Paleoamerican Odyssey (Peopling of the Americas Publications)
 1623491924, 9781623491925

Citation preview

[128.104.46.206] Project MUSE (2024-03-01 06:01 GMT) UW-Madison Libraries

PALEOAMERICAN ODYSSEY

Peopling of the Americas Publications Sponsored by the Center for the Study of the First Americans General Editors: Michael R. Waters and Ted Goebel

Paleoamerican odyssey Edited by Kelly E. Graf, Caroline V. Ketron, and Michael R. Waters

Center for the Study

TEXAS A&M PRESS of UNIVERSITY the First Americans College Station Department of Anthropology Texas A&M University

Copyright © 2014 by Kelly E. Graf, Caroline V. Ketron, and Michael R. Waters All rights reserved Original publication by Center for the Study of the First Americans. First Texas A&M University Press edition, 2014 Manufactured in the United States of America This paper meets the requirements of ANSI/NISO, Z39.48-1992 (Permanence of Paper). Binding materials have been chosen for durability.

Library of Congress Cataloging-in-Publication Data Paleoamerican Odyssey Conference (2013 : Santa Fe, N.M.) Paleoamerican odyssey / edited by Kelly E. Graf, Caroline V. Ketron, and Michael R. Waters. pages cm “Peopling of the Americas publications.” Contains 31 of 37 text versions of papers presented orally at the Paleoamerican Odyssey Conference held in Santa Fe, New Mexico in 2013. Includes bibliographical references and index. ISBN 978-1-62349-192-5 (pbk. : alk. paper) — ISBN 978-1-62349-233-5 (ebook) 1. Paleo-Indians—Congresses.  2. Human beings—Migrations—Congresses.  3. Clovis culture— Congresses.  4. Paleoecology—Pleistocene—Congresses.  5. America—Antiquities—Congresses. I. Graf, Kelly E., editor of compilation.  II. Ketron, Caroline V. , editor of compilation.  III. Waters, Michael R., editor of compilation.  IV. Title. E61.P15 2013 970.01—dc23 2014006259 Design and typesetting by C&C Wordsmiths, Lenoir, North Carolina Cover art by Heather L. Smith

Contents Preface . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Kelly E. Graf

vii

Part i Human dispersals in the old World and Beringia 1. Occupying New Lands: Global Migrations and Cultural Diversi ication with Particular Reference to Australia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Peter Hiscock 2. Human Habitation in Arctic Western Beringia Prior to the LGM . . . . . . . . . . . . . . . . . . . . Vladimir Pitulko, Pavel Nikolskiy, Aleksandr Basilyan, and Elena Pavlova 3. Human Technological and Behavioral Adaptation to Landscape Changes around the Last Glacial Maximum in Japan: A Focus on Hokkaido . . . . . . . . . . . . . . . . . . . . . . . . Masami Izuho 4. Siberian Odyssey . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Kelly E. Graf 5. Technology and Economy among the Earliest Prehistoric Foragers in Interior Eastern Beringia . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Ben A. Potter, Charles E. Holmes, and David R. Yesner 6. Biface Traditions of Northern Alaska and Their Role in the Peopling of the Americas . . . . . . Heather L. Smith, Jeffrey T. Rasic, and Ted Goebel

3 13

45 65

81 105

Part ii dispersal routes to the new World: archaeology and Genetics 7. After Clovis-First Collapsed: Reimagining the Peopling of the Americas . . . . . . . . . . . . . . . 127 Jon M. Erlandson 8. Locating Pleistocene-age Submerged Archaeological Sites on the Northwest Coast: Current Status of Research and Future Directions . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 133 Quentin Mackie, Loren Davis, Daryl Fedje, Duncan McLaren, and Amy Gusick 9. Vectors, Vestiges, and Valhallas—Rethinking the Corridor . . . . . . . . . . . . . . . . . . . . . . . . 149 John W. Ives, Duane Froese, Kisha Supernant, and Gabriel Yanicki 10. Three-Stage Colonization Model for the Peopling of the Americas . . . . . . . . . . . . . . . . . . 171 Connie J. Mulligan and Andrew Kitchen 11. The Late-Pleistocene Human Settlement of Interior North America: The Role of Physiography and Sea-Level Change. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 183 David G. Anderson, Thaddeus G. Bissett, and Stephen J. Yerka

Part iii clovis-era archaeology and ecology 12. Clovis across the Continent . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 207 D. Shane Miller, Vance T. Holliday, and Jordon Bright 13. The Clovis Landscape . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 221 Vance T. Holliday and D. Shane Miller 14. Imagining Clovis as a Cultural Revitalization Movement . . . . . . . . . . . . . . . . . . . . . . . . . 247 Bruce A. Bradley and Michael B. Collins 15. Clovis Caches: Current Perspectives and Future Directions . . . . . . . . . . . . . . . . . . . . . . . 257 J. David Kilby and Bruce B. Huckell 16. Complexities of the Colonization Process: A View from the North American West . . . . . . . 273 Charlotte Beck and George T. Jones 17. Clovis-era Subsistence: Regional Variability, Continental Patterning . . . . . . . . . . . . . . . . . 293 Gary Haynes and Jarod M. Hutson v

18. Pleistocene Extinctions: The State of Evidence and the Structure of Debate . . . . . . . . . . . 311 Nicole M. Waguespack

Part iV news from latin america 19. The First Human Settlers on the Yucatan Peninsula: Evidence from Drowned Caves in the State of Quintana Roo (South Mexico). . . . . . . . . . . . . . . . . . . . . . . . . . . . Arturo H. González, Alejandro Terrazas, Wolfgang Stinnesbeck, Martha E. Benavente, Jerónimo Avilés, Carmen Rojas, José Manuel Padilla, Adriana Velásquez, Eugenio Acevez, and Eberhard Frey 20. The Initial Colonization of South America Eastern Lowlands: Brazilian Archaeology Contributions to Settlement of America Models . . . . . . . . . . . . . . . . . . . . . Adriana Schmidt Dias and Lucas Bueno 21. Rethinking Early Objects and Landscapes in the Southern Cone: Fishtail-Point Concentrations in the Pampas and Northern Patagonia . . . . . . . . . . . . . . . . . . . . . . . . . Nora Flegenheimer, Laura Miotti, and Natalia Mazzia 22. Entangled Knowledge: Old Trends and New Thoughts in First South American Studies . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Tom D. Dillehay 23. Early Human Occupation of Lagoa Santa, Eastern Central Brazil: Craniometric Variation of the InitialSettlers of South America . . . . . . . . . . . . . . . . . . . Walter A. Neves, Mark Hubbe, Danilo Bernardo, André Strauss, Astolfo Araujo, and Renato Kipnis

323

339

359

377

397

Part V Pre-clovis archaeology 24. Fingerprinting Flake Production and Damage Processes: Toward Identifying Human Artifact Characteristics . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . William Andrefsky, Jr. 25. The Mammoth Steppe Hypothesis: The Middle Wisconsin (Oxygen Isotope Stage 3) Peopling of North America. . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Steven R. Holen and Kathleen Holen 26. The Late-Pleistocene Industries of Piauí, Brazil: New Data . . . . . . . . . . . . . . . . . . . . . . . Eric Boëda, Antoine Lourdeau, Christelle Lahaye, Gisele Daltrini Felice, Sibeli Viana, Ignacio Clemente-Conte, Mario Pino, Michel Fontugne, Sirlei Hoeltz, Niède Guidon, Anne-Marie Pessis, Amélie Da Costa, Marina Pagli 27. Pre-Clovis Megafauna Butchery Sites in the Western Great Lakes Region, USA . . . . . . . . . Daniel J. Joyce 28. Geochronology, Archaeological Context, and DNA at the Paisley Caves . . . . . . . . . . . . . . Dennis L. Jenkins, Loren G. Davis, Thomas W. Stafford, Jr., Paula F. Campos, Thomas J. Connolly, Linda Scott Cummings, Michael Hofreiter, Bryan Hockett, Katelyn McDonough, Ian Luthe, Patrick W. O’Grady, Karl J. Reinhard, Mark E. Swisher, Frances White, Bonnie Yates, Robert M. Yohe II, Chad Yost, Eske Willerslev 29. The Ones That Still Won’t Go Away: More Biased Thoughts on the Pre-Clovis Peopling of the New World . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . J. M. Adovasio and David R. Pedler 30. North America before Clovis: Variance in Temporal/Spatial Cultural Patterns, 27,000–13,000 cal yr BP . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Michael B. Collins, Dennis J. Stanford, Darrin L. Lowery, and Bruce A. Bradley 31. The First Americans: A Review of the Evidence for the Late-Pleistocene Peopling of the Americas . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . Michael R. Waters and Thomas W. Stafford, Jr. I

415

429 445

467 485

511

521

541

index . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . . 561

vi

Preface and Acknowledgments This volume, Paleoamerican Odyssey, was created as a companion to the Paleoamerican Odyssey Conference, held in Santa Fe, New Mexico, October 17–19, 2013. Over the past 14 years since the C ­ lovis and Beyond Conference, continued research in peopling of the Americas studies has taken us a long way. We now recognize that human dispersal to the Western Hemisphere was a complicated process. Renewed archaeological research in far-reaching places and contributions made in the studies of human molecular genetics and paleontology have transformed understanding of this process. The conference and companion volume bring together leading scholars from around the world currently researching the problem. With this volume, we present in full text 31 of the 37 original papers presented at the conference in October.

Caroline Ketron was instrumental in helping manage submissions and reviews and proof-reading chapters. Michael Waters developed the initial list of speakers and contributors and was especially helpful near the end of production while we were in the field. Ted Goebel contributed to the original concept and organization of the speaker list for the conference and hence the contributor list for the volume. He also provided invaluable advice throughout the production process. Jim and Char Chandler supplied excellent and muchappreciated copy-editing and page-layout skills and services. Without their devotion to detail, this volume would never have left the ground. In addition to being an excellent archaeologist, Heather Smith is gifted in graphic design and provided the cover design.

Special thanks are given to the list of contributors. Thanks to their diligence, chapters were prepared in a timely fashion, insuring external review and revision during the 12 months leading up to the conference. Without their attention, this book would never have been brought to fruition, especially in time to be distributed at the conference. Two reviewers commented on each paper so that more than 50 scholars participated in the review process. The editors are grateful to them for their willingness to contribute in this capacity. Without their timely responses the book would never have been completed in time for pre-conference print.

Many thanks go to the Center board and members for helpful suggestions and continued support. We are immensely grateful to Robert and Sharon ­Wilson for their generous donation, which helped to support the conference and publish this volume. Funding for the publication of this volume was also provided by the many donations received in ­memory of Lile Mecom Mullins.  Kelly E. Graf

vii



I. Human Dispersals in the Old World and Beringia

Chapter 1 Occupying New Lands: Global Migrations and Cultural Diversification with Particular Reference to Australia Peter Hiscock ABSTRACT In the Australasian region cultural differentiation, experimentation, and adaptation characterize the global dispersion of Homo sapiens. The migration of humans out of Africa and into Australia was not a singular process governed and guided by persistent traditions; normative and static images of social and economic organization cannot explain the diversity of cultural evidence associated with the dispersion. This paper reviews the evidence for a dynamic process of social, economic, and technological diversification associated with the spread of humans and their adaptation to new social and physical environments. The evidence is read as processes of adaptation, cultural connectedness, and isolation, interacting in complex and regionally varying ways. In the case of Australia the patterns of occupational history and cultural trajectories were not the product of conservative maintenance of social tradition or a diminished set of cultural practices following serial founder events. The evidence is best interpreted as a radiation of cultural practices and perspectives that reveals an adaptive response to the occupation of new lands. KEYWORDS:  Australia, Sahul, Colonization, Out of Africa, Adaptive radiation

Introduction

Australia was colonized by groups of Homo sapiens whose ancestors had migrated out of Africa many millennia earlier. Global dispersal of humans was a prolonged, complex process with multiple migrations in many regions. Archaeological and genetic evidence is currently revealing some of the complexity of that geographic expansion. Human settlement of Australia is a crucial case study of this dispersal process. The colonization of Australasia is one culminating event of the global expansion and must be understood in light of what we now know of the global process. At the same time epartment of Archaeology, School of Philosophical and Historical D Inquiry, University of Sydney, Sydney, NSW 2006, Australia; e-mail:  [email protected]

Australia provides a landscape that illuminates the timing and nature of human dispersals, particularly in the region of Oceania and East Asia. Australia enhances our understanding of early migratory processes because of a rare combination of qualities: It was colonized by one of the first waves of migrating H. sapiens moving eastward from Africa, the land was previously empty of other hominins thereby giving clarity to the initial occupational period, and there is a growing quantity of good-quality archaeological data available for parts of the continent. In this chapter I summarize current evidence for the arrival and dispersion of humans across the Pleistocene Australian continent, called Sahul, in the context of what we know of the broader eastward dispersion within the Old World. Current debates about the phase of colonization in Aus-

4 tralia focus on four critical questions of evidence. First, when did humans arrive and at what rate did they settle the diverse environments across the continent? Second, did human colonizers contribute substantially to environmental transformations in Australia? Third, what was the technological repertoire of foragers at and following colonization; and was that repertoire narrowing over time? Fourth and finally, what does public signaling, revealed archaeologically through art or ornamentation or ritual, tell us about the configuration of the colonizing societies. Answers to these questions have played a central role in discussions of not only the arrival of people in the Australian landscape, but more broadly in models of the dispersion process between Africa and Sahul. I will discuss key strands of evidence for each of those four questions.

Timing of Colonization and Rate of Settlement across Australia

Archaeological evidence indicates humans probably arrived in Australia between 45,000 and 60,000 years ago, and most likely somewhere around 50,000–55,000 years ago. This “window” of colonization is well established but imprecise, so that while some researchers prefer an antiquity in the low 50,000s others argue that humans arrived slightly later. These differences of opinions between Australian archaeologists are strongly held, and hence any statement about timing of colonization is considered controversial in a debate that increasingly shadows the Clovis/pre-Clovis controversy of the Americas. My view, as outlined in Hiscock (2008), is that 1) given the complexities and imprecision of both radiometric age estimates and stratigraphic associations the only statement that would be generally accepted today is that the earliest evidence for humans in Australia dates to 50,000 ± 5,000 years ago, 2) that given the extremely small sample of sites known the age of the earliest occupation identified thus far must be considered a minimum age for the colonization, and 3) the difference of a few thousand years may make very little difference to the process of arrival: either end of this “window” is long after H. sapiens exited Africa, and though the size of water barriers would change over time those variations need not have posed significant problems for colonizers. Much of the debate about the possibility of pre-50,000 occupation rests on the interpretation of sites in western Arnhem Land, near to the modern northern coastline. There, excavations at Nauwalabila and Malakunanja II yielded layers containing artifacts in association with sediments that have been estimated to be between 50–60 kya. While some archaeologists have accepted these age estimates as indications of human antiquity in the continent, others have rejected the stratigraphic association between dated materials and cultural materials and have argued that colonization was more recent. The most strident critique of association, and

Hiscock the strongest advocacy of a more recent age of colonization, has been advanced by Jim O’Connell and Jim Allen in a series of articles over the last decade (Allen and O’Connell 2003, 2008; O’Connell and Allen 2004, 2007, 2012; O’Connell et al. 2010). They argue humans can only be reliably dated to 44,000–46,000 years BP (O’Connell and Allen 2012:10), and that claims for earlier occupation have been refuted. Their claim is based on an interpretation that vertical movement within the Arnhem Land shelters, principally as a result of termite activity, has brought lithic artifacts and dated sand grains into a false association. In arriving at that conclusion they argue that all the lower artifacts have moved substantially downward through the sandy deposits and they imply that stratigraphic evidence that would document minimal vertical displacement, of a small pit dug more than 40,000 years ago claimed for Malakunanja II, is inadequately documented and perhaps that the pit may not exist (O’Connell and Allen 2004). These propositions are plausible, and O’Connell and Allen cite instances where substantial vertical displacement has been documented in Australian sandstone shelter deposits (e.g., Richardson 1992). And yet the evidence from Australian shelters also shows that where vertical movement occurs it is bidirectional, with artifacts moving up as well as down, and that substantial movement is not universal—some sandy deposits retain reasonably precise indications of the associations created at deposition (see Richardson 2011; Stockton 1973). O’Connell and Allen have not re-analyzed either the artifacts or the sediments recovered from Malakunanja II, and so while their models of taphonomic disturbance are relevant to the colonization debate they are not proof that artifacts had moved and are consequently not associated with the sediments well over 50,000 years old from which they were recovered. O’Connell and Allen seek to bolster their case by arguing that the inability of archaeologists to find more sites older than 50 kya indicates that there are no such sites to be found (e.g., Allen and O’Connell 2003:17). However, in the absence of discussions about the kinds of sites that have been dug in recent years and the probability of finding sites of high antiquity (see Langley et al. 2011; Surovell and Grund 2012), this argument is not compelling. We might equally argue that finding one or two sites of that vintage was remarkable luck given the small sample of early sites we had and still have. In any case, Malakunanja has recently been re-excavated by Chris Clarkson and his analysis of the site will shed light on its formation and antiquity in the near future. Whatever consensus emerges for the Arnhem Land sites there are a number of sites from the central and southern portions of the continent that are widely accepted as indicating occupation between 45,000 and 50,000 years ago. For instance, in southeastern Australia debris from human occupation was present at Lake Mungo, with the stratigraphically lowest artifacts bracketed by luminescence dates of 45.7 ± 2.3 and 50.1 ± 2.4 kya (Bowler et al. 2003). Although the stratigraphic integrity of these deposits have been queried (O’Connell and Allen 2004), the position of artifacts

Occupying New Lands: Global Migrations and Cultural Diversification with Particular Reference to Australia within the deposit has been coherently explained as reflecting ancient landsurface slope by the excavator (Shawcross 1998; Bowler et al. 2003), and there is little reason to doubt their association, leading researchers such as Smith (2013) to infer use of the palaeolake soon after 50,000 years ago. In southwestern Australia the Devil’s Lair cave was occupied before 46,000–47,000 years ago (Turney et al. 2001). A thick fan of alluvium in Layer 30 Lower is dated by radiocarbon at 44,000–47,000 BP. Artifacts and faunal remains are found far below this level, as far down as Layer 38. These layers are dominated by limestone rubble, which would have limited vertical movement of large objects. O’Connell and Allen (2004:840) emphasize that these layers have small channels indicating water flow, and they conclude that the artifacts contained within them may not be in primary deposition context—a point noted long ago for the bones (Balme 1980). Based on this concern O’Connell and Allen (2004:841) argue that a human presence at Devil’s Lair somewhere in the 41–46 ka range is now indicated, the argument for an earlier occupation remains equivocal.

That conclusion is unsustainable. I have examined the specimens from layers 31–38 and they are clearly artifactual, including a retouched flake, and the specimens are made on calcrete, which is uncommon in layers 30 and above. My observations agree with Dortch’s characterization of the assemblage (reported in O’Connell and Allen 2004) and reinforce Dortch’s point that the early assemblage in layers 31–38 is not likely to be derived from higher layers by vertical movement. Obviously the artifacts from the lower layers may have initially been deposited in sediments elsewhere within or near the cave, then eroded and redeposited into their levels, but if redeposition occurred it did so long before 44,000– 47,000 BP and the artifacts in question are substantially older than that date. The conclusion of Turney et al. (2001), that Devil’s Lair represents human occupation at close to or before 50,000 years BP, stands on current evidence. A number of other sites currently appear to have material dated to about or more than 45,000 years BP, including Carpenters Gap in the northwest (Fifield et al. 2001), Nawarla Gabarnmang in Arnhem Land (David et al. 2011), and Parnkupirti in central Australia (Veth et al. 2009). Taken together with the information described above from Devil’s Lair, Lake Mungo, Nauwalabila, and Malakunanja II, this suite of sites shows that human occupation of the continent, including southern regions, took place before the 45,000–46,000 BP date that O’Connell and Allen have hitherto accepted. However, given the imprecision of age estimates in many of these sites it is not currently clear whether the earliest evidence of occupation in these regions is slightly before or around 50,000 years BP or closer to 55,000 BP, and consequently the published interpretations (many of which ignore the statistical implications of error values) do not provide a basis for evaluating whether there was a difference in the antiquity of archaeological materials between different regions in the

5

north or south of the continent. This has rendered models of the rate of dispersion across the Australian mainland largely untestable (see Rindos and Webb 1992). An additional consideration is that these archaeological signals are unlikely to mark the arrival of humans on the continent and it is more probable that colonization occurred well before the dated occupation levels. One reason for this is that even with constant rates of site destruction there will be progressively fewer early sites preserved over time. This agerelated process of site destruction is non-linear, as demonstrated by generic modeling (Hiscock 2008), archaeological datasets from Australia (Williams 2013), and non-archaeological sediment records from North America (e.g., Surovell et al. 2009). The cumulative consequence of attritional processes, continued over a prolonged period, would be that very few sites from the period of colonization would have survived until today. Recent arguments by Surovell and Grund (2012), using calculations of predicted attrition rates, indicate if there was a phase of megafauna extinction in Australia dated to more than 40,000 years BP it is unlikely that any megafauna kill sites would be preserved or discovered. This conclusion could, and should, be directly converted into a statement of how improbable it is that sites from the even earlier phase of colonization would be discovered, reinforcing the conclusion that maximum site age so far established underestimates the age of colonization. Differential destruction of sites in different locations and landscapes has also been considered to have obscured the record of colonization. For instance, the Pleistocene coastline was hundreds of kilometers from the modern one, and substantial tracts of land that colonists presumably occupied have been lost to sea-level rise, potentially destroying the occupational debris of many generations of settlers—an argument rehearsed by Bowdler (1977) almost 40 years ago, but still relevant to this debate. That observation raises the question of the delay we might except between the initial colonization of the then exposed continental shelf and the appearance of people in distant locations: in the south or east, and in the desertic inland. We cannot exclude the possibility of extremely rapid settlement of all regions, perhaps by highly mobile foragers moving rapidly across landscapes using a narrow dietary focus, which might have minimised the gap between colonization and occupation of the sites mentioned here (see Beaton 1991). Such a rapid dispersal rate would be congruent with the calculations by Birdsell (1957) and is supported by a number of researchers including O’Connell and Allen (2012). Alternatively, in any region, it might have taken some time for population levels to grow to a point where material indicators of human activity became sufficiently common that they have been preserved and can still be found by archaeologists today, and hence humans may have been present in landscapes for a prolonged period before they become archaeologically visible at about 50,000 years ago. This proposition does not suggest any particular dispersal rate, but it implies that the current sample of dated assemblages across Australia is not so much an indication of the initial ap-

6 pearance of humans in each region as it is a record of demographic growth and residency length at each location. This proposition is applicable to not only dated artifact assemblages, but also to the onset of higher charcoal deposition rates in sediment cores. While those charcoal spikes were initially interpreted as indicators of the arrival of people, they are increasingly interpreted as indicating the emergence of altered landscape management strategies after the extinction of mega-marsupials, perhaps substantially after people arrive in that landscape.

Human Contributions to Transformations in Australia’s Flora and Fauna

The nature and extent of human impact on the Australian landscape is a subject of great contention, and, like debates in the Americas, arguments have polarized into those in favor of severe and rapid humanly induced changes and those that minimise the contribution of humans to shifts in fauna and flora communities. The environmental transformations most discussed include the reduction and extinction of megafaunal browsers and grazers across the continent. A suite of very large animals had lived in Australia at some time prior to the arrival of humans: giant kangaroos (such as Macropus rufus and Macropus giganteus titan) and giant wombat (Phascolonus gigas), tall flightless birds (Genyornis sp.), and four-legged marsupial browsers and grazers the same size as some species of hippopotamus and rhinoceros (such as Diprotodon optatum, Zygomaturus sp., Palorchestes sp.). Some studies have claimed a broad coincidence between the time at which species of large marsupials disappeared and the time that humans arrived (Miller et al. 1999; Roberts et al. 2001), and have argued that the human colonization of Australia might have triggered a trophic collapse in which particular kinds of animals were driven to extinction. As skillful predators whose hunting behaviors were unfamiliar to marsupial prey, the dispersing humans no doubt had the capacity to reduce the viability of vulnerable species, especially if hunters targeted the young of animals that reproduced slowly or if the prey species were limited in distribution and predictable in movement by, for example, being tethered to rare resources (Brook and Bowman 2002, 2004). The plausibility of inferences about the impact of humans on megafauna are inevitably linked to inferences about the chronology and demography of human occupation. For instance, O’Connell and Allen (2012:12) propose a colonization date so late that it implies megafaunal extinctions were well underway or even completed in many landscapes before humans appeared in the landscape, and their conclusion that human exploration of the continent was initially rapid and conducted by small groups with slowly increasing populations makes it impossible for humans to have had a significant impact. However as explained above neither a 45,000 year BP colonization nor a slow rate of demographic growth is convincingly demonstrated. Other researchers have argued for slightly earlier colonization ages and a rapid extinction of megafauna immediately around 46,000 years BP, a pattern

Hiscock that enhances the plausibility of human predation models (e.g., Miller et al. 1999; Roberts et al. 2001; Rule et al. 2012). If slower rates of dispersion across Australia and sustained population growth occurred in early populations the potential for human impacts, directly and indirectly, on megafauna would also be more plausible. Since there is uncertainty in the antiquity, behavior, and population growth of colonists, claims can be made that either of the opposing views is plausible. So what does the archaeological and palaeontological evidence suggest? Archaeologists in Australia have never found sites where large extinct animals were killed, and it seems likely that many species of megafauna had become extinct prior to the arrival of humans (Field and Wroe 2012; Field et al. 2008). This might mean that there was minimal coexistence between humans and megafauna; yet there are sites that may well display a temporal overlap, such as Cuddie Springs. If humans exploited megafauna, then why are there no kill sites? While Flannery (1990) argued that the rapid speed of the extermination may be responsible it is more likely that lack of preservation of material dating to more than 40,000 years ago is responsible (see Surovell and Gundy 2012). Nevertheless, environmental signatures consistently point to the reduction in range and density of large animals, if not their final extinction, between 40,000 and 50,000 years ago (Miller et al. 1999; Roberts et al. 2001; Rule et al. 2012), coincident with the arrival of people, and so the appearance of human hunters in the landscape is implicated in the extinction process. The timing of extinction trends in these mega-animals coincides with intensification of long-term continental drying, reductions in resource levels, and restructuring of the environment, and so even low levels of predation by the new human hunters may have tipped some species into terminal declines or accelerated declines already underway (see Field and Wroe 2012). Most likely it was not intensive or exclusive hunting of these large animals that reduced their populations perilously, but merely the addition of a new social carnivore to stressful ecological circumstances. While the extinctions themselves have captured the imagination of researchers it is the removal of those animals from the Australian landscape that shaped subsequent human occupation, as Flannery (1990) observed. The consequences of the reduction in the size of large marsupial populations, and their subsequent extinction, would have been dramatic. Removing large browsers and grazers from the ecosystem means they are no longer actively consuming vegetation, thereby potentially resulting in reduced openness within forests, reduced ecosystem patchiness, and increased fuel load that might facilitate altered fire regimes and subsequently nutrient cycles. Evidence of this chain of ecological transformation is recorded from several places, most notably from Lynch’s Crater, a swamp in northeast Australia. Cores drilled into the deep sediments of this swamp produced a record of pollen, charcoal, and spores of the fungus Sporormiella, which is passed through the bowel of large herbivores and in theory can be used as a proxy for

Occupying New Lands: Global Migrations and Cultural Diversification with Particular Reference to Australia their presence in a landscape (Feranec et al. 2011). Counts of Sporormiella spores, and by implication the abundance of large herbivores, declined markedly about 41,000 years ago (Rule et al. 2012). Immediately afterwards, charcoal fragments in the sediments, and by implication fire frequency and intensity, increased in response to increased fuel. The sequence from Lynch’s Crater is consistent with charcoal pulses in many sedimentary sequences across Australia dating to approximately 40,000–50,0000 years ago (Kershaw et al. 2006). Although these charcoal signals have often been discussed as possible signals of the arrival of humans it is more likely, given the earlier dates now available for occupation, that altered burning regimes mark the point at which new ecological relationships emerge in a land now largely devoid of very large herbivores. It is reasonable to conclude that while initial settlement of new territories across the continent may have been assisted by the exploitation of substantial meat packages represented by the large herbivores, those animals would have been found in small numbers, geographically variable in abundance, and for only a limited period. As a consequence early foraging practices would have been reasonably diverse, an inference consistent with the early archaeological assemblages of animal bones, which reflect flexible foraging strategies focused on hunting small- to medium-size game. It is likely that plant foods such as yams and seeds would have supplemented meat in the deserts and that these would have varied between environments, though the archaeological evidence for this foraging is rare. Regional differences in economic strategies, probably combined with disparate demographic histories, most likely underpinned regional differences in not only economic practices that emerged as each landscape was settled, but also in technological and social practices as groups adapted to the specific circumstance they encountered.

The Technological Repertoire of Foragers at and Following Colonization

There is a long tradition of thinking about Australian Pleistocene technologies as “simple” (see Hiscock 2008; White 1977). These views owe much to the early archaeological views of Australian Aborigines as “Palaeolithic survivals” and to the tendency of archaeologists to employ implementcategory richness as a convenient measure of technological “complexity.” Such views of the technological simplicity of humans arriving in Australia have sometimes been construed as measures of a lack of cognitive or cultural capacity or sophistication. Such views are frequently framed from European perspectives, typified in the characterization of technology in Sahul presented by Mellars: The earliest stone-tool technologies documented across the whole of Australasia are conspicuously lacking in any trace of distinctively ‘‘modern’’ or ‘‘Upper Palaeolithic,’’ bladebased technologies of the kind recorded from both the later African Middle Stone Age sites and the earliest modern human sites in southwest Asia and Europe (Mellars 2006:799).

7

This equation of blade-making and modernity, and also the equation of additional elements in a “package” of behaviors common in Upper Palaeolithic Europe, sets an expectation that is not met everywhere in Africa or East Asia as well as Australia. Hence Boivin and colleagues offer a hint of surprise when they write: “The Australian evidence is particularly interesting in that H. sapiens appears to have arrived without an African “package” of innovations, and to have acquired these independently and piecemeal” (Boivin et al in press).

The “failure” of modern human colonizers of Australia to display the European range of behaviors challenges normative views of both modernity and global dispersions. Mellars (2006) suggests that the reduced cultural variation he perceived resulted from serial founder effects, as groups moving eastward progressively changed or lost cultural knowledge and practices—a mechanism that continues to be invoked (e.g., Clarkson in press; O’Connor et al. 2011). Such a view has also created a paradox for researchers who accept the image of a narrow behavioral repertoire in colonizers, but perceive the appearance in the late Holocene of far greater behavioral diversity, perhaps the “full package” Mellars anticipated: blade technologies, many formal implement types, rich artistic representations, and so on (Brumm and Moore 2005; Franklin and Hapgood 2007; Hapgood and Franklin 2008). These researchers have attempted to explain an unexpectedly late, but still revolutionary transition to modern, complex technologies by citing delayed demographic increases or cultural intensifications, even though independent evidence for such events is scant (see Hiscock 2008). Curiously there is consistent and persistent characterization of early technologies as “simple” and unsophisticated. There are two important points that have been and can be made about such arguments. The first is that technologies develop in and are maintained as responses to specific economic/social situations, rather than being ancestral characteristics that were retained until lost through processes such as evolutionary bottlenecks. Consequently technological strategies can shift in response to even small changes in economic and socials functions, and such adaptive shifts need not be directional or “progressive.” We know that “blade technologies” and production of standardized hafted implements such as “microliths” have come and gone for at least 300,000–500,000 years (e.g., Faivre 2012; Hiscock and O’Connor 2006; Johnson and McBrearty 2010) and may have been made by hominids other than H. sapiens. Those lithic technologies and others are not reflections of cognitive capacity; they are strategies employed within a historical context in response to local conditions to obtain fitness or other advantages. Because the functioning of Pleistocene Australian technologies is only minimally described the relationship between technology and context remains difficult to specify, but one standard mechanism that has been proposed is that the technology functions to create tools of a kind and a “complexity”

8 that reflects foraging strategies in which they operated. A recent example of this principle is O’Connell and Allen’s (2012) proposal that the simplicity of early Australian technologies is correlated with diet breadth. They argue that colonizing foragers pursued high-ranked prey, requiring high search but low handling/processing investments, and that “simple” technologies were adequate for this task. While the argument for low handling times may offer an insight into the morphological irregularity and infrequent use of seed grinders, it is less persuasive in explicating the flaking technology that would be geared to near exclusive targeting of high-ranked game in Australia. It is worth noting that the application of such a model would be problematic in non-Australian contexts, especially in the Americas, where elaborate projectile weaponry is almost the norm for mobile foragers with narrow diet breadth. More critical for the proposed mechanism is the lack of a definition of the difference between “simple” and “complex” technologies, raising the question of how lithic industries reflect the toolkit that was in use. Even if narrow diet breadth was a key driver of low technological diversity and investment in production of handling tools, as O’Connell and Allen have hypothesized, it is not clear than this would necessarily be reflected, in any simple way, in the lithic assemblages of the time. The kinds of unifacially retouched flakes, often called “scrapers,” described for early Australian assemblages have not been well studied for wear or residues, but it is likely they are maintenance tools used to create organic artifacts of unknown form and complexity. In Arnhem Land the exceptionally well preserved Dynamic-phase rock paintings, probably dating to the terminal Pleistocene, depict elaborate multi-pronged spears, diverse organic hunting tools including boomerangs, and organic clothing and ornamentation. Since these organic tools were made at a time when the lithic industries consisted principally of flakes from bipolar and single-platform percussion cores, retouched “scrapers,” and edge-ground axes, it is clear that the typological impressiveness or diversity of the lithic assemblages is a poor measure of their functional effectiveness or the diversity of organic products they created. And since those terminal-Pleistocene assemblages are broadly comparable with ones in northern Australia dated to more than 35,000–40,000 years BP there is no basis in lithic patterns to think that early or colonizing foraging technology was limited in diversity or sophistication. For instance, we can be sure that elaborate composite artifacts were manufactured at or immediately after colonization, as revealed by the presence of edge-ground axes in deposits close to or exceeding 40,000 years BP (e.g., Chris Clarkson pers.com.; Geneste et al. 2010). If that is the case the question about technological and economic relationships might be re-expressed as, In what conditions is technological “complexity” expressed in organic materials, but not in the lithic assemblage creating them? While it is tempting to invoke a temporary unfamiliarity with lithic resources by colonizers, or perhaps cite “severe climatic and environmental constraints” as O’Connell and Allen (2012)

Hiscock have done, or “scarcity of high-quality, finegrained stone for tool production” as Mellars (2006:798) did, the longevity of these patterns, and the Holocene shift to different lithic patterns, belie such proposals. The characteristics of early lithic assemblages are not atypical of Australian prehistory; rather they are persistent and widespread, and reflect fundamental and effective functional relationships. An emphasis on organic tools, the employment of lithic tools for maintenance far more than for extractive tasks, and in some contexts low-diversity generic lithic-production strategies, is epitomized in the nineteenth-century observations of Aboriginal life across the country. To demand that the efficacy or sophistication of those stone-age technologies could be measured in terms of the presence of “typically Upper Palaeolithic tool forms such as end scrapers, backed blades, or burins” (Mellars 2006:798) would be futile, and illustrates the poverty of such a measure of “complexity” for early Australian assemblages. Recognizing this methodological difficulty poses the question of how Australian assemblages can be more productively characterised. Rather than use value-laden and ill-defined terms such as “simple” or “complex,” the question required is, Can we develop more informative depictions of the Pleistocene lithic technology? Without addressing that question at length there are two relevant observations that can be made here. The first is that while in terms of traditional typological classifications the variety of implements in early assemblages may appear limited, that must be balanced by an understanding that the retouched flakes in early assemblages were often relatively standardized. For instance, in the Willandra Lakes, which includes Lake Mungo, conventional categories of retouched flakes older than 30,000 years BP display comparatively low variability in size. The length of specimens recorded by Allen (1972) had coefficient of variation values ranging between 20 and 25, and values for width between 20 and 22. In comparison retouched flakes in the same region but later in time, 10,000–30,000 years BP, display consistently higher variability, with the specimens recorded by Allen (1972) having coefficient-of-variation values for length ranging between 22 and 42, and values for width between 26 and 39. These data are offered to make the points: 1) since artifact production in the earliest assemblages was producing more regular tool forms than later in prehistory, claims for the “simplicity” of that production should be closely examined, and 2) Pleistocene Australian assemblages are not unchanging; instead they evolved during all phases of Australian prehistory. Early technologies also displayed geographical differences. The best example of this is the presence of edgeground axes in northern Australia at, and probably well before, 35,000 years BP (Geneste et al. 2010), but not in southern Australia until the Holocene. These specimens are rarely complete and are therefore usually identified from fragments or flakes containing the ground edge, but we have evidence of

Occupying New Lands: Global Migrations and Cultural Diversification with Particular Reference to Australia a number of different production techniques such as biface production accomplished by non-invasive percussion flaking or pecking of a roughly lenticular cross section followed by bifacial grinding of one end. These axes would have been prepared in advance of use, hafted, and maintained over long periods. In short they were part of an elaborate technology for woodworking and other activities. That technology was restricted to the tropical north and perhaps evolved from early waisted but not ground tools known to be used in parts of tropical Sahul (Geneste et al. 2010). The presence of axes only in the north is one of the most obvious indicators of the emergence of regionally different cultural practices in the millennia following colonization. This development of regional distinctions in technology appears to have emerged during or immediately following the expansion of human groups across the continent. Those distinctions are also evident in the evidence for public signaling.

Public Signaling and the Configuration of the Colonizing Societies

The existence, at an early time, of regional-scale difference in public signally is observable in jewelry made 35,000–45,000 years ago. Ornaments, probably in the form of necklaces or bracelets, made of perforated shells or bones with mastic and ocher are found only in the northwestern portion of the continent. Their absence in the east and south is probably conditioned by issues of taphonomy and sample size (see Langley et al. 2011), but it is not a simple consequence of those factors alone since in some localities, especially the Tasmanian uplands, there are large, well-preserved faunal assemblages but no beads. At the very least this indicates regional traditions in the way ornamentation was produced, with only perishable plant materials being used for jewelry in the southeast, and it may well indicate the absence of ornamentation across a substantial portion of the continent in the millennia following settlement. A similar pattern of regional difference exists in the residues of painted-art production. Small ocher fragments have often been recovered from the sediments of occupied caves, often the only visible evidence of parietal art that disappeared long ago, and the changing abundance of ocher in different levels of the deposit may indicate changing intensities of rock painting. This phenomenon is most pronounced in the northern and western portions of the continent and has rarely been reported in the southeast. Furthermore, ocher pallets with ground facets are typical of northern Australia and it may be that paint was prepared in a different way in the south. Ocher was used in the southeast, such as in the burial of WLH3 at Lake Mungo in the southeast, where it was scattered around the interred body before the grave was closed, so we know these regional differences were not the presence or absence of symbol use or ritual, but different expressions of those activities. Hence a range of archaeological indicators reveal different symbolic expressions between north/northwestern regions and south/southeastern ones, and perhaps there are more local traditions that have not yet been defined.

9

Diversification and Regionalization of Cultural ­Activities

The patterns inferred above can be summarized as follows: 1) humans probably arrive well before 50,000 years BP and have spread across much of Sahul by shortly after 50,000 years BP; 2) in some landscapes declines in already stressed megafauna populations are likely to have coincided with the addition of human hunters, and the demise of those taxa may have prompted shifts in foraging practices; and in particular 3) anthropogenic burning strategies were in some environments altered in response to floral changes in lands without large browser/grazers; 4) technologies became regionally diversified as novel production systems and implement forms emerged and long-standing retouched forms were standardized in some locations, with a north/south regional distinctions being clearest; and 5) practices of public signaling, such as the creation of jewelry and painted or drawn art, also became regionally differentiated by about 35,000–40,000 years BP. These cultural differences emerged as human groups settled different environments, not only through a combination of founder effect and drift, but also as a result of the generation of novel practices as groups adjusted their social and cultural systems to historically contingent situations confronting them. Even during the early millennia of settlement it was geographical diversity and cultural adaptation rather than pan-continental uniformity and cultural stability that were the features of human occupation of Sahul. It is difficult to specify the context of these adaptive, diversifying processes given available archaeological evidence. Divergent selection acting on behaviors could promote cultural differentiation between regions over time, though the rate and magnitude of such diversification would depend on a variety of factors including the magnitude of differences in physical and social environments, and the rate of migration and demographic growth as proxies for, and mechanisms mediating, contact between populations. High levels of information flow and cultural linkages (as represented by material and social exchanges including exogamy rates) might constrain (though not prevent) cultural differentiation, while social isolation between dispersing groups might facilitate cultural divergence. Models of fast-dispersing but slow-growing populations advocated by O’Connell and Allen (2012) might suggest minimal cultural connectivity between distant regions in the colonizing phase, possibly creating a context that facilitated the emergence of regional difference in cultural systems. And yet the persistence of broad regional differences inferred here, at least between north or northwest versus south or southeast, indicate more complex mechanisms that maintained traditions and distinctions, and possibly contributed to the rapid growth of both populations and information networks.

10

Conclusion

This summary of the Australian archaeological evidence demonstrates that following human colonization, cultural variation was maintained and increased rather than diminished as range expansion proceeded. This process of broad-scale cultural diversification draws attention to the question of whether there are invariant material or behavioral markers of modernity and H.sapiens as many models of dispersal and species transition have presumed. Increasingly it seems that at a global level dispersing populations of Homo sapiens are not usefully characterized by a single or even a simple set of artifacts or behaviors. Instead, it is more productive to think of the unifying character of this migration as a consequence of the capacity of human groups to adapt social, economic, and technological activities to the different contexts they encountered. In the context of Australia it is clear that we have evidence of a dynamic process of social, economic, and technological diversification associated with the spread of humans and their adaptation to new social and physical environments. Such evidence reveals complex interactions of factors of adaptation and isolation during the dispersion. The cultural variation that emerged in Australia was not the result of conservative maintenance of social tradition or a diminished set of cultural practices following serial founder events, but a radiation of cultural practices and perspectives as an adaptive response to the occupation of new lands.

References Cited

Allen, H.  1972  Where the crow flies backwards. Ph.D. thesis, Australian National University. Allen, J., and J. F. O’Connell  2003  The long and the short of it: Archaeological approaches to determining when humans first colonised Australia and New Guinea, Australian Archaeology 57:5–19. ———  2008  Getting from Sunda to Sahul. In Islands of Inquiry: Colonization, Seafaring and the Archaeology of Maritime Landscapes, edited by G. Clark, F. Leach, and S. O’Connor, pp. 31–46. ANU EPress, Australian Natinal University. Balme, J.  1980  An analysis of charred bone from Devil’s Lair, Western Australia. Archaeology and Physical Anthropology in Oceania 15: 81–5. Beaton, J, M.  1991  Colonizing continents: Some problems from Australia and the Americas. In The First Americans: Search and Research, edited by T. D. Dillehay and D. J. Meltzer, pp. 209-230. CRC Press, Boca Raton. Birdsell, J. B.  1957  Some population problems involving Pleistocene Man. Cold Spring Harbor Symposia on Quantitative Biology 22:47– 69. Boivin, N., D. Q. Fuller, R. Dennell, R. Allaby, M. D. Petraglia  in press  Human dispersal across diverse environments of Asia during the Upper Pleistocene, Quaternary International. Bowdler, S.  1977  “The coastal colonization of Australia,” In Sunda and Sahul: Prehistoric studies in Southeast Asia, Melanesia and Australia, edited by J. Allen, J. Golson, and R. Jones, pp. 205–46, Academic Press, London. Bowler, J., H. Johnston, J. Olley, J. R. Prescott, R. G. Roberts, W. Shaw-

Hiscock cross, and N. A. Spooner  2003  New ages for human occupation and climatic change at Lake Mungo, Australia. Nature 421:837–40. Brook, B. W., and D. M. Bowman  2002  Explaining the Pleistocene megafaunal extinctions: Models, chronologies, and assumptions. Proceedings of the National Academy of Science 99:624–27. ———  2004  The uncertain blitzkrieg of Pleistocene megafauna. Journal of Biogeography 31:517–23. Brumm, A., and M. W. Moore  2005  Symbolic revolutions and the Australian archaeological record. Cambridge Archaeological Journal 15:157–75. Clarkson, C.  In Press  East of Eden: Founder effects and the archaeological signature of modern human dispersal. In East of Africa: Southern Asia, Australia and Human Origins, edited by M. Porr and R. Dennell. Cambridge University Press. David, B., J-M. Geneste, R. L. Whear, J-J. Delannoy, M. Katherine, R. G. Gunn, C. Clarkson, H. Plisson, P. Lee, F. Petchey, C. Rowe, B. Barker, L. Lamb, W. Miller, S. Hoerlé, D. James, E. Boche, K. Aplin, I. J. McNiven, T. Richards, A. Fairbairn, and J. Matthews  2011  Nawarla Gabarnmang, a 45,160 cal BP site in Jawoyn country, southwest Arnhem Land Plateau. Australian Archaeology 73:73–77. Faivre, J-P.  2012  A material anecdote but technical reality: Bladelet and small blade production during the recent middle Paleolithic at the Combe-Grenal Rock Shelter, Lithic Technology 37:5–25. Feranec, R. S., N. G. Miller, J. C. Lothrop, and R. W. Graham  2011 The Sporormiella proxy and end-Pleistocene megafaunal extinction: A perspective. Quaternary International 245-333–38. Field, J. and S. Wroe 2012 Aridity, faunal adaptations and Australian Late Pleistocene extinctions. World Archaeology 44:56–74. Field, J., M. Filios, and S. Wroe  2008  Chronological overlap between humans and megafauna in Sahul (Pleistocene Australia-New Guinea): A review of the evidence. Earth-Science Reviews 89:97–115. Fifield, L. K., M. I. Bird, C. S. M. Turney, P. A. Hausladen, G. M. Santos, and M. L. di Tada  2001  Radiocarbon dating of human occupation of Australia prior to 40 ka BP – successes and pitfalls. Radiocarbon 43:1139–45. Flannery, T. F.  1990  Pleistocene faunal loss: Implications of the aftershock for Australia’s past and future. Archaeology in Oceania 25:45–67. Franklin, N., and P. Hapgood  2007  Modern human behavior and Pleistocene Sahul in review. Australian Archaeology 65:1–16. Geneste, J-M., B. David, H. Plisson C. Clarkson, J-J Delannoy, F. Petchey, and R. Whear  2010  Earliest evidence for ground-edge axes: 35,400 ± 410 cal BP from Jawoyn country, Arnhem Land. Australian Archaeology 71:66–69. Hapgood, P. J., and N. Franklin  2008  The revolution that didn’t arrive: A review of Pleistocene Sahul, Journal of Human Evolution 55:187–222. Henrich, J.  2004  Demography and cultural evolution: How adaptive cultural processes can produce maladaptive losses–the Tasmanian case. American Antiquity 69:197–214. Hiscock, P.  2008  Archaeology of Ancient Australia. Routledge, London. Hiscock, P., and S. O’Connor  2006  An Australian perspective on modern behavior and artifact assemblages, Before Farming [online version] 2006/1 article 5. Johnson, C. R., and S. McBrearty  2010  500,000 year old blades from the Kapthurin Formation, Kenya. Journal of Human Evolution 58:193–200.

