

Palaeogeography, Palaeoclimatology, Palaeoecology 163(1–2): 33–47.

Ings, Thomas C., José M. Montoya, Jordi Bascompte, Nico Blüthgen, Lee Brown, Carsten F.

Dormann, François Edwards, David Figueroa, Ute Jacob, J. Iwan Jones, Rasmus B. Lauridsen,

Mark E. Ledger, Hannah M. Lewis, Jens M. Olesen, F. J. Frank van Veen, Phil H. Warren, and

Guy Woodward

2009 Ecological networks--beyond food webs. Journal of Animal Ecology 78(1): 253–

69.

221

Inman, Robert M., Audrey J. Magoun, Jens Persson, and Jenny Mattisson

2012 The wolverine’s niche: linking reproductive chronology, caching, competition,

and climate. Journal of Mammalogy 93(3): 634–644.

Jenkins, S. G., S. T. Partridge, T. R. Stephenson, S. D. Farley, and C. T. Robbins

2001 Nitrogen and carbon isotope fractionation between mothers, neonates, and nursing

offspring. Oecologia 129(3): 336–341.

Johnson, C. N.

2009 Ecological consequences of Late Quaternary extinctions of megafauna.

Proceedings of the Royal Society 276(1667): 2509–19.

Jones, Miriam C., and Zicheng Yu

2010 Rapid deglacial and early Holocene expansion of peatlands in Alaska.

Proceedings of the National Academy of Sciences 107(16): 7347–52.

Kartzinel, Tyler R., Patricia a. Chen, Tyler C. Coverdale, David L. Erickson, W. John Kress,

Maria L. Kuzmina, Daniel I. Rubenstein, Wei Wang, and Robert M. Pringle \

2015 DNA metabarcoding illuminates dietary niche partitioning by African large

herbivores. Proceedings of the National Academy of Sciences 112(26):

201503283.

Kirillova, Irina V., J. Argant, E. G. Lapteva, O. M. Korona, J. van der Plicht, E. V. Zinovyev, A.

A. Kotov, O. F. Chernova, Ye O. Fadeeva, O. A. Baturina, M. R. Kabilov, F. K. Shidlovskiy, and O. G. Zanina

2015 The diet and environment of mammoths in North-East Russia reconstructed from

the contents of their feces. Quaternary International 406(B): 147–161.

Kleynhans, Elizabeth J., Anna E. Jolles, Marjan R E Bos, and Han Olff

222

2011 Resource partitioning along multiple niche dimensions in differently sized

African savanna grazers. Oikos 120(4): 591–600.

Kurtén, Björn, and Elaine Anderson

1980 Pleistocene Mammals of North America. Columbia University Press, New York.

Lanoë, François Bertrand, and Charles E Holmes

2016 Animals as Raw Materials in Beringia: Insights from the Site of Swan Point

CZ4b, Alaska. American Antiquity 81(4): 682–696.

Leonard, Jennifer A, Carles Vilà, Kena Fox-Dobbs, Paul L. Koch, Robert K Wayne, and Blaire

Van Valkenburgh

2007 Megafaunal extinctions and the disappearance of a specialized wolf ecomorph.

Current Biology 17(13): 1146–50.

Leonard, Jennifer A., R. K. Wayne, and A. Cooper

2000 Population genetics of ice age brown bears. Proceedings of the National Academy

of Sciences 97(4): 1651–1654.

Leonard, Jennifer A, Robert K Wayne, and Jane Wheeler

2002 Ancient DNA evidence for Old World origin of New World dogs. Science 298:

1613–1616.

Llyod, A. H., M. E. Edwards, B. P. Finney, J. A. Lynch, V. Barber, and N. H. Bigelow

2006 Holocene Development of the Alaskan Boreal Forest. In Alaska’s Changing

Boreal Forest, edited by F.S. Chapin III, M. W. Oswood, K. Van Cleve, L. A.

Viereck, and D. L. Verbyla, pp. 62–80. Oxford University Press, New York.

MacDonald, G.M., D. W. Beilman, Y. V. Kuzmin, L. A. Orlova, K. V. Kremenetski, Beth

Shapiro, R. K. Wayne, and Blaire Van Valkenburgh

223

2012 Pattern of extinction of the woolly mammoth in Beringia. Nature

Communications 3: 893.