Occupying New Lands: Global Migrations and Cultural Diversification with Particular Reference to Australia Kershaw, P., S. van der Kaars, P. Moss, B. Opdyke, F. Guillard, S. Ruleand, and C. Turney  2006  Environmental change and the arrival of people in the Australian region. Before Farming (online version) 2006/1 article 2. Langley, M. C., C. Clarkson, and S. Ulm  2011  From small holes to grand narratives: The impact of taphonomy and sample size on the modernity debate in Australia and New Guinea. Journal of Human Evolution 61:197–208. Mellars, P.  2006  Going East: New genetic and archaeological perspectives on the modern human colonization of Eurasia. Science 313(5788):796–800. Miller, G. H., J. W. Magee, B. J. Johnson, M. L. Fogel, N. A. Spooner, M. T. McCulloch, and L. K. Ayliffe  1999  Pleistocene extinction of Genyornis newtoni: Human impact on Australian megafauna. Science 283:205–08. O’Connell, J. F., and J. Allen  2004  Dating the colonization of Sahul (Pleistocene Australia-New Guinea): A review of recent research. Journal of Archaeological Science 31:835–53. ———  2007  Pre-LGM Sahul (Pleistocene Australia-New Guinea) and the Archaeology of Early Modern Humans. In Rethinking the Human revolution, edited by P. Mellars, K. Boyle, O. Bar-Yosef, and C. Stringer, pp. 395–410. McDonald Institute for Archaeological Research, Cambridge University. Cambridge. ———  2012  The Restaurant at the End of the Universe: Modelling the colonization of Sahul. Australian Archaeology 74:5–17. O’Connell, J. F., J. Allen, and K. Hawkes  2010  Pleistocene Sahul and the Origins of Seafaring. In The global origins and development of seafaring, edited by A. Anderson, J. Barrett, and K. Boyle. McDonald Institute for Archaeological Research, Cambridge University, Cambridge. O’Connor, S., R. Ono, and C. Clarkson  2011  Pelagic fishing at 42,000 BP and the maritime skills of modern humans. Science 334:117–21. Richardson, N.  1992  Conjoin sets and stratigraphic integrity in a sandstone shelter: Kenniff Cave (Queensland, Australia). Antiquity 66:408–18. ———  2011  Site formation processes, Kenniff Cave, Queensland. Masters thesis, Australian National University. Rindos, D., and E. Webb  1992  Modelling the initial human colonization of Australia: Perfect adaptation, cultural variability and cultural change, Proceedings of the Australasian Society of Human Biology 5:441–54.

11

Roberts, R., R. Jones, and M. Smith  1993  Optical dating at Deaf Adder Gorge, Northern Territory, indicates human occupation between 53,000 and 60,000 years ago. Australian Archaeology 37:58–9. Roberts, R. G., T. F. Flannery, L. K. Ayliffe, H. Yoshida, J. M. Olley, G. J. Prideaux  2001  New ages for the last Australian megafauna: Continent-wide extinction about 46,000 years ago. Science 292:1888–92. Rule, S. B. W. Brook, S. G. Haberle, C. S. M. Turney, P. Kershaw, and C. N. Johnson  2012  The aftermath of megafaunal extinction: Ecosystem transformation in Pleistocene Australia. Science 335:1483–86. Shawcross, W.  1998  Archaeological excavations at Mungo. Archaeology in Oceania 33:183–200. Smith, M. A., M. J. Bird, C. S. M. Turney, L. K. Fifield, G. M. Santos, P. A. Hausladenand, and M. L. di Tada  2001  New ABOX AMS-14C ages remove dating anomalies at Puritjarra rock shelter. Australian Archaeology 53:45–7. Stockton, E. D.  1973  Shaw’s Creek Shelter: Human displacement of artifacts and its significance. Mankind 14:112–17. Surovell, T. A., and B. S. Grund  2012  The associational critique of Quaternary overkill and why it is largely irrelevant to the extinction debate, American Antiquity 77:672–88. Surovell, T. A., J. B. Finely, G. M. Smith, P. J. Brantingham, and R. L. Kelly  2009  Correcting temporal frequency distributions for taphonomic bias. Journal of Archaeological Science 36:1715–24. Turney, C. S. M., M. I. Bird, L. K. Fifield, R. G. Roberts, M. Smith, C. E. Dortch  2001  Early human occupation at Devil’s Lair, southwestern Australia, 50,000 years ago. Quaternary Research 55:3–13. Veth, P., M. Smith, J. Bowler  2009  Excavations at Parnkupirti, Lake Gregory, Great Sandy Desert: OSL ages for occupation before the Last Glacial Maximum. Australian Archaeology 69:1–10. White, J. P.  1977  Crude, colourless and unenterprising: Prehistorians and their views on the stone age of Sunda and Sahul. In Sunda and Sahul: Prehistoric Studies in Southeast Asia, Melanesia and Australia, edited by J. Allen, J. Golson, and R. Jones, pp.13–30. Academic Press, London. Williams, A. N.  2013  A new population curve for prehistoric Australia. Proceedings of the Royal Society B 280. Wroe, S., and J. Field  2006  A review of the evidence for a human role in the extinction of Australian megafauna and an alternative interpretation. Quaternary Science Reviews 25:2692–2703.

Chapter 2 Human Habitation in Arctic Western Beringia Prior to the LGM Vladimir Pitulko1, Pavel Nikolskiy2, Aleksandr Basilyan2, and Elena Pavlova3

ABSTRACT For years, the initial stage of human habitation within western Beringia was supposed to be not older than the Late Upper Paleolithic, with firm dates younger than the Last Glacial Maximum (LGM). Discovery of Yana RHS doubled the length of the record of human habitation in NE Asia. Human occupations at the Yana site pre-date the LGM and show that the area was inhabited almost 30,000 14C years ago, providing the earliest evidence for human habitation known in the Arctic. The site yielded unique evidence for Early Upper Paleolithic culture in this remote part of the world. Faunal remains that come from the site belong to almost all species of the local Late Pleistocene habitat. Reindeer, bison, and horse are most numerous. Three major contexts compose the Yana archaeological complex. Two of them are lithic contexts called correspondingly “macro tools” (cores, scrapers, large tools) and “micro tools” (small scrapers, chisel-like pieces, backed blades, but almost no burins). The third one comprises a well-developed bone/ivory industry that includes hunting equipment, sewing toolkits, and other implements. Numerous personal ornaments and decorated artifacts demonstrate highly developed complicated symbolic behavior. This article presents the data on geology, radiocarbon dating, and artifact collections of the Yana site. KEYWORDS: Arctic western Beringia, Northeast Asia, Upper Pleistocene, Upper Paleolithic, Human mgrations

Introduction

For a long time Northeast (NE) Asia remained almost unexplored in terms of archaeology. The archaeological record for Paleolithic habitation in NE Asia was significantly increased in the early 1960s and 1970s. At that time the Ushki archaeological complex was discovered by Dikov (1965) on the Kam­ chatka Peninsula, followed by findings in Aldan (Mochanov 1969) and at Berelekh (Vereschagin and Mochanov 1972). These discoveries, made within just 10 years, raised hopes 1

Institute for the Material Culture History, Russian Academy of Sciences, 18 Dvortsovaya nab., St Petersburg, 191186, Russia. 2 Geological Institute, Russian Academy of Sciences, Pyzhevsky per. 7, 119017 Moscow, Russia. 3 Arctic & Antarctic Research Institute, 38 Bering St., St Petersburg, 199397, Russia. Corresponding author e-mail: [email protected]

for fast progress in expanding the record of Paleolithic human habitation in western Beringia. However, for decades it remained restricted to those sites by time and space (Figure 2.1). Within the past two decades a number of presumably Paleolithic but undated sites were discovered on the Chukotka Peninsula as well as at several pre-Holocene sites in the Upper Kolyma and western Chukotka regions (Dikov 1993; Kiryak at el. 2003; Slobodin 1999). More of Late Paleolithic archaeological materials have been also found in the Lena River system, Khairgas Cave in Middle Lena (Stepanov et al. 2003), and especially in the Vitim drainage (Ineshin and Teten’kin 2010). All these sites have often been reviewed and discussed from the perspective of their importance for the peopling of the New World through the Bering Land Bridge (Dikov 1977, 1979; Mochanov 1977; West 1996; Bonnichsen and Turnmire 1999; Goebel and Buvit 2011). Generally speaking, this dataset postdates the Last Glacial Maximum (LGM)

14

Pitulko et al.

Figure 2.1  Radiocarbon-dated western Berin­ gia Upper Paleolithic sites (for the Kheta site the age estimate is given according to Slobodin, 1999).

and spans 18,000–17,000 radiocarbon years before present (14C yr BP) at best. Even by the end of the 1980s it was crystal clear that this is a very modest background on which to draw well-based conclusions. In addition, the Ushki site complex was redated and appeared to be even younger than it was thought (Goebel et al. 2003; 2010) and then any potential ancestor for the North American Stone Age disappeared in NE Asia from the areas adjacent to the Bering Land Bridge. Thus, by the beginning of the 21st century it might be concluded that for this vast territory almost nothing of the Paleolithic of NE Asia remained really known except for temporary human habitation episodes within the terminal Pleistocene. The question of LGM and pre-LGM habitation remained totally unanswered. The discovery of Yana RHS changed this situation, doubling the length of time known for human habitation in western Beringia.

Yana RHS Site Location

Yana RHS was discovered in 2001 (Pitulko et al. 2004). It is situated in the lower Yana River at 70° 43′ N latitude and 135° 25′ E longitude (Figure 2.2) in the westernmost portion of the extensive coastal lowland that begins with the Yana River in the west and extends to the Kolyma River in the east within the limits of present Arctic Siberia. In terms of paleo-

geography, this is the westernmost part of western Beringia. The site is located within the permafrost zone. Vegetation of this area is typical for the border of southern hypo­ arctic tundra and open northern woodland area (Atlas of the Arctic 1985; CAVM 2003; Yurtsev et al. 1978), including tundra landscapes with associations of tussock sedge, dwarf shrub, and moss tundra. Generally, the area is moist tundra, dominated by tussock cottongrass (Eriophorum ­vaginatum) and dwarf shrubs; scant Betula exilis shrub vegetation coexists with grassy hypnum bog vegetation and scarce larch trees (Larix gmelinii) with suffruticous plants (Vaccinium vitis-idaea, Arctous alpine, Empetrum nigrum, and Ledum palustre), lichen, and mossy river valley open woodlands. This area is located within the eastern part of the transitional climatic region of the coastal climate of the Arctic Siberian climatic zone (Atlas of the Arctic 1985; Gakkel and Korotkevich 1960). Annual average temperature varies within -13.9 to -14.2°C, mean January temperature is -37.4 to -38°C, mean July temperature reaches +11 to +11.5°C (Izyumenko 1966, 1968). This is harsh cold and dry environment with a short frostless period (57 days per year). Average precipitation varies from 200 to 240 mm per year. Where the site is located, the Yana River makes a near 90-degree turn and runs roughly from west to east, and then turns again and flows north (Figure 2.2). In this part of the

Human Habitation in Arctic Western Beringia Prior to the LGM

15

Localities that retain cultural material in situ —Yana B, Northern Point (NP), and TUMS 1

Position of surface concentrations of the material

Localities that yielded surface finds—ASN, Upstream Point, and Southern Point locality adjacent to Yana mass accumulation of mammoth (YMAM)

Part of the profile that is out of picture

Position for geology profiles

Presumed position of the top of T2 body near the Upstream Point (UP) locality

Position of the cultural layer

Figure 2.2  Yana RHS site location and relationship between its structural elements. A, schematic map of the area (a fragment of Google Earth satellite image is used); B, schematic geomorphology profile. Height is given as a.w.l. (above water level). T1, terrace 1 (10–11 m a.w.l.); T2, terrace 2 (16–18 m a.w.l.); T3, terrace 3 (40–45 m a.w.l.).

river valley, abandoned terraces have been preserved only as fragments. At the present time the 3rd terrace level (T3) is 40–45 m, the 2nd terrace (T2) is 16–18 m, and the 1st terrace (T1) is at 10–11 m above water level (a.w.l.). All geomorphological surfaces lower than 9–10 m a.w.l. are within the modern or historical floodplain. Permafrost deposits with high ice content and a polygonal grid of syngenetic ice wedges are present in three terrace bodies. These ice wedges span several generations (Pitulko et

al. 2007; Basilyan et al. 2009, 2011). Ice content varies from 30% to 70%. Such icy deposits are termed ice-complex deposits (Romanovskiy 1993) and are widespread within western Beringia. Sediments of this character, which have yielded excellent artifact preservation, are highly sensitive to natural agents such as running water, summer insolation and atmospheric heat that cause erosional processes and therefore demand a specific determined excavation strategy (Pitulko 2008; Pitulko and Pavlova 2012).

16

Pitulko et al.

Site Structure, Geology, and Dating

Although Yana RHS is termed a site, it is indeed a complex of geoarchaeological objects. These localities seem to represent separate but roughly contemporaneous archaeological sites. The site structure includes several known separate localities (at least seven of them) discovered within the T2 body (Figure 2.2). Their state of preservation varies considerably. Two localities, known as ASN (named after S. N. Astakhov) and the Upstream Point, are totally washed out and have yielded surface finds only. The Southern Point produced plenty of fragmented bones, with a high number of mammoth remains, and numerous lithic artifacts collected on the river bank. Yana Mass Accumulation of Mammoth (YMAM), which is also a part of the spatial structure of Yana RHS, was found next to it in 2008 (Basilyan et al. 2011). By its geology, that concentration of mammoth bones belongs to the deposits of a paleochannel that existed when Yana RHS was inhabited. Potentially

there is a portion of the Yana RHS occupation that remains in this area adjacent to YMAM. However, to date all efforts to locate it have been unsuccessful. The other three, TUMS 1 (named after V. E. Tumskoy), Northern Point, and Yana-B, have a well-preserved in situ cultural layer that belongs to the middle portion of the T2 body. All these localities are actually separate “sites” that compose the Yana RHS “site complex.” In T2, which bears the cultural layer of the Yana RHS, four geological members can be distinguished (Figure 2.3). These members are separated by stratigraphic unconformities and erosional surfaces that correspond to extensive breaks of sedimentation. The basal, erosional part of T2 (Figure 2.3: Members 1–2) occurs only in the upstream area of the exposure and is proposed to date to the early- to late-Pleistocene transition (Basilyan et al. 2009, 2011). This ancient alluvium is a remainder of the T3 fill that was partially side cut and down cut by T2 when it was formed.

A

B Bed number within Member 3

Cultural layer in situ

Ice-complex deposits of different geological age, with syncryogenic ice wedges in alluvial and proluvial deposits

Location of cryolithological profiles

Polygonal ice wedges

Sand with small pebbles

Eolian deposits

Location of bones of Pleistocene animals sampled for 14C dating

Sandy silt

Cross bedding

Clayey-sand silt

Ice-wedge cast

Sandy-clayey silt

Conglomerates

Location of 14C samples within cryolithological profiles

Interbedding of clayey silt bands and sandy-clayey silt with beds and lenses of peat

Radiocarbon date and lab code

Position of Upper Paleolithic Yana RHS cultural layer Number for geological member

Active layer

Location of 14C samples on the profile (plant remains dated)

Terrace edge (schematic) Baidzharakh—cone-shaped land forms resulting from partial thawing of polygonal ice-wedge net

Figure 2.3  Geology of Yana RHS site area. A, Soplivaya Gora (Yana-195) reference profile of Quaternary deposits; B, cryolithology for different localities within the Yana RHS – YMAM, Yana B, Northern Point, TUMS 1.

Human Habitation in Arctic Western Beringia Prior to the LGM

17

Table 2.1  Description of the upper (accumulative) part of Terrace II from bottom to top (Fig. 2B: Bed 1-5): Bed



1

2

3

Description

Thickness (m)

Member 1 4–5 Alluvium-lacustrine and thaw-lake deposits with icewedge casts; these deposits were formed in the end of Early Pleistocene and Middle Pleistocene. It expressed locally within a portion of T2 near its junction with T3 (Figure 21.3). Member 2 5–7 Alluvium-lacustrine and eolian deposits of Early Late Pleistocene age are developed within the area next to T2 and T3 junction. Member 3 5.2 Late Pleistocene alluvium Silt of greenish gray color with high ice content that increases from the bottom to the top and reaches 90%; spotted bottom part has spots of brownish color resulting from geochemical reactions; the upper part often has peat hummocks that can be up to 0.5 m wide. Squamate belt-like cryo textures are typical. Some of the schliren are up to 8 cm thick, curved up next to ice-wedge margins. Silt of light greenish gray color, sandy silt of brownish gray color, lenses of sand with sandy silt fill. Initial lamination is heavily damaged by cryo-turbations in the fossil seasonally melted layer and looks like verti0.6 cal linguiform interosculations of one sort of silt into another one. Squamate cryo textures are represented by 2- to 3-cm-thick sub-horizontal schliren, which can be up to 30 cm wide. The bottom of that layer cuts off cryogenic textures of underlying deposits. This border is traced by the exposures for about one kilometer Silt of brownish gray color, massive, with unclear wavy cleavage and numerous filiform roots of grass plants. Cryo textures are massive and occasionally in the form of thin schliren oriented according to 1.5–1.7 lamination direction. They are 2 m wide; their thickness varies from 2 to 5 mmt. In the middle part of that horizon, at 1.2–1.7 m above its bottom, there are fragmented and intact bones of Pleistocene animals and artifacts (cultural layer). Then, closer to the rear part of the terrace, the facies change from floodplain deposits to channel deposits. In addition, the lowering of the stratum to the water level and its breakout into underlying deposits become clearly visible. Here, these deposits fill a narrow erosion channel, 40 m wide, cut by the old stream. The fill is represented by cross-bedded sandy silt with diverse

Member 3 (Figure 2.3: Bed 1–4) and Member 4 compose the actual fill of the T2 body (Figure 2.3, see also Table 2.1). Alluvial deposits of T2 started to accumulate shortly after 40,000 years ago. Alluvial sedimentation forming its body stops at around 13,000 14C yr BP (when T1 starts building up) and changes to eolian sedimentation. At the top of T2 sequence are young erosional cuts filled with terminal-Pleistocene and Holocene deposits (Member 4). Bed 3 of Member 3, which constitutes the most typical alluvial body of T2, bears the cultural layer of Yana RHS. It occurs 7.5 m above the average summer water level and is overlain by 8–11 m of frozen sediments. It is easily recognized by bright reddish brown color in natural bank exposures between excavated portions of the site. Member 3 is composed of sandy-silt and silty sediments

Bed

Description

Thickness (m)

grain sand, gravels, and pebbles in the bottom, with lenses of grass plant macro remains. On the eastern side of this channel, a series of sandy silt lenses with bones of large animals and hair wisps were found. The bone-bearing lens is 20 m long and about 1.5 m thick. Its middle part consists almost exclusively of bone remains of Pleistocene fauna, mostly mammoths. Along with the bones, lithic artifacts, debitage, and ivory flakes were found here. 4



In this part of the profile, ice wedges up to 3 m wide 7.5–9.5 are developed. They make a polygonal pattern, each side 15–20 m long. Some of the ice wedges have terminations inside the stratum; some are cut down by overlying deposits; some have grown inside the overlying deposits. Silt and sandy silt sediments are of yellowish gray color with slightly visible inclined bedding. The bottom part of this stratum consists of sand with silt filling that also has sparse gravel and pebbles and rare bones of Pleistocene animals. It overlies the underlying deposits with a wash-out type of contact. An inter-bed of massive aleurolites of brownish color (0.2 m thick) is found in the middle part of the stratum. It is characterized by massive cryo textures; in the bottom part of the stratum, there appear belt-like cryo textures that follow the bedding pattern of the sediments. Ice wedges (up to 2 m wide) are terminated by overlying deposits. Member 4 ≤ 8 Alluvial and proluvial deposits, Terminal Pleistocene and Holocene. Deposits of this member consist of fine-grained sand with silt filling of gray color—cross-bedded, shoestring sands that fill erosion channels. Numerous unconformities these deposits are the result of cutting and filling during Terminal Pleistocene and Holocene. They are enriched with a large number of plant remains in a form of allochtonic concentrations and hummocks of grass, roots, twigs, and chunks of driftwood trunks. Ice wedges (up to 1.2 m wide) are well developed within this stratum. Sometimes these wedges are terminated on the boundary of the lenses. Massive, thick schliren cryo textures characterize these deposits. Ice schliren (up to 8 mm thick and 1 m long) are occasionally oriented in accordance with lamination.

whose sand content changes within the profile, decreasing from bottom to top. Beds 1 and 2, which form the bottom of the T2 profile, have higher sand and ice content than Beds 3 and 4, which are composed mostly of fine-grained silt. This is a typical pattern of alluvial sedimentation, with no visible significant breaks. It reflects the formation of these deposits under conditions of floodplain terrace that gradually changed to high floodplain terrace sedimentation before the cultural remains were deposited, and then remained under the same conditions until the alluvial sedimentation ended 14,000– 13,000 14C yr BP (Figure 2.3, Table 2.2). Multiple radiocarbon dating of cultural remains and sediments delivered a solid set of 14C dates to support the chronology of Yana RHS (Figures 2.3 and 2.4, see also Table 2.2). The age of the cultural material in general is deter-

18

Pitulko et al.

mined to be in the interval 28,900–26,900 14C yr BP by a series of 14C dates obtained on faunal remains and by direct dating of organic artifacts and hearth mass (Table 2.2, Figure 2.4). They are additionally controlled by dates of the underlying and overlying deposits (Figures 2.3 and 2.4; see also Table 2.2). Dates obtained on materials from collections correspond to each other, to the age of the layer, and to their geological position. Four individual areas (structural elements) of the Yana site are radiocarbon dated, that is, all the areas for which this is possible—TUMS 1, Northern Point, Yana B, and YMAM/Southern Point locality. Thus the

14

C date, deviation 1 sigma

In situ 14C dates of the cultural layer

chronological position of this archaeological material, the earliest in NE Asia, is consistent. The dates obtained on various materials (bone collagen, plant macro-remains, humic acids, charcoal, charred material from hearth features) are satisfactorily consistent with each other and also are revealing. Though the age determinations of plant organics, charcoal, and bone materials correspond on the whole to the picture observed, they show a notable spread. It was previously found that there are two potential chronologically different phases. One extends to the lower chronological boundary of occupation of the site (~28,500–

14

C dates 28,000–27,000 BP

“Young” 14C dates 26,900–25,100 BP (habitation phase 3).

Group of “old” dates (habitation phase 1); 28,900–28,000 BP, habitation phase 2

Figure 2.4  Distribution of radiocarbon dates obtained for different structural parts of the Yana RHS site. Chronological relationship is shown for all localities that retain cultural layer in situ (Yana B, Northern Point, TUMS 1), Yana mass accumulation of mammoth (YMAM), and for the occupation level in between. 14C dates are not calibrated; for each date the instrumental error is shown. See Table 21.2 for details on radiocarbon dates.

Human Habitation in Arctic Western Beringia Prior to the LGM

19

Table 2.2  Radiocarbon chronology of the Yana RHS geoarchaeological objects. No 14C date Laboratory no. Sediments that overlie the YMAM 1 15,910 ± 110 LU-5973 2 17,710 ± 140 LU-5968 3 18,550 ± 180 LE-8492 4 18,750 ± 100 Beta-250677 5 20,150 ± 120 Beta-250676 6 21,010 ± 500 LE-8502 7 21,220 ± 100 Beta-250661 8 21,570 ± 100 Beta-250664 9 21,580 ± 400 LE-8509 10 21,640 ± 250 LE-8510 11 22,400 ± 110 Beta-250663 12 23,230 ± 110 Beta-250662 13 23,330 ± 150 Beta-250640 14 23,450 ± 160 Beta-250639

Material dated

Reference

Grass and moss Bone collagen from woolly rhinoceros humerus Bone collagen from mammoth scapula Plant remains Plant remains Plant macro remains/grass Plant macro remains/grass Plant macro remains/grass Plant macro remains/grass Plant macro remains/grass Plant macro remains/grass Plant macro remains/grass Bone collagen from Pleistocene bison vertebra Bone collagen from caribou antler

Basilyan et al. 2011 Basilyan et al. 2011 Basilyan et al. 2011 Basilyan et al. 2011 Basilyan et al. 2011 Basilyan et al. 2011 Basilyan et al. 2011 Basilyan et al. 2011 Basilyan et al. 2011 Basilyan et al. 2011 Basilyan et al. 2011 Basilyan et al. 2011 Basilyan et al. 2011 Basilyan et al. 2011

Yana mass accumulation of mammoth (YMAM) and Southern Point (SP) locality (combined) 25,100 ± 1000 LE-8644 Bone collagen from mammoth mandible 15 16 25,800 ± 600 GIN-11465 Bone collagen from reindeer bone fragment 17 26,650 ± 200 Beta-250672 Collagen from woolly rhinoceros horn (direct dating of the foreshaft from bone accumulation at YMAM) 18 27,200 ± 1200 LE-8572 Bone collagen from mammoth mandible 19 27,400±600 GIN-11466 Bone collagen from mammoth mandible (juvenile or female individual by size) 20 27,600 ± 600 LE-8650 Bone collagen from mammoth mandible 21 27,600 ± 500 GIN-11467 Bone collagen from Pleistocene horse bone 22 27,740 ± 200 LE-8508 Plant macro remains/grass 23 27,800±500 GIN-11464 Bone collagen from mammoth tusk fragment (burnt fragment) 24 28,200 ± 400 LE-8568 Bone collagen from mammoth mandible 25 28,400 ± 430 LE-8565 Bone collagen from mammoth mandible 26 28,470 ± 210 Beta-257535 Bone collagen from brown bear limb Bone 27 28,600 ± 800 LE-8574 Bone collagen from mammoth mandible 28 28,900 ± 900 LE-8573 Bone collagen from mammoth mandible 29 31,200 ± 1200 LE-8569 Bone collagen from mammoth mandible Sediments that underlie the YMAM 30 31,500 ± 500 LE-8498

Peat

Basilyan et al. 2011 Pitulko et al. 2004 this publication Basilyan et al. 2011 Pitulko et al. 2004 Basilyan et al. 2011 Pitulko et al. 2004 Basilyan et al. 2011 Pitulko and Pavlova 2010 Basilyan et al. 2011 Basilyan et al. 2011 Basilyan et al. 2011 Basilyan et al. 2011 Basilyan et al. 2011 Basilyan et al. 2011 Basilyan et al. 2011

Reddish brown horizon (paleosol?) between SP locality and the Yana B area (corresponds to the stratigraphic position of cultural layer at ca. 7.5 m a.w.l.) 31 27,500 ± 210 Beta-250646 Bone collagen from Pleistocene horse mandible Basilyan et al. 2011 32 27,500 ± 350 LE-8471 Bone collagen from mammoth vertebra Basilyan et al. 2011 33 27,700 ± 270 LE-8808 Ivory Basilyan et al. 2011 34 27,970 ± 210 Beta-250638 Bone collagen from mammoth rib Basilyan et al. 2011 35 28,030 ± 160 Beta-250636 Bone collagen from reindeer mandible Basilyan et al. 2011 Yana B locality—cultural layer 36 27,670 ± 210 Beta-250634 37 27,910 ± 280 Beta-191323 38 28,060 ± 210 Beta-250637 39 28,210 ± 200 Beta-250635 40 28,250 ± 200 Beta-250633 

1

Bone collagen from Pleistocene horse pelvic bone Bone collagen from Arctic fox tibia Bone collagen from Pleistocene bison metacarpal bone Bone collagen from mandible Bone collagen from reindeer metacarpal bone

Pitulko and Pavlova 2010 Pitulko and Pavlova 2010 Basilyan et al. 2011 Pitulko and Pavlova 2010 Pitulko and Pavlova 2010

Date from artifacts collected next to the bank exposure of the cultural layer; other NP dates are from in situ position. Laboratory codes:: Beta, Beta Analytic, Inc. (Miami, Florida, USA); GIN, Geological Institute, Russian Academy of Sciences, Moscow; LE, Institute for the History of Material Culture, Russian Academy of Sciences (St. Petersburg, Russia); LU, St. Petersburg State University (St. Petersburg, Russia). All dates by Beta Analytic are 14C AMS; GIN, LU and LE dates are standard radiometric (conventional 14C) dates.

Pitulko et al.

20 Table 2.2 Cont’d. No

14

41

28,350 ± 250

C date

Laboratory no.

Material dated

Reference

Beta-204881

Bone collagen from reindeer metatarsal bone

Pitulko and Pavlova 2010

Northern Point locality—sediments that underlie the cultural layer 42 10,590 ± 300 LE-7615 Peat 43 11,950 ± 70 Beta-223406 Plant remains 44 14,010 ± 80 Beta-243115 Plant remains 45 17,970 ± 100 Beta-243116 Plant remains 46 19,770 ± 100 Beta-243117 Plant remains 47 22,290 ± 150 Beta-204858 Plant remains 48 26,450 ± 160 Beta-191331 Plant remains

Pitulko and Pavlova 2010 Pitulko and Pavlova 2010 Pitulko and Pavlova 2010 Pitulko and Pavlova 2010 Pitulko and Pavlova 2010 Pitulko and Pavlova 2010 Pitulko et al. 2007

Northern Point locality—cultural layer 49 26,680±160 Beta-191334 50 27,140 ± 180 Beta-191321 51 27,200 ± 2400 LE-7668 52 27,250 ± 230 Beta-223413 53 27,440 ± 210 Beta-162233 54 27,510 ± 180 Beta-191332 55 27,620 ± 240 Beta-204863 56 27,820 ± 190 Beta-191328 57 27,890 ± 190 Beta-191335 58 27,900 ± 200 Beta-191333 59 28,000 ± 190 Beta-191329 60 28,090 ± 200 Beta-191327 61 28,250 ± 170 Beta-173064 62 28,500 ± 200 Beta-191326 63 28,570 ± 300 Beta-191322 64 29,130 ± 410 Beta-204864

Pitulko and Pavlova 2010 Pitulko et al. 2007 Pitulko and Pavlova 2010 Pitulko and Pavlova 2010 Pitulko et al. 2004 Pitulko and Pavlova 2010 Pitulko and Pavlova 2010 Pitulko et al. 2007 Pitulko and Pavlova 2010 Pitulko and Pavlova 2010 Pitulko et al. 2007 Pitulko and Pavlova 2010 Pitulko and Pavlova 2010 Pitulko and Pavlova 2010 Pitulko et al. 2007 Pitulko and Pavlova 2010

Burnt bone fragment Bone collagen from musk ox metacarpal bone Charred materal from hearth Charcoal Collagen from woolly rhinoceros horn foreshaft1 Plant remains Bone collagen Plant remains Plant remains Charred material from the hearth Plant remains Bone collagen from Pleistocene bison phalange Mammoth ivory artifact/foreshaft1 Pleistocene bison hoof phalange Pleistocene hare humerus Soot

Sediments that underlie the cultural layer at the Northern Point locality 65 29,610 ± 230 Beta-191330 Plant remains 66 33,220 ± 520 Beta-204873 Plant organic material 67 34,820 ± 620 Beta-204875 Plant organic material

Pitulko et al. 2007 Pitulko and Pavlova 2010 Pitulko and Pavlova 2010

Sediments that overlie the cultural layer of the TUMS 1 locality 68   8960 ± 80 LE-6447 Peat 69 18,100 ± 340 LE-6445 Plant remains/ twigs 70 22,400 ± 300 LE-6446 Plant remains 71 25,900 ± 750 LE-6444 Plant remains

Pitulko et al. 2004 Pitulko et al. 2004 Pitulko et al. 2004 Pitulko et al. 2004

TUMS 1 locality of the Yana RHS (cultural layer) 72 27,300 ± 270 Beta-173067

Bone collagen from horse mandible

Pitulko et al. 2004

Radiocarbon dates shown on Figure 3 only Terrace 3 deposits 73 > 29,000 LE-6002 74 > 32,000 LE-6027 75 > 45,000 GIN-11696 76 > 45,000 GIN-11697

Plant remains Plant remains Bone collagen (Pleistocene horse) Bone collagen (Pleistocene bison)

Pitulko et al. 2004 Pitulko et al. 2004 Pitulko et al. 2004 Pitulko et al. 2004

Holocene deposits insert into T2 body (backfilling of erosion channels cut into the terrace body from above) 77 3010 ± 40 LE-6000 Peat 78 4350 ± 30 LE-6024 Wood

1

Pitulko et al. 2004 Pitulko et al. 2004

Date from artifacts collected next to the bank exposure of the cultural layer; other NP dates are from in situ position. Laboratory codes:: Beta, Beta Analytic, Inc. (Miami, Florida, USA); GIN, Geological Institute, Russian Academy of Sciences, Moscow; LE, Institute for the History of Material Culture, Russian Academy of Sciences (St. Petersburg, Russia); LU, St. Petersburg State University (St. Petersburg, Russia). All dates by Beta Analytic are 14C AMS; GIN, LU and LE dates are standard radiometric (conventional 14C) dates.

Human Habitation in Arctic Western Beringia Prior to the LGM

21

Table 2-2 Cont’d. No 79 80 81

14

C date

Laboratory no.

Material dated

Reference

4380 ± 20 4430 ± 60 6710 ± 50

LE-6023 LE-8480 LE -6025

Wood Wood Twigs

Pitulko et al. 2004 Pitulko et al. 2004 Pitulko et al. 2004

Radiocarbon dates from the cultural layer at Northern Point locality obtained on wood 82 44,010 ± 1100 Beta-216803 Wood 83 > 45,200 Beta-216802 Wood 84 > 47,000 LE-7446 Wood

Pitulko and Pavlova 2007 Pitulko and Pavlova 2007 Pitulko and Pavlova 2007

1

Date from artifacts collected next to the bank exposure of the cultural layer; other NP dates are from in situ position. Laboratory codes:: Beta, Beta Analytic, Inc. (Miami, Florida, USA); GIN, Geological Institute, Russian Academy of Sciences, Moscow; LE, Institute for the History of Material Culture, Russian Academy of Sciences (St. Petersburg, Russia); LU, St. Petersburg State University (St. Petersburg, Russia). All dates by Beta Analytic are 14C AMS; GIN, LU and LE dates are standard radiometric (conventional 14C) dates.

28,000 14C yr BP); the second corresponds to the upper boundary (~27,500–27,000 14C yr BP). Within the interval of these limits the site was possibly visited sporadically (Pitulko and Pavlova 2007, 2010). These phases are not really well pronounced in the dataset (Figure 2.4) and may not even exist, but it is clear that humans inhabited the area of the Yana site for at least two thousand years, ~29,000–27,000 14C yr BP, and visited it sporadically for some purpose. The number of such episodes is not identifiable. The youngest group of dates ( 35%), and numbers for other species (except for hare bones, which are only 5% of surface finds) are relatively close to those of in situ finds. The composition of in situ finds is

Graded density of finds per m2

Polygonal ice wedges, where finds are not present

Hearth

Unexcavated area Excavation grid numbers

Human Habitation in Arctic Western Beringia Prior to the LGM Table 2.3  Structure of the collection unearthed at Yana RHS—Northern point locality excavation (2002 through 2007) by major categories of finds, plotted material only. Category and type

No. of finds

Lithic artifacts 7058 Pebbles 160 Rock debris, roughly flaked pieces of raw material 267 Flakes 4975 Flakes of quartz crystal 84 Blades and bladelets 23 Cores (all types) 465 Artifacts of unclear morphology, but with secondary 46 retouch (no type and function found), mostly fragmented Flaked tool preforms 235 Sidescrapers (all types) 685 Core scrapers 9 Tools with expanded bifacial retouch 21 Chisel-like tools 37 Atypical forms 2 Micro-tools 47 Abrasives 2 Osseous (bone, antler & mammoth ivory) artifacts 1576 including flakes, slivers, shavings Fauna remains (bones and bone fragments) 21,016 Other, including groups listed below 1472 Hair 5 Wood 196 Red ocher paint (including pieces of raw material) 680 Fair-cracked rocks 350 Exfoliated rocks 32 Large pieces of slag-like caked substance 209 Total 31,122

completely reversed: Mammoth bones are 3.3% of the total, and hare bones exceed 20%. This illustrates the effects of taphonomic factors because small and light bones that erode out of exposures are easily carried off by the stream. The faunal collection is in the initial stage of analysis. This circumstance essentially limits possible final conclusions, forcing preliminary conclusions. Naturally the bone remains of large herbivores are present in the collection (Figure 2.6B, C; Table 2.4), including representatives of megafauna such as the mammoth and rhinoceros (all determinations are by Pavel Nikolskiy of Geological Institute, RAS). In the analyzed part of the collection, the proportion of unidentified bone is high. The undetermined bone amounts should therefore be expected to amount to at least 30–40% of the total collection (Figure 2.6C). The cultural layer of the site was filled with a large quantity of small bone pieces— ”crumbs”— which not only points toward the “kitchen use” of game animals, but also suggests bones were reduced by crushing for some purpose, perhaps for use as fuel. The majority of bones were damaged to some degree or broken. Some of them retain hunting lesions (Figure 2.7A, B). Study of the faunal collection has revealed a considerable number

23 of bone fragments with cutmarks, polished areas, and other traces of work or use. There are practically no large long-bones belonging to mammoth in the cultural layer (they occur sporadically), while the largest number of bone finds of this animal is contained in the surface collections. In the faunal complex of the site their presence is significantly less than that of bison, horse, reindeer, and hare (Figure 2.6C; Table 2.4). Besides the animals enumerated, there are small numbers of rhinoceros, musk ox, elk, red deer, beaver, and several other species. The finds of beaver and red deer point indirectly to the presence of a forested landscape at a relatively short distance, most probably in the intermountain basins to the south of the site. The predatory species are represented by wolf, Arctic fox, wolverine, and brown bear. In the collections of surface material there is a lion humerus, which, based on its radiocarbon age, is coeval with occupation of the site (Pitulko et al. 2004). The superficial characteristics of the faunal material of the site indicates that practically all species of large mammals that lived in the northern part of NE Asia at the end of the Karginian time are represented in the assemblage. In the initial publications it was reported that reindeer was the primary food resource of the inhabitants of the Yana site (Pitulko 2006; Pitulko et al. 2004). It now appears that bison was the predominant food resource, substantially supplemented by reindeer and horse. The hare evidently had important significance in the life of the occupants. Besides its food value, it was a source of lightweight and warm fur. Where the bones of this animal were especially numerous, planography has revealed mostly complete and articulated skeletons. This circumstance suggests that these animals were being procured for their skins, and this may point indirectly to seasonal hunting (at least in relation to this species of animal), since hunting it for skins makes sense when hares are in winter coat during late fall and early winter.

Material Culture: Lithic Tools

The characteristics of the industry given below, based on materials from excavations at the Northern Point locality for 2002–2007, are also valid as a general conclusion. In all, there are four distinctly different basic contexts in the industry from the Yana site: 1) production of multi-functional tools (sidescrapers) for processing hunted prey and different materials (Figures 2.8–10); 2) production of micro-tools for working tusk and bone (Figure 2.11); 3) manufacturing of artifacts from tusk and bone (Figures 2.7C; 2.12–14); and 4) production of red ocher (Figure 2.6D). A correspondingly limited number of categories of the inventory are also presented: cores and products of flaking associated with first and second stages of reduction, and final forms; objects of tusk, bone, and antler (hunting equipment, everyday

24

Pitulko et al.

Percentage of different species within the collection (100% = total number of the examined part of the collection) Percentage of the surface finds next to the excavated area (100% = total number of fauna remains collected on the surface) Percentage of the fauna remains found in situ (100% = total number of in situ finds of fauna remains mapped during excavations of the cultural layer)

Figure 2.6  General structure and statistics for the archaeological material excavated from Yana RHS—Northern Point locality. Counts of artifacts includes finds from 2002 through 2007; identification and statistics apply to approximately 4000 items of bone material. A, general structure of the collection and percentage for major categories of finds (100% = total number of finds); B, percentage of major groups of fauna remains (100% = total number of the examined part of the collection); C, percentage of the species composition; D, percentage of major cultural components of the assemblage (100% = total number of finds in situ, excluding the fauna remains); E, percentage of lithic artifacts found in situ (100% = total number of lithic artifacts found in situ).

implements, personal adornments and symbolic objects); and raw material for producing pigment and tools and lumps of prepared red ocher (Pitulko 2010; Pitulko et al. 2012). Cobbles of different varieties of gray and dark-gray silicified aleurites served as toolstone in the Yana industry (Table 2.3; Figure 2.6). Flint-like raw material is poorly represented in the collection, possibly because it was transported from some other source. A notable role in the production of micro-tools is played by quartz, though its proportion is insignificant in the total composition of the collection. Crystals of quartz could have been collected, even from the river cobbles, but most probably the source was on the slopes of the first spurs of the Kular Range 15–20 km south of the site, a day’s walk away.

The character of the stone industry is distinctly one of flakes (flakes amount to more than 70% of the collection of the stone inventory), which is normal for any industry in which access to raw material is unlimited. Its quality varies and on average is not very good. There are a quantity of elongated spalls. However, neither their quantity nor form permits us to draw the conclusion that Yana flintknappers were intentionally manufacturing elongated forms. Rather, it can be supposed that there was a preference for obtaining relatively short and broad blanks, their massiveness suggesting the use of a hard hammer. Blades and bladelets of irregular shape, exclusively double-edged, represented less than 1% of the artifacts. In many cases the flaking can be determined as situ-

Human Habitation in Arctic Western Beringia Prior to the LGM ational (or unsystematic), but on the whole the industry is based on principles of radial and orthogonal flaking. Correspondingly, the leading forms are orthogonal (some can be classified as uniface choppers and biface choppers) and discoidal cores, as well as pyramidal and unsystematic cores (Figure 2.8). Parallel flaking was definitely known to the occupants of the Yana site, but was rarely used, predominantly with end and planar work. In the collection there are only seven end cores, and two wedge-shaped ones made on large flakes—one is from the surface collection and one from the excavation (Figure 2.8: 3). End cores were formed, as a rule, on small elongated cobbles split lengthwise; in other cases it can be supposed that a depleted discoid core was reused. A rare method can be seen in the bipolar technique of flaking. The methods used for flaking quartz on the whole were evidently the same as used for all other materials. It is certainly possible to speak of bifacial radial flaking (Figure 2.11T), which was used for flaking quartz and possibly about planar parallel, which was predetermined by the form of source material and its ability to be flaked. A characteristic feature of the cobble technique of flaking in the Yana industry is the citrus-slice technique used to make blanks for backed sidescrapers, which are the most frequently encountered form in the inventory of the Yana site. Among the stone artifacts present is also a notable quantity of elongated cobbles with unifacial flattening (a core with flattening flaking?) and so-called reduced cobbles with a striking platform formed as a result of truncation and beveled in relation to the long axis of the source item of raw material (Figure 2.8B). In spite of the undoubted use of bifacial working, the

25 industry at the Yana site on the whole is unifacial. Among its tools the sidescraper category absolutely predominates (Table 2.3; Figure 2.6D–E) and includes single-, double-, and multi-working-edge varieties (Figure 2.9, 2.10). The most characteristic form is the simple backed sidescraper, with a single straight or weakly convex working edge. Often it was made on a large citrus flake and retained a cortex. At the same time, there is a significant group of backed scrapers that were formed by steep or vertical retouch and the working edge refined by ventral retouch. These large massive plano-convex tools make up perhaps the most characteristic group of tools of the Yana industry. Other characteristic forms are scrapers with multiple angular working edges (Figure 2.9D, 2.10D) and a carefully formed sharpened area made at the intersection of two converging working edges; pointed items (Figure 2.9A, C, E); and scrapers with an arc-shaped working edge (Figure 2.10E). Items less frequently seen are double-ended unifacially convex points (Figure 2.9B), scrapers with trim on the ventral surface, combination tools (Figures 2.9J; 2.10I), concave sidescrapers on flat cobbles (Figure 2.9H); pick-like tools (Figure 2.10J) of unclear function, and choppers and large chisel-like tools. There is also a substantial group of scrapers (primarily with multiple working edges) made on substantially less massive blanks, often rounded or oval in form, with working edges created by different methods of flaking—dorsal, ventral, and opposite, as well as bifacial (Figure 2.10A–C). In the collection are three bifaces round in plan. These are, however, probably residual discoid cores. Large tools show little evidence of secondary reduction. Most often the work-

Table 2.4  Fauna remains of the Yana RHS—Northern point excavation (based on processed part of the collection), counts by number and per cent. Species SF CL T SF (%) CL (%) T (%) Pleistocene hare 2 321 323 5.1 21.0 13.24 Beaver 0 1 1 0.0 0.1 0.04 Lemming 0 1 1 0.0 0.1 0.04 Wolf 1 62 63 2.6 4.0 2.58 0 6 6 0.0 0.4 0.25 Arctic fox Brown bear 0 3 3 0 0.2 0.12 Wolverine 0.04 Red deer 0 3 3 0.0 0.2 0.12 Elk 0 1 1 0.0 0.1 0.04 Reindeer 6 359 365 15.4 23.4 14.96 Pleistocene bison 10 595 605 25.6 38.9 24.80 Musk ox 0 1 1 0.0 0.1 0.04 Pleistocene horse 6 104 110 15.4 6.8 4.51 Woolly rhinoceros 0 2 2 0.0 0.1 0.08 Mammoth 14 67 81 35.9 4.4 3.32 Fishes 0 1 1 0.0 0.1 0.04 Total for identified bones 39 1531 1570 100.0 100.0 64.34 Unidentifiable bones 21 849 870 35.66 Total 60 2380 2440 100.0 SF, surface finds; CL, in situ finds; T, total.

26

Pitulko et al.

C

A

[128.104.46.206] Project MUSE (2024-03-01 06:01 GMT) UW-Madison Libraries

A1

B

B1

a

b

c

Figure 2.7  Yana RHS reindeer bones excavated in 2008 at Northern Point locality (main excavation area of the Yana site), showing hunting lesions, and osseous tools. A, proximal part of tibia; A1, embedded tip of the ivory point; B, pelvic bone fragment; B1, conical-shape hole made by hunting tool probably similar one shown in C. C, mammoth ivory points and their fragments: a, ivory shaft (medial fragment of large point?); b–c, ivory points with unilaterally beveled base.

ing edges were simply formed on a cobble already suitable in form and size (Figure 2.9G–I) instead of backing the artifact (naturally backed sidescrapers, normally with a single working edge formed by bifacial or, more often, unifacial retouch). Correspondingly, the basic method is predominantly edge retouch (ventral, dorsal, bifacial) of spalls fragmented by the simplest methods. The technique of removing burin spalls is uncharacteristic and present only in isolated (possibly involuntary) cases. Production of red ocher is, to certain degree, related to the large-tools category because it involves crushing and powdering polymict pink sandstone, which contains small

amounts of hematite (Pitulko and Ivanova 2010). Heavy-duty granite hammerstones, some of them broken, are found. The raw material comes from outcrops known upstream from the Yana site. Producing the pigment required four major steps 1) crushing the debris of the raw material to the smallest possible size; 2) powdering it by grinding; 3) extracting water from the pigment; followed by 4) drying and then mixing it with animal fat, which was needed to aggregate small particles of red ocher pigment into a consistency convenient to handle.

Human Habitation in Arctic Western Beringia Prior to the LGM Analyzed samples of fat residue most closely match reindeer fat (Pitulko et al. 2012). Micro-tools constitute a rather substantial series in the industry from the Yana site (Figure 2.11). Their length rarely exceeds 3 cm. Many of them are made from greenish gray argillite, or less often from quartz crystal or high-quality chert. Among them are chisel-like tools (Figure 2.11L, R), micro endscrapers (Figure 2.11E, Q), backed artifacts (Figure 2.11I), artifacts with a beveled working edge (Figure 2.11O, P), perforators, and micro-points, which are the leading form (Figure 2.11A–D, F). Surprisingly, burins, which are usually thought to be carving indicators, are rare, simple, and appear in nonstandardized forms. A rather rare but characteristic method

27 was used for chisel-like tools and micro endscrapers—forming the working edge in the area on the striking platform of the blank (Figure 2.11Q). Judging the form of the blanks for most micro-tools is difficult, but most probably elongated or lamellar spalls were used as well as fragments of flakes. Micro-points were made by applying vertical retouch on one side and sharpening the opposite edge. These tools co-occur spatially with ivory and bone waste products, indicating their use in manufacturing mammoth ivory, bone, and antler artifacts of all kinds.

Material Culture: Osseous Artifacts

Excavations yielded a number of osseous artifacts. In all, they

A

B

C

D

E

F Figure 2.8  Yana RHS lithic industry. B, reduced pebble. Flake cores: A, radial core; B, reduced pebble/core preform; C, edge-faceted wedge-shaped core; D–E, multidirectional cores (D possibly prepared for edge flaking); F, flat core.

28

Pitulko et al.

A

B

D

C

E

G

F

H

I

J

Figure 2.9  Yana RHS lithic industry. A, C, E, large points; B, double point; D, F–G, I–J, scrapers; H, massive tool on flat pebble, A–H made on cortical flakes or on “right-size” pebbles; A, E, symmetrical points; B, symmetrical double point; C, asymmetrical point; D, single straight sidescraper with sharp-angle tip; F, transverse scraper-knife with sharp-angle tip; G, single straight sidescraper with steep retouch; H, single concave sidescraper with abrupt retouch; I, naturally backed sidescraper with bifacial retouch; J, combination endscraper-sidescraper on massive elongated marginal flake.

belong to four categories: hunting equipment, everyday implements, personal adornments, and symbolic objects. There are about 1500 plotted items for 2002–2007 (Table 2.3), which is about 5% of the total, but constitutes ~16% within the collection of excavated artifacts; indeed, their number ex-

ceeds the number of lithic tools with secondary retouch. Two functional categories are osseous points and beveled ivory rods (hunting equipment), and bone/ivory punches, awls, eyed bone needles, and needle cases (sewing kits or everyday implements).

Human Habitation in Arctic Western Beringia Prior to the LGM Hunting equipment includes 15 beveled rods made mostly of mammoth ivory and 2 of woolly rhinoceros (Figure 2.12B, C; see also Pitulko and Nikolskiy 2012). Various proposals have been advanced in deciding the function of such rods; in most cases the researchers associated them with the complex of hunting equipment (Bradley 1995; Dixon 1999). Artifacts of like morphology and the same size are foreshafts, extensions of the striking part of the shaft of a spear/dart, which are found in the final-Pleistocene sites of the Clovis culture on the American continent, where they are validly as-

sociated with hunting megafauna, especially mammoth. Yana tools clearly resemble the foreshafts of much younger Clovis assemblages of North America. In addition, there are 15 large pointed ivory rods (Figure 2.7C) with unilaterally flattened bases (sagaie) and a number of basal fragments intermediate between these two types. Pointed rods (spear points, probably including projectile spears) appear prominently in the inventories of the Yana site. Whole objects and objects fragmented to various degrees are present (Figures 2.7C; 2.12A), and many pieces are

A

B

C

I

29

D

E

F

G

H

J

Figure 2.10  Yana RHS lithic industry. A–I, scrapers; J, pick-like tool, A–J all made on marginal flakes that retain cortex, except for H (massive flake preform used) and J (pebble preform); A–C, single convex sidescraper with bifacial retouch; D, convergent straight sidescraper with ventral retouch thinning that does not affect its working edge; E, naturally backed single convex sidescraper; F, angled scraper; G, naturally backed single convex sidescraper with bifacial retouch; H, straight transverse scraper; I, naturally backed combination endscraper-sidescraper (with convex working edge); J, pick-like tool.

Pitulko et al.

30

A

E

I

B

C

F

G

J

D

H

K

M

L

N

P

O

Q

S

R

T

Figure 2.11  Yana RHS lithic industry micro-tools. A–D, F, backed points; G, double-edged flake (bladelet?); M, blade fragment (?); E, Q, micro-endscrapers (E made on quartz crystal microblade with triangular cross section, Q, convex working edge formed on the proximal end by ventral retouching); O–P, asymmetrical points; I, backed tool; J, N, borer (  lateral tip formed by burin spalls); L, R, chisel-like tools (quartz crystal); S, combination endscrapersidescraper with notched tool; T, bifrontal radial core (quartz crystal).

Human Habitation in Arctic Western Beringia Prior to the LGM represented only by the basal part (with beveled base). For those whose size and shape can be confidently determined, they appear to fall into three groups of points—small (15–20 cm), medium (25–35 cm), and large (to 50 cm, and possibly more). The largest part of the cross section changes proportionally with the linear size of the tool, on large specimens attaining 2–2.5 cm. These are needle-like points with a beveled base whose surface was carefully polished, not so much for esthetic considerations as for an increase in penetrating capacity and to ensure quick withdrawal if thrust at close range). The beveled areas of the base exhibit transverse notches, which provided reliable hafting to the shaft or foreshaft. This simplest and obviously most numerous, universal type of hunting equipment was exceptionally widespread and evidently formed the basis for the complex of equipment of Paleolithic hunters over a vast period of time—from the Upper Paleolithic of Europe (Smith 1966) and Paleolithic of Siberia (Sitlivy et al. 1997: Figure 37:1–3, 61:1) to the pre-Holocene Clovis sites in America (Bradley and Stanford 2004: 467, Figure 2). Regardless of the way they were used (thrusting or throwing), they were formidable weapons capable of making damaging wounds, as seen from hunting lesions on bones (Figure 2.7). Everyday implements include first of all tools of a sewing kit. These are different punches made of sharp bones or bone fragments (Figure 2.12F–H), carefully manufactured awls of bone-shaft fragments, eyed needles, and needle cases (Figure 2.12). In all, the studied part of the collection includes 17 awls/punches and 26 eyed needles, of which 8 are intact and 18 are proximal eye fragments. These implements are made of long bones split lengthwise. Massive bone-shaft fragments were then shaped by cutting and shaving, and polished. Awl handles are heavily polished and covered with manufacturing cutmarks. Some of these tools are marked with finely cut, precisely spaced incisions (Figure 2.12K), possibly for making linear measurements. Needles are of two types: large and massive (90–110 mm long with eye diameter 2.2–2.7 mm), and smaller ones (50–60 mm long with eyes 1.0–1.4 mm). Most eyes were drilled from both sides, using conical drills. Needles and awls are similarly distributed spatially, indicating their use in both sewing and working skin. Most needles are undecorated, except for a few which bear a series of four or five dots or cuts symmetrically placed near the eye, and one with four and five dots on different surfaces of a single piece. These marks may have served as ownership marks (Figure 2.12I–J, L). Similar dot patterns were found on the needles of Denisova Cave, Altai Mountains (Derevianko and Rybin 2003). This group of finds also includes needle cases made of bone shafts. In one case a large needle was found next to an ornamented needle case made of possible wolf femur shaft. It is decorated by tightly placed narrow and short incisions organized as “bands” in a wavy/spiral pattern (Figure 2.12M–N). In general, the bone tools of the Yana site are quite varied. But even if they have clear and stable morphology,

31 their function may remain unclear. Thus, there are interesting wedge-shaped artifacts of massive long bones (most probably mammoth) split lengthwise, with heavy wear of the working ends (Figure 2.12D–E). They were probably used to split mammoth tusk chunks delivered to the living site for processing. In addition to these tools, many of them decorated, the Yana collection includes a number of finds in two categories of art objects. The first consists of diverse types of personal ornaments. The other comprises large three-dimensional decorated objects. Personal ornaments include small beads or sewn-on adornments (Figure 2.3), tooth pendants (Figures 2.13, 2.14), diverse flat decorations commonly known as hair-bands or “diadems” (Figure 2.14), and pendants of ivory and soft stone (Figure 2.14). Beads are most numerous within this category, numbering, for all years of work through 2011, about 6,000 finished specimens and more than 700 examples of semiproducts, preforms, and incomplete items (Pitulko et al. 2012). Finished specimens belong to two different types: type 1, simple rounded mammoth-ivory beads; and type 2, tubular beads with a deep concentric incision around the middle (Pitulko et al. 2012). Sub-quadrangular specimens (Figure 2.13As, v, z) are unfinished beads of type 1 and therefore do not warrant designation as a specific type. Some of them were colored with a red ocher pigment mixed with animal fat (Pitulko and Ivanova 2010; Pitulko et al. 2012). Although many of these finds come from water screening, hundreds of them have been found in situ both in linear arrangements and as single finds. The longest arrangement had 159 pieces (Figure 2.13B) with a repetitive symmetrical pattern which involves three round beads (type 1), followed by a single tubular one (type 2), and then by another group of three round beads, for a total of seven elements. Shorter fragments of arrangements consist of type 1 beads retaining their original position on the string. Type 1 beads (n = 5592), are simple round beads made of mammoth ivory (Figure 2.13Ag). The production sequence can be deduced from finished artifacts and by-products (Figure 2.13Ap–z). It started with making preforms of long, thin carefully shaved flat ivory plates (Figure 2.13Ap). Preforms were then shaped with side notches (Figure 2.13Aw) and subquadrangular bead blanks were detached (Figure 2.13Aq–r, t–u, x–y). After that, bead blanks were pierced with a biconical drill (Figure 2.13As, w, z). Sub-quadrangular beads were then tightly assembled on a thin stick or string for rounding and polishing. This final operation is needed to give the beads uniform size and balance, which is important for their future use. This technology is closely reminiscent of French Aurignacian (White 1995) and Russian Plain Upper Paleolithic Sunghir 1 (White 1997). Although the size of type 1 beads varies, it is clear that there existed a standard: 2.1 mm height/length, and 4.0 mm diameter. Internal hole diameter is 2.4 mm, but is sometimes as small as 1.6 mm, which indicates use of really thin string.

Pitulko et al.

32

F

D A

G

E

H

B

C

I

K J

L M

N

Human Habitation in Arctic Western Beringia Prior to the LGM It is also found that values of D (diameter) and L (length) for both types of beads demonstrate the ratio 1:2. Remarkably, the size standard for type 1 ivory beads (mean value D = 4.0 mm) perfectly corresponds to that of type 2 tubular “hare bone” beads (D = 3.8 mm). Obviously the diameter of hare bone determined the size of type 1 ivory beads. This suggests a high degree of skill, preference, and reasoning skills. Type 2 tubular beads (N = 299), made of bones from the Pleistocene hare (Lepus cf. tanaiticus Gureev), are represented by several hundred finished items and by-products (Figure 2.13Ae–f; h–o). Their dimensions vary. Finished items are 7.5 mm long (mean value) with a diameter of 3.8 mm. They were made by cutting a bone shaft into pieces, which were then decorated with a groove around the cylinder and polished. Another group of personal decorations includes all kinds of pendants, mostly tooth and mammoth-ivory pendants and unique forms made of minerals (Pitulko et al. 2012). Thus, a piece of anthraxolite, a soft jet-black mineral, is shaped like

33 from herbivore incisors and rarely of Arctic fox canines (Figure 2.13Ab). They number 81 items. All of them, except for one large notched horse-tooth pendant from Yana B locality (Figure 2.14A), are found at the Northern point excavation (Table 2.5). Tooth pendants are often found in groups whose composition indicates necklaces made from single species (bison or musk ox) or combined reindeer and bison necklaces. Most teeth have a drilled or cut (rarely) perforation through the root portion of the tooth (Figure 2.13Aa), although four items are notched (Figure 2.14A) and one is grooved (Figure 2.13Ad). It is interesting to note that in some cases there are caches of “preforms,” untouched incisors with no perforation or a mixed set of both drilled and blank items. In one case there is a set of seven reindeer milk incisors (Table 2.5). Most of the in situ finds consist of groups of seven, indicating great antiquity for this common “magic” number (Miller 1956:81–97). Another group of pendants includes eight fragments of ring-shaped

Table 2.5  Yana RHS site, excavations 2002 through 2010. Tooth pendants by species (single finds and groups of drilled pendants and groups of tooth prepared for processing). Animal species Artifact description No. of single finds No. of groups

Group description, type and no. of artifacts

Reindeer incisor Drilled tooth 23 2 7 Reindeer milk incisor (R-I) Drilled tooth 9 Reindeer milk incisor (R-MI) Untouched preform 6 2 7 Pleistocene bison incisor (Pb-I) Drilled tooth 3 3 7 2 Pb-I + 1 R-I 2 Pb-I + 5 (no ID) Pleistocene horse incisor (Ph-I) Drilled tooth 1 Pleistocene horse incisor (Ph-I) Tooth with a notch in the 2 root part Wolf fang (W-F) Tooth with a notch in the 1 root part Wolf Pm1 (low) Tooth with a notch in the 1 root part Arctic fox fang (AF-F) Drilled tooth 9

a horse head whose likeness has been enhanced by a conical drilled eye (Figure 2.14B). The same exotic raw material appears in decorations of Malta site, Siberia (Medvedev 1998). Another piece is a red amber pebble pendant with a double encircling cut. Since anthraxolite and amber have been reported from New Siberia Island, about 600 km north of Yana RHS area (Gakkel 1967), this can be taken as evidence for long-distance transport of exotic materials. Tooth pendants (Figures 2.13Aa, c–d; 2.14A) were made Figure 2.12 (opposite)  Yana RHS osseous tools. Point fragments and rods with beveled ends (foreshafts): A, ivory rod fragment; B, ivory foreshaft; C, rhinoceros-horn foreshaft. Point fragments, awls, punches, bone wedges: D–E, bone wedges made of lengthwise-split bone (function is unclear); F, H, awl made of bone; G, bone punch. Yana RHS osseous artifacts: I–J, bone needle fragments with dot marks near the eye; K, polished awl of massive bone fragment, with multiple regular incisions on the laterals; L, large eyed bone needle, with dot marks near the eye (ownership marks?); M, ornamented needle case of wolf-femur shaft (?); N, loft drawing of the decoration on the needle case.

Total 37 9 15 14

1 2 1 1 9

ivory artifacts (Figure 2.14C–E) termed Kurtak-type pendants (Pitulko et al. 2012). Mammoth-ivory “diadems,” found in many Upper Paleolithic sites across Eurasia, are also abundant at Yana. These are thin, flat plaques flat perforated on the ends. They probably served to decorate hair bands (Figure 2.14F–J). Only one was intact; 40 others were end fragments with a single or, in one case, double perforation. The level of fragmentation makes it hard to tell whether all of them were hair-band decorations or bracelets or some other flat sewn-on ornament. Hair-band ornaments and their fragments clearly fall into two metric classes: narrow (4–6 mm) and wide (9–12 mm). All are planoconvex in cross section and have either conical or biconical drilled holes. They are decorated with a linear arrangement of strokes and dots parallel to the long axis. Wavy linear designs are also present, as well as patterns composed of transverse lines and complex rectangular compositions. These design elements may be either drawn or formed by dots.

Pitulko et al.

34

A

b

c

d

a h

i

n l

f

e

j

o

k p

g Figure 2.13  Yana RHS personal ornaments, typical semi-products, and preforms. A, byproducts of bead production at Yana site: a, reindeer incisors with drilled holes; b, drilled arctic fox canines; c, drilled reindeer incisor; d, reindeer incisor with a circum cut; e–f, tubular beads with a circum-cut decoration (type-2 beads); g, bone beads (type-1 beads). Type-2 bead production h–o (by-products of bead production including distal and proximal parts of bones left after diaphysis was used). Type-1 bead production p–z (semi-products, preforms, by-products, and incomplete beads): p, initial preform—thin and long, carefully shaved flat ivory plate with rectangular cross section; w, a preform with side notches that mark the bead preform; q–r, t–u, x–y, bead preforms; s, v, z, drilled bead preforms with unsuccessful drilling. Pleistocene hare (Lepus cf. tanaiticus Gureev) cut bones h–o: h–k, o, distal fragments of metatarsal bones; l–n, proximal fragments of metatarsal bones; p–z, mammoth ivory. B, the largest concentration of beads (159 beads in linear arrangement) discovered in J40 unit in the excavation at Yana RHS—Northern Point locality (field drawing by A. Mashezerskaya). B1, upper part of the concentration; B2, lower part; a–r, groups within the concentration (collecting sequence). Repeated pattern 1-31 is clearly visible (tubular bead with central incision is followed by three simple rounded beads, and then repeats).

w r

c

m

q

r

s

t

u

v

x

y

z B

b f d

h

e a

p

g m

i j

Finally, a peculiar motif in the shape of the letter “A” should be noted. This is possibly a symbolic, very simplified representation of a human being. A small fragment of a thin ivory plaque also carries this motif. Its surface bears a complex geometrical design of widely spaced dotted lines separated with either diagonals or a double oblique cross (Figure 2.14O). Continuous crossing lines are decorated with short regular strokes set at a right angle from one side. Anthropomorphic symbols make a row parallel to the edge of the object.

k B1

n

l

o B2

Another fragment (Figure 2.14K) has a dashed line of small dots that form an anthropomorphic image with several strokes placed inside it. Four lines of paired dots run perpendicular to the long axis of the object. A line of widely spaced strokes separates the main field from the bottom one, with two symbols placed in it. The latter are formed by vertical strokes bordered by a horizontal line. These two fragments are probably bracelets. It should be noted that there are many other decorated objects in addition to personal ornaments, notably fragments

Human Habitation in Arctic Western Beringia Prior to the LGM of thin (3–5 mm in diameter) ivory shafts decorated with multiple perimeter incisions (Figure 2.14L–N).

Art Objects In Relief

Three-dimensional art objects are rare and less well preserved. This group includes 19 antler animal figurines, ornamented ivory vessels (two fragmented, one intact), and two engraved mammoth tusks. One of them bears a complex symbolic composition of anthropomorphic images, lines, and shaded zones, which demonstrates that the inhabitants of this remote part of the world had developed the faculty of visual perspective as early as 28,000 BP, which is one of the earliest records anywhere (Pitulko et al. 2012). At least four ivory vessels have been unearthed, in different states of preservation. Three are fragmented. The most complete one has a complicated surface decoration that entirely covers the artifact. The designs consist either of relatively sparse or dense impressions creating “knotty” lines arranged in wavy or zigzag patterns on the outside surface and in a sort of primitive meander on the front side (Figure 2.14P). The artifact has a pair of drilled holes near the edge that may have been an unsuccessful attempt at repair, and an opening pierced from the inside. There is no indication of the function of these artifacts, or whether they are objects of ritual, but ethnographic and historical records cite containers made of wood, birch-bark, walrus ivory, or bone that served a ritual purpose. For example, the ceremonial offering of water to a killed animal, a hunter’s way of asking for forgiveness, is one of the most common rituals in hunting societies elsewhere in Siberia, the Russian Far East, and the Arctic regions. Antler animal figurines were made of detached antler bases that have been whittled into animal forms. The bottom surfaces are ground flat. The rough texture of the antler base with its bony excrescences may have simulated fur or hair and explains why this material was chosen for figurines. Most figurines depict mammoth or bison; at least one is a horse head (Figure 2.14Q–T) . These very abstract, schematic zoomorphic images resemble the marl sculptures from Kostenki (Abramova 1995). Although their function is unclear, they may have served as counters for logical games, as is the case for similar small animal figures known from the Siberian ethnographic record (Ivanov 1970).

Discussion

The discovery of the Yana RHS site more than doubled the length of known human habitation in western Beringia and generally in the Arctic. Together with the Mamontovaya Kuria site in the north of European Russia (Pavlov et al. 2001), whose age is overestimated (Pitulko et al. 2011), this site confirms the oldest known phase of human dispersal into the Arctic regions of Eurasia that pre-dates the LGM. This stage is often also thought to be “initial” (Dolukhanov 2008; Gribchenko 2008; Slobodin 2011; Velichko and Vasil’iev 2008), which is most probably not the case for western Beringia. At this time, at around 30,000 14C yr BP, the presence of humans in the

35 Arctic, including western Beringia, becomes archaeologically visible as a result of growing population density, but the timing of the initial migration into unglaciated territories such as western Beringia has not yet been determined. In contrast to human migrations to unglaciated areas, initial human migrations to glaciated areas were obviously being controlled by the deglaciation of the European north. Nevertheless Yana RHS is the earliest known phase of human habitation within western Beringia. Perhaps some of the materials found in Vilui River valley (Mochanov and Fedoseeva 2002) belong to the same phase as Yana RHS. However, they are mostly surface finds. Judging by their morphology, fauna associations, topography, and the Quaternary geology of the region, they may be even slightly older than Yana RHS. Although no results of excavations of these sites have been professionally published, which excludes further discussion of them, their potential significance exceeds that of the Diring-Yuriakh site. Estimates of its age make it no older than 260,000 years ago (Waters et al. 1997, 1999). It must be noted that this is the age of the sediments that were supposedly a matrix for the artifacts associated with the deflation level on the profile. When sampling the sediments, Waters and his colleagues did not observe that (Waters et al. 1999); so this is a date for the sediments, not for the artifacts. Based on their geological context, which is easy to see from published data (Mochanov 1992), the Diring artifacts cannot be older than late Pleistocene and so likely post-date the LGM. Slobodin (2011) includes this site in the “Pebble-tool” tradition, which is, in his opinion, the oldest one in Beringia, though poorly known. However, it is hard to say whether it constitutes a unified tradition. It is clear that assemblages assigned to the Pebble-tool tradition (e.g., Diring-Yuriakh, Lopatka, Siberdik), although different in age, all date to post-LGM, which suggests that this “tradition” is a mixture of archaeological materials. If such unity really exists, it is not visible yet. These assemblages cannot be grouped together simply because they appear to reflect use of pebble/ cobbles. The decision to use pebbles and cobbles is dictated by raw-material availability, site function, and site activities, and may not reflect an archaeological tradition. This is well demonstrated by the Yana site industry (Pitulko 2010; Pitulko and Pavlova 2010; Pitulko et al. 2004). It developed using raw material that, although not of great quality, was abundant and served as a satisfactory toolstone. Yana RHS belongs to the oldest known member of the cultural stratigraphy within western Beringia, but it does not set a precedent in terms of artifact morphology, or lithic and bone technology. Experience in dating the Yana site once again demonstrates that radiocarbon dates obtained on bone collagen from animals unmistakably killed by humans are the most reliable indication of the age of an Upper Paleolithic site. This fact is essential in interpreting radiocarbon dating results. There must be serial dating of overlying and underlying sediments. A few dates that differ in age may not be wrong. Instead, they often reflect a period of human activity at the site

Pitulko et al.

36

B

A

C

D

E

F

G

I

H

P

K

J M

Q

O R

N

L S

T

Human Habitation in Arctic Western Beringia Prior to the LGM

37

location that may span 2000–5000 radiocarbon years, as previously shown by Sulerzhitsky (2004) and which is the case for the Yana site. In Yana, on the basis of serial dating, three habitation phases are found, but each consists of hard-todefine, repeated visits/episodes, sometimes visible through the microstratigraphy of the cultural layer. It must be stressed that dates on mammoth, particularly on mammoth ivory, can be much older than the habitation because humans may have collected old ivory, in which case such dates are not necessarily coeval to the habitation (Sulerzhitsky 2004; Pitulko and Pavlova 2007, 2010). Moreover, dating of wood collected from the cultural layer demonstrates the low reliability of such dates for Paleo­ lithic sites within permafrost deposits. Their radiocarbon date may be skewed as a result of having been water transported from their natural source (exposure of older deposits that existed at a time of habitation) or having been brought to the site by humans for use as fuel or as material for making artifacts because wood was not readily available. Since the landscape of most parts of western Beringia remained treeless for a significant time, dates on wood must be interpreted with care. Thus Abramova (1979), and Yi and Clark (1985) questioned the age of several Aldan sites that Mochanov claimed were older than 30,000 14C yr BP (Mochanov 1977). The experience of dating the Yana site and the study of the geology of the Aldan area (Pitulko and Pavlova 2010) amply demonstrate how such misinterpretation occurs and confirm that so-called “old” sites of the Dyuktai cultural group constitute a component of the post-LGM Microblade Beringian tradition. Although the radiocarbon dating dataset reveals three habitation phases, no part of the Yana site is associated with the youngest date of about 26,000 BP. Two habitation phases are clearly present in the studied part of the site at Yana B, Northern Point, and TUMS 1 localities, whose cultural layer formed during multiple habitation episodes. Presumably that was a seasonal habitation, as indicated by evidence of reindeer hunting (or multi-seasonal, perhaps sometimes even a year-round habitation). It is obvious that the Yana settlers kept returning to the same place for a substantial amount of time. Their migration range (or trade network?) extended at least several hundred kilometers, which is evident from exotic raw materials—­ anthraxolite and amber—used for manufacturing personal ornaments (Pitulko et al. 2012). The nearest known source for both of these materials is in the area of the New Siberian Islands. Although no dwelling structures were found, a

number of hearths were opened during the excavations. At Northern Point locality, hearths occur on a slightly elevated paleo-ground surface that was used for camping many times. Hearths lie in an irregular spatial pattern, not situated in a central area, but instead in two lines at almost a 90° angle. A single-line pattern of hearth planography was observed in the Transbaikal by A. Konstantinov (2001). It is known also, for instance, from the Malta site and at Kostenki, and normally is explained by long-house construction. But at these sites, this interpretation is supported by dwelling features observed in situ (e.g., mammoth bones, cobbles and rocks, piled antlers), which is not the case for the Yana features. No single large mammoth bone is associated with the habitation spots in Yana, despite the mass accumulation of mammoth skeletal remains (Basilyan et al. 2011), which is coeval with the habitation areas and constitutes a part of the site’s spatial structure. In situ ivory, rib fragments, and hyoid bones indicate the limited consumption of fresh meat and the wide use of ivory as a raw material. The preference for ivory as raw material is evidence that mammoths were hunted (Nikolskiy and Pitulko 2012). It is noteworthy that the Yana site area was abandoned at around 26,000 14C yr BP, when the mammoth population in western arctic Beringia experienced a significant decline (Nikolskiy et al. 2011). This suggests that the Yana site area was chosen because of its convenience for mammoth hunting, even though mammoths were not a primary food source. It is equally noteworthy that a study of the frequencies of distribution of dated mammoth remains in the northern Yana-Indigirka lowlands (Nikolskiy et al. 2011) demonstrates that the presence in these regions of people at 28,000 14C yr BP had no influence on the local mammoth population. Indeed, one of the peaks of the mammoth population occurred at this time. Mammoth ivory was an important raw material for manufacturing personal ornaments and, most important, as toolstock. The widespread use of bone implements in the Upper Paleolithic is well known and inspired Anikovich to term it “the Bone epoch” (Anikovich 2003; 2007), by analogy with the “bronze” and “iron” ones. The importance of ivory for making hunting weapons is quite comparable with the significance of metals in the same sense and was well demonstrated by Peterkin (1993). The Yana RHS site exemplifies this, since the Yana hunting equipment kit consists almost exclusively of ivory/bone tools, including long points and foreshafts. Some lithics have been also used, but only as a detail, perhaps only for tipping the ivory point. A fragment of one such tipping point is embedded in a mammoth scapula (Nikolskiy and Pitulko 2012). It is also well known that Upper Paleolithic humans were capable of producing shafts of ivory that exceeded the length of a full-size thrusting spear found in Sunghir (Bader 1998), Berelekh (Vereschagin 1977), and in several sites in the Russian Plain—Kostenki, Avdeevo, Eliseevichi, Khotylevo, and others as finished artifacts or by-products (Khlopatchev 2006). The maximum length of the ivory shafts from Yana is 63

Figure 2.14 (opposite)  Yana RHS personal ornaments, osseous artifacts, and zoomorphic figurines: A, tooth pendant (horse incisor with notch); B, “horse head” pendant made of black soft stone; C–E, fragments of ringshaped pendants (Kurtak-type pendants). Raw material: A, Pleistocene horse incisor; B, anthraxolite; C–E, mammoth ivory. F–J, decorated ivory “diadems” with drilled holes on their ends (end and medial fragments); K, fragment of ornamented bracelet; L–N, thin polished ivory rods with regular deep-cut circumferential incisions (not bead preforms); O, fragment of ornamented ivory bracelet; P, decorated ivory scoop. Zoomorphic figurines made of shed antler base parts: R–T, mammoth figurines; Q, horse head (?).

38 cm, but some Eurasian finds of this implement are much longer. The length of the Berelekh shaft (Vereschagin 1977), although incomplete, exceeds 90 cm. The famous ivory spears from Sunghir burials are about 2 m long (Bader 1998) and are thought to be ritual objects. It seems likely that both full-size ivory spears and the idea of hunting with spears equipped with detachable foreshafts made of ivory or other suitable material arose as an adaptation to a treeless landscape. The shortage of wood made this a highly desirable material and much more readily available than wood. Although wood was available as driftwood everywhere in the Eurasian Arctic even at 28,000 14C yr BP, its poor quality may have been a problem for toolmaking, especially for making shafts for spears. Pollen data from Yana locality indicate no trees at the habitation time (Pitulko and Pavlova 2010; Pavlova et al. 2012; Pitulko et al. 2007). Driftwood is commonly too flimsy to use for making shafts. Thus animal horn or ivory (strong, flexible, very thin for its strength) would have been used even when wood was available. Mammoth ivory suited this use admirably because it was tough, dense, and could be cut into long, thin shafts. Even a mid-size tusk could supply the stock for several shafts up to 2 m long. Historical Inuit of Thule, Greenland, for whom driftwood was scarce, make spears of walrus and narwhal tusks. Some of the latter were 2 m long (Malaurie 1989) and were used as hunting equipment, not for ritual or ceremony. Foreshafts of mammoth ivory and rhinoceros horn constitute a part of the Yana hunting weapon kit. This is a peculiar feature of the Yana assemblage. Without drawing direct analogies between the foreshafts of Yana RHS and those of North American Clovis (Dunbar et al. 1989), which are separated in space by 3,000–4,000 km and in time by > 15,000 years, we must nevertheless note that the hunting equipment used by both cultures employed a detachable spearhead. Somehow this technology must have reached the New World. In Northeast Asia, it is found at the Late Paleolithic locality of Berelekh (Vereschagin 1977) and at a number of Neolithic sites (Fedoseeva 1980, 1992). This technology persisted in NE Asia until the end of the Stone Age and during the Neolithic served as the foundation for the Eskimo harpoon complex. Decline of the mammoth population in western Beringia after the LGM probably led to important cultural changes that are visible in the archaeological record. This is the time of the dispersion of assemblages of the Beringian Microblade tradition. Its spread follows mammoth migration in a northerly direction when their habitat shrank, and then their dispersal into refuges separated from each other (Pitulko and Nikolskiy 2012). Thus the extinction of mammoth becomes a trigger for archaeologically visible cultural changes. Obviously Yana inhabitants did not depend on a single or rare resource. Fauna remains indicate intensive exploitation of many of large herbivores, including woolly rhinoceros, Pleistocene bison, Pleistocene horse, reindeer, and Pleistocene hare. Remains of woolly rhinoceros are found in rather notable quantity. It is quite interesting that they are present in practically all sites of the Upper Paleolithic, when

Pitulko et al. this animal lived together with humans, up to its extinction 15,000–14,000 years ago. Similar observations have been made rather widely at Yenisei sites (Abramova 1989), Transbaikal region sites (Klement’ev 2005), and in NE Asia (Mochanov 1977). In individual cases (primarily in Transbaikal) the great quantity of rhinoceros bones testifies that it was a basic hunted species. Collections may not have mammoth remains (and often do not), but rhinoceros remains are practically always present. The widespread association of rhinoceros and humans is evidence that people inhabited ecotopes similar to the habitat of the rhinoceros—low-hilly and medium-hilly areas with slopes with southern exposure, dry, moderately covered with woody-brushy vegetation. The rhinoceros itself, being an animal of predictable habits (it follows paths within its feeding area), was for this reason favored for hunting. Consequently, paleontological finds of this animal can help locate occupation sites dating to MIS3 and MIS4 times. The mass procurement of hares is visible from the fauna remains excavated from the site. It is quite remarkable that many bones are still articulated, frequently as a complete skeleton. The remains of hundreds of animals are evident, which suggests site occupation in the autumn. Hares are hunted best with snares. Judging by the abundance of fully articulated hare remains, this animal must have been hunted for its fur rather than its meat. Hare pelts are very light and warm, but not durable. Its fur was used by many northern peoples for underclothes, insoles, and inner socks. According to Malaurie (1989), Inuit hunters of the Thule District in Greenland trapped as many as 1,000 to 1,500 hares each season to make clothing, but rarely ate them as they considered their meat tasteless. The stone industry of Yana RHS has no direct analogies with any of known Siberian EUP assemblages. Sites of the Karginian period are scattered from Altai to Baikal, far south in Siberia. The typological and technological characteristics of these sites are extremely variable, and their occupants have no demonstrable affiliation with other cultures. This may be due in part to the diversity of environmental conditions that is typical for warm climatic epochs. On the other hand, environmental “leveling” during cold phases may be expected to produce a noticeable standardization in material culture, as exemplified in Siberia by the widespread distribution of assemblages based on wedge-shaped core technologies during the LGM and termed by Goebel “cold adaptation” (Goebel 2002). However, it was not adaptation just to cold but to above-mentioned changes in animal population driven by climatic changes (Pitulko and Nikolskiy 2012; MacDonald et al. 2012; Nikolskiy et al. 2011). Although the Yana RHS assemblage creates an impression of primitive technology owing primarily to the absence of blades, the bone and ivory specimens are anything but primitive. Following Bar-Yosef and Kuhn (1999), we therefore have to question the theoretical dictum that blade industries are more “advanced” or sophisticated than those based on flakes. The Yana inventory clearly demonstrates that this is

Human Habitation in Arctic Western Beringia Prior to the LGM not necessarily true, and perhaps this applies in many other cases as well. This is chiefly a question of taphonomy because such well preserved organics cannot be found everywhere in the world. Yana clearly shows that a flake-based industry can achieve a high level of technological excellence. In addition, it must be mentioned that experiments by Eren et al. (2008) confirm that a blade industry is not any better than a flake-based one. Any tool can be made of flakes. The exception is a situation where standardized preforms are essential or where use of such preforms is dictated by some extenuating factor. In all other cases, flakes are even better than blades because their production does not require a great amount of raw material. Flake technology is also lowrisk in terms of potential knapping errors. If toolstone is plentiful and standard preforms are not required, then both flake-based and blade-based technologies can be used by the same cultural group. Probably this is the key to the blade/non-blade dichotomy observed in eastern Beringia (Wygal 2011). It was also suggested for cultural variability in Yenisei (Vasil’ev 2001), where Afontovo and Kokorevo cultural groups are coeval and exist in the same territory (Abramova 1989). Re-investigation of the question led Graf (2011) to conclude that neither model (different cultural groups versus different site functions) can be confirmed in the lithic data. But probably this dichotomy can be explained by seasonality (Pitulko 2010) and also by different activities practiced at the site (i.e., by use of a different behavioral model) by the same group. Thus the Afontovo sites (Afontova Gora type sites ) yielded a number of micro-tools including chisel-like tools. These sites, particularly Afontova Gora, have many ivory artifacts and much ivory waste material, which suggests ivory processing. The same is observed in the industry at the Yana site (Pitulko and Pavlova 2010). Possibly this is a clue to the dichotomy. Other influences may lie in seasonal changes in the activity and in the toolkit. Yana lithic artifacts, both large tools and micro-tools, find certain analogies in different Upper Paleolithic sites far south, in the Yenisei and Transbaikal regions, Mongolia, and even in West Siberia. They are not all necessarily of the Yana age. Similar core technology and large-tool (scrapers) morphology and typology can be found in Irkutsk Military Hospital (Abramova 1989), Kurtak 4 (Lisitsyn 2000), and Tolbaga (Konstantinov M. 1996) whose ages are comparable to Yana RHS. At the same time, large sidescrapers of Afontovo and Kokorevo sites demonstrate sometimes even closer similarity (Abramova 1989; Lisitsyn 2000). The Yana RHS micro-tool toolkit is especially interesting for its similarity to artifacts found elsewhere. Several microtools that emulate those from Yana were found during excavations at the west Siberian site of Shestakovo (Derevianko et al. 2003a), which is somewhat younger than Yana. They were found in Horizons 7 and 6. Radiocarbon dating set the age of Horizon 7 at 22,000–21,000 14C yr BP, and the age of Horizon 6 at about 20,000 14C yr BP. The site is associated with a cluster of mammoth remains near a salt lick. Presumably the

39 complex of micro-tools, whose function was not determined (Derevianko et al. 2003a:78, 85–86, Figure 54), is related to ivory processing. It appears that toolmakers visited the site to collect ivory from the carcasses of sick animals that had died. Use-wear analysis of these tools revealed they had been used to work ivory (N. Skakun, pers. comm.). Similar sub­ triangular points are known from other sites, including some closer in time to the Yana site, for example, in the materials of the Kamenka B site (Lbova 2000:5) or Tolbor-4 (Rybin et al. 2007:146, Figure 2.3:5, 11, 21). The latter is not dated, but judging by analogies cited during the analysis of his materials, Rybin suggests its age is no less than 30,000–40,000 years ago. These remote and asynchronous analogies demonstrate that specious purpose- or function-related similarities may be found in the core technology, tool morphology, and typology of disparate assemblages. Their apparent similarity betokens a functional relatedness rather than a cultural continuity as once thought. Some technologies and tool types, of course, may serve as a genuine fossile directore, such as Clovis technology and points, Chindadn points, etc. A rare but real opportunity to trace cultural development and continuity comes when taphonomic factors do not complicate the preservation of organics. This is a primary advantage of practicing archaeology in permafrost regions, and the Yana site is an excellent example. Besides hunting tools and everyday implements, Yana yielded many organic artifacts of everyday implements such as needles and awls. The sewing toolkit in Yana is exceptionally well developed, featuring a variety of eyed needles used for different kinds of work. The collection of needles from Yana is one of the oldest in the world. Many Yana needles bear a sort of “decoration,” which is most probably an ownership mark. Needles from Denisova Cave in Altai (Derevianko and Rybin 2003) are similarly marked. The most impressive part of the Yana collection, however, is the personal ornaments, zoomorphic figurines, and diverse decorated objects exemplified by the fragment of mammoth tusk with complex symbolic composition (Pitulko et al. 2012). The collection of about 650 items (Medevedev 1998) exceeds by many times the number of all “Paleolithic art” artifacts previously found in Siberia. Personal ornaments in the Yana collection include many types known from early Upper Paleolithic sites across Eurasia: tooth pendants, small beads, tubular pendants with grooved cuts. In addition to such simple ornaments, Yana RHS contains a number of artifacts that are less common but are still known widely throughout Eurasia in time and space (Pitulko et al. 2012). Chronologically these types mostly predate LGM. For example, ring-shaped ivory pendants from Yana are similar to those found at the Upper Yenisei Kurtak 4 site, radiocarbon dated from ca. 26,000 to 23,000 14C yr BP (Lisitsyn 2000). They are termed Kurtak-type pendants (Pitulko et al. 2012). In Siberia, except for the Yana site and Kurtak 4, they are found in a number of places—in level 3 of Khotyk in Transbaikal area between 34,000 and 26,000 14C yr BP (Tashak 2009), in Denisova Cave (Derevianko and Rybin 2003;

40 Derevianko et al. 2003), in Kashtanka 1 (Lisitsyn 2000) and Ust’ Kova (Medvedev 1998). The degree of similarity among these finds varies. They can be made of different materials such as mammoth ivory, soft stone, and even fossil (Pliocene) ostrich eggshell, which is abundant in south Siberia. But among the sites listed, only Kashtanka 1 is significantly younger than Yana RHS, while others are of roughly the same age or even older than Yana, suggesting it is a widespread and significant Upper Paleolithic cultural element. Surprisingly, a similar find is known also from the Aurignacian context at Spy, Belgium (Vanhaeren and d’Errico 2006). The abstract mammoth and bison figurines from Yana have much in common with the marl sculptures at Kostenki 1, 4, and 11 (Abramova 1995). In Western Europe, similar objects come from Isturitz (Abramova 2003–2004). While the Kostenki figurines are somewhat younger than the Yana ones, the Isturitz sculpture is from the typical Aurignacian layer. Similar objects are also known from different European assemblages, either contemporaneous with Yana RHS or slightly younger (Abramova 1995; 2003–2004). Single finds at a particular site are usually more realistic, while multiple examples tend to be more abstract. The closest analogy to the Yana zoomorphic figurines are flat-based, abstract carvings from Kostenki 11, which are about the same size as the Yana sculptures. Some of the artifacts found at Yana do not have analogy at all. Decorated ivory vessels are not known anywhere else except for the Listvenka site in the Angara region of Siberia (Akimova and Drozdov 2005). Listvenka is much younger than Yana, however, and the piece is of poorer quality. Perhaps an ivory object found at the Eliseevichi site in the Russian Plain was a preform for such an artifact (Khlopachev 2006). Whatever function they served, they were rare artifacts that probably filled a ritual purpose. At the same time, Yana finds share certain features with traditional Siberian ethnographic designs. The decoration on a plaque (Figure 2.14O), for example, resembles a historic Yukaghir design signifying the relationship of a man and a woman (Iochelson 2005). The only difference between the Yukaghir pattern, which dates to the 19th–20th century, and the Yana piece is the transformation of the ornamental belt composed of zigzag anthropomorphic elements. Rows of instantly recognizable anthropomorphic symbols are usually thought to be associated with shamanism. Such designs are widespread in the traditional culture of Kets, Selkups, Evens and some other peoples of Siberia (Ivanov 1954), serving as decoration on shaman costumes. Such parallels make a plausible argument for the existence of some shamanic cult among the Yana people. Additionally, the engraved composition on the mammoth tusk has certain similarity with historical Yukaghir’ birch-bark pictograms called “tos” (Iochelson 2005). Also, the 1-and-3 pattern in bead decoration is very interesting, since it anticipates the unique Yukaghir counting system based on 1 and 3 (Tugolukov 1979). All these apparent similarities, however, do

Pitulko et al. not mean that the present authors are trying to infer a direct link between the Yana culture and modern Yukaghirs. These similarities nonetheless demonstrate that some cultural ideas may survive long enough to contribute to historical or even present-day culture. The many similarities existing between the Yana RHS finds and those from geographically and chronologically distant regions of Eurasia are not likely the result of convergence. There are too many of them to be explained by mere coincidence. These similarities are “patterns that connect” (Schuster and Carpenter 1996) to an earlier Paleolithic cultural base that includes Upper Paleolithic innovations and inherited lithic traditions from the Middle Paleolithic. They are a manifestation of cultural continuity across Eurasia. The Yana RHS culture has common ancestry with many Siberian cultures, and its roots definitely lie in the Yenisei and Transbaikal regions.