Manley, W. F.

2002 Postglacial Flooding of the Bering Land Bridge: A Geospatial Animation.

Institute of Arctic and Alpine Research, University of Colorado:

http://instaar.colorado.edu/QGISL/bering_land_bridge.

Manley, W. F., and Darrell S. Kaufman

2002 Alaska PaleoGlacier Atlas. Institute of Arctic and Alpine Research, University of

Colorado: http://instaar.colorado.edu/QGISL/ak_paleoglacier_atlas.

Mann, Daniel H., Pamela Groves, Michael L. Kunz, Richard E. Reanier, and Benjamin V.

Gaglioti

2013 Ice-age megafauna in Arctic Alaska: extinction, invasion, survival. Quaternary

Science Reviews 70: 91–108.

Mann, Daniel H., Richard E. Reanier, Dorothy M. Peteet, Michael L. Kunz, and Mark Johnson

2001 Environmental Change and Arctic Paleoindians. Arctic Anthropology 38(2): 119–

138.

Mann, Daniel H., Pamela Groves, Richard E. Reanier, and Michael L. Kunz

2010 Floodplains, permafrost, cottonwood trees, and peat: What happened the last time

climate warmed suddenly in arctic Alaska? Quaternary Science Reviews 29(27–

28): 3812–3830.

Mann, Daniel H., Pamela Groves, Richard E. Reanier, Benjamin V. Gaglioti, Michael L. Kunz, and Beth Shapiro

224

2015 Life and extinction of megafauna in the ice-age Arctic. Proceedings of the

National Academy of Sciences 112: 14301–14306.

Martinez-Meyer, Enrique, A. Townsend Peterson, and William W. Hargrove

2004 Ecological niches as stable distributional constraints on mammal species, with

implications for Pleistocene extinctions and climate change projections for

biodiversity. Global Ecology and Biogeography 13: 305–314.

Matheus, Paul E.

1995 Diet and Co-ecology of Pleistocene Short-Faced Bears and Brown Bears in

Eastern Beringia. Quaternary Research 44: 447–453.

McInnes, P. F., R. J. Naiman, J. Pastor, and Y. Cohen

1992 Effects of Moose Browsing on Vegetation and Litter of the Boreal Forest, Isle

Royale, Michigan, USA. Ecology 73(5): 2059–2775.

Meachen, Julie A., Adrianna C. Janowicz, Jori E. Avery, and Rudyard W. Sadleir

2014 Ecological Changes in Coyotes (Canis latrans) in Response to the Ice Age

Megafaunal Extinctions. PloS one 9(12): e116041.

Meiri, Meirav, Adrian M. Lister, Matthew J Collins, Noreen Tuross, Ted Goebel, Simon

Blockley, Grant D. Zazula, Nienke Van Doorn, R. Dale Guthrie, Gennady G Boeskorov,

Gennady F Baryshnikov, Andrei Sher, and Ian Barnes

2014 Faunal record identifies Bering isthmus conditions as constraint to end-

Pleistocene migration to the New World. Proceedings of the Royal Society 281:

20132167.

Milakovic, Brian, and Katherine L. Parker

225

2013 Quantifying carnivory by grizzly bears in a multi-ungulate system. Journal of

Wildlife Management 77(1): 39–47.

Morin, Peter J.

2011 Community Ecology. Wiley-Blackwell.

Newsome, Seth D., Carlos M. Del Rio, Stuart Bearhop, and Donald L. Philips

2007 A niche for isotopic ecology. Frontiers in Ecology and the Environment 5(8).

Newsome, Thomas M., Luigi Boitani, Guillaume Chapron, Paolo Ciucci, Christopher R.

Dickman, Justin A. Dellinger, José V. López-Bao, Rolf O. Peterson, Carolyn R. Shores, Aaron J.

Wirsing, and William J. Ripple

2016 Food habits of the world’s grey wolves. Mammal Review 46(4): 255–269.

Nogués-Bravo, David, Jesús Rodríguez, Joaquín Hortal, Persaram Batra, and Miguel B. Araújo

2008 Climate change, humans, and the extinction of the woolly mammoth. PLoS

biology 6(4): e79.