Conclusion

The Yana RHS testifies to the high level of spiritual and technological development attained by the peoples of western Beringia before the LGM. It is the oldest evidence of elaborate symbolic activity known to date north of the Arctic Circle. Some finds suggest the existence of shamanic cults among the Upper Paleolithic settlers of East Siberia. At the same time, it clearly demonstrates cultural unity with the Eurasian Upper Paleolithic in general. Yana RHS represents the oldest known phase of human habitation in the western Beringia. It is followed by a Beringian microblade phase beginning around the time of the LGM. Despite the harsh environmental conditions, western Beringia most likely did not become depopulated at the LGM. Instead, these conditions drove fauna and human migrations, which probably explains why the only visible pre-Holocene cultural link—the Chindadn connection—is located in the northern part of western Beringia. However, discovery of Yana RHS does not take us closer to answering the question, Who were the first Americans and when did they happen to discover the New World? This question still stands. Yana does not help to answer it directly, but it offers an opportunity to look for the answer because it is unequivocal evidence that a pre-LGM human migration into the New World was possible. By at least 29,000–28,000 14C yr BP humans were already in arctic western Beringia, and nothing would have prevented them from crossing the Bering Land Bridge. But did they?

Acknowledgments

This research had been performed as a part of the ZhokhovYana project, a long-term Russian–U.S. effort funded by Rock Foundation, New York, since 2000. As a part of the Fundamental Research Program run by Russian Academy of Sciences, it is also supported by the Academy through the Cultural Heritage Fundamental Research project and by the Russian Foundation for Basic Research (grant No. 11-06-12018). The authors are thankful to Sergey Kritsuk (topographic and

Human Habitation in Arctic Western Beringia Prior to the LGM

41

mapping data processing), Pavel Ivanov (photography), Alla Mashezerskaya (graphic drawings, bone & ivory), Anastasiya Abdul’manova (graphic drawings, lithic tools) and Veronika Stegantseva (image processing). Enormous thanks to Elena Chekhova and Svetlana Burshneva (conservation). And certainly authors are thankful to all participants of the fieldwork in Yana RHS. Special thanks to VICAAR/NORPOLEX company for logistics support.

Bradley, B., and D. Stanford  2004  The north-Atlantic ice-edge corridor: A possible Paleolithic route to the New World. World Archaeology 36(4):459–78.

References Cited

Derevianko, A. P., V. I. Molodin, V. N. Zenin, S. V. Leschinsky, and E. N. Maschenko 2003a Site du Paleolithique Recent Chestakovo. Institute of Archaeology and Ethnography Press, Novosibirsk. (In Russian.)

Abramova, Z. A.  1979  K voprosu o vozraste aldanskogo paleolita [On the question of the age of the Paleolithic in Aldan]. Soviet Archaeology 4:5–14. (In Russian.) ———  1989  Paleolit Severnoi Azii [The Paleolithic of Northern Asia]. In The Paleolithic of the Caucasus and Northern Asia, edited by P. I. Boriskovskiy, pp. 145–255. Nauka, Moscow. ———  1995  L’art Paléolitique d’Europe Orientale et de Sibérie. Gérôme Millon Editions, Grenoble. ———  2003–2004  A catalogue of depictions of mammoth in European Palaeolithic miniature art. Stratum Plus 1:23–67. (In Russian.) Akimova, E. V., and N. I. Drozdov  2005  Arkheologiya [Archaeology]. In Listvenka, edited by A. P. Derevianko, pp. 38–133. Institute for Archaeology and Ethnography SB RAS, Krasnoyarsk, (In Russian.) Anikovich, M. V.  2003  The Early Upper Paleolithic in Eastern Europe. Archaeology, Ethnography, and Anthropology of Eurasia 2(29):15– 29. (In Russian.) ———  2007  Upper Paleolithic Origins in Eastern Europe and in Gorny Altai. Archaeology, Ethnography, and Anthropology of Eurasia 1(14): 2–15. (In Russian.) Atlas Arktiki [Atlas of the Arctic]  1985  GUGK, Moscow. (In Russian.). Bar-Yosef, O., and S. L. Kuhn  1999  The big deal about blades: Laminar technologies and human evolution. American Anthropologist 101:322–38. Bader O. N.  1998  Sungirskie Paleoliticheskie pogrebeniya [Sunghir Paleolithic burials]. In Upper Paleolithic site Sunghir: Burials and environment, edited by N. N. Bader and Y. A. Lavrushin, pp. 5–150. Scientific World, Moscow (In Russian.) Basilyan, A. E., M. A. Anisimov, P. A. Nikolskiy, and V. V. Pitulko  2011 Woolly mammoth mass accumulation next to the Palaeolithic Yana RHS site, Arctic Siberia: Its geology, age, and relation to past human activity. Journal of Archaeological Science 38:2461–74. Basilyan, A. E., M. A. Anisimov, E. Y. Pavlova, V. V. Pitulko, and P. A. Nikolsky  2009  Base reference section of Quaternary deposits for Yana-Indighirka lowland in the Yana River downstream area. In Fundamental Problems of Quaternary: Results and Trends of Further Research, edited by A. E. Kontorovich, V. S. Volkova, I. V. Khazina and L. B. Khazin, pp. 63–65. Proceedings of the VI All-Russian Quaternary Conference. October 19–23, 2009. Russian Academy of Sciences, A. A. Trofimuk Institute for Oil & Gas Geology and Geophysics, Novosibirsk. (In Russian.) Bonnichsen, R., and K. L. Turnmire (eds)  1999  Ice Age Peoples of North America. Center for the Study of the First Americans, Oregon State University Press, Corvallis. Bradley, B.  1995  Clovis ivory and bone tools. In Le Travail et l’Usage de l’Ivoire au Paleolithique Supereiur, edited by J. Hahn, M. Menu, Y. Taborin, P. Walter, and F. Widemann, pp. 114–25. Centro Universitario Europeo per I Bene Culturali. Actes de la Table Ronde, Ravello.

CAVM Team  2003  Circumpolar Arctic Vegetation Map. Scale 1:7,500,000. Conservation of Arctic Flora and Fauna (CAFF) Map No. 1. U.S. Fish and Wildlife Service, Anchorage, Alaska. Derevianko, A. P., and E. P. Rybin  2003  The oldest evidence for symbolic behavior of the Palaeolithic men in Gorny Altai. Archaeology, Ethnography, and Anthropology of Eurasia 3:27–50 (In Russian.)

Derevianko, A. P., M. V. Shunkov, A. K. Agadjanian, G. F. Baryshnikov, E. M. Malaeva, V. A. Ulianov, N. A. Kulik, A. V. Postnov, and A. A. Anoikin  2003b  Paleoenvironment and Paleolithic human occupation of Gorny Altai. Subsistence and Adoptation in the Vicinity of Denisova Cave. Institute of Archaeology & Ethnography SB RAS Press, Novosibirsk. Dikov, N. N.  1965  The stone age of Kamchatka and the Chukchi Peninsula in the light of new archaeological data. Arctic Anthropology 3(1):10–25. ———  1977  Archaeological sites in Kamchatka, Chukotka, and Upper Kolyma. Nauka, Moscow. (In Russian.) ———  1979  Ancient Cultures of North-Eastern Asia. Nauka, Moscow. (In Russian.) ———  1993  Asia at the Juncture with America in Antiquity. Nauka, St Petersburg. (In Russian.) Dixon, J. E.  1999  Bones, Boats & Bison: Archaeology and the First Colonization of Western North America. University of New Mexico Press, Albuquerque. Dolukhanov, P. M.  2008  Evolution of environment and early human dispersal in northern Eurasia. In Way to North. Paleoenvironment and Earliest Inhabitants of Arctic and Sub-Arctic, edited by A. A. Velichko and S. A. Vasil’iev, pp. 33–47. Institute of Geography RAS, Moscow. Dunbar, J., S. Webb, and D. Cring  1989  Culturally and naturally modified bones from a Paleoindian site in the Aucilla River, north Florida. In Bone Modification, edited by R. Bonnichsen and M. Sorg, pp. 473–97. Center for the Study of the First Americans, Institute for Quaternary Studies, University of Main, Orono. Eren M.I ., A. Greenspan, and C. G. Sampson  2008  Are Upper Paleolithic blade cores more productive than Middle Paleolithic discoidal cores? A replication experiment. Journal of Human Evolution 55:952–61. Fedoseeva, S. A.  1980  Ymyayhtahskaya kul’tura Severo-Vostochnoi Azii [Ymyiakhtakh culture of North-East Asia]. Nauka, Novosibirsk. (In Russian.) Fedoseeva, S. A.  1992  Diring-Yuryahskiy mogil’nik [DiringYuryakh Cemetery]. In Archaeological research in Yakutiya, edited by Y. A. Mochanov, pp. 84–105. Nauka, Novosibirsk.(In Russian.) Gakkel, Y. Y. (ed.)  1967  Novosibirskie ostrova [New Siberian islands]. Gidrometeoizdat, Leningrad. (In Russian.) Gakkel, Y. Y., and E. S. Korotkevich  1960  Severnaya Yakutiya (fiziko-geograficheskaya kharakteristika [Northern Yakutia (physical geography)]. Morskoy transport, Leningrad. (In Russian.) Goebel, T.  2002  The “Microblade Adaptation” and Recolonization of Siberia during the late Upper Pleistocene. In Thinking Small: Global Perspectives on Microlithization, edited by R. G. Elston and S. L.

42

Pitulko et al.

Kuhn, pp. 117–31. Archaeological Papers of the AAA Vol. 12. American Anthropological Association, Arlington.

potheses)], edited by L. V. Lbova, pp. 126–33. Institute of Archaeology & Ethnography SB RAS Press, Novosibirsk. (In Russian.)

Goebel, T., and I. Buvit (eds  2011  From the Yenisei to the Yukon: Interpreting Lithic Assemblage Variability in Late Pleistocene/Early Holocene Beringia. Center for the Study of the First Americans, Texas A&M University Press, College Station.

Konstantinov, A. V.  2001  Drevnie zhilischa Zabaikalya [Ancient Dwelling of Zabaikalye]. Nauka, Novosibirsk (In Russian.)

Goebel, T., M. R. Waters, and M. Dikova  2003  The Archaeology of Ushki Lake, Kamchatka, and the Pleistocene peopling of the Americas. Science 301:501–05. Goebel, T., S. B. Slobodin, and M. R. Waters  2010  New dates from Ushki-1, Kamchatka, confirm 13,000 cal BP age for earliest Paleolithic occupation. Journal of Archaeological Science 37:2640–49. Graf, K. E.  2011  On Late Upper Paleolithic Variability in SouthCentral Siberia: Rethinking the Afontovo and Kokorevo Cultures. 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. 33–46. Texas A&M University Press, College Station. Gribchenko, Y. N.  2008  Initial colonization of Paleolithic and neolithic Humans on the NE of the east European plain . In Way to North. Paleoenvironment and earliest inhabitants of Arctic and Sub-Arctic, edited by A. A. Velichko and S. A. Vasil’iev, pp. 107–33. Institute of Geography RAS, Moscow. Ineshin, E. M., and A. V. Teten’kin  2010  Humans and environment in the North of Baikalian Siberia in the Late Pleistocene. Archaeological site Bol’shoi Yakor’ I. Nauka, Novosibirsk. (In Russian.) Iochelson, W. I.  2005  The Yukaghir and the Yukaghirized Tángus. Nauka, Novosibirsk (In Russian.) Ivanov, S. V.  1954  Materialy po izobrazitelnomu iskusstvu narodov Sibiri XIX–nachala XX veka. [Data on ornamental art of Siberia in XIX–beginng of the XXs century]. Izdatelstvo Akademii Nauk SSSR, Moskva–Leningrad. (In Russian.) ———  1970  Skulptura narodov Severa Sibiri [Sculpture of the North Siberian natives]. Nauka, Leningrad. (In Russian.) Izyumenko, S. A.  1966  Spravochnik po klimatu SSSR. Vypusk 24. Yakutskaya ASSR. Chast II. Temperatura vozdukha i pochvy [Handbook on the climate of the USSR. Issue 24. Yakut ASSR. Part II. Air and soil temperatures]. Gidrometeoizdat, Leningrad. (In Russian.) ———  1968  Spravochnik po klimatu SSSR. Vypusk 24. Yakutskaya ASSR. Chast IV. Vlaznost vozdukha, atmosfernye osadki i snezhnyi pokrov [Handbook on the climate of the USSR. Issue 24. Yakut ASSR. Part IV. Atmospheric moisture, precipitation and snow cover]. Gidrometeoizdat, Leningrad. (In Russian.) Katasonov, E. M. (ed.)  1979  Stroenie i absolyutnaya geokhronologiya alasnykh otlozhenii Tsentral’noi Yakutii [Structure and Absolute Geochronology of Alassy Deposits of Central Yakutia]. Nauka, Novosibirsk. (In Russian.) Kiryak M. A., O. Y. Glushkova, and T. A. Brown  2003  Verhnepaleoliticheskie kompleksy v doline reki Tytylvaam (Zapolyarnaya Churotka) [Upper Paleolithic complexes in Tytylvaam River valley, Arctic Chukotka]. Archaeology, Ethnography, and Anthropology of Eurasia 3:2–15. (In Russian.) Khlopatchev, G. A.  2006  Les Industries des Ivories du Paléolitique Supériore de l’Europe Orientale. Nauka, Saint Petersburg. (In Russian.) Klement’ev, A. M.  2005  Kopytnye mlekopitayushchie i mamont v paleolite zapadnogo Zabaikal’ya [Hoofed mammals and mammoths in the Paleolithic of Western Zabaikal’e]. In Paleoliticheskie kul’tury Zabaikal’ya i Mongolii (novye pamyatniki, metody, gipotezy) [Paleolithic Cultures of Zabaikal’e and Mongolia (New Sites, Methods, and Hy-

———  1994  Kamennyi vek Vostochnogo regiona Baikal’skoi Azii [The Stone Age of the Eastern Region of the Baikal Area of Asia]. BNTs SO RAN, Ulan Ude, Chita. (In Russian.) Lbova, L. V.  2000  Paleolit Severnoi zony Zapadnogo Zabaikal’ya [The Paleolithic of the Northern Zone of Western Zabaikal’e]. BNTs SO RAN, Ulan Ude. (In Russian.) Lisitsyn, N. F.  2000  Late Paleolithic of the Chulym-Yenisey interfluvial. Institute for the History of Material Culture RAS, St. Petersburg. (In Russian.) 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 Communnications 3:893 doi: 10.1038/ncomms1881(2012). Malaurie, J. P.  1989  Les Derniers Rois de Thulé. Plon, Paris. Medvedev, G. I.  1998  Art from central Siberian Palaeolithic sites. In The Paleolithic of Siberia: New Discoveries and Interpretations, edited by A. P. Derevianko, pp.132–37. Institute of Archaeology & Ethnography SB RAS, Novosibirsk; University of Illinois Press, Urbana (IL). Miller, G. A.  1956  The magical number seven, plus or minus two: Some limits on our capacity for processing information. Psychological Review 63:81–97. Mochanov, Y. A.  1969  Dyuktaiskaya verhnepaleoliticheskaya kultura I nekotorye aspekty ee genesisa [Dyuktai Upper Paleolithic culture and some questions of its origin]. Soviet Archaeology 4:235–39 (In Russian.). ———  1977  Drevneishie etapy zaseleniya chelovekom SeveroVostochnoi Azii [Earliest stages of the Peopling of North-Eastern Asia]. Nauka, Novosibirsk. (In Russian.) ———  1992  Drevneishii paleolit Diringa I problema vnetropicheskoi prarodiny chelovechestva [The Earliest Paleolithic of Diring and the Problem of Extratropical Emergence of Humankind]. Nauka, Novosibirsk. Mochanov, Y. A., and S. A. Fedoseeva  2002  Arkheologiya, paleolit Severo-Vostochnoi Azii, vnetropicheskaya prarodina cheloveka I drevneishie etapy zaseleniya chelovekom Ameriki [Archaeology, Paleolithic of Noreastern Asia, Extratropical Emergence of Humankind, and the Oldest Stages of the Peopling of the Americas]. Yakutia (Sakha Republic) Academy of Sciences, Yakutsk. Nikolskiy, P. A. and V. V. Pitulko  2012  Data from Upper Palaeolithic Yana RHS site, Arctic Siberia, suggest some clues to recognize mammoth hunt in those sites that have mammoth remains. Paper presented at the International Conference “European Middle Palaeolithic during MIS 8–MIS 3. Cultures–environment–chronology.” Wolbrom, Poland, 25–28 September 2012. Nikolskiy, P. A., L. D. Sulerzhitsky, and V. V. Pitulko  2011  Last straw versus Blitzkrieg overkill: Climate-driven changes in the Arctic Siberia mammoth population and the late Pleistocene extinction problem. Quaternary Science Reviews 30:2309–28. Pavlov, P., J. I. Svendsen, and S. Indrelid  2001  Human presence in the European Arctic nearly 40,000 years ago. Nature 413:64–67. Pavlova, E. Y., M. V. Dorozhkina, and V. V. Pitulko  2012  Environmental and climate changes in the west of Yana-Indigirka lowland in the end of late Pleistocene (reconstruction on palaebotanical data). In Proceedings of the Joint Conference “Geomorphology and Qua-

Human Habitation in Arctic Western Beringia Prior to the LGM ternary Palaeogeography of Polar Regions”, Symposium “Leopoldina” and the INQUA Peribaltic working group Workshop, edited by A. Zhirov, V. Kuznetsov, D. Subetto, and J. Thiede, pp. 310–12. Saint Petersburg, SPbSU, 9–17 September 2012. Saint Petersburg State University, Saint Petersburg. (In Russian.) Peterkin, G. L.  1993  Lithic and organic hunting technology in the French Upper Paleolithic. In Hunting and Animal Exploitation in the Later Palaeolithic and Mesolithic of Eurasia, edited by G. L. Peterkin, H. M. Bricker, and P. Mellars. Archaeological Papers of the AAA 4:49-68. American Anthropological Association, Arlington. Pitulko, V. V.  2006  The Stone Age cultural chronology of north eastern Asia. In Papers of the 2nd Northern Archaeological Congress, edited by A. V. Golovnev, pp. 307–32. Charoid Publishers, Ekaterinburg and Khanty-Mansiisk. ———  2008  Principal excavation techniques under permafrost conditions (based on Zhokhov and Yana sites, Northern Yakutia ). Archaeology, Ethnography, and Anthropology of Eurasia 2:26–33. ———  2010  Rasselenie i adaptatsii drevenego cheloveka v pozdnem pleistocene Severo-Vostochnoy Azii [Late Pleistocene habitation and adaptations of the Ancient Man across NE Asia]. In Adaptations of Peoples and Cultures to Natural Changes, Social and Man-caused Transformations, edited by A. P. Derevianko, A. B. Kudelin, and V. A. Tishkov, pp. 38–46. ROSSPEN, Moscow. (In Russian.) Pitulko, V. V., and V. V. Ivanova   2010  Production and use of the red mineral dye on the Paleolithic camp site Yanskaya, Eastern Yakutia. In Papers of the Third Northern Archaeological Congress. Abstracts, edited by A. V. Golovnev, pp.46–47. Charoid, Ekaterinburg and Khanty-Mansiisk. (In Russian.) Pitulko, V. V., and E. Y. Pavlova  2007  Radiocarbon age of the Paleolithic sites in Yana-Indigirka lowland and peculiarities of the radiocarbon dating of ice complex deposits. In Radiocarbon in Archaeological and Paleoenvironmental Research, edited by G. I. Zaitseva and M. N. Kulkova, pp. 155–63. Institute for the History of Material Culture RAS, St. Petersburg.(In Russian.) ——— 2010 Geoarchaeology and Radiocarbon Chronology of the Stone Age of the North-East Asia. Nauka, St. Petersburg. (In Russian.) ———  2012  Permafrost as an archaeological environment. In Geomorphic Processes and Geoarchaeology: From Landscape Archaeology to Archaeotourism, edited by M. Bronnikova and A. Panin, pp. 224–27. International conference held in Moscow-Smolensk, Russia, 20–24 August 2012. Extended abstracts Universum Publishers, MoscowSmolensk. Pitulko, V. V., and P. A. Nikolskiy  2012  The extinction of the ­woolly mammoth and the archaeological record in North-eastern Asia. World Archaeology 44(1):21–42. Pitulko, V. V., E. Y . Pavlova, and S. G. Kritsuk  2012  Alluvial complex structure and development dynamics in low-stream Yana River valley in the vicinity of Paleolithic Yana RHS site. In Proceedings of the Joint Conference “Geomorphology and Quaternary Palaeogeography of Polar Regions,” Symposium “Leopoldina” and the INQUA Peribaltic working group Workshop, edited by A. Zhirov, V. Kuznetsov, D. Subetto and J. Thiede, pp. 313–16. Saint Petersburg, SPbSU, 9–17 September 2012. Saint Petersburg State University, Saint Petersburg. (In Russian.) Pitulko, V. V., P. A. Nikolsky, A. E. Basilyan, and E. Y. Pavlova  2011 Dating mass accumulations of mammoth across Arctic Eurasia. In The Quaternary in all of its variety. Basic issues, results, and major trends of further research, edited by O. P. Korsakova, V. V. Kolka, and L. D. Chistyakova, pp.146–49.Proceeding of the VII All-Russian Quaternary Conference. Apatity, 12–17 September 2011. Volume 2 Geological Institute, Kola Research Center RAS, Apatity and St Petersburg.

43 Pitulko, V. V., E. Y. Pavlova, A. E. Basilyan, and S. G. Kritsuk  2011 Features of vertical distribution of the materials within the marginal zones of permafrost polygonal structures and its importance for dating of quaternary deposits in cryolitozone. In Proceeding of the VII AllRussian Quaternary Conference. The Quaternary in all of its variety. Basic issues, results, and major trends of further research, Apatity, edited by O. P. Korsakova, V. V. Kolka, and L. D. Chistyakova, pp. 149–53, 12–17 September 2011. Volume 2, Geological Institute, Kola Research Center, RAS, Apatity and St Petersburg. (In Russian.) Pitulko, V. V., E. Y. Pavlova, P. A. Nikolskiy, and V. V. Ivanova  2012 Simvolicheskaya deyatel’nost’ verkhnepaleoliticheskogo naseleniya Arkticheskoi Sibiri (busy I podveski Yanskoi stoyanki) [Symbolic behavior of the Upper Paleolithic inhabitants of Arctic Siberia (beads and pendants from the Yana site)]. In Historical & Cultural Heritage and Intellectual Values in Russia, edited by A. P. Derevianko, A. B. Kudelin, and V. A. Tishkov, pp. 35–51. ROSSPEN, Moscow. ———  2012  The oldest art of Eurasian Arctic. Antiquity 86(333):642–59. Pitulko, V. V., P. A. Nikolsky, E. Y. Girya, A. E. Basilyan, V. E. Tumskoy, S. A. Koulakov, S. N. Astakhov, E. Y. Pavlova, and M. A. Anisimov 2004  The Yana RHS site: Humans in the Arctic before the last glaciation. Science 303:52–56. Pitulko, V. V., E. Y. Pavlova, S. A. Kuzmina, P. A. Nikolskiy, A. E. Basilyan, V. E. Tumskoy, and M. A. Anisimov  2007  Natural–climatic changes in the Yana–Indigirka lowland during the terminal Kargino time and habitat of late Paleolithic man in northern part of east Siberia. Doklady Earth Sciences 417:1256–60. Romanovskiy, N. N.  1993  Fundamentals of Cryogenesis of Lithosphere. Moscow University Press, Moscow. (In Russian.) Rybin, E. P., S. A. Gladyshev, and A. A. Tsybankov  2007  Vozniknovenie i razvitie “otshchepovykh” industrii rannei pory verkhnego paleolita Severnoi Mongolii” [The emergence and development of “flaking” industries in the early part of the Upper Paleolithic of northern Mongolia]. In Northern Eurasia in the Anthropogene: Man, Paleotechnologies, Geoecology. Materials from the All-Russia Conference with International Participation, Dedicated to the 100th year of the birth of M. M. Gerasimov. Vol. 2., edited by G. I. Medvedev, pp. 137–53. Ottisk, Irkutsk. (In Russian.) Schuster, C., and E. Carpenter  1996  Patterns That Connect. Abrams Publishers, H. N. New York. Sher, A. V., and I. R. Plakht  1988  Radiouglerodnoe datirovanie i problemy stratigrafii pleistotsena nizmennostei Severo-Vostoka SSSR [Radiocarbon dating and problems of Pleistocene stratigraphy of the lowlands of the northeastern USSR]. Izv. AN SSSR. Geology Series 8:17–31. (In Russian.) Sitlivy, V., G. I. Medvedev, and E. A. Lipnina  1997  Le Paleolithique de la Rive Occidentale du Lac Baikal. Les Civilisations Prehistoriques D’Asie Centrale. N 1. Bruxelles. Slobodin, S. B.  1999  Archaeology of Kolyma and continental Priokhotye in Late Pleistocene and Early Holocene. North-East Interdisciplinary Scientific Research Institute, Magadan. (In Russian.) ———  2011  Late Pleistocene and early Holocene cultures of Beringia: The general and the specific. 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. 91–118. Texas A&M University Press, College Station. Smith, P. E.  1966  Le Solutrean en France. Bordeaux. Sokolov, D. D., L. D. Sulerzhitsky, and V. N. Tutubalin  2004  Vremya aktivnosti lyudei na paleoliticheskikh pamyatnikakh po dannym

44

Pitulko et al.

radiouglerodnogo datirovaniya [The time of activity of people at paleolithic sites based on data from radiocarbon dating]. Soviet Archaeology 3:99–102. (In Russian.)

Vereschagin, N. K., and Y. A. Mochanov  1972  Samye severnye v mire sledy verhnego paleolita [The world northernmost traces of the Upper Paleolithic]. Soviet Archaeology 3:332–36 (In Russian.)

Stepanov, A. D., A. S. Kirillin, S. A. Vorobei, E. N. Solovyeva, and N. N. Efimov  2003  Peschera Haiyrgas na Srednei Lene (rezultaty issledovaniy 1998-1999 g. [Khaiyrgas Cave on the Middle Lena (results of the 1998–1999 research program)]. In Ancient culture of Northeastern Asia. Astroarcheology. Paleoinformatics, edited by A. N. Alekseev, pp. 98–113. Nauka, Novosibirsk. (In Russian.)

Waters, M. R., S. L. Forman, and J. M. Pierson  1997  Diring Yuriakh: A lower Paleolithic site in central Siberia. Science 275:1281–84.

Sulerzhitsky, L. D.  2004  Vremya sushchestvovaniya nekotorykh pozdnepaleoliticheskikh poselenii po dannym radiouglerodnogo datirovaniya kostei megafauny [The time of existence of some late Paleolithic settlements based on data of radiocarbon dating of bones of megafauna]. Soviet Archaeology 3:103–12. (In Russian.) Tashak, V. I.  2009  Symbolism in the early Upper Palaeolithic of western Transbaikalia. Transactions of the Institute for the History of Material Culture 4:50–62. (In Russian.) Tugolukov, V. A.  1979  Kto vy, yukagiry? [Who are you, Yukaghirs?] Nauka, Moscow. (In Russian.) Vanhaeren, M., and F. d’Errico  2006  Aurignacian ethno-linguistic geography of Europe revealed by personal ornaments. Journal of Archaeological Science 33:1105–28. Vasil’ev, S. A.  2001  The final Paleolithic in northern Asia: Assemblage diversity and explanatory models. Arctic Anthropology 38(2):3–30. Velichko, A. A., and S. A. Vasil’iev (eds)  2008  Way to North: Paleo­ environment and the Earliest Inhabitants of Arctic and Sub-Arctic. Institute of Geography RAS, Moscow. Vereshchagin, N. K.  1977  Berelekh cemetery of mammoths. Proceedings of Zoological Institute 72:5–49. (In Russian.)

———  1999  Late Quaternary geology and geochronology of Diring Yuryakh, an Early Paleolithic site in central Siberia. Quaternary Research 51:195–211. West, F. H. (ed.)  1996  American Beginnings. University of Chicago Press, Chicago. White, R.  1995  Ivory personal ornaments of Aurignacian age: Technological, social, and symbolic perspective. In Le Travail et l’usage de l’ivoire au Paleolithique Supereiur, edited by J. Hahn, M. Menu, Y. Taborin, P. Walter, and F. Widemann, pp. 29–62. Centro Universitario Europeo per I Bene Culturali. Actes de la Table Ronde, Ravello. ———  1997  Intégrer la complexité sociale et opérationnalle: La construction matériale d’identité sociale à Sungir, edited by H. Camps-Fabrer, pp. 319–33. Préhistoire d’os. L’Universite de Provence, Provence. Wygal B. T.  2011  The microblade/non-microblade dichotomy: Climatic implications, toolkit variability, and the role of tiny tools in eastern Beringia. 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. 234–54. Texas A&M University Press, College Station. Yi, S., and G. Clark  1985  The “Dyuktai culture” and New World Origins. Current Anthropology 26(1):1–20. Yurtsev B. A., A. I. Tolmachev, and O. B. Rebristaya  1978  Floristicheskoye ogranichenie i razdelenie Arktiki [Floristic limitations of the Arctic region and its subdivisions]. In The Arctic Floristic Region, edited by B. A. Yurtsev, pp.9–104. Nauka, Leningrad. (In Russian.)

 Chapter 3 Human Technological and Behavioral Adaptation to Landscape Changes around the Last Glacial Maximum in Japan: A Focus on Hokkaido Masami Izuho ABSTRACT Here I discuss inter-assemblage variability of Last Glacial Maximum (LGM) sites on Hokkaido based on the integration of ecological and archaeological data. Of four geochronologically reliable sites relevant to the LGM, three, Minamimachi 2, Marukoyama, and Kawanishi-C, are concentrated at around 26,000 cal yr BP, while dates from Kashiwadai 1 range between 27,900 and 22,100 cal yr BP. Evaluation of both ecological and behavioral data (site location, assemblage size, toolkit diversity, as well as lithic reduction) lead to the following conclusions: (1) occupations at LGM sites were relatively brief, but detailed studies reveal that Kashiwadai 1 Lithic Concentration (LC) 11, Shimaki, and Kawanishi-C were relatively long-term processing sites, while Minamimachi 2 and Marukoyama were relatively short-term processing sites; (2) both blade and flake reduction concurrently contributed to inter-site variability; (3) it is possible that either these data represent two distinct coeval groups of hunter-gatherers practicing different behavioral strategies, one equipped with a blade/microblade-based toolkit and the other with a flake/blade-based toolkit, or they represent discrete behavioral strategies left behind at slightly different times. As such, a discussion of LGM hunter-gatherer adaptations in Hokkaido has important implications for the origin and route of the peopling of the Americas, as well as for the adaptational processes needed to survive in the New World. KEYWORDS: Last Glacial Maximum, Hokkaido, Upper Paleolithic, Geochronology, Flake-, blade-, and blade/ microblade-based toolkits.

Introduction

Technological studies of microblade industries in Hokkaido are thought to be of crucial importance to understanding the origin and dispersal of prehistoric human groups in northeastern Asia and North America (Hayashi 1968; Kato 1985; Kimura 1995; Morlan 1967). Recent work in Hokkaido, the northernmost Japanese island, has produced archaeological data conF aculty of Social Sciences and Humanities, Tokyo Metropolitan University. Minami-Osawa 1-1, Hachioji-shi, Tokyo 192-0397 Japan; e-mail: [email protected]

cerning local cultural sequences and changes in technological organization (Izuho and Akai 2005; Izuho and Takahashi 2005; Izuho et al. 2012; Nakazawa et al. 2005). Initially, smallflake industries, termed the Wakabanomori and ShukubaiSankakuyama (perhaps dating to 30,000 14C yr BP), emerged in deciduous broadleaf forests that supported a faunal community accompanied by Naumann’s elephant. In contrast, a later microblade industry (e.g., at Kashiwadai 1, LC 15), a blade industry (Kawanishi-C), and a flake industry (Kashiwadai 1, LC 11), all dating between 22,000 and 20,000 14C yr BP, were associated with sub-arctic forested steppe into which woolly mam-

46 moths migrated via a land bridge from the Russian Far East and Sakhalin. After the LGM around 18,000 14C yr BP, various microblade assemblages, associated with the pleniglacial period and Younger Dryas, emerged coincident with a gradual decrease of the forest and abrupt recovery of the grass steppe. However, the relationships between environmental changes and various cultural complexes are not solid, and debates among Paleolithic researchers in Hokkaido are ongoing (Sato 2003; Yamada 2006; Nakazawa et al. 2005). Recent progress with regard to questions about humanenvironment interaction has come from research of ancient dna. In Adachi et al. (2011), the control and coding regions of mtdna from prehistoric Jomon skeletons from Hokkaido were analyzed in detail, and haplogroups N9b (coalescence time ca. 22,000 cal yr BP), D4h2, G1b, and M7a were observed. Results suggest that most of Hokkaido’s Jomon people were direct descendants of Paleolithic Siberians and that their migration into Hokkaido increased during the LGM. In terms of Siberia, Graf (2008, 2009a,b) concluded that human groups may have migrated during the LGM to the south of Siberia and to areas such as Hokkaido, based on evaluation of radiocarbon dates of Paleolithic sites in the upper Yenisei River valley. Graf (2009b) also insisted that recolonization of Siberia by human groups equipped with true microblade technology only occurred after the LGM. These studies indicate that further work on the Upper Paleolithic of Hokkaido will contribute not only to the local prehistory of the island itself, but to the issues of human adaptation to northern environments and ultimately the peopling of the Americas. It is reasonable to suggest that future research on the rich Upper Paleolithic sequences from Hokkaido can contribute to discussion of the origins of the First

Izuho Americans and their routes to the New World, as well as shed light on how human groups technologically and behaviorally adapted to different ecological zones.

Regional Setting Modern Hokkaido Hokkaido, one of the biggest islands of the Japanese Archipelago, is situated along the eastern margin of the Asian continent. The geographic coordinates of the island lie between 41° 24′–45° 30′ N and 139° 20′–145° 48′ E (Figure 3.1). Today the climate of Hokkaido varies from temperate in the southwest to cool-temperate in the northeast. The temperature is affected by seasonal westerly winds from Asia during winter and seasonal easterly winds from the Pacific Ocean during summer. In addition, sea currents flowing north from the Pacific Ocean and south from the Sea of Okhotsk add variability to the local climate of Hokkaido. Distance in a straight line between northernmost Hokkaido and Yakutsk is about 2000 km. Likewise, Uelen, the eastern end of the Chukchi Peninsula, is about 3600 km from Hokkaido. The Last Glacial Maximum Hokkaido Analysis of sea-bottom sediment cores, including from the Oki Ridge in the Sea of Japan, suggests that the LGM started around 30,000 cal yr BP and ended about 19,000 cal yr BP (Lambeck et al. 2002; Yokoyama et al. 2007). Also, results of pollen analysis, and organic carbon and nitrogen from sediment cores in Lake Nojiri, central Honshu, show that the duration of LGM was ca. 29,000–18,000 cal yr BP (Kumon et al. 2003, 2009). In Siberia, evidence from continental glaciers and loess-

2 1

Figure 3.1  Map of Hokkaido and LGM sites mentioned in the text: 1, Kashiwadai 1; 2, Shimaki; 3, Minamimachi 2; 4, Marukoyama; 5, Kawanishi-C.

4

3 5

Human Technological and Behavioral Adaptation to Landscape Changes around the Last Glacial Maximum in Japan

47

paleosol sequences suggests that the LGM started close to 24,000 cal yr BP and ended at nearly 21,500 cal yr BP (Chlachula 2001; Svendsen et al. 2004). This relatively short LGM chronology roughly corresponds to the Gydansk Stadial (25,000–20,000 years ago) of the Sartan Glacial in Siberian Quaternary geochronology (Kind 1974). These data from Japan and Siberia suggest that a large gap exists in our knowledge of the timing of the onset of the LGM. Although further work on this topic is required, these differences are probably due to variation in the timing, duration, and intensity of LGM environments across the globe; moreover, climatic data from the Sea of Japan are of high resolution, whereas the same level of data are not yet available for the terrestrial environments of continental Siberia. In this paper, I consider LGM paleoenvironmental data from both regions and place them in the context of Hokkaido’s paleoclimatic change and its effects on past human behavior. The environmental setting of late-Pleistocene Hokkaido has been discussed in Nakazawa et al. (2005), Izuho and Takahashi (2005), and Izuho et al. (2012) in English. Here I mainly focus on the LGM. The Paleo-SHK Peninsula  Figure 3.2 shows the landscape and vegetation in and around Hokkaido during the LGM. The landscape of Hokkaido dramatically changed from the last interglacial to the Holocene owing to a eustatic drop in sea level of 105–130 m. During almost the entire the late Pleistocene, including the LGM, two major landmasses were maintained in and around the Japanese Archipelago. Hokkaido was connected with Sakhalin and the Asian continent at the mouth of Amur River (Ono 1990; Sato 2003; Vasilevski 2005, 2008; Izuho et al. 2012; Machida et al. 2004). This geographical district is called Paleo-Sakhalin/Hokkaido/Kurile (SHK) Peninsula. On the other hand, Shikoku, Kyushu, and Honshu formed the Paleo-Honshu Island, where land bridges did not emerge at either the Tsugaru Strait (between Hokkaido and Honshu) or the Korea Strait during the late Pleistocene. Below, I use “southern half of Paleo-SHK Peninsula” to designate Hokkaido (Sato 2003; Vasilevski 2008; Izuho et al. 2012; Sato et al. 2011). The ancient LGM shoreline is estimated to have extended about 10 km farther from the present coast. It is presumed that the width of the Tsugaru Strait was reduced to less than 2 km. A bifacial stemmed point and mammoth molars were caught with a trawl net by fishermen several kilometers off Erimo Cape and Kushiro of eastern Hokkaido (Nishi 1991; Takahashi et al. 2006), indicating that Paleolithic sites are probably preserved on the sea bottom along the Hokkaido coast. The LGM sea ice that covered the northern part of the Sea of Japan and coastal area of the Sea of Okhotsk is assumed to have had a different distribution from the present. Diatom analysis shows that the western part of the Sea of Okhotsk, including the northeastern coastal area of Hokkaido, was covered with perennial sea ice (Shiga and Koizumi 2000). The northern part of the Sea of Japan was also covered

Figure 3.2  Map showing the LGM geography and vegetation of Paleo-SHK Peninsula (modified from Izuho et al. 2012 and Igarashi 2008).

with perennial sea ice, because the much shallower Tsushima Strait blocked the northward penetration of the Tsushima current, resulting in remarkably decreased sea-surface temperature (Matsui et al. 1998). Since Pacific Ocean currents rarely influenced the Sea of Japan, it was a lake-like enclosed sea (Shiga and Koizumi 2000). Major currents, including the Kuroshio, were at quite a distance from the coasts. This resulted in a continental climate over the entire archipelago, with colder and drier conditions, unlike the present warm and humid Holocene climate (Igarashi 1990). LGM Flora of the Paleo-SHK Peninsula  LGM vegetation of the Paleo-SHK Peninsula has been reconstructed based on pollen records from seven long boring cores such as from Kenbuchi basin (Igarashi 1996; Igarashi et al. 1993), Nakafurano (Igarashi et al. 1993), Ishikari Bay (Ishii et al. 1981; Ozawa et al. 1995), Utasai (Sakaguchi 1989), Shiribeshitoshibetsugawa (Suzuki et al. 1999), and Koe (Igarashi et al. 2002), and nine cores which contain only LGM samples such as from Erimo Cape (Hoshino and Matsuzawa 1987). Igarashi (2008) presented a reconstruction map of LGM vegetation in Hokkaido based on these 16 datasets (Figure 3.2). The result suggests that vegetation across

48 the Paleo-SHK Peninsula comprised plant species that were more tolerant of cold climates than the present flora. LGM vegetation of the Paleo-SHK Peninsula is divided into four types by composition of tree taxa, from north to south: (1) open larch-pine taiga, (2) open larch-pine-spruce taiga, (3) open larch-pine-spruce-fir taiga, and (4) open spruce-larch-fir-oak taiga (Igarashi 2008). The prevalence of non-arboreal taxa in each vegetation zone suggests that these tree groups coexisted with patchy grasslands and/or high moors. The composition of non-arboreal taxa resembles Siberia, implying the “mammoth steppe” expanded into the southern end of the Paleo-SHK Peninsula (Igarashi 2008). In sum, vegetation zones (1), (2), and (3) in northern and central Paleo-SHK Peninsula were characterized by arctic, open forest and grassland, while vegetation zone (4) in the southern part of Paleo-SHK Peninsula included sub-arctic species of open forest and grassland. LGM Fauna of the Paleo-SHK Peninsula  The LGM faunal community of the Paleo-SHK Peninsula consisted of mammoth (Mammuthus primigenius), brown bear (Ursus arctos), bison (Bison priscus), reindeer (Rangifer tarandus), Siberian musk deer (Moschus moschiferus), horse (Equus caballus), moose (Alces alces), snow sheep (Ovis nivicola), panther (Panthera sp.), wolf (Canis lupus), and arctic fox (Alopex lagopus), collectively called the Mammoth Fauna Group (Kirillova 2003; Vasilevski 2008; Kuzmin et al. 2005; Iwase et al. 2011; Takahashi et al. 2006). At least after 50 ka, the Mammoth Fauna migrated to the Paleo-SHK Peninsula from Siberia (Takahashi and Izuho 2012; Iwase et al. 2011). In contrast, the Paleoloxodon-Sinomegaceros Faunal Complex was distributed in cool-temperate to temperate zones of Paleo-Honshu Island, having migrated from the west accompanied by Naumann’s elephant (Palaeoloxodon naumanni), tiger (Panthera tigris), wolf (Canis lupus), Asian black bear (Ursus thibetanus), giant deer (Sinomegaceros yabei), wild boar (Sus scrofa), and Shika deer (Cervus Nippon). In Hokkaido, the

Figure 3.3  Location of major primary sources of obsidian and geological distribution of “hard-shale” in the southern part of PaleoSHK Peninsula (modified from Morisaki et al. 2013).