Olofsson, J., P. E. Hulme, L. Oksanen, and O. Suominen

2004 Importance of large and small mammalian herbivores for the plant community

structure in the forest tundra ecotone. Oikos 106(2): 324–334.

Olofsson, Johan, Lauri Oksanen, Terry Callaghan, Philip E. Hulme, Tarja Oksanen, and Otso

Suominen

2009 Herbivores inhibit climate-driven shrub expansion on the tundra. Global Change

Biology 15(11): 2681–2693.

Oswald, W. Wyatt, L. B. Brubaker, F. S. Hu, and G. W. Kling

2014 Late Quaternary Environmental and Ecological History of the Arctic Foothills,

Northern Alaska. In Alaska’s Changing Arctic: Ecological Consequences for

226

Tundra, Streams, and Lakes, edited by J. E. Hobbie and G. W. Kling, pp. 81–89.

Oxford University Press, New York.

Owen-Smith, Norman, and M. G. L. Mills

2008 Predator-prey size relationships in an African large-mammal food web. Journal of

Animal Ecology 77: 173–183.

Palombo, Maria Rita, and Margherita Mussi

2006 Large mammal guilds at the time of the first human colonization of Europe: The

case of the Italian Pleistocene record. Quaternary International 149: 94–103.

Pardi, Melissa I., and Felisa A. Smith

2015 Biotic responses of canids to the terminal Pleistocene megafauna extinction.

Ecography: 141–151.

Pastor, J., B. Dewey, R. J. Naiman, P. F. Mcinnes, and Y. Cohen

1993 Moose Browsing and Soil Fertility in the Boreal Forests of Isle Royale National

Park. Ecology 74(2): 467–480.

Pérez-Claros, J. A., and P. Palmqvist

2008 How many potential prey species account for the bulk of the diet of mammalian

predators? Implications for stable isotope paleodietary analyses. Journal of

Zoology 275(1): 9–17.

Pitulko, Vladimir V.

2011 The Berelekh quest: A review of forty years of research in the mammoth

graveyard in northeast Siberia. Geoarchaeology 26(1): 5–32.

Post, E., and C. Pedersen

227

2008 Opposing plant community responses to warming with and without herbivores.

Proceedings of the National Academy of Sciences 105(34): 12353–12358.

Potter, Ben A.

2007 Models of faunal processing and economy in Early Holocene interior Alaska.

Environmental Archaeology 12(1): 3–23.

2008 Radiocarbon Chronology of Central Alaska: Technological Continuity and

Economic Change. Radiocarbon 50(2): 181–204.

Potter, Ben A., Charles E Holmes, and David R. Yesner

2013 Technology and Economy among the Earliest Prehistoric Foragers in Interior

Eastern Beringia. In Paleoamerican Odyssey, edited by Kelly E. Graf, C. V.

Ketron, and M. R. Waters, pp. 81–103. Center for the Study of the First

Americans, College Station, TX.

Potter, Ben A., Joel D Irish, Joshua D. Reuther, and Holly J. McKinney

2014 New insights into Eastern Beringian mortuary behavior: A terminal Pleistocene

double infant burial at Upward Sun River. Proceedings of the National Academy

of Sciences 111(48): 17060–17065.

Raghavan, Maanasa, Gonçalo Espregueira Themudo, Colin I. Smith, Grant D. Zazula, and Paula

F. Campos

2014 Musk ox (Ovibos moschatus) of the mammoth steppe: tracing palaeodietary and

palaeoenvironmental changes over the last 50,000 years using carbon and

nitrogen isotopic analysis. Quaternary Science Reviews 102: 192–201.

Reimer, P. J., E. Bard, A. Bayliss, J. Beck, P. G Blackwell, Christopher Bronk Ramsey, P. M.

Grootes, T. P. Guilderson, H. Hadlifason, I. Hajdas, C. Hatte, T. J. Heaton, D. L. Hoffmann, A.

228

G. Hogg, K. A. Hughen, K. F. Kaiser, B. Kromer, S. W. Manning, M. Niu, R. W. Reimer, D. A.

Richards, E. M .Scott, J. R. Southon, R. A. Staff, C. S. M. Turney, and J. van der Plicht

2013 IntCal13 and Marine13 Radiocarbon Age Calibration Curves 0-50,000 Years cal

BP. Radiocarbon 55(4): 1869–1887.

Reuther, Joshua D.

2013 Late Glacial and Early Holocene Geoarchaeology and Terrestrial Paleoecology

in the Lowlands of the Middle Tanana Valley, Subarctic Alaska. PhD Dissertation,

University of Arizona.

Reuther, Joshua D., Ben A. Potter, C. E. Holmes, J. K. Feathers, François Lanoë, and J.

Kielhofer

2016 The Rosa-Keystone Dunes Field: The geoarchaeology and paleoecology of a late

Quaternary stabilized dune field in Eastern Beringia. The Holocene 26(12): 1939–

1953.

Riginos, Corinna, and James B. Grace

2008 Savanna tree density, herbivores, and the herbaceous community: Bottom-up vs.

top-down effects. Ecology 89(8): 2228–2238.

Ritchie, Mark E., and Han Olff

2004 Resource Partitioning and Biodiversity in Fractal Environments with Applications

to Dryland Communities. In Biodiversity in Drylands: Towards a Unified

Framework, edited by M Schachak, J R Gosz, S T A Pickett, and A Perevolotsky,

pp. 206–219. Oxford University Press.

Rivals, Florent, Matthew C. Mihlbachler, Nikos Solounias, Dick Mol, Gina M. Semprebon, John de Vos, and Daniela C. Kalthoff

229

2010 Palaeoecology of the Mammoth Steppe fauna from the late Pleistocene of the

North Sea and Alaska: Separating species preferences from geographic influence

in paleoecological dental wear analysis. Palaeogeography, Palaeoclimatology,

Palaeoecology 286(1–2): 42–54.

Rivals, Florent, and G.M. Semprebon

2016 Latitude matters: An examination of behavioural plasticity in dietary traits

amongst extant and Pleistocene Rangifer tarandus. Boreas.

Rivals, Florent, and Gina M. Semprebon

2011 Dietary plasticity in ungulates: Insight from tooth microwear analysis. Quaternary

International 245(2): 279–284.

Rodríguez-Gómez, Guillermo, Paul Palmqvist, Jesús Rodríguez, Ana Mateos, M Patrocinio

Espigares, Sergio Ros-Montoya, and Bienvenido Martínez-Navarro

2016 On the ecological context of the earliest human settlements in Europe : Resource

availability and competition intensity in the carnivore guild of Barranco León-D

and Fuente Nueva-3 (Orce, Baza Basin, SE Spain). Quaternary Science Reviews

143: 69–83.

Rodríguez-Gómez, Guillermo, Jesús Rodríguez, Jesús Ángel Martín-González, Idoia

Goikoetxea, and Ana Mateos

2013 Modeling trophic resource availability for the first human settlers of Europe: the

case of Atapuerca TD6. Journal of human evolution 64(6): 645–57.

Sankaran, Mahesh, Niall P. Hanan, Robert J. Scholes, Jayashree Ratnam, David J. Augustine,

Brian S. Cade, Jacques Gignoux, Steven I. Higgins, Xavier Le Roux, Fulco Ludwig, Jonas Ardo,

Feetham Banyikwa, Andries Bronn, Gabriela Bucini, Kelly K. Caylor, Michael B. Coughenour,

230

Alioune Diouf, Wellington Ekaya, Christie J. Feral, Edmund C. February, Peter G. H. Frost,

Pierre Hiernaux, Halszka Hrabar, Kristine L. Metzger, Herbert H. T. Prins, Susan Ringrose,

William Sea, Jörg Tews, Jeff Worden, and Nick Zambatis

2005 Determinants of woody cover in African savannas. Nature 438(7069): 846–849.

Sattler, Robert A.

1997 Large mammals in Lower Rampart Cave 1, Alaska: interspecific utilization of an

eastern Beringian cave. Geoarchaeology 12(6): 657–688.

Sattler, Robert A., Vinson, and T. E. Gillispie

2001 Calibrated Radiocarbon Ages and Taphonomic Factors in Beringian Cave Faunas

at the End of the Pleistocene. In People and Wildlife in Northern North America,

edited by S. C. Gerlach and M. S. Murray, pp. 112–123. Archaeopress: BAR

International Series 944, Oxford.