Izuho boundary between the two faunal groups during MIS3 and MIS2 varied with the expansion and contraction of various vegetation zones, and some loss and replacement of species within and between the groups probably occurred (Takahashi 2007; Takahashi and Izuho 2012; Takahashi et al. 2013). Lithic Raw Materials  The unique tectonic setting of the arctrench systems in and around the Japanese Islands allowed human groups to procure various high-quality raw material from metamorphic, volcanic, and sedimentary sources (Izuho and Hirose 2010; Izuho et al. 2012; Morisaki et al. 2013; Nakazawa et al. 2005). Two major types of cryptocrystalline material, obsidian and “hard-shale,” exist in southern Paleo-SHK Peninsula shown in Figure 3.3. To date, at least 21 obsidian sources, including Shirataki, Oketo, Tokachi-Mitsumata and Akaigawa, have been discovered on Hokkaido (Izuho and Hirose 2010; Izuho and Sato 2008; Ferguson et al. 2013). High-quality “hard-shale” is widely distributed in the southwestern parts of the Paleo-SHK Peninsula. The quality of the material is usually poor, but adequate material for chipped-stone tools is scattered across some limited areas (Izuho 1998). In sum, foragers could procure various high-quality materials anywhere in the southwestern part of the Paleo-SHK Peninsula within a 150-km radius.

Research History and Nature of Upper Paleolithic Records on Hokkaido

The goal of Upper Paleolithic research in Japan, since the first substantial excavation at the Iwajuku site in central Honshu, has been to establish chronologies and to characterize Japanese Paleolithic cultures based on comparisons of manufacturing techniques and index-type artifacts from Upper Paleolithic sites. As the picture of the Japanese Upper Paleolithic was being established (Anbiru 1978; Serizawa 1960; Sugihara 1965; Tozawa 1990; see Nakazawa 2010 for more detail), several researchers were influenced by processual ar-

Human Technological and Behavioral Adaptation to Landscape Changes around the Last Glacial Maximum in Japan chaeologists from the United States and conducted research to explain the Paleolithic record from an evolutionary ecology view with emphasis on the process of Upper Paleolithic cultural evolution (Anzai 1990; Sato 1992; Tamura 1992). To date, however, traditional research is still occurring, and Japanese Upper Paleolithic researchers are coping with new issues, including advances in lithic analysis, to reconstruct hunter-gatherer mobility patterns and resource-acquisition strategies (Kunitake 2005; Morisaki 2011; Yamada 2006), to elucidate stone-tool function by micro use-wear analysis (Iwase 2012; Kanomata 2005; Iwase and Morisaki 2008; Tsutsumi 1997), and to characterize hunting strategies by synthesizing data from vertebrate paleontology, the geographic distribution of fossil records, radiocarbon dating (Iwase et al. 2012; Takahashi et al. 2006), and Upper Paleolithic human ecology (Izuho and Sato 2008; Izuho et al. 2012; Morisaki et al. 2013; Sato et al. 2011). Upper Paleolithic research in Hokkaido began in the 1950s (Yoshizaki 1956). To date, more than 861 Paleolithic sites have been found in Hokkaido, and over 360 of these have been excavated (Database committee of Japanese Paleolithic Research Association 2010). Many Upper Paleolithic sites in Hokkaido are associated with eolian or colluvial contexts on alluvial terraces, gentle slopes, and paleo-dunes (Izuho and Akai 2005; Izuho et al. 2013). These can be divided into two groups: those from the Ishikari Lowland and the Tokachi Plain that have established, relatively detailed geochronological contexts developed from late Quaternary tephra sequences; and those from the Tokoro River and Yubetsu River basins, where site stratigraphy and soils are shallow and poorly developed (less than 30 cm deep), and are subject to severe post-depositional disturbance, mainly by periglacial processes (Izuho and Oda 2008). As in mainland Japan, until the 1990s much effort was spent to establish cultural chronologies and traits of Upper Paleolithic sites. For example, Yoshizaki (1958) and Oba and Matsushita (1965) changed the Upper Paleolithic cultural chronology again and again as a consequence of misinterpreting stone-tool manufacturing techniques and index-type artifacts (Tsurumaru 2001; Izuho and Akai 2005). On the other hand, using Binford’s (1966) scheme, Kato (1970) attempted to explain diversity in Upper Paleolithic assemblages and differences in site function. Since the 1970s, Ueno and Kato (1973), Tsurumaru (1979), and Yonemura (1983) have undertaken detailed reconstruction of microblade-core manufacturing technology. Coinciding with increased excavation of Upper Paleolithic sites due to economic growth since the 1980s (Tsurumaru 2001), recent studies have focused on reconstructing entire reduction sequences from excavated sites (Yamada 1986; Shiraishi 1993; Kimura 1992). In the new millennium, studies using geoarchaeology and evolutionary ecology have significantly increased in Hokkaido (Nakazawa 2011; Sato 2003; Yamada 2006; Izuho and Akai 2005; Izuho et al. 2009). As Sato (2003), Izuho and Sato (2008), and Morisaki et al. (2010) suggested, Hokkaido’s Upper Paleolithic assemblages can be classified into trapezoid, backed point, flake,

49

or ­ microblade industries based on techno-morphological features and tool-type composition. These industries are geochronologically divided into three stages: (1) small-flake industries dated to older than 27,000–24,000 14C yr BP, and perhaps as old as 30,000 14C yr BP; (2) early microblade, blade, and flake industries, all dated between 22,000 and 18,000 14C yr BP; and (3) various microblade assemblages dated between 18,000 and 12,000 14C yr BP (Izuho and Akai, 2005; Izuho and Takahashi, 2005; Izuho et al. 2012; Nakazawa et al., 2005). It is also suggested that these stages are associated with the following environments: 1) cool-temperate deciduous broadleaf forests, where remains of Naumann’s elephant were found, 2) patchy sub-arctic coniferous forests and steppe into which woolly mammoths migrated from the Russian Far East via Sakhalin, as well as 3) increased forest and decreased grass steppe, respectively (Izuho and Takahashi 2005; Izuho et al. 2012). These reconstructed shifts of technology and landscape imply that the emergence of various lithic industries in Hokkaido resulted from hunter-gatherer behavioral adaptations to changing environments and subsequent food resources (Sato and Izuho 2011; Sato et al. 2011). Recently researchers have started to question how LGM flake, blade, and microblade assemblages can be explained in the context of late-Pleistocene environments (Nakazawa 2011; Yamada 2006; Terry et al. 2012). Moreover, current issues in Beringian prehistory discussed in Goebel and Buvit (2011) are similar to topics being researched in Hokkaido.

Material and Methods Methods The materials used in this paper are gleaned from previous research by Izuho and Akai (2005) and Izuho et al. (2012). A total of five sites (comprising 33 lithic concentrations) are covered in this paper: Kashiwadai 1 (LC 1 of Block excavation A, and LC2, 3, 6, 11, 12, 14, and 15 of Block Excavation B) (Hokkaido Center for Buried Cultural Property 1999), Shimaki (LC1 of the 2010-2011 campaign) (Terry et al. 2012), Minamimachi 2 (LC1 from the lower layer) (Obihiro Board of Education 1995), Marukoyama (LC1 from the lower layer) (Chitose Board of Education 1994), and Kawanishi-C (LC1 through LC12 of Layer VI) (Obihiro Board of Education 1998, 2000a, b). General descriptions of each site and their cultural assemblages are provided in the next section. The date of human occupation, site location, toolkit variability, and lithic reduction sequences are discussed. The techno-typology of these five geochronologically reliable sites resemble other sites in the area—Ogachi-Kato 2, Kyushirataki 16, Oketoazumi, Kamishirataki 7, Kamishirataki 8 sites, mainly composed of flake-reduction sequences, Ogachi-Kato 2, Pirika 1, Miyako, Yunosato 4, Kamishirataki 2, Midori 1, and Shinmichi 4 sites mainly composed of blade/ microblade-reduction sequences, as well as Kyushirataki 15

50 composed of a blade-reduction sequence. These latter sites are rejected from this study because they are not geochronologically reliable. However, they imply that these three reduction sequences were a consistent strategy of LGM huntergatherers on Hokkaido. Evaluation of Radiocarbon Dates  Because many of the Upper Paleolithic sites on Hokkaido are not in primary contexts, geochronological assessment using 14C AMS dates is essential (Izuho and Akai 2005; Izuho et al. 2012). For each radio­carbon date, the following factors are taken into consideration: 1) dating method (AMS or conventional), pretreatment procedure, and material analyzed, 2) geological context of the sample, and 3) archaeological context of the sample within the site. Calibration was done with OxCal (v.4.1.5) and all AMS dates are expressed as 2σ. Site Location  The landscape context of the archaeological record of Paleolithic Hokkaido is assessed through evaluation of site location and lithic raw-material environments. Unfortunately, prehistoric animal resources and vegetation in and around the sites at the time of occupation could not be discerned because organic remains are not preserved at any of the sites discussed here. This is common for all of Hokkaido (Izuho et al. 2012; Nakazawa et al. 2005). Toolkit Diversity  Hunter-gatherer toolkit diversity is affected by the constraints of foraging behavior (Binford 1980; Shott 1986; Torrence 1983), and this is no less the case in Hokkaido (Terry et al. 2012). I use toolkit diversity, measured by the number of tool types present (Richness), and the degree to which the tool types are equally represented (Evenness: Simpson’s Index 1/D), as proxy for the range of activities carried out at each site. Lithic-Reduction Sequences  The goal of reduction-sequence analysis is to reconstruct Upper Paleolithic huntergatherer lithic industries and raw-material use on Paleo-SHK. Each reconstruction is based on the removal order of detached pieces through refit studies, a common approach in studies of Upper Paleolithic Hokkaido (Akai 2005a, b, 2008; Izuho 1998; Naoe 2009; Oda 2009; Suzuki 2007; Takakura 2000; Yamada 1999; Naoe and Nagasaki 2005; Izuho et al. 2013MS). Each refit sequence is separated into primary and secondary reduction stages. For this study, I relied on excavation reports, which contained information on lithic toolkit assemblage structure and refit analysis. To corroborate the published data, I examined each assemblage. Kashiwadai 1 (LC 1, 2, 3, 6, 11, 12, 14, and 15) The Kashiwadai 1 site is located in the central part of the Ishikari lowland, central Hokkaido (42° 48′ N, 141° 41′ E) (Koaze et al. 2003) on a paleo-dune formed from reworked Spfa-1 tephra along the Chitose River at an altitude of 13 m

Izuho asl. The artifact-bearing layer was found in the lower part of a thick eolian loam 4 m below the surface between reworked Spfa-1 (50,000–40,000 cal yr BP [45,000–35,000 14C yr BP]) and En-a (21,000–19,000 cal yr BP [17,000 14C yr BP]) tephras (Hokkaido Center for Buried Cultural Property 1999; Izuho and Akai 2005). Paleolithic components were found in two block excavations, labeled A and B. A lithic concentration associated with a hearth feature was found in block excavation A. Several lithic concentrations with 11 associated hearths and 4 charcoal concentrations were found in block excavation B. A total of 31 14C AMS dates are reported from the site and shown in Figure 3.4A and Table 3.2. These dates can be divided into two intervals; 26 younger dates at 27,100– 22,100 cal yr BP (22,550–18,800 14C yr BP) and 5 older dates at 43,000–31,400 cal yr BP (38,500–27,200 14C yr BP), though all of the dates are consistent with the geological context of the site. The older dates, including three charcoal samples recovered from “hearths,” are rejected because these are thought to have resulted from natural fire during or just after the deposition of the Spfl pyroclastic flow. Thus, the ages of archaeological components at Kashiwadai 1 fall within the interval of 27,900–22,100 cal yr BP (22,550–18,800 14C yr BP) (Izuho and Akai 2005; Izuho et al. 2012). A total of 159 lithic specimens were recovered from a concentration in block excavation A. Raw material consists of non-local obsidian (n=2, 0.9g total) and “hard-shale” (n = 157, 267.8g total) (Hokkaido Center for Buried Cultural Property 1999). The lithic assemblage consists of blades, microblades, platform-forming spalls, and flakes. Primary reduction produced blade and microblades from wedge-shaped cores. Endscraper edge retouch was a common secondary reduction strategy. From block excavation B, a total of 32,822 lithic specimens were recovered from 14 concentrations. Several lithic concentrations produced blade/microblade and flake assemblages. Of these concentrations, LC11 is classified as a tightly clustered flake assemblage consisting of 9409 specimens. Lithic raw material in LC11 consists of obsidian (n = 6276, 2416.9g), “hard-shale” (n = 935, 2435.9g), andesite (n = 432, 5429.3g), agate (n = 298, 1017.9g), chert (n = 298, 1303.27g), and other material (n = 1,170, 2682.1g). Obsidian and “hardshale” were considered non-local, while the sources of agate, chert, and andesite are unknown (Hokkaido Center for Buried Cultural Property 1999). Primary reduction used bipolar percussion to produce triangular and trapezoid flakes from discoidal and conical cores. Endscraper edge and margin retouch are common secondary reduction strategies. Tools made on flakes include endscrapers, sidescrapers, wedgeshaped tools, retouched flakes, and utilized flakes with micro-flaking, as well as flake cores, mobile art, ocher, hammerstones, pebble tools, and pebbles (Figure 3.5B). A total of 3400 lithic specimens, consisting of blade/microblade artifacts, were recovered from the six other lithic concentrations in block excavation B. The lithic assemblage consists of microblades, blades, burins, endscrapers, Ranko-

Human Technological and Behavioral Adaptation to Landscape Changes around the Last Glacial Maximum in Japan

51

k calbp 45 a 40

b c d

35

e

30 f g h

I

j k

l m n o p q r s

25

u v w x

t

y z aa ab ac ad ae

20 IV Geological layer

Sp f

k calbp 35

En -

a-

a

1

A

k calbp 35

k calbp 35

k calbp 35 al

30

30

30

30

af

ak am an

ag 25

25 unknown

20

20

25

25

20

20

15

15

10

10

ao

ap

ar

ah ai

Geological layer

B

Geological layer

C

Geological layer

D

IVc

IVb

Geological layer

d

at

Ta -

VI

as

-a

-1 fa Sp

-a

fa

-1

VII

d

III

Ta -

-a En

-1 Sp

fa

1 Si

pf

a-

2 apf

VI

aq

En

10 V

Sp

10

Si

aj

15

En

15

A, Kashiwadai 1; B, Shimaki; C, Minamimachi 2; D, Marukoyama; E, Kawanishi-C. The laboratory numbers for radiocarbon dates follow. a Beta-126179 x Beta-120880 b Beta-126181 y Beta-112916 c Beta-126180 z Beta-120883 d Beta-126182 aa Beta-112919 e Beta-112918 ab Beta-126170 f Beta-126171 ac Beta-120882 g Beta-126173 ad Beta-112915 h Beta-126168 ae Beta-126177 i Beta-126169 af Gak-3262 j Beta-112913 ag NUTA2-7680 k Beta-126183 ah Gak-18248 l Beta-126174 ai TKa-15508 m Beta-120881 aj Gak-18247 n Beta-126178 ak NUTA-2801 o Beta-126175 al TKa-15536 p Beta-126176 am Beta-107731 q Beta-112922 an NUTA2-7677 r Beta-126184 ao TKa-15537 s Beta-112914 ap Beta-106506 t Beta-126167 aq Beta-127399 u Beta-112920 ar Beta-126151 v Beta-112921 as Beta-126150 w Beta-156337 at TKa-15535

E

Figure 3.4  Radiocarbon dates from the sites discussed in this chapter. Black dot = calendar age, white circle=conventional age.

shi-type microblade cores, flake cores, and flakes. Other material includes ocher, amber beads, hammerstones, and anvils (Figure 3.7). Lithic raw material consists mainly of “hard-shale” followed by small amounts of obsidian, andesite, agate, and chert. Obsidian and “hard-shale” were considered non-local (Hokkaido Center for Buried Cultural Property 1999). Primary reduction produced flakes, blades and microblades. Continuous retouch, endscraper edging, and burin manufacture are common secondary reduction strategies.

Shimaki Site (LC1 of the 2010–2011 Campaign)

The Shimaki site is located at the topographic division known as the Northern Terraces on the northeastern Tokachi Plain, eastern Hokkaido (43° 14′ N, 143° 18′ E) in an area with prominent, relatively flat to wavy tread surfaces (Koaze et al. 2003). The site is situated at the edge of the Kitaoribe II Terrace along the Otofuke River, a tributary of the Tokachi River at an altitude of 290 m asl. The artifact-bearing layer is in an eolian loam, 0.5–1.0 m below the surface, and positioned between primary fall of the Shikaribetsu Dai 2 (Sipfa-2,  20,000 BP) tephras (Izuho and Akai 2005; Kato and Yamada 1988).

Two conventional 14C dates are reported: 23,800 14C yr BP from the artifact-bearing layer; and 25,500 ± 1200 14C yr BP) from 40 cm below the cultural level, which provides a lower age limit (Kato and Yamada 1998; Kosaka and Nogawa 1972; Buvit et al. 2011) (Table 3.2 and Figure 3.4B). In addition to the radiocarbon ages, a fission track date (21,700 ± 1800 cal yr BP [BS-A-51]) and five obsidian hydration dates (19,000 ± 800 cal yr BP [BS.A1-26]; 18,200 ± 500 cal yr BP [A1-137]; 18,200 ± 700 cal yr BP [B1-43]; 17,200 ± 800 cal yr BP [B1-2896]; 17,500 ± 900 cal yr BP [B3-80]) were reported (Kato and Yamada 1998). Although the exact provenience of the charcoal is not designated in the excavation report, these dates are all consistent with the geological context of the site. A total of around 10,000 lithic specimens were recovered from 8 lithic concentrations from 1967–1988. Raw material consists of obsidian, andesite, “hard-shale,” agate, and other coarse igneous pebbles. Obsidian gravels, usually 10 cm in diameter, were the dominant raw material acquired from the bed of the Otofuke River near the site (Kato and Yamada 1988). Primary reduction produced flakes, which were used to produce triangular and trapezoidal flakes from discoidal

52

Izuho

A

B

C

0

10

cm

D

Figure 3.5  LGM flake assemblages: A, Shimaki; B, Kashiwadai 1; C, Marukoyama; D, Minamimachi 2.

and conical cores, as well as bladelets from wedge-shaped bladelet cores. Endscraper edge retouch, marginal retouch, beaked retouch, as well as flat flaking on ventral surfaces are clearly evident secondary reduction strategies. The tool assemblage consists of endscrapers, sidescrapers, perforators, wedge-shaped tools, retouched flakes, and utilized flakes with micro-flaking, as well as flake cores, ocher, hammerstones, pebble tools, and pebbles (Figure 3.5A). During the 2010–2011 excavations, a total of 637 lithic specimens were recovered from a single lithic concentration and subsequently analyzed. Raw material consists of obsidian, andesite, “hard-shale,” and other coarse igneous pebbles. The majority of the raw material was rounded ob-

sidian gravels exhibiting incipient cone cortex gathered from the Otofuke River roughly 2.3 km from the site (Terry et al. 2012). Primary reduction produced triangular and trapezoidal flakes from discoidal and conical cores, as well as bladelets from wedge-shaped bladelet cores. Endscraper edge retouch, marginal retouch, beaked retouch, as well as flat flaking on ventral surfaces were common secondary reduction strategies. The tool assemblage consists of endscrapers, sidescrapers, perforators, wedge-shaped tools, retouched flakes, retouched blades, utilized flakes with micro-flaking, as well as flake cores, wedge-shaped bladelet cores, ocher, hammerstones, pebble tools, and pebbles (Terry et al. 2012). The majority of on-site reduction was involved in producing

Human Technological and Behavioral Adaptation to Landscape Changes around the Last Glacial Maximum in Japan flakes, although several blade tools were transported to the site from another location.

Minamimachi 2 (LC1 of Lower Layer)

The Minamimachi 2 site is located on the Southern Terraces in the southeastern Tokachi Plain, eastern Hokkaido (42° 52′ N, 143° 10′ E) (Koaze et al. 2003) at the edge of the Motomatsu terrace of the Satsunai River at an altitude of 79 m asl. Three archaeological layers were found in eolian units classified from top to bottom as a Jomon component in the Black Humus soil above the Ta-d tephra (layer I), a microblade assemblage in an eolian loam between the Ta-d and En-a tephras (layer III), as well as a flake assemblage in an eolian loam between the En-a and Spfa-1 tephras (layer VI) (Obihiro Board of Education 1995; Izuho and Akai 2005; Izuho et al. 2013). Two conventional and two AMS 14C dates are reported (Obihiro Board of Education 1995; Nakamura 2005; Izuho et al. 2013). Charcoal samples were collected from different hearths in layers III and VI and dated by conventional methods and AMS, respectively. Although these four dates are consistent with the geological context of the site, there seems to be a significant difference between the conventional and AMS dates. Two conventional dates (19,610 ± 270 and 13,790 ± 190 14C yr BP) are rejected, and thus the layer VI flake assemblage is within the interval of 26,220–25,500 cal yr BP (21,610 ± 70 14 C yr BP) (Izuho et al. 2013) (Table 3.2 and Figure 3.4C). A total of 2228 lithic specimens were recovered from a flake concentration in geological layer VI. Raw material mainly consists of obsidian (54.9%), followed by andesite (34.7%), and agate (3.3%). The lithic assemblage consists of sidescrapers, retouched flakes, as well as flakes, flake cores, and pebbles (Figure 3.5D). Primary reduction produced triangular and trapezoidal flakes from discoidal cores. Endscraper edge and margin retouch are common secondary reduction strategies.

Marukoyama Site (LC1 of Lower Layer)

The Marukoyama site is located in the central part of the Ishikari Lowland, central Hokkaido (42° 51’ N, 141°42’ E) (Koaze et al. 2003) on a paleo-dune formed from reworked Spfa-1 material along the Chitose River at an altitude of 22 m asl. Three archaeological layers were found in eolian units classified from top to bottom as a Jomon component in the Black Humus soil above the Ta-d tephra, a microblade assemblage in an eolian loam between the Ta-d and En-a tephras, as well as a blade assemblage in an eolian loam between the En-a and Spfa-1 tephras (Chitose Board of Education 1994). A single 14C AMS date, 27,540-25,550 cal yr BP (21,940 ± 250 14 C yr BP), is reported (Table 3.2, Figure 3.4D). The sample was collected from a charcoal concentration, which overlaps with lithic concentration. The dates are consistent with the geological context of the site. In the upper part of the Paleolithic component, approximately 200 lithic specimens were recovered from five concentrations (Chitose Board of Education 1994; Nakazawa 2001). Raw material mainly consists of obsidian (63%), followed by “hard-shale” (27%), and minor amounts of andesite and chert.

53

The lithic assemblage consists of sidescrapers, endscrapers, burins, as well as flakes, flake cores, ocher, and hammerstones (Figure 3.5C). Primary reduction produced triangular and trapezoidal flakes from discoidal and conical cores. Endscraper edge and margin retouch, and burin technology were common secondary reduction strategies.

Kawanishi-C (LC1-12 of Layer VI)

The Kawanishi-C site is located on the southern terraces in the southeastern Tokachi Plain, eastern Hokkaido (42° 52′ N, 143° 11′ E) (Koaze et al. 2003) at the edge of the Kamisatsunai I terrace of the Satsunai River at an altitude of 70 m asl. The Obihiro Board of Education carried out three salvage excavations, uncovering a total area of 6856 m2 (Obihiro Board of Education 1998, 2000a, 2000b). Three archaeological layers were found in eolian units classified from top to bottom as a Jomon component in the Black Humus soil above the Ta-d tephra, a microblade assemblage in an eolian loam between the Ta-d and En-a tephras (layer IVb and IVc), as well as a blade assemblage in an eolian loam between the En-a and Spfa-1 tephras 0.7 m below the surface (layer VI) (Izuho and Akai 2005; Obihiro Board of Education 1998). A total of nine 14C AMS dates from Kawanishi-C are shown in Figure 3.4E and Table 3.2. Of the samples, one was collected from 0.2 m below the cultural level (layer VII), four were collected from layer VI, one from layer IVc, and three from layer IVb. A total of four charcoal dates, 26,680–25,820 cal yr BP (21,780 ± 90 14C yr BP), 26,650–25,690 cal yr BP (21,710 ± 70 14C yr BP), 26,100–25,150 cal yr BP (21,480 ± 120 14 C yr BP) and 26,150–25,640 cal yr BP (21,420 ± 190 14C yr BP) were recovered from hearths associated with lithic concentrations in layer VI. Likewise, samples collected from charcoal concentrations in layer IV dated to 20,350–19,850 cal yr BP (16,920 ± 50 14C yr BP) and 16,060–15,020 cal yr BP (12,900 ± 50 14C yr BP). Although the dates from layers IVb and IVc are somewhat inconsistent, probably owing to post-depositional disturbance that is observable in a profile at the site, nonetheless all the dates are consistent with key tephra layers. Thus, the age of archaeological component in Kawanishi-C falls within the interval of 27,000–25,000 cal yr BP (22,000–21,000 14C yr BP) (Izuho et al. 2013). In the upper part of the Paleolithic component, 2000 lithic specimens were recovered from five concentrations. In contrast, 19,000 lithic artifacts were recovered from 12 concentrations in the lower part of the Paleolithic component. Raw material mainly consists of obsidian, followed by a small amount of “hard-shale,” agate, andesite, and coarse-grained igneous cobbles and pebbles. The lithic assemblage is primarily blade-based and consists of sidescrapers, endscrapers, burins, perforators, and wedge-shaped tools, as well as flakes, flake cores, ocher, pebble tools, and hammerstones (Figure 3.6). Primary reduction produced blades. Endscraper edge forming and burin technology were common secondary reduction strategies (Obihiro Board of Education 1998, 2000a, 2000b).

54

Figure 3.6  LGM blade assemblage from Kawanishi-C.

Izuho

0

10

Results LGM Geochronology of Hokkaido Resulting 14C AMS dates of archaeological component from LGM sites on Hokkaido are shown in Figure 3.8. Kashiwadai 1 dates fall within 27,900 and 22,100 cal yr BP (22,550 and 18,800 14C yr BP); Minamimachi 2—26,220–25,500 cal yr BP (21,610 ± 70 14C yr BP); Marukoyama—27,540–25,550 cal yr BP (21,940 ± 250 14C yr BP); and Kawanishi-C—27,000–25,000 cal yr BP (22,000–21,000 14C yr BP). The dates from Shimaki are excluded because of inconsistencies in the radio­carbon chronology at the site. All remaining dates in Figure 3.8 fall between 30,000 and 21,000 cal yr BP, the range of the LGM from Sea of Japan data. Among them, the ages from Kashiwadai 1 span 5800 years, extending beyond the Siberian LGM. Dates from Paleolithic layers at Minamimachi 2, Marukoyama, and Kawanishi-C, on the other hand, overlap at 26,000 cal yr BP, and do not extend past the Siberian LGM. Site Location and Landscape Context Locations for each study site are shown in Table 3.1. All sites are situated on large, expansive plains—Kashiwadai 1 and Marukoyama (10–20 m asl) in the southern part of the Ishikari Lowland, Minamimachi 2 and Kawanishi-C (70–80 m asl) in the southern part of the Tokachi Plain, and Shimaki (290 m

cm

asl.) in the northern part of the Tokachi Plain. Kashiwadai 1 and Marukoyama are associated with paleo-dunes 5 m above the nearest active alluvial channel, while Shimaki, Minamimachi 2, and Kawanishi-C are associated with a middle alluvial terrace about 30 m above active channels. Although the specific vegetation at and around the sites during occupation could not be reconstructed, all the sites were generally situated in what was sub-arctic grassland and open spruce-larch-fir-oak taiga. Several members of the Mammoth Fauna Group would have been available as prey. Local lithic raw material varies greatly between the Ishikari lowland and the Tokachi Plain. In the Ishikari lowland, local high-quality raw material is currently unknown. It is likely that low-quality chert is found in the vicinity of the sites, although its precise location is not known. The nearest highquality obsidian source is Akaigawa 80 km away in a straight line; a “hard-shale” source lies in southwestern Hokkaido 100 km away. At Shimaki, on the other hand, high-quality large pebbles and cobble-sized obsidian gravel, from the Tokachi-Mitsumata source (35 km from the site), is locally available at the Otofuke River. Shimaki is also 50 km for Oketo source. Minamimachi 2 and Kawanishi-C are located at the confluence of the Otofuke and Tokachi rivers, where high-quality obsidian gravel derived from Tokachi-Mitsumata and Tokachi-Shikari-

Human Technological and Behavioral Adaptation to Landscape Changes around the Last Glacial Maximum in Japan

0

10

55

cm

Figure 3.7   LGM blade/microblade assemblage from Kashiwadai 1.

betsu is also locally available. The primary obsidian sources nearest these sites are Tokachi-Mitsumata (70 km), and Oketo (85 km). Toolkit Diversity Toolkit diversity for each study assemblage is shown in Table 3.3. Kashiwadai 1 is divided into two groups, flake assem-

blages at LC11 and blade/microblade assemblages from LC1, 2, 3, 6, 12, 14, and 15. The quantity of chipped-stone tools varies greatly between the sites—Kashiwadai 1 flake assemblage (n = 165), Kashiwadai 1 blade/microblade assemblage (n = 657), Shimaki (n = 63), Marukoyama (n = 34), Minamimachi 2 (n = 10), and Kawanishi-C (n = 252). The Kashiwadai 1 flake assemblage,

Table 3.1  Sites mentioned in the text. Site number corresponds with number in Figure 1. No. (Fig.1)

Type of assemblage Site

Geographical coordinates River basin Location

Altitude (m.a.s.l.)

Height above riverbed (m)

No. of lithic concentrations No. of lithics

1 Flake Kashiwadai 1 42˚ 48′ 58″ N, Chitose R. Paleo-dune 13 5 1 LC11 141˚41′ 6″ E (Spfa-1) 1 Blade/ Kashiwadai 1 42˚ 48′ 58″ N, Chitose R. Paleo-dune 13 5 7 microblade LC1,2,3,6,12, 141˚ 41′ 6″ E (Spfa-1) 14,15 2 Flake Shimaki 43˚ 14′ 2″ N, Otofuke R. River terrace 290 30–35 8 3 Flake Minamimachi 2 42˚ 52′ 45″ N, Satsunai R. River terrace 80 30–35 1 143˚ 10′ 22″ E River Terrace 4 Flake Marukoyama 42˚ 51′ 46″ N, Chitose R. Paleo-dune 22 10 4 141˚ 42′ 35″ E (Spfa-1) 5 Blade Kawanishi-C 47˚ 52′ 54″ N, Satsunai R. Kamisarabetsu I 70 25–30 12 143˚ 11′ 23″ E terrace Note: sedimentary environment of all sites is aeolian.

9409 3420

ca. 10,000 2228 ca. 200 19,326

56

Izuho

Table 3.2  List of radiocarbon dates from assemblages related to LGM Hokkaido. Site

Geological Archaeological Lab. Radiocarbon context context number Method δ13C age

Kashiwadai 1 between En-a Area M-66, and Spfa-1 KD1-32 between En-a Area I-66, and Spfa-1 KD1-34 between En-a M-68, KD1-33 and Spfa-1 between En-a Area P-6, Sb-1, and Spfa-1 hearth, KD1-35 between En-a Area O-7, and Spfa-1 KD1-6 between En-a Area F-59, Sb-9, and Spfa-1 hearth, KD1-24 between En-a Area H-58, Sb-4a, and Spfa-1 hearth, KD1-21 between En-a Area I-63, Sb-10, and Spfa-1 hearth, KD1-26 between En-a Area D-58, Sb-7, and Spfa-1 hearth, KD1-22 between En-a Area D-57, KD1-1 and Spfa-1 between En-a Area F-61, Sb-11, and Spfa-1 hearth, KD1-36 between En-a Area F-61, Sb-11, and Spfa-1 hearth, KD1-27 between En-a Area K-65, Sb-13 and Spfa-1 hearth, KD1-14 between En-a Area P-69, charcoal and Spfa-1 concentration, KD1-31 between En-a Area F-64, Sb-12, and Spfa-1 hearth, KD1-28 between En-a Area P-55, Sb-2, and Spfa-1 KD1-10 between En-a Area H-63, Sb-15, and Spfa-1 hearth, KD1-29 between En-a Area F-64, Sb-12, and Spfa-1 hearth, KD1-37 between En-a Area F-59, Sb-9, and Spfa-1 KD1-2 between En-a Area L-58, Sb-3, and Spfa-1 KD1-20 between En-a Area P-55, Sb-2, and Spfa-1 KD1-8 between En-a Spot-I, 24-12c, and Spfa-1 Sb-1, No. 203 between En-a Area P-55, Sb-2, and Spfa-1 KD1-9 between En-a Area K-65, Sb-13, and Spfa-1 hearth, KD1-13 between En-a Area N-63, Sb-14, and Spfa-1 hearth, KD1-16 between En-a Area H-63, Sb-10, and Spfa-1 KD1-4 between En-a Area P-6, Sb-1, and Spfa-1 KD1-7 between En-a Area E-56, Sb-6, and Spfa-1 hearth, KD1-23 between En-a Area N-63, Sb-14, and Spfa-1 hearth, KD1-15

cal yr BP (2σ) Evaluation2 Reference

Beta- AMS -24.9 37,350 ± 550 126179 Beta- AMS -24.6 33,030 ± 540 126181 Beta- AMS -25.8 32,490 ± 360 126180 Beta- AMS -25.2 31,350 ± 330 126182 Beta- AMS -25.0 28,200 ± 480 112918 Beta- AMS -24.7 22,550 ± 180 126171 Beta- AMS -24.7 22,340 ± 170 126168 Beta- AMS -24.5 22,340 ± 200 126173 Beta- AMS -24.7 22,300 ± 180 126169 Beta- AMS -23.9 22,210 ± 210 112913 Beta- AMS -23.6 22,200 ± 170 126183 Beta- AMS -24.7 21,790 ± 230 126174 Beta- AMS -25.2 21,000 ± 100 120881 Beta- AMS -25.9 20,900 ± 190 126178

42,950– 4 Hokkaido 41,350 Center for Buried Cultural 38,930– 4 Property (1999) 36,570 38,400– 4 36,440 36,540– 3 35,090 33,960– 4 31,400 27,910– 1 26,650 27,690– 1 26,240 27,720– 1 26,220 27,670– 1 26,210 27,660– 4 26,110 27,590– 1 26,150 26,910– 1 25,340 25,500– 1 24,610 25,540– 1 24,420

Beta- AMS -24.9 20,790 ± 160 126175 Beta- AMS -26.6 20,680 ± 210 112922 Beta- AMS -25.3 20,700 ± 150 126176 Beta- AMS -24.2 20,610 ± 160 126184 Beta- AMS -26.8 20,570 ± 120 112914 Beta- AMS -24.5 20,580 ± 160 126167 Beta- AMS -24.4 20,510 ± 160 112920 Beta- AMS -24.6 20,500 ± 200 156337 Beta- AMS -25.6 20,490 ± 130 112921 Beta- AMS -25.8 20,390 ± 070 120880 Beta- AMS -25.3 20,370 ± 070 120883 Beta- AMS -23.7 20,320 ± 150 112916 Beta- AMS -26.1 20,180 ± 120 112919 Beta- AMS -25.9 20,130 ± 150 126170 Beta- AMS -25.6 19,840 ± 70  120882

25,220– 24,340 25,230– 24,080 25,080– 24,310 25,050– 24,180 24,980– 24,220 25,030– 24,130 24,970– 24,000 24,990– 23,930 24,930– 24,010 24,680– 23,910 24,530– 23,930 24,760– 23,820 24,460– 23,770 24,470– 23,610 23,970– 23,360

1 1 1 1 1 1 1 1 1 1 1 1 1 1 1

Hokkaido Center for Buried Cultural Property (1999)

Human Technological and Behavioral Adaptation to Landscape Changes around the Last Glacial Maximum in Japan

57

Table 3.2 Cont'd. Site

Geological Archaeological Lab. Radiocarbon context context number Method δ13C age



between En-a Area F-64, KD1-3 and Spfa-1 between En-a Area H-63, Sb-15, and Spfa-1 hearth, KD1-30

Shimaki

between Sipfa-1 unknown unknown Conv. – 23,8001   – 2 and Sipfa-2 between Sipfa-1 Brown loam 40cm GaK- Conv. – 25,500 ± 12001   – 2 and Sipfa-2 below cultural level 3262

Minamimachi 2 between En-a and Spfa-1 between En-a and Spfa-1 between Ta-d and En-a between Ta-d and En-a

9, 8-6, burned soil in Sp-1 charcoal 9, 8-6, burned soil in Sp-1 charcoal 13, 14-7, pit 2 in Sp-3 charcoal 2 13, 14-7, pit 2 in Sp-3 charcoal 2

23,920– 22,980 23,260– 22,100

4 1

Hokkaido Center for Buried Cultural Property (1999) Kato and Yamada (1988) Kosaka and Nogawa (1972)

NUTA2- AMS -24.0 21,610 ± 70  26,230– 1 7680 25,500 Gak- Conv. – 19,610 ± 2701   – 18248 Gak- Conv. – 13,790 ± 1901   – 1 18247 TKa- AMS -25.1 14,450 ± 80  17,900– 1 15508 17,200

Nakamura (2005)

NUTA- AMS – 21,940 ± 250 2801

27,540– 1 25,550

Chitose Board of Education (1994)

17-15, layer VII, TKa- AMS -24.6 27,840 ± 200 scattered charcoal 15536 17-15, burned-soil Beta- AMS -26.3 21,780 ± 90  in Sp-4, geo-5932 107731 17-15, burned soil NUTA2- AMS -23.8 21,710 ± 70  in Sp-4 7677 Blade assemblage, TKa- AMS -25.6 21,480 ± 120 Sp-4, hearth 15537 charcoal 2 16-19, burned soil Beta AMS -26.1 21,420 ± 190 in Sp-3, geo-5805 106506 21-21, Sp-17, Beta- AMS -26.7 16,920 ± 50  (10325) No. 3 126151 21-21, Sp-17, Beta- AMS -28.1 13,020 ± 40  (10324) No. 2 126150 23-22, Sp-16, Beta- AMS -26.2 12,900 ± 50  (10913) No. 1 127399 19-17-d, layer IVb TKa- AMS -26.0 12,290 ± 80  scattered charcoal 15535

32,700– 4 31,450 26,680– 1 25,820 26,650– 1 25,690 26,100– 1 25,150

Izuho et al. (2013)

26,150– 1 25,640 20,350– 4 19,850 16,320– 4 15,160 16,060– 4 15,020 14,890– 4 13,930

Obihiro Board of Education (1998) Obihiro Board of Education (2000) Obihiro Board of Education (1998) Obihiro Board of Education (1998) Izuho et al. (2013)

Marukoyama between En-a Y-14, Hearth-30 and Spfa-1 Kawanishi-C between En-a and Spfa-1 between En-a and Spfa-1 between En-a and Spfa-1 between En-a and Spfa-1 between En-a and Spfa-1 between Ta-d and En-a between Ta-d and En-a between Ta-d and En-a between Ta-d and En-a

Beta- AMS -24.4 19,660 ± 130 112915 Beta- AMS -25.4 18,830 ± 150 126177

cal yr BP (2σ) Evaluation2 Reference

Obihiro Board of Education (1995) Obihiro Board of Education (1995) Izuho et al. (2013)

Obihiro Board of Education (1998) Nakamura (2005) Izuho et al. (2013)

All dates obtained from charcoal samples. 1 Measured radiocarbon age (not corrected for isotopic fractionation, calculated using the δ13C) 2 Evaluation code: 1 Best (AMS age consistent with geological and archaeological context); 2 Okay (AMS age consistent with geological context); 3 Bad (AMS age inconsistent with geological context); 4 Terrible (AMS age not associated with archaeological context or cultural horizon).

the Shimaki assemblage, and the Kawanishi-C assemblage express the highest diversity (richness = 6), followed by the Kashiwadai 1 blade/microblade assemblage (richness = 5), and the Marukoyama assemblage (richness = 4). The lowest diversity is in the Minamimachi 2 assemblage (richness = 2) (Figure 3.9 and Table 3.3). Tool evenness (Simpson’s Diversity Index [1/D]), on the other hand, is relatively higher at Kawanishi-C (evenness = 4.05) and Marukoyama (even­ness = 3.23), followed Shimaki

(evenness = 3.02), and the Kashiwadai 1 flake assemblage (evenness = 2.38). Evenness is relatively low at Minamimachi 2 (evenness = 1.55) and for the Kashiwadai 1 blade/microblade assemblage (evenness = 1.10). The low evenness for the Kashiwadai 1 blade/microblade assemblage is strongly affected by the microblade richness (n = 626; 95.3% of the tools). When microblades are excluded, tool richness dramatically changes from 1.01 to 3.04, approaching the middle range of the study sites. In terms of the assemblage types:

58

Izuho 1) flake assemblages exhibit relatively higher diversity because of the high numbers of endscrapers and retouched flakes,

k calbp (2σ)

2) blade assemblages are dominated by relatively high numbers of burins and scrapers, and

3) blade/microblade assemblages are dominated by numbers of standardized tools and microblades (where n = 626).

Lithic Reduction Sequences Three types of reduction sequences are apparent in the LGM assemblages: flake reduction, blade reduction, and blade/ microblade reduction. Schematic flows of each reduction sequence are shown in Figure 3.10. At Kashiwadai 1, flakes were reduced at LC-11, and blades/microblades at LC1, 2, 3, 6, 12, 14, and 15. On-site production at Shimaki is dominated by flake reduction of local high-quality obsidian gravel, with incidental production of scrapers from blades detached from large blocky materials. Blade reduction using large blocky obsidian dominates at Kawanishi-C. Flake reduction is highest at Minamimachi 2 and Marukoyama.

Discussion and Conclusions

Reexamining the dates from Kashiwadai 1, a site containing both flake and blade/microblade reduction sequences, illustrates that this site may have been occupied for a period of 5800 years from at least 3000 years prior to the Siberian LGM to near its end. Minamimachi 2, Marukoyama, and Kawanishi-C, on the other hand, are concentrated at about 26,000 cal yr BP, and predate the Siberian LGM. Discrete distribution of flake assemblages at Kashiwadai 1 LC-11 may or may not be contemporaneous with blade/microblade assemblages at LC1, 2, 3, 6, 12, 14, and 15. My colleagues and I are currently analyzing lithic technology and spatial distribution of artifacts to better understand this variability.

Figure 3.8  Radiocarbon chronology for the LGM sites in the southern part of Paleo-SHK Peninsula. Bars represent 2σ age range.

Site location, quantities of stone tools, and toolkit diversity suggest that all site occupations were relatively brief (Nakazawa 2007; Terry 2012). Sites, however, functioned in two different ways: as relatively long-term processing sites such as Kashiwa-

Table 3.3  Toolkit diversity for the study sites.

Primary Pièce reduction Endscraper Sidescraper Burin Notch Perforator esquillée Wedge Kashiwadai 1 Flake 48 LC11 (n=165) (29.1%) Kashiwadai 1 Rankoshi- 14 LC1,2,3,6,12,14,15 type blade/ (2.1%) (n=657) microblade Shimaki 2010– Flake 25 2011 campaign (blade) (39.7%) (n=63) Minamimachi 2 Flake 2 Lower LC1 (n=10) (20%) Marukoyama Flake 8 Lower LC1 (n=34) (23.5%) Kawanishi-C Blade 67 layer VI LC1–12 (flake) (26.6%) (n=252) 1 Simpson’s diversity index (1/D)

19 (11.5%) 2 (0.3%)

0 0 (0.0%) (0.0%) 4 0 (0.6%) (0.0%)

0 (0.0%) 0 (0.0%)

4 (2.4%) 11 (1.7%)

0 (0.0%) 0 (0.0%)

Retouched Tool flake/blade Microblade richness 94 (57.0%) 0 (0.0%)

0 4 (0.0%) 626 5 (95.3%)

Tool evenness1 2.38 1.10

4 0 6 1 0 1 26 0 6 3.02 (6.3%) (0.0%) (9.5%) (1.6%) (0.0%) (1.6%) (41.3%) (0.0%) 8 0 0 0 0 0 0 0 2 1.55 (80%) (0.0%) (0.0%) (0.0%) (0.0%) (0.0%) (0.0%) (0.0%) 11 1 0 0 14 0 0 0 4 3.23 (32.4%) (2.9%) (0.0%) (0.0%) (41.2%) (0.0%) (0.0%) (0.0%) 82 40 0 7 1 0 55 0 6 4.05 (32.5%) (15.9%) (0.0%) (2.8%) (0.4%) (0.0%) (21.8%) (0.0%)

Human Technological and Behavioral Adaptation to Landscape Changes around the Last Glacial Maximum in Japan

59

Figure 3.9  LGM toolkit diversity at the study sites.

dai 1 LC11, Shimaki, and Kawanishi-C, and relatively short-term processing sites such as Minamimachi 2 and Marukoyama. Sites where blade reduction dominates lack distinct hunting weapons. In contrast, the sites where flake reduction dominates lack both distinct hunting weapons and formal burin tools. In addition, while the Shimaki assemblage is dominated by flake reduction, there are some blade-based tools. In contrast, flake reduction is minimal in the blade-based assemblage at the short-term camp of Kawanishi-C. These lines of evidence suggest that blade reduction and flake reduction are products of site function. In contrast, the assemblage of Kashiwadai 1, dominated by microblades, likely represents a more complete huntergatherer toolkit. This suggests that hunter-gatherers with different behavioral strategies equipped with both blade/microblade-based toolkits and flake- and blade-based toolkits coexisted in Hokkaido. On the other hand, future research could establish that this lithic variability reflects people using these sites at different times, even perhaps during different seasons, but this might be difficult to discern given the poor preservation of seasonal indicators at Hokkaido Paleolithic sites. These conclusions will be tested by studying site formations and by lithic analysis, including such aspects as intensity of tool retouch, contribution of local and exotic raw material, and obsidian sourcing to better understand technological organization and the timing and duration of site occupations.