Schwartz-Narbonne, Rachel, Fred J. Longstaffe, Jessica Z. Metcalfe, and Grant D. Zazula

2015 Solving the woolly mammoth conundrum: amino acid 15N-enrichment suggests a

distinct forage or habitat. Scientific reports 5: 9791.

Seierstad, Inger K., Peter M Abbott, Matthias Bigler, Thomas Blunier, Anna J. Bourne, Edward

Brook, Susanne L. Buchardt, Christo Buizert, Henrik B. Clausen, Eliza Cook, Dorthe Dahl-

Jensen, Siwan M. Davies, S. Pedersen, Trevor J. Popp, Myriam Guillevic, Sigfús J. Johnsen,

Sune O. Rasmussen, Jeffrey P. Severinghaus, Anders Svensson, and Bo M. Vinther

2014 Consistently dated records from the Greenland GRIP , GISP2 and NGRIP ice

cores for the past 104 ka reveal regional millennial-scale d 18 O gradients with

possible Heinrich event imprint. Quaternary Science Reviews 106: 29–46.

231

Shapiro, Beth, Alexei J. Drummond, Andrew Rambaut, Michael C. Wilson, Paul E. Matheus,

Andrei V. Sher, Oliver G. Pybus, M Thomas P. Gilbert, Ian Barnes, Jonas Binladen, Eske

Willerslev, Anders J. Hansen, Gennady F. Baryshnikov, James A. Burns, Sergei Davydov,

Jonathan C. Driver, Duane G. Froese, C. Richard Harington, Grant Keddie, Pavel Kosintsev,

Michael L. Kunz, Larry D. Martin, Robert O. Stephenson, John Storer, Richard Tedford, Sergei

Zimov, and Alan Cooper

2004 Rise and fall of the Beringian steppe bison. Science 306: 1561–1565.

Simberloff, Daniel, and Tamar Dayan

1991 The Guild Concept and the Structure of Ecological Communities. Annual Review

of Ecology and Systematics 22: 115–143.

Sinclair, A. R. E.

2003 Mammal population regulation, keystone processes and ecosystem dynamics.

Philosophical transactions of the Royal Society of London 358: 1729–1740.

Sinclair, A. R. E., Simon Mduma, and Justin S. Brashares

2003 Patterns of predation in a diverse predator-prey system. Nature 425: 288–290.

Solow, Andrew R., David L. Roberts, and Karen M. Robbirt

2006 On the Pleistocene extinctions of Alaskan mammoths and horses. Proceedings of

the National Academy of Sciences 103(19): 7351–3.

Stephenson, Robert O., S. C. Gerlach, R. Dale Guthrie, C. R. Harington, R. O. Mills, and G.

Hare

2001 Wood Bison in Late Holocene Alaska and Adjacent Canada: Paleontological,

Archaeological and Historical Records. In People and Wildlife in Northern North

232

America, edited by S. C. Gerlach and M. S. Murray, pp. 125–133. Archaeopress:

BAR International Series, Oxford.

Stevens, Rhiannon E., and Robert E. M. Hedges

2004 Carbon and nitrogen stable isotope analysis of northwest European horse bone

and tooth collagen, 40,000 BP-present: Palaeoclimatic interpretations. Quaternary

Science Reviews 23(7–8): 977–991.

Stiner, Mary C.

1994 Honor among Thieves: A Zooarchaeological Study of Neandertal Ecology.

Princeton University Press.

2002 Carnivory, Coevolution, and the Geographic Spread of the Genus Homo. Journal

of Archaeological Research 10(1): 1–63.

2004 Comparative ecology and taphonomy of spotted hyenas, humans, and wolves in

Pleistocene Italy. Revue de Paléobiologie 23(2): 771–785.

Stiner, Mary C., Natalie D. Munro, and Todd A. Surovell

2000 The Tortoise and the Hare: Small-Game Use, the Broad-Spectrum Revolution,

and Paleolithic Demography. Current Anthropology 41(1): 39–73.

Stone, K. D., and J. A. Cook

2000 Phylogeography of black bears (Ursus americanus) of the Pacific Northwest.

Canadian Journal of Zoology 78: 1218–1223.

Stuart, Anthony J.