Acknowledgments

I would like to thank Drs. Michael Waters, Ted Goebel, and Kelly Graf for the opportunity to present this paper. I also

thank Drs. Caroline Ketron and Kelly Graf again for editing and correcting the English. A deep thanks also goes out to Drs. Ian Buvit and Karisa Terry for collaboration both with field and laboratory work on the LGM prehistory of Hokkaido. Noriyoshi Oda helped prepare figures and tables for this paper. Financial support was provided by the Japan Society for the Promotion of Science KAKENHI, grant number 24320157 (PI: Masami Izuho).

References Cited

Adachi, N., K. Shibata, K. Umetsu, T. Kitano, H. Matsumura, R. Fujiyama, J. Sawada, and M. Tanaka  2011  Mitochondrial dna analysis of Hokkaido Jomon skeletons: Remnants of Archaic maternal lineages at the southwestern edge of former Beringia. American Journal of Physical Anthropology 146:346–60. Akai, F.  2005a  Ishikari-Teichitai-Nanbu niokeru Saisekiji Sekkigun no Kenkyu [A study of the microblade industry on southern Ishikari low-land]. Hokkaido Kyusekki Bunka Kenkyu 10:59–78. (In Japanese.) ———  2005b  Chitose-shi Marukoyama Iseki Eniwa-a Tephra joi Sekkigun no Saikento [Reexamination of the assemblage from above-En-a horizon of the Marukoyama site, Hokkaido]. Ronshu Oshorokko I: 103–21. (In Japanese.) ———  2008  Kamihoronai-Moi Sekkigun no Hakurikatei [Lithic reduction processes of the Palaeolithic assemblage of the Kamihoronai-Moi site, Hokkaido]. Ronshu Oshorokko II: 83–97. (In Japanese.) Anbiru, M.  1978  Sendoki Jidai no Kenkyu [A study of the Palaeolithic in Japan]. In Nihon Kokogaku wo Manabu 1, edited by H. Otsuka, M. Tozawa, and M. Sahara, pp.64–78. Yubikaku, Tokyo. (In Japanese.) Anzai, M.  1990  Mumoji Shakai no Kokogaku [The archaeology of the prehistoric society]. Rokko Shuppan, Tokyo. (In Japanese.)

[128.104.46.206] Project MUSE (2024-03-01 06:01 GMT) UW-Madison Libraries

60

Izuho

Figure 3.10 Schematic lithic reduction sequences of the LGM hunter-gatherer in the southern part of Paleo-SHK Peninsula.

Binford, L.  1980  Willow smoke and dogs’ tails: Hunter-gatherer settlement systems and archaeological site formation. American Antiquity 45(1):4–20. Binford, L. R., and Binford, S. R.  1966  A preliminary analysis of functional variability in the Mousterian of Levallois facies. American Anthropologist 68(2):238–95.

Ferguson, J. R., M. D. Glascock, M. Izuho, M. Mukai, K. Wada, and H. Sato  2013  (Accepted) Multi-method characterization of obsidian source compositional groups on Hokkaido Island (Japan). In Methodological Issues of Obsidian Provenance Studies and the Standardization of Geologic Obsidian, edited by A. Ono, Y. Suda, and M. Glascock. British Archaeological Report, International Series, Oxford.

Buvit, I., K. Terry, M. Izuho, and K. Hamaguchi  2011  Recent Excavation at the Shimaki Paleolithic site, Hokkaido, Japan. Current Research in the Pleistocene 28:1–3.

Graf, K. E.  2008  Uncharted Territory: Late Pleistocene Hunter-Gather Dispersals in the Siberian Mammoth-Steppe. Unpublished Ph.D. dissertation, Department of Anthropology, University of Nevada, Reno.

Chitose Board of Education  1994  Marukoyama Iseki niokeru Kokogakuteki Chosa [Excavations of Marukoyama site, Hokkaido]. Chitose-shi Kyoiku Iinkai. Chitose. (In Japanese.)

———  2009a  “The good, the bad, and the ugly”: Evaluating the radiocarbon chronology of the middle and late Upper Paleolithic in the Enisei River valley, south-central Siberia. Journal of Archaeological Science 36 (3):694–707.

Chlachula, J.  2001  Pleistocene climate change, natural environments and Palaeolithic occupation of the Altai area, west-central Siberia. Quaternary International 80-81:131–67.

———  2009b  Modern human colonization of the Siberian mammoth steppe: A view from south-central Siberia. In Sourcebook of Pa-

Human Technological and Behavioral Adaptation to Landscape Changes around the Last Glacial Maximum in Japan

61

leolithic Transitions, edited by M. Camps, and P. R. Chauhan, pp.479– 501. Springer, New York.

type burins in Sugikubo blade industry, central Japan: New evidence for versatile lithic tool use. Current Research in the Pleistocene 25:47–50.

———  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 37:210–23.

Iwase, A., J. Hashizume, and M. Izuho  2010  Nihon Retto no Koki Koshinsei Kohan ni okeru Rikusei Honyu Dobutsu So Kenkyu no Genjo to Kadai [Current achievements and issues in the study of the Late Last Pleistocene terrestrial mammals in the Japanese Archipelago]. Ronshu Oshorokko III:89–121. (In Japanese with English abstract.)

Goebel, T. and I. Buvit (eds.)  2011  From the Yenisei to the Yukon. Texas A&M University Press, College Station. Hayashi, K.  1968  The Fukui microblade technology and its relationship in Northeast Asia and North America. Arctic Anthropology 5:128–90. Hokkaido Center for Buried Cultural Property  1999  Kashiwadai 1 Iseki [Kashiwadai 1 site]. Hokkaido Maizo Bunkazai Center, Sapporo. (In Japanese.) Hoshino, F., and I. Matsuzawa  1987  Paleoenvironments when Mammuthus primigenius lived in Hokkaido, Japan—The coldest stage during Würm. Professor Masaru Matsui Memproal Volume, pp. 79–89. (In Japanese.) Igarashi, Y.  1990  Kafun Kaseki kara Saguru Shinrin no Rekishi: Hokkaido no 3 Man’nenkan [Pollen analysis for the forest history of past 30,000 years in Hokkaido]. Nihon Ringakkai Hokkaido Shibu Ronbunshu 38:1–9. (In Japanese.) ———  1996  A late glacial climatic reversion in Hokkaido, Northeast Asia, inferred from the Larix pollen record. Quaternary Science Reviews 15:989–95. ———  2008  Climate and vegetation changes since 40,000 Years BP in Hokkaido and Sakhalin. In Human Ecosystem Changes in the Northern Circum Japan Sea Area (NCJSA) in Late Pleistocene, edited by H. Sato, pp. 27–41. Research Institute for Humanity and Nature. Kyoto. (In Japanese with English abstract.) ———  2011  Hokkaido to Sakhalin ni okeru Saisyuuhyoukisaiseiki no Syokusei: Tokuni Sogen no Hattatu ni tuite. [The flora of Last Glacial Maximum in Hokkaido and Sakhalin]. In No to Hara no Kankyoshi, edited by H. Sato and K. Iinuma, pp. 17–34. Bun’ichi Sogo Shuppan. Tokyo. (In Japanese.) Igarashi, Y., and T. Oba  2006  Fluctuations in the East Asian monsoon over the last 144ka in the northwest Pacific based on a highresolution pollen analysis of IMAGES core MD01-2421. Quaternary Science Reviews 25:1447–59. Igarashi, Y., T. Igarashi, H. Daimaru, O. Yamada, T. Miyagi, K. Matsushita, K. Hiramatsu  1993  Hokkaido no Kenbuchi Bonchi to Furano Bonchi niokeru 32,000 Nenkan no Shokusei Hensen [Vegetation history of Kenbuchi Basin and Furano Basin in Hokkaido, North Japan, since 32,000 yrs BP]. Daiyonki Kenkyu 32:89-105. (In Japanese with English abstract.) Igarashi, Y., M. Murayama, T. Igarashi, T. Higake, and M. Fukuda 2002 History of Larix forest in Hokkaido and Sakhalin, northeast Asia since the Last Glacial. Acat Palaeontologica Sinica 41:524–33. Ishii, J., Y. Igarashi, S. Sasaki, N. Mino, and K. Matumoto  1981   Ishikari Wan Tairikudana yori Saishu shita Deitansou ni tuite [On the peats collected from the continental shelf of the Isikari Bay, Hokkaido, Japan]. Tikyuu Kagaku, 35:231–39. (In Japanese with English abstract.) Iwase, A.  2012  Saishu Hyoki Saiseiki no Honshu Tohanbu Nihonkai gawa Chiiki ni okeru Sekki Siyo no Tokucho [Lithic tool use during the last glacial maximum in eastern Honshu, Japan: Use-wear analysis on burin of the Sugikubo blade industry]. Kyusekki Kenkyu 8:65–89. (In Japanese with English abstract.) Iwase, A., and K. Morisaki  2008  Use-wear analysis of Kamiyama-

Iwase, A., J. Hashizume, M. Izuho, T. Takahashi, and H. Sato  2011  The Timing of Megafaunal Extinction in the late Late Pleistocene on the Japanese Archipelago. Quaternary International 255:114–24. Izuho, M.  1997  Tokorogawa Ryuiki niokeru Sekkisekizai no Kiso Kenkyu [A fundamental study of the lithic raw material on the Tokoro River Basin]. Hokkaido Kyusekki Bunka Kenkyu 2:1–14. (In Japanese.) ———  1998  Hokkaido Tokoro-gun Oketo-cho Ogachi-Kato Iseki niokeru Saishu Shiryo [Collected artifacts from the surface of the Ogachi-Kato site, Hokkaido]. Hokkaido Kyusekki Bunka Kenkyu 3:13– 28. (In Japanese.) Izuho, M., and F. Akai  2005  Hokkaido no Kyusekki Hennen: Iseki Keisei Kateiron no Tekiyo [Geochronology of Palaeolithic sites on Hokkaido]. Kyusekki Kenkyu 1:39–55. (In Japanese with English abstract.) Izuho, M., and W. Hirose  2010  A review of archaeological obsidian studies on Hokkaido Island (Japan). In Crossing the Straits: Prehistoric Obsidian Source Exploitation in the North Pacific Rim, edited by Y. V. Kuzmin and M. D. Glascockpp, pp. 9–25. British Archaeological Report, International Series 2152, Oxford. Izuho, M., and H. Oda  2008  Hokkaido Yufutsu-gun Atsuma-cho Kamihoronai-Moi Iseki Kyusekki Chiten no Hoshasei Tanso Nendai [AMS 14C dates for the Upper Palaeolithic site of Kamihoronai-Moi, Hokkaido]. Roshu Oshorokko II:13–18. (In Japanese with English abstract.) Izuho, M., and H. Sato  2008  Landscape evolution and culture changes in the Upper Paleolithic of northern Japan. In The Current Issues of Paleolithic Studies in Asia; Proceedings of the International Symposium “The Current Issues of Paleolithic Studies in Asia and Contiguous Regions,” edited by A. P. Derevianko and M. V. Shunkov, pp. 69–77. Publishing Department of the Institute of Archaeology and Ethnography SB RAS. Novosibirsk. Izuho, M., and K. Takahashi  2005  Correlation of Paleolithic industries and Paleoenvironmental change in Hokkaido (Japan). Current Research in the Pleistocene 22:19–21. Izuho, M., F. Akai, Y. Nakazawa, and A. Iwase  2012  The Upper Paleolithic of Hokkaido: Current evidence and its geochronological framework. In Environmental Changes and Human Occupation in East Asia during OIS3 and OIS2, edited by A. Ono and M. Izuho, pp. 109– 28, British Archaeological Report, International Series 2352, Oxford. Izuho, M., J. R. Ferguson, M. D. Glascock, N. Oda, F. Akai, Y. Nakazawa, and H. Sato  2013a (in print)  Integration of obsidian compositional studies and lithic reduction sequence analysis at the Upper Paleolithic site of Ogachi-Kato 2, Hokkaido, Japan. In Methodological Issues of Obsidian Provenance Studies and the Standardization of Geologic Obsidian, edited by A. Ono, Y. Suda, and M. D. Glascock.British Archaeological Report, International Series, Oxford. Izuho, M., D. Kunikita, N. Oda, T. Yamahara, and M. Kitazazawa  2013b Hokkaido Tokachi Heiya no Koki Kyusekki Jidai Iseki no Chishitsu Hennen [New AMS 14C dates for the Upper Paleolithic sites in the Tokachi Plain on Hokkaido (Japan)]. Kyusekki Kenkyu 9:137–48. (In Japanese with English abstract.)

62 Izuho, M., Y. Nakazawa, F. Akai, T. Soda, and H. Oda  2009  Geoarchaeological investigations at the Upper Paleolithic site of Kamihoronai-Moi, Hokkaido, Japan. Geoarchaeology 24:492–517. Japanese Paleolithic Research Association  2010  Nihon Retto no Kyusekki Jidai Iseki [Paleolithic sites in the Japanese Islands: A database]. Nihon Kyusekki Gakkai, Tokyo. (In Japanese.) Kanomata, Y.  2005  Saisekizin no Sochakuho to Siyoho: Araya Iseki, Tachikarushunai Dai V Iseki C chiten Shutudoshiryo no Bunseki kara [A way of use and insertion of microblades]. Koukogakuzassi 88(4):1–27. (In Japanese.)

Izuho Kuzmin, V. Y., S. V. Gorbunov, L. A. Orlova, A. A. Vasilevky, E. V. Alekseeva, A. N. Tikhonov, I. V. Kirillova, and G. S. Burr  2005  14C dating of the late-Pleistocene faunal remains from Sakhalin Island (Russian Far East). Current Research in the Pleistocene 22:78–80. Lambeck, K., T. M. Esat, and E. K. Potter  2002  Links between climate and sea level for the past three million years. Nature 419:199– 206. Machida, H., T. Oba, A. Ono, H. Yamazaki, Y. Kawamura, and A. Momohara 2004 Daiyonkigaku [Quaternary]. Asakura Shoten, Tokyo. (In Japanese.)

Kato, S.  1970  Sendoki Jidai no Rekishisei to Chiikisei [Historical and regional characteristics in Palaeolithic]. In Kyodoshi Kenkyu to Kokogaku, edited by T. Wakamori, pp. 58–92, Asakura Shoten, Tokyo. (In Japanese.)

Matui, H., R. Tada, and T. Ohba  1998  Low-salinity isolation event in the Japan Sea in response to eustatic sea-level drop during LGM: Reconstruction based on salinity-balance model. Quaternary Research 37:221–33. (In Japanese.)

———  1985  Siberia no Senshi Bunka to Nihon [The prehistory of Siberia and Japan]. Rokko Shuppan, Tokyo. (In Japanese.)

Morlan, R. E.  1967  The palaeolithic period of Hokkaido: An outline. Arctic Anthropology 4:164–220.

Kato, S., and M. Yamada  1988  Hokkaido Kato-gun Kamishihorocho Shimaki Isekino Sekki Bunka [A paleolithic culture at Shimaki: The oldest Palaeolithic components in Hokkaido]. Rekishi Jinrui 16:252–340. (In Japanese.)

Morisaki, K.  2010  Kyusekki Shakai no Kozoteki Henka to Chiiki Tekio [Structural changes and regional adaptation of Palaeolithic Society]. Rokuichi Syobou, Tokyo. (In Japanese with English abstract.)

Kimura, H.  1992  Reexamination of the Yubetsu technique and study of the Horokozawa Toma lithic culture. (In Japanese and English) ———  1995  Shiberia no Kyusekki Bunka [The study of Palaeolithic culture in Siberia]. Hokkaido Daigaku Tosho Kankokai, Hokkaido. (In Japanese with English abstract.) Kind, N. V.  1974  Late Quaternary geochronology according to isotope data (Geokhronologia pozdniego antropogena po izotopnym dannym). Transactions of the Geological Institute of the Academy of Science (In Russian.). Kirillova, I. V.  2003  Remains of vertebrates from the Tronnyi Grotto, Central Sakhalin. Kraevedchesky Bulleten 14(2):128–37. (In Russian.) Koaze H., M. Nogami, Y. Ono, and K. Hirakawa  2003  Nihon no Chikei 2: Hokkaido [Regional geomorphology of the Japanese Islands vol. 2: Geomorphology of Hokkaido]. The University of Tokyo Press, Tokyo. (In Japanese.) Kobayashi, T.  1970  Microblade industries in Japanese archipel­ ago. Arctic Anthropology 2:38–58. Kosaka, T. and Nogawa, K. 1974 Hokkaido Tokachi Kamishihorocho niokeru Kyusekki Hoganso no 14C Nendai [14C date of Palaeolithic layer in Kamishihoro town, Tokachi, Hokkaido]. Chikyu Kagaku 26:137–38. Kumon, F., S. Kawai, and Y. Inouchi  2003 Nojiriko Kotei Taisekibutsuchu no Yukitanso Zenchisso Ganyuritsu oyobi Kafun Bunseki ni motozuku Yaku 25,000–6,000 nenmae no Kikohendo [Climate changes between 25,000 and 6000 yrs BP deduced from TOC, TN, and fossil pollen analyses of a sediment core from Lake Nojiri, Central Japan]. Daiyonki Kenkyu 42(1):13–46. (In Japanese with English abstract.)

Morisaki, K., M. Izuho, and H. Sato  2010  Upper Paleolithic organizations response to landscape changes in Northern Japan. In The Initial Human Habitation of the Continental and the Insular Parts of the North East Asia, edited by A. Vasilevskii. Yujno-Sakhalinsk. Morisaki K., M. Izuho, K. Terry, and H. Sato  2013  Upper Paleolithic technological organizational responses to landscape changes in Northern Japan. Manuscript on file. Nara National Research Institute for Cultural Properties. Nakazawa, Y.  2001  Hokkaido Saishu Hyoki Kyokusoki niokeru Sozai Riyo karamita Hakuhen Sekkigun no Keisei: Chitose-shi Marukoyama Iseki En-a Kaso Sekkigun no Zokusei Bunseki niyoru Keishikigaku teki Kento [A formation of flake assemblage of Marukoyama site, Hokkaido, in Last Glacial Maximum]. Hokkaido Kyusekki Bunka Kenkyu 6:1–18. (In Japanese.) ———  2007  Hearth-Centered Spatial Organization: A comparative Approach to the Study of Palimpsests in Late Upper Paleolithic Site in Hokkaido (Japan) and Cantabria (Spain). Unpublished Ph.D. dissertation, Department of Anthropology, Universal of New ­Mexico, Albuquerque. ———  2010  Dual nature in the creation of disciplinary identity: A socio-historical review of Palaeolithic archaeology in Japan. Asian Perspectives 49:231–50. ———  2011  Keitaisei Sekijin Sekki no Koyo: Patch Riyo Model to Sekki Shohi no Setten wo Saguru [Utility of mobile blade tools: On the relationship between patch choice model and stone tool consumption]. Kyusekki Kenkyu 7:107–25. (In Japanese with English abstract.) Nakazawa, Y., M. Izuho, J. Takakura, and S. Yamada  2005  Toward an understanding of technological variability in microblade assemblages in Hokkaido, Japan. Asian Perspectives 44:276–92.

———  2009  Nojiriko Taisekibutsu ni motozuku Chubu-Nihon no Kako 7.2 mannenkan no shosaina Kokiko Fukugen [High-resolution reconstruction of paleoclimate during the last 72 ka on the basis of the drilled sediments from Lake Nojiri, central Japan]. Kyusekki Kenkyu 5:3–10. (In Japanese with English abstract.)

Nakamura, T.  2005  Hokkaido no Senshi Jidai Iseki kara Shutsudoshita Tankamoppen oyobi Kurumikara no AMS14C Nendai [AMS 14C dates for the prehistoric sites on Hokkaido]. In Hinetsu Kokogakuteki Kokuyoseki no Fission Track Nendai niyoru 20,000 nengoe no 14C Nendai no Kosei, edited by M. Suzuki. Rikkyo University, Tokyo. (In Japanese.)

Kunitake, S.  2005  Kouki kyusekkijidai zenhanki no kyoju kodo no hensen to gijutu kozo no henyo [A study on the changes of settlement system and lithic technology in the Early Upper Palaeolithic]. Busshitsu bunka 78:1–25. (In Japanese.).

Naoe, Y.  2009  Hokkaido-Tobu niokeru Kokuyoseki Riyo [Obsidian procurement and consumption on northeastern Hokkaido]. In Kokuyoseki ga hiraku Jinrui Shakai no Koryu, edited by H. Sato, pp. 14–31. The University of Tokyo, Tokyo. (In Japanese.)

Human Technological and Behavioral Adaptation to Landscape Changes around the Last Glacial Maximum in Japan Naoe, Y., and J. Nagasaki  2005  Hokkaido Koki Kyusekki Jidai Zenhan Ki no Sekizai Shohi Senryaku; Shirataki Ia Gun to Shukubai Iseki Sankakuyama Chiten [Lithic raw material consumption strategy at Late Upper Palaeolithic on Hokkaido]. Hokkaido Kyusekki Bunka Kenkyu 10:45–58. (In Japanese.) Nishi, Y.  1991  Hokkaido Kushirooki Hakken no Yuzetsu Sentoki nitsuite [A tanged point from the sea bottom of Kushiro, Hokkaido]. Kushiro Ichiritsu Hakubutsukan Kiyo 16:21–24. (In Japanese.) Oba, T., and W. Matsushita  1965  Hokkaido no Sendoki Jidai [The Palaeolithic on Hokkaido]. In Sendoki Jidai, edited by S. Sugihara, pp.174–97. Kawade Shobo, Tokyo. (In Japanese.) Obihiro Board of Education  1995  Obihiro Minamimachi Iseki [Minamimachi Site, Obihiro]. Obihiro-shi Kyoiku Iinkai. Obihiro. (In Japanese.) ———  1998  Obihiro Kawanishi C Iseki [Kawanishi C site, Obihiro]. Obihiro-shi Kyoiku Iinkai. Obihiro. (In Japanese.) ——— 2000a Obihiro Kawanishi C Iseki 2 [Kawanishi C site, Obihiro volume 2]. Obihiro-shi Kyoiku Iinkai. Obihiro. (In Japanese.) ——— 2000b Obihiro Kawanishi C Iseki 3 [Kawanishi C site, Obihiro volume 3]. Obihiro-shi Kyoiku Iinkai. Obihiro. (In Japanese.) Oda, N.  2009  Kitami-shi Momijiyama Iseki Shutsudo Sekkigun no Saikento [Reevaluation of the Momijiyama site, Hokkaido]. In Nihonrettoh Hokubu no Koshinsei/Kanshinsei Ikoki niokeru Kyojukeitai to Bunkakeisei nikansuru Kenkyu, edited by H. Sato, pp. 139–221. The University of Tokyo, Tokyo. (In Japanese with English abstract.) Ono, Y. 1990 Kita no Rikkyo [Northern Land Bridge]. Daiyonki Kenkyu 29:183-192. (In Japanese with English abstract.) Ono, Y., and Y. Igarashi  1993  Hokkaido no Shizenshi: Hyoki no Shinrin wo Tabi suru [Natural history in Hokkaido: A journey of ice age forest]. Hokkaido University Press. Sapporo. (In Japanese.) Ozawa, S., S. Ohtsu, Y. Igarashi  1995  Quaternary paleoenvironmental changes from drilling core of Shinkouhigashi observation well on Ishikari Bay, Hokkaido, Japan. Proceeding of Geological Society of Japan 102 tailai, pp. 110. Sakaguchi, Y.  1989  Some pollen records from Hokkaido and Sakhalin. Bulletin of the Dept. Geography University of Tokyo 21:1–17. Sato, H.  1992  Nihon Kyusekki Bunka no Kozo to Shinka [Evolution and structure of Paleolithic culture in Japan]. Kashiwa Shobo, Tokyo. (In Japanese.) ———  2003  Hokkaido no Koki Kyusekki Jidai Zenhanki no Yoso [The Early Upper Paleolithic on Hokkaido]. Kodai Bunka 55:181–94. (In Japanese.) Sato, H., and M. Izuho  2011  Sokatsu: Kannihonkai Hokubu Chiiki niokeru Kokikoshinsei no Kankyohendo to Ningen no Sogosayo no Hensenshi [A summary: The history of human and environmental fluctuation in the Northern Circum Japan Sea Area in the Pleistocene]. In Kannihonkai Hokubu Chiiki niokeru Kokikoshinsei no Kankyohendo to Ningen no Sogosayo nikansuru Sogoteki Kenkyu, edited by H. Sato, pp. 225–32. The University of Tokyo, Tokyo. (In Japanese.)

63

Shiraishi, N.  1993  Hokkaido niokeru Saisekiji Sekkigun no Tenkai [The microblade industries on Hokkaido]. Busshitsu Bunka. 56:1– 22. (In Japanese.) Shott, M.  1986  Technological organization and settlement mobility: An ethnographic examination. Journhal of Anthropological Research 42(1):15–51. Sugihara, S. (ed.)  1965  Sendoki Jidai [The Palaeolithic in Japan]. Kawade shobo, Tokyo. (In Japanese.) Suzuki, H.  2007  Gensanchi niokeru Isekikanheni Kenkyu [A study of inter-site variability on the Shirataki site group]. Kokogaku Ronso, pp. 109–29, Tohoku University, Miyagi. (In Japanese.) Suzuki, M., M. Yoshikawa, T. Murata  1999  Toshiribetsugawa Ryuuiki Teiti ni okeru Koushinsei makki ikou no Kankyou hensen [Environmental changes in the Shiribetsu-Toshibetsugawa Lowland Hokkaido, Japan since the Latest Pleistocene]. Bulletin of the National Museum of Japanese History, 81:371–86. (In Japanese.) Svendsen, J. I., H. A. Alexanderson, V. I. Astakhov, I. Demidov, J. A. Dowdeswell, S. Funder, V. Gataullin,M. Henriksen C. Hjort, M. Houmark-Nielsen,H. W. Hubberten, O. Ingolfsson, M. Jakobsson, K. H. Kjaer, E. Larsen, H. Lokrantz, J. P. Lunka, A. Lysa, J. Mang, A. Matiouchkov, A. Murray, P. Moller, F. Niessen, O. Nikolskaya, L. Polyak, M. Saarnisto, C. Siegert, M. J. Siegert, R. F. Spielhagen, and R. Stein  2004  Late Quaternary ice sheet history of northern Eurasia. Quaternary Science Reviews 23:1229–71. Takahashi, K.  2007  Nihon Retto no Senshin- Koshin Sei niokeru Rikusei Honyu Dobutsuso no Keisei Katei [The formative history of the terrestrial mammalian fauna of the Japanese Islands during the Plio- Pleistocene]. Kyusekki Kenkyu, 3:5-14. (In Japanese with English abstract.) Takahashi, K., and M. Izuho  2012  Formative history of terrestrial fauna of the Japanese Islands during the Plio-Pleistocene. In Environmental Changes and Human Occupation in East Asia during OIS3 and OIS2, edited by A. Ono and M. Izuho, pp. 73–86, British Archaeological Report, International Series 2352, Oxford. Takahashi, K., A. Iwase, and M. Izuho  2013  (submitted) Formative process of the Japanese terrestrial mammalian fauna without large species during MIS3 and MIS2. In The Emergence and Diversity of Modern Human Behavior in Paleolithic Asia, edited by Y. Kaifu, M. Izuho, H. Sato, and A. Ono. Texas A&M University Press, College Station. Takahashi, K., Y. Soeda, M. Izuho, G. Yamada, M. Akamatsu, and C. Chun-Hsiang  2006  The chronological record of the woolly mammoth (Mammuthus primigenius) in Japan, and its temporary replacement by Paleoloxodon naumanni during MIS 3 in Hokkaido (northern Japan). Palaeogeography, Palaeoclimatology, Palaeoecology 233:1–10. Takakura, J.  2000  Hokkaido Kitami-shi Yoshiizawa Iseki B chiten Shutudo Saisekijin Sekkigun no Saikento [Reevaluation of the Yoshiizawa site Location B, Hokkaido]. Hokkaido Kyusekki Bunka Kenkyu 5:1–34. (In Japanese.)

Sato, H., S. Yamada, and M. Izuho  2011  Kyusekki Jidai no Shuryo to Dobutsu Shigen [Animal resource and hunting in Upper Paleolithic in Japanese Islands]. In No to Hara no Kankyoshi, edited by H. Sato and K. Iinuma. Bun’ichi Sogo Shuppan. Tokyo. (In Japanese.)

Tamura, T.  1992  Tooiyama Kuroiishi: musashino 2ki sekkigun no shakai seitaigakuteki ichi kousatsu [A socio-ecological study on the lithic assemblages of ‘Musashino 2nd stage’]. Senshi Kokogaku Ronshu 2: 1–46. (In Japanese.)

Serizawa, C.  1960  Sekkijidai no Nihon [The Stone Age in Japan]. Tsukiji Shokan, Tokyo. (In Japanese.)

Terry, K., M. Izuho, I. Buvit, K. Hamaguchi, and K. Hor  2012  Current investigations at the Shimaki Site—An LGM occupation in Hokkaido, Japan. Paper presented at 77th Society for American Archaeology, Memphis.

Shiga, K., and I. Koizumi  2000  Latest Quaternary oceanographic changes in the Okhotsk Sea based on diatom records. Marine Micropaleontology 38:91–117.

Torrence, R.  1983  Time budgeting and hunter-gatherer technol-

64

Izuho

ogy. Hunter-gatherer Economy in Prehistory: A European Perspective. pp. 11–22. Cambridge University Press.

(NCJSA) in Late Pleistocene, edited by H. Sato. pp. 44–67. Research Institute for Humanity and Nature. Kyoto. (In Japanese and Russian.)

Tozawa, M.  1990  Nihon Sendoki Bunka no Kozo [A structure of the Palaeolithic age in Japan]. Dohosha Shuppan, Tokyo. (In Japanese.)

Yamada, A.  1986  Hokkaido Koki Kyusekki Jidai niokeru Sekki Seisaku Gijutsu Kozo no Hensen nikansuru Yosatsu [A preliminary examination of lithic reduction sequences on the Upper Palaeolithic in Hokkaido]. Kokogaku Zasshi 71(4):1–29. (In Japanese.)

Tsurumaru, T.  1979  Hokkaido Chihono Saisekijin Bunka [Microblade culture in Hokkaido District]. Sundai Shigaku 47:23–50. (In Japanese.) Tsurumaru, T.  2001  Hokkaido Kyusekki Kokogaku no Ronten: Kon-nichiteki Shoten to Tenbo [Contemporary issues and perspectives of Palaeolithic archaeology on Hokkaido]. Hokkaido Kokogaku, 37:3–22. (In Japanese.) Tsutsumi, T.  1997  Araya-gata Chokokutogata Sekki no Kino Suitei. [A function of the Araya-type burin]. Kyusekki Kokogaku 54:17–35. (In Japanese with English abstract.) Ueno, S., and M. Kato  1973  Tohoku Chiho no Saisekijin Gijutsu to sono Hokkaido tono Kanren nitsuite [Microblade technologies of Tohoku region and relation to Hokkaido]. Hokkaido Kokogaku 9:25– 49. (In Japanese.) Vasilevski, A. A.  2005  Periodization of the Upper Paleolithic of Sakhalin and Hokkaido in the light of research conducted at the Ogonki-5 site. In The Middle to Upper Paleolithic Transition in Eurasia: Hypothesis and Facts, edited by A. P. Derevianko. Institute of Archaeology and Ethnography Press, Novosibirsk. ———  2008  Sakhalin niokeru Mammoth Dobutsugun to Jinrui no Tekio [Mammoth Fauna and Human Adaptation in Sakhalin]. In Human Ecosystem Changes in the Northern Circum Japan Sea Area

Yamada, S.  1999  Hokkaido no Zenhanki Saisekijin Sekkigun nituiteno Kenkyu [A study of the early microblade industry on Hokkaido]. Senshi Kokogaku Ronshu 8:1–70. (In Japanese.) ——— 2006 Hokkaido niokeru Saisekijin Sekkigun no Kenkyu [A study of microblade assemblages in Hokkaido, Japan]. Rokuichi Shobo. Tokyo. (In Japanese.) Yokoyama, Y., Y. Kudo, R. Tada, I. Minami, R. C. Finkel, and H. Matsuzaki  2007  Japan Sea oxygen isotope stratigraphy and global sealevel changes for the last 50,000 years recorded in sediment cores from the Oki Ridge. Palaeogeography, Palaeoclimatology, Palaeoecology 247:5–17. Yonemura, M.  1983  Hokkaido niokeru Sendokijidai Shumatsu no Yoso [The last stage of the Upper Palaeolithic on Hokkaido]. Sundai Shigaku 58:25–45. (In Japanese.) Yoshizaki, M.  1956  Hokkaido Kitami Ainonai-mura Toyoda Iseki Gaiho [A summary report of Toyoda site, Hokkaido]. Sekki Jidai 1:46– 48. (In Japanese.) Yoshizaki, M.  1958  Hokkaido no Mudoki Bunka nitsuite [The Palaeolithic sites on Hokkaido]. Kyodo no Kagaku 19:2–6. (In Japanese.)

Chapter 4 Siberian Odyssey Kelly E. Graf Abstract Human dispersal to the Americas was a complex process. Both place of origin and timing of this event are hotly debated. Based on genetics, geography, language, and cultural similarities, most researchers consider Siberia the homeland of the First Americans with migration via the Bering Land Bridge. Others, however, argue earliest colonizers originated in Western Europe, arriving via a trans-oceanic voyage. Some hold that this early colonization event took place before the Last Glacial Maximum (LGM), while others contend it happened much more recently during the Late Glacial. In this chapter, I review Siberian Upper Paleolithic archaeology. The Siberian record indicates two pulses of modern humans into far northeast Asia during the late Pleistocene, one before and one after the LGM. The colonization of Siberia by modern humans was an episodic process, taking over 10,000 years, setting the stage for the initial peopling of the Americas. Keywords:  Siberia, Upper Paleolithic, LGM

Introduction

Archaeologists have long looked to Siberia as the homeland of the First Americans. Despite resistance from some archaeologists (see Collins et al. this volume; Bradley and Stanford 2004; Stanford and Bradley 2012), mounting evidence mostly from molecular geneticists investigating dna from both ancient skeletons and living populations has very convincingly illustrated that all founding lineages and sub-lineages of First Americans originated in greater Northeast Asia and came to the Americas via a single migration (de Saint Pierre et al. 2012; Derenko et al. 2001; Fagundes et al. 2008; Fu et al. 2013; Gilbert et al. 2008; Kashani et al. 2013; Kemp et al. 2007; Kitchen et al. 2008; Mulligan and Kitchen, this volume; O’Rourke and Raff 2010; Tamm et al. 2007). Clearly if we are to understand where First Americans came from, then dna studies are the definitive way to do this because try as we may, we cannot make silent stones and bones speak about their makers’ origins. Because Siberia is likely the Pleistocene homeland of Center for the Study of the First Americans, Department of ­Anthropology, 234 Anthropology Building, 4352-TAMU, Texas A&M University, College Station, TX 77843-4352; e-mail:  [email protected]

Native Americans, I want to turn attention to the archaeological record of this vast region of Northeast Asia to look at the behaviors that conditioned the timing and process of dispersal to Beringia and the New World. The Siberian Upper Paleolithic record is traditionally divided into three phases: early, middle, and late Upper Paleolithic (Vasil’ev 1992). These phases are based on typological and chronological distinctions. The record is characterized by peaks and nadirs of dated cultural layers (or occupations). When compared with global climate records, high points in occupation numbers tend to align with warm intervals, while lows in occupation numbers tend to correspond with cold intervals during the second half of the late Pleistocene (Graf 2005). In the sections that follow, I provide a broad overview of the archaeological record of modern humans in Siberia during the late Pleistocene and follow up with a discussion of how this record can inform on dispersals north and to the Americas.

Early Upper Paleolithic

Anatomically modern Homo sapiens were not the first people to inhabit southern Siberia, evidenced by archaeological and skeletal remains of Neanderthals and Denisovans at Denisova and Okladnikov Caves in southwestern Siberia (Derevianko

66 2010; Green et al. 2010; Krause et al. 2007; Reich et al. 2010; Turner 1990). These occupations probably happened before 50,000 years ago, presumably during the last interglacial/ early glacial or marine isotope stages (MIS) 5–4 (Goebel 1993, 1999, 2002b; Derevianko et al. 1998), so their ages cannot be established using radiocarbon (14C) dating. A series of radio-thermoluminescence or RTL dates obtained on site sediments indicates the Middle Paleolithic layers date to late Middle Pleistocene times or before about 128,000 years ago (Derevianko et al. 1998, 2003). Despite not being the “first” to Siberia, early modern humans were certainly the first to disperse north and east into sub-arctic and eventually arctic Siberia, Beringia, and ultimately the New World (Goebel 1993, 1999; Mulligan and Kitchen this volume; Neves et al. this volume; Pitulko et al. this volume; Potter et al. this volume). Earliest modern humans were certainly capable of inhabiting a wide array of habitats because as early as about 47,000–40,000 years ago their cultural remains were distributed from sites in Southwest Asia (Bar-Yosef 2000; Kuhn 2002; Marks 1990) north and east to sites such as Kostenki and Kara Bom found in cold, dry environments of Eastern Europe and southern Siberia, respectively (Anikovich et al. 2007; Goebel et al. 1993). By about 38,000 years ago modern humans expanded as far north as the Arctic Circle (~66˚ N) as evidenced by a handful of lithic artifacts and faunal remains at Mamontovaya Kurya in the western foothills of the Ural Mountains of European Russia (Pavlov et al. 2001, 2004) and eventually arctic Northeast Asia, where thousands of lithic and osseous cultural materials were found at Yana RHS (71˚ N) in northwestern Beringia (Pitulko et al. 2004, this volume). The initial pulse into northern environments is recognized, not by skeletal remains, but by the durable artifacts left behind. Below, I explore the geographical and temporal framework of the early Upper Paleolitihc (EUP) and then review the archaeological record of these sites in Siberia. EUP Distribution and Chronology Hundreds of Paleolithic sites dot the Siberian landscape; however, EUP sites have been found sparsely distributed across just the territory south of 55˚ N from the upper Ob’ River in the west to Transbaikal in the east (Figure 4.1). Localities with EUP occupations are not numerous, and many of these have contextual problems. Though at least 25 EUP sites are reported, only 8 of these were found in secure contexts with meaningful data to review, and therefore provide the corpus of data on the EUP archaeological record. In this chapter, 14C dates were calibrated using the IntCal09 curve (Reimer et al. 2009) in Calib 6.0 14C calibration program. In the northern foothills of the Altai Mountains near the headwaters of the Ob’ River in southwestern Siberia, EUP artifacts were discovered at seven sites. Four sites, Anui-1, Anui-3, Ust’ Karakol, and Kara Bom, were found in open-air contexts on terrace-like surfaces mantled by colluvium and overlooking the Anui and Ursul rivers (Derevianko et al. 1998; Goebel et al. 1993). The other three sites, Denisova, Ust’

Graf Kanskaia, and Maloialomanskaia caves, overlook the Anui, Charysh, and Katun’ river drainages, respectively. Of these, Ust’ Karakol and Kara Bom represent the initial pulse of modern humans into this region with their clean EUP assemblages and relatively reliable 14C dates (Derevianko et al. 1998, 2003; Goebel 2002b, 2004a; Goebel et al. 1993). Kara Bom contains at least four discrete EUP cultural layers (2a–2d) within a single geological stratum (4) following Goebel et al. (1993) or from two geological strata (6–5) following Derevianko and Rybin (2005). Wood charcoal from these layers has produced eight 14C dates, placing EUP occupations between about 43,300 ± 1600 (GX-17596) and 30,990 ± 460 (GX-17594) 14 C years before present (yr BP) or about 49,740–34,770 calendar (cal) yr BP, with the lower two occupation horizons dated to 49,700–44,300 cal yr BP and the subsequent horizons 40,800–34,600 cal yr BP. Ust’ Karakol contains a single EUP cultural component found in four geological strata (11–8). Seven wood charcoal samples from three hearth features give an age range of 35,100 ± 2850 (SOAN-3259) to 29,720 ± 360 (SOAN-3359)14C (44,600–33,400 cal) yr BP for the component. The other five sites come from dubious contexts where either their small, yet diagnostic, assemblages have no associated chronometric dates (e.g., Anui-1) or from EUP artifacts found with Middle Paleolithic artifacts in stratigraphically mixed contexts in colluvial or cave deposits (e.g., Anui-3, Denisova, Ust’ Kanskaia, and Maloialomanskaia Caves) (Derevianko and Markin 1990; Derevianko et al. 1990, 1998; Goebel 1993, Goebel 2002b, 2004a; Tseitlin 1979). Several EUP sites were discovered in the northern foothills of the Saian Mountains and Lena-Angara Plateau in the upper reaches of the Enisei, Angara, and Lena rivers in southcentral Siberia. All archaeological sites found from the Enisei River east to Lake Baikal are discussed together below. In this region sites were discovered in open-air settings on fluvial terraces mantled by a combination of colluvial and eolian deposits. The best evidence of EUP was found at Malaia Syia and Makarovo-4 (Abramova et al. 1991; Drozdov et al. 1990: Goebel 2002b, 2004a; Larichev et al. 1988; Muratov et al. 1982). There are several more that could represent EUP occupations such as Ust’ Maltat-2, Sosnovyi Bor, Arembovskii, and Voennyi Gospital; however, these sites come either from geological contexts that lack datable materials or from artifacts found in surface contexts that cannot be reliably associated with dates from nearby profiles (Akimova et al. 2010a, 2010b; Goebel 1993, 2002a, 2004a; Medvedev 1983; Medvedev et al. 1990; Muratov et al. 1982, Semin et al. 1990; Tseitlin 1979). Malaia Syia is an open-air site that contains a 3-m-thick profile of loess with 5 stratigraphic layers and a single cultural layer containing an EUP artifact assemblage found in the paleosol of stratum 4 (about 2.5 m below the surface) (Muratov et al. 1982). Four dates are reported from this paleosol, three on bone samples from the faunal assemblage and one on a “grab-sample” of charcoal from the paleosol. Two bone dates of 34,500 ± 450 (SOAN-1286) and 34,420 ± 360 (SOAN1287) 14C yr BP place the age of the EUP component within the range of 40,800–38,600 cal yr BP. The other two dates are

Siberian Odyssey

67

Figure 4.1  Map of Northern Asia with locations of Paleolithic sites discussed in text: 1, Anui, Ust’ Karakol, Denisova Cave; 2, Kara Bom; 3, Ust’ Kanskaia Cave; 4, Maloialomanskaia Cave; 5, Shestakovo; 6, Chernoozer’e; 7, Malaia Syia; 8, Achinsk; 9, Afontova Gora; 10, Ust’ Maltat, Derbina, Listvenka; 11, Kurtak, Kashtanka; 12, Novoselovo, Kokorevo; 13, Sabanikha, Tashtyk; 14, Ui, Maina; 15, Golubaia, Nizhnii Idzhir’; 16, Ust’ Kova; 17, Igeteiskii Log; 18, Mal’ta, Buret’, Sosnovyi Bor; 19, Arembovskii, Voennyi Gospital; 20, Makarovo; 21, Nepa; 22, Alekseevsk; 23, Studenoe, Ust’ Menza, Priiskovoe, Chitkan, Kosaia Shivera; 24, Kunalei, Podzvonka; 25, Varvarina Gora, Kamenka; 26, Tolbaga, Masterov Kliuch, Masterov Gora; 27, Khotyk, Sannyi Mys, Sapun; 28, Sokhatina; 29, Diuktai Cave; 30, Ust’ Mil’, Verkhne Troitskaia; 31, Ikhine; 32, Yana; 33, Berelekh; 34, Ushki Lake; 35, Ogon’ki; 36, Kashiwadai.