1991 Mammalian extinctions in the late Pleistocene of northern Eurasia and North

America. Biological Reviews 66(4): 453–562.

Stuart, Anthony J., P. A. Kosintsev, T. F. G. Higham, and Adrian M. Lister

233

2004 Pleistocene to Holocene extinction dynamics in giant deer and woolly mammoth.

Nature 431: 684–9.

Surovell, Todd A., Spencer R. Pelton, Richard Anderson-Sprecher, and Adam D. Myers

2016 Test of Martin’s overkill hypothesis using radiocarbon dates on extinct

megafauna. Proceedings of the National Academy of Sciences 113: 886–91.

Szpak, Paul

2014 Complexities of nitrogen isotope biogeochemistry in plant-soil systems:

implications for the study of ancient agricultural and animal management

practices. Frontiers in Plant Science 5: 288.

Szpak, Paul, Darren R. Gröcke, Regis Debruyne, Ross D.E. MacPhee, R. Dale Guthrie, Duane

Froese, Grant D. Zazula, William P. Patterson, and Hendrik N. Poinar

2010 Regional differences in bone collagen δ13C and δ15N of Pleistocene mammoths:

Implications for paleoecology of the mammoth steppe. Palaeogeography,

Palaeoclimatology, Palaeoecology 286(1–2): 88–96.

Tomasik, Eric, and Joseph A. Cook

2005 Mitochondrial Phylogeography and Conservation Genetics of Wolverine (Gulo

Gulo) of Northwestern North America. Journal of Mammalogy 86(2): 386–396.

Turner, Alan, Mauricio Antón, and Lars Werdelin

2008 Taxonomy and evolutionary patterns in the fossil Hyaenidae of Europe. Geobios

41: 677–687.

Tuross, Noreen

2012 Comparative decalcification methods, radiocarbon dates, and stable isotopes of

the VIRI bones. Radiocarbon 54(3–4): 837–844.

234

Urton, Erin J. M., and Keith A. Hobson

2005 Intrapopulation variation in gray wolf isotope (d15N and d13C) profiles:

Implications for the ecology of individuals. Oecologia 145(2): 317–326.

Valentine, D. W., K. Kielland, F.S. Chapin III, A. D. McGuire, and K. Van Cleve

2006 Patterns of Biogeochemistry in Alaskan Boreal Forests. In Alaska’s Changing

Boreal Forest, edited by F.S. Chapin III, M. W. Oswood, K. Van Cleve, L. A.

Viereck, and D. L. Verbyla, pp. 241–268. Oxford University Press, New York.

Van Valkenburgh, Blaire

2001 The Dog-Eat-Dog World of Carnivores: A Review of Past and Present Carnivore

Community Dynamics. In Meat Eating and Human Evolution, edited by C. B.

Stanford and H. T. Bunn, pp. 101–121. Oxford University Press, New York.

Van Valkenburgh, Blaire, Matthew W. Hayward, William J. Ripple, Carlo Meloro, and V.

Louise Roth

2015 The impact of large terrestrial carnivores on Pleistocene ecosystems. Proceedings

of the National Academy of Sciences 113(4): 862–867.

Vanak, Abi Tamim, Daniel Fortin, Maria Thaker, Monika Ogden, Cailey Owen, Sophie

Greatwood, and Rob Slotow

2013 Moving to stay in place: Behavioral mechanisms for coexistence of African large

carnivores. Ecology 94(11): 2619–2631.

Viau, A. E., K. Gajewski, M. C. Sawada, and J. Bunbury

2008 Low- and high-frequency climate variability in eastern Beringia during the past

25 000 years. Canadian Journal of Earth Sciences 45(11): 1435–1453.

235

Weinstock, Jaco, Eske Willerslev, Andrei Sher, Wenfei Tong, Simon Y. W. Ho, Dan Rubenstein,

John Storer, James Burns, Larry Martin, Claudio Bravi, Alfredo Prieto, Duane Froese, Eric Scott,

Lai Xulong, and Alan Cooper

2005 Evolution, systematics, and phylogeography of pleistocene horses in the new

world: a molecular perspective. PLoS biology 3(8): e241.