aberrantly too young (Goebel 1993; Lisitsyn 2000; Muratov et al. 1982). Makarovo-4 is another open-air site that consists of a 4-m-thick profile of colluvial and aeolian deposits with four stratigraphic layers and a single EUP cultural layer. Sandblasted artifacts were found about 2 m below the surface lying atop a sandy scree, wind-deflated lag deposit, underlying stratum 2 and capping stratum 3. The cultural layer with overlying and underlying sediment was penetrated by a network of icewedge pseudomorphs beginning in stratum 2 and extending down through stratum 3 (Aksenov 1989). Despite these postdepositional site-formation problems, the EUP assemblage appears intact as a component, and three infinite AMS dates all > 39,000 (AA-8880, AA-8878, AA-8879)14C yr BP on three bone samples found in situ during excavations indicate an age > 45,000 cal yr BP (Goebel and Aksenov 1995). Similar to the Altai region, most proposed EUP site occupations in south-central Siberia come from problematic contexts. At the Ust’ Maltat-2 site, which is being actively eroded by the Krasnoiarsk Reservoir, EUP artifacts were found erod-

ing from a redeposited stratigraphic setting and not directly dated (Akimova et al. 2010b). Arembovskii, and Sosnovyi Bor produced characteristic EUP assemblages but remain undated (Goebel 1993, 2004a; Medvedev 1983). Voennyi Gospital was first excavated in 1871, and the original EUP assemblage was subsequently lost to a fire in 1879. Renewed excavations in the 1980s produced a small, unexpressive assemblage with a 14C date on horse bone of 29,700 ± 500 (GIN-4440) 14C (35,200– 33,100 cal) yr BP, which may or may not date the original EUP occupation (Goebel 2004a; Medvedev et al. 1990). In the Transbaikal in southeastern Siberia, several EUP sites have been reported in open-air settings on alluvial terrace landforms mantled by a combination of colluvial and eolian deposits. Three of these sites contain EUP cultural occupations that were found in reasonably reliable stratigraphic situations, including Varvarina Gora, Masterov Kliuch’, and Kamenka (Bazarov et al. 1982; Goebel 2004a; Goebel and Aksenov 1995; Goebel et al. 2000a; Lbova 1996, 2000, 2005; Okladnikov and Kirillov 1980). Varvarina Gora is an open-air

68 site that contains a 2.5-m-thick profile of colluvial and eolian deposits with four strata and a single cultural layer containing an EUP artifact assemblage found in stratum 3 (about 1.5 m below the surface) (Goebel 2004a; Okladnikov and Kirillov 1980). Three conventional 14C dates on bones from the faunal assemblage suggest an age range of 34,900 ± 780 (SOAN1524) to 29,895 ± 1790 (SOAN-3054) 14C (41,600–31,100 cal) yr BP (Bazarov et al. 1982; Lbova 2005), but two AMS bone dates were infinite (> 34,050 [AA-8875] and > 35,300 [AA-8893] 14C or > 38,500 cal yr BP) (Goebel and Aksenov 1995). The infinite AMS dates, plus the probability that the conventional bone dates are too young, suggests an age of > 38,500 cal BP for the cultural layer. Masterov Kliuch’ is situated in 160-cm-thick colluvial deposits with six stratigraphic units and two EUP cultural layers. Detailed geoarchaeological work found the lower EUP layer in a primary depositional context, but the upper component was not. Two AMS bone dates from the faunal assemblage associated with the lower component produced dates of 32,510 ± 1440 (AA-23640) 14 C and 29,860 ± 1000 (AA-23641) 14C yr BP, indicating an age range of 40,700–32,000 cal yr BP (Goebel et al. 2000a). Kamenka was found in a 12-m-thick set of colluvial deposits in which an EUP cultural component, complex A/C, was found underlying a middle Upper Paleolithic (MUP) complex B. Six conventional 14C dates on bones were reported for complex A/C (Lbova 2005). Four indicate an age range of 35,845 ± 695 (SOAN-2904) to 30,220 ± 270 (SOAN-3354) 14C yr BP or about 42,000–34,500 cal yr BP. The other two dates seem aberrant with one far older than the rest (40,500 ± 3800 [AA-26743] 14 C yr BP and the other (26,760 ± 265 [SOAN-3353] 14C yr BP) far too young, especially since it is statistically the same age as the dates from overlying complex B. Several other sites in the Transbaikal hint at EUP occupation, but they were found in problematic contexts. Sannyi Mys, Masterov Gora, Sapun, and Sokhatino have EUP assemblages that remain undated (Aseev and Kholiushkin 1985; Goebel 1993; Kirillov and Kasparov 1990; Okladnikov 1971; Okladnikov and Kirillov 1980), while both Podzvanka and Khotyk each have Middle and Upper Paleolithic artifacts found in the same colluvial stratum, suggesting mixed stratigraphic contexts (Buvit et al. 2011; Lbova 2000, 2005; Tashak 2000, 2002). Another problem is Tolbaga. This site contains a massive EUP component also found in colluvial sediments. All artifacts and bones were oriented downslope so it is fairly clear that the archaeological materials are secondarily deposited. The five 14C bone dates (three conventional and two AMS) span 34,860 ± 2100 (SOAN-1522) to 25,200 (AA-8874) 14 C (43,000–29,500 cal) yr BP (Bazarov et al. 1982; Goebel and Waters 2001; Orlova 1998). The very long time range reflected by the dates may mean the component consists of a mixture of EUP and later MUP artifacts. EUP Technologies EUP lithic assemblages are blade-based and flake-based technologies. Raw-material procurement and selection focused on use of fine-grained, high-quality stones such as argillite,

Graf chert, quartzite, and, to a lesser extent, basalt. Most raw materials were local stones found close to the sites in nearby stream alluvium (Goebel 2004a). Primary reduction strategies focused on production of blades as tool blanks. Blade cores were large, and either parallel (flat-faced) or sub-prismatic in character. Blade removal proceeded in either a unidirectional or bidirectional fashion so that cores either possessed a single striking platform or two platforms, respectively. Typically, a EUP core began its use life as a large parallel core, depending on the size of the original cobble. Through reduction, multiple sides (or fronts) of the core would be used, and eventually the core would take on a sub-prismatic shape. On occasion, more informal flake cores were used (Goebel 1993, 2004a). Secondary-reduction activities largely centered on manufacturing tools on elongated blades. Tool retouch was primarily unifacial with resharpening flakes removed from dorsal surfaces of the tool blank. Burination occurred in nearly all assemblages with production of angle burins on snaps. Bifacial reduction occurred in nearly every EUP assemblage; however, bifaces are present in low frequencies in individual assemblages. Typical EUP tool types include unifacial points, retouched blades, side scrapers, end scrapers, bifaces, notches, denticulates, gravers, burins and wedges (Goebel 1993, 2002b, 2004b). EUP assemblages contain osseous artifacts that include both utilitarian and nonutilitarian pieces. Nearly every EUP cultural occupation has an inventory of non-lithic tools. These items include small antler points, awls and needles made on bones, and ivory and bone retouching implements. By contrast, nonutilitarian pieces are rare in EUP assemblages, appearing in very low numbers (typically fewer than five) in the assemblages from Kara Bom, Voennyi Gospital, Kamenka, and Tolbaga (Derevianko and Rybin 2005; Goebel 2004a; Lbova 2000; Vasil’ev et al. 1987). Mostly these pieces consist of bone beads, bone-bead preforms (or cylindrical beads), and bone and teeth pendants. Furthermore, a single polished-stone pendant was found at each of the sites of Varvarina Gora and Ust’ Karakol (Derevianko and Rybin 2005). Osseous and stone jewelry pieces were also found in mixed levels at the sites of Denisova Cave, Maloialomanskaia Cave, Ust’ Kanskaia Cave, Podzvonka, and Khotyk (Derevianko and Rybin 2005; Derevianko et al. 2008; Lbova 2000; Tashak 2000, 2002) and likely belong with the EUP components of these sites. EUP Fauna Generally, Siberian Paleolithic faunal assemblages are few, and typically only presence or absence of data is available in the literature. This is certainly the case for the EUP. Of the sites discussed above, fauna from only half or 12 sites is reported (Vasil’ev 2003a). Most assemblages provide a long list of fauna representing a wide range of habitats such as forest, forest-steppe, and tundra. Typical taxa include bison, yak, woolly rhinoceros, horse, wild ass, red deer, roe deer, reindeer, Argali sheep, Mongolian gazelle, Kiakhta antelope, Siberian mountain goat, bear, wolf, fox, hare, and ground

Siberian Odyssey squirrel and other rodents. Incidentally, data on number of identified specimens (NISP) and minimum number of individuals (MNI) are available from five sites (Malaia Syia, Tolbaga, Varvarina Gora, Kamenka, and Podzvonka) (Germonpré and Lbova 1996; Ovodov 1987; Vasil’ev et al. 1987; Vasil’ev 2003a). Numbers from these assemblages demonstrate no preference for the different taxa procured. From these data it appears that in Siberia early modern humans were procuring animals when they encountered them and did not use a systematic hunting strategy (Goebel 2004a). EUP Features Hearth features were found at the sites of Kara Bom, Ust’ Karakol, Makarovo-4, Varvarina Gora, Tolbaga, Kamenka, and Sannyi Mys. Those features found at Kara Bom, Ust’ Karakol and Makarovo-4 were unprepared, but consisted of discrete concentrations of ash, charcoal, charred sediment, and bone and lithic debris (Aksenov 1989; Derevianko et al. 2003; Goebel et al. 1993), whereas sites in the Transbaikal had stonelined hearths (Konstantinov 1994; Lbvoa 2000; Okladnikov 1971). Storage pits were found at Varvarina Gora, Tolbaga, Kamenka, and Sannyi Mys. Dwellings were identified by oval-to-circular distributions of large cobbles, boulders, and stone slabs encompassing storage pits and central hearth features (Konstantinov 1994; Lbova 1992, 1996, 2000, 2005; Okladnikv 1971; Okladnikov and Kirillov 1980). An additional dwelling was reportedly found at Malaia Syia; however, no details were ever published (Vasil’ev 2003b). EUP Summary EUP sites, representing the initial pulse of modern humans into Siberia, are distributed from the Ob’ River to the Transbaikal and as far north as the uppermost reaches of the Lena River immediately west of Lake Baikal. Dating of EUP sites has been highly problematic since most are situated in complicated colluvial-depositional settings so there is a good chance of mixture of multiple archaeological components. Additionally, most dates from these sites were obtained on bones by conventional 14C methods. Because conventionally dated samples were not properly pretreated so bone apatite was dated and some of these were likely pooled samples, they should not be trusted unless they provide ages statistically the same as AMS dates on bone collagen. Sites with clean archaeological assemblages, found in understandable stratigraphic contexts, and associated with chronometric dates suggest modern humans were first in the Altai foothills region, and perhaps south-central Siberia, by about 50,000 cal yr BP. Certainly they were to Makarovo-4 before 45,000 years ago and possibly to the Transbaikal as early as 42,000– 41,000 years ago. Available chronological data indicate that the EUP may have lasted until about 33,000 years ago; however, the younger ages are somewhat suspect owing to their potential contamination by recent or modern carbon. Overall, EUP sites date to the middle part of MIS-3, climatically a period of global warmth. Though the age ranges for several sites are quite large, encompassing both cold and warm

69 intervals of MIS-3, sites with tighter dates seem to group into two periods. For instance, Kara Bom, Makarovo-4, and perhaps Varvarina Gora date to an early warm interval prior to the early stade of MIS-3 (50,000–45,000 years ago), while others such as Masterov Kliuch’ and Tolbaga seem to date to the warm Malokheta interstade (40,000–35,000 years ago). The archaeological record suggests these initial settlers were making sophisticated blade-based technologies. Projectiles were tipped with unifacial stone points, a variety of food-processing and clothing-manufacturing implements are reflected in the lithic and osseous industries, and presence of nonutilitarian objects at many sites indicates people were adorning themselves with jewelry items. Interestingly, the use of personal ornaments is known from across the EUP world; however, the presence of needles is known only to the northern contexts of Siberia and Eastern Europe, perhaps reflecting the need for manufacturing warm clothing in these northern environments (Hoffecker 2005). Substantial dwellings, storage pits, and highly varied faunal assemblages indicate people were living at these sites for relatively long periods of time, perhaps on a seasonal basis.

Middle Upper Paleolithic

Across western Eurasia, Upper Paleolithic sites dating between 30,000 and 20,000 14C (~34,000–24,000 cal) yr BP and containing elaborate burials, Venus figurines, and small bladelet tools are termed Gravettian (Roebroeks et al. 2000). In Siberia this phase is commonly called the MUP (Vasil’ev 1992, 2000). Many archaeologists have referred to MUP or groups of MUP sites as the Mal’ta Complex or Mal’ta Culture, named for the famous Mal’ta site (Derevianko 1998; Lisitsyn 2000; Okladnikov 1968). Siberian MUP sites are typically distributed wider than before, there are many more, and their assemblages are known for bladelet production, Venus figurines and other art, and impressive dwelling features. MUP Distribution and Chronology MUP sites have the same basic west-east distribution as the EUP. In fact, a few of the sites discussed above also contain later MUP occupation layers (e.g., Ust’ Karakol, Malaia Syia, and Kamenka). As with the EUP, most are located south of 55˚N; however, a handful (e.g., Nepa-1, Alekseevsk, Ust’ Kova, Igeteiskii Log, Achinsk, Kurtak-4, Kurtak-5, Kashtanka-1, and Novoselovo-13) were found farther north, between 55˚ N and 60˚ N, a range extending more than 500 km north of that of the EUP. Additionally, the Yana RHS site (Yana), was found another 1200 km north at ~71˚ N along the lower Yana River, only 150 km upstream from where the river flows into the Arctic Ocean (Pitulko et al. 2004, this volume). Unlike the record for the EUP, there are several dozen reliable MUP cultural occupations reported for Siberia. Since I am limited for space and there are so many more post-dating the EUP, the rest of my review will highlight not all, but only key sites from MUP and late Upper Paleolithic (LUP) contexts. At least three MUP sites have occupation layers that date from about 30,000–26,000 14C (35,000–30,500 cal) yr BP, dur-

70 ing the warm Lipovo-Novoselovo interstade. At Yana three spatially discrete localities with MUP archaeological remains were found and are called Yana B, Northern Point, and TUMS 1. Twenty-two samples of identified bones, wood charcoal from a hearth feature, and plant remains provided AMS dates all in good agreement (Pitulko et al. this volume; Pitulko and Pavlova 2010). Dates range from 28,570 ± 300 (Beta-191322) to 26,680 ± 160 (Beta-191334) 14C (34,100–31,000 cal) yr BP. Two bone dates from Nepa-1 of 33,100 ± 1500 (AA-27382) and 26,065 ± 300 (AA-8885) 14C (41,100–30,300 cal) yr BP indicate an age at least as old as Yana RHS (Goebel 2004b). Two hearth-charcoal samples from Ust’ Karakol produced three dates in good agreement, ranging from 27,020 ± 435 (SOAN-3356) 14C to 26,305 ± 280 (SOAN-3261) 14C (32,300– 30,500 cal) yr BP (Derevianko et al. 1998, 2003). All three of these sites have occupation layers in good agreement with each other, except for the old date from Nepa-1. Very little work has been undertaken and reported at this site (Goebel 2004b). It could be that there were two occupation events, one EUP followed by a later MUP occupation. Or perhaps the younger date is aberrant because it was an early AMS date run on bone, probably without pretreatment levels used on more recent bone samples. The other reasonably well dated MUP cultural occupations number at least 16, are distributed south of 60˚ N, and fall within the age range of about 26,000–20,000 14C (30,500–23,500 cal) yr BP. Many of these date to the last 3000 years of this period, corresponding to the MIS-3 to MIS-2 transition and increased cooling with the gradual onset of the last glacial maximum (LGM) and include occupations at Mal’ta, Buret’, Kunalei, Ui-1, Novoselovo-13, and Kashtanka-1, to name a few (Graf 2009; Medvedev et al. 1996; Goebel et al. 2000a). MUP Technologies MUP lithic assemblages are blade-based and flake-based technologies. Raw materials are typically fine-grained, high-quality stones, including chert, siltstone, meta-siltstone, quartzite, argillite, and mudstone. Fine-grained igneous stones such as basalt and andesite were also occasionally procured as well as coarser stones such as quartz, granite, gabbro, diorite, tuff and sandstone (Abramova et al. 1991; Buvit et al. 2011; Derevianko et al. 2003; Drozdov et al. 1990; Graf 2010; Lisitsyn 2000; Medvedev 1998b; Pitulko et al. this volume; Terry et al. 2009; Vasil’ev 1996, 2000). Most raw materials were local, found in nearby stream alluvium, as evidenced by the common presence of alluvial-cobble cortex on debitage and tool blanks; however, others were obviously nonlocal, some being procured from very distant sources (Buvit et al. 2011; Graf 2010; Pitulko et al. this volume). Primary reduction strategies were often split between production of blades, bladelets, and flakes as tool blanks (Graf 2010). Blade cores range in form from flat-faced to subprismatic. They also vary in size from large to quite small, so their detached tool blanks range in size from blade to bladelet, depending either on the stage of reduction when discarded or the size of the initial raw-material package (i.e.,

Graf cobble versus pebble) (Graf 2008, 2010; Terry et al. 2009). Multidirectional, bidirectional, and unidirectional flake cores evidence systematic removal of flakes from one or more fronts. Occasionally bifacial cores were produced (Graf 2010). Secondary-reduction activities largely centered on manufacturing unifacial, bifacial, and burin tools on blades, bladelets, and flakes. Bifacial tools are present in many site assemblages; however, their frequency within assemblages is typically low (i.e., in Enisei river assemblages they make up  20 cm) rodtype points such as those from Yana RHS, Mal’ta, Buret’ and Igeteiskii Log (for examples see Pitulko et al. this volume). Long ivory spear shafts were also found at Yana. Other utilitarian implements include mostly retouchers/billets, awls, and needles (Abramova et al. 1991; Kirillov and Derevianko 1998; Lisitsyn 2000; Medvedev 1998b; Pitulko et al. this volume; Vasil’ev 2000). Nonutilitarian, art forms are common and quite spectacular in the MUP. These include mostly carved mammothivory pieces that can be divided into personal adornment and symbolic “mobile” art. Personal adornment pieces are numerous; found in several MUP assemblages such as Shestakovo, Achinsk, Sabanikha, Kurtak-4, Ust’ Kova, Mal’ta, Buret’, Sokhatino-4 and Yana; and include undecorated and decorated beads, drop pendants, and flat-form rectangular and disk-shaped pendants made on ivory, and tooth pendants made on fox and cervid canines and incisors, respectively. Additionally, stone beads and pendants have also been found at Mal’ta, Yana, and Kurtak-5. Mobile art pieces include engraved ivory plaques or badges (Achinsk, Mal’ta), enigmatic rod-shaped pieces (Achinsk), zoomorphic figurines such as enigmatic “beasts” resembling the outlines of mammoths, bison, or bears (Ust’ Kova, Mal’ta, Yana RHS), swans or other birds (Mal’ta, Buret’), and anthropomorphic forms called Venus figurines (Mal’ta and Buret’). These Venus figurines date to the same time as most found in western Eurasia; however, unlike Western versions the female form on Siberian pieces is carved in 2D instead of 3D and full-body winter clothing with hoods is also carved on these pieces (Abramova 1995; Drozdov et al. 1990; Kirillov and Derevianko 1998; Lisitsyn 2000; Medvedev 1998a; Pitulko et al. 2012, this volume; Vasil’ev 2000).

Siberian Odyssey MUP Fauna Kitchen lists from the MUP tell us that a wide variety of fauna were utilized, from large to small taxa (e.g., mammoth, woolly rhinoceros, horse, steppe bison, auroch, Irish elk, Argali sheep, Siberian mountain goat, Saiga antelope, red deer, roe deer, reindeer, Arctic fox, red fox, and hare). From these faunal lists, one major pattern emerges that is worth noting here. Unlike the EUP, not all taxa are represented in all assemblages of the MUP. Investigating further with MNI data, Ui-1 and Kashtanka-1 assemblages indicate a focus on taxa specific to these locations. The assemblage from the upland site of Ui-1 contains mostly upland taxa (Siberian mountain goat and Argali sheep), and the assemblage from the Kashtanka-1 site, situated in a more lowland/plain location, is dominated by reindeer. Feasibly these two sites represent special-task locations where hunters extracted local ungulate resources, perhaps during the rut season. Numerous other sites without MNI data but with short faunal lists also hint at being shortterm, special-task locations (Lisitsyn 2000; Vasil’ev 2003a). In contrast to this pattern, the faunal assemblage from Mal’ta is highly varied with at least 13 taxa represented, but no one type of animal is present in high frequencies, except for reindeer in which counts are mostly of antler. A similar pattern is also true for Yana, where high numbers of varied taxa are present (Pitulko et al. this volume). Based on faunal data, Mal’ta and perhaps Yana may have been residential sites or base camps. Faunal assemblages from other sites such as Buret’, Kamenka, Sabanikha, and Kurtak-4 also hint at this pattern due to their wide variety of taxa (Ermolova 1978; Lisitsyn 2000; Vasil’ev 2000, 2003a). MUP Features Hearth features abound in sites of the MUP, in which both prepared and unprepared hearths were discovered. Some sites such as Ui-1, Kashtanka-1, Novoselovo-13, Kurtak-4, and Kunalei have just a few hearths associated with lithic and bone scatters, signaling them as sites used as shortterm, special-task locations. Dwellings have been proposed at several sites. Some were possibly substantial, consisting of centrally located fireplaces and storage pits surrounded by boulder and stone-slab construction materials (Achinsk, Mal’ta, Buret’, and Chitkan [Konstantinov 1994; Larichev et al. 1988; Medvedev 1998b]). Though no clear dwelling or hearth features were observed at Priiskovoe, the relatively tight distribution of several cobbles (24 m2) coupled with heavy concentration of lithic debris has led some to interpret this combination as a dwelling (Buvit et al. 2011; Konstantinov 1994). At Yana (Northern Point Locality) two linear alignments of hearth features may represent at least two “lightly-framed” dwelling structures (Pitulko et al. this volume). A slab-lined hearth feature at Ui-1, coupled with discrete distribution of lithic materials found around it, led Vasil’ev (1996, 2003b) to hypothesize that a possible light, above-ground structure once stood at this seasonal hunting camp. Finally, a double burial feature, containing the remains of two children under the age of four, was found at Mal’ta (Alekseev 1998). Together

71 with the very late Ushki Lake burial, these are the only burial features reported for the Siberian Paleolithic; all other human remains come from isolated finds and not associated with burial features (Akimova et al. 2005, 2010a; Gerasimova at al. 2007; Kuzmin et al. 2009). The Mal’ta children were interred with a wide array of grave goods, including a beaded necklace with pendants, osseous projectile point, unifacial stone tools, bone bracelet, and an ivory plaque and bird figurine (Okladnikov 1940). MUP Summary Siberian MUP sites have a wide distribution, forming a large triangle from southwestern Siberia (~51˚ N/85˚ E), north to western Beringia (~71˚ N/135˚ E), and back south to the Transbaikal in southeastern Siberia (~52˚ N/113˚ E). Interestingly, the oldest dated occupations, which emerge at roughly the same time (~34,000–31,000 years ago), are found along the southwest-to-northeast side of this scalene triangular distribution with Yana to the northeast, Nepa-1 in the middle, and Ust’ Karakol to the southeast. The current record lacks other MUP sites between Yana and Nepa-1, and for the 7500 years following this initial incursion north, MUP occupations remains south and east of the Yana-Nepa-Ust’ Karakol line. MUP artifact assemblages are based on flake-core and blade-core reduction. Strikingly, most sites indicate reliance on informal flake over blade production, and those sites dating to the final 3500 years of the MUP (~27,000–23,500 cal BP) evidence strong use of bladelets, as a result of economizing raw material or of consistently selecting small rawmaterial packages (i.e., small alluvial cobbles). Most assemblages have an elaborate osseous industry. Most projectiles are made on bone, antler, or ivory, except for a few bifacial projectiles found at Ust’ Kova and Derbina. Sites with small artifact assemblages tend to be short-term, logistical campsites, while sites with hordes of interesting decorative and artistic pieces reflect longer-term residential sites. Faunalassemblage compositions and types of domestic features also support this interpretation. The record suggests MUP huntergatherers were logistically mobile, perhaps seasonally revisiting residential bases and associated spike camps.

Late Upper Paleolithic

Across Siberia there are hundreds of sites reportedly containing LUP cultural occupations (Abramova et al. 1991; Derevianko 1998; Goebel 2002a). As can be expected, variations in LUP assemblages across regions exist and have led to development of different, regional archaeological “cultures” of the period. Perhaps most notable to American scholars are the Afontova, Kokorevo (Vasil’ev 1992), and Diuktai cultural traditions (Yi and Clark 1985). Some have abandoned such notions (Graf 2011; Pitulko et al. this volume; Vasil’ev 1992, 1996, 2000 but see 2011), but many still adhere to them (Akimova et al. 2005; Derevianko 2010; Lisitsyn 2000). In this chapter the LUP is treated as above, a chronological phase that has patterns of technological, subsistence, and overall landscape organization characteristic for the phase.

72 LUP Distribution and Chronology LUP sites are distributed east from the Ob’ to the Pacific, north from the Russian border to Beringia, and found in places previously uninhabited, specifically the Russian Far East and western Beringia east of the Yana River (Sakhalin Island, Kamchatka, and Chukotka) (Slobodin 2011; Vasilevskii 2008). Northern Siberia, north of about 60˚ N and west of the Lena River, is devoid of LUP sites, but seems to have been inhabited by the middle Holocene (Pitulko and Pavlova 2010). As you will see below, sites become younger from south to north. Ogon’ki-5, located on Sakhalin Island, may represent the earliest LUP site in Russian Northeast Asia. During the early half of the last glacial cycle (including LGM), Sakhalin was connected to both Hokkaido, Japan, and mainland Siberia via land bridges (Ono and Machida 1987). Four AMS dates on hearth charcoal are in good agreement and indicate an age range of 19,440 ± 140 (Beta-115987) to 18,920 ± 150 (AA-25343) 14C (23,600–22,200 cal) yr BP. The next youngest is Studenoe-2 in the Transbaikal (Buvit and Terry 2011; Goebel et al. 2000b; Goebel 2002a). Once thought to date to about 17,500 14C yr BP (Goebel 2002a), detailed geoarchaeological research coupled with new dates led Buvit and Terry (2011) to argue the occupation happened a little earlier, from 18,020 ± 230 (AA67845) to 17,550 ± 90 (AA-37964) 14C yr BP or 22,200–20,500 cal yr BP. The selected age range is based on four congruent dates on wood charcoal from four hearth features associated with a dwelling complex (Buvit and Terry 2011; Konstantinov 2001; Goebel et al. 2000). The next youngest site is Nizhnii Idzhir-1 in the Enisei valley, dating to 17,200 ± 70 (LE-1984) 14 C (21,100–20,100 cal) yr BP (Astakhov 2008). Though this is based on but one 14C date, the date comes from a clean context where the hearth feature is surrounded by a dense artifact scatter averaging 5 cm thick across the excavation (25 m2) (Astakhov 2008). Overwhelmingly, most other LUP occupations found in good geoarchaeological settings are situated south of 56˚ N and postdate these ages. A cluster of early LUP sites approaching western Beringia and situated along the Aldan River have long been discussed in the context of the peopling of Beringia and the Americas (Chard 1974; Goebel 1999; Holmes 2011; Mochanov 1977; Powers 1973; Yi and Clark 1985). The most reliable dates come from Diuktai Cave. The earliest LUP layer dates to about 13,200 ± 250 (GIN-405) to 13,090 ± 70 (LE-784) 14C (16,800–15,100 cal) yr BP. Though critical of the other early open-air Aldan LUP sites (e.g., Ikhine and Ust’ Mil’), Yi and Clark (1985) tentatively accept dates and stratigraphic context of artifacts from Verkhne Troitskaya, giving an average date of 16,615 14C yr BP for the occupation. The problem is that this average was calculated on four dates that range from 18,300 ± 180 (LE-905) to 14,530 ± 160 (LE-864) 14C yr BP (Mochanov 1977; Tseitlin 1979) and do not overlap, even at 2σ. The dating of the Verkhne Troitskaia artifacts is just as problematic as the dating of Ikhine and Ust’ Mil’ cultural materials because the samples were pieces of wood found dispersed in 80 cm of alluvial sediment over-

Graf lying the artifacts. Artifacts were found dispersed below in a 100-cm-thick zone of fluvial sediments that, from Mochanov’s (1977:59) description, represents an active stream bed. It is very difficult to accept that artifacts and dated samples were found in primary depositional contexts. Unfortunately, there are no characteristic LUP sites from dated contexts in northeastern Siberia, east of Diuktai Cave, that clearly predate the Clovis era in the Americas (Slobodin 2011; Waters and Stafford 2007). LUP Technologies LUP assemblages are flake-based, microblade-based, and blade-based technologies. Raw materials are typically of the same types found in MUP assemblages, chert, siltstone, meta-siltstone, quartzite, argillite, mudstone, basalt, andesite, quartz, granite, gabbro, diorite, tuff, and sandstone (Abramova 1989; Abramova et al. 1991; Buvit and Terry 2011; Graf 2010; Lisitsyn 2000; Vasil’ev 1996; Terry et al. 2009). In the Enisei, most raw materials were local from nearby alluvium with a few nonlocal stones used (Graf 2010). Data from the Transbaikal region, however, suggest most raw materials were procured from nonlocal sources (Terry et al. 2009). Primary reduction strategies led to production of flakes, blades, and microblades (Graf 2010). Blade cores are typically of the large, flat-faced variety. Flake cores are multi­directional, unidirectional, bidirectional, bifacial, or bipolar. Microblade cores are either wedge-shaped (manufactured on bifaces) or tortsovyi cores (manufactured on flakes or small cobbles and pebbles) (Abramova et al. 1991; Gómez Coutouly 2011; Graf 2008, 2010; Terry 2010; Terry et al. 2009). Secondary-reduction activities included manufacturing uni­facial, bifacial, and burin tools on blades, bladelets, and flakes (including bifacial thinning flakes). Typical LUP lithic-tool types include side scrapers, end scrapers, retouched flakes and blades, bifaces, gravers, burins, bifaces, and to a lesser extent notches, denticulates, and wedges (Abramova et al. 1991; Graf 2008, 2010; Lisitsyn 2000; Terry et al. 2009). Bifaces as tools are present in many site assemblages (Graf 2010). LUP sites in the Aldan river valley typically contain finished lanceolate points, some with over-face flake scars (Mochanov 1977). Osseous technologies centered on producing tools such as bone, antler, and ivory points, awls, needles, retouchers, and shaft straighteners (i.e., “baton de commandement”) (Abramova 1979a, 1979b; Abramova et al. 1991; Akimova et al. 2005; Gening and Petrin 1985). Projectile points are slotted along their lateral margins, probably with burins or gravers from associated toolkits. Microblade midsections were inserted into these slots to produce very sharp, lethal spear tips. A few examples of points with microblade inserts still in place exist across Siberia (e.g., Chernoozer’e and Listvenka (Akimova et al. 2005; Gening and Petrin 1985). We know these points were used to hunt because two examples of microblade-composite projectiles were found embedded in mammoth and bison bones at Lugovskoe (western Siberia) and Kokorevo-1 (Enisei), respectively (Abramova 1979b; Zenin et al. 2006). Despite the fact that many sites with well-

Siberian Odyssey

73

preserved faunal remains also have osseous artifacts in their assemblages, nonutilitarian examples are rare. When present, they typically consist of bone beads and pendants sometimes with abstract engravings (e.g., Chernoozer’e, Kokorevo, Afontova Gora, Tashtyk, Ui, Maina, Listvenka, and Studenoe-2) (Abramova 1979a, 1979b, Abramova et al. 1991, Akimova et al. 2005; Astakhov 2008; Konstantinov 2001). Finally, only one uncontested piece of non-personal-adornment art comes from the LUP, an anthropomorphic clay statuette from Maina (Vasil’ev 1996).

(Akimova et al. 2005; Konstantinov 1994, 2001; Vasil’ev 1996), providing the best evidence of dwellings during the LUP. Some have proposed additional dwelling features at Chernoozer’e, Kokorevo-1, Golubaia-1, Nizhnii Idzhir’, Ui-2 and Maina, but this is based on arrangements of chipped-stone lithic and faunal debris around hearths (Abramova 1979b; Astakhov 2008; Gening and Petrin 1985; Vasil’ev 1996, 2003b). These distributions could simply be explained as knapping stations or animal-processing areas, so their interpretation as dwelling features remains equivocal (Vasil’ev 2003b).

LUP Fauna Faunal remains from LUP assemblages are quite numerous (Vasil’ev 2003a). Again, however, analyses of these data are limited mostly to kitchen lists; however, a few sites do contain NISP and MNI data. At least 29 taxa are represented in LUP assemblages on whole. Using Vasil’ev’s (2003a) massive faunal lists for the Paleolithic, there are several patterns to point out. Of 256 LUP occupation assemblages, the leaders are clearly bison and reindeer. These two are present in 141 and 132 assemblages, respectively with red deer, caprids, mammoth, and horse coming in close behind them. There is some regional variation, with bison and caprids selected most often in southwestern Siberia; bison and reindeer in the Enisei; horse and reindeer in the Angara; red deer and reindeer in the Transbaikal; and horse, bison and reindeer in western Beringia. Another interesting pattern is that wolf/ dog is present in at least 37 assemblages, whereas during the MUP only 5 occupations have wolf and during the EUP they are present almost exclusively at cave sites. Perhaps this pattern of many wolves/dogs in LUP assemblages reflects their use as beasts of burden during the Late Glacial. Assemblages with NISP and MNI data indicate that a single species dominates any given assemblage. These taxa were bison, caprids, reindeer, hare, and red deer. Together, faunal data demonstrate a focused hunting strategy was in play during the LUP.

LUP Summary Siberian LUP sites outnumber those that came before. They adhere to the same basic distribution of the MUP; however they extend into reaches of the Russian Far East that were previously unoccupied by modern humans. The earliest appearance of microblade-bearing technologies (a diagnostic feature of LUP assemblages) appears first on the late-Pleistocene-aged Hokkaido/Sakhalin Peninsula, next to the west in the Transbaikal, and then farther west in south-central and central Siberia soon afterward. By about 20,000–19,000 cal BP, LUP sites were springing up as far north as 56˚ N, and in the Aldan river valley by 16,000 cal BP. Investigations of the frequencies of 14C-dated LUP occupations through the Late Glacial illustrate a trend of gradual increase through time with small population spikes during warm intervals (i.e., Bølling and Allerød interstades) and population nadirs during intervening cold periods (Oldest Dryas and Older Dryas interstades) (Goebel 2002a; Graf 2005, 2009). LUP artifact assemblages are internally consistent with lithic industries based on flake-core, microblade-core, and blade-core reduction with many assemblages containing more flake than blade production. Microblade-osseous composite-point production was prevalent throughout the LUP, evidenced by microblade cores, microblades, and slotted points. Though working of osseous materials centered on production of utilitarian implements, production of personal adornment did not cease, but production of other art forms did. The faunal record indicates a focus on specific, mostly large-game, taxa at most sites. This faunal pattern coupled with the near lack of any substantial dwellings suggests people of the LUP were on the move, perhaps frequently traveling between sites. Provisioning and tool-richness data from the Enisei demonstrate that individuals within LUP communities were being provisioned and most LUP sites reflect residential bases (Graf 2010, 2011). Combined archaeological data indicate LUP hunter-gatherer groups were highly mobile, frequently moving across the landscape. Interestingly, two sites do not fit this overall LUP pattern and should be mentioned because they were found in western Beringia and date to the Allerød interstade just before and during the Clovis era. The sites of Berelekh and Ushki Lake, located along the lower Indigirka River in northwestern Beringia and Kamchatka River on the Kamchatka Peninsula, respectively, have biface-based and flake-based assemblages with bifacial teardrop-shaped (Berelekh) and stemmed (Ushki)

LUP Features LUP hearth features are fairly common across Siberia; however, these typically appear as isolated, unprepared hearths associated with localized lithic scatters (Akimova et al. 2005; Astakhov 1999, 2008; Gening and Petrin 1985; Konstantinov 1994, 2001; Lisitsyn 2000). In the Enisei, 17 of 23 14C-dated LUP sites contain at least one hearth feature, yet 85 distinct cultural layers are documented at these sites (Graf 2008). Certainly, preservation issues and excavation techniques can hinder observation and documentation of hearth features, but this low proportion of hearths-to-occupation numbers indicates some sites were only visited for very short periods of time. Unlike western Eurasia, LUP dwellings are fairly rare in Siberia. The sites of Studenoe-1, Studenoe-2, Ust’ Menza-1, Ust’ Menza-2, Ust’ Menza-3, and Kosaia Shivera in the Transbaikal and Listvenka and Ui-2 in the Enisei have circular or semi-circular alignments of boulders and stone slabs, containing lithic debris and hearth features within these alignments

74 projectile points and preforms, unifacial tools on flakes and blades, and no clear association with microblades; however, the teardrop-shaped point from Berelekh was found in a surface context (Dikov 1968; Mochanov 1977; Pitulko 2011; Vereshchagin 1977). Dwelling and burial features were identified at Ushki Lake. Dates from these cultural occupations suggest an age of about 11,500–11,000 14C (13,700–12,700 cal) yr BP (Goebel et al. 2003, 2010; Pitulko 2011). Overall, data from these two sites do not fit the typical LUP pattern, but these sites are very late, dating to the terminal Pleistocene, so it is difficult to compare them with other LUP occupations of late-glacial age.

Discussion Were Eup Hunter-Gatherers Poised to People the North? Modern humans first dispersed to southern Siberia by about 50,000 years ago. These initial settlers were making bladebased Upper Paleolithic technologies, similar to those used by contemporary early modern humans in other regions of the Old World, and they were opportunistically hunting a wide array of fauna. Their technologies were sophisticated. They were making prepared lithic technologies, and presence of awls and needles suggests they were manufacturing and mending clothing and other items made of animals skins. It would appear that they were technologically equipped to push into northern landscapes. They were hunting a variety of northern fauna, and also fauna found in more moderate climates today, and they seem to have maintained their population in southern Siberia during a period of relative warmth. Though land-use strategy and provisioning data are scanty at best, they do suggest people were tied to local resources, not ranging long distances, and using a logistical mobility strategy. Early modern humans in southern Siberia seemed to have been just learning northern landscapes, populations were probably not large, and they were likely not poised to disperse to Beringia and the Americas. Why Did the MUP Expansion Happen and What Does It Mean for Human Dispersals to the Americas? Clearly the presence of Yana in northwestern Beringia at 32,000 years ago is quite magnificent. Material remains found so early so far north is truly eye-opening and indicates that by the MUP people were capable of expanding north into the Arctic and Beringia. MUP knappers expanded their toolkits to include a wider array of implements so they were better prepared. For instance, people used a range of techniques to produce projectile technologies, not just osseous points, but also bifacial points were now being made. MUP tool makers were selecting from a greater list of raw-material types, probably because they had grown accustomed to the sources in their territorial ranges and had learned where quality raw materials were located, perhaps even beyond typical territory limits (e.g., amber found in the Yana assemblage came from 600 km away [Pitulko et al. this volume]). There seems to have been greater vari-

Graf ability in producing osseous implements. The wider range and increased numbers of awls and needles may signal increased clothing manufacture and repair. The relative explosion in symbolic, “mobile” art across Siberia suggests maintenance of social networks. Faunal data indicate varied site functions among MUP sites, with large residential bases and small, short-term hunting camps. Largely the taxa in these assemblages are cold- and dry-adapted species, many herd animals that made up the mammoth-steppe of Eurasia (Guthrie 1990). These faunal data fit climatically since most MUP sites date to the 4000 years immediately preceding the LGM and the oldest sites (Yana and Nepa-1) are in northern settings (Goebel 1999). Settlement data indicate a wellestablished logistical system with large, long-term residential bases and resource-extraction sites, perhaps revisited annually or biannually on seasonal bases. During the height of the warm Lipovo-Novoselovo interstade, MUP populations expanded north to such places as the Nizhnaia Tunguska river valley in central Siberia and Indigirka river valley of western Beringia, where they thrived, hunting mammoth-steppe fauna, not just for food procurement, but also for toolmaking and creating mobile art that was visual and symbolic for maintaining distant social networks (Meltzer 2009; also see Flegenheimer et al. this volume). As temperatures began to fall and climate declined between 28,000 and 23,000 cal BP, the MUP range shrank south, reflected by the multitude of later sites scattered across the south and the abandoning of Yana by 31,000 years ago. As the record stands now, it is difficult to expect the MUP to have contributed in a direct way to dispersal into the Americas from a Beringian springboard, unless such a dispersal event took place between 34,000 and 31,000 cal BP. At this time, the eastern Beringian record simply does not support this (Holmes 2011; Potter et al. this volume). Where Did They Go? Siberia during the LGM was a harsh place, with countless paleoecological records including ice-wedges, cryoturbated loams, and various pollen and faunal records evidencing this phenomenon (Elias and Brigham-Grette 2007; Miller et al. 2010). There has been much debate centering on whether or not people were in Siberia during the LGM (Davis and Ranov 1999; Dolukhanov et al. 2002; Goebel 1999, 2002a; Graf 2005, 2008, 2009, 2010; Kuzmin and Keates 2005; Kuzmin 2008). In this debate, I have fallen clearly on the side of no. Data from Ogon’ki 5 coupled with recent work at Kashiwadai in Hokkaido, Japan (Izuho and Takahashi 2005; Nakazawa et al. 2005) suggest there were LGM refugia for humans just outside of Siberia proper, in the Russian Far East and northern Japan. The overwhelming pattern in interior Siberia, however, still suggests abandonment. Even if some populations found refuge in isolated, more productive areas during the LGM, the archaeological record does not support continued occupation of northern Siberia and Beringia at this time, despite recent modeling by geneticists (Mulligan and Kitchen this volume).

Siberian Odyssey Dispersal Back North The Siberian LUP may have begun in the Russian Far East during the LGM. One of the hallmarks of the LUP is microblade technology. This technological strategy probably originated in an area outside of southern Siberia because the best evidence to date for LGM-aged sites with microblade technologies in Northeast Asia comes from the Kashiwadai site and Ogon’ki-5 as mentioned above. Perhaps this region provided an LGM refugium for both people and mammals alike. Remains of LGM-aged gregarious herd fauna were found on the Hokkaido/Sakhalin Peninsula (Izuho and Takahashi 2005), signaling a productive mammoth-steppe-like landscape. Perhaps hunting concentrated large-herd fauna led to development of microblade-osseous composite projectile point technology. Determining the origins of Siberian microblade technology is an interesting problem because it seems to be intricately tied to when and where people were during the LGM and how parts of Siberia were repopulated via a late-glacial dispersal north. After the occupation of Ogon’ki and Studenoe, LUP sites begin emerging quickly to the west and eventually north of Lake Baikal, making it to western Beringia by 16,000 years ago and then to eastern Beringia by 14,500 years ago (Holmes 2011). The organization of technologies and subsistence indicates these people were highly mobile as a group and therefore could have quickly dispersed farther west, deeper into southern Siberia and north to Beringia. Why would they choose high residential mobility when their MUP predecessors did not? Perhaps it became an economical way to pursue mobile herd fauna, which were becoming more dispersed during the late-glacial demise of the mammoth steppe, especially if human populations were increasing between 20,000 and 17,000 cal BP (Goebel 2002a; Graf 2005). Could group mobility have been facilitated with the help of dogs carrying gear or even pulling sleds? Siberian LUP faunal assemblages, archaeological evidence of domesticated dog in European Russia at 17,000 cal BP (Sablin and Khlopachev 2003), and recent dna studies that contend East Asia was the homeland of dog domestication during the Late Glacial (Ding et al. 2012; Savolainen et al. 2002) may all hint at this as a possibility.

Concluding Remarks

The Siberian Upper Paleolithic record suggests at least two dispersal scenarios to Beringia and therefore provides two possible time frames for Beringians to disperse to the New World. Before Yana was discovered, geneticists predicted a pre-LGM dispersal to Beringia (Bonatto and Salzano 1997). Clearly people of the MUP dispersed to western Beringia at about 34,000–31,000 years ago and were therefore poised to head to Alaska at this time. Was the Bering Land Bridge open at this time? There are indications that the onset of the LGM began in the Arctic as early as 32,000–31,000 years ago so it was likely open (Elias and Brigham-Grette 2007; Elias and Crocker 2008). The land bridge may have been open, but by 31,000 years ago conditions may have become too harsh

75 to sustain human population this far north. After Yana, there is no trace of humans in or near Beringia until people camp at Diuktai Cave at least 14,000 years later. If MUP huntergatherers contributed directly to initial dispersals to America, then this dispersal event must have taken place before 31,000 years ago, very soon after Yana was reached. When reaching Alaska, however, it is highly likely that the northern end of the North American ice sheets had already coalesced. To date, we have no evidence of people in eastern Beringia before 14,500 years ago. Did people of the LUP disperse to the New World after the LGM? There is very good evidence of this position. LUP people had reached Diuktai Cave, the gateway to Beringia, no later than 16,000 years ago and perhaps as early as 17,000 years ago. Certainly, the “pre-Clovis” site of Swan Point provides excellent evidence of LUP people in eastern Beringia by 14,500 years ago. Geneticists tell us that the founding population of First Americans consisted of as many as 1000 people and was positioned in Beringia during the LGM (Kitchen et al. 2008; Mulligan and Kitchen this volume; Tamm et al. 2007). The archaeological record indicates a small population in western Beringia at 32,000 years ago, and paleoecologists think the LGM may have begun in Beringia soon afterward. Perhaps the climate was too cold and dry to maintain a recognizable human population through the LGM. The archaeological record indicates by 16,000 years ago people with LUP technologies were positioned in western Beringia to disperse east to the New World. Could the genetic clock be off? Is the archaeological record too coarse-grained to detect an LGM population? Either way, we still have a lot of work ahead to discover the answers.