Willerslev, Eske, John Davison, Mari Moora, Martin Zobel, Eric Coissac, Mary E. Edwards,

Eline D. Lorenzen, Mette Vestergård, Galina Gussarova, James Haile, Joseph Craine, Ludovic

Gielly, Sanne Boessenkool, Laura S. Epp, Peter B Pearman, Rachid Cheddadi, David Murray,

Kari Anne Bråthen, Nigel Yoccoz, Heather Binney, Corinne Cruaud, Patrick Wincker, Tomasz

Goslar, Inger Greve Alsos, Eva Bellemain, Anne Krag Brysting, Reidar Elven, Jørn Henrik

Sønstebø, Julian Murton, Andrei Sher, Morten Rasmussen, Regin Rønn, Tobias Mourier, Alan

Cooper, Jeremy Austin, Per Möller, Duane Froese, Grant D. Zazula, François Pompanon,

Delphine Rioux, Vincent Niderkorn, Alexei Tikhonov, Grigoriy Savvinov, Richard G. Roberts,

Ross D.E. MacPhee, M. Thomas P. Gilbert, Kurt H Kjær, Ludovic Orlando, Christian

Brochmann, and Pierre Taberlet

2014 Fifty thousand years of Arctic vegetation and megafaunal diet. Nature 506: 47–

51.

Wolf, Adam, Christopher E. Doughty, and Yadvinder Malhi

2013 Lateral Diffusion of Nutrients by Mammalian Herbivores in Terrestrial

Ecosystems. PLoS ONE 8(8): e71352.

Wooller, Matthew J, Grant D. Zazula, Mary Edwards, Duane G. Froese, Richard D. Boone,

Carolyn Parker, and Bruce Bennett

236

2007 Stable Carbon Isotope Compositions of Eastern Beringian Grasses and Sedges:

Investigating Their Potential as Paleoenvironmental Indicators. Arctic, Antarctic,

and Alpine Research 39(2): 318–331.

Yeakel, Justin D., Paulo R Guimarães, Hervé Bocherens, and Paul L. Koch

2013 The impact of climate change on the structure of Pleistocene food webs across the

mammoth steppe. Proceedings of the Royal Society 280: 2013.0239.

Yesner, David R.

2001 Human dispersal into interior Alaska: antecedent conditions, mode of

colonization, and adaptations. Quaternary Science Reviews 20(1–3): 315–327.

Yesner, David R., K. J. Crossen, and Norman A. Easton

2011 Gearchaeological and Zooarchaeological Correlates of Early Beringian

Assemblages: Insights from the Little John Site, Yukon. In From the Yenisei to

the Yukon: Interpreting Lithic Assemblage Variability in Late Pleistocene/Early

Holocene Beringia, edited by Ted Goebel and Ian Buvit, pp. 308–322. Texas

A&M University Press, College Station.

Youngman, P. M.

1993 The Pleistocene Small Carnivores of Eastern Beringia. Canadian Field Naturalist

107(2): 139–163.

Zazula, Grant D., Duane G. Froese, Charles E. Schweger, Rolf W. Mathewes, Alwynne B.

Beaudouin, Alice M. Telka, C. Richard Harington, and John A. Westgate

2003 Ice-age steppe vegetation in east Beringia. Nature 423: 603.

Zazula, Grant D., Ross D. E. MacPhee, Jessica Z. Metcalfe, Alberto V. Reyes, Fiona Brock,

Patrick S. Druckenmiller, Pamela Groves, C. Richard Harington, Gregory Hodgins, Michael L.

237

Kunz, Fred J. Longstaffe, Daniel H. Mann, H. Gregory McDonald, Shweta Nalawade-Chavan, and John R. Southon

2014 American mastodon extirpation in the Arctic and Subarctic predates human

colonization and terminal Pleistocene climate change. Proceedings of the

National Academy of Sciences 111(52): 18460–18465.

Zimov, S.A., N.S. Zimov, A.N. Tikhonov, and F.S. Chapin III

2012 Mammoth steppe: a high-productivity phenomenon. Quaternary Science Reviews 57: 26–45.

Zimov, S. A., V. I. Chuprynin, A. P. Oreshko, F.S. Chapin III, J. F. Reynolds, and C. Chapin

1995 Steppe-Tundra Transition : A Herbivore-Driven Biome Shift at the End of the Pleistocene. The American Naturalist 146(5): 765–794.

238