Acknowledgments

Thanks to two reviewers for providing insightful comments and suggestions. I want to thank Russian colleagues who have provided access to collections and Russian literature. I also acknowledge the Center for the Study of the First Americans and Department of Anthropology, Texas A&M University for logistical support and the Arctic Social Sciences Program, National Science Foundation (awards #0525828 and #1019190) for supporting travel to Russia to collect data used in this chapter.

References Cited

Abramova, Z. A.  1979a  Paleolit Eniseia: Afontovskaia Kul’tura [Paleolithic of the Enisei: Afontova Culture]. Nauka, Leningrad. (In Russian) ———  1979b  Paleolit Eniseia: Kokorevskaia Kul’tura [Paleolithic of the Enisei: Kokorevo Culture]. Nauka, Leningrad. (In Russian) ———  1989  Paleolit Severnoi Azii [The Paleolithic of Northern Asia]. In Paleolit Kavkaza I Severnoi Azii, pp. 143–243. Nauka, Leningrad. (In Russian) ———  1995  L’art Paléolithique d’Europe Orientale et de Sibérie [Art of the Paleolithic of Eastern Europe and Siberia]. Jerome Millon, Grenoble. (In French) Abramova, Z. A., S. N. Astakhov, S. A. Vasil’ev, N. M. Ermalova, and N.

76

Graf

F. Lisitsyn  1991  Paleolit Eniseia [Paleolithic of the Enisei]. Nauka, Leningrad. (In Russian)

Pleistocene/Early Holocene Beringia, edited by T. Goebel and I. Buvit, pp. 47–57. Texas A&M University Press, College Station.

Akimova, E. V., O. A. Gorel’chenkova, I. V. Stasiuk, E. A. Tomilova, and L. A. Orlova  2003  Mestonakhozhdeniia Rannesartanskogo Vremeni Podtaezhnoi Zony Srednego Eniseia In Drevnosti Prieniseiskoi Sibiri: Vypusk 2. Sbornik Nauchykh Trudov, pp. 20–28. KGU, Krasnoiarsk.

Buvit, I., K. Terry, V. K. Kolosov, and M. V. Konstantinov  2011  The alluvial history and sedimentary record of the Priiskovoe site and its place in the Paleolithic prehistory of Siberia. Geoarchaeology: An International Journal 26(5):616–48.

Akimova, E. V., T. Higham, I. Stasyuk, A. Buzhilova, M. Dobrovolskaya, and M. Mednikova  2010a  A new direct radiocarbon AMS date for an Upper Palaeolithic human bone from Siberia. Archaeometry 52:1122–30.

Chard, C. S.  1974  Prehistory of Northeast Asia. University of Wisconsin Press, Madison.

———  2010b  The Late Paleolith study of the Derbina Bay (Krasnoyarsk Reservoir, Siberia). In Quaternary Stratigraphy and Paleontology of the Southern Russia: Connections between Europe, Africa and Asia, pp. 14–18. Abstract Volume of the 2010 Annual Meeting INQUASEQS, June 21–26, 2010, Rostov-on-Don, Russia.

Derenko, M. V., T. Grzybowski, B. A. Malyarchuk, J. Czarny, D. Miścicka-Śliwka, and I. A. Zakharov  2001  The Presence of Mitochondrial Haplogroup X in Altaians from South Siberia. American Journal of Human Genetics 69:237–41.

Akimova, E. V., N. I. Drozdov, S. A. Laukhin, V. P. Chekha, L. A. Orlova, V. G. Kol’tsova, A. F. San’ko, and E. G. Shpakova  2005  Paleolit Eniseia: Listvenka [The Paleolithic of the Enisei: Listvenka]. Institut Arkheologii i Etnografii Sibirskogo Otdeleniia Rossiiskoi Akademii Nauk, Krasnoiarsk. (In Russian) Aksenov, M. P.  1989  Paleoli i Mesolit Verkhnei Leny. Avtoreferat. [Paleolithic and Mesolithic of the Upper Lena River. Synopsis]. Nauka, Novosibirsk. Alekseev, V.  1998  The physical specificities of Paleolithic hominids in Siberia. In The Paleolithic of Siberia: New Discoveries and Interpretations, edited by A. P. Derevianko, pp. 329–35. Anikovich, M. V., A. A. Sinitsyn, J. F. Hoffecker, V. T. Holliday, V. V. Popov, S. N. Lisitsyn, S. L. Forman, G. M. Levkovskaya, G. A. Pospelova, I. E. Kuz’mina, N. D. Burova, P. Goldberg, R. I. Macphail, B. Giaccio, N. D. Praslov  2007  Early Upper Paleolithic in Eastern Europe and implications for the dispersal of modern humans. Science 315(5809):223–26. Aseev, I. V., and Iu. P. Kholiushkin  1985  Paleoliticheskie mestonakhozhdeniia v raione Khorinska [Paleolithic locations in the region of Khorinska] (BASSR). In Archaeologicheskie Issledovaniia v Raionakh Novostroek Sibiri, pp. 7–11. Nauka Novosibirsk. Astakhov, S. N.  1999  Paleolit Eniseia: Paleoliticheskie Stoianki Afontovoi Gore v G. Krasnoiarske [Paleolithic of the Enisei: Paleolithic Site of Afontova Gora in the City of Krasnoiarsk]. RAN, St. Petersburg. ——— 2008 Paleoliticheskie Pamiatniki Tuvy [Paleolithic Sites of the Tuva]. Nestor-Istoriia Publishers, St. Petersburg. (In Russian) Bar-Yosef, O.  2000  The Middle and Early Upper Paleolithic in Southwest Asia and Neighboring Regions. Peabody Museum of Archaeology and Ethnology Bulletin 8:107-156. Bazarov, D-D. B., M. V. Konstantinov, A. B. Imetkhenov, L. D. Bazarova, and V. V. Savinova  1982  Geologiia i Kul’tura Drevnikh Poselenii Zapadnogo Zabaikal’ia [Geology and culture of the ancient settlements of West Transbaikal]. Sibsirskoe Otdelenie. Nauka, Novosibirsk. (In Russian) Bonatto, S. L., F. M. Salzano  1997  A Single and Early Migration for the Peopling of the Americas by Mitochondrial dna Sequence Data. Proceedings of the National Academy of Sciences 94:1866–71. Bradley, B., and D. Stanford  2004  The North Atlantic ice-edge corridor: A possible Paleolithic route to the New World. World Archaeology 36(4):459–78. Buvit, I., and K. Terry  2011  Last Glacial Maximum human populations in the southwest Transbaikal, southern Siberia. In From the Yenisei to the Yukon: Interpreting Lithic Assemblage Variability in Late

Davis, R. S., and V. A. Ranov  1999  Recent work on the Paleolithic of central Asia. Evolutionary Anthropology 8:1–8.

Derevianko, A. P.  1998  The Paleolithic of Siberia: New Discoveries and Interpretations. University of Illinois Press, Urbana. ———  2010  Three scenarios: Of the Middle to Upper Paleolithic Transition. Archaeology, Ethnology, and Anthropology of Eurasia 38/3:2–32. Derevianko, A. P., and S. V. Markin  1990  Predvaritel’nye itogi izucheniia Must’e Altaia [Preliminary Results from Studies of the Mousterian in the Altai]. In Arkheologicheskie Etnografischeskie i Antropologicheskie Issledovaniia v Mongolii [Archaeological, Ethnographical and Anthropological Studies in Mongolia]. Derevianko, A. P., and E. P. Rybin  2005  The earliest representations of symbolic behavior by Paleolithic humans in the Altai Mountains. In The Middle to Upper Paleolithic Transitions in Eurasia: Hypotheses and Facts, edited by A. P. Derevianko, pp. 232–55. Institute of Archaeology and Ethnography Press, Novosibirsk. Derevianko, A. P., M. V. Shunkov, and P. V. Volkov  2008  A Paleolithic bracelet from Denisova Cave. Archaeology, Ethnology, and Anthropology. 34/2:13–25. Derevianko, A. P., M. V. Shyn’kov, A. K. Agadzhanian, G. F. Baryshnikov, E. M. Maleva, V. A. Ul’ianov, N. A. Kulik, A. V. Postnov, A. A. Anoikin 2003 Prirodnaia Sreda i Chelovek v Paleolite Gornogo Altaia: Usloviia Obitaniia v Okrestnostiikh Denisovoi Peshchery [Paleoenvironment and Paleolithic Occupation of the Altai Mountains]. Izdatel’stvo Instituta Arkheologii i Etnografii SO RAN, Novosibirsk. (In Russian) Derevianko, A. P., Iu. V. Grichan, M. I. Dergacheva, A. N. Zenin, S. A. Laukhin, G. M. Levkovskaia, A. M. Maloletko, S. V. Markin, V. I. Molodin, N. D. Ovodov, V. T. Petrin, and M. V. Shun’kov  1990  Arkheologiia i Paleoekologiia Paleolita Gornogo Altaia [Archaeology and Paleoecology of the Paleolithic of the Altai Mountains]. Nauka, Novosibirsk. Derevianko, A. P., A. K. Agadzhanian, G. F. Baryshnikov, M. I. Dergacheva, T. A. Dupal, E. M. Malaeva, S. V. Markin, V. I. Molodin, S. V. Nikolaev, L. A. Orlova, V. T. Petrin, A. V. Postnov, V. A. Ul’ianov, I. K. Fedeneva, I. V. Foronova, and M. V. Shun’kov 1998 Arkheologiia, Geologiia i Paleogeografiia Pleistotsena i Golotsena Gornogo Altaia [Archaeology, Geology, and Paleogeography of the Pleistocene and Holocene in the Altai Mountains]. IIAE, Novosibirsk. (In Russian) De Saint Pierre, M. C. M. Bravi, J. M. B. Motti, N. Fuku, M. Tanaka, E. Llop, S. L. Bonatto, and M. Moraga  2012  An alternative model for the early peopling of southern South Americas revealed by analyses of three mitochondrial dna haplogroups. PLOS ONE 7(9):e43486. Dikov, N. N.  1968  The discovery of the Palaeolithic in Kamchatka and the problem of the initial occupation of America. Arctic Anthropology 5(1):191–203. Ding, Z, M. Oskarsson, A. Ardalan, H. Angleby, L. Dahlgren, C. Teipeli,

Siberian Odyssey

77

E. Kirkness, P. Sabolainen, and T. P. Zhang  2012  Origins of domestic dog in southern East Asia is supported by analysis of Y-chromo­ some dna. Heredity 108(5):507–14.

pological Papers of the University of Alaska, New Series, Volume 2, Issue 1 2002, pp. 94–114. APUA, NS Department of Anthropology, University of Alaska, Fairbanks.

­ aitseva Dolukhanov, P. M., A. M. Shukurov, P. E. Tarasov, and G. I. Z 2002  Colonization of northern Eurasia by modern humans: Radiocarbon chronology and environment. Journal of Archaeological Science 29:593–606.

———  2004a  The early Upper Paleolithic of Siberia. In The Early Upper Paleolithic Beyond Western Europe, edited by J. Brantingham, S. L Kuhn, and K. W. Kerry, pp. 162–95. University of California Press, Berkeley.

Drozdov, N. I., V. P. Chekha, S. A. Laukhin, V. G. Kol’tsova, E. V. Akimova, A. V. Ermolaev, V. P. Leont’ev, S. A. Vasil’ev, A. F. Iamkikh, G. A. Demidenko, E. V. Aretm’ev, A. A. Bikulov, A. A. Bokarev, I. V. Foronova, and S. D. Sidoras  1990  Khrono-stratigrafiia Paleoliticheskikh Pamiatnikov Srednei Sibiri: Bassein R. Yenisei [Chronostratigraphy of the Paleolithic sites of middle Siberia: Yenisei River basin]. Nauka, Novosibirsk.

———  2004b  The search for a Clovis progenitor in subarctic Siberia. In Entering the Americas: Northeast Asia and Beringia Before the Last Glacial Maximum, edited by D. B. Madsen, pp. 311–56. University of Utah Press, Salt Lake City.

Elias, S. A., and J. Brigham-Grette  2007  Glaciations: Late Pleistocene events in Beringia. In Encyclopedia of Quaternary Science, pp. 1057–66. Elsevier, Amsterdam.

Goebel, T., and S. B. Slobodin  2010  New dates from Ushki-1, Kam­chatka, confirm 13,000 cal BP age for earliest Paleolithic occupation. Journal of Archaeological Science 37:2640–49.

Elias, S., and B. Crocker  2008  The Bering Land Bridge: A moisture barrier to the dispersal of steppe-tundra biota? Quaternary Science Reviews 27:2473–83.

Goebel, T., and M. R. Waters  2001  New AMS 14C ages for the Tolbaga Upper Paleolithic site, Transbaikal, Siberia. Current Research in the Pleistocene 17:32–33.

Ermalova, N. M.  1978  Teriofauna Doliny Angary v Pozdnem Antopogene [Terrestrial fauna of the Angara Valley during the Upper Anthropogene]. Nauka, Novosibirsk. (In Russian)

Goebel, T., A. P. Derevianko, and V. T. Petrin  1993  Dating the Middle-to-Upper Paleolithic transition at Kara-Bom. Current Anthropology 34:452–58.

Fagundes, N. J. R., R. Kanitz, and S. L. A. Bonatto  2008  Reevaluation of the native American Mtdna genome diversity and its bearing on the models of early colonization of Beringia. PLOS ONE 3:e3157.

Goebel, T., M. R. Waters, and M. Dikova  2003  The archaeology of Ushki Lake, Kamchatka, and the Pleistocene peopling of the Americas. Science 301:501–05.

Fu, Q., M. Meyer, X. Gao, U. Stenzel, H. A. Burbano, J. Kelso, and S. Pääbo 2013 dna analysis of an early modern human from Tianyuan Cave, China. PNAS 110(6):223–27.

Goebel, T., M. R. Waters, and M. N. Meshcherin  2000a  Masterov Kliuch and the early Upper Paleolithic of the Transbaikal, Siberia. Asian Perspectives 39(1-2):47–70.

Gengin, V. F., and B. T. Petrin  1995  Pozdnepaleoliticheskaia Epokha na Iuge Zapadnoi Sibiri [Upper Paleolithic era of south western Siberia]. Nauka, Novosibirsk.

Goebel, T., M. R. Waters, I. Buvit, M. V. Konstantinov, and A. V. Konstantinov  2000b  Studenoe-2 and the origins of microblade technologies in the Transbaikal, Siberia. Antiquity 74:567–75.

Gerasimova, M. M., S. N. Astakhov, and A. A. Velichko  2007  Paleoliticheskii Chelovek, Ego Material’naia Yul’tura I Prirodnaia Sreda Obitaniia [Paleolithic man, his material culture and environment (Illustrated Catalogue of Pleistocene Hominids Discovered at Russia and Adjacent Regions)]. Nester Istoria, St. Petersburg.

Gomez Coutouly, Y. A.  2011  Identifying pressure flaking modes at Diuktai Cave: A case study of the Siberian Upper Paleolithic microblade tradition. 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. 75–90. Texas A&M University Press, College Station.

Germonpré, M., and L. Lbova  1996  Mammalian remains from the Upper Palaeolithic site of Kamenka, Buryatia (Siberia). Journal of Archaeological Science 23:35–57. Gilbert, M. T. P., D. L. Jenkins, A. Gotherstrom, N. Naveran, J. J. Sanchez, M. Hofreiter, P. F. Thomsen, J. Binladen, T. F. G. Higham, R. M. Yohe II, R. Parr, L. S. Cummings, and E. Willerslev  2008  dna from pre-Clovis human coprolites in Oregon, North American. Science 320:786–89. Goebel, F. E. (Ted)  1993  The Middle to Upper Paleolithic Transition in Siberia. Unpublished Ph.D. dissertation, Department of Anthropology, University of Alaska Fairbanks. Goebel, T.  1999  Pleistocene human colonization of Siberia and peopling of the Americas: An ecological approach. Evolutionary Anthropology 8:208–27. ———  2002a  The “Microblade Adaptation” and recolonization of Siberia during the Late Upper Pleistocene. In Thinking Small: Global Perspectives on Microlithization, edited by R. G. Elston and S. L. Kuhn, pp. 117–31. Archaeological Papers of the AAA, No. 12. American Anthropological Association, Arlington. ———  2002b  The middle to Upper Paleolithic transition in Siberia. In Archaeology and Human Paleoecology in the Holarctic, Anthro-

Goebel, T., and M. Aksenov  1995  Accelerator radiocarbon dating of the initial Upper Palaeolithic in southeast Siberia. Antiquity 69:349–57.

Graf, K. E.  2005  Abandonment of the Siberian mammoth-steppe during the LGM: Evidence from the calibration of 14C-dated archaeological occupations. Current Research in the Pleistocene 22:2–5. ———  2008  Uncharted Territory: Late Pleistocene HunterGatherer Dispersals in the Siberian Mammoth-Steppe. Unpublished Ph.D. dissertation, Department of Anthropology, University of Nevada, Reno. ———  2009  “The Good, the bad, and the ugly:” Evaluating the radiocarbon chronology of the middle and late Upper Paleolithic in the Enisei River valley, south-central Siberia. Journal of Archaeological Science 36:694–707. ———  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 37:210–23. ———  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 T. Goebel and I. Buvit, pp. 33–46. Texas A&M University Press, College Station. Green, R. E., J. Krause, A. W. Briggs, T. Maricic, U. Stenzel, M. Kricher,

78 N. Patterson, H. Li, W. Zhai, M. Fritz, N. F. Hansen, E. Y. Durand, A. Malaspinas, J. D. Jensen, T. Marques-Bonet, C. Alkan, K. Prüfer, M. Meyer, H. A. Burbano, J. M. Good, R. Schultz, A. Aximu-Petri, A. Butthof, B. Höber, B. Höffner, M. Sigemund, A. Weihmann, C. Nusbaum, E. S. Lander, C. Russ, N. Novod, J. Affourtit, M. Egholm, C. Verna, P. Rudan, D. Brajkovic, Z. Kucan, I. Gusix, v. B. Doronichev, L. V. Golovanova, P. L. F. Johnson, E. E. Eichler, D. Falush, E. Birney, J. C. Mullikin, M. Slatkin, R. Nielsen, J. Kelso, M. Lachmann, D. Reich, S. Pääbo  2010  A draft sequence of the Neanderthal genome. Science 328(5979):710–22. Guthrie, R. D.  1990  Frozen Fauna of the Mammoth Steppe. The University of Chicago Press, Chicago. Hoffecker, J. F.  2005  A Prehistory of the North: Human Settlement of the Higher Latitudes. Rutgers University Press, New Brunswick. Holmes, C. 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 T. Goebel and I. Buvit, pp. 179–91. Texas A&M University Press, College Station. Izuho, M., and K. Takahashi  2005  Correlation of Paleolithic industries and paleoenvironmental change in Hokkaido (Japan). Current Research in the Pleistocene 22:19–21. Kashani, B. H., U. A. Perego, A. Olivieri, N. Angerhofer, F. Gandini, V. Carossa, H. Lancioni, O. Semino, S. R. Woodward, A. Achilli, A. Torroni  2013  Mitochondrial haplogroup C4c: A rare lineage entering America through the Ice-Free Corridor? American Journal of Physical Anthropology 147:35–39. Kemp, B. M., S. R. Malhi, J. McDonough, D. A. Bolnick, J. A. Eshleman, O. Rickards, C. Martinez-Labarga, J. R. Johnson, J. G. Lorenz, E. J. Dixon, T. E. Fifield, T. H. Heaton, R. Worl, D. G. Smith  2007 ­Genetic analysis of early Holocene skeletal remains from Alaska and its implication for the timing of the peopling of the Americas. American Journal of Physical Anthropology 132:605–21. Kirillov, I., and A. Derevianko  1998  The Paleolithic of the TransBaikal area. In The Paleolithic of Siberia: New Discoveries and Interpretations, edited by A. P. Derevianko, pp. 137–50. University of Illinois Press, Urbana. Kirillov, I., and A. Kasparov  1990  Arkheologiia Zabaikal’ia: problem i perspectivy (epokha Paleolita) [Archaeology of Transbaikal: Problems and perspectives (Paleolithic era)]. In Khronostratigrafiia Paleolita Severnois, Tsentral’noi, I Vostochnoi Azii I Ameriki, pp. 194–98. Nauka, Novosibirsk. Kitchen, A., M. M. Miyamoto, and C. J. Mulligan  2008  A threestage colonization model for the peopling of the Americas. PLOS ONE 3(2):e1596. Konstantinov, A. V.  2001  Drevnie Zhilishche Zabaikal’ia [Ancient Dwellings of the Transbaikal]. Nauka, Novosibirsk. Konstantinov, M. V.  1994  Kamennyi Vek Vstochnogo Regiona Baikal’skoi Azii. SO RAN, Ulan-Ude—Chita.

Graf Kuzmin, Ya. V., and S. G. Keates  2005  Dates are not just data: Paleoltihic settlement patterns in Siberia derived from radiocarbon records. American Antiquity 70(4):773–89. Kuzmin, Ya. V., P. A. Kosintsev, D. I Razhev, and G. W. L. Hodgins  2009  The oldest directly-dated human remains in Siberia: AMS 14C age of talus bone from the Baigara locality, West Siberian Plain. Journal of Human Evolution 57:91–95. Larichev, V., U. Khol’ushkin, and I. Laricheva  1988  The Upper Paleolithic of Northern Asia: Achievements, Problems, and Perspectives in Western Siberia. Journal of World Prehistory 2(4):359–96. Lbova, L. V.  1992  Brianskii Paleoliticheskii kompleks (k obosnovaniiu arkheologicheskogo raiona) [Brian Paleolithic Complex (with explanation of the archaeological region). In Paleoekologiia I Rasselenie Drevnego cheloveka v Severnoi Azii I Amerike, pp. 162–64. Nauka Krasnoaiarsk. ———  1996  Paleoliticheskoe Mestonakhozhdenie Kamenka. In Novye Paleoliticheskie Pamiatniki Zabaikal’ia, pp. 24–47. Chitinskii Gosudarstvennyi Pedagogicheskii Institut. NG Chernyshevskogo, Chita. ——— 2000 Paleolit Severnoi Zony Zapadnogo Zabaikal’ia [Paleolithic of the northern zone of western Transbaikal]. BNts, Siberian Branch, Nauka, Ulan-Ude. ———  2005  The transition from the Middle to Upper Paleolithic in the western Trans-Baikal. In The Middle to Upper Paleolithic Transitions in Eurasia: Hypotheses and Facts, edited by A. P. Derevianko, pp. 355–71. Institute of Archaeology and Ethnography Press, Novosibirsk. Lisitsyn, N. F. 2000 Pozdnii Paleolit Chulymo-Eniseiskogo Mezhdurech’ia [Late Paleolithic of the Chulymo-Enisei Interfluve]. Trudy Tom II, Institut Istorii Material’noi Kul’tury, RAN, St. Petersburg. (In Russian) Marks, A. E.  1990  The Middle and Upper Palaeolithic of the Near East and the Nile Valley: The problem of cultural transformations. In The Emergence of Modern Humans: An Archaeological Perspective, edited by P. Mellars, pp. 56–80. Cornell University Press, Ithaca. Medvedev, G. I.  1983  Paleolit Iuzhnogo Priangar’ia. Avtoreferat. Doctor of Historical Sciences Dissertation. Institute of History, Philosophy, and Philology, Siberian Branch, Soviet Academy of Sciences, Novosibirsk. ———  1998a  Art from central Siberian Paleolithic sites. In The Paleolithic of Siberia: New Discoveries and Interpretations, edited by A. P. Derevianko, pp. 132–37. University of Illinois Press, Urbana. ———  1998b  Upper Paleolithic sites in south-central Siberia. In The Paleolithic of Siberia: New Discoveries and Interpretations, edited by A. P. Derevianko, pp. 122–32. University of Illinois Press, Urbana. Medvedev, G. I., N. A. Savel’ev, V. V. Svinin  1990  Stratigrafiia, Paleogeografiia, i Arkheologiia Iuga Srednei Sibiri: K XIII Kongressu INQUA 1990. Nauka, Irkutsk.

Krause, J., L. Orlando, D. Serre, B. Viola, K. Prüfer, M. P. Richards, J.-J. Hublin, C. Hänni, A. P. Derevianko, and S. Pääbo  2007  Neanderthals in central Asia and Siberia. Nature 449:902–04.

Medvedev, G., N. Cauwe, G. Vorob’eva, D. Coupe, L. Claes, E. Lipnina, S. Modrie, S. Mukharamov, S. Osadchii, P. Pettitt, P. Rebrikov, E. Rogovskii, V. Sitlivyi, L. Sulerzhitskii, and F. Khenzykhenova  1996  Paleoliticheskoye Mestonakhozhdenie Mal’ta [The Paleolithic location of Mal’ta]. ARCOM, Irkutsk.

Kuhn, S. L.  2002  Paleolithic archaeology in Turkey. Evolutionary Anthropology11:198–210.

Meltzer, D. J.  2009  First Peoples in a New World: Colonizing Ice Age America. University of California Press, Berkeley.

Kuzmin, Ya. V.  2008  Siberia at the Last Glacial Maximum: environment and archaeology. Journal of Archaeological Research 16(2):163–221.

Miller, G. H., J. Brigham-Grette, R. B. Alley, L. Anderson, H. A. Bauch, M. S. V. Douglas, M. E. Edwards, S. A. Elias, B. P. Finney, J. J. Fitzpatrick, S. V. Funder, T. D. Herbert, L. D. Hinzman, D. S. Kaugfam,

Siberian Odyssey G. M. MacDonarld, L. Plyak, A. Robock, M. C. Serreze, J. P. Smol, R. Spielhagen, J. W. C. White, A. P. Wolfe, E. W. Wolff  2010  Temperature and precipitation history of the Arctic. Quaternary Science Reviews 29:1679–1715. Mochanov, I. A.  1977  Drevneishie Etapy Zaseleniia Chelovekom Severo-Vostochnoi Azii [The earliest stages of the populations of man in northeast Asia]. Nauka, Novosibirsk. (In Russian) Muratov, V. M., N. D. Ovodov, V. A. Panychev, S. A. Safarova  1982 Obshchaia Kharakteristika Paleoliticheskoi Stoianki Malaia Syia v Khakasii. In Arkheologiia Severnoi Azii, pp. 33–48. Nauka, Novosibirsk. Nakazawa, Y., M. Izuho, J. Takakura, S. Yamada  2005  Toward an understanding of technological variability in microblade assemblages in Hokkaido, Japan. Asian Perspectives 44(2):276–92. Okladnikov, A. P.  1940  Buret’, Novaia Paleoliticheskaia Stoianka na Angare. Sovetsakaia Arkheologiia 5:290–93. ———  1968  Sibir v Drevnekamennom Veke: Epokha Paleolita. Istoriia Sibiri 1:37–93. ———  1971  Mnogosloinoe poselenie Sanny Mys na R. Ude, v 35 km nizhe Khorinksa, Buriatskoi ASSR (raskopki 1968 goda) [Multicomponent settlement of Sanny Mys on the Ude River, 35 km from Khorinska ]. In Materialy Polovykh Issledovanii Dal’nevostochnoi Arkheologisheskoe Ekspeditsi II, pp. 7–83. Nauka, Novosibirsk. Okladnikov, A. P., and I. I. Kirillov  1980  Iugo-Vostochnoe Zabaikal’e v Epokhu Kamnia I Rannei Bronzy [South-Eastern Transbaikal in the Stone Age and Early Bronze Age ]. Nauka, Novosibirsk. Ono, Y. H. Machida  1987  Quaternary sea-level changes in Japan. In Sea Level Changes, edited by M. J. Tooley and I. Shennan, pp. 182– 224. Institute of British Geographers, London. Orlova, L. A.  1998  Radiouglerodnoe Datirovanie Arkheologicheskikh Pamiatnikov Sibiri i Dal’nego Vostoka. In Statisticheskii Analiz Pozdnepaleoliticheskikh Kompleksov Severnoi Azii, edited by A. P. Derevianko, Iu. P. Kholiushkin, P. S. Rostovtsev, and V. T. Boronin, pp. 74–88. IAET SO RAN, Novosibirsk. O’Rourke, D. H., and J. A. Raff  2010  The human genetic history of the Americas: The final frontier. Current Biology 20:R202–07. Ovodov, N. D.  1987  Fauna Paleoliticheskikh poselenii Tobaga i Varvarina Gora v zapadnom Zabaikal’e [Fauna of Paleolithic settlements of Tolbaga and Varvarina Gora in western Transbaikal]. In Prirodnaia Sreda i Drevnii Chelovek v Pozdnem Antopogene, pp. 122–40. Nauka, Ulan-Ude. Pitulko, V. V.  2011  The Berelekh quest: A review of forty years of research in the mammoth graveyard in northeast Siberia. Geoarchaeology: An International Journal 26(1):5–32. Pitulko, V. V., E. Iu. Pavlova  2010  Geoarkheologiia i Radiouglerodnaia Khronologiia Kamennogo Veka Severo-Vostochnoi Azii [Geoarchaeology and Radiocarbon Chronology of the Stone Age of the North-East Asia]. Nauka, St. Petersburg. Pitulko, V. V., E. Yu. Pavlova, P. A. Nikolskiy, and V. V. Ivanova  2012 The oldest art of the Eurasian Arctic: Personal ornaments and symbolic objects from Yana RHS, Arctic Siberia. Antiquity 86:642–59. Pitulko, V. V., P. A. Nikolsky, E. Yu. Cirya, A. E. Basilyan, V. E. Tumskoy, S. A. Koulakov, S. N. Astakhov, E. Yu. Pavlova, and M. A. Anisimov  2004  The Yana RHS site: Humans in the Arctic before the Last Glacial Maximum. Science 303(5654):52–56. Powers, W. R. 1973 Paleolithic Man in Northeast Asia. Arctic Anthropology 10(2):1-106. Reich, D., R. E. Green, M. Kircher, J. Krause, N. Patterson, E. Y. Du-

79 rand, B. Viola, A. W. Briggs, U. Stnzel, P. L. F. Johnson, R. Maricic, J. M. Good, T. Marques-Bonet, C. Alkan, Q. Fu, S. Mallick, H. Li, M. Meyer, E. E. Eichler, M. Stoneking, M. Richards, S. Talamo, M. V. Shunkov, A. P. Derevianko, J. J. Hublin, J. Kelson, M. Slatkin, S. Pääbo  2010  Genetic history of an archaic hominin group from Denisova Cave in Siberia. Nature 468:1053–60. Reimer, P. J., M. G. L. Baillie, E. Bard, A. Bayliss, J. W. Beck, P. G. Blackwell, C. Bronk Ramsey, C. E. Buck, C. E. Burr, R. L. Edwards, M. Friedrich, P. M. Grootes, T. P. Guilderson, I. Hajdas, T. J. Heaton, A. G. Hogg, K. A. Hughen, K. F. Kaiser, B. Kromer, G. McCormac, S. Manning, R. W. Reimer, D. A. Richards, J. R. Southon, S. Talamo, C. S. M. Turney, J. van der Plicht, C. E. Weyhenmeyer  2009  IntCal09 and Marine09 radiocarbon age calibration curves, 0–50,000 years cal BP. Radiocarbon 51(4):1111–50. Roebroeks, W., M. Mussi, J. Svoboda, and K. Fennema  2000  Hunters of the Golden Age: The Mid Upper Palaeolithic of Eurasia 30,000– 20,000 BP. University of Leiden, Leiden. Savoleinin, P., Y. Zhang, J. Luo, J. Lundeberg, and T. Leitner  2002 Genetic evidence for an East Asian origin of domesticated dogs. Science 298:1610–13. Semin, M. Iu., S. O. Shelkovaia, and A. A. Chebortarev  1990  Novoe Paleoliticheskoe mestonakhozhenie v g. Irkutske (imeni I. V. Arembovskogo) [New Paleolithic site in the city of Irkutsk (named for I. V. Arembovskii ]. In Paleoetnologiia Sibiri, pp. 114–15. Gosudarstvennyi Universitet, Irkutsk. Seong, C.  2011  Evaluating radiocarbon dates and Late Paleolithic chronology in Korea. Arctic Anthropology 48(1):93–112. Slobodin, S. B.  2011  Late Pleistocene and early Holocene cultures of Beringia: The general and the specific. 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. 91–118. Texas A&M University Press, College Station. Stanford, D. J., and B. A. Bradley  2012  Across Atlantic Ice: The Origins of America’s Clovis Culture. University of California Press, Berkeley. Tamm, E., T. Kivisild, M. Reidla, M. Metspalu, D. G. Smith, C. J. Mulligan, C. M. Bravi, O. Richards, C. Martinez-Labarga, E. K. Khusnutdinova, S. A. Fedorova, M. V. Golubenko, V. A. Stepanov, M. A. Gubina, S. I. Zhadanov, L. P. Ossipova, L. Damba, M. I. Voevoda, J. E. Dipierri, R. Villems, and R. S. Malhi  2007  Beringian standstill and spread of native American founders. PLOS ONE September 2007(9):e828. Tashak, V. I.  2000  Datirovanie verkhnepaleoliticheskogo poselenia Podzvankaia i nachalo verkhnego paleolita na iuge Buriatii. In Problemy Istorii i Kultury Kochevykh Tsivilizatsii Tsentralnoi Asii: Akheologia, Etnologia, pp. 25–29. BNC SO RAN, Ulan-Ude. ———  2002  Podzvankaia: paleoliticheskie materially nizhnego kompleksa (Respublika Buriatia). In Arkheologia i Kulturnaia Antropologia Dalnego Vostoka i Tsentralnoi Azii, pp. 25–33. Institut Istorii, Arkheologii i Etnografii Narodov Dalnego Vostoka, DVO RAN, Vladivostok. Terry, K.  2010  Extreme Measures: Upper Paleolithic Raw Material Provisioning Strategies and Settlement of the Transbaikal Region, Siberia. Unpublished Ph.D. dissertation, Department of Anthropology, Washington State University. Terry, K., W. Andrefsky, Jr., and M. V. Konstantinov  2009  Raw material durability, function, and retouch intensity in the Upper Paleolithic of Transbaikal, Russia. In Lithic Materials in Upper Paleolithic Societies, edited by B. Blades and B. Adams, pp. 256–69. Blackwell Publishing, Hoboken.

80 Tseitlin, S. M.  1979  Geologiia Paleolita Severnoi Azii. Nauka, Moskow. Turner, C. G.  1990  Paleolithic Siberian dentition from Denisova and Okladnikov caves, Altayskiy Kray, U.S.S.R. Current Research in the Pleistocene 7:65–66. Vasil’ev, S. A.  1992  The Late Paleolithic of the Yenisei: A New Outline. Journal of World Prehistory 6(4):337-383. ———  1996  Pozdnii Paleolit Verkhnego Eniseia [The Late Paleolithic of the Upper Enisei]. RAN, IIMK, St. Petersburg. (In Russian) ———  2000  The Siberian mosaic: Upper Palaeolithic adaptations and change before the Last Glacial Maximum. In Hunters of the Golden Age: The Mid Upper Palaeolithic of Eurasia 30,000 – 20,000 BP, edited by W. Roebroeks, M. Mussi, J. Svoboda, and K. Fennema, pp. 173–95. University of Leiden, Leiden. ———  2003a  Faunal exploitation, subsistence practices and Pleistocene extinctions in Paleolithic Siberia. Deinsea 9:513–56. ———  2003b  The Upper Paleolithic domestic structures in Siberia: A critical review of relevant evidence. In Perceived Landscapes and Built Environments: The Cultural Geography of Late Paleolithic Eurasia, edited by S. A. Vasil’ev, O. Soffer, and J. Kozlowski, pp. 155–60. BAR International Series 1122, Archaeopress, Oxford. ———  2011  The earliest Alaskan archaeological record: A view from Siberia. In From the Yenisei to the Yukon: Interpreating Lithic Assemblage Variability in Late Pleistocene/Early Holocene Beringia, edited by T. Goebel and I. Buvit, pp. 119–27. Texas A&M University Press, College Station.

Graf Vasil’ev, S. G., O. V. Kuznetsov, and M. N. Meshcherin  1987  Poselenie Tolbaga (novyi etap issledovanii) [The settlement of Tolbaga (new phase of research]. In Prirodnaia Sreda I Drevnii chelovek v Pozdnem Antropogene, pp. 109–21.Nauka, Ulan-Ude. Vasilevski, A. A.  2005  Periodization of the Upper Paleolithic of Sakhalin and Hokkaido in the light of research conducted at the Ogonki-5 Site. In The Upper to Middle Paleolithic Transition in Eurasia: Hypotheses and Facts: Collection of Scientific Papers, edited by A. P. Derevianko, pp. 427–45. IAE Press, Novosibirsk. ——— 2008 Kamennyi Vek Ostrova Sakhalin [Stone age of Sakhalin Island]. Russian Academy of Science, Siberian Branch, Institute of Archaeology and Ethnography, Sakhalin State University, YuzhnoSakhalinsk. Vereshchagin, N. K.  1977  Berelekhskoe “kladbishche” mamontov [Berelekh mammoth “cemetery”]. Trudy Zoologisheskogo Instituta 72:5–50. Waters, M. R., and T. Stafford  2007  Redefining the age of Clovis: Implications for the peopling of the Americas. Science 315(5815):112–1126. Yi, S., and G. Clark  1985  The “Dyuktai Culture” and new world origins. Current Anthropology 26:1–20. Zenin, V. N., S. V. Leshchinskiy, K. V. Zolotarev, P. M. Grootes, M. –J. Nadeu  2006  Lugovskoe: geoarchaeology and culture of a Paleolithic site. Archaeology, Ethnology and Anthropology of Eurasia 1(25):41–53.

Chapter 5 Technology and Economy among the Earliest Prehistoric Foragers in Interior Eastern Beringia Ben A. Potter1, Charles E. Holmes1, and David R. Yesner2 Abstract In the past decade, the archaeological record of eastern interior Beringia has seen a transformation in our understanding of the earliest foragers. This presentation focuses on new sites, data, and interpretations of technology and economy from the region, including emerging models of landscape use and settlement systems. Patterns of technological continuity and discontinuity from adjacent regions are evaluated. Pre– and post–Younger Dryas occupations can be distinguished in eastern Beringia, although the signatures of these occupations relate more to changes in behavioral organization and land use than to stylistic changes in technology. Intrasite and intersite patterning in lithic assemblages appears to reflect seasonal or activity-specific variation. Regional variation (e.g., differences between interior and north Alaskan assemblages) may reflect colonization patterns on a larger scale, including distinct populations and timing/direction of colonization, though with some inter-regional technological linkages (e.g., microblade technology). Clovis ancestors may be present in Beringia, but they are not easily distinguished through material culture patterns. Faunal analyses presented here indicate subsistence economic change through time, including (1) relatively broad diet breadth in the Bølling-Allerød period, (2) increased diet breadth during the Younger Dryas, and (3) narrowing diet breadth during the post-Younger Dryas/early-Holocene period. This appears to be a Beringia-wide phenomenon that reflects broad effects of climate change and possibly episodic colonization. These data are used to evaluate technological and economic adaptations relating to the initial colonization of Beringia and subsequent expansion into different ecological niches during the Younger Dryas. Keywords:  Beringia, Younger Dryas, Faunal analysis, Paleoeconomy, Technological organization

Introduction

Central Alaska (in eastern Beringia) contains one of the densest concentrations of late-Pleistocene and early-Holocene sites in the Western Hemisphere (Potter 2011). A total of 46 components older than 10,000 cal yr BP have been reported at 33 sites within the Tanana basin, including 8 sites within the upper Nenana basin (Table 5.1, Figure 5.1). This data1 2

University of Alaska Fairbanks University of Alaska Anchorage Corresponding author e-mail:  [email protected]

set provides a counterpoint to Clovis complex analyses that tends to dominate the Paleoindian literature (Meltzer 2009; see also Hoffecker and Elias 2007). Given that many of the Beringian sites have been recently discovered and analyzed, these data have not been well integrated into a broader understanding of colonization of the New World. Ongoing investigations by Potter at Upward Sun River, Mead, and Gerstle River (Potter 2005; Potter et al. 2011a, 2011b), by Holmes at Swan Point and Eroadaway (Holmes 2011, Holmes et al. 2010), and by Yesner and Easton at Little John and Broken Mammoth (Easton et al. 2011; Yesner et al. 2011) have pro-

82

Potter et al.

Figure 5.1 Map of Tanana valley and surrounding ecoregions showing the locations of sites dating between ~14,000 and 10,000 cal yr BP. Ecoregions from Gallant et al. (1995), shaded areas above 1000 m asl.

vided substantial new information on technological and subsistence strategies and paleoenvironments of the first Alaskans. Recent work by Gelvin-Reymiller and Reuther (Wooller et al. 2012) has revealed multiple late-Pleistocene sites near Quartz Lake, central Tanana basin. Additional work in the upper Nenana basin at Owl Ridge and Dry Creek (Graf and Bigelow 2011, Graf and Goebel 2009) has largely substantiated and expanded upon earlier interpretations (Phippen 1988; Powers et al. 1983). In addition, a number of previously excavated sites have been re-analyzed, resulting in considerably older occupations: Teklanika West (Coffman and Potter 2011), Carlo Creek (Bowers and Reuther 2008), and Eroadaway (Holmes et al. 2010). In this paper, we summarize the state of current research at several early components at Mead, Swan Point, and Broken Mammoth sites, and offer our perspectives on subsistence economies and technological organization. We also offer a broad perspective on changes through time in these and other sites in central Alaska, with particular focus on the transition from the earliest components (pre-Clovis or Clovisage) in the Bølling-Allerød interstadial (~14,700–12,900 cal yr BP) to later components associated with the Younger Dryas chrono­zone (~12,900–11,500 cal yr BP) and post-Younger Dryas / early Holocene (~11,500–10,000 cal yr BP). While the effects of the Younger Dryas have been debated for central Alaska (see review in Kokorowski et al. 2008; see also Graf and Bigelow 2011), these periods reflect changing environmental conditions for Beringian inhabitants. Currently, interpretations of eastern Beringian assemblage variability for these periods generally follow two patterns:

1) multiple traditions derived from technological-typological assignments (e.g., Nenana, Chindadn, Mesa, and

Denali complexes) (Goebel et al. 1991; Hoffecker 2011; Hoffecker et al. 1993), or 2) a single broad technological tradition where land use and mobility patterns structure assemblage variation (Holmes 2001; Potter 2011; West 1996). The role of subsistence economy, land-use patterns, and seasonality generally remain poorly linked with technological datasets (Graf and Bigelow 2011; Potter 2008, Yesner 1996, 2006, 2007). We evaluate competing interpretations with new intrasite and intersite data from the lowest cultural zones (CZ4) at Mead, Swan Point, and Broken Mammoth, allowing for investigation of intrasite spatial patterning of lithics, fauna, and features to help elucidate conditioning factors of assemblage variability, and understand broader patterns of adaptation, including the initial colonization of Beringia.

Mead Site

Archaeological investigations at Mead (totaling 128 m2) have focused on understanding organizational properties, isolating activity areas, and evaluating recurring modes of organization (Potter et al. 2011b). The site is well stratified with a concordant suite of 37 radiocarbon dates, and the lowest components/cultural zones (CZ3b, CZ4, and CZ5) are relatively undisturbed by post-depositional disturbances (Potter et al. 2011b, see also Dilley 1998; Gilbert 2011). Cultural zones 4 and 5 date to the Bølling-Allerød interstadial (13,440–13,200 and 13,110–12,790 cal BP) while CZ3b dates to the Younger Dryas stadial (12,120–11,850 cal BP) (Potter et al. 2011b). A plan view for all cultural materials associated with CZ4, i.e., directly associated with the lower paleosol complex, is illustrated in Figure 5.2. Three hearths yielded statistically similar ages at ~11,100 14C yr BP (13,100–12,800 cal yr BP) (Table 5.2). Three lithic activity areas and four large but dis-

Technology and Economy among the Earliest Prehistoric Foragers in Interior Eastern Beringia

83

Table 5.1  Late-Pleistocene and early-Holocene components in the Tanana basin. Site Component Age, 14C yr1 No. 14C dates Age, cal yr BP2 Excavation area (m2) Reference Swan Point CZ4 12,160  ±  20 8 14,150–13,870 65 Holmes 2011; this paper Little John Sub–paleosol 12,020 ± 70 1 14,050–13,720 ~100 Easton et al. 2011 Mead CZ5 11,460 ± 50 1 13,440–13,200 156 Potter et al. 2011b; this paper Broken Mammoth CZ4 11,440 ± 60 2 13,430–13,160 326 Holmes 1996 Upward Sun River C1 11,320 ± 30 3 13,300–13,120 12 Potter et al. 2011a Walker Road C1 11,220 ± 90 3 13,310–12,850 200 Goebel et al. 1996 Moose Creek C1 11,190 ± 60 1 13,270–12,880 50 Pearson 1999 Linda’s Point 11,100 ± 40 2 13,120–12,780 13 Sattler et al. 2011 Dry Creek C1 11,120 ± 90 1 13,210–12,730 347 Powers et al. 1983 Owl Ridge C1 11,100 ± 60 2 13,140–12,750 27 Phippen 1988; Graf and Bigelow 2011 Mead CZ4 11,080 ± 20 5 13,110–12,790 156 Potter et al. 2011b; this paper Bachner C1 11,030 ± 70 1 13,100–12,700